The overall modeling approach is based on a first principle nonlinear state-space representation based on a set of differential-algebraic equations (DAEs) implemented in Matlab Simulink. The objective is to achieve the highest possible accuracy with the lowest possible computational effort. From a computational perspective, the simplest representation consists of lumped zero-dimensional models of the individual system components. These are based on energy and mass balance DAEs with additional quasi-static formulations regarding heat transfer or electrochemical reactions.
3.2. Balance of Plant (BoP) Component Models
BoP component modeling in the literature mainly varies in terms of the degree of component discretization, consideration of pressure drops, heat losses, and gas delays, as well as in the approach used to determine the amount of recirculated off-gas.
Engelbracht [
23] chose to discretize all major BoP components, such as the allothermal pre-reformer and afterburner, as a cascade of CSTRs similar to his approach with SOFCs, while Stiller only [
14] discretized the heat exchangers in order to capture their dynamic thermal characteristics. Other authors, like Carré [
12], Braun [
21], and Kupecki [
18], used a lumped approach for all BoP components. A distinct feature of Stiller’s model is the consideration of gas transport delays in between the lumped model components to account for the gas residence times. This is realized by adiabatic pipes with time delays determined by component volume and the ideal gas law. Heat losses apart from the SOFC stack are neglected in all models, except for those by Carré, either by simplification or the assumption of a high level of module integration. Pressure drops of the gas flows are considered to estimate the required electric power from the blowers affecting the system’s net efficiency. However, the pressure drop calculation in the anode loop is typically not utilized to determine the share of recirculated gas, as the recirculation ratio is taken as a plant input variable. On the contrary, Kupecki implemented the performance map of the recirculation blower in order to determine the required electric rotational blower frequency
as a function of the anode loop pressure drop and the recirculating mass flow
for stationary system simulation.
This publication aims to merge the above-mentioned model approaches to adequately depict the dynamic behavior of the targeted plant model. BoP components are implemented as lumped zero-dimensional models in order to decrease computational effort. This simplification is supported by the previous experimental characterization of the heat exchangers, a grid independence study on heat transfer, and the degree of pre-reforming in the allothermal pre-reformer, as well as the implementation of corresponding heat transfer correlations [
7]. Heat losses are considered in all BoP components as they have been determined in detail by means of a steady-state thermal analysis, considering the individual component insulation.
The spatial separation of the fuel cell stack modules from the central FPM represents a novelty compared to published system designs. This new design, dictated by ship construction boundary conditions, inevitably leads to higher gas volumes in the anode gas loop heavily affecting the transient system behavior. Therefore, this publication places considerable emphasis on the correct dynamic representation of the gas residence times following the approach of Stiller.
In order to achieve a realistic dynamic depiction of the AOG mass flow distribution, the operating characteristics of the recirculation blower need to be implemented according to the approach of Kupecki, as the blower represents the flow-determining component. Due to the closed loop structure, the overall component pressure drops in the loop are imposed on the blower to be overcome, which, together with a given rotational speed, results in the mass flow conveyed by the blower. This in turn defines the amount of recirculated AOG, thus the recirculation ratio
. Implementation of the blower characteristics into the plant model is also beneficial with regard to system control to estimate the state of recirculation without directly measuring the hot mass flow in the anode loop, as such sensors are considered cost-intensive and prone to failure [
16,
28].
In contrast, setting the inlet air mass flows at ambient temperature to their desired set points can be considered as a straightforward subordinate and independent control task, as mass flow meters in the immediate vicinity of the blowers are implemented to set the blower frequencies accordingly. It is assumed that the mass flows are set to their respective set point without significant control delay at all times so that these blower characteristics do not have to be considered and the blowers are only modeled in terms of their power demand and the rise in air temperature. Analogously, ideal control of the gaseous fuel mass flow is assumed with the aid of a conventional mass flow controller.
3.3. Governing Equations
Each component depicted inside the system boundaries in
Figure 2 is modeled using a set of the following equations stated with the associating assumptions. An overview is given in
Table 1. The fuel flow, treated as pure
, and the dry air flows enter the system at ambient conditions (
,
). Co-generation of useful heat in downstream exhaust gas heat exchangers is not considered in this study.
