Journal of Mechanical Systems for Transportation and Logistics

Similar documents
Development of Assist Steering Bogie System for Reducing the Lateral Force

Special edition paper

Abstract In this paper, we developed a lateral damper to improve the running safety of railway vehicles

Study on System Dynamics of Long and Heavy-Haul Train

Gauge Face Wear Caused with Vehicle/Track Interaction

Analysis of Torsional Vibration in Elliptical Gears

Results in rail research using SIMPACK

A study on the evaluation method of the characteristics of the contact point between wheel and rail

Analysis of steering performance of differential coupling wheelset

Analysis on Steering Gain and Vehicle Handling Performance with Variable Gear-ratio Steering System(VGS)

Dynamic Response of High-Speed-Moving Vehicle Subjected to Seismic Excitation Considering Passengers' Dynamics

Research on Skid Control of Small Electric Vehicle (Effect of Velocity Prediction by Observer System)

THE INFLUENCE OF THE WHEEL CONICITY ON THE HUNTING MOTION CRITICAL SPEED OF THE HIGH SPEED RAILWAY WHEELSET WITH ELASTIC JOINTS

Simulation of a Narrow Gauge Vehicle using SIMPACK, Model Validation using Scaled Prototypes on Roller-Rig

Dynamic Behavior Analysis of Hydraulic Power Steering Systems

Improvement of Vehicle Dynamics by Right-and-Left Torque Vectoring System in Various Drivetrains x

Suppression of chatter vibration of boring tools using impact dampers

Design and Test of Transonic Compressor Rotor with Tandem Cascade

Switch design optimisation: Optimisation of track gauge and track stiffness

Friction Characteristics Analysis for Clamping Force Setup in Metal V-belt Type CVTs

Relationship between steering torque and ease of driving with bar type steering in high speed range

inter.noise 2000 The 29th International Congress and Exhibition on Noise Control Engineering August 2000, Nice, FRANCE

Simulation of railway track maintenance trains at MATISA

Analysis and evaluation of a tyre model through test data obtained using the IMMa tyre test bench

Study of the Performance of a Driver-vehicle System for Changing the Steering Characteristics of a Vehicle

Development of a High Efficiency Induction Motor and the Estimation of Energy Conservation Effect

Dynamic characteristics of railway concrete sleepers using impact excitation techniques and model analysis

CALCULATING ROLLING RESISTANCE OF FREIGHT WAGONS USING MULTIBODY SIMULATION

MODELING SUSPENSION DAMPER MODULES USING LS-DYNA

Development of Motor-Assisted Hybrid Traction System

Rigid-Flexible Coupling Dynamics Simulation Analysis of Wheel/Rail Interaction in High-speed Turnout Zone

MULTIBODY DYNAMIC STABILITY ANALYSIS OF A DIESEL-HYDRAULIC LOCOMOTIVE

Experimental investigation on vibration characteristics and frequency domain of heavy haul locomotives

Oscillator Experiment of Simple Girder Bridge coupled with Vehicle

Influence of dynamic unbalance of wheelsets on the dynamic performance of high-speed cars

Preliminary Study on Quantitative Analysis of Steering System Using Hardware-in-the-Loop (HIL) Simulator

A Fundamental Study Concerning the Proper Performance of Lift Buffers in Revised JIS A 4306

Relative ride vibration of off-road vehicles with front-, rear- and both axles torsio-elastic suspension

Special edition paper

Pantograph and catenary system with double pantographs for high-speed trains at 350 km/h or higher

Design Calculation and Verification using SIMPACK Wheel/Rail

Fig.1 Sky-hook damper

Development of Integrated Vehicle Dynamics Control System S-AWC

A study on the vibration analysis of a maglev vehicle A theoretical investigation of the effect of magnetic damping on a vibration control system

Active Suspensions For Tracked Vehicles

Analysis on Steering Capability of a New Bogie with Independently Rotating Wheels

Vehicle Dynamic Simulation Using A Non-Linear Finite Element Simulation Program (LS-DYNA)

Vehicle Turn Simulation Using FE Tire model

Special edition paper

Featured Articles Utilization of AI in the Railway Sector Case Study of Energy Efficiency in Railway Operations

