Fact-checked by Grok 2 weeks ago

Optional stopping theorem

The optional stopping theorem, also known as Doob's optional stopping theorem or Doob's optional sampling theorem, is a fundamental result in that addresses the behavior of martingales under random s. It states that, for a martingale process and an appropriate , the of the process at the stopping time equals its initial , implying that strategic stopping cannot, on average, alter the fairness of the process. Named after American mathematician , who formalized it in his seminal 1953 book Stochastic Processes, the theorem generalizes properties of martingales—stochastic processes where the of the next value, given the current information, equals the current value—to scenarios where observation or intervention ceases at a determined by the process itself. A stopping time T is a such that the event \{T = n\} depends only on the history of the process up to time n, ensuring decisions to stop are non-anticipating. The precise statement requires conditions to ensure the result holds, as naive application can fail without them. For a discrete-time martingale \{X_n\}_{n \geq 0} and T, E[X_T] = E[X_0] if at least one of the following holds: (i) T is bounded (i.e., P(T \leq N) = 1 for some fixed N); (ii) the martingale is bounded (i.e., |X_n| \leq K for some K , and T < \infty ); or (iii) E[T] < \infty and the increments are bounded in expectation (i.e., E[|X_{n+1} - X_n| \mid \mathcal{F}_n] \leq K for some K). Extensions exist for submartingales and supermartingales, as well as continuous-time processes, often requiring uniform integrability or right-continuity assumptions. Proofs typically involve constructing a stopped process X_{n \wedge T} (the value at the minimum of n and T), which remains a martingale, and applying the dominated convergence theorem to pass to the limit as n \to \infty. The theorem has profound applications across probability and related fields. In gambling theory, it demonstrates that no betting strategy can yield positive expected profit in a fair game, such as a symmetric , where stopping upon reaching a goal or ruin preserves the initial expectation—famously applied to gambler's ruin problems. In finance, it underpins models assuming asset prices follow , showing that timing trades based on past information cannot generate expected gains under the efficient market hypothesis. Other uses include analyzing random walks (e.g., expected hitting times in biased cases) and sequential hypothesis testing, where it ensures unbiased estimators under optional sampling. These insights highlight the theorem's role in preventing fallacious inferences from selective data stopping in statistical analysis.

Prerequisites

Martingales

In probability theory, a martingale is a stochastic process \{X_t\} that satisfies the property of fairness in expectation conditional on the available information up to time t. This concept, formalized by , captures situations where future expectations do not systematically deviate from the current value given the past. For discrete-time processes, consider a stochastic process \{X_n\}_{n=0}^\infty adapted to a filtration \{\mathcal{F}_n\}_{n=0}^\infty, meaning X_n is \mathcal{F}_n-measurable for each n. The process is a martingale if \mathbb{E}[X_{n+1} \mid \mathcal{F}_n] = X_n almost surely for all n \geq 0, assuming integrability \mathbb{E}[|X_n|] < \infty. In continuous time, for a process \{X_t\}_{t \geq 0} adapted to a filtration \{\mathcal{F}_t\}_{t \geq 0}, it is a martingale if \mathbb{E}[X_t \mid \mathcal{F}_s] = X_s almost surely for all $0 \leq s < t, with \mathbb{E}[|X_t|] < \infty. Related variants include submartingales, where \mathbb{E}[X_{n+1} \mid \mathcal{F}_n] \geq X_n (or the continuous analog), modeling processes with non-decreasing conditional expectations, and supermartingales, where the inequality is reversed. Martingales exhibit several fundamental properties that underscore their role in preserving information and expectation. A key feature is the conservation of unconditional expectation: \mathbb{E}[X_n] = \mathbb{E}[X_0] for all n, which follows from the tower property of conditional expectations applied iteratively. Additionally, for fixed (non-random) times m \leq n, the optional sampling property holds: \mathbb{E}[X_n \mid \mathcal{F}_m] = X_m, extending the martingale equality to future fixed horizons. These properties ensure that martingales maintain a balanced outlook over time without bias from past observations. Classic examples illustrate martingales in simple settings. The position S_n = \sum_{i=1}^n \xi_i in a simple symmetric on the integers, where each \xi_i = \pm 1 with equal probability $1/2 independently, forms a martingale because \mathbb{E}[S_{n+1} \mid \mathcal{F}_n] = S_n + \mathbb{E}[\xi_{n+1}] = S_n. Similarly, in a fair gambling game where a player wagers on independent with even odds and probability $1/2 of winning or losing a fixed unit amount per bet, the player's fortune process after each round is a martingale, reflecting the absence of any edge.