Conservation of mass, including chemical reactions: For each species,
i, present in the respective flow, a DAE is formulated using the ideal gas law rearranged to the molar fraction
, which is equal to the outlet molar fraction
in a CSTR [
12]:
with
and
being the inlet molar fraction and overall molar flow, considering a total of
k reactions occurring simultaneously with the corresponding stoichiometric coefficients
and reaction rates
. In accordance with models in the literature, this formulation implies the assumption of a quasi-static total outlet molar flow. This assumption is permissible due to the high prevailing temperatures and significant thermal inertia of the components; see [
12,
24].
At the SOFC cathode, molecular oxygen is reduced to oxygen ions (ORRs), which mitigate through the electrolyte and lead to the oxidation of hydrogen (HOR) at the anode, resulting in the net electrochemical reaction and an associated electron flow:
The ORR and HOR reaction rates of the associated species are, thus, related to the local current density determined by the
submodule. The MSR and WGS reactions occur at the SOFC anode and in the pre-reforming reactors:
Reaction rate modeling differs with respect to the associated component assumptions. The WGS reaction is assumed to be in equilibrium, both in the SOFC control volumes and lumped pre-reformers. Formulating the sum of the forward and backward reaction rates and inserting the equilibrium constant
yields [
29]:
with
being the backward reaction rate constant and
the CSTR partial pressure of species
i. The equilibrium constant can be calculated by means of the standard molar Gibbs enthalpy of the WGS reaction:
Regarding the MSR reaction rate in the pre-reformers, chemical equilibrium is assumed due to the high activity of the used precious metal catalyst and the high gas residence times. Thus, a similar expression for the MSR equilibrium reaction rate is formulated:
The values of the backward reaction rate constants
and
are only of numerical significance and should be chosen large enough for the correct calculation of equilibrium composition [
29]. However, too large values lead to an increased computational effort. Thus, a sensitivity study of the backward reaction rate constants was performed for each equilibrium reaction in both the pre-reformers and the stack. A trade-off was achieved by setting the backward reaction rate constants for each reactor to the values depicted in
Table 2, which leads to a maximum absolute molar composition error of 0.2 mol-% in the investigated operating range.
In contrast, the area-specific kinetic Langmuir–Hinshelwood approach from van Biert et al. [
24] is used for each discretized anode control volume:
The values of the pre-exponential factor
, the activation energy
, as well as the adsorption constants
and
are directly adopted from [
27], whereas
is calculated analogously to Equation (
10). At the oxidation unit, a total conversion of the remaining combustible AOG species is assumed so that the reaction rates are determined by the respective molar flows:
Conservation of Energy, including chemical reactions: The general energy balance DAE is shown in Equation (
16), considering enthalpy changes of fluids entering and leaving as ideal gases, as well as reaction enthalpies, heat flows
, and electrical or mechanical power
:
where
is the lumped temperature according to the CSTR assumption,
m is the mass of the component under consideration and
its specific heat capacity.
represents the temperature-dependent molar isobaric heat capacity of the corresponding ideal gas mixture and
is the molar reaction enthalpy of reaction
k at reference temperature
.
According to the lumped model assumption, BoP components comprise one energy DEA per fluid stream, as depicted in
Table 1. Thus, it is assumed that the fluids leave at the solid component temperature, which couples the thermal gas behavior with the dominating thermal inertia of the solid. For components with two fluid streams, the masses are distributed among the two balance equations with respect to the fluid’s degree of component interaction. For plate-fin heat exchangers, the mass is distributed uniformly to the DAEs, whereas for shell-and-tube heat exchangers, including the allothermal pre-reformer, both the mass of the shell and half of the mass of the tube are assigned to the hot shell flow due to the larger heat transfer surface.
