You are on page 1of 40

Generalized Charts of Energy Storage Effectiveness for

Thermocline Heat Storage Tank Design and Calibration

Peiwen Li1, Jon Van Lew1, Wafaa Karaki1

Cholik Chan1, Jake Stephens2, Qiuwang Wang3

1
Department of Aerospace and Mechanical Engineering

The University of Arizona

Tucson, AZ 85721, USA

peiwen@email.arizona.edu

2
US Solar Holdings LLC.

1000 E. Water Street, Tucson, AZ 85719, USA

3
School of Energy and Power Engineering

Xi’an Jiaotong University

Xi’an Shaanxi, 710049, China

Abstract

Solar thermal energy storage is important to the daily extended operation and cost reduction of a concentrated

solar thermal power plant. To provide industrial engineers with an effective tool for sizing a thermocline heat

storage tank, this paper used dimensionless heat transfer governing equations for fluid and solid filler material and

studied all scenarios of energy charge and discharge processes. It has been found that what can be provided through

the analysis is a series of well-configured general charts bearing curves of energy storage effectiveness against four

dimensionless parameters grouped up from the storage tank dimensions, properties of the fluid and filler material,
and operational conditions (such as, mass flow rate of fluid, and energy charge and discharge periods). As the curves

in the charts are generalized, they are applicable to general thermocline heat storage systems. Engineers can

conveniently look up the charts to design and calibrate the dimensions of thermocline solar thermal storage tanks

and operational conditions, without doing complicated modeling and computations. It is of great significance that

the generalized charts will serve as tools for thermal energy storage system design and calibration in energy

industry.

1. Introduction

Power generation using concentrated solar thermal energy is one of the most promising renewable energy

technologies [1]. It has received a great amount of research and development work in the last ten years [2]. In

particular, solar trough and solar tower concentrated thermal power generation systems are becoming more and

more reliable and matured, and their cost also has been reduced with the increase of productivity and demand [3].

It has been widely accepted that further cost reduction of concentrated solar thermal power systems may be

accomplished by adding solar thermal storage system that provides heat for prolonged operation of the power plant

and thus increases the operational capacity of the power plant and, at the same time, improves the ability of power

dispatch [4].

If an energy-carrying fluid medium in a thermal storage system can be withdrawn at the same temperature at

which it had been originally stored, the system has the highest efficiency, or has zero exergy loss from the viewpoint

of the second law of thermodynamics [5]. On the basis of this fundamental understanding, the first generation solar

thermal storage system was developed in the earlier stage, which included two heat transfer fluid storage tanks, one

for hot fluid and the other for cold fluid [6]. During energy storage process fluid from the cold fluid tank is sent to

the solar field to be heated and then stored in the hot fluid tank; while during energy discharge process, fluid from

hot fluid tank is pumped out to release heat to the power plant and afterwards flows back to the cold fluid tank.

Since the heat transfer fluid is generally expensive [7], it was thus proposed that a single tank be used for thermal

storage [8]. Such a single tank is also named as a thermocline tank which requires hot fluid being on top of cold

fluid and essentially stratified during energy charge, storage, and discharge processes. The phenomenon of

stratification of fluid by maintaining a temperature gradient is generally referred to as a thermocline. During energy

charging processes, hot fluid is charged into the tank from top and at the same time, cold fluid at the bottom of the
tank is pumped out to the solar field for absorbing heat. During energy discharging processes, hot fluid in the tank is

pumped out from the top, which then releases its heat to the power plant before returning to the tank from the

bottom.

Essentially both the two-tank and single-tank strategies of thermal energy storage use heat transfer fluid as the

heat storage medium. The method of further reducing the use of the high-cost heat transfer fluid has to rely on a

secondary energy storage medium, which must be significantly cheaper than the heat transfer fluid [9]. This

mechanism of heat storage using solid material with heat transfer fluid features the third generation of solar thermal

energy storage technology that heat transfer fluid serves mainly as the energy carrying medium; while cheaper

materials such as rocks, salts, concrete, sands, and even soil serve as energy storage media [10]. Under such a

situation, obviously, the heat storage and retrieving process involves heat transfer between the heat-carrying fluid

medium and the heat storage solid medium. Also, it is important that a single storage tank be used and thermocline

phenomenon still be maintained in this third generation of thermal energy storage technique.

Due to the existence of thermal interaction between heat transfer fluid (heat carrying medium) and heat storage

solid material, a temperature difference between the two is inevitable, and thus the heat-carrying medium can hardly

reach the same temperature as it had when it was charged into the tank. Therefore, the actual temperature history of

the heat transfer fluid flowing out from the tank during energy charge and discharge can be complicated by the

material properties of energy storage medium and energy-carrying fluid, as well as their interaction behavior

(characterized by the fluid flow and heat transfer behavior in porous media). Obviously, in order to design such a

thermal energy storage system and to size the volume of the tank, one has to analyze the heat transfer behavior of

the system.

Through the analyses in this study a series of general charts of energy storage effectiveness against a variety of

parameters will be configured, which will help industrial engineers in designing and sizing a thermal storage system.

The formatting and illustration of the general charts will accommodate full spectrum of selections of properties of

solid filler and fluid materials, dimensions of storage tanks, as well as operational conditions. The objective function

in the charts is the required quantity of energy delivery in a required time period and a mass flow rate.

Since a large number of cases must be analyzed to configure the general charts of energy delivery effectiveness

versus variety of parameters of an energy storage system, an effective computational tool must be used. A number of

analyses and solutions to the heat transfer governing equations of a working fluid flowing through a filler-packed
bed have been presented in the past [11, 12, 13, 14, 15]. As the pioneering work, Schumann [12] in 1927 presented a

set of equations governing the energy conservation of fluid flow through porous media. Schumann’s equations have

been widely adopted in the analysis of thermocline heat storage which has solid filler material inside a tank.