Stopping Times

In stochastic processes, a stopping time \tau is a random variable taking values in the extended non-negative reals \mathbb{T} \cup \{\infty\}, where \mathbb{T} is the time index set (discrete or continuous), such that the event \{\tau \leq t\} belongs to the filtration \mathcal{F}_t for every t \in \mathbb{T}. This condition ensures that the decision to stop by time t depends only on information available up to that time, preventing lookahead bias in the observation of the process. The filtration \{\mathcal{F}_t\}_{t \in \mathbb{T}} is an increasing family of \sigma-algebras on the probability space, representing the progressively refined information sets as time evolves: \mathcal{F}_s \subseteq \mathcal{F}_t for s \leq t. Each \mathcal{F}_t captures the observable events up to time t, allowing stopping times to model flexible, data-dependent termination points in processes like random walks or diffusions without anticipating future outcomes. Common examples include the first of a barrier in a random walk, such as the first time a symmetric random walk reaches a positive level a > 0, defined as \tau = \inf\{n \geq 0: S_n \geq a\}, where S_n is the walk position; this satisfies the property because whether the barrier is hit by step n is determined by the path up to n. Bounded stopping rules, like stopping at the minimum of a and a fixed horizon N, \tau = \min(\inf\{n: S_n \geq a\}, N), also qualify, as the bound ensures the event \{\tau \leq t\} remains \mathcal{F}_t-measurable for t \leq N. Key properties of stopping times include almost sure finiteness, where \mathbb{P}(\tau < \infty) = 1, which holds for many applications like recurrent random walks hitting any finite barrier, ensuring the process terminates with probability one. For bounded stopping times, where \tau \leq K almost surely for some finite K, optional sampling applies to martingales, preserving expectations at stopping: if M_t is a martingale, then \mathbb{E}[M_\tau] = \mathbb{E}[M_0].

Formal Statement

Discrete-Time Version

The optional stopping theorem in its discrete-time formulation addresses the behavior of martingale processes at random stopping times. For a discrete-time martingale (X_n)_{n \geq 0} adapted to a filtration (\mathcal{F}_n)_{n \geq 0} and a stopping time \tau with respect to this filtration, the theorem asserts that \mathbb{E}[X_\tau] = \mathbb{E}[X_0] provided that \tau satisfies certain regularity conditions ensuring the expectation is well-defined and the martingale property is preserved at the stopping time. One fundamental condition is that \tau is bounded, meaning \tau \leq c almost surely for some constant c < \infty. Under this boundedness, the stopped process X_{\tau \wedge n} remains a martingale for each n, and taking limits as n \to \infty yields the equality of expectations since \tau is finite with probability 1. This condition is particularly useful in settings where the stopping time cannot exceed a fixed horizon, such as in finite-step analyses of random walks. The theorem extends to submartingales and supermartingales. For a submartingale (X_n)_{n \geq 0}, where \mathbb{E}[X_{n+1} \mid \mathcal{F}_n] \geq X_n almost surely, the optional stopping theorem states that \mathbb{E}[X_\tau] \geq \mathbb{E}[X_0] under analogous conditions, such as bounded \tau or uniform integrability of the stopped process. Similarly, for a supermartingale, where \mathbb{E}[X_{n+1} \mid \mathcal{F}_n] \leq X_n almost surely, \mathbb{E}[X_\tau] \leq \mathbb{E}[X_0] holds. These inequalities reflect the non-decreasing or non-increasing nature of expectations in these processes. When the stopping time \tau may take the value \infty with positive probability, the theorem handles this by defining X_\tau = \lim_{n \to \infty} X_{n \wedge \tau} on the event \{\tau = \infty\}, provided the limit exists almost surely (as for ). The equality or inequality then applies conditionally on \{\tau < \infty\}, with the overall expectation adjusted accordingly if \mathbb{P}(\tau < \infty) = 1.

