You are on page 1of 32

Accepted Manuscript

Design optimization method for tube and fin latent heat thermal energy storage
systems

Ralf Raud, Michael E. Cholette, Soheila Riahi, Frank Bruno, Wasim Saman, Geoffrey
Will, Theodore A. Steinberg

PII: S0360-5442(17)31005-8
DOI: 10.1016/j.energy.2017.06.013
Reference: EGY 11017

To appear in: Energy

Received Date: 11 January 2017


Revised Date: 7 May 2017
Accepted Date: 4 June 2017

Please cite this article as: Raud R, Cholette ME, Riahi S, Bruno F, Saman W, Will G, Steinberg TA,
Design optimization method for tube and fin latent heat thermal energy storage systems, Energy (2017),
doi: 10.1016/j.energy.2017.06.013.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT

1 Design optimization method for tube and fin latent heat thermal
2 energy storage systems

3 Ralf Rauda,, Michael E. Cholettea , Soheila Riahib , Frank Brunob , Wasim Samanb , Geoffrey Willa ,
4 Theodore A. Steinberga

PT
5
a QueenslandUniversity of Technology, 2 George St, Brisbane QLD 4000, Australia
6
b Barbara Hardy Institute, University of South Australia, Mawson Lakes Campus, SA 5095, Australia

RI
7 Abstract

SC
8 The search for optimal phase change materials (PCMs) for latent heat thermal energy storage

9 systems (LHTESS) focuses almost exclusively on the properties of the PCM. This neglects the

U
10 significant contribution of the cost of the containment vessel on the total cost of the LHTESS.

Thus, to accurately assess the thermoeconomic performance of various PCMs, the relationship
11

12
AN
between the cost of the containment vessel and the properties of the PCM must be understood.

13 This paper presents an analytical method for optimizing the design of tube and shell or tube
M
14 and fin containment vessels for the least cost, subject to geometric and performance constraints.

15 One of the key performance constraints is the time to charge, whose evaluation typically requires
D

16 computationally-expensive simulations which are unsuitable for optimization algorithms. To en-

able efficient optimization, a novel closed-form approximation of the charge time is developed and
TE

17

18 validated through numerical simulations.

Subsequently, the optimization methodology is used to investigate the relationship between the
EP

optimal vessel geometry and configuration, fin properties, PCM properties, and cost of two PCMs.

For these PCMs, the use of high thermal conductivity fins is shown to dramatically reduce the
C

total cost of the system. The new methodology is an efficient way to compare the thermoeconomic

performance of PCMs as latent heat storage solutions, allowing for the accurate assessment of
AC

PCMs.

Preprint submitted to Elsevier June 5, 2017


ACCEPTED MANUSCRIPT

common abbreviations
PCM phase change material
PV photovoltaic
LHTESS latent heat thermal energy storage system
HTF heat transfer fluid
Ste Stefan number
Re Reynolds number

PT
Pr Prandtl number
material properties
∆H latent heat of fusion
ρs PCM density

RI
λs PCM thermal conductivity
ρm containment metal density
λm containment metal thermal conductivity

SC
St containment metal yield strength
ρf fin density
λf fin thermal conductivity
cHT
p
F
specific heat of the HTF
HT F
ρ HTF density

U
P0 HTF pressure
design variables
S
Co
AN total storage
linear corrosion allowance
N number of tubes
L length of tube
M
R outer radius of tube
r inner radius of tube
v(f,s) volume fraction (fins, PCM)
D

ws fin half pitch


` width of PCM
∆T temperature difference between PCM and HTF
TE

P(m,f,P CM ) Price per kg (metal, fins, PCM)


calculated
∆He effective latent heat
ρe effective PCM density
EP

λe effective PCM thermal conductivity


CF cylinder factor
N EF non-equilibrium factor
tm time to melt
C

Q heat flow
v HTF velocity
AC

Hr (x, t) Required remaining energy to melt (at position x and time t)


etc.
x distance into the tube
N̄ number of infinitesimal elements
e natural exponent
z the parameters modified in the optimization (`, r, L, etc.)
∆He ρe `2
a 2λe ∆T0 · CF · N EF
2((R+`)2 −R2 )λe
b `2 r 2 cHT F ·ρHT F ·v HT F ·CF ·N EF
p

2
ACCEPTED MANUSCRIPT

19 1. Introduction

20 As climate change accelerates, the need for research into renewable energy has become critical.

21 Electrical generation creates more than a quarter of worldwide emissions [1]. In order to reduce

22 emissions, renewable energy sources that can drive large power grids are being investigated. Solar

PT
23 energy in particular has attracted intensive research over the last few decades. This research has

24 led to the widespread adoption of photovoltaic (PV) power sources, which poses problems for large

RI
25 grids. PV currently lacks cost-effective storage technologies; this negatively impacts the ability

26 to provide consistent power [2], and leads to large stresses and inefficiencies in baseload power

SC
27 generation. Large scale adoption of rooftop PV has caused rolling blackouts and brownouts in

28 interconnected power systems. Concentrated solar thermal power mitigates these disadvantages as

U
29 it has storage and controlled output built into the paradigm. In particular, the ability to cheaply

30 store thermal energy has driven significant investment into the research of thermal energy storage

31
AN
systems. Several reviews [3, 4, 5] conclude that latent heat thermal energy storage could be more

32 cost effective than directly storing electricity.


M
33 Latent heat thermal energy storage systems (LHTESS) use the phase change of the storage

34 medium to store large amounts of thermal energy over small temperature ranges. The discharge of
D

35 these systems provides consistent output temperatures, which is important for the efficient operation

36 of heat engines. Several reviews have examined a large range of phase change materials (PCM) as
TE

37 possible storage mediums. In particular, Khare et al. [5] conducted a study in which the properties

38 of numerous known metal alloy PCMs were compared. However, this study did not take into
EP

39 account the cost of the containment material, which is greater than half the cost of the storage

40 system [6, 7], compared to the PCM, which usually is less than 35% of the total cost. Wen et al. [8]

showed that an increase in thermal conductivity from 1 W/mK to 2 W/mK can reduce the amount
C

41

42 of containment material required by nearly 30%.


AC

43 As the PCM is a smaller component of the overall cost, changing the PCM to a more expensive,

44 but higher thermal conductivity option, could reduce the total system cost by reducing the cost

45 of the required containment vessel. This cost reduction balances decreased containment vessel

46 cost with increased PCM cost. This balancing requires a systematic optimization to achieve the

3
ACCEPTED MANUSCRIPT

47 lowest possible total system cost, a problem which has been articulated for some time [9]. Recently,

48 Rathgeber et al. [10] demonstrated that the storage system cost is not primarily a function of the

49 PCM cost, but rather the containment vessel. This further demonstrates that optimized design is

50 necessary. To date, no rigorous optimization has been undertaken. While limited parametric studies

PT
51 have been conducted [9, 11, 12, 13, 14, 15], these studies neglect the interaction effects between

52 design variables (such as thermal conductivity and density), making the resulting configuration

RI
53 suboptimal.