Schumann’s analysis and solutions were performed for the case of fixed fluid temperature at the inlet to a storage

system. Also, the initial temperature in the storage tank is assumed uniform. In most solar thermal storage

applications the inlet fluid temperature in charge and discharge processes may vary and the initial temperature can

also be strongly nonlinear. To overcome these limitations in Schumann’s analysis, Shitzer and Levy [13] employed

Duhamel’s theorem on the basis of Schumann’s solution to consider a transient inlet fluid temperature of the storage

system. However, Shitzer and Levy’s solution still assumed the initial temperature in the tank being uniform. For a

heat storage system in a solar thermal power plant, heat charge and discharge are cycled daily. The initial

temperature field of a heat charge process is dictated by the most recently completed heat discharge process, and

vise versa. Therefore, non-uniform and nonlinear temperature distribution is typical for both heat charge and

discharge processes. To accommodate the non-uniform initial temperature and time-varying inlet fluid temperature,

numerical methods to solve the Schumann equations were presented in literature by McMahan [14], Kolb [15],

Pacheco et al. [16], Van Lew et al. [17], and Karaki et al. [18]. After a rigorous evaluation and comparison, the

most effective and accurate numerical method and the efficient computational schemes given by Van Lew et al. [17]

was used in the current study.

2. Energy Storage Effectiveness in Thermocline Tanks

2.1 The scenario of energy charge and discharge for an ideal thermocline tank

If an energy-carrying fluid medium in a thermal storage system can be withdrawn at its temperature originally

being stored, the system has the highest efficiency, or has zero exergy loss from the viewpoint of the second law of

thermodynamics [5]. Such a thermal energy storage system may be idealized by using two separate storage tanks, or

by using a single storage tank with an ideal thermal insulation baffle (movable along the height of the tank) in

between the hot fluid and cold fluid, as illustrated in Fig. 1. For the single tank in Fig. 1 during energy charging

process, hot fluid flows into the tank from top and displaces the cold fluid out of tank from bottom; while during

energy discharging process, hot fluid is pumped out from top of the tank and cold fluid is charged in from bottom of
the tank. This type of energy storage and delivery using a single tank is named as the ideal thermocline storage. As

one single tank is used to achieve both the energy storage and delivery functions, it is obviously more economical

than a two-tank thermal energy storage system.

The ideal thermocline heat delivery effectiveness may be considered as 1.0 since it has no exergy loss. This is

explained in Fig. 2 by showing the fluid exit temperatures during energy charge and discharge processes. In heat

charge process, hot fluid charges in from top of the tank. The temperature of flowing-out fluid at the bottom is

shown in Fig. 2(a). When hot fluid is withdrawn from top of the tank, cold fluid enters in the tank from bottom, and

the same high-temperature fluid is discharged out as shown in Fig. 2(b).

The ideal heat rate of delivery from an ideal thermocline tank at a required mass flow rate and a period (both are

operational conditions required by the power plant) is defined as:

Q
Q T  P  m  C f (TH  TL ) (1)
 th

 is thermal energy rate, Q is electrical power, and  is thermal efficiency in thermal power plant. The
where QT P th

ideal thermocline energy delivery is calculated as:

QT  t  Q T (2)

and the volume of ideal thermocline storage tank is

Videal  t  m /  f (3)

2.2 Energy storage effectiveness in a thermocline tank having filler material

When a solid filler material is packed in a heat storage tank, as shown in Fig. 3(a), it leaves a void volume of

Vtan k , where Vtan k is the volume of the tank and  the void fraction. Under the same mass flow rate as established

by the system requirements of the power plant, the real fluid velocity in the charge/discharge processes of the tank

(assuming to have the same volume of Videal ) with filler material will be higher than that in an ideal thermocline

tank. Therefore, during a discharge process, the temperature of the fluid flowing out from top will decrease after a

time when the pre-existing hot fluid is completely discharged; from then on the hot fluid discharged out from top

originates from cold fluid coming into the tank that has been heated by the filler material. The more the discharge

process progresses, the more the temperature of the discharged fluid will decrease, as illustrated in Fig. 3(b).
When hot fluid is charged into the tank, the pre-existing cold fluid in the tank is displaced out from the bottom of

the tank in the meantime. After the pre-existing cold fluid is discharged, any further discharged fluid will be the

fluid which enters the tank at high temperature and gives its energy to the cold filler material. Since the hot fluid

cannot give the entirety of its thermal energy to the cold filler material, the discharged fluid temperature at the tank

bottom will gradually increase. The more the charge process progresses, the more the temperature of the discharged

fluid from the bottom of the tank will increase. Figure 3 (c) schematically illustrates this temperature variation of the

discharged fluid from the bottom of a tank.

Although the discharged fluid has temperature degradation/drop, as shown in Fig. 3 (b), a heat storage tank

having filler material is still being considered as a cost-effective and economical technology, since the heat transfer

fluid is expensive and shall not be used as the major heat storage medium. Therefore, the significance of using filler

material is to displace and minimize the use of heat transfer fluid in a heat storage tank.

No matter whether a filler material is being used or not in a storage tank, a power plant operation requires a

specified period of heat transfer fluid discharge time under a specified mass flow rate. Therefore, minimizing the

temperature degradation by proper design and operation of a thermal storage tank becomes the key issue of

thermocline energy storage, which is the focus of this study.

Before going into mathematical analysis about the heat transfer and energy conservation of fluid and filler

material in a thermocline tank, it is worthwhile to observe the following three scenarios.

(1) The energy charge period is the same as that of required discharge period. In this case, using the heat storage

tank having filler material one can never achieve the ideal quantity of heat discharge. This is because the energy

charged into the tank can not be larger than the ideal amount of energy, as seen from illustration in Fig. 3 (c), and

therefore, the energy discharge process cannot achieve the ideal quantity of energy, as seen from illustration in Fig.

3 (b).

(2) There is a longer period of energy charge, t ch arg e , than the period of required energy discharge, t disch arg e .

From the above analysis and its conclusion, it is obvious that in order to get a required amount of energy discharge,

using an increased charge time is inevitable. However, even if a longer heat charging time is applied, there is still the

possibility that an ideal quantity of heat discharge cannot be achieved under the particular conditions that the filler

material has ( C) filler  ( C) fluid . This is because that in such a situation, there is still no extra amount of energy,

compared to that of the ideal energy, being stored in the tank, and the best case scenario is that a tank is fully
charged in a very long time. On the other hand, the energy discharge process is not comparable to that of an ideal

thermocline tank, and thus the energy discharged cannot approach the ideal quantity, as seen from the illustration in

Fig. 3 (b).

