1. Introduction
Aircraft electrical power requirements have increased exponentially over the first hundred years of powered flight to support on-board technological advancements including flight computers, actuators, power generation and distribution systems, directed energy, and propulsion systems [
1]. This trend is projected to continue with the development of future high-power and power-dense electronic components [
2]. Next-generation electronic systems, such as radar, are expected to generate a heat flux greater than 1000
in waste heat [
3], which must be removed from an aircraft to ensure safe and continuous operation. Traditional aircraft thermal subsystems and heat sinks are approaching the limit of their ability to adequately transport and dissipate ever-increasing heat loads. To address this shortcoming in cooling capability, novel aircraft thermal management subsystem concepts, such as two-phase solutions, have been investigated and characterized to provide constant surface temperature heat rejection from a platform [
4].
A mechanically pumped two-phase loop (MPTL) is a thermal subsystem that leverages refrigerant phase change to provide a near-isothermal evaporator temperature for cooling electronics [
5]. MPTLs have been fielded in terrestrial applications [
6], microgravity flow boiling experiments [
7], and spacecraft thermal control systems [
8]. In addition, they have been studied for high-heat-flux aircraft avionics applications [
9]. MPTLs provide many benefits compared to single-phase pumped loop systems, including a lower required mass flow rate to maintain heat source surface temperature, a low-pressure differential throughout the cycle, a narrow temperature operation range, an ability to operate with physically distributed heat loads (i.e., multiple evaporators), and stable operation over a broad heat load range [
10,
11,
12]. Unfortunately, MPTLs have known issues with evaporator start-up and shut-down [
10,
13], which may lead to pressure oscillations [
14] and could result in hardware damage [
15] through evaporator dryout or partial dryout [
16].
These known issues led to the study of MPTL system dynamics with numerical and experimental methods by several researchers. A numerical model was developed and experimental data used to chart an increase in gas- and liquid-phase pressure when a 100 W step evaporator heat load was applied to an MPTL. This heat load decreased the refrigerant mass flow rate through the system as refrigerant mass was transferred to the accumulator throughout the heat load duration [
17]. Refrigerant mass exchange between the accumulator and the rest of the MPTL system has been identified by other research groups as influencing system dynamics [
18]. A research group described the importance of the accumulator in transient operation by equating it to the system ‘brain’ to manage temperature or pressure. This group found that temperature overshoot, pressure fluctuation, and evaporator temperature change are related to accumulator mass exchange [
5]. The impact of the accumulator volume on the transient response time was evaluated and it was found that a large-volume accumulator is less sensitive to changes in evaporator input power than a small accumulator [
19]. An analytical model was developed to describe the accumulator volume’s influence on MPTL system performance. This model showed that the accumulator volume has a large effect on boiling temperatures and evaporator exit vapor quality [
20]. Another study demonstrated that pressure oscillations within the accumulator are determined by the evaporator heating load, accumulator volume, and initial thermodynamic state of the loop and accumulator [
21]. Studies have been performed to evaluate the impact of accumulator volume [
20] and the location of the compressible volume provided by an accumulator within an MPTL system [
22]. Compressible volume exists in components where refrigerant vapor is present, such as an accumulator. The compressible volume allows a system to respond to internal pressure changes without transmitting the pressure change through the remainder of the system. Compressible volume has been passively implemented through the presence of a surge (expansion) tank to dampen flow instabilities [
3]. In summary, transient evaporator events such as start-up and shut down are proven to impact MPTL system pressure, temperature, and refrigerant mass flow rate due to mass exchange between the compressible volume accumulator and the remainder of the system. Thus, there is a need to further explore the integration of a compressible volume accumulator within an MPTL. It is imperative to address highly transient evaporator heat loads to maintain the evaporator setpoint temperature.
Various accumulator technologies have been developed to improve MPTL evaporator temperature control. Pressure-controlled accumulators [
7] and temperature-controlled accumulators [
8] have been used in spacecraft applications by NASA. Temperature-controlled accumulators add or remove heat from the accumulator to maintain the system saturation pressure. To achieve a desired saturation pressure, pressure-controlled accumulators modulate the system volume. Pressure-controlled accumulators have been reported to respond faster than temperature-controlled accumulators [
19], which is important for pulsed heat load applications. The fielded spacecraft accumulator technologies target evaporators with low magnitudes of heat input, at 10–100 W [
23]. Accumulator technologies require further investigation to assess the integration feasibility of an accumulator into an MPTL for future high-performance aircraft thermal subsystems [
2].
