rovnice v metrických prostorech

Similar documents
Dedicated to Prof. Jaroslav Kurzweil on the occasion of his 80th birthday

convergence theorem in abstract set up. Our proof produces a positive integrable function required unlike other known

Course 212: Academic Year Section 1: Metric Spaces

THEOREMS, ETC., FOR MATH 515

PERIODIC PROBLEMS WITH φ-laplacian INVOLVING NON-ORDERED LOWER AND UPPER FUNCTIONS

van Rooij, Schikhof: A Second Course on Real Functions

Continuous Functions on Metric Spaces

Metric Spaces and Topology

converges as well if x < 1. 1 x n x n 1 1 = 2 a nx n

Measures. Chapter Some prerequisites. 1.2 Introduction

The Heine-Borel and Arzela-Ascoli Theorems

Real Analysis Problems

Singularities and Laplacians in Boundary Value Problems for Nonlinear Ordinary Differential Equations

McGill University Math 354: Honors Analysis 3

Lecture Notes in Advanced Calculus 1 (80315) Raz Kupferman Institute of Mathematics The Hebrew University

Existence and Uniqueness

Analysis Finite and Infinite Sets The Real Numbers The Cantor Set

Introduction and Preliminaries

Functional Analysis. Franck Sueur Metric spaces Definitions Completeness Compactness Separability...

7 Complete metric spaces and function spaces

n [ F (b j ) F (a j ) ], n j=1(a j, b j ] E (4.1)

Math 209B Homework 2

1/12/05: sec 3.1 and my article: How good is the Lebesgue measure?, Math. Intelligencer 11(2) (1989),

3 (Due ). Let A X consist of points (x, y) such that either x or y is a rational number. Is A measurable? What is its Lebesgue measure?

FUNCTIONAL ANALYSIS LECTURE NOTES: COMPACT SETS AND FINITE-DIMENSIONAL SPACES. 1. Compact Sets

4th Preparation Sheet - Solutions

Applied Analysis (APPM 5440): Final exam 1:30pm 4:00pm, Dec. 14, Closed books.

2. Function spaces and approximation

Final. due May 8, 2012

MA651 Topology. Lecture 10. Metric Spaces.

MATH 202B - Problem Set 5

Math 118B Solutions. Charles Martin. March 6, d i (x i, y i ) + d i (y i, z i ) = d(x, y) + d(y, z). i=1

Introduction to Real Analysis Alternative Chapter 1

Recall that if X is a compact metric space, C(X), the space of continuous (real-valued) functions on X, is a Banach space with the norm

Metric Spaces Lecture 17

Tools from Lebesgue integration

Economics 204 Fall 2011 Problem Set 2 Suggested Solutions

SYMMETRIC INTEGRALS DO NOT HAVE THE MARCINKIEWICZ PROPERTY

for all subintervals I J. If the same is true for the dyadic subintervals I D J only, we will write ϕ BMO d (J). In fact, the following is true

INTRODUCTION TO MEASURE THEORY AND LEBESGUE INTEGRATION

The Arzelà-Ascoli Theorem

REAL ANALYSIS LECTURE NOTES: 1.4 OUTER MEASURE

Some Background Material

MATHS 730 FC Lecture Notes March 5, Introduction

Geometry and topology of continuous best and near best approximations

d(x n, x) d(x n, x nk ) + d(x nk, x) where we chose any fixed k > N

s P = f(ξ n )(x i x i 1 ). i=1

MAT 570 REAL ANALYSIS LECTURE NOTES. Contents. 1. Sets Functions Countability Axiom of choice Equivalence relations 9

Lebesgue-Stieltjes measures and the play operator

Analysis III. Exam 1

From now on, we will represent a metric space with (X, d). Here are some examples: i=1 (x i y i ) p ) 1 p, p 1.

Lebesgue Integration: A non-rigorous introduction. What is wrong with Riemann integration?

A CAUCHY PROBLEM ON TIME SCALES WITH APPLICATIONS

Locally Lipschitzian Guiding Function Method for ODEs.

Math 321 Final Examination April 1995 Notation used in this exam: N. (1) S N (f,x) = f(t)e int dt e inx.

ON PERIODIC BOUNDARY VALUE PROBLEMS OF FIRST-ORDER PERTURBED IMPULSIVE DIFFERENTIAL INCLUSIONS

Functional Differential Equations with Causal Operators

Measure and Integration: Solutions of CW2

P-adic Functions - Part 1

Optimization and Optimal Control in Banach Spaces

FEEDBACK DIFFERENTIAL SYSTEMS: APPROXIMATE AND LIMITING TRAJECTORIES

Existence Theorem for Abstract Measure. Differential Equations Involving. the Distributional Henstock-Kurzweil Integral

Lecture 19 L 2 -Stochastic integration

2 Statement of the problem and assumptions

On intermediate value theorem in ordered Banach spaces for noncompact and discontinuous mappings

Austin Mohr Math 730 Homework. f(x) = y for some x λ Λ

arxiv: v1 [math.ca] 7 Jul 2013

LINEAR SYSTEMS OF GENERALIZED ORDINARY DIFFERENTIAL EQUATIONS

LECTURE 15: COMPLETENESS AND CONVEXITY

POINTWISE PRODUCTS OF UNIFORMLY CONTINUOUS FUNCTIONS

I. The space C(K) Let K be a compact metric space, with metric d K. Let B(K) be the space of real valued bounded functions on K with the sup-norm

arxiv: v1 [math.fa] 18 Feb 2016

Topology. Xiaolong Han. Department of Mathematics, California State University, Northridge, CA 91330, USA address:

Chapter 2. Metric Spaces. 2.1 Metric Spaces

Fragmentability and σ-fragmentability

Math 117: Continuity of Functions

EXISTENCE OF SOLUTIONS TO A BOUNDARY-VALUE PROBLEM FOR AN INFINITE SYSTEM OF DIFFERENTIAL EQUATIONS

HOMEWORK ASSIGNMENT 6

Maths 212: Homework Solutions

Math 341: Convex Geometry. Xi Chen

Measures. 1 Introduction. These preliminary lecture notes are partly based on textbooks by Athreya and Lahiri, Capinski and Kopp, and Folland.

Bootcamp. Christoph Thiele. Summer As in the case of separability we have the following two observations: Lemma 1 Finite sets are compact.

An introduction to Mathematical Theory of Control

Assignment-10. (Due 11/21) Solution: Any continuous function on a compact set is uniformly continuous.

Chapter 2 Convex Analysis

Introduction to Topology

AW -Convergence and Well-Posedness of Non Convex Functions

MA5206 Homework 4. Group 4. April 26, ϕ 1 = 1, ϕ n (x) = 1 n 2 ϕ 1(n 2 x). = 1 and h n C 0. For any ξ ( 1 n, 2 n 2 ), n 3, h n (t) ξ t dt

Bounded Derivatives Which Are Not Riemann Integrable. Elliot M. Granath. A thesis submitted in partial fulfillment of the requirements

EXPOSITORY NOTES ON DISTRIBUTION THEORY, FALL 2018

The small ball property in Banach spaces (quantitative results)

EXISTENCE RESULTS FOR NONLINEAR FUNCTIONAL INTEGRAL EQUATIONS VIA NONLINEAR ALTERNATIVE OF LERAY-SCHAUDER TYPE

Variational Stability for Kurzweil Equations associated with Quantum Stochastic Differential Equations}

Economics 204 Summer/Fall 2011 Lecture 5 Friday July 29, 2011

(convex combination!). Use convexity of f and multiply by the common denominator to get. Interchanging the role of x and y, we obtain that f is ( 2M ε

Chapter 1. Measure Spaces. 1.1 Algebras and σ algebras of sets Notation and preliminaries

Continuous functions that are nowhere differentiable

Chapter 2 Metric Spaces

Lecture 10. Riemann-Stieltjes Integration

Some Applications to Lebesgue Points in Variable Exponent Lebesgue Spaces

Transcription:

Univerzita Karlova v Praze Matematicko-fyzikální fakulta DIPLOMOVÁ PRÁCE Břetislav Skovajsa Zobecněné obyčejné diferenciální rovnice v metrických prostorech Katedra matematické analýzy Vedoucí diplomové práce: Studijní program: Studijní obor: prof. RNDr. Jan Malý DrSc. Matematika Matematická analýza Praha 2014

I would like to thank my advisor, prof. RNDr. Jan Malý DrSc., for everything he has done for me. Especially for being strict when necessary, yet leaving me enough creative freedom, for always pointing me in the right direction and always giving me the right perspective to look at a given problem.

