3.1. Thermal Model
In order to simulate the transient thermal behaviour of the heat transfer, through the thickness of the Ablative Thermal Protection Systems (ATPSs) and combustion chamber, the FDM solution technique was chosen. The choice of this technique was based on its generality, versatility with different materials and geometries, and due to the background support that exists, in the form of previously developed work, such as those performed by Barato et al. [
4], Shi et al. [
10], Li et al. [
6] and Natali et al. [
5].
Some assumptions are made to simplify the model in the study. Only heat transfer in the thickness direction is considered; therefore, a one-dimensional FDM model is used. It is assumed that the ATPS is a material that has constant physical properties in the thickness direction; therefore, any anisotropic material, such as composite materials, can be analyzed, as long as their properties are uniform in the thickness direction (transverse isotropy) [
11], which is common for tubes made of fiber-reinforced matrices as seen on many ATPS materials. The thermal expansion of the materials is not modeled and is assumed to be negligible. Additionally, the thermal analysis of the ATPS assumes the worst-case scenarios, where the ATPS is exposed to the engine combustion during the entire burn time and is observed in the post-combustion chamber of a hybrid rocket engine.
Equation (
3), which is the transient heat equation, can be simplified according to the model at use, into
where
r denotes the radial or thickness direction and
h the specific enthalpy of the material. This is the energy equation that defines the heat transfer inside the material. When considering an ATPS charring material that undergoes phase change during pyrolysis, in which the virgin material changes into char material, the energy involved in this process is taken into consideration through two additional terms in the heat equation [
5,
6,
8,
10,
12]
The second term on the right side of the equation defines the energy involved in the decomposition of the material, being
Q the heat of the decomposition. The third term on the right side of Equation (
10) quantifies the energy absorbed by the gases released during pyrolysis (transpiration cooling), where
is the generated gas mass flow and can be calculated as
Equation (
10) can be extended to its final form by expanding the left side of the equation and the first term on the right side of the equation, giving the following energy equation for the ATPS material:
where
is the enthalpy of the gases released during pyrolysis and
is the enthalpy of the solid material. Both quantities can be defined, respectively, as [
5,
6,
8]
where
c and
are the specific heat capacities of the solid material and the released gases, respectively.
During the pyrolysis process, where the phase change can be observed, the virgin material is transformed into char material at a rate defined by the time derivative of
. To determine this rate, the Arrhenius degradation equation is used [
4,
5,
6,
8,
10], which states that the degradation rate of the material due to thermal phase change can be expressed as an equation of the
nth order of the form
where
A is the Arrhenius preexponential factor,
is the activation energy,
and
are the density of virgin material and charred material, respectively, and
R is the universal gas constant. The variables in the Arrhenius equation depend on the type of material that is being degraded, and it is the most common way to study the phase change in charring materials that undergo pyrolysis [
5,
6,
8,
13]. As can be seen, the fraction to the power of
n seen in the equation indicates the state of change that a portion of the material is in, taking the value of one when the material is still in the virgin state and taking a value of zero when the material is completely charred. This factor is therefore used in the mixture rule [
6] to determine the thermal characteristics of an ATPS during decomposition:
where
,
,
and
are the thermal conductivities and specific heat capacities of virgin and charred materials, respectively. To be precise, these are the thermal properties of the material both in its original state and after a complete thermal degradation, when the pyrolysis has ended and only charred residues remain.
Apply now the discretization methods mentioned in the FDM technique, in the form of Equations (
4)–(
6), and the approximated, discretized, second-order differential equation of energy becomes
Note that there are two FDM techniques, the explicit and implicit methods [
7]. The difference between the methods consists of whether the temperature in node
m at time
is obtained through the previous temperature value
(explicit method), or through the values of neighboring nodes
and
(implicit method). Although the explicit method can be easily solved step-by-step by solving the equation in each node, the implicit method can only be applied by solving the equation in all nodes simultaneously [
7] through the use of matrix inversion solvers. The implicit method is chosen here for two main reasons. Firstly, the explicit method is conditionally stable and therefore depends on a stability criterion to remain stable and provide valid solutions. Using the implicit method, the stability of the solutions is safeguarded. Secondly, the lack of criterion limitation, and the use of matrix systems of equations, improve the computational performance of the solver.
Equation (
18) can be rearranged in a more useful form:
A similar equation can be obtained for the nodes in the casing. By disregarding the terms related to material phase change and assuming a uniform temperature distribution throughout the casing, Equation (
20) can be obtained. Suppose a casing with uniform temperature distribution and consider the very low Biot number that is obtained for the boundary conditions of the casing, which are related to the casing materials’ high thermal conductivity.
