Anda di halaman 1dari 14

Simulation of wheelrail contact forces

S. IWNICKI
Manchester Metropolitan University, Chester Street, Manchester M1 5GD, UK Received in final form 4 August 2003

A B S T R A C T This paper summarizes the forces that develop in the contact patch between the wheel

and rail in a railway vehicle. The ways that these forces govern the behaviour of a vehicle running on straight and curved track are explained and the methods commonly used to calculate and utilize the forces summarized. As an illustration, the results from a computer simulation of a typical UK passenger train are presented and certain aspects examined. Keywords contact forces; profiles; railway vehicle dynamics; vehicle dynamics; wheel rail interaction.
NOMENCLATURE

a, b = the contact ellipse semi-axes C 11 , C 22 , C 23 , C 33 = constants calculated from approximate formulae given by Kalker5 E = Youngs modulus for the material in the contact patch Y l , Y r , Y w = lateral forces at left, right wheel contact patch, lateral force on wheelset Fx, Fy, Mz = longitudinal and lateral force and spin moment at contact patch Fx , Fy = forces at the contact patch (as above) modified by Johnson and Vermeulen4 f 11 , f 22 , f 23 , f 33 = linear creep coefficients defined by Kalker5 I = inertia of the wheelset about a central vertical axis l 0 = half the gauge m = mass of the wheelset N l , N r = normal force at left, right wheel contact patch P 0 = vertical force at the wheel due to static vehicle load P 1 , P 2 = dynamic vertical force response peaks at the wheel after a vertical irregularity Ql , Qr , Qw = vertical forces at left, right wheel contact patch, vertical force on wheelset r l , r r , = wheel radius at left, right wheel r 0 = wheel radius with wheelset in central position R = curve radius v = forward velocity of the wheelset U 1 , U 2 , 3 = actual velocity at the contact patch in lateral, longitudinal and spin directions U 1 , U 2 , 3 = velocity at the contact patch (as above) calculated from wheel motion W = wheelset weight = wheelset lateral displacement, velocity y, y 1 , 2 , 3 = lateral, longitudinal and spin creepage l , r , = conicity of wheel, left, right, effective

Correspondence: S. Iwnicki. E-mail: s.d.iwnicki@mmu.ac.uk

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

887

888

S. I W N I C K I

= coefficient of friction at the contact patch = roll angle of wheelset, angular velocity of wheelset rolling about , axle = yaw angle, yaw velocity of wheelset , = angular frequency of the kinematic oscillation of the rolling wheelset
INTRODUCTION

All the forces supporting and guiding a railway vehicle must be transmitted through the contact patches between the wheels and the rails. An understanding of the way that these forces are generated and the effect that they have on the behaviour of the vehicle has developed from trial and error in the early days to the use of computers to solve the complex equations developed in the 1960s and has resulted in the powerful computer packages currently in use. An understanding of the geometry of the wheel and rail are the foundations of this understanding.
THE GEOMETRY

The forces between railway wheels and rails are governed by the geometry of the wheel and the rail. In particular the geometry of a vertical cross section of the rail and a radial cross section of the wheel are critical. Wheel and rail profiles The earliest railway wheels were cylindrical and ran on flanged rails. They were usually fitted to an axle so that

both wheels could rotate independently. A good example of this is shown in Fig. 1 with probably the first locomotive in the world, built by Richard Trevithick, pulling trucks on the plate way at the Pen-y-Darren iron works in Wales in 1804. Fitting the flanges to the wheels instead of the rails must have made a considerable saving of material and probably allowed better guidance of the vehicle although it had the disadvantage of preventing the vehicles from running on the road. Adding a small amount of conicity to the wheels would have enhanced this guidance and the modern wheelset was formed when the two wheels were joined to the axle and fixed to the vehicle body through bearings in axleboxes. If a rolling wheelset moves away from the centre of the track the conicity at the wheels means that it will have a larger rolling radius on one side than on the other. As the wheels are stiffly linked in torsion they have to have the same rotational speed and the wheelset is forced to yaw about the vertical axis. This yaw angle tends to point the wheelset back towards the central rolling line and the wheelset will then naturally roll back to the centre of the track. In a curve the wheelset will tend to move outwards until the rolling radius difference between the two wheels

Fig. 1 Trevithicks tram engine in 1804 running on flanged rails at the Pen-y-Darren plateway. Painted by Terence Cuneo. Reproduced with the kind permission of the Cuneo Estate.

