The three components, i.e., adsorber, condenser and evaporator, are not thermally connected, but are only coupled by the vapor mass low
between the components, which is determined by the adsorption/desorption rate inside the adsorber. The adsorption/desorption rate in turn depends on the vapor pressure conditions in the condenser and evaporator, respectively, which again are affected by the vapor mass flow. Hence, the coupling is bidirectional. During the regeneration phase, the pressure at the outlet of the adsorber is set by the condenser pressure, as vapor flows from the adsorber into the condenser. During the cooling phase, the adsorber inlet pressure is defined by the evaporator pressure. Since the components are only coupled through their boundary conditions, the components can be modeled separately. All three components are panel-shaped and sufficient mechanical stiffness is achieved through internal structures, such as fins or pins known from evacuated flat plate solar thermal collectors [
35].
2.1. Adsorber
The design scheme and model domain of the adsorber is depicted in the cross-section side view in
Figure 2. The panel-shaped adsorber is enclosed by a metal casing at the back and sides as well as by a metal absorber sheet at the outer face. The adsorber is attached vertically to a southeast facade of the building. The adsorber is filled with a granular adsorbent, while a wire mesh separates the adsorbent from the vacuum gap. Due to this gap, the backside of the adsorber is thermally well insulated towards the building’s envelope. This insulation is further secured by additional vacuum insulation panels on the back of the casing. The applied adsorption pair is granular zeolite 13X and water. Water vapor can flow in and out of the adsorber through a central in-/outlet at the back of the adsorber, assuming an instant and homogenous distribution of the vapor inside the vacuum gap. To overcome the heat transport limitations of the zeolite packed bed, metal fins are integrated. In order to efficiently absorb the solar irradiation during the regeneration phase, the outer adsorber face is covered by a metal absorber sheet as well as a vacuum glazing. Between the absorber sheet and the vacuum glazing, there is an air channel, which is closed by a flap during the regeneration phase, while it is open during the cooling phase to efficiently transfer the released heat of adsorption to the ambient by free air convection through the air gap channel. To enhance the heat release in the cooling phase, the absorber sheet is equipped with vertical metal fins in the air gap to increase the heat transfer surface (not depicted in
Figure 2). Since geometry as well as boundary conditions are symmetrical and under the assumption that temperature variations over the absorber sheet are negligibly small, the examination can be reduced to the marked model domain. The geometry parameters of the reduced problem are noted in
Figure 2.
The adsorber model is based on a model for the adsorbent bulk, previously developed and published by co-author Schaefer [
32,
33]. Here, the existing model is extended by the internal metal fins in the adsorbent bulk and the absorber sheet, including the absorber fins. The assumptions of the underlying model are discussed in detail in the aforementioned publication, and thus, they are only mentioned in brief here. The main assumptions, referring to the processes in the adsorbent bulk, are as follows:
The vapor is an ideal gas and viscous Newton-fluid;
Local thermal equilibrium is assumed;
Due to the low pressure, rarefaction effects in the mass and heat transport in the adsorbent bulk are considered;
The adsorption kinetics can be described by the linear-driving-force approximation;
The heat of adsorption depends on the water uptake;
The temperature dependencies of the material parameters (viscosity, heat capacity, density, etc.) are accounted for.
Following these assumptions, the balance equations for vapor mass and adsorbent energy as well as the mass balance for the adsorbed water are derived, yielding:
where
,
and
describe the local pressure, temperature and water uptake in the zeolite adsorbent bulk, the fluxes
and
denote the local vapor mass und heat flux, respectively,
refers to the porosity of the adsorbent bulk,
refers to the specific gas constant of the vapor,
refers to the bulk density,
,
and
refer to the specific heat capacities of the dry zeolite adsorbent, the adsorbate and the vapor, respectively,
refers to the heat of adsorption,
refers to the adsorption kinetics parameter and
refers to the water uptake at equilibrium. Using the density of the dry zeolite adsorbent
, the bulk density is calculated with
. In [
33], the values and equations of the parameters and material functions are given in detail.
The vapor mass flux
and the heat flux
must be specified to close the system of balance equations. Darcy’s law is locally applied for the vapor mass flux, which reads as follows:
where
denotes the permeability tensor,
denotes the density and
denotes the dynamic viscosity of the vapor.
To calculate the heat flux in the adsorbent bulk, Fourier’s law is applied:
where
is the effective heat conductivity of the adsorbent bulk. For both the permeability and heat conductivity, so-called Knudsen correction factors are applied. These factors account for the rarefaction effects due to the low vapor pressure and are described by Schaefer in [
32,
33].
