You are on page 1of 8

446

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 18, NO. 2, MARCH 2010

A Reduced-Order Nonlinear Clutch Pressure Observer for Automatic Transmission


Bingzhao Gao, Hong Chen, Member, IEEE, Haiyan Zhao, and Kazushi Sanada

AbstractFor a novel type of automatic transmissions adopting clutch-to-clutch shift control technology with electro-hydraulic actuators, a clutch pressure observer method based on input-to-state stability (ISS) is proposed. Model uncertainties including steady state error and unmodelled dynamics are considered as additional disturbance inputs and the observer is designed in order that the error dynamics is input-to-state stable. Lookup tables, which are widely used to represent complex nonlinear characteristics of engine systems, appear in their original form in the designed reduced-order observer. The designed pressure observer is tested on an AMESim powertrain simulation model. Comparing with the sliding mode method, the designed pressure observer has the better performance. Index TermsAutomatic transmission (AT), clutch pressure estimation, input-to-state stability (ISS), nonlinearity, reduced-order observer.

I. INTRODUCTION

N RECENT years, in order to improve fuel economy, reduce emission, and enhance driving performance, many new technologies have been introduced in the transmission area, such as dual clutch transmission (DCT) and new automatic transmission (AT) controlling clutches independently [2]. Furthermore, smart proportional valves with large ow 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. In both DCTs and new ATs, the change of the speed ratio is regarded as the process of one clutch being engaged while the other being disengaged, namely, clutch-to-clutch shift. For the vehicles with hydraulic cylinder as clutch actuator, which is ubiquitous in the present transmissions, the cylinder pressure
Manuscript received March 27, 2008; revised March 03, 2009. Manuscript received in nal form May 31, 2009. First published August 11, 2009; current version published February 24, 2010. Recommended by Associate Editor F. Vasca. This work was supported in part by the National Science Fund of China for Distinguished Young Scholars under Grant 60725311 and by the National Nature Science Foundation China (90820302). This brief appeared in part at the IEEE-CDC, Cancun, Mexico, 2008. B.-Z. Gao is with the National Automobile Dynamic Simulation Laboratory, Jilin University, Changchun 130025, PR China, and also with the Department of Mechanical Engineering, Yokohama National University, Yokohama 240-8501, Japan (e-mail: gaobingzhao@hotmail.com). H. Chen is with the Department of Control Science and Engineering, Jilin University (Campus NanLing), Changchun 130025, PR China, and also with the National Automobile Dynamic Simulation Laboratory, 130025 Changchun, PR China, (e-mail: chenh@jlu.edu.cn; chenh98cn@hotmail.com). H. Zhao is with the Department of Control Science and Engineering, Jilin University (Campus NanLing), Changchun 130025, PR China. K. Sanada is with the Department of Mechanical Engineering, Yokohama National University, Yokohama 240-8501, Japan (e-mail: ksanada@ynu.ac.jp). Color versions of one or more of the gures in this paper are available online at http://ieeexplore.ieee.org. Digital Object Identier 10.1109/TCST.2009.2024758

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 [4]. Because of the complex nonlinearities in 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 large numbers of experiments in the steady state, are widely used to describe the nonlinear characteristics. There inherently exist model uncertainties, such as steady-state error and unmodelled 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. There have been some studies on the estimation of the transmission shaft torque and the clutch pressure. The sliding mode observer in [5] is designed to estimate the torque of an automotive driveshaft in [6]. The adaptive sliding mode algorithm is proposed to estimate the turbine torque of a torque converter in [7]. Furthermore, [4] uses the sliding mode method to estimate the clutch pressure in a hydraulically powered stepped AT. The extended algorithm in [8] is used to estimate the clutch pressure and the transmission output shaft torque simultaneously. In [9] and [10], a neural network is suggested to estimate the turbine torque, in which the engine speed, the turbine speed, and the oil temperature are inputs. [10] also designs a driving load observer by assuming that the driving load is slowly-varying. In [11], recursive least square method with multiple forgetting factors is used to estimate the road grade and the vehicle mass. In [12], a full-order observer is proposed for the pressure monitoring of a torque converters 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. 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 paper, in which electro-hydraulic actuators are adopted to control the clutches. A reduced-order clutch pressure observer based on the concept of input-to-state stability (ISS) [13], [14] is proposed, where the rotational speeds are the measured outputs and the special structure of the clutch pressure system is exploited. Model uncertainties are considered as additional disturbance inputs. Lookup tables of the nonlinear characteristics of powertrain systems appear in their original form. A systematic procedure is given to design the nonlinear clutch pressure observer such that it follows. The error dynamics is ISS, where modelling errors are as the inputs. This means that the initial estimation error de-

1063-6536/$26.00 2009 IEEE


Authorized licensd use limted to: IE Xplore. Downlade on May 13,20 at 1:407 UTC from IE Xplore. Restricon aply.

