9
2 IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 18, NO. 1, JANUARY 2003 Small-Signal Modeling and Control of ZVT-PWM Boost Converters Jong-Lick Lin, Member, IEEE, and Chin-Hua Chang Abstract—A small-signal model for zero-voltage-transition pulse width modulation (ZVT-PWM) boost converters is derived in this paper. The accuracy of the developed model is verified by experimental measurement. It shows that the ZVT-PWM boost converters exhibit better dynamical behavior than the conventional PWM boost converters. To achieve output voltage regulation and line voltage disturbance rejection of ZVT-PWM boost converters, a classical controller and a modified integral variable structure controller are designed. The output voltage regulation in the presence of line voltage and load variations for both controllers are compared by experimental results. Index Terms—Modified integral variable structure control, small-signal model, ZVT-PWM soft-switching boost converter. I. INTRODUCTION T HE ZVT-PWM soft-switching converters [1]–[3] solve the existing problems of high switching losses of conventional PWM converters and high voltage and current stresses of reso- nant converters. By taking advantages of pulse width modula- tion and tank resonance, the soft-switching boost power con- verters with constant frequency operation are considered in this paper. These kinds of converters are especially useful in the ap- plication of high efficiency power converter systems. In the literature, all research on ZVT-PWM soft-switching converters have been restricted to steady-state analysis. The as- sumption that the small-signal model should be the same as con- ventional PWM DC-DC converters is not confirmative. There- fore, a small-signal model for a ZVT-PWM boost converter is developed in this paper. The differences between ZVT-PWM and conventional PWM boost converters are shown by com- paring their dynamical models. The main purpose to derive a mathematical model is required for controller design to achieve output voltage regulation and line voltage disturbance rejection. In this paper, the averaging method for two-time-scale discontinuous system (AM-TTS-DS) [4] is used to derive a small-signal model of ZVT-PWM boost converters. The model is verified by experimental results. Then based on this model, a classical controller and a modified integral variable structure controller (MIVSC) [5] are designed. Finally, the overall system is implemented in the laboratory and the system performances for the two controllers are compared. Manuscript received September 3, 1999; revised October 1, 2002. This work was supported by the National Science Council of the Republic of China, under Grant Contract NSC 88-2213-E-006-095. Recommended by Associate Editor M. A. Rahman. The authors are with the Department of Engineering Science, National Cheng Kung University, Tainan, Taiwan 701, ROC (e-mail: [email protected]). Digital Object Identifier 10.1109/TPEL.2002.807176 II. ZVT-PWM BOOST CONVERTERS A ZVT-PWM soft-switching boost converter [3] is shown in Fig. 1. It includes a conventional PWM boost converter drawn with thin lines and some additional components drawn with thick lines. The additional components are composed of a res- onant tank, auxiliary diode and snubber capacitor . The resonant tank includes an auxiliary switch and resonant in- ductor and capacitor . The zero-voltage soft switching of the main switch is achieved by tank resonance which is controlled by the auxiliary switch . From a circuit point of view, the system operation in one switching period can be divided into eight linear circuit stages. The key waveforms at steady state are shown in Fig. 2. All switching components work in soft-switching conditions without extra voltage/current peaks except suffering from larger current spike in stage 4 of time interval . The detailed analysis of circuit operations was proposed in [3]. The steady-state analysis for stages 1 and 2 are shown as follows. Stage 1 (see Fig. 3): The auxiliary switch turns on at . Current starts to flow through the resonant tank and the reso- nance occurs. Circuit analysis gives (1) (2) (3) for with and . Stage 2 (see Fig. 4): When becomes greater than , turns off and stage 2 starts. The increasing current offered by leads to the decreasing of . Circuit analysis yields (4) (5) (6) 0885-8993/03$17.00 © 2003 IEEE

Small-signal modeling and control of ZVT-PWM boost

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Small-signal modeling and control of ZVT-PWM boost

2 IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 18, NO. 1, JANUARY 2003

Small-Signal Modeling and Control ofZVT-PWM Boost Converters

Jong-Lick Lin, Member, IEEE,and Chin-Hua Chang

Abstract—A small-signal model for zero-voltage-transitionpulse width modulation (ZVT-PWM) boost converters is derivedin this paper. The accuracy of the developed model is verifiedby experimental measurement. It shows that the ZVT-PWMboost converters exhibit better dynamical behavior than theconventional PWM boost converters. To achieve output voltageregulation and line voltage disturbance rejection of ZVT-PWMboost converters, a classical controller and a modified integralvariable structure controller are designed. The output voltageregulation in the presence of line voltage and load variations forboth controllers are compared by experimental results.

Index Terms—Modified integral variable structure control,small-signal model, ZVT-PWM soft-switching boost converter.

I. INTRODUCTION

T HE ZVT-PWM soft-switching converters [1]–[3] solve theexisting problems of high switching losses of conventional

PWM converters and high voltage and current stresses of reso-nant converters. By taking advantages of pulse width modula-tion and tank resonance, the soft-switching boost power con-verters with constant frequency operation are considered in thispaper. These kinds of converters are especially useful in the ap-plication of high efficiency power converter systems.

In the literature, all research on ZVT-PWM soft-switchingconverters have been restricted to steady-state analysis. The as-sumption that the small-signal model should be the same as con-ventional PWM DC-DC converters is not confirmative. There-fore, a small-signal model for a ZVT-PWM boost converter isdeveloped in this paper. The differences between ZVT-PWMand conventional PWM boost converters are shown by com-paring their dynamical models. The main purpose to derive amathematical model is required for controller design to achieveoutput voltage regulation and line voltage disturbance rejection.

In this paper, the averaging method for two-time-scalediscontinuous system (AM-TTS-DS) [4] is used to derivea small-signal model of ZVT-PWM boost converters. Themodel is verified by experimental results. Then based on thismodel, a classical controller and a modified integral variablestructure controller (MIVSC) [5] are designed. Finally, theoverall system is implemented in the laboratory and the systemperformances for the two controllers are compared.

Manuscript received September 3, 1999; revised October 1, 2002. This workwas supported by the National Science Council of the Republic of China, underGrant Contract NSC 88-2213-E-006-095. Recommended by Associate EditorM. A. Rahman.

The authors are with the Department of Engineering Science, National ChengKung University, Tainan, Taiwan 701, ROC (e-mail: [email protected]).

Digital Object Identifier 10.1109/TPEL.2002.807176

II. ZVT-PWM BOOSTCONVERTERS

A ZVT-PWM soft-switching boost converter [3] is shown inFig. 1. It includes a conventional PWM boost converter drawnwith thin lines and some additional components drawn withthick lines. The additional components are composed of a res-onant tank, auxiliary diode and snubber capacitor . Theresonant tank includes an auxiliary switch and resonant in-ductor and capacitor .

The zero-voltage soft switching of the main switch isachieved by tank resonance which is controlled by the auxiliaryswitch . From a circuit point of view, the system operation inone switching period can be divided into eight linear circuitstages. The key waveforms at steady state are shown in Fig. 2.All switching components work in soft-switching conditionswithout extra voltage/current peaks except suffering fromlarger current spike in stage 4 of time interval .

The detailed analysis of circuit operations was proposed in[3]. The steady-state analysis for stages 1 and 2 are shown asfollows.

Stage 1(see Fig. 3): The auxiliary switch turns on at .Current starts to flow through the resonant tank and the reso-nance occurs. Circuit analysis gives

(1)

(2)

(3)

for with and .Stage 2(see Fig. 4): When becomes greater than ,

turns off and stage 2 starts. The increasing currentoffered by leads to the decreasing of . Circuit analysisyields

(4)

(5)

(6)

0885-8993/03$17.00 © 2003 IEEE

Page 2: Small-signal modeling and control of ZVT-PWM boost

LIN AND CHANG: SMALL-SIGNAL MODELING AND CONTROL 3

Fig. 1. ZVT-PWM boost converter.

Fig. 2. Key waveforms at steady state.

Fig. 3. Circuit in Stage 1.

for with with ,

,

,

and .

Fig. 4. Circuit in Stage 2.

III. A PPROXIMATESOLUTIONS FORSTAGES1 AND 2

The solutions for in (4) and in (5) are too com-plicated so that it is difficult for small-signal analysis. The sim-plification first begins with stage 1 and then extends to stage 2.Calculation of is necessary in Table I.

Since and are viewed as short timeperiods comparing to the switching period, it is reasonableto assume that in stage 1 and in stage 2. Then,it follows from the fact and as that(1)–(3) can be approximated to

(7)

(8)

(9)

and (4)–(6) can be approximated to

(10)

(11)

(12)

It follows from (11) that

(13)

Page 3: Small-signal modeling and control of ZVT-PWM boost

4 IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 18, NO. 1, JANUARY 2003

TABLE IVOLTAGES AND CURRENTS ATEACH STAGE

Fig. 5. Exact and approximate solutions: (a)i (t) and (b)� (t).

by boundary condition . Substituting (13) into (10)yields

(14)

The approximation results with load are shownin Fig 5. The exact solutions for in (4) and in (5)are shown with solid lines, and the corresponding approximatesolutions in (10) and (11) are shown with dotted lines. It is seenthat the approximation results are very close to the exact solu-tions, and thus the simplification is reasonable.

IV. CALCULATIONS OF AND IN TABLE I

For the convenience in latter analysis, the capacitor voltagesand in Table I are now calculated. Since

elements , and are all power lossless, from an energyconservation point of view we conclude that

total energy stored in and at

total energy stored in and at

total energy stored in and at plus

energy supplied by from to (15)

It follows from the second equality in (15) and ,and, that

(16)

where

Page 4: Small-signal modeling and control of ZVT-PWM boost

LIN AND CHANG: SMALL-SIGNAL MODELING AND CONTROL 5

Fig. 6. Equivalent circuit model of the ZVT-PWM boost converter.

(17)

and thus we have

(18)

from (16). Moreover, following from the first equality in (15)and , and gives

(19)

Together with (14) and (18), we get

(20)

The resonant currents, voltages and duration of each stage aresummarized in Table I.

V. MATHEMATICAL MODEL DERIVATION

The small-signal model of the ZVT-PWM boost convertershown in Fig. 1 will be derived based on AM-TTS-DS [4]. TheZVT-PWM boost converter can be divided into two subcircuits,a fast subcircuit and a slow subcircuit. The former includes faststate variables , and . The latter includesslow state variables and . The equivalent circuitmodel of the ZVT-PWM boost converter is shown in Fig. 6where the part enclosed by dotted line is the equivalent fast sub-circuit. Current is the sum of diode currents and

in Fig. 1.The ZVT-PWM boost converters are used to obtain a well

regulated output voltage which is just the slow state vari-able of the capacitor voltage . From a fast subcircuit pointof view, the slow state variables in Fig. 1 can be treated as con-stants. This leads to the results in Section II thatand . On the contrary, from a slow subcircuitpoint of view, only averaging behaviors of the fast state variableshave effects on the slow state variables. As a result, substitutingthe fast variables by their moving averages, the averaged model

of the slow subcircuit is then obtained. Viewing this, the stateequation of the equivalent circuit model in Fig. 6 is given by

(21)

with output equation

(22)

Then, taking the moving average for and listed inTable I, the state-space averaged model of (21) can be expressedas

(23)

where , and are defined by

(24)

(25)

(26)

with

(27)

(28)

(29)

With the help of the notation, (23) can also be rewritten as anonlinear and time-invariant state equation

(30)

and output equation

(31)

It is interesting to note that (30) becomes the dynamical equationof the conventional PWM boost converters if both and

inside the last bracket in (30) vanish.To obtain the linearized state equation, the nonlinear equa-

tion (30) is expanded into the Taylor series around the nominaloperating point Q, and then the higher order nonlinear terms

Page 5: Small-signal modeling and control of ZVT-PWM boost

6 IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 18, NO. 1, JANUARY 2003

are neglected. The nominal operating point can be easily ob-tained by setting and in (30). Intro-

ducing , ,, , , and

, the mathematical manipulation gives

(32)

The solutions for and in (32) are given by

(33)

with , where

(34)

(35)

(36)

With the help of (33), it leads to the conclusion that

(37)

which approximates to the voltage gain of the conventionalPWM boost converter. Then, by introducing small perturbationsaround the operating point

(38)

with

(39)

it is verified that

(40)

where

(41)

Fig. 7. Root locus for ZVT-PWM converter as a function of switching periodT .

Fig. 8. Bode plot of ~� (s)= ~d(s): (1) ZVT-PWM boost converter and(2) conventional boost converter.

Fig. 9. Bode plot of~� (s)=~� (s): (1) ZVT-PWM boost converter and(2) conventional boost converter.

(42)

(43)

(44)

with

Page 6: Small-signal modeling and control of ZVT-PWM boost

LIN AND CHANG: SMALL-SIGNAL MODELING AND CONTROL 7

Fig. 10. Root locus for ZVT-PWM boost converter as a function of loadR .

Fig. 11. Root locus for conventional PWM boost converter as a function ofloadR .

and . Nowperforming Laplace transform of (40), the transfer functionsfrom line voltage and control voltage to output voltage

(45)

and

(46)

are obtained after some calculations, where the denominatorpolynomial is given by

(47)

and is the controller output signal applied to the inputof the PWM. The small-signal gain of the PWM is given by

with denoting the peak valueof the modulating signal used in PWM [6], [7].

Fig. 12. ZVT-PWM boost converter with driving circuit and feedbackcontroller.

Fig. 13. Frequency response of~� (s)=~� (s).

VI. COMPARISONWITH PWM BOOSTCONVERTERS

It is seen from definitions (41)–(44) that all of , ,and vanish if approaches infinity. Thus the av-

eraged mathematical model in (40) of the ZVT-PWM boost con-verters becomes equivalent to the averaged model of the conven-tional PWM boost converters.

Moreover, the pole polynomial in (47) is a functionof the switching period , and thus the root-locus plot of

with a variable parameter is drawn in Fig. 7with . It reveals that the poles of the aver-aged model of the conventional PWM boost converter are

with damping ratio andnatural frequency rad/s, as shown in Fig. 7 with

. For switching frequency 50 kHz, the poles ofthe averaged model of the ZVT-PWM boost converter are

and with damping ratio , asshown in Fig. 7 with .

The Bode plots of and of the twoconverters are compared, respectively, in Figs. 8 and 9 forswitching frequency 50 kHz. It is seen that the ZVT-PWMboost converter exhibits better dynamical behavior with largerdamping ratio.

On the other hand, variations of load resistance can also af-fect the poles of the boost converter. Figs. 10 and 11 show theroot-locus plots for load for ZVT-PWM and conventional

Page 7: Small-signal modeling and control of ZVT-PWM boost

8 IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 18, NO. 1, JANUARY 2003

Fig. 14. Time responses of ZVT-PWM converter with classical controllerK (s): (a) output voltage� (t) and (b) control signal� (t).

Fig. 15. Time responses of ZVT-PWM converter with MIVSC controllerK (s): (a) output voltagev (t) and (b) control signalv (t).

Fig. 16. Output responses of ZVT-PWM converter with classical controllerK (s) under load variations: (a)R = 200! 100, V = 5V and (b)R =100 ! 50 V = 5V; vertical: 50 mv/div, ac; horizontal: 5 ms/div.

PWM boost converters, respectively. With the variations of load, the poles of the ZVT-PWM boost converter

migrate from to andand damping ratio changes from 0.92 to 12.46. While polesof the conventional PWM boost converter migrate from

to , damping ratio changesfrom 0.365 to 0.348 and natural frequencychanges from 743to 732 rad/s. It reveals that the poles of the ZVT-PWM boostconverter migrate to the negative real axis in the s-plane whilethe poles of the conventional PWM boost converter are almostfixed. Although the poles are sensitive to the load variations,

ZVT-PWM boost converters still exhibit better dynamical char-acteristics due to larger damping ratio.

VII. M ODEL VERIFICATION

In this section, the derived small-signal model is verifiedby the experimental results. A small power prototype of theZVT-PWM boost converter for theoretical verification isimplemented with the following parameters :

nominal input voltage 5 V;reference voltage 10 V;

Page 8: Small-signal modeling and control of ZVT-PWM boost

LIN AND CHANG: SMALL-SIGNAL MODELING AND CONTROL 9

Fig. 17. Output responses of ZVT-PWM converter with MIVSC controllerK (s) under load variations: (a)R = 200! 100, V = 5V and (b)R =100 ! 50, V = 5V; vertical: 50 mv/div, ac; horizontal: 5 ms/div.

Fig. 18. Output responses of ZVT-PWM converter with classical controllerK (s) under line voltage variationsV = 5! 4V: (a)R = 200 (b)R =100, and (c)R = 50; vertical: 50 mv/div, ac; horizontal: 50 ms/div.

Fig. 19. Output responses of ZVT-PWM converter with MIVSC controllerK (s) under line voltage variationsV = 5! 4V: (a)R = 200 (b)R =100 and (c)R = 50; vertical: 50 mv/div, ac; horizontal: 50 ms/div.

peak of the applied rampwave

14 V;

load resistance 50–200 ;input inductor 1 mH;output capacitor 470 F;resonant inductor 7 H;resonant capacitor 40 nF;snubber capacitor 2 nF;inductor series resistance0.5 ;

switching frequency 50 kHz;nominal duty ratio 0.5.A driving circuit is designed so that the main switch oper-

ates under zero-voltage switching. Fig. 12 is the implementedZVT-PWM boost converter with driving circuit and feedbackcontroller. Substituting the above parameters and

into (40), the dynamical equation of the ZVT-PWM boost con-verter is given by

(48)

(49)

where and. It follows directly from (46) that the nominal transfer func-

tion is given by

(50)

Page 9: Small-signal modeling and control of ZVT-PWM boost

10 IEEE TRANSACTIONS ON POWER ELECTRONICS, VOL. 18, NO. 1, JANUARY 2003

The experimental results measured by transfer functionanalyzer R9211b are compared with theoretical predictions inFig. 13. It is seen that the developed small-signal model accu-rately predicts the dynamical characteristics up to 1000 rad/s.

VIII. C ONTROLLERSDESIGN

In this section, two kinds of controllers will be designed forthe ZVT-PWM boost converter. They are classical and MIVSCcontrollers.

A. Classical Controller Design

The root-locus method is now applied to design a classicalcontroller. A pole at is added to eliminate the steady-stateerror and a zero at is added to improve phase margin.Moreover, by adding a zero at and a pole at

, the root loci are migrated to the left half of s-plane.As a result, a phase lead-lag controller is given by

(51)

B. MIVSC Controller Design

The MIVSC control scheme is applied to design a slidingmode controller for the ZVT-PWM boost converter in (50) withtwo poles at and . The eigenvaluesof the closed-loop system in Fig. 12 are selected at ,

and . By applying eigenstructureassignment technique, the hyperplanes vector is found to be

which yields with

sgn(52)

The switching function is defined as, where is the output state of the integrator.

To examine the performance of the closed-loop system withrespect to controllers and , variations of load re-sistance and line voltage are taken into consideration. Simula-tion results for ZVT-PWM boost converter with classical con-troller are shown in Fig. 14 with V. The linevoltage is initially set to 5 V and load is set to 200 at

s, then the load is changed to at s,and at s, then line voltage is changed to Vat s, and finally V at s.

Simulation results for ZVT-PWM boost converter withMIVSC controller are also shown in Fig. 15 with

V. The line voltage is initially set to 5 V andload is set to 200 at s, then the load is changedto at s, and at s,then line voltage is changed to V at s, andfinally V at s. It is interesting to note fromthe simulation results that the ZVT-PWM boost converter withMIVSC controller exhibits excellent output voltageresponse which is almost independent of the variations of loadresistance and line voltage.

Experimental results for load and line voltage variations areshown in Figs. 16–19 for the ZVT-PWM boost converter with

classical controller and MIVSC controller. It shows that theMIVSC controller still exhibits better performance than theclassical controller.

IX. CONCLUSION

A small-signal model of ZVT-PWM boost converters is de-veloped in this paper. The analysis shows that ZVT-PWM boostconverters exhibit better dynamical behavior than the conven-tional PWM boost converters. The mathematical model is veri-fied by experimental results. With the help of this model, a clas-sical controller and an MIVSC controller are then designed toregulate the output voltage and reject the line voltage distur-bance. Results from simulations and experiments show that theMIVSC controller is better than the classical controller in thepresence of variations of line voltage and load resistance.

REFERENCES

[1] G. Hua, E. X. Yang, Y. Jiang, and F. C. Lee, “Novel zero-voltage-transi-tion PWM converter,”IEEE Trans. Power Electron, vol. 9, pp. 213–219,Mar. 1994.

[2] B. P. Divakar and A. Ioinovici, “PWM converter with low stresses andzero capacitive turn on losses,”IEEE Trans. Aerosp. Electron. Syst., vol.33, pp. 913–920, July 1997.

[3] C. J. Tseng and C. L. Chen, “Novel ZVT-PWM converters with activesnubbers,”IEEE Trans. Power Electron., vol. 13, pp. 861–869, Sept.1998.

[4] J. Sun and H. Grotstollen, “Averaged modeling of switching powerconverters: reformulation and theoretical basis,” inProc. IEEE PESC,Toledo, Spain, 1992, pp. 1162–1172.

[5] J. L. Lin and W. J. Yang, “Modified IVSC stabilization and swing upcontrol of an inverted pendulum,” inProc. R.O.C. Automat. Contr. Conf.,Taipei, Taiwan, 1997, pp. 150–153.

[6] T. L. Chern and Y. C. Wu, “An optimal variable structure control withintegral compensation for electrohydraulic position servo control sys-tems,”IEEE Trans. Ind. Electron, vol. 39, pp. 460–463, Oct. 1992.

[7] M. E. Aggoune, F. Boudjemaa, A. Bensenouce, A. Hellal, M. R. El-emesai, and S. V. Vadari, “Design of variable structure voltage regulatorusing pole assignment technique,”IEEE Trans. Automat. Contr., vol. 39,pp. 2106–2111, Oct. 1994.

Jong-Lick Lin (M’99) was born in Tainan, Taiwan,R.O.C., in 1948. He received the B.S. and M.S.degrees in electrical engineering from NationalTaiwan University, Taipei, Taiwan, in 1973 and1977, respectively, and the Ph.D. degree from theEngineering Department, Leicester University,Leicester, U.K., in 1992.

He is a Professor in the Department of EngineeringScience, National Cheng Kung University, Tainan,Taiwan. His research interests include robust controltheory, singular system analysis, sliding mode

control, and dynamics analysis and controller design of power electronics.

Chin-Hua Chang was born in Taichung, Taiwan,R.O.C., in 1975. He received the B.S. degree inelectrical engineering from National Sun Yat-SenUniversity, Kaoshiung, Taiwan, in 1997, and theM.S. degree in engineering science from the NationalCheng Kung University, Tainan, Taiwan, in 1999.

His research interests include system control,sliding mode control, and power electronics systemdesign.