(3) Having a filler material with ( C) filler  ( C) fluid together with a heat charging period, t ch arg e , longer than the

required heat discharging period, t disch arg e , is the prerequisite for a heat storage tank with filler material to deliver

heat that approaches the ideal heat delivery, seen in an ideal thermocline tank. A larger (C) filler compared to ( C) fluid

allows more energy than the required energy discharge being stored, if a longer period of charge than that of

discharge is applied. In conclusion, only when the two conditions— ( C) filler  ( C) fluid and t ch arg e  t disch arg e are

both satisfied is it possible that the non-ideal storage tank can contain more energy than the required amount for

discharging, and thus it is possible to approach an ideal amount of energy delivery.

Finally, it is also very interesting to note that under the situations discussed in the above paragraphs the volume

of the storage tank may be larger than that of an ideal thermocline tank, but it may also be smaller than that of an

ideal thermocline tank if ( C) filler is much larger than ( C) fluid .

The mathematical analysis hereafter will investigate and quantitatively show the above discussed phenomena for

a storage tank with filler material.

3. Heat Transfer and Energy Storage Modeling and Computation

3.1 Energy conservation in fluid and rocks

Through energy balance analysis for fluid and rocks, governing equations for the temperatures of fluid and rocks

will be constructed. Shown in Fig. 4 is a one-dimensional control volume of size dz in a rock-packed flow bed. Both

the energy conservation in the fluid and in the rocks in the control volume will be examined. For convenience of

analysis, the positive direction of coordinate z is set always identical to the fluid flow direction. In an energy charge

process hot fluid flows into the tank from the top, and thus z = 0 is at the top of the tank. In a heat discharge

process, cold fluid flows into the tank from bottom and this makes z = 0 for the bottom of the tank.

Several assumptions are typically made to simplify the analysis of a thermocline (with filler material) heat

storage process: (1) A uniform radial distribution of the fluid flow and rocks through the storage tank is assumed to
reduce the analysis to a one-dimensional problem along the axis, z, of the storage tank; (2) The contact between

rocks is point contact and therefore heat conduction between rocks are negligible; (3) It has been confirmed that the

Peclet number of fluid flow in a thermocline tank is large (Pe >> 100), and therefore the heat conduction in the

axial direction in the fluid is negligible [19]; (4) It is assumed that lumped heat capacitance method is applicable to

the transient heat conduction in a single rock; however, if the Biot number [40] of the rock, due to its property and

heat transfer with fluid, is large, a correction of large Biot number effects is considered later; (5) There is no heat

loss from the storage tank to the surroundings; this applies to both the energy charge and discharge process as well

as the resting time in between a charge and a discharge.

Heat loss from a thermal storage tank is inevitable; however, one needs to decide the dimensions of a storage

tank firstly in order to find the heat loss. Later on, to compensate for the heat loss will result in a larger volume of

the heat storage tank and a longer heat charge period. A simple way of refining the design is to increase both the

heat charge time and tank size with a percentage equal to the ratio of heat loss versus the projected heat delivery.

Nevertheless, to make this study focused, the current work concentrates on the determination of the dimensions of a

storage tank without considering heat loss. This assumption also assists the connection of the results from a heat

charging process to be the initial condition of the following on discharging process, and visa versa. By doing so,

multiple cyclic energy charges and discharges in the actual operation can be simulated from the current modeling

analysis.

Based upon the above modeling assumption (1), the cross-sectional area of the tank seen by the fluid flow is

assumed constant at all points along the axis of the tank, and is:

a f  R 2 (4)

The thermal energy balance of the fluid in the control volume dz is:

Tf
 f R 2 U( z   z  dz )  hS r (Tr  Tf )dz   f C f R 2 dz (5)
t

where the average fluid velocity in the packed bed is:

m
U (6)
f a f

With substitution for the definition of enthalpy change,  z  dz   z  C f T f , and rearrangement of Eq. (5), the

energy balance equation for fluid becomes:


hS r Tf T
(Tr  Tf )  U f (7)
 f C f R 2 t z

Introducing the following dimensionless variables,

 f  (Tf  TL ) (TH  TL ) (8.a)

 r  (Tr  TL ) (TH  TL ) (8.b)

z*  z / H (8.c)

t *  t /( H / U) (8.d)

The dimensionless governing equation for heat transfer fluid is finally shown to be

 f  f 1
  ( r   f ) (9)
t *
z * r

where

U  f C f R 2
r  (10)
H hS r

In the equation, S r is the heat transfer surface area of rocks per unit length of the tank, which is a function of the

radius of rocks as well as the schemes that rocks are packed. Based upon the modeling assumption (2), S r is

calculated as:

f s R 2 (1  )
Sr  (11)
r

where f s is the factor of surface shape and may vary between 2 and 3 depending upon the rock’s packing scheme.

In this study, it is chosen to be 3.0.

The heat transfer coefficient h (W/m2 oC) between solid and fluid in porous media in the above equations was

based upon the analysis provided by reference [21].

 Cf
m
h  0.191 Re  0.278 Pr  2 / 3 (12)
R 2

where the Reynolds number is the modified Reynolds number for porous media, defined as [21]:

4G rchar
Re  (13)
f

where G is the mass flux of fluid through the porous bed,



m
G (14)
R 2

and rchar is defined as the characteristic radius by Nellis and Klein [21] (sometimes defined as the hydraulic radius):

d r
rchar  (15)
4(1   )

For the energy balance of the filler material (rocks) in a control volume dz as shown in Fig. 4, it is understood

that the filler delivers or takes heat to or from the passing fluid at the cost of a change in the internal energy of the

filler. The equation is:

Tr
hS r (Tr  Tf )dz   r C r (1  )R 2 dz (16)
t

With substitution of dimensionless variables given in Eq. (8), the above governing equation for filler turns to be

 r H CR
 ( r   f ) (17)
t * r

f Cf 
where H CR  (18)
 r C r (1  )

In the charge and discharge processes, the rocks and heat transfer fluid will have a temperature difference at any

local location. Once the fluid comes to rest upon the completion of a charge or discharge process, the fluid will

equilibrate with the local rocks to reach the same temperature. The energy balance of this situation at a local location

is:

 f C f Tf initial  (1  ) r C r Tr initial   f C f Tfinal  (1  ) r C r Tfinal (19)

Here, the initial temperatures of rocks and fluid are from the results of their respective charge or discharge process.