The MPTL transient performance is also impacted by the selected control scheme. Throughout the aerospace industry, proportional–integral (PI) controllers are commonly used to regulate flow rate and temperature. PI controllers are effective in systems with linear behavior and where response rates are slow [
23]. However, they may not be suitable for systems that have highly nonlinear behavior or rapid dynamics, such as with rapidly pulsed evaporator heat loads. A more computationally intensive, and potentially improved, control strategy is model predictive control (MPC). MPC uses a representative mathematical plant model to predict future states by calculating the optimal control decision over a time horizon to minimize an objective function cost [
24,
25]. The performance impact that a control scheme has on an MPTL has been previously evaluated. Li et al. [
26] compared the ability of an MPTL to track an evaporator setpoint temperature with a PI controller and MPC. The model predictive control more closely tracked the evaporator setpoint temperature and reduced the average transient time by at least four times relative to a PI controller. The system evaluated used a fixed-volume receiver and did not actively control compressible volume to influence transient performance. Another research group compared the MPTL response with two different control schemes: an active disturbance rejection controller (ADRC) and a PI controller [
27]. The ADRC was able to improve temperature tracing and regulation relative to the PI controller. The evaluated MPTL systems were fixed-volume for both research groups. In another study seeking to improve the MPTL control methodology, the refrigerant flow rate was recommended as a process variable to include in controller design. The modulation of the refrigerant flow rate allowed the average wall temperature to be reduced by 0.74 °C relative to a fixed refrigerant flow rate [
28]. In summary, previous research has demonstrated MPTL performance improvements from MPC compared to PI control techniques. However, these MPTL architectures did not include components to modulate the system volume, such as a pressure-controlled accumulators. The combination of MPC and a pressure-controlled accumulator within an MPTL has the potential to improve evaporator isothermal transient performance under pulsed, high-heat-flux electronics cooling. However, this combination of technologies has not previously been evaluated.
The present study explored the transient performance resulting from the integration of an MPTL, a pressure-controlled accumulator, and an advanced MPC scheme relative to a PI- controlled MPTL with an uncontrolled accumulator volume. The study collected experimental data to validate the thermal and flow performance of a numerical model for three-step evaporator heat loads. The validated model was used to quantify transient behavior and compare the two control schemes under multiple applied evaporator heat load profiles, to maintain a near-constant refrigerant setpoint temperature at the evaporator outlet. As individual technologies, MPC and controlled accumulators have been proven to improve temperature setpoint tracking for MPTLs. However, the combination of these two technologies has not previously been evaluated.
3. Numerical Model Description and Validation
A transient numerical model was created in the physical modeling software MathWorks Simscape™ R2023a, to represent system behavior. Simscape domain libraries provide non-directional blocks where the equation formulation does not explicitly specify inlet and outlet ports. These blocks can be parameterized to represent physical components and track energy and mass flow throughout a system [
31]. Thermo-fluid properties are calculated in Simscape through the domain-specific ‘Fluid Properties’ function.
Figure 3 is the Simscape representation of the experimental hardware shown in
Figure 1.
A brief formulation of the underlying equations can be described for the two-phase pipe component (i.e., Pipe (2P)). In the MPTL model shown in
Figure 3, each subsystem denoted ‘PIPE_#’ contains a Pipe (2P) component. The assumptions of the Pipe (2P) component include fully developed flow, negligible influence of gravity, and no fluid inertia effects [
32]. An additional assumption was imposed through the modeling strategy to eliminate heat transfer to the environment. These assumptions resulted in the governing equations below.
Equation (1) below is the energy equation [
32].
In Equation (1),
represents the time derivative of the fluid internal energy within the pipe,
and
, are the mass flow rates into and out of the pipe,
is the fluid internal energy,
and
are the energy flow rates into and out of the pipe, and
is the heat transfer rate into the pipe through the pipe wall. The Gnielinski correlation was used to determine the Nusselt number for single-phase turbulent flow through a pipe [
33]. The Gnielinski correlation is valid for smooth tubes over the ranges
and
. The Cavallini and Zecchin correlation was used to determine the Nusselt number for a two-phase mixture within the pipe [
32]. The Cavallini and Zecchin correlation is a semi-empirical correlation, which has been demonstrated to show better accuracy compared to many other correlations [
34].
