1. Introduction
Optional semimartingales represent a class of stochastic processes with a well-established calculus that extends to both standard and nonstandard (or unusual) conditions. In this paper, we focus on optional semimartingales with càdlàg (right-continuous with left limits) and làdlàg (left-continuous with right limits) paths under nonstandard conditions. We explore their potential as a framework for modeling financial markets exhibiting price bubbles, extending the traditional approach by accommodating more general filtrations and capturing a wider range of market phenomena and information structures.
The market value of an asset is primarily driven by supply and demand, while its fundamental value is determined by internal factors such as the company’s products, management quality, and brand strength. Common methods for estimating fundamental value include discounted models based on dividends, cash flows, or residual income. Price bubbles occur when a significant gap develops between an asset’s market price and its fundamental value. However, the two prices are intertwined in various ways. It is conceivable that the disparity between them results from differences in the valuation of intrinsic value, as well as differing short- and long-term outlooks—whether positive or negative—on the firm’s market value.
Price bubbles, as the name suggests, are episodes of sharp price increases followed by sudden collapses. They have been extensively studied by economists and mathematicians, leading to the identification of various conditions under which bubbles can form. For instance, bubbles arise in infinite-horizon growing economies with rational traders (
O’Connell and Zeldes 1988;
Tirole 1985;
Weil 1989). Bubbles can also emerge in economies where rational traders hold divergent beliefs or behave myopically (
Tirole 1982) or in markets with irrational traders (
De-Long et al. 1990). Moreover, bubbles can occur when arbitrageurs are unable to synchronize their trades (
Abreu and Brunnermeier 2003) or in markets with constraints on borrowing (
Santos and Woodford 1997;
Scheinkman and Xiong 2004). Furthermore, price bubbles have been observed on a sector-wide scale, such as the housing bubbles in the U.S. (
Case and Shiller 2003) and earlier in Japan (
Stone and Ziemba 1993). Notably, in all scenarios where price bubbles arise, arbitrageurs cannot profit from them. That is to say that bubbles appear at and survive for unpredictable times and grow to unpredictable sizes such that consistent arbitrage opportunities are not possible to construct.
Mathematically, a stock price exhibits a bubble when the price process is a positive strict local martingale under the equivalent local martingale measure (
Jarrow 2015). A strict local martingale refers to a local martingale that is not a true martingale. The characterization of price bubbles and the pricing of derivatives in finite-horizon economies, under the No Free Lunch with Vanishing Risk (NFLVR) hypothesis, has been extensively explored in several studies (
Cox and Hobson 2005;
Heston et al. 2007;
Loewenstein and Willard 2000a,
2000b). However, bubbles have been shown to violate several classical option pricing theorems—most notably, the put–call parity, which is almost never violated empirically
Kamara and Miller (
1995);
Klemkosky and Resnick (
1980);
Ofek and Richardson (
2003). In (
Jarrow 2015), it was demonstrated that these violations arise due to insufficient structural assumptions about the economy within the NFLVR framework. The study further characterized potential price bubbles in incomplete markets under both the NFLVR and No-Dominance (ND) assumptions (
Merton 1973) and proposed a theory for bubble birth that involves a market exhibiting different local martingale measures across time.
A study by
Kardaras et al. (
2015) examined the impact of asset price bubbles on derivative pricing. It provided decomposition formulas for specific classes of European path-dependent options, demonstrating how stock price bubbles affect option values under the NFLVR condition. Furthermore,
Biagini et al. (
2014) studied the flow in the space of equivalent martingale measures and the corresponding shifts in perception of the fundamental value of a given asset. This framework captured the birth of a bubble as an initial submartingale that transitions into a supermartingale before eventually returning to its initial value of zero. Additionally,
Biagini et al. (
2023) demonstrated that price bubbles are filtration-dependent, meaning they may emerge in one informational setting while vanishing in another. Their findings also highlight how traders with limited information can misinterpret bubbles as arbitrage opportunities, further emphasizing the role of information asymmetry in financial markets.
Carr et al. (
2014) utilized the Föllmer measure to construct a pricing operator for market models where the exchange rate is driven by a strict local martingale. This construction allowed them to preserve put–call parity and provided the minimal joint replication price for a contingent claim. Furthermore,
Jarrow et al. (
2022) introduced invariance theorems to test asset price bubbles in markets where prices evolve as Markov diffusion processes. Their results show that the existence of a bubble can be identified solely through the quadratic variation of the price process, eliminating the need to estimate the drift term under an equivalent local martingale measure. For a comprehensive survey of recent literature on financial bubbles, we refer readers to (
Camerer 1989;
Hirano and Toda 2024;
Protter 2013;
Scheinkman and Xiong 2003).
This paper presents a new approach to the theory of price bubbles using the calculus of optional semimartingales in nonstandard probability spaces. Our goal is to expand the mathematical toolkit for addressing problems in mathematical finance. We propose two market models based on optional semimartingales to explore bubble dynamics. The first model characterizes asset prices as càdlàg semimartingales, illustrating how price bubbles arise from the progressive incorporation of external information—beyond that generated by the price process—into market filtration. The second model employs làdlàg optional semimartingales, necessitating a redefinition of market structures to align with nonstandard probability spaces. This framework effectively captures the emergence, evolution, and eventual dissolution of bubbles. To substantiate our theoretical results, we provide illustrative examples that demonstrate practical applications of the proposed models. Through the application of optional semimartingale calculus, this work delivers new insights into the formation and dynamics of price bubbles. In the next section, we offer a concise introduction to the stochastic calculus of optional processes, establishing the foundation for our analysis.
