Vous êtes sur la page 1sur 156

INFORMATION TO USERS

This manuscript has been reproduced from the microfilm master. UMI films the text directly from the original or copy submitted. Thus, some thesis and dissertation copies are in typewriter face, while others may be from any type of computer printer. The quality of this reproduction is dependent upon the quality of the copy submitted. Broken or indistinct print, colored or poor quality illustrations and photographs, print bleedthrough, substandard margins, and improper alignment can adversely afreet reproduction. In the unlikely event that the author did not send UMI a complete manuscript and there are missing pages, these will be noted. Also, if unauthorized copyright material had to be removed, a note will indicate the deletion. Oversize materials (e.g., maps, drawings, charts) are reproduced by sectioning the original, beginning at the upper left-hand corner and continuing from left to right in equal sections with small overlaps. Each original is also photographed in one exposure and is included in reduced form at the back of the book. Photographs included in the original manuscript have been reproduced xerographically in this copy. Higher quality 6" x 9" black and white photographic prints are available for any photographs or illustrations appearing in this copy for an additional charge. Contact UMI directly to order.

UMI
A Bell & Howell Information Company 300 North Zeeb Road, Ann Arbor MI 48106-1346 USA 313/761-4700 800/521-0600

NONLINEAR CONTROL AND OPERATION OF DC TO DC SWITCHING POWER CONVERTERS

BY PALLAB MIDYA
B. Tech., Indian Institute of Technology Kharagpur, 1988 M. S., Syracuse University, 1990

THESIS
Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Electrical Engineering in the Graduate College of the University of Illinois at Urbana-Champaign, 1995

Urbana, Illinois

UMI Number: 9812942

UMI Microform 9812942 Copyright 1998, by UMI Company. All rights reserved. This microform edition is protected against unauthorized copying under Title 17, United States Code.

300 North Zeeb Road Ann Arbor, MI 48103

UMI

UNIVERSITY OF ILLINOIS AT URBANA-CHAMPAIGN THE GRADUATE COLLEGE

MAY 1995

W E H E R E B Y R E C O M M E N D T H A T T H E T H E S I S BY

PALLAB MIDYA
ENTITLED. NONLINEAR CONTROL AND OPERATION OF

DC TO DC SWITCHING POWER CONVERTERS BE A C C E P T E D I N P A R T I A L F U L F I L L M E N T O F T H E R E Q U I R E M E N T S THE DEGREE O F . DOCTOR OF PHILOSOPHY FOR

ALA=^
r\J -HVf9U\*vv\"@\
Max)

Director of Thesis Research Head of Department

Committee on Final Examination!

Lk
-T^/^S.
\/ywvJ> P - K . VC
t Required for doctor's degree but not for master's.

Chairperson

NONLINEAR CONTROL AND OPERATION OF DC TO DC SWITCHING POWER CONVERTERS


Pallab Midya, Ph.D. Department of Electrical and Computer Engineering University of Illinois at Urbana-Champaign, 1995 Philip T. Krein, Advisor ABSTRACT Conventional control of switching power conversion is based on a linear model of the system, which limits its scope and performance. This thesis addresses a family of control and operation issues that arise from the nonlinear nature of the switching power conversion. The focus is on dc to dc power converters with a small number of states. Nonlinear noise analysis of conventional feedback has been performed to evaluate this important aspect of converter performance. Alternative schemes are developed that reduce noise susceptibility, reduce effects of source and load disturbance, and improve system operation. A nonlinear control scheme called sensorless current mode (SCM) control has been developed that emulates current mode control without current sensing. SCM has very significant advantages in dynamic range and noise susceptibility. Optimal control approaches to switching power conversion has been explored. Optimization of a nonmonotonic tuning problem has also been explored. A generalized tuning scheme has been developed and implemented for practical converters with excellent results.

ui

ACKNOWLEDGEMENTS
This work was supported in part by the University of Illinois Power Affiliates Program and in part by a contract with the Sorensen Company, Paxton. IL. The latter also provided a very challenging research problem that contributed to the thesis. I would like to thank my advisor Professor P. T. Krein for his expert guidance. I am most thankful to him for also providing me with the opportunity to work independently. His insight into theoretical issues contributed greatly to the success of the thesis. I would like to thank him for the long hours we spent in the laboratory; they were most educational. I would also like to thank the Ph.D. committee, Professor P. R. Kumar, Professor P. W. Sauer, and Professor B. S. Song for their comments that helped broaden the perspective of the thesis. I would like to thank Dr. S. Dharanipragada, Dr. G. W. Wright, Dr. S. Ponnapalli, and Mr. J. Locker for numerous technical discussions that were very helpful. Thanks are due to Mr. A. Kulkarni for maintaining the computer system and for software support. The staff in the Electrical and Computer Engineering Store were very helpful in providing parts for the experimental work. I would like to especially thank my wife Lipika for bearing with me while I was writing the thesis and for helping out with the numerous figures in the thesis.

iv

DEDICATION
This thesis is dedicated to my parents, Mr. Gunakar Midya and Mrs. Runu Midya, who gave me the most important ideas in life.

TABLE OF CONTENTS CHAPTER


1 INTRODUCTION 1.1 Control Issues in Power Converter 1.2 Description of a Switched Power Converter 1.3 Noise Analysis of Switching Control and its Mitigation 1.4 Optimal, Lyapunov and Energy Based Control 1.5 Nonmonotonic Tracking 1.6 Conclusions 2 LITERATURE SURVEY 2.1 Small Signal Analysis 2.2 Large Signal Analysis 2.3 Linear Feedback Control Schemes 2.4 Nonlinear Control Schemes 3 NOISE ANALYSIS OF THE PWM PROCESS 3.1 Model of the PWM Process 3.2 Description of the Noise Process 3.3 Duty Ratio Analysis 3.4 Interpretation of the Results 3.5 Conclusions 4 NOISE ANALYSIS OF PWM IN A CONTROL LOOP 4.1 Description of the PWM Control Loop 4.2 Time Domain Noise Analysis 4.3 Frequency Domain Analysis 4.4 Internal Noise Effects 4.5 Conclusions 5 SENSORLESS CURRENT MODE CONTROL 5.1 Sensorless Current Mode Control Theory 5.2 Discontinuous Mode 5.3 Flux Balance Control of Transformer Coupled Converter 5.4 Experimental Verification 5.5 SCM Performance Comparison 6 OPTIMAL CONTROL APPROACHES 6.1 Introduction 6.2 Approximate Optimal Control Methods 6.3 Lyapunov Function Control
VI

PAGE
1 2 4 8 10 11 12 13 13 14 15 16 20 21 23 26 28 32 33 34 37 40 46 47 49 50 54 56 58 65 73 73 78 82

6.4 6.5 6.6 6.6

Gradient Approximation to Optimal Control Fast Switching Optimal Control Experimental Results Conclusions

87 91 96 98 100 101 105 107 110 112 113 113 116 117 119 123 135

7 AUTO TUNING SCHEMES 7.1 Tuning Theory 7.2 Maximum Power Point Tracker 7.3 Experimental Results 7.4 Active Filter 7.5 Conclusions 8 CONCLUSIONS 8.1 Important Analytical Results 8.2 Performance Improvements 8.3 Future Work REFERENCES APPENDIX A: CIRCUIT DIAGRAMS APPENDIX B: DERIVATION OF NOISE IN PWM

APPENDIX C: DERIVATION OF NOISE IN CLOSED LOOP PWM ... 140 VITA 146

Vll

CHAPTER 1 INTRODUCTION
Power electronic systems control energy transfer between a source and a load by switch action. Switch action makes these systems discontinuous and fundamentally nonlinear. Conventional control is based on a linear model of a system, which limits the scope and performance of conventional methods for nonlinear applications. The usual advantages of linear state feedback control are not obtained because of the inherent nonlinearity of power converters. Secondary nonlinear effects such as saturation in magnetics, uncontrolled switch action in diodes and nonlinear input sources further constrain the scope of linear control in switching power converters. This thesis addresses a family of control and operation issues that arise from the nonlinear nature of switching power conversion. The focus is on dc to dc power converters with a small number of states. Three types of disturbances are considered here: disturbances at the source, disturbances at the load and disturbances due to noise. Noise, for example, affects converter control in unusual ways that have not been analyzed in previous work. Nonlinear noise analysis of conventional feedback has been performed to evaluate this important aspect of converter performance. Alternative schemes are developed that reduce noise susceptibility, reduce effects of source and load disturbance and improve system operation. The performance criteria for a power converter, as defined by user specifications, do not correspond to linear performance bounds. Thus a linear feedback scheme that satisfies the performance requirements around a given operating point does not satisfy the same 1

requirements around another operating point. Nonlinear controllers can be developed to meet given performance requirements over the entire range of operation. 1.1 Control Issues in Power Converters A power converter ideally transfers energy between a power source and a load, without loss, such that the output is unperturbed by transients in the source or the load. Usually the source and the load are at different current and voltage levels, and there is a range of values over which each is allowed to vary. The power converter output is required to stay within a specified tolerance level in response to these changes. The ability to handle large high speed disturbances is one measure of the performance of a power converter. In the absence of source and load transients, a disturbance due to noise makes the steady-state operation of a power converter nontrivial. The ability to maintain a constant operation and output in the presence of noise is another measure of the performance of a power converter. The switching frequency, hence the cycle time, of the switch is usually fixed. The ontime of the switch is a variable and is the principal control input to the system. Duty ratio is defined as the ratio of the on time to the switching cycle time. Duty ratio is a number between zero and one. The digital signal to the switch is called the switching function. It determines the switch action and is obtained by pulse width modulation (PWM) of an analog duty ratio signal. It varies the on-time keeping the cycle time a constant. The PWM process is a highly nonlinear process which is susceptible to noise. The analysis of this process allows one to predict performance degradation due to noise. It also

enables us to compare noise susceptibility of different systems and to synthesize control schemes that are robust in their resistance to noise. States that are used in feedback in power converters are capacitor voltage and inductor current. Voltage sensing is simple and relatively robust to noise. Current sensing is often done by introducing a small resistor in the path of the current. The voltage drop in this resistor is then sensed. To maintain high efficiency, this drop is a very small fraction of the voltages in the system. This makes the current signal prone to noise. To improve noise robustness, control schemes are presented here that do not require current sensing. Instead current estimators are built that do not introduce noise. Optimal control that minimizes a cost function is a reasonable target for nonlinear control. For example, some cost functions to be optimized are recovery time following a transient, overshoot in the output or the energy stored, or error during a transient. Owing to the switching action, a converter's output does not converge to an equilibrium point but moves around it. This produces the output ripple, which makes unbounded the traditional cost function based on the integral of the output error squared. An optimal control approach, modified to take these factors into account, is presented. Power and energy are fundamentally nonlinear quantities. A power converter can be controlled by monitoring the energy or the power in the system. This would address the nonlinear aspect of power conversion in the most fundamental way. Based on this idea, control schemes involving direct computation of a Lyapunov function or of energy stored have been developed.

A tracking problem focused on power is described as an example of a static optimization problem. A nonlinear tuning scheme has been developed that handles such tracking problems. Examples include output ripple cancellation and maximum power tracking of a solar panel. These nonmonotonic tuning problems originate in nonlinear!ties such as inductor saturation and sources with nonlinearities. The nonlinear tuning scheme developed here is quite general and could be used in a wide variety of static optimization problems. Chapter 2 is a survey of existing control techniques, both linear and nonlinear, used in switching power converters. Some of these methods are widely used in commercial power converters while others are relatively new but show considerable merit. The relationship between the methods developed here to those in the survey will be examined. The similarities and the differences between the methods will be brought out. In particular, the relative advantages and disadvantages of these methods will be analyzed, and the scope for performance improvement will be outlined. 1.2 Description of a Switching Power Converter Physically, a power converter is a network of switches and passive storage elementsinductors, capacitors and transformers. The passive elements store energy in order to smooth power transfer without loss. Figure 1.1 shows a boost converter, which is a typical example of a switching power converter for interfacing a dc source to a dc load. When the switch is on, the inductor stores energy from the source. When the switch is turned off, this energy is fed to the load and the output capacitor through the diode. The energy stored in the capacitor provides the load power when the switch is on. The output voltage obtained is higher than the input voltage by a factor of 1/(1-D).

/SMI
L

D>
DIODE C R
/

V.

SWITCH

Figure 1.1 Boost converter 1.2.1 Analytical model of converter There are idealized models of power converters with ideal passive components and switches. Reference [1] provides analytical models for typical power converters. Control action of a power converter must be performed solely through choice of switch position. Two or more switch configurations are possible for any power converter. In a particular switch configuration, the converter is a network of passive components connecting the input source to the load. Neglecting component nonlinearity, each configuration can be characterized by a set of linear differential equations. The system behavior in the i* configuration is determined by the differential equation: X(r) = AX{t) * fl (1.1)

The overall system can be characterized by a combination of the linear differential equations and the switching functions Hs. For the ^configuration H; is unity and the other terms are zero.

X(t) = / / ( 0 ) ( A * ( / ) + B)

(1.2)

Often the switching action is set to a predetermined frequency. Then the control input in each cycle is in effect the duty ratio D. In conventional power converter control analysis and design, duty ratio is taken as a linear input to the system. Reference [2] is a representative of the feedback design procedures that are used in the industry. Note that in general Equation (1.2) is a time-varying system. In general, this system does not satisfy a Lipschitz condition, since A(t,X) is neither continuous in time nor in state. However, the system is passive and lossy, and the system states (X) are continuous.

rosin
L C V.
in

ffrom
Vout R
i

L
c

Vout

v.
in

CONTINUOUS MODE (1)

CONTINUOUS MODE (2)

roi%1
L C r V. in R
v

out

DISCONTINUOUS MODE (3)

Figure 1.2 Different configurations of boost converter

Consider again the boost converter. It has two switches: a transistor, which is actively controlled, and a diode. Figure 1.2 shows the three possible configurations for this converter. In normal operation when the transistor is off the diode is on, and vice versa, and only two possible switch configurations occur. If the transistor is on, H, is 1 and H2 is 0. If the transistor is off, H, is 0 and H2 is 1. The A and B matrices are as follows.

1
\
=

1
0

RC 0 0

A, =

~RC ~C

-1 0
L

a, = s2 = v.
L

(1.3)

For the purpose of analysis, the system is normalized such that the desired operating point is [1,1]. The desired operating point before normalization for a boost converter is [Vref,V0Ut2/(RVref)]. The normalization is performed by dividing the state by the reference state. 1.2.2 Discontinuous mode of operation Under certain conditions, usually light loads, the diode current can go to zero and the diode can switch off prior to transistor turn-on. In the boost converter, this third configuration is described by the following: 1 0 RC 0 0

^3

*3

(1.4)

The different configurations are numbered from 1 to 3 and the circuits corresponding to each configuration are shown in Figure 1.2 It is possible to take the system from configuration 2 to configuration 1 and back by turning the switch on or off. However, the transition between configurations 2 and 3 is determined by the load. If a converter operates in configuration 3, it is said to be in discontinuous mode. This is often avoided because of the control problems it introduces through state reduction. However, the nonlinear control techniques developed here handle this mode without degradation in performance. Thus, the nonlinear control techniques offer new alternatives for operation of power converters. 1.3 Noise Analysis of Switching Control and Noise Mitigation PWM is a convenient technique for the generation of switching signals in commercial power converters. This process modulates any noise in the analog input into a timing error in the switching signal. Linear small signal analysis does not adequately capture this phenomenon. Detailed nonlinear analysis of the phenomenon is presented here to elucidate the time domain and frequency domain effects of this noise modulation. Various conventional techniques are analyzed and compared for noise susceptibility. Perhaps the most effective way to mitigate noise effects is to feed back only those states that can be sensed with low noise. In many dc-dc converters, this translates to sensing voltages but not currents. Information about the currents can be reconstructed from the voltages in the system. Power loss associated with the current sensing is avoided by this method. Chapter 3 deals with the noise analysis of the PWM process. This is a time domain analysis of the PWM process is presented. The statistics of the duty ratio variation are explored. An unusual result is uncovered in the analysis: it is found that the average duty ratio

is modified in the presence of noise even when the noise has an average of zero. The variation of the duty ratio is determined by the rms values of the noise and its derivative. The nonlinear nature of the noise modulation is also reflected in the statistics of the duty ratio. The probability density and distribution functions of the duty ratio are obtained in closed form. These theoretical results match Monte Carlo simulation data and experiments in which commercial PWM ICs were subjected to a noise generator. Chapter 4 deals with noise analysis in closed-loop PWM systems. Here conventional techniques such as voltage mode control, peak current mode control and current mode control with ramp are tested for noise susceptibility. The standard deviation of the duty ratio in noisy conditions is obtained in terms of ramp slopes and noise power. Frequency domain analysis of this process predicts that the noise is aliased to subsynchronous frequencies (frequencies lower than the switching frequency). This quantifies for the first time a commonly observed phenomenon in power electronics, namely, that a noisy converter has subsynchronous oscillations. The theoretical results are compared to Monte Carlo simulation data and to experimental results in which a PWM loop is subjected to random noise from a noise generator. There is very good agreement both in the time domain and frequency domain results. Chapter 5 describes a sensorless approach to current mode feedback control. This sensorless current mode (SCM) control method minimizes noise problems by avoiding current feedback. The current is estimated from the intermediate voltages in the system. The method retains most of the advantages of conventional current feedback approaches but has superior noise rejection properties. For low voltage applications where current sensing is harder this

method is particularly useful. A family of 2 V output converters has been built using the SCM control method. In certain transformer coupled converters, there is a possibility of transformer failure due to dc offset which saturates the transformer core. The flux in the transformer has positive and negative excursions, and the transformer flux must be balanced to keep the excursions centered about zero. Magnetic flux in a transformer is not readily measurable, and conventional control estimates the flux by the transformer current. SCM control estimates the flux and enforces flux balance using the transformer voltages. A flux balance approach is presented for push-pull forward converters. 1.4 Optimal, Lyapunov and Energy Based Control In converters where all states, including currents, are measurable, the dynamic optimal control problem of disturbance can be approached. Here the integral of the output error squared is defined to be the cost function. Minimizing this cost function is useful in converters that must respond to rapid load and source transients. The finite horizon optimal control problem and the infinite horizon optimal control problem are considered. The control law obtained is complex and difficult to implement directly. Boundary control for power electronics [3] is a large signal control method. In boundary control, a switching boundary is defined, which is a surface in state space. The switching boundary divides the state-space into regions. The switch action is determined by comparing the system state to the boundary. A switching boundary is denied which corresponds to the optimal control action. This transforms the optimal control law to a boundary control problem and simplifies the implementation.