Equation (2) is the mass equation [
32].
In Equation (2), is the fluid density, is the pipe internal pressure, is the fluid volume in the pipe, and are the mass flow rates into and out pf the pipe, and is a density smoothing correction term. This formulation accounts for compressible flow effects in the partial derivative terms. The density smoothing term, , is used to reduce numerical interpolation errors at phase boundaries for fluid properties.
Equation (3) is the momentum equation [
32].
In Equation (3),
is pressure,
is the mass flow rate,
is the cross-sectional area,
is the fluid specific volume inside the pipe, and
represents the effects of viscous friction forces. The viscous friction force depends on the flow regime and the Darcy friction factor, estimated by the explicit Haaland equation [
33]. The Haaland equation avoids the need for iteration, as required by use of a Moody diagram and, thus, is convenient for numerical calculations. The pipe component was parameterized with measured values, such as length and diameter. A similar process was followed for each component shown in the
Figure 3 MPTL representation.
A variable-step Simulink solver (daessc) with default tolerances was selected for model convergence [
31]. There were no noted issues with model convergence. The initial conditions were manually assigned to match the experiment data five seconds before the pulse, to prevent numerical initialization errors from impacting the results. To evaluate the control scheme’s impact on transient performance, a PI controller and an MPC controller were designed and implemented in the numerical model.
3.1. Control Design
The numerical model was used to create state-space representations of the dynamic system to develop the PI control and MPC logic. While the two control methods contained different effectors, a similar process was used to generate each state-space representation. The state-space representations were reduced to improve computational speed and stability. The process is described in more detail for each controller below.
3.1.1. Proportional–Integral (PI) Controller Design and Model Linearization
A single-input, single-output (SISO) feedback PI controller was designed to serve as a baseline control architecture. In this controller, the evaporator exit pressure () was the measured variable (MV) and the condenser water-side volumetric flow rate was the process variable (PV). As the refrigerant evaporator exit pressure increases, the condensing water-side valve will open to transfer heat from the MPTL to the water-side reservoir.
The representative Simscape model was linearized to develop a PI controller. To linearize the system, the Simscape numerical model was evaluated with a constant condenser water-side volumetric flow rate to arrive at a steady-state condition. Equation (4) shows the output vector,
, and derivative of the state equation,
, which were developed after the physical Simscape equations were linearized. From the state-space representation, shown in Equation (4), pole-zero pairs were canceled within the MATLAB default tolerance to remove unobservable and uncontrollable states [
29]. Then, unstable parts of the state-space representation were decomposed into a stable state-space representation, as shown in Equation (5) [
35]. In Equation (5),
and
are the state-space representation.
are the unstable modes, and
are the stable modes.
A pole map was generated to compare the original state-space representation with the reduced state-space representation. The reduced state-space representation contained many high-frequency, oscillatory poles and roots and fewer low-frequency poles and roots. The reduced state-space linear representations of the system gains,
and
, were determined to achieve the fastest stable response.
3.1.2. Model Predictive Controller (MPC) Design and Model Linearization
A multiple-input, multiple-output (MIMO) MPC was designed with three manipulated variables (MVs), one measured disturbance (MD), and two measured outputs (MOs). These manipulated variables were selected according to previous research, as refrigerant pump voltage (refrigerant volumetric flow rate), accumulator-nitrogen pressure regulator voltage (refrigerant saturation pressure), and condenser water-side throttle valve current (condenser water-side volumetric flow rate). The measured disturbance was the pulsed evaporator heat load, denoted
. The measured outputs were the evaporator outlet pressure (
) and pressure-controlled accumulator displacement (
). The MPC quadratic program objective function is
In Equation (6),
is the quadratic program decision at each timestep,
is the output reference tracker term,
is the manipulated variable tracking term,
is the manipulated variable move suppression to eliminate large changes in manipulated variables, and
is the term to prevent constraint violation. The output reference tracker term
is shown in Equation (7). This term evaluates each plant output variable
at each prediction step
, determines the difference between the reference value
and predicted value
, and applies a weight and scale term
. Each of the terms shown in Equation (6) has a similar form to the output reference tracker term
shown in Equation (7). Additional information about the optimization solver is provided by MathWorks [
25].
