logo资料库

Nonlinear model predictive control for autonomous vehicles-Falco....pdf

第1页 / 共15页
第2页 / 共15页
第3页 / 共15页
第4页 / 共15页
第5页 / 共15页
第6页 / 共15页
第7页 / 共15页
第8页 / 共15页
资料共15页,剩余部分请下载后查看
566 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 15, NO. 3, MAY 2007 Predictive Active Steering Control for Autonomous Vehicle Systems Paolo Falcone, Francesco Borrelli, Jahan Asgari, Hongtei Eric Tseng, and Davor Hrovat, Fellow, IEEE Abstract—In this paper, a model predictive control (MPC) approach for controlling an active front steering system in an autonomous vehicle is presented. At each time step, a trajectory is assumed to be known over a finite horizon, and an MPC controller computes the front steering angle in order to follow the trajectory on slippery roads at the highest possible entry speed. We present two approaches with different computational complexities. In the first approach, we formulate the MPC problem by using a nonlinear vehicle model. The second approach is based on suc- cessive online linearization of the vehicle model. Discussions on computational complexity and performance of the two schemes are presented. The effectiveness of the proposed MPC formulation is demonstrated by simulation and experimental tests up to 21 m/s on icy roads. Index Terms—Active steering, autonomous vehicles, model pre- dictive control, nonlinear optimization, vehicle dynamics control, vehicle stability. I. INTRODUCTION R ECENT trends in automotive industry point in the di- rection of increased content of electronics, computers, and controls with emphasis on the improved functionality and overall system robustness. While this affects all of the vehicle areas, there is a particular interest in active safety, which effectively complements the passive safety counterpart. Passive safety is primarily focused on the structural integrity of vehicle. Active safety on the other hand is primarily used to avoid accidents and at the same time facilitate better vehicle controllability and stability especially in emergency situations, such as what may occur when suddenly encountering slippery parts of the road [10]. Early works on active safety systems date back to the 1980s and were primarily focused on improving longitudinal dy- namics part of motion, in particular, on more effective braking (ABS) and traction control (TC) systems. ABS systems in- crease the braking efficiency by avoiding the lock of the braking wheels. TC systems prevent the wheel from slipping and at the same time improves vehicle stability and steerability by maxi- mizing the tractive and lateral forces between the vehicle’s tire and the road. This was followed by work on different vehicle Manuscript received November 10, 2006. Manuscript received in final form January 12, 2007. Recommended by Associate Editor K. Fishbach. P. Falcone and F. Borrelli are with the Universitá del Sannio, Dipartimento di Ingegneria, Università degli Studi del Sannio, 82100 Benevento, Italy (e-mail: falcone@unisannio.it; francesco.borrelli@unisannio.it). J. Asgari, H. E. Tseng, and D. Hrovat are with Research and Innovation Center, Ford Research Laboratories, Dearborn, MI 48124 USA (e-mail: jas- gari@ford.com; htseng@ford.com; dhrovat@ford.com). Color versions of Figs. 1, 2, and 4–12 are available online at http://ieeexplore. ieee.org. Digital Object Identifier 10.1109/TCST.2007.894653 stability control systems [34] (which are also known under different acronyms such as electronic stability program (ESP), vehicle stability control (VSC), interactive vehicle dynamics (IVD), and dynamic stability control (DSC)). Essentially, these systems use brakes on one side and engine torque to stabilize the vehicle in extreme limit handling situations through con- trolling the yaw motion. In addition to braking and traction systems, active front steering (AFS) systems make use of the front steering com- mand in order to improve lateral vehicle stability [1], [2]. Moreover, the steering command can be used to reject ex- ternal destabilizing forces arising from -split, asymmetric braking, or wind [21]. Four-wheel steer (4WS) systems follow similar goals. For instance, in [3], Ackermann et al. present a decoupling strategy between the path following and external disturbances rejection in a four-wheel steering setup. The automatic car steering is split into the path following and the yaw stabilization tasks, the first is achieved through the front steering angle, the latter through the rear steering angle. Research on the AFS systems has also been approached from an autonomous vehicle perspective. In [16], an automatic steering control for highway automation is presented, where the vehicle is equipped with magnetic sensors placed on the front and rear bumpers in order to detect a lane reference im- plemented with electric wire [13] and magnetic markers [36]. A more recent example of AFS applications in autonomous vehicles is the “Grand Challenge” race driving [5], [23], [30]. In this paper, it is anticipated that the future systems will be able to increase the effectiveness of active safety interventions beyond what is currently available. This will be facilitated not only by additional actuator types such as 4WS, active steering, active suspensions, or active differentials, but also by additional sensor information, such as onboard cameras, as well as in- frared and other sensor alternatives. All these will be further complemented by global positioning system (GPS) information including prestored mapping. In this context, it is possible to imagine that future vehicles would be able to identify obstacles on the road such as an animal, a rock, or fallen tree/branch, and assist the driver by following the best possible path, in terms of avoiding the obstacle and at the same time keeping the vehicle on the road at a safe distance from incoming traffic. An additional source of information can also come from surrounding vehicles and environments which may convey the information from the vehicle ahead about road condition, which can give a significant amount of preview to the controller. This is particular is useful if one travels on snow or ice covered surfaces. In this case, it is very easy to reach the limit of vehicle handling capabilities. Anticipating sensor and infrastructure trends toward in- creased integration of information and control actuation agents, 1063-6536/$25.00 © 2007 IEEE
FALCONE et al.: PREDICTIVE ACTIVE STEERING CONTROL FOR AUTONOMOUS VEHICLE SYSTEMS 567 it is then appropriate to ask what is the optimum way in con- trolling the vehicle maneuver for a given obstacle avoidance situation. We assume that a trajectory planning system is available and we consider a double lane change scenario on a slippery road, with a vehicle equipped with a fully autonomous guidance system. In this paper, we focus on the control of the yaw and lateral vehicle dynamics via active front steering. The control input is the front steering angle and the goal is to follow the desired trajectory or target as close as possible while fulfilling various constraints reflecting vehicle physical limits and design requirements. The future desired trajectory is known only over finite horizon at each time step. This is done in the spirit of model predictive control (MPC) [14], [26] along the lines of our ongoing internal research efforts dating from early 2000 (see [7] and references therein). In this paper, two different formulations of the AFS MPC problem will be presented and compared. The first one follows the work presented in [7] and uses a nonlinear vehicle model to predict the future evolution of the system [26]. The resulting MPC controller requires a nonlinear optimization problem to be solved at each time step. We will show that the computa- tional burden is currently an obstacle for experimental valida- tion at high vehicle speed. The second formulation tries to over- come this problem and presents a suboptimal MPC controller based on successive online linearization of the nonlinear vehicle model. This is linearized around the current operating point at each time step and a linear MPC controller is designed for the re- sulting linear time-varying (LTV) system. The idea of using time varying models goes back to the early 1970s in the process con- trol field although it has been properly formalized only recently. Studies on linear parameter varying (LPV) MPC schemes can be found in [9], [18], [20], [22], and [35]. Among them, the work in [18] and [20] is the closest to our approach and it presents an MPC scheme for scheduled LTV models which has been successfully validated on a Boeing aircraft. In general, the per- formance of such a scheme is highly dependant on the nonlin- earities of the model. In fact, as the state and input trajectories deviate from the current operating point, the model mismatch increases. This can generate large prediction errors with a con- sequent instability of the closed-loop system. We will show that, in our application, a state constraint can be introduced in order to significantly enhance the performance of the system. Exper- imental results show that the vehicle can be stabilized up to 21 m/s on icy roads. Finally, an LTV MPC with a one-step con- trol horizon is presented. This can be tuned in order to provide acceptable performance and it does not require any complex op- timization software. We implemented the MPC controllers in real time on a pas- senger car, and performed tests on snow covered and icy roads. The last part of this paper describes the experimental setup and presents the experimental and simulation results of the proposed MPC controllers. It should be noted that our early work in [7] focuses on the vehicle dynamical model and on simulation re- sults of the nonlinear MPC scheme only. This paper is structured as follows. Section II describes the used vehicle dynamical model with a brief discussion on tire models. Section III introduces a simplified hierarchical Fig. 1. Simplified vehicle “bicycle model.” framework for autonomous vehicle guidance. The contribution and the research topic of this paper are described in details and put in perspective with existing work and future research. Section IV formulates the control problem when the nonlinear and the linear prediction models are used. The double lane change scenario is described in Section V, while in Section VI, the experimental and simulation results