direct voltage control of dc–dc boost converters using enumeration-based model predictive control

10
Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected]. This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication. Direct Voltage Control of DC-DC Boost Converters Using Enumeration-Based Model Predictive Control Petros Karamanakos, Student Member, IEEE, Tobias Geyer, Senior Member, IEEE, and Stefanos Manias, Fellow, IEEE Abstract—This paper presents a model predictive control (MPC) approach for dc-dc boost converters. A discrete-time switched non-linear (hybrid) model of the converter is derived, which captures both the continuous and the discontinuous con- duction mode. The controller synthesis is achieved by formulating an objective function that is to be minimized subject to the model dynamics. The proposed MPC strategy, utilized as a voltage- mode controller, achieves regulation of the output voltage to its reference, without requiring a subsequent current control loop. Furthermore, a state estimation scheme is implemented that addresses load uncertainties and model mismatches. Simulation and experimental results are provided to demonstrate the merits of the proposed control methodology, which include a fast transient response and a high degree of robustness. Index Terms—DC-DC boost converter, model predictive con- trol (MPC), optimal control, voltage control, hybrid system. I. I NTRODUCTION O VER the past decades dc-dc conversion has matured into a ubiquitous technology, which is used in a wide variety of applications, including dc power supplies and dc motor drives [1]. Dc-dc converters are intrinsically difficult to control due to their switching behavior, constituting a (continuous-time) switched linear or hybrid system. To date, a plethora of control schemes has been proposed to address these difficulties. These control techniques range from linear techniques, such as proportional-integral (PI) controllers based on average models [2], [3] to fuzzy logic [4], [5], and from non-linear techniques [6], [7] and feedforward control [8], [9] to sliding mode [10], [11] and H methods [12]. Although existing control approaches have been shown to be reasonably effective, several challenges have not been fully addressed yet, such as ease of controller design and tuning, as well as robustness to load parameter variations. Furthermore, the computational power available today and the recent theoretical advances with regards to controlling hybrid systems allow one to tackle these problems in a novel way. The aim is not only to improve the performance of the closed-loop system, but to also enable a systematic design and implementation procedure. Model predictive control (MPC) is a particularly promising candidate to achieve this [13], [14], since it allows one to directly include constraints in the design phase and to address the switching or hybrid nature of dc-dc P. Karamanakos and S. Manias are with the Department of Electri- cal and Computer Engineering, National Technical University of Athens, 15780 Zografou, Athens, Greece (e-mail: [email protected]; ma- [email protected]). T. Geyer is with ABB Corporate Research, 5405 Baden-D¨ attwil, Switzer- land (e-mail: [email protected]). converters. MPC was developed in the 1970s in the process control industry, and has recently been introduced to the field of power electronics, including three-phase dc-ac and ac-dc systems [15]–[20], as well as dc-dc converters [21]–[30]. In MPC the control action is obtained by solving on- line at each time-step an optimization problem with a given objective function over a finite prediction horizon, subject to the discrete-time model of the system and constraints. The optimal sequence of control inputs is the one that minimizes the objective function. To provide feedback, allowing one to cope with model uncertainties and disturbances, only the first element of the sequence of control inputs is applied to the converter. At the next time-step, the optimization problem is repeated with updated measurements or estimates. This procedure is known as the receding horizon policy [14]. In this paper, MPC is employed as a voltage-mode controller for the dc-dc boost converter. The main control objective is the regulation of the output voltage to a commanded value, while rejecting the impact of variations in the input voltage and the load. The discrete-time model of the converter used by the controller is designed such that it accurately predicts the plant behavior both when operating in continuous (CCM) as well as in discontinuous conduction mode (DCM). As a result, the formulated controller is applicable to the whole operating regime, rather than just to a particular operating point. To address time-varying and unknown loads, a Kalman filter is added that estimates the converter states and provides offset- free tracking of the output voltage due to its integrating action, despite changes in the load [31]. In that way the robustness of the controller is ensured even when the converter operates under non-nominal conditions. The proposed scheme carries several benefits. The very fast dynamics achieved by MPC, combined with its inherent ro- bustness properties, are some of its key beneficial characteris- tics. Furthermore, thanks to the fact that the control objectives are expressed in the objective function in a straightforward manner, the design process is simple and laborious tuning is avoided. The inherent computational complexity is the most prominent drawback—the computational power required increases exponentially as the prediction horizon is extended. To address this issue, a move blocking strategy is adopted [32], which results in a significant reduction of the computations required and facilitates the real-time implementation of the controller. Finally, the absence of a modulator and the di- rect manipulation of the converter switches imply a variable switching frequency.

Upload: stefanos

Post on 08-Dec-2016

220 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Direct Voltage Control of DC–DC Boost Converters Using Enumeration-Based Model Predictive Control

Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected].

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication.

Direct Voltage Control of DC-DC Boost ConvertersUsing Enumeration-Based Model Predictive Control

Petros Karamanakos,Student Member, IEEE, Tobias Geyer,Senior Member, IEEE, andStefanos Manias,Fellow, IEEE

Abstract—This paper presents a model predictive control(MPC) approach for dc-dc boost converters. A discrete-timeswitched non-linear (hybrid) model of the converter is derived,which captures both the continuous and the discontinuous con-duction mode. The controller synthesis is achieved by formulatingan objective function that is to be minimized subject to the modeldynamics. The proposed MPC strategy, utilized as a voltage-mode controller, achieves regulation of the output voltagetoits reference, without requiring a subsequent current controlloop. Furthermore, a state estimation scheme is implemented thataddresses load uncertainties and model mismatches. Simulationand experimental results are provided to demonstrate the meritsof the proposed control methodology, which include a fasttransient response and a high degree of robustness.

Index Terms—DC-DC boost converter, model predictive con-trol (MPC), optimal control, voltage control, hybrid system.

