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

Arbitrage concepts under trading restrictions in discrete-time financial markets

Claudio Fontana  and  Wolfgang J. Runggaldier Department of Mathematics “Tullio Levi - Civita”, University of Padova, Italy. [email protected]; [email protected]
Abstract.

In a discrete-time setting, we study arbitrage concepts in the presence of convex trading constraints. We show that solvability of portfolio optimization problems is equivalent to absence of arbitrage of the first kind, a condition weaker than classical absence of arbitrage opportunities. We center our analysis on this characterization of market viability and derive versions of the fundamental theorems of asset pricing based on portfolio optimization arguments. By considering specifically a discrete-time setup, we simplify existing results and proofs that rely on semimartingale theory, thus allowing for a clear understanding of the foundational economic concepts involved. We exemplify these concepts, as well as some unexpected situations, in the context of one-period factor models with arbitrage opportunities under borrowing constraints.

Key words and phrases:
Trading constraints; market viability; arbitrage of the first kind; numéraire portfolio.
JEL classification: C02, C61, G11, G13.
Financial support from the University of Padova (research programme BIRD190200/19) and the Europlace Institute of Finance is gratefully acknowledged. Declarations of interest: none.

1. Introduction

The notions of arbitrage, market viability and state-price deflators are deeply connected and play a foundational role in financial economics and mathematical finance. Starting from the seminal works [Ros77, Ros78], the connections between these three concepts represent the essence of the fundamental theorem of asset pricing.111We refer the reader to [Sch10] for an excellent overview of the main steps in the development of discrete-time and continuous-time versions of the fundamental theorem of asset pricing. The terminology fundamental theorem of asset pricing has been introduced in [DR87]. In frictionless discrete-time financial markets, if no trading restrictions are imposed, the appropriate no-arbitrage concept takes the classical form of absence of arbitrage opportunities (no classical arbitrage). By the fundamental theorem of asset pricing of [HK79, HP81] (extended to general probability spaces in [DMW90]), this is equivalent to the existence of an equivalent martingale measure, whose density acts as a state-price deflator. Moreover, always in the absence of trading restrictions, the results of [RS06] imply that no classical arbitrage is equivalent to market viability, intended as the solvability of portfolio optimization problems. No classical arbitrage thus represents the minimal economically meaningful no-arbitrage requirement for a frictionless discrete-time financial market.

In the presence of trading restrictions, these results continue to hold true as long as the set of constrained strategies is a cone, provided that equivalent martingale measures are replaced by equivalent supermartingale measures (see [FS16, Theorem 9.9] and Theorem 2.12 below). However, many practically relevant trading restrictions, such as borrowing constraints or the possibility of limited short sales, correspond to convex non-conic constraints. In this case, as it will be shown below, market viability is no longer equivalent to no classical arbitrage, but rather to the weaker condition of no arbitrage of the first kind (NA1). Under convex trading restrictions, NA1 represents therefore the minimal economically meaningful concept of no-arbitrage and is equivalent to the existence of a numéraire portfolio or, more generally, a supermartingale deflator.

The NA1 condition, introduced under this terminology in [Kar10], corresponds to the absence of positive payoffs that can be super-replicated with an arbitrarily small initial capital and is equivalent to the no unbounded profit with bounded risk condition studied in the seminal work [KK07] (see also [Fon15] for an analysis of no-arbitrage conditions equivalent to NA1). In continuous-time, a complete theory based on NA1 has been developed in a general semimartingale setting starting with [KK07], also allowing for convex (non-conic) constraints. The connection between NA1 and market viability has been characterized in [CDM15] in an unconstrained semimartingale setting (see also [CCFM17] for further results in this direction).

Scarce attention has, however, been specifically paid to NA1 in discrete-time models, despite their widespread use in economic theory. This is also due to the fact that, for discrete-time markets with conic constraints, there is no distinction between NA1 and no classical arbitrage (see Remark 2.3 below). To the best of our knowledge, the only works that specifically address discrete-time models by relying on no-arbitrage requirements weaker than no classical arbitrage are [ES01] and [KS09]. In a one-period model on a finite probability space, [ES01] show that limited forms of arbitrage may coexist with market equilibrium under convex constraints (see Remark 2.7 below for a more detailed discussion). Closer to our setting, [KS09] derive the central results of [KK07] on the numéraire portfolio in a one-period setting.

The present paper intends to fill this gap in the literature, in the framework of general discrete-time models with convex (not necessarily conic) constraints. Compared to [ES01, KS09], we develop a complete theory of asset pricing based on NA1, also in the case of multi-period models with random convex constraints. We prove that market viability is equivalent to NA1, thereby showing that no classical arbitrage may pose unnecessary restrictions in the case of non-conic constraints. Building our analysis on this central result, we derive versions of the fundamental theorem of asset pricing, study the valuation of contingent claims and discuss non-trivial examples of our theory in the context of general factor models. We make a systematic effort to provide direct and self-contained proofs based on portfolio optimization arguments. The simplicity of the discrete-time structure allows for a clear understanding of the economic concepts involved, avoiding the technicalities of the continuous-time semimartingale setup.

The paper is divided into three sections, whose contents and contributions can be outlined as follows. In Section 2, we consider a general one-period setting. Extending the analysis of [KS09], we prove the equivalence between NA1 and the solvability of portfolio optimization problems (market viability), thus establishing the minimality of NA1 from an economic standpoint. This enables us to obtain a direct proof of the characterization of NA1 in terms of the existence of the numéraire portfolio or, more generally, a deflator. We show that NA1 leads to a dual representation of super-hedging values and a characterization of attainable claims, and permits to rely on several well-known hedging approaches in constrained incomplete markets, even in the presence of arbitrage opportunities. Besides its pedagogical value, the one-period setting introduces several techniques that will be important for the analysis of the multi-period case.

Section 3 illustrates the theory in the context of factor models with borrowing constraints. We introduce a general factor model, where a single factor is responsible of potential arbitrage opportunities. In this setting, the NA1 condition and the set of arbitrage opportunities admit explicit descriptions in terms of the factor loadings. When NA1 holds but no classical arbitrage does not, we show the existence of a maximal arbitrage strategy. These results can be easily visualized in a two-dimensional setting, which enables us to provide examples of situations where, despite the existence of arbitrage opportunities, it is not necessarily optimal to invest in them. The analysis of this section clarifies the interplay between the support of the asset returns distribution, their dependence structure and the borrowing constraints.

Finally, Section 4 generalizes the central results of Section 2 to a multi-period setting with random convex constraints. We derive several new characterizations of NA1, showing that it holds globally if and only if it holds in each single trading period, and prove its equivalence to market viability. The most general result on the solvability of portfolio optimization problems in discrete-time was obtained in [RS06], relying on no classical arbitrage. Our Theorem 4.5 extends this result by introducing trading restrictions and weakening the no-arbitrage requirement to the minimal condition of NA1 (in turn, our proofs of Theorems 2.5 and 4.5 are inspired from [RS06]). By generalizing the one-period analysis, we then give an easy proof of the equivalence between NA1, the existence of the numéraire portfolio and the existence of a supermartingale deflator, for general discrete-time models with random convex constraints.

We close this introduction by briefly reviewing some related literature, limiting ourselves to selected contributions that are specifically connected with the present discussion. Relying on the concept of no classical arbitrage, the fundamental theorem of asset pricing with constraints on the amounts invested in the risky assets is proved in [PT99] in the case of conic constraints (see also [KP00, Pha00] for valuation and hedging problems in that setting) and in [Bra97] in the case of convex constraints. The specific case of short-sale constraints is treated in the earlier work [JK95]. General forms of conic constraints have been considered in [Nap03], extending the analysis of [PT99]. In the case of convex constraints on the fractions of wealth invested, as considered in the present work, versions of the fundamental theorem of asset pricing based on the usual notion of no classical arbitrage are given in [CPT01, EST04, Rok05]. In comparison to the latter contributions, we choose to work with the weaker concept of NA1, due to its equivalence to market viability. In an unconstrained setting, the connection between no classical arbitrage and market viability is studied in [RS05, RS06], generalized in [Nut16] under model uncertainty. In the presence of model uncertainty and convex portfolio constraints, [BZ17] prove a version of the fundamental theorem of asset pricing based on a robust generalization of the notion of no classical arbitrage. Finally, we mention the recent work [BCL19], where super-hedging has been studied under a weak no-arbitrage condition, called absence of immediate profits. However, the latter condition does not suffice to ensure market viability.

2. The single-period setting

We consider a general financial market in a one-period economy, where dd risky assets are traded, together with a riskless asset with constant price equal to one. We assume that asset prices are discounted with respect to a baseline security and are represented by the vector St=(St1,,Std)+dS_{t}=(S^{1}_{t},\ldots,S^{d}_{t})^{\top}\in\mathbb{R}^{d}_{+}, for t=0,1t=0,1, expressed in terms of returns as

S1i=S0i(1+Ri), for all i=1,,d,S^{i}_{1}=S^{i}_{0}(1+R^{i}),\qquad\text{ for all }i=1,\ldots,d,

where R=(R1,,Rd)R=(R^{1},\ldots,R^{d})^{\top} is a dd-dimensional random vector on a given probability space (Ω,,P)(\Omega,\mathcal{F},P) such that Ri1R^{i}\geq-1 a.s., for all i=1,,di=1,\ldots,d. We denote by 𝒮\mathcal{S} the support of the distribution of RR, namely the smallest closed set AdA\subset\mathbb{R}^{d} such that P(RA)=1P(R\in A)=1 (see [FS16, Proposition 1.45]). We also denote by \mathcal{L} the smallest linear subspace of d\mathbb{R}^{d} containing 𝒮\mathcal{S} and by \mathcal{L}^{\perp} its orthogonal complement in d\mathbb{R}^{d}. The orthogonal projection of a vector xdx\in\mathbb{R}^{d} on \mathcal{L} is denoted by p(x){\rm p}_{\mathcal{L}}(x).

2.1. Trading restrictions

Trading strategies are denoted by vectors πd\pi\in\mathbb{R}^{d}. We write Vtπ(v)V^{\pi}_{t}(v) for the wealth at time tt generated by strategy π\pi starting from initial capital v>0v>0, with

V0π(v)=v and V1π(v)=v(1+π,R),V^{\pi}_{0}(v)=v\qquad\text{ and }\qquad V^{\pi}_{1}(v)=v(1+\langle\pi,R\rangle),

where ,\langle\cdot,\cdot\rangle denotes the scalar product in d\mathbb{R}^{d}. With this notation, a trading strategy π\pi represents fractions of wealth held in the dd risky assets, with the remaining fraction 1π,𝟏1-\langle\pi,\mathbf{1}\rangle being held in the riskless asset. For i=1,,di=1,\ldots,d, a negative value of πi\pi_{i} corresponds to a short position in the ii-th risky asset, whenever short sales are allowed. Similarly, π,𝟏<1\langle\pi,\mathbf{1}\rangle<1 corresponds to a positive investment in the riskless asset, while π,𝟏>1\langle\pi,\mathbf{1}\rangle>1 corresponds to borrowing from the riskless asset.

Note that Vtπ(v)=vVtπ(1)V^{\pi}_{t}(v)=vV^{\pi}_{t}(1), for all v>0v>0 and t=0,1t=0,1. In the following, we shall use the notation Vtπ:=Vtπ(1)V^{\pi}_{t}:=V^{\pi}_{t}(1). A trading strategy π\pi is said to be admissible if V1π0V^{\pi}_{1}\geq 0 a.s. Denoting by Θadm\Theta_{\rm adm} the set of all admissible trading strategies, it holds that (see, e.g., [KS99, Lemma 4.3])

Θadm={πd:π,z1 for all z𝒮}.\Theta_{\rm adm}=\{\pi\in\mathbb{R}^{d}:\langle\pi,z\rangle\geq-1\text{ for all }z\in\mathcal{S}\}.

In the terminology of [KK07, KS09], the set Θadm\Theta_{\rm adm} corresponds to the natural constraints ensuring non-negative wealth. Observe that, with the present parametrization, the notion of admissibility does not depend on the initial capital.

Besides the natural constraints, we assume that market participants face additional trading restrictions, represented by a convex closed set Θcd\Theta_{\rm c}\subseteq\mathbb{R}^{d}. Realistic examples of trading restrictions include the following situations (see also [CPT01, Section 4] for additional examples):

  1. (i)

    prohibition of short-selling: Θc=+d\Theta_{\rm c}=\mathbb{R}^{d}_{+};

  2. (ii)

    prohibition of short-selling and borrowing: Θc=Δd\Theta_{\rm c}=\Delta^{d}, where Δd:={π+d:π,𝟏1}\Delta^{d}:=\{\pi\in\mathbb{R}^{d}_{+}:\langle\pi,\mathbf{1}\rangle\leq 1\};

  3. (iii)

    limits to borrowing: Θc={πd:π,𝟏c}\Theta_{\rm c}=\{\pi\in\mathbb{R}^{d}:\langle\pi,\mathbf{1}\rangle\leq c\}, for some c1c\geq 1;

  4. (iv)

    limited positions in the risky assets: Θc=i=1d[αi,βi]\Theta_{\rm c}=\prod_{i=1}^{d}[-\alpha_{i},\beta_{i}], for some αi,βi>0\alpha_{i},\beta_{i}>0, i=1,,di=1,\ldots,d.

Market participants are restricted to choose strategies belonging to the set Θ:=ΘadmΘc\Theta:=\Theta_{\rm adm}\cap\Theta_{\rm c}. We refer to such strategies as allowed strategies. Observe that, as illustrated by examples (ii) and (iii) above, trading restrictions on the riskless asset can be enforced by specifying through the set Θc\Theta_{\rm c} suitable restrictions on the total fraction of wealth held in the dd risky assets.

In general, the financial market may contain redundant assets, meaning that different combinations of assets may generate identical portfolio returns. This happens whenever \mathcal{L}^{\perp} is strictly bigger than {0}\{0\}. Indeed, ρ\rho\in\mathcal{L}^{\perp} if and only if π,R=π+ρ,R\langle\pi,R\rangle=\langle\pi+\rho,R\rangle a.s. for every πΘ\pi\in\Theta. In other words, investing according to a strategy ρ\rho\in\mathcal{L}^{\perp} does not produce any loss or profit and, therefore, does not alter the outcome of any other allowed strategy π\pi. For this reason, we shall assume that investors are always allowed to choose trading strategies in the set \mathcal{L}^{\perp}, meaning that Θc\mathcal{L}^{\perp}\subset\Theta_{{\rm c}}. In turn, this implies that Θ\mathcal{L}^{\perp}\subset\Theta.

To the convex closed set Θ\Theta, we associate its recession cone Θ^\widehat{\Theta}, defined as the set of all vectors ydy\in\mathbb{R}^{d} such that π+λyΘ\pi+\lambda y\in\Theta for every λ0\lambda\geq 0 and πΘ\pi\in\Theta (see [Roc70, Chapter 8]). The set Θ^\widehat{\Theta} has a clear financial interpretation: it represents the set of all allowed strategies that can be arbitrarily scaled and added to any other strategy πΘ\pi\in\Theta without violating admissibility and trading restrictions. The cone Θ^\widehat{\Theta} is closed and, by [Roc70, Corollary 8.3.2], it holds that

Θ^={πd:a1πΘ for all a>0}=a>0aΘ.\widehat{\Theta}=\bigl{\{}\pi\in\mathbb{R}^{d}:a^{-1}\pi\in\Theta\text{ for all }a>0\bigr{\}}=\bigcap_{a>0}a\Theta.

As a consequence of the fact that \mathcal{L}^{\perp} is a linear subspace of Θ\Theta, it holds that Θ^\mathcal{L}^{\perp}\subseteq\widehat{\Theta}. In turn, the latter property can be easily seen to imply that p(Θ)Θ{\rm p}_{\mathcal{L}}(\Theta)\subseteq\Theta, i.e., p(π)Θ{\rm p}_{\mathcal{L}}(\pi)\in\Theta for all πΘ\pi\in\Theta.

2.2. Arbitrage concepts

We proceed to recall two important notions of arbitrage. First, we define the set222The definition of the set arb\mathcal{I}_{\rm arb} is coherent with the classical definition of arbitrage (see [FS16, Definition 1.2]). Indeed, arb\mathcal{I}_{\rm arb}\neq\emptyset if and only if there exists a portfolio (ϑ0,ϑ)×d(\vartheta_{0},\vartheta)\in\mathbb{R}\times\mathbb{R}^{d} such that ϑ0+ϑ,S0=0\vartheta_{0}+\langle\vartheta,S_{0}\rangle=0, ϑ0+ϑ,S10\vartheta_{0}+\langle\vartheta,S_{1}\rangle\geq 0 a.s. and P(ϑ0+ϑ,S1>0)>0P(\vartheta_{0}+\langle\vartheta,S_{1}\rangle>0)>0, with ϑ0\vartheta_{0} and ϑ\vartheta denoting respectively the number of shares of the riskless asset and of the dd risky assets. Assuming without loss of generality that S0i>0S^{i}_{0}>0, for all i=1,,di=1,\ldots,d, this equivalence follows in a straightforward way by setting ϑi=πi/S0i\vartheta_{i}=\pi_{i}/S^{i}_{0}, for i=1,,di=1,\ldots,d, and ϑ0=π,𝟏\vartheta_{0}=-\langle\pi,\mathbf{1}\rangle. This shows that absence of arbitrage can be equivalently understood as a property of the returns RR or of the price couple (S0,S1)(S_{0},S_{1}).

arb:={πd:π,z0 for all z𝒮}.\mathcal{I}_{\rm arb}:=\bigl{\{}\pi\in\mathbb{R}^{d}:\langle\pi,z\rangle\geq 0\text{ for all }z\in\mathcal{S}\bigr{\}}\setminus\mathcal{L}^{\perp}.

Under trading restrictions, the set of arbitrage opportunities is given by arbΘ\mathcal{I}_{\rm arb}\cap\Theta and consists of all allowed strategies π\pi that generate a non-negative and non-null return (see also [KS09, Definition 3.5]). We say that no classical arbitrage holds if arbΘ=\mathcal{I}_{\rm arb}\cap\Theta=\emptyset.

We now recall a second and stronger notion of arbitrage (see [Kar10, Definition 1]). To this effect, we define as follows the super-hedging value v(ξ)v(\xi) of a non-negative random variable ξ\xi:

(2.1) v(ξ):=inf{v>0:πΘ such that v(1+π,R)ξ a.s.}.v(\xi):=\inf\bigl{\{}v>0:\exists\;\pi\in\Theta\text{ such that }v(1+\langle\pi,R\rangle)\geq\xi\text{ a.s.}\bigr{\}}.

In the next definition, we denote by L+0L^{0}_{+} the family of non-negative random variables on (Ω,)(\Omega,\mathcal{F}).

Definition 2.1.

A random variable ξL+0\xi\in L^{0}_{+} with P(ξ>0)>0P(\xi>0)>0 is an arbitrage of the first kind if v(ξ)=0v(\xi)=0. We say that no arbitrage of the first kind (NA1) holds if v(ξ)=0v(\xi)=0 implies ξ=0\xi=0 a.s.

An arbitrage of the first kind consists of a non-negative non-null payoff that can be super-replicated starting from an arbitrarily small initial capital. Observe that NA1 is weaker than no classical arbitrage, as will be explicitly illustrated by the examples considered in Section 3. The next proposition provides three equivalent formulations of the NA1 condition.

Proposition 2.2.

The following are equivalent:

  1. (i)

    the NA1 condition holds;

  2. (ii)

    arbΘ^=\mathcal{I}_{\rm arb}\cap\widehat{\Theta}=\emptyset;

  3. (iii)

    Θ^=\widehat{\Theta}=\mathcal{L}^{\perp};

  4. (iv)

    the set Θ\Theta\cap\mathcal{L} is bounded (and, hence, compact).

Proof.

