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

Arbitrage Problems with Reflected Geometric Brownian Motion

Dean Buckner Kevin Dowd  and  Hardy Hulley Dean Buckner
The Eumaeus Project
London
United Kingdom
d.e.buckner@eumaeus.org Kevin Dowd
Durham University Business School
Mill Hill Lane
Durham DH1 3LB
United Kingdom
kevin.dowd@durham.ac.uk Hardy Hulley
Finance Discipline Group
University of Technology Sydney
P.O. Box 123
Broadway, NSW 2007
Australia
hardy.hulley@uts.edu.au
Abstract.

Contrary to the claims made by several authors, a financial market model in which the price of a risky security follows a reflected geometric Brownian motion is not arbitrage-free. In fact, such models violate even the weakest no-arbitrage condition considered in the literature. Consequently, they do not admit numéraire portfolios or equivalent risk-neutral probability measures, which makes them totally unsuitable for contingent claim valuation. Unsurprisingly, the published option pricing formulae for such models violate textbook no-arbitrage bounds.

We thank Johannes Ruf for valuable discussions that fine-tuned the arguments in this article. We also thank Claudio Fontana several insightful comments on a previous draft.

1. Introduction

In a seminal paper, Skorokhod (1961) studied the problem of inserting an instantaneously reflecting boundary into the state space of a one-dimensional Itô diffusion. The resulting process is described by an SDE that contains a new term—called a reflection term—that controls reflection off the boundary, along with the usual drift and diffusion terms. In addition to proving existence and uniqueness results for the solution to this SDE, Skorokhod (1961) derived an explicit expression for the reflection term. Today we recognise the reflection term as the local time of the diffusion at the reflecting boundary.

Several studies have considered financial market models in which the price of a risky security follows a reflected geometric Brownian motion (RGBM). Such a process is obtained by applying Skorokhod’s (1961) construction to a vanilla geometric Brownian motion, causing it to reflect off a lower boundary. For example, Veestraeten (2013), Hertrich (2015), Neuman and Schied (2016) and Hertrich and Zimmermann (2017) have used RGBMs to model exchange rates constrained by central bank target zone policies, while Gerber and Pafumi (2000) and Ko et al. (2010) modelled the value of an investment fund with a capital guarantee as an RGBM. Recently, Thomas (2021) modelled house prices as an RGBM, under the assumption that the government will support the property market if prices fall enough. Related models, where the price of a risky security is constrained by reflecting boundaries but the underlying dynamics is not geometric Brownian motion, have been studied by Feng et al. (2020) and Melnikov and Wan (2021).

Veestraeten (2008) claimed that the RGBM model does not offer any arbitrage opportunities. He justified this claim by noting that the security price follows a continuous process and that the time spent by it on the reflecting boundary has Lebesgue measure zero. Based on those observations, he reasoned that arbitrageurs cannot generate riskless profits by purchasing the security when its price reaches the boundary and selling it when its price is reflected off the boundary. On the strength of this heuristic argument, he concluded that the RGBM model is arbitrage-free and must therefore admit an equivalent risk-neutral probability measure. He then obtained an expression for the density of the putative equivalent risk-neutral probability measure and used it to derive pricing formulae for European puts and calls. Following an amendment to the put pricing formula by Hertrich and Veerstraeten (2013), and with some modifications to cater for dividends, these option pricing formulae have been used by Veestraeten (2013), Hertrich (2015), Hertrich and Zimmermann (2017) and Thomas (2021).

In this paper, we show that Veestraeten’s (2008) argument is unsound and his claim that the RGBM model is arbitrage-free is incorrect. In fact, we demonstrate that the model fails to satisfy even the weakest no-arbitrage condition considered in the literature. Consequently, it does not admit a numéraire portfolio or an equivalent risk-neutral probability measure. These deficiencies make the RGBM model unsuitable for contingent claim valuation and undermine the validity of the option pricing formulae in the articles cited above. In fact, those formulae are shown to behave quite pathologically when the price of the risky security is close to the reflecting boundary.

