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

The McCormick martingale optimal transport

Erhan Bayraktar Department of Mathematics, University of Michigan, Ann Arbor, Email: [email protected]. Erhan Bayraktar is partially supported by the National Science Foundation under grant DMS-2106556 and by the Susan M. Smith chair.    Bingyan Han Thrust of Financial Technology, The Hong Kong University of Science and Technology (Guangzhou), Email: [email protected]. Bingyan Han is partially supported by The Hong Kong University of Science and Technology (Guangzhou) Start-up Fund G0101000197 and the Guangzhou-HKUST(GZ) Joint Funding Program (No. 2024A03J0630).    Dominykas Norgilas Department of Mathematics, North Carolina State University, Email: [email protected].
Abstract

Martingale optimal transport (MOT) often yields broad price bounds for options, constraining their practical applicability. In this study, we extend MOT by incorporating causality constraints among assets, inspired by the nonanticipativity condition of stochastic processes. However, this introduces a computationally challenging bilinear program. To tackle this issue, we propose McCormick relaxations to ease the bicausal formulation and refer to it as McCormick MOT. The primal attainment and strong duality of McCormick MOT are established under standard assumptions. Empirically, using the lower and upper bounds derived from marginal constraints, the McCormick relaxations reduce the price gap by an average of 1% for stocks with liquid option markets and 4% for those with moderately liquid markets. When tighter bounds on probability masses are applied, the average reduction increases to 12.66%.
Keywords: Martingale optimal transport, causal optimal transport, bilinear constraints, McCormick relaxations, strong duality
Mathematics Subject Classification: 91G20, 60G42, 90C46.

1 Introduction

In the realm of option pricing, a crucial task is to establish model-independent arbitrage-free price bounds for derivatives. These bounds serve a direct application in identifying potential arbitrage opportunities in traded options. Hobson and Neuberger, (2012) pioneered the study on robust bounds for forward start options, while Martingale Optimal Transport (MOT), introduced by Beiglböck et al., (2013), presented a comprehensive framework utilizing tools from Monge–Kantorovich mass transport. The strong duality result has a financial interpretation, yielding sub/super hedging portfolios. Furthering the MOT research line, Beiglböck et al., (2017, 2019) considered dual attainment problems. Beiglböck and Juillet, (2016) identified a martingale coupling reminiscent of the classic monotone quantile coupling. Several numerical schemes for MOT were proposed by Guo and Obłój, (2019); Eckstein et al., (2021), employing techniques such as linear programming (LP) or neural networks. Investigations into MOT stability were carried out by Backhoff-Veraguas and Pammer, (2022); Brückerhoff and Juillet, (2022); Wiesel, (2023).

A significant hurdle in MOT arises from the wide price bounds, thereby constraining their practical applicability. This claim is empirically substantiated by Beiglböck et al., (2013, Figure 1) and Eckstein et al., (2021, Tables 4.2 and 4.3). Achieving more precise bounds typically requires the incorporation of additional information into the modeling framework. For example, Lütkebohmert and Sester, (2019) argued that tighter bounds can be derived when more information about the variance of returns is available, although at the cost of sacrificing model independence. Fahim and Huang, (2016) enhanced precision of price bounds by integrating liquid exotic options into the hedging portfolio. Furthermore, Eckstein et al., (2021, Section 4) refined the bounds by assuming knowledge of the marginals at earlier maturities.

In this study, we introduce another methodology inspired by recent strides in causal optimal transport (COT) (Lassalle,, 2013; Backhoff-Veraguas et al.,, 2017). COT posits a crucial condition: given the past of one process XX, the past of another process YY should be independent of the future of XX under the transport plan. Essentially, COT imposes the nonanticipativity condition. The versatility of COT is evidenced by its applications in various domains such as mathematical finance (Backhoff-Veraguas et al.,, 2020), mean field games (Acciaio et al.,, 2021; Backhoff-Veraguas and Zhang,, 2023), and stochastic optimization (Pflug and Pichler,, 2012, 2014; Acciaio et al.,, 2020).

In Section 3, we present the causality constraint in a two-asset setting. An essential challenge arises as the MOT incorporating causality transforms into a bilinear program that is non-convex. This inherent difficulty stems from the fact that only marginals at each maturity are known in MOT. However, causality requires conditional kernels that rely on the joint distribution across different maturities, detailed in Proposition 3.2. Consequently, to overcome this difficulty, we relax the original problem and consider the McCormick envelopes (McCormick,, 1976) for the bilinear equality constraints. We refer to it as the McCormick MOT. McCormick relaxations offer a straightforward implementation and are widely employed in optimization software, while we note that various alternative algorithms for bilinear problems exist; see Audet et al., (2000); Sherali and Fraticelli, (2002); Anstreicher, (2009).

In discrete scenarios, McCormick MOT is a finite-dimensional LP problem and thus straightforward. Our main technical contributions are for the continuous case with absolutely continuous densities. To derive McCormick envelopes, we assume that lower and upper bounds for densities are available, which are referred to as capacity constraints in Korman and McCann, (2015). The primal attainment is established in Theorem 4.5 after selecting a suitable weak topology. Analogously to the classic MOT, a natural question is the strong duality of the McCormick MOT. Lemma 4.6 and Theorem 4.7 prove the relevant result under standard assumptions in the MOT literature. The main idea underlying these proofs remains rooted in the Fenchel-Rockafellar duality theorem.

In our numerical study, we introduce a calibration method for risk-neutral densities that ensures convex order while addressing bid-ask spreads and multiple tenors. Given the discrete nature of empirical data, our approach formulates a finite-dimensional LP problem, offering a straightforward optimization process. We observe that the improvement in price bounds depends on the specific option payoffs and the liquidity of European options on the underlying assets. McCormick MOT proves to be effective for path-dependent payoffs, where the temporal structure plays a more pronounced role. First, using the lower and upper bounds that naturally arise from marginal constraints, the McCormick relaxations reduce the price gap by an average of 1% for stocks with liquid option markets and 4% for those with moderately liquid markets. When tighter bounds on probability masses are applied, the McCormick MOT achieves an average reduction of 12.66%. Importantly, our methodology is universal since it integrates seamlessly with other knowledge available to the agent. The source code is publicly available at https://github.com/hanbingyan/McCormick.

The rest of the paper is organized as follows. Section 2 introduces the notation and formulation of MOT. Section 3 presents the motivation and several implications of the causality constraint. In Section 4, we demonstrate the McCormick relaxations in both discrete and continuous cases. Section 5 shows the improvements of the price bounds with an illustrative example and empirical data. The appendix gives proofs of the main results.

2 Problem formulation

We consider a financial market comprising two risky assets, denoted XX and YY, while the extension to include more assets is straightforward. Suppose that the market has European options on each of the two stocks for the maturities {T1,,TN}\{T_{1},\ldots,T_{N}\}, where N2N\geq 2 and the maturities are not necessarily evenly spaced. Let Xt:=XTtX_{t}:=X_{T_{t}} and Yt:=YTtY_{t}:=Y_{T_{t}} represent the prices of risky assets at maturity TtT_{t}. The range of the first risky asset XX at time TtT_{t} is denoted as 𝒳t{\mathcal{X}}_{t}, assumed to be a closed subset of \mathbb{R}. Consequently, 𝒳:=𝒳1:N=𝒳1××𝒳N{\mathcal{X}}:={\mathcal{X}}_{1:N}={\mathcal{X}}_{1}\times\ldots\times{\mathcal{X}}_{N} forms a closed subset of N\mathbb{R}^{N}. The set of all Borel probability measures on 𝒳t{\mathcal{X}}_{t} is denoted as 𝒫(𝒳t){\mathcal{P}}({\mathcal{X}}_{t}). Similarly, for the second risky asset YY, we introduce another closed set 𝒴=𝒴1:N=𝒴1××𝒴N{\mathcal{Y}}={\mathcal{Y}}_{1:N}={\mathcal{Y}}_{1}\times\ldots\times{\mathcal{Y}}_{N}, with the corresponding notations defined analogously.

In this study, we assume that individual risk-neutral probabilities denoted Xtμt𝒫(𝒳t)X_{t}\sim\mu_{t}\in{\mathcal{P}}({\mathcal{X}}_{t}) and Ytνt𝒫(𝒴t)Y_{t}\sim\nu_{t}\in{\mathcal{P}}({\mathcal{Y}}_{t}), are known. Specifically, if the prices of European options are available for all strikes, Breeden and Litzenberger, (1978) provided a formula for the risk-neutral probability density. In cases involving a finite number of strikes, a linear interpolation method is described in Davis and Hobson, (2007). Additionally, an alternative cubic-spline method has been proposed by Fengler, (2009) for finitely many options prices.

However, the joint distribution among different maturities or assets is typically unknown. Let Π(μ¯)=Π(μ1,,μN)\Pi(\bar{\mu})=\Pi(\mu_{1},\ldots,\mu_{N}) represent the set of couplings μ𝒫(𝒳1:N)\mu\in{\mathcal{P}}({\mathcal{X}}_{1:N}) with μt\mu_{t} as marginals for each tt. Similarly, Π(ν¯)=Π(ν1,,νN)\Pi(\bar{\nu})=\Pi(\nu_{1},\ldots,\nu_{N}) is defined. Consider Π(μ¯,ν¯)\Pi(\bar{\mu},\bar{\nu}) as the set of couplings π𝒫(𝒳1:N×𝒴1:N)\pi\in{\mathcal{P}}({\mathcal{X}}_{1:N}\times{\mathcal{Y}}_{1:N}) with μt\mu_{t} and νt\nu_{t} as marginals. The model-independent and arbitrage-free framework in Beiglböck et al., (2013) involves examining a measure πΠ(μ¯,ν¯)\pi\in\Pi(\bar{\mu},\bar{\nu}) that satisfies the martingale condition:

𝔼π[Xt+1|X1:t,Y1:t]=Xt,𝔼π[Yt+1|X1:t,Y1:t]=Yt,1tN1.\mathbb{E}_{\pi}[X_{t+1}|X_{1:t},Y_{1:t}]=X_{t},\quad\mathbb{E}_{\pi}[Y_{t+1}|X_{1:t},Y_{1:t}]=Y_{t},\quad 1\leq t\leq N-1. (2.1)

For the sake of simplicity, we assume zero interest rates and no dividends in (2.1), while our empirical results utilize forward prices to incorporate these factors. Denote (μ¯,ν¯)Π(μ¯,ν¯){\mathcal{M}}(\bar{\mu},\bar{\nu})\subset\Pi(\bar{\mu},\bar{\nu}) as the set of all martingale measures satisfying marginal constraints. Suppose {μt}t=1N\{\mu_{t}\}^{N}_{t=1} possess identical finite first moments and increase in convex order in tt, that is, for all convex functions ff and 1tN11\leq t\leq N-1,

f𝑑μtf𝑑μt+1.\displaystyle\int fd\mu_{t}\leq\int fd\mu_{t+1}.

Denote the convex order condition as μtcxμt+1\mu_{t}\preceq_{cx}\mu_{t+1}. Similarly, assume {νt}t=1N\{\nu_{t}\}^{N}_{t=1} have the same finite moments and νtcxνt+1\nu_{t}\preceq_{cx}\nu_{t+1}. By Strassen, (1965), the assumptions above are necessary and sufficient conditions for (μ¯,ν¯){\mathcal{M}}(\bar{\mu},\bar{\nu})\neq\emptyset.

In the subsequent discussion, we examine an exotic option with the payoff function c(x1:N,y1:N)c(x_{1:N},y_{1:N}). The MOT framework (Beiglböck et al.,, 2013; Eckstein et al.,, 2021) investigates model-free bounds for this exotic option:

supπ(μ¯,ν¯)c(x1:N,y1:N)π(dx1:N,dy1:N) and infπ(μ¯,ν¯)c(x1:N,y1:N)π(dx1:N,dy1:N).\sup_{\pi\in{\mathcal{M}}(\bar{\mu},\bar{\nu})}\int c(x_{1:N},y_{1:N})\pi(dx_{1:N},dy_{1:N})\quad\text{ and }\inf_{\pi\in{\mathcal{M}}(\bar{\mu},\bar{\nu})}\int c(x_{1:N},y_{1:N})\pi(dx_{1:N},dy_{1:N}). (2.2)

(2.2) is known as the primal formulation. The dual problem unveils semi-static hedging strategies as outlined in Beiglböck et al., (2013, Theorem 1.1).

In the absence of modeling assumptions, the gap between lower and upper bounds in MOT can be substantial, thereby limiting its practical applicability. A common approach to address this challenge involves incorporating additional information. For instance, Fahim and Huang, (2016) augmented the model with market prices of specific exchange-listed exotic options, such as standardized digital and barrier options. Another strategy entails including information related to volatility. Joint calibration on SPX and VIX options leverages the structural link between the S&P 500 and VIX indices; see Guyon et al., (2017); De Marco and Henry-Labordere, (2015) for more details. Additionally, Bayraktar et al., (2021) examined MOT examples with volatility constraints, while Lütkebohmert and Sester, (2019) considered supplementary information regarding the variance of returns.

The conventional optimal transport framework fails to account for the temporal structure and information flow inherent in time series data. To address this limitation, recent research has introduced COT (Lassalle,, 2013; Backhoff-Veraguas et al.,, 2017). When dealing with options on multiple assets, we demonstrate that the nonanticipativity (causality) condition provides an alternative means of tightening bounds, typically without requiring new data or estimation. Section 3 presents the causality constraint and its implications for MOT.

3 Motivation and implication of causality

3.1 Nonanticipativity condition

Under a risk-neutral measure π\pi with discrete-time processes, the underlying assets XX and YY are typically represented as:

Xt+1Xt=XtWt+1,Yt+1Yt=YtZt+1.X_{t+1}-X_{t}=X_{t}W_{t+1},\qquad Y_{t+1}-Y_{t}=Y_{t}Z_{t+1}. (3.1)

Here, {Wt}t=1N1\{W_{t}\}^{N-1}_{t=1} and {Zt}t=1N1\{Z_{t}\}^{N-1}_{t=1} denote random returns for XX and YY, respectively, which are the only sources of randomness in the system (3.1). A standard assumption posits that {Wt}t=1N1\{W_{t}\}^{N-1}_{t=1} are independent and identically distributed, as commonly imposed in binomial tree models. {Zt}t=1N1\{Z_{t}\}^{N-1}_{t=1} also satisfy the same assumption. However, it should be noted that for each time 1tN11\leq t\leq N-1, WtW_{t} and ZtZ_{t} are allowed to be correlated. For example, a parsimonious assumption is that the correlation between WtW_{t} and ZtZ_{t} is a constant. Under these conditions, Xt+1X_{t+1} is independent of YtY_{t}, conditional on XtX_{t}. This equivalently implies that π(dxt+1|xt,yt)=π(dxt+1|xt)\pi(dx_{t+1}|x_{t},y_{t})=\pi(dx_{t+1}|x_{t}), since knowledge of YtY_{t} does not aid in predicting Wt+1W_{t+1} once XtX_{t} is known. It is crucial to emphasize that YtY_{t} and Xt+1X_{t+1} are not independent if XtX_{t} is unknown.

Aligned with the nonanticipativity condition that originates from stochastic difference (or differential) equations, Lassalle, (2013) imposes that a transport plan π\pi should satisfy

π(dxt+1|x1:t,y1:t)=π(dxt+1|x1:t),t=1,,N1,π-a.s.,\pi(dx_{t+1}|x_{1:t},y_{1:t})=\pi(dx_{t+1}|x_{1:t}),\quad t=1,...,N-1,\quad\text{$\pi$-a.s.}, (3.2)

which is equivalent to

π(dyt|x1:N)=π(dyt|x1:t),t=1,,N1,π-a.s.,\pi(dy_{t}|x_{1:N})=\pi(dy_{t}|x_{1:t}),\quad t=1,...,N-1,\quad\text{$\pi$-a.s.}, (3.3)

see Eckstein and Pammer, (2024, Remark 2.2). The property (3.2) (or (3.3)) is known as the causality condition. A transport plan that satisfies (3.2) (or (3.3)) is termed causal. If we interchange the positions of XX and YY the following condition holds:

π(dxt|y1:N)=π(dxt|y1:t),t=1,,N1,π-a.s.,\pi(dx_{t}|y_{1:N})=\pi(dx_{t}|y_{1:t}),\quad t=1,...,N-1,\quad\text{$\pi$-a.s.}, (3.4)

then the transport plan is anticausal. Couplings that are both causal and anticausal are referred to as bicausal. For later use, denote bc(μ¯,ν¯)(μ¯,ν¯){\mathcal{M}}_{bc}(\bar{\mu},\bar{\nu})\subset{\mathcal{M}}(\bar{\mu},\bar{\nu}) as the set of all bicausal and martingale transport plans, with μt\mu_{t} and νt\nu_{t} as marginals for each tt.