Modeling and Vibration Analysis of a Drum type Washing Machine

Study on Flow Fields in Variable Area Nozzles for Radial Turbines

Multiphysics Modeling of Railway Pneumatic Suspensions

Analytical Technology for Axial Piston Pumps and Motors

PROCEEDINGS. High Tech in Heavy Haul

Railway Bogies with Radial Elastic Wheelsets

Steering Dynamics of Tilting Narrow Track Vehicle with Passive Front Wheel Design

Defect Monitoring In Railway Wheel and Axle

Characteristics of wheel-rail vibration of the vertical section in high-speed railways

Interrelation between Wavelengths of Track Geometry Irregularities and Rail Vehicle Dynamic Properties

Maneuvering Experiment of Personal Mobility Vehicle with CVT-Type Steering Mechanism

The Application of Simulink for Vibration Simulation of Suspension Dual-mass System

EXPERIMENTAL INVESTIGATION OF THE FLOWFIELD OF DUCT FLOW WITH AN INCLINED JET INJECTION DIFFERENCE BETWEEN FLOWFIELDS WITH AND WITHOUT A GUIDE VANE

Forced vibration frequency response for a permanent magnetic planetary gear

What is model validation? Overview about DynoTRAIN WP5. O. Polach Final Meeting Frankfurt am Main, September 27, 2013

Optimisation of Railway Wheel Profiles using a Genetic Algorithm

Steering Actuator for Autonomous Driving and Platooning *1

Development of Advanced Computational Models of Railway Vehicles

Semi-Active Suspension for an Automobile

The track-friendly high-speed bogie developed within Gröna Tåget

High Capacity Tapered Roller Bearings

Development of a Multibody Systems Model for Investigation of the Effects of Hybrid Electric Vehicle Powertrains on Vehicle Dynamics.

Proceedings of the World Congress on Engineering 2008 Vol II WCE 2008, July 2-4, 2008, London, U.K.

MOTOR VEHICLE HANDLING AND STABILITY PREDICTION

Adaptive Power Flow Method for Distribution Systems With Dispersed Generation

III B.Tech I Semester Supplementary Examinations, May/June

Simulation of freight train during braking operation using SIMPACK

Performance Based Track Geometry: Optimizing Transit System Maintenance

Design and Calculation of Fast-Running Shunting Locomotives

A Novel Chassis Structure for Advanced EV Motion Control Using Caster Wheels with Disturbance Observer and Independent Driving Motors

B.TECH III Year I Semester (R09) Regular & Supplementary Examinations November 2012 DYNAMICS OF MACHINERY

Special edition paper

Study on Electromagnetic Levitation System for Ultrathin Flexible Steel Plate Using Magnetic Field from Horizontal Direction

APPLICATION OF A NEW TYPE OF AERODYNAMIC TILTING PAD JOURNAL BEARING IN POWER GYROSCOPE

NUMERICAL ANALYSIS OF IMPACT BETWEEN SHUNTING LOCOMOTIVE AND SELECTED ROAD VEHICLE

Optimization of Seat Displacement and Settling Time of Quarter Car Model Vehicle Dynamic System Subjected to Speed Bump

Comparative study on wheel rail dynamic interactions of side-frame cross-bracing bogie and sub-frame radial bogie

MODELS FOR THE DYNAMIC ANALYSIS OF THE SUSPENSION SYSTEM OF THE VEHICLES REAR AXLE

Improvements of Existing Overhead Lines for 180km/h operation of the Tilting Train

Comparative study between double wish-bone and macpherson suspension system

Development of Emission Control Technology to Reduce Levels of NO x and Fuel Consumption in Marine Diesel Engines

MARINE FOUR-STROKE DIESEL ENGINE CRANKSHAFT MAIN BEARING OIL FILM LUBRICATION CHARACTERISTIC ANALYSIS

The evaluation of endurance running tests of the fuel cells and battery hybrid test railway train

Development of Rattle Noise Analysis Technology for Column Type Electric Power Steering Systems

Mathematical Modelling and Simulation Of Semi- Active Suspension System For An 8 8 Armoured Wheeled Vehicle With 11 DOF

A Study on Design Factors of Gas Pedal Operation

ALS (Active Lateral Suspension) By Bernard GAUTIER SNCF

Development of a Clutch Control System for a Hybrid Electric Vehicle with One Motor and Two Clutches

An Evaluation of Active Knee Bolsters

Steering performance of an inverted pendulum vehicle with pedals as a personal mobility vehicle

Transcription:

A Potential of Rail Vehicle Having Bolster with Side Bearers for Improving Curving Performance on Sharp Curves Employing Link-Type Forced Steering Mechanism* Katsuya TANIFUJI **, Naoki YAEGASHI ** and Hitoshi SOMA ** ** Dept. of Advanced Mechanical Science and Engineering, Graduate School, Niigata University, 8050 Ikarashi 2-nocho, Nishi-ku, Niigata, 950-2181, Japan E-mail: tanifuji@eng.niigata-u.ac.jp Abstract The air spring of bolsterless bogie trucks, which have been widely employed in railway vehicles in recent years, undergoes a large distortion when the vehicles negotiate sharp curves in lines such as subway lines, and this can deteriorate the durability of air springs. Furthermore, bolsterless trucks tend to suffer from increased wheel lateral force around sharp curves with a radius of 100 m or less. In this paper we discuss the application of a link-type forced steering mechanism to bogie trucks with a bolster as a countermeasure against the above-mentioned situation. A numerical simulation is carried out using a MBS software, SIMPACK. As a result, under the condition of reduced longitudinal stiffness in the primary suspension, a bolster truck with the link-type steering mechanism exhibits the potential to suppress the wheel lateral force occurring around sharp curves. Also, the deterioration in running stability due to the application of the steering mechanism can be recovered by adding moderate lateral damping in the secondary suspension. In addition, the obtained wear index shows that the forced steering truck has decreased flange wear resulting from passing through sharp curves. Key words: Transport, Railway, Simulation, Curve Negotiation, Forced Steering Truck 1. Introduction *Received 14 Dec., 2009 (No. T2-08-0273) Japanese Original : Trans. Jpn. Soc. Mech. Eng., Vol.75, No.750, C (2009), pp.420-428 (Received 1 Apr., 2008) [DOI: 10.1299/jmtl.3.393] Copyright 2010 by JSME There are sharp curves with a radius of 160 m or less in conventional narrow-gauge lines, particularly in subway lines. When a railway vehicle negotiates such a sharp curve, a large lateral force acts between the wheel and the rail, which accelerates the wear on the wheel flange and on the side of the rail. In addition, in the case of vehicles with bolsterless trucks, the durability of the air spring deteriorates owing to the large longitudinal distortion caused by the increased bogie angle in sharp curves. As a measure to reduce the lateral force during the negotiation of a curve, a forced steering mechanism has been employed [1-6]. However, it has been reported that for a vehicle with bolsterless trucks the link-type forced steering mechanism increases the wheel lateral force in sharp curves with a radius of 100 m or less because the wheel is strongly pressed onto the rail by the steering mechanism [2, 4]. On the other hand, a vehicle with side-bearer-type bolster trucks causes the trucks to turn through the slippage of the side bearers as a result of the lateral force induced where the wheel flange comes in contact with the rail in entrance transition curves. Owing to the slippage of the side bearers, the large 393

distortion of air springs can be avoided because the bolster rotates little relative to the car body, although the friction occurring in the side bearers may increase the lateral force slightly. In addition, the wheel lateral force is reduced in circular curves since the restoring moment arising from the distorted air springs decreases. As mentioned above, it is expected that the large distortion of the air springs can be avoided, thus reducing the wheel lateral force in circular curves, by adding a forced steering mechanism to a bolster-type vehicle. In this paper, we demonstrate the suitability of a bolster-type vehicle for negotiating sharp curves by a numerical simulation using the MBS software SIMPACK. We also examine the possibility that a link-type forced steering mechanism can prevent an increase in the wheel lateral force around sharp curves [7]. 2. Comparison between Truck Structures 2.1 Analytical vehicle model An analytical model of a bolster-type vehicle running on a conventional narrow-gauge line [8] is developed using SIMPACK, which has a specialized module for the analysis of railway systems, allowing us to readily perform a valid calculation on the contact force between the wheel and the rail [9]. The vehicle model has a total of 36 degrees of freedom of motion. The vehicle with bolsterless trucks, which is used for comparison, is modeled by uniting the bolster with the truck frame as shown in Fig. 1. Also, the bolster anchor of the side-bearer-type truck is replaced with a traction device having an equivalent longitudinal stiffness and damping coefficient. Thus, the vehicle with bolsterless trucks has 34 degrees of freedom. The curving performances for each truck structure are compared using the two vehicle models. Fig. 1 Modeling of bolsterless truck using a truck model with a bolster and side bearings Table 1 Conditions of curved track 2.2 Comparison between curving performance of truck structures A curving simulation is carried out using SIMPACK for a bolster-type vehicle and a vehicle with bolsterless trucks, and the behaviors in a circular curve are compared. It was expected that the wheel lateral force would become small for the bolster-type vehicle. Curves with radii from 100 m to 400 m were investigated and the track conditions are listed 394

