Chapter 17

Thermochemical Energy Storage

Henner Kerskes    Research and Testing Centre for Solar Thermal Systems (TZS), Institute for Thermodynamics and Thermal Engineering (ITW), University of Stuttgart, Germany

Abstract

Thermochemical energy storage is a new technology which provides the advantage of high storage densities and minor thermal losses. This makes the technology attractive for low-temperature long-term storage as well as for high-temperature storage. The storage mechanisms range from physical adsorption to reversible chemical reactions and, therefore, a number of storage materials are available. New materials are under development to further improve the material properties with regard to storage quantity and heat transfer. In addition, innovative storage concepts are being developed for different applications and are being tested in the laboratory or in demonstration plants. The chapter presents the state of the art of thermochemical energy storage and describes the fundamental physical phenomena. An overview of common and recent storage material is given. Furthermore, the chapter introduces the principal concepts used for long-term and high-temperature storage. Finally, selected examples demonstrate implementation of the technology.

Keywords

thermochemical energy storage
sorption storage
hot-water store
radial stream adsorber
zeolite
solar thermal system

1. Introduction

The development of our society toward a sustainable society with a modern energy system based on a high ratio of renewable energies together with high energy efficiencies brings thermal energy storage into a new focus. The integration of renewable energies, such as solar or wind power, with technologies which have high efficiencies, such as combined heat and power, require new storage technologies and present new challenges to those working in the field of heat storage. Compared with today’s energy systems, compact and long-term storage processes will play an important role in future energy systems. High specific storage capacities and reduced heat losses are important technical aspects for future energy storage developments.
Numerous studies over the past few years have shown that thermochemical energy storage is a key technology to developing highly efficient short- and long-term thermal energy storage for various applications, such as solar thermal systems or cogeneration systems [1]. By storing energy in the form of chemical bonds of suitable materials, energy can be stored with almost no energy loss for long periods of time. At the same time, high energy storage density can be achieved. Both criteria are crucial for future energy storage applications.
Research activities in the field of low-temperature thermochemical energy storage (TCES) have developed strongly over the last few years—particularly in the field of material development and material optimization [25]. The main focus of this activity is on improving the chemical and thermal properties of materials such as increasing the energy storage density, enhancing the thermal conductivity, or improving cyclic stability. In addition, some attention has been paid to the design of the storage itself and its subcomponents such as the reactor. These topics are indispensable if the processes are to become commercially available.
High-temperature heat storage systems are used to improve the energy efficiency of power plants and the recovery of process heat. They are also required for continuous power supply in solar thermal applications. Thermochemical reactions offer an option for high storage capacity even at high temperatures [6].

2. Physical fundamentals of thermochemical energy storage

More than 90% of all thermal energy storage processes used in a wide range of applications are sensible heat storage processes. For temperatures below 100 °C, water is mainly used as the storage material. For higher temperature applications, solid storage materials such as ceramics or liquids in the form of molten salts are available. The technology of sensible heat storage is well understood and much experience has been gained from the vast range of examples that are available. The situation is completely different for TCES systems and few proven examples exist in spite of its promising potential.
In TCES systems, reversible chemical or physical reactions are used for storing and releasing energy and as such are very different from storing sensible heat. Only in the case of gas-phase adsorption processes is there any similarity to the mechanism of latent heat or phase change energy storage. In Fig. 17.1 a comprehensive overview of the different thermal energy storage mechanisms is given. Thermochemical processes are divided into two main branches: sorption processes, which again can be divided into adsorption and absorption; and reversible chemical reactions, which in turn can be divided into solid–gas reactions and solid–liquid reactions.
image
Figure 17.1 Classification of thermal energy storage mechanism.

2.1. Thermochemical Energy Storage (reaction)

Generally, thermochemical energy storage is based on the utilization of heat of reaction released by reversible chemical reactions. For example, a chemical compound of type A–B can be split reversibly into the components A and B by supplying heat.
In this process the supplied quantity of heat, denoted here by ∆RH, is used to break the A–B bond into independent species A and B. If the reverse reaction of turning products A and B into the compound A–B is avoided the energy stored in the chemical bonds can be stored without energy loss for any length of time.
An example of this type of reaction is the dehydration of salt hydrates, for example, magnesium hydroxide into magnesium oxide as shown in Eq. 17.1:

MgO+H2OMg(OH)2+RH

image(17.1)
The energy required for the endothermic reaction can be provided by heat from any source. If the salt oxide is stored in a hermetically sealed space, then the energetic state can be kept without energy loss for an unlimited period of time. This means that heat losses only occur during charging and discharging. If at any later time the oxide is brought into contact with water or water vapor, hydration occurs and the quantity of heat supplied during dehydration is released.
Typical reversible solid/gas reactions with potential for thermal energy storage are listed in Table 17.1. These reaction types cover a temperature range from below 100 °C up to very high temperatures of over 800 °C. Besides the potential for high-energy storage density this is a reason this technology is attractive for a wide range of energy storage processes.

Table 17.1

Examples of Reversible Solid/Gas Reaction for Thermal Energy Storage

Type of reaction Reaction Temperature range/°C
Dehydration of salt hydrates
MgSO47H2OMgSO4H2O+6H2O image
MgCl26H2OMgCl2H2O+5H2O image
CaCl26H2OCaCl2H2O+5H2O image
CuSO45H2OCuSO4H2O+4H2O image
CuSO4H2OCuSO4+H2O image
100–150
100–130
150–200
120–160
210–260
Deammoniation of ammonium chlorides
CaCl28NH3CaCl24NH3+4NH3 image
CaCl24NH3CaCl22NH3+2NH3 image
MnCl26NH3MnCl22NH3+4NH3 image
25–100
40–120
40–160
Dehydration of metal hydrides
MgH2Mg+H2 image
Mg2NiH4Mg2Ni+2H2 image
200–400
150–300
Dehydration of metal hydroxides
Mg(OH)2MgO+H2O image
Ca(OH)2CaO+H2O image
Ba(OH)2BaO+H2O image
250–350
450–550
700–800
Decarboxylation of metal carbonates
ZnCO3ZnO+CO2 image
MgCO3MgO+CO2 image
CaCO3CaO+CO2 image
100–150
350–450
850–950

2.2. Prototype of the Combined Hot Water and Sorption Store

A related process, also important in the field of TCES, is the adsorption process. Eq. 17.2 describes the basic principle of the charging and discharging process:

