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

00footnotetext: The authors wish to thank Camilo Hernández for fruitful discussions. L.T. is partially supported by the NSF grant DMS-2005832, the NSF CAREER award DMS-2143861 and the AMS Claytor-Gilmer fellowship. S.W. is partially supported by the NSF grant DMS-2005832.

Optimal Bubble Riding with Price-Dependent Entry:
a Mean Field Game of Controls with Common Noise

Ludovic Tangpi  and  Shichun Wang
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 NN-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 PtP_{t} starts at P0>0P_{0}>0 at time t=0t=0 when the bubble starts. Each player ii is characterized by pii.i.dνpp^{i}\stackrel{{\scriptstyle\text{i.i.d}}}{{\sim}}\nu_{p} on [P0,)[P_{0},\infty), a price threshold for the player to deem the bubble worth riding. The player enters the game at

tiinf{t0:Ptpi}(T+1).t^{i}\coloneqq\inf\{t\geq 0:P_{t}\geq p^{i}\}\wedge(T+1).

The value T+1T+1 is chosen arbitrarily, but strictly greater than TT, in order to bound 𝒯\mathcal{T} 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 pip^{i} due to their dependence on the common noise in PP. If the price trajectory is càdlàg and jumps are non-positive, then Pti=piP_{t^{i}}=p^{i} on the event that tiTt^{i}\leq T. We assume that there are initial players with thresholds P0P_{0} who are already in the game at t=0t=0. That is, νp({P0})>0\nu_{p}(\{P_{0}\})>0. Let p=(p1,,pN)\vec{p}=(p^{1},\dots,p^{N}) be the vector of thresholds. Then the number of players in the game by time tt is

Nin(t,P;p)=i=1N𝟙{tit}=i=1N𝟙{maxstPspi}=NFpN,p(maxstPs)N_{in}(t,P;\vec{p})=\sum_{i=1}^{N}\mathds{1}_{\{t^{i}\leq t\}}=\sum_{i=1}^{N}\mathds{1}_{\{\max_{s\leq t}P_{s}\geq p^{i}\}}=NF_{p}^{N,\vec{p}}(\max_{s\leq t}P_{s}) (1)

where FpN,pF_{p}^{N,\vec{p}} is the empirical CDF of the thresholds (p1,,pN)(p^{1},\dots,p^{N}).

2.1.2. Price Dynamics in the Bubble Phase

Let P+P^{+} denote the price process in the bubble phase before the burst. The growth of the bubble should depend on the inflow of players, Nin(t,P;p)N_{in}(t,P;\vec{p}), which by (1) is a function of the maximum process of the bubble price itself. This motivates the following price dynamics before burst

dPt+=b(t,maxstPs+,Pt+)dt+σ0dBt,P0+=P0,dP^{+}_{t}=b(t,\max_{s\leq t}P^{+}_{s},P^{+}_{t})dt+\sigma_{0}dB_{t},\quad P^{+}_{0}=P_{0}, (2)

where bb is called the bubble trend function. Because the price grows with entry, bb depends on the thresholds p\vec{p} and should be monotone increasing in its second argument. We present the generalized form of the two examples for bb given in [62] with price-dependent entry.

Example 2.1 (Exponential Bubble).

Abreu and Brunnermeier [1] assumed a fully deterministic model with exponential price trajectory. In our setting, this translates to

dPt+=tPt+dt+σ0dBt,t=Nin(t,P;p)N=FpN,p(maxstPs),dP^{+}_{t}=\ell_{t}P^{+}_{t}dt+\sigma_{0}dB_{t},\quad\ell_{t}=\ell\frac{N_{in}(t,P;\vec{p})}{N}=\ell F_{p}^{N,\vec{p}}(\max_{s\leq t}P_{s}),

where >0\ell>0 stands for the peak growth rate of the bubble. If we assume that everyone enters at t=0t=0, that is having pi=P0p^{i}=P_{0} for all ii, we obtain the model used in [1] with rate \ell.

Example 2.2 (LPPL Bubble).

The Johansen-Ledoit-Sornette (JLS) model proposed by Johansen et al. [40] uses an assumption on the hazard rate hth_{t} 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 dPt+=htPt+dt+σ0dBtdP^{+}_{t}=h_{t}P^{+}_{t}dt+\sigma_{0}dB_{t} with

ht=A(tct)t1+C(tct)t1cos(ωln(tct)ϕ),h_{t}=A(t_{c}-t)^{\ell_{t}-1}+C(t_{c}-t)^{\ell_{t}-1}\cos(\omega\ln(t_{c}-t)-\phi),

with parameters A,C,ω,ϕA,C,\omega,\phi and critical time tct_{c} set to TT. In particular, the vanilla JLS model uses t=(0,1)\ell_{t}=\ell\in(0,1) measures the power law acceleration of prices, which we generalize by incorporating the impact from the players’ entry, namely

t=FpN,p(maxstPs),(0,1).\ell_{t}=\ell F_{p}^{N,\vec{p}}(\max_{s\leq t}P_{s}),\quad\ell\in(0,1).

Similar to the previous example, if all players enter at t=0t=0, 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 bb 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 ii has initial endowment K0ii.i.dνKK^{i}_{0}\stackrel{{\scriptstyle\text{i.i.d}}}{{\sim}}\nu_{K} on \mathbb{R} 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 K0i/piK^{i}_{0}/p^{i} shares of the bubble asset. We also allow for negative values of KiK^{i}, 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 T>0T>0. By choosing her trading rate αi=(αti)titT\alpha^{i}=(\alpha^{i}_{t})_{t^{i}\leq t\leq T} after entry, the player can control her inventory trajectory by

dXti=αtidt+σdWti,Xti=0 on t<ti,Xtii=K0i/pi.dX_{t}^{i}=\alpha_{t}^{i}dt+\sigma dW^{i}_{t},\quad X_{t}^{i}=0\text{ on }t<t^{i},\quad X^{i}_{t^{i}}=K^{i}_{0}/p^{i}.

where σ>0\sigma>0 is fixed and Wi,,WNW^{i},\dots,W^{N} are independent 11–dimensional Brownian motions corresponding to random streams of demand [19, 52]. A positive αt\alpha_{t} corresponds to buying and a negative αt\alpha_{t} corresponds to selling. We require αti=0\alpha_{t}^{i}=0 on t<tit<t^{i} before entry for each i{1,,N}i\in\{1,\dots,N\}, and αt\alpha_{t} takes values in a compact interval AA\subset\mathbb{R}.

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 τ\tau that is independent from the market information (B,{Wi}i=1,,N)(B,\{W^{i}\}_{i=1,\dots,N}). 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 μtN\mu^{N}_{t} and the average inventory μ¯t\bar{\mu}_{t} as

μtN1Ni=1NδXti,μ¯tN𝟙{Nin(t,P,p)0}Nin(t,P;p)i=1NXti=𝟙{Nin(t,P,p)0}FpN,p(maxstPs)xμtN(dx).\mu^{N}_{t}\coloneqq\frac{1}{N}\sum_{i=1}^{N}\delta_{X^{i}_{t}},\quad\bar{\mu}_{t}^{N}\coloneqq\frac{\mathds{1}_{\{N_{in}(t,P,\vec{p})\neq 0\}}}{N_{in}(t,P;\vec{p})}\sum_{i=1}^{N}X_{t}^{i}=\frac{\mathds{1}_{\{N_{in}(t,P,\vec{p})\neq 0\}}}{F^{N,\vec{p}}_{p}(\max_{s\leq t}P_{s})}\int_{\mathbb{R}}x\mu_{t}^{N}(dx). (3)

Note that our assumption on the existence of initial players allows us to drop the indicator for all t[0,T]t\in[0,T] when NN\to\infty. For a given inventory threshold function ζ:[0,T]+\zeta:[0,T]\to\mathbb{R}_{+}, define the endogenous burst as

τ¯N(μN)inf{t>mini{1,2,,N}ti:infs[0,t]μ¯sNζt}T.\bar{\tau}^{N}(\mu^{N})\coloneqq\inf\left\{t>\min_{i\in\{1,2,\dots,N\}}t^{i}:\inf_{s\in[0,t]}\bar{\mu}^{N}_{s}\leq\zeta_{t}\right\}\wedge T.

The true burst time is defined as the first of the two events:

τ(μN)ττ¯N(μN).\tau^{*}(\mu^{N})\coloneqq\tau\wedge\bar{\tau}^{N}(\mu^{N}).

At burst time, the price drops by a fraction βt\beta_{t} of the bubble component γt\gamma_{t}, defined by

γt0tb(s,maxusPu+,Ps+)𝑑s.\gamma_{t}\coloneqq\int_{0}^{t}b(s,\max_{u\leq s}P^{+}_{u},P^{+}_{s})ds. (4)

The function β:[0,T]+\beta:[0,T]\to\mathbb{R}_{+} 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 ρ:A\rho:A\to\mathbb{R} 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 θN\theta^{N} and the permanent price impact term ρ,θtNFpN,p\left\langle\rho,\theta^{N}_{t}\right\rangle_{F_{p}^{N,\vec{p}}} as

θtN1Ni=1Nδαti,ρ,θtNFpN,p𝟙{Nin(t,P,p)0}Nin(t,P;p)i=1Nρ(αti)=𝟙{Nin(t,P,p)0}FpN,p(maxstPs)Aρ(a)θtN(da).\theta^{N}_{t}\coloneqq\frac{1}{N}\sum_{i=1}^{N}\delta_{\alpha^{i}_{t}},\quad\left\langle\rho,\theta^{N}_{t}\right\rangle_{F_{p}^{N,\vec{p}}}\coloneqq\frac{\mathds{1}_{\{N_{in}(t,P,\vec{p})\neq 0\}}}{N_{in}(t,P;\vec{p})}\sum_{i=1}^{N}\rho(\alpha_{t}^{i})=\frac{\mathds{1}_{\{N_{in}(t,P,\vec{p})\neq 0\}}}{F^{N,\vec{p}}_{p}(\max_{s\leq t}P_{s})}\int_{A}\rho(a)\theta^{N}_{t}(da). (5)

This is the second source of interaction among players currently in the game. Since by definition αi\alpha^{i} and XiX^{i} are both 0 before entry, there is a factor 1/FpN1/F^{N}_{p} in both (3) and (5) before the integral. After burst, the bubble trend is no longer present, so the price PP^{-} after burst follows on [τ,T][\tau^{*},T]

dPt=ρ,θtNFpN,p+σ0dBt,Pτ=Pτ+βτγτ.dP^{-}_{t}=\left\langle\rho,\theta^{N}_{t}\right\rangle_{F_{p}^{N,\vec{p}}}+\sigma_{0}dB_{t},\quad P^{-}_{\tau^{*}}=P^{+}_{\tau*}-\beta_{\tau^{*}}\gamma_{\tau^{*}}. (6)

Define Dt=𝟙{tτ}D^{*}_{t}=\mathds{1}_{\{t\geq\tau^{*}\}}. Using chain rule on Pt=Pt+(1Dt)+PtDtP_{t}=P^{+}_{t}(1-D_{t}^{*})+P^{-}_{t}D_{t}^{*} gives us the price dynamics

dPt=(1Dt)dPt+Pt+dDt+DtdPt+PtdDt=𝟙{t<τ}dPt++𝟙{tτ}dPtγτβτdDt.\begin{split}dP_{t}&=(1-D_{t}^{*})dP^{+}_{t}-P_{t}^{+}dD_{t}^{*}+D_{t}^{*}dP^{-}_{t}+P^{-}_{t}dD_{t}^{*}\\ &=\mathds{1}_{\{t<\tau^{*}\}}dP^{+}_{t}+\mathds{1}_{\{t\geq\tau^{*}\}}dP^{-}_{t}-\gamma_{\tau^{*}}\beta_{\tau^{*}}dD^{*}_{t}.\end{split} (7)

2.1.5. Objective and Equilibrium

Player ii’s cash process is modeled by

dKti=αtiPtκ(αti)dt,K0iνk.dK^{i}_{t}=-\alpha^{i}_{t}P_{t}-\kappa(\alpha^{i}_{t})dt,\quad K^{i}_{0}\sim\nu_{k}.

where κ()\kappa(\cdot) is a continuous, strictly convex function satisfying κ(0)=0\kappa(0)=0 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 κ\kappa being quadratic. See also [19, Section 2.1] for choosing κ\kappa as the antiderivative of ρ\rho. Note that the cash process remains at the initial endowment until the player enters the game, since αi\alpha^{i} is kept at 0. Under the usual self-financing condition, the pre-burst wealth ViV^{i} of this player follows

dVti\displaystyle dV^{i}_{t} =dKti+XtidPt+PtdXti\displaystyle=dK^{i}_{t}+X^{i}_{t}dP_{t}+P_{t}dX^{i}_{t}
=(κ(αti)+Xtib(t,maxstPs+,Pt+)𝟙{t<τ}+Xtiρ,θtNFpN,p𝟙{tτ})dt\displaystyle=\left(-\kappa(\alpha^{i}_{t})+X_{t}^{i}b(t,\max_{s\leq t}P^{+}_{s},P^{+}_{t})\mathds{1}_{\{t<\tau^{*}\}}+X_{t}^{i}\left\langle\rho,\theta^{N}_{t}\right\rangle_{F_{p}^{N,\vec{p}}}\mathds{1}_{\{t\geq\tau^{*}\}}\right)dt
XtiβtγtdDt+σ0XtidBt+σPtdWti.\displaystyle\qquad-X^{i}_{t}\beta_{t}\gamma_{t}dD^{*}_{t}+\sigma_{0}X_{t}^{i}dB_{t}+\sigma P_{t}dW_{t}^{i}.

The players are allowed to continue trading until TT, 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 c(XTi)2c(X^{i}_{T})^{2} with c>0c>0 to encourage selling. For a fixed ϕ>0\phi>0, we also impose a quadratic running inventory cost ϕ(Xti)2\phi(X_{t}^{i})^{2} 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 ii that she wants to minimize:

JN,i(𝜶,p)𝔼[(VTiVtii)+tiTϕ(Xti)2𝑑t+c(XTi)2]=𝔼[c(XTi)2+Xτiβτγτ]+𝔼[tiT(κ(αti)+ϕ(Xti)2Xtib(t,maxstPs+,Pt+)𝟙{t<τ}+Xtiρ,θtNFpN,p𝟙{tτ})𝑑t]\begin{split}J^{N,i}(\boldsymbol{\alpha},\vec{p})&\coloneqq\mathbb{E}\left[-(V_{T}^{i}-V_{t^{i}}^{i})+\int_{t^{i}}^{T}\phi(X_{t}^{i})^{2}dt+c(X_{T}^{i})^{2}\right]=\mathbb{E}\left[c(X^{i}_{T})^{2}+X^{i}_{\tau^{*}}\beta_{\tau^{*}}\gamma_{\tau^{*}}\right]\\ &+\mathbb{E}\left[\int_{t^{i}}^{T}\left(\kappa(\alpha^{i}_{t})+\phi(X^{i}_{t})^{2}-X_{t}^{i}b(t,\max_{s\leq t}P^{+}_{s},P^{+}_{t})\mathds{1}_{\{t<\tau^{*}\}}+X_{t}^{i}\left\langle\rho,\theta^{N}_{t}\right\rangle_{F_{p}^{N,\vec{p}}}\mathds{1}_{\{t\geq\tau^{*}\}}\right)dt\right]\end{split} (8)

for given vectors of strategies 𝜶=(α1,,αN)\boldsymbol{\alpha}=(\alpha^{1},\dots,\alpha^{N}) and entry thresholds p=(p1,,pN)\vec{p}=(p^{1},\dots,p^{N}). 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 NN 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 (Ω,,)(\Omega,\mathcal{F},\mathbb{P}) be a probability space that supports independent (K0,W,B,τ)(K_{0},W,B,\tau), whose law under \mathbb{P} is νK×𝒲×𝒲×ντ\nu_{K}\times\mathcal{W}\times\mathcal{W}\times\nu_{\tau} where 𝒲\mathcal{W} is the (one dimensional) Wiener measure. Let 𝔽=(t)t[0,T]\mathbb{F}=(\mathcal{F}_{t})_{t\in[0,T]} be a \mathbb{P}-completed filtration defined on this probability space such that W,BW,B are (𝔽,)(\mathbb{F},\mathbb{P}) Brownian motions, the initial wealth K0K_{0} is 0\mathcal{F}_{0}-measurable, and exogenous burst time τ\tau is not an 𝔽\mathbb{F}-stopping time. Let 𝔾\mathbb{G} be the smallest filtration, on which τ\tau is a stopping time, that contains 𝔽\mathbb{F}. We will see in later sections that by construction, τ\tau will in fact be a 𝔾\mathbb{G}-totally inaccessible stopping time under mild assumptions.

Let (E)\mathcal{B}(E) denote the Borel subsets of a Polish space (E,d)(E,d), and let 𝒫(E)\mathcal{P}(E) denote the set of all probability measures on (E)\mathcal{B}(E). Unless specified otherwise, 𝒫(E)\mathcal{P}(E) is equipped with the topology of weak convergence of measures, and 𝒫(E)\mathcal{P}(E) is also a Polish space. Denote the Wasserstein space (or order 1) by 𝒫1(E)\mathcal{P}_{1}(E), that is

𝒫1(E){μ𝒫(E):Ed(x0,x)μ(dx)<}.\mathcal{P}_{1}(E)\coloneqq\left\{\mu\in\mathcal{P}(E):\int_{E}d(x_{0},x)\mu(dx)<\infty\right\}.

where the choice of x0Ex_{0}\in E is arbitrary. Equip 𝒫1(E)\mathcal{P}_{1}(E) with the 1-Wasserstein distance

𝒲1(μ,μ)infπΠ(μ×μ)E×Ed(x,y)𝑑π(x,y).\mathcal{W}_{1}(\mu,\mu^{\prime})\coloneqq\inf_{\pi\in\Pi(\mu\times\mu^{\prime})}\int_{E\times E}d(x,y)d\pi(x,y).

Let D([0,T],)D([0,T],\mathbb{R}) denote the space of all càdlàg functions from [0,T][0,T] to \mathbb{R}. For a fixed tTt^{*}\leq T, define

𝒳t{𝒙tD([0,T],):𝒙tt=0 on [0,t), continuous on [t,T]} and 𝒳t[0,T]𝒳t.\mathcal{X}^{t^{*}}\coloneqq\Big{\{}\boldsymbol{x}^{t^{*}}\in D([0,T],\mathbb{R}):\boldsymbol{x}_{t}^{t^{*}}=0\text{ on }[0,t^{*}),\text{ continuous on }[t^{*},T]\Big{\}}\text{ and }\mathcal{X}^{*}\coloneqq\bigcup_{t^{*}\in[0,T]}\mathcal{X}^{t^{*}}.

For each 𝒙t𝒳\boldsymbol{x}^{t^{*}}\in\mathcal{X}^{*}, we require tt^{*} to be the largest value such that 𝒙t𝒳t\boldsymbol{x}^{t^{*}}\in\mathcal{X}^{t^{*}} to avoid redundancies. Suppose t1,t2[0,T]t_{1},t_{2}\in{[0,T]}. Let 𝒙t1,𝒚t2𝒳\boldsymbol{x}^{t_{1}},\boldsymbol{y}^{t_{2}}\in\mathcal{X}^{*}. Notice that the standard sup distance d(𝒙t1,𝒚t2)=𝒙t1𝒚t2d(\boldsymbol{x}^{t_{1}},\boldsymbol{y}^{t_{2}})=||\boldsymbol{x}^{t_{1}}-\boldsymbol{y}^{t_{2}}||_{\infty} is no longer suitable on 𝒳\mathcal{X}^{*} because it does not allow two processes to be close unless t1=t2t_{1}=t_{2}. Therefore, for each 𝒙t𝒳\boldsymbol{x}^{t^{*}}\in\mathcal{X}^{*}, we can define its continuous counterpart 𝒙¯tC([0,T],)\bar{\boldsymbol{x}}^{t^{*}}\in C([0,T],\mathbb{R}) as

𝒙¯tt{𝒙ttt[0,t)𝒙tttt.\bar{\boldsymbol{x}}_{t}^{t^{*}}\coloneqq\begin{cases}\boldsymbol{x}_{t^{*}}^{t^{*}}&t\in[0,t^{*})\\ \boldsymbol{x}_{t}^{t^{*}}&t\geq t^{*}.\end{cases}

We then define a metric on 𝒳D([0,T],)\mathcal{X}^{*}\subset D([0,T],\mathbb{R}) to be

d𝒳(𝒙t1,𝒚t2)𝒙¯t1𝒚¯t2+|t1t2|,and𝒙t𝒳𝒙¯t+t.d_{\mathcal{X}^{*}}(\boldsymbol{x}^{t_{1}},\boldsymbol{y}^{t_{2}})\coloneqq\|\bar{\boldsymbol{x}}^{t_{1}}-\bar{\boldsymbol{y}}^{t_{2}}\|_{\infty}+|t_{1}-t_{2}|,\quad\text{and}\quad\lVert\boldsymbol{x}^{t^{*}}\rVert_{\mathcal{X}^{*}}\coloneqq\lVert\bar{\boldsymbol{x}}^{t^{*}}\rVert_{\infty}+t^{*}. (9)

Define ΘA𝒫([0,T]×A)\Theta^{A}\coloneqq\mathcal{P}([0,T]\times A) with time marginal being Lebesgue, and define

{μ𝒫(𝒳):𝒳𝒙𝒳μ(d𝒙)<}=𝒫1(𝒳).\mathcal{M}\coloneqq\left\{\mu\in\mathcal{P}(\mathcal{X}^{*}):\int_{\mathcal{X}^{*}}\lVert\boldsymbol{x}\rVert_{\mathcal{X}^{*}}\mu(d\boldsymbol{x})<\infty\right\}=\mathcal{P}_{1}(\mathcal{X}^{*}). (10)

Equip \mathcal{M} and ΘA\Theta^{A} with the topology of 1-Wasserstein convergence and weak convergence, respectively. Each θΘA\theta\in\Theta^{A} can uniquely disintegrate into θ(dt,da)=θt(da)\theta(dt,da)=\theta_{t}(da) with some measurable map tθt𝒫(A)t\mapsto\theta_{t}\in\mathcal{P}(A). Let FpF_{p} be the CDF of the price threshold distribution νp\nu_{p}. Each μ\mu\in\mathcal{M} can also be viewed as a 𝒫()\mathcal{P}(\mathbb{R})-valued process tμtt\mapsto\mu_{t}, where μt=μt1\mu_{t}=\mu\circ\mathcal{E}_{t}^{-1} with t:𝒳\mathcal{E}_{t}:\mathcal{X}^{*}\to\mathbb{R} being the time coordinate mapping. Let μ\mu and θ\theta be the law of XX and α\alpha. Then the average inventory among the players in the game and their price impact are the natural limit of (3) and (5), namely:

μ¯t1Fp(maxstPs)x,μt,ρ,θtFp1Fp(maxstPs)ρ,θt.\bar{\mu}_{t}\coloneqq\frac{1}{F_{p}(\max_{s\leq t}P_{s})}\left\langle x,\mu_{t}\right\rangle,\quad\left\langle\rho,\theta_{t}\right\rangle_{F_{p}}\coloneqq\frac{1}{F_{p}(\max_{s\leq t}P_{s})}\left\langle\rho,\theta_{t}\right\rangle.

2.2.1. Burst Time

While the exogenous burst time is the same as the NN-player game, the endogenous burst now depends on μ¯\bar{\mu}. To ensure that there are already players in the game at time t=0t=0, we assume that νp({P0})>0\nu_{p}(\{P_{0}\})>0 and define

τ¯(μ)inf{t[0,T]:infs[0,t]μ¯sζt}T,\bar{\tau}(\mu)\coloneqq\inf\left\{t\in[0,T]:\inf_{s\in[0,t]}\bar{\mu}_{s}\leq\zeta_{t}\right\}\wedge T,

where ζ:[0,T]+\zeta:[0,T]\to\mathbb{R}^{+} is deterministic, continuous and strictly increasing with ζ0(0,𝔼[K0]/P0)\zeta_{0}\in(0,\mathbb{E}[K_{0}]/P_{0}). The upper bound is set so that the bubble at least survives the initial players. These conditions guarantee enough regularity of τ¯\bar{\tau} for the equilibrium to exist (see [62, Sections 2.1.5 and 6.1]). The actual burst time is again τ(μ)=τ¯(μ)τ\tau^{*}(\mu)=\bar{\tau}(\mu)\wedge\tau. Throughout the paper, we work under the following assumptions.

Assumption (E).
  1. (E1)

    AA is a compact interval that includes 0.

  2. (E2)

    ντ\nu_{\tau} is absolutely continuous with respect to Lebesgue measure on +\mathbb{R}^{+} satisfying ντ(τ>T)>0\nu_{\tau}({\tau>T})>0. Moreover, its deterministic, non-negative intensity process k:++k:\mathbb{R}_{+}\to\mathbb{R}_{+} is bounded by some Ck>0C_{k}>0 on [0,T][0,T].

  3. (E3)

    K0K_{0} has all moments. ρ:A\rho:A\to\mathbb{R} is locally bounded.

For a càdlàg process YY, define MtY=sup0stYsM^{Y}_{t}=\sup_{0\leq s\leq t}Y_{s}. Observe from Examples 2.1 and 2.2 that the bubble trend function bb depends on the running maximum of the price PP naturally through the CDF of entry thresholds. That is, the dynamics of PP depends on Fp(MtP)F_{p}(M^{P}_{t}) where FpF_{p} 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 bb below, which we make as standing assumptions on the bubble, allow to obtain well-posedness.

Assumption (B).
  1. (B1)

    The bubble function b:[0,T]×[0,1]×b:[0,T]\times[0,1]\times\mathbb{R}\to\mathbb{R} is non-negative and satisfies the assumptions in Proposition A.1.

  2. (B2)

    The bubble size β:[0,T]×Ω\beta:[0,T]\times\Omega\to\mathbb{R} is a positive, continuous, bounded 𝔽\mathbb{F}-progressively measurable process.