54 The following studies have attempted to optimize the containment vessel design, but fall short of

SC
55 achieving a rigorous optimization. Esen et al. [9] first articulated the problem of the coupled design

56 optimization and PCM properties forcing a joint analysis for efficient selection. They develop a

57 numerical enthalpy based model which considers heat flow in both the radial and axial directions.

U
58 This model is later validated experimentally [16]. A parametric study is conducted which optimizes

59

60
AN
the design for four low temperature PCMs. This study only varies one variable at a time, fixing

the others. This type of parametric analysis leaves large holes in the design space, thus the optimal

designs are not rigorous. Bechiri [17] develop an analytic solution to the tube and shell LHTESS
M
61

62 thermal behavior problem. This analytic solutions promises to be easily optimized, but because it

63 contains exotic functions, must be solved iteratively. Their solution closely matches experimental
D

64 results, but neglecting the addition of fins, and an iterative solution, makes it less than optimal for
TE

65 design optimization. This led them to conduct a parametric analysis to optimize design for paraffin

66 based LHTESS. Tehrani et al. [15] explored the feasibility of tube and shell heat exchangers for

LHTES in the temperature range of 286-565 °C. They sought to determine the optimal parameters
EP

67

68 for the design while minimizing the heat transfer area. This optimization was conducted with a

69 numerical model developed by Esen et al. and using a time step of approximately 0.01 seconds. To
C

70 determine optimal properties, 36 simulations were performed over a range of parameters. Unfortu-
AC

71 nately, only 12 simulations were performed for each length. For each length, L/D ratios from 10 to
R
72 100 were covered and similarly large ranges of r0 . Again, this porous exploration yields important

73 general trends in regards to system performance, but neglects to fully optimize the design variables.

74 Campos-Celador et al. [18] develop an analytic solution for the heat transfer behavior of a finned

4
ACCEPTED MANUSCRIPT

75 plate LHTESS. This analytic solution collapsed each cell to a single node, drastically reducing the

76 computation time as compared to a strictly numerical solution. Campos-Celador et al. recognized

77 that optimization requires solutions which are computationally cheap; they succeeded in creating

78 a model which agreed with numerical experiments and experimental data. Their model is limited,

PT
79 however, to rectangular enclosures which are not space efficient when the most expensive compo-

80 nent is the containment material. Nithyanandam and Pitchumani [11] have conducted a rigorous

RI
81 optimization. Their work studies the geometries of encapsulated PCMs and PCMs with embedded

82 heat pipes. For both geometries, their solution was a numerical model based on balanced heat flow

SC
83 equations discretized into a numerical grid. Their heat flow model takes into account all of the heat

84 transfer processes which occur in the system: convective heat transfer between the heat transfer

85 fluid (HTF) and the tank filler, the radial thermal conduction into the wall of the encapsulated

U
86 PCMs, and the conduction and convection inside the PCM. Once discretized, the behavior of the

87

88
AN
LHTESS is iteratively solved using a non-dimensional time step of 5 · 10−5 . Campos-Celador et al.

[18] have shown that iterative numerical solutions are extremely resource intensive, especially for

the fine grid structures that Nithyanandam et al. [11] uses. In addition, Nithyanandam’s optimiza-
M
89

90 tion is not entirely rigorous; they perform parametric studies by varying one variable at a time. In

91 addition, their search driven optimization uses advanced search space functions with parameters
D

92 which require careful fine tuning. Without a doubt, their results are accurate for the particular
TE

93 cases studied, but applying these results to a wider selection of PCMs would be difficult.

94 In addition to the aforementioned shortcomings, none of the studies mentioned so far attempt

to optimize fin and tube geometries. Optimizing tube and fin geometries is extremely important
EP

95

96 because several studies have found that fins dramatically increase the heat transfer over tube and

97 shell geometries [13, 14, 19, 20]. Few studies have attempted to optimize fin and tube geometries
C

98 [13, 14]. Almsater et al. [13] investigated the effectiveness of finning tubes via a numerical model.
AC

99 Their numerical model is based on the effective heat capacity method and utilizes a network of

100 resistors to determine the heat flow. Their results bear up the large increases in effectiveness and

101 charging capacity that can be obtained with the use of fins. However, their numerical model consists

102 of approximately 60,000 cells. As it will be shown later, this suggests computation times on the order

5
ACCEPTED MANUSCRIPT

103 of days for each simulation. Hubner et al. [14] consider the economic impact of longitudinal tree

104 branching fins in a tube and shell geometry. They perform an optimization to minimize the levelized

105 cost of electricity of a LHTESS coupled with a steam generator. Their optimization is unique in

106 several ways. Firstly, they consider multiple longitudinal fin geometries. Secondly, they discard

PT
107 the circular outer perimeter simplification present in many analyses. In addition, they consider

108 multiple differing fin and tube materials. However, they assume a fixed temperature HTF and fixed

RI
109 tube length. Both together present an unrealistic scenario with extremely poor effectiveness. This

110 simplification will fail to account for the complex association between cost and PCM properties

SC
111 such as density and thermal conductivity. Finally, this analysis, like many before, uses an iterative

112 solution which makes precise cost estimates difficult to ascertain.

113 Thus, no design methodology has been presented which can accurately optimize the performance

U
114 of a PCM heat exchanger with reasonable computational requirements. In a review of the known

115

116
AN
design methodologies Castell et al. [21] concluded that the effectiveness-NTU method showed the

greatest promise of the available methods. Amin et al. [22] and Tay et al. [23] have developed

empirical correlations which relate the effectiveness to the mass flow rate of the HTF. This correla-
M
117

118 tion is accurate for all design cases of tube in shell PCMs, and has a consistent systematic error of

119 approximately 11%. Thus, this correlation can be used to analytically optimize design constraints.
D

120 However, the effect of radial or longitudinal finned tubes is not completely characterized. To be
TE

121 precise: Using this P factor, the PCM system can be designed and optimized using the e-NTU

122 method developed. Further analysis is needed to develop a more generalized equation [for the P

factor] [19].
EP

123

124 Thus, the effectiveness-NTU method cannot compare PCMs with different thermal conduc-

125 tivities across all geometries. In addition, the characterization of tube and fin geometries via the
C

126 effectiveness-NTU method requires experimental correlations which do not exist for all design cases.
AC

127 The use of numerical methods also has limitations; these methods take time to develop for specific

128 geometries and require large amounts of computational power to accurately solve problems [18].

129 Attempting to use these methods to optimize heat exchanger design for large numbers of potential

130 PCM candidates would prove time and resource intensive.

6
ACCEPTED MANUSCRIPT

131 In this paper, a method to optimize the design of a PCM heat exchanger for least cost is

132 presented. This method relies on a novel closed-form analytic expression to determine the melting

133 time of a PCM in a tube and fin or tube in shell heat exchanger. This analytic expression, which

134 depends wholly on the geometry of the heat exchanger and the properties of the materials, is used

PT
135 as a non-linear constraint to minimize the volume of the fins and or tubes which comprise the heat

136 exchanger, to determine the least cost design for the given PCM and performance targets. This

RI
137 optimization is computationally cheap, rigorous, and precise. This method allows for PCMs to be

138 compared based on the total cost of the system, rather than simply the properties of the PCM

SC
139 itself, yielding a more accurate comparison.

