Vous êtes sur la page 1sur 14

A Coupled Dynamic Valve Spring and Engine Performance Simulation

Nigel Fleming, Richard Pearson, and Mike Bassett, Lotus Engineering Software, Lotus Engineering, UK

Abstract
Now that the use of simulation is an established part of the engine design and development process much current activity is directed
toward the coupling of codes to obtain greater model refinement. In this paper results from an engine cycle simulation code coupled
with a valve-spring dynamics code, with the facility to model a hydraulic lash adjuster, are presented. The coupled simulation enables
the examination of the influence of the force generated by the cylinder gas pressure on the valve motion and this modified behaviour
is then fed back in to the engine cycle simulation code to assess the impact on the engine performance. The most significant departure
of the valve motion from that generated by a kinematic analysis was found to be due to the valve-train flexibility. For both an
automotive engine and a high-speed racing engine the dynamically modified valve motion gives rise to only small differences in
predicted engine performance.

A model of a gas spring of the type used in some racing engines is also described in the paper. The model includes the effects of heat
and mass transfer and gas properties.

1. Introduction
In the past 30 years computer simulation techniques for mechanical analysis and thermodynamic cycle simulation have become firmly
established as an inextricable part of the powertrain design process. Engine performance simulation codes have evolved from the
status of research tools to fully supported commercial software packages. These thermodynamic cycle simulations, which include
non-steady gas dynamics, are used routinely to predict engine performance [1,2], thereby assisting in the design of the intake and
exhaust manifolds (and silencers), the specification of cam profiles, the prediction of the gas pressure forces and thermal loading on
the engine. Valve-train and crank-train models of varying levels of complexity are then used to perform the mechanical and thermal
design analysis of the base-engine components. Classical calculations for crank train components are performed at the concept stage
and increasingly sophisticated modelling techniques are applied, leading to full finite element models of the engine structure. Initial
design of valve-train systems is often performed using simple kinematic models of individual valve-train units that can then be refined
by conducting dynamic analyses and, ultimately, building a full model of an engine valve train.

Engine performance simulation codes and cam profile design / valve-train analysis codes have a symbiotic relationship with each
other within the engine concept design process. The results of parametric analyses using the former programs specify the lift and
duration targets for the profile design carried out using the latter. The resulting cam profiles may then be used as input to the engine
simulation program in a latter stage of the engine design to assess the effect of mechanical constraints of the valve-train upon the
performance of the engine.

Valve-train analysis has become increasingly sophisticated as the engine design process has evolved [3-6]. Models of components
such as hydraulic lash adjusters have been introduced [6] and the effects of gas pressure forces on valve motion have been modelled
[6]. This paper describes a cam profile design and analysis tool that can be used to rapidly generate profiles using a kinematic
approach. The resulting model of the valve-train system can then be used as the basis of a 1-D dynamic model that can be coupled at
run-time to an engine performance simulation code. This enables the effects on the valve motion of the varying gas pressure forces in
the engine cylinder to be accounted for whilst providing the engine simulation with a more realistic valve motion.

A brief description of the engine performance simulation code is given in Section 2, followed by a more detailed account of the valve-
train analysis software in Section 3, including a description of the gas spring model used in the simulation of the racing engine.
Section 4 describes the coupling of the two programs whilst Section 5 presents the results and discussion of the work.

2. The Engine Simulation Model


The engine cycle simulation program used in the present work was Lotus Engine Simulation. The governing equations of one-
dimensional unsteady gas flow in ducts are solved using a shock-capturing finite volume scheme [2] in order to capture the effects of
intake and exhaust manifold geometry on engine performance. Boundary models are available for poppet, reed, piston-ported, or disk
valves, pipe junctions, superchargers, turbocharger compressors and turbines, charge-coolers and other major engine components.
Heat release models are used to characterise the combustion event and a choice of in-cylinder heat transfer models is available.