The final temperatures of rock and fluid are the same after equilibrium is reached.

According to the assumption of no heat loss from the storage tank, it can be seen that the equilibrium temperature

at the end of one process (charge or discharge) will necessarily be the initial condition of the next process in the

cycle. This connects the discharge and charge processes so that an overall periodic result can be obtained.

The initial temperatures of rock and fluid in the storage tank should be known. Also, the inlet fluid temperature is

known as a basic boundary condition, with which the rock temperature at inlet location z=0 can be easily obtained

from Eq. (17). Therefore, as the boundary conditions, both temperatures of fluid and rock at the inlet z=0 are known.
3.2 Correction to lumped capacitance approximation for heat transfer in rocks

The modeling presented in section 3.1 used lumped capacitance method to treat the heat transfer inside rocks.

This method actually ignored the resistance of heat conduction inside a rock. This will result in discrepancy which

makes the calculated energy going into or coming out from a rock higher than that in the actual physical process.

From literature [20] it is known that when the Biot number of the heat transfer of a solid object is larger than 0.1, the

lumped capacitance assumption will result in increased inaccuracy for the heat transfer and energy conservation

analysis of a rock in a fluid. In order to correct the lumped capacitance approximation, Jeffreson [22] proposed to

give a correction to the convective heat transfer coefficient between rocks and fluid in the form of:

1 1  Bi / 5
 (20)
hp h

where h p is the new heat transfer coefficient to replace the h obtained in Eq.(12). The corrected heat transfer

coefficient h p will be used in equations wherever h is needed. Any other terms and properties in all governing

equations remain unchanged.

The Jeffreson correction allows for the thermocline model to remain in a one-dimensional system yet increases

the accuracy of the results by accounting for the internal thermal gradient in the packed bed filler material. A

comparison of the results of dimensionless energy going in or out from a single rock of the filler using exact

transient heat conduction solution from literature [20], using the lumped capacitance method [20], and using the

corrected lumped capacitance method by introducing a corrected heat transfer coefficient by Eq. (20) were given in

Fig. 5. The definition of the dimensionless energy, Q * , in Fig. 5 is the ratio of energy going in or out from the rock

compared to the ideal energy going in or out from a rock. The energy going in or out from a rock is counted based

on the change of internal energy of the rock. The ideal energy going in or out from a rock is the internal energy

change of a rock assuming its temperature completely changed from initial temperature to the fluid temperature

around it.

Parameters of rocks and heat transfer coefficient used for the comparison of the three methods in Fig. 5 were

listed in Table 1. The Bi number of the case is 2.54, which is close to the situation of the rocks in typical

thermocline systems. As shown in Fig. 5, the results from Jeffreson correction model agreed with the exact transient

heat conduction solution very well. Therefore the Jeffreson correction, Eq. (20), of the heat transfer coefficient

between rocks and fluid in the thermocline system is recommended.


3.3 Numerical method of solution to the governing equations

Using the approach of method of characteristic Eqs. (9) and (17) can be solved numerically with minimum

computation time and high accuracy. Details of the method of characteristics and numerical computation are given

in literature [17, 23]. However the key steps of solution to the governing equations are introduced in the following

section.

As shown in Fig. 6, the length of the storage tank is discretized (represented by node number i ) with the

maximum grid number of M, and time length is discretized (represented by node number j ) with the maximum

number of N. The fluid temperature at inlet is known as a function of time. The rock temperature at the inlet is

calculated based on Eq. (17), for which an explicit finite difference equation is obtained

 r 1, j 1   r 1, j H CR
 ( r 1, j 1   f 1, j 1 ) (21)
t *
r

where the fluid inlet condition  f 1,1 to  f 1, N and initial condition of rocks  r 1, 1 are both known and therefore,

 r 1, 2 up to  r 1, N can be solved explicitly. Therefore, in the matrix shown in Fig. 6 temperatures of fluid and rocks

at nodes 1,1 to 1, N and 1,1 to  M ,1 are known.

Using the method of characteristics, literature [17] obtained the following algebraic equation matrix to solve the

temperatures of rocks and fluid at the node  2, 2 .

 t * t *   f 2 , 2    t *  t * 
 1   
  f1,1  1  
   
 2 r  2 r
r1,1
 2 r 2 r     (22)
 H CR t
*
H CR t *      H CR t *   H CR t * 
  1        1  
 2 r 2 r   r2 , 2   f 2 ,1  2 r  r2 ,1  2 r 

This matrix in Eq. (22) includes an implicit condition of discretization of the time and length so that z *  t * .

Since the matrix only has two equations, Cramer’s rule can be applied to obtain the solution of  f 2, 2 and  r 2, 2

efficiently. Further, the matrix can be easily applied to other nodes by marching the time and space steps as given in

Fig. 6. While the marching of z * steps is limited to z *  1 , the marching of time t * has no limitation. The error
of such an implementation is not straightforwardly analyzed here in this study, but the formal accuracy of method of

characteristics is on the order of O( t *2 ) as described in literature [23].

It is important to note that all coefficient terms in Eq. (22) are one-time determined from z , t * ,
*
 r , and

H CR . Therefore the numerical computation takes minimum computing time, which is much more efficient than all

other methods seen in references [14, 15, 24].

4. General Charts of Energy Storage Effectiveness in Thermocline Tank Having Filler Material

4.1 Energy storage effectiveness

As was discussed in section 2.2 for a thermocline storage tank with filler material, the energy delivered in the

required time period at a required mass flow rate is always less than that of the ideal energy delivery in an ideal

thermocline tank. With the solution of the governing equations, the discharged fluid temperature from a storage tank

can be obtained. If the required heat discharging period is t ref , disch arg e , an energy storage effectiveness can be

defined as:

t ref , disch arg e

 [T f ( z  H , t )  TC ]dt
 0
(23)
(T H  TC )  t ref , disch arg e

where the numerator represents the energy discharged from the actual tank, and the denominator represents the

energy discharge from an ideal thermocline tank. The dimensionless form of the required time period of energy

discharge is defined as:

t ref,discharge
d  (24)
H U

Similarly a dimensionless form of the time period of energy charge is defined as:

t charge
c  (25)
H U

Substitute the dimensionless energy discharge period  d into Eq. (23), there is
d
1

d  [1  
0
f ( z *  1, t * )]dt * (26)

It is now known that  is essentially the function of the following three parameters—  c /  d ,  r , and H CR .

