Pathwise integration with respect to paths of finite quadratic variation Anna Ananova, Rama Cont

To cite this version:

Anna Ananova, Rama Cont. Pathwise integration with respect to paths of finite quadratic variation. 2016. ￿hal-01286515v3￿

HAL Id: hal-01286515 https://hal.archives-ouvertes.fr/hal-01286515v3 Preprint submitted on 29 Aug 2016

HAL is a multi-disciplinary open access L’archive ouverte pluridisciplinaire HAL, est archive for the deposit and dissemination of sci- destinée au dépôt et à la diffusion de documents entific research documents, whether they are pub- scientifiques de niveau recherche, publiés ou non, lished or not. The documents may come from émanant des établissements d’enseignement et de teaching and research institutions in France or recherche français ou étrangers, des laboratoires abroad, or from public or private research centers. publics ou privés. Pathwise integration with respect to paths of finite quadratic variation

Anna ANANOVA and Rama CONT

Revised: August 2016. To appear in: Journal de Math´ematiques Pures et Appliqu´ees.

Abstract We study a pathwise integral with respect to paths of finite quadratic variation, defined as the limit of non-anticipative Riemann sums for gradient-type integrands. We show that the integral satisfies a path- wise isometry property, analogous to the well-known Ito isometry for stochastic integrals. This property is then used to represent the inte- gral as a continuous map on an appropriately defined vector space of integrands. Finally, we obtain a pathwise ’signal plus noise’ decompo- sition for regular functionals of an irregular path with non-vanishing quadratic variation, as a unique sum of a pathwise integral and a com- ponent with zero quadratic variation. A. Ananova & R. Cont (2016) Pathwise integration with respect to paths of finite quadratic variation, Journal de Math´ematiquesPures et Appliqu´ees.

Contents

1 Non-anticipative functional calculus and pathwise integra- tion of gradient functionals 3

2 Isometry property of the pathwise integral 8 2.1 Pathwise isometry formula ...... 8 2.2 Partitions of hitting times ...... 14 2.3 The pathwise integral as a continuous isometry ...... 16

3 Pathwise nature of the integral 18

4 Rough-smooth decomposition for functionals of irregular paths 23

1 In his seminal paper ’Calcul d’Ito sans probabilit´es’[12], Hans F¨ollmer proved a change of variable formula for smooth functions of paths with infinite variation, using the concept of quadratic variation along a sequence 0 of partitions. A path ω ∈ C ([0,T ], R) is said to have finite quadratic m m m variation along the sequence of partitions π = (0 = t0 < t1 < ··· < m tk(m) = T ) if for any t ∈ [0,T ], the limit

X m m 2 [ω]π(t) := lim (ω(t ) − ω(t )) < ∞ m→∞ i+1 i m ti ≤t exists and defines a continuous increasing function [ω]π : [0,T ] → R+ called the quadratic variation of ω along π. Extending this definition to vector- valued paths (see Section 1) F¨ollmer[12] showed that, for integrands of the Z 2 d form ∇f(ω(t)), f ∈ C (R ) one may define a pathwise integral ∇f(ω(t))dω(t) as a pointwise limit of Riemann sums along the sequence of partitions πm. This construction has been extended in various directions, to less regular functions [1, 9, 23] and path-dependent functionals [4, 7, 8]. In particular, Cont & Fourni´e[4] constructed pathwise integrals of the type

Z T π ∇ωF (t, ω)d ω 0 where ∇ωF is a directional derivative (Dupire derivative) of a non-anticipative functional F , and proved a change of variable formulas for such integrals. This paper contributes several new results to the study of the pathwise approach to stochastic integration [12] and its extension to path-dependent functionals [4]. Our main result is an isometry formula for the pathwise integral: we give 0 conditions on the integrand φ and the path ω ∈ C ([0,T ], R) under which Z . Z t π 2 [ φd ω]π(t) = |φ| d[ω]π. 0 0 Theorem 2.1 gives a precise statement of this property for multidimensional paths. Our conditions notably allow for continuous paths with H¨older ex- ponent strictly less than 1/2 and apply to typical paths of and diffusion processes. This result may be understood as a pathwise version of the well-known Ito isometry formula [19], showing that the Ito isometry formula holds not just as an equality in expectation but in fact path by path.

2 This result has several interesting consequences. First, it implies that the pathwise integral may be defined as a a continuous mapping for the ’quadratic variation’ metric on an appropriate space of integrands (Propo- sition 2.8). This continuity property, together with its interpretation as a limit of Riemann sums, is what distinguishes our construction from other pathwise constructions of stochastic integrals [20, 21, 23, 25] which lack ei- ther a continuity property or an interpretation as a limit of non-anticipative Riemann sums. A second consequence of the pathwise isometry property is a determin- istic ’signal plus noise’ decomposition for functionals of an irregular path: we show that any regular functional of pathω ¯ with non-vanishing quadratic variation may be uniquely decomposed as the sum of a pathwise integral with respect toω ¯ and a ’smooth’ component with zero quadratic varia- tion (Proposition 4.1). This may be seen as a pathwise analogue of the well-known decomposition [10] (or, more precisely the de- composition of Dirichlet processes [13]). A similar result was obtained by Hairer and Pillai [18] in the context of rough path theory, using different techniques. Finally, we clarify the pathwise nature of the integral defined in [4]: we show (Theorem 3.2) that this integral is indeed a pathwise limit of non- anticipative Riemann sums, which is important for interpretation and in applications. Outline Section 1 recalls some key definitions and results on functional calculus from [4, 3] and recalls the definition of the F¨ollmerintegral [12] and its extension to path-dependent integrands by Cont & Fourni´e[4]. The ’isometry formula’ for this integral is derived in Section 2 (Theorems 2.1). In Section 2.2 we discuss the ’isometry formula’ for Lebesgue partitions. We use the of Section 2 in Section 2.3 to represent the integral as a continuous map (Proposition 2.8). In Section 3 we clarify the ’pathwise’ nature of this integral (Theorem 3.2) . Section 4 derives a pathwise ’signal plus noise’ de- composition for functionals of irregular paths (Proposition 4.1), which may be viewed as a deterministic analogue of the semimartingale decomposition for stochastic processes.

1 Non-anticipative functional calculus and path- wise integration of gradient functionals

Our approach relies on the Non-anticipative Functional Calculus [4, 3], a functional calculus which applies to non-anticipative functionals of c`adl`ag

3 paths with finite quadratic variation, in the sense of F¨ollmer[12]. We recall here some key concepts and results of this approach, following [3]. d 0 Let X be the canonical process on Ω = D([0,T ], R ), and (Ft )t∈[0,T ] be the filtration generated by X. We are interested in causal, or non- d anticipative functionals of X, that is, functionals F : [0,T ]×D([0,T ], R ) 7→ R such that ∀ω ∈ Ω,F (t, ω) = F (t, ωt), (1) where ωt := ω(t ∧ ·) is the stopped path of ω at time t. The process t 7→ 0 F (t, Xt) then only depends on the path of X up to t and is (Ft )-adapted. It is convenient to define such functionals on the space of stopped paths d [3]: a stopped path is an equivalence class in [0,T ] × D([0,T ], R ) for the following equivalence relation:

(t, ω) ∼ (t0, ω0) ⇐⇒ t = t0 and ω(t ∧ .) = ω0(t ∧ .)  . (2)

d The space of stopped paths is defined as the quotient of [0,T ]×D([0,T ], R ) by the equivalence relation (2):

d  d  ΛT = {(t, ω(t∧·)), (t, ω) ∈ [0,T ]×D([0,T ], R )} = [0,T ] × D([0,T ], R ) / ∼

We denote WT the subset of ΛT consisting of continuous stopped paths. We endow the set ΛT with a metric space structure by defining the following distance:

0 0 0 0 0 0 0 d∞((t, ω), (t , ω )) = sup |ω(u ∧ t) − ω (u ∧ t )| + |t − t | := kωt − ωt0 k∞ + |t − t | u∈[0,T ]

d (ΛT , d∞) is then a complete metric space. Any map F : [0,T ]×D([0,T ], R ) → R verifying the causality condition (1) can be equivalently viewed as a func- tional on the space ΛT of stopped paths. d We call non-anticipative functional on D([0,T ], R ) a measurable map 0,0 d F : (ΛT , d∞) → R and denote by C (ΛT ) the set of such maps which are continuous for the d∞) metric. Some weaker notions of continuity for non-anticipative functionals turn out to be useful [5]: Definition 1.1. A non-anticipative functional F is said to be: • continuous at fixed times if for any t ∈ [0,T ], F (t, ·): D([0,T ], k · d 0 k∞) → R in [0,T ], i.e. ∀ω ∈ D([0,T ], R ), ∀ > 0, ∃η > 0, ∀ω ∈ d D([0,T ], R ), 0 0 kωt − ωtk∞ < η =⇒ |F (t, ω) − F (t, ω )| < 

4 0 0 • left-continuous if ∀(t, ω) ∈ ΛT , ∀ > 0, ∃η > 0 such that ∀(t , ω ) ∈ ΛT , 0 0 0  0 0 (t < t and d∞ (t, ω), (t , ω ) < η) =⇒ |F (t, ω) − F (t , ω )| < 

0,0 d We denote by Cl (ΛT ) the set of left-continuous functionals. Simi- 0,0 d larly, we can define the set Cr (ΛT ) of right-continuous functionals. We also introduce a notion of local boundedness for functionals: Definition 1.2. A non-anticipative functional F is said to be boundedness- d preserving if for every compact subset K of R , ∀t0 ∈ [0,T ], ∃C(K, t0) > 0 such that:

∀t ∈ [0, t0], ∀(t, ω) ∈ ΛT , ω([0, t]) ⊂ K =⇒ |F (t, ω)| < C(K, t0).

d We denote by B(ΛT ) the set of boundedness-preserving functionals. We now recall some notions of differentiability for functionals following d d [4, 3, 11]. For e ∈ R and ω ∈ D([0,T ], R ), define the vertical perturbation e ωt of (t, ω) as the path obtained by shifting the path by e at t: e ωt = ωt + e1[t,T ]. Definition 1.3. A non-anticipative functional F is said to be:

• horizontally differentiable at (t, ω) ∈ ΛT if F (t + h, ω ) − F (t, ω ) DF (t, ω) = lim t t (3) h→0+ h

exists. If DF (t, ω) exists for all (t, ω) ∈ ΛT , then (3) defines a non- anticipative functional DF , called the horizontal derivative of F .

• vertically differentiable at (t, ω) ∈ ΛT if the map:

d g(t,ω) : R −→ R e 7→ F (t, ωt + e1[t,T ]) is differentiable at 0. Its gradient at 0 is called the Dupire derivative (or vertical derivative) of F at (t, ω):

d ∇ωF (t, ω) = ∇g(t,ω)(0) ∈ R (4)

i.e. ∇ωF (t, ω) = (∂iF (t, ω), i = 1, ··· , d) with

F (t, ωt + hei1[t,T ]) − F (t, ωt) ∂iF (t, ω) = lim h→0 h

5 d where (ei, i = 1, ··· , d) is the canonical basis of R . If F is vertically d d differentiable at all (t, ω) ∈ ΛT , ∇ωF :ΛT → R defines a non- anticipative functional called the vertical derivative of F .

2 We may repeat the same operation on ∇ωF and define similarly ∇ωF , 3 ∇ωF , etc. Remark 1. Note that DF (t, ω) is not the partial derivative in t:

F (t + h, ω) − F (t, ω) DF (t, ω) 6= ∂tF (t, ω) = lim . h→0 h

∂tF (t, ω) corresponds to a ’Lagrangian’ derivative which follows the incre- ment of F along the path from t to t + h, whereas in (3) the path is stopped at t and the increment is taken along the stopped path.

We consider the following classes of smooth functionals:

1,k d Definition 1.4. We define Cb (ΛT ) as the set of non-anticipative function- d als F : (ΛT , d∞) → R which are • horizontally differentiable with DF continuous at fixed times;

j 0,0 d • k times vertically differentiable with ∇ωF ∈ Cl (ΛT ) for j = 0, ··· , k; k d •DF, ∇ωF, ··· , ∇ωF ∈ B(ΛT ). n n n n Consider now a sequence π = (0 = t0 < t1 < ··· < tk(n) = T ) of n n n partitions of [0,T ]. |π | = sup{|ti+1 − ti |, i = 1 ··· k(n)} → 0 will denote the mesh size of the partition. A c`adl`agpath x ∈ D([0,T ], R) is said to have n finite quadratic variation along the sequence of partitions (π )n≥1 if for any t ∈ [0,T ] the limit

X n n 2 [x]π(t) := lim (x(t ) − x(t )) < ∞ (5) n→∞ i+1 i n ti+1≤t exists and the increasing function [x] has Lebesgue decomposition

c X 2 [x]π(t) = [x]π(t) + |∆x(s)| 0

c where [x]π is a continuous, increasing function. We denote the set of such paths Qπ([0,T ], R).

6 1 d d A d-dimensional path x = (x , ..., x ) ∈ D([0,T ], R ) is said to have finite i i j quadratic variation along π if x ∈ Qπ([0,T ], R) and x + x ∈ Qπ([0,T ], R) for all i, j = 1..d. Then for any i, j = 1, ··· , d and t ∈ [0,T ], we have

i j i j X i n i n j n j n n→∞ ij [x + x ]π(t) − [x ]π(t) − [x ]π(t) (x (tk+1)−x (tk )).(x (tk+1)−x (tk )) → [x]π (t) = . n n n 2 tk ∈π ,tk ≤t

+ The matrix-valued function [x] : [0,T ] → Sd whose elements are given by [xi + xj] (t) − [xi] (t) − [xj] (t) [x]ij(t) = π π π π 2 is called the quadratic covariation of the path x along the sequence of par- titions π. For further discussion of this concept, we refer to [3]. d Consider now a path ω ∈ Qπ([0,T ], R ) with finite quadratic variation along π. Since ω has at most a countable set of jump times, we may assume that the partition ’exhausts’ the jump times in the sense that

sup |ω(t) − ω(t−)| n→∞→ 0. (6) t∈[0,T ]−πn

Then the piecewise-constant approximation

k(n)−1 n X ω (t) = ω(ti+1−)1[ti,ti+1[(t) + ω(T )1{T }(t) (7) i=0 converges uniformly to ω: sup kωn(t) − ω(t)k → 0. Note that with the t∈[0,T ] n→∞ n n n n n n notation (7), ω (ti −) = ω(ti −) but ω (ti ) = ω(ti+1−). If we define

n n n,∆ω(ti ) n n n,∆ω(ti ) n n ω n = ω + ∆ω(t )1[tn,T ], then ω n (t ) = ω(t ). ti i i ti − i i Approximating the variations of the functional by vertical and horizontal increments along the partition πn, we obtain the following pathwise change 1,2 d of variable formula for C (ΛT ) functionals, derived in [4] under more gen- eral assumptions:

1,2 Theorem 1.5 (Pathwise change of variable formula for C functionals [4]). d 1,2 d Let ω ∈ Qπ([0,T ], R ) verifying (6). Then for any F ∈ Cb (ΛT ) the limit

k(n)−1 Z T n π X  n n,∆ω(ti ) n n ∇ωF (t, ωt−)d ω := lim ∇ωF ti , ω n .(ω(ti+1) − ω(ti )) (8) n→∞ ti − 0 i=0

7 exists and Z T Z T 1 t 2 c  F (T, ωT ) − F (0, ω0) = DF (t, ωt)dt + tr ∇ωF (t, ωt−)d[ω]π(t) 0 0 2 Z T π X + ∇ωF (t, ωt−)d ω + (F (t, ωt) − F (t, ωt−) − ∇ωF (t, ωt−).∆ω(t)) . 0 t∈]0,T ]

A consequence of this theorem is the ability to define the pathwise in- Z . π tegral ∇ωF (t, ωt−)d ω as a limit of ”Riemann sums” computed along 0 0 d the sequence of partitions π. For a continuous path ω ∈ C ([0,T ], R ) this simplifies to:

Z T k(n)−1 π X  n n  n n ∇ωF (t, ω)d ω := lim ∇ωF ti , ωtn .(ω(ti+1) − ω(ti )) (9) n→∞ i 0 i=0 This integral, first constructed in [4], extends H. F¨ollmer’sconstruction [12] 2 d for integrands of the form ∇f, f ∈ C (R ) to (path-dependent) integrands of the form 1,2 d ∇ωF,F ∈ Cb (ΛT ). This construction applies to typical paths of Brownian motion and semi- martingales, which satisfy the quadratic variaton property almost-surely, although they have infinite p-variation for p = 2 so (9) cannot be constructed as a Young integral [27]. The goal of this paper is to explore some properties of this integral.

2 Isometry property of the pathwise integral

2.1 Pathwise isometry formula d Let ω ∈ Qπ([0,T ], R ) be a given path with finite quadratic variation along a n nested sequence of partitions π = (π )n≥1. Our goal is to provide conditions under which the the following identity holds:

t Z t [F (·, ω·)]π (t) = h∇ωF (s, ωs−) ∇ωF (s, ωs−), d[ω]π(s)i, (10) 0 where we use the notation hA, Bi := tr(AB), for square matrices A, B.

8 This relation was first noted in [3] for functionals of the form F (ω) = Z t φ.dω with φ piecewise-constant. To extend this property to a more gen- 0 eral setting we assume a Lipschitz-continuity condition on the functional F and a H¨older-type regularity condition for the path ω. Assumption 1 (Uniform Lipschitz continuity).

0 d 0 0 ∃K > 0, ∀ω, ω ∈ D([0,T ], R ), ∀t ∈ [0,T ], |F (t, ω)−F (t, ω )| ≤ Kkωt−ωtk∞. d We denote by Lip(ΛT , k·k∞) the space of all non-anticipative functionals satisfying the above property. n n Consider a nested sequence of partitions π = {ti , i = 0, . . . , m(n)} of n d [0,T ]. We define the oscillation osc(f, π ) of a function f : [0,T ] → R between points of the partition πn as

n n osc(f, π ) = max sup |f(t) − f(tj )| j=0..m(n)−1 n n t∈(tj ,tj+1] We make the following assumptions on the sequence of partitions: Assumption 2 (Vanishing oscillation along π).

osc(ω, πn) n→→+∞ 0.

Assumption 3.

n n n→+∞ max |F (ti+1, ω) − F (ti , ω)| → 0. i∈{0,..., m(n)−1}

n d Remark 2 (Partitions with vanishing mesh). If |π | → 0 and F : (ΛT , d∞) → R is Lipschitz-continuous then Assumptions 2 and 3 are satisfied for any 0 d ω ∈ C ([0,T ], R ). To see this, set  > 0. By uniform continuity of ω on [0,T ] there exists δ() > 0 such that kω(t) − ω(s)k ≤  whenever |t − s| ≤ δ. Since |πn| → 0 n there exists N1 ≤ 1 such that |π | ≤  ∧ δ() for n ≤ N1. Then for n ≥ N1 n we have osc(ω, π ) ≤  so Assumption 2 holds. Furthermore, for n ≤ N1,

n n n n d ((t , ω), (t , ω)) = |t − t | + kω n − ω n k ∞ i+1 i i+1 i ti+1 ti ∞ n n ≤ |π | + sup |ω(t) − ω(ti )| ≤ 2 n n t∈[ti ,ti+1] n n so d∞((ti+1, ω), (ti , ω)) → 0 as n → ∞. Lipschitz-continuity of F then n n n→+∞ entails max |F (ti+1, ω) − F (ti , ω)| → 0. i∈{0,..., m(n)−1}

9 In section 2.2 we will also consider sequences of partitions whose mesh may not converge to zero. For such partitions Assumption 3 is not redun- dant. For 0 < ν < 1, denote by Cν([0,T ]) the space of ν−H¨oldercontinuous functions:

ν d 0 kf(t) − f(s)k C ([0,T ], R ) = {f ∈ C ([0,T ]), sup ν < +∞}, (t,s)∈[0,T ]2,t6=s |t − s|

ν− d α d and C ([0,T ], R ) = ∩ C ([0,T ], R ) 0≤α<ν the space of functions which are α−H¨olderfor every α < ν. For ω ∈ ν d C ([0,T ], R ), the following piecewise constant approximation

m(n)−1 n X n ω := ω(t −)1 n n + ω(T )1 (11) i+1 [ti ,ti+1) {T } i=0 n n ν satisfies: kω − ω k∞ ≤ |π | . Our main result is the following: n Theorem 2.1 (Pathwise Isometry formula). Let π = (π )n≥1 be a sequence√ 3 − 1 of partitions of [0,T ], and ω ∈ Q ([0,T ], d)∩Cν([0,T ], d) for ν > , π R R 2 satisfying Assumption 2. 1,2 d d 1,1 d Let F ∈ Cb (ΛT ) ∩ Lip(ΛT , k · k∞) with ∇ωF ∈ Cb (ΛT ) such that Assumption 3 holds. Then:

T Z .  Z π t [F (·, ω·)]π (T ) = ∇ωF (s, ωs)d ω (T ) = h∇ωF (s, ωs) ∇ωF (s, ωs), d[ω]πi. 0 π 0 (12)

Remark 3. We note that for typical paths of a Brownian diffusion or con- tinuous semimartingale with non-degenerate component, the assumptions of Theorem 2.1 hold almost surely as soon as Assumption 1 is 1 − d satisfied. More generally the result holds for ω ∈ C 2 ([0,T ], R ), which corresponds to the H¨olderregularity of Brownian paths [24]. We start with an important lemma which is essential in proving both the pathwise isometry formula and the uniqueness of pathwise integral. This result also establishes a connection between the functional derivatives intro- duced in Section 1 and controlled rough paths [17].