are presented. This is then followed by concluding remarks in Section VII which highlight future research directions. II. MODELING and are the forces in car body frame, This section describes the vehicle and tire model used for simulations and control design. This section has been extracted from [7] and it is included in this paper for the sake of complete- the longitudinal (or ness and readability. We denote by “tractive”) and lateral (or “cornering”) tire forces, respectively, is the normal tire and are the absolute car position are the car geometry (distance of is the gravita- is the is the car mass, are the longitudinal and lateral wheel veloc- are the local lateral and longitudinal coordinates is the slip angle, is the road friction coefficient, load, inertial coordinates, front and rear wheels from center of gravity), tional constant, and slip ratio, ities, in car body frame, is the wheel steering angle, is the vehicle speed, is the wheel radius, and is the car inertia, and and is the heading angle. The lower subscripts and particularize a variable at the front wheel and the rear wheel, respectively, e.g., is the front wheel longitudinal force. and A. Vehicle Model car under the assumption of a constant tire normal load, i.e., A “bicycle model” [25] is used to model the dynamics of the , . Fig. 1 depicts a diagram of the vehicle model, which has the following longitudinal, lateral, and turning or yaw degrees of freedom: (1a) (1b) (1c) The vehicle’s equations of motion in an absolute inertial frame are (2)
568 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 15, NO. 3, MAY 2007 The wheel’s equations of motion describe the lateral (or cor- nering) and longitudinal wheel velocities (3a) (3b) (3c) (3d) where tively, and and are front and rear wheel steering angle, respec- (4a) (4b) The following equations hold for rear and front axes by using the corresponding subscript for all the variables. Longitudinal and lateral tire forces lead to the following forces acting on the center of gravity: Tire forces and for each tire are given by (5a) (6) , , , and are defined next. The tire slip angle where and represents the angle between the wheel velocity vector the direction of the wheel itself, and can be compactly expressed as The slip ratio is defined as if if for braking for driving (7) (8) and where are the radius and the angular speed of the wheel, represents the road friction coef- respectively. The parameter ficient and is assumed equal for front and rear wheels. is the total vertical load of the vehicle and is distributed between the front and rear wheels based on the geometry of the car model (described by the parameters and ) (9) The nonlinear vehicle dynamics described in (1)–(9), can be rewritten in the following compact from: (10) and friction coefficient at each time instant has been explicitly highlighted. The , is assumed to be zero at any time where the dependence on slip ratio value state and input vectors are respectively. In this paper, instant. and Model (10) captures the most relevant nonlinearities associ- ated to lateral stabilization of the vehicle. Section II-B briefly describes the models of tire forces and . Fig. 2. Longitudinal and lateral tire forces with different  coefficient values. B. Tire Model With the exception of aerodynamic forces and gravity, all of the forces which affect vehicle handling are produced by the tires. Tire forces provide the primary external influence and, be- cause of their highly nonlinear behavior, cause the largest vari- ation in vehicle handling properties throughout the longitudinal and lateral maneuvering range. Therefore, it is important to use a realistic nonlinear tire model, especially when investigating large control inputs that result in response near the limits of the maneuvering capability of the vehicle. In such situations, the lateral and longitudinal motions of the vehicle are strongly cou- pled through the tire forces, and large values of slip ratio and slip angle can occur simultaneously. Most of the existing tire models are predominantly “semi-em- pirical” in nature. That is, the tire model structure is determined through analytical considerations, and key parameters depend on tire data measurements. Those models range from extremely simple (where lateral forces are computed as a function of slip and one measured angle, based on one measured slope at value of the maximum lateral force) to relatively complex al- gorithms, which use tire data measured at many different loads and slip angles. In this paper, we use a Pacejka tire model [4] to describe the tire longitudinal and cornering forces in (6). This is a complex, semi-empirical nonlinear model that takes into consideration the interaction between the longitudinal force and the cornering force in combined braking and steering. The longitudinal and cornering forces are assumed to depend on the normal force, slip angle, surface friction coefficient, and longitudinal slip. Fig. 2 depicts longitudinal and lateral forces versus longitudinal slip and slip angle, for fixed values of the friction coefficients. We remark that the front tire of the “bicycle” model represents the two front tires of the actual car.