While the dimensionless time period of discharge process and the mass flow rate must be prescribed, the

dimensionless time period of energy charge period is the variable that will be searched to determine the targeted

value of  , which should approximately equal to 1.0. It has been concluded in section 2.2 that a longer energy

charge time than that of energy discharge time is always needed in order to achieve the energy delivery effectiveness

of 1.0.

As the results of an example from the current computation, Fig. 7 shows the variation of temperatures of fluid

flowing out from the tank in an energy charge and following-up discharge process, respectively. The energy charge

process needs a longer time than that of the energy discharge process. Since more energy is charged into the tank

than needed for discharge, the temperature degradation of the discharged fluid is not significant, which yield an

energy discharge efficiency close to 1.0.

4.2 Energy storage effectiveness chart for tank size design

It is convenient that a general series of charts of energy delivery effectiveness  versus  c /  d ,  r , and

H CR be provided for the design of a storage tank. Since  is a function of three variables, the general charts can be

configured in different ways. Illustrated in Fig. 8 is a configuration of the charts for a given  d , in which multiple

graphs, each has a specific  r , are provided. In each graph, multiple curves, each has a given HCR, for the energy

storage effectiveness  versus  c /  d is provided. To have a full spectrum of the charts, more graphs in the same

configuration as shown in Fig. 8 must be provided at variety of different values of  d .

Figure 9 shows some results from the computation in this study, the graphs can form a general chart for a

constant  d of 4.0. In these graphs every data point in the curves were based on the final heat discharge process,

which occurs at the end of several cycles of charge and discharge, when the results of cyclic charge and discharge

are independent of the most initial condition in the tank. A most initial condition is assumed where the temperatures

of rocks and fluid are both  f   r  1 and the program then calculates the temperature distribution during a
discharge. The temperature results after a discharge are loaded as the initial condition of the next charge process

and the computation is repeated; similarly the results after a charge are loaded into the next discharge process. In

the case where the fluid energy density ( fCf ) is much lower than the rock energy density (  r C r ), the
discharge-charge cycle must be repeated many times to reach the end of the computation.

The graphs in Fig. 9 clearly agree with the following conclusions that have already been drawn from the

qualitative analysis and discussion in section 2.2:

(1) The energy delivery effectiveness does not reach 1.0 if  c /  d  1.0 .

(2) A decrease of  r corresponds to an increase of the volume of storage tank. Therefore with a decrease of

 r the effectiveness  can approach 1.0 more easily under the same values of H CR and  c /  d . For example, at a

ratio of  c /  d  1.5 , the energy delivery effectiveness easily approximate to 1.0 when  r changed from 0.2 to

0.01.

(3) A decrease of HCR corresponds to the case that (  r C r ) is increasing relatively to (  f C f ) , and therefore the

energy storage capability is improved so that   1.0 is also easier to achieve. This is seen in every graph in Fig. 9.

(4) There are cases that no matter what ratios of  c /  d are used, the energy storage effectiveness can hardly

approach 1.0. This occurs when either  r or H CR are too large, which physically corresponds to small tank size

and  r C r   f C f , respectively.

4.2 Chart for calibration analysis of a given storage tank

Designers for a thermal storage system often need to calibrate or confirm that a given storage tank can satisfy an

energy delivery requirement. As the dimensions of the storage tank and the power plant operational conditions are

known, the value of  r is essentially given. This requires that the general charts be configured in the form as shown

in Fig. 10. Engineers can look up different energy discharge time from multiple graphs and essentially look up the

energy storage capability of a given tank.

Based on computation results, Fig. 11 gives four representative graphs, each having multiple curves for the

energy delivery effectiveness versus the charge and discharge time ratio. The series of graphs are for a fixed value of
 r  0.03 , and each graph has a fixed  d . The results are calculated in the same manner as discussed for Fig. 9

for cyclic charges and discharges.

The size parameters of a storage tank are included in the parameter,  r . Under the same value of  r , a smaller

 d means smaller required energy storage, and therefore, the energy delivery effectiveness can reach to 1.0 easily

for most cases of H CR . At larger values of  d , a larger ratio of  c /  d is needed in order to achieve the energy

delivery effectiveness of 1.0.

5. Applications of the charts for design of thermocline storage tanks

5.1 Sizing a thermocline storage tank

To decide the dimensions of a thermocline storage tank, the required operational conditions from the power plant

include: the electrical power, the thermal efficiency of the power plant, the extended period of operation based on

stored thermal energy, the required high temperature of heat transfer fluid from the storage tank, the low temperature

of fluid returned from the power plant, the properties of the heat transfer fluid and the solid filler material including

the nominal radius of fillers, as well as the packing porosity in a thermocline tank. The design analysis using the

general charts provided in the present study will include the following steps:

(1) Decide an ideal volume for an ideal thermocline tank.

The total thermal energy is related to the required volume of an ideal thermocline tank in the form of:
QT

Cf

TH  TL   m fluid C fluid TH  TL t disch arg e


l
u
i
d

 Videal   
f
l
u
i
d

(27)

Once the volume of the ideal thermocline tank is determined, a chosen diameter, R, and corresponding height, H, of

the tank can be decided, which will be used in the first trial for energy storage effectiveness analysis. Following

these dimensions, the parameters—  d ,  r , H CR for a thermocline tank with filler material, can be decided, where

 d is determined by setting t ref ,disch arg e equal to t disch arg e .

(2) Look up the design charts (Fig. 9) and see if an energy storage effectiveness of 1.0 can be achieved with the

parameters decided in step (1). It might be common that the energy effectiveness can not be close to 1.0 during the
first trial. This is because the first trial uses dimensions from an ideal thermocline storage tank. However, with the

results from the first trial one can predict and guess a new  r and a new  d for the next trial. In fact, the increase of

the height of the storage tank can result in both decrease of  r and  d in the same proportion.

