Simulating Coupled Longitudinal, Pitch and Bounce Dynamics of Trucks with Flexible Frames

Abstract

Simulating the dynamic response of trucks requires that a model be constructed and subjected to road inputs. Inclusion or omission of flexible frame effects is often based on intuition or assumption. If frame vibration is assumed to be significant, it is typically incorporated in one of two ways. Either a complex finite element model of the frame is used, or a simplified linear modal expansion model (which assumes small motions) is employed. The typical low-order modal expansion model, while computationally efficient and easier to use, is limited by the fact that 1) large rigid body motions and road grade changes are not supported, and 2) longitudinal dynamics are not coupled to vertical and bounce dynamics. In this paper, a bond graph model is presented which includes coupled pitch and bounce motions, longitudinal dynamics, and transverse frame vibration. Large rigid body motions are allowed, onto which small flexible vibrations are superimposed. Frame flexibility is incorporated using modal expansion of a free-free beam. The model allows for a complete pitch-plane representation in which motive forces can propel the truck forward over varying terrain, including hills. The effect of frame flexibility on vehicle dynamics can then be studied. This is an extension of the typical half-car model in which suspension motion is assumed vertical, pitch angles are small, and longitudinal dynamics are completely decoupled or omitted. Model output shows the effect of frame flexibility on vehicle responses such as forward velocity, pitch angle, and payload acceleration. Participation of individual modes can be seen to increase as road input approaches their natural frequency. The bond graph formalism allows for any or all flexible frame modes to be easily removed from the model if their effects are negligible, and for inclusion of more complex submodels for components such as suspension and engine if desired.

Share and Cite:

D. Rideout, "Simulating Coupled Longitudinal, Pitch and Bounce Dynamics of Trucks with Flexible Frames," Modern Mechanical Engineering, Vol. 2 No. 4, 2012, pp. 176-189. doi: 10.4236/mme.2012.24023.

1. Introduction

The dynamic analysis of trucks requires a mathematical model of the vehicle structure (including engine, cab, and transmission), suspensions and tires, and the road excitation. While flexural vibration of the chassis can often be neglected in smaller, relatively stiff automobiles, large trucks and buses can experience significant “beaming mode” vibration. Beaming is response of the frame at its first modal transverse bending frequency, and for nonarticulated trucks this frequency can be on the order of bounce and pitch frequencies of a rigid vehicle [1-3]. Beaming response can be sizeable at the centre of the frame midway between the steered wheels and rear axle [4].

Approaches to modeling flexible vehicles range from 1) ignoring body flexibility by using a lumped mass model [5]; 2) modeling the frame as a regular free-free beam and calculating, estimating or measuring modal masses and stiffnesses [4,6]; and 3) modeling the entire vehicle using the finite element method [1,7-10]. As in most other dynamic systems, the analyst is faced with a spectrum of possible model complexity. The simplest models are easy to implement and computationally efficient but of limited accuracy and predictive ability. The most complex models present great computational burden but are potentially more accurate if parameters can be accurately determined. Low-order models with frame flexibility are typically limited to pitch-plane dynamics (bounce and pitch motions), and assume that frame angular motions are small and motion of any point can be assumed vertical. Longitudinal effects such as propulsion and braking forces, aerodynamic drag, tire rolling, slip resistance, and road inclination, when incorporated in order to predict forward speed, gradeability, or fuel economy; are usually decoupled from pitch and bounce modes. To predict vertical and pitch motions of the passenger compartment or payload, a vertical road input is typically applied to a pitch plane model, possibly with frame flexibility. Longitudinal motion determines when the vehicle encounters bumps, and therefore when suspension and body motions in the pitch plane are excited. Vertical road undulations and the resulting pitch plane response can also affect longitudinal motion. In other words, for certain vehicle parameters and road roughness, there can be two-way coupling between longitudinal and pitch/bounce motion [11-13]. To maximize the accuracy of vehicle response prediction when such coupling is present, and to further account for the effect of frame flexibility, the typical small-vertical-motion model with frame vibration must be extended to allow the large rigid body motions that arise from longitudinal motion and change in road inclination.

This paper presents a model which includes pitch and bounce motions, longitudinal dynamics, and transverse frame vibration. Forces and velocities are resolved along coordinate axes parallel and perpendicular to the undeformed frame. Large rigid body motions are allowed, onto which small flexible vibrations are superimposed. Frame flexibility is incorporated using modal expansion of a free-free beam. The model allows for complete pitch-plane representation in which motive forces can propel the truck forward over varying terrain, including hills. The effect of frame flexibility on vehicle dynamics can then be studied. The bond graph formalism is used to generate the model. Bond graphs, in addition to facilitating integration of flexible and rigid subsystems, allow for inclusion of more complex submodels for components such as suspension and engine if desired.