in Table 1. Here, V is the running speed, C is the cant, X TC is the length of the transition curve and S is the slack. The values in Table 1 are regarded as typical for conventional lines. Figure 2(a) shows the simulated longitudinal distortion of the air spring in the front truck, x AS, as the truck negotiates curves, and Fig. 2(b) shows the wheel lateral force acting on the outer side of the leading wheelset, Q 1. The displacement, x AS, obtained for the bolster-type vehicle is the relative displacement between the bolster and the car body, whereas it is the relative displacement between the truck frame and the car body for the vehicle with the bolsterless trucks. Here, it is assumed that the displacements on the left and right sides are identical. Figure 2(a) shows that the distortion of the air spring in the vehicle with the bolsterless trucks increases when the curve becomes sharper, while that in the bolster-type vehicle remains small. Furthermore, Fig. 2(b) shows that the wheel lateral force, Q 1, of the vehicle with the bolsterless trucks is greater than that of the bolster-type vehicle on a sharp curve of 100 m radius, although it is smaller on curves with radii of 200 m or more. These results show the superior curving performance of the bolster-type vehicle around curves with a radius of 100 m or less. Displacement x AS [mm] 60 40 20 0 Bolsterless With bolster 100 200 300 400 0 1 2 3 4 5 Curve radius R [m] Lateral force Q 1 [kn] 25 20 15 10 5 0 With bolster Bolsterless 100 200 300 400 0 1 2 3 4 5 Curve radius R [m] (a) Longitudinal displacement of air spring at front truck (b) Absolute value of lateral force of outer wheel on 1st axle Fig. 2 3. Addition of Steering Mechanism Comparison between curving performance of truck structures 3.1 Steering mechanism In this paper, as shown in Fig. 3(a), a link-type forced steering mechanism based on the bogie angle is used [10]. Figure 3(b) shows that the bolster and the axle box are connected to the steering lever, which rotates around a pin on the truck frame, via spring elements. When the vehicle negotiates a curve, the bogie angle generated by the turning of the truck causes the steering lever to rotate, which forces the front and rear wheelsets to turn in the opposite direction to each radial position. This allows the wheelset to decrease its attack angle and reduces the wheel lateral force during curving. The front and rear trucks each have two steering levers, each with one degree of freedom rotation; thus, the total number of degrees of freedom in the bolster-type vehicle with the steering mechanism is 40 because of the 4 degrees of freedom of the steering levers. 3.2 Conditions set in simulation Table 2 shows the four conditions on the steering mechanism added to the bolster-type vehicle. Here, in addition to evaluating the steering performance with or without the steering mechanism, the effect of lowering the longitudinal support stiffness of the axle box, k wx, is compared, where the stiffness is reduced from the baseline value by 70%. The friction coefficient, µ, between the wheel and the rail is varied around the baseline value of µ = 0.3 in the range from 0.1 to 0.55, considering that it changes greatly depending 395

on the conditions of the rail surface and the wheel tread. The curves considered have radii R = 100 m 400 m with the parameter values given in Table 1. Figure 4 shows an example of simulation results for the vehicle condition B0. The figure shows the waveform of the wheel lateral force acting on the outer leading wheel as it runs around a curve of R = 300 m with µ= 0.3. The comparison of the steering effect under various vehicle conditions is performed by considering the wheel lateral force, which becomes constant as the vehicle enters a circular curve as shown in the figure. Truck frame Connected to car-body through bolster-anchor Bolster Steering lever (a) Outline of forced steering truck Bolster Axle box k wx k LU Steering lever Truck frame Wheelset Pin connected to truck frame (b) Modeling k LL Fig.3 Modeling of forced steering link mechanism Table 2 Truck condition for simulation Fig. 4 Example waveform of wheel lateral force 396