(3) If the effectiveness from step (1) cannot approach 1.0 even if a large  c /  d is chosen, one has two ways to

improve the effectiveness during the second trial, which are to decrease H CR or decrease both  r and  d in the

same proportion. However, H CR is determined by properties of fluid and filler material, which has very limited

options, and therefore decrease of  r and  d is inevitable. The decrease of  r is actually due to the increase of the

height of a storage tank. This means that to get an effectiveness of 1.0 one has to increase the height of the storage

tank. Because the height of tank is included in both  r and  d , the decrease of  r must be accompanied by the

same proportional decrease of  d .

Although the situation of selecting a smaller  r in step (3) is common, the opposite scenario of selecting a larger

 r may also occur. This is corresponding to the case that the energy storage effectiveness  can approach 1.0 at

 c /  d slightly larger than 1.0. This also means that the storage tank may be oversized. One can increase  r and

 d in the same proportion to achieve an effectiveness of 1.0, but at the expense of a larger  c /  d than before.

Physically, this scenario corresponds to the situation that one can use a smaller tank but at a longer charge time to

obtain the same amount of required energy delivery. This may happen if the H CR is very small, which is due to a

much larger value of (  r C r ) in relation to (  f C f ).

5.2 Find the energy storage capacity and the period of operation for a given tank

If the dimensions of a storage tank and the operational conditions of power plant, including the mass flow rate

and the high and low fluid temperatures, are given, it is the task of designers to find a proper time period of energy

charge that can satisfy the needed operation time of the power plant. The known parameters will be  r and H CR at

a required operation period of  d . A series of graphs (as given by Fig. 11) for the given  r must be looked up.

From the set of graphs one needs to identify a particular graph for the given  d . In such a particular graph, the
energy storage effectiveness versus  c /  d at various values of H CR can be easily found. If at the required

 d the energy storage effectiveness cannot approach 1.0 at any value of  c /  d , a new  d must be selected. In

this way the storage tank is calibrated for whether or not it is suitable for the required  d .

5.3 An example of tank size and operation design

The following example shows the sizing procedure of a storage tank. A solar thermal power plant has 1.0 MW

electrical power output at a thermal efficiency of 20%. The heat transfer fluid used in the solar field is Therminol®

VP-1. The power plant requires the high and low fluid temperatures of 390oC and 310oC, respectively. River rocks

are used as the filler material and the void fraction of packed rocks in the tank is 0.33. Required time period of

energy discharge is 1 hour, the storage tank diameter is 6 m. The rock diameter is 4 cm.

Making use of Eq. (27), the above given details on the power plant, as well as the properties of Therminol® VP-1,

we can find a necessary mass flow rate of 23.76 kg/m3, and an ideal tank height of 4 m to deliver the required

power. Using Eqs. (12) and (20) we find a heat transfer coefficient to be 32.05 W/m2K. With this information, the

values of H CR ,  d , and  r are 0.451, 3.03, and 0.06, respectively. Given in Fig. 12 is a family of charts for

 d =3 and  r =0.06 at various values of H CR and  c /  d . Figure 12 shows that at given value of H CR there is

no time ratio  c /  d that will deliver an effectiveness value of 1.0 from this tank of an ideal volume.

One option to deliver more power and approach an effectiveness of 1.0, is to increase the height of the storage

tank. When the height is increased to 6 m, the values of  d and  r change to 2.03 and 0.0404, respectively. Figure

13 gives the family of curves for  d =2 and  r =0.04. In this figure, it is seen that at H CR of 0.45, when the time

period ratio,  c /  d , is approximately 1.5 an effectiveness of 0.99 is possible. This should be acceptable.

A design engineer may feel that such a large period of charge is not acceptable. In that case, the designer can

increase the height once again. When the height of the storage tank is increased to 8 m, the values of  d and  r

become 1.52 and 0.0303, respectively. Figure 14 shows the family of curves for  d =1.5 and  r =0.03. In Fig. 14

we clearly see that when  c /  d > 1.25, an effectiveness of almost 1.0 is possible for all values of H CR .
This above example demonstrates the power of having a data library containing the most common combinations

of H CR ,  d , and  r . The design engineer of this 1.0 MWe power plant thermocline solar thermal storage has

been given a clear idea of the impact of design choices and feasibility of the overall performance of the storage

system without the need for complicated and lengthy calculations in a simulation. Again, it is the expectation of the

authors that the currently proposed general charts be made available as a design tool in concentrated solar power

industry.

As to the calibration application, using generalized charts to find the energy storage capacity and the period of

operation for a given tank, the procedures described in section 5.2 is straightforward and easy to follow. It may not

be necessary to show an example here.

Finally, it is to be noted that heat loss from a thermal storage tank is inevitable. After the tanks size being

decided, it may be necessary to estimate a heat loss. To compensate the heat loss, a larger volume of the heat storage

tank and a longer heat charge period may be needed. The current authors propose that a simple way of refining the

design is to increase both the heat charge time and tank size with a percentage equal to the ratio of heat loss versus

the projected heat delivery.

6. Concluding Remarks

Thermocline solar thermal energy storage technology is receiving increased attention in the solar energy industry.

The energy storage capacity and energy delivery effectiveness in a thermocline tank is governed by the energy

conservation equations of both solid filler material and heat transfer fluid. As has been discussed in this paper that

the two governing equations are partial differential equations and solving them will need numerical computations,

which may not be convenient to engineers to repeatedly solve when they design or calibrate thermocline storage

tanks.

To provide a generalized tool for reference in designing a thermal storage tank, this study gave a comprehensive

analysis to the thermocline energy storage and delivery performance. The study used dimensionless heat transfer

governing equations for fluid and solid filler material and could analyze the energy charge and discharge

effectiveness in general. Consequently this paper proposed a series of general charts illustrating the energy storage

effectiveness as a function of four dimensionless parameters that are grouped and comprised of all the given

parameters of a thermocline tank including: properties of solid filler material and heat transfer fluid, dimensions of
the storage tank, mass flow rate of heat transfer fluid, filler material size and the void fraction of packing in the tank,

and energy charge and discharge time period. The charts are generalized and are applicable to any cases of

thermocline thermal storage.

Engineers can conveniently look up the general charts and decide the dimensions of thermocline solar thermal

storage tanks as well as operational conditions without doing complicated modeling computations. It is of great

significance to reducing the work load when they design and calibrate thermal energy storage systems. The authors

expect the currently proposed general charts be widely adopted as a benchmark design tool for application in

concentrated solar power industry.

