Optimal Bubble Riding with Price-Dependent Entry:
a Mean Field Game of Controls with Common Noise
Abstract.
In this paper we further extend the optimal bubble riding model proposed in [62] by allowing for price-dependent entry times. Agents are characterized by their individual entry threshold that represents their belief in the strength of the bubble. Conversely, the growth dynamics of the bubble is fueled by the influx of players. Price-dependent entry naturally leads to a mean field game of controls with common noise and random entry time, for which we provide an existence result. The equilibrium is obtained by first solving discretized versions of the game in the weak formulation and then examining the measurability property in the limit. In this paper, the common noise comes from two sources: the price of the asset which all agents trade, and also the exogenous bubble burst time, which we also discretize and incorporate into the model via progressive enlargement of filtration.
April 23, 2025
1. Introduction
Financial bubbles have become a topic of growing concern in the recent past. The classical view of Beaver [10] that investors only follow “information content” is clearly not applicable to bubbles. Empirical evidence demonstrates the substantial stock price premium following vacuous company announcements of broad intentions to enter the cryptocurrency market [3] and investors’ overreaction to description of Blockchain activities in firms’ 8-K disclosures [22]. Journeying back another thirty years, a similar “gold rush” occurs during the dot-com bubble. Instead of avoiding the overpriced technology segment, sophisticated investors such as hedge funds invest heavily in the bubble while acknowledging an inevitable burst in the future [42, 33]. The herding behavior is more prevalent now due to the increasing democratization of investing, as evidenced by the dramatic surge of retail traders during the “meme-stock” frenzy [24]. Recent advancements in generative artificial intelligence (AI) unleash a frenzy both on Wall Street and among retail investors, pushing up stock prices of big technology companies. As NVIDIA’s market capitalization marches across the trillion dollar line, many begin to suspect overvaluation in the semiconductor market. However, despite the unprecedented interest rate hikes and the recent turmoil in the cryptocurrency market, enthusiasm towards a potential AI bubble has not dampened. The intricate interplay between the intention to leverage rapid growth and the apprehension towards a future price adjustment provides the motivation for our equilibrium-based model proposed in [62], which we further investigate in this paper.
Substantial empirical evidence points to the inaccuracy of viewing bubbles merely as “irrational exuberance”. A wealth of literature exists on the topic of rationality behind bubbles. The famous “greater fool” model formulated by [5], as well as a more recent adaptation [53], pinpoints the driving factor behind bubble riding as the perception that others will acquire the overpriced asset in the future. Abreu and Brunnermeier [1] explore the idea of “information asymmetry” from another perspective by giving investors different entry times and various priors on the bubble formation time (see also [29, 30] for extensions of this model). Sotes-Paladino and Zapatero [61] use a dynamic trading model to show that sophisticated, risk-averse money managers can invest in overvalued non-benchmark asset due to the presence of convex incentives. The authors in [58, 7] argue that a “chain of middlemen” could also spur the escalation of the asset price.
Despite extensive debates surrounding how a bubble is formed, or even defined, it is commonly agreed that the influx of investors and capital is what sustains the overvaluation. In other words, it is natural to model these events in a large population setting. This is the motivation behind the use of a mean field game (MFG) in our previous paper [62], which should be understood as the infinite population limit of symmetric stochastic differential games [26, 51, 45, 16]. First introduced by Lasry and Lions [50, 48, 49] and also by Huang et al. [35, 36], mean field games provide tractable solutions compared to large but finite population games. We refer the readers to the monographs of Carmona and Delarue [17, 18] for the probabilistic approach and to the notes of Cardaliaguet [15] for the partial differential equation (PDE) approach to MFGs.
Our previous study [62] introduced a class of MFGs with varying entry times. Players begin to take advantage of the rising price trajectory at different times during the “awareness window”, a period viewed by Abreu and Brunnermeier [1] as a measure of heterogeneity among traders. The inflow of traders in turn fuels the price dynamics, whose drift is a function of the number of players currently in the game. We modeled the burst of the bubble as the minimum of exogenous and endogenous burst times. While an exogenous crash occurs due to events independent of trading, an endogenous crash happens when the average inventory of the players in the game falls below a threshold. We also included price impact as a second source of interactions among the agents through the controls, leading to an extended MFG. We proved existence of MFG equilibrium using the method initiated in [19]. Leveraging established methods on filtration enlargement, we were able to decompose the optimal strategy into before-and-after-burst segments, each part being progressively measurable with respect to the original filtration. Numerically, our model discovers that the equilibrium strategy attempts to delay the burst time and therefore sustain the growth if the bubble is large enough.
The aim of this paper is to remove two major limitations of the model in [62]. The first is that the entry times could not depend on the price of the traded asset. They were instead modeled as independent samples from the same pre-determined distribution. However, an intuitive entry criterion for bubble riders is the first time the price crosses a certain threshold, which we use to characterize player influx in the present work. The second improvement is that we allow for an unbounded drift term in the price dynamics. In particular, since players’ entry affects the price, the drift term will depend on the running maximum process of the price itself. We also provide a well-posedness result for this path-dependent dynamics.
As natural as these changes may seem, they require the model to incorporate a “common noise” which is famously challenging because the law of the population has to react to the realization of the noise. Just as in the case of a classical MFG, there are two approaches to deal with common noise. The analytic approach either reformulates the problem into a coupled system of stochastic PDEs or a deterministic, but infinite dimensional, PDE called the master equation (see Cardaliaguet et al. [16] for a careful presentation). Given a sufficiently smooth solution of the master equation, one can usually obtain strong results on the MFG equilibrium such as uniqueness or even regularity. However, almost all well-posedness results of the master equation require the Lasry-Lions monotonicity condition [49], or the “displacement monotonicity” condition [2, 32, 38] (see also the “anti-monotonicity” condition [55]). On the other hand, the probabilistic approach introduced by Carmona et al. [20] avoids making this assumption by a compactness argument. Since the monotonicity condition is too strong for our model, we take the latter route for constructing MFG equilibrium. A notable drawback to this compactness approach is that the controls might only be measurable with respect to a larger filtration. A well-known immersion property is enforced to ensure fairness in observing that additional information. Immersion is a crucial property in the theory of filtration enlargement [43, 65], stochastic control [44, 31], the theory of conditional McKean-Vlasov SDEs [47] and of course mean field games [20]. For an extensive discussion and generalization on both methods of tackling common noise, see [18]. Other recent extensions on related topics include MFGs with finite state space [9], restoring uniqueness of equilibrium [25], incorporating absorption [14], convergence from N-player games [26, 46], and MFGs with interactions through controls [28]. To our knowledge, there aren’t general existence results on the equilibrium of extended MFGs with common noise, which will be our main contribution in this work (Theorem 2.8). It is worth noting that although we provide a more intuitive model by incorporating common noise compared to our previous version in [62], the result is certainly weaker and less explicit, especially for numerical analysis.
The paper is structured as follows. In Section 2, we recall the features of the -player model for bubble riding from [62] and also introduce the new mechanism for price-dependent entry. Then we formulate the limit mean field game with common noise and state the assumptions as well as the main existence result. The proof is broken down into two parts. In Section 3 we first show an intermediate step with a weaker notion of admissibility. Then we strengthen the result in Section 4 by reducing the filtration for the controls.
2. Model Setup
2.1. The N-Player Game
2.1.1. Price-dependent Entry
Suppose the price trajectory starts at at time when the bubble starts. Each player is characterized by on , a price threshold for the player to deem the bubble worth riding. The player enters the game at
The value is chosen arbitrarily, but strictly greater than , in order to bound if the price never reaches the threshold. Note that in contrast to [62], the entry times are still random even if we fix the individual information due to their dependence on the common noise in . If the price trajectory is càdlàg and jumps are non-positive, then on the event that . We assume that there are initial players with thresholds who are already in the game at . That is, . Let be the vector of thresholds. Then the number of players in the game by time is
(1) |
where is the empirical CDF of the thresholds .
