Download as pdf or txt
Download as pdf or txt
You are on page 1of 25

Power-shaping of reaction systems: the CSTR

case study
A.Favache, D.Dochain 1
IMAP, Universite catholique de Louvain, 2 place Sainte Barbe,
B-1348 Louvain-la-Neuve, Belgium
CESAME, Universite catholique de Louvain, 4-6 avenue G. Lemaitre,
B-1348 Louvain-la-Neuve, Belgium

Abstract
The non-isothermal continuous stirred tank reactor (CSTR) is a classical yet complex case study of nonlinear dynamical systems. Power-shaping control is a recent
approach for the control of nonlinear systems based on the physics of the dynamical
system. In this paper we apply the power-shaping control approach to the nonisothermal CSTR study case. A global Lyapunov function is derived for the openloop exothermic CSTR. This Lyapunov function is then reshaped by the means of
a controller in order to stabilize the process at a desired temperature. Some considerations on the local and global convergence to the desired state are presented.
Key words: non-isothermal CSTR, power-shaping control

Introduction

Thermodynamic systems, and among them chemical reaction systems, are


usually nonlinear dynamical systems. They can therefore have a complex behaviour and be difficult to analyze and to control. Stability analysis of nonlinear systems requires the use of abstract mathematical tools such as the two
Lyapunov methods or the passivity theory. Over the past years, several works
have combined those abstract concepts with the underlying physical phenomena giving rise to the dynamical behaviour of the system. These works include
for instance the study of port-Hamiltonian systems [4][7], energy-balancing
Email address: audrey.favache@uclouvain.be,
denis.dochain@uclouvain.be (A.Favache, D.Dochain).
1 Corresponding author.

Preprint submitted to Automatica

13 October 2009

passivity based control (PBC) [13][19] or the introduction of the contact


formalism for expressing the dynamics of systems in which irreversible phenomena arise [5][6][11]. The non-isothermal continuous stirred tank reactor
(CSTR) is a classical study case of nonlinear systems. Indeed the dynamical behaviour exhibits complex features, such as multiple equilibrium points.
Up to now no precise physical interpretation of the complex behaviour of the
non-isothermal reactor has been found [10].
Power-shaping control [18] has been developed in the past years as an extension of energy-balancing passivity-based control [13][19]. In energy-balancing
passivity based control, the controller reshapes the energy function of the system so that it has a minimum at the desired equilibrium point. The controller
provides the system a finite amount of energy so as to drive the system to
the desired state. This concept has been applied to electro-mechanical systems [17][20] and also to thermodynamic systems where the storage function
is the entropy instead of the energy [1][21]. Nevertheless energy-balancing
passivity-based control cannot be applied to some systems (namely systems
with pervasive dissipation). To overcome this difficulty the concept of powershaping control has been introduced, firstly for the stabilization of nonlinear
RLC circuits [18]. Contrary to energy-balancing passivity-based control, the
storage function used for the control is related to the power and not to the
energy. Power-shaping control has subsequently been applied to the control of
mechanical and electromechanical systems [12].
This paper is dedicated to the analysis and control of the CSTR study case
by the means of the power-shaping approach. First we shall briefly introduce
the concepts of the power-shaping approach in Section 2. Then in Section 3
we shall explain the general methodology for writing the Brayton-Moser form
on the particular case of a CSTR. This methodology shall then be applied to
obtain the Brayton-Moser formulation of the dynamics of a CSTR with an
irreversible reaction (Section 4). This formulation shall be used to design a
control system for this CSTR using the power-shaping approach. Finally we
shall give some considerations about the generalization of this work to more
general cases. In Section 5 we shall extend the results obtained for the CSTR
with an irreversible reaction to more complex reactions and reaction kinetics.
Section 6.1 presents some primary results for the Brayton-Moser formulation
for a general CSTR.

Power-shaping control

In this section we briefly explain the principles of the power-shaping approach.


The statements are given without any proof. For more details, the reader can
refer to [12][14][15][18].
2

2.1

The Brayton-Moser formulation

Let us consider a dynamic system of dimension N with m inputs. The state


of the system is given by the vector x RN and the input is given by vector
uc Rm . The power-shaping control theory is based on the Brayton-Moser formulation of the system dynamics [3]. In this formulation the system dynamics
are of following form:
Q (x)

dx
= P (x) + G (x) uc
dt

(1)

where Q (x) : RN RN RN is a non-singular square matrix, P (x) : RN


R is a scalar function of the state and G (x) : RN RN Rm . Additionally
the symmetric part of the matrix Q (x) is negative semi-definite, i.e.:
Q (x) + Qt (x)  0

(2)

The function P (x) is called the potential function. In electrical and mechanical systems, the potential function has the units of power and is related to the
dissipated power in the system. In the first one it is related to the so-called
content and co-content of the resistances [18][15]; in the latter it is related to
the Rayleigh dissipation function [14].
Let us now assume the system dynamics is given by the following relation:
dx
= f (x) + g (x) uc
dt

(3)

where f (x) : RN RN and g (x) : RN RN Rm . The system (3) can be


written in the form (1) if there exists a non-singular matrix Q (x) fulfilling (2)
and that solves following partial differential equation 2 :
(Q (x) f (x)) = t (Q (x) f (x))

(4)

P (x) is the solution of the following partial differential equation system:


P (x) = Q (x) f (x)

(5)

and the function G (x) is given by G (x) = Q (x) g (x).


2.2

Power-shaping control

Let us assume that the system dynamics can be expressed by using the
Brayton-Moser equations presented before. The desired equilibrium state is
2

This condition is equivalent to the existence of the potential function P (x).

denoted by x . The principle of power-shaping control is to choose the input


uc (x) such that in closed loop the system dynamics are given by the following
relation:
dx
= Pd (x)
(6)
dt
where Pd (x) : RN R is the re-shaped potential function. The desired
equilibrium point x must be a local minimum of the potential function Pd (x)
in order to be locally asymptotically stable. The function Pd (x) can be used
as a Lyapunov function for the closed-loop system.
The function Pd (x) cannot be chosen arbitrarily since the following relation
must be fulfilled:
g (x) Q1 (x) Pa (x) = 0
(7)
where Pa (x) = Pd (x) P (x) and
g (x) : RN RN m RN


is a full-rank left annihilator of g (x) (i.e. g (x) g (x) = 0 with rank g (x) =
N m). Under these conditions, the control input u (x) that re-shapes P (x)
into Pd (x) is the following one:


uc (x) = g t (x) Qt (x) Q (x) g (x)

1

g t (x) Qt (x) Pa (x)

The Brayton-Moser PDE for a general CSTR

3.1

3.1.1

The dynamical model

The general CSTR

Let us consider a liquid-phase CSTR with constant volume V containing Nc


species and in which Nr reactions take place. The reactor is cooled/heated
by a surrounding jacket. As it has been shown in [10] the dynamics of such a
system are given as follows (i = 1, ..., Nc ):
Nr
 X
F  in
dni
=
Ci V ni +
il rl (T, n)
dt
V
l=1

dU
F  in
h V H + Q
=
dt
V

(8a)
(8b)

with ni the quantity of species i, n is a column vector with ni in the ith


position, U the internal energy of the mixture, F the volume flow rate, Ciin is
the concentration of i in the inlet flow, il is the stoichiometric coefficient of i
4

in reaction l, rl is the reaction rate of reaction l, hin is the volumetric enthalpy


of the inlet flow, T the temperature, H the enthalpy of the mixture and Q the
heat transferred into the reacting mixture. H and T are functions of U and
n depending on the thermodynamic model of the liquid (e.g. an ideal liquid
mixture):
(n, U ) , T = T (n, U )
H=H
Using the temperature function as a state transformation, (8b) can be replaced
by the following relation:
Nr
X
dT
in

(r H)l (T ) rl (T, n)
= Q + Fh +
Cp V
dt
l=1

(9)

where (r H)l (T ) is the molar reaction heat 3 of reaction l and Cp V =


PNc
i=1 Cpi ni is the total heat capacity of the mixture (with Cpi the volumetric
heat capacity of species i).
The aim is to control the reactor temperature by acting on the heat transfer.
As explained in the previous section, the change from the usual form of the
system dynamics (3) into its Brayton-Moser formulation (1) requires the solution of the partial differential equation system (4). Using (8) (4) is written
as follows:

xj

=
xi

Nc
X

Nr
 X
F  in
qik (x)
C i V ni +
il
V
k=1
l=1
Nc
X

"

#!

#!
Nr
 X
F  in
qjk (x)
Cj V nj +
jl
V
k=1
l=1
"

(10)

The solution of these equations turns out to be quite intricate and we do not
even know if a solution exists. In order to keep the resolution as simple as
possible we shall first work with a simplified version of the CSTR model.

3.1.2

The simplified CSTR

In order to simplify the model, we consider the following additional assumptions:


the specific heat Cp of the mixture are constant.
the reaction enthalpy (r H)l of reaction l is independent of the temperature.
Following usual thermodynamic notations (r H)l > 0 for exothermic reactions
and (r H)l < 0 for endothermal reactions.

the heat exchange between the reactor and the jacket is proportional to the
temperature difference between them, with the the heat transfer coefficient.
The aim is to control the reactor temperature by acting on the heat transfer,
and more precisely by acting on the cooling fluid flow rate. Changing the
cooling fluid flow rate influences the system dynamics via the heat transfer
coefficient which is dependent on the cooling fluid flow rate. If the jacket
dynamics are fast enough compared to the reactor dynamics, this relation is
time independent. Therefore we shall consider in the sequel that the control
input uc is the quantity CpV . However implementing this control law in practice
requires that the relation between the cooling fluid flow rate and the heat
transfer coefficient is know. The open-loop case corresponds to = 0, i.e. an
adiabatic reactor.
This additional assumptions allows us to rewrite (8) as follows [2][23]:

Nr
 X
F  in
dni
il rl (T, n)
=
Ci V ni +
dt
V
l=1

(11)

Nr
 X
dT
F  in
=
l rl (T, n) + uc (Tw T )
T T +
dt
V
l=1

with Tw the (constant) cooling fluid temperature and T in is the inlet temperature.
The time evolution equations of n and T have the same form and the dynamics
of the system can be written as follows:

F  in
dx
=
x x + r (x) + uc [0, ..., 0, Tw T ]t
dt
V

(12)

where x = [nA , . . . , nNc , T ]t . RNc +1 RNr is a matrix that contains the


stoichiometric coefficient of species i in the reaction l in position (i, l) if 1
i Nc and that contains l in the lth column if i = Nc + 1. Applying the
notations of Section 2 to (12), we have:
f (x) =

3.2


F  in
x x + r (x)
V

and

t

g (x) = 0, , 0, Tw T

Integrability condition: the Brayton-Moser partial differential equation

The first step to find the Brayton-Moser form of the dynamics is the solution
of a partial differential equation, i.e. find a matrix Q (x) that fulfills (2) and
6

(4). By using (12) we can write the following relation (for i = 1, . . . , Nc + 1):
(Q (x) f (x))i =

NX
c +1

qik

k=1



F  in
xk xk + kl rl (x)
V

By differentiating the above expression the condition (4) can be rewritten as


follows for all i, j = 1, . . . , Nc + 1:
NX
Nr
c +1
X
F
rl
qij +
qik
kl
V
xj
k=1
l=1

"

NX
c +1
k=1

qik
xj



F  in
xk xk + kl rl (x)
V

NX
Nr
c +1
X
rl
F
kl
= qji +
qjk
V
xi
l=1
k=1

"

NX
c +1
k=1

qjk
xi

(13)



F  in
xk xk + kl rl (x)
V

This partial differential equation is difficult to solve in general. So we shall


now restrict ourselves to a particular set of possible matrices Q(x) to find a
solution to (4). This set is such that the partial differential equation system
is transformed into an algebraic one. (13) is transformed into an algebraic
equation by setting:
i, j, k = 1, . . . , Nc + 1, i 6= j :

qik
=0
xj

(14)

(13) becomes then as follows:


NX
Nr
c +1
X
F
rl
qik
kl
qij +
V
xj
l=1
k=1

NX
Nr
c +1
X
F
rl
= qji +
qjk
kl
V
xi
k=1
l=1

"

"

(15)

The above set of algebraic equations is still complex to solve due to the general form of the kinetics. Therefore, we shall also assume that the kinetics of
reaction l can be written as a product as follows:
rl (x) = kl (T )

Nc
Y

(l)
m (nm )

m=1
(l)
The functions (l)
m (nm ) are smooth functions of their argument and m (0) = 0
if the species nm intervenes in the kinetics of reaction l; otherwise
(l)
m (nm ) = 1. kl (T ) is the kinetic coefficient of the reaction l. kl (T ) is such

that the reaction rate rl (x) fulfills the following conditions 4 :


kl (T ) > 0,

lim kl (T ) = 0,

dkl
> 0,
dT

lim kl (T ) = kl > 0

T 0

dkl
= 0,
T 0 dT

dkl
=0
T dT

lim

(16)

lim

Remark 1 This assumption restricts the set of reaction kinetics that are considered since it excludes for example the Michaelis-Menten kinetics for enzymatic reactions. However the considered ones still cover a large number of
practical cases, and among them the reactions with mass action law kinetics.
(15) is written as follows (for i, j = 1, . . . , Nc + 1 with i 6= j):
for i, j 6= Nc + 1:

(l)

Nr
X
j
F
kl (T ) (l) (ni )
qij (ni ) +
i
V
nj
l=1

NX
c +1

qik (ni ) kl

! N
Yc

(l)
m (nm )

m6=i,j

k=1
Nr
X

(17a)

(l)

F
i
kl (T ) (l) (nj )
= qji (nj ) +
j
V
ni
l=1

NX
c +1

qjk (nj ) kl

k=1

! N
Yc

(l)
m (xm )

m6=i,j

for j = Nc + 1:
Nr
X
dkl
F
(l)
i (ni )
qij (ni ) +
V
dT
l=1

"

NX
c +1

qik (ni ) kl

k=1

Nc
Y

(l)
m (nm )

m=1,m6=i

(17b)

(l)

Nr
X
F
i
kl (T )
= qji (T ) +
V
ni
l=1

NX
c +1

qjk (T ) kl

k=1

Nc
Y

(l)
m (nm )

m=1,m6=i

A general solution of (17a) and (17b) has not been found, yet. Despite of this,
some characteristics of the solution (if it exists) can be derived. Indeed (17a)
has to be satisfied for any (n, T ) and in particular for T tending to 0. By
taking the limit of (17a) for T tending to 0, we find that the following relation
4

These assumptions on k (T ) are fulfilled by the commonly used Arrhenius law.

has to be fulfilled for i, j 6= Nc + 1 and i 6= j:


F
F
qij (ni ) = qji (nj )
V
V
This means that qij = qji and qij and qji have to be constants. Thus the only
elements that can depend on the composition in the matrices Q (x) that solve
(17a) and (17b) are the diagonal terms. All non-diagonal elements (except for
the last line and the last column) are constant.

The non-isothermal CSTR with first order irreversible reaction

As already mentioned, a general solution of (17a) and (17b) has not been
found, yet. Furthermore to this stage we do not even know if a solution exists.
Therefore we shall in a first instance work with the most simple case, namely
the case of an irreversible reaction with one reactant.

4.1

The dynamical model

In order to simplify the system dynamics to reduce its dimension as much


as possible, we shall the particular case of an irreversible reaction A . . .
with reaction kinetics that follow the mass action law. As a consequence the
reaction kinetics can be written as follows:
r (nA , T ) V = k (T ) nA
where k (T ) is the kinetic coefficient. k (T ) fulfills (16).
From (11) the dynamical behaviour of the CSTR can be represented by a
two-dimensional system as follows [2][23]:

F  in
dnA
=
CA V nA k (T ) nA
dt
V


F  in
dT

=
T T + k (T ) nA + uc (Tw T )
dt
V

(18)

(r H)
. It can be shown that this system can have several equilibcp V
rium points depending on the numerical values of the parameters. In this work
we do not consider the cases where limit cycles can occur. In the endothermic case there can be only one asymptotically stable equilibrium, whereas in
the exothermic case an odd number of equilibria can occur. In this last case,
there is an alternance of asymptotically stable and unstable equilibrium points
with =

[2][8][23]. The most common case which can be observed for instance with
Arhenius type kinetics is the case of three equilibrium points where the temperature of the unstable equilibrium is intermediate to the two other (asymptotically stable) equilibrium temperatures.
Proposition 2 The dynamics of the CSTR can be expressed in the form of
the Brayton-Moser equations with:

Q (nA , T ) =

q12
2
F
F 2
k(T )
k(T ) +4( V ) +4 V

k(T )+4 V
k(T )

k(T )

(19)

where 0 < 1 and q12 are two real coefficients such that

q12

< 0.

PROOF. It is sufficient to check that the matrix Q (x) given in (19) is nonsingular and fulfills (2) and (17b) 5 . Indeed the determinant of Q (nA , T ) is
non-zero:
 
4 2 F 2
det Q (nA , T ) = 2 q12
>0
k
V
Concerning (2) it is sufficient to check the sign of the trace and the determinant
of the symmetric part of Q (nA , T ) (the dependence in T of k (T ) has not been
written for the sake of shortness of the notations):


tr Q + Qt = 2q12

k 2 ( 2 + 1) + 4

 2
F

V
k 2

+ 4 VF k

6= 0

2
F 2 q12
(1 ) 0
V
k2
As a consequence the condition (17b) is fulfilled [8].

det Q + Qt = 16

The potential function P (nA , T ) is the solution of (5):


P (nA , T )
"
 
2
F 1   in
= q12
CA V nA + T in T
V 2
F
+ 4
V


2

Z T

CAin V k ( ) +

F
V

k ( )2

T0

+ k ( ) (T in )

(20)

where T0 > 0 is an arbitrary reference temperature. The first term is related


to the convection whereas the other terms are linked to the reaction kinetics.
5

(17a) does not exist because Nc = 1.

10

The equilibrium points of the system n


A , T are the solutions of f n
A , T =
0. As a consequence:


Q n
A , T f n
A , T = P n
A , T = 0
Hence the equilibrium points of the system are the critical points of P (nA , T ).
Proposition 3 The asymptotically stable equilibria are local minima of the
function P (nA , T ) and the unstable equilibria are saddle points.

PROOF. First let us determine the nature of the critical points of P (nA , T )
by computing the eigenvalues of the Hessian matrix H (P) at the equilibrium
points:
 2 

q12 F
F
1 + 2 + 4
tr (H (P)) =  V 2 k T
V
k T
"

det (H (P)) = 4

2
q12

 3
F
V

 2

k T

F
dk
k T + n
A
V
dT
 

 
F
dk
+ k T +
n
A
dT
V

!#

Besides the eigenvalues of the linearized system around the equilibria 1 and
2 are given as follows:
1 =

F
V

and 2 = n
A

 
dk
F
k T
dT
V

At an unstable equilibrium 2 > 0 whereas at the stable ones 2 < 0 [8]. As


a consequence, H (P) is positive definite at the stable equilibria, whereas it
indefinite at the unstable ones.

As shown in [8] there exist a particular choice of the reference temperature


T0 such that P (nA , T ) > 0 for all (nA 0, T > 0). More precisely this value
is equal to one of the stable equilibrium temperatures. With this particular
choice of T0 , P (nA , T ) > 0 is a Lyapunov function of the open-loop system.

4.2

Controller design

The aim is to design a control law uc = uc (nA , T ) such that the closed loop
system has an asymptotically stable equilibrium at the desired state (nA , T ).
Let us apply the methodology of Section 2.2.
11

80

nA [mol]
unstable equilibrium point

70
60

stable equilibrium point

50
40
30
20
10
0
380

separatrix
400

420

440

460

480

500

520

540

T [K]

Fig. 1. Level curves of P (nA , T ) for the case with three equilibrium points. The
low temperature equilibrium lies outside the frame.

Lemma 4 The desired set-point (nA , T ) can be an equilibrium point of the


closed-loop system if and only if:
nA =

F
V
F
V

CAin V
+ k (T )

(21)

PROOF. (nA , T ) is an equilibrium of the closed loop system, if and only if:

dnA
F  in
=
CA V nA k (T ) nA = 0
dt
V

(21) can be deduced directly from this equation.


Lemma 5 With Q (nA , T ) given in Proposition 2 there exists a control input uc = uc (nA , T ) such that the closed-loop dynamics can be written in the
form (6) with the re-shaped potential function Pd (nA , T ) = Pa (nA , T ) +
P (nA , T ) if and only if Pa (nA , T ) is of the following form:
Pa (nA , T ) = w nA +
|

Z T

v ( ) d

(22)

T0

{z
z

F
k(T )2 +4( V
) +4 VF k(T )

where v (T ) =
differentiable function.

k(T )2

. w (z) : R R is some continuously

12

PROOF. The closed-loop dynamics can be written in the form (6) if and
only if there exists a solution uc (nA , T ) to the equation
Pa (nA , T ) = Q (nA , T ) g (nA , T ) uc (nA , T )
This is the case if and only if Pa(nA ,T ) is the solution of the differential
equation (7). In our case g (x) = 0 with R so that (7) becomes as
follows:

q12
Pa Pa
v (T )

=0
det (Q (x))
nA
T
The solution of this partial differential equation is of the form given in (22).
Proposition 6 Let us denote by
uc (Tw T ) =


F  in
T T k (T ) nA
V

the equilibrium value of the control input. If the conditions of Lemmas 4 and
5 are fulfilled and if:

dw
= |q12 | uc (Tw T )
dz z
with

(23)

k VF

1 d2 w
=

>

2

q12 dz 2 z
4 VF k + VF k 2

where

(24)

F
dk
= k +
+
n
V
dT T A
then (nA , T ) is an asymptotically stable equilibrium of the closed-loop dynamics.


If
F 2
F
2
k +
k < 0
4
V
V
then the closed-loop system has no asymptotically stable equilibrium.


(25)

PROOF. The closed-loop system (6) has an asymptotically stable equilibrium at the desired set-point (nA , T ) if and only if (nA , T ) is a local minimum
of Pd (nA , T ). Lemmas 4 and 5 combined with (23) ensure that (nA , T ) is
such that Pd (nA , T ) = 0. To have a minimum the Hessian matrix H (Pd )
of the function Pd (nA , T ) must be positive definite at x = (nA , T ).

H (Pd ) (x ) =

VF

VF v (T )
13

VF

v (T )
q12
H22

with H22 = kV2 k 4 VF v (T )2 . If (25) is fulfilled, then the


Hessian matrix H (Pd ) cannot be positive definite [8]. In the other cases
must fulfill the condition (24) so that the Hessian matrix H (Pd ) is definite
positive [8].
2

The simplest function w (z) that fulfills the above proposition is the following
one:
Pa (nA , T ) = w (z)
h

= |q12 | (z z )2 uc (Tw T ) (z z )

The final expression of the control input is finally deduced from (8):
uc (nA , T ) (Tw T ) =

nA

nA

Z T
T

v ( ) d + u (Tw T )

(26)

The control law has two parameters to be chosen: and . Note that the
possible values are constrained by (24) and (25). Finally let us mention that
the control action u (nA , T ) acts in (18) via the term

g (nA , T ) uc (nA , T ) = 


R
T

nA nA T v ( ) d + u (Tw T )
which does not depend on the difference (Tw T ) anymore. This means that
the actual control input is the transferred heat.

4.3

Considerations on local and global convergence

The controller designed in the previous section only guarantees a local convergence to the desired equilibrium point, i.e. the system will converge to desired
point if and only if the initial conditions are close enough to it. Global convergence is ensured if and only if the shaped potential function Pd (nA , T ) does
not have other local minima, i.e. if and only if x is a global minimum of the
function Pd (nA , T ).
By analogy with (21) let us define the function neq
A (T ) by the following relation:
F in
C V
eq
V A
nA (T ) = F
+ k (T )
V
14

Let us also consider the following function:


(T ) =


F
F  in
CA V neq
(Tin T )
A (T ) +
V
V

+ uc (Tw T ) neq
A (T ) nA +

ZT

v ( ) d (27)

{z

z (neq
A (T ),T )z

#
Lemma 7 A state n#
is an equilibrium state of the closed loop system
A, T
(18) if and only if it is contained in the following set:

eq
#
= T # = 0, n#
A = nA T

o

#
PROOF. To prove the only-if part, let n#
be an equilibrium state
A, T
of the closed loop system. Hence we have:

dnA
=0=
dt
dT
=0=
dt

F
V
F
V

#
CAin V n#
n#
A k T
A

T in T # + k T # n#
A

nA

n#
A

Z T#
T

(28)

v ( ) d

+ uc (Tw T )

The first equation can be rewritten as nA = neq


A (T ). When replacing this
expression
into the second equation of (28), the right-hand term is equal to


#
T .

#
that is contained in .
To prove the if part, let us consider a state n#
A, T

#
By replacing nA by the expression of neq
in the first equation of (28), it
A T

is obvious that
dynamics:

dnA
dt

#
= 0 if n#
. Let us now consider the temperature
A, T




dT
F  in
=
T T # + k T # n#
A
dt
V 


#
+ uc n#
Tw T #
A, T





F  in
#
=
T
T T # + k T # neq
A
V
!

nA

neq
A

+ uc (Tw T )

15

Z T#
T

v ( ) d

The right-hand term is equal to T # and consequently, we have

dT
dt

= 0.

The function (T ) is linear in the parameter . Therefore it can be written


as follows:
(T ) = 0 (T ) + (T )
As it can be deduced from the work in [10] the existence of the three open-loop
equilibria implies that 0 (T ) is decreasing for low and high temperatures, but
increasing on one determined interval (see also [8]). As a consequence, (T )
can have several zeros, depending on the term (T ).
Proposition 8 If the parameter fulfills the following condition:
T R+ : 1 >


(T ) k 2 (T )


4 VF k (T ) +

F
V

(29)

2

dk
where (T ) = k (T ) + VF + dT
neq
A (T ), then there exists a lower bound

on such that (T ) has only one zero.

PROOF. The derivative of (T ) with respect to


d
d

 

ZT

is given as follows:

k 2 ( ) ( ) + 4 VF k ( ) +
k 2 ( )

F
V

F
V

2

+ k ( )

If the numerator is positive, i.e. if (29) is fulfilled, then (T ) is of the sign


of (T T ). As a consequence for > 0, we have (T ) > 0 (T ) (resp.
(T ) < 0 (T )) for T < T (resp. T > T ). As it can be seen in Figure 2 if
the value of is sufficiently high, then (T ) has only one zero: the desired
closed-loop equilibrium.

Proposition 9 If (29) is fulfilled and


T R+ :

F
(T ) k 2

V


>

k 2 (T ) + 4 VF k + VF

(30)

then the controller (26) achieves global stability of the desired equilibrium
point.

PROOF. (29) and (30) imply (25) and (24), respectively. Using Proposition 6
the desired equilibrium is locally stable.
16

(T ) [J/min]

increasing

0 (T )
desired set point

T [K]
Fig. 2. Influence of

on (T )

Moreover, the derivative of (T ) is given as follows:


F
(T ) k (T )2
d
V


=
dT
k (T )2 k (T ) + VF

k (T )2 (T ) + 4 VF k +


k (T )2 k (T ) +

F
V

F
V



< 0 for all T > 0. (T ) is monotonConditions (29) and (30) imply that d
dT
ically increasing and hence has a unique zero, namely the desired closed-loop
equilibrium.
Remark 10 The conditions for the global convergence (29) and (30) are very
similar to the ones for the local convergence (25) and (24), respectively, but
instead of being satisfied only at the desired set point temperature T , they
have to be satisfied at any T .
The shape of the function Pd (nA , T ) is shown in Figures 3 and 4. In the
first case, the parameters and were chosen such that local convergence is
achieved (i.e. (24) is fulfilled but not (30)). The function Pd (nA , T ) has two
local minima and consequently the convergence to the desired state is only
local. In the second case the parameters and were chosen such that global
convergence is achieved (i.e. (30) is fulfilled). The function Pd (nA , T ) has a
single minimum and the convergence to the desired state is global.
Remark 11 Even if the conditions of Proposition 9 are satisfied, global convergence can be lost if the input is bounded. Local convergence will however be
kept. It would be interesting to discuss the size of the attraction region with the
proposed controller in case of input constraints compared to other controller
proposed in the literature (e.g. in [24]).

17

nA [mol]
other convergence point

70

#
(n#
A, T )

60

50

40

locus of possible
closed-loop equilibria

30

20

desired set point (nA , T )


10

390

400

410

420

430

440

450

460

470

480

490

500

T [K]

Fig. 3. Level curves of Pd (nA , T ) (local convergence)


80

nA [mol]

70

60

locus of possible
closed-loop equilibria

50

40

30

20

10

0
420

desired set point (nA , T )


440

460

480

500

520

540

560

T [K]

Fig. 4. Level curves of Pd (nA , T ) (global convergence)

Extension to more complex systems

The power-shaping approach has given interesting results on the case study
of a CSTR with an irreversible reaction. It is therefore of major interest to
see if this approach can be extended to more complex systems, and more
particularly to systems with more than one reacting chemical species and/or
with multiple reactions.
We shall now look at some particular cases of more complex reactions to get
18

a first intuition of the existence of non-singular and negative semi-definite


solutions of (17a) and (17b). The solution of (17a) and (17b) in the case of
parallel reactions, consecutive reactions and one reaction with two reactants is
given in [8][9]. In the three cases it has been assumed that the kinetic functions
(l)
m (nm ) can be expressed as powers of nm , i.e.:
(l)

m
(l)
m (nm ) = (nm )

It has been shown that in some cases (17a) and (17b) have no solution. This
does not mean that the dynamics cannot be put into the Brayton-Moser form.
But if the Brayton-Moser form exists, then the relation (14) is not fulfilled and
the partial differential equation (13) has to be solved. In the other cases a general solution has been found, but the obtained matrix may have a non-negative
symmetric part. Therefore the corresponding potential functions cannot be
used in general as a Lyapunov functions for the open-loop system because
they may be non-decreasing along the system trajectories. As shown in [12],
given a matrix Q (x) that fulfills (4) other solutions to (4) can be built departing from the first one. Applying this methodology on the matrices that have
been found could lead to another matrix Q (x) that would be negative-definite
and hence give a Lyapunov function for the open-loop system.

6.1

Some considerations on the general CSTR dynamics

General solution

As it has been explained in Section 3.1.1 the solution to (10) is quite intricate.
One of the difficulties is due to the term in VF . This difficulty can be overcome
by also considering the dilution coefficient VF as an input of the system. This
corresponds to the system studied in [22]. The Brayton-Moser formulation
of the CSTR has been found with the same assumptions as in Section 3.1.2.
However, as the authors mention, they did not manage to extend their result
to more general cases.
Contrary to [22] who used the simplified form of the dynamics (12) with mass
action law dynamics we shall use the more general form (8). Let us consider
19

a reversible reaction A
B. Using the notations of (3) we have:

f (nA , nB , U ) =

r (T, nA , nB ) V

r (T, nA , nB ) V

g (nA , nB , U ) =

CAin V

nA

CBin V nB

(nA , nB , U ) 1
V hin H

with uc =

F
V

t

Q . The partial differential system (4) becomes (for i, j =

1, 2, 3):

[ai (x) r (x) V ] =


[aj (x) r (x) V ]
xj
xi
where the following notations have been used:

(31)

ai (U, nA , nB ) = qi2 (U, nA , nB ) qi1 (U, nA , nB )




r (U, nA , nB ) = r T (U, nA , nB ) , nA , nB
and x = (nA , nB , U ). The general solution of (31) is given as follows:
a1 (x) r (x) V =
a2 (x) r (x) V =
a3 (x) r (x) V =

Z
Z

(U, nA , nB ) dnB + (U, nA )


(U, nA , nB ) dnA + (U, nB )

Z
d
d
dnA dnB +
dnA
dU
dU
Z

dnB + (U )
+
U
Z Z

where (U, nA , nB ), (U, nA ), (U, nB ) and (U ) are real scalar functions.


The corresponding matrix Q (nA , nB , U ) and the potential function P (x) are
given respectively by:

Q (x) =

q12 (x) a1 (x) q12 (x) q13 (x)

q22 (x) a2 (x) q22 (x) q23 (x)

q32 (x) a3 (x) q32 (x) q33 (x)

P (nA , nB , U ) =

ZZ

(nA , nB , U ) dnA dnB +


+

20

(nA , U ) dnA

(nB , U ) dnB +

(U ) dU +

where is some real constant. The functions qij (x) can be chosen arbitrarily as
far as Q (x) is non-singular and (2) is fulfilled. Finally let us mention that the
solution to the partial differential equation system has been obtained without
restricting to a particular subset of solutions, contrary to what has been done
in the previous sections.

6.2

Particular solution

Let us choose the following particular solution:


(nA , nB , U ) =

2 S
nA nB

=0

(U, nA ) = 0

(U, nB ) = 0

(U ) = 0

q32 = 0

q13 = 0

q23 = 0

where S (nA , nB , U ) is the entropy function. The potential P (x) is the opposite of the entropy function S (nA , nB , U ) with the following matrix:

Q (x) =

B
T rV

T rV





4
B 4

2
T rV
T rV
T rV

1
T rV

0
0

4(1)T rV

(32)

where 0 < 1 and 1 are two real constants and = B A . i is the


S
chemical potential of the species i with Ti = n
. Following thermodynamics
i
6
the reaction kinetics must fulfill the condition :
sign [
r (x) V ] = sign [A B ]
Using this property it can be seen that Q (x) in (32) fulfills (2). The dynamics
can be written as follows:
dx
= Q1 (x) S
dt

(33)

Using (2) it is possible to see that S (nA , nB , U ) is always decreasing along


the systems trajectories. This is in accordance with the second principle of
thermodynamics which claims that the entropy production is always positive.
Indeed the open-loop system uc = 0 is an isolated system and the entropy
variation is equal to the entropy production.
6

i.e. the reaction evolves in the direction of decreasing affinity A B .

21

The open-loop equilibria x = n


A, n
B , U are the set of points such that
A (
x) = B (
x). There are an infinity of points x contained in this set. The
open-loop equilibria are locally stable but not asymptotically stable. Indeed
let us consider an initial point (n0A , n0B , U 0 ). From (8) it can be seen easily that
along the trajectories we have:
nA + nB = n0A + n0B

and U = U 0

(34)

Using (34) as the invariant set, S (x) as a Lyapunov function candidate and
(33), the stability is deduced from La Salles theorem [16].
However the power-shaping methodology cannot be applied using the above
proposed matrix Q (x) and potential S (x). Indeed the matrix Q (x) is not
defined at the open-loop equilibria and Q1 (x) is singular at these points.
Contrary to what has been shown in Section 4 the gradient of the potential
function is not zero at the open-loop equilibria.
Remark 12 Note that the availability function proposed by Alonso, Ydstie
and coworkers (e.g. [1]) can also be used as a potential function for the BraytonMoser form of the dynamics:

ref

ref

P (x) = S ref S (x) TAref TBref

1
T ref

nref nA
 A

ref
nB

nB

U ref U

ref
ref
is a reference state and S ref = S xref . The
where xref = nref
A , nB , U
corresponding matrix Q (x) is very similar to (32) and is also not defined at
the open-loop equilibria.

Conclusion

In this paper we apply the power-shaping approach to the control of a CSTR.


This approach requires to write the system dynamics in the Brayton-Moser
form. This particular form of the dynamics also allows to find a Lyapunov
function for the open-loop system. However the dynamics an be written in the
Brayton-Moser form only if a partial differential equation system is solved. We
have given the form of this partial differential system in the case of a CSTR
with general kinetics and we have proposed a methodology for finding some
of its solutions. This methodology has been successfully applied to a CSTR
with irreversible kinetics with one reactant. The Lyapunov function for the
open-loop CSTR that has been found using the Brayton-Moser formulation
of the dynamics has been related to the physical phenomena. We have also
22

shown that this methodology needs to be extended in the case of more complex
reaction kinetics. Finally first results for applying the power-shaping approach
in the case of a general CSTR with two control inputs have been presented.
These results show how the power-shaping approach can be related to the
entropy of the system. The work presented in [12] offers promising tools to
extend these results, by instance for finding alternative potential functions
departing from the proposed one.

Acknowledgements
This paper presents research results of the Belgian Network DYSCO (Dynamical
Systems, Control, and Optimization), funded by the Interuniversity Attraction Poles
Programme, initiated by the Belgian State, Science Policy Office. The scientific
responsibility rests with its authors. The work of A. Favache is funded by a grant
of aspirant of Fonds National de la Recherche Scientifique (Belgium).

References
[1] A.A. Alonso, B.E. Ydstie, and J.R. Banga. From irreversible thermodynamics
to a robust control theory for distributed process systems. Journal of Process
Control, 12:507517, 2002.
[2] R. Aris and N.R. Amundson. An analysis of chemical reactor stability and
controlI : The possibility of local control, with perfect or imperfect control
mechanisms. Chemical Engineering Science, 7(3):121131, 1958.
[3] R.K. Brayton and J.K. Moser. A theory of nonlinear networks I. Quaterly of
applied mathematics, 22:133, 1964.
[4] M. Dalsmo and A.J. van der Schaft. On representations and integrability of
mathematical structures in energy-conserving physical systems. Journal on
Control and Optimization, 37(1):5491, 1998.
[5] D. Eberard.
Extension des syst`emes hamiltoniens ports aux syst`emes
irreversibles: une approche par la geometrie de contact. PhD thesis, Universite
Claude Bernard, Lyon 1, 2006.
[6] D. Eberard, B.M. Maschke, and A.J. van der Schaft. Port contact systems
for irreversible thermodynamical systems. In Proc. 44th IEEE Conference
on Decision and Control and European Control Conference, Sevilla (Spain),
December 2005.
[7] D. Eberard, B.M. Maschke, and A.J. van der Schaft. Energy-conserving
formulation of RLC-circuits with linear resistors. In Proc. 17th International

23

Symposium on Mathematical Theory of Networks and Systems, Kyoto (Japan),


July 2006.
[8] A. Favache. Thermodynamics and Process Control. PhD thesis, Universite
catholique de Louvain, 2009.
[9] A. Favache and D.Dochain. Power-shaping control of an exothermic continuous
stirred tank reactor (CSTR). International Symposium on Advanced Control of
Chemical Processes (ADCHEM), Koc (Turkey), pages 101110, 2009.
[10] A. Favache and D. Dochain. Thermodynamics and chemical systems stability:
The CSTR case study revisited. Journal of Process Control, 2008.
[11] A. Favache, B. Maschke, and D. Dochain. Contact structures: application
to interconnected thermodynamical systems. In Proc. European Control
Conference 2007, Kos (Greece), July 2007.
[12] E. Garcia-Canseco, D. Jeltsema, J.M.A. Scherpen, and R. Ortega. Powerbased control of physical systems: two case studies. In Proc. 17th IFAC World
Congress, Seoul (Korea), 2008.
[13] D. Jeltsema, R. Ortega, and J. M.A. Scherpen. An energy-balancing perspective
of interconnection and damping assignment control of nonlinear systems.
Automatica, 40(9):16431646, September 2004.
[14] D. Jeltsema and J.M.A. Scherpen. On mechanical mixed potential, content and
co-content. In Proc. European Control Conference, pages 7378, 2003.
[15] D. Jeltsema and J.M.A. Scherpen. A power-based description of standard
mechanical systems. Systems & Control Letters, 56(5):349356, May 2007.
[16] H. Khalil. Nonlinear Systems. Prentice Hall, Upper Saddle River, 3rd edition,
2002.
[17] B.M. Maschke, R. Ortega, and A.J. van der Schaft. Energy-based Lyapunov
functions for forced Hamiltonian systems with dissipation. IEEE Transactions
on Automatic Control, 45(8):14981502, 2000.
[18] R. Ortega, D. Jeltsema, and J.M.A. Scherpen. Power shaping: A new paradigm
for stabilization of nonlinear RLC-circuits. IEEE Transactions on Automatic
Control, 48(10):17621767, october 2003.
[19] R. Ortega, A.J. van der Schaft, I. Mareels, and B. Maschke. Putting energy
back in control. IEEE Control Systems Magazine, 21(2):1833, April 2001.
[20] R. Ortega, A.J. van der Schaft, B.M. Maschke, and G.Escobar. Interconnection
and
damping
assignment:
passivity-based control of port-controlled Hamiltonian systems. Automatica,
38(4):585596, 2002.
[21] I. Otero-Muras, G. Szederknyi, A.A. Alonso, and K.M. Hangos. Dynamic
analysis and control of chemical and biochemical reaction networks. In Proc.
International Symposium on Advanced Control of Chemical Processes, pages
165170, Gramado (Brazil), April 2006.

24

[22] H. Ramrez, D. Sbarbaro, and R. Ortega. On the control of non-linear processes:


an IDA-PBC approach. Journal of Process Control, 19:405414, 2009.
[23] A. Uppal, W. H. Ray, and A. B. Poore. On the dynamic behavior of continuous
stirred tank reactors. Chemical Engineering Science, 29(4):967985, April 1974.
[24] F. Viel, F. Jadot, and G. Bastin. Global stabilization of exothermic chemical
reactors under input constraints. Automatica, 33:14371448, 1997.

25

You might also like