Abstract

This paper presents a generalization of the concept and uses of projected dynamical systems to the case of nonpivot Hilbert spaces. These are Hilbert spaces in which the topological dual space is not identified with the base space. The generalization consists of showing the existence of such systems and their relation to variational problems, such as variational inequalities. In the case of usual Hilbert spaces these systems have been extensively studied, and, as in previous works, this new generalization has been motivated by applications, as shown below.

1. Introduction

In this paper we study the existence of solutions for a class of differential equations with discontinuous and nonlinear right-hand side on the class of nonpivot Hilbert spaces. This class of equations (called projected differential equations) was first introduced (in the form we use) in [1]; however have other studies of a similar formulation has been known since [2–4]. The formulation of the flow of such equations as dynamical systems in ℝ𝑛 is due to [1, 5], and it has been applied to study the dynamics of solutions of finite-dimensional variational inequalities in [5, 6].

Finite-dimensional variational inequalities theory provides solutions to a wide class of equilibrium problems in mathematical economics, optimization, management science, operations research, finance, and so forth (see, e.g., [4, 6–8] and the references therein). Therefore there has been a steady interest over the years in studying the stability of solutions to finite-dimensional variational inequalities (and consequently the stability of equilibria for various problems). In general, such a study is done by associating a projected dynamical system to a variational inequality problem; however in the past few years the applied problems, as well as the theoretical results, have progressed to a qualitative study of stability of solutions to variational inequality problems on Hilbert spaces and even on Banach spaces. Examples of the kind of variational problems (and their applications) can be found in see [9–19] and the references therein).

In this paper we present a new step in this study: we show that a projected differential equation has solutions on a non-pivot Hilbert space of any dimension. We prove the existence and uniqueness of integral curves and show they remain in a given constraint set of interest. As in the finite-dimensional case, a dynamics given by solutions to a projected differential equation is interesting because it describes these problems as dynamical systems. Moreover, as shown in this paper, the new results were needed to be developed for the study of the weighted traffic equilibrium problem (see [20]). Our goal in this paper is to present the mathematical techniques involved in proving the existence of solutions to projected differential equations in a non-pivot setting, which is in fact similar to the one in [21], but adapted to a non-pivot space; in addition, there are a number of preliminary results needed prior to obtaining our main result, which are remarkable since they also hold in a larger setting, namely, that of a reflexive Banach space (see the results in [22, 23]). Last but not least, we also present a projected system formulation called implicit. These kinds of systems have been introduced in the literature in [24], but without any existence result being presented in their case. We thus solve this additional problem in this paper as well.

2. Background Material

In this section we present several definitions and results pertinent to the reader and considered essential for the presentation of the later material.

2.1. Dual Realization of a Hilbert Space

Each time we work with a Hilbert space 𝑉, it is necessary to decide whether or not we identify the topological dual space π‘‰βˆ—=β„’(𝑉,ℝ) with 𝑉. Commonly this identification is made, one of the reasons for this being that the vectors of the polar of a set of 𝑉 are in 𝑉. In some cases the identification does not make sense. For clarity of presentation, we remind below of the basic results regarding the dual realization of a Hilbert space. The readers can refer to [25] for additional information.

First, consider a pre-Hilbert space 𝑉 with an inner product ((π‘₯,𝑦)), and its topological dual π‘‰βˆ—=β„’(𝑉,ℝ). It is well known that π‘‰βˆ— is a Banach space for the classical dual norm (β€–π‘“β€–βˆ—=supπ‘₯βˆˆπ‘‰(|𝑓(π‘₯)|/β€–π‘₯β€–)). It is also known that there exists an isometry π½βˆΆπ‘‰β†’π‘‰βˆ— such that 𝐽 is linear and for all π‘₯βˆˆπ‘‰, 𝐽(π‘₯)=grad(β€–π‘₯β€–2/2). This mapping 𝐽 is called a duality mapping of (𝑉,π‘‰βˆ—).

Theorem 2.1 (Theorem  1 page 68, [25]). Let 𝑉 be a Hilbert space with the inner product ((π‘₯,𝑦)) and π½βˆˆβ„’(𝑉,π‘‰βˆ—) the duality mapping above. Then J is a surjective isometry from 𝑉 to π‘‰βˆ—. The dual space π‘‰βˆ— is a Hilbert space with the inner product: ((𝑓,𝑔))βˆ—=π½ξ€·ξ€·βˆ’1𝑓,π½βˆ’1𝑔𝐽=π‘“βˆ’1𝑔.(2.1)

Theorem 2.2 (Theorem  2 page 69, [25]). Let V be a pre-Hilbert space. Then there exists a completion 𝑉 of V, that is, an isometry j from V to the Hilbert space 𝑉 such that 𝑗(𝑉) is dense in 𝑉.

Definition 2.3. Let 𝑉 be a Hilbert space. We call {𝐹,𝑗}, where(i)𝐹 is a Hilbert space,(ii)𝑗 is an isometry from 𝐹 to β„’(𝑉,ℝ), a dual realization of 𝑉. We then set βŸ¨π‘“,π‘₯⟩=π‘—βˆ˜π‘“(π‘₯),βˆ€π‘“βˆˆπΉ,βˆ€π‘₯βˆˆπ‘‰,(2.2) where βŸ¨π‘“,π‘₯⟩ is the duality pairing for 𝐹×𝑉.

Remark 2.4. The duality pairing is a nondegenerate bilinear form on 𝐹×𝑉 and ‖𝑓‖𝐹=supπ‘₯βˆˆπ‘‰(|βŸ¨π‘“,π‘₯⟩|/β€–π‘₯β€–). These properties permit us to prove that 𝐹 is isomorphic to π‘‰βˆ—.

We deduce from Theorems 2.1 and 2.2 that π‘˜=π‘—βˆ’1βˆ˜π½βˆˆβ„’(𝑉,𝐹) is a surjective isometry such that(π‘₯,𝑦)=βŸ¨π‘˜(π‘₯),π‘¦βŸ©.(2.3) We use the following convention here: when a dual realization {𝐹,𝑗} of a space has been chosen, we set 𝐹=π‘‰βˆ— and π‘—βˆ˜π‘“(π‘₯)=βŸ¨π‘“,π‘₯⟩. We say that the isometry π‘˜βˆΆπ‘‰β†’π‘‰βˆ— is the duality operator associated to the inner product on 𝑉 and to the duality pairing on π‘‰βˆ—Γ—π‘‰ by the relation(π‘₯,𝑦)=βŸ¨π‘˜(π‘₯),π‘¦βŸ©.(2.4) A special but most frequent case is to choose a dual realization of 𝑉 the couple {𝑉,𝐽}; in this case the Hilbert space 𝑉 is called a pivot space. To be more precise, we introduce the following definition.

Definition 2.5. A Hilbert space 𝐻 with an inner product (π‘₯,𝑦) is called a pivot space, if we identify π»βˆ— with 𝐻. In that case π»βˆ—=𝐻,𝑗=𝐽,⟨π‘₯,π‘¦βŸ©=(π‘₯,𝑦).(2.5)

Sometimes it does not make sense to identify the space itself with its topological dual, as the following example shows.

Let us consider 𝑉=𝐿2(ℝ,(1+|π‘₯|))βŠ‚πΏ2(ℝ) (dense subspace of 𝐿2(ℝ)) endowed with the inner product:(𝑒,𝑣)𝑉=ξ€œβ„(1+|π‘₯|)𝑒(π‘₯)𝑣(π‘₯)𝑑π‘₯.(2.6)An element πœ‘βˆˆπΏ2(ℝ)βˆ— is also an element of π‘‰βˆ—. If we identify πœ‘ to an element π‘“βˆˆπΏ2(ℝ), this function does not define a linear form on 𝑉, and the expression πœ‘(𝑣)=βŸ¨π‘“,π‘£βŸ©π‘‰ has no meaning on 𝑉. In this situation it is necessary to work in a non-pivot Hilbert space. We provide now some useful examples of non-pivot H-spaces.

Let Ξ©βŠ‚β„π‘› be an open subset of, π‘ŽβˆΆΞ©β†’π‘…+⧡{0}, a continuous and strictly positive function called β€œweight” and π‘ βˆΆΞ©β†’π‘…+⧡{0}, a continuous and strictly positive function called β€œreal time density.” The bilinear form defined on π’ž0(Ξ©) (continuous functions with compact support on Ξ©) by(π‘₯,𝑦)π‘Ž,𝑠=ξ€œΞ©π‘₯(πœ”)𝑦(πœ”)π‘Ž(πœ”)𝑠(πœ”)π‘‘πœ”(2.7) is an inner product. We remark here that if π‘Ž is a weight, then π‘Žβˆ’1=1/π‘Ž is also a weight. Let us introduce the following.

Definition 2.6. We call 𝐿2(Ξ©,π‘Ž,𝑠) a completion of π’ž0(Ξ©) for the inner product ⟨π‘₯,π‘¦βŸ©π‘Ž,𝑠.

