Experimental Implementation of an Electromagnetic Engine Valve by Tushar Anil Parlikar B.S., Swarthmore College (2001) B.A., Swarthmore College (2001) Submitted to the Department of Electrical Engineering and Computer Science in partial fulfillment of the requirements for the degree of Master of Science at the MASSACHUSETTS INSTITUTE OF TECHNOLOGY February 2003 © Massachusetts Institute of Technology, MMIII. All rights reserved. Author___ Department of Electrical Engineering and Computer Science January 31, 2003 Certified by. John G. Kassakian Professor of Electrical Engineering and Computer Science Thesis Supervisor Certified by ,;Thomas A. Keim Principal Research Scientist, Laboratory for ElecyromagnWic and Electronic Systems T e is -su-wvisor Accepted by Arthur C. Smith Chairman, Department Committee on Graduate Students MASSACHUSETTS INSTITUTE OF TECHNOLOGY MAY 12 2003 LIBRARIES BARKER -- I Experimental Implementation of an Electromagnetic Engine Valve by Tushar Anil Parlikar Submitted to the Department of Electrical Engineering and Computer Science on January 31, 2003, in partial fulfillment of the requirements for the degree of Master of Science Abstract A novel electromagnetic valve drive system (EMVD) for internal combustion engines was proposed by members of MIT's Laboratory for Electromagnetic and Electronic Systems in September 2001. Modeling and simulation results showed significant advantages of their EMVD over previously designed valve drives. The objective of this research was to evaluate the technical feasibility of the proposed EMVD. An experimental EMVD apparatus was designed, mathematically modeled and constructed. The apparatus was integrated into a computer-controlled experimental test stand, and preliminary experiments to characterize the EMVD were performed. The performance of the EMVD in the laboratory was comparable to that in simulations. The results obtained showed that the novel EMVD system is very promising technology. Thesis Supervisor: John G. Kassakian Title: Professor of Electrical Engineering and Computer Science Thesis Supervisor: Thomas A. Keim Title: Principal Research Scientist, Laboratory for Electromagnetic and Electronic Systems Acknowledgements I would like to thank Professor John Kassakian and Dr. Thomas Keim, my thesis supervisors, for their patience, guidance, and support during the course of this project. Woo Sok, Yihui, and Michael, the other project team members, and Wayne Ryan, the laboratory Engineering Specialist, were also key contributors to this work. Dr. David Turner of the Eaton Corporation, Dr. John Miller, formerly of the Ford Motor Corporation, Dr. Bruno Lesquene of Delphi Automotive Inc., the MIT Central Machine Shop, and Mr. Andrew Dunlap of dSPACE Inc. were instrumental at various stages of our research. Several members (and affiliates) of the Laboratory for Electronic and Electromagnetic Systems were invaluable while I conducted experiments, and wrote this report - Ale, Babak, Dave N., Dave P., Dave (and Sarah) W., Ernst, Frankie, Ivan (and Marina), John, Josh, Juan, Karin, Kiyomi, Lodewyk, Rob, Ross, Sandip, Tim, and Vivian - thanks for wonderful times and great memories. I am very grateful to my friends (NA, MS, VD, MR, LJM, TMW, AH, BD, and JW to name a few) for always being there for me. Above all, I owe my deepest gratitude to my family (my mother, my brothers Rajeev and Sanjeev, my sister-in-law Urmila, my in-laws Bruce and Barbara, and my dear wife Beth) without whom I would not be where I am today. -5- Contents 1 2 1.1 Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11 1.2 T hesis G oals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 1.3 Organization of this Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . 13 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.2 Conventional Engine Valve Trains . . . . . . . . . . . . . . . . . . . . . . . 15 2.3 Variable Valve Timing Systems . . . . . . . . . . . . . . . . . . . . . . . . . 16 2.4 Normal-Force Actuated Electromagnetic Valve Drive Systems . . . . . . . . 17 . . . . 19 Control Challenges for Electromagnetic Valve Drive Systems 21 The MIT EMVD 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 21 3.2 The Concept of the MIT EMVD . . . . . . . . . . . . . . . . . . . . . . . 21 3.3 System Modeling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22 3.4 3.5 4 15 Background 2.4.1 3 11 Introduction 3.3.1 The Nonlinear Mechanical Transformer . . . . . . . . . . . . . . . 22 3.3.2 Equations of Motion . . . . . . . . . . . . . . . . . . . . . . . . . . 23 . . . . . . . . . . . . . . . . . . . . 26 3.4.1 Control Constraints for the MIT EMVD . . . . . . . . . . . . . . . 27 3.4.2 Possible Controllers for the EMVD . . . . . . . . . . . . . . . . . . 28 . . . . . . . . . . . . . . . . 29 The Feedback-Controlled MIT EMVD Feedback-Controlled MIT EMVD Simulation The Experimental EMVD Test Stand 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . -7- 33 33 Contents 4.2 Preliminary Design of the Experimental Test Stand . . . . . . . . . . . . . . 33 4.3 Selection of Components for the Test Stand . . . . . . . . . . . . . . . . . . 35 4.4 Mechanical Components for EMVD Apparatus . . . . . . . . . . . . . . . . 37 4.4.1 General Structure of the EMVD Apparatus . . . . . . . . . . . . . . 37 4.4.2 Mechanical Component Design and 3-D Solid Modeling . . . . . . . 39 4.4.3 The Disk Cam - NTF . . . . . . . . . . . . . . . . . . . . . . . . . . 42 4.4.4 Construction and Assembly . . . . . . . . . . . . . . . . . . . . . . . 46 4.4.5 Mounting the Linear - z domain - Position Sensor . . . . . . . . . . 52 4.4.6 Adjusting the Valve Seat . . . . . . . . . . . . . . . . . . . . . . . . 54 The Experimental Test-Stand . . . . . . . . . . . . . . . . . . . . . . . . . . 55 4.5 5 6 The Motor and The Motor Drive 59 5.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 5.2 Design and Construction of the Motor Drive . . . . . . . . . . . . . . . . . . 59 5.3 Experiments with the Motor Drive . . . . . . . . . . . . . . . . . . . . . . . 65 5.3.1 Testing the Motor Drive Inverter Circuit . . . . . . . . . . . . . . . . 65 5.3.2 Characterization of the Motor Drive Inverter Circuit . . . . . . . . . 66 5.4 Modeling the dc Motor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69 5.5 The Dynamometer Apparatus . . . . . . . . . . . . . . . . . . . . . . . . . . 71 5.6 Experiments to Obtain Motor Parameters . . . . . . . . . . . . . . . . . . . 72 5.6.1 Dynamometer Tests . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 5.6.2 Transient Response Motor Tests . . . . . . . . . . . . . . . . . . . . 75 5.6.3 Inductance Measurements . . . . . . . . . . . . . . . . . . . . . . . . 80 5.6.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81 Controller Design and Experimental Results 83 6.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 83 6.2 Overview of Controller Design . . . . . . . . . . . . . . . . . . . . . . . . . . 83 6.3 Modeling the EMVD Plant 85 6.4 System Identification Experiments . . . . . . . . . . . . . . . . . . . . . . . . . . . -8- . . . . . . . . . . . . . . . . . . . . . . . 87 Contents . . . . . . . . . . . . . . . . . . . . . . 88 . . . . . . . . . . . . . . 90 6.5 Initial Mode Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 6.6 Holding Mode Control . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94 6.7 Reference Input Generation . . . . . . . . . . . . . . . . . . . . . . . . . . . 95 6.8 A Check on the NTF Characteristic Relation . . . . . . . . . . . . . . . . . 95 6.9 Transition Mode Controllers . . . . . . . . . . . . . . . . . . . . . . . . . . . 97 6.4.1 Free Oscillation Experiments 6.4.2 Open Loop Transfer Function Experiments 6.9.1 The Initial Attempt: a PD Compensator . . . . . . . . . . . . . . . 97 6.9.2 Lead Compensator Design . . . . . . . . . . . . . . . . . . . . . . . . 98 6.10 Linear Controller Implementation . . . . . . . . . . . . . . . . . . . . . . . . 103 6.10.1 PD Compensator . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 103 6.10.2 Lead Compensator . . . . . . . . . . . . . . . . . . . . . . . . . . . . 106 6.10.3 Valve Seating Velocity with the Lead Compensator . . . . . . . . . . 107 6.11 Robust Adaptive Controller Design . . . . . . . . . . . . . . . . . . . . . . . 108 . . . . . . . . . . . . . . . . . . . . . . . . . 108 6.11.2 Controller Stability . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111 . . . . . . . . . . . . . . . . . 112 6.11.4 Simulation Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . 113 6.11.5 Robust Adaptive Controller Implementation . . . . . . . . . . . . . . 115 6.11.1 Controller Development 6.11.3 Controller Implementation Algorithm 7 117 Conclusions and Future Work 7.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117 7.2 Evaluation of Thesis Objectives and Contributions . . . . . . . . . . . . . . 117 7.3 Future Work . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 118 A MATLAB Simulation of the EMVD in the 0 Domain 121 B Simulink Models for Experiments with the EMVD Apparatus 123 C Drawings of Parts for the Apparatus 135 -9 Contents D Drawings of Parts for the Dynamometer Test Stand 153 E dSPACE Models for Experiments with the EMVD 155 F Lab View File Used to Read Oscilloscope Data 159 G MATLAB Program for the Disk Cam Roller-Follower Profile 161 H Printed Circuit Board Schematics and Layout 165 I 167 Summary of Pacific Scientific 4N63 Data Sheet J Programs to Analyze the Motor and Motor Drive Tests 169 K MATLAB Simulation of the Adaptive Controller for the EMVD 181 L MATLAB Programs to Analyze EMVD Experimental Data 185 Bibliography 195 - 10 - Chapter 1 Introduction 1.1 Introduction A power requiretoday are moving towards higher systems electrical UTOMOTIVE increase in electrical load, the automotive industry estabresulting ments. Due to the lished a new 42V standard that will eventually replace the current 14V system [28]. The introduction of this standard, coupled with recent advances in power electronics, sensors and microprocessors, has led to several innovations in automotive systems. Many of these innovations significantly increase fuel economy, and some involve the replacement of automotive mechanical systems with electrical [1, 2]. Of particular relevance, the new voltage standard has made the electrification of internal combustion (IC) engine valves a technically and economically viable innovation. In conventional IC engines (see Fig. 1.1), engine valve displacements are fixed relative to the crankshaft position. The valves are actuated with cams that are located on a belt-driven camshaft, and the shape of these cams is determined by considering a tradeoff between engine speed, power, and torque requirements, as well as vehicle fuel consumption. This optimization results in an engine that is highly efficient only at certain velocities [3, 4]. If instead, the engine valves are actuated as a variable function of crankshaft angle, engine load, and other parameters, significant improvements in fuel economy - up to 20% - can be achieved [5]. In addition, improvements in torque and emissions are predicted [5]. IC engines in which both the duration (how long each valve is opened or closed) and the phase (when each valve is opened or closed) of the valves can be controlled are traditionally said to have variable valve timing (VVT) [6]. However, it is becoming increasingly common to refer to varying only the valve phase as VVT. Variable valve timing can be achieved using both mechanical and electromechanical actuation systems. In this thesis, the focus is on an electromechanical valve drive. With variable valve timing alone, a 10% improvement in fuel economy can be achieved [5]. Furthermore, if we control the lift (how much each valve is opened) of the valves, we can gain another 10% improvement. In the past year, a doctoral candidate at MIT - Woo Sok Chang, and Dr. Thomas Keim and - 11 - Introduction Camshaft Engine Valves Cylinder Crankshaf Figure 1.1 A section of a conventional internal combustion engine [7]. Prof. John Kassakian, members of MIT's Laboratory for Electronic and Electromagnetic Systems (LEES), have proposed an electromechanical valve drive (EMVD) incorporating a nonlinear mechanical transformer [3, 8, 9, 10]. Their proposal suggested significant advantages of their EMVD over previously designed actuation systems. The benefits range from lower average power consumption to a smaller seating velocity (velocity at which an engine valve engages its seat). 1.2 Thesis Goals The objective of this research was threefold: 1) To model the mechanical structure of the EMVD using 3-dimensional modeling software; 2) To construct this EMVD apparatus in the laboratory, and integrate it into a computer-controlled experimental test stand; and 3) To carry out experiments to verify the operation of the EMVD and compare experimental results to computer simulations and mathematical modeling. The objectives of this research were fulfilled. The EMVD apparatus that we have constructed was used to successfully prove the benefit of using a nonlinear mechanical transformer, as well as give us some powerful insights on how to optimize the system. During the course of this project, I worked closely with two doctoral candidates in the laboratory, Woo Sok Chang and Yihui Qiu, as well as an undergraduate, Michael Seeman'04. Without their collaboration, patience and guidance, this research would not have been possible. - 12 - 1.3 1.3 Organization of this Thesis Organization of this Thesis This document is organized as follows: in Chapter 2, we discuss previous EMVD designs and outline the challenges associated with these designs. In Chapter 3, the MIT EMVD is described in detail. Chapter 4 details the design and implementation of the experimental EMVD test stand. In Chapter 5, the design and construction of the motor drive inverter circuit, as well as the tests performed on the off-the-shelf motor we purchased are described. Chapter 6 describes the design and implementation of some controllers for the MIT EMVD, as well as experimental results obtained using these controllers. Finally, Chapter 7 concludes this report. Some of the text and figures in this document have been reprinted/adapted, with permission from the IEEE, from the paper "A New Electromagnetic Valve Actuator," @2002 IEEE (cited as reference [9] in this document). This paper was presented at the IEEE Workshop on Power Electronics in Transportation in October 2002. 13 - Chapter 2 Background 2.1 Introduction T systems. We on electromagnetic valve drive background some provides chapter HIS begin by describing conventional engine valve trains and normal-force actuated electromagnetic valve drive systems. In addition, the control challenges for normal-force actuated electromagnetic valve drive systems are discussed. 2.2 Conventional Engine Valve Trains An IC engine valve's kinematics profiles (such as valve position versus time, valve speed versus time, and so on) are of fixed shape and are timed relative to the engine crankshaft position. From a control systems perspective, we say the engine valves are not controllable. If instead, we could independently control the duration, phase and lift of the valves, a marked improvement in emissions, efficiency, maximum power, and fuel economy would be seen. The engine's mechanical design, although simple, compromises the efficiency and maximum power of the engine [6]. However, any variable valve actuation system must be able to offer a variable valve profile without compromising the essential characteristics of a conventional IC engine valve profile, which are described next. Let us examine the kinematics variables for a conventional IC engine valve, as shown in Fig. 2.1. In the figure, the valve stroke is defined as the displacement of the valve from fully-open to fully-closed positions [5]. Valve transition time is defined as the time taken for the valve to go from one end of its stroke to the other. The inertial power profile shown in Fig. 2.1 is obtained as the product of inertial force and valve velocity, and has an instantaneous peak value on the order of 2-3kW for each valve in a typical engine. The average power losses associated with driving the engine valves is approximately 3kW for 16 valves in a 2.OL, 4 cylinder engine at 6000rpm engine speed and wide open throttle [3]. There are a few important points to make about Fig. 2.1. First, although the valve inertial - 15 - Background Valve Closed Position Velocity Acceleration . .Inertial power Camshaft angles Figure 2.1 Conventional valve train profiles [8]. power is very large, it is also regenerative - after an initial input of inertial power, this inertial power is regenerated continuously. A spring is used to store the initial required energy and then the energy is transferred cyclically to the engine valve and cam. To be a competitive technology, any variable valve actuation system must be able to provide this large inertial power economically [3, 8]. Second, the seating velocity of the valve is small (less than 3cm/s at 600rpm engine speed, and less than 30cm/s at 6000rpm engine speed), which allows for the so-called soft landing of the valve. In order to prevent excessive wear of engine valves, any variable valve actuation system should allow for the soft landing of the valve. Third, an engine valve's kinematics profiles are inherently smooth. From a mechanical design perspective, discontinuities in valve kinematics profiles can generate undesirable impacts/losses and acoustical noise. 2.3 Variable Valve Timing Systems With an electromagnetically-driven variable valve timing system (VVT), one can independently control the phase and duration of the engine valve profiles, and carry out variable engine displacement (where certain cylinders in the engine are deactivated). In these VVT systems, the valve can be held in the open or closed positions for a variable time period (called the holding time), and it transitions from one end of the stroke to the other in the transition time. Prototype electromagnetically-actuated VVT systems have been proposed by several companies in the automotive industry, the first being proposed and patented by FEV Motorentechnik [11, 12]. Other companies that have worked on this technology include BMW [13, 14], GM [15], Renault [16, 17], Siemens [4, 18], and Aura [19]. - 16 - 2.4 Normal-Force Actuated Electromagnetic Valve Drive Systems Most electromagnetically-driven VVT systems have emulated one of the main characteristics of conventional IC valve profiles - that of regenerative inertial power. At the heart of these actuators is a valve-spring system, where an engine valve is coupled to two springs (with the same spring constant) as shown in Fig. 2.2. The equilibrium position for this massspring system is in the middle of the valve stroke [11]. Such a system has an inherent natural frequency (wn), mass (M), effective spring constant (k), and damping ratio ((). Assuming there was no damping, an initial displacement of the valve in the direction of either spring would result in sustained oscillations of the valve at the system's natural frequency (wn = ). Fixed Refereq<rame Spring Spring Dri'der (Attached to the VaIve Stem) -- pring Fixed Reference Fr.m Va1ve Figure 2.2 An engine valve-spring system. In the ideal frictionless case, considering only the dynamics of the valve, the electromechanical actuator for the valve-spring system only has to be able to hold the valve at either end of its stroke. In reality, due to gas forces in the engine, especially on the exhaust valves, additional work is required to reject the gas force disturbance. In addition, as the spring forces increase linearly with valve displacement, these forces are largest at the ends of the stroke, making it difficult to hold the valve in the open or closed position without using a large holding force, and thus a lot of electrical power [3]. 2.4 Normal-Force Actuated Electromagnetic Valve Drive Systems The most popular method of controlling the valve-spring system is to use two solenoids: one to hold the valve open and one to hold the valve closed [11, 12]. Fig. 2.3 shows a normal-force actuated valve-spring system [8]. Each electromagnet exerts a unidirectional normal force, and thus, the system employs two normal force actuators. The force exerted by these actuators is proportional to the square of the current input, but decreases as a function of the air gap between the actuator and the armature. Hence, these actuators - 17 - Background have a nonuniform force constant. For a fixed level of current, each solenoid can exert a large force when the valve is very close to the solenoid, but small forces when the valve is a short distance away from the solenoid. For example, when the valve is at the either end of the stroke, the relevant solenoid can produce a large force with a relatively small current. Thus, there is low holding current at both ends of the stroke. However, when the valve is at the lower end of its stroke, a large upward force requires a very large current in the upward-acting solenoid. Fixed Refere cc Frame pring Armature-... / normal Force Actuators 3Prng Fixed Reference F alve Figure 2.3 A normal-force actuated valve-spring system. Let us take a closer look at the free-flight dynamics for a normal-force actuated valvespring system without friction, and gravitational and gas forces, as shown in Fig. 2.4 [8]. The kinematics profiles in Fig. 2.4 can be easily explained. Suppose the valve is held closed by turning on the lower normal-force actuator. Ideally, if the valve were released, it would be accelerated by the springs past the equilibrium position of the system to its open position, where it would naturally stop. In reality, friction, and gravitational and gas forces prevent the valve from reaching the open position, and thus, near the open position, the second normal-force actuator is turned on, and the valve is pulled into its open position. Valve position Valve velocity Time Valve acceleration Figure 2.4 Kinematics profiles for a normal-force actuated valve-spring system [8]. - 18 - 2.4 2.4.1 Normal-Force Actuated Electromagnetic Valve Drive Systems Control Challenges for Electromagnetic Valve Drive Systems In the idealized motion described above, the springs play a large role because they provide the large inertial power to accelerate the valve at the beginning of its stroke, and then to absorb the inertial power to decelerate the valve at the ends of its stroke. As was the case in conventional IC engine valves, this inertial power is regenerative because energy is stored in the springs instead of being dissipated. In addition, due to the electromagnet-actuators' nonuniform force-displacement characteristics, the current required to hold the valve open or closed (holding current) is small [81. One of the other desirable characteristics for VVT systems is that of soft landing for the valves: the valves should reach either end of the stroke with very small velocity and acceleration. However, there are substantial control challenges to achieving soft landing with normal-force actuators. First, since the normal-force actuators are unidirectional, it is impossible to decelerate the valve as it approaches an end of its stroke - to arrive exactly at the end of the stroke with exactly zero velocity (defined as perfect soft landing), the receiving-end actuator must do exactly as much work as was done against friction and gas force over the entire transition. If the actuator does not do this much work, the valve will stop before the end of the stroke, and will be driven away again by the spring. If the actuator does any more than the exactly correct work, the valve arrives at the end of the stroke with non-zero velocity, and impacts the valve seat. A second control challenge is that the electromagnetic actuators have a nonuniform force constant, making it difficult to apply enough force to the valve when it is close to the equilibrium point of the system. Thus, it is difficult to counteract the effects of the gas force disturbance on the system. In the idealized free-flight valve-spring dynamics, as shown in Fig. 2.4, we can observe that the acceleration curve has discontinuities at both the end and beginning of the stroke. These discontinuities assume the instantaneous release of the valve at the beginning of the stroke and the instantaneous capture of the valve at the end of the stroke. These instantaneous actions require step changes in force. A true step force would create shock waves in the system and produce audible noise. To reduce this noise, it is possible to release the valves more slowly, but this lengthens transition time and increases the work which the capturing actuator must do. A possible solution to the control challenges in the normal-force actuated valve-spring system is to attempt to use a bi-directional shear force actuator (see [15, 20, 21]) to control the valve-spring system. An example of such an actuator is a rotary electric motor. Such actuators have uniform force constants and can exert bi-directional forces. Fig. 2.5 shows the - 19 - Background results from a simulation of a feedback-controlled valve-spring system with a rotary electric motor as the actuator [8]. As was the case before, the equilibrium position of the system is at the midpoint of the stroke. The reference input for this simulation was a smooth valve profile. 10 ......... ~.. 8 6 4 2 ............. ............. ............ ...........---~-...Valve position [mm] ........... .~ ~..... ....... ....... Valve velocity [m/sec] ............ .... ~.~... ~...... ............~ 0 -2 -4 -- ...... -6 -8 -10 0 Figure 2.5 system [8]. .. ~ ..... ~..... .~ ..- ....... 2 ..... 4 ~.... ........ Current [xlOOA] .... . ..................- .. -... ........ ......... .... Valve acceleration [km/sec2 ] ~ 6 ...... .......... ........... 10 8 12 Time [msec] Simulation of a feedback-controlled shear-force actuated engine valve-spring As expected, with active control and a bi-directional shear-force actuator, the effect of gas force is reduced by the controller, while the valve kinematics profiles are smooth - thereby eliminating the soft-landing problem inherent in normal-force actuated valve-spring systems. However, there are problems with this VVT system. First, the holding current is very high because the spring forces at the ends of the stroke are large. Second, the required driving current to follow smooth valve kinematics profiles is also large. Thus, the corresponding power loss of this VVT system is too large to be economically feasible [8]. - 20 - Chapter 3 The MIT EMVD 3.1 Introduction model for this system. MIT EMVD, and details a systemcontroller design issues. the introduces chapter HIS Simulations of the EMVD are discussed, as are preliminary The chapter concludes with a simulation result on which we based our component selection for the experimental EMVD apparatus. T 3.2 The Concept of the MIT EMVD In order to solve the problems associated with the previously discussed VVT systems, an EMVD incorporating a nonlinear mechanical transformer was proposed recently [8] by members of MIT's Laboratory for Electromagnetic and Electronic Systems. This EMVD comprises an electric motor that is coupled to a valve spring system with a nonlinear mechanical transformer (NTF) [3, 8, 9, 10]. Figure 3.1 shows a schematic of this EMVD and Fig. 3.2 shows a desirable nonlinear mechanical transformer characteristic between the z and 6 domains [3, 8, 9], where the valve stroke is 8 mm, corresponding to a rotational displacement of approximately 1 radian. In the proposed EMVD, the electric motor acts as a uniform-force-constant actuator, giving excellent control over valve position in the z domain [8, 9]. Using well-known active control techniques, small seating velocities, small position and velocity errors, and smooth kinematics variables can be achieved [8, 9]. In addition, the characteristic of the NTF can be designed such that the holding and driving currents in the system are reduced [8, 9]. 21 The MIT EMVD Electric NTF Pi..d Re~ferceXrarn Spring (Attaca t. V ve Stm Spring iier Si 1 -~1ring ]Fixed Reference IFre"n Figure 3.1 A schematic of the proposed EMVD. to. -. -30 -2'0 0 1 20 30 A desirable characteristic for the NTF. Figure 3.2 3.3 -1 System Modeling 3 -... .(.)(3.. =... .(.).... ......=. In this section, we derive the equations of motion for the MIT EMVD in both the 6 and z domains. The importance of the nonlinear mechanical transformer is then discussed. We conclude with simulations of the idealized (frictionless) MIT EMVD. 3.3.1 The Nonlinear Mechanical Transformer In Fig. 3.1, since 9 is a function of z and vice-versa, it is easy to show that the use of the nonlinear mechanical transformer implies that the following relations hold between 9 and z [3, 8]: - 22 - 3.3 dz- dzd dt dO dt d2 z d 2z cit2 2 dO' dO ( 2 (3.2) dzd d)+ System Modeling 2 0 (3.3) dO dt 2 The NTF provides a very desirable coupling (explained in detail below) between the z and 0 domains. By equating the energy in the z and 0 domains and using the NTF characteristic, the following relation results: (3.4) ro = dz do where 3.3.2 ro is torque in the 0 domain and fz is the force in the z domain. Equations of Motion The equations of motion for the proposed EMVD in Fig. 3.1 are as follows [8, 9]: d2 z mzd 2+ f= dz Bzd+ Kzz (3.5) 2 9d do JO d 2 + Bo- +o (3.6) = KTi where To is the transformer torque in the 0 domain, fz is the transformer force in the z domain, JO is the inertia in the 0 domain, m, is the mass in the z domain, BO is the friction in the 0 domain, Bz is the friction in the z domain, KT is the motor torque constant, Kz is the spring constant, 0 is the displacement in the rotational domain, and z is the displacement in the vertical domain. Equations (3.5) and (3.6) can be combined using the NTF characteristic relations (3.1), (3.2), (3.3), and (3.4). In this manner, we can obtain a single equation of motion in either the z or the 0 domains. This equation will be nonlinear, since linear equations such as (3.5) and (3.6) in one domain are transformed to nonlinear equations in the other domain because of the nonlinear nature of the transformer characteristic. Thus, by using (3.1), (3.2), (3.3), and (3.4), in (3.5), a nonlinear, time-varying, second-order differential equation of motion can be obtained in the 0 domain: JO 2 +Bo 0 -1Jod B9 Jo2 d2 0 +-t2BO dO dz + fz- = KTi do 2dt zd2+Bz + ymztj t± + -23- +Kzz (3.7) = KT dOA- (3.8) The MIT EMVD (dz 2) d20 T+o+ + B + B dz dz d2z 2 dO +B++ dz K f(0)d = Kji2. If we are to assume that a time-varying gas force disturbance also acts on the valve (typical of exhaust valves in an internal combustion engine), then the equation of motion becomes: ( dz d 2 d2 0 +dz +B dz d 2 z 2 +mzy 2 dO dz = Ki+g(t) (3.10) where g(t) is the gas force in the z domain reflected to the 0 domain. The system thus resembles a typical second-order differential equation with nonlinear mass/inertia and nonlinear damping. At either end of the stroke, the slope of the NTF characteristic, 4-, is very small. Thus, the reflected motor inertia in the z domain is very large, creating inherently smooth valve kinematics profiles, since the valve is slowed down by the large effective inertia when it is opened and closed. Moreover, the large spring force at the ends of the stroke when reflected to the 0 domain is small [9]. If this were not the case, the holding currents to keep the valve open or closed would be very large due to the large spring forces when the valve is moved away from the system's equilibrium position [9]. In effect, the NTF enables the use of small holding and driving currents when actuating the valve. In addition, because the gas force on the exhaust valve is largest at the opening end of the exhaust stroke, the reflected gas force in the 0 domain is also small. This characteristic makes it easy to open the valves against a large gas force. Thus, the proposed EMVD allows for the use of motors with small size. The benefit of using springs in the proposed EMVD is that they allow, ideally, for lossless low-power transitions of the valve from one end of the stroke to the other end. Once an initial amount of energy is injected into the system by compressing one of the springs, that energy is converted to kinetic energy and then transferred from that spring to the other spring continuously. If we assume no friction in (3.10), such that B, = 0 and Bo = 0, and we assume reasonable values for the other system parameters, as indicated in Table 3.1, the idealized "free-oscillation" response shown in Fig. 3.3 is obtained for the EMVD (see Appendix A for the MATLAB simulation program). This free-oscillation response emulates the idealized response of the normal-force actuated EMVD seen in Fig. 2.3. Although the MIT EMVD has profiles similar to the idealized normal-force actuated EMVD, these profiles come packaged with a system where one has more control over the valve's motion. In the non-ideal case where friction is included, it is - 24 - System Modeling 3.3 0.5 C 0 0 a- 0 0 0 0 0.01 0.005 0.015 0. 03 0.025 0.02 Time(s) 200 -. >2 (U -. -. . -.. 100 0 - -..-.-.-.-.-.- -100 -200 -300 0 Figure 3.3 0.02 0.015 Time(s) 0.01 0.005 0.025 0. 