The model presented so far solely describes the processes in the adsorbent bulk. Therefore, the model has to be extended to include the heat transport in the metal fins and absorber sheet. As the applied extension is already presented by Schaefer et al. for another application in [
34], only a short summary is given here. For the metal fins integrated into the adsorbent bulk, the one-dimensional transient heat equation [
36] is applied, which reads as follows:
where
refers to the local fin temperature, averaged over the fin cross-section and
,
,
and
denote the density, specific heat capacity, heat conductivity and the thickness of the metal fin, respectively. The heat exchange between metal fin and adsorbent bulk is given by the term
. Approximating the local temperature distribution over the cross-section of the fin by a parabolic function, this term yields:
where
is the contact temperature between fin and zeolite.
Similar to the metal fins, the metallic absorber sheet is modeled by the two-dimensional transient heat equation [
36]:
where
is the local temperature in the adsorber sheet and
,
and
denote the density, specific heat capacity and heat conductivity of the adsorber sheet, respectively. The thermal coupling of the sheet and the zeolite as well as of the internal metal fins and the zeolite is achieved by equating the heat fluxes at the surface, where the heat flux in the zeolite at the contact surfaces is calculated according to Fourier’s law from the local temperature gradient in the zeolite at the contact boundary.
Finally, the adsorber model has to be extended in order to incorporate the air gap channel, which can be opened or closed, as well as the vacuum glazing. This is achieved by applying the heat flux
, which depends on the operational phases, to the outer face of the absorber sheet. During the regeneration phase, the vacuum glazing and closed air gap channel act as a solar thermal collector. The efficiency
of such a collector is calculated by:
where
refers to the optical efficiency, which is the maximum efficiency of the solar collector,
to the linear loss factor,
to the quadratic loss factor and
to the solar irradiation ([
37], p. 356). The temperatures
and
denote the temperature at the surface of the absorber sheet and the ambient temperature, respectively. Thus, the heat flux
to the outer face of the absorber sheet during the regeneration phase reads as follows:
During the cooling phase, the air gap channel is opened by the flaps to cool the absorber sheet by free convection. The optional external fins on the absorber sheet are not spatially resolved, but are considered applying the geometry factor for the increased surface of the absorber sheet and for the fin efficiency, which takes into account the non-uniform temperature distribution along the fin. Furthermore, the additional heat capacity added by the mass of the external fins is incorporated by increasing the effective heat capacity of the metal in the adsorber.
The heat flux
, applied to the outer face of the absorber sheet during the cooling phase, is the heat flux released to the ambient by free convection and radiative emission reduced by the received solar irradiation. This yields:
where
refers to the emissivity factor of the adsorber sheet,
refers to the Stefan–Boltzmann constant and
refers to the convective heat transfer coefficient. The coefficient
is defined as:
and is introduced to pseudo-linearize the heat flux Equation (
9).
The air gap channel is divided into several small channels formed by the absorber sheet, the external fins and the vacuum glazing. It is assumed that the convective heat transfer coefficient
between the metallic structures and the air flow inside the channels can be determined from correlations for free convection in a vertical heated tube from ([
38], p. 703), applying the hydraulic diameter
of the rectangular channel between the fins. It is distinguished between laminar and turbulent boundary layers depending on the Grashof number
. For laminar boundary layers (
), the heat transfer coefficient is as follows:
where
denotes the heat conductivity of the air inside the channels. For turbulent boundary layers (
), the heat transfer coefficient is as follows:
2.2. Condenser
The design scheme and model domain of the condenser is depicted in
Figure 3. Similar to the adsorber, the condenser is enclosed by a metal casing and is attached vertically to the northwest facade of the building. Internal metal fins are included to enhance heat transport to the ambient, and thus, improve the condensation rate. The backside of the condenser is thermally insulated, whereby the insulation is assumed ideal. In the regeneration phase, vapor flows from the adsorber into the condenser and condenses on the metal fins or the water surface, releasing the heat of condensation to the fins or the water, respectively. The heat is then conducted through the internal fins to the external fins, where it is emitted to the ambient air. Furthermore, the water phase is in direct contact to the outer wall and heat is being transferred to the ambient through the outer wall. Due to symmetry of geometry and boundary conditions, the model domain can be reduced to the marked area (red dashed box).