2.1.2. Price Dynamics in the Bubble Phase
Let denote the price process in the bubble phase before the burst. The growth of the bubble should depend on the inflow of players, , which by (1) is a function of the maximum process of the bubble price itself. This motivates the following price dynamics before burst
(2) |
where is called the bubble trend function. Because the price grows with entry, depends on the thresholds and should be monotone increasing in its second argument. We present the generalized form of the two examples for given in [62] with price-dependent entry.
Example 2.1 (Exponential Bubble).
Example 2.2 (LPPL Bubble).
The Johansen-Ledoit-Sornette (JLS) model proposed by Johansen et al. [40] uses an assumption on the hazard rate of the burst time and arrives at a mean trajectory following the log-periodic power law (LPPL). While we model the burst time very differently, we can match the shape of the process by taking with
with parameters and critical time set to . In particular, the vanilla JLS model uses measures the power law acceleration of prices, which we generalize by incorporating the impact from the players’ entry, namely
Similar to the previous example, if all players enter at , the model reduces to a standard LPPL. See [41, 59, 60] for more detailed analysis of the JLS model.
It is not obvious that the path-dependent SDE (2) is strongly solvable, since is not necessarily bounded or Lipschitz, as in the case of empirical CDF. We defer the well-posedness result to the next section (Proposition A.1), where monotonicity of the drift is the key condition that ensures existence and uniqueness of a strong solution. The risk of the bubble bursting is not reflected in (2) since this is the dynamics given that the bubble is still alive. We take a constant diffusion as the time horizon is relatively short-term by nature of a bubble.
2.1.3. Inventory and Trading Rate
Each player has initial endowment on and goes “all in” when they enters the bubble ride. For simplicity, assume that there is no transaction cost when joining a bubble ride. That is, each player joins the game with shares of the bubble asset. We also allow for negative values of , which represent a initial short (attack) position on the asset. Note that allowing price-dependent entry fixes a shortcoming of the original model proposed in [62] where only the initial inventory is assumed to be i.i.d., which implies that the players who enter later will have more initial wealth due to the higher asset price at entry.
Suppose that there is a common horizon . By choosing her trading rate after entry, the player can control her inventory trajectory by
where is fixed and are independent –dimensional Brownian motions corresponding to random streams of demand [19, 52]. A positive corresponds to buying and a negative corresponds to selling. We require on before entry for each , and takes values in a compact interval .
2.1.4. Burst Time and Post-burst Price Dynamics
Following [1] and [62], we allow the bubble to burst for both exogenous and endogenous reasons. An exogenous burst will be modeled as a non-anticipative random time, more specifically a totally inaccessible stopping time that is independent from the market information . On the other hand, an endogenous burst occurs when the inventory of the active players (i.e. those who have entered) becomes too low to sustain the frenzy of the bubble. Define the empirical measure of the inventory and the average inventory as
(3) |
Note that our assumption on the existence of initial players allows us to drop the indicator for all when . For a given inventory threshold function , define the endogenous burst as
The true burst time is defined as the first of the two events:
At burst time, the price drops by a fraction of the bubble component , defined by
(4) |
The function is referred to as the “size” of the bubble [1], or the “loss amplitude” in the JLS model [60].
The setup becomes an optimal execution problem after the crash. Trades convey information which has a long-term impact to the price dynamics. When the bubble is present, the frenzy of the bubble growth dominates the impact from selling. However, after the crash, the asset price is governed by price impact within the short horizon. We use the game-theoretic extension of the model by Almgren and Chriss [6], where the aggregate trading rate determines the instantaneous price impact.
Let be a concave (hence also continuous) function for the instantaneous impact. See [13, 54, 57] for reasons of the concavity of price impact. Define the empirical measures of controls and the permanent price impact term as
(5) |
This is the second source of interaction among players currently in the game. Since by definition and are both before entry, there is a factor in both (3) and (5) before the integral. After burst, the bubble trend is no longer present, so the price after burst follows on
(6) |
Define . Using chain rule on gives us the price dynamics
(7) |
2.1.5. Objective and Equilibrium
Player ’s cash process is modeled by
where is a continuous, strictly convex function satisfying that measures the temporary price impact that affects only the individual trader and not the price itself. The well-known linear temporary impact [6] corresponds to being quadratic. See also [19, Section 2.1] for choosing as the antiderivative of . Note that the cash process remains at the initial endowment until the player enters the game, since is kept at . Under the usual self-financing condition, the pre-burst wealth of this player follows
The players are allowed to continue trading until , even if the burst has already happened. But by definition of riding a bubble, the players do not believe in the fundamental value of the asset. Therefore, we impose a quadratic terminal inventory penalty with to encourage selling. For a fixed , we also impose a quadratic running inventory cost which Cartea et al. [21] refer to as ambiguity aversion. Adding these costs to the negative of increase in wealth, we have the total cost of player that she wants to minimize:
(8) |
for given vectors of strategies and entry thresholds . The interaction among the players appears both in the price impact term through the average trading speed and also the burst time through the average inventory. We refer the readers to [62] for more details on the model.
It is well-known that finite-player games of this type quickly becomes intractable as increases. Since the phenomenon of bubble riding fits naturally in the large-population setting, we shift our focus directly to the mean field limit of the game described above.
2.2. Mean Field Game Setup
Let be a probability space that supports independent , whose law under is where is the (one dimensional) Wiener measure. Let be a -completed filtration defined on this probability space such that are Brownian motions, the initial wealth is -measurable, and exogenous burst time is not an -stopping time. Let be the smallest filtration, on which is a stopping time, that contains . We will see in later sections that by construction, will in fact be a -totally inaccessible stopping time under mild assumptions.
Let denote the Borel subsets of a Polish space , and let denote the set of all probability measures on . Unless specified otherwise, is equipped with the topology of weak convergence of measures, and is also a Polish space. Denote the Wasserstein space (or order 1) by , that is
where the choice of is arbitrary. Equip with the 1-Wasserstein distance
Let denote the space of all càdlàg functions from to . For a fixed , define
For each , we require to be the largest value such that to avoid redundancies. Suppose . Let . Notice that the standard sup distance is no longer suitable on because it does not allow two processes to be close unless . Therefore, for each , we can define its continuous counterpart as
We then define a metric on to be
(9) |
Define with time marginal being Lebesgue, and define
(10) |
Equip and with the topology of 1-Wasserstein convergence and weak convergence, respectively. Each can uniquely disintegrate into with some measurable map . Let be the CDF of the price threshold distribution . Each can also be viewed as a -valued process , where with being the time coordinate mapping. Let and be the law of and . Then the average inventory among the players in the game and their price impact are the natural limit of (3) and (5), namely:
2.2.1. Burst Time
While the exogenous burst time is the same as the -player game, the endogenous burst now depends on . To ensure that there are already players in the game at time , we assume that and define
where is deterministic, continuous and strictly increasing with . The upper bound is set so that the bubble at least survives the initial players. These conditions guarantee enough regularity of for the equilibrium to exist (see [62, Sections 2.1.5 and 6.1]). The actual burst time is again . Throughout the paper, we work under the following assumptions.
Assumption (E).
-
(E1)
is a compact interval that includes .
-
(E2)
is absolutely continuous with respect to Lebesgue measure on satisfying . Moreover, its deterministic, non-negative intensity process is bounded by some on .
-
(E3)
has all moments. is locally bounded.
For a càdlàg process , define . Observe from Examples 2.1 and 2.2 that the bubble trend function depends on the running maximum of the price naturally through the CDF of entry thresholds. That is, the dynamics of depends on where is not necessarily Lipschitz continuous. Therefore, the SDE defining the price process may not be well-posed. We will show in the appendix that appropriate growth and monotonicity conditions on below, which we make as standing assumptions on the bubble, allow to obtain well-posedness.
Assumption (B).
-
(B1)
The bubble function is non-negative and satisfies the assumptions in Proposition A.1.
-
(B2)
The bubble size is a positive, continuous, bounded -progressively measurable process.
Using Proposition A.1, the price dynamics (7) is well defined for a fixed , namely
where the pre-burst price follows (2) and the post-burst price follows
The bubble component is defined by (4). Before crash, the bubble component dominates the drift of the price dynamics, whereas the impact term takes over after the the crash. Note that the price has exactly one jump at , and the jump size is always negative.
