Vous êtes sur la page 1sur 6

Lagrangian Modeling and Control of Switching Networks with

Integrated Coupled Magnetics


Jacquelien M.A. Scherpen, Dimitri Jeltsema, J. Ben Klaassens
Fac. ITS, Dept. of Electrical Eng., Delft University of Technology,
P.O. Box 5031, 2600 GA Delft, The Netherlands.
E-mail: j.m.a.scherpen/d.jeltsema/j.b.klaassens@its.tudelft.nl
Phone: +31-15-278/-6152/-4516/-2928, Fax: +31-15-278-6679

Abstract - In this paper a method is presented to build an Euler-Lagrange As shown in [9, 11], the non-minimum phase character of the
model for electrical networks, including switches and integrated (non-ideal) converters above induces several difficulties for the applica-
coupled-magnetics, in a structured general way. One of the advantages of
emphasizing the physical structure of these systems is its functionality during tion of passivity-based control techniques. For this reason, the
the controller design stage. In case a switching network contains coupled- output is controlled indirectly through the regulation of the in-
inductor structures, an additional path for the energy transfer is introduced. put current. We can show that there exists a range of cou-
For this reason, a basic building block is proposed that describes the dynami- pling ‘mismatch’ adjustments for which all currents and volt-
cal behaviour of a pair of magnetically coupled-inductors. This building block
is applicable to all types of switching converters, and easily predicts the ex- ages show unstable zero-dynamics. This is mainly caused by
istence of reduced or zero-ripple current. The switches make the dynamic the additional magnetic coupling between the loops in the net-
models nonlinear. For using the Lagrangian structure for controller design, work. In this case it is not possible to design a globally stable
the zero-dynamics for such switching network had to be studied. It is shown
controller. In addition we show that for exact matching con-
that under certain coupling conditions it will not be possible to design a glob-
ally stable controller. The approach is illustrated by means of the coupled- ditions (zero-current ripple) the order of the zero-dynamics is
inductor Ćuk converter with zero-output ripple, in closed loop with an adap- reduced.
tive passivity-based controller.
In Section 2 we present the general procedure to develop an EL
1 Introduction model for switching and non-switching electrical networks.
Then, in Section 3 we show how to include coupled-magnetics
During the last thirty years major research has been devoted
into the EL model. Section 4 treats a study of the stability
to optimization of modeling, design and control techniques of
of the zero-dynamics, corresponding to the (non-) minimum
DC-to-DC switched-mode power converters. In recent devel-
phase behaviour of the output in relation to the coupling ad-
opments these systems are considered from a physical mod-
justments. In Section 5 an adaptive passivity-based controller
eling point of view. It is shown in [9, 11] that the conven-
for the Ćuk converter with zero-output ripple is developed.
tional average PWM (pulse width modulation) models of the
Simulations are presented. Finally, in Section 6 we end with
classical Buck, Boost, Buck-Boost and the more complex Ćuk
the conclusions and recommendations.
converter correspond to systems derived from classical Euler-
Lagrange (or Hamiltonian, [2]) dynamic considerations. The 2 Euler-Lagrange Modeling of Electrical (Switching)
approach consists of establishing a suitable set of average Networks
Euler-Lagrange (EL) parameters modulated by the duty ratio
The constraint EL dynamics of an electrical circuit Σe can in
function. Here, we develop a procedure that results in the EL
general be characterized as follows [11]:
parameters of a general switching electrical network structure,  
where we assume switches to be ON and OFF, and where the d ∂L ∂L ∂D
− =− + A(q)λ + Fq
EL parameters are extended with constraint equations stem- Σe : dt ∂ q̇ ∂q ∂ q̇ (1)
ming from Kirchhoff’s current laws. The major advantage of 
A(q) q̇ = 0
emphasizing the physical structure of the systems is its func-
tionality during the controller design stage. This will finally where q̇ is the vector of flowing current and q represents the
electric charge. The vector q constitutes the generalized coor-
lead to feedback (passivity-based) controlled systems that do
not involve cancellation of non-linearities and are globally de- dinates describing the circuit and is assumed to have N com-
fined, see e.g. [6, 10]. ponents, represented by q1 , . . . , qN . It is well-known that the
scalar function L is the Lagrangian of the system, defined as
The Lagrangian approach for switching networks as initiated the difference between the magnetic co-energy of the circuit,
in [9, 11] was up to now, build on the assumption that the con- denoted by T (q, q̇), and the electric field energy of the circuit,
verters, and electrical circuits in general, do not contain any denoted by V (q), i.e., L(q, q̇) = T (q, q̇) − V(q). The func-
magnetic couplings between its different loops. Here, this as- tion D(q̇) is the Rayleigh dissipation cofunction of the system.
sumption is no longer necessary, since we include essential The vector Fq = [Fq1 , . . . , FqN ] represents the ordered com-
terms that correspond to the magnetic energy of the coupling ponents of the set of generalized forcing functions, or volt-
into the Lagrangian description. age sources, associated with each generalized coordinate. The
constraints that follow from Kirchhoff’s current laws are cap- The above procedure, as well as the developments in e.g. [9],
tured in A(q). Following ([6]), we refer to the set of functions was initially build upon the assumption that the electric cir-
(T , V, D, Fq ) as the EL parameters of the circuits, and simply cuits do not contain any magnetic couplings between its dif-
express a circuit Σe by means of the ordered five-tuple: ferent loops. In the following section, we will show that this
assumption is no longer necessary.
Σe = (T , V, D, Fq , A(q)). (2)
3 Coupled-Magnetics
The EL equations actually form a balance of generalized In this section we treat the inclusion of coupled-magnetics
forces, or efforts, which in the electrical domain is given by the in the EL framework. First, a pair of magnetically coupled-
voltages. It involves both generalized position and generalized inductors is considered. Next, we generalize our developments
velocity coordinates for each separate physical element. In the to circuits containing several magnetic couplings between its
electrical domain that means that we attach to each separate different loops. As an example, we illustrate the potential of
element (inductor and capacitor) two state-variables, namely a the method using the well-known Ćuk converter in which both
charge and a current. Clearly this does not correspond to the the inductors are coupled, e.g. [1].
physical intuition, but it can be viewed as if for the inductor
the charge is an intermediate help variable, and for the capac- A pair of coupled-inductors may be considered as the non-
itor the current is. These help variables are finally removed ideal equivalent of an AC transformer, with a√rate of coupling,
using the constraint equation. k ∈ [0, 1) and an effective turns ratio n = L1 L2 . As a re-
sult of the coupling, both the magnetizing currents share the
This framework can be used to model electrical networks with same flux paths with an order or magnitude depending on k.
linear inductive and capacitive elements and with or without This involves an additional path for the energy using a mag-
ideal switches. The switches can be naturally involved in the netic field. In terms of the common fluxes φi, j = 1, 2, a pair
constraints that follow from Kirchhoff’s current laws. If we of coupled-inductors can be characterized as follows
denote by the scalar u the switch position, which is assumed     
to take values on a discrete set of the form {0, 1}. The com- φ1 L1 L12 iL1
= , (3)
plete procedure is as follows: φ2 L21 L2 iL2
Procedure: for which L12 ≥ 0 and L21 ≥ 0 satisfy
√the condition of reci-
1. Give all N dynamic elements in the network two coordi- procity, i.e., L12 = L21 = Lm = k L1 L2 denoted as the
nates, namely a charge and a current coordinate, qj , and mutual inductance.
q̇j , j = 1 . . . N .
2. Determine the corresponding energy for all ideal ele-
ments, i.e., the magnetic co-energy for the inductive el-
-
+ i
Ls1 Ls2

iL2 +
L1
ements, denoted by T (q, q̇), and the electric field energy vL1 vL2
φ1 φ2
~
for the capacitive elements, denoted by V(q). In case of a Lm
switching network, this step does not involve the position − ideal −
of the switch.
1:1

3. Determine the Rayleigh dissipation energy, denoted by -


+ i

iL2 +
D(q̇), for the resistive elements, which may involve the L1
switch position u, and the use of a Kirchhoff current law vL1 ΣT vL2
for determining the current through the resistive element − −
in terms of the dynamic elements as given in step 1.
4. Determine the generalized forcing functions Fq given by Figure 1: Equivalent topology of a pair of coupled-inductors
the voltage sources, possibly depending on the switch po-
sition. ¿From these relations it is clear that L1 and L2 would form two
separate inductors if k = 0, and thus Lm = 0. For non-zero
5. Give the constraint equations that are determined by values of k, Lm is representative for the interaction between
Kirchhoff’s currents laws, that do not include the laws the inductors. Now, in order to obtain expressions for the cur-
of step 3, and thus only involve the currents through the rents and voltages, equation (3) is rewritten as
dynamic elements. If there are no constraint equations     
for this step, then put A(q) = 0. d iL1 β −γ vL1
= , (4)
dt iL2 −γ α vL2
6. Plug the information of the previous steps in the con-
straint form of the EL equations (1) and determine a state dφj
where vLj = dt , j = 1, 2 and α, β and γ may be expressed
space model by choosing the currents corresponding with
as
the inductive elements, and either the charge or the volt-
age corresponding with the capacitive elements, as state n2 1 nk
α= ;β = ;γ = . (5)
variables. (1 − k 2 )L1 (1 − k 2 )L1 (1 − k 2 )L1
The term 1 − k 2 can be considered as the magnetic flux disper- be steered into the output inductor, or vice-versa, to result in
sal, which denotes the amount of flux not shared by both the practically zero ripple current on either the input or the out-
inductors. Notice that other parameterizations of (5) are also put of the converter. Following the general procedure as given
possible but, as will be illustrated later in the example, these in section 2, we start with defining the following (intermedi-
notations provide a straight forward insight in the magnetizing ate) state variables qj , q̇j , j = L1 , C1, L2 , C2 . We proceed
energy interconnections. A visual representation of (3) and (4) with equating the magnetic co-energy, T (q, q̇), and potential
may be given as in Fig. 1, in which Lsj = Lj − Lm are the energy, V(q), of the circuit
leakage inductances of the primary and the secondary wind-
ing, respectively. In view of the Lagrangian modeling proce- T (q̇L1 , q̇L2 ) = 12 L1 q̇L
2
1
+ 12 L2 q̇L
2
2
+ Lm q̇L1 q̇L2 ;
(8)
dure, we consider a pair of magnetically coupled inductors as V(qC1 , qC2 ) = 1 2 1 2
2C1 qC1 + 2C2 qC2 .
a single system ΣT for which the total amount of stored energy
T in terms of the currents is, using (3), given by The remaining EL properties are readily found to be
     D(q̇L2 , q̇C2 ) = 12 R(−q̇L2 − q̇C2 )2 ;
iL1 L1 Lm iL1
T (iL1 , iL2 ) = 12 . (6)
iL2 Lm L2 iL2 FqL1 = E; FqC1 = FqL2 = FqC2 = 0; (9)
For general descriptions of the magnetic energy of an electrical A(q) q̇ = q̇C1 + uq̇L2 − (1 − u)q̇L1 = 0.
network containing N inductors with N magnetically coupled
loops, it follows in a similar way that Plugging (8) and (9) into the constraint equation (1) yields for
  the coupled-inductor Ćuk converter:
L1 L12 · · · L1N
 L21 L2 · · · L2N  L1 q̈L1 + Lm q̈L2 = −(1 − u)λ + E
 
T (q, q̇) = 12 q̇   . .. .. ..  q̇, (7) 1
q = λ
 .. . . .  C1 C1
L2 q̈L2 + Lm q̈L1 = R(−q̇L2 − q̇C2 ) + uλ (10)
LN1 LN2 · · · LN 1
C2 q C2 = R(−q̇L2 − q̇C2 )
where i is subsequently replaced by its generalized coordinates 0 = q̇C1 + uq̇L2 − (1 − u)q̇L1 ,
q̇ = [q̇1 , . . . , q̇N ] . Notice that T (q, q̇) satisfies T (q, q̇) ≥ 0.
which results in the state equations for x = [x1 , x2 , x3 , x4] =
The separate inductors are denoted by L1 , . . . , LN , while the
[q̇L1 , C1−1 qC1 , q̇L1 , C2−1 qC2 ] , in terms of α, β and γ as
mutual inductances L12 , L21 , . . . , LN−1,N , LN,N−1 , denote
the various coupling paths between these inductors. If, for ex- ẋ1 = −uγx2 − (1 − u)βx2 − γx4 + βE
ample, there exists no coupling between the M − 1th and the ẋ2 = (1 − u) C11 x1 − u C11 x3
M th inductor the corresponding coefficients are set to zero. (11)
ẋ3 = uαx2 + (1 − u)γx2 + αx4 − γE
As before, notice that in case of a symmetrical network the
condition of reciprocity is satisfied. ẋ4 = − C12 x3 − RC1
2
x4 .

Example: Coupled-inductor Ćuk converter This is the model with discrete values for the switch, where
This example illustrates the potential of the proposed pro- x1 , x3 represent the inductor currents, and x2 , x4 represent the
cedure for switching networks, including coupled-magnetics. capacitor voltages. We have thus shown that the assumptions
We use the coupled-inductor Ćuk converter, which serves as a in e.g. [11] and [9] are no longer necessary. With help from
case study throughout the paper. We explicitly assume that the the definitions in (5) it is now easy to see that for the matching
converter operates at continues conduction mode. condition, n = k, α = γ, the state x3 does not depend on the
switch position function anymore, which results in zero ripple
-q̇
L1

q̇L2 output current. The same holds for x1 in case of the inverse
matching condition, n = k −1 , β = γ, which results in zero
ΣT q C2 ripple input current. A third practically relevant condition can
C2 be found for n = 1, 0 ≤ k < 1, α = β, which is denoted as
+ C1 6 the balanced ripple reduction. Both the current ripples can be
E

q C2 R reduced between 0 and approximately 50%.
C1
Remark: These conditions, unfortunately, are often not easy
C1 to acquire in practice, i.e., there will always be a certain ‘mis-
u=1 u=0
match’ between k and n. We come back to this in the next
section. Coupled-inductor extensions can also be applied to
Figure 2: Coupled-inductor Ćuk converter the classical Buck, Boost and Buck-Boost converters, but in
these cases an extra capacitor is needed to serve as a driven
The circuit topology of the coupled-inductor Ćuk converter is
voltage source for the secondary inductor, see e.g. [14].
depicted in Fig. 2. The capacitive energy transfer imposes
identical rectangular voltage waveforms on both the induc- As shown in [6, 9, 11], the switched EL equations are also
tors which has justified the magnetic extension, [1]. Providing closely related to the average PWM models. For the Ćuk con-
the right adjustments of k and n, the input current ripple can verter given by the dynamic equations (11) this means that the
state variable x is replaced by the average state z, represent- the unstable zero-dynamics (i.e., for linear systems this corre-
ing the average inductor currents and capacitor voltages, and sponds to zeros in the right half plane, or in other words, un-
the discrete signal u is replaced by the duty ratio function D desirable non-minimum phase behaviour), and therefore, had
that takes values in the open interval ]0,1[. The description to be indirectly controlled via the average input-current, which
of the dynamics as described in words above is given by the exposed stable zero-dynamics.
following more compact matrix representation
Example (continued) As in [11], given a desired equilibrium
Aż + DJ1 z + J2 z + Rz = E (12) value V2d for the output voltage, which correspond to a con-
stant value of the duty ratio function Dc = V2d /(V2d − E),
where A is invertible, and where Jj = −Jj and its elements the unique corresponding equilibrium values for the average
are given by {−1, 0, 1}, for j = 1, 2. For the coupled-inductor voltages and currents are

Ćuk converter we have E = [E, 0, 0, 0] and V22d V2
    I1d = ; I2d = − d ; V1d = E − V2d . (14)
L1 0 Lm 0 0 −1 0 0 RE R
 0 C1 0 0    This means that if we want to regulate z4 toward an equilib-
A= ;J =  1 0 1 0 ;
 Lm 0 L2 0  1  0 −1 0 0  rium value V2d which is known to correspond to a steady state
0 0 0 C2 0 0 0 0 value Dc of the duty ratio function D, then such a regulation
can be indirectly accomplished by stabilizing one of the other
   
0 0 0 0 0 1 0 0 average variables toward the corresponding equilibrium values
 0 0 0 0   −1 0 0 0  computed in (14).
R=
 0
 ; J2 =  .
0 0 0   0 0 0 −1 
Henceforth, for the uncoupled-case (k = 0) we know for the
0 0 0 R−1 0 0 1 0
Ćuk [11], Boost and Buck-Boost converters [9], that a feasi-
2 ble regulation of the output voltage is only achievable through
The presented Euler-Lagrange modeling technique for indirect regulation of the input current. But if k = 0, does
(switching) networks results in the same dynamical models as the input current still exhibit stable zero-dynamics under all
when the Hamiltonian framework is used, e.g. [2, 5], provided coupling conditions? From the previous section, we know we
that the same level of ideal physics is assumed. However, have two special cases of interest, k = n and k = n−1 , both
the Hamiltonian framework does not introduce the ‘semi’- with its related ‘mismatches’ 0 ≤ k < n, n < k < 1 and
physical intermediate help-variables. In case of two mutually 0 ≤ k < n−1 , n−1 < k < 1, respectively. 2
coupled-inductors for example, the magnetic energy is now
given by For the converter structures we consider here, we can bring
the system in normal form (see e.g., [3]) for the output y
    
φL1 β −γ φL1 given by the state zi , which has relative degree ri , then we
H(φ) = 1
, (13) obtain the new coordinates, ξ = [ξ1 , . . . , ξri ], and η =
2 φL2 −γ α φL2
[ηri+1 , . . . , ηN ] , with [ξ, η] = ϕ(z), and ϕ1 = y, . . . ϕri =
where α, β and γ are given in (5). Nevertheless, the La- y(ri ) . For the η coordinates, that are related to the state of the
grangian approach gives us the opportunity to apply the well- zero-dynamics we obtain
known passivity-based control techniques, as presented in
∂ϕi (z) −1
[6, 9] for the buck, boost and buck-boost converter, but this A J1 z = 0, i = ri + 1, . . . , N.
time for more general electrical network structures. ∂z
This implies that at least the total energy, given by
4 Zero-Dynamics
ϕri +1 (z) = 12 z  Az (15)
In the sequel, we continue with the dynamic PWM model of
the converter structures, as can be obtained from the previous is part of the zero-dynamics. If we drop the assumption that
section, where we now denote the averaged state space vari- the generalizing forces are independent of the switch, we can
ables by z, and the duty cycle by D. The design of passivity still do the above analysis, but then we have to add to the total
based control for the conventional buck, boost, buck-boost and energy of (15) an additional term which depends on the voltage
Ćuk converter can be found in [9, 11], whereas further anal- source.
ysis of the closed loop system is explored in some follow-up Example (continued) If we consider the average input current
work, e.g. [12, 13]. Since we have given a general procedure z1 as the output, for the matching condition n = k, α = β,
to build an Euler-Lagrange model for switching networks, we and thus that
can also generally apply the passivity based control design ξ1 := ϕ1 (z) = z1 ,
technique. However, one issue that remains is the choice of
the average state variable to be stabilized to a certain value, and that the state of the zero-dynamics is going to be deter-
in order to, possibly indirectly, regulate our average output to- mined by
ward a desired equilibrium value. For the conventional boost, η1 := ϕ2 (z) = 12 C1 z22 + 12 (β − α)−1 (z1 + z3 )2
buck-boost and Ćuk converters, it was shown in [9, 11] that the η2 := ϕ3 (z) = z3
average output voltage could not be directly controlled, due to η3 := ϕ4 (z) = z4 .
(a) 0 ≤ k<n, n=0.7 4 (b) n<k<1, n=0.7
It is seen that the relative degree equals r1 = 1. We then 2
x 10
4
2
x 10

proceed by calculating z = ϕ−1 [ξ, η], and form the zero-


dynamics by giving z1 its desired value I1d , i.e., then ξ1 = I1d , 1 1

and we consider the dynamics of η, see [3]. In order to deter-

Im
Im
0 0
mine whether or not the zero-dynamics are stable, we linearize
the dynamics at the equilibrium point η 0 , that corresponds to −1 −1

(14). In order to numerically compute the eigenvalues, we


−2 −2
need to complete the design of our converter first. In the se- −4000 −3000 −2000 −1000 0 −4000 −3000 −2000 −1000 0
Re Re
quel we use the following realistic values [11]: E = 100V,
x 10
5 (c) 0 ≤ k<n−1, n=(0.7)−1 5
x 10 (d) n−1<k<1, n=(0.7)−1
R = 40Ω, C1 = C2 = 10µF, and the coupling parameters are 1.5 1.5

set to: n = 0.7, 0 ≤ k < 1, where for the matching condition 1 1

we take k = n, L1 = 600µH, L2 ∝ k and for the inverse 0.5 0.5

matching we take L2 = 600µH, L1 ∝ k . In case for the

Im

Im
0 0

zero-dynamics of z1 for the matching condition this yields for −0.5 −0.5

the linearization the eigenvalues (−1655, −2248 ± 1538.9i). −1 −1

We conclude that for this condition z1 shows stable zero- −1.5 −1.5
−1 0 1 2 −1 0 1 2
dynamics. Re 5
x 10 Re 5
x 10

Next, consider the case that k = n−1 . In this case the normal
Figure 3: Eigenvalues of the zero-dynamics of the input current for: (a,b)
form is found as matching-condition mismatches, (c,d) inverse matching-condition
mismatches
ξ1 := ϕ1 (z) = z1 ,
ξ2 := ϕ2 (z) = −βz2 + βE − βz4 = ż1 ,
5 Adaptive Passivity-Based Stabilization of the
and determined the state of the zero-dynamics by Zero-Output Ripple Ćuk Converter
η1 := ϕ3 (z) = 12 C1 z22 + 12 (β − α)−1 (z1 + z3 )2 We now provide the only feasible regulation based on an in-
η2 := ϕ4 (z) = z4 . direct output capacitor voltage control, achievable through the
regulation of the input current. In order to account for para-
We now notice that the relative degree equals r1 = 2, which metric uncertainty and changes in load we define the following
tells us that that the order of the zero-dynamics is reduced from version of the control reference as
3 to 2. Following the same procedure as above, we calculate
2
the eigenvalues after the linearization as (−1655, −100850). z1d = I1d = θ
V2d
, (16)
Again, we conclude that the regulation through z1 remains fea- E
sible.
where the quantity θ denotes the estimate of R−1 . Corre-
Finally, we consider the (practical) ‘mismatch’ cases. The sponding to this objective for the input current z1 , the required
general description of the normal form is now given by input voltage, output voltage and output current may be repre-
sented by the functions z2d (t), z3d (t) and z4d (t), to be deter-
ξ1 := ϕ1 (z) = z1 , mined. Now we follow the procedures as given in [6, 9, 11],
and obtain the following implicit definition of our controller
and that the state of the zero-dynamics is going to be deter-
that preserves passivity of the closed loop:
mined by
Ażd + DJ1 zd + J2 zd + Θzd − Ri z̃ = E (17)
η1 := ϕ2 (z) = 12 (β − γ)−1 z12 + 12 C1 z22 + 12 (α − γ)−1 z32
η2 := ϕ3 (z) = − C11 z1 − C11 z3 where z̃ = z−zd , denotes the state error dynamics. The neces-
η3 := ϕ4 (z) = z4 . sary damping injection that is required
for asymptotic−1 stability

is accomplished through Ri = diag R1 , R−1 2 , R3, R4 . The
Again we perform similar calculations as before, but now for
estimation matrix is denoted Θ = diag[0, 0, 0, θ], where θ is
all values of k in the closed interval 0 ≤ k < 1, except for
given by
k = n and k = n−1 . The calculations are performed us- t
ing MATLAB, and we find the eigenvalues as a function of k, θ = θ0 − z2d z̃4 dτ. (18)
as depicted in Fig. 3, where ‘◦’ is denoting the corresponding 0
starting points for k. ¿From Fig. 3.(d) we conclude that in case Hence, we find an explicit definition of our controller by
of inverse matching condition ‘mismatch’ k > n−1 all states choosing the control rule
in the converter show non-minimum phase behaviour. For that,
2
V2d
it is not possible to apply the passivity-based controller tech- βz2d + γz4d − βE − βR1 z̃1 + γR3 z̃3 − θ̇ E
niques. Fortunately, this range of adjustments will not be of D= (19)
(β − γ)z2d
prime concern in a practical situation where we rather choose
for zero output current ripple, corresponding to the matching and let the functional relations z2d , z3d , z4d as in (17). For fur-
condition instead of the inverse matching condition. ther analysis, we refer to related considerations and remarks
as given in [6, 9]. The above developments lead to a full-state Future research is recommended towards the involvement of
feedback scheme, which is based on philosophies that can be more non-ideal physical effects, a more general statement
found in [4, 8]. about the non-minimum phase structure, and the influence of
more complex topologies that are given in the power electron-
Current i
L1
[A] Current i
L2
[A] ics literature. Furthermore, in general it remains a problem to
25 20
tune the controller in an optimal way, since the correspond-
15
20
ing equations are quite complex. It is therefore recommended
10
15 to search for smart algorithms to tackle this problem. Some
5
10 preliminary results on this topic can be found in [12, 13].
0

5
−5
Acknowledgement
0 −10
0 0.005
t [s]
0.01 0.015 0 0.005
t [s]
0.01 0.015 The first author likes to thank Romeo Ortega and Gerardo Es-
Voltage v
C1
[V] Voltage v
C2
[V] cobar for the preliminary discussions on the modeling proce-
350 100

300 50
dure and the zero-dynamics.
250 0

200 −50
References
150 −100 [1] S. Ćuk and R.D. Middlebrook, Coupled-Inductor and Other
100 −150 Extensions of a New Optimum Topology Switching DC-to-DC
50 −200 Converter, Advances in Switched-Mode Power Conversion,
0
0 0.005 0.01 0.015
−250
0 0.005 0.01 0.015
Vol. 2. Pasadena, CA: TESLAco, 331-347, 1981.
t [s] t [s] [2] G. Escobar, A.J. van der Schaft, R. Ortega, A Hamiltonian view-
Duty Ratio Function [] −−− Resistance [Ω], Θ [Ω]
1 80 point in the modeling of switching power converters, Automat-
ica, Vol. 35 (1999) 445–452.
0.8 70
[3] A. Isidori, Nonlinear Control Systems, Third Edition, Springer
0.6 60 Verlag, London (1995).
0.4 50 [4] J.G. Kassakian, M. Schlecht and G.C. Verghese, Principles of
power electronics, Addison-Wesley, Reading, MA (1991).
0.2 40
[5] B. Maschke, R. Ortega, A.J. van der Schaft, Energy-based Lya-
0
0 0.005 0.01 0.015
30
0 0.005 0.01 0.015 punov functions for forced Hamiltonian systems with dissipa-
t [s] t [s]
tion, to appear in IEEE Trans. Aut. Contr..
[6] R. Ortega, A. Lorı́a, P.J. Nicklasson and H. Sira-Ramı́rez,
Figure 4: Simulation of the closed-loop passivity-based Ćuk converter with Passivity-Based Control of Euler-Lagrange Systems; Mechan-
zero-output ripple
ical, Electrical and Electromechanical Applications, Springer-
Verlag London Limited, 1998.
In Fig. 4, some simulation results are depicted for the Ćuk con-
[7] H. Sira-Ramı́rez, A geometric approach to pulse-width-
verter (11) in closed-loop with the controller (17). The same
modulated control in nonlinear dynamical system, IEEE Trans.
values, with k = n, for the components as in the previous sec- Aut. Contr. AC-34 (1989) 184–187.
tion have been considered for the simulation. To depict the [8] H. Sira-Ramı́rez, P. Lischinsky-Arenas, Dynamical discontin-
robustness of the controller with respect to abrupt unmodeled uous feedback control of nonlinear system, IEEE Trans. Aut.
changes in the load resistance R, a temporary (step) change Contr. AC-35 (1990) 1373–1378.
in the load resistance of 20% of its nominal value is applied. [9] H. Sira-Ramı́rez, R.A. Perez-Moreno, R. Ortega, M. Garcia-
The control parameters are set to R1 = 1, R2 = 2, R3 = 10 Esteban, Passivity-Based Controllers for the Stabilization of
and R4 = 15. The initial conditions of the Ćuk converter DC-to-DC Power Converters, Automatica, Vol.33, 4 (1997)
where set x0 = [0, 0, 0, 0], while for the desired control state 499–513.
zd0 = [8, 300, 4, −200]. As can be seen from the figure, the [10] A.J. van der Schaft, L2 -gain and passivity techniques in non-
controller achieves the desired indirect stabilization of the out- linear control, Springer Verlag, Lecture Notes in Contr. & Inf.
put voltage around the desired equilibrium value. During the Sc. 218, 1996.
[11] J.M.A. Scherpen, J.B. Klaassens and L. Ballini, Lagrangian
sudden change in load, the controller rapidly manages to es-
Modeling and Control of DC-to-DC Converters, IEEE INT-
timate the (new) value of the load resistance. Notice that the
ELEC Copenhagen, June 1999, 99CH37007, Nr. 31-41.
output current ripple equals zero as expected. [12] J.M.A. Scherpen, R. Ortega, On nonlinear control of Euler-
Lagrange system: Disturbance attenuation properties, System
6 Conclusions & Control Letters 30,pp.49–56, 1997.
[13] J.M.A. Scherpen, R. Ortega, G. Escobar, On disturbance atten-
A method has been presented to build an Euler-Lagrange
uation properties of control schemes for Euler-Lagrange sys-
model for switching electrical networks, which is also appli- tems: theoretical and experimental results, Proc. Eur. Contr.
cable to ideal electrical networks without switches. In addi- Conf. Brussels, Vol. 6a, FR-A A6, 6pp., (July 1997).
tion, we have included coupled magnetics into the Lagrangian [14] J. Wang, W.G. Dunford and K. Maunch, Analysis of a Ripple-
framework. The method presented here can also be performed Free Input-Current Boost Converter with Discontinuous Con-
for Hamiltonian systems, which is a recent topic of study in duction Characteristics, IEEE Power Electronics, Vol. 12,
e.g., [2, 5]. No. 4, July 1997.

Vous aimerez peut-être aussi