140 The paper is organized as follows: First, the optimization problem statement is given. Subse-

141 quently, A novel closed-form solution for the time of melting of the PCM in the heat exchanger is

U
142 derived. This analytic melt time method is validated with CFD results. Finally, the optimization

143

144
AN
method is used to compare the cost performance of two PCMs for use in Solar Thermal Energy Stor-

age. These two PCMs are compared under a variety of different conditions, and the relationships

between their properties and the cost of the system is discussed.


M
145

146 2. Optimization Problem Statement


D

147 To determine the least cost design of the containment vessel, the primary cost is assumed to
TE

148 be the cost of the metal and the cost of the PCM. The cost is assumed to be primarily a function

149 of the cost of the raw materials of metal and the PCM, as previous studies have shown that the
EP

150 additional costs, such as manufacturing, are relatively constant and small [7]. Other studies have

151 found the manufacturing costs to be large [14], but proportional to the material cost and consistent

across different geometries and configurations. Thus, the material costs can be inflated to account
C

152

153 for this discrepancy.


AC

154 The general geometry used is the same as depicted by Esen et al. [9]1 The outer containment

155 shell is assumed to have a minimum thickness equal to the corrosion tolerance (Co). Thus the cost

1 Fig. 1, Esen et al., Sol Ener 62 (1998)

7
ACCEPTED MANUSCRIPT

156 is determined to be:

 †
vf ρf S h
Cost = Pf N Lπ(R2 − r2 ) + 2Co · Lπ(R + `)N 0.5
+
vs ρs ∆H
 i‡  S ?
+ 2Co · πN (R + `)2 · ρm · Pm + PP CM (1)

PT
∆H

157 where v is the volume fraction, ρ is the density, S is the total stored energy, ∆H is the latent

RI
158 heat, P is the cost per kilogram, N is the number of tubes, L is the length of tubes, R is the outer

159 radius of the tube, r is the inner radius of the tube, ` is the width of the PCM, N is the number

SC
160 of tubes, and the subscripts f , m, and P CM refer to the fins, containment material, and PCM

161 respectively. For clarity, the cost of the fins has been denoted with †, the cost of the tubes and shell

162 with ‡, and the cost of the PCM with ?. For a finless system, the total cost is given by:

U
AN
Cost = N · Lπ(R2 − r2 ) + 2Co · Lπ(R + `)N 0.5

+2Co · πN (R + `)2 · ρm · Pm +
 S
PP CM (2)
∆H
M
163 The optimization problem can be stated as follows;
D

min Cost(z)
z
TE

such that:

P (z) ≥ P0 (3)
EP

QP
in
CM
(z) = QHT F
out (z) (4)

v HT F = vmin
HT F
(5)
C

M elt(z) ≤ M eltmax (6)


AC

† The volume fraction of both the fins and PCM are known, so the ratio of these relates the volume of the PCM

to the volume of the fins. The PCM volume is fixed given a specific storage size S, so the volume of PCM multiplied
by the volume fraction ratio, multiplied by the density of the fins gives the mass of the fins.
‡ The cost of the containment vessel is given by multiplying the number of tubes times the volume of each tube,

then adding two terms for the outer containment vessel. All three terms are then multiplied by the density to
determine the mass.

8
ACCEPTED MANUSCRIPT

164 where z = (`, L, r) if cost is given by Eq. (2) or z = (`, L, r, vs , vf ) for cost given by Eq. (1).

165 The constraint Eq. (3) requires that yield strength of the tube, P (z) be greater than the oper-

166 tating pressure of the heat transfer fluid, P0 . Barlow’s formula is used to determine the necessary

167 thickness of the HTF tubes:

PT
St
R=r· + Co (7)
St − P0

168 where P0 is the pressure of the HTF and St is the yield strength of the containment material

RI
169 multiplied by a safety factor.

170 Eqs. (4) and (5) require that the HTF be able to expell heat faster than the PCM can accept it.

SC
171 This assures that the limiting heat flow condition will be the amount of heat that can be transferred

172 into the PCM. The HTF velocity is set via the Dittus-Boelter correlation (assuming turbulent flow):

U
((R + `)2 − R2 )vs ρs ∆H
QP
in
CM
(z) =

QHT F
AN
out (z) = 0.023Re
0.8
tm (z)

P r0.333 λHT F ∆T0 (8)


M
173 where tm (z) is the time to melt, Re is the Reynolds number, P r is the Prandtl number, and ∆T0

174 is the difference between inlet temperature of the HTF and the PCM melting temperature. Thus
D

175 constraints (4) and (5) implicitly set the velocity due to the presence of the Reynolds number in

QHT F
out .
TE

176

177 Eq. (6) constrains the time to melt, as recent reviews by the Sunshot Initiative suggest that

178 charging and discharge time is a primary consideration in the development of thermal energy storage
EP

179 for solar thermal power. This constraint can be constructed by numerical or CFD methods, however,

180 as previously discussed these are time and resource intensive. Instead, an analytic expression is
C

181 developed for the time of melting. The constraint can be modified for the time of solidification,

182 but the PCMs of interest to this study have significantly reduced thermal conductivity in the liquid
AC

183 state. This causes the melting process to take longer.

9
ACCEPTED MANUSCRIPT

184 3. Time of Melting Constraint

185 Bauer [24] developed an analytic equation to determine the time of solidification for a PCM in fin

186 and tube arrangements with isothermal walls. These results assume the PCM starts at the melting

temperature, the tube wall is isothermal, and the contribution of the specific heat is negligible.

PT
187

188 Bauer’s results are based on the thermophysical properties of the PCM and the geometry of the

189 containment design:

RI
∆He ρe `2
tm = N EF · CF · (9)
2λe ∆T

SC
190 where tm is the time to melt. The other variables are given by:

U
 1/4 
`ws 2 vf λf
N EF = 0.8 +1
R ` vs λs

CF = ln

`
R
+1
AN
R
`
2 
+1 −
1 R
2
+
`


ρe = vs ρs + vf ρf
M
vs ρs
∆He = ∆H
ρe
D

λe = vs λs + vf λf (10)
TE

191 where ` is the length of the fin, R is the radius of the pipe, ws is the half-pitch between the fins, vs

192 is the volume fraction of PCM, vf is the volume fraction of the fin, λs is the thermal conductivity
EP

193 of the PCM, λf is the thermal conductivity of the fin, ∆H is the latent heat of fusion of the PCM,

194 ∆T is the temperature difference between the PCM and the isothermal wall, ρs is the density of

195 the PCM, ρf is the density of the fins.


C

196 Fig. 1 visually demonstrates these variables. Bauer has verified the accuracy of his analytic
AC

197 expressions with numerical methods for a large range of input parameters. The numerical methods

198 were verified with comparison to experimental data. These results were accurate for Stefan numbers

199 < 0.1, requiring that ∆T be ≈ 5°C, however, as will be shown later, calculations with Ste< 0.25

200 produce consistently precise results.