FALCONE et al.: PREDICTIVE ACTIVE STEERING CONTROL FOR AUTONOMOUS VEHICLE SYSTEMS 569 We remark that the scheme in Fig. 3 is an oversimplified scheme and that additional hierarchical levels could be present both in the trajectory/mode replanning module and in the low-level control system module. The union of the first three modules is often referred to as guidance and navigation control (GNC) system. Typically, the trajectory replanner and the low-level control system modules do not share the same information on envi- ronment and vehicle. For instance, the replanning algorithms can use information coming from cameras or radars which may not be used at the lower level. Also, typically, the frequency at which the trajectory replanning module is executed is lower than the one of the lower level control system. The design of both modules makes use of vehicle and environment models with different levels of detail. The fidelity of the dynamical model used for the design of the two modules is dictated, among many factors, by a performance/computational resource compromise and, in the literature, there is no accepted standard on this. One of the possible control paradigms for the two modules con- sists in using a high-fidelity vehicle model for designing the lower level controller while the trajectory planner relies on a rougher/less detailed dynamical model of the vehicle. Clearly, the higher the fidelity of the models used at the higher level is, the easier the job for the lower level control algorithm becomes. Studies on GNC algorithms vary in 1) the focus (trajectory replanner and/or the low-level control system); 2) the type of vehicle dynamical model used; 3) the type of control design used; and 4) inputs and sensors choice. In [23], the trajectory replanner module is based on a receding horizon control design. The planning problem is formulated as a constrained optimization problem minimizing a weighted sum of arrival time, steering, and acceleration control efforts. The vehicle model is a simple rear-centered kinematic model with acceleration, speed, steering, steering rate, and rollover con- straints. The lower level control module uses two separated pro- portional–integral–differentials (PIDs) to control longitudinal and lateral dynamics. The longitudinal controller acts on throttle and brakes while the lateral controls on the steering angle. The GNC architecture in [30] is similar to [23]. The trajectory planning task is posed as a constrained optimization problem. The cost function penalizes obstacles collision, distance from the precomputed offline trajectory and the lateral offset from the current trajectory. At the lower level, a PI controller acts on brakes and throttle to control the longitudinal dynamics. A simple nonlinear controller, instead, is used to control the lateral dynamics through the steering angle. Details on the vehicle dy- namical model used in [30] are not disclosed. In [29], a scheme similar to the one in [23] is used to design a GNC systems for a flight control application. In [33], an explicit MPC scheme has been applied at the lower level control to allocate four wheel slips in order to get a desired yaw moment. The steering angle is not controlled. In this paper, we assume that the path can be generated with two different methods. In the first, the trajectory is established by simply driving a test vehicle slowly along the desired path, e.g., a double lane change manoeuvre. The actual path is recorded by differential GPS and then used as a desired path for subsequent tests at higher speed. This method has been used Fig. 3. Simplified architecture for fully autonomous vehicle guidance system. III. HIERARCHICAL FRAMEWORK FOR AUTONOMOUS GUIDANCE In this section, we borrow the simplified schematic architec- ture in Fig. 3 from the aerospace field [8], [24], [31], in order to explain our approach and contribution. The architecture in Fig. 3 describes the main elements of an autonomous vehicle guidance system and it is composed of four modules: the trajectory/mode generator, the trajectory/mode replanning, the low-level con- trol system, and the vehicle and the environmental model. The trajectory/mode planning module precomputes offline the ve- hicle trajectory together with the timing and conditions for op- eration mode change. In the aerospace field, examples of op- eration mode selection include aeroshell parachute deployment or heatshield release, in the automotive field this could include switching between two or more types of energy sources (i.e., gas, electricity, hydrogen) or (in a very futuristic scenario) mor- phing between different vehicle shapes. The trajectory and the mode of operation computed offline can be recomputed online during the drive by the trajectory/ mode replanning module based on current measurements, at fixed points or on the occurrence of certain events (such as tracking errors exceeding certain bounds, hardware failure, ex- cessive wind, the presence of a pop-up obstacle). The low-level control system commands the vehicle actua- tors such as front and rear steering angles, four brakes, engine torque, active differential, and active suspensions based on sensor measurements, states, and parameters estimations and reference commands coming from the trajectory/mode replan- ning module. Such reference commands can include lateral and longitudinal positions, pitch, yaw, and roll rates. The low-level control system objective is to keep the vehicle as close as pos- sible to the currently planned trajectory despite measurement noise, unmodeled dynamics, parameteric uncertainties, and sudden changes on vehicle and road conditions which are not (or not yet) taken into account by the trajectory replanner. In particular, when a vehicle is operating near its stability limit, these additional noises, disturbances, and uncertainties must be considered, possibly through detecting the vehicle’s internal state, and compensated for. For example, if rear tires saturate, a skillful driver would switch his/her steering input from the usual steering command for trajectory following to a counter-steering one for stabilizing the vehicle. It is conceivable that an automated steering would not produce the necessary stabilizing counter-steer if the commanded steering is only a function of the desired trajectory and vehicle’s current position and heading (without considering additional vehicle dynamic states).
570 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 15, NO. 3, MAY 2007 in [34] to generate the reference path for a steering robot on high . In this case, no trajectory replanning is needed and the contribution of the work presented in this paper is to facilitate systematic and repeatable tests of safety critical emergency manoeuvres during limit conditions, such as obstacle avoidance manoeuvres on slippery surfaces, i.e., snow and ice. In the second method, we assume that a trajectory replanning module is available and the trajectory is recomputed at a less frequent rate than the frequency of the lower level controller. For both cases, we focus on the lower level control design by means of nonlinear and LTV MPC for the specific scenario of an active steering system. As suggested in [28], there is a significant challenge involved in obtaining the steering required to accomplish the limit ma- neuver considered in this paper while maintaining vehicle sta- bility. By focusing on the lower level MPC controller, we also believe that the resultant steering may mimic a skillful driver who takes the full vehicle dynamic states into account. Com- pared to the lower level control algorithms presented in the aforementioned literature, our approach 1) is model based and uses the vehicle model (10) and the highly nonlinear Pacejka tire model described in Section II-B; 2) includes constraints on in- puts and states in the control design; 3) is systematic and multi- variable and can accommodate new actuators and higher fidelity models. Moreover, we have experimentally validated the con- troller presented in this paper with a dSPACE AutoBox system which is a standard rapid prototyping system used in automo- tive industries [11]. nonlinear vehicle model (10) and the Pacejka tire model are used to predict the future evolution of the system. The mini- mization of a quadratic performance index, subject to the non- linear vehicle dynamics, is a nonlinear optimization problem. Such optimization problem is solved online, at each time step. This can be computationally demanding, depending on the ve- hicles states and constraints. The second formulation, presented in Section IV-B, tries to overcome this problem. A LTV approx- imation of vehicle model (10) and the Pacejka tire model are used to predict the future evolution of the system. This leads to a suboptimal LTV MPC controller. In this case, a time varying convex quadratic optimization problem is formulated and solved at each time step, leading to the reduction of the computational burden with an acceptable loss of performance. We will show that the MPC performance is enhanced by including a constraint on the tire slip angle which stabilizes the vehicle at high speed. A. Nonlinear (NL) MPC In order to obtain a finite-dimensional optimal control problem, we discretize the system dynamics (10) with a fixed sampling time where the formulation is used, with . (11a) (11b) and IV. ACTIVE STEERING CONTROLLER DESIGN position states: In this section, we introduce the control design procedure for the proposed path following problem via an active steering system. (12) We define the following output map for yaw angle and lateral , the yaw rate Desired references for the heading angle , define a desired path over a finite and the lateral distance horizon. The nonlinear vehicle dynamics (10) and the Pacejka tire model are used to predict the vehicles behavior, and the front is chosen as control input. The rear steering steering angle and angle is assumed to be zero is estimated at each time instant. The approach used in [6] can be used for the online estimation of are measured, and the road friction , the tire slip ratios . and consider the following cost function: (13) where denote the corresponding reference signal. At each time step , the following finite horizon optimal control problem is solved: and subj. to A MPC scheme is used to solve the path following problem. The main concept of MPC is to use a model of the plant to pre- dict the future evolution of the system [6], [14], [17], [26], [27]. At each sampling time, starting at the current state of the ve- hicle, an open-loop optimal control problem is solved over a fi- nite horizon. The open-loop optimal control problem minimizes the deviations of the predicted outputs from their references over a sequence of future steering angles, subject to operating con- straints. The resulting optimal command signal is applied to the process only during the following sampling interval. At the next time step, a new optimal control problem based on new mea- surements of the state is solved over a shifted horizon. In the following two different formulations of the AFS MPC problem will be presented. Section IV-A describes the first MPC formulation as presented in the preliminary work [7]. There, the (14a) (14b) (14c) (14d) (14e) (14f) (14g) (14h) (14i)
FALCONE et al.: PREDICTIVE ACTIVE STEERING CONTROL FOR AUTONOMOUS VEHICLE SYSTEMS 571 is the optimization where denotes the output vector predicted vector at time , , at time and applying to system (11) and (12) the input sequence denote the output prediction obtained by starting from the state and . horizon and the control horizon, respectively. We use and the control signal is assumed constant for all . We assume slip and friction coefficient values constant and over the prediction equal to the estimated values at time horizon [constraint (14d)]. In (13), the first summand reflects the penalty on trajectory tracking error while the second summand is a measure of the are weighting matrices of appropriate steering effort. dimensions. and We denote by the se- by . Then, the first is used to compute the optimal control action quence of optimal input increments computed at time solving (14) for the current observed states sample of and the resulting state feedback control law is (15) , the optimization problem (14) is At the next time step solved over a shifted horizon based on the new measurements of the state . B. LTV MPC Let be the current time and be the current state and the previous input of system (11) and (12), respec- tively. We consider the following optimization problem: and subj. to (16a) (16b) (16c) (16d) (16e) (16f) (16g) (16h) (16i) (16j) (16k) (16l) (16m) , around the point , for the estimated and and model (16b) and where (16c) is obtained by linearizing model (11) at each time step . The variables denote the outputs of the linearized system and the corresponding reference signal, denotes the tire slip angle varia- respectively. The variable tion and it is an additional output of the linearized model which is only constrained and not tracked. Inequalities (16j) are soft is a slack variable. The constraints on the tire slip angle and term in (16b) penalizes the violation of the constraint on the slip angle and is a weight coefficient. , The optimization problem (16) can be recast as a quadratic program (QP) (details can be found in [7]). We denote by the sequence of optimal input by solving (16) for the current deviations computed at time observed states is used to compute the optimal control action and the resulting state feedback control law is . Then, the first sample of (17) the optimization problem (16) is At the next time step solved over a shifted horizon based on the new measurements and based on an updated linear model of the state (16b)–(16d) computed by linearizing the nonlinear vehicle model (11) around the new state, slip ratio, road friction coeffi- cient, and previous input. We remark that model (11) is linearized around an operating point that, in general, is not an equilibrium point. Therefore, is the linear time-invariant (LTI) model (16b)–(16d) at time used to predict the state and the output deviations from the tra- , respec- jectories as initial condition tively, computed by solving (11) with . Accordingly, the op- and timization variables represent the input variation with respect to the previous input for for , , . and state trajectory Alternatively, the vehicle model (11) can be linearized around . In this a nominal input case, (16e) would become and the would represent optimization variables the input variations around the nominal input. This approach and requires a nominal input and state trajectory, i.e., . Such trajectory could be computed from the higher level replanning algorithm described in Section III or from the lower level MPC controller. We remark that in this case an LTV model over the prediction horizon ( ), , , , could be used at each time step instead of the LTI model (16b)–(16d). An MPC scheme similar to the one presented in this paper can be found in [18], [19], and [20]. In these works, a similar MPC formulation is used. An LTI prediction model is used to predict the behavior of the system over the prediction horizon. The LTI model is updated according to the values of flight con- dition dependent scheduling parameters. In [19] and [20], the LTI model is obtained by interpolation over a precomputed data- base of linearized models, while in [18] the LTI model is ob- tained by linearizing the nonlinear kinematics around the cur- rent measurements.
572 IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 15, NO. 3, MAY 2007 When evaluating the online computational burden of the pro- posed scheme, in addition to the time required to solve the op- timization problem (16), one needs to consider the resources ) in (16b) spent in computing the linear models ( and (c) and translating (16) into a standard quadratic program- ming (QP) problem. Nevertheless, for the proposed application, complexity of the MPC (16) and (17) greatly reduces compared to the NL MPC presented in Section IV-A. This will be shown for a specific scenario in Sections VI-A and VI-B. , , , The stability of the presented control scheme is difficult to prove even under no model mismatch and it is a topic of cur- rent research. Also, robustness of nonlinear MPC schemes is an active area of research by itself. An analytical and meaningful study of the robustness of the proposed scheme would be even more difficult. The uncertainty of the tire characteristics and the road condition are often difficult to describe with a mathemat- ical formalism which is realistic and not too conservative. It should be noted that in the MPC scheme (16) and (17), the introduction of the state constraints (16j) is needed in order to obtain an acceptable performance and it is a contribution of this paper. As shown next in Section VI-A, such constraint arises from a careful study of the closed-loop behavior of the non- linear MPC presented in Section IV-A. In fact, extensive sim- ulations have shown that the nonlinear MPC never exceeds cer- tain tire slip angles under stable operations. By removing the constraints (16j) the performance of the LTV MPC controller (16) and (17) is not acceptable and the system becomes un- stable at high vehicle speeds. In fact, a simple linear model is not able to predict the change of slope in the tires characteristic (see Fig. 2). To overcome this issue, we included constraints (16j) in the optimization problem, in order to forbid the system from entering a strongly nonlinear and possibly unstable re- gion of the tire characteristic. In particular, by looking at the tire characteristics in Fig. 2, it is clear that a linear approxima- tion of the tire model around the origin is no longer valid if the slip angle exceeds certain bounds. Led by this observation and by a study on the closed-loop behavior of the nonlinear MPC presented in Section IV-A, we included the constraints (16j) in the optimization problem. In particular, for a given , the tire slip angle is constrained in the mostly linear region of the lat- eral tire force characteristic. By no means does the constraints (16j) enforce the dynamical system to operate in a linear region: system nonlinearities (11) and longitudinal tire nonlinearities are still relevant when constraints (16j) are included in the MPC formulation. Note that the constraints (16j) are implicit linear constraints on state and input and they can be handled systematically only in an MPC scheme. A soft constraint formulation is preferred to a hard constraint in order to avoid infeasibility. In fact, during experiments the tire slip angle is estimated from IMU and GPS measurements. Acceleration measurements are noisy and the GPS signal can be lost. Moreover, as shown in (3) and (7), the tire slip angle depends on the steering angle. The latter, as ex- plained in Section V-B, in our experimental setup is affected by the driver’s imposed steering angle. An additional tracking error on yaw rate is included in the per- formance index of the LTV MPC problem (16) and (17) (com- pare (16d) to (12)). Extensive simulations have shown that this additional term significantly improves the performance of the LTV MPC controller (16) and (17). V. DOUBLE LANE CHANGE ON SNOW USING ACTIVE STEERING The MPC steering controllers described in Sections IV-A and IV-B have been implemented to perform a sequence of double lane changes at different entry speeds. This test represents an obstacle avoidance emergency maneuver in which the vehicle is entering a double lane change maneuver on snow or ice with a given initial forward speed. The control input is the front steering angle and the goal is to follow the trajectory as close as possible by minimizing the vehicle deviation from the target path. The experiment is repeated with increasing entry speeds until the vehicle loses control. The same controller can be used to control the vehicle during different maneuvers in different scenarios [21]. The simulation and experimental results will be presented in Section VI. Next, we describe the reference generation and present the experimental setup in Section V-B. A. Trajectory Generation The desired path is described in terms of lateral position as function of the longitudinal position and yaw angle (18a) (18b) , where , and , 2, . , , The reference trajectories (18a) and (18b), can be used di- rectly only in the nonlinear MPC formulation, being a nonlinear . In the LTV MPC for- function of the longitudinal distance mulation, we generate the reference trajectories from (18a) and (18b) by assuming that the vehicle will travel a portion of the desired path at a constant speed in the next steps. Because of the assumption on constant travel velocity, the method for generating the previously described trajectory can affect the performance of the closed-loop system. In particular, in extreme handling situations, when the tracking errors are large due to spinning or side skidding, the computed reference could lead to aggressive maneuvers. As explained in Section III, more accurate methods could be used in order to generate a smoother reference for the LTV MPC scheme by taking into account the state of the vehicle. B. Experimental Setup Description The MPC controllers presented in Sections IV-A and IV-B have been tested through simulations and experiments on slip- pery surfaces. The experiments have been performed at a test center equipped with icy and snowy handling tracks. The MPC controllers have been tested on a passenger car, with a mass of 2050 Kg and an inertia of 3344 kg/m . The controllers were run
FALCONE et al.: PREDICTIVE ACTIVE STEERING CONTROL FOR AUTONOMOUS VEHICLE SYSTEMS 573 in a dSPACE Autobox system, equipped with a DS1005 pro- cessor board and a DS2210 I/O board, with a sample time of 50 ms. We used an Oxford Technical Solution (OTS) RT3002 sensing system to measure the position and the orientation of the vehicle in the inertial frame and the vehicle velocities in the vehicle body frame. The OTS RT3002, is housed in a small package that contains a differential GPS receiver, inertial measurement unit (IMU), and a DSP. It is equipped with a single antenna to receive GPS information. The IMU includes three accelerometers and three angular rate sensors. The DSP receives both the measurements from the IMU and the GPS, utilizes a Kalman filter for sensor fusion, and calculates the position, the orientation, and other states of the vehicle such as longitudinal and lateral velocities. The car was equipped with an AFS system which utilizes an electric drive motor to change the relation between the hand steering wheel and road wheel angles. This is done indepen- dently from the hand wheel position, thus the front road wheel angle is obtained by summing the driver hand wheel position and the actuator angular movement. Both the hand wheel posi- tion and the angular relation between hand and road wheels are measured. The sensor, the dSPACE Autobox, and the actuators communicate through a CAN bus. The autonomous steering test is initiated by the driver with a button. When the button is pushed, the inertial frame in Fig. 1 is initialized as follows: the origin is the current vehicle position, are directed as the current longitudinal and the axes lateral vehicle axes, respectively. Such inertial frame becomes also the desired path coordinate system. Once the initialization procedure is concluded, the vehicle executes the double lane change maneuver. and During the experiment, the hand wheel may deviate from its center position. This is caused by the difficulty the driver can have in holding the steering still, which was needed to facil- itate autonomous behavior with that particular test vehicle. In our setup, this is treated as a small bounded input disturbance. Furthermore, noise may affect the yaw angle measurement due to the single antenna sensor setup. Compared to a dual antenna setup, a single antenna system has to learn the vehicle orienta- tion and/or coordinate during vehicle motion. When the vehicle stands still the yaw angle is computed by integrating the yaw rate measurement from the IMU. This might cause the presence of a small offset in the orientation measurement, while traveling at low speed or being still. The effects of both input disturbance and measurement noise will be clear later in the presented ex- perimental results. VI. PRESENTATION AND DISCUSSION OF RESULTS In Section VI-A, three types of MPC controllers will be presented. These controllers have been derived by the MPC problem formulations presented in Sections IV-A and IV-B and will be referred to as Controller A, B, and C. • Controller A: Nonlinear MPC (14) and (15) with the fol- lowing parameters: • 0.05 s, , • , ; , ; . , , ; MAXIMUM COMPUTATION TIME OF CONTROLLERS A AND B PERFORMING A DOUBLE LANE CHANGE MANEUVER AT DIFFERENT VEHICLE SPEEDS TABLE I • Controller B: LTV MPC (16) and (17) with the following parameters: • 0.05 s, , , ; , , , ; • weighting matrices , , , , . . • Controller C: Same as Controller B with Next, the results obtained with the three controllers will be described and a comparison between the simulation and the ex- perimental results will be given for each of them. The actual road friction coefficient was set manually and constant for each experiment depending on the road conditions. This choice was driven by the study of the controller closed-loop perfor- estimation and its associated mance independently from the error and dynamics. For each controller more simulation, exper- iments, and comments can be found in [12]. A. Controller A The controller (14) and (15) with the parameters defined in Section VI has been implemented as a C-coded S-function in which the commercial NPSOL software package [15] is used for solving the nonlinear programming problem (14). The choice of NPSOL has been motivated by its performance and the avail- ability of the source C code. Limited by the computational complexity of the nonlinear programming solver and the hardware used, we could perform experiments at low vehicle speeds only. In fact, as the entry speed increases, larger prediction and control horizons are re- quired in order to stabilize the vehicle along the path. Larger pre- diction horizons involve more evaluations of the objective func- tion, while larger control horizons imply a larger optimization problem (14). In Table I, we report the maximum computation time required by the Controllers A and B to compute a solution to the problems (14) and (16), respectively, when the maneuver described in Section V is performed at different vehicle speeds. The selected control and prediction horizons in Table I are the shortest allowing the stabilization of the vehicle at each speed. The results have been obtained in simulation with a 2.0-GHz Centrino-based laptop running Matlab 6.5. During experiments, the maximum iterations number in NPSOL bas been limited in order to guarantee real-time com- putation. The bound was selected after preliminary tests on the real-time hardware. In Fig. 4, the simulation results for a maneuver at 7 m/s are presented. In Fig. 5, the corresponding experimental results are presented. In the upper plot of Fig. 5(b), the dashed line rep- resents the steering action from the driver (i.e., the input dis- turbance) that, in this test, is negligible. The actual road wheel
分享到:
收藏