Prohlašuji, že jsem tuto diplomovou práci vypracoval samostatně a výhradně s použitím citovaných pramenů, literatury a dalších odborných zdrojů. Beru na vědomí, že se na moji práci vztahují práva a povinnosti vyplývající ze zákona č. 121/2000 Sb., autorského zákona v platném znění, zejména skutečnost, že Univerzita Karlova v Praze má právo na uzavření licenční smlouvy o užití této práce jako školního díla podle 60 odst. 1 autorského zákona. V... dne... podpis

Název práce: Zobecněné obyčejné diferenciální rovnice v metrických prostorech Autor: Břetislav Skovajsa Katedra: Katedra matematické analýzy Vedoucí diplomové práce: prof. RNDr. Jan Malý DrSc. Abstrakt: Cílem této práce je vybudovat základy zobecněných obyčejných diferenciálních rovnic v metrických prostorech. Diferenciální rovnice v metrických prostorech již byly studovány dříve, avšak zvolený přístup není schopen zahrnout obecnější druhy integrálních rovnic. Práce nabízí definici, která kombinuje obecnost metrických prostorů se silou Kurzweilových zobecněných obyčejných diferenciálních rovnic. Dále prezentujeme věty o jednoznačnosti a existenci, které poskytují nové výsledky i v kontextu euklidovských prostorů. Klíčová slova: Zobecněné obyčejné diferenciální rovnice, metrické prostory, křivky, Henstock-Kurzweilův integrál. Title: Generalized ordinary differential equations in metric spaces Author: Břetislav Skovajsa Department: Department of mathematical analysis Supervisor: prof. RNDr. Jan Malý DrSc. Abstract: The aim of this thesis is to build the foundations of generalized ordinary differential equation theory in metric spaces. While differential equations in metric spaces have been studied before, the chosen approach cannot be extended to include more general types of integral equations. We introduce a definition which combines the added generality of metric spaces with the strength of Kurzweil s generalized ordinary differential equations. Additionally, we present existence and uniqueness theorems which offer new results even in the context of Euclidean spaces. Keywords: Generalized ordinary differential equations, metric spaces, curves, Henstock-Kurzweil integral.

Introduction Let f : R n (a, b) R n be a continuous function and consider the standard ordinary differential equation x = f(x, t). To solve it in the classical sense means finding an interval I (a, b) and a function x : I R n such that the derivative x (t) exists for all t I and x (t) = f(x(t), t). Given an initial condition x(t 0 ) = x 0 it is equivalent with x(t) = x 0 + (R) t 0 f(x(s), s) ds. (1) The next step in the evolution of ordinary differential equations is Carathéodory s approach, based on using the Lebesgue integral on the right side of (1) i.e. x(t 2 ) x(t 1 ) = (L) 2 t 1 f(x(s), s) ds. (2) While studying continuous dependence of (2) on a parameter, J. Kurzweil realized that all relevant information can be extracted from the function F (x, t) = t 0 f(x, s) ds. (3) In order to describe the notion of a solution of the differential equation (2) by means of the function (3) he introduced the concept of the generalized Perron integral in his paper [1]. Surprisingly, this new construction of the Perron integral was based on Riemann sums. The same construction was independently, and for unrelated reasons, discovered by R. Henstock. Let us interpret the classical ordinary diferential equation as the problem to find a function u which behaves infinitesimally as t u(τ) + f(u(τ), τ)(t τ) near each τ (a, b). J. Kurzweil used his generalized Perron integral to continue this notion and investigate a very general problem. Given a function F : R n [a, b] [a, b] R n, he looked for a function u : [a, b] R n satisfying the tangent behaviour u(t) u(τ) + F (u(τ), τ, t) F (u(τ), τ, τ) for t τ. (4) This problem is referred to as GODE (generalized ordinary differential equation). J. Kurzweil, together with J. Jarník, S. Schwabik and M. Tvrdý, continued developing this theory, even extending its methods into Banach spaces, as seen in [3], [2]. The use of generalized Perron integral allows to handle a much wider class of problems than the Carathéodory approach. Any attempt to study ordinary differential equations in the context of metric 1

spaces struggles against the lack of linear structure. The concept of derivative has no sense for functions with values in a metric space. However, if we normalize the function F (x, τ, t) from (4) by requiring F (x, τ, τ) = x, we can avoid addition in (4) and observe that the problem to find a function u which behaves infinitesimally as F (u(τ), τ, ) near each τ [a, b] can be posed in metric spaces. Using a completely different notation and terminology, a similar approach has been actually used to study differential equations (also called mutational equations ) in metric spaces, see e.g. [4, 5, 6] by J.-P. Aubin, [7, 8] by T. Lorenz and [9] by J. Tabor. However, this theory has been developed only in the classical or Carathéodory setting. So far, the theory of GODEs has not been considered in metric spaces. The reason for this is that the concept of a solution is based on integration, and therefore requires linear structure of the target. Our idea is to modify the concept of solution so that the new definition has a good sense in metric spaces. This enables us to build a theory which has the power to include both the theory of GODEs and the current theories in metric spaces. Our main achievements are uniqueness and existence theorems which show that existing results of the GODE theory can be carried over into metric spaces and even offer some improvements. The first chapter contains notation and a recollection of standard theorems. The second chapter presents the basics of the strong Kurzweil-Henstock integral and generalized ordinary differential equations. The third chapter describes methods of extending the definition of a generalized ordinary differential equation into metric spaces. The fourth and fifth chapter deal with uniqueness and existence of a solution respectively. The last chapter deals with comparison of provided and existing theorems in the context of Euclidean spaces. 2

Preliminaries We will use the following standard symbols: N R R + R + 0 The set of natural numbers. The set of real numbers. The set of positive real numbers. The set of nonnegative real numbers. f : X Y f is a mapping from X to Y. A B The set A is a subset of the set B. A B The intersection of sets A and B. A B The union of sets A and B. x M x is an element of M. φ(x) The image of the set X under the mapping φ. φ 1 (X) The preimage of the set X under the mapping φ. B(x, r) Closed ball with the center x and radius r. U(x, r) Open ball with the center x and radius r. C(K, L) The space of continuous mappings between K and L. f g The empty set. The composition mapping x f(g(x)). x Absolute value of x R. For x R we say that x is positive if x > 0 and nonnegative if x 0. A function f : R R is increasing if f(s) < f(t) for s < t and nondecreasing if f(s) f(t) for s < t. For a real function h : R R we will use h(x+) to denote the limit of h at the point x R from the right, if it exists. By D + f(τ) we denote the upper right derivative of the function f at the point τ R i.e. D + f(τ) = lim sup t τ + f(t) f(τ) t τ Theorem (Arzela-Ascoli). Let (K, ϱ) and (L, σ) be metric spaces, let K be compact and let M be a subset of C(K, L). respect to the topology of uniform convergence if Then M is relatively compact with (A1) The set M is uniformly equicontinuous i.e. for every ε > 0 there exists δ > 0 such that all f M and x, y K with ϱ(x, y) < δ satisfy σ(f(x), f(y)) < ε. (A2) The set {f(x) ; f M} has a compact closure for every x K. Theorem (Carathéodory). Let the function F : R n [a, b] R n satisfy the following conditions: (C1) The function x f(x, t) is continuous for almost all t [a, b]. 3

(C2) The function t f(x, t) is measurable for all x R n. (C3) There exists a Lebesgue integrable function g : [a, b] R + 0 such that f(x, t) g(t) for all x R n and almost all t [a, b]. Then for every x 0 R n and every t 0 [a, b] there exists > 0 and a function x : [a, b] [t 0, t 0 + ] R n such that x(t) = x 0 + (L) f(x(s), s) ds t 0 for t [a, b] [t 0, t 0 + ]. 4