A multiple-input, multiple-output (MIMO) MPC was designed following a similar linearization and state-space reduction process as that performed with the PI controller. The MPC was designed with three manipulated variables (MVs), two measured outputs (MOs), and one measured disturbance (MD). The variables captured by the MPC were selected from reference MPTL control methodologies, recommendations from prior studies [
28], and parameters enabled by the incorporation of the pressure-controlled accumulator. The numerical model was evaluated with a constant condenser water-side volumetric flow rate, refrigerant volumetric flow rate, and accumulator pressure to arrive at a steady-state condition for linearization. The linearized state-space representation was decomposed and reduced with the same methods as the PI controller. The reduced state-space representation contained many high-frequency, oscillatory poles and roots and fewer low-frequency poles and roots.
Each of the manipulated variables and measured outputs was assigned a hard constraint, according to hardware limitations, which the MPC controller objective function could not violate. Objective function weights were selected to modulate the system as desired and improve controller stability. Manipulated variable hard constraints and objective function weights are shown in
Table 2.
System dynamics dictated the sampling period to be 0.1 s, prediction horizon to be 60 s, and the control horizon to be 12 s. The controller performance was verified in terms of closed-loop internal stability, closed-loop nominal stability, steady-state gains, and constraints.
3.2. Statistical Metrics
To compare the experimental results with the numerical model for validation, a 25-s measurement window, , was defined. The measurement window was defined as a five-second pre-evaporator load pulse (), five-second evaporator load pulse (), and then fifteen seconds after the evaporator load pulse response ().
Over this measurement window, two metrics are defined: maximum percent error (MPE) and mean average percent error (MAPE). MPE reports the maximum percent error over the measurement window to determine the condition with the maximum difference between the experimental data and predictions. MAPE reports the average percent error over the measurement window to provide an estimate for transient model assessment. For a function
, with measured values
and predicted values
, MPE and MAPE are useful and defined by Equations (8) and (9).
Furthermore, in applications where multiple evaporator load pulses can be applied, it is useful to define a metric for which the system returns to a steady state and the evaporator load can be applied a second time. This parameter was chosen to be a 3% settling time, which is the time it takes for the system to reach
of the initial condition after an evaporator pulse load has been applied [
36].
To evaluate the impact of an evaporator heat pulse, the pressure and temperature increases are calculated across the duration of the pulse. Equation (10) shows the pressure rise at the evaporator outlet for the pulse duration.
3.3. Model Validation
Previous work has demonstrated that the system transient response is influenced by the total system volume and system charge [
17]. Prior to experimental data collection, the MPTL volume was measured. A nitrogen tank with a known initial pressure and volume was expanded into the MPTL. The measured final pressure was measured, and the final volume was calculated by the Ideal Gas Law. The experimentally measured volume was within 0.85% of the model-represented volume.
The purpose of the collected experimental data was to describe the impact a passive accumulator has on an MPTL transient response under a pulsed evaporator heat load. A brief description of pressure and temperature response is provided below. MPE and MAPE were used as metrics to validate the numerical model response, and the 1200 W evaporator step heat input was used for discussion. Within the MPTL cycle, system transient performance was strongly influenced by the pressure at each cycle point. Since an MPTL cycle is nearly isobaric, increased pressure at one cycle point will increase pressure throughout the system. As slightly subcooled refrigerant flowed through the evaporator and a 1200 W heat load was applied,
, the subcooled liquid changed phase to a two-phase liquid–vapor mixture. The two-phase liquid–vapor mixture had a lower density (107
) than the subcooled liquid (1225
), which increased the system volume through the accumulator and increased system pressure. Upon termination of the evaporator heat load,
, heat continued to be removed from the cycle via the condenser and the MPTL pressure returned to the initial state. This behavior is presented in
Figure 4A. Over the five-second, 1200 W pulsed evaporator load, the evaporator outlet pressure increased by 0.043 MPa. After the evaporator pulse was removed, the MPTL system required 6.8 seconds to reach a 3% settling time. The pressure-controlled accumulator had a constant setpoint of 0.66 MPa. As the evaporator outlet pressure increased, the pressure-controlled accumulator increased the system volume to maintain a system 0.66 MPa saturation pressure.