The code has a built-in ‘drag-and-drop’ model builder. A screen shot of a typical model is shown in Figure 1. This model includes
‘Sensor and Actuator’ elements that one used to vary the timing of the intake camshaft and control the operation of a variable
geometry intake manifold. Sensors and actuators can make use of more sophisticated signal-operator functions which are built in to
the code. As the software can operate in the Component Object Model (COM) environment, the input and output signals associated
with the sensors and actuators can also be passed externally so that the model is driven by, for example, a Simulink control model.
Fig. 1: Screen shot of model in Lotus Engine Simulation

3. The Valve-Train Simulation Model


The simulation program used in this study, Lotus Concept Valve Train, provides the facility to build valve-train models of various
degrees of complexity. Initial cam profile design can be executed using a wholly kinematic analysis with an empirical factor that is
used in the determination of float speed limits. The cam profile may be designed using a segmented-polynomial or Bezier-curve
approach, as described below. Later in the design process a dynamic model can be generated directly from the kinematic model. A
multi-mass representation of a single or double wire spring can be constructed or a gas spring of the type used in motorsport
applications may be modelled. It is also possible to include either solid or hydraulic lash adjusters in the dynamic model.

3.1. Cam Profile Generation


Cam profiles can be defined using a segmented polynomial approach. The particular polynomial form for each segment is dependent
on the number of points defined and the boundary conditions imposed on each segment. A typical polynomial for a profile segment is
2 8 14
æ q ö æ q ö æ q ö æ q ö
y = y 0 + C1 çç ÷÷ + C 2 çç ÷÷ + C3 çç ÷÷ + C 4 çç ÷÷ + ...... . (1)
è q ref ø è q ref ø è q ref ø è q ref ø
The user constructs the polynomial curve to define the valve/cam motion. A segmented polynomial curve is pre-defined which the
user can modify and impose additional constraints upon. The default curve is based on an eleven-point, six-segment, definition. Each
segment is represented by a polynomial having the required order to suit the number of points and boundary conditions specified. The
complete displacement curve and the first three derivatives (velocity, acceleration and jerk) are simultaneously displayed. Features
such as the ramp heights and velocities, the symmetry of the profile, the maximum lift, and the profile period may be modified.

‘Clipped velocity’ profiles can be created using 15 points and 10 segments, whilst ‘clipped acceleration’ profiles are based on 17
points with 12 segments. Clipped velocity or acceleration profiles enable the user to limit the maximum and minimum values of these
parameters to address problems such as eccentricity limitations. The points of extrema can be coupled so that, when modified, their
magnitudes are identical.
Existing profiles can be imported in to the software and analysed. During this import process Chebyshev polynomials [7] may be used
to interpolate the data on to a regular crank-angle basis, smooth the data if required, or produce undefined derivatives. Imported
profiles can be fitted to the polynomial format described above using a ‘feature-identification’ approach where the matching process is
performed according to user-defined criteria

An additional option for designing cam profiles is the use of piecewise Bezier curves [8] which are constructed as a sequence of cubic
segments [9]. The construction of Bezier curves uses a special case of cubic Hermite interpolation based on a construction due to
Bernstein [10] in which the interpolating polynomials depend upon certain control points. Figure 2 shows a Bezier curve defined by a
set of n+1 control points Z0, Z1,…,Zn. according to the equation
n
å
C(t ) = Zi Bi, n (t ) , (2)
t =0
where Bi , n (t ) is a Bernstein polynomial of the form
n!
Bi ,n (t ) = t i (1 - t ) n-i , (3)
(n - i )!i!
and t Î [0,1] . The Bezier curve passes through the first and last control points and is tangential at its end points to Z1-Z0 and Zn-Zn-1.
Performing translations and rotations on the control points gives rise to these effects on the Bezier curve.

Z1

Z2
Z0

Z4

Z3

Fig. 2: Bezier curve defined by control points Z0-Z4

The cam profile can be designed by using Bezier curves to define the acceleration, the velocity, or the lift. Six segments, each with a
minimum of four points, are used to define an acceleration diagram, with the ‘smoothness’ of the global curve being conditioned by
gradient continuity across the transitions. An additional requirement that the distance is identical between the last and first two points
on contiguous segments is imposed. Conventional ramps are added to the first and last points of the Bezier curve to create a complete
cam profile. The curves for velocity and lift are calculated via differentiation, and jerk from integration, of the global curve. This
dense but unevenly spaced data is then interpolated back onto the required regular cam angle increment. A number of constraints are
then applied to ensure that the basic shape of the piecewise curve follows that required for a coherent cam profile.