We begin, in Section 2, with a brief overview of a weak no-arbitrage condition, in the simplified setting of a financial market comprising a bank account and a single non-dividend-paying stock. The primary purpose of this overview is to serve as a roadmap for our subsequent analysis of the RGBM model. In particular, we derive a necessary and sufficient characterisation of the aforementioned no-arbitrage condition that subsequently allows us to pinpoint the exact source of the arbitrage trouble for the RGBM model.

We analyse the RGBM model in Section 3. After setting it up and cataloguing its basic properties, we construct a trading strategy that violates the weak no-arbitrage condition studied in Section 2, and explain how this result is intimately related to the reflecting behaviour of the stock price in the model. The weak no-arbitrage failure of the RGBM model means that it does not admit an equivalent risk-neutral probability measure, invalidating the risk-neutral approach to contingent claim valuation. Consequently, the published option pricing formulae for the model have no theoretical justification. We examine those formulae in detail, demonstrating that they violate textbook no-arbitrage bounds, and we explain those violations in terms of the reflecting behaviour of the stock price.

In summary, this paper makes two significant contributions. First, it establishes that the RGBM model provides an interesting and non-trivial example of a financial market model that violates a weak no-arbitrage condition, and relates this failure to the characteristics of the model.111Our analysis could potentially be extended to other financial market models with reflecting boundaries, such as the models studied by Feng et al. (2020) and Melnikov and Wan (2021). The second contribution is to highlight the erroneous reasoning about the arbitrage properties of the RGBM model in Veestraeten (2008), which has led to the publication of invalid option pricing formulae by Veestraeten (2008, 2013), Hertrich and Veerstraeten (2013), Hertrich (2015), Hertrich and Zimmermann (2017) and Thomas (2021). Of particular concern is Thomas (2021), which recommends the RGBM model to practitioners as a valuation framework for no-negative equity guarantees, a large and important class of insurance products.

2. A Weak No-Arbitrage Condition and Its Consequences

This section develops the theoretical prerequisites for our analysis of the reflected geometric Brownian motion model in Section 3. We begin by introducing a class of continuous financial market models that is general enough to encompass the reflected geometric Brownian motion model as a particular example. In the context of this modelling framework, we then formulate the so-called structure condition, which is expressed in terms of the characteristics of a model, before introducing a weak no-arbitrage condition, known as no increasing profit. Our main result establishes that the structure condition and the no-increasing profit condition are equivalent in our setting. Finally, we summarise the economic and modelling consequences of a failure of the no increasing profit condition.

2.1. A General Modelling Framework

Throughout this paper, we assume the existence of a filtered probability space (Ω,,𝔉,P)(\Omega,\mathscr{F},\mathfrak{F},\textsf{P}), whose filtration 𝔉=(t)t0\mathfrak{F}=(\mathscr{F}_{t})_{t\geq 0} satisfies the usual conditions of completeness and right-continuity. It is understood that all random variables and stochastic processes are defined on this space and all stochastic processes are adapted to its filtration. Given a continuous semimartingale XX, we shall write 𝖫(X)\mathsf{L}(X) to denote the family of predictable processes φ\varphi, such that the stochastic integral 0φsdXs\int_{0}^{\cdot}\varphi_{s}\,\mathrm{d}X_{s} exists.

Consider a financial market comprising a risk-free security and a single risky security. For convenience, we shall refer to the former as a bank account and to the latter as a stock. The value BB of the bank account is given by BtertB_{t}\coloneqq\mathrm{e}^{rt}, for all t0t\geq 0, where r0r\geq 0 is a continuously compounding risk-free interest rate. The stock price SS is determined by the SDE

dSt=StdAt+StdMt\mathrm{d}S_{t}=S_{t}\,\mathrm{d}A_{t}+S_{t}\,\mathrm{d}M_{t} (2.1)

for all t0t\geq 0, with initial value S0>0S_{0}>0, where AA is a continuous finite variation process and MM is a continuous local martingale. For convenience, we shall assume that Mt<\langle M\rangle_{t}<\infty, for all t0t\geq 0, which ensures that the stock price is strictly positive, by virtue of the law of large numbers for local martingales (see Revuz and Yor 1999, Exercise V.1.16).