(i)(ii)(i)\Rightarrow(ii): by way of contradiction, suppose that NA1 holds and there exists πarbΘ^\pi\in\mathcal{I}_{\rm arb}\cap\widehat{\Theta}. Then ξ:=π,RL+0\xi:=\langle\pi,R\rangle\in L^{0}_{+} and P(ξ>0)>0P(\xi>0)>0. For every v>0v>0, it holds that π/vΘ\pi/v\in\Theta and v(1+π/v,R)>ξv(1+\langle\pi/v,R\rangle)>\xi a.s. This implies that v(ξ)=0v(\xi)=0, yielding a contradiction to NA1.
(ii)(iii)(ii)\Rightarrow(iii): we already know that Θ^\mathcal{L}^{\perp}\subseteq\widehat{\Theta}. Conversely, since Θ^a>0aΘadm\widehat{\Theta}\subseteq\bigcap_{a>0}a\Theta_{\rm adm}, every element πΘ^\pi\in\widehat{\Theta} satisfies π,R0\langle\pi,R\rangle\geq 0 a.s. Condition (ii) then implies π,R=0\langle\pi,R\rangle=0 a.s., so that π\pi\in\mathcal{L}^{\perp}.
(iii)(iv)(iii)\Rightarrow(iv): the set Θ\Theta\cap\mathcal{L} is non-empty, closed and convex. Hence, by [Roc70, Theorem 8.4], Θ\Theta\cap\mathcal{L} is bounded if and only if its recession cone Θ^\widehat{\Theta\cap\mathcal{L}} consists of the zero vector alone. Since Θ^=Θ^\widehat{\Theta\cap\mathcal{L}}=\widehat{\Theta}\cap\mathcal{L}, condition (iii)(iii) implies that Θ^={0}\widehat{\Theta\cap\mathcal{L}}=\{0\}, thus establishing property (iv).
(iv)(i)(iv)\Rightarrow(i): by way of contradiction, let ξL+0\xi\in L^{0}_{+} with P(ξ>0)>0P(\xi>0)>0 and suppose that, for all nn\in\mathbb{N}, there exists πnΘ\pi^{n}\in\Theta such that n1(1+πn,R)ξn^{-1}(1+\langle\pi^{n},R\rangle)\geq\xi a.s. In this case, it holds that 1+p(πn),Rnξ1+\langle{\rm p}_{\mathcal{L}}(\pi^{n}),R\rangle\geq n\xi a.s., for all nn\in\mathbb{N}. Since P(ξ>0)>0P(\xi>0)>0 and p(πn)Θ{\rm p}_{\mathcal{L}}(\pi^{n})\in\Theta\cap\mathcal{L}, for every nn\in\mathbb{N}, this contradicts the boundedness of the set Θ\Theta\cap\mathcal{L}. ∎

The properties stated in Proposition 2.2 admit natural and direct interpretations, which can be formulated as follows:

  1. (ii)

    there do not exist arbitrage opportunities that can be arbitrarily scaled;

  2. (iii)

    all allowed strategies that can be arbitrarily scaled reduce to trivial strategies;

  3. (iv)

    all allowed strategies not containing degeneracies are bounded.

As shown in Sections 2.3 and 2.4 below, the compactness property (iv) is fundamental, since it allows solving optimal portfolio and hedging problems under NA1, even when no classical arbitrage fails to hold. The equivalence (i)(iv)(i)\Leftrightarrow(iv) is therefore the most important novel insight provided by Proposition 2.2. The condition arbΘ^=\mathcal{I}_{\rm arb}\cap\widehat{\Theta}=\emptyset appears in [KK07, KS09] under the name no unbounded increasing profit (NUIP), where the unboundedness refers to the fact that the arbitrage profit generated by an element of arbΘ^\mathcal{I}_{\rm arb}\cap\widehat{\Theta} can be scaled to arbitrarily large values.

Remark 2.3.

Under conic trading restrictions, no classical arbitrage holds if and only if there are no arbitrages of the first kind. This simply follows from the observation that, if Θc\Theta_{\rm c} is a cone, then arbΘ^=arbΘ\mathcal{I}_{\rm arb}\cap\widehat{\Theta}=\mathcal{I}_{\rm arb}\cap\Theta. This implies that the two arbitrage concepts differ only in the presence of additional restrictions beyond conic (and, in particular, natural) constraints.

Remark 2.4 (On relative arbitrage).

The arbitrage concepts introduced so far have been implicitly defined with respect to the riskless asset with constant price equal to one. More generally, in the spirit of [FK09, Definition 6.1], a strategy πΘ\pi\in\Theta is said to be an arbitrage opportunity relative to θΘ\theta\in\Theta if P(V1πV1θ)=1P(V^{\pi}_{1}\geq V^{\theta}_{1})=1 and P(V1π>V1θ)>0P(V^{\pi}_{1}>V^{\theta}_{1})>0 or, equivalently, if πθarb(Θθ)\pi-\theta\in\mathcal{I}_{\rm arb}\cap(\Theta-\theta). If θΘ^c\theta\in\widehat{\Theta}_{\rm c}, then arb(Θθ)=\mathcal{I}_{\rm arb}\cap(\Theta-\theta)=\emptyset implies no classical arbitrage (i.e., arbΘ=\mathcal{I}_{\rm arb}\cap\Theta=\emptyset). Conversely, if θΘ^c-\theta\in\widehat{\Theta}_{\rm c}, then arbΘ=\mathcal{I}_{\rm arb}\cap\Theta=\emptyset implies arb(Θθ)=\mathcal{I}_{\rm arb}\cap(\Theta-\theta)=\emptyset. It follows that, for every θΘ^c(Θ^c)\theta\in\widehat{\Theta}_{\rm c}\cap(-\widehat{\Theta}_{\rm c}), no classical arbitrage coincides with absence of arbitrage opportunities relative to θ\theta.333The condition θΘ^c(Θ^c)\theta\in\widehat{\Theta}_{\rm c}\cap(-\widehat{\Theta}_{\rm c}) amounts to saying that arbitrary long and short positions in the portfolio θ\theta are not precluded by the trading restrictions represented by Θc\Theta_{\rm c}. This condition is conceptually equivalent to the requirement appearing in the definition of numéraire adopted in [KST16] (see Definition 10 therein). However, there is no general implication between the two conditions arbΘ=\mathcal{I}_{\rm arb}\cap\Theta=\emptyset and arb(Θθ)=\mathcal{I}_{\rm arb}\cap(\Theta-\theta)=\emptyset. Observe that, unlike arbitrage opportunities, the notion of arbitrage of the first kind is universal, in the sense that it does not depend on a reference strategy θ\theta (see Definition 2.1). The relation between NA1 and relative arbitrage is further discussed in Remark 2.11.

2.3. Market viability and fundamental theorems

The economic relevance of the NA1 condition is explained by its equivalence with the solvability of optimal portfolio problems, as shown in the next theorem. We denote by 𝒰\mathcal{U} the set of all random utility functions, consisting of all functions U:Ω×+{}U:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{R}\cup\{-\infty\} such that U(,x)U(\cdot,x) is \mathcal{F}-measurable and bounded from below, for every x>0x>0, and U(ω,)U(\omega,\cdot) is continuous, strictly increasing and concave, for a.e. ωΩ\omega\in\Omega.444For simplicity of notation, we shall omit to denote explicitly the dependence of UU on ω\omega in the following. Besides allowing for the possibility of random endowments or state-dependent preferences, the extension to random utility functions will be needed in the proof of Theorem 2.9 as well as for the solution of certain hedging and valuation problems (see the last part of Section 2.4).

Theorem 2.5.

The following are equivalent:

  1. (i)

    the NA1 condition holds;

  2. (ii)

    for every U𝒰U\in\mathcal{U} such that supπΘ𝔼[U+(V1π)]<+\sup_{\pi\in\Theta}\mathbb{E}[U^{+}(V^{\pi}_{1})]<+\infty, there exists an allowed strategy πΘ\pi^{*}\in\Theta\cap\mathcal{L} such that

    (2.2) 𝔼[U(V1π)]=supπΘ𝔼[U(V1π)].\mathbb{E}\bigl{[}U(V^{\pi^{*}}_{1})\bigr{]}=\underset{\pi\in\Theta}{\sup}\,\mathbb{E}\bigl{[}U(V^{\pi}_{1})\bigr{]}.
Proof.

(i)(ii)(i)\Rightarrow(ii): note first that (2.2) can be equivalently stated by maximizing over Θ\Theta\cap\mathcal{L}, since for every πΘ\pi\in\Theta it holds that π,R=p(π),R\langle\pi,R\rangle=\langle{\rm p}_{\mathcal{L}}(\pi),R\rangle a.s. By Proposition 2.2, NA1 implies that Θ^=Θ^={0}\widehat{\Theta\cap\mathcal{L}}=\widehat{\Theta}\cap\mathcal{L}=\{0\}. Hence, in view of [Roc70, Theorem 27.3], it suffices to show that the proper concave function u:Θu:\Theta\cap\mathcal{L}\rightarrow\mathbb{R} defined by u:Θπu(π):=𝔼[U(1+π,R)]u:\Theta\cap\mathcal{L}\ni\pi\mapsto u(\pi):=\mathbb{E}[U(1+\langle\pi,R\rangle)] is upper semi-continuous. To this effect, we adapt some of the arguments of [RS06, Lemma 2.3] (see also [Nut16, Lemma 2.8]). Since the set Θ\Theta\cap\mathcal{L} is bounded under NA1 (see Proposition 2.2), there exists a bounded polyhedral set 𝒫span(Θ)\mathcal{P}\subset{\rm span}(\Theta\cap\mathcal{L}) such that Θ𝒫\Theta\cap\mathcal{L}\subseteq\mathcal{P} (see, e.g., [Roc70, Theorem 20.4]). Denote by {p1,,pN}\{p_{1},\ldots,p_{N}\} the set of extreme points of 𝒫\mathcal{P}. Since a linear function defined on a polyhedral set attains its maximum on the set of extreme points, it holds that

π,Rmaxj=1,,Npj,R, for all πΘ.\langle\pi,R\rangle\leq\max_{j=1,\ldots,N}\langle p_{j},R\rangle,\qquad\text{ for all }\pi\in\Theta\cap\mathcal{L}.

By monotonicity of UU, this implies that

U+(1+π,R)j=1NU+(1+pj,R)=:ζ, for all πΘ.U^{+}(1+\langle\pi,R\rangle)\leq\sum_{j=1}^{N}U^{+}(1+\langle p_{j},R\rangle)=:\zeta,\qquad\text{ for all }\pi\in\Theta\cap\mathcal{L}.

We proceed to show that 𝔼[ζ]<+\mathbb{E}[\zeta]<+\infty. Since U(1)U(1) is bounded from below, we can assume without loss of generality that U(1)0U(1)\geq 0. We recall from [RS06, Lemma 2.2] the inequality

(2.3) U+(λx)2λ(U+(x)+U(2)), for all x>0 and λ1.U^{+}(\lambda x)\leq 2\lambda\bigl{(}U^{+}(x)+U(2)\bigr{)},\qquad\text{ for all $x>0$ and $\lambda\geq 1$.}

Let ϕ\phi be an element of the relative interior of Θ\Theta\cap\mathcal{L} and εj(0,1]\varepsilon_{j}\in(0,1] such that ϕ+εj(pjϕ)Θ\phi+\varepsilon_{j}(p_{j}-\phi)\in\Theta\cap\mathcal{L}, for all j=1,,Nj=1,\ldots,N. By inequality (2.3) and monotonicity of UU, together with the fact that ϕΘΘadm\phi\in\Theta\cap\mathcal{L}\subseteq\Theta_{\rm adm}, we obtain

(2.4) U+(1+pj,R)\displaystyle U^{+}\bigl{(}1+\langle p_{j},R\rangle\bigr{)} =U+(1+ϕ,R+pjϕ,R)\displaystyle=U^{+}\bigl{(}1+\langle\phi,R\rangle+\langle p_{j}-\phi,R\rangle\bigr{)}
2εj(U+(εj(1+ϕ,R)+εjpjϕ,R)+U(2))\displaystyle\leq\frac{2}{\varepsilon_{j}}\left(U^{+}\bigl{(}\varepsilon_{j}(1+\langle\phi,R\rangle)+\varepsilon_{j}\langle p_{j}-\phi,R\rangle\bigr{)}+U(2)\right)
2εj(U+(1+ϕ+εj(pjϕ),R)+U(2)).\displaystyle\leq\frac{2}{\varepsilon_{j}}\left(U^{+}\bigl{(}1+\langle\phi+\varepsilon_{j}(p_{j}-\phi),R\rangle\bigr{)}+U(2)\right).

Due to the assumption that supπΘ𝔼[U+(V1π)]<+\sup_{\pi\in\Theta}\mathbb{E}[U^{+}(V^{\pi}_{1})]<+\infty, the first term on the last line of (2.4) is integrable, for each j=1,,Nj=1,\ldots,N. The same assumption implies that 𝔼[U(1)]<+\mathbb{E}[U(1)]<+\infty, from which 𝔼[U(2)]<+\mathbb{E}[U(2)]<+\infty follows by concavity of UU. This proves that the random variable ζ\zeta is integrable. Let now (πn)n(\pi^{n})_{n\in\mathbb{N}} be a sequence in Θ\Theta\cap\mathcal{L} converging to some element π0Θ\pi^{0}\in\Theta\cap\mathcal{L}. An application of Fatou’s lemma, together with the continuity of UU, yields that

lim supn+u(πn)𝔼[lim supn+U(1+πn,R)]=u(π0),\limsup_{n\rightarrow+\infty}u(\pi^{n})\leq\mathbb{E}\bigl{[}\limsup_{n\rightarrow+\infty}U(1+\langle\pi^{n},R\rangle)\bigr{]}=u(\pi^{0}),

thus proving the upper semi-continuity of the function uu.
(ii)(i)(ii)\Rightarrow(i): by way of contradiction, let πΘ\pi^{*}\in\Theta\cap\mathcal{L} be the maximizer in (2.2) and suppose that NA1 fails to hold. By Proposition 2.2, there exists θarbΘ^\theta\in\mathcal{I}_{\rm arb}\cap\widehat{\Theta}. It holds that π+θΘ\pi^{*}+\theta\in\Theta\cap\mathcal{L} and 𝔼[U(V1π+θ)]>𝔼[U(V1π)]\mathbb{E}[U(V^{\pi^{*}+\theta}_{1})]>\mathbb{E}[U(V^{\pi^{*}}_{1})], thus contradicting the optimality of π\pi^{*}. ∎

The above theorem asserts the equivalence between NA1 and market viability, intended as the existence of an optimal strategy for every well-posed expected utility maximization problem. In particular, the proof makes clear that one of the crucial consequences of NA1 is the compactness of the set Θ\Theta\cap\mathcal{L} of non-redundant allowed strategies (see Proposition 2.2).

Remark 2.6.

The proof of Theorem 2.5 relies on the fact that, under NA1, the set Θ\Theta\cap\mathcal{L} and the function uu have no common directions of recession. The relevance of this property in expected utility maximization problems has been first recognized in the early work [Ber74].

Remark 2.7.

In discrete-time models without trading constraints, it is well-known that market viability is equivalent to no classical arbitrage (see [RS05, RS06] and [FS16, Theorem 3.3]). In view of Remark 2.3, the same holds true in the case of conic constraints. In the case of convex non-conic constraints, Theorem 2.5 shows that market viability is equivalent to the weaker NA1 condition. From an economic standpoint, this result implies that assuming no classical arbitrage may pose unnecessary restrictions on the model. In the special case of a finite probability space, this insight already appeared in [ES01], where the authors proved the equivalence between the existence of a solution to optimal portfolio problems and the validity of a condition called by the authors no unlimited arbitrage. Translated in our context, no unlimited arbitrage corresponds to the existence of a strategy θΘ\theta\in\Theta such that there are no arbitrage opportunities relative to θ\theta, in the sense of Remark 2.4. As shown in Remark 2.11 below, this is equivalent to NA1 and, therefore, the results of [ES01] can be recovered as a special case.555The finiteness condition appearing in part (ii)(ii) of Theorem 2.5 is always satisfied on a finite probability space. In continuous-time semimartingale models, the connection between NA1 and the solvability of expected utility maximization problems is discussed in [KK07] and [CDM15] (in an Itô process setting, earlier results in this direction have been obtained in [LW00]).

The NA1 condition admits an equivalent characterization in terms of the existence of a (supermartingale) deflator or of a numéraire portfolio, defined as follows.

Definition 2.8.

A random variable ZL+0Z\in L^{0}_{+} with P(Z>0)=1P(Z>0)=1 is said to be a deflator if

(2.5) 𝔼[ZV1π]1, for all πΘ.\mathbb{E}[Z\,V^{\pi}_{1}]\leq 1,\qquad\text{ for all }\pi\in\Theta.

The set of all deflators is denoted by 𝒟\mathcal{D}.
An allowed trading strategy ρΘ\rho\in\Theta is said to be a numéraire portfolio if 1/V1ρ𝒟1/V^{\rho}_{1}\in\mathcal{D}, meaning that

(2.6) 𝔼[V1π/V1ρ]1, for all πΘ.\mathbb{E}[V^{\pi}_{1}/V^{\rho}_{1}]\leq 1,\qquad\text{ for all }\pi\in\Theta.

It is well-known (see, e.g., [Bec01]) that a numéraire portfolio is unique in the sense that if ρ1\rho^{1} and ρ2\rho^{2} satisfy (2.6), then ρ1ρ2\rho^{1}-\rho^{2}\in\mathcal{L}^{\perp}. The numéraire portfolio is therefore uniquely defined on Θ\Theta\cap\mathcal{L}. The next theorem shows that NA1 is necessary and sufficient for the existence of the numéraire portfolio. In a general semimartingale setting, the corresponding result has been proved in [KK07]. In the present context, Theorem 2.5 enables us to give a short and simple proof based on log-utility maximization, thus highlighting the central role of market viability. Besides simplifying the techniques employed in [KS09, Lemma 6.2 and Theorem 6.3], our proof can be easily generalized to the multi-period setting, as will be shown in Section 4.

Theorem 2.9.

The following are equivalent:

  1. (i)

    the NA1 condition holds;

  2. (ii)

    𝒟\mathcal{D}\neq\emptyset;

  3. (iii)

    there exists the numéraire portfolio.

Moreover, ρΘ\rho\in\Theta is the numéraire portfolio if and only if it is relatively log-optimal, in the sense that it satisfies 𝔼[log(V1π/V1ρ)]0\mathbb{E}[\log(V^{\pi}_{1}/V^{\rho}_{1})]\leq 0, for all πΘ\pi\in\Theta.

Proof.

(i)(iii)(i)\Rightarrow(iii): as a preliminary, similarly as in [Kar09, KS09], let (fn)n(f_{n})_{n} be a family of functions such that fn:d(0,1]f_{n}:\mathbb{R}^{d}\rightarrow(0,1] and 𝔼[log(1+R)fn(R)]<+\mathbb{E}[\log(1+\|R\|)f_{n}(R)]<+\infty, for each nn\in\mathbb{N}, and fn1f_{n}\nearrow 1 as n+n\rightarrow+\infty. A specific choice is for instance given by fn(x)=𝟏{x1}+𝟏{x>1}x1/nf_{n}(x)=\mathbf{1}_{\{\|x\|\leq 1\}}+\mathbf{1}_{\{\|x\|>1\}}\|x\|^{-1/n}. For each nn\in\mathbb{N}, define the function (ω,x)Un(ω,x):=log(x)fn(R(ω))(\omega,x)\mapsto U_{n}(\omega,x):=\log(x)f_{n}(R(\omega)), for (ω,x)Ω×(0,+)(\omega,x)\in\Omega\times(0,+\infty), with Un(ω,0):=limx0Un(ω,x)=U_{n}(\omega,0):=\lim_{x\downarrow 0}U_{n}(\omega,x)=-\infty. For each nn\in\mathbb{N}, it holds that Un𝒰U_{n}\in\mathcal{U} and

(2.7) 𝔼[Un+(1+π,R)]π+𝔼[log(1+R)fn(R)]<+, for all πΘ.\mathbb{E}\bigl{[}U^{+}_{n}(1+\langle\pi,R\rangle)\bigr{]}\leq\|\pi\|+\mathbb{E}\bigl{[}\log(1+\|R\|)f_{n}(R)\bigr{]}<+\infty,\qquad\text{ for all }\pi\in\Theta.

