Chapter 2 Pressure Estimation of a Wet Clutch 2.1 Introduction

Similar documents
Single-track models of an A-double heavy vehicle combination

Torsional vibrations in truck powertrains with dual mass flywheel having piecewise linear stiffness

Estimation of Tire-Road Friction by Tire Rotational Vibration Model

MECH 3140 Final Project

Manufacturing Equipment Control

Modelling the Dynamics of Flight Control Surfaces Under Actuation Compliances and Losses

Simple Car Dynamics. Outline. Claude Lacoursière HPC2N/VRlab, Umeå Universitet, Sweden, May 18, 2005

Methodology for modeling, parameter estimation, and validation of powertrain torsional vibration

Team-Exercises for DGC 100 Modelica Course

Dual-mode switched control of an electropneumatic clutch actuator with input restrictions

Road Vehicle Dynamics

NONLINEAR BACKSTEPPING DESIGN OF ANTI-LOCK BRAKING SYSTEMS WITH ASSISTANCE OF ACTIVE SUSPENSIONS

Final Exam April 30, 2013

ET3-7: Modelling I(V) Introduction and Objectives. Electrical, Mechanical and Thermal Systems

Structural Dynamics Lecture 2. Outline of Lecture 2. Single-Degree-of-Freedom Systems (cont.)

WORK SHEET FOR MEP311

METHOD FOR ANALYZING EPICYCLIC GEARBOXES

Dynamic analysis of the Empact CVT Ratio and slip dependent, non-minimum phase dynamics

IMPACT OF ROAD SURFACE ROUGHNESS AND ENGINE TORQUE ON THE LOAD OF AUTOMOTIVE TRANSMISSION SYSTEM

Experimental analysis and modeling of transmission torsional vibrations

Adaptive observer-based switched control for electropneumatic clutch actuator with position sensor

The single track model

Appendix A: Exercise Problems on Classical Feedback Control Theory (Chaps. 1 and 2)

Answers to questions in each section should be tied together and handed in separately.

EN40: Dynamics and Vibrations. Final Examination Wed May : 2pm-5pm

PY205N Spring The vectors a, b, and c. are related by c = a b. The diagram below that best illustrates this relationship is (a) I

Emil Ritzén. Reg. nr: LiTH-ISY-EX

A 954 C HD. Technical Description Hydraulic Excavator. Machine for Industrial Applications

CONTROL OF THE NONHOLONOMIC INTEGRATOR

Analysis of Nonlinear Characteristics of Turbine Governor and Its Impact on Power System Oscillation

Dynamic Modeling of Fluid Power Transmissions for Wind Turbines

Driveline Modeling and Principles for Speed Control and Gear-Shift Control

Dynamic Modeling and Controller Design for a Seamless Two-Speed Transmission for Electric Vehicles*

Fuzzy Proportional-Integral State Feedback Controller For Vehicle Traction Control System

Analysis and Design of an Electric Vehicle using Matlab and Simulink

An improved brake squeal source model in the presence of kinematic and friction nonlinearities

Vibrations and Properties of Inertia Continuously Variable Transmissions

The basic principle to be used in mechanical systems to derive a mathematical model is Newton s law,

Parameter Prediction and Modelling Methods for Traction Motor of Hybrid Electric Vehicle

Graph-theoretic Modeling and Dynamic Simulation of an Automotive Torque Converter

Modelling and Simulating the Efficiency and Elasticity of Gearboxes

Vehicle longitudinal speed control

The University of Melbourne Engineering Mechanics

Dr Ian R. Manchester

A STUDY ON THE BASIC CONTROL OF SPEED RATIO OF THE CVT SYSTEM USED FOR ELECTRIC VEHICLES

Adaptive Estimation of the Engine Friction Torque

Prediction and Prevention of Tripped Rollovers

Vane Type Rotary Actuators Series Variations

Dynamic Tests on Ring Shear Apparatus

2nd International Conference on Electronic & Mechanical Engineering and Information Technology (EMEIT-2012)

Automated Estimation of an Aircraft s Center of Gravity Using Static and Dynamic Measurements

CHAPTER 6 HEAT DISSIPATION AND TEMPERATURE DISTRIBUTION OF BRAKE LINER USING STEADY STATE ANALYSIS

NONLINEAR PATH CONTROL FOR A DIFFERENTIAL DRIVE MOBILE ROBOT

Powertrain Systems of the Future

Final Examination Thursday May Please initial the statement below to show that you have read it

Driveline Modeling and Control c 1997 Magnus Pettersson Department of Electrical Engineering Linkoping University S{ Linkopin

Multi-Robotic Systems

Application of Adaptive Sliding Mode Control with an Ellipsoidal Sliding Surface for Vehicle Distance Control

Vibrations Qualifying Exam Study Material

ELECTRONIC CONTROL OF CONTINUOUSLY VARIABLE TRANS- The technology of a Continuously Variable Transmission (CVT) has been

2108. Free vibration properties of rotate vector reducer

Electric Vehicle Performance Power and Efficiency

Mechatronics. MANE 4490 Fall 2002 Assignment # 1

Q2. A machine carries a 4.0 kg package from an initial position of d ˆ. = (2.0 m)j at t = 0 to a final position of d ˆ ˆ

Inertia Identification and Auto-Tuning. of Induction Motor Using MRAS

CHAPTER 3 QUARTER AIRCRAFT MODELING

RESEARCH ON AIRBORNE INTELLIGENT HYDRAULIC PUMP SYSTEM

Chapter 14 Oscillations

PHYSICS. Chapter 15 Lecture FOR SCIENTISTS AND ENGINEERS A STRATEGIC APPROACH 4/E RANDALL D. KNIGHT Pearson Education, Inc.

ELEC4631 s Lecture 2: Dynamic Control Systems 7 March Overview of dynamic control systems

Adaptive Control of Variable-Speed Variable-Pitch Wind Turbines Using RBF Neural Network

Selection Calculations For Linear & Rotary Actuators

e jωt = cos(ωt) + jsin(ωt),

Kinematics, Dynamics, and Vibrations FE Review Session. Dr. David Herrin March 27, 2012

Methods and Tools. Average Operating Point Approach. To lump all engine operating points into one single average operating point.

Supplementary Problems

UNIT-I (FORCE ANALYSIS)

Impact. m k. Natural Period of Vibration τ. Static load Gray area Impact load t > 3 τ. Absorbing energy. Carrying loads

Assignment 9. to roll without slipping, how large must F be? Ans: F = R d mgsinθ.

Robotics & Automation. Lecture 25. Dynamics of Constrained Systems, Dynamic Control. John T. Wen. April 26, 2007

Robot Control Basics CS 685

CHAPTER INTRODUCTION

Analysis and Control of Nonlinear Actuator Dynamics Based on the Sum of Squares Programming Method

Vehicle Dynamics CEE 320. Winter 2006 CEE 320 Steve Muench

NONLINEAR CONTROLLER DESIGN FOR ACTIVE SUSPENSION SYSTEMS USING THE IMMERSION AND INVARIANCE METHOD

Linear guide drives. Synchronous shafts The use of synchronous shafts enables several linear axes to be operated with one drive.