2.2.2. Entry Time
Since the underlying asset starts at a known value , the entry threshold should be at least this amount, that is, is a distribution on . Consider the product probability space . We naturally extend on the product space. The representative player enters the game at a random -stopping time , , where is given by
(11) |
The value is again arbitrarily chosen. Then is a bounded stopping time for every .
Lemma 2.3.
Consider in equation (11). For -almost every , the function is strictly increasing until and if it reaches . Moreover, it is left continuous with right limit on , and is jointly measurable.
Proof.
Monotonicity is obvious. To see that it is strict, we can first write . Note that the price process is -almost surely continuous except at where there is a non-positive jump. Therefore, is a monotone increasing, -almost surely continuous process, which implies strict monotonicity of .
For each such that is continuous, for . For , we have and . Take an increasing sequence . Then is also an increasing sequence which converges to some as . If , then we can find such that for all by monotonicity of . converging to implies which contradicts the definition of . The existence of right limit follows a similar argument. Joint measurability follows from Lemma A.2. ∎
2.2.3. Admissibility of Controls
By continuity of , given we can also recover the price threshold by . However, it is still useful to define admissibility of controls in two separate ways: one in terms of entry times, and the other in terms of entry thresholds. To simplify notation, we denote by (resp. ) the -algebra generated by the (resp. )-progressively measurable subsets of .
Definition 2.4.
Define the following sets for admissible controls:
-
•
For , let denote the set of square integrable, -measurable processes such that for . We also set to be the singleton of the constant process.
-
•
A time-admissible control is a process that is -measurable such that for almost all , . Let denote all such strategies.
-
•
A price-admissible control is a process that is -measurable such that for -almost all , on the random interval . Let denote all such strategies.
By Lemma 2.3, any time-admissible control induces a price-admissible control by . For each , the corresponding state process satisfies
(12) |
2.2.4. Objective and Equilibrium
Following the same derivation from the N-player game, using (8) we can define the running cost function :
(13) |
and the terminal cost function :
(14) |
Allowing to vary in each step, by Assumptions (Assumption (B)), (Assumption (E)) and Proposition A.1 we have
Given a price process , define the process . For a fixed and , the objective which the representative player minimizes over is:
If the player does not enter by time , her total cost is . This is also true by construction, see Remark (C4) below.
Remark 2.5.
We make a note that the following set of properties of the cost structure will be utilized in the proof.
-
(C1)
The running cost function is (jointly) Borel measurable and can be decomposed as
For each , , and are continuous. In addition, there exists such that for all with :
-
(C2)
The terminal cost function is almost surely continuous in . In addition, there exists such that
-
(C3)
is strictly convex in ; is convex in and .
-
(C4)
for any . for all .
Although we will focus on the specific case of the model with cost functions (13) and (14), most of our results remain true for arbitrary costs satisfying (C1) - (C4).
2.2.5. Identical Threshold Case
A special case is where everyone has the same threshold and thus enters all at the beginning. Then the bubble function does not depend on . Suppose further that also does not depend on . Then the game reduces to a fixed entry time case in [62, Proposition A.10].
2.2.6. Common Noise and Admissible Setup
Unlike idiosyncratic noise, the presence of common noise does not vanish even when the number of players goes to infinity. As a consequence, we need to consider “random versions” of , which we denote as , to represent the conditional probability measures given the common noise. Specifically, the probability setup should also support random variable . Therefore, for , the objective a representative agent minimizes is
(15) |
where follows (12).
In our setup, there are two sources of common noise to the players: a Brownian motion from the price process and a jump process for the exogenous burst. For any stochastic process and random variable , define their natural filtration where is the -completion of . Intuitively, are conditional laws given , so if we view as a -valued process, it should be -adapted. The natural filtration to work with is the completion of . An equilibrium of this type is called a strong solution, which is known to be very hard to obtain (see e.g. the monograph [18]). Instead, we look for a weak equilibrium in the sense of [44, 20] where we only require to be the conditional law of state and control processes given both the common noise and the law process itself.
We collect all the components from this section in the next definition in a more general setting where we do not assume that the underlying probability space has a product structure.
Definition 2.6.
An admissible probability setup is a filtered probability space satisfying the usual conditions that supports the following mutually independent random elements:
-
(1)
A two-dimensional Brownian motion .
-
(2)
-measurable initial data with law .
-
(3)
A -stopping time with law , from which we can define the jump process .
If an admissible probability setup additionally supports taking values in , we can then define , the price process and random entry time
Observe that may not be defined for every threshold value that takes, making this setup slightly weaker. Similarly, we will also weaken the notion of price-admissibility and let denote the set of processes that is -progressive measurable such that -almost surely, . In fact, Lemma 2.3 ensures that our -by- construction is also a particular case under this new definition.
It is worth noting that if is also independent from , then will be an -totally inaccessible stopping time (see Remark 3.5). In particular, if is just , this would be a desired feature for the exogenous burst time because the admissible controls can only react to it once occurs but cannot anticipate it.
Definition 2.7.
A weak MFG equilibrium with strong control is an admissible probability setup that supports a -measurable random variable , paired with optimal control and corresponding state process satisfying (12) such that
-
(1)
The filtration .
-
(2)
minimizes over the objective defined in (15).
-
(3)
is a version of the conditional law of given under . That is,
-
(4)
is a version of the conditional law of given under . That is,
Theorem 2.8.
Under Assumptions (Assumption (B)) and (Assumption (E)), there exists a weak MFG equilibrium with strong control.
3. Existence of MFG Solutions with Weak Control
3.1. Weak Controls
The term “strong control” in Definition 2.7 refers to the fact that is an valued process that is -progressive. We shall prove Theorem 2.8 by following the chain of arguments presented in [20]. Specifically, we use a fixed point and compactness argument by discretizing the common noise and then taking weak limit to obtain an equilibrium. To ensure that the limit exists, we first work with relaxed controls in a larger filtration.
3.1.1. Relaxed Controls
Since the space of uniformly bounded functions is not compact, a standard workaround when analyzing extended MFGs, especially in the presence of common noise, is to consider relaxed controls. A relaxed control is a randomized strategy taking values in where
Any can be characterized, with dt a.s. uniqueness, by the form where is a Borel measurable mapping. Therefore, we can view each as a -valued process. For a given admissible probability setup, the set of admissible relaxed controls is defined as
A strict control refers to the case where is -almost surely a Dirac measure almost everywhere. The state process corresponding to a relaxed control is
(16) |
Define as the subset of whose first projection is Lebesgue measure on . Any can be characterized, with a.s. uniqueness, by such that . We naturally extend any bounded measurable function to by
In particular, for . Recall from Remark (C1) that we have separability between and in the cost . Therefore, when evaluating (or rather its extension) on an element of , we can drop the underline from the notation to avoid further confusion. In particular, for a bounded measurable function , sometimes we slightly abuse the notation by using to mean if is in instead of . Endow with the stable topology, which is the weakest topology making the map continuous, for each bounded measurable function that is continuous in the measure variable for each . Since is convex, compact and metrizable, so is . See [39] for details.
3.1.2. Immersion Property and Lifted Environment
In this section, we will also weaken the first requirement in Definition 2.7 and work with a filtration that is potentially larger than . Allowing more information into the system immediately requires extra care to ensure fairness in observing that additional information. A widely-used procedure is to check that is immersed in . This notion of fairness is also called the (H)-hypothesis, natural extension, or compatibility. It is a crucial property in the theory of filtration enlargement [43, 65], stochastic control [44, 31], the theory of conditional McKean-Vlasov SDEs [47] and of course mean field games [20].
Definition 3.1.
A filtration is said to be immersed in another filtration defined on the same probability space if and every square integrable -martingale is a square integrable -martingale. An -adapted càd-làg process with values in a Polish space is compatible with if its natural filtration is immersed in .
The following proposition is a useful characterization of this property and explains how compatibility weakens the adaptedness to a conditional independence requirement, which is mainly a property of laws. See e.g. [18, Proposition 1.3] for a proof.
Proposition 3.2.
On probability space , consider two filtrations . The following statements are equivalent.
-
(1)
is immersed in .
-
(2)
is conditionally independent from given for every .
-
(3)
For any , .