2. Stochastic Calculus of Optional Processes
The study of stochastic processes without the usual conditions was initiated by Dellacherie and Meyer in 1970 (
Dellacherie 1975). Subsequent advancements were made by numerous mathematicians, with significant contributions by
Galtchouk (
1980,
1985). These works developed a stochastic calculus for processes in nonstandard probability spaces. In this section, we describe some aspects of this theory.
Let for be a complete probability space, where is a complete measure and contains all null sets. However, the family () is not assumed to be complete or right- or left-continuous. We introduce and as optional and predictable algebras on , respectively. is generated by all -adapted processes whose trajectories are right-continuous and have limits from the left. is generated by all -adapted predictable processes whose trajectories are left-continuous and have limits on the right. In addition to the filtration , we introduce , , and . is right-continuous filtration, and the family satisfies the usual conditions under .
A random process () is said to be optional if it is -measurable. An optional process has right and left limits but is not necessarily right- or left-continuous in . A random process (X) is predictable if , and strongly predictable if and . A predictable process has right and left limits but may not necessarily be right- or left-continuous in . For either optional or predictable processes, we can define the following processes: , , and such that and such that .
On an unusual stochastic basis, three canonical types of stopping times exist. Predictable stopping times () are such that is -measurable for all t. Totally inaccessible stopping times () are such that is -measurable for all t; however, we point out that is not necessarily -measurable, since may not be right-continuous. Finally, totally inaccessible stopping times in the broad sense () are such that is -measurable for all t, but since is right-continuous, is also -measurable. A process () belongs to the space if there is a localizing sequence of stopping times in the broad sense (, , , ) such that for all n, where is a space of processes and is an extension of by localization.
A process (
) is increasing if it is non-negative; its trajectories do not decrease; and, for any
t, the random variable (
) is
-measurable. Let
(
for short) be a collection of increasing processes. An increasing process (
A) is integrable if
and locally integrable if there is a sequence (
,
,
) such that
for all
. The collection of such processes is denoted by
(
). (A process
,
) is a finite variation process if it has finite variation on every segment (
,
), that is,
, for all
, where
(
for short) denotes a set of
-adapted finite variation processes. A process (
) of finite variation belongs to the
space of integrable finite variation processes if
. A process (
) belongs to
if there is a sequence (
,
,
) such that
for any
, i.e., for any
n,
. A finite variation or an increasing process (
A) can be decomposed as
, where
is continuous,
is right-continuous,
is discrete right-continuous, and
is discrete left-continuous such that
where the series converges absolutely (
Abdelghani and Melnikov (
2019)).
Definition 1. A process () is an optional martingale (supermartingale or submartingale) if and there exists an integrable random variable () such that, for any stopping time (), Let
(
for short) denote a set of optional martingales and
denote a set of optional local martingales. If
M is a (local) optional martingale, then it can be decomposed as
where
is a continuous,
is a right-continuous, and
is a left-continuous (local) optional martingale.
and
are orthogonal to each other and to any continuous local martingale. Moreover,
and
can be written as
Definition 1 tells us that all optional martingales are right-closed. Also, for a càdlàg martingale (Y) adapted to , the optional projection () is a làdlàg optional martingale under unusual conditions.
An optional semimartingale (
) can be decomposed to an optional local martingale and an optional finite variation process:
where
and
. A semimartingale (
X) is called special if the above decomposition exists with a strongly predictable process (
). Let
denote a set of optional semimartingales and
be a set of special optional semimartingales. If
, then the semimartingale decomposition is unique. Through optional martingale decomposition and decomposition of predictable processes (see
Galtchouk (
1977,
1985)), we can decompose a semimartingale further to
with
,
, and
, where
and
are finite variation processes that are right- and left-continuous, respectively.
are right-continuous local martingales,
are discrete right-continuous local martingales, and
are left-continuous local martingales. This decomposition is useful for defining integration with respect to optional semimartingales.
A stochastic integral with respect to an optional semimartingale was defined in (
Galtchouk 1985) as
The stochastic integral with respect to a finite variation process or a strongly predictable process (
over
and
over
) is interpreted as usual in the Lebesgue sense. The integral expressed as
over
is our usual stochastic integral with respect to a càdlàg local martingale, whereas
over
is a Galtchouk stochastic integral with respect to a left-continuous local martingale. In general, the stochastic integral with respect to an optional semimartingale (
X) can be defined as a bilinear form (
) such that
where
Y is, again, an optional semimartingale,
, and
. Therefore, the stochastic integral over optional semimartingales is defined on a much larger space of integrands, where
is the product space of predictable and optional processes. To understand this stochastic integral, one must keep in mind the definition of integration intervals and the relationship between integrands and integrators. Consider the following integral:
which is
-adapted and well defined according to
Galtchouk (
1985); the integration is taken over the interval of
and the integrand (
) is adapted to
, while
is adapted to
for all
.
The properties of optional stochastic integrals are described as follows: First, isometry is satisfied with
The quadratic variation
is defined as
Linearity is also satisfied with for any and in the space expressed as ; is , with its martingale part satisfying a.s. for any stopping time (T) in the broad sense, is , with its martingale part satisfying a.s. for any stopping time (T). Moreover, orthogonality is such that are orthogonal in the sense that their product is a local optional martingale. Also, differentials are independent: and . Lastly, for any semimartingale (Z), the quadratic projection is .
