Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Available online at www.sciencedirect.

com

Solar Energy 84 (2010) 974–985


www.elsevier.com/locate/solener

Thermal analysis of solar thermal energy storage


in a molten-salt thermocline
Zhen Yang, Suresh V. Garimella *
Cooling Technologies Research Center, NSF I/UCRC, School of Mechanical Engineering, Purdue University, West Lafayette, IN 47907-2088, USA

Received 19 June 2009; received in revised form 22 February 2010; accepted 8 March 2010
Available online 2 April 2010

Communicated by: Associate Editor Halime Paksoy

Abstract

A comprehensive, two-temperature model is developed to investigate energy storage in a molten-salt thermocline. The commercially
available molten salt HITEC is considered for illustration with quartzite rocks as the filler. Heat transfer between the molten salt and
quartzite rock is represented by an interstitial heat transfer coefficient. Volume-averaged mass and momentum equations are employed,
with the Brinkman–Forchheimer extension to the Darcy law used to model the porous-medium resistance. The governing equations are
solved using a finite-volume approach. The model is first validated against experiments from the literature and then used to systemat-
ically study the discharge behavior of thermocline thermal storage system. Thermal characteristics including temperature profiles and
discharge efficiency are explored. Guidelines are developed for designing solar thermocline systems. The discharge efficiency is found
to be improved at small Reynolds numbers and larger tank heights. The filler particle size strongly influences the interstitial heat transfer
rate, and thus the discharge efficiency.
Ó 2010 Elsevier Ltd. All rights reserved.

Keywords: Solar thermal energy; Energy storage; Thermocline; Molten salt

1. Introduction a liquid state at atmospheric pressures at temperatures


below 315 °C. Although higher HTF operating tempera-
Parabolic-trough solar thermal electric technology is tures are desired in order to achieve higher Rankine cycle
one of the promising approaches to providing the world efficiencies, currently available candidate oils are limited
with clean, renewable and cost-competitive power on a to operating below approximately 400 °C; the pressuriza-
large scale. In a solar parabolic-trough plant, solar thermal tion needed to operate at higher temperatures is prohibi-
energy is collected by troughs in a collector field, and then tively expensive. Hence, fossil fuels need to be used to
delivered by a heat transfer fluid (HTF) into a steam gen- further superheat the steam generated by the oils to the
erator to generate steam for producing electricity in a Ran- higher temperatures desired in the turbine.
kine steam turbine. In early studies and applications of this More viable candidates for high-temperature HTFs are
technology, synthetic oils with operating temperatures molten salts, such as the commercially available HITEC
below 400 °C were used as the HTF in prototype plants (binary) and HITEC XL (ternary). Molten salts can oper-
(Cohen, 2006; Frier, 1999; Kearney, 1989; Price, 2002). ate up to a higher temperature range of 450–500 °C with
For instance, the Solar Electric Generating Stations very low vapor pressures. The use of a molten salt as the
(SEGS) I plant used CaloriaÒ as the HTF. CaloriaÒ is in HTF can achieve a higher output temperature from the col-
lector field, resulting in the generation of steam at temper-
*
Corresponding author. +1 765 494 5621. atures above 450 °C for use in the turbine, which in turn
E-mail address: sureshg@purdue.edu (S.V. Garimella). raises the Rankine cycle efficiency to approximately 40%.

0038-092X/$ - see front matter Ó 2010 Elsevier Ltd. All rights reserved.
doi:10.1016/j.solener.2010.03.007
Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985 975

Nomenclature

CP specific heat, J kg1K1 umag velocity magnitude, ms1


d diameter of thermocline tank, m um mean velocity magnitude at the inlet of filler re-
d0 diameter of connecting tube at the inlet and out- gion, ms1
let of thermocline tank, m
ds diameter of filler particles, m Greek
*
er unit vector in the r direction, – a thermal diffusivity of molten salt, m2s1
*
ex unit vector in the x direction, – e porosity, –
F inertial coefficient, – l viscosity of molten salt, kgm1s1
g acceleration due to gravity, m/s2 m kinematic viscosity of molten salt, m2s1
h thermocline tank height, m q
* density, kgm3
*
h0 distributor region height, m s stress tensor, Nm2
hi interstitial heat transfer coefficient, W m3K1
K permeability, m2 Subscripts
k thermal conductivity, W m1K1 c at the inlet low temperature
p pressure, Pa h at the outlet high temperature
T temperature, K l molten salt phase
t time, s s solid filler phase
*
u velocity vector, ms1 e effective

In comparison, current high-temperature oils generate thermal storage wherein solar thermal energy delivered
steam at 393 °C with a corresponding cycle efficiency of by the Therminol oil from the collector field is transferred,
37.6% (Kearney et al., 2003). The use of molten salts as through a heat exchanger, to molten salt which serves as
thermal storage media allows for higher storage tempera- the storage medium. The expensive heat exchanger may
tures, thereby reducing the volume of the thermal storage be eliminated by employing direct thermal storage. In a
unit for a given storage capacity. Moreover, molten salts direct molten-salt thermal storage system, a single fluid,
are cheaper and more environmentally friendly than cur- e.g., the molten salt, serves as both the HTF and the stor-
rently available high-temperature oils. The major disadvan- age medium, and flows directly between the collector-field
tage of molten salts is their relatively high melting pipes and the thermal storage tanks. The direct solar ther-
temperature (149 °C for HITEC XL relative to 15 °C for mal energy storage approach is attractive for future para-
CaloriaÒ and 12 °C for Therminol VP-1Ò), which necessi- bolic-trough solar thermal power plants both in terms of
tates special measures such as the use of fossil fuels or elec- higher efficiency and lower cost.
tric heating to maintain the salts above their melting In both indirect and direct molten-salt thermal storage
temperatures in order to avoid serious damage to the systems, there are two prevailing design options: two-tank
equipment when solar power is unavailable at night or in storage, and single-tank thermocline storage. In a two-tank
poor weather conditions. storage system, the molten-salt HTF flows from a cold
Thermal energy storage (Gil et al., 2010; Medrano et al., tank, through the oil-salt heat exchange (indirect system)
2010; Esen and Ayhan, 1996) for solar thermal power or the collector field (direct system), to a hot tank during
plants (Laing et al., 2006; Lovegrove et al., 2004; Michels a charge cycle, and flows back from the hot tank, though
and Pitz-Paal, 2007; Luzzi et al., 1999) offers the potential the steam generator, to the cold tank during a discharge
to deliver electricity without fossil fuel backup as well as to cycle. The two-tank molten-salt storage design was used
meet peak demand, independent of weather fluctuations. in the Solar Two demonstration plant (Pacheco and Gil-
The current baseline design for SEGS plants uses Thermi- bert, 1999), and was shown to be a low-risk and cost-effec-
nol VP-1Ò as the heat transfer fluid in the collector field. tive approach. Compared to the two-tank storage system,
Therminol VP-1Ò has a low freezing point of 12 °C, and single-tank thermocline storage offers the potential for sig-
is stable up to 400 °C, which is higher than the operating nificantly reducing storage costs. The thermocline storage
temperature possible with CaloriaÒ. However, it is still dif- approach uses a packed bed (Sanderson and Cunningham,
ficult to use Therminol VP-1Ò as the HTF near or above 1995; Mawire and McPherson, 2009; Mawire et al., 2009;
400 °C in practical applications, due to its undesirably high Singh et al., 2010) in a single tank that is marginally larger
vapor pressure (>1 MPa) which can incur significant costs than one of the tanks in a two-tank thermal storage system
for pressurization of the system. A near-term solution for as used in Solar Two. Buoyancy forces help maintain stable
thermal storage in solar-trough plants is to use indirect thermal stratification between hot and cold molten salts in
976 Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985