03 Idealized free oscillation of the EMVD in the 0 domain. easier to control the MIT EMVD than the normal-force actuated EMVD [8]. If larger K were assumed, the idealized free-oscillation response would show faster transitions from one end of the stroke to the other. We can also write state equations for the EMVD in the z domain. Denoting position and velocity in the z domain by x1 and x 2 respectively, the following nonlinear state equations are obtained in the z domain [8, 9]: (3.11) ; i = X2 (3.12) z2 = F1(x1, x 2 ) + F 2 (x 1 , x 2 )i + d(t) where: -{b F1(x1, x2) + Jo dx± )2 + = mz + F 2 (x1, x 2 ) = Kdx1} (3.13) JO(d)2 dO K T mz + J) (3.14) and d(t) is the time-varying gas force disturbance acting on the valve. Figure 3.4 shows the idealized free-oscillation response of the valve-spring system in the - 25 - The MIT EMVD Parameter Value Jo BO Kz mz Bz KT z 7.08 .10-6 kgm 2 7.64- 10-4 Nm/(rad/s) 2 .49328.7 N/m 0.09 kg 1.29 kg/s 0.069 Nm/A 0.004- sin(346) m/rad 9997r/)- ____________sin(O. Table 3.1 Parameters for simulation of free-oscillation response in the 0 domain. z domain, where the parameters used are listed in Table 3.2. The Simulink model used to generate this response appears in Fig. B.1 in Appendix B. By comparing Fig. 3.3 to Fig. 3.4, one can clearly see an effect of the NTF on the system dynamics - smooth valve kinematics profiles because the valve is slowed down by the large effective inertia at both ends of the stroke. Parameter Value JO BO Kz mz Bz KT 6.9- 10-6 kgm 2 1.16 - 10-5 Nm/(rad/s) 2 - 100000 N/m 0.09 kg 5 kg/s 0.07 Nm/A z 0.004 - sin(3.460) m/rad ____________sin(O.9997r/2) Table 3.2 3.4 Parameters for simulation of free-oscillation response in the z domain. The Feedback-Controlled MIT EMVD In this section, will look closely at the MIT EMVD as a feedback-controlled system. The constraints the controller must adhere to are discussed first, followed by discussions of the various controllers we could implement. We conclude the section with feedback-controlled simulations of the MIT EMVD, which take into account motor losses, gas force on the valve, and friction. - 26 - The Feedback-Controlled MIT EMVD 3.4 x 10-3 4 - - - - . 2 0 0 - ..... - -. . . -2 -4 ) 0.002 0.004 0.006 0.01 Time(s) 0.008 0.012 0.014 0.016 0.018 0.02 0.012 0.014 0.016 0.018 0.02 6 .. .. .. ...... >4 > 0 > . .. . .. -2 . . . .. . . .. .. -4 -0 0 0.002 Figure 3.4 3.4.1 0.004 0.006 0.01 Time(s) 0.008 Idealized free oscillation of the EMVD in the z domain. Control Constraints for the MIT EMVD A block diagram of the feedback-controlled EMVD apparatus is shown in Fig. 3.5. The reference input is the desired valve position, and the system output is the actual valve position. The difference between the two is passed into a controller which provides an appropriate current input to a motor drive. This motor drive supplies the desired current to the motor. Note that, for simplicity, this model assumes a perfectly responding motor drive which supplies as much current to the motor as desired, and assumes nothing about the dynamics of the motor drive. Ctor o ( NVaTve Vatveoe rInputl Figure 3.5 Cnrle r' (MWtor, NTF, PSsiton alve-Spring Systern) The EMVD as a feedback control system. , and g(t) in (3.10) are either known, Since the nominal values of JO, mz, KT, Bz, B6 , d, I design and implement various types to possible is it bounded, and known bounded, or both - 27 - The MIT EMVD of controllers for the EMVD to track the desired motor angular position trajectory Od(t). For a conventional internal combustion engine, the maximum valve transition times required at engine speeds of 6000rpm are approximately 3-4ms [6]. Based on these transition times, the required natural frequency of the valve-spring system in the MIT EMVD can be calculated to be approximately 150Hz, which is the reciprocal of twice the required valve transition time. Thus, the feedback-controlled EMVD must be able to respond to inputs with frequencies of approximately 150Hz [9]. Considering this fact, we decided on an overall control system bandwidth of approximately 1kHz. This bandwidth in turn constrains the bandwidth of the position sensors and the motor drive to be approximately 10kHz, as it is desirable for the sensor dynamics not to affect the feedback-controlled EMVD dynamics [9]. There are several issues that must be considered when designing controllers for the proposed EMVD. First, it is important to note that the dynamic characteristics of the proposed EMVD change along the valve stroke [9]. At the ends of the stroke, the effective inertia in the z domain is large, while at the midpoint of the stroke, this effective inertia is small. Thus, in the z domain, the effective system gain of the valve-spring system decreases at the ends of the stroke and increases at the midpoint of the stroke [9]. Second, it is important to be able to minimize errors when the valve is almost open or almost closed, such that the valve reaches these positions with small velocity. The errors as the valve transitions from one end of the stroke to the other are not as important. Third, the controller must be able to track the desired motor angular position trajectory even in the presence of parameter uncertainty and gas force disturbances. 3.4.2 Possible Controllers for the EMVD As we mentioned earlier, there are several possible control laws that could be used to actuate the proposed EMVD. A linear control law, such as a fixed-gain PD controller, is not well-suited to the control of this EMVD [9]. Fixed-gain controllers cannot account for the changing dynamic characteristics of the MIT EMVD during the valve stroke. For instance, at the ends of the stroke, the effective inertia in the z domain is large, while at the midpoint of the stroke, this effective inertia is small. Thus, the effective system gain of the valve-spring system decreases at the ends of the stroke and increases at the middle of the stroke [8, 9]. Nonetheless, on our first attempt at designing controllers (described in Chapter 6), we used linear controllers to control the EMVD apparatus. A linear control law can work well if the controller gains are varied with valve position - 28 - 3.5 Feedback-Controlled MIT EMVD Simulation [8]. In this respect, one control technique that can be implemented is that of piecewise linearization. In this technique, the state space is divided into sections where the nonlinear system is approximately linear and a different control law is used to govern the motion when the system is in a particular region of the state space [22]. When applied to the proposed EMVD, this technique corresponds to dividing the z domain into regions such that the slope of the NTF characteristic (in Fig. 3.2) is approximately linear in each region. An appropriate controller can then be used to control the system in each region. Piecewise linearization is an approximation to a control technique where one uses a continuous nonlinear gain-varying function. In this method, a nonlinear mapping is used such that the controller gains are varied as the system moves from operating in one part of the state space to another. Such nonlinear controllers are easily implemented [23, 24]. 3.5 Feedback-Controlled MIT EMVD Simulation In order to show that the proposed EMVD was feasible, at least in terms of power consumption and ease of control law implementation, we carried out a feedback-controlled simulation of the MIT EMVD using a combination of the control laws described in the previous section. The control method we used was that of designing a nonlinear controller which takes into account the nonlinear system dynamics of the plant. The main control law we used was based on the feedback linearization technique [22], which is inherently different from both the piecewise linearization and continuous gain-varying techniques described in the previous section. The control law for a feedback-linearized nonlinear controller for the proposed EMVD is [8, 9]: X1,d - Fi(xi, x 2 ) - ko(xi - X1,d) - ki( F 2 (Xi, X2) -1,d) (3.15) where x1,d is the desired valve position, and ko and k, are appropriate controller gains [8]. These gains can be determined such that the closed-loop system has a 1kHz bandwidth. The NTF characteristic (see last line in Table 3.1) we chose for the simulation was completely flat at the ends of the stroke (when z -+ ±4mm in Fig. 3.2). This characteristic deteriorates the valve transition time for the EMVD because the acceleration of the valve when it is near the end of the stroke is small [8]. In order to solve this problem, a feedforward control technique can be used: pulses of current can be applied to the motor when the valve is near either end of the stroke [8, 9]. This current injection technique results in an almost 50% reduction in the transition time of the valve when compared to simulations where the - 29 - The MIT EMVD technique is not used [8]. Figure 3.6 shows results from the MATLAB simulation of the MIT EMVD, including the effects of gas force, friction and electric motor losses, using reasonable electrical and mechanical parameters [8]. A variety of control laws were used in this simulation. At the beginning of the simulation, a combination of current injection and a PD controller were used [8]. After the brief period of current injection, the feedback linearization-based control law in (3.15) was used. During the holding time period, another PD controller was used. The spikes seen in Fig. 3.6 are an artefact of the simulation - abrupt switching between the two control laws we used. These spikes can be eliminated in practice [3], possibly by using a filter on the control output to the plant. ~i 1 ft 1 U 4 r _________ 2 Acceleration [km/s TN. - I IAL)riF )r ~! ~LI - - Motor Torque [Nm] -Velocity [m/s] U -2 -4 -6 -8 - - 0 10 2 ~- 4 ~ ~- 10 8 6 -- Position [mm] _ 2 1 - Voltage K1OV] 4Current [x1OA] Power [kW] _____ _____ 1- -2 -4 -6 ______ .i ______ i ______ j ______ 2 6 -8 10 -= Figure 3.6 4 £______ uhf - L______ .1 ______ .i..J __ 8 10 12 14 Simulation of the MIT electromagnetic valve drive [8]. The simulation results showed that the holding and driving currents are small, and the overall power appears to be technically and economically feasible. In the simulation, which was carried out at 6000rpm engine speed (corresponding to fast transition times) and wide open throttle conditions, the transition time for the valve was approximately 4ms, the seating velocity was less than 0.5cm/s [8, 9]. The total average electrical power input was approximately 1.2kW for 16 valve actuators in a 2.OL, 4-cylinder engine. In comparison, the average power for the valve drive of a conventional IC engine is 3kW, while that for a IC engine with roller-follower type cams is 1.5kW [8]. These power losses include both the required power to compensate for gas forces and the electrical and mechanical power losses in the EMVD. - 30 - 3.5 Feedback- Controlled MIT EMVD Simulation The resulting valve and motor power profiles in Fig. 3.6 were used to select the components for the EMVD apparatus, as well as for the experimental test-stand. We turn to component selection for the EMVD apparatus, and EMVD apparatus modeling, design and construction in the next chapter. - 31 - Chapter 4 The Experimental EMVD Test Stand 4.1 Introduction IN order to prove the concept of the proposed EMVD, we have designed and constructed an experimental EMVD apparatus on a workbench. This apparatus has been integrated into a computer-controlled experimental test stand, thereby achieving one of the main objectives of this thesis research. This chapter details the design, modeling, construction, and assembly of this experimental test stand. We will discuss the design and construction of the EMVD apparatus, including the nonlinear mechanical transformer. 4.2 Preliminary Design of the Experimental Test Stand The simulation result in Fig. 3.6 showed, theoretically, that the concept of using a nonlinear mechanical transformer was valid. This simulation was based on the z domain nonlinear state-space model of the MIT EMVD. Having obtained satisfactory simulation results, we decided to construct an EMVD apparatus which would demonstrate the benefit of using a nonlinear mechanical transformer in a motor-driven engine valve-spring system. We wanted to integrate this EMVD apparatus into a test stand from which we could perform experiments on the EMVD. Initially, we envisioned a test-stand that was comprised of a computer-controlled digital signal processor (DSP) to control the EMVD, power electronic circuits to process the control signal from the DSP, an oscilloscope to measure and display relevant experimental parameters, an appropriate motor to drive the engine valve-spring system, appropriate displacement sensors, and an EMVD apparatus, which would be comprised of the motor, the engine valve-spring system, and the mechanical parts connecting these two. With this test stand, we wanted to perform experiments to characterize MIT's EMVD, and compare it to theoretical simulation results, such as those in Fig. 3.6. We also wanted to - 33 - The Experimental EMVD Test Stand collect data such as valve position, motor position, valve velocity, motor velocity, motor current, motor voltage, motor power, and motor drive circuit power in real-time, that is, while an experiment was being conducted. If the experimental results validated the concept of the MIT EMVD, we would have demonstrated an extremely viable candidate for electromagnetic engine valve drive systems which could then be implemented in automotive engines. A schematic of the test stand is shown in Fig. 4.1. DSP EMVD Apparatus Motor Motor Drive PC Oscilloscope Figure 4.1 """"""""" Rotary Position Sensor Linear Position Sensor Block diagram of the EMVD experimental test stand. There were three primary constraints on the design of the experimental test stand. First, we wanted the mass and inertia of the moving components in the apparatus, for example, of the valve, to be as small as possible, because the larger the mass/inertia of components in the system, the larger the strain on the motor. Thus, we tried to make the components as small and light as possible without compromising their mechanical capabilities. Second, we had wanted to achieve an overall system bandwidth of about 1kHz (see section 3.4.1), so we wanted the bandwidth of all the electrical components in the experimental test stand to be approximately 10kHz, such that their mechanical and electrical dynamics did not affect the overall feedback-controlled system dynamics significantly. Third, we wanted to obtain off-the-shelf components wherever possible in order to minimize the time required for designing our own components and for constructing the test stand. Keeping these constraints in mind, we carefully designed and constructed or purchased all the components that would be integrated into the experimental test stand. We decided to select components such as the motor, engine valve, position sensors, and springs first, before carefully designing other mechanical components for the test stand, such as the nonlinear mechanical transformer. In the next few sections, the selection of the off-the-shelf components for the test stand, and the design, modeling, construction, and assembly of the EMVD apparatus will be discussed. - 34 - 4.3 4.3 Selection of Components for the Test Stand Selection of Components for the Test Stand In this section, the selection of off-the-shelf equipment for the experimental test stand is described. In general, this equipment was selected by carefully considering the bandwidth, mass/inertia and cost constraints for the EMVD apparatus. In order to control the EMVD apparatus, we needed a DSP capable of processing signals from the position sensors and producing a control signal for the motor with reasonable sampling intervals. I selected dSPACE's DS1104 processor for this purpose, primarily because of its ability to integrate well with MATLAB and Simulink. This feature was particularly advantageous because all the simulations of the MIT EMVD had been done in either MATLAB or Simulink. The DS1104 processor is connected to a PC in a PCI slot. To use the DS1104 processor, one simply builds a Simulink file, which is then automatically compiled and run on the processor. The DS1104 comes with a Simulink toolbox, which contains Simulink blocks such as the DS1104 A/D and D/A channel blocks, that can be used with blocks in other Simulink toolboxes, to build Simulink files that will run in real time. In addition, the inputs and outputs of the A/D and D/A channels, as well as the values of variables in the compiled Simulink file, can be changed in real time using dSPACE Control Desk software. The speed at which the DS1104 works is based on the hardware specifications of the computer used. Based on the hardware requirements we obtained from dSPACE, we obtained a Dell PC with appropriate memory size (256MB) and sufficient processor speed (2.4GHz). Using this computer, we were able to achieve sampling rates of 80kHz on the DSP. We also wanted to collect data on the power consumption of the motor. Thus, we obtained a high-sampling rate digital oscilloscope with a GPIB card and Lab View software. The GPIB card was connected to another PCI slot in the PC, and using Lab View software, we were able to capture screens of data from the oscilloscope. To measure motor current, we obtained a high-bandwidth (several megahertz) hall-effect current probe, and to measure voltage, a high-bandwidth differential voltage probe was obtained. These probes could also be used to measure the power consumed by the motor drive. One of the most critical components in the experimental test stand was the motor. We obtained an off-the-shelf motor (Pacific Scientific's 4N63-100 low inertia permanent magnet dc motor) with a large torque-to-rotor inertia ratio, high power rating, and appropriate electrical and mechanical time constants for the proposed EMVD. In particular, we chose a motor with nominally low inductance (100pH) and low resistance (1Q) in order to make it easier to design the motor drive electronics since the electrical time constant of the motor - 35 - The Experimental EMVD Test Stand can then potentially be much larger than that of the motor drive. The motor was also chosen because it is able to respond with enough torque up to frequencies of 150Hz, and has an appropriate torque constant, KT [9]. The adequacy of these motor characteristics was determined using the simulation of Fig. 3.6. Unfortunately, the motor we chose was large in size - too large to be easily implemented in an actuation system on a engine head. Nonetheless, even if a custom-made motor had been smaller in size, it would have been much more expensive. When the proposed EMVD is implemented on a cylinder head, smaller motors will have to be custom-made [9]. So as to actuate the motor with an instantaneously-acting current source, which was assumed in the simulation of Fig. 3.6, we needed to obtain a high-bandwidth motor drive circuit. However, it was not possible to obtain a high-bandwidth motor drive for the EMVD apparatus at a reasonable cost. Therefore, we designed and constructed an appropriate motor drive circuit. The design, construction, and testing of this circuit is described in Chapter 5. Based on some rough calculation, we were not confident that the motor bearings alone would be adequate to support the side load of the valve drive, and thus, to provide additional support, we obtained an appropriately sized deep-grove ball bearing (SKF's 61901 12mm ball bearing). Standard exhaust engine valves from a Ford Zetec 16 valve, 2.OL cylinder head were acquired for the EMVD apparatus. These valves were removed from a cylinder head donated by the Ford Motor Corporation. The mass of the valves, each weighing approximately 60g, was determined to be similar to the value used in simulations. To meet the 150Hz natural frequency requirement of the proposed EMVD, we chose die springs with the appropriate z domain stiffness. These springs were also chosen such that their effective 0 domain inertia was small [9]. This effective inertia is comprised of their mass in the z domain reflected to the 0 domain through the NTF characteristic equations (3.1), (3.2), (3.3), and (3.4). Furthermore, the stiffness of the springs was determined after carefully choosing an appropriate characteristic for the NTF. It is important to note that the inertia of the springs and the NTF cannot be neglected because they have a strong effect on system dynamics [9]. Considering these constraints, we decided to obtain die springs from McMaster Carr Supply Company that had appropriate lengths and diameters: a soft (effective stiffness=28.71b/in) spring set; a stiffer (effective stiffness=3201b/in) spring set; and a stiffest (effective stiffness=8001b/in) spring set. We picked the stiffest springs to allow fast valve transitions, and we designed the EMVD apparatus for this set of springs. The soft and stiffer spring sets were picked so - 36 4.4 Mechanical Components for EMVD Apparatus that we could use them in our preliminary experiments with the apparatus - to verify the operation of the EMVD at lower effective engine speeds. Thus, these two sets of springs had the same free length and diameter as the stiffest spring set. For the rotary position sensor, we purchased US Digital's E6D differential optical encoder with a 8192-line resolution. This number of lines gives high enough resolution (approximately 0.0008 radians/line) and bandwidth for our application. In particular, the resolution of the sensor, when reflected from the 0 domain to the z domain through the NTF, becomes higher as the valve approaches an end of the stroke. From a control point of view, this high resolution implies that high precision position control can be carried out at the ends of the stroke, where the valve's seating velocity must be effectively controlled [9]. We chose a high-bandwidth, low mass, variable-reluctance type linear position sensor (Sentech's Fastar FS300) for measuring valve displacement in the z domain. This sensor, although not critical to the control of the EMVD, allowed for an investigation of the effects of the compliance between the motor shaft rotation and the valve's motion [9], and an accurate measurement of valve seating velocity. 4.4 Mechanical Components for EMVD Apparatus In this section, we will describe the design of various mechanical components for the EMVD apparatus. The design and solid modeling of the EMVD apparatus, including the nonlinear mechanical transformer, is discussed first. This discussion is followed by a description of the construction and assembly of the apparatus. We conclude with more detail on two components in the EMVD test stand: the linear position sensor and the valve seat. 4.4.1 General Structure of the EMVD Apparatus Having selected and obtained several components (three sets of springs, engine valves, a motor, position sensors, a DSP, an oscilloscope, Lab View, dSPACE, and MATLAB software) for the EMVD experimental test stand, we were able to envision a more accurate version of the EMVD apparatus. After carefully measuring the components we had obtained, we were able to draw a 2-D model of the apparatus. Figure 4.2 shows this model of the EMVD apparatus. The general structure of the apparatus is shown in Fig. 4.2, although the dimensions on the drawing are not completely accurate. The entire apparatus is mounted on a table, with two - 37 - The Experimental EMVD Test Stand Motor Optical Motor Mount Encode Disk Bearing Housinn Cam Bearing Valve Holder TopPat Codr mpring hSun pin t . TDivider Spring 11 Vave v ehh s e I I Figure 4.2 Valve Seat Table An accurate model of the experimental EMVD test stand. supporting columns, labeled column I and column II. The motor is mounted onto column I with two motor mounting brackets. The optical encoder is mounted on the rear end of the motor shaft, while the external motor bearing is placed at the front end of the motor shaft. This bearing is enclosed in a bearing housing, that is mounted in a bearing housing holder, which is in turn mounted onto column II. Two springs are placed on the engine valve stem with a spring divider separating them. There is a valve holder at the top of the valve stem, and this valve holder is connected to the NTF using a roller-follower (IKO's CFS-5-V rollers). For simplicity, we decided to use a disk-cam as the NTF, though the exact structure of the NTF will be discussed in detail in a later section of this chapter. The NTF is in turn mounted onto the motor shaft. The valve assembly (engine valve, springs, spring divider) is mounted between two plates, labeled top and bottom plates in the figure. In addition, there is a valve seat plate mounted to columns I and II. The linear position sensor is mounted to the bottom face of the engine valve. To mount the structure of Fig. 4.2 securely, we obtained a high load capacity 36-by-24in - 38 - 4.4 Mechanical Components for EMVD Apparatus steel table (McMaster Carr Supply Company's 4769T44 Table). This table is supplied with a working surface flat to 0.002in/ft. 4.4.2 Mechanical Component Design and 3-D Solid Modeling Having constructed a complete 2-D model of the EMVD apparatus, we decided to carefully design each mechanical part on paper and then construct a 3-D model of the apparatus in 3-D solid modeling software. The purpose of using 3-D solid modeling software was threefold: to be able to virtually assemble and disassemble the EMVD apparatus and get a good 3-D vision of the structure of the apparatus, from which one could quickly see design inaccuracies and deficiencies; to be able to quickly change dimensions and shapes of the mechanical parts in software; and, to be able to generate neat drawings of the mechanical parts for the EMVD apparatus, from which these parts could be constructed. For the 3-D solid modeling, we used SolidWorks software due to its simple user interface. There were three constraints on the design of the mechanical parts for the EMVD apparatus. First, the mass and inertia of the moving mechanical parts had to be as small as possible, without compromising the load capabilities of these parts. Second, the parts had to be designed such that assembling and disassembling the apparatus would be straightforward. Third, the parts had to be designed such that they could be constructed easily by an experienced machinist. The mechanical parts to be designed were the two columns, the two motor-mounting plates, the bearing housing, the bearing housing holder, the bearing sleeve, the top and bottom plates, the spring divider, the valve holder, and the linear position sensor mount. In addition, we had to specify the locations and sizes of mounting holes that had to be drilled in the steel table. The initial design of most of the mechanical parts was done by Woo Sok Chang1 , while I modeled and refined the designed parts in SolidWorks, and made sure these parts would be feasible in terms of assembly and disassembly of the apparatus. It took several months to learn how to use SolidWorks and construct a complete 3-D model of the EMVD apparatus. Although the modeling of each individual part was straightforward, the assembly of these parts into a complete 3-D model was difficult. Each pair of parts had to be fit together precisely, with exact constraints on their combined motion. So as to obtain as complete a model as possible, parts of the EMVD apparatus we had obtained earlier, such as the steel table and motor, also had to be modeled in SolidWorks and assembled with other parts we had designed and modeled ourselves. In addition, the nuts and bolts Some more details on these designs will be published in Woo Sok Chang's doctoral thesis. 39 - The Experimental EMVD Test Stand for each part had to be precisely located and incorporated into the SolidWorks model. Figure 4.3 is a cross-section of the 3-D model of the EMVD apparatus, showing the motor, disk cam (NTF), roller-follower, valve, valve holder, springs, and linear position sensor. In this model, we can also see the top and bottom plates connected together with three bolts and two 0.5in precision bars, and the valve seat plate, through which the linear position sensor is mounted. The exact manner in which this position sensor was mounted will be discussed in section 4.4.5, although one can immediately observe the change in the sensor's location when compared to its location in Fig. 4.2. C Motor Roller Valve Holder Disk Cam Spring Divider Springs Position Sensor Figure 4.3 Cross-section of the EMVD apparatus. The most difficult component to design was the disk cam as we had to mathematically derive the profile for the slot in the cam. The design of this part is discussed in detail in section 4.4.3. While virtually assembling the apparatus, I found several parts that did not assemble together correctly. Many of the parts had to be redesigned in order to make assembling the apparatus easier. These parts were redesigned and quickly remodeled in SolidWorks. Thus, we quickly realized the benefit of using this 3-D modeling software. In order to make the construction of the parts easier, wherever possible I standardized the sizes of all the holes to American drill bit sizes. In addition, appropriate tolerances (generally, 0.010in or less, depending on the accuracy needed) for the dimensions, flatness, parallel sides, and perpendicularity of each of the parts were specified. Tolerances were also specified for the holes in the parts. Each clearance hole was made at least 0.020in larger - 40 - 4.4 Mechanical Components for EMVD Apparatus than the nominal diameter. These tolerances were used to allow for more flexibility (more degrees of freedom) in the assembly of the apparatus. Several parts in the EMVD apparatus had to be press-fit together. Woo Sok Chang calculated the necessary press-fit dimensions on the parts that were to be press-fit together, such as the spring divider to the engine valve, the disk cam to the motor shaft, the bearing to the motor shaft, and the valve holder to the valve. Because the bearing we purchased had a diameter larger than the motor shaft, we had to design a bearing sleeve part that would be press-fit to the bearing before the bearing was press-fit to the motor shaft. Figure 4.4 shows a 3-D side view of the EMVD apparatus. The figure shows some nuts, bolts, and washers for the holes in the various mechanical parts, which were selected and purchased off-the-shelf once we were satisfied with the accuracy of the SolidWorks model. Encoder MoTr Dis Cam Bearing Springs Position Sensor Figure 4.4 3-D side-view of the EMVD apparatus. I generated final versions of the drawings for the parts in SolidWorks, including specifications for the hole locations in the steel table. Appendix C contains the final versions of these drawings, including tolerance specifications, from which parts for the EMVD apparatus were constructed. We constructed two sets of parts for the valve assembly (the spring dividers, valve holders, top plates, bottom plates, and valve seats) because we wanted to construct two different valve assemblies with the soft and stiffer sets of springs. - 41 - The Experimental EMVD Test Stand 4.4.3 The Disk Cam - NTF The disk cam was the most important part in the EMVD apparatus. In this section, we begin by discussing the design of the cam, and follow this discussion with a derivation of the roller-follower profile in the disk cam slot. In the next section, we detail the construction and assembly of the entire EMVD apparatus, including this disk cam. As we mentioned earlier, for simplicity, we chose a disk cam to be the NTF. The disk cam was intended to be press-fit to the motor shaft and has a slot with a nonlinear shape as shown in Fig. 4.3. As the disk cam rotates with the motor shaft, a roller-follower whose shaft is connected to the valve rolls over either the top or the bottom surface of this disk cam slot [9]. The valve and the roller-follower are free to move up and down, but constrained in all other possible directions of motion. The shape of the disk cam slot can determine any desired nonlinear mechanical transformer characteristic, such as that in Fig. 3.2, which was obtained using the following nonlinear relation: f (0) = (0.004m) sin(3.46 )(4.1) sin(0.9997r/2) z The disk cam design was carried out in several steps. First, for the profile of the center of the roller-follower in the disk cam slot, for simplicity, we decided to implement the relation in equation (4.1), which is not optimal in terms of power transfer between the 0 and z domains. However, an optimal mid-stroke linearized transformer modulus was calculated such that at mid-stroke the maximum available power in the 0 domain could be delivered to the load in the z domain. This value was used as a guide for the disk cam design. The optimal mid-stroke transformer modulus, r, was calculated by taking into account the inertia of the motor, disk cam, and motor bearing, approximately 6.9 -10- 6kgm 2 , and then using the estimated moving mass in the z domain (90g) as follows: Jo _6.9.