Investors in the market described above are able to trade self-financing portfolios comprising the bank account and the stock. The following definition makes this concept precise.

Definition 2.1.

A trading strategy is a predictable process ξ𝖫(S)\xi\in\mathsf{L}(S). Given a trading strategy ξ𝖫(S)\xi\in\mathsf{L}(S) and an initial endowment v0v\geq 0, the stochastic process Vv,ξV^{v,\xi}, determined by the SDE

dVtv,ξ=Vtv,ξξtStBtdBt+ξtdSt=r(Vtv,ξξtSt)dt+ξtStdAt+ξtStdMt,\mathrm{d}V^{v,\xi}_{t}=\frac{V^{v,\xi}_{t}-\xi_{t}S_{t}}{B_{t}}\,\mathrm{d}B_{t}+\xi_{t}\,\mathrm{d}S_{t}=r(V^{v,\xi}_{t}-\xi_{t}S_{t})\,\mathrm{d}t+\xi_{t}S_{t}\,\mathrm{d}A_{t}+\xi_{t}S_{t}\,\mathrm{d}M_{t}, (2.2)

for all t0t\geq 0, with initial value V0v,ξ=vV^{v,\xi}_{0}=v, is called the portfolio value generated by vv and ξ\xi.

Given a trading strategy ξ𝖫(S)\xi\in\mathsf{L}(S) and an initial endowment v0v\geq 0, we interpret Vtv,ξV^{v,\xi}_{t} as the value at time t0t\geq 0 of a self-financing portfolio that holds ξs\xi_{s} shares of the stock at each time s[0,t]s\in[0,t], with initial value V0v,ξ=vV^{v,\xi}_{0}=v. The SDE (2.2) follows from the self-financing requirement that the portfolio must hold (Vtv,ξξtSt)/Bt\nicefrac{{(V^{v,\xi}_{t}-\xi_{t}S_{t})}}{{B_{t}}} units of the bank account at time t0t\geq 0, if it holds ξt\xi_{t} shares of the stock at that time.

2.2. The Structure Condition

Following Jacod and Shiryaev (2003, Proposition II.2.9), there exists a continuous (and hence predictable) increasing process GG, such that

rt=0tρsdGs,At=0tbsdGsandMt=0tas2dGs,rt=\int_{0}^{t}\rho_{s}\,\mathrm{d}G_{s},\qquad A_{t}=\int_{0}^{t}b_{s}\,\mathrm{d}G_{s}\qquad\text{and}\qquad\langle M\rangle_{t}=\int_{0}^{t}a_{s}^{2}\,\mathrm{d}G_{s}, (2.3)

for all t0t\geq 0, where ρ\rho, bb and aa are some predictable processes. One possible choice for GG is obtained by setting Gtt+Var(A)t+MtG_{t}\coloneqq t+\textsf{Var}(A)_{t}+\langle M\rangle_{t}, for all t0t\geq 0. The next definition formulates the so-called structure condition in terms of the processes ρ\rho, bb and aa.

Definition 2.2.

The financial market satisfies the structure condition if there is a predictable process ϑ\vartheta, such that

bt(ω)ρt(ω)=ϑt(ω)at2(ω),b_{t}(\omega)-\rho_{t}(\omega)=\vartheta_{t}(\omega)a_{t}^{2}(\omega), (2.4)

for PG-a.a(ω,t)Ω×+\textsf{P}\otimes G\text{-a.a}\;(\omega,t)\in\Omega\times\mathbb{R}_{+}.