We now introduce an 𝑛-dimensional version of the previous space. If we denote by 𝑉𝑖=𝐿2(Ξ©,ℝ,π‘Žπ‘–,𝑠𝑖) and π‘‰βˆ—π‘–=𝐿2(Ξ©,ℝ,π‘Žπ‘–βˆ’1,𝑠𝑖), the space𝑉=π‘šξ‘π‘–=1𝑉𝑖(2.8) is a non-pivot Hilbert space with the inner product:(𝐹,𝐺)𝑉=(𝐹,𝐺)𝐚,𝐬=π‘šξ“π‘–=1ξ€œΞ©πΉπ‘–(πœ”)𝐺𝑖(πœ”)π‘Žπ‘–(πœ”)𝑠𝑖(πœ”)π‘‘πœ”.(2.9) The spaceπ‘‰βˆ—=π‘šξ‘π‘–=1π‘‰βˆ—π‘–(2.10) is clearly a non-pivot Hilbert space for the following inner product(𝐹,𝐺)π‘‰βˆ—=(𝐹,𝐺)πšβˆ’1,𝐬=π‘šξ“π‘–=1ξ€œΞ©πΉπ‘–(πœ”)𝐺𝑖(πœ”)𝑠𝑖(πœ”)π‘Žπ‘–(πœ”)π‘‘πœ”,(2.11) and the following bilinear formπ‘‰βˆ—Γ—π‘‰βŸΆβ„,βŸ¨π‘“,π‘₯βŸ©π‘‰βˆ—Γ—π‘‰=βŸ¨π‘“,π‘₯⟩𝐬=π‘šξ“π‘–=1ξ€œΞ©π‘“π‘–(πœ”)π‘₯𝑖(πœ”)𝑠𝑖(πœ”)π‘‘πœ”(2.12) defines a duality between 𝑉 and π‘‰βˆ—. More precisely we have the following (see [20] for a proof).

Proposition 2.7. The bilinear form (2.12) defines a duality mapping between π‘‰βˆ—Γ—π‘‰, given by π½ξ€·π‘Ž(𝐹)=1𝐹1,…,π‘Žπ‘šπΉπ‘šξ€Έ.(2.13)

For applications of these spaces, the reader can refer to [20].

2.2. Variational Analysis in Non-Pivot H-Spaces

Let 𝑋 be a Hilbert space of arbitrary (finite or infinite) dimension and let πΎβŠ‚π‘‹ be a nonempty, closed, convex subset. We assume the reader is familiar with tangent and normal cones to 𝐾 at π‘₯∈𝐾 (𝑇𝐾(π‘₯), respectively, 𝑁𝐾(π‘₯)), and with the projection operator of 𝑋 onto 𝐾, π‘ƒπΎβˆΆπ‘‹β†’πΎ given by ‖𝑃𝐾(𝑧)βˆ’π‘§β€–=infπ‘₯βˆˆπΎβ€–π‘₯βˆ’π‘§β€–. Moreover we use here the following characterization of 𝑃𝐾(π‘₯):π‘₯=𝑃𝐾𝐽(π‘₯)⟺π‘₯βˆ’π‘₯ξ€Έ,π‘¦βˆ’π‘₯≀0,βˆ€π‘¦βˆˆπΎ.(2.14) The properties of the projection operator on Hilbert and Banach spaces are well known (see e.g., [26–28]). The directional derivative of the operator 𝑃𝐾 is defined, for any π‘₯∈𝐾 and any element π‘£βˆˆπ‘‹, as the limit (for a proof see [26]):πœ‹πΎ(π‘₯,𝑣)∢=lim𝛿→0+𝑃𝐾(π‘₯+𝛿𝑣)βˆ’π‘₯𝛿;moreoverπœ‹πΎ(π‘₯,𝑣)=𝑃𝑇𝐾(π‘₯)(𝑣).(2.15) Let πœ‹πΎβˆΆπΎΓ—π‘‹β†’π‘‹ be the operator given by (π‘₯,𝑣)β†¦πœ‹πΎ(π‘₯,𝑣). Note that πœ‹πΎ is nonlinear and discontinuous on the boundary of the set 𝐾. In [1, 29] several characterizations of πœ‹πΎ are given.

The following theorem has been proven in the framework of reflexive strictly convex and smooth Banach spaces. We will use it to obtain a decomposition theorem in non-pivot Hilbert spaces (for a proof see [30, Th. 2.4]).

Theorem 2.8. Let 𝑋 be a real reflexive strictly convex and smooth Banach space, and let 𝐢 be a non-empty, closed and convex cone of 𝑋. Then for all π‘₯βˆˆπ‘‹ and for all π‘“βˆˆπ‘‹βˆ— the following decompositions hold: π‘₯=𝑃𝐢(π‘₯)+π½βˆ’1Π𝐢0𝐽(π‘₯),⟨Π𝐢0𝐽(π‘₯),𝑃𝐢(π‘₯)⟩=0,𝑓=𝑃𝐢0(𝑓)+π½Ξ πΆπ½βˆ’1(𝑓),βŸ¨π‘ƒπΆ0(𝑓),Ξ πΆπ½βˆ’1(𝑓)⟩=0.(2.16) Here 𝑃𝐢 is the metric projection operator on 𝐾, and Π𝐢0 is the generalized projection operator on 𝐢0 (for a definition of Π𝐢0 see [28]).

Remark 2.9. It is known that 𝑃𝐢 and Π𝐢 coincide whenever the cone 𝐢 belongs to a Hilbert space. This observation implies the following result.

Corollary 2.10. Let 𝐢 be a nonempty closed convex cone of a non-pivot Hilbert space 𝑋. Then for all π‘₯βˆˆπ‘‹ and π‘“βˆˆπ‘‹βˆ— the following decompositions hold: π‘₯=𝑃𝐢(π‘₯)+π½βˆ’1𝑃𝐢0𝐽(π‘₯),βŸ¨π‘ƒπΆ0𝐽(π‘₯),𝑃𝐢(π‘₯)⟩=0,𝑓=𝑃𝐢0(𝑓)+π½π‘ƒπΆπ½βˆ’1𝑃(𝑓),𝐢0(𝑓),π‘ƒπΆπ½βˆ’1(𝑓)=0.(2.17)

We highlight that Zarantonello has shown in [27] a similar decomposition result in reflexive Banach spaces.

Lemma 2.11 ([26, Lemma  4.5]). For any closed convex set 𝐾, 𝑃𝐾(π‘₯+β„Ž)=π‘₯+β„Ž+∘(β€–β„Žβ€–),π‘₯∈𝐾,β„Žβˆˆπ‘‡πΎ(π‘₯),(2.18)where ∘(β€–β„Žβ€–)/β€–β„Žβ€–β†’0 as β„Žβ†’0 over any locally compact cone of increments.

Remark 2.12. To prove Lemma 2.11 only the properties of the norm in Hilbert spaces are used; therefore the proof is valid in the non-pivot setting.

The following lemma has been proven in the pivot case in [26]. We give below a similar proof in non-pivot spaces.

Lemma 2.13. For any π‘₯∈𝐾, 𝑃𝐾(π‘₯+β„Ž)=π‘₯+𝑃𝑇𝐾(π‘₯)(β„Ž)+∘(β€–β„Žβ€–),(2.19) where ∘(β€–β„Žβ€–)/β€–β„Žβ€–β†’0 as β„Žβ†’0 over any locally compact cone of increments.

Proof. Clearly, we have in general that β€–π‘Ž+𝑏‖2=β€–π‘Žβ€–2+‖𝑏‖2+2(π‘Ž,𝑏).(2.20)Taking π‘ŽβˆΆ=π‘₯+β„Žβˆ’π‘ƒπ‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž),π‘βˆΆ=𝑃π‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž)βˆ’π‘ƒπΎ(π‘₯+β„Ž),(2.21)we get β€–β€–π‘₯+β„Žβˆ’π‘ƒπΎβ€–β€–(π‘₯+β„Ž)2=β€–β€–π‘₯+β„Žβˆ’π‘ƒπ‘₯+𝑇𝐾(π‘₯)β€–β€–(π‘₯+β„Ž)2+‖‖𝑃π‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž)βˆ’π‘ƒπΎβ€–β€–(π‘₯+β„Ž)2ξ€·+2π‘₯+β„Žβˆ’π‘ƒπ‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž),𝑃π‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž)βˆ’π‘ƒπΎξ€Έ,(π‘₯+β„Ž)(2.22) but ξ€·π‘₯+β„Žβˆ’π‘ƒπ‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž),𝑃π‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž)βˆ’π‘ƒπΎξ€Έ=𝐽(π‘₯+β„Ž)π‘₯+β„Žβˆ’π‘ƒπ‘₯+𝑇𝐾(π‘₯)(ξ€Έπ‘₯+β„Ž),𝑃π‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž)βˆ’π‘ƒπΎ(π‘₯+β„Ž)β‰₯0(2.23) using the variational principle (2.14) applied to 𝑃π‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž). By definition of the projection operator we have β€–β€–π‘₯+β„Žβˆ’π‘ƒπΎβ€–β€–(π‘₯+β„Ž)2≀‖‖π‘₯+β„Žβˆ’π‘ƒπΎξ€Ίπ‘ƒπ‘₯+𝑇𝐾(π‘₯)ξ€»β€–β€–(π‘₯+β„Ž)2.(2.24)Therefore we have β€–β€–π‘₯+β„Žβˆ’π‘ƒπ‘₯+𝑇𝐾(π‘₯)β€–β€–(π‘₯+β„Ž)2+‖‖𝑃π‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž)βˆ’π‘ƒπΎβ€–β€–(π‘₯+β„Ž)2≀‖‖π‘₯+β„Žβˆ’π‘ƒπΎξ€Ίπ‘ƒπ‘₯+𝑇𝐾(π‘₯)ξ€»β€–β€–(π‘₯+β„Ž)2.(2.25)As 𝑃π‘₯+𝑇𝐾(π‘₯)(π‘₯+β„Ž)=π‘₯+𝑃𝑇𝐾(π‘₯)(β„Ž) (just apply the definition and the variational principle (2.14)), we have β€–β€–β„Žβˆ’π‘ƒπ‘‡πΎ(π‘₯)β€–β€–(β„Ž)2+β€–β€–π‘₯+𝑃𝑇𝐾(π‘₯)(β„Ž)βˆ’π‘ƒπΎβ€–β€–(π‘₯+β„Ž)2≀‖‖π‘₯+β„Žβˆ’π‘ƒπΎξ€·π‘₯+𝑃𝑇𝐾(π‘₯)ξ€Έβ€–β€–(β„Ž)2,(2.26)but using the Corollary 2.10 we have β„Ž=𝑃𝑇𝐢(π‘₯)(β„Ž)+π½βˆ’1𝑃𝑁𝐾(π‘₯)(𝐽(β„Ž)), and therefore, ‖‖𝑃𝐾(π‘₯+β„Ž)βˆ’π‘₯βˆ’π‘ƒπ‘‡πΎ(π‘₯)β€–β€–(β„Ž)2β‰€β€–β€–π½βˆ’1𝑃𝑁𝐾(π‘₯)(𝐽(β„Ž))+π‘₯+𝑃𝑇𝐾(π‘₯)(β„Ž)βˆ’π‘ƒπΎξ€·π‘₯+𝑃𝑇𝐾(π‘₯)ξ€Έβ€–β€–(β„Ž)2βˆ’β€–β€–π½βˆ’1𝑃𝑁𝐾(π‘₯)β€–β€–(𝐽(β„Ž))2≀‖‖π‘₯+𝑃𝑇𝐾(π‘₯)(β„Ž)βˆ’π‘ƒπΎξ€·π‘₯+𝑃𝑇𝐾(π‘₯)ξ€Έβ€–β€–(β„Ž)2‖‖𝐽+2βˆ’1𝑃𝑁𝐾(π‘₯)β€–β€–β€–β€–(𝐽(β„Ž))π‘₯+𝑃𝑇𝐾(π‘₯)(β„Ž)βˆ’π‘ƒπΎξ€·π‘₯+𝑃𝑇𝐾(π‘₯)ξ€Έβ€–β€–.(β„Ž)(2.27)But by Lemma 2.11, π‘₯+𝑃𝑇𝐾(π‘₯)(β„Ž)βˆ’π‘ƒπΎ(π‘₯+𝑃𝑇𝐾(π‘₯)(β„Ž))=π‘œ(‖𝑃𝑇𝐾(π‘₯)(β„Ž)β€–), so we can write ‖‖𝑃𝐾(π‘₯+β„Ž)βˆ’π‘₯βˆ’π‘ƒπ‘‡πΎ(π‘₯)β€–β€–(β„Ž)2≀2β€–β€–π½βˆ’1𝑃𝑁𝐾(π‘₯)‖‖‖‖𝑃(𝐽(β„Ž))+π‘œπ‘‡πΎ(π‘₯)β€–β€–π‘œξ€·β€–β€–π‘ƒ(β„Ž)𝑇𝐾(π‘₯)β€–β€–ξ€Έ.(β„Ž)(2.28)Therefore we have, ‖‖𝑃𝐾(π‘₯+β„Ž)βˆ’π‘₯βˆ’π‘ƒπ‘‡πΎ(π‘₯)β€–β€–(β„Ž)2β‰€π‘œ(β€–β„Žβ€–)2.(2.29)