1. Current Theory 1.1 SKH Integral The term partition of [a, b] R will stand for any collection of closed intervals and tags {[t i 1, t i ], τ i } k i=1 such that t 0 = a, t k = b and τ i [t i 1, t i ]. Let δ : [a, b] R + be a positive real function defined on [a, b]. A partition of [a, b] is called δ-fine if for each i = 1,..., k it satisfies [t i 1, t i ] [τ i δ(τ i ), τ i + δ(τ i )]. Lemma 1.1 (Cousin). For every δ : [a, b] R + the set of all δ-fine partitions of [a, b] is nonempty. Proof. There exist finitely many τ 1,..., τ n such that n [a, b] U(τ i, δ(τ i )). i=1 Choose a minimal subset of τ 1,..., τ n such that the intervals U(τ i, δ(τ i )) still cover the whole interval [a, b] and rename it to τ 1 <... < τ m. From the minimality of the cover we can observe that for every i = 1,..., m 1 we have U(τ i, δ(τ i )) U(τ i+1, δ(τ i+1 )). Otherwise, every point of the interval (τ i + δ(τ i ), τ i+1 δ(τ i+1 )) would have to be covered by a ball containing either U(τ i, δ(τ i )) or U(τ i+1, δ(τ i+1 )), therefore contradicting the minimality of the cover. We set t 0 = a, t m+1 = b and choose t i U(τ i, δ(τ i )) U(τ i+1, δ(τ i+1 )) (τ i, τ i+1 ) arbitrarily. The resulting partition is obviously δ-fine. Definition 1.2 (J. Kurzweil). A function f : [a, b] R n is called KH integrable (Kurzweil-Henstock) over [a, b] if there exists A R n such that for every ε > 0 there exists δ : [a, b] R + such that every δ-fine partition {[t i 1, t i ], τ i } k i=1 of [a, b] satisfies k A f(τ i )(t i t i 1 ) < ε. i=1 We say that A is the definite KH integral of f over [a, b] and use the symbol (HK) b a f(x) dx. A function u : [a, b] R n will be called an indefinite KH integral of f on [a, b] if u(t) u(s) = (HK) s f(x) dx for [s, t] [a, b]. 5

While this might seem like a simple modification of the Riemann definition, the resulting integral is equivalent to the Perron integral. The following version, called Strong Kurzweil-Henstock integral, is better suited to dealing with values in more abstract spaces. In contrast with common habits, we no longer limit ourselves to closed intervals as domains for the indefinite integral. In the rest of this chapter E will denote a normed linear space and I R will denote an interval. Definition 1.3. A function u : I E is called an indefinite SKH integral of f : I E if for every ε > 0 and every [S, T ] I there exists δ : [S, T ] R + such that for every δ-fine partition {[t i 1, t i ], τ i } k i=1 of [S, T ] we have k u(t i ) u(t i 1 ) f(τ i )(t i t i 1 ) < ε. i=1 We define the definite SKH integral of f over [S, T ] I as T (SKH) f(x) dx = u(t ) u(s). S Obviously, any SKH integrable function is KH integrable. The reverse is only true in case of E = R n and is called the Saks-Henstock Lemma. However, for the definition of the generalized ordinary differential equations we will need an even more general type of integration. The idea is that instead of integrating with respect to an additive set function like (t i t i 1 ) or, in the Stieltjes case, g(t i ) g(t i 1 ), we use a fully nonadditive expression. Definition 1.4. A function u : I E is called an indefinite SKH integral of U : I I E if for every ε > 0 and every [S, T ] I there exists δ : [S, T ] R + such that for every δ-fine partition {[t i 1, t i ], τ i } k i=1 of [S, T ] we have k u(t i ) u(t i 1 ) U(τ i, t i ) + U(τ i, t i 1 ) < ε. i=1 We define the definite SKH integral of U over [S, T ] I as T (SKH) D t U(τ, t) = u(t ) u(s). S In this case, the integrand is referred to as a function of coupled variables in order to call attention to their fundamentally different roles. As we can easily observe, a function u : I E is an indefinite SKH integral of f : I E if it is an indefinite SKH integral of U(τ, t) = f(τ) t in the sense of coupled variables. 6

1.2 MC Integral The monotonically controlled (MC for short) integral was introduced in [10] by J. Malý and H. Bendová. Their aim was to build the foundations of integral theory at the generality of Perron integral while using unexpectedly simple definitions and proofs. They prove that the MC integral of a function with respect to a Lebesgue-Stieltjes measure is equivalent to the corresponding SKH integral. In this section we generalize the defnition of the MC integral and the equivalence result to the case of coupled variables, which will often prove useful. Definition 1.5. A function u : I E is called an indefinite MC integral of U : I I E if there exists an increasing function ξ : I R, called control function of (U, u) on I, such that u(t) u(τ) U(τ, t) + U(τ, τ) lim t τ, t I ξ(t) ξ(τ) We define the definite MC integral of U over [S, T ] I as (MC) T S D t U(τ, t) = u(t ) u(s). = 0 for τ I. Note that if ξ is a control function of (U, u) on I, α > 0 and ζ : I R is a nondecreasing function, then αξ + ζ is also a control function of (U, u) on I. Lemma 1.6. Let {a k } k N and {b k } k N be sequences of real numbers such that a k < b k for every k N. Let (a k ) a, (b k ) b and let u : (a, b) E be an indefinite MC integral of U : (a, b) (a, b) E on (a k, b k ) for each k N. Then u is an indefinite MC integral of U on (a, b) and the control function of (U, u) on (a, b) can be assumed to be bounded. Proof. For each k N let γ k be a control function of (U, u) on (a k+1, b k+1 ). Since γ k is bounded on (a k, b k ), we may assume that 0 γ k 1 on (a k, b k ). We define the functions ζ k as 0, x a k, ζ k (x) = γ k (τ), a k < τ < b k, 1, x b k. The bounded control function of (U, u) on (a, b) is then found in the form ζ = 2 k ζ k. k=1 Theorem 1.7 (Equivalence). A function u : I E is an indefinite MC integral of U : I I E on I if and only if it is an indefinite SKH integral of U on I. 7

Proof. First, we assume that I = [a, b] and that u is an indefinite MC integral of U on [a, b] with a control function satisfying 0 ξ 1. Choose ε > 0. Then for each τ [a, b] there exists δ(τ) > 0 such that every t (τ δ(τ), τ + δ(τ)) [a, b] satisfies u(t) u(τ) U(τ, t) + U(τ, τ) < ε ξ(t) ξ(τ). For a δ/2-fine partition {[t i 1, t i ], τ i } k i=1 we can see that u(t i ) u(τ i ) U(τ i, t i ) + U(τ i, τ i ) < ε(ξ(t i ) ξ(τ i )), u(τ i ) u(t i 1 ) + U(τ i, t i 1 ) U(τ i, τ i ) < ε(ξ(τ i ) ξ(t i 1 )). Therefore, we obtain k u(t i ) u(t i 1 ) U(τ i, t i ) + U(τ i, t i 1 ) < ε i=1 k (ξ(t i ) ξ(τ i ) + ξ(τ i ) ξ(t i 1 )) = ε(ξ(b) ξ(a)) ε. i=1 Let u be an indefinite SKH integral of U on [a, b]. For k N let δ k : [a, b] R + correspond to ε k = 2 k. For τ (a, b] we define P k (τ) as the set of all δ k -fine partitions of [a, τ] and ξ k (τ) = sup { P u(t i ) u(t i 1 ) U(τ i, t i ) + U(τ i, t i 1 ) ; P P k (τ)}, ξ(τ) = τ + k ξ k (τ). k=1 Choose ε > 0 and find k 0 N such that 1 k 0 t (τ δ k0 (τ), τ) we have < ε. It follows that for τ (a, b] and u(t) u(τ) U(τ, t) + U(τ, τ) ξ k0 (τ) ξ k0 (t) k (ξ k (τ) ξ k (t)) 1 (ξ(τ) ξ(t)) < ε(ξ(τ) ξ(t)). k 0 k 0 k=1 Similarly, for t (τ, τ + δ k0 (τ)) we have u(t) u(τ) U(τ, t) + U(τ, τ) ξ k0 (t) ξ k0 (τ) < ε(ξ(t) ξ(τ)). If I is not necessarily closed, we easily write it as a union of an increasing sequence of closed intervals and use Lemma 1.6. 8