Because it is possible to drag the points of one curve independent of the others, (except for enforced continuity), the overall curve will
not normally meet the fundamental requirements required by a particular profile and additional operations are necessary. For a
symmetrical profile it is sufficient to ensure that the zero velocity point at maximum lift occurs at the centre of the profile – this will
also ensure that the velocity and lift at the top of the closing ramp are correctly matched. For an asymmetrical profile, once the
position of maximum lift has been set, a further two matching operations must be carried out, ensuring the compatibility of both
velocity and lift at the top of the closing ramp. Two points in the closing segments of the profile are manipulated to give the
appropriate ramp velocity and the lift is then checked against the specified lift at the top of the closing ramp. If the latter parameter is
not correct the acceleration curve is re-manipulated to achieve the closing ramp velocity and the top-of-ramp lift is checked again.
This process is iterated automatically by the software according to prescribed criteria.

Four segments are used if the velocity curve is used to manipulate the cam profile and two segments are used if the user wishes to
define the lift curve. Additional points may be added within each Bezier segment up to a maximum of 20.

Figure 3a shows a screen-shot from Lotus Concept Valve Train where a control point has been selected in the acceleration diagram –
in this case the selected point can be freely manipulated in the vertical direction (see yellow buttons at top of control panel) in order to
modify the cam profile. An example of the output results from a kinematic analysis is shown in Figure 3b.
Fig. 3a: Manipulation of acceleration diagram during cam profile design

Fig. 3b: Example of output from valve-train static analysis.


3.2. Dynamic Model of Valve Spring
Once a cam profile has been designed it can be analysed in more detail by converting the kinematic model of the valve-train system in
to a dynamic model. Only single valve actuation assemblies can be modelled in Lotus Concept Valve Train and hence no camshaft
torsional vibration effects are included. The stiffness of the camshaft bearings, the cam lobe, the tappet, the valve stem, and the valve
seat-to-cylinder head contact region are used in the dynamic model. The component mass data is carried over from the kinematic
model and the valve spring element can be resolved in greater detail by sub-dividing it into multiple discrete masses, each with their
own stiffness and damping values, as shown in Figure 4. The simulation calculates the motion of the individual masses (shown to the
left of the mass elements in Figure 4) and is capable of modelling the ability of the spring to control the motion of the valve across the
engine speed range.

3.3. Model of Hydraulic Lash Adjuster


Hydraulic lash adjusters (HLA’s) are now common in modern automotive engines. As part of the valve-train dynamic modelling
facility Lotus Concept Valve Train contains a means of representing these devices which is based on that described in reference [11].
A screen-shot of an HLA model in Lotus Concept Valve Train is shown in Figure 4. It can be seen that the ball-check valve and its
spring, which control the flow of fluid between the high- and low-pressure reservoirs, is included in the model. This mass transfer and
the effect of aeration levels on the bulk modulus of elasticity of the oil are accounted for in the simulation. The results of incorporating
such a model into a coupled analysis with an engine performance simulation will be described later in the paper.

Hydraulic lash
adjuster

Discrete mass
representing
part of the
valve spring

Fig. 4: Model of hydraulic lash adjuster

3.3. Gas Spring Model


Current F1 racing engines and some GP1 motorcycle engines employ small chambers filled with a gas, usually nitrogen, which
serve the same function as a mechanical spring in a conventional valve train. These ‘gas springs’ circumvent the problem of spring
surge by eliminating the coil spring entirely. A schematic of such a system is shown in Figure 5. The gas spring itself is
represented by a chamber, 1, which is fed by the system supply connections, 2. The chamber also experiences mass transfer
through a calibrated orifice, 3, and via leakage of the gas to the ambient, 4, via a sealing ring. The volume of the chamber varies as
the valve is moved. The system is shown as part of a ‘finger-follower’ actuation mechanism which are common in high
performance engines due to the lower frictional penalty they impose and the greater freedom of valve lift range and profile shape
they facilitate.
5