3. Non-Pivot and Implicit PDS in Hilbert Spaces

3.1. PDS in Pivot H-Spaces

Let 𝑋 be a pivot Hilbert space of arbitrary (finite or infinite) dimension and let πΎβŠ‚π‘‹ be a nonempty, closed, convex subset. The following result has been shown (see [21]).

Theorem 3.1. Let 𝑋 be a Hilbert space and let 𝐾 be a nonempty, closed, convex subset. Let πΉβˆΆπΎβ†’π‘‹ be a Lipschitz continuous vector field and let π‘₯0∈𝐾. Then the initial value problem associated to the projected differential equation (PrDE) 𝑑π‘₯(𝜏)π‘‘πœ=πœ‹πΎ(π‘₯(𝜏)),βˆ’πΉ(π‘₯(𝜏)),π‘₯(0)=π‘₯0∈𝐾(3.1) has a unique absolutely continuous solution on the interval [0,∞).

This result is a generalization of the one in [6], where π‘‹βˆΆ=ℝ𝑛, 𝐾 was a convex polyhedron and 𝐹 had linear growth.

Definition 3.2. A projected dynamical system then is given by a mapping πœ™βˆΆβ„+×𝐾→𝐾 which solves the initial value problem: Μ‡πœ™(𝑑,π‘₯)=πœ‹πΎ(πœ™(𝑑,π‘₯),βˆ’πΉ(πœ™(𝑑,π‘₯)))a.a.𝑑, πœ™(0,π‘₯)=π‘₯0∈𝐾.

3.2. PDS in Non-Pivot H-Spaces

In this subsection we show that, with minor modifications, the existence of PDS in non-pivot H-spaces can be obtained. We first introduce non-pivot projected dynamical systems (NpPDSs) and then show their existence. In analogy with [21] we first introduce the following.

Definition 3.3. For πΉβˆΆπΎβ†’π‘‹βˆ—, a non-pivot projected differential equation (NpPrDE) is a discontinuous ODE given by 𝑑π‘₯(𝑑)𝑑𝑑=πœ‹πΎξ€·ξ€·π½π‘₯(𝑑),βˆ’βˆ’1ξ€Έξ€Έβˆ˜πΉ(π‘₯(𝑑))=𝑃𝑇𝐾(π‘₯(𝑑))ξ€·βˆ’ξ€·π½βˆ’1ξ€Έξ€Έ.∘𝐹(π‘₯(𝑑))(3.2)

Consequently the associated Cauchy problem is given by𝑑π‘₯(𝑑)𝑑𝑑=πœ‹πΎξ€·ξ€·π½π‘₯(𝑑),βˆ’βˆ’1ξ€Έξ€Έβˆ˜πΉ(π‘₯(𝑑)),π‘₯(0)=π‘₯0∈𝐾.(3.3) Next we define what we mean by a solution for a Cauchy problem of type (3.3).

Definition 3.4. An absolutely continuous function π‘₯βˆΆβ„βŠ‚β„β†’π‘‹, such that π‘₯(𝑑)∈𝐾,π‘₯(0)=π‘₯0∈𝐾,βˆ€π‘‘βˆˆβ„,Μ‡π‘₯(𝑑)=πœ‹πΎξ€·ξ€·π½π‘₯(𝑑),βˆ’βˆ’1ξ€Έξ€Έβˆ˜πΉ(π‘₯(𝑑)),a.e.onℐ(3.4) is called a solution for the initial value problem (3.3).

Finally, assuming that problem (3.3) has solutions as described above, then we are ready to introduce the following.

Definition 3.5. A non-pivot projected dynamical system (NpPDS) is given by a mapping πœ™βˆΆβ„+×𝐾→𝐾 which solves the initial value problem Μ‡πœ™(𝑑,π‘₯)=πœ‹πΎ(πœ™(𝑑,π‘₯),βˆ’(π½βˆ’1∘𝐹)(πœ™(𝑑,π‘₯))),a.a.𝑑,πœ™(0,π‘₯)=π‘₯0∈𝐾.

To end this section we show how problem (3.3) can be equivalently (in the sense of solution set coincidence) formulated as a differential inclusion problem. Finally, in Subsection 3.3 we show that solutions for this new differential inclusion problem exist. We introduce the following differential inclusion:Μ‡π‘₯(𝑑)βˆˆπ½βˆ’1ξ€·βˆ’πΉ(π‘₯)βˆ’π‘πΎξ€Έ(π‘₯),π‘₯(0)=π‘₯0∈𝐾,(3.5) and we call π‘₯βˆΆβ„βŠ‚β„β†’π‘‹ absolutely continuous a solution to (3.5) ifπ‘₯(𝑑)∈𝐾,π‘₯(0)=π‘₯0∈𝐾,βˆ€π‘‘βˆˆβ„,Μ‡π‘₯(𝑑)βˆˆπ½βˆ’1ξ€·βˆ’πΉ(π‘₯)βˆ’π‘πΎ(ξ€Έπ‘₯),a.a.𝑑.(3.6) We introduce also the following differential inclusion:Μ‡π‘₯(𝑑)βˆˆπ½βˆ’1ξ‚€ξ‚π‘βˆ’πΉ(π‘₯)βˆ’πΎξ‚(π‘₯),π‘₯(0)=π‘₯0∈𝐾,(3.7) where𝑁𝐾(π‘₯)=π‘›βˆˆπ‘πΎβ€–ξ€Ύ.(π‘₯)βˆ£β€–π‘›β€–β‰€β€–πΉ(π‘₯)(3.8)

Obviously, we call π‘₯βˆΆβ„βŠ‚β„β†’π‘‹ absolutely continuous a solution to (3.7) ifπ‘₯(𝑑)∈𝐾,π‘₯(0)=π‘₯0∈𝐾,βˆ€π‘‘βˆˆβ„,Μ‡π‘₯(𝑑)βˆˆπ½βˆ’1ξ‚€ξ‚π‘βˆ’πΉ(π‘₯)βˆ’πΎξ‚(π‘₯),a.a.𝑑.(3.9)

Proposition 3.6. The solution set of problem (3.3) coincides with the solution set of problem (3.9).