Specifically, the third statement in Proposition 3.2 allows us to eventually recover a strong control in Section 4 from a larger filtration. To ensure that we carry enough information in the smaller filtration for the immersion property to eventually hold, we will in a lifted environment [18]. Instead of , we require the admissible probability setup to support a random joint probability measure that represents the conditional law of given common noise. Let and denote its first and third marginals, which serves the same purpose of in the objective (17). Now we have all the ingredients to define a solution with weak control.
Definition 3.3.
A weak MFG equilibrium with weak control is an admissible probability setup that supports a -measurable random variable , paired with optimal relaxed control and corresponding state process satisfying (16) such that
-
(1)
The filtration is immersed in .
-
(2)
minimizes over the relaxed objective defined in (17).
-
(3)
is a version of the conditional law of given under .
-
(4)
is a version of the conditional law of given under .
It is worth noting that both definitions are weak in the probabilistic sense, where the probability space is part of the solution. They are also both weak in the sense of control theory, where the equilibrium strategy is not necessarily measurable with respect to the Brownian motions, but potentially depends on additional randomness.
The usual fixed point argument using compactness no longer applies to these conditional probability measures as their domain becomes too large. To combat the infinite dimensionality issue, Carmona et al. [20] discretizes time and space to reduce the common noise to a finite dimension process and then pass to the limit. We adapt the discretization scheme from [18], also used in [14]. In this section, our goal is to prove the following intermediate result.
Theorem 3.4.
Under Assumptions (Assumption (B)) and (Assumption (E)), there exists a weak MFG equilibrium with weak control.
3.2. Weak Formulation and Enlargement of Filtration
Since the probability space is part of the solution, it is convenient to work on the canonical space with the product structure in Section 2.2.2. We will also work under the weak formulation as in [19]. Define
and let be the corresponding identity maps. Let be a -algebra carrying the above random variables. Define the corresponding probability measures
Define entry time in a -by- way on as (11). Lemma 2.3 ensures that we have an admissible probability setup. Let denote the uncontrolled state variable on the product space :
(18) |
Given define
(19) |
By Girsanov’s theorem, and square integrability of , is a Brownian motion under and follows the state SDE (12) under . Given , the cost under the weak formulation is
(20) |
If we fix a price threshold , then is defined on by in a similar way as .
3.2.1. Progressive Enlargement of Filtration
We now recall some facts regarding filtration enlargement. Let be a filtration supporting and is independent from . Let denote the natural filtration of the exogenous burst time jump process . Define the progressively enlarged filtration
Note that is the smallest filtration which contains and such that is a -stopping time. Since is independent from , Proposition 3.2 implies that is immersed in . In particular, and remain ()–Wiener processes.
For any –predictable process , Assumption (E2) implies there exists a unique -predictable process such that (see [27, Page 186 (a)]). Since is a –submartingale, by Doob–Meyer decomposition we can find a unique, –predictable, increasing compensator process with and such that:
(21) |
is a –martingale. Under Assumption (E2), we have .
Remark 3.5.
The random time is a -inaccessible stopping time if either of the two following conditions is satisfied (see e.g. [8, 12]):
-
(1)
Every -martingale is a.s. continuous;
-
(2)
avoids all -stopping times. That is, for any -stopping time .
For example, if is just the -completed Brownian filtration with the initial enlargement of and , by martingale representation theorem (1) would be satisfied. Under the non-atomic condition in Assumption (E2), if there is independence between and , then (2) holds.
3.3. Proof of Theorem 3.4
In this section, we prove the existence of equilibrium with weak controls using backward stochastic differential equations (BSDE).
3.3.1. Backward SDEs with Random Entry Times
We begin by introducing a few notation of spaces and norms. For a filtration and probability measure on , define the following spaces of processes on :
-
•
Let denote the space of -valued -progressively measurable, càdlàg processes on satisfying
-
•
Let denote -valued -predictable processes on satisfying
-
•
Let denote -valued -predictable processes on satisfying
We drop from notation when is the probability measure. Respectively, for a probability measure on , define and in the same way for processes on . In particular, when , denotes -valued -predictable processes on satisfying
We drop from notation when considering the whole interval .
Since we take the weak formulation to MFGs, we can rewrite the objective function (20) using the solution to a BSDE. Define the Hamiltonian by
By Remark 2.5 and Assumption (E1), for each , there exists a unique element in that minimizes . For our model, the minimizer is a function of only, which we denote as . Let denote the minimized Hamiltonian, that is
(22) |
Remark 3.6.
We point out some properties of and that will be utilized later.
-
(S1)
For a general and satisfying the properties in Remark 2.5, is a jointly measurable function of and continuous in by Berge’s maximum theorem. In our case, the unique minimizer only depends on .
-
(S2)
The minimized Hamiltonian is Lipschitz in , and it is jointly continuous in for fixed .
Recall the definition of in (21). For a given , consider a generic type of BSDEs on the enlarged filtration solved on
(23) |
where is a martingale orthogonal to . A solution to the BSDE (23) is a process on the probability space . If the pre-enlarged filtration is generated by the Brownian motions, then . Note that the BSDE above is solved on a random interval even after conditioning on a . The following proposition addresses the solvability of this BSDE. To differentiate the two types of admissibility, we denote a time-admissible control in by and price-admissible control in by .
Proposition 3.7.
Suppose that and fix a -progressive . Given , for each , there exists a unique solution to the following BSDE
(24) |
where follows
If we define the process for each , then is time admissible and induces a price-admissible control . Moreover, minimizes (20) over .
Proof.
Well-posedness follows from [56, Theorem 53.1]. We need to show that is jointly measurable when composing the -by- solutions. We first show that is -almost surely left-continuous in .
Suppose we have a sequence , and let and be the corresponding control processes. Then we have
The first term goes to by dominated convergence theorem since is assumed to be bounded. To show the convergence of the second term, by continuity of it suffices to show in . By the stability of BSDE solutions (e.g. [56, Proposition 54.2]), we have
It is easy to check that for any , we have , and by (S2) left-continuity is proved. The second condition in Lemma A.2 is satisfied, and invoking the lemma yields time-admissibility.
Therefore, for each , we can solve (23) with random entry time on and obtain . Note that altering only affects the initial inventory and entry time . Since , and also by Lemma 2.3, the function is left-continuous for -almost every . Therefore, Lemma A.2 again implies joint measurability and therefore price-admissibility of .
Observe that (20) can be rewritten as for each price-admissible , where is the conditional objective given . That is
Then optimality of follows if we show conditional optimality of for each , which we fix from this point on. Take any price-admissible strategy . We can uniquely solve the following BSDE
(25) |
We can also solve (25) on with as input. Then by (22) and uniqueness, the solution coincides with the solution of (23) on the interval . Comparison principle of (25)([62, Proposition 4.3]) implies -almost surely. Optional stopping theorem and Remark (C4) imply
∎
For the remainder of this section up until Remark 3.16, we take , so Proposition 3.7 applies. This result implies that for a given we can find by solving the problem -by-, and it is well defined for every , not just almost every . On the other hand, by uniqueness of the optimizer, we can also obtain by solving the BSDE on the whole product space.
Corollary 3.8.
3.3.2. Fixed Point from Discretization
Instead of conditional measure flows given common noise , we look at a piecewise constant approximation process. Suppose for , a partition on and a finite set (some grid on space) are given. Define the -valued finite process on by
(29) |
where each is a measurable random variable taking values in the finite set, for . These knots are meant to approximate the Brownian common noise . We also have an additional source of common noise: the external burst time , which requires us to discretize the jump process .
Let denote the space of processes on of the form for some . Equip with the natural metric . For , define the valued process on by:
(30) |
Then it is obvious that for all :
(31) |
Let denote the (finite) -algebra generated by , and choose ’s such that for every . We now define the input domain for conditional laws of the state and control. At this stage, we can work with -valued controls. However, in anticipation of taking the limit in the space of relaxed controls, we switch to strict controls now. For , call its corresponding strict control. Define
(32) | ||||
(33) |
where for each , denotes the conditional law under given . Equip each coordinate of with the Wasserstein metric. Denote by an element in .
Lemma 3.9.