1. Gas spring.
2. System supply reservoir. 4
3. Calibrated orifice.
4. Ambient.
5. Finger follower.
1

2 3 2

Fig. 5. Schematic of gas spring model

The gas spring is modelled in Lotus Concept Valve Train by considering the change in properties of the gas in the spring chamber due
to its change in volume and the mass and heat transfers between the various connected reservoirs. An ideal gas is assumed, the
composition of which is defined by specifying the mole fractions of each constituent from a choice of 13 species. The variations of the
properties of the gas as a function of temperature are also included. For each species the enthalpy is given by

( )
5
h(T ) = h - h0 = Â a1T + a2T 2 + a3T 3 + a4T 4 + a5T 5 = Â å a j T j , (4)
j =1
where  is the Universal Gas Constant and the coefficients a1 to a5 are taken from Benson [12]. The enthalpy of the mixture is then
given by
Nspec
h (T ) = å x j h j (T ) , (5)
j =1
where Nspec is the number of species in the gas mixture and xj is the mole fraction of specie j.

The specific internal energy and mass of the gas in the spring chamber is integrated using a fourth-order Runge-Kutta scheme. The
internal energy at time level n+1 is given by
1
e n+1 = e n + (k 0 + 2k1 + 2k 2 + k3 ) (6)
6
where
de de de de
k0 = D t ; k1 = D t ; k2 = D t ; k3 = D t . (7)
dt t dt t dt t dt t
n ,en n +(1 / 2) D t ,en +(1 / 2) k0 n +(1 / 2 ) D t ,en +(1 / 2) k1 n + D t ,en+ k2

The chamber mass is integrated in the same way. The rate of change of internal energy with time is obtained from the First Law of
Thermodynamics for an unsteady, open system in the form

d e 1 ìd Q dV dm ü
= í -p -e + åin (m& h0 ) - åout (m& h0 )ý (8)
dt m î dt dt dt þ

where h0 is the specific stagnation enthalpy of the gas entering or leaving the chamber, and Q represents the heat transfer rate through
the gas chamber walls. The mass flow rates ( m & ) are obtained by solving the equations of compressible flow through an orifice [13].

Figure 6 shows a screen-shot from Lotus Concept Valve Train which depicts a model of a direct-acting valve mechanism with a gas
spring. The results of incorporating such a model into a coupled analysis with an engine performance simulation will be described
later in the paper.
Spring
chamber

Fig. 6. Screen-shot of gas spring model

4. Coupled Analysis of Valve Train Dynamics and Engine Performance


It is possible to link Lotus Engine Simulation externally to any other computer program which is capable of operating in the COM
environment. In the present work, however, a facility to link Lotus Engine Simulation directly with Lotus Concept Valve Train via a
dynamic link library (DLL) which shares dynamic memory was developed. This approach simplifies the procedure for setting up a
coupled simulation significantly and reduces the computational overhead associated with the COM environment.

The purpose of coupling the codes is to enable the force generated by the pressure differential across the valve head to be included in
the simulation of the valve-train dynamics and, conversely, to provide the engine performance simulation with a more realistic valve
lift curve. The dynamics of each valve in the engine simulation model can be modelled simply by selecting the ‘Dynamic’ option from
their property sheets, as shown in Figure 7 (inside dashed line). Clicking the option labelled ‘Dynamic Model Data’ then opens Lotus
Concept Valve Train, producing a window similar to that shown in Figure 4, so that the appropriate data can be entered or an existing
model can be loaded. After loading or setting up the appropriate valve-train data the engine simulation code can then be run in the
usual way.

Fig. 7. Screen-shot of model used for analysis


5. Results and Discussion
Figure 8 shows a comparison of measured and predicted valve spring strain for the direct-acting valve train. The measured results
were obtained using a test rig on which the cylinder-head of a two-litre, four-cylinder, 16-valve engine was motored at 6000 rev/min.
(3000 rev/min. camshaft speed). The valve-train system has single wire springs, and hydraulic tappets. The correlation between
measured and predicted test results is very good.

Typical Valve Spring Strain Correlation


6000

5000

4000

Microstrain
3000

2000

1000

-1000
Test
Predicted
- 2000

Time
Fig. 8. Comparison of measured and predicted valve spring strain

Figure 9 shows the difference in valve lift predicted by a dynamic model of a direct-acting valve train with a mechanical tappet
compared with the kinematically generated profile [(dynamic lift)-(kinematic lift)]. No tappet-to-valve clearance was used in this case
and the system has been set up to ensure that there is no valve float. Results for both 2000 rev/min and 6500 rev/min are shown. In the
valve-closed period there is a slight recession of the valve caused by the seating force applied by the spring. The two large peaks in
each curve are caused by the system flexibility due to loading from the two main acceleration events (see Figure 3a for the general
form of the acceleration curve). The effects caused by these main acceleration events are larger at 6500 rev/min. than at 2000 rev/min..
The higher momentum of the valve in the deceleration event at 6500 rev/min. causes the departure of the profile from the kinematic
profile to reduce below that of the level found at 2000 rev/min.
0.00E+00
0 90 180 270 360
-5.00E-03
Change in valve lift / [mm]

-1.00E-02

-1.50E-02

-2.00E-02

-2.50E-02

mechanical tappet: 2000 rev/min


-3.00E-02
mechanical tappet: 6500 rev/min

-3.50E-02
Cam angle / [deg]

Fig. 9. Effects of valve-train dynamics on valve lift for direct-acting system with mechanical tappet
Figure 10 shows results produced by modelling a system with a hydraulic tappet. In this case the change in valve lift is relative to a
kinematic profile which has not been modified in any way. The presence of the hydraulic tappet introduces additional flexibility into
the system due to the compressibility of the oil. Further compliance is caused by the flow of oil through the ball-check valve before it
seats and the leakage of oil past the tappet seals. The differential lift does not recover to the extent that it does with a mechanical
tappet as the spring force, which retains control of the profile at all times, causes leakage of oil out of the tappet.
0.00E+00
0 90 180 270 360

-2.00E-02
Change in valve lift / [mm]

-4.00E-02

-6.00E-02

-8.00E-02

Hydraulic tappet: 2000 rev/min


-1.00E-01
Hydraulic tappet: 6500 rev/min

-1.20E-01
Cam angle / [deg]

Fig. 10. Effect of valve-train dynamics on valve lift for direct-acting system with mechanical tappet

Figure 11 shows the effects of introducing increasing levels of complexity to a dynamic model of a valve-train system with a
hydraulic tappet. An inlet valve is considered operating at an engine speed of 6500 rev/min.. The kinematic profile against which the
dynamic profiles were compared had 0.05mm at the valve opening point and 0.06mm at the valve closing point (varying linearly
across the profile duration) subtracted from the profile in order to account for the effects of tappet collapse described above. The
results obtained using a rigid valve show that the flexibility of this element contributes significantly to the departure of the dynamic
valve motion from the kinematic profile. Changing other system stiffness values for components such as the cam lobe and shaft and
valve seat also have a large effect and the sum of these accounts for most of the change in valve lift.
3.00E-02
Basic dynamic model (no dynamic spring)
2.00E-02 Basic dynamic model with rigid valve
Basic dynamic model with gas pressure
Dynamic spring model
1.00E-02
Dynamic spring model + gas pressure
Change in valve lift / [mm]

Basic dynamic model with rigid valve, seat, and camshaft


0.00E+00
0 90 180 270 360
-1.00E-02

-2.00E-02

-3.00E-02

-4.00E-02

-5.00E-02

-6.00E-02
Cam angle / [deg.]