A(s)+xH2O(g)AxH2O(s)+adsH

image(17.2)
The basic reaction shown in Eq. 17.2 is very similar to the chemical reaction described in Eq. 17.1. During the energy-discharging process the solid reactant A is in contact with water vapor (H2O(g)). The solid adsorbs the water vapor to form the product A⋅x H2O(s) and the adsorption enthalpy (∆adsH) is released in the form of heat. As this is a reversible adsorption reaction the same amount of heat has to be applied to decompose the product A·x H2O(s) into water vapor and the adsorbent A. It should be noted that the supply heat during desorption is typically at a higher temperature than the heat released during adsorption.
As in the case of chemical reactions there is no loss of energy during long-term storage, so long as the adsorbent is kept in a dry, water vapor–free environment.
The technology of adsorption is widely used in the chemical industry for a range of processes which include waste water treatment and gas purification. The process of adsorption is very well known and intensively investigated [7,8]. Therefore, only a brief introduction will be given which is necessary to understand the adsorption process for thermal energy storage and to understand the concepts discussed in Section 4.
The process of adsorption describes the attachment of molecules or ions from a gas phase or a liquid to a solid surface. Adsorption is a phase transfer process and can be explained as an enrichment of chemical species from a fluid phase onto the (inner) surface of a solid material. The solid surface provides energy-rich active sites which interact with the gas-phase species. By adsorbing these species the energy level of the surface is reduced to a thermodynamically more stable condition as energy is released. The solid is referred to as the adsorbent, the gas-phase molecule is called the adsorptive, and the adsorbed molecule is the adsorbate. Fig. 17.2 illustrates the adsorption process. As shown in Eq. 17.2, adsorption is an exothermal process and heating up the solid reverses the process, the binding forces are overcome, and the adsorbed species are released from the adsorbent back into the fluid. Thereby the adsorbent is raised to a higher energy level. This process is referred to as desorption.
image
Figure 17.2 Schematic representation of adsorption and desorption.
The process of adsorption and desorption is an equilibrium controlled process. The amount of adsorbed mass per mass of adsorbent is referred to as loading X. Adsorption equilibrium depends on temperature and is defined by the partial pressure p* of the adsorptive in the gas phase and the adsorbent loading X. The equilibrium can be expressed as a function f (p*, T, X). Typically, the solid loading X is given as a function of the partial pressure p* at constant temperature T:

X=f(p*)T

image(17.4)
This relation is called the adsorption isotherm at temperature T. In Fig. 17.3, adsorption isotherms of H2O adsorption on zeolite 13X are depicted. Adsorption equilibrium is described by two dependencies:
at constant temperature the adsorbed amount of H2O increases with increasing water vapor pressure in the gas phase
the higher the temperature the lower the equilibrium at constant pressure.
image
Figure 17.3 Adsorption isotherms of water vapor on zeolite of type 13X [9].
Note: 1 mbar = 102 Pa.
This behavior of energy storage defines the boundary condition for charging and discharging of thermochemical energy. For the discharging process it is beneficial to operate at low temperatures and high partial pressure to achieve a high amount of adsorbed mass. In a thermal process, desorption is achieved by increasing the temperature of the storage material. Furthermore, low partial pressure supports desorption.
In addition to adsorption capacity, the heat or enthalpy of adsorption is one of the most important parameters in heat storage. The enthalpy of adsorption is the difference in energy between the adsorption process and the desorption process.
The enthalpy of adsorption can be described as the sum of heat of condensation (for gas adsorption) and additional bonding forces which are weak Van der Waals forces or electrostatic forces. The value of these bonding forces is characteristic of the adsorbent material. For example activated carbon shows rather small bonding forces while zeolites have higher bonding forces. Due to the fact that the surface energy of adsorbent is inhomogeneous, preferential sites of adsorption exist. The first molecules adsorb on the active sites with the highest bonding forces. Adsorption enthalpy is, in most cases, a function of the loading X. In Fig. 17.4 the heat of adsorption of water vapor on zeolite 13X is depicted as a function of loading X. Starting at a high energy state, enthalpy decreases continuously and ends up slightly above condensation enthalpy.
image
Figure 17.4 Dependency of heat of adsorption on loading X for zeolite 13X [10].
To take advantage of high energetic adsorption sites, almost complete desorption of the storage material is necessary. Compared with Fig. 17.3, very high temperatures or nearly zero water vapor pressure is mandatory to reach this energetic state. However, the adsorption processes of typical storage applications start at values of (0.05–0.1) g g–1 and, therefore, high energetic sites are not included in the process.
The thermal storage capacity provided by an adsorption process is the product of adsorption capacity and adsorption enthalpy. The specific mass amount of thermal energy stored can be expressed as:

q=Xadsh=(XadsXads)XdesXdasadsh

image
and taking the mass of adsorbents into account the absolute value can be calculated as:

Q=madsXadsh=mads(XadsXdes)adsh¯

image
where Q is stored energy; mads is the mass of adsorbents; Xads is the amount of adsorbate after adsorption; Xdes is the adsorbed amount after desorption; and ∆adsh is mean adsorption enthalpy in the interval (Xads – Xdes).
Storage density can be determined by dividing the value of Q by the volume of the storage material. Adsorption capacity and adsorption enthalpy are both dependent on the operating conditions during adsorption and desorption. Therefore, storage density must be specified in terms of the adsorption and desorption conditions.

3. Storage materials

Recently, there has been much interest shown in finding new materials for sorption or thermochemical storage for low- and high-temperature storage applications. This is largely due to the small thermal losses and the high storage density of thermochemical heat storage as well as the fact that it can be used for both low- and high-temperature applications. Even if the applications are very different there are general requirements a thermochemical storage material should fulfill. A high-performing storage material is characterized by:
high adsorption capacity (water uptake)
high heat of adsorption is required for high energy storage density
fast reaction kinetic is desirable for high thermal power output during charging and discharging
the desorption temperature should be at an appropriate level.
Compared with chemical reactions whose reaction kinetics strongly depend on the reaction temperature (Arrhenius law), some adsorption processes show high kinetics even at low temperatures. This makes the adsorption process attractive for low-temperature applications such as space heating and domestic hot-water preparation where heat is required at temperatures below 100 °C. The processes of water vapor adsorption on zeolite and silica gel are the most studied processes for thermochemical energy storage [1114].

3.1. Adsorption Materials

Adsorption materials are characterized by high porosity and a large inner surface. In technical applications the materials are used in different forms, for example, granules such as spheres or cylinders or similar forms with typical dimensions of (1–5) mm. Fig. 17.6 illustrates the pore structure inside a granule. The pores may be subdivided in macropores, mesopores, or micropores. A classification according to the pore diameter is given in Table 17.2. The diffusion of gas-phase molecules into the inner region of the particles occurs in the larger macropores and mesopores, while adsorption takes place in the micropores.