Motivated by the nonanticipativity condition, we propose to further restrict the MOT problem (2.2) by considering bicausal plans:

supπbc(μ¯,ν¯)c(x1:N,y1:N)π(dx1:N,dy1:N) and infπbc(μ¯,ν¯)c(x1:N,y1:N)π(dx1:N,dy1:N).\sup_{\pi\in{\mathcal{M}}_{bc}(\bar{\mu},\bar{\nu})}\int c(x_{1:N},y_{1:N})\pi(dx_{1:N},dy_{1:N})\,\text{ and }\inf_{\pi\in{\mathcal{M}}_{bc}(\bar{\mu},\bar{\nu})}\int c(x_{1:N},y_{1:N})\pi(dx_{1:N},dy_{1:N}). (3.5)

Before delving into the main results, we discuss some properties for the bicausal MOT (3.5). The first consequence of the bicausal condition is the equivalence of the martingale property under individual filtrations and the joint filtration. The proof of Proposition 3.1 directly applies the causality condition (3.2) and is therefore omitted.

Proposition 3.1.

If π\pi is a bicausal transport plan in Π(μ¯,ν¯)\Pi(\bar{\mu},\bar{\nu}), then the martingale condition (2.1) under the filtration generated by (X,Y)(X,Y) is equivalent to the following martingale condition (3.6) with the filtration generated by each asset:

𝔼π[Xt+1|X1:t]=Xt,𝔼π[Yt+1|Y1:t]=Yt,1tN1.\mathbb{E}_{\pi}[X_{t+1}|X_{1:t}]=X_{t},\qquad\mathbb{E}_{\pi}[Y_{t+1}|Y_{1:t}]=Y_{t},\quad 1\leq t\leq N-1. (3.6)

3.2 Non-convexity

When the joint distributions and, consequently, conditional kernels μ(dxt+1|x1:t)\mu(dx_{t+1}|x_{1:t}) are provided, the causality condition (3.2) (or (3.3)) imposes a linear constraint on π\pi; see Backhoff-Veraguas et al., (2017, Proposition 2.4 (3)). However, this is no longer applicable when the conditional kernels are not predetermined. The corresponding equivalent condition is outlined in Proposition 3.2. The proof follows a rationale similar to that of Backhoff-Veraguas et al., (2017, Proposition 2.4 (3)), and we include it in the Appendix for the sake of completeness.

Proposition 3.2.

πΠ(μ¯,ν¯)\pi\in\Pi(\bar{\mu},\bar{\nu}) is causal if and only if

ht(y1:t)[gt(x1:N)gt(x1:t,x¯t+1:N)π(dx¯t+1:N|x1:t)]π(dx1:N,dy1:N)=0,\int h_{t}(y_{1:t})\left[g_{t}(x_{1:N})-\int g_{t}(x_{1:t},\bar{x}_{t+1:N})\pi(d\bar{x}_{t+1:N}|x_{1:t})\right]\pi(dx_{1:N},dy_{1:N})=0, (3.7)

for all 1tN1\leq t\leq N, htCb(𝒴1:t)h_{t}\in C_{b}({\mathcal{Y}}_{1:t}), and gtCb(𝒳1:N)g_{t}\in C_{b}({\mathcal{X}}_{1:N}).

To observe that the causality constraint is no longer linear in π\pi, one can think about the discrete case. The conditional mass function π(xt+1:N|x1:t)=π(x1:t,xt+1:N)/π(x1:t)\pi(x_{t+1:N}|x_{1:t})=\pi(x_{1:t},x_{t+1:N})/\pi(x_{1:t}), where the denominator π(x1:t)>0\pi(x_{1:t})>0 is not uniquely determined. In the subsequent section, we demonstrate that this constraint is bilinear in π\pi and non-convex. This characteristic poses a primary challenge in the bicausal MOT and makes the program difficult to solve. Consequently, we propose adopting the McCormick relaxation (McCormick,, 1976) for the bilinear constraint. The resulting relaxed problem is convex, offering a computationally more tractable approach while maintaining the capacity to enhance bounds in various scenarios.

4 McCormick relaxation

4.1 The discrete case

To elucidate the main concept, we examine the case with discrete marginals, interpreting π(x1:N,y1:N)\pi(x_{1:N},y_{1:N}) as probability mass functions. The causality constraint (3.2) (or (3.3)) can be expressed equivalently as

π(x1:N,yt)π(x1:t)=π(x1:t,yt)π(x1:N),t=1,,N1.\pi(x_{1:N},y_{t})\pi(x_{1:t})=\pi(x_{1:t},y_{t})\pi(x_{1:N}),\quad t=1,\ldots,N-1. (4.1)

We stress that (4.1) holds pointwise, while (3.2) (or (3.3)) holds π\pi-a.s. Indeed, if π(x1:N)>0\pi(x_{1:N})>0, then π(x1:t)>0\pi(x_{1:t})>0. Both conditional probabilities in (3.3) are well-defined. Then (4.1) holds. If π(x1:N)=0\pi(x_{1:N})=0, then π(dyt|x1:N)\pi(dy_{t}|x_{1:N}) is not defined. (3.3) is not required to hold. However, we note that π(x1:N,yt)=0\pi(x_{1:N},y_{t})=0 since π(x1:N)=0\pi(x_{1:N})=0. Therefore, (4.1) is valid as π(x1:N,yt)π(x1:t)=π(x1:t,yt)π(x1:N)=0\pi(x_{1:N},y_{t})\pi(x_{1:t})=\pi(x_{1:t},y_{t})\pi(x_{1:N})=0.

(4.1) essentially constitutes a bilinear equality constraint on π\pi. Consequently, it is computationally challenging to find exact upper/lower bounds in the bicausal MOT problem (3.5). In the optimization literature, bilinear programming falls under quadratically constrained quadratic programming (QCQP). Several algorithms, including branch and cut with reformulation-linearization techniques (Audet et al.,, 2000), and semi-definite programming relaxations (Anstreicher,, 2009; Sherali and Fraticelli,, 2002), have been proposed. In this study, we opt for the classic McCormick relaxation in McCormick, (1976) for several reasons. It is widely employed as a built-in routine in optimization software and is also straightforward to implement manually. Moreover, the dual formulation remains tractable, allowing for the demonstration of the impact of causality constraints on hedging strategies.

McCormick relaxations construct the convex and concave envelopes as follows. The bilinear term on the left-hand side of (4.1) is substituted with w1(x1:N,yt)=π(x1:N,yt)π(x1:t)w_{1}(x_{1:N},y_{t})=\pi(x_{1:N},y_{t})\pi(x_{1:t}). For probability π(x1:N,yt)\pi(x_{1:N},y_{t}), we suppose the upper bound UN,t(x1:N,yt)U_{N,t}(x_{1:N},y_{t}) and lower bound LN,t(x1:N,yt)L_{N,t}(x_{1:N},y_{t}) are known. The corresponding upper and lower bounds for π(x1:t)\pi(x_{1:t}) are denoted as Ut,0(x1:t)U_{t,0}(x_{1:t}) and Lt,0(x1:t)L_{t,0}(x_{1:t}), respectively. Here, the subscripts in Lt,sL_{t,s} and Ut,sU_{t,s}, 0t,sN0\leq t,s\leq N, indicate that different functions are used when the variables (x,y)(x,y) have different indices. However, for the sake of brevity, we denote Lt,s=:LL_{t,s}=:L and Ut,s=:UU_{t,s}=:U, since the indices in (x,y)(x,y) imply the functions used. Then we simply write

L(x1:N,yt)\displaystyle L(x_{1:N},y_{t}) π(x1:N,yt)U(x1:N,yt) and L(x1:t)π(x1:t)U(x1:t).\displaystyle\leq\pi(x_{1:N},y_{t})\leq U(x_{1:N},y_{t})\quad\text{ and }\quad L(x_{1:t})\leq\pi(x_{1:t})\leq U(x_{1:t}).

Since (π(x1:N,yt)L(x1:N,yt))(π(x1:t)L(x1:t))0(\pi(x_{1:N},y_{t})-L(x_{1:N},y_{t}))(\pi(x_{1:t})-L(x_{1:t}))\geq 0, expanding the expression gives

w1(x1:N,yt)L(x1:N,yt)π(x1:t)+π(x1:N,yt)L(x1:t)L(x1:N,yt)L(x1:t).w_{1}(x_{1:N},y_{t})\geq L(x_{1:N},y_{t})\pi(x_{1:t})+\pi(x_{1:N},y_{t})L(x_{1:t})-L(x_{1:N},y_{t})L(x_{1:t}). (4.2)

Similarly, the same procedure applied to products involving U(x1:N,yt)π(x1:N,yt)U(x_{1:N},y_{t})-\pi(x_{1:N},y_{t}) and U(x1:t)π(x1:t)U(x_{1:t})-\pi(x_{1:t}) yields three additional inequalities. We also replace the right-hand side of (4.1) with w2(x1:N,yt)=π(x1:t,yt)π(x1:N)w_{2}(x_{1:N},y_{t})=\pi(x_{1:t},y_{t})\pi(x_{1:N}), and four corresponding inequalities are derived similarly. The causality constraint becomes w1(x1:N,yt)=w2(x1:N,yt)w_{1}(x_{1:N},y_{t})=w_{2}(x_{1:N},y_{t}).

Together with the relaxation on anticausality and the martingale condition, we introduce McCormick martingale couplings as follows:

Definition 4.1.

Given the upper bound U()U(\cdot) and lower bound L()L(\cdot) on probability masses, πΠ(μ¯,ν¯)\pi\in\Pi(\bar{\mu},\bar{\nu}) is called a McCormick martingale coupling if it satisfies

  1. (1)

    the capacity constraint: LπUL\leq\pi\leq U;

  2. (2)

    the martingale condition:

    xt+1xt+1π(xt+1|x1:t,y1:t)=xt,yt+1yt+1π(yt+1|x1:t,y1:t)=yt,1tN1;\sum_{x_{t+1}}x_{t+1}\pi(x_{t+1}|x_{1:t},y_{1:t})=x_{t},\quad\sum_{y_{t+1}}y_{t+1}\pi(y_{t+1}|x_{1:t},y_{1:t})=y_{t},\quad 1\leq t\leq N-1; (4.3)
  3. (3)

    the McCormick relaxation of the causality condition: for all 1tN11\leq t\leq N-1 and x1:N,ytx_{1:N},y_{t}, there exist variables w1(),w2()[0,1]w_{1}(\cdot),w_{2}(\cdot)\in[0,1] satisfying

    w1(x1:N,yt)\displaystyle w_{1}(x_{1:N},y_{t}) L(x1:N,yt)π(x1:t)+π(x1:N,yt)L(x1:t)L(x1:N,yt)L(x1:t),\displaystyle\geq L(x_{1:N},y_{t})\pi(x_{1:t})+\pi(x_{1:N},y_{t})L(x_{1:t})-L(x_{1:N},y_{t})L(x_{1:t}), (4.4)
    w1(x1:N,yt)\displaystyle w_{1}(x_{1:N},y_{t}) U(x1:N,yt)π(x1:t)+π(x1:N,yt)U(x1:t)U(x1:N,yt)U(x1:t),\displaystyle\geq U(x_{1:N},y_{t})\pi(x_{1:t})+\pi(x_{1:N},y_{t})U(x_{1:t})-U(x_{1:N},y_{t})U(x_{1:t}),
    w1(x1:N,yt)\displaystyle w_{1}(x_{1:N},y_{t}) U(x1:N,yt)π(x1:t)+π(x1:N,yt)L(x1:t)U(x1:N,yt)L(x1:t),\displaystyle\leq U(x_{1:N},y_{t})\pi(x_{1:t})+\pi(x_{1:N},y_{t})L(x_{1:t})-U(x_{1:N},y_{t})L(x_{1:t}),
    w1(x1:N,yt)\displaystyle w_{1}(x_{1:N},y_{t}) π(x1:N,yt)U(x1:t)+L(x1:N,yt)π(x1:t)L(x1:N,yt)U(x1:t),\displaystyle\leq\pi(x_{1:N},y_{t})U(x_{1:t})+L(x_{1:N},y_{t})\pi(x_{1:t})-L(x_{1:N},y_{t})U(x_{1:t}),
    w2(x1:N,yt)\displaystyle w_{2}(x_{1:N},y_{t}) L(x1:t,yt)π(x1:N)+π(x1:t,yt)L(x1:N)L(x1:t,yt)L(x1:N),\displaystyle\geq L(x_{1:t},y_{t})\pi(x_{1:N})+\pi(x_{1:t},y_{t})L(x_{1:N})-L(x_{1:t},y_{t})L(x_{1:N}),
    w2(x1:N,yt)\displaystyle w_{2}(x_{1:N},y_{t}) U(x1:t,yt)π(x1:N)+π(x1:t,yt)U(x1:N)U(x1:t,yt)U(x1:N),\displaystyle\geq U(x_{1:t},y_{t})\pi(x_{1:N})+\pi(x_{1:t},y_{t})U(x_{1:N})-U(x_{1:t},y_{t})U(x_{1:N}),
    w2(x1:N,yt)\displaystyle w_{2}(x_{1:N},y_{t}) U(x1:t,yt)π(x1:N)+π(x1:t,yt)L(x1:N)U(x1:t,yt)L(x1:N),\displaystyle\leq U(x_{1:t},y_{t})\pi(x_{1:N})+\pi(x_{1:t},y_{t})L(x_{1:N})-U(x_{1:t},y_{t})L(x_{1:N}),
    w2(x1:N,yt)\displaystyle w_{2}(x_{1:N},y_{t}) π(x1:t,yt)U(x1:N)+L(x1:t,yt)π(x1:N)L(x1:t,yt)U(x1:N),\displaystyle\leq\pi(x_{1:t},y_{t})U(x_{1:N})+L(x_{1:t},y_{t})\pi(x_{1:N})-L(x_{1:t},y_{t})U(x_{1:N}),
    w1(x1:N,yt)\displaystyle w_{1}(x_{1:N},y_{t}) =w2(x1:N,yt);\displaystyle=w_{2}(x_{1:N},y_{t});
  4. (4)

    the McCormick relaxation of the anticausality condition, by interchanging xx and yy in (4.4).

We denote the set of all McCormick martingale couplings as (μ¯,ν¯;L,U){\mathcal{M}}(\bar{\mu},\bar{\nu};L,U).

All of the inequalities in (4.4) are derived in the same way as (4.2). Note that (4.4) holds for all x1:N,ytx_{1:N},y_{t}, instead of merely π\pi-a.s. In condition (3), it is possible to simplify w1(x1:N,yt)=w2(x1:N,yt)=:w(x1:N,yt)w_{1}(x_{1:N},y_{t})=w_{2}(x_{1:N},y_{t})=:w(x_{1:N},y_{t}). The term w()w(\cdot) (or w1()w_{1}(\cdot) and w2()w_{2}(\cdot)) also serves as a variable in the optimization process. However, in (4.4), we can eliminate the dependence on w()w(\cdot) by imposing that the maximum of the lower bounds does not exceed the minimum of the upper bounds for w()w(\cdot). Therefore, we only write π(μ¯,ν¯;L,U)\pi\in{\mathcal{M}}(\bar{\mu},\bar{\nu};L,U) to signify that π\pi is a McCormick martingale coupling.