The process is -almost surely continuous for each for all . Consequentially, the mapping is -almost surely continuous on the closure of , where for , .
Proof.
Take a sequence in . Note that the event that does not converge to is, up-to a -null set, contained in . Then by dominated convergence theorem, it suffices to show that for all , for each , which is implied by . This follows from price dynamics (2) and the fact that the only jump of is negative by Assumptions (Assumption (B)). Then continuity of mean processes for each implies continuity of the endogenous burst mapping (see the proof of [62, Theorem 6.1]). Taking closure in Wasserstein space preserves the continuity of the mean processes. ∎
Lemma 3.10.
The set is Wasserstein pre-compact and convex in , and is Wasserstein compact and convex in .
Proof.
The statement on is immediate given that is compact and convex. For convexity of , take . It suffices to show that for all , there is such that
See [62, Lemma 4.7] for a proof. As for Wasserstein pre-compactness, we show that for each marginal. Since the dimensionality is finite, it suffices to check for each which follows from [62, Lemma 5.9] with the additional fact that for arbitrary non-negative random variable , and for all . ∎
Suppose we take any and . Define the inputs to the optimization problem: for each ,
(34) |
Again can be viewed as a -measurable random process taking values in . By Proposition 3.7 we obtain strict, optimal control with along with the probability measure on , and is a Brownian motion under . Denote by the optimal control in the strict relaxed form . Define output conditional measures :
(35) |
We have now defined our fixed point mapping:
Lemma 3.11.
For each , the mapping is continuous.
Proof.
We shall fix and drop the notation to avoid confusion with the proof steps for sequential continuity. Recall is obtained by solving the optimization problem -by-. Let denote the conditional probability measure on from Girsanov transform for a given . Take a sequence of vectors in the sense that each coordinate converges in Wasserstein distance, and define correspondingly for . Then almost surely, converges to in and converges to in . For define the discretized conditional probability measures and as in (34), and let be part of the unique solution to the BSDE (23) with and input . Let be the optimal control given .
We first show that in for every . Recall that optimal controls are continuous in , so it suffices to show in . Using the stability property of BSDE solutions [56, Proposition 54.2], convergence of is immediately implied if we show
where and
By continuity of we know that converges to in probability if both and in probability. Lemma 3.9 gives us convergence in burst time. Also, the inventory is continuous everywhere except at entry . Therefore, does not converge in probability to only if , which also has probability . Therefore,
By dominated convergence theorem, we have .
Lemma 3.9 implies almost surely. Under Assumption (E1) and Remark (C1), all terms converge to by dominated convergence theorem and Fubini’s theorem. Therefore, for all , in . Since is assumed to be bounded, this implies that in the sense on the product space .
For each , denote by the probability measure on from Girsanov transformation. By construction, for all and
Therefore, by boundedness of we can calculate the relative entropy
Pinsker’s inequality implies that converges to in total variation. By triangular inequality and the convergence in of controls, we have
Bounded convergence theorem yields in the stable topology. Since for each , this implies convergence of as well. Boundedness of ensures Wasserstein convergence as well. in total variation also implies weak convergence of for each . To show Wasserstein convergence, it suffices to show uniform integrability (see e.g. [63, Theorem 6.9]).
which follows from (28). Therefore, since is finite:
and continuity of holds for all . ∎
Proposition 3.12.
The mapping admits a fixed point for all .
Proof.
Let denote the closure of , which by Lemma 3.10 is convex and compact. Note that the input to the BSDE is still well-defined for , except that might not be the law of a strict control anymore. Therefore, we can define on the larger domain . Moreover, the Wasserstein closure preserves the continuity of the mean process. This implies that is still continuous on (see the proof of [62, Theorem 6.1]), so continuity of still holds. Applying Brouwer’s fixed point theorem (e.g. [4, Corollary 17.56]) yields a fixed point of . However, the range of is still strictly in since allowing to take values in doesn’t affect how we construct the optimal control and its corresponding state process. Then is also a fixed point of had we not enlarged the domain. ∎
For each , let be the equilibrium strategy from a fixed point from Proposition 3.12 and . Define the corresponding random laws as in (34). Then by construction, minimizes defined in (20) over .
Remark 3.13.
Proof.
We fix and recall that the uncontrolled state process is defined by
Recall from Remark 3.6 that is a continuous function that minimizes the Hamiltonian. For each defined on the probability space , by [56, Theorem 53.1] there exists a unique solution to the following BSDE
(36) |
Remark (S2) also implies well-posedness for the following BSDE solved on :
(37) |
whose unique solution we denote by . However, since are measurable, uniqueness of the solution implies that is almost surely zero, and coincides with the solution of (26) on . Recall from Corollary 3.8 and the construction of the fixed point mapping that almost surely, we must also have .
Recall from Remark (C4) that for any admissible control, the Hamiltonian is before entry. Then the generator of (36) is -almost surely greater than the generator of (37), and they are equal when we take . If a comparison principle for the general BSDE (36) holds, then Remark 3.13 follows from the argument in the proof of Proposition 3.7. To ease some notation, for we denote
Due to the presence of compensated martingale and orthogonal martingale , additional conditions are required for comparison principle to hold. In light of [23], a sufficient condition is the existence of an equivalent measure to such that
(38) |
is a martingale under . This probability can be constructed by defining
Since is Lipschitz in , we can apply Girsanov’s theorem and conclude by [23, Theorem 2] that comparison principle holds. Denote again by the probability measure corresponding to , which is defined as in (19). Recall that and all agree at time and the Hamiltonian is for any admissible strategy before . Using comparison principle and applying optional stopping theorem yield
The inequality holds for all -admissible . ∎
Remark 3.14.
Since the state variable is linearly controlled, for each we can define an -valued control in that corresponds to the same state process. Moreover, Remark (C3) and Jensen’s inequality imply that this control is at least as good as in terms of minimizing the objective value. See [34, Theorem A.9] and [31, Theorem 4.11] for more rigorous arguments on existence of strict controls achieving the same value and regarding the weak formulation of the problem. In other words, given inputs , the optimal -valued control is also optimal over relaxed controls for the relaxed objective (17).
3.3.3. Solutions as Weak Limit Points
Before taking , we shall “lift” the marginal laws of and to a joint law in order to carry more information along the way. Let be the strict control corresponding to the equilibrium strategy from the fixed point mapping. Recall that , taking values in denotes the initial wealth and entry threshold. Define on the lifted random law by
(39) |
Recall the notation from Definition 3.3 that and denote the first and third marginals of , which are the conditional law of under .
Lemma 3.15.
The sequence is tight.
Proof.
Assumption (E1) implies that
for some that does not depend on . Let be a stopping time and . Then
for a possibly different . These two conditions are sufficient for Aldous’ criterion for tightness of . For each , remain independent Brownian motions under , and also retain the same law throughout by independence. Compactness of implies the tightness of .
Now tightness of in the weak topology follows from the tightness of (see the proof of [18, Lemma 3.16]). As suggested by Lemma 3.9, we will need to equip with the Wasserstein metric in order to guarantee continuity of , where the product space is equipped with the metric. By [17, Corollary 5.6] and boundedness of , the proof of [18, Lemma 3.16] implies that it suffices to show uniform square-integrability of . Since is a Brownian motion under , we only need to show that
which is implied by (28). Finally, tightness of the marginals implies that of the joint law.
∎
Remark 3.16.
Let be a limit point in Lemma 3.15. Since we work with a weak limit, we only care about the law, not the processes themselves. Therefore, without loss of generality, we can take . Then let , be the respective canonical processes on , and a probability measure on . For each , define also on the law . We can obviously drop the from the notation (or even , since we can always work on the canonical space), but we keep it to avoid confusion.
Define the jump process , price process and entry time , thanks to the strong solvability from Proposition A.1. Take to be the -completed natural filtration generated by , which is again the progressively enlarged version of by , so is a -inaccessible stopping time. Let . Weak convergence of to implies that with entry time satisfies the relaxed state SDE (16) on .
Note that the existence of fixed points of the discretized problem and tightness result both hold for arbitrary choice of discretized process . Now we need to specify the time and space grids to ensure approximates well enough. We will prove Theorem 3.4 by verifying that defined on satisfies the three required conditions in Definition 3.3, namely consistency, compatibility and optimality.