Finally due to the limited space of this paper, it does not provide a large number of the graphs for the general

charts proposed in the work. However upon request, the current authors will provide a larger number of graphs and

charts covering a wide range of parameters for industrial application.

Acknowledgement

The authors are grateful to the supported by the US Department of Energy, National Renewable Energy

Laboratory, under DOE Award Number DE-FC36-08GO18155, and US Solar Thermal Storage LLC.

Nomenclature

Bi Biot number, specially defined as h  r / k r

Heat capacity [ J /(kg 


o
C C ) ].

dr Average diameter of rocks [ m ].

G Mass flux [ kg/(m2 s) ].

Enthalpy of fluid at a location along the main flow direction [ J /( kg 


o
 C) ]

h Heat transfer coefficient [ W /(m 2 oC ) ]

hp Corrected heat transfer coefficient [ W /(m 2 oC ) ]

H Overall height of the storage tank [ m ].

HCR Dimensionless parameter in heat transfer equations.

kr Thermal conductivity of rocks [W/(mK)]


m Mass flow rate [ kg/s ]

Q Energy [ J ].

Q Energy rate or power [ W ].

Q* Dimensionless energy or energy ratio.

r Average radius of the filler material [ m ]

R Radius of the storage tank [ m ].

Re Modified Reynolds number for flow in porous media

Sr Surface area of solid filler material per unit length of tank [ m ]

t Time [sec ]

o
T Temperature [ C ]

U Velocity of heat transfer fluid in the axial direction of a storage tank [ m/s ]

Volume of tank [ m ]
3
V

z Location of a fluid element along the axis of the tank (m)

Greek Symbols

t Time period of energy charging or discharging [ sec ].

z Length of a section of whole of a storage tank [ m ].

 Porosity of the storage tank.

 Energy delivery effectiveness defined by Eqs. (23) or (26).

 th Thermal efficiency of a solar power plant.

μ Viscosity [ Pa  s ].

 Dimensionless time period of energy charge or discharge.

Density [ kg / m ].
3
ρ

r Dimensionless parameter.

θ Dimensionless temperature.
Superscript

* Dimensionless variables

Subscript

c Energy charge charge process

d Energy discharge process

exit Exit of heat transfer fluid.

f Thermal fluid.

ideal For ideal thermocline (no solid filler material in tank)

r Filler material (rocks).

ref Reference time period—a requirement condition of power plant.

z Location along the axis of the tank

References

[1] Herrmann, U., and Kearney, D. W., 2002. “Survey of thermal energy storage for parabolic trough power plants,”

Journal of Solar Energy Engineering, 124(2), pp. 145–152.

[2] Antoni Gil, Marc Medrano, Ingrid Martorell, Ana Lazaro, Pablo Dolado, Belen Zalba, and Luisa F. Cabeza.

2010, “State of the art on high temperature thermal energy storage for power generation. part 1—concepts,

materials and modellization,” Renewable and Sustainable Energy Reviews, 14(1): 31-55, 2010.

[3] Hank Price, Eckhard Lupfert, David Kearney, Eduardo Zarza, Gilbert Cohen, Randy Gee, Rod Mahoney, 2002,

“Advances in Parabolic Trough Solar Power Technology, Journal of Solar Energy Engineering, 124, pp. 109–125.

[4] Robert Pitz-Paal, Jurgen Dersch, Barbara Milow, Felix Tellez, Alain Ferriere, Ulrich Langnickel, Aldo Steinfeld,

Jacob Karni, Eduardo Zarza, Oleg Popel, 2007, “Development Steps for Parabolic Trough Solar Power

Technologies With Maximum Impact on Cost Reduction, Journal of Solar Energy Engineering, 131, pp. 371-377.

[5] Adrian Bejan, Advanced Engineering Thermodynamics, 2006, John Wiley & Sons, Inc., New Jersey.

[6] Moens L, Blake D M, Rudnicki D L, Hale M J, 2003, ”Advanced thermal storage fluids for solar parabolic

trough systems,” Journal of Solar Energy Engineering, 125, 112–6.


[7] Kearney D, Herrmann U, Nava P, B. Kelly, R. Mahoney, J. Pacheco, R. Cable, N. Potrovitza, D. Blake, H. Price,

2003, “Assessment of a molten salt heat transfer fluid in a parabolic trough solar field, “ J Solar Energy Eng-Trans

ASME 2003;125:170–6.

[8] R. Gabbrielli, C. Zamparelli, 2009, “Optimal Design of a Molten Salt Thermal Storage Tank for Parabolic

Trough Solar Power Plants,” Journal of Solar Energy Engineering, 131, 041001.

[9] Hasnain SM.Reviewonsustainable thermal energy storage technologies, part I: heat storage materials and

techniques. Energy Convers Manage 1998;39: 1127–38.

[10] Brosseau D, Kelton JW, Ray D, Edgar M, Chisman K, Emms B. Testing of thermocline filler materials and

molten-salt heat transfer fluids for thermal energy storage systems in parabolic trough power plants. J Solar Energy

Eng Trans ASME 2005;127:109–16.

[11] Beasley, D. E., and Clark, J. A., 1984. “Transient response of a packed bed for thermal energy storage”.

International Journal of Heat and Mass Transfer, 27(9), pp. 1659 – 1669.

[12] Schumann, T. E.W., 1929. “Heat transfer: A liquid flowing through a porous prism”. Journal of the Franklin

Institute, 208(3), pp. 405 – 416.

[13] Shitzer, A., and Levy, M., 1983. “Transient behavior of a rock-bed thermal storage system subjected to variable

inlet air temperatures: Analysis and experimentation”. Journal of Solar Energy Engineering, 105(2), May, pp. 200–

206.

[14] McMahan, A. C., 2006. “Design and optimization of organic rankine cycle solar-thermal powerplants”.

Master’s thesis, University of Wisconsin-Madison.

[15] Kolb, G. J., and Hassani, V., 2006. “Performance analysis of thermocline energy storage proposed for the 1 mw

saguaro solar trough plant”. ASME Conference Proceedings, 2006(47454), pp. 1–5.