1.3 Generalized Ordinary Differential Equations Definition 1.8. Let F : E [a, b] [a, b] E be an arbitrary mapping. We say that a function u : [a, b] E is a solution of the equation x = D t F (x, τ, t) on [a, b] if u(t) = u(s) + (SKH) s D t F (u(τ), τ, t) for [s, t] [a, b]. (1.1) However, when working with the definition, we will find it more convenient to interpret (1.1) in terms of indefinite SKH integral i.e. for every ε > 0 there exists δ : [a, b] R + such that every δ-fine partition P of [a, b] satisfies u(t i ) u(t i 1 ) F (u(τ i ), τ i, t i ) + F (u(τ i ), τ i, t i 1 ) < ε. (1.2) P It can always be arranged for the function F to satisfy F (x, τ, τ) = x without changing the equation x = D t F (x, τ, t). This is done by considering the function F (x, τ, t) = x + F (x, τ, t) F (x, τ, τ), since then we will have F (u(τ i ), τ i, t i 1 ) F (u(τ i ), τ i, t i ) = F (u(τ i ), τ i, t i 1 ) F (u(τ i ), τ i, t i ). This process is called normalization and is one of the main ideas behind extending generalized ordinary differential equations into metric spaces. It is what allows us to interpret the equation as u(t) F (u(τ), τ, t) for t τ. To be more specific, it will allow us to express the sum (1.2) in terms of distance of two elements by using the identity u(τ) = F (u(τ), τ, τ). Lemma 1.9. Let F : E [a, b] [a, b] E satisfy F (x, τ, τ) = x. A function u : [a, b] E is a solution of the equation x = D t F (x, τ, t) on [a, b] if and only if for every ε > 0 there exists δ : [a, b] R + such that every δ-fine partition P of [a, b] satisfies ( ) u(t i ) F (u(τ i ), τ i, t i ) + u(t i 1 ) F (u(τ i ), τ i, t i 1 ) < ε. (1.3) P Proof. Fix ε > 0 and find δ : [a, b] R + such that (1.2) holds for every δ-fine partition P = {[t i 1, t i ], τ i } k i=1 of [a, b]. However, if {[t i 1, t i ], τ i } k i=1 is δ-fine, then the partition {[s j 1, s j ], σ j } 2k j=1 = {..., [t i 1, τ i ], τ i, [τ i, t i ], τ i,...} is also δ-fine, and therefore k ( u(t i ) u(τ i ) F (u(τ i ), τ i, t i ) + F (u(τ i ), τ i, τ i ) i=1 ) + u(τ i ) u(t i 1 ) F (u(τ i ), τ i, τ i ) + F (u(τ i ), τ i, t i 1 ) < ε. 9

By using u(τ i ) = F (u(τ i ), τ i, τ i ) we obtain k i=1 ( ) u(t i ) F (u(τ i ), τ i, t i ) + u(t i 1 ) F (u(τ i ), τ i, t i 1 ) < ε. Let us now assume that for every ε > 0 there exists δ : [a, b] R + such that every δ-fine partition P of [a, b] satisfies (1.3). Then we have u(t i ) u(t i 1 ) F (u(τ i ), τ i, t i ) + F (u(τ i ), τ i, t i 1 ) P P = P ( u(t i ) u(τ i ) F (u(τ i ), τ i, t i ) + F (u(τ i ), τ i, τ i ) ) + u(τ i ) u(t i 1 ) F (u(τ i ), τ i, τ i ) + F (u(τ i ), τ i, t i 1 ) ( ) u(t i ) F (u(τ i ), τ i, t i ) + u(t i 1 ) F (u(τ i ), τ i, t i 1 ) < ε. As this was the desired result, the proof is finished. Many works concerned with generalized ordinary differential equation theory only study the equation x = D t G(x, t), defined by u(t) = u(s) + (SKH) s D t G(u(τ), t) for [s, t] [a, b]. The reason for this is that a large class of GODEs (see [2]) can be simplified by putting G(x, t) = (SKH) a D s F (x, σ, s). However, this type of equation offers no analogy to normalization. As an example, consider Carathéodory s equation x(ν) x(σ) = (L) It is equivalent with x = D t G(x, t), where G(x, t) = (L) ν σ f(x(s), s) ds. t 0 f(x, s) ds for t 0 fixed (see [3], Theorem 5.14), while the normalized form is F (x, τ, t) = x + (L) τ f(x, s) ds. Note that the previously mentioned choice F (x, τ, t) = f(x, τ) t and its normalized form F (x, τ, t) = x + f(x, τ)(t τ) are equivalent with x(ν) x(σ) = (SKH) ν σ f(x(s), s) ds, as seen, among other options, in Theorem 17.3 of [2]. 10

2. Solutions in metric spaces In the remainder of the text X always denotes a metric space. We adopt the convention that x y stands for the distance between x and y. Definition 2.1 (J. Malý and B. Skovajsa). Let I R be an interval and assume that F : X I I X satisfies F (x, τ, τ) = x. We say that u : I X is a solution of x = D t F (x, τ, t) on I if T (SKH) D t u(t) F (u(τ), τ, t) = 0 for [S, T ] I. (2.1) S Equivalently, the indefinite SKH integral of U(τ, t) = u(t) F (u(τ), τ, t) is equal to zero on I. As a trivial consequence of the equivalence theorem we have that substituting the MC integral into 2.1 results in an equivalent definition. We expand the definition of a solution into more practical forms: (i) We say that u : (a, b) X is a solution of x there exists an increasing function ξ : (a, b) R such that u(t) F (u(τ), τ, t) lim t τ ξ(t) ξ(τ) In this case we will refer to ξ as control function of (F, u). = D t F (x, τ, t) on (a, b) if = 0 for τ (a, b). (2.2) (ii) Alternatively, we say that u : [a, b] X is a solution of x = D t F (x, τ, t) on [a, b] if for every ε > 0 there exists δ : [a, b] R + such that every δ-fine partition P of [a, b] satisfies ( ) u(t i ) F (u(τ i ), τ i, t i ) + u(t i 1 ) F (u(τ i ), τ i, t i 1 ) < ε. (2.3) P Proof of (i). The form (2.2) is a direct interpretation of (2.1). The indefinite MC integral of u(t) F (u(τ), τ, t) being equal to zero on (a, b) is defined by the property lim t τ u(t) F (u(τ), τ, t) u(τ) F (u(τ), τ, τ) ξ(t) ξ(τ) = 0. However, we assumed F to be normalized, which means u(τ) = F (u(τ), τ, τ). Therefore, u(τ) F (u(τ), τ, τ) = 0 and we obtain (2.2). 11

Proof of (ii). Let us assume that the indefinite SKH integral of u(t) F (u(τ), τ, t) is zero on [a, b] i.e. for every ε > 0 there exists δ : [a, b] R + such that for every δ-fine partition {[t i 1, t i ], τ i } k i=1 of [a, b] we have k u(t i ) F (u(τ i ), τ i, t i ) u(t i 1 ) F (u(τ i ), τ i, t i 1 ) < ε. i=1 However, if {[t i 1, t i ], τ i } k i=1 is δ-fine, then the partition is also δ-fine, and therefore 2k j=1 {[s j 1, s j ], σ j } 2k j=1 = {..., [t i 1, τ i ], τ i, [τ i, t i ], τ i,...} u(s j ) F (u(σ j ), σ j, s j ) u(s j 1 ) F (u(σ j ), σ j, s j 1 ) < ε. (2.4) For j = 2i 1 we have u(s j ) F (u(σ j ), σ j, s j ) = u(τ i ) F (u(τ i ), τ i, τ i ) = 0, u(s j 1 ) F (u(σ j ), σ j, s j 1 ) = u(t i 1 ) F (u(τ i ), τ i, t i 1 ). (2.5) Similarly, for j = 2i we have u(s j ) F (u(σ j ), σ j, s j ) = u(t i ) F (u(τ i ), τ i, t i ), u(s j 1 ) F (u(σ j ), σ j, s j 1 ) = u(τ i ) F (u(τ i ), τ i, τ i ) = 0. (2.6) By substituting (2.5) and (2.6) back into (2.4) we get (2.3). The reverse implication is trivial. We can now observe that on normed linear spaces our definition is equivalent to Kurzweil s as a trivial consequence of (2.3) and Lemma 1.9. This can also be easily achieved by using (2.2). Let E be a normed linear space and let F : E (a, b) (a, b) E satisfy F (x, τ, τ) = x. We can see that u : (a, b) E is an indefinite MC integral of F (u(τ), τ, t), and therefore a solution of x = D t F (x, τ, t), if and only if there exists an increasing function ξ : (a, b) R such that u(t) u(τ) F (u(τ), τ, t) + F (u(τ), τ, τ) lim t τ ξ(t) ξ(τ) = 0 for τ (a, b). However, this is equivalent with (2.2) due to u(τ) = F (u(τ), τ, τ). 12