10

Instead of minimizing an integral of the output error, it is simpler to explicitly minimize a Lyapunov function. This would not necessarily minimize the integral of the output error. However, this ensures large signal stability and a well-defined response to large signal transients. Control based on the quantity of stored energy is another way to address the problem of optimal power transfer in a power converter. These techniques exploit the fundamental nature of power converters as opposed to attempting to fit a linear feedback scheme to a nonlinear circuit. Chapter 6 approaches the optimal control problem for switching power converters. Switching boundaries are developed that correspond to the finite horizon and the infinite horizon optimal control problem. Implementation of some simple optimal control problems has been attempted. The issue of system complexity and other limitations for optimal control are discussed. 1.5 Nonmonotonic Tracking In the steady-state optimization problem, the output being optimized could be a nonmonotonic function of the controlled variable. Often the output is dependent on the temperature, the load impedance, and other unknown and variable quantities. Thus, it is very useful to have an automatic tuning scheme that would take the system to the desired operating point using only input and output information. A generalized tuning scheme has been developed that uses the correlation between changes in the input and corresponding changes in the output to tune the operating point. When the system reaches the desired operating point, the correlation goes to zero and the system converges. This corresponds to the point at which the derivative of the output with

11

respect to the input is zero. This tuning scheme is appropriate for any tuning problem which has a single maximum or minimum. A variety of tuning problems in power electronics fall into this category. A tuning scheme based on correlation usually requires an excitation input. The switching action perturbs all the states and provides this excitation. Thus, this tuning scheme is appropriate for switching power applications. Chapter 7 introduces the tuning theory and establishes criteria for stability and convergence. Two tuning problems, adaptive ripple cancellation and maximum power point tracking, are implemented using this method. Adaptive ripple cancellation [4] uses a feedforward scheme to estimate the current ripple shape and the tuning scheme to determine the amplitude. This method is used to establish a new form of an efficient active filter for high performance applications. 1.6 Conclusions This thesis analyzes a family of steady-state and dynamic nonlinear operation and control problems of dc to dc switching converters and develops appropriate nonlinear control techniques. A family of converters has been designed to test these techniques. Chapter 9 summarizes the important theoretical results, the performance improvements and areas for future work. Novel analytical results have been obtained in the area of noise analysis of PWM, feedforward control and large signal optimal control. Significant improvements have been attained in the areas of noise immunity, transient response, ripple reduction and automatic tracking.

12

CHAPTER 2 LITERATURE SURVEY


Switched dc-dc power converters are nonlinear, discontinuous and time-varying systems. The only control input is the switch duty ratio. Switch action is a bounded and discrete input. This makes the analysis, simulation and control of switched power converters a complex and challenging problem. Much of the prior analysis of power converters is small signal analysis, which is obtained by linearizing the nonlinear system. The design of the feedback structure for switched power converters is often limited to linear feedback. This is an artificial constraint and often a restricting one. Some nonlinear control schemes have been developed that exhibit significant performance improvements over those for linear control schemes. In addition there are nonlinear tracking or optimization problems that are beyond the scope of linear feedback control. 2.1 Small Signal Analysis The circuit level model of switched power converters, including switch characteristics, is complex and offers limited scope for analysis. The model is simplified by assuming ideal switching, and this analysis provides insight into converter action in open loop [1]. There are excellent small signal analysis techniques for converter action that include the action of the feedback network [5,6,7]. Reference [5] combines sampled data modelling with a pole-zero model of the control loop. It provides a very accurate small signal model of current mode control. Reference [6] provides an accurate small signal model for average current mode control. Reference [7] analyzes existing models for switched power converters, including both linear and nonlinear models. 13

The interaction between the converter and the feedback network introduces nonlinearities in addition to the nonlinearities due to switch action. Furthermore, the disturbances are often large. Thus, the accuracy and validity of small signal analysis are limited. 2.2 Large Signal Analysis A large signal analysis technique that provides the most intuitive understanding of converter action is geometric control. In this technique the converter action is observed as the movement of the system state in the state space. The control law is expressed as a boundary in state space. The switch is toggled when the system state crosses the boundary. Reference [8] presents geometric control and does a large signal analysis of linear and nonlinear control laws in state space. This provides a pictorial representation of the operation of a switched power converter. Novel control laws have been developed as boundaries in state space, and even traditional control laws can be interpreted using the idea of boundary control [8]. The optimal control approaches developed in this thesis have been interpreted in terms of boundary control. Averaging theory has been developed as an alternative to small signal approximation [9]. Averaging analysis allows one to predict some converter characteristics that are lost in small signal analysis. The added complexity of averaging analysis is justified because of this feature. Reference [9] develops a framework for applying averaging theory to the analysis of switched power converters and shows that this analysis recovers an important phenomenon such as output ripple that is not captured by linear analysis.

14

PWM (pulse width modulation) is a process used to convert from a duty ratio signal to the switch function. The characteristics of its noise properties and noise interaction with the modulation process are well-understood from a communication point of view. Reference [10] provides a detailed analysis of band limited Gaussian noise. References [11,12] present detailed analyses of the noise behavior of the PWM process and other pulse modulation schemes. The analyses are linear and the quantity of interest are signal to noise ratio. There is no comparable existing analysis from the point of view of power electronics in which the duty ratio and switching function are the noise issues rather than the transmission channel noise. 2.3 Linear Feedback Control Schemes Linear feedback considers duty ratio to be a continuous and unbounded input to a linear system. The duty ratio signal is obtained by a linear combination of the output and the states of the system. Perhaps the simplest of feedback schemes is direct output feedback. For voltage output converters, this is called voltage mode control and is used extensively. Current mode control is a relatively recent control technique that has greatly improved transient performance [13]. A ramp signal is introduced in current mode control to maintain a fixed switching frequency and avoid subharmonic instability. The design of feedback networks for current mode control has been optimized for performance. Current mode control, however, is prone to noise and difficult to implement for low voltage converters. To improve noise performance, average current mode control has been introduced [1]. Further, some of the noise associated with current sensing has been avoided using alternative techniques [14]. System complexity

15

has been increased for reduced noise in the system. In this thesis, sensorless current mode control has been developed that shares certain advantages of current mode control but has improved noise performance and dynamic range and does not require current sensing. 2.4 Nonlinear Control Schemes Stability criteria and performance requirements for switched power converters for most electronic loads are different for small signal and large signal disturbances. These requirements do not readily translate to linear criteria, as does pole placement. This provides an incentive to introduce nonlinear control schemes. Nonlinear control techniques have been developed that address the nonlinear control problem of switched power conversion. Some of these techniques attack the problem using nonlinear feedforward and feedback techniques [15,16,17]. Reference [16] implements a nonlinear control law by sensing the inductor current and integrating it in a capacitor. The charge in the capacitor is used to determine the switch position. This provides a simple control law similar to current mode control. Reference [17] analyzes a nonlinear difference equation and uses the results to construct control laws for dc to dc converters. The control laws are shown to have good small signal behavior in theory and simulations. Others attempt the nonlinear control problem using formal Lyapunov theory. Reference [18] develops a formal Lyapunov function based on the energy in a converter and constructs control laws that minimize the Lyapunov function. The optimal control problem has also been addressed for some converters [19]. Reference [19] computes an error based on linear state feedback in addition to the output feedback . This scheme has been simulated for a buck

16

converter and was shown to improve performance. These nonlinear techniques have much potential. However, much of commercial power converter control relies on linear feedback [2]. One important class of nonlinear optimization problems attempts to maximize or minimize a nonmonotonic function of the states. These problems require nonlinear control schemes. An autotuning technique that has been used traditionally for online tuning is the cross-correlation technique. It has been implemented in the digital domain for slow tuning of controllers [20]. Reference [20] considers a general linear system and tunes the gain of the PID feedback for optimum performance. The scheme imposes a pseudo-random binary sequence noise signal to probe the system. The algorithm is implemented in the digital domain and is computationaly intensive. In the analog domain, tuning of a scalar quantity is feasible. This has been attempted for the maximum power transfer problem [21]. In reference [21]. a simple analog tuning scheme has been developed and implemented. A small sinusoidal signal is injected into the duty ratio and the response is used to tune the system. The scheme maximizes the charging current to the battery and thus achieves maximum power point tracking. It is difficult to guarantee the stability of such a nonlinear control algorithm. Even so, stability has been achieved by constructing a very slow tuning process whose dynamics do not interact with the other dynamics of the system. These systems track the maximum power transfer problem with a delay of about a second. A fast tracking scheme has been developed in Chapter 7 that tracks with a delay of only 2 milliseconds. References [22,23] describe a high efficiency inverter being driven from a solar array through a maximum power tracking inverter. The maximum power point is obtained by

17

introducing a small step and computing the power digitally. Response times on the order of a few seconds are obtained. Reference [24] describes a system that computes the maximum power point by calculating a small signal model of the system. The maximum power point is again obtained by using the response to a small perturbance to the system. Another control problem of considerable interest is output ripple cancellation which is implemented by injection of a cancellation current, which is an inverse of the ripple current. There are a number of active filters that cancel the harmonics introduced into the power system by inverters and rectifiers [25,26,27,28]. The cancellation current is obtained using feedback, feedforward, or a combination of the two. These controllers have to react to the power line frequency, and this allows for the computation and cancellation of individual harmonics [28]. By contrast, the calculation and cancellation of a high frequency ripple require a circuit with small delay in computation. Reference [29] describes a converter in which coupled inductors are used to obtain and cancel current waveforms. The cancellation is achieved almost perfectly for a high performance converter. Since the cancellation is based on coupled inductors, the precision requirements on the magnetics are extreme. Reference [30] describes a simple feedback active filter that can cancel ripple for a variety of power converters. It obtains the cancellation current signal by monitoring the output. The scope of this filter is limited because the ripple at the output of a high performance converter is hard to monitor accurately. As the cancellation progresses, the signal being monitored disappears, causing signal to noise problems.

18

Reference [31] describes an active filter that cancels line harmonics at the input of an inverter. Since it has to respond to the line frequency and its higher harmonics, it is very slow and is implemented digitally. The active filter developed here suppresses ripple frequencies on the order of 100 kHz. The gain of the cancellation current is optimized by minimizing the cross-correlation between the cancellation current and the output ripple. There is a significant reduction in hardware by implementing this in the analog domain.

19

CHAPTER 3 NOISE ANALYSIS OF THE PWM PROCESS


In steady state, disturbances due to noise perturb converter operation and distort the output. The converter output is a function of the input voltage (or current) and the duty ratio. Noise can affect the converter output through either of these two variables. The propagation of noise from the input to the output has been studied extensively. Quantitatively, the effect is represented by audio susceptibility, and converters are designed to have low or zero audio susceptibility. Audio susceptibility is a small signal analysis that does not consider the duty ratio variation and nonlinear!ty of the PWM process. Therefore, it has a limited ability to predict the noise performance of a power converter under operating conditions. The variation in duty ratio due to noise is very significant to the converter output. This is a common phenomenon that has not been quantified in previous work. The variation in duty ratio often degrades performance, and it is important to be able to quantify this phenomenon. Understanding of this process would help in the comparison of noise performance of feedback schemes and the synthesis of robust schemes. The PWM process uses a comparison between an analog signal and a ramp time base to generate a switching function. This process has been analyzed from noise performance criteria in the communication literature for large SNR (signal to noise ratio) [32] as well as for small SNR [33,34]. This analysis of PWM assumes uniform sampling, which is an oversimplification in the context of many power converters. In a power converter, using a conventional PWM IC, the sample time is the same as the switch on-time. Thus variation of

20

the sample time is of special interest. A rigorous noise analysis of PWM in the context of power electronics is required. In previous work, noise analysis of PWM-based dc-dc converters almost always addressed small-signal performance measures such as audio susceptibility [35] and control-tooutput response [36]. These give an indication of the robustness of a given control circuit but do not give any insight into the important nonlinear noise effects. Consider, for instance, that audio susceptibility and other measures do not show dramatic distinctions between voltagemode and current-mode control loops. This contrasts with qualitative observations made by many designers, who report clear distinctions in the noise behavior of different control methods [37,38]. Noise effects would be expected to depend on switching frequency, the nominal duty ratio, and control loop properties; this is not captured in small-signal analysis. In this chapter a nonlinear model of the PWM process is constructed and the PWM process is analyzed. Bandlimited Gaussian noise is imposed and the variation in duty ratio is estimated. The cumulative probability distribution function and the probability density function are obtained analytically. 3.1 Model of the PWM Process A schematic of the PWM process, based on a commercial converter control IC, is illustrated in Figure 3.1. The analog input is compared to an internally generated ramp. A latch prevents multiple transitions per cycle. The standard practice is to use the comparator to reset the latch and a clock signal to set the latch. The set and reset operations could be exchanged, and the results obtained here apply with minor modifications.

21

AV
RAMP ANALOG SIGNAL

^v COMPARATOR R

LATCH

SWITCHING FUNCTION Q
i

CLOCK

Figure 3.1 Schematic of PWM IC The switch is set by a synchronous clock signal and reset when the analog input signal crosses the ramp signal. The presence of a noise pulse affects the reset timing and, hence, the duty ratio. Figure 3.2 shows the switching function and associated waveforms obtained with and without noise in the analog input.
RAMP ANALOG SIGNAL RAMP ANALOG SIGNAL

-r>rZ-vJA
1 0 SWITCH ING FUNCTION i 0 1 1

1
SWITCHING FUNCTION

::

CLOCK

:':'

CLOCK

(a) Waveforms without noise

(b) Waveforms with noise in analog input Figure 3.2 PWM waveforms
22

3.2 Description of the Noise Process Let us assume that noise interferes with the analog signal at the PWM input. As a reasonable model, let us further assume that the noise has a Gaussian probability density function and limited bandwidth. The bandwidth limitation is realistic given that low pass filtering is normally applied explicitly or implicitly by the comparator speed limitation. Further, the noise process is assumed to be stationary. This means that the statistics of the noise process (mean, standard deviation, etc.) are not a function of time. The noise and its derivative are assumed to have zero mean and a defined standard deviation. Further, the noise and its derivative are assumed to be uncorrected. 3.2.1 Definitions Random variables: Duty ratio = X Noise = Y Noise derivative = Z Duty ratio in absence of noise = D0 Noise standard deviation = cY Noise derivative standard deviations a z Ramp amplitude = V^, Switch cycle time period = Ts The ramp is linear and has zero reset time. In the absence of noise, the duty ratio is a function of the signal and ramp level. After noise is added to the process, the duty ratio is stochastic, and it is possible to find only its statistical nature. Noise is added to the analog

23

input. The analysis, simulation and the experiment use this model for the noise. The schematic is shown in Figure 3.3.

AV\
RAfA?
1

+\^

COMPARATOR R

LATCH

SWITCHING FUNCTION Q

rui
WHITE NOISE

CLOCK

ANALOC SIGNAL

-m

LOW PASS FILTER

Figure 3.3 Noise injection to the PWM process Probability density functions (pdf) are represented as lower case 'p' while cumulative probability distribution functions (PDF) are represented by upper case letter 'P.' Recall that a cumulative probability distribution function P(X) is the probability of the given random variable being less than X. Since X cannot be less than minus infinity, P(-=) is always zero. Similarly, X is always less than infinity and P(>) is always unity. P(X) is an increasing function of X. The probability density function p(X) is a measure of the likelihood of the random variable being equal to X. The density function p(X) is a derivative of the cumulative distribution function P(X). The total area under the density function p(X) is always equal to unity. Gaussian noise is uniquely characterized by its mean and standard deviations. The Gaussian probability density function p(X) with zero mean and standard deviation c is given by the following equation. 24

xx)=__L_f2tr
Tv/2T

(3.1)

Figure 3.4 is a plot of cumulative and probability distribution functions for the standard Gaussian distribution (mean=0, standard deviation =1). Note that pdf is the density function and PDF is the distribution function.

0.8 0.6 0.4 0.2


0

'pdf 'PDF'

-4

-3

-2

-1

Figure 3.4 Gaussian distribution: cumulative and density functions Let us define two separate Gaussian distributions. The noise variable is denoted Y, while its derivative is denoted Z. The corresponding pdf s for Y and Z are as follows.
-r:

p,m=

.TV

(3.2)

o-/2rT

25

-z1

p2(Z)=

e^

(3.3)

3.3 Duty Ratio Analysis The duty ratio D is a random variable. Consider the probability that D is between X and (X+AX). The probability of this event can be expressed as [P(X+AX)-P(X)]. This event occurs if the analog signal is larger than the ramp in the interval (0 to X) and the analog signal is smaller than the ramp voltage in the interval (X to X+AX). For the analysis of duty ratio, we have to introduce conditional probability. Given two events A and B, the conditional probability of A given B is defined as [P(AIB)], which is defined as the ratio of the probability of both A and B occurring relative to the probability of B occurring. This definition is valid for discrete probability as well as for continuous probability functions. P(Af]B)=P(B)P(A\B) OR P(A\B)=P(Af]B) P(B) (3.4)

Define A to be the event in which the analog signal is less than the ramp voltage in the interval (X to X+AX). Define the event B in which the analog signal has not become less than the ramp in the interval (0 to X). The switch is reset in the interval (X, AX) if and only if both events occur. Thus P(AnB) is by definition [P(X+AX)-P(X )]. Let Q denote the value of the conditional probability P(AIB) of the comparator toggling in the interval (X to X+AX) given that there has not been a transition between 0 and X.