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

S I M U L AT I O N O F W H E E L R A I L C O N TAC T F O R C E S

889

Fig. 2 An idealized wheelset displaced laterally.

matches the yaw velocity needed for the curve (Fig. 2). This lateral displacement is known as the rolling line offset and the wheelset will curve perfectly as long as there is sufficient clearance for the required lateral movement. If the flangeway clearance is exceeded before the rolling line offset is reached then perfect curving will not be possible. The following equation links the lateral displacement, y, and the curve radius R: R l0 r0 y , = r0 + y R + l0 and the rolling line offset is therefore: y= r0 l 0 , R (2) (1)

can lead to derailment. The speed at which hunting occurs is known as the critical speed and vehicle designers must ensure that the critical speed is above the maximum running speed. In fact the kinematic behaviour is usually moderated by the creep forces, which are discussed below. Most railway organizations cant the rails inwards by a small angle and this usually matches the conicity of the wheel so that the normal force with the wheelset in the central position is directed along the web of the rail. In the United Kingdom this is 1 in 20 but 1 in 30 (for example in Sweden) and 1 in 40 (many countries including Germany) are also common. Wear at the wheel will tend to change the wheel tread from a cone to a more complex concave shape. Many railway organizations have designed worn profiles, which are intended to maintain a constant geometry as the wheel wears. In the United Kingdom the P8 profile was constructed from an average measured worn profile and is shown in Fig. 3. In use the wheel and rail will wear and the profiles must be measured to obtain accurate geometrical information. This can be done with reasonable accuracy using mechanical or laser devices and Fig. 4 shows a typical measured wheel and rail profile from the MiniProf measuring device.
T H E C O N TA C T

where r 0 is the radius at the contact point when the wheelset is central, l 0 represents half the gauge, R the radius of the curve and is the effective conicity. In fact the wheelset will tend to overshoot its equilibrium position (due to the developed yaw angle) and an oscillation known as the kinematic oscillation will be set up. This kinematic oscillation is also observed on straight track after any deviation from the natural rolling line. This oscillation was observed by George Stephenson in 1827 and analyzed by Klingel1 in 1873. The angular frequency of the kinematic oscillation can be found by assuming the motion to be sinusoidal: =v , r0 l 0 (3)

where v is the forward velocity of the wheelset. The greater the conicity of the wheelset, the smaller the curve radius for which perfect curving will be possible given a particular flangeway clearance. The other side of this engineering compromise is that the greater the conicity, the lower the rolling speed at which the wheelset becomes unstable. This instability is caused by the wheelset overshooting the equilibrium rolling line and is known as hunting. Hunting will be limited by flange contact but

At the point or points where the wheel contacts the rail a contact patch develops. The size and shape of this contact patch can be calculated from the normal force, the material properties and the geometry of the wheel and the rail in this region. As the wheel and the rail are both bodies of revolution it is possible to describe this geometry by using the radii of curvature in the direction of rolling and for the cross sectional geometry. In predicting the contact, the theory of Hertz based on uniform elastic properties of contacting bodies of revolution is often used giving an elliptical contact patch with semiaxes that can be calculated. Although this is an approximation based on full elasticity it is widely used and generally gives acceptable results. An alternative is to split the contact patch up into strips and to evaluate the contact conditions and the contact stress for each strip finally ensuring a balance between the wheel load and the total normal force at the contact patch.
THE FORCES

The forces acting in the contact patch can be split into normal and tangential components. The tangential force is usually split further into longitudinal (in the direction of the rail axis) and lateral (in the plane normal to the rail axis). The normal force and the lateral force can be

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

890

S. I W N I C K I

Fig. 3 The P8 worn profile.