Proof. (3.3)β‡’(3.9). Let π‘₯(β‹…) be an absolutely continuous function on 𝐾 such that π‘₯(β‹…) is a solution to (3.3). Then π‘₯(𝑑)∈𝐾, for all π‘‘βˆˆπΌ and Μ‡π‘₯(𝑑)=πœ‹πΎ(π‘₯(𝑑),βˆ’(π½βˆ’1∘𝐹)(π‘₯(𝑑))),a.e.onℐ; therefore using Corollary 2.10 we get Μ‡π‘₯(𝑑)=βˆ’π½βˆ’1(𝐹(π‘₯))βˆ’π½βˆ’1𝑃𝑁𝐾(π‘₯)(βˆ’πΉ(π‘₯)), a.e.∈𝐼. Evidently, 𝑃𝑁𝐾(π‘₯)(βˆ’πΉ(π‘₯))βˆˆπ‘πΎ(π‘₯). Moreover as 𝑁𝐾(π‘₯) is a closed, convex cone, we get that ‖𝑃𝑁𝐾(π‘₯)(βˆ’πΉ(π‘₯))β€–π‘‹βˆ—β‰€β€–βˆ’πΉ(π‘₯)β€–π‘‹βˆ—)𝑁0𝐾(π‘₯)=𝑇𝐾(π‘₯) and both contains 0). Therefore βˆƒΜƒπ‘›πΎξ‚π‘(π‘₯)∈𝐾(π‘₯),̃𝑛𝐾(π‘₯)∢=𝑃𝑁𝐾(π‘₯)(βˆ’πΉ(π‘₯)) such that Μ‡π‘₯(𝑑)=βˆ’π½βˆ’1(𝐹(π‘₯(𝑑))βˆ’Μƒπ‘›πΎ(π‘₯)) for a.a π‘‘βˆˆπΌ, so we have Μ‡π‘₯(𝑑)βˆˆβˆ’π½βˆ’1𝑁(𝐹(π‘₯(𝑑))βˆ’πΎ(π‘₯)) for a.a π‘‘βˆˆπΌ, and π‘₯(β‹…) is a solution to (3.9).
(3.9)β‡’(3.3). As the trajectory remains in 𝐾 it is clear that Μ‡π‘₯(𝑑)βˆˆπ‘‡πΎ(π‘₯(𝑑)). First we show that for almost all π‘‘βˆˆπΌ we have Μ‡π‘₯(𝑑)βˆˆπ‘βŸ‚πΎ(π‘₯(𝑑)).(3.10) Let us consider three different cases; first suppose that π‘₯(𝑑)∈int(𝐾), we have then 𝑁𝐾(π‘₯(𝑑))={0π‘‹βˆ—} and then π‘βŸ‚πΎ(π‘₯(𝑑))=π‘‹βˆ— and (3.10) is automatically satisfied. Suppose now that π‘₯(𝑑)βˆˆπœ•πΎ and in π‘₯(𝑑), πœ•πΎ is smooth. In that case 𝑇𝐾(π‘₯(𝑑)) is flat and π‘βŸ‚πΎ(π‘₯(𝑑))βŠŠπ‘‡πΎ(π‘₯(𝑑)) with π‘βŸ‚πΎ(π‘₯(𝑑)) not reduced to {0π‘‹βˆ—}, if Μ‡π‘₯(𝑑)βˆ‰π‘βŸ‚πΎ(π‘₯(𝑑)); then in a neighbourhood 𝒱(𝑑) the trajectory π‘₯(𝑑′),π‘‘β€²βˆˆπ’±(𝑑) goes in int(𝐾), so we are in the first case and we can exclude time 𝑑. Suppose now that π‘₯(𝑑)βˆˆπœ•πΎ and π‘₯(𝑑) is in a corner point. In that case π‘βŸ‚πΎ(π‘₯(𝑑))={0}; therefore if Μ‡π‘₯(𝑑)=0, (3.10) is satisfied. If Μ‡π‘₯(𝑑)β‰ 0, it means that π‘₯(𝑑′)β‰ π‘₯(𝑑) for π‘‘β€²βˆˆπ’±(𝑑), with π‘₯(𝑑′) in one of the two previous cases; as we can β€œexclude” time 𝑑, we have (3.10). As we can write Μ‡π‘₯(𝑑)=π½βˆ’1(βˆ’πΉ(π‘₯)βˆ’Μƒπ‘›πΎ(π‘₯)), we have ⟨𝐽(Μ‡π‘₯(𝑑))βˆ’π½π½βˆ’1(βˆ’πΉ(π‘₯)),Μ‡π‘₯(𝑑)⟩=0.(3.11)Using the polarity between 𝑁𝐾(π‘₯(𝑑)) and 𝑇𝐾(π‘₯(𝑑)) and the variational principle (2.14) we deduce (3.3).

3.3. Existence of NpPDS

In this section we show that problem (3.3) has solutions and consequently that NpPDSs exist in the sense of Definition 3.5, by showing that problem (3.7) has solutions, in the sense of Definition 3.4. To obtain the main result of this paper, we need some preliminary ones, according to the following steps.(1)We first prove the existence of a sequence of approximate solutions with β€œgood” properties such that βˆ€π‘˜β‰₯π‘˜0,ξ€·π‘₯π‘˜(𝑑),Μ‡π‘₯π‘˜ξ€Έξ‚€π½(𝑑)∈graphβˆ’1ξ‚€ξ‚π‘βˆ’πΉβˆ’πΎξ‚ξ‚+β„³,(3.12) for any neighbourhood β„³ of 0 in 𝑋×𝑋. This step constitutes Theorem 3.9.(2)we prove next that the sequence obtained in the first step converges to a solution of problem (3.7) and that it has a weakly convergent subsequence whose derivative converges to Μ‡π‘₯(β‹…).

The methodology of the proofs is completely analogous to that used for pivot Hilbert spaces in [21]. Therefore we present the results with summary proofs, pointing out where they need to be updated for the case of a non-pivot H-space. The main difference in all proofs is made by the presence of the linear mapping 𝐽.

The main result can be stated as follows.

Theorem 3.7. Let 𝑋 be a Hilbert space and π‘‹βˆ— its topological dual and let πΎβŠ‚π‘‹ be a nonempty, closed and convex subset. Let πΉβˆΆπΎβ†’π‘‹βˆ— be a Lipschitz continuous vector field with Lipschitz constant 𝑏. Let π‘₯0∈𝐾. Then the initial value problem (3.3) has a unique solution on ℝ+.

Proof Existence of a solution on an interval [0,𝑙],𝑙<∞
For this part of the proof, we need two major results, as follows.
Proposition 3.8. Let 𝑋 be a nonpivot H-space, let π‘‹βˆ— be its topological dual, and let πΎβŠ‚π‘‹ be a non-empty, closed and convex subset. Let πΉβˆΆπΎβ†’π‘‹βˆ— be a Lipschitz continuous vector field with Lipschitz constant 𝑏, so that on πΎβˆ©π΅π‘‹(π‘₯0,𝐿), with 𝐿>0 and π‘₯0∈𝐾 arbitrarily fixed, we have ‖𝐹(π‘₯)β€–β‰€π‘€βˆΆ=‖𝐹(π‘₯0)β€–+𝑏𝐿.
Then the set-valued mapping π’©π‘βˆΆπΎβˆ©π΅π‘‹(π‘₯0,𝐿)→ℝ given by 𝑁π‘₯βŸΌπΉβˆ’πΎξ‚­(π‘₯),𝑝(3.13) has a closed graph.

Proof . The proof is similar to the one in [21].
We show first that the mapping π’©π‘βˆΆπΎβˆ©π΅π‘‹(π‘₯0,𝐿)→ℝ given by 𝑁π‘₯β†¦βŸ¨βˆ’πΎ(π‘₯),π‘βŸ© has a closed graph. It is clear that for each π‘βˆˆπ‘‹, the set-valued map π’©π‘βˆΆπΎβˆ©π΅π‘‹(π‘₯0,𝐿)→ℝ maps πΎβˆ©π΅π‘‹(π‘₯0,𝐿) into 2[βˆ’π‘€β€–π‘β€–,𝑀‖𝑝‖]. Let {(π‘₯𝑛,𝑧𝑛)}π‘›βˆˆgraph(𝒩𝑝) such that (π‘₯𝑛,𝑧𝑛)β†’(π‘₯,𝑧)βˆˆπ‘‹Γ—2[βˆ’π‘€β€–π‘β€–,𝑀‖𝑝‖]. We want to show that (π‘₯,𝑦)∈graph(𝒩𝑝). From π‘§π‘›βˆˆgraph(𝒩𝑝), for all 𝑛, we deduce that there exists π‘¦π‘›ξ‚π‘βˆˆβˆ’πΎ(π‘₯𝑛) such that 𝑧𝑛=βŸ¨π‘¦π‘›,π‘βŸ©. Since the set βˆ’ξ‚π‘πΎ(π‘₯)βŠ‚π΅π‘‹βˆ—(0,𝑀) and π΅π‘‹βˆ—(0,𝑀) is weakly compact, then there exists a subsequence π‘¦π‘›π‘˜ and π‘¦βˆˆπ‘‹βˆ— such that π‘¦π‘›π‘˜β‡€π‘¦(3.14) for the weak topology 𝜎(π‘‹βˆ—,π‘‹βˆ—βˆ—)byreflexivity=𝜎(π‘‹βˆ—,𝑋), which is equivalent to ξ«π‘¦π‘›π‘˜ξ¬,π›½βŸΆβŸ¨π‘¦,π›½βŸ©,βˆ€π›½βˆˆπ‘‹.(3.15) Suppose now that ξ‚π‘π‘¦βˆ‰βˆ’πΎ(π‘₯). This implies that at least one of the following two alternatives should be satisfied.(1)There exists π‘€βˆˆπΎ such that βŸ¨π‘¦,π‘€βˆ’π‘₯⟩<πœ†<0.(2)‖𝑦‖>πœ‡>‖𝐹(π‘₯)β€–. In the first case as βŸ¨π‘¦π‘›π‘˜,π›½βŸ©β†’βŸ¨π‘¦,π›½βŸ©,βˆ€π›½βˆˆπ‘‹ for π‘˜>π‘˜0 we have βŸ¨π‘¦π‘›π‘˜,π‘€βˆ’π‘₯⟩<πœ†/2. But βŸ¨π‘¦π‘›π‘˜,π‘€βˆ’π‘₯π‘›π‘˜βŸ©=βŸ¨π‘¦π‘›π‘˜,π‘€βˆ’π‘₯⟩+βŸ¨π‘¦π‘›π‘˜,π‘₯βˆ’π‘₯π‘›π‘˜βŸ© and as π‘₯π‘›π‘˜β†’π‘₯, there exists π‘˜1>0 such that βˆ€π‘˜β‰₯π‘˜1, we have βŸ¨π‘¦π‘›π‘˜,π‘₯βˆ’π‘₯π‘›π‘˜βŸ©β‰€β€–π‘₯βˆ’π‘₯π‘›π‘˜β€–β€–π‘¦π‘›π‘˜β€–<(|πœ†|/4𝑀)𝑀=|πœ†|/4. Thus βŸ¨π‘¦π‘›π‘˜,π‘€βˆ’π‘₯π‘›π‘˜βŸ©<πœ†/4<0, for all π‘˜>π‘šπ‘Žπ‘₯(π‘˜0,π‘˜1). But this contradicts the fact that π‘¦π‘›π‘˜ξ‚π‘βˆˆβˆ’πΎ(π‘₯π‘›π‘˜).
In the second case as βŸ¨π‘¦π‘›π‘˜,π›½βŸ©β†’βŸ¨π‘¦,π›½βŸ©,βˆ€π›½βˆˆπ‘‹, we have ([31, Proposition III.12]) ‖𝐹(π‘₯)β€–<‖𝑦‖≀liminfπ‘˜β†’βˆžβ€–π‘¦π‘›π‘˜β€– which is a contradiction because π‘¦π‘›ξ‚π‘βˆˆβˆ’πΎ(π‘₯𝑛),βˆ€π‘›βˆˆβ„•. The continuity of 𝐹 and the first part of the proof implies that 𝑁π‘₯βŸΌβŸ¨πΉβˆ’πΎ(π‘₯),π‘βŸ©(3.16) has non-empty, closed and convex values for each π‘₯∈𝐾 and has a closed graph.