The following section reviews pitch plane models and approaches to including flexibility effects. Section 3 provides an overview of the bond graph modeling language. Section 4 gives schematics and equations for the rigid and flexible portions of the vehicle model, describes how terrain undulations are incorporated, and presents the final bond graph. Section 5 contains model output, including a study of the effect of road roughness on the coupling between frame flexibility, vehicle response in the pitch plane, and longitudinal motion. Discussion, conclusions and future work comprise Section 6.

2. Literature Review

Traditionally, linear pitch plane models have been the starting point for low-order modeling of truck dynamics including frame flexibility. Margolis and Edeal [4] created a 2 degree-of-freedom, small vertical motion bus frame model to which the engine, cab, and load were added in addition to suspension forces. The bond graph approach facilitated addition of external components to the flexible substructure. Frequency response of the frame showed a dominant beaming mode at approximately 10 Hz, very near the unsprung mass frequency. Margolis and Edeal [5] extended the aforementioned model to a five-axle tractor-trailer with submodels for engine, cab, and sleeper module/fuel tank. Non-linear elements were included in the model. Dynamic motions, in which the relative velocity across the suspension remained nearly zero, were identified as a significant effect of interaction between vehicle dynamics and frame vibration.

Michelberger et al. [14] identified a discrete transfer function for a free-free beam model of a two-axis bus, and then estimated modal characteristics based on measured data. For an air suspension bus, frequencies of 6.7 and 12 Hz were associated with the first two bending modes of the frame. Interaction of frame motion with the other vehicle dynamic elements is foreseeable given the 9.1 Hz engine mount natural frequencies and 10.3 - 10.4 Hz wheel-hop frequencies of the front and rear axles. Ibrahim et al. [9] modeled a truck frame using modal superposition, with modal properties calculated using a finite element model. The linear model included truck longitudinal velocity and a cab suspended by two linear suspension systems. The model assumed constant longitudinal velocity and no aerodynamic effects or tire lift-off. Frame flexibility was found to strongly affect driver’s vertical acceleration and cab’s pitch acceleration for a truck with frame natural frequencies of 7.25, 13, and 18 Hz. Yi [15] generated a 20-node finite element model of a frame modeled as a block, with both vertical and lateral vibration degrees of freedom. Response of rigid vs. flexible models to a pulse steering input showed significant discrepancies in predictions of lateral motion and tire deformation. Cao [16] developed a modal superposition representation of a truck frame rail with five segments—a front rail, kick-down rail, mid rail, kick-up rail, and rear tail. The frame rails were assumed to be the primary contributors to beaming mode. In the model 312 Nastran elements were used. The frame was modeled in isolation rather than as part of a vehicle model, without considering the effects of engine, cab, box and cab mounts that were acknowledged to affect the beaming frequency and nodes.

Truck frames have also been represented in ways other than free-free beams and finite element models for transverse vibration. Lumped-parameter subsystems are presented in [8,17]. In the former, torsional compliance was of primary concern for predicting roll angles, load transfer and yaw stability. The model in Aurell [17], in contrast to prior models in which the frame was represented as two rigid masses joined by a roll-axis torsional spring, introduced a “warp model” in which a pitch-axis torsional spring connected two longitudinal frame rails. A transverse bending degree of freedom was added by discretizing the longitudinal elements into two rigid masses connected by a pitch-axis torsional spring. Goodarzi and Jalali [8] formulated a 13 degree of freedom model with three rigid transverse beams joined by two massless flexible beams. The flexible beams have torsional and transverse compliance. Measured data and the influence coefficient method were used to populate the mass, stiffness, and damping matrices of a linear model. Flexibility was found to have a significant effect on pitch and roll angle predictions for a bus. While discretized bending representations are easy to implement and offer more insight into vehicle response than a rigid frame, a large number of elements are typically required to give a very close approximation of the lowest natural frequencies. Modal expansion has the advantage of giving the correct natural frequencies of the first n modes that are retained. Given the uncertainties in determining natural frequencies analytically for a complex irregular beam with multiple attachment points, either a discretized or modal expansion model can be tuned to match measured natural frequencies in practical applications.

The previously cited works show a range of modeling complexities in accounting for flexibility of heavy truck frames, and verify the potential importance of that vibration in vehicle dynamics models. However, prior models assume small angular motions of the frame, and do not include longitudinal dynamics along with pitch and bounce dynamics. As stated in the introduction, this paper combines longitudinal, rigid pitch and bounce, and transverse flexural dynamics of a truck into a single, computationally efficient model implemented using the bond graph graphical modeling language. Background on bond graphs follows, after which the model details are presented.