10 ν d Lemma 2.2. Let ω ∈ C ([0,T ], R ) for some ν ∈ (1/3, 1/2] and F ∈ 1,2 d n 1,1 d n×d d Cb (ΛT , R ) with ∇ωF ∈ Cb (ΛT , R ) and F ∈ Lip(ΛT , k · k∞). De- fine F Rs,t(ω) := F (s, ωs) − F (t, ωt) − ∇ωF (t, ωt)(ω(s) − ω(t)). (13)

Then there exists a constant CF,T,kωkν increasing in T and kωkν, such that

F 2βν |Rs,t(ω)| ≤ CF,T,kωkν |s − t| , with 2βν = ν(1 + ν). Proof. We will prove only the case when the values of F are scalar, i.e. n = 1; the extension to the general case is straightforward. We start by recalling the following result from [3, Proposition 5.26]: 1,1 d Lemma 2.3. Let G ∈ Cb (ΛT ) and λ :[t, s] → R a continuous function with finite variation. Then Z s Z s G(s, λs) − G(t, λt) = DG(u, λu)du + ∇ωG(u, λu)dλ(u). (*) t t Here the integrals are defined in the Riemann-Stieltjes sense. d Now, let us fix a Lipschitz continuous path λ: [0,T ] → R , using the above lemma repeatedly for G = F, ∇ωF, ··· we will obtain an integral F formula for Rt,s(λ) in terms of the derivatives F . For the sake of convenience i we denote by ∂iF and λ respectively the i-th coordinates of ∇ωF and λ. We also use Einstein’s convention of summation in repeated indexes and the following notation δλt,s := λ(s) − λ(t). Using (*) for G = F , we have

Z s Z s F i Rt,s(λ) = DF (u, λu)du + (∂iF (u, λu) − ∂iF (t, λt)) dλ (u) (14) t t For the second term on the right-hand side of the above identity we use the (*) with G = ∂iF and then Fubini’s theorem to get Z s Z s Z u i i (∂iF (u, λu) − ∂iF (t, λt)) dλ (u) = D∂iF (r, λr)drdλ (u) t t t Z s Z u 2 j i + ∂ijF (r, λr)dλ (r)dλ (u) t t Z s Z s i i 2 ˙ j i i = D∂iF (r, λr)(λ (s) − λ (r))dr + ∂ijF (r, λr)λ (r)(λ (s) − λ (r))dr. t t (15)

11 Thus Z s Z s F i 2 i j Rt,s(λ) = (DF (u, λu) + D∂iF (u, λu)δλu,s)du + ∂ijF (r, λr)δλr,sdλ (r) t t (16) To use the above formula for estimate the error term RF for H¨older continuous paths we will use a piecewise linear approximation. For ω ∈ Cν let ωN be the piecewise linear approximation of ω on [t, s] defined by • ωN (r) = ω(r), ∀r ∈ [0, t],

s − t • ωN (τ N ) = ω(τ N ), k = {0, ..., N}, where τ N = t + k , k k k N

N N N • ω is linear on each interval [τk , τk+1]. Then, we have

|s − t|ν kωN − ωk ≤ Ckωk , kωN k ≤ Ckωk , ∞ ν N ν ν ν

X |δωτi,τi+1 | |ω˙ N | = 1 ≤ N 1−νkωk |s − t|ν−1, (17) |τ − τ | (τi,τi+1) ν i i+1 i ν |δωa,b| ≤ kωkν|b − a| . 2 Using the local boundedness property of DF, D∇ωF and ∇ωF , and the representation (16) for λ = ωN , we obtain

F N N 1+ν N 2 1−ν 2ν |Rt,s(ω )| ≤ CF |s − t| + CF kω kν|s − t| + CF kω kνN |s − t| . (18)

N N On the other hand since ωt = ωt and ω (s) = ω(s) F N F N N −ν ν |Rt,s(ω )−Rt,s(ω)| = |F (s, ω )−F (s, ω)| ≤ CF kω −ωk∞ ≤ CF kωkνN |s−t| . (19) from above estimates (18), (19) and triangle inequality F 1+ν 2 1−ν 2ν −ν ν |Rt,s(ω)| ≤ CF |s−t|+CF kωkν|s−t| +CF kωkνN |s−t| +N |s−t| . 2 1−ν 2ν To optimize the above bound, we choose N so that kωkνN |s − t| ≈ −ν ν −2 −ν N |s − t| i.e. N ≈ kωkν |s − t| . Hence the result F 2ν ν+ν2 |Rt,s(ω)| ≤ CF,T (1 + kωkν)|s − t| + CF,νkωkν |s − t| .

12 We are now ready to prove the isometry property:

Proof of Theorem 2.1. To prove the isometry formula we note that

2  n n  t t F (t , ωtn ) − F (t , ωtn ) − h∇ωF (t, ωtn ) ∇ωF (t, ωtn ), δωtn,tn δωtn,tn i i+1 i+1 i i i i i i+1 i i+1 F 2 F ≤ |R n n (ω)| + CF |R n n (ω)||δωtn,tn | ti ,ti+1 ti ,ti+1 i i+1

F From the Assumption 2 it follows that Mn := max Rtn,tn (ω) →n→∞ 0 i i i+1 F n n ν2+ν and from part a) of Theorem (3.1) we have R n n (ω) ≤ C|t − t | . ti ,ti+1 i+1 i √ 1 Since ν > ( 3 − 1)/2 is equivalent to ν2 + ν > , we get 2

2 2− 1 2− 1 X F ν2+ν X n n ν2+ν R n n (ω) ≤ CMn |t − t | ≤ CTMn → 0. ti ,ti+1 i+1 i i i

X 2 π Consequently, since |δω n n | → tr([ω] ), we have by Cauchy -Schwarz ti ,ti+1 i inequality

s 2s X F X F X 2 |R n n (ω)||δωtn,tn | ≤ R n n (ω) |δωtn,tn | → 0. ti ,ti+1 i i+1 ti ,ti+1 i i+1 i i i

We conclude

2 X  n n  X t t F (t , ωtn ) − F (t , ωtn ) − h∇ωF (t, ωtn ) ∇ωF (t, ωtn ), δωtn,tn δωtn,tn i i+1 i+1 i i i i i i+1 i i+1 i i X F 2 X F ≤ |R n n (ω)| + CF |R n n (ω)||δωtn,tn | → 0 ti ,ti+1 ti ,ti+1 i i+1 i i which together with

T Z X t t t h∇ F (t, ω n ) ∇ F (t, ω n ), δω n n δω n n i → h∇ F (s, ω ) ∇ F (s, ω ), d[ω] i. ω ti ω ti ti ,ti+1 ti ,ti+1 ω s ω s π i 0 implies

T Z .  Z π t ∇ωF (s, ωs)d ω (T ) = h∇ωF (s, ωs) ∇ωF (s, ωs), d[ω]πi. 0 π 0

13 Z .  π To prove the equality [F (·, ω·)]π (T ) = ∇ωF (s, ωs)d ω (T ) we 0 π note, using the change of variable formula (Theorem (1.5)) it is enough to Z · Z · 2 prove that the paths DF (t, ωt)dt, ∇ωF (t, ωt−), d[ω]π(t) have zero 0 0 quadratic variation along πn. For the second path it follows from Assump- n n tion 2 as it implies that max |[ω](ti+1) − [ω](ti )| → 0. For the first path note i that since it is obviously a path of bounded variation we just need to prove tn Z i+1 that max DF (t, ωt)dt → 0. For that we write i n ti n n n Z ti+1 Z ti+1 Z ti+1 DF (t, ωt)dt = [DF (t, ωt) − DF (t, ωtn )]dt + DF (t, ωtn )dt n n i n i ti ti ti n Z ti+1 n n = [DF (t, ωt) − DF (t, ωtn )]dt + F (ti+1, ωtn ) − F (ti , ωtn ) n i i i ti n Z ti+1  n n  = [DF (t, ωt) − DF (t, ωtn )]dt + F (ti+1, ωtn ) − F (ti+1, ωtn ) n i i i+1 ti  n n  + F (t , ω n ) − F (t , ω n ) . i+1 ti+1 i ti By the continuity of F, DF and Assumption 2 the first two summands in the right hand side of above equality uniformly converge to 0 as n → ∞, by Assumption 3 the third summand also converges uniformly to 0.

Remark 4 (Relation with Ito isometry formula). Let P be the Wiener mea- Z t 0 π sure on C ([0,T ], R). Then the pathwise integral (t, ω) → ∇ωF (u, ω)d ω 0 is defined P−almost surely and defines a process which is a version of the Z . Ito integral ∇ωF (t, W )dW . Integrating (12) with respect to P yields the 0 well-known Ito isometry formula [19]:  Z .   Z T  Z T  P P 2 P 2 E [ ∇ωF (t, W )dW ](T ) = E | ∇ωF (t, W )dW | = E |∇ωF (t, W )| dt . 0 0 0 So, Theorem 2.1 reveals that underlying the Ito isometry formula is the pathwise isometry (12) which does not rely on the Wiener measure.

2.2 Partitions of hitting times An alternative approach to the definition of the pathwise integral is to con- sider quadratic variation and Riemann sums computed along ’Lebesgue par-

14 titions’ i.e. partitions of hitting times of uniformly spaced levels [2, 9, 20, 22]:

n n n n −n τ0 (ω) = 0, τk+1(ω) = inf{t > τk , |ω(t) − ω(τk )| ≥ 2 } ∧ T. (20) Although this is different in spirit from the original approach of F¨ollmer, there are some arguments for using such ’intrinsic’ sequences of partitions for a given path, one of them being that quadratic variation computed along (20) is invariant under a time change. Thus a natural question is whether the results above apply to such partitions. We will now present a version of Theorem 2.1 for the sequence of partitions defined by (20). n n n Define m(n) = inf{k ≥ 1, τk (ω) = T } and denote τ (ω) = (τk (ω), k = 0, . . . , m(n)). It is easy to see that the vanishing oscillation conditon (As- sumption 2) automatically holds for the sequence (τ n(ω), n ≥ 1) since by construction osc(ω, τ n(ω)) ≤ 2−n. However, as the example of a constant path shows, for a general path ω ∈ 0 d n C ([0,T ], R ), the partitions τ (ω) may fail to have a vanishing mesh size. We will show now that, if a path ω has strictly increasing quadratic variation along its Lebesgue partition τ n(ω) then |τ n| → 0. 0 d Lemma 2.4. Assume ω ∈ C ([0,T ], R ) has finite quadratic variation along the Lebesgue partition τ n(ω) defined by (20):

X n n 2 ∀t ∈ (0,T ], lim |ω(τ ) − ω(τ )| = [ω]τ(ω)(t) > 0. (21) n→∞ k+1 k n τk ≤t n If t → [ω]τ(ω)(t) is a strictly increasing function then |τ (ω)| → 0. Proof. Let h > 0, t ∈ [0,T − h]. Denote by m(n, t, t + h) the number of partition points of τ n(ω) in the interval [t, t + h]. Then

X n n 2 −n |ω(τk+1) − ω(τk )| = 4 (m(n, t, t + h) − 1). n t≤τk ≤t+h

Since t → [ω]τ(ω)(t) is strictly increasing,

X n n 2 −n |ω(τ )−ω(τ )| = 4 (m(n, t, t+h)−1) → [ω]τ(ω)(t+h)−[ω]τ(ω)(t) > 0 k+1 k n→∞ n t≤τk ≤t+h so in particular m(n, t, t + h) ∼ 4n → ∞. This holds for any interval t ∈ [0,T − h] which implies that the |τ n| < h for n large enough. Since this is true for any h > 0, this implies |τ n| → 0.

15 The condition that [ω]τ(ω) is strictly increasing is an ’irregularity’ condi- tion on ω: it means that the quadratic variation over any interval is non-zero. For paths verifying this condition, Theorem 2.1 may be extended to Lebesgue partitions:

Theorem 2.5 (Pathwise Isometry formula along Lebesgue partitions). Let √ ! 3 − 1 1 ω ∈ Q ([0,T ], d) ∩ Cν([0,T ], d) for some ν ∈ , and F ∈ τ(ω) R R 2 2 1,1 d d 1,1 d Cb (ΛT ) ∩ Lip(ΛT , k · k∞) such that ∇ωF ∈ Cb (ΛT ). If ω has strictly increasing quadratic variation [ω]τ(ω) with respect to τ(ω) then the isometry formula holds along the partition of hitting times defined by (20):

t Z .  Z π t [F (·, ω·)]τ(ω) (t) = ∇ωF (s, ωs−)d ω (t) = h∇ωF (s, ωs) ∇ωF (s, ωs), d[ω]τ(ω)i. 0 0 Proof. First note that the sequence of partitions (τ n(ω), n ≥ 1) automati- −n cally satisfies Assumption 2 since osc(ω, τn(ω)) ≤ 2 . n Since [ω]τ(ω) is strictly increasing, Lemma 2.4 implies that |τ | → 0. ν d Since ω ∈ C ([0,T ], R ) which together with the continuity of F implies that n n n→+∞ max |F (ti+1, ω) − F (ti , ω)| → 0. i∈{0,..., m(n)−1} so the sequence of partitions τ n(ω) satisfies Assumption 3. The result then follows from the Theorem 2.1.

2.3 The pathwise integral as a continuous isometry Theorem 2.1 suggests that the existence of an isometric mapping underlying the pathwise integral. We will now proceed to make this structure explicit. A consequence of Theorem 2.1 is that, if ω is an irregular path with strictly increasing quadratic variation, any regular functional F (., ω) has zero quadratic variation along π if and only if ∇ωF vanishes along ω:

d Proposition 2.6 (Preservation of irregularity). Let ω ∈ Qπ([0,T ], R ) ∩ √ # 3 − 1 1 d[ω]π Cν([0,T ], d) for some ν ∈ , such that := a(t) > 0 is a R 2 2 dt 1,2 d d right-continuous, positive definite function, and F ∈ Cb (ΛT ) ∩ Lip(ΛT , k · 1,1 d k∞) be a non-anticipative functional with ∇ωF ∈ Cb (ΛT ).

16 If Assumptions 2 and 3 hold, then the path t 7→ F (t, ω) has zero quadratic variation along the partition π if and only if ∇ωF (t, ω) = 0, ∀t ∈ [0,T ]. Proof. Indeed, from Theorem 2.1

T Z t [F (·, ω·)] (T ) = ∇ωF (s, ω)a(s)∇ωF (s, ω) ds. 0 Since a(t) is positive definite the integrand on the right hand side is non- negative and strictly positive unless ∇ωF (s, ωs) = 0. So by right-continuity of the integrand the integral is zero if and only if ∇ωF (·, ω·) ≡ 0. 1,2 Let F ∈ Cb (ΛT ) be a functional satisfying assumptions of Theorem 2.1. We denote R(ΛT ) the set of such regular functionals. Proposition 2.6 may be understood in the following way: if a path is ’irregular’ in the sense of having strictly increasing quadratic variation, this property is locally preserved by any regular functional transformation F as long as ∇ωF does not vanish. 1 − d 1 − We note that for ω ∈ C 2 ([0,T ], R ), we have F (·, ω) ∈ C 2 ([0,T ], R). d Let a: [0,T ] → S+ be a continuous function taking values in positive- definite symmetric matrices.

Definition 2.7 (Harmonic functionals). F ∈ R(ΛT ) is called a−harmonic if 1 ∀(t, ω) ∈ Λ , DF (t, ω ) + h∇2 F (t, ω ), a(t)i = 0. T t 2 ω t

We denote Ha(ΛT ) the space of a(·)-harmonic functionals.

Note that for any F ∈ R(ΛT ) the functional defined by Z t Z t 1 2 G(t, ω) = F (t, ω) − DF (s, ω)ds − h∇ωF (s, ω), a(s)ids 0 2 0 is a(·)-harmonic. The class of harmonic functionals plays a key role in probabilistic inter- pretation of the Functional Ito calculus [14, 6, 3]. We will now see that this class also plays a role in the extension of the pathwise integral. d ν d Letω ¯ ∈ Qπ([0,T ], R ) ∩ C ([0,T ], R ) such that d[¯ω]π/dt = a. The functional change of variable formula (Theorem 1.5 ) then implies that

Z t π ∀F ∈ Ha(ΛT ), ∀t ∈ [0,T ],F (t, ω¯) = F (0, ω¯) + ∇ωF (u, ω¯)d ω.¯ 0

17 By Theorem 2.1, we have Z t t 2 [F (., ω¯)]π(t) = ∇ωF (u, ω¯)a(u)∇ωF (u, ω¯)du = k∇ωF (., ω¯)kL2([0,T ],a) < ∞. 0 Consider the vector spaces  Ha(¯ω) := F (·, ω¯·) F ∈ Ha(ΛT ) ⊂ Qπ([0,T ], R),  2 Va(¯ω) := ∇ωF (·, ω¯·) F ∈ Ha(ΛT ) ⊂ L ([0,T ], a). which are the images of ω under all harmonic functionals and their vertical derivatives. Proposition 2.6 then implies that the map p ω ∈ Ha(¯ω) → kωkπ = [ω]π(T ) defines a norm on Ha(¯ω). The pathwise integral can then be lifted to a continuous map  Iω¯ : Va(¯ω), k · kL2([0,T ],a) → (Ha(¯ω), k · kπ) Z · π φ = ∇ωF (., ω¯) → Iω¯ (φ) := φd ω.¯ (22) 0 which is in fact an isometry:  Proposition 2.8 (Isometry property). Iω¯ : Va(¯ω), k · kL2([0,T ],a) → (Ha(¯ω), k · kπ) is an injective isometry.

3 Pathwise nature of the integral Z π The definition (9) of the integral ∇ωF (., ω)d ω involves values of the integrand ∇ F along piecewise constant approximations of ω. To be able Zω π to intepret ∇ωF (., ω)d ω as a pathwise integral, we must show that it only depends on the values ∇ωF (t, ω), t ∈ [0,T ] of the integrand along the path ω. Our goal is to provide conditions under which the following property holds: 1,2 d Property 1 (Pathwise nature of the integral). Let F1,F2 ∈ Cb (ΛT ), and ν ω ∈ Qπ([0,T ], R) ∩ C ([0,T ]) for some 0 < ν < 1/2. If ∇ωF1(t, ωt) = ∇ωF2(t, ωt), ∀t ∈ [0,T ] then Z T Z T π π ∇ωF1 · d ω = ∇ωF2 · d ω. 0 0

18 To give a precise statement, let us introduce the following assumption:

Assumption 4 (Horizontal local Lipschitz property). d A non-anticipative functional F :ΛT → V with values in a finite dimensional real vector space V , satisfies the horizontal locally Lipschitz property if

d 0 d ∀ω ∈ D([0,T ], R ), ∃ C > 0, η > 0, ∀h ≥ 0, ∀t ≤ T − h, ∀ω ∈ D([0,T ], R ),

0 0 0 kωt − ωtk∞ < η, ⇒ |F (t + h, ωt) − F (t, ωt)| ≤ Ch. The above assumption asserts the Lipschitz regularity of F in time as the value of the path ω is frozen. Note that this property is weaker than horizontal differentiability. We start with a useful lemma:

ν d Lemma 3.1 (Expansion formula for regular functionals). Let ω ∈ C ([0,T ], R ) 1,1 d n for some ν ∈ (1/3, 1/2] and F ∈ Cb (ΛT , R ) be a non-anticipative func- 1,1 d n×d 2 1,1 d n×d×d 3 tional such that ∇ωF ∈ Cb (ΛT , R ), ∇ωF ∈∈ Cb (ΛT , R ),F, DF, ∇ωF ∈ d 3 Lip(ΛT , k · k∞) and ∇ωF horizontally locally Lipschitz. Then

Z s F (s, ωs) − F (t, ωt) = ∇ωF (t, ωt)(ω(s) − ω(t)) + DF (u, ωu)du t 1 2 + h∇2 F (t, ω ), (ω(s) − ω(t)) t(ω(s) − ω(t))i + O(|s − t|3ν +ν), 2 ω t as |s − t| → 0, uniformly in t, s ∈ [0,T ].

Proof. For the result we will need to further expand the representation (16). First, note that the second term in the third line of (15) can be written as Z s Z u Z s Z u 2 j i 2 j i ∂ijF (r, λr)dλ (r)dλ (u) = ∂ijF (t, λt) dλ (r)dλ (u) t t t t Z s Z u (23) 2 2  j i + ∂ijF (r, λr) − ∂ijF (t, λt) dλ (r)dλ (u) t t

2 Using the Lemma (*) for G = ∂ijF and then Fubini’s theorem, we get

19 Z s Z u 2 2  j i ∂ijF (r, λr) − ∂ijF (t, λt) dλ (r)dλ (u) t t Z s Z u Z r 2 j i = D∂ijF (τ, λτ )dτdλ (r)dλ (u) t t t Z s Z u Z r (24) 3 j i + ∂ijkF (τ, λτ )dλk(τ)dλ (r)dλ (u) t t t Z s Z s 2 ji 3 ji ˙ k = D∂ijF (τ, λτ )Λτ,sdτ + ∂ijkF (τ, λτ )Λτ,sλ (τ)dτ, t t where Z b Z u Z b ji j i j j  ˙ i Λa,b = dλ (r)dλ (u) = λ (u) − λ (a) λ (u)du. a a a Combining (14), (15), (23) and (24) we obtain Z s Z s F i 2 ji Rt,s(λ) = DF (u, λu)du + D∂iF (r, λr)δλr,sdr + ∂ijF (t, λt)Λt,s t t Z s Z s (25) 2 ji 3 ji ˙ k + D∂ijF (τ, λτ )Λτ,sdτ + ∂ijkF (τ, λτ )Λτ,sλ (τ)dτ. t t Finally exploiting the symmetry of second and third order derivatives: 2 2 3 3 ∂ijF = ∂jiF , ∂ijkF = ∂jikF , from (25) we obtain:

Z s Z s F i 2 ˜ ji Rt,s(λ) = DF (u, λu)du + D∂iF (r, λr)δλr,sdr + ∂ijF (t, λt)Λt,s t t Z s Z s (26) 2 ˜ ji 3 ˜ ji ˙ k + D∂ijF (τ, λτ )Λτ,sdτ + ∂ijkF (τ, λτ )Λτ,sλ (τ)dτ, t t where Λ˜ is the symmetric part of matrix Λ. By integration by parts it is easy to compute

ji ij Λ + Λ 1 Λ˜ ji = a,b a,b = (λj(b) − λj(a))(λi(b) − λi(a)). a,b 2 2 Therefore, Z s Z s F i 1 2 i j Rt,s(λ) = DF (u, λu)du + D∂iF (r, λr)δλr,sdr + ∂ijF (t, λt)δλt,sδλt,s t t 2 Z s Z s 1 2 i j 1 3 i j ˙ k + D∂ijF (τ, λτ )δλτ,sδλτ,sdτ + ∂ijkF (τ, λτ )δλτ,sδλτ,sλ (τ)dτ, 2 t 2 t (27)

20 we decompose the last term in the above as follows Z s Z s 3 i j ˙ k 3 i j ˙ k ∂ijkF (τ, λτ )δλτ,sδλτ,sλ (τ)dτ = ∂ijkF (t, λt)δλτ,sδλτ,sλ (τ)dτ t t Z s 3 3  i j ˙ k + ∂ijkF (τ, λτ ) − ∂ijkF (t, λt) δλτ,sδλτ,sλ (τ)dτ t

3 Exploiting the cyclical symmetry of tensor ∂ijkF in the indices i, j, k: d ∂3 F (t, λ )δλi δλj λ˙ k(τ) = −∂3 F (t, λ )δλi δλj λk ijk t τ,s τ,s ijk t τ,s τ,s dτ τ,s 1 d d d = ∂3 F (t, λ )[δλi δλj λk + δλj δλk λi + δλk δλi λj ] 3 ijk t τ,s τ,s dτ τ,s τ,s τ,s dτ τ,s τ,s τ,s dτ τ,s 1 d   = − ∂3 F (t, λ ) δλi δλj λk 3 ijk t dτ τ,s τ,s τ,s hence Z s 3 i j ˙ k 1 3 i j k ∂ijkF (t, λt)δλτ,sδλτ,sλ (τ)dτ = ∂ijkF (t, λt)δλt,sδλt,sλt,s. t 3 Thanks to this (27) can be written as Z s Z s F i 1 2 i j Rt,s(λ) = DF (u, λu)du + D∂iF (r, λr)δλr,sdr + ∂ijF (t, λt)δλt,sλt,s t t 2 Z s 1 2 i j 1 3 i j k + D∂ijF (τ, λτ )δλτ,sδλτ,sdτ + ∂ijkF (t, λt)δλt,sδλt,sλt,s 2 t 6 Z s 1 3 3  i j ˙ k + ∂ijkF (τ, λτ ) − ∂ijkF (t, λt) δλτ,sδλτ,sλ (τ)dτ. 2 t (28)

N F N Now, as in the proof of part a), we take λ = ω and estimate Rt,s(ω ). To 3 estimate the last term in (28) we use that ∇ F is in Lip([0,T ], k · k∞) and is locally horizontally Lipschitz:

3 N 3 N 3 N 3 N |∂ijkF (τ, ωτ ) − ∂ijkF (t, ωt )| ≤ |∂ijkF (τ, ωτ ) − ∂ijkF (τ, ωt,τ−t)| 3 N 3 N N N +|∂ijkF (τ, ωt,τ−t) − ∂ijkF (t, ωt )| ≤ CF kωt,τ−t − ωτ k∞ + CF |τ − t| ν ν ≤ CF,kωkν |τ − t| + CF |τ − t| ≤ CF,kωkν ,T |s − t| .

Plugging this estimate into the formula (28) and using the local boundedness 2 2 3 of the derivatives DF, D∂iF, ∂ijF, D∂ijF, ∂ijkF and the estimates (17) we

21 obtain Z s F N N 1 2 i j Rt,s(ω ) − DF (u, ωu )du − ∂ijF (t, ωt)δωt,sδωt,s t 2 1+ν 1+2ν ≤ CF,kωkν |s − t| + CF,kωkν |s − t| (29) 3ν 1−ν 4ν +CF,kωkν |s − t| + CF,kωkν N |s − t| 3ν 1−ν 4ν ≤ CF,kωkν ,T |s − t| + CF,kωkν N |s − t| ,

N N where we have used also that ωt = ωt, ω (s) = ω(s). Recalling (19) we conclude by triangle inequality Z s F 1 2 i j Rt,s(ω) − DF (u, ωu)du − ∂ijF (t, ωt)δωt,sδωt,s t 2 Z s N −ν ν (30) ≤ DF (u, ωu ) − DF (u, ωu) du + CF N |s − t| t 3ν 1−ν 4ν +CF,kωkν ,T |s − t| + CF,kωkν N |s − t| ,

The above inequality holds for any N > 1,thus we can take N ≈ |s − t|−3ν to get Z s F 1 2 i j Rt,s(ω) − DF (u, ωu)du − ∂ijF (t, ωt)δωt,sδωt,s t 2 Z s N ν+3ν2 ≤ DF (u, ωu ) − DF (u, ωu) du + CF,kωkν |s − t| t 1+ν ν+3ν2 ≤ CF kωkν|s − t| + CF,kωkν |s − t| , where in the last inequality we used that DF is Lipschitz continuous. Hence the result.

We are now ready to prove the main result of this section, which gives sufficient conditions on the functional F under which the pathwise integral Z π ∇ωF (t, ω)d ω is a (pathwise) limit of Riemann sums computed along ω, rather than approximations of ω: √ 13 − 1 Theorem 3.2. Let ω ∈ Q ([0,T ], d) ∩ Cν([0,T ], d) with ν > π R R 6 1,2 d 2 satisfying Assumption 2. Assume F ∈ Cb (ΛT ) is such that ∇ωF, ∇ωF, ∈

22 1,1 d 3 d 3 Cb (ΛT ), F, DF, ∇ F ∈ Lip(ΛT , k·k∞) and ∇ωF is horizontally locally Lip- schitz. Then the pathwise integral (9) is a limit of non-anticipative Riemann sums: Z T π X ∇ωF (u, ω)d ω(u) = lim ∇ωF (t, ω)(ω(s) − ω(t). n→+∞ 0 [t,s]∈πn

In particular, its value only depends on F (., ω) and

Z T π ∇ωF (·, ω) = 0 =⇒ ∇ωF (u, ω)d ω(u) = 0. 0 Proof. Let us denote

n n n n n A := F (t , ω n ) − F (t , ω n ) − ∇ F (t, ω n )(ω(t ) − ω(t )) i i+1 ti+1 i ti ω ti i+1 i n Z ti+1 1 2 t − DF (u, ω)du − h∇ωF (t, ωtn ), ωtn,tn ωtn,tn i n 2 i i i+1 i i+1 ti n From Assumption (2) max |Ai | →n→∞ 0 and by part b) of Theorem (3.1) i n n n 3ν2+ν 2 |Ai | ≤ C|ti+1 − ti | . From the condition on ν, 3ν + ν > 1, therefore

1− 1 1− 1 X n n 3ν2+ν X n n n 3ν2+ν |Ai | ≤ C(max |Ai |) |ti+1 − ti | ≤ CT (max |Ai |) → 0, i i i i

X n which implies Ai → 0. Hence the limit of Riemann sums exist and is i equal to

Z T X n n lim ∇ωF (t, ωtn )(ω(t ) − ω(t )) = F (T, ω) − F (0, ω) − DF (u, ω)du n→∞ i i+1 i i 0 Z T Z T 1 2 π π − h∇ωF (t, ω), d[ω] (t)i = ∇ωF (t, ω)d ω(t). 2 0 0

4 Rough-smooth decomposition for functionals of irregular paths

As an application of the previous results, we now derive a decomposition theorem for functionals of irregular paths. Let π = { πn } be a sequence of

23 partitions with |πn| → 0. Consider an irregular path with strictly increasing quadratic variation along π:

d[¯ω] ω¯ ∈ Q ([0,T ], d) ∩ C1/2−([0,T ], d) with π > 0 dt − a.e. (31) π R R dt We define the vector space of paths obtained as the image ofω ¯ under regular non-anticipative functionals:  U(¯ω) := F (·, ω¯) F satisfies the assumptions of Theorem 3.2 ⊂ Qπ([0,T ], R).

Denote

 d Va(¯ω) := ∇ωF (·, ω¯) F satisfies the assumptions of Theorem 3.2 ⊂ Qπ([0,T ], R ).

The following result gives a ’signal plus noise’ decomposition for such paths, in the sense of Dirichlet processes [13]:

Proposition 4.1 (Rough-smooth decomposition of paths). Any path ω ∈ U(¯ω) has a unique decomposition

Z t π ω(t) = ω(0) + φd ω¯ + s(t) where φ ∈ Va(¯ω), [s]π = 0. (32) 0

Z t The term φ.dπω¯ is the ’rough’ component of ω which inherits the ir- 0 regularity ofω ¯ while s(.) represents a ’smooth’ component with zero quadratic variation.

Remark 5 (Pathwise ’Doob-Meyer’ decomposition). This result may be viewed as a pathwise analogue of the well-known decomposition of a continuous semimartingale into a local martingale (a process with stricly increasing quadratic variation A similar pathwise decomposition result was obtained by Hairer and Pillai [18] using rough path techniques (see also [16, The- orem 6.5] and [15]). Unlike the semimartingale decomposition, the result of Hairer and Pillai [18] involves H¨older-type regularity assumptions on the components, as well as a uniform H¨olderroughness condition on the path. Our setting is closer to the original semimartingale decomposition in that the components are distinguished based on quadratic variation and the irreg- ularity condition (31) on ω is also expressed in terms of quadratic variation.

24 d Proof. Let ω ∈ U(¯ω). Then there exists F ∈ R(ΛT ) verifying the assump- tions of Theorem 3.2, with ω(t) = F (t, ω¯). The functional change of variable formula applied to F then yields the decomposition with φ = ∇ωF (., ω¯) ∈ Z t 1 2 Va(¯ω) and s(t) = F (0, ω0)+ du(DF (u, ω)+ < ∇ωF (u, ω), d[ω] >). The 0 2 continuity of ths integrand implies that s has finite variation so [s]π = 0. Consider now two different decompositions Z t Z t π π ω(t) − ω(0) = φ1d ω¯ + s1(t) = φ2d ω¯ + s2(t). 0 0 Z π so s1 −s2 = (φ1 −φ2)d ω.¯ Since φ1 −φ2 ∈ Va(¯ω) this representation shows 0 that s1 − s2 ∈ Qπ([0,T ], R) so the pathwise quadratic covariation [s1, s2]π is well defined and we can apply the polarization formula to conclude that Z π [s1 − s2]π = 0. Applying Proposition 2.6 to s1 − s2 = (φ1 − φ2)d ω.¯ then 0 implies that that φ1 = φ2. Z t π Finally, applying Theorem 3.2 to (φ1−φ2)d ω¯ then shows that s1 = s2 0 which yields uniqueness of the decomposition (32).

Remark 6. As noted by Schied [26], Qπ([0,T ], R) is not a vector space and, given two paths (ω1, ω2) ∈ Qπ([0,T ], R) the quadratic covariation along π cannot be defined in general. By contrast, the space U(¯ω) introduced above is a vector space of paths with finite quadratic variation along π. Moreover, 2 for any pair of elements (ω1, ω2) ∈ U(¯ω) , the quadratic covariation along π Z π is well defined; if ωi = φi.d ω + si is the rough-smooth decomposition of 0 ωi the quadratic covariation is given by Z t t [ω1, ω2]π(t) = < φ1φ2, d[ω] > . (33) 0 References

[1] J. Bertoin, Temps locaux et int´egration stochastique pour les processus de Dirichlet, in S´eminairede Probabilit´es,XXI, vol. 1247 of Lecture Notes in Math., Springer, Berlin, 1987, pp. 191–205.

[2] R. V. Chacon, Y. Le Jan, E. Perkins, and S. J. Taylor, Gener- alised arc length for Brownian motion and L´evyprocesses, Z. Wahrsch. Verw. Gebiete, 57 (1981), pp. 197–211.

25 [3] R. Cont, Functional Ito Calculus and functional Kolmogorov equa- tions, in Stochastic Integration by Parts and Functional Ito Calculus (Lecture Notes of the Barcelona Summer School in Stochastic Analysis, July 2012), Advanced Courses in Mathematics, Birkhauser Basel, 2016, pp. 115–208.

[4] R. Cont and D.-A. Fournie´, Change of variable formulas for non- anticipative functionals on path space, J. Funct. Anal., 259 (2010), pp. 1043–1072.

[5] R. Cont and D.-A. Fournie´, A functional extension of the Ito for- mula, C. R. Math. Acad. Sci. Paris, 348 (2010), pp. 57–61.

[6] R. Cont and D.-A. Fournie´, Functional Ito calculus and stochastic integral representation of martingales, Annals of Probability, 41 (2013), pp. 109–133.

[7] R. Cont and Y. Lu, Weak approximation of martingale representa- tions, . Appl., 126 (2016), pp. 857–882.

[8] R. Cont and C. Riga, Pathwise analysis and robustness of hedg- ing strategies, working paper, Laboratoire de Probabilit´eset Mod`eles Al´eatoires,2015.

[9] M. Davis, J. Ob loj,´ and P. Siorpaes, Pathwise with local times, working paper, Imperial College London, 2015.

[10] C. Dellacherie and P.-A. Meyer, Probabilities and potential, vol. 29 of North-Holland Mathematics Studies, North-Holland Publish- ing Co., Amsterdam, 1978.

[11] B. Dupire, Functional Itˆocalculus, Bloomberg Portfolio Research pa- per, (2009).

[12] H. Follmer¨ , Calcul d’Itˆosans probabilit´es, in Seminar on Probability, XV (Univ. Strasbourg, Strasbourg, 1979/1980) (French), vol. 850 of Lecture Notes in Math., Springer, Berlin, 1981, pp. 143–150.

[13] H. Follmer¨ , Dirichlet processes, in Stochastic Integrals: Proceedings of the LMS Durham Symposium, July 7 – 17, 1980, D. Williams, ed., Springer, Berlin, 1981, pp. 476–478.

[14] D.-A. Fournie, Functional Ito calculus and applications, ProQuest LLC, 2010. PhD Thesis, Columbia University.

26 [15] P. Friz and A. Shekhar, Doob-Meyer for rough paths, Bulletin of the Institute of Mathematics. Academia Sinica., 8 (2013), pp. 73–84.

[16] P. K. Friz and M. Hairer, A course on rough paths, Universitext, Springer, 2014.

[17] M. Gubinelli, Controlling rough paths, J. Funct. Anal., 216 (2004), pp. 86–140.

[18] M. Hairer and N. S. Pillai, Regularity of laws and ergodicity of hypoelliptic SDEs driven by rough paths, Ann. Probab., 41 (2013), pp. 2544–2598.

[19] N. Ikeda and S. Watanabe, Stochastic differential equations and diffusion processes, vol. 24 of North-Holland Mathematical Library, North-Holland Publishing Co., Amsterdam; Kodansha, Ltd., Tokyo, second ed., 1989.

[20] R. L. Karandikar, On pathwise stochastic integration, Stochastic Process. Appl., 57 (1995), pp. 11–18.

[21] M. Nutz, Pathwise construction of stochastic integrals, Electron. Com- mun. Probab., 17 (2012), pp. no. 24, 7.

[22] N. Perkowski and D. J. Promel¨ , Local times for typical price paths and pathwise Tanaka formulas, Electronic Journal of Probability, 20 (2015), pp. no. 46, 15.

[23] , Pathwise stochastic integrals for model free finance, (2015).

[24] D. Revuz and M. Yor, Continuous martingales and Brownian mo- tion, vol. 293 of Grundlehren der Mathematischen Wissenschaften, Springer-Verlag, Berlin, third ed., 1999.

[25] F. Russo and P. Vallois, Elements of stochastic calculus via regu- larization, in S´eminairede Probabilit´es,XL, vol. 1899 of Lecture Notes in Math., Springer, Berlin, 2007, pp. 147–185.

[26] A. Schied, On a class of generalized Takagi functions with linear path- wise quadratic variation, J. Math. Anal. Appl., 433 (2016), pp. 974–990.

[27] L. C. Young, An inequality of the H¨oldertype, connected with Stieltjes integration, Acta Math., 67 (1936), pp. 251–282.

27