GAO et al.: REDUCED-ORDER NONLINEAR CLUTCH PRESSURE OBSERVER FOR AUTOMATIC TRANSMISSION

447

is transferred from clutch A to clutch B and the inertia phase where clutch B is synchronized [15]. By selecting the turbine speed , the speed difference of clutch B , and the pressure of cylinder B as state variables , respectively, the inertia phase of the rst to second gear up shift process is described in the following state space form: (1a) (1b) (1c) with

(2a)
Fig. 1. Schematic graph of automatic transmission.

cays exponentially and the estimation error is guaranteed to be bounded for the bounded modelling 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 benets from the reduced order and the time-invariant gains of the observer. Lower observer gains are obtained through convex optimization, which increases the robustness against noises and reduces the estimated upper bound of the error offset. The rest of this paper is organized as follows. In Section II, a dynamic model of the considered transmission is derived for the shift process. In Section III, the proposed reduced-order clutch pressure observer is described and the property of the observer is investigated in the concept of ISS. Based on the theoretic analysis, a systematic procedure to design the pressure observer is then given, following by a design example. In Section IV, the proposed observer is tested on a complete powertrain simulation model, and comparison results with an existing sliding mode algorithm are given as well. II. CLUTCH SYSTEM MODELLING AND PROBLEM STATEMENT We consider the powertrain in passenger vehicles with a twospeed AT, as schematically shown in Fig. 1. A planetary gear set is adopted as the shift gear. Two clutches are used as the actuators. Two proportional pressure valves are used to control the two clutches respectively. When clutch A is engaged and clutch B disengaged, the powertrain operates on the 1st gear , where is the and the speed ratio is given out by 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 on the second gear with a speed ratio of During the shift process, the oncoming and offgoing clutches are controlled by the two valves through a separate controller, which is assumed to be well-designed. The power-on rst to second up shift is considered here as an example. The gear shift process is divided into the torque phase where the turbine torque
Authorized licensd use limted to: IE Xplore. Downlade on May 13,20 at 1:407 UTC from IE Xplore. Restricon aply.

(2b) is the current of valve B; is the turbine torque where is the equivalent resistant torque delivered from [16], [17]; is the constant coefcients dethe tire to the drive shaft; termined by the inertia moments of vehicle and transmission denotes the return spring force of clutch B. Other pashafts; rameters are dened in Table I and referred to [1] for the detailed modeling process. In order to estimate the pressure of clutch B, the rotational speeds of the transmission are used as the mea. sured outputs, i.e., Similarly the torque phase of the rst to second gear up shift process is described in the following state space form: (3a) (3b) with the measured output and

(4) are different coefcients from in (2). MoreNote that over, although there is no obvious change of clutch speed during is difthe torque phase, the speed difference of the clutch is also considferent for various driving maneuvers. Hence, ered as an input for the torque phase model. Due to the extreme complexity of the torque converter and the aerodynamic drag, the nonlinear functions in (2) and (4) are generally given as lookup tables, which are obtained by a series of steady-state experimentations and contain inherently errors. Other modelling uncertainties include variations of parameters, such as the vehicle mass, the road grade and the damping coefcient of shafts. Hence, the problem considered in this paper is to estimate the pressure of clutch B in the presence of model errors, given the valve electric current , and the measured rotational speeds of the transmission , , and .

448

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 18, NO. 2, MARCH 2010

III. REDUCED-ORDER NONLINEAR STATE OBSERVER A. 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 , and rewrite the dynamics system for estimating the clutch pressure as follows: (5a) (5b) where is the measured output, uncertainties and in particular summarizes model

Upon

multiplication

of

(13)

by

, it becomes . Integrating it over

leads to

(6a) (6b) for the inertia phase. The expressions for the torque phase is omitted. Because the shaft torque affects the related shaft accelerations directly, the difference between the true accelerations and the is used to constitute the estimated values correction term. The observer is then designed in the form of (7) (or ) is the observer gain to be where and the error determined. Dene the observer error as dynamics is then described by (8) Let to obtain and differentiate it along the solution of (8) (9) By the use of Youngs Inequality [14], the above equality becomes (10) where . Choose to satisfy the following inequality: (11) with , then we arrive at (12) Replacing leads to (13)
Authorized licensd use limted to: IE Xplore. Downlade on May 13,20 at 1:407 UTC from IE Xplore. Restricon aply.