The stochastic calculus of optional processes and its applications have witnessed significant advancements in recent years. In
Abdelghani and Melnikov (
2020), a solution to the nonhomogeneous linear stochastic equation of optional semimartingales was provided. The work reported in (
Gasparyan 1985) established the existence and uniqueness of solutions under Lipschitz conditions for nonlinear stochastic equations driven by optional semimartingales, while
Abdelghani and Melnikov (
2020) proved the same under monotonicity conditions. A comparison of solutions for stochastic equations of optional semimartingales was presented in
Abdelghani and Melnikov (
2020), and
Jarni and Ouknine (
2021) proved the existence and uniqueness of solutions of reflected stochastic equations driven by optional semimartingales. The study reported in
Falkowski (
2025) further addressed the existence, uniqueness, and approximation of solutions of stochastic differential equations with two time-dependent reflecting barriers driven by optional semimartingales.
In the context of filtering theory,
Gasparyan (
1988) derived nonlinear filtering equations for optional semimartingales under a restricted filtration that is neither right-continuous nor complete, while
Abdelghani and Melnikov (
2019) derived a nonlinear filtering equation for optional supermartingales under various conditions, utilizing a version of the optional decomposition of local optional supermartingales. In the field of statistics,
Abdelghani et al. (
2021) investigated regression problems where the observed process is an optional semimartingale.
In mathematical finance,
Kühn and Stroh (
2009) characterized a tractable domain of general integrands for optional semimartingales that possesses desirable properties for modeling dynamic trading gains when price processes follow optional semimartingales. The work reported in (
Abdelghani and Melnikov 2017) employed optional semimartingales with random times to model derivative contracts with default. Additionally, the optional decomposition of local optional supermartingales derived in (
Abdelghani and Melnikov 2019) aids in describing the capital evolution for corresponding minimal hedging portfolios. A criterion for the existence of local optional martingale deflators in financial markets following optional semimartingales was provided in (
Abdelghani and Melnikov 2020). Optional semimartingale markets have found application in energy markets, where prices exhibit spikes (
Abdelghani et al. 2022). For a comprehensive review of optional semimartingale calculus and its applications, we recommend referring to (
Abdelghani and Melnikov 2020).
3. Càdlàg Semimartingale Market in Unusual Probability
Space
Here, we develop a model that describes a financial market with an asset price following a càdlàg semimartingale in an unusual probability space. In this model, we demonstrate that price bubbles emerge as a result of incorporating additional external information, beyond the information generated by the price process, into the market’s filtration. In real-world economies, fluctuations in optional local martingale measures naturally occur, reflecting regime changes influenced by diverse beliefs, risk aversion attitudes, institutional structures, technological advancements, tax policy changes, and fluctuations in endowments.
Let
be a
complete probability space. On this space, we define the following filtrations:
,
, and
.
may not be right-continuous nor complete.
is right-continuous.
satisfies the usual hypotheses. Let the market price of the risky asset be given by
, where
T is a stopping time in
that represents the termination time of the risky asset. Let
be the cumulative dividend process of the risky asset and
be the risk-free rate of return of a money market account. The money market account serves as a numeraire to make the spot interest rate zero. We assume that
S,
D, and
r are càdlàg semimartingales that are
-measurable and remain càdlàg semimartingales under
. Furthermore, we assume that
,
, and
. Let
be the terminal payoff or the liquidation value of the asset at time
. Therefore, the wealth process (
X) associated with the market price of the risky asset is
The market value of wealth is the current position in stock plus the accumulated dividends and the terminal payoff. Since the asset does not exist after T, we stop the wealth at T.
Remark 1. Since S and D are càdlàg semimartingales with respect to and , the wealth (X) is a càdlàg semimartingale with respect to and . In other words, S, D, and X remain càdlàg semimartingales, even after expansion of to .
Remark 2. The assumption that S, D, and r are càdlàg semimartingales and that the probability space is complete is needed for us to be able to use the results of the theory of no free lunch with vanishing risk.
Next, we establish the following criteria for allowable trading strategies: Firstly, all trading strategies must be self-financing to prevent doubling strategies. Secondly, trading strategies must be admissible, imposing restrictions on traders’ ability to borrow without bounds. Thirdly, our market must satisfy NFLVR to safeguards against the emergence of arbitrage opportunities. Furthermore, we impose the no-dominance assumption, which ensures that no cash flow generated by a trading strategy can result in one portfolio dominating another over any trading time intervals.
3.1. Trading Strategies
A trading strategy is a pair (
) of processes adapted to
representing the number of units in the risky asset (
S) and money market account (
r) held at time
t. The corresponding portfolio value process (
) of the trading strategy
is
Let us suppose
is a semimartingale; then, with
, the value process (
) is given by
where
Discarding the temporary assumption that
is a semimartingale, we can define a self-financing trading strategy
to be a pair of processes, with
being predictable and
being optional such that
where
. As noted, a self-financing trading strategy starts with zero dollars (
), and all proceeds from purchases/sales of the risky asset are financed/invested in the money market account because Equation (
4) shows that
is uniquely determined by
if a trading strategy is self-financing.
To avoid doubling strategies by indefinite borrowing of S, we need to restrict the class of self-financing trading strategies further. The notion of admissibility corresponds to a lower bound on the wealth process—an inability to borrow without bound.
Definition 2. Admissibility. Let be a wealth process given by Equation (3). We say that the trading strategy (π) is a-admissible if it is self-financing and for all , almost surely. We say a trading strategy is admissible if it is self-financing and there exists an such that for all t, almost surely. We denote the collection of admissible strategies as . Now, we can introduce the notion of an arbitrage-free market. Let
and
where
is the set of a.s. bounded random variables and
is the set of non-negative, finite-valued random variables.