Figure 4A shows the pressure response measured in the experiments, as captured by the numerical model. The predicted pressure response has a similar slope and maximum pressure when compared to the experimental data. Deviations between the model and experimental data are within the experimental measurement uncertainty (
0.13 MPa) and can be attributed to minor differences in the model initial conditions and heat rejection via the condenser.
Like the pressure response, the evaporator outlet temperature increased while the evaporator pulse was applied, and then it decreased once the evaporator load was removed.
Figure 4B shows the evaporator outlet temperature transient response under a 1200 W evaporator pulse. The evaporator outlet refrigerant temperature increased by 2.5 K over the evaporator load pulse duration. In addition, the model underpredicted the maximum temperature relative to the experimental results. However, the difference between the measured and calculated maximum temperatures was relatively small, at roughly 1 K.
Figure 4B shows that the numerical model captured the general shape of the measured temperature response.
The predicted refrigerant mass flow rate agreed within 3% of the experimentally measured mass flow rate across the evaporator pulse duration. The numerical model captured a decrease in refrigerant mass flow rate upon evaporator heat load application, in accordance with a previous study [
17]. However, the refrigerant volumetric flow meter was placed upstream of the recuperator in a subcooled liquid refrigerant region, and therefore did not capture these dynamics.
Table 3 shows the measured evaporator outlet refrigerant pressure and temperature for each of the three evaporator heat pulses in
Figure 2A. As the magnitude of heat input increased, the associated pressure and temperature rise increased.
Table 4 compares the measured and predicted pressures and temperatures at each cycle point inlet and outlet for the 1200 W heating pulse. The cycle points correspond to those in
Figure 3. Relatively small values of the MPE and MAPE indicate that the model predicts the measured pressure reasonably well. In addition, the MPE and MAPE were averaged across all cycle points to create a single value to represent the model fit for each of the three pulse load values, shown in
Table 4. With this metric, the MPE agrees within 3.45% and the MAPE agrees within
. This indicates that the model predicted a reasonable transient pressure and temperature performance under multiple pulsed evaporator loads relative to the experimental data.
5. Summary and Conclusions
In this study, experimental methods were used to characterize a system response for an MPTL with a pressure-controlled accumulator evaluated for three different pulsed evaporator heat loads. The pressure-controlled accumulator contained a flexible bladder and was actuated by compressed nitrogen. The experimental results were compared with numerical simulation data for model validation. With the validated model, traditional PI control and an advanced MPC were developed. Both control strategies were numerically evaluated against three evaporator heat load profiles, which varied heat input magnitude, duration, cycles, and spacing to determine the impacts that the control scheme and compressible volume have on MPTL performance for isothermal () evaporator operation.
In the current work, a representative, transient numerical model well-represented the pressure, temperature, and mass flow rate response for multiple cycle points and multiple step evaporator heat loads. Under pulsed evaporator heat loads, the MPE agreed within 3.45% and the MAPE agreed within . The model-predicted refrigerant mass flow rate agreed within 3% of the experimentally measured mass flow rate across the evaporator pulse duration. The model was leveraged for control development.
The transient performance levels of a PI- and MPC-controlled MTPL were evaluated for three simulated evaporator heat input profiles. In all profiles, the MPC scheme produced less saturation pressure variability than the PI control scheme, which led to a more uniform saturation temperature across the evaporator load duration. As evaporator heat input transients increased (Profiles 1 and 3), the MPC-controlled MPTL was able to maintain temperature requirements, whereas the PI-controlled MPTL did not. However, for low-heat-flux applications (e.g., Profile 2), the two control schemes showed similar thermodynamic responses. The addition of the refrigerant mass flow rate and nitrogen-side pressure to control the accumulator compressible volume as effectors improved the MPC thermodynamic response compared to the PI-controlled response. These additional effectors enabled the MPC system to decrease the generated refrigerant liquid–vapor mixture upon evaporator heat input and provided actuation to re-compress the accumulator for heat rejection.
From this work, it can be concluded that the incorporation of a controlled compressible volume, via an accumulator, into an MPTL has the potential to provide near-isothermal () evaporator operation throughout a pulsed heat load duration. Furthermore, the incorporation of advanced control algorithms and novel manipulated variables, such as nitrogen-side pressure, to actively control accumulator compressible volume offers performance advantages for pulsed evaporator heat loads. As future aerospace designs demand increased thermal capability, an MPTL with an accumulator should be evaluated, especially when design constraints require isothermal evaporator performance.