(14) Hence, the properties of the error dynamics of the designed observer (7) are described as follows. Theorem 1: Suppose the following: , ; the observer gain is chosen to satisfy (11). Then, the error dynamics of the observer (7) is as follows: (i) input-to-state stable, if is bounded in amplitude, i.e., ; for . (ii) exponentially stable with Proof: It follows from (12) that which shows that the error dynamics admits the input-to-state stability property [14, p. 503] if the model error is supposed to be bounded in amplitude, as property (i) required. , we obtain from (14) that By taking which proves property (ii). Remark 3.1: Now we give some discussions on the parameters and . From property (ii), is chosen according to the required decay rate of the error. If is bounded in amplitude, , then (14) becomes i.e., (15) Hence, one may choose larger to reduce the offset. From (11), however, one should also notice that the larger the , the higher the observer gain. Remark 3.2: Equation (15) gives just an upper bound of the estimation error offset, if the bound of the model error is given. The real offset could be much smaller, due to the multiple use of inequalities in the above derivation. Remark 3.3: Besides satisfying (11), we do not impose other assumptions on the observer gain . This implies that can , while one may be designed theoretically to depend on choose it as time-invariant (constant) in practice. Solution of time-invariant will be discussed in the following Subsection III-B. B. Implementation Issues In order to avoid taking derivatives of the measured variables, let (16) then, for a time-invariant (17) Equations (16) and (17) constitute the reduced-order observer for the nonlinear clutch slip control system. 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 characteristics of the engine

GAO et al.: REDUCED-ORDER NONLINEAR CLUTCH PRESSURE OBSERVER FOR AUTOMATIC TRANSMISSION

449

TABLE I PARAMETERS FOR OBSERVER DESIGN

and torque converter, is represented in the form of lookup tables, which is easy to be processed in computer. According to Theorem 1 and Remark 3.1, a systematic procedure is given to design the reduced-order nonlinear clutch pressure observer in the form of (16) and (17) as follows: S1) choose the parameter according to the required decay rate of the estimation error; S2) choose the parameter , where it is suggested to start from some smaller values (according to Remark 3.1); S3) determine the observer gain such that (11) is satised; S4) for a given model error bound, use (15) to compute the estimated upper bound of the offset and check if the offset bound is acceptable; S5) if the offset bound is acceptable, end the design procedure. If not acceptable, go to S2. It is well known that getting model error bounds is in general very difcult, if not impossible. As mentioned in Remark 3.2, for a given model error bound, (15) gives just an upper bound of the estimation error offset, which might be much larger than the real offset. Hence, the stop of iterating S1-S5 is somehow a rule of thumb. We now give a solution of (11) for choosing to be time-invariant, where the requirement for low observer gains can be and in (11) considered through optimization. If vary in a polytope with vertices, i.e.,

and vary in a polyTheorem 2: Suppose that tope as (18). Then, any time-invariant satisfying the following linear matrix inequalities (LMIs): (21) meets the observer gain condition (11). Proof: The substitution of (19) into (11) leads to (22) Due to (20), the satisfaction of (21) guarantees (22) and hence (11). In (21), and are known and bounded, and are selected to bounded and (where is the number and ) is bounded, of time-varying parameters in too. Hence, some constant is always found to render it satisfying. Moreover, we prefer to low observer gains, for being robust against noises and reducing the upper bound of the error offset, which is estimated by (15). Hence, is obtained through the following convex optimization: subject to LMIs (21) and (23)

and , the solution of (23) gives then a constant obGiven server gain with the lowest possible gains satisfying the condition (11). C. Observer Design for Clutch Pressure

(18) where exist denotes the convex hull of the polytope. Then, there such that (19) with being functions of and satisfy (20) Hence, conclusions are given as follows.
Authorized licensd use limted to: IE Xplore. Downlade on May 13,20 at 1:407 UTC from IE Xplore. Restricon aply.