If NA1 holds, Proposition 2.2 implies that Θ\Theta\cap\mathcal{L} is bounded and, therefore, it holds that supπΘ𝔼[Un+(1+π,R)]<+\sup_{\pi\in\Theta}\mathbb{E}[U^{+}_{n}(1+\langle\pi,R\rangle)]<+\infty. For each nn\in\mathbb{N}, Theorem 2.5 gives then the existence of an element ρnΘ\rho^{n}\in\Theta\cap\mathcal{L} which is the maximizer in (2.2) for U=UnU=U_{n}. For an arbitrary element πΘ\pi\in\Theta and ε(0,1)\varepsilon\in(0,1), let πε:=επ+(1ε)ρnΘ\pi^{\varepsilon}:=\varepsilon\pi+(1-\varepsilon)\rho^{n}\in\Theta. The optimality of ρn\rho^{n} together with the elementary inequality log(x)(x1)/x\log(x)\geq(x-1)/x, for x>0x>0, implies that

0\displaystyle 0 1ε(𝔼[Un(1+πε,R)]𝔼[Un(1+ρn,R)])\displaystyle\geq\frac{1}{\varepsilon}\Bigl{(}\mathbb{E}\bigl{[}U_{n}(1+\langle\pi^{\varepsilon},R\rangle)\bigr{]}-\mathbb{E}\bigl{[}U_{n}(1+\langle\rho^{n},R\rangle)\bigr{]}\Bigr{)}
(2.8) =1ε𝔼[log(V1πε/V1ρn)fn(R)]𝔼[πρn,R1+ρn,R+επρn,Rfn(R)].\displaystyle=\frac{1}{\varepsilon}\mathbb{E}\bigl{[}\log(V^{\pi^{\varepsilon}}_{1}/V^{\rho^{n}}_{1})f_{n}(R)\bigr{]}\geq\mathbb{E}\left[\frac{\langle\pi-\rho^{n},R\rangle}{1+\langle\rho^{n},R\rangle+\varepsilon\langle\pi-\rho^{n},R\rangle}f_{n}(R)\right].

Noting that xy+εxxy+x/22\frac{x}{y+\varepsilon x}\geq\frac{x}{y+x/2}\geq-2, for all ε(0,1/2)\varepsilon\in(0,1/2), y>0y>0 and xyx\geq-y, we can let ε0\varepsilon\searrow 0 and apply Fatou’s lemma, thus obtaining

(2.9) 𝔼[πρn,R1+ρn,Rfn(R)]0, for all πΘ and n.\mathbb{E}\left[\frac{\langle\pi-\rho^{n},R\rangle}{1+\langle\rho^{n},R\rangle}f_{n}(R)\right]\leq 0,\qquad\text{ for all }\pi\in\Theta\text{ and }n\in\mathbb{N}.

Since Θ\Theta\cap\mathcal{L} is compact (see Proposition 2.2), we may assume that the sequence (ρn)n(\rho^{n})_{n\in\mathbb{N}} converges to some ρΘ\rho\in\Theta\cap\mathcal{L} as n+n\rightarrow+\infty. Therefore, since πρn,R/(1+ρn,R)1\langle\pi-\rho^{n},R\rangle/(1+\langle\rho^{n},R\rangle)\geq-1 a.s. and recalling that fn1f_{n}\nearrow 1 as n+n\rightarrow+\infty, another application of Fatou’s lemma gives that

𝔼[πρ,R1+ρ,R]0, for all πΘ.\mathbb{E}\left[\frac{\langle\pi-\rho,R\rangle}{1+\langle\rho,R\rangle}\right]\leq 0,\qquad\text{ for all }\pi\in\Theta.

Equivalently, it holds that 𝔼[V1π/V1ρ]1\mathbb{E}[V^{\pi}_{1}/V^{\rho}_{1}]\leq 1, for all πΘ\pi\in\Theta. In view of Definition 2.8, we have thus shown that NA1 implies the existence of the numéraire portfolio.
(iii)(ii)(iii)\Rightarrow(ii): this implication is immediate by Definition 2.8.
(ii)(i)(ii)\Rightarrow(i): let Z𝒟Z\in\mathcal{D} and consider a random variable ξL+0\xi\in L^{0}_{+} with P(ξ>0)>0P(\xi>0)>0 such that, for every nn\in\mathbb{N}, there exists πnΘ\pi^{n}\in\Theta such that V1πn(1/n)ξV^{\pi^{n}}_{1}(1/n)\geq\xi a.s. Definition (2.8) implies that

𝔼[Zξ]𝔼[ZV1πn(1/n)]=1n𝔼[ZV1πn]1n, for all n.\mathbb{E}[Z\,\xi]\leq\mathbb{E}\bigl{[}Z\,V^{\pi^{n}}_{1}(1/n)\bigr{]}=\frac{1}{n}\mathbb{E}\bigl{[}Z\,V^{\pi^{n}}_{1}\bigr{]}\leq\frac{1}{n},\qquad\text{ for all }n\in\mathbb{N}.

Since Z>0Z>0 a.s., letting n+n\rightarrow+\infty yields that ξ=0\xi=0 a.s., thus proving the validity of NA1.
It remains to prove the last assertion of the theorem. If ρΘ\rho\in\Theta satisfies (2.6), then its relative log-optimality is a direct consequence of Jensen’s inequality. Conversely, if ρΘ\rho\in\Theta is relatively log-optimal, then (2.6) follows by the same arguments used in (2.8)-(2.9). ∎

Remark 2.10.

If there exists a log-optimal portfolio, i.e., an allowed strategy ρΘ\rho\in\Theta satisfying 𝔼[log(V1π)]𝔼[log(V1ρ)]<+\mathbb{E}[\log(V^{\pi}_{1})]\leq\mathbb{E}[\log(V^{\rho}_{1})]<+\infty, for all πΘ\pi\in\Theta, then ρ\rho is also relatively log-optimal and, therefore, coincides with the numéraire portfolio. The numéraire property of the log-optimal portfolio can also be directly deduced from the proof of Theorem 2.9. In applications, computing the log-optimal portfolio typically represents a simple way to determine the numéraire portfolio (see for instance Examples 2.14 and 3.11).

Remark 2.11.

NA1 is equivalent to the existence of a strategy θΘ\theta\in\Theta with V1θ>0V^{\theta}_{1}>0 a.s. such that there are no arbitrage opportunities relative to θ\theta, in the sense of Remark 2.4. Indeed, suppose there exists θΘ\theta\in\Theta with V1θ>0V^{\theta}_{1}>0 a.s. and let πΘ^\pi\in\widehat{\Theta}. Then π+θ\pi+\theta is an arbitrage opportunity relative to θ\theta if and only if πarb\pi\in\mathcal{I}_{\rm arb}. Conversely, if NA1 holds, then there do not exist arbitrage opportunities relative to the numéraire portfolio ρ\rho, as a consequence of (2.6). However, absence of arbitrage opportunities relative to some strategy θΘ\theta\in\Theta with V1θ>0V^{\theta}_{1}>0 a.s. does not suffice to conclude that θ\theta is the numéraire portfolio (see Example 3.10 for an explicit counterexample).

Theorems 2.5 and 2.9 represent the central results of arbitrage theory based on NA1. For completeness, we now state the fundamental theorem of asset pricing based on no classical arbitrage, in the general version of [Rok05, Theorem 4] specialised for a one-period setting. We give a simple proof inspired by [KaS09, Proposition 2.1.5] and [Kar09, Theorem 3.7], which in turn follow an original idea of [Rog94]. Similarly to Theorem 2.9, the proof is based on utility maximization arguments. For a set AdA\subseteq\mathbb{R}^{d}, we denote by coneA{\rm cone}\,A its conic hull.

Theorem 2.12.

Suppose that the set coneΘ{\rm cone}\,\Theta is closed. Then no classical arbitrage holds if and only if there exists a probability measure QPQ\sim P such that 𝔼Q[V1π]1\mathbb{E}^{Q}[V^{\pi}_{1}]\leq 1, for all πconeΘ\pi\in{\rm cone}\,\Theta.

Proof.

Observe first that arbΘ=\mathcal{I}_{\rm arb}\cap\Theta=\emptyset if and only if arb(coneΘ)=\mathcal{I}_{\rm arb}\cap({\rm cone}\,\Theta)=\emptyset. In turn, this implies that no classical arbitrage holds if and only if arbC=\mathcal{I}_{\rm arb}\cap C=\emptyset, where C:=(coneΘ)C:=({\rm cone}\,\Theta)\cap\mathcal{L}. Define the proper convex function f:Cπf(π):=𝔼[exp(1π,R)]f:C\ni\pi\mapsto f(\pi):=\mathbb{E}^{\prime}[\exp(-1-\langle\pi,R\rangle)], where 𝔼\mathbb{E}^{\prime} denotes expectation with respect to the probability measure PP^{\prime} defined by dP/dP=eR2/𝔼[eR2]\mathrm{d}P^{\prime}/\mathrm{d}P=e^{-\|R\|^{2}}/\mathbb{E}[e^{-\|R\|^{2}}]. By Fatou’s lemma, the function ff is lower semi-continuous. Since CC is closed by assumption, [Roc70, Theorem 27.3] implies that the function ff admits a minimizer πC\pi^{*}\in C if it has no directions of recession in common with the cone CC. By [Roc70, Theorem 8.5], this amounts to verifying that

(2.10) f^(π):=limγ+f(γπ)γ>0, for all πC{0}.\hat{f}(\pi):=\lim_{\gamma\rightarrow+\infty}\frac{f(\gamma\pi)}{\gamma}>0,\qquad\text{ for all }\pi\in C\setminus\{0\}.

We now show that (2.10) is always satisfied under no classical arbitrage. Arguing by contradiction, let πC{0}\pi\in C\setminus\{0\} such that f^(π)0\hat{f}(\pi)\leq 0. In this case, by Fatou’s lemma, it holds that

0\displaystyle 0 f^(π)𝔼[lim infγ+e1γπ,Rγ]𝔼[lim infγ+e1γπ,Rγ𝟏{π,R<0}].\displaystyle\geq\hat{f}(\pi)\geq\mathbb{E}^{\prime}\left[\liminf_{\gamma\rightarrow+\infty}\frac{e^{-1-\gamma\langle\pi,R\rangle}}{\gamma}\right]\geq\mathbb{E}^{\prime}\left[\liminf_{\gamma\rightarrow+\infty}\frac{e^{-1-\gamma\langle\pi,R\rangle}}{\gamma}\mathbf{1}_{\{\langle\pi,R\rangle<0\}}\right].

This implies that necessarily π,R0\langle\pi,R\rangle\geq 0 a.s. Since π\pi\in\mathcal{L}, this contradicts no classical arbitrage. [Roc70, Theorem 27.3] then yields the existence of an element πC\pi^{*}\in C such that f(π)f(π)f(\pi^{*})\leq f(\pi), for all πC\pi\in C. The definition of PP^{\prime} implies that differentiation and integration can be interchanged, so that the gradient of the function ff at π\pi^{*} is given by f(π)=𝔼[exp(1π,R)R]\nabla f(\pi^{*})=-\mathbb{E}^{\prime}[\exp(-1-\langle\pi^{*},R\rangle)R]. Therefore, since CC is a cone and ff is finite on CC, [Roc70, Theorem 27.4] implies that

0π,f(π)=𝔼[e1π,Rπ,R].0\geq\bigl{\langle}\pi,-\nabla f(\pi^{*})\bigr{\rangle}=\mathbb{E}^{\prime}\bigl{[}e^{-1-\langle\pi^{*},R\rangle}\langle\pi,R\rangle\bigr{]}.

Setting dQ/dP=eV1πR2/𝔼[eV1πR2]\mathrm{d}Q/\mathrm{d}P=e^{-V^{\pi^{*}}_{1}-\|R\|^{2}}/\mathbb{E}[e^{-V^{\pi^{*}}_{1}-\|R\|^{2}}] yields a probability measure QPQ\sim P such that 𝔼Q[V1π]1\mathbb{E}^{Q}[V^{\pi}_{1}]\leq 1, for all πC\pi\in C and, hence, for all πconeΘ\pi\in{\rm cone}\,\Theta.

Conversely, suppose there exists a probability measure QPQ\sim P such that 𝔼Q[V1π]1\mathbb{E}^{Q}[V^{\pi}_{1}]\leq 1, for all πconeΘ\pi\in{\rm cone}\,\Theta. Then, for every πΘ\pi\in\Theta, it holds that 𝔼Q[π,R]0\mathbb{E}^{Q}[\langle\pi,R\rangle]\leq 0. If πarbΘ\pi\in\mathcal{I}_{\rm arb}\cap\Theta, this implies that π,R0\langle\pi,R\rangle\leq 0 QQ-a.s. However, since QPQ\sim P, this contradicts the fact that πarb\pi\in\mathcal{I}_{\rm arb}. ∎

Remark 2.13.

Theorem 2.12 does not hold without the assumption of closedness of coneΘ{\rm cone}\,\Theta.666The same assumption is required in the fundamental theorem of asset pricing in the formulation of [CPT01]. [Rok05, Theorem 4] requires the closedness of p(coneΘ){\rm p}_{\mathcal{L}}({\rm cone}\,\Theta), the set of all vectors in d\mathbb{R}^{d} that are projections onto \mathcal{L} of elements of coneΘ{\rm cone}\,\Theta. In our setting, since Θ^\mathcal{L}^{\perp}\subseteq\widehat{\Theta}, it holds that p(coneΘ)=(coneΘ){\rm p}_{\mathcal{L}}({\rm cone}\,\Theta)=({\rm cone}\,\Theta)\cap\mathcal{L}. This implies that p(coneΘ){\rm p}_{\mathcal{L}}({\rm cone}\,\Theta) is closed if and only if coneΘ{\rm cone}\,\Theta is closed. Indeed, one can construct a counterexample along the lines of [Rok05, Example 1] where no classical arbitrage holds but there does not exist a probability measure QPQ\sim P such that 𝔼Q[V1π]1\mathbb{E}^{Q}[V^{\pi}_{1}]\leq 1, for all πconeΘ\pi\in{\rm cone}\,\Theta. Observe that, in comparison to no classical arbitrage, NA1 has the additional advantage of not requiring any extra technical condition on the model.

The probability measure QQ appearing in Theorem 2.12 represents an equivalent supermartingale measure (ESMM). If NA1 holds and the numéraire portfolio ρ\rho satisfies 𝔼[1/V1ρ]=1\mathbb{E}[1/V^{\rho}_{1}]=1, then an ESMM QQ can be defined by setting dQ/dP=1/V1ρ\mathrm{d}Q/\mathrm{d}P=1/V^{\rho}_{1}. However, this is not always possible, even when coneΘ{\rm cone}\,\Theta is closed and no classical arbitrage holds, as the following simple example illustrates (see also [Bec01, Example 6] for a related example in an unconstrained setting).

Example 2.14.

Let d=1d=1 and suppose that R=eY1R=e^{Y}-1, with Y𝒩(0,1)Y\sim\mathcal{N}(0,1). In this case, it holds that 𝒮=[1,+)\mathcal{S}=[-1,+\infty) and Θadm=[0,1]\Theta_{\rm adm}=[0,1] (i.e., short-selling and borrowing from the riskless asset are prohibited). Suppose that Θc=[0,c]\Theta_{\rm c}=[0,c], for some c[0,1]c\in[0,1], so that Θ=[0,c]\Theta=[0,c]. Clearly, no classical arbitrage holds and, therefore, there exists an ESMM QQ. For instance, it can be easily checked that dQ/dP=exp(αYα2/2)\mathrm{d}Q/\mathrm{d}P=\exp(\alpha Y-\alpha^{2}/2) defines an ESMM, for any α1/2\alpha\leq-1/2. However, if c<1/2c<1/2, the numéraire portfolio ρ\rho cannot be used to construct an ESMM, since 𝔼[1/V1ρ]<1\mathbb{E}[1/V^{\rho}_{1}]<1. Indeed, it can be easily checked that the function h:[0,1]h:[0,1]\rightarrow\mathbb{R} defined by h(π):=𝔼[log(V1π)]h(\pi):=\mathbb{E}[\log(V^{\pi}_{1})] is finite-valued, strictly concave and achieves its maximum at 1/21/2, so that h(π)>0h^{\prime}(\pi)>0 for all π<1/2\pi<1/2. Therefore, if c<1/2c<1/2, the log-optimal portfolio and, therefore, the numéraire portfolio ρ\rho (see Remark 2.10) are given by ρ=c\rho=c and it holds that h(ρ)>0h^{\prime}(\rho)>0 or, equivalently, 𝔼[1/V1ρ]<1\mathbb{E}[1/V^{\rho}_{1}]<1.

2.4. Hedging and valuation of contingent claims

The pricing of contingent claims is traditionally based on the paradigm of no classical arbitrage. In this section, we show that the weaker NA1 condition suffices to develop a general and effective theory for the hedging and valuation of contingent claims in the presence of convex constraints. We first prove the fundamental super-hedging duality. Recall that for a random variable ξL+0\xi\in L^{0}_{+} (contingent claim) its super-hedging value v(ξ)v(\xi) is defined as in (2.1), with the usual convention inf=+\inf\emptyset=+\infty.

Theorem 2.15.

Suppose that NA1 holds and let ξL+0\xi\in L^{0}_{+}. Then

(2.11) v(ξ)=supZ𝒟𝔼[Zξ].v(\xi)=\sup_{Z\in\mathcal{D}}\mathbb{E}[Z\xi].

Moreover, there exists a pair (v,π)+×Θ(v,\pi)\in\mathbb{R}_{+}\times\Theta such that ξ=V1π(v)\xi=V^{\pi}_{1}(v) a.s. and 𝔼[Zξ]=v\mathbb{E}[Z\xi]=v, for some Z𝒟Z\in\mathcal{D}, if and only if there exists an element Z𝒟Z^{*}\in\mathcal{D} such that 𝔼[Zξ]=supZ𝒟𝔼[Zξ]<+\mathbb{E}[Z^{*}\xi]=\sup_{Z\in\mathcal{D}}\mathbb{E}[Z\xi]<+\infty.

Proof.

Let 𝒱(ξ):={v>0:πΘ such that vV1πξ a.s.}\mathcal{V}(\xi):=\{v>0:\exists\;\pi\in\Theta\text{ such that }vV_{1}^{\pi}\geq\xi\text{ a.s.}\} and 𝒞:={V1π:πΘ}L+0\mathcal{C}:=\{V^{\pi}_{1}:\pi\in\Theta\cap\mathcal{L}\}-L^{0}_{+}. If v𝒱(ξ)v\in\mathcal{V}(\xi), there exists πΘ\pi\in\Theta such that vV1πξvV^{\pi}_{1}\geq\xi a.s. Then, for every Z𝒟Z\in\mathcal{D} it holds that

𝔼[Zξ]v𝔼[ZV1π]v.\mathbb{E}[Z\xi]\leq v\,\mathbb{E}[ZV^{\pi}_{1}]\leq v.

By taking the supremum over all Z𝒟Z\in\mathcal{D} and the infimum over all v𝒱(ξ)v\in\mathcal{V}(\xi), we obtain that v(ξ)supZ𝒟𝔼[Zξ]=:vv(\xi)\geq\sup_{Z\in\mathcal{D}}\mathbb{E}[Z\xi]=:v^{*}. The converse inequality is trivial if v=+v^{*}=+\infty. Assuming therefore that 0<v<+0<v^{*}<+\infty, we will show that v(ξ)>vv(\xi)>v^{*} cannot hold. Indeed, if v(ξ)>vv(\xi)>v^{*}, then ξv𝒞\xi\notin v^{*}\mathcal{C}. Let ρ\rho be the numéraire portfolio (which exists by Theorem 2.9). Being closed in L0L^{0} (see Lemma 2.16 below) and bounded in L1L^{1}, the set v𝒞/V1ρv^{*}\mathcal{C}/V^{\rho}_{1} is closed in L1L^{1}. Therefore, by the Hahn-Banach theorem (see, e.g., [FS16, Theorem A.58]), there exists a bounded random variable Z¯\bar{Z} such that

(2.12) +>1v𝔼[Z¯ξV1ρ]>supX𝒞𝔼[Z¯XV1ρ]=:s.+\infty>\frac{1}{v^{*}}\mathbb{E}\left[\bar{Z}\frac{\xi}{V^{\rho}_{1}}\right]>\sup_{X\in\mathcal{C}}\mathbb{E}\left[\bar{Z}\frac{X}{V^{\rho}_{1}}\right]=:s.

Since n𝟏{Z¯<0}𝒞-n\mathbf{1}_{\{\bar{Z}<0\}}\in\mathcal{C}, for all n0n\geq 0, inequality (2.12) implies that Z¯0\bar{Z}\geq 0 a.s. and P(Z¯>0)>0P(\bar{Z}>0)>0. Moreover, since 1𝒞1\in\mathcal{C}, it holds that s>0s>0. For ε(0,1)\varepsilon\in(0,1), we define

(2.13) Zε:=(ε+(1ε)Z¯s)1V1ρ.Z^{\varepsilon}:=\left(\varepsilon+(1-\varepsilon)\frac{\bar{Z}}{s}\right)\frac{1}{V^{\rho}_{1}}.

It holds that P(Zε>0)=1P(Z^{\varepsilon}>0)=1 and, for every πΘ\pi\in\Theta,

𝔼[ZεV1π]=ε𝔼[V1πV1ρ]+1εs𝔼[Z¯V1πV1ρ]1,\mathbb{E}[Z^{\varepsilon}V^{\pi}_{1}]=\varepsilon\mathbb{E}\left[\frac{V^{\pi}_{1}}{V^{\rho}_{1}}\right]+\frac{1-\varepsilon}{s}\mathbb{E}\left[\bar{Z}\frac{V^{\pi}_{1}}{V^{\rho}_{1}}\right]\leq 1,

thus showing that Zε𝒟Z^{\varepsilon}\in\mathcal{D}, for all ε(0,1)\varepsilon\in(0,1). Moreover, for a sufficiently small ε\varepsilon, (2.12) together with (2.13) implies that 𝔼[Zεξ]>v=supZ𝒟𝔼[Zξ]\mathbb{E}[Z^{\varepsilon}\xi]>v^{*}=\sup_{Z\in\mathcal{D}}\mathbb{E}[Z\xi], which is absurd. Therefore, we must have ξv𝒞\xi\in v^{*}\mathcal{C} , thus proving that v(ξ)v=supZ𝒟𝔼[Zξ]v(\xi)\leq v^{*}=\sup_{Z\in\mathcal{D}}\mathbb{E}[Z\xi].
To prove the last assertion of the theorem, observe that the first part of the proof yields that vV1πξv^{*}V^{\pi}_{1}\geq\xi a.s., for some πΘ\pi\in\Theta. If there exists Z𝒟Z^{*}\in\mathcal{D} such that v=𝔼[Zξ]v^{*}=\mathbb{E}[Z^{*}\xi], then we have that

v=𝔼[Zξ]v𝔼[ZV1π]v.v^{*}=\mathbb{E}[Z^{*}\xi]\leq v^{*}\mathbb{E}[Z^{*}V^{\pi}_{1}]\leq v^{*}.

Since Z>0Z^{*}>0 a.s., this implies that ξ=V1π(v)\xi=V^{\pi}_{1}(v^{*}) a.s. Conversely, if ξ=V1π(v)\xi=V^{\pi}_{1}(v) a.s. for some (v,π)+×Θ(v,\pi)\in\mathbb{R}_{+}\times\Theta with v=𝔼[Zξ]v=\mathbb{E}[Z^{*}\xi], for some Z𝒟Z^{*}\in\mathcal{D}, then (2.5) implies that 𝔼[Zξ]=supZ𝒟𝔼[Zξ]\mathbb{E}[Z^{*}\xi]=\sup_{Z\in\mathcal{D}}\mathbb{E}[Z\xi]. ∎

Lemma 2.16.

If NA1 holds, then the set 𝒞:={V1π:πΘ}L+0\mathcal{C}:=\{V^{\pi}_{1}:\pi\in\Theta\cap\mathcal{L}\}-L^{0}_{+} is closed in L0L^{0}.

Proof.

Let (Xn)n𝒞(X_{n})_{n\in\mathbb{N}}\subseteq\mathcal{C} be a sequence converging in L0L^{0} to a random variable XX as n+n\rightarrow+\infty. For each nn\in\mathbb{N}, it holds that Xn=1+πn,RAnX_{n}=1+\langle\pi_{n},R\rangle-A_{n}, for (πn,An)(Θ)×L+0(\pi_{n},A_{n})\in(\Theta\cap\mathcal{L})\times L^{0}_{+}. By Proposition 2.2, NA1 implies that the set Θ\Theta\cap\mathcal{L} is compact and, therefore, there exists a subsequence (πnm)m(\pi_{n_{m}})_{m\in\mathbb{N}} converging to an element πΘ\pi\in\Theta\cap\mathcal{L}. In turn, this implies that the sequence (Anm)m(A_{n_{m}})_{m\in\mathbb{N}} converges in probability to a random variable AL+0A\in L^{0}_{+}, thus proving the closedness of 𝒞\mathcal{C} in L0L^{0}. ∎

Whenever the quantity supZ𝒟𝔼[Zξ]\sup_{Z\in\mathcal{D}}\mathbb{E}[Z\xi] is finite, it provides the super-hedging value of ξ\xi. In a general semimartingale setting, the duality relation (2.11) has been stated in [KK07, Section 4.7]. We contribute by providing a transparent and self-contained proof in a one-period setting. In addition, Theorem 2.15 provides a necessary and sufficient condition for the attainability of a contingent claim ξ\xi. When perfect hedging is not possible, one may resort to several alternative hedging approaches, which are all feasible under NA1 even if no classical arbitrage fails to hold. A first possibility is represented by hedging with minimal shortfall risk, corresponding to

(2.14) 𝔼[(ξvV1π)]=min! over all (v,π)(0,v0]×Θ,\mathbb{E}\bigl{[}\ell(\xi-vV^{\pi}_{1})\bigr{]}=\min\,!\qquad\text{ over all }(v,\pi)\in(0,v_{0}]\times\Theta,

for some initial capital v0>0v_{0}>0, where :\ell:\mathbb{R}\rightarrow\mathbb{R} is an increasing convex loss function such that (x)=0\ell(x)=0, for all x0x\leq 0, and 𝔼[(ξ)]<+\mathbb{E}[\ell(\xi)]<+\infty (see [FS16, Section 8.2]). Problem (2.14) can be solved by first minimizing 𝔼[(ξY)]\mathbb{E}[\ell(\xi-Y)] over all random variables YL+0Y\in L^{0}_{+} such that supZ𝒟𝔼[ZY]v0\sup_{Z\in\mathcal{D}}\mathbb{E}[ZY]\leq v_{0} and then considering the pair (v(Y),π)(v(Y^{*}),\pi^{*}) which super-replicates the minimizing random variable YY^{*} (if \ell is strictly increasing on [0,+)[0,+\infty), then v(Y)=v0v(Y^{*})=v_{0}). As long as NA1 holds, the feasibility of this approach is ensured by Theorem 2.15.

An alternative way to hedge and compute the value of a contingent claim ξ\xi is provided by utility indifference valuation. For a given utility function uu and an initial capital v>0v>0, this corresponds to finding the solution p=p(ξ)p=p(\xi) to the equation

(2.15) supπΘ𝔼[u(vV1π)]=supπΘ𝔼[u((vp)V1π+ξ)].\sup_{\pi\in\Theta}\mathbb{E}\bigl{[}u(vV^{\pi}_{1})\bigr{]}=\sup_{\pi\in\Theta}\mathbb{E}\bigl{[}u((v-p)V^{\pi}_{1}+\xi)\bigr{]}.

Defining Upη(x,ω):=u((vηp)x+ηξ(ω))U_{p}^{\eta}(x,\omega):=u((v-\eta p)x+\eta\xi(\omega)), for η{0,1}\eta\in\{0,1\}, Theorem 2.5 with U=UpηU=U^{\eta}_{p} shows that NA1 is sufficient for the solvability of the two maximization problems appearing in (2.15). Whenever it exists, p(ξ)p(\xi) represents a (buyer) value for ξ\xi, while the strategy π\pi^{*} that achieves the supremum on the right-hand side of (2.15) with p=p(ξ)p=p(\xi) provides a hedging strategy for ξ\xi.

As a variant of the latter approach, one can consider marginal utility indifference valuation, in the sense of [Dav97]. This corresponds to finding the value p=p(ξ)p=p^{\prime}(\xi) which solves

limη0𝔼[Upη(V1π)]𝔼[Up0(V1π)]η=0.\lim_{\eta\downarrow 0}\frac{\mathbb{E}[U_{p}^{\eta}(V^{\pi^{*}}_{1})]-\mathbb{E}[U_{p}^{0}(V^{\pi^{*}}_{1})]}{\eta}=0.

where UηU^{\eta} is defined as above, for η[0,1]\eta\in[0,1], and πΘ\pi^{*}\in\Theta is the strategy solving problem (2.2) with U=uU=u. Similarly as in [FR13], if NA1 holds and u(x)=log(x)u(x)=\log(x), it can be shown that

(2.16) p(ξ)=𝔼[ξ/V1ρ],p^{\prime}(\xi)=\mathbb{E}[\xi/V^{\rho}_{1}],

as long as the expectation is finite, where ρ\rho denotes the numéraire portfolio (see Theorem 2.9). In the context of the Benchmark Approach (see [BP03, PH06]), formula (2.16) corresponds to the well-known real-world pricing formula, which is applicable as long as NA1 is satisfied.

3. Factor models with arbitrage under borrowing constraints

In this section, we study the arbitrage concepts discussed above in the context of a one-period factor model, under constraints on the fraction of wealth that can be borrowed/invested on the riskless asset. We start from a general model and then consider more specific cases.

3.1. A general factor model

In the setting of Section 2, we assume that asset returns are generated by the factor model

(3.1) R=QY,R=QY,

where Qd×Q\in\mathbb{R}^{d\times\ell} and Y=(Y1,,Y)Y=(Y_{1},\ldots,Y_{\ell})^{\top} is an \ell-dimensional random vector with independent components, for some \ell\in\mathbb{N}. A non-diagonal matrix QQ permits to introduce general correlation structures among the dd asset returns. Without loss of generality, we assume that rank(Q)=d{\rm rank}(Q)=d. Under this assumption, it holds that ={0}\mathcal{L}^{\perp}=\{0\}.

Remark 3.1.

Multi-factor models are widely employed in financial economics and econometrics, the Arbitrage Pricing Theory of [Ros76] and its extensions representing some of the most notable instances (see [BF17, Chapter 5], [CK95] and [CLM97, Chapter 6] for overviews on the topic). Multi-factor asset pricing models can always be written in the form (3.1), modulo the assumption of independent factors.777Under this assumption, representation (3.1) enables us to reduce the analysis to \ell independent sources of randomness. We stress that any correlation structure among the asset returns RR can be generated by a suitable specification of the matrix QQ. To this effect, recall first that the random vector RR represents the excess returns of dd risky assets with respect to a baseline security, usually chosen as a riskless asset. Multi-factor models are typically stated in the form

(3.2) R=𝔼[R]+BF+ϵ,R=\mathbb{E}[R]+BF+\epsilon,

where 𝔼[R]\mathbb{E}[R] is the vector of risk premia, FF is a kk-dimensional random vector of common risk factors, for some k<dk<d, Bd×kB\in\mathbb{R}^{d\times k} is the matrix of factor loadings and ϵ\epsilon is a dd-dimensional random vector of idiosyncratic (asset-specific) risk factors. Depending on the modeling choices, FF can represent a vector of economic factors or statistical factors. In the standard formulation (see, e.g., [Ing87, Chapter 7]), all components of FF and ϵ\epsilon are assumed to be uncorrelated. Notice now that factor model (3.2) can be written in the form (3.1) by setting Q=(𝔼[R],B,𝕀d)Q=(\mathbb{E}[R],B,\mathbb{I}_{d}) and Y=(1,F,ϵ)Y=(1,F^{\top},\epsilon^{\top})^{\top}, where 𝕀d\mathbb{I}_{d} denotes the (d×d)(d\times d) identity matrix. In the special case of absence of idiosyncratic risk, the vector YY can be directly identified with FF. Equation (3.1) therefore provides the simplest unifying representation of multi-factor asset pricing models.

For k=1,,k=1,\ldots,\ell, we denote by 𝒴k\mathcal{Y}_{k} the support of YkY_{k} and let ykinf:=inf𝒴ky_{k}^{\inf}:=\inf\mathcal{Y}_{k} and yksup:=sup𝒴ky^{\sup}_{k}:=\sup\mathcal{Y}_{k}. In this section, we work under the following standing assumption:

(3.3) y1inf=0,y1sup=+ and ykinf<0<yksup, for all k=2,,.y_{1}^{\inf}=0,\quad y_{1}^{\sup}=+\infty\qquad\text{ and }\qquad y_{k}^{\inf}<0<y_{k}^{\sup},\quad\text{ for all }k=2,\ldots,\ell.

As will become clear in the sequel, condition (3.3) corresponds to viewing the first factor Y1Y_{1} as the driving force of possible arbitrage opportunities, while the remaining factors cannot be exploited to generate arbitrage.888The only requirement in order to allow for arbitrage opportunities is the existence of a linear combination of factors with positive support. The assumption that Y1Y^{1} has positive support is only made for convention. In the context of the factor model (3.1)-(3.3), the following lemma gives a necessary and sufficient condition to ensure positive asset prices. For i=1,,di=1,\ldots,d and k=1,,k=1,\ldots,\ell, we denote by qi,kq_{i,k} the element on the ii-th row and kk-th column of QQ.

Lemma 3.2.

In the context of the model of this section, for each i=1,,di=1,\ldots,d, it holds that Ri1R^{i}\geq-1 a.s. if and only if the following condition is satisfied:

(3.4) qi,10 and k=2(qi,k+ykinfqi,kyksup)1,q_{i,1}\geq 0\qquad\text{ and }\qquad\sum_{k=2}^{\ell}\bigl{(}q^{+}_{i,k}y_{k}^{\inf}-q^{-}_{i,k}y^{\sup}_{k}\bigr{)}\geq-1,

with the convention 0×()=00\times(-\infty)=0 and 0×(+)=00\times(+\infty)=0.

Proof.

Condition (3.4) is obviously sufficient to ensure that Ri1R^{i}\geq-1 a.s., for all i=1,,di=1,\ldots,d. Conversely, let i{1,,d}i\in\{1,\ldots,d\} and suppose that Ri1R^{i}\geq-1 a.s. For all nn\in\mathbb{N} and k=1,,k=1,\ldots,\ell, let

ykinf(n):=(ykinf+1n)(n) and yksup(n):=(yksup1n)n,y^{\inf}_{k}(n):=\Bigl{(}y^{\inf}_{k}+\frac{1}{n}\Bigr{)}\vee(-n)\qquad\text{ and }\qquad y^{\sup}_{k}(n):=\Bigl{(}y^{\sup}_{k}-\frac{1}{n}\Bigr{)}\wedge n,

where xz:=max{x,z}x\vee z:=\max\{x,z\} and xz:=min{x,z}x\wedge z:=\min\{x,z\}, for any (x,z)2(x,z)\in\mathbb{R}^{2}. With this notation, it holds that P(Ykykinf(n))>0P(Y_{k}\leq y^{\inf}_{k}(n))>0 and P(Ykyksup(n))>0P(Y_{k}\geq y^{\sup}_{k}(n))>0, for all nn\in\mathbb{N} and k=1,,k=1,\ldots,\ell. Let Ki+:={k{1,,}:qi,k0}K^{+}_{i}:=\{k\in\{1,\ldots,\ell\}:q_{i,k}\geq 0\} and Ki:={1,,}Ki+K^{-}_{i}:=\{1,\ldots,\ell\}\setminus K^{+}_{i}. Since k=1qi,kYk1\sum_{k=1}^{\ell}q_{i,k}Y_{k}\geq-1 a.s. and due to the independence of the factors {Y1,,Y}\{Y_{1},\ldots,Y_{\ell}\}, it holds that

0\displaystyle 0 <P(Ykykinf(n) and Yjyjsup(n);kKi+,jKi)\displaystyle<P\biggl{(}Y_{k}\leq y^{\inf}_{k}(n)\text{ and }Y_{j}\geq y^{\sup}_{j}(n);\;\forall k\in K^{+}_{i},\forall j\in K^{-}_{i}\biggr{)}
=P(kKi+qi,kYk1jKiqi,jYj and Ykykinf(n) and Yjyjsup(n);kKi+,jKi).\displaystyle=P\Biggl{(}\sum_{k\in K^{+}_{i}}q_{i,k}Y_{k}\geq-1-\sum_{j\in K^{-}_{i}}q_{i,j}Y_{j}\;\text{ and }Y_{k}\leq y^{\inf}_{k}(n)\text{ and }Y_{j}\geq y^{\sup}_{j}(n);\;\forall k\in K^{+}_{i},\forall j\in K^{-}_{i}\Biggr{)}.

In turn, this necessarily implies that kKi+qi,kykinf(n)1jKiqi,jyjsup(n),\sum_{k\in K^{+}_{i}}q_{i,k}y^{\inf}_{k}(n)\geq-1-\sum_{j\in K^{-}_{i}}q_{i,j}y^{\sup}_{j}(n), for each nn\in\mathbb{N}. Condition (3.4) follows by letting n+n\rightarrow+\infty and using condition (3.3). ∎

In particular, condition (3.4) requires that qi,k0q_{i,k}\geq 0 if yksup=+y^{\sup}_{k}=+\infty and qi,k0q_{i,k}\leq 0 if ykinf=y^{\inf}_{k}=-\infty, for all i=1,,di=1,\ldots,d and k=1,,k=1,\ldots,\ell. Observe that condition (3.4) relates the support of the random factors to the dependence structure of the asset returns, represented by the off-diagonal elements of QQ. Arguing similarly as in Lemma 3.2, it can be shown that the set Θadm\Theta_{\rm adm} of admissible strategies can be represented as follows:

(3.5) Θadm={πd:πQ,10 and k=2((πQ,k)+ykinf(πQ,k)yksup)1},\Theta_{\rm adm}=\left\{\pi\in\mathbb{R}^{d}:\pi^{\top}Q_{\bullet,1}\geq 0\text{ and }\sum_{k=2}^{\ell}\left((\pi^{\top}Q_{\bullet,k})^{+}y^{\inf}_{k}-(\pi^{\top}Q_{\bullet,k})^{-}y^{\sup}_{k}\right)\geq-1\right\},

where Q,kQ_{\bullet,k} denotes the kk-th column of the matrix QQ, with the same convention as in (3.4).

We now introduce additional trading restrictions, as considered in Section 2.1. More specifically, we assume the presence of borrowing constraints:

(3.6) Θc:={πd:π,𝟏c},\Theta_{\rm c}:=\{\pi\in\mathbb{R}^{d}:\langle\pi,\mathbf{1}\rangle\leq c\},

for some fixed c>0c>0. If c(0,1)c\in(0,1), this corresponds to requiring that at least a proportion 1c1-c of the initial wealth is invested in the riskless asset, while, if c1c\geq 1, at most a proportion c1c-1 of the initial wealth can be borrowed from the riskless asset. Note that, since the set Θc\Theta_{\rm c} is not a cone, the notions of arbitrage opportunity and arbitrage of the first kind differ (see Remark 2.3). As in Section 2.1, the set Θ\Theta of allowed strategies is defined as Θ:=ΘadmΘc\Theta:=\Theta_{\rm adm}\cap\Theta_{\rm c}.

The following proposition summarizes the arbitrage properties of the factor model under consideration, in the presence of borrowing constraints. We denote by (Q)\mathcal{R}(Q^{\top}) the range of the matrix QQ^{\top} and by ek{\rm e}_{k} the kk-th vector of the canonical basis of \mathbb{R}^{\ell}, for k=1,,k=1,\ldots,\ell.

Proposition 3.3.

In the context of the model of this section, the following hold:

  1. (i)

    there are arbitrage opportunities if and only if e1(Q){\rm e}_{1}\in\mathcal{R}(Q^{\top}). In that case, it holds that

    (3.7) arbΘ={λ(QQ)1Q,1:λ>0 and λ(QQ)1Q,1,𝟏c};\mathcal{I}_{\rm arb}\cap\Theta=\bigl{\{}\lambda(QQ^{\top})^{-1}Q_{\bullet,1}:\lambda>0\text{ and }\lambda\langle(QQ^{\top})^{-1}Q_{\bullet,1},\mathbf{1}\rangle\leq c\bigr{\}};
  2. (ii)

    if e1(Q){\rm e}_{1}\in\mathcal{R}(Q^{\top}), then NA1{\rm NA}_{1} holds if and only if (QQ)1Q,1,𝟏>0\langle(QQ^{\top})^{-1}Q_{\bullet,1},\mathbf{1}\rangle>0.

Proof.

(i): let πd\pi\in\mathbb{R}^{d} such that π,QY0\langle\pi,QY\rangle\geq 0 a.s. The same argument used to prove Lemma 3.2 and representation (3.5) implies that the vector π\pi satisfies πQ,10\pi^{\top}Q_{\bullet,1}\geq 0 and

k=2((πQ,k)+ykinf(πQ,k)yksup)0.\sum_{k=2}^{\ell}\left((\pi^{\top}Q_{\bullet,k})^{+}y^{\inf}_{k}-(\pi^{\top}Q_{\bullet,k})^{-}y^{\sup}_{k}\right)\geq 0.