Using Proposition A.1, the price dynamics (7) is well defined for a fixed (μ,θ)×ΘA(\mu,\theta)\in\mathcal{M}\times\Theta^{A}, namely

Pt=𝟙{t<τ)}Pt++𝟙{tτ}Pt,P_{t}=\mathds{1}_{\{t<\tau^{*})\}}P^{+}_{t}+\mathds{1}_{\{t\geq\tau^{*}\}}P^{-}_{t},

where the pre-burst price P+P^{+} follows (2) and the post-burst price PP^{-} follows

Pt\displaystyle P_{t}^{-} =Pτ+βτγτ+τtρ,θsFp𝑑s+σ0(BtBτ)\displaystyle=P_{\tau^{*}}^{+}-\beta_{\tau^{*}}\gamma_{\tau^{*}}+\int_{\tau^{*}}^{t}\left\langle\rho,\theta_{s}\right\rangle_{F_{p}}ds+\sigma_{0}(B_{t}-B_{\tau^{*}})
=P0+γτ+σ0Bτβτγτ+τtρ,θsFp𝑑s+σ0(BtBτ)\displaystyle=P_{0}+\gamma_{\tau^{*}}+\sigma_{0}B_{\tau^{*}}-\beta_{\tau^{*}}\gamma_{\tau^{*}}+\int_{\tau^{*}}^{t}\left\langle\rho,\theta_{s}\right\rangle_{F_{p}}ds+\sigma_{0}(B_{t}-B_{\tau^{*}})
=P0+τtρ,θsFp𝑑s+σ0Bt+(1βτ)γτ.\displaystyle=P_{0}+\int_{\tau^{*}}^{t}\left\langle\rho,\theta_{s}\right\rangle_{F_{p}}ds+\sigma_{0}B_{t}+(1-\beta_{\tau^{*}})\gamma_{\tau^{*}}.

The bubble component γ\gamma 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 τ\tau^{*}, and the jump size βτγτ-\beta_{\tau^{*}}\gamma_{\tau^{*}} is always negative.

2.2.2. Entry Time

Since the underlying asset starts at a known value P0P_{0}, the entry threshold should be at least this amount, that is, νp\nu_{p} is a distribution on [P0,)[P_{0},\infty). Consider the product probability space (Ω×[P0,),([P0,)),νp)(\Omega\times[P_{0},\infty),\mathcal{F}\otimes\mathcal{B}([P_{0},\infty)),\mathbb{P}\otimes\nu_{p}). We naturally extend (K0,W,B,τ)(K_{0},W,B,\tau) on the product space. The representative player enters the game at a random 𝔾\mathbb{G}-stopping time 𝒯(p)\mathcal{T}(p^{*}), p[P0,)p^{*}\in[P_{0},\infty), where 𝒯:Ω×[P0,)[0,T]{T+1}\mathcal{T}:\Omega\times[P_{0},\infty)\to[0,T]\cup\{T+1\} is given by

𝒯(p)inf{t[0,T]:Ptp}(T+1),a.s.\mathcal{T}(p^{*})\coloneqq\inf\{t\in[0,T]:P_{t}\geq p^{*}\}\wedge(T+1),\quad\mathbb{P}-\operatorname*{\textit{a.s.}} (11)

The value T+1T+1 is again arbitrarily chosen. Then 𝒯(p)\mathcal{T}(p^{*}) is a bounded 𝔽B\mathbb{F}^{B} stopping time for every p[P0,)p^{*}\in[P_{0},\infty).

Lemma 2.3.

Consider 𝒯:Ω×[P0,)[0,T]{T+1}\mathcal{T}:\Omega\times[P_{0},\infty)\to[0,T]\cup\{T+1\} in equation (11). For \mathbb{P}-almost every ω\omega, the function 𝒯(ω,)\mathcal{T}(\omega,\cdot) is strictly increasing until and if it reaches T+1T+1. Moreover, it is left continuous with right limit on [P0,)[P_{0},\infty), and 𝒯\mathcal{T} is jointly measurable.

Proof.

Monotonicity is obvious. To see that it is strict, we can first write 𝒯(p)=inf{t0:MtPp}(T+1)\mathcal{T}(p^{*})=\inf\{t\geq 0:M^{P}_{t}\geq p^{*}\}\wedge(T+1). Note that the price process PP is \mathbb{P}-almost surely continuous except at τ\tau^{*} where there is a non-positive jump. Therefore, MPM^{P} is a monotone increasing, \mathbb{P}-almost surely continuous process, which implies strict monotonicity of 𝒯\mathcal{T}.

For each ω\omega such that MPM^{P} is continuous, 𝒯(p)=T+1\mathcal{T}(p^{*})=T+1 for p>MTPp^{*}>M^{P}_{T}. For pMTPp^{*}\leq M^{P}_{T}, we have 𝒯(p)[0,T]\mathcal{T}(p^{*})\in[0,T] and M𝒯(p)=pM_{\mathcal{T}(p^{*})}=p^{*}. Take an increasing sequence pnpp_{n}\uparrow p^{*}. Then 𝒯(pn)\mathcal{T}(p_{n}) is also an increasing sequence which converges to some t𝒯(p)t\leq\mathcal{T}(p^{*}) as nn\to\infty. If t<𝒯(p)t<\mathcal{T}(p^{*}), then we can find t(t,𝒯(p))t^{\prime}\in(t,\mathcal{T}(p^{*})) such that pnMtPpp_{n}\leq M^{P}_{t^{\prime}}\leq p^{*} for all nn by monotonicity of MPM^{P}. pnp_{n} converging to pp^{*} implies MtP=pM^{P}_{t^{\prime}}=p^{*} which contradicts the definition of 𝒯(p)\mathcal{T}(p^{*}). 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 PP, given 𝒯=t\mathcal{T}=t^{*} we can also recover the price threshold by p=Ptp^{*}=P_{t^{*}}. 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 𝒫(𝔾)\mathcal{PM}(\mathbb{G}) (resp. 𝔽\mathbb{F}) the σ\sigma-algebra generated by the 𝔾\mathbb{G} (resp. 𝔽\mathbb{F})-progressively measurable subsets of Ω×[0,T]\Omega\times[0,T].

Definition 2.4.

Define the following sets for admissible controls:

  • For t[0,T]t^{*}\in[0,T], let 𝔸(t)\mathbb{A}({t^{*}}) denote the set of square integrable, 𝒫(𝔾)\mathcal{PM}(\mathbb{G})-measurable processes α:Ω×[0,T]A\alpha:\Omega\times[0,T]\to A such that αt=0\alpha_{t}=0 for t[0,t)t\in[0,t^{*}). We also set 𝔸(T+1)\mathbb{A}(T+1) to be the singleton of the constant 𝟎\mathbf{0} process.

  • A time-admissible control α\alpha is a process Ω×[0,T]×[0,T](ω,t,t)αtt(ω)A\Omega\times[0,T]\times[0,T]\ni(\omega,t,t^{*})\mapsto\alpha_{t}^{t^{*}}(\omega)\in A that is 𝒫(𝔾)([0,T])\mathcal{PM}(\mathbb{G})\otimes\mathcal{B}([0,T])-measurable such that for almost all tt^{*}, αt𝔸(t)\alpha^{t^{*}}\in\mathbb{A}(t^{*}). Let 𝔸\mathbb{A} denote all such strategies.

  • A price-admissible control α\alpha is a process Ω×[0,T]×[P0,)(ω,t,p)αt(ω,p)A\Omega\times[0,T]\times[P_{0},\infty)\ni(\omega,t,p^{*})\mapsto\alpha_{t}(\omega,p^{*})\in A that is 𝒫(𝔾)([P0,))\mathcal{PM}(\mathbb{G})\otimes\mathcal{B}([P_{0},\infty))-measurable such that for νp\nu_{p}-almost all pp^{*}, αt(,p)=0a.s.\alpha_{t}(\cdot,p^{*})=0\operatorname*{\textit{a.s.}} on the random interval [0,𝒯(p)T)[0,\mathcal{T}(p^{*})\wedge T). Let 𝔸\mathbb{A}^{*} denote all such strategies.

By Lemma 2.3, any time-admissible control α𝔸\alpha\in\mathbb{A} induces a price-admissible control by αt(,p)=αt𝒯(,p)()\alpha_{t}(\cdot,p^{*})=\alpha_{t}^{\mathcal{T}(\cdot,p^{*})}(\cdot). For each α𝔸\alpha\in\mathbb{A}^{*}, the corresponding state process satisfies

Xt𝒯,α=𝟙{t𝒯}K0/𝒫+0tαs𝑑s+σ(Wt𝒯W𝒯),t[0,T].X^{\mathcal{T},\alpha}_{t}=\mathds{1}_{\{t\geq\mathcal{T}\}}K_{0}/\mathscr{P}+\int_{0}^{t}\alpha_{s}ds+\sigma(W_{t\vee\mathcal{T}}-W_{\mathcal{T}}),\quad t\in[0,T]. (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 f:[0,T]×××[0,T]××Af:[0,T]\times\mathbb{R}\times\mathbb{R}\times[0,T]\times\mathbb{R}\times A\to\mathbb{R}:

f(t,x,𝔟,η,ϱ,a)=κ(a)+ϕx2x(𝔟𝟙{t<η}+ϱ𝟙{tη})f(t,x,\mathfrak{b},\upeta,\varrho,a)=\kappa(a)+\phi x^{2}-x\big{(}\mathfrak{b}\mathds{1}_{\{t<\upeta\}}+\varrho\mathds{1}_{\{t\geq\upeta\}}\big{)} (13)

and the terminal cost function g:Ω×××[0,T]g:\Omega\times\mathbb{R}\times\mathbb{R}\times[0,T]\to\mathbb{R}:

g(x1,x2,η)=cx12+βηγηx2.g(x_{1},x_{2},\upeta)=cx_{1}^{2}+\beta_{\upeta}\gamma_{\upeta}x_{2}. (14)

Allowing C>0C>0 to vary in each step, by Assumptions (Assumption (B)), (Assumption (E)) and Proposition A.1 we have

𝔼[supη[0,T],α𝔸|g(XT𝒯,α,Xη𝒯,α,η)|2]\displaystyle\mathbb{E}\biggl{[}\sup_{\upeta\in[0,T],\alpha\in\mathbb{A}^{*}}|g(X^{\mathcal{T},\alpha}_{T},X^{\mathcal{T},\alpha}_{\upeta},\upeta)|^{2}\biggr{]} C𝔼[1+(K0p)4+σ4WT4+supη[0,T]γη2+supη[0,T]Xη2]\displaystyle\leq C\mathbb{E}\left[1+\left(\frac{K_{0}}{p^{*}}\right)^{4}+\sigma^{4}W_{T}^{4}+\sup_{\upeta\in[0,T]}\gamma^{2}_{\upeta}+\sup_{\upeta\in[0,T]}X^{2}_{\upeta}\right]
C𝔼[1+(0Tb(t,MtP,Pt)𝑑t)2]\displaystyle\leq C\mathbb{E}\left[1+\left(\int_{0}^{T}b(t,M^{P}_{t},P_{t})dt\right)^{2}\right] C𝔼[1+0T|Mt|P||2𝑑t]\displaystyle\leq C\mathbb{E}\left[1+\int_{0}^{T}|M^{|P|}_{t}|^{2}dt\right]
C𝔼[1+Tsupt[0,T]Pt2]<.\displaystyle\leq C\mathbb{E}\left[1+T\sup_{t\in[0,T]}P_{t}^{2}\right]<\infty.

Given a price process PP, define the process btPb(t,MtP,Pt)b^{P}_{t}\coloneqq b(t,M^{P}_{t},P_{t}). For a fixed θ=(θt)t[0,T]\theta=(\theta_{t})_{t\in[0,T]} and μ=(μt)t[0,T]\mu=(\mu_{t})_{t\in[0,T]}, the objective which the representative player minimizes over 𝔸\mathbb{A}^{*} is:

Jμ,θ(α)=𝔼[g(XT𝒯,α,Xτ(μ)𝒯,α,τ(μ))+𝒯TTf(s,Xs𝒯,α,bsP,τ(μ),ρ,θs,αs)𝑑s].J^{\mu,\theta}(\alpha)=\mathbb{E}\left[g(X^{\mathcal{T},\alpha}_{T},X^{\mathcal{T},\alpha}_{\tau^{*}(\mu)},\tau^{*}(\mu))+\int_{\mathcal{T}\wedge T}^{T}f(s,X^{\mathcal{T},\alpha}_{s},b^{P}_{s},\tau^{*}(\mu),\left\langle\rho,\theta_{s}\right\rangle,\alpha_{s})ds\right].

If the player does not enter by time TT, her total cost is 0. 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.

  1. (C1)

    The running cost function f:[0,T]×××[0,T]××Af:[0,T]\times\mathbb{R}\times\mathbb{R}\times[0,T]\times\mathbb{R}\times A\to\mathbb{R} is (jointly) Borel measurable and can be decomposed as

    f(t,x,𝔟,η,ϱ,a)=fa(t,x,a)+fb(t,x,𝔟)𝟙{0t<η}+fc(t,x,ϱ)𝟙{tη}.f(t,x,\mathfrak{b},\upeta,\varrho,a)=f_{a}(t,x,a)+f_{b}(t,x,\mathfrak{b})\mathds{1}_{\{0\leq t<\upeta\}}+f_{c}(t,x,\varrho)\mathds{1}_{\{t\geq\upeta\}}.

    For each tt, fa(t,,)f_{a}(t,\cdot,\cdot), fb(t,,)f_{b}(t,\cdot,\cdot) and fc(t,,)f_{c}(t,\cdot,\cdot) are continuous. In addition, there exists f>0\ell_{f}>0 such that for all (t,𝔟,ϱ,x)[0,T]×××(t,\mathfrak{b},\varrho,x)\in[0,T]\times\mathbb{R}\times\mathbb{R}\times\mathbb{R} with pmp\leq m:

    |fa(t,x,a)|+|fb(t,x,𝔟)|+|fc(t,x,ϱ)|f(1+|x|2+|𝔟|2).|f_{a}(t,x,a)|+|f_{b}(t,x,\mathfrak{b})|+|f_{c}(t,x,\varrho)|\leq\ell_{f}\left(1+|x|^{2}+|\mathfrak{b}|^{2}\right).
  2. (C2)

    The terminal cost function g:Ω×××[0,T]g:\Omega\times\mathbb{R}\times\mathbb{R}\times[0,T]\to\mathbb{R} is almost surely continuous in (x1,x2,η)(x_{1},x_{2},\upeta). In addition, there exists C>0C>0 such that 𝔼[supη[0,T],α𝔸|g(XT𝒯,α,Xη𝒯,α,η)|2]C.\mathbb{E}\left[\sup_{\upeta\in[0,T],\alpha\in\mathbb{A}^{*}}|g(X^{\mathcal{T},\alpha}_{T},X^{\mathcal{T},\alpha}_{\upeta},\upeta)|^{2}\right]\leq C.

  3. (C3)

    ff is strictly convex in (a,x)(a,x); gg is convex in x1x_{1} and x2x_{2}.

  4. (C4)

    f(t,0,𝔟,η,ϱ,0)=0f(t,0,\mathfrak{b},\upeta,\varrho,0)=0 for any (t,𝔟,η,ϱ)[0,T]××[0,T]×(t,\mathfrak{b},\upeta,\varrho)\in[0,T]\times\mathbb{R}\times[0,T]\times\mathbb{R}. g(0,0,η)=0g(0,0,\upeta)=0 for all η[0,T]\upeta\in[0,T].

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 p=P0p^{*}=P_{0} and thus enters all at the beginning. Then the bubble function bb does not depend on MpM^{p}. Suppose further that bb also does not depend on PP. 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 (μ,θ)(\mu,\theta), which we denote as (μ,ϑ)(\upmu,\upvartheta), to represent the conditional probability measures given the common noise. Specifically, the probability setup should also support random variable 𝔓(μ,ϑ):Ω×ΘA\mathfrak{P}\coloneqq(\upmu,\upvartheta):\Omega\to\mathcal{M}\times\Theta^{A}. Therefore, for α𝔸\alpha\in\mathbb{A}^{*}, the objective a representative agent minimizes is

Jμ,ϑ(α)=𝔼[g(XT𝒯,α,Xτ𝒯,α,τ(μ))+𝒯TTf(s,Xs𝒯,α,bsP,τ(μ),ρ,ϑs,αs)𝑑s],J^{\upmu,\upvartheta}(\alpha)=\mathbb{E}\left[g(X^{\mathcal{T},\alpha}_{T},X^{\mathcal{T},\alpha}_{\tau^{*}},\tau^{*}(\upmu))+\int_{\mathcal{T}\wedge T}^{T}f(s,X^{\mathcal{T},\alpha}_{s},b^{P}_{s},\tau^{*}(\upmu),\left\langle\rho,\upvartheta_{s}\right\rangle,\alpha_{s})ds\right], (15)

where X𝒯,αX^{\mathcal{T},\alpha} follows (12).

In our setup, there are two sources of common noise to the players: a Brownian motion BB from the price process PP and a jump process Dt=𝟙{τt}D_{t}=\mathds{1}_{\{\tau\leq t\}} for the exogenous burst. For any stochastic process ZZ and random variable ξ\xi, define their natural filtration 𝔽Z,ξ(tZ,ξ)t[0,T]\mathbb{F}^{Z,\xi}\coloneqq(\mathcal{F}^{Z,\xi}_{t})_{t\in[0,T]} where tZ,ξ\mathcal{F}^{Z,\xi}_{t} is the \mathbb{P}-completion of σ((ξ,Zs)s[0,t])\sigma((\xi,Z_{s})_{s\in[0,t]}). Intuitively, (μ,ϑ)(\upmu,\upvartheta) are conditional laws given (B,D)(B,D), so if we view 𝔓\mathfrak{P} as a 𝒫()×𝒫(A)\mathcal{P}(\mathbb{R})\times\mathcal{P}(A)-valued process, it should be 𝔽B,D\mathbb{F}^{B,D}-adapted. The natural filtration to work with is the completion of 𝔽K0,𝒫,W,B,D\mathbb{F}^{K_{0},\mathscr{P},W,B,D}. 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 (μ,ϑ)(\upmu,\upvartheta) to be the conditional law of state and control processes given both the common noise (B,D)(B,D) and the law process 𝔓\mathfrak{P} 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 (Ω,,𝔾=(𝒢t)t[0,T],)(\Omega,\mathcal{F},\mathbb{G}=(\mathcal{G}_{t})_{t\in[0,T]},\mathbb{P}) satisfying the usual conditions that supports the following mutually independent random elements:

  1. (1)

    A two-dimensional Brownian motion (W,B)(W,B).

  2. (2)

    𝒢0\mathcal{G}_{0}-measurable initial data (K0,𝒫)×[P0,)\mathcal{I}\coloneqq(K_{0},\mathscr{P})\in\mathbb{R}\times[P_{0},\infty) with law νKνp\nu_{K}\otimes\nu_{p}.

  3. (3)

    A 𝔾\mathbb{G}-stopping time τ\tau with law ντ\nu_{\tau}, from which we can define the jump process Dt𝟙{τt}D_{t}\coloneqq\mathds{1}_{\{\tau\leq t\}}.

If an admissible probability setup additionally supports 𝔓=(μ,ϑ)\mathfrak{P}=(\upmu,\upvartheta) taking values in ×ΘA\mathcal{M}\times\Theta^{A}, we can then define τ(μ)\tau^{*}(\upmu), the price process PP and random entry time

𝒯inf{t[0,T]:Pt𝒫0}(T+1).\mathcal{T}\coloneqq\inf\{t\in[0,T]:P_{t}-\mathscr{P}\geq 0\}\wedge(T+1).

Observe that 𝒯\mathcal{T} may not be defined for every threshold value p[P0,)p^{*}\in[P_{0},\infty) that 𝒫\mathscr{P} takes, making this setup slightly weaker. Similarly, we will also weaken the notion of price-admissibility and let 𝔸\mathbb{A}^{*} denote the set of processes Ω×[0,T](ω,t)αt(ω)A\Omega\times[0,T]\ni(\omega,t)\mapsto\alpha_{t}(\omega)\in A that is 𝔾\mathbb{G}-progressive measurable such that \mathbb{P}-almost surely, αt𝟙{t[0,𝒯)}=0\alpha_{t}\mathds{1}_{\{t\in[0,\mathcal{T})\}}=0. In fact, Lemma 2.3 ensures that our pp^{*}-by-pp^{*} construction is also a particular case under this new definition.

It is worth noting that if τ\tau is also independent from 𝔓\mathfrak{P}, then τ\tau will be an 𝔽,B,W,D,𝔓\mathbb{F}^{\mathcal{I},B,W,D,\mathfrak{P}}-totally inaccessible stopping time (see Remark 3.5). In particular, if 𝔾\mathbb{G} is just 𝔽,B,W,D,𝔓\mathbb{F}^{\mathcal{I},B,W,D,\mathfrak{P}}, this would be a desired feature for the exogenous burst time because the admissible controls can only react to it once τ\tau occurs but cannot anticipate it.

Definition 2.7.

A weak MFG equilibrium with strong control is an admissible probability setup (Ω,,𝔾,)(\Omega,\mathcal{F},\mathbb{G},\mathbb{P}) that supports a 𝒢T\mathcal{G}_{T}-measurable random variable 𝔓=(μ^,ϑ^):Ω×ΘA\mathfrak{P}=(\hat{\upmu},\hat{\upvartheta}):\Omega\to\mathcal{M}\times\Theta^{A}, paired with optimal control α^𝔸\hat{\alpha}\in\mathbb{A}^{*} and corresponding state process X𝒯,α^X^{\mathcal{T},\hat{\alpha}} satisfying (12) such that

  1. (1)

    The filtration 𝔾=𝔽,B,W,D,𝔓\mathbb{G}=\mathbb{F}^{\mathcal{I},B,W,D,\mathfrak{P}}.

  2. (2)

    α^\hat{\alpha} minimizes over 𝔸\mathbb{A}^{*} the objective Jμ^,ϑ^J^{\hat{\upmu},\hat{\upvartheta}} defined in (15).

  3. (3)

    μ^\hat{\upmu} is a version of the conditional law of X𝒯,α^X^{\mathcal{T},\hat{\alpha}} given (B,D,𝔓)(B,D,\mathfrak{P}) under \mathbb{P}. That is,

    μ^t()=(Xt𝒯,α^|tB,D,𝔓) for almost all t[0,T].\hat{\upmu}_{t}(\cdot)=\mathbb{P}\left(X_{t}^{\mathcal{T},\hat{\alpha}}\in\cdot|\mathcal{F}_{t}^{B,D,\mathfrak{P}}\right)\text{ for almost all }t\in[0,T].
  4. (4)

    ϑ^\hat{\upvartheta} is a version of the conditional law of α^\hat{\alpha} given (B,D,𝔓)(B,D,\mathfrak{P}) under \mathbb{P}. That is,

    ϑ^t()=(α^t|tB,D,𝔓) for almost all t[0,T].\hat{\upvartheta}_{t}(\cdot)=\mathbb{P}\left(\hat{\alpha}_{t}\in\cdot|\mathcal{F}_{t}^{B,D,\mathfrak{P}}\right)\text{ for almost all }t\in[0,T].
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 α^\hat{\alpha} is an AA valued process that is 𝔽,B,W,D,𝔓\mathbb{F}^{\mathcal{I},B,W,D,\mathfrak{P}}-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 (B,D)(B,D) 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 Γ\Gamma where

Γ{γ𝒫([0,T]×A) with time marginal being the Lebesgue measure}.\Gamma\coloneqq\{\upgamma\in\mathcal{P}([0,T]\times A)\text{ with time marginal being the Lebesgue measure}\}.

Any γΓ\upgamma\in\Gamma can be characterized, with dt a.s. uniqueness, by the form γ(dt,da)=(γt(da)dt)t[0,T]\upgamma(dt,da)=(\upgamma_{t}(da)dt)_{t\in[0,T]} where tγt𝒫(A)t\mapsto\upgamma_{t}\in\mathcal{P}(A) is a Borel measurable mapping. Therefore, we can view each γΓ\upgamma\in\Gamma as a 𝒫(𝒫(A))\mathcal{P}(\mathcal{P}(A))-valued process. For a given admissible probability setup, the set of admissible relaxed controls is defined as

IΓ{γΓ that is 𝔾-progressive such that -almost surely γt𝟙{t[0,𝒯)}=δ0t[0,T]}.\mathrm{I}\Gamma\coloneqq\left\{\upgamma\in\Gamma\text{ that is }\mathbb{G}\text{-progressive such that $\mathbb{P}$-almost surely }\upgamma_{t}\mathds{1}_{\{t\in[0,\mathcal{T})\}}=\delta_{0}\ \forall t\in[0,T]\right\}.

A strict control refers to the case where γt\upgamma_{t} is \mathbb{P}-almost surely a Dirac measure almost everywhere. The state process corresponding to a relaxed control γIΓ\upgamma\in\mathrm{I}\Gamma is

Xt𝒯,γ=𝟙{t𝒯}K0/𝒫+0tAaγ(ds,da)+σ(Wt𝒯W𝒯),t[0,T].X^{\mathcal{T},\upgamma}_{t}=\mathds{1}_{\{t\geq\mathcal{T}\}}K_{0}/\mathscr{P}+\int_{0}^{t}\int_{A}a\upgamma(ds,da)+\sigma(W_{t\vee\mathcal{T}}-W_{\mathcal{T}}),\quad t\in[0,T]. (16)

Define Θ\Theta as the subset of 𝒫([0,T]×𝒫(A))\mathcal{P}([0,T]\times\mathcal{P}(A)) whose first projection is Lebesgue measure dtdt on [0,T][0,T]. Any θΘ\theta\in\Theta can be characterized, with dtdt a.s. uniqueness, by {θt𝒫(𝒫(A))}t[0,T]\{\theta_{t}\in\mathcal{P}(\mathcal{P}(A))\}_{t\in[0,T]} such that θ(dt,dq)=θt(dq)dt\theta(dt,dq)=\theta_{t}(dq)dt. We naturally extend any bounded measurable function F:𝒫(A)F:\mathcal{P}(A)\to\mathbb{R} to F¯:𝒫(𝒫(A))\underline{F}:\mathcal{P}(\mathcal{P}(A))\to\mathbb{R} by

F¯(θ)𝒫(A)θ(dq)F(q).\underline{F}(\theta)\coloneqq\int_{\mathcal{P}(A)}\theta(dq)F(q).