10
ACCEPTED MANUSCRIPT

201 Based on previous experiments [25] relating the times of melting and solidification of PCMs,

202 neglecting the influence of convection2 in the PCM on the result (as Bauer does), the relationship

203 between the melting and solidification times is dominated by the relationship of the thermal con-

204 ductivity and densities of the liquid and solid phases. In this paper, the liquid thermal conductivity

PT
205 and liquid density are used in Bauer’s analytic results to yield the melting time. These properties

206 are assumed to be constant as the temperature of the PCM does not vary significantly. In addition,

RI
207 the properties of the HTF are also assumed to be constant with temperature. In the case of the

208 HTF used in Section 4 this is clearly the case [26].

SC
209 In the following derivation Bauer’s results will be extended to heat exchangers with non-

210 isothermal walls. The objective is to describe how the melting proceeds along the tube. In order to

211 determine the time of melting at the exit of the tube, a model of the HTF temperature is required.

U
212 To develop this model, a number of assumptions must be made. First, the heat flow is assumed to

213

214
AN
be independent of the melt fraction [27, 28]. Second, the PCM is assumed to start at the melting

temperature, but solid. Then define x is as the distance into the tube (x = 0 is the inlet). As

the HTF moves through the tube, energy is transferred from the HTF to the PCM, reducing the
M
215

216 difference in temperature between the HTF and the immediately adjacent PCM.

217 In order to utilize Bauer’s results, the length of the tube L is split into discrete elements with
D

218 length ∆x, as shown in Fig. 2. Thus, each element at position x can be assumed to be an annulus
TE

219 with an isothermal wall. Then, Eqs. (9),(10) can be used to find the “time to melt” for that element

220 tm (x, t). However, this “time to melt” is simply an approximation of the remaining time for the

element to completely melt, as the temperature of the HTF at a point is not constant in time.
EP

221

222 Instead, the heat transfer rate from the HTF to the PCM is related to the temperature difference

223 (∆T (x)). Assuming that all of the heat lost by the HTF is transferred into the PCM, then the
C

224 energy flowing into each PCM element reduces the temperature of the corresponding HTF element.
AC

225 Letting z = (x, 0), QP


in
CM
in Eq. (8) describes the heat flow into a single PCM element:

2 This assumption is discussed further in Section 3.2

11
ACCEPTED MANUSCRIPT

Total Energy ((R + `)2 − R2 )ρs ∆He · dx


QP
in
CM
= = = r2 cHT
p
F
· ρHT F · v HT F · d∆T (x, 0) (11)
Time to melt tm (x, 0)

PT
226 rearranging yields the temperature profile of the HTF:

d∆T (x, 0) ((R + `)2 − R2 )ρs ∆He


= 2 HT F HT F HT F (12)

RI
dx r cp ·ρ ·v · tm (x, 0)

227 at the beginning of the charging process (t = 0).

SC
228 Since tm (x, t) depends on ∆T (x, t), via Eq. (9), (12) is found to be a constant coefficient

229 differential equation in x, which has the solution:

U
∆T (x, 0) = ∆T0 · e−bx

b=
`2 r2 cHT
p
AN
2((R + `)2 − R2 )λe
F · ρHT F · v HT F · CF · N EF
(13)

Rearranging (9) and (13) the time of melting at t = 0 can be found;


M
230

tm (x, 0) = a · ebx (14)


D
TE

∆He ρe `2
a= · CF · N EF (15)
2λe ∆T0
EP

231 where a is simply the time for the first element, x = 0 to melt.3 Note that this expression only

232 describes the “instantaneous” time to melt, assuming that the HTF temperature profile stays

233 constant. This is not true, as once the first element completes melting, the HTF temperature
C

234 profile will change.


AC

235 Assuming that the PCM starts at the melting temperature (i.e sensible heat is negligible), the

3 This is the time of melting (Eq (9)) expressed using the inlet temperature differential (∆T ), which is the
0
temperature differential at x = 0.

12
ACCEPTED MANUSCRIPT

236 melting at x = 0 is clearly independent of what happens further down the tube. Thus, prior to the

237 melting of the first element, the “instantaneous” time remaining to melt for the elements further

238 down the tube can be expressed as:

PT
tm (x, t) = a · e−bx − t ∀t≤a (16)

After the melting is complete at x = 0, the sensible heat stored is neglected by assumption,

RI
239

240 implying that there is no transfer of energy out of the HTF into the PCM. Thus, the temperature

gradient of the HTF will start to move. Fig. 3 shows that, for t > a, if the sensible heat is neglected

SC
241

242 the temperature differential will not change as the HTF crosses melted elements.

243 Because the optimization is based on the time to melt of the entire PCM, the time taken for the

U
244 last element to melt must be determined. However, the melting of the last element is complicated

245

246
AN
by the fact that after t = a the HTF temperature at x = L changes constantly with time. Instead,

the focus will be on the progression of melting through the tube. Clearly, once the first element

247 melts, the subsequent elements will complete melting in order. The time when the “melt front”
M
248 reaches the end of the tube yields the time of melting.

249 To describe the progression of the melt front, the time ∆t1 to move the melt front from x = 0
D

250 to the element at ∆x is described. Assuming a left continuous partition4 , Eq. (16) can be written
TE

251 as:

Hr (x, a) ρe `2
EP

tm (x, a) = · · CF · N EF
∆T (x, a) 2λe

Hr (x, a) = ∆He (1 − e−bx ) (17)


C

∆T = ∆T0 e−bx
AC

252 This is developed by understanding that for t ≤ a the time remaining to melt is proportional

4 ∆T (x, t) for each element is assumed to be ∆T (x, t) at the left edge of the element

13
ACCEPTED MANUSCRIPT

253 to the amount of energy required to melt5 . Thus, the amount of energy required to complete the

254 melting of any element ∆x, at time t = a, is given by Hr (x, a), called the “heat required.” For the

255 melt front to move from x = 0 to x = ∆x, some amount of time is required for the HTF to deliver

256 this heat. This can be computed by substituting Eq. (17) into Eq. (9) in place of ∆He yielding:

PT
∆t1 = a · (1 − e−b·∆x ) (18)

RI
257 where ∆t1 is the amount of time required for the melt front to move from x = 0 to x = ∆x. To

understand how the melt front moves with time, the additional time (∆t2 ) for the next element(2∆x)

SC
258

259 to complete melting must be computed. To determine this, the amount of heat remaining at

260 t = a + ∆t1 must be computed. Note that at t = a, the amount of heat required to complete

U
261 melting of 2∆x can be given by Eq. (17).