4. Running Simulation 4.1 Curving performance Figure 5 shows a comparison between wheel lateral forces, Q 1, obtained from the simulation by varying the curve radius R. In the case of R = 300 m, shown in Fig. 5(a), the wheel lateral force is decreased upon adding the steering mechanism when the friction coefficient µ is higher than the baseline value of 0.3, that is, by changing the vehicle condition from B0 to A0, and from B1 to A1. The effect on the reduction of Q 1 is particularly large in the case of vehicle condition A1, in which the longitudinal support stiffness of the axle box, k wx, is lower. On the other hand, Q 1 remains about the same magnitude in the case of the low friction coefficient, µ = 0.1, even though the steering mechanism is added and the support stiffness k wx is lower. Fig. 5 Comparison of wheel lateral force in sharp curves between truck conditions In the case of R = 200 m, shown in Fig. 5(b), the effect of the steering mechanism on the reduction of Q 1 is small for vehicle condition A0, in which the support stiffness k wx remains unchanged from the baseline value. In this case, Q 1 increases to more than that for vehicle condition B0, which has no steering mechanism, when the friction coefficient is decreased. However, vehicle condition A1, in which the steering mechanism is added and the support stiffness is lower, still has the effect of reducing Q 1. When the curve radius decreases to R = 100 m, as shown in Fig. 5(c), Q 1 for the vehicle conditions with the steering mechanism (A0 and A1) becomes larger than that for the vehicle conditions without the steering mechanism (B0 and B1) when the friction coefficient is less or equal to 0.3. However, the wheel lateral force for vehicle condition A1, in which the support stiffness k wx is lower, is about the same magnitude as that for condition B1 when µ has the baseline value of 0.3, and still remains about 20% higher than that for vehicle condition B1 when µ is decreased to 0.1. In contrast, when µ is increased to 0.55 the wheel lateral force decreases the most for vehicle condition A1. In addition, there is little change in Q 1 for conditions B0 and B1, which have no steering mechanism, even though the support stiffness k wx is lower. 397

Consequently, it is suggested that for the bolster-type vehicle Q 1 does not increase even in a sharp curve with radius 100 m unless µ decreases greatly when the longitudinal support stiffness, k wx, is lowered and the link-type forced steering mechanism is added. 4.2 Effect of leverage ratio The steering leverage ratio L 2 / L 1 is the ratio of the lower length L 1 to the upper length L 2 of the steering lever as shown in Fig. 6. L 1, the baseline value of which is L 10 = 0.046 m, is the length from the centers of rotation of the steering lever to the position where it is connected with the wheelset. L 2, the baseline value of which is L 20 = 0.3 m, is the length from the center of rotation to the position where it is connected with the bolster. Considering the spatial limitation, it is assumed that the steering leverage ratio varies in the range from 0.6 to 1.4 times the baseline value L 20 /L 10 by the scaling factor α. In the examination of the leverage ratio, only vehicle condition A1 is considered, because under this condition the leverage ratio has a large effect on the steering performance. L 2/ L 1 = α (L 20/ L 10) Scaling factor Fig. 6 Leverage length of steering mechanism Fig. 7 Effect of leverage ratio (truck A1) Figure 7 shows a comparison of Q 1 in a circular curve of 100 m radius, that is, in the steady state, with the friction coefficient µ between the wheel and the rail varied as a parameter. When µ is less than the baseline value of 0.3, Q 1 varies slightly with changing leverage ratio. Also Q 1 tends to increase with increasing µ. This is because the spring force required to steer the wheelset increases with the leverage ratio; thus its reaction force acts in the opposite direction, inhibiting the rotation of the truck. On the other hand, when µ is 0.55, Q 1 is decreased and a steering effect is obtained with increasing leverage ratio. When α is 1.4, however, Q 1 starts to increase. Figure 8 shows the behavior of the front wheelset (the first axle) and the rear wheelset (the second axle) in the case of µ = 0.55 shown in Fig. 7. The wheel lateral forces on the outer and inner sides of the rail in each axle are shown in Fig. 8(a), the lateral displacements of the front and rear axles are shown in Fig. 8(b), and the attack angles of the truck frame and both axles are shown in Fig. 8(c). For the wheel lateral forces shown in Fig. 8(a), positive values mean that the force acts from the outer side of the rail to the inner side, and vice versa for negative values. The subscripts of Q denote the axle number (1 or 2) and the wheel position (I for the inner rail side or O for the outer side). The lateral force Q 1O of the outer wheel on the first axle is always positive for all values of α, which means that the wheel is subjected to a reaction force acting from the outer rail at the flange contact. This corresponds to wheelset behavior in which the lateral displacement y w1 of the first axle reaches the flange clearance δ of about 10 mm. In contrast, the lateral displacement y w2 of the second axle remains within the flange clearance, and the wheel lateral forces of both the inner and outer sides of the rail are negative. Here, a lateral creep force acts toward the outer side of the rail. The difference between the wheel lateral force acting on the inner and outer wheels depends on the 398