the same tank in this one-tank approach. A low-cost filler


material is used to fill most of the thermocline tank volume
and acts as the primary thermal storage medium; this helps
reduce the quantity of the relatively more expensive molten
salt, and presents a significant cost advantage over the two-
tank approach. It was shown using system-level models
(Price, 2003; Kearney and Associates, 2001) that thermo-
cline storage may offer the lowest-cost energy storage
option, saving 35% of the cost relative to the two-tank stor-
age system.
Ideal filler materials for thermocline thermal energy
storage should meet several requirements: low cost, wide
availability, high heat capacity, and compatibility with
the molten-salt HTFs. A wide range of materials, including
quartzite, taconite, marble, NM limestone, apatite, corun- Fig. 1. Schematic diagram of the thermocline thermal energy storage
dum, scheelite and cassiterite, have been considered as can- system under analysis and the axisymmetric coordinate system used.
didates for the filler material in a HITEC XL molten-salt
thermocline storage system (Brosseau et al., 2005). Quartz-
ite rock and silica sand were found to withstand the molten tom of the tank as shown in the figure. HITEC is a eutectic
salt environment with no significant deterioration that mixture of water-soluble, inorganic salts: potassium nitrate
would impact the performance or operability of a thermo- (53 wt.%), sodium nitrite (40 wt.%) and sodium nitrate
cline thermal storage. A demonstration on such a thermo- (7 wt.%). It is in a liquid state above 149 °C (its melting
cline on a pilot-scale (2.3 MW h) was reported in Pacheco temperature) and very stable up to 538 °C. Its physical
et al. (2002). properties, such as viscosity and thermal conductivity,
Although a few studies of molten-salt thermocline ther- change with temperature. HITEC is nonflammable, non-
mal energy storage for parabolic-trough solar thermal explosive and evolves no toxic vapors under recommended
plants have been reported, the thermal behavior and effi- conditions of use, and therefore is considered a potential
ciency of these systems under different operating conditions candidate for molten-salt HTFs used in parabolic-trough
is not yet well-understood. A model that is capable of pre- solar thermal electric plants (Costal Chemical Co.).
dicting the charge/discharge efficiency is needed, as are During the charging (heating) period, hot molten salt
guidelines for the design of molten-salt thermocline systems from the collector field enters the storage tank from the
for parabolic-trough solar thermal electric plants. upper port, transfers heat to the cold filler material, and
The present work develops a comprehensive analysis of exits the storage tank at a lower temperature through the
the discharge dynamics of molten-salt thermocline thermal bottom port. Thermal energy from the collector field is
energy storage for parabolic-trough solar thermal electric thus stored in the filler medium of the storage tank. During
plants. HITEC molten salt is considered as the HTF and discharge, cold liquid is pumped into the storage tank
quartzite rock as the filler in the computations, although through the bottom port, heated by the hot filler medium,
the analysis methodology is valid for any combination of and drawn from the tank at a higher temperature through
salt and filler. The thermal behavior, including temperature the upper port. The thermal energy stored in the filler med-
profiles and discharge efficiency, are specifically investi- ium is thus retrieved by the cold fluid for further use.
gated. Based on results from the model, guidelines are Since energy storage in a thermocline depends on buoy-
developed for the design of thermocline thermal energy ancy to maintain thermal stratification, a uniform flow at
storage systems. the inlet and outlet contributes to effective thermal stratifi-
cation and for improved performance of the thermocline.
2. Development of a thermocline model Therefore, measures for maintaining a uniform flow condi-
tion at the inlet and outlet are needed in practice. In Fig. 1,
The thermocline unit considered for analysis is schemat- two distributor regions (free of filler material) are included
ically illustrated in Fig. 1. The height of the filler region is at the upper and lower ends of the filler region in the stor-
denoted h, h0 is the height of the distributor region, d the age tank. Due to their low flow resistance (compared to the
diameter of the cylindrical tank, and d0 the diameter of filler material region), the distributor regions are expected
the ports, as shown in the figure. An axisymmetric coordi- to lead to a uniform distribution of the flow at the inlet
nate system is used as indicated. The cylindrical thermo- and outlet.
cline thermal storage tank has inlet/exit ports at the The volume-averaged governing continuum and momen-
center of the top and bottom surfaces. The bulk of the tank tum equations for the HTF phase are:
is occupied by a filler material, quartzite rock, at a porosity
of e. A molten-salt heat transfer fluid, HITEC, fills the pore @ðeql Þ *
þ r  ðql u Þ ¼ 0 ð1Þ
volume as well as the unfilled portions at the top and bot- @t
Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985 977