A simple choice of bounds is L=0L=0 and U(x1:N,yt)=min{μ1(x1),,μN(xN),νt(yt)}U(x_{1:N},y_{t})=\min\{\mu_{1}(x_{1}),\ldots,\mu_{N}(x_{N}),\nu_{t}(y_{t})\}, which are derived from the marginal conditions and do not require additional modeling assumptions. Under this choice, (μ¯,ν¯;L,U){\mathcal{M}}(\bar{\mu},\bar{\nu};L,U) is also not empty. Indeed, we can first construct (or rather arbitrarily pick) the marginal laws, μΠ(μ¯)\mu\in\Pi(\overline{\mu}) and νΠ(ν¯)\nu\in\Pi(\overline{\nu}), such that X1:NX_{1:N} and Y1:NY_{1:N} are martingales under μ\mu and ν\nu, respectively. Then the independent coupling μν\mu\otimes\nu serves as a McCormick martingale coupling. In the general finite and discrete cases, many LP solvers can check whether (μ¯,ν¯;L,U){\mathcal{M}}(\bar{\mu},\bar{\nu};L,U) is non-empty.

Remark 4.2.

In Definition 4.1, the martingale condition is imposed under joint filtration, as the bicausal condition is relaxed and Proposition 3.1 is not applicable in this context.

Remark 4.3.

The total mass of LL or UU may not be equal to one. Both LL and UU are functions only, rather than probability measures or signed measures. Even when UU is a measure, the product π()U()\pi(\cdot)U(\cdot) is not a measure. Specifically, for a set A=BCA=B\cup C where BB and CC are disjoint, π(A)U(A)=[π(B)+π(C)][U(B)+U(C)]π(B)U(B)+π(C)U(C)\pi(A)U(A)=[\pi(B)+\pi(C)][U(B)+U(C)]\neq\pi(B)U(B)+\pi(C)U(C). Furthermore, it is crucial to interpret (4.4) pointwise in the discrete case, given that the McCormick relaxation is derived pointwise in (4.2).

We focus on addressing the minimization problem associated with the McCormick MOT in (4.5), as the maximization problem can be treated similarly:

infπ(μ¯,ν¯;L,U)c(x1:N,y1:N)π(dx1:N,dy1:N).\inf_{\pi\in{\mathcal{M}}(\bar{\mu},\bar{\nu};L,U)}\int c(x_{1:N},y_{1:N})\pi(dx_{1:N},dy_{1:N}). (4.5)

In finite and discrete scenarios, the infimum in the primal problem is attained when the set (μ¯,ν¯;L,U){\mathcal{M}}(\bar{\mu},\bar{\nu};L,U) is not empty. The dual problem is derived through classic finite-dimensional LP and is omitted here.

4.2 The absolutely continuous case with capacity constraints

In the continuous scenario similar to Korman and McCann, (2015); Bogachev, (2022), we make the assumption that the domains 𝒳t=𝒴t={\mathcal{X}}_{t}={\mathcal{Y}}_{t}=\mathbb{R} and the marginals μt\mu_{t} and νt\nu_{t} are absolutely continuous with respect to the Lebesgue measure on \mathbb{R}. Similarly, we consider couplings π\pi that are absolutely continuous with respect to the Lebesgue measure λ\lambda on N×N\mathbb{R}^{N}\times\mathbb{R}^{N}. When the context is evident, we denote λ(dx1:N,dy1:N)\lambda(dx_{1:N},dy_{1:N}) by dλd\lambda and omit the arguments. Let f(x1:N,y1:N)f(x_{1:N},y_{1:N}) represent the density of a coupling π\pi. Although probability measures with densities are absolutely continuous with respect to the Lebesgue measure, we restrict them to the Borel σ\sigma-algebra on the Euclidean space and interpret them as Borel probability measures. Additionally, nonnegative stock prices result in zero densities in negative domains.

Due to absolute continuity, the causality constraint (3.2) (or (3.3)) transforms into the following density equality:

f(x1:N,yt)f(x1:t)=f(x1:t,yt)f(x1:N),λ-a.e.,t=1,,N1.f(x_{1:N},y_{t})f(x_{1:t})=f(x_{1:t},y_{t})f(x_{1:N}),\quad\text{$\lambda$-a.e.},\quad t=1,\ldots,N-1. (4.6)

Crucially, (4.6) holds λ\lambda-a.e., instead of merely π\pi-a.s. It can be verified in the same spirit of (4.1), by noting that both sides of (4.6) are zero when regular conditional kernels are not defined.

With a slight abuse of notation, let w1(x1:N,yt)=f(x1:N,yt)f(x1:t)w_{1}(x_{1:N},y_{t})=f(x_{1:N},y_{t})f(x_{1:t}) still be the auxiliary variable. With the same convention on omitting the subscripts as in the discrete case, we suppose that the upper bound u()u(\cdot) and lower bound l()l(\cdot) for the density f()f(\cdot) are known, such that

l(x1:N,yt)\displaystyle l(x_{1:N},y_{t}) f(x1:N,yt)u(x1:N,yt),λ-a.e. and l(x1:t)f(x1:t)u(x1:t),λ-a.e.,\displaystyle\leq f(x_{1:N},y_{t})\leq u(x_{1:N},y_{t}),\;\text{$\lambda$-a.e.}\quad\text{ and }\quad l(x_{1:t})\leq f(x_{1:t})\leq u(x_{1:t}),\;\text{$\lambda$-a.e.},

which are referred to as capacity/density constraints (Korman and McCann,, 2015; Bogachev,, 2022). Similarly, the inequality (4.2) becomes

w1(x1:N,yt)l(x1:N,yt)f(x1:t)+f(x1:N,yt)l(x1:t)l(x1:N,yt)l(x1:t),λ-a.e.w_{1}(x_{1:N},y_{t})\geq l(x_{1:N},y_{t})f(x_{1:t})+f(x_{1:N},y_{t})l(x_{1:t})-l(x_{1:N},y_{t})l(x_{1:t}),\quad\text{$\lambda$-a.e.} (4.7)

Other inequalities can be derived in a similar fashion. It is important to emphasize that l()l(\cdot) and u()u(\cdot) are not necessarily density functions.

Denote gtg_{t} as the density of μt\mu_{t}, and g¯:=(g1,,gN)\bar{g}:=(g_{1},\ldots,g_{N}) as the individual density vectors. Similarly, let hth_{t} be the density of νt\nu_{t}, and h¯:=(h1,,hN)\bar{h}:=(h_{1},\ldots,h_{N}) be the individual density vectors. We define the continuous version of McCormick martingale couplings similarly. Note that (4.9) holds λ\lambda-a.e.

Definition 4.4.

Given the upper bound u()u(\cdot) and lower bound l()l(\cdot) on densities, a coupling π\pi with density fΠ(g¯,h¯)f\in\Pi(\bar{g},\bar{h}) is called a McCormick martingale coupling if it satisfies

  1. (1)

    the capacity constraint: lful\leq f\leq u, λ\lambda-a.e.;

  2. (2)

    the martingale condition:

    xt+1xt+1f(xt+1|x1:t,y1:t)λ(dxt+1)\displaystyle\int_{x_{t+1}}x_{t+1}f(x_{t+1}|x_{1:t},y_{1:t})\lambda(dx_{t+1}) =xt, 1tN1,\displaystyle=x_{t},\;1\leq t\leq N-1, (4.8)
    yt+1yt+1f(yt+1|x1:t,y1:t)λ(dyt+1)\displaystyle\int_{y_{t+1}}y_{t+1}f(y_{t+1}|x_{1:t},y_{1:t})\lambda(dy_{t+1}) =yt, 1tN1;\displaystyle=y_{t},\;1\leq t\leq N-1;
  3. (3)

    the McCormick relaxation of the causality condition on densities: for all 1tN11\leq t\leq N-1, there exist w1(x1:N,yt),w2(x1:N,yt)0w_{1}(x_{1:N},y_{t}),w_{2}(x_{1:N},y_{t})\geq 0 such that the following inequality holds λ\lambda-almost everywhere:

    w1(x1:N,yt)\displaystyle w_{1}(x_{1:N},y_{t}) l(x1:N,yt)f(x1:t)+f(x1:N,yt)l(x1:t)l(x1:N,yt)l(x1:t),\displaystyle\geq l(x_{1:N},y_{t})f(x_{1:t})+f(x_{1:N},y_{t})l(x_{1:t})-l(x_{1:N},y_{t})l(x_{1:t}), (4.9)
    w1(x1:N,yt)\displaystyle w_{1}(x_{1:N},y_{t}) u(x1:N,yt)f(x1:t)+f(x1:N,yt)u(x1:t)u(x1:N,yt)u(x1:t),\displaystyle\geq u(x_{1:N},y_{t})f(x_{1:t})+f(x_{1:N},y_{t})u(x_{1:t})-u(x_{1:N},y_{t})u(x_{1:t}),
    w1(x1:N,yt)\displaystyle w_{1}(x_{1:N},y_{t}) u(x1:N,yt)f(x1:t)+f(x1:N,yt)l(x1:t)u(x1:N,yt)l(x1:t),\displaystyle\leq u(x_{1:N},y_{t})f(x_{1:t})+f(x_{1:N},y_{t})l(x_{1:t})-u(x_{1:N},y_{t})l(x_{1:t}),
    w1(x1:N,yt)\displaystyle w_{1}(x_{1:N},y_{t}) f(x1:N,yt)u(x1:t)+l(x1:N,yt)f(x1:t)l(x1:N,yt)u(x1:t),\displaystyle\leq f(x_{1:N},y_{t})u(x_{1:t})+l(x_{1:N},y_{t})f(x_{1:t})-l(x_{1:N},y_{t})u(x_{1:t}),
    w2(x1:N,yt)\displaystyle w_{2}(x_{1:N},y_{t}) l(x1:t,yt)f(x1:N)+f(x1:t,yt)l(x1:N)l(x1:t,yt)l(x1:N),\displaystyle\geq l(x_{1:t},y_{t})f(x_{1:N})+f(x_{1:t},y_{t})l(x_{1:N})-l(x_{1:t},y_{t})l(x_{1:N}),
    w2(x1:N,yt)\displaystyle w_{2}(x_{1:N},y_{t}) u(x1:t,yt)f(x1:N)+f(x1:t,yt)u(x1:N)u(x1:t,yt)u(x1:N),\displaystyle\geq u(x_{1:t},y_{t})f(x_{1:N})+f(x_{1:t},y_{t})u(x_{1:N})-u(x_{1:t},y_{t})u(x_{1:N}),
    w2(x1:N,yt)\displaystyle w_{2}(x_{1:N},y_{t}) u(x1:t,yt)f(x1:N)+f(x1:t,yt)l(x1:N)u(x1:t,yt)l(x1:N),\displaystyle\leq u(x_{1:t},y_{t})f(x_{1:N})+f(x_{1:t},y_{t})l(x_{1:N})-u(x_{1:t},y_{t})l(x_{1:N}),
    w2(x1:N,yt)\displaystyle w_{2}(x_{1:N},y_{t}) f(x1:t,yt)u(x1:N)+l(x1:t,yt)f(x1:N)l(x1:t,yt)u(x1:N),\displaystyle\leq f(x_{1:t},y_{t})u(x_{1:N})+l(x_{1:t},y_{t})f(x_{1:N})-l(x_{1:t},y_{t})u(x_{1:N}),
    w1(x1:N,yt)\displaystyle w_{1}(x_{1:N},y_{t}) =w2(x1:N,yt);\displaystyle=w_{2}(x_{1:N},y_{t});
  4. (4)

    the McCormick relaxation of the anticausality condition on densities, by interchanging xx and yy in (4.9).

We use f(g¯,h¯;l,u)f\in{\mathcal{M}}(\bar{g},\bar{h};l,u) to denote that ff corresponds to a McCormick coupling with capacity constraints.

With continuous densities, we denote the risk-neutral price of the exotic option as

P(f):=c(x1:N,y1:N)f(x1:N,y1:N)𝑑λ.P(f):=\int c(x_{1:N},y_{1:N})f(x_{1:N},y_{1:N})d\lambda. (4.10)

The primal problem, incorporating McCormick relaxation and capacity constraints, can be expressed as

P=inff(g¯,h¯;l,u)P(f).P=\inf_{f\in{\mathcal{M}}(\bar{g},\bar{h};l,u)}P(f). (4.11)

This formulation is referred to as the continuous version of McCormick MOT.

Theorem 4.5 demonstrates the primal attainment of the problem (4.11).

Theorem 4.5.

Suppose the following conditions hold:

  1. 1.

    (g¯,h¯;l,u){\mathcal{M}}(\bar{g},\bar{h};l,u) is not empty;

  2. 2.

    Probability densities gt,htL1()g_{t},h_{t}\in L^{1}(\mathbb{R}) and have finite first moments. The bounds l,uL1(N×N)l,u\in L^{1}(\mathbb{R}^{N}\times\mathbb{R}^{N}) and l,u0l,u\geq 0;

  3. 3.

    The cost (payoff) function c:N×N(,]c:\mathbb{R}^{N}\times\mathbb{R}^{N}\rightarrow(-\infty,\infty] is measurable and cC(1+i=1N|xi|+j=1N|yj|)c\geq-C(1+\sum^{N}_{i=1}|x_{i}|+\sum^{N}_{j=1}|y_{j}|) for some constant C0C\geq 0.

Then the infimum in (4.11) is attained.

4.2.1 The strong duality

It is well known that the dual formulation of the classic MOT corresponds to a semi-static hedging strategy, comprising the sum of static vanilla portfolios and dynamic delta positions (Beiglböck et al.,, 2013). A natural question arises regarding the impact of the McCormick relaxation on the hedging portfolio.

In an informal manner, one can derive the dual problem by considering the Lagrange multipliers method and interchanging the infimum with the supremum. We introduce the multipliers as follows, having reduced w1(x1:N,yt)=w2(x1:N,yt)=w(x1:N,yt)w_{1}(x_{1:N},y_{t})=w_{2}(x_{1:N},y_{t})=w(x_{1:N},y_{t}):

  1. (1)

    ϕt\phi_{t} and φt\varphi_{t}, 1tN1\leq t\leq N denote potential functions testing the marginal constraints for μt\mu_{t} and νt\nu_{t}, respectively. In financial terms, they can be interpreted as a portfolio of static vanilla options;

  2. (2)

    αt\alpha_{t} and βt\beta_{t}, 1tN11\leq t\leq N-1 test the martingale condition for XX and YY, respectively. They represent self-financing trading strategies in the risky assets;

  3. (3)

    γt,i\gamma_{t,i}, 1tN11\leq t\leq N-1, 1i81\leq i\leq 8, are multipliers for the McCormick relaxation of the causality condition;

  4. (4)

    ηt,i\eta_{t,i}, 1tN11\leq t\leq N-1, 1i81\leq i\leq 8 are multipliers for the McCormick relaxation of the anticausality condition;

  5. (5)

    κ\kappa and θ\theta are the multipliers for the upper and lower bounds, respectively.