Table 17.2

Classification of Pore Size

Pore size/nm Classification Specific inner surface/(m2 g–1)
<2 Micropores >400 image
2 < d < 50 Mesopores 10–400
>50 Macropores 0.5–2

Classical adsorption materials are silica gel, zeolite, and active carbon. Their suitability depends upon the individual application. Active carbon is a cheap and robust adsorption material with rather small adsorption capacity and low adsorption enthalpy. It is often used for the adsorption process in purification, heat pump, or cooling applications. Silica gel and zeolite show better performance for energy storage because the product of adsorption capacity and adsorption enthalpy is much higher (compare Table 17.3). Most adsorption storage systems described in the literature use silica gel/water or zeolite/water as working pairs [15]. Compared with zeolite, silica gel shows a lower desorption temperature (t < 120 °C) which results from the lower adsorption enthalpy. From the technical point of view this may have advantages. On the other hand, the adsorption kinetics of silica gel are slower than for zeolites. Adsorption measurements carried out by Jähnig et al. [12] showed that the achievable temperature lift decreases with increasing loading of the silica gel.

Table 17.3

Characteristic Values of Thermochemical Storage Materials: Active Carbon, Silica Gel, and Zeolite

Characteristic values Active carbon Silica gel Zeolite
Inner surface/(m2 g–1) 650–750
Mean adsorption enthalpy/(kJ kg–1H2O) ∼2400 ∼2600 ∼3500
Specific heat capacity cp /(kJ K–1 kg–1) 0.709 0.9–1.0 0.8–0.9
Heat conductivity/(W m–1 K–1) 1.2–1.6 0.14–0.2 0.58

Temperature lift denotes the temperature increase occurring in the bed during adsorption. If temperature lift becomes too small for effective heat transfer the remaining adsorption capacity is lost.
An important advantage of zeolite is its high adsorption kinetics. The fast adsorption of gas-phase molecules in combination with high adsorption enthalpy yields high heat release. Steep adsorption fronts occur which is beneficial for process control. Even very low adsorptive concentrations can be completely adsorbed which makes zeolites perfect for working with moist air.
Of the wide range of different zeolite types, types A, NaX, and NaY have been tested successfully for energy storage. Type A zeolite (technically known as 4A and 5A) is a robust and inexpensive material with good hydrothermal stability. NaX (13X) and NaY-type zeolites show higher adsorption enthalpies than 4A and are very good candidates for thermal energy storage. Further improvements have been achieved by synthesizing novel binderless zeolites of type 4A and 13X [16]. This new class of zeolite is characterized by improved adsorption capacity, sufficient hydrothermal stability, and higher adsorption enthalpies.
In addition to classical adsorption materials, new materials are under development. The material class of aluminophosphate (AlPOs) and silicoaluminophosphate (SAPOs) represents a group of microporous materials with a zeolite-like crystalline structure. These materials are less hydrophilic than zeolite and may have the potential to fill the gap between zeolite and silica gel regarding optimization of adsorption strength and the ability to be desorbed at low temperatures. The structural types of ALPO-17 and ALPO-18 as well as SAPO-34 look very promising for heat storage with respect to the adsorption equilibrium. Experimental investigations [17] have shown that under moderate operating conditions the water uptake is much higher compared with zeolites. Even at desorption temperatures below 100 °C, significant adsorbed quantities of 25% in mass have been obtained. However, the application of AlPOs and SAPOs as heat storage material is at present not suitable due to cost reasons. The high production costs of the material are caused by the organic template which disappears during synthesis.
Very promising adsorption results has been obtained with another class of microporous materials called metal–organic framework (MOFs) materials. This material class is characterized by a huge internal surface and outstanding adsorption capacity. Water uptake values up to 1.4 g of water per gram of MOF have been measured [18]. MOFs exist in a wide range of different compositions. The large variety and adjustability of the pore structures enables tailored modification to improve the material with respect to heat storage. Due to these attractive properties the technology of MOFs is undergoing fast development. The component MIL-101 has been identified as material with excellent water adsorption properties. Important improvements have already been achieved concerning hydrothermal stability. Some MOFs are already commercially available but are still very expensive.

3.2. Salt Hydrates

Reversible solid–gas reactions offer the potential of even higher storage density than adsorption processes as a result of high enthalpies of reaction. Solid–gas reactions are easy to handle if the gas-phase reactant is present as a natural part of the environment, such as water vapor. In an open process neither condensing nor storing of the gas-phase reactant is necessary. This is the case for the reaction of salt hydration. Table 17.1 illustrates that the temperature range of reaction equilibrium is well suited for low-temperature applications. Furthermore, heat transformation is high and the theoretically achievable storage density is six to ten times higher than a hot-water store (temperature change ∆T = 50 K). However, handling and practical use of these materials are complex. In recent research projects, different salt hydration systems have been investigated by several researchers. Favorably reviewed were sulfates such as magnesium sulfate (MgSO4), chlorides such as calcium chloride (CaCl2), or magnesium chloride and strontium bromide. [19]. Besides the theoretical potential some drawbacks do exist.
In the following, typical difficulties encountered during hydration and dehydration are discussed using the magnesium sulfate as an example. Magnesium sulfate monohydrate (MgSO4 . H2O) has high potential as a chemical storage material. Taking into account the enthalpy of condensation of water vapor a theoretical storage density of 2.3 GJ m–3heptahydrate (633 kW h m–3heptahydrate) can be obtained. This is about 11 times higher than the storage density of water storage with the same volume (∆T = 50 K).
Experimental investigations have been carried out by Bertsch et al. to analyze the reaction behavior of MgSO4 in a fixed bed reactor [20]. The reactor had a length of 20 cm and a diameter of 3.5 cm and moist air was passed through it. The inlet conditions of a fixed bed reactor can be adjusted with regard to mass flow rate, temperature, and water vapor partial pressure. In experimental investigations it was found that hydration does not reach the heptahydrate state. Starting with magnesium monohydrate, water uptake was much smaller than expected. The uptake of only three to four moles of water per mole of MgSO4. . H2O has been measured. There are different hydrates of magnesium sulfate, that is, mono-, bi-, tetra-, penta-, hexa-, and heptahydrate, even though not all of these hydrates are stable or crystalline. The equilibrium curves of mono-, hexa-, and heptahydrate are depicted in Fig. 17.5. To calculate equilibrium curves at atmospheric pressure, thermodynamic data have been taken from [21], [22], and [23]. Fig. 17.4 also depicts the reactor inlet conditions of the air flow (temperature, partial pressure of water vapor) of each experiment (black dots) and the maximum temperature lift achieved (bars to the right). In experiments 1, 2, and 3 the reactor inlet temperature was comparatively low. Furthermore, the inlet conditions were very close to the equilibrium of magnesium sulfate hexahydrate to heptahydrate. Both effects result in a slow reaction rate and a moderate temperature lift. In experiment 4 and 5 the higher inlet temperature favors hydration of the anhydrate and a higher reaction rate, a shorter hydration time, and a higher temperature lift was observed. The highest temperature lift of 30 K was obtained in experiment 6. However, a fully hydrated state of the magnesium sulfate cannot be achieved under these conditions as the reactor inlet condition is to the right of the equilibrium curve of hexahydrate to heptahydrate.
image
Figure 17.5 Equilibrium curves of MgSO4 and its hydrates and reactor inlet conditions of the air flow (point 1 to 6) for the different experiments.
Note: 1 mbar = 102 Pa.
With increasing hydration of the salt a decreasing reactor temperature was observed in all experiments. This is due to the approach of reaction equilibrium as hydration of the salts proceeds. A decreasing reaction rate is the result and the lower enthalpy of reaction yields a small temperature lift. Furthermore, the maximum achievable reactor temperature is limited by the equilibrium of the intermediates. Magnesium sulfate monohydrate can absorb water at a higher temperature than a more hydrated salt such as magnesium sulfate hexahydrate. The storage densities achieved were below the theoretical value for all experiments.
Similar results have been reported by other researchers [24]. Their experiments show that the practical use of pure magnesium sulfate is difficult because of its low power density. They concluded that the application of magnesium sulfate as a thermochemical storage material is problematic. In addition, many authors report that MgSO4.7H2O melts or boils in its crystal water before degrading to lower hydrates [25,26]. This leads to crystal growth and bonding which hinders the hydration process by increasing diffusion resistance.
These findings are transferable to most other salt hydrates investigated for thermochemical energy storage. However, good experiences with pure strontium bromide as storage material were reported by Wyttenbach et al. [27].

3.3. Composite Materials

Much better performance has been achieved for highly dispersed salt on a carrier matrix [28,29]. The carrier matrix can be of passive or active type. By active, we mean that the supporting matrix takes part in the adsorption process.
In most cases zeolite is used as an active matrix for this purpose. A passive matrix can be any porous medium. Materials with a well-defined pore structure and pore size distribution have shown interesting results. Some of these storage materials already show good characteristics in terms of reaction kinetics, energy storage density, and mechanical stability. Composite salt and zeolite materials are prepared by impregnating commercially available zeolites (e.g., zeolite 13X or zeolite 4A particles) with a salt solution. Early experiments showed an increase in energy storage density compared with pure zeolite of approximately 20% [28,3032]. The composite material showed similar behavior to that of pure zeolite—a high reaction rate associated with high temperature lift even at low water vapor pressures. To date these new composite materials have not been characterized in detail and the mechanism of interaction between salt and zeolite is not fully understood. Deeper insight into the reaction mechanism is necessary to develop composite materials which fulfill the expected targets which include tailored properties for energy storage.
Fig. 17.6 compares the experimentally achieved storage density of some thermal energy storage materials. Commercially available materials are silica gel, zeolite 4A, zeolite 13X, and a new type of binder-free zeolite (13XBF). In addition, the results of two noncommercially available composite materials are depicted. Composite A is made of a passive carrier matrix impregnated with MgCl2 and reported by Zondag [33]. Composite B is zeolite 13X matrix impregnated with MgCl2. The volume of the probes was 200 mL. All materials were desorbed at 140 °C in a dry air stream (p* < 102 Pa or 1 mbar). Adsorptions were carried out at 30 °C and water vapor pressure of 20 × 102 Pa (20 mbar) For comparison reasons the storage density of water (sensible heat storage) for a change in temperature of ∆T = 50 K is also depicted.
image
Figure 17.6 Comparison of storage materials.
Operating conditions: desorption 140 °C in dry air; adsorption 30 °C and 2-kPa (20 mbar) water vapor pressure.
In recent years a wide range of thermochemical storage materials have been analyzed and tested and new materials are under development [34]. Good progress has been made regarding increasing performance, hydrothermal stability, and cost reduction.

4. Thermochemical storage concepts

The technology of TCES is expected to have high technical potential not only for compact short-term storage applications but also long-term and high-temperature storage applications. Besides the development of a wide range of new or improved storage materials, innovative storage concepts have been developed and tested under laboratory or pilot scale conditions [35,36]. At present no commercial thermochemical heat storage material is available on the market. Nevertheless, the positive outcomes of recent research activities point to expected market entry in the near future. In the section a discussion of the principal operation modes using an adsorption process for thermal energy storage is given.
The operation of thermochemical heat storage systems can be divided into open- and closed-system designs (Fig. 17.7). An open system works under atmospheric pressure and is in contact with the environment. This means the gas-phase reactant, in most cases water vapor, is extracted from the environment for the adsorption process and is released to the environment during desorption. By contrast, in a closed system the water vapor circulates in a hermetically closed loop typically under negative pressure. The evaporation and condensation of water has to be enforced by technical devices, and an additional water reservoir is necessary. From the design point of view the open system has the advantage of lower technical effort: no condenser, no evaporator, no water reservoir is necessary, it works at normal pressure and process control is not complex.
image
Figure 17.7 Comparison of open- and closed-system concept.
The open- and closed-system operation modes are presented in more detail to explain the difference and to discuss the advantages and disadvantages of both system designs.

4.1. Closed-System Operation Mode

The closed system consists of a vessel, containing the adsorbent, which is referred to as the adsorber. The adsorber is equipped with a heat exchanger for heat input during charging and to extract the heat of adsorption during discharging. Evaporation and condensation of the adsorptive can be performed in the same unit (heat exchanger) because the processes do not take place simultaneously. Furthermore, a water reservoir is needed to store the condensed water. The volume of the water reservoir should be dimensioned according to the maximum water uptake of the adsorbents (i.e., approximately 30% of the adsorber volume).
The process of TCES can be subdivided into three steps. These three steps are illustrated in Fig. 17.8 and are described as follows:
1. Desorption: charging the store
During the charging process the adsorber is heated to a temperature high enough to almost dry the storage material. The heat Qdes has to be supplied to the storage material at a high temperature. Water vapor is released from the adsorbent and condensed in the condenser at a low temperature. The heat of condensation can either be used as a low-temperature heat source or has to be rejected to the environment. The liquid is stored in the reservoir.
One of the most crucial points for attaining high storage density is achieving almost complete drying out of the adsorbent. The higher the temperature inside the adsorber and the lower the temperature of the condenser the better the desorption.
To give an example let us assume zeolite 13X and water as a working pair. To attain a minimal water loading of 80 g kg–1 at a desorption temperature of 180 °C, the temperature of the condenser should be below 7 °C, which is the dew point temperature of water at 103 Pa (10 mbar) (compare with Fig. 17.3). In other words, good desorption requires either a very high desorption temperature or a very low condenser temperature. This depends on the application. After charging, the storage material and components will cool down to ambient temperature. If sensible energy cannot be used to meet the load it occurs as a heat loss.
2. Storage.
From the moment the system is at ambient temperature no further energy losses occur. As long as the storage material is hermetically sealed the energy of adsorption can be stored over an arbitrary time without any losses.
3. Adsorption: discharging the store.
Heat Qin is supplied to the evaporator at a low temperature level to evaporate the liquid water. The resulting steam is adsorbed in the adsorber and the heat of adsorption is released. The adsorption material (adsorbent) is thus heated up. The released heat Qads can be used at a higher temperature level for heating purposes. This process can be driven as long as adsorption equilibrium is reached.
image
Figure 17.8 Schematic sketch of closed adsorption process.
This means low-temperature energy for evaporation is supplied to gain useful heat at a higher temperature level during adsorption. In fact, from the thermodynamic point of view, adsorption energy storage is a heat pump cycle.
Due to the fact that adsorption is a completely reversible process the amount of heat supplied for desorption Qdes is exactly equal to the heat Qads gained back during adsorption. To fulfill the energy balance the heat of condensation must be equal to the energy of evaporation. In fact, the lower the adsorbent temperature and the higher the water vapor pressure in the evaporator the higher the maximum water loading of the adsorbent (compare Fig. 17.3).
The challenge is to extract the heat from the adsorber. Furthermore, a low adsorbent temperature is a requirement to keep the adsorption process upright and to extract constant power from the store. The main heat transport mechanism is heat conduction. Unfortunately, the heat conductivity of adsorption materials like zeolite is very low (0.1–0.5 W m–1 K–1). Therefore, the heat exchanger design plays an important role in storage performance.
The second power limitation aspect is the performance of the evaporator. The power extracted from the adsorption process equals the power of the evaporator multiplied by the ratio of adsorption enthalpy to evaporation enthalpy (for zeolite it is approximately a factor of 1.5). In Section 4.1 an example of a closed adsorption heat storage system is given.

4.2. Open-System Operation Mode

The open adsorption system is less complex in terms of apparatus und design. It consists mainly of an adsorber, a fan, and one or more heat exchangers. The open mode requires an adsorptive which is naturally present in the atmosphere, such as the water vapor within ambient air.
In Fig. 17.9 the operation of an open adsorption store is depicted schematically. Again the process can be divided into three steps:
1. Desorption: charging the store. For charging the store no heat exchanger inside the adsorber is required—just hot air blown through the store to heat up the storage material to the desired desorption temperature. Using zeolites, which are very efficient in combination with water vapor as the adsorptive, temperatures between (180 and 200) °C are favorable. The desorbed moisture is transported out of the adsorber by the air flow and released to ambient. Referring to Fig. 17.3, using ambient air heated to 200 °C (water vapor pressure 10 × 102 Pa or 10 mbar) will lead to a minimum water loading of 80 g kg–1.
2. Storage. During the storing phase the storage material must be kept free of moisture.
3. Adsorption: discharging the store. Control of the discharging process is very simple. A fan is switched on to blow moist air through the storage system. Adsorption takes place, the solid is heated up by the release of adsorption enthalpy, and in turn the solid heats up the air flow, which now leaves the store very dry and at a higher temperature. This hot air flow can be used for heating purposes.
image
Figure 17.9 Schematic sketch of open adsorption process.
The open adsorption process is illustrated in Fig. 17.9. The adsorber, or more generally, the reactor is the key element of the system. The reactor is the apparatus where the chemical reaction/adsorption takes place. Two general approaches of reactor design can be found in the literature: the integrated reactor concept if the material reservoir is also the reactor (Fig. 17.10 left) and the external reactor concept if the material is stored separately from the reactor (Fig. 17.10 right).
image
Figure 17.10 (a) Integrated and (b) external reactor concept for a thermochemical energy store.

4.2.1. Integrated Reactor Concept

In the integrated reactor concept the material is stationary inside the storage system. The main advantage is its simplicity. The type of reactor is very similar to a conventional fixed bed reactor. This is a known technology and many experiences of designing and operating of fixed bed reactors are available [37]. However, depending on the capacity of the store the size of the reactor may become very large. Reactors with large quantities of storage material (m > 300 kg) require subdivision of the volume for several reasons. An important aspect is the reduction in pressure drop when blowing moist air through the reactor. Low pressure drop is essential to limiting the electrical consumption of the fan/blower and is a necessity for high storage efficiency. Furthermore, dividing the storage mass into smaller parts is favorable and improves the heat and mass transfer inside the storage system. Reduction of the thermal mass heated up immediately reduces heat losses during desorption and adsorption. Additionally, this measure decreases the thermal response time of the system. This is an important improvement under transient operation conditions such as solar thermal applications. However, high temperatures (t > 120 °C) are needed for endothermic regeneration reaction/desorption. This implies the use of temperature-resistant materials throughout the entire TCES system.
In contrast to the external reactor design, material transport is not required. This implies less material stress as well as no technical or energetic effort for transporting the storage material.

4.2.2. External Reactor Concept

The external reactor concept separates the reactor from the storage material reservoir. Material transportation (e.g., vacuum conveying system) is required to move the material from the material reservoir to the reactor and vice versa. By separating the material reservoir and reactor the reaction is reduced to only a small part of the total storage material amount at any one time. Thermal heat capacities and heat losses, especially during the regeneration process, are reduced. Furthermore, the high temperature during material regeneration is restricted to the reactor only. As a result, the power of the storage system and storage capacity are decoupled. The design of the reactor defines the power of the system while storage capacity depends on the size of the storage vessel only. This insures a large degree of freedom in the ratio of power to capacity and enables good scalability. The storage vessel can be simple in design and no insulation is required.
Comparison of the integrated and external reactor concepts is given in [35].

5. Selected examples

At present a number of concepts have been developed and some prototypes have been built, which shows dynamic development in this technology. This section describes selected examples of TCES concepts introduced in Section 4. This includes open and closed adsorption systems, systems using salt hydration as the storage mechanism, as well as high-temperature storage via dehydration of metal hydroxides. A detailed review of long-term sorption energy storage is given by N’Tsoukpoe [38] and Bales et al. [15].