262
AN
By Eq. (11), the heat flow rate into element 2∆x can be given by:

Hr (2∆x, a)
Qin (2∆x, t) = ∀ t ∈ (a, a + t1 ] (19)
tm (2∆x, a)
M
263 Recall that for a < t < a + ∆t1 the melt front has moved by one element, thus the new
D

264 temperature profile is given by:


TE

∆T (x, t) = ∆T0 e−b(x−∆x) ∀ t ∈ (a, a + t1 ], x ≥ ∆x (20)

The heat required to complete the melting of element 2∆x at t = a + ∆t1 is given by the heat
EP

265

266 flow rate (Eq. (19)), utilizing the new temperature profile (Eq. (20)), and multiplied by the time

267 elapsed (∆t1 ) and can be computed:


C

∆t1
Hr (2∆x, a + t1 ) = ∆He − ∆He · e−b·2∆x − · Hr (2∆x, a)
AC

(21)
tm (2∆x, a)

5 Consider how Eq. (9) is a linear function of the latent heat. Thus as the latent heat decreases, the time

decreases in a linear manner. In this case, the assumption is that a partially melted PCM is equivalent to a PCM
with proportionally less latent heat, at least in regards to the time required to complete the melting process.

14
ACCEPTED MANUSCRIPT

268 Using a similar development as used for Eq. (18), ∆t2 can be evaluated:

∆t2 = a − (a · e−b·2∆x + a(1 − e−b∆x )e−b∆x )

∆t2 = a · (1 − e−b·∆x ) (22)

PT
269 Note that Eqs. (18) and (22) are identical and the time for the melt front to pass between any

RI
270 two elements is constant. To determine the time taken for the melt front to go a distance L, let

271 ∆x = L/N̄ , and the time to melt at a distance L can be written as:

SC
L
M elt(L) = lim a · N̄ (1 − e−b N̄ ) (23)
N̄ →∞

U
272 where M elt(L) is the time for the melt front to pass to a distance L. Eq. (23) simplifies to:

AN
M elt(L) = a · b · L + a (24)
M
273 via an application of L’Hopital’s rule. This shows that the melt front moves linearly as a function

274 of time. Thus, the constraint Eq (6) is fulfilled by use of Eq. (24).
D

275 3.1. CFD Validation


TE

276 In order to validate Eq. (24), numerical modeling of the melting of a low temperature PCM

277 with pure conduction heat transfer is performed using ANSYS FLUENT [29]. The time dependent

278 Navier-Stokes equations are solved including the latent heat transfer and phase change at solid-
EP

279 liquid interface, and zero velocity in the solid region as introduced in a previous work by Riahi et

280 al. [30].


C

281 The following simplifications are assumed: i) density is constant ii) same density for the solid
AC

282 and liquid phases, iii) the natural convection flow is ignored, incompressible and inviscid flow, iv)

283 internal thermal radiation is negligible. These assumptions were used in a model developed in

284 previous work by Riahi et al. [30], using the solidification and melting option in FLUENT[29].

285 The model, and the assumptions, were validated with experimental data from a study by Jones

15
ACCEPTED MANUSCRIPT

286 et al. [31] and readers are referred to these studies for further details. The same algorithm,

287 schemes, and discretization methods implemented in the validated model are used in this study.

288 The convergence thresholds for all calculations were at least 10−4 for the continuity equation and

289 10−6 for other equations. A two dimensional symmetric grid with 17500 cells is generated from

PT
290 half of the geometry with higher resolution at the boundaries. The time step was set to 0.1 seconds

291 to capture details in the temperature fields in a feasible computation time. The properties of the

RI
292 geometry, PCM, and HTF, used in this correlation are outlined in Table 1. The initial temperature

293 is set to one °C below the melting temperature in order for the PCM to start in the solid phase.

SC
294 As shown in Table 1, the error is consistent across several geometries and HTF velocities.

295 This systematic error is consistent across variable geometries, suggesting that it has little effect on

296 comparisons. In addition, the systematic error is consistent with Bauer’s sensitivity analysis. As

U
297 these CFD tests take on the order of 29-450 hours to complete with standard desktop computing

298

299
AN
power, evaluation of the melt time using CFD would be intractable for optimization, as a large

number of such simulations would be needed. Thus, the analytic solution enables rapid evaluation

and optimization with a modest systematic approximation error.


M
300

301 3.2. Natural Convection


D

302 The assumption that natural convection does not play a large role is not uncommon [9, 14, 32],
TE

303 but engenders some limitations. In certain orientations, the effect of natural convection for tube

304 and shell geometries is large [23, 33], but not in others. Additionally, the effect of natural convection

305 in systems with long and densely packed fins is small [19, 34].
EP

306 In particular, Tay et al. have studied the effect of natural convection in finned tube geometries

307 [19]. They found that neglecting natural convection did not effect the accuracy of their CFD
C

308 simulation. To account for the effect of natural convection in vertically orientated tube and shell

configurations, Tay et al. [23] used an “effective thermal conductivity”. The effective thermal
AC

309

310 conductivity is just the thermal conductivity multiplied by a constant factor. This was successfully

311 applied to a number of geometries, with an uncertainty of 8.2%.

312 To determine if this effective thermal conductivity is a viable means of accounting for natural

16
ACCEPTED MANUSCRIPT

313 convection, Tay et al.’s example is followed, and Eq. (24) is compared to the results of Lacroix’s

314 [27] simulations and experiments. Lacroix developed a computational model using the standard

315 heat balance equations and included the effect of natural convection. Their results were extremely

316 accurate when compared against experiments, and they proceeded to simulate the behavior of

PT
317 numerous geometries. In this study, an effective thermal conductivity factor of 1.33 is utilized and

318 the melt time model (Eq. (24)) is compared to Lacroix’s results. With this factor, an uncertainty of

RI
319 12% is achieved. This compares favorably with Tay et al.’s results. If required, an effective thermal

320 conductivity can be substituted for λe in Eq. (24). However, ignoring natural convection gives

SC
321 more consistent uncertainty, as seen in the previous section, and thus is better for comparison of

322 similar PCMs.

U
323 4. Comparison of Eutectic Salt PCMs

324
AN
The optimization formulation developed above will now be used to compare the economic per-

325 formance of two systems with different eutectic salt PCMs. This optimization is performed via
M
326 the Matlab function “fmincon.” No upper bound is set. The lower bound is set to 0.1mm for each

327 variable to prevent divide by zero errors. The function and constraint tolerances are set to 1e-8.
D

328 The properties of these PCMs are listed in Table 2. The two PCMs are referred to by their differing

329 anion. This investigation focuses on the potential of a supercritical CO2 Brayton cycle heat engine,
TE

330 thus, sCO2 operating at 120 bar is chosen as the heat transfer fluid. The charging inlet temperature

331 is set to 660 °C. Both potential PCMs, referred to herein by the differing anion, have melting points
EP

332 around 630 °C. Thus, the difference between the PCM and the HTF is set to 30 °C. Inconel 601 is

333 chosen as the tube and containment materials for its high fatigue strength at high temperature, low

creep, and corrosion resistance. The yield strength listed in Table 2 includes a 0.72 safety factor.
C

334

335 The corrosion allowance is an estimate determined from a recent review [35]. The effect of the
AC

336 corrosion allowance on the cost is elaborated on later. The time to melt is set at 6 hours based on

337 SunShot targets.

338 Fig. 4 shows the relationship between cost per kWh and the total desired storage. Clearly,

339 at small storage capacities, the external containment cost has a large impact on the total cost.