The next result is constructing the sequence of approximate solutions for the problem (3.7).
Theorem 3.9. Let 𝑋 be a Hilbert space and π‘‹βˆ— its topological dual, and let πΎβŠ‚π‘‹ be a non-empty, closed and convex subset. Let πΉβˆΆπΎβ†’π‘‹βˆ— be a Lipschitz continuous vector field so that on πΎβˆ©π΅π‘‹(π‘₯0,𝐿), with 𝐿>0 and π‘₯0∈𝐾, we have ‖𝐹(π‘₯)β€–β‰€π‘€βˆΆ=‖𝐹(π‘₯0)β€–+𝑏𝐿. Let π‘™βˆΆ=𝐿/𝑀 and β„βˆΆ=[0,𝑙]. Then there exists a sequence {π‘₯π‘˜(β‹…)} of absolutely continuous functions defined on ℐ, with values in 𝐾, such that for all π‘˜β‰₯0,π‘₯π‘˜(0)=π‘₯0 and for almost all π‘‘βˆˆβ„, {π‘₯π‘˜(𝑑)} and {Μ‡π‘₯π‘˜(𝑑)} (the sequence of its derivatives) have the following property: for every neighbourhood β„³ of 0 in 𝑋×𝑋 there exists π‘˜0=π‘˜0(𝑑,β„³) such that βˆ€π‘˜β‰₯π‘˜0,ξ€·π‘₯π‘˜(𝑑),Μ‡π‘₯π‘˜ξ€Έξ‚€ξ‚π‘(𝑑)∈graphβˆ’πΉβˆ’πΎξ‚+β„³.(3.17)
Proof. The proof, based on topological properties of the space 𝑋, can be found in [21]. However, given we are now working in non-pivot H-spaces, then instead of π‘§π‘βˆΆ=𝑃𝐾(π‘₯βˆ’β„Žπ‘πΉ(π‘₯)) we now construct π‘§π‘βˆΆ=𝑃𝐾(π‘₯βˆ’β„Žπ‘π½βˆ’1∘𝐹(π‘₯)).
Next we show that the sequence {π‘₯π‘˜(β‹…)} built in Theorem 3.9 is uniformly convergent to some π‘₯(β‹…). Again, following closely [21], by Theorem 3.9 there exists a pair (π‘’π‘˜,βˆ’πΉ(π‘’π‘˜)βˆ’π‘›π‘˜ξ‚π‘)∈graph(βˆ’πΉβˆ’πΎ) such that π‘₯π‘˜(𝑑)βˆ’π‘’π‘˜(𝑑)=πœ–1,π‘˜(𝑑),Μ‡π‘₯π‘˜(𝑑)+π½βˆ’1ξ€·πΉξ€·π‘’π‘˜ξ€Έ(𝑑)+π‘›π‘˜ξ€Έ=πœ–2,π‘˜(𝑑),(3.18)where πœ–1,π‘˜(𝑑) and πœ–2,π‘˜(𝑑) are vector functions, not necessarily continuous, satisfying β€–πœ–1,π‘˜(𝑑)β€–<πœ–π‘˜ and β€–πœ–2,π‘˜(𝑑)β€–<πœ–π‘˜ where πœ–π‘˜β†’0 as π‘˜β†’βˆž and π‘›π‘˜βˆˆξ‚π‘πΎ(π‘’π‘˜) and π‘›π‘šβˆˆξ‚π‘πΎ(π‘’π‘š).
Let π‘˜,π‘š be two indexes. Then we evaluate 12𝑑‖‖π‘₯π‘‘π‘‘π‘˜(𝑑)βˆ’π‘₯π‘šβ€–β€–(𝑑)2=𝐽̇π‘₯π‘˜(𝑑)βˆ’Μ‡π‘₯π‘šξ€Έ(𝑑),π‘₯π‘˜(𝑑)βˆ’π‘₯π‘šξ¬=𝑒(𝑑)βˆ’πΉπ‘˜ξ€Έξ€·π‘₯(𝑑)+πΉπ‘˜ξ€Έξ€·π‘’(𝑑)+πΉπ‘šξ€Έξ€·π‘₯(𝑑)βˆ’πΉπ‘šξ€Έ(𝑑),π‘₯π‘˜(𝑑)βˆ’π‘₯π‘šξ¬+π‘₯(𝑑)βˆ’πΉπ‘˜ξ€Έξ€·π‘₯(𝑑)+πΉπ‘šξ€Έ(𝑑),π‘₯π‘˜(𝑑)βˆ’π‘₯π‘šξ¬(𝑑)+βŸ¨βˆ’π‘›π‘˜+π‘›π‘š,π‘’π‘˜(𝑑)βˆ’π‘’π‘š(𝑑)⟩+βŸ¨βˆ’π‘›π‘˜+π‘›π‘š,βˆ’π‘’π‘˜(𝑑)+π‘₯π‘˜(𝑑)+π‘’π‘š(𝑑)βˆ’π‘₯π‘š+ξ«π½ξ€·πœ–(𝑑)⟩1,π‘˜(𝑑)βˆ’πœ–2,π‘š(𝑑),π‘₯π‘˜(𝑑)βˆ’π‘₯π‘š(.𝑑)(3.19) But using the monotonicity of π‘₯↦𝑁𝐾(π‘₯), the isometry property of 𝐽, and the b-Lipschitz continuity of 𝐹 we get that 12𝑑‖‖π‘₯π‘‘π‘‘π‘˜(𝑑)βˆ’π‘₯π‘šβ€–β€–(𝑑)2β€–β€–π‘₯β‰€π‘π‘˜(𝑑)βˆ’π‘₯π‘šβ€–β€–(𝑑)2+ξ€·πœ–π‘˜+πœ–π‘šξ€Έβ€–β€–π‘›π‘˜βˆ’π‘›π‘šβ€–β€–ξ€·πœ–+(1+𝑏)π‘˜+πœ–π‘šξ€Έβ€–β€–π‘₯π‘˜(𝑑)βˆ’π‘₯π‘šβ€–β€–.(𝑑)(3.20)We now let πœ™(𝑑)∢=β€–π‘₯π‘˜(𝑑)βˆ’π‘₯π‘š(𝑑)β€–, so from the previous inequalities we get Μ‡πœ™(𝑑)πœ™(𝑑)β‰€π‘πœ™(𝑑)2+ξ€·πœ–π‘˜+πœ–π‘šξ€Έ[](1+𝑏)πœ™(𝑑)+2𝑀.(3.21) Using the same technique as in [21] we get πœ™(𝑑)2β‰€π‘Žπ‘ξ€·πœ–π‘˜+πœ–π‘šπ‘’ξ€Έξ€·2π‘π‘‘ξ€Έβ‰€π‘Žβˆ’1π‘ξ€·πœ–π‘˜+πœ–π‘šπ‘’ξ€Έξ€·2π‘π‘™ξ€Έβˆ’1,(3.22) where 𝑙 is the length of ℐ. So the Cauchy criteria are satisfied uniformly and we get the conclusion.
From the previous step we know that {π‘₯π‘˜(β‹…)} is uniformly convergent to π‘₯(β‹…) and as (π‘₯π‘˜(𝑑),Μ‡π‘₯π‘˜ξ‚π‘(𝑑))∈graph(βˆ’πΉβˆ’πΎ)+β„³, we now deduce that there exists a πœƒ such that β€–Μ‡π‘₯π‘˜(𝑑)β€–β‰€πœƒ. Using the arguments in [21] and the result of [32], we deduce the existence of a subsequence of {Μ‡π‘₯π‘˜}weakly*-convergent to Μ‡π‘₯(β‹…)∈𝐿∞(𝐼,𝑋).
Finally, we finish this part of the proof by showing that π‘₯(β‹…) is indeed a solution of the differential inclusion (3.7). From Theorem 3.9, for each π‘˜β‰₯π‘˜0 and almost every π‘‘βˆˆβ„ there exists a pair ξ€·π‘’π‘˜(𝑑),π‘£π‘˜ξ€Έξ‚€ξ‚π‘(𝑑)∈graphβˆ’πΉβˆ’πΎξ‚(3.23) such that β€–π‘₯π‘˜(𝑑)βˆ’π‘’π‘˜(𝑑)β€–<πœ–π‘˜ and β€–Μ‡π‘₯π‘˜(𝑑)βˆ’π‘£π‘˜(𝑑)β€–<πœ–π‘˜, where πœ–π‘˜β†’0 when π‘˜β†’βˆž. Let π‘βˆˆπ‘‹ arbitrarily fixed. Then for almost all π‘‘βˆˆβ„ξ€·π‘’π‘˜(𝑑),βŸ¨π‘£π‘˜ξ€Έξ‚π‘(𝑑),π‘βŸ©βˆˆgraphξ‚€ξ‚¬βˆ’πΉβˆ’πΎ,β€–β€–,π‘ξ‚­ξ‚βŸ¨Μ‡π‘₯π‘˜(𝑑),π‘βŸ©βˆ’βŸ¨π‘£π‘˜β€–β€–(𝑑),π‘βŸ©β‰€β€–π‘β€–πœ–π‘˜.(3.24) So π‘’π‘˜(𝑑)β†’π‘₯(𝑑) for every π‘‘βˆˆβ„ and βŸ¨π‘£π‘˜(𝑑),π‘βŸ©β†’βŸ¨Μ‡π‘₯π‘˜(𝑑),π‘βŸ© for almost all π‘‘βˆˆβ„. By Proposition 3.8, we know that 𝑁graph(βŸ¨βˆ’πΉβˆ’πΎ,π‘βŸ©) is closed, so it follows that for almost all π‘‘βˆˆβ„, ξ€·π‘₯(𝑑),βŸ¨Μ‡π‘₯π‘˜ξ€Έξ‚π‘(𝑑),π‘βŸ©βˆˆgraphξ‚€ξ‚¬βˆ’πΉβˆ’πΎ,𝑝.(3.25) Since the set 𝑁𝐹(π‘₯(𝑑))βˆ’πΎ(π‘₯(𝑑)) is convex and closed, it follows that Μ‡π‘₯(𝑑)βˆˆπ½βˆ’1ξ‚€ξ‚€ξ‚π‘βˆ’πΉπ‘₯(𝑑)βˆ’πΎξ‚ξ‚(π‘₯(𝑑)).(3.26) By Proposition 3.6, π‘₯(𝑑) is a solution of problem (3.3).
Uniqueness of Solutions on [0,𝑙]
Step 1 (π‘₯(β‹…) is the unique solution). Suppose that we have two solutions π‘₯1(β‹…) and π‘₯2(β‹…) starting at the same initial point. For any fixed π‘‘βˆˆβ„ we get 12𝑑‖‖π‘₯𝑑𝑑1(𝑑)βˆ’π‘₯1β€–β€–(𝑑)2=𝐽̇π‘₯1(𝑑)βˆ’Μ‡π‘₯2ξ€Έ(𝑑),π‘₯1(𝑑)βˆ’π‘₯2=𝐽(𝑑)Μ‡π‘₯1ξ€Έξ€·(𝑑)βˆ’π½Μ‡π‘₯2ξ€Έ(𝑑),π‘₯1(𝑑)βˆ’π‘₯2≀‖‖𝐽(𝑑)βˆ’1ξ€·ξ€·π‘₯βˆ’πΉ1ξ€Έ(𝑑)βˆ’π‘›1ξ€·π‘₯+𝐹2ξ€Έ(𝑑)+𝑛2ξ€Έ,π‘₯1(𝑑)βˆ’π‘₯2β€–β€–β€–β€–π‘₯(𝑑)>≀𝑏1(𝑑)βˆ’π‘₯2β€–β€–(𝑑)2,(3.27) because the metric projection is a nonexpansive operator in 𝑋, 𝐽 is a linear isometry, and 𝐹 is b-Lipschitz. By Gronwall’s inequality we obtain β€–π‘₯1(𝑑)βˆ’π‘₯2(𝑑)β€–2≀0, so we have π‘₯1(𝑑)=π‘₯2(𝑑) for any π‘‘βˆˆβ„.