Regarding the discretized stack, the four energy DAEs per control volume are formulated for the gas temperatures
and
, the solid temperatures of the overall cell
, and the interconnect
identical to [
24]. Heat is generated through reaction in the cell, and heat conduction between two neighboring spatial solid control volumes occurs in accordance with Fourier’s law, governed by the solid thermal conductivity of the cell and the interconnect. Four convective heat flows occur per control volume between the solid–fluid pairs
, and are expressed in the following form:
as the local heat transfer coefficient, depending on the Nusselt number
, the thermal conductivity of the respective fluid
, and the hydraulic channel diameter
. In alignment with the authors’ concerns regarding stack modeling mentioned above, a constant value of
is chosen for the Nusselt number due to a fully developed rectangular laminar channel flow.
Heat transfer in the BoP heat exchangers is modeled using the NTU method by calculating the maximum possible heat flow and scaling it with the heat exchange effectiveness
[
30]:
Apart from the flow configuration and the fluid’s heat capacity rates, the effectiveness depends on the overall heat transfer coefficient
U, determined by the convective heat transfer on the cold and hot sides
, as well as the thermal wall conduction resistance:
Nusselt correlations for the implemented heat exchangers in the form
were determined experimentally and are documented in [
7]. Heat losses in all BoP components, as well as the outer interconnect DAEs of the stack model, are implemented analogously to Equation (
17):
Heat loss transfer coefficients are parameterized according to the thermal simulation results of the individual components.
Electrochemical Modeling:
For each active cell control volume identified by the spatial index
z, the local voltage losses and electric current are determined with the quasi-static voltage breakdown:
using the local Nernst potential, depending on temperature and the molar fractions of reactants:
The overall ASR is typically formulated as an Arrhenius type to address the dominating temperature-dependent ionic conductivity in the electrolyte [
23]. The equation is expanded by a linear degradation rate
given in mΩcm
2 kh
−1 provided by the cell manufacturer [
31] to account for degraded cell operation after a specific operating time
in hours:
The reference ASR at the reference temperature
and begin of life, the activation energy
, as well as the offset contact resistance
were determined experimentally by evaluation of polarization curves from a single cell in a high-temperature test bench and are given in
Table 2. The test setup and the experimental operating procedure are explained in detail in [
32]. Coupling the electrochemical equations of each active control volume is realized by means of the following algebraic constraints to set the requested cell current
I and the isopotential boundary condition:
Finally, the overall electrical stack power is determined by linearly extrapolating the single-cell behavior:
Gas residence time:
Following the pipe modeling of Stiller, the time delay associated with the respective component and pipe volumes are calculated by means of the ideal gas law:
considering the component volume
filled with gas at temperature
and pressure
at a molar flow rate
. To avoid an unnecessary number of consecutive gas delays, volumes of adjacent pipes and components are merged into one without sacrificing modeling details.
Laminar pressure drop:
As a rough approximation, laminar pipe flow is assumed in all components enabling the use of the Hagen–Poiseuille equation, depending on gas density
, flow velocity
w, geometry, and the Reynolds number
. By inserting the definition of the Reynolds number, a linear dependence on the mass flow
is obtained:
depending on the dynamic viscosity
, density
, and the pressure loss coefficient
. Values for
are assigned based on the individual component design nominal pressure drops, which were chosen to reach the maximum anode overpressure of 50
at maximum electrical current and begin of life to avoid unnecessary high BoP power demand.
The thermodynamic behavior of all implemented blowers is based on an ideal gas isentropic pressure change from
and
to the desired outlet pressure
complemented by the isentropic efficiency
yielding the blower outlet temperature:
Electrical power demand is modeled by using the corresponding enthalpy difference and considering a mechanical blower efficiency
accounting for friction and electric motor irreversibilities:
Regarding the recirculation blower, the operating map of the implemented side channel blower is utilized, which relates the mass flow with the imposed pressure drop of the anode loop and a specified electrical blower rotation frequency
:
The recirculation ratio is, thus, determined by the blower characteristics and the molar fuel flow:
Finally, the system electrical net power is calculated by subtracting the BoP blower power demand (Equation (
29)) from the SOFC power output (Equation (
25)):
considering a generalized efficiency of the SOFC power electronics
. The system’s net efficiency rates the net power with the lower heating value of supplied fuel: