arXiv:1508.05984v2 [math.PR] 25 Jan 2017 osdrfatoa rwinmto ihHrtindex Nu Hurst Coutin, with also see motion Brownian processes; of Dirichlet fractional context of [3], consider Bertoin class the to large in related a are seen for results be our First, should works. contribution nected Our time. local disptws oa time local pathwise admits aaaMyrfruaadso htapt facniuu sem continuous a of path a that show and formula Tanaka-Meyer itsai[7 wocnie ls fGusa processes Gaussian of class a consider (who [27] Viitasaari us hw oeitwhenever exist to shown sums, hr h integral the where tcatcitga geswt i ahieitga a.s. integral pathwise his path with the agrees in integral variation stochastic quadratic admits a.s. aclsfauiglcltms hc stefis ancontr devel time main to local first of us the notion led is a It which define sake. times, own local its featuring for calculus pursuing worth (s is finance believe, mathematical in applications from arose question (1.1) oml for formula tcatcitga n t tosfruat functions to Itˆo’s formula its and integral stochastic hsi acls o function a For calculus. chastic variation epu omnsadavc tvrossae fti work. this of stages thank various authors at The advice support. and financial comments Q their helpful of for Institute Oxford Oxford-Man in lege the to grateful Ob also l´ojJan (FP7/2007-2013) is Programme Framework Seventh Union’s pean h neligmtvto eidorcretsuywst ex to was study current our behind motivation underlying The nasmnlppr ¨lmr[3 inee non-probabili a pioneered F¨ollmer [13] paper, seminal a In hsrsac a engnrul upre yteEuropean the by supported generously been has research This Date AHIESOHSI ACLSWT OA TIMES LOCAL WITH CALCULUS STOCHASTIC PATHWISE is eso:1 a 02 hsvrin aur 6 2017. 26, January version: This 2012. May 16 version: First : hntepriin r osrce rmsopn times. stopping from constructed are mot partitions degenerate Brownian the such however, when standard partitions; b a (random) can of partiti of process sequence variation of non-decreasing quadratic choice given pathwise the arbitrary the on an depend that time, show limiting we the local how the detail in and study f variation we Finally, conditions equivalent time. provide local We wise co formulae. ch a Tanaka-Meyer, time of the of paths of change version for pathwise defi a.s. establish We times existing sequ local gale. other general (stochastic) subsumes usual a approach the along Our with quantities interval. discrete time suitable the of limit a Abstract. h x i f t ∈ ln euneo attos( partitions of sequence a along C f MARK DAVIS, JAN OB OB L JAN DAVIS, MARK 2 ( esuyanto flcltm o otnospt,dfie a defined path, continuous a for time local of notion a study We R : x 0 t t ) f − ′ ( x f s ( ) x dx L 0 L = ) X t s t x ( ( x ω sdfie stelmt fnnatcptv Riemann non-anticipative of limits the as defined is u h 1. ) x fra aibe eitoue oino quadratic of notion a introduced he variable, real of Z o otnosfunction continuous a for ) ( i 0 u t Introduction t hc hnare ihteuul(stochastic) usual the with agrees then which ) f xss ¨lmras bevdta aho a of path a that observed F¨ollmer also exists. ′ ( x JADPER SIORPAES PIETRO AND OJ ´ s 1 ) dx s π + n ) 2 1 n atttv iac n tJh’ Col- John’s St and Finance uantitative Z n rvdteascae Itˆo’s associated the proved and 0 asFole n ldmrVv for Vovk Vladimir F¨ollmer and Hans t R rn gemn o 335421. no. agreement grant ERC / f f eerhCucludrteEuro- the under Council Research > H ′′ hc r o in not are which ( h hwdsmlrresults similar showed who bet,tequadratic the objects, eaiu sexcluded is behaviour x tnossemimartin- ntinuous eDvse l 9)bt we but, [9]) al. et Davis ee neo aibe and variables of ange .Scn,rltdresults related Second, ). neo attosof partitions of ence reitneo path- of existence or lr n uo 8 (who [8] Tudor and alart iesneadteusual the and sense wise s bto forwr.We work. our of ibution n.I particular, In ons. ) iin n agrees and nitions x civdas by a.s. achieved e d 1 pptws stochastic pathwise op o o suitable a for ion rv h associated the prove , / h tcapoc osto- to approach stic x )adStie and Sottinen and 3) he rvoscon- previous three i s imartingale edtepathwise the tend , C s 2 This . X a.s. 2 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES appeared in the unpublished diploma thesis of Wuermli [31]. Our approach is simi- lar, however the proof in [31] was complicated and applied only to square integrable martingales. We also have a slightly different definition of which includes continuity in time and importantly we consider convergence in Lp for p [1, ) ∈ ∞ instead of just p = 2. This allows us to capture the tradeoff between the general- ity of paths considered and the scope of applicability of the Tanaka-Meyer formula. π ′′ π Indeed, as the term R Lt (u)f (du) suggests, there is a natural duality between Lt and f ′′, so the smaller the space to which Lπ belongs, the more general f ′′ one can R t take. This fact was already powerfully exploited by our third main reference, the recent paper by Perkowski and Pr¨omel [22] (and, to a much lesser extent, [11]), in π ′′ which Lt belongs to the space of continuous functions, and thus f can be a gen- eral measure (i.e. f ′ has bounded variation and f is the difference of two arbitrary convex functions), recovering Tanaka-Meyer formula in full generality (the authors π also consider the case where Lt is continuous and also has bounded p-variation, and thus f ′ can be any function with bounded q-variation). In particular, the conclusion X(ω) of their main theorem (Theorem 3.5 in [22]) on the existence of Lt (u) for a.e. ω is stronger than ours; however, since their local time has to be continuous, results in [22] apply if X is a either under the original probability P or under some Q P (see [22, Remark 3.6]), whereas our Theorem 6.1 applies to a general ≫ semimartingale X. Further, we investigate several questions not considered in [3], [31] and [22], as we explain now. The main advantage of our definition, as compared with these previous works, is that we are able to characterise the existence of pathwise local time with a number of equivalent conditions (see Theorem 3.1). This feature seems to be entirely new. It allowed us in particular to build an explicit example of a path which admits a quadratic variation but no local time. Also, while [9] and [31] (following [13]) consider partitions πn whose mesh is going to zero (which are well suited for changing variables), the results [3, Theorem 3.1] and [22, Theorem 3.5] of existence of pathwise local time consider Lebesgue1 partitions. Since neither type of partition is a special case of the other, this makes the results in these papers hard to compare. We solve this conundrum by proving our existence result (Theorem 6.1) for a general type of partition, which subsumes both types considered above. 1 Finally, we prove that the existence of Lt(u) is preserved by a C change of variables (improving on [9, Proposition B.6]) and by time changes, and that g t g(x )dx → 0 s s is continuous (similarly2 to [22]). R Finally, we investigate how the limiting objects, quadratic variation and local time, depend on the choice of partitions. We show that for a path which oscil- lates enough, with a suitable choice of partitions, its quadratic variation can attain essentially any given non-decreasing function. From this, taking care of null sets and measurability issues, one can deduce that for a W and a given increasing [0, ]-valued measurable process A with A = 0 there exist refining ∞ 0 partitions (πn)n made of random times such that W πn := (W W )2 A a.s. for all t 0. h it tj ∈πn tj+1∧t − tj ∧t → t ≥ This result illustrates,P in the most stark way possible, the dependence of the pathwise quadratic variation (and thus of the pathwise local time) on the partitions (πn)n. This may push the reader to dismiss the notion of pathwise quadratic variation (and local time). However, it worth recalling that if we restrict ourselves to partitions

1Which we define in (2.2). 2Note that due to difference in definitions of local time, we use different topologies. PATHWISE WITH LOCAL TIMES 3 constructed from stopping times, the limit of W πn exists and is independent of the h it choice of partitions, and it always equals t. Analogously, our Theorem 6.1 states that, if we only consider partitions constructed from stopping times, the pathwise X(ω) local time Lt (u) of a semimartingale X exists and is independent of the choice of partitions, and it coincides with the classical local time. As already known by L´evy, inf x π = 0 for every continuous function x and πh i1 sup W (ω) π = for a.e. ω. Our analysis builds on these facts and answers πh i1 ∞ in particular two questions which they leave open: whether for the general path x = W (ω) one can make x πn converge to any chosen C = C(ω) R, and what h i1 ∈ dependence in t we can expect for lim x πn . Specifically it is clear that it must nh it be an increasing function, and we wondered whether it is automatically continuous; indeed, while we followed F¨ollmer [13], who carefully required that lim x πn be nh it continuous3, several authors who cite [13] do not (see for example [3], [9], [26]) and our results show that this is a significant omission. The plan for rest of the paper is as follows. In Section 2 we introduce most of the notations and definitions, and recall parts of [13]. In Section 3 we iden- tify several conditions equivalent to the existence of pathwise local time, prove the Tanaka-Meyer formula and the continuity of g t g(x )dx . In Section 4 we → 0 s s consider change of variable and time, and in Section 5 we extend Tanaka-Meyer R formula from the case of a Sobolev function f to the case where f is a difference of convex functions. In Section 6 we prove that a path of a semimartingale a.s. ad- mits the pathwise local time, and relate this to the downcrossing representation of semimartingale local time proved by L´evy. Finally in Section 7 we state the results about dependence of quadratic variation on the sequence of partitions, including the convergence W πn A mentioned above. We only give the proof for one path h it → t avoiding the (non-trivial) technicalities related to measurability and null sets. The latter are given in the appendix.

2. Pathwise stochastic calculus In this section we introduce most notations and definitions used throughout the article, and we revisit the part of [13] which deals with continuous functions, slightly refining its results to include uniformity in t and more general partitions. By measure we mean sigma-additive positive measure; a Radon measure will be the difference of two measures which are finite on compact sets. With µ we | | will denote the total-variation measure relative to a ‘real measure’ µ (i.e. µ is the difference of two measures), and with max(µ, 0) the measure (µ + µ )/2 (i.e. the | | positive part in the Hahn-Jordan decomposition of µ). We will say that gn g p → fast in Lp(µ) if g g < ; this trivially implies that g g a.s. n∈N || n − ||Lp(µ) ∞ n → and in Lp(µ). We will denote by (resp. ) the Borel sets of [0, ) (resp. [0,T]). P B BT ∞ For a continuous function x = (xs)s≥0, xt and xt are respectively the minimum and maximum of x over s [0,t]. We set x = 0, denote by δ the Dirac measure at t, s ∈ ∞ t and by π a partition of [0, ), i.e. π = (t ) N where t [0, ],t = 0,t < t ∞ k k∈ k ∈ ∞ 0 k k+1 if t < , and lim t = . For such x and π, we set k+1 ∞ k→∞ k ∞ (2.1) O (x,π) := max x(b) x(a) : a, b [t ,t ) [0,t] for some k N . t {| − | ∈ k k+1 ∩ ∈ }

3 More precisely [13] deals with c`adl`ag x and requires that µn (defined later in (2.3)) converge 2 weakly to a measure µ which assigns mass (∆xt) to the singleton {t}; if x is continuous this implies Π continuity of hxi· . 4 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES

F¨ollmer works with a sequence of finite partitions (πn)n whose step on compacts converges to zero. This excludes very commonly used partitions: the Lebesgue par- titions, i.e., those of the form πP = πP (x) = (tk)k∈N, (2.2) where t := 0, t := inf t>t : x P,x = x 0 k+1 { k t ∈ t 6 tk } for some P partition of R, i.e., P = (pk)k∈Z with

pk [ , ], lim pk = , p0 = 0, and pk

6 n ∗ ∗ For example |Ht |≤ Xt with X t := sup s≤t |Xs| . 6 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES where X = M+V is the canonical semimartingale decomposition of X, [M] := M 2 t t − 2 t M dM is the quadratic variation of M and V is the variation of (V ) . 0 s s | |t s s∈[0,t] We recall the inequality R (2.10) sup Xt Lp(P) Cp X Sp , k t≤T | | k ≤ k k which holds for local martingales (this being one side of the celebrated Burkholder- Davis-Gundy inequalities) and thus trivially extends to X p. We will also use ∈ S without further mention that if H is locally-bounded and predictable then the canon- · · · ical decomposition of 0 HdX is 0 HdM + 0 HdV and so · R TR R1/2 T 2 HdX = Ht d[M]t + Ht d V t . p p 0 S 0 p 0 | | | | L Z Z  L Z

Proposition 2.4. If in Proposition 2.3 we make the stronger assumption that

O (X,π ) < a.s. for all T < then X(ω) Π and X(ω) Π = [X](ω) for n T n ∞ ∞ ∈ Q h i a.e. ω. P Proof. Fix a compact time interval [0, T ] on which we will work. By prelocalizing we can assume that X 4 (see7 Emery´ [10, Th´eor`eme 2]) and passing to an equivalent ∈ S probability we can moreover8 assume that K := O (X,π ) L4. Take Hn as n T n ∈ in Proposition 2.3, Kn := Hn X and notice that sup Kn O (X,π ), so − P t | t | ≤ T n sup Kn K and in particular sup Kn 2 K2 L2. Using (2.10) and n t | t | ≤ n t | t | ≤ ∈ (a + b)2 2(a2 + b2) gives that sup t KndX converges to zero fast in L2 if P ≤ t | P0 | E( (Kn)2Rd[M] + ( Knd V )2) n | | 2 2 is finite, which is true sinceP itR is bounded aboveR by E(K ([M]T + V )), which is | |T finite by H¨older inequality since K L4, X 4. Since sup t KndX 0 fast ∈ ∈ S t≤T | 0 |→ in L2 and thus a.s., (2.9) yields sup X πn [X] 0 a.s..  t≤T |h it − t|→ R Theorem 2.5 (F¨ollmer [13]). If x , g C1 and t [0, ) the limit ∈ Q ∈ ∈ ∞ (2.11) lim g(xt )(xt ∧t xt ∧t) n j j+1 − j t ∈π jXn t exists uniformly on compacts and defines a continuous function of t denoted 0 g(xs)dxs. This integral satisfies Itˆo’s formula: for f C2(R), ∈ R t 1 t (2.12) f(x ) f(x )= f ′(x )dx + f ′′(x )d x . t − 0 s s 2 s h is Z0 Z0 Notice that the series defining the F¨ollmer integral in (2.11) and later in this paper are in fact finite sums, since every partition is finite on compacts. Proof. By using the second order Taylor’s expansion write (2.13) f(x ) f(x ) tj+1∧t − tj ∧t t ∈π jXn as 1 (2.14) f ′(x )(x x )+ f ′′(x )(x x )2 + C (t) , tj tj+1∧t − tj ∧t 2 tj tj+1∧t − tj ∧t n t ∈π t ∈π jXn jXn

7This statement also appears in [23, Chapter 5, Theorem 14], without proof. 8If X ∈ S4(P) and (dQ/dP)(ω) := C exp(−K(ω)) then K ∈ L4(Q), and X ∈ S4(Q) since dQ/dP ∈ L∞(P). PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 7 where the correction term Cn(t) is bounded by (2.15) φ( x x )(x x )2 | tj+1∧t − tj ∧t| tj+1∧t − tj ∧t t ∈π jXn for some increasing function φ which is continuous at 0 and such that φ(0) = 0. Since x (πn)n , the term (2.15) converges to 0 (for t = T , and thus also uniformly ∈ Q in t T ). Since (2.4) states that the second term of (2.14) converges to the last ≤ term of (2.12) uniformly on compacts, by difference the first term of (2.14) also converges, uniformly on compacts; moreover (2.12) holds since the telescopic sum (2.13) equals f(x ) f(x ).  t − 0 Remark that F¨ollmer [13] considers sums of the form g(x )(x x ) and g(x )(x x )2, tj tj+1 − tj tj tj+1 − tj πnX∋tj ≤t πnX∋tj ≤t whereas we consider (2.16) g(x )(x x ) and g(x )(x x )2. tj tj+1∧t − tj ∧t tj tj+1∧t − tj ∧t t ∈π t ∈π jXn jXn Since the difference between these two sums is g(x )(x x ) and g(x )((x x )2 (x x )2) ti ti+1 − t ti ti+1 − ti − ti+1 − t (where i := max j : π t t ), which goes to zero as O (x,π ) 0, these { n ∋ j ≤ } T n → expressions are equal in the limit. The reason we prefer to use (2.16) is that it involves only non-anticipative quantities (i.e. their value of time t does not depend on the value of x at later times), which better fits with the theory of stochastic integration and thus allows us to obtain formulae like (2.9) and (6.7).

3. Pathwise local time

As already suggested in [13], there should be an extension of ‘Itˆoformula’ valid also when f ′′ is not a continuous functions, as it is in (2.12). In the theory of continuous semimartingales, such an extension proceeds via local times and the Tanaka-Meyer formula; what follows is a pathwise version. ′ ′′ If f− is the left-derivative of a convex function f, and f is the second derivative of f in the sense of distributions (i.e. the unique positive Radon measure which satisfies f ′′([a, b)) = f ′ (b) f ′ (a)) we obtain for a b − − − ≤ b b f(b) f(a) = f ′ (y)dy = f ′ (a)+ f ′′(du) dy − − − Za Za Z[a,y) ! = f ′ (a)(b a)+ (b u)f ′′(du) , so that − − − Z[a,b) ∞ f(b) f(a) = f ′ (a)(b a)+ 1 (u) b u f ′′(du), a, b R. − − − [a∧b,a∨b) | − | ∀ ∈ Z−∞ So if given a function x and a partition π = (t ) , we set for u, v R j j ∈ [u, v), if u v, Ju, vM := ≤ ([v, u), if u>v, and define the discrete local time (along π) as π (3.1) L (u) := 2 1Jx ,x M(u) xtj ∧t u , t tj ∈π tj ∧t tj+1∧t | +1 − | P 8 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES then, if f equals the difference of two convex functions, we have the following discrete Tanaka-Meyer formula 1 (3.2) f(x ) f(x )= f ′ (x )(x x )+ Lπ(u)f ′′(du). t − 0 − tj tj+1∧t − tj ∧t 2 t t ∈π R Xj Z A simple but important remark is that only the values of f in the compact interval [x , x ] are relevant. Note that Lπ(u)=0 for u / [x , x ] and Lπ( ) is c`adl`ag, thus it t t t ∈ t t t · is bounded; in particular Lπ( ) is f ′′-integrable. t · In the remainder of this section we will restrict our attention to those functions whose second derivative is not a general Radon measure but instead one which ad- mits a density with respect to the Lebesgue measure. Thus, the underlying measure space will be R with its Borel sets, endowed with the Lebesgue measure 1(du) L (sometimes denoted simply by du). We will consider Lπn ( ) as a function in Lp, and t · denote by W k,p the (Sobolev) space of functions whose kth derivative in the sense of distributions is in Lp; i.e., W 1,p is the set of absolutely continuous functions whose classical derivative (which exists a.e.) belongs to W 0,p = Lp, and W 2,p is the set of C1 functions whose classical derivative belongs to W 1,p. The following is our main theorem in this section. Theorem 3.1. Let x be continuous function and fix a sequence of partitions Π = (π ) such that O (x,π ) 0 as n for all t [0, ). Then, for 1

9 2 Indeed [31] does not require t 7→ Lt to be continuous, and considers strong convergence in L instead of weak convergence in Lp. PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 9

Example 3.6 below we show that the two notions are strictly different and give an explicit construction of a path which admits quadratic variation but not a pathwise local time. We will henceforth denote by the space of continuous functions x for which the Lp equivalent conditions of Theorem 3.1 hold. We will call Lt(u) the pathwise local time of x at time t at level u. Observe that and L (u) a priori depend on Π = (π ) , Lp t n n and that L (u) depends on x. We will write Π, LΠ(u), Lx(u) or Lx,Π(u) only t Lp t t t when we want to highlight these dependencies; as in the remainder of this section n πn Π = (πn)n will be fixed, we will never do that, and we will simply write Lt for Lt . n n Notice that since Lt (u)=0for u / [xt, xt], we can consider Lt (u) as an element of p ∈ L (µ) with µ being the restriction of the Lebesgue measure to [xt, xt]. In particular, k,q k,q Theorem 3.1 holds if W is replaced with Wloc . Moreover, ifp ˆ p, since µ is p pˆ ≤ πn finite, L (µ) embeds continuously in L (µ), and so p pˆ and the limits of (Lt )n p pˆ L ⊆L in the weak L and L topology coincide, so Lt does not really depend on p. Note also that for x , using standard regularisation techniques, we can define ∈L1 a modification (lt)t of the pathwise local time (Lt)t which is c`adl`ag and increasing in t for a.e. u. The occupation time formula then extends to all Borel bounded h T T (3.5) h(t,x )d x = h(t, u)dl (u)du t h it t Z0 ZR Z0 Finally, we show that if x the F¨ollmer integral is a continuous linear func- ∈ Lp tional on W 1,q. This fact could have been used to define F¨ollmer’s integral for g W 1,q as the continuous extension of the F¨ollmer’s integral for g C1 defined ∈ ∈ in Theorem 2.5, as done in [3]. Note that the following result would not hold if we only assumed uniform convergence on compacts of gn to g. Proposition 3.2. Let p (1, ] (resp. p = 1) with conjugate exponent q. If x , ∈ ∞ ∈Lp g , g W 1,q, g (x ) g(x ) and g′ g′ in the weak (resp. weak∗) topology of Lq, n ∈ n 0 → 0 n → then t g (x )dx t g(x )dx for all t [0, ), and the convergence is uniform 0 n s s → 0 s s ∈ ∞ on compacts if moreover g′ g′ weakly (resp. weakly∗) in Lq. R R | n| → | | Proof. Define f(u) := u g(y)dy and analogously f from g , and notice that x0 n n f (u) f(u) for all u R, so Tanaka-Meyer formula (3.3) gives the thesis. If n → R∈ moreover g′ g′ weakly in Lq then since the positive part max(h, 0) of h equals | n| → | | (h + h )/2, Polya’s scholium 2.2 shows local uniformity of the convergence | | L (u) max(g′ (u), 0)du L (u) max(g′(u), 0)du ; t n → t ZR ZR working analogously with the negative parts we get the thesis.  In the rest of this section we establish Theorem 3.1 via a series of lemmas; if not explicitly stated otherwise, p is assumed to be in (1, ). ∞ Lemma 3.3. x iff Ln(u)du converges to a continuous function ψ of t ∈ Q R t t ∈ [0, ). In this case the convergence is uniform on compacts and x = ψ. ∞ R h i Proof. Applying formula (3.2) with f(x)= x2 L1([x , x ]) we obtain ∈ t t (3.6) x2 x2 2x (x x )= Ln(u)du tj ∈πn tj+1∧t − tj ∧t − tj tj+1∧t − tj ∧t R t The statementP follows rewriting the left side of (3.6) as R (x x )2.  tj ∈πn tj+1∧t − tj ∧t Given x , ν will denote the occupation measure ofP(x ) (along Π), defined ∈ Q t s s≤t on the Borel sets of [0,t] by ν (A) := t 1 (x )d x . t 0 A s h is R 10 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES

Lemma 3.4. If x and t [0, ) the following are equivalent. ∈ Q ∈ ∞ (1) For every g W 1,q the following sequence converges ∈ (3.7) g(x )(x x ). tj ∈πn tj tj+1∧t − tj ∧t n p (2) The sequence (L (P))n converges in the weak topology of L (to a quantity t · which we denote by L ( )). t · The above conditions imply that (Ln( )) is bounded in Lp and (3.3) holds. Con- t · n versely, if (Ln( )) is bounded in Lp and x then items (1) and (2) hold, and ν t · n ∈ Q t has a density L with respect to 1. t L Proof. The equivalence between items (1) and (2), and the fact that these imply (3.3), follows immediately applying (3.2) with f(u) := u g(y)dy. That item (2) x∗ implies the boundedness of (Ln( )) follows from Banach-Steinhaus Theorem. For t · n R the opposite implication notice that since x we can use Theorem 2.5, which ∈ Q together with (3.2) shows that (3.8) lim Ln(u)h(u)du = t h(x )d x = hdν for all h C0 . ∃ n R t 0 s h is t ∈ Since Lp is reflexiveR (see [6, TheoremR 4.10]), its unitR ball is sequentially compact n in the weak topology [6, Theorem 3.18], so we can get convergence of Lt along some subsequence (of any subsequence) to some Lt and all we have to show is that the n limit does not depend on the subsequence. Considering (Lt )n as elements of the measure space ([x , x ], 1) we have that C0 Lp, so g(u)ν (du)= g(u)L (u)du t t L ⊆ t t for all continuous g. Thus L (u)du = ν (du); in particular the limit L does not t t R R t depend on the subsequence, proving item (2).  Lemma 3.5. If the equivalent conditions 1 and 2 of Lemma 3.4 are satisfied for all t [0, ), the following conditions are equivalent. ∈ ∞ (1) For every g W 1,q the function t g(x )dx is continuous in t [0, ). ∈ 0 s s ∈ ∞ (2) For every g W 1,q the convergence in (3.7) is uniform on compacts. ∈ R (3) The map [0, ) t L ( ) Lp is continuous in the weak topology of Lp. ∞ ∋ 7→ t · ∈ (4) For every h Lq the convergence Ln(u)h(u)du L (u)h(u)du is uni- ∈ R t → R t form on compacts. R R Proof. The identity (3.2) shows that items (2) and (4) are equivalent. The identity (3.3) shows that (1) and (3) are equivalent. Trivially item (2) implies item (1). Finally scholium 2.2 shows that item (3) implies item (4).  Proof of Theorem 3.1. If item 5 holds, since the last term in the decomposition (4.3) n is bounded by Ot(x,π) and the two sums are increasing in t, (Lt )n is bounded in Lp for all t [0, ); moreover Lemma 3.4 shows that for all h Lq ∈ ∞ ∈ (3.9) lim Ln(u)h(u)du = L (u)h(u)du = hdν = t h(x )d x ; ∃ n R t t t 0 s h is Since (3.9) shows thatR Lt(u)h(u)duRis a continuous functionR of Rt, Lemma 3.5 implies that item 3 holds. That item 3 implies x follows applying Lemma 3.3 since R ∈ Q 1 Lp and Ln = 0 outside [x , x ]. Lemma 3.4 states that ν has a density L ; [xt,xt] ∈ t t t t t thus, formula (3.4) holds. All other assertions follow directly from Lemmas 3.4 and 3.5. If p = 1 or p = the proofs hold with the following minor modification in the ∞ part of the proof of Lemma 3.4 which deals with the sequential compactness of (Ln) . If p = , the unit ball of L∞ is sequentially compact since it is compact t n ∞ (and metrizable) in the weak∗ topology because of Banach-Alaoglu Theorem (and since L1 is separable), see [6, Theorem 3.16] (and see [6, Theorems 3.28 and 4.13]). PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 11

n n 1 If p = 1, since Lt 0, Lemma 3.3 implies that (Lt )n is bounded in L , so if n ≥ (Lt )n is equintegrable then it is weakly sequentially compact (by the Dunford-Pettis Theorem, see [6, Theorem 4.30]).  We end this section with the following Example 3.6. There exists a function that admits pathwise quadratic variation but no pathwise local time. Put differently, we show that the inclusion Π Π can be strict. This proves L1 ⊂ Q that the additional requirement in (5) in Theorem 3.1 is not automatic and is indeed needed. More precisely, we now construct a continuous function x : [0, 1] R and → a sequence of refining partitions Π = (πn)n of [0, 1] whose mesh is going to zero and πn such that x t converges for all t to the (continuous!) Cantor function c(t) (a.k.a. h i πn 1 the Devil staircase) but (L1 )n does not converge weakly in L (du); in particular x 10 has no pathwise local time along (πn)n, no matter which definition we use . Our construction was inspired by a remark by Bertoin on page 194 of [3]. 1 1 2 Divide [0, 1] into three equal subintervals and remove the middle one I1 := ( 3 , 3 ). 1 2 Divide each of the two remaining closed intervals [0, 3 ] and [ 3 , 1] into three equal 2 1 2 2 7 8 subintervals and remove the middle ones I1 := ( 32 , 32 ) and I2 := ( 32 , 32 ). Continu- i i ing in this fashion, at each step i we remove the middle intervals I1,...,I2i−1 , each of length 1/3i. The Cantor set is defined as i−1 C := [0, 1] ∞ 2 Ii , \∪i=1 ∪j=1 j and the function which we will consider is x(t) := 2min s t . To construct s∈C | − | our partitions π of [0, 1] we define first a refining sequence (πi ) of Lebesgue n p n,j n i i k,i 2ni+1 0,i i 0,i partitions of Ij setting πn,j = (tn,j)k=0 with tn,j = inf Ij (so that x(tn,j) = 0), and k+1,i k,i −ni (3.10) tn,j := inf t>tn,j : xt (2 supt∈Ii x(t))Z, xt = x k,i , { ∈ j 6 tn,j } 2ni+1,i i −ni n i so that tn,j = sup Ij and x k+1,i x k,i = 2 supt∈Ii x(t) = 1/(√32 ) and so | tn,j − tn,j | j i πn,j 2ni+1−1 2 ni+1 −i −2ni −i 1−ni (3.11) x i := k=0 (xtk+1,i xtk,i ) = 2 3 2 = 3 2 . h iIj n,j − n,j Then define our refiningP sequence (πn)n of partitions of [0, 1] whose mesh is going n 2i−1 i 2 to zero setting π := 0, 1 π and we set ǫ := n so that as n n { }∪∪i=1 ∪j=1 n,j n 2 3 →∞ i−1 πi n+1 πn n 2 n,j n i 1−(ǫn) (3.12) x 1 = i=1 j=1 x i = i=1(ǫn) = 1−ǫ 1 . h i h iIj n → Now, the Cantor functionP cPis defined on [0P, 1] to be the only continuous extension ∞ 2i−1 ¯i 11 of the function f which is defined on the set D := 0, 1 i=1 j=1 Ij in this way : { }∪∪ ∪ i f(0) = 0,f(1) = 1, and each time we remove the middle third Ij from a parent i ¯i i interval Jj, f is defined on the closure Ij of Ij to be the average of its values at the i ¯1 ¯2 ¯2 extremes of Jj (so f = 1/2 on I1 , f = 1/4 on I1 and f = 3/4 on I2 etc.). Since the difference between x πn and the increasing function (x h it πn∋tj ≤t tj+1 − x )2 is going to zero for all t as n , and since c is continuous and increasing, tj → ∞ P to conclude that x πn c(t) for all t it is enough to show it for all t in the h it → dense set D, see also Lemma 7.2 below. We already know this for t = 1 and

10Meaning that if one replaced the weak topology of L1 with any stronger topology (e.g. the weak/strong topology of Lp, or the uniform topology as done in [22, Definition 2.5]) one would still πn not obtain convergence of Lt . 11Such extension exists and is unique since f is continuous and D is dense in [0, 1]. 12 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES

πn πn πn (trivially) for t = 0. Since x 1 = x 2 , and (3.11) shows that x 1 2 0, h i(0, 3 ] h i( 3 ,1] h i( 3 , 3 ] → πn 1 2 ¯1 (3.12) and (2.6) give that x t 1/2 = c(t) for all t [ 3 , 3 ] = I1 . Analogously πn πn h i → πn ∈ πn x 1 = x 2 1 , and (3.11) shows that x 1 2 0, so x 1 1/2 and (2.6) give h i(0, 9 ] h i( 9 , 3 ] h i( 9 , 9 ] → h i 3 → that x πn 1/22 = c(t) for all t [ 1 , 2 ]= I¯2. In this way we see that x πn c(t) h it → ∈ 9 9 1 h it → for all t D and thus for all t [0, 1]. ∈ ∈ πn To conclude, let us prove that the pathwise local time L1 (u) converges to 0 for all u = 0, so that (Lπn ) does not converge weakly in L1(du) (because otherwise, by 6 1 n the Dunford-Pettis Theorem [6, Theorem 4.30], it would be uniformly integrable and 1 1 πn would thus converge to zero strongly in L (du), whereas we know that 0 L1 (u)du = πn πn x 1 c(1) = 1). Since x(t) 0 for all t, L1 (u) = 0 for all n if u < 0. For each h i → ≥ 12 R i, j the function (x(t)) i crosses each level u > 0 at most twice, and since t∈πn∩Ij i−1 sup x(t) : t / k 2 Ii = 1/√3k+1 is strictly smaller than any u > 0 for big { ∈ ∪i=1 ∪j=1 j} enough i = i(u), the number of times (x(t))t∈πn crosses level u> 0 is bounded above independently of n; since O (x,π ) = 1/(√3n) 0 as n , this implies that 1 n → → ∞ Lπn (u) 0. 1 → 4. Change of variables and time-change In applications to the study of variance derivatives, for example [9], one starts with a continuous positive price function S, and the ‘variance’ is defined as the quadratic variation of the log price x = log S. In this connection it is useful to be able to change variables, and to relate for example the local time of log(x) with the one of x. We recall that, although being a semimartingale is preserved only by C2 transformations, possessing a quadratic variation (in the sense of Definition 2.1) is more generally invariant under C1 transformations; indeed f C1 and x Π t ′ ∈ ∈ Q imply f(x) Π and f(x) Π = f (x )2 x Π (see [26, Proposition 2.2.10]). We ∈ Q h it 0 s h is prove below a similar result for the pathwise local time (if f is monotone), extending R the C2 case treated in [9]; then we show that time-change preserves the pathwise local time. For Propositions 4.1 and 4.2 we consider a fixed sequence of partition (πn)n such that O (x,π ) 0 as n for all t [0, ). t n → →∞ ∈ ∞ Proposition 4.1. Let x and let f : R R be C1 and strictly monotone. Then ∈Lp → f(x) and the pathwise local times of x and f(x) are related by ∈Lp (4.1) Lf(x)(f(u)) = f ′(u) Lx(u). t | | t In Proposition 4.1 one considers the same sequence of partitions (πn)n for x and for f(x). This seems to be problematic, since ideally we would like Proposition 4.1 to hold also for Lebesgue partitions, and clearly if P is a partition of R then πP (f(x)) differs from πP (x). However Proposition 4.1 does apply to suitably chosen Lebesgue partitions since πf(P )(f(x)) = πP (x) if f is strictly increasing. π To prove Proposition 4.1 and better understand the behavior of L , let tJ := max t π : t t and { j ∈ j ≤ } (4.2) πU (u) := t π : x u

Proof of Proposition 4.1. Since adding t to any partition π does not change the value π of Lt (u) and insures that the last term in (4.3) is zero, we assume without loss of generality that our partitions contain t. If f is strictly increasing and a b then ≤ x u

(4.4) U (f(xt ) f(u)) + D (f(u) f(xt )). tj ∈πt (u) j+1 − tj ∈πt (u) − j+1 If t πU (u),P since f C1 there exists z (uP) (u, x ) such that j ∈ t ∈ j ∈ tj+1 f(x ) f(u)= f ′(z (u))(x u), tj+1 − j tj+1 − so we can write the first sum in (4.4) as

′ ′ ′ (4.5) (f (zj (u)) f (u))(xtj+1 u)+ f (u) (xtj+1 u). U − − U − tj ∈Xπt (u) tj ∈Xπt (u) Treating analogously the second sum in (4.4) we get that

(4.6) Lf(x),π(f(u)) f ′(u)Lx,π(u) t − t is bounded by

′ ′ (4.7) 2 f (zj(u)) f (u) xtj+1 u . U D | − || − | tj ∈πt (Xu)∪πt (u) Now define (4.8) R (g,π) := max g(c) g(d) : c, d [x , x ], c d O (x,π) . t {| − | ∈ t t | − |≤ t } ′ x,πn x,πn Clearly (4.7) with π = πn is bounded by Rt(f ,πn)Lt (u), so since Lt converges to Lx and R (f ′,π ) 0 we get that (4.6) with π = π converges to 0, proving the t t n → n thesis. If f is strictly decreasing then the argument is the same save for the sign change, which comes from the fact that upcrossings are now transformed in downcrossings and conversely, so x u needs to be replaced by u x .  tj+1 − − tj+1 Proposition 4.2. Let τ : [0, ) [0, ) be an increasing c`adl`ag function such ∞ → ∞ that xτ is continuous and τ(0) = 0. Given Π = (πn)n, let τ(Π) := (τπn )n where, given π = (t ) , τ denotes the partition (τ ) . If O (x,τ ) 0 for all t [0, ) j j π tj j τt πn → ∈ ∞ τ(Π) Π and x p then O (x τ,π ) 0 for all t [0, ), x and the pathwise ∈ L t ◦ n → ∈ ∞ τ ∈ Lp local times are related by x◦τ,Π x,τ(Π) Lt (u)= Lτt (u). 13 −1 Moreover if τ is bijective then x τ is continuous, O −1 (x τ,τπ )= Ot(x,π) for ◦ τt ◦ any partition π, and if P is a partition of R then the Lebesgue partitions of x τ −1 ◦ and x satisfy πP (x τ)= τ . ◦ πP (x) Proof. Even if τ is not strictly increasing, the identity τ : s [t ,t ) [0,t] = [τ ,τ ) [0,τ ] τ([0, )), { s ∈ i i+1 ∩ } ti ti+1 ∩ t ∩ ∞ holds, and it trivially implies that Ot(x τ,πn) Oτt (x,τπn ), with equality if τ is a x◦τ,π x,τπ ◦ ≤ bijection. Trivially Lt (u)= Lτt (u) holds for every partition π, and everything else follows easily. 

13 I.e. if τ is strictly increasing, continuous and such that τ(0) = 0, limt→∞ τ(t)= ∞. 14 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES

Note that Propositions 4.1 and 4.2 hold (with the same proof) with other defini- tions of existence of the pathwise local time; for example if one replaced the weak topology of Lp for p [1, ) (resp. the weak∗ topology on L∞) with the strong one ∈ ∞ in item 3 of Theorem 3.1, or if one considered Definition 2.5 in [22].

5. Extension to convex functions The choice of how to define the existence of the pathwise local time is intrinsically linked to the class of functions for which one is able to establish the pathwise Tanaka- Meyer formula (3.3). To establish it for all convex functions one needs to restrict significantly the set of paths for which the local time exists; nonetheless, in [22] it is shown that this approach works for general enough paths (namely, for the ‘typical path’ in the sense of Vovk [30]). It is natural to ask if the above can be extended even further, to all continuous functions. As already remarked in [13], the next proposition shows that, if one wants to consider a generic path of a local martingale, the answer is no – to define stochastic integrals in a pathwise manner for more general integrands one has to consider partitions which depend both on the integrator and the integrand as in [4, Theorem 7.14] and [16]. Proposition 5.1. (Stricker [28]) Let x C[0, T ]. If for every sequence of parti- ∈ tions (π ) with O (x,π ) 0 and every bounded continuous function f on R the n n T n → Riemann sums f(x )(x x ) converge, then x has finite variation. ti∈πn ti ti+1 − ti In what followsP we take a different route from [22] to further extend F¨ollmer’s integral and Tanaka-Meyer formula beyond f C2. We consider f which is a ′ ∈ difference of two convex functions and write f− for its left-continuous derivative and f ′′ for the second distributional derivative of f. In a way somewhat reminiscent of t ′ t ′ [3, Proposition 1.2], we define 0 f−(xs)dxs as the limit of 0 fn(xs)dxs, where fn are some special C2 functions converging to f and t f ′ (x )dx is defined in Theorem R 0 n s R s 2.5 as a limit of Riemann sums. R We now fix Π=(π ) such that O (x,π ) 0 as n for all t [0, ), and n n t n → →∞ ∈ ∞ we consider a function g which is C2, positive and with compact support in [0, ), ∞ and such that R g(x)dx = 1. We will then approximate the target fuction f with f := g f, where denotes the convolution between a function and a measure (or a n n ∗ R ∗ function), g is the mollifier g (u) := ng(nu). Recall that, if x , L ( ) is seen an n n ∈L1 t · element of L1(du); the following theorem assumes that there exists a modification of Lt which is c`adl`ag in u, i.e., a function L˜t(u) c`adl`ag in u and such that, for each t, the set u : L˜ (u) = L (u) has zero Lebesgue measure; this is not an unreasonable { t 6 t } assumption, as it is satisfied by a.e. path of a semimartingale (indeed the local time of a continuous semimartingale has a modification which is jointly c`adl`ag in u and continuous in t).

Theorem 5.2. Assume that x 1 and there exists a modification Lt(u) of the ∈ L 2 pathwise local time which is c`adl`ag in u for all t. If f is convex then fn is C and for all t [0, ) the F¨ollmer integral t f ′ (x )dx converges to a finite limit, ∈ ∞ 0 n s s denoted by t f ′ (x )dx , which is independent of the choice of g and satisfies 0 − s s R R t 1 (5.1) f(x ) f(x )= f ′ (x )dx + L (u)f ′′(du). t − 0 − s s 2 t Z0 ZR Moreover if Lt(u) is jointly c`adl`ag in u and continuous in t then the convergence is uniform on compacts and t t f ′ (x )dx is continuous. 7→ 0 − s s R PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 15

14 t Theorem 5.2 allows to define 0 g−(xs)dxs for any function g of finite variation on compacts, since then f(u) := u g(y)dy is the difference of convex functions. We xR0 now study the continuity properties of g · g (x )dx . R 7→ 0 − s s Proposition 5.3. Let x and assume thatR there exists a modification L (u) of ∈L1 t the pathwise local time which is continuous in u. If gn, and g are functions of finite variation on compacts, g (x ) g(x ) and g′ g′ weakly (seen as measures), then n 0 → 0 n → t g (x )dx t g(x )dx for all t [0, ). Moreover, if g′ g′ weakly then 0 n s s → 0 s s ∈ ∞ | n| → | | the convergence is uniform on compacts. R R Proof. Define f(u) := u g(y)dy and analogously f from g , and notice that x0 n n f (u) f(u) for all u R, so Tanaka-Meyer formula (5.1) gives the thesis. If n → R∈ moreover g′ g′ weakly then, since the positive part max(h, 0) of h equals | n| → | | (h + h )/2, Polya’s scholium 2.2 shows local uniformity of the convergence | | L (u) max(g′ , 0)(du) L (u) max(g′, 0)(du) ; t n → t ZR ZR working analogously with the negative parts we get the thesis.  t ′ It is then natural to ask for which paths the above given definition of 0 f−(xs)dxs coincides with the one used in Theorem 3.1 for f W 2,q. The answer is that the limit ∈ R t of the Riemann sums f ′ (x )(x x ) exists and equals f ′ (x )dx tj ∈πn − tj tj+1∧t − tj ∧t 0 − s s iff Lx,πn (u)f ′′(du) converges to Lx(u)f ′′(du), as it follows from (3.2) and (5.1). t P t R In particular this holds if x , so the definition of the F¨ollmer’s integral R ∈ LpR⊆ L1 given in Theorem 5.2 is indeed an extension of the one given in Theorem 3.1. Proof of Theorem 5.2. Since f is uniformly continuous on compacts, f f point- n → wise. Thus, if we can prove that L df ′′ L df ′′, the thesis follows applying R t n → R t (3.3) to f and taking limits; indeed (5.1) shows that t f ′ (x )dx does not de- n R R 0 − s s pend on g. Defineg ˆ (u) := g ( u) and apply Fubini’s theorem and the identity n n − R f ′′ = g f ′′ to get that n n ∗ (5.2) L df ′′ = (ˆg L ) df ′′. t n n ∗ t ZR ZR Since L is zero outside [x , x ] and g has compact support, L , g andg ˆ L are all 0 t t t t n ∗ t outside a common compact interval [ A, A]. In particular since L ( ) is c`adl`ag it is − t · bounded; since sup gˆ L (u) sup L (u) , the thesis follows from the dominated u | n ∗ t |≤ u | t | convergence theorem and (5.2) if we prove thatg ˆ L (u) L (u) for all u. Notice n ∗ t → t that (5.3) (ˆg L L )(u)= gˆ (y)(L (u y) L (u))dy . n ∗ t − t n t − − t ZR Since L ( ) is right continuous, for every ǫ> 0 and u there is an n such that t · (5.4) L (u y) L (u) <ǫ if y [ A/n, 0]; | t − − t | ∈ − since g = 0 outside [0, A], the integral on the right side of (5.3) is actually over [ A/n, 0], so gˆ L (u) L (u) <ǫ. − | n ∗ t − t | Finally if Lt(u) is jointly c`adl`ag in u and continuous in t then n such that (5.4) holds can be chosen as to hold simultaneously for all t in any given compact set.

14 t ′ As pointed out to us by F¨ollmer [12] another possible definition of R0 f−(xs)dxs for non-smooth t ′ 2 ′′ convex f is as the limit of R0 fk(xs)dxs for any (fk)k ⊆ C such that fk (x)dx (considered as a measure) converges weakly to f ′′(x)dx. It follows from (3.3) that this definition makes sense (i.e. the limit exists and is independent of the approximating sequence (fk)k), and agrees with ours, if Lt(u) has a modification which is continuous in u). 16 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES

· ′ This implies that the convergence is uniform on compacts, and so 0 f−(xs)dxs is continuous.  R 6. Upcrossing representations of local time

In this section we will consider a continuous semimartingale X = (Xt)t (with t [0, ) or t [0, T ]) with canonical semimartingale decomposition X = M + V ∈ ∞ ∈ 15 and with (classical ) local time ℓt(u) which is (jointly) continuous in t and c`adl`ag in u (such a version exists, see [17, Chapter 3, Theorem 7.1]). Some of our results specialise to the case where ℓ is jointly continuous in t and u; this holds in the important case when dV is absolutely continuous with respect to d M (this follows h i from (6.6) below, see also [32, Example 2.2.3]), in particular if X is a local martingale (under the original probability P or a Q such that P Q). ≪ The following is the main theorem of this section. It essentially says that the pathwise local time sampled along optional partitions (πn)n exists on a.e. path of a semimartingale, and that a.e. it equals the (classical) local time (in particular, it does not depend on (πn)n). Theorem 6.1. Assume that f : R R is the difference of two convex functions, that → π are optional partitions such that O (X,π ) 0 a.s. and that X = (X ) n T n → t t∈[0,∞) is a continuous semimartingale. If X has a jointly continuous local time ℓ, or if f 1 is C , then there exists a subsequence (nk)k such that, for ω outside a P-null set (which may depend on f ′′),

X(ω),πnk (ω) p ′′ (6.1) sup Lt (u) ℓt(ω, u) 0 in L ( f (du)) as k t≤T − → | | →∞ simultaneously for all p [1, ), T < . ∈ ∞ ∞ Note that applying Theorem 6.1 with f(x) = x2/2 C1 gives in particular ∈ that a.e. path of a continuous semimartingale is in p for all p < ; indeed, X,π L ∞ L nk (u) ℓ (u) strongly (and thus weakly) in Lp(du) a.s., locally uniformly in t. t → t The previous theorem follows from the following technical statement. Theorem 6.2. Let π be optional partitions such that O (X,π ) 0 a.s., p n T n → ∈ [1, ),T < , X p, µ be a sigma-finite positive Borel measure on R, and define ∞ ∞ ∈ S

πn X(ω),πn(ω) (6.2) h (u) := sup Lt (u) ℓt(ω, u) , u R. t≤T − p ∈ L (P(dω))

Then hπn ( ) is bounded and (hπn ( )) converges pointwise (resp. µ a.e.) to 0 if ℓ is · · n jointly continuous (resp. if µ is a measure with no atoms). The fact that (hπn ( )) converges pointwise to zero was given an involved proof16 · n in [31] in the case where X is a continuous martingale bounded in L2, p = 2 and π are deterministic partitions such that π 0. In the case where X is in a n || n|| → class of continuous Dirichlet processes which includes 2 semimartingales and the S partitions are of Lebesgue-type, it is shown in [3, Theorem 2.5 and Proposition 2.7] that LX(ω),πn(ω)(u) ℓ (ω, u) weakly in L1(dP du) for each t. t → t × Moreover, Lemieux [18, Theorem 2.4] has derived a version of Theorem 6.1 where the Lp( f ′′ (du)) convergence is replaced by the uniform convergence, in the special | | 15We refer to the semimartingale local time, i.e. the one for which the Tanaka-Meyer formula holds; this is in general different from the parallel notion of local time for Markov processes. 16The uniformity in t, not stated in [31], follows easily by Doob’s L2-inequality since (6.7) shows πn,X 2 that (Lt (u) − ℓt(u))t is a L bounded martingale for each u, n. PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 17 case where the partitions are of Lebesgue-type. For the case of continuous local martingales one can also consult Perkins [21] or Chacon et al. [7, Theorem 2 and Remark 2], or Perkowski and Pr¨omel [22, Theorem 3.5 and Remark 3.6], who actually prove convergence not only for P a.e. ω but even quasi surely with respect to the set of all local martingale measures. Although our approach yields a weaker type of convergence, it has a simple proof and it works for continuous semimartingales and general optional partitions such that O (X,π ) 0 a.s.. T n → In the special case of Lebesgue partitions π = πǫZ, Theorem 6.2 closely relates to the downcrossing representation of local time conjectured by L´evy (proved by Itˆo and McKean for Brownian motion, extended by El Karoui to semimartingales, and found in [24, Theorem VI.1.10]), which states that, for an X p, ∈ S

ǫ lim sup ǫDt (ω, 0) ℓt(ω, 0) = 0, ǫ→0 t≤T | − | p L (P(dω))

ǫ where Dt (ω, 0) (defined in (6.3) below) is the number of downcrossings at level 0. Indeed, as we now explain, L´evy’s representation above is equivalent to the fact that hπǫnZ (0) 0 whenever 0 <ǫ 0. → n → Given a continuous path x = (x ) and a < b, we set σa,b := 0, τ a,b = inf t : s s≤t 0 0 { x = b and, for k 1, we define t } ≥ a,b a,b a,b a,b σ := inf t>τ : xt = a , τ := inf t>σ : xt = b , (6.3) k { k−1 } k { k } Dǫ(u) := max k : σu,u+ǫ t . t { k ≤ } ǫ It turns out that the downcrossings Dt (u) of (Xs)s≤t from u+ǫ to u are closely related ǫ u+ǫ,u to the local time along πǫZ. Indeed, the upcrossings Ut (u) := max k : τk t ǫ { ≤ } of (Xs)s≤t from u to u + ǫ differ from Dt (u) by at most 1, so using (4.3) we get that

πǫZ πǫZ πǫZ (6.4) L (u)/2= U (u)(ǫ u)+ D (u)u + 1[x ,x )(u) xt u . t t − t tJ t | − | The last term is bounded by O (x,π Z) ǫ and, considering u = 0, we get that t ǫ ≤ (6.5) LπǫZ (0)/2 ǫDǫ(0) 2ǫ , | t − t |≤ which concludes the proof of equivalence. We recall the following fact, for which we refer to [24, Chapter 6, Theorem 1.7]: · · (6.6) 2 1 dX = 2 1 dV = ℓ (u) ℓ (u ) a.s., u R. {Xs=u} s {Xs=u} s · − · − ∀ ∈ Z0 Z0 Proof of Theorem 6.2. Consider the convex function f(x) := x u and let sign(x | − | − u) be its left-derivative and 2δu its second (distributional) derivative. Subtracting from the discrete-time Tanaka-Meyer formula (3.2) its continuous-time stochastic counterpart we get that t (6.7) 0= (Hπn (u) H (u))dX + (Lπn,X (u) ℓ (u))/2, s − s s t − t Z0 n where using πn = (τi )i we define the predictable processes

πn H (u) := sign(Xτ n u)1(τ n,τ n ](s) and Hs(u) := sign(Xs u). s i − i i+1 − Xi πn · πn Now h (u) 0 follows from (2.10) and (6.7) if we show that 0 Hs (u)dXs · → → H (u)dX in p. To this end notice that 0 s s S R (6.8)R Hπn (u) H (u) Kπn (u) := 2 1 , | s − s |≤ s × {Os(X,πn)≥|Xs−u|} 18 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES

πn and that since X· and O·(X,πn) are continuous adapted processes, K· (u) is pre- dictable, so it is enough to prove that · Kπn (u)dX 0 in p. Since Oπn 0 0 s s → S T → a.s. implies that Kπn (u) 0 a.s. on X = u for all t T , and since Kπn t → R{ t 6 } ≤ ≤ 2, the thesis follows from the (deterministic) dominated convergence theorem if · 1 dX p = 0, which by (6.6) holds for all u if ℓ is continuous. Since k 0 {Xs=u} skS Minkowski inequality for integrals says that R T T T 1/p p 1{Xs=u}d V s 1{Xs=u} L (µ)d V s = (µ( Xs )) d V s , | | p ≤ k k | | { } | | Z0 L (µ) Z0 Z0

p P which is zero for µ which has no atoms. Using (6.6), considering L (µ ) norm and · ⊗ π using Fubini, we conclude that 1 dX p = 0 for µ a.e. u, and so h n 0 k 0 {Xs=u} skS → µ a.e.. Finally (2.10), (6.7) and (6.8) imply that R · πn π h (u) Cp 2(Hs (u) Hs(u))dXs 4Cp X Sp for all u R, ≤ − p ≤ k k ∈ Z0 S concluding the proof. 

Proof of Theorem 6.1. Let (τm)m a sequence of stopping times which prelocalizes X to p (see Emery [10, Theoreme 2]), i.e. τ a.s. and Xτm− p for all m. Let S m ↑∞ ∈ S µ (A) := f ′′ (A [ i, i]) and set i | | ∩ − G (ω, T, u) := sup LX(ω),πn(ω)(u) ℓ (u, ω) n t≤T | t − t | m and Gn := 1{T <τm}Gn. Since µi is a finite measure, Theorem 6.2 implies that, as m p n , Gn converges to 0 in L (P µi), for all m, i N and T 0. Passing to → ∞ × ∈ ≥ p a subsequence (without relabelling) we can get convergence fast in L (P µi) and p,T m p × so, for ω outside a P-null set Ni,m , Gn (ω,T, ) converges to 0 in L (µi). Then along m · p a diagonal subsequence we obtain that Gn (ω,T, ) converges to 0 in L (µi) for all · p,T i,m,p,T N 0 for every ω outside the null set N ′′ := N . Since ∈ \ { } f ∪i,m,T,p∈N\{0} i,m G = Gm on T <τ , G 0 in Lp(µ ) for all i, p, T N 0 for every ω n n { m} n → i ∈ \ { } outside N ′′ . Since outside a compact set G (ω,T, ) = 0 for all n, convergence in f n · Lp(µ ) for arbitrarily big i, p implies convergence in Lp( f ′′ ) for all p [1, ). Since i | | ∈ ∞ G (ω, , u) = 0 is increasing, convergence for arbitrarily big T implies convergence n · for all T [0, ).  ∈ ∞ 7. Dependence on the partitions In this section we investigate the extent to which the pathwise quadratic variation x Π := lim x πn depends on the sequence of partitions Π := (π ) . Instead of h i nh i n n constructing explicit examples we show that, for functions with a highly oscillatory behavior, the pathwise quadratic variation depends in the most extreme way possible on (πn)n. We then build on this fact and state how this applies to the general path of a Brownian motion; since taking care of all the thorny technicalities which arise from the dependence in ω (i.e. tracking the null sets and ensuring measurability) requires a long technical proof, we relegate this to the appendix. Our work builds on two facts already mentioned (without proof) by L´evy in [20, Pag. 190]: that inf x π = 0 for every continuous function x and that for a.e. πh i1 path B(ω) of a Brownian motion sup B(ω) π = . The corresponding proofs πh i1 ∞ can be found in Freedman [14, Pag. 47 and 48]; the second fact can be found in a strengthened form and with an alternative proof in Taylor [29, Corollary in Section 4]. Our first result combines and generalises the above: we show that, with a suitable choice of (πn)n, the pathwise quadratic variation may be equal to an arbitrary given PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 19 increasing process a. Notice that we do not even make assume a is right–continuous. We will denote with D the dyadics of order n in [0, 1], i.e D := [0, 1] N2−n. n n ∩ Theorem 7.1. Let x : [0, 1] R be a continuous function such that for every → 0 c < d 1 there exist partitions (ˆπ ) of [c, d] such that lim x πˆn = . Then, ≤ ≤ n n nh i(c,d] ∞ if a : [0, 1] [0, ) is an increasing function such that a = 0, there exist refining → ∞ 0 partitions (π ) of [0, 1] such that D π for all n and n n n ⊆ n (7.1) x πn a for all t [0, 1] as n , h it → t ∈ →∞ and the convergence is uniform if (at)t is continuous. Moreover, given arbitrary partitions (¯π ) , one can choose the (π ) such that π¯ π for all n. n n n n n ⊆ n To prove that convergence occurs at all times simultaneously, we will need the following simple lemma. Lemma 7.2. Let a : [0, 1] [0, ) be increasing, x : [0, 1] R be continuous and → ∞ → (π ) be partitions of [0, 1] such that O(x,π ) 0, and assume that n n n → x πn a(t) for all t F [0, 1] as n . h it → ∈ ⊆ →∞ If F is dense in [0, 1] and contains the times of jump of a then x πn a pointwise h i· → · on [0, 1], and if a is continuous the convergence is uniform.

Proof. Although x πn is not necessarily an increasing function, it differs from the h i n increasing function a (t) := µn([0,t]) (where µn is as in (2.3)) by at most O(x,πn), and so it is enough to prove the statement with an replacing x πn . By hyphothesis h i x πn a(t) for all t at which a is not continuous. If a is continuous at t then h it → for each ε > 0 there exist s ,s F s.t. s 0 there existπ ˜′ such that x π˜ [1, 1 + 1/i]. As we assumed above, h is ⊆ there exist a partitionπ ˆ such that x πˆ is arbitrarily large. Now using Freedman’s h is construction with k equal to be the integer part of x πˆ we obtainπ ˜′ = F (ˆπ, k) such h is 17This requires that π contains each endpoint of the subintervals; this does no harm, as it only means the π we have to build must contain these points. 20 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES

π′ that x s [1, 1 + 1/k] [1, 1 + 1/i], concluding our proof of the existence of some h i ∈ ′ ⊆ π′ such that x π a(t) 1/2n. |h it − |≤ Now, given anyπ ˜n, by applying the above reasoning to the increments of a on each subinterval ofπ ˜ , we can find a π π˜ such that x πn a(t) 1/2n n n ⊇ n |h it − | ≤ simultaneously for all t π˜ . If we define (˜π ,π ) by induction settingπ ˜ := ∈ n n n 0 0, 1 =: π and takingπ ˜ to be the union of D π¯ with the times when a has { } 0 n n ∪ n a jump of size bigger than 1/n and with π , and then building π fromπ ˜ as ∪k

One can apply Theorem 7.1 to the paths of Brownian motion; indeed, as L´evy first remarked, for a.e. ω there exist partitions π = π (ω) s.t. lim B(ω) πn = . n n nh i(0,1] ∞ However, to obtain an interesting result one needs to show that the partitions can be chosen in a measurable way. This requires first to correspondingly strengthen Levy’s result in the following way. Lemma 7.3. If 0 c < d there exist random partitions πn of [c, d] such that ≤ n B π a.s. as n . h i(c,d] →∞ →∞ To prove Lemma 7.3, one needs to revisit the proof of [29, Theorem 1 and its Corollary in Section 4] and delve into the proof of the existence of a Vitali subcover to show how one can choose a measurable one (on a set of probability arbitrarily close to 1); although this essentially follows from an application of the section theorem, the proof is involved and we relegate it to the appendix. Having established Lemma 7.3, one can follow the logic of the proof of Theorem 7.1 and with laborious but entirely elementary proofs18 one can check measurability to obtain a similar result for the paths of Brownian motion, which we state below. To slightly generalize Theorem 7.1 to include the case of a positive but potentially non- finite process A, we identify [0, 1] (with the Euclidean topology) with [0, ] using the ∞ bijection 1 exp( x) (where e−∞ := 0), and thus we endow [0, ] with the distance − − ∞ d(a, b) := e−a e−b , which makes it homeomorphic to [0, 1] and for all M < | − | ∞ satisfies d(a, b) a b Cd(a, b) for all a, b [0, M] and some C = C(M). Of ≤ | − | ≤ ∈ course, if A (ω) is finite valued the convergence under the Euclidean distance a b · | − | is equivalent to the convergence under the distance d(a, b). In all that follows, if π = (τn)n is a random partition we denote by π(ω) the sequence (τn(ω))n∈N. Given sets C, D which depend on ω, we write that C D if C(ω) D(ω) for a.e. ω, ⊆ ⊆ and in particular we say that a sequence of random partition (πn)n∈N is refining if π π for all n. n ⊆ n+1 Theorem 7.4. Let B be a Brownian motion and A a jointly-measurable19 increasing process with values in [0, ] and such that A = 0. Then, there exist refining random ∞ 0 partitions π of [0, ) such that N2−n π and for all ω outside a null set n ∞ ⊆ n (7.2) B(ω) πn(ω) A (ω) for all t [0, ) as n ; h it → t ∈ ∞ →∞ 18The proofs rest entirely on Borel Cantelli’s lemma and on the fact that, given a c`adl`ag adapted process D, its jumps of size bigger than a given constant are stopping times, see [25, Theorem 3.1]. 19Of course any c`adl`ag increasing process A is jointly-measurable. However this is not true for general increasing processes: for example if A := Y 1{τ} + 1(τ,∞) where τ is an exponentially distributed random time and Y is a non-measurable function with values in (0, 1) then A is a process with respect to the completed sigma-algebra (At is measurable since At = 0 a.e.), yet A is clearly not jointly-measurable. PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 21 if 0 c

It is insightful to contrast this result with the well known fact that, if (πn)n is a sequence of optional partitions such that O (B,π ) 0 a.e., B(ω) πn(ω) converges T n → h it to t uniformly on compacts in probability (see the proof of Proposition 2.3). The i random times (τn)i making up πn do not need to look far into the future to break the convergence of B(ω) πn(ω) to t: as the theorem states, one could take the (τ i ) to h it n i look only an arbitrarily small amount of time into the future. Notice that the random times making up π are bounded (since (τ i ) = π N2−n implies τ i i/2n). n n i n ⊇ n ≤ Although we stated Theorem 7.4 only for Brownian motion, it holds for any con- tinuous stochastic process with an oscillatory behavior wild enough to have infinite 2-variation on any interval, in the sense that for every 0 c

Appendix A. Proof of Theorem 7.4 In order to deal with the technicalities involved in tracking the dependence in ω, we need to introduce a number of new definitions; these boil down to asking that, when evaluated at each ω, random partitions are (deterministic) partitions and the operations defined on them correspond to the analogous operations for (deterministic) partitions. Given two random times σ τ, with slight abuse of ≤ notation we denote by [σ, τ] the set (ω,t) Ω [0, ) : σ(ω) t τ(ω) . Given { ∈ × ∞ ≤ ≤ } random times σ τ we will say that π = (τ ) N is a random partition of [σ, τ] if ≤ k k∈ τ are random times such that τ = σ, τ τ τ with τ <τ on τ <τ , k 0 k ≤ k+1 ≤ k k+1 { k+1 } and for a.e. ω there exists some k = k(ω) such that τk(ω)(ω)= τ(ω); we then denote by K(π) the (finite) random variable

(A.1) K(π) := min k N : τ = τ if π = (τ ) N. { ∈ k } k k∈ We denote by [σ, τ] the set of random partitions of [σ, τ], with σ the constant P { } partition (i.e. σ = (σ ) with σ = σ for all i N), and with [0, ) the set of { } i i i ∈ P ∞ random partitions of [0, ) defined shortly before Definition 2.1 (one could more ∞ generally define the random partitions of [σ, τ); notice that for (τ ) to be in [σ, τ) n n P it is not required that τ = τ for big enough n, unlike [σ, τ], so the set τ (ω) n P ∪n{ n } does not need to be finite). We now introduce several operations that one can perform on random partitions. Given random times α σ τ β and (τ ) = π ≤ ≤ ≤ k k ∈ [α, β] we define π [σ, τ] to be the random partition ((τ τ) σ) of [σ, τ]; notice P ∩ k ∧ ∨ k that (π [σ, τ])(ω) equals (π(ω) [σ(ω),τ(ω)]) σ(ω),τ(ω) . Given a measurable ∩ ∩ ∪ { } partition (An)n∈N of Ω and for each n a random quantity τn defined on An, one can define on Ω a random quantity τ by setting τ := τn on An; we will sometimes use this construction to define random times (and thus random partitions). Given random partitions π = (τn)n of [σ, τ] andπ ˜ = (˜τn)n of [˜σ, τ˜], we define by induction

20 i −n This is only used to obtain that τn +2 are stopping times; otherwise B measurable is enough. 22 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES the random partition π π˜ = (ρ ) of [σ σ,τ˜ τ˜] as follows: ρ := σ σ˜, if ∪ n n ∧ ∨ 0 ∧ ρ (ω)= τ(ω) τ˜(ω) then ρ (ω) := τ(ω) τ˜(ω), and if ρ (ω) <τ(ω) τ˜(ω) then n ∨ n+1 ∨ n ∨ ρ (ω) := min t > ρ (ω) : t π(ω) π˜(ω) τ(ω) τ˜(ω) ; n+1 { n ∈ ∪ ∪ { ∨ }} notice that the ρn are indeed random times, as it follows from the following repre- sentation, where given a random time τ and a measurable set A we denote by τ A the random time τ A := τ on A and τ A := on Ω A: ∞ \ {τk>ρn} {τ˜k>ρn} ρn+1 = min(τ τ˜ (τ τ˜)). k k ∧ k ∧ ∨ i k i Given π [σi,τi] for i k we can analogously define i=0π [mini σi, maxi τi]; ∈ P i ≤ ∪ k ∈ P in particular when π = σ for each i this defines σ = (ρ ) N as an { i} ∪i=0{ i} n n∈ element of [min σi, max σi] (the point being that the (ρ ) are ordered whereas P i i n n the (σi)i in general are not). Notice that we cannot reasonably define the random i i i partition i∈Nπ for general (π )i∈N; indeed in general the set i∈Nπ (ω) is not ∪ ′ ′ ∪i finite, so there is no random partition π such that π (ω) = i∈Nπ (ω) for a.e. ω. i ∪ However, if π = (τi)i [σ, τ] and π [τi,τi+1] for each i N, we can define i ∈ P ∈ P ∈ Nπ = (ρ ) [σ, τ] by induction like above: ρ := σ, if ρ (ω) = τ(ω) then ∪i∈ n n ∈ P 0 n ρn+1(ω) := τ(ω), and if ρn(ω) <τ(ω) then i (A.2) ρ (ω) := min t > ρ (ω) : t Nπ (ω) τ(ω) ; n+1 { n ∈∪i∈ ∪ { }} i since for fixed ω the set π(ω) is finite, also Nπ (ω) is finite; thus the minimum ∪i∈ in (A.2) exists, and σK = τ for some K = K(ω), so i∈Nπi [σ, τ]. Analogously i ∪ ∈ Pi if π = (τ ) [0, ) and π [τ ,τ ] we can define Nπ = (ρ ) [0, ) i i ∈ P ∞ ∈ P i i+1 ∪i∈ n n ∈ P ∞ such that ρ τ for all n by setting ρ := 0 and n ≤ n 0 i ρ (ω) := min t > ρ (ω) : t Nπ (ω) , n+1 { n ∈∪i∈ } i where the minimum exists and ρ since Nπ (ω) is finite on compacts. n ↑∞ ∪i∈ Recall the definitions given in (2.5)–(2.8). Note that given finite random times α σ τ β and π [α, β] the random variable B π is defined by (2.5) path ≤ ≤ ≤ ∈ P h i(σ,τ] by path, i.e., B π (ω) := B(ω) π(ω) . Thanks to the next simple lemma, h i(σ,τ] h i(σ(ω),τ(ω)] in the rest of this section we only need to consider [0, 1]-valued random times; in particular π(ω) will be a finite set for a.e. ω for any random partition π and B π h i(σ,τ] will always be well defined. Scholium A.1. It is enough to prove Theorem 7.4 on t [0, 1] . { ∈ } Proof. For k N applying Theorem 7.4 on the time interval [0, 1] to the Brownian k ∈ k k motion Bt := Bt+k Bk, the increasing process At := At+k Ak andπ ¯n := − k − k k (¯πn [k, k + 1]) k produces a refining sequence (πn)n [0, 1] such that πn π¯n ∩ −πn ⊆ P ⊇ for all n and B t converges a.s. to At uniformly on t [0, 1]. Since (k)k∈N is a h i k ∈ k ‘random’ partition of [0, ) and π +k [k, k+1], we can define π := Nπ +k, ∞ n ∈ P n ∪k∈ n which is a random partition of [0, ) that trivially gives Theorem 7.4 on the time ∞ interval [0, ).  ∞ In the proof of the next lemma we will use the following notation: given a subset E of (0, 1) Ω and ω Ω, we set E(ω) := t : (t,ω) E , E(t) := ω : (t,ω) E × ∈ { ∈ } { ∈ } and Π (E) := ω : (t,ω) E for some t (0, 1) . With we will denote Ω { ∈ ∈ } B1 × F the product sigma algebra of the Borel sets of (0, 1) with the underlying sigma B1 algebra on Ω; whenever a function of (t,ω) (or a subsets of (0, 1) Ω) is - F × B1 × F measurable, we will simply say that is measurable. We will assume that contains F all null sets; this is without loss of generality because of [15, Chapter 1, Lemma PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 23

1.19]. The Lebesgue measure on (0, 1) will be denoted with 1, and the dyadics L (resp. the dyadics of order n) in (0, 1) with D (resp. D ), i.e D := (0, 1) N2−n n n ∩ and D = D . As usual the inf (resp. sup) of the empty set is defined to be ∪n≥0 n ∞ (resp. ). −∞ Proof of Lemma 7.3. Step 1. Notice that Brownian motion satisfies (7.3) with ψ(h) := h2/(2 log log(1/h)) since φ(h) := 2h log log(1/h) is asymptotically inverse to ψ (meaning that φ(ψ(h))/h 1 and ψ(φ(h))/h 1 as h 0) and by the the iterated log law → p → ↓ B B lim sup | t+h − t| = 1 a.s. . h↓0 φ(h) Moreover also the process X := (B B )/√b a satisfies (7.3) (with the t t(b−a)+a − a − same ψ), so we can without loss of generality take a = 0, b = 1. For k,n N, ∈ t,ǫ (0, 1) and h F (0, 1] define Y h := ψ( B B )/h, ∈ ∈ ⊂ t | t+h − t| (A.3) E(F ) := (t,ω) (0, 1) Ω : Y h > 1 ǫ for some h F , { ∈ × t − ∈ } En := E((0, 1/2k ] D ) and E := E((0, 1/2k ]). Since Y h and Y (n) := max Y h : k ∩ n k t t { t h (0, 1/2k] D are continuous in t and measurable in ω, Y h and Y (n) are ∈ ∩ n} · · measurable. It follows that En = Y (n) > 1 ǫ is measurable , and so also is k { · − } En = E((0, 1/2k ] D). Since Y h is continuous in h, E equals E((0, 1/2k ] D) ∪n k ∩ t k ∩ and thus it is measurable, and in particular E := E is measurable. Notice that ∩k k (7.3) shows that P(Ek(t)) = 1 for each t, k, and so P(E(t)) = 1 for each t. Fubini’s theorem applied to the product of P with 1 shows that 1(E (ω))=1= 1(E(ω)) L L k L for P a.e. ω, and in particular Π (E) = ω : E(ω) = has probability 1. Define Ω { 6 ∅} for every ω Π (E) and n N ∈ Ω ∈ n(ω) := [t,t + h] : h (0, 1/2n], h + t< 1, (t,ω) E and Y h(ω) > 1 ǫ ; J { ∈ ∈ t − } since by definition E is the set of (t,ω) for which there exist arbitrarily small h> 0 such that Y h > 1 ǫ, n(ω) is a Vitali cover of E(ω) for every ω Π (E). It t − J ∈ Ω follows from Vitali’s covering theorem [19, Theorem 1.31] that for every ω ΠΩ(E) n n n n N (ω) ∈ there exist N (ω) < and ((ti , hi )(ω))i=1 such that for all i = j and ω Ω ∞ 6 n ∈ (A.4) n(ω) [tn, tn + hn](ω) =: In(ω) , In(ω) In(ω)= , N hn(ω) > 1 ǫ . J ⊇ i i i i i ∩ j ∅ i=1 i − n n n Assume for the moment that ti , hi ,N depended measurablyP in ω. Since on ΠΩ(E) n n 2 Y := ψ( Btn hn Btn ) h (1 ǫ) (1 ǫ) > 0 | i + i − i | ≥ i − ≥ − Xi Xi and any interval in n(ω) has length at most 1/2n, lim ψ(s)/s2 = 0 implies that J s↓0 2 (Btn hn Btn ) on Π (E) and so a.s., i i + i − i →∞ Ω n n and so if π = (sk)k is the random partition made of 0, 1 and the points t and P n i tn + hn for i = 1,...,N n we get that B π a.s.. i i h i(0,1] →∞ Step 2. Thus, to conclude the proof it is enough to show that one can choose n n n ti , hi ,N which depend measurably in ω and satisfy (A.4) for all ω. While we cannot quite do that, by revisiting the proof of Vitali’s covering theorem and applying the section theorem (for an elementary proof of which we refer to [1, Theorem 3.1], [2]) and its immediate corollary [24, Chapter 1, Theorem 4.14] we obtain measurable tn, hn,N n which satisfy (A.4) for all ω V , where V is a large set, and this i i ∈ n n allows us to conclude the proof as we explain after (A.6). While on Vn necessarily tn, hn (0, 1] for i N n, in general our tn, hn may also take the values 0 and . i i ∈ ≤ i i ∞ Indeed, in Steps 3,4 we will construct by induction on i 1 random times tn, hn ≥ i i 24 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES which satisfy the properties stated below in (A.6) relative to the objects which we will now define. Set V := Π (E) ( tn + hn < ) n Ω ∩ ∩i≥1{ i i ∞} and define the decreasing family of random intervals ( n) setting n := n = Ji i J1 J 6 ∅ on Π (E) and n := otherwise and, given Ω J1 ∅ (A.5) Cn := tn + hn < n = , i { i i ∞}∩{Ji 6 ∅} we define by induction n := [t,t + h] n : [t,t + h] [tn, tn + hn]= on Cn , n := otherwise , Ji+1 { ∈Ji ∩ i i i ∅} i Ji+1 ∅ and then we set Si := sup h (0, 1] : [t,t + h] n,t (0, 1) 0 , { ∈ ∈Ji ∈ }∨ i n n n so that S = 0 = i = . In Steps 3,4 we will construct ti , hi such that { } {Jn n ∅}n n n i n [ti , ti + hi ] i and 2hi >S > 0 on Ci (A.6) ∈J hn =0 on n = = tn = 0 , P(V ) 1 1/2n; i {Ji ∅} { i } n ≥ − n n once obtained such random times (ti , hi )i≥0, the proof proceeds as follows. From (A.6), the proof of Vitali’s covering theorem (see [19, Theorem 1.31]) shows that 1 n n n (E(ω) i≥1[ti , ti + hi ](ω)) = 0 for ω Vn, and so (since we proved that L1 \ ∪ n k n ∈ (E(ω)) = 1) N := inf k : i=1 hi > 1 ǫ is a finite random variable on L n { n − }n n Vn. Notice that hi > 0 for all i N , since hi = 0 = i = is increas- n n n NPn≤ { } {J ∅} ing in i. Thus N and (ti , hi )i=1 satisfy (A.4) on Vn for all i = j, and thus if n n6 n n π = (sk)k is the random partition made of 0, 1 and the points ti and ti + hi for n n 2 i = 1,...,N , reasoning as in Step 1 we have that Y (1 ǫ) on Vn. It follows πn ≥ − that B (0,1](ω) if ω Vn for infinitely many n’s, and so by Borel Cantelli’s h i πn → ∞ ∈ n lemma B (0,1](ω) a.s. since P(Ω Vn) 1/2 . h i →∞ \ ≤ n n Step 3. To conclude the proof we need, for fixed n, to define random times ti , hi which satisfy (A.6). We will so do in Step 4, using the auxiliary processes Li which we introduce in this step. Given n (which so far we only defined for i = 1) and Ji F (0, 1/2n], define Li(F ) : (0, 1) Ω [0, 1] as ⊆ × → (A.7) Li(F )(ω) := sup h F : [t,t + h] n(ω) 0 , t { ∈ ∈Ji }∨ and set Li := Li((0, 1/2n]). We now need to prove that L1 is measurable. Notice h 1 1 n that, since Yt is continuous in h, L = supk≥n L (Dk (0, 1/2 ]). As we proved E h ∩ and Y· are measurable, and so such is A1 := (t,ω) E : t< 1 h, Y h > 1 ǫ . h { ∈ − t − } k−n Since L1(D (0, 1/2n]) = i/2k on A1 ( 2 A1 ) for i = 1,..., 2k−n and k ∩ i/2k \ ∪j=i+1 j/2k L1(D (0, 1/2n]) = 0 otherwise, L1(D (0, 1/2n]) and thus L1 are measurable. k ∩ k ∩ In Step 4 we will build random times tn, hn from L1. From tn, hn one can define n 1 1 1 1 J2 as done after equation (A.5) and thus L2 as specified in (A.7). One can then iterate the above procedure and define tn, hn by induction on i 1: from a measurable L2 i i ≥ build random times tn, hn as explained in Step 4, and from them build n and a 2 2 J3 measurable L3 etc. For this to work we need to show that Li built from n (and Ji thus from n and the random times (tn, hn)i−1 ) is measurable for all i 2; we J1 j j j=1 ≥ now do so for i = 2, the general case being only notationally more complicated. If F (0, 1/2n], from the definition of n it follows that L2(F ) equals ⊆ J2 t L¯2(F ) := sup h F : [t,t + h] n and [t,t + h] [tn, tn + hn]= 0 t { ∈ ∈J ∩ 1 1 1 ∅} ∨ PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 25

n n 2 n n on C1 , whereas on Ω C1 we have Lt (F ) = 0. Since t1 , h1 are random times and n \ 21 n 1 = is -measurable (as proved in Step 4), C1 is -measurable; thus, it is {J 6 ∅} F ¯2 F enough to prove that Lt (F ) is measurable. The proof is basically the same as for L1: since A1 is measurable, such is A2 := A1 B1 where h h h ∩ h B1 := (t,ω) E : t> tn + hn (t,ω) E : t + h< tn , h { ∈ 1 1 } ∪ { ∈ 1 } k−n and since L¯2(D (0, 1/2n]) = i/2k on A2 ( 2 A2 ) for i = 1,..., 2k−n and k ∩ i/2k \ ∪j=i+1 j/2k L¯2(D (0, 1/2n]) = 0 otherwise, L¯2(D (0, 1/2n]) is measurable and thus so is k ∩ k ∩ L¯2 = sup L¯2(D (0, 1/2n]). k≥n k ∩ Step 4. In this step we explain how to use a measurable Li to build random n n times ti , hi which satisfy the first 3 statements of (A.6) and (A.8) P(tn + hn = ) 1/2n+i; i i ∞ ≤ since P(Π (E)) = 1, it follows from (A.8) that P(V ) 1 1/2n,so(tn, hn) satisfy Ω n ≥ − i i i≥1 (A.6), concluding the proof. Notice that, despite the fact that (0, 1) is uncountable, i i S = supt∈(0,1) Lt is also measurable with respect to the (complete) sigma algebra : this follows from22 [24, Chapter 1, Theorem 4.14] and the identity Si > λ = F i i { i } ΠΩ( L > λ ), which holds for all λ R. It follows that ΠΩ( L > S /2 ) = i { } n ∈ i { } S > 0 = i = is -measurable, and since L is measurable we can apply the { } {J 6 ∅} F i i section theorem (with the constant filtration t := ) to L >S /2 and obtain a ¯n ¯n n F F n+{i+1 i} i random time ti such that P(ti = and i = ) 1/2 and L¯tn >S /2 (and ∞ J 6 ∅ ≤ 1 n ¯n ¯n n ¯n in particular i = and ti > 0) on ti < . We then define ti to equal ti on n J 6 ∅ { ∞} ¯n n n i = and to equal 0 otherwise. In particular ti = i = = ti = {J 6 ∅} n n+i+1 n { ∞}∩{Jn ¯n6 ∅}i { i ∞} and so P(ti = ) 1/2 , on ti (0, ) we have ti = ti , Ltn > S /2 and ∞ ≤ { ∈ ∞ } 1 n = , and finally tn = 0 = n = . Define {Ji 6 ∅} { i } {Ji ∅} G := (h, ω) (0, 1/2n] Ω : tn (0, ), [tn, tn + h] n,h>Si/2 , i { ∈ × i ∈ ∞ i i ∈Ji } which is measurable since Si, tn and (tn, hn)i−1 are -measurable and Y · is mea- i j j j=1 F t surable. Since tn (0, ) = Π (G ), by applying the section theorem to G { i ∈ ∞ } Ω i i we find a random time h¯n such that P(h¯n = and tn (0, )) 1/2n+i+1 and i i ∞ i ∈ ∞ ≤ (h¯n(ω),ω) G for ω h¯n < . We then define hn to equal 0 on tn = 0 and to i ∈ i ∈ { i ∞} i { i } equal h¯n otherwise. In particular hn = tn < = h¯n = tn (0, ) i { i ∞}∩{ i ∞} { i ∞}∩{ i ∈ ∞ } has probability at most 1/2n+i+1, so (A.8) holds. Notice that hn = 0 = tn = { i } { i 0 = n = , and on tn (0, ), hn < = Cn we have h¯n = hn < and } {Ji ∅} { i ∈ ∞ i ∞} i i i ∞ so hn > Si/2 > 0 and [tn, tn + h] n; thus we have defined random times tn, hn i i i ∈ Ji i i which satisfy (A.8) and the first 3 statements of (A.6), concluding the proof. 

We now strengthen the previous result as to make the quadratic variation to be exploding in all (non-trivial) intervals simultaneously.

Lemma A.2. There exist π [0, 1] such that n ∈ P lim B πn a.s. on σ<τ as n for all random times 0 σ τ 1 . n h i(σ,τ] →∞ { } →∞ ≤ ≤ ≤ Note that the previous equality is required to hold only on σ<τ since trivially { } B π =0 on σ = τ . h i(σ,τ] { }

21 i n i In Step 4 we use the measurability of L to prove that {Ji 6= ∅} = {S > 0} is F-measurable. 22For a proof of this result one can consult [5, Theorem A.5.10]. 26 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES

n−1 i Proof. Lemma 7.3 gives for each n N 0 and i = 0, 1,..., 2 some πn n n ∈ n\ { } −n 2n−1 i ∈ [i/2 , (i + 1)/2 ] such that P( B i 2 ) 2 . Define π := π [0, 1]. P h iπn ≤ ≤ n ∪i=0 n ∈ P Now given σ τ take ω τ σ> 2/2n , so there exists i = i(ω) such that ≤ ∈ { − } [i/2n, (i + 1)/2n](ω) [σ, τ](ω), ⊆ and hence, using (2.7), we see that B πn (ω) B πn (ω) h i[i/2n,(i+1)/2n] ≤ h i(σ,τ] and since π [i/2n, (i + 1)/2n]= πi [i/2n, (i + 1)/2n] we get that n ∩ n ∩ i πn πn B (ω)= B (ω)= B i (ω), h i[i/2n,(i+1)/2n] h i[i/2n,(i+1)/2n] h iπn Putting everything together we get that n πn n n 2 −1 n n B 2 and τ σ> 2/2 B i (ω) 2 and τ σ> 2/2 , {h i(σ,τ] ≤ − }⊆∪i=0 {h iπn ≤ − } and so the Borel-Cantelli lemma gives the result.  The following lemma states in probabilistic terms the fact that quadratic variation along the partition πd (resp. πu) is only slightly bigger than 0 (resp. 1). Lemma A.3. Given random times 0 σ τ 1, π [σ, τ] and a random ≤ ≤ ≤ ∈ P variable Z with values in N 0 , there exists π′ [σ, τ] such that π π′ and \ { } ∈ P ⊆ B ′ = B /Z. h iπ h iπ In particular for any ε> 0 there exist πd,πu [σ, τ] such that πd π πu, ∈ P ⊇ ⊆ 1 1 P B d > < ε and P B u / 1, 1+ and σ<τ < ε. h iπ Z h iπ ∈ Z       ′ Proof. By working separately on each subinterval [σi,σi+1] of π = (σi)i, to find π we can assume w.l.o.g. that σ = σ, σ = τ for all i N. Define 0 i+1 ∈ (B B )(i Z) σ′ := min t σ : B = τ − σ ∧ + B on B = B i { ≥ t Z σ} { τ 6 σ} ′ ′ ′ ′ and σ := σ, σ := τ on B = B ; then π π := (σ ) N [σ, τ] and 0 i+1 { τ σ} ⊆ i i∈ ∈ P 2 2 Bτ Bσ B π′ = (Bσ′ Bσ′ ) = − Z = B π/Z. h i i+1 − i Z h i Xi   ′ ′ Now fix ε> 0, and apply the previous result to find some π = πn such that

B ′ = B /n 0; h iπn h iπ → d ′ 1 taking π := π for n big enough shows that P( B d > ) < ε. n h iπ Z Finally let π be as in Lemma A.2 and let π′ [σ, τ] be such that π′ π and n n ∈ P n ⊇ n B ′ = B /Y where Y := max k N : k B 1. h iπn h iπn n n { ∈ ≤ h iπn }∨ ′ ′ Notice that B πn 1 + 1/Yn and on B πn 1 we have that B πn 1; moreover h i ≤ h i ≥ u ′ h i ≥ Yn B πn 1 a.s. on σ<τ . Taking π := πn for n big enough it follows ≥ h i − →∞ 1 { } that P B u / 1, 1+ and σ<τ < ε.  h iπ ∈ Z We now essentially prove the convergence at any fixed time. Lemma A.4. Given random times 0 σ τ 1 and a random variable Y with ≤ ≤ ≤ values in [0, ], there exist π [σ, τ] such that ∞ n ∈ P B πn Y a.s. on σ<τ as n . h i(σ,τ] → { } →∞ PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 27

Proof. On σ = τ we define π = σ for all n, and on Y = ,σ<τ we { } n { } { ∞ } set πn equal to the random partition πn given by lemma A.2; this clearly gives the thesis on Y = σ = τ . To conclude we will define π separately on each { ∞}∪{ } n Y [i/2n, (i + 1)/2n),σ<τ , i N. We want to buildπ ˜ [σ, τ] such that for { ∈ } ∈ n ∈ P all i N 0 , ∈ \ { } (A.9) P ( B / [i, i + 1),Y [i/2n, (i + 1)/2n) and σ<τ) < 1/2n+i ; h iπ˜n ∈ ∈ then if we define π to be for i N 0 the random partition π′ given by Lemma n ∈ \ { } A.3 with Z = 2n, and for i = 0 the random partition πd given by Lemma A.3 with ε = 1/2n and Z = 2n, since trivially B Y > 1/2n,Y [i/2n, (i + 1)/2n) B / [i/2n, (i + 1)/2n) Y , {|h iπn − | ∈ } ⊆ {h iπn ∈ ∋ } it follows that ∞ P( B Y > 1/2n,Y < and σ<τ) < 1/2n + 1/2n+i = 2/2n, |h iπn − | ∞ Xi=1 and so Borel-Cantelli’s lemma yields the thesis. n n We will construct suchπ ˜n separately on each Y [i/2 , (i+1)/2 ),σ<τ , i N { ∈ i,k } ∈ as the union over k = 0,...,i 1 of some partitions πn of some subintervals − [σi,k,σi,k+1]. First, we define the random times τ σ σi,k := − k + σ, i N 0 , k = 0,...,i i ∈ \ { } i,k i,k+1 i,k and notice that σ < σ on σ<τ , so we can use Lemma A.3 to find πn { } ∈ [σi,k,σi,k+1] such that P 1 i P B i,k / 1, 1+ and σ<τ < , h iπn ∈ Z 2n+i   n   where we take Z := j on Y [j/2n, (j + 1)/2n) for j N 0 and23 Z := 1 on n { ∈ } ∈ \ { } n Y [0, 1/2n) . Intersecting with Y [i/2n, (i + 1)/2n) shows in particular that { ∈ } { ∈ } n n i (A.10) P B i,k / [1, 1 + 1/i),Y [i/2 , (i + 1)/2 ) and σ<τ < . h iπn ∈ ∈ 2n+i  i i−1 i,k i,k i,k +1 Then we defineπ ˜ := πn , which belongs to [min σ , max σ ]= [σ, τ], n ∪k=0 P k k P and we set π˜ :=π ˜i on Y [i/2n, (i + 1)/2n) and σ<τ , i N 0 n n { ∈ } ∈ \ { } andπ ˜ := σ τ on Y [0, 1/2n) Y = σ = τ , so triviallyπ ˜ [σ, τ]. n { }∪{ } { ∈ }∪{ ∞}∪{ } n ∈ P Since (2.8) gives that for i N 0 ∈ \ { } i−1 n n B π˜ = B i,k on Y [i/2 , (i + 1)/2 ) and σ<τ h i n h iπn { ∈ } Xk=0 and since i−1 a / [i, i + 1) implies that a / [1, 1 + 1/i) for some k, from (A.10) k=0 k ∈ k ∈ summing over k and majorizing we obtain thatπ ˜ satisfies (A.9) for all i N 0 , P n ∈ \ { } concluding the proof.  To deal with the fact that A may take the value , we have decided to work ∞ with the distance d(a, b)= exp( a) exp( b) on [0, ], which is not invariant by | − − − | ∞ translations yet satisfies the following property.

23 n On {Y ∈ [0, 1/2 )} we can define Zn arbitrarily, as long as it takes values in N \ {0}. 28 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES

Lemma A.5. Given a , b [0, ], i = 1,...,n, we have i i ∈ ∞ n n n (A.11) d ai, bi d(ai, bi) . ! ≤ Xi=0 Xi=0 Xi=0 Proof. It is enough to prove this for n = 2 and then iterate. Since exp( t) is positive a − decreasing if 0 b a v and u 0 from d(a, b)= exp( t)dt we obtain ≤ ≤ ≤ ≤∞ ≥ b − (A.12) d(a + u, b + u) d(a, b) and d(a, b) d(Rv, b) . ≤ ≤ Assume w.l.o.g. that a b ; then we claim that 1 ≥ 1 if a < b then d(a + a , b + b ) max(d(a , b ), d(a , b )) : 2 2 1 2 1 2 ≤ 1 1 2 2 indeed if a + a b + b using (A.12) one has 1 2 ≥ 1 2 d(a + a , b + b ) d(a + b , b + b ) d(a , b ) , 1 2 1 2 ≤ 1 2 1 2 ≤ 1 1 and the other case is analogous. If a b by the linearity of the integral and (A.12) 2 ≥ 2 d(a + a , b + b )= d(a + a , a + b )+ d(a + b , b + b ) d(a , b )+ d(a , b ). 1 2 1 2 1 2 1 2 1 2 1 2 ≤ 2 2 1 1 

We can finally stitch all the pieces together. We define := 0, so in the ∞−∞ following proof the quantities A A (with t t ) are always well defined ti+1 − ti i ≤ i+1 and satisfy n−1 A A = A A . i=0 ti+1 − ti tn − t0 Proof of TheoremP 7.4. Thanks to Scholium A.1 it is enough to work on the time interval [0, 1]. For simplicity, we will first build (in step 1 and 2) πn which may fail to be refining and to includeπ ¯n but does satisfy the other assertions of the theorem. Step 1. To isolate the main idea from the technicalities we first deal with the case of continuous A, using the same notation as for the general case so as to be i able to refer back to this case later; denote byπ ˜n = (σn)i∈N the ‘random’ partition n 2 i/2n and notice that σi =1 if i ˜i := 2n, so it is enough to consider from ∪i=0{ } n ≥ now on i ˜i 1. Given π [σi ,σi+1] define ≤ − ∈ P n n i i π (A.13) ∆ (π)=∆ (π, A) := d( B i+1 , A i+1 A i ) n n i σ σn h i[σn,σn ] n − Notice that by definition ∆i (π, A)=0on σi = σi+1 ; thus, thanks to Lemma A.4, n { n n } we can find πi [σi ,σi+1] such that n ∈ P n n i i −n ˜ −n ˜ P(∆n(πn) > 2 /i) < 2 /i. ˜ Setting π := i−1πi and using (2.6) and Lemma A.5 we get, writing B πn and A n ∪i=0 n h it t as the sums of their increments over the subintervals ofπ ˜n

πn k−1 i i ˜i−1 i i d( B k , Aσk ) i=0 ∆n(πn) i=0 ∆n(πn) ; h iσn n ≤ ≤ if the sum over ˜i positive terms isP greater than 2−nPthen at least one summand is greater than 2−n/˜i and so

πn n ˜i−1 i i −n ˜ (A.14) max d( B t , At) > 1/2 i=0 ∆n(πn) > 2 /i , {t∈π˜n h i }⊆∪ { } and so we obtain that for k = n (A.15) P(max d( B πn , A ) > 1/2n) 1/2n. t∈π˜k h it t ≤ Since (˜πk)k is refining

πn πn max d( B t , At) max d( B t , At) , t∈π˜k h i ≤ t∈π˜n h i PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 29 and so (A.15) holds for all k n; thus for each fixed k we can apply Borel-Cantelli’s ≤ πn lemma to get that max d( B , A ) 0 a.s. as n , and since F := Nπ˜ t∈π˜k h it t → →∞ ∪k∈ k is dense in [0, 1] and contains the set of jumps of A (which in the case of step 1 is empty), Lemma 7.2 gives24 the convergence for all t [0, 1], uniformly over every ∈ interval where A is continuous (which in the case of step 1 is everywhere). Notice that sinceπ ˜ N/2n by construction τ i + 2−n is a stopping time, where π = (τ i ) . n ⊇ n n n i Step 2. We deal now with a general increasing process A. Consider the positive increasing process Dt := lims∈Q,s↓t 1 exp( As), which is c`adl`ag, bounded by 1, and − − ′ 0 has the same times of jump as A. Construct πn by setting τn := 0, i+1 i ′ i (A.16) τn := inf t>τn : Dt lim Ds > 1/n 1 , and πn := (τn)i. { − s↑t }∧

i ′ Notice that the τn are random times and πn are random partitions of [0, 1] (for an 25 ′ elementary proof see [25, Lemma 3.3]) and nπn contains all the times of jumps ∪ i of D, i.e. of A. Since D is increasing and D1 1 we have that τn = 1 for any i n. i ≤ ′ 2n n ≥ Now defineπ ˜ = (σ ) N [0, 1] as the random partition π ( i/2 ), notice n n i∈ ∈ P n ∪ ∪i=0{ } that σi = 1 for any i ˜i := n + 2n and that (˜π ) is refining. The proof given for n ≥ n n continuous A then applies word by word, giving (A.15) and the thesis. Step 3. Finally, we will now improve on the above proof and show that πn can be chosen to be refining and to includeπ ¯n. We will define (πn, π˜n)n by induction; more precisely we set π0 :=π ˜0 :=π ¯0, and for n 1 we will defineπ ˜n given (πk)k 0 there exists πn [σ ,σ ] such that n ∈ P n n i i,˜i n˜ ˜ n˜ P(∆n(πn , A) > 1/2 i) < P(K(˜πn)= i)/2 i.

i i,˜i ˜ ˜ ˜ Then we set πn := πn on K(˜πn) = i for each i such that P(K(˜πn) = i) > 0; i { } i this defines πn on a set of full measure, and on its complement we can define πn := σi σi+1 . Then πi belongs to [σi ,σi+1] and for every ˜i N 0 { n} ∪ { n } n P n n ∈ \ { } (A.17) P(∆i (πi , A n) > 1/2n˜i and K(˜π )= ˜i ) P(K(˜π )= ˜i)/2n˜i . n n ∧ n ≤ n i ˜ 26 Now we set πn := i∈Nπn and notice that πn on K(˜πn) = i equals the finite ˜ ∪ { } union i−1πi , and so by the same argument as for (A.14) we get that ∪i=0 n ˜i πn n ˜ Mn := max d( B t , At) > 1/2 and K(˜πn)= i {t∈π˜n h i } is a subset of ˜ i−1 ∆i (πi , A) > 1/2n˜i and K(˜π )= ˜i , ∪i=0 { n n n }

24As [0, ∞] is homeomorphic to [0, 1], Lemma 7.2 holds if a has values in [0, ∞] instead of [0, 1]. 25The cited lemma deals with stopping times and c`adl`ag adapted processes; these reduce to random times and c`adl`ag processes when considering a constant filtration. 26 i i+1 i ˜i−1 Indeed for i ≥ ˜i on {K(˜πn)= ˜i} we have σn =1= σn and so πn = {1}⊆ πn . 30 MARK DAVIS, JAN OBL OJ´ AND PIETRO SIORPAES

˜i ˜ n and thus (A.17) shows that Mn has probability smaller than P(K(˜πn)= i)/2 . The proof is concluded since

πn n ˜i ˜ n n P(max d( B t , At) > 1/2 )= P(Mn) P(K(˜πn)= i)/2 = 1/2 . t∈π˜n h i ≤ X˜i∈N X˜i∈N As before our construction gives that τ i + 2−n is a stopping time asπ ˜ N/2n.  n n ⊇

References

[1] R. F. Bass. The measurability of hitting times. Electronic Communications in Probability, 15:99–105, 2010. [2] R. F. Bass. Correction to the measurability of hitting times. Electronic Communications in Probability, 16:189–191, 2011. [3] J. Bertoin. Temps locaux et int´egration stochastique pour les processus de Dirichlet. In S´eminaire de Probabilit´es XXI, volume 1247 of Lecture Notes in Mathematics, pages 191–205. Springer, 1987. [4] K. Bichteler. Stochastic integration and Lp-theory of semimartingales. Annals of Probability, 9(1):49–89, 1981. [5] K. Bichteler. Stochastic integration with jumps, volume 89. Cambridge University Press, 2002. [6] H. Brezis. Functional Analysis, Sobolev Spaces and Partial Differential Equations. Springer, 2010. [7] R. Chacon, Y. Le Jan, E. Perkins, and S. J. Taylor. Generalized arc length for Brownian motion and L´evy processes. Z. Wahrsch. ver. Geb., 57:197–211, 1981. [8] L. Coutin, D. Nualart, and C. A. Tudor. Tanaka formula for the fractional Brownian motion. Stochastic Processes and their Applications, 94(2):301–315, 2001. [9] M. Davis, J. Ob l´oj, and V. Raval. Arbitrage bounds for prices of weighted variance swaps. Mathematical Finance, 24:821–854, 2014. [10] M. Emery.´ Une topologie sur l’espace des semimartingales. In S´eminaire de Probabilit´es XIII, volume 721 of Lecture Notes in Mathematics, pages 260–280. Springer, 1979. [11] C. Feng and H. Zhao. Two-parameter p, q-variation paths and integration of local times. Po- tential Analysis, 25(2):165–204, 2006. [12] H. F¨ollmer. Private communication. [13] H. F¨ollmer. Calcul d’Itˆosans probabilit´es. In S´eminaire de Probabilit´es XV, volume 850 of Lecture Notes in Mathematics, pages 143–150. Springer, 1981. [14] D. Freedman. Brownian Motion and Diffusion. Springer, 2nd edition, 1983. [15] J. Jacod and A. N. Shiryaev. Limit theorems for stochastic processes, volume 288 of Grundlehren der Mathematischen Wissenschaften [Fundamental Principles of Mathematical Sciences]. Springer-Verlag, Berlin, second edition, 2003. [16] R. Karandikar. On pathwise stochastic integration. Stochastic Processes and their Applications, 57(1):11–18, 1995. [17] I. Karatzas and S. E. Shreve. Brownian Motion and Stochastic Calculus, volume 113 of Graduate Texts in Mathematics. Springer, 2nd edition, 1991. [18] M. Lemieux. On the quadratic variation of semimartingales. PhD thesis, University of British Columbia, 1983. [19] G. Leoni. A first course in Sobolev spaces, volume 105 of Graduate Studies in Mathematics. American Mathematical Society, Providence, RI, 2009. [20] P. L´evy. Processus Stochastiques et Mouvement Brownien. Paris, 1965. [21] E. Perkins. A non-standard approach to Brownian local time. PhD thesis, University of Illinois at Urbana-Champaign, 1979. [22] N. Perkowski and D. J. Pr¨omel. Local times for typical price paths and pathwise Tanaka formulas. Electronic Journal of Probability, 20(46):1–15, 2015. [23] P. Protter. Stochastic Integration and Differential Equations, volume 21 of Applications of Mathematics. Springer, 2nd edition, 2004. [24] D. Revuz and M. Yor. Continuous Martingales and Brownian Motion, volume 293 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, Berlin, 3rd edition, 1999. [25] P. Siorpaes. On a dyadic approximation of predictable processes of finite variation. Electronic Communications in Probability, 19:1–12, 2014. PATHWISE STOCHASTIC CALCULUS WITH LOCAL TIMES 31

[26] D. Sondermann. In Introduction to Stochastic Calculus for Finance, volume 579 of Lecture Notes in Economics and Mathematical Systems. Springer, 2006. [27] T. Sottinen and L. Viitasaari. Pathwise integrals and Itˆo–Tanaka formula for Gaussian pro- cesses. Journal of Theoretical Probability, 29:590–616, 2016. [28] C. Stricker. Quasi-martingales et variations. In S´eminaire de Probabilit´es XV, volume 850 of Lecture Notes in Mathematics, pages 493–498. Springer, 1981. [29] S. J. Taylor. Exact asymptotic estimates of Brownian path variation. Duke Math. J., 39:219– 241, 1972. [30] V. Vovk. Continuous-time trading and the emergence of probability. Finance and Stochastics, 16(4):561–609, 2012. [31] M. W¨urmli. Lokalzeiten f¨ur Martingale. Diplomarbeit, ETH Z¨urich, 1980. [32] J.-Y. Yen and M. Yor. Local Times and Excursion Theory for Brownian Motion, volume 2088 of Lecture Notes in Mathematics. Springer, 2013.

Davis: Department of Mathematics, Imperial College London, London SW7 2AZ E-mail address: [email protected] URL: http://www.imperial.ac.uk/people/mark.davis

Obl oj,´ Siorpaes: Mathematical Institute, University of Oxford, Oxford OX2 6GG E-mail address: [email protected], [email protected] URL: www.maths.ox.ac.uk/people/jan.obloj, www.maths.ox.ac.uk/people/pietro.siorpaes