Continuous-Time Version

The continuous-time version of the optional stopping theorem addresses stochastic processes indexed by continuous time, requiring adaptations to handle path regularity and the potential for stopping times to be infinite with positive probability. It applies to martingales, submartingales, and supermartingales adapted to a filtered probability space (\Omega, \mathcal{F}, (\mathcal{F}_t)_{t \geq 0}, P) where the filtration (\mathcal{F}_t) is right-continuous, meaning \mathcal{F}_t = \bigcap_{s > t} \mathcal{F}_s for each t \geq 0. The processes are assumed to have (right-continuous with left limits) paths , ensuring well-defined values at stopping times. For a martingale (X_t)_{t \geq 0}, and any \tau, the stopped process (X_{t \wedge \tau})_{t \geq 0} is also a martingale. Consequently, \mathbb{E}[X_{\tau \wedge t} \mid \mathcal{F}_s] = X_{s \wedge \tau}, \quad 0 \leq s \leq t, and, unconditionally, \mathbb{E}[X_{\tau \wedge t}] = \mathbb{E}[X_0], \quad t \geq 0. This holds without further restrictions on \tau, as the right-continuity of the and property of the paths guarantee the optional sampling property for the minimum of deterministic and random times. To extend the result to the unrestricted stopping time, additional conditions are required for the equality \mathbb{E}[X_\tau] = \mathbb{E}[X_0]. These include \tau being bounded (i.e., there exists T < \infty such that P(\tau \leq T) = 1), or the family of random variables \{X_{t \wedge \tau} : t \geq 0\} being , meaning \sup_{t \geq 0} \mathbb{E}[|X_{t \wedge \tau}| \mathbf{1}_{\{|X_{t \wedge \tau}| > K\}}] \to 0 as K \to \infty. of the original martingale (X_t) also suffices, as it implies uniform integrability of the stopped process. Under these conditions, X_\tau is integrable and \mathbb{E}[X_\tau] = \mathbb{E}[X_0]. Analogous results hold for submartingales and supermartingales, but with inequalities reflecting their respective properties. For a submartingale (X_t)_{t \geq 0} and \tau, \mathbb{E}[X_{\tau \wedge t}] \geq \mathbb{E}[X_0], \quad t \geq 0, with equality \mathbb{E}[X_\tau] \geq \mathbb{E}[X_0] under the same conditions as above (bounded \tau or of the stopped process). For a supermartingale, the inequalities reverse: \mathbb{E}[X_{\tau \wedge t}] \leq \mathbb{E}[X_0], \quad t \geq 0, and \mathbb{E}[X_\tau] \leq \mathbb{E}[X_0] under the additional conditions. These versions follow from the fact that the stopped process inherits the sub- or super-martingale property.

Conditions and Variations

Bounded and Finite Expectation Conditions

In the discrete-time setting, the optional stopping theorem holds under certain sufficient conditions that ensure the expectation at the stopping time equals the initial expectation for a martingale (X_n)_{n \geq 0}. One basic condition is that the stopping time \tau is almost surely bounded, meaning there exists a constant c < \infty such that \tau \leq c with probability 1. Under this boundedness, \tau takes only finitely many values, allowing direct application of the optional sampling theorem: the stopped process X_{\tau} is itself a martingale, so \mathbb{E}[X_\tau] = \mathbb{E}[X_0]. This condition simplifies the analysis by avoiding infinite horizons and ensuring integrability without additional assumptions. Another standard condition is that the martingale is bounded, meaning there exists a constant K < \infty such that |X_n| \leq K almost surely for all n, and the stopping time \tau is almost surely finite (P(\tau < \infty) = 1). Under these assumptions, the truncated processes X_{\tau \wedge n} are uniformly bounded by K, so X_{\tau \wedge n} \to X_\tau almost surely and |X_{\tau \wedge n}| \leq K, which is integrable. By the bounded convergence theorem, \mathbb{E}[X_{\tau \wedge n}] \to \mathbb{E}[X_\tau]. Since \mathbb{E}[X_{\tau \wedge n}] = \mathbb{E}[X_0] for each n by the martingale property, it follows that \mathbb{E}[X_\tau] = \mathbb{E}[X_0]. A slightly more general condition relaxes the boundedness of \tau while imposing restrictions on the martingale increments. Specifically, if \mathbb{E}[\tau] < \infty, \tau is almost surely finite, and the increments are bounded such that |X_{n+1} - X_n| \leq c for some constant c < \infty and all n, then the theorem applies: \mathbb{E}[X_\tau] = \mathbb{E}[X_0]. To see why, consider the truncated stopping times \tau \wedge n, which are bounded for each fixed n and thus satisfy the previous condition, yielding \mathbb{E}[X_{\tau \wedge n}] = \mathbb{E}[X_0]. As n \to \infty, X_{\tau \wedge n} \to X_\tau almost surely since \tau < \infty a.s. Moreover, |X_{\tau \wedge n}| \leq |X_0| + c \tau, and the right-hand side has finite expectation because \mathbb{E}[|X_0| + c \tau] < \infty. By the dominated convergence theorem, \mathbb{E}[X_{\tau \wedge n}] \to \mathbb{E}[X_\tau], so \mathbb{E}[X_\tau] = \mathbb{E}[X_0]. This truncation approach extends the theorem to handle potentially unbounded stopping times under the finite expectation and bounded increment assumptions, providing a practical tool for verification in discrete martingale problems. These conditions are introductory and sufficient for many cases, though more advanced versions invoke uniform integrability for broader applicability.