The inertia phase is taken as an example to show the detailed design procedure. The parameters are regarded as constants for simplicity, which are listed in Table I, together and with the other parameters. Nonlinear functions , are given as lookup tables in the observer. These parameters are derived from the nominal setting of an AMESim simulation model of the AT shown in Fig. 1, which will be discussed later in Section IV-A . Following the procedure given in Section III-B , is chosen to meet the requirement for the desired decay rate of the estimation error. It is desired that the error converges in 0.1 s, and the settling time as 4 time constants [18, p. 221], which implies and results in . is chosen with the purpose of achieving a smaller Then, offset of the estimation error. Start with and obtain

450

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 18, NO. 2, MARCH 2010

Fig. 2. Left: Engine torque characteristics with speed and throttle opening. Middle: Capacity factor and torque ratio of the torque converter. Right: Friction characteristics of clutch plates.

and 0.175 MPa (see the following for the detailed calculation). The offset bound is too large for real applications. According to S2 of the procedure given in Section III-B, we enlarge the value of , and nally the value . being used is We now solve the optimization problem (23) to obtain the is considlowest possible observer gain. Since ered as constant, the polytope in (18) is given by , where two vertices are computed by (6a) with . The solution reads . In order to check if the estimation offset is acceptable, we now compute roughly the bound of modeling errors. Since powertrain systems admit highly nonlinear, complex dynamics and various uncertainties, it is indeed difcult, if not impossible, to obtain a comprehensive estimation of the modeling error bound. Hence, some major uncertainties are taken as the examples to estimate the value of . The major uncertainties here are calculation errors of in (5), which contains the turbine torque and the vehicle driving load . The change of the vehicle mass in . From numerous simaffects also the coefcients ulations of different powertrain settings, a bound of is determined as 1600 rad/s . According to (15), an upper bound of the offset is obtained for the designed observer. The result is 0.05 MPa, which is less that 10% of the variation range of the working pressure of the valve and is acceptable. Similarly following the procedure given in Section III-B, the observer gain for the torque phase is calculated, and the result . is IV. SIMULATION RESULTS A. Powertrain Simulation Model The powertrain simulation model is established by commercial simulation software AMESim. Except for the simplied 2-speed transmission, the parameters used in this paper represent a typical front-wheel-drive mid-size passenger car equipped with a 2000cc 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, time-delay in control and time-varying parameters are also considered in the simulation model of the proportional valves [19], which are neglected in the observer design.
Authorized licensd use limted to: IE Xplore. Downlade on May 13,20 at 1:407 UTC from IE Xplore. Restricon aply.

1) Engine: The work reported in this paper is primarily concerned with shift transients and therefore a simple engine model is used. The dynamic equation of the engine speed is represented by (24) where is the engine output torque and is the output torque of the converter pump. The engine output torque is simplied as and the engine a lookup table of the engine rotational speed , showing in the left plot throttle angle , i.e., of Fig. 2. 2) Torque Converter: When the turbine is driven forward, the dynamics of the torque converter are often characterized as and [7], where the capacity factor and the torque ratio are given in the middle plot of Fig. 2. 3) Planetary Gear Set: Using the submodel provided by 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 ; the inertia moment of the ring gear ; the teeth number of the sun gear and the teeth number of the ring gear . 4) Differential Gear Box and Drive Shaft: The gear ratio of the differential gear box is denoted as . The two drive shafts between the differential gear and the front wheels are represented as a torsion spring with stiffness coefcient and a torsion damping with damping coefcient . 5) Tires: The longitudinal tire force , which is usually simplied as a function of the longitudinal slip ratio , rises increases under a threshold and declines slowly fast when after that [20]. Here it is represented approximately as a function of . The longitudinal slip is calculated as , where is the wheel rotary velocity and is the car body velocity. 6) Road Loads: The road load consists of three parts: the grade force , the rolling resistant moment of tires and the aerodynamics drag . The resistant moment of tires is regarded as constant here. The grade force can be calculated , where is the vehicle mass, is the as grade angle of the road. The aerodynamics drag is described as [20]. 7) Clutches and Valves: The friction coefcient is a nonshown in the right plot of Fig. 2. In the linear function of design of the pressure observer in Section III-C , we assume the