Fig. 4 Wheel and rail profiles as measured by a miniprof device.

replaced with a vertical and lateral force where the vertical force is truly vertical and the lateral force acts in the horizontal plane. These are known as V (or Q)vertical and L (or Y)lateral and the ratio L/V or Y/Q is often used as an indicator of the nearness to derailment. Gravitational stiffness force As the wheelset moves laterally the direction of the normal force between the wheel and rail changes and a component of this force is directed towards the track centreline and

helps to centre the wheelset. This effect is known as the gravitational stiffness and the force depends on the lateral displacement and the roll angle of the wheelset. When the lateral displacement is small the gravitational stiffness force can be calculated ignoring differences in conicity across the wheels: Referring to Fig. 5 Yl = Nl sin(l ), Yr = Nr sin(r + ),

(4)

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

S I M U L AT I O N O F W H E E L R A I L C O N TAC T F O R C E S

891

Fig. 5 An idealized wheelset.

where Y l , Y r are the lateral forces, N l , N r are the normal forces, l r represent the conicity at left and right wheel and roll angle of wheelset equating vertical forces Nl cos(l ) = Q l , Nr cos(r + ) = Q r , so the total lateral force Yw = Yr Yl = [ Q l tan(l ) Q r tan(r + )] , for small angles this can be simplified to: Yw = W, where W = Ql + Qr = the total vertical load acting on wheelset. But = y rr rl = , 2l 0 l0 Yw = Wy . l0 (7) (8) (6) (5)

so finally

Creep forces When a railway wheel deviates from pure rolling, that is during acceleration, braking or curving or when subject to lateral forces through the suspension, forces tangential to the normal force are transmitted to the rail at the contact patch. These are called creep forces and are due to microslippage or creepage in the area of contact. If a cylindrical wheel rolls along a straight, flat rail with no tangential force being transmitted between the wheel and the rail the horizontal distance covered in one revolution of the wheel will be exactly equal to its circumference. If, however, a torque is applied to the axle to accelelerate the wheel then it will be found that in one revoultion the horizontal movement is less than the circumference of the wheel. This is due to the material behaviour within the contact patch as material is compressed at entry before a section where adhesion takes place then a section where the material slips out of compression and finally exits in tension.

Carter2 was the first to study creep in railway wheels and he looked to the earlier work of Reynolds in belt drives. He considered the wheel to be a thick cylinder rolling on a flat plate and only examined creepage in the longitudinal direction. He assumed without proof that the area of adhesion was at the leading edge of the contact patch. This was extended to the three-dimensional case of two rolling spheres in contact by Johnson3 who included consideration of lateral and longitudinal creepage. He assumed elliptical Hertzian contact and predicted an elliptical adhesion region within this. Slip only occurred in the area between the two regions. As a result of experimental work by Johnson and Vermeulen4 this theory was extended for non-spherical bodies and calculated tangential creep forces with an error consistently less than 25%. In a railway wheel the creepage can be calculated from the attitude of the wheelset and the resulting creep forces may then be evaluated. The relationship between creepage and creep force has been studied thoroughly by Kalker5 and his equations are used in almost all simulations. The theory of wheel rail creepage was only truly considered in three dimensions for the first time by Johnson3 who included spin of the wheel about a vertical axis in his theory. Kalker then developed a numerical method of predicting creep forces for arbitrary creepage and spin, the first available for predicting these forces. This was subsequently verified experimentally by Brickle6 who also looked at the result of having a narrow contact ellipse as is the case during flange contact. From this work Kalker has produced an exact numerical theory and a linear theory for use when creepage and spin are small. He has also made available several computer programs to predict creep forces for given creepages and spin. Creepage occurs in all three directions in which relative motion can occur and it is defined as follows: Longitudinal creepage 1 = Lateral creepage 2 = Spin creepage 3 = v1 v1 , v (9a) (9b) (9c)