Sliding Mode Control: A Comparison of Sliding Surface Approach Dynamics

California Partners for Advanced Transportation Technology UC Berkeley

Investigation of Steering Feedback Control Strategies for Steer-by-Wire Concept

Theory of Vibrations in Stewart Platforms

1820. Selection of torsional vibration damper based on the results of simulation

Tutorial 1 - Drive fundamentals and DC motor characteristics

Funnel control in mechatronics: An overview

Robust Control For Variable-Speed Two-Bladed Horizontal-Axis Wind Turbines Via ChatteringControl

Real-time Motion Control of a Nonholonomic Mobile Robot with Unknown Dynamics

NUMERICAL INVESTIGATION OF A THREE-DIMENSIONAL DISC-PAD MODEL WITH AND WITHOUT THERMAL EFFECTS

Electronic Throttle Valve Control Design Based on Sliding Mode Perturbation Estimator

The present invention refers to an improved. Various types of turbines are known in the. art, with higher or lower efficiencies, that the

Homework Assignment 4 Root Finding Algorithms The Maximum Velocity of a Car Due: Friday, February 19, 2010 at 12noon

Engineering Science OUTCOME 2 - TUTORIAL 3 FREE VIBRATIONS

Transcription:

Chapter 2 Pressure Estimation of a Wet Clutch 2. Introduction In both DCTs and new ATs [9], the change of the speed ratio is regarded as the process of one clutch being engaged with the other being disengaged, namely, the clutch-to-clutch shift. Furthermore, smart proportional valves with a large flow rate are developed for direct clutch pressure control, without using the pilot duty solenoid valve [3]. These valves can be used in new ATs to improve the ability of adapting to different driving conditions, as well as to reduce cost and to improve packaging. For vehicles with a hydraulic cylinder as clutch actuator, the cylinder pressure control becomes important for good shift quality. Sensors measuring the clutch cylinder pressure, however, are seldom used because of the cost and durability. Hence, it is required to estimate the shaft torque or the cylinder pressure, in order to enhance control performance [22]. There have been some studies on the estimation of the transmission shaft torque and the clutch pressure. A sliding mode observer is designed to estimate the torque of an automotive drive shaft in [3, 4]. An adaptive sliding mode algorithm is proposed to estimate the turbine torque of a torque converter in [23]. Furthermore, [22] uses the sliding mode method to estimate the clutch pressure in a hydraulically powered stepped AT. The extended algorithm in [2] is used to estimate the clutch pressure and the transmission output shaft torque simultaneously. In [9, 7], a neural network is suggested to estimate the turbine torque, in which the engine speed, the turbine speed and the oil temperature are inputs. Reference [9] also designs a driving load observer by assuming that the driving load is slowly-varying. In [20], a recursive least squares method with multiple forgetting factors is used to estimate the road grade and the vehicle mass. In [0], a full-order observer is proposed for the pressure monitoring of a torque converter s lock up clutch, where a state-dependent term is appended in the conventional Luenberger state observer to eliminate the effect of possible parameter variations in some sense. The question of how to design this term is crucial for the performance and the implementation of the observer. A new AT with clutch-to-clutch shift technology is considered in this chapter, in which electro-hydraulic actuators are adopted to control the clutches independently. H.Chen,B.Gao,Nonlinear Estimation and Control of Automotive Drivetrains, DOI 0.007/978-3-642-4572-2_2, Science Press Beijing and Springer-Verlag Berlin Heidelberg 204 37

38 2 Pressure Estimation of a Wet Clutch Fig. 2. Schematic graph of an automatic transmission Because of the complex nonlinearities in an automotive powertrain, such as the speed torque relationship of engines and the characteristics of torque converters, it is very hard to model the whole dynamics with physical principles. Lookup tables, which are obtained from many experiments in the steady state, are widely used to describe the nonlinear characteristics. There inherently exist model uncertainties, such as steady-state error and unmodeled dynamics. Moreover, the variation of the vehicle mass and the road grade also bring uncertainties to the powertrain dynamics. Therefore, the clutch pressure/torque estimator must be robust against the variation of powertrain parameters and the uncertainties. 2.2 Description and Modeling of a Powertrain System We consider the powertrain in passenger vehicles with a two-speed AT, as schematically shown in Fig. 2.. A planetary gear set is adopted as the shift gear. Two clutches are used as the actuators, and two proportional pressure valves are used to control the two clutches. When clutch A is engaged and clutch B disengaged, the This chapter uses the content of [4], with permission from IEEE.

2.2 Description and Modeling of a Powertrain System 39 Fig. 2.2 Engine torque map with speed and throttle opening powertrain operates in the first gear and the speed ratio is given by i = + γ, (2.) where γ is the ratio of the teeth number of the sun gear to that of the ring gear. While clutch A is disengaged and clutch B engaged, the vehicle is driven in the second gear with a speed ratio of i 2 =. (2.2) The powertrain simulation model is established by the commercial simulation software AMESim. Except for the simplified 2-speed transmission, the simulation model represents a typical front-wheel-drive mid-size passenger car equipped with a 2000 cc injection gasoline engine. The constructed model captures the important transient dynamics during the vehicle shift process, such as the drive shaft oscillation and the tire slip. Moreover, the time-delays in control and time-varying parameters are also considered in the simulation model of the proportional valves [6], which are neglected in the observer design. Engine The work reported here is primarily concerned with shift transients, and therefore a simple engine model is used. The dynamic equation of the engine speed is represented by I e ω e + C e ω e = T e T p, (2.3) where T e is the engine output torque and T p is the output torque of the converter pump. The engine output torque is simplified as a nonlinear function of the engine rotational speed ω e and the engine throttle angle θ th, i.e., T e = T e (ω e,θ th ). This map isshowninfig.2.2. Torque Converter The capacity factor C(λ) and the torque ratio t(λ) of the considered torque converter are given in Fig. 2.3.

40 2 Pressure Estimation of a Wet Clutch Fig. 2.3 Capacity factor C(λ) and torque ratio t(λ) of the torque converter Planetary Gear Set Using the submodel provided by commercial software, such as AMESim, the planetary gear set can be modeled conveniently. The following parameters are required for the modeling setting: the inertia moment of the torque converter turbine I t ; the inertia moment of the ring gear I r ; the teeth number of the sun gear Z s and the teeth number of the ring gear Z r. Differential Gear Box and Drive Shaft The gear ratio of the differential gear box is denoted as R df. The two drive shafts between the differential gear and the front wheels are represented as a torsion spring with stiffness coefficient K s and a torsion damping with damping coefficient C s. Tires The longitudinal tire force F x, which is usually simplified as a function of the longitudinal slip ratio S x, rises fast when S x increases under a threshold ds x and declines slowly after that [7]. Here it is represented approximately as a tanh function of F x = F x max tanh( 2S x ds x ). The longitudinal slip S x has been defined in Sect... Road Loads The road load consists of three parts: the grade force F G, the rolling resistant moment of tires T w and the aerodynamic drag F A, which has been introduced in Sect... Clutches and Valves The friction coefficient μ is a nonlinear function of ω showninfig.2.5. In the design of the pressure observer in Sect. 2.3.2, we assume the parameters (τ cv,k cv ) of the proportional pressure control valve as constant, and we also ignore the time-delay of the valve. Actually, the valve has a time-delay and the parameters vary according to different operating points [6]. Hence, the dynamics of the proportional valve in the powertrain simulation is given by τ cv ṗ cb (t) = p cb (t) + K cv i b (t L cv ). (2.4) Finally, the values of the parameters used in the powertrain simulation are listed in Table 2.. Nonlinear functions T e (ω e,θ th ), C(λ), t(λ), μ( ω), τ cv, L cv and K cv are given in the lookup tables.

2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness 4 Table 2. Values of simulation model parameters A A Front area of vehicle 2 m 2 C D Aerodynamic drag coefficient 0.3 C e Damping coefficient of engine crane 0.047 Nm/rad C l Damping coefficient of drive shaft 0 Nm/rad ds x Longitudinal slip threshold of tire 0. F s Return spring force of clutch B 600 N F x max Maximum longitudinal force of tire 3200 N i df Gear ratio of the differential gear box 3.0 I e Inertia of crane and pump 0.7 kg m 2 I r Inertia moment of ring gear 0.0 kg m 2 I t Inertia moment of turbine 0.06 kg m 2 K cv Gain of valve B 0.7 MPa/A K l Stiffness of drive shaft 6500 Nm/rad L cv Time-delay of valve B 0 0.03 s m Vehicle mass 500 kg R w Tire radius 0.3 m T w Moment of resistance of tires 0 Nm Z r Teeth number of ring gear 60 Z s Teeth number of sun gear 40 θ g Road grade 0 deg ρ Air density.2 kg/m 3 τ cv Time constant of valve B 0.02 0.20 s 2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness A reduced-order clutch pressure observer based on the concept of input-to-state stability (ISS) [2, 8] is proposed, where the rotational speeds are the measured outputs and the special structure of the clutch pressure system is exploited. The lookup tables of the nonlinear characteristics of powertrain systems appear in their original map form, and the model uncertainties are considered as additional disturbance inputs. A systematic procedure is given to design the nonlinear clutch pressure observer such that The error dynamics is input-to-state stable, where modeling errors are the inputs. This means that the initial estimation error decays exponentially and the estimation error is guaranteed to be bounded for the bounded modeling errors; The requirements on estimation performance, such as decay rate and error offset, are easily and explicitly considered during the design process; The implementation of the designed observer benefits from the reduced order and the time-invariant gains of the observer;

42 2 Pressure Estimation of a Wet Clutch Fig. 2.4 st-to-2nd gear shift (st gear driving torque phase inertia phase 2nd gear driving) Lower observer gains are obtained through convex optimization, which increases the robustness against noises and reduces the estimated upper bound of the error offset. 2.3. Clutch System Modeling and Problem Statement During the shift process, the on-coming and off-going clutches are controlled by the two valves through separate controllers, which are assumed to be well-designed. The controllers discussed in Chap. 3 and Chap. 4 can be viewed as such controllers. The power-on st-to-2nd upshift is considered here as an example of the shift process. The gear shift process is generally divided into the torque phase, as shown in Fig. 2.4, where the turbine torque is transferred from clutch A to clutch B and the inertia phase where clutch B is synchronized [8]. Note that 4 8-speed ATs are extensively used in production cars, a 2-speed AT is just adopted here as an example to exploit the design process, and it can be applied to other ATs if the parameters of the clutch-to-clutch shift process are replaced. We start to describe the modeling of the inertia phase and then obtain a model for the torque phase by taking into account the general fact that there is no gear ratio change in the torque phase. Inertia Phase In the inertia phase, where the two clutches are both slipping, because the planetary gear set is a two-degree-of-freedom system, two states variables, such as the turbine speed ω t and the speed ω r of the ring gear, can be used to describe its movement.

2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness 43 The basic kinematic equation of the planetary gear set is ω r ω 0 = γ (2.5) ω t ω 0 where γ is the ratio of the teeth number of the sun gear to that of the ring gear, ω r is the ring gear speed, ω 0 is the planetary gear carrier speed, i.e., the output speed of the transmission, ω t is the sun gear speed, i.e., the turbine speed of the torque converter. The torque of the planetary gear set should also satisfy the torque balance equations: T SG = γ + γ T 0, (2.6a) T RG = + γ T 0, (2.6b) where T SG is the sun gear torque, T RG is the ring gear torque, T 0 is the planetary gear carrier torque. With the driving traction of torque T 0, the vehicle body will move according to the equation I ve ω 0 = T 0 T ve. (2.7) where I ve is the equivalent inertia of the vehicle body, T ve is the equivalent driving resistant torque. From Fig. 2., the equations of the sun gear and the ring gear are given as I t ω t = T t T cb T SG, I r ω r = T cb + T ca T RG, (2.8a) (2.8b) where I t is the inertia moment of the torque converter turbine, I r is the inertia moment of the ring gear and parts connected, T ca is the torque delivered by clutch A, T cb is the torque delivered by clutch B. Substituting (2.6a), (2.6b) and (2.7)into(2.8a), (2.8b), we have I t ω t + R g I ve ω 0 = T t T cb R g T ve, (2.9a) I r ω r + R g R g I ve ω 0 = T cb + T ca R g R g T ve (2.9b) with R g = + γ. Using (2.5) and rearranging the above equations, the dynamic equations of the transmission can be obtained as follows: ω t = C T t + C 2 T ca + C 3 T cb + C 4 T ve, ω r = C 2 T t + C 22 T ca + C 23 T cb + C 24 T ve, (2.0a) (2.0b)

44 2 Pressure Estimation of a Wet Clutch with the coefficients C = R 2 g I r + (R g ) 2 I ve R 2 g I ti r + (R g ) 2 I t I ve + I r I ve, (R g )I ve C 2 = C Rg 2I r + (R g ) 2, I ve [ ] (R g )I ve C 3 = C Rg 2I r + (R g ) 2 +, I ve R g I r C 4 = C Rg 2I r + (R g ) 2, I ve (R g )I ve C 2 = Rg 2I ti r + (R g ) 2, I t I ve + I r I ve Rg 2 C 22 = C I t + I ve 2, (R g )I ve [ R 2 ] g I t + I ve C 23 = C 2 +, (R g )I ve R g I t C 24 = C 2. I ve The turbine torque T t is calculated from the steady-state characteristics of the torque converter as T t = t(λ)c(λ)ωe 2, (2.) where C(λ) denotes the capacity factor of the torque converter, t(λ) is the torque ratio, ω e is the engine speed and λ is the speed ratio defined as λ = ω t. (2.2) ω e On the other hand, the transferred torque T cb during clutch slipping is determined by the cylinder pressure. If the force of the return spring is treated as constant, the relationship between the clutch torque and the cylinder pressure is described as T cb = μ( ω)rn (Ap cb F s ), (2.3) where R is the effective radius of the push force acted on the plates of clutch B, N and A are the plate number and the piston area of clutch B, p cb is the pressure of cylinder B, F s is the return spring force of clutch B and μ is the friction coefficient of clutch plates depending on the speed difference. The speed difference ω is defined as ω = ω t ω r. (2.4)

2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness 45 The cylinder pressure is determined by the input current of the proportional pressure control valve. The dynamics of the proportional valve can be simplified as a first-order system [6], τ cv ṗ cb = p cb + K cv i b, (2.5) where τ cv is the time constant of valve B, K cv is the gain of valve B, and i b is the electric current of valve B. Moreover, if the torsion dynamics of the drive shaft, the tire slip and the road grade are ignored, the resistant torque T ve, delivered from the tire to the drive shaft, can be calculated as T ve = T w + C ARw 3 ω0 2 R, (2.6) df where T w denotes the rolling resistance moment of the tire, R w is the tire radius, R df is the gear ratio of the differential gear box, ω 0 is the output speed of the transmission and C A is a constant coefficient depending on air density, aerodynamic drag coefficient and the front area of the vehicle. By selecting the turbine speed ω t, the speed difference ω of clutch B, and the pressure p cb of cylinder B as state variables, denoted as x,x 2,x 3, respectively, and ignoring the pressure of clutch A because it is small enough, the inertia phase of the st-to-2nd gear upshift process is described in the following state space form: R 3 df ẋ = C 3 μ(x 2 )RNAx 3 + f (ω e,x,x 2 ), ẋ 2 = (C 3 C 23 )μ(x 2 )RNAx 3 + f 2 (ω e,x,x 2 ), ẋ 3 = τ cv x 3 + K cv τ cv u, (2.7a) (2.7b) (2.7c) with f (ω e,x,x 2 ) = C T t (ω e,x ) + C 4 T ve (x,x 2 ) C 3 μ(x 2 )RNF s, f 2 (ω e,x,x 2 ) = (C C 2 )T t (ω e,x ) + (C 4 C 24 )T ve (x,x 2 ) (C 3 C 23 )μ(x 2 )RNF s, (2.8a) (2.8b) where u = i b is the current of valve B and viewed as control input. In order to estimate the pressure of clutch B, the rotational speeds of the transmission are used as the measured outputs, i.e., The parameter values can be found in Table 2.2. y =[x x 2 ] T. (2.9)