The developed model of the condenser describes the condensation and the heat transport processes in the metal structures under the following physical assumptions:
The energy and mass of the vapor are neglectable;
The vapor is an ideal gas;
The vapor is always saturated;
Incoming vapor fully condenses;
The water is incompressible;
Temperature variations in the condensed water are negligibly small, but natural convection is indirectly captured through correlations for the heat transfer;
Temperature variations in height direction of the walls are negligibly small;
Boundary effects of the top and bottom of the casing are neglected;
The temperature dependencies of the water (density, heat capacity, enthalpy of evaporation, etc.) are accounted for.
As it is assumed that incoming vapor fully condenses, the mass change of the condensed water
is equal to modulus of the incoming vapor mass flow
. The mass change is the sum of condensation on the metallic structures, where it is assumed that solely laminar film condensation
takes place, and the condensation at the free water surface
. Thus, the mass balance reads as follows:
Note that
is negative during desorption according to Equation (
26). The mass flow
is calculated from the Nußelt theory of laminar film condensation [
36,
39] with the following physical assumptions:
The condensate film flow is always laminar and steady;
The surface temperatures of fins and walls are isotherm;
The vapor bulk is stagnant.
In case the mass flow exceeds the incoming vapor flow , the condensation at the water surface becomes negative, which means that evaporation takes place to maintain the state of saturation in the vapor phase. It is assumed that a water circulation inside the condenser occurs in this case, with water evaporating at the water surface and condensing again on the metallic fins and casing.
Due to the assumption of a homogenous water temperature, no spatial discretization of the water is required. Thus, the energy balance of the water yields:
where
denotes the heat fluxes to the internal fins and casing and
the enthalpy of evaporation. The term
accounts for the sensible heat of the vapor, which is coming in from the adsorber and is generally cooled to the temperature of the condenser. It is assumed that this sensible heat of the inflowing vapor is first transferred to the water phase before the vapor condenses.
Furthermore, the heat transport in the metallic fins and walls has to be modeled. This is achieved in a similar way as for the internal fins of the adsorber, applying the one-dimensional transient heat equation. For the fins, this is as follows:
where
is the the local temperature, averaged over the fin cross-section, and
,
,
and
are the density, specific heat capacity, heat conductivity and the thickness of the metal fin, respectively. The heat transfer from the water or vapor to the internal fins as well as from the external fins to the ambient air is
. This coupling term is derived by approximating the local temperature distribution over the cross-section of the fin, applying a parabolic function. Determining the temperature gradient from the parabolic temperature profile at the fin surface and evaluating Fourier’s law, the coupling heat flux term reads as follows:
where
denotes the contact temperature between the fin and the water, vapor or ambient air. The contact temperatures are determined by equating the heat fluxes at the internal boundaries.
The heat transport inside the casing is modeled with the same approach, but the local temperature distribution over the cross-section of the casing is approximated by a piecewise linear function instead of a parabolic function. Moreover, the fins and the casing are thermally coupled by equating the heat fluxes at the contact surface. For better readability, the heat transfer equations are only given for the fins in the following. Yet, the given equations are also valid for the casing.
The heat transfer between the metallic structures and the water or vapor phase, respectively, is modeled with:
where
is the heat transfer coefficient and
is the contact temperature for the respective contact surface. In the case of contact with the vapor phase, the Nußelt theory of laminar film condensation [
39] is applied to model the heat transfer and the heat transfer coefficient
, averaged over the whole fin area, is calculated from:
where
refers to the heat conductivity of liquid water,
g refers to the gravitational acceleration,
refers to the density of liquid water,
refers to the density of vapor,
refers to the temperature of the water in the condenser, which equals the saturation temperature, and
refers to the contact temperature of vapor and fin. The variable
is the kinematic viscosity of the water and
is the height of the vapor phase.
In the case of contact with the water phase, the heat transfer is modeled, assuming free convection inside the water phase and the heat transfer coefficient is estimated with an empirical equation given in the VDI Heat Atlas ([
38], p. 675):
where
denotes the Prandtl number,
denotes the Rayleigh number,
denotes the height of the water phase inside the condenser and
denotes the distance between the fins.
Finally, the heat transfer from the external fins and casing to the ambient air has to be modeled. Here, both convective heat transfer and thermal radiation are considered:
The convective heat transfer coefficient
is approximated conservatively by a constant value of
, which is an averaged value for the expected temperature ranges and geometry parameters applying the procedure for free convection of external flows at vertical surfaces described in the VDI Heat Atlas ([
38], p. 667). As the external heat transfer is very sensitive to the fluctuating wind conditions, it is not necessary to apply a precise correlation function, which is only valid for steady conditions. Moreover, this might give a wrong impression on the model accuracy of the aspect. The radiative emissivity
is set to
. Note that the second term in Equation (
20), which accounts for thermal radiation, is not applied to the external fin areas, which are opposite of each other, assuming that they mainly radiate on each other rather than into the ambient. Only the radiation from the base surface of the condenser is taken into the account.