Constraints in the dual problem consist of the following components:

  1. (1)

    The subhedging condition in the sense that

    c(x1:N,y1:N)t=1Nϕt(xt)t=1Nφt(yt)+κ(x1:N,y1:N)θ(x1:N,y1:N)\displaystyle c(x_{1:N},y_{1:N})-\sum^{N}_{t=1}\phi_{t}(x_{t})-\sum^{N}_{t=1}\varphi_{t}(y_{t})+\kappa(x_{1:N},y_{1:N})-\theta(x_{1:N},y_{1:N}) (4.12)
    +t=1N1αt(x1:t,y1:t)(xt+1xt)+t=1N1βt(x1:t,y1:t)(yt+1yt)\displaystyle+\sum^{N-1}_{t=1}\alpha_{t}(x_{1:t},y_{1:t})(x_{t+1}-x_{t})+\sum^{N-1}_{t=1}\beta_{t}(x_{1:t},y_{1:t})(y_{t+1}-y_{t})
    +Ψ(x1:N,y1:N,γ,η;u,l)0,\displaystyle+\Psi(x_{1:N},y_{1:N},\gamma,\eta;u,l)\geq 0,

    where

    Ψ(x1:N,y1:N,γ,η;u,l)\displaystyle\Psi(x_{1:N},y_{1:N},\gamma,\eta;u,l) (4.13)
    :=t=1N1γt,1(x1:N,yt)(l(x1:N,yt)+l(x1:t))+t=1N1γt,2(x1:N,yt)(u(x1:N,yt)+u(x1:t))\displaystyle:=\sum^{N-1}_{t=1}\gamma_{t,1}(x_{1:N},y_{t})\big{(}l(x_{1:N},y_{t})+l(x_{1:t})\big{)}+\sum^{N-1}_{t=1}\gamma_{t,2}(x_{1:N},y_{t})\big{(}u(x_{1:N},y_{t})+u(x_{1:t})\big{)}
    t=1N1γt,3(x1:N,yt)(u(x1:N,yt)+l(x1:t))t=1N1γt,4(x1:N,yt)(l(x1:N,yt)+u(x1:t))\displaystyle\quad-\sum^{N-1}_{t=1}\gamma_{t,3}(x_{1:N},y_{t})\big{(}u(x_{1:N},y_{t})+l(x_{1:t})\big{)}-\sum^{N-1}_{t=1}\gamma_{t,4}(x_{1:N},y_{t})\big{(}l(x_{1:N},y_{t})+u(x_{1:t})\big{)}
    +t=1N1γt,5(x1:N,yt)(l(x1:t,yt)+l(x1:N))+t=1N1γt,6(x1:N,yt)(u(x1:t,yt)+u(x1:N))\displaystyle\quad+\sum^{N-1}_{t=1}\gamma_{t,5}(x_{1:N},y_{t})\big{(}l(x_{1:t},y_{t})+l(x_{1:N})\big{)}+\sum^{N-1}_{t=1}\gamma_{t,6}(x_{1:N},y_{t})\big{(}u(x_{1:t},y_{t})+u(x_{1:N})\big{)}
    t=1N1γt,7(x1:N,yt)(u(x1:t,yt)+l(x1:N))t=1N1γt,8(x1:N,yt)(l(x1:t,yt)+u(x1:N))\displaystyle\quad-\sum^{N-1}_{t=1}\gamma_{t,7}(x_{1:N},y_{t})\big{(}u(x_{1:t},y_{t})+l(x_{1:N})\big{)}-\sum^{N-1}_{t=1}\gamma_{t,8}(x_{1:N},y_{t})\big{(}l(x_{1:t},y_{t})+u(x_{1:N})\big{)}
    +t=1N1ηt,1(xt,y1:N)(l(xt,y1:N)+l(y1:t))+t=1N1ηt,2(xt,y1:N)(u(xt,y1:N)+u(y1:t))\displaystyle\quad+\sum^{N-1}_{t=1}\eta_{t,1}(x_{t},y_{1:N})\big{(}l(x_{t},y_{1:N})+l(y_{1:t})\big{)}+\sum^{N-1}_{t=1}\eta_{t,2}(x_{t},y_{1:N})\big{(}u(x_{t},y_{1:N})+u(y_{1:t})\big{)}
    t=1N1ηt,3(xt,y1:N)(u(xt,y1:N)+l(y1:t))t=1N1ηt,4(xt,y1:N)(l(xt,y1:N)+u(y1:t))\displaystyle\quad-\sum^{N-1}_{t=1}\eta_{t,3}(x_{t},y_{1:N})\big{(}u(x_{t},y_{1:N})+l(y_{1:t})\big{)}-\sum^{N-1}_{t=1}\eta_{t,4}(x_{t},y_{1:N})\big{(}l(x_{t},y_{1:N})+u(y_{1:t})\big{)}
    +t=1N1ηt,5(xt,y1:N)(l(xt,y1:t)+l(y1:N))+t=1N1ηt,6(xt,y1:N)(u(xt,y1:t)+u(y1:N))\displaystyle\quad+\sum^{N-1}_{t=1}\eta_{t,5}(x_{t},y_{1:N})\big{(}l(x_{t},y_{1:t})+l(y_{1:N})\big{)}+\sum^{N-1}_{t=1}\eta_{t,6}(x_{t},y_{1:N})\big{(}u(x_{t},y_{1:t})+u(y_{1:N})\big{)}
    t=1N1ηt,7(xt,y1:N)(u(xt,y1:t)+l(y1:N))t=1N1ηt,8(xt,y1:N)(l(xt,y1:t)+u(y1:N));\displaystyle\quad-\sum^{N-1}_{t=1}\eta_{t,7}(x_{t},y_{1:N})\big{(}u(x_{t},y_{1:t})+l(y_{1:N})\big{)}-\sum^{N-1}_{t=1}\eta_{t,8}(x_{t},y_{1:N})\big{(}l(x_{t},y_{1:t})+u(y_{1:N})\big{)};
  2. (2)

    The dual inequality for the McCormick relaxation of the causality condition:

    γt,1(x1:N,yt)γt,2(x1:N,yt)+γt,3(x1:N,yt)+γt,4(x1:N,yt)\displaystyle-\gamma_{t,1}(x_{1:N},y_{t})-\gamma_{t,2}(x_{1:N},y_{t})+\gamma_{t,3}(x_{1:N},y_{t})+\gamma_{t,4}(x_{1:N},y_{t}) (4.14)
    γt,5(x1:N,yt)γt,6(x1:N,yt)+γt,7(x1:N,yt)+γt,8(x1:N,yt)0,t=1,,N1.\displaystyle-\gamma_{t,5}(x_{1:N},y_{t})-\gamma_{t,6}(x_{1:N},y_{t})+\gamma_{t,7}(x_{1:N},y_{t})+\gamma_{t,8}(x_{1:N},y_{t})\geq 0,\quad t=1,\ldots,N-1.
  3. (3)

    The dual inequality for the McCormick relaxation of the anticausality condition:

    ηt,1(xt,y1:N)ηt,2(xt,y1:N)+ηt,3(xt,y1:N)+ηt,4(xt,y1:N)\displaystyle-\eta_{t,1}(x_{t},y_{1:N})-\eta_{t,2}(x_{t},y_{1:N})+\eta_{t,3}(x_{t},y_{1:N})+\eta_{t,4}(x_{t},y_{1:N}) (4.15)
    ηt,5(xt,y1:N)ηt,6(xt,y1:N)+ηt,7(xt,y1:N)+ηt,8(xt,y1:N)0,t=1,,N1.\displaystyle-\eta_{t,5}(x_{t},y_{1:N})-\eta_{t,6}(x_{t},y_{1:N})+\eta_{t,7}(x_{t},y_{1:N})+\eta_{t,8}(x_{t},y_{1:N})\geq 0,\quad t=1,\ldots,N-1.

Define the set of feasible multipliers as

Φc:={(ϕ,φ,α,β,γ,η,κ,θ)|\displaystyle\Phi_{c}:=\Big{\{}\Big{(}\phi,\varphi,\alpha,\beta,\gamma,\eta,\kappa,\theta\Big{)}\Big{|} ϕt,φt,θ,κ,αt,βt,γt,i,ηt,i are continuous and bounded;\displaystyle\phi_{t},\varphi_{t},\theta,\kappa,\alpha_{t},\beta_{t},\gamma_{t,i},\eta_{t,i}\text{ are continuous and bounded};
γt,i0,ηt,i0,κ0,θ0;\displaystyle\gamma_{t,i}\geq 0,\eta_{t,i}\geq 0,\kappa\geq 0,\theta\geq 0;
(4.12), (4.14), (4.15) hold}.\displaystyle\text{\eqref{simple_f_dual}, \eqref{dual_Mc}, \eqref{dual_AntiMc} hold}\Big{\}}. (4.16)

The objective function in the dual problem is

D(ϕ,φ,α,β,γ,η,κ,θ)\displaystyle D(\phi,\varphi,\alpha,\beta,\gamma,\eta,\kappa,\theta) (4.17)
:=t=1Nϕt(xt)gt(xt)λ(dxt)+t=1Nφt(yt)ht(yt)λ(dyt)\displaystyle:=\sum^{N}_{t=1}\int\phi_{t}(x_{t})g_{t}(x_{t})\lambda(dx_{t})+\sum^{N}_{t=1}\int\varphi_{t}(y_{t})h_{t}(y_{t})\lambda(dy_{t})
κ(x1:N,y1:N)u(x1:N,y1:N)𝑑λ+θ(x1:N,y1:N)l(x1:N,y1:N)𝑑λ\displaystyle\quad-\int\kappa(x_{1:N},y_{1:N})u(x_{1:N},y_{1:N})d\lambda+\int\theta(x_{1:N},y_{1:N})l(x_{1:N},y_{1:N})d\lambda
t=1N1[γt,1(x1:N,yt)l(x1:N,yt)l(x1:t)+γt,2(x1:N,yt)u(x1:N,yt)u(x1:t)]λ(dx1:N,dyt)\displaystyle\quad-\sum^{N-1}_{t=1}\int\Big{[}\gamma_{t,1}(x_{1:N},y_{t})l(x_{1:N},y_{t})l(x_{1:t})+\gamma_{t,2}(x_{1:N},y_{t})u(x_{1:N},y_{t})u(x_{1:t})\Big{]}\lambda(dx_{1:N},dy_{t})
+t=1N1[γt,3(x1:N,yt)u(x1:N,yt)l(x1:t)+γt,4(x1:N,yt)l(x1:N,yt)u(x1:t)]λ(dx1:N,dyt)\displaystyle\quad+\sum^{N-1}_{t=1}\int\Big{[}\gamma_{t,3}(x_{1:N},y_{t})u(x_{1:N},y_{t})l(x_{1:t})+\gamma_{t,4}(x_{1:N},y_{t})l(x_{1:N},y_{t})u(x_{1:t})\Big{]}\lambda(dx_{1:N},dy_{t})
t=1N1[γt,5(x1:N,yt)l(x1:t,yt)l(x1:N)+γt,6(x1:N,yt)u(x1:t,yt)u(x1:N)]λ(dx1:N,dyt)\displaystyle\quad-\sum^{N-1}_{t=1}\int\Big{[}\gamma_{t,5}(x_{1:N},y_{t})l(x_{1:t},y_{t})l(x_{1:N})+\gamma_{t,6}(x_{1:N},y_{t})u(x_{1:t},y_{t})u(x_{1:N})\Big{]}\lambda(dx_{1:N},dy_{t})
+t=1N1[γt,7(x1:N,yt)u(x1:t,yt)l(x1:N)+γt,8(x1:N,yt)l(x1:t,yt)u(x1:N)]λ(dx1:N,dyt)\displaystyle\quad+\sum^{N-1}_{t=1}\int\Big{[}\gamma_{t,7}(x_{1:N},y_{t})u(x_{1:t},y_{t})l(x_{1:N})+\gamma_{t,8}(x_{1:N},y_{t})l(x_{1:t},y_{t})u(x_{1:N})\Big{]}\lambda(dx_{1:N},dy_{t})
t=1N1[ηt,1(xt,y1:N)l(xt,y1:N)l(y1:t)+ηt,2(xt,y1:N)u(xt,y1:N)u(y1:t)]λ(dxt,dy1:N)\displaystyle\quad-\sum^{N-1}_{t=1}\int\Big{[}\eta_{t,1}(x_{t},y_{1:N})l(x_{t},y_{1:N})l(y_{1:t})+\eta_{t,2}(x_{t},y_{1:N})u(x_{t},y_{1:N})u(y_{1:t})\Big{]}\lambda(dx_{t},dy_{1:N})
+t=1N1[ηt,3(xt,y1:N)u(xt,y1:N)l(y1:t)+ηt,4(xt,y1:N)l(xt,y1:N)u(y1:t)]λ(dxt,dy1:N)\displaystyle\quad+\sum^{N-1}_{t=1}\int\Big{[}\eta_{t,3}(x_{t},y_{1:N})u(x_{t},y_{1:N})l(y_{1:t})+\eta_{t,4}(x_{t},y_{1:N})l(x_{t},y_{1:N})u(y_{1:t})\Big{]}\lambda(dx_{t},dy_{1:N})
t=1N1[ηt,5(xt,y1:N)l(xt,y1:t)l(y1:N)+ηt,6(xt,y1:N)u(xt,y1:t)u(y1:N)]λ(dxt,dy1:N)\displaystyle\quad-\sum^{N-1}_{t=1}\int\Big{[}\eta_{t,5}(x_{t},y_{1:N})l(x_{t},y_{1:t})l(y_{1:N})+\eta_{t,6}(x_{t},y_{1:N})u(x_{t},y_{1:t})u(y_{1:N})\Big{]}\lambda(dx_{t},dy_{1:N})
+t=1N1[ηt,7(xt,y1:N)u(xt,y1:t)l(y1:N)+ηt,8(xt,y1:N)l(xt,y1:t)u(y1:N)]λ(dxt,dy1:N).\displaystyle\quad+\sum^{N-1}_{t=1}\int\Big{[}\eta_{t,7}(x_{t},y_{1:N})u(x_{t},y_{1:t})l(y_{1:N})+\eta_{t,8}(x_{t},y_{1:N})l(x_{t},y_{1:t})u(y_{1:N})\Big{]}\lambda(dx_{t},dy_{1:N}).

Finally, we can state the dual problem as

D=supΦcD(ϕ,φ,α,β,γ,η,κ,θ),D=\sup_{\Phi_{c}}D(\phi,\varphi,\alpha,\beta,\gamma,\eta,\kappa,\theta), (4.18)

where we omit the time subscript in ϕ,φ,α,β\phi,\varphi,\alpha,\beta, etc., for simplicity.

Loosely speaking, McCormick relaxations modify the subhedging portfolio (ϕ,φ,α,β)(\phi,\varphi,\alpha,\beta) in (4.12) by considering

c(x1:N,y1:N)+κ(x1:N,y1:N)θ(x1:N,y1:N)+Ψ(x1:N,y1:N,γ,η;u,l)c(x_{1:N},y_{1:N})+\kappa(x_{1:N},y_{1:N})-\theta(x_{1:N},y_{1:N})+\Psi(x_{1:N},y_{1:N},\gamma,\eta;u,l) (4.19)

in lieu of c(x1:N,y1:N)c(x_{1:N},y_{1:N}). Additionally, the objective function (4.17) is augmented with terms originating from capacity constraints and McCormick relaxations. However, elucidating these additional terms poses challenges due to the imposition of numerous inequalities in the McCormick envelopes.

We establish strong duality P=DP=D, as defined in (4.11) and (4.18), through a two-step proof. First, we extend the capacity-constrained case presented in Korman et al., (2015, Theorem 1) to encompass non-compact supports, where neither martingale conditions nor McCormick relaxations are imposed. Although our Lemma 4.6 is acknowledged in the literature (Korman et al.,, 2015), a precise proof has not been provided in previous work. For later use, let Π(g,h;u)\Pi(g,h;u) represent the set of couplings with density 0fu0\leq f\leq u and marginals gg and hh on \mathbb{R}. While the proof focuses on one-dimensional cases for simplicity, the extension to multi-dimensional cases is analogous. The main idea involves a modification of Villani, (2003, Proposition 1.22), specifically tailored to address the capacity constraint. Unlike Theorem 4.5, Lemma 4.6 assumes lower semicontinuous (l.s.c.) costs. This condition is needed since the proof employs continuous and bounded functions for cost (payoff) approximation.

Lemma 4.6.

Suppose the following conditions hold:

  1. 1.

    The set of couplings Π(g,h;u)\Pi(g,h;u) is not empty;

  2. 2.

    The upper bound uu satisfies 0uL(×)0\leq u\in L^{\infty}(\mathbb{R}\times\mathbb{R}). Marginal probability densities g,hL()g,h\in L^{\infty}(\mathbb{R}) and have finite first moments;

  3. 3.

    The cost c:×(,]c:\mathbb{R}\times\mathbb{R}\rightarrow(-\infty,\infty] is l.s.c. and cC(1+|x|+|y|)c\geq-C(1+|x|+|y|) for some constant C0C\geq 0.

Then strong duality holds:

inffΠ(g,h;u)P(f)=supΨcF(ϕ,φ,κ),\inf_{f\in\Pi(g,h;u)}P(f)=\sup_{\Psi_{c}}F(\phi,\varphi,\kappa), (4.20)

where P(f)P(f) is defined in (4.10),

F(ϕ,φ,κ):=ϕ(x)g(x)𝑑x+φ(y)h(y)𝑑yκ(x,y)u(x,y)𝑑x𝑑y,F(\phi,\varphi,\kappa):=\int\phi(x)g(x)dx+\int\varphi(y)h(y)dy-\int\kappa(x,y)u(x,y)dxdy,

and

Ψc:={(ϕ,φ,κ)|\displaystyle\Psi_{c}:=\Big{\{}(\phi,\varphi,\kappa)\Big{|} ϕ(x)+φ(y)c(x,y)+κ(x,y),κ(x,y)0,\displaystyle\phi(x)+\varphi(y)\leq c(x,y)+\kappa(x,y),\,\kappa(x,y)\geq 0,
κL1(udxdy),ϕL1(gdx),φL1(hdy)}.\displaystyle\kappa\in L^{1}(udxdy),\phi\in L^{1}(gdx),\varphi\in L^{1}(hdy)\Big{\}}.

The infimum is also attained. Moreover, it does not change the value of the supremum on the right-hand side if we restrict (κ,ϕ,φ)(\kappa,\phi,\varphi) to be continuous and bounded.

To establish the strong duality result using Lemma 4.6, we additionally assume that the upper and lower bounds are continuous and bounded in Theorem 4.7.

Theorem 4.7.

Suppose the following conditions hold:

  1. 1.

    (g¯,h¯;l,u){\mathcal{M}}(\bar{g},\bar{h};l,u) is not empty;

  2. 2.

    The upper bound uu satisfies 0uCb(N×N)0\leq u\in C_{b}(\mathbb{R}^{N}\times\mathbb{R}^{N}). The lower bound ll satisfies 0lCb(N×N)0\leq l\in C_{b}(\mathbb{R}^{N}\times\mathbb{R}^{N}). The probability densities gt,htL()g_{t},h_{t}\in L^{\infty}(\mathbb{R}) have finite first moments;

  3. 3.

    The cost (payoff) function c:N×N(,]c:\mathbb{R}^{N}\times\mathbb{R}^{N}\rightarrow(-\infty,\infty] is l.s.c. and cC(1+i=1N|xi|+j=1N|yj|)c\geq-C(1+\sum^{N}_{i=1}|x_{i}|+\sum^{N}_{j=1}|y_{j}|) for some constant C0C\geq 0.

Then the strong duality holds: P=DP=D.

5 Numerical study

In this section, we investigate the effectiveness of McCormick MOT using both synthetic and empirical data. Given the typical unavailability of risk-neutral densities, our empirical study includes the development of a calibration procedure capable of accommodating bid-ask spreads. We focus on basket options with Asian-style payoffs. Initially, using the lower and upper bounds that naturally arise from marginal constraints, the McCormick relaxations reduce the price gap by an average of 1% for stocks with liquid option markets and 4% for those with moderately liquid markets. When tighter bounds on probability masses are applied, the McCormick MOT achieves an average reduction of 12.66%.

5.1 An illustrating example

Assuming a zero risk-free rate for simplicity, we explore a scenario involving two stocks, XX and YY, covering two periods. An exotic option has a payoff given by max{(X2X1)2,(Y2Y1)2}\max\{(X_{2}-X_{1})^{2},(Y_{2}-Y_{1})^{2}\}.

Setting the initial stock prices as X0=10X_{0}=10 and Y0=20Y_{0}=20 for illustrative purposes, we posit that each stock can have only three different prices at each maturity. Specifically:

  • (1)

    At time t=1t=1, the stock price X1=11,10,9X_{1}=11,10,9 with probabilities 0.2,0.6,0.20.2,0.6,0.2, respectively. Y1=24,20,16Y_{1}=24,20,16 with probabilities 0.3,0.4,0.30.3,0.4,0.3, respectively;

  • (2)

    At time t=2t=2, the stock price X2=20,10,0X_{2}=20,10,0 with probabilities 0.1,0.8,0.10.1,0.8,0.1, respectively. Y2=26,20,14Y_{2}=26,20,14 with probabilities 0.2,0.6,0.20.2,0.6,0.2, respectively.

It is straightforward to solve this problem with optimization programs such as Gurobi. The classic MOT yields option price bounds of [20.93,24.40][20.93,24.40]. To isolate the impact of McCormick relaxations, we utilize the lower and upper bounds implied by the marginal conditions: L=0L=0 and U(x1:N,yt)=min{μ1(x1),,μN(xN),νt(yt)}U(x_{1:N},y_{t})=\min\{\mu_{1}(x_{1}),\ldots,\mu_{N}(x_{N}),\nu_{t}(y_{t})\}. Employing the McCormick MOT refines the bounds to [21.50,24.40][21.50,24.40]. Despite the non-convex nature of bicausal MOT, (due to the small size of the problem) the program successfully addresses the problem by brute-force, providing bounds of [21.64,24.40][21.64,24.40]. This simple example highlights the potential of McCormick MOT, prompting further exploration in more general cases.

5.2 A calibration method of risk-neutral densities

In the empirical case, McCormick MOT also needs risk-neutral distributions as input parameters. When calibrating μ1\mu_{1} and μ2\mu_{2} independently, there is no guarantee of maintaining convex order. To address this issue, we propose a methodology for calibrating the risk-neutral densities of a risky asset while maintaining the convex order between marginals. An alternative approach involves the separate calibration of μ1\mu_{1} and μ2\mu_{2}, followed by convexification using the method outlined in Alfonsi et al., (2017, Equation 3.1). However, this approach often results in the modification of the support of μ2\mu_{2}, raising uncertainties regarding the consistency of the modified μ2\mu_{2} with the options data.

Our construction is outlined as follows. The current time is denoted as 0. At present, a finite number of European call options with known bid and ask prices are available. Recall that the set of maturities is represented as T1<<Tt<<TNT_{1}<\ldots<T_{t}<\ldots<T_{N}. For each maturity TtT_{t}, there are n(t)n(t) options available, where n(t)n(t) is not required to be the same across all maturities. Strike prices at a given maturity TtT_{t} are sorted as K1,t<<Ki,t<<Kn(t),tK_{1,t}<\ldots<K_{i,t}<\ldots<K_{n(t),t}. In the context of considering the risk-neutral density of a single asset, we denote the underlying stock price at TtT_{t} as St:=STtS_{t}:=S_{T_{t}}. Furthermore,

  1. (1)

    FtF_{t} represents the forward price of the stock, where the forward contract is initiated at time 0 and delivered at time TtT_{t};

  2. (2)

    DtD_{t} denotes the discount factor for time TtT_{t}, i.e., the price of a zero-coupon bond at time 0 with maturity at time TtT_{t};

  3. (3)

    Ai,tA_{i,t} is the ask price of the call with the strike Ki,tK_{i,t} and maturity TtT_{t}. Similarly, Bi,tB_{i,t} denotes the bid price.

We refer to Ci,tC_{i,t} as a feasible price if Ci,t[Bi,t,Ai,t]C_{i,t}\in[B_{i,t},A_{i,t}]. To streamline the framework, we assume that the interest rate is deterministic and that any dividends paid by the underlying assets are also deterministic if applicable. Following the approach of Davis and Hobson, (2007), we introduce scaled prices denoted as

ci,t=Ci,tDtFt,ai,t=Ai,tDtFt,bi,t=Bi,tDtFt,ki,t=Ki,tFt.c_{i,t}=\frac{C_{i,t}}{D_{t}F_{t}},\quad a_{i,t}=\frac{A_{i,t}}{D_{t}F_{t}},\quad b_{i,t}=\frac{B_{i,t}}{D_{t}F_{t}},\quad k_{i,t}=\frac{K_{i,t}}{F_{t}}. (5.1)

We write μ(S1:N=s1:N)\mu(S_{1:N}=s_{1:N}) as a joint risk-neutral measure encompassing all maturities. For simplicity, we denote μt(St=st)\mu_{t}(S_{t}=s_{t}) as the marginal of μ\mu on StS_{t}. Similarly, μ(S1:t=s1:t)\mu(S_{1:t}=s_{1:t}) refers to the marginal of μ\mu on S1:tS_{1:t}. Suppose that the support of the stock price at time TtT_{t}, denoted as 𝒮t{\mathcal{S}}_{t}, is given by the set of possible strikes at time TtT_{t}, i.e., 𝒮t={K1,t,,Kn(t),t}{\mathcal{S}}_{t}=\{K_{1,t},\ldots,K_{n(t),t}\}. Consequently, μ\mu is an n(1)×n(2)××n(T)n(1)\times n(2)\times\ldots\times n(T) array.

To find the risk-neutral measure μ\mu, we consider the following optimization problem:

minci,t,μ1tN1in(t)|ci,tai,t|+|ci,tbi,t|\displaystyle\min_{c_{i,t},\;\mu}\sum_{\begin{subarray}{c}1\leq t\leq N\\ 1\leq i\leq n(t)\end{subarray}}|c_{i,t}-a_{i,t}|+|c_{i,t}-b_{i,t}| (5.2)
subject tost𝒮t(stFtki,t)+μt(St=st)=ci,t,\displaystyle\textrm{ subject to}\sum_{s_{t}\in{\mathcal{S}}_{t}}\left(\frac{s_{t}}{F_{t}}-k_{i,t}\right)^{+}\mu_{t}(S_{t}=s_{t})=c_{i,t}, (5.3)
st+1𝒮t+1st+1Ft+1μ(s1:t,st+1)=stFtμ(S1:t=s1:t),s1:t, 1tN1,\displaystyle\qquad\qquad\sum_{s_{t+1}\in{\mathcal{S}}_{t+1}}\frac{s_{t+1}}{F_{t+1}}\mu(s_{1:t},s_{t+1})=\frac{s_{t}}{F_{t}}\mu(S_{1:t}=s_{1:t}),\quad\forall\;s_{1:t},\;1\leq t\leq N-1, (5.4)
st𝒮tstμt(St=st)=Ft, 1tN,\displaystyle\qquad\qquad\sum_{s_{t}\in{\mathcal{S}}_{t}}s_{t}\mu_{t}(S_{t}=s_{t})=F_{t},\quad\forall\;1\leq t\leq N, (5.5)
μ(S1:N=s1:N)0,st𝒮tμt(St=st)=1, 1tN.\displaystyle\qquad\qquad\mu(S_{1:N}=s_{1:N})\geq 0,\quad\sum_{s_{t}\in{\mathcal{S}}_{t}}\mu_{t}(S_{t}=s_{t})=1,\quad\forall\;1\leq t\leq N. (5.6)

The objective function (5.2) aims to place the feasible price ci,t[bi,t,ai,t]c_{i,t}\in[b_{i,t},a_{i,t}]. The constraint (5.3) represents the risk-neutral pricing formula, while (5.4) enforces the martingale condition. (5.5) stems from the fact that the expected value of stock prices under a risk-neutral measure equals the forward price. The final condition ensures that μt\mu_{t}, the marginal of μ\mu at time TtT_{t}, acts as a probability mass function. A notable advantage of our framework is that the program (5.2) is still an LP problem, although the solution is not necessarily unique. When there is a risk-neutral measure μ\mu that guarantees ci,t[bi,t,ai,t]c_{i,t}\in[b_{i,t},a_{i,t}], the objective value is the lowest and equals i,t|ai,tbi,t|\sum_{i,t}|a_{i,t}-b_{i,t}|. When the optimized objective value exceeds i,t|ai,tbi,t|\sum_{i,t}|a_{i,t}-b_{i,t}|, it indicates the existence of a risk-neutral measure μ\mu, but certain options lack a feasible price ci,t[bi,t,ai,t]c_{i,t}\in[b_{i,t},a_{i,t}] under μ\mu. Although only marginals (μ1,,μN)(\mu_{1},\ldots,\mu_{N}) are required, our program in (5.2) actually obtains a joint martingale probability distribution. Consequently, it ensures the convex order of marginals automatically. Besides, it addresses bid-ask spreads without solely relying on the mid-price, i.e., the average of bid and ask prices.

5.3 Basket options

The effectiveness of McCormick relaxations depends on the characteristics of the option payoff. Since causality imposes conditional independence, we find that McCormick relaxations are more effective when the payoff function is path-dependent for both assets. Consider a basket call option with an Asian-style payoff, averaged across two stocks (X,Y)(X,Y) and two periods:

max{X1+X2+Y1+Y24K,0}.\max\left\{\frac{X_{1}+X_{2}+Y_{1}+Y_{2}}{4}-K,0\right\}. (5.7)

First, we employ probability bounds implied by marginal constraints. Although these bounds are typically loose, the McCormick MOT can still yield tighter price intervals compared to the classic MOT framework.

We obtained the zero-coupon yield curve, forward prices, and European option prices from OptionMetrics via Wharton Research Data Services (WRDS). In the first example, consider an investor interested in stocks from two pharmaceutical companies, namely Gilead Sciences (Ticker: GILD) and GSK plc (Ticker: GSK). According to OptionMetrics data, the liquidity of European options for these stocks is considered moderate. For example, on 28 February 2023, GSK options show an open interest of 4458 and a volume of 167401, while GILD options have an open interest of 4989 and a volume of 233003. We utilize Gurobi to solve the linear program (5.2) and derive the risk-neutral marginals (μ1,μ2)(\mu_{1},\mu_{2}) and (ν1,ν2)(\nu_{1},\nu_{2}). To assess the reduction in basket option price bounds, we introduce the following ratio:

McCormick MaxMcCormick MinMOT MaxMOT Min.\frac{\text{McCormick Max}-\text{McCormick Min}}{\text{MOT Max}-\text{MOT Min}}. (5.8)

For the tenors, we select T1T_{1} as the maturity date closest to the current time T0=0T_{0}=0. The second maturity, T2T_{2}, is approximately one month after T1T_{1}. In the context of the payoff (5.7), the strike KK is determined as the integer rounded off from the average of forward prices with delivery dates T1T_{1} and T2T_{2}. Consequently, the basket option is close to the at-the-money level. We employ Gurobi to solve both the classic and McCormick MOT, treating them as LP problems. For the computational complexity, our problems typically have around 10,000 variables and 1,000 linear inequality constraints, which can be efficiently solved in less than a second on a laptop. Table 1 provides examples of price limits for the basket option and the ratios defined in (5.8). Considering T0T_{0} ranging from February 28, 2022, to February 28, 2023 (the most recent one-year horizon in WRDS), we derive ratio values and present a histogram in Figure 1. The average ratio is approximately 96%, which signifies a 4% reduction in price bounds. In the best-case scenario, the bounds are reduced by 20%. Our code and Excel files detailing the ratios are accessible at https://github.com/hanbingyan/McCormick.

When considering stocks with a more liquid option market, the reduction in the price gap is less pronounced. As an illustration, on February 28, 2023, European options for JPMorgan Chase (JPM) reported an open interest of 54,034 and a volume of 1,189,048, while European options for Morgan Stanley (MS) exhibited an open interest of 22,885 and a volume of 669,638. The corresponding results are presented in Table 2, and Figure 2 provides the histogram of the ratios. On average, the application of the McCormick MOT results in a 1% reduction in the bounds. This outcome can be attributed to larger supports of risk-neutral densities for JPM and MS.