Additionally, four boundary conditions equations have to be defined in order to fully describe the system. The boundary conditions are divided into two solid/fluid boundaries and one solid/solid boundary. The solid/fluid boundaries are present in the inner and outer layers of the combustion chamber, where the material is in contact with the gases from combustion and the ambient air. The solid/solid boundary condition exists in the contact between the ATPS and the casing, where two equations are obtained, one for each boundary node of each material.
Equation (
21) and Equation (
22) are associated with node 1 and node
n, respectively, where node 1 is part of the ATPS, while node
n is part of the casing. The equations are obtained using the same process as Equations (
19) and (
20), only taking into account two additional energy terms, which come from heat transfer in the form of convection and radiation with the surrounding gases.
The boundary condition for the contact between the two materials, which is represented by nodes
k and
, for the ATPS and the casing, respectively, is defined by an additional term that corresponds to the heat transfer coefficient between the two materials. This term has an equivalence with thermal circuits analysis, where we have two thermal resistances in series [
7]. Equations (
23) and (
24) correspond to nodes
k and
, respectively.
According to energy conservation, the heat transferred from node
k to node
must be equal to that received from
, so the term is symmetric in both equations. Note that for all equations, the terms
and
are functions of the temperature in the respective node and are specific to the material of the ATPS. These functions are obtained by integrating the specific heat capacity in temperature [
5,
6,
8]. Furthermore, the specific heat capacity of the gases of pyrolysis (
) is also a function of temperature.
Now that the equations of thermal behavior are defined for both materials and boundaries, a matrix system of equations can be obtained. The implicit method of FDM can be solved using a system of matrix equations, which allows the computation of the state of all nodes simultaneously. Starting by defining the matrix
A as
where
A is a tri-diagonal matrix of the terms of
, with the exception of the terms
, which are defined in the matrix
B.
and
and
and
are terms related to solid/fluid boundary conditions, as seen in Equation (
21) and Equation (
22), respectively. Equivalently, the entries of the matrix
A with indices of
k and
(which correspond to the nodes that define the boundary between ATPS and casing) are related to the solid/solid boundary and are seen in Equations (
23) and (
24), respectively. Finally, the matrix entries
,
, and
and the entries
,
, and
are the terms of Equation (
19) and Equation (
20), respectively, with the subscripts 0, 1 and
indicating the terms of
,
and
, respectively.
Additionally, the diagonal matrix
B is defined as
where the diagonal entries change for each line related with nodes 1 to
k, while for lines
to
n the material properties are constant, and therefore, the diagonal is fixed.
Finally, the vectors
C,
, and
are defined:
where
C is the vector of the constants of the equations of energy, taking into account the terms related to both material phase change and heat transfer with the exterior. Vectors
and
are the vectors of the temperature values in each node, for the current and following time steps, respectively.
Using Equation (
19) to Equation (
24), the following matrix system corresponds to the complete system of all equations of energy in all nodes:
Equation (
25) is the matrix function that gives the temperature value in all nodes in the next time step, knowing the temperature of all nodes in the current time step. This is the matrix solver that is computed to simulate the thermal behaviour of the system at each time step.
In order to estimate the heat convection coefficient inside the combustion chamber
, which controls the rate at which thermal energy is transferred from the combustion to the ATPS, Bartz’s equation (Equation (
26)) [
14,
15] was used, which gives a semi-empirical correlation between the heat transfer coefficient and the flow properties and the geometry of the chamber.
where
is given as:
where
is the diameter of the nozzle throat,
is the radius of curvature of the nozzle,
and
A are the areas of the nozzle throat and chamber, respectively,
is the characteristic velocity of the engine,
is the Prandtl number,
is the specific heat capacity at constant pressure of the gases of the combustion reaction,
is the dynamic viscosity of the gases,
is the pressure inside the combustion chamber,
and
are the temperature of the wall in contact with the combustion and the temperature of the combustion gases, respectively,
M is the Mach number,
is the specific heat ratio of the gases and finally
m is the exponent of the viscosity dependence on temperature (assumed to be 0.6).
In order to reduce the complexity of the information required to run the simulation, some assumptions were made to acquire an approximation of the value of the flow Mach number. The velocity will be different throughout the engine length, and also due to the fact that the post-combustion chamber of a hybrid engine is the location where the flow conditions are more severe [
16,
17] (due to several factors such as, being the section where the ATPS is in direct contact with combustion during the entire burn, the flow has the highest velocity in the region before the nozzle, and in this zone, the oxidizer and liquid/gaseous fuel are well mixed, allowing the complete combustion reaction to occur), the location of the engine chosen as the simulation point is the entrance of the nozzle. This location allows us not only to approximately estimate the flow velocity using well-known propulsion equations but also to assume a simulation for a worst-case scenario, where the conditions of the combustion are more severe.