Existence of Solutions on ℝ+
From above we can assert the existence of a solution to problem (3.3) on an interval [0;𝑙], with 𝑏>0 fixed and 𝐿>0 arbitrary. We note that we can choose 𝐿 such that 𝑙β‰₯1/(1+𝑏) in the following way: if ‖𝐹(π‘₯0)β€–=0, we let 𝐿=1, and if ‖𝐹(π‘₯0)β€–β‰ 0, then we let 𝐿β‰₯‖𝐹(π‘₯0)β€–. In both cases we obtain 𝑙β‰₯1/(1+𝑏). Therefore beginning at each initial point π‘₯0∈𝐾, problem (3.3) has a solution on an interval of length at least [0;1/(1+𝑏)]. Now if we consider problem (3.3) with π‘₯0=π‘₯(1/(1+𝑏)), applying again all the above, we obtain an extension of the solution on an interval of length at least 1/(1+𝑏). By continuing this solution we obtain a solution on [0,∞).

3.4. Implicit PDS

In this section we consider a generic Hilbert space 𝑋, where generic is taken to mean that the dimensionality could be either finite or infinite, and the space could be either a pivot or a non-pivot space. Let us introduce the following definition.

Definition 3.10. Let 𝑋 be a generic H-space and let πΎξ…žβŠ‚π‘‹ be a non-empty, closed subset. Consider a pair (𝑔,𝐾) such that 𝐾 is convex and π‘”βˆΆπΎξ…žβ†’πΎ=π‘Ÿ(πΎξ…ž)βŠ‚π‘‹, is continuous, injective, and π‘”βˆ’1 is Lipschitz continuous.
Consider πΉβˆΆπ‘‹β†’π‘‹βˆ— satisfying (πΉβˆ˜π‘”)(𝑦)=𝐹(𝑦),βˆ€π‘¦βˆˆπΎξ…ž. Then the pair (𝑔,𝐾) is called a convexification pair of (𝐹,πΎξ…ž).

Example 3.11. Here is an example of such a convexification pair in ℝ2. Let πΎξ…ž={(π‘₯,𝑦)βˆˆβ„2∣0≀π‘₯≀1,0≀𝑦≀π‘₯} and let 𝑔 be the map of πΎξ…ž into 𝐾=[0,1]Γ—[0,1], namely: ξ‚€2𝑔(π‘₯,𝑦)=π‘₯,1+π‘₯𝑦+1βˆ’π‘₯.1+π‘₯(3.28) We can easily check that 𝑔 is continuous and monotone. Now take 𝐹 to be 𝐹(π‘₯,𝑦)=(π‘₯,π‘Ž), where π‘Ž is an arbitrary constant in ℝ. Then we have πΉβˆ˜π‘”(π‘₯,𝑦)=(π‘₯,π‘Ž)=𝐹(π‘₯,𝑦).

We now introduce another type of a projected equation as follows.

Definition 3.12. Let 𝑋 be a generic H-space and let πΎξ…žβŠ‚π‘‹ be a non-empty, closed subset. An implicit projected differential equation (ImPrDE) is a (PrDE) given by (3.2) where π‘₯(𝑑)∢=𝑔(𝑦(𝑑)),π‘”βˆΆπΎξ…žβ†’πΎβŠ‚π‘‹, that is: 𝑑𝑔(𝑦(𝑑))𝑑𝑑=𝑃𝑇𝐾(𝑔(𝑦(𝑑)))ξ€·βˆ’π½βˆ’1ξ€Έ.βˆ˜πΉβˆ˜π‘”(𝑦(𝑑))(3.29)

The motivation for the introduction of such an equation comes from the desire to study the dynamics on a set πΎξ…žβŠ‚π‘‹, where πΎξ…ž could be nonconvex, and to study as well some dynamic problems on a so-called translated set (see Section 4 below).

Considering now (3.29) and a convexification pair (𝑔,𝐾) of a nonempty, closed πΎξ…žβŠ‚π‘‹, then the Cauchy problem associated to (3.29) and the pair (𝑔,𝐾) is given by𝑑𝑔(𝑦(𝑑))𝑑𝑑=πœ‹πΎξ€·ξ€·π½π‘”(𝑦(𝑑)),βˆ’βˆ’1∘𝐹|𝐾′(𝑦(𝑑)),𝑔(𝑦(0))=π‘₯0∈𝐾.(3.30) Next we define what we mean by a solution for a Cauchy problem of type (3.30).

Definition 3.13. An absolutely continuous function π‘¦βˆΆβ„βŠ‚β„β†’π‘‹, such that 𝑦(𝑑)βˆˆπΎξ…ž,𝑔(𝑦(0))=π‘₯0∈𝐾,βˆ€π‘‘βˆˆβ„,𝑑𝑔(𝑦(𝑑))𝑑𝑑=πœ‹πΎξ€·ξ€·π½π‘”(𝑦(𝑑)),βˆ’βˆ’1∘𝐹|𝐾′(𝑦(𝑑)),a.e.onℐ(3.31) is called a solution for the initial value problem (3.30).

We claim that problem (3.30) has solutions by Theorem 3.9. It is obvious that by a change of variable π‘₯(β‹…)∢=𝑔(𝑦(β‹…)), problem (3.30) has solutions on 𝐾, in the sense of Definition 3.4. But since 𝑔 is assumed continuous and strictly monotone, then 𝑔 is invertible and so 𝑦(β‹…)=π‘”βˆ’1(π‘₯(β‹…)); moreover, we see that such a 𝑦 is a solution to problem (3.30) in the above sense.

Now we are ready to introduce the following.

Definition 3.14. An implicit projected dynamical system (ImPDS) is given by a mapping πœ™βˆΆβ„+Γ—πΎξ…žβ†’πΎ which solves the initial value problem: Μ‡πœ™(𝑑,𝑔(𝑦(𝑑)))=πœ‹πΎξ€·ξ€·π½πœ™(𝑑,𝑔(𝑦(𝑑))),βˆ’βˆ’1βˆ˜πΉξ€Έξ€Έ(πœ™(𝑑,𝑦(𝑑))),a.a.𝑑,πœ™(0,𝑔(𝑦(0)))=π‘₯0∈𝐾,(3.32) where (𝑔,𝐾) is a convexification pair.