T0T_{0} T1T_{1} T2T_{2} Strike MOT Max MOT Min McCormick Max McCormick Min Ratio
2022-02-28 2022-03-04 2022-04-01 50 1.88583 1.35447 1.88414 1.35823 0.98974
2022-04-26 2022-04-29 2022-05-27 52 2.03708 1.44786 2.03535 1.44919 0.99482
2022-06-23 2022-06-24 2022-07-22 52 1.30432 0.87923 1.26458 0.88730 0.88752
2022-08-26 2022-09-02 2022-09-30 47 1.18972 0.87830 1.17176 0.89783 0.87965
2022-10-24 2022-10-28 2022-11-25 50 1.48902 0.66430 1.48362 0.66841 0.98848
2022-12-20 2022-12-23 2023-01-20 59 1.66154 1.08860 1.64943 1.09253 0.97201
2023-02-17 2023-02-24 2023-03-24 60 1.19966 0.68178 1.19715 0.68867 0.98185
Table 1: Basket option on GILD and GSK. Time is in the Year-Month-Day format.
T0T_{0} T1T_{1} T2T_{2} Strike MOT Max MOT Min McCormick Max McCormick Min Ratio
2022-02-28 2022-03-04 2022-04-01 116 3.51744 1.60546 3.51189 1.61620 0.99148
2022-04-26 2022-04-29 2022-05-27 102 3.31127 1.35499 3.30457 1.37144 0.98817
2022-06-23 2022-06-24 2022-07-22 93 2.91588 1.22635 2.90865 1.23232 0.99219
2022-08-19 2022-08-26 2022-09-23 104 2.55403 0.72075 2.53698 0.73453 0.98318
2022-10-17 2022-10-21 2022-11-18 96 3.07063 0.81261 3.06671 0.83110 0.99007
2022-12-13 2022-12-16 2023-01-13 113 2.89605 0.69713 2.89492 0.71601 0.99090
2023-02-10 2023-02-17 2023-03-17 120 2.57027 0.38493 2.56400 0.39660 0.99179
Table 2: Basket option on JPM and MS.
Refer to caption
Figure 1: Distribution of ratios, with GILD and GSK as underlying assets.
Refer to caption
Figure 2: Distribution of ratios, with JPM and MS as underlying assets.

5.4 Tighter bounds on probability masses

In previous instances, the natural lower and upper bounds LL and UU on probability masses may have been relatively loose, necessitating tighter constraints using additional information. Our methodology integrates seamlessly with other knowledge available to the agent. In this subsection, we focus on the basket option (5.7) on GILD and GSK. Suppose the agent imposes the following constraints on the coupling π\pi:

π(x1:N,y1:N)\displaystyle\pi(x_{1:N},y_{1:N}) 0.01,\displaystyle\leq 0.01, (5.9)
π(x1:N,y1:N)\displaystyle\pi(x_{1:N},y_{1:N}) ζmin{μ1(x1),,μN(xN),ν1(y1),,νN(yN)}, for (x1:N,y1:N)𝒜.\displaystyle\geq\zeta\min\{\mu_{1}(x_{1}),\ldots,\mu_{N}(x_{N}),\nu_{1}(y_{1}),\ldots,\nu_{N}(y_{N})\},\text{ for }(x_{1:N},y_{1:N})\in\mathcal{A}. (5.10)

Here, ζ>0\zeta>0 serves as a shrinking parameter, and 𝒜\mathcal{A} is a subset of paths where (5.10) holds. Empirically, there are typically more than 2,000 paths available for (x1:N,y1:N)(x_{1:N},y_{1:N}). The upper bound of 0.01 in (5.9) ensures that the coupling π\pi does not concentrate excessively on too few paths; this value is chosen arbitrarily for demonstration purposes. For the lower bound (5.10), the agent specifies that certain paths of interest should have a sufficiently large probability. To make (5.9) and (5.10) feasible, we can set ζ0.01\zeta\leq 0.01. Lacking prior knowledge of the underlying assets GILD and GSK, the set 𝒜\mathcal{A} is chosen arbitrarily for demonstration. For each support 𝒳1{\mathcal{X}}_{1}, 𝒳2{\mathcal{X}}_{2}, 𝒴1{\mathcal{Y}}_{1}, and 𝒴2{\mathcal{Y}}_{2}, we select one-third of the possible stock prices in an equally spaced manner, resulting in the set 𝒜\mathcal{A} containing at most 1/811.23%1/81\approx 1.23\% of the paths. The set of McCormick martingale couplings (μ¯,ν¯;L,U){\mathcal{M}}(\bar{\mu},\bar{\nu};L,U) can be empty after imposing (5.9) and (5.10). However, the existence of a McCormick martingale coupling can be numerically checked using LP solvers such as Gurobi.

Table 3 presents the statistics of the ratios obtained under different shrinking parameters ζ\zeta and maturities T2T_{2}. We consider ζ{0.005,0.01}\zeta\in\{0.005,0.01\} for simplicity. The second maturity date T2T_{2} can be 1, 2, 3 or 4 weeks after the first maturity date T1T_{1}, where T1T_{1} is the maturity date closest to the current time T0=0T_{0}=0. Overall, the McCormick MOT yields much better price intervals when tighter bounds LL and UU are available. On average, the eight cases in Table 3 achieve a reduction of 12.66%. The best instance, with a minimum ratio of 0.1908, indicates a reduction of more than 80%. Notably, improved constraints LL and UU have already led to tighter price intervals with the MOT method. The reductions in Table 3 demonstrate that McCormick relaxations can further enhance these intervals, primarily because the inequalities like (4.4) become more effective. Besides, the performance of McCormick relaxations is better when the maturity T2T_{2} is shorter since the number of paths is smaller, and the upper bound (5.9) is more likely to be binding. Generally, a larger parameter ζ\zeta results in greater improvements in price bounds. However, this is not always the case, as the set (μ¯,ν¯;L,U){\mathcal{M}}(\bar{\mu},\bar{\nu};L,U) is more likely to be empty with larger ζ\zeta.

ζ\zeta Index for T2T_{2} Ratios Max Ratios Mean Ratios Min
0.005 1 0.9827 0.8767 0.5749
0.005 2 0.9642 0.8398 0.3441
0.005 3 0.9929 0.8985 0.5748
0.005 4 0.9892 0.9165 0.4718
0.01 1 0.9802 0.8476 0.1908
0.01 2 0.9559 0.8452 0.3796
0.01 3 0.9895 0.8604 0.3247
0.01 4 0.9852 0.9025 0.4542
Table 3: Basket option on GILD and GSK with tighter bounds (5.9) and (5.10).

Although Eckstein et al., (2021) used a different approach by incorporating additional maturities, a comparative analysis of reductions with their findings provides information on the significance of our results. In Eckstein et al., (2021, Table 4.3), focusing on uniform marginals and spread options, the reported price intervals are [0.335,8.337][0.335,8.337] for a single period, [0.416,8.254][0.416,8.254] for two periods and [0.776,7.920][0.776,7.920] for four periods. Consequently, in their four-period scenario, the calculated ratio is (7.9200.776)/(8.3370.335)=89.28%(7.920-0.776)/(8.337-0.335)=89.28\%, falling within the range observed in our Table 3.

Acknowledgments

This research began when Bingyan Han was a postdoctoral researcher in the Department of Mathematics at the University of Michigan. He expresses gratitude to the University of Michigan for providing support and an atmosphere conducive to this work.

References

  • Acciaio et al., (2021) Acciaio, B., Backhoff-Veraguas, J., and Jia, J. (2021). Cournot–Nash equilibrium and optimal transport in a dynamic setting. SIAM Journal on Control and Optimization, 59(3):2273–2300.
  • Acciaio et al., (2020) Acciaio, B., Backhoff-Veraguas, J., and Zalashko, A. (2020). Causal optimal transport and its links to enlargement of filtrations and continuous-time stochastic optimization. Stochastic Processes and their Applications, 130(5):2918–2953.
  • Alfonsi et al., (2017) Alfonsi, A., Corbetta, J., and Jourdain, B. (2017). Sampling of probability measures in the convex order and approximation of martingale optimal transport problems. Available at SSRN 3072356.
  • Anstreicher, (2009) Anstreicher, K. M. (2009). Semidefinite programming versus the reformulation-linearization technique for nonconvex quadratically constrained quadratic programming. Journal of Global Optimization, 43:471–484.
  • Audet et al., (2000) Audet, C., Hansen, P., Jaumard, B., and Savard, G. (2000). A branch and cut algorithm for nonconvex quadratically constrained quadratic programming. Mathematical Programming, 87:131–152.
  • Backhoff-Veraguas et al., (2020) Backhoff-Veraguas, J., Bartl, D., Beiglböck, M., and Eder, M. (2020). Adapted Wasserstein distances and stability in mathematical finance. Finance and Stochastics, 24(3):601–632.
  • Backhoff-Veraguas et al., (2017) Backhoff-Veraguas, J., Beiglböck, M., Lin, Y., and Zalashko, A. (2017). Causal transport in discrete time and applications. SIAM Journal on Optimization, 27(4):2528–2562.
  • Backhoff-Veraguas and Pammer, (2022) Backhoff-Veraguas, J. and Pammer, G. (2022). Stability of martingale optimal transport and weak optimal transport. The Annals of Applied Probability, 32(1):721–752.
  • Backhoff-Veraguas and Zhang, (2023) Backhoff-Veraguas, J. and Zhang, X. (2023). Dynamic Cournot-Nash equilibrium: the non-potential case. Mathematics and Financial Economics, 17(2):153–174.
  • Bayraktar et al., (2021) Bayraktar, E., Zhang, X., and Zhou, Z. (2021). Transport plans with domain constraints. Applied Mathematics & Optimization, 84(1):1131–1158.
  • Beiglböck et al., (2013) Beiglböck, M., Henry-Labordere, P., and Penkner, F. (2013). Model-independent bounds for option prices – a mass transport approach. Finance and Stochastics, 17:477–501.
  • Beiglböck and Juillet, (2016) Beiglböck, M. and Juillet, N. (2016). On a problem of optimal transport under marginal martingale constraints. Annals of Probability, 44(1):42–106.
  • Beiglböck et al., (2019) Beiglböck, M., Lim, T., and Obłój, J. (2019). Dual attainment for the martingale transport problem. Bernoulli, 25(3):1640–1658.
  • Beiglböck et al., (2017) Beiglböck, M., Nutz, M., and Touzi, N. (2017). Complete duality for martingale optimal transport on the line. The Annals of Probability, 45(5):3038–3074.
  • Bertsekas and Shreve, (1978) Bertsekas, D. and Shreve, S. E. (1978). Stochastic optimal control: The discrete-time case. Academic Press.
  • Bogachev, (2007) Bogachev, V. I. (2007). Measure Theory, volume II. Springer Science & Business Media.
  • Bogachev, (2022) Bogachev, V. I. (2022). Kantorovich problems with a parameter and density constraints. Siberian Mathematical Journal, 63(1):34–47.
  • Breeden and Litzenberger, (1978) Breeden, D. T. and Litzenberger, R. H. (1978). Prices of state-contingent claims implicit in option prices. Journal of Business, pages 621–651.
  • Brückerhoff and Juillet, (2022) Brückerhoff, M. and Juillet, N. (2022). Instability of martingale optimal transport in dimension d2\mathrm{d}\geq 2. Electronic Communications in Probability, 27(none):1 – 10.
  • Davis and Hobson, (2007) Davis, M. H. and Hobson, D. G. (2007). The range of traded option prices. Mathematical Finance, 17(1):1–14.
  • De Marco and Henry-Labordere, (2015) De Marco, S. and Henry-Labordere, P. (2015). Linking vanillas and VIX options: A constrained martingale optimal transport problem. SIAM Journal on Financial Mathematics, 6(1):1171–1194.
  • Eckstein et al., (2021) Eckstein, S., Guo, G., Lim, T., and Obłój, J. (2021). Robust pricing and hedging of options on multiple assets and its numerics. SIAM Journal on Financial Mathematics, 12(1):158–188.
  • Eckstein and Pammer, (2024) Eckstein, S. and Pammer, G. (2024). Computational methods for adapted optimal transport. The Annals of Applied Probability, 34(1A):675–713.
  • Fahim and Huang, (2016) Fahim, A. and Huang, Y.-J. (2016). Model-independent superhedging under portfolio constraints. Finance and Stochastics, 20:51–81.
  • Fengler, (2009) Fengler, M. R. (2009). Arbitrage-free smoothing of the implied volatility surface. Quantitative Finance, 9(4):417–428.
  • Guo and Obłój, (2019) Guo, G. and Obłój, J. (2019). Computational methods for martingale optimal transport problems. The Annals of Applied Probability, 29(6):3311–3347.
  • Guyon et al., (2017) Guyon, J., Menegaux, R., and Nutz, M. (2017). Bounds for VIX futures given S&P 500 smiles. Finance and Stochastics, 21:593–630.
  • Hobson and Neuberger, (2012) Hobson, D. and Neuberger, A. (2012). Robust bounds for forward start options. Mathematical Finance, 22(1):31–56.
  • Korman and McCann, (2015) Korman, J. and McCann, R. (2015). Optimal transportation with capacity constraints. Transactions of the American Mathematical Society, 367(3):1501–1521.
  • Korman et al., (2015) Korman, J., McCann, R. J., and Seis, C. (2015). An elementary approach to linear programming duality with application to capacity constrained transport. Journal of Convex Analysis, 22(3):797–808.
  • Lassalle, (2013) Lassalle, R. (2013). Causal transference plans and their Monge-Kantorovich problems. arXiv preprint arXiv:1303.6925.
  • Lütkebohmert and Sester, (2019) Lütkebohmert, E. and Sester, J. (2019). Tightening robust price bounds for exotic derivatives. Quantitative Finance, 19(11):1797–1815.
  • McCormick, (1976) McCormick, G. P. (1976). Computability of global solutions to factorable nonconvex programs: Part I – Convex underestimating problems. Mathematical Programming, 10(1):147–175.
  • Pflug and Pichler, (2012) Pflug, G. C. and Pichler, A. (2012). A distance for multistage stochastic optimization models. SIAM Journal on Optimization, 22(1):1–23.
  • Pflug and Pichler, (2014) Pflug, G. C. and Pichler, A. (2014). Multistage stochastic optimization, volume 1104. Springer.
  • Rogers and Williams, (1993) Rogers, L. and Williams, D. (1993). Diffusions, Markov Processes, and Martingales: Volume 1, Foundations. Cambridge University Press.
  • Sherali and Fraticelli, (2002) Sherali, H. D. and Fraticelli, B. M. (2002). Enhancing RLT relaxations via a new class of semidefinite cuts. Journal of Global Optimization, 22:233–261.
  • Strassen, (1965) Strassen, V. (1965). The existence of probability measures with given marginals. The Annals of Mathematical Statistics, 36(2):423–439.
  • Strasser, (1985) Strasser, H. (1985). Mathematical theory of statistics: Statistical experiments and asymptotic decision theory, volume 7. Walter de Gruyter.
  • Villani, (2003) Villani, C. (2003). Topics in Optimal Transportation. Number 58. American Mathematical Society.
  • Wiesel, (2023) Wiesel, J. (2023). Continuity of the martingale optimal transport problem on the real line. The Annals of Applied Probability, 33(6A):4645–4692.

Appendix A Proofs of results

Proof of Proposition 3.2.

Define fh(x1:N):=ht(y1:t)π(dy1:t|x1:N)f^{h}(x_{1:N}):=\int h_{t}(y_{1:t})\pi(dy_{1:t}|x_{1:N}) with htCb(𝒴1:t)h_{t}\in C_{b}({\mathcal{Y}}_{1:t}). It is worth noting that fh(x1:N)f^{h}(x_{1:N}) is defined π\pi-almost surely. A coupling πΠ(μ¯,ν¯)\pi\in\Pi(\bar{\mu},\bar{\nu}) is causal if and only if

fh(x1:N)=fh(x1:t,x¯t+1:N)π(dx¯t+1:N|x1:t),π-a.s.f^{h}(x_{1:N})=\int f^{h}(x_{1:t},\bar{x}_{t+1:N})\pi(d\bar{x}_{t+1:N}|x_{1:t}),\quad\pi\text{-a.s.} (A.1)

for all 1tN1\leq t\leq N and all fhf^{h} defined with htCb(𝒴1:t)h_{t}\in C_{b}({\mathcal{Y}}_{1:t}). We emphasize that (A.1) holds π\pi-almost surely, and the conditional kernel π(dx¯t+1:N|x1:t)\pi(d\bar{x}_{t+1:N}|x_{1:t}) depends on the choice of π\pi since the joint distribution of X1:NX_{1:N} is not fixed.

