You are on page 1of 8

HIGH TEMPERATURE HEAT STORAGE USING GAS-SOLID

REACTIONS

F. Schaube*, A. Wrner*, H. Mller-Steinhagen*+


*
German Aerospace Center - DLR, Institute of Technical Thermodynamics (DLR)
Pfaffenwaldring 38-40, 70569 Stuttgart, Germany
Franziska.Schaube@dlr.de
+
University of Stuttgart, Institute for Thermodynamics and Thermal Engineering (ITW)

ABSTRACT

Heat storage systems are used to improve energy efficiency of power plants and the recovery
of process heat. They are also required for continuous power supply in solar thermal
applications. Thermo-chemical reactions offer an option for high storage capacity even at high
temperatures. In an initial study, system requirements and suitability of the decomposition
reactions of calcium carbonate CaCO3 and calcium hydroxide Ca(OH)2 were identified. Main
concerns were the correct system integration of the complex energy and mass flows and
problems with cycling stability observed during thermo-gravimetric measurements. To
understand mechanisms and limitations on transport in a packed bed reactor a 2D-model has
been applied to describe heat and mass transfer during the reaction Ca(OH)2 CaO + H2O.
This paper reports on the storage characteristics in case a carrier gas is used as the medium to
transport the heat of reaction as well as the gaseous reactant.

1. INTRODUCTION

Thermal storage systems are usually restricted in terms of temperature and system conditions.
To extend operative ranges thermo-chemical storage has been proposed. Compared to
sensible or phase change systems not only heat transfer but also mass transfer has to be
considered. At first reactants must be heated to reaction temperature. When this temperature is
reached the exothermic or endothermic reaction takes place assuming that temperature and
pressure are in a range where the reaction is not limited by kinetics. Afterwards reactants have
to be stored. In case of solids or liquids the storage temperature may be the same as during
reaction whereas gaseous products have to be cooled down and pressurized or liquified.

Within a joint research project, ITW+ research is focused on low-temperature domestic


applications while DLR* is investigating reactions in temperature ranges suitable for power
generation. High temperature, high reaction enthalpy and thus high reaction entropy are
required (Teq~H/S). Therefore decomposition reactions are proposed for such systems.
Because of their high energy density and low cost gas-solid reactions are promising
candidates, especially decomposition reactions of metal hydroxides and carbonates. Table 1
shows achievable capacities. The storage capacity related to the solid is calculated with a bulk

Table 1: Achievable storage capacities


Storage capacity Storage capacity
Teq[1 bar] H [1 bar]
related to solid related to all reactants
C kJ/mol
kWh/m3 kWh/m3
Ca (OH ) 2 + H CaO + H 2 O 521 100 410 323
CaCO 3 + H CaO + CO 2 896 167 626 113
density being half the solid density. Storage capacity decreases if it is related to the volume of
all reactants. It is assumed that H2O is stored as a liquid and CO2 at a pressure of 50 bar.

Decomposition reactions of calcium carbonate, calcium hydroxide and magnesium hydroxide


were often investigated regarding heat pump applications as reaction temperature can be
controlled by the reaction pressure. In case of the carbonate system an additional loss arises
from compressing the reactant CO2 to gas storage pressure. When using hydroxides there is a
considerable need for low temperature heat for the evaporation of H2O. Hence, specific
system integration into the heat supplying and demanding process is necessary which results
in different efficiencies depending on system conditions. Kato proposed storing heat from a
cogeneration process at around 250-400C and discharge the heat from a magnesium
hydroxide system at around 100-170C [1]. Fujimoto suggests heat supply at 400C while
condensing water at 20C. The heat would be released at 80C during hydration of calcium
oxide with simultaneous evaporation at 17C [2]. Thinking of daily high temperature heat
storage for power plants discharge temperature should be close to charge temperature. For an
isobaric process this implies that decomposition (discharge) temperature is higher than the
temperature during the forming (charge) reaction. Figure 1 shows the additional amounts of
energy in a system of 100 kWh heat stored in the reaction calculated with thermo-chemical
data given by Barin [3]. Basically, it would be favorable to implement a storage that can be
operated at high temperature. Thus the decomposition is proposed to take place at 900C
(550C) and the formation at 800C (450C) in the CaCO3/CaO system (Ca(OH)2/CaO,
respectively).