2.3. Evaporator
The design scheme and model domain of the evaporator is shown in
Figure 4. The evaporator is also enclosed by a metal casing and is installed inside the building as a cooling ceiling. The applied box-shaped configuration is a very simple configuration, chosen in this work to determine the potential of the system with respect to the cooling rate. Nevertheless, more complex solutions such as pipe coils or active ventilation are possible and could be investigated with slight modifications to the evaporator model presented in the following. Vapor is extracted from the evaporator by the adsorption in the adsorber during the cooling phase, which reduces the pressure in the evaporator. Thus, evaporation is induced and the remaining water is cooled due to extracted heat of evaporation. The water then cools the metallic bottom sheet of the evaporator, which is in direct contact with the air inside the room. The model domain can be limited to the depicted section (red dashed box) as the horizontal dependency of the surface temperature of the bottom metal casing can be neglected.
For the water, the mass and energy balance equations are evaluated, applying the following assumptions:
Energy and mass of the vapor phase are neglectable;
The vapor is an ideal gas;
The water is incompressible;
Vapor pressure in the evaporator is assumed to be equal to saturation pressure;
Outflowing vapor instantly evaporates;
The temperature dependencies of the water (density, heat capacity, enthalpy of evaporation, etc.) are accounted for.
As the mass of the vapor phase is neglected, the change of water mass inside the evaporator
is determined by the vapor mass flux to the adsorber:
It is assumed that the natural convection in the water phase can be approximated by Rayleigh–Benard convection. The energy balance for the water phase in the evaporator yields:
where
refers to the contact area between water phase and bottom sheet. The contact temperature
is calculated by evaluating the heat fluxes at the surface between water and the metal bottom sheet. The heat transfer coefficient
between the bottom sheet and the water phase is determined by applying empirical correlations given in the VDI Heat Atlas ([
38], p. 674), where it is distinguished between laminar and turbulent boundary layers depending on the Rayleigh number
. For laminar boundary layers (
), the heat transfer coefficient is as follows:
where
denotes the height of the water phase. For turbulent boundary layers (
), the heat transfer coefficient is as follows:
In order to calculate the temperature change inside the metallic bottom of the evaporator, the one-dimensional heat conduction in x-direction is modeled. This yields:
where
,
and
are the density, specific heat capacity and thickness of the metallic bottom sheet, respectively. Again, the contact temperature
is determined by equating the heat fluxes at the surface between the bottom sheet and the air inside the room. The temperature of the room
is set to a constant value of
. The heat transfer coefficient
between the bottom sheet and the air is assumed to a constant value of
, which is a conservative estimation and accounts for both convective heat exchange and radiation. No correlation is applied as
depends on many parameters, such as ventilation of the room. Here, in contrast to the adsorber and condenser model, heat radiation is not directly modeled, since the surface temperatures inside the room are unknown, but is included indirectly in the heat transfer coefficient.
2.4. Coupling and Boundary Conditions
The three components are connected by pipes and the distinct component models are coupled through the mass flows of vapor or water. The vapor flow
between adsorber and condenser or evaporator, respectively, is determined by the ad-/desorption rate inside the adsorber:
where
denotes the mass of the dry zeolite in the adsorber and
denotes the spatially averaged water uptake of the adsorber. During desorption, the water uptake decreases and
becomes negative. The vapor flows out of the adsorber and into the condenser. During adsorption, the water uptake increases and
becomes positive, accordingly. The vapor flows out of the evaporator and into the adsorber. Furthermore, the adsorber model is coupled to the other component models through the pressure. The outlet pressure at the adsorber is set to the pressure of the condenser in the regeneration phase and the inlet pressure of the adsorber is set to the pressure of the evaporator in the cooling phase.
As the cooling demand increases with ambient temperatures and solar irradiation, a representative day in summer for a site in Stuttgart, Germany, where the proposed adsorption system will be realized within the CRC 1244 [
4], is chosen. The 26th of August in 2016 is identified as such a day and is defined as the reference day for the simulation study. The respective hourly values for ambient temperature and solar radiation are taken from the database PVGIS-SARAH of the Photovoltaic Geographical Information System [
40,
41]. The azimuth angle of the vertical adsorber facade is
(south-east). In
Figure 5, the time curves of the applied ambient conditions are given. For the solar collector efficiency
, an evacuated flat plate collector [
42] and a flat plate collector [
43] are chosen. In addition, the average between the two is calculated and used as third case. The coefficients for efficiency Equation (
7) are given in
Table 2. Furthermore, the temperature
of the room to be cooled is assumed to a constant value of
. This allows for a more general examination of the proposed adsorption cooling system, independent of specific room and building conditions.