- m;z 10- 6 kgm 2 =8.76- 10- 3m 090kg 0.090kg (4.2) For the relation (4.1), the mid-stroke (0 = 0) modulus is actually 13.6mm. Nonetheless, still using the optimal r, and the total displacement (the valve stroke) in the z domain (8mm), we determined the (linearized) angle, Onominal, which the disk cam slot must span as follows: 0-3 Znmial__8 0.913 rad = 52 degrees (4.3) 8.76 -10-33 r After determining the angle of rotation in the 0 domain, the next step was to determine points for the slot in the disk cam. This slot would essentially be a profile traced out by the roller-follower, and thus, we had to mathematically derive this profile. We intended to Onominal = Znominal - 42 - 4.4 Mechanical Components for EMVD Apparatus give these points to the machinist who would be making this cam. For the profile of the center of the roller-follower in the disk cam slot, we decided to implement the relation in equation (4.1) for -26' < 0 < 260. In order to realize this function and be able to create a slot in the disk cam with a precise machine, we had to derive this roller-follower profile in planar coordinates from its current (z, 0) coordinate system. In the next several paragraphs, we will carry out this derivation. Our eventual goal in the derivation of the roller-follower profile was to have the center of the roller-follower trace out the NTF characteristic in (4.1). Figure 4.5 is a plot of this NTF characteristic. X 10-3Z=f(0) 4- E 0 -2 -3. -30 Figure 4.5 -20 -10 0 0 (degrees) 10 20 30 The desired nonlinear mechanical transformer characteristic. Based on space constraints, we decided to locate the (0, 0) point in Fig. 4.5 16.75mm below the center of the motor shaft. Thus, we first shifted the roller-follower profile in Fig. 4.5 in the positive z coordinate direction by 16.75mm. Figure 4.6 shows the resulting translated profile. After shifting the roller-profile, we changed coordinates from polar (z - 0) to rectangular 2 2 (x - y), being careful in defining the x and y coordinates such that z = Ix + y . For convenience, we decided to define y = -z, and then define x such that we were using a righthand coordinate system. The following relations were used to accomplish the coordinate change: x = (z + ro) - sin(0) (4.4) y = -(z + ro) - cos(0) (4.5) - 43 - The Experimental EMVD Test Stand Z=l(0)+r 0021 0.02 - ..- .----.-. -. .-.---.-- .. -.- -.-.- ..- 0.019F-+.- 0.018F ?0.017 [ - - - --. . -. .... - ---- - 0.016- 0.015- 0.014 0.013-20 -30 Figure 4.6 0 -10 0 (degrees) 20 10 The translated NTF characteristic (polar coordinates). where z is given by equation (4.1) and ro = 16.75mm. The negative sign in the equation for y is necessary because of our having defined y = -z, and also because we shifted the original roller-follower profile in the positive z direction. The resulting roller-follower profile is shown in Fig. 4.7. The points used to plot this profile were also used by a machinist later to mill the slot in the disk cam. Y=g(x) -10 -12 - - .... -- .. ..... ...... --. - ...-..-.. -. -. -14 -16 - .. --. . -. .---.. -. -. . -- -..- -- -- -- -.-. .. -.... -.... -- -. .- -18 -20 -4 Figure 4.7 -2 0 2 4 6 8 The translated NTF characteristic (rectangular coordinates). To ensure the accuracy of the roller-follower profile (for the center of the roller-follower) we had just derived, we decided to plot the points of contact of the roller-follower with the upper and lower surfaces of the disk cam slot. In order to plot these contact points, we had to do another mathematical derivation. First, for each point (Xcenter, Ycenter) on the roller-follower profile in Fig. 4.7, the gradient of the roller-follower profile at that point was - 44 - 4.4 Mechanical Components for EMVD Apparatus calculated 2 . This gradient is the same as that of a line tangent to (Xcenter, Ycenter). Using this gradient, the gradient of the line through (Xcenter, Ycenter) and perpendicular to the tangent at (Xcenter, Ycenter) was determined. Using standard trigonometric relations, the upper (Xupper, Yupper) and lower (Xiower, Yiower) contact points were then calculated using the following relations: Xupper Xcenter - rroller -sin(a) (4.6) Yupper Ycenter + rroller -cos(a) Xlawer Xcenter + rroller sin(a) (4.7) (4.8) Ylower Ycenter - rroller cos(a) (4.9) = where rroller denotes the roller-follower radius and a denotes the angle between the tangent line at (Xcenter, Ycenter) and the x axis. The profiles for the center, top, and bottom contact points of the roller-follower that were obtained in this manner are illustrated in Fig. 4.8. Center of ro Top contact Point Bottom contact point 0 .-......... nter -5 ~-10... . 00_ -15 -t0 5 0 -5 Horizontal Displacement Figure 4.8 - t0 15 x (meters) Roller-follower profiles for the disk cam. After gaining complete satisfaction at the accuracy of the roller-follower profile derivation, we modeled the disk cam in SolidWorks. Figure 4.9 shows the SolidWorks model of the disk cam. The slot for this disk cam was created using points for the roller-follower profile generated in MATLAB. At the ends of the stroke, the slot was extended to give an approximately 1mm margin in the z domain when assembling the EMVD apparatus. In addition, at all points on the valve stroke, the disk cam slot was made a little wider than the roller-follower radius to give some clearance when assembling the apparatus (see the MATLAB program used to generate these points in Appendix G). 2 This calculation was done both analytically and numerically and the same results were obtained. - 45 - The Experimental EMVD Test Stand Figure 4.9 2-D side-view of the disk cam. In the next section, the construction and assembly of the EMVD apparatus, including this disk cam, will be discussed. 4.4.4 Construction and Assembly In this section, we will discuss the construction and assembly of the EMVD apparatus, including procedural details on how to assemble and disassemble the apparatus. In effect, it took three months to get the parts for the EMVD apparatus constructed and assembled into a working model. The drawings for all the parts were sent to the MIT Central Machine Shop for construction (see Appendix C for these drawings). It took several weeks to obtain the constructed parts, mainly because we asked that all the parts be constructed in steel (including the two columns), and that they be machined using very precise (more than 0.001in precision) milling machines. From the perspective of an experienced machinist, the disk cam was the most difficult component to construct because it required precise computer-controlled milling of the NTF characteristic. We provided the MIT Central Machine Shop with the SolidWorks model of the disk cam, from which the tool-path coordinator extracted the points needed to mill the disk cam slot with a computer-controlled milling machine. We also requested that two out of the four cams be case-hardened. Figure 4.10 shows a picture of the disk cam with the roller-follower in the disk cam slot. When we received the parts from the MIT Central Machine Shop, we checked them to ensure that they met the specifications listed on the drawings. For many of the parts we intended to assemble by thermal shrink fit, to assure that assembled interface pressures remained within acceptable limits, we specified dimensional error tolerances which were beyond the abilities of the shop. Except for the above-mentioned bearing housing holder, we accepted the parts as machined. With the exception of the valve holder, we had no difficulty in - 46 - 4.4 Mechanical Components for EMVD Apparatus Motor Shaft Hole RollerFollower NTF Profile Figure 4.10 The nonlinear mechanical transformer - disk cam. assembling the shrink fit parts, and to date, we have seen no evidence of movement at the interfaces of these joints. The latter observation suggests that the parts have adequate interface pressure in the assembled state, and the former supports the conjecture that the interface load is not excessive, although the possibility exists that the parts have plastically yielded. In the case of the valve holders, the fit was obviously too loose. For reasons cited, re-making the parts offered little assurance that better results could be achieved. Instead, the sockets of the valve holders were drilled and tapped for radial set screws. The modified valve holders have proven totally adequate for work to date. Once we received all the mechanical parts for the EMVD apparatus, we began the assembling process. During the initial assembly of the EMVD apparatus, we spent time experimenting to find the ideal assembly procedure. We soon discovered several "tricks" that were useful to know when assembling the apparatus. The assembly procedure can be described as a multi stage process. The first stage in assembly was to make the clearance holes in the steel table we had purchased. The second task in the assembly procedure was to assemble the engine valve, a set of the stiffer springs, the spring divider, the top plate, the bottom plate, and the valve seat (the valve assembly) together. We assembled two valve assemblies, the first with the stiffer springs, and the second with the soft springs. Each assembly was done in several steps. First, the valve was passed through the valve seat. Then the bottom plate was placed over the valve stem. After the bottom plate, one of the stiffer springs was loosely placed over the valve stem. The spring divider was then press-fit onto the valve stem using a drill press. We were careful to press the spring divider to a precisely pre-marked location on the valve stem. Another one of the stiffer springs was then loosely placed on the valve stem. We then placed the top plate over this spring. The top plate, bottom plate, and valve seat were then vertically aligned, and we placed three 2in-length, 0.25in-diameter, partially-threaded bolts - 47 - The Experimental EMVD Test Stand (with appropriate washers), and two 0.5in precision bars through holes in the top plate and the bottom plate. We screwed nuts onto the bottom of these bolts underneath the bottom plate. Then, by tightening the nuts below the bottom plate slowly and evenly, we pre-compressed the two springs such that the length between the top and bottom plates was approximately 47mm (a design decision made when modeling the EMVD apparatus), thereby compressing each spring by approximately 5 - 6mm. It is important to note that while tightening these bolts, the top and bottom plates must remain as parallel as possible so as to keep the entire valve assembly well-aligned. To achieve this alignment, at almost every turn while tightening the nuts, we ensured that the precision bars were free to rotate in their clearance holes through the top and bottom plates. Finally, the valve holder was pushed onto the tip of the valve stem, and the set screws on the valve holder were tightened. The resulting valve assembly is shown in Fig. 4.11. Spring Divider Spring Valve Precision Alignment Bars Valve Seat Figure 4.11 The valve assembly with the stiffer spring set. Using the procedure just described, we assembled another valve assembly with the set of soft springs. A different spring divider was used to accomodate the slightly smaller diameter of the stiffer springs. After examining the constructed valve assemblies, we realized that the valves in each assembly were not completely constrained to move only vertically - the valves could still move around significantly within the holes in the top and bottom plates. To solve this problem, I designed four bushings, one for each top/bottom plate, to be press-fit into these plates, keeping an approximate 0.005in clearance between the valve stem and the internal diame- 48 - 4.4 Mechanical Components for EMVD Apparatus ter of the bushings. The drawings for these bushings are included in Appendix C. These bushings were press-fit into the top and bottom plates. Furthermore, we observed that the bottom plates in the valve assemblies were not held rigidly enough by the three partially-threaded 0.25in bolts. Thus, we ordered 0.25in fullythreaded bolts and used these bolts with several extra nuts to fully secure the bottom plates. To this end, we first removed the old bolts from the valve assemblies. Then, for each valve assembly, we placed three bolts through the top plate. We secured these bolts to the top plate with three nuts and washers tightened underneath the top plate. We then screwed three more nuts and washers onto the bolt. The bolts were then passed through the bottom plate, and three more nuts and washers were used below the bottom plate. The two springs were pre-compressed as before, however, after obtaining the desired 47mm distance between the top and bottom plates, we tightened the nuts that were on top of the bottom plate. In this manner, the rigidity of the entire valve assembly was significantly improved. Figure 4.12 shows a picture of this modified valve assembly with the soft springs. The mounting of the linear position sensor, labeled in this picture, will be described in detail in section 4.4.5. The valve assembly with the stiffer springs was constructed in a similar manner, but since we only had one linear position sensor, we incorporated it into the valve assembly with the soft springs. 0.251n Valve Holder Bolt TOP Plate Precision Aiignment Rar Spring Divider Bottom Plate Valve Linear Position Sensor Figure 4.12 Nut d Valve Seat Modified valve assembly with the soft spring set. After constructing the valve assemblies, the motor, optical encoder, and motor mount plates (the motor assembly) were put together. We only constructed one motor assembly, and this construction was done in several steps. First, the front motor mount was bolted loosely - 49 - The Experimental EMVD Test Stand onto the front of the motor. Then, screws were passed through the optical encoder case and the rear motor mount and screwed onto the rear of the motor. Second, the disk cam was press-fit onto the motor shaft to the pre-determined location using a drill press. The front and rear motor mounts were then loosely bolted to Column I. All the bolts on the motor assembly were then tightened (using shims whenever necessary) while ensuring the motor was well-aligned. To achieve this alignment, at almost every turn of the bolts, we rotated the motor shaft to see if there was an excessive amount of friction acting on the shaft. If we felt a lot of friction, we would loosen the bolts a little and then re-tightened them slowly and evenly. Finally, the optical encoder disk and case cover were mounted on the rear of the motor shaft. We were very careful when handling the encoder disk because it can very easily be damaged. When the encoder disk was mounted (following instructions from US Digital), we again made sure that the motor did not get misaligned. Figures 4.13 and 4.14 show two views of the motor assembly. Front tr Motor Rear Motor Mount Figure 4.13 Disk V Cm Picture of the motor assembly (front view). Having constructed the valve and motor assemblies, the next task was to combine these two assemblies and column II together, using the valve assembly with the soft spring set. First, we rotated the disk cam such that it pointed vertically up - in the direction away from column I. Then we placed the valve assembly onto the top of columns I and II. The disk cam was then rotated back to its normal location. We bolted the top plate onto columns I and II. We then inserted the roller-follower into the valve holder and checked to see if the roller-follower was located directly below the motor shaft in the disk cam slot. If this location was not precise, we would have had to unbolt the top plate and either remount the valve assembly or the motor assembly or both assemblies such that the roller-follower - 50 - 4.4 Mechanical Components for EMVD Apparatus Motr/ Motor Optimt Figure 4.14 Picture of the motor assembly (rear view). would be located directly below the motor shaft. During this process, we made sure that the linear position sensor was aligned with its core which was mounted on the steel table. We had to be careful that the valve assembly and the motor assembly did not become misaligned while assembling them together. We used shims behind the front and rear motor mounts to ensure the motor was aligned. After aligning the motor and valve assemblies, we bolted the valve seat to columns I and II. The exact manner in which we mounted the valve seat will be discussed in more detail in section 4.4.6. Finally, columns I and II were bolted tightly onto the steel table. We wanted to use the set of soft springs in our first group of experiments with the EMVD apparatus. After checking to see that the motor bearings could support the side load of the valve drive, we decided not to attach the additional bearing to the motor shaft. Thus, the bearing sleeve, bearing housing, and bearing housing holder parts were set aside for use in experiments with the stiffer springs. Pictures of the assembled bench-top apparatus, without the motor bearing, the bearing sleeve, and bearing housing holder, are displayed in Figs. 4.15 and 4.16. Figure 4.15 shows the apparatus without column II obstructing the view. - 51 - The Experimental EMVD Test Stand Front Motor Disk Cam Valve Holder Roller Plate Spring Spring Divider Bottom- Precision Bar , Plate Valve Seat Position Sensor Mmant Figure 4.15 Position Sensor Picture of the assembled EMVD apparatus (cross-section view). Rear Motor Mount Front Motor Mount Motor Disk Cam , Valve Holder Optical Encder Column II Column I Top Plate Bottom Plate Figure 4.16 Valve Seat P icture of the assembled EMVD apparatus (side-view). In the next two sections, we will give more details on how the linear position sensor and the valve seat were mounted on the EMVD apparatus. 4.4.5 Mounting the Linear - z domain - Position Sensor In this section, we will give details on how the linear position sensor was mounted to the EMVD apparatus. The linear position sensor is comprised of the sensor armature and the sensor core. In order - 52 - 4.4 Mechanical Components for EMVD Apparatus to mount this sensor, we had to mount the core onto the steel table, and then mount the armature onto one of the moving parts in the valve assembly. Initially (see Fig. 4.2) we envisioned this position sensor being mounted to the valve face, but we soon realized that there would not be enough space to do this. Therefore, we decided to mount the sensor armature on the spring divider. In this respect, we had to modify the spring divider by extending its shape and adding a threaded throughhole to the part. We also modified the bottom plate and the valve seat to allow the position sensor core to pass through these parts. Another mechanical part, the linear position sensor holder, was then constructed to allow us to mount the position sensor core onto the steel table. Top Plate Spring Divider Lock Nuts Bottom Plate Valve Seat Linear / Position Sensor Armature Figure 4.17 Linear, 1 Valve Position Sensor Core The mounted linear position sensor (with column II removed from view). Figures 4.17 and 4.18 are two views of the mounted position sensor, showing the linear position sensor armature and core. After the parts mentioned above were constructed, the position sensor was easily incorporated into the apparatus. In order to mount the sensor, we unbolted the top plate and valve seat from column II, so that we had more room to do the mounting. First, the sensor core was mounted to the table using the sensor holder. Second, the sensor armature was secured to the spring divider using a size 6 screw and two size 6 lock nuts. We were careful to ensure that the center of the sensor armature was aligned with the center of the sensor core, which was mounted onto the steel table. The sensor was thus rigidly connected to the spring divider, making it less susceptible to excessive vibration noise. - 53 - The Experimental EMVD Test Stand \Top Spring- . Nub .uo.l sent Aogui.,. Figure 4.18 4.4.6 The mounted linear position sensor (close-up view). Adjusting the Valve Seat In this section, we will give details on how the valve seat was correctly located in the EMVD apparatus. It is important to note that we carried out this process after the entire EMVD apparatus was assembled and after carrying out some preliminary experiments with the apparatus. ColTMn H Bottom Plate Valve Seat Pieces of Shim Sinck Figure 4.19 The valve seat adjusted to allow for firm valve seating. In order to mount the valve seat correctly, we unbolted the valve seat (assembled to the columns earlier) and inserted shims between columns I and II and the top of the valve seat until we were sure the valve was sealing the hole in the valve seat. To aid in ensuring that - 54 - 4.5 The Experimental Test-Stand the valve was sealing the hole, we used measurements from both the optical encoder and the linear position sensor. We also inserted a thin sheet of paper between the valve and the valve seat - when the valve seals this sheet of paper cannot be pulled out from between these two parts. The sensor measurements also allowed for a determination of the accuracy of the position sensors. Figure 4.19 shows a picture of the adjusted valve seat. 4.5 The Experimental Test-Stand In this section, we will describe the integration of the various components of the experimental EMVD test stand. We begin with a discussion of the components in the test stand including a brief discussion of how the EMVD apparatus works. We follow this discussion with a description of how experiments can be carried out using this test stand. Figure 4.20 shows another picture of the assembled bench-top EMVD apparatus. In this picture, we can clearly view the motor, disk cam, and valve assembly. Motor Mount Motor Drive Isk Cam and RollerFolower A Moto Valve Assembly Figure 4.20 Picture of the assembled EMVD apparatus. From Fig. 4.20, we can quickly understand the operation of the MIT EMVD. As the motor shaft is rotated either clockwise or counterclockwise, the disk cam, which is rigidly connected to the motor shaft also rotates, causing the roller-follower to move within the disk cam slot. When the roller-follower moves in the disk cam slot, the engine valve, which is connected to the roller-follower by the valve holder, moves vertically up or down. The engine valve is constrained by design to move only vertically. - 55 - The Experimental EMVD Test Stand At either end of the disk cam slot, the roller-follower lies directly below the motor shaft on a fairly "flat" part (see Fig. 4.7) of the disk cam slot, and thus the engine valve can be held open and closed without providing any electrical input to the motor - the zero holding current characteristic of the MIT EMVD. The benefit of using springs in the apparatus becomes clearer when considering the motion described above. Once an initial amount of energy is injected into the system by compressing one of the springs, that energy is converted to kinetic energy and then transferred continuously from that spring to the other spring as the valve moves up and down. Hence, ideally, the engine-valve spring system would transition from one end of the stroke to the other without any electrical input. However, in reality, the electric motor has to provide power to overcome friction and gas force in the engine valve-spring system. One of the primary objectives of this thesis was to construct an experimental EMVD test stand. To achieve this goal, the assembled EMVD apparatus was integrated with the PC, the DSP, the oscilloscope (with current and voltage probes), the motor drive, and several power supplies. Figure 4.21 shows a picture of the resulting experimental EMVD test stand. Figure 4.21 Picture of the EMVD experimental test stand. The integration of the EMVD apparatus into the experimental test stand was straightforward. The PC was connected to the oscilloscope with a GPIB card. The dSPACE DSP was inserted into a PCI slot on the PC. Both Lab View and dSPACE Control Desk software were installed on the PC. The current and voltage probes were connected to the oscilloscope - 56 - 4.5 The Experimental Test-Stand at their transmitting ends, and connected to the motor at their sensing ends. Whenever desired for a particular experiment, other inputs were also connected to the oscilloscope. The motor drive, which will be discussed in detail in the next chapter, was connected to the motor in the EMVD apparatus. In addition, an ADC channel on the dSPACE DSP was connected to the motor drive input terminals, to allow the current command input from the controller in the DSP to reach the dc motor. The power supplies were connected to the motor drive and the linear position sensor. The outputs of the optical encoder and linear position sensor were connected to two separate input channels (an encoder receiver channel and an ADC channel respectively) on the DSP. All the inputs to and outputs from the DSP were actually connected to the DSP I/O channels box. This box is physically attached to the DSP that resides inside the PC. To run an experiment with the EMVD apparatus, one begins by modeling a controller in Simulink software on the PC. When this model is compiled, it automatically runs on the DSP. As noted earlier, the DSP comes with its own Simulink toolbox, which contains blocks such as the ADC and DAC channels, which can be used to build Simulink models that will run in real time. Thus, in the Simulink models, we can use inputs from the ADC channels on the DSP, and provide outputs to the DAC channels on the DSP. The speed at which the DSP processes these signals is dependent on both the speed of the PC as well as the complexity of the Simulink model. In addition to implementing controllers in Simulink, one can also design and implement filters, and implement various mathematical operations. Generally, during an experiment, motor and valve displacements from the EMVD apparatus are sensed through the ADC channels on the dSPACE board to which the position sensor outputs are connected. These displacements can be displayed in dSPACE Control Desk or on the oscilloscope, or on both devices. Motor current, voltage, and/or motor power are displayed on the oscilloscope during the experiment 3 at a 2.5MHz sampling rate. The values of variables in the Simulink model can be viewed and modified in real time using dSPACE Control Desk. With the experimental set-up described above, we can obtain and save experimental data in two ways. First, any of the variables in the Simulink model that are being displayed in dSPACE Control Desk can be "captured" (the term used by dSPACE) and saved to a MATLAB .mat data file. These files can then be opened in MATLAB and the data arrays in these files can then be processed and plotted (see, for example, the MATLAB file emvddataprocess.m in Appendix L) . Second, any data displayed on the 4-channel oscilloscope can be saved on the PC using a Lab View software file. The data from the oscilloscope is saved in IEEE binary format by 3 The sampling rate on the DSP is not high enough to display these variables in dSPACE Control Desk. - 57 - The Experimental EMVD Test Stand Lab View, and thus must be converted to MATLAB ASCII format before processing and plotting the data. The MATLAB file readbin.m in Appendix L was used to carry out this conversion. We discussed the design, construction, and assembly of the experimental EMVD test stand in this chapter. In the next chapter, we turn to the design and construction of the motor drive circuit. We also discuss experiments that were carried out to test the motor and motor drive. - 58 - Chapter 5 The Motor and The Motor Drive 5.1 Introduction IN this chapter, we will describe the design, construction, modeling, and testing of the motor drive circuit that was used to drive the motor in the EMVD apparatus. In addition, we will describe the modeling and testing of the Pacific Scientific permanent magnet dc motors we purchased. We will begin by describing the motor drive circuit, and follow this description with a discussion of the dc motor, including tests we performed to extract motor parameters. At the outset, it is important to note that the MIT EMVD poses significant challenges in the area of electrical and electronic component design, including, but not limited to, the motor drive circuit. 5.2 Design and Construction of the Motor Drive In this section, the design and construction of the motor drive circuit is described. In the next two sections, we will give details on the testing of this circuit. There were five primary design constraints on the design of the motor drive: because of the use of the current injection technique we wanted to use to control the EMVD, we needed a motor drive with a high slew rate capability (70A/ms) [9]; the bandwidth of the motor drive had to be approximately 10kHz; the motor drive had to be able to source approximately 1kW of average power; the motor drive had to be able to connect directly to the ±10V output from the ADC channel on the dSPACE DSP; and finally, the motor drive had to be able to provide bi-directional current, so that the motor shaft could be rotated in either direction. The reasons underlying most of these constraints were described in section 3.4.1 of this thesis report. As was mentioned earlier, the simulation of Fig. 3.6 assumed that a high-bandwidth current - 59 - The Motor and The Motor Drive source was instantaneously supplying current to the motor. Because it was not possible to buy a high-bandwidth motor drive for the EMVD apparatus at a reasonable cost, we had to design and construct an appropriate motor drive circuit. Furthermore, when simulating the feedback-controlled MIT EMVD, we had selected motor current to act as the control input. This control method provides a direct relationship between input signal and motor torque, allowing fast response to changes in the desired valve profile. Thus, the motor drive circuit had to be able to take a given reference current input and provide a current output to the motor approximately equal to this input, using an appropriate controller to guarantee this equality. Bearing this current relationship in mind, we chose a full-bridge (bi-directional) PWM inverter topology with hysteretic current control to implement the motor drive circuit [25, 26]. A block diagram of the hysteretic current-controlled full-bridge motor drive circuit appears in Fig. 5.11. current command Ripple Current +50v Current Sense Reference Hysteresis Band Set High Side Gate Drive Current Comparators PWM + Delay - Motor - Logic Low Side Gate Drive Figure 5.1 A hysteretic current-controlled motor drive. The hysteresis band size (or, effectively, the ripple current), and switching frequency are related by the following equation: 1 Iripple,pk 2 di T dt 1lVbus 2f L (5.1) 'For more details on the circuit topology, design, schematics, or layout, please see "Design and Implementation of a Motor Drive Amplifier", a report by Michael Seeman'04. This item is available on request from the Laboratory of Electromagnetic and Electronic Systems at MIT. - 60 - 5.2 Design and Construction of the Motor Drive where Iripple,pk is half the peak-to-peak ripple current, L is the motor inductance, Vbus is the bus voltage, T is the switching period, f is the switching frequency, and ! is the desired inverter slew rate. In our application, the ripple current is not given by equation (5.1) because of the presence of the back EMF from the dc motor. The hysteresis band size was determined after selecting the appropriate bus voltage and slew rate. This band was chosen to obtain a reasonable ripple current at a switching frequency where power dissipation is not too high. Although the hysteresis current-control method does not have a fixed frequency, a worst-case frequency can be found for a particular hysteresis band. Thus, in our motor drive inverter circuit, the ripple current does not change, but the switching frequency does vary. For our motor drive, we assumed a worstcase switching frequency of 300kHz for the MOSFETs and gate drivers. The hysteresis current control is carried out by a combination of the hysteresis band set sub-circuit, the current comparator sub-circuit, and the PWM/logic sub-circuit, all labeled in Fig. 5.1. This control is thus achieved in several stages. First, the current command (actually a voltage signal from the DSP) is input to the motor drive circuit. A ripple current reference is generated and then both added to and subtracted from this current command using adder and subtractor circuits. This process effectively creates upper and lower hysteresis bands around the current command input. The ripple current reference was generated using a zener diode together with National Semiconductor's LM4040-2.5 2.5V dc supply. The adder and subtractor circuits were designed using standard non-inverting op-amp configurations. We selected National Semiconductor's LF411 op-amps for these circuits. These op-amps were selected primarily because they could use a ±12V supply, they had high slew rates (up to 70V/ts), and they had high gain-bandwidth products. The resistors used in these op-amp circuits were accurate to within 1%. Figure 5.2 shows expected hysteresis bands for our motor drive inverter circuit with the effective ripple current reference, Vref, equal to 47mV - this ripple reference was generated using a circuit with a potentiometer connected to the zener diode. This figure was generated by the MATLAB file hysteresis.m in Appendix J. Second, the motor current is sensed (see "motor sense" in Fig. 5.1) and compared to the upper and lower hysteresis bands simultaneously. This comparison was