The structure condition was first identified by Schweizer (1995), who formulated it somewhat differently. To demonstrate that the two formulations are equivalent, let S^\hat{S} denote the discounted stock price, defined by S^tSt/Bt=ertSt\hat{S}_{t}\coloneqq\nicefrac{{S_{t}}}{{B_{t}}}=\mathrm{e}^{-rt}S_{t}, for all t0t\geq 0. It satisfies the SDE

dS^t=rS^tdt+S^tdAt+S^tdMt,\mathrm{d}\hat{S}_{t}=-r\hat{S}_{t}\,\mathrm{d}t+\hat{S}_{t}\,\mathrm{d}A_{t}+\hat{S}_{t}\,\mathrm{d}M_{t},

for all t0t\geq 0, with S^0=S0>0\hat{S}_{0}=S_{0}>0, by an application of Itô’s formula. Next, define the continuous finite variation process A^\hat{A} and the continuous local martingale M^\hat{M}, by setting

A^t0trS^udu+0tS^udAuandM^t0tS^udMu,\hat{A}_{t}\coloneqq-\int_{0}^{t}r\hat{S}_{u}\,\mathrm{d}u+\int_{0}^{t}\hat{S}_{u}\,\mathrm{d}A_{u}\qquad\text{and}\qquad\hat{M}_{t}\coloneqq\int_{0}^{t}\hat{S}_{u}\,\mathrm{d}M_{u}, (2.5)

for all t0t\geq 0. If the structure condition holds, for some predictable process ϑ\vartheta satisfying (2.4), then

dA^t(ω)=rS^t(ω)dt+S^t(ω)dAt(ω)=(bt(ω)ρt(ω))S^t(ω)dGt(ω)=ϑt(ω)at2(ω)S^t(ω)dGt(ω)=ϑt(ω)S^t(ω)dMt(ω)=ϑt(ω)S^t(ω)dM^t(ω),\begin{split}\mathrm{d}\hat{A}_{t}(\omega)=-r\hat{S}_{t}(\omega)\,\mathrm{d}t+\hat{S}_{t}(\omega)\,\mathrm{d}A_{t}(\omega)&=\bigl{(}b_{t}(\omega)-\rho_{t}(\omega)\bigr{)}\hat{S}_{t}(\omega)\,\mathrm{d}G_{t}(\omega)\\ &=\vartheta_{t}(\omega)a_{t}^{2}(\omega)\hat{S}_{t}(\omega)\,\mathrm{d}G_{t}(\omega)\\ &=\vartheta_{t}(\omega)\hat{S}_{t}(\omega)\,\mathrm{d}\langle M\rangle_{t}(\omega)\\ &=\frac{\vartheta_{t}(\omega)}{\hat{S}_{t}(\omega)}\,\mathrm{d}\langle\hat{M}\rangle_{t}(\omega),\end{split}

for PG-a.a(ω,t)Ω×+\textsf{P}\otimes G\text{-a.a}\;(\omega,t)\in\Omega\times\mathbb{R}_{+}. On the other hand, if there is a predictable process λ\lambda, such that

dA^t(ω)=λt(ω)dM^t(ω),\mathrm{d}\hat{A}_{t}(\omega)=\lambda_{t}(\omega)\,\mathrm{d}\langle\hat{M}\rangle_{t}(\omega), (2.6)

for PG-a.a(ω,t)Ω×+\textsf{P}\otimes G\text{-a.a}\;(\omega,t)\in\Omega\times\mathbb{R}_{+}, then the same reasoning shows that the structure condition holds, with the predictable process ϑλS^\vartheta\coloneqq\lambda\hat{S} satisfying (2.4). So, the structure condition is equivalent to the existence of a predictable process λ\lambda that satisfies (2.6), which is how Schweizer (1995) originally formulated it.

The reformulation of the structure condition above offers a useful insight that will help us interpret the arbitrage properties of the reflected geometric Brownian motion model, studied in Section 3. First, note that the finite variation processes A^\hat{A} and M^\langle\hat{M}\rangle may be regarded as (possibly signed) random measures on (+,(+))(\mathbb{R}_{+},\mathscr{B}(\mathbb{R}_{+})). With that interpretation in mind, (2.6) states that A^M^\hat{A}\ll\langle\hat{M}\rangle, with predictable density λ\lambda. This means that if 0𝟏UdM^t(ω)=0\int_{0}^{\infty}\mathbf{1}_{U}\,\mathrm{d}\langle\hat{M}\rangle_{t}(\omega)=0 then 0𝟏UdA^t(ω)=0\int_{0}^{\infty}\mathbf{1}_{U}\,\mathrm{d}\hat{A}_{t}(\omega)=0, for P-a.a.ωΩ\textsf{P}\text{-a.a.}\;\omega\in\Omega and any Borel-measurable set U(+)U\in\mathscr{B}(\mathbb{R}_{+}). In particular, if there are measurable subsets of +\mathbb{R}_{+} over which the sample paths of A^\hat{A} increase or decrease but the sample paths of M^\langle\hat{M}\rangle remain constant, then the structure condition cannot hold.