17
ACCEPTED MANUSCRIPT

340 However, this effect is quickly reduced as the storage capacity grows. As shown in Fig. 4 the cost

341 of the CO3 PCM system is less, regardless of storage capacity. This implies that the 20% higher

342 thermal conductivity of the CO3 PCM provides a benefit which outweights the fact that its cost

343 is double that of the other PCM. Important to note is that the energy density, in Jm−3 , for both

PT
344 PCMs is nearly identical.

345 Fig. 5 demonstrates that the superiority of CO3 carries over to a tube and fin configuration

RI
346 wherein the fin is made of the same material of the tube. For this optimization, note that the
ws
347 analytic expressions by Bauer require that ` < 0.5 and vf > 0.01, so this solution does not

SC
348 terminate in the finless system. Nevertheless, of note is that the cost actually increases for this

349 configuration, but this cost increase is not proportionally the same between the two PCMs. The

350 CO3 system experiences a cost increase of approximately 5%, while the SO4 system cost only

U
351 increases by 1.4%, on average. This is the influence of the thermal conductivity on the cost; the

352

353 CO3 system.


AN
SO4 system benefits from the effectively increased thermal conductivity significantly more than the

Fig. 6 shows the cost per kWh for a tube and fin system with Inconel tubes and Aluminum
M
354

355 fins6 . The cost is less than 35% of the cost of the finless system. This is due to the influence of the

356 cheap thermal conductivity enhancement of the aluminum. The fins, with a thermal conductivity
D

357 500 times greater, greatly enhance the heat transfer. This leads to a smaller required number of
TE

358 HTF tubes, which form the bulk of the total cost of the system. The most interesting result from

359 this series of optimizations is the reversal of the more cost effective system. For this geometry the

SO4 system is more cost effective, by as much as 13%. This is a much greater cost advantage than
EP

360

361 the CO3 system has in the finless configuration. Again, this is the effect of a significantly cheaper

362 PCM when the thermal conductivity advantage is reduced.


C

363 Table 3 enumerates the design parameters for different PCMs and configurations with one MWh
AC

364 of storage. The results show that the optimum ` is similar for the two different PCMs for each

365 configuration. The differences in cost are made up for in large differing L values for tube and

6 Aluminum fins are used as an example. Clearly, in this situation, they would melt. However, other similar

materials may exist, or slightly lower temperatures may be selected

18
ACCEPTED MANUSCRIPT

366 shell configurations. In contrast, for finned configurations the fin pitch appears to be the dominant

367 factor in optimizing the geometry for cost. This demonstrates that, even for geometries designed

368 to increase the effective thermal conductivity, the thermal conductivity difference between the

369 two PCMs has a large impact on the optimum parameters. However, when the effective thermal

PT
370 conductivity is dramatically increased by the aluminum fins, the thermal conductivity difference

371 of the two PCMs is much less influential. This can be seen in the parameters for aluminum fins,

RI
372 which are very close. For this configuration, the Na2 SO4 PCM is more cost effective because it is

373 cheaper.

SC
374 Fig. 7 demonstrates the effect of the linear corrosion allowance on the cost of the system.

375 This linear corrosion allowance is the variable “Co” in Eqs (1) and (2). The corrosion allowance

376 increases the thickness of the tubes and outer shell to account for losses due to corrosion. For this

U
377 comparison, the SO4 system with 104 kWh of storage was used in conjunction with the aluminum fin

378

379
AN
configuration. Clearly, the dominant cost variable in this optimization is the effect of the corrosion

allowance, as varying this parameter from one to three millimeters increases the cost per kWh by

84%. However, Fig. 7(b) is a more interesting result. This line relates the linear corrosion rate to a
M
380

381 cost of metal such that the total system cost remains at $8.87 per kWh. Here it is demonstrated that

382 a tube material which costs 14.42$/kg yields the same total system cost as a material which costs
D

383 5.07$/kg, if the more expensive material has a third the corrosion allowance. This cost relationship
TE

384 demonstrates that expensive and exotic containment materials can be cost effective if they reduce

385 the corrosion rate. In addition, while uncharacterized, the effect on the thermophysical properties

of the corrosion products is undoubtedly greater, and worse for system performance, with a larger
EP

386

387 corrosion allowance. Thus, the investigation of containment materials should focus on solutions

388 which present as little corrosion as possible, with lower cost is a secondary consideration.
C
AC

389 5. Conclusions

390 A new method is presented to optimize for least cost the geometry of a tube in shell, and tube

391 and fin, heat exchangers for latent heat thermal energy storage systems. This optimization is based

392 on analytic solutions for the melting time of a PCM in the aforementioned geometries, and gives

19
ACCEPTED MANUSCRIPT

393 specific insight into the effect of the PCM’s thermophysical properties on the cost of the system.

394 This optimization, because it is based on analytic expressions, is very fast to compute, allowing for

395 easy integration into a system wide parameter optimization. This is in contrast to optimizations

396 presented previously, which are based on CFD or numerical methods, both of which take signif-

PT
397 icantly longer to compute. However, this optimization is limited by two important assumptions.

398 First, the neglect of natural convection implies that the results for tube and shell containment are

RI
399 not rigorous for all arrangements. The effect of convection on these results is a subject for futher

400 study. Secondly, the assumption that the PCM starts at the melting temperature limits the ap-

SC
401 plicability of these results to storage systems which include a significant amount of sensible heat

402 storage. Again, the results contained herein can be expanded, but this is a subject for further study.

403 A comparison was undertaken between two potential eutectic salt PCMs for use in a LHTESS in

U
404 a concentrated solar thermal power plant. This comparison showed that for tube and shell geome-

405

406
AN
tries, the dominant property is the thermal conductivity of the PCM, with a 20% larger thermal

conductivity being enough advantage to overcome a 100% price difference. However, for tube and

fin geometries with large thermal conductivity enhancement, the optimum geometry is based on
M
407

408 the energy density, with the thermal conductivity of the PCM playing a smaller role. The effect

409 of the corrosion allowance on the cost of the system is discussed. The cost of the tube material is
D

410 found to play a large role, and because of the uncertain effects of corrosion products on PCM per-
TE

411 formance, expensive containment materials with large costs are suggested to be further researched,

412 in preference to cheaper containment materials with potentially less corrosion resistance.
EP

413 6. Acknowledgments

The authors would like to thank acknowledge the support of the Australian Government for this
C

414

415 study, through the Australian Renewable Energy Agency (ARENA) and within the framework of
AC

416 the Australian Solar Thermal Research Initiative (ASTRI).

20
ACCEPTED MANUSCRIPT

417 7. Bibliography

418 [1] R.K Pachauri and L.A Meyer. Climate change 2014: Synthesis report. Technical report, IPCC,

419 2014.

PT
420 [2] M.A Eltawil and Z Zhao. Grid-connected photovoltaix power systems: technical and potential

421 problems- a review. Renew Sustain Ener Rev, 14:112–129, 2010.

RI
422 [3] Murat M Kenisarin. High-temperature phase change materials for thermal energy storage.

423 Renewable and Sustainable Energy Reviews, 14(3):955–970, 2010.

SC
424 [4] Ming Liu, Wasim Saman, and Frank Bruno. Review on storage materials and thermal perfor-

425 mance enhancement techniques for high temperature phase change thermal storage systems.