3. Uniqueness Our focus when dealing with uniqueness and existence of a solution will be to generalize theorems from [3] by S. Schwabik. We restrict our attention to the continuous case. Ultimately, we hope to treat solutions in the space of regulated functions. However, this leads to additional technical complications which will be pursued in future research. Definition 3.1. Let X be a metric space, I R an interval and x : I X a solution of x = D t F (x, τ, t) on [τ, τ + δ] I with x(τ) = x τ. We say that x is locally unique in the future at the point (x τ, τ) if for any solution y : I X of x = D t F (x, τ, t) on [τ, τ + η] such that y(τ) = x τ there exists ζ > 0 such that y(t) = x(t) for t [τ, τ + ζ]. Definition 3.2. We say that G : Ω R n belongs to the class F(Ω, h, ω) if G(x, t 2 ) G(x, t 1 ) h(t 2 ) h(t 1 ) (3.1) G(x, t 2 ) G(x, t 1 ) G(y, t 2 ) + G(y, t 1 ) ω( x y ) h(t 2 ) h(t 1 ) (3.2) for (x, t 2 ), (x, t 1 ), (y, t 2 ), (y, t 1 ) Ω, where Ω = U R (a, b), h : [a, b] R is nondecreasing and ω : R + 0 R + 0 is continuous and nondecreasing with ω(0) = 0 and ω(ν) > 0 for ν > 0. The symbol U R stands for {x R n ; x < R}. In the case of uniqueness, our motivation comes from the following theorem: Theorem 3.3 (S. Schwabik, [3], page 122 ). Let G : Ω R n belong to F(Ω, h, ω), where the function h is nondecreasing and continuous from the left, ω : R + 0 R + 0 is continuous, nondecreasing, ω(0) = 0 and for ν > 0 we have ω(ν) > 0 and ν lim r 0 + r 1 ds = +. (3.3) ω(s) Then every solution x : (a, b) R n of x = D t G(x, t) such that x(τ) + lim t τ+ G(x(τ), t) G(x(τ), τ) U R is locally unique in the future at (x(τ), τ). Lemma 3.4. Let f : (a, b) R be a real function which at every point x (a, b) satisfies (T1) D + f(x) 0, (T2) lim sup f(x h) f(x). h 0 + Then f is nondecreasing on (a, b). Proof. See [12], page 135. 13

Theorem 3.5. Let X be a metric space, let F : X (a, b) (a, b) X fulfil F (x, τ, τ) = x and let us assume the following: (U1) The function t F (x, τ, t) is continuous for every x X and τ (a, b). (U2) There exists a nondecreasing and continuous function ω : R + 0 R + 0 such that ω(0) = 0 and for ν > 0 we have ω(ν) > 0 and ν lim r 0 + r 1 ds = +. (3.4) ω(s) (U3) There exists ξ : (a, b) R increasing such that all x, y X and all τ (a, b) satisfy lim inf t τ + F (x, τ, t) F (y, τ, t) x y ξ(t) ξ(τ) ω( x y ). (3.5) Then every solution of x = D t F (x, τ, t) on (a, b) is locally unique in the future. Proof. Let u, v be two solutions and α, β (a, b) such that α < β, u(α) = v(α) and u(β) v(β). We also assume that ξ already controls both (F, u) and (F, v) and that for t > s we have ξ(t) ξ(s) t s. Condition (U1) implies that any solution of x = D t F (x, τ, t) is continuous. Indeed, for a fixed τ (a, b) and ε > 0 we can choose δ 1 > 0 such that F (u(τ), τ, t) u(τ) < ε for t τ < δ 1. Then, we choose δ 2 (0, δ 1 ) such that u(t) F (u(τ), τ, t) < ε ξ(t) ξ(τ) for t τ < δ 2. Since ξ is increasing, we can choose δ 3 (0, δ 2 ) such that ξ(t) ξ(τ) < ξ(τ+) ξ(τ ) + 1 for t τ < δ 3. Finally, we obtain u(t) u(τ) u(t) F (u(τ), τ, t) + F (u(τ), τ, t) u(τ) < ε(ξ(τ+) ξ(τ ) + 2). Since the solutions are continuous, we can achieve u v everywhere on (α, β) by simple modification of α. Now fix ν > 0 and define (t) = u(t) v(t) and Φ(r) = ν r 1 ω(s) ds. Consider the function t Φ( (t)) + ξ(t) on (α, β). It is well defined everywhere on (α, β), since > 0 on (α, β). Since ξ is bounded on [α, β] and (α) = 0, we use (3.4) to easily deduce that lim t α + Φ( (t)) + ξ(t) = +. (3.6) We now employ Lemma 3.4 to show that t Φ( (t)) + ξ(t) is nondecreasing on (α, β), which obviously contradicts (3.6). Since ξ is increasing and Φ is continuous, we can easily see that their sum satisfies (T2). The rest of the proof will focus on showing that all τ (α, β) satisfy D + (Φ( (τ)) + ξ(τ)) 0. 14

We notice that Φ is decreasing and for every s > 0 we have Φ (s) = 1 ω(s) (3.7) We fix τ (α, β) and distinguish two cases: 1) For every ζ > 0 there exists t (τ, τ + ζ) such that (t) (τ). This implies Φ( (t)) + ξ(t) Φ( (τ)) + ξ(τ), and therefore D + (Φ( (τ)) + ξ(τ)) 0. 2) There exists ζ > 0 such that for every t (τ, τ + ζ) we have (t) > (τ). We notice that Φ ( (τ)) = Φ(s) Φ( (τ)) lim s (τ) s (τ) = lim t τ+ Φ( (t)) Φ( (τ)) (t) (τ) (3.8) Fix ε > 0. We combine (3.7) and (3.8) in order to find δ (0, ζ) such that for every t (τ, τ + δ) we have 1 Φ( (t)) Φ( (τ)) ( ε)( (t) (τ)) (3.9) ω( (τ)) Since ξ controls both (F, u) and (F, v), we can find η (0, δ) such that all t (τ, τ + η) satisfy u(t) F (u(τ), τ, t) ξ(t) ξ(τ) < ε, v(t) F (v(τ), τ, t) ξ(t) ξ(τ) < ε. (3.10) Now, fix σ (0, η) and use (3.5) to find T (τ, τ + σ) such that F (u(τ), τ, T ) F (v(τ), τ, T ) (τ) ξ(t ) ξ(τ) ω( (τ)) + ε. (3.11) We define E u (t) := u(t) F (u(τ), τ, t), E v (t) := v(t) F (v(τ), τ, t) and notice that (t) F (u(τ), τ, t) F (v(τ), τ, t) + E u (t) + E v (t). Applying this to (3.11) results in We add (3.10) and obtain (T ) (τ) E u (T ) E v (T ) ξ(t ) ξ(τ) ω( (τ)) + ε. (T ) (τ) ξ(t ) ξ(τ) ω( (τ)) + 3 ε. In combination with (3.9) we have 1 Φ( (T )) Φ( (τ)) ( + ε)(ω( (τ)) + 3 ε)(ξ(t ) ξ(τ)), ω( (τ)) Φ( (T )) Φ( (τ)) ξ(t ) ξ(τ) (1 + εc(τ)), 15

Φ( (T )) Φ( (τ)) + ξ(t ) ξ(τ) ξ(t ) ξ(τ) We recall that ξ(t) ξ(s) t s for t > s, and therefore εc(τ). Φ( (T )) Φ( (τ)) + ξ(t ) ξ(τ) T τ εc(τ). Since such T will be in every right neighbourhood of τ, we finally arrive at D + (Φ( (τ)) + ξ(τ)) 0 and the proof is finished. 16

4. Existence In the case of existence, our aim is to generalize the following theorem: Theorem 4.1 (S. Schwabik, [3], page 114). Let G : Ω R n belong to F(Ω, h, ω) and let (x τ, τ) Ω satisfy x τ + lim t τ+ G(x τ ), t) G(x τ, τ) U R. Then there exists > 0 and a solution x : [τ, τ + ] R n of x = D t G(x, t) on [τ, τ + ] with x(τ) = x τ. Once again, we add assumptions of continuity to avoid additional technical complications. As previously mentioned, this is a temporary restriction, since our ultimate goal is to find solutions among regulated functions. Lemma 4.2. Let the functions u k : [a, b] X, k N, be continuous on [a, b] and assume that for every ε > 0 and every τ [a, b] there exists k 0 N and η > 0 such that for k N, k k 0 we have that t τ < η implies u k (t) u k (τ) < ε. Then the functions {u k } k N are uniformly equicontinuous. Proof. Fix ε > 0. Since each of the functions u k is uniformly continuous, we find the corresponding δ k > 0. First, we notice that for every m N the functions {u k } k m are uniformly equicontinuous, as we can set δ := min {δ 1,..., δ m }. We can cover the interval [a, b] by the system {U(τ j, η(τ j )/2) ; j = 1,..., J}. Every τ j corresponds to a certain k j N from the assumption on functions u k. We set k 0 := max {k j ; j = 1,..., J} and δ := min {η(τ j )/2 ; j = 1,..., J}. We already know that the functions u k are uniformly equicontinuous for k k 0. For k k 0 and t, s [a, b] such that t s < δ there exists m {1,..., J} such that t U(τ m, η(τ m )/2), and therefore s U(τ m, η(τ m )). Since k k 0 k m, we obtain u k (t) u k (s) u k (t) u k (τ m ) + u k (τ m ) u k (s) < 2ε. Theorem 4.3. Let X be a metric space in which every closed ball is compact, let F : X [a, b] [a, b] X satisfy F (x, τ, τ) = x and let us assume the following: (E1) For every τ, t [a, b] the function x F (x, τ, t) is continuous. (E2) There exists ζ : R + 0 R + 0 increasing and continuous such that ζ(0) = 0 and for every x X and every τ, t, s [a, b] we have F (x, τ, t) F (x, τ, s) < ζ( t s ). 17

(E3) For every y X and τ, t [a, b] there exists z X such that F (z, τ, t) = y. (E4) There exists ξ : [a, b] R increasing such that for every ε > 0 there exists η : [a, b] R + such that all τ, σ [a, b], all t, s U(τ, η(τ)) U(σ, η(σ)) and all x, y X with x y < max {η(τ), η(σ)} satisfy F (x, τ, t) F (y, σ, t) < F (x, τ, s) F (y, σ, s) + ε ξ(t) ξ(s). Then for every x 0 X and τ 0 [a, b] there exists a solution u : [a, b] X of x = D t F (x, τ, t) on [a, b] with u(τ 0 ) = x 0. Proof. Step 1: Find a solution candidate. Fix x X and a partition {[t i 1, t i ], τ i } of [a, b]. We set v(a) := x. For t [t i 1, t i ] let v(t) := F (z, τ i, t), where z X is chosen so that F (z, τ i, t i 1 ) = v(t i 1 ). This can always be arranged due to condition (E3). For fixed δ : [a, b] R + any function constructed from a δ-fine partition will also be called δ-fine. For arbitrary τ [a, b] and δ-fine partition {[t i 1, t i ], τ i } with τ [t m, t m+1 ] the partition {... [t m 1, t m ], τ m, [t m, τ], τ, [τ, t m+1 ], τ, [t m+1, t m+2 ], τ m+1,...} is also δ-fine. Therefore, we can begin the construction from τ [a, b] and build analogously on the left side. Set ε k := 2 k, find the corresponding η k : [a, b] R + from condition (E4) and put γ k (τ) := min{2 k, η 1 (τ)/2,..., η k (τ)/2, ζ 1 (η 1 (τ)/2),..., ζ 1 (η k (τ)/2)}. (4.1) Since Lemma 1.1 ensures the existence of γ k -fine partitions P k of [a, b], we can use these partitions to construct γ k -fine approximate solutions u k : [a, b] X with u k (τ 0 ) = x 0. We now verify the assumptions of Lemma 4.2. The most important fact to remember is that the approximate solutions u k consist of segments of the form t F (u k (τ i ), τ i, t) for t [t i 1, t i ]. Fix τ [a, b) and ε > 0. First, find k 0 N such that 2 k 0 < ε and k 1 k 0 such that 2 ζ(2 k1+1 ) < η k0 (τ). Then, find β > 0 such that 2 ζ(β) < η k0 (τ) 2 ζ(2 k1+1 ) and α > 0 such that for t (τ, τ + α) we have ξ(τ+) ξ(t) < min {1/6, η k0 (τ) 2 ζ(2 k1+1 ) 2 ζ(β)}. (4.2) Set δ := min {α, β, η k0 (τ), ζ 1 (ε/6)}. 18

Fix t (τ, τ + δ(τ)) and k k 1. Since δ ζ 1 (ε/6), we notice that u k (t) u k (τ) u k (t) F (u k (τ), τ, t) + F (u k (τ), τ, t) u k (τ) u k (t) F (u k (τ), τ, t) + ε 6. (4.3) The second inequality of (4.3) holds by u k (τ) = F (u k (τ), τ, τ) and (E2). Note that hereafter we will be using the uniform continuity of F (x, τ, t) in the third variable, i.e. condition (E2), without explicit mention. We finish the estimate (4.3) by dealing with u k (t) F (u k (τ), τ, t). Find j, m N, j m such that for the corresponding intervals [t j 1, t j ] P k and [t m 1, t m ] P k we have τ [t j 1, t j ) and t [t m 1, t m ]. The necessity to forbid τ = t j comes from the possibility of τ being a point of discontinuity of the function ξ. The case j = m is trivial, since t [t j 1, t j ] implies u k (t) u k (τ) < ε/6. Now, assume that m j + 1 and use induction by i to proceed. Our aim will be to show that if F (u k (τ), τ, s) u k (s) < F (u k (τ), τ, t j ) u k (t j ) + ε k0 ξ(s) ξ(t j ) (4.4) holds for all s [t i 1, t i ] [τ, t], then it holds for all s [t i, t i+1 ] [τ, t], where i = j + 1,..., m 1. First, we need to verify that (4.4) holds for s [t j, t j+1 ] [τ, t]. We recall that (4.1) ensures that u k are 2 n fine for n k. Together with k k 1 we have t j τ 2 k 1+1, and therefore u k (τ) u k (t j ) ζ(2 k 1+1 ). Similarly, u k (t j ) u k (τ j+1 ) ζ(2 k 1+1 ). Since k 1 was chosen so that 2 ζ(2 k 1+1 ) < η k0 (τ), we obtain u k (τ) u k (τ j+1 ) < η k0 (τ). We also recall that u k are constructed to be η n /2-fine for n k, in particular η k0 /2-fine. Together with τ s τ t < δ η k0 (τ) we have that [t j, s] U(τ, η k0 (τ)) U(τ j+1, η k0 (τ j+1 )). This allows us to use condition (E4) with F (u k (τ j+1 ), τ j+1, s) = u k (s) and F (u k (τ j+1 ), τ j+1, t j ) = u k (t j ) to obtain F (u k (τ), τ, s) u k (s) < F (u k (τ), τ, t j ) u k (t j ) + ε k0 ξ(s) ξ(t j ). Now, assume that j +2 i m 1 and that (4.4) holds for all s [t i 1, t i ] [τ, t]. Since i < m, we have [t i 1, t i ] [τ, t]. This means we can use (4.4) with s = t i to obtain F (u k (τ), τ, t i ) u k (t i ) < F (u k (τ), τ, t j ) u k (t j ) + ε k0 ξ(t i ) ξ(t j ). 19

Fix s [t i, t i+1 ] [τ, t] and note that [t i, s] U(τ, η k0 (τ)) U(τ i+1, η k0 (τ i+1 )). In order to justify the use of condition (E4), we compute u k (τ) u k (τ i+1 ) u k (τ) F (u k (τ), τ, t i ) + F (u k (τ), τ, t i ) u k (t i ) + u k (t i ) u k (τ i+1 ) ζ(δ) + F (u k (τ), τ, t i ) u k (t i ) + ζ(2 k1+1 ) ζ(δ) + ζ(2 k1+1 ) + F (u k (τ), τ, t j ) u k (t j ) + ε k0 ξ(t i ) ξ(t j ) ζ(δ) + ζ(2 k1+1 ) + F (u k (τ), τ, t j ) u k (τ) + u k (τ) u k (t j ) + ε k0 ξ(t i ) ξ(t j ) 2 ζ(δ) + 2 ζ(2 k1+1 ) + ε k0 ξ(t i ) ξ(t j ) 2 ζ(δ) + 2 ζ(2 k1+1 ) + ξ(t i ) ξ(τ+) However, 2 ζ(δ) + 2 ζ(2 k1+1 ) + ξ(t i ) ξ(τ+) < η k0 (τ) due to (4.2). Therefore, we can use (E4) again to get F (u k (τ), τ, s) u k (s) < F (u k (τ), τ, t i ) u k (t i ) + ε k0 ξ(s) ξ(t i ) < F (u k (τ), τ, t j ) u k (t j ) + ε k0 ξ(t i ) ξ(t j ) + ε k0 ξ(s) ξ(t i ) = F (u k (τ), τ, t j ) u k (t j ) + ε k0 ξ(s) ξ(t j ). This completes the induction step and by setting s = t we obtain the estimate F (u k (τ), τ, t) u k (t) < F (u k (τ), τ, t j ) u k (t j ) + ε k0 ξ(t) ξ(t j ). (4.5) We recall that we have u k (t) u k (τ) u k (t) F (u k (τ), τ, t) + ε 6 We combine this with ξ(t) ξ(τ+) < 1/6, ε k0 < ε and (4.5) to obtain We finish by using u k (t) u k (τ) F (u k (τ), τ, t j ) u k (t j ) + 2 ε 6. F (u k (τ), τ, t j ) u k (t j ) F (u k (τ), τ, t j ) u k (τ) + u k (τ) u k (t j ) < 2 ε 6. Thus, it becomes clear that u k (t) u k (τ) < ε. An analogous procedure can be done on the left neighbourhood of every τ (a, b], and therefore {u k } k N satisfy the assumptions of Lemma 4.2. As a result, they are uniformly equicontinuous i.e. there exists ω : R + 0 R + 0 increasing and continuous such that ω(0) = 0 and for every t, s [a, b] and every k N we have u k (t) u k (s) < ω( t s ). 20

Set δ := ω 1 (1) and find M N such that M (b a)/δ. Then we can see that {u k } k N B(x 0, M). Note that we will be using the uniform equicontinuity of the solutions with the continuity module ω without further mention. The Arzela-Ascoli theorem now ensures the existence of a subsequence of {u k } uniformly converging to u : [a, b] X. Step 2: We now prove that u solves x = D t F (x, τ, t). For an arbitrary partition P of [a, b] we can see that ( ) u(t i 1 ) F (u(τ i ), τ i, t i 1 ) + u(t i ) F (u(τ i ), τ i, t i ) P = ( lim u k (t i 1 ) lim F (u k (τ i ), τ i, t i 1 ) k k P ) + lim u k (t i ) lim F (u k (τ i ), τ i, t i ) k k = ( ) lim u k(t i 1 ) F (u k (τ i ), τ i, t i 1 ) + lim u k (t i ) F (u k (τ i ), τ i, t i ) k k P ( ) = lim u k (t i 1 ) F (u k (τ i ), τ i, t i 1 ) + u k (t i ) F (u k (τ i ), τ i, t i ). k P Therefore, it suffices to show that for every ε > 0 there exists δ : [a, b] R + and k 0 N such that for every k N, k k 0 and every δ-fine partition P of [a, b] we have u k (t i 1 ) F (u k (τ i ), τ i, t i 1 ) + u k (t i ) F (u k (τ i ), τ i, t i ) < ε. P Fix k N and recall the construction of u k from P k. For τ [a, b] we distinguish four cases: (i) There exists α > 0 and i N such that [τ α, τ + α] [t i 1, t i ]. (ii) For τ = a there exists α > 0 such that [a, a + α] [t 0, t 1 ]. (iii) For τ = b there exists α > 0 and i N such that [b α, b] [t i 1, t i ]. (iv) There exists i N such that t i = τ. Therefore, we can find α > 0 such that [τ α, τ] [t i 1, t i ] and [τ, τ + α] [t i, t i+1 ]. For fixed k N and τ [a, b] we will denote the corresponding α by α(k, τ). Fix ε > 0 and find k 0 N such that 2 k 0 < ε/(ξ(b) ξ(a)). Fix τ [a, b] and let the function η correspond to ε k0 = 2 k 0. Find k 1 k 0 such that ζ(2 k 1 ) < η(τ). For h 0 define β(h) := η(τ) ζ(2 k 1 ) ω(h). This function is decreasing, continuous and β(0) = η(τ) ζ(2 k 1 ) > 0. These properties imply the existence of h 0 (0, η(τ)) such that β(h) > 0 for h [0, h 0 ]. Define δ(τ) = min {h 0, α(k 0, τ), α(k 0 + 1, τ),..., α(k 1, τ)}. 21

Now, fix k k 0 and r, s [a, b] such that τ δ(τ) r τ s τ + δ(τ). First, let k 0 k k 1. Since δ(τ) α(k, τ), there exists τ j P k such that u k (τ) = F (u k (τ j ), τ j, τ) and u k (s) = F (u k (τ j ), τ j, s). We recall that (4.1) yields τ τ j ζ 1 (η(τ j )/2), and therefore u k (τ) u k (τ j ) < η(τ j ). We also recall that τ s δ(τ) < η(τ) and τ, s B(τ j, η(τ j )/2) U(τ j, η(τ j )). Condition (E4) now implies F (u k (τ), τ, s) F (u k (τ j ), τ j, s) < F (u k (τ), τ, τ) F (u k (τ j ), τ j, τ) + ε k0 ξ(s) ξ(τ). By using F (u k (τ), τ, τ) = u k (τ) = F (u k (τ j ), τ j, τ) and F (u k (τ j ), τ j, s) = u k (s) we get F (u k (τ), τ, s) u k (s) < ε k0 ξ(s) ξ(τ). An analogous procedure on the left neighbourhood of τ gives us F (u k (τ), τ, r) u k (r) < ε k0 ξ(τ) ξ(r). Now, let k k 1. Again, the condition τ t < η(τ) is obviously satisfied for t [τ, s] and the use of condition (E4) is only limited by u k (τ) u k (τ i ) for τ i P k. However, unlike Step 1, we do not need the conclusion of condition (E4) to hold on [t i 1, t i ] in order to verify its assumptions on [t i, t i+1 ]. Find j, m N such that τ [t j 1, t j ] and s [t m 1, t m ] and note that u k (τ) u k (τ j ) ζ(2 k 1 ) < η(τ), u k (τ) u k (τ j+1 ) ω( τ τ j+1 ) < η(τ), u k (τ) u k (τ j+2 ) ω( τ τ j+2 ) < η(τ),. u k (τ) u k (τ m 1 ) ω( τ τ m 1 ) < η(τ), u k (τ) u k (τ m ) u k (τ) u k (s) + u k (s) u k (τ m ) ω( τ s ) + ζ(2 k 1 ) < η(τ). Persistent use of condition (E4) gives us F (u k (τ), τ, s) u k (s) < F (u k (τ), τ, t m 1 ) u k (t m 1 ) + ε k0 ξ(s) ξ(t m 1 ), F (u k (τ), τ, t m 1 ) u k (t m 1 ) < F (u k (τ), τ, t m 2 ) u k (t m 2 ) + ε k0 ξ(t m 1 ) ξ(t m 2 ),. F (u k (τ), τ, t j+1 ) u k (t j+1 ) < F (u k (τ), τ, t j ) u k (t j ) + ε k0 ξ(t j+1 ) ξ(t j ), F (u k (τ), τ, t j ) u k (t j ) < F (u k (τ), τ, τ) u k (τ) + ε k0 ξ(t j ) ξ(τ) = ε k0 ξ(t j ) ξ(τ). 22

Therefore, F (u k (τ), τ, s) u k (s) < ε k0 ξ(s) ξ(t m 1 ) +... + ε k0 ξ(t j ) ξ(τ) = ε k0 ξ(s) ξ(τ). For any δ-fine partition {[s i 1, s i ], σ i } n i=1 of [a, b] and k k 0 we obtain n u k (s i 1 ) F (u k (σ i ), σ i, s i 1 ) + u(s i ) F (u k (σ i ), σ i, s i ) i=1 < 2 k 0 n ξ(s i 1 ) ξ(σ i ) + ξ(σ i ) ξ(s i ) = 2 k 0 (ξ(b) ξ(a)) < ε, i=1 and the proof is finished. 23

5. Euclidean Case First of all, we present an example that demonstrates the difference between the generalized and standard ordinary differential equation theory. Consider the functions G(x, t) = C(t), F (x, τ, t) = x + C(t) C(τ), on the domain R [0, 1] and R [0, 1] [0, 1] respectively, where C(t) is the Cantor stair function. These functions satisfy the assumptions of theorems 3.3, 3.5, 4.1 and 4.3 and solutions of the associated GODEs are vertical translations of the Cantor function itself. However, it is not covered by Carathéodory s approach (much less by the classical theory), as the solutions are not absolutely continuous. Uniqueness Let us recall the uniqueness theorem 3.3. Our main contribution to the theory of generalized ordinary differential equations on Euclidean spaces is that we can replace condition (3.1) (bounded variation) with continuity of t G(x, t) at every x U R. To this end, we verify the assumptions of Theorem 3.5 for the function F (x, τ, t) = x + G(x, t) G(x, τ). Indeed, if t G(x, t) is continuous, then the function t x + G(x, t) G(x, τ) is also continuous. Conditions (3.3) and (3.4) are identical and require no further comment. Notice that F (x, τ, t) F (y, τ, t) = x y + G(x, t) G(x, τ) + G(y, t) G(y, τ). Therefore, F (x, τ, t) F (y, τ, t) x y G(x, t) G(x, τ) + G(y, t) G(y, τ). If (3.2) holds, then we have G(x, t) G(x, τ) + G(y, t) G(y, τ) ω( x y ) h(t) h(τ), and therefore lim inf t τ + F (x, τ, t) F (y, τ, t) x y h(t) h(τ) ω( x y ). Another notable difference between Theorem 3.3 and Theorem 3.5 is that we have managed to weaken (3.2) to a local version. 24

Existence Similarly to uniqueness, we can replace bounded variation in Theorem 4.1 with continuity of t G(x, t). Once again, we verify the assumptions of Theorem 4.3 for F (x, τ, t) = x + G(x, t) G(x, τ). We notice that F (x, τ, t) F (y, τ, t) = x + G(x, t) G(x, τ) y G(y, t) + G(y, τ) x y + G(x, t) G(x, τ) G(y, t) + G(y, τ) x y + ω( x y ) h(t) h(τ). Condition (E1) is therefore satisfied. Since Theorem 4.1 only gives local existence, we can confine the problem to a compact set K around the initial condition. As a result, F (x, τ, t) is uniformly continuous on K [a, b] [a, b], and therefore satisfies (E2). The condition (E3) is then a simple result of topological degree theory, as for t τ small enough the mapping x F (x, τ, t) is uniformly close to the identity mapping x F (x, τ, τ). Trading bounded variation for continuity might seem impractical at first. However, consider the following existence theorem from [11] by R. Henstock. We will show that while it is not contained in Theorem 4.1, it does follow from Theorem 4.3. Theorem 5.1. Assume that f : R n [a, b] R n satisfies the following conditions: (H1) The function x f(x, t) is continuous for almost all t [a, b]. (H2) The function t f(x, t) is SKH integrable over [a, b] for every x R n. (H3) There exists S R n compact and δ : [a, b] R + such that all δ-fine partitions {α = α 0, τ 1, α 1,..., τ k, α k = β} of [α, β] [a, b] and all functions w : [a, b] R n satisfy k f(w(τ i ), τ i )(α i α i 1 ) S. i=1 Then for every v R n and τ [a, b] there exists y : [a, b] R n such that y(t) = v + (SKH) τ f(y(s), s) ds for t [a, b]. We prove that (H1) (H3) for f : R n [a, b] R n imply (E1) (E4) for F (x, τ, t) = x + (SKH) τ f(x, s) ds. Since condition (H3) ensures that the solution stays in a compact set around the initial condition (v, τ), we can limit ourselves to studying the function f on B R := {x R n ; x R} for R > 0 sufficiently large. We make use of the following decomposition theorem ([3], page 78). 25

Theorem 5.2. A function f : R n [a, b] R n satisfies (H1) (H3) if and only if f(x, t) = g(t) + h(x, t), where g : [a, b] R n is SKH integrable over [a, b] and h : R n [a, b] R n satisfies the Carathéodory conditions (C1) (C3). As an immediate result of this theorem we have that x F (x, τ, t) is continuous, since the Carathéodory conditions are exactly what we need to use the Lebesgue dominated convergence theorem for s h(x n, s) with x n x and get lim n τ h(x n, s) ds = τ h(x, s) ds. Since F (x, τ, t) is obviously continuous in τ and t and B R is compact, we easily get condition (E2). Therefore, the functions x F (x, τ, t) are uniformly close to the identity mapping for τ t small enough and from the topological degree theory we get (E3). In order to verify condition (E4), we apply the following theorem ([3], page 135). Theorem 5.3. Let h : R n [a, b] R n satisfy (C1) (C3) on B R [a, b]. Then there exists p : [a, b] R Lebesgue integrable and ω : [0, 2R] R + 0 and increasing with ω(0) = 0 such that 2 G(x, t 2 ) G(x, t 1 ) G(y, t 2 ) + G(y, t 1 ) ω( x y ) p(s) ds t 1 for x, y B R and t 1, t 2 [a, b], where G(x, t) is defined as We notice that and therefore G(x, t) = t 0 h(x, s) ds. F (x, τ, t) = x + G(x, t) G(x, τ) + (SKH) F (x, τ, t 2 ) F (x, τ, t 1 ) = G(x, t 2 ) G(x, t 1 ) + (SKH) F (y, σ, t 1 ) F (y, σ, t 2 ) = G(y, t 1 ) G(y, t 2 ) + (SKH) We arrive at the result τ g(s) ds, 2 t 1 t1 t 2 g(s) ds g(s) ds continuous F (x, τ, t 2 ) F (x, τ, t 1 ) F (y, σ, t 2 ) + F (y, σ, t 1 ) ω( x y ) ξ(t 2 ) ξ(t 1 ) where ξ is is defined as ξ(t) = a p(s) ds. 26

For any ε > 0 we can find η > 0 such that ω(η) < ε. Since ω is increasing, x y < η implies ω( x y ) < ε. We finish by applying a b a b to get that for x y < η we have F (x, τ, t 2 ) F (y, σ, t 2 ) F (x, τ, t 1 ) F (y, σ, t 1 ) + ε ξ(t 2 ) ξ(t 1 ) Therefore, F (x, τ, t) satisfies (E4) and the case of Theorem 5.1 is covered by Theorem 4.3. To see that Theorem 5.1 theorem is not contained in Theorem 4.1, consider the following example: { x 2 cos(π/x 2 ) 0 < x 1, H(x) = 0 x = 0. The derivative H (x) = { 2x cos(π/x 2 ) + (2π/x) sin(π/x 2 ) 0 < x 1, 0 x = 0, exists everywhere on the interval [0, 1]. Since H is SKH integrable, Theorem 5.1 can be applied to the the case f(x, t) = H (t) due to Theorem 5.2. However, the corresponding function G(x, t) = (SKH) 0 H (t) = H(t) does not satisfy the assumptions of Theorem 4.1, since H is not of bounded variation. While Theorem 5.1 is able to handle functions of unbounded variation, it remains within the bounds of the standard ordinary differential equation theory. It allows for the solution to be an indefinite SHK integral of the right hand side, but not in the sense of coupled variables. Therefore, the example involving the Cantor function is not covered. Thus, we can see that Theorem 4.1 and Theorem 5.1 are incomparable. However, our existence theorem 4.3 contains both a large portion of Theorem 4.1 and the entire Theorem 5.1. 27

Bibliography [1] J. Kurzweil, Generalized ordinary differential equations and continuous dependence on a parameter, Czechoslovak Math. J. 7 (82) 1957 418 449. [2] J. Kurzweil, Generalized ordinary differential equations, Series in Real Analysis, 11. World Scientific Publishing Co. Pte. Ltd., Hackensack, NJ, 2012. [3] S. Schwabik, Generalized ordinary differential equations, Series in Real Analysis, 5. World Scientific Publishing Co., Inc., River Edge, NJ, 1992. [4] J.-P. Aubin, A note on differential calculus in metric spaces and its applications to the evolution of tubes, Bull. Pol. Acad. Sci., Math. 40 (1992), no. 2, 151 162. [5] J.-P. Aubin, Mutational equations in metric spaces, Set-Valued Analysis, 1 (1993), no. 1, 3 46. [6] J.-P. Aubin, Mutational and Morphological Analysis: Tools for Shape Evolution and Morphogenesis, Birkhäuser Boston, Inc., Boston, MA, 1999. [7] T. Lorenz, Mutational Analysis: A Joint Framework for Cauchy Problems in and Beyond Vector Spaces, Lecture Notes in Mathematics, 1996. Springer- Verlag, Berlin, 2010. [8] T. Lorenz Mutational inclusions: differential inclusions in metric spaces, Discrete Contin. Dyn. Syst. Ser. B 14 (2010), no. 2, 629 654. [9] Jacek Tabor, Differential equations in metric spaces, Proceedings of EQUADIFF, 10 (Prague, 2001). Math. Bohem. 127 (2002), no. 2, 353 360. [10] J. Malý, H. Bendová, An elementary way to introduce a Perron-like integral, Ann. Acad. Sci. Fenn. Math. 36 (2011), no. 1, 153 164. [11] R. Henstock, Lectures on the theory of integration, Series in Real Analysis, 1. World Scientific Publishing Co., Singapore, 1988. [12] B. Thomson, Real functions, Lecture Notes in Mathematics, 1170. Springer- Verlag, Berlin, 1985. 28