* * *
! *
@ðql u Þ u u * * The physical properties of HITEC change with temper-
þ r  ql ¼ erp þ r  s þeql g ature, and can be calculated according to the following
@t e
  curvefits to experimental measurements reported by Costal
l * F * Chemical Co.:
þe u þ pffiffiffiffi ql umag u ð2Þ
K K
* * * Density : ql ¼ 1938:0  0:732ðT l  200:0Þ ð6Þ
*
* * * * *
* r u þðr u ÞT * @ Viscosity : l ¼ exp½4:343  2:0143ðlnT l  5:011Þ ð7Þ
where s ¼ 2l S  23 lS kk I , S ¼ 2
and r ¼ e r @r þ
*
eh @ * @ Thermal conductivity : k l ¼ 6:53  104 ðT 1  260:0Þ þ 0:421 ð8Þ
r @h
þ e x @x.
In the axisymmetric coordinate system shown
in Fig. 1, the problem is two-dimensional, and the velocity The fits in Eqs. (6)–(8) represent the experimental data with
* * *
vector is u ¼ ur e r þ ux e x and its derivatives in the h-direc- maximum errors of 0.2%, 0.5% and 0.8%, respectively. The
tion are all zeros, i.e., @u
@h
r
¼ @u
@h
x
¼ 0. heat capacity of HITEC is relatively constant at 1561.7 J/
Since the HTF and the filler material may be at different kg-K according to the published data. Properties of quartz-
temperatures due to their distinct thermal conductivities ite rock are treated as constants, with a specific heat capac-
and heat capacities, the energy equation is applied separately ity of 830 J/kg-K and density of 2500 kg/m3 (Heat
to the two phases. For the HTF, the energy equation is: Capacities of Some Common Substances).
In practice, flow distributors are generally employed to
@½eql C P ;l ðT l  T c Þ *
þ r  ½ql u C P ;l ðT l  T c Þ ensure a uniform flow condition at the inlet/outlet of the
@t
" *
! # * * filler region. The characteristic length of the distributor
*
* u * uu (represented by diameter d or height h0 as shown in
¼ r  ðk e rT l Þ  pr  u þtr r s þ
e 2e Fig. 1) is much larger than the particle size in the filler
region. Also, the molten salt is pumped into the distribu-
@ql
 þ hi ðT s  T l Þ ð3Þ tor through the port at a velocity much larger than that
@t in the filler region because of the small cross-sectional
and for the filler, it is: area of the port relative to the open frontal area of the
@ filler region. For these two reasons, a high-Reynolds-
½ð1  eÞqs C P ;s ðT s  T c Þ ¼ hi ðT s  T l Þ ð4Þ number turbulent flow is present in the distributor region;
@t
the flow in the filler region, in contrast, is laminar, at a
The heat transfer between the HTF and the filler is ac- much lower local Reynolds number defined based on
counted for with a volumetric interstitial heat transfer coef- the particle size. To model the turbulent flow in the dis-
ficient hi, which appears as a source term on the right side tributor region, the standard k  e model with a standard
of Eqs. (3) and (4). Heating of the HTF caused by compres- wall function (Launder and Spalding, 1972) is employed.
sion work due to volume expansion/shrinkage, viscous ef- At the interface between the distributor and the filler
fects and kinetic energy changes, as respectively shown in regions, the scalars, i.e., pressure and temperature, and
the second, third and fourth terms on the right side of the fluxes, i.e., mass, momentum and energy, are all held
Eq. (3), is small and less than 104 times the conduction continuous; the turbulent energy kt and its rate of dissi-
or convection terms, but is included for completeness. pation et are set in the usual way for standard wall
In a thermocline using quartzite rock as the filler, the fil- functions, i.e., @k t =@n ¼ 0 and et ¼ 0:090:75 k 1:5
t =ð0:42DlÞ,
ler particles are completely surrounded by the HTF (a con- where o/on is the gradient in the wall normal direction
tinuous phase) and have poor thermal contact with and Dl the distance from the wall to its adjacent cell
neighboring particles; therefore, the filler is treated as a dis- center.
persed phase embedded in a continuous HTF phase. The In the validation presented in the next section, it will be
effective thermal conductivity ke of the HTF-filler mixture shown that the distributor regions do effectively render the
in Eq. (3) can then be expressed as (Gonzo, 2002) flow uniform at the entrance and exit to the filler region. A
1 þ 2b/ þ ð2b3  0:1bÞ/2 þ /3 0:05 expð4:5bÞ thermocline thermal energy storage tank with distributors
ke ¼ kl ð5Þ is thus thermally equivalent to one without distributors
1  b/
k l
but with a uniform flow imposed at the inlet and outlet
where / ¼ 1  e and b ¼ kkssþ2k l
. Eq. (5) provides a good esti- of the filler region.
mate for the effective conductivity of liquid-saturated por- In order to generalize the model and the corresponding
ous media with / ranging from 0.15 to 0.85 and b from results, Eqs. (1)–(4) are non-dimensionalized as follows.
0.499 to 1. Thermal conduction of the filler phase is, how- Continuity equation:
ever, neglected in Eq. (4) because of the large thermal con-
tact resistance between filler particles; conduction within @ðUq Þ *
the filler particles is also neglected due to their small size e þ r~  ðUq U Þ ¼ 0 ð9Þ
@s
(<0.1 m). Temperature changes in the filler particles are
thus solely due to heat transfer from the HTF phase. Momentum equation:
978 Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985

* * *!
@ðUq U Þ Uq U U 0:732ðT h  T c Þ
þ Rer~  Uq ¼ 1  Hl
@s e 2084:4  0:732T c
exp½4:343  2:0143 ln½ðT h  T c ÞHl þ T c  þ 10:094
~
* Ul ¼
~ þ r~  T  eUq Gr*
¼ erP ex exp½4:343  2:0143 ln T c þ 10:094
* ! 6:53  104 ½ðT h  T c ÞHl þ T c  þ 0:5908
Ul U FRe * Ukl ¼
e þ Uq U mag U ð10Þ 6:53  104 T c þ 0:5908
Da2 Da
1 þ 2b/ þ ð2b3  0:1bÞ/2 þ /3 0:05 expð4:5bÞ
Uke ¼ Ukl
1  b/
Energy equation for the molten salt: UCpl ¼ 1; UCps ¼ 1; Uqs ¼ 1