46 2 Pressure Estimation of a Wet Clutch Table 2.2 Parameters for observer design C Coefficient in (2.7a) (2.7c) 5.52 C 3 Coefficient in (2.7a) (2.7c) 25.85 C 4 Coefficient in (2.7a) (2.7c) 0.0 C 2 Coefficient in (2.7a) (2.7c) 0.33 C 23 Coefficient in (2.7a) (2.7c) 7.38 C 24 Coefficient in (2.7a) (2.7c) 0.0 R Effective radius of plates of clutch B 0.3 m N Plate number of clutch B 3 A Piston area of clutch B 0.0 m 2 kg m 2 kg m 2 kg m 2 kg m 2 kg m 2 kg m 2 τ cv Time constant of valve B 0.04 s K cv Gain of valve B.0 MPa/A μ min Minimum friction coefficient 0.0 μ max Maximum friction coefficient 0.6 c Coefficient in (2.2a), (2.2b) 0.40 c 3 Coefficient in (2.2a), (2.2b) 0.40 c 4 Coefficient in (2.2a), (2.2b) 0.6 kg m 2 kg m 2 kg m 2 Torque Phase In the st-to-2nd upshift torque phase, there is no slip in clutch A, hence, there is no gear ratio change of the transmission. The motion of the drive line during this phase can be described by the following equation: with the constant coefficients ω t = c T t + c 3 T cb + c 4 T ve, (2.20) Rg 2 c = Rg 2I, t + I ve c 3 = c, c 4 = c R g, where ω t is the turbine speed, T cb is the torque delivered by clutch B, T ve is the resistant torque delivered from the tire to the drive shaft, I t is the inertia moment of the torque converter turbine, I ve is the equivalent inertia moment of the vehicle body and R g = + γ again. The torque phase of the st-to-2nd gear upshift process is then described in the following state space form:

2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness 47 ẋ = c 3 μ( ω)rnax 3 + f t (ω e,x, ω), ẋ 3 = τ cv x 3 + K cv τ cv u, (2.2a) (2.2b) with the measured output and y = x (2.22) f t (ω e,x, ω)= c T t (ω e,x ) + c 4 T ve (x, ω) c 3 μ( ω)rnf s. (2.23) Note that c ij are different coefficients from C ij in (2.8a), (2.8b). Moreover, although there is no obvious change of the clutch speed during the torque phase, the speed difference ω of the clutch is different for various driving maneuvers. Hence ω is also considered as an input for the torque phase model. Estimation Problem Due to the extreme complexity of the torque converter and the aerodynamic drag, the nonlinear functions in (2.8a), (2.8b) and (2.23) are generally given as lookup tables (i.e., maps), which are obtained by a series of steady-state experiments and inherently contain errors. Other modeling uncertainties include variations of parameters, such as the vehicle mass, the road grade and the damping coefficient of shafts. Hence, the problem considered here is to estimate the pressure of clutch B in the presence of model errors, given the valve electric current i b and the measured rotational speeds of the transmission ω e, ω t and ω. 2.3.2 Reduced-Order Nonlinear State Observer Reduced-Order Nonlinear Observer with ISS Property In this section, the special structure of the clutch pressure system is considered to derive a reduced-order pressure observer. The robustness of the designed observer with respect to model errors is achieved in the sense of ISS property. To do this, we denote the variable to be estimated as z, and rewrite the system dynamics for estimating the clutch pressure as follows: ẏ = F (y,u)+ G(y,u)z+ w(y,u,z), ż = A 22 (u, p)z + B 2 (u, p), (2.24a) (2.24b) where y is the measured output, u is the control input and p is the vector of parameters which may include t and others, w(y,u,z) summarizes model uncertainties,

48 2 Pressure Estimation of a Wet Clutch and in particular ( f (ω F (y,u)= e,x,x 2 ) f 2 (ω e,x,x 2 ) ( G(y,u)= A 22 (u, p) = τ cv, C 3 μ(x 2 )RNA (C 3 C 23 )μ(x 2 )RNA ), (2.25a) ), (2.25b) (2.25c) B 2 (u, p) = K cv τ cv u (2.25d) for the inertia phase. The expressions for the torque phase are F (y,u)= f t (ω e,x, ω), G(y,u)= c 3 μ( ω)rna, A 22 (u, p) =, τ cv B 2 (u, p) = K cv τ cv u. (2.26a) (2.26b) (2.26c) (2.26d) Remark 2. We exploit the more general form of (2.24a), (2.24b) to derive the pressure observer such that the suggested design method might be useful if the timevarying property of the proportional pressure valves is taken into account (where p = t), or if other kinds of valves are used to control clutch pressures. For example, if a PWM valve is used, the pressure dynamics can be described by [22] ż = C z 0.0Pl u z, (2.27) where C z is a positive constant, u is the pulse duty cycle, and P l is the main line pressure. Then, for all given u, we may linearize (2.27) at a fixed operating point of z = P s to approximate the pressure dynamics in the form of (2.24b) with In this case, p =[P s P l ] T. C z A 22 (p) = 2, 0.0P l u P s B 2 (p) = C z 0.0Pl u P s + C z P s 2 0.0P l u P s. Because the shaft torque affects the related shaft accelerations directly, the difference between the true accelerations ẏ and the estimated values F (y,u)+ G(y,u)ẑ is used to constitute the correction term. Hence, let the observer be designed in the

2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness 49 form of ẑ = A 22 (u, p)ẑ + B 2 (u, p) + L ( ẏ F (y,u) G(y,u)ẑ ), (2.28) where L R 2 in the inertia phase (or L R in the torque phase) is the observer gain to be determined. By defining the observer error as the error dynamics can then be described by e = z ẑ, (2.29) ė = A 22 (u, p)z + B 2 (u, p) (2.30) ( A 22 (u, p)ẑ + B 2 (u, p) + L ( ẏ F (y,u) G(y,u)ẑ )) = ( A 22 (u, p) LG(y,u) ) e Lw. (2.3) We define V = 2 et e and differentiate it along the solution of (2.30) to infer V = e T ė = e T ( A 22 (u, p) LG(y,u) ) e e T Lw. (2.32) Applying Young s Inequality [2] (see Lemma B. in Appendix B) to the last term of the above equality leads to with κ > 0. Then, (2.32) becomes e T Lw κ e T e + 4κ w T L T Lw (2.33) V e T ( A 22 (u, p) LG(y,u)+ κ ) e + 4κ w T L T Lw. (2.34) We now choose L to satisfy the following inequality with κ 2 > 0, then we arrive at A 22 (u, p) LG(y,u)+ κ κ 2 (2.35) V κ 2 e T e + 4κ w T L T Lw, (2.36) which implies that the error dynamics admits the input-to-state stability property if the model error w is supposed to be bounded in amplitude (see Lemma B.2 in Appendix B). Moreover, it follows from (2.36) that V 2κ 2 V + 4κ w T L T Lw. (2.37)

