Next Article in Journal
Detail Explanation of Coordinate Transformation Procedures Used in Geodesy and Geoinformatics on the Example of the Territory of the Republic of Croatia
Previous Article in Journal
Basic Composition, Antioxidative Properties, and Selected Mineral Content of the Young Shoots of Nigella (Nigella sativa L.), Safflower (Carthamus tinctorius L.), and Camelina (Camelina sativa L.) at Different Stages of Vegetation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Path Tracking Control with Constraint on Tire Slip Angles under Low-Friction Road Conditions

1
Department of Intelligent Systems and Robotics, Chungbuk National University, Cheongju 28644, Chungbuk, Republic of Korea
2
Research Center for Electrical and Information Technology, Seoul National University of Science and Technology, 232 Gongneung-ro, Nowon-gu, Seoul 01811, Republic of Korea
*
Author to whom correspondence should be addressed.
Appl. Sci. 2024, 14(3), 1066; https://doi.org/10.3390/app14031066
Submission received: 26 December 2023 / Revised: 17 January 2024 / Accepted: 24 January 2024 / Published: 26 January 2024
(This article belongs to the Special Issue Advances in Vehicle System Dynamics and Control)

Abstract

:

Featured Application

Path Tracking Control for Autonomous Driving.

Abstract

This paper presents a method to design a path tracking controller with a constraint on tire slip angles under low-friction road conditions. On a low-friction road surface, a lateral tire force is easily saturated and decreases as a tire slip angle increases by a large steering angle. Under this situation, a path tracking controller cannot achieve its maximum performance. To cope with this problem, it is necessary to limit tire slip angles to a value where the maximum lateral tire force is achieved. The most commonly used controllers for path tracking, linear quadratic regulator (LQR) and model predictive control (MPC), are adopted as a controller design methodology. The control inputs of LQR and MPC are front and rear steering angles and control yaw moment, which have been widely used for path tracking. The constraint derived from tire slip angles is imposed on the steering angles of LQR and MPC. To fully verify the performance of the path tracking controller with the constraint on tire slip angles, a simulation is conducted on vehicle simulation software. From the simulation results, it is shown that the path tracking controller with the constraint on tire slip angles presented in this paper is quite effective for path tracking on low-friction road surface.

1. Introduction

Over the past decade, research groups and the automobile industry have intensively explored autonomous driving because it is predicted to have impacts on several aspects of driving such as traffic flow, ride comfort, and road safety [1,2,3]. With respect to safety, autonomous driving has been expected to reduce social costs and damages incurred from drivers by preventing traffic accidents. As a result, a huge amount of papers on autonomous driving have been published in the meantime. According to the survey, it has been widely known that autonomous driving operations include perception, localization, planning, control and navigation, and system management [3]. Among them, planning or path planning generates a reference or target path based on information obtained from environments, and path tracking controls a vehicle, following the reference states with several actuators. This paper focuses on path tracking control (PTC). For the last decade, PTC has been intensively explored [4,5,6]. Consequently, a huge deal of papers have been published to date in the field of PTC [4,5,6,7,8,9,10].
To date, most path tracking controllers for autonomous driving have been designed on a high-friction road surface where the tire-road friction coefficient, μ, was set between 0.8 and 1.0. Since 2019, some papers concentrating on PTC under low-friction road conditions have been published [11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]. These papers are classified according to controller design methodologies, input configurations, actuator combinations and tire-road friction coefficient, as given in Table 1. As shown in Table 1, the lower limits of μ were set between 0.25 and 0.4. In this paper, a low friction road condition means that μ is 0.4 or less.
In Table 1, δf, δr, ∆Fx and ∆Mz are the control inputs, which represent the front and rear steering angles, the control longitudinal force, and the control yaw moment, respectively. An input configuration is composed of those control inputs. In Table 1, seven input configurations were given. With those input configurations, a linear quadratic regulator (LQR), model predictive control (MPC), and sliding mode control (SMC) were adopted as a controller design methodology.
From the point of view of an actuator for PTC, front wheel steering (FWS), rear wheel steering (RWS) and 4-wheel steering (4WS) have been selected as an actuator for PTC to date. For the last decade, by virtue of development of in-wheel motor or e-corner module systems, 4-wheel independent steering/braking/driving (4WIS/4WIB/4WID) has been available for vehicle stability control (VSC) and PTC. Several actuator combinations can be composed of those actuators, as shown in Table 1. If ∆Mz or ∆Fx is selected as a control input, a control allocation is needed to distribute it the control tire forces generated by RWS, 4WID and 4WIB.
Naturally, on a low-friction road surface, the magnitudes of longitudinal and lateral tire forces, Fx and Fy, become small because μ is low [33,34]. Among them, Fy’s generated at front wheels have significant effects on path tracking performance in cornering. Figure 1 shows Fy with respect to the tire slip angle, α, and μ. In Figure 1, the values of the tire slip angles, αm and α, are 5° and 20°, which give the maximum and saturated lateral tire forces, Fy,max and Fy,∞, respectively. As shown in Figure 1, Fy is saturated and decreased to Fy,∞ as α increases to α over αm where Fy,max is generated. In other words, Fy becomes Fy,∞, which is smaller than Fy,max due to larger slip angle α over αm. For example in Figure 1, 16% of Fy,max is reduced at α when μ is 0.6. Moreover, Fy is reduced as μ decreases, as shown in Figure 1. On a low-friction road surface, Fy is more easily saturated than on high-friction one. For the above two reasons, a path tracking controller, designed on a high-friction road surface, shows a poor performance on a low-friction one. To overcome this problem, it is necessary to design a path tracking controller with the consideration of low-friction road surface and lateral tire force saturation. In this paper, a constraint on the tire slip angle (CTSA) is derived from the definition of α, which results in the bound constraint on steering angles.
Research on vehicle motion control under low-friction road conditions has been performed in the field of vehicle stability control (VSC). In this paper, VSC means a control which tries to make a vehicle track a reference yaw rate generated by a steering angle, and lateral stability control (LSC) means a control which tries to keep the side-slip angle, β, as small as possible [34,35]. To date, a lot of papers have been published in the field of VSC and LSC [36,37,38]. Among the literature, some papers have been published to limit the tire slip angle such that Fy is not saturated or Fy is maintained at its maximum on low-friction road surface [39,40,41,42,43,44,45].
These papers can be classified into two categories. The first approach is to limit the slip angle of the front wheel when α is over αm at control allocation stage [39,40]. If Fy reaches its maximum, then the optimization variable Fy of front wheel is set to the constant Fy,max, and an extra optimization problem is formulated with remaining variables and modified cost function. Generally, quadratic programming (QP) has been adopted for control allocation. In the research, QP is applied twice so as to limit α. Another method is to use a daisy chaining method (DCM) for control allocation, which is a cascade optimization under element-wise constraint [41]. The second approach is to use an extra controller for the purpose of making α be αm [42,43,44,45]. In the research, various methods have been applied to keep Fy at its maximum so as to generate the maximum lateral acceleration. For example, some compensation coefficients were adopted to tune the magnitude of Fy,max. However, in these research projects, α was not directly manipulated by a method. Different from the previous two research projects, CTSAs are set as the bound on steering angles in this paper.
The aim of this paper is to present a method to impose a CTSA for the purpose of making a path tracking controller better under low friction conditions. From the definition of tire slip angle, CTSA is derived. To consider CTSA in controller design procedure, LQR and MPC are adopted as a controller design methodology. δf and δr and ∆Mz are selected as a control input, which make three input configurations. When imposing CTSA in LQR, the control inputs, δf and δr, are clipped by the constraint. In MPC, the constraint on steering angle is simply added into the optimization procedure. To distribute ∆Mz into tire forces generated by various actuator combinations, a control allocation is applied. Through the control allocation, the effects of various actuator combinations composed from FWS, 4WS, 4WID and 4WIB, on path tracking performance is analyzed. A simulation is done on a vehicle simulation software, CarSim 8.02, to verify path tracking performance with CTSAs. From simulation results, it is concluded that a path tracking controller with CTSA shows quite a better performance in terms of path tracking than one without CTSA on low-friction road surface.
The contributions of this paper can be summarized as follows:
  • This paper presents a method to impose a CTSA on steering angles in order to generate the maximum lateral force on a low-friction road surface. The constraint is simply derived from the definition of tire slip angle and easily applied into steering angles, i.e., control inputs of LQR and MPC. From simulation results, it is verified that the CTSA can greatly improve the path tracking performance on low-friction surface.
  • The proposed CTSA makes the most of the lateral tire force. This means that there are little tire forces left to improve the control performance with extra actuators. Under that condition, the effects of various combinations composed from the actuators such as RWS, 4WID, and 4WIB on control performance are discussed from simulation results.
  • From the discussion on simulation results, it is recommended that LQR is better than MPC in terms of computation time and simplicity of controller structure, and that the actuator combination, FWS + RWS, shows the best performance under CTSA.
This paper consists of five sections. In Section 2, design procedures for LQR and MPC with various actuator combinations are presented. Moreover, the derivation procedure for CTSAs is also presented. In Section 3, performance measures on path tracking are presented for the purpose of tuning LQR and MPC. In Section 4, simulation is conducted and simulation results are discussed in view of the performance measures. The conclusion of this paper is drawn in Section 5.

