In this paper, the model for the supervisory control approach can be divided into two sub-models: engine thermal model and TWC thermal model. The steady-state engine model only outputs hot engine data, so establishing the engine thermal model that takes the engine’s coolant temperature into account to predict cold engine outputs is necessary, especially for engine-out emissions during cold-start. The TWC thermal model includes catalyst temperature dynamics, which are required for computing the conversion efficiency of the TWC.
3.1. Engine Thermal Model
The engine thermal model can be further divided into two sub-blocks, coolant temperature dynamics, and cold-engine correction factor. A lumped-capacitance thermal network model, depicted in
Figure 2, was defined for the coolant temperature dynamics model. The heat transfer equations for calculating the coolant temperature are listed in
Appendix A.
For control purposes, a simplified model that predicts cold engine outputs should be built. One approach is to simply multiply hot engine outputs by a cold factor, which is a function of the coolant temperature. This method was investigated by Murrell et al. [
31], which can be expressed as follows:
where
is the engine’s coolant temperature factor,
is the engine cooling system’s thermostat set point, and
is the coolant temperature. These temperatures are expressed in degrees centigrade.
,
,
, and
are the hot engine fuel consumption rate and hot engine outputs rate of HC, CO, and NOx, respectively. These values can be obtained by looking up maps by engine speed and engine torque, shown in
Figure 3. From
Figure 3a, it can be seen that fuel enriches at high speed or torque. This can be explained by looking at the fundamentals of the internal combustion engine. For the high speed or torque, the only way to increase power is to richen the mixture after the wide-open-throttle. Thus, after the wide-open-throttle, the mixture becomes richer and richer, the air/fuel ratio gets smaller and smaller and fuel efficiency is lower and lower as the required power increases.
,
,
, and
are the cold engine fuel consumption rate and cold engine outputs rate of HC, CO, and NOx, respectively; and
,
,
,
,
,
,
, and
are the curve-fitting parameters. The cold-start test data of fuel consumption and emissions are not available to the author, so above parameters come from Advanced Vehicle Simulator (ADVISOR). Their values are 1, 3.1, 7.4, 3.072, 9.4, 3.21, 0.6 and 7.3, respectively.
3.2. TWC Thermal Model
The schematic diagram of the TWC is shown in
Figure 4. The TWC can be simplified as the catalyst monolith (A), catalyst internal shell (B), and catalyst external shell (C). The TWC thermal model uses the exhaust gas flow rate (
) and the exhaust gas temperature (
) at the TWC inlet as its inputs. The value of these inputs can be obtained by looking up maps indexed vertically by engine speed and horizontally by engine torque. The exhaust gas flow rate map, shown in
Figure 5a, is available from the fuel map and air-fuel ratio (A/F) according to the following equation:
The map of the exhaust gas temperature is shown in
Figure 5b, which can also be obtained from the fuel map through Equations (3)–(5):
where
is the engine’s waste heat,
is the mass of fuel from fuel map,
is the lower heating value of the fuel,
is the speed of the engine with the unit of rad/s,
is the engine torque,
is the power of the exhaust gas, and
is the fraction of waste heat that goes to the exhaust. This parameter can be estimated by engine speed, it is shown in
Figure 5.
is the temperature of the exhaust gas,
is the exhaust gas flow with unit g/s,
is the capacitance of the exhaust gas, and
is the ambient temperature. It is known from
Figure 6b, the exhaust gas temperature decreases with the increasing engine torque in most regions. According to Equation (5), since
and
are constants, the exhaust gas temperature is determined by the power of the exhaust gas and the exhaust gas flow. To explain the above phenomenon in
Figure 6b, we take 3000 rpm as an example. The change curves of the exhaust gas power, flow and temperature with engine torque when engine speed is 3000 rpm are shown in
Figure 7. According to this figure, when the engine torque is below 100 Nm, the exhaust gas flow’s increasing rate is greater than the rate of the exhaust gas power; therefore, the exhaust gas temperature decreases with the increasing torque in this torque range. When the engine torque is between 100 Nm and 140 Nm, the line segment AB is nearly parallel to CD, and BE is nearly parallel to DF, these mean that the exhaust gas flow’s increasing rate is nearly the same as the rate of the exhaust gas power; thus, the exhaust gas temperature is almost unchanged during this torque range. When the engine torque is above 140 Nm, the exhaust gas power’s increasing rate is obvious greater than the rate of the exhaust gas flow, so the exhaust gas temperature increases with the increasing engine torque during this torque range.
Neglecting the exhaust gas heat loss to the exhaust manifold and catalyst inlet/outlet pipes, the thermal network representing the catalyst and associated thermal elements is shown in
Figure 8. The TWC was modeled through a three-node lumped capacitance model, which includes monolith (
), internal shell (
) and external shell (
). Heat is exchanged from the exhaust gas to the node of the monolith and the node of the internal shell through convection. A part of the monolith’s heat transmits to the internal shell via conduction. Heat is exchanged from the node of the internal shell to the node of the external shell via conduction and radiation. Heat transfers from the node of the external shell to the ambient air through convection and radiation.
and
are the resistance of thermal conduction,
and
are the resistance of convective heat transfer, and
and
are the resistance of radiative heat transfer. These thermal resistances can be obtained by the following equations:
where
is thermal conductivity,
is the corresponding surface area,
is the representative distance between nodes,
is the convective heat transfer coefficient,
and
are the surface temperatures,
is the emissivity, and
is the Steffan–Boltzman constant.