done using two of National Semiconductor's LM319N comparators. The supply voltage for this comparator (12V) was ground-referenced, making the comparators compatible with the PWM/logic sub-circuit they were feeding. The motor current was sensed by passing wires carrying this motor current through a LEM LA 55-P 50A 200kHz hall-effect current sensor. Third, the outputs of the comparators are processed by a logic circuit, and the gate drivers - 61 - The Motor and The Motor Drive Hysteresis Bands in the Motor Drive with V =47mV 0.02 ihVe oo . . .rv h ~~~~ysteresis ... .. . . and h~ 5.2 Figure Appropriate -0.02 .. . . . . .. if current the motor-0.04 is larger the . . . . . ... .. . . .. ..than 0 Figure 5.2 1 2 urpen ytrss 4 Time 3 ... .. . . badonmpmandM 5 6 7 7V SFTS 8 7 Appropriate hysteresis bands for the motor drive with V ref 47mV. for the MOSFETs are activated based on the outputs of this logic circuit. For instance, if the motor current is larger than the upper hysteresis band, one pair of MOSFETS in the bridge is turned on and the other pair is turned off. Similarly, if the motor current is smaller than the lower hysteresis band level, the other pair of MOSFETS in the bridge is turned on and this pair is turned off. If the motor current is between the upper and lower hysteresis bands, the state of the MOSFETs in the bridge is not modified. We selected Texas Instrument's CD4011BE NAND gate chip to implement an RS flip-flop to perform as the logic circuit. We added R-C-D delay circuits at the gate outputs to prevent current shoot-through in the MOSFET bridge - the diodes in these circuits were used to delay the turn-on of the MOSFETs. In fact, the resistances in the delay circuits were eventually implemented as potentiometers to allow for precise adjustment to prevent shootthrough. The outputs from the delay circuit were passed into two International Rectifier IR2110 gate drivers. Each of these gate drivers was used to activate a pair of MOSFETS on the bridge. After carefully considering the thermal characteristics of several MOSFETs, we selected International Rectifier's IRF2807 N-channel MOSFETs (in a TO-220 package) for the bridge, and appropriate heat sinks (Redpoint Thermalloy's KM150-1 heat sinks) for the expected level of power dissipation. We used one heat sink for each pair of MOSFETs. An R-D delay circuit was also used at the gate driver outputs to additionally delay the turn-on of the MOSFETs in the bridge. In particular, we chose the IRF2807 to minimize the power dissipation at the assumed - 62 - 5.2 Design and Construction of the Motor Drive worst-case switching frequency of 300kHz. At a load current of 15A, switching frequency of 300kHz, and ripple current amplitude of 0.95A (corresponding to a motor torque ripple amplitude of 0.067Nm), an IRF2807 MOSFET at nearly 100% duty cycle would have a switching loss of 3.09W and a conduction loss of 5.35W, totalling 8.44W for each of the four MOSFETS in the bridge. For power supplies, we selected a Hewlett Packard 60V 9A power supply to provide the bus voltage, and a Tektronix 30V 3A power supply to act as the control circuit power supply. To reduce EMI in the bus voltage power supply, we used two 22000pF electrolytic capacitors connected in parallel across the voltage bus. These capacitors were rated for IGA of ripple current through the voltage bus at the worst-case (300kHz) switching frequency. In effect, hysteresis current control keeps the actual motor current within a certain hysteresis band of the desired motor current by switching on diagonal pairs of MOSFETs in the bridge. This control method features a simple control loop, fast response time, well-defined ripple current, and variable switching frequency, which is a function of load and input signal [9]. However, due to the non-integrating nature of the feedback loop, the controller has a nonzero tracking error. This tracking error is bounded by the magnitude of the hysteresis band. In laboratory tests carried out before we constructed the motor drive inverter circuit, we observed that the control circuit needs to be very precise, especially in the generation of the hysteresis bands. Small errors in the precision of different parts of the control circuit can add up incrementally and cause the overall control loop response to deteriorate significantly. The hysteretic current-control design is also highly sensitive to the applied load - in our case, the resistance and inductance of the motor in series with the motor's back EMF [25]. Solving Kirchoff's voltage law for this load yields: dim Vbus = imR + KwWm + Lm dt (5.2) where im is the motor current, Rm is the motor resistance, K, is the motor back EMF constant, Lm is the motor inductance, and d is the inverter slew rate. In effect, the slew rate of the inverter circuit determines how fast large amounts of current can be driven into the load, and thus, how large a bus voltage is required. A higher slew rate can make a hysteresis current controlled circuit unstable because higher slew rates cause an increase in switching frequency for fixed ripple current [9]. To determine the required bus voltage for the inverter circuit, we first obtained the necessary motor parameters for use in equation (5.2). For our dc motor, the inductance is 100pH, the - 63 - The Motor and The Motor Drive a/, and the rated rms current armature resistance is 1Q, the back EMF constant is 0.07 rad/s~ is 15A (for a complete set of motor parameters, please see the data sheet in Appendix I). From simulations of the feedback-controlled EMVD (see Fig. 3.6) where the current injection technique was used, we determined that, optimally, the current in the motor must be able to rise at a rate of 70A/ms [9]. In addition, we determined that the maximum motor velocity in simulations was 3 6 Oad. Using this information in equation (5.2), the required bus voltage was calculated as being 57V. Thus, we would have to provide this bus voltage during the current injection periods to achieve the desired high slew rate. Since the motor we purchased was rated for 42V at continuous duty, we decided to use the larger 57V bus voltage, when required, only for short periods of time. Appendix H contains a schematic for the motor drive inverter circuit. Michael Seeman'04 created a printed circuit board (PCB) layout based on this schematic, and then purchased several PCBs. All the components were also purchased and three working models of the motor drive inverter circuit were constructed by hand. Although we only needed one circuit, we completed two additional boards as reserves. A photograph of a completed motor drive circuit appears in Fig. 5.3. dc bus Current input Command Input Capacitors on the Power Bus Control Circuit Heat Sinks Current Sensor Output to Motor Figure 5.3 The motor drive inverter circuit. This motor drive inverter circuit was easily incorporated into the experimental EMVD test stand. The output from the DSP was connected to the "current command" input in Fig. 5.3 on the motor drive, and the motor was connected across the terminals next to the "current sensor" in Fig. 5.3. In the next section, we will discuss the experiments we carried out on the motor drive. - 64 - 5.3 5.3 Experiments with the Motor Drive Experiments with the Motor Drive We will discuss the experimental testing of the motor drive inverter circuit in this section, beginning with a discussion of the preliminary testing carried out on each motor drive circuit that was constructed. Then, we will discuss the modeling of this motor drive, as well as the experimental verification of this model. We had to test the motor drive for two reasons: to ensure there were no mistakes in the PCB layout and construction and to obtain and verify a mathematical model for the motor drive inverter circuit. 5.3.1 Testing the Motor Drive Inverter Circuit In this section we will discuss the testing of the constructed motor drive inverter circuits. For each motor drive, the same testing procedure was used, and this procedure is described first. We follow this description with some inherent design problems with the motor drive circuit. To test the motor drive, the control sub-circuit was checked first. The control circuit was connected to its power supply (±12V), and a reference signal from the DSP was input to the motor drive circuit. Then, the pins of the hysteresis band op-amps as well as the comparators were checked for correct signal processing. The ripple current reference signal level was set (using a potentiometer) to approximately 50mV. If these op-amps and comparators are functioning well, the inputs and outputs of the CD4011BE logic chip are then checked to ensure consistency with the comparator outputs. During this testing process, the voltage bus was disconnected from the motor drive. If the control sub-circuit is functioning satisfactorily, the power sub-circuit can be tested. This test was carried out without the MOSFETS in place on the motor drive PCB. However, the voltage bus (set at 25V unless the motor is being air cooled) and control circuit power supplies were connected to the motor drive and turned on. A signal generator was used to provide 50% duty cycle 0-to-12V and 12-to-OV voltage pulses to the two inputs on the CD4011BE logic chip. The outputs of this logic chip were then checked for consistency with these input voltage pulses. At this point, the potentiometers in the R-C-D delay sub-circuits were adjusted such that there was no shoot-through in the signals going to the gate drivers. Once the shoot-through was eliminated, the inputs and outputs of the gate drivers were checked. - 65 - The Motor and The Motor Drive If the power sub-circuit minus the MOSFETS and gate drivers is functioning well, these components can be placed in the motor drive PCB, and the power sub-circuit can then be retested to ensure that the MOSFETS are turning on and off, however, a very small bus voltage was used first, to ensure that any remaining shoot-through problems do not destroy the MOSFETs. By placing differential voltage probes (because the high-side gate driver in the motor drive is not ground-referenced) across each MOSFET's drain and source, the potentiometers in the R-C-D delay sub-circuits were again fine-tuned to ensure no current shoot-through in the MOSFETS. If the results from these procedures are satisfactory, the motor drive can then be connected to the motor. While testing our three motor drive circuits with the procedure outlined above, we encountered a few notable issues. First, we observed the sensitivity of the control circuit in the motor drive to the hysteresis band level. If this level is too high (> 150mV), the motor drive control sub-circuit will not function because the MOSFET switching frequency can become unstable (this was observed in experiments). Second, due to a few incorrect traces and incorrect resistor values on the motor drive PCB, several logic chips, op-amps, and comparators were destroyed when we first turned on one of the motor drives. These components had to be replaced, and the incorrect traces were repaired. Third, when we initially constructed the motor drives, we did not use potentiometers in the R-C-D delay sub-circuits, and this led to a very serious shoot-through problem in our first motor drive. Appendix H contains an updated version of the original motor drive inverter circuit schematic, with the updates pertaining to the errors in the printed circuit board. In the next section, we will describe how other motor drive circuit parameters were obtained. 5.3.2 Characterization of the Motor Drive Inverter Circuit We will describe some of the measured motor drive inverter circuit parameters in this section. In addition, we will discuss the modeling of the motor drive, and the verification of the model. We wanted to characterize the motor drive inverter circuit in terms of its thermal capability, bandwidth, slew rate, and ripple current. We quantified each of these parameters in the laboratory. To measure the thermal performance of the motor drive, we used the motor drive to supply a 112H inductor with 30A sinusoidal current at a 50V bus voltage. When this experiment was carried out for 25 minutes, the heat sinks on the motor drive PCB got noticeably warmer, but none of the circuit components were destroyed. Thus, the motor drive inverter 66 5.3 Experiments with the Motor Drive circuit can source more than 1kW of power without any thermal breakdown. The current ripple on the motor drive was set to approximately IA by varying the ripple current reference on the motor drive circuit. The motor drive was then connected to the motor and operated with various command currents waveforms (dc, sinusoidal) and frequencies. We observed only tiny changes (50mA) in the ripple current amplitude. This result was expected because we were still using an inductor connected to the motor drive. With motor back EMF, the ripple current amplitude can change more significantly. To measure the motor drive bandwidth, the motor drive was connected to the same inductor mentioned above, and sinusoidal currents of increasing frequency, and varying amplitude, were used as reference currents for the motor drive. The maximum frequency at which the motor drive tracked (where tracking implies that the motor current is limited between the hysteresis bands around the reference current) the reference current was recorded as the motor drive bandwidth - for our motor drives, this number was approximately 9.5kHz. The breakdown of the motor drive performance occurs because the gate drivers in the inverter circuit cannot switch fast enough to track the reference current. Motor drive slew rate was measured with the motor connected to the motor drive. A step current command from the DSP was input to the motor drive with a 42V bus voltage, and the slope of the actual motor current was determined. Figure 5.4 shows the step response of the motor drive/motor combination with a 7A step in current, from which the slew rate can be estimated at 400A/ms. 10 Time(s) Figure 5.4 Step response of the motor drive/motor combination with a 7A step input. It is important to note that although the motor dynamics, including the changing resistance and back EMF, contribute greatly to some of the motor drive characteristics, these dynamics - 67 - The Motor and The Motor Drive do not affect all the motor drive characteristics. For instance, the use of an inductor to quantify the thermal performance of the motor drive itself, and to measure the motor drive inverter circuit bandwidth, is valid because these characteristics should not be affected by motor dynamics, at least with the assumption that the bus voltage is large enough. A few of the design parameters we measured in the laboratory, together with the design targets are shown in Table 5.1. Quantity Slew Rate Bandwidth Power Current Ripple Measured Value - 400A/ms at 42V bus voltage 9.5kHz > 1kW 0.98A Desired Value - 70A/ms 10kHz 1kW < 1A Table 5.1 Characteristics of the motor drive. Figure 5.5 shows the time response of the motor drive/motor combination with a 7A, 3kHz sinusoidal current command from the DSP, with Vbs = 42V. We can clearly observe that this sinusoidal current command is actually a sampled (with a zero-order hold circuit in the DAC) version of a sine wave. 10 -6 Time(s) Figure 5.5 command. X 0- Current waveforms for the motor drive with a 7A, 3kHz sinusoidal current The experimental results in Figs. 5.4 and 5.5 compare very well with simulation results in [9] that were obtained using a Simulink model where the motor drive circuit was modeled with a hysteresis block. Based on these simulations and experimental results, we concluded that at least in terms of controller design, the hysteresis current controlled motor drive can be treated as a linear gain between the reference motor current and the actual motor - 68 - 5.4 Modeling the dc Motor current. This approximation is justified because the ripple current in the motor drive output is always above the 200kHz frequency range, and can thus be treated as a low-amplitude high-frequency disturbance that would be easily filtered out by the motor. In the rest of this chapter, we will turn to the modeling and characterization of the dc motor we purchased. As we mentioned earlier, this motor was probably one of the most important components in the EMVD apparatus. 5.4 Modeling the de Motor In the next two sections, we will describe laboratory tests that were carried out to characterize the motor. In this section, we will describe the theory underlying these tests. This theory is a review of dc motor modeling, and is only included here for completeness. We can dynamically describe dc motors with the following two relations: Vm = imRm + Kwwm + Lm dt rm = JmL + BmW + Text ; ; (5.3) (5.4) where Vm is motor voltage, im is motor current, Rm is motor resistance, K, is the motor back EMF constant, Wm is motor speed, Lm is motor inductance, rm is torque exerted by the motor, Jm is the motor rotor's inertia, Bm is the viscous friction coefficient, and Text is the external torque applied to the motor. Furthermore, the motor current and torque are related by: Tm = KT - im (5.5) where KT is the motor torque constant. This torque constant is equal in magnitude to K, in a consistent set of units (for instance, KT in m and K, in rad/s) Equation (5.3) describes the electrical dynamics of the dc motor, while equation (5.4) describes its mechanical dynamics. In steady state, when the motor speed, voltage and current are constantdim=-0 and W = 0, and thus, these equations reduce to: areco sta t7dt - Vm = imRm + Kwowm; Tm = BmW + Text. (5.6) (5.7) By substituting equation (5.5) into equation (5.7), and then solving equation (5.6) for im - 69 - The Motor and The Motor Drive and also substituting the result into equation (5.7), the following relation is obtained: Text = KT Rm Vm - (KTK' Bm + IW. Rm (5.8) Thus, by applying an external torque to the motor, and supplying the motor with a constant voltage such that it reaches a steady state speed, we can do a least squares curve fit of the experimental data obtained to equations (5.6) and (5.8), from which we can obtain numerical values for most of the motor parameters. Although KT and K, are constrained by physical laws to be equal in magnitude (in a consistent set of units), the values of these two constants may be slightly different because the two equations are curve fit independently. In addition, the friction coefficient Bm obtained includes some friction that may be applied by the external torque source. This procedure is the basis for the dynamometer test described in section 5.6.1. To find the rotor inertia, Jm, of a dc motor, as well as the viscous friction coefficient, Bm, one must carry out transient response motor tests. The theory underlying these tests is fairly simple. By substituting equation (5.5) into equation (5.4), we obtain the following relation: KTim = JmD + BmW +Text. (5.9) If we assume there is no external torque acting on the motor, we have that: KTim = Jm + Bmw. (5.10) Assuming that the motor starts from rest (P = 0, W = 0), we can take the Laplace Transform of equation (5.10), and find the transfer function between motor velocity, Wm(s), and motor current, im(s): Wi=(S) KT (5.11) im(s) Jms + Bm Assuming that the dc motor is initially at rest, and a step in motor current of amplitude A is applied to the motor (such that im(s) = j), equation (5.11) reduces to: WM(S) = AKT . + BMs JmS 2 (5.12) By carrying out a partial fraction expansion on equation (5.12), and taking the inverse Laplace transform of the result, we have that: Wm(0) AKT AKT _lat = B me- im .(5.13) Bm ~ Bm From this relation, we observe that given a step current input of A, the motor velocity - 70 - 5.5 The Dynamometer Apparatus rises exponentially to a steady state velocity equal to AKT. The time constant, r, for this Bm response is given by: =m (5.14) Bm If a known external inertia, Jf, is incorporated into the motor before the transient response test is carried out, the effective inertia of the motor in equations (5.9) to (5.13) equals Jm plus Jf, and the equation above changes to: Jm + Jf Jm Bm .J (5.15) Therefore, by applying a step current to a dc motor with, and then without, a known inertia Jf, one can solve equations (5.14) and (5.15) simultaneously for the two unknown motor parameters, Jm and Bm. Usually, the inertia Jf is designed to be much larger than Jm. To obtain a more accurate time constant value, we can also apply pulses of current and then find time constants for both the rising and falling parts of the motor velocity response. This procedure was the basis for the transient response tests described in section 5.6.2. To measure the remaining unknown motor parameter, Lm, one can simply use an impedance analyzer, although this inductance could also be obtained from the rise time of the motor current waveform in response to step changes in motor voltage. In the next section, we will describe the design and construction of a dynamometer apparatus which we used to test our dc motors. After this discussion, we will describe experimental results that allowed us to characterize these dc motors. 5.5 The Dynamometer Apparatus The design and construction of the dynamometer apparatus will be discussed in this section. The laboratory dynamometer we used was a Magtrol HD-700 series hysteresis dynamometer. It can be used to apply up to 800oz-in of torque, though we only used it to apply up to 50oz-in. The diameter of this dynamometer's shaft is 0.5in, while our motor shaft diameter was din. To obtain a good connection between the motor and the dynamometer, we used a 0.5in-to-lin flexible coupling (DKN's 45/41-9.51H7-12.66H7 metal bellows coupling). In addition, we needed to construct mechanical parts that would be used to mount the motor and appropriately interface it to the dynamometer. To this end, we designed a mounting block and mounting plate on which to mount the motor. These two parts were designed to be strong enough to support the motor. In particular, the motor mounting plate was designed to be thick enough (1cm thickness) so as to provide enough stiffness to prevent motor - 71 The Motor and The Motor Drive vibration during testing. These two parts were constructed in the laboratory machine shop by Yihui Qiu and Wayne Ryan, the laboratory Engineering Specialist. To ensure we had an accurate model of the dynamometer apparatus (in terms of assembly and disassembly) and to be able to generate drawings of these parts for construction, I modeled these parts in SolidWorks. Appendix D contains the final SolidWorks drawings that were used to construct the parts for the dynamometer test stand. The assembly of this stand was straightforward and will not be discussed here. Figure 5.6 is a picture of the assembled dynamometer test stand, showing the flexible coupling, motor, and dynamometer. Figure 5.6 Picture of the dynamometer apparatus. In the next section, we will describe experiments that were carried out to test the motors, including experiments carried out with this dynamometer test stand. 5.6 Experiments to Obtain Motor Parameters We will discuss the motor tests we performed in this section. We carried out three sets of tests: a motor-dynamometer test to obtain a torque-speed curve for the motor; a transient response test to obtain the motor speed response to a pulsed current input; and a motor inductance measurement test. The first set of tests was used to find all the motor parameters except for Lm, Jm, and Bm. The latter two parameters were found from the second set of tests, while Lm was determined from the third set of tests. - 72 - 5.6 Experiments to Obtain Motor Parameters For each motor test, we carried out a set of two experiments because we purchased two 4N63-100 dc motors from Pacific Scientific. We had planned to use one of these motors in the EMVD apparatus, and keep the other motor on reserve. However, before carrying out strenuous experiments with any of these motors, we wanted to verify that the motors we received had parameters close to the manufacturer's specifications. From here on, the motor on which the disk cam had been placed will be called motor A, while the motor we had on reserve will be referred to as motor B. The experimental motor tests we performed also gave us an accurate measurement of the motor transfer function with speed as the output and current as the input. I carried out all the experimental tests with Michael Seeman and Yihui Qiu. In the next two subsections, we will discuss the dynamometer tests and the motor transient response tests respectively. We will follow these discussions with a description of the motor inductance measurement tests. 5.6.1 Dynamometer Tests Figure 5.7 shows the experimental set up for the dynamometer tests. Figure 5.7 Picture of the dynamometer test stand. The experimental set-up was fairly simple. The motor was connected to the HP power supply we were using for the motor drive, and the dynamometer was connected to a separate Tektronix 30V 6A dc power supply. A current probe was connected to the motor, so as to allow a display of the actual motor current on the oscilloscope. A strobe was purchased and - 73 - The Motor and The Motor Drive used to measure the motor shaft speed (see "strobe" in Fig. 5.7). The motor drive and the DSP were not used in this set-up. The testing procedure was the same for both motor A and motor B. We carried out the tests at 18V motor voltage because the manufacturer had specified results from a torque-speed test performed at this voltage. We first measured the no-load motor speed and current. We then measured the motor speed and current at dynamometer torque load increments of 5ozin, until we were close to the motor's uncooled torque limit of 70oz-in. The dynamometer torque increments were obtained by supplying the dynamometer with increasing magnitude dc currents. The experimental data was then plotted against the manufacturer's specifications (with Rm = 0.99Q) and least-squares curve fit to equations (5.6) and (5.8) using the MATLAB program dynotests.m in Appendix J. Figure 5.8 shows the plotted experimental results in SI units. Clearly, both motor A and motor B are extremely close to the manufacturer's specifications. Torque-Speed Curve at 18V 0.4 0.35-+-- Least Squares Curve Fit - Motor A -0- Least Squares Curve Fit - Motor B Manufacturer Specification 0.3 -- 0.25- z 0.20 0.15- 0.1 - 0.05 - 0 170 Figure 5.8 180 190 200 210 Speed (rad/s) 220 230 240 250 Experimental results from the motor-dynamometer tests. The least squares curve fit parameters are displayed in Tables 5.2 and 5.3. As mentioned earlier, KT and K, are not equal in magnitude because they were obtained from two equations that were curve fit independently. In addition, the least squares curve fit value of the viscous friction coefficient, Bm, was about three times larger than expected. The main reason for this discrepancy was that the least squares curve fit value of Bm includes the - 74 - 5.6 Experiments to Obtain Motor Parameters viscous friction in the dynamometer. It was impossible for us to separate these two frictions during the dynamometer tests. However, since there was no applied torque or torsional damping in the transient response tests (to be described in the next subsection), we were able to obtain a more accurate value of Bm with these tests. Parameter Manufacturer's Specifications Resistance, Rm (Q) Torque Constant, KT Back EMF Constant, ( ""M) K (rad Viscous Friction, Bm (rad/s) 0 ~) Table 5.2 Parameter 0.89@25 C and 1.310155 C 0.07 0.07 7.64 10-5 Viscous Friction, Bm ( Nm ) Table 5.3 Least Squares Value I 0.993 0.0694 0.0696 1.85 -10- 4 Motor parameters for motor A. Manufacturer's Specifications Resistance, Rm (Q) Torque Constant, KT Back EMF Constant, K, ( 5.6.2 " 0 0.89 25 C and 1.31@155'C 0.07 ) 0.07 jj7.64 -10- 5 Least Squares Value 0.916 0.0716 0.0703 2.45 10-4 Motor parameters for motor B. Transient Response Motor Tests In this section, we will describe the transient response motor tests we carried out on motors A and B. We carried out these tests to determine the rotor inertias and viscous friction coefficients for these two motors. The theory underlying these tests was discussed in section 5.4. For these experimental tests, we first constructed an aluminum flywheel for the motors. Although the exact dimensions of this flywheel were not critical, we wanted the flywheel inertia to be much larger than the rotor inertia. The flywheel we used could essentially be modeled as two circular disks, and thus, the flywheel inertia was easily calculated using: 1 Jf = 2 mR 1 where Jf is the flywheel inertia, m, is the of the first circular disk, m 2 is the mass of the second circular disk. Figure 5.9 shows to attach the flywheel's inner circular disk + 1 2 M2R2 (5.16) mass of the first circular disk, R 1 is the radius the second circular disk, and R 2 is the radius of the constructed flywheel. Set screws were used to the motor shaft. The experimental set-up for these transient response tests was also fairly simple. For these - 75 - The Motor and The Motor Drive Alumainum Figure 5.9 Picture of the flywheel used. tests, the motor drive was connected to the motor as well as the DSP. We ensured later that the average motor current corresponded to the commanded motor current. The motor current was sensed using a current probe and displayed on the oscilloscope. The experimental set-up for these tests is shown in Fig. 5.10. The picture shows the set up when the flywheel was not attached to the motor shaft. DSP I/O ChannpeI Oscilioscope EMVD Motor Apparatus Drive Figure 5.10 Picture of the motor test set-up. A dSPACE Control Desk layout program was used to display pulses in commanded motor current and actual motor velocity (see Fig. E.1 in Appendix E). This model used variables from a Simulink file (see Fig. B.11 in Appendix B) that was used to build a DSP program which would command pulses in motor current and read motor velocity using the optical encoder on the motor. The experimental data was saved using dSPACE Control Desk software, as well as transferred from the oscilloscope via Lab View software (see Fig. F.1 in Appendix F for the Lab View file we used). - 76 - 5.6 Experiments to Obtain Motor Parameters For this experiment, the procedures were slightly different for motor A and motor B. Four experiments were carried out - the responses of each motor's speed to step changes in that motor's current were observed both with, and without, attaching the flywheel to the motor shaft. We decided to measure the time constants for both a positive current step and a negative current step, and thus used current pulses with 50% duty ratio. For motor A with the flywheel attached, the current pulses were ±1A each, with a 15s period at 50% duty ratio, while for motor A without the flywheel attached (but with the disk cam attached), the current pulses were ±1.25A each, with a 0.5s period at 50% duty ratio. For motor B with the flywheel attached, the current pulses were ±0.9A each, with a 15s period at 50% duty ratio, while for motor B without the flywheel attached, the current pulses were ±0.9A each, with a 0.35s period at 50% duty ratio. The current pulses did not have to be as large in amplitude for motor B because it did not have a disk cam attached to its shaft. Figures 5.11 and 5.12 show the experimental results we obtained for motor A. Current Waveforms for Motor A with flywheel attached 1 - ---- - -- --- -- - --- - - 0 4~ -- - - - Average Current-- Commanded Current C.) -1 -2 L0 10 5 15 Time (s) Motor Velocity for Motor A with flywheel attached 400 200 (U 0 S=1.325s --- -- d--- -200 p=1.117s: - - - -p - - u--- - - 10 0 15 Time (s) Figure 5.11 step response of motor A with the disk cam and flywheel attached. To make the experimental plots in this section, the program motortimetests.m in Appendix J was used. This program also numerically calculated the time constants that are displayed - 77 - The Motor and The Motor Drive Current Waveforms for Motor A without flywheel attached 0 -Average Current C omm anded Current .-.- -.-. .-.-.-.-.--.-.-.-.-.-.-.-.-. -.-.-.--.-.-.-.-. -. I 0.1 0.05 0 . . . .. . . . 0.2 0.15 -.. -. - 0.25 0.35 0.3 0.4 0.45 0.5 Time (s) Motor Velocity for Motor A without flywheel attached 4An.r00I 2001 (U 0 T a =61.90ms 57.16ms -200 - An 0 - - - 0.05 Figure 5.12 0.1 - - 0.15 - - - - - - - 0.25 0.2 0.3 - - 0.35 0.4 - 0.45 0.5 Time (s) Step response of motor A with the disk cam attached. on the graphs. It is important to note that the current pulse amplitudes shown in the plots were chosen by first observing whether the motor velocity was actually rising exponentially to some maximum in several test run experiments. Before analyzing the experimental data and calculating Jm and Bin, we had to calculate a few inertias. The outer diameter of the flywheel was 9.91cm, and its inner diameter was 2.59cm. Thus, for the flywheel in Fig. 5.9, the inertia, assuming the density of aluminum is 28004, was calculated as being 0.000024334kgrn 2 . For the disk cam, which was still attached to motor A when we carried out these motor tests, the inertia was numerically calculated (in SolidWorks) as being 0.728 - 10-6kgm 2 . It would have been very difficult to calculate the disk cam inertia analytically. Figures 5.13 and 5.14 show the experimental results we obtained for motor B. - 78 - Experiments to Obtain Motor Parameters 5.6 Current Waveforms for Motor B with ftywheel attached 2 a -- - - - - -- -61 Average Motor Current otor Currant Cc mmanded M -1 0 15 10 5 Time (a) Motor Velootty for Motor B with 300 0 .... . .... . ----.... 