2. Design of Path Tracking Controller with Constraints on Tire Slip Angles

2.1. Vehicle Model

In the context of PTC, a 2-DOF bicycle model has been widely employed to represent yaw and lateral behaviors of a vehicle, as discussed in the literature [25,30,31,32,33,34,35,39,40]. Figure 2 shows the coordinates, geometry, and variables of the bicycle model. This models the lateral and yaw motions of a vehicle under the assumption of a constant longitudinal velocity. The state variables for this model are the yaw rate, γ, and the side-slip angle, β. With these variables, the equations of motion (1) are obtained [32]. The tire slip angles, denoted as αf and αr, are defined as Equation (2) for the front and rear wheels. The linear lateral tire forces, Fyf and Fyr, are calculated as Equation (3). By combining Equations (1)–(3), the state-space equation for the bicycle model is derived as Equation (4).
m v x β ˙ t + γ t = F y f t + F y r t I z γ ˙ t = l f F y f t l r F y r t + Δ M z t
α f t = δ f t β t l f γ t v x , α r t = δ r t β t + l r γ t v x
F y f t = C f α f t , F y r t = C r α r t
β ˙ t = C f + C r m v x β t + C f l f + C r l r m v x 2 1 γ t + C f m v x δ f t + C r m v x δ r t γ ˙ t = C f l f + C r l r I z β t + l f 2 C f + l r 2 C r I z v x γ t + l f C f I z δ f t l r C r I z δ r t + 1 I z Δ M z t
Figure 3 shows the coordinates and variables derived from the target path for path tracking. The literature on PTC introduced the concepts of lateral offset and heading errors, denoted as ey and eφ, at a specific point C as depicted in Figure 3 [11,12,13,14,15,16,17,18,19,20,21,22,23,24,25,26,27,28,29,30,31,32]. To enhance path tracking performance, this paper introduces a preview function, as used in the previous works [24,26,30,31,32]. The preview distance, Lp, is calculated using Equation (5), where kv is the velocity gain representing the preview interval. In Figure 3, the point C is located at the vehicle’s center of gravity. The preview point, P, and the point Q on the target path are determined with Lp. At the point Q, the time derivatives of ey and eφ are calculated as Equation (6), under the assumption that eφ is limited to an angle less than 10 degrees and, consequently, can be approximated as sineφeφ.
L p = k v v x
e ˙ y t = v x sin e φ t v x β t L p γ t v x e φ t v x β t L p γ t e ˙ φ t = φ ˙ d t φ ˙ t = v x χ t γ t
The vectors x, w, and u of the states, the disturbances, and control inputs are defined as Equation (7), respectively [11,12,14,15,16,17,19,20,21,22,23,28,29,31,32]. With those vectors and Equations (4), (6) and (7), the state–space equation for PTC is derived as Equation (8), where the matrices A, B1 and B2 are defined as Equation (9) [32].
x t = e y t e φ t β t γ t T w t = χ t u t = δ f t δ r t Δ M z t T
x ˙ t = A x t + B 1 w t + B 2 u t
A = 0 v x v x L p 0 0 0 1 0 0 C f + C r m v x C f l f + C r l r m v x 2 1 0 0 C f l f + C r l r I z l f 2 C f + l r 2 C r I z v x , B 1 = 0 v x 0 0 , B 2 = 0 0 0 0 0 0 C f m v x C r m v x 0 l f C f I z l r C r I z 1 I z
Three types of control input, named u1, u2 and u3, corresponding to input configurations IC#1, IC#2 and IC#3, are composed of three elements, δf, δr and ∆Mz, as given in Equation (10) [31,32]. In Equation (10), B2{k} representing the k-th column of the matrix B2. In Equation (10), δf is available in all input configurations, while δr is exclusively selected in IC#2 and IC#3. In the case of IC#3, RWS, 4WIB and 4WID are used to generate ∆Mz. With those actuators, δr and the braking and traction torques (TBi and TDi) for each wheel are obtained from ∆Mz. This paper assumes that TBi and TDi are generated through the use of 4WIB and 4WID, respectively.
u 1 t = δ f t B 21 = B 2 1 for I C # 1 u 2 t = δ f t δ r t T B 22 = B 2 1 B 2 2 for I C # 2 u 3 t = δ f t Δ M z t T B 23 = B 2 1 B 2 3 for I C # 3

2.2. Design of LQR

When designing LQR, a linear quadratic cost function (LQCF) is to be defined. LQCFs, J1, J2 and J3, for PTC with three input configurations, IC#1, IC#2, and IC#3, are defined as Equation (11), respectively. These LQCFs are transformed into the vector-matrix form as Equation (12). The weighting matrices Q and Ri in Equation (12) are defined in Equation (13). The weight ρi in LQCFs is determined with Bryson’s rule, as given in Equation (14), where ξi represents the maximum allowable value for the associated terms [31,32,46]. For better path tracking performance, ξi and kv are tuned. The control inputs u1, u2 and u3 for IC#1, IC#2 and IC#3 are obtained as Equation (15). In Equation (15), Pi stands for the solution to the Riccati equation for IC#i.
J 0 = 0 ρ 1 e y 2 t + ρ 2 e φ 2 t + ρ 3 β 2 t + ρ 4 γ 2 t d t J 1 = J 0 + 0 ρ 5 δ f 2 t d t for I C # 1 J 2 = J 0 + 0 ρ 5 δ f 2 t + ρ 6 δ r 2 t d t for I C # 2 J 3 = J 0 + 0 ρ 5 δ f 2 t + ρ 7 Δ M z 2 t d t for I C # 3
J i = 0 x t u i t T Q 0 0 R i x t u i t d t , i = 1 , 2 , 3
Q = d i a g ρ 1 , ρ 2 , ρ 3 , ρ 4 , R 1 = ρ 5 for I C # 1 R 2 = d i a g ρ 5 , ρ 6 for I C # 2 R 3 = d i a g ρ 5 , ρ 7 for I C # 3
ρ i = 1 ξ i 2
u i t = K i x t = R i 1 B 2 i T P i x t , i = 1 , 2 , 3

2.3. Constraint on Tire Slip Angles

Generally, there are physical limits on the front and rear steering angles, δf and δr. For example, δf and δr are limited to 30° and 5°, respectively. Under the limitations, δf is easily saturated on a low-friction road surface [33,34]. Moreover, as shown in Figure 1, Fyf is also saturated and decreased as αf is over αm where Fy,max is generated. In case of that condition, a path tracking controller cannot achieve the maximum performance. For the reason, α should be limited to αm, as given in Equation (16). By combining Equation (2) and Equation (16), Equation (17) is obtained. From Equation (17), the constraints on αf and αr are obtained as Equation (18). Let this denote the constraint on the tire slip angle, Equation (18), as CTSA.
α f t α m , α r t α m
α f t = δ f t β t l f γ t v x α m , α r t = δ r t β t + l r γ t v x α m
α m + β t + l f γ t v x δ f t α m + β t + l f γ t v x α m + β t l r γ t v x δ r t α m + β t l r γ t v x
In order to calculate the lower and upper bounds of Equation (18), it is essential to measure or estimate β. However, β is hard to measure with an on-board sensor. For this reason, β should be estimated with an observer. For the purpose, an extended Kalman Filter is employed in this paper [47].
From the point of view of LQR, Equation (18) is the constraint on the control inputs, δf and δr. For this reason, it is necessary to design LQR with the consideration of the constraints. The simplest way to impose the constraints, Equation (18), is to clip the control inputs obtained from LQR with the constraints. This does not need any additional procedures. The second method is to design with LQR under the constraints, Equation (18). The third method is to use MPC because it can explicitly handle any constraints in optimization procedure. The second method is not considered in this paper. Let this denote LQR without and with the CTSA as LQRN and LQRC, respectively.

2.4. Design of MPC with Constraints on Tire Slip Angles

As shown in Table 1, MPC has been widely adopted as a path tracking controller for the last decade because it can handle nonlinear or time-varying systems and nonlinear constraints besides linear time-invariant ones [11,12,14,16,17,19,20,21,22,27,28,29,30]. Generally, MPC is designed in the discrete-time domain. For the reason, the first step of MPC is to discretize Equation (8) with the sampling time, Ts. As a result, the discrete–time version of Equation (8) is obtained as Equation (19). In Equation (19), the subscript k represents a particular time instant. If k is 0, it means the current time instant. The matrices F, G and H in Equation (19) are obtained as an approximated form of Equation (20). In this paper, it is assumed that the disturbance wk in Equation (19) is set to 0. The discrete–time version of LQCF, Equation (11), is obtained as Equation (21). In Equation (21), N is the prediction horizon. Let this denote the current state vector as x0. The state vector at the time instant k is obtained as Equation (22) from Equation (19). In Equation (22), x0 is the current state. By expanding Equation (22) from the first to N-th time instant, new vectors and matrices are defined as Equation (23). With these definitions, the state–space equation and LQCF from the first to N-th time instant are calculated as Equation (24) and Equation (25), respectively. By replacing X of Equation (25) with Equation (24), the LQCF, Equation (25), is obtained by Equation (26). In Equation (27), the bound on steering angle, δmax, is imposed on the control input U. Moreover, instead of Equation (27), CTSAs can be imposed on the control input, as set in Equation (28). In Equation (28), Uf and Ur are the elements of U corresponding to δf and δr, respectively. β0 and γ0 are the values of β and γ, obtained from the current state x0. With this manner, the quadratic programming with the optimization variable U, the cost function JMPC, and the bound constraints of Equation (27) and Equation (28) is formulated [7]. In this paper, the solver, quadprog(), provided in MATLAB R2019a optimization toolbox, was applied to find the optimum U at every time instant k. After finding the optimum, MPC usually selects the first element of the optimum U as the control input. In this paper, Ts and N are set to, respectively, 0.01 s and 30.
x k + 1 = F x k + G w k + H u k
F = I + A T s , G = B 1 T s , H = B 2 T s
J M P C = k = 0 N 1 x k T Q x k + u k T R u k
x k + 1 = F k x 0 + i = 0 k 1 F i 1 H u k i 1
X = x 1 x 2 x 3 x N , F ¯ = F F 2 F 3 F N , H ¯ = H 0 0 0 0 F H H 0 0 0 F 2 H F H H 0 0 0 F N 1 H F N 2 H F N 3 H H , U = u 0 u 1 u 2 u N 1 Q ¯ = d i a g Q , , Q , R ¯ = d i a g R , , R
X = F ¯ x 0 + H ¯ U
J M P C = x 0 T Q x 0 + X T Q ¯ X + U T R ¯ U
J M P C = U T H ¯ T Q ¯ H ¯ + R ¯ U + 2 x 0 T F ¯ T Q ¯ H ¯ U
δ max U δ max
α m + β 0 + l f γ 0 v x U f α m + β 0 + l f γ 0 v x α m + β 0 l r γ 0 v x U r α m + β 0 l r γ 0 v x
Different from LQR, MPC can directly handle the constraints on the control inputs, i.e., δf and δr, as given in Equation (27) and Equation (28). However, MPC is to solve the optimization problem every time step, which requires a large amount of computations, compared to LQR with clipping. For the reason, there are hard limitations in implementing MPC on real vehicles. Let this denote MPC without and with the CTSA as MPCN and MPCC, respectively.

2.5. Control Allocation for LQR and MPC with IC#3

As shown in Equation (10), ΔMz is calculated from IC#3. After the controllers calculate ΔMz, it should be converted into the longitudinal forces applied to wheels. This force is generated by various combinations of the actuators, RWS, 4WIB, and 4WID. Typically, this conversion is accomplished with a process known as control allocation. In this paper, a weighted least square (WLS)-based method is employed for control allocation, which has been adopted in previous works [31,32].
Figure 4 illustrates the tire forces at each wheel and ΔMz when ΔMz is positive [25,30,31,32,35,39,40]. In Figure 4, four wheels are numbered as 1, 2, 3 and 4 in the order of the front left, front right, rear left and rear right wheels, respectively. In Figure 4, ∆Fyr represents Fy generated by the rear wheels through RWS. ∆Fx1, ∆Fx2, ∆Fx3 and ∆Fx4 denote Fx‘s generated by 4WID and 4WIB. If ∆Fxi is positive, it corresponds to TDi generated by 4WID; otherwise, it corresponds to TBi generated by 4WIB. Those five tire forces need to be determined so as to generate ΔMz. For the purpose, WLS-based method is selected as a control allocation method.
The equilibrium condition among ∆Mz and tire forces is derived as Equation (29) from the geometric information depicted in Figure 4 [36,39,43,46]. The components of vector z are given in Equation (30).
z 1 t z 2 t z 3 t z 4 t z 5 t z t Δ F y r t Δ F x 1 t Δ F x 2 t Δ F x 3 t Δ F x 4 t q t = z t q t = Δ M z t
z 1 t = 2 l r cos δ r t , z 2 t = l f sin δ f t + t f cos δ f t , z 3 t = l f sin δ f t t f cos δ f t , z 4 t = l r sin δ r t + t r cos δ r t , z 5 t = l r sin δ r t t r cos δ r t
The quadratic cost function for WLS is given as Equation (31). Within Equation (31), ζi is the product of μ and Fzi at the wheel i, i.e., μFzi, which represents the friction circle radius. The vector of virtual weights, denoted as κ, is introduced in Equation (31) for the purpose of selecting the combination of actuators [25,30,31,32,35,39]. The quadratic cost function, JEC in Equation (32), is obtained from the constraint Equation (29). In previous research, the equilibrium condition Equation (29) should be satisfied for the purpose of generating ∆Mz. The cost functions, Equations (31) and (32), are summed into a single one (33) with the Lagrange multiplier, η. In Equation (33), it is important to set η to a value of 1 or higher; otherwise, the equilibrium condition Equation (29) will not be met. The optimal solution for Equation (33) is algebraically calculated as Equation (34) by taking the derivative of Equation (33) with respect to q [31,32].
J Q = κ 2 Δ F x 1 2 t ζ 1 2 + κ 3 Δ F x 2 2 t ζ 2 2 + κ 1 Δ F y r 2 t + κ 4 Δ F x 3 2 t ζ 3 2 + κ 1 Δ F y r 2 t + κ 5 Δ F x 4 2 t ζ 4 2 = q T t Θ q t Θ = diag 1 ζ 3 2 + 1 ζ 4 2 , 1 ζ 1 2 , 1 ζ 2 2 , 1 ζ 3 2 , 1 ζ 4 2 κ κ = diag κ 1 κ 2 κ 3 κ 4 κ 5
J E C = z t q t Δ M z t T z t q t Δ M z t
J C A = J Q + η J E C = q T t Θ q t + η z t q t Δ M z t T z t q t Δ M z t
q o p t t = η Θ + η z T t z t 1 z T t Δ M z t
When WLS is applied to control allocation for IC#3, an arbitrary combination of actuators comprising RWS, 4WID, and 4WIB can be configured. To account for several actuator combinations, the virtual weights, denoted as κi, can be set to match a particular combination of those actuators [31,32]. The detailed usage of the virtual weights and the procedure to derive δr, TDi and TBi of each wheel can be also found in the literature [31,32].
In the field of VSC or PTC, the friction circle constraint, as defined in Equation (35), has been included into the optimization procedure for control allocation [48,49,50]. This constraint can be transformed to define the tire force margin (TFM), as defined in Equation (36). TFM indicates how much extra tire force can be generated under VSC or PTC. If TFM is small, Fx and Fy are near its maximum, which indicates the fact that an extra actuator selected to generate these forces becomes useless. If TFM is zero, this indicates the fact that both Fx and Fy are saturated and that there are no margins to generate extra tire force for VSC or PTC. As shown in Equation (36), TFM becomes much smaller under low-friction road conditions. If CTSA is imposed, then it is expected that TFM easily becomes zero.
μ F z t 2 F x 2 t + F y 2 t
TFM t = μ F z t F x 2 t + F y 2 t

3. Performance Measures for Path Tracking Control

In the domain of PTC, ey and eφ have been widely used as indicators for assessing path tracking performance. However, different from conventional PTC literature, this paper adopts a double lane change maneuver as the target path for collision avoidance [13,16,17,20,21,31,32]. This maneuver was employed to evaluate the reachability and agility of a path tracking controller [31,32]. Figure 5 shows both the target path and the trajectory of a vehicle. In Figure 5, the points A, B and C are on the target path, which represent the peak, zero-crossing and convergence points, respectively. The points D, E, F and G are on the vehicle trajectory, which represent the peak, zero-crossing, over-shoot and convergence points, respectively. Five metrics, MX, MY, MOS, MDX and MSX, representing path tracking performance, are defined as Equation (37), based on the points along the target path and vehicle trajectory in Figure 5. In Equation (37), the subscripts X and Y correspond to the x- and y-positions of the corresponding point, respectively.
M X = D X A X = D X 73.20 M Y = D Y A Y = D Y 3.53 M D X = E X B X = E X 91.50 M S X = G X C X = G X 190.00 M O S = F Y 1.65 1.65 + 3.53 × 100
MX and MY represent the peak’s center and lateral offsets, respectively. In this study, agility and reachability are represented by MX and MY, respectively. If MY exceeds −0.02 m, the controller’s path tracking performance can be regarded as satisfactory. MOS represents the percentage of overshoot, which is identical to that of automatic control theory. This indicates lateral damping, reflecting agility. In this paper, a satisfactory performance is achieved if MOS remains below 16%, which is equivalent to an overshoot of 0.85 m. MDX and MSX represent the response delay and the settling delay, respectively. MDX corresponds to a delay, standing for the agility of longitudinal motion. MDX highly depends on μ. MSX can be interpreted as the settling time, indicating the convergence speed of the vehicle’s lateral motion toward a target y-position. MSX is closely related to MOS. In essence, smaller absolute values of these metrics correspond to improved path tracking performance.
As mentioned previously, VSC and LSC aim for yaw rate tracking and lateral stability, respectively. Typically, yaw rate tracking performance is assessed by the difference between real and reference yaw rates, while lateral stability performance is evaluated by the magnitude of β. It has been considered satisfactory if β is less than 3 degrees [51]. Let the maximum absolute value of β be referred to as MASSA. The side-slip angular rate, when combined with β, has been also employed as a measure of lateral stability [30]. Let this denote the maximum absolute value of side-slip angular rate as MASSAR.
In this paper, the tuning parameters are the maximum allowable value ξi given in Equation (14) and the velocity gain kv given in Equation (5). There are three tuning objectives when tuning these parameters in this paper. The first objective is collision avoidance, which is achieved by tuning parameters such that MY is larger than −0.02 m. Generally, the best MSX is attained when MOS is around 0.9, provided that MY exceeds −0.02 m. The second objective is path tracking, which is achieved by tuning parameters such that MX, MDX and MSX are maintained as small as possible. Among them, the minimum of MDX clearly depends on μ. The third objective is lateral stability, which is achieved by tuning parameters such that MASSA is less than 2°.

4. Simulation and Discussion

To verify the effects of the CTSA on path tracking performance under low-friction road condition, a simulation was conducted. The path tracking controllers, LQRN, LQRC, MPCN and MPCC, were implemented with MATLAB/Simulink R2019a, connected to CarSim [52]. Test scenario was given in Figure 5. In the simulation, F-segment sedan model in CarSim was selected as a test vehicle [52]. From simulation with the model under step steering or steady-state cornering, the parameters and its values of the 2-DOF bicycle model were obtained as given in the reference [25,30,31,32]. In this paper, a first-order system was adopted to model actuators such as FWS and 4WS, and 4WIB and 4WID, where the bandwidths of these actuators are set to 5 Hz and 2 Hz, respectively. The initial vx and μ were set to 60 km/h and 0.4, respectively. For the purpose of maintaining vx as constant as possible, a built-in speed controller provided in CarSim was applied. The simulation horizon was set to 15 s.
Generally, the larger δr, the larger β. For the reason, it is necessary to limit ∆Mz to a certain value so as to make δr small for IC#3. Fy,max from F-segment sedan model in CarSim is 7500 N where μ is 1 [52]. This will be 3000 N if μ is set to 0.4. As shown in Equation (29), Fy at front wheels, i.e., ∆Fyf, was not used to generate ∆Mz because δf is included in the control input. From the geometric information of the test vehicle, the maximum available yaw moment at rear wheels is calculated as 3000 × 2 × lr = 11,400 Nm. After ∆Mz was obtained from LQR with IC#3, it is limited by clipping it into 11,400 Nm in this paper. δmax of front and rear wheels was set to 30°. αm was set to 5°, which is referred from CarSim tire model as given in Figure 1.
The first simulation is done for LQRN and LQRC. Table 2 and Table 3 show the simulation results of LQRN and LQRC. As mentioned earlier, the control parameters were tuned such that MY is larger than −0.02 m and MOS is near 0.9. These two measures are nearly the same as shown in Table 2 and Table 3. For the reason, these are not used for comparison. The percentages inside the parentheses in Table 3 represent the percentage reductions of MX, MDX and MSX relative to the corresponding values given in Table 2 for each actuator combination. As shown in those tables, MX, MDX and MSX of LQRC are improved, compared to those of LQRN. This indicates that the CTSA can effectively limit α to αm, which generates Fy,max.
In case of LQRN, the actuator combination, FWS + 4WIB of IC#3, shows the best performance in terms of MDX and MSX. On the contrary, in case of LQRC, the actuator combination, FWS + RWS of IC#3, shows the best performance in terms of MX, MDX and MSX. Moreover, this combination is superior to IC#2 of LQRC, which used the identical actuators, FWS and RWS. This means that any actuator combinations using 4WID and 4WIB are inferior to FWS + RWS of IC#3. The input configurations, IC#1 and IC#2, of LQR do not require a control allocation. In IC#1 and IC#2, δf and δr, are directly computed and applied into steering actuators. However, IC#3 of LQR requires a control allocation, which is complex and time-consuming. Due to its simplicity, IC#1 and IC#2 of LQR are recommended if one do not think the difference between IC#1/IC#2 and IC#3 is significant. From these results, it can be concluded that CTSA is strongly recommended for better path tracking. Moreover, it is also concluded that FWS + RWS of IC#3 is recommended as an actuator combination instead of IC#2 in terms of path tracking performance, and that IC#1 or IC#2 is recommended as an input configuration instead of IC#3 in terms of simplicity and ease of implementation.
Figure 6 shows the plots drawn from the simulation results of IC#1 of LQRN and LQRC. In Figure 6d,e, the legends, FL, FR, RL and RR, stand for front left, front right, rear left and rear right wheels, respectively. As shown in Figure 6a, δf was limited to a certain value near αm. Compared to the limits on the steering angles, ±30°, αm, i.e., 5°, is quite small. As a result, Fy‘s at front wheels were not saturated and were generated to its own maximum, as shown in Figure 6d. For the reason, TFMs of front wheels become 0, as shown in Figure 6e, and the path tracking performance of LQRC was improved, compared to LQRN, as shown in Figure 6b. Because the maximum Fy‘s are generated by FWS with CTSA of LQRC, TFMs of LQRC at front wheels became zero, which indicates the other actuators are not needed for FWS to generate extra Fy’s. This can be checked in Figure 6e. On the other hand, as shown in Figure 6d, TFMs of LQRN at front wheels did not become zero, which means the other actuators can generate extra Fy‘s at front wheels.
The second simulation is done for MPCN and MPCC. Table 4 and Table 5 show the simulation results of MPCN and MPCC. As shown in those tables, MX, MDX and MSX of MPCC are improved, compared to those of MPCN. This coincides with the results of LQRC. This means that CTSA can effectively limit α to αm, which cause the maximum Fy.
Figure 7 show the plots drawn from the simulation results of IC#1 of MPCN and MPCC. In Figure 7d,e, the legends, FL, FR, RL and RR, stand for front left, front right, rear left and rear right wheels, respectively. As shown in Figure 7c, Fy’s of MPCN were saturated due to the large α. On the contrary, Fy’s of MPCC were not saturated due to the CTSA, as shown in Figure 7c. Because the maximum Fy’s are generated by FWS with the CTSA of MPCC, TFM of MPCC at front wheels became zero, which means the other actuators are not needed for FWS to generate extra Fy’s. This can be checked in Figure 7d. On the other hand, as shown in Figure 6d, TFM of MPCN at front wheels did not become zero, which means the other actuators can generate extra Fy’s at front wheels.

5. Conclusions

This paper presented a path tracking controller with CTSA on low-friction road surface. The contents of this paper are summarized as follows:
  • This paper presented a method to impose a CTSA such that the maximum lateral force is generated on low-friction road surface. The constraint is derived simply from the definition of tire slip angle and easily applied into LQR and MPC. From simulation results, the CTSA can greatly improve the path tracking performance on low-friction surface.
  • The proposed CTSA makes the most of the lateral tire force. This means that there are little tire forces left to improve the control performance. Under that condition, it was shown that various combinations composed from RWS, 4WID, and 4WIB have little effects on control performance.
  • From the simulation results, it is recommended that LQR is better than MPC in terms of computation time and simplicity of controller structure. Moreover, RWS calculated from the control yaw moment in IC#3 provided the best path tracking result. However, IC#3 requires a control allocation, which is complex and time-consuming, compared to IC#1 and IC#2. For the reason, IC#1 and IC#2 are preferred to IC#3 due to its simplicity and ease of implementation if one do not think the difference between IC#1/IC#2 and IC#3 is significant.
The CTSA adopted in this paper are easy to derive and simple to apply to LQR or MPC. Moreover, a procedure to design LQR with bound constraints on the control inputs are not needed. For the reason, it is recommended that a path tracking controller should include the CTSA.
The limitations of the proposed method are to measure the yaw rate and velocities and to estimate the side-slip angle with available sensors. To overcome these issues, sensor specification and estimation schemes are to be investigated. In further research, experiments on a scaled vehicle will be done with the proposed method. Moreover, extra experiments will be performed on real vehicles on low friction conditions.

Author Contributions

J.L. conceptualized the main idea, and designed this study. S.Y. supported fund for this study. S.Y. participated in formulating the idea, as well as validating the proposed method and results. S.Y. implemented the methodology and obtained results. J.L. and S.Y. analyzed the results and drafted the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Research Program funded by Seoul National University of Science and Technology (2023-0885).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the references [31,32].

Conflicts of Interest

The authors declare no conflicts of interest.

Nomenclature

4WS4-wheel steering
4WIB4-wheel independent braking
4WID4-wheel independent drive
CTSAconstraint on tire slip angle
FWSfront wheel steering
LQCFLQ cost function
LQRlinear quadratic regulator
LQRCLQR with CTSAs
LQRNLQR without CTSAs
LSClateral stability control
MASSAmaximum absolute value of the side-slip angle over simulation horizon
MASSARmaximum absolute value of side-slip angular rate over simulation horizon
MPCmodel predictive control
MPCCMPC with CTSAs
MPCNMPC without CTSAs
RWSrear wheel steering
SMCsliding mode control
TFMtire force margin derived from the friction circle (N)
VSCvehicle stability control
WLSweighted least square
Cf, Crcornering stiffness of front and rear tires (N/rad)
ey, eφlateral offset error (m) and heading error (rad)
Fx, Fy, Fzlongitudinal, lateral and vertical tire forces of a tire (N)
Fxi, Fyi, Fzilongitudinal, lateral and vertical tire forces of i-th wheel (N)
Fyf, Fyrfront and rear lateral tire forces in the 2-DOF bicycle model (N)
Fy,maxmaximum lateral tire force of a tire at α = αm (N)
zmatrix used for the constraint on RWS and 4WS in WLS based method
Izyaw moment of inertial (kg·m2)
JiLQ cost function for the input configuration IC#i
KLQR,igain matrix of LQR for input configuration IC#i
kvvelocity gain for preview distance (s−1)
Lppreview distance (m)
lf, lrdistance from a center of gravity to front and rear axles (m)
MX, MYdifferences between x- and y-positions at the peak points of the target path (m)
MDX, MSXresponse and settling delays of vehicle trajectory with respect to target path (m)
MOSpercentage overshoot along y-direction in the lower lane of the target path
Nprediction horizon of MPC
mvehicle total mass (kg)
qvector of tire forces as a solution of WLS based method
TBi, TDibraking and traction torques generated at i-th wheel from ∆Mz (N·m)
Tssampling interval of discrete-time system (s)
tf, trhalf of track widths of front and rear axles (m)
vx, vylongitudinal and lateral velocities of CoG of a vehicle (m/s)
αtire slip angle (rad)
αmtire slip angle where the maximum lateral tire force comes out (rad)
αf, αrtire slip angles of front and rear wheels (rad)
βside-slip angle of CoG of a vehicle (rad) = tan−1(vy/vx) ≈ (vy/vx)
δf, δrfront and rear steering angles as a control input in LQR and MPC (rad)
Fyrcontrol lateral tire force of rear wheels generated by an actuator (N)
Fxicontrol longitudinal tire force at i-th wheel, generated by an actuator (N)
Mzcontrol yaw moment as a control input in LQR and MPC (N·m)
ηtuning parameter on relaxation term of equality constraint
γ, γdreal and reference yaw rates (rad/s)
χcurvature at a particular point on a target path (m−1)
κivirtual weight on the longitudinal and lateral tire forces
κthe vector of virtual weights
ξithe maximum allowable value of i-th term in LQ cost function
φheading angle of a vehicle (rad)
φddesired heading angle obtained at preview point (rad)
μtire-road friction coefficient
ρiweight on i-th term in LQCF
Θweighting matrix of WLS based method
ζthe product of μ and Fz = μFz

References

  1. Montanaro, U.; Dixit, S.; Fallaha, S.; Dianatib, M.; Stevensc, A.; Oxtobyd, D.; Mouzakitisd, A. Towards connected autonomous driving: Review of use-cases. Veh. Syst. Dyn. 2019, 57, 779–814. [Google Scholar] [CrossRef]
  2. Yurtsever, E.; Lambert, J.; Carballo, A.; Takeda, K. A survey of autonomous driving: Common practices and emerging technologies. IEEE Access 2020, 8, 58443–58469. [Google Scholar] [CrossRef]
  3. Omeiza, D.; Webb, H.; Jirotka, M.; Kunze, M. Explanations in autonomous driving: A survey. IEEE Trans. Intell. Transp. Syst. 2022, 23, 10142–10162. [Google Scholar] [CrossRef]
  4. Paden, B.; Cap, M.; Yong, S.Z.; Yershov, D.; Frazzoli, E. A survey of motion planning and control techniques for self-driving urban vehicles. IEEE Trans. Intell. Veh. 2016, 1, 33–55. [Google Scholar] [CrossRef]
  5. Sorniotti, A.; Barber, P.; De Pinto, S. Path tracking for automated driving: A tutorial on control system formulations and ongoing research. In Automated Driving; Watzenig, D., Horn, M., Eds.; Springer: Cham, Switzerland, 2017. [Google Scholar]
  6. Amer, N.H.; Hudha, H.Z.K.; Kadir, Z.A. Modelling and control strategies in path tracking control for autonomous ground vehicles: A review of state of the art and challenges. J. Intell. Robot. Syst. 2017, 86, 225–254. [Google Scholar] [CrossRef]
  7. Bai, G.; Meng, Y.; Liu, L.; Luo, W.; Gu, Q.; Liu, L. Review and comparison of path tracking based on model predictive control. Electronics 2019, 8, 1077. [Google Scholar] [CrossRef]
  8. Yao, Q.; Tian, Y.; Wang, Q.; Wang, S. Control strategies on path tracking for autonomous vehicle: State of the art and future challenges. IEEE Access 2020, 8, 161211–161222. [Google Scholar] [CrossRef]
  9. Rokonuzzaman, M.; Mohajer, N.; Nahavandi, S.; Mohamed, S. Review and performance evaluation of path tracking controllers of autonomous vehicles. IET Intell. Transp. Syst. 2021, 15, 646–670. [Google Scholar] [CrossRef]
  10. Stano, P.; Montanaro, U.; Tavernini, D.; Tufo, M.; Fiengo, G.; Novella, L.; Sorniotti, A. Model predictive path tracking control for automated road vehicles: A review. Annu. Rev. Control 2022, 55, 194–236. [Google Scholar] [CrossRef]
  11. Ren, Y.; Zheng, L.; Khajepour, A. Integrated model predictive and torque vectoring control for path tracking of 4-wheeldriven autonomous vehicles. IET Intell. Transp. Syst. 2019, 13, 98–107. [Google Scholar] [CrossRef]
  12. Zhang, B.; Zong, C.; Chen, G.; Li, G. An adaptive-prediction-horizon model prediction control for path tracking in a four-wheel independent control electric vehicle. Proc. Inst. Mech. Eng. Part D J. Automob. Eng. 2019, 233, 3246–3262. [Google Scholar] [CrossRef]
  13. Chen, X.; Peng, Y.; Hang, P.; Tang, T. Path tracking control of four-wheel independent steering electric vehicles based on optimal control. In Proceedings of the 2020 39th Chinese Control Conference (CCC), Shenyang, China, 27–30 July 2020; pp. 5436–5442. [Google Scholar]
  14. Guo, L.; Ge, P.; Yue, M.; Li, J. Trajectory tracking algorithm in a hierarchical strategy for electric vehicle driven by four independent in-wheel motors. J. Chin. Inst. Eng. 2020, 43, 807–818. [Google Scholar] [CrossRef]
  15. He, Z.; Nie, L.; Yin, Z.; Huang, S. A two-layer controller for lateral path tracking control of autonomous vehicles. Sensors 2020, 20, 3689. [Google Scholar] [CrossRef] [PubMed]
  16. Wu, H.; Si, Z.; Li, Z. Trajectory tracking control for four-wheel independent drive intelligent vehicle based on model predictive control. IEEE Access 2020, 8, 73071–73081. [Google Scholar] [CrossRef]
  17. Wu, H.; Li, Z.; Si, Z. Trajectory tracking control for four-wheel independent drive intelligent vehicle based on model predictive control and sliding mode control. Adv. Mech. Eng. 2021, 13, 9. [Google Scholar] [CrossRef]
  18. Hang, P.; Chen, X. Path tracking control of 4-wheel steering autonomous ground vehicles based on linear parameter-varying system with experimental verification. Proc. Inst. Mech. Eng. Part I J. Syst. Control Eng. 2021, 235, 411–423. [Google Scholar] [CrossRef]
  19. Xie, J.; Xu, X.; Wang, F.; Tang, Z.; Chen, L. Coordinated control based path following of distributed drive autonomous electric vehicles with yaw-moment control. Cont. Eng. Prac. 2021, 106, 104659. [Google Scholar] [CrossRef]
  20. Xiang, C.; Peng, H.; Wang, W.; Li, L.; An, Q.; Cheng, S. Path tracking coordinated control strategy for autonomous four in-wheel-motor independent-drive vehicles with consideration of lateral stability. Proc. Inst. Mech. Eng. Part D J. Automob. Eng. 2021, 235, 1023–1036. [Google Scholar] [CrossRef]
  21. Yang, K.; Tang, X.; Qin, Y.; Huang, Y.; Wang, H.; Pu, H. Comparative study of trajectory tracking control for automated vehicles via model predictive control and robust H-infinity state feedback control. Chin. J. Mech. Eng. 2021, 34, 74. [Google Scholar] [CrossRef]
  22. Wang, G.; Liu, L.; Meng, Y.; Gu, Q.; Bai, G. Integrated path tracking control of steering and braking based on holistic MPC. IFAC PapersOnLine 2021, 54, 45–50. [Google Scholar] [CrossRef]
  23. Tong, Y.; Jing, H.; Kuang, B.; Wang, G.; Liu, F.; Yang, Z. Trajectory tracking control for four-wheel independently driven electric vehicle based on model predictive control and sliding model control. In Proceedings of the 2021 5th CAA International Conference on Vehicular Control and Intelligence (CVCI), Tianjin, China, 29–31 October 2021. [Google Scholar]
  24. Sun, X.; Wang, Y.; Hu, W.; Cai, Y.; Huang, C.; Chen, L. Path tracking control strategy for the intelligent vehicle considering tire nonlinear cornering characteristics in the PWA form. J. Frankl. Inst. 2022, 359, 2487–2513. [Google Scholar] [CrossRef]