50 2 Pressure Estimation of a Wet Clutch Upon multiplication of (2.37)bye 2κ 2t, it becomes Integrating it over [0,t] leads to d ( Ve 2κ 2 t ) w T L T Lwe 2κ2t. (2.38) dt 4κ V(t) V(0)e 2κ2t + t e 2κ2(t τ) w(τ) T L T Lw(τ) dτ. (2.39) 4κ 0 Hence, the properties of the error dynamics of the designed observer (2.28) are described as follows: Theorem 2. Suppose that κ > 0,κ 2 > 0; The observer gain L is chosen to satisfy (2.35). Then, the error dynamics of the observer (2.28) is (a) Input-to-state stable, if w is bounded in amplitude, i.e., w L ; (b) Exponentially stable with κ 2 for w = 0. Proof It follows from (2.36) that V κ 2 e 2 + 4κ λ max (L T L) w 2, which shows that the error dynamics admits the input-to-state stability property [2, p. 503] (see Lemma B.2 in Appendix B) if the model error w is supposed to be bounded in amplitude, as property (a) required. By taking w = 0, we obtain from (2.39) that e(t) e(0) e κ2t, t 0 which proves property (b). Remark 2.2 By the equivalences of the ISS property listed in Appendix B, property (a) implies that the error dynamics of the designed observer (2.28) isrobustly stable if w is viewed as the effect of model uncertainties. This is the case in Sect. 2.3.3. Remark 2.3 Now we give a discussion on the parameters κ and κ 2. From property (b), κ 2 is chosen according to the required decay rate of the error. If w is bounded in amplitude, i.e., w L, then (2.39) becomes e(t) 2 e(0) 2 e 2κ 2 t + 2 2 which implies that w 2 sup λ max (L T L) t [0,t] e 2κ2(t τ) dτ, (2.40) 4κ 0 e( ) 2 w 2 sup(λ max(l T L)) t lim e 2κ2(t τ) dτ, (2.4) 2κ t 0

2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness 5 and furthermore, e( ) 2 w 2 sup(λ max(l T L)). (2.42) 4κ κ 2 Hence, one may choose a larger κ to reduce the offset. From (2.35), however, one should also notice that the larger the κ, the higher the observer gain. Remark 2.4 Inequality (2.42) gives just an upper bound on the estimation error offset if a bound on the model error is given. The real offset could be much smaller, due to the multiple use of inequalities in the above derivation. Remark 2.5 Besides satisfying (2.35), we do not impose other assumptions on the observer gain L. This implies that L can be designed theoretically to depend on (y,u), while one may choose it as time-invariant (constant) in practice. A solution with time-invariant L will be discussed in the next section. Implementation Issues In order to avoid taking derivatives of the measured variables, let then, we can infer for a time-invariant L that to arrive at η = ẑ Lẏ η =ẑ Ly, (2.43) = A 22 (u, p)ẑ + B 2 (u, p) + L ( ẏ F (y,u) G(y,u)ẑ ) Lẏ, (2.44) η = ( A 22 (u, p) LG(y,u) ) (η + Ly) + B 2 (u, p) LF (y,u). (2.45) Equations (2.43) and (2.45) constitute the reduced-order observer for the nonlinear clutch slip control system. We notice that the nonlinearities of the powertrain system appear in their original form in the observer. Therefore, the merits arise the characteristics of powertrain mechanical systems, such as the characteristics of the engine and torque converter, are represented in the form of lookup tables which are easy to process on a computer. According to Theorem 2. and Remark 2.3, a systematic procedure is given to design the reduced-order nonlinear clutch pressure observer in the form of (2.43) and (2.45) as follows: Step Choose parameter κ 2 according to the required decay rate of the estimation error; Step 2 Choose parameter κ, where it is suggested to start from some smaller values (according to Remark 2.3);

52 2 Pressure Estimation of a Wet Clutch Step 3 Determine the observer gain L such that (2.35) is satisfied; Step 4 For a given model error bound, use (2.42) to compute the estimated upper bound of the offset and check if the offset bound is acceptable. Step 5 If the offset bound is acceptable, end the design procedure. If not, go to Step 2. It is well known that getting model error bounds is in general very difficult, if not impossible. As mentioned in Remark 2.4, for a given model error bound, (2.42) gives just an upper bound of the estimation error offset, which might be much larger than the real offset. Hence, the stopping rule of iterating Step Step 5 is somehow a rule of thumb. We now give a solution of (2.35) for choosing L to be time-invariant, where the requirement for low observer gains can be considered through optimization. If A 22 (u, p) and G(y,u)in (2.35) vary in a polytope with r vertices, i.e., ( A22 (u, p) G(y,u) ) Co {( A 22, G ), ( A22,2 G 2 ),..., ( A22,r G r )}, (2.46) where Co{ } denotes the convex hull of the polytope, then, there exist satisfying such that β 0,β 2 0,...,β r 0 (2.47) r β i = (2.48) i= ( A22 (u, p) G(y,u) ) = Hence, the result is given as follows. r ( ) β i A22,i G i. (2.49) Theorem 2.2 Suppose that A 22 (u, p) and G(y,u) vary in a polytope as (2.46). Then, any time-invariant L satisfying the following Linear Matrix Inequalities (LMIs) i= A 22,i LG i + κ + κ 2 0, i =, 2,...,r (2.50) meets the observer gain condition (2.35). Proof Since A 22 (u, p) and G(y,u) vary in a polytope as (2.46), then, we have (2.49) with (2.47) and (2.48). By the convexity of A 22 (u, p) LG(y,u)+ κ + κ 2

2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness 53 Fig. 2.5 Friction characteristics of clutch plates in A 22 and G, and by the use of Jensen s inequality (see Appendix E), we infer r A 22 (u, p) LG(y,u)+ κ + κ 2 β i (A 22,i LG i + κ + κ 2 ). (2.5) i= Hence, β i 0 and the satisfaction of (2.50) guarantees A 22 (u, p) LG(y,u)+ κ + κ 2 0, (2.52) as required. In (2.50), G i and A 22,i are known and bounded, κ and κ 2 are selected to be bounded and r = 2 m (where m is the number of time-varying parameters in G(y,u) and A 22 ) is bounded, too. Hence, some constant L is always found to render make it hold. Moreover, we prefer low observer gains, due to robustness against noises and the reduction of the upper bound of the error offset, which is estimated by (2.42). Hence, by the use of Schur complement (see Appendix E), L is obtained through the following LMI optimization min α,l α subject to LMIs (2.50) and ( α L L T I ) 0. (2.53) Given κ and κ 2, the solution of (2.53) gives then a constant observer gain with the lowest possible gains satisfying condition (2.35). Observer Design for Clutch Pressure The inertia phase is taken as an example to show the detailed design procedure. The parameters (τ cv,k cv ) are regarded as constants for simplicity, which are listed in Table 2.2 together with the other parameters. Nonlinear functions f,f 2 and μ are given as lookup tables for the observer. The map of μ isshowninfig.2.5, while f,f 2 are given by third-order maps and examples when ω e = 500 rad/s are shown in Fig. 2.6. These parameters are derived from the nominal setting of an AMESim simulation model of the AT shown in Fig. 2..