2.5. Operating Phases and Switching Criteria
The adsorption cooling system runs through four operational phase, which differ regarding the opening or closing of the connecting pipes and the air gap channel of the adsorber, see
Figure 2. The operational phases and the corresponding switching criteria are depicted in
Figure 6. The switching criteria depend on the solar irradiation and the pressures in the three components.
1. Regeneration phase: In this phase, the water uptake of the adsorber is reduced. This is realized by heating up the adsorber with solar irradiation and opening the connection between adsorber and condenser, so that the released vapor can flow to the condenser and condense. The air channel is closed in this case to facilitate the heat-up of the adsorber. The outlet pressure of the adsorber corresponds to the saturation pressure in the condenser, which strongly effects the desorption rate. Low pressures are beneficial as the water uptake at equilibrium decreases with decreasing pressure. Therefore, an efficient cooling of the condenser is required.
The regeneration phase takes place while the solar radiation shines on the adsorber and lasts until decreases below due to the course of the sun. For lower values of , the rate of desorption becomes small as the adsorber is almost not heated up further. Furthermore, the longer the regeneration phase continues, the later the cooling phase starts. Thus, the end of the regeneration phase strongly depends on the compass orientation of the facade. Alternatively, the facade can be shaded actively. The second switching criteria regarding the negative slope of the solar irradiation is necessary to avoid switching to the cooling phase in the early morning hours when is still below . Furthermore, switching should only take place if the adsorber can be effectively cooled by opening the air channel. This is ensured by applying the third switching criteria, which evaluates if the emitted heat flux with an opened air channel exceeds the incoming solar irradiation, reduced by the optical losses of the solar collector.
2. Inter-cooling phase: At the end of the regeneration phase the pressure inside the adsorber, which is almost equal to that of the condenser, is above the pressure inside the evaporator due to a higher temperature of the condenser. Thus, immediately connecting the adsorber with the evaporator would lead to a vapor flow from the adsorber into the evaporator and heat up the evaporator, which is contrary to the intention during the cooling phase. Therefore, it is necessary to reduce the pressure inside the adsorber first by closing all connecting pipes and opening the air channel. Then, the adsorber cools down and the remaining vapor inside the adsorber is re-adsorbed, which reduces the pressure. Once the pressure is equal to or lower than the pressure of the evaporator, the system can switch to the cooling phase.
3. Cooling phase: In this phase, the room is cooled by evaporating water in the evaporator. Therefore, the connection between evaporator and adsorber is opened, so that the vapor can be adsorbed, which leads to continous evaporation. This also means that the inlet pressure at the adsorber is determined by the pressure inside the evaporator. However, a control valve, which allows for the reduction of the adsorber inlet pressure, is considered to control the temperature, and thus, the cooling power of the evaporator. The details of the control are explained in the next
Section 2.6. Again, the vapor flow is determined from the adsorption rate in the adsorber, refer to Equation (
26). This rate can be increased by improving the cooling of the adsorber, as the water uptake at equilibrium increases with decreasing temperatures. Thus, the air channel is opened. The cooling phase is active until the solar radiation increases again in the early morning hours. The second switching criteria regarding the positive slope of the solar irradiation is necessary to avoid switching to the regeneration phase in the afternoon hours when
is below
, but still exceeds
.
4. Inter-heating phase: At the end of the cooling phase, the pressure inside the adsorber is below that of the condenser, which would lead to vapor flowing from the condenser to the adsorber if the regeneration phase would start immediately after the cooling phase. Therefore, the pressure inside the adsorber must first increase, which is achieved by closing all valves in the connecting pipes and also the air channel. Through the solar irradiation the adsorber is heated up and a relatively small amount of water is desorbed, which increases the pressure. Once the pressure is equal to or higher than the pressure of the condenser, the system can switch to the regeneration phase.
Note that the adsorber can only be connected to either the condenser or the evaporator at the same time, as adsorption and desorption cannot take place at the same time. To close the internal water cycle, liquid water is pumped from the condenser into the evaporator in the early morning hours to compensate for the water, which has been transferred from the evaporator to the condenser via the adsorber. This pumping procedure is considered to require only a short time compared to the described phases and is conducted at the end of the inter-heating phase.