@ * According to Wakao and Kaguei (1982), the interstitial


Pr ðeUq UCpl Hl Þ þ PrRer~  ðUq UCpl Hl U Þ Nusselt number for liquid flow through particle beds can
@s
" * # be expressed by
~ ~ ~
*
~
* * UU @Uq
¼ r  ðUke rHl Þ þ PrA P r  U þtrðr U  T =eÞ þ 1=3
2 @s Nui ¼ 6ð1  eÞ½2 þ 1:1Re0:6
L Pr L  ð13Þ
þ Ukl Nui ðHs  Hl Þ ð11Þ where ReL and PrL are the local Reynolds and Prandtl
numbers, respectively.
Assuming the distributor regions are properly designed
Energy equation for the filler phase: so that uniform flows at the inlet and outlet of the filler
region have been achieved, the appropriate boundary con-
@ ditions are as follows.
Pr ½ð1  eÞXUqs UCps Hs  ¼ Ukl Nui ðHs  Hl Þ ð12Þ At the inlet:
@s
U X ¼ 1; U R ¼ 0; Hl ¼ 0 ð14Þ
The non-dimensional parameters included in Eqs. (9)–(12) and at the outlet:
are defined as follows:
@U X @U R @Hl
¼ ¼ ¼0 ð15Þ
* @X @X @X
tmc x r * u h
s¼ ; X ¼ ; R¼ ; U¼ ; H¼ ;
d 2s ds ds um ds Eqs. (9)–(13) show that heat transfer and fluid flow in a
h 0
d d 0 thermocline storage tank is decided by Re and material
H0 ¼ ; D¼ ; D0 ¼ ; properties, i.e., Uq, Ul, Ukl, Uks, Uke, UCpl, UCps and Pr.
ds ds ds
pffiffiffiffi Once the HTF and the filler particles are selected and the
um d s pd s gd 2s K material properties are determined, the characteristics of
Re ¼ ; P¼ ; Gr ¼ ; Da ¼ ;
mc lc um m c um ds the thermal energy storage process are solely determined
u2m hi d 2s by Re.
A¼ ; Nui ¼ ; The computational domain is discretized into finite
C P ;l;c ðT h  T c Þ k l;c
volumes. All the variables are stored at the centers of
mc C P ;l;c lc Tl  Tc Ts  Tc
Pr ¼ ¼ ; Hl ¼ ; Hs ¼ ; the square mesh cells. A second-order upwind scheme is
ac k l;c Th  Tc Th  Tc
used for the convective fluxes, while a central-differencing
*
qs;c C P ;s;c
* * scheme is used for discretizing the diffusion fluxes. A sec-
* s ds
X¼ ; T ¼ ond-order implicit scheme is used for time discretization.
ql;c C P ;l;c lum Pressure–velocity coupling is implemented through the
*
* @ eh @ * @ PISO algorithm (Issa, 1986). Iterations at each time step
r~ ¼ e x þ þ er : are terminated when the dimensionless residuals for all
@X R @h @R
equations drop below 104. The computations are per-
formed using the commercial software FLUENT (FLU-
Coefficients Uq, Ul, UCpl, Ukl, Uke, Uqs and UCps account for ENT 6.1 Documentation). User-defined functions
the temperature dependence of the density, viscosity, spe- (UDFs) are developed to account for Eqs. (11) and
cific heat, thermal conductivity, effective thermal conduc- (12). Grid and time-step dependence are checked by
tivity of the molten salt, and the density and specific heat inspecting results from different grid densities and time
of the filler material, respectively. These coefficients can intervals. Based on this, DX = DR = 0.01 and
be expressed as follows, according to the data in Costal Ds = 1  103 are chosen as this setting results in a tem-
Chemical Co.: perature along the line R = 0 throughout the discharge
Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985 979

process that is within 1% of that for the case with HTF and the filler particles takes values of as much as
DX = DR = 0.005 and Ds = 5  104. 0.1; this is consistent with the extent of scatter in the exper-
imental measurements and the deviation (0.1) between
3. Model validation the experimental and the numerical temperature profiles
in Fig. 2. Within the experimental uncertainty, therefore,
The experimental results of Pacheco et al. (2002) are the results from the simulations are seen to agree well with
used here to validate the numerical model. A small pilot- the experiments. The flow distributor regions are seen to be
scale, 2.3 MW h, thermocline system was designed and quite effective from the results of simulation 1, since the
built for their experiments. The storage tank was filled with axial temperature profiles at different times in simulation
a mixture of quartzite rock and silica sand resulting in a 1 are almost identical to those for the ideal uniform inlet
porosity of 0.22. A nearly eutectic mixture of soldium and exit flow assumed in simulation 2. It is clear from these
nitrate and potassium nitrate was used as the HTF. The results that fluid flow and heat transfer in a thermocline
non-dimensional parameters for the experiments were as thermal energy storage tank with well-designed distributor
follows: H = 67, H’ = 1.1, D = 33, D’ = 3.3, Re = 220, regions are equivalent to those under uniform inlet and
Pr = 13.4, Gr = 9.59  107, Da = 0.01, A = 1.21  1012. outlet flow conditions.
The numerical results for the axial temperature profiles
are compared with the experimental ones in Fig. 2. Simula- 4. Results and discussion
tion 1 uses the same conditions as in the experiment. Prop-
erty parameters Uq, Ul, Ukl, Uke, UCpl and Pr are taken as The validated numerical code discussed above is
those of HITEC salt, since properties for the exact eutectic employed here to systematically investigate the discharge
mixture (whose composition is close to that of HITEC) characteristics of a thermocline energy storage unit. Based
used in Pacheco et al. (2002) were not provided. In order on the numerical results, a procedure is proposed for
to understand the effectiveness of the flow distributor in designing thermocline thermal storage systems. In the
rendering uniform flow to the filler region, another case results presented in this section, it is assumed that effective
(simulation 2) is considered; conditions for simulation 2 flow distributors have led to the establishment of uniform
are identical to those for simulation 1, except that the mol- flow of molten salt in the filler region in all cases. As dis-
ten-salt flow field is not solved for in the distributor region, cussed above, this condition is readily achieved in practice.
and instead, the flow is set to be uniform at the entrance
and exit of the filler region. 4.1. Temperature profiles
As shown in Fig. 2, the experimental results display
some scatter in the temperature profiles. This may have Typical temperature histories of the molten salt and the
been caused by the contact of some of the thermocouples filler material during a discharge cycle are shown in Fig. 3.
with the rock while others may have been located squarely The temperature profiles at any given time, e.g., s = 1.5,
in the pore centers. Results from the model show that the can be divided into three zones. In the constant, low-tem-
non-dimensional temperature difference DH between the

