Sie sind auf Seite 1von 13

THIS DOCUMENT IS PROTECTED BY U.S.

AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

Experimental Validation of a Dynamic Waste Heat Recovery System Model for Control Purposes
Emanuel Feru
Technische Universiteit Eindhoven

2013-01-1647
Published 04/08/2013

Frank Kupper, Chepa Rojer, Xander Seykens, Fabio Scappin and Frank Willems
TNO Automotive

Jeroen Smits
DAF Trucks

Bram De Jager and Maarten Steinbuch


Technische Universiteit Eindhoven
Copyright 2013 SAE International doi:10.4271/2013-01-1647

ABSTRACT
This paper presents the identification and validation of a dynamic Waste Heat Recovery (WHR) system model. Driven by upcoming CO2 emission targets and increasing fuel costs, engine exhaust gas heat utilization has recently attracted much attention to improve fuel efficiency, especially for heavy-duty automotive applications. In this study, we focus on a Euro-VI heavy-duty diesel engine, which is equipped with a Waste Heat Recovery system based on an Organic Rankine Cycle. The applied model, which combines first principle modelling with stationary component models, covers the two-phase flow behavior and the effect of control inputs. Furthermore, it describes the interaction with the engine on both gas and drivetrain side. Using engine dynamometer measurements, an optimal fit of unknown model parameters is determined for stationary operating points. From model validation, it is concluded that the identified model shows good accuracy in steady-state and can reasonably capture the most important dynamics over a wide range of operating conditions. The resulting real-time model is suitable for model-based control.

INTRODUCTION
Stringent CO2 emission legislation and fuel costs directed automotive industry to focus on using engine exhaust gas heat recovery systems [1]. These energy recovery systems are known as Waste Heat Recovery (WHR) systems. For heavyduty applications, the technology used in WHR systems is based on the Organic Rankine Cycle (ORC) [2], which is a suitable solution for energy recovery from a low-grade heat source such as engine exhaust gas. The potential of WHR systems was investigated in [3] for light-duty automotive applications and has been found to increase the engine efficiency by 2-3 % compared to a conventional engine. For heavy-duty applications [4], the increase in engine efficiency is even more significant, up to 4%. The net fuel economy strongly depends on the drive cycle [5] and the WHR system configuration [6]. Thermodynamic analysis on the WHR system [7], [8], demonstrate that 5% fuel saving can be achieved for a configuration where energy is recovered from the Exhaust Gas Recirculation (EGR) line only. Control of engines with a WHR system is challenging, especially for configurations where energy is recovered form the Exhaust Gas Recirculation (EGR) line as well as the exhaust line. These complex applications are characterized by multivariable control actions and strong coupling between

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

engine and WHR system on both energy level and emissions level. Accurate engine torque control is required in order to realize the required drivability while emission levels are kept low. Moreover, information on states, e.g., enthalpy and vapor fraction, is crucial for control and damage prevention. Therefore, for control development a model-based approach is essential. The work and results in the literature are mostly oriented on modeling and control of steam cycles for large-scale power plants, with considerably different configurations in comparison with the small-scale types used in automotive. In [9] and [10] a Rankine cycle system model is presented for a biomass-fired power plant. The model is intended to be used for equipment dimensioning and control design. Furthermore, in [11], [12] the focus is on ORC basic scheme modeling for energy recovery from a variable waste heat source. However, for automotive applications, only a few results are presented in the literature. Although these applications are characterized by highly transient conditions, validation results of dynamic WHR models are still lacking. This also holds for the combination of engine and WHR system. In this paper, a dynamic model of the complete WHR system is identified and validated. For control purposes, this model has to reasonably capture the real WHR system behavior and has to be computationally efficient. Compared to other work, the main contribution of this study is the combination of: Complex WHR system with two parallel evaporators, which is installed on an automotive diesel engine; Systematic identification of WHR model parameters; Model validation of steady-state as well as dynamic behavior over a wide operating range using engine dynamometer results. This work is organized as follows. First, the examined engine platform is presented. Second, the modeling of each component within the WHR recovery system is discussed. Using the Matlab/Simulink environment, the overall WHR model is implemented. Third, the model identification and validation is performed over a wide range of operating points. Finally, conclusions are drawn and directions for future research are discussed.

catalyst (AMOX) to avoid unwanted NH3 slip. To meet the upcoming CO2 targets, a WHR system is integrated with the engine.