  25. Jeong, Y.; Yim, S. Integrated path tracking and lateral stability control with four-wheel independent steering for autonomous electric vehicles on low-friction roads. Machines 2022, 10, 650. [Google Scholar] [CrossRef]
  26. Hu, H.; Bei, S.; Zhao, Q.; Han, X.; Zhou, D.; Zhou, X.; Li, B. Research on trajectory tracking of sliding mode control based on adaptive preview time. Actuators 2022, 11, 34. [Google Scholar] [CrossRef]
  27. Barari, A.; Afshari, S.S.; Liang, X. Coordinated control for path-following of an autonomous four in-wheel motor drive electric vehicle. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2022, 236, 6335–6346. [Google Scholar] [CrossRef] [PubMed]
  28. Wang, G.; Liu, L.; Meng, Y.; Gu, Q.; Bai, G. Integrated path tracking control of steering and differential braking based on tire force distribution. Int. J. Control Autom. Syst. 2022, 20, 536–550. [Google Scholar] [CrossRef]
  29. Wang, W.; Zhang, Y.; Yang, C.; Qie, T.; Ma, M. Adaptive model predictive control-based path following control for four-wheel independent drive automated vehicles. IEEE Trans. Intell. Transp. Syst. 2022, 23, 14399–14412. [Google Scholar] [CrossRef]
  30. Lee, J.; Yim, S. Comparative study of path tracking controllers on low-friction roads for autonomous vehicles. Machines 2023, 11, 403. [Google Scholar] [CrossRef]
  31. Park, M.; Yim, S. Comparative study on effects of input configurations of linear quadratic controller on path tracking performance under low-friction condition. Actuators 2023, 12, 153. [Google Scholar] [CrossRef]
  32. Park, M.; Yim, S. Comparative study on coordinated control of path tracking and vehicle stability for autonomous vehicles on low-friction roads. Actuators 2023, 12, 398. [Google Scholar] [CrossRef]
  33. Wong, H.Y. Theory of Ground Vehicles, 3rd ed.; John Wiley and Sons, Inc.: New York, NY, USA, 2001. [Google Scholar]
  34. Rajamani, R. Vehicle Dynamics and Control; Springer: New York, NY, USA, 2006. [Google Scholar]
  35. Nah, J.; Yim, S. Vehicle stability control with four-wheel independent braking, drive and steering on in-wheel motor-driven electric vehicles. Electronics 2020, 9, 1934. [Google Scholar] [CrossRef]
  36. Ahangarnejad, A.H.; Radmehr, A.; Ahmadian, M. A review of vehicle active safety control methods: From antilock brakes to semiautonomy. J. Vib. Control 2021, 27, 1683–1712. [Google Scholar] [CrossRef]
  37. Katsuyama, E.; Yamakado, M.; Abe, M. A state-of-the art review: Toward a novel vehicle dynamics control concept taking the driveline of electric vehicles into account as promising control actuators. Veh. Syst. Dyn. 2021, 59, 976–1025. [Google Scholar] [CrossRef]
  38. Peng, H.; Chen, X. Active safety control of X-by-wire electric vehicles: A survey. SAE Int. J. Veh. Dyn. Stab. NVH 2022, 6, 115–133. [Google Scholar] [CrossRef]
  39. Yim, S.; Kim, S.; Yun, H. Coordinated control with electronic stability control and active front steering using the optimum yaw moment distribution under a lateral force constraint on the active front steering. Proc. Inst. Mech. Eng. Part D J. Automob. Eng. 2016, 230, 581–592. [Google Scholar] [CrossRef]
  40. Nah, J.; Yim, S. Vehicle dynamic control with 4WS, ESC and TVD under constraint on front slip angles. Energies 2021, 14, 6306. [Google Scholar] [CrossRef]
  41. Park, G.; Han, K.; Nam, K.; Kim, H.; Choi, S.B. Torque vectoring algorithm of electronic-four-wheel drive vehicles for enhancement of cornering performance. IEEE Trans. Veh. Technol. 2020, 69, 3668–3679. [Google Scholar] [CrossRef]
  42. Heo, H.; Joa, E.; Yi, K.; Kim, K. Integrated chassis control for enhancement of high speed cornering performance. SAE Int. J. Commer. Veh. 2015, 8, 102–109. [Google Scholar] [CrossRef]
  43. Her, H.; Joa, E.; Yi, K.; Kim, K. Integrated chassis control for optimized tyre force coordination to enhance the limit handling performance. Proc. Inst. Mech. Eng. Part D J. Automob. Eng. 2016, 230, 1011–1026. [Google Scholar] [CrossRef]
  44. Her, H.; Koh, Y.; Joa, E.; Yi, K.; Kim, K. An integrated control of differential braking, front/rear traction, and active roll moment for limit handling performance. IEEE Trans. Veh. Technol. 2016, 65, 4288–4300. [Google Scholar] [CrossRef]
  45. Joa, E.; Yi, K.; Sohn, K.; Bae, H. Four-wheel independent brake control to limit tire slip under unknown road conditions. Cont. Eng. Prac. 2018, 76, 79–95. [Google Scholar] [CrossRef]
  46. Bryson, A.E.; Ho, Y.C. Applied Optimal Control; Hemisphere: New York, NY, USA, 1975. [Google Scholar]
  47. Kim, H.H.; Ryu, J. Sideslip angle estimation considering short-duration longitudinal velocity variation. Int. J. Automot. Technol. 2011, 12, 545–553. [Google Scholar] [CrossRef]
  48. de Castro, R.; Tanelli, M.; Araújo, R.E.; Savaresi, S.M. Design of safety-oriented control allocation strategies for overactuated electric vehicles. Veh. Syst. Dyn. 2014, 52, 1017–1046. [Google Scholar] [CrossRef]
  49. Rezaeian, A.; Zarringhalam, R.; Fallah, S.; Melek, W.; Khajepour, A.; Chen, S.-K.; Moshchuck, N.; Litkouhi, B. Novel tire force estimation strategy for real-time implementation on vehicle applications. IEEE Trans. Veh. Technol. 2015, 64, 2231–2241. [Google Scholar] [CrossRef]
  50. Wang, J.; Wang, R.; Jing, H.; Chen, N. Coordinated active steering and four-wheel independent driving/braking control with control allocation. Asian J. Control 2016, 18, 98–111. [Google Scholar] [CrossRef]
  51. National Highway Traffic Safety Administration. FMVSS No. 126, Electronic Stability Control Systems: NHTSA Final Regulatory Impact Analysis; National Highway Traffic Safety Administration: Washington, DC, USA, 2007. [Google Scholar]
  52. Mechanical Simulation Corporation. VS Browser: Reference Manual, the Graphical User Interfaces of BikeSim, CarSim, and TruckSim; Mechanical Simulation Corporation: Ann Arbor, MI, USA, 2009. [Google Scholar]