Furthermore, (A.1) holds if and only if for any gtCb(𝒳1:N)g_{t}\in C_{b}({\mathcal{X}}_{1:N}), we have

gt(x1:N)[fh(x1:N)fh(x1:t,x¯t+1:N)π(dx¯t+1:N|x1:t)]π(dx1:N)=0.\displaystyle\int g_{t}(x_{1:N})\left[f^{h}(x_{1:N})-\int f^{h}(x_{1:t},\bar{x}_{t+1:N})\pi(d\bar{x}_{t+1:N}|x_{1:t})\right]\pi(dx_{1:N})=0. (A.2)

Interchanging the expectations on x1:Nx_{1:N} and x¯t+1:N\bar{x}_{t+1:N} yields

fh(x1:N)[gt(x1:N)gt(x1:t,x¯t+1:N)π(dx¯t+1:N|x1:t)]π(dx1:N)=0.\displaystyle\int f^{h}(x_{1:N})\left[g_{t}(x_{1:N})-\int g_{t}(x_{1:t},\bar{x}_{t+1:N})\pi(d\bar{x}_{t+1:N}|x_{1:t})\right]\pi(dx_{1:N})=0. (A.3)

Subsequently, utilizing the tower property, (A.3) is equivalent to

ht(y1:t)[gt(x1:N)gt(x1:t,x¯t+1:N)π(dx¯t+1:N|x1:t)]π(dx1:N,dy1:N)=0.\displaystyle\int h_{t}(y_{1:t})\left[g_{t}(x_{1:N})-\int g_{t}(x_{1:t},\bar{x}_{t+1:N})\pi(d\bar{x}_{t+1:N}|x_{1:t})\right]\pi(dx_{1:N},dy_{1:N})=0. (A.4)

Proof of Theorem 4.5.

If we substitute the cost cc with c+C(1+i=1N|xi|+j=1N|yj|)c+C(1+\sum^{N}_{i=1}|x_{i}|+\sum^{N}_{j=1}|y_{j}|), it only introduces a finite constant to the original problem, thanks to the finite first moments and marginal conditions. Consequently, we can assume c0c\geq 0 henceforth.

We endow L1(N×N)L^{1}(\mathbb{R}^{N}\times\mathbb{R}^{N}) with the weak topology, where fnf_{n} converges to ff in the weak topology if and only if

fnv𝑑λfv𝑑λ,vL(N×N).\int f_{n}vd\lambda\rightarrow\int fvd\lambda,\quad\forall\,v\in L^{\infty}(\mathbb{R}^{N}\times\mathbb{R}^{N}).

Since every function in (g¯,h¯;l,u){\mathcal{M}}(\bar{g},\bar{h};l,u) is bounded by the same integrable function uu, Bogachev, (2007, Theorem 4.7.20 (v)) shows that (g¯,h¯;l,u){\mathcal{M}}(\bar{g},\bar{h};l,u) has a compact closure under the weak topology of L1(N×N)L^{1}(\mathbb{R}^{N}\times\mathbb{R}^{N}). If we can demonstrate that (g¯,h¯;l,u){\mathcal{M}}(\bar{g},\bar{h};l,u) is closed, then it is also compact.

Consider a sequence of functions {fn}\{f_{n}\} in (g¯,h¯;l,u){\mathcal{M}}(\bar{g},\bar{h};l,u) that converges to ff in the weak topology of L1(N×N)L^{1}(\mathbb{R}^{N}\times\mathbb{R}^{N}). This convergence is expressed as:

limnfnv𝑑λ=fv𝑑λ.\lim_{n\rightarrow\infty}\int f_{n}vd\lambda=\int fvd\lambda.

First, select v=𝟏Av=\mathbf{1}_{A} with a measurable set AA in N×N\mathbb{R}^{N}\times\mathbb{R}^{N}. Since it holds that

fn(x1:N,y1:N)v(x1:N,y1:N)𝑑λu(x1:N,y1:N)v(x1:N,y1:N)𝑑λ,\int f_{n}(x_{1:N},y_{1:N})v(x_{1:N},y_{1:N})d\lambda\leq\int u(x_{1:N},y_{1:N})v(x_{1:N},y_{1:N})d\lambda,

we can obtain that fuf\leq u λ\lambda-a.e. Similarly, we can show flf\geq l, λ\lambda-a.e. Next, consider v(x1:N,y1:N)=ψ(x1)L()v(x_{1:N},y_{1:N})=\psi(x_{1})\in L^{\infty}(\mathbb{R}). Then, f(x1:N,y1:N)f(x_{1:N},y_{1:N}) has the marginal g1g_{1} on x1x_{1}. Other marginal constraints can also be verified. To establish that ff satisfies the martingale constraint, we observe that

|N×Nfn(x1:N,y1:N)αt(x1:t,y1:t)(xt+1xt)dλ\displaystyle\Big{|}\int_{\mathbb{R}^{N}\times\mathbb{R}^{N}}f_{n}(x_{1:N},y_{1:N})\alpha_{t}(x_{1:t},y_{1:t})(x_{t+1}-x_{t})d\lambda
N×Nf(x1:N,y1:N)αt(x1:t,y1:t)(xt+1xt)dλ|\displaystyle\qquad-\int_{\mathbb{R}^{N}\times\mathbb{R}^{N}}f(x_{1:N},y_{1:N})\alpha_{t}(x_{1:t},y_{1:t})(x_{t+1}-x_{t})d\lambda\Big{|}
|Kfn(x1:N,y1:N)αt(x1:t,y1:t)(xt+1xt)𝑑λKf(x1:N,y1:N)αt(x1:t,y1:t)(xt+1xt)𝑑λ|\displaystyle\leq\Big{|}\int_{K}f_{n}(x_{1:N},y_{1:N})\alpha_{t}(x_{1:t},y_{1:t})(x_{t+1}-x_{t})d\lambda-\int_{K}f(x_{1:N},y_{1:N})\alpha_{t}(x_{1:t},y_{1:t})(x_{t+1}-x_{t})d\lambda\Big{|}
+|(N×N)\Kfn(x1:N,y1:N)αt(x1:t,y1:t)(xt+1xt)dλ\displaystyle\quad+\Big{|}\int_{(\mathbb{R}^{N}\times\mathbb{R}^{N})\backslash K}f_{n}(x_{1:N},y_{1:N})\alpha_{t}(x_{1:t},y_{1:t})(x_{t+1}-x_{t})d\lambda
(N×N)\Kf(x1:N,y1:N)αt(x1:t,y1:t)(xt+1xt)dλ|,\displaystyle\qquad\quad-\int_{(\mathbb{R}^{N}\times\mathbb{R}^{N})\backslash K}f(x_{1:N},y_{1:N})\alpha_{t}(x_{1:t},y_{1:t})(x_{t+1}-x_{t})d\lambda\Big{|},

where K:=[a,a]N×[a,a]NK:=[-a,a]^{N}\times[-a,a]^{N} is a compact set in N×N\mathbb{R}^{N}\times\mathbb{R}^{N}, and αtL\alpha_{t}\in L^{\infty}. With a given compact set KK, the first term converges to zero since xtx_{t} and xt+1x_{t+1} are bounded on KK. The second term is bounded by αtε\|\alpha_{t}\|_{\infty}\varepsilon since fnf_{n} and ff have marginals (g¯,h¯)(\bar{g},\bar{h}) with finite first moments. When the side width aa\rightarrow\infty, we have ε0\varepsilon\rightarrow 0. Consequently, ff also satisfies the martingale constraint. In summary, the limit f(g¯,h¯;l,u)f\in{\mathcal{M}}(\bar{g},\bar{h};l,u), and the compactness follows.

The proof of the lower semicontinuity for the objective is standard. When cc is bounded, the functional

fcf𝑑λf\mapsto\int cfd\lambda

is continuous, following from the definition of weak topology. In the case where cc is nonnegative but unbounded, there exists a sequence of bounded measurable functions {cn}\{c_{n}\} converging increasingly to cc. Using the monotone convergence theorem, we have

fc𝑑λ=fsupncndλ=supnfcn𝑑λ,\int fcd\lambda=\int f\sup_{n}c_{n}d\lambda=\sup_{n}\int fc_{n}d\lambda,

demonstrating that fc𝑑λ\int fcd\lambda is a supremum of continuous functionals and, consequently, l.s.c. The claim then follows as the infimum is attained for an l.s.c. functional on a compact set. ∎

Proof of Lemma 4.6.

Similarly to the proof of Villani, (2003, Proposition 1.22), we can substitute the cost cc for c+C(1+|x|+|y|)+εc+C(1+|x|+|y|)+\varepsilon for an arbitrarily large ε>0\varepsilon>0. With the finite first moments and marginal conditions, this modification adds the same finite constant to both sides of the strong duality. The result for the modified cost implies the result for the original cost function cc. Hence, we proceed with the assumption cε>0c\geq\varepsilon>0 in the subsequent analysis.

Define the functionals Ξ\Xi and Θ\Theta as follows:

Ξ:\displaystyle\Xi: (b(x,y),κ(x,y))Cb(×)×Cb(×)\displaystyle(b(x,y),\kappa(x,y))\in C_{b}(\mathbb{R}\times\mathbb{R})\times C_{b}(\mathbb{R}\times\mathbb{R}) (A.5)
{ϕ(x)g(x)𝑑x+φ(y)h(y)𝑑yκ(x,y)u(x,y)𝑑x𝑑y, if b(x,y)=ϕ(x)+φ(y) and κ(x,y)0;+, else. \displaystyle\mapsto\left\{\begin{array}[]{lll}&\int\phi(x)g(x)dx+\int\varphi(y)h(y)dy-\int\kappa(x,y)u(x,y)dxdy,\\ &\text{ if }b(x,y)=\phi(x)+\varphi(y)\text{ and }\kappa(x,y)\leq 0;\\ &+\infty,\text{ else. }\\ \end{array}\right.

and

Θ:\displaystyle\Theta: (b(x,y),κ(x,y))Cb(×)×Cb(×)\displaystyle(b(x,y),\kappa(x,y))\in C_{b}(\mathbb{R}\times\mathbb{R})\times C_{b}(\mathbb{R}\times\mathbb{R}) (A.6)
{0, if b(x,y)c(x,y)+κ(x,y) and κ(x,y)0;+, else. \displaystyle\mapsto\left\{\begin{array}[]{lll}&0,\text{ if }b(x,y)\geq-c(x,y)+\kappa(x,y)\text{ and }\kappa(x,y)\leq 0;\\ &+\infty,\text{ else. }\\ \end{array}\right.

Ξ\Xi is well-defined. Specifically, if ϕ(x)+φ(y)=ϕ~(x)+φ~(y)\phi(x)+\varphi(y)=\tilde{\phi}(x)+\tilde{\varphi}(y) for all xx and yy, then ϕ=ϕ~+s\phi=\tilde{\phi}+s and φ=φ~s\varphi=\tilde{\varphi}-s with a constant ss; see Villani, (2003, Section 1.1, p.27) for the same argument. Analogously, Θ\Theta is also well-defined.

To verify the assumptions in Fenchel-Rockafellar duality theorem (Villani,, 2003, Theorem 1.9), we observe that Ξ\Xi and Θ\Theta are finite at (b,κ)=(0,1)(b,\kappa)=(0,-1). Moreover, under conditions bε\|b\|_{\infty}\leq\varepsilon and κ+112\|\kappa+1\|_{\infty}\leq\frac{1}{2}, it follows that κ0\kappa\leq 0 and c+κεb-c+\kappa\leq-\varepsilon\leq b. Consequently, Θ(0,1)=0\Theta(0,-1)=0, and therefore Θ\Theta is continuous at (0,1)(0,-1).

A direct calculation yields

inf(b,κ)Cb×Cb[Θ(b,κ)+Ξ(b,κ)]\displaystyle\inf_{(b,\kappa)\in C_{b}\times C_{b}}\Big{[}\Theta(b,\kappa)+\Xi(b,\kappa)\Big{]}
=inf(ϕ,φ,κ)Cb×Cb×Cb[ϕ(x)g(x)dx+φ(y)h(y)dyκ(x,y)u(x,y)dxdy|\displaystyle=\inf_{(\phi,\varphi,\kappa)\in C_{b}\times C_{b}\times C_{b}}\Big{[}\int\phi(x)g(x)dx+\int\varphi(y)h(y)dy-\int\kappa(x,y)u(x,y)dxdy\Big{|}
ϕ(x)+φ(y)c(x,y)+κ(x,y),κ0]\displaystyle\hskip 113.81102pt\phi(x)+\varphi(y)\geq-c(x,y)+\kappa(x,y),\kappa\leq 0\Big{]}
=sup(ϕ,φ,κ)Cb×Cb×Cb[ϕ(x)g(x)dx+φ(y)h(y)dyκ(x,y)u(x,y)dxdy|\displaystyle=-\sup_{(\phi,\varphi,\kappa)\in C_{b}\times C_{b}\times C_{b}}\Big{[}\int\phi(x)g(x)dx+\int\varphi(y)h(y)dy-\int\kappa(x,y)u(x,y)dxdy\Big{|}
ϕ(x)+φ(y)c(x,y)+κ(x,y),κ0],\displaystyle\hskip 113.81102pt\phi(x)+\varphi(y)\leq c(x,y)+\kappa(x,y),\kappa\geq 0\Big{]},

where in the last equality, we substitute (ϕ,φ,κ)(\phi,\varphi,\kappa) with (ϕ,φ,κ)(-\phi,-\varphi,-\kappa).

A challenge arises due to the fact that the topological dual of Cb(×)C_{b}(\mathbb{R}\times\mathbb{R}), denoted as (Cb(×))(C_{b}(\mathbb{R}\times\mathbb{R}))^{*}, is greater than M(×)M(\mathbb{R}\times\mathbb{R}), which represents the set of all finite signed measures. To calculate the Legendre-Fenchel transforms of Θ\Theta and Ξ\Xi, we take advantage of Villani, (2003, Lemmas 1.24 and 1.25) to deal with continuous linear functionals on Cb(×)C_{b}(\mathbb{R}\times\mathbb{R}). In addition, we note that the topological dual of the product space Cb(×)×Cb(×)C_{b}(\mathbb{R}\times\mathbb{R})\times C_{b}(\mathbb{R}\times\mathbb{R}) is the product of dual spaces, namely (Cb(×))×(Cb(×))(C_{b}(\mathbb{R}\times\mathbb{R}))^{*}\times(C_{b}(\mathbb{R}\times\mathbb{R}))^{*}.

With linear forms (l1,l2)(Cb(×))×(Cb(×))(l_{1},l_{2})\in(C_{b}(\mathbb{R}\times\mathbb{R}))^{*}\times(C_{b}(\mathbb{R}\times\mathbb{R}))^{*}, the Legendre-Fenchel transform of Θ\Theta is defined as follows:

Θ(l1,l2)\displaystyle\Theta^{*}(-l_{1},-l_{2})
=sup(b,κ)Cb×Cb[l1,b+l2,κ|b(x,y)c(x,y)+κ(x,y),κ(x,y)0]\displaystyle=\sup_{(b,\kappa)\in C_{b}\times C_{b}}\Big{[}\left\langle-l_{1},b\right\rangle+\left\langle-l_{2},\kappa\right\rangle\Big{|}b(x,y)\geq-c(x,y)+\kappa(x,y),\;\kappa(x,y)\leq 0\Big{]}
=sup(b,κ)Cb×Cb[l1,bκ+l1+l2,κ|b(x,y)c(x,y)+κ(x,y),κ(x,y)0],\displaystyle=\sup_{(b,\kappa)\in C_{b}\times C_{b}}\Big{[}\left\langle l_{1},b-\kappa\right\rangle+\left\langle l_{1}+l_{2},\kappa\right\rangle\Big{|}b(x,y)\leq c(x,y)+\kappa(x,y),\;\kappa(x,y)\geq 0\Big{]},

where ,\left\langle\cdot,\cdot\right\rangle denotes the duality bracket (Villani,, 2003, p. xiv).

Next, we identify the conditions for Θ(l1,l2)<+\Theta^{*}(-l_{1},-l_{2})<+\infty and compute Θ(l1,l2)\Theta^{*}(-l_{1},-l_{2}). If there exists v0v\geq 0 satisfying l1+l2,v>0\left\langle l_{1}+l_{2},v\right\rangle>0, then we can choose b=κ=kvb=\kappa=kv and send kk\rightarrow\infty. Similarly, if there exists v0v\geq 0 and l1,v<0\left\langle l_{1},v\right\rangle<0, then we can consider κ=0\kappa=0 and b=kvb=-kv with kk\rightarrow\infty. Therefore, Θ(l1,l2)<+\Theta^{*}(-l_{1},-l_{2})<+\infty only when l1l_{1} is nonnegative and l1+l2l_{1}+l_{2} is nonpositive.