In particular, F¯(δq)=F(q)\underline{F}(\delta_{q})=F(q) for q𝒫(A)q\in\mathcal{P}(A). Recall from Remark (C1) that we have separability between aa and qq in the cost ff. Therefore, when evaluating ff (or rather its extension) on an element of 𝒫(𝒫(A))\mathcal{P}(\mathcal{P}(A)), we can drop the underline from the notation to avoid further confusion. In particular, for a bounded measurable function ρ:A\rho:A\to\mathbb{R}, sometimes we slightly abuse the notation by using ρ,θt\left\langle\rho,\theta_{t}\right\rangle to mean ρ,𝒫(A)θt(dq)\left\langle\rho,\int_{\mathcal{P}(A)}\theta_{t}(dq)\right\rangle if θ\theta is in Θ\Theta instead of ΘA\Theta^{A}. Endow Θ\Theta with the stable topology, which is the weakest topology making the map θϕ𝑑θ\theta\to\int\phi d\theta continuous, for each bounded measurable function ϕ:[0,T]×𝒫(A)\phi:[0,T]\times\mathcal{P}(A)\to\mathbb{R} that is continuous in the measure variable for each tt. Since AA is convex, compact and metrizable, so is Θ\Theta. See [39] for details.

The version of objective function 15 for relaxed controls is

Jμ,ϑ(γ)=𝔼[g(XT𝒯,γ,Xτ𝒯,γ,τ(μ))+𝒯TTAf(s,Xs𝒯,γ,bsP,τ(μ),ρ,ϑs,a)γ(da,ds)].J^{\upmu,\upvartheta}(\upgamma)=\mathbb{E}\left[g(X^{\mathcal{T},\upgamma}_{T},X^{\mathcal{T},\upgamma}_{\tau^{*}},\tau^{*}(\upmu))+\int_{\mathcal{T}\wedge T}^{T}\int_{A}f(s,X^{\mathcal{T},\upgamma}_{s},b^{P}_{s},\tau^{*}(\upmu),\left\langle\rho,\upvartheta_{s}\right\rangle,a)\upgamma(da,ds)\right]. (17)

Notice from (17) and (16) that AA-valued controls are naturally embedded in the space of relaxed controls in the form of strict controls.

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 𝔾\mathbb{G} that is potentially larger than 𝔽,B,W,D,𝔓\mathbb{F}^{\mathcal{I},B,W,D,\mathfrak{P}}. 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 𝔽,B,W,D,𝔓\mathbb{F}^{\mathcal{I},B,W,D,\mathfrak{P}} is immersed in 𝔾\mathbb{G}. 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 \mathbb{H} is said to be immersed in another filtration 𝔽\mathbb{F} defined on the same probability space if 𝔽\mathbb{H}\subset\mathbb{F} and every square integrable 𝔽\mathbb{F}-martingale is a square integrable \mathbb{H}-martingale. An 𝔽\mathbb{F}-adapted càd-làg process μ=(μt)t0\upmu=(\upmu_{t})_{t\geq 0} with values in a Polish space is compatible with 𝔽\mathbb{F} if its natural filtration 𝔽μ\mathbb{F}^{\upmu} is immersed in 𝔽\mathbb{F}.

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 (Ω,,)(\Omega,\mathcal{F},\mathbb{P}), consider two filtrations =(t)t[0,T]𝔽=(t)t[0,T]\mathbb{H}=(\mathcal{H}_{t})_{t\in[0,T]}\subset\mathbb{F}=(\mathcal{F}_{t})_{t\in[0,T]}. The following statements are equivalent.

  1. (1)

    \mathbb{H} is immersed in 𝔽\mathbb{F}.

  2. (2)

    T\mathcal{H}_{T} is conditionally independent from t\mathcal{F}_{t} given t\mathcal{H}_{t} for every t[0,T]t\in[0,T].

  3. (3)

    For any ζ𝕃1(t)\zeta\in\mathbb{L}^{1}(\mathcal{F}_{t}), 𝔼[ζ|T]=𝔼[ζ|t]\mathbb{E}[\zeta|\mathcal{H}_{T}]=\mathbb{E}[\zeta|\mathcal{H}_{t}].

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 𝔓=(μ,ϑ)\mathfrak{P}=(\upmu,\upvartheta), we require the admissible probability setup to support a random joint probability measure 𝔐\mathfrak{M} that represents the conditional law of (X𝒯,W,γ,)(X^{\mathcal{T}},W,\upgamma,\mathcal{I}) given common noise. Let 𝔐x\mathfrak{M}^{x} and 𝔐γ\mathfrak{M}^{\upgamma} denote its first and third marginals, which serves the same purpose of (μ,ϑ)(\upmu,\upvartheta) 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 (Ω,,𝔾,)(\Omega,\mathcal{F},\mathbb{G},\mathbb{P}) that supports a 𝒢T\mathcal{G}_{T}-measurable random variable 𝔐:Ω𝒫(𝒳×𝒳×Γ×2)\mathfrak{M}:\Omega\to\mathcal{P}(\mathcal{X}^{*}\times\mathcal{X}\times\Gamma\times\mathbb{R}^{2}), paired with optimal relaxed control γ^IΓ\hat{\upgamma}\in\mathrm{I}\Gamma and corresponding state process X𝒯,γ^X^{\mathcal{T},\hat{\upgamma}} satisfying (16) such that

  1. (1)

    The filtration 𝔽,B,W,D,𝔐\mathbb{F}^{\mathcal{I},B,W,D,\mathfrak{M}} is immersed in 𝔾\mathbb{G}.

  2. (2)

    γ^\hat{\upgamma} minimizes over IΓ\mathrm{I}\Gamma the relaxed objective J𝔐x,𝔐γJ^{\mathfrak{M}^{x},\mathfrak{M}^{\upgamma}} defined in (17).

  3. (3)

    𝔐x\mathfrak{M}^{x} is a version of the conditional law of X𝒯,γ^X^{\mathcal{T},\hat{\upgamma}} given (B,D,𝔐)(B,D,\mathfrak{M}) under \mathbb{P}.

  4. (4)

    𝔐γ\mathfrak{M}^{\upgamma} is a version of the conditional law of γ^\hat{\upgamma} given (B,D,𝔐)(B,D,\mathfrak{M}) under \mathbb{P}.

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

Ω1×𝒳,Ω0=𝒳×+,ΩΩ1×Ω0,ΩcΩ×[P0,)\Omega_{1}\coloneqq\mathbb{R}\times\mathcal{X},\quad\Omega_{0}=\mathcal{X}\times\mathbb{R}_{+},\quad\Omega\coloneqq\Omega_{1}\times\Omega_{0},\quad\Omega_{c}\coloneqq\Omega\times[P_{0},\infty)

and let (K0,W,B,τ,𝒫)(K_{0},W,B,\tau,\mathscr{P}) be the corresponding identity maps. Let \mathcal{F} be a σ\sigma-algebra carrying the above random variables. Define the corresponding probability measures

1νK𝒲,0𝒲ντ,10,νp.\mathbb{Q}_{1}\coloneqq\nu_{K}\otimes\mathcal{W},\quad\mathbb{Q}_{0}\coloneqq\mathcal{W}\otimes\nu_{\tau},\quad\mathbb{Q}\coloneqq\mathbb{Q}_{1}\otimes\mathbb{Q}_{0},\quad\mathbb{P}\coloneqq\mathbb{Q}\otimes\nu_{p}.

Define entry time 𝒯\mathcal{T} in a pp^{*}-by-pp^{*} way on Ωc\Omega_{c} as (11). Lemma 2.3 ensures that we have an admissible probability setup. Let X𝒯X^{\mathcal{T}} denote the uncontrolled state variable on the product space [0,T]×Ωc[0,T]\times\Omega_{c}:

Xt𝒯K0/𝒫+σ(WtW𝒯) for t𝒯andXt𝒯0 for t[0,𝒯).X^{\mathcal{T}}_{t}\coloneqq K_{0}/\mathscr{P}+\sigma(W_{t}-W_{\mathcal{T}})\text{ for }t\geq\mathcal{T}\quad\text{and}\quad X^{\mathcal{T}}_{t}\coloneqq 0\text{ for }t\in[0,\mathcal{T}). (18)

Given α𝔸\alpha\in\mathbb{A}^{*} define

dαd(0σ1αs𝑑Ws)T,WtαWt0tσ1αs𝑑s.\frac{d\mathbb{P}^{\alpha}}{d\mathbb{P}}\coloneqq\mathcal{E}\Big{(}\int_{0}^{\cdot}\sigma^{-1}\alpha_{s}dW_{s}\Big{)}_{T},\quad W^{\alpha}_{t}\coloneqq W_{t}-\int_{0}^{t}\sigma^{-1}\alpha_{s}ds. (19)

By Girsanov’s theorem, and square integrability of α\alpha, WαW^{\alpha} is a Brownian motion under α\mathbb{P}^{\alpha} and X𝒯X^{\mathcal{T}} follows the state SDE (12) under α\mathbb{P}^{\alpha}. Given (μ,ϑ)(\upmu,\upvartheta), the cost under the weak formulation is

Jweakμ,ϑ(α)𝔼α[g(XT𝒯,Xτ(μ)𝒯,τ(μ))+𝒯TTf(s,Xs𝒯,bsP,τ(μ),ρ,ϑs,αs)𝑑s].J^{\upmu,\upvartheta}_{weak}(\alpha)\coloneqq\mathbb{E}^{\mathbb{P}^{\alpha}}\left[g(X^{\mathcal{T}}_{T},X^{\mathcal{T}}_{\tau^{*}(\upmu)},\tau^{*}(\upmu))+\int_{\mathcal{T}\wedge T}^{T}f(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu),\left\langle\rho,\upvartheta_{s}\right\rangle,\alpha_{s})ds\right].\\ (20)

If we fix a price threshold pp^{*}, then α(p)\mathbb{Q}^{\alpha(p^{*})} is defined on Ω\Omega by dα(p)d\frac{d\mathbb{Q}^{\alpha(p^{*})}}{d\mathbb{Q}} in a similar way as α\mathbb{P}^{\alpha}.

3.2.1. Progressive Enlargement of Filtration

We now recall some facts regarding filtration enlargement. Let 𝔽=(t)t[0,T]\mathbb{F}=(\mathcal{F}_{t})_{t\in[0,T]} be a filtration supporting (W,B,)(W,B,\mathcal{I}) and is independent from τ\tau. Let (𝒟t)t[0,T](\mathcal{D}_{t})_{t\in[0,T]} denote the natural filtration of the exogenous burst time jump process (Dt)t[0,T](D_{t})_{t\in[0,T]}. Define the progressively enlarged filtration

𝒢tt𝒟t,𝔾=(𝒢t)t[0,T].\mathcal{G}_{t}\coloneqq\mathcal{F}_{t}\vee\mathcal{D}_{t},\qquad\mathbb{G}=(\mathcal{G}_{t})_{t\in[0,T]}.

Note that 𝔾\mathbb{G} is the smallest filtration which contains 𝔽\mathbb{F} and such that τ\tau is a 𝔾\mathbb{G}-stopping time. Since τ\tau is independent from 𝔽\mathbb{F}, Proposition 3.2 implies that 𝔽\mathbb{F} is immersed in 𝔾\mathbb{G}. In particular, WW and BB remain (𝔾,\mathbb{G},\mathbb{P})–Wiener processes.

For any 𝔾\mathbb{G}–predictable process hh, Assumption (E2) implies there exists a unique 𝔽\mathbb{F}-predictable process ff such that ht𝟙{tτ}=ft𝟙{tτ}h_{t}\mathds{1}_{\{t\leq\tau\}}=f_{t}\mathds{1}_{\{t\leq\tau\}} (see [27, Page 186 (a)]). Since DD is a 𝔾\mathbb{G}–submartingale, by Doob–Meyer decomposition we can find a unique, 𝔽\mathbb{F}–predictable, increasing compensator process KK with K0=0K_{0}=0 and such that:

Mtτ=Dt0t(1Ds)𝑑KsM^{\tau}_{t}=D_{t}-\int_{0}^{t}(1-D_{s-})dK_{s} (21)

is a (,𝔾)(\mathbb{P},\mathbb{G})–martingale. Under Assumption (E2), we have dKt=ktdtdK_{t}=k_{t}dt.

Remark 3.5.

The random time τ\tau is a 𝔾\mathbb{G}-inaccessible stopping time if either of the two following conditions is satisfied (see e.g. [8, 12]):

  1. (1)

    Every 𝔽\mathbb{F}-martingale is a.s. continuous;

  2. (2)

    τ\tau avoids all 𝔽\mathbb{F}-stopping times. That is, (τ=L)=0\mathbb{P}(\tau=L)=0 for any 𝔽\mathbb{F}-stopping time LL.

For example, if 𝔽\mathbb{F} is just the \mathbb{P}-completed Brownian filtration with the initial enlargement of KK and 𝒫\mathscr{P}, by martingale representation theorem (1) would be satisfied. Under the non-atomic condition in Assumption (E2), if there is independence between τ\tau and 𝔽\mathbb{F}, 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 \mathbb{H} and probability measure ~\widetilde{\mathbb{Q}} on Ω\Omega, define the following spaces of processes on [s,t][0,T][s,t]\subseteq[0,T]:

  • Let 𝒮,~2[s,t]\mathcal{S}^{2}_{\mathbb{H},\widetilde{\mathbb{Q}}}[s,t] denote the space of \mathbb{R}-valued \mathbb{H}-progressively measurable, càdlàg processes YY on Ω×[s,t]\Omega\times[s,t] satisfying

    Y𝒮,~2[s,t]𝔼~[supu[s,t]|Yu|2]12<.||Y||_{\mathcal{S}_{\mathbb{H},\widetilde{\mathbb{Q}}}^{2}[s,t]}\coloneqq\mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\sup_{u\in[s,t]}\left|Y_{u}\right|^{2}\right]^{\frac{1}{2}}<\infty.
  • Let ,~2[s,t]\mathcal{H}^{2}_{\mathbb{H},\widetilde{\mathbb{Q}}}[s,t] denote \mathbb{R}-valued \mathbb{H}-predictable processes ZZ on Ω×[s,t]\Omega\times[s,t] satisfying

    Z,~2[s,t]𝔼~[st|Zu|2𝑑u]12<.||Z||_{\mathcal{H}_{\mathbb{H},\widetilde{\mathbb{Q}}}^{2}[s,t]}\coloneqq\mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\int_{s}^{t}|Z_{u}|^{2}du\right]^{\frac{1}{2}}<\infty.
  • Let ,~,D2[s,t]\mathcal{H}_{\mathbb{H},\widetilde{\mathbb{Q}},D}^{2}[s,t] denote \mathbb{R}-valued \mathbb{H}-predictable processes UU on Ω×[s,t]\Omega\times[s,t] satisfying

    U,~,D2[s,t]𝔼~[st|Uu|2𝑑Du]12<.||U||_{\mathcal{H}^{2}_{\mathbb{H},\widetilde{\mathbb{Q}},D}[s,t]}\coloneqq\mathbb{E}^{\widetilde{\mathbb{Q}}}\left[\int_{s}^{t}|U_{u}|^{2}dD_{u}\right]^{\frac{1}{2}}<\infty.

We drop ~\widetilde{\mathbb{Q}} from notation when \mathbb{Q} is the probability measure. Respectively, for a probability measure ~c\widetilde{\mathbb{Q}}_{c} on Ωc\Omega_{c}, define 𝒮,c,~c2[s,t],,c,~c2[s,t]\mathcal{S}^{2}_{\mathbb{H},c,\widetilde{\mathbb{Q}}_{c}}[s,t],\mathcal{H}^{2}_{\mathbb{H},c,\widetilde{\mathbb{Q}}_{c}}[s,t] and ,c,~c,D2[s,t]\mathcal{H}_{\mathbb{H},c,\widetilde{\mathbb{Q}}_{c},D}^{2}[s,t] in the same way for processes on Ωc×[s,t]\Omega_{c}\times[s,t]. In particular, when ~c=\widetilde{\mathbb{Q}}_{c}=\mathbb{P}, ,c2[s,t]\mathcal{H}^{2}_{\mathbb{H},c}[s,t] denotes \mathbb{R}-valued \mathbb{H}-predictable processes ZZ on Ωc×[s,t]\Omega_{c}\times[s,t] satisfying

Z,c2[s,t]𝔼[st|Zu|2𝑑u]12=𝔼[[P0,)]st|Zu(p)|2𝑑uνp(dp)]12<.||Z||_{\mathcal{H}_{\mathbb{H},c}^{2}[s,t]}\coloneqq\mathbb{E}^{\mathbb{P}}\left[\int_{s}^{t}|Z_{u}|^{2}du\right]^{\frac{1}{2}}=\mathbb{E}^{\mathbb{Q}}\left[\int_{[P_{0},\infty)]}\int_{s}^{t}|Z_{u}(p^{*})|^{2}du\nu_{p}(dp^{*})\right]^{\frac{1}{2}}<\infty.

We drop [s,t][s,t] from notation when considering the whole interval [0,T][0,T].

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

H(t,x,𝔟,η,ϱ,z,a)=f(t,x,𝔟,η,ϱ,a)+σ1az.H(t,x,\mathfrak{b},\upeta,\varrho,z,a)=f(t,x,\mathfrak{b},\upeta,\varrho,a)+\sigma^{-1}az.

By Remark 2.5 and Assumption (E1), for each (t,x,m,p,η,ϱ,z)(t,x,m,p,\upeta,\varrho,z), there exists a unique element in AA that minimizes H(t,x,m,η,ϱ,z,)H(t,x,m,\upeta,\varrho,z,\cdot). For our model, the minimizer is a function of zz only, which we denote as a^(z)\hat{a}(z). Let hh denote the minimized Hamiltonian, that is

h(t,x,𝔟,η,ϱ,z)H(t,x,𝔟,η,ϱ,z,a^(z))=κ(a^(z))+ϕx2x(𝔟𝟙{t<η}+ϱ𝟙{tη})+σ1a^(z)z.h(t,x,\mathfrak{b},\upeta,\varrho,z)\coloneqq H(t,x,\mathfrak{b},\upeta,\varrho,z,\hat{a}(z))=\kappa(\hat{a}(z))+\phi x^{2}-x\big{(}\mathfrak{b}\mathds{1}_{\{t<\upeta\}}+\varrho\mathds{1}_{\{t\geq\upeta\}}\big{)}+\sigma^{-1}\hat{a}(z)z. (22)
Remark 3.6.

We point out some properties of a^\hat{a} and hh that will be utilized later.

  1. (S1)

    For a general ff and gg satisfying the properties in Remark 2.5, a^\hat{a} is a jointly measurable function of (t,x,z)(t,x,z) and continuous in zz by Berge’s maximum theorem. In our case, the unique minimizer a^()\hat{a}(\cdot) only depends on zz.

  2. (S2)

    The minimized Hamiltonian hh is Lipschitz in zz, and it is jointly continuous in (x,z,ϱ)(x,z,\varrho) for fixed (t,m,p,η)(t,m,p,\upeta).

Recall the definition of MτM^{\tau} in (21). For a given pp^{*}, consider a generic type of BSDEs on the enlarged filtration 𝔾\mathbb{G} solved on [𝒯(p),T][\mathcal{T}(p^{*}),T]

Yt=g(XT𝒯(p),Xτ𝒯(p),τ(μ))+tTh(s,Xs𝒯(p),bsP,τ(μ),ρ,ϑs,Zs)𝑑stTZs𝑑WstTs𝑑BstTUs𝑑MsτtT𝑑Ms,t[𝒯(p),T].\begin{split}Y_{t}&=g(X^{\mathcal{T}(p^{*})}_{T},X^{\mathcal{T}(p^{*})}_{\tau^{*}},\tau^{*}(\upmu))+\int_{t}^{T}h(s,X^{\mathcal{T}(p^{*})}_{s},b^{P}_{s},\tau^{*}(\upmu),\left\langle\rho,\upvartheta_{s}\right\rangle,Z_{s})ds\\ &-\int_{t}^{T}Z_{s}dW_{s}-\int_{t}^{T}\mathfrak{Z}_{s}dB_{s}-\int_{t}^{T}U_{s}dM^{\tau}_{s}-\int_{t}^{T}dM_{s},\quad t\in[\mathcal{T}(p^{*}),T].\end{split} (23)

where MM is a martingale orthogonal to (W,B,Mτ)(W,B,M^{\tau}). A solution to the BSDE (23) is a process (Y,Z,,U,M)𝒮𝔾2[𝒯(p),T]×𝒢𝔾2[𝒯(p),T]×𝔾2[𝒯(p),T]×𝔾,D2[𝒯(p),T]×𝒮𝔾2[𝒯(p),T](Y,Z,\mathfrak{Z},U,M)\in\mathcal{S}^{2}_{\mathbb{G}}[\mathcal{T}(p^{*}),T]\times\mathcal{G}^{2}_{\mathbb{G}}[\mathcal{T}(p^{*}),T]\times\mathcal{H}^{2}_{\mathbb{G}}[\mathcal{T}(p^{*}),T]\times\mathcal{H}^{2}_{\mathbb{G},D}[\mathcal{T}(p^{*}),T]\times\mathcal{S}^{2}_{\mathbb{G}}[\mathcal{T}(p^{*}),T] on the probability space (Ω,,𝔾,)(\Omega,\mathcal{F},\mathbb{G},\mathbb{P}). If the pre-enlarged filtration 𝔽\mathbb{F} is generated by the Brownian motions, then M0M\equiv 0. Note that the BSDE above is solved on a random interval even after conditioning on a pp^{*}. The following proposition addresses the solvability of this BSDE. To differentiate the two types of admissibility, we denote a time-admissible control in 𝔸\mathbb{A} by α\alpha and price-admissible control in 𝔸\mathbb{A}^{*} by α𝒯\alpha^{\mathcal{T}}.

Proposition 3.7.

Suppose that 𝔾=𝔽,W,B,D\mathbb{G}=\mathbb{F}^{\mathcal{I},W,B,D} and fix a 𝔾\mathbb{G}-progressive 𝔓=(μ,ϑ)\mathfrak{P}=(\upmu,\upvartheta). Given pp^{*}, for each t[0,T]t^{*}\in[0,T], there exists a unique solution (Yt,Zt,t,Ut)(Y^{t^{*}},Z^{t^{*}},\mathfrak{Z}^{t^{*}},U^{t^{*}}) to the following BSDE

Yt=g(XTt,Xτt,τ(μ))+tTh(s,Xst,bsP,τ(μ),ρ,ϑs,Zs)𝑑stTZs𝑑WstTs𝑑BstTUs𝑑Msτ,t[t,T].\begin{split}Y_{t}&=g(X^{t^{*}}_{T},X^{t^{*}}_{\tau^{*}},\tau^{*}(\upmu))+\int_{t}^{T}h(s,X^{t^{*}}_{s},b^{P}_{s},\tau^{*}(\upmu),\left\langle\rho,\upvartheta_{s}\right\rangle,Z_{s})ds\\ &-\int_{t}^{T}Z_{s}dW_{s}-\int_{t}^{T}\mathfrak{Z}_{s}dB_{s}-\int_{t}^{T}U_{s}dM^{\tau}_{s},\quad t\in[t^{*},T].\end{split} (24)

where XtX^{t^{*}} follows

Xtt=𝟙{tt}(K0/p+σ(WtWt)).X^{t^{*}}_{t}=\mathds{1}_{\{t\geq t^{*}\}}\left(K_{0}/p^{*}+\sigma(W_{t}-W_{t^{*}})\right).

If we define the process α^tt=𝟙{tt}a^(Ztt)𝔸t\hat{\alpha}^{t^{*}}_{t}=\mathds{1}_{\{t\geq t^{*}\}}\hat{a}(Z^{t^{*}}_{t})\in\mathbb{A}^{t^{*}} for each tt^{*}, then α^\hat{\alpha} is time admissible and induces a price-admissible control α^𝒯()𝔸\hat{\alpha}^{\mathcal{T}(\cdot)}\in\mathbb{A}^{*}. Moreover, α^𝒯\hat{\alpha}^{\mathcal{T}} minimizes (20) over 𝔸\mathbb{A}^{*}.

Proof.

Using Assumption (E2) and (21), we can rewrite (24) as

Yt=g(XTt,Xτt,τ(μ))+tTh(s,Xst,bsP,τ(μ),ρ,ϑs,Zs)+Usks𝟙{0s<τ}dstTZs𝑑WstTs𝑑BstTUs𝑑Ds,t[t,T].\begin{split}Y_{t}&=g(X^{t^{*}}_{T},X^{t^{*}}_{\tau^{*}},\tau^{*}(\upmu))+\int_{t}^{T}h(s,X^{t^{*}}_{s},b^{P}_{s},\tau^{*}(\upmu),\left\langle\rho,\upvartheta_{s}\right\rangle,Z_{s})+U_{s}k_{s}\mathds{1}_{\{0\leq s<\tau\}}ds\\ &-\int_{t}^{T}Z_{s}dW_{s}-\int_{t}^{T}\mathfrak{Z}_{s}dB_{s}-\int_{t}^{T}U_{s}dD_{s},\quad t\in[t^{*},T].\end{split}

Well-posedness follows from [56, Theorem 53.1]. We need to show that α^\hat{\alpha} is jointly measurable when composing the tt^{*}-by-tt^{*} solutions. We first show that tα^tt^{*}\mapsto\hat{\alpha}^{t^{*}} is \mathbb{P}-almost surely left-continuous in 𝔾2\mathcal{H}^{2}_{\mathbb{G}}.

