A PIE Representation of Coupled Linear 2D PDEs and Stability Analysis using LPIs
Abstract
We introduce a Partial Integral Equation (PIE) representation of Partial Differential Equations (PDEs) in two spatial variables. PIEs are an algebraic state-space representation of infinite-dimensional systems and have been used to model 1D PDEs and time-delay systems without continuity constraints or boundary conditions – making these PIE representations amenable to stability analysis using convex optimization. To extend the PIE framework to 2D PDEs, we first construct an algebra of Partial Integral (PI) operators on the function space , providing formulae for composition, adjoint, and inversion. We then extend this algebra to and demonstrate that, for any suitable coupled, linear PDE in 2 spatial variables, there exists an associated PIE whose solutions bijectively map to solutions of the original PDE – providing conversion formulae between these representations. Next, we use positive matrices to parameterize the convex cone of 2D PI operators – allowing us to optimize PI operators and solve Linear PI Inequality (LPI) feasibility problems. Finally, we use the 2D LPI framework to provide conditions for stability of 2D linear PDEs. We test these conditions on 2D heat and wave equations and demonstrate that the stability condition has little to no conservatism.
1 INTRODUCTION
In this paper, we consider the problem of representation and stability analysis of linear Partial Differential Equations (PDEs) with multiple states evolving in 2 spatial dimensions.
First, consider how a PDE is defined. When we refer to a PDE, we are actually referring to 3 separate governing equations: The partial differential equation itself; a continuity constraint on the solution; and a set of boundary conditions (BCs). Any solution of the PDE is required to satisfy all three constraints at all times – leading to challenging questions of existence and uniqueness of solutions. Furthermore, suppose we seek to examine whether all solutions to a PDE exhibit a common evolutionary trait, such as stability or -gain. How does each of the 3 governing equations affect this property? The fact that we have 3 governing equations significantly complicates the analysis and control of PDEs.
For comparison, consider the stability question for a system defined by a linear Ordinary Differential Equation (ODE) in state-space form, , where the ODE itself is the only constraint on the solutions of the system. In this case, a necessary and sufficient (N+S) condition for stability of the solutions of the system is the existence of a quadratic measure of energy (Lyapunov Function (LF)), with , such that for any solution of the ODE, is decreasing for all . A N+S condition for stability will then be existence of a matrix such that for all . This constraint is in the form of a Linear Matrix Inequality (LMI), and can be solved using convex optimization algorithms [1].
Now, let us consider the problems with extending this state-space approach to stability analysis of 2D PDEs. In this case, the state at time of the PDE is a function of 2 spatial variables – raising the question of how one can parameterize the convex cone of LFs which are positive on this 2D function space without introducing significant conservatism. Moreover, we recall that solutions to the ‘PDE’ are required to satisfy not only the PDE itself (e.g. ), but are also required to satisfy both continuity constraints (e.g. ) and boundary conditions (e.g. ). The second problem is then, given a Lyapunov function of the form , how to determine whether for all satisfying all three constraints. In particular, how do the BCs and continuity constraints influence ?
Regarding the first problem, it is known that for linear PDEs, as was the case for ODEs, existence of a decreasing quadratic LF is N+S for stability of solutions [2] - so that we may assume the LF has the form for some positive operator . However, it is unclear how to parameterize a set of linear operators which is suitably rich so as to avoid significant conservatism, whilst still allowing positivity of the operators to be efficiently enforced. As a result, most prior work has been restricted to employing variations of the identity operator for . For example, in [3], a Lypanunov function of the form was used. Meanwhile, in [4], [5], and [6], the authors assumed to be a multiplier operator of the form , with positivity implied by the matrix inequality . In [7], the authors extended these functionals somewhat, using polynomial multipliers with the Sum-of-Squares (SOS) constraint . However, in each of these cases, the use of multiplier operators (analagous to the use of diagonal matrices in an LMI) implies significant conservatism in any stability analysis using such results.
We now turn to the second problem with stability analysis of PDEs: enforcing negativity of the derivative. As stated, for linear ODEs , this condition is easy to enforce: for all solutions and all if and only if is a negative definite matrix - an LMI constraint. However, for a PDE , the state space has more structure. That is, while it is true that if then , negativity of the operator is not a N+S condition for stability, as need only hold for solutions satisfying the BCs and continuity constraints. To account for this, in [4], [5], and [6], the authors consider particular types of BCs, allowing the use of known integral inequalities such as Poincare, Wirtinger, etc. to prove negativity. In [7], it was proposed to use a more general set of inqualities defined by Green’s functions. In all these cases, however, the process of identification and application of useful inequalities to prove negativity requires significant expertise and insight.
To avoid having to manually integrate boundary conditions and continuity constraints into the expression for , [8, 9, 10] suggest representing the PDE as a Partial Integral Equation (PIE). A PIE is a unitary representation of the PDE whose solution is defined on and hence does not require boundary conditions or continuity constraints.
For autonomous systems, PIEs are parameterized by the Partial Integral (PI) operators and , and take the form , where is the so-called “fundamental state”. This fundamental state is free of boundary and continuity constraints, and is associated to the PDE state through the transformation . PI operators form a Banach *-algebra, with analytic expressions for composition, adjoint, etc. Furthermore, positive PI operators can be parameterized by positive matrices – allowing us to solve Linear PI Inequality (LPI) optimization problems using semidefinite programming. Thus, the use of PIEs and PI operators also resolves the difficulty with parameterizing positive LFs, taking these to be of the form , where is a PI operator. In 1D, the Matlab toolbox PIETOOLS [8] can be used for parsing and solving LPI optimization problems.