5.1. Closed Adsorption Storage Systems

Thermochemical storage for solar space heating in a single-family house has been developed at the Institute for Sustainable Technologies, (Austria) (AEE Intec) [12]. In a project called MODESTORE a closed adsorption system has been developed. This system operates under vacuum conditions with silica gel and water as the working pair. It consists of a storage material vessel with an internal spiral heat exchanger, a water reservoir, and an evaporator/condenser unit. A schematic drawing of the sorption store is depicted in Fig. 17.11.
image
Figure 17.11 Drawing of prototype of the MODESTORE project [12].
Charging the storage system (material desorption) is done by heating the storage material via the internal heat exchanger. Desorbed water vapor is condensed in the condenser and pumped into the water reservoir. During the discharging process (adsorption) water from the water reservoir is pumped into the evaporator to produce vapor. The vapor is adsorbed by the storage material and the heat of adsorption is released. The heat is transferred to the space-heating loop via the internal heat exchanger.
A pilot system of the sorption store was installed and monitored in a building located in Gleisdorf. The system consists of a 32 m2 flat plate collector area, a 900 L water buffer store, two 500 kg sorption stores, a separate water store for hot-water preparation, and an auxiliary heater (wood pellets). Such a system has been successfully tested. However, experimental results remained below expected values for storage density. One main reason for the insufficient storage density originates from the working pair of silica gel and water vapor. The achieved temperature lift decreases with time during the adsorption process. A necessary temperature lift of at least 10 K was only reached during the first 35% of total storage capacity.
This research has been continued within the framework of the European project COMTES (Combined Development of Compact Thermal Energy Storage Technologies) [39]. The aim was to overcome these drawbacks using storage materials with higher performance and further improved system technology. A hydraulic scheme of the system is depicted in Fig. 17.12. A storage volume double the 1000 L adsorption store was realized in stainless steel vessels. Binderless 13X zeolite was used as the storage material which is characterized by high adsorption capacity and fast adsorption kinetics [9,40]. High-performance vacuum flat plate solar thermal collectors serve as the heat source for desorbing the storage material during summer. The heat for evaporation is taken from the environment, using an air heat exchanger unit. In addition, the solar thermal collectors deliver low-temperature heat for evaporation in case of insufficient ambient conditions.
image
Figure 17.12 Scheme of the closed adsorption storage system developed within the project COMTES. (Source AEE Intec.)
The focus has been heat and mass transfer inside the sorption unit. A heat exchanger design was developed which provides a high heat transfer rate. First experiments have shown that utilization of the storage material and the achieved temperature lift inside the store were successfully increased. The steam generated by the developed evaporator/condenser unit was continuously controlled and allowed constant power extraction of several kilowatts from the storage system. The demonstration plant was able to analyze thermal performance under realistic conditions. Fig. 17.13 shows the experimental setup and the solar thermal collectors installed at AEE Intec.
image
Figure 17.13 Experimental setup of 2 m2 closed sorption storage at AEE-Intec (left) and vacuum flat plate collector for heat production (right). (Pictures AEE Intec.)
A similar closed adsorption storage system has been developed in another European project [41]. ZeoSys together with the Fraunhofer Institute for Interfacial Engineering and Biotechnology (IGB) has developed a compact adsorption storage system with the aim of storing industrial waste heat and heat from combined heat and power systems [41]. Basic investigations have been carried out in a small 1.5 L storage system where the performance of different sorption materials has been tested. Important process parameters, such as temperatures and pressure ranges, have been analyzed with respect to storage density and power output. Power is a crucial issue when using a solid storage medium with very low thermal conductivity such as zeolite (λ ≈ 0.1 W m–1 K–1). Several designs of heat exchangers have been tested in a medium-scale reactor of 15 L volume. A good solution has been found with a heat exchanger configuration that showed a heat power rate more than 60% higher than measured with a parallel copper plate heat exchanger in the bulk. In the final step the improved system concept has been tested successfully in a fully functional storage system with 750 L of storage volume.
Both projects are promising examples for a technology that appears to have a number of advantages for heating applications as well as for industrial process heat applications.