Suppose we have a sequence tnt[0,T]t_{n}^{*}\uparrow t^{*}\in[0,T], and let αtn\alpha^{t^{*}_{n}} and αt\alpha^{t^{*}} be the corresponding control processes. Then we have

αtnαt𝔾2\displaystyle\lVert\alpha^{t^{*}_{n}}-\alpha^{t^{*}}\rVert_{\mathcal{H}_{\mathbb{G}}^{2}} =𝔼[0T|αttnαtt|2𝑑t]=𝔼[tnt|αttn|2𝑑t]+𝔼[tT|αttnαtt|2𝑑t]\displaystyle=\mathbb{E}\left[\int_{0}^{T}|\alpha^{t^{*}_{n}}_{t}-\alpha^{t^{*}}_{t}|^{2}dt\right]=\mathbb{E}\left[\int_{t^{*}_{n}}^{t^{*}}|\alpha^{t^{*}_{n}}_{t}|^{2}dt\right]+\mathbb{E}\left[\int_{t^{*}}^{T}|\alpha^{t^{*}_{n}}_{t}-\alpha^{t^{*}}_{t}|^{2}dt\right]
=𝔼[tnt|αttn|2𝑑t]+𝔼[tT|a^(Zttn)a^(Ztt)|2𝑑t].\displaystyle=\mathbb{E}\left[\int_{t^{*}_{n}}^{t^{*}}|\alpha^{t^{*}_{n}}_{t}|^{2}dt\right]+\mathbb{E}\left[\int_{t^{*}}^{T}\left|\hat{a}(Z^{t^{*}_{n}}_{t})-\hat{a}(Z^{t^{*}}_{t})\right|^{2}dt\right].

The first term goes to 0 by dominated convergence theorem since AA is assumed to be bounded. To show the convergence of the second term, by continuity of a^\hat{a} it suffices to show ZtnnZtZ^{t^{*}_{n}}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}Z^{t^{*}} in 𝔾2[t,T]\mathcal{H}^{2}_{\mathbb{G}}[t^{*},T]. By the stability of BSDE solutions (e.g. [56, Proposition 54.2]), we have

ZtnZt𝔾2[t,T]2=𝔼[tT|ZttnZtt|2𝑑t]C𝔼[|g(XTtn,Xτtn,τ)g(XTt,Xτt,τ)|2]\displaystyle\lVert Z^{t^{*}_{n}}-Z^{t^{*}}\rVert^{2}_{\mathcal{H}_{\mathbb{G}}^{2}[t^{*},T]}=\mathbb{E}\left[\int_{t^{*}}^{T}\left|Z^{t^{*}_{n}}_{t}-Z^{t^{*}}_{t}\right|^{2}dt\right]\leq C\mathbb{E}\left[\left|g(X^{t^{*}_{n}}_{T},X^{t^{*}_{n}}_{\tau^{*}},\tau^{*})-g(X^{t^{*}}_{T},X^{t^{*}}_{\tau^{*}},\tau^{*})\right|^{2}\right]
+C𝔼[tT|h(s,Xstn,bsP,τ,ρ,ϑs,Zst)h(s,Xst,bsP,τ,ρ,ϑs,Zst)|2𝑑s].\displaystyle\qquad+C\mathbb{E}\left[\int_{t^{*}}^{T}\left|h(s,X^{t^{*}_{n}}_{s},b^{P}_{s},\tau^{*},\left\langle\rho,\upvartheta_{s}\right\rangle,Z^{t^{*}}_{s})-h(s,X^{t^{*}}_{s},b^{P}_{s},\tau^{*},\left\langle\rho,\upvartheta_{s}\right\rangle,Z^{t^{*}}_{s})\right|^{2}ds\right].

It is easy to check that for any tt, we have XttnnXttX^{t^{*}_{n}}_{t}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}X_{t}^{t^{*}}, 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 p[P0,)p^{*}\in[P_{0},\infty), we can solve (23) with random entry time on [𝒯(p),T][\mathcal{T}(p^{*}),T] and obtain α^𝒯(p)\hat{\alpha}^{\mathcal{T}(p^{*})}. Note that altering pp^{*} only affects the initial inventory K0/pK_{0}/p^{*} and entry time 𝒯(p)\mathcal{T}(p^{*}). Since P0>0P_{0}>0, and also by Lemma 2.3, the function pα^t𝒯(p)(ω)p^{*}\mapsto\hat{\alpha}^{\mathcal{T}(p^{*})}_{t}(\omega) is left-continuous for ×λ\mathbb{Q}\times\lambda-almost every (ω,t)(\omega,t). Therefore, Lemma A.2 again implies joint measurability and therefore price-admissibility of α^𝒯()\hat{\alpha}^{\mathcal{T}(\cdot)}.

Observe that (20) can be rewritten as Jweakμ,ϑ(α𝒯)=𝔼νp[Jweakμ,ϑ(α𝒯|p)]J^{\upmu,\upvartheta}_{weak}(\alpha^{\mathcal{T}})=\mathbb{E}^{\nu_{p}}[J^{\upmu,\upvartheta}_{weak}(\alpha^{\mathcal{T}}|p^{*})] for each price-admissible α𝒯\alpha^{\mathcal{T}}, where Jweakμ,ϑ(|p)J^{\upmu,\upvartheta}_{weak}(\cdot|p^{*}) is the conditional objective given 𝒫=p\mathscr{P}=p^{*}. That is

Jweakμ,ϑ(α𝒯|p)𝔼α^𝒯(p)[g(XT𝒯(p),Xτ𝒯(p),τ(μ))+𝒯(p)TTf(s,Xs𝒯(p),bsP,τ(μ),ρ,ϑs,αs𝒯(p))𝑑s].J^{\upmu,\upvartheta}_{weak}(\alpha^{\mathcal{T}}|p^{*})\coloneqq\mathbb{E}^{\mathbb{Q}^{\hat{\alpha}^{\mathcal{T}(p^{*})}}}\left[g(X^{\mathcal{T}(p^{*})}_{T},X^{\mathcal{T}(p^{*})}_{\tau^{*}},\tau^{*}(\upmu))+\int_{\mathcal{T}(p^{*})\wedge T}^{T}f(s,X^{\mathcal{T}(p^{*})}_{s},b^{P}_{s},\tau^{*}(\upmu),\left\langle\rho,\upvartheta_{s}\right\rangle,\alpha^{\mathcal{T}(p^{*})}_{s})ds\right].

Then optimality of α^𝒯\hat{\alpha}^{\mathcal{T}} follows if we show conditional optimality of α^𝒯(p)\hat{\alpha}^{\mathcal{T}(p^{*})} for each pp^{*}, which we fix from this point on. Take any price-admissible strategy β𝒯\beta^{\mathcal{T}}. We can uniquely solve the following BSDE

Ytβ,p=g(XT𝒯(p),Xτ𝒯(p),τ(μ))+tTH(s,Xs𝒯(p),bsP,τ(μ),ρ,ϑs,Zsβ,p,βs𝒯(p))𝑑stTZsβ,p𝑑WstTsβ,p𝑑BstTUsβ,p𝑑Msτ,t[0,T].\begin{split}Y^{\beta,p^{*}}_{t}&=g(X^{\mathcal{T}(p^{*})}_{T},X^{\mathcal{T}(p^{*})}_{\tau^{*}},\tau^{*}(\upmu))+\int_{t}^{T}H(s,X^{\mathcal{T}(p^{*})}_{s},b^{P}_{s},\tau^{*}(\upmu),\left\langle\rho,\upvartheta_{s}\right\rangle,Z^{\beta,p^{*}}_{s},\beta^{\mathcal{T}(p^{*})}_{s})ds\\ &-\int_{t}^{T}Z^{\beta,p^{*}}_{s}dW_{s}-\int_{t}^{T}\mathfrak{Z}^{\beta,p^{*}}_{s}dB_{s}-\int_{t}^{T}U^{\beta,p^{*}}_{s}dM^{\tau}_{s},\quad t\in[0,T].\end{split} (25)

We can also solve (25) on [0,T][0,T] with α^\hat{\alpha} as input. Then by (22) and uniqueness, the solution coincides with the solution of (23) on the interval [𝒯(p),T][\mathcal{T}(p^{*}),T]. Comparison principle of (25)([62, Proposition 4.3]) implies Y0α^𝒯,pY0β𝒯,pY^{\hat{\alpha}^{\mathcal{T}},p^{*}}_{0}\leq Y^{\beta^{\mathcal{T}},p^{*}}_{0} \mathbb{Q}-almost surely. Optional stopping theorem and Remark (C4) imply

Jweakμ,ϑ(α^𝒯|p)=𝔼α^𝒯(p)[Y0α^,p]=𝔼[Y0α^,p]𝔼[Y0β,p]=𝔼β𝒯(p)[Y0β,p]=Jweakμ,ϑ(β𝒯|p).J^{\upmu,\upvartheta}_{weak}(\hat{\alpha}^{\mathcal{T}}|p^{*})=\mathbb{E}^{\mathbb{Q}^{\hat{\alpha}^{\mathcal{T}(p^{*})}}}\left[Y^{\hat{\alpha},p^{*}}_{0}\right]=\mathbb{E}^{\mathbb{Q}}\left[Y^{\hat{\alpha},p^{*}}_{0}\right]\leq\mathbb{E}^{\mathbb{Q}}\left[Y^{\beta,p^{*}}_{0}\right]=\mathbb{E}^{\mathbb{Q}^{\beta^{\mathcal{T}(p^{*})}}}\left[Y^{\beta,p^{*}}_{0}\right]=J^{\upmu,\upvartheta}_{weak}(\beta^{\mathcal{T}}|p^{*}).

For the remainder of this section up until Remark 3.16, we take 𝔾=𝔽,W,B,D\mathbb{G}=\mathbb{F}^{\mathcal{I},W,B,D}, so Proposition 3.7 applies. This result implies that for a given (μ,ϑ)(\upmu,\upvartheta) we can find α^\hat{\alpha} by solving the problem pp^{*}-by-pp^{*}, and it is well defined for every pp^{*}, not just νp\nu_{p} almost every pp^{*}. On the other hand, by uniqueness of the optimizer, we can also obtain α^𝒯\hat{\alpha}^{\mathcal{T}} by solving the BSDE on the whole product space.

Corollary 3.8.

Given (μ,ϑ):Ω0(,Θ)(\upmu,\upvartheta):\Omega_{0}\to(\mathcal{M},\Theta) that is 𝔽B,D,\mathbb{F}^{B,D,\mathcal{I}}-progressive, there exists a unique solution (Y,Z,,U)𝒮𝔾,c2×𝒢𝔾,c2×𝔾,c2×𝔾,c,D2(Y,Z,\mathfrak{Z},U)\in\mathcal{S}^{2}_{\mathbb{G},c}\times\mathcal{G}^{2}_{\mathbb{G},c}\times\mathcal{H}^{2}_{\mathbb{G},c}\times\mathcal{H}^{2}_{\mathbb{G},c,D} to the following BSDE on the product space in the 𝔾\mathbb{G} filtration

Yt=g(XT𝒯,Xτ(μ)𝒯,τ(μ))+tTh(s,Xs𝒯,bsP,τ(μ),ρ,ϑs,Zs)𝑑stTZs𝑑WstTs𝑑BstTUs𝑑Msτ,t[𝒯T,T].\begin{split}Y_{t}&=g(X^{\mathcal{T}}_{T},X^{\mathcal{T}}_{\tau^{*}(\upmu)},\tau^{*}(\upmu))+\int_{t}^{T}h(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu),\left\langle\rho,\upvartheta_{s}\right\rangle,Z_{s})ds\\ &-\int_{t}^{T}Z_{s}dW_{s}-\int_{t}^{T}\mathfrak{Z}_{s}dB_{s}-\int_{t}^{T}U_{s}dM^{\tau}_{s},\quad t\in[\mathcal{T}\wedge T,T].\end{split} (26)

Additionally, the process defined by α^t𝟙{t𝒯}a^(Zt)𝔸\hat{\alpha}_{t}\coloneqq\mathds{1}_{\{t\geq\mathcal{T}\}}\hat{a}(Z_{t})\in\mathbb{A}^{*} is dt\mathbb{P}\otimes dt almost surely identical to the one constructed in Proposition 3.7 and minimizes (20).

For any price-admissible α𝔸\alpha\in\mathbb{A}^{*}, let X~α\widetilde{X}^{\alpha} denote the solution of (12) defined on Ωc×[0,T]\Omega_{c}\times[0,T]. Then by Girsanov’s theorem

(X~α,,𝒯,W,B)1=α(X𝒯,,𝒯,Wα,B)1.\mathbb{P}\circ(\widetilde{X}^{\alpha},\mathcal{I},\mathcal{T},W,B)^{-1}=\mathbb{P}^{\alpha}\circ(X^{\mathcal{T}},\mathcal{I},\mathcal{T},W^{\alpha},B)^{-1}. (27)

Under Assumption (E1), for any p>0p>0 we have

supα𝔸𝔼α[supp[P0,)X𝒯(p)𝒳p]=supα𝔸𝔼[supp[P0,)X~α𝒯(p)𝒳p]<.\sup_{\alpha\in\mathbb{A}^{*}}\mathbb{E}^{\mathbb{P}^{\alpha}}\left[\sup_{p^{*}\in[P_{0},\infty)}\lVert X^{\mathcal{T}(p^{*})}\rVert_{\mathcal{X}^{*}}^{p}\right]=\sup_{\alpha\in\mathbb{A}^{*}}\mathbb{E}\left[\sup_{p^{*}\in[P_{0},\infty)}\lVert\widetilde{X}^{\alpha^{\mathcal{T}(p^{*})}}\rVert_{\mathcal{X}^{*}}^{p}\right]<\infty. (28)

3.3.2. Fixed Point from Discretization

Instead of conditional measure flows given common noise (B,D)(B,D), we look at a piecewise constant approximation process. Suppose for NN\in\mathbb{N}, a partition {0=t0<t1<<tN1<tN=T}\{0=t_{0}<t_{1}<\cdots<t_{N-1}<t_{N}=T\} on [0,T][0,T] and a finite set (some grid on space) ΛN\Lambda_{N}\subset\mathbb{R} are given. Define the ΛN\Lambda_{N}-valued finite process VNV^{N} on 𝒳\mathcal{X} by

VtN(β)i=1Nvi1(β)𝟙{t[ti1,ti)}+vN1(β)𝟙{t=T},V^{N}_{t}(\beta)\coloneqq\sum_{i=1}^{N}v_{i-1}(\beta)\mathds{1}_{\{t\in[t_{i-1},t_{i})\}}+v_{N-1}(\beta)\mathds{1}_{\{t=T\}}, (29)

where each vi:𝒳ΛNv_{i}:\mathcal{X}\to\Lambda_{N} is a tiB\mathcal{F}^{B}_{t_{i}} measurable random variable taking values in the finite set, for i=0,,N1i=0,\dots,N-1. These knots are meant to approximate the Brownian common noise Bti(β)B_{t_{i}}(\beta). We also have an additional source of common noise: the external burst time τ\tau, which requires us to discretize the jump process DD.

Let 𝒳D\mathcal{X}_{D} denote the space of processes on [0,T][0,T] of the form Dt(η)=𝟙{ηt}D_{t}(\upeta)=\mathds{1}_{\{\upeta\leq t\}} for some η[0,T]\upeta\in[0,T]. Equip 𝒳D\mathcal{X}_{D} with the natural metric d(D(η),D(η))=|ηη|d(D(\upeta),D(\upeta^{\prime}))=|\upeta-\upeta^{\prime}|. For NN\in\mathbb{N}, define the 𝒳D\mathcal{X}_{D} valued process on +\mathbb{R}_{+} by:

DtN(η)i=1N𝟙{ηti1}𝟙{t[ti1,ti)}+𝟙{ηtN1}𝟙{t=T}.D^{N}_{t}(\upeta)\coloneqq\sum_{i=1}^{N}\mathds{1}_{\{\upeta\leq t_{i-1}\}}\mathds{1}_{\{t\in[t_{i-1},t_{i})\}}+\mathds{1}_{\{\upeta\leq t_{N-1}\}}\mathds{1}_{\{t=T\}}. (30)

Then it is obvious that for all ε>0\varepsilon>0:

limNN(d(DN,D)ε)=1.\lim_{N\to\infty}\mathbb{P}^{N}\left(d(D^{N},D)\leq\varepsilon\right)=1. (31)

Let 𝒱N{A1,,A|𝒱N|}\mathcal{V}_{N}\coloneqq\{A_{1},\cdots,A_{|\mathcal{V}_{N}|}\} denote the (finite) σ\sigma-algebra generated by (VN,DN)(V^{N},D^{N}), and choose viv_{i}’s such that (Ak)>0\mathbb{P}(A_{k})>0 for every kk. We now define the input domain for conditional laws of the state and control. At this stage, we can work with AA-valued controls. However, in anticipation of taking the limit in the space of relaxed controls, we switch to strict controls now. For α𝔸\alpha\in\mathbb{A}^{*}, call γ(α)IΓ\upgamma(\alpha)\in\mathrm{I}\Gamma its corresponding strict control. Define

N\displaystyle\mathcal{M}_{N} {(α,1(X𝒯),,α,|𝒱N|(X𝒯)): for some α𝔸}𝒫1(𝒳)|𝒱N|\displaystyle\coloneqq\left\{\left(\mathcal{L}^{\alpha,1}(X^{\mathcal{T}}),\dots,\mathcal{L}^{\alpha,|\mathcal{V}_{N}|}(X^{\mathcal{T}})\right):\text{ for some }\alpha\in\mathbb{A}^{*}\right\}\subset\mathcal{P}_{1}(\mathcal{X}^{*})^{|\mathcal{V}_{N}|} (32)
ΘN\displaystyle\Theta_{N} {(α,1(γ(α)),,α,|𝒱N|(γ(α))): for some α𝔸}Θ|𝒱N|\displaystyle\coloneqq\left\{\left(\mathcal{L}^{\alpha,1}(\upgamma(\alpha)),\dots,\mathcal{L}^{\alpha,|\mathcal{V}_{N}|}(\upgamma(\alpha))\right):\text{ for some }\alpha\in\mathbb{A}^{*}\right\}\subset\Theta^{|\mathcal{V}_{N}|} (33)

where for each α𝔸,k1,2,,|𝒱N|\alpha\in\mathbb{A}^{*},k\in 1,2,\dots,|\mathcal{V}_{N}|, α,k\mathcal{L}^{\alpha,k} denotes the conditional law under α\mathbb{P}^{\alpha} given AkA_{k}. Equip each coordinate of N\mathcal{M}_{N} with the Wasserstein metric. Denote by 𝔪=(𝔪1,,𝔪|𝒱N|)\vec{\mathfrak{m}}=(\mathfrak{m}_{1},\dots,\mathfrak{m}_{|\mathcal{V}_{N}|}) an element in N\mathcal{M}_{N}.

Lemma 3.9.

The process t𝔼α[Xt𝒯|Ak]t\mapsto\mathbb{E}^{\mathbb{P}^{\alpha}}[X^{\mathcal{T}}_{t}|A_{k}] is 0\mathbb{Q}_{0}-almost surely continuous for each kk for all α𝔸\alpha\in\mathbb{A}^{*}. Consequentially, the mapping N𝔪τ¯(μN)\mathcal{M}_{N}\ni\vec{\mathfrak{m}}\mapsto\bar{\tau}\left(\upmu^{N}\right) is 0\mathbb{Q}_{0}-almost surely continuous on the closure of N\mathcal{M}_{N}, where for ω0=(β,η)Ω0=𝒳×+\omega_{0}=(\beta,\upeta)\in\Omega_{0}=\mathcal{X}\times\mathbb{R}_{+}, μN((β,η))k=1|𝒱N|𝔪k𝟙{(β,D(η))Ak}\upmu^{N}((\beta,\upeta))\coloneqq\sum_{k=1}^{|\mathcal{V}_{N}|}\mathfrak{m}_{k}\mathds{1}_{\{(\beta,D(\upeta))\in A_{k}\}}.

Proof.

Take a sequence tnntt_{n}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}t_{\infty} in [0,T][0,T]. Note that the event that Xtn𝒯X^{\mathcal{T}}_{t_{n}} does not converge to Xt𝒯X^{\mathcal{T}}_{t} is, up-to a α\mathbb{P}^{\alpha}-null set, contained in {𝒯=t}\{\mathcal{T}=t_{\infty}\}. Then by dominated convergence theorem, it suffices to show that for all tt, α({𝒯=t}|Ak)=0\mathbb{P}^{\alpha}(\{\mathcal{T}=t_{\infty}\}|A_{k})=0 for each AkA_{k}, which is implied by α(𝒯=t)=(𝒯=t)=0\mathbb{P}^{\alpha}(\mathcal{T}=t_{\infty})=\mathbb{P}(\mathcal{T}=t_{\infty})=0. This follows from price dynamics (2) and the fact that the only jump of PP is negative by Assumptions (Assumption (B)). Then continuity of mean processes for each kk 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 N\mathcal{M}_{N} is Wasserstein pre-compact and convex in 𝒫(𝒳)|𝒱N|\mathcal{P}(\mathcal{X}^{*})^{|\mathcal{V}_{N}|}, and ΘN\Theta_{N} is Wasserstein compact and convex in Θ|𝒱N|\Theta^{|\mathcal{V}_{N}|}.

Proof.

The statement on ΘN\Theta_{N} is immediate given that AA is compact and convex. For convexity of N\mathcal{M}_{N}, take α1,α2𝔸\alpha^{1},\alpha^{2}\in\mathbb{A}^{*}. It suffices to show that for all λ[0,1]\lambda\in[0,1], there is α𝔸\alpha\in\mathbb{A}^{*} such that

dαd=λdα1d+(1λ)dα2d.\frac{d\mathbb{P}^{\alpha}}{d\mathbb{P}}=\lambda\frac{d\mathbb{P}^{\alpha^{1}}}{d\mathbb{P}}+(1-\lambda)\frac{d\mathbb{P}^{\alpha^{2}}}{d\mathbb{P}}.

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 k{1,,|𝒱N|}k\in\{1,\dots,|\mathcal{V}_{N}|\} which follows from [62, Lemma 5.9] with the additional fact that 𝔼[ξ|Ak]𝔼[ξ]/(Ak)\mathbb{E}[\xi|A_{k}]\leq\mathbb{E}[\xi]/\mathbb{P}(A_{k}) for arbitrary non-negative random variable ξ\xi, and (Ak)>0\mathbb{P}(A_{k})>0 for all kk. ∎

Suppose we take any 𝔪N=(𝔪1N,,𝔪|𝒱N|N)N\vec{\mathfrak{m}}^{N}=(\mathfrak{m}^{N}_{1},\dots,\mathfrak{m}^{N}_{|\mathcal{V}_{N}|})\in\mathcal{M}_{N} and 𝔞N=(𝔞1N,,𝔞|𝒱N|N)ΘN\vec{\mathfrak{a}}^{N}=(\mathfrak{a}^{N}_{1},\dots,\mathfrak{a}^{N}_{|\mathcal{V}_{N}|})\in\Theta_{N}. Define the inputs to the optimization problem: for each (β,η)=ω0Ω0=𝒳×+(\beta,\upeta)=\omega_{0}\in\Omega_{0}=\mathcal{X}\times\mathbb{R}_{+},

μN(ω0)k=1|𝒱N|𝔪kN𝟙{(β,D(η))Ak}𝒫1(𝒳),ϑN(ω0)k=1|𝒱N|𝔞kN𝟙{(β,D(η))Ak}Θ.\upmu^{N}(\omega_{0})\coloneqq\sum_{k=1}^{|\mathcal{V}_{N}|}\mathfrak{m}^{N}_{k}\mathds{1}_{\{(\beta,D(\upeta))\in A_{k}\}}\in\mathcal{P}_{1}(\mathcal{X}),\quad\upvartheta^{N}(\omega_{0})\coloneqq\sum_{k=1}^{|\mathcal{V}_{N}|}\mathfrak{a}^{N}_{k}\mathds{1}_{\{(\beta,D(\upeta))\in A_{k}\}}\in\Theta. (34)

Again (μN,ϑN)(\upmu^{N},\upvartheta^{N}) can be viewed as a 𝔽B,D\mathbb{F}^{B,D}-measurable random process taking values (μtN,ϑtN)(\upmu_{t}^{N},\upvartheta_{t}^{N}) in 𝒫()×𝒫(𝒫(A))\mathcal{P}(\mathbb{R})\times\mathcal{P}(\mathcal{P}(A)). By Proposition 3.7 we obtain strict, optimal control α^𝒯,N𝔸\hat{\alpha}^{\mathcal{T},N}\in\mathbb{A}^{*} with α^t𝒯,N=a^(Zt𝒯,N)\hat{\alpha}^{\mathcal{T},N}_{t}=\hat{a}(Z^{\mathcal{T},N}_{t}) along with the probability measure Nα^𝒯,N\mathbb{P}^{N}\coloneqq\mathbb{P}^{\hat{\alpha}^{\mathcal{T},N}} on Ωc\Omega_{c}, and WtNWt0tσ1α^s𝒯,N𝑑sW^{N}_{t}\coloneqq W_{t}-\int_{0}^{t}\sigma^{-1}\hat{\alpha}^{\mathcal{T},N}_{s}ds is a Brownian motion under N\mathbb{P}^{N}. Denote by γN\upgamma^{N} the optimal control in the strict relaxed form γ(α^𝒯,N)=(δα^t𝒯,Ndt)t[0,T]\upgamma(\hat{\alpha}^{\mathcal{T},N})=(\delta_{\hat{\alpha}_{t}^{\mathcal{T},N}}dt)_{t\in[0,T]}. Define output conditional measures (𝔪N,out,𝔞N,out)N×ΘN(\vec{\mathfrak{m}}^{N,out},\vec{\mathfrak{a}}^{N,out})\in\mathcal{M}_{N}\times\Theta_{N}:

𝔪kN,out()N(Ak{X𝒯})N(Ak),𝔞kN,out()N(Ak{γN})N(Ak).\mathfrak{m}^{N,out}_{k}(\cdot)\coloneqq\frac{\mathbb{P}^{N}(A_{k}\cap\{X^{\mathcal{T}}\in\cdot\})}{\mathbb{P}^{N}(A_{k})},\quad\mathfrak{a}^{N,out}_{k}(\cdot)\coloneqq\frac{\mathbb{P}^{N}(A_{k}\cap\{\upgamma^{N}\in\cdot\})}{\mathbb{P}^{N}(A_{k})}. (35)

We have now defined our fixed point mapping:

ΦN:N×ΘN(𝔪N,𝔞N)(𝔪N,out,𝔞N,out)N×ΘN.\Phi^{N}:\mathcal{M}_{N}\times\Theta_{N}\ni(\vec{\mathfrak{m}}^{N},\vec{\mathfrak{a}}^{N})\mapsto(\vec{\mathfrak{m}}^{N,out},\vec{\mathfrak{a}}^{N,out})\in\mathcal{M}_{N}\times\Theta_{N}.
Lemma 3.11.