To determine the Mach number at the entrance of the nozzle, an analysis of the nozzle is performed, assuming that the flow through the nozzle is isentropic [
2]. Additionally, it is assumed that the nozzle is ideally expanded; therefore, the pressure at the nozzle exit is equal to the ambient pressure, which is given as an input [
2]. With these assumptions, the following analysis process can be executed to determine the velocity of the flow at the nozzle entrance.
Firstly, the flow velocity at the nozzle exit is determined by using the following correlation obtained from assuming an isentropic flow through the nozzle.
where
and
are the chamber pressure and the exit pressure of the nozzle (equal to the ambient pressure), respectively, and
is the Mach number at the exit of the nozzle.
Then, the mass flow rate through the nozzle is calculated using the equation obtained from the analysis of the converging/divergent nozzle [
2], with an isentropic flow with area variation, where the flow at the nozzle throat is assumed to be under sonic conditions.
where
is the exit area of the nozzle and
is the temperature inside the chamber.
Assuming a constant mass flow rate through the nozzle, Equation (
29) can now be inverted to determine the Mach number at the nozzle entrance, using the area of the nozzle entrance, which is assumed to be equal to the area of the combustion chamber section. Note that inverting Equation (
29) requires solving a nonlinear equation, for which no analytical solution can be obtained, and so the equation is solved numerically. Additionally, two solutions for the Mach number are obtained from solving the non-linear equation, one corresponding to the subsonic (converging) section of the nozzle and another to the hypersonic (diverging) section; therefore, only the subsonic solution is chosen as the result and applied to Bartz’s equation.
3.2. Mechanical Model
With the thermal model developed, the capability to estimate the temperature in each node of the casing at each time step of the burn was achieved. However, knowledge of the temperature alone is not sufficient to predict the occurrence of a critical engine failure. Although failure detection uses upper limit values of material temperature (related to the maximum service temperature of the casing material), additionally, a mechanical model of the combustion chamber is needed to predict a mechanical failure of the casing. To address this problem, a simple pressure vessel analysis would not be enough, since this type of analysis would not take into account the decrease in yield strength with temperature of the casing material. Knowledge of the temperature of the casing at each moment in time is also necessary in order to determine the mechanical properties of the casing at each moment and evaluate the capability of the casing to withstand the applied instantaneous pressure load.
Similarly to the thermal model, some assumptions were made to simplify the analytical model used. Firstly, only the elastic deformation of the material is considered [
18,
19], since, for the study of reusable rocket engines, it is essential that the structural integrity of the materials does not exceed the elastic range to prevent permanent deformations.
Secondly, only isotropic materials are considered for the casing. This assumption significantly simplifies the analytical models employed and is justified by a third assumption. The casing material has high thermal conductivity. More specifically, it is assumed that the Biot number at the interface between the casing and the external air is very low [
7], which implies that the temperature distribution across the thickness of the casing is approximately uniform. Under this condition, thermal stress can be neglected, which is a desirable characteristic for combustion chamber casings [
2]. The use of a metallic material with high thermal conductivity minimizes temperature gradients, thus substantially reducing thermal stresses. This helps prevent one of the primary causes of structural degradation in the casing: fissure formation [
2].
When studying the use of metals as a material choice for the casing of the combustion chamber, the isotropic analytical equations for the mechanical model can be used, and the thermal stress loads can be neglected [
9] (note that, as explained in the thermal model, the thermal expansion is also neglected), simplifying the calculations. The final assumption is made, which has already been used in the thermal model, which considers that the geometry of the combustion chamber is axisymmetric.
Following the work developed by Habib et al. [
19] and Wang et al. [
18], the development of the mechanical model starts with the stress equations for an uncoupled elastic steady-state system.
where the subscripts
r,
, and
z indicate the cylindrical coordinates of the stress, radial, hoop, and axial. The stress and the displacement are represented by
and
, respectively. The Young’s modulus is represented by
E, while
is the Poisson’s ratio. Finally,
and
are the temperature in function of the radius and the thermal expansion coefficient, respectively. Note that the assumption of a uniform temperature distribution across the casing gives a temperature function that is constant; therefore,
T does not vary in the radial direction.
Starting by solving the equilibrium Equation (
7), by replacing the stresses with the corresponding correlations given in Equation (
30), the overall equilibrium equation is given as
Replace the strains in the previous equation with the correlations for the axisymmetric geometry given in Equation (
8), we obtain the following equilibrium equation for axisymmetric vessels
Equation (
31) is a second-order differential equation. Notice that the format of this differential equation is of the type of Euler–Cauchy differential equations for which there are some known solutions. Since Equation (
31) has two real roots, the solution of the differential equation is given as
, where
and
are the real roots. For Equation (
31), the roots are given as
and
. Therefore, the solution to the Euler–Cauchy differential equation is the displacement function, given as
where
A and
B are the integration constants to be determined.