Definition 3. We say that a market satisfies NFLVR ifwhere denotes the closure of in the sup-norm topology on . The NFLVR excludes all self-financing trading strategies with zero initial investment that guarantee non-negative cash flows or offer strictly positive cash flows with positive probability. The NFLVR condition also excludes sequences of trading strategies that approach arbitrage opportunities. Therefore, we assume the following:
Assumption 1. The market satisfies NFLVR.
A key concept to characterize a market that satisfies the NFLVR is the equivalent local martingale measure:
Definition 4. Equivalent Local Martingale Measure. Let be a probability measure equivalent to such that the wealth process (X) is a -local martingale. We call an Equivalent Local Martingale Measure (ELMM), and we denote the set of ELMMs as .
The first fundamental theorem of asset pricing (
Delbaen and Schachermayer 1994,
1995,
1998,
1999) posits that a market is free of arbitrage in the NFLVR sense if and only if there is an equivalent probability measure (
) under which the asset price process (
S) is a
martingale. Any
martingale that is bounded from below is also a local martingale. Since
S is non-negative, it is also a local martingale with respect to
. This leads us to the following theorem:
Theorem 1. First Fundamental Theorem. A market satisfies NFLVR if and only if there exists an ELMM.
Next, we define what we mean by no dominance. For each admissible trading strategy (), its wealth process () is given by , where is a local martingale under each .
Definition 5. Set of Super-Replicated Cash Flows. Consider a fixed future time denoted by v and let be the collection of all payoffs () associated with an asset derived from an admissible trading strategy. Here, is the asset’s cumulative dividend process. It is a non-negative, non-decreasing càdlàg semimartingale adapted to . is a non-negative random variable representing the asset’s terminal payoff at time v. Now, let us define the set of super-replicating trading strategies (Φ): Set represents the cash flows of assets that can be super-replicated through trading in the risky asset and the money market account. As we illustrate later, this particular set of cash flows is relevant for our assumption of no dominance. Our initial demonstration involves establishing that this subset of asset cash flows comprises a convex cone.
Lemma 1. Φ is closed under addition and multiplication by positive scalars, i.e., it is a convex cone.
If
, then for each
,
The first inequality follows because is a wealth process of admissible trading strategies. The second inequality follows because is a non-negative (because both and are non-negative) -local martingale bounded below and, hence, a supermartingale such that . Therefore, each asset () is integrable under any ELMM. This is the reason for restricting our attention to the set of cash flows (). This set () is large enough to contain many of the assets of interest in derivatives pricing.
We do not expect to see in a well-functioning market or any dominated assets or portfolios. To formalize this idea, let us denote the market price of
at time
t as
and fix
. Consider a pair of stopping times (
) and define the net gain (
) by purchasing
and selling at
as follows:
Definition 6. Dominance. Let and be two assets. If a pair of stopping times () exists such thatand a.s., then we say that asset 2 dominates asset 1 at time σ. Therefore, we impose the following condition:
Assumption 2. (No Dominance): Let the market price be represented by a function () such that there are no dominated assets in the market.
The addition of a no-dominance assumption prevents the occurrence of price bubbles in a complete market. Therefore, studying bubbles necessitates looking at an incomplete market, which, according to the second fundamental theorem of asset pricing, is characterized by numerous local martingale measures in the absence of arbitrage. The choice of one of these measures defines the fundamental price and can lead to a bubble. Notably, the theory of bubbles in incomplete markets suggests that bubbles can burst but cannot re-emerge, contradicting economic observations. Accommodating the possibility of bubble rebirth requires a significant adaptation of the martingale pricing framework to allow the market to exhibit different local martingale measures over time. These shifts correspond to changes in underlying economic fundamentals and can create new bubbles. Implementing this change, which contradicts the NFLVR theory’s assumption of a consistent local martingale measure, demands a noteworthy extension of this theory, as carried out in
Jarrow et al. (
2010) under usual conditions with the method of filtration expansion. In the subsequent section, we outline a comparable extension approach, leveraging the calculus of optional processes within nonstandard probability spaces, where filtration naturally progresses from
to
.
3.2. Regime-Change Processes
We define regime-shift processes as follows. Let
denote an increasing sequence of random times with
such that
are
-measurable for any
i and
t. In other words,
values are stopping times in the wide sense whenever
; otherwise, they are regular stopping times. Let
be a sequence of
-measurable random variables that characterize the state of the economy at those times. Furthermore, we define the two stochastic processes (
and
) as
counts the number of regime shifts up to and including time t, whereas identifies the characteristics of the regime at time t. Obviously, N and Y are -measurable. Let be a natural filtration generated by N and Y and the filtration be defined as . According to the definition of , is an increasing sequence of -stopping times. We just take so that is extended to only by N and Y. To avoid confusion that may arise from “” in the symbol, we set .
Remark 3. We need not suppose that Y and N are independent of the filtration () to which the price process (S) and wealth process (X) are adapted. Allowing for dependence would enable the birth of bubbles to be influenced by intrinsic uncertainty (see Froot and Obstfeld (1991)). However, if Y and N are independent of , an extrinsic source of randomness is introduced into our economic framework. 3.3. Fundamental Price
In classical mathematical finance, the fundamental price is equivalent to the market price. But in finite-horizon markets, the difference between the market price and fundamental price is nil. This is because the market price is a martingale for any EMM’s and is equal to the arbitrage-free price, which is equal to the conditional expectation of the asset’s payoffs under . The conditional expectation of the stock’s payoffs is the present value of the asset’s cash flow, which is its fundamental value. In these cases, there are no price bubbles.
In contrast, the local martingale approach to mathematical finance allows for the existence of price bubbles (see, for example,
Cox and Hobson (
2005);
Delbaen and Schachermayer (
1994,
1995,
1998,
1999);
Loewenstein and Willard (
2000a,
2000b)). In a market adhering to the no free lunch with vanishing risk (NFLVR) condition, the arbitrage-free price and the market price for a primary asset are indistinguishable, yet they may differ from the conditional expectation of the asset’s payoffs, which represents the fundamental price. Indeed, for any
, if the asset’s price is a strict local martingale, then a bubble exists.
We take the fundamental price as the asset’s discounted expected cash flows, given a local martingale measure (
), as is in (
Jarrow et al. 2010). The local martingale measure (
) selected from
for valuation is a measure consistent with market prices of traded derivatives.
Schweizer and Wissel (
2008) and
Jacod and Protter (
2010) showed that if enough derivative securities of a certain type are traded, then
can be determined. Similar to
Jarrow et al. (
2010), we assume that the selection of the measure can depend on the current economic regime, and as the regime shifts, so does the local martingale measure selected by the market. This selection process determines the fundamental value and the birth of price bubbles.
First, let us consider the ELMMs of the wealth process (X) in . With respect to this restricted set, the Radon–Nikodym derivative is . Its density process is defined as . Z is , an -adapted process.
Let the local martingale measure in our extended economy depend on the state of the economy at time t, as represented by the filtration (), the number of regime shifts be , and the state variable be . Suppose at t, ; let be the ELMM selected by the market at time t, given . Given , the fundamental price of an asset or a portfolio is the asset’s expected discounted cash flows at time t:
Definition 7. Fundamental Price. Let be an asset with a maturity of v and payoff of . The fundamental price () of asset ϕ is defined by, where . In particular, the fundamental price of the risky asset (
) is the discounted future cash flow on
under
for all
i, given by
To understand this definition, let us focus on the risky asset’s fundamental price. At any time (
), given that we are in the
ith regime (
), the right side of expression (
13) simplifies to
Note that the payoff of the asset at infinity () does not contribute to the fundamental price, since agents cannot consume it. Furthermore, when the time is , the fundamental price is . We emphasize that under NFLVR and no-dominance, the market price () equals the arbitrage-free price but needs not equal the fundamental price ().
Next, we present an alternative formulation of the fundamental price (Equation (
12)) using a conditional expectation with respect to an equivalent probability measure (
) instead of the sum of conditional expectation with respect to measures
.
Theorem 2. There exists an equivalent probability measure () such that Proof. Let
be a Radon–Nikodym derivative of
with respect to
,
, and
. We define
Then,
, almost surely, and
at any moment of time (
t). Therefore, we can define an equivalent measure (
) on
as
. The Radon–Nikodym density (
) on
is
Then,
and we observe that
and continue to
□
is known as the valuation measure at t, and collection is known as the valuation system. If for all t, then the valuation system is static with , , and there are no regime shifts. If the market is not static, then it is dynamic, with the valuation system expressed as , where, in the ith regime , the valuation measure coincides with .
Given the definition of the asset’s fundamental price (
13), the fundamental wealth process is expressed as follows:
Thus,
, where
Therefore, we can rewrite
as
Then, given the definition of the asset’s fundamental price (
13), an asset price bubble (
B) is defined as the difference between the market price of the asset (
S) and its fundamental value (
):
Remark 4. So far, we have presented a model of an economy that exhibits regime changes. In this market model, the random times responsible for creating regime shifts in the economy are stopping times that are in the filtration (). By doing so, we have avoided the requirement that every local martingale should also be a local martingale, where represents the extended filtration that incorporates random regime-shift times. However, we have made an important assumption that the market price process remains a càdlàg optional semimartingale under both and . This assumption allows us to utilize the theory of NFLVR. Other results concerning the decomposition of bubbles and derivatives with bubbles, among others, can be easily extended to optional extended financial markets with some modifications.
4. Làdlàg Optional Semimartingale Market with Bubbles
In this section, we introduce alternative financial markets where assets are characterized by làdlàg optional semimartingales in unusual probability spaces. We provide a version of NFLVR specifically tailored for these markets. Additionally, we develop a theory of local martingale deflators for làdlàg optional semimartingales and study the formation of price bubbles in these markets. Finally, we present illustrative examples.
4.1. Absence of Arbitrage
Research on no-arbitrage arguments has culminated in the Fundamental Theorem of Asset Pricing, which states that under the usual conditions, for a real-valued semimartingale (
X), there exists a probability measure (
) equivalent to
under which
X is a
martingale if and only if
X does not permit a free lunch with vanishing risk (NFLVR). Given
{
admissible and
exists a.s.} and
,
X is said to satisfy NFLVR if
, where
is the closure of
in the norm topology of
Delbaen and Schachermayer (
1994).
While, the comprehensive development of NFLVR or NA1 (no-arbitrage of the first kind), as well as the equivalence relations—ELMM (equivalent local martingale measure) and ELMD (equivalent local martingale deflator)—for optional semimartingales in unusual probability spaces, are important and achievable, such topics exceed the purview of this paper. Nevertheless, we present a compelling case that demonstrates the plausibility of financial markets in unusual probability spaces and establish that they are free of arbitrage under certain conditions.
Consider a market on the
unusual probability space expressed as
. Let
be a real-valued optional semimartingale and
be a
càdlàg semimartingale. Suppose
satisfies the NFLVR condition with the admissible portfolio expressed as
and
a.s.; then there, is ELMM
such that
is a local martingale in
. Given that
is a local martingale under
, we aim to recover
Y and identify the portfolio (
) as a process in
using the following optional projection:
where “·” is the stochastic integral with respect to the càdlàg semimartingale (
) and predictable integrand and “∘” is the stochastic integral with respect to the làdlàg optional semimartingales (
Y) with the optional integrand, “
”.