Fig. 2. Comparison between the numerical and experimental (Pacheco


et al., 2002) axial temperature profiles during discharge of a thermocline Fig. 3. Axial temperature profiles at different times during a discharge
thermal energy storage unit (2.3 MW h): Simulation 1 – with distributors, cycle at Re = 50 and H = 250: Hl – molten salt temperature, Hs – filler
Simulation 2 – with uniform flow at the inlet and outlet of the filler region. temperature.
980 Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985

perature zone near the inlet (X = 0) at the bottom of the


storage tank, both Hl and Hs take values of zero. The next
zone is characterized by significant temperature changes in
both phases (0 < Hl < 1 and 0 < Hs < 1). The final zone is
the constant, high-temperature zone near the exit at the
top of the storage tank, where both Hl and Hs take values
of unity. In the first zone, the filler is completely cooled by
the cold molten salt (Hl = 0) pumped into the storage tank,
and thermal equilibrium exists between the two phases. In
the second zone (referred to hereafter as the heat-exchange
zone), the temperature of the quartzite rock is higher than
that of the molten salt, and thermal energy is transferred to
the cold salt. In the third zone, the salt is fully heated up to
Hl = 1 and is once again in thermal equilibrium with the fil-
ler material.
As time progresses, the intermediate, heat-exchange
zone advances from the inlet towards the outlet, leaving
behind an expanding constant-low-temperature zone and
causing the constant-high-temperature zone in front to
shrink, as shown in Fig. 3. It is convenient to track the
location Xm, where the molten salt is at temperature
Hl = 0.5, as being representative of the heat-exchange zone. Fig. 5. Control volume fixed at the mean-temperature location of
Fig. 4 illustrates the change in the position of the heat- Hl = 0.5.
exchange zone with time at different Reynolds numbers.
In this moving coordinate system, the filler enters the con-
All the results are seen to be well represented by a single
trol volume at the high temperature Th and exits at the low
linear fit passing through the origin, when plotted against
temperature Tc; the molten salt enters the control volume
the product of s and Re. The slope of this line obtained
in the opposite direction as the filler at the low temperature
by linear regression is 1.29.
Tc and exits at the high temperature Th. The net thermal
The slope of the line in Fig. 4 can also be obtained via a
energy flux of the salt and the filler changes the total ther-
simple energy balance on a control volume that covers the
mal energy in the control volume, according to
molten salt and filler in the entire heat-exchange zone, as
shown in Fig. 5. The moving coordinate system has its ori- DE ¼ ql;c u0 Ac e Cpl;c ðT c  T h Þ þ ð1  eÞqs v Ac Cps ðT h  T c Þ
gin fixed at the location Xm, and travels with the control ð16Þ
volume from the bottom to the top of the storage tank.
where DE is the thermal energy change in the control vol-
ume, Ac is the cross-section area of the storage tank, u0
(=um/e  v) is the relative speed of molten salt in the mov-
ing coordinate system, and v is the speed of the traveling
coordinate system which is equal in magnitude to the rela-
tive speed of the filler. Although temperature profiles in the
molten salt in the heat-exchange zone change with time,
they are essentially symmetric about the mid-temperature
point Xm, as will be discussed later. This indicates that
the thermal energy of the salt in this zone changes little
with time, which is also true for the thermal energy of
the filler material in this zone. Therefore, DE is close to
zero so that Eq. (16) becomes
u 
m
 v Ac e Cpl;c ql;c ðT h  T c Þ ¼ ð1  eÞv Ac Cps qs ðT h  T c Þ
e
ð17Þ
which yields
Cpl;c ql;c um
v¼ ð18Þ
e Cpl;c ql;c þ ð1  eÞCps qs
Fig. 4. Representative location Xm in the heat-exchange zone for different
Reynolds numbers at different times. The results are well represented by a Since v is the speed of the traveling coordinate system with
single straight line passing through the origin. its origin fixed in the heat-exchange zone, it is also the
Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985 981