Uniform Integrability Condition

The uniform integrability condition provides a key generalization of the , enabling its application to unbounded stopping times \tau without requiring \mathbb{E}[\tau] < \infty or boundedness of the process. For a martingale (X_t)_{t \geq 0} and stopping time \tau, the family \{X_{\tau \wedge t} : t \geq 0\} is said to be uniformly integrable if \sup_{t \geq 0} \mathbb{E}\left[ |X_{\tau \wedge t}| \mathbf{1}_{\{|X_{\tau \wedge t}| > K\}} \right] \to 0 \quad \text{as} \quad K \to \infty. This condition ensures that the tails of the distributions of the stopped processes are controlled uniformly in t. Under this , since X_{\tau \wedge t} \to X_\tau as t \to \infty and \{X_{\tau \wedge t} : t \geq 0\} is uniformly integrable, Vitali's convergence theorem implies that X_{\tau \wedge t} \to X_\tau in L^1, so \mathbb{E}[X_{\tau \wedge t}] \to \mathbb{E}[X_\tau]. But \mathbb{E}[X_{\tau \wedge t}] = \mathbb{E}[X_0] for each t by the martingale property, yielding \mathbb{E}[X_\tau] = \mathbb{E}[X_0]. Examples of uniformly integrable families include those from bounded martingales, where |X_t| \leq M for some constant M, making the indicator term zero for K > M. Another case arises for L^1-bounded martingales, where \sup_t \mathbb{E}[|X_t|] < \infty, which implies uniform integrability by de la Vallée Poussin's theorem applied to the family. Unlike conditions requiring bounded stopping times or finite \mathbb{E}[\tau], uniform integrability accommodates cases where \mathbb{E}[\tau] = \infty, broadening applicability to processes like certain random walks or diffusions with potentially infinite stopping times.

Applications and Examples

Gambling and Fair Games

In the context of gambling, the optional stopping theorem underscores the fairness of games where the fortune process forms a martingale, meaning the expected future fortune equals the current fortune given past outcomes. This property models scenarios where each bet has zero expected gain, such as fair coin flips with equal win and loss probabilities, preventing any systematic advantage through strategic stopping. A classic illustration is the gambler's ruin problem, where a player starts with initial stake k and bets $1 per round on fair coin flips, stopping upon reaching 0 (ruin) or a goal N > k (success). The fortune X_t at time t is a martingale because \mathbb{E}[X_{t+1} \mid \mathcal{F}_t] = X_t, where \mathcal{F}_t is the filtration of information up to t. By the optional stopping theorem applied to the bounded stopping time \tau = \min\{t : X_t = 0 \text{ or } X_t = N\}, the expected fortune at stopping satisfies \mathbb{E}[X_\tau] = k, implying no net expected gain despite optional cessation rules. When the house imposes a finite limit, such as total N, the remains bounded, ensuring the 's conditions hold and averting paradoxes from prolonged play. In contrast, an infinite house limit allows unbounded s, potentially violating and leading to apparent contradictions where intuitive strategies seem to yield gains, though the highlights the necessity of conditions for validity. The theorem's insights trace to historical developments by Joseph Doob, whose martingale framework in the 1950s resolved variants of the , where repeated tosses yield potentially infinite payoffs but bounded betting systems fail to guarantee positive expected returns under optional stopping. In such games, martingale transforms of winnings demonstrate that the probability of reaching a high (e.g., doubling initial stake) is at most 1/2, reinforcing the impossibility of beating a fair game through timing.