For each NN\in\mathbb{N}, the mapping ΦN\Phi^{N} is continuous.

Proof.

We shall fix NN and drop the notation to avoid confusion with the proof steps for sequential continuity. Recall α^𝒯\hat{\alpha}^{\mathcal{T}} is obtained by solving the optimization problem pp^{*}-by-pp^{*}. Let α^𝒯(p)\mathbb{P}^{\hat{\alpha}^{\mathcal{T}(p^{*})}} denote the conditional probability measure on Ω\Omega from Girsanov transform for a given p[P0,)p^{*}\in[P_{0},\infty). Take a sequence of vectors (𝔪n,𝔞n)n(𝔪,𝔞)N×ΘN(\vec{\mathfrak{m}}^{n},\vec{\mathfrak{a}}^{n})\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}(\vec{\mathfrak{m}}^{\infty},\vec{\mathfrak{a}}^{\infty})\in\mathcal{M}_{N}\times\Theta_{N} in the sense that each coordinate converges in Wasserstein distance, and define correspondingly (μn,ϑn)(\upmu^{n},\upvartheta^{n}) for n{}n\in\mathbb{N}\cup\{\infty\}. Then 0\mathbb{Q}_{0} almost surely, μn\upmu^{n} converges to μ\upmu^{\infty} in \mathcal{M} and ϑn\upvartheta^{n} converges to ϑ\upvartheta^{\infty} in Θ\Theta. For n{}n\in\mathbb{N}\cup\{\infty\} define the discretized conditional probability measures μn\upmu^{n} and ϑn\upvartheta^{n} as in (34), and let Z𝒯(p),nZ^{\mathcal{T}(p^{*}),n} be part of the unique solution to the BSDE (23) with M0M\equiv 0 and input (μn,ϑn)(\upmu^{n},\upvartheta^{n}). Let α^t𝒯(p),n=a^(Zt𝒯(p),n)\hat{\alpha}^{\mathcal{T}(p^{*}),n}_{t}=\hat{a}(Z^{\mathcal{T}(p^{*}),n}_{t}) be the optimal control given 𝒫=p\mathscr{P}=p^{*}.

We first show that α^𝒯(p),nnα^𝒯(p),\hat{\alpha}^{\mathcal{T}(p^{*}),n}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\hat{\alpha}^{\mathcal{T}(p^{*}),\infty} in 𝔾2\mathcal{H}^{2}_{\mathbb{G}} for every p[P0,)p^{*}\in[P_{0},\infty). Recall that optimal controls are continuous in Z𝒯(p)Z^{\mathcal{T}(p^{*})}, so it suffices to show Z𝒯(p),nnZ𝒯(p),Z^{\mathcal{T}(p^{*}),n}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}Z^{\mathcal{T}(p^{*}),\infty} in 𝔾2\mathcal{H}^{2}_{\mathbb{G}}. Using the stability property of BSDE solutions [56, Proposition 54.2], convergence of Z𝒯(p),nZ^{\mathcal{T}(p^{*}),n} is immediately implied if we show

limn𝔼[|Δng|2+𝒯(p)T|Δnhs|2𝑑s]=0,\lim_{n\to\infty}\mathbb{E}\left[\left|\Delta_{n}g\right|^{2}+\int_{\mathcal{T}(p^{*})}^{T}\left|\Delta_{n}h_{s}\right|^{2}ds\right]=0,

where Δngg(XT𝒯(p),Xτ(μn)𝒯(p),τ(μn))g(XT𝒯(p),Xτ(μ)𝒯(p),τ(μ))\Delta_{n}g\coloneqq g\left(X^{\mathcal{T}(p^{*})}_{T},X^{\mathcal{T}(p^{*})}_{\tau^{*}(\upmu^{n})},\tau^{*}(\upmu^{n})\right)-g\left(X^{\mathcal{T}(p^{*})}_{T},X^{\mathcal{T}(p^{*})}_{\tau^{*}(\upmu^{\infty})},\tau^{*}(\upmu^{\infty})\right) and

Δnhsh(s,Xs𝒯(p),bsP,τ(μn),ρ,ϑsn,Zs𝒯(p),)h(s,Xs𝒯(p),bsP,τ(μ),ρ,ϑs,Zs𝒯(p),).\Delta_{n}h_{s}\coloneqq h(s,X^{\mathcal{T}(p^{*})}_{s},b^{P}_{s},\tau^{*}(\upmu^{n}),\left\langle\rho,\upvartheta^{n}_{s}\right\rangle,Z^{\mathcal{T}(p^{*}),\infty}_{s})-h(s,X^{\mathcal{T}(p^{*})}_{s},b^{P}_{s},\tau^{*}(\upmu^{\infty}),\left\langle\rho,\upvartheta^{\infty}_{s}\right\rangle,Z^{\mathcal{T}(p^{*}),\infty}_{s}).

By continuity of gg we know that Δng\Delta_{n}g converges to 0 in probability if both τ(μn)nτ(μ)\tau^{*}(\upmu^{n})\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\tau^{*}(\upmu^{\infty}) and Xτ(μn)nXτ(μ)X_{\tau^{*}(\upmu^{n})}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}X_{\tau^{*}(\upmu^{\infty})} in probability. Lemma 3.9 gives us convergence in burst time. Also, the inventory is continuous everywhere except at entry 𝒯(p)\mathcal{T}(p^{*}). Therefore, Xτ(μn)X_{\tau^{*}(\upmu^{n})} does not converge in probability to Xτ(μ)X_{\tau^{*}(\upmu^{\infty})} only if 𝒯(p)=τ(μ)\mathcal{T}(p^{*})=\tau^{*}(\upmu^{\infty}), which also has probability 0. Therefore,

({|Δng|2n0})1({𝒯(p)=τ(μ)})=1.\mathbb{P}(\{|\Delta_{n}g|^{2}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}0\})\geq 1-\mathbb{P}\left(\{\mathcal{T}(p^{*})=\tau^{*}(\upmu^{\infty})\}\right)=1.

By dominated convergence theorem, we have 𝔼[|Δng|2]n0\mathbb{E}[\left|\Delta_{n}g\right|^{2}]\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}0.

Now let InI_{n} denote the random interval [τ(μn)τ(μ),τ(μn)τ(μ)][\tau^{*}(\upmu^{n})\wedge\tau^{*}(\upmu^{\infty}),\tau^{*}(\upmu^{n})\vee\tau^{*}(\upmu^{\infty})]. Observe that

|𝟙{0t<τ(μn)}𝟙{0t<τ(μ)}|2=|1𝟙{tτ(μn)}1+𝟙{tτ(μ)}|2=𝟙{tIn}.\left|\mathds{1}_{\{0\leq t<\tau^{*}(\upmu^{n})\}}-\mathds{1}_{\{0\leq t<\tau^{*}(\upmu^{\infty})\}}\right|^{2}=\left|1-\mathds{1}_{\{t\geq\tau^{*}(\upmu^{n})\}}-1+\mathds{1}_{\{t\geq\tau^{*}(\upmu^{\infty})\}}\right|^{2}=\mathds{1}_{\{t\in I_{n}\}}.

Define InpIn[𝒯(p),T]I^{p^{*}}_{n}\coloneqq I_{n}\cap[\mathcal{T}(p^{*}),T]. Remark (C1) implies

𝔼[𝒯(p)T|Δnhs|2𝑑s]\displaystyle\mathbb{E}\left[\int_{\mathcal{T}(p^{*})}^{T}\left|\Delta_{n}h_{s}\right|^{2}ds\right] 𝔼[𝒯(p)T|fb(s,Xs𝒯(p),bsP)𝟙{0t<τ(μn)}fb(s,Xs𝒯(p),bsP)𝟙{0t<τ(μ)}|2𝑑s]\displaystyle\leq\mathbb{E}\left[\int_{\mathcal{T}(p^{*})}^{T}\left|f_{b}(s,X^{\mathcal{T}(p^{*})}_{s},b^{P}_{s})\mathds{1}_{\{0\leq t<\tau^{*}(\upmu^{n})\}}-f_{b}(s,X^{\mathcal{T}(p^{*})}_{s},b^{P}_{s})\mathds{1}_{\{0\leq t<\tau^{*}(\upmu^{\infty})\}}\right|^{2}ds\right]
+𝔼[𝒯(p)T|fc(s,Xs𝒯(p),ρ,ϑsn)𝟙{sτ(μn)}fc(s,Xs𝒯(p),ρ,ϑs)𝟙{sτ(μ)}|2𝑑s]\displaystyle+\mathbb{E}\left[\int_{\mathcal{T}(p^{*})}^{T}\left|f_{c}(s,X^{\mathcal{T}(p^{*})}_{s},\left\langle\rho,\upvartheta^{n}_{s}\right\rangle)\mathds{1}_{\{s\geq\tau^{*}(\upmu^{n})\}}-f_{c}(s,X^{\mathcal{T}(p^{*})}_{s},\left\langle\rho,\upvartheta^{\infty}_{s}\right\rangle)\mathds{1}_{\{s\geq\tau^{*}(\upmu^{\infty})\}}\right|^{2}ds\right]
𝔼[Inp|fb(s,Xs𝒯(p),bsP)|2𝑑s]+2𝔼[Inp|fc(s,Xs𝒯(p),ρ,ϑs)|2𝑑s]\displaystyle\leq\mathbb{E}\left[\int_{I_{n}^{p^{*}}}|f_{b}(s,X^{\mathcal{T}(p^{*})}_{s},b^{P}_{s})|^{2}ds\right]+2\mathbb{E}\left[\int_{I_{n}^{p^{*}}}|f_{c}(s,X^{\mathcal{T}(p^{*})}_{s},\left\langle\rho,\upvartheta^{\infty}_{s}\right\rangle)|^{2}ds\right]
+2𝔼[𝒯(p)T|fc(s,Xs𝒯(p),ρ,ϑsn)fc(s,Xs𝒯(p),ρ,ϑs)|2𝟙{sτ(μn)}𝑑s].\displaystyle+2\mathbb{E}\left[\int_{\mathcal{T}(p^{*})}^{T}\left|f_{c}(s,X^{\mathcal{T}(p^{*})}_{s},\left\langle\rho,\upvartheta^{n}_{s}\right\rangle)-f_{c}(s,X^{\mathcal{T}(p^{*})}_{s},\left\langle\rho,\upvartheta^{\infty}_{s}\right\rangle)\right|^{2}\mathds{1}_{\{s\geq\tau^{*}(\upmu^{n})\}}ds\right].

Lemma 3.9 implies 𝟙{tInp}n𝟙{t=τ(μ)}\mathds{1}_{\{t\in I_{n}^{p^{*}}\}}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\mathds{1}_{\{t=\tau^{*}(\upmu^{\infty})\}} almost surely. Under Assumption (E1) and Remark (C1), all terms converge to 0 by dominated convergence theorem and Fubini’s theorem. Therefore, for all p[P0,)p^{*}\in[P_{0},\infty), α^𝒯(p),nnα^𝒯(p),\hat{\alpha}^{\mathcal{T}(p^{*}),n}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\hat{\alpha}^{\mathcal{T}(p^{*}),\infty} in 𝔾2\mathcal{H}^{2}_{\mathbb{G}}. Since AA is assumed to be bounded, this implies that α^nnα^\hat{\alpha}^{n}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\hat{\alpha}^{\infty} in the 𝔾,c2\mathcal{H}^{2}_{\mathbb{G},c} sense on the product space Ωc\Omega_{c}.

For each n{}n\in\mathbb{N}\cup\{\infty\}, denote by nα^n\mathbb{P}^{n}\coloneqq\mathbb{P}^{\hat{\alpha}^{n}} the probability measure on Ωc\Omega_{c} from Girsanov transformation. By construction, n\mathbb{P}^{n}\ll\mathbb{P} for all nn and

dnd=(0σ1(α^tnα^t)𝑑Wt)T.\frac{d\mathbb{P}^{n}}{d\mathbb{P}^{\infty}}=\mathcal{E}\left(\int_{0}^{\cdot}\sigma^{-1}\left(\hat{\alpha}^{n}_{t}-\hat{\alpha}^{\infty}_{t}\right)dW_{t}\right)_{T}.

Therefore, by boundedness of AA we can calculate the relative entropy

(|n)=𝔼[logdnd]=12𝔼[0Tσ2|α^tnα^t|2𝑑t]n0,\mathcal{H}(\mathbb{P}^{\infty}|\mathbb{P}^{n})=-\mathbb{E}^{\mathbb{P}^{\infty}}\left[\log\frac{d\mathbb{P}^{n}}{d\mathbb{P}^{\infty}}\right]=\frac{1}{2}\mathbb{E}^{\mathbb{P}^{\infty}}\left[\int_{0}^{T}\sigma^{-2}\left|\hat{\alpha}^{n}_{t}-\hat{\alpha}^{\infty}_{t}\right|^{2}dt\right]\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}0,

Pinsker’s inequality implies that n\mathbb{P}^{n} converges to ,p\mathbb{P}^{\infty,p^{*}} in total variation. By triangular inequality and the convergence in 𝔾,c2\mathcal{H}^{2}_{\mathbb{G},c} of controls, we have

n(α^tn)1n(α^t)1 in dt-measure.\mathbb{P}^{n}\circ\left(\hat{\alpha}^{n}_{t}\right)^{-1}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\mathbb{P}^{\infty}\circ\left(\hat{\alpha}_{t}^{\infty}\right)^{-1}\text{ in }dt\text{-measure}.

Bounded convergence theorem yields n(γ(α^n))1n(γ(α^))1\mathbb{P}^{n}\circ(\upgamma(\hat{\alpha}^{n}))^{-1}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\mathbb{P}^{\infty}\circ(\upgamma(\hat{\alpha}^{\infty}))^{-1} in the stable topology. Since (Ak)>0\mathbb{P}(A_{k})>0 for each kk, this implies convergence of 𝔞kn,outn𝔞k,out\mathfrak{a}^{n,out}_{k}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\mathfrak{a}^{\infty,out}_{k} as well. Boundedness of AA ensures Wasserstein convergence as well. nn\mathbb{P}^{n}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\mathbb{P}^{\infty} in total variation also implies weak convergence of 𝔪kn,outn𝔪k,out\mathfrak{m}^{n,out}_{k}\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\mathfrak{m}^{\infty,out}_{k} for each kk. To show Wasserstein convergence, it suffices to show uniform integrability (see e.g. [63, Theorem 6.9]).

limRlim supnsupk{1,,|𝒱N|}{𝒙𝒳>R}𝒙𝒳𝔪kn,out(d𝒙)\displaystyle\lim_{R\to\infty}\limsup_{n\to\infty}\sup_{k\in\{1,\dots,|\mathcal{V}_{N}|\}}\int_{\{\lVert\boldsymbol{x}\rVert_{\mathcal{X}^{*}}>R\}}\lVert\boldsymbol{x}\rVert_{\mathcal{X}^{*}}\mathfrak{m}^{n,out}_{k}(d\boldsymbol{x})
limRsupα𝔸supk{1,,|𝒱N|}1(Ak)𝔼α[X𝒯𝒳𝟙{X𝒯𝒳>R}]=0\displaystyle\leq\lim_{R\to\infty}\sup_{\alpha\in\mathbb{A}^{*}}\sup_{k\in\{1,\dots,|\mathcal{V}_{N}|\}}\frac{1}{\mathbb{P}(A_{k})}\mathbb{E}^{\mathbb{P}^{\alpha}}\left[\lVert X^{\mathcal{T}}\rVert_{\mathcal{X}^{*}}\mathds{1}_{\{\lVert X^{\mathcal{T}}\rVert_{\mathcal{X}^{*}}>R\}}\right]=0

which follows from (28). Therefore, since |𝒱N||\mathcal{V}_{N}| is finite:

ΦN(𝔪n,𝔞n)nΦN(𝔪,𝔞) in N×ΘN,\Phi^{N}(\vec{\mathfrak{m}}^{n},\vec{\mathfrak{a}}^{n})\stackrel{{\scriptstyle n\to\infty}}{{\longrightarrow}}\Phi^{N}(\vec{\mathfrak{m}}^{\infty},\vec{\mathfrak{a}}^{\infty})\text{ in }\mathcal{M}_{N}\times\Theta_{N},

and continuity of ΦN\Phi^{N} holds for all NN. ∎

Proposition 3.12.

The mapping ΦN\Phi^{N} admits a fixed point (𝔪N,𝔞N)N×ΘN(\vec{\mathfrak{m}}^{N},\vec{\mathfrak{a}}^{N})\in\mathcal{M}_{N}\times\Theta_{N} for all NN\in\mathbb{N}.

Proof.

Let ¯N\overline{\mathcal{M}}_{N} denote the closure of N\mathcal{M}_{N}, which by Lemma 3.10 is convex and compact. Note that the input (μN,ϑN)(\upmu^{N},\upvartheta^{N}) to the BSDE is still well-defined for 𝔪N¯N\vec{\mathfrak{m}}^{N}\in\overline{\mathcal{M}}_{N}, except that ϑN\upvartheta^{N} might not be the law of a strict control anymore. Therefore, we can define ΦN\Phi^{N} on the larger domain ¯N×ΘN\overline{\mathcal{M}}_{N}\times\Theta_{N}. Moreover, the Wasserstein closure preserves the continuity of the mean process. This implies that τ()\tau^{*}(\cdot) is still continuous on ¯N\overline{\mathcal{M}}_{N} (see the proof of [62, Theorem 6.1]), so continuity of ΦN\Phi^{N} still holds. Applying Brouwer’s fixed point theorem (e.g. [4, Corollary 17.56]) yields a fixed point (𝔪N,𝔞N)(\vec{\mathfrak{m}}^{N},\vec{\mathfrak{a}}^{N}) of ΦN\Phi^{N}. However, the range of ΦN\Phi^{N} is still strictly in N×ΘN\mathcal{M}_{N}\times\Theta_{N} since allowing μ\upmu to take values in ¯N\overline{\mathcal{M}}_{N} doesn’t affect how we construct the optimal control and its corresponding state process. Then (𝔪N,𝔞N)N×ΘN(\vec{\mathfrak{m}}^{N},\vec{\mathfrak{a}}^{N})\in\mathcal{M}_{N}\times\Theta_{N} is also a fixed point of Φ\Phi had we not enlarged the domain. ∎

For each NN, let αN𝔸\alpha^{N}\in\mathbb{A}^{*} be the equilibrium strategy from a fixed point (𝔪N,𝔞N)(\vec{\mathfrak{m}}^{N},\vec{\mathfrak{a}}^{N}) from Proposition 3.12 and N=αN\mathbb{P}^{N}=\mathbb{P}^{\alpha^{N}}. Define the corresponding random laws (μN,ϑN)(\upmu^{N},\upvartheta^{N}) as in (34). Then by construction, αN\alpha^{N} minimizes JweakμN,ϑNJ^{\upmu^{N},\upvartheta^{N}}_{weak} defined in (20) over 𝔸\mathbb{A}^{*}.

Remark 3.13.

For each NN\in\mathbb{N}, we work with the same filtration 𝔽,W,B,D\mathbb{F}^{\mathcal{I},W,B,D}. In fact, by the argument in [62, Proposition A.10], α^N\hat{\alpha}^{N} also minimizes (20) had we allowed for a larger filtration 𝔽~𝔽,W,B,D\widetilde{\mathbb{F}}\supseteq\mathbb{F}^{\mathcal{I},W,B,D}. We provide a proof to keep the paper self-contained.

Proof.

We fix NN\in\mathbb{N} and recall that the uncontrolled state process is defined by

Xt𝒯K0/𝒫+σ(WtW𝒯) for t𝒯andXt𝒯0 for t[0,𝒯).X^{\mathcal{T}}_{t}\coloneqq K_{0}/\mathscr{P}+\sigma(W_{t}-W_{\mathcal{T}})\text{ for }t\geq\mathcal{T}\quad\text{and}\quad X^{\mathcal{T}}_{t}\coloneqq 0\text{ for }t\in[0,\mathcal{T}).

Recall from Remark 3.6 that a^()\hat{a}(\cdot) is a continuous function that minimizes the Hamiltonian. For each β𝔸\beta\in\mathbb{A}^{*} defined on the probability space (Ω,,,𝔽~)(\Omega,\mathcal{F},\mathbb{P},\widetilde{\mathbb{F}}), by [56, Theorem 53.1] there exists a unique solution (Yβ,Zβ,β,Uβ,Nβ)𝒮𝔽~2×𝒢𝔽~2×𝔽~2×𝔽~,D2×𝒮𝔽~2(Y^{\beta},Z^{\beta},\mathfrak{Z}^{\beta},U^{\beta},N^{\beta})\in\mathcal{S}^{2}_{\widetilde{\mathbb{F}}}\times\mathcal{G}^{2}_{\widetilde{\mathbb{F}}}\times\mathcal{H}^{2}_{\widetilde{\mathbb{F}}}\times\mathcal{H}^{2}_{\widetilde{\mathbb{F}},D}\times\mathcal{S}^{2}_{\widetilde{\mathbb{F}}} to the following BSDE

Yt=g(XT𝒯,Xτ(μN)𝒯,τ(μN))+tTH(s,Xs𝒯,bsP,τ(μN),ρ,ϑsN,Zs,βs)𝑑stTZs𝑑WstTs𝑑BstTUs𝑑MsτtT𝑑Ns,t[0,T].\begin{split}Y_{t}&=g(X^{\mathcal{T}}_{T},X^{\mathcal{T}}_{\tau^{*}(\upmu^{N})},\tau^{*}(\upmu^{N}))+\int_{t}^{T}H(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu^{N}),\left\langle\rho,\upvartheta^{N}_{s}\right\rangle,Z_{s},\beta_{s})ds\\ &-\int_{t}^{T}Z_{s}dW_{s}-\int_{t}^{T}\mathfrak{Z}_{s}dB_{s}-\int_{t}^{T}U_{s}dM^{\tau}_{s}-\int_{t}^{T}dN_{s},\quad t\in[0,T].\end{split} (36)

Remark (S2) also implies well-posedness for the following BSDE solved on (Ω,,,𝔽~)(\Omega,\mathcal{F},\mathbb{P},\widetilde{\mathbb{F}}):

Yt=g(XT𝒯,Xτ(μN)𝒯,τ(μN))+tT𝟙{s𝒯}h(s,Xs𝒯,bsP,τ(μN),ρ,ϑsN,Zs)𝑑stTZs𝑑WstTs𝑑BstTUs𝑑MsτtT𝑑Ns,t[0,T].\begin{split}Y_{t}&=g(X^{\mathcal{T}}_{T},X^{\mathcal{T}}_{\tau^{*}(\upmu^{N})},\tau^{*}(\upmu^{N}))+\int_{t}^{T}\mathds{1}_{\{s\geq\mathcal{T}\}}h(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu^{N}),\left\langle\rho,\upvartheta^{N}_{s}\right\rangle,Z_{s})ds\\ &-\int_{t}^{T}Z_{s}dW_{s}-\int_{t}^{T}\mathfrak{Z}_{s}dB_{s}-\int_{t}^{T}U_{s}dM^{\tau}_{s}-\int_{t}^{T}dN_{s},\quad t\in[0,T].\end{split} (37)

whose unique solution we denote by (Y^,Z^,^,U^,N^)(\hat{Y},\hat{Z},\hat{\mathfrak{Z}},\hat{U},\hat{N}). However, since (μN,ϑN)(\mu^{N},\upvartheta^{N}) are 𝔽B,D\mathbb{F}^{B,D} measurable, uniqueness of the solution implies that N^\hat{N} is Ndt\mathbb{P}^{N}\otimes dt almost surely zero, and (Y^,Z^,^,U^)(\hat{Y},\hat{Z},\hat{\mathfrak{Z}},\hat{U}) coincides with the solution of (26) on [𝒯,T][\mathcal{T},T]. Recall from Corollary 3.8 and the construction of the fixed point mapping ΦN\Phi^{N} that Ndt\mathbb{P}^{N}\otimes dt almost surely, we must also have αtN=a^(Z^t)𝟙{t𝒯}\alpha^{N}_{t}=\hat{a}(\hat{Z}_{t})\mathds{1}_{\{t\geq\mathcal{T}\}}.

Recall from Remark (C4) that for any admissible control, the Hamiltonian is 0 before entry. Then the generator of (36) is dt\mathbb{P}\otimes dt-almost surely greater than the generator of (37), and they are equal when we take β=αN\beta=\alpha^{N}. 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 t[0,T]t\in[0,T] we denote

ΔHt(β,αN)H(t,Xt𝒯,btP,τ(μN),ρ,ϑtN,Ztβ,βt)H(t,Xt𝒯,btP,τ(μN),ρ,ϑtN,ZtαN,αtN).\Delta H_{t}(\beta,\alpha^{N})\coloneqq H(t,X^{\mathcal{T}}_{t},b^{P}_{t},\tau^{*}(\upmu^{N}),\left\langle\rho,\upvartheta^{N}_{t}\right\rangle,Z^{\beta}_{t},\beta_{t})-H(t,X^{\mathcal{T}}_{t},b^{P}_{t},\tau^{*}(\upmu^{N}),\left\langle\rho,\upvartheta^{N}_{t}\right\rangle,Z^{\alpha^{N}}_{t},\alpha^{N}_{t}).

Due to the presence of compensated martingale MτM^{\tau} and orthogonal martingale NN, additional conditions are required for comparison principle to hold. In light of [23], a sufficient condition is the existence of an equivalent measure ~\widetilde{\mathbb{P}} to \mathbb{P} such that

SttTΔHs(β,αN)𝑑s+tT(ZsβZsαN)𝑑Ws+tT(sβsαN)𝑑Bs+tT(UsβUsαN)𝑑Msτ+tT𝑑NsβtT𝑑NsαN,t[0,T]\begin{split}S_{t}&\coloneqq-\int_{t}^{T}\Delta H_{s}(\beta,\alpha^{N})ds+\int_{t}^{T}(Z^{\beta}_{s}-Z^{\alpha^{N}}_{s})dW_{s}+\int_{t}^{T}(\mathfrak{Z}^{\beta}_{s}-\mathfrak{Z}^{\alpha^{N}}_{s})dB_{s}\\ &\qquad+\int_{t}^{T}(U^{\beta}_{s}-U^{\alpha^{N}}_{s})dM^{\tau}_{s}+\int_{t}^{T}dN^{\beta}_{s}-\int_{t}^{T}dN^{\alpha^{N}}_{s},\quad t\in[0,T]\end{split} (38)

is a martingale under ~\widetilde{\mathbb{P}}. This probability can be constructed by defining

d~d(0ΔHt(β,αN)ZtβZtαN𝟙{ZtβZtαN0}𝑑Wt)T.\frac{d\widetilde{\mathbb{P}}}{d\mathbb{P}}\coloneqq\mathcal{E}\left(\int_{0}^{\cdot}\frac{\Delta H_{t}(\beta,\alpha^{N})}{Z^{\beta}_{t}-Z^{\alpha^{N}}_{t}}\mathds{1}_{\{Z^{\beta}_{t}-Z^{\alpha^{N}}_{t}\neq 0\}}dW_{t}\right)_{T}.

Since HH is Lipschitz in zz, we can apply Girsanov’s theorem and conclude by [23, Theorem 2] that comparison principle holds. Denote again by β\mathbb{P}^{\beta} the probability measure corresponding to β\beta, which is defined as in (19). Recall that N,β\mathbb{P}^{N},\mathbb{P}^{\beta} and \mathbb{P} all agree at time t=0t=0 and the Hamiltonian is 0 for any admissible strategy before 𝒯\mathcal{T}. Using comparison principle and applying optional stopping theorem yield

JweakμN,ϑN(αN)=𝔼N[Y𝒯αN]=𝔼N[Y0αN]=𝔼[Y0αN]𝔼β[Y0β]=𝔼β[Y𝒯β]=JweakμN,ϑN(β).J^{\upmu^{N},\upvartheta^{N}}_{weak}(\alpha^{N})=\mathbb{E}^{\mathbb{P}^{N}}\left[Y^{\alpha^{N}}_{\mathcal{T}}\right]=\mathbb{E}^{\mathbb{P}^{N}}\left[Y^{\alpha^{N}}_{0}\right]=\mathbb{E}^{\mathbb{P}}\left[Y^{\alpha^{N}}_{0}\right]\leq\mathbb{E}^{\mathbb{P}^{\beta}}\left[Y^{\beta}_{0}\right]=\mathbb{E}^{\mathbb{P}^{\beta}}\left[Y^{\beta}_{\mathcal{T}}\right]=J^{\upmu^{N},\upvartheta^{N}}_{weak}(\beta).

The inequality holds for all 𝔽\mathbb{F}-admissible β\beta. ∎

Remark 3.14.

Since the state variable is linearly controlled, for each γIΓ\upgamma\in\mathrm{I}\Gamma we can define an AA-valued control αt=Aaγt(da)\alpha_{t}=\int_{A}a\upgamma_{t}(da) in 𝔸\mathbb{A}^{*} that corresponds to the same state process. Moreover, Remark (C3) and Jensen’s inequality imply that this control α\alpha is at least as good as γ\upgamma 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 (μ,ϑ)(\upmu,\upvartheta), the optimal AA-valued control is also optimal over relaxed controls for the relaxed objective (17).

3.3.3. Solutions as Weak Limit Points

Before taking NN\to\infty, we shall “lift” the marginal laws of 𝒳𝒯\mathcal{X}^{\mathcal{T}} and γ\upgamma to a joint law in order to carry more information along the way. Let γN=γ(αN)\upgamma^{N}=\upgamma(\alpha^{N}) be the strict control corresponding to the equilibrium strategy αN\alpha^{N} from the fixed point mapping. Recall that =(K0,𝒫)\mathcal{I}=(K_{0},\mathscr{P}), taking values in ×[p,)\mathbb{R}\times[p^{*},\infty) denotes the initial wealth and entry threshold. Define on Ω0\Omega_{0} the lifted random law 𝔐N𝒫(𝒳×𝒳×Γ×2)\mathfrak{M}^{N}\in\mathcal{P}(\mathcal{X}^{*}\times\mathcal{X}\times\Gamma\times\mathbb{R}^{2}) by

𝔐N(ω0)k=1|𝒱N|𝟙{(β,η)Ak}αN,k((X𝒯,WN,γN,)),ω0=(β,η)Ω0.\mathfrak{M}^{N}(\omega_{0})\coloneqq\sum_{k=1}^{|\mathcal{V}_{N}|}\mathds{1}_{\{(\beta,\upeta)\in A_{k}\}}\mathcal{L}^{\alpha^{N},k}\left((X^{\mathcal{T}},W^{N},\upgamma^{N},\mathcal{I})\right),\quad\omega_{0}=(\beta,\upeta)\in\Omega_{0}. (39)

Recall the notation from Definition 3.3 that 𝔐N,x\mathfrak{M}^{N,x} and 𝔐N,γ\mathfrak{M}^{N,\upgamma} denote the first and third marginals of 𝔐N\mathfrak{M}^{N}, which are the conditional law of (X𝒯,γN)(X^{\mathcal{T}},\upgamma^{N}) under (VN,DN)(V^{N},D^{N}).

Lemma 3.15.

The sequence N(B,τ,𝔐N,X𝒯,WN,γN,)1\mathbb{P}^{N}\circ(B,\tau,\mathfrak{M}^{N},X^{\mathcal{T}},W^{N},\upgamma^{N},\mathcal{I})^{-1} is tight.

Proof.

Assumption (E1) implies that

𝔼N[supt[0,T]|Xt𝒯|]C\mathbb{E}^{\mathbb{P}^{N}}\left[\sup_{t\in[0,T]}|X^{\mathcal{T}}_{t}|\right]\leq C

for some C>0C>0 that does not depend on NN. Let ss be a 𝔾\mathbb{G} stopping time and δ>0\delta>0. Then

𝔼N[|Xs+δ𝒯Xs𝒯|]𝔼N[ss+δ|α^tN|𝑑t+σ|Wρ+δNWρN|]Cδ1/2\mathbb{E}^{\mathbb{P}^{N}}\left[|X^{\mathcal{T}}_{s+\delta}-X^{\mathcal{T}}_{s}|\right]\leq\mathbb{E}^{\mathbb{P}^{N}}\left[\int_{s}^{s+\delta}|\hat{\alpha}^{N}_{t}|dt+\sigma|W^{N}_{\rho+\delta}-W^{N}_{\rho}|\right]\leq C\delta^{1/2}

for a possibly different CC. These two conditions are sufficient for Aldous’ criterion for tightness of N(X𝒯)1\mathbb{P}^{N}\circ(X^{\mathcal{T}})^{-1}. For each NN, (WN,B)(W^{N},B) remain independent Brownian motions under N\mathbb{P}^{N}, and (τ,)(\tau,\mathcal{I}) also retain the same law throughout by independence. Compactness of AA implies the tightness of N(γN)1\mathbb{P}^{N}\circ(\upgamma^{N})^{-1}.

Now tightness of N(𝔐N)1\mathbb{P}^{N}\circ(\mathfrak{M}^{N})^{-1} in the weak topology follows from the tightness of N(X𝒯,WN,γN,)1\mathbb{P}^{N}\circ(X^{\mathcal{T}},W^{N},\upgamma^{N},\mathcal{I})^{-1} (see the proof of [18, Lemma 3.16]). As suggested by Lemma 3.9, we will need to equip 𝒫(𝒳×𝒳×Γ×2)\mathcal{P}(\mathcal{X}^{*}\times\mathcal{X}\times\Gamma\times\mathbb{R}^{2}) with the Wasserstein metric in order to guarantee continuity of τ¯\bar{\tau}, where the product space is equipped with the l1l_{1} metric. By [17, Corollary 5.6] and boundedness of AA, the proof of [18, Lemma 3.16] implies that it suffices to show uniform square-integrability of X𝒯𝒳+WN\lVert X^{\mathcal{T}}\rVert_{\mathcal{X}^{*}}+\lVert W^{N}\rVert_{\infty}. Since WNW^{N} is a Brownian motion under N\mathbb{P}^{N}, we only need to show that

limRsupN𝔼N[X𝒯𝒳𝟙{X𝒯𝒳R}]<,\lim_{R\to\infty}\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\left[\lVert X^{\mathcal{T}}\rVert_{\mathcal{X}^{*}}\mathds{1}_{\{\lVert X^{\mathcal{T}}\rVert_{\mathcal{X}^{*}}\geq R\}}\right]<\infty,

which is implied by (28). Finally, tightness of the marginals implies that of the joint law.

Remark 3.16.

Let ¯(B,τ,𝔐,X𝒯,W,γ,)1\bar{\mathbb{P}}\coloneqq\mathbb{P}^{\infty}\circ(B,\tau,\mathfrak{M}^{\infty},X^{\mathcal{T}},W^{\infty},\upgamma^{\infty},\mathcal{I})^{-1} 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 Ω¯0+×𝒳×𝒫1(𝒳×𝒳×Γ×2),Ω¯1𝒳×𝒳×Γ×2\bar{\Omega}_{0}\coloneqq\mathbb{R}_{+}\times\mathcal{X}\times\mathcal{P}_{1}(\mathcal{X}^{*}\times\mathcal{X}\times\Gamma\times\mathbb{R}^{2}),\ \bar{\Omega}_{1}\coloneqq\mathcal{X}^{*}\times\mathcal{X}\times\Gamma\times\mathbb{R}^{2}. Then let (τ,B,𝔐)(\tau,B,\mathfrak{M}^{\infty}), (X,W,γ,)(X,W^{\infty},\upgamma^{\infty},\mathcal{I}) be the respective canonical processes on Ω¯Ω¯0×Ω¯1\bar{\Omega}\coloneqq\bar{\Omega}_{0}\times\bar{\Omega}_{1}, and ¯\bar{\mathbb{P}} a probability measure on Ω¯\bar{\Omega}. For each NN, define also on Ω¯\bar{\Omega} the law ¯NN(B,τ,𝔐N,X𝒯,WN,γN,)1\bar{\mathbb{P}}^{N}\coloneqq\mathbb{P}^{N}\circ(B,\tau,\mathfrak{M}^{N},X^{\mathcal{T}},W^{N},\upgamma^{N},\mathcal{I})^{-1}. We can obviously drop the \infty from the notation (or even NN, since we can always work on the canonical space), but we keep it to avoid confusion.

Define the jump process DD, price process PP and entry time 𝒯\mathcal{T}, thanks to the strong solvability from Proposition A.1. Take 𝔾\mathbb{G} to be the ¯\bar{\mathbb{P}}-completed natural filtration generated by (B,D,𝔐,X𝒯,W,γ,)(B,D,\mathfrak{M}^{\infty},X^{\mathcal{T}},W^{\infty},\upgamma^{\infty},\mathcal{I}), which is again the progressively enlarged version of 𝔽B,𝔐,X,W,γ,\mathbb{F}^{B,\mathfrak{M}^{\infty},X,W^{\infty},\upgamma^{\infty},\mathcal{I}} by Dt=𝟙{τt}D_{t}=\mathds{1}_{\{\tau\leq t\}}, so τ\tau is a 𝔾\mathbb{G}-inaccessible stopping time. Let =𝒢Tσ(τ)\mathcal{F}=\mathcal{G}_{T}\vee\sigma(\tau). Weak convergence of γN\upgamma^{N} to γ\upgamma^{\infty} implies that X𝒯X^{\mathcal{T}} with entry time 𝒯\mathcal{T} satisfies the relaxed state SDE (16) on (Ω¯,,¯,𝔾)(\bar{\Omega},\mathcal{F},\bar{\mathbb{P}},\mathbb{G}).

Note that the existence of fixed points of the discretized problem and tightness result both hold for arbitrary choice of discretized process VNV^{N}. Now we need to specify the time and space grids to ensure VNV^{N} approximates BB well enough. We will prove Theorem 3.4 by verifying that 𝔐\mathfrak{M}^{\infty} defined on (Ω¯,,¯,𝔾)(\bar{\Omega},\mathcal{F},\bar{\mathbb{P}},\mathbb{G}) 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 NN, 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 NN\in\mathbb{N}, let ti=iT2Nt_{i}=\frac{iT}{2^{N}} for i=0,,2Ni=0,\dots,2^{N}. Set v0=0v_{0}=0 and vi=vi1+Π(N)(BtiBti1)v_{i}=v_{i-1}+\Pi^{(N)}(B_{t_{i}}-B_{t_{i-1}}), where the projection map Π(N):\Pi^{(N)}:\mathbb{R}\to\mathbb{R} is defined as

Π(N)(x)={4N4Nx|x|4N4Nsign(x)|x|>4N.\Pi^{(N)}(x)=\begin{cases}4^{-N}\lfloor 4^{N}x\rfloor&|x|\leq 4^{N}\\ 4^{N}\operatorname*{sign}(x)&|x|>4^{N}\end{cases}.

Then on the event EN{B4N1}E_{N}\coloneqq\{\lVert B\rVert_{\infty}\leq 4^{N}-1\} the process VNV^{N} satisfies

|VtiNBti|12N,N\left|V^{N}_{t_{i}}-B_{t_{i}}\right|\leq\frac{1}{2^{N}},\quad\forall N\in\mathbb{N}

and so we always have

VNB12N+sups,t[0,T]:|st|2N|BsBt|.\lVert V^{N}-B\rVert_{\infty}\leq\frac{1}{2^{N}}+\sup_{s,t\in[0,T]:\ |s-t|\leq 2^{-N}}|B_{s}-B_{t}|.

The right hand side converges to 0 in \mathbb{P} probability. Since BB remains a Brownian motion under each N\mathbb{P}^{N}, N(EN)\mathbb{P}^{N}(E_{N}) converges to 11. Therefore, for all ε>0\varepsilon>0 we have

limNN(VNBε)=1.\lim_{N\to\infty}\mathbb{P}^{N}\left(\lVert V^{N}-B\rVert_{\infty}\leq\varepsilon\right)=1. (40)

With this choice of VNV^{N}, we have the following consistency property in the limit.

Lemma 3.17.

For all bounded, uniformly continuous functions h0:𝒳×𝒫1(Ω¯1)×𝒳Dh^{0}:\mathcal{X}\times\mathcal{P}_{1}(\bar{\Omega}_{1})\times\mathcal{X}_{D}\to\mathbb{R} and h1:𝒳×𝒳×Γ×2h^{1}:\mathcal{X}^{*}\times\mathcal{X}\times\Gamma\times\mathbb{R}^{2}\to\mathbb{R}, we have

𝔼¯[h0(B,𝔐,D)h1(X𝒯,W,γ,)]=𝔼¯[h0(B,𝔐,D)Ω¯1h1(x,w,q,ι)𝑑𝔐(x,w,q,ι)].\bar{\mathbb{E}}\left[h^{0}(B,\mathfrak{M}^{\infty},D)h^{1}(X^{\mathcal{T}},W^{\infty},\upgamma^{\infty},\mathcal{I})\right]=\bar{\mathbb{E}}\left[h^{0}(B,\mathfrak{M}^{\infty},D)\int_{\bar{\Omega}_{1}}h^{1}(x,w,q,\iota)d\mathfrak{M}^{\infty}(x,w,q,\iota)\right].
Proof.

This equality holds at the discretization level by (39), namely

𝔼N[h0(VN,𝔐N,DN)h1(X𝒯,\displaystyle\mathbb{E}^{N}[h^{0}(V^{N},\mathfrak{M}^{N},D^{N})h^{1}(X^{\mathcal{T}}, WN,γN,)]\displaystyle W^{N},\upgamma^{N},\mathcal{I})]
=𝔼N[h0(VN,𝔐N,DN)Ω¯1h1(x,w,q,ι)𝑑𝔐N(x,w,q,ι)].\displaystyle=\mathbb{E}^{N}\left[h^{0}(V^{N},\mathfrak{M}^{N},D^{N})\int_{\bar{\Omega}_{1}}h^{1}(x,w,q,\iota)d\mathfrak{M}^{N}(x,w,q,\iota)\right].

By (40) and (31), uniform continuity of the function h0h^{0} allows us to swap (VN,DN)(V^{N},D^{N}) with (B,D)(B,D) on both sides of the equality above without changing the limits, if they exist. Boundedness of h0,h1h_{0},h_{1} and weak convergence enables us to take NN\to\infty and retain equality in the limit. ∎

Lemma 3.17 says that 𝔐\mathfrak{M}^{\infty} is a version of conditional distribution of (X𝒯,W,γ,)(X^{\mathcal{T}},W^{\infty},\upgamma^{\infty},\mathcal{I}) given (B,𝔐,D)(B,\mathfrak{M}^{\infty},D). which implies the consistency of marginals 𝔐,x\mathfrak{M}^{\infty,x} and 𝔐,γ\mathfrak{M}^{\infty,\upgamma}. We want to carry the conditional joint laws because of the compatibility condition.

3.3.5. Optimality

For each N{}N\in\mathbb{N}\cup\{\infty\}, define 𝔓N(μN,ϑN)=(𝔐N,x,𝔐N,γ)\mathfrak{P}^{N}\coloneqq(\upmu^{N},\upvartheta^{N})=(\mathfrak{M}^{N,x},\mathfrak{M}^{N,\upgamma}) to be the first and third marginals of 𝔐N\mathfrak{M}^{N}. Define the state process corresponding to relaxed control any γIΓ\upgamma\in\mathrm{I}\Gamma as in (16) but in the environment N\mathbb{P}^{N} using WNW^{N}, namely

XN,γ𝟙{t𝒯}K0/𝒫+0tAaγ(ds,da)+σ(Wt𝒯NW𝒯N),t[0,T].X^{N,\upgamma}\coloneqq\mathds{1}_{\{t\geq\mathcal{T}\}}K_{0}/\mathscr{P}+\int_{0}^{t}\int_{A}a\upgamma(ds,da)+\sigma(W^{N}_{t\vee\mathcal{T}}-W^{N}_{\mathcal{T}}),\quad t\in[0,T].

Then in particular, XN,γN=X𝒯X^{N,\upgamma^{N}}=X^{\mathcal{T}}. Recalling (17) the objective function for relaxed controls, we now define for each NN the objective under the environment N\mathbb{P}^{N}:

JN(γ)𝔼¯N[g(XTN,γ,Xτ(μN)N,γ,τ(μN))+𝒯TTAf(s,XsN,γ,bsP,τ(μN),ρ,ϑsN,a)γ(da,ds)].J^{N}(\upgamma)\coloneqq\mathbb{E}^{\bar{\mathbb{P}}^{N}}\left[g(X^{N,\upgamma}_{T},X^{N,\upgamma}_{\tau^{*}(\upmu^{N})},\tau^{*}(\upmu^{N}))+\int_{\mathcal{T}\wedge T}^{T}\int_{A}f(s,X^{N,\upgamma}_{s},b^{P}_{s},\tau^{*}(\upmu^{N}),\left\langle\rho,\upvartheta^{N}_{s}\right\rangle,a)\upgamma(da,ds)\right].
Lemma 3.18.

limNJN(γN)=J(γ)\lim_{N\to\infty}J^{N}(\upgamma^{N})=J^{\infty}(\upgamma^{\infty}).

Proof.

Since the convergence in Lemma 3.15 is weak, we need to uniformly approximate f,gf,g by bounded functions. For kk\in\mathbb{N} and xx\in\mathbb{R}, denote by x¯k\underline{x}_{k} the projection of xx on [k,k][-k,k]. Define fk:[0,T]××××[0,T]××Af^{k}:[0,T]\times\mathbb{R}\times\mathbb{R}\times\mathbb{R}\times[0,T]\times\mathbb{R}\times A\to\mathbb{R} and g:Ω×××[0,T]g:\Omega\times\mathbb{R}\times\mathbb{R}\times[0,T]\to\mathbb{R} by:

fk(t,x,𝔟,η,ϱ,a)=κ(a)+ϕx¯k2x¯k(𝔟¯k𝟙{t<η}+ϱ𝟙{tη}),\displaystyle f^{k}(t,x,\mathfrak{b},\upeta,\varrho,a)=\kappa(a)+\phi\underline{x}_{k}^{2}-\underline{x}_{k}\left(\underline{\mathfrak{b}}_{k}\mathds{1}_{\{t<\upeta\}}+\varrho\mathds{1}_{\{t\geq\upeta\}}\right), (41)
gk(x,y,η)=cx¯k2+βηγη¯ky¯k.\displaystyle g^{k}(x,y,\upeta)=c\underline{x}_{k}^{2}+\beta_{\upeta}\underline{\gamma_{\upeta}}_{k}\underline{y}_{k}. (42)

Recall that γ\gamma here is the bubble component defined in (4). Since we only care about γ\gamma at the burst time, we can equivalently take γt=0tbsP𝑑s\gamma_{t}=\int_{0}^{t}b^{P}_{s}ds. Since the price impact functions κ\kappa and ρ\rho are continuous, compactness of AA implies that for each kk\in\mathbb{N}, there exists some Ck>0C_{k}>0 such that

|gk(XT𝒯,Xτ(μN)𝒯,τ(μN))|+|𝒯TTAfk(s,Xs𝒯,bsP,τ(μN),ρ,ϑsN,a)γ(da,ds)|Ck.\left|g^{k}(X^{\mathcal{T}}_{T},X^{\mathcal{T}}_{\tau^{*}(\upmu^{N})},\tau^{*}(\upmu^{N}))\right|+\left|\int_{\mathcal{T}\wedge T}^{T}\int_{A}f^{k}(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu^{N}),\left\langle\rho,\upvartheta^{N}_{s}\right\rangle,a)\upgamma(da,ds)\right|\leq C_{k}.

For N{}N\in\mathbb{N}\cup\{\infty\} and kk\in\mathbb{N}, define the approximated objective JN,kJ^{N,k} on IΓ\mathrm{I}\Gamma by

JN,k(γ)𝔼N[gk(XT𝒯,Xτ(μN)𝒯,τ(μN))+𝒯TTAfk(s,Xs𝒯,bsP,τ(μN),ρ,ϑsN,a)γ(da,ds)].J^{N,k}(\upgamma)\coloneqq\mathbb{E}^{\mathbb{P}^{N}}\left[g^{k}(X^{\mathcal{T}}_{T},X^{\mathcal{T}}_{\tau^{*}(\upmu^{N})},\tau^{*}(\upmu^{N}))+\int_{\mathcal{T}\wedge T}^{T}\int_{A}f^{k}(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu^{N}),\left\langle\rho,\upvartheta^{N}_{s}\right\rangle,a)\upgamma(da,ds)\right].

Then weak convergence implies that limN|JN,k(γN)J,k(γ)|=0\lim_{N\to\infty}|J^{N,k}(\upgamma^{N})-J^{\infty,k}(\upgamma^{\infty})|=0. To shorten the notation, let τNτ(μN)\tau^{N}\coloneqq\tau^{*}(\upmu^{N}). Using (28), we have

supN𝔼N[supt[0,T]|Xt𝒯Xt𝒯¯k|2]\displaystyle\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\left[\sup_{t\in[0,T]}\left|X^{\mathcal{T}}_{t}-\underline{X^{\mathcal{T}}_{t}}_{k}\right|^{2}\right] =supN𝔼N[supt[0,T]||Xt𝒯|k|2𝟙{|Xt𝒯|>k}]\displaystyle=\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\left[\sup_{t\in[0,T]}\left||X^{\mathcal{T}}_{t}|-k\right|^{2}\mathds{1}_{\{|X^{\mathcal{T}}_{t}|>k\}}\right]
supN𝔼N[X𝒯𝒳2𝟙{X𝒯𝒳>k}]k0.\displaystyle\leq\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\left[\lVert X^{\mathcal{T}}\rVert^{2}_{\mathcal{X}^{*}}\mathds{1}_{\{\lVert X^{\mathcal{T}}\rVert_{\mathcal{X}^{*}}>k\}}\right]\stackrel{{\scriptstyle k\to\infty}}{{\longrightarrow}}0.

Recall that PP has the same law under N\mathbb{P}^{N} for each NN. Then similarly, Assumption (Assumption (B)) and Proposition A.1 together imply

supN𝔼N[supt[0,T]|btPbtP¯k|2]\displaystyle\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\left[\sup_{t\in[0,T]}|b^{P}_{t}-\underline{b^{P}_{t}}_{k}|^{2}\right] =𝔼1[supt[0,T]|btPbtP¯k|2]𝔼1[bP𝟙{bP>k}]k0.\displaystyle=\mathbb{E}^{\mathbb{P}^{1}}\left[\sup_{t\in[0,T]}|b^{P}_{t}-\underline{b^{P}_{t}}_{k}|^{2}\right]\leq\mathbb{E}^{\mathbb{P}^{1}}\left[\lVert b^{P}\rVert_{\infty}\mathds{1}_{\{\lVert b^{P}\rVert_{\infty}>k\}}\right]\stackrel{{\scriptstyle k\to\infty}}{{\longrightarrow}}0.
supN𝔼N[supt[0,T]|γtγt¯k|2]\displaystyle\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\left[\sup_{t\in[0,T]}|\gamma_{t}-\underline{\gamma_{t}}_{k}|^{2}\right] =𝔼1[supt[0,T]|γtγt¯k|2]𝔼1[γ2𝟙{γ>k}]\displaystyle=\mathbb{E}^{\mathbb{P}^{1}}\left[\sup_{t\in[0,T]}|\gamma_{t}-\underline{\gamma_{t}}_{k}|^{2}\right]\leq\mathbb{E}^{\mathbb{P}^{1}}\left[\lVert\gamma\rVert^{2}_{\infty}\mathds{1}_{\{\lVert\gamma\rVert_{\infty}>k\}}\right]
𝔼1[TbP𝟙{TbP>k}]k0.\displaystyle\leq\mathbb{E}^{\mathbb{P}^{1}}\left[T\lVert b^{P}\rVert_{\infty}\mathds{1}_{\{T\lVert b^{P}\rVert_{\infty}>k\}}\right]\stackrel{{\scriptstyle k\to\infty}}{{\longrightarrow}}0.

These uniform integrability properties, along with Assumption (Assumption (B)) and the separability condition in (C1), imply that there exists C>0C>0 such that

supN𝔼N[|ggk|(XT𝒯,XτN𝒯,τN)]csupN𝔼N[|(XT𝒯)2(XT𝒯)¯k2|]+CsupN𝔼N[|γτNγτN¯k|2]+CsupN𝔼N[|XτN𝒯XτN𝒯¯k|2]k0,\begin{gathered}\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\left[|g-g^{k}|(X^{\mathcal{T}}_{T},X^{\mathcal{T}}_{\tau^{N}},\tau^{N})\right]\leq c\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\left[\left|(X^{\mathcal{T}}_{T})^{2}-\underline{(X^{\mathcal{T}}_{T})}_{k}^{2}\right|\right]\\ +C\sqrt{\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}[|\gamma_{\tau^{N}}-\underline{\gamma_{\tau^{N}}}_{k}|^{2}]}+C\sqrt{\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}[|X^{\mathcal{T}}_{\tau^{N}}-\underline{X^{\mathcal{T}}_{\tau^{N}}}_{k}|^{2}]}\ \stackrel{{\scriptstyle k\to\infty}}{{\longrightarrow}}0,\end{gathered} (43)

and also

supN𝔼N[𝒯TTA|ffk|(s,Xs𝒯,bsP,τN,ρ,ϑsN,a)γN(da,ds)]|ϕ|0TsupN𝔼N[|(Xs𝒯)2(Xs𝒯)¯k2|]ds+C0TsupN𝔼N[|Xs𝒯(Xs𝒯)¯k|2]ds+0TsupN𝔼N[|bsPbsP¯k|2]+supN𝔼N[|Xs𝒯Xs𝒯¯k|2]dsk0.\begin{gathered}\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\Bigl{[}\int_{\mathcal{T}\wedge T}^{T}\int_{A}|f-f^{k}|(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{N},\left\langle\rho,\upvartheta^{N}_{s}\right\rangle,a)\upgamma^{N}(da,ds)\Bigr{]}\\ \leq|\phi|\int_{0}^{T}\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\left[\left|(X^{\mathcal{T}}_{s})^{2}-\underline{(X^{\mathcal{T}}_{s})}_{k}^{2}\right|\right]ds+C\int_{0}^{T}\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}\left[\left|X^{\mathcal{T}}_{s}-\underline{(X^{\mathcal{T}}_{s})}_{k}\right|^{2}\right]ds\\ +\int_{0}^{T}\sqrt{\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}[|b^{P}_{s}-\underline{b^{P}_{s}}_{k}|^{2}]}+\sqrt{\sup_{N\in\mathbb{N}}\mathbb{E}^{\mathbb{P}^{N}}[|X^{\mathcal{T}}_{s}-\underline{X^{\mathcal{T}}_{s}}_{k}|^{2}]}\ ds\stackrel{{\scriptstyle k\to\infty}}{{\longrightarrow}}0.\end{gathered} (44)

Therefore, for any fixed kk\in\mathbb{N},

|JN(γN)J(γ)|\displaystyle\left|J^{N}(\upgamma^{N})-J^{\infty}(\upgamma^{\infty})\right| |JN(γN)JN,k(γN)|+|JN,k(γN)J,k(γ)|+|J,k(γ)J(γ)|.\displaystyle\leq\left|J^{N}(\upgamma^{N})-J^{N,k}(\upgamma^{N})\right|+\left|J^{N,k}(\upgamma^{N})-J^{\infty,k}(\upgamma^{\infty})\right|+\left|J^{\infty,k}(\upgamma^{\infty})-J^{\infty}(\upgamma^{\infty})\right|.

Taking limit NN\to\infty gives

limN|JN(γN)J(γ)|supN|JN(γN)JN,k(γN)|+|J,k(γ)J(γ)|.\lim_{N\to\infty}\left|J^{N}(\upgamma^{N})-J^{\infty}(\upgamma^{\infty})\right|\leq\sup_{N\in\mathbb{N}}\left|J^{N}(\upgamma^{N})-J^{N,k}(\upgamma^{N})\right|+\left|J^{\infty,k}(\upgamma^{\infty})-J^{\infty}(\upgamma^{\infty})\right|.

Taking kk\to\infty on the right hand side and using (43) and (44) give the result. ∎

Let βIΓ\beta\in\mathrm{I}\Gamma be another 𝔾\mathbb{G}-admissible relaxed strategy. Following the proof of Lemma 3.18, we also have JN(β)NJ(β)J^{N}(\beta)\stackrel{{\scriptstyle N\to\infty}}{{\longrightarrow}}J(\beta). Remarks 3.13 and 3.14 together imply that JN(γN)JN(β)J^{N}(\upgamma^{N})\leq J^{N}(\beta) for each NN. Taking NN\to\infty on both sides we have J(γ)J(β)J^{\infty}(\upgamma^{\infty})\leq J^{\infty}(\beta) for all βIΓ\beta\in\mathrm{I}\Gamma, so optimality is proved.

Now recall that for an AA-valued control α𝔸\alpha\in\mathbb{A}^{*}, we denote by γ(α)\upgamma(\alpha) its corresponding strict control in the space of relaxed controls, where each time marginal is the Dirac measure at αt\alpha_{t}. Using the optimality lemma above, we can in fact show that γ\upgamma^{\infty} must be a strict control.

Lemma 3.19.

There is a version of γ\upgamma^{\infty} that is 𝔽B,X𝒯,W,,D\mathbb{F}^{B,X^{\mathcal{T}},W^{\infty},\mathcal{I},D}-progressively measurable that is a strict control taking the form γ=γ(α^)\upgamma^{\infty}=\upgamma(\hat{\alpha}^{\infty}) for some α^𝔸\hat{\alpha}^{\infty}\in\mathbb{A}^{*}.

Proof.

Define αtAaγt(da)\alpha^{\infty}_{t}\coloneqq\int_{A}a\upgamma^{\infty}_{t}(da) for t[0,T]t\in[0,T] and γ~γ(α)\tilde{\upgamma}\coloneqq\upgamma(\alpha^{\infty}). Then α𝔸\alpha^{\infty}\in\mathbb{A}^{*} and γ~IΓ\tilde{\upgamma}\in\mathrm{I}\Gamma is a strict control. It is obvious that γ~\tilde{\upgamma} and γ\upgamma^{\infty} both give rise to the same state process X𝒯X^{\mathcal{T}} according to (16). Using strict convexity of ff in aa and Jensen’s inequality, we have

J(γ~)\displaystyle J^{\infty}(\tilde{\upgamma}) =𝔼¯[g(XT𝒯,Xτ(μ)𝒯,τ(μ))+𝒯TTAf(s,Xs𝒯,bsP,τ(μ),ρ,ϑs,a)γ~(da,ds)]\displaystyle=\bar{\mathbb{E}}\left[g(X^{\mathcal{T}}_{T},X^{\mathcal{T}}_{\tau^{*}(\upmu^{\infty})},\tau^{*}(\upmu^{\infty}))+\int_{\mathcal{T}\wedge T}^{T}\int_{A}f(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu^{\infty}),\left\langle\rho,\upvartheta^{\infty}_{s}\right\rangle,a)\tilde{\upgamma}(da,ds)\right]
=𝔼¯[g(XT𝒯,Xτ(μ)𝒯,τ(μ))+𝒯TTf(s,Xs𝒯,bsP,τ(μ),ρ,ϑs,αs)𝑑s]\displaystyle=\bar{\mathbb{E}}\left[g(X^{\mathcal{T}}_{T},X^{\mathcal{T}}_{\tau^{*}(\upmu^{\infty})},\tau^{*}(\upmu^{\infty}))+\int_{\mathcal{T}\wedge T}^{T}f(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu^{\infty}),\left\langle\rho,\upvartheta^{\infty}_{s}\right\rangle,\alpha^{\infty}_{s})ds\right]
𝔼¯[g(XT𝒯,Xτ(μ)𝒯,τ(μ))+𝒯TTAf(s,Xs𝒯,bsP,τ(μ),ρ,ϑs,a)γ(da,ds)]\displaystyle\leq\bar{\mathbb{E}}\left[g(X^{\mathcal{T}}_{T},X^{\mathcal{T}}_{\tau^{*}(\upmu^{\infty})},\tau^{*}(\upmu^{\infty}))+\int_{\mathcal{T}\wedge T}^{T}\int_{A}f(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu^{\infty}),\left\langle\rho,\upvartheta^{\infty}_{s}\right\rangle,a)\upgamma^{\infty}(da,ds)\right]
=J(γ).\displaystyle=J^{\infty}(\upgamma^{\infty}).

The inequality is strict (which contradicts with optimality of γ\upgamma^{\infty}) unless γ=γ(α)\upgamma^{\infty}=\upgamma(\alpha^{\infty}).

Lebesgue differentiation theorem allows us to define α^𝔸\hat{\alpha}^{\infty}\in\mathbb{A}^{*} by

α^t={limnn(t1/n)+tαs𝑑s if the limit exists0 otherwise.\hat{\alpha}^{\infty}_{t}=\begin{cases}\lim_{n\to\infty}n\int_{(t-1/n)+}^{t}\alpha^{\infty}_{s}ds&\text{ if the limit exists}\\ 0&\text{ otherwise}.\end{cases}

Then ¯dt\bar{\mathbb{P}}\otimes dt almost surely, α^t=αt\hat{\alpha}^{\infty}_{t}=\alpha^{\infty}_{t}. Note that α^\hat{\alpha}^{\infty} shares the same measurability with 0αs𝑑s\int_{0}^{\cdot}\alpha^{\infty}_{s}ds, which by (12) is 𝔽X𝒯,W,B,,D\mathbb{F}^{X^{\mathcal{T}},W^{\infty},B,\mathcal{I},D} measurable. ∎

A consequence of the lemma above is that we can drop either γ\upgamma^{\infty} or X𝒯X^{\mathcal{T}} from the definition of 𝔾\mathbb{G} and simply consider 𝔾=𝔽B,𝔐,X𝒯,W,,D=𝔽B,𝔐,γ,W,,D\mathbb{G}=\mathbb{F}^{B,\mathfrak{M}^{\infty},X^{\mathcal{T}},W^{\infty},\mathcal{I},D}=\mathbb{F}^{B,\mathfrak{M}^{\infty},\upgamma^{\infty},W^{\infty},\mathcal{I},D}. Moreover, both optimality and consistency still hold for γ=γ(α^)\upgamma^{\infty}=\upgamma(\hat{\alpha}^{\infty}). 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 𝔽¯𝔽,𝔐,W,B,D\bar{\mathbb{F}}\coloneqq\mathbb{F}^{\mathcal{I},\mathfrak{M}^{\infty},W^{\infty},B,D} is immersed in 𝔾=(𝒢t)t[0,T]\mathbb{G}=(\mathcal{G}_{t})_{t\in[0,T]} defined above. We need to keep in mind that while \mathcal{I} is 𝒢0\mathcal{G}_{0}-measurable, it is not 0𝒳𝒯\mathcal{F}^{\mathcal{X}^{\mathcal{T}}}_{0}-measurable due to random entry, which is why we need to treat \mathcal{I} separately.

Lemma 3.20.

The filtration 𝔾\mathbb{G} is compatible with (,𝔐,W,B,D)(\mathcal{I},\mathfrak{M}^{\infty},W^{\infty},B,D).

Proof.

By Proposition 3.2, it suffices to show that for all t[0,T]t\in[0,T], tX𝒯\mathcal{F}^{X^{\mathcal{T}}}_{t} is conditionally independent from ¯T=T,𝔐,W,B,D\bar{\mathcal{F}}_{T}=\mathcal{F}^{\mathcal{I},\mathfrak{M}^{\infty},W^{\infty},B,D}_{T} given ¯t=t,𝔐,W,B,D\bar{\mathcal{F}}_{t}=\mathcal{F}^{\mathcal{I},\mathfrak{M}^{\infty},W^{\infty},B,D}_{t}. We follow the proof of [20, Lemma 3.7].

Lemma 3.17 implies that WW^{\infty} is a ¯\bar{\mathbb{P}}-Brownian motion independent from (B,𝔐,D)(B,\mathfrak{M}^{\infty},D). Fix t[0,T]t\in[0,T]. Consider three bounded functions ϕtm,ϕt+w\phi^{m}_{t},\phi^{w}_{t+} and ϕt1\phi^{1}_{t} where ϕtm:𝒫1(Ω¯1)\phi^{m}_{t}:\mathcal{P}_{1}(\bar{\Omega}_{1})\to\mathbb{R} is t𝔐\mathcal{F}_{t}^{\mathfrak{M}^{\infty}} measurable, ϕt+w:𝒳\phi^{w}_{t+}:\mathcal{X}\to\mathbb{R} is σ(WsWt:s[t,T])\sigma(W_{s}-W_{t}:s\in[t,T]) measurable, and ϕt1:Ω¯1\phi^{1}_{t}:\bar{\Omega}_{1}\to\mathbb{R} is t,X𝒯,W\mathcal{F}^{\mathcal{I},X^{\mathcal{T}},W^{\infty}}_{t} measurable. By Lemma 3.17 and property of Brownian motion we have

𝔼¯[ϕtm(𝔐)Ω¯1ϕt1𝑑𝔐]𝔼¯[ϕt+w(W)]\displaystyle\bar{\mathbb{E}}\left[\phi_{t}^{m}(\mathfrak{M}^{\infty})\int_{\bar{\Omega}_{1}}\phi^{1}_{t}d\mathfrak{M}^{\infty}\right]\bar{\mathbb{E}}\left[\phi^{w}_{t+}(W^{\infty})\right] =𝔼¯[ϕtm(𝔐)ϕt1(X𝒯,W,γ,)ϕt+w(W)]\displaystyle=\bar{\mathbb{E}}\left[\phi_{t}^{m}(\mathfrak{M}^{\infty})\phi^{1}_{t}(X^{\mathcal{T}},W^{\infty},\upgamma^{\infty},\mathcal{I})\phi^{w}_{t+}(W^{\infty})\right]
=𝔼¯[ϕtm(𝔐)Ω¯1ϕt+w(w)ϕt1(x,w,q,ι)𝑑𝔐(x,w,q,ι)].\displaystyle=\bar{\mathbb{E}}\left[\phi_{t}^{m}(\mathfrak{M}^{\infty})\int_{\bar{\Omega}_{1}}\phi^{w}_{t+}(w)\phi^{1}_{t}(x,w,q,\iota)d\mathfrak{M}^{\infty}(x,w,q,\iota)\right].

Since this holds for all ϕtm\phi_{t}^{m}, ¯\bar{\mathbb{P}} almost surely we have

𝔼¯[ϕt+w(W)]𝔼𝔐[ϕt1(X𝒯,γ,,W)]=𝔼𝔐[ϕt+w(W)ϕt1(X𝒯,γ,,W)]\bar{\mathbb{E}}\left[\phi^{w}_{t+}(W^{\infty})\right]\mathbb{E}^{\mathfrak{M}^{\infty}}[\phi^{1}_{t}(X^{\mathcal{T}},\upgamma^{\infty},\mathcal{I},W^{\infty})]=\mathbb{E}^{\mathfrak{M}^{\infty}}[\phi^{w}_{t+}(W^{\infty})\phi^{1}_{t}(X^{\mathcal{T}},\upgamma^{\infty},\mathcal{I},W^{\infty})] (45)

where by 𝔼𝔐[ϕ(X𝒯,γ,,W)]\mathbb{E}^{\mathfrak{M}^{\infty}}[\phi(X^{\mathcal{T}},\upgamma^{\infty},\mathcal{I},W^{\infty})] we mean the integral Ω¯ϕ𝑑𝔐\int_{\bar{\Omega}}\phi d\mathfrak{M}^{\infty} for ϕ:Ω¯1\phi:\bar{\Omega}_{1}\to\mathbb{R}. Note that this expectation is ¯t\bar{\mathcal{F}}_{t}-measurable if ϕ\phi is t,X𝒯,W\mathcal{F}^{\mathcal{I},X^{\mathcal{T}},W^{\infty}}_{t}-measurable.

Additionally, consider bounded functions ϕι,ϕtx,φt,φT\phi^{\iota},\phi^{x}_{t},\varphi_{t},\varphi_{T} where ϕι:2\phi^{\iota}:\mathbb{R}^{2}\to\mathbb{R} is Borel measurable, ϕtx:𝒳\phi^{x}_{t}:\mathcal{X}^{*}\to\mathbb{R} is tX𝒯\mathcal{F}^{X^{\mathcal{T}}}_{t} measurable, φt\varphi_{t} and φT\varphi_{T} are functions from 𝒳×𝒫1(Ω¯1)×𝒳D\mathcal{X}\times\mathcal{P}_{1}(\bar{\Omega}_{1})\times\mathcal{X}_{D} to \mathbb{R} that are tB,𝔐,D\mathcal{F}^{B,\mathfrak{M}^{\infty},D}_{t} and TB,𝔐,D\mathcal{F}^{B,\mathfrak{M}^{\infty},D}_{T} measurable, respectively. Using (45) and Lemma 3.17, we have

𝔼¯[ϕtx(X𝒯)φT(B,𝔐,D)ϕι()ϕt+w(W)ϕtw(W)φt(B,𝔐,D)]=𝔼¯[𝔼𝔐[ϕtx(X𝒯)ϕι()ϕt+w(W)ϕtw(W)](φTφt)(B,𝔐,D)]=𝔼¯[𝔼𝔐[ϕtx(X𝒯)ϕι()ϕtw(W)](φTφt)(B,𝔐,D)]𝔼¯[ϕt+w(W)]=𝔼¯[𝔼¯[ϕtx(X𝒯)ϕι()ϕtw(W)|¯t](φTφt)(B,𝔐,D)]𝔼¯[ϕt+w(W)]=𝔼¯[𝔼¯[ϕtx(X𝒯)ϕι()ϕtw(W)|¯t]φt(B,𝔐,D)𝔼¯[φT(B,𝔐,D)|¯t]]𝔼¯[ϕt+w(W)]=𝔼¯[𝔼¯[ϕtx(X𝒯)|¯t]𝔼¯[φT(B,𝔐,D)|¯t]φt(B,𝔐,D)ϕι()ϕtw(W)ϕt+w(W)],\begin{split}\bar{\mathbb{E}}[&\phi^{x}_{t}(X^{\mathcal{T}})\varphi_{T}(B,\mathfrak{M}^{\infty},D)\phi^{\iota}(\mathcal{I})\phi^{w}_{t+}(W^{\infty})\phi_{t}^{w}(W^{\infty})\varphi_{t}(B,\mathfrak{M}^{\infty},D)]\\ &=\bar{\mathbb{E}}\left[\mathbb{E}^{\mathfrak{M}^{\infty}}[\phi^{x}_{t}(X^{\mathcal{T}})\phi^{\iota}(\mathcal{I})\phi^{w}_{t+}(W^{\infty})\phi_{t}^{w}(W^{\infty})](\varphi_{T}\cdot\varphi_{t})(B,\mathfrak{M}^{\infty},D)\right]\\ &=\bar{\mathbb{E}}\left[\mathbb{E}^{\mathfrak{M}^{\infty}}[\phi^{x}_{t}(X^{\mathcal{T}})\phi^{\iota}(\mathcal{I})\phi_{t}^{w}(W^{\infty})](\varphi_{T}\cdot\varphi_{t})(B,\mathfrak{M}^{\infty},D)\right]\bar{\mathbb{E}}[\phi^{w}_{t+}(W^{\infty})]\\ &=\bar{\mathbb{E}}\biggl{[}\bar{\mathbb{E}}\left[\phi^{x}_{t}(X^{\mathcal{T}})\phi^{\iota}(\mathcal{I})\phi_{t}^{w}(W^{\infty})|\bar{\mathcal{F}}_{t}\right](\varphi_{T}\cdot\varphi_{t})(B,\mathfrak{M}^{\infty},D)\biggr{]}\bar{\mathbb{E}}[\phi^{w}_{t+}(W^{\infty})]\\ &=\bar{\mathbb{E}}\biggl{[}\bar{\mathbb{E}}\left[\phi^{x}_{t}(X^{\mathcal{T}})\phi^{\iota}(\mathcal{I})\phi_{t}^{w}(W^{\infty})|\bar{\mathcal{F}}_{t}\right]\varphi_{t}(B,\mathfrak{M}^{\infty},D)\bar{\mathbb{E}}\left[\varphi_{T}(B,\mathfrak{M}^{\infty},D)|\bar{\mathcal{F}}_{t}\right]\biggr{]}\bar{\mathbb{E}}[\phi^{w}_{t+}(W^{\infty})]\\ &=\bar{\mathbb{E}}\left[\bar{\mathbb{E}}[\phi_{t}^{x}(X^{\mathcal{T}})|\bar{\mathcal{F}}_{t}]\bar{\mathbb{E}}\left[\varphi_{T}(B,\mathfrak{M}^{\infty},D)|\bar{\mathcal{F}}_{t}\right]\varphi_{t}(B,\mathfrak{M}^{\infty},D)\phi^{\iota}(\mathcal{I})\phi_{t}^{w}(W^{\infty})\phi^{w}_{t+}(W^{\infty})\right],\end{split} (46)

where the last equality follows from the independence of ϕt+w(W)\phi^{w}_{t+}(W^{\infty}) and ¯t\bar{\mathcal{F}}_{t}. Since ϕι\phi^{\iota} and ϕtw\phi_{t}^{w} are arbitrary, we can replace them with bounded ϕιφι\phi^{\iota}\cdot\varphi^{\iota} and ϕtwφtw\phi^{w}_{t}\cdot\varphi_{t}^{w}, each with the same corresponding mesurability requirements. Then by definition of conditional expectation we have

𝔼¯[ϕtx(X𝒯)\displaystyle\bar{\mathbb{E}}\bigl{[}\phi^{x}_{t}(X^{\mathcal{T}}) φT(B,𝔐,D)ϕι()ϕt+w(W)ϕtw(W)|¯t]\displaystyle\varphi_{T}(B,\mathfrak{M}^{\infty},D)\phi^{\iota}(\mathcal{I})\phi^{w}_{t+}(W^{\infty})\phi_{t}^{w}(W^{\infty})|\bar{\mathcal{F}}_{t}\bigr{]}
=𝔼¯[ϕtx(X𝒯)|¯t]𝔼¯[φT(B,𝔐,D)ϕι()ϕt+w(W)ϕtw(W)|¯t].\displaystyle=\bar{\mathbb{E}}\left[\phi^{x}_{t}(X^{\mathcal{T}})|\bar{\mathcal{F}}_{t}\right]\bar{\mathbb{E}}\left[\varphi_{T}(B,\mathfrak{M}^{\infty},D)\phi^{\iota}(\mathcal{I})\phi^{w}_{t+}(W^{\infty})\phi_{t}^{w}(W^{\infty})|\bar{\mathcal{F}}_{t}\right].

We conclude by noting that TW\mathcal{F}^{W^{\infty}}_{T} is generated by ϕt+w(W)ϕtw(W)\phi^{w}_{t+}(W^{\infty})\phi_{t}^{w}(W^{\infty}) with arbitrary ϕt+w\phi^{w}_{t+} and ϕtw\phi^{w}_{t}. ∎

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 γ(α)\upgamma(\alpha^{\infty}), and α\alpha^{\infty} is 𝔽B,X𝒯,W,𝔐,,D\mathbb{F}^{B,X^{\mathcal{T}},W^{\infty},\mathfrak{M}^{\infty},\mathcal{I},D} progressive. In order to obtain an equilibrium with strong control, we will show that α\alpha^{\infty} is 𝔽B,W,𝔓,,D\mathbb{F}^{B,W^{\infty},\mathfrak{P}^{\infty},\mathcal{I},D} measurable after bringing the lifted environment 𝔐\mathfrak{M}^{\infty} back to the “original” environment 𝔓=(μ,ϑ)=(𝔐,x,𝔐,γ)\mathfrak{P}^{\infty}=(\upmu^{\infty},\upvartheta^{\infty})=(\mathfrak{M}^{\infty,x},\mathfrak{M}^{\infty,\upgamma}).