54 2 Pressure Estimation of a Wet Clutch Fig. 2.6 MAPs of f,f 2 when ω e = 500 rad/s Following the procedure given in Sect. 2.3.2, κ 2 is chosen to meet the requirement for the desired decay rate of the estimation error. It is desired that the error converges in 0. s. Then, taking the settling time as 4 times the time constant [5, p. 22] leads to κ 4 2 = 0. and results in κ 2 = 40. Then, κ is chosen with the purpose of achieving a smaller offset of the estimation error. Start with κ = and obtain L = ( 78 20) and e( ) 0.75 MPa (see the following for the detailed calculation). The offset bound is too large for real applications. According to Step 2 of the procedure given in the above subsection, we enlarge the value of κ, and finally, the value being used is κ = 5. We now solve the optimization problem (2.53) to obtain the lowest possible observer gain. Since A 22 = τ cv is considered as constant, the polytope in (2.46) is given by G(y,u)= Co{G, G 2 }, where the two vertices are computed by (2.25b) with μ min μ(x 2 ) μ max. The solution reads L = ( 783 30). In order to check if the estimation offset is acceptable, we now roughly compute the bound of modeling errors. Since powertrain systems admit highly nonlinear, complex dynamics and various uncertainties, it is indeed difficult, if not impossible, to obtain a comprehensive estimate of the modeling error bound. Hence, some major uncertainties are taken as examples to estimate the value of w. The major uncertainties here are calculation errors of F (y,u) in (2.24a), (2.24b), which contains the turbine torque T t and the vehicle driving load T ve. The change of the vehicle mass affects also the coefficients C ij in F (y,u). From numerous simulations of different powertrain settings, a bound on w is determined as w = 600 rad/s 2. According to (2.42), an upper bound of the offset is obtained for the designed observer.

2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness 55 The result is e( ) 0.05 MPa, which is less that 0 % of the variation range of the working pressure of the valve and is acceptable. Similarly, following the procedure given in the above subsection, the observer gain for the torque phase is calculated, and the result is L = 5.02 0 4. 2.3.3 Simulation Results The proposed clutch pressure observer is programmed using MATLAB/Simulink and combined with the above complete powertrain simulation model through cosimulations. The two clutch valves are controlled by a pre-designed clutch slip controller to ensure a rapid and smooth shift process. In this study, the major concern is put on the power-on st-to-2nd gear upshift process. Figure 2.7 gives the simulation results of the shift process with the driving condition of Table 2., i.e., the condition for the observer design. During the shift process, the engine throttle angle is adjusted to cooperate with the transmission shift. In both torque and inertia phases, the pressure of cylinder B is estimated by the designed observers. After the inertia phase (after 8.34 s), because the clutch B has been locked up, the pressure is computed from the simplified control valve dynamics (2.7c). During the torque phase (between 7.7 and 7.94 s), the rotational speeds do not change much, whereas during the inertia phase (between 7.94 and 8.34 s), the rotational speeds change intensively because of the clutch slip. Hence, the estimation performance in the inertia phase is much better, although it is also acceptable in the torque phase. The estimation error is plotted in the bottom of Fig. 2.7 as the solid line, where the result for L = 0 is also given for comparison. The error peak is reduced by about 35 % and the average error is reduced by about 3 %. Note that the shift process operates in the nominal driving condition, but the stiffness of the drive shaft and the tire slip are considered in the simulation model, while these are ignored in the model for designing the observer. Moreover, the time-delay in control and time-varying parameters are also considered in the simulation model of the proportional valve. The proposed observer is now tested under the driving conditions which deviate from the nominal setting, where the vehicle mass, road grade, torque characteristics of the engine and the torque converter are varied. We increase or decrease each of the items, and carry out simulations under different combination of these changes. The results with relatively large errors are shown in Fig. 2.8, where the driving condition setting is as follows: the torque characteristic of the engine is enlarged by 5 %, and subsequently the capacity of the torque converter is also enlarged; the vehicle mass is increased from 500 to 725 kg, and the road grade angle is varied from 0 to 5 degrees. Due to the large model errors, the pressure estimation error becomes larger in the torque phase. The reason is that there is no slip in clutch A during the torque

56 2 Pressure Estimation of a Wet Clutch Fig. 2.7 Results of the nominal driving condition phase, and no large change of the transmission speeds for the large vehicle inertia. Therefore, the torsion of the drive shaft and the tire slip play important roles in the drive line. The omission of these terms in the observer design deteriorates the estimation performance. In the inertia phase, because of the clutch slip, the designed observer still works well and the pressure estimation error is acceptable. 2.3.4 Design of Full-Order Sliding Mode Observer and Comparison As a comparison, a full-order sliding mode observer is designed according to [3, 22]. Taking the inertia phase as an example, we rewrite system equa-

2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness 57 Fig. 2.8 Results of different driving condition tions (2.24a), (2.24b) as ẏ = C 3 μ(y 2 )RNAz + f, ẏ 2 = (C 3 C 23 )μ(y 2 )RNAz + f 2, ż = τ cv z + K cv τ cv u. (2.54a) (2.54b) (2.54c) Following [3], the sliding mode observer can be designed in the following form: ŷ = C 3 μ(ŷ 2 )RNAẑ + fˆ + κ s sign(ỹ ), ŷ 2 = (C 3 C 23 )μ(ŷ 2 )RNAz + fˆ 2 + κ s2 sign(ỹ 2 ), ẑ = ẑ + K cv u + κ s3 sign(ỹ ) + κ s4 sign(ỹ 2 ), τ cv τ cv (2.55a) (2.55b) (2.55c)

58 2 Pressure Estimation of a Wet Clutch where κ s,κ s2,κ s3 and κ s4 are observer gains, and ỹ = y ŷ, ỹ 2 = y 2 ŷ 2. Gains κ s and κ s2 should satisfy the following sliding condition: κ s > f + C 3 μrna z 920, κ s2 > f 2 + (C 3 C 23 )μrna z 380 (2.56a) (2.56b) with f = f fˆ, f 2 = f 2 fˆ 2, z = z ẑ. Thus, κ s and κ s2 are selected as κ s = 2000, κ s2 = 3200. (2.57a) (2.57b) According to the desired estimation offset and error decay rate, gains κ s3 and κ s4 can be calculated as κ s3 = 4.2 0 6, κ s4 = 7 0 6. (2.58a) (2.58b) Similarly, the sliding mode observer for the torque phase can be designed in the following form ŷ = c 3 μ( ω)rnaẑ + fˆ t + κ st sign(ỹ ), ẑ = ẑ + K cv u + κ st2 sign(ỹ ), τ cv τ cv and the gains are calculated as κ st = 00, κ st2 = 7.5 0 6. (2.59a) (2.59b) (2.60a) (2.60b) The sampling frequency of the sliding mode observer is chosen to be 00 Hz, in order to test the feasibility of the resulting observer for real applications [9]. In the discrete implementation, the observer gains have to be reduced in order to restrain oscillations resulting from sampling and two sets of the tuned values are given in

2.3 Clutch Pressure Estimation Without Consideration of Drive Shaft Stiffness 59 Table 2.3 Gains of discrete observers ISS Torque phase Inertia phase L =.8 0 4 L = ( 320 540) Sliding (large gains) Sliding 2 (small gains) κ st = 80 κ s = 000 κ st2 =.5 0 6 κ s2 = 600 κ s3 = 8 0 5 κ s4 =.2 0 6 κ st = 80 κ s = 000 κ st2 = 3.5 0 5 κ s2 = 600 κ s3 = 4 0 5 κ s4 = 7 0 5 Fig. 2.9 Comparison between ISS observer and sliding mode observer (torque converter capacity is enlarged by 5 %; m = 725 kg; θ g = 5 ) Table 2.3. Hence, the proposed ISS observer is also discretized by the same sampling frequency and the tuned gains are also listed in Table 2.3. The comparison results of these three observers are shown in Fig. 2.9, where the driving condition is the same as that of Fig. 2.8.InFig.2.9, the solid line represents the error of the reduced-order observer, while the dotted and dashed lines represent the error of the full-order sliding mode observers with the large and small gains, respectively. It is seen that the proposed reduced-order observer works well in the inertia phase. The sliding mode observer with large gains (Sliding ) tracks true values without large errors but with chatters, while the other sliding mode observer (Sliding 2) achieves few chatters at the cost of the large estimation errors. As for robustness, the proposed observer achieves robustness in the sense of input-to-state stability, where the model errors are represented as external inputs.