3.3.4. Consistency
Under a slight abuse of notation of re-indexing , we take the same processes used in [18, 14] where time is discretized to the dyadic mesh and space projected to a more refined grid. Specifically, for a fixed , let for . Set and , where the projection map is defined as
Then on the event the process satisfies
and so we always have
The right hand side converges to in probability. Since remains a Brownian motion under each , converges to . Therefore, for all we have
(40) |
With this choice of , we have the following consistency property in the limit.
Lemma 3.17.
For all bounded, uniformly continuous functions and , we have
Proof.
This equality holds at the discretization level by (39), namely
By (40) and (31), uniform continuity of the function allows us to swap with on both sides of the equality above without changing the limits, if they exist. Boundedness of and weak convergence enables us to take and retain equality in the limit. ∎
Lemma 3.17 says that is a version of conditional distribution of given . which implies the consistency of marginals and . We want to carry the conditional joint laws because of the compatibility condition.
3.3.5. Optimality
For each , define to be the first and third marginals of . Define the state process corresponding to relaxed control any as in (16) but in the environment using , namely
Then in particular, . Recalling (17) the objective function for relaxed controls, we now define for each the objective under the environment :
Lemma 3.18.
.
Proof.
Since the convergence in Lemma 3.15 is weak, we need to uniformly approximate by bounded functions. For and , denote by the projection of on . Define and by:
(41) | |||
(42) |
Recall that here is the bubble component defined in (4). Since we only care about at the burst time, we can equivalently take . Since the price impact functions and are continuous, compactness of implies that for each , there exists some such that
For and , define the approximated objective on by
Then weak convergence implies that . To shorten the notation, let . Using (28), we have
Recall that has the same law under for each . Then similarly, Assumption (Assumption (B)) and Proposition A.1 together imply
These uniform integrability properties, along with Assumption (Assumption (B)) and the separability condition in (C1), imply that there exists such that
(43) |
and also
(44) |
Therefore, for any fixed ,
Taking limit gives
Taking on the right hand side and using (43) and (44) give the result. ∎
Let be another -admissible relaxed strategy. Following the proof of Lemma 3.18, we also have . Remarks 3.13 and 3.14 together imply that for each . Taking on both sides we have for all , so optimality is proved.
Now recall that for an -valued control , we denote by its corresponding strict control in the space of relaxed controls, where each time marginal is the Dirac measure at . Using the optimality lemma above, we can in fact show that must be a strict control.
Lemma 3.19.
There is a version of that is -progressively measurable that is a strict control taking the form for some .
Proof.
Define for and . Then and is a strict control. It is obvious that and both give rise to the same state process according to (16). Using strict convexity of in and Jensen’s inequality, we have
The inequality is strict (which contradicts with optimality of ) unless .
Lebesgue differentiation theorem allows us to define by
Then almost surely, . Note that shares the same measurability with , which by (12) is measurable. ∎
A consequence of the lemma above is that we can drop either or from the definition of and simply consider . Moreover, both optimality and consistency still hold for . In fact, this is the case for every limit point of the sequence in Lemma 3.15.
3.3.6. Compatibility
Following Definition 3.3, we need to show that is immersed in defined above. We need to keep in mind that while is -measurable, it is not -measurable due to random entry, which is why we need to treat separately.
Lemma 3.20.
The filtration is compatible with .
Proof.
By Proposition 3.2, it suffices to show that for all , is conditionally independent from given . We follow the proof of [20, Lemma 3.7].
Lemma 3.17 implies that is a -Brownian motion independent from . Fix . Consider three bounded functions and where is measurable, is measurable, and is measurable. By Lemma 3.17 and property of Brownian motion we have
Since this holds for all , almost surely we have
(45) |
where by we mean the integral for . Note that this expectation is -measurable if is -measurable.
Additionally, consider bounded functions where is Borel measurable, is measurable, and are functions from to that are and measurable, respectively. Using (45) and Lemma 3.17, we have
(46) |
where the last equality follows from the independence of and . Since and are arbitrary, we can replace them with bounded and , each with the same corresponding mesurability requirements. Then by definition of conditional expectation we have
We conclude by noting that is generated by with arbitrary and . ∎
We have then finished the proof of Theorem 3.4.
4. Strong Control and Separability by Burst
4.1. Strong Control in Original Environment
Recall from Lemma 3.19 that the weak control found in the previous section is in fact a strict control , and is progressive. In order to obtain an equilibrium with strong control, we will show that is measurable after bringing the lifted environment back to the “original” environment .
4.1.1. Back to Original Environment
The reason for lifting the environment is solely for the proof of the compatibility lemma 3.20, in particular the first and third equality in (46). Recall from (39) that we took to be the joint conditional law of given under . We did this to ease the notation in the consistency and compatibility lemmas. Notice that we did not need the full fledged joint law in deriving (46), but only the product of the marginals. This implies that for fixed , we could alternatively define for each :
This version still carries the necessary inputs to the BSDE (26) as its first and third marginals, and tightness of follows immediately from that of in Lemma 3.15. Then we take a limit point and follow the same argument in Remark 3.16 to work on the canonical space. In particular, is the canonical process on . Following the argument in Lemma 3.17, the fixed point property for each now leads to a weaker consistency in the limit. Namely, for all bounded, uniformly continuous, -valued functions with respective domains , we have
(47) |
which results from taking of the following equalities by construction
This is also sufficient for the consistency requirement in Definition 3.3. Similarly, with the equality (45) holds only for taking the form of a product, separable in each coordinate. This weaker property, however, is sufficient for (46) and hence the compatibility requirement. Since the optimality property only depends on the marginals and thus is not influenced, we can replace with in the final filtration .
Note from (47) that almost surely, is a product measure of its four marginals by uniqueness of measures on the product space. More importantly, its second and fourth are almost surely the Wiener measure and , respectively, since for each , are mutually independent under . Being complete, the filtration in the original environment coincides with from the lifted environment. Therefore, we can equivalently take , and the compatibility condition reads that is immersed in .
4.1.2. Strong Control via Optional Projection
To further strengthen the measurability property of from to , we follow the proof of [20, Proposition 4.4]. Recall that the state equation (12) with as the Brownian motion is satisfied by and . By optional projection we can find -optional processes and such that for any finite -stopping time :
(48) |
Since is immersed in , Proposition 3.2 implies that for each ,
(49) |
Using Fubini’s theorem for conditional expectation along with (49) on (12), we can replace by a modification such that almost surely
Notice that given , the bubble burst time is a -stopping time, and the bubble component is measurable. Recall also from (C1) that the running cost depends on only through , which is -adapted. Then by conditional Jensen’s inequality, Remark (C3) and (48),
By strict convexity of in , the inequality is strict unless and are both already adapted. Strict inequality would lead to a contradiction to optimality of among -progressive controls, since is -optional, hence also -progressive.
4.1.3. Exogenous Burst Time as Totally Inaccessible Stopping Time
The section above implies that we can take to begin with. This concludes the proof for the existence statement of Theorem 2.8. We now mention a desired feature for the bubble model as a corollary.
Corollary 4.1.
The exogenous burst time is a -totally inaccessible stopping time.
5. Concluding Remarks
In this paper we proposed a more realistic extension of the bubble riding game introduced in [62]. In contrast to [62] where agents were assumed to enter the game at independent and identically distributed times on an awareness window , here we allow players to enter the game when the price trajectory of the bubble asset reaches a given threshold. We also allow the initial inventory to depend on the initial (cash) investment and the price level at time of entry. Due to these improvements on the model, the resulting MFG in the limit is one with common noise in addition to non-standard features such as random entry times, interaction through the controls and possible jump of the state processes. Because the coefficients of the game do not satisfy the usual monotonicity conditions assumed in common noise MFG theory, we have to settle for existence of equilibria in a suitable weak form (see Definition 2.7). In short, the weaker, more realistic model assumptions made in the present paper result in weak, abstract equilibrium strategies whereas the stronger model assumptions made in [62] result in stronger equilibrium strategies that can be numerically simulated thus providing interesting economical insights.
Appendix A Two Auxiliary results
For a càdlàg process , denote by its running maximum. Recall from (1) and the price dynamics of the N-player game that the bubble trend function naturally depends on , which is not Lipschitz in . In general, the dynamics of asset price in the bubble phase is not well-posed. However, as the bubble is fueled by players’ entry, should be increasing in , hence also increasing in at each time since is a CDF. This monotonicity property of the path-dependent SDE (2) restores unique solvability.
Proposition A.1.
The following path-dependent SDE
(50) |
has a unique strong solution satisfying if for each fixed :
-
(1)
There exists such that for all :
-
(2)
is increasing (not necessarily strictly) in each argument.
Proof.
We adapt the proof of [11, Theorem 4.1]. The first condition guarantees a weak solution satisfying the integrability condition that is unique in law (see [37, Proposition 5.3.6 and Remark 5.3.8]). By the well-known result of Yamada and Watanabe [64], we only need to show pathwise uniqueness. Suppose and are two solutions on the same probability space with respect to the same Brownian motion . Observing that is absolutely continuous, by Tanaka’s formula we get
(51) |
We can equate the above expressions for all , implying that for almost every we have
(52) |
We now show that if , then . Define
The case is trivial if .
On the event , continuity of and implies that for all . Suppose , then there must exist where . Then define
By continuity again, for all . By definition of , we must also have for all . Monotonicity of leads to a contradiction
Therefore, and in particular, . We can then rewrite (51) as
where the last line vanishes by (52). Therefore, also satisfies (50). Similarly, one can show is also a solution. Then by uniqueness of law, we have which leads to pathwise-uniqueness and completes the proof. The integrability property easily follows from Grönwall’s inequality. ∎
The following measure theoretic result is probably well known. We give a proof since we could not find a directly citable reference.
Lemma A.2.
Let be a complete measurable space. A function is jointly measurable if for all :
-
(1)
is measurable.
-
(2)
converges to in -measure for any increasing sequence .
Proof.
First let be any closed set and let be a countable, dense subset of . For , denote by the open set . We claim that for -almost every and any , if and only if for each , there is such that . Note that we can always approximate any by an increasing sequence with elements in such that the functions converge -almost everywhere to . The claim follows almost immediately. Denoting by the preimage of , joint measurability is proved by writing
∎
References
- Abreu and Brunnermeier [2003] Dilip Abreu and Markus K. Brunnermeier. Bubbles and crashes. Econometrica, 71(1):173–204, 2003.
- Ahuja [2016] Saran Ahuja. Wellposedness of mean field games with common noise under a weak monotonicity condition. SIAM Journal on Control and Optimization, 54(1):30–48, 2016.
- Akyildirim et al. [2020] Erdinç Akyildirim, Shaen Corbet, Douglas Cumming, Brian Lucey, and Ahmet Sensoy. Riding the wave of crypto-exuberance: The potential misusage of corporate blockchain announcements. Technological Forecasting and Social Change, 159:120191, 2020. ISSN 0040-1625.
- Aliprantis and Border [2006] Charalambos D. Aliprantis and Kim C. Border. Infinite Dimensional Analysis: a Hitchhiker’s Guide. Springer, 3rd edition, 2006.
- Allen et al. [1993] F. Allen, S. Morris, and A. Postlewaite. Finite bubbles with short sale constraints and asymmetric information. Journal of Economic Theory, 61(2):206–229, 1993. ISSN 0022-0531.
- Almgren and Chriss [2001] Robert Almgren and Neil Chriss. Optimal execution of portfolio transactions. Journal of Risk, pages 5–39, 2001.
- Awaya et al. [2022] Yu Awaya, Kohei Iwasaki, and Makoto Watanabe. Rational bubbles and middlemen. Theoretical Economics, 17(4):1559–1587, 2022.
- Azéma et al. [1993] J. Azéma, T. Jeulin, F. Knight, and M. Yor. Le théorème d’arrêt en une fin d’ensemble prévisible. In Séminaire de Probabilités XXVII, pages 133–158. Springer Berlin Heidelberg, Berlin, Heidelberg, 1993.
- Bayraktar et al. [2021] Erhan Bayraktar, Alekos Cecchin, Asaf Cohen, and François Delarue. Finite state mean field games with wright–fisher common noise. Journal de Mathématiques Pures et Appliquées, 147:98–162, 2021.
- Beaver [1968] William H. Beaver. The information content of annual earnings announcements. Journal of Accounting Research, 6:67–92, 1968.
- Belfadli et al. [2009] Rachid Belfadli, Said Hamadéne, and Youssef Ouknine. On one-dimensional stochastic differential equations involving the maximum process. Stochastics and Dynamics, 9(02):277–292, 2009.
- Blanchet-Scalliet and Jeanblanc [2004] Christophette Blanchet-Scalliet and Monique Jeanblanc. Hazard rate for credit risk and hedging defaultable contingent claims. Finance and Stochastics, 8(1):145–159, 2004.
- Bouchaud et al. [2009] Jean-Philippe Bouchaud, J. Doyne Farmer, and Fabrizio Lillo. Chapter 2 - how markets slowly digest changes in supply and demand. In Thorsten Hens and Klaus Reiner Schenk-Hoppé, editors, Handbook of Financial Markets: Dynamics and Evolution, Handbooks in Finance, pages 57–160. North-Holland, 2009.
- Burzoni and Campi [2021] Matteo Burzoni and Luciano Campi. Mean field games with absorption and common noise with a model of bank run. arXiv preprint, 2021.
- Cardaliaguet [2012] Pierre Cardaliaguet. Notes on mean field games. In Technical report, Université de Paris - Dauphine, 2012.
- Cardaliaguet et al. [2019] Pierre Cardaliaguet, François Delarue, Jean-Michel Lasry, and Pierre-Louis Lions. The Master Equation and the Convergence Problem in Mean Field Games: (AMS-201). Princeton University Press, 2019. ISBN 9780691193717.
- Carmona and Delarue [2018a] René Carmona and François Delarue. Probabilistic Theory of Mean Field Games with Applictaions, volume I. Springer, 1st edition, 2018a.
- Carmona and Delarue [2018b] René Carmona and François Delarue. Probabilistic Theory of Mean Field Games with Applictaions, volume II. Springer, 1st edition, 2018b.
- Carmona and Lacker [2015] René Carmona and Daniel Lacker. A probabilistic weak formulation of mean field games and applications. The Annals of Applied Probability, 25(3):1189–1231, 2015.
- Carmona et al. [2016] René Carmona, François Delarue, and Daniel Lacker. Mean field games with common noise. The Annals of Probability, 44(6):3740–3803, 2016.
- Cartea et al. [2017] Alvaro Cartea, Ryan Francis Donnelly, and Sebastian Jaimungal. Algorithmic trading with model uncertainty. SIAM Journal on Financial Mathematics, 8(1):635–671, 2017. ISSN 1945-497X.
- Cheng et al. [2019] Stephanie F. Cheng, Gus De Franco, Haibo Jiang, and Pengkai Lin. Riding the blockchain mania: Public firms’ speculative 8-k disclosures. Management Science, 65(12):5901–5913, 2019.
- Cohen et al. [2010] Samuel N. Cohen, Robert J. Elliott, and Charles E. M. Pearce. A general comparison theorem for backward stochastic differential equations. Advances in Applied Probability, 42(3):878–898, 2010.
- Credit Suisse Corporate Insights Group [2022] Credit Suisse Corporate Insights Group. The investor landscape: Four evolving themes and their implications. Technical report, Credit Suisse, New York, NY, 2022.
- Delarue [2019] François Delarue. Restoring uniqueness to mean-field games by randomizing the equilibria. Stochastics and Partial Differential Equations: Analysis and Computations, 7(4):598–678, 2019.
- Delarue et al. [2020] François Delarue, Daniel Lacker, and Kavita Ramanan. From the master equation to mean field game limit theory: Large deviations and concentration of measure. Annals of Probability, 48:211–263, 2020.
- Dellacherie et al. [1992] C. Dellacherie, B. Maisonneuve, and Meyer P.A. Probabilités et potentiel, chapitres XVII-XXIV, Processus de Markov (fin). Compléments de calcul stochastique, volume 5. Hermann, 1992.