difference between the gravitational forces acting on the contact points of each wheel tread. Observing the attack angle ψ w1 of the front wheelset in Fig. 8(c), it can be seen that the marked reduction of ψ w1 resulting from the increase in α to 1.2 leads to a decrease in the wheel lateral force Q 10 in Fig. 8(a). When α reaches 1.4, ψ w1 becomes positive and the wheelset is in the state of oversteering, then the wheel lateral force Q 10 starts to increase since the lateral creep force acts toward the inner side of the rail. For the second axle, it is confirmed that the lateral creep force, which acts toward the outer side of the rail, increases with an excess leverage ratio. As a result, when the attack angle ψ w2 increases the wheel lateral force of the inner wheel Q 2I increases markedly. Scaling factor Scaling factor Scaling factor Fig. 8 Effect of leverage ratio on front truck (truck A1, R =100 m, µ = 0.55) Consequently, it can be concluded that an increase in the leverage ratio does not cause a reduction of the wheel lateral force when the friction coefficient µ between the wheel and the rail is 0.3 or less, although a reduction in the wheel lateral force due to an increased leverage ratio can be anticipated when µ is large. 4.3 Wear index There is considerable wear of the wheel at the flange when a vehicle negotiates a sharp curve. Figure 9 shows a simulated waveform for the lateral displacement y w1 of the first wheelset for vehicle condition A1 with µ = 0.3. It is shown that the wheelset moves by about 10 mm toward the outer rail side (i.e. a value of -10 mm was obtained) and that the wheelset is in balance condition with the rail reaction force when it enters the transition curve. In this situation, the flange maintains contact with the rail, and wear on the wheel flange and the side of the rail occurs. The curving performance is compared for a sharp curve of radius R = 100 m using the wear index WI F proposed by Heumann [11] given by WI = µ F ψ (1) F F F 399

Here, µ F is the friction coefficient where the flange is in contact with the rail, F F is the flange reaction force, and ψ is the attack angle of the wheelset. Vehicle steering performance is compared for condition A1, which has the steering mechanism and lowered longitudinal support stiffness k wx of the axle box, and condition B0, where the support stiffness k wx has the baseline value and there is no steering mechanism. Here, we set the baseline friction coefficient of µ = 0.3 for vehicle condition A1, and µ = 0.1 for condition B0 assuming the presence of a friction modifier which has been employed to reduce the wheel lateral force around sharp curves [12]. Displacement y w1 [mm] 20 10 0-10 -20 Fig. 9 Transition curve Circular curve 0 100 200 300 Traveling distance [m] Lateral displacement of 1st axle Flange reaction force F F1 [kn] 40 30 20 10 0 B0 A1 0 100 200 300 Attack angle ψ w1 [mrad] 20 0-20 -40 A1 B0 0 100 200 300 Traveling distance [m] Traveling distance [m] (a) Flange reaction force (b) Attack angle Wear index WI F [N] 150 100 50 0 B0 A1 0 100 200 300 Traveling distance [m] (c) Wear index on flange Fig. 10 Comparison of factors on flange wear in a sharp curve (R = 100 m) Simulated waveforms of the flange reaction force F F1, attack angle ψ w1 and wear index WI F of the first wheelset are shown in Fig. 10. It is shown in Figs. 10(a) and 10(b) that ψ w1 of vehicle condition A1 is less than that for vehicle condition B0 because of the steerage mechanism, although the flange reaction forces, F F1, for conditions A1 and B0 have almost the same magnitude. As a result, the wear index WI F for condition A1 is smaller than that for condition B0 in the sharp curve and the exit transition curve as shown in Fig. 10(c). Thus, less flange wear is predicted for condition A1. 400