GAO et al.: REDUCED-ORDER NONLINEAR CLUTCH PRESSURE OBSERVER FOR AUTOMATIC TRANSMISSION

451

TABLE II VALUES OF SIMULATION MODEL PARAMETERS

parameters 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 [19]. Hence, the dynamics of the proportional valve in the powertrain simulation is given by (25) Finally, the values of the parameters used in the powertrain simulation are listed in Table II. Nonlinear functions , , , , and are given as lookup tables. B. 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 predesigned clutch slip controller to ensure a rapid and smooth shift process. In this study, the major concern is put on the power-on rst to second gear up shift process. Left plots in Fig. 3 give the simulation results of the shift process with the driving condition of Table II, 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 by the simplied control valve dynamics (1c). During the torque phase (between 7.7 and 7.94 s) the rotational speeds do not change greatly, 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. 3 as the solid line, where the result for is also given as a comparison. The performance is improved. The error peak is reduced about 35% and the average error is reduced about 31%. Note that the shift process operates in the nominal driving condition, but the stiffness of the drive shaft and
Authorized licensd use limted to: IE Xplore. Downlade on May 13,20 at 1:407 UTC from IE Xplore. Restricon aply.

the tire slip are considered in the simulation model, while these are ignored in the model for designing the observer. Moreover, 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, the road grade, torque characteristics of the engine and the torque converter are varied. The result with the relatively large error is shown in the right of Fig. 3, where the driving condition setting is as follows: the torque characteristics of the engine is enlarged by 15%, and subsequently the capacity of the torque converter is also enlarged; the vehicle mass is increased from 1500 to 1725 kg, and the road grade angle is varied from 0 to 5 deg. 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 phase, and no large change of the transmission speeds for the large vehicle inertia. Therefore, the torsion of the drive shaft and the tires slip play important roles in the drive line. Omission of these terms in the observer design deteriorate 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. As a comparison, a full-order sliding mode observer is designed according to [4], [6], given as follows: (26a) (26b) (26c) for the inertia phase, and (27a) (27b) for the torque phase. The sampling frequency of the sliding mode observer is chosen to be 100 Hz, in order to test the feasibility of the resulting observer for real applications [10]. In the discrete implementation, the observer gains have to be

452

IEEE TRANSACTIONS ON CONTROL SYSTEMS TECHNOLOGY, VOL. 18, NO. 2, MARCH 2010

Fig. 3. Left: Results of the nominal driving condition. Right: Results of different driving condition.

TABLE III GAINS OF DISCRETE OBSERVERS

reduced in order to restrain oscillations resulting from sampling and two sets of the tuned values are given in Table III. Hence, the proposed ISS observer is also discretized by the same sampling frequency and the tuned gains are also listed in Table III. The comparison results of these three observers are shown in Fig. 4, where the driving condition is the same as that in the right of Fig. 3. In Fig. 4, the solid line represents the error of the reduced-order observer, while the dotted and dashed lines
Authorized licensd use limted to: IE Xplore. Downlade on May 13,20 at 1:407 UTC from IE Xplore. Restricon aply.

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 1) 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. In the viewpoint of robustness, the proposed observer achieves robustness in the sense of ISS, where the model errors are represented as external inputs.

GAO et al.: REDUCED-ORDER NONLINEAR CLUTCH PRESSURE OBSERVER FOR AUTOMATIC TRANSMISSION

453

Fig. 4. Comparison between ISS observer and sliding mode observer (torque . 1725 kg; converter capacity is enlarged by 15%;

m=

 =5 )

V. CONCLUSION A reduced-order observer of clutch pressure is proposed in the concept of ISS, for a new kind of AT adopting clutch-toclutch shift technology. The proposed observer uses the easily and precisely measured rotational speeds to constitute the correction term, while the lookup tables include the nonlinear characteristics of engine systems. Hence, it is easily implemented. Model uncertainties including steady state errors and unmodelled dynamics are considered as additive disturbance inputs and the observer is designed such that the error dynamics is input-to-state stable. The designed observer is tested on an AMESim powertrain simulation model, which contains the complete drivetrain and clutch actuators. Simulation results show that, in the inertia phase, the estimation error is restricted in the required bound even if the model error is large. Comparing with the sliding mode observer, the proposed ISS observer has the better estimation performance and ability in eliminating chatters. In the torque phase, however, the peak error becomes larger when the driving condition deviates from the nominal setting. This deciency might be improved by using more precise models for the observer design, such as a model in consideration of the torsion of the drive shaft. It is going to be one of our future works. REFERENCES
[1] B.-Z Gao, H. Chen, H.-Y Zhao, and K. Sanada, A reduced-order nonlinear clutch pressure observer for automatic transmission using ISS, in Proc. 47th IEEE Conf. Dec. Control, Cancun, Mexico, 2008, pp. 57125717.