- Djete [2021] Mao Fabrice Djete. Mean field games of controls: on the convergence of nash equilibria, 2021.
- Doblas-Madrid [2012] Antonio Doblas-Madrid. A robust model of bubbles with multidimensional uncertainty. Econometrica, 80(5):1845–1893, 2012.
- Doblas-Madrid [2016] Antonio Doblas-Madrid. A finite model of riding bubbles. Journal of Mathematical Economics, 65:154–162, 2016.
- el et al. [1987] Karoui Nicole el, Nguyen Du’hŪŪ, and Jeanblanc-Picqué Monique. Compactification methods in the control of degenerate diffusions: existence of an optimal control. Stochastics, 20(3):169–219, 1987. doi: 10.1080/17442508708833443.
- Gangbo et al. [2022] Wilfrid Gangbo, Alpár R. Mészáros, Chenchen Mou, and Jianfeng Zhang. Mean field games master equations with nonseparable hamiltonians and displacement monotonicity. The Annals of Probability, 50(6):2178 – 2217, 2022.
- Griffin et al. [2011] John M. Griffin, Jeffrey H. Harris, Tao Shu, and Selim Topaloglu. Who drove and burst the tech bubble? The Journal of Finance, 66(4):1251–1290, 2011.
- Haussmann and Lepeltier [1990] U. G. Haussmann and J. P. Lepeltier. On the existence of optimal controls. SIAM Journal on Control and Optimization, 28(4):851–52, 05 1990.
- Huang et al. [2006] Minyi Huang, Roland P. Malhamé, and Peter E. Caines. Large population stochastic dynamic games: closed-loop McKean-Vlasov systems and the Nash certainty equivalence principle. Communications in Information & Systems, 6(3):221–252, 2006.
- Huang et al. [2007] Minyi Huang, Peter E. Caines, and Roland P. Malhame. The nash certainty equivalence principle and mckean-vlasov systems: An invariance principle and entry adaptation. In 2007 46th IEEE Conference on Decision and Control, pages 121–126, 2007.
- Ioannis Karatzas [1991] Steven E. Shreve Ioannis Karatzas. Brownian Motion and Stochastic Calculus. Graduate Texts in Mathematics. Springer New York, NY, 2nd edition, 1991.
- Jackson and Tangpi [2023] Joe Jackson and Ludovic Tangpi. Quantitative convergence for displacement monotone mean field games with controlled volatility, 2023.
- Jacod and Mémin [1981] Jean Jacod and Jean Mémin. Sur un type de convergence intermédiaire entre la convergence en loi et la convergence en probabilité. Séminaire de probabilités de Strasbourg, Tome, 15:529–546, 1981.
- Johansen et al. [1999] Anders Johansen, Didier Sornette, and Olivier Ledoit. Predicting financial crashes using discrete scale invariance. Journal of Risk, 1(4):5–32, 1999.
- Johansen et al. [2000] Anders Johansen, Olivier Ledoit, and Didier Sornette. Crashes at critical points. International Journal of Theoretical and Applied Finance, 3:219–255, 2000.
- K. Brunnermeier and Nagel [2004] Markus K. Brunnermeier and Stefan Nagel. Hedge funds and the technology bubble. The Journal of Finance, 59(5):2013–2040, 2004.
- Kharroubi and Lim [2014] Idris Kharroubi and Thomas Lim. Progressive enlargement of filtrations and Backward SDEs with jumps. Journal of Theoretical Probability, 27:683–724, 2014.
- Kurtz [2014] Thomas Kurtz. Weak and strong solutions of general stochastic models. Electronic Communications in Probability, 19:1 – 16, 2014. doi: 10.1214/ECP.v19-2833.
- Lacker [2016] Daniel Lacker. A general characterization of the mean field limit for stochastic differential games. Probability Theory and Related Fields, 165(3):581–648, 2016.
- Lacker and Flem [2022] Daniel Lacker and Luc Le Flem. Closed-loop convergence for mean field games with common noise, 2022.
- Lacker et al. [2020] Daniel Lacker, Mykhaylo Shkolnikov, and Jiacheng Zhang. Superposition and mimicking theorems for conditional mckean vlasov equations. Journal of the European Mathematical Society, 2020.
- Lasry and Lions [2006a] Jean–Michel Lasry and Pierre–Louis Lions. Jeux à champ moyen. II – horizon fini et contrôle optimal. Comptes Rendus Mathematique, 343(10):679–684, 2006a.
- Lasry and Lions [2007] Jean–Michel Lasry and Pierre–Louis Lions. Mean field games. Japanese Journal of Mathematics, 2(1):229–260, 2007.
- Lasry and Lions [2006b] Jean–Michel Lasry and Pierre-Louis Lions. Jeux à champ moyen. I – le cas stationnaire. Comptes Rendus Mathematique, 343(9):619–625, 2006b.
- Laurière and Tangpi [2022] Mathieu Laurière and Ludovic Tangpi. Convergence of large population games to mean field games with interaction through the controls. SIAM Journal on Mathematical Analysis, 54(3):3535–3574, 2022.
- Leal [2022] Laura S. Leal. Topics in High–Frequency Optimal Execution and Microstructure of Product Repricings. PhD thesis, Princeton University, 2022.
- Liu and Conlon [2018] Feng Liu and John R. Conlon. The simplest rational greater-fool bubble model. Journal of Economic Theory, 175:38–57, 2018. ISSN 0022-0531.
- Mastromatteo et al. [2014] I. Mastromatteo, B. Tóth, and J.-P. Bouchaud. Anomalous impact in reaction-diffusion financial models. Phys. Rev. Lett., 113:268701, Dec 2014. doi: 10.1103/PhysRevLett.113.268701.
- Mou and Zhang [2022] Chenchen Mou and Jianfeng Zhang. Mean field game master equations with anti-monotonicity conditions, 2022.
- Pardoux [1997] E. Pardoux. Generalized discontinuous backward stochastic differential equations. In Nicole El Karoui and Laurent Mazliak, editors, Backward Stochastic Differential Equations, chapter IV, pages 207–219. Addison Wesley Longman Limited, 1997.
- Pohl et al. [2018] Mathias Pohl, Alexander Ristig, Walter Schachermayer, and Ludovic Tangpi. The amazing power of dimensional analysis: Quantifying market impact. Market Microstructure and Liquidity, 3(3-4):1850004, 2018.
- Rubinstein and Wolinsky [1987] Ariel Rubinstein and Asher Wolinsky. Middlemen. The Quarterly Journal of Economics, 102(3):581–593, 08 1987.
- Sornette [2003] D Sornette. Critical market crashes. Physics Reports, 378(1):1–98, Apr 2003.
- Sornette et al. [2013] Didier Sornette, Ryan Woodard, Wanfeng Yan, and Wei-Xing Zhou. Clarifications to questions and criticisms on the Johansen–Ledoit–Sornette financial bubble model. Physica A: Statistical Mechanics and its Applications, 392(19):4417–4428, Oct 2013.
- Sotes-Paladino and Zapatero [2018] Juan Sotes-Paladino and Fernando Zapatero. Riding the Bubble with Convex Incentives. The Review of Financial Studies, 32(4):1416–1456, 07 2018.
- Tangpi and Wang [2022] Ludovic Tangpi and Shichun Wang. Optimal bubble riding: A mean field game with varying entry times. arXiv preprint, 2022. URL https://arxiv.org/abs/2209.04001.
- Villani [2009] Cédric Villani. Optimal Transport. Springer Berlin, Heidelberg, 2009.
- Yamada and Watanabe [1971] Toshio Yamada and Shinzo Watanabe. On the uniqueness of solutions of stochastic differential equations. Journal of Mathematics of Kyoto University, 11(1):155–167, 1971.
- Yor [1980] Marc Yor. Application d’un lemme de Jeulin au grossissement de la filtration brownienne. Séminaire de probabilités de Strasbourg, 14:189–199, 1980.
Princeton University
Operations Research and Financial Engineering
Email address: shichun.wang@princeton.edu
Princeton University
Operations Research and Financial Engineering
Bendheim Center for Finance
Email address: ludovic.tangpi@princeton.edu