3. Bond Graph Modeling Language

In bond graphs, generalized inertias and capacitances store energy as a function of the system state variables (momentum and displacement, respectively), sources provide inputs from the environment, and generalized resistors remove energy from the system. The time derivatives of generalized momentum p and displacement q are generalized effort e and flow f. Power is the product of effort and flow. For example, force and voltage are efforts, velocity and current are flows, linear momentum and flux linkage are generalized momenta, and translational displacement and charge are generalized displacement variables.

Power-conserving elements allow changes of state to take place. Such elements include power-continuous generalized transformer (TF) and gyrator (GY) elements that algebraically relate elements of the effort and flow vectors into and out of the element. In certain cases, such as large motion of rigid bodies in which coordinate transformations are functions of the geometric state, the constitutive laws of these power-conserving elements can be state-modulated. Generalized series and parallel connections are represented by 1 and 0 junctions. All elements bonded to a 1-junction have common flow, and their efforts sum algebraically to zero. All elements bonded to a 0-junction have common effort, and all flows algebraically sum to zero. Sources represent ports through which the system interacts with its environment. See for example Figure 1 where the effort source represents either the force source or battery, and generalized flow associated with the 1-junction is either velocity or current.

The power conserving bond graph elements—TF, GY, 1-junctions, 0-junctions, and the bonds that connect them —are collectively referred to as “junction structure”. Figure 2 defines the symbols and constitutive laws of sources, storage and dissipative elements, and powerconserving elements in scalar form. Bond graphs may also be constructed with the constitutive laws and junction structure in matrix-vector form, in which case the bond is indicated by a double-line. Power bonds contain a half-arrow that indicates the direction of algebraically positive power flow, and a causal stroke normal to the bond that indicates whether the effort or flow variable is the input or output from the constitutive law of the connected elements. See Figure 3 where the effect of causal stroke location on the constitutive law form of two generic elements A and B is illustrated.

The constitutive laws in Figure 2 are consistent with the placement of the causal strokes. Full arrows are reserved for modulating signals that represent powerless information flow such as orientation angles that determine the transformation matrix between a body-fixed and inertial reference frame. Bond graphs, because they use the same small set of symbols for energy storage, dissipate and exchange in any energy domain (electrical, mechanical, hydraulic, etc.), the assembly of submodels from various disciplines is straightforward. For example, connecting a motor model to a linkage model requires simply “bonding” the 1-junction for motor output rotational speed to the 1-junction for linkage input link rotational speed.

Causality (equation input-output structure) automatically propagates through the entire model upon assembly of submodels. The graphical representation of causality

Figure 1. Example bond graph.

Figure 2. Bond graph symbols.

Figure 3. Bond graph causality.

allows for visual detection of input-output conflicts between subsystems, and immediately indicates any numerical simulation issues such as algebraic loops or differential-algebraic equations. The reader is supposed to refer to [18] for more details on bond graph modeling.

The truck model is implemented using bond graphs in this paper, and uses 20 sim [19] commercial software to enter the graph, generate and simulate equations of motion, and plot results. The implementation could be done (albeit with more tedious equation derivation) in software such as Matlab.

4. Vehicle Model

Figure 4 shows a schematic of the pitch plane vehicle model. The rigid aspects of the model are based on [20] and [13]. The model permits large angular motions of the sprung mass and uses nonlinear constitutive laws for aerodynamic drag and tire slip and rolling resistance. The aerodynamic drag constitutive law assumes that drag coefficient and frontal area are constant, and the effect of crosswinds is not considered. After the rigid model aspects are reviewed, the superposition of transverse beam vibrations onto the motion of the rigid frame will be described.

Conflicts of Interest

The authors declare no conflicts of interest.