v2 v2 , v
3

where v1 , v2 and 3 are the actual velocities of the wheel; v1 , v2 and 3 are the pure rolling velocities (velocity when no creep occurs at the same forward velocity) calculated from the wheel motion and v is the forward velocity of the wheelset. The creepage can be calculated from the attitude of the wheelset using equations based on the geometry and derived by Wickens.7 The velocities at each wheel can then be derived in terms of the lateral displacement of the wheelset centre of gravity

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

892

S. I W N I C K I

and the yaw angle of the wheelset, both with respect to the track centreline: for the right-hand wheel : for the left-hand wheel : rr rl v1r = v v1l = v r0 r0 (10) + v + v v2r = y v2l = y v3r = 0
3r

with Duvorol at low values of spin creepage and less than 18% at all values of spin creepage. The creepage creep/force relationship is further complicated by the fact that the three creepages do not act independantly. Kalker5 has shown that the creep forces depend on the creepages as follows: F x = f11 1 , F y = f22 2 f23 3 , Mz = f23 2 f33 3 , where f 11 , f 22 , f 23 and f 33 are the linear creep coefficients. The equations of Johnson and Vermeulen then modify the tangential forces: Fx = Fx Fs Fs 1 N 3 Fs N
2

v3l = 0
3r

+ =

+ =

(12)

for the right-hand rail : for the left-hand rail : 2l 0 v1r = v l0 2R v2r = 0 v3r = 0
3r

2l 0 + l0 v1l = v 2R v2l = 0 v3l = 0


3l

=0

=0 and

1 27

Fs N

N (13)

are the lateral displacement and velocity of the where y , y are the yaw angle and yaw velocity of the wheelset, , is the angular velocity of wheelset rolling wheelset and (=V /r o ) Note: If the track is considered to be flexible rather than rigid there may be a lateral component of track velocity to consider. These velocities can then be used in the previous equations to give the creepages for both wheels: l0 l0 y 1r = v r0 R y 21r = v 3r = r0 v l l0 y + 1l = + 0 + v r0 R y 21l = v . 3l = r0 v

(for F s 3 N), F x = N

Fx (for F s > 3 N), Fs 1 2 where F s = ( F x + F y 2 ) 2 , is the coefficient of friction at the contact patch and N is the normal force at the contact patch and similarly for FY from FY . The linear creep coefficients are calculated as follows: f11 = E (a , b )C11 , f22 = E (a , b )C22 , f23 = E (a , b )3/2 C23 , (14)

(11)

f33 = E (a , b )2 C33 , where E is the Youngs modulus, a, b represent the contact ellipse semi-axes and C 11 , C 22 , C 23 , C 33 are constants calculated from approximate formulae given by Kalker.5 The creep forces thus arrived at for the lateral and longitudinal direction at each wheel are then combined to give a lateral force and a yaw torque acting on each wheelset: Yw = 2 f22 Mw = 2 f23 y f23 v v 2 2 f11 l 0 y 2 f11 l 0 y . 2 f33 v v v r0 (15)

After determining the creepages it is necessary to find the related creep forces. At small values of creepage the relationship can be considered to be linear and linear coefficients can be used in calculations. However, at larger values of creepage, for example during flange contact, the relationship becomes highly non-linear and the creep force approaches a limiting value determined by the normal force and the coefficient of friction in the contact area. When working in this region it is necessary to use a different calculation method. It may be appropriate to use one of the programs based on the Kalker theory described above (e.g. Duvorol, Contact and Fastsim) but a simpler method based on the cubic saturation theory of Johnson and Vermeulen can also be used with generally good results. This is a heuristic method and involves calculating the creep force expected from the linear coefficient and modifying it by a factor derived from this value divided by the limiting creep force. The inaccuracy of this method has been shown by Shen, Hedrick and Elkins8 to be less than 10% when compared