Theorem 3. Let be a -local martingale; then, a.s. .
Proof. Consider the sequence (
) of stopping times in
, where
is a martingale for all
k and
. Since
for all
s, then
Observe that
and if
is evolving in the interval of
, then
is evolving in the interval of
. Consequently,
□
After establishing that is the optional projection of on , we proceed to demonstrate that is a local optional martingale with respect to .
Lemma 2. is a local optional martingale under .
Proof. Utilizing the aforementioned theorem, for any
, we have the following:
□
By applying the same approach that led us to establish NFLVR under unusual conditions, we observe that if NA1 holds for on , then it also holds for Y on .
Consequently, we can confidently assert that if NFLVR or NA1 is satisfied for càdlàg semimartingales on , it must also hold for their optional version through optional projection on . Thus, we can conclude that optional markets are free of arbitrage opportunities under the appropriate conditions on and .
4.2. Market and Portfolios
Consider the market price of a risky asset (
) a non-negative làdlàg optional semimartingale adapted to
. Once more, let
be a càdlàg semimartingale process adapted to
representing the cumulative dividend of the asset. Let
be a terminal payoff of the asset at some future time (
). We assume that both
,
,
and
,
. As usual, the money market account serves as a numeraire that we assume to be a strictly positive process, and we suppose that its value has already been incorporated in the different components of our market (
S,
D, and
L). The wealth process associated with
S,
D, and
L is
where
. Consider the integral form of
X,
The integral is
, since
and
, and zero otherwise. Moreover,
Remark 5. S can be decomposed to , where is càdlàg semimartingale and is a left-continuous optional semimartingale. Theorem (1.14) in Galtchouk (1980) tells us that there are three sequences of stopping times that absorb all the jumps of S: a predictable sequence of stopping times () measurable in , an inaccessible sequence of stopping times () that are -measurable, and wide-sense stopping times () that are -measurable. As we develop this theory, we see that stopping times in the wide sense (), which absorb the left-optional jumps of S, are the ones that can lead to regime shifts in the economy. Let the portfolio expressed as
consists of the optional processes (
and
). The value process of the portfolio is given by
. We restrict the portfolio (
) to be
self-financing in
X, meaning that there is no inflow or outflow of wealth beyond the initially invested amount, and any dividends gained from
S are reinvested in it. In other words, the change in the value process (
) is only due to change in
X –
or
with
. Hence, it follows that
where
Since is uniquely determined by , the portfolio is, indeed, self-financing. The process (X) is an optional semimartingale, and is X-integrable, i.e., . Thus, evolves in the space expressed as , which is not our usual predictable portfolio space but contains predictable and optional components. On the other hand, belongs to .
The optional integral (
) can be written as
where the trading strategy (
) is of two components: a predictable component (
) and an optional component (
). The integral (
) is the usual stochastic integral over the càdlàg semimartingale (
). Similarly, the stochastic integral,
is well-defined (see
Galtchouk (
1985)) and is
-measurable, as the integration is taken over the interval of
. Moreover, the integrand (
) is
-measurable, while
is
-measurable for any
.
Now, for convenience, let
X be given by
where
is the stochastic exponential,
is an optional semimartingale adapted to
and
and is
H-integrable. Additionally, suppose that the dividend process is given by
, where
, and is
S-integrable. Then, we can write
X in terms of
S as
Since
, it follows that
Next, we proceed to construct a local optional martingale deflator for X.
4.3. Transforming Optional Semimartingales to Local Optional Martingales
Our objective is to find a transformation ( a.s. ) belonging to that will render . This Z is known as a local martingale deflator. For Z, we can define a local optional martingale () as the optional stochastic integral () which allows us to write Z as the stochastic exponential (). Next, we present a way to find N.
Given
and
,
is a local optional martingale if
belongs to
. Here,
is derived from
If we also have and , then .
Since,
H is an optional semimartingale, it can be decomposed as
, where
A is a finite-variation process and
M is a local martingale. Let
and write
as
is a local optional martingale under
if
where
represents the compensators of
. By finding all
, i.e., the process (
g) such that Equation (
28) holds and
, we obtain the set (
) of all appropriate local optional martingale deflators (
Z) such that
is a local optional martingale.
Lemma 3. A solution to is given byif it exists, where and g is M-integrable. Proof. The result follows directly by solving the equation for g, assuming the existence of the Radon–Nikodym derivative (). □
Lemma 4. If Z is a local martingale deflator of X and π is a self-financing portfolio, then is a local optional martingale. Conversely, if is a local optional martingale for some , then Z is a local martingale deflator of X.
Proof. Suppose that
is a local martingale deflator of
X and
is a self-financing,
X-integrable portfolio. Then,
can be written as
from which it follows that
is a local optional martingale. Conversely, if
is a local optional martingale for some
and
is a self-financing,
X-integrable portfolio, then
is the difference of two local optional martingales and, therefore, is, itself, a local optional martingale for any optional process (
). □
Lemma 5. If is a local optional martingale, then is a local optional martingale.
Proof. Since
if
and 0 otherwise and
and finite for that, the dividend cannot explode in finite time; then, if
X is a local optional martingale, so is
S. Let us now consider a representation of
in terms of
:
Since and Z are local optional martingales, is also a local optional martingale. □
Remark 6. If and , then we can define as a new measure equivalent to , i.e., , and . The set of all optional local martingale measures () corresponding to Z in is denoted as as .
4.4. Bubbles and Fundamental Values
For a làdlàg optional semimartingale market, the definition of fundamental price does not require the valuation system of local martingale measures used in extended economies to explicitly define regime shifts. In optional semimartingale markets, regime shifts are implicitly encoded in the set of local optional martingale deflators () of X, in which a deflator (Z) incorporates market-regime shifts in its left-continuous part ().
At time
T or thereafter, the total accumulated wealth (
) of an investor in
S is
If
, then the present value of
scaled by the local martingale deflator (
Z) is
. The difference between
X and
(
) is an optional semimartingale. Deflating the difference by the local martingale deflator (
Z), we obtain
where
is an optional local martingale and
is an optional martingale; hence,
is a local optional martingale. Since
is a positive local optional martingale, it is an optional supermartingale and
. Therefore,
.
Next, we decompose
to attained wealth plus the remaining wealth up to time
T. To do this, consider
, where
and
is what we may call the fundamental price, as it is the present value of all unrealized future dividends and the final asset value. Therefore, the fundamental wealth is
where
is the realized wealth.
Granted a local optional martingale deflator (), the above discussion leads us to the following definitions of fundamental price and wealth under the corresponding equivalent local optional martingale measure ():
Definition 8. Optional Fundamental Price. For an asset with a maturity of T and payoff of , the fundamental price () is defined by where . In particular, the fundamental price of the risky asset (
) is the discounted future cash flow on
under
, as given by
Its fundamental wealth (
) is
which is a uniformly integrable optional martingale.
Knowing the fundamental value, we can define the price and wealth bubbles as follows:
Definition 9. An asset price bubble (B) is , and the associated wealth bubble is .
Moreover, knowing that and are positive local martingales, we have the following lemma:
Lemma 6. Under , bubbles Y and B are non-negative local optional martingales.
Bubbles in optional semimartingale markets share similarities with those in càdlàg markets but with a key difference: in optional semimartingale markets, regime shifts naturally arise due to the presence of wide-sense random times. These shifts introduce additional complexity to the market dynamics. Before delving into the properties of bubbles as optional local martingales, let us first revisit some foundational concepts.
A càdlàg stochastic process (M) is a martingale under the usual conditions if for any t with , and for any t and s with , the a.s., where and satisfies the usual hypothesis. Furthermore, M is a uniformly integrable martingale if converges to zero as uniformly in t – , where the supremum is over for a finite time interval and if the process is considered on . Also, any martingale (M) on a finite time interval () is uniformly integrable and is closed by . The martingale convergence theorem tells us that if M is a martingale on and , then there exists an almost sure limit (, where Y is an integrable random variable). The optional sampling theorem allows us to express the martingale property in terms of stopping time. Let represent bounded stopping times belonging to . Then, for any càdlàg martingale, random variables and are integrable and . M is a càdlàg local martingale if there exists a sequence of stopping times () such that , and for each n, the stopped processes () is a uniformly integrable martingale in t.
In contrast, a process (
) is defined as a strong optionalmartingale, as per
Dellacherie (
1975), if
, the random variable (
) is integrable for any bounded stopping time (
) and for every pair of bounded stopping times (
S and
T with
), and we have
. Also, every càdlàg martingale is a strong optional martingale, but not all strong optional martingales are càdlàg. However, optional martingales were also defined in
Galtchouk (
1980) as follows: a process (
M) is an optional martingale if it is optional and there exists an integrable random variable (
) such that
a.s. for any stopping time (
with
). With either definition of optional martingale, the definition of local optional martingale coincides. Moreover, every càdlàg local martingale in
is an optional local martingale in
.
In the following discussion, we adopt the definition of optional martingales from
Galtchouk (
1980), where a true optional martingale is characterized by the existence of an integrable random variable (
) such that
a.s. for any stopping time (
with
). However, our focus remains on local optional martingales, and we explore the conditions under which they qualify as strict local martingales or true martingales.
4.4.1. Fundamental Value Invariance
Let
denote a set of local optional martingale deflators such that for any
, the
family is uniformly integrable for some localizing sequence (
). In other words,
or equivalently,
(see
Abdelghani and Melnikov (
2024)). Additionally, let
be a set of local optional martingale deflators that are not uniformly integrable. It is evident that the choice of an equivalent local martingale deflator impacts the fundamental value. However, for the
class, fundamental values are invariant. We characterize this invariant class with the following lemmas:
Lemma 7. Let . Suppose that at T, the fundamental wealth () is equal to its market value (); then, the fundamental wealth process () and the fundamental price () do not depend on the choice of .
Proof. Suppose
Z and
. Let
be the fundamental wealth given
Z, and let
be the fundamental wealth given
. Since
and
are uniformly integrable local martingales under
Z and
, respectively, for any
,
The difference between and does not depend on the choice of Z. Therefore, since , . □
The following lemma describes the relationship between the fundamental price under a deflator (); a non-uniform integrable local optional martingale; and , a uniformly integrable martingale.
Lemma 8. Let , , and . Then, even if the fundamental wealth () is equal to its market value (), we have Proof. For any
, we have that
since
is a submartingale—
. □
Remark 7. In (Abdelghani and Melnikov 2024), it was demonstrated that a local optional martingale deflator (, , and ) can be a true optional martingale if one can find a localizing sequence () such that for any n and the family is uniformly integrable on . This leads to the conclusion that . To verify the uniform integrability of , it is sufficient to check that . 4.4.2. Decomposition Theorems
According to the Riesz decomposition theorem (see Theorem 2.1 in
Galtchouk (
1982)), any optional supermartingale can be decomposed into two components: an optional martingale, which is uniformly integrable by definition, and a potential, which is a non-negative optional supermartingale whose expectation converges to zero. Applying this decomposition to a price bubble (
B), which is an optional supermartingale, we can express it as the sum of a uniformly integrable bubble
with
a.s.) and a potential bubble (
–
). Therefore, the price (
S) admits the following unique decomposition:
.
Theorem 4. Under , if the price bubble (B) is a local optional martingale, which is also an optional supermartingale, then the price can be decomposed as follows:where is a uniformly integrable martingale with a.s. and is a potential. Proof. This result follows from the Riesz decomposition theorem for optional supermartingales, which allows for the decomposition of B into a uniformly integrable martingale () and a potential (). □
On the other hand, if the deflated price bubble (
) is a strict local optional supermartingale, it can be decomposed into a uniformly integrable optional martingale, a local optional martingale, and a local optional potential. We demonstrate this in the following section. It is important to note that
Kazamaki (
1970) showed that a local supermartingale can be decomposed into a local martingale and a potential under standard conditions. The proof we present for the decomposition of a local optional martingale extends Kazamaki’s result to local optional supermartingales under nonstandard conditions.
Let
be a localizing sequence such that
for all
n. The localized
is an optional supermartingale. Using the decomposition of Riesz, we obtain
where
is a uniformly integrable optional martingale and
is an optional potential for every
n. Let us further assume that
but this assumption is true for price bubbles, as they are always
0 by construction. Since
implying that
is decreasing in
m for any
k and
t. Thus, we are compelled to set
where
is an
set of
-measure zero, which may depend on
t and
n, and
is its compliment. From the monotone convergence theorem, we have that for each pair (
) and each
n,
It follows from condition (
32) that
which implies that
is integrable. Hence,
is integrable as well. Therefore, for each
n,
is an optional martingale. Next, we consider the relationship between
and
. We know that
so
and
coincide and are
-measurable for any
p. As a result, we may put forth the definition of
and observe that
is an optional martingale for every
n. Therefore,
is a local optional martingale, and
for which
for all
n. Therefore,
is a local potential. Furthermore, for each
, there exist
such that
. However, note that the
family may not be uniformly integrable. Let
; then, we may write
as
where
is a uniformly integrable optional martingale and the
family comprises local optional martingales. Therefore,
and we have the following lemma:
Theorem 5. If condition (32) is satisfied, then the deflated price bubble has the following decomposition:where is an optional martingale, is a local optional martingale, and is a local optional potential. Thus, the deflated price process is Proof. This follows from the application of the Riesz decomposition of local optional supermartingales, as discussed above. □
5. Illustrative Examples
In an optional semimartingale market, bubble birth is possible, as we show with the following examples.
Example 1. Regime-shifted geometric Brownian motion.
Consider an asset price described by the following optional semimartingale:
where
W is Brownian motion,
is a random regime-shift time, and
is a continuously increasing process that is
-measurable. The asset (
S) pays no dividends and expires at
, with a terminal payoff of
.
We can write . This shows that S has two phases; one phase is before , and the other is after . Before , S is a martingale free of bubbles, whereas after , it is not. Let , where , the compensator of , is continuous and -measurable. is a left-continuous, -measurable optional martingale.
We can choose a local martingale deflator (
Z) in the form of
such that
is a local martingale, as follows:
since
and
.
For
to be a local martingale, we must have the following:
Ideally, the integral equation (
33) admits at least one solution for
and
in terms of
v,
, and
. However, it is possible that Equation (
33) has infinitely many solutions for
, resulting in many possible deflators (
Z) with various properties. In addition to satisfying (
33), the deflator must also satisfy
, imposing further constraints on the permissible values of
.
Let us consider several choices of solutions for Equation (
33). Setting
, we find that
, and the deflated price becomes
, which depends on the regime-shift time (
). On the other hand, if
, the resulting deflator is
, which subsumes the regime-shift time (
). In this case, the deflated price simplifies to
, becoming independent of
. Thus, depending on the choice of
and the properties of the
v,
, and
, the characteristics of
Z and
are determined.
Since the asset pays no dividends, the fundamental price under
is given by
. A price bubble, defined as
, is an optional semimartingale. However, after deflating by
Z, the bubble becomes a local optional martingale:
Depending on the choice of and the properties of the v, , and , may be either a true martingale or a strict local optional martingale.
Example 2. Jump diffusion exhibiting a regime shift.
Here, we present a concrete example of a regime-shifted diffusion process. Suppose the market price of an asset is given by
where the constants are
,
, and
;
W is Brownian motion; and
N is a regime-shift Poisson process adapted to
with a constant intensity of
. The price process (
S) reflects the price of the asset in all regimes. Suppose the asset pays no dividends so that the wealth process (
X) is equal to
S. To find a local martingale deflator (
Z) of
S, we write
Let
, where
. Then,
and
To transform
S to a local optional martingale using
Z, we must choose
a and
b such that
There are many solutions to these equations; therefore, there are many possible local martingale deflators (Z). Let us consider few possible solutions to these equations.
Suppose
; then,
,
and
In this case, is a uniformly integrable martingale under the Novikov condition.
On the other hand, if
, then
and
. Therefore,
is not a uniformly integrable martingale, depending on
N, which jumps at wide-sense random times, causing a regime shift in the market.
Example 3. Regime change and bubble rebirth.
Suppose the market’s valuation measure shifts at time
from
to
. Then, according to Lemma (8),
The fundamental price of the asset is then given by
and the bubble is
Thus, a bubble is born at time . As shown in Lemma (7), a switch between uniformly integrable deflators does not alter the value of X, ensuring that no bubble exists. Bubble formation occurs only when the valuation measure transitions from a uniformly integrable martingale to a non-uniformly integrable martingale.