Furthermore, the Legendre-Fenchel transform of Ξ\Xi is given by

Ξ(l1,l2)=sup(b,κ)Cb×Cb[\displaystyle\Xi^{*}(l_{1},l_{2})=\sup_{(b,\kappa)\in C_{b}\times C_{b}}\Big{[} l1,b+l2,κϕ(x)g(x)𝑑xφ(y)h(y)𝑑y\displaystyle\left\langle l_{1},b\right\rangle+\left\langle l_{2},\kappa\right\rangle-\int\phi(x)g(x)dx-\int\varphi(y)h(y)dy
+κ(x,y)u(x,y)dxdy|b(x,y)=ϕ(x)+φ(y) and κ(x,y)0].\displaystyle+\int\kappa(x,y)u(x,y)dxdy\Big{|}b(x,y)=\phi(x)+\varphi(y)\text{ and }\kappa(x,y)\leq 0\Big{]}.

To avoid Ξ(l1,l2)=+\Xi^{*}(l_{1},l_{2})=+\infty, we first need

l1,ϕ+φ=ϕ(x)g(x)𝑑x+φ(y)h(y)𝑑y.\left\langle l_{1},\phi+\varphi\right\rangle=\int\phi(x)g(x)dx+\int\varphi(y)h(y)dy. (A.7)

It is noteworthy that l1l_{1} should be nonnegative, as argued above. Moreover, l1l_{1} also acts continuously on the subset C0(×)C_{0}(\mathbb{R}\times\mathbb{R}) of Cb(×)C_{b}(\mathbb{R}\times\mathbb{R}), where C0(×)C_{0}(\mathbb{R}\times\mathbb{R}) denotes the set of all continuous functions approaching 0 at infinity. Since the topological dual of C0(×)C_{0}(\mathbb{R}\times\mathbb{R}) is M(×)M(\mathbb{R}\times\mathbb{R}), there exists a unique measure πM(×)\pi\in M(\mathbb{R}\times\mathbb{R}) such that

l1,v=v(x,y)𝑑π,vC0(×).\left\langle l_{1},v\right\rangle=\int v(x,y)d\pi,\quad\forall\,v\in C_{0}(\mathbb{R}\times\mathbb{R}).

We can then decompose l1=π+Rl_{1}=\pi+R, where RR is a continuous linear functional with

R,v=0,vC0(×).\left\langle R,v\right\rangle=0,\quad\forall\,v\in C_{0}(\mathbb{R}\times\mathbb{R}).

Under the condition (A.7), Villani, (2003, Lemma 1.25) shows that R=0R=0, and l1=πl_{1}=\pi is a nonnegative measure with Borel marginals gdxgdx and hdyhdy.

Another requirement to avoid Ξ(l1,l2)=+\Xi^{*}(l_{1},l_{2})=+\infty is

l2,κ+κ(x,y)u(x,y)𝑑x𝑑y0,κ(x,y)0,κCb(×).\left\langle l_{2},\kappa\right\rangle+\int\kappa(x,y)u(x,y)dxdy\leq 0,\quad\forall\,\kappa(x,y)\leq 0,\,\kappa\in C_{b}(\mathbb{R}\times\mathbb{R}). (A.8)

Since l1+l2l_{1}+l_{2} is nonpositive and l1l_{1} is a measure, then

l1+l2,v=v(x,y)𝑑π+l2,v0,v(x,y)0,vCb(×).\displaystyle\left\langle l_{1}+l_{2},v\right\rangle=\int v(x,y)d\pi+\left\langle l_{2},v\right\rangle\leq 0,\quad\forall\,v(x,y)\geq 0,\,v\in C_{b}(\mathbb{R}\times\mathbb{R}). (A.9)

Together with (A.8), setting κ=v\kappa=-v yields

v(x,y)𝑑πl2,vv(x,y)u(x,y)𝑑x𝑑y,v(x,y)0,vCb(×).\int v(x,y)d\pi\leq-\left\langle l_{2},v\right\rangle\leq\int v(x,y)u(x,y)dxdy,\quad\forall\,v(x,y)\geq 0,\,v\in C_{b}(\mathbb{R}\times\mathbb{R}). (A.10)

This implies that π\pi is absolutely continuous with respect to the Lebesgue measure λ\lambda on the Borel σ\sigma-algebra of ×\mathbb{R}\times\mathbb{R}. Otherwise, suppose that the opposite is true. Then, there exists a Borel set BB such that λ(B)=0\lambda(B)=0 and π(B)>0\pi(B)>0. Since π\pi is regular (Bertsekas and Shreve,, 1978, Proposition 7.17), there exists a closed subset FBF\subset B and π(F)>δ>0\pi(F)>\delta>0 for a small δ\delta. Similar to the proof of Bertsekas and Shreve, (1978, Proposition 7.18), let Gn={(x,y)×|d((x,y),F)<1/n}G_{n}=\{(x,y)\in\mathbb{R}\times\mathbb{R}|d((x,y),F)<1/n\} with a sufficiently large nn. By Urysohn’s lemma, there exist continuous functions 0vn10\leq v_{n}\leq 1 such that vn=0v_{n}=0 on GncG^{c}_{n} and vn=1v_{n}=1 on FF. Then

π(F)vn(x,y)𝑑πvn(x,y)u(x,y)𝑑λuvn𝑑λuλ(Gn),\displaystyle\pi(F)\leq\int v_{n}(x,y)d\pi\leq\int v_{n}(x,y)u(x,y)d\lambda\leq\|u\|_{\infty}\int v_{n}d\lambda\leq\|u\|_{\infty}\lambda(G_{n}),

which implies

π(F)uλ(n=1Gn)=uλ(F)uλ(B)=0,\displaystyle\pi(F)\leq\|u\|_{\infty}\lambda(\cap^{\infty}_{n=1}G_{n})=\|u\|_{\infty}\lambda(F)\leq\|u\|_{\infty}\lambda(B)=0,

leading to a contradiction with π(F)>δ>0\pi(F)>\delta>0. By the Radon-Nikodým theorem (Rogers and Williams,, 1993, Theorem 9.3, p.98), the Radon-Nikodým density exists: f(x,y)=dπdλ,λ-a.e.f(x,y)=\frac{d\pi}{d\lambda},\,\lambda\text{-a.e.}

Furthermore, (A.10) implies f(x,y)u(x,y),λ-a.e.f(x,y)\leq u(x,y),\,\lambda\text{-a.e.} Otherwise, suppose λ(B)>0\lambda(B)>0, where B:={(x,y)×|f(x,y)>u(x,y)}B:=\{(x,y)\in\mathbb{R}\times\mathbb{R}|f(x,y)>u(x,y)\}. BB is a Borel set. Since λ\lambda is regular, there exists a closed subset FBF\subset B and λ(F)>0\lambda(F)>0. Introducing continuous and bounded functions vnv_{n} similarly as before, (A.10) and the dominated convergence theorem lead to

1Ff𝑑λvnf𝑑λvnu𝑑λ1Gnu𝑑λ1Fu𝑑λ,\displaystyle\int\textbf{1}_{F}fd\lambda\leq\int v_{n}fd\lambda\leq\int v_{n}ud\lambda\leq\int\textbf{1}_{G_{n}}ud\lambda\rightarrow\int\textbf{1}_{F}ud\lambda,

which contradicts the fact that f>uf>u on FF and λ(F)>0\lambda(F)>0.

In summary, Θ(l1,l2)+Ξ(l1,l2)<+\Theta^{*}(-l_{1},-l_{2})+\Xi^{*}(l_{1},l_{2})<+\infty requires that l1l_{1} is a probability measure in Π(g,h)\Pi(g,h) with density 0fu0\leq f\leq u and l1+l2l_{1}+l_{2} is nonpositive. Clearly, the reverse direction is also true: If l1l_{1} is a probability measure in Π(g,h)\Pi(g,h) with density 0fu0\leq f\leq u and l1+l2l_{1}+l_{2} is nonpositive, then Θ(l1,l2)+Ξ(l1,l2)<+\Theta^{*}(-l_{1},-l_{2})+\Xi^{*}(l_{1},l_{2})<+\infty.

Moreover, when l1l_{1} is a probability measure in Π(g,h)\Pi(g,h) with density 0fu0\leq f\leq u and l1+l2l_{1}+l_{2} is nonpositive, it follows that Ξ(l1,l2)=0\Xi^{*}(l_{1},l_{2})=0 and

Θ(l1,l2)\displaystyle\Theta^{*}(-l_{1},-l_{2}) =supbCb[b(x,y)f(x,y)𝑑x𝑑y|b(x,y)c(x,y)]=c(x,y)f(x,y)𝑑x𝑑y,\displaystyle=\sup_{b\in C_{b}}\Big{[}\int b(x,y)f(x,y)dxdy\Big{|}b(x,y)\leq c(x,y)\Big{]}=\int c(x,y)f(x,y)dxdy,

because the l.s.c. cost cεc\geq\varepsilon can be approximated pointwise by a monotonically increasing sequence of continuous and bounded functions. Then the equality above follows from the monotone convergence theorem. Therefore, the right-hand side of the Fenchel-Rockafellar duality reduces to

max(l1,l2)(Cb)×(Cb)[Θ(l1,l2)Ξ(l1,l2)]\displaystyle\max_{(l_{1},l_{2})\in(C_{b})^{*}\times(C_{b})^{*}}\Big{[}-\Theta^{*}(-l_{1},-l_{2})-\Xi^{*}(l_{1},l_{2})\Big{]} =minfΠ(g,h;u)c(x,y)f(x,y)𝑑x𝑑y.\displaystyle=-\min_{f\in\Pi(g,h;u)}\int c(x,y)f(x,y)dxdy.

We have proved the strong duality (4.20), albeit with continuous and bounded (ϕ,φ,κ)(\phi,\varphi,\kappa). In the L1L^{1} case, the claim follows from

inffΠ(g,h;u)P(f)=supΨcCbF(ϕ,φ,κ)supΨcF(ϕ,φ,κ)inffΠ(g,h;u)P(f),\displaystyle\inf_{f\in\Pi(g,h;u)}P(f)=\sup_{\Psi_{c}\cap C_{b}}F(\phi,\varphi,\kappa)\leq\sup_{\Psi_{c}}F(\phi,\varphi,\kappa)\leq\inf_{f\in\Pi(g,h;u)}P(f),

where the last inequality is a result of weak duality, similar to Villani, (2003, Proposition 1.5). In addition, in the first equality, the notation ΨcCb\Psi_{c}\cap C_{b} indicates that (ϕ,φ,κ)(\phi,\varphi,\kappa) are also continuous and bounded. ∎

Proof of Theorem 4.7.

Similarly to the proof of Lemma 4.6, we can assume the cost c0c\geq 0.

Consider a continuous and bounded cost cc. Maximization in the dual problem can be achieved in two steps. First, we maximize over (ϕ,φ,κ)(\phi,\varphi,\kappa) when other multipliers (α,β,γ,η,θ)(\alpha,\beta,\gamma,\eta,\theta) are given. This reduces to a dual problem with the upper capacity constraint in Lemma 4.6, but with a new cost given by

c~(x1:N,y1:N,α,β,γ,η,θ;u,l)\displaystyle\tilde{c}(x_{1:N},y_{1:N},\alpha,\beta,\gamma,\eta,\theta;u,l) (A.11)
=c(x1:N,y1:N)θ(x1:N,y1:N)\displaystyle=c(x_{1:N},y_{1:N})-\theta(x_{1:N},y_{1:N})
+t=1N1αt(x1:t,y1:t)(xt+1xt)+t=1N1βt(x1:t,y1:t)(yt+1yt)\displaystyle\quad+\sum^{N-1}_{t=1}\alpha_{t}(x_{1:t},y_{1:t})(x_{t+1}-x_{t})+\sum^{N-1}_{t=1}\beta_{t}(x_{1:t},y_{1:t})(y_{t+1}-y_{t})
+Ψ(x1:N,y1:N,γ,η;u,l).\displaystyle\quad+\Psi(x_{1:N},y_{1:N},\gamma,\eta;u,l).

On the basis of our assumptions, we have c,θ,α,β,γ,η,u,lCbc,\theta,\alpha,\beta,\gamma,\eta,u,l\in C_{b}. Hence, c~\tilde{c} is continuous, and |c~|C(1+t=1N|xt|+t=1N|yt|)|\tilde{c}|\leq C(1+\sum^{N}_{t=1}|x_{t}|+\sum^{N}_{t=1}|y_{t}|). An application of Lemma 4.6 leads to

D\displaystyle D =supαt,βtγt,i,ηt,i0θ0supΦc with (αt,βt,γt,i,ηt,i,θ) given D(ϕ,φ,α,β,γ,η,κ,θ)\displaystyle=\sup_{\begin{subarray}{c}\alpha_{t},\beta_{t}\\ \gamma_{t,i},\,\eta_{t,i}\geq 0\\ \theta\geq 0\end{subarray}}\sup_{\begin{subarray}{c}\Phi_{c}\text{ with }(\alpha_{t},\beta_{t},\gamma_{t,i},\,\eta_{t,i},\theta)\text{ given }\end{subarray}}D(\phi,\varphi,\alpha,\beta,\gamma,\eta,\kappa,\theta) (A.12)
=supαt,βtγt,i,ηt,i0θ0inffΠ(g¯,h¯;u)c~(x1:N,y1:N,α,β,γ,η,θ;u,l)f(x1:N,y1:N)𝑑λ\displaystyle=\sup_{\begin{subarray}{c}\alpha_{t},\beta_{t}\\ \gamma_{t,i},\,\eta_{t,i}\geq 0\\ \theta\geq 0\end{subarray}}\inf_{f\in\Pi(\bar{g},\bar{h};u)}\int\tilde{c}(x_{1:N},y_{1:N},\alpha,\beta,\gamma,\eta,\theta;u,l)f(x_{1:N},y_{1:N})d\lambda (A.13)
=inffΠ(g¯,h¯;u)supαt,βtγt,i,ηt,i0θ0c~(x1:N,y1:N,α,β,γ,η,θ;u,l)f(x1:N,y1:N)𝑑λ\displaystyle=\inf_{f\in\Pi(\bar{g},\bar{h};u)}\sup_{\begin{subarray}{c}\alpha_{t},\beta_{t}\\ \gamma_{t,i},\,\eta_{t,i}\geq 0\\ \theta\geq 0\end{subarray}}\int\tilde{c}(x_{1:N},y_{1:N},\alpha,\beta,\gamma,\eta,\theta;u,l)f(x_{1:N},y_{1:N})d\lambda (A.14)
=inff(g¯,h¯;l,u)c(x1:N,y1:N)f(x1:N,y1:N)𝑑λ=P.\displaystyle=\inf_{f\in{\mathcal{M}}(\bar{g},\bar{h};l,u)}\int c(x_{1:N},y_{1:N})f(x_{1:N},y_{1:N})d\lambda=P. (A.15)

(A.14) follows from the minimax theorem, given in Strasser, (1985, Theorem 45.8, p. 239) or Beiglböck et al., (2013, Theorem 3.1). In fact, Π(g¯,h¯;u)\Pi(\bar{g},\bar{h};u) is compact in the weak topology of L1L^{1}. Since c~\tilde{c} has a linear growth rate and the marginals have finite first moments, we can also prove that the objective c~f𝑑λ\int\tilde{c}fd\lambda is continuous in fΠ(g¯,h¯;u)f\in\Pi(\bar{g},\bar{h};u) using the weak topology of L1L^{1}, similar to the approach in Beiglböck et al., (2013, Lemma 2.2).

The final equality (A.15) is derived from the observation that any violation of the martingale condition, the McCormick relaxations, or the lower bound results in an infinite value.

In the case of a nonnegative and l.s.c. cost function cc, the strong duality can be established through a standard approximation argument. For example, see the last part of the proof presented in Beiglböck et al., (2013, Theorem 1.1). ∎