Next, the wear index integrated over the whole curved section including the transition curves is compared, as shown in Fig. 11. The integrated wear index (WI F ) SUM for condition A1 with the steering mechanism remains at a low level, while that for condition B0 increases with µ. Therefore, from the viewpoint of the wear index, the steering mechanism is effective for reducing the wear of the wheel flange; although it was shown above that the steering mechanism has a limited effect on reducing the wheel lateral force in a sharp curve. Fig. 11 Comparison of integrated wear index 4.4 Running stability For a vehicle operating under condition A1 with the lowered longitudinal support stiffness of the axle box, k wx, it is concerned that the running stability may deteriorate because of the reduced restriction of the wheelset motion and also because of the effect of the steering mechanism itself. Thus, the convergence property of the car-body lateral vibration is examined by assuming a step-shaped alignment irregularity of 5 mm on a tangent track. The simulation results are compared in Fig. 12. µ is assumed to be 0.3, and the target speed is set to V = 180 km/h, considering the maximum design speed for a commuter train with some leeway. Fig. 12 Running stability in each condition 401

Figure 12(a) shows the result for condition B0 at a speed of 180 km/h, Fig. 12(b) shows the result for condition A1 at a speed of 135 km/h and Fig. 12(c) shows the result for condition A1 at a speed of 180 km/h. The waveforms shown are for car-body lateral displacements y BF and y BR on the floor above the front and rear truck center positions, respectively. The lateral vibration for condition B0 is stable and converges at the target speed of 180 km/h as shown in Fig. 12(a). On the other hand, the lateral vibration for condition A1 does not converge, even at 135 km/h, and becomes a lasting vibration as shown in Fig. 12(b). Also, the vibration grows with increasing running speed and diverges at 180 km/h as shown in Fig. 12(c). Figure 13 shows the car-body lateral vibration of the vehicle with added lateral dampers running at the target speed of 180 km/h. The lateral vibration converges rapidly at the target speed, and it is confirmed that the running stability can be improved by adding lateral dampers with a relatively small damping coefficient. Here, the effect of adding the lateral dampers on the curving performance is small. Fig. 13 Improvement of running stability: truck condition A1 with lateral damper Accordingly, it is shown that the addition of lateral dampers is effective for improving the running stability, which may deteriorate in a vehicle with a steering mechanism and lowered longitudinal support stiffness of the axle box. 5. Conclusions We examined the curving performance of a bolster-type vehicle with side bearers having a link-type forced steering mechanism around sharp curves with a radius of 100 m, similar to those found in subway lines. (1) The bolster-type vehicle with an added steering mechanism can suppress the increase in the wheel lateral force occurring around a sharp curve when the longitudinal support stiffness of the axle box is reduced. When the friction coefficient between the wheel and the rail increases, it is even possible to reduce the wheel lateral force. (2) When the friction coefficient between the wheel and the rail is high, the wheel lateral force occurring around sharp curves can be reduced by increasing the leverage ratio. However, it is possible that the large leverage ratio may increase the wheel lateral force when the friction coefficient is µ = 0.3, which is the baseline value assumed in this study. (3) From the viewpoint of the wear index on the wheel flange, it is expected that the flange wear can be suppressed by the addition of the steering mechanism through the reduction of the attack angle as the curve is negotiated. (4) For a vehicle with the link-type steering mechanism and a low longitudinal support stiffness of the axle box, it is possible to ensure running stability up to the target speed by adding lateral dampers with a relatively small damping coefficient. As mentioned above, it is expected that the large distortion of air springs generated in the vehicle with bolsterless trucks can be eliminated by employing a bolster-type vehicle, and that the flange wear can be decreased for bolster-type vehicles. 402