This equation can be seen to explain the observed motion of the wheelset described earlier. The lateral creep force is proportional to the yaw angle of the wheelset and the yaw torque acting on the wheelset about a vertical axis is proportional to its lateral displacement. The effect of this is to steer the wheelset towards the centre of the track in decaying oscillations at all speeds up to a critical speed at which the oscillations continue laterally and in yaw. At higher speed the behaviour is unstable and the oscillations increase until limited by flange contact.

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

S I M U L AT I O N O F W H E E L R A I L C O N TAC T F O R C E S

893

The equations for creep force and gravitational stiffness derived above can now be combined with the inertia force terms for a wheelset to give a general equation of motion. The wheelset is taken as having the two degrees of freedom of lateral translation motion and yaw rotation y Wy + 2 f22 =0 + f23 + my v v l0 Iz 2 f23
2 2 f11 l 0 y y 2 f11 l 0 + 2 f33 + + = 0. v v v r0 (16)

Vertical forces The vertical forces that develop between the wheel and rail are made up of a force that supports the static load of the vehicle and a dynamically varying force in response to the vehicle motion along track with irregularities. These forces are usually referred to as P 0 force: the static load on the wheel P 1 force: the high frequency dynamic force where the wheel vibrates on the contact stiffness P 2 force: the lower frequency dynamic force caused by the wheel and rail vibrating on the substructure stiffness (pad, sleeper, ballast) Figure 7 shows an example of these forces as a vehicle runs over a vertical dip (such as a fishplated joint between rails).

These equations are used in simulations of complete railway vehicles. A note on approximations Some approximations are made due to the Hertz contact (mentioned in Section 3), due to the longitudinal shift of the contact patch and the change in the angle of the contact plane. It is also possible for multiple contact to occur between the wheel and the rail for some wheel rail profile combinations. These may be significant in flange contact and a fuller treatment is given by Brickle.6 The Kalker methods are widely used within computer simulation packages but they tend to be complicated and a new heueristic model has been proposed by Shen et al.8 which is much simpler to evaluate and is becoming more popular. Figure 6 shows a comparison of the Shen et al.8 heuristic model with Kalkers Duvorol and simplified Fastsim predictions.

Fig. 7 Vertical forces after a dipped rail joint.

Fig. 6 Comparison of creep force models (from Ref. [8]).

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

894

S. I W N I C K I

The frequency of the P 1 force is influenced by the unsprung mass (wheel and associated axle mass, bearings and brake gear) and the Hertzian stiffness at the contact patch. The P 1 force can be excited by a rail irregularity or defect or a wheel flat and its peak value occurs at around 1 ms after the disturbance. The P 2 force peak occurs at around 10 ms and is influenced by the rail and sleeper mass as well as the unsprung mass and the stiffness of the rail pads and substructure.
FORCE LIMITS

Derailment Likelihood of wheelclimb derailment is indicated by the ratio of the lateral to vertical force as mentioned earlier. The theory of Nadal10 is used to establish limits for the Y/V derailment ratio with 0.8 as the limiting value in UIC 518. Wheel unloading Very low vertical forces at the contact patch can indicate that a vehicle is susceptible to derailment by rolling over or by failing to follow twists in the track. A lower limit of 60% of the static wheel load (i.e., unloading by over 40%) is set in the United Kingdom.
S I M U L AT I O N PA C K A G E S