U
426 Renewable and Sustainable Energy Reviews, 16(4):2118–2132, 2012.

427

428
AN
[5] Sameer Khare, Mark DellAmico, Chris Knight, and Scott McGarry. Selection of materials

for high temperature latent heat energy storage. Solar Energy Materials and Solar Cells,

107:20–27, 2012.
M
429

430 [6] R Muren, D A Arias, and B Luptowski. performance based cost modeling of phase change
D

431 thermal energy storage for high temperature concentrating solar power systems. In ASME

432 2009 internationa mechanical engineering congress and exposition, 2009.


TE

433 [7] B R Nandi, S Bandyopadhyay, and R Banerjee. Analysis of high temperature thermal energy

434 storage for solar power plant. pages 438–444, 2012.


EP

435 [8] Shaoyi Wen, Evan Fleming, Li Shi, and Alexandre K da Silva. Numerical Optimization and

436 Power Output Control of a Hot Thermal Battery with Phase Change Material. Numerical
C

437 Heat Transfer, Part A: Applications, 65(9):825–843, 2014.


AC

438 [9] Mehmet Esen, Aydin Durmu, and Ayla Durmu. Geometric design of solar-aided latent heat

439 store depending on various parameters and phase change materials. Solar Energy, 62(1):19 –

440 28, 1998.

21
ACCEPTED MANUSCRIPT

441 [10] Christoph Rathgeber, Stefan Hiebler, Eberhard Lvemann, Pablo Dolado, Ana Lazaro, Jaume

442 Gasia, Alvaro de Gracia, Laia Mir, Luisa F. Cabeza, Andreas Knig-Haagen, Dieter Brgge-

443 mann, lvaro Campos-Celador, Erwin Franquet, Benjamin Fumey, Mark Dannemand, Thomas

444 Badenhop, Jan Diriken, Jan Erik Nielsen, and Andreas Hauer. Iea shc task 42 / eces annex 29

PT
445 a simple tool for the economic evaluation of thermal energy storages. Energy Procedia, 91:197

446 – 206, 2016.

RI
447 [11] K Nithyanandam and R Pitchumani. Design of a latent thermal energy storage system with

448 embedded heat pipes. Applied Energy, 126:266–280, 2014.

SC
449 [12] H.J. Xu and C.Y. Zhao. Thermodynamic analysis and optimization of cascaded latent heat

450 storage system for energy efficient utilization. Energy, 90, Part 2:1662 – 1673, 2015.

U
451 [13] Saleh Almsater, Wasim Saman, and Frank Bruno. Performance enhancement of high temper-

452
AN
ature latent heat thermal storage systems using heat pipes with and without fins for concen-

453 trating solar thermal power plants. Renewable Energy, 89:36 – 50, 2016.
M
454 [14] Stefan Hbner, Markus Eck, Christoph Stiller, and Markus Seitz. Techno-economic heat transfer

455 optimization of large scale latent heat energy storage systems in solar thermal power plants.
D

456 Applied Thermal Engineering, 98:483 – 491, 2016.


TE

457 [15] S. Saeed Mostafavi Tehrani, Robert A. Taylor, Pouya Saberi, and Gonzalo Diarce. Design and

458 feasibility of high temperature shell and tube latent heat thermal energy storage system for

solar thermal power plants. Renewable Energy, 96, Part A:120 – 136, 2016.
EP

459

460 [16] Mehmet Esen. Thermal performance of a solar-aided latent heat store used for space heating

by heat pump. Solar Energy, 69(1):15 – 25, 2000.


C

461

[17] Mohammed Bechiri and Kacem Mansouri. Analytical solution of heat transfer in a shell-and-
AC

462

463 tube latent thermal energy storage system. Renewable Energy, 74:825 – 838, 2015.

464 [18] A Campos-Celador, G Diarce, I González-Pino, and J M Sala. Development and compara-

22
ACCEPTED MANUSCRIPT

465 tive analysis of the modeling of an innovative finned-plate latent heat thermal energy storage

466 system. Energy, 58:438–447, 2013.

467 [19] N H S Tay, M Belusko, A Castell, L F Cabeza, and F Bruno. An effectiveness-NTU technique

for characterising a finned tubes PCM system using a CFD model. Applied Energy, 131:377–

PT
468

469 385, 2014.

RI
470 [20] M. Costa, D. Buddhi, and A. Oliva. Numerical simulation of a latent heat thermal energy stor-

471 age system with enhanced heat conduction. Energy Conversion and Management, 39(34):319

– 330, 1998.

SC
472

473 [21] A. Castell and C Solé. An overview on design methodologies for liquid-solid pcm storage

systems. Renew Sustain Ener Rev, 52:289–307, 2015.

U
474

[22] N.A.M. Amin, F. Bruno, and M. Belusko. Effectivenessntu correlation for low temperature
475

476
AN
{PCM} encapsulated in spheres. Applied Energy, 93:549 – 555, 2012. (1) Green Energy;

477 (2)Special Section from papers presented at the 2nd International Enery 2030 Conf.
M
478 [23] N.H.S. Tay, F. Bruno, and M. Belusko. Experimental validation of a {CFD} and an -ntu model

479 for a large tube-in-tank {PCM} system. International Journal of Heat and Mass Transfer,
D

480 55(2122):5931 – 5940, 2012.


TE

481 [24] Thomas Bauer. Approximate analytical solutions for the solidification of PCMs in fin ge-

482 ometries using effective thermophysical properties. International Journal of Heat and Mass
EP

483 Transfer, 54(23-24):4923–4930, 2011.

484 [25] J.P Kotzé, T.W von Backstr´’om, and P.J Erens. Simulation and testing of a latent heat
C

485 htermal energy storage unit with metallic phase change material. In SOLARPACES 2013:

International Conference on Concentrating Solar Power and Chemical Energy Systems, 2013.
AC

486

487 [26] S. Gupta. Developing 1-d heat transfer correlations for supercritical water and carbon dioxide

488 in vertical tubes. Master’s thesis, University of Ontario Institute of Tehcnology.

23
ACCEPTED MANUSCRIPT

489 [27] M Lacroix. Numerical simulation of a shell and tube latent heat thermal energy storage system.

490 Sol Ener, 50(4):357–367, 1993.

491 [28] Y. Kozak, B. Abramzon, and G Ziskind. Experimental and numerical investigation of a hybrid

pcm-air heat sink. Appl Therm Engin, 59:142–152, 2013.

PT
492

493 [29] ANSYS Fluent Users Guide, 15.0 edition, 2015.

RI
494 [30] S. Riahi and et al. Numerical modeling of inward and outward melting of high temperature

495 pcm in a vertical cylinder. In SOLARPACES 2015: International Conference on Concentrating

SC
496 Solar Power and Chemical Energy Systems, 2016.

497 [31] B.J Jones and et al. Experimental and numberical study of melting in a cylinder. Int J Heat

U
498 Mass Trans, 49(15):2724–2738, 2006.