speed of advance of this zone, i.e., dxm/dt = v, where xm is changes little, i.e., DE  0 in Eq. (16), which supports the
the dimensional form of Xm. The slope in Fig. 4 can then be earlier assumption in the derivation of Eq. (17).
rewritten as Increasing the Reynolds number results in an expansion
of the heat-exchange zone. For instance, the temperature
dX m dxm v Cpl;c ql;c
¼ ¼ ¼ ¼ 1:29 profiles at Re = 10 change more gradually in the heat-
dðs  ReÞ dðtum Þ um eCpl;c ql;c þ ð1  eÞCps qs exchange zone than at Re = 1 when compared at the same
ð19Þ position Xm, resulting in a wider heat-exchange zone, as
shown in Fig. 6. At the higher Reynolds number, a longer
This result depends only on the physical properties, and flow distance is needed for the fluid to be heated by the fil-
is identical to the slope obtained from the linear fit to ler phase, leading to a more gradual temperature rise and a
the numerical data in Fig. 4. This validates the assump- corresponding increase in the extent of the heat-exchange
tion made earlier that the total thermal energy in the zone. Since the molten salt in the heat-exchange zone is
heat-exchange zone is essentially invariant, i.e., DE = 0, at a relatively lower temperature, an expanded heat-
during the discharge cycle. Both the analysis above and exchange zone can lead to significant waste of thermal
the data in Fig. 4 indicate that the heat-exchange zone energy if the salt delivered at lower temperatures is not use-
advances at a constant speed from the inlet on the bot- ful for further application. This points to the important
tom to the outlet at the top in a thermocline thermal effects of Reynolds number on the design of a thermocline
storage tank. thermal energy storage system.
Fig. 6 shows the development of the axial temperature The effect of tank height is illustrated in Fig. 7 in terms
profiles plotted in the moving coordinate system, with the of its effect on the temperature history of the molten salt.
horizontal axis being (X  Xm). All the temperature pro- Prior to the heat-exchange zone reaching the tank outlet,
files pass through the point (0, 0.5) in Fig. 6, and appear salt at a constant high-temperature level is available at
to be symmetrical about this point. As the discharge pro- the outlet, i.e., Hl = 1. As the heat-exchange zone arrives
cess proceeds (and the position of the heat-exchange zone at the outlet, the salt temperature begins to drop, finally
Xm increases in Fig. 6), the thermal energy decrease (tem- reaching the constant low temperature level (Hl = 0) when
perature drop) in the region to the right of the point (0, the thermal energy stored in the filler particles has been
0.5) is effectively compensated by an increase (temperature completely depleted. Thermocline tanks with a larger
rise) in the region to left of the point, as shown by a com- height can effectively extend the discharge stage wherein
parison of the profiles at Xm = 65 and 322 at Re = 1 or 10. the salt temperature is maintained at a high level. For
This causes the thermal energy of the molten salt in the instance, the salt temperature begins to drop at s = 5 when
heat-exchange zone to be maintained at a near-constant tank height H is 450, whereas this time is prolonged to
level. This conclusion also holds for the filler phase. There- twice the period (s = 10) with H = 850, as shown in
fore, the total thermal energy in the heat-exchange zone Fig. 7. Since the quality of low-temperature salt is not

Fig. 6. Molten salt temperature profiles in the heat-exchange zone. Fig. 7. Output temperature history of the molten salt with Re = 50 at
Sharper changes in temperature profile occur at lower Reynolds number different tank heights. A thermocline tank of a larger height exhibits a
and small Xm. prolonged constant-high-temperature-discharge stage.
982 Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985

acceptable for generating superheated steam in the tur- tank with a larger height extends the constant high-temper-
bines, it is desired that most of the stored thermal energy ature discharge stage, as shown earlier in Fig. 7, thus
be retrieved at a high-temperature level in order to meet increasing the fraction of initial stored energy that is recov-
the design conditions; this also helps to maintain higher ered as high-temperature thermal energy. At a higher Rey-
thermal-to-electrical conversion efficiency of the turbine nolds number, the heat-exchange zone expands greatly, as
generator. It may be noted that only thermal energy with shown in Fig. 6. For instance, the heat-exchange zone for
temperature above a certain level, e.g., Hl > 0.95 as chosen Re = 10 at Xm = 579 extends over an X  Xm of 200; for
for this work and shown in Fig. 7, is usually considered as a particle diameter of 5 cm, this would imply a zone length
“useful” energy. of 10 m. Since the salt temperature in the heat-exchange
zone is lower than the constant high-temperature level,
4.2. Discharge efficiency an expanded heat-exchange zone reduces the amount of
high-temperature molten salt delivered, and thus decreases
It is of interest to quantify the amount of useful energy the discharge efficiency g.
that a thermocline system can deliver during a discharge The numerical results for the efficiency in Fig. 8 are well
cycle. The discharge efficiency of a thermocline thermal represented by the following correlation:
energy storage system is defined in this work as follows g ¼ 1  0:1807Re0:1801 ðH =100Þm ð21Þ
Output energy with Hl > H0 where m ¼ 0:00234Re0:6151 þ 0:00055Re  0:485.

Total energy initially stored in the thermocline tank This correlation can predict the numerical data within a
ð20Þ maximum error of 1% for Reynolds number between 1 and
50 and H between 10 and 800, as shown by the solid-line
where H0 is a threshold value determined by the applica- predictions from the equation included in Fig. 8.
tion of interest. A value of 0.95 for H0 is chosen in this Three other important parameters which capture the
work, implying that thermal energy delivered at tempera- performance of the thermocline system, i.e., discharge
tures greater than (Tc + 0.95(Th  Tc)) qualifies as useful power per unit cross-sectional area ( P/A), useful thermal
energy. If Th = 450 °C and Tc = 250 °C, for example, HI- energy per unit cross-sectional area (Q/A), and total stored
TEC liquid delivered at temperatures above 440 °C is con- thermal energy per unit cross-sectional area (Qt/A), may be
sidered useful in generating superheated steam for the defined as follows:
steam turbine.
The efficiency defined by Eq. (20) varies depending on P Th  Tc
¼ um ql;c Cpl;c ðT h  T c Þ ¼ k l;c  RePr ð22Þ
the construction and working conditions of the thermo- A ds
cline system. Fig. 8 shows the discharge efficiency calcu- Q
lated for different Re for thermoclines of different heights ¼ ½eql Cpl þ ð1  eÞqs Cps T h ðT h  T c Þhg
A
H. It is clear that the efficiency increases with tank height
¼ ½eql Cpl þ ð1  eÞqs Cps T h ðT h  T c Þd s  H g ð23Þ
H, and decreases with a rise in the Reynolds number. A
Qt
¼ ½eql Cpl þ ð1  eÞqs Cps T h ðT h  T c Þh
A
¼ ½eql Cpl þ ð1  eÞqs Cps T h ðT h  T c Þd s  H ð24Þ

where Q is the useful thermal energy delivered at a temper-


ature above H0, and the thermal properties included in the
square brackets subscripted by Th are calculated at temper-
ature Th. The importance of parameters (P/A) and (Q/A) in
designing a thermocline storage system will be demon-
strated in the following.
From Eqs. (22)–(24), a non-dimensional discharge
power may be defined as RePr, a non-dimensional useful
energy as Hg, and a non-dimensional total energy as H
(equaling the non-dimensional height of the storage tank).
Fig. 9 plots the efficiency under conditions of different dis-
charge power and total energy, which can serve as a guide-
line for the design of thermocline storage systems. The
discharge efficiency is seen to increase with an increase in
the total stored thermal energy and decrease with an
increase in discharge power. If high discharge efficiency is
Fig. 8. Discharge efficiency g of a thermocline at different H and Re. g desired, a thermocline storage unit should be designed to
increases with H and decreases as Re increases. have a large height and operate at a low discharge power.
Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985 983

required to meet the discharge power and thermal energy


output demands may be determined. For instance, a ther-
mocline storage unit with a non-dimensional discharge
power of 600 and a non-dimensional useful thermal energy
of 400 calls for a non-dimensional total thermal energy (or
non-dimensional tank height) of 470.