Thus, through use of the LPI and PIE framework, stability analysis, as well as tasks like -optimal controller design [11], can be efficiently performed for almost any 1D PDE using convex optimization. However, as of yet, none of this architecture exists for 2D PDEs. Specifically, no concept of fundamental state has been defined for 2D PDEs, and there is no known algebra of 2D PI operators which could be used to parameterize a PIE representation, or be incorporated into some form of LPI optimization algorithm.
The goal of this paper, then, is to recreate the PIE framework for linear, 2nd order PDEs on a 2D domain with non-periodic boundary conditions. To this end, the paper makes the following contributions.
1) Identify an algebra of 2D PI operators.
In Section 3, we parameterize a set of PI operators with domain , which we combine with the algebra of 4-PI operators on (representing the boundary of the domain) to yield a Banach *-algebra of PI operators on . We demonstrate that this set of PI operators is closed under, addition, adjoint, and composition – deriving analytic expressions for the result of each operation. In Section 4, we further derive analytic expressions for the inverse of a suitable PI operator on , and the composition of a differential operator with a suitable 2D PI operator, showing that the result of each is a PI operator.
2) Identify the fundamental state for a 2D PDE.
In Subsection 6.1, we differentiate () the PDE state up to the maximal degree allowed per the continuity constraints. The resulting will then be free of boundary and continuity constraints.
3) Reconstruct the PDE state from the fundamental state.
Having defined a fundamental state , we isolate a set of “core” and ”full” boundary values of the PDE state as and respectively. Using the fundamental theorem of calculus, we can then express and , where , , and are 2D PI operators. Next, we impose the boundary conditions as , where is a PI operator, allowing us to write and thus . Finally, we retrieve the PDE state as
, where is a 2D PI operator.
4) Derive a PIE representation for a standardized PDE.
In Section 5, we present a standardized format for writing coupled PDEs. In Subsection 6.2, we then use the transformation and the composition rules for differential operators with PI operators, to derive an equivalent PIE representation as .
5) Derive an LPI stability test.
In Subsection 7.1, we parameterize a LF as a PI operator . Using this LF, we prove that existence of a PI operator satisfying the LPI certifies stability of the PDE.
6) Parameterize the convex cone of positive PI operators.
In Subsection 7.2, we introduce a PI operator , defined by monomial basis functions . For any positive matrix , then, the product will be a PI operator satisfying for any .
7) Implement this methodology in PIETOOLS.
In Section 8, we implement a class of 2D-PI operators in the MATLAB toolbox PIETOOLS 2021b, along with the formulae for constructing the PIE representation of a PDE. This allows an arbitrary PDE to be converted to an equivalent PIE, at which point stability may be tested by solving the LPI for a positive operator . Parameterizing the cone of positive operators as positive matrices, this problem may then be solved using semidefinite programming. We test this implementation on a heat equation and a wave equation in Section 9.
Parameter space | Explicit notation | Associated PI mapping |
---|---|---|
2 Notation
For a given domain and , let denote the set of -valued square-integrable functions on , with the standard inner product. is defined similarly and we omit the domain when clear from context. For any , we denote . Using this notation, we define as a Sobolev subspace of , where
As for , we occasionally use or when the domain is clear from context. For , we use the norm
For any , we denote the Dirac delta operators
3 Algebras of PI Operators
In [10], we parameterized an algebra of PI operators whose domain was functions of a single spatial variable, . These operators took the form
(1) | ||||
with polynomial parameters . In [9], these PI operators were generalized, yielding operators with domain . These extended operators then had the form
with parameters , and – where we note that the third parameter is itself a parameterized PI operator on the domain . In this paper, we are tasked with further generalizing the algebra of PI operators to functions of two spatial variables – i.e. . Operators in this rather more complicated algebra take the form of
(2) | ||||
where we now have 9 polynomial parameters . Making matters worse, we will also need to include cross-terms from , and in our algebra. In the following subsections, to make presentation of this class of operators somewhat tractable, we will make heavy use of an “operator parameterization of operators” approach, whereby we parameterize simpler algebras of operators using polynomials and embed these simpler algebras in the more complex ones. To aid in keeping track of these algebras, we use, e.g. the term 0112-PI algrebra to indicate its domain and range include (indicated by ‘0’), (indicated by ), (indicated by ), and (indicated by ).
Note that throughout this article we overload the notation for a PI operator with parameters , where the structure and parameterization of the operator varies depending on the specific parameter space . The different parameter spaces we consider are listed in Table 1.
3.1 An Algebra of 2D to 2D PI Operators
We start by parameterizing operators on , for which we define the parameter space
Then, for any , we let the associated 2D-PI operator be as in Eqn. (2). Defining addition and scalar multiplication of the parameters in the obvious manner, it immediately follows that , and , for any . Further defining multiplication of 2D-PI operators as in the following lemma, we conclude that the set of 2D-PI operators is an algebra.
3.2 An Algebra of 011-PI Operators
Having defined an algebra of operators on , we now consider a parameterization of operators on . To this end, we first let for any , the associated 1D-PI operator be as in (1). Next, we define a parameter space for 011-PI operators as
Then, for any , we let the associated PI operator be given by
where M is the multiplier operator and is the integral operator, so that (through some abuse of notation)
and
Clearly then, and for any and . Moreover, we can also compose 011-PI operators, as per the following lemma.
3.3 An Algebra of 0112-PI Operators
We now combine the 011-PI algebra and the 2D-PI algebra to obtain an algebra of operators on . Specifically, for any
where
we may define the 0112-PI operator as
where for , we have
and for , we have
Clearly, the set of operators parameterized in this manner is closed under addition and scalar multiplication. By the following lemma, then, the set of 0112-PI operators is an algebra.
Lemma 3
For the purpose of implementation in Section 8, we will be considering only PI operators parameterized by polynomial functions, exploiting the following result:
Corollary 4
For any polynomial parameters and , the composite parameters are also polynomial.
4 Useful Properties of PI Operators
Having defined the different algebras of PI operators to be used in later sections, we now derive several critical properties of such operators. Specifically, we focus on obtaining an analytic expression for the inverse of a 011-PI operator, using a generalization of the formula in [12]. This result will be used to enforce the boundary conditions when deriving the mapping from fundamental state to PDE state in Section 6. In addition, we consider composition of differential operators with a PI operator, proving that the result is a PI operator, and obtaining an analytic expression for this operator. This result will be used to relate differential operators in the PDE to PI operators in the equivalent PIE representation in Section 6. Finally, we give a formula for the adjoint of a PI operator, necessary for deriving and enforcing LPI stability conditions in Section 7.
4.1 Inverse of 011-PI Operators
First, given a 011-PI operator defined by parameters , we prove that is a 011-PI operator and obtain an analytic expression for the parameters of the inverse . As is typical, we restrict ourselves to the case where the operator has a separable structure and the parameters are polynomial (and hence have a finite-dimensional parameterization).
Lemma 5
Proof 4.1.
See Appendix 12.1 for a proof.
4.2 Differentiation and 2D-PI Operators
While differentiation is an unbounded operator and PI operators are bounded, we now show that for a PI operator with no multipliers, composition of the differential operator with this PI operator is a PI operator and hence bounded. This will be used in Section 6 to show that the dynamics of a PDE can be equivalently represented using only bounded operators (the PIE representation).
Lemma 6.
Suppose
and let
(3) |
where
Then, for any ,
Proof 4.2.
To prove this result, we exploit the linearity of the PI operator, splitting
Recall now the Leibniz integral rule, stating that, for arbitrary ,
Then, for arbitrary ,
Repeating these steps, it also follows that
Finally, by linearity of the derivative and PI operators,
Lemma 7.
Suppose
and let
(4) |
where
Then for any ,
Proof 4.3.
The proof follows along the same lines as that of Lemma 6.
4.3 Adjoint of 2D-PI Operators
Finally, we give an expression for the adjoint of a 2D-PI operator.
Lemma 8.
Suppose and define such that
(5) |
Then for any and ,
Proof 4.4.
See Appendix 12.2 for a proof.
5 A Standardized PDE Format in 2D
Having formulated the required hierarchy of algebras of PI operators, we now introduce a class of linear 2D PDEs, the solutions of which may be represented using 2D PIEs. These 2D PDEs are represented in a standardized format, allowing for efficient construction of a general mapping between the PDE and PIE state spaces – this construction being found in Section 6.
We consider a coupled PDE in the following compact representation
(6) |
where the matrices
allow us to partition the states according to differentiability, so that for all , where also includes boundary conditions, parameterized by the 011-PI operator as
(7) |
where and where allows us to list all the possible boundary values for the state components and , and as limited by differentiability. In particular,
(8) |
where
where we use the Dirac operators defined as
This formulation is very general and allows us to express almost any linear 2D PDE. For examples of this representation, see the examples in Section 9.
Definition of Solution
For a given initial condition , we say that a function satisfies the PDE defined by if is Frechét differentiable, , for all and Equation (6) is satisfied for all .
Definition 9.
We say that a solution with initial condition of the PDE defined by is exponentially stable in if there exist constants such that
We say the PDE defined by is exponentially stable in if any solution of the PDE is exponentially stable in .
6 The Fundamental State on 2D
In this section, we provide the main technical result of the paper, wherein we show that for a suitably well-posed set of boundary conditions, , there exists a unitary 2D-PI operator (where is defined in Eqn. (7)) such that if we define the differentiation operator
(9) |
then for any and , we have
This implies that for any , there exists a unique where the map from to is defined by a 2D-PI operator. Because differentiation of a PI operator is a PI operator (Section 4.2), this implies that derivatives of can be expressed in terms of a PI operator acting on . Using these results, in Thm. 13, we show that for any suitable PDE defined by , there exist 2D-PI operators such that satisfies
if and only if satisfies the PDE.
6.1 Map From Fundamental State to PDE State
As mentioned above, given the boundary-constrained “PDE state” , we associate a corresponding “fundamental state” , defined as
(12) |
In the following lemma, we temporarily ignore boundary conditions and use the fundamental theorem of calculus to express any in terms of , and a set of “core” boundary values.
Proof 6.1.
The proof follows directly from the identities
and
which follow from the fundamental theorem of calculus.
Corollary 11.
With these definitions, we can express an arbitrary PDE state in terms of a corresponding state and . In the following theorem, we describe this relation as a PI operator, incorporating the boundary conditions to describe a map from to . In doing so, we will require the operator to be well-posed, defining sufficient boundary conditions for the solution to the PDE to be uniquely defined. We express this restriction through invertibility of a 011-PI operator.
Definition 12.
Define
(13) |
where
with the functions as defined in (10), and | ||||
(22) | ||||
with | ||||
(26) | ||||
and , where is defined as in Equation (54) in 12.1, and | ||||
(33) |
where the functions are as defined in (11).
Theorem 13.
Proof 6.2.
Suppose , and define . Furthermore, let and be as defined in Eqn. (10), and and be as defined in Eqn. (11), such that (by Lemma 10 and Corollary 11)
(34) |
where , , , and . Enforcing the boundary conditions , we may use the composition rules of PI operators to express
where and , with
and
defined as in Equations (26) and (33). By well-posedness of the boundary conditions, operator is invertible, so that the boundary state may be expressed directly in terms of the fundamental state as
where with
defined as in Equations (22). Finally, substituting this expression into Equation (6.2), we obtain
as desired.
Corollary 14.
Let be as defined in Theorem 13. Then is unitary with respect to
Proof 6.3.
By Thm. 13, for any , there exists such that , hence is surjective. Furthermore, for any ,
concluding the proof.
6.2 PDE to PIE conversion
We now demonstrate that, given a PDE defined by , for appropriate choice of , we may define a Partial Integral Equation (PIE) whose solutions are equivalent to those of the PDE. Specifically, for given PI operators , and an initial condition , we say solves the PIE defined by for initial condition if , for all and for all
(35) |
The following result shows that if is as defined in Theorem 13, then satisfies the PDE defined by if and only if satisfies the PIE defined by .
Lemma 15.
Proof 6.4.
Specific examples of PDEs and their PIE equivalents are given in Section 9. In the following section, we propose stability conditions for the PIE which can be enforced using LMIs.
7 Stability as an LPI
Having derived an equivalent PIE representation of PDEs, we now show how this representation can be used for stability analysis. First, we show that existence of a quadratic Lyapunov function for a PIE can be posed as a convex Linear PI Inequality (LPI) optimization problem, with variables of the form for , and inequality constraints of the form . Next, we show how to use LMIs to parameterize the cone of positive semidefinite 2D-PI operators - allowing us to test the Lyapunov stability criterion. Finally, we will discuss a PIETOOLS numerical implementation of this stability test, which will be applied to several numerical examples in Section 9.
7.1 Lyapunov Stability Criterion
We first express the problem of existence of a quadratic Lyapunov function as an LPI, whose feasibility implies stability of the associated PIE and PDE. Specifically, the following theorem tests for existence of a quadratic Lyapunov function of the form such that for any solution of the PDE defined by .
Theorem 16.
Proof 7.1.
Let be an arbitrary solution to the PDE defined by , so that is a solution to the PIE defined by . Consider the candidate Lyapunov function defined as
Since , this function is bounded from above as
In addition, since is a solution to the PIE, the temporal derivative of along satisfies
Applying the Grönwell-Bellman inequality, it immediately follows that
implying
and thus
In this stability condition, the decision variable is and the constraints are operator inequalities on the inner product space . While the decision variables may be readily parameterized using polynomials, to numerically enforce the inequality constraints, we need to parameterize the cone of operators in which are positive semidefinite. This problem will be addressed in the following subsection.
7.2 A Parameterization of Positive PI Operators
Having posed the PDE stability problem as an LPI, we now show how to parameterize the cone of positive 2D-PI operators using positive matrices. Specifically, we have the following result.
Proposition 17.
Outline of Proof
Given and scalar function , a 2D-PI operator is defined as in Eqn. (91), where each of the defining parameters is a product of and . Then, if for some matrix , by definition of the map , the associated PI operator is such that . It follows that, for any ,
for any , as desired. For the full proof of this proposition, please see Appendix 12.4.
Using Prop. 17, we may enforce an LPI with variable using an LMI constraint on . For this test, we use a monomial basis, for , thus implying the parameters in will be polynomial. For our choice of , we may choose (implying the inequality is valid over any domain), or , implying the operator is positive only on the domain .
In the following section, we will apply Prop 17 to obtain an LMI for stability of a given PDE. For this section we use the notation
where now is an LMI constraint which implies is a positive operator on .
8 PIETOOLS Implementation
In this section, we show how the PIETOOLS 2021b toolbox may be used to perform stability analysis of PDEs. This toolbox offers a framework for implementation and manipulation of PI operators in MATLAB, allowing e.g. Lyapunov stability analysis [9], robust stability analysis [13], and -optimal control [11] of PDEs. For a detailed manual of the PIETOOLS toolbox we refer to [8].
To implement PI operators in MATLAB, the dpvar class of polynomial objects is used to define the polynomial functions parameterizing PI operators . A class of 0112-PI operators is then defined as opvar2d objects, and overloaded with standard operations such as multiplication (*), addition (+) and adjoint (’) presented in earlier sections. Defining decision operators dopvar2d in terms of positive matrices, we may also enforce positivity conditions , allowing stability to be tested with any LMI solver.
An overview of the steps performed in this process is provided below.
-
1.
Define the independent polynomial variables. These are the spatial variables in the PDE. Also define the “dummy” variables and .
pvar x y tt nu;
-
2.
Initialize an optimization program structure X.
X = sosprogram([x y tt nu]);
-
3.
Construct the PDE, defining the sizes of the state variables, the matrices defining the PDE, and an opvar2d object Ebb defining . Convert these to a corresponding PIE using convert_PIETOOLS_PDE, and extract opvar2d objects and .
PDE.n.n_pde = [n0,n1,n2]; PDE.dom = [a,b;c,d]; PDE.PDE.A = ...;ΨPDE.BC.Ebb = ...; PIE = convert_PIETOOLS_PDE(PDE); T = PIE.T; A = PIE.A;
-
4.
Declare a positive operator as a dopvar2d object, using maximal monomial degree , and spatial domain dom. Add a small constant to ensure strict positivity. Impose the additional requirement for some .
[X, P] = poslpivar_2d(X,n,dom,d); P = P + eps; D = -del*(T’*T) - A’*P*T - T’*P*A; X = lpi_ineq_2d(X,D);
-
5.
Call the SDP solver.
X = sossolve(X);
-
6.
Get the solution , certifying stability.
Psol = getsol_lpivar_2d(X,P);
9 Illustrative Examples
To illustrate the techniques described in the previous sections, we will apply them to several simple examples. In each case, we will show how the problem may be expressed in the standardized format, as well as how the corresponding PIE is defined, and we will numerically test stability.
9.1 Heat Equation
As a first example, we consider a 2D heat equation
To describe this system in the standardized Format (6), we use PDE state with , and define , setting for all other . To enforce the boundary conditions, we require , as well as , which can be expressed as where
Here we note that, any matrix may be represented as a diagonal PI operator through appropriate choice of the parameters , and , so that the boundary conditions may also be written in the standardized format . Then, we may describe the system as a PDE defined by (or really ), for which we obtain a corresponding PIE representation
where is the fundamental state.
For the purpose of testing accuracy of the stability analysis, we consider the following system, as presented in [14],
where are positive constants. This system may be represented in the standardized format by letting , as well as , with all other matrices . For the boundary conditions, we require , as well as , which may be enforced as , where
Implementing this system in PIETOOLS 2021b, we obtain a PIE representation
where we define
Letting and , stability of this system can be proven analytically whenever . Using PIETOOLS 2021b, performing a bisection search over and using monomial degree , exponential stability can be verified for any .
9.2 Wave Equation
As a second example, we consider a 2D wave equation
To write this system in the standardized form, we define and , so that the PDE may be denoted as
Requiring , and , we may write the boundary conditions as , where
Letting , the associated PIE representation is
where is the fundamental state. Simulation suggests the system is neutrally stable ( in Defn. 9) with the alternative boundary conditions . Setting (see Thm. 16), and using a monomial degree , this can be verified with PIETOOLS.
9.3 Coupled ODE-PDE System
As a final example, we consider a diffusion equation coupled to an ODE, a s appearing in [15],
which is stable whenever the matrix is Hurwitz. An equivalent PIE representation of this system may be obtained by first deriving a PIE representation of the PDE subsystem. To this end, we use as PDE state with , and let to describe the PDE in the standardized format (6). For the boundary conditions, we enforce , as well as , enforcing the conditions as , where
Using PIETOOLS 2021b , the associated PIE representation is found to be
where . Now, to incorporate the ODE dynamics, we note that the boundary value may be written in terms of the PIE state as
allowing the ODE dynamics to described using 0112-PI operators as
where . In this representation, stability analysis can also be performed as discussed in Section 7, using a Lyapunov function based on the 0112-PI operators describing the system. This was done for a simple scalar case where , and , in which case stability can be proven analytically whenever . Using degree and performing bisection on , stability could be verified for .
10 Conclusion
In this paper, we have shown that any well-posed, linear, second order 2D PDE can be converted to an equivalent PIE, and we have provided the formulae describing this conversion. To derive these formulae, we have introduced different PI operators in 2D, showing that the product, inverse, adjoint, and composition with differential operators of these operators are described by PI operators as well. Exploiting these relations, we derived a mapping between the PDE state , constrained by the boundary conditions as described by , and a fundamental state , free of any such constraints. Accordingly, the solution to the PIE is not constrained by boundary or continuity constraints, which allowed us to derive LPI conditions for stability of the system. Finally, by paramaterizing PI operators as matrices, we showed how stability may be tested using semidefinite programming, as implemented in the MATLAB toolbox PIETOOLS 2021b.
Having demonstrated that the PIE representation can be extended to 2D PDEs, an obvious next step would be to extend it to 3D, and possibly ND systems. In addition, we may expand upon our results to include more complicated systems, such as those involving higher order spatial derivatives or alternative boundary conditions. Moreover, different LPIs may also be introduced for -gain analysis and (robust) controller synthesis, as was done for 1D systems. Finally, alternative spatial domains may also be considered, extending the PIE methodology to more complex geometries.
References
- [1] S. Boyd, L. El Ghaoui, E. Feron, and V. Balakrishnan, Linear matrix inequalities in system and control theory. SIAM, 1994.
- [2] R. F. Curtain and H. Zwart, An Introduction to Infinite-Dimensional Linear Systems Theory. Berlin, Heidelberg: Springer-Verlag, 1995.
- [3] M. Ahmadi, G. Valmorbida, D. Gayme, and A. Papachristodoulou, “A framework for input-output analysis of wall-bounded shear flows,” arXiv preprint arXiv:1802.04974, 2018.
- [4] E. Fridman and M. Terushkin, “New stability and exact observability conditions for semilinear wave equations,” Automatica, vol. 63, pp. 1–10, 2016.
- [5] O. Solomon and E. Fridman, “Stability and passivity analysis of semilinear diffusion PDEs with time-delays,” International Journal of Control, vol. 88, no. 1, pp. 180–192, 2015.
- [6] M. Wakaiki, “An LMI approach to stability analysis of coupled parabolic systems,” IEEE Transactions on Automatic Control, vol. 65, no. 1, pp. 404–411, 2019.
- [7] G. Valmorbida, M. Ahmadi, and A. Papachristodoulou, “Convex solutions to integral inequalities in two-dimensional domains,” in 2015 54th IEEE Conference on Decision and Control (CDC). IEEE, 2015, pp. 7268–7273.
- [8] S. Shivakumar, A. Das, and M. M. Peet, “PIETOOLS: A MATLAB toolbox for manipulation and optimization of partial integral operators,” in 2020 American Control Conference (ACC). IEEE, 2020, pp. 2667–2672.
- [9] S. Shivakumar, A. Das, S. Weiland, and M. M. Peet, “A generalized LMI formulation for input-output analysis of linear systems of ODEs coupled with PDEs,” in 2019 IEEE 58th Conference on Decision and Control (CDC). IEEE, 2019, pp. 280–285.
- [10] M. M. Peet, S. Shivakumar, A. Das, and S. Weiland, “Discussion paper: A new mathematical framework for representation and analysis of coupled PDEs,” IFAC-PapersOnLine, vol. 52, no. 2, pp. 132–137, 2019.
- [11] S. Shivakumar, A. Das, S. Weiland, and M. M. Peet, “Duality and -optimal control of coupled ODE-PDE systems,” in 2020 59th IEEE Conference on Decision and Control (CDC). IEEE, 2020, pp. 5689–5696.
- [12] G. Miao, M. M. Peet, and K. Gu, “Inversion of separable kernel operator and its application in control synthesis,” in Delays and Interconnections: Methodology, Algorithms and Applications. Springer, 2019, pp. 265–280.
- [13] A. Das, S. Shivakumar, M. M. Peet, and S. Weiland, “Robust analysis of uncertain ODE-PDE systems using PI multipliers, PIEs and LPIs,” in 2020 59th IEEE Conference on Decision and Control (CDC), 2020, pp. 634–639.
- [14] E. E. Holmes, M. Lewis, J. Banks, and R. R. Veit, “Partial differential equations in ecology: Spatial interactions and population dynamics,” Ecology, vol. 75, pp. 17–29, 1994.
- [15] M. Krstic, “Compensating actuator and sensor dynamics governed by diffusion PDEs,” Systems & Control Letters, vol. 58, no. 5, pp. 372–377, 2009.
- [16]
11 Algebras of PI Operators
11.1 Notation
In this section, we will prove the composition rules of PI operators in 2D, as outlined in the article. In describing these results, we will (for the sake of compactness) use subscripts and superscripts to denote the free variables of a function , so that:
For the sake of simplicity, we will also let , though the results presented in this section extend to any domain . We denote the integral of a function accordingly as:
In addition, we will rely on the functions , , and to a lesser extent , to limit the domains of integration in subsequent results. Here, denotes a Dirac delta function, such that, for any ,
for . Similarly, denotes an indicator function,
so that
and
Finally, we let denote a “rectangular” function,
so that
Based on these definitions, the following identities follow trivially:
(37) |
11.2 Preliminaries
Using the notation introduced in the previous subsection, we may compactly define different parameterizations of PI operators, as well as their compositions. For example, recall that for , we define an associated PI operator as
for arbitrary . Using the Dirac delta function , and the indicator function as introduced earlier, we may equivalently denote this operation as
Then, defining functions for as
(38) |
we can describe the operation using a single sum as
This notation will allow us to compactly write the composition rules for different PI operators, using the following corollary.
Corollary 18.
Let for be as defined in Eqn. 38. Then, for any ,
In addition, we will rely heavily on the following proposition.
Proposition 19.
Proof 11.1.
Expanding the sums, invoking the definition of , and applying identities (11.1), we find
as desired.
Applying this result, the composition rules of 1D-PI operators follow immediately.
Lemma 20.
For any and , there exists a unique such that . Specifically, we may choose , where the linear parameter map is such that
where, defining functions as in Eqn. (19),
for each .
Proof 11.2.
Let be arbitrary. Then, applying the results from Proposition 19, we find
Expanding the terms in this expression for the composition, it is easy to verify that these composition rules for 1D-PI operators match those (for 3-PI operators) presented in [16]. Using this same approach, we can also derive composition rules for PI operators on additional dimensions.
11.3 An Algebra of 011-PI Operators
Recall that we defined a parameter space for 011-PI operators as
(40) |
Then, for any
we may write the associated PI operation using the functions as defined in Eqn. (38) as
for any .
Lemma 21.
For any and , where , and , , there exists a unique such that . Specifically, we may choose , where the linear parameter map is defined such that
(41) |
where
and where is defined such that, for arbitrary , , where
and is defined such that, for arbitrary and , , where
Proof 11.3.
To prove this lemma, we will exploit the linear structure of 011-PI operators, allowing us to express
Then, maps a vector , and functions and to as
and
and finally
Similarly, such states get mapped to functions in as
and
and finally
Similarly, we attain a mapping to as
and
and finally
Combining these results, we conclude .
11.4 An Algebra of 2D-PI Operators
We now consider 2D-PI operators, acting on functions . Defining a parameter space for these operators as
(42) |
for any
we may write the associated PI operation using the functions as defined in Eqn. (38) as
Lemma 22.
For any and , there exists a unique such that . Specifically, we may choose , where the linear parameter map is such that
(43) |
where, defining functions as in Eqn. (19),
for each .
Proof 11.4.
Let be arbitrary. Then, applying the results from Proposition 19, we find
11.5 An Operator from 2D to 011
Having defined algebras of operators on and , it remains to define operators mapping and back. For this first mapping, we first define a parameter space
with associated operator defined such that, for arbitrary ,
This operator maps functions on two variables to functions on a single variable, allowing us to build a mapping from to . In particular, letting
(44) |
an associated operator may be defined such that, for any and ,
This operator allows us to map functions to functions , which is necessary to map state variables living on the interior of a 2D domain to state variables living on the boundary. An important property of this operator is also that its composition with 011- and 2D-PI operators returns another 2D011-PI operator, as described in the following lemmas.
Lemma 23.
For any and , where
there exists a unique such that . Specifically, we may choose , where the linear parameter map is defined such that
(45) |
with
where is defined such that, for arbitrary , , with
is defined such that, for arbitrary , , with
for , and is defined such that, for arbitrary , , , where
Proof 11.5.
To prove this lemma, we will exploit the linear structure of 011-PI operators, allowing us to express
Considering each of these terms separately, we may invoke the definitions of the different operators to find that, for arbitrary ,
In addition, using Corollary 18 and Proposition 19, it follows that
Finally, by the same approach,
Combining the results, we conclude that .
Lemma 24.
For any and , where
there exists a unique such that . Specifically, we may choose , where the linear parameter map is defined such that
(46) |
where
and where
with
for .
Proof 11.6.
To prove this result, we once again note that, by linearity of the PI operators,
Considering each of these terms separately, we find that, for arbitrary ,
Similarly, using Proposition 19, we find
and
Combining the results, we conclude that .
11.6 An Operator from 011 to 2D
In addition to the operator mapping to defined in the previous section, we also define an operator performing the inverse of this mapping. For this, we first define a parameter space
with associated operator defined such that, for arbitrary and ,
Building upon this, we define yet another space
(47) |
with associated PI operator such that, for arbitrary and ,
This operator maps functions in to functions , allowing us to map state variables living on the boundary of a 2D domain to state variables living on its interior. As was the case for 2D011-PI operators, the composition of 0112D-PI operators with 011- and 2D-PI operators too can be expressed as a PI operator, as described in the following lemmas.
Lemma 25.
For any and , where
there exists a unique such that . Specifically, we may choose , where the linear parameter map is defined such that
(48) |
where
and where is defined such that, for arbitrary , , with
and is defined such that, for arbitrary , , with
for , and is defined such that, for arbitrary , , , where
Proof 11.7.
To prove this lemma, we will exploit the linear structure of 011-PI operators, allowing us to express
Considering each of these terms separately, we find that for an arbitrary function ,
Furthermore, using Corollary 18 and Proposition 19, we find that
Finally, performing the same steps
Combining the results, we conclude that .
Lemma 26.
For any and , where
there exists a unique such that . Specifically, we may choose , where the linear parameter map is defined such that
(49) |
where
and
with
for .
Proof 11.8.
Applying Proposition 19, and the definitions of the operators and , it follows that, for arbitrary ,
as desired.
In addition to these compositions with 011- and 2D-PI operators, the compositions of 0112D-PI and 2D011-PI can also be expressed as PI operators, as described in the following lemmas.
Lemma 27.
For any and , where
there exists a unique such that . Specifically, we may choose , where the linear parameter map is defined such that
(50) |
where
and where
with
for .
Proof 11.9.
To prove this result, we exploit the linear structure of the PI operators, allowing us to decompose
Focusing first on the terms involving , we find that, for arbitrary ,
and
and finally,
Similarly, for the terms involving , we find that, for arbitrary ,
and
and finally, using Proposition 19,
Finally, for the terms involving , for arbitrary ,
and
and, once more using Proposition 19,
Combining the results, we conclude that .
Parameter space | Explicit notation | Associated PI operation |
---|---|---|
Lemma 28.
For any and , where
there exists a unique such that . Specifically, we may choose , where the linear parameter map is defined such that
(51) |
where , and
Proof 11.10.
Linear Parameter Map | Associated Parameter Spaces | Defined in |
---|---|---|
Equation (41), Appendix 11.3 | ||
Equation (43), Appendix 11.4 | ||
Equation (45), Appendix 11.5 | ||
Equation (46), Appendix 11.5 | ||
Equation (48), Appendix 11.6 | ||
Equation (49), Appendix 11.6 | ||
Equation (50), Appendix 11.6 | ||
Equation (51), Appendix 11.6 |
11.7 An Algebra of 0112-PI Operators
Having described PI operators mapping functions in and and the corresponding composition rules, we can now combine our results to describe a 0112-PI operator, mapping functions in . Letting
(52) |
for arbitrary
we define an associated PI operator as
where the different operators ,…, are as defined in the previous sections. Using the results from these sections, it is also easy to see that the set of operators parameterized in this manner also forms an algebra.
Theorem 29.
For any and , there exists a unique such that . Specifically, we may choose , where the linear parameter map defined such that
(53) |
where
where the different parameter maps are listed in Table 3.
Proof 11.11.
Exploiting the linear structure of the 0112-PI operator, and applying the results from the previous sections, it follows that
12 Additional Proofs
12.1 Inverse of 011-PI Operators
Lemma 30.
with | (54) |
and with
where
and where
with
Proof 12.1.
Let be as defined. Then, by the composition rules of PI operators, we have , where
with and , and where
and
To demonstrate that the operator defines an inverse of the operator , we show that describes an identity operation on . In particular, we show that each of the functions is constantly equal to zero, except for , and , which are identity matrices of appropriate sizes. To this end, we first note that and , from which it immediately follows that and . To see that also , we expand each of the terms in its definition, obtaining
Adding these terms, we immediately find .
For the remaining functions, we also expand the different terms in their definitions. Starting with , we find
from which it follows that
Expanding the terms in the expression of in the same way, we obtain
suggesting also
Next, we consider the expression for , for which
Adding these terms, it is clear that also .
Similarly, for ,
We note that each of these terms may be described as a constant matrix, premultiplied by the function , and postmultiplied by the function . Hence, we may also express for some matrix . In particular, adding the different terms, we find this matrix to be given by
proving that also . Finally, for , we find
Studying these terms, it is clear that (similar to ) we may express , where
from which it follows that also .
Performing the same steps as for , and , we can also show that , and are constantly equal to zero. This leaves only , and as nonzero functions defining . By definition of the 011-PI operator, it immediately follows that for any , proving the desired result.
12.2 Adjoint of 2D-PI operator
Lemma 31.
Suppose and define such that
(55) |
Then for any and ,
Proof 12.2.
Let and be as defined. Recall the definition of the indicator function
and let the Dirac delta function be defined such that, for any ,
Then, as described in Appendix 11.4, defining
we may describe the 2D-PI operation as
Noting that , and , it immediately follows that
as desired.
12.3 Map From Fundamental to PDE State
For the following theorem, recall the definition
(56) |
for the space of solutions to a standardized PDE. In particular, recall that any such solution must satisfy the boundary conditions , where describes the solution along the boundary, and for some .
(57) |
where
(58) |
and
(67) |
with
(68) |
and , where is defined as in Equation (54), and and , with
(69) |
where,
(76) |
Theorem 32.
Proof 12.3.
We will first proof the first identity, . To this end, suppose , and define . Furthermore, let and (for appropriate ) be as defined in Equations (4) and (4), and let
where
Then, by Lemma 10 and Corollary 11,
(77) |
where , , , and . Enforcing the boundary conditions , we may use the composition rules of PI operators to express
where and with
and
defined as in Equations (4) and (4). Assuming is of sufficient rank, we can then invert the 011-PI operator , allowing us to write
where with
defined as in Equations (4). Finally, substituting this expression into Equation (12.3), and once more using the composition rules of PI operators, we obtain
as desired.
Suppose now and . To prove the second identity, , we split into its different components , and , using the matrices
and
so that
By definition of the operator , we then have to prove that
(78) | ||||
(79) | ||||
(80) |
To prove Relation (78), we use the definition of the parameters (Equations (4)), suggesting , whilst for . It immediately follows that
For the remaining relations, we first note that , so that
where we define for . Since the multiplier terms ( and ) of this PI operator are all zero, by Lemmas 6 and 7, we may write the composition of the differential operator with this PI operator as a PI operator,
where,
with , , and .
Studying Equations (4) defining , and , it is easy to see that , and therefore also
so that
For the remaining parameters , we consider the products . Once more studying Equations (4), we note that , , and are all constant matrices. By the definitions (Equations (4)) of functions , for each , this implies also
Finally, for , it is easy to see that
Combining the results, we obtain
proving Relation (79). To prove the final Relation (80), we note that , and therefore
where we define for . The resulting operator contains no multiplier terms, allowing once more the composition with the differential operator to be taken, yielding (by Lemmas 6 and 7)
where
with , , and . Once again, it is clear from the definitions of functions , and that the derivatives , and , , evaluated at and respectively, will all be equal to zero, suggesting
In addition, studying the definitions of the functions , , , and , we find that their derivatives , , , and are all constant matrices. By definition of the functions and , it follows that for each , and thus
Finally, from the definition of , it follows that
Combining these results, we obtain Relation (80)
12.4 A Parameterization of Positive PI Operators
Proposition 33.
Proof 12.4.
Let be an arbitrary matrix of appropriate size, and . It is easy to see that, by definition of the functions , the PI operator defined by is self-adjoint. Furthermore, if we define a 2D-PI operator as
(91) |
by the composition rules of 2D-PI operators (as discussed in Appx. 11.4), it follows that . Since , we may split for some , and thus
for any , concluding the proof.