Recalling condition (3.3), this implies that πQ,k=0\pi^{\top}Q_{\bullet,k}=0, for all k=2,,k=2,\ldots,\ell. It follows that π,QY0\langle\pi,QY\rangle\geq 0 a.s. if and only if Qπ=λe1Q^{\top}\pi=\lambda{\rm e}_{1}, for some λ0\lambda\geq 0. Since rank(Q)=d{\rm rank}(Q)=d, it holds that arb={λ(QQ)1Qe1:λ>0}\mathcal{I}_{\rm arb}=\{\lambda(QQ^{\top})^{-1}Q{\rm e}_{1}:\lambda>0\}, from which representation (3.7) of the set arbΘ\mathcal{I}_{\rm arb}\cap\Theta follows directly from the definition of the set Θc\Theta_{\rm c} in (3.6).
(ii): by Proposition 2.2, NA1 holds if and only if arbΘ^=\mathcal{I}_{\rm arb}\cap\widehat{\Theta}=\emptyset. Representation (3.7) implies that arbΘ^=\mathcal{I}_{\rm arb}\cap\widehat{\Theta}=\emptyset if and only if (QQ)1Q,1,𝟏>0\langle(QQ^{\top})^{-1}Q_{\bullet,1},\mathbf{1}\rangle>0. ∎

Remark 3.4.

The vector (QQ)1Q,1(QQ^{\top})^{-1}Q_{\bullet,1} corresponds to the strategy replicating the factor Y1Y_{1}. While exact replication of Y1Y_{1} may be precluded by borrowing constraints, (3.7) shows that any allowed strategy that replicates a positive fraction of Y1Y_{1} is an arbitrage opportunity. The factor Y1Y_{1} can be (super-)replicated at zero cost if (QQ)1Q,1,𝟏0\langle(QQ^{\top})^{-1}Q_{\bullet,1},\mathbf{1}\rangle\leq 0, in which case NA1 fails.

Remark 3.5.

The proof of Proposition 3.3 shows that a strategy πarbΘ\pi\in\mathcal{I}_{\rm arb}\cap\Theta necessarily satisfies πQ,k=0\pi^{\top}Q_{\bullet,k}=0, for all k=2,,k=2,\ldots,\ell. When =d\ell=d, this corresponds to a set of d1d-1 linear equations in dd variables. This set defines a line in d\mathbb{R}^{d}, which we call arbitrage line. This concept will be illustrated in the two-dimensional model considered in Section 3.3.

In view of Theorem 2.5, NA1 ensures the well-posedness of optimal portfolio problems. In the presence of arbitrage opportunities, the borrowing constraint (3.6) is binding for every optimal allowed strategy. This is a direct consequence of the following simple result.

Lemma 3.6.

In the context of the model of this section, suppose that e1(Q){\rm e}_{1}\in\mathcal{R}(Q^{\top}) and NA1{\rm NA}_{1} holds. Then, for every πΘ\pi\in\Theta, there exists an element π^Θ\hat{\pi}\in\Theta such that

π^,QYπ,QY a.s. and π^,𝟏=c.\langle\hat{\pi},QY\rangle\geq\langle\pi,QY\rangle\text{ a.s.}\qquad\text{ and }\qquad\langle\hat{\pi},\mathbf{1}\rangle=c.

Moreover, there exists a strategy πmax\pi^{\max}, explicitly given by

(3.8) πmax=c(QQ)1Q,1,𝟏(QQ)1Q,1,\pi^{\max}=\frac{c}{\langle(QQ^{\top})^{-1}Q_{\bullet,1},\mathbf{1}\rangle}(QQ^{\top})^{-1}Q_{\bullet,1},

such that πmax,𝟏=c\langle\pi^{\max},\mathbf{1}\rangle=c and πmax,QYπ,QY\langle\pi^{\max},QY\rangle\geq\langle\pi,QY\rangle a.s., for all πarbΘ\pi\in\mathcal{I}_{\rm arb}\cap\Theta.

Proof.

Let π\pi be an arbitrary allowed strategy. Letting λ:=(cπ,𝟏)(QQ)1Q,1,𝟏10\lambda:=(c-\langle\pi,\mathbf{1}\rangle)\langle(QQ^{\top})^{-1}Q_{\bullet,1},\mathbf{1}\rangle^{-1}\geq 0, define the strategy π^:=π+λ(QQ)1Q,1\hat{\pi}:=\pi+\lambda(QQ^{\top})^{-1}Q_{\bullet,1}. Clearly, it holds that π^,𝟏=c\langle\hat{\pi},\mathbf{1}\rangle=c and, in addition, π^,QY=π,QY+λe1Q(QQ)1QY=π,QY+λY1π,QY\langle\hat{\pi},QY\rangle=\langle\pi,QY\rangle+\lambda{\rm e}_{1}^{\top}Q^{\top}(QQ^{\top})^{-1}QY=\langle\pi,QY\rangle+\lambda Y_{1}\geq\langle\pi,QY\rangle a.s. The second part of the lemma follows as a direct consequence of the characterization (3.7) of the set arbΘ\mathcal{I}_{\rm arb}\cap\Theta. ∎

We call maximal arbitrage strategy the strategy πmax\pi^{\max} given in (3.8). Whenever NA1 fails to hold (i.e., (QQ)1Q,1,𝟏0\langle(QQ^{\top})^{-1}Q_{\bullet,1},\mathbf{1}\rangle\leq 0), a maximal arbitrage strategy does not exist, because arbitrage opportunities can be arbitrarily scaled. Note that πmax\pi^{\max} is not necessarily the optimal strategy in an expected utility maximization problem of type (2.2). Similarly, πmax\pi^{\max} does not necessarily coincide with the numéraire portfolio ρ\rho. This will be explicitly illustrated in Examples 3.93.11.

Remark 3.7 (On relative arbitrage).

(1) In the context of the model of this section, let us assume that NA1{\rm NA}_{1} holds and e1(Q){\rm e}_{1}\in\mathcal{R}(Q^{\top}). Then, for θΘ\theta\in\Theta, there exist arbitrage opportunities relative to θ\theta if and only if θ,𝟏<c\langle\theta,\mathbf{1}\rangle<c. Indeed, if θ,𝟏<c\langle\theta,\mathbf{1}\rangle<c, then the existence of an arbitrage opportunity relative to θ\theta follows from Lemma 3.6. Conversely, suppose that θ,𝟏=c\langle\theta,\mathbf{1}\rangle=c and let πd\pi\in\mathbb{R}^{d} with πθarb\pi-\theta\in\mathcal{I}_{\rm arb}. By Proposition 3.3, this holds if and only if πθ=η(QQ)1Q,1\pi-\theta=\eta(QQ^{\top})^{-1}Q_{\bullet,1}, for some η>0\eta>0. However, since π,𝟏=θ,𝟏+η(QQ)1Q,1,𝟏>c,\langle\pi,\mathbf{1}\rangle=\langle\theta,\mathbf{1}\rangle+\eta\langle(QQ^{\top})^{-1}Q_{\bullet,1},\mathbf{1}\rangle>c, the strategy π\pi is not an allowed trading strategy. This shows that there cannot exist arbitrage opportunities relative to θ\theta if θ,𝟏=c\langle\theta,\mathbf{1}\rangle=c. In particular, there do not exist arbitrage opportunities relative to πmax\pi^{\max}.

(2) One can also study the existence of arbitrage opportunities relative to the market portfolio πmkt\pi^{\rm mkt} defined by πimkt:=S0i/S0,𝟏\pi^{\rm mkt}_{i}:=S^{i}_{0}/\langle S_{0},\mathbf{1}\rangle, for i=1,,di=1,\ldots,d (see [FK09, Section 2]). As a consequence of part (1) of this remark, arbitrage opportunities relative to the market exist if and only if c>1c>1. The financial intuition is that, if c>1c>1, then it is possible to invest the whole initial capital vv in the market portfolio, borrow an amount v(c1)v(c-1) from the riskless asset and invest that amount in the strategy πmax\pi^{\max}, thus improving the performance of the market portfolio. The strategy πΘ\pi^{*}\in\Theta which best outperforms the market portfolio is given by π=πmkt+c1cπmax\pi^{*}=\pi^{\rm mkt}+\frac{c-1}{c}\pi^{\max}.

3.2. The case of a unit triangular matrix QQ

Let us consider the special case where QQ is a (d×d)(d\times d) upper triangular matrix with qi,i=1q_{i,i}=1, for all i=1,,di=1,\ldots,d. In this case, the results presented in Section 3.1 can be stated explicitly in terms of the elements of QQ. First, condition (3.4) ensuring the positivity of asset prices can be rewritten in the following recursive form:

(3.9) ydinf1 and yiinf1k=i+1d(qi,k+ykinfqi,kyksup), for all i=1,,d1.y^{\inf}_{d}\geq-1\qquad\text{ and }\qquad y^{\inf}_{i}\geq-1-\sum_{k=i+1}^{d}\bigl{(}q^{+}_{i,k}y^{\inf}_{k}-q^{-}_{i,k}y^{\sup}_{k}\bigr{)},\quad\text{ for all }i=1,\ldots,d-1.

In view of (3.5), the set Θadm\Theta_{\rm adm} of admissible strategies takes the form

(3.10) Θadm={πd:π10 and k=2d((i=1k1πiqi,k+πk)+ykinf(i=1k1πiqi,k+πk)yksup)1}.\Theta_{\rm adm}=\left\{\pi\in\mathbb{R}^{d}:\pi_{1}\geq 0\text{ and }\sum_{k=2}^{d}\left(\Biggl{(}\sum_{i=1}^{k-1}\pi_{i}q_{i,k}+\pi_{k}\Biggr{)}^{+}y^{\inf}_{k}-\Biggl{(}\sum_{i=1}^{k-1}\pi_{i}q_{i,k}+\pi_{k}\Biggr{)}^{-}y^{\sup}_{k}\right)\geq-1\right\}.

Since rank(Q)=d{\rm rank}(Q)=d, the condition e1(Q){\rm e}_{1}\in\mathcal{R}(Q^{\top}) is automatically satisfied and, therefore, there exist arbitrage opportunities (see Proposition 3.3). More specifically, it holds that

(3.11) arbΘ={λQ1,1:λ>0 and λQ1,1,𝟏c},\mathcal{I}_{\rm arb}\cap\Theta=\bigl{\{}\lambda Q^{-1}_{1,\bullet}:\lambda>0\text{ and }\lambda\langle Q^{-1}_{1,\bullet},\mathbf{1}\rangle\leq c\bigr{\}},

where Q1,1Q^{-1}_{1,\bullet} denotes the first row of the matrix Q1Q^{-1}, written as a column vector. The following lemma gives an explicit representation of the vector Q1,1Q^{-1}_{1,\bullet}, which determines all the arbitrage properties of the model under consideration.

Lemma 3.8.

In the context of the model of this section, suppose that QQ is a unit triangular matrix. Then, for all k=1,,dk=1,\ldots,d, it holds that Q1,k1=αkQ^{-1}_{1,k}=\alpha_{k}, where αk\alpha_{k} is defined by

α1:=1andαk:=JA(k)(1)|J|1l=1|J|1qjl,jl+1,for k=2,,d,\alpha_{1}:=1\qquad\text{and}\qquad\alpha_{k}:=\sum_{J\in A(k)}(-1)^{|J|-1}\prod_{l=1}^{|J|-1}q_{j_{l},j_{l+1}},\quad\text{for }k=2,\ldots,d,

where A(k)A(k) denotes the family of all subsets J={j1,,jr}{1,,k}J=\{j_{1},\ldots,j_{r}\}\subseteq\{1,\ldots,k\}, with rkr\leq k, such that j1=1j_{1}=1, jr=kj_{r}=k and jl<jl+1j_{l}<j_{l+1}, for all l=1,,r1l=1,\ldots,r-1, and |J||J| denotes the cardinality of JJ.

Proof.

The vector Q1,1Q^{-1}_{1,\bullet} is the unique solution πd\pi\in\mathbb{R}^{d} to the linear system Qπ=e1Q^{\top}\pi={\rm e}_{1}. Since QQ is a unit triangular matrix, the solution π\pi is characterized by π1=1\pi_{1}=1 and by the recursive relation

(3.12) πk=i=1k1πiqi,k, for all k=2,,d.\pi_{k}=-\sum_{i=1}^{k-1}\pi_{i}q_{i,k},\qquad\text{ for all }k=2,\ldots,d.

To prove the lemma, it suffices to show that the vector α=(α1,,αd)\alpha=(\alpha_{1},\ldots,\alpha_{d})^{\top} satisfies (3.12). To this effect, notice that, for every k=2,,dk=2,\ldots,d,

i=1k1αiqi,k=q1,ki=2k1JA(i)(1)|J|1l=1|J|1qjl,jl+1qi,k=αk.-\sum_{i=1}^{k-1}\alpha_{i}q_{i,k}=-q_{1,k}-\sum_{i=2}^{k-1}\sum_{J\in A(i)}(-1)^{|J|-1}\prod_{l=1}^{|J|-1}q_{j_{l},j_{l+1}}q_{i,k}=\alpha_{k}.

This shows that α=(α1,,αd)\alpha=(\alpha_{1},\ldots,\alpha_{d})^{\top} satisfies (3.12) and, therefore, it holds that Q1,1=αQ^{-1}_{1,\bullet}=\alpha. ∎

In view of (3.11), the vector α\alpha introduced in Lemma 3.8 generates all arbitrage strategies, up to a multiplicative factor depending on the borrowing constraint cc. More precisely, every arbitrage strategy π\pi is necessarily of the form π=λα\pi=\lambda\alpha, with λ>0\lambda>0 satisfying λα,𝟏c\lambda\langle\alpha,\mathbf{1}\rangle\leq c, and is such that V1π=1+λY1V_{1}^{\pi}=1+\lambda Y_{1}. Furthermore, by (3.12), all such strategies π\pi belong to the arbitrage line (see Remark 3.5). As an example, for d=4d=4, all arbitrage strategies are proportional to

α=(1q1,2q1,3+q1,2q2,3q1,4+q1,2q2,4+q1,3q3,4q1,2q2,3q3,4).\alpha=\begin{pmatrix}1\\ -q_{1,2}\\ -q_{1,3}+q_{1,2}\,q_{2,3}\\ -q_{1,4}+q_{1,2}\,q_{2,4}+q_{1,3}\,q_{3,4}-q_{1,2}\,q_{2,3}\,q_{3,4}\end{pmatrix}.

In the model considered in this subsection, the condition characterizing the validity of NA1 takes the simple form Q1,1,𝟏>0\langle Q^{-1}_{1,\bullet},\mathbf{1}\rangle>0 (see Proposition 3.3). As a consequence of Lemma 3.8, this implies the following explicit characterization of NA1:

(3.13) NA1 holds 1+J{1,,d}(1)|J|1l=1|J|1qjl,jl+1>0,\text{NA${}_{1}$ holds }\qquad\Longleftrightarrow\qquad 1+\sum_{J\subseteq\{1,\ldots,d\}}(-1)^{|J|-1}\prod_{l=1}^{|J|-1}q_{j_{l},j_{l+1}}>0,

where the summation is taken over all sets J={j1,,jr}J=\{j_{1},\ldots,j_{r}\}, with 2rd2\leq r\leq d, such that j1=1j_{1}=1 and jl<jl+1j_{l}<j_{l+1}, for all l=1,,r1l=1,\ldots,r-1. In view of (3.8), the same quantity appearing on the right of (3.13) represents the denominator of the maximal arbitrage strategy πmax\pi^{\max}.

3.3. A two-dimensional example with arbitrage

We now present a two-dimensional model that allows for a geometric visualization of the concepts introduced above. Let d=2d=2 and consider a pair (Y1,Y2)(Y_{1},Y_{2}) of independent random variables such that 𝒴1=[0,+)\mathcal{Y}_{1}=[0,+\infty) and y2inf<0<y2supy^{\inf}_{2}<0<y^{\sup}_{2}. Let

Q=(1γ01),Q=\begin{pmatrix}1&\gamma\\ 0&1\end{pmatrix},

with γ\gamma\in\mathbb{R}, and suppose that the asset returns (R1,R2)(R_{1},R_{2}) are generated as in (3.1). To ensure positive asset prices, condition (3.9) needs to be satisfied. In this example, the largest possible support of the distribution of the random factor Y2Y_{2} is given by

{y2inf=1 and y2sup=+, if γ[0,1);y2inf=1/γ and y2sup=+, if γ1;y2inf=1 and y2sup=1/γ, if γ<0.\begin{cases}y^{\inf}_{2}=-1\text{ and }y^{\sup}_{2}=+\infty,&\text{ if }\gamma\in[0,1);\\ y^{\inf}_{2}=-1/\gamma\text{ and }y^{\sup}_{2}=+\infty,&\text{ if }\gamma\geq 1;\\ y^{\inf}_{2}=-1\text{ and }y^{\sup}_{2}=-1/\gamma,&\text{ if }\gamma<0.\end{cases}

In view of (3.10), a strategy π=(π1,π2)\pi=(\pi_{1},\pi_{2}) is admissible if and only if

(3.14) {π10 and γπ1π21γπ1, if γ[0,1);π10 and γπ1π2γγπ1, if γ1;π10 and γγπ1π21γπ1, if γ<0.\begin{cases}\pi_{1}\geq 0\text{ and }-\gamma\pi_{1}\leq\pi_{2}\leq 1-\gamma\pi_{1},&\text{ if }\gamma\in[0,1);\\ \pi_{1}\geq 0\text{ and }-\gamma\pi_{1}\leq\pi_{2}\leq\gamma-\gamma\pi_{1},&\text{ if }\gamma\geq 1;\\ \pi_{1}\geq 0\text{ and }\gamma-\gamma\pi_{1}\leq\pi_{2}\leq 1-\gamma\pi_{1},&\text{ if }\gamma<0.\end{cases}

In this two-dimensional setting, the borrowing constraint (3.6) takes the form π1+π2c\pi_{1}+\pi_{2}\leq c. Together with (3.14), this constraint determines the set Θ\Theta of allowed strategies. Regardless of the values of γ\gamma and cc, arbitrage opportunities always exist. More specifically, it holds that

(3.15) arbΘ={π2:π1>0,π2=γπ1 and π1(1γ)c}.\mathcal{I}_{\rm arb}\cap\Theta=\bigl{\{}\pi\in\mathbb{R}^{2}:\pi_{1}>0,\,\pi_{2}=-\gamma\pi_{1}\text{ and }\pi_{1}(1-\gamma)\leq c\bigr{\}}\neq\emptyset.

The arbitrage line (see Remark 3.5) is described by the equation π2=γπ1\pi_{2}=-\gamma\pi_{1}. Figure 1 provides a visualization of the set Θ\Theta, with the arbitrage line highlighted in red.

π1\pi_{1}π2\pi_{2}π2=1γπ1\pi_{2}=1-\gamma\pi_{1}π2=cπ1\pi_{2}=c-\pi_{1}1cc(c1γ,cγ1γ)\quad(\frac{c}{1-\gamma},-\frac{c\gamma}{1-\gamma})\quadπ2=γπ1{\color[rgb]{0,0,0}\definecolor[named]{pgfstrokecolor}{rgb}{0,0,0}\pgfsys@color@gray@stroke{0}\pgfsys@color@gray@fill{0}\pi_{2}=-\gamma\pi_{1}}
Figure 1. Geometric illustration of the set Θ\Theta (yellow area), for c=2.5c=2.5 and γ=0.5\gamma=0.5.

The NA1 condition is satisfied if and only if Q1,1,𝟏>0\langle Q^{-1}_{1,\bullet},\mathbf{1}\rangle>0. Therefore, we have that

 NA1 holds γ<1.\text{ NA${}_{1}$ holds }\qquad\Longleftrightarrow\qquad\gamma<1.

Indeed, from (3.15) we have that arbΘ^=\mathcal{I}_{\rm arb}\cap\widehat{\Theta}=\emptyset if and only if γ<1\gamma<1. Graphically, this condition corresponds to requesting that the arbitrage line intersects the borrowing constraint line (see Figure 1), i.e., the line of equation π2=cπ1\pi_{2}=c-\pi_{1}. Observe also that the set Θ\Theta is compact if and only if such an intersection occurs (compare with condition (iv) in Proposition 2.2).

For γ<1\gamma<1, all arbitrage strategies are contained in the line segment passing through the origin and the point (π1max,π2max)(\pi^{\max}_{1},\pi^{\max}_{2}) characterizing the maximal arbitrage strategy and given by

(3.16) π1max=c1γ and π2max=cγ1γ,\pi^{\max}_{1}=\frac{c}{1-\gamma}\qquad\text{ and }\qquad\pi^{\max}_{2}=-\frac{c\gamma}{1-\gamma},

as follows from (3.8). Graphically, the strategy πmax\pi^{\max} corresponds to the point of intersection between the arbitrage line and the borrowing constraint line. If the two lines do not intersect, then every arbitrage opportunity can be arbitrarily scaled (i.e., NA1 fails to hold).

In view of Theorem 2.9, the numéraire portfolio ρ\rho exists if and only if γ<1\gamma<1. The numéraire portfolio may or may not coincide with the maximal arbitrage strategy πmax\pi^{\max}, depending on the distributional properties of Y1Y_{1} and Y2Y_{2}. For illustration, we present three simple examples.

Example 3.9.

Let γ[0,1)\gamma\in[0,1) and suppose that 𝔼[Y2]=0\mathbb{E}[Y_{2}]=0. In this case, it holds that ρ=πmax\rho=\pi^{\max}. Indeed, let π=(π1,π2)\pi=(\pi_{1},\pi_{2}) be an arbitrary strategy satisfying (3.14) and π1+π2c\pi_{1}+\pi_{2}\leq c. By Lemma 3.6, there exists a strategy of the form π^=(π^1,cπ^1)\hat{\pi}=(\hat{\pi}_{1},c-\hat{\pi}_{1}) such that V1π^V1πV^{\hat{\pi}}_{1}\geq V^{\pi}_{1} a.s. Due to (3.14), it necessarily holds that 0π^1c/(1γ)0\leq\hat{\pi}_{1}\leq c/(1-\gamma). Therefore, using the independence of Y1Y_{1} and Y2Y_{2} and the fact that 𝔼[Y2]=0\mathbb{E}[Y_{2}]=0, we have that

𝔼[V1πV1πmax]𝔼[V1π^V1πmax]=𝔼[1+π^1Y11+c1γY1]1,\mathbb{E}\left[\frac{V^{\pi}_{1}}{V^{\pi^{\max}}_{1}}\right]\leq\mathbb{E}\left[\frac{V^{\hat{\pi}}_{1}}{V^{\pi^{\max}}_{1}}\right]=\mathbb{E}\left[\frac{1+\hat{\pi}_{1}Y_{1}}{1+\frac{c}{1-\gamma}Y_{1}}\right]\leq 1,

where the last inequality follows from the fact that Y10Y_{1}\geq 0 a.s. This shows that the numéraire portfolio ρ\rho coincides with the maximal arbitrage strategy πmax\pi^{\max} given in (3.16).

Example 3.10.

Let γ=1/2\gamma=1/2 and c=1c=1. Suppose that Y1Exp(1)Y_{1}\sim{\rm Exp}(1) and 1+Y2Exp(β)1+Y_{2}\sim{\rm Exp}(\beta), with β>0\beta>0. In this case, for suitable values of β\beta, the maximal arbitrage strategy is not the numéraire portfolio. Indeed, considering the strategy (0,1)Θ(0,1)\in\Theta, we have that

𝔼[V1(0,1)V1πmax]=𝔼[1+Y21+2Y1]=1β𝔼[11+2Y1]=e2β1/2+exxdx0.461β.\mathbb{E}\left[\frac{V^{(0,1)}_{1}}{V^{\pi^{\max}}_{1}}\right]=\mathbb{E}\left[\frac{1+Y_{2}}{1+2Y_{1}}\right]=\frac{1}{\beta}\mathbb{E}\left[\frac{1}{1+2Y_{1}}\right]=\frac{\sqrt{e}}{2\beta}\int_{1/2}^{+\infty}\frac{e^{-x}}{x}\mathrm{d}x\approx\frac{0.461}{\beta}.

For any sufficiently small value of β\beta, it holds that 𝔼[V1(0,1)/V1πmax]>1\mathbb{E}[V^{(0,1)}_{1}/V^{\pi^{\max}}_{1}]>1 and, therefore, the strategy πmax\pi^{\max} cannot be the numéraire portfolio in that case. Furthermore, since V1πmaxV1πV^{\pi^{\max}}_{1}\geq V^{\pi}_{1} a.s. for all πarbΘ\pi\in\mathcal{I}_{\rm arb}\cap\Theta (see Lemma 3.6), the numéraire portfolio ρ\rho does not belong to the set of arbitrage opportunities (i.e., ρarbΘ\rho\notin\mathcal{I}_{\rm arb}\cap\Theta).999Taking for instance β=0.3\beta=0.3 in the example under consideration, the numéraire portfolio ρ\rho can be numerically computed as ρ(1.335,0.335)(2,1)=πmax\rho\approx(1.335,-0.335)\neq(2,-1)=\pi^{\max}. In view of Remark 2.10, the log-optimal strategy ρ\rho is therefore not an arbitrage strategy. Moreover, since the trading constraint (3.6) is binding for ρ\rho (as a consequence of Lemma 3.6), it is not allowed to improve the strategy ρ\rho by adding to it a fraction of any arbitrage strategy.

This example shows that, even in the presence of arbitrage, it is not necessarily optimal to invest in an arbitrage opportunity. The financial intuition is that, for a logarithmic investor and sufficiently small β\beta, the risk-reward profile of the strategy ρ\rho is more attractive than any arbitrage opportunity. Indeed, in the present example every allowed strategy π=(π1,π2)\pi=(\pi_{1},\pi_{2}) satisfies

V1π=1+π1Y1+(π1/2+π2)Y2.V^{\pi}_{1}=1+\pi_{1}Y^{1}+(\pi_{1}/2+\pi_{2})Y^{2}.

Since π1/2+π20\pi_{1}/2+\pi_{2}\geq 0 by (3.14), losses can only occur on the event {Y2<0}\{Y^{2}<0\}, which happens with probability 1exp(β)1-\exp(-\beta). In view of (3.15), arbitrage strategies satisfy π1/2+π2=0\pi_{1}/2+\pi_{2}=0 and therefore eliminate the influence of the risk factor Y2Y^{2}, with consequently no risk of losses. On the contrary, for sufficiently small β\beta the log-optimal strategy ρ\rho does not belong to the set arb\mathcal{I}_{\rm arb}, thus implying a positive exposure to the factor Y2Y^{2}. The financial explanation is that the log-optimal strategy can tolerate the risk of losses in order to profit from potentially large values of Y2Y^{2}, which are most likely for small values of β\beta.

Example 3.11.

Let γ<0\gamma<0 and suppose that 𝔼[Y1]<+\mathbb{E}[Y_{1}]<+\infty and 𝔼[Y2]<+\mathbb{E}[Y_{2}]<+\infty. Under these assumptions, the log-optimal portfolio π\pi^{*} exists and, therefore, it coincides with the numéraire portfolio ρ\rho. Lemma 3.6 together with (3.14) implies that π\pi^{*} is of the form (π1,cπ1)(\pi^{*}_{1},c-\pi^{*}_{1}), with π1D(c,γ):=[(c1)+1γ,cγ1γ]\pi^{*}_{1}\in D(c,\gamma):=[\frac{(c-1)^{+}}{1-\gamma},\frac{c-\gamma}{1-\gamma}]. Consider the function g:D(c,γ)g:D(c,\gamma)\rightarrow\mathbb{R} defined by

g(π1):=𝔼[log(V1(π1,cπ1))]=𝔼[log(1+π1(Y1+(γ1)Y2)+cY2)],g(\pi_{1}):=\mathbb{E}\bigl{[}\log\bigl{(}V^{(\pi_{1},c-\pi_{1})}_{1}\bigr{)}\bigr{]}=\mathbb{E}\bigl{[}\log\bigl{(}1+\pi_{1}\bigl{(}Y_{1}+(\gamma-1)Y_{2}\bigr{)}+cY_{2}\bigr{)}\bigr{]},

for π1D(c,γ)\pi_{1}\in D(c,\gamma). Since the function gg is concave and π1max=c/(1γ)\pi^{\max}_{1}=c/(1-\gamma) belongs to the interior of the interval D(c,γ)D(c,\gamma), the log-optimal portfolio π\pi^{*} is given by πmax\pi^{\max} if and only if g(π1max)=0g^{\prime}(\pi^{\max}_{1})=0. The latter condition is equivalent to

(3.17) 𝔼[Y11+c1γY1]=(1γ)𝔼[Y21+c1γY1].\mathbb{E}\left[\frac{Y_{1}}{1+\frac{c}{1-\gamma}Y_{1}}\right]=(1-\gamma)\mathbb{E}\left[\frac{Y_{2}}{1+\frac{c}{1-\gamma}Y_{1}}\right].

In the present example, ρ=πmax\rho=\pi^{\max} holds if and only if condition (3.17) is satisfied. In particular, unlike in Example 3.9 where γ[0,1)\gamma\in[0,1), note that (3.17) cannot be satisfied if 𝔼[Y2]=0\mathbb{E}[Y_{2}]=0.

4. The multi-period setting

In this section, we extend the analysis of Section 2 to the multi-period case. We allow for convex trading constraints evolving randomly over time and prove that NA1 holds in a dynamic setting if and only if it holds in each single trading period. This fundamental fact enables us to address the multi-period case by relying on arguments similar to those employed in Section 2. For brevity of presentation, we prove multi-period versions of only the central results characterizing market viability and NA1, the remaining results and remarks admitting analogous extensions.

4.1. Setting and trading restrictions

Let (Ω,,𝔽,P)(\Omega,\mathcal{F},\mathbb{F},P) be a filtered probability space, where 𝔽=(t)t=0,1,,T\mathbb{F}=(\mathcal{F}_{t})_{t=0,1,\ldots,T} and 0\mathcal{F}_{0} is the trivial σ\sigma-field completed by the PP-nullsets of \mathcal{F}, for a fixed time horizon TT\in\mathbb{N}. Similarly to Section 2, we consider dd risky assets and a riskless asset with constant price equal to one. The discounted prices of the dd risky assets are represented by the dd-dimensional adapted process S=(St)t=0,1,,TS=(S_{t})_{t=0,1,\ldots,T}. For each i=1,,di=1,\ldots,d, we assume that

Sti=St1i(1+Rti), for all t=1,,T,S^{i}_{t}=S^{i}_{t-1}(1+R^{i}_{t}),\qquad\text{ for all }t=1,\ldots,T,

where each random variable RtiR^{i}_{t} is t\mathcal{F}_{t}-measurable, satisfies Rti1R^{i}_{t}\geq-1 a.s. and represents the return of asset ii on the period [t1,t][t-1,t]. For each t=1,,Tt=1,\ldots,T, we denote by 𝒮t\mathcal{S}_{t} the t1\mathcal{F}_{t-1}-conditional support of the random vector Rt=(Rt1,,Rtd)R_{t}=(R^{1}_{t},\ldots,R^{d}_{t})^{\top} (i.e., the support of a regular version of the t1\mathcal{F}_{t-1}-conditional distribution of RtR_{t}, see [BCL19, Definition 2.2]). We also denote by t\mathcal{L}_{t} the smallest linear subspace of d\mathbb{R}^{d} containing 𝒮t\mathcal{S}_{t} and by t\mathcal{L}_{t}^{\perp} its orthogonal complement. Conditional expectations are to be understood in the generalized sense (see, e.g., [HWY92, Section 1.4]).

A set-valued process A=(At)t=1,,TA=(A_{t})_{t=1,\ldots,T} is said to be predictable if, for each t=1,,Tt=1,\ldots,T, the correspondence (set-valued mapping) AtA_{t} from Ω\Omega to d\mathbb{R}^{d} is t1\mathcal{F}_{t-1}-measurable.101010We recall that a correspondence AtA_{t} from Ω\Omega to d\mathbb{R}^{d} is t1\mathcal{F}_{t-1}-measurable if, for every open subset GdG\subset\mathbb{R}^{d}, it holds that {ωΩ:At(ω)G}t1\{\omega\in\Omega:A_{t}(\omega)\cap G\neq\emptyset\}\in\mathcal{F}_{t-1}, see [RW98, Definition 14.1]. The processes 𝒮=(𝒮t)t=1,,T\mathcal{S}=(\mathcal{S}_{t})_{t=1,\ldots,T}, =(t)t=1,,T\mathcal{L}=(\mathcal{L}_{t})_{t=1,\ldots,T} and =(t)t=1,,T\mathcal{L}^{\perp}=(\mathcal{L}_{t}^{\perp})_{t=1,\ldots,T} are all predictable (see [BCL19, Lemma 2.4] and [RW98, Exercise 14.12-(d)]). For each t=1,,Tt=1,\ldots,T, the orthogonal projection of a vector xdx\in\mathbb{R}^{d} on t\mathcal{L}_{t} is denoted by pt(x){\rm p}_{\mathcal{L}_{t}}(x) and it is t1\mathcal{F}_{t-1}-measurable (see [RW98, Exercise 14.17]).

We describe trading strategies via predictable processes π=(πt)t=1,,T\pi=(\pi_{t})_{t=1,\ldots,T}, with πt=(πt1,,πtd)\pi_{t}=(\pi^{1}_{t},\ldots,\pi^{d}_{t})^{\top} representing fractions of wealth held in the dd risky assets between time t1t-1 and time tt. We denote by Vtπ(v)V^{\pi}_{t}(v) the wealth at time tt generated by strategy π\pi starting from capital v>0v>0, with

V0π(v)=v and Vtπ(v)=vk=1t(1+πk,Rk),for t=1,,T.V^{\pi}_{0}(v)=v\qquad\text{ and }\qquad V^{\pi}_{t}(v)=v\prod_{k=1}^{t}(1+\langle\pi_{k},R_{k}\rangle),\quad\text{for }t=1,\ldots,T.

As in Section 2.1, we define Vtπ:=Vtπ(1)V^{\pi}_{t}:=V^{\pi}_{t}(1). A predictable strategy π\pi is said to be admissible if Vtπ0V^{\pi}_{t}\geq 0 a.s., for all t=1,,Tt=1,\ldots,T. Equivalently, introducing the random set

(4.1) Θadm,t:={πd:π,z1 for all z𝒮t}, for t=1,,T,\Theta_{{\rm adm},t}:=\{\pi\in\mathbb{R}^{d}:\langle\pi,z\rangle\geq-1\text{ for all }z\in\mathcal{S}_{t}\},\qquad\text{ for }t=1,\ldots,T,

a predictable strategy π\pi is admissible if and only if πtΘadm,t\pi_{t}\in\Theta_{{\rm adm},t} holds a.s. for all t=1,,Tt=1,\ldots,T. Note that, for every (ω,t)Ω×{1,,T}(\omega,t)\in\Omega\times\{1,\ldots,T\}, the set Θadm,t(ω)\Theta_{{\rm adm},t}(\omega) is a non-empty, closed and convex subset of d\mathbb{R}^{d}. Arguing similarly as in [RW98, Exercise 14.12-(e)], it can be shown that the predictability of 𝒮\mathcal{S} implies that the set-valued process Θadm=(Θadm,t)t=1,,T\Theta_{\rm adm}=(\Theta_{{\rm adm},t})_{t=1,\ldots,T} is predictable.

Trading constraints are modelled through a set-valued predictable process Θc=(Θc,t)t=1,,T\Theta_{\rm c}=(\Theta_{{\rm c},t})_{t=1,\ldots,T} such that Θc,t(ω)\Theta_{{\rm c},t}(\omega) is a convex closed subset of d\mathbb{R}^{d}, for all (ω,t)Ω×{1,,T}(\omega,t)\in\Omega\times\{1,\ldots,T\}. Similarly as in Section 2.1, we assume that t(ω)Θc,t(ω)\mathcal{L}^{\perp}_{t}(\omega)\subset\Theta_{{\rm c},t}(\omega), for all (ω,t)Ω×{1,,T}(\omega,t)\in\Omega\times\{1,\ldots,T\}. The family of allowed strategies is given by all d\mathbb{R}^{d}-valued predictable processes π=(πt)t=1,,T\pi=(\pi_{t})_{t=1,\ldots,T} such that πt\pi_{t} belongs a.s. to Θt:=Θadm,tΘc,t\Theta_{t}:=\Theta_{{\rm adm},t}\cap\Theta_{{\rm c},t}, for all t=1,,Tt=1,\ldots,T. Note that, as a consequence of [RW98, Proposition 14.11], the set-valued process Θ=(Θt)t=1,,T\Theta=(\Theta_{t})_{t=1,\ldots,T} is predictable. For brevity of notation, we shall simply write πΘ\pi\in\Theta to denote that a trading strategy π\pi is allowed. For each (ω,t)Ω×{1,,T}(\omega,t)\in\Omega\times\{1,\ldots,T\}, the set Θ^t(ω)\widehat{\Theta}_{t}(\omega) is defined as the recession cone of Θt(ω)\Theta_{t}(\omega). The set-valued process Θ^=(Θ^t)t=1,,T\widehat{\Theta}=(\widehat{\Theta}_{t})_{t=1,\ldots,T} is predictable, as a consequence of the predictability of Θ\Theta together with [RW98, Exercise 14.21], and admits the same financial interpretation as the recession cone Θ^\widehat{\Theta} introduced in a single-period setting in Section 2.1.

Remark 4.1.

Trading constraints evolving randomly over time arise naturally as a consequence of the admissibility requirement and are not purely motivated by mathematical generality, as pointed out also in [KK07]. Indeed, admissibility requires that, for each t=1,Tt=1\ldots,T, the strategy πt\pi_{t} is chosen at time t1t-1 in such a way that πt,Rt1\langle\pi_{t},R_{t}\rangle\geq-1 a.s., conditionally on the information available up to time t1t-1. Therefore, as becomes apparent from (4.1), the randomness of Θadm,t\Theta_{{\rm adm},t} is due to the fact that the t1\mathcal{F}_{t-1}-conditional support 𝒮t\mathcal{S}_{t} of RtR_{t} and, therefore, the set of admissible strategies πt\pi_{t} may depend on the realizations of the asset returns (R1,,Rt1)(R_{1},\ldots,R_{t-1}). Consequently, even in the presence of deterministic trading constraints Θc\Theta_{\rm c}, the set-valued process Θ\Theta of allowed strategies is a deterministic process only in the special case where the asset returns (Rt)t=1,,T(R_{t})_{t=1,\ldots,T} form a sequence of serially independent random vectors.

4.2. Arbitrage concepts

An allowed strategy πΘ\pi\in\Theta is said to be an arbitrage opportunity if

(4.2) P(VTπ1)=1 and P(VTπ>1)>0.P(V^{\pi}_{T}\geq 1)=1\qquad\text{ and }\qquad P(V^{\pi}_{T}>1)>0.

We say that no classical arbitrage holds if there does not exist a strategy πΘ\pi\in\Theta satisfying (4.2). For t=1,,Tt=1,\ldots,T, we denote by L+0(t)L^{0}_{+}(\mathcal{F}_{t}) the family of non-negative t\mathcal{F}_{t}-measurable random variables. Definition 2.1 can be naturally extended to a multi-period setting as follows.

Definition 4.2.

A random variable ξL+0(T)\xi\in L^{0}_{+}(\mathcal{F}_{T}) with P(ξ>0)>0P(\xi>0)>0 is said to be an arbitrage of the first kind if v(ξ)=0v(\xi)=0, where v(ξ):=inf{v>0:πΘ such that VTπ(v)ξ a.s.}v(\xi):=\inf\{v>0:\exists\;\pi\in\Theta\text{ such that }V^{\pi}_{T}(v)\geq\xi\text{ a.s.}\}.
No arbitrage of the first kind (NA1) holds if, for every ξL+0(T)\xi\in L^{0}_{+}(\mathcal{F}_{T}), v(ξ)=0v(\xi)=0 implies ξ=0\xi=0 a.s.

As a preliminary to the statement of the next proposition, we define, for each t=1,,Tt=1,\ldots,T,

arb,t:={πd:π,z0 for all z𝒮t}t.\mathcal{I}_{{\rm arb},t}:=\{\pi\in\mathbb{R}^{d}:\langle\pi,z\rangle\geq 0\text{ for all }z\in\mathcal{S}_{t}\}\setminus\mathcal{L}_{t}^{\perp}.

By [RW98, Exercise 14.12-(e)], the random set arb,t\mathcal{I}_{{\rm arb},t} is t1\mathcal{F}_{t-1}-measurable, for all t=1,,Tt=1,\ldots,T. For a random variable ζL+0(t)\zeta\in L^{0}_{+}(\mathcal{F}_{t}), we define its super-hedging value at time t1t-1 by

vt1(ζ):=essinf{xL+0(t1):hL0(t1;Θt) such that x(1+h,Rt)ζ a.s.},{\rm v}_{t-1}(\zeta):=\operatorname{ess\,inf}\bigl{\{}x\in L^{0}_{+}(\mathcal{F}_{t-1}):\exists\;h\in L^{0}(\mathcal{F}_{t-1};\Theta_{t})\text{ such that }x(1+\langle h,R_{t}\rangle)\geq\zeta\text{ a.s.}\bigr{\}},

where L0(t1;Θt)L^{0}(\mathcal{F}_{t-1};\Theta_{t}) denotes the family of t1\mathcal{F}_{t-1}-measurable random vectors h:Ωdh:\Omega\rightarrow\mathbb{R}^{d} such that P(hΘt)=1P(h\in\Theta_{t})=1.

For the usual concept of no classical arbitrage, it is well-known that absence of arbitrage in a multi-period setting is equivalent to absence of arbitrage opportunities in each single trading period (see, e.g., [FS16, Proposition 5.11]). In the next proposition, we prove that an analogous property holds for NA1 and we also provide several equivalent characterizations.

Proposition 4.3.

The following are equivalent:

  1. (i)

    the NA1 condition holds;

  2. (ii)

    there does not exist a strategy πΘ^\pi\in\widehat{\Theta} satisfying (4.2);

  3. (iii)

    for every t=1,,Tt=1,\ldots,T and ζL+0(t)\zeta\in L^{0}_{+}(\mathcal{F}_{t}), vt1(ζ)=0{\rm v}_{t-1}(\zeta)=0 a.s. implies ζ=0\zeta=0 a.s.;

  4. (iv)

    arb,tΘ^t=\mathcal{I}_{{\rm arb},t}\cap\widehat{\Theta}_{t}=\emptyset a.s., for all t=1,,Tt=1,\ldots,T;

  5. (v)

    Θ^t=t\widehat{\Theta}_{t}=\mathcal{L}^{\perp}_{t} a.s., for all t=1,,Tt=1,\ldots,T;

  6. (vi)

    the set Θtt\Theta_{t}\cap\mathcal{L}_{t} is a.s. bounded (and, hence, compact), for all t=1,,Tt=1,\ldots,T.

Proof.

(i)(iii)(i)\Rightarrow(iii): by way of contradiction, assume that NA1 holds and suppose that, for some t=1,,Tt=1,\ldots,T, there exists ζL+0(t)\zeta\in L^{0}_{+}(\mathcal{F}_{t}) such that vt1(ζ)=0{\rm v}_{t-1}(\zeta)=0 a.s. and P(ζ>0)>0P(\zeta>0)>0. In this case, for every v>0v>0, one can find hL0(t1;Θt)h\in L^{0}(\mathcal{F}_{t-1};\Theta_{t}) such that v(1+h,Rt)ζv(1+\langle h,R_{t}\rangle)\geq\zeta a.s. Define then the strategy π=(πs)s=1,,T\pi=(\pi_{s})_{s=1,\ldots,T} by πs:=h\pi_{s}:=h if s=ts=t and πs:=0\pi_{s}:=0 otherwise. With this definition, it holds that πΘ\pi\in\Theta and VTπ(v)=v(1+h,Rt)ζV^{\pi}_{T}(v)=v(1+\langle h,R_{t}\rangle)\geq\zeta a.s., contradicting the validity of NA1.
(iii)(iv)(iii)\Rightarrow(iv): we adapt to the present setting the arguments of [KK07, Section 5]. By way of contradiction, assume that (iii)(iii) holds and let P(arb,tΘ^t)>0P(\mathcal{I}_{{\rm arb},t}\cap\widehat{\Theta}_{t}\neq\emptyset)>0, for some t=1,,Tt=1,\ldots,T. For each nn\in\mathbb{N}, define the t1\mathcal{F}_{t-1}-measurable random set

arb,tn:={πd:π,z0 for all z𝒮t and 𝔼[π,Rt1+π,Rt|t1]1/n}arb,t.\mathcal{I}^{n}_{{\rm arb},t}:=\left\{\pi\in\mathbb{R}^{d}:\langle\pi,z\rangle\geq 0\text{ for all $z\in\mathcal{S}_{t}$ and }\mathbb{E}\left[\frac{\langle\pi,R_{t}\rangle}{1+\langle\pi,R_{t}\rangle}\bigg{|}\mathcal{F}_{t-1}\right]\geq 1/n\right\}\subset\mathcal{I}_{{\rm arb},t}.

We have that arb,tΘ^t\mathcal{I}_{{\rm arb},t}\cap\widehat{\Theta}_{t}\neq\emptyset if and only if arb,tnΘ^t\mathcal{I}^{n}_{{\rm arb},t}\cap\widehat{\Theta}_{t}\neq\emptyset for all large enough nn\in\mathbb{N} (see [KK07, Lemma 5.1]). Hence, there exists a sufficiently large nn\in\mathbb{N} such that P(arb,tnΘ^t)>0P(\mathcal{I}^{n}_{{\rm arb},t}\cap\widehat{\Theta}_{t}\neq\emptyset)>0. It can be easily checked that the set arb,tn(ω)Θ^t(ω)\mathcal{I}^{n}_{{\rm arb},t}(\omega)\cap\widehat{\Theta}_{t}(\omega) is closed and convex, for all ωΩ\omega\in\Omega. Therefore, by [RW98, Corollary 14.6], there exists an t1\mathcal{F}_{t-1}-measurable random vector πtn:Ωd\pi^{n}_{t}:\Omega\rightarrow\mathbb{R}^{d} such that πtn(ω)arb,tn(ω)Θ^t(ω)\pi^{n}_{t}(\omega)\in\mathcal{I}^{n}_{{\rm arb},t}(\omega)\cap\widehat{\Theta}_{t}(\omega) when arb,tn(ω)Θ^t(ω)\mathcal{I}^{n}_{{\rm arb},t}(\omega)\cap\widehat{\Theta}_{t}(\omega)\neq\emptyset and πtn(ω)=0\pi^{n}_{t}(\omega)=0 when arb,tn(ω)Θ^t(ω)=\mathcal{I}^{n}_{{\rm arb},t}(\omega)\cap\widehat{\Theta}_{t}(\omega)=\emptyset. The random variable ζ:=πtn,Rt\zeta:=\langle\pi^{n}_{t},R_{t}\rangle belongs to L+0(t)L^{0}_{+}(\mathcal{F}_{t}) and satisfies P(ζ>0)>0P(\zeta>0)>0. Moreover, since πtnΘ^t\pi^{n}_{t}\in\widehat{\Theta}_{t} a.s., it holds that πtn/vΘt\pi^{n}_{t}/v\in\Theta_{t} a.s., for all v>0v>0. Noting that v(1+πtn/v,Rt)>ζv(1+\langle\pi^{n}_{t}/v,R_{t}\rangle)>\zeta a.s., this implies that vt1(ζ)=0{\rm v}_{t-1}(\zeta)=0 a.s., thus contradicting property (iii)(iii).
(ii)(iv)(ii)\Leftrightarrow(iv): this equivalence follows by the same arguments used in [FS16, Proposition 5.11], together with the construction of πtn\pi^{n}_{t} performed in the previous step of the proof.
(iv)(v)(vi)(iv)\Rightarrow(v)\Rightarrow(vi): these implications can be proved as in Proposition 2.2.
(vi)(i)(vi)\Rightarrow(i): by way of contradiction, let ξL+0(T)\xi\in L^{0}_{+}(\mathcal{F}_{T}) with P(ξ>0)>0P(\xi>0)>0 and suppose that, for all nn\in\mathbb{N}, there exists an allowed strategy πnΘ\pi^{n}\in\Theta such that VTπn(1/n)ξV^{\pi^{n}}_{T}(1/n)\geq\xi a.s. Then, it holds that 1+t=1Tpt(πtn),Rtnξ1+\prod_{t=1}^{T}\langle{\rm p}_{\mathcal{L}_{t}}(\pi^{n}_{t}),R_{t}\rangle\geq n\xi a.s., for all nn\in\mathbb{N}. Similarly as in the proof of Proposition 2.2, the fact that P(ξ>0)>0P(\xi>0)>0 contradicts the a.s. boundedness of the sets Θtt\Theta_{t}\cap\mathcal{L}_{t}, for t=1,,Tt=1,\ldots,T. ∎

Proposition 4.3 shows that, in a multi-period setting, NA1 is equivalent to the absence of arbitrarily scalable arbitrage opportunities (property (ii)(ii)) as well as to the absence of arbitrage of the first kind in each single trading period (property (iii)(iii)). Properties (iv)(iv)(vi)(vi) can be interpreted similarly to the analogous properties discussed in Section 2.2. Note also that NA1 is equivalent to no classical arbitrage if the constraint process Θc\Theta_{{\rm c}} is cone-valued (see Remark 2.3).

Remark 4.4.

Property (vi)(vi) in Proposition 4.3 implies that, for each t=1,,Tt=1,\ldots,T, there exists an t1\mathcal{F}_{t-1}-measurable random variable HtH_{t} such that πHt\|\pi\|\leq H_{t} a.s., for all πL0(t1;Θtt)\pi\in L^{0}(\mathcal{F}_{t-1};\Theta_{t}\cap\mathcal{L}_{t}). The t1\mathcal{F}_{t-1}-measurability of HtH_{t} follows from the closedness and t1\mathcal{F}_{t-1}-measurability of Θtt\Theta_{t}\cap\mathcal{L}_{t}.

4.3. Market viability and fundamental theorems

We proceed to characterize NA1 in terms of the solvability of portfolio optimization problems, extending Theorem 2.5 to the multi-period setting. In view of Proposition 4.3, the NA1 condition admits a local description. By employing a dynamic programming approach, this allows reducing a portfolio optimization problem to a sequence of one-period problems, to which we can apply techniques analogous to those used in the proof of Theorem 2.5. This approach is inspired by [RS06], where the implication (i)(ii)(i)\Rightarrow(ii) of the following theorem has been proved under no classical arbitrage for an unconstrained market. In comparison to [RS06], we allow for convex trading constraints and base our analysis on the minimal NA1 condition. Similarly as in Section 2.3, we denote by 𝒰\mathcal{U} the set of all functions U:Ω×+{}U:\Omega\times\mathbb{R}_{+}\rightarrow\mathbb{R}\cup\{-\infty\} such that U(,x)U(\cdot,x) is T\mathcal{F}_{T}-measurable and bounded from below, for every x>0x>0, and U(ω,)U(\omega,\cdot) is continuous, strictly increasing and concave, for a.e. ωΩ\omega\in\Omega.

Theorem 4.5.

The following are equivalent:

  1. (i)

    the NA1 condition holds;

  2. (ii)

    for every U𝒰U\in\mathcal{U} such that supπΘ𝔼[U+(VTπ)]<+\sup_{\pi\in\Theta}\mathbb{E}[U^{+}(V^{\pi}_{T})]<+\infty, there exists an allowed strategy πΘ\pi^{*}\in\Theta\cap\mathcal{L} such that

    𝔼[U(VTπ)]=supπΘ𝔼[U(VTπ)].\mathbb{E}\bigl{[}U(V^{\pi^{*}}_{T})\bigr{]}=\underset{\pi\in\Theta}{\sup}\,\mathbb{E}\bigl{[}U(V^{\pi}_{T})\bigr{]}.
Proof.

(i)(ii)(i)\Rightarrow(ii): suppose that NA1 holds and let U𝒰U\in\mathcal{U} be such that supπΘ𝔼[U+(VTπ)]<+\sup_{\pi\in\Theta}\mathbb{E}[U^{+}(V^{\pi}_{T})]<+\infty. Since U𝒰U\in\mathcal{U}, it holds that supπΘ𝔼[U+(xVTπ)]<+\sup_{\pi\in\Theta}\mathbb{E}[U^{+}(xV^{\pi}_{T})]<+\infty for all x0x\geq 0. The existence of an optimal strategy πΘ\pi^{*}\in\Theta\cap\mathcal{L} will be shown in a constructive way by applying dynamic programming. For all (ω,x)Ω×+(\omega,x)\in\Omega\times\mathbb{R}_{+}, define UT(ω,x):=U(ω,x)U_{T}(\omega,x):=U(\omega,x) and, for t=0,1,,T1t=0,1,\ldots,T-1,

(4.3) Ut(ω,x):=esssupπt+1L0(t;Θt+1t+1)𝔼[Ut+1(ω,x(1+πt+1,Rt+1(ω)))|t](ω),U_{t}(\omega,x):=\underset{\pi_{t+1}\in L^{0}(\mathcal{F}_{t};\Theta_{t+1}\cap\mathcal{L}_{t+1})}{\operatorname{ess\,sup}}\mathbb{E}\left[U_{t+1}\bigl{(}\omega,x(1+\langle\pi_{t+1},R_{t+1}(\omega)\rangle)\bigr{)}\big{|}\mathcal{F}_{t}\right](\omega),

taking a regular version of the conditional expectation (the existence of the conditional expectation will follow from the proof below).111111In the following, for simplicity of notation, we shall omit to denote explicitly the dependence on ω\omega in Ut(ω,x)U_{t}(\omega,x). Proceeding by backward induction, let t<Tt<T and suppose that Ut+1𝒰U_{t+1}\in\mathcal{U} and

(4.4) supπt+1L0(t;Θt+1t+1)𝔼[Ut+1+(x(1+πt+1,Rt+1))]<+, for all x0.\sup_{\pi_{t+1}\in L^{0}(\mathcal{F}_{t};\Theta_{t+1}\cap\mathcal{L}_{t+1})}\mathbb{E}\left[U^{+}_{t+1}\bigl{(}x(1+\langle\pi_{t+1},R_{t+1}\rangle)\bigr{)}\right]<+\infty,\qquad\text{ for all }x\geq 0.

These hypotheses are satisfied by assumption for t=T1t=T-1 and will be shown inductively for all t<T1t<T-1. Since the family {𝔼[Ut+1(x(1+πt+1,Rt+1))|t];πt+1L0(t;Θt+1t+1)}\{\mathbb{E}[U_{t+1}(x(1+\langle\pi_{t+1},R_{t+1}\rangle))|\mathcal{F}_{t}];\pi_{t+1}\in L^{0}(\mathcal{F}_{t};\Theta_{t+1}\cap\mathcal{L}_{t+1})\} is directed upward, for all x>0x>0 there exists a sequence (πt+1n(x))n(\pi_{t+1}^{n}(x))_{n\in\mathbb{N}} with values in Θt+1t+1\Theta_{t+1}\cap\mathcal{L}_{t+1} such that

(4.5) limn+𝔼[Ut+1(x(1+πt+1n(x),Rt+1))|t]=Ut(x) a.s.\lim_{n\rightarrow+\infty}\mathbb{E}\left[U_{t+1}\bigl{(}x(1+\langle\pi^{n}_{t+1}(x),R_{t+1}\rangle)\bigr{)}\big{|}\mathcal{F}_{t}\right]=U_{t}(x)\text{ a.s.}

As a consequence of NA1, the set Θt+1t+1\Theta_{t+1}\cap\mathcal{L}_{t+1} is closed and a.s. bounded (see Proposition 4.3). Therefore, by [FS16, Lemma 1.64], there exists a subsequence (πt+1nk(x))k(\pi^{n_{k}}_{t+1}(x))_{k\in\mathbb{N}} converging a.s. to an element π^t+1(x)L0(t;Θt+1t+1)\hat{\pi}_{t+1}(x)\in L^{0}(\mathcal{F}_{t};\Theta_{t+1}\cap\mathcal{L}_{t+1}). By the same arguments used in the proof of the implication (i)(ii)(i)\Rightarrow(ii) of Theorem 2.5 (but carried out conditionally on t\mathcal{F}_{t}, see also [RS06, Lemma 2.3]), the boundedness of Θt+1t+1\Theta_{t+1}\cap\mathcal{L}_{t+1} (see Remark 4.4), the properties of Ut+1U_{t+1} and (4.4) together imply the existence of an t+1\mathcal{F}_{t+1}-measurable integrable random variable ζt+1\zeta_{t+1} such that

(4.6) Ut+1+(x(1+πt+1,Rt+1))ζt+1, for all πt+1L0(t;Θt+1t+1).U^{+}_{t+1}\bigl{(}x(1+\langle\pi_{t+1},R_{t+1}\rangle)\bigr{)}\leq\zeta_{t+1},\qquad\text{ for all }\pi_{t+1}\in L^{0}(\mathcal{F}_{t};\Theta_{t+1}\cap\mathcal{L}_{t+1}).

Therefore, an application of Fatou’s lemma, together with the continuity of Ut+1U_{t+1}, yields that

lim supk+𝔼[Ut+1(x(1+πt+1nk(x),Rt+1))|t]\displaystyle\limsup_{k\rightarrow+\infty}\mathbb{E}\bigl{[}U_{t+1}\bigl{(}x(1+\langle\pi^{n_{k}}_{t+1}(x),R_{t+1}\rangle)\bigr{)}\big{|}\mathcal{F}_{t}\bigr{]} 𝔼[lim supk+Ut+1(x(1+πt+1nk(x),Rt+1))|t]\displaystyle\leq\mathbb{E}\Bigl{[}\limsup_{k\rightarrow+\infty}U_{t+1}\bigl{(}x(1+\langle\pi^{n_{k}}_{t+1}(x),R_{t+1}\rangle)\bigr{)}\Big{|}\mathcal{F}_{t}\Bigr{]}
=𝔼[Ut+1(x(1+π^t+1(x),Rt+1))|t].\displaystyle=\mathbb{E}\left[U_{t+1}\bigl{(}x(1+\langle\hat{\pi}_{t+1}(x),R_{t+1}\rangle)\bigr{)}\big{|}\mathcal{F}_{t}\right].

Together with (4.5), this shows that

(4.7) Ut(x)=𝔼[Ut+1(x(1+π^t+1(x),Rt+1))|t].U_{t}(x)=\mathbb{E}\bigl{[}U_{t+1}\bigl{(}x(1+\langle\hat{\pi}_{t+1}(x),R_{t+1}\rangle)\bigr{)}\big{|}\mathcal{F}_{t}\bigr{]}.

Condition (4.4) implies that Ut(x)<+U_{t}(x)<+\infty a.s., for all x0x\geq 0, thus proving the well-posedness of (4.3). Moreover, the same arguments employed in [RS06, Lemma 2.5] allow to show that the optimizer π^t+1(x)\hat{\pi}_{t+1}(x) can be chosen t(+)\mathcal{F}_{t}\otimes\mathcal{B}(\mathbb{R}_{+})-measurable.121212While [RS06] work under no classical arbitrage and do not consider trading constraints, an inspection of the proof of their Lemma 2.5 shows that only the a.s. boundedness of the set of allowed strategies is needed. In our context, the latter property holds under NA1 as a consequence of Proposition 4.3. Since the set Θt+1t+1\Theta_{t+1}\cap\mathcal{L}_{t+1} is convex and we assumed that Ut+1𝒰U_{t+1}\in\mathcal{U}, the function Ut(ω,)U_{t}(\omega,\cdot) inherits the strict increasingness and concavity of Ut+1(ω,)U_{t+1}(\omega,\cdot), for a.e. ωΩ\omega\in\Omega. Furthermore, Ut(x)𝔼[Ut+1(x)|t]U_{t}(x)\geq\mathbb{E}[U_{t+1}(x)|\mathcal{F}_{t}] and, therefore, Ut(x)U_{t}(x) is a.s. bounded from below, for every x>0x>0. In particular, this implies that Ut(x)U_{t}(x) is a.s. finite valued for all x>0x>0 and, by concavity, continuous on (0,+)(0,+\infty). To prove continuity at x=0x=0, note that Ut(0)lim infn+Ut(1/n)U_{t}(0)\leq\liminf_{n\rightarrow+\infty}U_{t}(1/n). On the other hand, using (4.7), it holds that

lim supn+Ut(1/n)\displaystyle\limsup_{n\rightarrow+\infty}U_{t}(1/n) =lim supn+𝔼[Ut+1((1/n)(1+π^t+1(1/n),Rt+1))|t]\displaystyle=\limsup_{n\rightarrow+\infty}\mathbb{E}\bigl{[}U_{t+1}\bigl{(}(1/n)(1+\langle\hat{\pi}_{t+1}(1/n),R_{t+1}\rangle)\bigr{)}\big{|}\mathcal{F}_{t}\bigr{]}
𝔼[lim supn+Ut+1((1/n)(1+π^t+1(1/n),Rt+1))|t]=𝔼[Ut+1(0)|t]=Ut(0),\displaystyle\leq\mathbb{E}\Bigl{[}\limsup_{n\rightarrow+\infty}U_{t+1}\bigl{(}(1/n)(1+\langle\hat{\pi}_{t+1}(1/n),R_{t+1}\rangle)\bigr{)}\Big{|}\mathcal{F}_{t}\Bigr{]}=\mathbb{E}[U_{t+1}(0)|\mathcal{F}_{t}]=U_{t}(0),

where, similarly as above, the inequality follows from Fatou’s lemma using (4.6) and the second equality follows from the continuity of Ut+1U_{t+1} together with the a.s. boundedness of Θt+1t+1\Theta_{t+1}\cap\mathcal{L}_{t+1}. We have thus shown that Ut𝒰U_{t}\in\mathcal{U}. To complete the proof of the inductive hypothesis, it remains to show that (4.4) holds true for each t<T1t<T-1. For every x>0x>0 and πtL0(t1;Θtt)\pi_{t}\in L^{0}(\mathcal{F}_{t-1};\Theta_{t}\cap\mathcal{L}_{t}), using repeatedly (4.7) and iterated conditioning, we have that

(4.8) 𝔼[Ut+(x(1+πt,Rt))]𝔼[U+(x(1+πt,Rt)k=1Tt(1+π^t+k(Vt+k1),Rt+k))],\mathbb{E}\bigl{[}U_{t}^{+}\bigl{(}x(1+\langle\pi_{t},R_{t}\rangle)\bigr{)}\bigr{]}\leq\mathbb{E}\biggl{[}U^{+}\biggl{(}x(1+\langle\pi_{t},R_{t}\rangle)\prod_{k=1}^{T-t}(1+\langle\hat{\pi}_{t+k}(V_{t+k-1}),R_{t+k}\rangle)\biggr{)}\biggr{]},

with Vt:=x(1+πt,Rt)V_{t}:=x(1+\langle\pi_{t},R_{t}\rangle) and Vt+k:=Vt+k1(1+π^t+k(Vt+k1),Rt+k)V_{t+k}:=V_{t+k-1}(1+\langle\hat{\pi}_{t+k}(V_{t+k-1}),R_{t+k}\rangle), for k=1,,Ttk=1,\ldots,T-t. Since supπΘ𝔼[U+(xVTπ)]<+\sup_{\pi\in\Theta}\mathbb{E}[U^{+}(xV^{\pi}_{T})]<+\infty, inequality (4.8) implies the validity of (4.4), for all t=0,1,,T2t=0,1,\ldots,T-2. Finally, the optimal strategy π=(πt)t=1,,TΘ\pi^{*}=(\pi^{*}_{t})_{t=1,\ldots,T}\in\Theta\cap\mathcal{L} is defined recursively by

πt:=π^t(Vt1π),where Vtπ=Vt1π(1+πt,Rt), for all t=1,,T, and V0π=1.\pi^{*}_{t}:=\hat{\pi}_{t}(V^{\pi^{*}}_{t-1}),\qquad\text{where }\;V^{\pi^{*}}_{t}=V^{\pi^{*}}_{t-1}(1+\langle\pi^{*}_{t},R_{t}\rangle),\text{ for all }t=1,\ldots,T,\,\text{ and }\,V^{\pi^{*}}_{0}=1.

The optimality of π\pi^{*} follows by noting that, for any strategy πΘ\pi\in\Theta,

𝔼[U(VTπ)]𝔼[UT1(VT1π)]U0(1)=𝔼[U1(V1π)]==𝔼[U(VTπ)].\mathbb{E}[U(V^{\pi}_{T})]\leq\mathbb{E}[U_{T-1}(V^{\pi}_{T-1})]\leq\ldots\leq U_{0}(1)=\mathbb{E}[U_{1}(V^{\pi^{*}}_{1})]=\ldots=\mathbb{E}[U(V^{\pi^{*}}_{T})].

(ii)(i)(ii)\Rightarrow(i): in view of Proposition 4.3, this implication follows by the same argument used for proving the implication (ii)(i)(ii)\Rightarrow(i) in Theorem 2.5. ∎

To the best of our knowledge, Theorem 4.5 provides the most general characterization of market viability for discrete-time models under random convex constraints.

In the following definition, for πΘ\pi\in\Theta, we denote by VπV^{\pi} the stochastic process (Vtπ)t=0,1,,T(V^{\pi}_{t})_{t=0,1,\ldots,T}.

Definition 4.6.

An adapted stochastic process Z=(Zt)t=0,1,,TZ=(Z_{t})_{t=0,1,\ldots,T} satisfying Zt>0Z_{t}>0 a.s. for all t=1,,Tt=1,\ldots,T and Z0=1Z_{0}=1 is said to be a supermartingale deflator if ZVπZV^{\pi} is a supermartingale, for all πΘ\pi\in\Theta. The set of all supermartingale deflators is denoted by 𝒟\mathcal{D}. An allowed strategy ρΘ\rho\in\Theta is said to be a numéraire portfolio if 1/Vρ𝒟1/V^{\rho}\in\mathcal{D}, i.e., if Vπ/VρV^{\pi}/V^{\rho} is a supermartingale.

We now prove a version of the fundamental theorem of asset pricing based on NA1 in the presence of convex constraints, extending Theorem 2.9 to the multi-period case. In a continuous-time semimartingale setting, the general version of this result is given in [KK07, Theorem 4.12]. By relying on the same approach adopted in the proof of Theorem 2.9, we can give a simple and short proof in a general discrete-time setting.

Theorem 4.7.

The following are equivalent:

  1. (i)

    the NA1 condition holds;

  2. (ii)

    𝒟\mathcal{D}\neq\emptyset;

  3. (iii)

    there exists the numéraire portfolio.

Proof.

(i)(iii)(i)\Rightarrow(iii): let t{1,,T}t\in\{1,\ldots,T\} and consider a family (fn)n(f_{n})_{n\in\mathbb{N}} of measurable functions such that fn:d(0,1]f_{n}:\mathbb{R}^{d}\rightarrow(0,1] and 𝔼[log(1+Rt)fn(Rt)]<+\mathbb{E}[\log(1+\|R_{t}\|)f_{n}(R_{t})]<+\infty, for each nn\in\mathbb{N}, and fn1f_{n}\nearrow 1 as n+n\rightarrow+\infty (see the proof of Theorem 2.9). For each nn\in\mathbb{N}, let Ut,n(ω,x):=log(x)fn(Rt(ω))U_{t,n}(\omega,x):=\log(x)f_{n}(R_{t}(\omega)), for all (ω,x)Ω×(0,+)(\omega,x)\in\Omega\times(0,+\infty). For each nn\in\mathbb{N}, it holds that Ut,n𝒰U_{t,n}\in\mathcal{U}. By Proposition 4.3, NA1 implies that Θtt\Theta_{t}\cap\mathcal{L}_{t} is a.s. bounded and, therefore, inequality (2.7) conditionally on t1\mathcal{F}_{t-1} implies that esssupπtL0(t1;Θtt)𝔼[Ut,n+(1+πt,Rt)|t1]<+\operatorname{ess\,sup}_{\pi_{t}\in L^{0}(\mathcal{F}_{t-1};\Theta_{t}\cap\mathcal{L}_{t})}\mathbb{E}[U^{+}_{t,n}(1+\langle\pi_{t},R_{t}\rangle)|\mathcal{F}_{t-1}]<+\infty a.s. Using again the boundedness of Θtt\Theta_{t}\cap\mathcal{L}_{t}, this can be shown to imply the existence of an element ρtnL0(t1;Θtt)\rho^{n}_{t}\in L^{0}(\mathcal{F}_{t-1};\Theta_{t}\cap\mathcal{L}_{t}) such that

𝔼[Ut,n(1+ρtn,Rt)|t1]=esssupπtL0(t1;Θtt)𝔼[Ut,n(1+πt,Rt)|t1] a.s.\mathbb{E}\bigl{[}U_{t,n}(1+\langle\rho_{t}^{n},R_{t}\rangle)\big{|}\mathcal{F}_{t-1}\bigr{]}=\underset{\pi_{t}\in L^{0}(\mathcal{F}_{t-1};\Theta_{t}\cap\mathcal{L}_{t})}{\operatorname{ess\,sup}}\mathbb{E}\bigl{[}U_{t,n}(1+\langle\pi_{t},R_{t}\rangle)\big{|}\mathcal{F}_{t-1}\bigr{]}\text{ a.s.}

By the same reasoning as in (2.8)-(2.9) (now conditionally on t1\mathcal{F}_{t-1}), we obtain that

𝔼[πtρtn,Rt1+ρtn,Rtfn(Rt)|t1]0 a.s., for all πtΘt and n.\mathbb{E}\left[\frac{\langle\pi_{t}-\rho_{t}^{n},R_{t}\rangle}{1+\langle\rho_{t}^{n},R_{t}\rangle}f_{n}(R_{t})\bigg{|}\mathcal{F}_{t-1}\right]\leq 0\text{ a.s.},\qquad\text{ for all }\pi_{t}\in\Theta_{t}\text{ and }n\in\mathbb{N}.

Since Θtt\Theta_{t}\cap\mathcal{L}_{t} is bounded and closed, we can assume that (ρtn)n(\rho_{t}^{n})_{n\in\mathbb{N}} converges a.s. to an element ρtL0(t1;Θtt)\rho_{t}\in L^{0}(\mathcal{F}_{t-1};\Theta_{t}\cap\mathcal{L}_{t}) as n+n\rightarrow+\infty (up to passing to a suitable subsequence, see [FS16, Lemma 1.64]). Since fn1f_{n}\nearrow 1 as n+n\rightarrow+\infty, an application of Fatou’s lemma gives that

𝔼[πtρt,Rt1+ρt,Rt|t1]0 a.s., for all πtL0(t1;Θt).\mathbb{E}\left[\frac{\langle\pi_{t}-\rho_{t},R_{t}\rangle}{1+\langle\rho_{t},R_{t}\rangle}\bigg{|}\mathcal{F}_{t-1}\right]\leq 0\text{ a.s.},\qquad\text{ for all }\pi_{t}\in L^{0}(\mathcal{F}_{t-1};\Theta_{t}).

Let π=(πt)t=1,,TΘ\pi=(\pi_{t})_{t=1,\ldots,T}\in\Theta. Then, for each t{1,,T1}t\in\{1,\ldots,T-1\}, the last inequality implies that

𝔼[VtπVtρ|t1]=Vt1πVt1ρ𝔼[1+πt,Rt1+ρt,Rt|t1]Vt1πVt1ρ a.s.,\mathbb{E}\left[\frac{V^{\pi}_{t}}{V^{\rho}_{t}}\bigg{|}\mathcal{F}_{t-1}\right]=\frac{V^{\pi}_{t-1}}{V^{\rho}_{t-1}}\,\mathbb{E}\left[\frac{1+\langle\pi_{t},R_{t}\rangle}{1+\langle\rho_{t},R_{t}\rangle}\bigg{|}\mathcal{F}_{t-1}\right]\leq\frac{V^{\pi}_{t-1}}{V^{\rho}_{t-1}}\text{ a.s.},

thus proving that the strategy ρ=(ρt)t=1,,T\rho=(\rho_{t})_{t=1,\ldots,T} corresponds to the numéraire portfolio.
(iii)(ii)(iii)\Rightarrow(ii): this implication is immediate by Definition 4.6.
(ii)(i)(ii)\Rightarrow(i): this implication follows by the same argument used in the proof of Theorem 2.9. ∎

Finally, we mention that the proof of Theorem 2.12 can be similarly extended to the multi-period case, thus providing a utility maximization proof of the fundamental theorem of asset pricing for no classical arbitrage, in the spirit of [Rog94] (see also [KaS09, Section 2.1.4]). Theorem 2.15 also admits a direct extension to the multi-period setting, with an identical statement.

References

  • [BCL19] J. Baptiste, L. Carassus, and E. Lépinette. Pricing without martingale measure. Preprint (available at https://arxiv.org/abs/1807.04612), 2019.
  • [BF17] E. Barucci and C. Fontana. Financial Markets Theory: Equilibrium, Efficiency and Information. Springer, London, second edition, 2017.
  • [BZ17] E. Bayraktar and Z. Zhou. On arbitrage and duality under model uncertainty and portfolio constraints. Mathematical Finance, 27(4):988–1012, 2017.
  • [Bec01] D. Becherer. The numeraire portfolio for unbounded semimartingales. Finance and Stochastics, 5(3):327–341, 2001.
  • [Ber74] D. Bertsekas. Necessary and sufficient conditions for existence of an optimal portfolio. Journal of Economic Theory, 8(2):235–247, 1974.
  • [Bra97] W. Brannath. No Arbitrage and Martingale Measures in Option Pricing. PhD thesis, University of Vienna, 1997.
  • [BP03] H. Bühlmann and E. Platen. A discrete time benchmark approach for insurance and finance. ASTIN Bulletin, 33(2):153–172, 2003.
  • [CLM97] J.Y. Campbell, A.W. Lo, and A.C. MacKinlay. The Econometrics of Financial Markets. Princeton University Press, Princeton (NJ), 1997.
  • [CPT01] L. Carassus, H. Pham, and N. Touzi. No arbitrage in discrete time under portfolio constraints. Mathematical Finance, 11(3):315–329, 2001.
  • [CCFM17] H.N. Chau, A. Cosso, C. Fontana, and O. Mostovyi. Optimal investment with intermediate consumption under no unbounded profit with bounded risk. Journal of Applied Probability, 54(3):710–719, 2017.
  • [CDM15] T. Choulli, J. Deng, and J. Ma. How non-arbitrage, viability and numéraire portfolio are related. Finance and Stochastics, 19:719–741, 2015.
  • [CK95] G. Connor and R.A. Korajczyk. The arbitrage pricing theory and multifactor models of asset returns. In R.A. Jarrow, V. Maksimovic, and W.T. Ziemba, editors, Handbooks in Operations Research and Management Science, volume 9 (Finance), pages 87–144. North-Holland, Amsterdam, 1995.
  • [DMW90] R.C. Dalang, A. Morton, and W. Willinger. Equivalent martingale measures and no-arbitrage in stochastic securities market models. Stochastics and Stochastic Reports, 29(2):185–201, 1990.
  • [Dav97] M.H.A. Davis. Option pricing in incomplete markets. In M.A.H. Dempster and S.R. Pliska, editors, Mathematics of Derivative Securities, pages 227–254. Cambridge University Press, Cambridge, 1997.
  • [DR87] P.H. Dybvig and S.A. Ross. Arbitrage. In J. Eatwell, M. Milgate, and P. Newman, editors, The New Palgrave Dictionary of Economics, volume 1, pages 100–106. Macmillan, London, 1987.
  • [ES01] H. Elsinger and M. Summer. Arbitrage and portfolio choice with financial constraints. Österreichische Nationalbank working paper no. 49, 2001.
  • [EST04] I.V. Evstignveev, K. Schürger, and M.I. Taksar. On the fundamental theorem of asset pricing: random constraints and bang-bang no-arbitrage criteria. Mathematical Finance, 14(2):201–221, 2004.
  • [FK09] R. Fernholz and I. Karatzas. Stochastic portfolio theory: an overview. In A. Bensoussan and Q. Zhang, editors, Mathematical Modeling and Numerical Methods in Finance, volume XV of Handbook of Numerical Analysis, pages 89–167, Oxford, 2009. North-Holland.
  • [FS16] H. Föllmer and A. Schied. Stochastic Finance: An Introduction in Discrete Time. De Gruyter, Berlin - New York, fourth edition, 2016.
  • [Fon15] C. Fontana. Weak and strong no-arbitrage conditions for continuous financial markets. International Journal of Theoretical and Applied Finance, 18(1):1550005, 2015.
  • [FR13] C. Fontana and W.J. Runggaldier. Diffusion-based models for financial markets without martingale measures. In F. Biagini, A. Richter, and H. Schlesinger, editors, Risk Measures and Attitudes, EAA Series, pages 45–81. Springer, London, 2013.
  • [HK79] J.M. Harrison and D.M. Kreps. Martingales and arbitrage in multiperiod securities markets. Journal of Economic Theory, 20(3):381–408, 1979.
  • [HP81] J.M. Harrison and S.R. Pliska. Martingales and stochastic integrals in the theory of continuous trading. Stochastic Processes and their Applications, 11(3):215–260, 1981.
  • [HWY92] S.-W. He, J.-G. Wang, and J.-A. Yan. Semimartingale Theory and Stochastic Calculus. Science Press - CRC Press, Beijing, 1992.
  • [Ing87] J.E. Ingersoll. Theory of Financial Decision Making. Rowman & Littlefield, Bollman Place, 1987.
  • [JK95] E. Jouini and H. Kallal. Arbitrage in securities markets with short-sales constraints. Mathematical Finance, 5(3):197–232, 1995.
  • [KaS09] Y. Kabanov and M. Safarian. Markets with Transaction Costs: Mathematical Theory. Springer, Berlin - Heidelberg, 2009.
  • [KK07] I. Karatzas and C. Kardaras. The numeraire portfolio in semimartingale financial models. Finance and Stochastics, 11(4):447–493, 2007.
  • [Kar09] C. Kardaras. No-free-lunch equivalences for exponential Lévy models under convex constraints on investment. Mathematical Finance, 19(2):161–187, 2009.
  • [Kar10] C. Kardaras. Finitely additive probabilities and the fundamental theorem of asset pricing. In C. Chiarella and A. Novikov, editors, Contemporary Quantitative Finance: Essays in Honour of Eckhard Platen, pages 19–34. Springer, Berlin - Heidelberg, 2010.
  • [KST16] I. Klein, T. Schmidt, and J. Teichmann. No arbitrage theory for bond markets. In J. Kallsen and A. Papapantoleon, editors, Advanced Modelling in Mathematical Finance: In Honour of Ernst Eberlein, volume 189 of Springer Proceedings in Mathematics & Statistics, pages 381–412. Springer, Cham, 2016.
  • [KP00] P.-F. Köehl and H. Pham. Sublinear price functionals under portfolio constraints. Journal of Mathematical Economics, 33(3):339–351, 2000.
  • [KS99] R. Korn and M. Schäl. On value preserving and growth optimal portfolios. Mathematical Methods of Operations Research, 50(2):189–218, 1999.
  • [KS09] R. Korn and M. Schäl. The numeraire portfolio in discrete time: existence, related concepts and applications. In H. Albrecher, W.J. Runggaldier, and W. Schachermayer, editors, Advanced Financial Modelling, volume 8 of Radon Series on Computational and Applied Mathematics, pages 303–326. De Gruyter, Berlin - New York, 2009.
  • [LW00] M. Loewenstein and G.A. Willard. Local martingales, arbitrage, and viability: free snacks and cheap thrills. Economic Theory, 16:135–161, 2000.
  • [Nap03] C. Napp. The Dalang-Morton-Willinger theorem under cone constraints. Journal of Mathematical Economics, 31(1-2):111–126, 2003.
  • [Nut16] M. Nutz. Utility maximization under model uncertainty in discrete time. Mathematical Finance, 26(2):252–268, 2016.
  • [Pha00] H. Pham. Dynamic Lp{L^{p}}-hedging in discrete time under cone constraints. SIAM Journal on Control and Optimization, 38(3):665–682, 2000.
  • [PT99] H. Pham and N. Touzi. The fundamental theorem of asset pricing under cone constraints. Journal of Mathematical Economics, 31(2):265–279, 1999.
  • [PH06] E. Platen and D. Heath. A Benchmark Approach to Quantitative Finance. Springer, Berlin - Heidelberg, 2006.
  • [RS05] M. Rásony and L. Stettner. On utility maximization in discrete-time financial markets models. Annals of Applied Probability, 15(2):1367–1395, 2005.
  • [RS06] M. Rásony and L. Stettner. On the existence of optimal portfolios for the utility maximization problem in discrete time financial market models. In Y. Kabanov, R. Liptser, and J. Stoyanov, editors, From Stochastic Calculus to Mathematical Finance, pages 589–608. Springer, Berlin - Heidelberg, 2006.
  • [Roc70] T. Rockafellar. Convex Analysis. Princeton University Press, Princeton (NJ), 1970.
  • [RW98] T. Rockafellar and R. Wets. Variational Analysis. Springer, Berlin - Heidelberg, 1998.
  • [Rog94] L.C.G. Rogers. Equivalent martingale measures and no-arbitrage. Stochastics and Stochastic Reports, 51(1–2):41–50, 1994.
  • [Rok05] D.B. Rokhlin. An extended version of the Dalang-Morton-Willinger theorem under portfolio constraints. Theory of Probability and its Applications, 49(3):429–443, 2005.
  • [Ros76] S.A. Ross. The arbitrage theory of capital asset pricing. Journal of Economic Theory, 13(3):341–360, 1976.
  • [Ros77] S.A. Ross. Return, risk, and arbitrage. In I. Friend and J. Bicksler, editors, Risk and Return in Finance, pages 189–217. Ballinger, Cambridge, 1977.
  • [Ros78] S.A. Ross. A simple approach to the valuation of risky streams. Journal of Business, 51(3):453–475, 1978.
  • [Sch10] W. Schachermayer. The fundamental theorem of asset pricing. In R. Cont, editor, Encyclopedia of Quantitative Finance, pages 792–801. Wiley, Chichester, 2010.