4.3. Design procedure and examples

The analyses presented thus far are used in this section


to develop procedures for designing a thermocline thermal
energy storage system with HITEC as the molten salt and
quartzite rock for the filler particles, although the proce-
dure itself is generally applicable to other materials sys-
tems. It is assumed that discharge power P and useful
thermal energy Q are predetermined by the application,
and that the rock can be packed to a porosity of 0.22 in
the filler region (Pacheco et al., 2002). The recommended
Fig. 9. Discharge efficiency at different discharge powers (RePr) and total design procedure follows.
thermal energies H. A high discharge efficiency occurs at a low discharge
power and a high total thermal energy. 1. Choose tank diameter d and filler particle size ds based
on practical requirements.
However, this may not be practically feasible as the dis- 2. Calculate the cross-sectional area of the storage tank
charge power would need to be maintained above a certain A = 0.25d2, and then the discharge power per unit
value and the tank height would need to be limited for cost cross-sectional area (P/A) and total thermal energy per
considerations as well as to reduce heat loss from the cor- unit cross-sectional area (Q/A).
respondingly higher tank surface area. The discharge 3. Calculate the non-dimensional discharge power RePr
power and the amount of useful thermal energy would typ- and useful thermal energy Hg, using Eqs. (22) and
ically be determined by the application, leaving other (23), respectively.
parameters to be decided during the design of a storage 4. Calculate Re from the value of RePr and assuming
unit. To facilitate the design under such conditions, H = Hg.
Fig. 10 shows the total thermal energy H under conditions 5. Use the Re and H obtained in step 4 to calculate g from
of different discharge power and useful thermal energy. Eq. (21).
From this figure, the total thermal energy (or tank height) 6. Obtain H by dividing Hg with the efficiency g obtained
in step 5.
7. Repeat steps 5 and 6 until the difference between the
newly obtained H and that in last iteration is smaller
than 0.1%.
8. The final H is the required height for the thermocline
storage tank; also obtained is the discharge efficiency g.
9. The dimensional height of the tank h is calculated as
h = ds  H.

Table 1 shows some examples of thermocline designs


based on this procedure. The storage tank is initially at
450 °C, and cold HITEC at 250 °C (Tc) is fed into the tank.
The output HITEC is at a temperature level of 450 °C (Th)
during the early discharge phase, and later drops in temper-
ature as the thermal energy stored in the tank is depleted.
Thermal energy delivered at a temperature exceeding
440 °C (H0 = 0.95) is regarded as useful energy.
It is observed that an increase in discharge power P
decreases the discharge efficiency g, as evident from a
comparison of cases 1, 2, 3, and 4 in Table 1 with cases
Fig. 10. Total thermal energy H at different discharge power RePr and 5, 6, 7, and 8, respectively. This trend is due to expansion
useful thermal energy Hg. The value of the useful thermal energy is always of the heat-exchange zone at larger powers (also larger
lower than that of the total thermal energy. Re values) as shown in Fig. 6, which reduces the amount
984 Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985