Then, the equation for the lumped capacitor model is described as:
where
is the mass of the catalyst monolith (ceramic);
is the lumped thermal capacitance of the catalyst monolith
is the temperature of the catalyst monolith;
is the convective heat transfer coefficient between exhaust gas and catalyst monolith, which is a function of exhaust gas flow, and the function, from ADVISOR, is expressed as follows.
is the inner (honeycomb) surface area of the catalyst monolith;
is the temperature of the catalyst internal shell;
is the change rate of the catalyst monolith’s thermal energy;
is the net heat flow from the exhaust gas to the catalyst monolith through convection;
is the net heat flow from the catalyst monolith to the catalyst internal shell via conduction;
is the net heat flow from chemical reactions of the exhaust gas; and
,
, and
are the flow rates of HC, CO, and NOx in the exhaust gas, these parameters can be calculated with Equation (1). According to Equation (1),
,
and
, also expressed as
,
, and
, are the function of engine torque, engine speed, and the engine’s coolant temperature, while the engine’s coolant temperature is the function of the engine torque and engine speed. Therefore, these parameters are dependent on engine torque and engine speed.
,
, and
are the molar masses of HC, CO, and NOx, respectively;
,
, and
are the heat production of HC, CO, and NOx, respectively, by chemical reaction.
,
, and
are the conversion efficiencies of HC, CO, and NOx, respectively. Catalyst conversion efficiencies are the function of catalyst temperature. Additionally, there is a catalyst efficiency adjustment (decrease), which is made at high exhaust flow rates. The relationship between conversion efficiency and the catalyst temperature can be described by arctan functions.
where
represents each type of emission;
is the light off temperature of each type of emissions;
is a tuning parameter that represent a slope of the efficiency function;
is a correction factor for the exhaust gas flow rate, this factor can be approximated by a linear function as shown below [
32].
where
is flow rate of each emission;
,
and
are the parameters estimated by experimental data. Finally, the list of parameters is shown in
Table 2.
The list of equations used for the catalyst internal shell (B) is described as follows:
where
is the mass of the catalyst internal shell;
is the lumped thermal capacitance of the catalyst internal shell,
is the convective heat transfer coefficient between the exhaust gas and the catalyst internal shell;
is the surface area of the catalyst internal shell;
is the temperature of the catalyst external shell;
is the change rate of the catalyst internal shell’s thermal energy;
is the net heat flow from the exhaust gas to the catalyst internal shell through convection;
is the net heat flow from the catalyst internal shell to the external shell via conduction and radiation;
is the emissivity; and
is the Stefan–Boltzmann constant
.
The list of equations used for catalyst external shell (C) is as follows:
where
is the mass of the catalyst external shell;
is the lumped thermal capacitance of the catalyst external shell, and the value is 460 J/kgK;
is the ambient temperature;
is the change rate of the catalyst’s external shell’s thermal energy;
is the net heat flow from the catalyst external shell to the ambient air via convection and radiation; and
is the surface area of the catalyst’s external shell.
After converting the above equations, the state equation of the catalyst monolith’s temperature was obtained:
where
is the engine torque and
is the engine speed.
3.3. Parameter Estimation and Model Validation
We focus on parameter estimation and model validation of the engine and TWC thermal model in this section. Some of these thermal models’ parameters, shown in
Table 3, are from ADVISOR, which was developed by the American National Renewable Energy Laboratory (NREL) for rapid analysis of the performance and fuel economy of conventional, electric, and hybrid vehicles and contains a component data file (its directory is “:\ADVISOR 2002\date”), this file contains several sub files, such as fuel converter data files, exhaust after treatment files, transmission data files, and driving cycle files. We can obtain the parameters in
Table 3 by inspecting the fuel converter data files and exhaust after treatment files. The remaining parameters, more sensitive to coolant and catalyst temperature, come from calibration. These remaining parameters are calibrated by comparing the model’s temperature response with the test data. The parameters of the engine thermal model are tuned to match the coolant temperature responses of the model to those of the road test data, after the engine thermal model is tuned properly, parameters of the TWC model are then tuned to generate the catalyst temperature responses that match with those of road test data. Note that only limited the coolant and catalyst temperatures test data are available to the author due to the difficulty of experiment set-ups for the engine out emissions and tail out emissions, and thus, the tail-pipe emission of the model response with those of test data is not performed. The process of the real vehicle on road test and parameter estimation are shown as follows. Firstly, we chose a route in our university campus for the plug-in hybrid electric vehicle’s road testing. Then, the vehicle starts by the ISG motor. After starting, the operating mode of the vehicle was switched from pure electric mode to only engine driving mode. The signals of the vehicle speed, the engine speed, the engine torque, the coolant temperature and the catalyst temperature are collected during the test. The vehicle speed is shown in
Figure 9a, and the engine torque and motor torque are shown in
Figure 9b. Finally, using the vehicle speed, the engine speed, and the engine torque test data as inputs to the thermal model of engine and TWC, the parameters of the engine thermal model are tuned to match the coolant temperature responses of the model to those of the road test data, the comparison between model coolant temperature and test data temperature is shown in
Figure 9c. The parameters of the TWC thermal model are tuned to match the catalyst temperature responses of the model to those of the road test data, the comparison between model catalyst temperature and test data temperature is shown in
Figure 9d. The list of parameters, obtained from the tuning process, is shown in
Table 4.
After parameter estimation, another road test date is used to validate above engine and TWC thermal model. This road test’s vehicle speed is shown in
Figure 10a, the engine torque and motor torque are shown in
Figure 10b, the comparison between model temperature and test data temperature is shown in
Figure 10c,d. From
Figure 10c,d, the thermal model of the engine and TWC can predict the coolant and catalyst temperature well, despite some discrepancy existence.