499

500
AN
[32] M. Belusko, N.H.S. Tay, M. Liu, and F. Bruno. Effective tube-in-tank {PCM} thermal storage

for {CSP} applications, part 1: Impact of tube configuration on discharging effectiveness. Solar

501 Energy, 139:733 – 743, 2016.


M
502 [33] Martin Longeon, Adle Soupart, Jean-Franois Fourmigu, Arnaud Bruch, and Philippe Marty.

Experimental and numerical study of annular {PCM} storage in the presence of natural con-
D

503

504 vection. Applied Energy, 112:175 – 184, 2013.


TE

505 [34] Nourouddin Sharifi, Theodore L. Bergman, and Amir Faghri. Enhancement of {PCM} melting

506 in enclosures with horizontally-finned internal surfaces. International Journal of Heat and Mass
EP

507 Transfer, 54(1920):4182 – 4192, 2011.

508 [35] M. Liu, J.C. Gomez, C.S. Turchi, N.H.S. Tay, W. Saman, and F. Bruno. Determination of
C

509 thermo-physical properties and stability testing of high-temperature phase-change materials


AC

510 for csp applications technologies. Sol Ener Mat Sol Cell, 139:81–87, 2015.

511 [36] R. Raud, M.E Cholette, G. Will, and T.A Steinberg. Analysis of the economic effect of

512 nanoparticle suspension in phase change materials for latent heat thermal energy storage. In

513 Proceedings of the 10th Australasian Heat and Mass Transfer Conference, 2016.

24
ACCEPTED MANUSCRIPT

2wf 2ws

PCM

T
L

HTF

IP
R

CR
`

Figure 1: Geometry of heat exchanger with variables included.

US
AN
M
tm(x, a)
tm(x, 0)

D
TE
EP

0 2∆x 4∆x 6∆x 8∆x L 0 2∆x 4∆x 6∆x 8∆x L


C

Figure 2: Discretization of tube into elements with unique time to melt and length ∆x, at time t = 0 and t = a.
AC

25
ACCEPTED MANUSCRIPT

Inlet Temperature Differential 30 °C


HTF specific heat 4180 J/kgK
HTF density 1000 kg/m3
PCM latent heat 248 J/g
PCM thermal conductivity (l) 0.146 W/mK
PCM density 769 kg/m3

PT
Test # 1 r 0.0025 m ` 0.0175 m L 0.5m v HT F 0.1 m/s 12.8% error
Test # 2 r 0.0025 m ` 0.0175 m L 0.5m v HT F 0.01 m/s 11.8% error
Test # 3 r 0.0025 m ` 0.0175 m L 0.25m v HT F 0.01 m/s 10.9% error

RI
Test # 4 r 0.005 m ` 0.035 m L 0.5m v HT F 0.01 m/s 13.6% error

Table 1: Constants used for the CFD correlation. The wall thickness is set to 0.001 m.

30 °C

SC
Inlet Temperature Differential
HTF pressure 120 bar
HTF specific heat 2000 J/kgK
HTF density 150 kg/m3

U
Inconel Aluminum NaCl + Na2 CO3 NaCl + Na2 SO4
Price ($/kg) 7.5 1.5 0.22 0.11
Density(kg/m3 )
Thermal Conductivity (W/mK)
7780
22
AN 2560
225
1898
0.60
2020
0.494
Latent Heat (J/g) - - 283 266
M
Table 2: Constants used for the optimization[36]
ED

t≤a
∆T0 a + ∆t1 < t < a + ∆t1 + ∆t2
a + ∆t1 + ∆t2 < t < a + ∆t1 +
+∆t2 + ∆t3
∆T (x, t)

PT
CE

∆x 2∆x L

Figure 3: Temperature differential as a function of x, for t ≤ a (solid line). After the first element has melted, ∆T (x, t)
AC

=∆T0 in the range 0 < x ≤ ∆x, and then begins to decrease again. The dotted and dashed lines demonstrate the
change in the temperature differential as a function of x and for the times as listed in the legend.

26
ACCEPTED MANUSCRIPT

37

36

PT
35

34
Cost ($\kWh)

RI
33

32

SC
31

30

U
29
10 100 AN
1000 10000
Total Storage (kWh)
100000 1000000
M

Figure 4: Optimal cost per kWh as a function of total stored energy for two PCM systems in the tube and shell
D

geometry, with the Na2 CO3 PCM cost shown as a dotted line and the Na2 SO4 PCM cost shown as a solid line.
TE
C EP
AC

27
ACCEPTED MANUSCRIPT

38

37

PT
36
Cost ($\kWh)

35

RI
34

SC
33

32

U
31
10 100 AN
1000 10000
Total Storage (kWh)
100000 1000000
M

Figure 5: Optimal cost per kWh as a function of total stored energy for two PCM systems in the tube and fin
D

geometry with Inconel 601 fins, with the Na2 CO3 PCM cost shown as a dotted line and the Na2 SO4 PCM cost
shown as a solid line.
TE
C EP
AC

28
ACCEPTED MANUSCRIPT

20

18

PT
16
Cost ($\kWh)

14

RI
12

SC
10

U
6
10 100 AN
1000 10000
Total Storage (kWh)
100000 1000000
M

Figure 6: Optimal cost per kWh as a function of total stored energy for two PCM systems in the tube and fin
D

geometry with aluminum fins, with the Na2 CO3 PCM cost shown as a dotted line and the Na2 SO4 PCM cost shown
as a solid line.
TE

16
12
a) 14 b)
Affordable Cost ($/kg)

10
EP

12
Cost ($\kWh)

8 10
6 8
6
4
4
C

2 2
0 0
AC

0.5 1.5 2.5 3.5 0.5 1.5 2.5 3.5


Corrosion Allowance (mm) Corrosion Allowance (mm)

Figure 7: Effect of corrosion allowance on optimal cost (a) and the equivalent optimal cost of the containment
material to maintain the same sytem cost as a fuction of corrosion allowance (b)

29
ACCEPTED MANUSCRIPT

PT
RI
SC
Tube in Shell
r ` L N Total Volume $
(m3 )
CO3 0.0021 0.0147 0.1732 32600 7.24 30.603

U
SO4 0.0021 0.0136 0.1589 40100 7.31 32.273

r ` L
AN Inconel Fins
ws N Total Volume
(m3 )
$

CO3 0.00256 0.01784 0.19509 0.00751 20700 7.25 32.12


M
SO4 0.00254 0.0176 0.19231 0.00595 21600 7.29 32.661

Aluminum Fins
r ` L ws N Total Volume $
D

(m3 )
CO3 0.00318 0.05358 0.72778 0.00134 880 7.33 10.83
TE

SO4 0.00321 0.05386 0.72785 0.00132 871 7.34 9.54

Table 3: Optimum parameters based on geometry for 1 MWh of storage. All units are in (m).
C EP
AC

30
ACCEPTED MANUSCRIPT

• New method to minimize the cost of latent heat energy storage systems is developed
• An analytic solution for the time to melt of a PCM in a heat exchanger is developed
• The relationships between the optimal cost and design choices are explored

PT
RI
U SC
AN
M
D
TE
C EP
AC

You might also like