0 0 - - . .. - - - ----. -- 20 1 attached - - .. . -. - ---... --.-.-.-.-.----.-.-.....--..-. . -- - - -.-. - - - -- - - . --. - . . . . . .... ..> -200 ...-.. -000iml flywheel - - --.-- .-.-- I i 12 154 Time (s) Figure 5.13 Step response of motor B with the flywheel attached. Current Waveforms 0 0.1 0.0 for Motor B without flywheal attached 0.15 0.2 0.3 0.25 0.35 Time (a) Motor Vetooity for Motor S without -La -- - 3 00 - =2-- 3 va - r - - - - -- - ttywheat - attached - - -- 200 0 - - S 10 - - - - -100 -- - - - - -- - -- - - - - - -- - - m20:40mw - - -200 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 Time (a) Figure 5.14 Step response of motor B. From the experimental results in Figs. 5.11 through 5.14, we can observe slightly different time constants for the rising and falling motor velocity waveforms. The reason for this discrepancy was probably some unmodeled (and possibly nonlinear) friction in the motor shaft. To calculate the rotor inertia and viscous friction coefficients from this experimental data, we first averaged the time constants (since they were all almost equal) displayed on each graph and then solved equations (5.14) and (5.15) simultaneously for Jm and Bmn. We were careful to account for the inertias of the disk cam and flywheel in these calculations. Table 5.4 - 79 - The Motor and The Motor Drive shows the calculated rotor inertias and viscous friction coefficients versus the manufacturer's specifications. The program JBvalues.m in Appendix J was used to calculate these values. Parameter 2 Rotor Inertia, Jm (kgmn ) Viscous Friction, Bm (rm) Table 5.4 Manufacturer's Specifications Motor A Motor B 3.6- 10-6 7.64 - 10-5 5.185- 102.095- 10-4 4.441 10-6 2.177. 10-4 Motor inertia and viscous friction coefficient. The calculated viscous friction coefficients were lower than those obtained from the torquespeed curve tests, confirming that the dynamometer added friction to the system. However, these coefficients are almost 3 times larger than the manufacturer's specifications, implying that the manufacturer's specifications may have been incorrect, or that the motor was somehow misaligned during the experiment, thereby increasing the friction in the motor. The calculated rotor inertia for motor A is 1.4 times that specified by the manufacturer, while the rotor inertia for motor B is 1.2 times that specified by the manufacturer. Again, these results might have been obtained because of a misalignment in the motor test assembly, or because the manufacturer's specifications were incorrect. In the next section, we will briefly discuss the manner in which we measured motor inductance. We will then conclude this chapter with some comments on the motor testing experiments. 5.6.3 Inductance Measurements The motor inductances and inductance quality factors were measured using a National Instruments impedance analyzer in the laboratory. The inductances were measured at the motor terminals, and at frequencies up to 100kHz. The data from the experiment was plotted using the program motorinduct.m Appendix J. The resulting motor inductances and quality factors for motor A and motor B are shown in Fig. 5.15. The inductance at low frequencies corresponds very closely to the manufacturer's specifications. Thus, the motor drive circuit design was carried out with an acceptable value for motor inductance. - 80 - 5.6 Experiments to Obtain Motor Parameters Motor Inductance [ i -oo 100- A -- 80 60 40 20 0 10 20 Figure 5.15 5.6.4 30 40 Frsqoeocy (kHz) 0 s0 70 80 Inductances for motors A and B. Conclusions Table 5.5 shows a summary of the motor parameters for both motors A and B obtained with the experimental tests described in the previous three sections. From the table, we can see that, in general, the characteristics of motor derived from the experimental data are very close to the manufacturer's specifications. Parameter Manufacturer's Specifications Motor A Motor B Resistance, Rm (Q) Torque Constant, KT (m) 0.89@250 0.07 0.993 0.0694 0.916 0.0716 0.07 0.0696 0.0703 7.64 -10-5 2.095. 10-4 2.177- 10-4 100 3.6 - 10-g_ 120 5.185- 10-6 95 4.441 _10-6 ) Back EMF Constant, K ( Viscous Friction, Bm ( ) Inductance, Lm (ILH) Inertia, Jm (kgm'2 ) Table 5.5 Summary of the motor parameters. In this chapter, we discussed the design, construction, and testing of the motor drive inverter circuit for the EMVD apparatus. We also described experimental tests that were performed to extract motor parameters. In the next chapter, we will use some of these motor parameters to design controllers for the MIT EMVD. - 81 - Chapter 6 Controller Design and Experimental Results 6.1 Introduction IN this chapter, the penultimate of this thesis report, we will discuss controller design for the EMVD, as well as experimental results obtained using these controllers. Some of the motor parameters obtained in the previous chapter were used for this design. Our discussion herein will also include system identification experiments to extract parameters for the EMVD plant, especially for the engine valve-spring system in the z domain. We needed these parameters to design the controllers. Furthermore, we will discuss experiments that were done to determine the actual nonlinear relation between the z and 0 domains, and experiments that were performed to measure the engine valve's seating velocity. We will begin with an overview of the controller design for the MIT EMVD. This discussion will depend strongly upon that in Chapter 3. The need for system identification will then be motivated. We will follow this discussion with some EMVD plant modeling, and a derivation of the theory underlying the system identification experiments. Then, we will describe the system identification experiments we performed, and follow this description with details on the controllers that were designed. The implementation of these controllers using the experimental test stand and the results from these experiments will then be discussed. Finally, we will conclude this chapter with details on the design of a robust adaptive controller for the MIT EMVD. 6.2 Overview of Controller Design In this section, we give an overview of controller design similar to the discussion in Chapter 3. We will also motivate the need for system identification experiments. A block diagram of the feedback-controlled EMVD apparatus is shown in Fig. 6.1. The - 83 - Controller Design and Experimental Results reference input is the desired valve position, and the system output is the actual valve position. In the experimental test stand in the laboratory, we implemented the controllers in the 9 domain, and thus the reference input was the desired motor position, and the system output was actual motor position. Since we knew (approximately) the relation between the motions in the 9 and z domains, an appropriate reference input could easily be generated for motion in the 9 domain. In order to ensure our assumed z - 9 NTF characteristic was correct, we had to perform experiments to determine the amount of compliance between the z and 9 domain motions. Current Reference aet" Input Figure 6.1 Motor VlePsto L[orrie otele (otor. EoDjPan NTF. Velne-Spring Systenm) The EMVD as a feedback control system. The difference between the actual motor position and the desired motor position is passed into a controller (implemented using the dSPACE DSP), which provides a current control input to the motor drive. The motor drive then supplies the desired current to the motor. In the previous chapter, we verified experimentally that this motor drive could be modeled as a linear gain (between the desired motor current and the actual motor current) and a high-frequency (200kHz) disturbance representing the ripple current in the motor drive. As we mentioned earlier, a linear control law, such as a fixed-gain PD controller, is not well-suited to the control of the MIT EMVD. Fixed-gain controllers cannot account for the changing dynamic characteristics of the MIT EMVD during the valve stroke. For instance, at the ends of the stroke, the effective inertia in the 9 domain is small, while at the midpoint of the stroke, this effective inertia is large. Thus, in the 9 domain, the effective system gain of the valve-spring system increases at the ends of the stroke and decreases at the middle of the stroke. Nonetheless, on our first attempt at designing controllers, which is described in this chapter, we used linear controllers to control the MIT EMVD - making sure that these controllers can control the EMVD at both the middle and the ends of the stroke. As we shall see, although not optimal, these controllers can perform reasonably well. In terms of controller performance, we noted earlier that it is important to be able to minimize errors when the valve is almost open or almost closed, such that the valve reaches these positions with small velocity. The errors as the valve transitions from one end of the stroke to the other are not as important. Furthermore, the controller must be able to track the desired motor angular position trajectory even in the presence of parameter uncertainties and gas force disturbances. In the current EMVD apparatus, there is no physical gas force - 84 - 6.3 Modeling the EMVD Plant simulator, however, such a simulator will be incorporated into the apparatus in the near future by the MIT EMVD project team. We planned to implement our controllers on the MIT EMVD with the soft set of springs. Once these controllers were refined, they would be implemented with the stiffer set of springs so as to gain feedback control with smaller transition times. However, the implementation of these controllers with stiffer springs and the final evaluation of the MIT EMVD will not be described here - these topics will be described in detail in Woo Sok Chang's doctoral thesis. There were three modes of engine valve motion in the MIT EMVD for which we had to design controllers: initial mode; holding mode; and transition mode. The initial mode controller moves the engine valve from its resting position at the middle of the stroke to an extreme end of the stroke. The holding mode controller is used to hold the valve at either end of the stroke (opened or closed) with a variable holding time. The transition mode controller is used to smoothly and quickly move the valve from one end of the stroke to the other end. The transition mode controller also minimizes error during valve transition such that the valve is closed with a small seating velocity. For simplicity, we designed and implemented a single transition mode controller to control valve motion in each of these modes, and then generate a refined motor position reference input that allowed this transition mode controller to carry out initial and holding mode control. The two types of transition mode controllers we designed, implemented, and compared were a PD controller and a lead compensator. In fact, we implemented several versions of these two types of controllers. In addition, we designed a robust adaptive controller, whose implementation will be discussed in section 6.11.5. Before designing any controllers, we had to extract parameters for the EMVD plant. The motor parameters were already experimentally determined (as described in the previous chapter), but z domain parameters for the MIT EMVD, such as mass, friction, and spring stiffness, were not exactly known. Hence, we first performed some system identification experiments to extract these parameters. The theory underlying these experiments will be discussed in the next section, where we turn to the modeling of the EMVD plant. 6.3 Modeling the EMVD Plant In this section, we will describe the modeling of the EMVD plant, and explain the theory underlying the system identification experiments, which we will discuss in the next section. - 85 - Controller Design and Experimental Results A large part of this modeling was discussed in Chapter 3, but will be repeated here for completeness. The equations of motion for the MIT EMVD are as follows [8]: fz = mrz + Bz' + Kzz (6.1) Jo + Bo +,To (6.2) = KT im where To is the transformer torque in the 9 domain, fz is the transformer force in the z domain, JO is the inertia in the 9 domain, mz is the mass in the z domain, BO is the friction in the 9 domain, Bz is the friction in the z domain, KT is the motor torque constant, im is the motor current, Kz is the effective z domain spring constant, 0 is the displacement in the rotational domain, and z is the displacement in the vertical domain. Equations (6.1) and (6.2) can be combined using the NTF characteristic relations (see Chapter 3 to obtain a single second-order nonlinear time-varying differential equation of motion in the 0 domain: SJo+ mz( dz + mz-d + (Bo + Bz () dz 19±$Kz f()d6 Krijm (6.3) If we are to assume that a time-varying gas force disturbance also acts on the valve (typical of exhaust valves in an internal combustion engine), then the equation of motion becomes: d Jo+mz d , 2 tdz~2 + + Zdzjj dz (Bo+Bz ( + z 0 dz d+ = ~m+ g(t) (6-4) where g(t) is the gas force in the z-domain reflected to the 0 domain through the NTF. Since the nominal values of Jo, mz, KT, Bz, Bo, d, d-0, and g(t) in (6.4) are either known (or can be determined experimentally), bounded, or both known and bounded, it is possible to design and implement various types of controllers for the EMVD to track the desired motor position trajectory, Od(t). As was given earlier, the NTF relation we implemented in the disk cam slot was: z= f () = 0.004 sin(3.469) m. sin(O.9997r/2) (6.5) For now, let us assume there is no gas force acting on the engine valve. Then, by using this relation in equation (6.4), we can observe that at the ends of the stroke (9 = ±0.456 rad) where the motor is essentially decoupled from the engine valve-spring system (because the - 86 - 6.4 System Identification Experiments slope of the relation (6.5) is approximately 0), the EMVD equation of motion reduces to: Jo0+ Bo = KTim, (6.6) while in the middle of the stroke (0 = 0 rad), assuming the slope of the NTF relation (6.5) is equal to r, the EMVD equation of motion reduces to: (JO + mzr 2 )0 + (Bo + Bzr 2 )0 + Kzr 20 = KTim. (6.7) For the NTF relation (6.5) we implemented, the slope at the point 0=0 rad is 0.0136 m/rad. By taking the Laplace Transform of equations (6.6) and (6.7), we obtain linearized transfer functions for the EMVD plant at the ends of the stroke and at the middle of the stroke respectively: _ (s) KT( Gend-of-stroke(s) = im(s) J0 82 + Bos (6.8) Gmd~stok mid-stroke (S) - 0s KT. im(s) (Jo + mzr 2 )s 2 + (Bo + Bzr 2 )s + Kzr2 (69 (6.9) where r=13.6mm. The linearized transfer function in (6.8) is only valid for small displacements near the open or closed engine valve positions, while that of (6.9) is only valid for small displacements about the middle of the stroke. In the context of the EMVD apparatus, where the full stroke is approximately 0.91 rad, small displacements are considered to be displacements less than 0.1 rad. The two linearized transfer functions we just derived were the basis for system identification experiments described in Section 6.4.2. With these experiments, we wanted to determine values for the EMVD plant parameters (such as Kz, m,, and so on). As with any system identification experiments, the system identification is only as accurate as the model used, however, we were fairly confident of the accuracy of these two transfer functions. In the next section, we will describe these system identification experiments. 6.4 System Identification Experiments In this section, we will describe experiments that were carried out to extract parameters for the EMVD plant. There were two experiments that were carried out: free oscillation experiments to characterize the z domain engine valve-spring system; and open loop transfer function experiments to match the experimentally obtained EMVD parameters to the - 87 - Controller Design and Experimental Results mathematical models in (6.8) and (6.9). In the next two subsections, we will describe the results from each experiment. The free oscillation experiments were carried out to determine the quality factor, Q, and the approximate damped natural frequency, Wd for both the engine valve-spring system and the entire EMVD plant. In addition, we wanted to know the damping ratio for the engine valve-spring system in the z domain, so that we could approximate the losses in the system as the valve moves from one end of the stroke to the other end (transition mode losses). We carried out these experiments with both of the valve assemblies we had - with the set of soft springs and stiffer springs. It is important to note that the higher the Q for the EMVD plant, the lower the losses and the less the strain on the motor driving the engine valve-spring system. Thus, with the free oscillation experiments, we hoped to confirm that the EMVD apparatus had been soundly assembled and had high enough Q to proceed with more experiments. In addition, we planned to use the experimental value of wd to generate a reference input for the feedbackcontrolled EMVD. We carried out the open loop transfer function experiments to obtain two linearized transfer functions on which to base our controller design. We especially needed to know the EMVD plant parameters (such as friction in the z domain system) that could not be easily estimated on paper. In the next subsection, we will describe the free oscillation experiments, and in the following subsection, we will describe the open loop transfer function experiments. 6.4.1 Free Oscillation Experiments In this section, we will describe the free oscillation experiments. These experiments were performed with two valve assemblies: one with a set of stiffer springs (effective stiffness=3201b/in) and one with a set of soft springs (effective stiffness=28.71b/in). The eventual goal for this project is to use the set of stiffest springs (effective stiffness=8001b/in) we have purchased to provide the necessary valve transition times characteristic at higher engine speeds. Our first experiment was to characterize the free oscillation dynamics of the valve assembly after releasing the valve from one end of the stroke. At first we obtained data without lubricating the apparatus, but the damping in the valve assembly was very large, leading to higher losses. After lubricating the valve stem (with 3-in-1 and WD-40), these losses were much lower. Figure 6.2 and Table 6.1 show the data from this experiment with associated - 88 - System Identification Experiments 6.4 values of quality factor (Q), damped natural frequency (wd) and damping ratio ((). Tek un 1 Trig? Thk.prevu T iV a K Soft Spring ____________~ 5~2 IV=1.06mm M140.Om '. A ChS- . 32Wm 00 V Figure 6.2 - .6m t 4 Free oscillation of valve assembly with the soft and stiffer spring sets. Table 6.1 Soft Spring Stiffer Spring Q 13.33 15.63 Wd 51Hz 185Hz ( 0.0332 0.0375 Parameters obtained for the valve assemblies. The second experiment we carried out was to characterize the same free oscillation dynamics, but this time, the motor assembly coupled to the valve assembly (the EMVD plant), without any electrical input from the motor. The valve was again moved to one end of the stroke and released. Again, the disk cam surface, roller and valve stem were lubricated before performing the experiment. Figure 6.3 and Table 6.2 show the results from this experiment with associated values of Q, Wd, and (. Q Wd ( Table 6.2 Soft Spring Stiffer Spring 2.9 38Hz 0.174 6.0 105Hz 0.084 Parameters obtained for EMVD with no electrical input. As expected, Q and wd were higher with the stiffer spring set than with the soft spring set. As the losses were not too high (approximately 27% with the stiffer spring set) and the quality factors were reasonable, we concluded that the EMVD apparatus had high enough Q and low enough damping that it would be feasible to continue with further experimentation. - 89 - Controller Design and Experimental Results T T.~~~~kT T RC T~ \At t r~ bRuh i 6. Trig? 4:4. -- --- - - - _ _ ---- J - - - - - .... ..... Soft Spring 1V=1.06mm Stiff Spring 1V=1.06mm j .Sj-M0 M[20.Gm~~. f Figure 6.3 6.4.2 Ar khN , Ch-3 r -1.20 V Free oscillation of EMVD with no electrical input. Open Loop Transfer Function Experiments In this section, we will describe the open loop transfer function experiments that were carried out to extract EMVD plant parameters. The first experiment was carried out on the motor assembly only, and was used to determine the approximate transfer function of the EMVD plant at the ends of the stroke. The second experiment was carried out on the entire EMVD plant, and was used to determine the approximate transfer function of the EMVD plant at the middle of the stroke. For these system identification experiments, only the valve assembly with the set of soft springs was used. The experimental set-up for the first experiment was as follows: the motor drive was connected to the DSP and the motor. A dSPACE Control Desk file was used to command and display the actual motor current (measured using the current probe) and to display the actual motor position. A sinusoidal current command was input to the motor drive. The frequency of this sinusoidal current command was increased (using the dSPACE Control Desk file) in increments of a few Hz, and the amplitude of this commanded current was adjusted such that motor displacement amplitude was less than 0.1 rad about the middle of the stroke. The amplitudes of the motor current and motor position were recorded. The phase difference between these two signals was also recorded. The experimental data was processed and plotted using the MATLAB program openlooptests.m in Appendix L. The experimental data was curve fit to a transfer function of the form given in (6.8). In this transfer function, the value of KT was assumed to be 0. 0699. This value was determined (for motor A) in the motor testing experiments described in the previous chapter. - 90 - System Identification Experiments 6.4 Figure 6.4 and Table 6.3 show the experimental results obtained - the numerical values in the table were obtained by comparison to the model in (6.8). The discrepancy between the experimental data and the curve fit is detailed below. Equation (6.10) below shows the curve fitted transfer function that was found: Goend-of-stroke (S) = (s) _ im(s) 0.069 4.87 - 10-6s2 + 0.0099s (6.10) Open Loop Response: Motor and Motor Drive 30 0 10E r a' a nta -30eFi -20 - - --- 10 - - -* - -- - - 10 102 10' Frequency (radians) Figure 6.4 Linearized transfer function of the EMVD plant at the ends of the stroke. Parameter Experimental Value Expected/Calculated Value Jmotor J_ 4.15- 10- 6 kgm 2 Assumed to be ~ 0.728 - 10- 6 kgm 2 4.87 -10- 6 kgm 2 3.5- 10- 6 kgm2 0.728. 10- 6 kgm 2 4.22 -10- 6 kgm2 KT 0.069r 0.07 m 0.0099 Nm 7.64 -10 Jdisk cam B9 a______rad/s Table 6.3 76.i-radNm Parameters obtained from curve fitting to experimental data. As mentioned earlier, at the ends of the stroke, the engine-valve spring system can be assumed to be decoupled from the motor, and thus the parameters in Table 6.3 are essentially motor parameters. As we found in the previous chapter, both the viscous friction coefficient and the rotor inertia obtained here were larger than we expected. One probable reason for these results is that there was some misalignment in the motor assembly. However, it is also possible, as mentioned in the previous chapter, that the motor we were using was damaged during assembly causing excessive friction in the internal motor bearings. Another prob- 91 - Controller Design and Experimental Results lem we had while carrying out the experiment was motor drift - at certain frequencies of input current it was difficult to maintain a zero-centered sinusoidal motor position - which could have made some of the experimental data, especially that at the lower frequencies, questionable. The experimental set-up for the second open loop transfer function experiment was essentially the same as the first experiment described above, except that the roller-follower was placed in the disk cam slot and connected to the valve assembly. As we did in the first experiment, the motor drive was connected to the DSP and the motor, and a dSPACE Control Desk file was used to command and display the actual motor current (measured using the current probe) and to display the actual motor position. Again, a sinusoidal current command was input to the motor drive, and its frequency was increased incrementally while the amplitude of motor displacement was adjusted such that it remained less than or equal to 0.1 rad about the mid-stroke valve position. The amplitudes of the motor current and motor position were recorded. The phase difference between these two signals was also recorded. The experimental data was processed and plotted using the same MATLAB program openlooptests.m in Appendix L that was used for the data from the first experiment. This experimental data was curve fit to a transfer function of the form given in (6.9). Again, in this transfer function, the value of KT was assumed to be 0.069N. Figure 6.5 and Table 6.4 show the experimental results obtained - the numerical values in the table were obtained by comparison to the model in (6.9). Again, the discrepancy between the experimental data and the curve fit is detailed below. Equation (6.11) below shows the curve fitted transfer function that was found: mid-stroke( _ 0(s) im(S) 0.069 _ 2.26 . 10-5S2 + 0.0029s + 1.0935 The extracted parameters do, in general, match the expected values, except for the lower frequencies (w < 200 rad/s). The reasons for the discrepancies in rotor inertia and motor viscous friction coefficient have been discussed above. The actual mass in the z domain closely matched the expected mass, while the friction in the z domain was about 1.6 times larger than expected. Furthermore, the spring constant in the z domain, Kz, was much smaller than that expected. One possible reason for these results is that the disk cam had been worn out when this experiment was performed, causing the deterioration of the desired NTF relation (6.5). Another probable reason is the misalignment of either the valve or motor assemblies, or both assemblies. However, it is also possible that the motor we were using was damaged during assembly causing excessive friction in the internal motor - 92 System Identification Experiments 6.4 Open Loop Response: EMVD at the middle of the stroke . -.-.- -20 ......... . - -- -25 I -35 -..-...- Curve-FiC.... -.-.-.-.-..-. .-.- -40 10' 100 W 102 Frequency (radians) Figure 6.5 Linearized transfer function of the entire EMVD. Parameter Experimental Value Jmotor Jo 4.15. 10-6kgm2 Assumed to be 4.87. 10 6 kgm 2 KT 0.069m BO 9 mz Bz K1Z 95.9g 37.8M 1.182 10ii Jdisk cam Table 6.4 bearings. Expected/Calculated Value 6 2 0.728. 10- kgm 3.5- 10- 6 kgm2 0.72. 10- 6 kgm2 4.22- 10- 6 kgm 2 0.07' 7.64 -10-5 .9 - 1 0-3 90g 20 1.653. 10i EMVD parameters obtained from curve fitting to experimental data. As was the case with the first experiment, the motor drift made some of the experimental data, especially that at the lower frequencies, questionable. In the next section, we will begin discussing the controller designs for the MIT EMVD, which were heavily based on these system identification experiments. We will begin with the initial mode and holding mode controllers, and follow this discussion with the design of the transition mode controllers. As was mentioned earlier, all the controller design discussed in this thesis report was done assuming that the valve assembly with soft springs was assembled in the EMVD apparatus. - 93 - Controller Design and Experimental Results 6.5 Initial Mode Control In this section, we will briefly discuss the initial mode controller that was implemented in the experimental EMVD test stand. When the EMVD is at rest, the engine valve is at the 0 = 0 position. The objective of the initial mode controller is to move the valve from this resting position to one end of the stroke (0 =- 0.46 rad) in as short a time period as possible. For the first design of the initial mode controller, a sinusoidal motor position reference input with linearly increasing amplitude was created in Simulink. The frequency of this sinusoid, 24Hz (corresponding to 20ms transition times), was designed to be close to the damped natural frequency of the EMVD plant that was determined experimentally (see Section 6.4.1). In effect, the EMVD's initial mode does not require its own controller, but only a welldesigned motor position reference input. Using this reference input, any well-designed transition mode controller will apply pulses of current into the motor (because the controller would try to apply large amplitude sinusoidal pulses that would be clipped by a currentlimiter block in the Simulink file), causing the disk cam to oscillate sinusoidally at a linearly increasing amplitude, thereby enabling initial mode control. Optimally, for the fastest EMVD start-up transients, this current would be applied at the exact moment when the motor velocity changes sign, such that the number of cycles required to increase the motor position amplitude to full-stroke amplitude was minimized. The need for a fast start-up transient is particularly important when using very stiff springs, because the initialization time must be much smaller than the thermal time constant of the uncooled motor. Such an initial mode controller will be implemented in the near future, and discussed in Woo Sok Chang's doctoral thesis. 6.6 Holding Mode Control In this section, we will discuss the holding mode controller that was implemented in the experimental EMVD test stand. The objective of the holding mode controller is to hold the engine valve at either the open or the closed position for a variable time period. As is the case with conventional IC engine valves, electromagnetically-actuated engine valves, when implemented in real engines, will be held closed or open most of the time. In addition, one of the key features of variable valve timing is to be able to easily vary the time the valve is held open or closed. Thus, it is important to have a good holding mode controller for the MIT EMVD. - 94 - 6.7 Reference Input Generation The holding mode controller was implemented in a Simulink file by using a "latch" block together with the sinusoidal reference input that was used for initial mode control, such that this reference input was latched at one end of the stroke for a fixed (but variable in Control Desk software) period of time (the "holding" time) and then allowed to transition sinusoidally from this end of the stroke to the other end (in the fixed "transition" time), before being held again at the other end of the stroke. Thus, holding mode control also does not require its own controller, but only a wellgenerated motor position reference input. In this respect, using a sinusoidal reference input (which is latched/held at one extreme amplitude during the holding period), any welldesigned transition mode controller would track this reference input with small position errors, thus enabling holding mode control. 6.7 Reference Input Generation The Simulink file used to generate the reference input that was used to implement both the initial and the holding mode control described in the previous two sections appears in Fig. B.2 in Appendix B. This reference input was the same reference input used for the transition mode controllers. As mentioned above, in effect, we designed only transition mode controllers, and relied on the carefully generated position reference input to obtain both initial and holding mode control for the EMVD using these transition mode controllers. In the next section, an experiment to determine the nonlinear transformer relation between the z and 0 domains will be described. Following this description, we will begin discussing the design of the transition mode controllers we implemented. 6.8 A Check on the NTF Characteristic Relation In this section, we briefly discuss an experiment to determine the actual relation between the z and the 0 domains. We needed to determine this value for two reasons: first, to be able to check that the NTF relation (6.5) discussed in Chapter 4 was implemented correctly in the disk cam slot; second, to be able to measure seating velocity with the rotary position sensor we needed to have an accurately measured relationship between displacements in the linear and rotary domains. - 95 - ControllerDesign and Experimental Results To carry out this experiment in the experimental EMVD test stand, the Simulink model in Fig. B.12 in Appendix B was implemented on the DSP. The DSP was connected to the motor drive, which was in turn connected to the motor. The motor assembly was connected to the valve assembly using the roller-follower. The outputs of both the linear position sensor and the rotary optical encoder were displayed on the oscilloscope. The valve was then made to transition (under feedback control) from one end of the stroke to the other, and the data from both the linear position sensor and the rotary optical encoder was captured on the oscilloscope and transferred to the PC using Lab View (see the Lab View file used in Appendix F). The lead compensator, whose design will be discussed in Section 6.9.2, was used to control the EMVD. Figure 6.6 is a plot of the actual z-0 relation (thick line in the plot) and the expected z-0 relation (thin line in the plot), which was obtained by evaluating the NTF relation (6.5), versus actual motor position (obtained from the optical encoder). The figure was plotted using the MATLAB program emvdseatvel.m in Appendix L for an open-to-closed valve transition. A 4 3 -0.5 Figure 6.6 comparison of the theoretical and experimental NTF characteristic rolation ( - -0.4 -0.3 .. . .. . .. . . . .. . . . .. . .. . t -0.2 . . .. . . . .. . a)relatn 0 -0.1 01 02 03 04 0. A comparison of the theoretical and experimental NTF characteristic relation. From Fig. 6.6, we can see that there is very little difference between the actual and expected z -0 relations - less than 200pam - much smaller than the machining tolerances at the MIT Central Machine Shop. One possible reason for this difference is that the surfaces of the disk cam slots were worn when the experiment was carried out. Nonetheless, this experimental result shows that the NTF relation (6.5) was implemented very accurately in the disk cam slot. In addition, the result shows that one could use the rotary optical encoder together with the NTF relation (6.5) to measure engine valve seating velocity. - 96 - 6.9 6.9 Transition Mode Controllers Transition Mode Controllers In the rest of this chapter, we will discuss the design and implementation of transition mode controllers for the MIT EMVD. As we mentioned earlier, we only designed transition mode controllers for the MIT EMVD and relied on the carefully generated position reference input to obtain both initial and holding mode control. We will begin by describing a PD compensator that was used for the first few experiments with the EMVD apparatus. We will follow this discussion, with a description of the design of a better lead compensator that is still in use in the experimental EMVD test stand. After discussing controller design, we will discuss experimental results obtained using these two types of controllers. 6.9.1 The Initial Attempt: a PD Compensator We will describe the design of a PD compensator for the MIT EMVD in this section. In the next section, we will discuss the design of a lead compensator for this EMVD. Considering the open loop transfer functions that were obtained experimentally, we designed a PD compensator for the EMVD plant. We decided to use a PD compensator for two reasons: first, PD compensators are easy to design using well-known classical control techniques; second, at the ends of the stroke, the EMVD plant's linearized transfer function resembles that of a dc motor, having an integrator to eliminate steady-state errors in position, thus we did not need to use an integrator in our controller. PD compensators can dramatically improve the transient response of a system, however, due to the derivative term in their transfer function, they are usually characterized by highfrequency noise when the reference input is changing rapidly. In addition, PD compensators require active circuits for physical implementation. There were three constraints on the design of the PD compensator: we wanted to have minimal (if any) overshoot in the feedback-controlled EMVD's transient response; we wanted the controller to have fairly high bandwidth (at least 300Hz for the system with soft springs); we wanted the controller to perform well even with a current-limiter on its output'. In general, we do not want to use a current-limiter during the transition mode, however, we used such a limiter in our experiments in case our controllers did not work as intended. 'We needed this current-limiter to prevent the motor from reaching its thermal limit. - 97 - ControllerDesign and Experimental Results The transfer function for a PD compensator, GPD(s), is given by: GPD(s) = Kp (1 + KDs) Kp (6.12) where Kp is the controller gain, and -KD is the controller's zero location. These two constants are chosen to obtain the desired transient response for the feedback-controlled plant. The design of our PD compensator was straightforward. The design of such compensators is fairly basic knowledge in classical control theory, and most undergraduate control theory textbooks can serve as a reference for this design. The process was essentially empirical after carefully considering the desired transient response of the feedback-controlled EMVD, we made intial guesses for the controller variables. We chose the zero location for the PD compensator to be at approximately 45Hz (or 283rad/s), and we chose the controller gain to be approximately 250 to contribute to the required loop-gain value. These values were chosen by bearing in mind the EMVD characteristics at both the middle and the ends of the stroke. This PD compensator was implemented in a Simulink model (see Fig. B.3 in Appendix B for this model), and run on the DSP. Using a dSPACE Control Desk model (see Fig. E.2 in Appendix E for this model), we were able to vary the chosen zero location and controller gain in real-time until we were satisfied with the feedback-controlled EMVD response. The PD compensator that gave the best feedback-controlled performance is given by the following transfer function: GPD(s) = 313.2(1 + 0.003455s) , (6.13) which corresponds to a controller gain of 313.2 and a zero location of 46Hz (or 290rad/s). The experimental results obtained using this controller will be discussed later in this chapter. In the next section, we will discuss the design of a lead compensator for the MIT EMVD. 6.9.2 Lead Compensator Design In this section, we will describe the design of a lead compensator for the MIT EMVD. In the next section, we will discuss the experimental results obtained using both the PD compensator from the previous section and the lead compensator from this section. In effect, a lead compensator is an approximation to a PD compensator. In general, PD compensators have to be implemented with either rate feedback, or with an additional pole that has greater magnitude than the PD compensator zero. The latter implementation is - 98 - Transition Mode Controllers 6.9 a lead compensator. Lead compensators can be implemented with passive circuits, and do not suffer from high-frequency noise because they do not have a differentiator. The constraints on the design of the lead compensator were the same as those on the PD compensator (see the previous section). For a typical lead compensator, the transfer function, GLEAD(s), is given by: GLEAD(s) = K 1+ s 1 ZLEAD 1+ PLEAD (6.14) s where K is the controller gain, ZLEAD is the lead compensator zero location, and PLEAD is the lead compensator pole location. These three constants are chosen to obtain the desired transient response for the feedback-controlled EMVD plant. The design of the lead compensator was fairly simple, and is also fairly basic knowledge in classical control theory. To obtain excellent feedback-controlled responses at both the middle and the ends of the stroke, we designed a lead compensator using MATLAB's ritool controller design GUI. We carefully considered the dynamics of the EMVD plant we had obtained from the open loop transfer function experiments. Bode Digram a 5 . . . . .. .. .. . .. .. . 60 . .. . .. .. ...... .. ... . .. . .... . 55 . ............... . .... . for the Lead Compensator a .... . ... . . .. .. ... .. . .. .. . .. .. ... . ... .. .. .. .. .. . .. . .. .. .. . .. .. .. . ... . .. ... .. . .. . ... ....... ............ ....... .. ... .. . .. . . .. .. . . .. .. . q j30 .. .. . .. .. .. . .. . .. . .. .. .. .. . .. . .... . ... ......... ......... U 1e 1e Frequency (rad/ec) Figure 6.7 Bode diagram for the designed lead compensator. Figure 6.7 shows the Bode diagram for the lead compensator we designed. The MATLAB program openlooptests.m in Appendix L was used to plot this figure. For the initial design of the lead compensator, we set K=400, ZLEAD ~ 272Hz, and PLEAD ~ 1234Hz, so as - 99 - Controller Design and Experimental Results to obtain the desired feedback-controlled response. Using rltool was particularly useful in quickly comparing lead compensators with different gains and zero/pole locations. Figure 6.8 shows the root locus for the lead compensated EMVD at the middle of the stroke, while Fig. 6.9 shows the same root locus at the ends of the stroke. These root loci were obtained using the linearized transfer functions from Section 6.4.2, and were plotted using MATLAB's ritool GUI. RooI L . le ftr In L..dCo."M&.BM 80 M~d-Soo .....I ...... 0.5 -= -7000 0 -0 - -4000 -20 -1 Root locus for the compensated EMVD (middle of the stroke). Figure 6.8 Rool L ON #o0.Lem Cw" em - at 1*5 &49.of #0 so P.00old W 4000 20M I .......... 0 -2000 .... - 4MM -9000 Figure 6.9 4000 -7000 -000 5000 -4000 -30W0 -200M -1000 Root locus for the compensated EMVD (ends of the stoke). These root loci show that the compensated EMVD dynamics at both the middle and the ends of the stroke are dominated by faster closed loop poles, indicating a better transient response for the feedback-controlled EMVD. - 100 - Transition Mode Controllers 6.9 Recall from Section 6.4.2 that the linearized transfer functions of the EMVD plant at the middle and ends of the stroke are given by: 0(s) Gmid-stroke(s) _0.069 ()- - .6 im(s) 2.26 -10-5s2 + 0.0029s + 1.0935 ' (6.15) and, = im(s) 0(s) Gend-of-stroke~) _ 0.069+ 0.0099s 4.87 -10-6s2 (6.16) Using these transfer functions with the designed lead compensator given by: GLEAD(S) = (6-17) 4001 ± 1 + 1.29 - 10-4s we can make Bode diagrams for the uncompensated and lead compensated EMVD plant at both the middle and the ends of the stroke. Figure 6.10 shows these Bode diagrams at the ends of the stroke, while Fig. 6.11 shows these Bode diagrams at the middle of the stroke. Bode Diagrams for the Lead Compensated EMVD at the Ends of the Stroke 100"7 50 - - - 0 : - : ::::..-. Plant Loop Gain -Plant*Controller Loop Gain Closed-Lop System - - - - - -50. 10 10 * 10 10 10 10 Frequency (rad/aec) Figure 6.10 Bode diagrams for the compensated EMVD (ends of the stoke). These Bode diagrams show that the closed-loop EMVD plant acts as a "perfect" low-pass filter at both the middle and the ends of the stroke. The bandwidth of this filter is high (at least 450Hz), and the transient responses seen in simulation were extremely desirable. The lead compensator that was designed was implemented in a Simulink model (see Fig. B.9 in Appendix B for this model), and run on the DSP. Using a dSPACE Control Desk model - 101 - Controller Design and Experimental Results Bode Diagrams for toh Lead Compensated + 5.8 -50~ 10 EMVD at the Middle of the Stroke .. s ......- 110 4 102 5 10 Frequency (rad/e) Figure 6.11 Bode diagrams for the compensated EMVD (middle of the stoke). (see Fig. E.3 in Appendix E for this model), we were able to vary the chosen zero/pole locations and the controller gain until we were satisfied with the feedback-controlled EMVD response. The lead compensator that gave the best feedback-controlled performance is given by the following transfer function: GLEAD(s) = 3501... 1 + 1.22.- 1-4s ( (.8 which corresponds to a controller gain of 350, a zero location of 270Hz, and a pole location of 1305Hz. At this point, it is imperative that we make some general comments on the controller design methods we have used. The linear controllers we designed were based on linearized transfer functions for the EMVD plant that were obtained from the open loop transfer function experiments described in Section 6.4.2. Although the dynamics of the EMVD plant do change between the middle and end of the stroke, the change is not great enough to affect the linear controller performance, at least while using the set of soft springs - this observation was made in the laboratory. It should be clear that a better control method would be to find linearized transfer functions for the EMVD at several points of the engine valve stroke and then design different controllers for each point - indeed, such controllers will be implemented and described in Woo Sok Chang's doctoral thesis. The experimental results obtained using both the PD and lead compensators will be discussed in the next section. -- 102 - 6.10 6.10 Linear Controller Implementation Linear Controller Implementation In this section, experimental results obtained using the PD and lead compensators will be presented. We will begin with an overview of the section, followed by detailed discussions of the feedback-controlled EMVD's performance with these two compensators. In the experimental EMVD test stand, two controllers were implemented: the PD compensator described in Section 6.9.1 and the lead compensator described in Section 6.9.2. For each compensator the reference input was a 24Hz sinusoid (corresponding to a 20.8ms transition time or an effective 1200rpm IC engine speed), generated using the method described in Section 6.7. In addition to a feedback controller, +/ - 2.25A pulses of (feed-forward) current were injected at the start of the valve transitions from the open-to-closed or closedto-open positions respectively. No current was injected at the ends of the stroke because we observed that current injection did not significantly speed up the valve transition time. Furthermore, the engine valve was held for 0.25s at each end of the stroke. The performance of these controllers was measured in the laboratory, and the two controllers were compared. In each case, the experimental set-up was simple. The dSPACE DSP was connected to the motor drive, and the motor drive was connected to the motor in the EMVD apparatus. The output of the motor drive was limited to 8A (in the Simulink model for each controller). The rotary optical encoder and linear position sensor were also connected to the DSP. A Simulink model was constructed for each controller, and then compiled and run on the DSP. A Control Desk file was used to monitor key experimental variables in real-time. The motor current and voltage were measured and displayed on the oscilloscope (at a 2.5GHz sampling rate). In addition, the motor position and position error were output to the oscilloscope from the DSP, such that all the data could be viewed synchronously on the oscilloscope. In the next three sections, we will describe the experimental results obtained using the two types of compensators. 6.10.1 PD Compensator In this section, we will present the experimental results obtained using the PD compensator to control the EMVD apparatus. Figure E.2 in Appendix E contains the dSPACE Control Desk file used to view the PD controller performance in real-time, while Fig. B.4 in Appendix B contains the Simulink model used to implement this PD Controller with current injection. Figures B.8, B.5, B.6, B.7, and B.8 in Appendix B contain other Simulink models that relate to the PD controller. - 103 - ControllerDesign and Experimental Results The MATLAB programs emvddataprocess.m and readbin.m in Appendix L were used to process and plot the experimental data. The Lab View file in Appendix F was used to transfer the experimental data to the PC in the experimental EMVD test stand. Figure 6.12 shows the motor position and position error for the PD compensated EMVD. Figure 6.13 shows the actual and commanded motor current for the PD compensated EMVD. 0.e W 0.6 0.4 -.-.-.-- 0. 2 --- -0.2 -0.2 .. .. .... . -0.4 - - ---- - - 0 0.4 W I- .-.--- -0.01 0 Time(s) -0.2 rror - ---- 0.01 .. .. -0.01 0.02 - - -- - - -- -- -0 .4 - 0 Time(s) 0.01 0.02 0 TWne(s) 0.01 0.02 0.02 0.02 -- 0.01 - . 0 w Error - ' n 0.01 u 0 - D-s-0.01 -0.02 -001 0 0.01 -0.02 0.02 -0.01 Time(s) Figure 6.12 Experimental results for the pd controller - motor position and position error. Commanded Motor Current Commanded Motor Current - Down Transition 6[ ..... .. 6 . .. Current n - Up Transition 4. o -2 - - - -CurreaL Injection-4 -4 - -- - -6 . -0.01 0.01 0 Time(s) Actual Motor Current - -- - - - - - -6 0.02 - - - -0.01 0 0.01 0.02 Time(s) Actual Motor Current Up Transition - Down Transition 4, -- - 0 -2 - - -- - .-. -.-.- -4 -4 -6 .---0.01 Figure 6.13 - 0 U -2 ..- --. -- ---. ..-..-..-.0 Time(s) -6 -0.01 0.02 0.01 -I- I-0 Time(s) 0.01 0 02 Experimental results for the pd controller - motor current. From these two figures, we can observe that there is less than 3% position error at any point in the stroke. In addition, we can see that the peak motor current is approximately - 104 - 6.10 Linear Controller Implementation 6A, which was below the current-limited value of 8A. However, the motor current and commanded motor current appear to be very noisy (thus, these two signals have a lot of "spikes" of current), probably because of the action of the numerical differentiator in Simulink. Although the motor does not move in response to these high-frequency spikes in motor current, the spikes do lead to unnecessary increases in motor temperature due to the additional motor current harmonics. Thus, we quickly realized the benefit of using a current-limiter during the transition mode. Figure 6.14 shows the actual instantaneous and average motor powers (averaged over each holding period and each transition period) for the PD compensated EMVD. The instantaneous power plot was obtained by sampling motor current and voltage at 2.5GHz, and then multiplying the samples (at each point) together in MATLAB. Due to the large number of samples being processed in MATLAB, this plot appears to be heavily aliased - MATLAB did not use the correct number of pixels when making the plot. Nonetheless, the samples themselves were not aliased, and the average power was calculated with an appropriate number of samples. The average power during the valve transition period is approximately 15W, while that during the holding period is approximately 5W. In terms of average power consumption, this result is comparable to the power consumption in conventional IC engines at lower speeds. The holding current for the MIT EMVD was not OW as predicted in simulations because of misalignments between the motor and valve assemblies. If we had used a 35Hz reference input, the average power consumption figures reported here would have been slightly smaller. 2I0nstantaneous Motor Power - Up Transition 200 100 - - - - - - 2Instantaneous 200 Motor Power - Down Transition 100 ... 0 a if 100 -200 - - -0.01 0 Time(s) - 0.01 -100 - - 0.02 -200 -0.01 0 Tirne(s) 0.01 0.02 Average Motor Power - Down Transition Average Motor Power - Up Transition 20 20 15, 15 0- 1 0' Figure 6.14 power. -0.01 0 Time(s) 0.01 -0.01 0.02 0 Time(s) 0.01 0.02 Experimental results for the pd controller - instantaneous and average motor - 105 - Controller Design and Experimental Results In the next section, we will discuss similar experimental results for the lead compensator. 6.10.2 Lead Compensator In this section, we will discuss the experimental results obtained using the lead compensator. We will discuss a measurement of valve seating velocity made using this compensator in the next section. Figure E.3 in Appendix E contains the dSPACE Control Desk file used to view the lead compensator performance in real-time, while Fig. B.10 in Appendix B contains the Simulink model used to implement this compensator with current injection. The MATLAB programs emvddataprocess.m and readbin.m in Appendix L were used to process and plot the experimental data. The Lab View file in Appendix F was used to transfer the experimental data to the PC in the experimental EMVD test stand. Figure 6.15 shows the motor position and current for the lead compensated EMVD. Other plots of experimental results for the lead compensated EMVD have been excluded here since they do not add significantly to the discussion. Motor Current in Amperes (2A/di ison) . Current Injection Figure 6.15 . T. Curren njecti ... .. - Time (4ms/divisio Time (4ms/division) Motor Current in Amperes Motor Position in Radians (2Aldivision) (0.216 rad/division) Experimental results for the lead compensator - motor position and current. Although not shown in this figure, the position error when using this lead compensator was also less than 3%. From Fig. 6.15, we can observe that the peak motor current was approximately 4A, which is approximately 2A less than the peak current observed for the PD compensated EMVD. In addition, the motor current is drastically less noisy than that for the PD compensated EMVD. Although not shown in the figure, the average power during the valve transition period was approximately 12W, while that during the holding period - 106 - Linear Controller Implementation 6.10 was approximately 5W. Thus, in terms of average power during the transition periods, the lead compensated EMVD outperformed the PD compensated EMVD. In later experiments with the lead compensator, the reference input was changed to a 35Hz sinusoid, and 14.3ms transition times were observed with a small increase in peak motor current. These transition times correspond to 1500rpm IC engine speeds. In effect, in all the performance criteria that were quantified, the lead compensated EMVD performed better than the PD compensated EMVD. We decided to discontinue the use of the PD compensator and carry out more experiments with the lead compensator. 6.10.3 Valve Seating Velocity with the Lead Compensator Using the lead compensator we designed, we measured the feedback-controlled valve seating velocity for the MIT EMVD. This measurement will be detailed in this section. The experimental set-up for this experiment was the same as that for the lead compensator experiments. Figure B.12 in Appendix B contains the Simulink model used to carry out the valve seating velocity measurements. The experimental data was captured on the oscilloscope and transferred to the PC using the Lab View file in Appendix F.1. The data was processed and plotted using the MATLAB program emvdseatvel.m in Appendix L. Before carrying out this experiment, the valve seat was adjusted as described in Section 4.4.6. Measurement of Seating Velocity E Valve Seated -S E - -V--a--l-- 50 -0.015 -0.02 -0.025 -0.005 -001 0.005 0 Time(s) a -0.025 002 -0.02 SO -0.015 Figure 6.16 Seated -Valve -0.01 0.025 - - -- -. - 200 0.015 0.01 -0.005 0.005 0 0.01 0.015 0.02 Time(s) An estimate of seating velocity. - 107 - 0.025 Controller Design and Experimental Results Figure 6.16 on the previous page shows experimental data (actual valve position and velocity) for an open-to-closed valve transition. From this figure, we can estimate that the valve seating velocity ranges between 3 - 21cm/s. During later experiments in the laboratory, we determined that 21cm/s was an upper bound on the valve seating velocity because the valve seat had been lowered such that the valve was touching its seat well before the actual "closed" end of the stroke. This range of seating velocities is very encouraging because it shows that the MIT EMVD does indeed have small seating velocities, thereby preventing excessive wear of the engine valves and valve seats. 6.11 Robust Adaptive Controller Design In the rest of this chapter, we will discuss the design and implementation of a robust adaptive controller for the MIT EMVD. This controller is a nonlinear controller that directly takes into account the nonlinear dynamics of the EMVD plant. It is an alternative to the timevarying feedback gains control method discussed in Section 3.4.2. One reason for using this controller is to be able to have a fixed controller that may be easier to implement than a switching controller, typical of the time-varying feedback gain controller. Another reason for using this type of controller is to be able to counteract parametric uncertainties in our system model, as well account for the unmodeled system dynamics. We will begin with the design and development of this robust adaptive controller, followed by feedback-controlled EMVD simulations using this controller. We will conclude the chapter with some discussion on the implementation of this controller. 6.11.1 Controller Development In this section, we will develop the idea of the robust adaptive controller. From a control systems perspective, a robust controller is one that will perform its intended function even in the presence of parametric uncertainties or unmodeled dynamics in the system model. An adaptive controller is one that will adapt to changes in the system model itself. We will begin the controller development with the EMVD model given in equation (6.4. Since the nonlinear mass/inertia given in equation (6.4) is always non-zero, equation (6.4) - 108 6.11 Robust Adaptive Controller Design can be rewritten by dividing through by the non-zero nonlinear mass term to obtain: S+ .. Bo +B,(z $)2 +mTz d Jo + mz - (d)2 d d2 + Kz (0) dm Jo + mz(L)2 = KT + d(t) (6.19) Jo + mz(d)2 where d(t) is g(t) divided by the nonlinear mass term. Since the nominal values of Jo, mz, KT, B, Bo, d, , and d(t) in (6.19) are either known or bounded, or both known and bounded, it was possible to design and implement a robust adaptive controller for the EMVD to track the desired motor angular position trajectory Od(t) [27, 22]. The controller can be made robust, in that by having bounds on the above mentioned parameters, we can take them into account when designing the controller. In addition, we can adaptively alter some controller variables to obtain improved controlled response. An excellent reference for this controller design is [22], however, we will give an overview of our design in this section. The development of the robust adaptive controller in [22] is similar to that of the general adaptive controller described in [27], except that the time-varying parameters add uncertainty to the system model which must be taken into consideration when designing the control law. As is the case in many nonlinear systems [22], after substituting the NTF relation (6.5) for f (0), this system can be expressed in the form: 0 + (as, + af (t))O + (as2 + af2(t)) sin(3.460) = (bsbf)im + d(t) (6.20) where asi, as2, and bs are constant parameters, while afi(t), af2(t), and bf are "fast" timevarying parameters. The design objective for our robust adaptive controller was to account for the uncertainty inherent in the slow-varying parameters asi, as2, and bs, and adapt to the changes in the fast-varying parameters afl (t), af2 (t), and bf. While taking into account these parameters, the nonlinear controller we designed had to track the desired/reference motor position, Od(t), as closely as possible. In the next few paragraphs, we will outline the design of our controller. It is important to note that in the literature, the robust adaptive controller designed for a nonlinear system of the form given in (6.20) is called hybrid because it is a combination of adaptive control (for the estimation of the slow-varying parameters) and robust control (to take into account the fast-varying parametric uncertainties). - 109 - Controller Design and Experimental Results In the case of the MIT EMVD modeled by equation (6.19), it is possible to formulate expressions for asi, as2, bs, afl(t), af2(t), d(t) and bf, as follows: | d(t) j< D (6.21) asi 1 Aasi (6.22) as2 |< Aas2 (6.23) af, <_ Aafl (6.24) af2 < Aaf2 (6.25) 1< bf bf - (6.26) 0 where D, Aas2, Aafl, Aaf2, and 0 are the bounds on the various parameters. The main idea behind the robust adaptive nonlinear controller is simple. To design such a control law, we began by defining a "sliding variable", s, by: s0=# + 2AO + 10 (6.27) where 6 is the error in the motor position 0, and s is defined such that the nonlinear system in equation (6.19) was effectively converted to a second-order linear system. From equation (6.27), we see that if we make s = 0, and fix A > 0 such that the system given by: +2A 1j+ =0 (6.28) is a stable second-order system with an exponentially-decaying zero-input response, then 0 will exponentially decay to zero. In this manner, our nonlinear controller can (ideally) achieve almost perfect tracking of the reference motor position input. To prevent discontinuities and high-frequency switching in the control input to the plant as s approaches the s = 0 plane, a boundary layer of thickness <D was defined such that the sliding variable s tends to s < ±D instead of tending towards 0. The boundary layer thickness <D was made time-varying to account for the uncertainties in the "slow" time-varying system parameters [22]. The variable A is effectively the feedback-controlled system bandwidth. The development of the control law itself was more complicated than the idea of defining a sliding variable. The control law was developed such that the feedback-controlled EMVD plant is asymptotically stable, a fact that can be proved using a Lyapunov-like function, V(t) and an important mathematical fact referred to as Barbalat's Lemma. The actual proof of this stability is beyond the scope of this thesis and will only be outlined in Section 6.11.2. - 110 - 6.11 Robust Adaptive Controller Design Using well-known ideas from robust adaptive control, we derived a control law (defining motor current) for the EMVD plant (see [22, pp 1642-1645]) as follows: in = bf (h10 + h 2 sin(3.460)) - ((bl 1bf1)(u* + k(6)sat(j) + 0-y (t)/sA)) where (6.29) h1 adaptively estimates 2-, h 2 adaptively estimates g, b, adaptively estimates bs, bf is a static estimate for bf obtained from (6.26). Furthermore the other parameters in equation (6.29) are given by the following equations: u* = -Od + 2A k(O) = = (t)b 1(u* - Aasi (6.30) (6.31) hi = -'Y(t)hinOsA ; (6.32) h2= bS ; ); sa = s +A - <Dsat( -- y(t)h 2n sin(3.460 )sA; 0 +Aas 2 sin(3.460) I +D (6.33) + rsat(j) + "y(t)sA)sA; /(j bs(hiO + h 2 sin(3.460)) - u* (1 - ,31) + Aasi 16 +As2 I sin(3.460) kd(O) = k(Od); k(6) = k(G) - kd(O) + Ab1. I +D (6.34) + ,); (6.35) (6.36) (6.37) It is not particularly relevant to discuss the exact functions of all the variables in these equations, nonetheless, these equations are presented here in order to be able to explain the controller implementation algorithm described in Section 6.11.3. Of particular relevance, the adaptation rate, 7(t), is time-varying to allow for convergence to the boundary layer[22]. Differential equations that determine both -y(t) and <D can be found in [22]. The parameters hij, h 2 n, and b, also control the adaptation rates of some of the parameters. 6.11.2 Controller Stability As we mentioned earlier, the control law defined in equation (6.29) was derived such that the feedback-controlled EMVD plant is asymptotically stable. In this section, we will outline the proof of this stability. The details of the stability analysis will not be carried out here as they are not very intuitive. For more details, please see [22]. In order to prove the stability of the controller, the following Lyapunov function can be - 111 Controller Design and Experimental Results used: + () hi - a/bs +2 h s+ V(t) = a2/bs 2 - 2 + ± + 2 ,_1 - b--1 2 ) . (6.38) By substituting equations from the previous section, we can obtain the following two results: V(t) -ny(t) I saI 1 d 2 dtsA<- sAI ; (6.39) . (6.40) In effect, these two results prove the asymptotic stability of the nonlinear robust adaptive controller we designed. Equation (6.40) is often referred to as the "sliding" condition in the literature. Controller Implementation Algorithm 6.11.3 The step-by-step algorithm to implement the hybrid adaptive controller can be enumerated as follows: " INPUTS: 0, Od, 0, A, 77, )3, D, Zasi, Aas2, Aafl, Aaf2, and bf " Calculate s " Calculate u* * Calculate k(9) and kd(0) * Calculate <P " Calculate k(9) * Calculate -y * Update the values of h1 , h2, bs " Calculate control input = motor current im " OUTPUT: motor current im This algorithm was implemented in a MATLAB program (see Appendix K). The parameters used for the adaptive controller were as follows: * A 4800 " hin= v2 -r f =h n; b, = 0.01 2 * r = 0.2 112 6.11 ((= * D = *Aas * /(J+m.-3.46 2 -0.0042))/(K /(J2+m Robust Adaptive Controller Design -3.462.0.0042)))+1 200 -3.46 -0.004 = 10 -Bo/Jo Aas2 = 10 - Kz/Jo SAaf = 10000 * Aaf2 = 10000 * bf = (KT2/(J +m.. 2 3.462 .0.0042))+1 2 The system constants and nonlinear mechanical transformer characteristic used in the MATLAB simulation were the same parameters summarized in Table 3.1 of Chapter 3. In the next section, we will briefly discussion simulation results obtained using the nonlinear control law derived in this section. In the last section of this chapter, we will discuss the implementation of this control law in the experimental EMVD test stand. 6.11.4 Simulation Results The nonlinear feedback-controlled EMVD response to a 60Hz sinusoidal reference signal is shown in Figs. 6.17 and 6.18. 10 /q & 0* 5 -0.51 0 0.005 0.01 0.015 0.02 0.025 Time~s) 0.03 0.035 0.04 0.045 0.05 1500 500-500- - 0 Figure 6.17 0.005 0.01 0.015 0.02 - -- . - -. --- - -- - -1000 - - 0.025 Time(s) 0.03 0.035 0.04 .-. 0.045 0.05 Controlled response to a 60hz sinusoidal reference signal. From these figures we can observe that the controller performed reasonably well, considering the fact that the tracking error only has to be small when the valve is almost open (0 = - 113 - Controller Design and Experimental Results Wor Postion and Tracking Error (in radians) 0.3- 40. 1 -0.12\ 0.4 - 0.3-0.4 0 0.005 0.01 0,015 0.02 0.025 0.03 0.035 0.04 0.045 0.05 Figure 6.18 Tracking error and motor position for controlled response to a 60hz sinusoidal reference signal. -0.456) or almost closed (8 = 0.456). However, the peak motor velocity was larger than the rated peak velocity for the motor we have in the laboratory. In simulations, I observed that if the initial parameter estimates were incorrect by an order of magnitude (especially that for bf), the controlled EMVD system became unstable possibly because the parameter estimates cannot be updated fast enough to keep the motor position bounded. For all the simulations, the motor current input was under 42A, which is the maximum allowable current for the dc motor in the laboratory. Because the motor velocity in simulation was too high, more refined simulations were carried out try to decrease this velocity. The results from the next set of simulations were more promising. In the next set of simulations, I observed that if the reference signal frequency was made to match the undamped natural frequency of the frictionless EMVD in Fig. 3.3 in Chapter 3, the system response improved significantly. The motor velocity has more "spikes", but these are artifacts of the simulation and not of the system response itself. In the simulation, the parameters continuously adapted, leading to changes in their values even when tracking error was small - this issue would have to be resolved with a "dead-zone" in the adaptation laws [22] when the controller is implemented in the experimental EMVD test stand. Figures 6.19 and 6.20 on the next two pages show the state variables and tracking error for a 42Hz sinusoidal reference signal. On the page following that, Fig. 6.21 shows the adaptation of the various parameters for a 42Hz sinusoidal reference signal. From these - 114 - 6.11 Robust Adaptive ControllerDesign figures, we can see that the tracking error is very small, and that parameter convergence never occurs, except for b6. 14 I0 0 1 0.005 0.01 0.015 0.02 0.025 Time(s) 0.03 0.035 0.04 0.045 0.0 0.035 0.04 0.045 0.05 200 . .-.-. . . -. . .. . -200 -400 0.005 Figure 6.19 0.01 0.015 0.02 0.025 TkMe(s) 0.03 Controlled response to a 42hz sinusoid. Kle 6 4 2 .-.. -.-.- . --. -.-. - - -.-.. -4 ... .. ..... ...... -8 -80 Figure 6.20 6.11.5 0.005 0.01 0.015 0.02 ..... 0.025 limeS) 0.03 0.035 0.04 0.045 005 Tracking error for controlled response to a 42hz sinusoid. Robust Adaptive Controller Implementation In this section, we will discuss some implementation issues for the robust adaptive controller that was developed in the previous two sections. The nonlinear robust adaptive controller was preliminarily implemented on the dSPACE DSP using a Simulink model, however, the controller did not perform as expected, because - 115 - Controller Design and Experimental Results 6 5X105 6 X105 2-2w i2 .1 0.2 W 00 o4 00 O04 0.05 0.01 2 00 ~ 4 0 -- -4 4 0 o0 ~.0 -2 0.02 0.03 Time(s) 1 - -.... 0.01 0 -...-...-. 0.02 0.03 Time(s) 0.4 0.05 0.04 0.05 0 10 .. .. 1 .5 01 0 j 0 Figure 6.21 0.01 0.02 0.03 Times) 0.04 0.05 0.01 . . 0.02 0.03 Time(s) Parameter adaptation for controlled response to a 42hz sinusoid. the adaptation did not seem to be as fast as it was in simulation. The main reason for this slow rate was that the sampling rate of the DSP in the experimental EMVD test stand was limited to less than 15 kHz when implementing this complicated nonlinear controller. Such a low sampling rate is inadequate for our purposes. In the near future, we will attempt to construct better Simulink models that may allow us to speed up the DSP's sampling rate because the manufacturer of the DSP, dSPACE Inc., claims that the DSP sampling rate cannot be increased without reducing the complexity of the Simulink model. Furthermore, assuming that the implementation of this nonlinear controller is possible using the DSP, the controller will have to be optimized for an appropriate reference input, and appropriate adaptation rates. In addition, a dead-zone will be added to the parameter estimation equations such that the parameters do not update too often. - 116 - Chapter 7 Conclusions and Future Work 7.1 Introduction IN this chapter, this thesis will be concluded with an evaluation of the original thesis objectives, and a discussion of future work on the MIT EMVD project. As noted earlier, in September 2001, a novel EMVD for internal combustion engines was proposed by members of MIT's Laboratory for Electronic and Electromagnetic Systems (LEES). This MIT EMVD is an electromechanical valve drive incorporating a nonlinear mechanical transformer [3, 8, 9, 10]. The proposed MIT EMVD suggested significant benefits over previously designed engine valve actuation systems, including lower average power consumption and smaller seating velocities. The goal of this thesis research was to implement the MIT EMVD in a laboratory test stand and carry out preliminary experiments to confirm these benefits. In the next section, we will evaluate the objectives of this thesis, and in the following section, we will discuss future work that will be done on the MIT EMVD project. 7.2 Evaluation of Thesis Objectives and Contributions As noted in the introductory chapter of this thesis report, the objectives of this thesis were: first, to model the mechanical structure of the EMVD using 3-D modeling software; second, to construct this EMVD apparatus in the laboratory, and integrate it into a computer-controlled experimental test stand; and, third, to carry out experiments to verify the operation of the EMVD and compare experimental results to computer simulations and mathematical modeling. All three primary objectives have been fulfilled. An EMVD apparatus was modeled using 3-D modeling software, constructed, and then - 117 - Conclusions and Future Work assembled in the laboratory. This work was described in Chapters 4 and 5. This EMVD apparatus was also mathematically modeled in Chapter 3. The first objective was thereby fulfilled. The assembled EMVD apparatus was integrated into a fully functional computer-controlled experimental EMVD test stand, thereby fulfilling the second objective of this thesis research. This experimental test stand is described in Chapter 4. Furthermore, a hysteresis currentcontrolled motor drive was designed and constructed for the MIT EMVD. This motor drive was incorporated into the experimental EMVD test stand. Using the experimental EMVD test stand, various controllers for the MIT EMVD can be easily implemented, and the experimental data obtained can be quickly transferred to the PC, where it can be processed and plotted in MATLAB. In this manner, several experiments to verify the operation of the MIT EMVD were performed, thereby fulfilling the third objective of this thesis research. These experiments included those carried out to characterize the motor and motor drive (described in Chapter 5), as well as the open loop characteristics of the EMVD plant (described in Chapter 6). Two linear controllers and a nonlinear controller were also designed for the MIT EMVD. The linear controllers were implemented in the experimental EMVD test stand, and their performances were compared. The experimental results we obtained confirmed the benefit of using a nonlinear mechanical transformer in a motor-driven engine-valve spring system - as seen in the small average power consumption, reasonable transition times, and low seating velocity results from Chapter 6. The experiments we performed also gave us some powerful insights on how to improve the MIT EMVD in the future. In addition to the contributions mentioned above, this thesis report will also contribute significantly to several parts of Woo Sok Chang's doctoral thesis. I also sincerely hope this report will be useful to anyone who works on the MIT EMVD project in the future. 7.3 Future Work The MIT EMVD project is an ongoing undertaking, and there are several significant challenges to be faced before this EMVD can be implemented in an IC engine. The future work on the MIT EMVD project can be summarized in terms of plans in the short term and goals for the longer term. In the near future, many experiments will have to be done on the EMVD apparatus, including experiments with the set of stiffer and stiffest springs - experiments that will aid - 118 - 7.3 Future Work in evaluating the MIT EMVD concept at higher effective engine speeds. In addition, the implementation of an optimized initial mode controller, the generation of a more refined reference input, the implementation of the robust adaptive nonlinear controller mentioned in the previous chapter, and the design and implementation of a time-varying-gain feedback controller (see [23, 24]) must be carried out. All of these experiments will probably be described in Woo Sok Chang's doctoral thesis, as these experiments are an integral part of his doctoral work. In the long term, the dc motor we are using in the experimental EMVD test stand will have to be replaced with a smaller motor with similar or superior torque and inertia characteristics. In addition, the motor drive we constructed will have to be optimized for its size. An appropriate motor cooling system will also have to be purchased for the dc motor that is currently in the EMVD apparatus before we can completely simulate the MIT EMVD at an effective 6000rpm engine speed. - 119 - Appendix A MATLAB Simulation of the EMVD in the 0 Domain % Tushar Parlikar % Simulation of EMVD's Free Oscillation - Globally Stable at theta=O cdc; clear all; tO=0; tf=0.03; amp=26*pi/180; X0 .45378560551853 thetaO=[amp 0]' ; f=120; options = odeset('MaxStep',5e-6); [to tf], theta0, options); y=theta'; [t,theta]=odel5s('adaptsimple', motorpos=y(1,:); motorvel=y(2,:); figure; subplot(2,1,1) plot(t,motorpos); grid; xlabel('Time(s) ') ,ylabel('Motor Position'); subplot (2,1,2) plot (t,motorvel); grid; xlabel('Time(s) ') ,ylabel('Motor Velocity'); function dtheta=adaptsimple(t,theta) t % Constant System Parameters Jm=2*3.54e-6;% rotor inertia + inertia in x-domain reflected back to the motor Bm=0;rotor friction K=2*49328.7;% Spring Constant mv=0.090; % valve, spring, spring divider, etc mass bv=O; % valve friction KT=0.069;7Xmotor torque constant % NTF function alpha=3.46; beta=0.999; lift=0.008; ntf=h*sin(alpha* (theta(1))); h=lift/(2*sin(beta*pi/2)); - 121 - MATLAB Simulation of the EMVD in the 0 Domain % Derivative of NTF function derivntf=(alpha * h) * cos(alpha * theta(i)); % Double Derivative of NTF function dderivntf=(-(alpha^2) * h) * sin(alpha * theta(1)); % Calculating/Updating Coeffs of the state derivatives den= Jm + (mv * (derivntf^2)); aa= Bm + (by * (derivntf^2)) + (mv * derivntf * dderivntf * theta(2)); ab= K * derivntf * (ntf); % Calculating control law i=Q; % UPDATING STATE EQUATIONS dthetaa=theta(2); dthetab= (1/den) * KT*i); dtheta=[dthetaa;dthetab]; C- (aa*theta(2)) - 122 - - ab + Appendix B Simulink Models for Experiments with the EMVD Apparatus This appendix contains Simulink models that were constructed to carry out experiments with the EMVD apparatus, including, but not limited to, the implementation of initial mode control, the generation of an appropriate reference input, the implementation of the PD and lead compensators, the measurement of valve seating velocity, and the time response tests of the motor. F1 Ff(u) X d2f/dz2 K- X - Inertia FrictionA -- f(u) * X1 Inertia2 .0 -fu) -df/dz F2 f(u) Inertia Integrator p ml'''Il''I'l','l''I'l'I Integratc r1 sition To Work~space x2 O F To Worksipace1 Figure B.1 Simulink Model for the MIT EMVD in the z domain. - 123 - - U) C) C,) 0 0 0 a C--A 0 "I X 0 0 0- 0 CO- > + 0 0 C\1J a) CO) 0 CL) Simulink Models for Experiments with the EMVD Apparatus x F 0 mC ) 124 - 0 0 CO) C) CO) 0 Figure B.2 Simulink Model Used to Generate the Reference Input (Including Initial Mode Control) for the Controllers. - P Gain 0.87 K- Ga1 swth SMIth2 Current Gal. 0.92 D Gain K- Gain5 ++ 1 Gainw4 0.05 VA Ratio DAC Saturation dit Derivative 1 -002 Constant Current Offset 1 Sisporsh"" 4a Position Error 51D- ;whh D Latch Sequence 0 x t2 win Wavel input Offlost -0.025e~oo~ebc Ene d*l Oe Mr Poito Gain1 2601WO.17678112 DS1104ENC-POS-C1 1 Slept MUX ADC DS1 104MUXAD DS1104ENC-SETUP Motor Velocity 10.90 Gain Vahve Position Motor Current Cocrmand GaW x7 Motor Volmpg Figure B.3 Simulink Model Used to Implement the PD Controller. - 125 - DS1104DAC-C1 0.0 3 * LUJ F iI 126 - 0.1 Simulink Model Used to Implement the PD Controller with Current Injection. I Simulink Models for Experiments with the EMVD Apparatus Figure B.4 - I i CLi a I tt I I -A *~I EU; 127 - S 0S L Figure B.5 Simulink Model Used to Implement the PD Controller allowing for easier online Parameter Changes. - I a x e 02 I IIa 128 Ia - - I R A SIL Simulink Model Used to Implement the PD Controller and measure Motor xl Simulink Models for Experiments with the EMVD Apparatus Figure B.6 Power. - Figure B.7 tive Block. a Ell 129 - 0 Fill Simulink Model Used to Implement the PD Controller with a Separate Deriva- - 0 I J H I&J -I ~L!i iii Simulink Models for Experiments with the EMVD Apparatus I 130 - Figure B.8 Simulink Model Used to Implement the PD Controller with a Separate Derivative Block and Filter. - Figure B.9 I I I I I I I ~j -j ~i I I I I ~i "! ~ll I *1 Lr~i;I 131 - I J Ii 11 Ig I Simulink Model Used To Implement the first Lead Compensator. - I iO 132 - - LI~ Simulink Model Used To Implement the second Lead Compensator. I Simulink Models for Experiments with the EMVD Apparatus Figure B.10 - (D 0 1.5 +_ ulse Generatori I One1 0 (D 0.05 +, VA Ratio Current Gain DAC DS1 104DAC-C1 Saturation 20 1 Motor Current Command 0.053 Current Offset One2Gain3 1 ENCODER MASTER SETUP DS1 104ENCSETUP du/dt Derivative CD 0 (D 0 I-tl 0 Enc position. DS11E04ENCPOS-Cl, Motor Velocity _J02*pi/2048 Gan1--- .001 DAC VelocityNolts Ratio DS1 104DACC2 6d Et I -4 134 - I I 2 4 1 .1I ~ I I .1 I i 21 ]I~IiI Ii Simulink Model Used To Measure Valve Seating Velocity. 2 Simulink Models for Experiments with the EMVD Apparatus Figure B.12 - Appendix C Drawings of Parts for the Apparatus This appendix comprises the final versions of the drawings that were used to make the parts for the EMVD apparatus. The bushings were manufactured after the apparatus was assembled. W (trz) u-i w N~U) EL CO 05 S a 0 '0 C', C,)! I I - CN oo9C917 Figure C.1 The Hole Locations for the Table. - 135 - 0:) C 13 4X05/16-18 TAPPED (Drill Size F)W 1.250 441 .. LO LO /1.535 01 -0 0 0 I LV 0 r-. r-j 1.878 1.820 N Cb ~1 8.00 0787) Top and Bottom should be arallel to within 0.001" iIt 0 3.878 1 0.535 0. 5 8 0 Cc 10 4XO5/16-18 TAPPED (Dill Size F)W 1.250 4.665r S Left and Right should be parallel to within 0.001" 11 Wcl I,-04 I 6X0.4040 2] 0 -o (Drill Size 4XO5/16-18 TAPPED (Drill Size F) 4.425 " CJ C64 VTHRU 1.25 (2 HOLES EACH ON THE TOP AND ON THE BOT OM) Quantity: 1 TOP VIEW IN D~NSE~PJ DVESOSARE 0.787 NOPIMIAW AND CONIFIDRNIAL SolidWorks Educational LicenseC Instructional Use Only ~cHE ININCHES - - LEES EMVD Project RACTONAL. I/laiD NDi IC21 ANGUJLAR%MACHM TWO PLACE DECIMAL (01 HuED PLACE DECMAL ODDS MAEERLAL COLUMN ONE (1) -- Mild Steel May2MTLartoyfo , APPUCAnION DO NOT SCALE DRAWING , Questions Call Tush art x8-8494 SIEIDWG, ,cAm N. 0ws 2me C[ 6.000 ili.o) j '0 co) 0e) -2.362 N 0630 1.102I 04t4--.L- W THRU 4XYf)4040 (Drill Size Y) 0 -~1 0 0 -0 0d '0 041*< 2X05/16-18 TAPPED (Drill Size F) T 1.25 (1 Hole on the Top, 1 on the Bottom) 4X b5/16-18 TAPPED (Drill Size F) T 1.25 APPED (Drill Size F) W1. 25 (1 Hole on the Top, 1on the Bottom) Quantity: 1 -,Top and Bottom should be parallel to within 0.001 ? 0.591 1.1021 LI 1.181 TOL.ERANCOES FRACflOCLAL: Iiie ANGULAR: MACH.001816."01 ENG AFFPR. WO PLACE DECIMAL 0.010 COLUMN TWO (11) THREEPLACE DECIMAL 0000 Fz- SolidWorks Educational LicenseCPvha -- lnstructional Use Only LEES EMVD Project DMENSIONSARE IN INCHES 0.'70TOP VIEW J EfAY AND CONFIDENIIAL MW 2=. MgfLm.c ElcIWNc IAATEROL Mild Steel or Syftms COLM NEXT ASY USED ON APPUCAT1ON DO NOT SCALE DRAWING AUEDME Questions: Call Tushar at x8-8494 SCALF: 1.2.6 1 WEtGW: .644 00.7969 (Drill Size 51/64)THRU .644 .669 .706 0 OI1 2X 0.0890 (Drill Size 43)WTHRU 6-M 5.750 4XO0.3390 (Drill Size R)VTHRU I-I- 0 .394 .34 1.575 00 0 -1.820 .402- - (.16) 0q .16 Quantity: 1 K1 2 .625 DIMENSIONSARE IN INOES TOLERANCES: *A; EMVD Project 1.ILEES ANGULAR: MACHl (Dl BP(D.001 TWO PLACE DECIMAL 010 71I1E PLACE DECIMAL 0/XI5 MOPUETARY AND CONFIDEA SolidWorks Educational Instructional Use Only mA'BIAL MOTOR MOUNT (BACK) Mild Steel LicenseCpwg May2=. MI Ltafr NEXT ASSY USED ON APPUCARON ~ ~ ~ 1 DO NOT SCALE DRAWING J L±__________± Questions: Call Tush at x8-8494 ~ _______ ~ ~CAIR 2 CL1 IWFiGI4T ISHELT I OF I 0.866 8o, 0.866 Lo.234-J IF 01.53TTHRU Ij;__10 I 4X00.1520 (Drill Size 24)T THRU LO CN 0-0 4XO0.3390 (Drill Size R)T THRU 0 0 0- 1.5354 0 0j Quantity: 1 2.750 1 DIMENSONSARE IN INCHES TOL.ERANCES: IRACZTIOftAL 1/1000 ANGULAR: MAC-1E BNO TWO PACE DECIMAL 0.010 71HREE PLACE DECIMAL. 0JO5 AME DRW uPrPria DTEL M0 LEES EMVD Project 001 SolidWorks Educational Instructional Use Only LICs PROPIETARY AND CONRDENRAL KLaboratoryfor Coyrigt May T20M MTERIAL COM IDAU EDGES NEXT ASSY Al rights mwrved. MOTOR MOUNT (FRONT) Mild Steel USED ON APPUICATION APPlICATiON I DO NOT I SCALE DRAWING DO NOT SCALE DRAWiNG Questions: Call Tushar at x8-8494 SCALE:12 SCALEED 1WEIGHT: IWEIGET lawEE ISHEET IICEOF I *1 0 315 2X 0O.50025±0.00025THRU .866 0 0 co 4X 0.3390 (Drill Size R)WTHRU cco L- (A) 0O.5022 ±0.00025T7THRU 0 (B) 00.3770± 0.00025THRU 80 ,6 CD 0. 3X O.2770 (Drill Size J) WTHRU - 00. 0 0.551 H It 0.709 (0.25) 0.15~ Quantity: 1 of (A) and 1 of (B) LI (A) and (B) differ in the dimension of the center hole. Note: Clamp together with a "bottom plate" before drilling the thru holes. LEES EMVD Project DIMENSIONS ARE ININCHES TOLERANCES: 2.953 SPOPE!ARY ANGULAR- MACH.D BN .001 ThOPLACEDECIMAL .01 TIREE PLACE DECIMAL S, AND CONFIDENIAL SolidWorks Educational Instructional Use Only Licensecpydgt My a =-.IT Laborotoryr . andElecrrncSyftms. A"LMild Steel NXASYUSEDON APPUCATION I DO NOT SCALE DRAWNG LQuEstK at A8-49 al TOP PLATE sa ADWG SCALE No 8 (A) |WIRC and (B) " |I SHUE I U. I0/! 0.039 0.381 0.328 0.512 0 00 2X00.50025 ±0.00025WTHRU 00.5781 (Drill Size 37/64) I7THRU -0 CD VY) 8 "6 0 T-1-7n CD \ rl'*)i 3X00.277 (Drill Size J)T THRU 0 0 00.5625 (Drill Size 9/16) I7THRU co) CI Cs cv, -o Clamp together with either "top plate" or Uanother top plate" before drilling thru holes. 0.25 Quantity: 2 | O NAME DMENDONS ARE IN INCHES TOLERSANCES: FRACTIONAL: I/T00 ANGULAR: MACA E01ND .01 TWO PLACE DECIMAL 0.010 THREE PLACE DECIMAL 0.005 E 1,634 I 3MOETAY ANDICOEFIDRN1IAL SolidWorks Educational License cvwo- Instructional Use Only ReEEroneioweSes rserved. IAN Ngs MAERIAL NEXT ASSY USED APPLICATION APPLICATION f DO NOT I LEES EMVD Project -ATE MFG APPR. COMIWENIS: ON SCALE DRAWING DO NOT SCALE DRAWSNG 11 CHECKED BOTTOM PLATE Mild Steel mALoGTwZryw - at x8-8494 Ci Uh SCALE: 1: 1.26 1VVDGHT: -$ME F, 1SDES~ GE 10 2.3 3 1.100 at the bottom (Taper at a 45 Degree Angle) ~ THRU 0 0 4X O.3390 (Drill Size R)17 THRU 0 8 00.787-T THRU 0 60 Ce TOP VIEW 0.861 H CD1 A45.- 0.399 1.063 ID IN 0.315 0 11______ 1 f 2.953 NOMEAIY AND COIDENTIAL A] rights sese LEES EMVD Project CHECKED ENG APPR. VALVE SEAT MFG APPR. Mild Steel M17 Lobwo Y (w s " =. SolidWorks Educational LIcenseC Ealesfoosognetic and Electord Systems. Instructional Use Only 1 _Quantity: DIENSONS ARE IN INCHES TOLERANCES: FRAINAL- ShOE0D A :GULAR MACH W1 BND .0I TWO PLACE DECIMAL 0.010 THREE PLACE DECIMAL 0.015 CO*EMMU NEXT ASSY USDON APPLICATION I DO NOT SCALE DRAWING L EDGES Quesions: Call Tusha at x8-8494 SM SAE scALE:2 NDWG. IWlG 11 r2 ISE14MIOF I Ioos W.193 4236 0 %uM.002 e"4I) ku'i' 0.236 irir~u 1 S0.236 0. 0 cJ 0 I01 C..4 6j (D tC 0.472 -0.472 0 CN di 0.236 0.472 CD 01 CV) 2 04 10 co) I 04J 0. _ _ _ _ _ Quantity: 2 C 0 0.23355 ±0.00025T7 0. 19685\ DIMENSIONS ARE IN INCHES TOLERANCES: FRACTIONAL- 1/1000 ANGULAR: MACH00 8i BPO Efi TWO PLACE DECIMAL ai THRE PLACE DECIMAL sins LEES EMVD Project CLZZLKW IbOPRETAIY AND CONFDENTIAL M La SolidWorks Educational License copydgh my Instructional Use Only MATIRAL x otaylo NEXT ASSY USED Mild Steel ON APPn(AION DO NOT SCALE Questions: Call Tushai at x8-8494 ,DRAWING VALVE HOLDER mw. IN 7 pSCLE2 1 g~ suur i LQ4 * (A) 00.5 0 0.375 0;l*) Lco d 0 00.2366 ±0.00025TTHRU RO.492 0 0* (b 0 -*These lines should be tangent to the 0.492" Radius Circle- Hn co 0.197 ed (Drill 3ize 6-32)TTHRU OTa 0.706 d Lzi~ Note (1): (A) and (B) differ in the outer diameter *. Note (2): The arm of this part does not have to be precise _.7-1 00 0.375- - 1 -- Quantity: 1 of (A) and 1 of (B) D"NSIONS ARE IN INCES TOLERANCES. FRACTIONAL' 1/00 . LEES EMVD Project ANIGUAR MACH .01 W0.001 0.3( 15 TOPACE DECIMAL 0.010 DIRE0 PLACE DECIMAL 0.00 PNOPRIEARY AND CONFIDENTIAL SolidWorks Educational LicenseCw omoy.MT Instructional Use Only . "mAL' -CHEMLZ SPRING DIVIDER Mild Steel or S UETA EDLSMON APULCATION IDO NOT DO NOT SCALE DRAWING Questions: Call Tushai at x8-8494 m D.09 (A) and 9(B) j"8 scW-4:: JWEG ~ME2A 1w0010 ISHM I F 00.3737 t0,)025 CV) Round AJI Edges RO.387 61 Mv 01 C,' C0 Is hole should be drilled on a CNC milling machine. e hole diameter is 0.398" +/- 0.0005". The profile is included in the files on our zip disk - "curve.txt" ere should be atleast 0.200" <> perpendicularly) of material on cV ach side of the slot. I F QUANTITY: 2 with case-hardening 2 without case-hardnening Nk Please refer to the Solidworks Drawing when making this part. SolidWorks Educational LicensecopyvotMay-. Mubnal.oyfor and Elcrorgc Systems. Instructional Use Only 1 1SO PLACE DECIMAL 11010 1HE PLACE DECIMAL oms I OPEUARY AND CONFIDENIAL M&^"AL NEXT ASSY USED APPUCATON Mild Steel ON RNW I DO LEES EMVD Project ONS ARE IN INCHES OANCNES ENG AMP. DISK CAM CIA, )uestlons: Cl NOT SCALE DRAWING at x8-8494 uhrIA 19 11 E r rV TRUE RO. 1869 ±0.00025 L0 This hole has diameter 0.398" +/-0.0005". It should be drilled with a CNC milling machine. The profile is included in our zip disk file -"curve.txt" -t) (D) TOEANCES: N AGLk AH01 SolidWorks Educational Instructional Use Only Use Only Instructional FRlOMiETARY AND CONFIDENIIAL Laat"YfMit Licensecpydaht May2. Ek~.*offw3Wwtc and ElsckoNhc Svasana. DISK CAM MAERLd StQeDSKCA 4RM tRN Cl Tushar sA ueons: NX APPLICATION DO NOT LEES EMVD Project BND alWdI SCALE DRAWNG at X8-8494 6 o :|9HO Le| LO 0 8 LL C3 CN 6 *0 L -S CNI Oq (D (0.625) (0.380) (D1 LD 04 10 C+ 65 DIENSONS ARE IN MM TOLERANCES. FRACTIONAL: W/OOD ANGULAR: MACH.001 SM.001N EGAPPR. 1WOPLACEDECIMAL a10 TEE PLACE DECNAL 0.05 MFG APPR. I OPMEIARY AND CONFORMAL SolidWorks Educational Licensecovgvm Instructional Use Only c 8a All d ts mev.. MATEMA 2= wLabwatoryfO, I APPUCAION I Do NOT BUSHING A Mild Steel NOSH NEX ASSY USED ON and EtrIcSylwm LEES EMVD Project ROUN ALL EDGES SCALE DRAIWING MZE A G..- I SC.E5; I EGT WISHE OF '1 U') 0 C*4 Cf C 04 Iq CN IN CIO. Oq (0.503) (0.625) 00 0 0 LO DRENSONS ARE IN MM TOLERANCES, FRATIONAL- 1/100 ANGULAR: MACH1.00 BND TWOPLACECIMAL 0.01 T1H PLACE DECIMAL SM6 .I IOPIIETARY AND CONFIDO(IM SolidWorks Educational Licensec Instructional Use Only AS May 20D2.MIr rIV* sm-ved. NEXTASSY - MAMAL Mild Steel Loratmy fo ~IVIFGAPP APR ENG DO NOT SCALE DRAVANG BUSHING B QA. zazw-m: ROUND AM 8)M USEDON APPUCATON I LEES EMVD Project DWG. No. A SCIE-5:1 !WCf REV S;E _ |2%E7 Lf C5 C 04 L0, El S U3 (0.0938 I-dea LfO N. v C? 1.ENSONS AE IN MM TOLERANCES: FRACTIONAL 1/10 BND.D1I ANGULAR: MACH.001 TWOPLACEDECIMAL 0.010 THREE PLACE DECIMAL OSin I IOPWMETARY MITLboratoryfor SolidWorks Educational LicenseopydghwMay2x2. y E"m"ngneftad Instructional Use Only AND CONDENIIAL MAIBMA NEXT ASSY USED I ENO LEES EMVD Project APR BUSHING C Mild Steel ON APPUICATON APPUCATTON NAME DAW CHECKED aIDwG. DO NOT SCALE DRAWING I DO NOT SCALE DRMMNG No-.- mSC ES: IWGT IWEGEET ESAJ~SI ISHEET I OF Shift I OF I I i~ZZZZ ) ItI 04 L() C LfO cc 0 CNi %_0 0_ 1 0 0102 6 O (9 LIMENDONSAJED IN MM DRA - T" ata -*% SolidWorks Educational Instructional Use Only LEES EMVD Project TOLERANCES: 1/1000 -NAL0ANGULAR MACHI.0d 814D .0 TWO PLACE DECIMAL 0.010 0X5 DECIMAL. TSM PLACE PIOPETARY AND CONFIDENIIAL Licensecpsiu "sasLov f MATERIAL Mild Steel RML &L M cowem* NEXTASSY IUSIEDON APPUCATION DO NOT SCALE DRAWING BUSHING D S~~ scALIEZA:I WBGH0: IWGVND ISHE OF 1 (1.386) 0.606 0.236 0.606 1.457 r-. '01 '0 10C Large Hole 1 S0.94425 ±0.000257 THRU Base Piece 2.598 Oq 1.693 1.850 C C.,' C 1 , 4X0.3390 (Drill Size R)W THRU 0.374- 0 -6 00so 0q I--0 10 coIN 0 0. =0 This part can be made out of a weld-ment, or even a bent plate. Please make sure that the large hole is perpendicular to the base piece. No0.606 Quantity: 1 Q.>06 D. ESNSAEINICE ME SPROPIETARY AND CONFIDGEN iI SolidWorks Educational License C-y'sm2=Lorn-m Instructional Use Only mm&Mild NEXT ASSY USED APPLICATION LEES EMVD Project SIOSARE ININCHES TOLERANCES: RACTIONAL~ 1/100) ANGULAR: MACH.1 BNDM TAR PLACE ECIMAL 0.0 1T-REE PLACE DECIMAL OM5 E IZZLI NTAM 1 BEARING HOUSING Steel ON 1 DO NOT SCALE DRAWING _ ___ _ ____ I J__ ___ Qu='0all at x8-8494 ASCAIR Tushh ar aI - SCALE:l:1S o : WEIG 11.5'IWiIIHT; 1 ISHEET I OF I 3.780 3.386 3.386, 14-1 01 0 M Iq i C0 I U") IN 0 01 0 5.136 .4 7 0"4. 2.362 09 (b 0.591 2X 00.3390 (Drill Size R)! THRU 1.181 0.768 A n 551 OR00 -1 o 0 I.- Quantity: 1 0 000 In 0 Top and Bottom Sides should be parallel to within 0.001' -0 4X 5/16-18 Tapped (Drill Size F) W0.500(It is OK ifyou tap thru the 0.581" secti on) 0D 2X00.3390 (Drill Size R)W THRU !I 1- 1.181 0.591 TOP VIEW C95 F--- SolidWorks Educational Licensec-nvogy.M and Ak Ig. Only Instructional Use E-A-' TCLERANCES. FREACTIONEAL l/lIDM ANGULAR: MACH1.001 8ND.00I TVWOPLACE DECIMAL 0.010 THIlE PEACE DECIMAL 0.005 M~EILMild LEES EMVD Project BEARING HOUSING HOLDER Steel COMRNUS ENDALLEDGES Elt NEXT ASSY USED ON APPUCATON a -8494 Tushar [eI Call No 14 T7 JOAOSARO GET OF1 Appendix D Drawings of Parts for the Dynamometer Test Stand This appendix comprises the final versions of the drawings that were used to make the parts for the Dynamometer Test Stand. SOz z G ii C" 0j It d 0S XX LO 7I0 Li: 99R7777 T HN-F6IW1 Figure D.1 The Dynamometer Motor Mounting Plate. - 153 - AY / 1A-1 A TA PPFl (Fn i F3.833 rI1 3.833 Si7 F Ti non 2.500 Lr c. M 171 LID C14 t -1.11.19.200 (b A C* 0 0 2X0.281 (Drill Size K) -7THRU 8.000 -1 DfNSONS:ARE IN INCHES TOLERANCES: FRACTIONAL 1/1000 AhGULAW MACH. I ENDu rnCHECKE TWOPRACE DECIMAL 0.010 EN AP THRDEE PLACE DECIMAL 0.00S 0.6000 IOPIoETARY AND CONFIDENTIAL GEI R Alumlnum SolidWorks Educational LicenseCopyightmay2002. WLabratoryf Systems. Electromagnefic and Electronic NEXTr Assy USED ON Instructional Use Only AN''ots"re*e'ed. DO NOT SCALE ORAWANG APPUICATION LEES EMVD Project DYNO COLUMN K4G AP O.A Questions: Call Tushar I SIZ. A at x8-8494 SCAL1 1 I IWEGH IsHEET I OF I Appendix E dSPACE Models for Experiments with the EMVD This appendix contains experimental layouts from dSPACE Control Desk Software that were used, together with the Simulink models in Appendix B, to carry out experiments with the EMVD. Figure E.1 dSPACE Control Desk Real-Time Window for Time-Response Motor Tests. - 155 - dSPACE Models for Experiments with the EMVD Figure E.2 pensator. dSPACE Control Desk Real-Time Window for Experiments with the PD Com- - 156 - Figure E.3 dSPACE Control Desk Real-Time Window for Experiments with the Lead Compensator. - 157 - Appendix F Lab View File Used to Read Oscilloscope Data This appendix contains the Lab View file that was used to transfer oscilloscope data to the PC after running an experiment with the EMVD apparatus. TDS754D Read Scope.vi C:\Documents and Settings\Administrator\My Documents\abviewfiles\TDS754D Read Scope.vi Last modified on 10/18/2002 at 11:41 AM Page 1 Printed on 1/16/2003 at 6:29 PM Figure F.1 dSPACE Control Desk Real-Time Window for Time-Response Motor Tests. - 159 - Appendix G MATLAB Program for the Disk Cam Roller-Follower Profile This Appendix contains the MATLAB file that was used to generate the profile for the disk cam roller-follower. The points generated by this program were used in SolidWorks to create a 3-dimensional model for the disk cam. %%%%%%%%%%%77 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % EMVD Project % Tushar Parlikar % March 2002 % Name of file: diskcamprofile.m % Purpose: to generate a slot in the disk cam for the roller %7%777777777777777%777%777777%% % % % % % 7777%%%%%%%%%%%%%%%%%%%%%%%%777 clc; clear all; % Enter the equation for the NTF relating the z and theta domains lift=0.008; alpha=3.46; beta=0.99; r=0.01675; % desired mid stroke gear ratio rO=0.004; % roller radius h=lift/(2*sin(beta*pi/2)); endpt=26; n=(endpt*2)+1; theta=linspace(-endpt ,endpt ,n); zl=h*sin(alpha*theta*pi/180); plot(theta,zl); title('Z=f(\theta)') (degrees)'),ylabel('Z (mm)'); grid; xlabel('\theta % Translate by +r z2=h*sin (alpha*theta*pi/180)+r; pause;clf; plot(theta,z2); title(' Z=f (\theta)+r') xlabel('\theta (degrees)') ,ylabel('Z (mm)'); grid % Convert to Polar Coordinates and "flip" x and y. % Change y to -y x=1*z2.*cos(theta*pi/180); y=z2.*sin(theta*pi/180); xO=y; yO=-1*x; pause;clf; plot(xO,yO) title('y=g(x)') xlabel('x'),ylabel('y'); axis equal 161 - MATLAB Programfor the Disk Cam Roller-Follower Profile grid % Generation of Top and Bottom Surface Profiles: % Find Derivative- numerically and analytically % Numerical Gradient dy-dxnum=gradient(yO)./gradient(x0); % Analytical Gradient - using quotient of parametric derivatives dxdtheta=(h*sin(alpha*theta*pi/180)+r).... *cos(theta*pi/180)+h*alpha*sin(theta*pi/180).*cos(alpha*theta*pi/180); dydtheta=(-h*sin(alpha*theta*pi/180)+r).... *sin(theta*pi/180)-h*alpha*cos(theta*pi/180).*cos(alpha*theta*pi/180); dy-dxanaly=dydtheta./dxdtheta; % The Equations below plot the profiles for the top and bottom slots. zetal=atan(dy-dxanaly); zeta2=atan(dy-dxnum); for i=1:length(xO) xlanaly(i)=x0(i)-rO*sin(zetal(i)); ylanaly(i)=y0(i)+rO*cos(zetal(i)); x2analy(i)=x0(i)+r0*sin(zetal(i)); y2analy(i)=y0(i)-rO*cos(zetal(i)); xlnum(i)=x0(i)-rO*sin(zeta2(i)); ylnum(i)=y0(i)+r0*cos(zeta2(i)); x2num(i)=x0(i)+r*sin(zeta2(i)); y2num(i)=y0(i)-r0*cos(zeta2(i)); end % Plot Roller Profiles pause;clf plot(xO,yO,'k',xlanaly,ylanaly,'b',x2analy,y2analy,'r',xlnum,ylnum,... 'b-',x2num,y2num,'r-') axis([-0.008 0.012 -0.024 0.002]) axis equal grid title('Cam Roller Profiles') legend('Center of roller', 'Top contact... point(analytical)','Bottom contact point(analytical)','Top... contact point(numerical)','Bottom contact point (numerical)') xlabel('Horizontal Displacement - x (meters)'),ylabel('Vertical... Displacement - y (meters)') gtext('\theta=0') - 162 - % Generate extra points to get 1mm of "roll" at the end of the stroke grdl=(yO(n)-yO(n-1))/(xO(n)-x(n-1)); grd2=(yO(1)-yO(2))/(xO(1)-xO(2)); [a,b]=solve(' (a-0.0091)/(b+0.0187)=0.4519' ,' ((a-0.0091)^2)+((b+0.0187))^2)=0.000001'); % Generate Profile Data and Save in a Text File for i=1:length(xO) profile1(:,i)=1000*[xO(i); yO(i); 0]; %convert to mm end fid=fopen('disk.txt','w'); fprintf(fid,'%6.2f %6.2f %6.2f\n',profilel); fclose(fid); 163 - Appendix H Printed Circuit Board Schematics and Layout This appendix contains a circuit schematic and some PCB layout diagrams for the motor drive that was constructed in the course of this project. The original circuit schematic was updated in August 2002 to reflect changes that needed to be made after testing the motor drive. Figure H.1 Circuit Schematic for the Motor Drive Circuit. - 165 - Appendix I Summary of Pacific Scientific 4N63 Data Sheet This appendix contains a summary of information from the Pacific Scientific 4N63 DC motor data sheet. For the EMVD apparatus, we used the 4N63 - 100 DC motor. 4N Ratings and Characteristics Coo-ing 4N63-000-1 None 4N63-100-1 Type 4N63-000-2 4N63-100-2 "None Type 1 Parameter Rated Torque Umts oz-in. // Nm 537T// T,67T3W 767/F,4 Rated Current (RMS) amps Catalog Listing Thermal Resistence (Rotor-Ambient) Continuous Power Dissipation (Power In - Power Out) Rated Voltage t ee Rated Power Out Pulse Current Continuous Stall Torque No Load Speed at Rated Voltage Torque Constant 14.2 6.7 14.1 1.81 0.47 1.81 0.47 72 30 3250 132 48 275 42 72 36 275 48 deg C/watt watts volts RPM watts amps max. oz-mn Nm Weight Figure I.1 3.5 x kg-mA2 oz-in/kRPM Nm/kRPM oz-in. // Nm microhenries 2.7/ 0.89 Too .- .089 1.29 - .00M8 3.5 x 10^-6 6.3 x 10A-6 0,008 0,011 0,011 2.f //l,014 7.T7/ TA1 1.1 TTT 7MI --- 1.5 6.3 x 10A-6 0.6 0.7 14.8 / 2.2 0.7 0.2 0.11 0.11 4.8 //2.2 0.2 4.8 / 2.2_ Data Sheet for the Pacific Scientific 4N63 DC Motor. - 167 - 1.5 -7- 775 0.6 4.8 // 2.2 0.89 1.29 U.M0 -- mS 9.7 9.47 1.31 ,1 5000 12.8 0,090 0,090 7.33 1.31 1.1 174 1,23 3750 12.8 5640 9.9 . 10^-6 47 77 0,544 0,070 4 47 134 7.33 0,008 mS lbs. /kg T 4 4025 9.9 oz-inYA-S^ 3 3 321 0,946 0,070 (back EMF constant x 1.353' Nm/amp V / kRPM Back EMF Constant ohms @ 25 deg. C Motor Terminal Resistance ohms @ 155 deg. C Rotor Inertia Viscous Damping Coefficient Static Friction Torque, max. Rotor Inductance Mechanical Time Constant @ 25 deg. C ([rotor inertia x terminal resistance x 105/torque constant]/ back EMF constant Electrical Time Constant @ 25 deg.C (rotor inductance/terminal resistance) 3 60 0,424 RPM oz-in./amp T63W/1T3 6.8 Appendix J Programs to Analyze the Motor and Motor Drive Tests This appendix contains MATLAB programs that were used to analyze the experimental data from the various motor and motor drive tests we performed, including the dynamometermotor tests, the motor time response tests, and the motor inductance tests. XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX%%XXXXXXXX % Program to Plot Motor Time Response Data obtained in LabView (from the Oscilloscope) % Tushar Parlikar % motortimetests.m % EMVD Project % LEES at MIT % Done from November 15-17, 2002 XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX clear all;clc % MOTOR B with flywheel [tla,motcurrla]=readbin('motcurrentia.bin'); [tla,motvella]=readbin('motvelocityla.bin'); [tla,currcomla]=readbin('currentcomla.bin'); % Filter Motor Current avgmotcurr1a=f ilter(1/1000*ones (1, 1000) ,1,motcurr1a); % Search for Time Constants for i=1:length(tla) if tla(i)<=0 indextla(1)=i; end if tla(i)<=7.5 indextla(2)=i; end if tla(i)<=15 indextla(3)=i; end - 169 - Programs to Analyze the Motor and Motor Drive Tests end step=motvella(indextla(2))-motvella(indextla(l)); perct=1-(1/exp(1)); tauvalue=step*perct; for j=indext1a(1):indext1a(2) if motvelia(j)>tauvalue+motvella(indextla(1)) tauBdownwithfwheel=tla(j)-tla(indextla(l)); end end step=motvella(indextla(3))-motvella(indextla(2)); perct=1-(1/exp(1)); tauvalue=step*perct; for j=indext1a(2):indext1a(3) if motvella(j)<tauvalue+motvella(indextla(2)) tauBupwithfwheel=tla(j)-tla(indextla(2)); end end % Plot Waveforms figure; subplot(2,1,1) plot(tla,avgmotcurrla,'g-',tla,currcomla*1.5,'k-.') grid;xlabel('Time (s)');ylabel('Current (A)');title('Current Waveforms for Motor B with flywheel attached'); axis([O 15 -2 2]); legend('Average Motor Current','Commanded Motor Current'); subplot(2,1,2) plot(tla,motvella*100) grid;xlabel('Time (s)');ylabel('Velocity (rad/s)');title('Motor Velocity for Motor B with flywheel attached') axis([0 15 -300 300]); gtext('\tau_{up}=1.122s');gtext('\tau_{down}=1.154s'); %% MOTOR B [tlb,motcurrlb]=readbin('motcurrentlb.bin'); [tlb,motvellb]=readbin('motvelocitylb.bin'); [tlb,currcomlb]=readbin('currentcomlb.bin'); % Filter Motor Current avgmotcurrlb=filter(1/1000*ones(1,1000),1,motcurrlb); % Search for Time Constants for i=1:length(tlb) if tlb(i)<=0 indextlb(1)=i; end if t1b(i)<=(0.35/2) indextlb(2)=i; - 170 - end if tlb(i)<=0.35 indextlb(3)=i; end end step=motvellb(indextlb(2))-motvella(indextlb(1)); perct=1-(1/exp(1)); tauvalue=step*perct; for j=indext1b(1):indext1b(2) if motvellb(j)>tauvalue+motvellb(indextlb(1)) tauBdown=tlb(j)-tlb(indextlb(1)); end end step=motvellb(indextlb(3))-motvellb(indextlb(2)); perct=1-(1/exp(1)); tauvalue=step*perct; for j=indext1b(2):indextlb(3) if motvellb(j)<tauvalue+motvellb(indextlb(2)) tauBup=t1b(j)-tlb(indext1b(2)); end end % Plot Waveforms figure; subplot(2,1,1) plot(tlb,avgmotcurrlb,'g-',tlb,currcomb*1.5,'k-.') grid;xlabel('Time (s)');ylabel('Current (A)');title('Current Waveforms for Motor B without flywheel attached'); axis([0 0.35 -2 2]); legend('Average Current','Commanded Current'); subplot(2,1,2) plot(tlb,motvellb*100) grid;xlabel('Time (s)');ylabel('Velocity (rad/s)');title('Motor Velocity for Motor B without flywheel attached'); axis([0 0.35 -320 320]); gtext('\tau_{up}=20.40ms');gtext('\tau_{down}=20.39ms'); U% MOTOR A with flywheel [t2a,motcurr2a]=readbin('motcurrent2a.bin'); [t2a,motvel2a]=readbin('motvelocity2a.bin'); [t2acurrcom2a]=readbin('currentcom2a.bin'); % Filter Motor Current avgmotcurr2a=filter(1/1000*ones(1,1000),1,motcurr2a); % Search for Time Constants for i=1:length(t2a) if t2a(i)<=0 - 171 Programs to Analyze the Motor and Motor Drive Tests indext2a(1)=i; end if t2a(i)==7.5 indext2a(2)=i; end if t2a(i)==15 indext2a(3)=i; end end step=motvel2a(indext2a(2))-motvel2a(indext2a(1)); perct=1-(1/exp(1)); tauvalue=step*perct; for j=indext2a(1):indext2a(2) if motvel2a(j)>tauvalue+motvel2a(indext2a(1)) tauAdownwithfwheel=t2a(j)-t2a(indext2a(1)); end end step=motvel2a(indext2a(3))-motvel2a(indext2a(2)); perct=1-(1/exp(1)); tauvalue=step*perct; for j=indext2a(2):indext2a(3) if motvel2a(j)<tauvalue+motvel2a(indext2a(2)) tauAupwithfwheel=t2a(j)-t2a(indext2a(2)); end end % Plot Waveforms figure; subplot(2,1,1) plot(t2a,avgmotcurr2a,'g-',t2a,currcom2a*1.5,'k-.') grid;xlabel('Time (s)');ylabel('Current (A)');title('Current Waveforms for Motor A with flywheel attached'); axis([0 15 -2 2]); legend('Average Current','Commanded Current'); subplot(2,1,2) plot(t2a,motvel2a*100) grid;xlabel('Time (s)');ylabel('Velocity (rad/s)');title('Motor Velocity for Motor A with flywheel attached') axis([O 15 -400 400]); gtext('\tau_{up}=1.117s');gtext('\tau_{down}=1.325s'); %% MOTOR A without flywheel [t2b,motcurr2b]=readbin('motcurrent2b.bin'); [t2b,motvel2b]=readbin('motvelocity2b.bin'); [t2b,currcom2b]=readbin('currentcom2b.bin'); % Filter Motor Current avgmotcurr2b=filter(1/1000*ones(1,1000),1,motcurr2b); - 172 - % Search for Time Constants for i=1:length(t2b) if t2b(i)<=0 indext2b(1)=i; end if t2b(i)==0.25 indext2b(2)=i; end if t2b(i)==0.50 indext2b(3)=i; end end step=motvel2b (indext2b (2)) -motvel2b(indext2b (1)); perct=1-(1/exp(1)); tauvalue=step*perct; for j=indext2b(1):indext2b(2) if motvel2b(j)>tauvalue+motvel2b(indext2b(1)) tauAdown=t2b(j)-t2b(indext2b(1)); end end step=motvel2b(indext2b(3))-motvel2b(indext2b(2)); perct=1-(1/exp(1)); tauvalue=step*perct; for j=indext2b(2):indext2b(3) if motvel2b(j)<tauvalue+motvel2b(indext2b(2)) tauAup=t2b(j)-t2b(indext2b(2)); end end % Plot Waveforms figure; subplot(2,1,1) plot(t2b,avgmotcurr2b,'g-',t2b,currcom2b*1.5,'k:') grid;xlabel('Time (s)');ylabel('Current (A)');title('Current Waveforms for Motor A without flywheel attached'); axis([0 0.5 -2 2]); legend ('Average Current' , 'Commanded Current'); subplot(2,1,2) plot(t2b,motvel2b*100) grid;xlabel('Time (s)');ylabel('Velocity (rad/s)');title('Motor Velocity for Motor A without flywheel attached'); axis([0 0.5 -400 400]); gtext('\tau_{up}=57.16ms');gtext('\tau_{down}=61.90ms'); % TIME CONSTANTS: fprintf('tauBup = %f s\n', tauBup); fprintf('tauBdown = %f s \n', tauBdown); fprintf('tauBdownwithflywheel = %f s \n', tauBdownwithfwheel); 173 - Programs to Analyze the Motor and Motor Drive Tests fprintf('tauBupwithflywheel = %f s \n', tauBupwithfwheel); fprintf('tauAup = %f s\n', tauAup); fprintf('tauAdown = %f s \n', tauAdown); fprintf('tauAupwithflywheel = %f s \n', tauAupwithfwheel); fprintf('tauAdownwithflywheel = %f s \n', tauAdownwithfwheel); %%RESULTS from MATLAB %% tauBup = 0.020400 s %% tauBdown = 0.020390 s UX tauBdownwithflywheel = 1.154000 s XX tauBupwithflywheel = 1.122400 s XX tauAup = 0.057160 s XX tauAdown = 0.061900 s XX tauAupwithflywheel = 1.116800 s XX tauAdownwithflywheel = 1.325000 s - 174 - X Data Analysis for Motor Inductance Tushar Parlikar 7 motorinduct.m % EMVD Project X LEES Laboratory X November 20, 2002 % Experiments done by Yihui and Mike % Data Analysis done by Tushar clear all; cdc; % In the program below, f=frequency, L=inductance, Q=quality factor % Data for Motor A (with disk cam) fA=[1 2 3 4 5 6 7 8 9 10 15 20 30 40 50 60 70 80 90 100 150 200 300]; LA=[110.1 118.8 91.4 99.8 83.3 82.4 80.56 75.42 75.6 73.92 65.97 60.32 52.96 48 44.3 41.5 39.15 37.26 35.66 34.28 29.45 26.5 22.95]; QA=[0.5 0.9 0.9 1.2 1.2 1.3 1.4 1.3 1.4 1.5 1.6 1.6 1.6 1.6 1.5 1.5 1.5 1.5 1.5 1.5 1.5 1.4 1.3]; % Data for Motor B fB=[1 1.4 1.5 1.6 1.7 1.8 1.9 2 3 4 5 6 7 8 9 10 20 30 40 50 60 70 80 90 100 150 200 250 300]; LB=[53 84.9 86.2 88.1 90.5 92.5 92.7 93 95 92.5 90.2 70.3 80.15 78.75 76 74.5 60.8 53.6 48.6 45 42.1 39.87 38.02 36.44 35.08 30.33 27.42 25.4 23.9]; QB=[0.2 0.5 0.6 0.1 0.7 0.8 0.8 0.9 1.2 1.4 1.5 1.2 1.5 1.6 1.6 1.6 1.7 1.7 1.6 1.6 1.6 1.6 1.6 1.6 1.5 1.5 1.4 1.4 1.3]; % plots figure; subplot(2,1,1) plot(fA,LA,'bx-',fB,LB,'k:');Xgrid; title('Inductance (L)'); xlabel('Frequency (kHz)') ,ylabel('Inductance (\muH)'); legend('Motor A','Motor B'); axis([0 80 0 120]); title('Motor Inductance and Inductance Quality Factor'); subplot(2,1,2) plot(fA,QA,'bx-',fB,QB,'k:');%grid; xlabel('Frequency (kHz)') ,ylabel('Quality Factor'); A','Motor B'); axis([0 80 0 1.8]); - 175 - legend('Motor Programs to Analyze the Motor and Motor Drive Tests %XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX% Program to Analyze Data from the Dynamometer Tests dynotests.m Tushar Parlikar November 19, 2002 % LEES EMVD Project % % % % % % % XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX clear all; cdc; % Experimental Data from Motor B (Experiment Done by Yihui, Mike, and Tushar on Nov 12, 2002) torqueB=(0:5:50)*0.0071; speedB=[2323 2280 2224 2168 2111 2057 1989 1930 1863 1813 1737]*(2*pi/60); currentB=[0.886 1.252 1.822 2.22 2.75 3.22 3.686 4.145 4.625 5.105 5.585]; voltageB=[18 18 18 18 18 18 18 18 18 18 18]; n=length(torqueB); % Least Squares Fit for Torque-Speed Curve Beta1B=(sum(torqueB.*speedB) - ((sum(torqueB)*sum(speedB)) /n) ) /(sum(speedB.*speedB) ... -(((sum(speedB))^2)/n)) Beta2B=mean(torqueB)-BetalB*mean(speedB) leastsquarestorqueB=BetalB.*speedB+Beta2B; % Least Squares Fit for Voltage Equation XB=[currentB' ,speedB']; BetaB=(inv(XB'*XB))*(XB'*voltageB'); % From MATLAB % BetaiB = -0.0057 % Beta2B = 1.4077 % BetaB % = 0.9160 0.0703 % Calculate and Display Kt,Kw, R, B RB=BetaB(1); KtB=Beta2B*(BetaB(1)/voltageB(1)); KwB=BetaB(2); BB=abs(BetalB)-((KtB*KwB)/RB); fprintf('R = %f Ohms\n', RB); fprintf('Kt = %f N.m/A \n', KtB); fprintf('Kw = %f V/(rad/s) \n', KwB); fprintf('B = %f N.m/(rad/s) \n', BB); - 176 - % R = 0.915954 Ohms % Kt = 0.071631 N.m/A % Kw = 0.070327 V/(rad/s) % B = 0.000245 N.m/(rad/s) % Experimental Data from November 12th - Experiment done by Mike and Yihui torque=(0:5:50)*0.0071; speed=[247 241.4 234 229 223 215.9 208.2 200.1 192.4 185.3 177.2]; current=[0.69 1.184 1.659 2.158 2.594 3.112 3.603 4.08 4.59 5.08 5.66]; voltage=18*ones(1,length(speed)); n=length(torque); % Least Squares Fit for Torque-Speed Curve Beta1=(sum(torque.*speed)-((sum(torque)*sum(speed))/n))/(sum(speed.*speed). -(((sum(speed))^2)/n)) Beta2=mean(torque)-Betal*mean(speed) leastsquarestorque=Betal.*speed+Beta2; % Least Squares Fit for Voltage Equation X=[current' ,speed']; Beta=(inv(X'*X))*(X'*voltage'); % From MATLAB % Betal % Beta2 % Beta % = = = -0.0050 1.2575 0.9930 0.0696 % Calculate and Display Kt,Kw, R, B R=Beta(1); Kt=Beta2*(Beta(1)/voltage(1)); B=abs(Betal)-((Kt*Kw)/R); Kw=Beta(2); fprintf('R = %f Ohms\n', R); fprintf('Kt = %f N.m/A \n', Kt); fprintf('Kw = %f V/(rad/s) \n', Kw); fprintf('B = %f N.m/(rad/s) \n', B); %R = 0.992970 Ohms % Kt = 0.069372 N.m/A % Kw = 0.069611 V/(rad/s) % B = 0.000185 N.m/(rad/s) % Manufacturer's Specifications - 177 - Programs to Analyze the Motor and Motor Drive Tests mfrtorque=[0 50]*0.0071; mfrspeed =[2700 1500]*(2*pi/60); mfrKt=0.07; mfrR=0.99; mfrB=7.64e-5; mfrtorqueadj=(mfrKt/mfrR).*voltage - (mfrB + ((mfrKt^2)/mfrR)).*speed; figure; plot(speed,leastsquarestorque,'b+-',speedB,leastsquarestorqueB,'ro-. speed,mfrtorqueadj,'k-'); title('Torque-Speed Curve at 18V'); xlabel('Speed (rad/s)'),ylabel('Torque (Nm)'); legend('Least Squares Curve Fit Motor A','Least Squares Curve Fit - Motor B','Manufacturer Specification'); 178 XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX7XXXXXXXXXXXXXXXXXXXXXXXXXXX % % % % % Program to Calculate Motor Drive Hysteresis Bands Tushar Parlikar hysteresis.m EMVD Project LEES Laboratory % % % % % % August 7, 2002 % 7XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX cdc; clear x=0:0.01:6*pi; Vref=0.305*sin(x); Vtha=0.047; Vthb=Vtha/6.5041; Vlo=0.1333*Vref-0.1333*6.5041*Vthb; Vhi=0. 1333*Vref+0. 1333*6. 5041*Vthb; figure; plot (x,Vhi, 'k: ',x,0.1333*Vref , 'k-' ,x,Vlo, 'k: '); legend('Hysteresis Band','Current Command'); grid; xlabel('Time') ,ylabel('Amplitude (V)') title('Hysteresis Bands in the Motor Drive with V_{ref}=47mV'); axis([0 8 -0.06 0.06]); %%%%%%%%%%%%%%%%%%%%%%%%%%%%X%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % Tushar Parlikar % % EMVD Project % % LEES Laboratory % % JBvalues.m % % December 15, 2002 % %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% clc;clear % Average the time constants: tauA=0.0595; tau-fA=1.2210; tauB=0.0204; tau.fB=1.1380; % Enter the disk cam and flywheel inertias: Jcam=7.28e-6; Jf=0.00024334; % Calculated J and B for each motor Bma=Jf/(taufA-tauA) Jmb=tauB*Bmb Jma=tauA*Bma-Jcam Bmb=Jf/(taufB-tauB) % Results from MATLAB % MOTOR A % Bma = 2.0950e-004 % Jma = 5.1855e-006 - 179 - Programs to Analyze the Motor and Motor Drive Tests % MOTOR B % Bmb = 2.1773e-004 % Jmb = 4.4418e-006 - 180 - Appendix K MATLAB Simulation of the Adaptive Controller for the EMVD X X Tushar Parlikar Simulation of Adaptive Controller for EMVD cdc; clear all; global k ti current p t0=0; tl=tO; k=1; p=1; tf=0.05; amp=26*pi/180; f=42; alpha=3.46; beta=0.999; lift=0.008; h=lift/(2*sin(beta*pi/2)); KT=0.069; Jm=2*3.54e-6; mv=0.090; K=2*49328.7; Bm=7.639e-4; lambda=4800*pi*f; eta=.2; beta=(KT^2/(Jm^2+mv*alpha^2*h^2))/(((KT^2/(Jm^2+mv*alpha^2*h^2))+1)/2); D=200*(alpha*h); delafl=1e6; thetaO_3=beta*((D + eta)/lambda); thetaO_7=(lambda^2)*((beta*(((sin(alpha*0))2))) thetaO=[amp 0 theta&_3 + eta^2); 1 1 KT/Jm thetaO_7 amp 0]'; options = odeset('MaxStep' ,le-6); tf], thetaO, options); [t, theta]=odel5s ('adapt', y=theta'; motorpos=y(1,:); motorvel=y(2,:); trackerror=y(1,:)-y(8,:); velerror=y(2,:)-y(9,:); motorposd=y(8,:); motorveld=y(9,:); s=velerror +lambda* (trackerror); doubderivthetad=-amp.*((2*pi*f)^2).*cos(2*pi*f.*t); ustar=-doubderivthetad' + lambda. * (trackerror) ; phi=y(3,:); bshat=y(4,:); hlhat=y(5,:); h2hat=y(6,:); gamma_t=y(7,:); figure; subplot(2,1,1) plot (t' ,motorpos); grid; 181 - [to MATLAB Simulation of the Adaptive Controllerfor the EMVD xlabel('Time(s)'),ylabel('Motor Position'); subplot(2,1,2) plot(t',motorvel); grid; xlabel('Time(s)'),ylabel('Motor Velocity'); figure; subplot(2,1,1) plot(t',motorpos,t',trackerror); grid; xlabel('Time(s)'),ylabel('Motor Position and Tracking Error'); subplot(2,1,2) plot(current); grid; ylabel('Motor Current (A)'); figure; plot(t',s,t',phi,t',-1*phi); grid; legend('s','\phi' ,'-\phi') xlabel('Time(s)'),ylabel('Sliding Variable s and Boundary Layer \phi'); figure; subplot(2,2,1) plot (t',h1hat); grid; xlabel('Time(s)'),ylabel('Estimate of h_{1}'); subplot(2,2,2) plot(t',h2hat); grid; xlabel('Time(s)'),ylabel('Estimate of h_{2}'); subplot(2,2,3) plot(t',bshat); grid; xlabel('Time(s)'),ylabel('Estimate of b_{s}'); subplot(2,2,4) plot (t' ,gamma-t); grid; xlabel('Time(s)'),ylabel('Estimate of \gamma (t)'); function dtheta=adapt(t,theta) global k ti t dt=t-tl; tl=t; thetatilda(k)=theta(1)-theta(8); error=theta(1)-theta(8); k=k+1; - 182 % Constant System Parameters Jm=2*3.54e-6; % rotor inertia Bm=7.639e-4; % rotor friction K=2*49328.7; % Spring Constant mv=0.090; % valve, spring, spring divider, etc mass bv=1.2945; % valve friction KT=0.069; % motor torque constant % Desired Trajectory f=60; amp=26*pi/180; thetad=amp*cos(2*pi*f*(t)); derivthetad=-amp*2*pi*f*sin(2*pi*f*(t)); doubderivthetad=-amp* ((2*pi*f) ^2)*cos(2*pi*f* (t)); % NTF function alpha=3.46; beta=0.999; lift=0.008; h=lift/(2*sin(beta*pi/2)); ntf=h*sin(alpha*(theta(1))); % Derivative of NTF function derivntf=(alpha*h)*cos(alpha*theta(1)); % Double Derivative of NTF function dderivntf=-1*((alpha^2)*h)*sin(alpha*theta(1)); % Calculating/Updating Coeffs of the state derivatives den= Jm + (mv*(derivntf^2)); al= Bm + (bv*(derivntf^2)) + (mv*derivntf*dderivntf*theta(2)); X a2= K*derivntf*(ntf); Calculating control law % Constants lambda=4800*pi*f; eta=0.2; beta=(KT^2/(Jm^2+mv*alpha^2*h^2))/(((KT^2/(Jm^2+mv*alpha^2*h^2))+1)/2); D=200*(alpha*h); del-asl=10*Bm/Jm; del_as2=10*K/Jm; del_af1=1e4; del_af2=1e4; bfhat=(((KT^2/(Jm^2+mv*alpha^2*h^2))+1)/2); % Calculations s=(theta(2)-theta(9))+ 2*lambda*(theta(1)-theta(8))... + (lambda^2)*y(size); % for use with "integral form" ustar=-doubderivthetad + 2*lambda*(theta(2)-theta(9))... + (lambda^2)*(theta(1)-theta(8)); % for use with "integral form" if abs((s/theta(3)))>=1 sdelta=s-(theta(3)*sign(s/theta(3))); else sdelta=0; end - 183 - MATLAB Simulation of the Adaptive Controller for the EMVD k_theta=beta*(((abs(theta(4)*(theta(5)*theta(2) + theta(6)*(sin(alpha*theta(1))))... - ustar))*(1-(1/beta)))+ delafl*abs(theta(2)) + del-af2*abs(sin(alpha*theta(1))) + D + eta); kdtheta=beta*(((abs(theta(4)*(theta(5)*theta(9) + theta(6)*(sin(alpha*theta(8))))... - ustar))*(1-(1/beta)))+ delafl*abs(theta(9)) + del-af2*abs(sin(alpha*theta(8))) + D + eta); kbartheta=k_theta - kd_theta + lambda*theta(3); gamma-prime=max([2*lambda;(-lambda*theta(7) + + (sin(alpha*theta(8)))^2)) + (lambda^3)*(/((beta*(theta(9)^2... eta^2)))/theta(7)]); if abs((s/theta(3)))>=1 curr=(1/bfhat)*(theta(5)*theta(2) + theta(6)*sin(alpha*theta(1)))... - (1/(theta(4)*bfhat))*(ustar - ktheta.*sign(s/theta(3)) + beta*gamma-prime.*sdelta); else curr=(1/bfhat)*(theta(5)*theta(2) + theta(6)*sin(alpha*theta(1)))... - (1/(theta(4)*bfhat))...*(ustar - k-theta*(s/theta(3)) + beta*gamma-prime.*sdelta); end % Set up derivative matrix u=KT*curr; dthetal=theta(2); dtheta2= (1/den) * (- (al*theta(2)) - a2 - u); % with control if kdtheta>=(lambda*(theta(3)/beta)) dtheta3=beta*kdtheta - lambda*theta(3); else dtheta3=(kdtheta/beta) - lambda*(theta(3)/(beta^2)); end if abs((s/theta(3)))>=1 dtheta4=0.01*(1./(theta(7)*(ustar ((del-asl*abs(theta(2)) ... + delas2*abs(sin(alpha*theta(1))))*dtheta3*sign(s/theta(3))) + gamma-prime*sdelta))); else dtheta4=0.01*(1./(theta(7)*(ustar ((del-as1*abs(theta(2))... + delas2*abs(sin(alpha*theta(1))))*dtheta3*(s/theta(3))) + gammaprime*sdelta))); end dtheta5=-2.25*theta(7)*theta(2)*sdelta; dtheta6=-2.25*theta(7)*sin(alpha*theta(1))*sdelta; dtheta7=-lambda*theta(7) + (lambda^3)*(1/((beta*(theta(9)^2 + ... (sin(alpha*theta(8)))^2)) + eta^2)); dtheta= [dthetal; dtheta2; dtheta3; dtheta4; dtheta5; dtheta6; dtheta7; derivthetad; doubderivthetad]; - 184 - Appendix L MATLAB Programs to Analyze EMVD Experimental Data This appendix contains MATLAB programs that were used to analyze the experimental data from the various experiments we performed, including those carried out with the PD compensator. %X%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%y % Program to Plot Data obtained in dSPACE and LabView % Final Version % Tushar Parlikar % EMVD Project % emvddataprocess.m % LEES Laboratory at MIT % Completed October 30, 2002 % % % % % % % %X%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% clear all;clc % First, we load the .mat data file from dSPACE Control Desk. load fourthsec003.mat t1=fourthsec003.X.Data; [currenti posi poserrorl]=deal (fourthsec003 .Y.Data); load fourthsec004.mat t2=fourthsec004.X.Data; [current2 pos2 poserror2]=deal(fourthsec004.Y.Data); % Below are figures that I could have plotted but did not end up plotting % figure; % subplot(2,2,1) % plot(tl,posl,'r',tl,poserrorl,'k') % grid;ylabel('Position (rad)');title('Closed-to-Open Transition at 24Hz with... % 250ms holding time'); % axis([-0.02 0.02 -0.6 0.6]); - 185 - MATLAB Programs to Analyze EMVD Experimental Data % subplot(2,2,2) % plot(t2,pos2,'r',t2,poserror2,'k') % grid;ylabel('Position (rad)');title('Open-to-Closed Transition at 24Hz with... % 250ms holding time'); % axis([-0.02 0.02 -0.6 0.6]); % subplot(2,2,3) % plot(tl,currentl) % grid;xlabel('Time(s) '),ylabel('Current % axis([-0.02 0.02 -7 7]); (A)');title('Commanded Motor Current'); % subplot(2,2,4) % plot(t2,current2) % grid;xlabel('Time(s)') ,ylabel('Current (A)');title('Commanded Motor Current'); % axis([-0.02 0.02 -7 7]); % Use the file readbin.m to enter data from LabView into MATLAB [tla,voltagela]=readbin('waveformla.bin'); [tlb,voltagelb]=readbin('waveformlb.bin'); [tla,motcurrla]=readbin('waveform2a.bin'); [tlb,motcurrlb]=readbin('waveform2b.bin'); [tla,posla]=readbin('waveform3a.bin'); [tlb,poslb]=readbin('waveform3b.bin'); [tla,curcommla]=readbin('waveform4a.bin'); [tlb,curcommlb]=readbin('waveform4b.bin'); powerla=motcurrla.*voltagela; powerlb=motcurrlb.*voltagelb; % Plot the Data figure; subplot(2,2,1) plot(tla,(posla/10),t2,poserror2) grid;xlabel('Time(s)'),ylabel('Position and Position Error(rad)');title('Position'); axis([-0.015 0.02 -0.6 0.6]); subplot(2,2,2) plot(tlb,(posib/10),tl,poserrorl) grid;xlabel('Time(s)'),ylabel('Position and Position Error(rad)');title('Position'); axis([-0.015 0.02 -0.6 0.6]); subplot(2,2,3) plot(t2,poserror2) grid;xlabel('Time(s)'),ylabel('Position Error (rad)');title('Position Error'); axis([-0.015 0.02 -0.02 0.02]); subplot(2,2,4) plot(tl,poserrorl) grid;xlabel('Time(s)'),ylabel('Position Error (rad)');title('Position Error'); axis([-0.015 0.02 -0.02 0.02]); - 186 - figure; subplot(2,2,1) plot(tla,curcommla*2) grid;xlabel('Time(s) '),ylabel('Current (A) ');title('Commanded Motor Current'); axis([-0.015 0.02 -7 7]); subplot(2,2,2) plot(tlb,curcommlb*2) grid;xlabel('Time(s)') ,ylabel('Current (A) ');title('Actual Motor Current'); axis([-0.015 0.02 -7 7]); subplot(2,2,3) plot(tla,motcurrla) grid;xlabel('Time(s) '),ylabel('Current (A) ');title('Commanded Motor Current'); axis([-0.015 0.02 -7 7]); subplot(2,2,4) plot(tlb,motcurrib) grid;xlabel('Time(s) ') ,ylabel('Current(A) ');title('Actual Motor Current'); axis([-0.015 0.02 -7 7]); % Calculate Average Power for "UP" transition for i=1:length(tia) if tla(i)<=-0.01082 indexla=i; end if tla(i)<=0.0130 indexlb=i; end end avgholddownpower=0; for j=i:indexla-1 avgholddownpower=avgholddownpower+( (power1a(j))/(index1a-1)); end avgtruppower=0; for j=indexla:indexib avgtruppower=avgtruppower+ ((poweria(j)) /(indexib-index1a)); end avgholduppower=0; for j=indexlb+1: length(t1a) avgholduppower=avgholduppower+( (power1a(j) )/(length(t1a)-indexlb-1)); end % Calculate Average Power for the "DOWN" transition for i=1:length(tlb) if tlb(i)<=-0.01012 index2a=i; end if tlb(i)<=0.0122 index2b=i; - 187 - MATLAB Programs to Analyze EMVD Experimental Data end end avgholddownpower2=0; for j=1:index2a-1 avgholddownpower2=avgholddownpower2+((powerib(j))/(index2a-1)); end avgtruppower2=0; for j=index2a:index2b 2 avgtruppower2=avgtruppower +((powerib(j))/(index2b-index2a)); end avgholduppower2=0; for j=index2b+1:length(tib) avgholduppower2=avgholduppower2+ ((powerib (j)) /(length(tlb) -index2b-1)); end figure; subplot(2,2,1) plot(tla,powerla) grid;xlabel('Time(s)'),ylabel('Power(W)');title('Instantaneous Motor Power'); axis([-0.015 0.02 -200 200]); subplot(2,2,2) plot(tib,powerlb) grid;xlabel('Time(s)'),ylabel('Power(W)');title('Instantaneous Motor Power'); axis([-0.015 0.02 -200 200]); subplot(2,2,3) plot([tla(1) tla(indexla-1) tla(indexla) tla(indexlb) t1a(index1b+1) tla(length(tla))],[avgholddownpower avgholddownpower avgtruppower avgtruppower avgholduppower avgholduppower]) grid;xlabel('Time(s)'),ylabel('Power(W)');title('Average Motor Power'); axis([-0.015 0.02 0 20]); subplot(2,2,4) plot([tlb(1) tlb(index2a-1) tlb(index2a) tib(index2b) tib(index2b+1) tib(length(tlb))],[avgholddownpower2 avgholddownpower2 avgtruppower2 avgtruppower2 avgholduppower2 avgholduppower2]) grid;xlabel('Time(s)'),ylabel('Power(W)');title('Average Motor Power'); axis([-0.015 0.02 0 20]); % This is another set of plots that I didn't want to display % % subplot(3,2,3) % % plot(tia,voltagela) X% grid;xlabel('Time(s)'),ylabel('Voltage(V)'); % % axis([-0.015 0.02 -35 35]); % % subplot(3,2,4) % % plot(tlb,voltagelb) % % grid;xlabel('Time(s)'),ylabel('Voltage(V)'); % % axis([-0.015 0.02 -35 35]); 188 - X November 6, 2002 % Tushar Parlikar % LEES EMVD PROJECT % openlooptests.m % This program plots the open loop responses for % A) The Motor and Motor Drive % B) The EMVD % In the second part of the program, the controller designed in rltool % is used to plot the open and closed loop control responses. X Data Analysis for Experiments done by Tushar and Mike/Yihui on November 4 and 5, 2002 clear all; cdc; Td=0.00; % dSPACE sampling delay time % A) MOTOR + MOTOR DRIVE - the end of the stroke wl=2*pi*([1 10:10:150]); f=wl./(2*pi); position=0.1*[10.44 9.08 6.48 4.32 3.76 3.16 2.72 2.1 1.68 1.6 1.38 1.208 1.07 .96 .55 .54]; current=[1.0 1.3 1.5 1.8 2.5 3 3.5 3.5 3.5 4 4 4 4 4 2 2]; tphase=([200 31.6 18.08 14.16 11.52 9.44 8 7 6.12 5.48 4.8 4.32 4 3.68 3.48 3.28]*0.001)-(2/25000); G1=position./current; P1=-1*tphase.*f*360; s = tf([1 0], [1]); Kt=0.051; J=3.5e-6; B=7.64e-4; System=Kt/(J*s^2+B*s); [Hlmag,Hlphase]=bode(System,wl); i=l:length(wi) H1(i)=H1mag(:,:,i); for Hlp(i)=Hlphase(:, :,i)-(wl(i)*Td); end figure;semilogx(wl,20*log10(G1) , '--',wl,20*log10(H1)) ;grid;title('Magnitude versus ... Frequency - Motor + Motor Drive'); legend('Experimental','Curve Fit'); figure;semilogx(wl,P1,'--',wl,Hlp);grid;title('Phase versus Frequency - Motor... + Motor Drive'); legend('Experimental','Curve Fit'); % B) MOTOR + MOTOR DRIVE + VALVE ASSEMBLY - the middle of the stroke w=[6.28 62.8 94.2 125.6 138.2 150.8 157.1 163.4 169.6 176.9 182.2 188.5 207.3 226.2 251.3 288.74 314.15 377 439 502 628]; t=[56 9.6 7.2 6.8 6 6 5.8 6 6 5.8 5.6 5.92 7.04 7.92 8.64 9.04 8.48 7.04 4.4 4.24 2.4]*1e-3 - 2/25000; G=[1.11/4 2.22/4 2.59/4 2.76/4 2.77/4 2.6/4 2.56/4 2.66/4 2.79/4 2.88/4 3.02/4 3.07/4 4.44/4 4.24/4 - 189 - MATLAB Programs to Analyze EMVD Experimental Data 3.8/4 2.36/4 1.62/4 1.2/6 .71/6 .68/6 .35/6]/10; P=-360*(w.*t)./(2*pi); s = tf([1 0], [1]); wn=220; zeta=0.2942; G2=10^(-24/20)*wn^2/(s^2+2*zeta*wn*s+wn^2); [G2mag,G2phase]=bode(G2,w); for i=1:length(w) H(i)=G2mag(:,:,i); Hp(i)=G2phase(:,:,i)-(w(i)*Td); end figure;semilogx(w,20*loglO(G),'--',w,20*loglO(H));grid;title('Magnitude versus... Frequency - Motor + Motor Drive + Valve Assembly'); legend('Experimental','Curve Fit'); figure;semilogx(w,P,'--',w,Hp);grid;title('Phase versus Frequency - Motor... + Motor Drive + Valve Assembly'); legend('Experimental','Curve Fit'); % Compensator Design - one compensator for both transfer functions... %C1=zpk([-506.4 -189.9],[0 -8477 -2781],980033.3479); C2=zpk([-1709],[-8198],1918.2384); figure;bode(System,'k:',C2*System,'b',(C2*System)/(1+(C2*System)),'r-.',{10 10e5});title('EMVD at the Ends of the Stroke'); legend('Plant Loop Gain','Plant*Controller Loop Gain','Closed-Loop System');grid;grid; figure;bode(G2,'k:',C2*G2,'b',(C2*G2)/(1+(C2*G2)),'r-.',{lO 10e5});title('EMVD at the Middle of the Stroke'); legend('Plant Loop Gain','Plant*Controller Loop Gain','Closed-Loop System');grid;grid; - 190 - %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % Tushar Parlikar % % % % % EMVD Project % LEES Laboratory % December 15, 2002 %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% clc;clear % Average time constants: tauA=0.0595; tau_fA=1.2210; tauB=0.0204; tau-fB=1.1380; % Enter the disk cam and flywheel inertias: Jcam=0.728e-6; Jf=0.000024334; % Calculated J and B for each motor Bma=Jf/(tau-fA-tauA) Jmb=tauB*Bmb Jma=tauA*Bma-Jcam Bmb=Jf/(tau-fB-tauB) % Results from MATLAB % MOTOR A % Bma = 2.0950e-004 % Jma = 5.1855e-006 % MOTOR B % Bmb = 2.1773e-004 % Jmb = 4.4418e-006 191 - MATLAB Programs to Analyze EMVD Experimental Data % This Program was Obtained from David D. Wentzloff, a former SM student in LEES. % Function to read LabVIEW binary file % [t, data] =readbin() % readbin.m %%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%%% % Routine for loading experimental data in program %% DO NOT ALTER function [t ,data]=readbin(wavefile) if nargin==O wavefile=uigetfile('*.bin' ,'Select file'); end [wavefid,message]=fopen(wavefile, 'r','ieee-be'); , 'float32' ,O); fclose(wavefid); dataM=fread(wavefid, [2,inf] t=dataM(1,:); data=dataM(2,:); return 192 - XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX % Program to Plot Data for seating velocity obtained using LabView % Tushar Parlikar % EMVD Project % LEES Laboratory at MIT % emvdseatvel.m % Completed January 17, 2003 XXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXXX clear all;clc % Use the file readbin.m to enter data from LabView into MATLAB [t,valpos]=readbin('valvepos.bin'); [t,valvel]=readbin('valvevel.bin'); Lt,filtvalpos] =readbin('valveposdsp.bin'); [t,motpos]=readbin('motorpos.bin'); % Conversion Factors factorl=1.06; factor2=21.2; factor3=1/10; % 1V =1.06mm for the position data A 1V=21.2cm/s for the velocity data % 1V=0.i radians for the motor position data % Convert to correct units valpos=filtvalpos*factorl-.13; % position in mm % Note: the -2 is subtracted because the oscilloscope signal was off-center. valvel=valvel*factor2; % velocity in cm/s motpos=motpos*factor3; A motor position in rad % Nonlinear Mechanical Transformer -[tia,thetaA]=readbin('ch2a.bin'); Compliance Check [t1b,xA]=readbin('ch3a.bin'); z-pred = 8e-3./2.*sin(3.46.*thetaA./10); figure; plot(thetaA/10,xA,'k', thetaA/10, z_pred*factor1*1e3-.2,'b'); xlabel('\theta (radians)'),ylabel('z (mm)') grid; axis([-0.505 0.505 -4.6 4.2]); title('A comparison of the theoretical and experimental NTF characteristic relation') legend('Actual (experimental) relation', 'Expected relation'); % Plot for seating velocity figure; subplot(2,1,1) plot(t,valpos,'k') grid;xlabel('Time(s)'),ylabel('Valve Position (mm)');title('Measurement of Seating Velocity'); 0.025 -5 5]); - 193 - axis([-0.025 MATLAB Programs to Analyze EMVD Experimental Data subplot(2,1,2) plot(t,valvel,'k') grid;xlabel('Time(s)'),ylabel('Valve Velocity (cm/s)'); axis([-0.025 0.025 0 205 ]); gtext('Valve Seated'); - 194 - Bibliography [1] J. G. Kassakian, "The Role of Power Electronics in Future 42V Automotive Electrical Systems," in EPE-PECM Conference, pp. 3-11, Dubrovnik, Croatia, Sept. 2002. [2] J. G. Kassakian, H-C. Wolf, J. M. Miller, and C. J. Hurton, "The Future of Automotive Electrical Systems," in IEEE Workshop on Power Electronics in Transportation,pp. 312, Dearborn, MI, October 24-25, 1996. [3] W. S. Chang, An Electromechanical Valve Drive Incorporating a Nonlinear Mechanical Transformer. Ph.D. thesis proposal, Massachusetts Institute of Technology, 2001, unpublished. [4] M. B. Levin, and M. M. Schlecter, "Camless Engine," SAE Technical Paper Series, Paper 960581, 1996. [5] P. Barkan, and T. Dresner, "A Review of Variable Valve Timing Benefits and Modes of Operation," SAE Technical Paper Series, Paper 891676, 1989. [6] T. Ahmad, and M. A. Theobald, "A Survey of Variable-Valve-Actuation Technology," SAE Technical Paper Series, Paper 891674, 1989. [7] J. Heywood, Internal Combustion Engine Fundamentals. Mc-Graw Hill, 1988. [8] W. S. Chang, T. A. Parlikar, J. G. Kassakian, and T. A. Keim, "An Electromechanical Valve Drive Incorporating a Nonlinear Mechanical Transformer," in SAE World Congress, Detroit, MI, March 2003, in press. [9] W. S. Chang, T. A. Parlikar, M. D. Seeman, D. J. Perreault, J. G. Kassakian, and T. A. Keim, "A New Electromagnetic Valve Actuator," in IEEE Workshop on Power Electronics in Transportation,pp. 109-118, Auburn Hills, MI, October 24-25, 2002. [10] W. S. Chang, J. G. Kassakian, and T. A. Keim "An Electromechanical Valve Drive Incorporating a Nonlinear Mechanical Transformer," US ProvisionalPatent 60/322,813, September 17, 2001. [11] F. Pischinger et al., "Electromechanical Variable Valve Timing," Automotive Engineering International,1999. [12] F. Pischinger et al., "Arrangement for Electromagnetically Operated Actuators," US Patent #4,515,343, 1985. - 195 BIBLIOGRAPHY (13] "Camless BMW Engine Still Faces Hurdles," Automotive Industries, pp. 34, October 1999. [14] R. Flierl, and M. Kliting, "The Third Generation of Valvetrains - New Fully Variable Valvetrains for Throttle-Free Load Control," SAE Technical Paper Series, Paper 200001-1227, 2000. [15] M. A. Theobald, B. Lesquesne, and R. R. Henry, "Control of Engine Load via Electromagnetic Valve Actuators," SAE Technical Paper Series, Paper 940816, 1994. [16] "Renault Research," Automotive Engineering International,pp. 114, March 2000. [17] "Emission Control," Automotive World, pp. 10-15, April 2000. [18] S. Butzmann, et al., "Sensorless Control of Electromagnetic Actuators for Variable Valve Train," SAE Technical Paper Series, Paper 2000-01-1225, 2000. [19] M. Gottschalk, "Electromagnetic Valve Actuator Drives Variable Valvetrain," Design News, November 1993. [20] R. R. Henry, and B. Lesquesne, "A Novel, Fully-Flexible, Electro-mechanical Engine Valve Actuation System," SAE Technical Paper Series, Paper 970249, 1997. [21] R. R. Henry, and B. Lesquesne, "Single-cylinder Tests of a Motor-driven, Variable-valve Actuator," SAE Technical Paper Series, Paper 2001-01-0241, 2001. [22] J-J. Slotine, and W. Li, Applied Nonlinear Control Prentice-Hall, 1991. [23] M. F. Schlecht, "Time-Varying Feedback Gains for Power Circuits with Active Waveshaping," in IEEE Power Electronics Specialists Conference, pp. 15-22, July 1981. [24] A. M. Stankovic, D. J. Perreault, and K. Sato, "Synthesis of Dissipative Nonlinear Controllers for Series Resonant DC/DC Converters," IEEE Transactions on Power Electronics, Vol. 14, No. 4, pp. 673-682, July 1999. [25] A. B. Plunkett, "A Current-Controlled PWM Transistor Inverter Drive," IEEE/IAS Annual Meeting Conference Record, pp. 789-792, 1979. [26] D. M. Brod, and D. W Novotny, "Current Control of VSI-PWM Inverters," IEEE Transactions on IndustrialApplications, Vol. IA-21, No. 4, pp. 562-570, June 1985. [27] J-J. Slotine, and J. A. Coetsee, "Adaptive Sliding Controller Synthesis for Nonlinear Systems," InternationalJournal of Control, Volume 43, 6, 1986. [28] J. G. Kassakian, H-C. Wolf, J. M. Miller, and C. J. Hurton, "Automotive Electrical Systems Circa 2005," IEEE Spectrum, pp. 22-27, August 1996. 196 -