Theorem 3.15. Let 𝑋 be a generic Hilbert space, and let πΎξ…ž be a non-empty closed subset of 𝑋. Let 𝐾 be non-empty, closed and convex, let π‘”βˆΆπΎξ…žβ†’πΎ be continuous and strictly monotone, and let πΉβˆΆπΎξ…žβ†’π‘‹ be Lipschitz continuous such that (πΉβˆ˜π‘”)|𝐾′=𝐹. Let also π‘₯0∈𝐾 and 𝐿>0 such that β€–π‘₯0‖≀𝐿. Then the initial value problem (3.30) has a unique solution on the interval [0,𝑙], where 𝑙=𝐿/(‖𝐹(π‘₯0)β€–+𝑏𝐿).

Proof. The proof consists in the modification of a few easy steps of the proof given in [21] combined with the results of the present paper.

4. Applications

4.1. NpPDS, ImPDS, and Variational Inequalities

It is worth noting at this point that, as in the pivot case, a NpPDS is also related to a variational inequality (VI) problem. To show this relation, we first define what is meant by a critical point of NpPDS.

Definition 4.1. A point π‘₯βˆ—βˆˆπΎ is called a critical point for (3.2) if πœ‹πΎξ€·π‘₯βˆ—ξ€·π½,βˆ’βˆ’1π‘₯βˆ˜πΉξ€Έξ€·βˆ—ξ€Έξ€Έ=0.(4.1)

Theorem 4.2. Let X be a generic Hilbert space and let πΎβŠ‚π‘‹ be a non-empty, closed and convex subset. Let πΉβˆΆπ‘‹β†’π‘‹βˆ— be a vector field. Consider the variational inequality problem: π‘₯∈𝐾∢⟨𝐹(π‘₯),π‘£βˆ’π‘₯⟩β‰₯0,βˆ€π‘£βˆˆπΎ.(4.2) Then the solution set of (4.2) coincides with the set of critical points of the non-pivot projected dynamical system (3.2).

Proof. It follows from the decomposition Theorem 2.8 (see also [23]).

The relation between an ImPDS and a VI problem is more interesting, as has been considered before in the literature, but with superfluous conditions on the projection operator 𝑃𝐾 we describe this relation next.

Definition 4.3. Let 𝑋 be a generic H-space and let πΎξ…žβŠ‚π‘‹ be a non-empty, closed subset. Let πΉβˆΆπ‘‹β†’π‘‹βˆ— be a mapping. Then we call g-variational inequality on the set 𝐾′ the problem of findingπ‘¦βˆˆπΎξ…ž,βŸ¨πΉβˆ˜π‘”(𝑦),π‘§βˆ’π‘”(𝑦)⟩β‰₯0,βˆ€π‘§βˆˆπΎ,(4.3) where (𝑔,𝐾) is a convexification pair of (𝐹,πΎξ…ž).

We highlight the importance of the relation πΉβˆ˜π‘”(𝑦)=𝐹(𝑦) from Definition 3.10 in order for (4.3) to make sense. Under (3.5) we can rewrite (4.3) asfindπ‘¦βˆˆπΎβ€²,⟨𝐹(𝑦),π‘§βˆ’π‘”(𝑦)⟩β‰₯0,βˆ€π‘§βˆˆπΎ.(4.4)

Remark 4.4. In [24], (4.4) is considered in a pivot H-space and is called a β€œgeneral variational inequality.” We prefer to use the term β€œg-variational inequality” in relation to (4.4), in order to avoid confusion with the commonly accepted β€œgeneralized variational inequality” which involves multimappings.

Theorem 4.5. If the problems (4.4) and (3.30) admit a solution, then the equilibrium points of (4.4) coincide with the critical points of (3.30).

Proof. Suppose π‘¦βˆ—βˆˆπΎξ…ž is a solution of (4.4); then by definition we have ξ«πΉξ€·π‘¦βˆ—ξ€Έξ€·π‘¦,π‘§βˆ’π‘”βˆ—ξ€Έξ¬β‰₯0,βˆ€π‘§βˆˆπΎ.(4.5) So by multiplying by a strictly positive constant πœ† and using the bilinearity of the inner product, we get ξ«ξ€·π‘¦βˆ’πΉβˆ—ξ€Έξ¬,𝑦≀0,βˆ€π‘¦βˆˆπ‘‡πΎξ€·π‘”ξ€·π‘¦βˆ—ξ€Έξ€Έ.(4.6) So we deduce that βˆ’πΉ(π‘¦βˆ—)βˆˆπ‘πΎ(𝑔(π‘¦βˆ—)); using the decomposition Theorem 2.8 we get 𝑃𝑇𝐾(𝑔(π‘¦βˆ—))(βˆ’π½βˆ’1𝐹(π‘¦βˆ—))=0, and so π‘¦βˆ— is a critical point of (3.30).
Now suppose that π‘¦βˆ— is a critical point of (3.30); then by definition we have π‘ƒπ‘‡βˆ—πΎ(𝑔(𝑦))ξ€·βˆ’π½βˆ’1πΉξ€·π‘¦βˆ—ξ€Έξ€Έ=0,(4.7) and by the decomposition theorem we get βˆ’πΉ(π‘¦βˆ—)βˆˆπ‘πΎ(𝑔(π‘¦βˆ—)). By the definition of the normal cone to 𝐾 in 𝑔(π‘¦βˆ—), the following inequality is satisfied: ξ«ξ€·π‘¦βˆ’πΉβˆ—ξ€Έξ€·π‘¦,π‘§βˆ’π‘”βˆ—ξ€Έξ¬β‰€0,βˆ€π‘§βˆˆπΎ,(4.8) which is exactly (4.4).

5. Examples and Applications

5.1. Weighted Traffic Problem

Let us introduce a network 𝒩, that means a set 𝒲 of origin-destination pair (origin/destination node) and a set β„› of routes. Each route π‘Ÿβˆˆβ„› links exactly one origin-destination pair π‘€βˆˆπ’². The set of all π‘Ÿβˆˆβ„› which link a given π‘€βˆˆπ’² is denoted by β„›(𝑀). For each time π‘‘βˆˆ(0,𝑇) we consider vector flow 𝐹(𝑑)βˆˆβ„π‘›. Let us denote by Ξ© an open subset of ℝ, by 𝑛=card(β„›), 𝐚={π‘Ž1,…,π‘Žπ‘›}, and by πšβˆ’1={π‘Ž1βˆ’1,…,π‘Žπ‘›βˆ’1} two families of weights such that for each 1≀𝑖≀𝑛,π‘Žπ‘–βˆˆπ’ž(Ξ©,ℝ+⧡{0}). We introduce also the family of real time traffic densities 𝐬={𝑠1,…,𝑠𝑛} such that for each 1≀𝑖≀𝑛, π‘ π‘–βˆˆπ’ž(Ξ©,ℝ+⧡{0}).

Let π‘Ÿπ‘– correspond to an element of π‘Ž and 𝑠, newly to π‘Žπ‘– and 𝑠𝑖. If we denote by 𝑉𝑖=𝐿2(Ξ©,ℝ,π‘Žπ‘–,𝑠𝑖) and π‘‰βˆ—π‘–=𝐿2(Ξ©,ℝ,π‘Žπ‘–βˆ’1,𝑠𝑖), the space𝑉=𝑛𝑖=1𝑉𝑖(5.1) is a Hilbert space for the inner product⟨𝐹,𝐺⟩𝐚,𝐬=𝑛𝑖=1ξ€œΞ©πΉπ‘–(πœ”)𝐺𝑖(πœ”)π‘Žπ‘–(πœ”)𝑠𝑖(πœ”)π‘‘πœ”.(5.2) The space π‘‰βˆ—=βˆπ‘›π‘–=1π‘‰βˆ—π‘– is a Hilbert space for the following inner product⟨𝐹,πΊβŸ©πšβˆ’1,𝐬=𝑛𝑖=1ξ€œΞ©πΉπ‘–(πœ”)𝐺𝑖(πœ”)𝑠𝑖(πœ”)π‘Žπ‘–(πœ”)π‘‘πœ”,(5.3) and the following bilinear form defines a duality between 𝑉 and π‘‰βˆ—:π‘‰βˆ—Γ—π‘‰βŸΆβ„,(5.4)βŸ¨π‘“,π‘₯⟩𝐬=𝑛𝑖=1ξ€œΞ©π‘“π‘–(πœ”)π‘₯𝑖(πœ”)𝑠𝑖(πœ”)π‘‘πœ”.(5.5) More exactly we have the following.

Proposition 5.1. The bilinear form (5.5) is defined over π‘‰βˆ—Γ—π‘‰ and defines a duality between π‘‰βˆ—Γ—π‘‰. The duality mapping is given by 𝐽(𝐹)=(π‘Ž1𝐹1,…,π‘Žπ‘›πΉπ‘›).