References

[1] R. V. Field, et al., “Structural Dynamics Modeling and Testing of the Department of Energy Tractor/Trailer Combination,” Sandia National Laboratories Report SAND—96-2576C; CONF-970233, International Modal Analysis Conference, Orlando, 1997.
[2] R. V. Field, et al., “Analytical and Experimental Assessment of Heavy Truck Ride,” Sandia National Laboratories Report SAND—97-2667C; CONF-980224, International Modal Analysis Conference, Santa Barbara, 1998.
[3] M. Ahmadian and P. Patricio, “Dynamic Influence of Frame Stiffness on Heavy Truck Ride Evaluation,” SAE Paper 2004-01-2623, Society of Automotive Engineers, Warrendale, 2004. doi:10.4271/2004-01-2623
[4] D. Margolis and D. Edeal, “Modeling and Control of Large Flexible Frame Vehicles Using Bond Graphs,” SAE Paper 892488, Society of Automotive Engineers, Warrendale, 1989. doi:10.4271/892488
[5] A. Dhir, “Nonlinear Ride Analysis of Heavy Vehicle Using Local Equivalent Linearization Technique,” International Journal of Vehicle Design, Vol. 13, No. 5, 1992, pp. 580-606.
[6] D. Margolis and D. Edeal, “Towards an Understanding of ‘Beaming’ in Large Trucks,” SAE Paper 902285, Society of Automotive Engineers, Warrendale, 1990.
[7] A. Costa Neto, et al., “A Study of Vibrational Behavior of a Medium Sized Truck Considering Frame Flexibility with the Use of ADAMS,” Proceedings of 1998 International ADAMS User Conference, Ann Arbor, 1998.
[8] A. Goodarzi and A. Jalali, “An Investigation of Body Flexibility Effects on the Ride Comfort of Long Vehicles,” Proceedings of CSME Canadian Congress of Applied Mechanics, CANCAM 2006, 2006.
[9] I. M. Ibrahim, et al., “Effect of Frame Flexibility on the Ride Vibration of Heavy Trucks,” Computers and Structures, Vol. 58, No. 4, 1996, pp. 709-713. doi:10.1016/0045-7949(95)00198-P
[10] I. M. Ibrahim, “A Generally Applicable 3D Truck Ride Simulation with Coupled Rigid Bodies and Finite Element Models,” International Journal of Heavy Vehicle Systems, Vol. 11, No. 1, 2004, pp. 67-85. doi:10.1504/IJHVS.2004.004032
[11] D. G. Rideout, J. L. Stein and L. S. Louca, “System Partitioning and Improved Bond Graph Model Reduction Using Junction Structure Power Flow,” Proceedings of ICBGM’05, International Conference on Bond Graph Modeling, New Orleans, 2005, pp. 43-50.
[12] D. G. Rideout and J. L. Stein, “Breaking Subgraph Loops for Bond Graph Model Partitioning,” Proceedings of ICBGM’07, International Conference on Bond Graph Modeling, San Diego, 2007, pp. 241-249.
[13] D. G. Rideout, J. L. Stein and L. S. Louca, “Extension and Application of an Algorithm for Systematic Identification of Weak Coupling and Partitions in Dynamic System Models,” Simulation Modelling Practice and Theory, Vol. 17, 2009, pp. 271-292. doi:10.1016/j.simpat.2007.10.004
[14] P. Michelberger, et al. “Dynamic Modelling of Commercial Road Vehicle Structures from Test Data,” SAE Paper 845120, Society of Automotive Engineers, Warrendale, 1984. doi:10.4271/845120
[15] T. Y. Yi, “Vehicle Dynamic Simulations Based on Flexible and Rigid Multibody Models,” SAE Paper 2000- 01-0114, Society of Automotive Engineers, Warrendale, 2000.
[16] C. Cao, “Approaches to Reduce Truck Beaming,” SAE Paper 2005-01-0829. Society of Automotive Engineers, Warrendale, 2005. doi:10.4271/2005-01-0829
[17] J. Aurell, “The Influence of Warp Compliance on the Handling and Stability of Heavy Commercial Vehicles,” Proceedings of AVEC 2002, Hiroshima, 2002.
[18] D. Karnopp, et al. “System Dynamics—Modeling and Simulation of Mechatronic Systems,” 4th Edition, John Wiley and Sons, New York, 2006.
[19] 20sim v.4.1.3.8 (2011), Controllab Products b.v., Enschede, 2011.
[20] L. S. Louca, et al., “Generating Proper Dynamic Models for Truck Mobility and Handling,” International Journal of Heavy Vehicle Systems, Vol. 11, No. 3-4, 2004, pp. 209-236. doi:10.1504/IJHVS.2004.005449
[21] D. Karnopp and D. Margolis, “Analysis and Simulation of Planar Mechanism Systems Using Bond Graphs,” Journal of Dynamic Systems, Measurement, and Control, Vol. 101, No. 2, 1979, pp. 187-191.
[22] S. S. Rao, “Mechanical Vibrations,” 4th Edition, Pearson-Prentice Hall, Upper Saddle River, 2004.
[23] H. Lee, “New Dynamic Modeling of Flexible-Link Ro- bots,” Journal of Dynamic Systems, Measurement, and Control, Vol. 127, No. 2, 2005, pp. 307-309. doi:10.1115/1.1902843
[24] A. Yigit, et al. “Flexural Motion of a Radially Rotating Beam Attached to a Rigid Body,” Journal of Sound and Vibration, Vol. 121, No. 2, 1988, pp. 201-210. doi:10.1016/S0022-460X(88)80024-5
[25] L. S. Louca, J. L. Stein and D. G. Rideout, “Generating Proper Integrated Dynamic Models for Vehicle Mobility Using a Bond Graph Formulation,” Proceedings of International Conference on Bond Graph Modeling ICBGM’01, Phoenix, 2001, pp. 339-345.

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.