5.2. Open Adsorption Storage Systems

Within the research project SolSpaces taking place at the University of Stuttgart (Germany), a new solar heating system has been developed with the aim of supplying the demand for space heating of energy-efficient buildings completely on a solar basis. The core element of the system is a sorption store, serving as seasonal energy store. In Fig. 17.14 the test building at the University of Stuttgart is shown. The building consists of a living space of 48 m2 and has an annual heat demand of approximately 2,500 kW h.
image
Figure 17.14 SolSpaces building at University of Stuttgart.
The solar heating concept implemented is based on a solar thermal system in combination with a sorption heat store and is designed for very high solar fraction (up to 100%). Solar fraction is the amount of energy provided by the solar thermal system divided by the total energy required. Fig. 17.15 illustrates the operating mode of the solar heating system during summer and winter schematically. The sorption store is integrated between the controlled ventilation of the building and the indoor air leaving the space-heating zone. In this case the required moisture to achieve space heating by adsorption is supplied by indoor air. In winter, indoor air at room temperature (∼20 °C) is allowed to pass through the sorption storage. The storage material adsorbs moisture from the air. The heat of adsorption is released and air flow heats up significantly above 20 °C. This warm air is then allowed to pass through the heat exchanger, where it heats up incoming fresh ambient air. The incoming air flow affects room heating. The power of the heating system is limited by the humidity of indoor air and the air change ratio. Under typical operating conditions the system delivers power of (1–1.2) kW. This is sufficient approximately 85% of the time in the heating season.
image
Figure 17.15 Summer and winter operation mode of the solar thermal heating system with an open sorption store.
In the summer months, when solar radiation is present in excess, the storage material is regenerated to charge the sorption store (desorption). Vacuum tube solar collectors are used to heat the ambient air. The hot air is then allowed to flow through the sorption store system to regenerate the storage material. The warm air leaving the sorption system is then passed through the heat exchanger to preheat incoming ambient air, before it enters the solar collectors.
A segmented storage system, subdivided into individual sorption units, has been developed. The sorption store has a cubic shape and a volume of 4.2 m3. For efficient operation, it is necessary to divide the sorption store into individual segments. Due to segmentation a significant reduction in pressure drop is achieved. At the same time, the mass of the storage material involved in the adsorption/desorption processes is reduced. This leads to reduced heat losses due to smaller thermal capacity. Furthermore, regular and homogeneous flow distribution within each segment is an important requirement for the efficiency of the store. These considerations result in the storage concept, which is schematically depicted in Fig. 17.16. Four areas arise as a result of two vertical partitions on the diagonals, each of which is subdivided by horizontal planes.
image
Figure 17.16 Concept of the SolSpaces segmented open sorption store.
At INSA de Lyon, the concept of an open adsorption system using exhaust air from the building as the water vapor source has been built. Hot air with a temperature between (80 and 150) °C will be provided in summer by air collectors for desorption. A new composite storage material, consisting of zeolite 13X as carrier matrix and 15% mass percent of magnesium sulfate, has been developed. Composite materials are characterized by simultaneous adsorption and hydration reaction. In a laboratory test reactor with 200 g of composite material a temperature lift of over 25 °C has been measured during the combined adsorption/hydration process (air flow rate of 8 L min–1 at 25 °C and relative humidity of 50%). A volumetric energy storage density of 166 kWh m–3 has been achieved. This is 65% of the theoretical storage density of the material (257 kW h m–3) and a 27% increase compared with pure zeolite 13X (131 kW h m–3). Microcalorimetry experiments revealed that energy density can be maintained over at least three charge and discharge cycles. Further research will focus on improving the carrier matrix and enhancing the energy storage density.
An open sorption system for long-term solar heat storage with magnesium chloride on a carrier matrix as storage material is under investigation at the Energy Research Centre of the Netherlands (ECN) [33,42]. A prototype reactor, containing 15 L of storage material has been successfully tested. The temperature for dehydrating the material was set to 130 °C. During hydration experiments a temperature lift of the airflow from (50 to 64) °C has been measured in the material bed [12 × 102 Pa (12 mbar) partial pressure of water vapour]. With further improvement of the reactor design and the heat exchangers it is expected that high energy density of approximately 280 kW h m–3 of the material can be technically reached.
The external reactor concept has been developed at the University of Stuttgart’s Institute of Thermodynamics and Thermal Engineering (ITW). A new process design for solar thermal long-term heat storage has been developed [43] and a solar thermal combination system has been extended by incorporating a thermochemical energy store. In Fig. 17.17 a schematic drawing of the system concept is depicted. Similar to the concepts of the Solar Spaces project, it is an open adsorption/hydration system using ambient or exhaust air to provide humidity for the reaction.
image
Figure 17.17 Sketch of the external reactor design developed at ITW.
The thermochemical energy store works as a low-power heating system and is connected to the combistore of the solar system via the collector loop heat exchanger. The thermochemical energy store consists of a material reservoir for the storage material and a reactor where heat and mass transfer take place during the reaction. The external reactor concept separates the storage material (a composite material of zeolite and salt) from the reactor. This has the advantage that the reaction is reduced to only a small part of the total storage material amount at a time. Thermal heat capacities and heat losses especially during the regeneration process are reduced. Furthermore, only the reactor has to withstand high temperatures whereas for the material reservoir low-cost materials can be used. Storage material transport between the material reservoir and reactor is done by a vacuum conveying system allowing very gentle material transport with low energetic expense.
The reactor is designed as a cross-flow reactor. The material enters the reactor from the top and runs gravity driven through the reactor. The air enters the reactor from the side and transports the humidity and heat into or out of the reactor. In heating mode, the heat released is transferred from the air flow to the water loop by an air-to-water heat exchanger. For material regeneration the air flow direction is reversed and the heat exchanger is used for transferring regeneration heat from the solar loop into the reactor via the air flow. A sketch of the reactor design and a picture of the laboratory prototype are depicted in Fig. 17.18. A detailed description of the reactor design and reactor operation mode is given in [44].
image
Figure 17.18 External cross-flow reactor of the CWS project: (a) pictures of reactor, (b) 3D illustration of the reactor, and (c) sketch of heat and mass fluxes of the reactor.

References

[1] Aydin D, Casey SP, Riffat S. The latest advancements on thermochemical heat storage systems. Renew Sust Energ Rev, 41, 356–67.

[2] Posern K, Kaps C. Calorimetric studies of thermochemical heat storage materials based on mixtures of MgSO4 and MgCl2. Thermochim Acta. 2010;502:7376.

[3] Henninger S, Schmidt F, Henning H-M. Water adsorption characteristics of novel materials for heat transformation applications. Appl Therm Eng. 2010;30:16921702.

[4] Ristić A, Henninger S, Kaučič V. Two-component water sorbents for thermo-chemical energy storage—a role of the porous matrix. Proceedings of Innostock 2012, 12th International Conference on Energy Storage, Llleida, Spain; 2012.

[5] Aristov YI, Tokarev MM, Restuccia G, Cacciola G. Selective water sorbents for multiple applications, 2. CaCl2 confined in micropores of silica gel: sorption properties. React Kinet Catal Lett. 1996;59:335342.

[6] Schmidt P, Bouché M, Linder M, Wörner A. Pilot plant development of high temperature thermochemical heat storage. Proceedings of Innostock 2012, 12th International Conference on Energy Storage, Llleida, Spain; 2012.

[7] Yang RT. Adsorbents: fundamentals and applications. Hoboken, NJ: Wiley-Interscience; 2003.

[8] Ruthven DM. Principles of adsorption and adsorption processes. New York: Wiley; 1984.

[9] Mette B, Kerskes H, Drück H, Müller-Steinhagen H. Experimental and numerical investigations on the water vapour adsorption isotherms and kinetics of binderless zeolite 13X. Int J Heat Mass Transf. 2014;71:555561.

[10] Jänchen J, Stach H. Shaping adsorption properties of nano-porous molecular sieves for solar thermal energy storage and heat pump applications. Sol Energy. 2014;104:1618.

[11] Jänchen J, Ackermann D, Weiler E, Hellwig U. Optimization of thermochemical storage by dealumination of zeolitic storage materials. 10th international conference on thermal energy storage. Pomona, NJ: EcoStock 2006; 2006.

[12] Jähnig D, Wagner W, Isaksson C. Thermo-chemical storage for solar space heating in a single-family house. 10th international conference on thermal energy storage. Pomona, NJ: EcoStock 2006; 2006.

[13] Kerskes H, Asenbeck S, Mette B, Bertsch F, Müller-Steinhagen H. Low temperature chemical heat storage – an investigation of hydration reactions. Effstock 2009, thermal energy storage for efficiency and sustainability.11th international conference on thermal energy storage. Stockholm, Sweden: Energi- och Miljötekniska Föreningen/EMTF Förlag; 2009.

[14] Cuypers R. MERITS: more effective use of renewables including compact seasonal thermal energy storage. Proceedings of Innostock 2012, 12th International Conference on Energy Storage, Llleida, Spain, 2012.

[15] Bales C. Laboratory tests of chemical reactions and prototype sorption storage units. Report of IEA solar heating and cooling programme—Task 32: advanced storage concepts for solar and low energy buildings. Paris, France: International Energy Agency; 2008.