Railway organizations around the world have set limits on the various forces existing between the wheel and the rail. Some brief comments on the main limits are given below to provide a context for the force levels. Vertical In the vertical direction high forces can cause damage to the rails and supporting structures and can cause rolling contact fatigue when combined with high tangential forces such as occur during traction, braking or curving. In the United Kingdom a limit of 322 kN is set for the P 2 force (based on the maximum load measured for a Deltic locomotive running at maximum speed over a dipped rail joint). UIC 5189 sets a maximum static load of 112.5 kN per wheel and a maximum dynamic vertical force per wheel of between 160 kN and 200 kN, depending on maximum speed (provided this values does not exceed the static wheel load plus 90 kN). In small radius curves (less than 600 m) UIC 518 sets a limit of 145 kN for the quasi-static vertical force. Lateral In the lateral direction high forces can cause distortion of the track on the bed of ballast. This is normally protected against by using the simple but widely established PrudHomme limit for the track shifting force at one wheelset, which can be calculated from the static load (P 0 force): Y <= 10 + P0 /3, where Y and P 0 are in kN. Lateral forces of very short duration are less likely to shift the track and therefore only forces that act for more than 2 m of track length are usually counted. The European standard for vehicle acceptance UIC518 reduces the PrudHomme limit by a factor of 0.85 for freight vehicles. In small radius curves (less than 600 m) UIC 518 sets a limit of 60 kN for the quasi-static lateral force.

Using modern computer packages it is possible to carry out realistic simulation of the dynamic behaviour of railway vehicles. The theoretical basis of the mathematical modelling used is now mature and reliable and programs originally written by research institutes have been developed into powerful, validated and user-friendly packages. Examples are: ADAMS/Rail, GENSYS Nucars, Simpack and Vampire and the recent Manchester Benchmark exercise (see Iwnicki11 ) compared the results from these five packages in simulating a typical freight vehicle and a typical passenger vehicle on four different track cases. The first stage in setting up a computer model is to prepare a set of mathematical equations that represent the vehicle. These are called the equations of motion and are usually formed as a set of matrices. The equations of motion can be prepared automatically by the computer package, a user interface requiring the vehicle parameters to be described in graphical form or by entering a set of coordinates describing all the important aspects of the bodies and suspension components. The vehicle is represented by a network of bodies connected to each other by flexible elements. This is called a multibody system and the complexity of the system can be varied to suit the vehicle and the results required. The bodies are usually rigid but can be flexible with given modal stiffness and damping properties. Masses and moments of inertia need to be specified. Points on the bodies, or nodes, are defined as connection locations and dimensions are specified for these. Springs, dampers, links, joints, friction surfaces or wheelrail contact elements can be selected from a library and connected between any of the nodes. All of these interconnections may include nonlinearities such as occurs with rubber or air spring elements or as in damper blow off valves. Inputs to the model are usually made at each wheelset. Typical inputs are cross level, gauge and vertical and lateral alignment of the track. These can be idealized

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

S I M U L AT I O N O F W H E E L R A I L C O N TAC T F O R C E S

895

discrete events representing, for example, dipped joints or switches, or can be measured values from a real section of track taken from a recording vehicle. In the United Kingdom the High Speed Track Recording Coach (HSTRC) runs over the whole network collecting track data at regular intervals. Additional forces may be specified such as wind loading or powered actuators (e.g. in tilting mechanisms). Depending on the purpose of the simulation a wide range of outputs for example displacements, accelerations, forces at any point can be extracted.
C A S E S T U DY

Class 43 power car The Class 43 HST power car is a four-axle, two-bogie diesel engine locomotive. The main characteristics are: Axle load Total weight Unsprung mass per axle Bogie wheelbase Bogie pivot spacing Maximum operating speed Tractive effort at 125 mph 171.67 kN 70 000 kg 2175 kg 2.6 m 10.3 m 125 mph 1368 kW

In order to demonstrate the nature and effect of the forces described above a short case study is presented here. Computer models of a locomotive and a passenger coach have been set up using the computer package ADAMS/Rail. The Class 43 power car and the Mk.3 passenger coach are chosen as making up the widely used HST train running on UK main lines. Various results from the two vehicle models are then presented as examples of what can be obtained from simulations such as these. Figure 8 shows the Mk.3 coach model in ADAMS/Rail format.