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 𝔐N\mathfrak{M}^{N} to be the joint conditional law of (X𝒯,WN,γN,)(X^{\mathcal{T}},W^{N},\upgamma^{N},\mathcal{I}) given (B,D)(B,D) under N\mathbb{P}^{N}. 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 NN\in\mathbb{N}, we could alternatively define for each (β,η)𝒳×+(\beta,\upeta)\in\mathcal{X}\times\mathbb{R}_{+}:

𝔐~N(β,η)k=1|𝒱N|𝟙{(β,D(η))Ak}αN,k(X𝒯)αN,k(WN)αN,k(γN)αN,k().\widetilde{\mathfrak{M}}^{N}(\beta,\upeta)\coloneqq\sum_{k=1}^{|\mathcal{V}_{N}|}\mathds{1}_{\{(\beta,D(\upeta))\in A_{k}\}}\mathcal{L}^{\alpha^{N},k}(X^{\mathcal{T}})\otimes\mathcal{L}^{\alpha^{N},k}(W^{N})\otimes\mathcal{L}^{\alpha^{N},k}(\upgamma^{N})\otimes\mathcal{L}^{\alpha^{N},k}(\mathcal{I}).

This version still carries the necessary inputs 𝔓N\mathfrak{P}^{N} to the BSDE (26) as its first and third marginals, and tightness of N(𝔐~N)1\mathbb{P}^{N}\circ(\widetilde{\mathfrak{M}}^{N})^{-1} follows immediately from that of N(𝔐N)1\mathbb{P}^{N}\circ(\mathfrak{M}^{N})^{-1} in Lemma 3.15. Then we take a limit point ¯(B,τ,𝔐~,X𝒯,W,γ,)1\bar{\mathbb{P}}\coloneqq\mathbb{P}^{\infty}\circ(B,\tau,\widetilde{\mathfrak{M}}^{\infty},X^{\mathcal{T}},W^{\infty},\upgamma^{\infty},\mathcal{I})^{-1} and follow the same argument in Remark 3.16 to work on the canonical space. In particular, 𝔐~\widetilde{\mathfrak{M}}^{\infty} is the canonical process on 𝒫1(Ω¯1)\mathcal{P}_{1}(\bar{\Omega}_{1}). Following the argument in Lemma 3.17, the fixed point property for each NN\in\mathbb{N} now leads to a weaker consistency in the limit. Namely, for all bounded, uniformly continuous, \mathbb{R}-valued functions h0,hx1,hw1,hγ1,hι1h^{0},h^{1}_{x},h^{1}_{w},h^{1}_{\upgamma},h^{1}_{\iota} with respective domains 𝒳×𝒫1(Ω¯1)×𝒳D,𝒳,𝒳,Γ,2\mathcal{X}\times\mathcal{P}_{1}(\bar{\Omega}_{1})\times\mathcal{X}_{D},\mathcal{X}^{*},\mathcal{X},\Gamma,\mathbb{R}^{2}, we have

𝔼¯[h0(B,𝔐~,D)𝒳hx1(x)d𝔐~(x)𝒳hw1(w)d𝔐~(w)Γhγ1(q)d𝔐~(q)2hι1(ι)d𝔐~(ι)]=𝔼¯[h0(B,𝔐~,D)hx1(X𝒯)hw1(W)hγ1(γ)hι1()]=𝔼¯[h0(B,𝔐~,D)Ω¯1hx1(x)hw1(w)hγ1(q)hι1(ι)𝑑𝔐~(x,w,q,ι)],\begin{split}\bar{\mathbb{E}}\biggl{[}&h^{0}(B,\widetilde{\mathfrak{M}}^{\infty},D)\int_{\mathcal{X}^{*}}h^{1}_{x}(x)d\widetilde{\mathfrak{M}}^{\infty}(x)\int_{\mathcal{X}}h^{1}_{w}(w)d\widetilde{\mathfrak{M}}^{\infty}(w)\int_{\Gamma}h^{1}_{\upgamma}(q)d\widetilde{\mathfrak{M}}^{\infty}(q)\int_{\mathbb{R}^{2}}h^{1}_{\iota}(\iota)d\widetilde{\mathfrak{M}}^{\infty}(\iota)\biggr{]}\\ &=\bar{\mathbb{E}}[h^{0}(B,\widetilde{\mathfrak{M}}^{\infty},D)h^{1}_{x}(X^{\mathcal{T}})h^{1}_{w}(W^{\infty})h^{1}_{\upgamma}(\upgamma^{\infty})h^{1}_{\iota}(\mathcal{I})]\\ &=\bar{\mathbb{E}}\left[h^{0}(B,\widetilde{\mathfrak{M}}^{\infty},D)\int_{\bar{\Omega}_{1}}h^{1}_{x}(x)h^{1}_{w}(w)h^{1}_{\upgamma}(q)h^{1}_{\iota}(\iota)d\widetilde{\mathfrak{M}}^{\infty}(x,w,q,\iota)\right],\end{split} (47)

which results from taking NN\to\infty of the following equalities by construction

𝔼¯[h0(VN,𝔐~N,DN)𝒳hx1(x)d𝔐~N(x)𝒳hw1(w)d𝔐~N(w)Γhγ1(q)d𝔐~N(q)2hι1(ι)d𝔐~N(ι)]=𝔼¯[h0(VN,𝔐~N,DN)hx1(X𝒯)hw1(WN)hγ1(γN)hι1()]=𝔼¯[h0(VN,𝔐~N,DN)Ω¯1hx1(x)hw1(w)hγ1(q)hι1(ι)𝑑𝔐~N(x,w,q,ι)].\begin{split}\bar{\mathbb{E}}\biggl{[}&h^{0}(V^{N},\widetilde{\mathfrak{M}}^{N},D^{N})\int_{\mathcal{X}^{*}}h^{1}_{x}(x)d\widetilde{\mathfrak{M}}^{N}(x)\int_{\mathcal{X}}h^{1}_{w}(w)d\widetilde{\mathfrak{M}}^{N}(w)\int_{\Gamma}h^{1}_{\upgamma}(q)d\widetilde{\mathfrak{M}}^{N}(q)\int_{\mathbb{R}^{2}}h^{1}_{\iota}(\iota)d\widetilde{\mathfrak{M}}^{N}(\iota)\biggr{]}\\ &=\bar{\mathbb{E}}[h^{0}(V^{N},\widetilde{\mathfrak{M}}^{N},D^{N})h^{1}_{x}(X^{\mathcal{T}})h^{1}_{w}(W^{N})h^{1}_{\upgamma}(\upgamma^{N})h^{1}_{\iota}(\mathcal{I})]\\ &=\bar{\mathbb{E}}\left[h^{0}(V^{N},\widetilde{\mathfrak{M}}^{N},D^{N})\int_{\bar{\Omega}_{1}}h^{1}_{x}(x)h^{1}_{w}(w)h^{1}_{\upgamma}(q)h^{1}_{\iota}(\iota)d\widetilde{\mathfrak{M}}^{N}(x,w,q,\iota)\right].\end{split}

This is also sufficient for the consistency requirement in Definition 3.3. Similarly, with 𝔐~\widetilde{\mathfrak{M}}^{\infty} the equality (45) holds only for ϕt1\phi^{1}_{t} 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 𝔐\mathfrak{M}^{\infty} with 𝔐~\widetilde{\mathfrak{M}}^{\infty} in the final filtration 𝔾=𝔽B,X𝒯,W,𝔐~,,D\mathbb{G}=\mathbb{F}^{B,X^{\mathcal{T}},W^{\infty},\widetilde{\mathfrak{M}}^{\infty},\mathcal{I},D}.

Note from (47) that ¯\bar{\mathbb{P}} almost surely, 𝔐~\widetilde{\mathfrak{M}}^{\infty} 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 (νKνp)(\nu_{K}\otimes\nu_{p}), respectively, since for each NN\in\mathbb{N}, (WN,,B,D)(W^{N},\mathcal{I},B,D) are mutually independent under N\mathbb{P}^{N}. Being complete, the filtration 𝔽𝔓\mathbb{F}^{\mathfrak{P}^{\infty}} in the original environment coincides with 𝔽𝔐~\mathbb{F}^{\widetilde{\mathfrak{M}}^{\infty}} from the lifted environment. Therefore, we can equivalently take 𝔾=𝔽B,X𝒯,W,𝔓,,D\mathbb{G}=\mathbb{F}^{B,X^{\mathcal{T}},W^{\infty},\mathfrak{P}^{\infty},\mathcal{I},D}, and the compatibility condition reads that 𝔽¯𝔽B,W,𝔓,,D\bar{\mathbb{F}}\coloneqq\mathbb{F}^{B,W^{\infty},\mathfrak{P}^{\infty},\mathcal{I},D} is immersed in 𝔾\mathbb{G}.

4.1.2. Strong Control via Optional Projection

To further strengthen the measurability property of γ\upgamma from 𝔾\mathbb{G} to 𝔽¯\bar{\mathbb{F}}, we follow the proof of [20, Proposition 4.4]. Recall that the state equation (12) with WW^{\infty} as the Brownian motion is satisfied by X𝒯X^{\mathcal{T}} and α^\hat{\alpha}^{\infty}. By optional projection we can find 𝔽¯\bar{\mathbb{F}}-optional processes X¯𝒯\bar{X}^{\mathcal{T}} and α¯\bar{\alpha} such that for any finite 𝔽¯\bar{\mathbb{F}}-stopping time ρ\rho:

X¯ρ𝒯𝔼¯[Xρ𝒯|¯ρ],α¯ρ𝔼¯[α^ρ|¯ρ],a.s.\bar{X}^{\mathcal{T}}_{\rho}\coloneqq\bar{\mathbb{E}}[X^{\mathcal{T}}_{\rho}|\bar{\mathcal{F}}_{\rho}],\quad\bar{\alpha}_{\rho}\coloneqq\bar{\mathbb{E}}[\hat{\alpha}^{\infty}_{\rho}|\bar{\mathcal{F}}_{\rho}],\quad\operatorname*{\textit{a.s.}} (48)

Since 𝔽¯\bar{\mathbb{F}} is immersed in 𝔾\mathbb{G}, Proposition 3.2 implies that for each 0stT0\leq s\leq t\leq T,

X¯s𝒯=𝔼¯[Xs𝒯|¯t],α¯s=𝔼¯[α^s|¯t],a.s.\bar{X}^{\mathcal{T}}_{s}=\bar{\mathbb{E}}[X^{\mathcal{T}}_{s}|\bar{\mathcal{F}}_{t}],\quad\bar{\alpha}_{s}=\bar{\mathbb{E}}[\hat{\alpha}^{\infty}_{s}|\bar{\mathcal{F}}_{t}],\quad\operatorname*{\textit{a.s.}} (49)

Using Fubini’s theorem for conditional expectation along with (49) on (12), we can replace X¯𝒯\bar{X}^{\mathcal{T}} by a modification such that ¯\bar{\mathbb{P}} almost surely

X¯t𝒯=𝟙{t𝒯}K0/𝒫+0tα¯s𝑑s+σ(Wt𝒯W𝒯),t[0,T].\bar{X}^{\mathcal{T}}_{t}=\mathds{1}_{\{t\geq\mathcal{T}\}}K_{0}/\mathscr{P}+\int_{0}^{t}\bar{\alpha}_{s}ds+\sigma(W^{\infty}_{t\vee\mathcal{T}}-W^{\infty}_{\mathcal{T}}),\quad t\in[0,T].

Notice that given 𝔓=(𝔐~,x,𝔐~,γ)(μ,ϑ)\mathfrak{P}^{\infty}=(\widetilde{\mathfrak{M}}^{\infty,x},\widetilde{\mathfrak{M}}^{\infty,\upgamma})\equiv(\upmu^{\infty},\upvartheta^{\infty}), the bubble burst time τ(μ)\tau^{*}(\upmu^{\infty}) is a 𝔽¯\bar{\mathbb{F}}-stopping time, and the bubble component γτ(μ)\gamma_{\tau^{*}(\upmu^{\infty})} is ¯τ(μ)\bar{\mathcal{F}}_{\tau^{*}(\upmu^{\infty})} measurable. Recall also from (C1) that the running cost ff depends on τ(μ)\tau^{*}(\upmu^{\infty}) only through DD, which is 𝔽¯\bar{\mathbb{F}}-adapted. Then by conditional Jensen’s inequality, Remark (C3) and (48),

Jμ,ϑ(α^)\displaystyle J^{\upmu^{\infty},\upvartheta^{\infty}}(\hat{\alpha}^{\infty}) =𝔼¯[𝒯TTf(s,Xs𝒯,bsP,τ(μ),ρ,ϑs,α^s)𝑑s+c(XT𝒯)2+βτ(μ)γτ(μ)Xτ(μ)𝒯]\displaystyle=\bar{\mathbb{E}}\left[\int_{\mathcal{T}\wedge T}^{T}f(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu^{\infty}),\left\langle\rho,\upvartheta_{s}^{\infty}\right\rangle,\hat{\alpha}_{s})ds+c(X^{\mathcal{T}}_{T})^{2}+\beta_{\tau^{*}(\upmu^{\infty})}\gamma_{\tau^{*}(\upmu^{\infty})}X^{\mathcal{T}}_{\tau^{*}(\upmu^{\infty})}\right]
=𝔼¯[𝒯TT𝔼¯[f(s,Xs𝒯,bsP,τ(μ),ρ,ϑs,α^s)|¯s]𝑑s]\displaystyle=\bar{\mathbb{E}}\left[\int_{\mathcal{T}\wedge T}^{T}\bar{\mathbb{E}}\left[f(s,X^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu^{\infty}),\left\langle\rho,\upvartheta_{s}^{\infty}\right\rangle,\hat{\alpha}_{s})|\bar{\mathcal{F}}_{s}\right]ds\right]
+c𝔼¯[𝔼¯[|XT𝒯|2|¯T]]+𝔼¯[βτ(μ)γτ(μ)𝔼¯[Xτ(μ)𝒯|¯τ(μ)]]\displaystyle\qquad\qquad+c\bar{\mathbb{E}}\left[\bar{\mathbb{E}}\left[|X^{\mathcal{T}}_{T}|^{2}|\bar{\mathcal{F}}_{T}\right]\right]+\bar{\mathbb{E}}\left[\beta_{\tau^{*}(\upmu^{\infty})}\gamma_{\tau^{*}(\upmu^{\infty})}\bar{\mathbb{E}}\left[X^{\mathcal{T}}_{\tau^{*}(\upmu^{\infty})}|\bar{\mathcal{F}}_{\tau^{*}(\upmu^{\infty})}\right]\right]
𝔼¯[𝒯TTf(s,X¯s𝒯,bsP,τ(μ),ρ,ϑs,α¯s)𝑑s+c(X¯T𝒯)2+βτ(μ)γτ(μ)X¯τ(μ)𝒯]\displaystyle\geq\bar{\mathbb{E}}\left[\int_{\mathcal{T}\wedge T}^{T}f(s,\bar{X}^{\mathcal{T}}_{s},b^{P}_{s},\tau^{*}(\upmu^{\infty}),\left\langle\rho,\upvartheta_{s}^{\infty}\right\rangle,\bar{\alpha}_{s})ds+c(\bar{X}^{\mathcal{T}}_{T})^{2}+\beta_{\tau^{*}(\upmu^{\infty})}\gamma_{\tau^{*}(\upmu^{\infty})}\bar{X}^{\mathcal{T}}_{\tau^{*}(\upmu^{\infty})}\right]
=Jμ,ϑ(α¯).\displaystyle=J^{\upmu^{\infty},\upvartheta^{\infty}}(\bar{\alpha}).

By strict convexity of ff in (x,a)(x,a), the inequality is strict unless α^\hat{\alpha}^{\infty} and X𝒯X^{\mathcal{T}} are both already 𝔽¯\bar{\mathbb{F}} adapted. Strict inequality would lead to a contradiction to optimality of α^\hat{\alpha}^{\infty} among 𝔾\mathbb{G}-progressive controls, since α¯\bar{\alpha} is 𝔽¯\bar{\mathbb{F}}-optional, hence also 𝔾\mathbb{G}-progressive.

4.1.3. Exogenous Burst Time as Totally Inaccessible Stopping Time

The section above implies that we can take 𝔾=𝔽¯=𝔽,B,W,D,𝔓\mathbb{G}=\bar{\mathbb{F}}=\mathbb{F}^{\mathcal{I},B,W^{\infty},D,\mathfrak{P}^{\infty}} 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 τ\tau is a 𝔽,B,W,D,𝔓\mathbb{F}^{\mathcal{I},B,W^{\infty},D,\mathfrak{P}^{\infty}}-totally inaccessible stopping time.

Proof.

In light of Remark 3.5 and Assumption (E2), it suffices to remark that τ\tau is independent from (,B,W,𝔓)(\mathcal{I},B,W^{\infty},\mathfrak{P}^{\infty}), which follows from the independence between τ\tau and (,B,WN,𝔓N)(\mathcal{I},B,W^{N},\mathfrak{P}^{N}). ∎

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 [0,η][0,\eta], 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 NN\to\infty 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 YY, denote by MtY=sup0stYsM^{Y}_{t}=\sup_{0\leq s\leq t}Y_{s} its running maximum. Recall from (1) and the price dynamics of the N-player game that the bubble trend function bb naturally depends on Fp(MtP)F_{p}(M^{P}_{t}), which is not Lipschitz in MPM^{P}. In general, the dynamics of asset price in the bubble phase is not well-posed. However, as the bubble is fueled by players’ entry, bb should be increasing in Fp(MtP)F_{p}(M^{P}_{t}), hence also increasing in MtPM^{P}_{t} at each time t[0,T]t\in[0,T] since FpF_{p} is a CDF. This monotonicity property of the path-dependent SDE (2) restores unique solvability.

Proposition A.1.

The following path-dependent SDE

Xt=x+0tb~(s,MsX,Xs)𝑑s+σ0BtX_{t}=x+\int_{0}^{t}\tilde{b}(s,M^{X}_{s},X_{s})ds+\sigma_{0}B_{t} (50)

has a unique strong solution satisfying 𝔼[X2]<\mathbb{E}[\lVert X\rVert_{\infty}^{2}]<\infty if for each fixed t[0,T]t\in[0,T]:

  1. (1)

    There exists C>0C>0 such that for all 𝒙C([0,T];)\boldsymbol{x}\in C([0,T];\mathbb{R}):

    |b~(t,Mt𝒙,𝒙t)|C(1+Mt|𝒙|).\left|\tilde{b}\left(t,M^{\boldsymbol{x}}_{t},\boldsymbol{x}_{t}\right)\right|\leq C\left(1+M^{|\boldsymbol{x}|}_{t}\right).
  2. (2)

    b~(t,,)\tilde{b}(t,\cdot,\cdot) 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 XX and YY are two solutions on the same probability space with respect to the same Brownian motion BB. Observing that XYX-Y is absolutely continuous, by Tanaka’s formula we get

XtYt=Yt+(XtYt)+=Yt+0t𝟙{Xs>Ys}d(XsYs)=x+σ0Bt+0t𝟙{Xs>Ys}b~(s,MsX,Xs)𝑑s+0t𝟙{XsYs}b~(s,MsY,Ys)𝑑s.YtXt=x+σ0Bt+0t𝟙{Ys>Xs}b~(s,MsY,Ys)𝑑s+0t𝟙{YsXs}b~(s,MsX,Xs)𝑑s.\begin{split}X_{t}\vee Y_{t}&=Y_{t}+(X_{t}-Y_{t})^{+}=Y_{t}+\int_{0}^{t}\mathds{1}_{\{X_{s}>Y_{s}\}}d(X_{s}-Y_{s})\\ &=x+\sigma_{0}B_{t}+\int_{0}^{t}\mathds{1}_{\{X_{s}>Y_{s}\}}\tilde{b}(s,M^{X}_{s},X_{s})ds+\int_{0}^{t}\mathds{1}_{\{X_{s}\leq Y_{s}\}}\tilde{b}(s,M^{Y}_{s},Y_{s})ds.\\ Y_{t}\vee X_{t}&=x+\sigma_{0}B_{t}+\int_{0}^{t}\mathds{1}_{\{Y_{s}>X_{s}\}}\tilde{b}(s,M^{Y}_{s},Y_{s})ds+\int_{0}^{t}\mathds{1}_{\{Y_{s}\leq X_{s}\}}\tilde{b}(s,M^{X}_{s},X_{s})ds.\end{split} (51)

We can equate the above expressions for all tt, implying that for almost every tt we have

𝟙{Xt=Yt}(b~(t,MtY,Yt)b~(t,MtX,Xt))=0.\mathds{1}_{\{X_{t}=Y_{t}\}}\left(\tilde{b}(t,M^{Y}_{t},Y_{t})-\tilde{b}(t,M^{X}_{t},X_{t})\right)=0. (52)

We now show that if Xs>YsX_{s}>Y_{s}, then MsXMsYM^{X}_{s}\geq M^{Y}_{s}. Define

s0sup{u[0,s]:Xu=Yu}.s_{0}\coloneqq\sup\{u\in[0,s]:X_{u}=Y_{u}\}.

The case is trivial if s0=0s_{0}=0.

On the event {s0>0}\{s_{0}>0\}, continuity of XX and YY implies that Xt>YtX_{t}>Y_{t} for all t(s0,s]t\in(s_{0},s]. Suppose MsX<MsYM^{X}_{s}<M^{Y}_{s}, then there must exist s[0,s0)s^{*}\in[0,s_{0}) where Ys=MsY>MsXXsY_{s^{*}}=M^{Y}_{s}>M^{X}_{s}\geq X_{s^{*}}. Then define

s1inf{u[s,s0]:Xu=Yu}.s_{1}\coloneqq\inf\{u\in[s^{*},s_{0}]:X_{u}=Y_{u}\}.

By continuity again, Yt>XtY_{t}>X_{t} for all t[s,s1)t\in[s^{*},s_{1}). By definition of ss^{*}, we must also have MtY>MtXM^{Y}_{t}>M^{X}_{t} for all t[s,s1)t\in[s^{*},s_{1}). Monotonicity of b~\tilde{b} leads to a contradiction

0>XsYs=ss1b~(t,MtY,Yt)b~(t,MtX,Xt)dt0.0>X_{s^{*}}-Y_{s^{*}}=\int_{s^{*}}^{s_{1}}\tilde{b}(t,M^{Y}_{t},Y_{t})-\tilde{b}(t,M^{X}_{t},X_{t})dt\geq 0.

Therefore, MsXMsYM^{X}_{s}\geq M^{Y}_{s} and in particular, MsXY=MsXM^{X\vee Y}_{s}=M^{X}_{s}. We can then rewrite (51) as

XtYt\displaystyle X_{t}\vee Y_{t} =x+σ0Bt+0tb~(s,MsXY,XsYs)𝑑s\displaystyle=x+\sigma_{0}B_{t}+\int_{0}^{t}\tilde{b}(s,M^{X\vee Y}_{s},X_{s}\vee Y_{s})ds
+0t𝟙{Xs=Ys}(b~(s,MsY,Ys)b~(s,MsXY,XsYs))𝑑s\displaystyle+\int_{0}^{t}\mathds{1}_{\{X_{s}=Y_{s}\}}\left(\tilde{b}(s,M^{Y}_{s},Y_{s})-\tilde{b}(s,M_{s}^{X\vee Y},X_{s}\vee Y_{s})\right)ds
=x+σ0Bt+0tb~(s,MsXY,XsYs)𝑑s\displaystyle=x+\sigma_{0}B_{t}+\int_{0}^{t}\tilde{b}(s,M^{X\vee Y}_{s},X_{s}\vee Y_{s})ds
+0t𝟙{{Xs=Ys}{MsX>MsY}}(b~(s,MsY,Ys)b~(s,MsXY,XsYs))𝑑s.\displaystyle+\int_{0}^{t}\mathds{1}_{\{\{X_{s}=Y_{s}\}\cap\{M^{X}_{s}>M^{Y}_{s}\}\}}\left(\tilde{b}(s,M^{Y}_{s},Y_{s})-\tilde{b}(s,M_{s}^{X\vee Y},X_{s}\vee Y_{s})\right)ds.

where the last line vanishes by (52). Therefore, XYX\vee Y also satisfies (50). Similarly, one can show XYX\wedge Y is also a solution. Then by uniqueness of law, we have 𝔼[|XY|]=𝔼[XYXY]=0\mathbb{E}[|X-Y|]=\mathbb{E}[X\vee Y-X\wedge Y]=0 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 (S,Σ,μ)(S,\Sigma,\mu) be a complete measurable space. A function f:S×f:S\times\mathbb{R}\to\mathbb{R} is jointly measurable if for all xx\in\mathbb{R}:

  1. (1)

    f(,x)f(\cdot,x) is measurable.

  2. (2)

    f(,xn)f(\cdot,x_{n}) converges to f(,x)f(\cdot,x) in μ\mu-measure for any increasing sequence xnxx_{n}\uparrow x.

Proof.

First let EE\subseteq\mathbb{R} be any closed set and let X={xm}m1X=\{x_{m}\}_{m\geq 1} be a countable, dense subset of \mathbb{R}. For ε>0\varepsilon>0, denote by 𝒪ϵ(E)\mathcal{O}_{\epsilon}(E) the open set {x:infeE|xe|<ε}\{x\in\mathbb{R}:\inf_{e\in E}|x-e|<\varepsilon\}. We claim that for μ\mu-almost every sSs\in S and any xx\in\mathbb{R}, f(s,x)Ef(s,x)\in E if and only if for each nn\in\mathbb{N}, there is xmX(x1n,x]x_{m}\in X\cap(x-\frac{1}{n},x] such that f(s,xm)𝒪1n(E)f(s,x_{m})\in\mathcal{O}_{\frac{1}{n}}(E). Note that we can always approximate any xx\in\mathbb{R} by an increasing sequence {xmk}k1\{x_{m_{k}}\}_{k\geq 1} with elements in XX such that the functions f(,xmk)f(\cdot,x_{m_{k}}) converge μ\mu-almost everywhere to f(,x)f(\cdot,x). The claim follows almost immediately. Denoting by f1f^{-1} the preimage of ff, joint measurability is proved by writing

f1(E)=n=1m=1{sS:f(s,xm)𝒪1n(E)}×[xm,xm+1n).f^{-1}(E)=\bigcap_{n=1}^{\infty}\bigcup_{m=1}^{\infty}\left\{s\in S:f(s,x_{m})\in\mathcal{O}_{\frac{1}{n}}(E)\right\}\times\left[x_{m},x_{m}+\frac{1}{n}\right).

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