The feasible flows have to satisfy the time-dependent capacity constraints and demand requirements; namely, for all π‘Ÿβˆˆβ„›, π‘€βˆˆπ’² and for almost all π‘‘βˆˆΞ©,πœ†π‘Ÿ(𝑑)β‰€πΉπ‘Ÿ(𝑑)β‰€πœ‡π‘Ÿξ“(𝑑),π‘Ÿβˆˆβ„›(𝑀)πΉπ‘Ÿ(𝑑)=πœŒπ‘€(𝑑),(5.6) where 0β‰€πœ†β‰€πœ‡ are given in 𝐿2([0,𝑇],ℝ𝑛), 𝜌∈𝐿2([0,𝑇],β„π‘š) where π‘š=card(𝒲),πΉπ‘Ÿ,π‘Ÿβˆˆβ„›, denotes the flow in the route π‘Ÿ. If Ξ¦=(Φ𝑀,π‘Ÿ) is the pair route incidence matrix, with π‘€βˆˆπ’² and π‘Ÿβˆˆβ„›, that is,Φ𝑀,π‘ŸβˆΆ=πœ’β„›(𝑀)(π‘Ÿ),(5.7) the demand requirements can be written in matrix-vector notation asΦ𝐹(𝑑)=𝜌(𝑑).(5.8) The set of all feasible flows is given by𝐾∢={πΉβˆˆπ‘‰βˆ£πœ†(𝑑)≀𝐹(𝑑)β‰€πœ‡(𝑑),a.e.inΞ©;Φ𝐹(𝑑)=𝜌(𝑑),a.einΞ©}.(5.9) We provide now the definition of equilibrium for the traffic problem. First we need to define the notion of equilibrium for a variational inequality. A variational inequality (VI) in a Hilbert space 𝑉 is to determineπ‘₯∈𝐾∢⟨𝐢(π‘₯),π‘¦βˆ’π‘₯⟩𝐬β‰₯0,βˆ€π‘¦βˆˆπΎ,(5.10) where 𝐾 is a closed convex subset of 𝑉, and πΆβˆΆπΎβ†’π‘‰βˆ— is a mapping.

Definition 5.2. π»βˆˆπ‘‰ is an equilibrium flow if and only if 𝐻∈𝐾,⟨𝐢(𝐻),πΉβˆ’π»βŸ©π¬β‰₯0,βˆ€πΉβˆˆπΎ.(5.11)

It is possible to prove the equivalence between condition (5.11) and what we will call a weighted Wardrop condition (5.13).

Theorem 5.3. 𝐻∈𝐾 is an equilibrium flow in the sense of (5.11) if and only if π‘ βˆ€π‘€βˆˆπ’²,βˆ€π‘ž,π‘šβˆˆβ„›(𝑀),a.e.inΞ©,(5.12)π‘ž(𝑑)πΆπ‘ž(𝐻(𝑑))<π‘ π‘š(𝑑)πΆπ‘š(𝐻(𝑑)),βŸΉπ»π‘ž(𝑑)=πœ‡π‘ž(𝑑)orπ»π‘š(𝑑)=πœ†π‘š(𝑑).(5.13)

Proof. see [20].

Based on previous results [20], this solution coincides with the set of critical points of the associated projected dynamical system.

5.2. Quasivariational Inequalities on Translated Sets
5.2.1. QVI

Let 𝑋 be a generic H-space, 𝐷 closed, convex, nonempty in 𝑋. Let π’¦βˆΆπ·β†’2𝑋 with 𝒦(π‘₯) convex for all π‘₯∈𝐷 and πΉβˆΆπ’¦β†’2π‘‹βˆ— a mapping.

Let us introduce the following variational inequality:findπ‘₯βˆˆπ’¦(π‘₯),⟨𝐹(π‘₯),π‘¦βˆ’π‘₯⟩β‰₯0,βˆ€π‘¦βˆˆπ’¦(π‘₯).(5.14) Note that in this case the set in which we are looking for the solution depends on π‘₯. For problem (5.14) we can provide the following existence result (see [17] or [33]).

Theorem 5.4. Let 𝐷 be a closed convex subset in a locally convex Hausdorff topological vector space 𝑋. Let us suppose that(i)π’¦βˆΆπ·β†’2𝐷 is a closed lower semicontinuous correspondence with closed, convex, and nonempty values,(ii)πΆβˆΆπ·β†’2𝑋′ is a monotone, finite continuous, and bounded single-valued map,(iii)there exist a compact, convex, and nonempty set π‘βŠ‚π· and a nonempty subset π΅βŠ‚π‘ such that(a)𝒦(𝐡)βŠ‚π‘,(b)𝒦(𝑧)βˆ©π‘β‰ βˆ…,βˆ€π‘§βˆˆπ‘, (c)for every π‘§βˆˆπ‘β§΅π΅, there exist Μ‚π‘§βˆˆπ’¦(𝑧)βˆ©π‘ with ⟨𝐢(𝑧),Μ‚π‘§βˆ’π‘§βŸ©<0.
Then there exists π‘₯ such that π‘₯βˆˆπ’¦(π‘₯)∢⟨𝐢(π‘₯),π‘¦βˆ’π‘₯⟩β‰₯0,βˆ€π‘¦βˆˆπ’¦(π‘₯).(5.15)

In order to study the disequilibrium behavior of (5.14), we introduce now the following projected differential equation.

Definition 5.5. We call projected dynamical system associated to the quasivariational inequality (5.14) the solution set of the projected differential equation: 𝑑π‘₯(𝑑)𝑑𝑑=𝑃𝑇𝒦(π‘₯)(π‘₯)ξ€·βˆ’π½βˆ’1𝐹(π‘₯),π‘₯(0)=π‘₯0βˆˆπ’¦(π‘₯).(5.16)

Remark 5.6. In general there are no existence results for problem (5.16). An existence result for a particular case of (5.16) has been given in [24], assuming the following fact.
Assumption 5.7. Let 𝑋 be a pivot H-space. For all 𝑒,𝑣,π‘€βˆˆπ‘‹, 𝑃𝒦(𝑒) satisfies the condition‖‖𝑃𝒦(𝑒)(𝑀)βˆ’π‘ƒπ’¦(𝑣)β€–β€–(𝑀)β‰€πœ†β€–π‘’βˆ’π‘£β€–,(5.17) where πœ†>0 is a constant.
However, this assumption fails to be true. One counterexample is as follows. We denote by 𝐢 a closed convex set and we take 𝑒,π‘£βˆˆπΆ; we denote by 𝒦(𝑒)=𝑇𝐢(𝑒) and by 𝒦(𝑣)=𝑇𝐢(𝑣) the tangent cones of 𝐢 at 𝑒 and 𝑣.
In fact, π‘€βˆˆπ‘‹ can only be chosen in one of the following four situations:(1)π‘€βˆˆπ’¦(𝑒)βˆ©π’¦(𝑣), (2)π‘€βˆˆπ’¦(𝑒)⧡𝒦(𝑣), (3)π‘€βˆˆπ’¦(𝑣)⧡𝒦(𝑒), (4)π‘€βˆˆπ‘‹β§΅(𝒦(𝑒)βˆͺ𝒦(𝑣)). Suppose now that we have π‘€βˆˆπ’¦(𝑒)⧡𝒦(𝑣); then by Moreau’s decomposition theorem we get ‖‖𝑃𝒦(𝑒)(𝑀)βˆ’π‘ƒπ’¦(𝑣)β€–β€–=β€–β€–(𝑀)π‘€βˆ’π‘ƒπ’¦(𝑣)β€–β€–=‖‖𝑃(𝑀)𝑁𝐢(𝑣)β€–β€–(𝑀)β‰€πœ†β€–π‘’βˆ’π‘£β€–,(5.18) where 𝑁𝐢(𝑣) is the normal cone of 𝐢 at 𝑣. Consider now 𝑋=ℝ2, 𝐢=[0,πœ–]2, 𝑒=(0,0) and 𝑣=(πœ–,πœ–). It is clear that we have the following: 𝑇𝐢(𝑒)=ℝ2+,𝑇𝐢(𝑣)=ℝ2βˆ’,𝑁𝐢(𝑣)=ℝ2+=𝑇𝐢(𝑒).(5.19) So for any π‘€βˆˆπ‘πΆ(𝑣) we get βˆšβ€–π‘€β€–β‰€πœ†β€–π‘’βˆ’π‘£β€–=2πœ–πœ†.(5.20) Since 𝑀 is arbitrary, let now π‘€βˆΆ=πœ‡π‘€, for any πœ‡>0. Then, βˆšβ€–πœ‡π‘€β€–β‰€πœ†β€–π‘’βˆ’π‘£β€–=2πœ–πœ†(5.21) should be true for any πœ‡>0. However this does not hold.

Consider now the special case of a set-valued mapping 𝒦 which is the translation of a closed, convex subset 𝐾:π’¦βˆΆπ‘₯⟢𝐾+𝑣(π‘₯),(5.22)

where 𝑣(π‘₯) is a vector linearly dependant on π‘₯; then problems (5.14) and (5.16) can be studied, under certain conditions, respectively, as a g-VI and an implicit PDS as shown below.

If 𝒦(π‘₯)=𝐾+𝑝(π‘₯) as done by Noor for type B PDS [24], we have the following equivalent formulations:𝑑π‘₯(𝑑)𝑑𝑑=𝑃𝑇𝐾+𝑝(π‘₯)(π‘₯)ξ€·βˆ’π½βˆ’1𝐹(π‘₯)=𝑃𝑇𝐾(𝑔(π‘₯))ξ€·βˆ’π½βˆ’1𝐹(π‘₯),π‘₯(0)=π‘₯0∈𝐾,(5.23) where 𝑔(π‘₯)=π‘₯βˆ’π‘(π‘₯), assuming 𝐹(𝑔(π‘₯))=𝐹(π‘₯βˆ’π‘(π‘₯))=𝐹(π‘₯). We can observe that if (𝑑𝑝(π‘₯))/𝑑𝑑=0, then (5.23) is equal to the implicit projected differential equation (3.29), and therefore Theorem 3.15 provides an existence result without assuming any kind of Lipschitz condition of the projection operator.

6. Conclusions

We show in this paper that previous results of existence of projected dynamical systems can be generalized to two new classes, namely, the non-pivot and the implicit PDS. The generalizations came as needed to study a more realistic traffic equilibrium problem, as well as to study the relations between an implicit PDS and a class of variational inequalities as previously introduced in [24] as an open problem.

Acknowledgments

The work has been supported by the first author’s NSERC Discovery Grant. The support is gratefully acknowledged.