The primary suspension includes (per axle): 4 coil springs 2 vertical dampers 4 traction links with end bushes 2 vertical bumpstops The secondary suspension includes (per bogie): 4 coil springs 1 traction centre equivalent bush 2 vertical dampers

Fig. 8 The computer model of the Mk3 passenger coach in ADAMS/Rail.

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

896

S. I W N I C K I

2 lateral dampers 2 yaw dampers 2 lateral bumpstops 2 vertical bumpstops Measured worn wheel profiles from a typical vehicle are used and left and right profiles are different. Mark 3 passenger coach The Mark 3 coach is often coupled with the Class 43 power car and is modelled here in the laden configuration: Axle load Total weight Unsprung mass per axle Bogie wheelbase Bogie pivot spacing Maximum operating speed 101.2 kN 41 281 kg 1595 kg 2.6 m 16.0 m 125.0 mph

1 anti roll bar equivalent bush 2 vertical viscous dampers 1 lateral viscous damper 2 vertical bump-stops + rebound-stops 2 lateral bump-stops 2 yaw friction elements to represent the friction pads. The wheel profiles used in the model are new P8 profiles. The vehicle models have been validated against measured track data provided by Corus Rail Technologies. These data were in the form of forces at the rail head derived from strain gauges on the rail web at a number of test sites where the Class 43 locomotive and Mk3 passenger coach was running. More details of this work are given in Jaiswal.12

Results The vehicles have been run on a range of track configurations and the behaviour in steady state curves as well as the dynamic response to track forces simulated. The track irregularities are taken from track recording coach data for the particular site. Figure 9 shows the lateral forces for the locomotive running into and around the 760 m radius curve at Aycliffe and Fig. 10 shows the position of the contact patch on the wheel and the rail for the same curve. The tables in Figs 11 and 12 present a summary of the displacements and forces acting on the rail as the two vehicles run in curves of 700 m and 1500 m at different speeds and with varying coefficients of friction. The speed

The primary suspension includes (per axle): 2 trailing arm bushes 2 vertical coil springs 2 vertical viscous damper 1 bush to represent the panhard link The secondary suspension includes (per bogie): 2 spring elements to represent the airspring + spring plank system 1 traction centre equivalent bush + damping

Fig. 9 Simulated lateral forces on the rail at the front bogie of a Class 43 locomotive running around a 760 m radius curve.

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

S I M U L AT I O N O F W H E E L R A I L C O N TAC T F O R C E S

897

Fig. 10 Simulation results showing contact position on rail and wheel for the Class 43 locomotive running around a 760 m radius curve.

Fig. 11 Curving behaviour of the Mk3 passenger coach 700 m radius curve with varying friction and cant deficiency.

differences are shown as cant deficienciesdeviations in the height difference across the rails from a balancing speed at which the lateral components of the centrifugal and gravitational forces balance. It can be seen that the lateral displacement of the leading wheelset does not

change greatly as the vehicles are running close to flange contact in all cases. The second wheelset moves outwards with increasing cant deficiency and on the coach this is more pronounced at the lower value of the coefficient of friction. The high stresses, sometimes evident on the

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

898

S. I W N I C K I

Fig. 12 Curving behaviour of the Class 43 locomotive in a 1500 m radius curve with varying friction and cant deficiency.

Finally, Fig. 14 shows the frequency distribution of the contact position for the two vehicles on straight and curved track. Observed running bands are superimposed on these plots and can be seen to match reasonably well with the tread contact of the vehicles (on the right rail). Flange contact (on the cess rail) is much more variable and is influenced greatly by the dynamic behaviour.