Figure 1. Schematic representation of a heavy-duty diesel engine equipped with a WHR system. In this WHR system, the heat is recovered from both the EGR line and exhaust line using two evaporators, one for each of the exhaust circuits. The recovered exhaust heat is converted to mechanical power, by means of a piston expander [13]. The piston expander and pumps are mechanically coupled with the engine crankshaft. Thus, the recovered power is directly transmitted to the engine. The selected working fluid is pure ethanol because of its physical and thermodynamic properties, which are suitable for this kind of applications. The ethanol mass flow is controlled by two valves such that the ethanol state of matter is vapor at the evaporators outlet. The necessity of a low-level controller is crucial for correct system operation and expander damage prevention. In order to close the Rankine cycle, the ethanol is cooled back into the tank by means of a condenser. To keep the cooling capacity of the WHR system under a certain bound, an exhaust bypass valve is installed before the exhaust evaporator. This valve bypasses the exhaust gas and therefore manipulates the exhaust evaporator recovered heat. Due to drivability requirements, the power delivered by the WHR system must be interrupted temporarily, e.g., during braking or gear shifting. In this case an electric bypass valve is activated which closes the vapor flow to the expander and therefore the generated torque is reduced. For safety reasons, a 60 bar pressure relief valve is used. The electric bypass valve and the pressure relief valve are located in the valve box. The considered experimental set-up is furthermore equipped with various sensors, e.g., temperature sensors, pressure and flow sensors. These sensors are used for low-level controls, monitoring and validation of the developed WHR model.

EXPERIMENTAL SET-UP CONFIGURATION


The experimental set-up (Fig. 1) consists of a 340 kW, 12.9 liter, 6 cylinder diesel engine, which is equipped with a 2500 bar common rail fuel injection system, a cooled high pressure Exhaust Gas Recirculation (EGR) system, Variable Turbine Geometry (VTG) and Charge Air Cooling (CAC). The exhaust aftertreatment system contains a Diesel Oxidation Catalyst (DOC), a Diesel Particulate Filter (DPF), a 32.6 liter Cu-Zeolite SCR catalyst (SCR) and an Ammonia Oxidation

MODEL DESCRIPTION
An overview of the dynamic WHR model along with the measurement system is shown in Fig. 2. The components of the WHR system corresponding to Fig. 1 can be identified: tank, pumps, valves, evaporators, condenser and expander

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

models are illustrated. The piping before the evaporators and after the evaporators are represented as volumes. They are modeled as incompressible pressure volumes and compressible pressure volumes for liquid case and vapor case, respectively. Furthermore, a compressible mixing junction is used to combine the vapor generated by the two evaporators.

Pumps
The pump model is characterized by the required pumping power, a mass flow through the pump and an outlet fluid temperature. These values are computed for the pressure ratio and rotational speed of the pump. The ideal mass flow rate is computed as a function of the pump rotational speed Np and fluid density using (1) where Vd is the pump displaced volume. The fluid density = f(T, p) is a function of temperature and pressure obtained from a look-up table. The real mass flow is then computed using (2) where the volumetric efficiency ol is given by

(3) Figure 2. Schematic representation of the WHR system model with applied sensors (p - pressure, T temperature, q - mass flow). The overall WHR model is developed based on the following assumptions: maximum system pressure, i.e., the opening pressure of the pressure relief valve, is 60 bar; working fluid flow is always in the positive direction, no reversion of the working fluid is possible; working fluid in the tank has a constant temperature of 65C and constant pressure of 1 bar; pipe transport delays and pressure drops along the pipes are negligible; pressure losses, evaporation and condensation in pipe sections are negligible; In what follows, the modeling of each component within the WHR system is presented. where pi and po are the pressures at the inlet and outlet of the pump, respectively. The parameters col1, col2, col3 and col4 are determined based on experimental data obtained for different operating conditions. The pump power is calculated using

(4) where is represents the isentropic efficiency of the pump. The isentropic efficiency is introduced to account for internal pump energy loss caused by friction and external thermal loss. For simplicity, the isentropic efficiency is is considered to be constant, equal to 90 %. The outgoing working fluid temperature is computed from the energy balance equation

(5) where cp = f(Ti, pi) is the specific heat capacity of the working fluid.

Tank
The tank is modeled as a volume which connects the condenser and the pumps. The tank volume is considered large enough such that the outgoing fluid conditions are not affected by the inlet conditions. As a consequence, the tank model behaves as a fluid source with predefined outgoing pressure and temperature. For this application, the tank pressure is set equal to ambient pressure of 1 bar and fluid temperature is set to 65C.

Valves
In the overall WHR model, two types of valves are used: incompressible flow valves for flows in liquid phase and compressible flow valves for vapor and two-phase flows. The evaporators bypass valves are modeled as incompressible flow valves, while the expander bypass valve and the

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

pressure relief valve are modeled as compressible flow valves.

temperature To = f(po, ho) is extracted from the fluid property table as a function of pressure and enthalpy.

Incompressible Flow Valve


The ideal mass flow rate for a fluid in liquid state through an orifice (Fig. 3) is given by Bernoulli's equation as:

Compressible Flow Valve


Compressible fluid flows through the valves are modeled using the compressible valve flow equation (10) (6) In (10), is the compressibility coefficient defined as

Here, subscripts 1 and 2 refer to the upstream and downstream sections, respectively. (11) where = cp/c is the ration of specific heat at constant pressure and volume ratio. The parameter is determined after a comparison of the fluid pressure ratio p2/p1 with the Figure 3. Incompressible fluid valve. Practically, it is advantageous to refer to the orifice area A0 rather than to the area A2 by means of a contraction coefficient . Then, Equation (6) becomes (12) (7) The flow through the orifice determines an energy loss, entailing a drop in the fluid flow by means of a coefficient c. Let us define a discharge factor cd given by In case the pressure ratio through the valve reaches the critical pressure ratio value, supersonic flow condition is reached and Equation (11) does not depend on the pressure ratio across the valve but only on the gas properties. Accordingly, the mass flow dependency, given in (10), becomes only a function of the inlet pressure. This effect is associated with so-called chocked flow. In the compressible flow valve models, and given in Equation (11) and (12), respectively, are evaluated and the mass flow is computed by means of Equation (10). The inlet fluid enthalpy is evaluated from fluid tables at the specified inlet state variables critical pressure ratio , in order to decide whether the flow through the orifice is subsonic or supersonic, i.e.,

(8) and let DCcbp represent the duty cycle applied to the input of the bypass valve actuator. The duty cycle DCcbp is a signal between 0 % and 100 % corresponding to a closed valve and fully opened valve, respectively. Then the incompressible fluid mass flow through the valve can be expressed as (9)

(13) The characterization of the valve requires to determine the discharge coefficient cd. Such parameter can be estimated using (9) through a representative set of flow rate and pressure difference p1 - p2 measurements, at a specified DCcbp value. This procedure is performed using steady-state measurements across the valve. The thermodynamic state of the outgoing fluid is calculated assuming an isenthalpic process through the valve. Therefore, the outgoing enthalpy is equal to the inlet enthalpy, i.e. ho = hi. The outgoing where f(Ti,pi) denotes the working fluid properties which is used for both working fluid states, i.e., liquid when xi = 0 and vapor when xi = 1. The outgoing vapor fraction and enthalpy calculations arise from isenthalpic process assumption, i.e., hi = ho. The outgoing vapor quality can then be computed as

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

regions: the working fluid, the heat exchanger wall and the exhaust gas. (14) where hl(po) and h(po) are the saturated liquid enthalpy and saturated vapor enthalpy at pressure po, respectively. The outgoing flow temperature is obtained based on the fluid properties as (15) Next, the compressible flow pressure relief valve model is discussed. The pressure relief valve is used to decrease the pressure in the circuit in case it reaches a predefined level. In the model, the orifice starts to open when the upstream and downstream pressure difference reaches 60 bar. Then the orifice opening increases linearly until the maximum flow area is reached at 80 bar pressure difference. Further increase of the upstream and downstream pressure difference maintains the valve at fully opened position, as represented in Fig. 4. Figure 5. Orifice area as a function of duty cycle signal DCcbp.

Figure 6. Schematic representation of a counter-flow heat exchanger. In [14], a mathematical model of the EGR evaporator is presented. The model is given by a set of nonlinear partial differential equations which describe the conservation of mass and energy.

Figure 4. Pressure relief valve orifice area as a function of pressure difference. The orifice area is extracted from look-up tables as a function of the upstream and downstream valve pressure difference. Similar to compressible flow valve model, the outgoing fluid vapor quality and temperature are obtained using (14) and (15), respectively. In case the power of WHR system needs to be reduced rapidly, typically during gear shifting, the compressible flow electric bypass valve is used. This valve is actuated by means of a duty cycle signal DCcbp, as represented in Fig. 5. The duty cycle signal DCcbp determines the valve flow area as a percentage of the maximum orifice area.

Conservation of mass (working fluid):

(16) Conservation of energy:

(17a)

(17b) Conservation of energy at the wall:

Evaporators and Condenser


The WHR system consists of two evaporators, one on the EGR line and one downstream of the aftertreatment system, and a condenser. The thermal dynamic behavior of the WHR system is mainly influenced by the evaporators and condenser. Therefore, the modeling of these three components is crucial for the overall WHR model. A simplified schematic representation of a counter-flow heat exchanger is shown in Fig. 6. It consists of three separate enthalpy, , are the mass flow rates,

(18) where hf is the working fluid enthalpy, hg is the exhaust gas , , of the working fluid and exhaust gas, respectively. Similar to [12], the heat transfer coefficient on the working fluid side f is characterized by three heat transfer coefficients, i.e., for liquid, for two-phase and

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

for vapor. To avoid any inconsistency and negative effects in the solution process the transition between two heat exchange coefficients is performed based on a sinus interpolation function. The change in sign of equation (17b) compared to equation (17a) refers to a counter flow application. To account for phase change mathematical equations for the working fluid properties (Fig. 7) are derived.

computed enthalpies of the inlet flows, the outgoing fluid enthalpy is calculated with the energy balance equation as

(20) The outgoing vapor quality xo(po) and outgoing flow temperature To(po, ho) are computed using similar expressions to (14) and (15), respectively. In order to avoid any inconsistency, note that in (20) the outgoing mass flow is considered be to be small but different from zero for extreme cases where there is small or no flow through the system.

Pressure Volumes
The pressure volume model assumes the working fluid to be in a superheated vapor state and that it behaves as an ideal gas. The laws for the mass conservation (21a) and energy balance (21b) are then applied to this ideal gas.

Figure 7. Ethanol temperature as a function of specific enthalpy and pressure. The model given by (16), (17) and (18) is discretized with respect to time and space based on a finite difference approximation. The resulting expressions are a set of difference dynamic equations, which are used to compute the evaporator model output variables. The predicted output variables are the outlet temperatures for the exhaust gas and working fluid side. Additionally, the model outputs the vapor fraction and working fluid density. These variables are important for control of the evaporation process within the WHR system. Using similar mathematical description as for the EGR evaporator, models for the EXH evaporator and condenser are developed. The EGR evaporator, EXH evaporator and condenser parameters are identified based on the procedure given in [14].

(21a)

(21b) Next, assume the ideal gas law

(22) and define

Mixing Junctions
These components are modeled as a constant pressure volume where the conservation laws for energy and mass flow are applied. The outlet fluid thermodynamic properties are obtained by means of the mixture of several homogeneous fluids with different states and flows. The outgoing mass flow is computed through the mass flow balance equation (23) Then, equation (21a) and equation (21b) become

(24a) (19) where n is the total number of inlet fluid streams and denotes the k-th inlet fluid mass flow. For each inlet stream k, the inlet fluid enthalpy hk,i is evaluated using (13) and fluid property tables at the specified inlet state variables, i.e., temperature Tk,i, pressure pi and vapor fraction xk,i. Using the

(24b) where V is the volume of gas and R is the ideal gas constant. Note that these equations both depend on temperature and mass flow rates and for steady state conditions Equation

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

(24a) and Equation (24b) simplify to respectively.

and

and the outlet vapor quality and temperature are obtained using similar expressions to (14) and (15), respectively.

Expander
The conversion of the thermal energy to mechanical energy is performed through the piston expander which is a map based component. Mass flow rate of the outgoing fluid is calculated from a look-up table for a specified pressure ratio and expander rotational speed, i.e.,

EXPERIMENTAL RESULTS
This section is divided into three subsections. The first subsection is dedicated to data processing, which implies data filtering and energy balance check. In the second subsection, the WHR model parameters are identified for stationary operating conditions. The third subsection discusses the model validation for transient operating conditions. The WHR model identification and validation are performed using data recorded from a state-of-the-art Euro-VI heavyduty diesel engine which is equipped with a WHR system.

(25) The WHR recovery system net power output which includes the pumps power effect is given by

Data Processing
First, the experimental data is filtered using a Butterworth filter in order to reduce the noise variance in the recorded data. The Butterworth filter is chosen to be a 4th order lowpass digital filter with a cutoff frequency of 0.5 Hz. We then apply this filter in both the forward and reverse directions such that a zero-phase digital filtering of the data is obtained. Second, assuming that the system is well isolated from the exterior, the measured signals from experiments are checked if they satisfy the overall energy balance. The energy balance check is performed on both evaporators as well as the condenser. For the evaporators, the stationary energy equations for the working fluid side and exhaust gas side are given by: (29a) (29b) The inlet and outlet exhaust gas and working fluid enthalpies are obtained based on corresponding measured temperatures, exhaust gas properties and working fluid properties. In the working fluid case, enthalpy is computed using the fluid properties given in Fig. 7. In the exhaust gas case, there is a different temperature range between the EGR evaporator and EXH evaporator. Thus, the exhaust gas enthalpies are obtained using two different exhaust gas specific heat capacities cpg,1 and cpg,2, respectively. The gas specific heat capacities can be found in Table 1. By assuming an ideally isolated system, Equation (29a) should equal Equation (29b) in absolute values. However, it follows from calculations that an energy unbalance is present in both the EGR evaporator and exhaust evaporator. In the top and bottom graph of Fig. 9, this unbalance is shown in blue for the EGR evaporator and for the EXH evaporator, respectively.

(26) In Fig. 8, the WHR system net power characteristic as a function of expander rotational speed and pressure ratio is illustrated.

Figure 8. WHR system net power characteristic. The WHR system net power characteristic and mass flow are obtained from the manufacturer. Moreover, the power losses due to mechanical friction and external losses are included in the characteristic given in Equation (26). Since Equation (26) includes the absorbed power by the pumps, the expander mechanical power Pexp is computed by adding to the WHR system net power the pumps power 1 and 2, as shown in equation (27).

(27) The inlet fluid enthalpy is evaluated from fluid tables using an expression similar to (13). The outlet flow enthalpy is calculated using

(28)

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

From investigations, it follows that the exhaust gas and working fluid mass flow measurements are less accurate compared to temperature measurements.

Model Identification
In the WHR model, various parameters are used. These parameters can be divided into two categories: measurable and empirical. The measurable parameters are usually physical or geometrical parameters given by the manufacturer. This is not the case for the empirical parameters, which are obtained using a fitting procedure or routine. The similarity between measurable and empirical parameters is that once known, they remain fixed over different experimental conditions for the same system set-up. The resulting parameters are summarized in Table 1. In the sequel, the identification of the emperical parameters is discussed in more detail for the individual components.

Pump
Figure 9. Results of the absolute power balance error for the EGR evaporator (top) and EXH evaporator (bottom). An effective method to correct for these inaccuracies is to use data reconciliation techniques [15]. The data reconciliation considers the correction of measured values, such that estimates from all of the conflicting readings are obtained. It is performed by minimizing a least-squares error objective function, subject to energy conservation principles. The data reconciliation approach is then used to derive generic corrections, which lead to an improved overall energy balance over the entire operating range. For the working fluid side of the EGR evaporator and EXH evaporator, the following corrections are obtained: The pump model parameters are found by fitting the volumetric efficiency (3) as a function of pump speed and pressure ratio. In the WHR system, there are two pumps, which are assumed to have identical characteristics. Thus, both pump model parameters are obtained using the same procedure. However, the obtained parameters can sightly differ to each other according to the steady-state data obtained from the experimental set-up.

Bypass Valves
The incompressible valve model identification requires to determine the flow area A0 at fully opened valve and a discharge coefficient cd as

(32) (30) where c1, c3 are the gain and c2, c4 are the offset correction coefficients, respectively. Similarly, for the exhaust gas side of the EGR evaporator we obtain (31) where c5 and c6 are gain and offset correction coefficients, respectively. For the EXH evaporator, the applied data reconciliation method shows insignificant adjustment of the exhaust gas mass flow rate. As a result, a correction for it is not necessary. In Fig. 9, the effect of the mass flow rates corrections are plotted in green. Energy balance check is performed over ten different operating points of the engine, such that obtained mass flow corrections are valid for a broad range of working conditions of the WHR system. Note that used corrections can worsen the energy balance for some steady state points, e.g, point 1 for both evaporators and point 2 for EGR evaporator. However, it can be seen that overall, the power balance error is reduced by approximately a factor two which indicates an improvement of the mass flows rates. This function is obtained based on a set of steady-state measurements of the mass flow rate through the valve for a specific duty cycle and pressure ratio along the valve.

Expander Valve
Characterization of the compressible valve model is performed by fitting a discharge coefficient for a representative set of flow and pressure difference at a specified flow area and fluid properties.

Evaporators and Condenser


For the heat exchangers, i.e., EGR evaporator, exhaust evaporator and condenser, model identification is performed using the procedure described in [14]. The procedure consists of solving an optimization problem which minimizes the error between a set of steady-state measurements and model output using a least-squares approach. The optimization problem returns the nominal heat transfer coefficients for the exhaust gas side and working fluid side, respectively. Similar to [12], the working fluid side is characterized by three different heat transfer coefficients: for fluid phase, two phase and vapor phase, respectively. In order to make the heat

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

transfer coefficient function continuous the three identified levels are interpolated using a sinus function. The resulting heat transfer coefficient for the EGR evaporator is given in Fig. 10 as a function of vapor ratio and working fluid mass flow rate. Similarly, the heat transfer coefficients for the EXH evaporator are determined. The condenser is characterized by a single average heat transfer coefficient on the working fluid side and a heat transfer coefficient on the coolant side. The obtained parameters are given in Table. 1.

Table 1. WHR system model parameters Measurable

Figure 10. Heat transfer coefficient as a function of vapor ratio and mass flow rate.

Expander
The expander model is given by the expander mass flow rate and net power output data as a function of pressure ratio and net power output in the form of look-up tables. These look-up tables are provided by the manufacturer and are validated against measured data. A summary of the WHR model parameters along with a short description is given in Table 1. Within the empirical parameters, subscripts f, g and c are associated with the working fluid, the exhaust gas and the coolant, respectively. Furthermore, the superscripts l, tp and stand for the liquid region, two-phase region and vapor region of the evaporators, respectively. Note that the evaporators and condenser internal geometries are complex and gives us an incomplete description of the wall cross sectional areas. Thus, in Table 1 we define length L of the evaporators and condenser and select the wall cross sectional areas accordingly. However, this procedure gives a rough estimate of these parameters. Therefore, the wall cross sectional areas are finally determined by fine-tuning using transient experimental data.

Model Validation
For the WHR system model validation, the measured signals shown in Fig. 11 are used as input. These time varying signals include the expander speed, by-pass valves duty cycle, exhaust gas mass flow and temperature for the EGR evaporator and EXH evaporator. The resulting model mass flows, pressures, temperatures are compared with measurements to evaluate model prediction accuracy and decide upon model validity. Note that this set of input data gives a good indication about the WHR model validity due to wide range of operating conditions captured during the measurements. Furthermore, to examine the model dynamic behavior during a change in the operating point, a time selection is made in the right hand side of each of the following figures. The time selection is

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

chosen at the first engine speed change for a representative transient analysis (see top graph of Fig. 11).

can be seen that also the temperature after the condenser is well reproduced.

Figure 12. Working fluid mass flow through the EGR evaporator, EXH evaporator and expander. Figure 11. Experimental WHR model input signals.

Bypass Valves and Evaporators


The valve models are validated after a comparison of the experimental mass flow rates with the simulation mass flow rates. Since a mass flow sensor for the expander is not available the mass flow rate through the expander is obtained based on the sum of the two measured evaporators mass flow rates. As it can be seen from Fig. 12, similar behavior between the model and experimental data is achieved. In Fig. 13, the predicted working fluid temperatures at the evaporators outlet are shown. For certain operating points, the accuracy of the model temperatures can have degraded accuracy. The discrepancies between the experimental data and model values show the sensitivity of the model to the energy unbalance and predicted mass flow accuracy. However, there are several actions that can be taken for improving the results which are given in more detail in a section dedicated for discussions. Figure 13. Working fluid temperatures at the EGR evaporator and EXH evaporator outlet.

Expander and Condenser


Fig. 14 indicates that the model provides an accurate representation of the working fluid temperature before the expander despite reduced accuracy of the evaporators temperature for certain operating conditions. This improved accuracy follows from the mixing junction, located after the evaporators, which has an averaging effect upon the temperature before the expander (see Equation (20)). The accuracy of this temperature is important for control and expander damage prevention. Furthermore, from Fig. 14 it

Figure 14. Working fluid temperature before the expander and before and after the condenser. In Table 2, the model accuracy is quantified through the average and maximum error of predicted mass flows, temperatures and pressure in all conditions by means of steady-state analysis. The steady-state error analysis shows that the model can capture well the steady-state behavior of

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

the system. However, more significant deviation can be seen in the predicted working fluid temperature after the evaporators. These discrepancies are a consequence of the energy unbalance, mass flow sensitivity and some effects which were not considered in the computed heat transfer coefficients, e.g., pressure dependence. Table 2. Model accuracy for stationary conditions

In summary, the resulting accuracies are appropriate for model-based control; particularly, the temperature accuracy before expander, which is an important variables to be monitored for expander damage prevention. However, there is space for model improvements in steady-state and during transient conditions. These improvements can be achieved by including additional thermal effects which are not consider in the current model representation.

Figure 15. Energy flow through the EGR evaporator, exhaust evaporator and condenser.

Overall WHR System Performance


The estimated energy flows through the EGR evaporator, exhaust evaporator and condenser are shown in Fig. 15. These energy flows are of interest, since it is important to have knowledge about the amount of heat present in the WHR system. Using this energy flow information the cooling capacity of the WHR system can be tracked in such a way that the engine cooling capacity in not affected. In Fig. 15 the calculated energy flows are obtained using a combination of measurements and fluid properties. This calculation is required since a sensor to directly measure energy flows within the system is not available. In Fig. 16, the net WHR system power and efficiency are illustrated. The calculated net power and efficiency are obtained using the expander maps with measurements as input. Then the calculated values are compared with the ones coming from the model. As expected a good prediction of the vapor temperature and pressure before the expander, within 2 % (see Table. 2), provides a good estimate of the expander generated power. The power and efficiency predictions are important for a possible high-level torque control implementation and maximization of the amount of heat recovered by the WHR system, respectively.

Figure 16. WHR system power and efficiency. The results shown in this section highlight that the WHR system output signals are reasonably well captured by the model. Mass flow and energy flow dynamics are reproduced correctly over the operating range of interest for control purposes. However, the evaporators thermal behavior can be improved for a better performance of the overall system.

Discussion
From the model validation, it is seen that there can still be significant differences between model and experimental results in stationary conditions, see also Table 2. Partly, this error is caused by the applied measurement corrections; as illustrated in Fig. 9, this correction is not perfect and an energy unbalance will still occur. Various component inaccuracies will also affect the overall WHR system result. Furthermore, the evaporator outlet temperature is found to be a function of pressure, see Fig. 17 for the EXH evaporator. This is not captured by the currently applied WHR system

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

model. Including a pressure dependency in the heat transfer coefficients could overcome this issue. Furthermore, on one hand, the geometry of the evaporators and condenser which can influence the steady state results as well as dynamic results is simplified. A complex geometry dictated by the design leads to differences between the heat transfer in the working fluid side and that on the exhaust gas side. In practice, these effects can be taken into account by means of geometrical correction factors. On the other hand, errors can also result from the tank model which in current representation is ideal, i.e., constant temperature and pressure. In real world, the tank temperature is not necessarily constant, but it depends on the condenser cooling capacity.

moderate: up to 8%. As seen from the error analysis, main model improvements we recommend include: pressure dependency of the heat transfer coefficients wall specific heat capacity as a function of temperature thermal model of the tank expander rotational inertia effects The validated WHR system model can run in real-time, such that it is suitable for control design and optimization. Besides WHR system model improvements, future research will focus on the development of a low-level WHR control strategy and on the extension of the supervisory control strategy developed in [16], which combines energy and emission management.

ACKNOWLEDGEMENTS
This research is done in the framework of the FERVENT research program (Further Emission Reduction, Vehicle Efficiency gains and Neutral Thermal loading). This program is partially funded by the Dutch Ministry of Economical affairs. Figure 17. EXH evaporator outlet temperature as a function of pressure. For automotive applications, dynamic behavior is important. It is seen that errors in dynamic model predictions can be significant. First, the corrections computed in steady-state will be inaccurate for transient conditions. Second, also model inaccuracies related to the wall heat capacity become prone in dynamic conditions. Based on the applied materials, it is suggested to implement a temperature dependent wall specific heat capacity. Improvements of the dynamical results can also be obtained by considering the expander rotational inertia effects which can induce changes in the mass flow rate through the expander as well as generated power during transients. Additionally, the implementation of dynamic parameter identification procedures could considerably improve the WHR model behavior during transients.

CONTACT INFORMATION
Emanuel Feru Eindhoven University of Technology P.O. Box 513 5600 MB Eindhoven The Netherlands e.feru@tue.nl Tel: +31 40 247 2072

REFERENCES
1. Hounsham, S., Stobart, R., Cooke, A., and Childs, P., Energy Recovery Systems for Engines, SAE Technical Paper 2008-01-0309, 2008, doi: 10.4271/2008-01-0309. 2. Teng, H., Regner, G., and Cowland, C., Achieving High Engine Efficiency for Heavy-Duty Diesel Engines by Waste Heat Recovery Using Supercritical Organic-Fluid Rankine Cycle, SAE Technical Paper 2006-01-3522, 2006, doi: 10.4271/2006-01-3522. 3. Edwards, K., Wagner, R., and Briggs, T., Investigating Potential Light-duty Efficiency Improvements through Simulation of Turbo-compounding and Waste-heat Recovery Systems, SAE Technical Paper 2010-01-2209, 2010, doi: 10.4271/2010-01-2209. 4. Teng, H. and Regner, G., Improving Fuel Economy for HD Diesel Engines with WHR Rankine Cycle Driven by EGR Cooler Heat Rejection, SAE Technical Paper 2009-01-2913, 2009, doi: 10.4271/2009-01-2913.

CONCLUSIONS
In this paper, a dynamic WHR system model was identified and validated for a heavy-duty automotive application. This model is based on first principles modelling and on empirical correlations from the manufacturer. Using engine dynamometer data, it is shown that the WHR model has good accuracy for stationary operating points: average modelling error between 2% and 6%. However, the prediction accuracy of the evaporators models within the WHR system is

THIS DOCUMENT IS PROTECTED BY U.S. AND INTERNATIONAL COPYRIGHT It may not be reproduced, stored in a retrieval system, distributed or transmitted, in whole or in part, in any form or by any means. Downloaded from SAE International by Oxford Brookes Univ, Monday, July 29, 2013 03:32:15 PM

5. Howell T.. Development of an ORC system to improve HD truck fuel efficiency. Technical Report RD.11/353805.1., DEER Conference, Ricardo Inc, October 2011. 6. Nelson C.. Exhaust Energy Recovery. Technical Report DE-FC26-05NT42419, DEER Conference, Cummins, August 2009. 7. Park, T., Teng, H., Hunter, G., van der Velde, B. et al., A Rankine Cycle System for Recovering Waste Heat from HD Diesel Engines - Experimental Results, SAE Technical Paper 2011-01-1337, 2011, doi: 10.4271/2011-01-1337. 8. Teng, H., Klaver, J., Park, T., Hunter, G. et al., A Rankine Cycle System for Recovering Waste Heat from HD Diesel Engines - WHR System Development, SAE Technical Paper 2011-01-0311, 2011, doi: 10.4271/2011-01-0311. 9. Colonna P. and van Putten H.. Dynamic modeling of steam power cycles. Part I - Modeling paradigm and validation. Applied Thermal Engineering, 27(2-3):467-480, 2007. 10. van Putten H. and Colonna P.. Dynamic modeling of steam power cycles. Part II - Simulation of a small simple Rankine cycle system. Applied Thermal Engineering, 27(14-15):2566-2582, 2007. 11. Wei D., Lu X., Lu Z., and Gu J.. Dynamic modeling and simulation of an organic rankine cycle (orc) system for waste heat recovery. Applied Thermal Engineering, 28(10): 1216-1224, 2008. 12. Quoilin S.. Sustainable Energy Conversion Through the Use of Organic Rankine Cycles for Waste Heat Recovery and Solar Applications, Ph.D. Thesis. University of Liege, 2011. 13. Badami M. and Mura M.. Preliminary design and controlling strategies of a small-scale wood waste Rankine Cycle (RC) with reciprocating steam engine (SE). Energy, 34(9):1315-1324, 2009. 14. Feru E., Willems F., Rojer C., de Jager B., and Steinbuch M.. Heat Exchanger Model Identification for Control of Waste Heat Recovery System in Diesel Engines. Accepted to 2013 American Control Conference, Washington DC, USA. 15. Bagajewicz M. J. and Cabrera E.. Data reconciliation in gas pipeline systems. Ind. Eng. Chem. Res., 42(22): 5596-5606, 2003. 16. Willems F., Kupper F., and Cloudt R.. Integrated Powertrain Control for optimal CO2-NOx tradeoff in an
The Engineering Meetings Board has approved this paper for publication. It has successfully completed SAE's peer review process under the supervision of the session organizer. This process requires a minimum of three (3) reviews by industry experts. All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or otherwise, without the prior written permission of SAE. ISSN 0148-7191

Euro-VI diesel engine with Waste Heat Recovery system. In Proceedings of the 2012 American Control Conference, pages 1296-1301, Montreal, Canada, 2012.

Positions and opinions advanced in this paper are those of the author(s) and not necessarily those of SAE. The author is solely responsible for the content of the paper. SAE Customer Service: Tel: 877-606-7323 (inside USA and Canada) Tel: 724-776-4970 (outside USA) Fax: 724-776-0790 Email: CustomerService@sae.org SAE Web Address: http://www.sae.org Printed in USA

Das könnte Ihnen auch gefallen