[2] Z. Sun and K. Hebbale, Challenges and opportunities in automotive transmission control, in Proc. Amer. Control Conf., 2005, vol. 5, pp. 32843289. [3] H. Shioiri, Drivetrain, in Recent 10 Years of Automotive Engineering (in Japanese). Tokyo, Japan: Society of Automotive Engineers of Japan, 2007, pp. 134137. [4] S. Watechagit and K. Srinivasan, Implementation of on-line clutch pressure estimation for stepped automatic transmissions, in Proc. Amer. Control Conf., 2005, vol. 3, pp. 16071612. [5] E. Misawa and J. Hedrick, Nonlinear observersA state-of-the-art survey, ASME J. Dyn. Syst., Meas., Control, vol. 111, pp. 344352, 1989. [6] R. A. Masmoudi and K. Hedrick, Estimation of vehicle shaft torque using nonlinear observers, ASME J. Dyn. Syst., Meas., Control, vol. 114, pp. 394400, 1992. [7] K. Yi, B. K. Shin, and K. I. Lee, Estimation of turbine torque of automatic transmissions using nonlinear observers, ASME J. Dyn. Syst., Meas., Control, vol. 122, pp. 276283, 2000. [8] S. Watechagit and K. Srinivasan, On-line estimation of operating variables for stepped automatic transmissions, in Proc. IEEE Conf. Control Appl. (CCA 2003), Istanbul, Turkey, 2003, vol. 1, pp. 279284. [9] B. K. Shin, J. O. Hahn, and K. I. Lee, Development of shift control algorithm using estimated turbine torque, SAE, Warrendale, PA, Tech. Paper 2000-01-1150, 2000. [10] J. O. Hahn and K. I. Lee, Nonlinear robust control of torque converter clutch slip system for passenger vehicles using advanced torque estimation algorithms, Veh. Syst. Dyn., vol. 37, no. 3, pp. 175192, 2002. [11] A. Vahidi, A. Stefanopoulou, and H. Peng, Recursive least squares with forgetting for online estimation of vehicle mass and road grade: Theory and experiments, Veh. Syst. Dyn., vol. 43, no. 1, pp. 3155, 2005. [12] J. O. Hahn, J. W. Hur, Y. M. Cho, and K. I. Lee, Robust observerbased monitoring of a hydraulic actuator in a vehicle power transmission control system, Control Eng. Practice, vol. 10, no. 3, pp. 327335, 2002. [13] E. D. Sontag, Input to state stability: Basic concepts and results, in Lecture Notes in Mathematics. Berlin, Germany: Springer-Verlag, 2005. [14] M. Krstic, I. Kanellakopoulos, and P. Kokotovic, Nonlinear and Adaptive Control Design. New York: Wiley-Interscience, 1995. [15] M. Goetz, M. C. Levesley, and D. A. Crolla, Dynamics and control of gearshifts on twin-clutch transmissions, Proc. Inst. Mech. Eng., Pt. D: J. Auto. Eng., vol. 219, no. 8, pp. 951963, 2005. [16] H. Xia and P. Oh, A dynamic model for automotive torque converters, Int. J. Veh. Des., vol. 21, no. 45, pp. 344354, 1999. [17] H. S. Jo, W. S. Lim, Y. I. Park, and J. M. Lee, A study of the shifting transients of a torque converter equipped vehicle, Int. J. Heavy Veh. Syst., vol. 7, no. 4, pp. 264280, 2000. [18] K. Ogata, Modern Control Engineering, 4th, Ed. Englewood Cliffs, NJ: Prentice-Hall, 2001. [19] K. Sanada and A. Kitagawa, A study of two-degree-of-freedom control of rotating speed in an automatic transmission, considering modeling errors of a hydraulic system, Control Eng. Practice, vol. 6, pp. 11251132, 1998. [20] T. D. Gillespie, Fundamentals of Vehicle Dynamics. New York: Society of Automotive Engineers, 1992.

Authorized licensd use limted to: IE Xplore. Downlade on May 13,20 at 1:407 UTC from IE Xplore. Restricon aply.

You might also like