Figure 1: Cycling characteristics, 1 bar CO2/H2O


The investigations are carried out for a partial pressure of CO2/H2O of 1 bar. Higher reaction
pressure is feasible but would demand a higher temperature during decomposition. On the
other hand lower (partial) pressure leads to a lower discharge temperature and could only be
acceptable for the hydroxide system. This is due to the complexity of gas separation and the
resulting constraint not to use a carrier gas. Without segregation the compression is more
energy-intense which is difficult to compensate by lowering the CO2 tank pressure as this
reduces the energy density. Also, lower total reaction pressure during decomposition would
require cumulative work for compression of carbon dioxide.
In comparison the Ca(OH)2/CaO system can be applied to a wider temperature range by
varying the total pressure or concentration of steam in an inert gas. However, the heat of
evaporation can not be neglected and may be only available at a certain temperature. As a
consequence the evaporation pressure determines the temperature level of discharge.
Nevertheless, if system conditions are suitable both of the reactions investigated will provide
a basis for high temperature heat storage systems with high energy density.
2. EXPERIMENTAL

Even if thermodynamic considerations reveal future potential of gas-solid reactions for


thermo-chemical heat storage the reactions may not meet demands regarding cycling stability
and reaction kinetics. Therefore first tests were carried out with commercially available
materials to learn more about their behaviour under the conditions specified in Figure 1.

Calcium carbonate (CaCO3, 99.6% pure, d50=5 m, Magnesia 449) and dolomite powder
(CaMg(CO3)2, 99.1% pure, d50=5.5 m, Magnesia 4179) were supplied by Magnesia GmbH,
Germany. Calcium hydroxide powder (Ca(OH)2, 100.4% pure, d50=4.78 m, CAL 54) was
obtained from Dr. Paul Lohmann GmbH KG, Germany. All powders were stored in a
desiccator before use. The reactions were carried out in a thermo-gravimetric analyzer of
Netzsch (STA Jupiter 449). About 10 mg of starting material were tabbed into the crucible.
Tests consisted of eight or four cycles in CO2 or H2O atmosphere, respectively.
CaCO3/CaMg(CO3)2 was heated to 1000C at a rate of 20C/min, and held at that temperature
for 10 min. Afterwards the sample was cooled down to 800C at the same rate and held there
for 90 min. Ca(OH)2 was heated to 600C at 5C/min and held at that temperature for 15 min.
Then it was cooled down to 450C at 5C/min and held there for 90 min.
Both the decarbonation and dehydration
occur rapidly and complete decomposition is
achieved in all cycles. Though carbonation
and hydration are not fully completed. At
first the formation reaction is controlled by
kinetics and takes place very quickly. Then
it is controlled by diffusion so that the
reaction rate is very low. Figure 2 shows the
degree of conversion after each cycle. While
the Ca(OH)2/CaO system is almost stable at
0.93 during four cycles there is less
conversion of CaO/CaOMgO to
CaCO3/CaCO3MgO with increasing cycling
number. An improvement is observed using
dolomite instead of calcium carbonate due to
its material structure but conversion also
Figure 2: Conversion of back reaction decreases after three cycles. In general the
reactions are reversible but diffusion in the
second step is slow. Therefore even for extended isothermal periods the reaction can not be
completed in an amount of time that is reasonable for a heat storage system.

Problems regarding carbonation have been reported before. Bhatia found the second stage
controlled by diffusion in the product layer [4] whereas Mess concluded that grain boundary
diffusion and diffusion through the carbonate crystal acted in parallel [5]. Varying the heating
regime Deutsch could observe the change of morphology to be responsible for thermal
characteristics [6]. The effect of pore closure on reaction kinetics could probably be an
important factor for other gas-solid reactions as well [7].
The hydration of calcium oxide has not been examined in such detail yet. An experiment of
1100 cycles was carried out by Bauerle which showed over 90% utilization [8]. The reactor
material has to be selected carefully in terms of deterioration though.
Looking at temperature and pressure ranges, applicability and material problems detailed
investigations of the hydroxide system appears to be more promising at this point.
3. MATHEMATICAL MODEL

Besides system integration and material development the reactor design is a main concern.
For use as a heat storage system minimal heat and parasitic losses and an adequately high
reaction and heat extraction rate are required. Ogura found the effective thermal conductivity
e of a fixed bed of pressed calcium hydroxide with a density of 1300 kg/m3 to be 0.38 W/mK
at 27C [9]. Enhancement could be achieved by compressing or integrating copper fins.
However, low thermal conductivity leads to the evaluation of other concepts rather than
indirect heat transfer. In this paper a fixed bed is investigated where a gas is used as a medium
to transport the heat of reaction H.

3.1. Governing equations Equation System

A continuous model is applied where the


solid is assumed to behave as a continuous (1 ) s = R (1)
t
medium and not as a medium of independent
particles. With the additional assumptions
r
that variation of porosity during reaction ( g ) + ( g u) = R (2)
and compression work are negligible, mass t
and energy balances in each phase are
described as follows: r K (3)
u = P
g
Mass balance for the solid phase r
r K c F K g u
Variation of the solid density s (Eq. 1) is due u = P /(1+ ) (4)
to the reaction of steam with the calcium g g
oxide or the decomposition of calcium
hydroxide at reaction rate R. The reaction dp 2wh
rate for both hydration and dehydration is c F = 0.551 5.5 , D e =
De w+h
derived from the kinetic measurements of
Matsuda as a function of temperature Ts, d 2p 3
conversion X, partial pressure PH2O and K= (5)
equilibrium pressure [10]. The latter is 180(1 - ) 2
calculated from the Vant Hoff relationship
derived from thermo-chemical data given by Tg r
g c pg + g c pg (Tg )u
Barin [3]. t
= ( g Tg ) + Rc pg Tg + h sg (Ts - Tg ) (6)
Mass balance for the gaseous phase
Variation of gas density g (Eq. 2) is rdue to
gas flow in the bed at velocity u and Ts
(1 - ) s c ps ( e Ts )
simultaneous hydration/dehydration. The gas t
is a mixture of nitrogen and steam with = R( H c ps Ts ) + h sg (Tg - Ts ) (7)
constant fraction PH2O/P with pressure P
deduced from the ideal gas law. Dynamic c ps = Xc pH + (1 X)c pO (8)
viscosity ug, heat capacity cpg and thermal
conductivity g are dependent on temperature
and concentration. Nu = 2 + 1.1Re 0.6
p Pr
1/ 3
(9)

Momentum balance for the gaseous phase


Gas velocity in the bed can be expressed by Darcys law (Eq. 3), adding the Forchheimer term
(Eq. 4) if Rep>1. Permeability K is defined according to Karman-Cozeny (Eq. 5) assuming
that the bed has width w, height h and consists of spherical particles with diameter dp.
Energy balance for the gas and solid phase
If the gas is not only reactant but is also used as a heat transfer medium the temperature of gas
T and solid Ts are not equal in the reaction zone. In that case two energy balances are required
(Eq. 6 and 7). The heat capacity for the solid phase cps is a function of conversion with the
heat capacities cpH of the hydroxide and cpO of the oxide at 500C (Eq. 8). The heat transfer
coefficient hsg is calculated according to Wakao (Eq. 9, [11]) and particle surface area.

3.2. Reference storage tank

As a reference a symmetric surface of a tank with radius r=0.273m, height h=0.273m and a
capacity of 26.3 kWh is investigated. Walls are assumed to be adiabatic. Initially, the
temperature, pressure and density are assumed to be uniform in the whole tank:
s (t = 0) = s0 , P(t = 0) = P0 and Ti (t = 0) = T0 .

Boundary conditions are given below.


On the lateral surfaces:
r rr r r
n( s )(x = 0, r) = 0 , nu (x = 0) = 0 , u (x = r ) = 0 and n( k i Ti )(x = 0, r) = 0 .

At the tank inlet:


r rr r
n( s )(z = h) = 0 , nu (z = h) = u in , Tg (z = h) = Tin and n( k s Ts )(z = h) = 0 .

At the tank outlet:


r r
n( s )(z = 0) = 0 , P(z = 0) = P0 and n( k i Ti )(z = 0) = 0 .

The system of 2D partial differential equations together with the initial and boundary
conditions is solved numerically by finite element method using the software COMSOL
Multiphysics. Time steps and grid are refined until the results do not depend on the number
of calculation points anymore. The properties used are indicated in Table 2. The simulation
results in a spatial solution of the evolution of temperature, pressure and solid density over
time.

3.3. Validity and assumptions


Table 2: Solid properties
The modeling approach and assumptions Density at end of discharge, 2200 kg/m3
ss
were theoretically and experimentally proved Density at end of charge, 1665 kg/m3
s0
for metal hydride applications [12]. Porosity, 0.5
Matsuda studied kinetics for particles with a Specific heat, cpO 934 J/kgK
diameter of 5 m [10]. Later modification for Specific heat, cpH 1530 J/kgK
0.7-1 mm particles was reported [13]. Thermal conductivity, e 0.38 W/mK
The ranges for temperature and volume Reaction heat of formation, H 100 kJ/mol
fraction of steam in per cent are 83 to 338C
and 2 to 15.7% or 420 to 450C and 1.5 to 6% for discharge and charge, respectively [10].
Variations will be made to clarify general dependencies.
The steam fraction is set constant because in this study there will always be an excess of
steam in the reacting zone so that the reaction rate will not decrease as a result of low mass
transfer rate.
4. RESULTS

In the beginning of discharge the storage has temperature T0. The gas mixture enters the
reactor at temperature Tin and starts to cool down the solid. Hydration begins after equilibrium
temperature Teq is reached. Then the gas is heated up in the reaction zone to Teq by the heat of
reaction. After full conversion this area is cooled down to Tin while the reaction zone is
moving through the reactor. To achieve a maximum temperature difference between in- and
outlet the pressure should be set so that Teq=T0. Dehydration proceeds accordingly.

To achieve conditions as specified in Figure 1 a pressure of 1.78 bar is required for discharge.
Steam enters the reactor at 450C at which the reaction preferably takes place and leaves the
reactor at 550C. Figure 3 shows the outlet temperature of the reactor during discharge for
two different pressures of H2O and same inlet velocities. Temperature decreases very quickly
to Teq and an overall conversion of 0.99 is achieved before the temperature drop at the outlet
exceeds 20C. The extension of the
charging process is a result of less
temperature difference between in-
and outlet and unequal mass flow rates
due to varied pressure. It was observed
that the overall conversion rate is
constant.
In Figure 4 the overall conversion after
ten minutes for different inlet
velocities and temperature differences
between in- and outlet are illustrated.
For this temperature and pressure
range the change of conversion rate is
almost linear. It is concluded that the
overall conversion rate is limited by
the maximum possible enthalpy
Figure 3: Outlet gas temperature Tout over time t change of the gas mixture.

For the following investigations conditions


are set to T0=420C, Tin=320C, P0=1 bar,
PH2O/P0=0.098 and uin=2m/s. In that case an
overall conversion of 0.94 in 47 minutes is
achieved before the temperature drop at the
outlet exceeds 20C.
Pre-exponential factor and heat transfer
coefficient are varied according to h*=khhsg
or R*=kRR. Changes in local conversion rate
and solid temperature distribution at x=0.1m
and z=0.25m are taken as a reference to
examine the impact of kinetics and heat
transport (Figure 5).
In case kinetics is slowed down (kR=0.1) the
reaction provides less heat than the gas
passing the reaction zone can take up. The
solid cools down and the reaction zone
Figure 4: Conversion X10 after 10 minutes
expands. As a consequence conversion rate
and also outlet temperature decrease significantly. On the other hand fast kinetics (kR=10)
results in a narrow zone of reaction being close to equilibrium. The local conversion rate
increases but the influence on the overall conversion rate and distribution of outlet
temperature is less pronounced.
In case the heat transfer coefficient is diminished (kh=0.01) the gas can not take up as much
heat and the reaction takes place at a higher solid temperature. Thus the reaction rate slows
down and the reaction zone expands which results in a higher temperature drop at the outlet.
On the other hand a higher heat transfer coefficient (kh=10) has no impact on the process due
to the limited thermal capacity of the gas flow.
The reaction is slightly controlled by kinetics but mainly restricted by heat transport.

Figure 5: Solid temperature Ts and local conversion X(0.1,0.25) at x=0.1m, z=0.25m over time t

These results can be translated to some extent to larger storage systems. A bed with the same
particle size shows the same conversion rates per length. However, absolute quantities are not
transferable because reactor height changes. With increasing height the pressure drop across
the system rises. Thinking of an energy storage system pumping losses have to be
compensated. Hence, particles have to grow. As outlined before this will have an impact on
the system characteristics dependent on the mass and heat transfer resistance inside the
particles. Also, material development is necessary to provide stable particle structures.

5. CONCLUSION

The decomposition reaction of calcium hydroxide is a promising candidate regarding the use
of a high temperature thermo-chemical storage system. As a result of high evaporation
enthalpy specific system integration is required which at the same time has influence on
reaction conditions. Material problems in terms of cycling stability and minimum particle
diameters are the main issues. If this could be solved a fixed bed reactor with direct heat
transfer could be applied to overcome low thermal conductivity. A mixture of nitrogen and
steam is proposed as the heat transfer medium.
A non - local thermal equilibrium continuous model was used to simulate heat and mass
transfer inside the reactor. Results show that the limiting factor on the reaction time is the heat
transport even though the process could be slightly improved by enhancing the rate of
reaction. For temperature and pressure ranges specified in this study the reaction is suitable to
be used in a storage system.
Reaction time and outlet temperature are sensitive to both heat transport and changes in
reaction rate. Therefore additional parameters for kinetics and heat resistance have to be
integrated into the model if larger particles or different material structures are used in major
storage systems.

ACKNOWLEDGEMENTS

This research is funded by the German Federal Ministry of Economics and Technology
(BMWi) under contract number 0327468A. The authors gratefully acknowledge this support.

REFERENCES

[1] Kato, Y., Sasaki, Y. & Yoshizawa, Y. (2003). Thermal Performance Measurement of a Packed Bed Reactor
of a Magnesium Oxide/Water Chemical Heat Pump. Journal of Chemical Engineering of Japan 36(7): 833-
839.

[2] Fujimoto, S., Bilgen, E. & Ogura, H. (2002). CaO/Ca(OH)2 Chemical Heat Pump System. Energy
Conversion and Management 43(7), 947-960.

[3] Barin, I. (1993). Thermochemical Data of Pure Substances. Weinheim, VCH Verlagsgesellschaft.

[4] Bhatia, S. K. & Perlmutter, D. D. (1983). Effect of the Product Layer on the Kinetics of the CO2-Lime
Reaction. AIChE Journal 29: 79-86.

[5] Mess, D., Sarofim, A. F. & Longwell, J. P. (1999). Product Layer Diffusion during the Reaction of Calcium
Oxide with Carbon Dioxide. Energy & Fuels 13(5): 999-1005.

[6] Deutsch, Y. & Heller-Kallai, L. (1991). Decarbonation and Recarbonation of Calcites heated in CO2 : Part
1. Effect of the Thermal Regime. Thermochimica Acta 182(1): 77-89.

[7] Beruto, D., Kim, M. G. & Searcy, A. W. (1989). Microstructure and Reactivity of Porous and Ultrafine
CaO Particles with CO2. High Temperatures - High Pressures 20: 25-30.

[8] Bauerle, G., Chung, D. K., Ervin, G., Hajela, G. P., Guon, J., Springer, T. H. & Rosemary, J. Storage of
Thermal Energy in Inorganic Oxides/Hydroxides. Energy System Group, Rockwell International, Canoga
Park.

[9] Ogura, H., Miyazaki, M., Matsuda, H., Hasatani, M., Yanadori, M. & Hiramatsu, M. (1991). Experimental
Study on Heat Transfer Enhancement of the Solid Reactant Particle Bed in a Chemical Heat Pump using
Ca(OH)2/CaO Reaction. Kagaku Kogaku Ronbunshu 17(5), 916-923.

[10] Matsuda, H., Ishizu, T., Lee, S. K. & Hasatani, M. (1985). Kinetic Study of Ca (OH)2/CaO Reversible
Thermochemical Reaction for Thermal Energy Storage by Means of Chemical Reaction. Kagaku Kogaku
Ronbunshu 11(5), 542-548.

[11] Wakao, N., Kaguei, S. & Funazkri, T. (1978). Effect of Fuid Dispersion Coefficients on Particle-To-Fluid
Heat Transfer Coefficients in Packed Beds. Chemical Engineering Science 34(1), 325-336.

[12] Joppich, F. (2009). Private Communication. PhD Thesis 2005-2009, not yet published.

[13] Fujimoto, S., Bilgen, E. & Ogura, H. (2002). Dynamic Simulation of CaO/Ca(OH)2 Chemical Heat Pump
System. Exergy 2: 6-14.

You might also like