60 2 Pressure Estimation of a Wet Clutch 2.4 Clutch Pressure Estimation when Considering Drive Shaft Stiffness In the above, a reduced-order clutch pressure observer was proposed when considering the concept of input-to-state stability (ISS). However, it is pointed out that during the torque phase, the estimation error becomes somehow unacceptable. During the torque phase, the engine torque is transferred from the off-going clutch to the on-coming clutch. If the clutch pressure can be estimated accurately, the precise timing of releasing and applying clutches can be guaranteed to prevent the clutches from tying-up and the traction interruption. Hence, in order to improve the estimation precision of the clutch pressure during the shift torque phase, the methodology proposed in the above section is extended to design an observer when considering the driveline stiffness. Because the drive axle shafts are the main components of the whole driveline, the rotational freedom of the drive shaft is introduced into the model-based design. The newly designed observer can simultaneously estimate the drive shaft torque as well as improve the accuracy of the clutch pressure estimation [6]. 2.4. Clutch System Modeling when Considering the Drive Shaft The power-on st-to-2nd upshift is still considered as the example, and the pressure observer is designed to estimate the clutch pressure during the shift process. When considering the drive shaft compliance, the system models can be constructed as follows. Torque Phase In the st-to-2nd upshift torque phase, it is assumed that there is no slip in clutch A, and the motion of the drive line during this phase is represented by the following equations: T s ω t = c T t + c 3 μ( ω)rn(ap cb F s ) + c 4, i df ω w = c 34 T s + c 35 T l, (2.6a) (2.6b) Ṫ s = K s i df i ω t K s ω w, ṗ cb = τ cv p cb + K cv τ cv u, (2.6c) (2.6d) where ω t is the turbine speed, ω w is the speed of the driving wheel (front wheel), T s is the drive shaft torque, p cb is pressure of cylinder B, T t is the turbine torque,

2.4 Clutch Pressure Estimation when Considering Drive Shaft Stiffness 6 Table 2.4 Parameters for observer design c 3 Coefficient of clutch torque in (2.69c).90 c 4 Coefficient of clutch torque in (2.69c) 4.76 c 34 Coefficient of clutch torque in (2.69c) 0.0074 C 3 Coefficient of clutch torque in (2.70b) 24.5 C 4 Coefficient of clutch torque in (2.70b) 0.98 C 23 Coefficient of clutch torque in (2.70b) 29.4 C 24 Coefficient of clutch torque in (2.70b) 8.82 C 34 Coefficient of clutch torque in (2.70b) 0.0074 γ Gear ratio of sun gear to ring gear 0.667 R Effective radius of plates of clutch B 0.3 m N Plate number of clutch B 3 A Piston area of clutch B 0.0 m 2 kg m 2 kg m 2 kg m 2 kg m 2 kg m 2 kg m 2 kg m 2 kg m 2 τ cv Time constant of valve B 0.04 s K cv Gain of valve B.0 MPa/A μ min Minimum friction coefficient 0.0 μ max Maximum friction coefficient 0.6 i df Gear ratio of the differential box 3 K s Stiffness of drive shaft 3000 Nm/rad ω t Normalization of ω t 00 rad/s ω w Normalization of ω w 0 rad/s ω Normalization of ω 00 rad/s T s Normalization of T s 000 Nm p cb Normalization of p cb 0 5 Pa T l is the resistant torque delivered from the tires, F s denotes the return spring force of clutch B and μ is the friction coefficient of clutch B depending on the speed difference ω. The definition of the other parameters can be found in Table 2.4. The turbine torque T t and resistance torque T l in (2.6a) (2.6d) are calculated as follows [7]: T t = t(λ)c(λ)ω 2 e, T l = T w + C A R 3 w ω2 w, (2.62a) (2.62b) where C(λ) denotes the capacity factor of the torque converter, t(λ) is the torque ratio, ω e is the engine speed and λ is the speed ratio defined as λ = ω t ω e, T w denotes the rolling resistance moment of tires, R w is the tire radius, and C A is a constant coefficient depending on air density, aerodynamic drag coefficient and the front area of the vehicle.

62 2 Pressure Estimation of a Wet Clutch Inertia Phase In the inertia phase, the pressure of cylinder A is greatly reduced, and the pressure of cylinder B increases so that the speed difference between ring gear and turbine can be reduced to zero, i.e., we have the engagement of clutch B. The dynamic motion of this phase can be described by the following equations if the drive axle shaft compliance is considered: ω t = C T t + C 3 μ( ω)rn(ap cb F s ) + C 4 T s i df, (2.63a) ω = (C C 2 )T t + (C 3 C 23 )μ( ω)rn(ap cb F s ) + (C 4 C 24 ) T s i df, ω w = C 34 T s + C 35 T l, Ṫ s = K s i df ( ω t + γ ω ṗ cb = τ cv p cb + K cv τ cv u, ) K s ω w, (2.63b) (2.63c) (2.63d) (2.63e) where ω is the slip speed of clutch B, i.e., the speed difference between the turbine and the ring gear, C ij are constant coefficients determined by inertia moments of the vehicle and transmission shafts; note that C ij are different from c ij of (2.6a) (2.6d). The models in consideration of the drive shaft stiffness are constructed for the observer design. State variables are selected as x = ω t ω t, x 2 = ω ω, x 3 = ω w ω w, x 4 = T s T s, x 5 = p cb p cb, so that the variables are normalized to have the same level of magnitude. The driveline motion of the upshift torque phase is then expressed in the following state space form: ẋ = c 4 T s ω t i df x 4 + c 3μ( ω t x )RNA p cb ω t x 5 + ω t f t (ω e,x ), (2.64a) ẋ 3 = c 34 T s ω w x 4 + ω w f t2 (x 3 ), (2.64b) ẋ 4 = K s ω t i df i T s x K s ω w T s x 3, (2.64c) ẋ 5 = x 5 + K cv u, τ cv τ cv p cb (2.64d)

2.4 Clutch Pressure Estimation when Considering Drive Shaft Stiffness 63 where u = i b is the control input and f t (ω e,x ) = c T t c 3 μ( ω)rnf s, f t2 (x 3 ) = c 35 T l. (2.65a) (2.65b) Similarly, the inertia phase can also be described in the following state space form with state variables of x to x 5 : ẋ = C 4 T s ω t i df x 4 + C 3μ(x 2 )RNA p cb ω t x 5 + ω t f (ω e,x ), (2.66a) ẋ 2 = (C 4 C 24 ) T s ωi df ẋ 3 = C 34 T s ω w x 4 + ω w f 3 (x 3 ), x 4 + C 3 C 23 μ(x 2 )RNA p cb x 5 + ω ω f 2(ω e,x ), (2.66b) (2.66c) ẋ 4 = K s ω t K s ω x x 2 K s ω w x 3, (2.66d) i df T s i df ( + γ) T s T s ẋ 5 = x 5 + K cv u, τ cv τ cv p cb with the nonlinear functions (2.66e) f (ω e,x ) = C T t C 3 μ( ω)rnf s, f 2 (ω e,x ) = (C C 2 )T t (C 3 C 23 )μ( ω)rnf s, f 3 (x 3 ) = C 35 T l. (2.67a) (2.67b) (2.67c) The problem considered here is to estimate the pressure of clutch B x 4 (drive shaft torque x 5, too) both in the torque and inertia phases, in the presence of model errors, given the measured rotational speeds of transmission x,x 2,x 3,ω e and valve electric current u. 2.4.2 Design of Reduced-Order Nonlinear State Observer Reduced-Order Nonlinear Observer Denote the variable to be estimated as z, and rewrite the dynamics of the system for estimating the clutch pressure as follows: ẏ = F (y,u)+ G(y,u)z + Hw(y,u,z), ż = A 2 y + A 22 z + B 2 (u), (2.68a) (2.68b)

