Vous êtes sur la page 1sur 10

A Dynamic Model of an Electropneumatic Valve Actuator for Internal Combustion Engines

Jia Ma
Delphi Powertrain Systems, Auburn Hill, MI 48326 e-mail: jia.ma@delphi.com

Guoming G. Zhu
e-mail: zhug@egr.msu.edu

Harold Schock
e-mail: schock@egr.msu.edu Michigan State University, East Lansing, MI 48824

This paper presents a detailed model of a novel electropneumatic valve actuator for both engine intake and exhaust valves. The valve actuators main function is to provide variable valve timing and variable lift capabilities in an internal combustion engine. The pneumatic actuation is used to open the valve and the hydraulic latch mechanism is used to hold the valve open and to reduce valve seating velocity. This combination of pneumatic and hydraulic mechanisms allows the system to operate under low pressure with an energy saving mode. It extracts the full pneumatic energy to open the valve and use the hydraulic latch that consumes almost no energy to hold the valve open. A system dynamics analysis is provided and followed by mathematical modeling. This dynamic model is based on Newtons law, mass conservation, and thermodynamic principles. The air compressibility and liquid compressibility in the hydraulic latch are modeled, and the discontinuous nonlinearity of the compressible ow due to choking is carefully considered. Provision is made for the nonlinear motion of the mechanical components due to the physical constraints. Validation experiments were performed on a Ford 4.6 l four-valve V8 engine head with different air supply pressures and different solenoid pulse inputs. The simulation responses agreed with the experimental results at different engine speeds and supply air pressures. DOI: 10.1115/1.4000816 Keywords: automotive systems, powertrain systems, camless valve actuation, hydraulic and pneumatic system model

Introduction

In a camless valvetrain, the motion of each valve is controlled by an independent actuator. There is no camshaft or other mechanisms coupling the valve to the crankshaft as in a conventional valvetrain. This provides the possibility to control the valve events, i.e., timing, lift, and duration, independent of crankshaft rotational angle. Various studies have shown that an engine with variable valve actuation reduces pumping losses, adjusts the cycle-to-cycle internal residual gas recirculation RGR, and reduces nitrogen oxide NOx emissions with improved performance over a wide operating range. A signicant amount of research has been contributed to demonstrate the advantage of variable valve actuation VVA over the traditional cam-based valvetrain for both gasoline and diesel engines. The investigation of intake valve timing control of a spark ignited SI engine was conducted in Ref. 1. It was found that at low and partial load conditions, engine pumping loss was reduced by 2080% due to throttless operation. Fuel consumption was improved up to 10% at idle. Through simulation and experiments, Negurescu et al. 2 showed that SI engine efciency can be improved up to 29% due to variable valve timing VVT, compared with a classic throttled engine. The engine torque output was also improved by up to 8% at low speed with wide open throttle. Research carried out in Ref. 3 demonstrated how VVT and variable valve lift VVL affect the partial load fuel economy of a light-duty diesel engine. In this study, the indicated and brakespecic fuel consumptions were improved up to 6% and 19%, respectively. The operation of an OttoAtkinson cycle engine by late intake valve closing to have a larger expansion ratio than compression ratio was studied in Ref. 4. A signicant reduction
Contributed by the Dynamic Systems Division of ASME for publication in the JOURNAL OF DYNAMIC SYSTEMS, MEASUREMENT, AND CONTROL. Manuscript received November 15, 2008; nal manuscript received November 20, 2009; published online February 3, 2010. Assoc. Editor: Bin Yao.

of carbon monoxide CO and NOx emissions was obtained. Urata et al. 5 also showed that the operational range of a homogeneously charged compression ignition HCCI engine can be expanded to both high and low load ranges through the adoption of VVT and VVL. The advantages of VVT and VVL engines lead to combustion optimization over a broad engine operational range. For example, Trask et al. 6 developed the VVT and VVL optimization methodologies for an I4 2.0 l camless ZETEC engine at various operational conditions including cold starts, cylinder deactivation, full load, idle, and transient operations. Three primary types of camless valve actuators are electromagnetic, hydraulic, and pneumatic actuators. Sugimoto et al. 7, Theobald et al. 8, and Pischinger and Kreuter 9 presented the results of electromagnetic actuators. A hydraulic actuator was discussed in Ref. 10. A pneumatic actuator incorporated with a permanent magnet control latch was presented in Ref. 11. The advantages and disadvantages of a pneumatic actuator over a hydraulic actuator were addressed in Ref. 12, where a pneumatic valve actuator with a physical motion stopper was presented and the simulations of the valve actuation system were shown. Implementation of various camless valve actuators was studied in 13. In order to provide an insight to the pneumatic actuator design and its control requirements, mathematical modeling was performed to the engine and its various actuation systems. A variable valve timing engine was modeled in Ref. 14, along with its engine control strategy. Tessler et al. 15 analyzed and modeled the dynamics of a pneumatic system consisting of a double-acting or single-acting cylinder and servo valve. A mathematical model of a pneumatic force actuator was presented in Ref. 16. In this article, an electropneumatic valve actuator EPVA is employed to replace the traditional camshaft in an internal combustion engine. The EPVA is capable of varying valve lift, timing, and opening duration as desired in a variable valve timing engine. Different from the pneumatic valve discussed in Ref. 12, the EPVA is designed to extract the maximum work from the air ow by incorporating a hydraulic latch mechanism to hold valve open MARCH 2010, Vol. 132 / 021007-1

Journal of Dynamic Systems, Measurement, and Control Copyright 2010 by ASME

Downloaded 27 Feb 2010 to 220.225.248.2. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 1 System dynamics at the air charging stage

Fig. 2 System dynamics at the expansion and dwell stage

and to reduce the power consumption; a hydraulic damper mechanism is also enabled to produce a desirable slow and smooth seating velocity when the valve returns to its seat. It has been shown that less than 4 kW power was required at 6000 rpm to run a 16 valve 2 l turbocharged four-cylinder engine 17. This is well in line with the electromagnetic valve systems and well below the hydraulic ones. Nonlinear mathematical model was developed in this paper to help establish the design and control criteria and to be used as a base to develop a control oriented model required for model-based control strategy development. This paper is organized as follows. System dynamics is described in Sec. 2, and a mathematical model is developed component by component in Sec. 3. The experimental validation of the developed model is provided in Sec. 4. Finally, conclusions are drawn in Sec. 5.

now charges the cylinder, the actuator piston starts moving down and opens the poppet valve. Although the right side of the outlet port valve is subject to the high pressure air, it remains closed due to the area difference between the two sides of the port valve. The on and off valve S1 is deactivated at the moment when solenoid 1 is energized. This only allows the oil to ow down through the check valve parallel to S1 and prevents the oil from returning to the reservoir. Note that the oil supply pressure is the same as the air supply pressure. 2.2 Expansion and Dwell Stage. In the expansion and dwell stage as shown in Fig. 2, both solenoids 1 and 2 are energized. The time delay between the activation of two solenoids is usually chosen between 2 ms and 5 ms depending on the desired valve lift height. The spool valve 2 is pushed slightly to the left so that the high pressure air can be sent to the left of the inlet port valve through the same spool valve. The on and off valve S2 is closed at the same time when solenoid 2 is energized to prevent the high pressure air from escaping to the atmosphere through the rst spool valve. The inlet port valve is closed mainly due to the spring force applied on the inlet port valve, which cuts off the air supply to the piston cylinder. Meanwhile, solenoid 1 remains energized; therefore, the outlet port valve remains closed. The air that was drawn into the actuator cylinder during the previous air charging stage is able to expand completely. The actuator piston and poppet valve both reach their maximum displacement. The high pressure oil dark gray trapped in the hydraulic latch on and off valve S2 remains closed balances the valve spring force and keeps the poppet valve open at its maximum lift height. This is called the energy saving mode. It allows the system to extract the full expansion work from the air, which has entered the cylinder without losing the capability of varying the valve open duration. 2.3 Air Discharging Stage. In the air discharging stage, the air leaves the actuator cylinder and the valve returns to its seat. As displayed in Fig. 3, both solenoids are de-energized. Conse-

System Dynamics

The EPVA consists of two solenoids, two spool valves, two port valves, an actuator piston, an actuator cylinder, and a hydraulic latch-damper system. An actuator piston pushes the back of the engine poppet valve stem, causing the valve to open. Solenoidcontrolled spool valves are used to control the ow of the air that enters and exits the actuator cylinder. In order to reduce the energy consumption, EPVA uses a hydraulic latch, which allows the actuator to extract the full expansion work out of the air that is drawn into the actuator cylinder. Meanwhile, the actuator is still capable of holding the valve open over the desired opening duration. A hydraulic damping mechanism is added to provide a soft seating velocity for the valve. According to the events taking place in the actuator cylinder, the system dynamics are divided into three stages: air charging, expansion and dwell, and air discharging stages. Figure 6 illustrates three stages over the valve lift prole. 2.1 Air Charging Stage. Figure 1 depicts the system dynamics when the actuator cylinder is operated at the air charging stage. The light gray color represents the high pressure supply pressure air, the white color represents the low pressure atmospheric pressure air, and the dark gray color represents the oil in the hydraulic latch damper. S1 and S2 are two on and off valves controlled by solenoids 1 and 2 through their mechanical linkages to the corresponding spool valves 1 and 2, respectively. Each controlled on and off valve is parallel to a check valve, see Fig. 1. When a solenoid is energized, its corresponding on and off valve is closed, and when the solenoid is deactivated, the on and off valve is tune on to allow two-way ow. An energized solenoid is shown in black while a de-energized one is transparent. During the charging stage, solenoid 1 is energized pushing the spool valve 1 slightly to the right. As a result, the high pressure air is sent to two locations, the left of the outlet port valve and the right of the inlet port valve. The left side of the inlet port valve is subject to the low pressure. Therefore, the high pressure air closes the outlet port valve and opens the inlet port valve. The supply air 021007-2 / Vol. 132, MARCH 2010

Fig. 3 System dynamics at the air discharging stage

Transactions of the ASME

Downloaded 27 Feb 2010 to 220.225.248.2. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

constant. The difference between the cylinder pressure and the supply pressure decreases as the pressure in the cylinder chamber builds up over time. The air then becomes unchoked and ows through the inlet with decreasing pressure. The ow exiting the outlet switches between a choked and an unchoked pattern as well for the same reason. This discontinuous nonlinearity of the ow has to be taken into consideration in the actuator piston model. As shown in Fig. 4, considering the control volume above the actuator piston in the cylinder chamber including the inlet and outlet, the rst law of thermodynamics can be written as

E v2 v2 W +m i hi + i m e he + e = Q 2 2 t

and W are the heat transfer rates into the control volume where Q and the work rate delivered by the control volume to the actuator e are the mass ow rates entering i and m piston, respectively; m and h are the enthalpies of the and exiting the control volume; h i e gas entering and exiting the cylinder chamber; and nally, E / t is the rate of change of the control volume total energy. on the actuator piston by the The rate of the work done W control volume is =A P y W p p 2

Fig. 4 Actuator piston model

quently, both on and off valves, S1 and S2, are activated. Both air and the oil ows are able to travel in two directions. Since both solenoids are off, the springs inside the two spool valves return the spools to their original positions. The high pressure air is then sent to both sides of the inlet port valve. The port valve remains closed mainly due to the spring force. Meanwhile, the low pressure air is applied on both sides of the outlet port valve. Since the oil in the hydraulic latch is now able to ow back up to its reservoir, the actuator piston returns to its rest position due to the valve spring force. The spring force pushes the actuator piston as well as the poppet valve back and the volume of the air in the actuator cylinder is then reduced. This results in an increase in the air pressure in the actuator cylinder and an increase in the air pressure at the right side of the outlet port valve. Therefore, the outlet port valve is pushed open, the air in the actuator cylinder is able to discharge, and its pressure decreases immediately. The poppet valve continues on its return course. The hydraulic damper activates when the poppet valve moves close to its seat. Due to the decreasing ow area where the oil leaves the passage, the velocity of the valve is reduced greatly, providing a smooth return.

where A p is the area of the actuator piston, P p is the pressure of the control volume the pressure applied on the actuator piston, is the velocity of the actuator piston movement. and y The supply air entering the cylinder chamber from the inlet can be viewed as a gas from a reservoir. The gas in the reservoir has zero velocity; therefore, its enthalpy is the stagnation enthalpy of the inlet supply air hin. Modeling the air leaving the cylinder chamber is simplied by considering the case that the pressure is balanced and the dynamics during the pressure balancing is ignored. In this case, the air leaving the cylinder chamber from the outlet can be viewed as a gas leaving a reservoir, which is the control volume inside the chamber. Hence, the enthalpy of the air leaving the chamber can be represented by the stagnation enthalpy of the air in the actuator cylinder h p: hi +
vi2 = hin = C pTin 2 v2 e = h p = C pT p 2

he +

Treating air as an ideal gas leads to P p = pRT p Replacing T p in Eq. 4 by P p / R p results in he +


v2 CpPp e = 2 R p

Mathematical Modeling

The purpose of this section is to derive governing equations of the individual components of the pneumatic-hydraulic valve actuator, which consists of the actuator piston, the hydraulic latch damper, the inlet and outlet port valves, two solenoids, and two spool valves, as displayed in Fig. 1. These equations were used to model the behavior of the valve under different sets of operating conditions. 3.1 Actuator Piston. In this subsection, energy conservation, mass conservation, and Newtons second law were used to determine the following variables: the rate change of the gas pressure , the rate change of the gas inside of the cylinder chamber P p , see Fig. 4. p, and the acceleration of the actuator piston y density A sudden reduction in pressure occurs at the inlet port when it opens. This causes the air ow to expand in an explosive fashion. The ow is choked and the pressure at the port is assumed to be Journal of Dynamic Systems, Measurement, and Control

where Tin and C p are the temperature of the air at the inlet, which equals to the ambient temperature Tatm = 295 K, and the specic heat of the air at constant pressure, respectively; R and p are the gas constants of the air and the density of the air in the cylinder chamber above the actuator piston; and P p and T p are the pressure and temperature of the air in the cylinder chamber above the actuator piston. In order to derive the equations for the mass ow rate when the air ow enters the inlet or leaves the outlet, two cases are considered, and they are choked and unchoked gas ows. The proof of the derivation of the mass ow equation is shown in Ref. 15. The following are the two main assumptions: a the gas ow in = 0 and b the ow is isentropic the valve actuator is adiabatic Q everywhere except across normal shock waves. Also a term pro is subtracted from the total power that is delivered portional to W MARCH 2010, Vol. 132 / 021007-3

Downloaded 27 Feb 2010 to 220.225.248.2. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

to the actuator piston to compensate the heat loss 16. The ow pattern at the inlet depends on the cylinder pressure P p and the supply pressure Psupply. Applying the compressible ow equation yields the expression for the mass ow rate at inlet i as follows: m i = in m

= 1 C A w P 3 P p din in supplyin k RTin CdoutAoutzout A py pk P py y


k3 P3 p p 18

k PsupplyAin RTin

For the unchoked case, where P p 0.53Psupply based on the results from Refs. 15,16,

in =


2 Pp k 1 Psupply
k+1/2k

Pp

1k/k

Psupply

1/2

where Cdin and Cdout are the ow discharge coefcients at the inlet and outlet, respectively; p is multiplied by the rate change of as it is assumed that part of the work is dissipated as heat work W loss from the system. Parameter p is chosen to be between 0 and 1 depending on the actual heat loss during the process. This formulation is studied in Ref. 16. Applying the law of mass conservation to the control volume above the actuator piston in the cylinder results in im e = A p p y + py m 19 e by Eqs. 7 and 11 leads to the expression i and m Replacing m p, for p = 1 CdinAinw Psupplyin A py py y

and for the choked case, where P p 0.53Psupply,

in = 0.58

Note that k = C p / Cv is the specic heat ratio, where Cv is the specic heat of air at constant volume. Ain is the area of the inlet. Since the port valves open and close at a very fast rate, the effective ow area Ain is approximated by Ain =

r2 1, w 0 , 0, w = 0

k CdoutAoutzoutkP p p RTin

10

20

where r1 is the inner radius of the inlet port valve and w is the displacement of the inlet port valve, see Fig. 7. The mass ow rate e equation can be derived as follows: m e = out m

can be obtained by Now the acceleration of the actuator piston y applying Newtons second law, see below: + Cfy + K py + p = A p P p + AcapPoil A p Acap Patm My M = M piston + M valve + 3 M spring + M cap
1

k P pAout RT p
1k/k

11

21

For the unchoked case, where Pout 0.53P p,

out =


2 Pout k 1 Pp
k+1/2k

Pout Pp

1/2

12

and for the choked case, where Pout 0.53P p,

out = 0.58

13

Here, Aout is the area of the outlet. It follows the same expression as Ain except that it is dependent on z. The Aout expression can be approximated as follows: Aout =

where M piston, M valve, M spring, and M cap are the mass of the actuator piston, intake valve, valve spring the effective spring mass equals one-third of the total spring mass, see Ref. 18, and the cap on the top of the valve stem; Acap is the area of the cap on the 2 top of the actuator piston stem; A p = r2 p roil is the actuator piston area, where r p and roil are the radius of the actuator piston and oil passage; C f is the damping coefcient approximating the energy dissipation due to the friction; and nally K p and p are the stiffness and preload of the valve spring. Rearranging Eq. 21 results in = y 1 K py + p A p P p + AoilPoil A p + Aoil Patm C f y M 22 Note that for exhaust valves, there is an additional force F p due to in-cylinder pressure: F p = AexhPcyl 23 where Aexh is the area of the exhaust valve and Pcyl is the incylinder pressure. Incorporating F p into Eq. 22, the acceleration becomes of the actuator piston y = y 1 K py + p A p P p + AoilPoil F p A p + Aoil Patm C f y M 24 3.2 Hydraulic Latch and Damper. Another mechanism that has a direct impact on the dynamics of the actuator piston is the hydraulic latch damper. The compressibility of the uid in the hydraulic latch is considered and the mechanism of adjusting the valve seating velocity is modeled in detail. Figure 5 illustrates the hydraulic latch function. The oil sits on the top of the actuator piston stem with the same supply pressure as the air pressure. Fluid enters or exits through area Aoilin / Aoilout. When the air, drawn in during the air charging stage, is fully expanded in the actuator cylinder, the actuator piston reaches to its maximum displacement. The on and off valve S1 is closed due to Transactions of the ASME

r2 2, 0,

z0 z=0

14

where r2 is the outer radius of the outlet port valve and z is the displacement of the outlet port valve, see Fig. 8. The rate change of the total energy of the control volume is the summation of the rate change of the internal energy, the kinetic energy and the potential energy. The kinetic and potential energies of the control volume are negligible. Hence, the rate change of the total energy is approximated as the rate of change of the internal energy, that is,

E U d = = mCvT p t t dt

15

where m is the mass of air inside the control volume and Cv is the is specic heat of air at a constant volume. The expression for m = pA p y m Expanding Eq. 15 and using Eqs. 16 and 5 result in 16

E A pC v y +P = P py p t R

17

can be derived by substituting Eqs. 2, 3, The expression for P p 6, 7, 11, and 17 into Eq. 1. That is, 021007-4 / Vol. 132, MARCH 2010

Downloaded 27 Feb 2010 to 220.225.248.2. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 5 Hydraulic latch/damper model

activating solenoid 1 to prevent the oil from returning, recalling system dynamics at the air charge, expansion and dwell stages. The pressurized oil is trapped in the passage and keeps the actuator piston at the maximum displacement until solenoid 1 is turned off, see Sec. 2.3. Hence, this hydraulic latch provides a holding force to keep the valve opening. Another function of this mechanism is to provide a low seating velocity for the valve. When the actuator piston approaches the original position, the cap on the top of the stem will partially block the exit area A. The actuator piston encounters a large resistant force due to the reduced ow area, which decreases the air ow velocity tremendously. The resistant force increases as the exit area reduces. Figure 6 shows a valve lift prole along with the solenoid action chart. The solenoid itself has about 23 ms electromagnetic delay upon activation. These delays were not shown in this chart. As was explained earlier, one valve cycle consists of three stages: air charging, expansion and dwell, and air discharging stage. So-

lenoid 1 is turned on at the beginning of air charging stage and turned off at the end of expansion and dwell stage. Solenoid 2 is turned on before expansion and dwell stage. Solenoid 2 operates at the same frequency and duty cycle as these of solenoid 1 with a time delay. Both inlet and outlet are closed during the overlap of solenoids 1 and 2. The oil is modeled as an incompressible ow at air charge and discharge stages due to relatively low oil pressure, while in the dwelling region, it is modeled as a slightly compressible ow under high oil pressure. The slight compressibility is what causes the volume change of the oil in the passage, hence the swing on the top of the valve lift prole. Apply the incompressible ow model at air charging stage to calculate the ow rate through the oil passage as follows:

qoil = CdoilinAoilin

Psupply Poil = Acapoily oil

25

Fig. 6 Valve lift prole with the solenoid action chart

Journal of Dynamic Systems, Measurement, and Control

MARCH 2010, Vol. 132 / 021007-5

Downloaded 27 Feb 2010 to 220.225.248.2. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 7 Inlet port valve model

Fig. 8 Outlet port valve model

Therefore, the pressure of the oil at air charging stage is Acapy Poil = Psupply CdoilinAoilin

oil

26

where qoil and Cdoilin are the volumetric ow rate of the uid and the discharge coefcient as the uid enters the passage, respectively; Aoilin and oil are the area where the uid enters the passage it will be calculated later and the density of the uid; Psupply and Poil are the supply and oil pressures, where Psupply = Poil. Apply the compressible ow model at expansion and dwell stages to derive the oil pressure. The state equation PVc = K = const is used here by choosing c large enough to represent the high level of incompressibility: PoillockVc = PiVic 27 Note that Eq. 27 is true only for an isothermal process. Substituting V = Acapy and Vi = Acapy i into Eq. 27 to obtain Poillock = Piy ic yc 28

air ow from the spool valve with pressure PcupR and the supply pressure Psupply. Pressure PcupR alternates between atmosphere and supply pressure, which is regulated by the spool valve. The port valve remains closed when PcupR equals Psupply due to the spring force and the valve opens when PcupR decreases to the atmosphere pressure. The supply air is treated as a stagnant ow with constant pressure. The equation of motion can be obtained using Newtons second law: + CcRw + KcRw = Psupply PcupRAcR mcRw AcR = r2 2. 32 where 0 w wmax and r2 is the outer radius of the inlet and outlet port valve see Fig. 4. Variables mcR and CcR are the mass of the inlet port valve and the damping coefcient compensating for the friction loss of the valve, and KcR is the spring , and w are the displacement, the velocconstant. Parameters w, w ity, and the acceleration of the inlet port valve; and wmax is the maximum distance, which the inlet port valve is allowed to travel. The discontinuous nonlinearity in the port valve dynamics caused by this physical limitation was considered. Rearranging Eq. 32 , leads the expression for w = w 1 KcRw AinletPsupply PcupRAcR CcRw mcR 33

where Poillock is the pressure of the oil at the expansion and dwell lock stages, y i is the maximum valve displacement, Vi is the volume of the uid at y i, and Pi is the oil pressure Poil at y i. Similarly, the equation of motion at air charging stage was obtained using the incompressible ow model as follows: qoil = Cd_oiloutAoilout

Poil Psupply = Acapy oil

29

Rearranging Eq. 29 results in Poil = Psupply +

Acapy Cd_oiloutAoilout

oil

30

3.4 Outlet Port Valve. The outlet port valve functions in a similar way to the inlet port valve, except that the air that pushes the port valve open is with the actuator cylinder pressure. The pressure in the actuator cylinder is unsteady; thus, the ow dynamic behavior is modeled. The modeling process is similar to the actuator piston. The control volume used here is shown in Fig. 8. Applying conservation of energy as shown in Eq. 1, we evalu , E / t, h + v2 / 2, m i, h e + v 2 e as follows: ate W i i e / 2, and m =A P z W cL out and AcL = r2 2 34

where Cd_oilout is the discharge coefcient as the uid exits the passage and Aoilout is the area where the uid exits the passage Aoilin = Aoilout = A, where A=

y 2 2

2 2rpass + 2

Aoil Acap,

y p1 y p1

+ Aoil Acap,

where Pout is the pressure on the outlet port valve in the control volume. Then,

31

E U d AcLCv y +P = = mCvTout = Pouty out t t dt R

35

The variables rpass, Aoil, Acap, and p1 are shown in Fig. 6. The seating velocity is greatly reduced when the stem enters the area where y p1. By adjusting p1 during the hydraulic latch design process, the slope of the response can be altered by adjusting the timing of entering the region where y p1. 3.3 Inlet Port Valve. As illustrated in Fig. 7, the inlet port valve is modeled as a mass-spring-damper system driven by the 021007-6 / Vol. 132, MARCH 2010

where z is the displacement of the outlet port valve, Tout is the gas temperature in the control volume, and Pout is the gas pressure in the control volume. The ideal gas law, Eq. 5, is used to derive Eq. 35. Treating the air ow from the actuator cylinder as stagnant ow leads to hi +
vi2 CpPp = h p = C pT p = 2 pR

36

Transactions of the ASME

Downloaded 27 Feb 2010 to 220.225.248.2. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 9 Spool valve model

i = inL m

k P pAout = AoutinLk p P p RT p

37

im e = AcLoutz + outz m

47

where Aout is the inlet area of the control volume. Similar to the evaluation of Ain in Eq. 14, the equivalent Aout can be approximated by the following equation: Aout =

e using Eq. 42 into the i using Eq. 37 and m Substituting m out can be written as below: equation above, out = 1 CdinLAoutzinLk p P p CdoutLALzoutLkPoutout AcLz outz z 48

r2 2, 0,

z0 z=0

38

Similar to the actuator piston modeling in Sec. 3.1, for the unchoked case, where Pout 0.53P p,

inL =


2 Pout k 1 Pp
k+1/2k

Pout Pp

1k/k

1/2

Finally, Newtons second law yields the equation of motion for the outlet port valve, + CcLz + KcLz = PoutAoutlet PcupLAcL mcLz 0 z zmax, Aoutlet = r2 1, AcL = r2 2. 49

39

and for the choked case where Pout 0.53P p,

inL = 0.58
Treating ambient air as stagnant ow results in he + e = outL m
v2 e = hatm = C pTatm 2

40

41 42

k PoutAL = ALoutLkoutPout RTout

and Parameters mcL and where CcL are the masses of the outlet port valve and the damping coefcient compensating for the friction loss of the valve, respec , and z are the tively, and KcL is the spring constant. Parameters z, z displacement, the velocity, and the acceleration of the outlet port valve, and zmax is the maximum distance, which the outlet port valve is allowed to travel. The discontinuous nonlinearity in the port valve dynamics was considered in the simulation. Rearrang results in ing Eq. 49 to obtain an expression for z = z 1 KcLz AoutletPout PcupLAcL CcLz mcL 50

where Tout is the temperature of the gas in the control volume, Pout is the inlet air pressure of the control volume, out is the density in the control volume, and AL is the outlet area of the control volume and also a function of geometry and the displacement of the outlet port valve, where A L = 2 r 1z If Patm 0.53Pout, for the unchoked case, 43 Patm Pout
1k/k

All the discharge coefcients that are involved in the ow equations were determined numerically and experimentally. 3.5 Spool Valve. The armature of the solenoid pushes the stem of the spool valve with the magnetic force Fs when the solenoid is energized and a precompressed spring returns the spool valve when the solenoid is de-energized. The spool valve is pressure balanced at two ends, as shown in Fig. 9. The spool valve equation of motion is + C sx + K s x + s = F s, mspoolx 0 x x0 51

outL =


2 Patm k 1 Pout
k+1/2k

1/2

44

and if Pout 0.53P p, for the choked case

outL = 0.58

45

Here, the gas was assumed ideal and the nonlinearity of the ow was considered in Eqs. 39, 40, 44, and 45. One can obtain in the following form by substituting Eqs. 3445 into Eq. P out as it was treated in the actuator piston = W 1 and letting Q L model. Then = 1 C A z P kP P dinL out p inL p out AcLz

where mspool is the mass of the spool valve, Cs is the damping coefcient used to model the frictional loss, and Ks and s are the stiffness and preload of the spring, respectively. 3.6 Solenoid Model. A solenoid can be modeled as a resistance and inductance RL circuit, as shown in Fig. 10. Note that when a peak and hold drive circuit is used, Vin becomes a function of time when the solenoid is activated. The relationship between the current i in the coil and the magnetic force Fs on the armature is assumed to take the following form: Fs = L bi2 x 1+ a 52

k Pp

CdoutLALzoutLRkTatmkoutPout

Poutz 46 Lk z

where L is a number between 0 and 1 depending on heat loss, and CdinL and CdoutL are the discharge coefcients. Applying mass conservation law to the control volume results in Journal of Dynamic Systems, Measurement, and Control

where x is the displacement of solenoid actuator, see Fig. 10. Coefcients a and b are chosen by curve tting the empirical data provided by the manufacture. MARCH 2010, Vol. 132 / 021007-7

Downloaded 27 Feb 2010 to 220.225.248.2. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Simulations and Experiments

Fig. 10 Solenoid model

Table 1 Test matrix Test No. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 Psupply psi 30 30 30 30 30 30 40 40 40 40 40 40 40 40 40 40 40 40 Period ms 100 100 40 40 24 24 100 100 100 100 40 40 40 40 24 24 24 24 Duty % 30 30 30 30 30 30 30 30 25 25 30 30 25 25 30 30 25 25 Delay ms 3 5 3 5 3 5 3 5 3 5 3 5 3 5 3 5 3 5

4.1 Experiment Setup. A Ford 4.6 l four-valve V8 engine head was used for the valve test. The camshaft was removed from the intake valve side and an EPVA was installed above one of the intake valves. A Micro-Epsilon point range laser sensor was used to measure the intake valve displacement. The laser sensor was mounted on an angle such that the laser beam from the emitter of the laser sensor would be perpendicular to the surface of the end of the valve stem. A dSPACE PCI board was used for both control and data acquisition. A low side switch drive circuit, made from two insulated gate bipolar transistors IGBTs, was used as a driver circuit for both solenoids. Two STP2416-015 small pushpull solenoids were used to drive two spool valves in the EPVA. The experiments were conducted under the combinations of various control parameters, where two supply pressures 30 psi and 40 psi were used; three solenoid periods, 100 ms, 40 ms, and 24 ms, corresponding to engine speed at 1200 rpm, 3000 rpm, and 5000 rpm were selected and two solenoid duty cycles 30% and 25% were applied. The delay between the rst and second solenoids was selected to be 3 ms and 5 ms. Table 1 lists all 18 experiments conducted. The valve responses were compared with the simulation responses in Sec. 4.2. The EPVA is aimed to tailor the engine intake ow without throttling. Therefore, in the experiments and simulations, the engine intake manifold pressure is considered to be close to the atmospheric pressure. No pressure loads are included on the valve head F p = 0. In future studies where the exhaust valve dynamics are investigated, the valve will have to open against a high engine cylinder pressure. This model can be expended to perform the exhaust valve simulation, see Eq. 24. 4.2 Simulation Results. The equations of motion derived previously were written in state space form and programed in MATLAB/SIMULINK. The simulations were performed under the same parameter sets as these used in experiments. Eighteen experimental and simulation responses of the valve displacement are presented in Figs. 1113 for all 18 tests. The dashed lines represent the experimental responses and the solid lines represent the simulation responses.

Fig. 11 Plots with 30 psi supply pressure and 30% duty cycle

021007-8 / Vol. 132, MARCH 2010

Transactions of the ASME

Downloaded 27 Feb 2010 to 220.225.248.2. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Fig. 12 Plots with 40 psi supply pressure and 30% duty cycle

Figure 11 compares the experimental and simulation responses with 30 psi supply pressure. The top two graphs of Fig. 11 display the responses of 30% solenoid duty cycle at 100 ms period with 3 ms test 1 and 5 ms test 2 delay between two solenoids. Note that the 100 ms solenoid period corresponds to the engine speed at 1200 rpm. The response with 5 ms delay test 2 had about 6 ms rising time, and the response with 3 ms delay test 1 had about 5 ms rising time. The maximum valve lift height was 6 mm for the response with 5 ms delay test 2 and 3.8 mm for the response with 3 ms delay test 1. The swing motion on the top of the prole indicates that the valve is in the dwell stage when the

hydraulic latch is utilized to hold the valve open. The slight compressibility of the oil in the hydraulic latch causes a few oscillations of the valve responses. The hydraulic damper is engaged when the valve is roughly 1 mm away from its seat which can be adjusted by adjusting the actuator piston stem clearance, where the slope of the response is largely decreased. The valve approaches to the original position gradually afterwards. The responses in the middle two graphs of Fig. 11 were obtained under the same operating conditions as those on the top graphs except that the solenoid period was reduced to 40 ms corresponding to 3000 rpm. The rising times of the responses with 5

Fig. 13 Plots with 40 psi supply pressure and 25% duty cycle

Journal of Dynamic Systems, Measurement, and Control

MARCH 2010, Vol. 132 / 021007-9

Downloaded 27 Feb 2010 to 220.225.248.2. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

ms test 4 and 3 ms test 3 delay between two solenoids were 6 ms and 5 ms. As the solenoid period was reduced, the dwell stage was shorter. The maximum valve lift height was 6 mm for the response with 5 ms delay test 4 and 4 mm for the response with 3 ms delay test 3. The solenoid period was further reduced to 24 ms corresponding to 5000 rpm. The responses tests 5 and 6 are shown in the bottom two graphs of Fig. 11. In this case, the maximum valve lift was 5 mm and the rising time was 6 ms for the response with 5 ms delay between two solenoids test 6; the maximum valve lift was 4 mm and the rising time was 5 ms for the response with 3 ms delay test 5. The maximum valve lift height in the 5 ms delay case was decreased from 6 mm to 5 mm. This happened because the solenoid was de-energized before the actuator piston can fully expand to its maximum displacement; the valve started returning before it reached its maximum lift. Moreover, the valve never entered the dwelling region in these two cases. The solenoid period was so short that the valve entered the air discharging stage immediately after the air charging stage. Hence, the swing motion disappeared on the top of the prole. The experiment and simulation responses at 40 psi supply pressure with 30% and 25% solenoid duty cycles are presented in Fig. 12 and 13, respectively. The rising time of the valve varied from 4 ms to 6 ms. The maximum valve lift was around 8 mm for the responses with 5 ms delay and 6 mm for the response with 3 ms delay in this case. As expected, the valve lift height could be controlled by regulating the supply pressure and/or varying the delay between two solenoids, and the valve open duration could be controlled by controlling the activation duration of the solenoid. The mathematical model was able to capture the dynamics of the EPVA closely.

Acknowledgment
The authors gratefully acknowledge the support for this work from the U.S. Department of Energy, National Energy Technology Laboratory, Energy Efciency and Renewable Energy Division, Samuel Taylor, Project Manager. The authors also like to acknowledge the support from Urban Carlson, Anders Hoglund, and Mats Hedman of Cargine Engineering. This work was completed during the Ph.D. study of J. Ma at Michigan State University.

References
1 Lenz, H. P., Wichart, K., and Gruden, D., 1988, Variable Valve TimingA Possibility to Control Engine Load Without Throttle, SAE Paper No. 880388. 2 Negurescu, N., Pana, C., Popa, M. G., and Racovitza, A., 2001, Variable ValveControl Systems for Spark Ignition Engine, SAE Paper No. 2001-010671. 3 Lanceeld, T., 2003, The Inuence of Variable Valve Actuation on the Part Load Fuel Economy of a Modern Light-Duty Diesel Engine, SAE Paper No. 2003-01-0028. 4 Boggs, D. L., Hilbert, H. S., and Schechter, M. M., 1995, The Otto-Atkinson Cycle Engine-Fuel Economy and Emissions Results and Hardware Design, SAE Paper No. 950089. 5 Urata, Y., Awasaka, M., Takanashi, J., Kakinuma, T., Hakozaki, T., and Umemoto, A., 2004, A Study of Gasoline-Fuelled HCCI Engine Equipped With an Electromagnetic Valve Train, SAE Paper No. 2004-01-1898. 6 Trask, N. R., Hammoud, M., Haghgooie, M., Megli, T. W., and Dai, W., 2003, Optimization Techniques and Results for the Operating Modes of a Camless Engine, SAE Paper No. 2003-01-0033. 7 Sugimoto, C., Sakai, H., Umemoto, A., Shimizu, Y., and Ozawa, H., 2004, Study on Variable Valve Timing System Using Electromagnetic Mechanism, SAE Paper No. 2004-01-1869. 8 Theobald, M. A., Lequesne, B., and Henry, R. R., 1994, Control of Engine Load Via Electromagnetic Operating Actuator, SAE Paper No. 940816. 9 Pischinger, F., and Kreuter, R. P., 1984, Electromagnetically Operating Actuator, U.S. Patent No. 4,455,543. 10 Lenz, H. P., Geringer, B., and Smetana, G., 1989, Initial Test Results of an Electro-Hydraulic Variable Valve Actuation System on a Firing Engine, SAE Paper No. 890678. 11 Richeson, W. E., and Erickson, F. L., 1989, Pneumatic Actuator With Permanent Magnet Control Valve Latching, U.S. Patent No. 4,852,528. 12 Watson, J. P., and Wakeman, R. J., 2005, Simulation of a Pneumatic Valve Actuation System for Internal Combustion Engine, SAE Paper No. 2005-010771. 13 Turner, J. W. G., Bassett, M. D., Pearson, R. J., Pitcher, G., and Douglas, K. J., 2004, New Operating Strategies Afforded by Fully Variable Valve Trains, SAE Paper No. 2004-01-1386. 14 Bobrow, J. E., and McDonell, B. W., Modeling and Control of a Variable Valve Timing Engine, Proceedings of the 2000 American Control Conference, Chicago, IL, Jun. 15 Tressler, J. M., Clement, T., Kazerooni, H., and Lim, M., 2002, Dynamic Behavior of Pneumatic Systems for Lower Extremity Extenders, Proceedings of the 2002 IEEE International Conference on Robotics & Automation, Washington, DC, May. 16 Richer, E., and Hurmuzlu, Y., 2000, A High Performance Pneumatic Force Actuator System: Part INonlinear Mathematical Model, ASME J. Dyn. Syst., Meas., Control, 122, pp. 416425. 17 2009, http://www.cargine.com/tech2.html 18 Thomson, W. T., 1998, Theory of Vibration With Applications, 5th ed., Prentice-Hall, Upper Saddle River, NJ.

Conclusions

This article presented a dynamic model for an EPVA. This model is a key element for developing control strategies of the EPVA for an internal combustion engine. Two solenoids and two spool valves, a single-acting cylinder, an inlet port valve, an outlet port valve, a hydraulic latch-damper, and an intake valve with its valve spring are included in this model. The mathematical model employs Newtons law, mass conservation, and principle of thermodynamics. The nonlinearity of the ow, incompressibility, and compressibility of the hydraulic uid and the nonlinearity of the motion due to the physical constraint was carefully considered in the modeling process. The model was implemented in MATLAB/ SIMULINK under different combinations of operation conditions. Validation experiments were performed on a Ford 4.6 l four-valve V8 engine head with various air supply pressures, solenoid periods, solenoid duty cycles, and time delay between two solenoids. The numerical simulation results were compared and showed excellent agreement with the experimental data.

021007-10 / Vol. 132, MARCH 2010

Transactions of the ASME

Downloaded 27 Feb 2010 to 220.225.248.2. Redistribution subject to ASME license or copyright; see http://www.asme.org/terms/Terms_Use.cfm

Vous aimerez peut-être aussi