I. I NTRODUCTION

OVER the past decades dc-dc conversion has maturedinto a ubiquitous technology, which is used in a wide

variety of applications, including dc power supplies and dcmotor drives [1]. Dc-dc converters are intrinsically difficultto control due to their switching behavior, constituting a(continuous-time) switched linear orhybrid system. To date,a plethora of control schemes has been proposed to addressthese difficulties. These control techniques range from lineartechniques, such as proportional-integral (PI) controllers basedon average models [2], [3] to fuzzy logic [4], [5], and fromnon-linear techniques [6], [7] and feedforward control [8], [9]to sliding mode [10], [11] andH∞ methods [12].

Although existing control approaches have been shown tobe reasonably effective, several challenges have not beenfully addressed yet, such as ease of controller design andtuning, as well as robustness to load parameter variations.Furthermore, the computational power available today andthe recent theoretical advances with regards to controllinghybrid systems allow one to tackle these problems in a novelway. The aim is not only to improve the performance of theclosed-loop system, but to also enable a systematic design andimplementation procedure. Model predictive control (MPC)isa particularly promising candidate to achieve this [13], [14],since it allows one to directly include constraints in the designphase and to address the switching or hybrid nature of dc-dc

P. Karamanakos and S. Manias are with the Department of Electri-cal and Computer Engineering, National Technical University of Athens,15780 Zografou, Athens, Greece (e-mail: [email protected]; [email protected]).

T. Geyer is with ABB Corporate Research, 5405 Baden-Dattwil, Switzer-land (e-mail: [email protected]).

converters. MPC was developed in the 1970s in the processcontrol industry, and has recently been introduced to the fieldof power electronics, including three-phase dc-ac and ac-dcsystems [15]–[20], as well as dc-dc converters [21]–[30].

In MPC the control action is obtained by solving on-line at each time-step an optimization problem with a givenobjective function over a finite prediction horizon, subject tothe discrete-time model of the system and constraints. Theoptimal sequence of control inputs is the one that minimizesthe objective function. To provide feedback, allowing one tocope with model uncertainties and disturbances, only the firstelement of the sequence of control inputs is applied to theconverter. At the next time-step, the optimization problemis repeated with updated measurements or estimates. Thisprocedure is known as thereceding horizon policy [14].

In this paper, MPC is employed as a voltage-mode controllerfor the dc-dc boost converter. The main control objective isthe regulation of the output voltage to a commanded value,while rejecting the impact of variations in the input voltageand the load. The discrete-time model of the converter used bythe controller is designed such that it accurately predictstheplant behavior both when operating in continuous (CCM) aswell as in discontinuous conduction mode (DCM). As a result,the formulated controller is applicable to the whole operatingregime, rather than just to a particular operating point. Toaddress time-varying and unknown loads, a Kalman filter isadded that estimates the converter states and provides offset-free tracking of the output voltage due to its integrating action,despite changes in the load [31]. In that way the robustnessof the controller is ensured even when the converter operatesunder non-nominal conditions.

The proposed scheme carries several benefits. The very fastdynamics achieved by MPC, combined with its inherent ro-bustness properties, are some of its key beneficial characteris-tics. Furthermore, thanks to the fact that the control objectivesare expressed in the objective function in a straightforwardmanner, the design process is simple and laborious tuningis avoided. The inherent computational complexity is themost prominent drawback—the computational power requiredincreases exponentially as the prediction horizon is extended.To address this issue, a move blocking strategy is adopted [32],which results in a significant reduction of the computationsrequired and facilitates the real-time implementation of thecontroller. Finally, the absence of a modulator and the di-rect manipulation of the converter switches imply a variableswitching frequency.

Page 2: Direct Voltage Control of DC–DC Boost Converters Using Enumeration-Based Model Predictive Control

Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected].

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication.

vs

iL RL L

S

D

CovCo

io

R vo

Fig. 1. Topology of the dc-dc boost converter.

This paper is organized as follows. In Section II thecontinuous-time model of the converter, suitable for both CCMand DCM, is presented. Furthermore, the discrete-time modelthat will be used as the prediction model is derived. Thecontrol problem is stated in Section III. In Section IV, theMPC problem is formulated and solved, using a move blockingscheme and enumeration, and a Kalman filter is added toaddress load variations. Section V presents simulation resultsillustrating the performance of the proposed control approach.In Section VI, the experimental validation of the introducedcontrol strategy is provided. The paper is summarized inSection VII, where conclusions are drawn.

II. M ODEL OF THEBOOST CONVERTER

A. Continuous-Time Model

The dc-dc boost converter shown in Fig. 1 is a converterthat increases the (typically uncontrolled) dc input voltagevs(t) to a higher (controlled) dc output voltagevo(t). The con-verter consists of two power semiconductors—the controllableswitch S, and the diodeD. The inductorL with the internalresistorRL is used to store and deliver energy depending onthe operating mode of the converter, while the filter capacitorCo is connected in parallel with the load resistorR so as toensure a constant output voltage during steady-state operationof the converter.

Three different linear dynamics are associated with theswitch positions. When the switchS is on (S = 1), energyis stored in the inductorL and the inductor currentiL(t)increases. When the switchS is off (S = 0), the inductor isconnected to the output and energy is released through it tothe load, resulting in a decreasingiL(t). Furthermore, whenthe switchS remainsoff and iL(t) = 0, then bothS andDare off ; the topology is reduced to the mesh formed by thecapacitorCo and the load. In this case, the converter operatesin DCM.

The state-space representation of the converter in thecontinuous-time domain is given by the following equa-tions [33]

dx(t)

dt=

(

A1 +A2u(t))

x(t) +Bvs(t) (1a)

y(t) = Cx(t) , (1b)

wherex(t) = [iL(t) vo(t)]

T (2)

is the state vector, encompassing the inductor current andthe output voltage across the output capacitor. The output

x(t) =

x(t) =

x(t) =

A1x(t)+

A1x(t)+

Bvs(t)

Bvs(t)

Bvs(t)(A1 +A2)x(t)+

daux = 1

daux = 1

daux = 0

u = 1

u = 1

u = 1

u = 0

u = 0

u = 0

u = 0

iL(t) = 0

iL(t) > 0

Fig. 2. Dc-dc converter presented as a continuous-time automaton.

y = vo(t) is given by the output voltage. The system matricesare

A1 =

[

− dauxRL

L− daux

Ldaux

Co

− 1

CoR

]

, A2 =

[

0 1

L

− 1

Co

0

]

,

B = [dauxL

0]T , and C = [0 1] .

The variableu denotes the switch position, withu = 1 imply-ing that the switchS is on, andu = 0 referring to the casewhere the switchS is off. Finally, daux is an auxiliary binaryvariable [34] that isdaux = 1 when the converter operates inCCM, i.e. eitheru = 1 or u = 0 and iL(t) > 0. When theconverter operates in DCM, i.e.u = 0 and iL(t) = 0, thendaux = 0 holds.

daux(t) =

1 if u(t) = 1, or u(t) = 0 and iL(t) > 00 if u(t) = 0 and iL(t) = 0

(3)

For a graphical summary, representing the boost converteras an automaton, see Fig. 2.

B. Discrete-Time Model

The derivation of an adequate model of the boost converterto serve as an internal prediction model for MPC is offundamental importance. As can be seen in Fig. 3, after thediscretization of the model in time, the converter can operatein four different modes, depending on the shape of the inductorcurrent. In order to precisely describe the operating modesofthe converter, the matricesΓ1 = A1 for daux = 1, Γ2 = A1 fordaux = 0, Γ3 = A2, and∆ = B for daux = 1 are introduced.Then, based on the continuous-time state-space model (1) andusing the forward Euler approximation approach, the followingdiscrete-time model of the converter is derived.

x(k + 1) =

E1x(k) + F1vs(k) Mode “1”E2x(k) + F2vs(k) Mode “2”E3x(k) + F3vs(k) Mode “3”E4x(k) Mode “4”

(4a)

y(k) = Gx(k) (4b)

where the matrices are E1 = 1+ (Γ1 + Γ3)Ts,E2 = 1+ Γ1Ts, E3 = 1

Ts

(τ1E2 + τ2E4), E4 = 1+ Γ2Ts,

Page 3: Direct Voltage Control of DC–DC Boost Converters Using Enumeration-Based Model Predictive Control

Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected].

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication.

Time Steps

iL

k k + 1 k + 2 k + 8 t

111 2 2 3 44

(a) Current.

Time Steps

u

k k + 1 k + 2 k + 8 t

111 2 2 3 44

(b) Switch position.

Fig. 3. Operation modes used in the mathematical model to describe theboost converter. Depending on the shape of the current four different modesare used.

x(k + 1) = x(k + 1) =

x(k + 1) =x(k + 1) =

F2vs(k)E1x(k)+ E2x(k)+F1vs(k)

E3x(k)+E4x(k)F3vs(k)

u = 1

u = 1

u = 1

u = 1

u = 0

u = 0

u = 0

u = 0

u = 0

iL(k) > 0

iL(k + 1) = 0

iL(k + 1) > 0

&

Fig. 4. Discrete-time mathematical model of the dc-dc converter representedas a discrete-time automaton.

F1 = ∆Ts, F2 = F1, F3 = ∆τ1, andG = C. Furthermore,τ1denotes the time-instant within the sampling interval, whenthe inductor current reaches zero, i.e.iL(k + τ1/Ts) = 0, andτ1 + τ2 = Ts. Finally, 1 is the identity matrix andTs is thesampling interval. Note thatE3 is derived by averaging overmodes “2” and “4”.

The four different operating modes of the converter’s math-ematical model are illustrated in Fig. 4. The transitions fromone mode to another are specified by conditions, such as theswitch position and the value of the current.

III. C ONTROL PROBLEM

For the dc-dc converter, the main control objective is forthe output voltage to accurately track its given reference—orequivalently to minimize the output voltage error—by appro-priately manipulating the switch. This is to be achieved despite

changes in the input voltage and load. During transients, theoutput voltage is to be regulated to its new reference value asfast and with as little overshoot as possible.

IV. M ODEL PREDICTIVE CONTROL

In this section an MPC scheme for dc-dc boost convertersis introduced, which directly controls the output voltage bymanipulating the switchS. Using an enumeration technique,the user-defined objective function is minimized subject totheconverter dynamics.

A. Objective Function

The objective function is chosen as

J(k) =

k+N−1∑

ℓ=k

(

|vo,err(ℓ+ 1|k)|+ λ|∆u(ℓ|k)|)

(5)

which penalizes the absolute values of the variables of concernover the prediction horizonN , which is of finite length. Thefirst term penalizes the absolute value of the output voltageerror

vo,err(k) = vo,ref − vo(k) . (6)

By penalizing the difference between two consecutive switch-ing states, the second term aims at decreasing the switchingfrequency and avoiding excessive switching

∆u(k) = u(k)− u(k − 1) . (7)

The weighting factorλ > 0 sets the trade-off between outputvoltage error and switching frequency,fsw. Note that thesampling intervalTs implicitly imposes an upper bound onthe switching frequency, i.e.fsw < 1/(2Ts). This value cor-responds to the case whenλ = 0, the output voltage is twicethe input voltage, i.e.vo = 2vs, and when the inductor is idealwith RL = 0.

B. Optimization Problem

The optimization problem underlying MPC at time-stepkamounts to minimizing the objective function (5) subject tothe converter model dynamics

U∗(k) = argmin J(k)subject to eq. (4).

(8)

The optimization variable is the sequence ofswitching states over the horizon, which isU(k) = [u(k) u(k + 1) . . . u(k +N − 1)]T . Minimizing (8)yields the optimal switching sequenceU∗(k). Out ofthis sequence, the first elementu∗(k) is applied to theconverter. The procedure is repeated atk + 1, based on newmeasurements acquired at the following sampling instance.

Minimizing (8) is a challenging task, since it is a mixed-integer non-linear optimization problem1. A straightforwardalternative is to solve (8) using enumeration, which involvesthe following three steps. First, by considering all possible

1It should be noted that the mathematical model of the converter givenby (4a) for modes “1” and “2” is affine (linear plus offset), and for mode “4”is linear, while the expression for mode “3” is non-linear.

Page 4: Direct Voltage Control of DC–DC Boost Converters Using Enumeration-Based Model Predictive Control

Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected].

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication.

Prediction steps

vo

k k + 3 k + 7 k + 8 k + 9 k + 10

Ts nsTs

(a)

Prediction steps

iL

k k + 3 k + 7 k + 8 k + 9 k + 10

Ts nsTs

(b)

Prediction steps

u

k k + 3 k + 7 k + 8 k + 9 k + 10

Ts nsTs

(c)

Fig. 5. Prediction horizon with move blocking: a) output voltage, b) inductorcurrent, and c) control input. The prediction horizon hasN = 10 time-steps,but the prediction interval is of length19Ts, sincens = 4 is used for thelastN2 = 3 steps.

combinations of the switching states (u = 0 or u = 1) over theprediction horizon, the set of admissible switching sequencesis assembled. For each of the2N sequences, the correspondingoutput voltage trajectory is predicted and the objective functionis evaluated. The optimal switching sequence is obtained bychoosing the one with the smallest associated cost.

C. Move Blocking

A fundamental difficulty associated with boost convertersarises when controlling their output voltage without an inter-mediate current control loop, since the output voltage exhibitsa non-minimum phase behavior with respect to the switchingaction. For example, when increasing the output voltage, theduty cycle of switchS has to be ramped up, but initially theoutput voltage drops before increasing. This implies that thesign of the gain (from the duty cycle to the output voltage) isnot always positive.

To overcome this obstacle and to ensure closed-loop sta-bility, a sufficiently long prediction intervalNTs is required,so that the controller can “see” beyond the initial voltagedrop when contemplating to increase the duty cycle. On theone hand, increasingN leads to an exponential increase inthe number of switching sequences to be considered andthus dramatically increases the number of calculations needed.On the other hand, long sampling intervalsTs reduce theresolution of the possible switching instants, since switchingcan only be performed at the sampling instants.

Prediction steps

v o

Past

k k + 4 k + 8 k + 12 k + 16 k + 20

(a)

Prediction steps

v o

Past

k k + 4 k + 7 k + 8 k + 9 k + 10 k + 11

(b)

Fig. 6. Effect of the move blocking scheme. In (a), without move blocking, aprediction horizon ofN = 20 steps of equal time-intervals is needed. In (b),with the move blocking strategy employed, anN = 11 prediction horizonis sufficient to achieve the same closed-loop result (N1 = 7, N2 = 4, andns = 4, total length23Ts).

A long prediction intervalNTs with a small N and asmall Ts can be achieved by employing amove blockingtechnique [32]. For the first steps in the prediction horizon,the prediction model is sampled withTs, while for steps far inthe future, the model is sampled more coarsely with a multipleof Ts, i.e. nsTs, with ns ∈ N

+ [35]. As a result, differentsampling intervals are used within the prediction horizon,asillustrated in Fig. 5. We useN1 to denote the number ofprediction steps in the first part of the horizon, which aresampled withTs. Accordingly, N2 refers to the number ofsteps in the last part of the horizon, sampled withnsTs. Thetotal number of time-steps in the horizon isN = N1 +N2.

An illustrative example of the effectiveness of the moveblocking strategy is depicted in Fig 6. Assume that at timeinstantkTs the output voltage reference increases in a stepwisemanner and the output voltage is to follow that change.However, as mentioned above, because of the non-minimumphase nature of the system, the output voltage initially tendsto decrease. In order to ensure that MPC is able to predictthe final voltage increase and will thus pick the correspond-ing switching sequence that achieves this, in this example,a prediction interval of twenty time-steps is required, i.e.NTs = 20.

By employing the move blocking scheme, the eleven-stephorizonN = 11, with N1 = 7, N2 = 4, andns = 4 suffices,resulting in a prediction interval of a 23 time-steps. In this way,the computational cost is significantly reduced. Without moveblocking, the number of switching sequences to be examinedis 220 = 1048576, and the state evolution has to be predictedfor 20 steps into the future. In contrast to this, when usingthe move blocking scheme, the total number of sequences

Page 5: Direct Voltage Control of DC–DC Boost Converters Using Enumeration-Based Model Predictive Control

Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected].

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication.

is 211 = 2048, and the evolution of the state needs to becalculated only for11 steps. As a result, the computationsrequired are decreased by three orders of magnitude, or99.9%.

It is important to point out that a high timing resolution isrequired only around the current time-step and the very nearfuture. Further ahead, a rough timing resolution suffices, dueto the receding horizon policy. The coarse plan of the secondpart of the prediction horizon is step by step shifted towardsthe beginning of the prediction horizon and simultaneouslyrefined.

D. Load Uncertainty

In most applications the load is unknown and time varying.Thus, an external estimation loop should be added, whichallows the elimination of the output voltage error in the pres-ence of load uncertainties. This additional loop is employedto provide state estimates to the previously derived optimalcontroller, where the load was assumed to be known andconstant. The output voltage reference will be adjusted so asto compensate for the deviation of the output voltage from itsactual reference.

To achieve this, a discrete-time Kalman filter [36] is de-signed similar to [25]; thanks to its integrating nature theKalman filter provides a zero steady-state output voltage error.Two integrating disturbance states,ie andve, are introduced inorder to model the effect of the load variations on the inductorcurrent and output voltage, respectively. The measured statevariables, iL and vo, together with the disturbance statevariables form the augmented state vector

xa = [iL vo ie ve]T . (9)

The Kalman filter is used to estimate the state vector givenby (9). Depending on the operating mode of the converter,as shown in Fig. 3, four different affine systems result.The respective stochastic discrete-time state equations of theaugmented model are

xa(k + 1) = Ezaxa(k) + Fzavs(k) + ξ(k) , (10)

wherez = 1, 2, 3, 4 corresponds to the four operating modesof the converter.

The measured state vector is given by

x(k) =

[

iL(k)vo(k)

]

= Gaxa(k) + ν(k) (11)

and the matrices are

Eza =

[

Ez 0

0 1

]

, F1a =

F1

00

= F2a =

F2

00

, F3a =

F3

00

,

F4a =[

0 0 0 0]T

, and Ga =[

1 1]

,(12)

where,1 is the identity matrix of dimension two and0 aresquare zero matrices of dimension two. The variablesξ ∈ R

4

and ν ∈ R2 denote the process and the measurement noise,

respectively. These terms represent zero-mean, white Gaussiannoise sequences with normal probability distributions. Their

covariances are given byE[ξξT ] = Q andE[ννT ] = R, andare positive semi-definite and positive definite, respectively.

A switched discrete-time Kalman filter is designed based onthe augmented model of the converter. The active mode of theKalman filter (one out of four) is determined by the switchingposition and the operating mode of the converter.

Due to the fact that the state-update for each operating modeis different, four Kalman gainsKz need to be calculated.Consequently, the equation for the estimated statexa(k) is

xa(k+1) = Ezaxa(k) +KzGa

(

xa(k)− xa(k))

+Fzavs(k) .(13)

The noise covariance matricesQ and R are chosen suchthat high credibility is assigned to the measurements of thephysical states (iL andvo), whilst low credibility is assignedto the dynamics of the disturbance states (ie and ve). TheKalman gains are calculated based on these matrices. Theestimated disturbances, provided by the resulting filter, canbe used to remove their influence from the output voltage.Hence, the disturbance stateve is used to adjust the outputvoltage referencevo,ref

vo,ref = vo,ref − ve . (14)

To this end, the estimated states,iL and vo, are used asinputs to the controller, instead of the measured states,iL andvo.

E. Control Algorithm

The proposed control concept is summarized in Algo-rithm 1. The functionf stands for the state-update given

Algorithm 1 MPC algorithm

function u∗(k) = MPC (x(k), u(k − 1))J∗(k) = ∞; u∗(k) = ∅; x(k) = x(k)for all U overN do

J = 0for ℓ = k to k +N − 1 do

if ℓ < k +N1 thenx(ℓ + 1) = f1(x(ℓ), u(ℓ))

elsex(ℓ + 1) = f2(x(ℓ), u(ℓ))

end ifvo,err(ℓ+ 1) = vo,ref − vo(ℓ+ 1)∆u(ℓ) = u(ℓ)− u(ℓ− 1)J = J + |vo,err(ℓ+ 1)|+ λ|∆u(ℓ)|

end forif J < J∗(k) then

J∗(k) = J , u∗(k) = U(1)end if

end forend function

by (4), with the subscripts1 and 2 corresponding to thesampling interval being used, i.e.Ts andnsTs, respectively.Fig. 7 depicts the flowchart of the introduced MPC algorithm,

Page 6: Direct Voltage Control of DC–DC Boost Converters Using Enumeration-Based Model Predictive Control

Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected].

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication.

Predict evolution ofx, ∆u andvo,err

J i ≤ J∗?Yes

Yes

J∗ = J i, U∗ = U i

i = 1

i = i+ 1

Build all switchingsequencesU overN .

based on moveblocking scheme.

Evaluate objectivefunctionJ i.

i ≤ 2N?No Stop!

Outputu∗(k) = U∗(1)

Fig. 7. Flowchart of the MPC algorithm.

DC/DCBoost Converter

KalmanFilter

MPCAlgorithm 1

vo

iL

vs

iLvove

vo,ref

vo,ref

u(k − 1)SwitchingSequences

S

Fig. 8. Block diagram of the MPC scheme and Kalman filter.

while the block diagram of the entire control scheme is shownin Fig. 8.

V. SIMULATION RESULTS

In this section simulation results are presented to demon-strate the performance of the proposed controller under severaloperating conditions. Specifically, the closed-loop converterbehavior is examined in both CCM and DCM. The dynamicperformance is investigated during start-up. Moreover, theresponses of the output voltage to step changes in the com-manded voltage reference, the input voltage and the load areillustrated.

The circuit parameters areL = 450µH, RL = 0.3Ω andCo = 220µF. The nominal load resistance isR = 73Ω. Ifnot otherwise stated, the input voltage isvs = 10V and thereference of the output voltage isvo,ref = 15V.

Time [ms]

v o[V

]

0 1 2 3 40

5

10

15

20

(a)

Time [ms]

i L[A

]

0 1 2 3 40

1

2

3

4

(b)

Fig. 9. Simulation results for nominal start-up: a) output voltage (solid line)and output voltage reference (dashed line), b) inductor current.

The weight in the objective function isλ = 0.1, the pre-diction horizon is N = 14 and the sampling interval isTs = 2.5µs. A move blocking scheme is used withN1 = 8,N2 = 6 andns = 4, i.e. the sampling interval for each of thelast six steps in the prediction interval isTs = 10µs2. Finally,the covariance matrices of the Kalman filter are chosen asQ = diag(0.1, 0.1, 50, 50) andR = diag(1, 1).

A. Start-Up

The first case to be examined is that of the start-up behaviorunder nominal conditions. As can be seen in Fig. 9, theinductor current is very quickly increased until the capacitoris charged to the desired voltage level. The output voltagereaches its reference value in aboutt ≈ 1.8ms, without anynoticeable overshoot. Subsequently, the converter operates inDCM with the inductor current reaching zero.

B. Step Changes in the Output Reference Voltage

Next, step changes in the reference of the output voltageare considered. First, a step-up change in the output referencevoltage is examined: at timet = 2ms the reference is doubledfrom vo,ref = 15V to vo,ref = 30V. As can be seen in Fig. 10,the controller increases the current temporarily in order toquickly ramp up the output voltage. Note that this favorablechoice is made by the controller thanks to its long predictionhorizon anddespite the non-minimum phase behavior of the

2The length of the prediction horizon in time should be as longas possible.A horizon of about80 µs is sufficient. The first part of the prediction horizonshould be finely sampled, since switching is possible only atthe samplinginstants. As such, the sampling intervalTs should be as small as possible.The number of steps in the prediction horizonN = N1 +N2 determines thecomputational complexity. To ensure that the control law can be computedwithin Ts, N should be relatively small, leading to the choice made above.

Page 7: Direct Voltage Control of DC–DC Boost Converters Using Enumeration-Based Model Predictive Control

Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected].

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication.

Time [ms]

v o[V

]

0 1 2 3 4 5 610

15

20

25

30

35

(a)

Time [ms]

i L[A

]

0 1 2 3 4 5 60

0.5

1

1.5

2

2.5

3

3.5

(b)

Fig. 10. Simulation results for a step-up change in the output voltagereference: a) output voltage (solid line) and output voltage reference (dashedline), b) inductor current.

converter. Once the output voltage has reached its reference,the inductor current is decreased to the level that correspondsto the steady-state power balance. The controller exhibitsan excellent behavior during the transient, reaching the newoutput voltage in aboutt ≈ 1.8ms, without any overshoot.

Furthermore, the behavior of the controller is tested un-der a step-down change in the output reference voltage. Attime t = 2ms, the output voltage reference changes fromvo,ref = 20V to vo,ref = 15V; the segment of interest is de-picted in Fig. 11. Since the proposed MPC strategy is formu-lated as a voltage-mode controller effort is put into decreasingthe voltage to its new desired level as quickly as possible.To do so, the controllable switch is turned off, the currentinstantaneously reaches zero, and the capacitor dischargesthrough the load until it reaches its new demanded value inaboutt ≈ 1.2ms.

C. Step Change in the Input Voltage

Operating at the steady-state operating point correspondingto vo,ref = 30V, the input voltage is changed in a step-wisefashion. At time t = 0.4ms the input voltage is increasedfrom vs = 10V to vs = 15V. The transient response of theconverter is depicted in Fig. 12. The output voltage remainspractically unaffected, with no undershoot observed, while thecontroller settles very quickly at the new steady-state operatingpoint.

D. Load Step Change

The last case examined is that of a drop in the loadresistance. As can be seen in Fig. 13, a step-down change inthe load fromR = 73Ω to R = 36.5Ω occurs att = 1ms (the

Time [ms]

v o[V

]

0 1 2 3 4 5 610

12.5

15

17.5

20

22.5

25

(a)

Time [ms]

i L[A

]

0 1 2 3 4 5 60

0.5

1

1.5

(b)

Fig. 11. Simulation results for a step-down change in the output voltagereference: a) output voltage (solid line) and output voltage reference (dashedline), b) inductor current.

Time [ms]

v o[V

]

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.829

29.5

30

30.5

31

(a)

Time [ms]

i L[A

]

0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.80

0.5

1

1.5

2

(b)

Fig. 12. Simulation results for a step-up change in the inputvoltage: a) outputvoltage (solid line) and output voltage reference (dashed line), b) inductorcurrent.

input voltage isvs = 15V, and the output voltage reference isvo,ref = 30V). The Kalman filter adjusts the output voltagereference to its new value so as to avoid any steady-statetracking error. This can be observed in Fig. 13(a); after theconverter has settled at the new operating point, the outputvoltage accurately follows its reference.

Page 8: Direct Voltage Control of DC–DC Boost Converters Using Enumeration-Based Model Predictive Control

Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected].

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication.

Time [ms]

v o[V

]

0 1 2 3 4 5 629

29.5

30

30.5

31

(a)

Time [ms]

i L[A

]

0 1 2 3 4 5 60

0.5

1

1.5

2

2.5

3

(b)

Fig. 13. Simulation results for a step-down change in the load: a) outputvoltage (solid line) and output voltage reference (dashed line), b) inductorcurrent.

VI. EXPERIMENTAL VALIDATION

To further investigate the potential advantages of the pro-posed algorithm, the controller was implemented on a dSpaceDS1104 real-time system. A boost converter was built using anIRF60 MOSFET and a MUR840 diode as active and passiveswitches, respectively. The values of the circuit elementsarethe same as in Section V. Moreover, the nominal input andoutput voltages and the nominal load resistance are the sameas previously. The voltage and current measurements wereobtained using Hall effect transducers.

Due to computational restrictions imposed by the com-putational platform, a six-step prediction horizon was im-plemented, i.e.N = 6 and the sampling interval was set toTs = 10µs. The prediction horizon was split intoN1 = 4 andN2 = 2 with ns = 2. The weight in the objective function waschosen asλ = 0.5. The covariance matrices of the Kalmanfilter are the same as previously.

A. Start-up

In Fig. 14 the output voltage and the inductor current of theconverter are depicted during start-up. The inductor currentrapidly increases to charge the output capacitor to the referencevoltage level as fast as possible. The output voltage reaches itsdesired value in aboutt ≈ 1.8ms. Subsequently, the inductorcurrent reaches its nominal value and the converter operatesin DCM.

B. Step Changes in the Output Reference Voltage

The second case to be analyzed is that of the transientbehavior during step changes in the output reference volt-age. A step-up change in the output reference voltage from

Time [ms]

v o[V

]

0 1 2 3 40

5

10

15

20

(a)

Time [ms]

i L[A

]

0 1 2 3 40

0.5

1

1.5

2

2.5

3

(b)

Fig. 14. Experimental results for nominal start-up: a) output voltage, and b)inductor current.

Time [ms]

v o[V

]

0 1 2 3 4 5 610

15

20

25

30

35

(a)

Time [ms]

i L[A

]

0 1 2 3 4 5 60

0.5

1

1.5

2

2.5

3

3.5

(b)

Fig. 15. Experimental results for a step-up change in the output voltagereference: a) output voltage, and b) inductor current.

vo,ref = 15V to vo,ref = 30V occurs at t ≈ 1.7ms. The re-sponse of the converter is illustrated in Fig. 15. The inductorcurrent instantaneously increases, enabling the output voltageto reach its new desired level as fast as possible. This happensin aboutt ≈ 1.9ms, without a significant overshoot. Moreover,a step-down change, illustrated in Fig. 16, is investigated.The output reference voltage changes fromvo,ref = 20V tovo,ref = 15V at t ≈ 1.9ms. As can be seen, the controller

Page 9: Direct Voltage Control of DC–DC Boost Converters Using Enumeration-Based Model Predictive Control

Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected].

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication.

Time [ms]

v o[V

]

0 1 2 3 4 5 610

12.5

15

17.5

20

22.5

25

(a)

Time [ms]

i L[A

]

0 1 2 3 4 5 60

0.5

1

1.5

(b)

Fig. 16. Experimental results for a step-down change in the output voltagereference: a) output voltage, and b) inductor current.

exhibits a favorable performance; the inductor current is in-stantly reduced to zero so as to allow the capacitor to dischargethrough the resistor, and the converter reaches the new steady-state operating point in aboutt ≈ 1.2ms.

C. Ramp Change in the Input Voltage

Subsequently, the input voltage is manually increased fromvs = 10V to vs = 15V (the output reference voltage isvo,ref = 30V), resulting in a voltage ramp fromt ≈ 16msuntil t ≈ 38ms, see Fig. 17. During the transient, the inductorcurrent changes accordingly in a ramp-like manner down to itsnew steady-state value. It can be seen that the output voltageremains unaffected and is kept equal to its reference value,implying that input voltage disturbances are very effectivelyrejected by the controller and the Kalman filter.

D. Load Step Change

The last case examined is that of a step-down change inthe load resistance occurring att ≈ 1.2ms. With the converteroperating at the previously attained operating point, the loadresistance is halved, i.e. fromR = 73Ω toR = 36.5Ω. As canbe observed in Fig. 18, the Kalman filter quickly adjusts thevoltage reference accordingly, resulting in a zero steady-stateerror in the output voltage, thanks to its integrating nature.

VII. C ONCLUSION

A model predictive control approach based on enumerationfor dc-dc boost converter is proposed that directly regulatesthe output voltage along its reference, without the use ofan underlying current control loop. This enables very fastdynamics during transients. Since the converter model isincluded in the controller, the time-consuming tuning of

Time [ms]

v s[V

]

0 10 20 30 40 50 608

10

12

14

16

18

(a)

Time [ms]

v o[V

]

0 10 20 30 40 50 6028

29

30

31

32

(b)

Time [ms]

i L[A

]

0 10 20 30 40 50 600

0.5

1

1.5

2

(c)

Fig. 17. Experimental results for a ramp change in the input voltage: a)input voltage, b) output voltage, and c) inductor current.

controller gains is avoided. The computational complexityis somewhat pronounced, but kept at bay by using a moveblocking scheme. In addition to that, the switching frequencyis variable. A load estimation scheme, namely a discrete-time switched Kalman filter, is implemented to address loadvariations and to ensure robustness to parameter variations.Simulation and experimental results demonstrate the potentialadvantages of the proposed methodology.

REFERENCES

[1] N. Mohan, T. M. Undeland, and W. P. Robbins,Power Electronics:Converters, Applications and Design, 3rd ed. Hoboken, NJ: Wiley,2003.

[2] R. W. Erickson, S.Cuk, and R. D. Middlebrook, “Large-signal modellingand analysis of switching regulators,” inProc. IEEE Power Electron.Spec. Conf., Cambridge, MA, Jun. 1982, pp. 240–250.

[3] J. Alvares-Ramırez, I. Cervantes, G. Espinosa-Perez, P. Maya, andA. Morales, “A stable design of PI control for dc-dc converters withan RHS zero,”IEEE Trans. Circuits Syst. I, vol. 48, no. 1, pp. 103–106,Jan. 2001.

[4] T. Gupta, R. R. Boudreaux, R. M. Nelms, and J. Y. Hung, “Implementa-tion of a fuzzy controller for dc-dc converters using an inexpensive 8-bmicrocontroller,”IEEE Trans. Ind. Electron., vol. 44, no. 5, pp. 661–669,Oct. 1997.

Page 10: Direct Voltage Control of DC–DC Boost Converters Using Enumeration-Based Model Predictive Control

Copyright (c) 2013 IEEE. Personal use is permitted. For any other purposes, permission must be obtained from the IEEE by emailing [email protected].

This article has been accepted for publication in a future issue of this journal, but has not been fully edited. Content may change prior to final publication.

Time [ms]

v o[V

]

0 1 2 3 4 5 626

28

30

32

34

(a)

Time [ms]

i L[A

]

0 1 2 3 4 5 60

0.5

1

1.5

2

2.5

3

(b)

Fig. 18. Experimental results for a step-down change in the load: a) outputvoltage, and b) inductor current.

[5] P. Mattavelli, L. Rossetto, G. Spiazzi, and P. Tenti, “General-purposefuzzy controller for dc-dc converters,”IEEE Trans. Power Electron.,vol. 12, no. 1, pp. 79–86, Jan. 1997.

[6] H. J. Sira-Ramırez, “Nonlinear P-I controller design for switchmode dc-to-dc power converters,”IEEE Trans. Circuits Syst. I, vol. 38, no. 4, pp.410–417, Apr. 1991.

[7] S. Hiti and D. Borojevic, “Robust nonlinear control forthe boostconverter,” IEEE Trans. Power Electron., vol. 10, no. 6, pp. 651–658,Nov. 1995.

[8] M. K. Kazimierczuk and A. Massarini, “Feedforward control dynamicof dc/dc PWM boost converter,”IEEE Trans. Circuits Syst. I, vol. 44,no. 2, pp. 143–149, Feb. 1997.

[9] M. K. Kazimierczuk and L. A. Starman, “Dynamic performance ofPWM dc/dc boost converter with input voltage feedforward control,”IEEE Trans. Circuits Syst. I, vol. 46, no. 12, pp. 1473–1481, Dec. 1999.

[10] P. Mattavelli, L. Rossetto, and G. Spiazzi, “Small-signal analysis of dc-dc converters with sliding mode control,”IEEE Trans. Power Electron.,vol. 12, no. 1, pp. 96–102, Jan. 1997.

[11] S.-C. Tan, Y. M. Lai, and C. K. Tse, “General design issues of sliding-mode controllers in dc-dc converters,”IEEE Trans. Ind. Electron.,vol. 55, no. 3, pp. 1160–1174, Mar. 2008.

[12] A. Kugi and K. Schlacher, “NonlinearH∞ controller design for a dc-to-dc power converter,”IEEE Trans. Contr. Syst. Technol., vol. 7, no. 2,pp. 230–237, Mar. 1999.

[13] D. Q. Mayne, J. B. Rawlings, C. V. Rao, and P. O. M. Scokaert, “Con-strained model predictive control: Stability and optimality,” Automatica,vol. 36, no. 6, pp. 789–814, Jun. 2000.

[14] J. B. Rawlings and D. Q. Mayne,Model Predictive Control: Theory andDesign. Madison, WI: Nob Hill, 2009.

[15] T. Geyer, “Low complexity model predictive control in power electronicsand power systems,” Ph.D. dissertation, Autom. Control Lab. ETHZurich, Zurich, Switzerland, 2005.

[16] P. Correa, M. Pacas, and J. Rodrıguez, “A predictive torque control forinverter-fed induction machines,”IEEE Trans. Ind. Electron., vol. 54,no. 2, pp. 1073–1079, Apr. 2007.

[17] T. Geyer, G. Papafotiou, and M. Morari, “Model predictive direct torquecontrol—Part I: Concept, algorithm and analysis,”IEEE Trans. Ind.Electron., vol. 56, no. 6, pp. 1894–1905, Jun. 2009.

[18] P. Cortes, M. P. Kazmierkowski, R. M. Kennel, D. E. Quevedo, andJ. Rodrıguez, “Predictive control in power electronics and drives,” IEEETrans. Ind. Electron., vol. 55, no. 12, pp. 4312–4324, Dec. 2008.

[19] P. Cortes, J. Rodrıguez, P. Antoniewicz, and M. Kazmierkowski, “Directpower control of an AFE using predictive control,”IEEE Trans. PowerElectron., vol. 23, no. 5, pp. 2516–2523, Sep. 2008.

[20] A. Nami, F. Zare, A. Ghosh, and F. Blaabjerg, “A hybrid cascadeconverter topology with series-connected symmetrical andasymmetricaldiode-clamped H-bridge cells,”IEEE Trans. Power Electron., vol. 26,no. 1, pp. 51–65, Jan. 2011.

[21] S. Bibian and H. Jin, “High performance predictive dead-beat digitalcontroller for dc power supplies,”IEEE Trans. Power Electron., vol. 17,no. 3, pp. 420–427, May 2002.

[22] J. Chen, A. Prodic, R. W. Erickson, and D. Maksimovic,“Predictivedigital current programmed control,”IEEE Trans. Power Electron.,vol. 18, no. 1, pp. 411–419, Jan. 2003.

[23] F. M. Oettmeier, J. Neely, S. Pekarek, R. DeCarlo, and K.Uthaichana,“MPC of switching in a boost converter using a hybrid state model witha sliding mode observer,”IEEE Trans. Ind. Electron., vol. 56, no. 9, pp.3453–3466, Sep. 2009.

[24] T. Geyer, G. Papafotiou, and M. Morari, “On the optimal control ofswitch-mode dc-dc converters,” inHybrid Syst.: Comput. and Control,ser. LNCS, R. Alur and G. J. Pappas, Eds. Berlin, Germany: Springer-Verlag, 2004, vol. 2993, pp. 342–356.

[25] T. Geyer, G. Papafotiou, R. Frasca, and M. Morari, “Constrained optimalcontrol of the step-down dc-dc converter,”IEEE Trans. Power Electron.,vol. 23, no. 5, pp. 2454–2464, Sep. 2008.

[26] A. G. Beccuti, G. Papafotiou, R. Frasca, and M. Morari, “Explicit hybridmodel predictive control of the dc-dc boost converter,” inProc. IEEEPower Electron. Spec. Conf., Orlando, FL, Jun. 2007, pp. 2503–2509.

[27] A. G. Beccuti, S. Mariethoz, S. Cliquennois, S. Wang, and M. Morari,“Explicit model predictive control of dc-dc switched-modepower sup-plies with extended Kalman filtering,”IEEE Trans. Ind. Electron.,vol. 56, no. 6, pp. 1864–1874, Jun. 2009.

[28] Y. Xie, R. Ghaemi, J. Sun, and J. S. Freudenberg, “Implicit modelpredictive control of a full bridge dc-dc converter,”IEEE Trans. PowerElectron., vol. 24, no. 12, pp. 2704–2713, Dec. 2009.

[29] P. Karamanakos, G. Papafotiou, and S. Manias, “Model predictivecontrol strategies for dc-dc boost voltage conversion,” inProc. Eur.Power Electron. Conf., Birmingham, UK, Aug./Sep. 2011, pp. 1–9.

[30] P. Karamanakos, T. Geyer, and S. Manias, “Direct model predictivecurrent control of dc-dc boost converters,” inProc. Int. Power Electron.and Motion Control Conf. and Expo., Novi Sad, Serbia, Sep. 2012, pp.DS2c.11–1–DS2c.11–8.

[31] T. Geyer, G. Papafotiou, and M. Morari, “Hybrid model predictivecontrol of the step-down dc-dc converter,”IEEE Trans. Contr. Syst.Technol., vol. 16, no. 6, pp. 1112–1124, Nov. 2008.

[32] R. Cagienard, P. Grieder, E. C. Kerrigan, and M. Morari,“Move blockingstrategies in receding horizon control,”J. of Process Control, vol. 17,no. 6, pp. 563–570, Jul. 2007.

[33] R. W. Erickson and D. Maksimovic,Fundamentals of Power Electronics,2nd ed. Norwell, MA: Kluwer Academic, 2001.

[34] A. Bemporad and M. Morari, “Control of systems integrating logic,dynamics and constraints,”Automatica, vol. 35, no. 3, pp. 407–427,Mar. 1999.

[35] T. Geyer, G. Papafotiou, and M. Morari, “Model predictive control inpower electronics: A hybrid systems approach,” inProc. IEEE Conf.Decis. Control, Seville, Spain, Dec. 2005, pp. 5606–5611.

[36] G. Pannocchia and J. B. Rawlings, “Disturbance models for offset-freemodel-predictive control,”AIChE J., vol. 49, no. 2, pp. 426–437, Feb.2003.