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

Arbitrage of the first kind and filtration enlargements in semimartingale financial models

Beatrice Acciaio Beatrice Acciaio, Statistics Department, London School of Economics and Political Science b.acciaio@lse.ac.uk Claudio Fontana Claudio Fontana, Laboratoire de Probabilités et Modèles Aléatoires, Paris Diderot University fontana@math.univ-paris-diderot.fr  and  Constantinos Kardaras Constantinos Kardaras, Statistics Department, London School of Economics and Political Science k.kardaras@lse.ac.uk
(Date: August 16, 2025)
Abstract.

In a general semimartingale financial model, we study the stability of the No Arbitrage of the First Kind (NA1) (or, equivalently, No Unbounded Profit with Bounded Risk) condition under initial and under progressive filtration enlargements. In both cases, we provide a simple and general condition which is sufficient to ensure this stability for any fixed semimartingale model. Furthermore, we give a characterisation of the NA1 stability for all semimartingale models.

Key words and phrases:
Progressive enlargement of filtrations; initial enlargement of filtrations; arbitrage of the first kind; martingale deflator
2010 Mathematics Subject Classification:
60G44, 91G10
The research of the second author was supported by a Marie Curie Intra European Fellowship within the 7th European Community Framework Programme under grant agreement PIEF-GA-2012-332345.

Introduction

In financial mathematics, market models with different sets of information have been widely studied, especially in relation to insider trading and credit risk modeling (see e.g. [JYC09] and the references therein). Typically, one starts by postulating a model with respect to a given information set and then enlarges that set with some additional information not originally present in the market. From a mathematical point of view, this corresponds to considering an enlargement of the original filtration on a given filtered probability space. Since the model aims at representing a financial market, a fundamental question is whether the additional information allows for arbitrage profits.

The present paper aims at answering the above question in the context of models driven by general semimartingales, both in the case where the additional information is added in a progressive way through time, and in the case where the additional information is fully added at the initial time. Referring to the terminology of the theory of enlargement of filtrations (see [Jeu80] for a complete account of the theory and [JYC09, § 5.9] and [Pro04, Ch. VI] for a presentation of the main results), this corresponds to considering a filtration obtained as a progressive or as an initial enlargement, respectively, of the original filtration.

Our analysis focuses on the No Arbitrage of the First Kind (NA1) condition (see [Kar10]), which is equivalent to the No Unbounded Profit with Bounded Risk (NUPBR) condition (see [Kar10, Proposition 1]). Mathematically, condition NA1 is equivalent to existence of strictly positive local martingale deflators, and can be shown to be the minimal condition ensuring the well-posedness of expected utility maximisation problems (see [KK07, Proposition 4.19]). In the case of a progressive enlargement with respect to a random time τ\tau, we study the stability of NA1 on the random time horizon [0,τ][0,\tau], showing that the existence of arbitrages of the first kind in the enlarged filtration is crucially linked to the possibility of the asset-price process exhibiting a jump at the same time when a particular nonnegative local martingale in the original filtration jumps to zero. In turn, we show that the possibility of the latter event is intimately related to how local martingales from the original filtration behave in the enlarged filtration, up to a suitable normalisation. In the case of an initial enlargement of the original filtration, and under the classical density hypothesis of [Jac85], we establish an analogous set of results, showing that the validity of NA1 in the enlarged filtration is linked to the possibility of the asset-price process jumping at the same time when a family of nonnegative martingales in the original filtration jumps to zero. In turn, as in the case of progressive enlargements, the latter possibility also fully characterises how local martingales from the original filtration behave in the enlarged filtration, up to a suitable normalisation.

In both cases of progressive and of initial enlargement, these results allow us to provide an easy sufficient condition ensuring the NA1 stability for a fixed semimartingale model, as well as to explicitly characterise the stability of NA1 for all semimartingale models. Although absent in the statements of our main results, an inspection of their proofs reveals a hands-on approach to the problem: using local martingale deflators in the original filtration, we explicitly construct local martingale deflators in the enlarged filtration in order to show validity of condition NA1. In the process, we obtain some interesting new results on progressive as well as initial filtration enlargement, showing how the super/local martingale property of a process can be transferred from the original filtration to the enlarged one by suitably deflating the process.

For progressive filtration enlargement with respect to an honest time τ\tau (see [Pro04, Ch. VI]), examples of arbitrage profits are provided in [Imk02], [Zwi07] and [FJS14]. In the context of continuous semimartingale models, as shown in [FJS14, Theorem 4.1] (see also [Kre13, Lemma 6.7]), condition NA1 is always valid in the enlarged filtration on the random time horizon [0,τ][0,\tau]. In the case of general semimartingale models, this is no longer true, see the example in § 1.5.1. In that context, the recent paper [ACDJ14] addresses the issue of NA1 stability in progressively enlarged filtrations and represents one of the sources of inspiration for the present work. In particular, the key role of conditions equivalent to those given in Theorem 1.4 and Remark 1.5 has been first pointed out and proved in [ACDJ14] (see Remark 1.6) and the characterisation we obtain in Theorem 1.7 turns out to be equivalent to the one already established in [ACDJ14] (see Remark 1.8). However, in comparison with the latter paper, we follow here a totally different approach and provide original and rather simple proofs to those results, avoiding the use of the compensated stochastic integral (see e.g. [HWY92, Definition 9.7]) and, somewhat surprisingly, not relying on the classical Jeulin-Yor decomposition formula (see [Jeu80, Proposition 4.16]). In contrast, we exploit the properties of an optional decomposition of the Azéma supermartingale associated to τ\tau recently established in [Kar15]. We also want to mention that, in the case of the classical No Free Lunch with Vanishing Risk (NFLVR) condition (see [DS94, DS98]), a study of its stability and of the relation with the preservation of the martingale property in progressively enlarged filtrations has been carried out in [CJN12].

In the initial filtration enlargement case, the possibility of realising arbitrage profits in the enlarged filtration has been studied in [GP98], [GP01] and [IPW01], among others. Concerning the classical NFLVR condition, it is well-known that it is stable under an initial enlargement with respect to a random variable JJ if the conditional law of JJ for all times is equivalent to the unconditional one (see e.g. [GP98]). However, to the best of our knowledge, the issue of NA1 stability with respect to an initial enlargement has never been studied so far. Interestingly, we show that both the progressive and the initial case can be treated by relying on the same methodological approach.

The paper is organised as follows. Section 1 contains the framework and statements of our main results. In Section 2 we consider progressive enlargement of filtrations. We study the crucial stopping times that will be then used to pinpoint local martingales and to prove stability of the NA1 condition in the enlarged filtrations. In Section 3 we perform the same analysis and obtain analogous results, mutatis mutandis, in the case of initially enlarged filtrations.

1. Main Results

1.1. Probabilistic set-up

In all that follows, we work on a filtered probability space (Ω,,𝐅,)(\Omega,\,\mathcal{F},\,\mathbf{F},\,\mathbb{P}), where 𝐅=(t)t+\mathbf{F}=(\mathcal{F}_{t})_{t\in\mathbb{R}_{+}} is a filtration satisfying the usual hypotheses of right-continuity and saturation by \mathbb{P}-null sets. In general, \mathcal{F}_{\infty}\subseteq\mathcal{F} holds, with the last set-inclusion being potentially strict.

We shall be using standard notation from the general theory of stochastic processes. For any unexplained notation and results, the reader can consult [HWY92] or [JS03].

1.2. The market model

Fix d={1,2,}d\in\mathbb{N}=\left\{1,2,\ldots\right\}, and let S(Si)i{1,,d}S\equiv(S^{i})_{i\in\left\{1,\ldots,d\right\}} be a collection of nonnegative semimartingales on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) 111We want to mention that the nonnegativity assumption is not crucial for the following results to hold, provided that the notion of local martingale is suitably replaced by the notion of sigma-martingale (see [DS98] and [TS14]).. Each SiS^{i}, i{1,,d}{i\in\left\{1,\ldots,d\right\}}, models the price process of an asset, discounted by a baseline security in the market. Starting with initial capital x[0,)x\in[0,\infty) and following a dd-dimensional, 𝐅\mathbf{F}-predictable and SS-integrable strategy HH, an investor’s discounted wealth process is given by Xx,H:=x+0(Ht,dSt)X^{x,H}\,:=\,x+\int_{0}^{\cdot}\left(H_{t},\mathrm{d}S_{t}\right). It should be noted that we are using vector stochastic integration throughout. Define 𝒳(𝐅,S)\mathcal{X}(\mathbf{F},S) to be the class of all nonnegative processes Xx,HX^{x,H} in the previous notation. (In the definition of the class 𝒳(𝐅,S)\mathcal{X}(\mathbf{F},S), the initial capital x[0,)x\in[0,\infty) and dd-dimensional, 𝐅\mathbf{F}-predictable and SS-integrable strategies HH are arbitrary, as long as Xx,H0X^{x,H}\geq 0.)

Definition 1.1.

For T(0,)T\in(0,\infty), an arbitrage of the first kind with information 𝐅\mathbf{F} and assets SS on [0,T][0,T] is χT𝕃+0(T)\chi_{T}\in{\mathbb{L}^{0}_{+}}(\mathcal{F}_{T}) with [χT>0]>0\mathbb{P}\left[\chi_{T}>0\right]>0 and with the property that for all x(0,)x\in(0,\infty) there exists X𝒳(𝐅,S)X\in\mathcal{X}(\mathbf{F},S) with X0=xX_{0}=x (where the wealth process XX may depend on xx) such that [XTχT]=1\mathbb{P}\left[X_{T}\geq\chi_{T}\right]=1. If no arbitrage of the first kind with information 𝐅\mathbf{F} and assets SS exists on any interval [0,T][0,T] for T(0,)T\in(0,\infty), we say that condition NA(𝐅,S)1{}_{1}(\mathbf{F},S) holds.

Whenever \mathbb{Q}\sim\mathbb{P}, we use 𝒴(𝐅,S,)\mathcal{Y}(\mathbf{F},S,\mathbb{Q}) to denote the class of all strictly positive 𝐅\mathbf{F}-adapted càdlàg processes YY with Y0=1Y_{0}=1, such that YY and YSYS are local martingales on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{Q}). The elements in 𝒴(𝐅,S,)\mathcal{Y}(\mathbf{F},S,\mathbb{Q}) are called strictly positive local martingale deflators (for SS on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{Q})). When strict positivity is replaced by nonnegativity, we simply talk of local martingale deflators. If YY^{\mathbb{Q}} denotes the density process of \mathbb{Q} with respect to \mathbb{P}, note that 𝒴(𝐅,S,)={Y(Y0/Y)|Y𝒴(𝐅,S,)}\mathcal{Y}(\mathbf{F},S,\mathbb{Q})=\big{\{}Y{(Y^{\mathbb{Q}}_{0}/Y^{\mathbb{Q}})}\ |\ Y\in\mathcal{Y}(\mathbf{F},S,\mathbb{P})\big{\}} holds. It comes as a consequence of [TS14, Theorem 2.6] that condition NA(𝐅,S)1{}_{1}(\mathbf{F},S) is equivalent to 𝒴(𝐅,S,)\mathcal{Y}(\mathbf{F},S,\mathbb{Q})\neq\emptyset (where, of course, \mathbb{Q}\sim\mathbb{P} is arbitrary). For our purposes (see Remark 1.3 below), we need a more precise statement.

Theorem 1.2.

Condition NA(𝐅,S)1{}_{1}(\mathbf{F},S) holds if and only if there exist \mathbb{Q}\sim\mathbb{P} and strictly positive X^𝒳(𝐅,S)\widehat{X}\in\mathcal{X}(\mathbf{F},S) such that (1/X^)𝒴(𝐅,S,)(1/\widehat{X})\in\mathcal{Y}(\mathbf{F},S,\mathbb{Q}).

Note that, even though the statement of Theorem 1.2 is sharper than [TS14, Theorem 2.6], it actually follows from the proof of the latter. Indeed, [TS14] prove that NA(𝐅,S)1{}_{1}(\mathbf{F},S) implies the existence of \mathbb{Q}\sim\mathbb{P} and strictly positive X^𝒳(𝐅,S)\widehat{X}\in\mathcal{X}(\mathbf{F},S) such that Y/(X^Y0)𝒴(𝐅,S,)Y^{\mathbb{Q}}/(\widehat{X}Y^{\mathbb{Q}}_{0})\in\mathcal{Y}(\mathbf{F},S,\mathbb{P}), with YY^{\mathbb{Q}} denoting the density process of \mathbb{Q} with respect to \mathbb{P}.

The main purpose of the paper is the study of stability of the NA1 condition when enlarging the filtration 𝐅\mathbf{F} in a progressive or initial way. Naturally, the first issue to be settled is the preservation of the semimartingale property of processes, which is typically referred to in the literature as the \mathcal{H}^{\prime}-hypothesis. In the case of progressive filtration enlargement by a random time τ\tau, it comes as a consequence of the Jeulin-Yor theorem that this always holds up to time τ\tau (and that for honest times it holds on all [0,)[0,\infty)); see [JY78]. For the case of initial filtration expansion, one well-known situation where the preservation of the semimartingale property holds is when Jacod’s density hypothesis is satisfied; see [Jac85]. We want to remark that these facts will also come as consequences of our analysis in Section 2 for the progressive enlargement case (see Corollary 2.9) and Section 3 for the initial enlargement case (see Remark 3.5).

Remark 1.3.

Theorem 1.2 will play a key role in the proof of our main theorems. In fact, it shows that NA1(𝐅,S)(\mathbf{F},S) is equivalent to the existence of Y𝒴(𝐅,S,)Y\in\mathcal{Y}(\mathbf{F},S,\mathbb{Q}) such that {ΔS0}={ΔY0}\left\{\Delta S\neq 0\right\}=\left\{\Delta Y\neq 0\right\}, for some \mathbb{Q}\sim\mathbb{P}. As shown below (see Sections 2.4 and 3.3), this property turns out to be crucial in order to construct local martingale deflators in enlarged filtrations starting from local martingale deflators from the original filtration (compare also with Remarks 2.11 and 3.8).

1.3. Main results under progressive filtration enlargement

We first study the stability of the NA1 condition under a progressive enlargement of the filtration 𝐅\mathbf{F} with respect to an \mathcal{F}-measurable random time τ:Ω[0,]\tau:\Omega\mapsto[0,\infty] such that [τ=]=0\mathbb{P}\left[\tau=\infty\right]=0. (We refer the reader to [Pro04, Chapter VI] for a textbook account of the theory of enlargement of filtrations.) The progressively enlarged filtration 𝐆=(𝒢t)t+\mathbf{G}=(\mathcal{G}_{t})_{t\in\mathbb{R}_{+}} is defined via

(1.1) 𝒢t={B|B{τ>t}=Bt{τ>t} for some Btt},t+.\mathcal{G}_{t}=\left\{B\in\mathcal{F}\ |\ B\cap\left\{\tau>t\right\}=B_{t}\cap\left\{\tau>t\right\}\text{ for some }B_{t}\in\mathcal{F}_{t}\right\},\quad\forall t\in\mathbb{R}_{+}.

In particular, 𝐆\mathbf{G} is a right-continuous filtration that contains 𝐅\mathbf{F} and makes τ\tau a stopping time, but note that 𝐆\mathbf{G} is not the smallest right-continuous filtration that contains 𝐅\mathbf{F} and makes τ\tau a stopping time, compare e.g. the discussion in [GZ08].

It comes as a consequence of the Jeulin-Yor theorem that Sτ:=(Sτt)t+S^{\tau}\,:=\,\left(S_{\tau\wedge t}\right)_{t\in\mathbb{R}_{+}} is a semimartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}) (see, for example, [JY78]; actually, we shall provide an alternative simple proof of this fundamental fact in Corollary 2.9). Then, the class 𝒳(𝐆,Sτ)\mathcal{X}(\mathbf{G},S^{\tau}) can be defined exactly in the same way as the corresponding class 𝒳(𝐅,S)\mathcal{X}(\mathbf{F},S) of § 1.2. The notation NA(𝐆,Sτ)1{}_{1}(\mathbf{G},S^{\tau}) used in the sequel refers to absence of arbitrage of the first kind with information 𝐆\mathbf{G} and assets SτS^{\tau}.

A key role in the study of progressive enlargement of filtrations is played by the Azéma supermartingale associated with τ\tau (given by the optional projection of 𝕀[[0,τ[[\mathbb{I}_{[\kern-1.22911pt[0,\tau[\kern-1.22911pt[} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), see [Jeu80] and references therein), that we denote by ZZ. This means that [τ>σ|σ]=Zσ\mathbb{P}\left[\tau>\sigma\ |\ \mathcal{F}_{\sigma}\right]=Z_{\sigma} for all finite stopping times σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}), and note that Z:=limtZt=0Z_{\infty}:=\lim_{t\rightarrow\infty}Z_{t}=0 holds in view of [τ=]=0\mathbb{P}\left[\tau=\infty\right]=0 (note that the limit ZZ_{\infty} always exists due to the supermartingale convergence theorem). Furthermore, if AA denotes the dual optional projection of 𝕀[[τ,[[\mathbb{I}_{[\kern-1.22911pt[\tau,\infty[\kern-1.22911pt[}, it follows that μ:=A+Z\mu:=A+Z is a nonnegative uniformly integrable martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) with μt=𝔼[A|t]\mu_{t}=\mathbb{E}\left[A_{\infty}|\mathcal{F}_{t}\right], for all t0t\geq 0 (see e.g. [Nik06, Section 8.2]). Moreover, by the general properties of the dual optional projection (see e.g. [HWY92, Theorem 5.27]), for any stopping time σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}), it holds that ΔAσ=[τ=σ|σ]\Delta A_{\sigma}=\mathbb{P}\left[\tau=\sigma\ |\ \mathcal{F}_{\sigma}\right] on {σ<}\left\{\sigma<\infty\right\}.

For all nn\in\mathbb{N}, let ζn:=inf{t+|Zt<1/n}\zeta_{n}\,:=\,\inf\left\{t\in\mathbb{R}_{+}\ |\ Z_{t}<1/n\right\}. Furthermore, set

(1.2) ζ:=limnζn=inf{t+|Zt=0 or Zt=0}=inf{t+|Zt=0},\zeta\,:=\,\lim_{n\to\infty}\zeta_{n}=\inf\left\{t\in\mathbb{R}_{+}\ |\ Z_{t-}=0\text{ or }Z_{t}=0\right\}=\inf\left\{t\in\mathbb{R}_{+}\ |\ Z_{t}=0\right\},

where the last equality holds from the fact that ZZ is a nonnegative supermartingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). We now introduce a stopping time that will be of major importance in the sequel. Consider the ζ\mathcal{F}_{\zeta}-measurable event Λ:={ζ<,Zζ>0,ΔAζ=0}\Lambda\,:=\,\left\{\zeta<\infty,\,Z_{\zeta-}>0,\,\Delta A_{\zeta}=0\right\}, and define

(1.3) η:=ζΛ=ζ𝕀Λ+𝕀ΩΛ.\eta\,:=\,\zeta_{\Lambda}=\zeta\mathbb{I}_{\Lambda}+\infty\mathbb{I}_{\Omega\setminus\Lambda}.

Clearly, η\eta is a stopping time on (Ω,𝐅)(\Omega,\,\mathbf{F}), and it satisfies [η>τ]=1\mathbb{P}\left[\eta>\tau\right]=1. Indeed, [τ>η|η]=Zη=0\mathbb{P}\left[\tau>\eta|\mathcal{F}_{\eta}\right]=Z_{\eta}=0 and [τ=η<|η]=ΔAη𝕀{η<}=ΔAζ𝕀Λ=0\mathbb{P}\left[\tau=\eta<\infty|\mathcal{F}_{\eta}\right]=\Delta A_{\eta}\mathbb{I}_{\{\eta<\infty\}}=\Delta A_{\zeta}\mathbb{I}_{\Lambda}=0 (remember that [τ=]=0\mathbb{P}\left[\tau=\infty\right]=0 by assumption). In § 1.5, it is shown that η\eta may be totally inaccessible or accessible. However, Lemma 2.5 shows that [η=σ<|σ]<1\mathbb{P}\left[\eta=\sigma<\infty\ |\ \mathcal{F}_{\sigma-}\right]<1 holds for all predictable times σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}).

The results below establish stability of condition NA1 in the current setting of progressive filtration enlargement. Together with their counterparts for initially enlarged filtrations (Theorems 1.11 and 1.12), they are the main results of this paper.

The first result is concerned with stability of the NA1 condition for a fixed semimartingale model.

Theorem 1.4.

If NA(𝐅,S)1{}_{1}(\mathbf{F},S) holds and [η<,ΔSη0]=0\mathbb{P}\left[\eta<\infty,\Delta S_{\eta}\neq 0\right]=0, then NA(𝐆,Sτ)1{}_{1}(\mathbf{G},S^{\tau}) holds.

Remark 1.5.

The message of the above theorem is that, to ensure the preservation of NA1 under progressive filtration enlargement, one only needs to check whether the price process jumps at time η\eta. It is then clear that, if NA1(𝐅,S~)(\mathbf{F},\widetilde{S}) holds for S~:=Sη=SηΔSη𝕀[[η,[[\widetilde{S}:=S^{\eta-}=S^{\eta}-\Delta S_{\eta}\mathbb{I}_{[\kern-1.22911pt[\eta,\infty[\kern-1.22911pt[}, then NA(𝐆,Sτ)1{}_{1}(\mathbf{G},S^{\tau}) holds as well, since [η>τ]=1\mathbb{P}\left[\eta>\tau\right]=1. Actually, in order to have NA(𝐆,Sτ)1{}_{1}(\mathbf{G},S^{\tau}), it is sufficient that NA1(𝐅,S~ζn)(\mathbf{F},\widetilde{S}^{\zeta_{n}}) holds for all nn\in\mathbb{N}. Indeed, note that NA1(𝐅,S~ζn)(\mathbf{F},\widetilde{S}^{\zeta_{n}}) implies NA1(𝐆,Sτζn)(\mathbf{G},S^{\tau\wedge\zeta_{n}}), and that the intervals [[0,τζn]][\kern-1.49994pt[0,\tau\wedge\zeta_{n}]\kern-1.49994pt] exhaust [[0,τ]][\kern-1.49994pt[0,\tau]\kern-1.49994pt], since [ζτ]=1\mathbb{P}\left[\zeta\geq\tau\right]=1. Now the claim follows since the NA1 condition can be given locally222Here we provide a proof by way of contradiction. Assume there are T(0,)T\in(0,\infty) and χT𝕃+0(𝒢T)\chi_{T}\in{\mathbb{L}^{0}_{+}}(\mathcal{G}_{T}) such that [χT>0]>0\mathbb{P}\left[\chi_{T}>0\right]>0, satisfying the condition that, for all x(0,)x\in(0,\infty), there exists Xx=x+(HxSτ)𝒳(𝐆,Sτ)X^{x}=x+(H^{x}\cdot S^{\tau})\in\mathcal{X}(\mathbf{G},S^{\tau}) with [XTxχT]=1\mathbb{P}\left[X^{x}_{T}\geq\chi_{T}\right]=1. Consider the set A:={χT>0}A:=\{\chi_{T}>0\} and take nn big enough such that B:={ζnTτT}A𝒢TB:=\{\zeta_{n}\wedge T\geq\tau\wedge T\}\cap A\in\mathcal{G}_{T} satisfies [B]>0\mathbb{P}\left[B\right]>0. Note that ψT:=χT𝕀B𝕃+0(𝒢T)\psi_{T}:=\chi_{T}\mathbb{I}_{B}\in{\mathbb{L}^{0}_{+}}(\mathcal{G}_{T}) is such that [ψT>0]>0\mathbb{P}\left[\psi_{T}>0\right]>0. Now, for every x(0,)x\in(0,\infty), define the process Yx:=x+(HxSτζn)𝒳(𝐆,Sτζn)Y^{x}:=x+(H^{x}\cdot S^{\tau\wedge\zeta_{n}})\in\mathcal{X}(\mathbf{G},S^{\tau\wedge\zeta_{n}}). By definition of admissibility, [YTx0]=1\mathbb{P}\left[Y^{x}_{T}\geq 0\right]=1. Moreover, on BB we have YTx=x+(HxSτζn)T=x+(HxSτ)TχT=ψTY^{x}_{T}=x+(H^{x}\cdot S^{\tau\wedge\zeta_{n}})_{T}=x+(H^{x}\cdot S^{\tau})_{T}\geq\chi_{T}=\psi_{T}. Altogether this gives [YTxψT]=1\mathbb{P}\left[Y^{x}_{T}\geq\psi_{T}\right]=1, which is in contradiction to NA1(𝐆,Sτζn)(\mathbf{G},S^{\tau\wedge\zeta_{n}})..

Remark 1.6.

Define Z~\widetilde{Z} to be the optional projection of 𝕀[[0,τ]]\mathbb{I}_{[\kern-1.22911pt[0,\tau]\kern-1.22911pt]} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) (see also [Jeu80, Section IV.1]); in other words, for any stopping time σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}), Z~σ=[τσ|σ]\widetilde{Z}_{\sigma}=\mathbb{P}\left[\tau\geq\sigma\ |\ \mathcal{F}_{\sigma}\right] holds on {σ<}\left\{\sigma<\infty\right\}, so that Z~=Z+ΔA\widetilde{Z}=Z+\Delta A. It is then straightforward to see that condition [η<,ΔSη0]=0\mathbb{P}\left[\eta<\infty,\,\Delta S_{\eta}\neq 0\right]=0 is equivalent to evanescence of the set {Z>0,Z~=0,ΔS0}\{Z_{-}>0,\,\widetilde{Z}=0,\,\Delta S\neq 0\}. Hence, Theorem 1.4 corresponds exactly to the result proved in [ACDJ14, Corollary 2.20, part (b)], by means of different techniques. Moreover, when SS is a quasi-left-continuous semimartingale (see [JS03, Definition I.2.25]), [ACDJ14, Theorem 2.8] shows that the validity of NA1(𝐅,S~ζn)(\mathbf{F},\widetilde{S}^{\zeta_{n}}), for all nn\in\mathbb{N}, is actually necessary and sufficient for the preservation of the NA1 property in 𝐆\mathbf{G} (see also [ACDJ14, Remark 2.9]).

Theorem 1.4 recovers the already-known fact that condition NA1 is stable under progressive enlargement for all continuous semimartingales; see [FJS14] and [Kre13]. Moreover, it implies that the condition [η<]=0\mathbb{P}\left[\eta<\infty\right]=0 is sufficient to guarantee NA1 stability for any collection of asset-price processes. In the next result we show that this condition is also necessary in order to have this general stability. In fact, for [η<]>0\mathbb{P}\left[\eta<\infty\right]>0, we provide an explicit example of arbitrage of the first kind, which further shows how condition [η<,ΔSη0]=0\mathbb{P}\left[\eta<\infty,\,\Delta S_{\eta}\neq 0\right]=0 in Theorem 1.4 cannot be dropped; see also § 1.5.1. Statement (1) of the following theorem is an immediate consequence of Theorem 1.4, while the proof of statement (2) is given in Section 2.4.

Theorem 1.7.

The following statements hold true:

  1. (1)

    If [η<]=0\mathbb{P}\left[\eta<\infty\right]=0, then for any SS such that NA(𝐅,S)1{}_{1}(\mathbf{F},S) holds, NA(𝐆,Sτ)1{}_{1}(\mathbf{G},S^{\tau}) also holds.

  2. (2)

    Suppose that [η<]>0\mathbb{P}\left[\eta<\infty\right]>0. Then, with DD being the predictable compensator of 𝕀[[η,[[\mathbb{I}_{[\kern-1.22911pt[\eta,\infty[\kern-1.22911pt[} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), the nonnegative process S:=(D)1𝕀[[0,η[[S\,:=\,\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[} is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), and SτS^{\tau} is nondecreasing with [Sτ>1]>0\mathbb{P}\left[S_{\tau}>1\right]>0. In particular, condition NA(𝐅,S)1{}_{1}(\mathbf{F},S) holds but condition NA(𝐆,Sτ)1{}_{1}(\mathbf{G},S^{\tau}) fails.

Remark 1.8.

Similarly to the discussion in Remark 1.6, condition [η<]=0\mathbb{P}\left[\eta<\infty\right]=0 is equivalent to evanescence of the set {Z>0,Z~=0}={Z>0,Z=0,ΔA=0}\{Z_{-}>0,\,\widetilde{Z}=0\}=\{Z_{-}>0,\,Z=0,\,\Delta A=0\}. Therefore, the characterisation we obtain in Theorem 1.7 is equivalent to that proved in [ACDJ14, Theorem 2.22], by means of different techniques.

Section 2 is devoted to the proof of Theorem 1.4 and Theorem 1.7; several interesting side results are also included there. In § 1.5 that follows, a couple of illustrative examples are given.

1.4. Main results under initial filtration enlargement

We now study the stability of condition NA1 under an initial enlargement of the filtration 𝐅\mathbf{F} with respect to an \mathcal{F}-measurable random variable JJ taking values in a Lusin space (E,E)(E,\mathcal{B}_{E}), where E\mathcal{B}_{E} denotes the Borel σ\sigma-field of EE. With some abuse of notation, we denote by 𝐆=(𝒢t)t+\mathbf{G}=(\mathcal{G}_{t})_{t\in\mathbb{R}_{+}} the right-continuous augmentation of the filtration 𝐆0=(𝒢t0)t+\mathbf{G}^{0}=(\mathcal{G}^{0}_{t})_{t\in\mathbb{R}_{+}} defined by 𝒢t0:=tσ(J)\mathcal{G}^{0}_{t}:=\mathcal{F}_{t}\vee\sigma(J), for all t+t\in\mathbb{R}_{+}. Let γ:E[0,1]\gamma:\mathcal{B}_{E}\mapsto[0,1] be the law of JJ (so that γ[B]=[JB]\gamma\left[B\right]=\mathbb{P}\left[J\in B\right] holds for all BEB\in\mathcal{B}_{E}). Furthermore, for all t+t\in\mathbb{R}_{+}, let γt:Ω×E[0,1]\gamma_{t}:\Omega\times\mathcal{B}_{E}\mapsto[0,1] be a regular version of the t\mathcal{F}_{t}-conditional law of JJ, which exists since (E,E)(E,\mathcal{B}_{E}) is Lusin.

Assumption 1.9.

Throughout §1.4, we work under the following condition:

  • (J)

    for all t+t\in\mathbb{R}_{+}, γtγ\gamma_{t}\ll\gamma holds in the \mathbb{P}-a.s. sense.

Assumption 1.9 is the classical density hypothesis introduced in [Jac85]. Indeed, as shown in [Jac85, Proposition 1.5] (see also [Pro04, Theorem VI.11]), condition (J) holds if and only if, for all t+t\in\mathbb{R}_{+} there exists a σ\sigma-finite measure νt\nu_{t} on (E,E)(E,\mathcal{B}_{E}) such that γtνt\gamma_{t}\ll\nu_{t} holds in the \mathbb{P}-a.s. sense. Jacod’s density hypothesis plays a prominent role in financial mathematics, notably in relation to the modeling of additional information (see e.g. [AIS98, GP98, GP01, Bau03, GVV06, KH07, KHOL11]).

The next auxiliary result (the proof of which is postponed to Section 3) implies the existence a good version of conditional densities. It essentially corresponds to [Jac85, Lemma 1.8] (see also [Ame00, Appendix A.1]). Note that 𝒪(𝐅)\mathcal{O}(\mathbf{F}) denotes the 𝐅\mathbf{F}-optional σ\sigma-field on Ω×+\Omega\times\mathbb{R}_{+}.

Lemma 1.10.

There exists a (E𝒪(𝐅))\left(\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})\right)-measurable function E×Ω×+(x,ω,t)ptx(ω)[0,)E\times\Omega\times\mathbb{R}_{+}\ni(x,\omega,t)\mapsto p^{x}_{t}(\omega)\in[0,\infty), càdlàg in t+t\in\mathbb{R}_{+} and such that:

  • (i)

    for every t+t\in\mathbb{R}_{+}, γt(dx)=ptxγ(dx)\gamma_{t}(\mathrm{d}x)=p^{x}_{t}\,\gamma(\mathrm{d}x) holds \mathbb{P}-a.s;

  • (ii)

    for every xEx\in E, the process px=(ptx)t+p^{x}=(p^{x}_{t})_{t\in\mathbb{R}_{+}} is a martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}).

For every xEx\in E and nn\in\mathbb{N}, define families of stopping times on (Ω,𝐅)(\Omega,\,\mathbf{F}) via

(1.4) ζnx:=inf{t+|ptx<1/n}andζx:=inf{t+|ptx=0}.\zeta^{x}_{n}:=\inf\{t\in\mathbb{R}_{+}\ |\ p^{x}_{t}<1/n\}\qquad\text{and}\qquad\zeta^{x}:=\inf\{t\in\mathbb{R}_{+}\ |\ p^{x}_{t}=0\}.

For all xEx\in E, it holds that (ζnx)n(\zeta^{x}_{n})_{n\in\mathbb{N}} is a nondecreasing sequence, [limnζnx=ζx]=1\mathbb{P}\left[\lim_{n\to\infty}\zeta^{x}_{n}=\zeta^{x}\right]=1, and px=0p^{x}=0 on [[ζx,[[\left[\negthinspace\left[\right.\right.\!\!\zeta^{x},\infty\!\left[\negthinspace\left[\right.\right.\! (see also [Jac85, Lemma 1.8]). Note also that, due to [Jac85, Corollary 1.11], it holds that [ζJ<]=0\mathbb{P}\left[\zeta^{J}<\infty\right]=0, with ζJ(ω):=ζJ(ω)(ω)\zeta^{J}(\omega)\,:=\,\zeta^{J(\omega)}(\omega) for all ωΩ\omega\in\Omega. For every xEx\in E, we consider the ζx\mathcal{F}_{\zeta^{x}}-measurable event Λx:={ζx<,pζxx>0}\Lambda^{x}:=\{\zeta^{x}<\infty,p^{x}_{\zeta^{x}-}>0\}. Define

(1.5) ηx:=ζΛxx=ζx𝕀Λx+𝕀ΩΛx,xE,\eta^{x}:=\zeta^{x}_{\Lambda^{x}}=\zeta^{x}\mathbb{I}_{\Lambda^{x}}+\infty\mathbb{I}_{\Omega\setminus\Lambda^{x}},\quad\forall x\in E,

which is a stopping time on (Ω,𝐅)(\Omega,\,\mathbf{F}) and represents the time at which pxp^{x} jumps to zero.

Under Assumption 1.9, we now discuss counterparts to Theorems 1.4 and 1.7 on the validity of NA1 in initially enlarged filtrations. Note that Assumption 1.9 guarantees that SS is a semimartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}), by [Jac85, Theorem 1.1], which is proved by relying on the Bichteler-Dellacherie characterisation of semimartingales. (In this respect, see also Remark 3.5 of the present paper.) This allows us to define the class 𝒳(𝐆,S)\mathcal{X}(\mathbf{G},S) and the condition NA(𝐆,S)1{}_{1}(\mathbf{G},S) as done in § 1.2 with respect to the filtration 𝐅\mathbf{F}. The first result is concerned with stability of condition NA1 for a fixed semimartingale model.

Theorem 1.11.

Under Assumption 1.9, suppose further that the space 𝕃1(Ω,,)\mathbb{L}^{1}(\Omega,\mathcal{F},\mathbb{P}) is separable and [ηx<,ΔSηx0]=0\mathbb{P}\left[\eta^{x}<\infty,\Delta S_{\eta^{x}}\neq 0\right]=0 holds for γ\gamma-a.e. xEx\in E. If NA(𝐅,S)1{}_{1}(\mathbf{F},S) holds, then NA1(𝐆,S)(\mathbf{G},S) holds.

Note that separability is a mild technical assumption that only allows us to use the results of [SY78, Proposition 4]; as the authors of the latter paper mention, it is satisfied in all cases of practical interest.

In § 1.5.3 we will provide an example showing how condition [ηx<,ΔSηx0]=0\mathbb{P}\left[\eta^{x}<\infty,\Delta S_{\eta^{x}}\neq 0\right]=0, for γ\gamma-a.e. xEx\in E, cannot be dropped.

As was the case for progressively enlarged filtrations, Theorem 1.11 has the following consequence: if [ηx<]=0\mathbb{P}\left[\eta^{x}<\infty\right]=0 for γ\gamma-a.e. xEx\in E, condition NA(𝐅,S)1{}_{1}(\mathbf{F},S) implies condition NA1(𝐆,S)(\mathbf{G},S)  for any asset-price process SS. In order to formulate the counterpart to statement (2) of Theorem 1.7 (regarding stability of the NA1  condition for all semimartingale models) in the case of initially enlarged filtrations, we have to slightly depart from our original setting. More precisely, the explicit example of an arbitrage of the first kind in the enlarged filtration when [ηx<]>0\mathbb{P}\left[\eta^{x}<\infty\right]>0 will involve a potentially infinite collection of semimartingales. (However, see Remark 1.13.) To wit, with DxD^{x} denoting the predictable compensator of 𝕀[[ηx,[[\mathbb{I}_{[\kern-1.22911pt[\eta^{x},\infty[\kern-1.22911pt[} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) for all xEx\in E, define the collection (Sx)xE(S^{x})_{x\in E} via

(1.6) Sx:=(Dx)1𝕀[[0,ηx[[,xE.S^{x}\,:=\,\mathcal{E}(-D^{x})^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta^{x}[\kern-1.22911pt[},\quad\forall x\in E.

In Section 3, under separability assumption on the space 𝕃1(Ω,,)\mathbb{L}^{1}(\Omega,\mathcal{F},\mathbb{P}), it is established that one can obtain a version of the function E×Ω×+(x,ω,t)Stx(ω)E\times\Omega\times\mathbb{R}_{+}\ni(x,\omega,t)\mapsto S^{x}_{t}(\omega) which is E𝒪(𝐅)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})-measurable. The process SJS^{J} defined via SJ(ω,t):=StJ(ω)(ω)S^{J}(\omega,t)\,:=\,S^{J(\omega)}_{t}(\omega) for all (ω,t)Ω×+(\omega,t)\in\Omega\times\mathbb{R}_{+} is a semimartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}), and has the following financial interpretation: an insider with knowledge of JJ and unit initial capital takes at time zero a position on a single unit of the stock with index JJ, and keeps it indefinitely. Although this strategy may involve an infinite number of assets, it is of the simplest possible buy-and-hold nature. Statement (1) of the following theorem is an immediate consequence of Theorem 1.11, while the proof of statement (2) is given in Section 3.3.

Theorem 1.12.

Under Assumption 1.9, the following statements hold true:

  1. (1)

    If [ηx<]=0\mathbb{P}\left[\eta^{x}<\infty\right]=0 holds for γ\gamma-a.e xEx\in E, then for any SS such that NA(𝐅,S)1{}_{1}(\mathbf{F},S) holds, NA1(𝐆,S)(\mathbf{G},S) also holds.

  2. (2)

    Suppose that the space 𝕃1(Ω,,)\mathbb{L}^{1}(\Omega,\mathcal{F},\mathbb{P}) is separable and that E[ηx<]γ[dx]>0\int_{E}\mathbb{P}\left[\eta^{x}<\infty\right]\gamma\left[\mathrm{d}x\right]>0. Then, the family (Sx)xE(S^{x})_{x\in E} in (1.6) consists of local martingales on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), and SJS^{J} is nondecreasing with [StJ=S0J,t+]<1\mathbb{P}\left[S^{J}_{t}=S^{J}_{0},\,\forall t\in\mathbb{R}_{+}\right]<1. In particular, NA1(𝐅,Sx)(\mathbf{F},S^{x}) holds, for every xEx\in E, but NA1(𝐆,SJ)(\mathbf{G},S^{J}) fails.

Loosely speaking, in part (2) of Theorem 1.12, the insider identifies from the beginning a single asset in the family (Sx)xE\left(S^{x}\right)_{x\in E} which will not default and can therefore arbitrage.

Remark 1.13.

If k[J=xk]=1\sum_{k\in\mathbb{N}}\mathbb{P}\left[J=x_{k}\right]=1 holds for a family {xk|k}E\left\{x_{k}\ |\ k\in\mathbb{N}\right\}\subseteq E, one can find a single asset that will lead to arbitrage of the first kind. Indeed, E[ηx<]γ[dx]>0\int_{E}\mathbb{P}\left[\eta^{x}<\infty\right]\gamma\left[\mathrm{d}x\right]>0 implies that there exists κ\kappa\in\mathbb{N} such that [ηxκ<]>0\mathbb{P}\left[\eta^{x_{\kappa}}<\infty\right]>0. Since [ζJ<]=0\mathbb{P}\left[\zeta^{J}<\infty\right]=0, [J=xκ,ηxκ<]=0\mathbb{P}\left[J=x_{\kappa},\,\eta^{x_{\kappa}}<\infty\right]=0 follows in a straightforward way; therefore, the buy-and-hold strategy 𝕀{J=xκ}\mathbb{I}_{\left\{J=x_{\kappa}\right\}} results in the arbitrage 𝕀{J=xκ}Sxκ\mathbb{I}_{\left\{J=x_{\kappa}\right\}}\cdot S^{x_{\kappa}}.

When the law γ\gamma has a diffuse component the previous argument may not work; however, one can still obtain an arbitrage of the first kind using a single asset under an assumption that is stronger (more precisely, at least not weaker) than E[ηx<]γ[dx]>0\int_{E}\mathbb{P}\left[\eta^{x}<\infty\right]\gamma\left[\mathrm{d}x\right]>0 as in part (2) of Theorem 1.12. To wit, for BEB\in\mathcal{B}_{E} with γ[B]>0\gamma\left[B\right]>0, define ηB\eta^{B} in the obvious way, as the time that the martingale (γt[B])t+\left(\gamma_{t}\left[B\right]\right)_{t\in\mathbb{R}_{+}} jumps to zero. Note the equality γt[B]=Bptxγ[dx]\gamma_{t}\left[B\right]=\int_{B}p^{x}_{t}\gamma\left[\mathrm{d}x\right], for all t+t\in\mathbb{R}_{+}; in particular, [ηB<]>0\mathbb{P}\left[\eta^{B}<\infty\right]>0 implies that E[ηx<]γ[dx]>0\int_{E}\mathbb{P}\left[\eta^{x}<\infty\right]\gamma\left[\mathrm{d}x\right]>0. (It is an open question whether the converse implication is also true for some set BEB\in\mathcal{B}_{E}.) Under the assumption [ηB<]>0\mathbb{P}\left[\eta^{B}<\infty\right]>0 for some BEB\in\mathcal{B}_{E} with γ[B]>0\gamma\left[B\right]>0, upon defining S:=(DB)1𝕀[[0,ηB[[S\,:=\,\mathcal{E}(-D^{B})^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta^{B}[\kern-1.22911pt[} where DBD^{B} denotes the predictable compensator of 𝕀[[ηB,[[\mathbb{I}_{[\kern-1.22911pt[\eta^{B},\infty[\kern-1.22911pt[} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), it can be shown that SS is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), and 𝕀{JB}S\mathbb{I}_{\left\{J\in B\right\}}\cdot S is nondecreasing with [St=S0,t+]<1\mathbb{P}\left[S_{t}=S_{0},\,\forall t\in\mathbb{R}_{+}\right]<1, that is, NA(𝐅,S)1{}_{1}(\mathbf{F},S) holds while NA1(𝐆,S)(\mathbf{G},S) fails.

Remark 1.14.

It is interesting to observe that the necessary and sufficient conditions given in Theorem 1.7 and in Theorem 1.12 for the preservation of the NA1 property under filtration enlargements bear resemblance to the necessary and sufficient condition obtained in [Fon14] for the preservation of the NA1 property under absolutely continuous (but not necessarily equivalent) changes of measure. This similarity is not a coincidence, given the deep link existing between filtration enlargements and non-equivalent changes of measure, as shown in [Yoe85].

The proof of Lemma 1.10 as well as of Theorems 1.11 and 1.12 is given in § 3. An example in the initial enlargement framework involving the Poisson process is given in § 1.5 below.

1.5. Examples

The first two examples are in the progressive filtration enlargement framework. In the first one, the stopping time η\eta is totally inaccessible and assertion (2) of Theorem 1.7 is illustrated by explicit computations; the second example contains a set-up where η\eta is accessible. The last example shows how condition [ηx<,ΔSηx0]=0\mathbb{P}\left[\eta^{x}<\infty,\Delta S_{\eta^{x}}\neq 0\right]=0, for γ\gamma-a.e. xEx\in E, cannot be dropped in Theorem 1.11.

1.5.1. An example under progressive filtration enlargement where η\eta is totally inaccessible

Let (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) be a complete probability space supporting an \mathcal{F}-measurable random variable ζ:Ω+\zeta:\Omega\mapsto\mathbb{R}_{+} such that [ζ>t]=exp(t)\mathbb{P}\left[\zeta>t\right]=\exp(-t) holds for all t+t\in\mathbb{R}_{+}. Set 𝐅=(t)t+\mathbf{F}=(\mathcal{F}_{t})_{t\in\mathbb{R}_{+}} to be the smallest filtration that satisfies the usual hypotheses and makes ζ\zeta a stopping time. Define τ:=ζ/2\tau\,:=\,\zeta/2, and consider the filtration 𝐆\mathbf{G} obtained as the progressive enlargement of 𝐅\mathbf{F} with respect to τ\tau. Let ZZ and AA be defined as in § 1.3.

Note that Zt=0Z_{t}=0 holds on {ζt}\left\{\zeta\leq t\right\}, while Zt=exp(t)Z_{t}=\exp(-t) holds on {t<ζ}\left\{t<\zeta\right\}, the last fact following from τ=ζ/2\tau=\zeta/2 and the memoryless property of the exponential law. Therefore, Zt=exp(t)𝕀{t<ζ}Z_{t}=\exp(-t)\mathbb{I}_{\left\{t<\zeta\right\}} is true for all t+t\in\mathbb{R}_{+}. Similarly, ΔAσ=[τ=σ|σ]=[ζ=2σ|σ]=0\Delta A_{\sigma}=\mathbb{P}\left[\tau=\sigma\ |\ \mathcal{F}_{\sigma}\right]=\mathbb{P}\left[\zeta=2\sigma\ |\ \mathcal{F}_{\sigma}\right]=0 is true for all bounded stopping times σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}), which implies that ΔA=0\Delta A=0. Note that ζ=inf{t+|Zt=0 or Zt=0}\zeta=\inf\left\{t\in\mathbb{R}_{+}\ |\ Z_{t-}=0\text{ or }Z_{t}=0\right\} and Zζ=exp(ζ)>0Z_{\zeta-}=\exp\left(-\zeta\right)>0. Since ΔA=0\Delta A=0, for η\eta defined as in (1.3), we obtain that η=ζ\eta=\zeta. The predictable compensator of 𝕀[[η,[[\mathbb{I}_{[\kern-1.22911pt[\eta,\infty[\kern-1.22911pt[} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) is equal to D:=(ηt)t+D\,:=\,(\eta\wedge t)_{t\in\mathbb{R}_{+}}; in particular, ζ=η\zeta=\eta is totally inaccessible on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}).

Here we have [η<]=1\mathbb{P}\left[\eta<\infty\right]=1, hence we can proceed to construct a local martingale SS as in Theorem 1.7-(2). To wit, S:=(D)1𝕀[[0,η[[=exp(D)𝕀[[0,η[[S\,:=\,\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[}=\exp(D)\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[}, that is, St=exp(t)𝕀{t<ζ}S_{t}=\exp(t)\mathbb{I}_{\left\{t<\zeta\right\}} for t+t\in\mathbb{R}_{+}. Note that SS is a quasi-left-continuous nonnegative martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), so that NA(𝐅,S)1{}_{1}(\mathbf{F},S) trivially holds. However, since SS is strictly increasing up to τ\tau, NA(𝐆,Sτ)1{}_{1}(\mathbf{G},S^{\tau}) fails.

1.5.2. An example under progressive filtration enlargement where η\eta is accessible

Let (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) be a complete probability space that supports an \mathcal{F}-measurable random variable ζ:Ω\zeta:\Omega\mapsto\mathbb{N} such that pk:=[ζ=k](0,1)p_{k}\,:=\,\mathbb{P}\left[\zeta=k\right]\in(0,1) holds for all kk\in\mathbb{N}, where k=1pk=1\sum_{k=1}^{\infty}p_{k}=1. Set 𝐅=(t)t+\mathbf{F}=(\mathcal{F}_{t})_{t\in\mathbb{R}_{+}} to be the smallest filtration that satisfies the usual hypotheses and makes ζ\zeta a stopping time. Since ζ\zeta is \mathbb{N}-valued, it is an accessible time on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). Define τ:=ζ1\tau\,:=\,\zeta-1, and consider the progressively enlarged filtration 𝐆\mathbf{G}. Let ZZ and AA be defined as in § 1.3.

Again, one may compute ZZ explicitly. In fact, Zt=0Z_{t}=0 holds on {ζt}\left\{\zeta\leq t\right\}; furthermore, upon defining qk=n=k+1pnq_{k}=\sum_{n=k+1}^{\infty}p_{n} for all k{0,1,}k\in\left\{0,1,\ldots\right\}, and denoting by \lceil\cdot\rceil the integer part, we have

Zt=[τ>t|t]=[ζ>t+1|t]=[ζ>t+1|t]=qt+1qt,on {t<ζ}.Z_{t}=\mathbb{P}\left[\tau>t\ |\ \mathcal{F}_{t}\right]=\mathbb{P}\left[\zeta>t+1\ |\ \mathcal{F}_{t}\right]=\mathbb{P}\left[\zeta>\lceil t+1\rceil\ |\ \mathcal{F}_{t}\right]=\frac{q_{\lceil t+1\rceil}}{q_{\lceil t\rceil}},\quad\text{on }\left\{t<\zeta\right\}.

Note that ζ=inf{t+|Zt=0 or Zt=0}\zeta=\inf\left\{t\in\mathbb{R}_{+}\ |\ Z_{t-}=0\text{ or }Z_{t}=0\right\} and Zζ=qζ/qζ1>0Z_{\zeta-}=q_{\lceil\zeta\rceil}/q_{\lceil\zeta-1\rceil}>0. Furthermore, ΔAζ=[τ=ζ|ζ]=0\Delta A_{\zeta}=\mathbb{P}\left[\tau=\zeta\ |\ \mathcal{F}_{\zeta}\right]=0 holds true. It follows that, for η\eta defined as in (1.3), η=ζ\eta=\zeta; in particular, η\eta is accessible on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}).

1.5.3. An example under initial filtration enlargement

Let us consider a probability space (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) supporting a Poisson process NN with intensity λ>0\lambda>0 stopped at time T(0,)T\in(0,\infty). Let 𝐅\mathbf{F} be the right-continuous filtration generated by NN and consider the random variable J:=NTJ\,:=\,N_{T}. As in [GVV06, § 4.2] (compare also with [GP01, § 4.3]), it can be checked that

ptx=eλt(λ(Tt))xNt(λT)xx!(xNt)!𝕀{Ntx},for all t[0,T),p^{x}_{t}=\mathrm{e}^{-\lambda t}\frac{\bigl{(}\lambda(T-t)\bigr{)}^{x-N_{t}}}{(\lambda T)^{x}}\frac{x!}{(x-N_{t})!}\mathbb{I}_{\{N_{t}\leq x\}},\quad\textrm{for all $t\in[0,T)$},

and pTx=eλTx!/(λT)x𝕀{NT=x}p^{x}_{T}=\mathrm{e}^{-\lambda T}x!/(\lambda T)^{x}\mathbb{I}_{\{N_{T}=x\}}, so that Jacod’s criterion (Assumption 1.9) is satisfied.

Consider then the process SS defined by St:=exp(Ntλt(e1))S_{t}\,:=\,\exp\bigl{(}N_{t}-\lambda t(\mathrm{e}-1)\bigr{)}, for all t[0,T]t\in[0,T]. The process SS is a strictly positive 𝐅\mathbf{F}-martingale (see e.g. [JYC09, Proposition 8.2.2.1]), so that NA1(𝐅,S)(\mathbf{F},S) holds. However, NA1(𝐆,S)(\mathbf{G},S) does not hold. To see this, define the 𝐆\mathbf{G}-stopping time σ:=inf{t[0,T]|Nt=NT}\sigma\,:=\,\inf\left\{t\in[0,T]\ |\ N_{t}=N_{T}\right\} and consider the strategy 𝕀]]σ,T]]-\mathbb{I}_{]\kern-1.22911pt]\sigma,T]\kern-1.22911pt]}. Then, for all t[0,T]t\in[0,T], we get

(𝕀]]σ,T]]S)t=𝕀{t>σ}exp(Nσλσ(e1))(1exp(λ(tσ)(e1))).(-\mathbb{I}_{]\kern-1.22911pt]\sigma,T]\kern-1.22911pt]}\cdot S)_{t}=\mathbb{I}_{\{t>\sigma\}}\exp\bigl{(}N_{\sigma}-\lambda\sigma(\mathrm{e}-1)\bigr{)}\Bigl{(}1-\exp\bigl{(}-\lambda(t-\sigma)(\mathrm{e}-1)\bigr{)}\Bigr{)}.

In particular, the process 𝕀]]σ,T]]S-\mathbb{I}_{]\kern-1.22911pt]\sigma,T]\kern-1.22911pt]}\cdot S is nondecreasing and [σ<T]=1\mathbb{P}\left[\sigma<T\right]=1, thus implying that NA1(𝐆,S)(\mathbf{G},S) fails to hold. Indeed, in the context of the present example, the processes pxp^{x} have a positive probability to jump to zero and this event occurs exactly in correspondence of the jump times of the Poisson process NN, thus showing that the condition [ηx<,ΔSηx0]=0\mathbb{P}\left[\eta^{x}<\infty,\Delta S_{\eta^{x}}\neq 0\right]=0 for γ\gamma-a.e. xEx\in E fail to hold.

2. Arbitrage of the First Kind in Progressively Enlarged Filtrations

In this section, the proof of Theorem 1.4 and Theorem 1.7 will be given. In the process, we will also obtain certain interesting results concerning the behaviour (up to the random time τ\tau) of nonnegative super/local martingales on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) in the enlarged filtration 𝐆\mathbf{G} (see Section 2.3). In particular, these results do not follow from classical results of enlargement of filtrations theory.

2.1. Representation pair associated with τ\tau

The next result is [Kar15, Theorem 1.1].

Theorem 2.1.

For any random time τ\tau on (Ω,𝐅)(\Omega,\,\mathbf{F}) satisfying [τ=]=0\mathbb{P}\left[\tau=\infty\right]=0 there exists a pair of processes (K,L)(K,L) with the following properties:

  1. (1)

    KK is 𝐅\mathbf{F}-adapted, right-continuous, nondecreasing, with 0K10\leq K\leq 1.

  2. (2)

    LL is a nonnegative local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), with L0=1L_{0}=1.

  3. (3)

    For any nonnegative optional processes VV on (Ω,𝐅)(\Omega,\,\mathbf{F}), we have

    (2.1) 𝔼[Vτ]=𝔼[+VtLtdKt].\mathbb{E}[V_{\tau}]=\mathbb{E}\left[\int_{{\mathbb{R}_{+}}}V_{t}\,L_{t}\mathrm{d}K_{t}\right].
  4. (4)

    +𝕀{Kt=1}dLt=0\int_{{\mathbb{R}_{+}}}\mathbb{I}_{\left\{K_{t-}=1\right\}}\mathrm{d}L_{t}=0 and +𝕀{Lt=0}dKt=0\int_{{\mathbb{R}_{+}}}\mathbb{I}_{\left\{L_{t}=0\right\}}\mathrm{d}K_{t}=0 hold \mathbb{P}-a.s.

It also comes as part of the results in [Kar15, §1.1] that Z=L(1K)Z=L(1-K), which gives a particular multiplicative optional decomposition of ZZ. In general, there are many possible optional multiplicative decompositions; the properties described in Theorem 2.1 specify the pair (K,L)(K,L) in a unique way. Note also that, in the special case where [τ=σ]=0\mathbb{P}\left[\tau=\sigma\right]=0 for every stopping time σ\sigma on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), the decomposition Z=L(1K)Z=L(1-K) coincides with the multiplicative Doob-Meyer decomposition of the supermartingale ZZ (see [Kar15, Remark 1.6]).

Remark 2.2.

Let σ\sigma be a stopping time on (Ω,𝐅)(\Omega,\,\mathbf{F}). For any BσB\in\mathcal{F}_{\sigma}, (2.1) applied to the process V=𝕀B𝕀]]σ,[[V=\mathbb{I}_{B}\mathbb{I}_{]\kern-1.22911pt]\sigma,\infty[\kern-1.22911pt[}, combined with Z=L(1K)Z=L(1-K) and the definition of ZZ, implies that

𝔼[Lσ(1Kσ)𝕀B]=𝔼[Zσ𝕀B]=𝔼[Vτ]=𝔼[𝕀B(σ,)LtdKt].\mathbb{E}\left[L_{\sigma}(1-K_{\sigma})\mathbb{I}_{B}\right]=\mathbb{E}\left[Z_{\sigma}\mathbb{I}_{B}\right]=\mathbb{E}\left[V_{\tau}\right]=\mathbb{E}\left[\mathbb{I}_{B}\int_{(\sigma,\infty)}L_{t}\mathrm{d}K_{t}\right].

Since the above equality holds for all BσB\in\mathcal{F}_{\sigma}, it follows that

(2.2) Lσ(1Kσ)=𝔼[(σ,)LtdKt|σ].L_{\sigma}(1-K_{\sigma})=\mathbb{E}\left[\int_{(\sigma,\infty)}L_{t}\mathrm{d}K_{t}\ \Big{|}\ \mathcal{F}_{\sigma}\right].
Remark 2.3.

Another use of (2.1) gives

[Lτ=0]=𝔼[+𝕀{Lt=0}LtdKt]=0.\mathbb{P}\left[L_{\tau}=0\right]=\mathbb{E}\left[\int_{{\mathbb{R}_{+}}}\mathbb{I}_{\left\{L_{t}=0\right\}}L_{t}\mathrm{d}K_{t}\right]=0.

Since LL is a nonnegative local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), it follows that [[0,τ]]{L>0}[\kern-1.49994pt[0,\tau]\kern-1.49994pt]\subseteq\left\{L>0\right\}.

Lemma 2.4.

For ζ\zeta defined in (1.2), and AA denoting the dual optional projection of 𝕀[[τ,[[\mathbb{I}_{[\kern-1.22911pt[\tau,\infty[\kern-1.22911pt[}, the following set equality holds:

(2.3) {ζ<,Zζ>0,ΔAζ=0}={ζ<,Kζ<1,Lζ>0,ΔKζ=0}.\left\{\zeta<\infty,\,Z_{\zeta-}>0,\,\Delta A_{\zeta}=0\right\}=\left\{\zeta<\infty,\ K_{\zeta-}<1,\,L_{\zeta-}>0,\,\Delta K_{\zeta}=0\right\}.

Furthermore, Lζ=0L_{\zeta}=0 holds on the above event.

Proof.

Since Z=L(1K)Z=L(1-K), {ζ<,Zζ>0}={ζ<,Kζ<1,Lζ>0}\left\{\zeta<\infty,\,Z_{\zeta-}>0\right\}=\left\{\zeta<\infty,\ K_{\zeta-}<1,\,L_{\zeta-}>0\right\} is immediate. According to the definition of KK in [Kar15, equation (1.1)], it follows that, on {ζ<}\left\{\zeta<\infty\right\}, ΔAζ=0\Delta A_{\zeta}=0 implies ΔKζ=0\Delta K_{\zeta}=0. Furthermore, on {ζ<,Zζ>0}\left\{\zeta<\infty,\,Z_{\zeta-}>0\right\}, ΔKζ=0\Delta K_{\zeta}=0 implies that Kζ=Kζ<1K_{\zeta}=K_{\zeta-}<1, which gives that ΔAζ=0\Delta A_{\zeta}=0 upon using [Kar15, equation (1.1)] again. The set-equality (2.3) has been established. Finally, note that the fact that 0=Zζ=Lζ(1Kζ)0=Z_{\zeta}=L_{\zeta}(1-K_{\zeta}) implies that Lζ=0L_{\zeta}=0 has to hold on {ζ<,Kζ<1,Lζ>0,ΔKζ=0}\left\{\zeta<\infty,\ K_{\zeta-}<1,\,L_{\zeta-}>0,\,\Delta K_{\zeta}=0\right\}. ∎

2.2. Results regarding the stopping time η\eta

Recall that η=ζ𝕀Λ+𝕀ΩΛ\eta=\zeta\mathbb{I}_{\Lambda}+\infty\mathbb{I}_{\Omega\setminus\Lambda}, where Λ:={ζ<,Zζ>0,ΔAζ=0}\Lambda\,:=\,\left\{\zeta<\infty,\,Z_{\zeta-}>0,\,\Delta A_{\zeta}=0\right\}. In view of (2.3), Λ={ζ<,Kζ<1,Lζ>0,ΔKζ=0}\Lambda=\left\{\zeta<\infty,\ K_{\zeta-}<1,\,L_{\zeta-}>0,\,\Delta K_{\zeta}=0\right\}. In the proof of the next result, it is established inter alia that η\eta is not predictable, when finite.

Lemma 2.5.

Let DD be the predictable compensator of 𝕀[[η,[[\mathbb{I}_{[\kern-1.22911pt[\eta,\infty[\kern-1.22911pt[} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). Then:

  1. (1)

    ΔD<1\Delta D<1, \mathbb{P}-a.s.; in particular, (D)\mathcal{E}(-D) is nonincreasing and strictly positive;

  2. (2)

    the nonnegative process (D)1𝕀[[0,η[[\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[} is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}).

Proof.

For any predictable time σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}), it holds that ΔDσ=[η=σ|σ]\Delta D_{\sigma}=\mathbb{P}\left[\eta=\sigma\ |\ \mathcal{F}_{\sigma-}\right] on {σ<}\left\{\sigma<\infty\right\} (see e.g. [HWY92, Theorem 5.27]). In the next paragraph, we shall show that ΔDσ<1\Delta D_{\sigma}<1 holds on {σ<}\left\{\sigma<\infty\right\} for any predictable time σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}). Then, the predictable section theorem implies that ΔD<1\Delta D<1 \mathbb{P}-a.s.; in particular, the process (D)1𝕀[[0,η[[\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[} will be well-defined. This will establish part (1).

We proceed in showing that [η=σ<|σ]<1\mathbb{P}\left[\eta=\sigma<\infty\ |\ \mathcal{F}_{\sigma-}\right]<1 holds for any fixed predictable time σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}). Suppose that Σ:={[η=σ<|σ]=1}σ\Sigma\,:=\,\left\{\mathbb{P}\left[\eta=\sigma<\infty\ |\ \mathcal{F}_{\sigma-}\right]=1\right\}\in\mathcal{F}_{\sigma-} is such that [Σ]>0\mathbb{P}\left[\Sigma\right]>0. Upon replacing σ\sigma by the predictable time σΣ:=σ𝕀Σ+𝕀ΩΣ\sigma_{\Sigma}\,:=\,\sigma\mathbb{I}_{\Sigma}+\infty\mathbb{I}_{\Omega\setminus\Sigma}, we infer the existence of a predictable time σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}) such that [σ<]>0\mathbb{P}\left[\sigma<\infty\right]>0 and {σ<}={[η=σ<|σ]=1}\left\{\sigma<\infty\right\}=\left\{\mathbb{P}\left[\eta=\sigma<\infty\ |\ \mathcal{F}_{\sigma-}\right]=1\right\} hold. From the previous set-equality it follows that [η=σ<|σ]=𝕀{η=σ<}\mathbb{P}\left[\eta=\sigma<\infty\ |\ \mathcal{F}_{\sigma-}\right]=\mathbb{I}_{\{\eta=\sigma<\infty\}}, which in particular implies that {η=σ<}σ\left\{\eta=\sigma<\infty\right\}\in\mathcal{F}_{\sigma-}. Therefore, since 𝔼[Δ(A+Z)σ|σ]=0\mathbb{E}\left[\Delta(A+Z)_{\sigma}\ |\ \mathcal{F}_{\sigma-}\right]=0 holds on {σ<}\left\{\sigma<\infty\right\} (because A+ZA+Z is a martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) and σ\sigma is predictable on (Ω,𝐅)(\Omega,\,\mathbf{F})),

𝔼[ΔAσ|σ]=𝔼[ΔZσ|σ]=𝔼[ΔZη|σ]=𝔼[Zη|σ],on {η=σ<},\mathbb{E}\left[\Delta A_{\sigma}\ |\ \mathcal{F}_{\sigma-}\right]=-\mathbb{E}\left[\Delta Z_{\sigma}\ |\ \mathcal{F}_{\sigma-}\right]=-\mathbb{E}\left[\Delta Z_{\eta}\ |\ \mathcal{F}_{\sigma-}\right]=\mathbb{E}\left[Z_{\eta-}\ |\ \mathcal{F}_{\sigma-}\right],\quad\text{on }\left\{\eta=\sigma<\infty\right\},

where in the last equality we have used the definition of η\eta. On the other hand, using again the definition of η\eta, we obtain that 𝔼[ΔAσ|σ]=𝔼[ΔAη|σ]=0\mathbb{E}\left[\Delta A_{\sigma}\ |\ \mathcal{F}_{\sigma-}\right]=\mathbb{E}\left[\Delta A_{\eta}\ |\ \mathcal{F}_{\sigma-}\right]=0 holds on {η=σ<}\left\{\eta=\sigma<\infty\right\}. It follows that 𝔼[Zη|σ]=0\mathbb{E}\left[Z_{\eta-}\ |\ \mathcal{F}_{\sigma-}\right]=0 on {η=σ<}\left\{\eta=\sigma<\infty\right\}. Since Zη>0Z_{\eta-}>0 holds on {η<}\left\{\eta<\infty\right\}, the equality 𝔼[Zη𝕀{η=σ<}|σ]=0\mathbb{E}\left[Z_{\eta-}\mathbb{I}_{\left\{\eta=\sigma<\infty\right\}}\ |\ \mathcal{F}_{\sigma-}\right]=0 implies that [η=σ<]=0\mathbb{P}\left[\eta=\sigma<\infty\right]=0, which contradicts the fact that [σ<]>0\mathbb{P}\left[\sigma<\infty\right]>0 and {σ<}={[η=σ<|σ]=1}\left\{\sigma<\infty\right\}=\left\{\mathbb{P}\left[\eta=\sigma<\infty\ |\ \mathcal{F}_{\sigma-}\right]=1\right\} hold. Therefore, [η=σ<|σ]<1\mathbb{P}\left[\eta=\sigma<\infty\ |\ \mathcal{F}_{\sigma-}\right]<1 holds for any predictable time σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}).

We continue in establishing part (2). Let I=𝕀[[η,[[I=\mathbb{I}_{[\kern-1.22911pt[\eta,\infty[\kern-1.22911pt[}, so that IDI-D is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). Integration-by-parts gives

(D)1𝕀[[0,η[[=10(D)t1dIt+0(1It)d(D)t1=0(D)t1dIt+(D)1,\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[}=1-\int_{0}^{\cdot}\mathcal{E}(-D)_{t}^{-1}\mathrm{d}I_{t}+\int_{0}^{\cdot}\left(1-I_{t-}\right)\mathrm{d}\mathcal{E}(-D)_{t}^{-1}=-\int_{0}^{\cdot}\mathcal{E}(-D)_{t}^{-1}\mathrm{d}I_{t}+\mathcal{E}(-D)^{-1},

where the second equality follows from the facts that 1I=𝕀[[0,η]]1-I_{-}=\mathbb{I}_{[\kern-1.22911pt[0,\eta]\kern-1.22911pt]} and (D)1\mathcal{E}(-D)^{-1} is constant on [[η,[[[\kern-1.49994pt[\eta,\infty[\kern-1.49994pt[. Using Itô’s formula (actually, integration theory for finite-variation processes is sufficient), it is straightforward to check that

(D)1=1+0(D)t1dDt.\mathcal{E}(-D)^{-1}=1+\int_{0}^{\cdot}\mathcal{E}(-D)^{-1}_{t}\mathrm{d}D_{t}.

It then follows that

(D)1𝕀[[0,η[[=10(D)t1d(ID)t,\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[}=1-\int_{0}^{\cdot}\mathcal{E}(-D)_{t}^{-1}\mathrm{d}\left(I-D\right)_{t},

which concludes the argument in view of the fact that IDI-D is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). ∎

We write \mathbb{Q}\sim\mathbb{P} whenever \mathbb{Q} is a probability that is equivalent to \mathbb{P} on \mathcal{F}. Note that all the quantities that we have defined and depend on τ\tau (in particular, η\eta) depend on the underlying probability measure. For establishing Theorem 1.4, it is important that η\eta remains invariant under equivalent changes of probability. The next result ensures that this is indeed the case.

Lemma 2.6.

Let \mathbb{Q}\sim\mathbb{P}, and let η\eta^{\mathbb{Q}} be the stopping time on (Ω,𝐅)(\Omega,\,\mathbf{F}) defined under \mathbb{Q} in analogy to ηη\eta\equiv\eta^{\mathbb{P}} defined in (1.3) under \mathbb{P}. Then η=η\eta^{\mathbb{Q}}=\eta holds almost surely (under both \mathbb{P} and \mathbb{Q}).

Proof.

Denote by ZZ^{\mathbb{Q}} the Azéma supermartingale associated with τ\tau on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{Q}). We claim that {Z>0}={Z>0}\{Z^{\mathbb{Q}}>0\}=\{Z>0\} holds modulo evanescence. Indeed, this follows from the optional section theorem, upon noting that

{Zσ=0}={[τ>σ|σ]=0}={[τ>σ|σ]=0}={Zσ=0}\left\{Z_{\sigma}=0\right\}=\left\{\mathbb{P}\left[\tau>\sigma\ |\ \mathcal{F}_{\sigma}\right]=0\right\}=\left\{\mathbb{Q}\left[\tau>\sigma\ |\ \mathcal{F}_{\sigma}\right]=0\right\}=\big{\{}Z^{\mathbb{Q}}_{\sigma}=0\big{\}}

holds for all bounded stopping times σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}), where the second set-equality holds because \mathbb{Q}\sim\mathbb{P}. In particular, Zη=0Z_{\eta}=0 and Zη>0Z_{\eta-}>0 imply Zη=0Z^{\mathbb{Q}}_{\eta}=0 and Zη>0Z^{\mathbb{Q}}_{\eta-}>0. Now denote by AA^{\mathbb{Q}} the dual optional projection of 𝕀[[τ,[[\mathbb{I}_{[\kern-1.22911pt[\tau,\infty[\kern-1.22911pt[} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{Q}). Since \mathbb{Q}\sim\mathbb{P} and [τ=η]=0\mathbb{P}\left[\tau=\eta\right]=0, it follows that ΔAη=[τ=η|η]=0\Delta A^{\mathbb{Q}}_{\eta}=\mathbb{Q}\left[\tau=\eta\ |\ \mathcal{F}_{\eta}\right]=0. Together with the previous observation, this implies ηη\eta^{\mathbb{Q}}\leq\eta. Upon interchanging the roles of \mathbb{P} and \mathbb{Q}, one obtains the reverse inequality, completing the proof. ∎

2.3. Super/local martingales in the progressively enlarged filtration

The next result, which will be key in the development, is also of independent interest.

Proposition 2.7.

The following statements hold true:

  1. (1)

    Let XX be a nonnegative supermartingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). Then, the process Xτ/LτX^{\tau}/L^{\tau} is a supermartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

  2. (2)

    Let XX be a nonnegative local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) such that [[η,[[{X=0}[\kern-1.49994pt[\eta,\infty[\kern-1.49994pt[\,\subseteq\left\{X=0\right\} holds (modulo evanescence). Then, the process Xτ/LτX^{\tau}/L^{\tau} is a local martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

Proof.

Note first that, by Remark 2.3, 1/Lτ1/L^{\tau} is well defined. If XX is a nonnegative supermartingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), the Doob-Meyer decomposition gives that X=NBX=N-B, where NN is a (non-negative) local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) and BB is an increasing predictable process on (Ω,𝐅)(\Omega,\,\mathbf{F}) with B0=0B_{0}=0. Let s<ts<t and G𝒢sG\in\mathcal{G}_{s}. By (1.1) there exists a set GssG_{s}\in\mathcal{F}_{s} such that G{τ>s}=Gs{τ>s}G\cap\left\{\tau>s\right\}=G_{s}\cap\left\{\tau>s\right\}. Define then the nonnegative optional process Y:=𝕀Gs𝕀]]s,[[(Xt/Lt)𝕀{Lt>0}Y\,:=\,\mathbb{I}_{G_{s}}\mathbb{I}_{]\kern-1.22911pt]s,\infty[\kern-1.22911pt[}(X^{t}/L^{t})\mathbb{I}_{\left\{L^{t}>0\right\}} on (Ω,𝐅)(\Omega,\,\mathbf{F}), so that 𝕀G{τ>s}Xtτ/Ltτ=Yτ\mathbb{I}_{G\cap\left\{\tau>s\right\}}X^{\tau}_{t}/L^{\tau}_{t}=Y_{\tau}. In view of Theorem 2.1, it follows that

(2.4) 𝔼[Yτ]\displaystyle\mathbb{E}\left[Y_{\tau}\right] =𝔼[[0,)YuLudKu]\displaystyle=\mathbb{E}\left[\int_{[0,\infty)}Y_{u}L_{u}\mathrm{d}K_{u}\right]
=𝔼[𝕀Gs(s,t]XuLu𝕀{Lu>0}LudKu+XtLt𝕀Gs{Lt>0}(t,)LudKu]\displaystyle=\mathbb{E}\left[\mathbb{I}_{G_{s}}\int_{(s,t]}\frac{X_{u}}{L_{u}}\mathbb{I}_{\{L_{u}>0\}}L_{u}\mathrm{d}K_{u}+\frac{X_{t}}{L_{t}}\mathbb{I}_{G_{s}\cap\left\{L_{t}>0\right\}}\int_{(t,\infty)}L_{u}\mathrm{d}K_{u}\right]
=𝔼[𝕀Gs(s,t]Xu𝕀{Lu>0}dKu+XtLt𝕀Gs{Lt>0}Lt(1Kt)]\displaystyle=\mathbb{E}\left[\mathbb{I}_{G_{s}}\int_{(s,t]}X_{u}\mathbb{I}_{\{L_{u}>0\}}\mathrm{d}K_{u}+\frac{X_{t}}{L_{t}}\mathbb{I}_{G_{s}\cap\left\{L_{t}>0\right\}}L_{t}(1-K_{t})\right]
=𝔼[𝕀Gs((s,t]XudKu+Xt𝕀{Lt>0}(1Kt))],\displaystyle=\mathbb{E}\left[\mathbb{I}_{G_{s}}\left(\int_{(s,t]}X_{u}\mathrm{d}K_{u}+X_{t}\mathbb{I}_{\left\{L_{t}>0\right\}}(1-K_{t})\right)\right],

where (2.2) was used in the third equality above. Noting that {Lt>0}{Ls>0}\left\{L_{t}>0\right\}\subseteq\left\{L_{s}>0\right\}, integration-by-parts then implies that

(2.5) 𝔼[Yτ]\displaystyle\mathbb{E}\left[Y_{\tau}\right] 𝔼[𝕀Gs{Ls>0}((s,t]XudKu+Xt(1Kt))]\displaystyle\leq\mathbb{E}\left[\mathbb{I}_{G_{s}\cap\left\{L_{s}>0\right\}}\left(\int_{(s,t]}X_{u}\mathrm{d}K_{u}+X_{t}(1-K_{t})\right)\right]
=𝔼[𝕀Gs{Ls>0}(Xs(1Ks)+(s,t](1Ku)dXu)].\displaystyle=\mathbb{E}\left[\mathbb{I}_{G_{s}\cap\left\{L_{s}>0\right\}}\left(X_{s}(1-K_{s})+\int_{(s,t]}(1-K_{u-})\mathrm{d}X_{u}\right)\right].

Furthermore, since 0K10\leq K\leq 1 and the process BB is increasing, it holds that

(2.6) (s,t](1Ku)dXu=(s,t](1Ku)dNu(s,t](1Ku)dBu(s,t](1Ku)dNu\int_{(s,t]}(1-K_{u-})\mathrm{d}X_{u}=\int_{(s,t]}(1-K_{u-})\mathrm{d}N_{u}-\int_{(s,t]}(1-K_{u-})\mathrm{d}B_{u}\leq\int_{(s,t]}(1-K_{u-})\mathrm{d}N_{u}

and [(1K)N,(1K)N][N,N]\bigl{[}(1-K_{-})\cdot N,\,(1-K_{-})\cdot N\bigr{]}\leq[N,N]. Suppose first that N1N\in\mathcal{H}^{1}, i.e., 𝔼[[N,N]1/2]<\mathbb{E}\bigl{[}[N,N]_{\infty}^{1/2}\bigr{]}<\infty, from which it follows that

(2.7) 𝔼[[(1K)N,(1K)N]1/2]𝔼[[N,N]1/2]<.\mathbb{E}\left[\bigl{[}(1-K_{-})\cdot N,\,(1-K_{-})\cdot N\bigr{]}^{1/2}_{\infty}\right]\leq\mathbb{E}\left[[N,N]_{\infty}^{1/2}\right]<\infty.

Together with (2.6), this implies that 𝔼[𝕀Gs{Ls>0}(s,t](1Ku)dXu]0\mathbb{E}\left[\mathbb{I}_{G_{s}\cap\left\{L_{s}>0\right\}}\int_{(s,t]}(1-K_{u-})\mathrm{d}X_{u}\right]\leq 0. Hence, due to (2.5),

𝔼[𝕀Gs{τ>s}XtτLtτ]\displaystyle\mathbb{E}\left[\mathbb{I}_{G_{s}\cap\left\{\tau>s\right\}}\frac{X^{\tau}_{t}}{L^{\tau}_{t}}\right] =𝔼[Yτ]𝔼[𝕀Gs{Ls>0}Xs(1Ks)]=𝔼[𝕀Gs{Ls>0}XsLsLs(1Ks)]\displaystyle=\mathbb{E}\left[Y_{\tau}\right]\leq\mathbb{E}\left[\mathbb{I}_{G_{s}\cap\left\{L_{s}>0\right\}}X_{s}(1-K_{s})\right]=\mathbb{E}\left[\mathbb{I}_{G_{s}\cap\left\{L_{s}>0\right\}}\frac{X_{s}}{L_{s}}L_{s}(1-K_{s})\right]
=𝔼[𝕀Gs{Ls>0}XsLs𝕀{τ>s}]=𝔼[𝕀Gs{τ>s}XsLs],\displaystyle=\mathbb{E}\left[\mathbb{I}_{G_{s}\cap\left\{L_{s}>0\right\}}\frac{X_{s}}{L_{s}}\mathbb{I}_{\left\{\tau>s\right\}}\right]=\mathbb{E}\left[\mathbb{I}_{G_{s}\cap\left\{\tau>s\right\}}\frac{X_{s}}{L_{s}}\right],

where Ls(1Ks)=Zs=[τ>s|s]L_{s}(1-K_{s})=Z_{s}=\mathbb{P}[\tau>s\ |\ \mathcal{F}_{s}] and [[0,τ]]{L>0}[\kern-1.49994pt[0,\tau]\kern-1.49994pt]\subseteq\left\{L>0\right\} were used in the last line. Since

𝔼[𝕀GXtτLtτ]\displaystyle\mathbb{E}\left[\mathbb{I}_{G}\frac{X^{\tau}_{t}}{L^{\tau}_{t}}\right] =𝔼[𝕀Gs{τ>s}XtτLtτ]+𝔼[𝕀G{τs}XsτLsτ]\displaystyle=\mathbb{E}\left[\mathbb{I}_{G_{s}\cap\left\{\tau>s\right\}}\frac{X^{\tau}_{t}}{L^{\tau}_{t}}\right]+\mathbb{E}\left[\mathbb{I}_{G\cap\left\{\tau\leq s\right\}}\frac{X^{\tau}_{s}}{L^{\tau}_{s}}\right]
𝔼[𝕀Gs{τ>s}XsτLsτ]+𝔼[𝕀G{τs}XsτLsτ]=𝔼[𝕀GXsτLsτ],\displaystyle\leq\mathbb{E}\left[\mathbb{I}_{G_{s}\cap\left\{\tau>s\right\}}\frac{X^{\tau}_{s}}{L^{\tau}_{s}}\right]+\mathbb{E}\left[\mathbb{I}_{G\cap\left\{\tau\leq s\right\}}\frac{X^{\tau}_{s}}{L^{\tau}_{s}}\right]=\mathbb{E}\left[\mathbb{I}_{G}\frac{X^{\tau}_{s}}{L^{\tau}_{s}}\right],

we have thus proved that Xτ/LτX^{\tau}/L^{\tau} is a supermartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}). The general case follows by localization. In fact, by [Pro04, Theorem IV.51], every local martingale NN on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) admits a nondecreasing sequence (σn)n(\sigma_{n})_{n\in\mathbb{N}} of stopping times (under 𝐅\mathbf{F} and, a fortiori, under 𝐆\mathbf{G}) \mathbb{P}-a.s. converging to infinity such that Nσn1N^{\sigma_{n}}\in\mathcal{H}^{1} for all nn\in\mathbb{N}. The preceding arguments imply that Xσnτ/LτX^{\sigma_{n}\wedge\tau}/L^{\tau} is a supermartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}) for all nn\in\mathbb{N} and statement (1) of the proposition then follows by Fatou’s lemma.

In order to prove part (2), define the nondecreasing sequence (σn)n(\sigma_{n})_{n\in\mathbb{N}} of stopping times via σn:=inf{t+|[X,X]t>n}ζn\sigma_{n}\,:=\,\inf\left\{t\in\mathbb{R}_{+}\ |\ [X,X]_{t}>n\right\}\wedge\zeta_{n}, for all nn\in\mathbb{N}. For future reference, note that σnζη\sigma_{n}\leq\zeta\leq\eta holds for all nn\in\mathbb{N}. It is straightforward to check that limn[ζn<τ]=0\lim_{n\to\infty}\mathbb{P}\left[\zeta_{n}<\tau\right]=0; therefore, in order to prove the result, it suffices to show that Xτσn/LτσnX^{\tau\wedge\sigma_{n}}/L^{\tau\wedge\sigma_{n}} is a martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}) for all nn\in\mathbb{N}. By part (1), the process Xτσn/LτσnX^{\tau\wedge\sigma_{n}}/L^{\tau\wedge\sigma_{n}} is a supermartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}) for all nn\in\mathbb{N}. It follows that it suffices to show that 𝔼[Xτσn/Lτσn]=𝔼[X0]\mathbb{E}\left[X_{\tau\wedge\sigma_{n}}/L_{\tau\wedge\sigma_{n}}\right]=\mathbb{E}\left[X_{0}\right] holds for all nn\in\mathbb{N}. Similarly as in the first part of the proof, set Y:=(X/L)𝕀{L>0}Y\,:=\,(X/L)\mathbb{I}_{\left\{L>0\right\}}, and note that YσnY^{\sigma_{n}} is optional on (Ω,𝐅)(\Omega,\,\mathbf{F}) and Xτσn/Lτσn=YτσnX_{\tau\wedge\sigma_{n}}/L_{\tau\wedge\sigma_{n}}=Y^{\sigma_{n}}_{\tau} holds for all nn\in\mathbb{N}. Computations analogous to (2.4) allow then to show that

(2.8) 𝔼[XτσnLτσn]=𝔼[Yτσn]=𝔼[[0,σn]XtdKt+Xσn𝕀{Lσn>0}(1Kσn)].\mathbb{E}\left[\frac{X_{\tau\wedge\sigma_{n}}}{L_{\tau\wedge\sigma_{n}}}\right]=\mathbb{E}\left[Y^{\sigma_{n}}_{\tau}\right]=\mathbb{E}\left[\int_{[0,\sigma_{n}]}X_{t}\mathrm{d}K_{t}+X_{\sigma_{n}}\mathbb{I}_{\left\{L_{\sigma_{n}}>0\right\}}(1-K_{\sigma_{n}})\right].

Note that Xσn𝕀{Lσn=0}(1Kσn)=0X_{\sigma_{n}}\mathbb{I}_{\left\{L_{\sigma_{n}}=0\right\}}(1-K_{\sigma_{n}})=0 holds for all nn\in\mathbb{N}; indeed, this follows from Lemma 2.4 since {Lσn=0,Kσn<1}={σn=η}\left\{L_{\sigma_{n}}=0,\,K_{\sigma_{n}}<1\right\}=\left\{\sigma_{n}=\eta\right\} holds for all nn\in\mathbb{N}. Therefore, similarly as in (2.5), integration-by-parts yields that

𝔼[XτσnLτσn]=𝔼[[0,σn]XtdKt+Xσn(1Kσn)]=𝔼[[0,σn](1Kt)dXt]=𝔼[X0],\mathbb{E}\left[\frac{X_{\tau\wedge\sigma_{n}}}{L_{\tau\wedge\sigma_{n}}}\right]=\mathbb{E}\left[\int_{[0,\sigma_{n}]}X_{t}\mathrm{d}K_{t}+X_{\sigma_{n}}(1-K_{\sigma_{n}})\right]=\mathbb{E}\left[\int_{[0,\sigma_{n}]}(1-K_{t-})\mathrm{d}X_{t}\right]=\mathbb{E}\left[X_{0}\right],

where the last equality makes use of inequality (2.7) (now applied with respect to the martingale XσnX^{\sigma_{n}}, with the convention K0=0K_{0-}=0), for all nn\in\mathbb{N}. This completes the argument. ∎

Proposition 2.7 shows that, up to a normalisation with respect to 1/Lτ1/L^{\tau}, the supermartingale property can always be transferred from the original filtration 𝐅\mathbf{F} to the enlarged filtration 𝐆\mathbf{G} and provides a sufficient criterion for transforming 𝐅\mathbf{F}-local martingales into 𝐆\mathbf{G}-local martingales. As shown in Section 2.4, this result will play a key role in proving Theorem 1.4.

In the rest of this section we provide a couple of interesting side results which, though not used in the sequel, are intimately connected to Proposition 2.7. The first one provides a characterisation of the local martingale property of Xτ/LτX^{\tau}/L^{\tau} on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}) for every nonnegative local martingale XX on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}).

Proposition 2.8.

The following statements are equivalent:

  1. (1)

    For every nonnegative local martingale XX on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), the process Xτ/LτX^{\tau}/L^{\tau} is a local martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

  2. (2)

    The process 1/Lτ1/L^{\tau} is a local martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

  3. (3)

    [η<]=0\mathbb{P}[\eta<\infty]=0.

Proof.

Implication (1) \Rightarrow (2) is trivial, while (3) \Rightarrow (1) follows from part (2) of Proposition 2.7. In order to prove (2) \Rightarrow (3), note that the sequence {τn}n\{\tau_{n}\}_{n\in\mathbb{N}} defined by τn:=inf{t+| 1/Ltτ>n}\tau_{n}\,:=\,\inf\left\{t\in\mathbb{R}_{+}\ |\ 1/L^{\tau}_{t}>n\right\}, for all nn\in\mathbb{N}, is a localising sequence for 1/Lτ1/L^{\tau} on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}). Define the sequence {νn}n\{\nu_{n}\}_{n\in\mathbb{N}} of stopping times on (Ω,𝐅)(\Omega,\,\mathbf{F}) via νn:=inf{t+|Lt<1/n}\nu_{n}\,:=\,\inf\left\{t\in\mathbb{R}_{+}\ |\ L_{t}<1/n\right\}, for all nn\in\mathbb{N}, and observe that τn=νn𝕀{νnτ}+𝕀{νn>τ}\tau_{n}=\nu_{n}\mathbb{I}_{\left\{\nu_{n}\leq\tau\right\}}+\infty\mathbb{I}_{\left\{\nu_{n}>\tau\right\}}. Then, by computations analogous to (2.8), we obtain

1=𝔼[1Lττn]=𝔼[1Lτνn]=𝔼[Kνn+𝕀{Lνn>0}(1Kνn)]=1𝔼[𝕀{Lνn=0}(1K)],1=\mathbb{E}\left[\frac{1}{L_{\tau\wedge\tau_{n}}}\right]=\mathbb{E}\left[\frac{1}{L_{\tau\wedge\nu_{n}}}\right]=\mathbb{E}\left[K_{\nu_{n}}+\mathbb{I}_{\left\{L_{\nu_{n}}>0\right\}}(1-K_{\nu_{n}})\right]=1-\mathbb{E}\left[\mathbb{I}_{\left\{L_{\nu_{n}}=0\right\}}(1-K_{\infty})\right],

where in the last equality we have used the fact that KK does not increase on {L=0}\left\{L=0\right\}. In turn, this implies that {K<1}{Lνn=0}=\left\{K_{\infty}<1\right\}\cap\left\{L_{\nu_{n}}=0\right\}=\emptyset holds (modulo evanescence). Due to Lemma 2.4 and since {ΔK>0}{L>0}\left\{\Delta K>0\right\}\subseteq\left\{L>0\right\} holds modulo evanescence (see [Kar15]), this implies that [η<]=0\mathbb{P}\left[\eta<\infty\right]=0. ∎

Part (1) of Proposition 2.7 leads to a quick and easy proof of the classical result of [JY78] on the semimartingale property of XτX^{\tau} on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}) for any semimartingale XX on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}).

Corollary 2.9.

For any semimartingale XX on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), the process XτX^{\tau} is a semimartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

Proof.

Let XX be a semimartingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), so that X=X0+B+NX=X_{0}+B+N, for some adapted process of finite variation BB and a local martingale NN on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). By [JS03, Proposition I.4.17], it holds that N=N+N′′N=N^{\prime}+N^{\prime\prime}, where NN^{\prime} and N′′N^{\prime\prime} are two local martingales on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) such that |ΔN|a|\Delta N^{\prime}|\leq a \mathbb{P}-a.s. for some a>0a>0 and N′′N^{\prime\prime} is of finite variation. In order to prove the claim it suffices to show that (N)τ(N^{\prime})^{\tau} is a semimartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}). To this effect, let σn:=inf{t0:|Nt|n}\sigma_{n}:=\inf\{t\geq 0:|N^{\prime}_{t}|\geq n\}, for nn\in\mathbb{N}, so that Ntσn(a+n)N^{\prime}_{t\wedge\sigma_{n}}\geq-(a+n) \mathbb{P}-a.s. for all t0t\geq 0. Hence, by part (1) of Proposition 2.7, the process (a+n+N)σnτ/Lσnτ(a+n+N^{\prime})^{\sigma_{n}\wedge\tau}/L^{\sigma_{n}\wedge\tau} is a supermartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}). In turn, this implies the semimartingale property of (N)σnτ(N^{\prime})^{\sigma_{n}\wedge\tau} on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}). Since semimartingales are stable by localization (see e.g. [JS03, Proposition I.4.25]), this shows the semimartingale property of (N)τ(N^{\prime})^{\tau} on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}). ∎

2.4. Condition NA1 in the progressively enlarged filtration

As a consequence of Proposition 2.7, a sufficient condition for NA(𝐆,Sτ)1{}_{1}(\mathbf{G},S^{\tau}) to hold is immediate. The proof of the following result is straightforward, hence omitted. The notation 𝒴(𝐆,Sτ,)\mathcal{Y}(\mathbf{G},S^{\tau},\mathbb{P}) is self-explanatory.

Proposition 2.10.

Suppose that there exists a local martingale deflator MM for SS on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) such that {M>0}=[[0,η[[\left\{M>0\right\}=[\kern-1.49994pt[0,\eta[\kern-1.49994pt[. Then, Mτ/Lτ𝒴(𝐆,Sτ,)M^{\tau}/L^{\tau}\in\mathcal{Y}(\mathbf{G},S^{\tau},\mathbb{P}).

In particular, observe that Proposition 2.10 provides an explicit procedure for transforming a local martingale deflator for SS on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) into a local martingale deflator for SτS^{\tau} on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}). We are now ready to present the proofs of our results on NA1 stability under progressive filtration enlargement.

Proof of Theorem 1.4.

In view of Lemma 2.6 and Theorem 1.2, we may assume without loss of generality (replacing \mathbb{P} with \mathbb{Q} if necessary) the existence of a strictly positive X^𝒳(𝐅,S)\widehat{X}\in\mathcal{X}(\mathbf{F},S) such that Y:=(1/X^)𝒴(𝐅,S,)Y\,:=\,(1/\widehat{X})\in\mathcal{Y}(\mathbf{F},S,\mathbb{P}). Since [η<,ΔSη0]=0\mathbb{P}\left[\eta<\infty,\Delta S_{\eta}\neq 0\right]=0 holds, we obtain [η<,ΔYη0]=0\mathbb{P}\left[\eta<\infty,\Delta Y_{\eta}\neq 0\right]=0; in particular, [η<,Δ(YS)η0]=0\mathbb{P}\left[\eta<\infty,\Delta(YS)_{\eta}\neq 0\right]=0 holds. In the notation of Lemma 2.5, define M:=Y(D)1𝕀[[0,η[[M\,:=\,Y\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[}. Note that M0=1M_{0}=1 and {M>0}=[[0,η[[\left\{M>0\right\}=[\kern-1.49994pt[0,\eta[\kern-1.49994pt[. By Lemma 2.5, it follows that MSi[(D)1𝕀[[0,η[[,YSi]MS^{i}-\left[\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[},YS^{i}\right] is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) for all i{1,,d}{i\in\left\{1,\ldots,d\right\}}. Furthermore,

[(D)1𝕀[[0,η[[,YSi]=[(D)1,YSi][(D)1𝕀[[η,[[,YSi]=[(D)1,YSi],\left[\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[},YS^{i}\right]=\left[\mathcal{E}(-D)^{-1},YS^{i}\right]-\left[\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[\eta,\infty[\kern-1.22911pt[},YS^{i}\right]=\left[\mathcal{E}(-D)^{-1},YS^{i}\right],

where [(D)1𝕀[[η,[[,YSi]=0\left[\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[\eta,\infty[\kern-1.22911pt[},YS^{i}\right]=0 follows from the fact that (D)1𝕀[[η,[[=(D)η1𝕀[[η,[[\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[\eta,\infty[\kern-1.22911pt[}=\mathcal{E}(-D)^{-1}_{\eta}\mathbb{I}_{[\kern-1.22911pt[\eta,\infty[\kern-1.22911pt[} is a single-jump process, jumping at η\eta. Since (D)1\mathcal{E}(-D)^{-1} is predictable, it follows that

[(D)1𝕀[[0,η[[,YSi]=[(D)1,YSi]=0Δ(D)t1d(YSi)t\left[\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[},YS^{i}\right]=\left[\mathcal{E}(-D)^{-1},YS^{i}\right]=\int_{0}^{\cdot}\Delta\mathcal{E}(-D)^{-1}_{t}\mathrm{d}\left(YS^{i}\right)_{t}

is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) for all i{1,,d}{i\in\left\{1,\ldots,d\right\}}. Therefore, MSiMS^{i} is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) for all i{1,,d}{i\in\left\{1,\ldots,d\right\}}, and Theorem 1.4 follows from Proposition 2.10. ∎

Proof of Theorem 1.7.

Statement (1) follows directly from Theorem 1.4.

For statement (2), let DD be as in Lemma 2.5, and define S=(D)1𝕀[[0,η[[S=\mathcal{E}(-D)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta[\kern-1.22911pt[}. Then S0=1S_{0}=1 and SS is a nonincreasing process up to τ\tau, thus Sτ1S_{\tau}\geq 1. Moreover, by Lemma 2.5, SS is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), hence NA1(𝐅,S)(\mathbf{F},S) holds. From (2.1) and Z=L(1K)Z=L(1-K), and using integration by parts and the definition of DD, we have

𝔼[Dτ]\displaystyle\mathbb{E}[D_{\tau}] =𝔼[0DtLtdKt]=𝔼[0DtdZt]=𝔼[0ZtdDt]=𝔼[0Ztd𝕀{ηt}]\displaystyle=\mathbb{E}\left[\int_{0}^{\infty}D_{t}L_{t}\mathrm{d}K_{t}\right]=-\mathbb{E}\left[\int_{0}^{\infty}D_{t}\mathrm{d}Z_{t}\right]=\mathbb{E}\left[\int_{0}^{\infty}Z_{t-}\mathrm{d}D_{t}\right]=\mathbb{E}\left[\int_{0}^{\infty}Z_{t-}\mathrm{d}\mathbb{I}_{\{\eta\leq t\}}\right]
=𝔼[Zη𝕀{η<}].\displaystyle=\mathbb{E}[Z_{\eta-}\mathbb{I}_{\{\eta<\infty\}}].

Therefore, if [η<]>0\mathbb{P}\left[\eta<\infty\right]>0, then [Dτ>0]>0\mathbb{P}\left[D_{\tau}>0\right]>0, hence [Sτ>1]>0\mathbb{P}\left[S_{\tau}>1\right]>0. This means that NA1(𝐆,Sτ)(\mathbf{G},S^{\tau}) fails, concluding the proof. ∎

Note that, in view of Proposition 2.8, Theorem 1.7 implies that NA1 is stable for all semimartingale models if and only if the process 1/Lτ1/L^{\tau} is a local martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

Remark 2.11.

Proposition 2.8 allows to give a direct proof of statement (1) of Theorem 1.7. Indeed, in view of [TS14, Theorem 2.6], NA1(𝐅,S)(\mathbf{F},S) is equivalent to the existence of a process Y𝒴(𝐅,S,)Y\in\mathcal{Y}(\mathbf{F},S,\mathbb{P}). Due to Proposition 2.8, if [η<]=0\mathbb{P}\left[\eta<\infty\right]=0, then Yτ/LτY^{\tau}/L^{\tau} and (Yτ/Lτ)Sτ(Y^{\tau}/L^{\tau})S^{\tau} are local martingales on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}), so that Yτ/Lτ𝒴(𝐆,Sτ,)Y^{\tau}/L^{\tau}\in\mathcal{Y}(\mathbf{G},S^{\tau},\mathbb{P}). Hence, by [TS14, Theorem 2.6], NA1(𝐆,Sτ)(\mathbf{G},S^{\tau}) holds.

Note, however, that this line of reasoning cannot be applied to prove Theorem 1.4. In fact, in order to construct a strictly positive local martingale deflator in the enlarged filtration 𝐆\mathbf{G} starting from an element of 𝒴(𝐅,S,)\mathcal{Y}(\mathbf{F},S,\mathbb{P}) and relying on Proposition 2.7, one needs to show that NA1(𝐅,S)(\mathbf{F},S) and [η<,ΔSη0]=0\mathbb{P}\left[\eta<\infty,\Delta S_{\eta}\neq 0\right]=0 together imply the existence of a strictly positive local martingale deflator which does not jump at η\eta. For this property to hold, we need a more precise statement of the main result of [TS14] in the form of Theorem 1.2.

2.5. A partial converse to Proposition 2.10

While Proposition 2.10 is sufficient for establishing the NA1 stability under progressive enlargement in Theorem 1.4, here we address the inverse problem. Precisely, we seek conditions ensuring the existence of a deflator for SS in 𝐅\mathbf{F} once a deflator for SτS^{\tau} exists in the enlarged filtration 𝐆\mathbf{G}. Additionally, we want the deflator in 𝐅\mathbf{F} to vanish on [[η,[[[\kern-1.49994pt[\eta,\infty[\kern-1.49994pt[, in order to end up in the setting of Proposition 2.10. The next result shows that this is indeed the case when τ\tau avoids all stopping times on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), meaning that [τ=σ<]=0\mathbb{P}\left[\tau=\sigma<\infty\right]=0 holds for all stopping times σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}).

Theorem 2.12.

Suppose that τ\tau avoids all stopping times on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). If 𝒴(𝐆,Sτ,)\mathcal{Y}(\mathbf{G},S^{\tau},\mathbb{P})\neq\emptyset, then there exists a local martingale deflator YY for SS on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), with Y=0Y=0 on [[η,[[[\kern-1.49994pt[\eta,\infty[\kern-1.49994pt[.

Proof.

Let CC be the predictable compensator of 𝕀[[τ,[[\mathbb{I}_{[\kern-1.22911pt[\tau,\infty[\kern-1.22911pt[} on (Ω,𝐆)(\Omega,\,\mathbf{G}), and note that for every predictable time σ\sigma in (Ω,𝐆)(\Omega,\,\mathbf{G}) it holds that ΔCσ=[τ=σ|𝒢σ]\Delta C_{\sigma}=\mathbb{P}\left[\tau=\sigma\ |\ \mathcal{G}_{\sigma-}\right] on {σ<}\left\{\sigma<\infty\right\}. Now, by assumption τ\tau avoids all stopping times on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), hence in particular all the predictable ones, which is equivalent to say that τ\tau is a totally inaccessible stopping time on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}); see [Jeu80, p.65]. From this fact it follows that ΔCσ=0\Delta C_{\sigma}=0 holds on {σ<}\left\{\sigma<\infty\right\} for every predictable time σ\sigma in (Ω,𝐆)(\Omega,\,\mathbf{G}). The predictable section theorem then implies that CC is continuous, thus, in particular, the process (C)1𝕀[[0,τ[[\mathcal{E}(-C)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\tau[\kern-1.22911pt[} is well-defined. Now, by the same arguments used in the proof of Lemma 2.5, it holds that (C)1𝕀[[0,τ[[\mathcal{E}(-C)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\tau[\kern-1.22911pt[} is a local martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}). Take M𝒴(𝐆,Sτ,)M\in\mathcal{Y}(\mathbf{G},S^{\tau},\mathbb{P}). Since τ\tau avoids all stopping times on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), then ΔSτ=0\Delta S_{\tau}=0 and, as in the proof of Theorem 1.4, we can assume without loss of generality that Δ(MS)τ=0\Delta(MS)_{\tau}=0 as well. These two facts allow us to repeat the same steps as in the proof of Theorem 1.4 to show that U:=M(C)1𝕀[[0,τ[[U:=M\mathcal{E}(-C)^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\tau[\kern-1.22911pt[} is a local martingale deflator for SτS^{\tau} on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

Now, define YY as the optional projection of UU on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). Note that Y0=1Y_{0}=1 and that Y=0Y=0 on [[η,[[[\kern-1.49994pt[\eta,\infty[\kern-1.49994pt[, since [τ<η]=1\mathbb{P}\left[\tau<\eta\right]=1 (see the discussion after (1.3)). Let (σn)n(\sigma^{\prime}_{n})_{n\in\mathbb{N}} be a localising sequence for UU on (Ω,𝐆)(\Omega,\,\mathbf{G}), and let (σn)n(\sigma_{n})_{n\in\mathbb{N}} be a sequence of stopping times on (Ω,𝐅)(\Omega,\,\mathbf{F}) such that σnτ=σnτ\sigma^{\prime}_{n}\wedge\tau=\sigma_{n}\wedge\tau for nn\in\mathbb{N}. Then it is easily verified that YY is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), with (σn)n(\sigma_{n})_{n\in\mathbb{N}} as a localising sequence. Moreover, for any stopping time σ\sigma in (Ω,𝐅)(\Omega,\,\mathbf{F}) we have

𝔼[SσσniYσσn]=𝔼[SσσniUσσn]=𝔼[(Si)σσnτUσσn]=𝔼[(Si)σσnτUσσn]=S0i,i{1,,d}.\mathbb{E}[S^{i}_{\sigma\wedge\sigma_{n}}Y_{\sigma\wedge\sigma_{n}}]=\mathbb{E}[S^{i}_{\sigma\wedge\sigma_{n}}U_{\sigma\wedge\sigma_{n}}]=\mathbb{E}[(S^{i})^{\tau}_{\sigma\wedge\sigma_{n}}U_{\sigma\wedge\sigma_{n}}]=\mathbb{E}[(S^{i})^{\tau}_{\sigma\wedge\sigma^{\prime}_{n}}U_{\sigma\wedge\sigma^{\prime}_{n}}]=S^{i}_{0},\quad\forall{i\in\left\{1,\ldots,d\right\}}.

This shows that YSiYS^{i} is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) for all i{1,,d}{i\in\left\{1,\ldots,d\right\}} and concludes the proof. ∎

3. Arbitrage of the First Kind in Initially Enlarged Filtrations

In this section, the proof of Theorem 1.11 and Theorem 1.12 will be given, and interesting side results will also be discussed. The validity of Jacod’s criterion (Assumption 1.9) is tacitly assumed throughout. We start by proving the existence of a good version of conditional densities for JJ.

Proof of Lemma 1.10.

Denote by 𝒪(𝐅¯)\mathcal{O}(\overline{\mathbf{F}}) the optional σ\sigma-field associated to the filtration 𝐅¯=(¯t)t+\overline{\mathbf{F}}=(\overline{\mathcal{F}}_{t})_{t\in\mathbb{R}_{+}} on E×ΩE\times\Omega defined by ¯t:=s>t(Es)\overline{\mathcal{F}}_{t}:=\bigcap_{s>t}\left(\mathcal{B}_{E}\otimes\mathcal{F}_{s}\right), t+t\in\mathbb{R}_{+}. Note that E𝒪(𝐅)𝒪(𝐅¯)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})\subseteq\mathcal{O}(\overline{\mathbf{F}}) (see [Jac85]). By [Jac85, Lemma 1.8], Assumption 1.9 implies the existence of an 𝒪(𝐅¯)\mathcal{O}(\overline{\mathbf{F}})-measurable nonnegative function p~:(x,ω,t)p~tx(ω)\tilde{p}:(x,\omega,t)\mapsto\tilde{p}^{x}_{t}(\omega) such that (i)(i)-(ii)(ii) hold. Since, for every xEx\in E, the process p~x\tilde{p}^{x} is 𝐅\mathbf{F}-optional, being 𝐅\mathbf{F}-adapted and càdlàg, Remark 1 after Proposition 3 of [SY78] gives the existence of a E𝒪(𝐅)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})-measurable version pp of p~\tilde{p}. ∎

The following consequence of Lemma 1.10 will be used in several places: for any t+t\in\mathbb{R}_{+} and (Et)\left(\mathcal{B}_{E}\otimes\mathcal{F}_{t}\right)-measurable function E×Ω×+(x,ω,t)ftx(ω)+E\times\Omega\times\mathbb{R}_{+}\ni(x,\omega,t)\mapsto f^{x}_{t}(\omega)\in\mathbb{R}_{+}, it holds that

(3.1) 𝔼[ftJ]=𝔼[Eftxptxγ[dx]]=E𝔼[ftxptx]γ[dx].\mathbb{E}\left[f_{t}^{J}\right]=\mathbb{E}\left[\int_{E}f_{t}^{x}\,p^{x}_{t}\,\gamma[\mathrm{d}x]\right]=\int_{E}\mathbb{E}\left[f_{t}^{x}\,p^{x}_{t}\right]\gamma[\mathrm{d}x].

3.1. Results regarding the stopping times (ηx)xE\left(\eta^{x}\right)_{x\in E}

The next result can be regarded as a counterpart to Lemma 2.5 in the case of initially enlarged filtrations. Note that 𝒫(𝐅)\mathcal{P}(\mathbf{F}) denotes the 𝐅\mathbf{F}-predictable σ\sigma-field on Ω×+\Omega\times\mathbb{R}_{+} in all that follows.

Lemma 3.1.

Fix xEx\in E, and let DxD^{x} be the predictable compensator of 𝕀[[ηx,[[\mathbb{I}_{\left[\negthinspace\left[\right.\right.\!\eta^{x},\infty\left[\negthinspace\left[\right.\right.\!} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), with ηx\eta^{x} defined in (1.5). Then:

  1. (1)

    ΔDx<1\Delta D^{x}<1, \mathbb{P}-a.s.; in particular, (Dx)\mathcal{E}(-D^{x}) is nonincreasing and strictly positive;

  2. (2)

    the nonnegative process (Dx)1𝕀[[0,ηx[[\mathcal{E}(-D^{x})^{-1}\mathbb{I}_{[\kern-1.22911pt[0,\eta^{x}[\kern-1.22911pt[} is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}).

Suppose moreover that the space 𝕃1(Ω,,)\mathbb{L}^{1}(\Omega,\mathcal{F},\mathbb{P}) is separable. Then, the function E×Ω×+(x,ω,t)(Dx)t(ω)E\times\Omega\times\mathbb{R}_{+}\ni(x,\omega,t)\mapsto\mathcal{E}(-D^{x})_{t}(\omega) can be chosen E𝒫(𝐅)\mathcal{B}_{E}\otimes\mathcal{P}(\mathbf{F})-measurable.

Remark 3.2.

Note that separability of 𝕃1(Ω,,)\mathbb{L}^{1}(\Omega,\mathcal{F},\mathbb{P}) is only needed to ensure that the collection ((Dx))xE\left(\mathcal{E}(-D^{x})\right)_{x\in E} introduced in Lemma 3.1 admits a version with good measurability properties.

Proof.

Fix xEx\in E. For any 𝐅\mathbf{F}-predictable time σ\sigma, it holds that ΔDσx=[ηx=σ|σ]\Delta D^{x}_{\sigma}=\mathbb{P}\left[\eta^{x}=\sigma|\mathcal{F}_{\sigma-}\right] on {σ<}\{\sigma<\infty\}. As in the proof of Lemma 2.5, if the event {[ηx=σ<|σ]=1}\{\mathbb{P}\left[\eta^{x}=\sigma<\infty|\mathcal{F}_{\sigma-}\right]=1\} has positive probability, one can find an predictable time σ~\tilde{\sigma} on (Ω,𝐅)(\Omega,\,\mathbf{F}) such that [ηx=σ~<]>0\mathbb{P}\left[\eta^{x}=\tilde{\sigma}<\infty\right]>0 and {ηx=σ~<}σ~\{\eta^{x}=\tilde{\sigma}<\infty\}\in\mathcal{F}_{\tilde{\sigma}-}. Then, by the martingale property of pxp^{x} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) and the definition of ηx\eta^{x},

0=𝔼[Δpσ~x|σ~]=𝔼[Δpηxx|σ~]=𝔼[pηxx|σ~],on {ηx=σ~<}.0=\mathbb{E}\left[\Delta p^{x}_{\tilde{\sigma}}|\mathcal{F}_{\tilde{\sigma}-}\right]=\mathbb{E}\left[\Delta p^{x}_{\eta^{x}}|\mathcal{F}_{\tilde{\sigma}-}\right]=-\mathbb{E}\left[p^{x}_{\eta^{x}-}|\mathcal{F}_{\tilde{\sigma}-}\right],\qquad\text{on }\{\eta^{x}=\tilde{\sigma}<\infty\}.

In turn, since pηxx>0p^{x}_{\eta^{x}-}>0 holds on {ηx<}\{\eta^{x}<\infty\}, this implies that [ηx=σ~<]=0\mathbb{P}\left[\eta^{x}=\tilde{\sigma}<\infty\right]=0, thus leading to a contradiction and showing that [ηx=σ<|σ]<1\mathbb{P}\left[\eta^{x}=\sigma<\infty|\mathcal{F}_{\sigma-}\right]<1 holds in the \mathbb{P}-a.s. sense for any predictable time σ\sigma on (Ω,𝐅)(\Omega,\,\mathbf{F}). Part (1) then follows by the predictable section theorem, while part (2) can be proved by relying on the same arguments used in the proof of Lemma 2.5.

Finally, since 𝕃1(Ω,,)\mathbb{L}^{1}(\Omega,\mathcal{F},\mathbb{P}) is assumed separable, [SY78, Proposition 4] gives the existence of a E(+)\mathcal{B}_{E}\otimes\mathcal{F}\otimes\mathcal{B}(\mathbb{R}_{+})-measurable function (x,ω,t)Dtx(ω)(x,\omega,t)\mapsto D^{x}_{t}(\omega) such that, for all xEx\in E, DxD^{x} is the predictable compensator of 𝕀[[ηx,[[\mathbb{I}_{\left[\negthinspace\left[\right.\right.\!\eta^{x},\infty\left[\negthinspace\left[\right.\right.\!} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). Due to [SY78, Remark 1, after Proposition 3], the function DD can also be chosen E𝒫(𝐅)\mathcal{B}_{E}\otimes\mathcal{P}(\mathbf{F})-measurable and the same measurability property is inherited by the function (x,ω,t)(Dx)t(ω)(x,\omega,t)\mapsto\mathcal{E}(-D^{x})_{t}(\omega) (see also [SY78, § 12]). ∎

In order to establish our main results, we need to ensure that the collection (ηx)xE(\eta^{x})_{x\in E} of stopping times on (Ω,𝐅)(\Omega,\,\mathbf{F}) remains invariant under equivalent changes of measure, for γ\gamma-a.e. xEx\in E.

Lemma 3.3.

Let \mathbb{Q} be a probability measure on (Ω,)(\Omega,\mathcal{F}) with \mathbb{Q}\sim\mathbb{P}. For xEx\in E, let η,x\eta^{\mathbb{Q},x} be the stopping time on (Ω,𝐅)(\Omega,\,\mathbf{F}) defined under \mathbb{Q} in analogy to η,x:=ηx\eta^{\mathbb{P},x}:=\eta^{x} defined in (1.5) under \mathbb{P}. Then η,x=ηx\eta^{\mathbb{Q},x}=\eta^{x} holds almost surely (under both \mathbb{P} and \mathbb{Q}) for γ\gamma-a.e. xEx\in E.

Proof.

As can be readily checked, Assumption 1.9 is invariant under equivalent changes of probability. Hence, there exists a nonnegative E𝒪(𝐅)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})-measurable function E×Ω×+(x,ω,t)qtx(ω)E\times\Omega\times\mathbb{R}_{+}\ni(x,\omega,t)\mapsto q_{t}^{x}(\omega) satisfying the properties of Lemma 1.10 under \mathbb{Q}. Moreover, due to [Jac85, Corollary 1.11] (now applied under the probability \mathbb{Q}), it holds that [qtJ=0]=0\mathbb{Q}\left[q^{J}_{t}=0\right]=0 and also [qtJ=0]=0\mathbb{P}\left[q^{J}_{t}=0\right]=0, since \mathbb{Q}\sim\mathbb{P}, for all t+t\in\mathbb{R}_{+}. Hence, by using formula (3.1) applied to the Et\mathcal{B}_{E}\otimes\mathcal{F}_{t}-measurable function ftx=𝕀{qtx=0}f^{x}_{t}=\mathbb{I}_{\{q^{x}_{t}=0\}}, for t+t\in\mathbb{R}_{+}, we obtain

0=[qtJ=0]=E𝔼[𝕀{qtx=0}ptx]γ[dx],for all t+,0=\mathbb{P}\left[q^{J}_{t}=0\right]=\int_{E}\mathbb{E}\left[\mathbb{I}_{\{q^{x}_{t}=0\}}p^{x}_{t}\right]\gamma\left[\mathrm{d}x\right],\qquad\text{for all }t\in\mathbb{R}_{+},

so that {qtx=0}{ptx=0}\{q^{x}_{t}=0\}\subseteq\{p^{x}_{t}=0\} \mathbb{P}-a.s. for γ\gamma-a.e. xEx\in E. In a similar way, due to the symmetric role of \mathbb{P} and \mathbb{Q}, one can show that {ptx=0}{qtx=0}\{p^{x}_{t}=0\}\subseteq\{q^{x}_{t}=0\} holds \mathbb{Q}-a.s. for γ\gamma-a.e. xEx\in E and for all t+t\in\mathbb{R}_{+}. By right-continuity, {qx=0}={px=0}\{q^{x}=0\}=\{p^{x}=0\} holds (up to evanescence), for γ\gamma-a.e. xEx\in E. Hence, by definition, η,x=ηx\eta^{\mathbb{Q},x}=\eta^{x} holds almost surely (under both \mathbb{P} and \mathbb{Q}) for γ\gamma-a.e. xEx\in E. ∎

3.2. Super/local martingales in the initially enlarged filtration

The next result is a counterpart to Proposition 2.7 in the case of initially enlarged filtrations. Recall that [ζJ=]=1\mathbb{P}\left[\zeta^{J}=\infty\right]=1, as explained after (1.4), so that the optional process 1/pJ1/p^{J} on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}) is well-defined.

Proposition 3.4.

Let X:E×Ω×++X:E\times\Omega\times\mathbb{R}_{+}\mapsto\mathbb{R}_{+} be E𝒪(𝐅)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})-measurable, and such that XxX^{x} is càdlàg for every xEx\in E. Then the following statements hold.

  1. (1)

    If XxX^{x} is a supermartingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) for γ\gamma-a.e. xEx\in E, then, XJ/pJX^{J}/p^{J} is a supermartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

  2. (2)

    If XxX^{x} is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) and [[ηx,[[{Xx=0}\left[\negthinspace\left[\right.\right.\!\!\eta^{x},\infty\!\left[\negthinspace\left[\right.\right.\!\subseteq\{X^{x}=0\} (modulo evanescence) hold for γ\gamma-a.e. xEx\in E, then, XJ/pJX^{J}/p^{J} is a local martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

Proof.

We first prove part (1). For any sts\leq t, AsA\in\mathcal{F}_{s} and h:(E,E)(+,+)h:(E,\mathcal{B}_{E})\rightarrow(\mathbb{R}_{+},\mathcal{B}_{\mathbb{R}_{+}}), using the fact that [ζJ=]=1\mathbb{P}\left[\zeta^{J}=\infty\right]=1 together with formula (3.1) (with ft(x)=𝕀A{ζx>t}g(x)Xtx/ptxf_{t}(x)=\mathbb{I}_{A\cap\{\zeta^{x}>t\}}\,g(x)X^{x}_{t}/p^{x}_{t}) and the supermartingale property of XxX^{x} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), for γ\gamma-a.e. xEx\in E, we obtain

𝔼[𝕀Ag(J)XtJptJ]\displaystyle\mathbb{E}\left[\mathbb{I}_{A}\,g(J)\frac{X_{t}^{J}}{p^{J}_{t}}\right] =𝔼[𝕀A{ζJ>t}g(J)XtJptJ]\displaystyle=\mathbb{E}\left[\mathbb{I}_{A\cap\{\zeta^{J}>t\}}g(J)\frac{X_{t}^{J}}{p^{J}_{t}}\right]
=Eg(x)𝔼[𝕀A{ζx>t}Xtx]γ[dx]\displaystyle=\int_{E}g(x)\mathbb{E}\left[\mathbb{I}_{A\cap\{\zeta^{x}>t\}}X^{x}_{t}\right]\gamma[\mathrm{d}x]
Eg(x)𝔼[𝕀A{ζx>s}Xtx]γ[dx]\displaystyle\leq\int_{E}g(x)\mathbb{E}\left[\mathbb{I}_{A\cap\{\zeta^{x}>s\}}X^{x}_{t}\right]\gamma[\mathrm{d}x]
Eg(x)𝔼[𝕀A{ζx>s}Xsx]γ[dx]=𝔼[𝕀Ag(J)XsJpsJ].\displaystyle\leq\int_{E}g(x)\mathbb{E}\left[\mathbb{I}_{A\cap\{\zeta^{x}>s\}}X^{x}_{s}\right]\gamma[\mathrm{d}x]=\mathbb{E}\left[\mathbb{I}_{A}\,g(J)\frac{X_{s}^{J}}{p^{J}_{s}}\right].

By the monotone class theorem, this shows that XJ/pJX^{J}/p^{J} is a supermartingale on (Ω,𝐆0,)(\Omega,\,\mathbf{G}^{0},\,\mathbb{P}). By right-continuity, this implies the supermartingale property on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}), thus proving part (1).

To prove part (2), note first that, since XxX^{x} is a nonnegative local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), hence a supermartingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), the sequence (σnx)n\left(\sigma^{x}_{n}\right)_{n\in\mathbb{N}} defined by σnx:=inf{t+|Xtx>n}\sigma^{x}_{n}:=\inf\{t\in\mathbb{R}_{+}\,|\,X^{x}_{t}>n\} for nn\in\mathbb{N} is localising for XxX^{x} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), for γ\gamma-a.e. xEx\in E. Moreover, since XX is E𝒪(𝐅)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})-measurable, the function E×Ω(x,ω)σnx(ω)tE\times\Omega\ni(x,\omega)\mapsto\sigma_{n}^{x}(\omega)\wedge t is Et\mathcal{B}_{E}\otimes\mathcal{F}_{t}-measurable for all t+t\in\mathbb{R}_{+} and nn\in\mathbb{N}, and, as a composition of measurable mappings, the function E×Ω(x,ω)Xσnx(ω)tx(ω)E\times\Omega\ni(x,\omega)\mapsto X^{x}_{\sigma^{x}_{n}(\omega)\wedge t}(\omega) is also Et\mathcal{B}_{E}\otimes\mathcal{F}_{t}-measurable, for all t+t\in\mathbb{R}_{+} and nn\in\mathbb{N} (compare also with [SY78], Remark 1 after Theorem 2). Since pp is E𝒪(𝐅)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})-measurable (see Lemma 1.10), the same reasoning allows to show that the function E×Ω(x,ω)Xσnx(ω)ζnx(ω)tx(ω)/pσnx(ω)ζnx(ω)tx(ω)E\times\Omega\ni(x,\omega)\mapsto X^{x}_{\sigma^{x}_{n}(\omega)\wedge\zeta^{x}_{n}(\omega)\wedge t}(\omega)/p^{x}_{\sigma^{x}_{n}(\omega)\wedge\zeta^{x}_{n}(\omega)\wedge t}(\omega) is Et\mathcal{B}_{E}\otimes\mathcal{F}_{t}-measurable for all t+t\in\mathbb{R}_{+} and nn\in\mathbb{N}, where the stopping time ζnx\zeta^{x}_{n} on (Ω,𝐅)(\Omega,\,\mathbf{F}) is defined in (1.4). Then, for any t0t\geq 0, formula (3.1) gives

𝔼[XσnJζnJtJpσnJζnJtJ]\displaystyle\mathbb{E}\left[\frac{X^{J}_{\sigma^{J}_{n}\wedge\zeta^{J}_{n}\wedge t}}{p^{J}_{\sigma^{J}_{n}\wedge\zeta^{J}_{n}\wedge t}}\right] =E𝔼[Xσnxζnxtxpσnxζnxtx𝕀{pσnxζnxtx>0}ptx]γ[dx]\displaystyle=\int_{E}\mathbb{E}\left[\frac{X^{x}_{\sigma^{x}_{n}\wedge\zeta^{x}_{n}\wedge t}}{p^{x}_{\sigma^{x}_{n}\wedge\zeta^{x}_{n}\wedge t}}\mathbb{I}_{\left\{p^{x}_{\sigma^{x}_{n}\wedge\zeta^{x}_{n}\wedge t}>0\right\}}p^{x}_{t}\right]\gamma[\mathrm{d}x]
=E𝔼[Xσnxζnxtx𝕀{pσnxζnxtx>0}]γ[dx]\displaystyle=\int_{E}\mathbb{E}\left[X^{x}_{\sigma^{x}_{n}\wedge\zeta^{x}_{n}\wedge t}\mathbb{I}_{\left\{p^{x}_{\sigma^{x}_{n}\wedge\zeta^{x}_{n}\wedge t}>0\right\}}\right]\gamma[\mathrm{d}x]
=E𝔼[Xσnxζnxtx]γ[dx]\displaystyle=\int_{E}\mathbb{E}\left[X^{x}_{\sigma^{x}_{n}\wedge\zeta^{x}_{n}\wedge t}\right]\gamma[\mathrm{d}x]
=E𝔼[X0x]γ[dx]=𝔼[X0Jp0J],\displaystyle=\int_{E}\mathbb{E}\left[X^{x}_{0}\right]\gamma[\mathrm{d}x]=\mathbb{E}\left[\frac{X^{J}_{0}}{p^{J}_{0}}\right],

where the second equality follows from the martingale property of pxp^{x} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) for all xEx\in E, the third equality from the fact that {pσnxζnxtx=0}={ηx=σnxζnxt}\{p^{x}_{\sigma^{x}_{n}\wedge\zeta^{x}_{n}\wedge t}=0\}=\{\eta^{x}=\sigma^{x}_{n}\wedge\zeta^{x}_{n}\wedge t\} together with the assumption that [[ηx,[[{Xx=0}\left[\negthinspace\left[\right.\right.\!\!\eta^{x},\infty\!\left[\negthinspace\left[\right.\right.\!\subseteq\{X^{x}=0\} for γ\gamma-a.e. xEx\in E, the fourth equality from the martingale property of XσnxxX^{x}_{\sigma^{x}_{n}\wedge\cdot} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}), for γ\gamma-a.e. xEx\in E and nn\in\mathbb{N}, and the last equality from all the previous steps in reverse order. In turn, since by part (1) the process (XJ/pJ)σnJζnJ(X^{J}/p^{J})^{\sigma^{J}_{n}\wedge\zeta^{J}_{n}} is a supermartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}) for all nn\in\mathbb{N}, this implies that (XJ/pJ)σnJζnJ(X^{J}/p^{J})^{\sigma^{J}_{n}\wedge\zeta^{J}_{n}} is a martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}) for all nn\in\mathbb{N}. Since [limnσnx=]=1\mathbb{P}\left[\lim_{n\to\infty}\sigma^{x}_{n}=\infty\right]=1 holds for every xEx\in E, and [ζJ=]=1\mathbb{P}\left[\zeta^{J}=\infty\right]=1, the sequence (σnJζnJ)n\left(\sigma^{J}_{n}\wedge\zeta^{J}_{n}\right)_{n\in\mathbb{N}} of stopping times on (Ω,𝐆)(\Omega,\,\mathbf{G}) satisfies [limn(σnJζnJ)=]=1\mathbb{P}\left[\lim_{n\to\infty}\left(\sigma^{J}_{n}\wedge\zeta^{J}_{n}\right)=\infty\right]=1, thus proving that XJ/pJX^{J}/p^{J} is a local martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}). ∎

A result analogous to part (1) of Proposition 3.4 has been recently established in [IP13] (see their Proposition 5.2). More specifically, according to their terminology, the process 1/pJ1/p^{J} is a universal supermartingale density for 𝐆\mathbf{G}.

Remark 3.5.

Proposition 3.4 can be used to establish that any semimartingale XX on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) remains a semimartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}). As was the case in Corollary 2.9, it suffices to show the result whenever XX is a nonnegative and bounded local martingale, thus a supermartingale, on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). By part (1) of Proposition 3.4, the process X/pJX/p^{J} is a a semimartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}); since also 1/pJ1/p^{J} is a strictly positive semimartingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}), the result follows.

We proceed with a result that is a ramification of Proposition 3.4 (this side result will not be used in other places). In the same spirit of Proposition 2.8, we can characterise the local martingale property of XJ/pJX^{J}/p^{J} on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}) for every E𝒪(𝐅)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})-measurable non-negative function XX such that XxX^{x} is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) for γ\gamma-a.e. xEx\in E.

Proposition 3.6.

The following statements are equivalent:

  1. (1)

    For any X:E×Ω×++X:E\times\Omega\times\mathbb{R}_{+}\mapsto\mathbb{R}_{+} that is E𝒪(𝐅)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})-measurable, and such that XxX^{x} is càdlàg for every xEx\in E, XxX^{x} is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) and [[ηx,[[{Xx=0}\left[\negthinspace\left[\right.\right.\!\!\eta^{x},\infty\!\left[\negthinspace\left[\right.\right.\!\subseteq\{X^{x}=0\} (modulo evanescence) hold for γ\gamma-a.e. xEx\in E, the process XJ/pJX^{J}/p^{J} is a local martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

  2. (2)

    The process 1/pJ1/p^{J} is a local martingale on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

  3. (3)

    [ηx<]=0\mathbb{P}\left[\eta^{x}<\infty\right]=0 holds for γ\gamma-a.e. xEx\in E.

Proof.

Implication (1) \Rightarrow (2) is trivial, while (3) \Rightarrow (1) follows from part (2) of Proposition 3.4. In order to prove (2) \Rightarrow (3), note that the sequence (ζnJ)n\left(\zeta^{J}_{n}\right)_{n\in\mathbb{N}} of stopping times on (Ω,𝐆)(\Omega,\,\mathbf{G}) is localising for 1/pJ1/p^{J} (see (1.4)), so that 𝔼[1/pζnJTJ]=𝔼[1/p0J]\mathbb{E}[1/p^{J}_{\zeta^{J}_{n}\wedge T}]=\mathbb{E}[1/p^{J}_{0}], for any T+T\in\mathbb{R}_{+}. Hence, due to formula (3.1) applied first to the E0\mathcal{B}_{E}\otimes\mathcal{F}_{0}-measurable function E×Ω(x,ω)𝕀{p0x(ω)>0}(1/p0x(ω))E\times\Omega\ni(x,\omega)\mapsto\mathbb{I}_{\{p^{x}_{0}(\omega)>0\}}\big{(}1/p^{x}_{0}(\omega)\big{)} and then to the Et\mathcal{B}_{E}\otimes\mathcal{F}_{t}-measurable function E×Ω(x,ω)𝕀{pζnxTx>0}(1/pζnxTx)E\times\Omega\ni(x,\omega)\mapsto\mathbb{I}_{\{p^{x}_{\zeta^{x}_{n}\wedge T}>0\}}\big{(}1/p^{x}_{\zeta^{x}_{n}\wedge T}\big{)},

E𝔼[𝕀{p0x>0}]γ(dx)=𝔼[1p0J]=𝔼[1pζnJTJ]\displaystyle\int_{E}\mathbb{E}\left[\mathbb{I}_{\{p^{x}_{0}>0\}}\right]\gamma(\mathrm{d}x)=\mathbb{E}\left[\frac{1}{p^{J}_{0}}\right]=\mathbb{E}\left[\frac{1}{p^{J}_{\zeta^{J}_{n}\wedge T}}\right] =E𝔼[1pζnxTx𝕀{pζnxTx>0}pTx]γ[dx]\displaystyle=\int_{E}\mathbb{E}\left[\frac{1}{p^{x}_{\zeta^{x}_{n}\wedge T}}\mathbb{I}_{\{p^{x}_{\zeta^{x}_{n}\wedge T}>0\}}p^{x}_{T}\right]\gamma[\mathrm{d}x]
=E𝔼[𝕀{pζnxTx>0}]γ[dx],\displaystyle=\int_{E}\mathbb{E}\left[\mathbb{I}_{\{p^{x}_{\zeta^{x}_{n}\wedge T}>0\}}\right]\gamma[\mathrm{d}x],

where in the last equality we have used the martingale property of pxp^{x} on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) for every xEx\in E. This implies that {p0x>0}{pζnxTx=0}=\{p^{x}_{0}>0\}\cap\{p^{x}_{\zeta^{x}_{n}\wedge T}=0\}=\emptyset holds (modulo evanescence) for γ\gamma-a.e. xEx\in E, for all T+T\in\mathbb{R}_{+}. Equivalently, it holds that [ηx=]=1\mathbb{P}\left[\eta^{x}=\infty\right]=1 for γ\gamma-a.e. xEx\in E. ∎

3.3. Condition NA1 in the initially enlarged filtration

In the spirit of Proposition 2.10, we can then establish a sufficient condition for the validity of NA1 in the initially enlarged filtration 𝐆\mathbf{G}. The proof of the next proposition is a straightforward application of Proposition 3.4. The notation 𝒴(𝐆,S,)\mathcal{Y}(\mathbf{G},S,\mathbb{P}) is clear.

Proposition 3.7.

Suppose that there exists a E𝒪(𝐅)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})-measurable function M:E×Ω×++M:E\times\Omega\times\mathbb{R}_{+}\mapsto\mathbb{R}_{+} such that M0x=1M^{x}_{0}=1 and MxM^{x} is càdlàg, for every xEx\in E, MxM^{x} and MxSM^{x}S are local martingales on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) and {Mx>0}[[0,ηx[[\{M^{x}>0\}\subseteq\left[\negthinspace\left[\right.\right.\!\!0,\eta^{x}\!\left[\negthinspace\left[\right.\right.\! hold for γ\gamma-a.e. xEx\in E. Then, MJ/pJ𝒴(𝐆,S,)M^{J}/p^{J}\in\mathcal{Y}(\mathbf{G},S,\mathbb{P}).

We are now in the position to prove our first main theorem in the framework of initial filtration enlargement.

Proof of Theorem 1.11.

We follow the proof of Theorem 1.4 in the case of progressively enlarged filtrations. In view of Theorem 1.2 and Lemma 3.3, we may assume without loss of generality the existence of a strictly positive X^𝒳(𝐅,S)\widehat{X}\in\mathcal{X}(\mathbf{F},S) such that Y:=1/X^𝒴(𝐅,S,)Y:=1/\widehat{X}\in\mathcal{Y}(\mathbf{F},S,\mathbb{P}). Since [ηx<,ΔSηx0]=0\mathbb{P}\left[\eta^{x}<\infty,\,\Delta S_{\eta^{x}}\neq 0\right]=0 holds for γ\gamma-a.e. xEx\in E, we obtain [ηx<,ΔYηx0]=0\mathbb{P}\left[\eta^{x}<\infty,\,\Delta Y_{\eta^{x}}\neq 0\right]=0 and [ηx<,Δ(YS)ηx0]=0\mathbb{P}\left[\eta^{x}<\infty,\,\Delta(YS)_{\eta^{x}}\neq 0\right]=0 for γ\gamma-a.e. xEx\in E. In the notation of Lemma 3.1, define the function E×Ω×+(x,ω,t)Mtx(ω):=Yt(ω)(Dx)t1(ω)𝕀{ηx(ω)>t}E\times\Omega\times\mathbb{R}_{+}\ni(x,\omega,t)\mapsto M^{x}_{t}(\omega)\,:=\,Y_{t}(\omega)\mathcal{E}(-D^{x})_{t}^{-1}(\omega)\mathbb{I}_{\{\eta^{x}(\omega)>t\}}. For all xEx\in E, the process MxM^{x} is càdlàg and satisfies M0x=1M^{x}_{0}=1 and {Mx>0}=[[0,ηx[[\{M^{x}>0\}=[\kern-1.49994pt[0,\eta^{x}[\kern-1.49994pt[. By part (2) of Lemma 3.1 and proceeding as in the proof of Theorem 1.4, it can be shown that MxM^{x} and MxSM^{x}S are local martingales on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}) for γ\gamma-a.e. xEx\in E. Moreover, due to the separability of 𝕃1(Ω,,)\mathbb{L}^{1}(\Omega,\mathcal{F},\mathbb{P}), Lemma 3.1 shows that (D)\mathcal{E}(-D) admits a E𝒫(𝐅)\mathcal{B}_{E}\otimes\mathcal{P}(\mathbf{F})-measurable version. Since 𝒫(𝐅)𝒪(𝐅)\mathcal{P}(\mathbf{F})\subseteq\mathcal{O}(\mathbf{F}), the conclusion then follows from Proposition 3.7. ∎

Finally, we provide the proof of our last main result.

Proof of Theorem 1.12.

Statement (1) follows directly from Theorem 1.11, by Remark 3.2 and since [ηx<]=0\mathbb{P}\left[\eta^{x}<\infty\right]=0 implies that DxD^{x} is indistinguishable from 11. We proceed with the proof of statement (2). Due to Lemma 3.1, the function E×Ω×+(x,ω,t)Stx(ω):=(Dx)t1(ω)𝕀{ηx(ω)>t}E\times\Omega\times\mathbb{R}_{+}\ni(x,\omega,t)\mapsto S^{x}_{t}(\omega)\,:=\,\mathcal{E}(D^{x})^{-1}_{t}(\omega)\mathbb{I}_{\{\eta^{x}(\omega)>t\}} is E𝒫(𝐅)\mathcal{B}_{E}\otimes\mathcal{P}(\mathbf{F})-measurable, and, therefore, also E𝒪(𝐅)\mathcal{B}_{E}\otimes\mathcal{O}(\mathbf{F})-measurable. Moreover, for all xEx\in E, SxS^{x} is a local martingale on (Ω,𝐅,)(\Omega,\,\mathbf{F},\,\mathbb{P}). Recall that [ηJ=ζJ=]=1\mathbb{P}\left[\eta^{J}=\zeta^{J}=\infty\right]=1 (see § 1.4), so that the process SJS^{J} is nondecreasing. Moreover, using in sequence formula (3.1), integration by parts and the properties of predictable compensators, we get, for any T(0,)T\in(0,\infty),

𝔼[DTJ]=E𝔼[DTxqTx]γ[dx]=E𝔼[(0,T]qtx𝑑Dtx]γ[dx]=E𝔼[qηxx𝕀{ηxT}]γ[dx].\mathbb{E}\left[D^{J}_{T}\right]=\int_{E}\mathbb{E}\left[D^{x}_{T}q^{x}_{T}\right]\gamma[\mathrm{d}x]=\int_{E}\mathbb{E}\left[\int_{(0,T]}q^{x}_{t-}dD^{x}_{t}\right]\gamma[\mathrm{d}x]=\int_{E}\mathbb{E}\left[q^{x}_{\eta^{x}-}\mathbb{I}_{\{\eta^{x}\leq T\}}\right]\gamma[\mathrm{d}x].

Hence, if E[ηx<]γ[dx]>0\int_{E}\mathbb{P}\left[\eta^{x}<\infty\right]\gamma\left[\mathrm{d}x\right]>0, then [DTJ>0]>0\mathbb{P}\left[D^{J}_{T}>0\right]>0 holds for some T(0,)T\in(0,\infty), which implies that [StJ=S0J,t+]<1\mathbb{P}\left[S^{J}_{t}=S^{J}_{0},\,\forall t\in\mathbb{R}_{+}\right]<1. ∎

Note that, in view of Proposition 3.6, the NA1 stability (in the sense of Theorem 1.12) in the enlarged filtration 𝐆\mathbf{G} is also equivalent to the local martingale property of 1/pJ1/p^{J} on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}).

Remark 3.8.

We want to point out that, analogously to Remark 2.11, Proposition 3.6 allows to give a direct proof of statement (1) of Theorem 1.12. Indeed, in view of [TS14, Theorem 2.6], NA1(𝐅,S)(\mathbf{F},S) is equivalent to the existence of a process Y𝒴(𝐅,S,)Y\in\mathcal{Y}(\mathbf{F},S,\mathbb{P}). Due to Proposition 3.6, if [ηx<]=0\mathbb{P}\left[\eta^{x}<\infty\right]=0 holds for γ\gamma-a.e. xEx\in E, then Y/pJY/p^{J} and (Y/pJ)S(Y/p^{J})S are local martingales on (Ω,𝐆,)(\Omega,\,\mathbf{G},\,\mathbb{P}), meaning that Y/pJ𝒴(𝐆,S,)Y/p^{J}\in\mathcal{Y}(\mathbf{G},S,\mathbb{P}). [TS14, Theorem 2.6] then implies that NA1(𝐆,S)(\mathbf{G},S) holds. However, as for the case of progressive enlargement, this argument fails to provide a direct proof of Theorem 1.11 (compare with Remark 2.11).

References

  • [ACDJ14] Anna Aksamit, Tahir Choulli, Jun Deng, and Monique Jeanblanc. Non-arbitrage up to random horizon for semimartingale models. Electronic preprint available at http://arxiv.org/abs/1310.1142v2, 2014.
  • [AIS98] Jürgen Amendinger, Peter Imkeller, and Martin Schweizer. Additional logarithmic utility of an insider. Stoch. Proc. Appl., 75(2):263–286, 1998.
  • [Ame00] Jürgen Amendinger. Martingale representation theorems for initially enlarged filtrations. Stoch. Proc. Appl., 89(1):101–116, 2000.
  • [Bau03] Fabrice Baudoin. Modeling anticipations in financial markets. In Peter Bank et al., editor, Paris-Princeton Lectures in Mathematical Finance 2002, pages 43–94. Springer, Berlin Heidelberg, 2003.
  • [CJN12] Delia Coculescu, Monique Jeanblanc, and Ashkan Nikeghbali. Default times, no-arbitrage conditions and changes of probability measures. Finance Stoch., 16(3):513–535, 2012.
  • [DS94] Freddy Delbaen and Walter Schachermayer. A general version of the fundamental theorem of asset pricing. Math. Ann., 300(3):463–520, 1994.
  • [DS98] Freddy Delbaen and Walter Schachermayer. The fundamental theorem of asset pricing for unbounded stochastic processes. Math. Ann., 312(2):215–250, 1998.
  • [FJS14] Claudio Fontana, Monique Jeanblanc, and Shiqi Song. On arbitrages arising with honest times. Finance Stoch., 18(3):515–543, 2014.
  • [Fon14] Claudio Fontana. No-arbitrage conditions and absolutely continuous changes of measure. In Caroline Hillairet, Monique Jeanblanc, and Ying Jiao, editors, Arbitrage, Credit and Informational Risks, volume 5 of Peking University Series in Mathematics, pages 3–18. World Scientific, Singapore, 2014.
  • [GP98] Axel Grorud and Monique Pontier. Insider trading in a continuous time market model. Int. J. Theor. Appl. Finance, 1(3):331–347, 1998.
  • [GP01] Axel Grorud and Monique Pontier. Asymmetrical information and incomplete markets. Int. J. Theor. Appl. Finance, 4(2):285–302, 2001.
  • [GVV06] Dario Gasbarra, Esko Valkeila, and Lioudmila Vostrikova. Enlargement of filtration and additional information in pricing models: Bayesian approach. In Yuri Kabanov, Robert Liptser, and Jordan Stoyanov, editors, From Stochastic Calculus to Mathematical Finance, pages 257–285. Springer, Berlin, 2006.
  • [GZ08] Xin Guo and Yan Zeng. Intensity process and compensator: a new filtration expansion approach and the Jeulin-Yor theorem. Ann. Appl. Probab., 18(1):120–142, 2008.
  • [HWY92] Sheng Wu He, Jia Gang Wang, and Jia An Yan. Semimartingale theory and stochastic calculus. Kexue Chubanshe (Science Press), Beijing, 1992.
  • [Imk02] Peter Imkeller. Random times at which insiders can have free lunches. Stoch. Stoch. Rep., 74(1-2):465–487, 2002.
  • [IP13] Peter Imkeller and Nicolas Perkowski. The existence of dominating local martingale measures. Forthcoming in Finance Stoch., 2013.
  • [IPW01] Peter Imkeller, Monique Pontier, and Ferenc Weisz. Free lunch and arbitrage possibilities in a financial market model with an insider. Stoch. Proc. Appl., 92(1):103–130, 2001.
  • [Jac85] Jean Jacod. Grossissement initial, hypothèse (H’), et théorème de Girsanov. In Thierry Jeulin and Marc Yor, editors, Grossissements de Filtrations: Exemples et Applications, volume 1118 of Lecture Notes in Mathematics, pages 15–35. Springer, Berlin - Heidelberg, 1985.
  • [Jeu80] Thierry Jeulin. Semi-martingales et grossissement d’une filtration, volume 833 of Lecture Notes in Mathematics. Springer, Berlin, 1980.
  • [JS03] Jean Jacod and Albert N. Shiryaev. Limit theorems for stochastic processes, volume 288 of Grundlehren der Mathematischen Wissenschaften. Springer-Verlag, Berlin, second edition, 2003.
  • [JY78] Thierry Jeulin and Marc Yor. Grossissement d’une filtration et semi-martingales: formules explicites. In Séminaire de Probabilités, XII (Univ. Strasbourg, Strasbourg, 1976/1977), volume 649 of Lecture Notes in Math., pages 78–97. Springer, Berlin, 1978.
  • [JYC09] Monique Jeanblanc, Marc Yor, and Marc Chesney. Mathematical methods for financial markets. Springer Finance. Springer-Verlag London Ltd., London, 2009.
  • [Kar10] Constantinos Kardaras. Finitely additive probabilities and the fundamental theorem of asset pricing. In Carl Chiarella and Alexander Novikov, editors, Contemporary Quantitative Finance: Essays in Honour of Eckhard Platen, pages 19–34. Springer, Berlin Heidelberg, 2010.
  • [Kar15] Constantinos Kardaras. On the stochastic behaviour of optional processes up to random times. Ann. Appl. Probab., 25(2):429–464, 2015.
  • [KH07] Arturo Kohatsu-Higa. Models for insider trading with finite utility. In René Carmona et al., editor, Paris-Princeton Lectures in Mathematical Finance 2004, pages 103–171. Springer, 2007.
  • [KHOL11] Arturo Kohatsu-Higa and Salvador Ortiz-Latorre. Modeling of financial markets with inside information in continuous time. Stoch. Dynam., 11(2):415–438, 2011.
  • [KK07] Ioannis Karatzas and Constantinos Kardaras. The numéraire portfolio in semimartingale financial models. Finance Stoch., 11(4):447–493, 2007.
  • [Kre13] Dörte Kreher. Change of measure up to a random time: details. Electronic preprint available at http://arxiv-web3.library.cornell.edu/abs/1309.6141, 2013.
  • [Nik06] Ashkan Nikeghbali. An essay on the general theory of stochastic processes. Probability Surveys, 3:345–412, 2006.
  • [Pro04] Philip Protter. Stochastic integration and differential equations. Applications of Mathematics (New York). Springer-Verlag, Berlin, 2.1 edition, 2004.
  • [SY78] Christophe Stricker and Marc Yor. Calcul stochastique dépendant d’un paramètre. Z. Wahrsch. Verw. Gebiete, 45(2):109–133, 1978.
  • [TS14] Koichiro Takaoka and Martin Schweizer. A note on the condition of no unbounded profit with bounded risk. Finance Stoch., 18(2):393–405, 2014.
  • [Yoe85] Chanta Yoeurp. Théorème de Girsanov généralisé et grossissement d’une filtration. In Thierry Jeulin and Marc Yor, editors, Grossissements de Filtrations: Exemples et Applications, volume 1118 of Lecture Notes in Mathematics, pages 172–196. Springer-Verlag, 1985.
  • [Zwi07] Jakub Zwierz. On existence of local martingale measures for insiders who can stop at honest times. Bull. Pol. Acad. Sci. Math., 55(2):183–192, 2007.