Now, by replacing the strains in the system of equations in Equation (
30) with the corresponding strain–displacement correlations given by Equation (
8), the equations of stress can be presented as
Note that these are the stress values that correspond uniquely to the mechanical loads. Because the thermal stress load can be neglected due to the uniform temperature distribution, by applying the linear-elastic and uncoupled thermo-mechanical conduction conditions, the combined stress will be just the mechanical stress [
18]. Using the obtained solution for the displacement function
, and noting that
, the previous system of equations can be rearranged, giving
with six unknown variables
,
,
,
A,
B and
, and a system of three equations, the system is under-defined. Three additional equations are needed, which are obtained from the boundary conditions of the axisymmetric vessel under pressure load. These boundary conditions can be defined as
where
p is the internal pressure of the combustion chamber, while
and
are the inner and outer radius of the casing, respectively. Note that the ATPS is assumed to have no structural utility, being simply applied as a thermal device, therefore, its contribution to the structural integrity of the engine is neglected, hence the use of the inner radius of the casing is chosen, and not the inner radius of the combustion chamber as a whole. By applying these boundary conditions, and after some solving and rearrangement of the system of equations, the values of
A,
B, and
can be determined
Finally, by replacing the values obtained in the stress system of equations in Equation (
34) and simplifying the results, the final equations of stress for the cylindrical coordinate system can be obtained
The mechanical model predicts the margin of safety that is expected for the engine design being studied from a structural point of view. For this, the formula for the mechanical margin of safety in stress loads is introduced as
where
is the factor of safety, which is a design factor that defines the relation between the maximum applied stress load and the maximum stress the material can withstand. For
, it is the strength of the material’s yield, the upper limit of the stress the material can withstand in the range of elastic deformation [
3,
9]. Lastly,
is the equivalent Von-Mises stress for the principal stresses of the system, given as
3.3. Materials
The first step in the process of testing and applying the developed model is the choice of materials to be studied and their corresponding properties. Six different materials were chosen as study subjects—three materials for the ATPS and three materials for the casing—providing a total of nine possible combinations of material designs.
For the case of the ATPS, the chosen materials are:
Glass fiber phenolic composite;
Carbon fiber epoxy composite;
Twaron fiber EPDM composite.
The glass and carbon fiber composites are categorized as fiber Reinforced Polymeric Ablators (FRPAs), while the twaron fiber composite is categorized as an Elastomeric Heat Shielding Material (EHSM). The glass fiber phenolic was chosen due to the available data and studies used in this article, with an emphasis on the work developed by Li et al. [
6], whose analytical thermal model was one of the bases for the work developed in
Section 3.1, and the experimentation data performed by the author using glass fiber phenolic composites are valuable for model validation. In terms of the carbon fiber epoxy, it was chosen because it is a very common composite in the aerospace industry. Composite materials using carbon fibers have been extensively studied, developed and used in a wide range of applications during the past years [
20]. It is therefore of interest to study its applicability for this specific function. Lastly, the twaron EPDM composite is chosen for a state-of-the-art comparison with the rest of the materials. EPDM matrix-based composites have been used for many years as the go-to advanced material for ATPS in solid rocket motors [
5,
21], and are therefore chosen to provide an industry standard comparison with the remaining materials.
As for the materials chosen for the casing of the combustion chamber, these are:
Note that only metals were chosen for the casing material. This is due to the assumptions and simplifications that were applied in the development of the mechanical model, mentioned in
Section 3.2. One of the design requirements for the casing material is a high maximum operating temperature, a characteristic in which most composite materials cannot compete with metals. Moving to the materials themselves, the stainless steel alloy AISI 304 was chosen because it is one of the most common stainless steel alloys available in the industry. The wide availability of this alloy will be used as the strength point for this material, providing a low-cost alternative to compare with the other metals. The 6061-T6 aluminum alloy has improved mechanical properties compared to other aluminium alloys [
22]. Due to its high yield strength and maximum operating temperature, this alloy is chosen to represent a high-performance choice among aluminum alloys. Finally, titanium alloy, similar to the choice of EPDM twaron, is the state-of-the-art choice for the casing [
23]. Of the three choices, the material that possesses the better balance of the design requirements is the material with the highest yield strength and the second highest maximum operating temperature. However, it is also the most expensive of the materials. While, on average, the aluminum alloy can be about twice as expensive as the stainless steel per unit of mass, the titanium alloy can be ten times more expensive than the aluminum alloy, even more for other advanced alloys [
24]. This specific alloy of titanium was chosen because it is one of the most commonly used titanium alloys in engine components in the aerospace industry [
25].