64 2 Pressure Estimation of a Wet Clutch where y is the measured outputs, w(y,u,z) summarizes model uncertainties which is normalized by H as w, and in particular y =[x,x 3 ] T, z =[x 4,x 5 ] T, u= i b, (2.69a) ( ) F (y,u)= ωt f t (ω e,y ), (2.69b) ω w f t2 (y 2 ) c 4 T s c 3 μ(x 2 )RNA p cb G(y,u)= ω t R df ω t, (2.69c) c 34 T s ω w 0 ( Ks ω t R df R T s ) K s ω w A 2 = T s, (2.69d) 0 0 ( ) 0 0 A 22 = 0 τ, (2.69e) cv ( ) 0 B 2 (u) = K cv u. (2.69f) τ cv p cb for the torque phase. For the inertia phase, y =[x,x 2,x 3 ] T is the measurement. Hence, (2.69b) (2.69e) are replaced by F (y,u)= G(y,u)= A 2 = ω t f (ω e,y ) ω f 2(ω e,y ) ω w f 3 (y 3 ) C 4 T s ω t R df (C 4 C 24 ) T s ωr df, C 3 μ(y 2 )RNA p cb ω t C 3 C 23 μ(y 2 )RNA p cb ω C 34 T s ω w 0 ( Ks ω t K s ω R df T s R df (+γ) T s 0 0 0 ( 0 0 A 22 = 0 τ cv The observer is then designed in the form of K s ω w T s, (2.70a) (2.70b) ), (2.70c) ). (2.70d) ẑ = A 2 y + A 22 ẑ + B 2 (u) + L ( ẏ F (y,u) G(y,u)ẑ ), (2.7) where L R 2 2 (L R 2 3 for the inertia phase) is the constant observer gain to be determined [4].

2.4 Clutch Pressure Estimation when Considering Drive Shaft Stiffness 65 In order to avoid taking derivatives of the measurements y, the following transformation is made. Let η =ẑ Ly, (2.72) then, we can infer for a time-invariant L that η = ( A 22 LG(y,u) ) (η + Ly) + A 2 y + B 2 (u) LF (y,u). (2.73) Equations (2.72) and (2.73) constitute then the reduced-order observer of the clutch pressure for the nonlinear driveline system. Obviously, the nonlinearities of the powertrain system appear in the observer in their original form. Therefore, the characteristics of powertrain mechanical systems, such as characteristics of the engine and the aerodynamic drag, are represented in the form of lookup tables, which is easily processed in computer control. Then the error dynamics of the designed shaft torque observer is analyzed using the concept of ISS (input-to-state stability) [, 2, 8]. By defining the observer error as e = z ẑ, (2.74) the error dynamics can then be described by ė = ( A 22 LG(y,u) ) e LH w. (2.75) We define V(e) = 2 et e and differentiate it along the solution of (2.75) to obtain and then V = e T ( A 22 LG(y,u) ) e e T LH w, (2.76) V e T ( A 22 LG(y,u)+ κ I ) e + w T H T L T LH w, (2.77) 4κ where κ > 0. We now choose L to satisfy the following matrix inequality: with κ 2 > 0, then we arrive at A 22 LG(y,u) (κ + κ 2 )I (2.78) V κ 2 e T e + w T H T L T LH w, (2.79) 4κ which implies that the error dynamics admits the input-to-state stability property if the model error w is supposed to be bounded in amplitude. It follows from (2.79) that e(t) 2 e(0) 2 e 2κ 2t + w 2 sup [0,t] λ max (H T L T LH ) 2κ t 0 e 2κ 2(t τ) dτ, (2.80)

66 2 Pressure Estimation of a Wet Clutch which implies that e(t) 2 w 2 sup(λ max(h T L T LH )) 4κ κ 2 as t. (2.8) For a more detailed deduction, please refer to Sect. 2.3. Gain Determination Now we discuss how to choose parameters κ, κ 2, and finally, the observer gain L. κ and κ 2 It follows from (2.79) that κ 2 can be chosen according to the required decay rate of the error. If it is desired that the error converges in 0.05 s, then 4 2κ 2 = 0.05, which results in κ 2 = 40. According to (2.8), one may choose a larger κ to reduce the offset. From (2.78), however, one should notice that the larger the κ, the higher the observer gain. Optimization of L We now give a solution of (2.78) for constant L through solving a set of linear matrix inequalities (LMIs). If A 22 (u) and G(y,u) in (2.78) vary in a polytope with r vertices, i.e., ( A22 G(y,u) ) Co {( A 22, G ), ( A22,2 G 2 ),..., ( A22,r G r )}, (2.82) where Co{ } denotes the convex hull of the polytope. Then, a constant observer gain L satisfying the following Linear Matrix Inequalities (LMIs): A 22,i LG i (κ + κ 2 )I, i =, 2,...,r, (2.83) meets the observer gain condition (2.78). Moreover, we prefer low observer gains, due to robustness against noises and also the reduction of the estimation error offset estimated as (2.8). Hence, L can be obtained through the following optimization: min trace(α) α,l subject to LMIs (2.83) and ( α LH H T L T I ) 0, (2.84) where α is a 2 2 positive diagonal matrix for both the torque and inertia phase. Given κ and κ 2, the solution of (2.84) gives then the lowest possible gains. Solution and Evaluation To calculate the observer gain and the error offset, the bound of the modeling error should be calculated first. It is indeed difficult, if not impossible, to obtain a comprehensive estimate of the modeling error bound. Hence some major uncertainties are taken into consideration to estimate the value of the modeling errors. If the estimation error of the turbine torque T t is bounded within

2.4 Clutch Pressure Estimation when Considering Drive Shaft Stiffness 67 5 %, the variation of vehicle mass is ±500 kg and the variation of road slope is ±5, the modeling error for the torque phase can be calculated as ( w = (2.85) ) with the normalization matrix ( ) 2.38 0 H =. (2.86) 0 0.3 Given the above modeling error bound and the system parameters (shown in Table 2.4), (2.84) and (2.8) can be used to calculate L and then check the error offset under the calculated gain. The final tuned results for the torque phase are κ = 5 and ( ) 2.08 30.49 L =, (2.87) 6.46 52.48 and the calculated error offset is e( ) 0.5, which means that the error offset of the clutch pressure p cb is not larger than 0.05 MPa, which is considered acceptable [4]. Similarly, following the procedure given above, the observer gain for the inertia phase can also be calculated, and the result reads ( ) 0.023 0.22 56.9 L =. (2.88).74 0.23.73 2.4.3 Simulation Results Besides the continuous simulation, discrete implementations are carried out as well to get an in-vehicle assessment of the proposed observer. The sampling rate is chosen to be 00 Hz in order to test the feasibility of implementing the resulting observer for real applications [9]. The discrete characteristics and random noise of the speed sensor are included as well. The major concern is put on the power-on st-to-2nd gear upshift process. Figure 2.0 gives the simulation results of the shift process with the nominal driving condition, i.e., the condition for the observer design. The continuous and discrete results are listed simultaneously. During the shift process, the engine throttle angle is adjusted to cooperate with the transmission shift. It can be seen that during the torque phase (between 7.74 and 7.94 s) the rotational speeds of the shafts do not change much, whereas during the inertia phase (between 7.94 and 8.24 s), the rotational speeds change extensively because of the clutch slip. In the torque and inertia phase, the pressure of cylinder B is estimated by the designed observers. After the inertia phase, i.e., after 8.24 s, because of the engagement of the clutch, the observer is not valid any more, and the pressure is estimated