Fig. 11. Effect of valve-train dynamics on valve lift for direct-acting system with mechanical tappet
The large dip in the change-in-valve-lift curve is due to the inclusion of the gas pressure forces in the cylinder which cause significant
valve recession in to the seat when the cylinder pressure is large (during the valve-closed period). The cylinder pressure makes only a
small difference to the intake valve in the valve-open period as the cylinder pressure is low during this time. Adding the dynamic
spring model to the system has the effect of slightly increasing the change in valve lift.

Figure 12a shows the kinematic profile generated for use on the direct-acting, four-cylinder, 16-valve automotive engine for which the
correlation between measured and predicted results shown in Figure 8 was produced. Values of 0.05mm and 0.1mm (varying linearly
across the profile duration) have been subtracted from the original profile in order to account for the hydraulic tappet collapse. Figure
12b shows that the major influence on the dynamic valve-lift curve is from the system stiffness and the collapse of the hydraulic
tappet. The rise in the change-in-valve-lift curve in the later part of the opening period is due to the over-compensation in the
kinematic profile of subtracting the predefined tappet collapse from the kinematic profile. The engine speed at which the results were
obtained was 6500 rev/min..

10

8
Valve lift / [mm]

2 LES - Fixed lift

0
0 90 180 270 360 450 540 630 720
Crank angle / [deg.]
Fig. 12a. Kinematic valve lift for direct-acting system with hydraulic tappet tappet – automotive engine

0.04
LES + Hydraulic tappet model
0.02 LES + Tappet + Spring
LES + Tappet + Spring + Gas loads
Change in valve lift / [mm]

0.00
0 90 180 270 360 450 540 630 720
Crank angle / [deg.]
-0.02

-0.04

-0.06

-0.08

7000 rev/min
-0.10

Fig. 12b. Effect of valve-train dynamics on valve lift for direct-acting system with hydraulic tappet. Automotive engine – 6500
rev/min.
A comparison of measured and predicted brake torque for the engine is shown in Figure 13. This prediction was performed using the
kinematic profile modified as described above. Figures 14 and 15 show the effects of introducing various degrees of sophistication
into the valve-train model, running Lotus Engine Simulation coupled with Lotus Concept Valve Train. In Figure 15 the change in
torque relative to that predicted using the kinematic profile is shown. It is clear that the largest effect is caused by the assumption
regarding the tappet collapse used in the kinematic curve.

200

190

180
Torque

10 Nm
Measured
170

Predicted - LES - fixed valve event


160

150

5000 5500 6000 6500 7000 7500


Engine speed/ [rev/min]

Fig. 13. Comparison of measured and predicted bake torque

200

190

180
Torque

10 Nm
Measured
170
LES - fixed valve event
LES + Hydraulic tappet model
160 LES + Tappet + Spring
LES + Tappet + Spring + Gas loads
150

5000 5500 6000 6500 7000 7500


Engine speed / [rev/min]
Fig. 14. Comparison of predicted bake torque for various valve-train models
-1.2

LES + Hydraulic tappet model


-1.0 LES + Tappet + Spring
LES + Tappet + Spring + Gas loads
Torque change (Nm)

-0.8

-0.6

-0.4

-0.2

5000 5500 6000 6500 7000 7500


0.0
Engine speed / [rev/min]

Fig. 15. Comparison break torque differences due to various valve-train models

Figure 16 shows a simulation model of a high-speed five-cylinder racing engine. The engine performance was modelled using a fixed
kinematic valve lift profile and by coupling the engine and valve-train simulation models. The valve-train-dynamic models included a
gas spring element of the type described in Section 3.3; the effect on the valve motion of the variation of cylinder pressure was also
included in one of the simulation runs.

Fig. 16. Simulation model of high-speed racing engine


The effects on the valve motion, at an engine speed of 16000 rev/min., of introducing the gas spring model and the variation of
cylinder gas pressure are shown in Figure 17. Again, the major contribution to the change in valve lift if due to the elasticity of the
valve-train components. It can be seen that, due to the very high engine speed, the effect of the valve acceleration is very large
compared with the valve head and seat deflection due to the gas pressure forces.
5.00E-02
Air spring model
Air spring + gas loads
0.00E+00
0 90 180 270 360
Change in valve lift / [mm]

-5.00E-02

-1.00E-01

-1.50E-01

-2.00E-01

-2.50E-01

-3.00E-01
Cam angle / [deg]

Fig. 17. Effect of valve-train dynamics on valve lift for direct-acting system with mechanical tappet and gas spring. Racing engine
– 16000 rev/min.

Figure 18 shows the impact of the dynamic valve motion on the results from engine performance simulation. At the peak-power speed
of 16000 rev/min. there is a reduction of 7.5 hp compared with the prediction obtained with the kinematic profile – this is consistent
with the reduction in valve lift across the profile caused by the dynamic effects in the valve train.
11

10

1 bar
BMEP

Fixed valve event


7
Air spring model
Air spring + gas loads
6

5000 7500 10000 12500 15000 17500 20000


Engine speed / [rev/min]

Fig. 18. Effect of valve-train dynamics on engine performance simulation.


Figure 19 shows that the inclusion of the gas pressure forces in the valve-train model make a considerable difference to the calculation
of valve seat forces for the five-cylinder racing engine. The peak seat force is approximately five times higher than the case when the
gas pressure forces are not included. The valve seat force gives and indication of the quality of the valve seating and affects valve seat
recession.
6000

5000
Valve seat force / [N]

4000
Air spring model + gas loads
Air spring model
3000

2000

1000

0
Time

Fig. 19. Effect of valve-train dynamics on engine performance simulation

6. Conclusions
The coupled engine and valve-train simulation enables the examination of the effects of the force generated by the cylinder gas
pressure on the valve motion. For both an automotive engine and a high-speed racing engine the dynamically modified valve motion
gives rise to only small differences in engine performance. The most significant departure of the valve motion from that generated by
a kinematic analysis was found to be due to the valve-train flexibility, with smaller effects being attributable to the collapse of the
hydraulic lash adjuster, when included. Omitting the gas pressure forces causes a considerable under-estimate of the peak valve seat
force.

7. References
1. Winterbone, D.E. and Pearson, R.J., Design techniques for engine manifolds – Wave action methods for IC engines. Professional Engineering
Publications, London, 1999.
2. Winterbone, D.E. and Pearson, R.J., Theory of engine manifold design – Wave action methods for IC engines. Professional Engineering
Publications, London, 2000.
3. Turlay, J.D., Smooth acceleration cam development.Trans. SAE, p.715, 61, 1953.
4. Kosugi, T., and Seino, T., Valve motion simulation method for high-speed internal combustion engine. SAE paper no. 850179, 1985.
5. Nishiura, H., and Akahane, H., Valve gear movement simulation. I.Mech.E. paper no. C430/006, International Conference: Computers in Engine
Technology, Robinson College, University of Cambridge, 10-12 September, 1991.
6. Watson, H.C., and Chow, C.H., Inclusion of hydraulic lash adjuster behaviour and gas dynamic forces in camshaft design. I.Mech.E. paper no.
C430/064, International Conference: Computers in Engine Technology, Robinson College, University of Cambridge, 10-12 September, 1991
7. Press, W.H., Teukolsky, S.A., Vetterling, W.T., and Flannery, B.P., Numerical Recipes in FORTRAN. 2nd Edition. Cambridge University Press,
Cambridge, 1992.
8. Chiorescu, C., Cristea, G., Oprean, M., and Jebelean, P., Computer analysis of the automotive valve gear. 3rd ATA International Cnference on
Innovation and Reliability in Automotive Design and Testing, Firenze, Italy, April 8-10, 1992.
9. Piegl, L., Fundamental Developments of Computer Aided Geometric Design. Academic Press, San Diego, 1993.
10. Lorentz, G.G., Bernstein polynomials. University of Toronto Press, Toronto, 1953.
11. ADAMS/Engine Product Guide, v12, MSC Software, 2002.
12. Benson, R.S., Advanced engineering thermodynamics. 2nd Edition. Pergamon Press, Oxford, 1977.
13. Benson, R.S., The thermodynamics and gas dynamics of internal combustion engines. Clarendon Press, Oxford, 1982.

Vous aimerez peut-être aussi