Arxiv Download

Download as pdf or txt
Download as pdf or txt
You are on page 1of 24

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/370816849

Control of distributed-parameter systems using normal forms: An


introduction

Preprint · May 2023

CITATIONS READS

0 120

5 authors, including:

Nicole Gehring Abdurrahman Irscheid


Johannes Kepler University Linz Universität des Saarlandes
46 PUBLICATIONS   246 CITATIONS    14 PUBLICATIONS   35 CITATIONS   

SEE PROFILE SEE PROFILE

Joachim Deutscher Frank Woittennek


Ulm University Private University for Health Sciences, Medical Informatics and Technology GmbH
184 PUBLICATIONS   1,573 CITATIONS    111 PUBLICATIONS   741 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Output regulation for distributed-parameter systems View project

Control of Infinite-Dimensional Systems View project

All content following this page was uploaded by Nicole Gehring on 17 May 2023.

The user has requested enhancement of the downloaded file.


Survey

Nicole Gehring*, Abdurrahman Irscheid, Joachim Deutscher, Frank Woittennek, and Joachim
Rudolph

Control of distributed-parameter systems using


normal forms: An introduction
Abstract: This paper gives an overview of the control of exists for any controllable system. It is well known that the
distributed-parameter systems using normal forms. Con- design of a stabilizing state feedback for (1) is straightfor-
arXiv:submit/4897546 [eess.SY] 16 May 2023

sidering linear controllable PDE-ODE systems of hyper- ward as the characteristic coefficients 𝑎𝑖 , 𝑖 = 0, . . . , 𝑛 − 1,
bolic type, two methods derive tracking controllers by are simply replaced by some desired ones of a stable closed-
mapping the system into a form that is advantageous for loop system (e.g. [28]). This simplicity motivates the use
the control design, analogous to the finite-dimensional of normal forms for a design. As illustrated in Figure 1,
case. A flatness-based controller makes use of the hy- instead of a potentially very involved synthesis based on
perbolic controller canonical form that follows from a the original representation, a system is first mapped into,
parametrization of the system’s solutions. A backstepping e.g., CCF, with a stabilizing feedback of the normal form
design exploits the strict-feedback form of the system to state implied by the obtained simple system structure.
recursively stabilize and transform the subsystems. Together with the inverse state transformation this re-
sults in the sought feedback of the original state. This
Keywords: distributed-parameter systems, state feedback,
design strategy using normal forms is not limited to linear
backstepping, flatness, canonical forms
finite-dimensional systems.
The nonlinear counterpart of (1) also involves a chain
1 Introduction of integrators, with a nonlinear function in the last dif-
ferential equation (e.g. [39]). As in the linear setting, the
In control theory, a normal or canonical form usually refers existence of such a nonlinear CCF is closely related to the
to a special state-space representation of a dynamical controllability of the system. Flatness or more precisely
system that is well-suited for a certain analysis or synthesis differential flatness (see, e.g., [14, 23, 27]) is a system
purpose. Thereby, the label canonical is usually used if property that can be interpreted as one possible extension
a form is unique. The basic idea of a controller design of the notion of controllability to nonlinear systems. If a
using normal forms is sketched in Figure 1. For linear finite- system is flat, there exists a (fictitious) flat output that
dimensional single-input single-output (SISO) systems, the allows to express all system variables – in particular the
pioneering work [20] introduced the controller canonical state and the input – in terms of the flat output and
form (CCF) and the observer canonical form (OCF) that a finite number of its time derivatives. Not only is this
are also sometimes called the controllable and observable flatness-based parametrization known to allow for an easy
canonical form, respectively. While any observable system trajectory planning and feedforward design, but the flat
can be mapped into OCF, the CCF output and its derivatives also form a nonlinear CCF state.
⎡ ⎤ ⎡ ⎤ Consequently, the controller synthesis for flat systems is
0 1 0 ··· 0 0
⎢ . .. .. .. .. ⎥ ⎢.⎥ straightforward. For example, all nonlinear SISO systems
⎢ .. . . . . ⎥ ⎢ .. ⎥
⎢ ⎥ ⎢ ⎥ in strict-feedback form
𝜂˙ = ⎢ .. .. .. 𝜂 + ⎢ .. ⎥ 𝑢 (1)
⎢ ⎥ ⎢ ⎥
⎢ . . . 0 ⎥

⎢.⎥ 𝜂˙ 1 = 𝑓1 (𝜂1 ) + 𝑔1 (𝜂1 )𝜂2 (2a)
··· ··· 0
⎢ ⎥ ⎢ ⎥
⎣ 0 1 ⎦ ⎣0⎦
𝜂˙ 2 = 𝑓2 (𝜂1 , 𝜂2 ) + 𝑔2 (𝜂1 , 𝜂2 )𝜂3 (2b)
−𝑎0 −𝑎1 · · · · · · −𝑎𝑛−1 1
..
.

*Corresponding author: Nicole Gehring, Johannes Kepler Univer- 𝜂˙ 𝑛 = 𝑓𝑛 (𝜂1 , . . . , 𝜂𝑛 ) + 𝑔𝑛 (𝜂1 , . . . , 𝜂𝑛 )𝑢 (2c)
sity Linz, Austria
(see, e.g., [21]) are flat due to their structure, with the
Abdurrahman Irscheid, Joachim Rudolph, Saarland University,
Germany
differential equation for 𝜂𝑖 (𝑡), 𝑖 = 1, . . . , 𝑛, only depending
Joachim Deutscher, Ulm University, Germany on 𝜂1 (𝑡), . . . , 𝜂𝑖+1 (𝑡) (with 𝜂𝑛+1 (𝑡) = 𝑢(𝑡)). Thus, 𝜂1 (𝑡)
Frank Woittennek, UMIT TIROL, Austria constitutes a flat output. The well-known backstepping
2 N. Gehring et al., Control of distributed-parameter systems using normal forms

feedback of stable based and the backstepping approach, here, the focus is
system
original state closed loop on linear heterodirectional hyperbolic systems, where the
partial differential equations (PDEs) are bidirectionally
state inverse state coupled with ordinary differential equations (ODEs) at
1 3 one boundary and actuated at the other one. Due to the
transformation transformation
interconnection between an ODE and a PDE subsystem,
2 feedback of these DPSs are also referred to as PDE-ODE systems.
stable
normal form They arise, for example, from technical processes like the
normal form state closed loop
axial and torsional vibrations of drilling systems, heavy
Fig. 1: Design of a stabilizing feedback of the original state based ropes with a load as well as networks of open channels
on the three steps involving a normal form. and transmission lines (see, e.g., [4, 26, 17]). Backstep-
ping controllers for hyperbolic PDE-ODE systems are
derived in many works, including [40, 8, 11, 1, 10, 2].
method (see, e.g., [21]) takes advantage of this recursive
Here, the multi-step design suggested in [8] is used that
system structure in order to successively stabilize the 𝑛
exploits the strict-feedback form of the PDE-ODE sys-
subsystems by choice of the virtual control input 𝜂𝑖+1 (𝑡)
tem for a recursive stabilization (see also [16, 2]). The
in the 𝑖-th design step. As the virtual feedback induces
flatness-based approach essentially corresponds to [35]
a state transformation for the next step, with the actual
and is presented using a constructive and illustrative per-
state feedback in the final step, this too can be interpreted
spective. Ultimately, a survey of the control design for
as an approach using normal forms (cf. Figure 1).
DPSs using normal forms is given in this paper, with a fo-
Importantly, both the backstepping design and the
cus on an introductory presentation and leaving in-depth
flatness-based approach have counterparts in the infinite-
mathematical background to cited references.
dimensional setting. By applying the recursive backstep-
In what follows, Section 2 introduces the considered
ping design to a spatially discretized (infinite-dimensional)
system class and specifies the control objective. Based on
reaction-diffusion equation, it is shown in [3] that
two preliminary transformations, the system is mapped
a special choice of the virtual feedbacks corresponds
into a simpler form for the subsequent design of two track-
to a Volterra integral transformation for the original
ing controllers. First, the flatness-based design in Section 3
distributed-parameter system (DPS). Again, the idea is
makes use of a parametrization of the system’s solutions
to map a given system into a form, from which a stabiliz-
in order to obtain a feedforward controller and the HCCF,
ing feedback is easily deduced. Nowadays, backstepping
which in turn yields the flatness-based tracking controller.
for parabolic and hyperbolic DPSs (see, e.g., [22, 25]) is
The backstepping method in Section 4 capitalizes on the
oftentimes used synonymous for a transformation-based
strict-feedback form of the system to successively stabi-
design of controllers, with transformations not limited to
lize it. While both approaches are discussed independent
those of Volterra type but also of Fredholm type or even
of one another, Section 5 goes into commonalities and
more involved ones. In contrast to this well-established
differences. The example of a heavy rope with a load in
and highly popular method, the flatness-based design in
Section 6 serves to illustrate the results and to give in-
[35, 34] is lesser known. It builds on the works [30] and
sight into the implementation of both controllers. Finally,
[29] that consider the controllability of linear hyperbolic
some details on existing as well as possible extensions are
DPSs and generalize the CCF to the respective hyperbolic
presented in Section 7.
controller canonical form (HCCF). The HCCF in [29] is
Notation: The elements of a vector 𝑣 ∈ R𝑛 are denoted
introduced based on an equivalent representation of the
by 𝑣𝑖 = 𝑒⊤ 𝑖 𝑣, 𝑖 = 1, . . . , 𝑛, with the canonical unit vectors
hyperbolic system as a neutral differential equation, i.e.,
𝑒𝑖 ∈ R𝑛 . Similarly, double indices indicate elements 𝑀𝑖𝑗
one with delays occurring in the highest time derivative
of a matrix 𝑀 = [𝑀𝑖𝑗 ]. The identity matrix is denoted
(see [15]). This delay representation directly follows from
by 𝐼. As always, R and N are the set of real and natural
a flatness-based parametrization (see [38]). In the end,
numbers, respectively, with the latter including zero. The
similar to the CCF (1), the control design based on the
notation 𝑓 ∈ 𝐶 𝑛 (Ω) means that a function 𝑓 : Ω → R is
HCCF only requires replacing the systems’ characteristic
𝑛 times piecewise continuously differentiable, 𝑓 ∈ 𝐿2 (Ω)
with a desired one for the closed loop.
that it is absolutely square integrable.
This paper gives an introduction to the systematic
design of tracking controllers for DPSs using normal forms.
In order to have a common set-up for both the flatness-
N. Gehring et al., Control of distributed-parameter systems using normal forms 3

The equations in (5) are arranged from 𝑧 = 0 to 𝑧 = 1 to


2 Problem statement reflect the strict-feedback form of the system, with 𝜉(𝑡)
essentially taking the role of 𝜂1 (𝑡) in (2) and 𝑥(𝑧, 𝑡) that
The class of linear hyperbolic PDE-ODE systems under
of 𝜂2 (𝑡). Specifically, the ODE subsystem (5a) is driven by
consideration is visualized in Figure 2 by their coupling
the PDE state at 𝑧 = 0 only. In turn, the PDE subsystem
structure. Preliminary transformations map the system
(5b)–(5d) involves the ODE and the PDE state as well
into a simplified representation that is advantageous for
as the input. Due to the bidirectional coupling between
the design of tracking controllers in Sections 3 and 4.
both subsystems, (5) is a PDE-ODE system. Figure 2
illustrates the system structure, with dedicated colors for
the ODE and PDE subsystem throughout the paper that
2.1 System class
also highlight the interconnection of both subsystems via

The central element of the system is a second-order het- 𝑏 and 𝑐 . Introduce positive-definite functions
erodirectional hyperbolic PDE given in terms of the two ∫︁𝑧
coupled transport equations d𝜁
𝜑𝑖 (𝑧) = , 𝑧 ∈ [0, 1], 𝑖 = 1, 2 (6)
𝜆𝑖 (𝜁)
𝜕𝑡 𝑥1 (𝑧, 𝑡) = 𝜆1 (𝑧)𝜕𝑧 𝑥1 (𝑧, 𝑡) + 𝐴11 (𝑧)𝑥1 (𝑧, 𝑡) 0

+ 𝐴12 (𝑧)𝑥2 (𝑧, 𝑡), 𝑧 ∈ [0, 1) (3a) with the inverses 𝜓𝑖 (𝑧) satisfying 𝜓𝑖 (𝜑𝑖 (𝑧)) = 𝑧 as well as
𝜕𝑡 𝑥2 (𝑧, 𝑡) = −𝜆2 (𝑧)𝜕𝑧 𝑥2 (𝑧, 𝑡) + 𝐴21 (𝑧)𝑥1 (𝑧, 𝑡) delays 𝜏𝑖 = 𝜑𝑖 (1). Then it is also apparent from Figure 2
that the input 𝑢(𝑡) at 𝑧 = 1 acts on the ODE at 𝑧 = 0
+ 𝐴22 (𝑧)𝑥2 (𝑧, 𝑡), 𝑧 ∈ (0, 1]. (3b)
only after a time delay 𝜏1 induced by the finite speed 𝜆1 (𝑧)
Assuming 𝜆1 (𝑧), 𝜆2 (𝑧) > 0, the component 𝑥1 (𝑧, 𝑡) = of propagation. Given appropriate initial conditions (ICs)
𝑒⊤ 2
1 𝑥(𝑧, 𝑡) of the PDE state 𝑥(𝑧, 𝑡) ∈ R propagates in the 𝜉(0) = 𝜉0 ∈ R𝑛 and 𝑥(𝑧, 0) = 𝑥0 (𝑧) ∈ R2 , with piecewise
negative direction of the normalized spatial domain [0, 1] continuous 𝑥0 , the system (5) with 𝑡 > 0 is well posed1
and 𝑥2 (𝑧, 𝑡) = 𝑒⊤2 𝑥(𝑧, 𝑡) in the positive 𝑧-direction, from (e.g. [4, App. A]).
0 to 1. The transport velocities 𝜆1 and 𝜆2 are assumed
to be piecewise continuously differentiable, with 𝑎𝑖𝑗 ∈
𝐶([0, 1]) for the four coupling functions. Due to a dynamic 2.2 Control objective
boundary condition at 𝑧 = 0, the transport equations (3a)
and (3b) are bidirectionally coupled with the ODE The control objective is to make the states 𝑥(𝑧, 𝑡) and
˙ = 𝐹 𝜉(𝑡) + 𝑏𝑥1 (0, 𝑡) 𝜉(𝑡) converge to the corresponding, predefined reference
𝜉(𝑡) (3c)
trajectories in
by means of the 𝑥1 (0, 𝑡) acting on (3c) and the ODE state
𝜉(𝑡) ∈ R𝑛 affecting the boundary condition (BC) (𝑥r (𝑧, 𝑡), 𝜉r (𝑡), 𝑢r (𝑡)). (7)

𝑥2 (0, 𝑡) = 𝑞0 𝑥1 (0, 𝑡) + 𝑐⊤ 𝜉(𝑡) (3d) While the specification of (7) is discussed in the context
of Section 3.3, in general, the reference is only assumed to
of (3b). The dimensions of 𝐹 , 𝑏 and 𝑐⊤ follow from that
satisfy the dynamics (5). The following two assumptions
of 𝜉(𝑡). The remaining BC
are imposed for the design of a (static) state-feedback
𝑥1 (1, 𝑡) = 𝑞1 𝑥2 (1, 𝑡) + 𝑢(𝑡) (3e) tracking controller for the system (5) with input 𝑢(𝑡):
of (3a) introduces the control input 𝑢(𝑡) ∈ R. (A1) The pair (𝐹 , 𝑏) is controllable.
Using the matrix (A2) In the BC (5b), 𝑞0 ̸= 0 holds.
[︃ ]︃
𝜆1 (𝑧) 0 Roughly speaking, (A1) ensures the controllability of the
Λ(𝑧) = (4)
0 −𝜆2 (𝑧) ODE subsystem and (A2) of the PDE subsystem. Noting
that the boundary value 𝑥1 (0, 𝑡) takes the role of an input
that comprises the transport velocities and 𝐴(𝑧) =
w.r.t. the ODE (5a), the necessity of (A1) is evident from
[𝐴𝑖𝑗 (𝑧)], (3a)–(3e) can be written in the compact form
classical theory of linear finite-dimensional systems (e.g.
˙ = 𝐹 𝜉(𝑡) + 𝑏𝑥1 (0, 𝑡)
𝜉(𝑡) (5a)

𝑥2 (0, 𝑡) = 𝑞0 𝑥1 (0, 𝑡) + 𝑐 𝜉(𝑡) (5b)
𝜕𝑡 𝑥(𝑧, 𝑡) = Λ(𝑧)𝜕𝑧 𝑥(𝑧, 𝑡) + 𝐴(𝑧)𝑥(𝑧, 𝑡) (5c) 1 This paper does not discuss the abstract notion of state spaces.
Still, a typical choice for (5) with 𝜉(𝑡) and 𝑥(𝑧, 𝑡) would be
𝑥1 (1, 𝑡) = 𝑞1 𝑥2 (1, 𝑡) + 𝑢(𝑡). (5d) R𝑛 × (𝐿2 ([0, 1]))2 .
4 N. Gehring et al., Control of distributed-parameter systems using normal forms

b x1 (z, t) b x̄1 (z, t)


u(t) ū(t)
ODE ODE q0
q0 q1 q̄1
ξ(t) ξ(t)

c⊤ x2 (z, t) c⊤ x̄2 (z, t)


0 1 z 0 1 z

Fig. 2: Visualization of the PDE-ODE system (5). Fig. 3: Visualization of the transformed PDE-ODE system (13).

[28, Thm. 9.5]). Assumption (A2) guarantees exact control- has zero diagonal elements, with the remaining entries
lability of the PDE subsystem (e.g. [30, Thm. 3.2]). This defined by 𝐴˜12 (𝑧) = 𝐴12 (𝑧)e𝛼1 (𝑧)+𝛼2 (𝑧) and 𝐴˜21 (𝑧) =
too is necessary as 𝑞0 = 0 only gives null controllability, a 𝐴21 (𝑧)e−𝛼1 (𝑧)−𝛼2 (𝑧) .
weaker property that is closely related to stabilizability In the second preliminary step, a Volterra integral
and merely allows to steer the system to zero2 . In con- transformation
trast, exact controllability means that all states can be ∫︁𝑧
reached from the origin. It can also be interpreted as null 𝑥(𝑧, ˜ 𝑡) −
¯ 𝑡) = 𝑥(𝑧, ˜ 𝑡) d𝜁
𝐾(𝑧, 𝜁)𝑥(𝜁, (11)
controllability in forward and backward time. In the end,
0
(A1) and (A2) guarantee controllability of the PDE-ODE
system (5), an obvious prerequisite for the existence of an is used to remove the remaining in-domain coupling at-
˜
tributed to 𝐴(𝑧) in (10) altogether, thus further sim-
infinite-dimensional CCF.
plifying the system representation. For that, the kernel
𝐾(𝑧, 𝜁) ∈ R2×2 , defined on the triangular domain
2.3 Preliminary transformations 𝒯 = {(𝑧, 𝜁) ∈ [0, 1]2 |𝜁 ≤ 𝑧}, (12)
Two preliminary transformations are invoked to map the has to be chosen such that (11) maps (9) into the form
system (5) into a form that is advantageous for the con-
troller designs in Sections 3 and 4. First, the transforma- ˙ = 𝐹 𝜉(𝑡) + 𝑏¯
𝜉(𝑡) 𝑥1 (0, 𝑡) (13a)
tion ⊤
𝑥
¯2 (0, 𝑡) = 𝑞0 𝑥
¯1 (0, 𝑡) + 𝑐 𝜉(𝑡) (13b)
[︂ 𝛼 (𝑧) ]︂
e 1 0 ¯ 𝑡) = Λ(𝑧)𝜕𝑧 𝑥(𝑧,
𝜕𝑡 𝑥(𝑧, ¯ 𝑡) + 𝐶(𝑧)𝜉(𝑡) (13c)
˜ 𝑡) = 𝐸(𝑧)𝑥(𝑧, 𝑡) =
𝑥(𝑧, 𝑥(𝑧, 𝑡) (8)
0 e−𝛼2 (𝑧) 𝑥
¯1 (1, 𝑡) = 𝑞¯1 𝑥
¯2 (1, 𝑡) + 𝑢
¯(𝑡), (13d)
∫︀ 𝑧
𝐴𝑖𝑖 (𝜁)
with 𝛼𝑖 (𝑧) = 0 𝜆𝑖 (𝜁)
d𝜁, 𝑖 = 1, 2, allows to rewrite (5c) where the matrix 𝐶(𝑧) and the boundary value 𝑢 ¯(𝑡) are
without the source terms coefficients 𝐴𝑖𝑖 (𝑧) (e.g. [5]), thus disregarded for the moment. It is apparent from Figure 3
simplifying the system structure. Applying the scaling (8) that the structure of (13) is less complex than that of (5)
to (5) yields (see also Figure 2), as the PDE subsystem essentially only
˙ = 𝐹 𝜉(𝑡) + 𝑏˜ comprises two cascaded transport equations, without any
𝜉(𝑡) 𝑥1 (0, 𝑡) (9a)
couplings.
𝑥 ˜1 (0, 𝑡) + 𝑐⊤ 𝜉(𝑡)
˜2 (0, 𝑡) = 𝑞0 𝑥 (9b) In order to determine 𝐾(𝑧, 𝜁), the integral transfor-
˜ 𝑡) = Λ(𝑧)𝜕𝑧 𝑥(𝑧,
𝜕𝑡 𝑥(𝑧, ˜ 𝑡) + 𝐴(𝑧) ˜ 𝑥(𝑧,
˜ 𝑡) (9c) mation (11) is inserted in (13) while keeping in mind that
˜1 (1, 𝑡) = 𝑞1 e𝛼1 (1)+𝛼2 (1) 𝑥
𝑥 ˜ 𝑡) satisfies (9). As 𝑥(0,
˜2 (1, 𝑡) + e𝛼1 (1) 𝑢(𝑡). (9d) 𝑥(𝑧, ¯ 𝑡) = 𝑥(0,
˜ 𝑡), (13a) and (13b)
follow from (9a) and (9b) without any conditions being
Therein, as a consequence of the transformation (8), the imposed on 𝐾(𝑧, 𝜁). Similarly, a comparison of (9d) and
in-domain coupling matrix (13d) yields 𝑞¯1 = 𝑞1 e𝛼1 (1)+𝛼2 (1) as well as the new input
[︃ ]︃
0 𝐴˜12 (𝑧) ∫︁1
˜
𝐴(𝑧) = (10)
˜ 𝛼1 (1)
𝑞1 𝑒⊤ ⊤
𝐴21 (𝑧) 0 𝑢
¯(𝑡) = e 𝑢(𝑡) + (¯ 2 − 𝑒1 )𝐾(1, 𝑧)𝑥(𝑧,
˜ 𝑡) d𝑧 (14)
0

2 Stabilization is still possible if (A2) is not met and even if (𝐹 , 𝑏)


that is introduced for convenience of notation. Note that
is only stabilizable (e.g. [8, 11, 10]). The choice of closed-loop 𝑢
¯(𝑡) is intentionally introduced such that the boundary
dynamics is more restricted in this case. term 𝑞¯1 𝑥
¯2 (1, 𝑡) in (13d) is retained. By that, both (5) and
N. Gehring et al., Control of distributed-parameter systems using normal forms 5

(13) can be interpreted as time-delay systems of neutral ζ ζ


type if 𝑞1 ̸= 0 (e.g. [15]), while (13) would never be neutral K11 (z, ζ) K12 (z, ζ)
1 1
if 𝑢
¯(𝑡) had compensated 𝑞¯1 𝑥 ¯2 (1, 𝑡).
˜ 𝑡)
Differentiating (11) w.r.t. 𝑡 and substituting 𝜕𝑡 𝑥(𝑧,
˜ 𝑡) by means of (11), an inte-
using (9c) as well as 𝜕𝑧 𝑥(𝑧,
gration by part yields
0 0

0 1 z 0 1 z
¯ 𝑡) = Λ(𝑧)𝜕𝑧 𝑥(𝑧,
𝜕𝑡 𝑥(𝑧, ¯ 𝑡) + 𝐾(𝑧, 0)Λ(0)𝑒2 𝑐 𝜉(𝑡)
ζ ζ
∫︁𝑧 K21 (z, ζ) K22 (z, ζ)
[︀ (︀ )︀ 1 1
+ Λ(𝑧)𝜕𝑧 𝐾(𝑧, 𝜁) + 𝜕𝜁 𝐾(𝑧, 𝜁)Λ(𝜁)
0
˜
]︀
− 𝐾(𝑧, 𝜁)𝐴(𝜁) ˜ 𝑡) d𝜁
𝑥(𝜁,
˜
[︀ ]︀
+ Λ(𝑧)𝐾(𝑧, 𝑧) − 𝐾(𝑧, 𝑧)Λ(𝑧) + 𝐴(𝑧) 𝑥(𝑧,
˜ 𝑡)
0 0
+ [𝐾(𝑧, 0)Λ(0)(𝑒1 + 𝑞0 𝑒2 )] 𝑥
˜1 (0, 𝑡) (15) 0 1 z 0 1 z

Fig. 4: Typical characteristic curves visualize the kernel equations


in view of (9b). For (15) to equal (13c) for arbitrary values
(17), with BCs (17b) and (17e) at 𝜁 = 0 highlighted in magenta,
˜ 𝑡), the matrix
of 𝑥(𝑧, and BCs (17c) and (17d) on the diagonal 𝜁 = 𝑧 in green.

𝐶(𝑧) = 𝐾(𝑧, 0)Λ(0)𝑒2 𝑐⊤ (16)


(15), it is evident that the PDE (5c) can be generalized to
in (13c) is defined based on 𝐾(𝑧, 𝜁) and the terms in the
square brackets in (15) have to vanish. The latter results 𝜕𝑡 𝑥(𝑧, 𝑡) = Λ(𝑧)𝜕𝑧 𝑥(𝑧, 𝑡) + 𝐴(𝑧)𝑥(𝑧, 𝑡) (18)
in the so-called kernel equations that comprise essentially ∫︁𝑧
four transport equations in + 𝐷0 (𝑧, 𝜁)𝑥(𝜁, 𝑡) d𝜁 + 𝑑1 (𝑧)𝑥1 (0, 𝑡) + 𝐷2 (𝑧)𝜉(𝑡),
˜ 0
(︀ )︀
Λ(𝑧)𝜕𝑧 𝐾(𝑧, 𝜁) + 𝜕𝜁 𝐾(𝑧, 𝜁)Λ(𝜁) = 𝐾(𝑧, 𝜁)𝐴(𝜁)
(17a) with additional suitable integral, local and ODE terms at
the actuated boundary (5d). Essentially, the system is only
for the elements 𝐾𝑖𝑗 (𝑧, 𝜁) of 𝐾(𝑧, 𝜁) on the triangular
required to be in strict feedback form (see [16]).
domain 𝒯 as well as the corresponding BCs

𝑞0 𝜆2 (0) As the preliminary transformations3 (8) and (11) make


𝐾11 (𝑧, 0) = 𝐾12 (𝑧, 0) (17b) use of Assumption (A2) (see division by 𝑞0 in (17e)), (13)
𝜆1 (0)
𝐴˜12 (𝑧) exists for all PDE-ODE systems with an exactly control-
𝐾12 (𝑧, 𝑧) = − (17c) lable PDE. Let 𝜉(0) = 𝜉0 and 𝑥(𝑧, ¯ 0) = 𝐸(𝑧)𝑥0 (𝑧) −
𝜆1 (𝑧) + 𝜆2 (𝑧) ∫︀ 𝑧
𝐾(𝑧, 𝜁)𝐸(𝜁)𝑥 0 (𝜁) d𝜁 be the ICs of (13) defined based
𝐴˜21 (𝑧) 0
𝐾21 (𝑧, 𝑧) = (17d) on those of (5). Then, the following lemma holds.
𝜆1 (𝑧) + 𝜆2 (𝑧)
𝜆1 (0) Lemma 1 (Equivalence of (5) and (13)). With the input
𝐾22 (𝑧, 0) = 𝐾21 (𝑧, 0). (17e)
𝑞0 𝜆2 (0) 𝑢
¯(𝑡) defined via (14), (13) is an equivalent representation
of every system (5) with input 𝑢(𝑡) that satisfies (A2).
The transformation (11) and the kernel equations (17) are
well-known in the context of backstepping control (see, Equivalence results from the invertibility of the trans-
e.g., [22]). Based on 𝐴𝑖𝑗 ∈ 𝐶([0, 1]) and 𝜆𝑖 ∈ 𝐶 1 ([0, 1]), formations (8) and (11), the first being obvious from the
𝑖, 𝑗 = 1, 2, it is shown in [32, 5] that (17) admits a regularity of the scaling matrix 𝐸(𝑧), 𝑧 ∈ [0, 1]. The kernel
unique continuous solution. It can be determined using the 𝐾I (𝑧, 𝜁) ∈ R2×2 of the inverse map
method of characteristics and a successive approximation.
∫︁𝑧
This seems quite intuitive based on Figure 4, which illus-
˜ 𝑡) = 𝑥(𝑧,
𝑥(𝑧, ¯ 𝑡) + ¯ 𝑡) d𝜁
𝐾I (𝑧, 𝜁)𝑥(𝜁, (19)
trates the evolution of the solution 𝐾(𝑧, 𝜁) along typical
0
characteristic curves on the triangular domain 𝒯 , starting
at the colorized boundaries specified by (17b)–(17e).

Remark 1 (General system class). Based on the 3 The two transformations could be combined into a single one.
Volterra integral transformation (11) and in particular However, this would yield more complex kernel equations.
6 N. Gehring et al., Control of distributed-parameter systems using normal forms

of (11) follows from the so-called reciprocity relation derivatives up to order 𝑘, which is also the CCF state for
∫︁𝑧 𝑘 = 𝑛 − 1. Then, based on the state transformation
𝐾I (𝑧, 𝜁) = 𝐾(𝑧, 𝜁) + ¯ I (𝜁,
𝐾(𝑧, 𝜁)𝐾 ¯ 𝜁) d𝜁¯ (20) ℎ⊤
⎡ ⎤
𝜁 ⎢ ℎ⊤ 𝐹 ⎥
𝑦 [𝑛−1] (𝑡) = 𝑇𝑐 𝜉(𝑡) with 𝑇𝑐 = ⎢ (22)
⎢ ⎥
(e.g. [10]). For every 𝜁 ≤ 𝑧 ∈ [0, 1], this matrix-valued .. ⎥
⎣ . ⎦
Volterra integral equation admits a unique continuous ℎ⊤ 𝐹 𝑛−1
solution 𝐾I (𝑧, 𝜁) (e.g. [24, Thm. 3.11]), as the elements
of 𝐾(𝑧, 𝜁) are 𝐶([0, 1]2 ) functions. (e.g. [28, 23]), standard calculations confirm that (21)
The representation (13) of system (5) serves as a provides the differential parametrization
common basis for the controller designs in Sections 3 and
𝜉(𝑡) = 𝑇𝑐−1 𝑦 [𝑛−1] (𝑡) (23a)
4. Based on the references (7) for (5), the input 𝑢 ¯(𝑡) is
determined such that the dynamics (13) is stabilized along ¯1 (0, 𝑡) = 𝑦 (𝑛) (𝑡) − 𝑒⊤
𝑥 −1 [𝑛−1]
𝑛 𝑇𝑐 𝐹 𝑇𝑐 𝑦 (𝑡)
a corresponding reference trajectory (𝑥 ¯ r (𝑧, 𝑡), 𝜉r (𝑡), 𝑢
¯r (𝑡)). = 𝑝⊤ [𝑛]
1 𝑦 (𝑡) (23b)
The function 𝑥 ¯ r (𝑧, 𝑡) follows from substituting (7) into
the transformations (8) and (11), with 𝑢 ¯r (𝑡) implied by of the quantities in (13a), i.e., the ODE state 𝜉(𝑡) and the
(14), (8) and (7). Lemma 1 ensures that any tracking boundary value 𝑥 ¯1 (0, 𝑡) are expressed in terms of 𝑦(𝑡) and
controller for (13) also guarantees the stabilization of the its derivatives. Substituting these into the BC (13b) gives
corresponding reference for (5). ¯2 (0, 𝑡) = 𝑞0 𝑝⊤
𝑥 [𝑛] 𝑇 −1 [𝑛−1]
1 𝑦 (𝑡) + 𝑐 𝑇𝑐 𝑦 (𝑡)
= 𝑝⊤ [𝑛]
2 𝑦 (𝑡). (23c)

3 Flatness-based design By (23), all boundary values at 𝑧 = 0 are parametrized.


In order to obtain a flatness-based parametrization of
The design in [35] makes use of a special state-space the distributed state 𝑥(𝑧, ¯ 𝑡), the PDE (13c) is solved by
representation of the PDE-ODE system (13), the HCCF, integration along the characteristic curves. This is signifi-
that is well-suited for the control task. The basis for the cantly facilitated due to the preliminary transformation
HCCF is a flatness-based parametrization of the system (11). Starting at the boundary 𝑧 = 0, in view of (6), the
¯ 𝑡), 𝜉(𝑡) and 𝑢
variables 𝑥(𝑧, ¯(𝑡) in terms of a flat output solution of the Cauchy problem w.r.t. 𝑧 reads
𝑦(𝑡). In contrast to finite-dimensional systems, due to 𝑥
¯1 (𝑧, 𝑡) = 𝑥¯1 (0, 𝑡 + 𝜑1 (𝑧)) (24a)
the hyperbolic nature of (13), parametrizing all system
∫︁𝑧 ⊤
solutions involves not only derivatives of 𝑦(𝑡) but also time 𝑒1 𝐶(𝜁)
− 𝜉(𝑡 + 𝜑1 (𝑧) − 𝜑1 (𝜁)) d𝜁
shifts. The HCCF of (13) is implied by the flatness-based 𝜆1 (𝜁)
0
parametrization of the input 𝑢 ¯(𝑡). From that, the choice
of a desired stable closed-loop dynamics directly gives the 𝑥
¯ 2 (𝑧, 𝑡) = 𝑥
¯ 2 (0, 𝑡 − 𝜑2 (𝑧)) (24b)
𝑧
flatness-based controller. ∫︁ ⊤
𝑒2 𝐶(𝜁)
+ 𝜉(𝑡 − 𝜑2 (𝑧) + 𝜑2 (𝜁)) d𝜁.
𝜆2 (𝜁)
0

3.1 Flatness-based parametrization Then, simply substituting (23) into (24) yields the flatness-
based parametrization
It is well known that the ODE (13a), which is controllable
by Assumption (A1), can be transformed into a CCF ¯1 (𝑧, 𝑡) = 𝑝⊤
𝑥 [𝑛]
1 𝑦 (𝑡 + 𝜑1 (𝑧)) (25a)
analogous to (1). Due to the chain of integrators in this 𝑧
∫︁ ⊤
special state-space representation, with the input acting 𝑒1 𝐶(𝜁) −1 [𝑛−1]
− 𝑇 𝑦 (𝑡 + 𝜑1 (𝑧) − 𝜑1 (𝜁)) d𝜁
only on the last state component, the first state component 𝜆1 (𝜁) 𝑐
0
of the CCF state is a flat output. Using the (invertible) ⊤ [𝑛]
𝑥
¯2 (𝑧, 𝑡) = 𝑝2 𝑦 (𝑡 − 𝜑2 (𝑧)) (25b)
Kalman controllability matrix 𝑀𝑐 = [𝑏, 𝐹 𝑏, . . . , 𝐹 𝑛−1 𝑏],
∫︁𝑧 ⊤
a flat output can be written as 𝑒2 𝐶(𝜁) −1 [𝑛−1]
+ 𝑇 𝑦 (𝑡 − 𝜑2 (𝑧) + 𝜑2 (𝜁)) d𝜁
𝜆2 (𝜁) 𝑐
𝑦(𝑡) = 𝑒⊤ −1
𝑛 𝑀𝑐 𝜉(𝑡)

= ℎ 𝜉(𝑡) (21) 0

(e.g. [28, 23]). Denote by 𝑦 [𝑘] (𝑡) = [𝑦(𝑡), 𝑦(𝑡),


˙ . . . , 𝑦 (𝑘) (𝑡)]⊤ , ¯ 𝑡), which involves 𝑦(𝑡) and
of the distributed state 𝑥(𝑧,
𝑘 ∈ N, the vector containing the flat output 𝑦(𝑡) and its its derivatives as well as time shifts thereof. As 𝜑𝑖 (𝑧) ≥
N. Gehring et al., Control of distributed-parameter systems using normal forms 7

Z Z Z
ū(t) ηn+1 (τ1 , t) ηn+1 (−τ2 , t) ηn (t) η3 (t) η2 (t) η1 (t) = y(t)
transport PDE ...

ηn+1 (τ, t)
...

⟨a,·⟩

an−1
an

a2

a1

a0
...
∫︀ 𝜏1
Fig. 5: Visualization of the HCCF (28) by means of a signal flow diagram, where ⟨𝑎, 𝜂𝑛+1 (·, 𝑡)⟩ = −𝜏2 𝑎(𝜏 )𝜂𝑛+1 (𝜏, 𝑡) d𝜏 .

𝜑𝑖 (𝜁) ≥ 0, 𝑖 = 1, 2, for all 𝜁 ∈ [0, 𝑧], any time shift in (25a) in (13). Since 𝜏1 and 𝜏2 are the time delays induced by the
is positive, thus corresponding to a prediction, with only transport velocities 𝜆1 (𝑧) and 𝜆2 (𝑧) in (13), the definition
negative time shifts, i.e. delays, occurring in (25b). The (27b) relates to the PDE state 𝑥 ¯(𝑧, 𝑡) where 𝑧 ∈ [0, 1].
flatness-based parametrization of the system variables in Moreover, the domain [−𝜏2 , 𝜏1 ] of the distributed state
(13) is completed by expressing the input in terms of 𝑦(𝑡). 𝜂𝑛+1 (𝜏, 𝑡) is on par with the time shifts in (26), i.e., 𝜏 can
To this end, recalling that 𝜓𝑖 (𝜑𝑖 (𝑧)) = 𝑧 and 𝜏𝑖 = 𝜑𝑖 (1), be interpreted as another time argument of 𝜂𝑛+1 (𝜏, 𝑡).
𝑖 = 1, 2, (25) is used in the BC (13d) at 𝑧 = 1 to give In accordance with [38], the HCCF for a system (5)
satisfying Assumptions (A1) and (A2) is defined as
¯(𝑡) = 𝑝⊤
𝑢 [𝑛]
1 𝑦 (𝑡 + 𝜏1 ) − 𝑞 ¯1 𝑝⊤ [𝑛]
2 𝑦 (𝑡 − 𝜏2 ) (26)
∫︁𝜏1 𝜂˙ 𝑖 (𝑡) = 𝜂𝑖+1 (𝑡), 𝑖 = 1, . . . , 𝑛 − 1 (28a)
− 𝑒⊤ −1 [𝑛−1]
1 𝐶(𝜓1 (𝜏1 − 𝜏 ))𝑇𝑐 𝑦 (𝑡 + 𝜏 ) d𝜏 𝜂˙ 𝑛 (𝑡) = 𝜂𝑛+1 (−𝜏2 , 𝑡) (28b)
0
𝜕𝑡 𝜂𝑛+1 (𝜏, 𝑡) = 𝜕𝜏 𝜂𝑛+1 (𝜏, 𝑡), 𝜏 ∈ [−𝜏2 , 𝜏1 ) (28c)
∫︁𝜏2
∫︁𝜏1 𝑛
− 𝑞¯1 𝑒⊤
2 𝐶(𝜓2 (𝜏2 − 𝜏 ))𝑇𝑐−1 𝑦 [𝑛−1] (𝑡 − 𝜏 ) d𝜏 ∑︁
𝜂𝑛+1 (𝜏1 , 𝑡) = − 𝑎(𝜏 )𝜂𝑛+1 (𝜏, 𝑡) d𝜏 − 𝑎𝑖−1 𝜂𝑖 (𝑡)
0
−𝜏2 𝑖=1
after a transformation of the integration variable, in order
to better highlight the different time shifts. Due to the time − 𝑎𝑛 𝜂𝑛+1 (−𝜏2 , 𝑡) + 𝑢
¯(𝑡) (28d)
shifts in the highest-order derivative of 𝑦(𝑡) in the func- and generalizes the well-known CCF (1) for linear ODEs to
tional differential equation (26), the input parametrization linear hyperbolic PDE-ODE systems. Therein, (28a)–(28c)
also constitutes a neutral-type time-delay system. Impor- directly follow from the definition (27) of the HCCF state.
tantly, (26) is equivalent to (13) in that it is simply a It is illustrated in Figure 5 that the HCCF comprises a
different way of writing the dynamics (e.g. [31]). chain of 𝑛 integrators (see (28a)–(28b)) that is also typical
for CCFs in the finite-dimensional case. The integrators are
3.2 Hyperbolic controller canonical form preceded by the transport PDE (28c), the state 𝜂𝑛+1 (𝜏, 𝑡)
of which propagates with a normalized velocity from 𝜏 = 𝜏1
In accordance with [31] and [38], the HCCF is introduced to the boundary at 𝜏 = −𝜏2 , where 𝜂𝑛+1 (−𝜏2 , 𝑡) serves as
based on the input parametrization (26). For that, a thor- the input of the last integrator (28b). The normalization
ough inspection of (26) reveals that 𝑢
¯(𝑡) explicitly depends of the velocity as well as the action of the input 𝑢 ¯(𝑡) at
(𝑛)
on 𝑦 (𝑡 + 𝜏1 ), which is apparent from the definition of the boundary 𝜏 = 𝜏1 , analogously to the input vector of
𝑝⊤
1 in (23b). Moreover, time shifts cover delays with a the CCF (1), justifies the classification of the HCCF (28)
maximum amplitude of 𝜏2 and predictions up to 𝜏1 . This as a normal form5 .
motivates the definition of the HCCF state4 The coefficients 𝑎𝑖 , 𝑖 = 0, . . . , 𝑛 and the continuous
𝜂𝑖 (𝑡) = 𝑦 (𝑖−1) (𝑡 − 𝜏2 ), 𝑖 = 1, . . . , 𝑛 (27a) function 𝑎(𝜏 ) in the BC (28d) are characteristic of the
(𝑛) system (cf. [31]), similar to the characteristic coefficients of
𝜂𝑛+1 (𝜏, 𝑡) = 𝑦 (𝑡 + 𝜏 ), 𝜏 ∈ [−𝜏2 , 𝜏1 ], (27b)
a finite-dimensional CCF (1). They follow from applying
where the dimension of the finite-dimensional part 𝜂(𝑡) =
[𝜂1 (𝑡), . . . , 𝜂𝑛 (𝑡)]⊤ corresponds to the highest-order deriva-
tive in (26) as well as the dimension of the ODE state 𝜉(𝑡) 5 It could be argued that (28) is not canonical and only consti-
tutes a hyperbolic controller form (HCF), as the domain [−𝜏2 , 𝜏1 ]
of the distributed state is fixed only up to a scaling and an off-
4 A typical state space would be R𝑛 × 𝐿2 ([−𝜏2 , 𝜏1 ]). See [35] for set, with [−1, 1] in [29] and [31]. However, it is only possible to
details on state spaces in the context of flatness-based control. normalize either the transport velocity in (28c) or the domain.
8 N. Gehring et al., Control of distributed-parameter systems using normal forms

the transformation between the states 𝜂(𝑡), 𝜂𝑛+1 (𝜏, 𝑡) and τ


¯ 𝑡) to the system representation (13). This HCCF
𝜉(𝑡), 𝑥(𝑧, τ1
state transformation is detailed in Appendix A. In order
to illustrate the underlying idea of the transformation,
η1 (τ, t)
the following simple example considers a hyperbolic PDE
without an ODE subsystem. 0 x̄(z, t)

Example 1 (HCCF for 𝑛 = 0). For a hyperbolic PDE −τ2


0 1 z
system without an ODE, the HCCF (28) only involves
the transport PDE Fig. 6: Hyperbolic PDE without an ODE (𝑛 = 0): The relation
(30) between the HCCF state 𝜂1 (𝜏, 𝑡) and the PDE state 𝑥 ¯ (𝑧, 𝑡)
𝜕𝑡 𝜂1 (𝜏, 𝑡) = 𝜕𝜏 𝜂1 (𝜏, 𝑡), 𝜏 ∈ [−𝜏2 , 𝜏1 ) (29a) of (13) for a fixed 𝑡 is visualized by the graphs in orange that
𝜏
∫︁ 1 correspond to characteristic curves of the PDE.
𝜂1 (𝜏1 , 𝑡) = − 𝑎(𝜏 )𝜂1 (𝜏, 𝑡) d𝜏 − 𝑎0 𝜂1 (−𝜏2 , 𝑡) + 𝑢
¯(𝑡)
−𝜏2
Importantly, (32) can always be solved for 𝑦 (𝑛) (𝑡 + 𝜏1 )
(29b)
due to the definition of 𝑝⊤ 1 in (23b) and thus directly
for the state 𝜂1 (𝜏, 𝑡) = 𝑦(𝑡 + 𝜏 ), 𝜏 ∈ [−𝜏2 , 𝜏1 ] (cf. (27) yields the BC (28d) in view of 𝑦 (𝑛) (𝑡 + 𝜏1 ) = 𝜂𝑛+1 (𝜏1 , 𝑡).
and [29]). With the boundary value 𝑦(𝑡) = 𝑥1 (0, 𝑡) taking While 𝑎𝑛 = −𝑞0 𝑞¯1 is obvious from 𝑝⊤ 2 in (23c), the explicit
the role of a flat output, the solution (24) of the Cauchy definition of 𝑎𝑖 , 𝑖 = 0, . . . , 𝑛−1, and 𝑎(𝜏 ) is omitted, as the
problem reads implementation of the flatness-based tracking controller
is ultimately not reliant on them.
𝑥
¯1 (𝑧, 𝑡) = 𝑥
¯1 (0, 𝑡 + 𝜑1 (𝑧)) = 𝜂1 (𝜑1 (𝑧), 𝑡) (30a) Let 𝜂(0) = 𝜂0 ∈ R𝑛 and 𝜂𝑛+1 (𝜏, 0) = 𝜂𝑛+1,0 (𝜏 ) ∈ R
𝑥 ¯2 (0, 𝑡 − 𝜑2 (𝑧)) = 𝑞0 𝜂1 (−𝜑2 (𝑧), 𝑡).
¯2 (𝑧, 𝑡) = 𝑥 (30b) be the ICs of (28) implied by the ICs of (13) and the HCCF
state transformation (see Lemma 5 in Appendix A). Then,
As 𝑞0 ̸= 0 by Assumption (A2), this PDE solution rep- the following result is consequence of Lemmas 1 and 5.
resents the transformation between the HCCF state and
¯ 𝑡), and vice versa. For a fixed time 𝑡, Figure 6 il-
𝑥(𝑧, Lemma 2 (HCCF). The HCCF (28) with the state (27)
¯ 𝑡)
lustrate the relation of the spatial domain [0, 1] of 𝑥(𝑧, is an equivalent representation of every system (5) that
and the time interval [−𝜏2 , 𝜏1 ] of 𝜂1 (𝜏, 𝑡). Based on 𝜑1 (𝑧) satisfies Assumptions (A1) and (A2).
and 𝜑2 (𝑧) in (30), the graphs (in orange) in Figure 6 cor- It is noteworthy that the stability of the PDE-ODE system
respond to characteristic curves of the hyperbolic PDE. with its equivalent representations (13), (28) and (32) is
Moreover, in view of the input parametrization implied by the solutions of the characteristic equation
¯(𝑡) = 𝑦(𝑡 + 𝜏1 ) − 𝑞0 𝑞¯1 𝑦(𝑡 − 𝜏2 )
𝑢 (31) ∫︁𝜏1
⎡ ⎤
𝑛−1
∑︁
0 = a(𝑠) = 𝑠𝑛⎣e𝜏1 𝑠 + 𝑎𝑛 e−𝜏2 𝑠 + 𝑎(𝜏 )e𝜏 𝑠 d𝜏 ⎦+ 𝑎𝑖 𝑠𝑖
(cf. (26)), it is easy to see that 𝑎(𝜏 ) = 0 and 𝑎0 = −𝑞0 𝑞¯1
−𝜏2 𝑖=0
in the HCCF (29). The input parametrization can be in-
(33)
terpreted as a neutral delay system, which is stable if
(see, e.g., [15] or [31, Thm. 4.1]), where a is apparent from
|𝑞0 𝑞¯1 | < 1 (see [15]).
both (28) and (32). It is shown in [31, Thm. 4.1] that
(28) is a Riesz spectral system if 𝑎𝑛 = ̸ 0, assuming an
Alternatively, 𝑎𝑖 , 𝑖 = 0, . . . , 𝑛 and 𝑎(𝜏 ) can be determined
appropriate state space for (27). While the location of
based on the input parametrization, which is simply a
the roots 𝑠𝑖 , 𝑖 ∈ N, of a is insufficient to deduce stability
representation of the dynamics (13) as a time-delay sys-
in general, for Riesz spectral systems, sup𝑖∈N Re(𝑠𝑖 ) < 0
tem. Applying the general relations (77) and (78) from
ensures exponential stability (see [6, Thm. 3.2.8]). For the
Appendix A to (26), the input parametrization reads
stabilization of (28), the characteristic function a would
¯(𝑡) = 𝑦 (𝑛) (𝑡 + 𝜏1 ) + 𝑎𝑛 𝑦 (𝑛) (𝑡 − 𝜏2 )
𝑢 (32) basically be replaced by a desired one for the closed loop,
𝑛−1 ∫︁𝜏1 analogous to the finite-dimensional setting in (1). However,
∑︁
+ (𝑖)
𝑎𝑖 𝑦 (𝑡 − 𝜏2 ) + (𝑛)
𝑎(𝜏 )𝑦 (𝑡 + 𝜏 ) d𝜏, while this is the underlying idea of the following control
𝑖=0 design, Section 3.3 will also make it clear that the HCCF
−𝜏2
is not explicitly required.
where all terms on the right-hand side could easily be
substituted using the definition (27) of the HCCF state.
N. Gehring et al., Control of distributed-parameter systems using normal forms 9

3.3 Tracking controller t


t∗ + τ2
In order to design a tracking controller, first, an appropri-
t∗
ate reference is specified.

yr (t)
3.3.1 Reference trajectories and feedforward ūr (t)
controller
t0
It is shown in Section 3.1 that the flat output (21)
parametrizes all system variables of (13) (cf. (23a), (25)
and (26)). Therefore, instead of specifying a reference for t0 − τ1
¯ 𝑡) that satisfies (13), an arbitrary (but suffi-
𝜉(𝑡) and 𝑥(𝑧, 0 1 z
ciently smooth) desired trajectory for 𝑦(𝑡) is chosen. For
instance, the transition of a system between two steady Fig. 7: Characteristic curves of the PDE subsystem illustrate the
states in arbitrary, finite time 𝑡* − 𝑡0 can be characterized relation between the transition times for 𝑦r (𝑡) and 𝑢
¯(𝑡). Gray
areas correspond to steady states.
by a piecewise-defined reference function

⎨𝑦0 , 𝑡 < 𝑡0

⎪ ¯ 𝑡), 𝑢
Lemma 3 (References). The references 𝜉r (𝑡), 𝑥(𝑧, ¯r (𝑡)
𝑦r (𝑡) = Ξ(𝑡), 𝑡 ∈ [𝑡0 , 𝑡* ] (34) implied by 𝑦r (𝑡) and (23a), (25), (26) satisfy (13).


⎩𝑦 , 𝑡 > 𝑡* .
*
The result is immediately proven since (23a), (25) and
As all time derivatives of 𝑦r (𝑡) vanish for 𝑡 < 𝑡0 and 𝑡 > 𝑡* , (26) parameterize all system solutions of (13) (see [35]).
the chosen initial and terminal values, 𝑦0 and 𝑦* , imply Substituting 𝑦(𝑡 + 𝜏 ) = 𝑦r (𝑡 + 𝜏 ) for 𝜏 ∈ [−𝜏2 , 𝜏1 ] in
stationary solutions for 𝜉(𝑡) and 𝑥(𝑧, ¯ 𝑡) by the flatness- the parametrization (26) or equivalently (32) yields the
based parametrizations (23a), (25) and (26). Conversely, reference
an equilibrium solution of (13) can be reformulated in (𝑛) (𝑛)
¯r (𝑡) = 𝑦r (𝑡 + 𝜏1 ) + 𝑎𝑛 𝑦r (𝑡 − 𝜏2 )
𝑢 (37)
terms of 𝑦(𝑡) using the HCCF state transformation (see
𝑛−1 𝜏1
Appendix A and Lemma 5). In view of the 𝑛-dimensional
∫︁
(𝑖) (𝑛)
∑︁
+ 𝑎𝑖 𝑦r (𝑡 − 𝜏2 ) + 𝑎(𝜏 )𝑦r (𝑡 + 𝜏 ) d𝜏
ODE state 𝜉(𝑡), the function Ξ(𝑡) is oftentimes chosen as
𝑖=0 −𝜏2
a polynomial
2𝑛+1 (︂ )︂𝑖 for the input 𝑢 ¯(𝑡) and thus a feedforward controller for (13).
∑︁ 𝑡 − 𝑡0
Ξ(𝑡) = 𝑐𝑖 (35) Due to the time shift in (26), a transition time 𝑡* − 𝑡0 for
𝑡* − 𝑡0
𝑖=0 𝑦r (𝑡) between two steady states results in a non-constant
(𝑛) ¯r (𝑡) for 𝑡 ∈ [𝑡0 − 𝜏1 , 𝑡* + 𝜏2 ]. This is illustrated by the
𝑢
of degree at least 2𝑛 + 1 so that 𝑦r (𝑡) is continuous for
characteristic curves in Figure 7 that showcase that a
all 𝑡 (e.g. [23, 27]). The coefficients 𝑐𝑖 , 𝑖 = 0, . . . , 2𝑛 + 1,
control action at 𝑧 = 1 takes the time 𝜏1 to affect the
are computed by requiring
opposite boundary at 𝑧 = 0 where the reference for the flat
(𝑖) output is specified. Thus, control action is necessary before
𝑦r (𝑡0 ) = 𝑦0 , 𝑦r (𝑡0 ) = 0, 𝑖 = 1, . . . , 𝑛 (36a)
(𝑖) the start of the transition at 𝑡0 . In turn, post action on
𝑦r (𝑡* ) = 𝑦* , 𝑦r (𝑡* ) = 0, 𝑖 = 1, . . . , 𝑛. (36b)
[𝑡* , 𝑡* + 𝜏2 ] in 𝑢
¯r (𝑡) is required to account for the transport
They are independent of the times 𝑡0 and 𝑡* as a result of delay 𝜏2 in the positive 𝑧-direction. Ultimately, while the
the special polynomial structure in (35). In the end, by the set-point change for the ODE subsystem and all quantities
choice of a polynomial (35) of degree 2𝑛 + 1, the transition at 𝑧 = 0 takes place in [𝑡0 , 𝑡* ], the overall system is in
time 𝑡* − 𝑡0 , the initial value 𝑦0 and the terminal value 𝑦* steady state only for 𝑡 < 𝑡0 − 𝜏1 and 𝑡 > 𝑡* + 𝜏2 .
remain as design parameters for the reference trajectory. References 𝜉r (𝑡) and 𝑥 ¯ r (𝑧, 𝑡) directly follow from re-
Based on 𝑦r (𝑡), references for all system variables placing 𝑦(𝑡) in (23a) and (25) by 𝑦r (𝑡). They in turn give
in (13) are found by substituting 𝑦r (𝑡) for 𝑦(𝑡) in their the remaining references in (7).
respective flatness-based parametrizations. This, in turn,
yields the references (7) by Lemma 1.
10 N. Gehring et al., Control of distributed-parameter systems using normal forms

3.3.2 Controller and closed loop for 𝑖 = 0, . . . , 𝑛 − 1 and 𝑎 ¯𝑛 = 𝛾. While the representation
(38) of the closed loop is advantageous for the stability
Analogous to the control design based on the CCF (1) analysis, it is more apparent from (40) that the open-loop
for finite-dimensional systems, a stabilizing controller for parameters 𝑎𝑖 , 𝑖 = 0, . . . , 𝑛, and 𝑎(𝜏 ) will be replaced
the HCCF (28) replaces the coefficients 𝑎𝑖 , 𝑖 = 0, . . . , 𝑛, by some desired ones 𝑎 ¯𝑖 , 𝑖 = 0, . . . , 𝑛 and 𝑎¯(𝜏 ) in the
and the function 𝑎(𝜏 ) by desired ones that are associated controlled system. This is verified based on the flatness-
with a stable closed-loop dynamics. Augmenting such a based tracking controller
feedback with the feedforward part 𝑢 ¯r (𝑡) in (37) allows to (𝑛)
𝑢
¯(𝑡) = 𝑢 ¯r (𝑡) + (𝑎𝑛 − 𝑎 ¯𝑛 )𝑒𝑦 (𝑡 − 𝜏2 )
track the reference 𝑦r (𝑡).
𝑛−1
In view of the relation between the characteristic equa- ∑︁ (𝑖)
+ (𝑎𝑖 − 𝑎 ¯𝑖 )𝑒𝑦 (𝑡 − 𝜏2 )
tion (33) of the system and the input parametrization (32)
𝑖=0
as well as the BC (28d) of the HCCF, the choice of a
∫︁𝜏1
desired closed-loop dynamics translates into specifying an (︀ )︀ (𝑛)
+ 𝑎(𝜏 ) − 𝑎 ¯(𝜏 ) 𝑒𝑦 (𝑡 + 𝜏 ) d𝜏 (42)
appropriate functional differential equation for the track-
−𝜏2
ing error 𝑒𝑦 (𝑡) = 𝑦(𝑡) − 𝑦r (𝑡). With the general case dis-
cussed in Remark 2, for simplicity, the infinite-dimensional that follows from solving (40) for 𝑦 (𝑛) (𝑡 + 𝜏1 ), i.e. the
error dynamics boundary value of the HCCF (28), and substituting the
𝑛−1
∑︁ result into (32) together with (37). Hence, applying (42)
0 = 𝜖(𝑛) (𝑡) + 𝜅𝑖 𝜖(𝑖) (𝑡) (38) to (32) yields the desired closed loop (40). Even though
𝑖=0 the integral in (42) involves positive time shifts, with
is considered where 𝜖(𝑡) = 𝑒𝑦 (𝑡 + 𝜏1 ) + 𝛾𝑒𝑦 (𝑡 − 𝜏2 ). From the reference 𝑦r (𝑡) known in advance, no predictions are
the corresponding characteristic equation required for the implementation of (42). In fact, in view
[︃ 𝑛−1
]︃ of the definition of the tracking error 𝑒𝑦 (𝑡) and the HCCF
∑︁
𝜏1 𝑠 −𝜏2 𝑠 𝑛 𝑖
0 = acl (𝑠) = (e + 𝛾e ) 𝑠 + 𝜅𝑖 𝑠 , (39) state (27), (42) is easily be rewritten as the state feedback
𝑖=0 (︀ (𝑛) )︀
𝑢
¯(𝑡) = 𝑢¯r (𝑡) + (𝑎𝑛 − 𝑎 ¯𝑛 ) 𝜂𝑛+1 (−𝜏2 , 𝑡) − 𝑦r (𝑡 − 𝜏2 )
it can easily be verified by explicit calculation that the
𝑛−1
choice 0 < |𝛾| < 1 with coefficients 𝜅𝑖 , 𝑖 = 0, . . . , 𝑛 − 1, ∑︁ (︀ (𝑖) )︀
+ (𝑎𝑖 − 𝑎 ¯𝑖 ) 𝜂𝑖+1 (𝑡) − 𝑦r (𝑡 − 𝜏2 ) (43)
of a Hurwitz polynomial results in Re(𝑠𝑖 ) < 0, 𝑖 ∈ N, for 𝑖=0
the roots of acl . Recalling that the HCCF (28) is a Riesz ∫︁𝜏1
spectral system (see text below (33) and [31, Thm. 4.1]),
(︀ )︀(︀ (𝑛) )︀
+ 𝑎(𝜏 ) − 𝑎 ¯(𝜏 ) 𝜂𝑛+1 (𝜏, 𝑡) − 𝑦r (𝑡 + 𝜏 ) d𝜏
these parameters ensure an exponentially stable closed- −𝜏2
loop system. Moreover, exponential stability can even be
shown for |𝛾| < 1 and coefficients 𝜅𝑖 , 𝑖 = 0, . . . , 𝑛 − 1, based on 𝜂(𝑡) and 𝜂𝑛+1 (𝜏, 𝑡). The overall closed-loop sys-
of a Hurwitz polynomial, as the solution of (38) decays tem is visualized by the signal flow diagram in Figure 8.
exponentially for 𝑡 > 𝜏1 if 𝛾 = 0. The following theorem asserts the stability of the
The link between the open-loop system represented closed-loop dynamics achieved by the flatness-based con-
by (32) and the closed loop (38) becomes clearer if (78) troller (43) and thus the tracking of a desired reference
(𝑖) trajectory.
in Appendix A is adapted to 𝑒𝑦 (𝑡 + 𝜏2 ), 𝑖 = 0, . . . , 𝑛 − 1,
in order to rewrite (38) in a form Theorem 1 (Closed-loop stability). Let Assumptions (A1)
0=
(𝑛)
𝑒𝑦 (𝑡 + 𝜏1 ) + 𝑎
(𝑛)
− 𝜏2 )
¯𝑛 𝑒𝑦 (𝑡 (40) and (A2) hold and |𝛾| < 1 with coefficients 𝜅𝑖 , 𝑖 =
𝜏 1
0, . . . , 𝑛 − 1, of a Hurwitz polynomial. Then, the state
𝑛−1 ∫︁
feedback controller (43) exponentially stabilizes the sys-
∑︁ (𝑖) (𝑛)
+ ¯𝑖 𝑒𝑦 (𝑡 − 𝜏2 ) +
𝑎 𝑎
¯(𝜏 )𝑒𝑦 (𝑡 + 𝜏 ) d𝜏
𝑖=0 tem (5) along reference trajectories (7).
−𝜏2

analogous to (32), with The result directly follows from Lemmas 1–3. Based on As-
𝑛−1
sumptions (A1) and (A2), the HCCF (28) is an equivalent
∑︁ (𝜏1 − 𝜏 )𝑛−𝑖−1 state-space representation of (5) by Lemma 1 and Lemma
𝑎
¯(𝜏 ) = 𝜅𝑖 (41a)
(𝑛 − 𝑖 − 1)! 2. The feedback (43) of the HCCF state ensures conver-
𝑖=0
𝑖 gence of the flat output 𝑦(𝑡) towards its reference 𝑦r (𝑡) (cf.
∑︁ (𝜏1 + 𝜏2 )𝑖−𝑘
𝑎
¯𝑖 = 𝜅𝑖 𝛾 + 𝜅𝑘 (41b) (38)), which in turn implies the same for the references
(𝑖 − 𝑘)!
𝑘=0 (7) by Lemma 3 and Lemma 1, thus completing the proof.
N. Gehring et al., Control of distributed-parameter systems using normal forms 11

reference yr controller ū(t) input transf. u(t)


system (5)
generator (43) (14)

η(t), ηn+1 (·, t) x̃(·, t)

state transf. x̄(·, t) state transf. x(·, t)


Lemma 5 ξ(t) (8) & (11) ξ(t)

Fig. 8: Signal flow diagram of the closed-loop system using the flatness-based controller (43) to track a reference (7) implied by 𝑦r (𝑡)
for the PDE-ODE system (5). Blocks with a stronger saturation highlight the main design components and the differences to the
backstepping-based controller in Figure 11.

Theorem 1 holds analogously for any other appropriate 𝑦(𝑡) to the interval [𝑡 − 𝜏2 , 𝑡 + 𝜏1 ] constitutes a state of the
choice of closed-loop dynamics (see Remark 2). The closed system at time 𝑡 in an appropriate state space6 . As such,
loop (38) is chosen for simplicity of presentation only. solving (38) for 𝑦 (𝑛) (𝑡 + 𝜏1 ) and substituting the result
in the input parametrization (26) already yields a state
Remark 2 (Choice of closed-loop dynamics). The feedback (see [35]). Using the transformation in Appendix
flatness-based design allows for a very general choice A (see also Lemma 5) as well as the preliminary transfor-
of closed-loop dynamics. For example, any functional mations (8) and (11), it can be written as a feedback of
differential equation the states 𝑥(𝑧, 𝑡) and 𝜉(𝑡) of (5). It should be noted that
𝑛 (︁ this abbreviated control design better reflects the solution-
(𝑖) (𝑖)
∑︁
0= 𝑎˜+ ˜−
𝑖 𝑒𝑦 (𝑡 + 𝜏1 ) + 𝑎 𝑖 𝑒𝑦 (𝑡 − 𝜏2 )
oriented perspective of the flatness-based approach, which
𝑖=0 has a strong connection to the behavioral approach in,
∫︁𝜏1

e.g., [33]. However, in contrast to the algorithmic design
(𝑖)
+ 𝑎
˜𝑖 (𝜏 )𝑒𝑦 (𝑡 + 𝜏 ) d𝜏 ⎠ , ˜+
𝑎𝑛 = 1, (44) based on the HCCF, such an approach requires a rigorous
−𝜏2 discussion of state spaces, which is largely omitted in this
paper to keep the presentation as simple as possible.
with design parameters 𝑎 ˜+ ˜−
𝑖 , 𝑎𝑖 , 𝑎
˜𝑖 (𝜏 ), 𝑖 = 0, . . . , 𝑛 such
that the tracking error 𝑒𝑦 (𝑡) converges to zero can be cho-
sen. Importantly, 𝑎 ˜+
𝑛 = 1 ensures that (44) can be solved
(𝑛)
for 𝑦 (𝑡 + 𝜏1 ) to obtain the controller corresponding to 4 Backstepping-based design
(44) by substitution into (32). Note that (44) has to be at
least of order 𝑛, with a choice of order larger than 𝑛 result- In order to design a tracking controller for (13) using
ing in a dynamic feedback, as opposed to the static one in backstepping, the control problem is reformulated as a
(43). Moreover, it is even possible to choose a nonlinear stabilization task. For that, introduce the tracking errors
functional differential equation for the closed loop.
𝑒𝜉 (𝑡) = 𝜉(𝑡) − 𝜉r (𝑡) and ¯ 𝑡) − 𝑥
𝜀(𝑧, 𝑡) = 𝑥(𝑧, ¯ r (𝑧, 𝑡)
The presented derivation of the state feedback (43) em- (45)
phasizes the parallels between the controller design based that satisfy the dynamics
on the finite-dimensional CCF (1) and the HCCF (28), re-
𝑒˙ 𝜉 (𝑡) = 𝐹 𝑒𝜉 (𝑡) + 𝑏𝜀1 (0, 𝑡) (46a)
spectively. However, it is possible to do without the HCCF

and thus without explicitly determining 𝑎𝑖 , 𝑖 = 0, . . . , 𝑛, 𝜀2 (0, 𝑡) = 𝑞0 𝜀1 (0, 𝑡) + 𝑐 𝑒𝜉 (𝑡) (46b)
and 𝑎(𝜏 ) in (28), which may be advantageous for the 𝜕𝑡 𝜀(𝑧, 𝑡) = Λ(𝑧)𝜕𝑧 𝜀(𝑧, 𝑡) + 𝐶(𝑧)𝑒𝜉 (𝑡) (46c)
implementation of a flatness-based tracking controller. ¯(𝑡) − 𝑢
𝜀1 (1, 𝑡) = 𝑞¯1 𝜀2 (1, 𝑡) + 𝑢 ¯r (𝑡) (46d)
Those familiar with the flatness-based control design for
finite-dimensional systems already know that a stabilizing based on (13) as well as references 𝜉r (𝑡), 𝑥 ¯ r (𝑧, 𝑡) and
feedback is usually directly determined based on an input 𝑢
¯r (𝑡) that solve (13) (e.g. flatness-based references, see
parametrization (e.g. [23, 27]). This too is possible for
the DPSs considered here, where both the HCCF and
6 The restriction belongs to the Sobolev space 𝐻 𝑛 ([−𝜏2 , 𝜏1 ]) of
the controller directly follow on the basis of the input 𝑛-times weakly differentiable functions in 𝐿2 ([−𝜏2 , 𝜏1 ]), which
parametrization. Looking at (26), it is interesting to ob- is isomorphic to the HCCF state space R𝑛 × 𝐿2 ([−𝜏2 , 𝜏1 ]) (see,
serve that the restriction of the trajectory of the flat output e.g., [35, Sec. 5.1] for details).
12 N. Gehring et al., Control of distributed-parameter systems using normal forms

z=1 z=0 may be interpreted as a desired value in the sense that


system (46): ū ε-PDE eξ -ODE
𝜀1 (0, 𝑡) → 𝑘⊤ 𝑒𝜉 (𝑡) for 𝑡 → ∞ is to be achieved.
This motivates the definition of the PDE error state
1st step: transformation (48)
˜ 𝑡) = 𝜀(𝑧, 𝑡) − 𝑁 (𝑧)𝑒𝜉 (𝑡)
𝜀(𝑧, (48)
system (51): ū ε̃-PDE eξ -ODE
with a matrix 𝑁 (𝑧) ∈ R2×𝑛 yet to be determined that
satisfies at least 𝑒⊤ 1 𝑁 (0) = 𝑘
⊤ in order for the error
2nd step: transformation (52) & feedback (59)
𝜀˜1 (0, 𝑡) = 0 to imply (47). Using this condition on 𝑁 (𝑧),
the backstepping transformation (48) maps the dynamics
system (57a)–(57c), (58): ε̄-PDE eξ -ODE (46) into the form

Fig. 9: Overview of the two steps involved in the backstepping 𝑒˙ 𝜉 (𝑡) = (𝐹 + 𝑏𝑘⊤ )𝑒𝜉 (𝑡) + 𝑏˜
𝜀1 (0, 𝑡) (49a)
design. Green blocks symbolize stabilized subsystems. [︀ ⊤ ⊤ ⊤
]︀
𝜀˜2 (0, 𝑡) = 𝑞0 𝜀˜1 (0, 𝑡) + 𝑐 + 𝑞0 𝑘 − 𝑒2 𝑁 (0) 𝑒𝜉 (𝑡)
(49b)
Lemma 3). The ICs 𝑒𝜉 (0) = 𝑒𝜉,0 = 𝜉0 − 𝜉r (0), 𝜀(𝑧, 0) = 𝜕𝑡 𝜀(𝑧, ˜ 𝑡) − 𝑁 (𝑧)𝑏˜
˜ 𝑡) = Λ(𝑧)𝜕𝑧 𝜀(𝑧, 𝜀1 (0, 𝑡) (49c)
𝜀0 (𝑧) = 𝑥¯ 0 (𝑧) − 𝑥
¯ r (𝑧, 0) are defined accordingly. [︀ ′ ⊤
]︀
+ Λ(𝑧)𝑁 (𝑧) − 𝑁 (𝑧)(𝐹 + 𝑏𝑘 ) + 𝐶(𝑧) 𝑒𝜉 (𝑡)
Systems (13) and (46) share the same coupling struc-
ture that is illustrated in Figure 3. Importantly, both are 𝜀˜1 (1, 𝑡) = 𝑞¯1 𝜀2 (1, 𝑡) − 𝑒⊤ ¯(𝑡) − 𝑢
1 𝑁 (1)𝑒𝜉 (𝑡) + 𝑢 ¯r (𝑡),
in strict-feedback form (see [16]) as the ODE subsystem (49d)
(46a) is driven only by the boundary value 𝜀1 (0, 𝑡) of the
with the sought matrix 𝐹 + 𝑏𝑘⊤ in (49a). The remaining
PDE, with the control input 𝑢 ¯(𝑡) acting on the opposite
freedom in the choice of 𝑁 (𝑧) is used to decouple the
boundary at 𝑧 = 1 of the PDE subsystem (46b)–(46d).
PDE from the ODE subsystem. This allows to consider
This structure is exploited for a recursive control design. In
the stabilization of the PDE subsystem (49b)–(49d) inde-
a first step, the ODE subsystem (46a) is stabilized by an
pendent of the ODE subsystem (49a) in Section 4.2. By
appropriate choice of its virtual input 𝜀1 (0, 𝑡). Analogous
requiring 𝑁 (𝑧) to satisfy the initial value problem
to the backstepping design for finite-dimensional systems,
this virtual feedback induces a state transformation into Λ(𝑧)𝑁 ′ (𝑧) = 𝑁 (𝑧)(𝐹 + 𝑏𝑘⊤ ) − 𝐶(𝑧) (50a)
error variables that are to be driven to zero by means
𝑘⊤
[︂ ]︂
of the input 𝑢 ¯(𝑡) in the second and final step. Figure 9 𝑁 (0) = ⊤ (50b)
𝑐 + 𝑞0 𝑘 ⊤
roughly illustrates the coupling structure between the
ODE and PDE subsystem following each of the two de- with (50a) for 𝑧 ∈ (0, 1], the ODE state 𝑒𝜉 (𝑡) only impacts
sign steps. It is important to note that the method chosen the actuated boundary (49d). There, it can (and will) be
for the stabilization of any of the subsystems is arbitrary. compensated by an appropriate choice of the control input
As such, in the classical sense of (2) where 𝜂1 (𝑡) relates 𝑢
¯(𝑡). For the same reason, the right-hand side of the
to 𝑒𝜉 (𝑡) in (46) and 𝜂2 (𝑡) to 𝜀(𝑧, 𝑡), this design strategy actuated boundary (49d) is not transformed into the new
involves only the one backstepping step in Section 4.1. coordinates. Motivated by its purpose to decouple the
PDE from the ODE subsystem, (48) is also referred to
as a decoupling transformation, (50) as the decoupling
4.1 Stabilization of the ODE subsystem equations (e.g. [10]).
The solution of (50) is found analogously to [10,
First, only the ODE subsystem (46a) is considered, Thm. 4.1] by multiplying (50) with the eigenvectors of
wherein the boundary term 𝜀1 (0, 𝑡) takes the role of an 𝐹 + 𝑏𝑘⊤ from the right. This breaks down the matrix-
input. In light of the controllability of the pair (𝐹 , 𝑏) (see valued initial value problem into 𝑛 vector-valued ones
(A1)), it is always possible to choose a (virtual) feedback with varying coefficients due to Λ(𝑧), which can be solved
explicitly7 . These unique, continuously differentiable solu-
𝜀1 (0, 𝑡) = 𝑘⊤ 𝑒𝜉 (𝑡) (47)
tions (e.g. [28, Thm. 3.3]) verify the existence of a unique
such that 𝐹 + 𝑏𝑘⊤ is Hurwitz, i.e., all eigenvalues of the
matrix have a negative real part, which ensures exponen-
7 This is due to the preliminary transformations in Section 2.3,
tial stability of 𝑒𝜉 (𝑡). However, such a control law cannot as applying the decoupling transformation to (5) directly would
be implemented, as 𝜀1 (0, 𝑡) is not an input. In fact, (47) result in an initial value problem that cannot be solved explicitly.
N. Gehring et al., Control of distributed-parameter systems using normal forms 13

solution 𝑁 ∈ (𝐶 1 ([0, 1]))2×𝑛 for (50). Consequently, the (53) if the kernel 𝑃 (𝑧, 𝜁) satisfies
backstepping transformation (48) maps (46) into (︀ )︀
Λ(𝑧)𝜕𝑧 𝑃 (𝑧, 𝜁) + 𝜕𝜁 𝑃 (𝑧, 𝜁)Λ(𝜁) = 0
𝑒˙ 𝜉 (𝑡) = (𝐹 + 𝑏𝑘⊤ )𝑒𝜉 (𝑡) + 𝑏˜
𝜀1 (0, 𝑡) (51a)
𝜀˜2 (0, 𝑡) = 𝑞0 𝜀˜1 (0, 𝑡) (51b) (54a)

𝜕𝑡 𝜀(𝑧, ˜ 𝑡) − 𝑁 (𝑧)𝑏˜
˜ 𝑡) = Λ(𝑧)𝜕𝑧 𝜀(𝑧, 𝜀1 (0, 𝑡) (51c) 𝑃 (𝑧, 𝑧)Λ(𝑧) − Λ(𝑧)𝑃 (𝑧, 𝑧) = 0
(54b)
𝜀˜1 (1, 𝑡) = 𝑞¯1 𝜀2 (1, 𝑡) − 𝑒⊤
1 𝑁 (1)𝑒𝜉 (𝑡) ¯(𝑡) − 𝑢
+𝑢 ¯r (𝑡),
(51d) ∫︁𝑧
𝑃 (𝑧, 0)Λ(0)(𝑒1 + 𝑞0 𝑒2 ) + 𝑃 (𝑧, 𝜁)𝑁 (𝜁)𝑏 d𝜁 = 𝑁 (𝑧)𝑏,
which is simply (49) in light of the conditions (50) on 0
𝑁 (𝑧). Importantly, the transformation (48) corresponds (54c)
to a classical backstepping step in the sense of (2) and
preserves the strict-feedback form in (51), with stability with (54a) defined on the triangular domain 𝒯 . In contrast
of the ODE subsystem (51a) ensured by the choice of 𝑘. to (17), the kernel equations (54) involve an integral BC
As (51) only involves a cascade of a PDE subsystem and for 𝑃 (𝑧, 𝜁) at 𝜁 = 0. By tracing (54) back to Volterra
a stable ODE subsystem (see Figure 9), at least if 𝑢 ¯(𝑡) integral equations, it is shown in [9, Lem. 6] that this type
compensates the ODE state in (51d), stabilizing the PDE of kernel equations admits a unique piecewise continuously
subsystem in the next step guarantees an overall stable differentiable solution. In fact, the solution is straightfor-
system. Note that the method chosen for the stabilization ward for the simple case of a 2 × 2 matrix 𝑃 (𝑧, 𝜁) consid-
of (51b)–(51d) is arbitrary. ered here. Based on the four scalar transport equations in
(54a), the two BCs 𝑃12 (𝑧, 𝑧) = 𝑃21 (𝑧, 𝑧) = 0 contained in
(54b) yield 𝑃12 (𝑧, 𝜁) = 𝑃21 (𝑧, 𝜁) = 0. Making use of the
4.2 Stabilization of the PDE subsystem resulting diagonal structure of 𝑃 (𝑧, 𝜁) and the method of
characteristics, the remaining kernel elements
The previous transformation (48) introduced the local 1
term −𝑁 (𝑧)𝑏˜ 𝜀1 (0, 𝑡) in (51c) that may have a destabi- 𝑃𝑖𝑖 (𝑧, 𝜁) = 𝑝𝑖 (𝜑𝑖 (𝑧) − 𝜑𝑖 (𝜁)), 𝑖 = 1, 2 (55)
𝜆𝑖 (𝜁)
lizing effect on the PDE subsystem. As a consequence,
the choice of a stabilizing feedback is not obvious from are given by the solution of the Volterra integral equations
(51b)–(51d). In order to facilitate the stabilization of the ∫︁𝜏
PDE subsystem, the Volterra integral transformation 𝑒⊤ ⊤
1 𝑁 (𝜓1 (𝜏 − 𝜎))𝑏𝑝1 (𝜎) d𝜎 + 𝑝1 (𝜏 ) = 𝑒1 𝑁 (𝜓1 (𝜏 ))𝑏
∫︁𝑧 0
𝜀(𝑧, ˜ 𝑡) −
¯ 𝑡) = 𝜀(𝑧, ˜ 𝑡) d𝜁
𝑃 (𝑧, 𝜁)𝜀(𝜁, (52) (56a)
0 ∫︁𝜏
𝑒⊤ ⊤
2 𝑁 (𝜓2 (𝜏 − 𝜎))𝑏𝑝2 (𝜎) d𝜎 − 𝑞0 𝑝2 (𝜏 ) = 𝑒2 𝑁 (𝜓2 (𝜏 ))𝑏
with 𝑃 (𝑧, 𝜁) ∈ R2×2
on the triangular domain 𝒯 in (12)
0
is used to recover the simple transport equations (56b)
¯ 𝑡) = Λ(𝑧)𝜕𝑧 𝜀(𝑧,
𝜕𝑡 𝜀(𝑧, ¯ 𝑡) (53) with 𝜏 ∈ [0, 𝜑1 (1)] in (56a) and 𝜏 ∈ [0, 𝜑2 (1)] in (56b) that
follow from (54c) after a transformation of the integration
in the new coordinates.
variable and a substitution of 𝜑1 (𝑧) (resp. 𝜑2 (𝑧)) by 𝜏 .
For that, analogous to the preliminary transformation
Because (54) guarantees (53) and 𝜀(0,¯ 𝑡) = 𝜀(0,˜ 𝑡), the
(11) and the calculations done in the context of (15), (52)
transformation (52) maps (51) into
˜ 𝑡) by
is differentiated w.r.t. 𝑡 and 𝑧. Substituting 𝜕𝑡 𝜀(𝑧,
means of (51c) and using an integration by parts together 𝑒˙ 𝜉 (𝑡) = (𝐹 + 𝑏𝑘⊤ )𝑒𝜉 (𝑡) + 𝑏¯
𝜀1 (0, 𝑡) (57a)
with the BC (51b), it is revealed that (52) maps (51c) into
𝜀¯2 (0, 𝑡) = 𝑞0 𝜀¯1 (0, 𝑡) (57b)
¯ 𝑡) = Λ(𝑧)𝜕𝑧 𝜀(𝑧,
𝜕𝑡 𝜀(𝑧, ¯ 𝑡) (57c)
¯(𝑡) − 𝑢
𝜀¯1 (1, 𝑡) = 𝑞¯1 𝜀2 (1, 𝑡) + 𝑢 ¯r (𝑡) − 𝑒⊤
1 𝑁 (1)𝑒𝜉 (𝑡)
∫︁1
− 𝑒⊤ ˜ 𝑡) d𝑧, (57d)
1 𝑃 (1, 𝑧)𝜀(𝑧,
0
14 N. Gehring et al., Control of distributed-parameter systems using normal forms

with ICs 𝑒𝜉 (0) = 𝑒𝜉,0 and 𝜀(𝑧,¯ 0) = 𝜀¯0 (𝑧) = 𝜀0 (𝑧) − b ε̄1 (z, t)
∫︀ 𝑧
𝑁 (𝑧)𝑒𝜉,0 − 0 𝑃 (𝑧, 𝜁)[𝜀0 (𝜁) − 𝑁 (𝜁)𝑒𝜉,0 ] d𝜁 for (57) de- 0
stable
fined based on those of (46). In view of the invertible ODE q0
transformations (48) and (52), where the inverse map of eξ (t)
(52) and the reciprocity relation yielding 𝑃I (𝑧, 𝜁) ∈ R2×2 ε̄2 (z, t)
are defined analogous to (19) and (20), respectively, (46)
0 1 z
and (57) are equivalent.

Lemma 4 (Equivalence of (46) and (57)). The form (57) Fig. 10: Visualization of the closed-loop system (57a)–(57c), (58)
for 𝑞¯1,cl = 0.
is an equivalent representation of (46).

Owing to the coupling structure of (57) and the stable


Theorem 2 (Closed-loop stability). Let Assumptions (A1)
ODE subsystem (57a), a state feedback for 𝑢 ¯(𝑡) only has
and (A2) hold, with 𝑘 such that 𝐹 + 𝑏𝑘⊤ is Hurwitz and
to ensure the stability of the PDE subsystem (57b)–(57d)
𝑞¯1,cl such that |𝑞0 𝑞¯1,cl | < 1. Then, the state feedback con-
in order for the overall system to be stable. With the
troller (59) exponentially stabilizes the system (5) along
PDE subsystem described in its most simple form by
reference trajectories (7).
two cascaded transport equations, this choice is very easy
(recall Figure 1). It is important to note that choosing Based on Lemma 4, exponential stability of the closed-
an appropriate feedback would no longer be obvious if loop system (57a)–(57c), (58) implies convergence of 𝜉(𝑡)
the ODE subsystem (57a) were not already stable, as a ¯ 𝑡) towards 𝑥
towards 𝜉r (𝑡) and of 𝑥(𝑧, ¯ r (𝑧, 𝑡) pointwise in
controller only compensates the potentially destabilizing space (see [8, Thm. 4] for the case 𝑞¯1,cl = 0). This ensures
terms at the actuated boundary (57d). the tracking of a reference (7) for the original system (5)
For example, stability of the closed loop is ensured if by Lemma 1 and Lemma 3.
the BC (57d) at 𝑧 = 1 takes the form
Remark 3 (Choice of closed-loop dynamics). Any
𝜀¯1 (1, 𝑡) = 𝑞¯1,cl 𝜀¯2 (1, 𝑡), (58)
choice of a controller that results in an exponentially
stable closed-loop dynamics for the PDE subsystem is
with 𝑞¯1,cl such that |𝑞0 𝑞¯1,cl | < 1. The exponential stabil-
admissible. For example, an additional transformation
ity of the PDE subsystem and thus of (57a)–(57c) with
of the form (8) may be used to introduce reaction terms
(58) is a classical result (see, e.g., [15, Thm. 2.2] in the
in the transport PDE (57c) and to have the solutions on
context of time-delay systems). Although Remark 3 ad-
the characteristic curves decay, with the feedback again
dresses alternative closed-loop dynamics, most commonly
following from the BC at 𝑧 = 1. Oftentimes, a back-
𝑞¯1,cl = 0 is chosen. The corresponding coupling structure
stepping design makes use of the additional assumption
of the closed-loop system (57) with (58) instead of (57d)
|𝑞0 𝑞¯1 | < 1 (e.g. [1]), which allows to set 𝑞¯1,cl = 𝑞¯1 in
is sketched in Figure 10. Therein, it is apparent that the
view of the stability condition |𝑞0 𝑞¯1,cl | < 1. In the context
input 𝜀¯1 (0, 𝑡) of the stable ODE vanishes for 𝑡 > 𝜏1 if
of time-delay systems, this choice is known to be robust
𝑞¯1,cl = 0 and the PDE state 𝜀(𝑧, ¯ 𝑡) itself is zero after the
w.r.t. small delays in the feedback loop that may arise in
finite time 𝜏1 + 𝜏2 . A comparison of (57d) and (58) then
a controller implementation due to the cancellation of the
directly yields the backstepping-based tracking controller
boundary term (see coefficient of 𝜀2 (1, 𝑡) in (59)).
𝑢
¯(𝑡) = 𝑢 𝑞1,cl − 𝑞¯1 )𝜀2 (1, 𝑡)
¯r (𝑡) + (¯ (59)
∫︁1
+ (𝑒⊤ ¯1,cl 𝑒⊤
1 −𝑞 2 )𝑃 (1, 𝑧)𝜀(𝑧, 𝑡) d𝑧 5 A comparison of both designs
0
⎡ ⎤
∫︁1 The controller designs in Sections 3 and 4 are clearly
+ (𝑒⊤ ¯1,cl 𝑒⊤
1 −𝑞 2 ) 𝑁 (1) −
⎣ 𝑃 (1, 𝑧)𝑁 (𝑧) d𝑧 ⎦ 𝑒𝜉 (𝑡),
based on normal forms (see Figure 1). Although they both
0 share the idea of using state transformations to map a
given system representation into a form from which a
written as a feedback of the original error states 𝑒𝜉 (𝑡) and
stabilizing feedback is easily inferred, they have quite a
𝜀(𝑧, 𝑡) of (46) by use of the transformations (48) and (52).
few differences, at least at first glance.
The overall closed-loop system is visualized by the
With an additional transport equation to incorporate
signal flow diagram in Figure 11. Its stability is asserted
the infinite-dimensional dynamics, the HCCF (28) used
in the following theorem.
N. Gehring et al., Control of distributed-parameter systems using normal forms 15

reference yr references ūr (t) controller ū(t) input transf. u(t)


system (5)
generator Lemma 3 (59) (14)

ε(·, t), eξ (t) x̃(·, t)

x̄r (·, t) error def. x̄(·, t) state transf. x(·, t)


ξr (t) (45) ξ(t) (8) & (11) ξ(t)

Fig. 11: Signal flow diagram of the closed-loop system using the backstepping controller (59) to track a reference (7) associated with
¯ r (𝑧, 𝑡), 𝜉r (𝑡) (and potentially obtained from (7)) for the PDE-ODE system (5). Blocks with a stronger saturation highlight the main
𝑥
design components and the differences to the flatness-based controller in Figure 8.

in the flatness-based design is a direct generalization of Using multiple state transformations for the deriva-
the CCF (1) to hyperbolic PDE-ODE systems (5). The tion of a controller also simplifies the numerical imple-
feedback simply follows from replacing the system’s char- mentation. The backstepping controller (59) requires the
acteristics by some desired one for the closed loop. In solutions 𝑁 (𝑧) and 𝑃 (𝑧, 𝜁) of the initial value problem
contrast, the backstepping approach in Section 4 transfers (50) and the kernel equations (54), respectively. As both
the design of the same name to infinite-dimensional sys- are standard equations encountered in multi-step designs
tems (5) in strict-feedback form. The controller is derived for different type of systems (see, e.g., [9, 16]), a toolbox
by successively stabilizing the subordinate subsystems and (e.g. [13]) with appropriate modular functions can be used.
in the process mapping the dynamics into a simpler form. In contrast, a one-step design (e.g. [11]) yields bilaterally
In the end, similar but not identical to the flatness-based coupled decoupling and kernel equations that are hard to
design based on the HCCF (28), the choice of a desired solve. More importantly, such a design paradigm leads to
BC (57d) (analogous to (28d)) for the closed-loop system different kernel equations for every new system structure,
directly yields a stabilizing feedback. Due to the cascaded which impedes the implementation of their solution.
structure of (57), this only assigns a stable dynamics to While the complexity of the backstepping design is
the PDE subsystem, with the ODE subsystem in closed mainly hidden in the explicit solution of (50) and (54),
loop already fixed by the choice of 𝑘 in the backstepping which is not discussed here in detail, the flatness-based
transformation (48). perspective uses the explicit system solutions and, thus,
If the two transformations (48) and (52) involved in may mistakenly be perceived as more complicated. How-
the backstepping design are combined into a single one, ever, the calculations for the HCCF state transformation
the result actually strongly resembles the structure (76) in Appendix A basically correspond to the explicit solution
of the integral transformation into HCCF. Both are es- of (50) and (54) in the backstepping design. The temporal
sentially of Volterra type with an additional dependence integrals in Appendix A are related to the spatial ones
on the ODE state (although one uses the tracking errors involved in the solution of (50) and (54) via the character-
(45)). The main difference between both transformations istic curves of (13c). Still, recall that the HCCF and thus
is that (76a) allows to map the ODE subsystem (57a) into the corresponding state transformation is not required for
the chain of integrators in (28a), which is fundamental the implementation of the flatness-based feedback.
in the HCCF. On the other hand, Section 4 does with- Ultimately, both controllers (42) and (59) guarantee
out a transformation of the ODE state, as the design exponential stability of the closed-loop system in an ap-
only relies on the strict-feedback form of the system, and propriate sense (see Theorems 1 and 2). In fact, both state
choosing a stabilizing virtual feedback (47) is easy enough feedbacks are identical if 𝛾 and 𝜅𝑖 , 𝑖 = 0, . . . , 𝑛 − 1, in the
without a special structure of 𝐹 . At this point, it should flatness-based design and 𝑞¯1,cl and 𝑘 for the backstepping
be emphasized that the preliminary transformations in approach are chosen such that
Section 2.3 are not necessary for either of the controller
designs. Rather, they simplify solving the Cauchy problem 𝛾 = −𝑞0 𝑞¯1,cl (60a)
w.r.t. 𝑧 (see (24)) for the derivation of the flatness-based 𝑛−1
∑︁
parametrization and allow for an explicit solution of the 𝑠𝑛 + 𝜅𝑖 𝑠𝑖 = det(𝑠𝐼 − 𝐹 − 𝑏𝑘⊤ ), (60b)
𝑖=0
initial value problem (50) in the backstepping approach.
This further underlines the advantage of a multi-step de- which also reflects Assumptions (A2) and (A1). This fol-
sign over a single transformation. lows from a comparison of the closed-loop systems (38)
16 N. Gehring et al., Control of distributed-parameter systems using normal forms

and (57), with (58) instead of (57d) in the latter. Corre-


spondences like (60) may no longer be obvious or even
exist if a different dynamics is chosen for the closed loop in
either Section 3 or 4 (recall Remarks 2 and 3). An equiv-
alence analysis related to (60) can also be found in [19]
in the context of the solution-based control of PDE-ODE
systems with a nonlinear ODE subsystem. The equiva-
lence of the closed-loop systems in Figures 8 and 11 for
appropriate design parameters is also discussed in the
following example.

6 Example: Heavy rope with load Fig. 12: Visualization of the desired transition between the two
steady states 𝑤r (𝑠, 0) = 0 and 𝑤r (𝑠, 5) = 5 of the heavy rope
with load. The transition is specified by the reference 𝑦r (𝑡) for the
The controller designs in Sections 3 and 4 are applied to load position, which is a flat output of the system.
the classical, linear model of a heavy rope with a load
(e.g. [26]) that also serves as the example for the back-
2
stepping approach in [8]. Numerical results illustrate the and 𝑛 = 2 due to 𝜉(𝑡) ∈ R yields (5). Therein,
control performance for different choices of the closed-loop [︃ ]︃ [︃ ]︃ [︃ ]︃
0 1 0 0
dynamics. In that context, further details on the imple- 𝐹 = 𝑔 , 𝑏 = 2𝑔 , 𝑐= , (64)
mentation of both state feedback controllers are provided. 0 − 𝜆(0) 𝜆(0) 1

𝜆(𝑧)
𝑞0 = 𝑞1 = −1 as well as 𝜆1 (𝑧) = 𝜆2 (𝑧) = ℓ and

𝜆 (𝑧)
6.1 System specification 𝐴𝑖1 (𝑧) = −𝐴𝑖2 (𝑧) = 𝑖 = 1, 2. It is easily verified
2ℓ ,
that this PDE-ODE in strict-feedback form satisfies As-
Consider the planar motion of a homogeneous rope of sumptions (A1) and (A2). Moreover, it is shown in [26]
length ℓ with constant cross-sectional area and line density that the heavy rope with a point mass admits the flat
𝜌 in the earth’s gravitational field, with acceleration 𝑔. output 𝑦(𝑡) = 𝑤(0, 𝑡) = [1, 0]𝜉(𝑡) (cf. (21)).
Using the curvilinear coordinate 𝑠 ∈ [0, ℓ] along the rope, Inspired by an experimental setup, the parameters in
𝑤(𝑠, 𝑡) denotes the horizontal displacement of the rope (61) are chosen as 𝜌 = 0.3 kg/m, ℓ = 3 m, 𝑔 = 9.81 kg/ms2
relative to a fixed vertical reference at time 𝑡. As sketched and 𝑚 = 0.25 kg. This implies delays 𝜏1 = 𝜏2 ≈ 0.67 s
in the context of Figure 12, a point load of mass 𝑚 is by (6) as well as eigenvalues 0 1/s and −3.43 1/s of 𝐹 .
attached to the lower end of the rope at 𝑠 = 0. The Consequently, the uncontrolled ODE is not asymptotically
velocity at 𝑠 = ℓ serves as control input. Using small-angle stable. In the following, all numerical values are given in
approximations, a momentum balance yields the model appropriate SI units, which are omitted to simplify the
𝜌𝜕𝑡 𝑤(𝑠, 𝑡) = 𝜕𝑠 𝑔(𝜌𝑠 + 𝑚)𝜕𝑠 𝑤(𝑠, 𝑡) , 𝑠 ∈ (0, ℓ) (61a) presentation.
2
(︀ )︀

𝜕𝑡2 𝑤(0, 𝑡) = 𝑔𝜕𝑠 𝑤(0, 𝑡) (61b)


𝜕𝑡 𝑤(ℓ, 𝑡) = 𝑢(𝑡) (61c) 6.2 Tracking problem
for the motion of a heavy rope with a load. In order to
rewrite (61) in the form (5), first, the spatial domain [0, ℓ] The objective is to horizontally move the rope from one rest
is normalized by setting 𝑧 = 𝑠ℓ ∈ [0, 1]. Then, in view of position to another (see Figure 12). For that, a reference
the dynamical BC (61b), choosing states 𝑦r (𝑡) for the flat output 𝑦(𝑡) = 𝜉1 (𝑡), i.e. the position of
[︂ ]︂ [︂ ]︂ the load, is specified according to (34), with
1 𝜆(𝑧) 1 𝜕𝑠 𝑤(ℓ𝑧, 𝑡)
𝑥(𝑧, 𝑡) = , 𝑧 ∈ [0, 1] (62a)
2 −𝜆(𝑧) 1 𝜕𝑡 𝑤(ℓ𝑧, 𝑡) 𝑐0 = 𝑦0 , 𝑐1 = 0, 𝑐2 = 0 (65a)
[︂ ]︂
𝑤(0, 𝑡) 𝑐3 = 10(𝑦* − 𝑦0 ), 𝑐4 = −15(𝑦* − 𝑦0 ), 𝑐5 = 6(𝑦* − 𝑦0 )
𝜉(𝑡) = (62b)
𝜕𝑡 𝑤(0, 𝑡) (65b)
with √︂
𝑔 for the polynomial of degree 5 in (35) due to 𝑛 = 2.
𝜆(𝑧) = (𝜌ℓ𝑧 + 𝑚), 𝑧 ∈ [0, 1] (63) More precisely, the load is transitioned between the initial
𝜌
N. Gehring et al., Control of distributed-parameter systems using normal forms 17

Fig. 13: Case 𝛾 = 0: The components of the ODE state 𝜉(𝑧, 𝑡)


converge to their respective reference for 𝑡 > 𝜏1 ≈ 0.67.

position 𝑦r (𝑡0 ) = 𝑦0 = 0 at time 𝑡0 = 𝜏1 and the final one


𝑦r (𝑡* ) = 𝑦* = 5 at 𝑡* = 5 − 𝜏2 . Note that the time interval
[𝑡0 , 𝑡* ] ≈ [0.67, 4.33] is chosen such that the feedforward
controller (37) achieves the transition over [0, 5] (recall
Fig. 14: Case 𝛾 = 0: The components of the PDE error state
Figure 7). Applying the designs in Sections 3 and 4 yields
𝑥(𝑧, 𝑡) − 𝑥r (𝑧, 𝑡) converge to zero for 𝑡 > 𝜏1 ≈ 0.67.
two tracking controllers.

𝑥(𝑧, 𝑡) of (5). In fact, both controllers require 𝐾(𝑧, 𝜁). A


6.3 Controller design and implementation numerical solution of the associated kernel equations (17)
is calculated offline on a fixed spatial grid with the help
Three different sets of controller parameters are consid- of the toolbox [13]. Similarly, the initial value problem
ered, each such that the flatness-based design and the (50) and the kernel equations (54) are solved offline using
backstepping method yield the same feedback (see (60)). the same spatial grid. The numerical solutions for 𝐾(𝑧, 𝜁),
For the flatness-based controller, 𝜅0 = 20, 𝜅1 = 9 and 𝑁 (𝑧) and 𝑃 (𝑧, 𝜁) form the basis for the implementation
𝛾 ∈ {−0.3, 0, 0.3} are chosen in the error dynamics (38). of the backstepping controller (59) in Figure 11.
By (60), this fixes 𝑞¯1,cl = − 𝑞𝛾0 = 𝛾 ∈ {−0.3, 0, 0.3} in
(57d) and the vector 𝑘 that determines the solution of
the initial value problem (50) in the backstepping design. 6.4 Simulation results
The division by 𝑞0 in the determination of 𝑞¯1,cl further
emphasizes the necessity of Assumption (A2) for choosing The implementation of system (5) makes use of the method
arbitrary closed-loop dynamics when using backstepping. of characteristics, with the explicit Euler method used
The implementation of the controllers follows the for numerical integration. As such, the step size 2.5 ·
−3
signal flow diagrams in Figures 8 and 11. Both 𝑦r (𝑡) and 10 chosen for the time 𝑡 implies the spatial grid of the
the references in (7) defined by it (see Lemmas 3 and distributed state 𝑥(𝑧, 𝑡). The ICs
[︃ ]︃ [︃ ]︃
1) are calculated offline. Note that the determination of − 21 2 1
3
the references required for the backstepping controller is 𝜉0 = , 𝑥0 (𝑧) = sin (2𝜋𝑧) (66)
0 5 −1
comparatively extensive. In general, all integrals appearing
in the implementation of the two controllers are discretized of the system are such that the rope is neither in steady
by the trapezoidal rule. state nor is the load at the position used as the starting
The flatness-based feedback (42) of the HCCF state position for the transition. The flatness-based and the
makes use of the transformation (76). The functions backstepping-based controllers are implemented as feed-
therein can be determined either analytically offline, which backs of the states 𝜉(𝑡) and 𝑥(𝑧, 𝑡) of (5) (cf. Figure 8 and
may be cumbersome, or simply numerically online. In any 11). Simulation results confirm that both controllers yield
case, there is no need for an online solution of Volterra the same closed-loop behavior in view of (60), apart from
integral equations. As seen in Figure 8, moreover, the very minor deviations that are solely due to numerical
design uses the kernel 𝐾(𝑧, 𝜁) of the preliminary transfor- issues associtated with the chosen step size. Therefore, the
mation (11) to obtain a feedback of the states 𝜉(𝑡) and following results reflect both designs equally, even though
only values for 𝛾 are referenced as 𝛾 = 𝑞¯1,cl .
18 N. Gehring et al., Control of distributed-parameter systems using normal forms

7 Concluding remarks
Both control strategies presented rely on normal forms and
state transformations to determine a stabilizing (static)
state feedback. The flatness-based design in Section 3
maps the PDE-ODE system (5) into the HCCF to find
a controller. In contrast, the backstepping technique in
Section 4 starts off with the strict-feedback form and
Fig. 15: Deviation of the ODE state component 𝜉1 (𝑡), which is a recursively stabilizes the two subsystems of (5). As such,
flat output, from its reference for different values of 𝛾. both approaches directly generalize respective designs in
the finite-dimensional setting.
The controllers (42) and (59) can be augmented to
For 𝛾 = 0, Figures 13 and 14 show that the states 𝜉(𝑡)
output feedback tracking controllers. For that, usually a
and 𝑥(𝑧, 𝑡) of (5) converge to their respective references
state observer is designed that provides estimates of the
𝜉r (𝑡) and 𝑥r (𝑧, 𝑡), with initial errors compensated and
PDE and ODE states based on a boundary measurement.
almost no error left for 𝑡 > 2.5. In particular, the conver-
Analogous to the controllability assumptions (A1) and
gence is exponential in time for 𝑡 > 𝜏1 ≈ 0.67 because of
(A2), this presumes observability or at least detectability
the closed-loop dynamics in (38) with 𝛾 = 0 as well as
of the corresponding PDE-ODE system. In the context
those in (57) with (58) instead of (57d).
of an observer design using normal forms, it is shown in
In Figure 15, the control performance for different
[36, 34] that the hyperbolic observer canonical form is
values of 𝛾 is compared based on the error 𝜉1 (𝑡) − 𝜉1,r (𝑡).
dual to the HCCF and allows for a very simple observer
This is especially interesting as 𝜉1 (𝑡) is a flat output of
design. The observer is derived from an input-output
the system and the convergence of a flat output to its
equation that takes the role of the flatness-based input
corresponding reference implies the same for all system
parametrization (32) as an equivalent representation of
variables. Looking at Figure 15, it becomes apparent that
the system dynamics. A flatness-based output feedback
the evolution of 𝜉1 (𝑡) for 𝑡 ∈ [0, 𝜏1 ] is independent of
tracking control design for the example of a pneumatic
𝛾. This is unsurprising as the control action at 𝑧 = 1
DPS can be found in [17]. Backstepping controllers have
takes the transport time 𝜏1 to affect the ODE at 𝑧 =
been designed using different collocated, anti-collocated
0. Overall, the case 𝛾 = 0, which corresponds to the
or pointwise in-domain measurements. While the state
results in Figures 13 and 14, demonstrates the fastest error
observation based on measuring 𝑥2 (1, 𝑡) can be shown to
convergence. Instead, the smallest rise time is observed for
be simply dual to the state feedback design for (5), the
𝛾 = 0.3, which is due to the alternating behavior of the
observers in [10] and [12] are more involved. In any case,
error that is evident based on the associated closed-loop
existing backstepping results rarely make use of a normal
dynamics. For 𝛾 = −0.3 and 𝛾 = 0.3, it takes the same
form or system structure to facilitate the observer design,
amount of time for the error to be zero, about twice as
with observability assumptions at time more related to a
long as in the case 𝛾 = 0. Still, a non-zero value for 𝛾
specific approach rather than a system property.
may be advantageous when looking at the control effort.
On the other hand, it is well known, that the de-
The root mean square 𝑢rms for 𝑢(𝑡) − 𝑢r (𝑡) is 𝑢rms ≈ 0.12
sign of backstepping controllers is not limited to linear
for 𝛾 = −0.3, 𝑢rms ≈ 0.16 for 𝛾 = 0 and 𝑢rms ≈ 0.21 for
hyperbolic PDE-ODE systems (5) with a scalar input.
𝛾 = 0.3, i.e., it increases with larger values of 𝛾. The upside
Rather, extensions exist towards multi-input multi-output
of 𝛾 = −0.3 is apparent from a comparison of (32) and
ODE-PDE-ODE systems with a general heterodirectional
(40) or alternatively of (57d) and (58) in open and closed
hyperbolic subsystem of arbitrary order and an additional
loop, respectively. With −𝑞0 𝑞¯1 = −1 the counterpart of
ODE subsystem at the actuated boundary (e.g. [9, 12, 16]),
𝛾 in open loop, a negative choice for 𝛾 essentially means
interconnected systems that are underactuated (e.g. [2])
that 𝑢(𝑡) does less compensation of the backward-traveling
or ODE-PDE-ODE systems of parabolic type (e.g. [7, 16]).
wave (see local terms 𝜂𝑛+1 (−𝜏2 , 𝑡) in (42) and 𝜀2 (1, 𝑡) in
Almost all of the cited references exploit certain system
(59)), thus requiring less control effort. Note that the delay-
structures, e.g. a strict-feedback form, to allow for a sim-
robust feedback mentioned in Remark 3 is not possible
plified, recursive design. Still, stabilizability and controlla-
here because 𝑞0 𝑞¯1 = 1.
bility assumptions differ, not only depending on whether
stabilization of an equilibrium is sufficient or a trajectory
N. Gehring et al., Control of distributed-parameter systems using normal forms 19

is to be tracked. This further emphasizes the importance (69) as well as a change of the order of integration yields
of a thorough system analysis ahead of a controller design, the Volterra integral equation
e.g. by checking for the property of flatness. Although 𝜑∫︁2 (𝑧)
extensions in the context of the flatness-based approach 𝜉(𝑡 − 𝜑2 (𝑧)) − ¯2 (𝜑2 (𝑧) − 𝜏 )𝜉(𝑡 − 𝜏 ) d𝜏
𝐶 (70)
for DPSs are sparse, in [37], the design in Section 3 is
0
generalized to PDE-ODE systems with a nonlinear ODE 𝜑∫︁2 (𝑧)
subsystem. In fact, beyond what is already said in Sec- 𝑏
= e−𝐹 𝜑2 (𝑧) 𝜉(𝑡) − e−𝐹 (𝜑2 (𝑧)−𝜏 ) 𝑥
¯2 (𝜓2 (𝜏 ), 𝑡) d𝜏
tion 5, there seems to be a fluid transition between the 𝑞0
0
backstepping design and the flatness-based perspective.
Both are combined in [18] to solve the problem of output for 𝜉(𝑡 − 𝜑2 (𝑧)) with 𝑧 ∈ [0, 1], where 𝑡 taking the role of
regulation for PDE-ODE systems with a nonlinear ODE a parameter. The convolution kernel
subsystem. In general, there appears to be a lot of un- ⊤
∫︁𝜏
−𝐹 𝜏 𝑏𝑐 𝑏
tapped potential in the systematic design of state feedback ¯
𝐶2 (𝜏 ) = e + e−𝐹 (𝜏 −𝜎) 𝑒⊤ 𝐶(𝜓2 (𝜎)) d𝜎
𝑞0 𝑞0 2
controllers and observers using normal forms for DPSs. 0
(71)
is defined for 𝜏 ∈ [0, 𝜑2 (𝑧)]. The existence of a unique
solution of (70) is shown, e.g., in [24, Thm. 3.11] and
A HCCF state transformation allows to express the delayed state 𝜉(𝑡 − 𝜑2 (𝑧)) for all
𝑧 ∈ [0, 1] in terms of the state at time 𝑡. Thus, together
In what follows, the mapping between the state (27) of with (67), the finite-dimensional part 𝜂(𝑡) of the HCCF
the HCCF (28) and the states 𝑥(𝑧,¯ 𝑡) and 𝜉(𝑡) of (13) is follows from 𝜉(𝑡) and the profile 𝑥¯2 (𝑧, 𝑡).
derived. For that, (23a) reveals that In order to calculate 𝜂𝑛+1 (𝜏, 𝑡) = 𝑦 (𝑛) (𝑡 + 𝜏 ) based
¯ 𝑡), first, apply a time shift 𝑡 ↦→ 𝑡 + 𝜏 to
on 𝜉(𝑡) and 𝑥(𝑧,
𝜂(𝑡) = 𝑇𝑐 𝜉(𝑡 − 𝜏2 ), (67)
(23b) to get
the finite-dimensional part of the HCCF state, corresponds ¯1 (0, 𝑡 + 𝜏 ) + 𝑒⊤
𝜂𝑛+1 (𝜏, 𝑡) = 𝑥 𝑛 𝑇𝑐 𝐹 𝜉(𝑡 + 𝜏 ). (72)
to the delayed ODE state of (13). As a relation with 𝜉(𝑡)
While 𝜏 ∈ [−𝜏2 , 𝜏1 ] for the distributed state (27b) of the
is sought, using the solution of the ODE (13a) backwards
HCCF, for 𝜏 ∈ [−𝜏2 , 0], the delayed terms on the right-
in time, the delayed state is expressed by
hand side of (72) are already expressed in terms of 𝜉(𝑡) and
¯ 𝑡) (see (70) and (69)). For 𝜏 ∈ [0, 𝜏1 ], future values
𝑥(𝑧,
𝜉(𝑡 − 𝜑2 (𝑧)) = e−𝐹 𝜑2 (𝑧) 𝜉(𝑡)
𝑥
¯1 (0, 𝑡 + 𝜏 ) and 𝜉(𝑡 + 𝜏 ) are required in (72). For that,
𝜑∫︁2 (𝑧)
similarly to the previous calculation of delayed values, the
− e−𝐹 (𝜑2 (𝑧)−𝜏 ) 𝑏¯
𝑥1 (0, 𝑡 − 𝜏 ) d𝜏 (68) ODE (13a) is solved forwards in time, while the shifted
0 boundary values
∫︁𝜏
for 𝑧 ∈ [0, 1], which in turn depends on delayed values
𝑥
¯1 (0, 𝑡+𝜏 ) = 𝑥¯1 (𝜓1 (𝜏 ), 𝑡)+ 𝑒⊤ 1 𝐶(𝜓1 (𝜎))𝜉(𝑡+𝜏 −𝜎) d𝜎
¯1 (0, 𝑡−𝜏 ), 𝜏 ∈ [0, 𝜑2 (𝑧)]. It had previously been observed
𝑥
from the solution (24) of the Cauchy problem that pre- 0
(73)
dictions of the boundary value 𝑥 ¯1 (0, 𝑡) relate to the state
with 𝜏 ∈ [0, 𝜑1 (𝑧)] follow from (24a). In the end, together
component 𝑥 ¯1 (𝑧, 𝑡) and delays of 𝑥¯2 (0, 𝑡) to 𝑥
¯2 (𝑧, 𝑡). Based
with (73), the unique solution (see again [24, Thm. 3.11])
on that, solving the BC (13b) for 𝑥 ¯1 (0, 𝑡), which is only
of the resulting Volterra integral equation
possible because of the assumption 𝑞0 ̸= 0 (see (A2)), and
𝜑∫︁1 (𝑧)
substituting the result in a slightly rewritten form of (24a)
𝜉(𝑡 + 𝜑1 (𝑧)) − 𝐶¯1 (𝜓1 (𝑧) − 𝜏 )𝜉(𝑡 + 𝜏 ) d𝜏 (74)
gives the delayed boundary value
0
1 (︀
¯2 (𝜓2 (𝜏 ), 𝑡) − 𝑐⊤ 𝜉(𝑡 − 𝜏 )
)︀
¯1 (0, 𝑡 − 𝜏 ) =
𝑥 𝑥 (69) 𝜑∫︁1 (𝑧)
𝑞0 𝐹 𝜑1 (𝑧)
∫︁𝜏 =e 𝜉(𝑡) + e𝐹 (𝜑1 (𝑧)−𝜏 ) 𝑏¯
𝑥1 (𝜓1 (𝜏 ), 𝑡) d𝜏
1
− 𝑒⊤
2 𝐶(𝜓2 (𝜎))𝜉(𝑡 − 𝜏 + 𝜎) d𝜎
0
𝑞0
0
for 𝑧 ∈ [0, 1], with the kernel
∫︁𝜏
for 𝜏 ∈ [0, 𝜑2 (𝑧)] in terms of the PDE state at 𝑡 and the
𝐶1 (𝜏 ) = e𝐹 (𝜏 −𝜎) 𝑏𝑒⊤
¯
1 𝐶(𝜓1 (𝜎)) d𝜎, (75)
delay ODE state. In the end, the combination of (68) and
0
20 REFERENCES

yields the right-hand side of (72) for 𝜏 ∈ [0, 𝜏1 ]. get


Based on the previous calculations, the transformation ∫︁𝜏¯
¯ 𝑡) into 𝜂(𝑡), 𝜂𝑛+1 (𝜏, 𝑡) takes the form
from 𝜉(𝑡), 𝑥(𝑧, 𝑓 (𝜏 )𝑦 (𝑖) (𝑡 + 𝜏 ) d𝜏 (77)
∫︁1 𝜏
¯
𝜏¯
𝜂(𝑡) = 𝐺0 𝜉(𝑡) − 𝑔0 (𝑧)¯
𝑥2 (𝑧, 𝑡) d𝑧 (76a) 𝑛−1
∑︁ ∫︁ (−1)𝑘−𝑖 (𝑘)
0 = 𝜏 − 𝜏 )𝑘−𝑖 𝑓 (𝜏 ) d𝜏
(¯ 𝑦 (𝑡 + 𝜏¯)
(𝑘 − 𝑖)!
𝑘=𝑖 𝜏
¯1 (𝜓1 (𝜏 ), 𝑡) − 𝑔1⊤ (𝜏 )𝜉(𝑡)
𝜂𝑛+1 (𝜏, 𝑡) = 𝑥 (76b) ¯
𝜓∫︁1 (𝜏 ) ∫︁𝜏¯ ∫︁𝜏
(−1)𝑛−𝑖 (𝑛)
+ (𝜏 − 𝜎)𝑛−𝑖−1 𝑓 (𝜎) d𝜎 𝑦 (𝑡 + 𝜏 ) d𝜏
− 𝑔1 (𝜏, 𝑧)¯
𝑥1 (𝑧, 𝑡) d𝑧, 𝜏 ∈ [0, 𝜏1 ] (𝑛 − 𝑖 − 1)!
𝜏 𝜏
0 ¯ ¯
1 for 𝑖 = 0, . . . , 𝑛 − 1 and any continuous function 𝑓 (𝜏 ),
𝜂𝑛+1 (−𝜏, 𝑡) = ¯2 (𝜓2 (𝜏 ), 𝑡) − 𝑔2⊤ (𝜏 )𝜉(𝑡)
𝑥 (76c)
𝑞0 −𝜏2 ≤ 𝜏 ≤ 𝜏 ≤ 𝜏¯ ≤ 𝜏1 . In (25), following a transformation
¯
𝜓∫︁2 (𝜏 ) of the integration variable, (77) allows to replace integrals
− 𝑔2 (𝜏, 𝑧)¯
𝑥2 (𝑧, 𝑡) d𝑧, 𝜏 ∈ [0, 𝜏2 ]. over 𝑦 (𝑖) (𝑡 + 𝜏 ), 𝑖 = 0, . . . , 𝑛 − 1, by ones over 𝑦 (𝑛) (𝑡 + 𝜏 ) =
0
𝜂𝑛+1 (𝜏, 𝑡). Similarly, use
𝑛−1
∑︁ (𝜏 + 𝜏2 )𝑘−𝑖
Specifically, (76a) follows from inserting the solution of
𝑦 (𝑖) (𝑡 + 𝜏 ) = 𝑦 (𝑘) (𝑡 − 𝜏2 )
the Volterra integral equation (70) into (67). The map in (𝑘 − 𝑖)!
𝑘=𝑖
(76b) (resp. (76c)) is obtained by using (73) (resp. (69)) as ∫︁𝜏
well as the solution 𝜉(𝑡 + 𝜑1 (𝑧)) (resp. 𝜉(𝑡 − 𝜑2 (𝑧))) of (74) (𝜏 − 𝜎)𝑛−𝑖−1 (𝑛)
+ 𝑦 (𝑡 + 𝜎) d𝜎 (78)
8
(resp. (70)) in (72). The partitioning in (76b) and (76c) (𝑛 − 𝑖 − 1)!
−𝜏2
is done to better highlight that 𝜂𝑛+1 (𝜏, 𝑡) is independent
(𝑖)
¯1 (𝑧, 𝑡) for 𝜏 ∈ [−𝜏2 , 0] and independent of 𝑥
of 𝑥 ¯2 (𝑧, 𝑡) for for 𝑖 = 0, . . . , 𝑛 − 1 and 𝜏 ∈ [−𝜏2 , 𝜏1 ] to express 𝑦 (𝑡 + 𝜏 )
𝜏 ∈ [0, 𝜏1 ], similar to the case of a hyperbolic PDE without in terms of the HCCF state in (27). Thus, the right-hand
an ODE depicted in Figure 6. In (76), the explicit defini- sides of (23b) and (25) only depend on 𝜂(𝑡) and 𝜂𝑛+1 (𝜏, 𝑡).
tion of 𝐺0 ∈ R𝑛×𝑛 as well as of the continuously differ- This allows to write the flatness-based parametrizations
entiable functions therein, where 𝑔0 (𝑧), 𝑔1 (𝜏 ), 𝑔2 (𝜏 ) ∈ R𝑛 as the transformation
and 𝑔1 (𝜏, 𝑧), 𝑔2 (𝜏, 𝑧) ∈ R, is omitted due to their length ∫︁0
and complexity. The following lemma summarizes the 𝜉(𝑡) = 𝐻0 𝜂(𝑡) + ℎ0 (𝜏 )𝜂𝑛+1 (𝜏, 𝑡) d𝜏 (79a)
results on the HCCF state transformation. −𝜏2

Lemma 5 (HCCF state transformation). The map (76) be- ¯1 (𝑧, 𝑡) = 𝜂𝑛+1 (𝜑1 (𝑧), 𝑡) + ℎ⊤
𝑥 1 (𝑧)𝜂(𝑡) (79b)
tween the states 𝜉(𝑡), 𝑥(𝑧,¯ 𝑡), 𝑧 ∈ [0, 1] of (13) and 𝜂(𝑡), 𝜑∫︁1 (𝑧)
𝜂𝑛+1 (𝜏, 𝑡), 𝜏 ∈ [−𝜏2 , 𝜏1 ] of the HCCF (28) is bijective. + ℎ1 (𝑧, 𝜏 )𝜂𝑛+1 (𝜏, 𝑡) d𝜏, 𝑧 ∈ [0, 1]
The inverse map follows from the flatness-based −𝜏2

parametrizations (23a) and (25) of 𝜉(𝑡) and 𝑥(𝑧, ¯ 𝑡), re- ¯2 (𝑧, 𝑡) = 𝑞0 𝜂𝑛+1 (−𝜑2 (𝑧), 𝑡) + ℎ⊤
𝑥 2 (𝑧)𝜂(𝑡) (79c)
spectively. For that, repeatedly apply integration by parts ∫︁0
as well as Cauchy’s formula for repeated integration to + ℎ2 (𝑧, 𝜏 )𝜂𝑛+1 (𝜏, 𝑡) d𝜏, 𝑧 ∈ [0, 1],
−𝜏2

which is the inverse of (76). The lengthy definition of


𝐻0 ∈ R𝑛×𝑛 as well as of the continuously differentiable
functions in (79), where ℎ0 (𝜏 ), ℎ1 (𝑧), ℎ2 (𝑧) ∈ R𝑛 and
ℎ1 (𝑧, 𝜏 ), ℎ2 (𝑧, 𝜏 ) ∈ R, is omitted.

References
8 Note that 𝑔1⊤ (0) = 𝑔2⊤ (0) + 1
𝑐⊤ , which allows to include [1] J. Auriol, F. Bribiesca-Argomedo, D. Bou Saba, M.
𝑞0
𝜏 = 0 in both (76b) and (76c). Di Loreto, and F. Di Meglio. “Delay-robust sta-
REFERENCES 21

bilization of a hyperbolic PDE-ODE system”. In: problems. Version 1.1. 2021. url: https://zenodo.
Automatica 95 (2018), pp. 494–502. org/record/6420876.
[2] J. Auriol, F. Bribiesca-Argomedo, S.-I. Niculescu, [14] M. Fliess, J. Lévine, P. Martin, and P. Rouchon.
and J. Redaud. “Stabilization of a hyperbolic PDEs- “Flatness and defect of non-linear systems: introduc-
ODE network using a recursive dynamics intercon- tory theory and examples”. In: Int. J. Contr. 61.6
nection framework”. In: Proc. European Control (1995), pp. 1327–1361.
Conference (ECC), Rotterdam, Netherlands. IEEE. [15] E. Fridman. Introduction to Time-Delay Systems:
2021, pp. 2493–2499. Analysis and Control. Birkhäuser Cham, 2014.
[3] A. Balogh and M. Krstic. “Infinite dimensional [16] N. Gehring. “A systematic design of backstepping-
backstepping-style feedback transformations for a based state feedback controllers for ODE-PDE-ODE
heat equation with an arbitrary level of instability”. systems”. In: IFAC-PapersOnLine 54.9 (2021). 24th
In: Eur. J. Contr. 8.2 (2002), pp. 165–175. International Symposium on Mathematical Theory
[4] G. Bastin and J.-M. Coron. Stability and Boundary of Networks and Systems (MTNS 2020), pp. 410–
Stabilization of 1-D Hyperbolic Systems. Birkhäuser 415.
Cham, 2016. [17] N. Gehring and F. Woittennek. “Flatness-based
[5] J. Coron, R. Vazquez, M. Krstic, and G. Bastin. “Lo- output feedback tracking control of a hyperbolic
cal exponential 𝐻 2 stabilization of a 2×2 quasilinear distributed-parameter system”. In: IEEE Control
hyperbolic system using backstepping”. In: SIAM J. Systems Letters 6 (2022), pp. 992–997.
Contr. Optim. 51.3 (2013), pp. 2005–2035. [18] A. Irscheid, J. Deutscher, N. Gehring, and J.
[6] R. Curtain and H. Zwart. Introduction to Infinite- Rudolph. “Output regulation for general heterodirec-
Dimensional Systems Theory: A State-Space Ap- tional linear hyperbolic PDEs coupled with nonlinear
proach. Springer New York, NY, 2020. ODEs”. In: Automatica 148 (2023), p. 110748.
[7] J. Deutscher and N. Gehring. “Output feedback [19] A. Irscheid, N. Gehring, J. Deutscher, and J.
control of coupled linear parabolic ODE-PDE-ODE Rudolph. “Tracking control for 2 × 2 heterodirec-
systems”. In: IEEE Trans. Automat. Contr. 66.10 tional hyperbolic PDEs that are bidirectionally cou-
(2021), pp. 4668–4683. pled with nonlinear ODEs”. In: Advances in Dis-
[8] J. Deutscher, N. Gehring, and R. Kern. “Back- tributed Parameter Systems. Ed. by J. Auriol, J.
stepping control of linear 2×2 hyperbolic sys- Deutscher, G. Mazanti, and G. Valmorbida. Vol. 14.
tems with dynamic boundary conditions”. In: Springer Cham, 2022, pp. 117–142.
IFAC-PapersOnLine 50.1 (2017). 20th IFAC World [20] R. Kalman. “Mathematical description of linear dy-
Congress, Toulouse, France, pp. 4522–4527. namical systems”. In: J. SIAM Contr. 1.2 (1963),
[9] J. Deutscher, N. Gehring, and R. Kern. “Output pp. 152–192.
feedback control of general linear heterodirectional [21] M. Krstic, I. Kanellakopoulos, and P. Kokotovic.
hyperbolic ODE-PDE-ODE systems”. In: Automat- Nonlinear and Adaptive Control Design. John Wiley
ica 95 (2018), pp. 472–480. & Sons Inc., 1995.
[10] J. Deutscher, N. Gehring, and R. Kern. “Out- [22] M. Krstic and A. Smyshlyaev. Boundary Control of
put feedback control of general linear heterodirec- PDEs. SIAM Philadelphia, 2008.
tional hyperbolic PDE-ODE systems with spatially- [23] J. Lévine. Analysis and Control of Nonlinear Sys-
varying coefficients”. In: Int. J. Contr. 92.10 (2019), tems: A Flatness-based Approach. Springer Berlin,
pp. 2274–2290. Heidelberg, 2009.
[11] F. Di Meglio, F. Bribiesca Argomedo, L. Hu, and [24] P. Linz. Analytical and Numerical Methods for
M. Krstic. “Stabilization of coupled linear heterodi- Volterra Equations. SIAM Philadelphia, 1987.
rectional hyperbolic PDE-ODE systems”. In: Auto- [25] T. Meurer. Control of Higher-Dimensional PDEs:
matica 87 (2018), pp. 281–289. Flatness and Backstepping Designs. Springer Berlin,
[12] F. Di Meglio, P.-O. Lamare, and U. Aarsnes. Heidelberg, 2012.
“Robust output feedback stabilization of an [26] N. Petit and P. Rouchon. “Flatness of heavy chain
ODE–PDE–ODE interconnection”. In: Automatica systems”. In: SIAM J. Control Optim. 40.2 (2001),
119 (2020), p. 109059. pp. 475–495.
[13] F. Fischer, J. Gabriel, and S. Kerschbaum. coni - [27] J. Rudolph. Flatness-Based Control: An Introduc-
a Matlab toolbox facilitating the solution of control tion. Shaker Verlag Aachen, 2021.
22 REFERENCES

[28] W. Rugh. Linear System Theory. Prentice Hall, A. Isidori. IFAC Symposia Series. Oxford: Pergamon,
1996. 1990, pp. 33–38.
[29] D. Russell. “Canonical forms and spectral determi- [40] Z. Zhou and S. Tang. “Boundary stabilization of
nation for a class of hyperbolic distributed parame- a coupled wave-ODE system with internal anti-
ter control systems”. In: J. Math. Anal. Appl. 62.1 damping”. In: Int. J. Contr. 85.11 (2012), pp. 1683–
(1978), pp. 186–225. 1693.
[30] D. Russell. “Controllability and stabilizability the-
ory for linear partial differential equations: recent
progress and open questions”. In: SIAM Rev. 20.4
(1978), pp. 639–739. List of contributors
[31] D. Russell. “Neutral FDE canonical representations
of hyperbolic systems”. In: J. Integr. Equat. Appl. Nicole Gehring
3.1 (1991), pp. 129–166. Institute of Automatic Control and Control Systems Technology,
[32] R. Vazquez, M. Krstic, and J. Coron. “Backstepping Johannes Kepler University Linz, Austria
boundary stabilization and state estimation of a [email protected]

2 × 2 linear hyperbolic system”. In: 2011 IEEE 50th Nicole Gehring received the Dipl.-Ing. degree in electrical engineer-
Conference on Decision and Control and European ing from Dresden University of Technology, Germany, in 2007 and
Control Conference (CDC-ECC). 2011, pp. 4937– the Dr.-Ing. degree in automatic control from Saarland University,
4942. Germany, in 2015. During her two-year stint in the industrial sector,
she gained experience in the control of power plants and motor
[33] J. Willems. “Paradigms and puzzles in the theory
vehicles. As a Postdoc, she was with Technical University of Mu-
of dynamical systems”. In: IEEE Trans. Automat. nich, Germany. She is currently with Johannes Kepler University
Contr. 36.3 (1991), pp. 259–294. Linz, Austria, where her research mainly focuses on the design of
[34] F. Woittennek. “Beobachterbasierte Zustandsrück- controllers and observers for linear distributed-parameter systems.
führungen für hyperbolische verteiltparametrische
Systeme (Observer based state feedback design for
hyperbolic distributed parameter systems)”. Ger- Abdurrahman Irscheid
Chair of Systems Theory and Control Engineering, Saarland
man. In: at - Automatisierungstechnik 60.8 (2012),
University, Germany
pp. 462–474. [email protected]
[35] F. Woittennek. “Flatness based feedback design for
hyperbolic distributed parameter systems with spa- Abdurrahman Irscheid received the B.Sc. degree in Mechatronics
and the M.Sc. degree with honors in Computational Engineering
tially varying coefficients”. In: IFAC Proceedings Vol-
of Technical Systems from Saarland University, Germany, in 2015
umes 46.26 (2013). 1st IFAC Workshop on Control of and 2017, respectively. He is currently working towards his Dr.-Ing.
Systems Governed by Partial Differential Equations degree at the Chair of Systems Theory and Control Engineering at
(CPDE), pp. 37–42. Saarland University. His main research topics concern the controller
[36] F. Woittennek. “On the hyperbolic observer canoni- and observer design for nonlinear distributed-parameter systems as
well as control-theoretic methods for convergence in prescribed finite
cal form”. In: Proc. 8th Int. Workshop on Multidi-
time.
mensional Systems (nDS’13). Erlangen, Germany,
2013, pp. 1–6.
[37] F. Woittennek, A. Irscheid, and N. Gehring. Joachim Deutscher
“Flatness-based analysis and control design for 2 × 2 Institute of Measurement, Control and Microtechnology, Ulm
hyperbolic PDEs with nonlinear boundary dynam- University, Germany
ics”. In: IFAC-PapersOnLine 55.26 (2022). 4th IFAC [email protected]
Workshop on Control of Systems Governed by Par- Joachim Deutscher received the Dr.-Ing. and the Dr.-Ing. habil. de-
tial Differential Equations (CPDE), pp. 13–19. grees both in automatic control from Friedrich-Alexander-Universität
[38] F. Woittennek and J. Rudolph. “Controller canoni- Erlangen-Nürnberg (FAU), Germany, in 2003 and 2010, respectively.
cal forms and flatness based state feedback for 1D In 2011 he was appointed Associate Professor at FAU and in 2017
hyperbolic systems”. In: IFAC Proceedings Volumes he became a Professor at the same university. Since April 2020 he
is a Full Professor at the Institute of Measurement, Control and
45.2 (2012). 7th Vienna Int. Conference on Mathe-
Microtechnology at Ulm University. His research interests include
matical Modelling (MATHMOD), pp. 792–797. control of distributed-parameter and multi-agent systems as well as
[39] M. Zeitz. “Canonical forms for nonlinear systems”. data-based control with applications in mechatronics. At present he
In: Nonlinear Control Systems Design 1989. Ed. by serves as Associate Editor for Automatica.
REFERENCES 23

Frank Woittennek
Institute of Automation and Control Engineering, UMIT TIROL –
Private University for Health Sciences and Health Technology,
Austria
[email protected]

Frank Woittennek received the Dr.-Ing. degree in electrical engineer-


ing from Technische Universität Dresden (TU Dresden), Germany, in
2007. Since 2015 he has been a Full Professor and the Head of the
Institute of Automation and Control Engineering at UMIT TIROL.
His current research interests include analysis, identification, control,
and observer design for distributed parameter systems and nonlinear
systems as well as applications in the fields of mechatronics, robotics,
process engineering, and energy systems.

Joachim Rudolph
Chair of Systems Theory and Control Engineering, Saarland
University, Germany
[email protected]

Joachim Rudolph received the doctorate degree from Université


Paris XI, Orsay, France, in 1991, and the Dr.-Ing. habil. degree
from Technische Universität Dresden, Germany, in 2003. Since
2009, he has been the Head of the Chair of Systems Theory and
Control Engineering at Saarland University, Saarbrücken, Germany.
His current research interests include controller and observer design
for nonlinear and infinite dimensional systems, algebraic systems
theory, and the solution of demanding practical control problems.

View publication stats

You might also like