Figure 1. Fy with respect to α and μ.
Figure 1. Fy with respect to α and μ.
Applsci 14 01066 g001
Figure 2. 2-DOF dynamic bicycle model.
Figure 2. 2-DOF dynamic bicycle model.
Applsci 14 01066 g002
Figure 3. Coordinates and variables derived from target path.
Figure 3. Coordinates and variables derived from target path.
Applsci 14 01066 g003
Figure 4. Geometric information and coordinate system of tire forces and ∆Mz.
Figure 4. Geometric information and coordinate system of tire forces and ∆Mz.
Applsci 14 01066 g004
Figure 5. Measures used to tune PTC.
Figure 5. Measures used to tune PTC.
Applsci 14 01066 g005
Figure 6. Simulation results with IC#1 of LQRN and LQRC. (a) Steering angles; (b) Trajectories; (c) Side-slip angles; (d) slip angles vs. lateral tire forces; (e) Tire force margin.
Figure 6. Simulation results with IC#1 of LQRN and LQRC. (a) Steering angles; (b) Trajectories; (c) Side-slip angles; (d) slip angles vs. lateral tire forces; (e) Tire force margin.
Applsci 14 01066 g006aApplsci 14 01066 g006b
Figure 7. Simulation results with IC#1 of MPCN and MPCC. (a) Steering angles; (b) Trajectories; (c) Side-slip angles; (d) slip angles vs. lateral tire forces; (e) Tire force margin.
Figure 7. Simulation results with IC#1 of MPCN and MPCC. (a) Steering angles; (b) Trajectories; (c) Side-slip angles; (d) slip angles vs. lateral tire forces; (e) Tire force margin.
Applsci 14 01066 g007aApplsci 14 01066 g007b
Table 1. Summary of path tracking controllers on low friction roads.
Table 1. Summary of path tracking controllers on low friction roads.
ControllerInput ConfigurationActuatorsμReference
LQR, MPC, SMC[δf], [δf δr]FWS, 4WS0.4[30]
MPC[δf]FWS0.4, 0.5[12]
0.2, 0.5[14]
0.3, 0.8[15]
0.2, 0.85[21]
0.3[22]
MPC[δfFx]FWS, 4WID0.4, 0.5[11]
FWS, 4WID, 4WIB0.6, 0.7[16]
FWS, 4WID, 4WIB0.4, 0.8[17]
FWS, 4WIB0.3[28]
MPC[δf]FWS, 4WID0.4,0.7[19]
[δfFxMz]FWS, 4WID0.3[20]
[δf], [∆Mz]FWS, 4WID, 4WIB0.3[23]
[δfMz]FWS, 4WID0.5, 0.6[29]
LQR[δf δr]4WS, 4WIS0.4, 0.65[13]
[δf δr]4WS0.25, 0.5[18]
[δf]FWS0.34[24]
LQR[δf], [δf δr], [δf δrMz]FWS, 4WS, 4WIB, 4WID0.4[31,32]
SMC[∆Mz]4WIS, 4WID, 4WIB0.4[25]
[δf]FWS0.5[26]
[δfMz]FWS, 4WID, 4WIB0.3[27]
Table 2. Summary of the simulation results with LQRN for IC#1, IC#2, IC#3.
Table 2. Summary of the simulation results with LQRN for IC#1, IC#2, IC#3.
Actuator
Combinations
MX
(m)
MY
(m)
MOSMDX
(m)
MSX
(m)
MASSA
(deg)
MASSAR
(deg/s)
IC#1FWS1.25−0.0181.08.584.390.5813.19
IC#24WS1.58−0.0170.98.944.971.2613.99
IC#3
FWS
+RWS1.15−0.0190.98.514.310.7913.16
+RWS + 4WID1.24−0.0180.98.474.540.7713.13
+RWS + 4WIB1.62−0.0180.98.664.110.8012.97
+RWS + 4WID + 4WIB1.47−0.0180.99.445.520.8013.25
+4WID1.67−0.0180.99.045.100.6013.38
+4WIB1.69−0.0170.98.143.570.6812.71
+4WID + 4WIB1.66−0.0170.99.645.710.6313.47
Table 3. Summary of the simulation results with LQRC for IC#1, IC#2, IC#3.
Table 3. Summary of the simulation results with LQRC for IC#1, IC#2, IC#3.
Actuator
Combinations
MX
(m)
MY
(m)
MOSMDX
(m)
MSX
(m)
MASSA
(deg)
MASSAR
(deg/s)
IC#1FWS0.77
(38%)
−0.0170.95.95
(31%)
1.67
(62%)
0.8713.14
IC#24WS1.01
(36%)
−0.0170.96.03
(33%)
1.42
(71%)
0.7012.90
IC#3
FWS
+RWS0.29
(75%)
−0.0170.95.28
(38%)
0.28
(94%)
0.8313.73
+RWS + 4WID0.36
(71%)
−0.0180.95.78
(32%)
1.24
(73%)
0.7913.52
+RWS + 4WIB1.07
(34%)
−0.0180.96.36
(27%)
1.71
(58%)
0.9113.76
+RWS + 4WID + 4WIB0.46
(69%)
−0.0180.95.57
(41%)
0.50
(91%)
0.8714.48
+4WID1.13
(32%)
−0.0180.97.63
(16%)
4.30
(16%)
0.9613.36
+4WIB1.44
(15%)
−0.0180.97.38
(9%)
3.50
(2%)
1.6313.61
+4WID + 4WIB0.99
(40%)
−0.0180.96.16
(36%)
1.40
(76%)
0.8313.23
Table 4. Summary of the simulation results with MPCN for IC#1, IC#2, IC#3.
Table 4. Summary of the simulation results with MPCN for IC#1, IC#2, IC#3.
Actuator
Combinations
MX
(m)
MY
(m)
MOSMDX
(m)
MSX
(m)
MASSA
(deg)
MASSAR
(deg/s)
IC#1FWS2.10−0.0180.99.425.670.5712.92
IC#24WS2.17−0.0190.99.425.951.4813.32
IC#3
FWS
+RWS2.07−0.0180.99.445.360.7813.23
+RWS + 4WID2.45−0.0180.99.785.840.8013.25
+RWS + 4WIB2.35−0.0190.99.424.850.8612.82
+RWS + 4WID + 4WIB2.17−0.0190.910.166.340.8313.01
+4WID2.35−0.0190.99.805.830.6113.38
+4WIB2.05−0.0190.98.253.770.6611.77
+4WID + 4WIB2.17−0.0190.910.166.340.8313.01
Table 5. Summary of the simulation results with MPCC for IC#1, IC#2, IC#3.
Table 5. Summary of the simulation results with MPCC for IC#1, IC#2, IC#3.
Actuator
Combinations
MX
(m)
MY
(m)
MOSMDX
(m)
MSX
(m)
MASSA
(deg)
MASSAR
(deg/s)
IC#1FWS1.54
(27%)
−0.0180.96.72
(29%)
2.69
(53%)
0.8312.83
IC#24WS1.54
(29%)
−0.0190.96.68
(29%)
2.49
(58%)
1.3313.45
IC#3
FWS
+RWS1.22
(41%)
−0.0180.96.20
(34%)
1.32
(75%)
0.8313.84
+RWS + 4WID1.45
(41%)
−0.0180.97.12
(27%)
2.88
(51%)
0.8013.15
+RWS + 4WIB1.85
(21%)
−0.0180.96.87
(27%)
2.63
(46%)
0.8412.57
+RWS + 4WID + 4WIB1.41
(35%)
−0.0190.96.47
(36%)
1.57
(75%)
0.8413.73
+4WID1.85
(21%)
−0.0190.98.53
(13%)
5.33
(9%)
0.9013.26
+4WIB2.08
(2%)
−0.0180.97.73
(6%)
3.40
(10%)
1.8811.31
+4WID + 4WIB1.41
(35%)
−0.0190.96.47
(36%)
1.57
(75%)
0.8413.73
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Lee, J.; Yim, S. Path Tracking Control with Constraint on Tire Slip Angles under Low-Friction Road Conditions. Appl. Sci. 2024, 14, 1066. https://doi.org/10.3390/app14031066

AMA Style

Lee J, Yim S. Path Tracking Control with Constraint on Tire Slip Angles under Low-Friction Road Conditions. Applied Sciences. 2024; 14(3):1066. https://doi.org/10.3390/app14031066

Chicago/Turabian Style

Lee, Jaepoong, and Seongjin Yim. 2024. "Path Tracking Control with Constraint on Tire Slip Angles under Low-Friction Road Conditions" Applied Sciences 14, no. 3: 1066. https://doi.org/10.3390/app14031066

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop