This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

A Note on BIBO Stabilitythanks: The research leading to these results has received funding from the Swiss National Science Foundation under Grant 200020-162343/1.

Michael Unser Biomedical Imaging Group, École polytechnique fédérale de Lausanne (EPFL), Station 17, CH-1015 Lausanne, Switzerland ([email protected]).
Abstract

The statements on the BIBO stability of continuous-time convolution systems found in engineering textbooks are often either too vague (because of lack of hypotheses) or mathematically incorrect. What is more troubling is that they usually exclude the identity operator. The purpose of this note is to clarify the issue while presenting some fixes. In particular, we show that a linear shift-invariant system is BIBO-stable in the LL_{\infty}-sense if and only if its impulse response is included in the space of bounded Radon measures, which is a superset of L1()L_{1}(\mathbb{R}) (Lebesgue’s space of absolutely integrable functions). As we restrict the scope of this characterization to the convolution operators whose impulse response is a measurable function, we recover the classical statement.

I Introduction

A statement that is made in most courses on the theory of linear systems as well as in the English version of Wikipedia111 https://en.wikipedia.org/wiki/BIBO_stability. Accessed November 2019. is that a convolution operator is stable in the BIBO sense (bounded input and bounded output) if and only if its impulse response is absolutely summable/integrable. While the proof of this equivalence is fairly straightforward for discrete-time systems, there seems to be some confusion in the continuous domain (see Appendix B for specific references), especially since the above statement excludes the identity operator, whose impulse response is the Dirac distribution δ\delta. Since δ\delta is not a measurable function in the sense of Lebesgue (see explanations in Appendix A) and hence not a member of L1()L_{1}(\mathbb{R}), does this mean that the identity operator is not BIBO-stable? Obviously not; this is what we want to clarify here. The argument, which is somewhat technical, rests on the shoulders of two giants: Laurent Schwartz and Lars Hörmander, who were awarded the Fields medal in 1950 and 1962, respectively, for their fundamental contributions to the theory of distributions and partial differential equations.

In the sequel, we shall revisit the topic of BIBO stability with the help of appropriate mathematical tools. In Section II, we recall the classical integral definition of a convolution operator. We then present a correction to the standard characterization of BIBO-stable filters (Proposition 1) together with a new upgraded proof. Since the underlying assumption that the impulse response should be a measurable function excludes the identity operator, we first explain in Section III the extended (distributional) form of convolution supported by Schwartz’ kernel theorem (Theorem 1). Based on this formalism, we present two Banach-space extensions of the classical result that should settle the issue: a first one (Theorem 2) that imposes that the result of the convolution be continuous, and a second (Theorem 3) that characterizes the BIBO-stable filters in full generality. The mathematical derivations are presented in Section IV, where we also make the connection with known results in harmonic analysis.

We like to mention a similar clarification effort by Hans Feichtinger, who proposes to limit the framework to convolution operators that are operating on C0()C_{0}(\mathbb{R}) (a well-behaved subclass of bounded functions) in order to avoid pathologies [3]. This is another interesting point of view that is complementary to ours, as discussed in Section IV.

II BIBO Stability: The Classical Formulation

The convolution of two functions h,f:h,f:\mathbb{R}\to\mathbb{R} is the function usually specified by

t(hf)(t)=h(τ)f(tτ)dτ\displaystyle t\mapsto(h\ast f)(t)\stackrel{{\scriptstyle\vartriangle}}{{=}}\int_{\mathbb{R}}h(\tau)f(t-\tau){\rm d}\tau (1)

under the implicit assumption that the integral in (1) is well defined for any tt\in\mathbb{R}. This latter point will be clarified as we develop the mathematics. In particular, this requires that the functions ff and hh both be measurable222 A function f:f:\mathbb{R}\to\mathbb{R} is said to be Lebesgue-measurable if the preimage f1(E)f^{-1}(E) of any Borel set EE in \mathbb{R} is a Borel set [15]. The property is preserved through pointwise multiplication and translation. in the sense of Lebesgue. Here, instead of designating the continuous-time signal by f(t)f(t) and its convolved (or filtered) version by h(t)f(t)h(t)\ast f(t), as engineers usually do, we are using the less ambiguous mathematical notations tf(t)t\mapsto f(t) or fLp()f\in L_{p}(\mathbb{R}) and t(hf)(t)t\mapsto(h\ast f)(t) or hfL()h\ast f\in L_{\infty}(\mathbb{R}).

If we fix hh and consider ff as the input signal, then (1) defines a linear shift-invariant (LSI) operator (or system) denoted by Th:fhf{\mathrm{T}}_{h}:f\mapsto h\ast f. Its impulse response hh is then formally described as h=Th{δ}h={\mathrm{T}}_{h}\{\delta\}, where δ𝒟()\delta\in{\mathcal{D}}^{\prime}(\mathbb{R}) is the Dirac distribution and 𝒟(){\mathcal{D}}^{\prime}(\mathbb{R}) Schwartz’ space of distributions [17]. This interpretation is backed by Schwartz’ kernel theorem, as explained in Section III-A.

An important practical requirement for an LSI system is that its response to any bounded input remains bounded. There is one mathematical aspect, however, that makes the formulation of BIBO stability nontrivial in the continuous domain: Depending on the context, the input and output boundedness requirements can be strict, with fL=fsup=supt|f(t)|<{\|f\|_{L_{\infty}}}=\|f\|_{\sup}\stackrel{{\scriptstyle\vartriangle}}{{=}}\sup_{t\in\mathbb{R}}|f(t)|<{\infty}, which arises when the function ff is continuous (i.e., fC()f\in C(\mathbb{R})), or in the looser sense of Lebesgue: |f(t)|fL<|f(t)|\leq\|f\|_{L_{\infty}}<\infty for almost any tt\in\mathbb{R} (see Section III-B for additional explanations). This latter condition is often expressed as fL()f\in L_{\infty}(\mathbb{R}) where L()={f:s.t.f is measurable and fL<}L_{\infty}(\mathbb{R})=\{f:\mathbb{R}\to\mathbb{R}\ \ \mbox{s.t.}\ \ f\mbox{ is measurable and }\|f\|_{L_{\infty}}<\infty\} is Lebesgue’s space of bounded functions.

Definition 1.

The linear operator T:fT{f}{\mathrm{T}}:f\mapsto{\mathrm{T}}\{f\} is said to be BIBO-stable if

  1. 1.

    T{f}{\mathrm{T}}\{f\} is well-defined for any fL()f\in L_{\infty}(\mathbb{R}), and,

  2. 2.

    there exists a constant C>0C>0 independent of ff such that

    T{f}LCfL\|{\mathrm{T}}\{f\}\|_{L_{\infty}}\leq C\|f\|_{L_{\infty}}

    for all fL()f\in L_{\infty}(\mathbb{R}).

The standard condition for the BIBO stability of the continuous-time convolution operator Th:fhf{\mathrm{T}}_{h}:f\mapsto h\ast f that is found in engineering textbooks is hL1<\|h\|_{L_{1}}<\infty, where the L1L_{1}-norm is defined by

hL1=|h(t)|dt.\displaystyle\|h\|_{L_{1}}\stackrel{{\scriptstyle\vartriangle}}{{=}}\int_{\mathbb{R}}|h(t)|{\rm d}t. (2)

A slightly more precise statement is hL1()h\in L_{1}(\mathbb{R}), where

L1()={f:s.t.f is measurable and fL1<}L_{1}(\mathbb{R})=\{f:\mathbb{R}\to\mathbb{R}\ \ \mbox{s.t.}\ \ f\mbox{ is measurable and }\|f\|_{L_{1}}<\infty\}

is Lebesgue’s space of absolutely integrable functions.

The sufficiency of the condition hL1()h\in L_{1}(\mathbb{R}) is deduced from the standard estimate

|h(τ)f(tτ)dτ|\displaystyle\left|\int_{\mathbb{R}}h(\tau)f(t-\tau)\,{\rm d}\tau\right| |h(τ)||f(tτ)|dτ\displaystyle\leqslant\int_{\mathbb{R}}|h(\tau)|\cdot|f(t-\tau)|\,{\rm d}\tau
(|h(τ)|dτ)fL,\displaystyle\leqslant\left(\int_{\mathbb{R}}\left|h(\tau)\right|{\rm d}\tau\right)\,\|f\|_{L_{\infty}},

which is valid for any tt\in\mathbb{R}. The convolution integral (1) is therefore well defined if fL()f\in L_{\infty}(\mathbb{R}), which then also yields the classical bound on BIBO stability

hfLhL1fL<.\displaystyle{\|h\ast f\|_{L_{\infty}}}\leq\|h\|_{L_{1}}\,\|f\|_{L_{\infty}}<\infty. (3)

By adapting the argument that is used in the discrete-time formulation of BIBO stability, many authors (see Appendix B) claim that the condition hL1()h\in L_{1}(\mathbb{R}) is also necessary. To that end, they apply the convolution system to a “worst-case” signal

f0(t)=sign(h(t))\displaystyle f_{0}(t)={\rm sign}\big{(}h(-t)\big{)} (4)

in order to produce the strongest response at t=0t=0,

(hf0)(0)=+h(τ)sign(h(τ))dτ=+|h(τ)|dτ,\displaystyle(h\ast f_{0})(0)=\int_{-\infty}^{+\infty}h(\tau){\rm sign}\big{(}h(\tau)\big{)}\,{\rm d}\tau=\int_{-\infty}^{+\infty}\left|h(\tau)\right|{\rm d}\tau,

which is then claimed to saturate the stability bound (3) with hf0L=hL1f0L\|h\ast f_{0}\|_{L_{\infty}}=\|h\|_{L_{1}}\,\|f_{0}\|_{L_{\infty}}. Unfortunately, this simple reasoning has two shortcomings. First, unlike in the discrete setting, the characterization of what happens at t=0t=0, which is a set of measure zero, is not sufficient to deduce that hf0L(hf0)(0)\|h\ast f_{0}\|_{L_{\infty}}\geq(h\ast f_{0})(0), unless one invokes the continuity of t(hf0)(t)t\mapsto(h\ast f_{0})(t), which is not yet known at this stage (see Theorem 2). Second, one cannot ensure that the Lebesgue convolution integral (1) is well defined for f0L()f_{0}\in L_{\infty}(\mathbb{R}), unless hh is Lebesgue-integrable333Any measurable function h:h:\mathbb{R}\to\mathbb{R} admits a unique decomposition as h=h+hh=h^{+}-h^{-} with h+,h:0h^{+},h^{-}:\mathbb{R}\to\mathbb{R}_{\geq 0}. It is Lebesgue integrable if min(h+L1,hL1)<\min(\|h^{+}\|_{L_{1}},\|h^{-}\|_{L_{1}})<\infty [4]., which then considerably limits the scope of the claim about necessity.

Our first practical fix is an extension of the argumentation to the larger space L1,loc()L_{1,\rm loc}(\mathbb{R}) of measurable functions that are locally integrable, meaning that 𝕂|h(t)|dt<\int_{\mathbb{K}}|h(t)|{\rm d}t<\infty over any compact domain 𝕂\mathbb{K}\subset\mathbb{R}. The reassuring outcome, which conforms with the practice in the field, is that one can determine the stability of an LSI system by integrating the absolute value of its impulse response—even if hh is not globally Lebesgue integrable, as in the case of an increasing and possibly oscillating exponential.

Proposition 1.

If hL1()h\in L_{1}(\mathbb{R}), then the convolution operator fhff\mapsto h\ast f defined by (1) is BIBO-stable with hfLhL1fL\|h\ast f\|_{L_{\infty}}\leq\|h\|_{L_{1}}\|f\|_{L_{\infty}}. Conversely, if the impulse response hh is measurable and locally integrable with |h(t)|dt=\int_{\mathbb{R}}|h(t)|{\rm d}t={\infty}, then the system is not BIBO-stable, in which case it is said to be unstable.

Proof.

The first statement is a paraphrasing of (3). For the converse part, we assume that hL1,loc()h\in L_{1,\rm loc}(\mathbb{R}) with |h(t)|dt=\int_{\mathbb{R}}|h(t)|{\rm d}t={\infty}. Because of the local integrability of hh, one can then still rely on the definition of the convolution given by (1), but only if the input function ff is bounded and compactly supported. By considering the truncated versions f0,T=f0𝟙[T,T]f_{0,T}=f_{0}\cdot\mathbbm{1}_{[-T,T]} of the worst-case signal (4), we can therefore determine the maximal value of the output signal as

(hf0,T)(0)=T+Th(τ)sign(h(τ))dτ=T+T|h(τ)|dτ.(h\ast f_{0,T})(0)=\int_{-T}^{+T}h(\tau){\rm sign}\big{(}h(\tau)\big{)}{\rm d}\tau=\int_{-T}^{+T}|h(\tau)|{\rm d}\tau.

The additional ingredient is the continuity of t(hf0,T)(t)t\mapsto(h\ast f_{0,T})(t) in the neighborhood of t=0t=0 (see Proposition 4 in Appendix D), which allows us to conclude that (hf0,T)(0)supt|hf0,T(t)|=hf0,TL(h\ast f_{0,T})(0)\leq\sup_{t\in\mathbb{R}}|h\ast f_{0,T}(t)|=\|h\ast f_{0,T}\|_{L_{\infty}}. While the latter quantity is finite for any fixed value of TT, we have that limT(hf0,T)(0)=|h(t)|dt=\lim_{T\to\infty}(h\ast f_{0,T})(0)=\int_{\mathbb{R}}|h(t)|{\rm d}t={\infty}, which indicates that the output signal becomes unbounded in the limit. This shows that the underlying system is unstable.

Another way of obtaining Proposition 1 is as a corollary of Theorem 3 (the complete characterization of BIBO-stable systems) and Proposition 3 in Section IV. The important examples of unstable filters that fall within the scope of Proposition 1 are the systems ruled by differential equations with at least one pole in the right-half complex plane; for instance, h(t)=𝟙+(t)eαth(t)=\mathbbm{1}_{+}(t)\mathrm{e}^{\alpha t} with Re(α)0{\rm Re}(\alpha)\geq 0 [13]. The derivative operator with h=δh=\delta^{\prime} and the Hilbert transform with h(t)=1/(πt)h(t)=1/(\pi t) are unstable as well (as asserted by Theorem 3), but they fall outside the scope of Proposition 1: the first because δ\delta^{\prime} is not a function (but a distribution), and the second because the function 1/t1/t is not locally integrable—in fact, the impulse response of the Hilbert transform is the distribution “1/(πt)1/(\pi t)” that requires the use of a “principal value” for the proper definition of the convolution integral [18].

In the stable scenario, where hL1()h\in L_{1}(\mathbb{R}), we are able to characterize the underlying filter by its frequency response

h^(ω)={h}(ω)=h(t)ejωtdt,\displaystyle\widehat{h}(\omega)\stackrel{{\scriptstyle\vartriangle}}{{=}}\mathcal{F}\{h\}(\omega)=\int_{\mathbb{R}}h(t)\mathrm{e}^{-\mathrm{j}\omega t}{\rm d}t, (5)

which is the “classical” Fourier transform of hh. Moreover, the Riemann-Lebesgue lemma ensures that h^C0()\widehat{h}\in C_{0}(\mathbb{R}) with h^suphL1\|\widehat{h}\|_{\sup}\leq\|h\|_{L_{1}}. We recall that C0()C_{0}(\mathbb{R}) is the Banach space of continuous and bounded functions that decay at infinity, equipped with the sup\sup-norm.

III Banach Formulations of BIBO Stability

The classical textbook statements on continuous-time BIBO stability, including our reformulation in Proposition 1, have two limitations. First, they exclude the identity operator with h=δh=\delta, as explained in Appendix A. Second, they are often evasive concerning the hypotheses under which the condition hL1()h\in L_{1}(\mathbb{R}) is necessary (see Appendix B). In this section, we show how this can be corrected by considering appropriate Banach spaces.

III-A Extension of the Notion of Convolution

The scope of our mathematical statements relies on Schwartz’ famous kernel theorem [5, 16] which delineates the complete class of linear operators that continuously map 𝒟()𝒟(){\mathcal{D}}(\mathbb{R})\to{\mathcal{D}}^{\prime}(\mathbb{R}). We recall that 𝒟()=𝒞c(){\mathcal{D}}(\mathbb{R})={\mathcal{C}}_{\rm c}^{\infty}(\mathbb{R}) is the space of smooth and compactly supported test functions equipped with the usual Schwartz topology444A sequence of functions φk𝒟()\varphi_{k}\in{\mathcal{D}}(\mathbb{R}) is said to converge to φ\varphi in 𝒟(){\mathcal{D}}(\mathbb{R}) if (i) there exists a compact domain 𝔽\mathbb{F} that includes the support of φ\varphi and of all φk\varphi_{k}, and (ii) φkφn0\|\varphi_{k}-\varphi\|_{n}\to 0 for all nn\in\mathbb{N}, where φn=DnφL\|\varphi\|_{n}\stackrel{{\scriptstyle\vartriangle}}{{=}}\|{\rm D}^{n}\varphi\|_{L_{\infty}} with Dn:𝒟()𝒟(){\rm D}^{n}:{\mathcal{D}}(\mathbb{R})\to{\mathcal{D}}(\mathbb{R}) the nnth derivative operator.. Its topological dual 𝒟(){\mathcal{D}}^{\prime}(\mathbb{R}) is the space of generalized functions also known as distributions. In essence, a distribution f𝒟()f\in{\mathcal{D}}^{\prime}(\mathbb{R}) is a linear map—more precisely, a continuous linear functional—that assigns a real number to each test function φ𝒟()\varphi\in{\mathcal{D}}(\mathbb{R}); this is denoted by f:φf,φf:\varphi\mapsto\langle f,\varphi\rangle. For instance, the definition of Dirac’s impulse as a distribution is δ:φδ,φ=φ(0)\delta:\varphi\mapsto\langle\delta,\varphi\rangle\stackrel{{\scriptstyle\vartriangle}}{{=}}\varphi(0).

Beside linearity, the property that defines an LSI operator is TLSI{φ(t0)}(t)=TLSI{φ}(tt0){\mathrm{T}}_{\rm LSI}\{\varphi(\cdot-t_{0})\}(t)={\mathrm{T}}_{\rm LSI}\{\varphi\}(t-t_{0}) for any t0t_{0}\in\mathbb{R}. Schwartz’ theorem then tells us that there is a one-to-one correspondence between continuous LSI operators 𝒟()C(){\mathcal{D}}(\mathbb{R})\to C(\mathbb{R}) and distributions, with the defining distribution h𝒟()h\in{\mathcal{D}}^{\prime}(\mathbb{R}) being the impulse response of the operator. The relevant space of continuous functions here is C()C(\mathbb{R}) with the topology of uniform convergence over compact sets, which involves the system of seminorms fN=sup|t|N|f(t)|,N\|f\|_{N}=\sup_{|t|\leq N}|f(t)|,N\in\mathbb{N}. The latter is an extended functional setup that tolerates arbitrary growth at infinity.

Theorem 1 (Schwartz’ kernel theorem for LSI operators).

For any given h𝒟()h\in{\mathcal{D}}^{\prime}(\mathbb{R}), the operator Th:φhφ{\mathrm{T}}_{h}:\varphi\mapsto h\ast\varphi with

t(hφ)(t)=h,φ(t)\displaystyle t\mapsto(h\ast\varphi)(t)\stackrel{{\scriptstyle\vartriangle}}{{=}}\langle h,\varphi(t-\cdot)\rangle (6)

is LSI and continuously maps 𝒟() c. C(){\mathcal{D}}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C(\mathbb{R}). Conversely, for every LSI operator TLSI:𝒟() c. C(){\mathrm{T}}_{\rm LSI}:{\mathcal{D}}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C(\mathbb{R}), there is a unique h𝒟()h\in{\mathcal{D}}^{\prime}(\mathbb{R}) such that TLSI=Th:φhφ{\mathrm{T}}_{\rm LSI}={\mathrm{T}}_{h}:\varphi\mapsto h\ast\varphi where the convolution is specified by (6).

Then, depending on the decay properties of hh, it is generally possible to extend the domain of the convolution operator Th{\mathrm{T}}_{h} to some appropriate Banach space according to the procedure described in Section IV. For instance, if hL1()h\in L_{1}(\mathbb{R}), then Th{\mathrm{T}}_{h} has a continuous extension L()Cb()C()L_{\infty}(\mathbb{R})\to C_{\rm b}(\mathbb{R})\subset C(\mathbb{R}) that coincides with the classical definition given by (1).

Finally, we note that, for the cases where the Dirac impulse δ\delta is in the domain of the extended operator (for instance, when hC()h\in C(\mathbb{R})), the distributional definition of the convolution given by (6) yields Th{δ}=hδ=h{\mathrm{T}}_{h}\{\delta\}=h\ast\delta=h, which explains the term “impulse response.”

III-B Banach Spaces of Bounded Functions

In order to investigate the issue of BIBO stability, it is helpful to describe the boundedness and continuity properties of functions via their inclusion in appropriate Banach subspaces of 𝒟(){\mathcal{D}}^{\prime}(\mathbb{R}). The three relevant function spaces are

C0()Cb()L().C_{0}(\mathbb{R})\subset C_{\rm b}(\mathbb{R})\subset L_{\infty}(\mathbb{R}).

The central space consists of the subset of bounded functions that are continuous:

Cb()={f: s.t. f is continuous and fsup<}.\displaystyle C_{\rm b}(\mathbb{R})=\left\{f:\mathbb{R}\to\mathbb{R}\ \mbox{ s.t. }f\mbox{ is continuous and }\|f\|_{\sup}<\infty\right\}.

It is a classical example of Banach space—a complete normed vector space [12]. The smaller space C0()C_{0}(\mathbb{R}), which is also equipped with the sup{\sup}-norm, imposes the additional constraint that f(t)f(t) should vanish at t=±t=\pm\infty. It is best described as the completion of 𝒟(){\mathcal{D}}(\mathbb{R}) equipped with the sup\sup-norm, which will have its importance in the sequel. This property is indicated by C0()=(𝒟(),sup)¯C_{0}(\mathbb{R})=\overline{({\mathcal{D}}(\mathbb{R}),\|\cdot\|_{\sup})}. The concept is also valid for L1()L_{1}(\mathbb{R}), which can be described as L1()=(𝒟(),L1)¯L_{1}(\mathbb{R})=\overline{({\mathcal{D}}(\mathbb{R}),\|\cdot\|_{L_{1}})}, where the L1L_{1}-norm is defined by (2) with f𝒟()f\in{\mathcal{D}}(\mathbb{R}) and the integral being classical—in the sense of Riemann. This completion property applies to Lp()=(𝒟(),Lp)¯L_{p}(\mathbb{R})=\overline{({\mathcal{D}}(\mathbb{R}),\|\cdot\|_{L_{p}})} with p[1,)p\in[1,\infty) as well [4, Proposition 8.17, p. 254], but not for p=p=\infty, which explains the importance of the space C0()C_{0}(\mathbb{R}), which is distinct from L()L_{\infty}(\mathbb{R}).

In order to properly identify L()L_{\infty}(\mathbb{R}) as a subspace of 𝒟(){\mathcal{D}}^{\prime}(\mathbb{R}), we shall exploit the property that the LL_{\infty}-norm is the dual of the L1L_{1}-norm. We therefore choose to define the LL_{\infty}-norm as

fL=supφ𝒟():φL11f,φ=supφL1():φL11f,φ,\displaystyle\|f\|_{L_{\infty}}\stackrel{{\scriptstyle\vartriangle}}{{=}}\sup_{\varphi\in{\mathcal{D}}(\mathbb{R}):\,\|\varphi\|_{L_{1}}\leq 1}\langle f,\varphi\rangle=\sup_{\varphi\in L_{1}(\mathbb{R}):\,\|\varphi\|_{L_{1}}\leq 1}\langle f,\varphi\rangle, (7)

where the central part of (7) takes advantage of the denseness555This means that, for any fL1()f\in L_{1}(\mathbb{R}) and any ϵ>0\epsilon>0, there exists a function φϵ𝒟()\varphi_{\epsilon}\in{\mathcal{D}}(\mathbb{R}) such that fφϵL1<ϵ\|f-\varphi_{\epsilon}\|_{L_{1}}<\epsilon. It is a direct consequence of L1()=(𝒟(),L1)¯L_{1}(\mathbb{R})=\overline{({\mathcal{D}}(\mathbb{R}),\|\cdot\|_{L_{1}})}. of 𝒟(){\mathcal{D}}(\mathbb{R}) in L1()L_{1}(\mathbb{R}). This yields a definition that is valid not only for (measurable) functions, but also for all f𝒟()f\in{\mathcal{D}}^{\prime}(\mathbb{R}). Consequently, we can redefine our target space as

L()={f𝒟():fL<},\displaystyle L_{\infty}(\mathbb{R})=\{f\in{\mathcal{D}}^{\prime}(\mathbb{R}):\|f\|_{L_{\infty}}<\infty\}, (8)

which is readily identified as the topological dual of L1()L_{1}(\mathbb{R}); that is, L()=(L1())L_{\infty}(\mathbb{R})=\big{(}L_{1}(\mathbb{R})\big{)}^{\prime} due to the dual specification of the LL_{\infty}-norm given by the right-hand side of (7).

While (8) defines L()L_{\infty}(\mathbb{R}) as a subspace of 𝒟(){\mathcal{D}}^{\prime}(\mathbb{R}), we can also identify its elements as (bounded) measurable functions f:f:\mathbb{R}\to\mathbb{R} via the classical association

φf,φ=φ(t)f(t)dt,\displaystyle\varphi\mapsto\langle f,\varphi\rangle\stackrel{{\scriptstyle\vartriangle}}{{=}}\int_{\mathbb{R}}\varphi(t)f(t){\rm d}t, (9)

where the right-hand side of (9) is a standard Lebesgue integral. Now, the main difference between the sup\sup-norm and the LL_{\infty}-norm is that, for fL()f\in L_{\infty}(\mathbb{R}) (identified as a function), the inequality |f(t)|fL|f(t)|\leq\|f\|_{L_{\infty}} holds for almost every tt\in\mathbb{R}. This means that it holds over the whole real line except, possibly, on a set of measure zero. This is often indicated as f=esssupt|f(t)|\|f\|_{\infty}=\operatorname*{ess\,sup}_{t\in\mathbb{R}}|f(t)|, using the notion of essential supremum. In other words, the LL_{\infty}-norm is more permissive than the sup\sup-norm with ffsup\|f\|_{\infty}\leq\|f\|_{\sup}. However, the two norms are equal whenever the function ff is continuous, which translates into the isometric inclusion C0()iso.Cb()iso.L()C_{\rm 0}(\mathbb{R})\xhookrightarrow{\mbox{\tiny\rm iso.}}C_{\rm b}(\mathbb{R})\xhookrightarrow{\mbox{\tiny\rm iso.}}L_{\infty}(\mathbb{R}).

III-C Extended Results on BIBO Stability

Remarkably, the combination of the two latter function spaces enables us to formulate a first Banach extension of the classical statement on BIBO stability. To that end, we shall restrict the distributional framework covered by Theorem 1 to the case where the impulse response hh is identifiable as a measurable function (i.e., hL1,loc()𝒟()h\in L_{1,{\rm loc}}(\mathbb{R})\subset{\mathcal{D}}^{\prime}(\mathbb{R})). The linear functional on the right-hand side of (6) then has an explicit integral description given by (1) with f=φ𝒟()f=\varphi\in{\mathcal{D}}(\mathbb{R}). Within this class of convolution operators, we now identify the ones whose domain can be extended to L()L_{\infty}(\mathbb{R}).

Theorem 2.

The convolution operator Th:fhf{\mathrm{T}}_{h}:f\mapsto h\ast f with hL1,loc()h\in L_{1,\rm loc}(\mathbb{R}) has a continuous extension L() c. Cb()L_{\infty}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{\rm b}(\mathbb{R}) if and only if hL1()h\in L_{1}(\mathbb{R}). Moreover,

hfsuphL1fL\|h\ast f\|_{\sup}\leq\|h\|_{L_{1}}\|f\|_{L_{\infty}}

with the bound being sharp in the sense that it also yields the norm of the underlying operator: ThLCb=hL1\|{\mathrm{T}}_{h}\|_{L_{\infty}\to C_{\rm b}}=\|h\|_{L_{1}} (see Definition 2).

The proof of this result is deferred to Section IV (see Item 2) and the final statement of Theorem 4.

It is of interest to compare Proposition 1 and Theorem 2 because they address the problem of stability from different but complementary perspectives. Proposition 1 is focused primarily on the well-posedness of the convolution integral (1) for fL()f\in L_{\infty}(\mathbb{R}). It can be paraphrased as: “Let hh be a measurable (and locally integrable) function. Then, the Lebesgue integral (1) defines a convolution operator that is BIBO-stable if and only if hL1()h\in L_{1}(\mathbb{R}).” By contrast, Theorem 2 considers the complete family of “classical” convolution operators Th:𝒟()C(){\mathrm{T}}_{h}:{\mathcal{D}}(\mathbb{R})\to C(\mathbb{R}) with hL1,loc()h\in L_{1,\rm loc}(\mathbb{R}) and precisely identifies the subset of operators that have a continuous extension from L()Cb()L_{\infty}(\mathbb{R})\to C_{\rm b}(\mathbb{R}). Since Cb()C_{\rm b}(\mathbb{R}) is isometrically embedded in L()L_{\infty}(\mathbb{R}), this is more informative than Proposition 1 because it also tells us that (hf)(t)(h\ast f)(t) is a continuous function of tt\in\mathbb{R}. In that respect, we note that the requirement that the convolution of any bounded function ff be continuous excludes the use of the identity operator with h=δh=\delta at the onset, even if we extend the framework to h𝒟()h\in{\mathcal{D}}^{\prime}(\mathbb{R}).

To obtain a more permissive characterization of BIBO stability, we need to extend the range of the operator from Cb()C_{\rm b}(\mathbb{R}) to L()L_{\infty}(\mathbb{R}), which should then also translate into a corresponding enlargement of the class of admissible impulse responses. We shall delineate the latter in a way that parallels our definition of L()L_{\infty}(\mathbb{R}), with the roles of the L1L_{1}- and sup\sup- (or LL_{\infty}-) norms being interchanged. To that end, we first define the {\mathcal{M}}-norm as

f=supφ𝒟():φsup1f,φ.\displaystyle\|f\|_{{\mathcal{M}}}\stackrel{{\scriptstyle\vartriangle}}{{=}}\sup_{\varphi\in{\mathcal{D}}(\mathbb{R}):\,\|\varphi\|_{\sup}\leq 1}\langle f,\varphi\rangle. (10)

This then yields the Banach space

()={f𝒟():f<},\displaystyle{\mathcal{M}}(\mathbb{R})=\{f\in{{\mathcal{D}}^{\prime}(\mathbb{R})}:\|f\|_{{\mathcal{M}}}<\infty\}, (11)

which also happens to be the space of bounded Radon measures666We adhere with Bourbaki’s nomenclature to distinguish the two complementary interpretations of a measure: either as a continuous linear functional on 𝒟(){\mathcal{D}}(\mathbb{R}) (Radon measure), or as a set-theoretic additive rule that associates a real number to any Borel set of \mathbb{R} (signed Borel measure) [2, 1]. on C0()C_{0}(\mathbb{R}). In other words, (){\mathcal{M}}(\mathbb{R}) is the topological dual of C0()C_{0}(\mathbb{R}). Moreover, we can invoke the Riesz-Markov theorem to identify ()=(C0()){\mathcal{M}}(\mathbb{R})=\big{(}C_{0}(\mathbb{R})\big{)}^{\prime} with the space of bounded signed Borel measures on \mathbb{R} [15]. Concretely, this means that any h()h\in{\mathcal{M}}(\mathbb{R}) is associated with a unique Borel measure μh\mu_{h}, which then gives a concrete definition of the linear functional

fh,f=f(τ)dμh(τ)\displaystyle f\mapsto\langle h,f\rangle\stackrel{{\scriptstyle\vartriangle}}{{=}}\int_{\mathbb{R}}f(\tau){\rm d}\mu_{h}(\tau) (12)

for any measurable function ff, while the total-variation norm of the measure μh\mu_{h} is given by μhTV=d|μh|=h\|\mu_{h}\|_{\rm TV}\stackrel{{\scriptstyle\vartriangle}}{{=}}\int_{\mathbb{R}}{\rm d}|\mu_{h}|=\|h\|_{{\mathcal{M}}} (see Section IV-C). The main point for us is that (){\mathcal{M}}(\mathbb{R}) is a superset of L1()L_{1}(\mathbb{R}), with f=fL1\|f\|_{{\mathcal{M}}}=\|f\|_{L_{1}} for all fL1()f\in L_{1}(\mathbb{R}). Moreover, we have that δ(t0)()\delta(\cdot-t_{0})\in{\mathcal{M}}(\mathbb{R}) for any t0t_{0}\in\mathbb{R} with δ(t0)=1\|\delta(\cdot-t_{0})\|_{{\mathcal{M}}}=1, as can be readily inferred from (10) by considering a non-negative test function that achieves its maximum φ(t0)=1\varphi(t_{0})=1 at t=t0t=t_{0}.

For the cases where the impulse response h()h\in{\mathcal{M}}(\mathbb{R}) is not an L1L_{1} function, we extend our definition of the original (Lebesgue) convolution integral as

t(hf)(t)=h,f(t)=f(tτ)dμh(τ),\displaystyle t\mapsto(h\ast f)(t)=\langle h,f(t-\cdot)\rangle\stackrel{{\scriptstyle\vartriangle}}{{=}}\int_{\mathbb{R}}f(t-\tau){\rm d}\mu_{h}(\tau), (13)

which is the same as (1) when we can write dμh(τ)=h(τ)dτ{\rm d}\mu_{h}(\tau)=h(\tau){\rm d}\tau, which happens when the corresponding measure μh\mu_{h} is absolutely continuous777Another way to put it is that hh is the Radon-Nikodym derivative of μh\mu_{h}. with respect to the Lebesgue measure. A standard manipulation then yields that

|(hf)(t)|\displaystyle\left|(h\ast f)(t)\right| |f(tτ)|d|μh|(τ)\displaystyle\leq\int_{\mathbb{R}}|f(t-\tau)|\,{\rm d}|\mu_{h}|(\tau)
fLd|μh|=fLh,\displaystyle\leq\|f\|_{L_{\infty}}\int_{\mathbb{R}}{\rm d}|\mu_{h}|=\|f\|_{L_{\infty}}\|h\|_{{\mathcal{M}}}, (14)

which is the basis for the direct (easy) part of Theorem 3, where the complete class of BIBO-stable systems is identified, including the identity operator.

Theorem 3.

The convolution operator Th:fhf{\mathrm{T}}_{h}:f\mapsto h\ast f with h𝒟()h\in{\mathcal{D}}^{\prime}(\mathbb{R}) has a continuous extension L() c. L()L_{\infty}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{\infty}(\mathbb{R}) if and only if h()h\in{\mathcal{M}}(\mathbb{R}). Moreover,

hfhfL\|h\ast f\|_{\infty}\leq\|h\|_{{\mathcal{M}}}\;\|f\|_{L_{\infty}}

with the bound being sharp in the sense that ThLL=h\|{\mathrm{T}}_{h}\|_{L_{\infty}\to L_{\infty}}=\|h\|_{{\mathcal{M}}}.

This result, which is also valid in dimensions higher than 11, is known in harmonic analysis [6, p. 140 Corollary 2.5.9], [18] but much less so in engineering circles. It can be traced back to an early paper by Hörmander that provides a comprehensive treatment of convolution operators on LpL_{p} spaces [7]. The reminder of the paper is devoted to the proof of the two theorems on BIBO stability and of some interesting variants (see Theorem 4). To that end, we shall rely on Schwartz’ powerful distributional formalism which, as we shall see, allows for a rather soft derivation, once the prerequisites have been laid out.

IV Mathematical Derivations

IV-A Extension of Convolution Operators

The most general form of a convolution operator backed by Schwartz’ kernel theorem (see Theorem 1) is Th:𝒟()C()𝒟(){\mathrm{T}}_{h}:{\mathcal{D}}(\mathbb{R})\to C(\mathbb{R})\xhookrightarrow{}{\mathcal{D}}^{\prime}(\mathbb{R}) with h𝒟()h\in{\mathcal{D}}^{\prime}(\mathbb{R}), where Th{φ}{\mathrm{T}}_{h}\{\varphi\} is defined by (6) for any φ𝒟()\varphi\in{\mathcal{D}}(\mathbb{R}). The two complementary ingredients at play there are: (i) the restriction of the domain to 𝒟(){\mathcal{D}}(\mathbb{R})—the “nicest” class of functions in terms of smoothness and decay—and (ii) the extension of the range to 𝒟(){\mathcal{D}}^{\prime}(\mathbb{R}), which can accommodate an arbitrary degree of growth (polynomial, or even exponential) at infinity. In other words, the theoretical framework is such that it can deal with the very worst scenarios, including unstable differential systems whose impulse response is exponentially increasing.

Then, depending on the smoothness and decay properties of hh, it is usually possible to extend the domain of Th{\mathrm{T}}_{h} to some Banach space 𝒳𝒟(){\mathcal{X}}\supseteq{\mathcal{D}}(\mathbb{R}) that is continuously embedded in 𝒟(){\mathcal{D}}^{\prime}(\mathbb{R}), which is denoted by 𝒳𝒟(){\mathcal{X}}\xhookrightarrow{}{\mathcal{D}}^{\prime}(\mathbb{R}). For this to be feasible, we require that 𝒳\|\cdot\|_{\mathcal{X}} be a valid norm on 𝒟(){\mathcal{D}}(\mathbb{R}) and that 𝒟(){\mathcal{D}}(\mathbb{R}) be dense in 𝒳{\mathcal{X}}, which is equivalent to 𝒳=(𝒟(),𝒳)¯{\mathcal{X}}=\overline{({\mathcal{D}}(\mathbb{R}),\|\cdot\|_{{\mathcal{X}}})}. In other words, 𝒳{\mathcal{X}} is the completion of 𝒟(){\mathcal{D}}(\mathbb{R}) equipped with the 𝒳\|\cdot\|_{{\mathcal{X}}}-norm.

We start by recalling the definition of the norm of a bounded operator.

Definition 2.

Let (𝒳,𝒳)({\mathcal{X}},\|\cdot\|_{{\mathcal{X}}}) and (𝒴,𝒴)({\mathcal{Y}},\|\cdot\|_{{\mathcal{Y}}}) be two Banach spaces and T{\mathrm{T}} a linear operator 𝒳𝒴{\mathcal{X}}\to{\mathcal{Y}}. Then, the operator is said to be bounded if

T𝒳𝒴=supf𝒳\{0}T{f}𝒴f𝒳<.\|{\mathrm{T}}\|_{{\mathcal{X}}\to{\mathcal{Y}}}\stackrel{{\scriptstyle\vartriangle}}{{=}}\sup_{f\in{\mathcal{X}}\backslash\{0\}}\frac{\|{\mathrm{T}}\{f\}\|_{{\mathcal{Y}}}}{\|f\|_{{\mathcal{X}}}}<\infty.

A direct consequence of Definition 2 is that a bounded operator T:𝒳𝒴{\mathrm{T}}:{\mathcal{X}}\to{\mathcal{Y}} continuously maps 𝒳{\mathcal{X}} into 𝒴{\mathcal{Y}}, as indicated by T:𝒳 c. 𝒴{\mathrm{T}}:{\mathcal{X}}\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{Y}}.

Theorem 2 then describes a functional mechanism that allows us to extend an operator initially defined on 𝒟(){\mathcal{D}}(\mathbb{R}). It is a particularization of a fundamental extension theorem in the theory of Banach spaces [14, Theorem I.7, p. 9].

Proposition 2 (Extension of a linear operator).

Let 𝒳{\mathcal{X}} and 𝒴{\mathcal{Y}} be two Banach subspaces of 𝒟(){\mathcal{D}}^{\prime}(\mathbb{R}) with the additional property that 𝒟(){\mathcal{D}}(\mathbb{R}) is dense in 𝒳{\mathcal{X}}. Then, the linear operator T:𝒟() c. 𝒟(){\mathrm{T}}:{\mathcal{D}}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{D}}^{\prime}(\mathbb{R}) has a unique continuous extension 𝒳=(𝒟(),𝒳)¯ c. 𝒴{\mathcal{X}}=\overline{({\mathcal{D}}(\mathbb{R}),\|\cdot\|_{\mathcal{X}})}\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{Y}} with T𝒳𝒴C\|{\mathrm{T}}\|_{{\mathcal{X}}\to{\mathcal{Y}}}\leq C if and only if

(i)\displaystyle(i)\ \ T{φ}𝒴,and\displaystyle{\mathrm{T}}\{\varphi\}\in{\mathcal{Y}},\quad\mbox{and} (15)
(ii)\displaystyle(ii)\ \ T{φ}𝒴Cφ𝒳\displaystyle\|{\mathrm{T}}\{\varphi\}\|_{{\mathcal{Y}}}\leq C\|\varphi\|_{{\mathcal{X}}} (16)

for all φ𝒟()\varphi\in{\mathcal{D}}(\mathbb{R}) and some constant C>0C>0.

Since a convolution operator Th:𝒟() c. 𝒟(){\mathrm{T}}_{h}:{\mathcal{D}}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{D}}^{\prime}(\mathbb{R}) is uniquely characterized by its impulse response h𝒟()h\in{\mathcal{D}}^{\prime}(\mathbb{R}), the same holds true for its extension Th:𝒳 c. 𝒴{\mathrm{T}}_{h}:{\mathcal{X}}\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{Y}}, which justifies the use of the same symbol. Rather than defining Th{f}=hf{\mathrm{T}}_{h}\{f\}=h\ast f through a Lebesgue integral as in (1) or (13), we can therefore rely on (6) and define our extended convolution operator Th:𝒳𝒴{\mathrm{T}}_{h}:{\mathcal{X}}\to{\mathcal{Y}} through a limit process. Specifically, we pick a Cauchy sequence (φn)(\varphi_{n}) in (𝒟(),𝒳)({\mathcal{D}}(\mathbb{R}),\|\cdot\|_{{\mathcal{X}}}) such that limnφn=f𝒳\lim_{n\to\infty}\varphi_{n}=f\in{\mathcal{X}}. Then, the sequence of functions (gn=hφn)(g_{n}=h\ast\varphi_{n}) with

t(hφn)(t)=h,φn(t)\displaystyle t\mapsto(h\ast\varphi_{n})(t)=\langle h,\varphi_{n}(t-\cdot)\rangle (17)

is Cauchy in 𝒴{\mathcal{Y}} and converges to a limit g=limn(hφn)𝒴g=\lim_{n\to\infty}(h\ast\varphi_{n})\in{\mathcal{Y}}, independently of the choice of the φn\varphi_{n} since the space 𝒴{\mathcal{Y}} is complete. We now recapitulate this process in the form of a definition.

Definition 3 (Banach extension of a distributional convolution operator).

Let 𝒳{\mathcal{X}} and 𝒴{\mathcal{Y}} be two Banach subspaces of 𝒟(){\mathcal{D}}^{\prime}(\mathbb{R}) with the additional property that 𝒟(){\mathcal{D}}(\mathbb{R}) is dense in 𝒳{\mathcal{X}}. When the two conditions in Theorem 2 hold, the unique continuous extension Th:𝒳 c. 𝒴{\mathrm{T}}_{h}:{\mathcal{X}}\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{Y}} of the convolution operator specified by (17) with h𝒟()h\in{\mathcal{D}}^{\prime}(\mathbb{R}) is defined by

Th:fhf=limn(hφn)𝒴,\displaystyle{\mathrm{T}}_{h}:f\mapsto h\ast f\stackrel{{\scriptstyle\vartriangle}}{{=}}\lim_{n\to\infty}(h\ast\varphi_{n})\in{\mathcal{Y}}, (18)

where (φn)(\varphi_{n}) is any sequence in 𝒟(){\mathcal{D}}(\mathbb{R}) such that limnfφn𝒳=0\lim_{n\to\infty}\|f-\varphi_{n}\|_{\mathcal{X}}=0.

Also important for our purpose is the adjoint operator T:𝒴𝒳{\mathrm{T}}^{\ast}:{\mathcal{Y}}^{\prime}\to{\mathcal{X}}^{\prime}, which is the unique linear operator such that

g,T{f}𝒴×𝒴=T{g},f𝒳×𝒳\langle g,{\mathrm{T}}\{f\}\rangle_{{\mathcal{Y}}^{\prime}\times{\mathcal{Y}}}=\langle{\mathrm{T}}^{\ast}\{g\},f\rangle_{{\mathcal{X}}^{\prime}\times{\mathcal{X}}}

for any g𝒴g\in{\mathcal{Y}}^{\prime} and f𝒳f\in{\mathcal{X}}^{\prime}, where the spaces 𝒳{\mathcal{X}}^{\prime} and 𝒴{\mathcal{Y}}^{\prime} are the duals of the topological vector spaces 𝒳{\mathcal{X}} and 𝒴{\mathcal{Y}}, respectively. If T:𝒳 c. 𝒴{\mathrm{T}}:{\mathcal{X}}\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{Y}} is bounded with operator norm T\|{\mathrm{T}}\|, then the adjoint T:𝒴 c. 𝒳{\mathrm{T}}^{\ast}:{\mathcal{Y}}^{\prime}\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{X}}^{\prime} is bounded with T=T\|{\mathrm{T}}^{\ast}\|=\|{\mathrm{T}}\|. In particular, the adjoint of the convolution operator Th:𝒟() c. 𝒟(){\mathrm{T}}_{h}:{\mathcal{D}}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{D}}^{\prime}(\mathbb{R}) is Th:𝒟() c. 𝒟(){\mathrm{T}}_{h^{\vee}}:{\mathcal{D}}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{D}}^{\prime}(\mathbb{R}), where hh^{\vee} is the time-reversed impulse response such that h,φ=h,φ\langle h^{\vee},\varphi\rangle=\langle h,\varphi^{\vee}\rangle, where φ(t)=φ(t)\varphi^{\vee}(t)\stackrel{{\scriptstyle\vartriangle}}{{=}}\varphi(-t).

We now briefly show how we make use of these two mechanisms to specify the continuous extension Th:L()L(){\mathrm{T}}_{h}:L_{\infty}(\mathbb{R})\to L_{\infty}(\mathbb{R}) with h()h\in{\mathcal{M}}(\mathbb{R}) (or, hL1()h\in L_{1}(\mathbb{R})) that is implicitly referred to in Theorems 2 and 3. The enabling ingredient there is the continuity bound hφL1hφL1\|h^{\vee}\ast\varphi\|_{L_{1}}\leq\|h^{\vee}\|_{{\mathcal{M}}}\|\varphi\|_{L_{1}} (see proof of Theorem 4, Item 4), which also yields hφL1()h^{\vee}\ast\varphi\in L_{1}(\mathbb{R}) for all φ𝒟()\varphi\in{\mathcal{D}}(\mathbb{R}). We then apply Definition 3 to specify the unique extension Th:L1() c. L1(){\mathrm{T}}_{h^{\vee}}:L_{1}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{1}(\mathbb{R}). An important point for our argumentation is that this (pre-adjoint) convolution operator also has a concrete implementation as

t(hφ)(t)=h,φ(t)=φ(τ+t)dμh(τ),\displaystyle t\mapsto(h^{\vee}\ast\varphi)(t)=\langle h^{\vee},\varphi(t-\cdot)\rangle=\int_{\mathbb{R}}\varphi(\tau+t){\rm d}\mu_{h}(\tau), (19)

which is supported by the same continuity bound with φ\varphi now ranging over L1()L_{1}(\mathbb{R}) instead of the smaller space 𝒟(){\mathcal{D}}(\mathbb{R}). The existence and uniqueness of Th:L1() c. L1(){\mathrm{T}}_{h^{\vee}}:L_{1}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{1}(\mathbb{R}) then guarantees the existence and unicity of the adjoint Th:L() c. L(){\mathrm{T}}^{\ast}_{h^{\vee}}:L_{\infty}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{\infty}(\mathbb{R}). To show that Th=Th{\mathrm{T}}^{\ast}_{h^{\vee}}={\mathrm{T}}_{h}, we use the explicit representation of Th{\mathrm{T}}_{h^{\vee}} given by (19) with h()h^{\vee}\in{\mathcal{M}}(\mathbb{R}) and invoke Fubini’s theorem to justify the interchange of integrals in

Th{f},g\displaystyle\langle{\mathrm{T}}_{h^{\vee}}\{f\},g\rangle =(f(τ+t)dμh(τ))g(t)dt\displaystyle=\int_{\mathbb{R}}\left(\int_{\mathbb{R}}f(\tau+t){\rm d}\mu_{h}(\tau)\right)g(t){\rm d}t
=f(x)g(xτ)dμh(τ)dx\displaystyle=\int_{\mathbb{R}}\int_{\mathbb{R}}f(x)g(x-\tau){\rm d}\mu_{h}(\tau){\rm d}x
=f(x)(g(xτ)dμh(τ))dx\displaystyle=\int_{\mathbb{R}}f(x)\left(\int_{\mathbb{R}}g(x-\tau){\rm d}\mu_{h}(\tau)\right){\rm d}x
=f,Th{g}\displaystyle=\langle f,{\mathrm{T}}_{h}\{g\}\rangle

for any fL1()f\in L_{1}(\mathbb{R}) and gL()g\in L_{\infty}(\mathbb{R}). This proves that the original convolution operator defined by (13) coincides with the adjoint of Th:L1() c. L1(){\mathrm{T}}_{h^{\vee}}:L_{1}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{1}(\mathbb{R}), which is also consistent with the property h=(h)h=(h^{\vee})^{\vee}. Since 𝒟()L(){\mathcal{D}}(\mathbb{R})\subset L_{\infty}(\mathbb{R}), we can therefore uniquely identify Th:L() c. L(){\mathrm{T}}_{h}:L_{\infty}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{\infty}(\mathbb{R}) as the extension of Th:𝒟()𝒟(){\mathrm{T}}_{h}:{\mathcal{D}}(\mathbb{R})\to{\mathcal{D}}^{\prime}(\mathbb{R}) that preserves the adjoint relation Th=Th{\mathrm{T}}^{\ast}_{h^{\vee}}={\mathrm{T}}_{h}.

IV-B Proof of Banach Variants of BIBO Stability

The Banach spaces of interest for us are 𝒳=C0(),Lp(){\mathcal{X}}=C_{0}(\mathbb{R}),L_{p}(\mathbb{R}) and 𝒴=C0(),Cb(),Lp(){\mathcal{Y}}=C_{0}(\mathbb{R}),C_{\rm b}(\mathbb{R}),L_{p}(\mathbb{R}) with p1p\geq 1.

Theorem 4.

Depending on the functional properties of its impulse response h𝒟()h\in{\mathcal{D}}^{\prime}(\mathbb{R}), the convolution operator Th:𝒟() c. 𝒟(){\mathrm{T}}_{h}:{\mathcal{D}}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}{\mathcal{D}}^{\prime}(\mathbb{R}) defined by (6) admits the following (unique) continuous extensions888See Definition 3 and accompanying explanations. The bottom line is that the definition of these operators is compatible with the convolution integral (1) or (13) depending on whether hh is a function or a Radon measure.

  1. 1.

    Let p,q(1,)p,q\in(1,\infty) be conjugate exponents with 1p+1q=1\frac{1}{p}+\frac{1}{q}=1. Then, hLq()Th:Lp() c. C0()h\in L_{q}(\mathbb{R})\ \Rightarrow\ {\mathrm{T}}_{h}:L_{p}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{0}(\mathbb{R}) with ThLpC0hLq\|{\mathrm{T}}_{h}\|_{L_{p}\to C_{0}}\leq\|h\|_{L_{q}}.

  2. 2.

    hL1()Th:L() c. Cb()h\in L_{1}(\mathbb{R})\quad\Rightarrow\quad{\mathrm{T}}_{h}:L_{\infty}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{\rm b}(\mathbb{R}) with ThLCb=hL1\|{\mathrm{T}}_{h}\|_{L_{\infty}\to C_{\rm b}}=\|h\|_{L_{1}}.

  3. 3.

    h()Th:C0() c. Cb()h\in{\mathcal{M}}(\mathbb{R})\quad\Leftrightarrow\quad{\mathrm{T}}_{h}:C_{0}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{\rm b}(\mathbb{R}).

  4. 4.

    h()Th:L1() c. L1()h\in{\mathcal{M}}(\mathbb{R})\quad\Leftrightarrow\quad{\mathrm{T}}_{h}:L_{1}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{1}(\mathbb{R}).

  5. 5.

    h()Th:L() c. L()h\in{\mathcal{M}}(\mathbb{R})\quad\Leftrightarrow\quad{\mathrm{T}}_{h}:L_{\infty}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{\infty}(\mathbb{R}).

Moreover, the operator norms for Items 3-5, characterized by an equivalence relation, are ThC0Cb=ThL1L1=ThLL=h\|{\mathrm{T}}_{h}\|_{C_{0}\to C_{\rm b}}=\|{\mathrm{T}}_{h}\|_{L_{1}\to L_{1}}=\|{\mathrm{T}}_{h}\|_{L_{\infty}\to L_{\infty}}=\|h\|_{{\mathcal{M}}}. Finally, under the hypothesis of local integrability hL1,loc()h\in L_{1,\rm loc}(\mathbb{R}), the continuity of Th:L() c. Cb(){\mathrm{T}}_{h}:L_{\infty}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{\rm b}(\mathbb{R}) implies that hL1()h\in L_{1}(\mathbb{R}), which is the converse part of Item 2.

Proof:

Item 1. Under the assumption that hLq()h\in L_{q}(\mathbb{R}) with q1q\geq 1, we invoke Hölder’s inequality

|(hφ)(t)||h(τ)||φ(tτ)|dτhLqφ(t)Lp\displaystyle\left|(h\ast\varphi)(t)\right|\leq\int_{\mathbb{R}}|h(\tau)|\cdot|\varphi(t-\tau)|{\rm d}\tau\leq\|h\|_{L_{q}}\|\varphi(\cdot-t)\|_{L_{p}}

for any φ𝒟()\varphi\in{\mathcal{D}}(\mathbb{R}), which yields the required upper bound hφLhLqφLp{\|h\ast\varphi\|_{L_{\infty}}}\leq\|h\|_{L_{q}}\|\varphi\|_{L_{p}}. Likewise, by linearity, we get that

|(hφ)(t)(hφ)(tΔt)|\displaystyle\left|(h\ast\varphi)(t)-(h\ast\varphi)(t-\Delta t)\right| =|h(φ(t)φ(tΔt))|\displaystyle=\left|h\ast\big{(}\varphi(t)-\varphi(t-\Delta t)\big{)}\right|
hLqφφ(Δt)Lp.\displaystyle\leq\|h\|_{L_{q}}\cdot\|\varphi-\varphi(\cdot-\Delta t)\|_{L_{p}}.

Due to the constraining topology of 𝒟(d){\mathcal{D}}(\mathbb{R}^{d}), limΔt0φφ(Δt)Lp=0\lim_{\Delta t\to 0}\|\varphi-\varphi(\cdot-\Delta t)\|_{L_{p}}=0 for any p1p\geq 1, which proves the continuity of the function t(hφ)(t)t\mapsto{(h\ast\varphi)(t)}. This leads to the intermediate outcome hφCb()h\ast\varphi\in C_{\rm b}(\mathbb{R}) for all φ𝒟()\varphi\in{\mathcal{D}}(\mathbb{R}).

If we now replace hh by ϕ𝒟()\phi\in{\mathcal{D}}(\mathbb{R}), we readily deduce that Tϕ{φ}=ϕφ{\mathrm{T}}_{\phi}\{\varphi\}=\phi\ast\varphi is compactly supported; hence, Tϕ{φ}C0(){\mathrm{T}}_{\phi}\{\varphi\}\in C_{0}(\mathbb{R}) for all φ𝒟()\varphi\in{\mathcal{D}}(\mathbb{R}) with ϕφLϕLqφLp{\|\phi\ast\varphi\|_{L_{\infty}}}\leq\|\phi\|_{L_{q}}\|\varphi\|_{L_{p}}. We then invoke Theorem 2 with 𝒳=(𝒟(),p)¯{\mathcal{X}}=\overline{({\mathcal{D}}(\mathbb{R}),\|\cdot\|_{p})} to deduce that Tϕ:Lp() c. C0(){\mathrm{T}}_{\phi}:L_{p}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{0}(\mathbb{R}) for p[1,)p\in[1,\infty) and Tϕ:C0() c. C0(){\mathrm{T}}_{\phi}:C_{0}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{0}(\mathbb{R}) for any ϕ𝒟()\phi\in{\mathcal{D}}(\mathbb{R}). Since the convolution is commutative, this implies that ϕh=hϕC0()\phi\ast h=h\ast\phi\in C_{0}(\mathbb{R}) for any hLq()h\in L_{q}(\mathbb{R}) with q(1,)q\in(1,\infty) (resp., hC0()h\in C_{0}(\mathbb{R})) and ϕ𝒟()Lp()\phi\in{\mathcal{D}}(\mathbb{R})\subset L_{p}(\mathbb{R}) which, by completion with respect to the Lp\|\cdot\|_{L_{p}} norm with p(1,)p\in(1,\infty), gives Th:Lp() c. C0(){\mathrm{T}}_{h}:L_{p}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{0}(\mathbb{R}) with ThLpC0hLq\|{\mathrm{T}}_{h}\|_{L_{p}\to C_{0}}\leq\|h\|_{L_{q}} (resp., Th:C0() c. C0(){\mathrm{T}}_{h}:C_{0}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{0}(\mathbb{R}) with ThC0C0=hL1\|{\mathrm{T}}_{h}\|_{C_{0}\to C_{0}}=\|h\|_{L_{1}}).

Item 3. Since Cb()iso.L()C_{\rm b}(\mathbb{R})\xhookrightarrow{\mbox{\tiny\rm iso.}}L_{\infty}(\mathbb{R}), the relevant duality bound there is (III-C), which yields hφLhφL\|h\ast\varphi\|_{L_{\infty}}\leq\|h\|_{{\mathcal{M}}}\;\|\varphi\|_{L_{\infty}}. This allows us to use the same argument as in Item 1 to show that Th{φ}Cb(){\mathrm{T}}_{h}\{\varphi\}\in C_{\rm b}(\mathbb{R}) for all φ𝒟()\varphi\in{\mathcal{D}}(\mathbb{R}). Since C0()=(𝒟(d),L)¯C_{0}(\mathbb{R})=\overline{({\mathcal{D}}(\mathbb{R}^{d}),\|\cdot\|_{L_{\infty}})}, we then apply the proven completion technique to specify the unique operator Th:C0()Cb(){\mathrm{T}}_{h}:C_{0}(\mathbb{R})\to C_{\rm b}(\mathbb{R}) with ThC0Cbh\|{\mathrm{T}}_{h}\|_{C_{0}\to C_{\rm b}}\leq\|h\|_{{\mathcal{M}}}. Conversely, let Th:C0() c. Cb(){\mathrm{T}}_{h}:C_{0}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{\rm b}(\mathbb{R}) with operator norm ThC0Cb<\|T_{h}\|_{C_{0}\to C_{\rm b}}<{\infty}. Then, for any φC0()\varphi\in C_{0}(\mathbb{R}),

(hφ)(0)=h,φThC0CbφL(h\ast\varphi)(0)=\langle h,\varphi^{\vee}\rangle\leq\|T_{h}\|_{C_{0}\to C_{\rm b}}\;\|\varphi\|_{L_{\infty}}

with φC0()\varphi^{\vee}\in C_{0}(\mathbb{R}) and φL=φL\|\varphi^{\vee}\|_{L_{\infty}}=\|\varphi\|_{L_{\infty}}. By substituting φ\varphi for φ\varphi^{\vee} and by recalling that 𝒟(){\mathcal{D}}(\mathbb{R}) is dense in C0()C_{0}(\mathbb{R}), we get that

supφC0()\{0}h,φφL\displaystyle\sup_{\varphi\in C_{0}(\mathbb{R})\backslash\{0\}}\frac{\langle h,\varphi\rangle}{\|\varphi\|_{L_{\infty}}} =supφ𝒟()\{0}h,φφL\displaystyle=\sup_{\varphi\in{\mathcal{D}}(\mathbb{R})\backslash\{0\}}\frac{\langle h,\varphi\rangle}{\|\varphi\|_{L_{\infty}}}
=hThC0Cb,\displaystyle=\|h\|_{{\mathcal{M}}}\leq\|{\mathrm{T}}_{h}\|_{C_{0}\to C_{\rm b}}, (20)

which then also proves that the bound is sharp.

Item 4. The key here is the estimate

|(hf)(t)|dt\displaystyle\int_{\mathbb{R}}\big{|}(h\ast f)(t)\big{|}\,{\rm d}t |f(tτ)|d|μh|(τ)dt\displaystyle\leq\int_{\mathbb{R}}\int_{\mathbb{R}}|f(t-\tau)|\,{\rm d}|\mu_{h}|(\tau)\,{\rm d}t
=(|f(x)|dx)d|μh|(τ)\displaystyle=\int_{\mathbb{R}}\left(\int_{\mathbb{R}}|f(x)|{\rm d}x\right){\rm d}|\mu_{h}|(\tau) (by Fubini)
=(|f(x)|dx)(d|μh|)\displaystyle=\left(\int_{\mathbb{R}}|f(x)|{\rm d}x\right)\left(\int_{\mathbb{R}}{\rm d}|\mu_{h}|\right)
=fL1h,\displaystyle=\|f\|_{L_{1}}\,\|h\|_{{\mathcal{M}}},

from which we deduce the boundedness of Th:L1()L1(){\mathrm{T}}_{h}:L_{1}(\mathbb{R})\to L_{1}(\mathbb{R}) with ThL1L1h\|{\mathrm{T}}_{h}\|_{L_{1}\to L_{1}}\leq\|h\|_{{\mathcal{M}}}. (The extension technique is essentially the same as in Item 1 with p=1p=1 and L1()=(𝒟(),L1)¯L_{1}(\mathbb{R})=\overline{({\mathcal{D}}(\mathbb{R}),\|\cdot\|_{L_{1}})}.) The converse implication and the sharpness of the bound will be deduced from Item 5 by duality.

Item 5. Since L()=(L1())L_{\infty}(\mathbb{R})=\big{(}L_{1}(\mathbb{R})\big{)}^{\prime}, the adjoint of Th:L1() c. L1(){\mathrm{T}}_{h}:L_{1}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{1}(\mathbb{R}) is Th=Th:L() c. L(){\mathrm{T}}_{h}^{\ast}={\mathrm{T}}_{h^{\vee}}:L_{\infty}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{\infty}(\mathbb{R}). The equivalence h()h()h\in{\mathcal{M}}(\mathbb{R})\Leftrightarrow h^{\vee}\in{\mathcal{M}}(\mathbb{R}) implies the continuity of Th:L() c. L(){\mathrm{T}}_{h}:L_{\infty}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{\infty}(\mathbb{R}) with ThLLh=h\|{\mathrm{T}}_{h}\|_{L_{\infty}\to L_{\infty}}\leq\|h\|_{\mathcal{M}}=\|h^{\vee}\|_{{\mathcal{M}}}. As for the converse implication, we take advantage of the embedding C0()iso.L()C_{0}(\mathbb{R})\xhookrightarrow{\mbox{\tiny\rm iso.}}L_{\infty}(\mathbb{R}), which allows us to reuse the argument of Item 3.

Item 2 and Its Converse. The first part follows from the beginning of the proof of Item 1, the application of the extension principle for p=1p=1, and the commutativity of the convolution integral, which yields fh=hfCb()f\ast h=h\ast f\in C_{\rm b}(\mathbb{R}) with hfLhL1fL\|h\ast f\|_{L_{\infty}}\leq\|h\|_{L_{1}}\|f\|_{L_{\infty}} for any fL()f\in L_{\infty}(\mathbb{R}). We show that the bound is sharp by applying the convolution operator to the “worst-case” signal f0f_{0} identified in (4). Conversely, let Th:L() c. Cb(){\mathrm{T}}_{h}:L_{\infty}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{\rm b}(\mathbb{R}) with ThLCb<\|T_{h}\|_{L_{\infty}\to C_{\rm b}}<{\infty}. Taking advantage of the isometric embedding Cb()iso.L()C_{\rm b}(\mathbb{R})\xhookrightarrow{\mbox{\tiny\rm iso.}}L_{\infty}(\mathbb{R}), we then invoke the equivalence in Item 5 to deduce that h<\|h\|_{{\mathcal{M}}}<{\infty}, which implies that h()h\in{\mathcal{M}}(\mathbb{R}). The announced equivalence then follows from Proposition 3 in Section IV-C. ∎

The result in Item 1 is discussed in most advanced treatises on the Fourier transform (e.g., [4, Proposition 8.8, p 241]). We are including it here in a self-contained form—at the expense of a few more lines in the proof of Item 2—because it nicely characterizes the regularization effect of convolution. The equivalences stated in Item 4 and Item 5 are known in the context of the theory of LpL_{p} Fourier multipliers [6, Section 2.5], even though the latter does not seem to have permeated to the engineering literature. The equivalence in Item 4 may also be identified as a special instance of Wendel’s theorem in the abstract theory of multipliers on locally compact Abelian groups [11, Theorem 0.1.1, p. 2]. Interestingly, the condition h()h\in{\mathcal{M}}(\mathbb{R}) is also sufficient for the continuity of Th:Lp() c. Lp(){\mathrm{T}}_{h}:L_{p}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{p}(\mathbb{R}), a claim that is supported by the Young-type norm inequality

hfLphfLp,\displaystyle\|h\ast f\|_{L_{p}}\leq\|h\|_{{\mathcal{M}}}\,\|f\|_{L_{p}}, (21)

which holds for any fLp()f\in L_{p}(\mathbb{R}) with p1p\geq 1. However, (21) is only sharp at the two end points p=1,+p=1,+\infty, in conformity with the statements in Items 4 and 5. In fact, the only other case where the complete class of convolution operators Th:Lp() c. Lp(){\mathrm{T}}_{h}:L_{p}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{p}(\mathbb{R}) has been characterized is for p=2p=2, with the necessary and sufficient condition being h^L()\widehat{h}\in L_{\infty}(\mathbb{R}) (bounded frequency response) [18, Theorem 3.18, p. 28], which is slightly more permissive than the BIBO requirement. Indeed, h()h^L()h\in{\mathcal{M}}(\mathbb{R})\Rightarrow\widehat{h}\in L_{\infty}(\mathbb{R}), whereas the reverse implication does not hold.

We like to single out Item 3 in Theorem 4 as the pivot point that facilitates the derivation of the (nontrivial) reverse implications—namely, the necessity of the condition h()h\in{\mathcal{M}}(\mathbb{R}). While the listed property is sufficient for our purpose, we can refer to a recent characterization by Feichtinger [3, Theorem 2, p. 499] which, in the present context, translates into the refined statement “h()Th:C0() c. C0()h\in{\mathcal{M}}(\mathbb{R})\Leftrightarrow{\mathrm{T}}_{h}:C_{0}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{0}(\mathbb{R}).” The additional element there is the vanishing of (hf)(t)(h\ast f)(t) at infinity, which calls for a more involved proof.

While the statement in Item 2 is a special case of Item 5, as made explicit in Section IV-C, the interesting part of the story is that this restriction induces a smoothing effect on the output, ensuring that the function t(hf)(t)t\mapsto(h\ast f)(t) is continuous. There is obviously no such effect for the case h=δ()h=\delta\in{\mathcal{M}}(\mathbb{R}) (identity) or, by extension, hd=na[n]δ(n)()h_{\rm d}=\sum_{n\in\mathbb{Z}}a[n]\delta(\cdot-n)\in{\mathcal{M}}(\mathbb{R}) with hd=a1\|h_{d}\|_{{\mathcal{M}}}=\|a\|_{\ell_{1}}, which corresponds to the continuous-time transposition of a digital filter.

IV-C Explicit Criterion for BIBO Stability

We now show how to determine h\|h\|_{{\mathcal{M}}} (our extended criterion for BIBO stability) under the assumption that hL1,loc()h\in L_{1,{\rm loc}}(\mathbb{R}). Any such impulse response can be identified with a distribution by considering the linear form

h:φh,φ=h(t)φ(t)dt,\displaystyle h:\varphi\mapsto\langle h,\varphi\rangle=\int_{\mathbb{R}}h(t)\varphi(t){\rm d}t, (22)

which continuously maps 𝒟(){\mathcal{D}}(\mathbb{R})\to\mathbb{R}. It turns out that the latter is a special instance of a real-valued Radon measure, which is an extended type of measure whose \|\cdot\|_{{\mathcal{M}}}-norm is not necessarily finite.

Definition 4 ([17]).

A distribution f𝒟()f\in{\mathcal{D}}^{\prime}(\mathbb{R}) is called a real-valued Radon measure if, for any compact subset 𝕂\mathbb{K}\subset\mathbb{R}, there exists a constant C𝕂>0C_{\mathbb{K}}>0 such that

f,φC𝕂supt𝕂|φ(t)|\displaystyle\langle f,\varphi\rangle\leq C_{\mathbb{K}}\sup_{t\in\mathbb{K}}|\varphi(t)| (23)

for all φ𝒟(𝕂)={φ𝒟():φ(t)=0,t𝕂}\varphi\in{\mathcal{D}}(\mathbb{K})=\big{\{}\varphi\in{\mathcal{D}}(\mathbb{R}):\varphi(t)=0,\forall t\notin\mathbb{K}\big{\}}.

A distribution f+𝒟()f^{+}\in{\mathcal{D}}^{\prime}(\mathbb{R}) is said to be positive if f+,φ0\langle f^{+},\varphi\rangle\geq 0 for all φ𝒟+()={φ𝒟():φ(t)0,t}\varphi\in{\mathcal{D}}^{+}(\mathbb{R})=\big{\{}\varphi\in{\mathcal{D}}(\mathbb{R}):\varphi(t)\geq 0,t\in\mathbb{R}\big{\}}.

The connection between the two kinds of distributions in Definition 4 is that a positive distribution is a special instance of a Radon measure, while any real-valued Radon measure ff admits a unique decomposition as f=(f+f)f=(f^{+}-f^{-}), where both f+,f0f^{+},f^{-}\geq 0 are positive distributions [19, Theorem 21.2, p. 218]. One then also defines the corresponding “total-variation measure” |f|=f++f|f|=f^{+}+f^{-}, which is positive by construction.

It turns out that the Dirac impulse δ\delta is a positive Radon measure with a universal bounding constant C𝕂=1C_{\mathbb{K}}=1. Likewise, the minimal constant in (23) for fL1,loc()f\in L_{1,{\rm loc}}(\mathbb{R}) is 𝕂=𝕂|f(t)|dt\mathbb{C}_{\mathbb{K}}=\int_{\mathbb{K}}|f(t)|{\rm d}t, which is essentially what is expressed in Proposition 3.

Proposition 3 (Total-variation norm for measurable functions).

Let hL1,loc()h\in L_{1,{\rm loc}}(\mathbb{R}). Then, h=hL1=+|h(t)|dt\|h\|_{{\mathcal{M}}}=\|h\|_{L_{1}}=\int_{-\infty}^{+\infty}|h(t)|{\rm d}t. Consequently, h()h\in{\mathcal{M}}(\mathbb{R}) if and only if +|h(t)|dt<\int_{-\infty}^{+\infty}|h(t)|{\rm d}t<\infty.

Proof:

In accordance with Definition 4, we view hL1,loc()h\in L_{1,{\rm loc}}(\mathbb{R}) as a real-valued Radon measure with h+(t)=max(h(t),0)h^{+}(t)=\max\big{(}h(t),0\big{)} and h(t)=max(0,h(t))h^{-}(t)=\max\big{(}0,-h(t)\big{)}, while the corresponding total-variation measure is |h|=h++hL1,loc()|h|=h^{+}+h^{-}\in L_{1,{\rm loc}}(\mathbb{R}) with |h|:t|h(t)||h|:t\mapsto|h(t)|, which is consistent with the notation. We then distinguish between two cases.

(i) Bounded Scenario. When h<\|h\|_{{\mathcal{M}}}<\infty, we can invoke the classical Jordan decomposition of a measure (see [4]),

f():f=f++f=|f|<,\displaystyle\forall f\in{\mathcal{M}}(\mathbb{R}):\ \|f\|_{{\mathcal{M}}}=\|f^{+}\|_{{\mathcal{M}}}+\|f^{-}\|_{{\mathcal{M}}}=\||f|\|_{{\mathcal{M}}}<\infty,

which allows us to reduce the problem to the easier determination of |h|\||h|\|_{\mathcal{M}}. Accordingly, for any given T>0T>0, we define hT=|h|𝟙[T,T]0h_{T}=|h|\cdot\mathbbm{1}_{[-T,T]}\geq 0 and observe that

hT=supφ𝒟():φL1hT,φ\displaystyle\|h_{T}\|_{{\mathcal{M}}}=\sup_{\varphi\in{\mathcal{D}}(\mathbb{R}):\,\|\varphi\|_{L_{\infty}}\leq 1}\langle h_{T},\varphi\rangle =hT,1\displaystyle=\langle h_{T},1\rangle
=hTL1<,\displaystyle=\|h_{T}\|_{L_{1}}<\infty,

where the supremum is achieved by considering any test function φT𝒟()\varphi_{T}\in{\mathcal{D}}(\mathbb{R}) such that φT(t)=1\varphi_{T}(t)=1 for all t[T,T]t\in[-T,T]. In the limit, we get that limThTL1=limThT=|h|<\lim_{T\to\infty}\|h_{T}\|_{L_{1}}=\lim_{T\to\infty}\|h_{T}\|_{{\mathcal{M}}}=\||h|\|_{{\mathcal{M}}}<\infty, from which we conclude that |h|=h=hL1\||h|\|_{{\mathcal{M}}}=\|h\|_{{\mathcal{M}}}=\|h\|_{L_{1}}.

(ii) Unbounded Scenario. The condition h=\|h\|_{{\mathcal{M}}}=\infty can be formalized as: for any nn\in\mathbb{N}, there exists φn𝒟()\varphi_{n}\in{\mathcal{D}}(\mathbb{R}) with φnL=1\|\varphi_{n}\|_{L_{\infty}}=1 such that h,φn>n\langle h,\varphi_{n}\rangle>n. However,

h,φn|h(t)φn(t)dt||h(t)|dtφnL=hL1.\displaystyle\langle h,\varphi_{n}\rangle\leq\big{|}\int_{\mathbb{R}}h(t)\varphi_{n}(t){\rm d}t\big{|}\leq\int_{\mathbb{R}}|h(t)|{\rm d}t\;\|\varphi_{n}\|_{L_{\infty}}=\|h\|_{L_{1}}.

Therefore, hL1>n\|h\|_{L_{1}}>n for all nn\in\mathbb{N}, leading to hL1=\|h\|_{L_{1}}=\infty.

Let us now conclude with a few more observations.

Since L1,loc()L_{1,{\rm loc}}(\mathbb{R}) can be identified as the subspace of measures that are absolutely continuous (see [17, p. 18]), the result in Proposition 3 is consistent with the well-known property in probability theory that L1()L_{1}(\mathbb{R}) coincides with the subset of bounded measures that are absolutely continuous.

Under the minimalistic assumption that hL1,loc()h\in L_{1,{\rm loc}}(\mathbb{R}), the convolution integral (1) is well defined for any tt\in\mathbb{R} provided that the input function f:f:\mathbb{R}\to\mathbb{R} is bounded and compactly supported. Equation (1) then even yields an output function t(hf)(t)t\mapsto(h\ast f)(t) that is continuous, as shown in Appendix D. However, the trouble comes from the fact that the output then inherits the potential lack of decay of hh when hL1()h\notin L_{1}(\mathbb{R}).

One can also make a connection between the result in Proposition 3 and the standard argument that is presented to justify the necessity of hL1()h\in L_{1}(\mathbb{R}). When the latter condition is fulfilled, we have that

h\displaystyle\|h\|_{{\mathcal{M}}} =supφ𝒟():φL1h,φ\displaystyle=\sup_{\varphi\in{\mathcal{D}}(\mathbb{R}):\,\|\varphi\|_{L_{\infty}}\leq 1}\langle h,\varphi\rangle
=hL1=supϕL():ϕL1h,ϕ=h(t)ϕ0(t)dt,\displaystyle=\|h\|_{L_{1}}=\sup_{\phi\in L_{\infty}(\mathbb{R}):\,\|\phi\|_{L_{\infty}}\leq 1}\langle h,\phi\rangle=\int_{\mathbb{R}}h(t)\phi_{0}(t){\rm d}t, (24)

where ϕ0(t)=sign(h(t))\phi_{0}(t)={\rm sign}\big{(}h(t)\big{)}. While the supremum is achieved exactly over L()L_{\infty}(\mathbb{R}) by taking ϕ=ϕ0\phi=\phi_{0}, it is a bit trickier over 𝒟(){\mathcal{D}}(\mathbb{R}) because of the additional smoothness requirement. Yet, due to the definition of the supremum, for any ϵ>0\epsilon>0 there exists a function φϵ𝒟()\varphi_{\epsilon}\in{\mathcal{D}}(\mathbb{R}) with φϵ=1\|\varphi_{\epsilon}\|_{\infty}=1 such that h(t)φϵ(t)dt=(1ϵ)hh=hL1\int h(t)\varphi_{\epsilon}(t){\rm d}t=(1-\epsilon)\|h\|_{{\mathcal{M}}}\leq\|h\|_{{\mathcal{M}}}=\|h\|_{L_{1}}. By taking ϵ\epsilon arbitrarily small, we end up with φϵ\varphi_{\epsilon} being a “smoothed” rendition of ϕ0\phi_{0}, so that the spirit of the initial argument is retained.

Appendix

A. Is the Dirac Distribution a Member of L1()L_{1}(\mathbb{R})?

Let us start with the historical observation that the eponymous impulse δ\delta is already present in the (early) works of both Fourier and Heaviside [10]. The former, as one would expect, defined it via an “improper” integral (the inverse Fourier transform of “11”), while the latter identified δ\delta as the “formal” derivative of the unit step (a.k.a. the Heaviside function). However, the mathematics for giving a rigorous sense to these identifications were missing at the time; one had to wait for the development Schwartz’ distribution theory in the 1950s [17], which already shows that the mere process of obtaining a rigorous definition of δ\delta was far from trivial.

From the pragmatic point of view of an engineer, the title question is at the heart of the matter to understand the scope of Proposition 1, and the source of some confusion, too. Let us start by listing the elements that could suggest that the answer to the question is positive.

  1. 1.

    It is common practice to make liberal use of what mathematicians consider abusive notations; in particular, equations such as f(t)=δ(τ)f(tτ)dτf(t)=\int_{\mathbb{R}}\delta(\tau)f(t-\tau){\rm d}\tau, which could suggest that δ(τ)\delta(\tau) can be manipulated as if it were a classical function of τ\tau.

  2. 2.

    Dirac’s δ\delta has the unit “integral” δ,1=1\langle\delta,1\rangle=1, which is indicated formally as δ(τ)dτ=1\int_{\mathbb{R}}\delta(\tau){\rm d}\tau=1. Moreover, δ0\delta\geq 0 in the sense that it is a positive distribution (see Definition 4).

  3. 3.

    The Dirac impulse is often described as the limit of φn(t)=n2πe(tn)2/2\varphi_{n}(t)=\frac{n}{\sqrt{2\pi}}\mathrm{e}^{-(tn)^{2}/2} as nn\to\infty, with φn𝒮()\varphi_{n}\in{\mathcal{S}}(\mathbb{R}). Since φnL1=1\|\varphi_{n}\|_{L_{1}}=1 for any n>0n>0, this could suggest that limnφnL1=1\lim_{n\to\infty}\|\varphi_{n}\|_{L_{1}}=1 as well.

In order to convince the reader that the answer to the title question is actually negative, we now refute these intuitive arguments one by one.

  1. 1.

    The explicit description of the Dirac impulse as a centered Gaussian distribution whose standard deviation σn=1/n\sigma_{n}=1/n tends to zero suggests that δ=limnφn\delta=\lim_{n\to\infty}\varphi_{n} must be entirely localized at t=0t=0. The best attempt at describing this limit in Lebesgue’s world of measurable functions would be

    p0(t)={+,t=00,otherwise,p_{0}(t)=\left\{\begin{array}[]{ll}+\infty,&t=0\\ 0,&\mbox{otherwise},\end{array}\right.

    which is equal to zero almost everywhere. However, since the width of the impulse is zero, we get that p0(t)dt=0\int_{\mathbb{R}}p_{0}(t){\rm d}t=0, which is incompatible with the property that δ(t)dt=1\int_{\mathbb{R}}\delta(t){\rm d}t=1. This points to the impossibility of representing δ\delta by a function in L1()L_{1}(\mathbb{R}) or even in L1,loc()L_{1,{\rm loc}}(\mathbb{R}). Strictly speaking, δ\delta is defined as a continuous linear functional on 𝒟(){\mathcal{D}}(\mathbb{R})—or, by extension, C0()C_{0}(\mathbb{R})—which precludes the application of any nonlinear operation (such as ||p|\cdot|^{p}) to it.

  2. 2.

    The generalized Fourier transform of δ\delta is {δ}=1\mathcal{F}\{\delta\}=1, which is bounded, but not decreasing at infinity. If δ\delta was included in L1()L_{1}(\mathbb{R}), this would contradict the Riemann-Lebesgue Lemma, which is equivalent to :L1() c. C0()\mathcal{F}:L_{1}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}C_{0}(\mathbb{R}) with L1C0=1\|\mathcal{F}\|_{L_{1}\to C_{0}}=1. By contrast, the inclusion δ()\delta\in{\mathcal{M}}(\mathbb{R}) is compatible with the (dual) continuity property of the Fourier transform ,:() c. L()\mathcal{F}^{\ast},\mathcal{F}:{\mathcal{M}}(\mathbb{R})\xrightarrow{\mbox{\tiny\ \rm c. }}L_{\infty}(\mathbb{R}) with L0=1\|\mathcal{F}^{\ast}\|_{{\mathcal{M}}\to L_{0}}=1.

  3. 3.

    While the sequence of rescaled Gaussians (φn)(\varphi_{n}) converges to δ𝒮()𝒟()\delta\in{\mathcal{S}}^{\prime}(\mathbb{R})\xhookrightarrow{}{\mathcal{D}}^{\prime}(\mathbb{R}) in the (weak) topology of 𝒮(){\mathcal{S}}^{\prime}(\mathbb{R}) (Schwartz’ space of tempered distributions), the problem is that it fails to be a Cauchy sequence in the (strong) norm topology of L1()L_{1}(\mathbb{R}). Hence, there is no guarantee that δ=limnφn\delta=\lim_{n\to\infty}\varphi_{n} stays in L1()L_{1}(\mathbb{R}).

B. Examples of Inaccurate Statements on BIBO Stability

This list is far from exhaustive and not intended to downplay the important contributions of the listed people who are internationally recognized leaders in the field. Its sole purpose is to illustrate the omnipresence of the misconception in the engineering literature, including in some of the most popular and authoritative textbooks in the theory of linear systems and signal processing.

As a start, one can read in the English version of Wikipedia that a necessary and sufficient condition for the BIBO stability of a convolution operator is that its impulse response be absolutely integrable, formulated as |h(τ)|dτ=hL1<\int_{\mathbb{R}}|h(\tau)|{\rm d}\tau=\|h\|_{L_{1}}<\infty. In view of the discussion around Proposition 1, this is only correct if one restricts the scope of the statement to those impulse responses that are Lebesgue-measurable and locally integrable.

Kailath mentions in [9, p. 175] that the equivalence between BIBO stability and hL1()h\in L_{1}(\mathbb{R}) is well known, and attributes the result to James, Nichols, and Phillips [8]. It turns out that the pioneers of the theory on control and linear systems were focusing their attention on analog systems ruled by ordinary differentiable equations whose impulse responses are sums of causal exponentials and, therefore, Lebesgue-measurable. Kailath then presents a proof on p. 176 that is essentially the one we used for Proposition 1, except that he neither considers a limit process nor explicitly says that hh must be (locally) integrable.

Oppenheim and Willsky discuss the property in [13, p. 113-114]. To justify the BIBO stability of the pure time-shift operator (including the identity), they then present an argument in support of the inclusion of δ(t0)\delta(\cdot-t_{0}) in L1()L_{1}(\mathbb{R}) (Example 2.13) which, in view of the discussion in Appendix A, is flawed.

Vetterli et al. claim in [20, Theorem 4.8, p. 357] that the operator Th{\mathrm{T}}_{h} is BIBO-stable from L()L()L_{\infty}(\mathbb{R})\to L_{\infty}(\mathbb{R}) if and only if hL1()h\in L_{1}(\mathbb{R}), a statement that is incompatible with Theorem 3. This can be corrected by limiting the scope of the equivalence as in the statement of Theorem 2.

IV-D Convolution in the “Unstable” Scenario

Here, we characterize the output of a potentially “unstable” filter when the input signal is compactly supported. The enabling hypothesis is the local integrability of the impulse response.

Proposition 4.

Let fL()f\in L_{\infty}(\mathbb{R}) be compactly supported and hL1,loc()h\in L_{1,{\rm loc}}(\mathbb{R}). Then, the function t(hf)(t)t\mapsto(h\ast f)(t) defined by (1) is bounded on any compact set 𝕂\mathbb{K}\subset\mathbb{R} and continuous; that is, hfC()h\ast f\in C(\mathbb{R}).

Proof.

Because of the local integrability of hh, the convolution integral (1) is well-defined for any tt\in\mathbb{R} with

(hf)(t)=h(τ)f(tτ)dτ=𝕄f(x)h(tx)dx\displaystyle(h\ast f)(t)=\int_{\mathbb{R}}h(\tau)f(t-\tau){\rm d}\tau=\int_{\mathbb{M}}f(x)h(t-x){\rm d}x (by change of variable)

and

|(hf)(t)|fL𝕄|h(t±τ)|dτ<,\displaystyle|(h\ast f)(t)|\leq\|f\|_{L_{\infty}}\int_{\mathbb{M}}|h(t\pm\tau)|{\rm d}\tau<\infty,

where 𝕄\mathbb{M} is the smallest symmetric interval such that f(t)=f(t)=0f(t)=f(-t)=0 for all t𝕄t\notin\mathbb{M}. For any given open bounded set 𝕂\mathbb{K}\subset\mathbb{R}, we then observe that

t𝕂:(hf)(t)=(h𝕂+𝕄f)(t)\displaystyle\forall t\in\mathbb{K}:\quad(h\ast f)(t)=(h_{\mathbb{K}+\mathbb{M}}\ast f)(t)

where h𝕂+𝕄=h𝟙𝕂+𝕄h_{\mathbb{K}+\mathbb{M}}=h\cdot\mathbbm{1}_{\mathbb{K}+\mathbb{M}} is the restriction of the original impulse response to the set 𝕂+𝕄={t+τ:t𝕂,τ𝕄}\mathbb{K}+\mathbb{M}=\{t+\tau:t\in\mathbb{K},\tau\in\mathbb{M}\}. Since h𝕂+𝕄L1()h_{\mathbb{K}+\mathbb{M}}\in L_{1}(\mathbb{R}), one has that

supt𝕂|hf(t)|fLh𝕂+𝕄L1<.\displaystyle\sup_{t\in\mathbb{K}}|h\ast f(t)|\leq\|f\|_{L_{\infty}}\|h_{\mathbb{K}+\mathbb{M}}\|_{L_{1}}<\infty. (25)

Likewise, for any t,t0𝕂t,t_{0}\in\mathbb{K}, we have that

|hf(t)\displaystyle|h\ast f(t) hf(t0)|\displaystyle-h\ast f(t_{0})|
fLh𝕂+𝕄(t)h𝕂+𝕄(t0)L1.\displaystyle\leq\|f\|_{L_{\infty}}\|h_{\mathbb{K}+\mathbb{M}}(t-\cdot)-h_{\mathbb{K}+\mathbb{M}}(t_{0}-\cdot)\|_{L_{1}}. (26)

Next, we invoke Lebesgue’s dominated-convergence theorem and the property that C0()C_{0}(\mathbb{R}) is dense in L1()L_{1}(\mathbb{R}) to show that h𝕂+𝕄(t)h𝕂+𝕄(t0)L10\|h_{\mathbb{K}+\mathbb{M}}(t-\cdot)-h_{\mathbb{K}+\mathbb{M}}(t_{0}-\cdot)\|_{L_{1}}\to 0 as tt0t\to t_{0}. This, together with (26), implies that limtt0|(hf)(t)(hf)(t0)|=0\lim_{t\to t_{0}}|(h\ast f)(t)-(h\ast f)(t_{0})|=0, which expresses the continuity of thf(t)t\mapsto h\ast f(t) at t=t0t=t_{0} for any t0𝕂t_{0}\in\mathbb{K}. ∎

Acknowledgments

The author is extremely thankful to Julien Fageot, Shayan Aziznejad and Hans Feichtinger for having spotted inconsistencies in earlier versions of the manuscript and for their helpful advice. He is also appreciative of Thomas Kailath and Vivek Goyal’s feedback, as well as that of the three anonymous reviewers.

References

  • [1] Jean-Michel Bony. Cours d’Analyse: Théorie des Distributions et Analyse de Fourier. Editions Ecole Polytechnique, 2001.
  • [2] Nicolas Bourbaki. Elements of Mathematics: 6. Integration. Springer, 2004.
  • [3] Hans G. Feichtinger. A novel mathematical approach to the theory of translation invariant linear systems. In Recent Applications of Harmonic Analysis to Function Spaces, Differential Equations, and Data Science, pages 483–516. Springer, 2017.
  • [4] Gerald B. Folland. Real Analysis: Modern Techniques and their Applications. John Wiley & Sons, 2013.
  • [5] I. M. Gelfand and N. Ya. Vilenkin. Generalized Functions. Vol. 4. Applications of Harmonic Analysis. Academic Press, New York, USA, 1964.
  • [6] Loukas Grafakos. Classical and Modern Fourier Analysis. Prentice Hall, 2004.
  • [7] Lars Hörmander. Estimates for translation invariant operators in LpL_{p} spaces. Acta Mathematica, 104(1):93–140, December 1960.
  • [8] Hubert M. James, Nathaniel B. Nichols, and Ralph S. Phillips. Theory of Servomechanisms, volume 25. McGraw-Hill New York, 1947.
  • [9] Thomas Kailath. Linear Systems, volume 156. Prentice-Hall Englewood Cliffs, NJ, 1980.
  • [10] Hikosaburo Komatsu. Fourier’s hyperfunctions and Heaviside’s pseudodifferential operators. In Takahiro Kawai and Keiko Fujita, editors, Microlocal Analysis and Complex Fourier Analysis, pages 200–214. World Scientific, 2002.
  • [11] Ronald Larsen. An Introduction to the Theory of Multipliers. Springer-Verlag, Berlin, 1970.
  • [12] Robert E. Megginson. An Introduction to Banach Space Theory. Springer, 1998.
  • [13] Alan V. Oppenheim and Alan S. Willsky. Signal and Systems. Prentice Hall, Upper Saddle River, NJ, 1996.
  • [14] Michael Reed and Barry Simon. Methods of Modern Mathematical Physics. Vol. 1: Functional Analysis. Academic Press, 1980.
  • [15] Walter Rudin. Real and Complex Analysis. McGraw-Hill, New York, 3rd edition, 1987.
  • [16] Laurent Schwartz. Théorie des noyaux. In Proc. International Congress of Mathematics 1950, volume 1, pages 220–230, Providence, RI, 1952. American Mathematical Society.
  • [17] Laurent Schwartz. Théorie des Distributions. Hermann, Paris, 1966.
  • [18] Elias M. Stein and Guido Weiss. Introduction to Fourier Analysis on Euclidean Spaces. Princeton University Press, Princeton, NJ, 1971.
  • [19] François Trèves. Topological Vector Spaces, Distributions and Kernels. Dover Publications, 2006.
  • [20] Martin Vetterli, Jelena Kovačević, and Vivek K. Goyal. Foundations of Signal Processing. Cambridge University Press, 2014.