CONCLUSIONS

Fig. 13 Predicted contact stress against position on the rail head for the Class 43 locomotive in 700 m to 1800 m radius curves.

The forces acting between the wheel and rail in a railway vehicle have been analyzed and can be simulated with modern computer packages to simulate the behaviour of the vehicle on any track configuration. A wealth of information about the vehicle motions and forces within the suspension and on the rails can be obtained from these simulations.

inside rail, are probably not as damaging as those on the outside wheel in the curve as the creep force is generally much lower at the inside contact. The highest stress occurs at the flange contact in the locomotive when the area can drop as low as 10 mm2 . Figure 13 shows how the contact stress at the high (outside) rail varies with the position on the rail head for the Class 43 locomotive on a range of curves. The trend of increasing contact stress as the contact moves towards the flange can clearly be seen.

Acknowledgements The author thanks Yann Bezin and the rest of the Rail Technology Unit team for the computer models of the Class 43 locomotive and the Mk3 passenger coach. Thanks also to Jay Jaiswal, Andy Stephens, Stephen Blair and Tom Kay at Corus Rail Technologies for the wheel and rail profiles and the trackside measurements used to validate the models; and Network Rail

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

S I M U L AT I O N O F W H E E L R A I L C O N TAC T F O R C E S

899

Fig. 14 Frequency of predicted contact positions and observed running band for the class 43 locomotive in a 760 m radius curve.

Fig. 15 Frequency of predicted contact positions and observed position of the running band for the Mk3 coach on straight track.

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

900

S. I W N I C K I

for their support for the Corus Rail Technologies and MMU work.
REFERENCES
1 Klingel (1883) Uber den Lauf der Eisenbahnwagen auf gerader Bahn. Organ Fortsch. Eisenb-wes 38, 113123. 2 Carter, F. W. (1926) On the action of a locomotive driving wheel. Proc. R. Soc. A 1125, 151157. 3 Johnson, K. L. (1958) The effect of spin upon the rolling motion of an elastic sphere upon a plane. J. Appl. Mech. 25, 332338. 4 Johnson, K. L. and Vermeulen, P. J. (1964) Contact of nonspherical bodies transmitting tangential forces. J. Appl. Mech. 31, 338340. 5 Kalker, J. J. (1979) The computation of three dimensional rolling contact with dry friction. Int. J. Numer. Methods Eng. 14, 12931307. 6 Brickle, B. V. (1973) The steady state forces and moments on a railway wheelset including flange contact conditions. Doctoral

Thesis, Loughborough University 1973. 7 Wickens, A. H. (1965) The dynamic stability of a railway vehicle wheelset and bogies having profiled wheels. Int. J. Solids Struct. 1, 319341. 8 Shen, Z. Y., Hedrick, J. K. and Elkins, J. A. (1989) A comparison of alternative creep force models for rail vehicle dynamic analysis. In: Proceedings of the 8th IAVSD Symposium Cambridge U.K. Swets and Zeitlinger B.V. Lisse 1989. 9 UIC. (1999) Leaflet 518 (draft). Test and approval of railway vehicles from the points of view of dynamic behaviour, safety, track fatigue and ride quality. International Union of Railways January 1999. 10 Nadal, M. J. (1908) Locomotives a Vapeur. Collection Encyclopedie Scientific, Biblioteque de Mechanique Appliquee et Genie, 186 (Paris). 11 Iwnicki, S. D. (1999) The Manchester Benchmarks for Rail Vehicle Simulation. Suppl. Veh. Syst. Dyn. 31. 12 Jaiswal, J., Blair, S., Stevens, A., Iwnicki, S. and Bezin, Y. (2002) A systems approach to evaluating rail life. In: Railway Engineering 2002 (Edited by M. Forde). Glasgow University, London.

c 2003 Blackwell Publishing Ltd. Fatigue Fract Engng Mater Struct 26, 887900

Anda mungkin juga menyukai