Random Walks

The optional stopping theorem finds a natural application in the analysis of simple symmetric s on the integers, where the position process serves as a martingale. Consider a X_0 = 0 with steps \xi_i = \pm 1 each with probability $1/2, so X_n = \sum_{i=1}^n \xi_i. Define the \tau = \min\{n \geq 0 : |X_n| = m\} for some positive integer m, which is the first of the barriers at \pm m. Since \{X_n\} is a martingale with bounded increments and \mathbb{E}[\tau] < \infty, the optional stopping theorem implies \mathbb{E}[X_\tau] = \mathbb{E}[X_0] = 0. At stopping, X_\tau = m or X_\tau = -m with equal probability $1/2, confirming the martingale property preserves the expected position at the boundary. To derive the expected stopping time \mathbb{E}[\tau], consider the quadratic process Y_n = X_n^2 - n. This is a martingale because \mathbb{E}[X_{n+1}^2 \mid \mathcal{F}_n] = \frac{(X_n + 1)^2 + (X_n - 1)^2}{2} = X_n^2 + 1, so \mathbb{E}[Y_{n+1} \mid \mathcal{F}_n] = Y_n. The conditions for the optional stopping theorem hold, yielding \mathbb{E}[Y_\tau] = \mathbb{E}[Y_0] = 0, or \mathbb{E}[X_\tau^2] = \mathbb{E}[\tau]. Since X_\tau^2 = m^2 almost surely, it follows that \mathbb{E}[\tau] = m^2. In the unbounded case without barriers, the one-dimensional symmetric random walk is recurrent, meaning it hits any fixed level, such as +1 starting from $0, with probability $1. However, the expected hitting time \mathbb{E}[\tau_1] = \infty, where \tau_1 = \min\{n \geq 1 : X_n = 1\}. This illustrates the necessity of conditions like finite expectation or uniform integrability in the optional stopping theorem, as the unbounded stopping time violates them, preventing direct application to yield finite expectations. Variations arise with biased random walks, where steps are +1 with probability p \neq 1/2 and -1 with probability q = 1-p. If p > 1/2, the position process \{X_n\} is a submartingale, as \mathbb{E}[X_{n+1} \mid \mathcal{F}_n] = X_n + (2p - 1) > X_n. The optional stopping for submartingales then provides \mathbb{E}[X_\tau] \geq \mathbb{E}[X_0] under appropriate conditions, such as bounded stopping times. To obtain exact expectations, center the process by the drift: Z_n = X_n - (2p - 1)n forms a martingale, allowing application of the theorem to hitting times like \mathbb{E}[\tau_b] = b / (2p - 1) for b > 0.

Mathematical Finance

In , the optional stopping theorem plays a crucial role in pricing under the , where discounted asset prices form martingales, ensuring no-arbitrage conditions. Specifically, if the discounted price process S_t e^{-\int_0^t r_u du} of a traded asset is a martingale under this measure, the theorem implies that for an admissible \tau, the expected discounted price at \tau equals the initial price: \mathbb{E}^\mathbb{Q} [S_\tau e^{-\int_0^\tau r_u du}] = S_0. This preservation of expectation allows for consistent pricing of contingent claims exercised at optional times, verifying that the market remains arbitrage-free as long as the stopping time satisfies the theorem's conditions. A key application arises in the pricing of American options, which permit early exercise at any up to maturity. The value of such an option is the supremum over all \tau of the risk-neutral expectation of the discounted payoff: V_0 = \sup_{\tau} \mathbb{E}^\mathbb{Q} [e^{-\int_0^\tau r_u du} g(S_\tau)], where g is the payoff function. If holds for the family of discounted payoffs, the optional stopping theorem ensures that suboptimal early exercise does not increase the option's value beyond the optimal strategy, aligning it with the counterpart under certain conditions. is essential in these models to prevent deviations in infinite-horizon or perpetual option settings. The theorem connects directly to the Snell envelope, defined as the smallest supermartingale dominating the discounted payoff process, which equals the American option value: V_t = \esssup_{\tau \geq t} \mathbb{E}^\mathbb{Q} [e^{-\int_t^\tau r_u du} g(S_\tau) \mid \mathcal{F}_t]. This envelope characterizes the optimal stopping time as the first of the continuation region boundary, enabling computational methods like in discrete time or PDE solutions in continuous time for pricing. However, the theorem's applicability in is limited without its conditions, such as bounded expectations or . In highly volatile markets, where stopping times can be unbounded—such as in models with or jumps—these conditions may fail, potentially leading to inconsistencies in pricing or apparent opportunities if the martingale property is violated at optional times.

Proofs

Discrete-Time Proof

The optional stopping theorem in discrete time applies to martingales and submartingales under suitable conditions on the stopping time \tau. For a martingale (X_n)_{n \geq 0} adapted to a (\mathcal{F}_n)_{n \geq 0} and a \tau, the stopped process is defined as X_{n \wedge \tau} = X_{\min(n, \tau)} for each n \geq 0. This stopped process inherits the martingale property: \mathbb{E}[X_{n \wedge \tau} \mid \mathcal{F}_m] = X_{m \wedge \tau} for m \leq n, which follows from the optional sampling theorem for bounded stopping times applied iteratively, ensuring the conditional expectations align with the original martingale increments. When \tau is bounded, say \tau \leq N almost surely for some fixed N, the expectation equality holds directly: \mathbb{E}[X_\tau] = \mathbb{E}[X_0]. To see this, note that for n \geq N, X_{n \wedge \tau} = X_\tau . The martingale property of the stopped process implies \mathbb{E}[X_{n \wedge \tau}] = \mathbb{E}[X_0] for all n, so taking n \geq N yields the result. This can also be derived via the telescoping sum representation: X_{n \wedge \tau} = X_0 + \sum_{k=1}^{n \wedge \tau} (X_k - X_{k-1}), where the increments X_k - X_{k-1} are martingale differences with conditional expectation zero, so \mathbb{E}[X_{n \wedge \tau}] = \mathbb{E}[X_0] by linearity. As n \to \infty, the boundedness ensures to X_\tau. For the case of finite expectation \mathbb{E}[\tau] < \infty, assuming bounded increments |X_n - X_{n-1}| \leq K almost surely for some constant K > 0, the equality \mathbb{E}[X_\tau] = \mathbb{E}[X_0] follows by taking limits. Specifically, X_{n \wedge \tau} \to X_\tau almost surely as n \to \infty since \tau < \infty almost surely under the finite expectation condition. Moreover, |X_{n \wedge \tau}| \leq |X_0| + K \tau, which is integrable because \mathbb{E}[|X_0| + K \tau] < \infty. By the dominated convergence theorem, \mathbb{E}[X_{n \wedge \tau}] \to \mathbb{E}[X_\tau], and since \mathbb{E}[X_{n \wedge \tau}] = \mathbb{E}[X_0] for all n, the limit equals \mathbb{E}[X_0]. The monotone convergence theorem applies similarly if the martingale is nonnegative, ensuring the limit passes inside the expectation without domination. In the submartingale case, where \mathbb{E}[X_n \mid \mathcal{F}_m] \geq X_m for m \leq n, the theorem yields an inequality: \mathbb{E}[X_\tau] \geq \mathbb{E}[X_0] under analogous conditions. For bounded \tau \leq k almost surely, the stopped process X_{n \wedge \tau} is a submartingale, so \mathbb{E}[X_{n \wedge \tau}] \geq \mathbb{E}[X_0] for all n, and taking n \geq k gives \mathbb{E}[X_\tau] \geq \mathbb{E}[X_0]. The upper bound \mathbb{E}[X_\tau] \leq \mathbb{E}[X_k] follows by considering the predictable process K_n = \mathbf{1}_{\{\tau < n\}}, which makes (K \cdot X)_n = X_n - X_{n \wedge \tau} a submartingale, leading to \mathbb{E}[X_k - X_\tau] \geq 0 via the submartingale property. For unbounded cases with finite \mathbb{E}[\tau], Doob's maximal inequality strengthens the argument: for a nonnegative submartingale, \mathbb{P}(\sup_{m \leq n} X_m \geq \lambda) \leq \frac{1}{\lambda} \mathbb{E}[X_n^+] for \lambda > 0, which bounds the tail and justifies the limit inequality through or domination.

Continuous-Time Proof

The proof of the optional stopping theorem in continuous time for martingales relies on the right-continuous paths of the process and builds upon foundational ideas from the discrete-time setting by incorporating continuous-time properties. Consider a martingale (X_t)_{t \geq 0} adapted to a (\mathcal{F}_t)_{t \geq 0} satisfying the usual conditions, with cadlag (right-continuous with left limits) paths . For a \tau, the key is to first localize the problem to bounded stopping times and then extend the result using . To establish the result for bounded stopping times, define \tau_n = \tau \wedge n for each n \in \mathbb{N}. Each \tau_n is bounded, and the stopped process (X_{t \wedge \tau_n})_{t \geq 0} is itself a martingale, since the optional sampling theorem holds for deterministic times and extends to bounded stopping times via the right-continuity of paths, ensuring X_{\tau_n} is \mathcal{F}_{\tau_n}-measurable. Consequently, \mathbb{E}[X_{t \wedge \tau_n}] = \mathbb{E}[X_0] for all t \geq 0. As t \to \infty, the (right-continuous with left limits) property of the paths guarantees that X_{t \wedge \tau_n} \to X_{\tau_n} , and by the (applicable under the martingale's integrability), \mathbb{E}[X_{\tau_n}] = \mathbb{E}[X_0]. For unbounded stopping times, approximate \tau by the increasing sequence \tau_n. If the family \{X_{\tau_n} : n \in \mathbb{N}\} is uniformly integrable, then X_{\tau_n} \to X_\tau almost surely (again by càdlàg paths), and Vitali's convergence theorem implies \mathbb{E}[X_{\tau_n}] \to \mathbb{E}[X_\tau] in L^1, yielding \mathbb{E}[X_\tau] = \mathbb{E}[X_0]. Without uniform integrability, the expectation may fail to converge, but under additional conditions like bounded expectation \mathbb{E}[|X_{\tau \wedge t}|] \leq K for some constant K and all t, the result holds by applied to the non-negative and negative parts. This localization and extension via uniform integrability ensures the theorem's validity in the continuous-time framework.

References

  1. [1]
    [PDF] Doob's Optional Stopping Theorem
    The essential content of the theorem is that you can't make money (in expectation) by buying and selling an asset whose price is a martingale. Precisely, the.
  2. [2]
    [PDF] The Martingale Stopping Theorem - Dartmouth Mathematics
    Feb 27, 2013 · We present a proof of the Martingale Stopping Theorem (also known as. Doob's Optional Stopping Theorem). We begin with some preliminaries on.
  3. [3]
    [PDF] Lecture 8: The Optional Stopping Theorem
    This is equivalently to say that no matter how complex is our stopping strategy, if it is reasonable enough, then in expectation ZT have the same value than Z0.
  4. [4]
    [PDF] The Martingale Stopping Theorem, and Applications
    Nov 18, 2019 · The Martingale Stopping Theorem, also sometimes referred to as the Optional Stopping Theorem provides one set of sufficient such con- ditions. ...
  5. [5]
    [PDF] Martingales, Stopping Times, and the Optional Stopping Theorem
    I will finish this paper by providing proofs of the Optional Stopping Theorem which will help explain when it is impossible to beat a fair game. Contents. 1 Two ...
  6. [6]
    Stochastic Processes (Doob) | PDF - Scribd
    Rating 5.0 (4) Stochastic Processes(Doob) (1) - Free download as PDF File (.pdf), Text File (.txt) or read online for free. Stochastic Processes by Doob.
  7. [7]
    [PDF] conditional expectation and martingales
    DISCRETE-TIME MARTINGALES. 1.1. Definition of a Martingale. Let {Fn}n≥0 be an increasing sequence of σ−algebras in a probability space (Ω,F,P).
  8. [8]
    [PDF] Martingales - Purdue Math
    Xn ∈ L1(Ω) for all n ≥ 0. Then. X is a martingale if Xn = E[Xn+1|Fn]. X is a supermartingale if Xn ≥ E[Xn+1|Fn]. X is a submartingale if Xn ≤ E[Xn+1|Fn].
  9. [9]
    [PDF] What is a Martingale?
    May 22, 2013 · Martingale theory illustrates the history of mathematical probability: the basic definitions are inspired by crude notions of gambling, but the ...
  10. [10]
    [PDF] Introduction to Martingales - Stat@Duke
    Nov 19, 2015 · If φ is harmonic, i.e., Gφ ≡ 0, then φ(Xt) is a martingale. 4.2.1 Example: Simple Random Walks. For the symmetric random walk on Z, for example, ...
  11. [11]
    [PDF] martingales in gambling and finance - UChicago Math
    Dec 8, 2023 · Consider a simple game with a fair coin, which has a probability of 1. 2 of landing on its head and an equal probability of 1. 2 of landing on ...
  12. [12]
    stopping time - PlanetMath.org
    Mar 22, 2013 · The first time that an adapted process Xt X t hits a given value or set of values is a stopping time. The inclusion of ∞ ∞ into the range of ...
  13. [13]
    [PDF] An essay on the general theory of stochastic processes - arXiv
    Definition 2.1. A stopping time is a mapping T : Ω → R+ such that {T ≤ t} ∈. Ft for all t ≥ 0. To a given stopping time T, we associate the σ-field FT defined ...
  14. [14]
    CONTROL AND STOPPING OF A DIFFUSION PROCESS ON AN ...
    Under either of these two conditions, one can restrict attention to stopping times in SX that are almost surely finite, without changing the value of the.
  15. [15]
    [PDF] Martingale inequalities - arXiv
    Apr 26, 2024 · We abbreviate Eτ f := E(f|Fτ ). The optional sampling theorem says that, for every discrete time martingale f, bounded stopping time τ, and ...
  16. [16]
    [PDF] 1 IEOR 4701: Martingales I: Discrete time
    Theorem 1.4 (Submartingale Optional Stopping Theorem) If X = {Xn : n ≥. 0} is a SUBMG and τ is a stopping time w.r.t. X such that the stopped process X is UI,.
  17. [17]
    [PDF] Continuous martingales and stochastic calculus
    Mar 11, 2018 · We should now like to establish conditions under which we have an optional stopping theorem for continuous martingales. As usual, our starting ...<|control11|><|separator|>
  18. [18]
    [PDF] An Introduction to Stochastic Processes in Continuous Time
    2.3.5 Optional stopping. We now come to the continuous-time version of the optional stopping theorem. Theorem 2.3.12 (Optional stopping theorem). Let M be a ...
  19. [19]
    Stochastic Processes - Joseph L. Doob - Google Books
    The theory of stochastic processes has developed so much in the last twenty years that the need for a systematic account of the subject has been felt, ...
  20. [20]
    Probability with Martingales - David Williams - Google Books
    Feb 14, 1991 · This book is a modern, lively and rigorous account ... Probability with Martingales Cambridge mathematical textbooks. Author, David Williams.
  21. [21]
    [PDF] Martingales - Books by Rene Schilling
    Jul 20, 2014 · A martingale is the mathematical description of a fair game: The expected net gain or loss from further play, independent of the history, is 0.
  22. [22]
    [PDF] Lecture 18 : Stopping Times and Martingales
    These are optional stopping theorems (maximal inequalities) and convergence theorems. The last equality holds only if Xn and XnHn are integrable. Since we ...
  23. [23]
    [PDF] [CS3958: Lecture 4] Optional Stopping Theorem - Chihao Zhang
    Jan 4, 2023 · We want to compute the ex- pected value of E [𝜏], that is, the average stopping time of the walk. We want to construct a martingale {𝑌𝑡 }𝑡 ≥0 ...
  24. [24]
    [PDF] 18.445 Introduction to Stochastic Processes - MIT OpenCourseWare
    Apr 13, 2015 · X is a submartingale if E[Xn |Fm] ≥ Xm a.s. for all n ≥ m. Today's Goal stopping time. Optional stopping theorem : E[XT ] = E[X0]? ... Suppose ...
  25. [25]
    Martingales and stochastic integrals in the theory of continuous trading
    This paper develops a general stochastic model of a frictionless security market with continuous trading.
  26. [26]
    [PDF] On the pricing of American options - Columbia Math Department
    This process is the Snell envelope of Q, i.e., the smallest supermartingale with ... Consequently, Theorem 6.4 applies to such American options and we also have Q ...
  27. [27]
    [PDF] 1 Discrete time martingales
    Theorem 1. Let (Xn)n be a submartingale and let N be a bounded stopping time, i.e. N ≤ k a.s. for some k ∈ N. Then. EX0 ≤ EXN ≤ EXk. Proof.