Appendix The components of a vehicle and their respective degrees of freedom are shown in Fig. 14, along with an outline drawing of the analytical model. Car body DOF x Bolster x, y, z φ, θ, ψ 1 = 6 ψ θ y φ ψ 2 = 2 z Truck frame Wheelset x, y, z φ, θ, ψ 2 =12 x, y θ, ψ 4 =16 Total 36DOF Fig. 14 Modelling of full vehicle with bolster trucks in SIMPACK and degrees of freedom (without steering mechanism) The parameter values used in the calculation are those assumed for a subway vehicle used on conventional narrow-gauge lines. Mass Car-body: 14.62 t, Truck frame: 1.35 t, Wheelset: 0.85 t Moment of inertia Car-body roll: 2.04 m, Car-body pitch: 7.27 m, Car-body yaw: 5.60 m Truck frame roll: 0.50 m, Truck frame pitch: 0.24 m, Truck frame yaw: 0.50 m Wheelset roll: 0.50 m, Wheelset pitch: 0.24 m, Wheelset yaw: 0.50 m Longitudinal distance between Axles in truck: 2.1 m, Truck centers: 13.8 m Lateral distance between Axle springs: 1.6 m, Air springs: 2.0 m, Side bearers: 2.0 m, Bolster anchors: 2.45 m Height from rail top Center of gravity of truck frame: 0.72 m, Center of gravity of car-body: 1.97 m Air spring center: 0.85 m, Bolster anchor: 0.55 m, Car-body floor: 1.175 m Primary suspension Lateral stiffness per axle:9 800 kn/m, Vertical stiffness per axle: 2 440 kn/m Longitudinal stiffness per axle box k wx : 7 350 kn/m (baseline value) 2 000 kn/m (lowered value) Lateral damping per axle: 9.8 kns/m, Longitudinal damping per axle: 9.8 kns/m Secondary suspension Lateral stiffness of air spring: 200 kn/m, Vertical stiffness of air spring: 185 kn/m 403

Longitudinal stiffness of air spring: 200 kn/m, Vertical damping of air spring: 20kNs/m Bolster anchor Longitudinal stiffness: 2 250 kn/m, Longitudinal damping: 5 kns/m Friction coefficient of side bearer: 0.20 Wheel Rolling radius: 0.43 m, Tread profile: modified arc Rail profile: 50 kg-n Steering lever Height of rotation center from rail top: 0.43 m, Lateral distance: 2.0 m Length of upper lever: 0.046 m (baseline), Length of lower lever: 0.3 m (baseline) Upper lever stiffness: 50 000 kn/m, Upper lever damping: 5 kns/m Lower lever stiffness: 50 000 kn/m, Lower lever damping: 5 kns/m Acknowledgements The authors would like to thank Mr. Takuji Nakai and Mr. Satoshi Kikko of Sumitomo Metal Industries, Ltd., for their useful discussions during the present study. References (1) Matsuo, M. et al., A Study on Forced Steering Truck by Simulation, Proceedings of D&D 93, No.930-42, 1993, pp.500-505, JSME. (2) Matsuo, M. et al., Development of High Speed Steering Locomotive Bogie for Narrow Gauge, Proceedings of The 2nd TRANSLOG, No.930-81, 1993, pp. 245-249, JSME. (3) Komeda, G. and Hirosawa, S., Curving Simulation of Pendulum Vehicle with Car Body Tilt Control and Guided Steering Device by Z-Link, Trans. of JSME, Series C, Vol. 59, No.564, 1993, pp. 2494-2500 (4) Fujimoto, H. et al., Development of Steering Bogies for Speed-Up of Freight Train, RTRI report, Vol.8 No.3, 1994, pp. 23-28, RTRI. (5) Okamoto, I. and Kobayashi, H., Newly Developed Tilt Truck with Link-Type Forced Wheelset Steering, Proceedings of WCRR 97, Vol.D, 1997, pp. 615-621 (6) Sato, E. et al., Lateral Force between Wheel and Rail during Curve Negotiation of a Limited Express Diesel Car with Forced Steering Bogie, Trans. JSME, Series C, Vol. 64 No.625, 1998, pp. 3563-3570 (7) Schupp, G. et al., Multibody System Simulation of Railway Vehicles with SIMPACK, Vehicle System Dynamics, supplement to Vol.31 (1999), pp.101-109 (8) The Japan Society of Mechanical Engineers ed., Dynamics of Railway Vehicles, 1994, pp. 7-13, Denkisha- kenkyukai. (9) Netter, H. et al., New Aspects of Contact Modelling and Validation within Multibody System Simulation of Railway Vehicles, Vehicle System Dynamics, supplement to Vol.28, 1998, pp.246-269 (10) pp. 130-131 in Reference (8) (11) Kumar, S. and Rao, D. L., Wheel-Rail Contact Wear, Work, and Lateral Force for Zero Angle of Attack - A Laboratory Study, Trans. ASME, J. Dyn. Sys. Meas. Control,Vol.106, 1984, pp. 319-326 (12) Ishida, M. et al., Study on Friction Moderating for Wheel/Rail Interface, Proceedings of J-RAIL 2005, pp. 345-346, JSCE. 404