Table 1 Discharge efficiency of the thermocline storage tank is


Results for various thermocline design examples. well predicted by the correlation developed in Eq. (21)
Case No. Q (MW h) P (MW) d (m) ds (m) g (–) (%) h (m) for Reynolds numbers in the range of 1–50 and non-dimen-
1 5 1 2 0.05 83.6 15.2 sional tank heights of 10–800. The efficiency increases with
2 5 1 2 0.1 75.4 16.8 tank height and decreases as Reynolds number increases.
3 5 1 5 0.05 73.4 2.77 Procedures for designing thermocline storage tanks are
4 5 1 5 0.1 61.4 3.31
5 5 2 2 0.05 81.6 15.6
proposed. The use of smaller filler particles can greatly
6 5 2 2 0.1 72.4 17.5 increase the discharge efficiency. For instance, a thermo-
7 5 2 5 0.05 70.5 2.88 cline storage unit (2 MW, 5 MW h and d = 5 m) with a fil-
8 5 2 5 0.1 57.6 3.52 ler particle size of 5 cm has a discharge efficiency that
9 10 1 2 0.05 88.0 28.8 exceeds that with a particle size of 10 cm by 12.9%.
10 10 1 2 0.1 81.6 31.1
11 10 1 5 0.05 80.1 5.07 References
12 10 1 5 0.1 70.5 5.76
13 10 2 2 0.05 86.4 29.4 Brosseau, D., Kelton, J.W., Ray, D., Edgar, M., 2005. Testing of
14 10 2 2 0.1 79.1 32.1 thermocline filler materials and molten-salt heat transfer fluids for
15 10 2 5 0.05 77.8 5.22 thermal energy storage systems in parabolic trough power plants. J.
16 10 2 5 0.1 67.3 6.03 Sol. Energy Eng. 127, 109–116.
Cohen, G., 2006. Nevada first solar electric generating system. In: IEEE
of useful thermal energy delivered. For a specified May Technical Meeting. Solargenix Energy, Las Vegas, Nevada.
<http://ewh.ieee.org/r6/las_vegas/IEEELASVEGASMAY2006.pdf>.
amount of useful thermal energy Q, choosing a larger
Esen, M., Ayhan, T., 1996. Development of a model compatible with solar
tank diameter can effectively reduce the required tank assisted cylindrical energy storage tank and variation of stored energy
height; however, this also decreases the discharge effi- with time for different phase-change materials. Energy Convers.
ciency. This decrease in discharge efficiency is related to Manage. 37 (12), 1775–1785.
the importance of the relatively large extent of the tank FLUENT 6.1 Documentation. <http://www.fluent.com/>.
Frier, S., 1999. An overview of the kramer junction SEGS recent
occupied by the heat-exchange zone in a short storage
performance. In: Parabolic Trough Workshop. KJC Operating Com-
tank. Increasing the useful thermal energy Q, with the pany, Ontario, California. <http://www.nrel.gov/csp/troughnet/pdfs/
discharge power, the tank diameter and the filler size 1999_kjc.pdf>.
being fixed, needs an increase in the tank height, as can Gil, A., Medrano, M., Martorell, I., Lazaro, A., Dolado, P., Zalba, B.,
be seen from a comparison of cases 1–8 with cases 9– Cabeza, L.F., 2010. State of the art on high temperature thermal
energy storage for power generation. Part 1 – Concepts, materials and
16, respectively. The discharge efficiency is also increased
modellization. Renew Sust. Energy Rev. 14, 56–72.
due to an increase in the storage tank height (as shown in Gonzo, E.E., 2002. Estimating correlations for the effective thermal
Fig. 8). The efficiency of a design with a small tank diam- conductivity of granular materials. J. Chem. Eng. 90, 299–302.
eter can be increased by using a larger height. However, Specific Heat Capacities of Some Common Substances. The Engineering
such a design can be more expensive in terms of materials ToolBox. <http://www.EngineeringToolBox.com>.
HITEC Heat Transfer Salt. Costal Chemical Co., L.L.C., Brenntag
cost, and also offers more surface area for heat loss to
Company. <http://www.coastalchem.com>.
the environment. These practical considerations would Issa, R.I., 1986. Solution of implicitly discretized fluid flow equations by
further inform design trade-offs. operator splitting. J. Comput. Phys. 62, 40–65.
It is also noted that the filler particle size strongly affects Kearney, D., 1989. Solar electric generating stations (SEGS). IEEE Power
the efficiency. The use of small-sized filler particles Eng. Rev. 9, 4–8.
Kearney and Associates. 2001. Engineering evaluation of a molten salt
increases the efficiency greatly, as can be seen by comparing
HTF in a parabolic through solar field. NREL Contract No. NAA-1-
cases 1, 3, 5, 7 with cases 2, 4, 6, 8, respectively. The contact 30441-04. <http://www.nrel.gov/csp/troughnet/pdfs/
area between HITEC and quartzite rock is increased with ulf_herrmann_salt.pdf>.
smaller particles, which increases the heat exchange rate Kearney, D., Herrmann, U., Nava, P., 2003. Assessment of a molten salt
between HITEC and quartzite rock, leading to increased heat transfer fluid in a parabolic through solar field. J. Sol. Energy
Eng. 125, 170–176.
discharge efficiency.
Laing, D., Steinmann, W.D., Tamme, R., Richter, C., 2006. Solid media
thermal storage for parabolic trough power plants. Sol. Energy 80,
5. Conclusions 1283–1289.
Launder, B.E., Spalding, D.B., 1972. Lectures in Mathematical Models of
Turbulence. Academic Press, London, England.
A two-temperature model is developed for investigating Lovegrove, K., Luzzi, A., Soldiani, I., Kreetz, H., 2004. Developing
energy discharge from a thermocline thermal energy stor- ammonia based thermochemical energy storage for dish power plants.
age system using molten salt as the heat transfer fluid Sol. Energy 76, 331–337.
and inexpensive rock as the filler. Thermal characteristics, Luzzi, A., Lovegrove, K., Filippi, E., Fricker, H., Schmitz-Goeb, M.,
including temperature profiles and discharge efficiency of Chandapillai, I., Kaneff, S., 1999. Techno-economic analysis of a
10 MWe solar thermal power plant using ammonia-based thermo-
the storage tank, are systematically explored. chemical energy storage. Sol. Energy 66, 91–101.
During discharge, the heat-exchange zone expands with Mawire, A., McPherson, M., 2009. Experimental and simulated temper-
time and Reynolds number, and its rate of travel is con- ature distribution of an oil-pebble bed thermal energy storage system
stant and can be precisely predicted by Eq. (18). with a variable heat source. Appl. Therm. Eng. 29, 1086–1095.
Z. Yang, S.V. Garimella / Solar Energy 84 (2010) 974–985 985

Mawire, A., McPherson, M., van den Heetkamp, R.R.J., Mlatho, S.J.P., Pacheco, J.E., Showalter, S.K., Kolb, W.J., 2002. Thermocline thermal storage
2009. Simulated performance of storage materials for pebble bed system for parabolic trough plants. J. Sol. Energy Eng. 124, 153–159.
thermal energy storage (TES) systems. Appl. Energy 86, Price, H., 2002. Parabolic Trough Technology Overview. Trough Tech-
1246–1252. nology – Algeria, NREL. <http://www.ornl.gov/sci/engineering_sci-
Medrano, M., Gil, A., Martorell, I., Potau, X., Cabeza, L.F., 2010. State ence_technology/world/renewable/>.
of the art on high-temperature thermal energy storage for power Price, H., 2003. A parabolic trough solar power plant simulation model.
generation. Part 2 – case studies. Renew Sust. Energy Rev. 14, <http://www.nrel.gov/docs/fy03osti/33209.pdf>.
31–55. Sanderson, T.M., Cunningham, G.T., 1995. Performance and efficient
Michels, H., Pitz-Paal, R., 2007. Cascaded latent heat storage for design of packed bed thermal storage systems part 1. Appl. Energy 50,
parabolic trough solar power plants. Sol. Energy 81, 829–837. 119–132.
Pacheco, J.E., Gilbert, R., 1999. Overview of recent results of the solar two Singh, H., Saini, R.P., Saini, J.S., 2010. A review on packed bed solar
test and evaluations program. In: Hogan, R., Kim, Y., Kleis, S., energy storage systems. Renew. Sust. Energy Rev. 14 (3), 1059–1069.
O’Neal, D., Tanaka, T. (Eds.), Renewable and Advanced Energy Wakao, N., Kaguei, S., 1982. Heat and Mass Transfer in Packed Beds.
Systems for the 21st Century. ASME Int., New York. Gordon and Beach, New York.

You might also like