[16] Jänchen J, Schumann K, Thrun E, Brandt A, Unger B, Hellwig U. Preparation, hydrothermal stability and thermal adsorption storage properties of binderless zeolite beads. Int J Low-Carbon Technol. 2012;00:15.

[17] Schmidt FP. Optimizing Adsorbents for Heat Storage Applications. Dissertation, Freiburg im Breisgau, 2004.

[18] Henninger SK, Habib HA, Janiak C. MOFs as adsorbents for low temperature heating and cooling applications. J Am Chem Soc. 2009;131:27762777.

[19] Dawoud B, Aristov Y. Experimental study on the kinetics of water vapour sorption on selective water sorbents, silica gel and alumina under typical operating conditions of sorption heat pumps. Int J Heat Mass Transf. 2003;46:273281.

[20] Bertsch F, Mette B, Asenbeck S, Kerskes H, Müller-Steinhagen H. Low temperature chemical heat storage—an investigation of hydration reactions. Effstock 2009—11th international conference on thermal energy storage, 2009.

[21] Chou I-M, Seal II R. Determination of epsomite–hexahydrite equilibria by the humidity–buffer technique at 0.1 MPa with implications for phase equilibria in the system MgSO4–H2O. Astrobiology. 2003; 3:3, 619–630.

[22] Knovel Database. Thermochemical properties of inorganic chemicals; 2008.

[23] Patnaik P. Handbook of inorganic chemicals. New York: McGraw-Hill Professional; 2002.

[24] Whiting G, Grondin D, Bennici S, Auroux A. Heats of water sorption studies on zeolite–MgSO4 composites as potential thermochemical heat storage materials. Sol Energy Mat Sol Cells. 2013;112:112119.

[25] van Essen VM, Zondag H, Schuitema R, van Helden W, Rindt CCM. Materials for thermochemical storage: characterization of magnesium sulfate, First international congress on heating, cooling and buildings, Lisbon, Portugal; 2008.

[26] Emons H-H, Ziegenbalg G, Naumann R, Paulik F. Thermal decomposition of the magnesium sulfate hydrates under quasi-isothermal and quasi-isobaric conditions. J Therm Anal Calorim. 1990;36:12651279.

[27] Wyttenbach J, Tanguy G, Stephan L. Thermochemical seasonal storage demonstrator for a single-family house: design. Conference Proceedings Eurosun 2014, Aix-les-Bains, France; September 16–19, 2014.

[28] Hongois S, Kuznik F, Stevens P, Roux J-J. Development and characterisation of a new MgSO4–zeolite composite for long-term thermal energy storage. Sol Energy Mat Sol Cells. 2011;95:18311837.

[29] Aristov YI, Gordeeva LG. Salt in a porous matrix adsorbent: design of the phase composition and sorption properties. Kinet Catal+. 2009;50:6572.

[30] Chan KC, Chao CYH, Sze-To GN, Hui KS. Performance predictions for a new zeolite 13X/CaCl2 composite adsorbent for adsorption cooling systems. Int J Heat Mass Transf. 2012;55:32143224.

[31] Aristov YI, Tokarev MM, Gordeeva LG, Snytnikov VN, Parmon VN. New composite sorbents for solar-driven technology of fresh water production from the atmosphere. Sol Energy. 1999;66:165168.

[32] von Beek T, Rindt C, Zondag H. Performance analysis of an atmospheric packed bed thermo-chemical heat storage system. Proceedings of Innostock 2012, 12th international conference on energy storage, Llleida, Spain; 2012.

[33] Zondag H, van Essen V, Bleijendaal LPJ, Kikkert B, Bakker M. Application of MgCl2 H2O for thermochemical seasonal solar heat storage. 5th international renewable energy storage conference (IRES 2010). SEMINARIS CampusHotel Berlin, Science & Conference Center, Berlin, Germany; 2010 November 22–24.

[34] Casey SP, Elvins J, Riffat S, Robinson A. Salt impregnated desiccant matrices for “open” thermochemical energy storage—selection, synthesis and characterisation of candidate materials. Energ Buildings. 2014;84:412425.

[35] Zondag HA, Schuitema R, Bleijendaal LPJ, Gores JC, van Essen M, van Helden W, Bakker M. R&D of thermochemical reactor concepts to enable heat storage of solar energy in residential houses. Paper presented at the 3rd International Conference on Energy Sustainability, July 19–23, 2009, San Francisco, CA. Proceedings of the ASME 3rd International Conference on Energy Sustainability. New York: American Society of Mechanical Engineers; 2009.

[36] Hauer A. Thermochemical energy storage for heating and cooling—first results of a demonstration project. Terrastock 2000, 8th International Conference on Thermal Energy Storage; 2000, p. 641–646.

[37] Eigenberger G, Ruppel W. Catalytic fixed-bed reactors. Ullmann’s encyclopedia of industrial chemistry. Weinheim, Germany: Wiley–VCH Verlag; 2000.

[38] N’Tsoukpoe KE, Liu H, Le Pierrès N, Luo L. A review on long-term sorption solar energy storage. Renew Sust Energy Rev. 2009;13:23852396.

[39] van Helden W, Thür A, Weber R, Furbo S, Gantenbein P, Heinz A, Salg F, Kerskes H, Williamson T, Sörensen H, Isaksen K, Jänchen J. COMTES: parallel development of three compact systems for seasonal solar thermal storage: introduction. Proceedings of Innostock 2012, 12th international conference on energy storage, Llleida, Spain; 2012.

[40] Mehlhorn D, Valiullin R, Kärger J, Schumann K, Brandt A, Unger B. Transport enhancement in binderless zeolite X- and A-type molecular sieves revealed by PFG NMR diffusometry. Micropor Mesopor Mat. 2014;188:126132.

[41] Lass-Seyoum A, Blicker M, Borozdenko D, Friedrich T, Langhof T. Transfer of laboratory results on closed sorption thermo-chemical energy storage to a large-scale technical system. Energy Procedia. 2012;30:310320.

[42] Zondag H, Kikkert B, Smeding S, de Boer R, Bakker M. Prototype thermochemical heat storage with open reactor system. Proceedings of Innostock 2012, 12th international conference on energy storage, Llleida, Spain; 2012.

[43] Kerskes H, Mette B, Bertsch F, Asenbeck S, Drück H, Chemical energy storage using reversible solid/gas-reactions (CWS)—results of the research project, 1st International Conference on Solar Heating and Coolingfor Buildings and Industry (SHC 2012) 2012, 30, 294–304.

[44] Mette B, Kerskes H, Drück H, Müller-Steinhagen H. New highly efficient regeneration process for thermochemical energy storage. Appl Energy. 2016;109:352359.

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset