You are on page 1of 14

Extremum Seeking Control: Convergence Analysis∗

Dragan Nešić

Abstract— This paper summarizes our recent work on dy- extremum seeking as one of the most promising adaptive
namical properties for a class of extremum seeking (ES) control methods [1, Section 13.3].
controllers that have attracted a great deal of research attention There are two main approaches to extremum seeking: (i)
in the past decade. Their local stability properties were already
investigated, see [2]. We first show that semi-global practical adaptive control extremum seeking; (ii) nonlinear program-
convergence is possible if the controller parameters are care- ming based extremum seeking. Adaptive control methods
fully tuned and the objective function has a unique (global) provide a range of adaptive controllers that solve the ex-
extremum. An interesting tradeoff between the convergence tremum seeking problem for a large class of systems [2].
rate and the size of the domain of attraction of the scheme The controller makes use of a certain excitation (dither)
is uncovered: the larger the domain of attraction, the slower
the convergence of the algorithm. The amplitude, frequency and signal which provides the desired sub-optimal behaviour if
shape of the dither signal are important design parameters in the controller parameters are tuned appropriately. On the
the extremum seeking controller. In particular, we show that other hand, nonlinear programming based extremum seek-
changing the amplitude of the dither adaptively can be used to ing methods combine the classical nonlinear programming
deal with global extremum seeking in presence of local extrema. methods for numerical optimization with an approximate on-
Moreover, we show that the convergence of the algorithm is
proportional to the power of the dither signal. Consequently, line generation of the gradient of the objective function by
the square-wave dither yields the fastest convergence among applying constant probing inputs successively [20].
all dithers of the same frequency and amplitude. We consider The main goal of this paper is to report on our recent
extremum seeking of a class of bioprocesses to demonstrate our results on stability properties of a class of adaptive extremum
results and motivate some open research questions for multi- seeking controllers. The first local stability analysis of this
valued objective functions.
class of controllers was reported in 2000 by Krstić and Wang
I. I NTRODUCTION [10]. This seminal paper used techniques of averaging and
singular perturbations to show that if the adaptive extremum
In many engineering applications the system needs to seeking controller is tuned appropriately, then sub-optimal
operate close to an extremum of a given objective (cost) func- extremum seeking is achieved if the system is initialized
tion during its steady-state operation. Moreover, the objective close to the extremum.
function is often not available analytically to the designer but We introduced a simplified adaptive scheme in [17] where
instead one can measure the value of the objective function it was shown under slightly stronger conditions that non-
by probing the system. local (even semi-global) extremum seeking is achieved if
Extremum seeking is an optimal control approach that the controller is tuned appropriately. Moreover, by using the
deals with situations when the plant model and/or the cost to singular perturbations techniques and averaging, we demon-
optimize are not available to the designer but it is assumed strated that this simplified scheme operates on average in its
that measurements of plant input and output signals are slow time scale as the steepest descent optimization scheme.
available. Using these available signals, the goal is to design We reported a detailed analysis of this simplified scheme
an extremum seeking controller that dynamically searches for in [17]. In [19] we analysed the flexibility in choosing
the optimizing inputs. This situation arises in a range of clas- the shape of the excitation dither signal to ensure faster
sical, as well as certain emerging, engineering applications. convergence. It was shown for static maps that a square wave
Indeed, this method was successfully applied to biochemical dither yields fastest convergence over all dither signals with
reactors [9], [4], ABS control in automotive brakes [8], the same amplitude and frequency. We reported conditions
variable cam timing engine operation [14], electromechanical that ensure global extremum seeking in the presence of
valves [13], axial compressors [21], mobile robots [11], local extrema in [18]. Adaptive schemes with multi-valued
mobile sensor networks [5], [12], optical fibre amplifiers objective functions that arise, for instance, in bioprocesses,
[7] and so on [2]. A good survey of the literature on this were investigated in [4]. Multi-valued functions pose some
topic prior to 1980 can be found in [16] and a more recent open research questions that we briefly mention in the last
overview can be found in [2]. Åström and Wittenmark rated section. In the sequel, we present an overview of our recent
∗ This work was supported by the Australian Research Council under the
results in [4], [17], [18], [19].
Discovery Grants and Australian Professorial Fellow schemes. The author Mathematical preliminaries: We denote the set of real
would like to thank Y. Tan, I. M. Y. Mareels and G. Bastin for the fruitful numbers as R. Given a sufficiently smooth function h :
collaboration that has lead to this work. Rp → R, we denote its ith derivative with respect to j th
D. Nešić is with Department of Electrical and Electronic Engineer-
ing, University of Melbourne, Parkville, 3052, Victoria, Australia. E-mail: variable as Dji h(x1 , . . . , xp ). When i = 1 and j = 1 we write
d.nesic@ee.unimelb.edu.au simply Dh(x1 , . . . , xp ) := D11 (x1 , . . . , xp ). The continuous
function β : R≥0 × R≥0 → R≥0 is of class KL if it is where θ ∈ R is a scalar parameter. The closed-loop system
nondecreasing in its first argument and converging to zero (3) with (4) is then
in its second argument. Given a measurable function x, we
define its L∞ norm k · k = ess supt≥0 |x(t)|. ẋ = f (x, α(x, θ)). (5)
We will show in the next section that the closed loop
systems with an adaptive extremum seeking controller can The requirement that θ is scalar and that (3), (4) is SISO is to
be written as a parameterized family of systems: simplify presentation. Multidimensional parameter situations
can be tackled, see [2].
ẋ = f (t, x, ε) , (1) We proposed in [17] a first order extremum seeking
scheme (see Figure1) that yields the following closed loop
where x ∈ Rn , t ∈ R≥0 and ε ∈ R`>0 are respectively dynamics:
the state of the system, the time variable and the parameter
vector. The stability of the system (1) can depend in an ẋ = f (x, α(x, θ̂ + a sin(ωt))) (6)
intricate way on the parameter ε and we will need the ˙
following definition (see [17] for motivating examples): θ̂ = kh(x)b sin(ωt), (7)

where (k, a, b, ω) are tuning parameters. Compared with the


Definition 1 The system (1) with parameter ε is said to extremum seeking scheme in [10], the proposed extremum
be semi-globally practically asymptotically (SPA) stable, seeking scheme in Figure 1 is simpler, containing only an
uniformly in (ε1 , . . . , εj ), j ∈ {1, . . . , `}, if there exists integrator (without low-pass and high-pass filters that are
β ∈ KL such that the following holds. For each pair used in [10]).
of strictly positive real numbers (∆, ν), there exist real
numbers ε∗k = ε∗k (∆, ν) > 0, k = 1, 2, . . . , j and for
each fixed εk ∈ (0, ε∗k ), k = 1, 2, . . . , j there exist εi =
εi (ε1 , ε2 , . . . , εi−1 , ∆, ν), with i = j + 1, j + 2, . . . , `, such
that the solutions of (1) with the so constructed parameters
ε = (ε1 , . . . , ε` ) satisfy: θ x& = f ( x , α( x , θ )) y
y = h( x )
|x(t)| ≤ β(|x0 |, (ε1 · ε2 · · · · · ε` )(t − t0 )) + ν, (2)

for all t ≥ t0 ≥ 0, x(t0 ) = x0 with |x0 | ≤ ∆. If we have that


j = `, then we say that the system is SPA stable, uniformly
in ε. θˆ
k x
+
Note that in Definition 1 we can construct a small “box” s
around the origin for the parameters εk , k = 1, 2, . . . , j so
that the stability property holds uniformly for all parameters
in this box, whereas at the same time we can not do so for a sin(ωt ) b sin (ω t )
the parameters εk , k = j + 1, . . . , l. Sometimes we abuse
terminology and refer to (ε1 · · · ε` ) in the estimate (2) as the
“convergence speed” (although the real convergence speed
depends also on the function β).
Fig. 1. A first order extremum seeking controller
II. A N ADAPTIVE CONTROL SCHEME

Consider the system:


III. G LOBAL EXTREMUM SEEKING IN ABSENCE OF
ẋ = f (x, u), y = h(x) , (3)
LOCAL EXTREMA
where f : Rn × R → Rn and h : Rn → R are sufficiently
In this section we summarize results from [17] that provide
smooth. x is the measured state, u is the input and y
guarantees for SPA stability under the following assump-
is the output. We suppose that there exists a unique x∗
tions:
such that y ∗ = h(x∗ ) is the extremum of the map h(·).
Due to uncertainty, we assume that neither x∗ nor h(·) is
precisely known to the control designer. The main objective Assumption 1 There exists a function l : R → Rn such that
in extremum seeking control is to force the solutions of the f (x, α(x, θ)) = 0 if and only if x = l(θ).
closed loop system to eventually converge to x∗ and to do
so without any precise knowledge about x∗ or h(·). Assumption 2 For each θ ∈ R, the equilibrium x = l(θ) of
Consider a family of control laws of the following form: (5) is globally asymptotically stable, uniformly in θ.
u = α(x, θ), (4)
Assumption 3 Denoting Q(·) = h ◦ l(·), there exists a parameter a2 · k affects the convergence speed. The smaller
unique θ∗ maximizing Q(·) and, the following holds1 : a2 · k, the slower the convergence and the larger the domain
of attraction.
DQ(θ∗ ) = 0 D2 Q(θ∗ ) < 0 (8)
Note that since h(·) is continuous, then for any ν > 0,

DQ(θ + ζ)ζ < 0 ∀ζ 6= 0 . (9) there exists ν1 > 0 such that
Note that (9) in Assumption 3 guarantees that there do not |x̃| ≤ ν1 =⇒ |h(x̃ + x∗ ) − y ∗ | ≤ ν . (14)
exist any local extrema. We will consider the case with local
extrema in the next section. Theorem 1 can be interpreted as follows. For any (∆, ν)
Introduce the change of the coordinates, x̃ = x − x∗ , we can adjust ε so that for all |z| ≤ ∆ we have that
θ̃ = θ̂ − θ∗ and the system takes the form: lim sup |y(t) − y ∗ | ≤ ν. In other words, the output of the
t→∞
system can be regulated arbitrarily close to the extremum
x̃˙ = ∗ ∗ ∗
f (x̃ + x , α(x̃ + x , θ̃ + θ + a sin(ωt)))
value y ∗ from an arbitrarily large set of initial conditions
˙
θ̃ = kh(x̃ + x∗ )b sin(ωt). (10) by adjusting the parameters ε in the controller. In particular,
the parameters ε are chosen so that Definition 1 holds with
Note that the point (x∗ , θ∗ ) is in general not an equilibrium
4 4 (∆, ν1 ) and ν1 is defined in (14).
point of the system (6), (7). We introduce k = ωδK, σ = ωt, Theorem 1 is a stronger result than [10, Theorem 1] since
where ω and δ are small parameters and K > 0 is fixed. The we prove SPA stability, as opposed to local stability in [10].
system equations expanded in time σ are: However, our results are stated under stronger assumptions
dx̃ (Assumptions 1-3) than those in [10]. Assumptions 1-3
ω = f (x̃ + x∗ , α(x̃ + x∗ , θ̃ + θ∗ + a sin(σ)))
dσ appear to be natural when non-local stability is investigated.
dθ̃ Moreover, we note that it is not crucial in Assumptions 1 –
= δKh(x̃ + x∗ )b sin(σ). (11) 3 that all conditions hold globally. For instance, instead of

requiring (9) in Assumption 3, we can assume:
The system (11) has the form (1) where the parameter vector
is defined as ε := [a b δ ω]T . For simplicity of presentation DQ(θ∗ + ζ)ζ < 0 ∀ζ ∈ D, ζ 6= 0 , (15)
in the sequel we let b = a and
where D is a bounded neighborhood of θ∗ . We note that
ε := [a2 δ ω]T . (12) these conditions are not very restrictive, whereas their global
The system (11) has a two-time-scale structure and our first version is (Assumptions 2 and 3). Indeed, if the maximum
main result is proved by applying the singular perturbations is isolated and all functions are sufficiently smooth, we can
and averaging methods (see [17]). conclude that the condition (8) implies that there exists a
set D satisfying (15). Similarly, we could assume only local
Theorem 1 Suppose that Assumptions 1, 2 and 3 hold. Then, stability in Assumption 2. If all of our assumptions were
the system (10) is SPA stable, uniformly in (a2 , δ). regional (as opposed to global) we could still state SPA
 stability with respect to the given bounded region.
The proof of Theorem 1 in [17] provides an interesting
Theorem 1 provides the parameter tuning guidelines since it insight into the way the extremum seeking controller oper-
shows that to achieve a certain domain of attraction one first ates. The parameter ω is used to separate time scales be-
needs to reduce a and δ sufficiently and then for fixed values tween the plant (boundary layer) and the extremum seeking
of these parameters reduce ω sufficiently. Hence, one can controller (reduced system), where the plant states are fast
achieve any given domain of attraction but the convergence and they quickly die out (Assumption 2). Using the singular
speed will be reduced simultaneously (cf. Definition 1). This perturbation method, we obtain that the reduced system in
tradeoff was first observed in [17]. In the convergence speed the variable “θr ” in time “σ = ωt” is time varying and it
analysis of the extremum seeking scheme, the “worst case” has the form:
convergence speed is considered. That is, the convergence dθr
speed of the overall system depends on the convergence = KaδQ(θ∗ + θr + a sin(σ)) sin(σ) , (16)

speed of the slowest sub-system. The first order extremum
for which we introduce an “averaged” system:
seeking controller (10), according to Theorem 1, yields the
following stability bound: dθav K
= a2 δ · DQ(θ∗ + θav ) . (17)
dσ 2
|z(t)| ≤ β(|z(t0 )|, (a2 δω)(t − t0 )) + ν,
  Hence, the averaged system (17) can be regarded as the “gra-
2 (t − t0 )
= β |z(t0 )|, (a k) + ν, (13) dient system” whose globally asymptotically stable equilib-
K rium θ∗ corresponds to the global maximum of the unknown
4
for all t ≥ t0 ≥ 0 and |z(t0 )| ≤ ∆, where z = (x̃T , θ̃T )T map Q (Assumption 3).
and k, K were defined before. Since K > 0 is fixed, the As already indicated, the first order extremum seeking
scheme works on average as a “gradient search” method.
1 Without loss of generality we assume that the extremum is a maximum. Both the excitation signal and the integrator are necessary
to achieve this. The excitation signal a sin(ωt) is added to through the decrease of γε2 as γε2 decreases to zero as a
system (3) to get probing while the multiplication (modula- decreases. Moreover, it is sometimes possible to achieve this
tion) of output and the excitation signal extracts the gradient in a manner that will guarantee SPA stability, uniform in
of the unknown mapping Q(·). The role of the integrator is ε, see [17]. A condition that is needed for this to hold is
to get on average the steepest decent along the gradient of summarized in the next assumption.
Q(·). Hence, the first order scheme is the simplest controller
structure that achieves extremum seeking. Assumption 5 Let the gains γε1 , γε2 come from Propositions
Note that we did not prove SPA stability, uniform in the 1 and 2. Assume that there exists γ ∈ K∞ such that for
whole vector ε in Theorem 1 for system (10) with parameter any 0 < s1 < s2 there exist ω ∗ and a∗ such that for all
ε. To prove such a result we need a stronger assumption: ω ∈ (0, ω ∗ ), a ∈ (0, a∗ ) and s ∈ [s1 , s2 ] we have that the
following small gain conditions hold:
Assumption 4 We have that (8) holds and there exists αQ ∈
γε1 ◦ γε2 (s) ≤ γ(s) < s, γε2 ◦ γε1 (s) ≤ γ(s) < s. (22)
K∞ such that DQ(θ∗ + ζ)ζ ≤ −αQ (|ζ|) for all ζ ∈ R.
The following result was proved in [17]:
Next we will use Assumptions 4 and 5 that will be stated
below after some auxiliary results are presented to prove
Theorem 2 Suppose that Assumptions 1, 2, 4 and 5 hold.
semi-global stability results uniform in the parameter ε.
Then, the closed-loop system (10) is SPA stable, uniformly
We introduce “boundary layer” using x̄ := x̃ − l(θ∗ + θ̃ +
in ε = (a2 , δ, ω). 
a sin(σ)) =: x̃ − l̃(θ̃ + a sin(σ)) and rewrite (11) in the time
scale “t” as follows: Note that the conditions (22) do not imply each other,
e1
x̄˙ = f̃ (x̄ + l̃(θ̃ + a sin(σ)), θ̃ + a sin(σ)) + ωa∆ (18) as can be easily seen from the case when γ2θ ◦ γ1z (s) =
sq , q > 1 and γ2z ◦ γ1θ (s) = sp , p > 1 in which case the
˙ e 2, q
θ̃ = aδK h̃ ◦ l̃(θ̃ + a sin(σ)) sin(σ) + aδ ∆ (19) conditions (22) becomerespectively γ1θ ωaq−1 (γ2z (s)) <
z 1−p θ p
where f̃ (x, θ) := f (x+x∗ , α(x+x∗ , θ∗ +θ)), h̃(x) := h(x+ s and γ1 ωa γ2 (s) < s. It is obvious, that in the first
x∗ ) and ∆e 1, ∆
e 2 are appropriate functions that depend on the case we can choose ω and a independent of each other so
state variables, parameters and time (see [17]). We consider that the first condition in (22) holds, whereas it is impossible
the system (18), (19) as a feedback connection of two to do so for the second condition in (22). This also illustrates
systems. It was shown in [17] that under our assumptions, the that conditions of Assumption 5 may not hold for some gains.
two systems are input to state stable (ISS) in an appropriate If all the gains γ1θ , γ2θ , γ1z , γ2z are linear then Assumption 5
sense (see [15]): holds. In particular, the small gain conditions (22) become
independent of a and can be achieved by reducing ω only.
Proposition 1 Suppose that Assumption 4 holds. Then, there It is an open question whether there is a genuine gap
exist β1 ∈ KL and for any ∆1 > ν1 > 0 and ω ∗ > 0 there between Theorems 1 and 2, that is whether there exists
exist γ1θ , γ2θ ∈ K∞ , a∗ > 0 and δ ∗ > 0 such that for all a ∈ an example satisfying all conditions of Theorem 1 but not
(0, a∗ ), δ ∈ (0, δ ∗ ), ω ∈ (0, ω ∗ ) and max{|θ̃0 |, ||x̄||} ≤ ∆1 , conditions of Theorem 2 that is not SPA stable uniformly in
with θ̃0 := θ̃(t0 ) we have that the solutions of the subsystem ε.
(19) satisfy:
n o Example 1 Consider the system:
|θ̃(t)| ≤ max β1 (|θ̃0 |, (a2 δω)(t − t0 )), γε1 (||x̄||), ν1 , (20)
ẋ = −x + u2 + 4u; y = −(x + 4)2 . (23)

for all t ≥ t0 ≥ 0, where γε1 (s) := γ1θ a1 γ2θ (s) .  Let the control input be u = θ. Then, using our notation we
have θ∗ = −2, x∗ = −4 and y ∗ = 0. We choose the initial
Proposition 2 Suppose that Assumptions 1 and 2 hold. value x(0) = 2 that is far away from the optimal value
Then, there exist β2 ∈ KL and for any positive ∆2 , ν2 , x∗ = −4. Let θ̂(0) = 0. We present simulations for various
a∗ and δ ∗ there exist γ1z , γ2z ∈ K∞ and ω ∗ > 0, such values of a, δ, ω to illustrate that the speed of convergence,
that for all a ∈ (0, a∗ ), δ ∈ (0, δ ∗ ) ω ∈ (0, ω ∗ ) and the domain of attraction and the accuracy of the algorithm
max{|x̄0 |, ||θ̃||} ≤ ∆, with x̄0 := x̄(t0 ), we have that the indeed depend on the parameters as suggested by Theorems
solutions of the subsystem (18) satisfy: 1 and 2.
n o We apply the first order scheme in Figure 1. By choosing
|x̄(t)| ≤ max β2 (|x̄0 |, t − t0 ), γε2 (||θ̃||), ν2 , (21)
a = 0.3, δ = 0.5(K = 4) and ω = 0.5, the performance of
for all t ≥ t0 ≥ 0, where γε2 (s) := γ1z (ωaγ2z (s)).  the scheme is shown in Figure 2, where |z| = |(x̃, θ̃)| and x̃
and θ̃ are the same as in Equation (10). It can be seen that,
Note that the ISS gains γε1 and γε2 in Propositions 1 and the state z converges to the neighborhood of the origin. The
2 depend on a and ω. Moreover, the gain γε1 increases to output also converges to the vicinity of the extremum value.
infinity as a is reduced to zero. Typically, this behavior We increase a such that a = 0.6 while keeping δ = 0.5 and
leads to lack of stability in the interconnection. However, ω = 0.5. Increasing a will get a fast convergent speed, while
in this case, it is possible to counteract this increase of γε1 the domain of the attraction would be smaller. It can be seen
clearly from Figure 2 that, though both y(t) and |z| converge The extremum seeking feedback scheme in Figure 4 was
very fast, they converge to a much larger neighborhood of proposed in [18] to deal with this problem. One of the main
their optimal values. differences between this scheme and the one in the previous
section is that the amplitude of the excitation signal in Figure
a=0.3,δ=0.5, ω=0.5 a=0.6,δ=0.5, ω=0.5
4 is time varying, whereas in Figure 1 the amplitude is fixed.
y
0 0 The model of the system in Figure 4 can be written as
y
−0.05 −0.05

ẋ = f (x, α(x, θ̂ + a · sin(ω · t)))


−0.1 −0.1
˙
θ̂ = ω · δ · h(x) · sin(ω · t)
0 100 200 300 0 100 200 300

ȧ = −ω · δ · ε · g(a), a(0) = a0 , (25)


0.5 0.5
|z|
0.4
|z|
0.4
where g(·) is a locally Lipschitz function that is zero at zero
0.3 0.3

0.2 0.2
and positive otherwise. The simplest choice is g(a) = a.
0.1 0.1
Here ε, ω, δ and a0 are to be chosen by the designer.
0 0
0 100 200 300 0 100 200 300

θ & = f ( x, α( x, θ))
x y
Fig. 2. The performance of the simplest extremum seeking scheme
y = h( x)

θˆ ωδ
0.05
a=0.3,δ=0.25, ω=0.5
0.05
a=0.3,δ=0.75, ω=0.5 + ×
s
0 0

−0.05 −0.05 ×
−0.1
y
−0.1
y sin( ω t )
−0.15 −0.15

−0.2 −0.2
a& = −εωδ g ( a ), a ( 0) = a0
0 200 400 600 0 200 400 600

0.5 0.5
|z| |z|
0.4 0.4 Fig. 4. A global extremum seeking feedback scheme.
0.3 0.3

0.2 0.2
Denoting σ = ω · t, the equations (25) in time “σ” are:
0.1 0.1

0 0 dx
0 200 400 600 0 200 400 600 ω· dσ = f (x, α(x, θ̂ + a sin(σ)))
dθ̂
Fig. 3. The performance of the simplest extremum seeking scheme dσ = δ · h(x) · sin(σ)
da
dσ = −ε · δ · g(a), a(0) = a0 . (26)
Next, we fix a = 0.3 and ω = 0.5. First, we let δ = 0.25
and observe that the state z converges to the neighborhood The system (26) is in standard singular perturbation form,
of the origin, see Figure 3. The output also converges to the where the singular perturbation parameter is ω. To obtain
vicinity of its extremum value. Then, we increase δ to be 0.75 the fast and slow systems, we set ω = 0 and “freeze” x at
and observe that the convergence speed has increased, see its “equilibrium”, x̃ = l(θ̂+a·sin(σ)) to obtain the “reduced”
Figure 3. However, when we further increase δ to be larger system in the variables (θ, a) in the time scale σ = ω · t:
than 1.40 unstable performance is observed. dθ
dσ = δ · Q(θ + a · sin(σ)) · sin(σ) = δ · µ(σ, θ, a) (27)
IV. G LOBAL EXTREMUM SEEKING IN PRESENCE OF da
dσ = −ε · δ · g(a), a(0) = a0 , (28)
LOCAL EXTREMA
In this section, we summarize the results from [18] where where Q(·) = h ◦ l(·) is the output equilibrium map (cf.
global extremum seeking was investigated for objective Assumption 6). We can write the “averaged” system of (27)
functions that have local extrema. In particular, we use the by using:
following assumption: 1 R 2π
µav (θ, a) := µ(t, θ, a)dt, (29)
2π 0
Assumption 6 There exists a unique global maximum θ∗ ∈
R of Q(·) such that where µ(·, ·, ·) comes from (27). Indeed, using the above
definition, we can analyze the closed loop system ((27)-(28))
Q(θ∗ ) > Q(θ), ∀θ ∈ R, θ 6= θ∗ . (24)
via the auxiliary “averaged” system:
Assumption 6 implies that besides the global maximum dθ
θ∗ there may also exist local maxima θ̃∗ . It is weaker than dσ = δ · µav (θ, a) (30)
da
Assumption 3 that does not allow for local maxima. dσ = −δ · ε · g(a), a(0) = a0 > 0 . (31)
By introducing the new time τ := ε · δ · σ, we can rewrite Theorem 3 Suppose that Assumptions 1, 2, 6 and 7 hold.
the above equations as follows: Then, for any strictly positive (∆, ν) and a0 > a∗ there

exist KL functions βx , βθ and ε∗ = ε∗ (a0 , ∆, ν) > 0 and
ε· dτ = µav (θ, a) for any ε ∈ (0, ε∗ ) there exists δ ∗ = δ ∗ (ε) > 0, for any
da
dτ = −g(a), a(0) = a0 > 0 , (32) δ ∈ (0, δ ∗ (ε)) there exists ω ∗ = ω ∗ (δ) > 0 and for any
such a0 , ε, δ ∈ (0, δ ∗ ) and ω ∈ (0, ω ∗ ), we have that for all
that are in standard singular perturbation form. However,
(x(t0 ), θ̂(t0 ), a(t0 )) satisfying a(t0 ) = a0 , |θ̂(t0 ) − `(a0 )| ≤
there are several differences with the classical singular per-
∆, |x(t0 ) − l(θ̂(t0 ))| ≤ ∆ and all t ≥ t0 ≥ 0 the solutions
turbation literature. In our case the equation:
of the system (25) satisfy:
0 = µav (θ, a) (33)  
|x(t) − l(θ̂(t))| ≤ βx |x(t0 ) − l(θ̂(t0 ))|, (t − t0 ) + ν ,
may not have k isolated real roots θ = `i (a). Indeed, some  
of the real roots may only be defined for a ∈ [0, ā] and such |θ̂(t) − `(a(t))| ≤ βθ |θ̂(t0 ) − `(a(t0 ))|, ωδ(t − t0 ) + ν .
that for some i and j we have `i (a) 6= `j (a), a ∈ [0, ā) and

`i (ā) = `j (ā). Moreover, it shown in [18] that there exists a
continuous function p(θ, a) such that: Theorem 3 presents a tuning mechanism for the controller
µav (θ, a) = a · p(θ, a) (34) parameters (choice of ω, δ, ε) and its initialization (choice
of a0 ) that guarantees semi-global practical convergence to
and this means that we will be unable to prove stability of the global extremum despite the presence of local extrema.
the “boundary layer” system uniform in a that is a standard Simulations in [18] illustrate that such convergence is indeed
assumption in the singular perturbation literature. Note also achieved. We note that since the static mapping Q(·) is
that we are interested in convergence properties of this not known, it is in general not possible to check a priori
system initialized from a set of initial conditions satisfying whether Assumption 7 holds, let alone analytically compute
a(0) = a0 which is a weaker property from the standard the values of a∗ , ε∗ , δ ∗ and ω ∗ . However, our result suggests
stability properties considered in the singular perturbation that if there is some evidence that Assumptions 6 and 7 may
literature. hold, then increasing sufficiently a0 and reducing sufficiently
Another assumption that characterizes solutions of the ε, δ and ω will indeed result in global convergence. In
equation (33) is needed in the main result. practice, determining how large a0 , ε, δ, ω should be, may
have to be determined through experiments.
Assumption 7 There exists an isolated real root θ = The stability result of Theorem 3 is different from the
`(a) : R≥0 → R of the equation (33) such that: stability result in [17, Theorem 1], where the closed-loop
1) ` is continuous and D1 p(`(a), a) < 0, ∀a ≥ 0, where system is proved to be SPA stable, uniformly in the tuning
p(θ, a) is defined in (34). parameters: the amplitude of the sine wave dither a and δ.
2) There exists a > 0 such that for all a ≥ a , θ = `(a) First of all, in Theorem 3, the initial value of the amplitude
∗ ∗

is the unique real root of (33). of the dither plays an important role in the proposed ES feed-
3) `(0) = θ∗ , where θ∗ is the global extremum, which back scheme to guarantee that the output of the system (25)
comes from Assumption 6. converges to the global extremum semi-globally practically.
Such an initial value a0 is also a tuning parameter in the
Before showing the stability properties of the closed proposed ES feedback scheme. Secondly, a0 does affect the
loop system (25), the following proposition shows stability convergence speed as βθ and βx are dependent on the choice
properties of the “reduced” system (27)-(28). of a0 , though analyzing how a0 affects the convergence
speed is much more difficult than other parameters. Thirdly,

Proposition 3 Suppose that Assumptions 6 and 7 hold. Theorem 3 clearly indicates that the choice of ε depends on

Then, for any strictly positive (∆, ν) and a0 > a there the choice of a0 and the choice of δ depends on the choice


exist β = βa0 ,∆,ν ∈ KL and ε∗ = ε∗ (a0 , ∆, ν) > 0 and for of ε and the choice of ω depends on the choice of δ.
∗ ∗ ∗ A consequence of Theorem 3 is that we can tune the
any ε ∈ (0, ε ) there exists δ = δ (ε) > 0 such that for any
such a0 , ε and δ ∈ (0, δ ∗ ) we have that for all (θ(σ0 ), a(σ0 )) extremum seeking controller to achieve lim supσ→∞ |θ(σ)−
satisfying a(σ0 ) = a0 and |θ(σ0 ) − `(a0 )| ≤ ∆ and all `(a(σ))| ≤ ν from an arbitrarily large set of initial conditions
t ≥ t0 ≥ 0 the solutions of the system (27), (28) satisfy: and for arbitrarily small ν > 0. Moreover, from (28) it is
obvious that there exists βa ∈ KL with βa (s, 0) = s, such
|θ(σ) − `(a(σ))| ≤ β(|θ(σ0 ) − `(a(σ0 ))|, δ(σ − σ0 )) + ν . (35) that for all a(σ0 ) = a0 ∈ R>0 we have:

From the semi-global practical asymptotical (SPA) sta- |a(σ)| ≤ βa (a(σ0 ), ε · δ · (σ − σ0 )), ∀σ ≥ σ0 ≥ 0 , (36)
bility properties of the “reduced” system ((27)- (28)), the
stability properties of the overall system (25) are stated in and since `(·) is continuous and `(0) = θ∗ , it follows that
the following theorem.
lim `(a(σ)) = θ∗ ⇒ lim sup |θ(σ) − θ∗ | ≤ ν ,
σ→∞ σ→∞
11
which implies semi-global practical extremum seeking since
10.5
θ∗ is the global extremum of Q(·). In the time “t”, (36)

Q(θ)
10
becomes 9.5

−1 −0.5 0 0.5 1 1.5 θ


|a(t)| ≤ βa (a(t0 ), ω · ε · δ · (t − t0 )), ∀t ≥ t0 ≥ 0. (37)
5

Note that since Q(·) is continuous, then for any ν > 0, 4


l(a)
3
there exists ν1 > 0 such that

a
2
1
|θ̂| ≤ ν1 =⇒ |Q(θ̂(t)) − θ∗ | = |y(t) − y ∗ | ≤ ν . (38) 0
−1 −0.5 0 0.5 1 1.5 θ

Theorem 3 can be interpreted as follows. For any (a0 , ∆, ν1 ),


Fig. 5. A 4th -order polynomial and its bifurcation diagram for which
where a0 > a∗ , ν1 is defined in (38), we can adjust ε, Assumption 7 holds.
δ and ω appropriately
 so that for all |z(t0 )| ≤ ∆, where
4 x − l(θ̂)
z = , we have that lim sup |y(t) − y ∗ | ≤ ν. 10.5
10
θ̂ − `(a) t→∞
In other words, the output of the system can be regulated 8

arbitrarily close to the global extremum value y ∗ from an −0.5

arbitrarily large set of initial conditions by adjusting the

θ (t)
y(t)
parameters (ω, δ, ε, a0 ) in the controller. This is despite the 5

possible presence of local extrema (cf. Assumption 6).


The system model (26) suggests a three-time-scale dynam-
ics when ω, δ and ε are very small. Indeed, the solutions
−1
first converge to a small neighborhood of the set X := 0
t
5
x 10
4
0
t
5
x 10
4

{(x, θ̂) : x − l(θ̂) = 0} (fast transient) and then with the


speed proportional to ω · δ to a neighborhood of the set Fig. 6. The performance of the extremum seeking feedback scheme when
L := {(θ̂, a) : θ̂ − `(a) = 0} (medium transient) and then a = 0.1 is fixed.

with the speed propositional to ω · δ · ε to a neighborhood of


the point (θ, a) = (θ∗ , 0) (slow transient). Moreover, during
the slow transient, the solutions stay in a ν-neighborhood of All conditions in Theorem 3 hold and hence the conclusion
the set X . of Theorem 3 holds. Let g(a) = a, a0 = 3, which turns
out to be sufficiently large (see Figure 5), using the same
Example 2 Assumption 7 is crucial to prove the global parameters as above, θ̂(t) converges to a small neighborhood
convergence of the proposed scheme. From the result of of the global maximum θ∗ = 1 (the global extremum) as seen
Theorem 3, θ̂ converges to `(0). If `(0) is not the global in Figure 7. This example shows that the proposed scheme
extremum, the output of the overall system would converge to can ensure that global extremum seeking is achieved despite
a local extremum. An illustrative example is used to show that the existence of local extrema.
global extremum seeking can be achieved by the proposed
scheme in Figure 4 in the presence of local extrema. Consider 10 1
2

the following dynamic system


0.8
5
ẋ1 = −x1 + x2 , ẋ2 = x2 + u, y = h(x), (39)
θ(t)

0.6
y(t)

a(t)

8
where h(x) = −(x1 + 3x2 )4 + 15 (x1 + 3x2 )3 + 56 (x1 +
1
0

2 0.4
3x2 ) + 10. The control input is chosen as
−5
0.2

u = −x1 − 4x2 + θ . (40)


−10 0 0
0 2 4 0 2 4 0 2 4
8 3
Moreover, we have Q(θ) = −θ4 + 15 θ + 56 θ2 + 10 that t x 10
4 t
x 10
4 t x 10
4


has a global maximum at θ = 1 and a local maximum at
Fig. 7. The performance of the proposed ES feedback scheme.
θ̃∗ = −0.6 as seen in Figure 5. Hence, Assumption 6 holds.
The bifurcation diagram in Figure 5 implies that Assumption
7 also holds. First, we use the extremum seeking scheme from
Section III, in which the amplitude of the sinusoidal signal V. D ITHER SHAPE EFFECTS
is fixed to be small a = 0.1. The initial condition is chosen In this section, we consider a static mapping h(·) with
to be θ̂(0) = −1 such that the local maximum θ̂ = −0.6 the first order extremum seeking controller from Section
lays between the initial condition and the global maximum. III, see Figure 8. Our goal is to consider the effects of the
By choosing ε = 1, δ = 0.005, ω = 0.1, x1 (0) = x2 (0) = 0, shape of the dither signal d(t) on convergence properties
the simulations show that the extremum seeking scheme is of the algorithm. We present the main result in [19] that
stuck in the local maximum θ̃∗ = −0.6, see Figure 6. describes in detail how different dithers affect the domain of
 2 1
attraction and speed of convergence, as well as the accuracy  π (t − 2πk), t ∈ [2πk, π(2k + 2 ))
2 1 3
of extremum seeking control. It is shown that the square wave tri(t) :=
 π (−t − π(2k − 1)), t ∈ [π(2k + 2 ), π(2k + 2 ))
2 3
produces the fastest convergence among all signals with the π (t − π(2k + 2)), t ∈ [π(2k + 2 ), 2π(k + 1))
same amplitude and frequency, if the amplitude a and the Note that by definition, the signals sin(t), sq(t) and tri(t)
parameter δ in the controller are sufficiently small. Moreover, are of unit amplitude and period 2π. We can generate similar
it is shown that in the limit as the amplitude is reduced to signals of arbitrary amplitude a and frequency ω, e.g., a ·
zero, all dithers yield almost the same domain of attraction sq(ωt). We will often use the “power” (average of the square
and accuracy. of the signal) of the normalized dithers (unit amplitude and
period 2π):
1 1
x y Psq = 1; Psin = ; Ptri = (43)
y = h(x) 2 3
In general, we use Pd to denote the power of dithers d(·) with
amplitude equal to 1 and period 2π. The power for dither d(·)
with amplitude a 6= 1 is equal to a2 Pd . We emphasize that
our results apply to arbitrary dithers satisfying Assumption
x̂ ωδ 9.
+ ×
s
Remark 1 Assumption 9 is needed in our analysis that is
based on averaging of (41). We note that most extremum
seeking literature (see [2]) uses dither signals of the form
d(t) = a sin(ωt) which obviously satisfy our Assumption 9.
d(t )
Introducing the coordinate change x̃ = x − x∗ , we can
rewrite (41) as follows:

Fig. 8. A peak seeking feedback scheme with arbitrary dither. x̃˙ = δωh(x̃ + x∗ + d(t)) · d(t) =: δω f¯(t, x̃, d). (44)
It was shown in Section III that under a stronger version
The model of the closed loop system in Figure 8 is: of Assumption 8 (uniqueness of the maximum) and with
ẋ = δ · ω · h(x + d(t)) · d(t), (41) the sinusoidal dither d(t) = a · sin(ωt), where ω = 1 (that
satisfies Assumption 9) we have that for any compact set D
where h : R → R is sufficiently smooth. The signal and any ν > 0 we can chose the amplitude of the dither
d(·) is referred to as “dither” and δ > 0 and ω > 0 a > 0 and δ > 0 and find a class KL function β, which
are parameters that the designer can choose. We use the depends on δ and d(·), such that the solutions of the closed
following assumptions: loop system (44) satisfy:
|x̃(t0 )| ∈ D =⇒ |x̃(t)| ≤ β (|x̃(t0 )|, t − t0 ) + ν, (45)
Assumption 8 : There exists a maximum x∗ of h(·) such
that for all t ≥ t0 ≥ 0.
Note that D, ν and β are performance indicators since
Dh(x∗ ) = 0; D2 h(x∗ ) < 0 . (42) they quantify different aspects of the performance of the
extremum seeking algorithm. We will show later that each
Assumption 9 Dither signals d(·) are periodic functions of of these indicators is affected by our choice of dither d(·)
period T > 0 (and frequency ω = 2π
T ) that satisfy: and the parameter δ > 0. In particular, we have that:
Z T Z • Speed of convergence of the algorithm is captured by
1 T 2
d(s)ds = 0; d (s)ds > 0; max |d(s)| = a; the function β. Obviously, we would like convergence
0 T 0 s∈[0,T ]
to be as fast as possible.
where a > 0 is the amplitude of the dither. • Domain of convergence is quantified by the set D.
In particular, we would like to make the domain of
The parameter ω in (41) is chosen to be the same as the convergence (attraction) as large as possible.
frequency of the dither signal. • Accuracy of the algorithm is quantified by the number
For comparison purposes, three special kinds of dither are ν > 0 since all trajectories starting in the set D
used repeatedly in our examples: sine wave, square wave and eventually end up in the ball Bν , where we have that
triangle wave. The sine wave is defined in the usual manner. |x(t) − x∗ | ≤ ν. Indeed, the smaller the number ν,
The square wave and triangle wave of unit amplitude and the closer we eventually converge to the maximum x∗
period 2π are defined as follows: (hence, the accuracy of the algorithm is better).

1, t ∈ [2πk, π(2k + 1)) It turns out that a direct analysis of the system (44) to
sq(t) :=
−1, t ∈ [π(2k + 1), 2π(k + 1)) estimate D, ν, β is hard but the system can be analyzed via
an appropriate auxiliary averaged system. We will carry out where δ is a controller parameter, a and ω are respectively
such an analysis in the next section. the amplitude and frequency of dither and Pd is the power of
Consider the following auxiliary gradient system: the normalized dither (with unit amplitude and period 2π).
Note also that ωδ is the integrator constant in Figure 8. Also,
ζ̇ = Dh(ζ + x∗ ) . (46)
note that Theorem 4 holds for sufficiently small a and δ that
Because of Assumption 8, the system (46) has the property are independent of ω which is an arbitrary positive number.
that x∗ is an asymptotically stable equilibrium2. Let D Obviously, if the product (49) is larger than 1 then the closed
denote the domain of attraction of x∗ for the system (46) loop system (41) converges faster than the auxiliary gradient
and note that since h(·) is assumed smooth, the set D is system (46). Similarly, if the product (49) is smaller than 1,
a neighborhood of x∗ . In other words, a consequence of the system (41) converges slower than the gradient system
Assumption 8 is that there exists β ∈ KL and a set D such (46).
that for all t ≥ 0 the solutions of (46) satisfy: The first observation is that for sufficiently small a and δ
the bound in Theorem 4 holds for any ω. Hence, for fixed
ζ0 ∈ D ⇒ |ζ(t)| ≤ β(|ζ0 |, t) (47)
a, δ and Pd we have that the larger the ω, the faster the
Using this auxiliary system, we can state our main result: convergence. In other words, Theorem 4 shows that in our
case study we can achieve arbitrarily fast convergence of
Theorem 4 Suppose that Assumption 8 holds and consider the extremum seeking closed loop by making ω sufficiently
the closed loop system (41) with an arbitrary dither d(·) large. Simulations in Example 3 verify our analysis. We
for which Assumption 9 holds, where a > 0 is the dither also emphasize that this result is in general not possible to
amplitude. Let D and β come from (47). Then, for any strict prove for general dynamic plants. For instance, the results
compact subset D̂ of D and any ν > 0, there exists a∗ > 0 in Sections III and IV that are stated for general dynamical
and δ ∗ > 0 such that for any a ∈ (0, a∗ ], δ ∈ (0, δ ∗ ] and systems provide a similar bound as in (48) under the stronger
any ω > 0 we have that solutions of (41) satisfy: assumption that ω is sufficiently small.
 Suppose now that a, δ and ω are fixed and we are only
x̃0 ∈ D̂ ⇒ |x̃(t)| ≤ β |x̃0 |, δωa2 Pd (t − t0 ) + ν (48)
interested in how the shape of dither affects the convergence.
A sketch of proof of Theorem 4 can be found in [19]. As we change dither, its (normalized) power Pd changes and
We emphasize that the auxiliary gradient system (46) as we can see from (43) that the square wave will yield
plays a crucial role in terms of achievable performance twice larger normalized power than the sine wave and three
of the extremum seeking controller. Indeed, D and β are times larger power than the triangle wave. Consequently, we
independent of the choice of dither and Theorem 4 specifies can expect twice faster convergence with the square wave
how they affect the achievable domain of attraction D̂ of the than with the sine wave and three times faster convergence
closed loop (41), as well as the speed of convergence via the than with the triangle wave. Simulation results in Example
function β. 4 that we present in the sequel are consistent with the above
We now discuss Theorem 4 in more detail to explain how analysis.
dither shape affects the domain of attraction, accuracy and Note that Theorem 1 is a weaker version of Theorem 4
convergence speed of the closed loop system. We note that for the sine wave dither only. Indeed, Theorem 1 does not
the controller parameter (a, δ) needs to be tuned appropri- consider arbitrary dither and the domain of attraction and
ately in order for Theorem 4 to hold. convergence estimates are not as sharp as in Theorem 4.
Domain of attraction: It is shown that any dither satisfying For instance, the relationship of convergence rate and the
Assumption 9 can yield a domain of attraction D̂ that is domain of attraction to the auxiliary system (46) was not
an arbitrary strict subset of the domain of attraction of the shown in Section III as this was impossible to do using the
gradient system (46) if a and δ are sufficiently small. We Lyapunov based proofs used to prove Theorem 1. On the
emphasize that ω > 0 can be arbitrary and a and δ do not other hand, using the trajectory based proofs adopted in this
depend on it. paper, we can prove tight estimates as outlined in Theorem
Accuracy: The ultimate bound that is quantified by the 4. Moreover, in Theorem 1 it was not clear how the dither
number ν can be made arbitrarily small by any dither power Pd affects the convergence rate of the average system.
satisfying Assumption 9 if a and δ are sufficiently small. Note that the function β in (48) is the same for any dither
Hence, in the limit, all dithers perform equally well in terms satisfying Assumption 9 and the only difference comes the
of domain of attraction and accuracy. parameters in (49). However, the values of a∗ and δ ∗ are
Convergence speed: We emphasize that β in (48) is the typically different for different dithers.
same as β in (47) for any dither d(·). The main difference in We note that one can state and prove a more general
speed of convergence comes from the scaling factor within version of Theorem 4 that applies to general dynamical plants
the function β: and in this case h(·) is an appropriate reference-to-output
map. With extra assumptions on the plant dynamics, one can
δ · ω · a2 · Pd , (49) use singular perturbation theory to prove this more general
2 Moreover, all local maxima of h(·) are asymptotically stable equilibria result (see for instance [17] for a Lyapunov based proof in
of (46) and all local minima of h(·) are unstable equilibria of (46). the case of sine wave dither). However, in this case we will
need to require that ω is sufficiently small. Next, we will discuss the situation when h(·) is a quadratic
The following proposition is obvious and it states that mapping, where we assume ω = 1. Consider the simplest
the power of the normalized square wave is larger than or possible case of quadratic maps:
equal to the power of any other normalized dither satisfying
Assumption 9. In other words, for fixed δ and a for which h(x) = −x2 + a1 x + a0 ,
(48) holds, the square wave is guaranteed to produce the we have that
fastest convergence over all dithers with the same amplitude
Dh(x) = −2x + a1
and frequency.
a1
and since x∗ = 2 , we can write with x̃ := x − x∗ :
Proposition 4 Consider arbitrary d(·) satisfying Assump-
Dh(x̃ + x∗ ) = −2x̃ .
tion 9. Then, we have that the power of the normalized dither
satisfies: Hence, the auxiliary gradient system (46) takes the following
0 < Pd ≤ Psq = 1 . form:
ζ̇ = −2ζ .
It has been shown in Theorem 4 that the convergence
speed of the extremum seeking systems depends on the It is not hard to show that in this case for arbitrary dither d(·)
choice the dither shape Pd , amplitude a and frequency ω satisfying Assumption 9 we have that the average system is
as well as δ. It also is shown that the domain of the of the form:
attraction
 and accuracy of all dithers are almost the same
a→0 x̃˙ = −2a2 Pd δx̃ . (51)
as . In this part, we use examples to illustrate
δ→0 Hence, in this special case we have that the average system
various behaviors and simulations to confirm our theoretical
for any dither satisfying Assumption 9 is globally expo-
findings. Our results should motivate the users of extremum
nentially stable. Indeed, for square wave, sine wave and
seeking control to experiment with different dithers in order
triangular wave we have from (43) that the following holds
to achieve the desired trade-off between convergence, domain
for all x0 ∈ R, t ≥ 0, respectively:
of attraction or accuracy.

The following example illustrates that increasing the fre- sq : |x̃(t)| = exp −2a2 δt |x̃0 |
quency of dither while keeping a, δ and Pd the same yields 
sin : |x̃(t)| = exp −a2 δt |x̃0 |
faster convergence.  
2
tri : |x̃(t)| = exp − a2 δt |x̃0 | .
3
Example 3 Consider the quadratic mapping
The square wave produces the fastest speed of convergence
h(x) = −(x + 4)2 (50) for the average system among all dithers with the same
amplitude. The same can be concluded for the actual system
where Dh(x̃+x∗ ) = −2x̃. It is trivial to see that in this case using the proof of Theorem 4. This suggests that the square-
D = R and β(s, t) = se−2t (see Theorem 4). The dither is wave dither should be the prime candidate to use in the ES
chosen to be d(t) = a sin(ωt). Hence, from (43) we have system for fast convergence speed, although this dither is
Psin = 1/2. The initial condition is chosen as x0 = −2. rarely considered in the literature [2]. Indeed, all references
When we fix a = 0.5 and δ = 0.1, the output response with that we are aware of use a sinusoidal dither signal. The
different frequencies is shown in Figure 9. It is clear that the simulation results shown in Example 4 illustrates that the
larger the ω is, the faster the convergence is. convergence speed of extremum seeking controller with the
0
square wave is fastest among all dithers with the same
amplitude.
−0.5

−1
Example 4 The simulation is done for the following system
−1.5
where ω = 1:
−2
ẋ = δh(x + d(t))d(t)
−2.5

−3 where h(x) = −(x + 4)2 . In the new coordinate, x̃ = x −


−3.5 ω=1
x∗ = x + 4, we have
ω=10

x̃˙ = δh(x̃ + x∗ + d(t))d(t) = −δ(x̃ + d(t))2 d(t).


−4

−4.5

−5
0 10 20 30 40 50 60 70 80 90 100
The averaged system is given in (51). The simulation result is
shown in Figure 10, where a = 0.1 and δ = 0.5. Simulations
Fig. 9. The output of the ES with different frequencies show that the extremum seeking controller with the square
wave dither converges fastest.
square wave dither sine dither triangular dither

0 0 0

−0.2 −0.2 −0.2 1.0


−0.4
y
y*
−0.4 −0.4
JP
−0.6 −0.6
y
y* −0.6
0.8
y
y*

−0.8 −0.8 −0.8

0.6
−1 −1 −1

−1.2 −1.2 −1.2


0.4
−1.4 −1.4 −1.4
JY
−1.6
0 5000
−1.6
0 5000
−1.6
0 5000 0.2

Fig. 10. The performance of the extremum seeking schemes of different 0.0

excitation signal 0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

Fig. 11. Productivity and yield for system (53) with c = 3, Km = 0.1.
VI. E XTREMUM SEEKING OF BIOPROCESSES
In this section we consider the steady-states of biopro-
cesses that need to achieve an optimal trade-off between
yield and productivity maximization, see [4]. To illustrate 0.9

the ideas, consider a single irreversible enzymatic reaction JT


of the form X1 −→ X2 with X1 the substrate (or reactant) 0.8
and X2 the product. The reaction takes place in the liquid
phase in a continuous stirred tank reactor. The substrate is
0.7
fed into the reactor with a constant concentration c at a
volumetric flow rate u. The reaction medium is withdrawn
at the same volumetric flow rate u so that the liquid volume 0.6

V is kept constant. The process dynamics are described by


the following standard mass-balance state space model: 0.5
u∗ ū
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

ẋ1 = −r(x1 ) + (u/V )(c − x1 ) (52a)


ẋ2 = r(x1 ) − (u/V )x2 (52b) Fig. 12. Overall performance index JT for system (53) with c = 3,
Km = 0.1 and λ = 0.5.
where x1 is the substrate concentration, x2 is the product
concentration and r(x1 ) is the reaction rate (called kinetics).
Obviously this system makes physical sense only in the non-
negative orthant x1 > 0, x2 > 0. Moreover the flow rate u state performance criteria are considered : the productivity
(which is the control input) is non-negative by definition and JP is the amount of the product harvested in the outflow per
physically upper-bounded (by the feeding pump capacity) unit of time JP := ūx̄2 = ūϕ2 (ū); the yield is the amount of
0 6 u 6 umax . In [4] we investigated two different cases product made per unit of substrate fed to the reactor: JY :=
depending on the form of the rate function r(x1 ). The first x̄2 ϕ2 (ū)
c = c . The sensitivity of JP and JY with respect to
one is the Michaelis-Menten kinetics which is the most basic ū is illustrated in Figure 11. The process must be operated
model for enzymatic reactions r(x1 ) = Kvmm+x x1
1
with vm the at a steady-state that achieves a trade-off between yield and
maximal reaction rate and Km the half-saturation constant. productivity. To this end we define an overall performance
−1
To normalise the model we use vm V and vm as the units index as a convex combination of JP and JY :
of u and time respectively. We also assume that the process
 
is equipped with an on-line sensor that measures the product 1−λ
concentration x2 in the outflow. So the normalised model JT (ū) , λJP + (1 − λ)JY = ϕ2 (ū) λū + (54)
c
becomes
x1 for λ ∈ [0, 1]. This cost function is illustrated in Figure
ẋ1 = − + u(c − x1 ) (53a)
Km + x1 12 where it is readily seen that it has a unique global
x1 maximum u∗ . The corresponding optimal steady-state is
ẋ2 = − ux2 . (53b)
Km + x1 naturally defined as x∗1 = ϕ1 (u∗ ), x∗2 = ϕ2 (u∗ ).
It can be readily verified that, for any positive constant input The kinetic rate function r(x1 ), the dynamical model (52)
flow rate ū ∈ (0, umax ], there is a unique steady-state x̄1 = and the function JT (ū) are unknown to the user while the
ϕ1 (ū), x̄2 = ϕ2 (ū) that is globally asymptotically stable in goal is to maximize the composite cost JT = λux2 + (1 −
the non-negative orthant. λ)c−1 x2 but he does not know that JT is a function of ū of
The industrial objective of the process is the production of the form (54) shown in Figure 12. We apply the extremum
the reaction product. For process optimization, two steady- seeking scheme from Section III with the following definition
of the output and input: We consider again the simple model (52) but we now
−1 assume that, in addition to the Michaelis-Menten kinetics,
y = h(x) := λux2 + (1 − λ)c x2 ; u := α(θ̂ + a sin(ωt))
the reaction rate is subject to exponential substrate inhibition.
(55)
The rate function is as follows:
where x2 is measured, λ chosen by the designer and α(·)
vm x1 −bxp1
is a sigmoid function (see [4]). In Figure 13 the operation r(x1 ) = e
Km + x1
of the extremum seeking control algorithm is illustrated for
appropriately tuned parameters a = 0.02, Kδ = 1, ω = 0.1. where b and p are two positive constant parameters. The
We see that there is a time scale separation between the dynamical model is written:
system itself and the climbing mechanism se predicted by vm x1 −bxp1
ẋ1 = − e + u(c − x1 ) (56a)
Theorem 1. Km + x1
vm x1 −bxp1
ẋ2 = e − ux2 (56b)
Km + x1
0.9
Depending on the value of ū ∈ (0, umax ], the system may
0.8
JT have one, two or three steady-states (x̄1 , x̄2 ) with x̄1 solution
0.7
of:
vm x̄1 −bx̄p1
0.6
e = ū(c − x̄1 )
0.5 Km + x̄1
0.4 and x̄2 = c − x̄1 .
0.3 The productvity JP = ūx̄2 is represented in Figure 14 as
0.2 a function of ū. In this example, JP is clearly a multivalued
0.1 function of ū. However it can be seen that it has a unique
0.0
ū global maximum for ū = u∗ . Moreover, the graph of Figure
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

Fig. 13. Extremum seeking for system (53) with a = 0.02, k = 1, 1.0
ω = 0.1. JP
0.8

The following corollary holds for all bioprocesses that


0.6
satisfy the following assumption:
0.4
Assumption 10 The following conditions hold: (i) For each
admissible value of the flow rate ū the system must have 0.2
a single globally asymptotically stable equilibrium; (ii) The
performance cost function must be single-valued and “well- 0.0

0.0 0.2 0.4 u∗ 0.6 0.8
shaped” in the sense that, for the admissible range of flow
rate values 0 6 ū 6 umax , it must have a single maximum
value JT (u∗ ) without any other local extrema. Fig. 14. Productivity JP for system (56) with c = 3, vm = 2, Km = 1,
b = 0.08, p = 3.4.
A direct consequence of Theorem 1 is the following:

Corollary 1 For any initial condition (x1 (0) > 0, x2 (0) > 14 can also be regarded as a bifurcation diagram with respect
0, θ0 (0)) and for any ν > 0, there exist parameters (a, k, ω) to the parameter ū where the solid branches correspond to
such that, for the closed-loop system (53), (55), (7), x1 (t) > stable equilibria and the dashed branch to unstable equilibria.
0, x2 (t) > 0, θ0 (t) bounded and Hence it can be seen that the maximum point is located on
  a stable branch.
lim sup |x1 (t) − x∗1 | + |x2 (t) − x∗2 | + |u(t) − u∗ | 6 ν. Here we assume that the industrial objective is to achieve
t→∞ the maximization of the productivity JP . A fully satisfactory
Note that the above corollary can be restated in the spirit of operation of the extremum seeking control law (with y(t) =
Theorem 1 to guarantee semi-global practical convergence. u(t)x2 (t)) can be observed in Figure 15 and Figure 16.
This is a stronger result than the main result in [22] that The result of Figure 15 is expected since we are in
deals only with local convergence. conditions quite similar to the previous case of Section IV.
It was shown in [4] that some bioprocesses may not satisfy The result of Figure 16 is more informative since here the
Assumption 10 and, in particular, JT may turn out to be convergence towards the maximum of the cost function is
multivalued. Such situation is quite natural in the context of operated in two successive stages. In a first stage, there
bioreactors and yet it is not covered with any of the presented is a fast convergence to the nearest stable state which is
results. A preliminary analysis of this situation was given in located on the lower stable branch followed by a quasi-
[4] and we summarize below. steady-state progression along that branch. Then, when the
y(t)
1.0 1.0
JP
0.8 0.8
a = 0.003

0.6 0.6

0.4 0.4
a = 0.0015
0.2 0.2

0.0
ū 0.0
t
0.0 0.2 0.4 0.6 0.8 0 1 2 3 4 5 6 7 8

Fig. 15. Extremum seeking for system (56) with a = 0.003, k = 10, Fig. 17. Output signal y(t) : when a is too small, the trajectory is stuck
ω = 0.01. on the lower branch.

1.0 1.0
JP JP
0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0.0
ū 0.0

0.0 0.2 0.4 0.6 0.8 0.0 0.2 0.4 0.6 0.8

Fig. 16. Extremum seeking for system (56) with a = 0.003, k = 6, Fig. 18. Extremum seeking for system (56) with a = 0.015, k = 6,
ω = 0.01. ω = 0.01.

state reaches the bifurcation point, there is a fast jump up reactor. Our analysis shows how various tuning parameters
to the good upper branch and a final climbing up to the in the extremum controller affect the overall convergence
maximum point. It is very important to emphasize here that, properties of the algorithm. Such results will be useful to
in order to get the result of Figure 16, the amplitude a of the practitioners since they provide controller tuning guidelines
dither signal must be large enough. Otherwise, the trajectory that can ensure larger domains of attraction, faster conver-
of the closed loop system definitely remains stuck on the gence or better accuracy of the extremum seeking algorithms.
lower branch at the bifurcation point as shown in Figure17. Moreover, it was shown that global extremum seeking in
On the other side, too large values of the dither amplitude presence of local extrema can be achieved using appropriate
are also prohibited because they produce cyclic trajectories tuning of controller parameters.
as shown in Figure18. From all these observations, we can
R EFERENCES
conclude that by tuning the amplitude of the dither signal
properly, it is possible to pass through the discontinuities [1] K. J. Åström and B. Wittenmark, Adaptive control (2nd ed.) Reading,
MA: Addison-Wesley, 1995.
of the stable branches of the cost function and to converge [2] K. B. Ariyur and M. Krstić, Real-Time Optimization by Extremum-
to the global maximum. This further motivates tuning the Seeking Control. Hoboken, NJ: Wiley-Interscience, 2003.
amplitude of dither in the extremum seeking controller. [3] A. Banaszuk, K.B. Ariyur, M. Krstić, C.A. Jacobson, “An adaptive
algorithm for control of combustion instability”, Automatica, vol. 40
While a preliminary analysis of this issue was presented (2004), 1965-1972.
in [4], a careful analysis and tuning guidelines in this case [4] G. Bastin, D. Nešić, Y. Tan and I.M.Y Mareels, “On extremum seek-
remain an open research problem. ing in bioprocesses with multivalued cost functions”, Biotechnology
Progress, in print, 2009.
[5] E. Biyik and M. Arcak, “Gradient climbing in formation via extremum
VII. C ONCLUSIONS seeking and passivity-based coordination rules”, Proc. 46th IEEE
A summary our recent results on dynamical properties of a Conf. Decis. Contr., New Orleans, USA, 2007, pp. 3133–3138.
[6] P. F. Blackman, “Extremum-seeking regulators”, In J. H. Westcott, An
class of adaptive extremum seeking controllers was presented exposition of adaptive control New York: The Macmillan Company,
and applied to a various models of a continuously stirred 1962.
[7] P.M. Dower, P. Farrell and D. Nešić, “Extremum seeking control of
cascaded Raman optical amplifiers”, IEEE Transactions on Control
Technology, vol. 16 (2008), pp. 396–407.
[8] S. Drakunov, Ü. Özgüner, P. Dix and B. Ashra, “ABS control using
optimum search via sliding modes”, IEEE Transactions on Control
Systems Technology, vol. 3 (1995), 79–85.
[9] M. Guay, D. Dochain, M. Perrier, “Adaptive extremum seekingcontrol
of continuous stirred tank bioreactors with unknown growth kinetics”,
Automatica, vol. 40 (2004), 881-888.
[10] M. Krstić and H.-H. Wang, “Stability of extremum seeking feedback
for general nonlinear dynamic systems”, Automatica, vol. 36 (2000),
595–601.
[11] C.G. Mayhew, R.G. Sanfelice and A.R. Teel, “Robust source-seeking
hybrid controllers for autonomous vehicles”, Proc. Amer. Contr. Conf.,
New York, July 11-13, 2007, 1185–1190.
[12] P. Ögren, E. Fiorelli and N.E. Leonard, “Cooperative control of
mobile sensor networks: adaptive gradient climbing in a distributed
environment”, IEEE Trans. Automat. Contr., vol. 49 (2004), 1292–
1302.
[13] K. S. Peterson and A. G. Stefanopoulou, “Extremum seeking control
for soft landing of an electromechanical valve actuator”, Automatica,
vol. 40 (2004), 1063-1069.
[14] D. Popović, M. Janković, S. Manger and A.R.Teel, “Extremum seeking
methods for optimization of variable cam timing engine operation”,
Proc. American Control Conference, Denver, Colorado, June 4-6,
3136–3141, 2003.
[15] E. Sontag, “Comments on integral variants of ISS”, Systems & Control
Letters, 34 pp. 93–100, 1998.
[16] J. Sternby, “Extremum control systems: An area for adaptive control?”,
Proc. Joint Amer. Contr. Conf., San Francisco, CA, (1980), WA2-A.
[17] Y. Tan, D. Nešić and I.M.Y. Mareels, “On non-local stability properties
of extremum seeking control”, Automatica, vol. 42 (2006), pp. 889–
903.
[18] Y. Tan, D. Nešić, I.M.Y. Mareels and A. Astolfi, “On global extremum
seeking in the presence of local extrema”, Automatica 45 (2009), pp.
245–251.
[19] Y. Tan, D. Nešić and I.M.Y. Mareels, “On the choice of dither in
extremum seeking systems: A case study”, Automatica, vol. 44 (2008),
pp. 1446–1450.
[20] A.R. Teel and D. Popović, “Solving smooth and nonsmoooth mul-
tivariable extremum seeking problems by the methods of nonlinear
programming”, Proc. Amer. Contr. Conf., Arlington, VA, June 25-27,
(2001), 2394-2399.
[21] H.-H. Wang, S. Yeung, and M. Krstić, “Experimental Application of
Extremum Seeking on an Axial-Flow Compressor”, IEEE Transac-
tions on Control Systems Technology, 8, pp. 300–309, 2000.
[22] H-H. Wang, M. Krstić and G. Bastin, “Optimizing bioreactors by
extremum seeking”, International Journal of Adaptive Control and
Signal Processing, Vol. 13, pp. 651-669, 1999.
[23] D. J. Wilde, Optimum Seeking Methods. Englewood NJ: Prentice-Hall
Inc., 1964.

You might also like