[16] Pacheco, J. E., Showalter, S. K., and Kolb, W. J., 2002. “Development of a molten-salt thermocline thermal

storage system for parabolic trough plants”. Journal of Solar Energy Engineering, 124(2), pp. 153–159.

[17] Jon T. Van Lew, Peiwen Li, Cho Lik Chan, Wafaa Karaki, Jake Stephens, “Transient Heat Delivery and

Storage Process in a Thermocline Heat Storage System,” IMECE2009-11701 Proceedings of the ASME 2009

International Mechanical Congress and Exposition, IMECE 2009, November 13-19, 2009, Lake Buena Vista,

Florida, USA.
[18] Wafaa Karaki, Jon T. Van Lew, Peiwen Li, Cho Lik Chan, Jake Stephens, “Heat Transfer in Thermocline

Storage System with Filler Materials: Analytical Model,” ES2010-90209, Proceedings of the ASME 2010 4th

International Conference on Energy Sustainability ES2010, May 17-22, 2010, Phoenix, Arizona, USA.

[19] Kays, W. M., Crawford, M. E., and Weigand, B., 2005. Convective Heat and Mass Transfer, fourth ed.

McGraw Hill.

[20] Incropera, F. P., and DeWitt, D. P., 2002. Introduction to Heat Transfer, fourth ed. John Wiley and Sons, Inc.

[21] Nellis, G., and Klein, S., 2009. Heat Transfer. Cambridge University Press.

[22] C.P. Jeffreson. Prediction of breakthrough curves in packed beds: 1. applicability of single parameter models.

American Institute of Chemical Engineers, 18(2):409-416, 1972.

[23] Polyanin, A. D., 2002, Handbook of Linear Partial Differential Equations for Engineers and Scientists, Boca

Raton: Chapman & Hall/CRC Press, ISBN 1-58488-299-9.

[24] McMahan, A., Klein, S. A., and Reindl, D. T., 2007. “A finite-time thermodynamic framework for optimizing

solar-thermal power plants”. Journal of Solar Energy Engineering, 129(4), pp. 355–362.
Table 1 Properties of the rock and heat transfer coefficient used in section 3.2
Figure Captions

Figure 1 Illustration of an ideal thermocline heat storage tank

Figure 2 Illustration of temperature variation of the fluid flowing out of an ideal thermocline storage tank ( (a)
Heat charge process; (b) Heat discharge process. )

Figure 3 Illustration of temperatures of heat transfer fluid during energy charge and discharge processes from a
thermocline tank with filler material. ( (a) Thermocline tank with filler material; (b) Heat discharge from top; (c)
Heat charge from top. )

Figure 4 One control volume/element for energy balance analysis

Fig. 5 Comparison of Jeffreson correction model, exact transient heat conduction solution, and lumped capacitance
solution for a solid object heat transfer in fluid

Figure 6 The matrix of nodes assignment for method of characteristic solution for Eqs. (9) and (17)

Figure 7 Dimensionless temperatures of the fluid flowing out from a tank in a charge and a following-on discharge
( the case has H CR  0.1 ,  r  0.04 ,  d  4.0 , and  c /  d  1.54 .)

Figure 8 Configuration of the charts including multiple graphs of  versus  c / d at a fixed  d

Figure 9 Computed results of multiple graphs of energy storage effectiveness versus  c / d at  d =4.0

Figure 10 Illustration of a configuration of the charts for energy delivery effectiveness using multiple graphs at
different  d but at a constant  r.

Figure 11 A series charts of energy effectiveness versus  c /  d at different  d and a constant value  r =0.03

Figure 12 Energy effectiveness versus  c /  d at  d =3 and  r =0.06

Figure 13 Energy effectiveness versus  c /  d at  d =2 and  r =0.04

Figure 14 Energy effectiveness versus  c /  d at  d =1.5 and  r =0.03


Hot fluid

Movable ideal
thermal
insulation baffle

Cold fluid

Figure 1 Illustration of an ideal thermocline heat storage tank


T exit (bottom) T exit (top)
TH TH

TL
TL

t t (hour)
t t
(a) (b)

Figure 2 Illustration of temperature variation of the fluid flowing out of an ideal thermocline storage tank

( (a) Heat charge process; (b) Heat discharge process )


(a)

T exit (top) T exit (bottom)


TH TH
T cut-off T cut-off

TL TL

t t
(  t) t (  t) t

(b) (c)

Figure 3 Illustration of temperatures of heat transfer fluid during energy charge and discharge processes from a
thermocline tank with filler material. ( (a) Thermocline tank with filler material; (b) Heat discharge from top; (c)
Heat charge from top. )
z

H dz

2R

Figure 4 One control volume/element for energy balance analysis


Fig. 5 Comparison of Jeffreson correction model, exact transient heat conduction solution, and lumped capacitance
solution for a solid object heat transfer in fluid
Figure 6 The matrix of nodes assignment for method of characteristic solution for Eqs. (9) and (17)
Figure 7 Dimensionless temperatures of the fluid flowing out from a tank in a charge and a following-on discharge
( the case has H CR  0.1 ,  r  0.04 ,  d  4.0 , and  c /  d  1.54 )
1.0  r1
HCR-n
CRn
 r2
 HCR-1
CR1
 r3
 r4
 r5
0.0
0.5  c / d 2.0 ri

Figure 8 Configuration of the charts including multiple graphs of  versus  c / d at a fixed  d


(a)  r  0.01 (b)  r  0.05

(c)  r  0.1 (d)  r  0.2

Figure 9 Computed results of multiple graphs of energy storage effectiveness versus  c / d at  d =4.0
1.0  d1
HCR-n
CRn
 d2
 HCR-1
CR1
 d3
 d4
 d5
0.0
0.5  c / di 2.0  di

Figure 10 Illustration of a configuration of the charts for energy delivery effectiveness using
multiple graphs at different  d but at a constant  r.
(a)  d  2.0 (b)  d  3.0

(c)  d  4.0 (d)  d  5.0

Figure 11 A series charts of energy effectiveness versus  c /  d at different  d and a constant value  r =0.03
Figure 12 Energy effectiveness versus  c /  d at  d =3 and  r =0.06
Figure 13 Energy effectiveness versus  c /  d at  d =2 and  r =0.04
Figure 14 Energy effectiveness versus  c /  d at  d =1.5 and  r =0.03

You might also like