26

P(X^6X) -P(X) =( 1 -P(X))Q

(3.5)

The event Q is the event that the switch is reset in the interval (X to X+AX). Since the switch is reset only once in a cycle, the switch has not been reset in the interval (0 to X). This constrains the noise (Y) to certain values only. Starting from these values of noise (Y), there are further constraints on the noise (Y) and its derivative (Z), such that the switch is rest in the interval (X to X+AX). The value of Q is obtained by calculating the probability of all noise values (Y) and noise derivative values (Z) that will result in a switch transition in the interval (X to X+AX). This is an area integral in noise and its derivative in the Y-Z plane. It is given by the integral in the Y-Z plane of the density functions p,(Y) and p2(Z) over values of Y and Z that result in a transition in the interval(X to X+AX). The noise (Y) determines the difference between the ramp and the sum of the signal and the noise. The noise derivative (Z) determines the rate at which the ramp and the sum of the signal and the noise approach each other. The area integral computes the probability of the switch being reset in the interval(X to X+AX).

g- ' W * -

M, V AX

(3.6)

The value of Q is substituted in Equation (3.6) and the differential equation obtained is solved analytically. It turns out that results can be expressed in closed form as follows. The detailed derivation is provided in Appendix B.
27

P(X)=1 -[1 -f,(y*/X-Do))]"'

(3-7)

M, is a dimensionless quantity obtained from the computation of Q. It is a function of the ramp rate (V^/T^) and the standard deviation of the noise derivative oz. It is a measure of the high frequency content of the noise normalized to the ramp rate.
RP T a

M=P2(Roz)+J. sJ2KR

where R=

(3.8)

sz

The quantity M, is in effect the modulation index of the PWM noise process. M, can also be interpreted as the ratio of expectation of the switch transition due to the ramp slope plus noise to the expectation of a transition due to ramp slope only. In the limiting case when the noise is zero, the quantity M, is equal to unity. However, in all other cases it is larger than unity. 3.4 Interpretation of the Results 3.4.1 Shift in average duty ratio For values of M, significantly larger than 1, the PDF of the duty ratio is quite different from the PDF of the noise. Also we expect a shift in the average duty ratio. The mean is difficult to calculate. The existence of a bias in the distribution is detected by a shift in the median value of the duty ratio (variable X). The median corresponds to the duty ratio with cumulative PDF equal to 0.5.

28

P(%) = 0 . 5 = 1 -[l-P^V^-D^f-

(3-9)

Define XM to be the median duty ratio. In absence of any shift, XM is equal to D0 . With noise the expression for XM is

P,(Vs/1(XAf-D0))=l-(0.5)I/w' <0.5

(3-10)

From the plot of cumulative PDF of the Gaussian distribution (Figure 3.2) we can see that all points where the PDF is less than half correspond to values less than the median. Thus the median duty ratio has moved from D0 to a new value less than D0. This result can be qualitatively explained as follows. As noise bandwidth increases, the number of uncorrected noise samples increases. Any noise spike that causes a ramp crossing causes a false trigger as noise bandwidth increases. Thus, there is a higher chance of a false trigger. Since a false trigger resets the switch, we expect the average and median duty ratios to drop in the presence of noise. Consider the dual of this PWM scheme in which the comparator is used to set the switch and the clock is used for the reset. For this scheme, an increase in the average duty ratio is expected. The shift in average duty ratio due to noise has numerous implications. For a singleloop system, noise would change the average duty ratio. For example, in an inverter this would create a dc imbalance. A dc imbalance in an ac system can cause saturation in the

29

magnetics and prevent successful operation. In a system with an outer integral loop, a dc offset in the outer loop would compensate for this effect. 3.4.2 Comparison with simulation and experimental data Figures 3.5(a) and 3.5(b) show the pdf of the duty ratio in the presence of noise with dq=0.5 . The analytical formula is compared to Monte Carlo simulation. Figure 3.5(a) corresponds to small Gz (standard deviation of noise derivative) and consequently, the noise process is fairly linear. Figure 3.5(b) corresponds to large o~z, and the mean duty ratio has shifted significantly. The analytical formulae predict this shift accurately.

'THEORY' SIMULATION'

THEORY SIMULATION'

1 -

0 'eooflfl 0.2 0.4 0.6 0.8

48oi 1

o 'oooecra 0 0.2

0.4

J 0.6

^ o o o o o o ooooo oe 0.8

(a) Small Noise Bandwidth X-axis: Duty ratio, Y-axis: pdf of duty ratio

(b) Large Noise Bandwidth

Figure 3.5 Probability density function: theoretical and Monte Carlo simulation results

30

Figures 3.6(a) and 3.6(b) show the PDF of the duty ratio in the presence of noise. The theoretical results are compared to the experimental values. The nominal duty ratio in the absence of noise is 50%. Figure 3.6(a) corresponds to a small noise derivative (Mpl); consequently the mean duty ratio is unchanged due to the presence of noise. Figure 3.6(b) corresponds to the large noise derivative (M,l) and the mean duty ratio has shifted significantly. The analytical formula predicts this shift accurately. Appendix A describes the circuit used to obtain the experimental data used.

: a
: 6

: <

(a) Small Noise Bandwidth


'THEORY'

(b) Large Noise Bandwidth X-axis: Duty ratio, Y-axis: 1- PDF of duty ratio Figure 3.6 Cumulative probability distribution: theoretical and experimental results 31

3.5 Conclusions In this chapter the noise analysis of the PWM-based power analysis has been undertaken. The statistical nature of the duty ratio variation has been explored. Analytical results have been obtained for the probability density function of the duty ratio and compared to Monte Carlo simulation and experimental results with good agreement. It is observed that in the presence of noise the duty ratio exhibits variation from its nominal value. This is expected from a small signal analysis of the PWM process. In addition to this, there is a shift in the average duty ratio, even though the noise has zero mean. This result has not been observed or analyzed in the literature. Small signal analysis cannot capture this phenomenon since it is caused by the nonlinear nature of the PWM process. The analysis introduced here quantifies PWM noise effects in the context of power electronics. The analysis is applicable to any converter topology that uses PWM to obtain its switching function. This chapter dealt with the PWM process without an outer loop. The following chapter builds on these results and analyzes the PWM process inside an integral loop. This brings the results into the practical domain; conventional feedback schemes are then compared.

32

CHAPTER 4 NOISE ANALYSIS OF PWM IN A CONTROL LOOP


Noise is an important concern in the design of any switching power converter. The pulse width modulation (PWM) process is the "gateway" by which noise influences converter behavior. To track a desired output voltage or current, the PWM process is embedded inside an analog control loop. In particular, for dc-dc converters, this uses either voltage or current feedback. Each type of loop adds its own noise properties. Current-mode and voltage-mode methods are compared here on the basis of how they process noise signals. The results provide a basis for understanding nonlinear noise effects. Control loops that exhibit a "memory" effect can propagate noise from a given switching period to subsequent periods - with detrimental results. In previous work, noise analysis of PWM-based dc-dc converters almost always addressed small-signal performance measures such as audio susceptibility [35] and control-tooutput response [36]. These measures provide some measure of the relative merit of various control schemes. However, there is no analysis to predict the noise performance in a power converter under operating conditions. Consider, for instance, that audio susceptibility and other measures do not show dramatic distinctions between voltage-mode and current-mode control loops. This contrasts with qualitative observations made by many designers, who report clear distinctions in the noise behavior of different control methods [37]. It is expected that noise effects depend on the switching frequency, the nominal duty ratio, and the control loop properties; this is modelled accurately by the following analysis.

33

The analysis here also predicts subharmonics and increased ripple in the output due to noise in the PWM process: phenomena often observed in noisy power converters. The noise spectrum below the switching frequency is obtained, and there is good agreement with the values predicted by theory. The issue of internally coupled noise due to commutation is also addressed. Conditions for proper operation of switching power converters in the presence of internally coupled noise are obtained. The analysis provides quantitative data for variance (or standard deviation) of the duty ratio. In the frequency domain, subharmonics in the output spectrum are calculated. The results obtained match experimental and simulation data. 4.1 Description of the PWM Control Loop The schematic of the PWM process with an outer control loop is shown in Figure 4.1. The feedback signal is the inductor current or the capacitor voltage, corresponding to current mode or voltage mode feedback, respectively. The analysis is done for current feedback, and the results for voltage feedback are obtained as a special case.
RAMP

SWITCHING FUNCTION
=>l

POWER CONVERTER

ANALOG FEEDBACK SIGNAL

Figure 4.1 Schematic of PWM process with outer control loop

34

In current mode control, the inductor current is used for feedback. The current has a waveform like a ramp with one slope when the switch is on and another when the switch is off. The switch is reset when the current becomes more than the reference current. Current mode control can be divided into two categories depending on whether an external ramp is used to generate the duty ratio. In peak current mode control, the reference current is a dc quantity. In current mode control with an external ramp, the reference current is a sawtooth. Figure 4.2 shows the waveforms in current mode control with an external ramp.

Figure 4.2 Current mode control waveforms 4.1.1 Noise propagation effect If noise alters the switching time, the duty ratio perturbation will alter the starting signal value during subsequent cycles as control action corrects the operation. This effect propagates into subsequent cycles. The basic process, based on an idealized current-mode control ramp, is illustrated in Figure 4.3. In the figure, a brief noise spike changes the timing of the comparator trip point. The duty ratio is disturbed, but in this particular case recovers to its nominal value after about three cycles. 35

In voltage mode feedback the output voltage is used for feedback. The output voltage is approximately a constant dc quantity because of the action of the output filter. By setting the current ramp rate to zero, the current mode control results can represent the results for voltage mode control.
NO NOISE RAMP ANALOG SIGNAL WtTHNOISE

m
SWITCHING FUNCTION , ' 1

NO NOISE

I i
1

i ,
1

WITH NOISE

I o I

Figure 4.3 Noise propagation waveforms 4.1.2 Definitions Ramp slope = SR Ramp amplitude = VRP Rate of change of current feedback signal with switch on = SN Rate of change of current feedback signal with switch off = SF Nominal duty ratio = D0 Noise value at the trip point of cycle N = EN Standard deviation of EN = aE Standard deviation of injected noise = o\ Difference between duty ratio in cycle N and nominal duty ratio = FN 36

Standard deviation of FN = o> Standard deviation of duty ratio = CTD Switching frequency = Fs Switching cycle time period = Ts 4.2 Time Domain Noise Analysis The PWM ramp starts from a value of zero at the beginning of each cycle and rises linearly until it reaches the end of the cycle, when it is reset to zero. The current signal is the negative of the inductor current and ramps down when the switch is on and ramps up when the switch is off. The switch turns on at the beginning of each cycle and is reset when the current equals the ramp value. The expression for the current at the switch transition is given by the following: /((AN%)=%S* (4.1)

Taking the difference between the equations at cycle N and cycle N+l, the following recursive equation is obtained.

/WV^MF+o^-ff)
Noise is introduced into the system and the expression for the current is modified. W+DJTJ+EN-DJfK

(4-2)

(4.3)

The duty ratio is obtained in recursive form as follows:

37

AM

(4.4)

The error in duty ratio (FN) is obtained in closed form as follows.

4.2.1 Standard deviation of duty ratio The deviation of the duty ratio can be obtained in terms of the individual noise values (EN). The individual noise samples are uncorrected and the standard deviation of the duty ratio can be obtained in closed form. Appendix B has the detailed mathematical derivation. The standard deviation of the duty ratio is

<v "' r/VW

i+r

where K= R~ F

(4.6)

"V^

In the case of voltage mode PWM, the quantity 'K' tends to unity. This makes the duty ratio variation the same as in the open-loop case. This result is similar to Equation (3.7). In current mode control with appropriate ramp, the value of 'K' can be set to zero and the standard deviation of the duty ratio variation is V2 times the open-loop standard deviation. In

38

the case of peak current mode with no ramp, the quantity 'K* is a function of the duty ratio. The result for peak current mode control is function of the average duty ratio is given by: 2d-DJ 1-2A,

~0=- % + S J \

(4.7)

Figure 4.4 shows the standard deviation of the duty ratio as obtained from Equations (4.6) and (4.7) compared with Monte Carlo simulation results. Note that for voltage mode control and current mode control with optimal ramp the results are not a function of the nominal duty ratio. For peak current mode control the performance degrades as the nominal duty ratio increases, and the system becomes unstable at D0=l/2.
THEORYSIMULATION' P E CURMXT KOOE ,

X axis: Nominal duty ratio Yaxis:Normalized aD


CURReKT HOOE WITH OPTIMAL (UUCP u o

VOLTAGE HOOe

0.2

04

06

08

Figure 4.4 Standard deviation of duty ratio: theoretical and experimental results 4.2.2 Effect of noise on ripple The ripple value represents variation of the load voltage in steady state. In the absence of any significant noise effects, the ripple is obtained from the switching waveform and the
39

output filter transfer function. The PWM process aliases high frequency components of noise in the control signal to subsynchronous frequencies. The additional ripple due to noise can be obtained from the noise spectrum of the switching function and the output filter transfer function. The output filter is designed for rejecting the switching components and higher harmonics. It normally does not provide significant attenuation at subharmonics of the switching frequencies. Thus a small duty ratio variation, due to noise, results in a

disproportionate increase in the ripple. 4.3 Frequency Domain Analysis The energy spectral density of a random process can be obtained from the autocorrelation function of the process. From Equation (4.5), we know that the duty ratio variation in any cycle is correlated to the duty ratio variation of previous cycles. An approximate expression can be obtained for the spectrum by treating the duty ratio as a discrete variable and by obtaining the autocorrelation in the discrete time domain and then taking its Fourier transform to obtain the energy spectral density in the frequency domain. While these results are not exact, they provide useful approximate closed-form expressions for the spectrum. The autocorrelation function and the power spectral density are a Fourier transform pair. The autocorrelation function is the expected value of the product of the signal with a delayed version of itself. The function is defined as follows. Rx(x)=<x(t)x(t+x)> (4.8)

40

The autocorrelation of the error (FN) at the origin is simply the power of the signal:

RJ0)=-1

5:

(4.9)

For a time difference of 'm' cycles the correlation can be obtained from Equation (4.4).

R^mT^K^lLLoy2 ^1

_ J _ r/(^^):

(4.10)

The power spectral density is obtained by taking a Fourier transform of the discrete time autocorrelation function. The detailed mathematical derivation is provided in Appendix C.

2av2(l-cos0) Y SJf) = where 0 =2rc//F, (l+K2-2KcosQ)Ts(SR+SN)2

(4.11)

The maximum of this spectrum occurs at half the switching frequency, which corresponds to the first subharmonic. The energy spectrum at this frequency is 4a2 __I
(!+#%+.?/

SJFS IT) =

(4.12)

Figures 4.5(a), 4.5(b) and 4.5(c) show the noise spectrum of the switching function for different values of K. The simulation results are compared to the analytical result given by Equation (4.11). The curve for K=0 corresponds to current mode control with optimal ramp.

41

The curve for K=I corresponds to voltage mode. The curve with K=-0.5 corresponds to peak current mode control with average duty ratio of 1/3.

(a) Peak current mode X axis: f/F5, Y axis:Noise spectrum (dB)


0 5

0 4

05

06

0 7

08

0 9

(b) Current mode (optimal ramp) X axis: f/Fs, Y axis:Noise spectrum (dB)

(c) Voltage mode X axis: f/Fs, Y axis:Noise spectrum (dB)

0 1

0 2

0 J

0 4

0 ,

0 '

0 7

0 8

Figure 4.5 Noise spectrum: theory and Monte Carlo simulation 42

The Figures 4.6(a) and 4.6(b) show the noise spectrum obtained experimentally from peak current mode control with nominal duty ratios of 0.25 and 0.45, respectively. The spectrum is measured in the subsynchronous frequency range. The switching frequency is 1.0 kHz and the spectrum is measured from 250 Hz to 750 Hz.
!

i
i

"TMFOPV

-20 -40 -60 -80

1 '
1-

ii H i 1 _|T' i
! !

I 1 i'

(a) D0=0.25 - 2 5 X axis: f/Fs, Y axis:Noise spectrum(dB) 0 -20 Vum

0.5

0.75

T XEQEY:

-j*tpr
-40
i

-60
-80

_ _

(b) D0=0.45 X axis: f/Fs. 0.25 Y axis:Noise spectrum(dB)

0.5

0.75

Figure 4.6 Noise spectrum for peak current mode: theory and experimental data

43

Figures 4.7(a) and 4.7(b) show the spectrum for current mode control with optimal ramp. The nominal duty ratios are 0.25 and 0.75, respectively. Current mode control without inherent noise is realized by using a current estimator, then Gaussian noise with known statistics is added.
-20
i

i iiOiiy-l

-40 ! ! . 1 -80

jytffy

>i ArtMJjr

-.iaijl

-60

r
,

1
!

1
5

1 1
1
0-75

(a)D 0 =0.25 X axis: f/Fs, Y axis:Noise spectrum(dB)


o -

25

XdeORY.' -20

-40

mjwp

nim u i > i \wiinMJili.iJii I L . I iAktlUljfo

fmmmfflfftt(pl ^Mipvffe!}&#
h i I4I AA AAJFAAI 0.4 0.45 0.5 0.55 0.6 0.65 0.7 0 75

utet

-60

-80

1
(b) D0=0.75 0.25 0.3 0.35

X axis: f/Fs, Y axis:Noise spectrum(dB) Figure 4.7 Noise spectrum current mode(optimaI ramp) : theory and experimental data

44

Figures 4.8(a) and 4.8(b) show the noise spectrum obtained for voltage mode control. The nominal duty ratios are 0.25 and 0.75, respectively. In all cases, the results are compared to theoretical values obtained from Equation (4.11), and agreement is good. Circuit diagrams of the experimental setup are shown in Appendix A. 1 o

... . "TMFnRV -.

-20

H rfkttkrtftilk
-40

;^;'^%i^y^j^^^trki!^fU^iVii.^.'ii tfc 1,, i v in,r r | " J ri F|fff/ '[l|f'r Vl ' i7' '!"
ff r r
i !

-60

ii
4

!il:
I i 1 .

'
!

-80

1 (a)D0=0.25 X axis: f/Fs, Y axis:Noise spectrum(dB)


0

i
0.5

25

l
i

r-

f ' T - 'F.ORY'

-20

-40

k_LI

WW frflW Iff r rw fflOTW in n M 1 ""' f '1


4iVWnHr

it i kk

ILlULi jiltJJL/ \\kf\i W r W , tihu&LkAh


1

-60

1 i

"r

i "irmr
i

i
!

"! ]
1

i
1

-80

1
i

(b) D0=0.75 X axis: f/Fs, 0.25 Y axis:Noise spectrum(dB)

0.5

0.75

Figure 4.8 Noise spectrum for voltage mode: theory and experimental data

45

4.4 Internal Noise Effects Switching action in the converter is itself a significant source of noise. This internal noise is distinct from thermal noise, which is nondeterministic, and external noise, which originates outside the converter. Internal noise affects the control signal by electromagnetic coupling between the switched current and the control circuit. In addition to this coupled noise, there is also a spike in the current signal ( defined If ) due to stray capacitance in the inductors [37]. This noise is present only at the switch turn on-time. However, it could be large enough to cause the switch to reset. The following analysis shows the noise margin of current mode control and the range of possible duty ratios. Noise (nTs) =/^Z I(nTs) (4.13)

'Z' is the stray coupling coefficient from the power circuit to the control circuit. There is no false trigger due to commutation noise only if the commutation noise is less than the difference between the signal and the ramp. For a nominal duty ratio of D0 , ramp slope of SR, and current ramp rate of SN, this difference is D0(SR+SN)TS. Thus the condition for preventing a false trigger at the beginning of the cycle is given by

The current at the beginning of the cycle has been approximated by the nominal dc current 1^. If the duty ratio required is less than this minimum, either in steady state or during a transient, the duty ratio will be set to zero. If the nominal duty ratio of the converter is less than this value, the converter will not operate properly. This result also illustrates that current 46

mode control has an internal noise problem at large current levels. This is also compounded by dynamic range problems as observed in [37]. This analysis and reference [37] point to the fact that current mode control, in particular, peak current mode control, is not an appropriate choice if the current variation is small compared to the dc current level. To reduce loss in the inductor, small current variation is considered to be good design practice. 4.5 Conclusions To obtain good response in steady state, PWM conversion is embedded in an outer integral loop. This modifies the noise properties of the overall system. The noise enhancement/ suppression properties of this outer loop were examined. In particular, voltage mode, peak current mode, and current mode with optimal ramp were considered. For the same noise level, the three systems were compared at various nominal duty ratios. It was observed that the voltage mode has the least noise enhancement and the peak current mode has the most noise enhancement. The results confirm qualitative observations that current mode controls are "noisier" than voltage mode controls. Analytical formulae for these schemes have been obtained and compared with simulation results. In the absence of noise, the spectrum of a switching function is identically zero below the switching frequency. The PWM process aliases any noise in the system to subharmonic frequencies. Analytical expressions have been obtained for these subharmonics. It is observed that the spectra for voltage mode control, peak current mode control and current mode control with optimal ramp are quite different both in amplitude and distribution. There is good correlation between the analytical results, simulation and experimental data.

47

The variation in duty ratio leads to an increase in output ripple; this phenomenon was also explored. The spectrum of a switching function in the presence of noise was obtained for voltage and current mode controls. By multiplying this spectrum by the transfer function of the output filter, the spectrum at the load is easily obtained. Apart from random Gaussian noise, there is commutation noise in a power converter. A brief analysis was done to explore this issue. This analysis showed that commutation noise is a critical issue, especially in current mode control. This confirms various observations in the literature.

48

CHAPTER 5 SENSORLESS CURRENT MODE CONTROL


It is desirable that the control schemes of power converters be robust to noise, have good dynamic performance and be simple to design. A control scheme has further merit if it requires minimal sensing of states and can be implemented using commercial PWM ICs. With these objectives in mind sensorless current mode (SCM) control was developed as part of this project. Current mode control is based on sensing of an inductor current with the objective of maintaining it at a desired level. This scheme has good dynamic response to disturbances at the input and output of the converter. It relies on current sensing, which can be prone to noise and can be hard to implement for low voltage power converters. It is possible to obtain information about the currents in the converter using the voltage information. For example, the current through an inductor is the integral of the voltage across it. In principle, it is possible to reconstruct the current entirely from the voltage. In practice, it is simple to construct the ac part of the current from the inductor voltage. More importantly, it is possible to formulate control laws in terms of the internal voltages in the converter. It is not necessary to explicitly reconstruct the current signal and then implement current mode control. Sensorless current mode control is a scheme to control the output voltage and current in a power converter using the intermediate voltages in the converter. In periodic steady-state operation, the average voltage across an inductor must be zero. In a buck converter, the SCM control law sets the voltage at the switch side of the inductor to a desired reference value. This indirectly sets the output voltage to the reference voltage. 49

In other converters, the SCM scheme sets the inductor voltage such that its average is zero when the output voltage is equal to the reference voltage. 5.1 Sensorless current mode control theory In a dc to dc converter, the output is connected to the input voltage by a network of switches, inductors and capacitors. Initially, we consider converters that have an inductor connected to an output capacitor for all or part of the switching period. The inductor is also connected to the input source for all or part of the switching period. The switch action partially determines the voltage across the inductor and it is possible to set the average inductor voltage over a cycle. Let us define switching function H0 for the converter. H0 is unity when the inductor is connected to the output voltage and zero otherwise. Define the voltage at the output side of the inductor to be V^ and the voltage at the other side of the inductor to be VL1. Define the voltage VA whose average is being set to zero. Also define VR the reference voltage desired at the output.

To set the average voltage VA to zero, it is necessary to integrate the voltage and to take switching action based on the integral. Define V, to be the integral of VA. Thus a hysteresis based control scheme would turn on the switch at a certain value of V, and turn it off at another. Figure 5.1 illustrates this control scheme for a generalized converter.

50

SWITCHING

"LI

J7RRTL

"L2

FUNCTION

VL2

Ho

L.

+X

INTEGRATOR SCHMITT TRIGGER

LI

J
Figure 5.1 Schematic of SCM with hysteresis When the output voltage is equal to the reference voltage, VA is the same as the inductor voltage. The voltage V, is the negative integral of the inductor voltage and is, therefore, an estimate of the current. The circuit.is effectively an estimator, although it reconstructs only the ac part of the current and is different from a true current observer. The scheme shares many operating characteristics with current mode control. To enforce a switching frequency, the control scheme is modified slightly. The switch position is determined by a set-reset latch. The latch is set by a synchronous clock pulse. It is reset when the value of V, is less than a ramp level. Introducing a ramp is a standard technique in current mode and voltage mode PWM control to ensure stable duty ratio. Figure 5.2 illustrates the modified control scheme.

51

SWITCHING FUNCTION
V

LI

JTRRTL

L2 CLOCK

L2

Ho
S 0

LI

/ -

COMPARATOR

LATCH

LI VR

Figure 5.2 Schematic of SCM with PWM The duty ratio at cycle N is defined to be DN and the cycle time is defined to be T. The magnitude of the ramp is equal to V%p. It is assumed to be zero at the beginning of the cycle and equal to VRP at the end of the cycle. The ramp slope SR = V%p/T. At the reset time, the signals into the comparator are equal. Thus V, is

y/(w+D,)n=%r

(5.2)

Define the derivative of V, to be SN when the switch is on and SF when the switch is off. A recursive equation can be obtained for the duty ratio by taking the difference of Equation (5.2) at cycle N from cycle N+l. V^N^D^m-Vji^D^Ti^l'D^S^^J^D^-D^T (5.3)

Figure 5.3 shows the waveforms of the voltages and the switching functions. A difference equation for the duty ratio can be obtained as

52

(5.4)
S

R~SN

^ - ^

The pole of this difference equation is at (SR-SF)/(SR -SN). The slope SN is always negative and Sp is always positive. The minimum ramp slope for a stable system is therefore (SN+SF)/2. A good choice is to select SR = SF. This sets the pole of the difference equation to zero. A pole at the origin of the Z-plane corresponds to the most stable system. This is analogous to the optimal ramp rate in current mode control.

RAMP 1

1 ! 1 Slope = S^ y ^

ANALOG SIGNAL

V,

\ v v ^ % o p e . ^

/Slope = ^ ON

OFF SWITCH POSITIOf 4

Figure 5.3 Waveforms in SCM For a stable system, the steady-state duty ratio is independent of the ramp slope. It is obtained by taking the limit N tending to infinity in Equation (5.4).

a,=

(5.5)

sFs
53

The minimum ramp slope for stability can be interpreted in terms of the steady state duty ratio using Equation (5.5). For steady-state duty ratio equal to half the minimum ramp slope for stability is zero. For smaller duty ratios the minimum ramp slope is negative. This means that the system can work without a ramp for duty ratio less than half. This result is analogous to the stability criterion of peak current mode control. The ramp stability criterion for SCM does not depend on values of the storage components of the power converter. This makes the design robust to variations in storage components. The slopes are determined by the input and the reference voltage. The latter is a constant and the variations in the former are known a priori. Thus designing a stable or optimal loop is simple. Any real inductor has some parasitic series resistance. Thus there is a steady-state error in the output voltage, which is equal to the resistive drop in the inductor. For a power converter with high efficiency, this drop is of the order of about 1% of the output voltage. For some applications, this is acceptable. In other cases, a Proportional-Integral (PI) loop can correct for this error. Figure 5.4 shows the schematic of the scheme with the outer PI loop. 5.2 Discontinuous mode In a conventional switched dc to dc power converter, passive diodes are used for many of the switch elements. It is possible to have diode turn-off prior to closing of the active switch because the diode current goes to zero. The behavior of a dc to dc converter in this discontinuous conduction mode is dramatically different from normal operation. Conventional design of dc to dc converters tends to avoid this mode of operation. The SCM control scheme, in contrast, handles the discontinuous mode of operation in a manner very similar to the

54

continuous operation. Since the inductor voltage is sensed and fed back into the control law, the control reacts immediately if the converter makes a transition between the continuous and discontinuous modes of operation.

LI

i"0
CLOCK

SWITCHING FUNCTION

"L2

Ho
S Q

LI

COMPARATOR RAMP

LATCH

"LiK%
V

Figure 5.4 Schematic of SCM with outer PI loop The stability analysis is similar except that now we have three circuit configurations: switch on, switch off and a third mode in which both the switch and the diode are off. In the last case the inductor is open circuited. This corresponds to zero current and a slope of V, equal to zero. Figure 5.5 shows the waveforms. Equation (5.2) is still valid. However, the ramp rates are affected and Equation (5.3) is modified since there is an additional interval of zero slope. We define a time weighted average slope S^ for the total off-period of the active switch. For all positive SF, the average S ro is less than SF. The original criterion for stability is SR > (SN + SF)/2 implies that SR > (SN + Sn))/2 is also true. Thus stability in discontinuous mode is ensured if the system is stable in continuous mode. 55

RAMP

ANALOG SIGNAL

OFF

SWITCH POSITION

Figure 5.5 Waveforms in SCM (discontinuous mode) 5.3 Flux Balance of Transformer Coupled Converter Magnetic flux in a transformer can be treated as a state variable. It is necessary to estimate this flux to avoid transformer saturation. Direct measurement of flux in the transformer of a power converter is rarely attempted. In conventional design, an estimator is used to estimate the flux. This estimate is used to control the excursion of the flux. The goal is to set the duty ratios such that the flux excursion is balanced and centered around zero. In current mode control, the transformer current is used to estimate the flux, which is proportional to the current. A dual of this scheme is to use voltage to estimate the flux. This eliminates the problem of current sensing. The flux is proportional to the integral of the voltage. The integration process provides a degree of high frequency noise suppression which makes it preferable to current mode control. Figure 5.6 shows the schematic for flux balancing in a transformer coupled converter. It is possible to balance flux and control the output voltage of the converter by modifying the flux boundaries. The following schematic shows such a control strategy. 56

CLOCK

21

I
FLUX ESTIMATE

COMPARATOR
S

PI TRANSFORM!
LATCH CLOCK

P2

INTEGRATOR

12

!
"22

1^
H|2

"r>

"
COMPARATOR

%
LATCH

Figure 5.6 Schematic for flux balancing Using this strategy the flux estimate O is maintained between 3>mm and Omax. The time base is obtained from two phase clock pulses applied to the set pins of the SR(set-reset) latches. The average duty ratio is set by the difference between 4>max and <Pmin and the frequency of the clock. The duty ratio controls the output at the secondary side of the transformer. For very precise output control, the outer PI loop can be set to monitor the secondary voltage. This PI loop would set the flux boundaries <J>max and 4>min. To ensure that there is no direct current path across the input voltage source during switch transition a small dead time of about 200 ns has been introduced between the complementary switch functions H,,,H2I and H12,H22. 57

5.4 Experimental verification 5.4.1 Buck converter A buck converter operating from 5 V to 2 V with load current from 0 to 15 A is considered here. It is built with an inner SCM loop and an outer PI loop. The SCM scheme is particularly simple for a buck converter because the output inductor is always connected to the load. When the control law sets the voltage on the input side of the inductor equal to the reference voltage the output becomes insensitive to the converter supply. Figure 5.7 is a schematic of the SCM control without the outer loop. A complete circuit diagram with the outer PI loop is given in Appendix A.

i
ViN

rTy;z

Vs

HJolTL

LOAD

BUCK CONVERTER CLOCK

VS

INTEGRATOR S Q

^U^l
RAMP

COMPARATOR LATCH

Figure 5.7 SCM schematic for buck converter Table 5.1 compares steady-state operation for the two cases: with and without the outer PI loop. The source regulation is less than 0.3% without the outer loop over a 20% input range. Load regulation and source regulation improve with addition of the outer loop. 58

With the outer loop, combined load and source regulation is less than 0.02% over the entire range. Table 5.1 Buck converter data

Input Voltage (V) 4.0 4.0 4.0 4.0 5.0 5.0 5.0 5.0 6.0 6.0 6.0 6.0

Output Voltage (V) (No Output Loop) 1.9958 1.9538 1.9133 1.8720 1.9958 1.9526 1.9121 1.8675 1.9958 1.9510 1.9107 1.8680

Output Voltage (V) (With Outer Loop) 1.9974 1.9973 1.9973 1.9972 1.9974 1.9973 1.9972 1.9973 1.9972 1.9971 1.9971 1.9971

Output Current (A) 0.04 5 10 15 0.04 5 10 15 0.04 5 10 15

Figures 5.8 and 5.9 show steady-state waveforms with the SCM scheme. Waveforms for the switch voltage, the output ripple, inductor current and its estimate V, are shown in Figure 5.8. The switch voltage and the inductor current estimate V, for continuous and discontinuous modes are shown in Figure 5.9. Note that the current estimate V, has a sign inversion from the inductor current because it is the negative integral of VA .

59

Switch Voltage (2 V/div) (upper), Output Voltage Ripple(20 mV/div)(lower) X axis: Time (2 us/div)

Inductor Current (1 A/div) (upper), V, (1 V/div)(lower) X axis: Time (2 us/div)

Figure 5.8 SCM control for buck converter: steady state behavior

(continuous mode)

(discontinuous mode)

X axis: Time (2 us/div) Y axis :V, (1 V/div) (upper), Switch Voltage (2 V/div) (lower) Figure 5.9 SCM control for buck converter: current estimate waveforms

60

5.4.2 Boost converter A boost converter for 70 V to 210 V step-up with a load current of 0 to 0.3 A is considered here. It is built with a single loop SCM. The converter worked reliably for an input voltage range of 60 V to 100 V. It can operate with no more than a 5% change in output voltage over its entire load range. Figure 5.10 provides a simplified schematic for the control scheme. Appendix A provides a complete circuit diagram of the converter.

foWL
v,
Hi

-W-

D2

LOAD

BOOST CONVERTER CLOCK V| H2 INTEGRATOR

_%/_
H2

-(I.

COMPARATOR RAMP

LATCH

Figure 5.10 SCM schematic for boost converter The source regulation was found to be better than 5%. Table 5.2 shows the variation in output voltage with input voltage and load impedance. Note the regulation is better than +/4% over the entire range. This performance is obtained without using an outer PI loop.

61

Table 5.2 Boost converter data


Input Voltage (V) 60 70 80 90 100 Output Voltage (Load=710 Q.) 217 213 210 211 212 Output Voltage (Load=5 k#) 219 217 215 214 212 Output Voltage (Load=200 kfi) 226 223 220 218 216

Figures 5.11 and 5.12 show characteristic steady-state waveforms for the boost converter both in continuous and discontinuous modes of operation. Figure 5.11 shows the switch voltage and the voltage integrator output V,. Figure 5.12 shows the current waveform and compares it to the voltage integrator output V,.

11.1

n
+ ,

rH

N-

'Vi

i7

/N

(continuous mode) (discontinuous mode) X axis:Time(5 us/div) Y axis:Inductor Current (0.5 A/div)(upper), Switch Voltage(50 V/div)(lower) Figure 5.11 SCM control for boost converter: steady state behavior

62

(continuous mode) Inductor current (1 A/div) (upper) V, (1 V/div)(lower) X axis: Time (5 us/div)

(discontinuous mode) Inductor current (0.5 A/div) (upper) V, (0.5 V/div)(lower) X axis: Time (5 us/div)

Figure 5.12 SCM control for boost converter: comparison of current and its estimate SCM control is very effective in handling load and source transients. For the boost converter, which is designed for an automotive application, it is necessary for the converter to maintain reliable output voltage during large signal transients. Figure 5.13 shows the response of the converter to load and source transients. During the load transient the converter load is changed from full load to no load (710 Q to 100 kQ.) and the converter responds with no more than 10% overshoot in the output voltage. Note that this also takes the converter from the continuous mode of operation to the discontinuous mode of operation. During the source transient the input voltage is changed from 55 V to 80 V and there is no more than 3% change in the output voltage.

63

'
!

!
i

! / 1
1

< 1

-1

1 |

.J..
I
1' 1
T T

r i j

i i i
J

1 i j

i i ' i
t J . . t ,..^i -r-ii

i^L

'
!

i
1

tTT
. i 7
T

,1 !
i
! i i

'"II 1
1

\
1

i
'

i
'

i
i
T

1I 1

M i ! ! ' 1

i!

! i

! '

! i

H
i

^ ^--^-^T-" ....
+

1 !

1 !

(load transient) 100% to no-load step Output voltage (20 V/div) (upper) Output current (0.1 A/div)(lower) X axis: Time (2 ms/div)

(source transient) Input voltage (10 V/div) (upper) Output voltage (5 V/div)(Iower) X axis: Time (50 ms/div)

Figure 5.13 SCM control for boost converter: Transient Response 5.4.3 Pushpull forward converter A simplified flux balance scheme for a pushpull converter is presented here. The application provides gate drive power for a 5 V to 2 V converter. To turn on the MOSFET switch VGS > 10 V is desired. A typical solution is to build a voltage tripler to provide a gatedrive supply above the 5 V rail. Instead a very simple pushpull converter was built that required very little hardware and a very small core. The circuit diagram is shown in Figure 5.14 . The power conversion efficiency obtained was 67%. The efficiency of a tripler is no more than 33%. This pushpull converter uses small BJTs for switches. The SCM scheme provides selfoscillation and does not need a separate oscillator. The frequency of this circuit varied with input voltage. Note that the transformer voltage is integrated and used as feedback. This 64

ensures flux balance and prevents saturation. The only analog component required for the control circuit is an opamp. The bandwidth requirement on the opamp is modest even though the switching frequency is 300 kHz. V
2N2222A 266CTI25-3E2Acore
V

PI
6 Turns

L&60J

(W5
g2
2N2907

36 Turns

2
luF

?i
V

IN582I

^
620pF 47k 10k
v

p2
pl

-vWr
-AW
I Ok I Ok l500pF

-vWV-

-Cx> H>o
\,
v

^
620pF

MC34071

Figure 5.14 Circuit for pushpull forward converter with SCM control 5.5 SCM Performance Comparison Here, SCM control is compared to current mode control and voltage mode control for a 5 V to 2 V buck converter. Dynamic performance is compared for the three control schemes. The circuits for each control schemes are shown in Appendix A. 5.5.1 Noise and dynamic range issues in SCM In current mode control, it is necessary to sense the inductor current. This is often done by a small sense resistor. To maintain efficiency, the sense voltage is small compared to the output voltage. To ensure that the loss in the sense resistor is at most 1%, the sense resistor 65

voltage has to be 1 % of the output voltage. By contrast, in SCM the feedback signal is a switch voltage which is of the order of the output voltage. Thus, the signals in the two cases are different by a factor of 100, or 40 dB. If the converter is running at 10% load, then the difference is a factor of 1000, or 60 dB. Thus SCM has a very significant advantage over current mode control in terms of signal magnitude. For low voltage converters, this is more significant since the current mode signal could be less than 50 mV at full load.

Transformer voltage (2 V/div) (upper) Transformer Current (0.1 A/div)(lower) X axis: Time (1 us/div)

V, (2 V/div) (upper) Transformer Current (0.1 A/div)(lower) X axis: Time (1 us/div)

Figure 5.15 SCM control for pushpull forward converter: steady state behavior

For current mode control with an external ramp, the current ramp rate is usually matched to the PWM ramp. The latter is fixed by design in most commercial PWM ICs. The gain of the current signal is fixed to create this match. For large load currents, the current signal gain has to be limited to maintain the signals within the I C s dynamic range. This problem is accentuated when the ratio of current ripple to its peak value is small, a condition

66

that is actually quite a common design choice since it allows for low loss in the magnetics. SCM on the other hand does not have these problems. The inductor voltages being fed back are almost independent of the load current. The outer loop only corrects for the dc drop in the magnetics and needs a very small dynamic range of operation. 5.5.2 Implementation without current sensing Current sensing usually increases losses and system complexity. The construction of precision noninductive current sense resistors can present serious problems, especially for low voltage power converters. Hall effect sensors and other nondissipative methods of sensing current offer a relatively expensive alternative. SCM bypasses these problems and matches the performance of current mode control using only voltage measurement. 5.5.3 Source regulation SCM has excellent steady-state source regulation as illustrated experimentally. Tables 5.1 and 5.2 show the steady-state performance of test buck and boost converters in the absence of outer loop action for source and load regulation tests. Good regulation of transients in the source voltage is expected for SCM because the input voltage is used in the feedback (Equation (5.2)). The input voltage is essentially being monitored on a cycle by cycle basis. Since there is only one control input per cycle (the duty ratio), this is the fastest control action feasible. Figures 5.16, 5.17 and 5.18 compare performance of SCM control, voltage mode control and current mode control during a source transient. Both continuous mode and discontinuous mode of operation are considered. SCM control and current mode control show excellent response, while voltage mode control does not respond effectively.

67

^ j y .*.i'.?-rr "'i1 *' if^wyW!* i.


X axis: Time (50 us/div) Y axis :Input VoItage(l V/div) (upper), Output Voltage(20 mV/div) (lower) Figure 5.16 SCM control for buck converter: source transient

X axis: Time (50 us/div) Y axis:Input Voltage(l V/div)(upper), Output Voltage(20 mV/div)(lower) Figure 5.17 Voltage mode control for buck converter: source transient

68

""ft s**1

( ^

j*M
>

StjiL. " T jll


1

1
A_*.

1 T<

*>*ts

A. 1= ^V/L/kAJ T P"

#u

FIFTH* lp T 'F rl l PffW H ' * r r i H l*f "\

#Uj

jvjjj,

JJ*U* , t ^ T ' t * WfMw ' t ' . - i f M

*4..*-*l

i
X axis: Time (50 us/div)

1-

Y axis : Input Voltage( 1 V/div)(upper), Output Voltage(20 mV/div)(lower) Figure 5.18 Current mode control for buck converter: source transient 5.5.4 Load Regulation Steady-state load regulation is good for SCM even without the action of the outer loop. In contrast, current mode control requires the outer loop to set the current level and cannot handle load variation without the outer loop. Tables 5.1 and 5.2 show the steady-state load regulation data for a buck converter and a boost converter using SCM control. Dynamic load regulation is improved by the action of the outer control loop. The outer loop for the SCM control and current mode control are the same. Figures 5.19, 5.20 and 5.21 show response to a 5 A load transient. In continuous mode, the performances of three schemes are comparable. However, the results in discontinuous mode are quite different. 5.5.5 Discontinuous mode of operation In discontinuous mode, the action of the converter changes. For example, a voltage mode control that performs well in continuous mode performs very poorly during a load 69

transient that takes the converter to the discontinuous mode of operation. In contrast, the SCM control scheme can handle the discontinuous mode of operation very well. Figures 5.18 and 5.19 highlight the difference. The regime of discontinuous mode of operation is usually avoided due to problems of conventional control. However, SCM facilitates the design of converters that allow discontinuous mode of operation. There are many advantages of operating a converter in the discontinuous mode for at least a part of the load range. These include operating the converter at a lower switching frequency, smaller inductor values and elimination of internal loads to force continuous mode. This results in an overall gain in power conversion efficiency.

(continuous mode)

(discontinuous mode)

X axis:Time(50 us/div) Y axis:Output Voltage (50 mV/div) (upper), Output Current(2 A/div)(lower) Figure 5.19 SCM control for buck converter: load transient

70

(continuous mode) (discontinuous mode) Xaxis: Time (100 us/div) Y axis:Output Voltage (50 mV/div) (upper), Output Current(2 A/div)(lower) Figure 5.20 Voltage mode control for buck converter: load transient

"

vVvv

^r
j "

>w\

W A A ; /^AAW * V * ^

A\

"
J W W

" ~ w \u

j*^

V iv
v

\
VAA/
VWVA,

y
=.

^ ! . /

j
1

(continuous mode) (discontinuous mode) X axis:Time(50 us/div) Y axis:Output Voltage (50 mV/div) (upper;, Output Current(2 A/div)(lower) Figure 5.21 Current mode control for buck converter: load transient

71

Note that the dynamic response to a load transient is comparable for the three control schemes. The dynamic response to a source transient is superior for SCM and conventional current mode control. SCM is a very attractive option for the control of dc to dc converters. It offers the performance merits of current mode control without the need for current sensing. The complexity level is almost the same as in voltage mode control except for an additional opamp. There is a dramatic performance improvement in source regulation and load regulation during discontinuous mode compared to conventional converter control methods. Eliminating the need for current sensing also saves the power loss associated with current sensing which may be as high as 2.5% for a 2 V converter. The difference in signal level between SCM and current mode controls is large. For example, in current mode control the current sense resistor may have a drop of 50 mV compared to more than 1 V across the inductor. A difference in signal level of 40 dB to 60 dB is not unusual. Numerous low voltage converters have been built by the author (48 V to 2 V; 12 V to 2 V and 5 V to 2 V) with output currents up to 60 A. In such circuits there is a problem of noise from the power stage coupling with the control signals. This increase in signal level was critical for reliable operation of these converters.

72

CHAPTER 6 OPTIMAL CONTROL APPROACHES


Minimization of a specified cost function is a standard approach to optimal control. Control of this type unifies large-signal and small-signal performances. In this thesis, quadratic output-error cost functions are considered for power converters. The conventional form of quadratic function is intractable for many converters. Several cost functions which

approximate it, but are easier to apply, are introduced. Examples include a cycle-by-cycle quadratic function, a cost function based on the stored energy in a converter, and fast-switching approximations to quadratic functions. These functions are studied for dynamic performance and stability. Results are shown to provide good performance, with little

distinction between large and small disturbances. 6.1 Introduction Any given power converter has a specific ultimate operating requirement. For example, a power supply is intended to produce an output voltage which tracks some reference signal subject to current and thermal limits. The supply must show good regulation, low output impedance, and fast recovery from load changes or other large transients. In conventional power electronics design, problems of small-signal performance and large-signal control are attacked separately. Since the only power control mechanism in any converter is through operation of switches, it is natural to examine possible unified approaches to control design. One might ask whether there is some "best way" to control the switches to meet the converter requirements. When the problem is approached from this perspective, there is no distinction between a large signal or small signal disturbance.
73

The theory of deterministic optimal control [39] is often applied to linear systems. Performance requirements are used to define a cost function, which is to be minimized. The method works for some nonlinear systems. However, power electronic systems are difficult to handle with many nonlinear methods because of the discontinuous control. This thesis explores control methods for power converters based on optimization of cost functions. Examples for dc-dc converters are discussed. It is shown that the approach achieves good performance for both large signal and small signal dynamics. Conventional controllers based on small-signal behavior at best can approach optimal performance only in the small signal limit. 6.1.1 Initial definitions An optimal control problem can be formulated when there is a set of dynamic equations, a control input which might be bounded, and a cost function to be minimized [39]. Power converters are well-characterized as linear circuits combined with ideal switches [40,41]. Switching functions h(t) allow the m circuit configurations to be described as a compact differential equation

x =][A,(f)Kr + B,]

(6.1)

All components of x are capacitor voltages or inductor currents. The desired operation of a converter corresponds to a point or consistent closed curve in the state space defined by the variables x. Consider an application such as a dc-dc converter or a PWM inverter. The only control is the position of the switches. If the switching frequency is very fast, the switch duty ratios

74

can be considered as control inputs, bounded between zero and one. An appropriate cost function might penalize error between a (possibly time-varying) reference value and the desired output voltage. An example of such a function is the simple quadratic integral 7, of the form

There should be a unique set of time-dependent control inputs that drives the converter from a given initial state to the final operating point so as to minimize /. Of course, there are many reasonable choices for a cost function other than the quadratic error integral. An exponent higher than two might be desired to strongly penalize output error. Or a less symmetrical function might penalize overshoot while allowing wider tolerance for undershoot. In [42], series-resonant converter control based directly on stateplane trajectories is explored as a geometric approach to optimal response. In [43], PWM control waveforms optimized for harmonic interference properties are discussed. It should be noted that J as written in (6.2) is not useful for power electronic control. The ripple allowed in a real circuit makes the error integral unbounded when Vref is the nominal output. For the purposes of this discussion, J is based on the true desired output signal, including ripple, rather than on an idealized control reference value. Here, we study converters on a per-unit basis, so that all states and inputs are unity at the desired operating point. These per-unit state variables will be given symbols y to distinguish them from the true operating variables x. The system is always organized so that the first state variable is the output voltage. Thus J is given by Equation (6.3).

75

J) = f~(y\(s) - \)2ds

(6.3)

6.1.2 Basic converters As a basis for exploring the cost function minimization approach to power converter optimization, consider three simple dc-dc converters: the buck, boost, and buck-boost circuits. The parameters and state variables are defined in Figure 6.1.

Buck

-D*
Boost
'In

+
%

Buckboost

Figure 6.1 Basic converters Continuous conduction mode will be used to start the analysis, although the controllers are robust enough to handle the discontinuous case as well. The three basic converters can be 76

written as a combination of on-state and off-state circuit equations. The on-state and off-state dynamics, respectively, are y
>on

= A _y + B
on? on an

(6.4)

The state matrices for these converters are Buck:


1 1

off

X 1 X

T~c
0

0 , Bon = '

k , 2off =
L_ ~r~,

(6.5)

Boost:
-

-0
\ n

Tc

Tc
0

s =

kT.

(6.6)

Bon = Boff =

k2TL

77

Buck-boost:
p -

-0
\ n

1+fc

Tc 0 0

Tc
# =

1 (6+1)7",

(6.7)

B =

* * '

%6+ur,

where TL = L/R and 7*c = RC are the inductor and capacitor time constants, respectively. The parameter k is the ratio of dc output to dc input for each converter. 6.2 Approximate Optimal Control Methods The definition of optimal control in terms of the cost function and system description is straightforward. However, implementation involves formidable calculations. In effect, the complete future behavior of y,(t) is needed to determine the control action at any given time. In a deterministic linear system, this does not create special problems. In a power electronic converter, we are faced with the problem that the best switching time depends on all subsequent switching times. In this section, various alternatives are explored that seek simpler controls while offering an approximate minimization of J. 6.2.1 Cycle by cycle optimization The integral cost function (6.3) cannot be written in closed form, in general, for nonlinear systems. An alternative is to break it into a "first cycle" portion, and ignore the remaining "future" portion. This first portion, J,, is
78

/,(') = f ^ w - ^V'^

(*-*)

where T is the converter operating period and t the time at which the cycle starts. This function provides an indication of error accumulated through a single operating cycle. Its minimization represents an approach to cycle-by-cycle control. This control philosophy has been addressed in other ways [44,45] for specific converters. The cost function (6.3) is sometimes called the infinite horizon cost function. Minimization of (6.8) represents the solution of the finite horizon problem [39]. Let us consider how a controller for Equation (6.8) could be implemented. The value of J, at any given time depends only on the system state at the cycle start time t and on the switching time t + DT. Minimization requires

fi = 0
dD

(6.9)

A system with one state variable can be reduced considerably. The scheme has a straightforward interpretation: The switching cycle begins with the switch off. The cost function is monitored from the beginning of the cycle with an analog integrator. The integral over the remainder of the cycle time is computed by assuming switch turn-on at time DT. When the rate of change of the total crosses zero, the switch is turned on. This integral comparison is similar to the method of [46]. Figure 6.2 shows the output voltage transient during start-up for a buck converter which uses this integral-comparison cycle-by-cycle control. The capacitor has been set small

79

enough to allow easy viewing of theripplebehavior. Steady state operation is reached rapidly and smoothly.
i i i 1 r

X axis: Time (20 us/div) Y axis: Normalized output voltage

,. I

Figure 6.2 Buck converter startup under minimization of J, It is not easy to generalize the integral comparison method to systems with many state variables. The method also requires complete knowledge of the system so that the integrals can be computed. A simplification allows cycle-by-cycle control to function for a broader range of converters. Consider a converter operated so that The off-time is fixed. The active switch turns off at the beginning of each cycle. Error accumulates in approximately a fixed amount during the off time, since its length is fixed. To minimize, therefore, we only have to look at the switch on time interval between t and f+DT. Minimization gives simply

0 = (^-V

(6.10)

80

Keep in mind that Vref varies with time and includes the ripple value. The switch should be turned off when the output voltage crosses the desired output waveform. Then the cycle repeats. This process is a variation of deadbeat control similar to a technique used in adaptive control systems [46]. It is known that such a scheme does not work properly for systems with nonminimum phase (i.e., an unstable zero), such as boost or buck-boost converters. Figure 6.3 shows a buck-boost converter operating to minimize J, with the constant off-time approach. The output looks reasonable, but the inductor current increases without bound.
' / '

X axis: Time(20 us/div) Y axis: (0.5 /div) Normalized voltage (Yl)

/ ./

1
-V

J u

/ /

r
/ '

Normalized current (Y2)

^fL_

Figure 6.3 Buckboost converter instability under J, This example shows that cycle-by-cycle control has significant drawbacks. Although it might be expected that minimization of error during each cycle will minimize the total error (6.3), this clearly is not the case in general. J, minimization will work for only a very limited class of converters. The cost functions (6.3) and (6.8) are positive definite functions of the converter output. Therefore, they are valid Lyapunov functions [47] for the associated converter. If a 81

controller forces such a function to decrease over time, then the system will be stable. This important property means that the cost function properties can be used to evaluate stability of the closed-loop system. The controller for the buck-boost converter simulated in Figure 6.3 does not force the error to decrease near the operating point. It does not meet the sufficient condition, and indeed is not stable. 6.3 Lyapunov Function Control The function J at any given point in the state space is determined by an output storage element with a voltage or current different from the desired operating value. In the cycle-bycycle case, only the output state was considered. It seems likely that a better control would include more state information. Furthermore, the idea that J is a Lyapunov function suggests that a more conventional Lyapunov function the total system stored energy might be of use in control. The idea of Lyapunov control for power electronics was discussed by Sanders and Verghese [18]. Here, we consider a weighted excess energy function as a possible basis for converter control. When two state variables are present, this function, JE, is written

where Vref and Iref are nominal operating values, and w, and w2 are arbitrary weighting parameters. JE is a valid Lyapunov function for the converter. It can be extended to any number of state variables. A controller design based on this function should have weights selected to obtain the desired performance. If the controller drives JE ever smaller, the converter will be stable. Not all values of w will bring about this result.

82

6.3.1 Implementation To simplify the control, we consider a scheme which minimizes JE at the end of each operating cycle. An observer is used to permit computation of the energy at the end of any given cycle. As mme cycle-by-cycle case, me switch is turned on at atimewmchm^^^^ the excess energy. This control approach is interesting because of its simplicity and performance. The normalized output voltage of a boost converter during a start-up transient is given in Figure 6.4 for this control and for two control schemes developed in Sections 6.4 and 6.5.

X axis: Time (50 us/div) Y axis: Normalized output voltage (0.1 /div)

Figure 6.4 Boost converter start-up transient under minimization of J E , J g , J ~ The output recovery of a buck-boost converter from a 50% load decrease is shown in Figure 6.5 for the same three control schemes, and also for a more conventional linear state feedback controller. All weights are set to unity. In both cases, the transition is fast and
83

smooth. In the case of the load transient, the controller acts to minimize the overshoot energy, rather than the peak value. In many converters, overshoot energy must be absorbed in the semiconductor switches. Minimization of this energy therefore is likely to make a converter more reliable than a conventional alternative.

X axis: Time(50 us/div) Y axis: Normalized output voltage (0.1 /div)

Figure 6.5 Buckboost converter 50% load transient under minimization of JE, Jg, J 6.3.2 High frequency limit A high-frequency approximation is one alternative to the cycle-by-cycle method for JE. It also provides an insightful geometric interpretation of the energy method. Consider the state space velocity vector V = yon - yoff, the difference between on-state and off-state system

84

velocities. We should choose this velocity so that its average over time carries the converter output closer to the operating point. In the state space, contours of constant JE can be constructed. These contours form a set of concentric ellipses centered on the operating point. The smaller the ellipse, the lower the excess energy associated with it. To minimize JE, we choose the switch configuration that moves most quickly toward the smallest ellipse. In the case of two switching devices, this is simply a matter of finding the angle between V and the gradient of JE.

f = 94'0L -)V

(6.12)

The active switch is on when c < 0 and off when c > 0. A state-plane example is given in Figure 6.6.

X axis: Y 1(0.2 /div) Y axis:Y2(0.2 /div)

Figure 6.6 Boost converter state plane contour at JE=0.4 and difference velocity V 85

6.3.3 Stability of energy function control Stability under energy control is assured if the function JF decreases with time. The time derivative is given by the following equation.

4r = w ^ - "-4 * w ^' '-4

<613>

To evaluate stability, the contours where dJ/dt = 0 are computed or plotted as a function of the weights for each configuration of a converter circuit. A region in which dJ/dt < 0 for all configurations will produce stable operation for any control. If there is at least one configuration for which dJ/dt < 0, stability can be guaranteed with the proper choice of switch action. An example is shown in Figure 6.7. This plot represents behavior of a buck converter with w, = w2 = 1. The upper curve is the zero contour for dJ^dt when the active switch is on. The region below the curve has dJ/dt < 0 with the switch on. The lower curve is for the off state. The region above the lower curve shows dj/dt < 0 if the switch is off. Thus, in the subspace between the two curves, JE is always decreasing, and operation is stable regardless of switch behavior. For all other points in the state space, dJ^dt can be made negative through the proper choice of switch setting. For converters other than the buck, some combinations of weights will give dJ^dt > 0 in any switch configuration. Such a converter might become unstable in this "rising error" region. The three dc-dc converter types under study here can be compared on the basis that Wj = 1. Stability regions for these converters become Buck converter: stable for all w2 > 1.

86

Boost converter: stable for vv, > 1, except unstable if w, is higher than the converter dc input-output gain. Buck-boost: stable for w2 > 1. It is important to recognize that these are sufficient conditions for stability. Weights outside these limits might still lead to a stable controller.

X axis: Yl(0.2/div) Y axis: Y2(0.2/div)

Figure 6.7 Stability regions for buck converter 6.4 Gradient Approximation to Optimal Control Although energy-based control is straightforward and easy to analyze for stability, it is not optimal in the sense of minimizing (6.3). Let us now attempt a more complete approximation of the cost function J. The next method to be described focusses on the gradient of the cost function and yields an approximation to J which we will call Jg.

87

The integral function in Equation (6.3) is continuous, and has continuous first derivatives. In the limit of fast switching, the velocity dy/dt can be written in terms of the on and off states and the duty ratio as I y, = y ^ + D(y,_, - y ^ ) r-*o (6.14)

This allows the time derivative of Equation (6.3) to be rewritten with the chain rule as

- = -(f, - 1): = ^ [ y ^ + D(y_, - y ^ ) ]

(6.15)

Since the only control is through switching, we should choose the configuration which minimizes dJ/dt at any moment. The derivative dJ/dD can be set to zero to accomplish this. The derivative is

dD yd'j The function g(y) = 0 is the surface in state space where the cost function is minimized. The polarity of g(y) is the key control parameter in this instance. For example, there is a region of state space in which g(y) is negative when the active switch is on. If the switch is kept on in this region, / will decrease. Similarly, in a region where g(y) > 0 with the switch on, the objective of optimality is best served by shutting off the switch. Along the surface g(y) = 0, the duty ratio should take on some intermediate value such that the converter states remain on the optimal surface. The function g(y) thus represents a geometric control element. It would be well-suited for sliding mode or direct boundary control techniques [11].

88

The only difficulty is that g(y) in Equation (6.16) is a function of the unknown J. There is no second equation here to permit a solution. An approximate solution for (6.16) is needed to generate a useful control. Here, we consider that the change of J along a particular variable should depend on the time rate of change of that variable. There might be a constant of proportionality, but the rates should still be related. If this holds for ail the states, g(y) can be rewritten as

g(y)=Z%.,,-W
1*1

(*-'7)

Results in the two-state case are easy to plot, although the concept extends to an arbitrary number of states. A control based on this g(y) will not minimize / , but will minimize some "gradient cost function" JK. For the three basic dc-dc converters, the g(y) - 0 contours obtained for minimization of Js are Buck: g(y) = y, - 1 Boost: g(y) = y2 - (y,)\ where z = TL/TC Buck-boost: g(y) = y2 - [(1 + ky,)/(l+k)]z, where k is the dc voltage step-up ratio. The simple result in the buck converter case will keep g(y) = 0 when the inductor current is kept at its nominal value. A hysteresis control based on inductor current will be optimal for the buck converter in the sense of Jg. The family of contours g(y) for the boost converter is given in Figure 6.8. A practical boost converter often has T c > TL, so z < 1 is typical.

89

X axis: YI(0.2/div) Y axis: Y2(0.2/div) Z=TL/TC

Figure 6.8 Plot of g(y)=0 for boost converter This gradient approximation approach can be analyzed for stability from the geometric conditions given in [48]. In contrast to the previous cases, this method is not based on a Lyapunov function. The performance of this approach for a boost converter with TL = T c was illustrated in Figures 6.4 and 6.5. In Figure 6.4, a boost converter start-up trace was shown. The Jg trace had a slightly higher cost function than JE, so it is not as optimal as the energy control method. For the 50% buck-boost load transient in Figure 6.5, the methods gave almost identical results. A view of system state space trajectories for a boost converter under large-signal transients is given in Figure 6.9. This shows the state behavior of the boost converter from various initial conditions. The dark line is g(y) = 0. Below the line, the active switch is off. Above the line, the switch is turned on. The controller keeps the states on the line once it is reached.

90

X axis: Yl(0.5/div) Y axis: Y2(0.5/div)

Figure 6.9 Boost converter large signal trajectories in state space 6.5 Fast Switching Optimal Control If the duty ratio D is truly a continuous variable, then second derivatives of J are defined, and additional equations can be developed to solve (6.16) for g(y). Continuity of D will apply in the fast switching limit. Let us reexamine (6.16) with this assumption in mind. Again, the function we compute will not be J, but will be a suboptimal approximation J_. The surface g(y) = 0 in principle will give an excellent approximation to true optimal control. 6.5.1 Derivation of optimal boundary Once the system trajectories reach the surface g(y) = 0, we want to keep them there. The control resembles a sliding mode in g(y). Because of this behavior, we will assume that the controller successfully forces the states to remain on g(y) = 0 once they reach that surface. The assumptions of a control along g(y) = 0 means that the total derivative dg/dy = 0. Once again, the chain rule can be used to write out the derivative as

91

o=E-^Lv^ -D(y_,-W]
1=1 "y,

(6.18)

To simplify the analysis, two operators will be defined:

4 ( ) = Z [ y ^ + ^(y,,, - & ^ ) ] f - ( )
(6.19)

4() = Z %^^-()
i-1

y,

First, L, is applied to (18) and L, is applied to (16). The results are subtracted to give a partial differential equation in /,

ay,Kon)
yjiaffl-

-i

dy, y-i

ay.

yJinn)

a^ >J7>
ay,

= 2(y,-l)(y',^,-y.m)

(6.20)

For the two-state case, these expressions can be solved to obtain the g(y) = 0 contour. Results for the three simple converters are Buck: g(y) = y, - / Boost: g(y) = y2(y2 - y,) + ay,(y, - J), where a = Tc/(2TLk2) and k is the dc voltage gain Buck-boost: g(y) = 2(k+l)y2(y2-y,) + (a+akyj+y2)(yrl),

where a = T(V[(l+k)TJ and k is the voltage gain. In this case, the surface g(y) = 0 will minimize the sub-optimal approximation 7_. Figures 6.10 and 6.11 show the contour g(y) = 0 for various parameters a for the boost and buck-boost converters, respectively. The interesting new feature of these contours compared to those of the gradient method is that they are closed. In geometric terms, this permits

92

unambiguous evaluation of stability. A converter operating with one of these contours will be stable as long as there is no stable equilibrium where the system could "lock."
' ' ' 1 1 1 1 1 1 j

Xaxis: Yl(0.2/div) Xaxis: Y2(0.2/div)

Figure 6.10 Boost converter g(y)=0 contours


i I 1 1 1 1 1 r

Xaxis: Yl(0.2/div) Yaxis: Y2(0.2/div)


BBO

_l

'

'

Figure 6.11 Buckboost converter g(y)=0 contours 93

6.5.2 Performance Large-signal transients imposed on boost and buck-boost converters under 7_ control were shown in Figures 6.4 and 6.5. The 7_ version gave the best results of these schemes in minimizing the actual error accumulation between the output and the reference. Several additional transients are simulated in Figures 6.12-14 for the /_ controller. In Figure 6.12, the start-up transient of a buck converter is compared with that for a more conventional controller that generates the D signal with linear state feedback. The fast-switching g(y) contour gives excellent performance, with no overshoot and a fast transition.

Xaxis: Time (10 us/div) Yaxis: Normalized output voltage

Figure 6.12: Buck converter start-up In Figure 6.13, a 5% load increase is imposed on the buck-boost converter of Figure 6.5. The controller is significantly faster than a conventional linear feedback controller and shows about the same time behavior as for the large-signal transient of Figure 6.5. However. the abrupt undershoot of the new controller might not be desirable. This is an indication that the quadratic cost function is not the best choice for a switching converter.

94

Xaxis: Time (5us/div) Yaxis: Normalized Output(0.02/div)

linear

Figure 6.13: Buckboost converter small signal load change All the controllers discussed here were designed based on continuous conduction mode. However, each scheme should be able to handle both large and small transients without difficulty.

Xaxis: Time (lOOus/div) Yaxis: Yl, Y2 (0.2/div)

Figure 6.14: Buckboost converter performance under J minimization 95

Figure 6.14 shows simulation results for a light load on the buck-boost converter under 7_ control. A load transient severe enough to drive the converter into discontinuous mode was imposed at t = 400 us. Within the limitations of the output filter capacitance, the controller permits virtually no output change. The converter recovers almost immediately, and maintains the desired output level. At this point we can compare the results with true optimal control, as based on the mimmum principal of Pon try agin [1]. For a converter of n state variables this involves the solution of a system of 2n nonlinear differential equations, n of which have initial values. The other n have final values. Even for n = 2 it is not possible, in general, to obtain an analytical solution. The fast-switching optimal controller derived here carries restrictions on the control performance and the switching speed. In a real system, it will lead to higher J compared to a true optimal control. However the system has clear stability properties and is relatively easy to implement. The simulation results suggest that it is quite useful as an approximation to the true optimal controller. 6.6 Experimental Results The implementation of optimal control is attempted for a practical low voltage converter that converts from 5 V to 2 V. This is a high performance converter that has aggressive load and source transient requirements. The converter is subjected to large signal input voltage and output current changes, and the results are compared to those for conventional control and new control techniques developed in this thesis. In particular, the comparison is between voltage mode control, current mode control and SCM (sensorless current mode) control. The test conditions are identical for

96

comparison. Figure 6.15 shows a source transient which is compared to Figures 5.16, 5.17 and 5.18.

iJjL i, j J pW* yVr "Wjr

,1 . k / " ^ |

yrn^'

1 i/r* fjr'i 1
i

. l._. _ A ^

k Mr 1
1 1 1 1 1JWA

1 1

'tAAAAA/ MAAAA UAAAAA /WWWM\W ff#\ \IMskfMiLAAAAA/ WNi fvVWV rYWvYY


V

WTV

T T T Y+*

Xaxis: Time (50 us/div) Yaxis :Input Voltage(l V/div) (upper), Output Voltage(20 mV/div) (lower) Figure 6.15 Optimal Control: source transient Figure 6.16 shows a load transient and the output response is measured. The results are compared to those in Figures 5.19, 5.20 and 5.21. The control law being implemented here is a boundary control. The boundary is simply y2-l=0. The implementation is quite complex. The variable y2 is the normalized inductor current. The reference current is a function of the load current and the load voltage. The load current and the inductor current have to be measured, and some multiplication and divisions have to be performed to obtain y2.

97

#w w\

rtVrVV \0t

AiiiM AMA/V AAW/^ ryir'ri ttT Tmrr

M*k &WdU,
vw

UUUUJUJUllli

..MIAAA/WAAAAAMIMA/MAAAA/

L,x/*A#

N
\ / ' k

(continuous mode) (discontinuous mode) Xaxis:Time(50 us/div) Yaxis:Output Voltage (50 mV/div) (upper), Output Current(2 A/div)(lower) Figure 6.16 Optimal Control: load transient 6.7 Conclusions A cost function approach to optimal power converter control has been presented. Conventional quadratic cost function optimizations cannot be solved in general for power electronic systems because of their nonlinearity. Useful approximations to the optimal control problem can be solved instead. Of these approximations, a Lyapunov approach based on energy functions and a near-optimal approach valid for fast switching appear to be the most promising. All of the methods consider a unified converter control problem. No special distinction is made between small-signal and large-signal controls. The performance in the small-signal case is comparable to that obtained with conventional techniques. The near-optimal approach appears to be robust enough to cope with extreme changes such as the transition between 98

continuous and discontinuous conduction modes with minimal sacrifice in performance. Final implementation results can be interpreted in terms of geometric variable-structure control methods. Although the examples considered dc-dc converters, the method has no inherent limitations on converter type or complexity.

99

CHAPTER 7 AUTO TUNING SCHEMES


The steady state optimization problem is of considerable interest in power electronics. This problem is non-trivial when the quantity being optimized is a nonmonotonic function of a controllable variable. The control problem is primarily to determine the optimal value for the controllable variable. The optimal value is a function of external factors such as temperature and load current, many of which are not readily measurable. Further, the functional relationship is not known precisely. A general tuning scheme would work by monitoring the relationship between the controllable variable and the output being optimized. In this chapter, a general tuning theory is developed that optimizes a nonmonotonic function. Power electronics has numerous applications for such optimization. This tuning theory has been used to solve the maximum power tracking problem, in which a nonlinear source, such as a solar array, is connected to a load, such as a storage battery. Power transfer between the two sources is controlled using a dc to dc converter. The power out of the source is maximized at a particular output current. The tuning scheme sets this input current to maximize the power out of the source. A feedforward ripple cancellation scheme presented in [4] performs ripple cancellation in switched power converters with the general tuning theory. It estimates the ripple current and injects a ripple cancellation current. The ripple current is estimated using a simple feedforward technique. However, the amplitude of the ripple is affected by the filter inductance, which varies with temperature and load current. The tuning scheme adjusts the gain of the cancellation current to minimize the output ripple. 100

7.1 Tuning Theory Correlation is used for tuning in many control problems. This usually involves the injection of an excitation input; then, the correlation between the excitation and the response is used to estimate system behavior. This is usually implemented digitally and results in a fairly complex circuit. This is not desirable in the context of a switching power converter. Here a simple analog tuning scheme is developed. It has very low parts count and is suitable for integration in an IC. Switching perturbs all the states in a converter. This serves as an excitation input for the tuning process. The tuning process in principle could update at the switching frequency. In the power electronics literature, there are a few tuning circuits that perturb the duty ratio at a low frequency [21]. Such schemes are slower than the switching frequency by more than one order of magnitude. The theory developed here is applicable for tuning a parameter in a switched power converter subject to some restrictions. 7.1.1 Definitions Define a cost function J being optimized and a controllable variable x. The controllable variable x can be considered as an input to the optimization process since it determines the output J. The objective is to maximize the output J by setting the input x to the optimal value.

From maxima-minima theory we know that at the maxima the derivative goes to zero. \ =0 dxx' (7.2)

101

where x0 is the optimal value of x that results in the maximum value of the output JQ. We define a correlation between the controllable variable x and the cost function J.

C,*=r^d,
JX

(7.3)

J dt dt

The variable x is assumed to be a controllable state. This implies that there exists some control circuit that can set the value of x. Define xR to be the reference value for x. Due to the switch action, x is not constant even in steady state but has ripple at the switching frequency. The average value for x is equal to xR in steady state. We define a general control law for this tuning problem, whereby k is a positive constant that sets the rate of convergence. ^=*C (7.4)

7.1.2 Mathematical analysis The control law can be rewritten as dJ_

fl=^()^=^(^)^
dt dx dt dx dt

(7.5)

~dl At the optimal value of x the quantity dJ/dx = 0. Thus the desired operating point is an equilibrium point. We impose bounds on xR to exclude any other points where dJ/dx is zero. These would correspond to any local minima or maxima of J.

102

Note that (dx/dt)2 is a nonnegative quantity. Owing to the switching action, the state x is n ever a constant. Thus (dx/dt)2 is positive definite. Thus, we have established the

optimal solution as the only equilibrium point. We further need to establish that the equilibrium point is stable. Consider a case in which there is a single maxima of J and no local minima of J as a function of x. J Figure 7.1 is a plot of J as a function of x. Note that the point (x0,J0) is the desired optimal point.

( % -U
/I

(J)
rv\ (X) -. >

Figure 7.1 Nonmonotonic optimization problem We consider two initial conditions (a) x < x0 and (b) x > x0. Note that in region (a) dJ/dx > 0. Thus dxR/dt > 0 by Equation (7.4). This causes x to increase until x=x0. At this point dJ/dx goes to zero and the system converges. Thus, if the system starts in region (a), it moves to the equilibrium point. If the system starts in region (b), dJ/dx < 0. This sets dxR/dt < 0. Consequently, x decreases till the system reaches the equilibrium point. This analysis assumes that k is small but positive.

103

Due to switching, x varies with a ripple at the switching frequency. The change in x over a cycle may be a significant fraction of its average value. It is necessary to check for convergence in steady state. Ideally the average of dJ/dx should go to zero over a cycle. Consider the system in steady state. The change in Xr over a cycle is zero.
T

((x)2dt=0
<*

(76)

Define the minimum value of x over the cycle to be XI and the maximum to be X2. Assume that the rate of change of x is determined only by the switch position. The value of dx/dt with the switch is defined to be Xn and Xf when the switch is off. Xn > 0 and Xf < 0. The integral in Equation (7.6) is transformed from the time domain to the x-domain. The resulting integral can be written as follows.
X2 XI

&dk+rfx/k=0
I dx "
Ldx ^

(77)

This can be approximated by assuming that Xn and Xf are nearly constant and can be taken out of the integral. This simplifies Equation (7.7) to

(X-Xf) ldx=0

(7-8)

Note that (X-Xf) is always nonzero. Thus the integral of dJ/dx over a cycle must go to zero.

104

7.1.3 Theoretical limitations This analysis assumes that J is a continuous and differentiable function of x. Thus this scheme would only work for systems that have this property. This analysis also assumes that there are no other points where dJ/dx = 0. If there are such points, the value of x is constrained to eliminate the region. For exact convergence, we assumed that the derivative of x is determined exclusively by switch position. This may not be exactly true. When this is not the case, the system will not converge to the exact optimal solution. In such situations, the convergence can be improved by increasing the switching frequency. This decreases the ripple in X over a cycle. This would make X and Xf more nearly constant. 7.2 Maximum Power Point Tracker (MPPT) 7.2.1 Definition of the problem A typical example of tuning is the maximum power point tracking problem from a solar array. The power output of a solar array is maximum at a particular current. The power output is a continuous, differentiable, nonmonotonic function of the current. Figure 7.2 shows typical V-I characteristics of a solar array under various illumination levels. The power is defined to be J and the current is x for this problem. The entire analysis of Section 7.1 is valid for this problem. The solar panel characteristics are primarily a function of the illumination and temperature. The exact functional relationship is not known and varies from cell to cell. The solar array is being used to provide power to a load, such as a battery for energy storage. The test case uses 40 cells in series to provide a nominal output voltage of 20 V and

105

a nominal current of 4 A. There are six lead acid batteries in series resulting in a nominal output voltage of 72 V. The boost converter is used to control power transfer.

10

15

20

25

VOLTAGE (V)

Figure 7.2 Solar array characteristics 7.2.2 Implementation The output current and voltage are sensed, and their time derivatives are computed. Using analog multipliers, the correlation is computed. The correlation output is used to determine the duty ratio, which, in turn, controls the solar array current. Figure 7.3 shows the schematic of the MPPT. A capacitor is connected at the input of the boost converter. This makes the current out of the solar array nearly a constant. This improves the tracking since the current out of the array becomes more like a dc quantity. In addition, it offsets the effect of the capacitance of the solar array. In absence of this capacitor, the capacitance of the solar array introduces a phase shift and interferes with the tracking algorithm. A detailed circuit diagram is given in Appendix A. 106

I BATTER>Y
5ULAK /x

s v^

I V

SOURCEI

BOOST CONVERTER

' \d/dt/ dl/dt

\d/dt /
dP/dt

Duty Ratio

"iT

(dP/dt).(dI/dt)

Figure 7.3 Schematic of MPPT 7.3 Experimental results 7.3.1 Static data For the purpose of verification of the MPPT, we construct a dummy source consisting of a voltage source in series with a resistor. The maximum power point is given by the maximum power transfer theorem. The input and output power of the MPPT are compared to the theoretical maximum and the results have been tabulated. The ratio between the power into the MPPT and the theoretical maximum is defined to be the tracking efficiency. The data are taken at three different output voltages and five different input impedances. The tracking efficiency is found to be better than 99% from full load to 30% of the full load. Below that, the tracking deteriorates to 98% at 20% load. At the lowest loading, the tracking algorithm does not converge, but the tracking efficiency is about 97% . The power conversion efficiency is about 95% at full load and close to that at other loads. 107

Table 7.1 Tracking and conversion efficiency data


Input Voltage (V) 19.5 19.4 19.2 20.8 20.7 20.5 20.8 20.2 20.3 16.3 16.1 16.1 16.2 16.0 16.0 Input Power (W) 70.9 70.9 71.0 50.8 50.7 50.7 28.1 28.2 28.1 16.3 16.3 16.2 9.1 9.1 9.1 Input Impedance
(Q)

Output Voltage (V) 70.9 67.5 62.6 74.5 68.2 63.4 74.5 68.6 63.9 73.9 68.4 63.6 72.8 67.3 62.5

Output Power (W) 67.3 67.2 67.2 48.6 48.6 48.6 26.9 27.1 27.0 15.6 15.6 15.5 8.7 8.7 8.7

=|

Tracking Efficiency
(%)

5.2 5.2 5.2 7.4 7.4 7.4 12.8 12.8 12.8 21.9 21.9 21.9 38.5 38.5 38.5

100 99.9 99.9 99.6 99.5 99.6 99.6 100 99.5 98.9 98.6 98.3 97.0 97.0 97.0

7.3.2 Dynamic data Consider a solar panel on a moving vehicle. Rapid changes in illumination or battery voltage might occur. These are classified as source and load transients, respectively. During a load transient, the input current changes initially due to the inductance of the boost converter. However, after the tracking algorithm takes effect, the input current from the solar array goes back to its optimal value. Figure 7.4 shows this result. The response time of the

108

tracker is on the order of 2 ms. The load transient is about 20% of the load voltage. The load voltage and the input current are plotted in Figure 7.4.

Xaxis: Time (0.5 ms/div), Yaxis: Output voltage (10 V/dv ac)(upper), Input current (1 A/div ac)(lower) Figure 7.4 Load transient response A fast transient in the illumination when the array is momentarily under shade corresponds to a source transient. In the dummy source this is realized by keeping the voltage source constant and changing the resistor. The optimal input current changes but the optimal input voltage stays the same. Due to the inductance, initially the input current does not change. The input voltage changes and the system moves from the optimal operating point. Due to the action of the tracking algorithm, the input voltage and current return to the optimal value. Figure 7.5 has plots of the input current and voltage. The system response time is on the order of 4 ms for a 35% change in input power.

109

1
m"*
1

"' I I
]

j
MxtisSii v & t o * . Vivwvx'. 'M&V&l

:.;w/j 1

T')~^

,i,

* ' * - W W , - * , *H<M-*V.'iv.-.M.w,l-.

X-axis: Time 2 ms/div Y-axis: Input Voltage 5 V/div (upper), Input Current 1 A/div (lower) Figure 7.5 Source transient response 7.4 Active Filter A feedforward active filter has been developed to cancel ripple in dc to dc converters [4]. The canceUation current, wWch is me negative of e ripple current, is injected at m^ output node. The ripple current is derived fmm me switch voltage wmch can be ^ ^ ^ The cancellation current also depends on the filter inductance which is varies with load current and temperature. A tuning scheme is used to track this variation. Details of the circuit and operation are presented in [4]. The expression for the cancellation current is as follows.

comp

= _ UV/,

(7.9)

110

The value of L is not accurately known and is estimated to be L,. The theoretical reduction in the ripple with this estimate is obtained. inpple(comp)
'nm*("0
com

m({__L_

(7.10)

P)

To obtain good ripple cancellation, a tuning scheme is developed. The correlation between the cancellation current and the ripple is used to tune the estimate for L. Figure 7.6

SWITCHED POWER CONVERTER

LOAD

comp Current Buffer

Figure 7.6 Tuned active filter for ripple cancellation The active filter works to cancel the ripple at the 100 kHz frequency range. The speed of the tuning process is slower. Fast tuning is required in response to a large change in the load current that changes the effective inductance. There is an output transient in response to the load change and the tuning circuit responds on a similar time scale. Figure 7.7 shows the output following a 2 A load transient. Note that the ripple cancellation is detuned by the load transient and tunes in less than 0.6 ms. Ill

X axis: 100 us/div Yaxis: Output voltage (100 mV/div AC) (upper), Load current (5 A/div) (lower) Figure 7.7 Transient response of tuned active filter 73 Conclusion The auto tuning scheme developed here is a practical, simple and fast method for a variety of tuning problems in switched power converters. The scheme is easily implemented with simple hardware with excellent performance. The hardware required is simple and readily integrable in an I C The scheme uses a characteristic of switched power converters, the ripple in all the state variables, to provide excitation input for the tuning process. This simplifies the implementation and makes the tuning process faster than comparable tuning schemes. The theory and design of the tuning scheme are fairly simple and applicable to many tuning problems associated with electronic power conversion.

112

CHAPTER 8 CONCLUSIONS
One of the primary objectives of this thesis is the analysis of simple dc to dc converters and their control circuits, with particular emphasis on the nonlinear nature of the systems. This analysis yields important insight into the control issues, and nonlinear control laws have been developed that improve performance. Control of practical converters has been implemented with significant improvement over existing systems. Section 8.1 outlines analytical contributions of this thesis. Section 8.2 highlights some of the practical results, in which the systems developed performed significantly better than conventional systems. Section 8.3 outlines directions for future work. These include system designs that seem promising but have not been implemented owing to system complexity. 8.1 Important Analytical Results The important contributions include analysis of the PWM noise process and synthesis of a general adaptive tuning process. Sensorless current mode control was developed as an alternative to conventional linear feedback, with with the result of an improved noise sensitivity and dynamic range. Important analytical results have been developed for boundary control. Control boundaries have been developed for geometric control that implement optimal control or minimize a Lyapunov function. 8.1.1 PWM noise analysis The PWM process was analyzed for noise behavior from the point of view of switching power conversion. The noise is modeled as band-limited Gaussian noise. The usual

113

linearizing assumption was avoided. Consequently, the results obtained predict behavior not observed in earlier work. These results are confirmed by experiment and simulation. 8.1.2 Closed-loop noise analysis Noise sensitivity of different control methods has been quantified for the first time in this thesis. The noise performance of different methods has been known qualitatively. These observations have been validated, and a better understanding of the process has been achieved. The noise sensitivity difference between peak current mode control and voltage mode control is often more than an order of magnitude. These differences have very significant implications for the design of converters. Subharmonics have been observed in the output of switched power converters due to noise in the control loop. This phenomenon has been analyzed and closed-form expressions have been obtained for the subharmonics. Again, the results are consistent with simulation and experiment. 8.1.3 Sensorless current mode control Sensorless current mode control has been developed as an alternative to current mode control. It reconstructs the current waveform from the internal voltages in the system. It has improved noise performance and dynamic range and eliminates the need to sense current. The improvement in signal strength is more than 40 dB over current mode control. Sensorless current mode control has been developed as a general control method which can be applied to a number of different converters. Control laws have also been developed for transformer coupled converters.

114

8.1.4 Optimal boundary control Switching boundaries have been developed for two state converters that optimize an infinite horizon cost function. The cost function has been chosen to be the integral of the square of the output error. Optimal boundaries have been developed that approach optimal control in the high switching frequency limit. 8.1.5 Lyapunov control Control laws have been developed that explicitly minimize an arbitrary Lyapunov function. The control law is expressed as a switching boundary in state space. Stability analysis has been performed for this control strategy and a stability criterion has been developed. It is a necessary criterion; a family of Lyapunov functions exists for which the explicit minimization is guaranteed to be stable in the large. Quadratic Lyapunov functions have been considered for two-state converters. A range of quadratic functions exist for which this control is guaranteed to be stable. These control laws minimize the Lyapunov function in each cycle, and the control constrains the system state during a large signal transient. 8.1.6 Tuning theory A generalized tuning procedure has been developed that can optimize a nonmonotonic function in a switched power converter. The ripple in the converter states is used as the excitation input for the tuning process. The tuning process is guaranteed to converge to the optimal solution provided the system does not have local maxima or minima.

115

8.2 Performance Improvements The implementation of the control laws developed in this thesis resulted in significant performance improvements. Sensorless current mode control was implemented with excellent performance in a variety of converter applications. A maximum power tracker has been developed for a solar vehicle application. An active filter has been developed for ripple cancellation that improves ripple performance in a variety of power converters. 8.2.1 Sensorless current mode control Sensorless current mode control has been implemented for buck, boost, and forward converters. A family of 2 V converters has been built that uses this control method. At this low output voltage, the elimination of the current sensing resulted in improved power conversion efficiency. A boost converter converting from 72 V to 210 V, using this control method, has been implemented for a solar vehicle application. 8.2.2 Maximum power point tracking Maximum power point tracking is required to obtain maximum power from a variable energy source such as a solar array. The maximum power point tracking problem has been implemented for a solar vehicle application. The scheme used here results in a tracking speed of about 2 ms and compared to 1 s for conventional trackers. 8.2.3 Active filter for ripple cancellation A feedforward active filter has been implemented that reduces ripple by about 15 dB and has been implemented for various power converters. The active filter consumes less than 1% of typical rated power and is appropriate for low voltage applications where ripple specifications are hard to meet.

116

8.3 Future Work A few ideas developed here were not implemented because of system complexity. However, many of these are practical, especially if integrated into the power controller IC. The implementation requires simple analog building blocks like op-amps, multipliers and analog switches. 8.3.1 Least-Squares current estimator A least-squares estimator can be used to reconstruct current waveforms or other states. This would be more robust to noise than the direct current sensing in current mode control. It would be able to reconstruct the dc part of the current, unlike sensorless current mode control. 8.3.2 Excess energy control The Lyapunov-based control minimizes a Lyapunov function. A more direct approach is to minimize the excess energy in the converter. This is appropriate for power converters since energy is conserved in these systems. The stability theory for this method has not been developed. It is expected that such control would have good, large-signal transient response. 8.3.3 Active filter with two-level tuning The active filter for ripple cancellation assumes a perfect inductor. However, any inductor has a significant series resistance. It is possible to cancel the ripple more accurately by taking into account this resistance. This resistance is nonlinear, and it is essential to tune for it. This would result in a two-level tuning active filter, and it is expected that such a filter would reduce ripple by 30 dB.

117

8.3.4 Nondissipative active clamp An active clamp has been developed that sets a hard limit on the output voltage during a large load transient. This active filter works by dissipating the excess energy. It should be possible to build an active clamp that would transfer this energy back to the source with little loss. Such a clamp would improve large signal transients without much loss in efficiency. 8.3.5 PI control with tuned gains Many power converters use simple PI or PID control for output feedback. Since a power converter is a nonlinear circuit, a single set of gains does not optimize performance at all operating conditions. A system that uses some kind of gain tuning could optimize performance over the entire range. A tuning scheme using correlation could be used for this application and significant performance improvement would be expected.

118

REFERENCES
[1] J. Kassakian, M. Schultz, and G. Verghese, Principles of Power Electronics. New York: Addison-Wesley, 1991. [2] L. H. Dixon Jr., "Closing the feedback loop," in Unitrode Power Supply Design Seminar Manual., 1986, pp. Cl-1 to Cl-7. [3] P. Midya and P. T. Krein, "Optimal Control Approaches to Switching Power Converters," in Power Electronics Specialists Conference Record, 1992, pp. 741-748. [4] P. Midya and P. T. Krein, "Feed-Forward Active Filter for Output Ripple Cancellation," International Journal of Electronics, vol. 77, no. 5, pp. 805-818, 1994. [5] R. B. Ridley, "A new, Continuous-Time Model for Current-Mode Control," in IEEE Transactions on Power Electronics vol. 6, no. 2, pp. 271-280, 1991. [6] W. Tang, F. C. Lee, and R. B. Ridley, "Small-Signal Modeling of Average Current-Mode Control," in IEEE Transactions on Power Electronics, vol. 8, no. 2, pp. 112-119, 1993. [7] R. M. Bass and P. T. Krein, "Continuous Time Formulation, Classification and Methods for Power Electronic Systems," in Proceeding of the Midwest Circuits and Systems Symposium, 1989, pp. 785-789. [8] R. M. Bass and P. T. Krein, "Phase-Plane Analysis of DC-DC Converters," in Proceeding of the North American Power Symposium, 1988, pp. 216-225. [9] P. T. Krein, J. Bentsman , R. M. Bass, and B. C. Lesieutre, "On the use of Averaging for the Analysis of Power Electronic System," in IEEE Transactions on Power Electronics, vol. 5, no. 2, pp. 182-190, 1990. [10] N. M. Blachman, "Gaussian Noise: Prediction based on its value and N Derivatives," in IEE Proceedings on Radar and Signal Processing, vol. 140, no. 2, pp. 98-102, 1993. [11] J. C. Candy and O. J. Benjamin, "The structure of Quantization Noise from Sigma-Delta Modulation," in IEEE Transactions on Communication, Vol. Com-29, pp. 1316-1323, 1981. [12] J. M. Goldberg and M B . Sandler, "Pseudo-Natural Pulse Width Modulation for High Accuracy Digital to Analogue Conversion," in IEEE Electronics Letters, Vol. 27, no. 16, pp. 1491-1492, 1991. [13] K. W. Ma and Y. S. Lee, "Technique for Sensing Inductor and dc Output Currents of PWM DC-DC Converter," in IEEE Transactions on Power Electronics, vol. 9, no. 3, pp. 346354, 1989. 119

[14] C. W. Deisch, "Simple switching control method changes power converter into a current source," in IEEE Power Electronics Specialists Conference Record, 1979, pp. 300-306. [15] L. Calderoni, L. Pinola, and V. Varoli, "Optimal feedforward compensation for PWM dc/dc converters," in IEEE Power Electronics Specialists Conference Record, 1992, pp. 235241. [16] W. Tang, R. B. Ridley, and I. Cohen, "Charge Control: Modeling, Analysis and Design," in IEEE Transactions on Power Electronics, vol. 8, no. 4, pp. 396-403, 1993. [17] C. K. Tse and K. M. Adams, "A nonlinear Large-Signal Feedforward-Feedback Control for two state DC to DC Converters," in Power Electronics Specialists Conference, 1991, pp. 722-729. [18] S. R. Sanders and G. C. Verghese, "Lyapunov-Based Control for Power Converters," in Power Electronics Specialists Conference Record, 1990, pp. 51-58. [19] L. B. Sobolev, "Optimal Control of Transients in DC-DC Converters," in Proceedings of the Power Conversion Conference, 1993, pp. 194-199. [20] C. C. Hang and K. K. Sin, "On-line Auto Tuning of PID Controllers based on the CrossCorrelations Technique," in IEEE Transactions on Industrial Electronics, vol. 38, no. 6, pp. 428-437, 1991. [21] C. R. Sullivan and M. J. Powers, "A High-Efficiency Maximum Power Point Tracker for Photovoltaic Arrays in a Solar-Powered Race Vehicle," in Power Electronics Specialists Conference Record, 1993, pp. 574-580. [22] J. H. R. Enslin and D. B. Snyman, "Combined Low-Cost, High-Efficiency Inverter, Peak Power Tracker and Regulator for PV Applications," in IEEE Transactions on Power Electronics, vol. 6, no. 1, pp. 73-82, 1991. [23] D. B. Snyman and J. H. R. Enslin, "Combined Low-Cost, High-Efficiency Inverter, Peak Power Tracker and Regulator for PV Applications," in Power Electronics Specialists Conference Record, 1990, pp. 67-74. [24] A. S. Kislovski, "Dynamic Behavior of a Constant-Frequency Buck Power Cell in a Photovoltaic Battery Charger with Maximum Power Tracker," in Proceedings of the IEEE Applied Power Electronics Conference, 1990, pp. 212-220. [25] G. Choe and M. Park, "Analysis and control of active power filter with optimized injection," in IEEE Transactions on Power Electronics, vol. 4, no. 4, pp. 427-433, 1989.

120

[26] R. Hudson, S. Hong and R. Hoft, "Modeling and simulation of a digitally controlled active rectifier for power conditioning," in Proceedings of the Sixth Applied Power Electronics Conference, 1991, pp. 423-429. [27] H. Jin, S. B. Dewan, and J. D. Lavers, "A new feedforward control technique for ac/dc switchmode power supplies," in Proceedings of the Seventh Applied Power Electronics Conference, 1992, pp. 376-382. [28] H. L. Jou, H. Y. Chu, and J. C. Wu, "A novel active power filter for reactive power compensation and harmonic suppression," in International Journal of Electronics, vol. 3, no. 2, pp. 577-587, 1993. [29] P. E. Martnelli and C. Ashley, "Coupled inductor boost converter with input and output ripple cancellation," in Proceedings of the Sixth Applied Power Electronics Conference, 1991, pp. 567-572. [30] P. Midya and F. Schlereth, "Dual switched mode power converter," in Proceedings of the Fifteenth IEEE Industrial Electronics Conference, 1989, pp. 155-158. [31] A. Nakajima, K. Oku, J. Nishidai, T. Shiraishi, Y. Ogihara, K. Mizuki, and M. Kumazawa, "Development of active filter with series resonant circuit," in IEEE Power Electronics Specialists Conference Record, 1988, pp. 1168-1173. [32] N. M. Blachman, "The SNR Threshold in PPM Reception," in IEEE Transactions on Communications, vol. Com-22, no. 8, pp. 1094-1098, 1974. [33] S. Haykin, Communication Systems. New Delhi: Wiley Eastern Limited, 1985. [34] R. E. Steele, Delta Modulation Systems. New York: Wiley, 1975. [35] R. B. Ridley, B.H. Cho, and F.C.Y. Lee, "Analysis and Interpretation of Loop Gains of Multiloop-Controlled Switching Regulators," in IEEE Transactions on Power Electronics, vol. 3, no. 4, pp. 489-497, 1988. [36] R. D. Middlebrook, "Small-Signal Modeling of Pulse-Width Modulated Switched-Mode Power Converters," in IEEE Transactions on Power Electronics, vol. 76, no. 4, pp. 343-354, 1988. [37] D. M. Sable, R. B. Ridley, and B. H. Cho, "Comparison of Performance of Single-Loop and Current-Injection Control for PWM Converters that Operate in Both Continuous and Discontinuous Modes of Operation," in IEEE Transactions on Power Electronics, vol. 7, no. 1, pp. 136-142, 1992.

121

[38] R. D. Middlebrook, "Modelling Current-Programmed Regulators," presented at IEEE Applied Power Electronics Conference, 1987. [39] H. Kwakernaak and R. Sivan, Linear Optimal Control Systems. New York: Wiley, 1972. [40] P. T. Krein and R. M. Bass, "Geometric formulation, classification and methods for power electronic systems," in IEEE Power Electronics Specialists Conference Record, 1990, pp. 499-505. [41] R. M. Bass and P. T. Krein, "Limit-Cycle Geometry and Control in Power Electronic Systems," in Proceeding of the Midwest Circuits and Systems Symposium, 1989, pp. 785-789. [42] R. Oruganti, J. J. Yang, and F. C. Lee, "Implementation of optimal trajectory control of series resonant converter," in IEEE Transactions on Power Electronics, vol. 3, no. 3, pp. 318-327, 1988. [43] J. Holtz and J. O. Krah, "Adaptive optimal pulse-width modulation for the line-side converter of electric locomotives," in IEEE Transactions on Power Electronics, vol. 7, no. 1, pp. 205-211, 1992. [44] w . W. Bums, m and T. G. Wilson, "Analytic derivation and evaluation of a state-trajectory control law for dc-to-dc converters," in Power Electronics Specialists Conference Record, 1977, pp. 70-85. [45] K. M. Smedley and S. Cuk, "One-cycle control of switching converters," in Power Electronics Specialists Conference Record, 1991, pp. 888-896. [46] G. C. Goodwin and K. S. Sin, Adaptive Filtering Prediction and Control. Englewood Cliffs, NJ: Prentice Hall, 1984. [47] K. Ogata, Modern Control Engineering. Englewood Cliffs, NJ: Prentice Hall, 1970. [48] R. M. Bass and P. T. Krein, "Large-signal design alternatives for switching power converter control," in IEEE Power Electronics Specialists Conference Record, 1991, pp. 882887.

122

APPENDIX A CIRCUIT DIAGRAMS


Appendix A contains circuit diagrams of some important circuits. These include circuits which have significant performance improvements over conventional practice as well as circuits used in experimental verification.

123

100

0-5V

Z7

Figure A.l Duty ratio analysis in presence of noise

124

Vsw SPECTRUM O
V+

HP4I95A

ANALYZER

50 100
77

UC 3843
0-5V

-vwIk
O-SIc R

77

-vws
Ik

OWVr
Vsw

BANDLIMITED _ INJECTED NOISE

IOOnF RANDOM NOISE GENERATOR G-R TYPE 13908

Ik Ik
77

IOOnF

IOOnF

77

IOOnF

Figure A.2 Spectral analysis of peak current mode control in presence of noise

125

100
VvV^-

Vsw

-r-CH
100

HP4195A SPECTRUM 50 ANALYZER

ft V+

v+
16 15 14 I]

77

200nF
18 17 12 II 10

UC 3526A

0-5V
I 2 3 4 5 6 7 8 9

BANDLIMITED INJECTED NOISE

. j

0-5k Vsw

IOCnF Ik

-VWV-

-W-, Ik R

8.8k 77

VOLTAGE MODE

Ik
Ik

Ik

Ik

CURRENT MODE

IOOnFIOOnF
77

IOOnF

RANDOM NOISE GENERATOR G-R TYPE1390B

Figure A.3 Spectral analysis of voltage and current mode control with ramp in presence of noise

126

Vf

ICU667 1 2 3 4

5V

(OUTPUT) I _i_ IS 17

16

^ $V
I) u
4 <

r^
I] 12 II 10
6 7 g 9

-y
SM

-i

w,?r_p5W,_^
7uH

UC 3526A

isoopr I9T0CI)
I Ok -TAW

ru
100k

5K '--AW 1

Figure A.4 SCM control: 5 V to 2 V buck converter

127

Vicmrt.)

Figure A.5 Voltage mode control 5 V to 2 V buck converter

128

luF
5V (OUTPUT:

fHH
18 17 16 IS 14 1 2 ) 4 5 6

rW-

= ffr
II 12 II 10

r
lOOOpF

MUR460

MUR4M)

200OpF

UC 3526A

LOAD

i
I Ok

r
V

-^Wr-

^
2000pF CD4066

Vffr

-vwvI00K -^W,i I Ok

7 HP22II

1 V "

-AW1 2 )

I
4| 10k

W-

-vWr-

100K
NVW

2001

2001

Figure A.6 SCM control 72 V to 210 V boost converter

129

mil

Core ll-crrnx Cuhcl 21HXT2VI H-2A

./?'
7T
i

-5-

I I lull 220uf"

Lojd

g
28 j |

mil.

22ul

lllk

68opr

Lr\
i (ik

opjmp 0 luf"

-JI 2 2ul

\
<>
47 100

Unity Gain Buffer

opjmp LH400I

2 2uF

Figure A.7 Feedforward active filter for ripple cancellation

130

1
-fli-

= 0
-ISHI5 -wvjl-

-WC-HI-

iS

-wIi? i

W-Hi-

fi

-*|i-

id

-^f|i

l0

lulu
131

:u

w^-l ..I

I^H

1 ^ 1

_^ I I ' U . U - - - L ' I

? I

r^n

IS _

I '"LJ - ' U

,. I LJJ,.l f LJi

Figure A.9 48 V to 2 V converter (single stage)

132

i n . 1667

-i
?

J\V
rft |\ 14

T
R?
It i:

i
II

1,
io

UC 3526A

1
6 >
V

;if55in.

-it

7h

'

J66CTI2) J E l A t o r t

n
" i_| ''

j 000 J

Um

<T5oTT

f4
TV

2 $

10k

t>o

DX^j-1

620pf

U =T

IJOOpf

Figure A.10 5 V to 2 V Converter

133

Vmt
O

vgate (m place of doubter)

Hi'

LuuLJ

IMSLSUJLMQ.}

TTI
y MUP120

^rmrrinmr
MUP120 V

Gate A

\/m

j ^ - ^ Gale 8 ^ > Gate A

l8v

Figure A . l l 48 V to 12 V Converter

134

APPENDIX B DERIVATION OF NOISE IN PWM


We start at Equation (3.5) of Chapter 3. The detailed mathematical analysis is presented here. P(X+AX)-P(X)=(1-P(X))Q (B.l)

The comparator resets the switch when the sum of the noise and signal is smaller than or equal to the ramp value. The signal value is given by D0VRP, which is the ramp voltage at duty ratio D0. To satisfy the condition that the duty ratio is X or more, the sum of the noise and signal is less than XV^p. Thus the noise (Y) is constrained to values greater than (X D)Vw, P(Af)B)=P(B)P(A\B) OR P(A\B)=P(Af]B) P(B) (B.2)

To compute Q, it is necessary to analyze the noise process in some detail. Equation (B.2) is applicable to continuous probability distributions as well. Note that Equation (B.2) is the same as Equation (3.4) of Chapter 3. Define A to be the unconstrained probability density function of the noise Y. Let B be the event in which the noise (Y) is constrained to values greater than (X -D0)VRP. Thus P(B) is equal to 1-P2((X -D^)\^). Define p3(Y) to be the conditional pdf of noise given that there is no transition between 0 and X. This conditional probability can be represented in terms of the original pdf of the noise using Equation (B.2). 135

^(n=.I-f,(^(%-D^)

(B.3)

There is a comparator transition between X and X+AX if the sum of the noise and the signal crosses the ramp voltage in this interval. The random variables Y and Z must satisfy some conditions for this to occur. This corresponds to a triangular region in the Y-Z plane as shown in Figure B.l. The conditional probability Q is obtained as an integral of the pdf of Y and Z over the area.

Figure B.l Area integral in noise

136

The area in the Y-Z plane can be explained as follows. The ramp rises at the rate of
VRP/TS

and if the noise derivative is larger than that there cannot be an intersection. This

eliminates all Z >VRP/TS. There is no transition the between the duty ratios of 0 and X. This eliminates the region Y less than (X -D0)VRP from the integral. There is a transition in the interval X to (X+AX) if the ramp equals the signal plus noise. The signal has zero slope, the ramp has slope (VRJ/TS) and the noise has slope equal to Z. Thus they approach each other at the rate of (V%p/Ts-Z). If the starting value of noise is such that there will be an intersection between X and (X+AX), then the duty ratio lies between X and (X+AX). This corresponds to the triangular region in Figure B.l.

Q=p2(Y)jjp2(Z)dydz where dy=AX(-Jl-Z) Ts

(B.4)

The variation of Y is limited since AX and, consequently, AT is a quantity tending to zero. Therefore, it is possible to neglect variations in p(Y) and take this quantity out of the integral. This allows us to represent Q in a simpler form. This is exact in the limit as AX tends to zero.

-as%rK-.*'-

137

The integrand can be evaluated and the result is as follows. It consists of two parts. One corresponds to the integral of the Gaussian probability density function and hence is the Gaussian cumulative probability distribution function. The other term is the integral of [Zp2(Z)], which is integrable in closed form.

^ * S % -' *- M, is a dimensionless quantity obtained from the integral in Equation (B.5). It turns out to be the noise Modulation Index of the PWM process.

M=P,(Rcz)+JL
J2KR

where /?=_JL T a sz

(B.7)

The variable R is a ratio of the slope of the ramp and the standard deviation of the noise derivative. It is a normalized measure of the high frequency content of the noise. Equation (B.l) can now be expanded to form the following equation.

P(X+AX) -P(X) =AX VRP Pl{y^X~D^ Ml 1 -P(X)) " i -f,(y*,(x-Do)) '

(B.8)

138

The quantity p,(Z) is the derivative of P,(Z), which allows us to integrate the above equation in closed form.

' dx i-p(X)

' dx ^

-M,4;l. nJ,^J i-p^v^x-D^y

Integrating both sides of the equation results in the following PDF in closed form:

log(l-P(X))=M, log(l-?,(y*/X-Do)))

(B.10)

Taking the antilog of the equation, we obtain the following result: P(X)=1 -[1 -PJV^X-D^f(B.l 1)

The quantity M, has been defined to be the modulation index of the PWM noise process. The PDF of the duty ratio is modulated from the PDF of the noise by the modulation index M,. Physically M, can be represented as the ratio of expectation of the switch transition due to the ramp slope plus noise, to the expectation of a transition due to ramp slope only. In the limiting case, when the noise is zero, the quantity M, is equal to unity. However, in all other cases it is larger than unity. Thus, we have arrived at Equation (B.l 1), which is the same as Equation (3.9) of Chapter 3.

139

APPENDIX C DERIVATION OF NOISE IN CLOSED LOOP PWM


Starting from Equation (4.1) of Chapter 4, the detailed mathematical analysis of the equation is presented here.

/((N+D^r,)^^/)^^

(c.i)

The results from two consecutive cycles can be subtracted to obtain the following relation.

/W^-V^V^-.?,)

(C.2)

Let the difference between the signal and the ramp value at the switch transition be called EN. This is a local error contributing to the duty ratio error. We can also define a global error FN, which is the difference between the duty ratio and its nominal value. In a system with no memory beyond a cycle, this would be the only term. /((W+D^+C^D^, (C.3)

Thus the current at the end of the cycle can be calculated as follows. KTJLN+l))~(TfipN-Et}+U-DJTffF (C4)

Combining Equation (C.3) with Equation (C.4) for step N+l, we obtain the following equation:

D _ = ^ ^ ^ ' ^ +_A_
140

(C.5)

Expressing the result in terms of the variation of the duty ratio from its nominal value, we obtain the following equation:

The overall errors FN can be expressed in terms of the local errors E^. Equation (C.7) is obtained by substituting consecutive values of FN in terms of E^. &.,= ' SEH.x<^'^^KEN_^K1EN.^KiEN.r )] (C.7)

The quantities E which are the local errors in cycle I, are independent of each other. Thus, the standard deviation of the error in duty ratio FN can be calculated in terms of the standard deviation of the voltage error EN from the summation. Using the linear PWM assumption, the standard deviation of EN is the same as the standard deviation of the noise o \ The duty ratio in the absence of noise is a constant. Therefore, the standard deviation of FN is the same as the standard deviation of the duty ratio (o*D). \ll+(K-l)2m-K)2K2+(l-K)2K*+

o0=aF*>

^
TS(SR+SN)

. . .

(C.8)

Taking the sum of the geometric series we obtain the following expression.

*o=

(C.9)

TJS+SJs
141

l+K

In the case of voltage mode PWM or current mode PWM with small current feedback, the quantity 'K' tends to unity and the duty ratio variation is as it is in open loop. In current mode control with the optimal ramp, the value of 'K' is zero and the standard deviation of the duty ratio variation is V2 times the open-loop standard deviation. In the case of peak current mode control with no ramp, the quantity 'K' is a function of the duty ratio. When SR is equal to zero:

K-^T

% = ( ! - %

- ? = - - % -

(CIO)

The result in Equation (C.9) can be represented in terms of the average duty ratio as follows.

aB-.

"r

%'-Oo)

(CI1)

* % + . V \ | !-%>o

Equation ( C l l ) provides the time domain result of closed loop noise in PWM. The frequency domain result is then obtained from the time domain result. The energy spectral density of a random process can be obtained from the autocorrelation function of the process. From Equation (C.7) we know that the duty ratio variation in any cycle is correlated to the duty ratio variation of the previous cycles. An approximate expression can be obtained for the spectrum by treating the duty ratio as a discrete variable and obtaining the autocorrelation in the discrete time domain, and then taking its fourier transform to obtain the energy spectral

142

density in the frequency domain. While these results are not exact they provide an approximate closed-form expression for the spectrum. The autocorrelation function and the power spectral density are a Fourier transform pair. The autocorrelation function is the expected value of the product of the signal with a delayed version of itself. The function is defined as follows. Rx(x)=<x(t)x(t+T)> (CI 2)

For a standard deviation in the duty ratio o>, the autocorrelation for zero time difference can be obtained as follows. #/0)=<f(n)->=oy This can be expanded to the following form using Equation (C.12).
2 cY

RJ0)=

(C.14)

r^i

%+&/

For a time difference of 'm' cycles the correlation can be obtained from Equation (C.7).

Lys('->K+'VJ

We note that E, from different time intervals are uncorrected to each other. Thus, in calculating the expected value, the cross terms have an expected value of zero. This simplifies the expression to a simple geometric series. 143

RfinTJ*

<[EN2(K-l)Km-lHK-l)\EN_2K^^K^EN^...)]>

(C.16)

The result can be summed in closed form as follows.

^ ^ '

Td^r

(CI7)

The Fourier transform of the discrete autocorrelation function is the power spectral density. However, calculating its Fourier transform is not trivial. Still, this is obtainable in closed form using the following theorem for a Fourier transform of a geometric series. If the autocorrelation sequence is a geometric series of the form (1, K, K2, K 3 . . . ) , then the power spectral density is given by the following equation.

SKK(f)=

L (1 -Kcos(2Kfi)2 +(/s:sin(27r/))2

(C18)

The time scale in Equation (C.17) is Ts and not unity. This is normalized and the infinite geometric series translates to a continuous frequency spectrum as follows:

2oV(l-cos8) SJf)= _ where Q=2nf/F, s (l+K2-2KcosB)Ts(SR+SN)2

_ n (CI?)

144

The maximum of this spectrum occurs at half the switching frequency. This corresponds to the first subharmonic. The energy spectrum at this frequency is as follows. 4<J2 12) = 1 (l+#%+f/

S ^

(C.20)

Thus, we have arrived at Equation (4.11) and Equation (4.12) of Chapter 4.

145

VITA
Pallab Midya was bom in Bhilai, India, on January 31, 1967. He received his Bachelor of Technology (Hons.) in Electronics and Electrical Communication Engineering from the Indian Institute of Technology, Kharagpur, India, in 1988. He received his Master of Science in Electrical Engineering from Syracuse University, Syracuse N.Y., in 1990. From 1988 to 1990 he worked as a Graduate Research Assistant in the area of Computational Electromagnetics. During 1992 he worked as an Intern for IBM, at the T. J. Watson Research Center, Yorktown, NY. Since 1990 he has been pursuing doctoral studies at the University of Illinois. He held the positions of Graduate Research Assistant and Graduate Teaching Assistant during this period. He also received the National Talent Scholarship from N.CE.R.T. in India from 1984 to 1988.

146

Vous aimerez peut-être aussi