2.3. The No Increasing Profit Condition

An increasing profit is the strongest form of arbitrage for continuous-time financial market models. This concept of arbitrage was introduced by Karatzas and Kardaras (2007) and thoroughly investigated by Fontana (2015). The following definition provides a formulation that is appropriate for our setting.

Definition 2.3.

The financial market admits an increasing profit if there is a trading strategy ξ𝖫(S)\xi\in\mathsf{L}(S), such that

  1. (1)

    V0,ξV^{0,\xi} is non-decreasing; and

  2. (2)

    P(V0,ξ>0)>0\textsf{P}(V^{0,\xi}_{\infty-}>0)>0.

The market satisfies the no increasing profit (NIP) condition if no such strategy exists.

Putting the previous definition into words, a model admits an increasing profit if there is a trading strategy, for which the value of the associated portfolio with an initial endowment of zero is a non-decreasing process with a positive probability of ultimately becoming strictly positive. Such arbitrage opportunities are unequivocally pathological, since they offer investors a chance of making something from nothing, without incurring any risk. Equilibrium asset prices cannot exist in a model that offers such opportunities, since there would be insatiable demand for the portfolios that exploit them. Consequently, a viable financial market model must satisfy the NIP condition.

The next theorem reveals that the NIP condition is equivalent to the structure condition. Fontana (2015, Theorem 3.1) proved essentially the same result, but the formulation and proof in our setting are quite different and the proof is sufficiently instructive to merit inclusion. Karatzas and Shreve (1991, Lemma 1.4.6) and Delbaen and Schachermayer (1995, Theorem 3.5) proved related results in the “only iff” direction.

Theorem 2.4.

The NIP condition is satisfied if and only if the structure condition is satisfied.

Proof.

(\Rightarrow) Suppose the NIP condition holds. Consider a portfolio with initial endowment v=1v=1 that implements a trading strategy ξ𝖫(S)\xi\in\mathsf{L}(S), specified by

ξt𝟏{at=0}(𝟏{bt>ρt}𝟏{bt<ρt})Vt1,ξSt,\xi_{t}\coloneqq\mathbf{1}_{\{a_{t}=0\}}\bigl{(}\mathbf{1}_{\{b_{t}>\rho_{t}\}}-\mathbf{1}_{\{b_{t}<\rho_{t}\}}\bigr{)}\frac{V^{1,\xi}_{t}}{S_{t}},

for all t0t\geq 0. In that case, the quadratic variation of the local martingale 0ξuSudMu\int_{0}^{\cdot}\xi_{u}S_{u}\,\mathrm{d}M_{u} satisfies

0ξuSudMut=0tξu2Su2dMu=0tξu2au2Su2dGu=0t𝟏{au=0}(𝟏{bu>ρu}𝟏{bu<ρu})2au2(Vu1,ξ)2𝑑Gu=0,\begin{split}\biggl{\langle}\int_{0}^{\cdot}\xi_{u}S_{u}\,\mathrm{d}M_{u}\biggr{\rangle}_{t}=\int_{0}^{t}\xi_{u}^{2}S_{u}^{2}\,\mathrm{d}\langle M\rangle_{u}&=\int_{0}^{t}\xi_{u}^{2}a_{u}^{2}S_{u}^{2}\,\mathrm{d}G_{u}\\ &=\int_{0}^{t}\mathbf{1}_{\{a_{u}=0\}}\bigl{(}\mathbf{1}_{\{b_{u}>\rho_{u}\}}-\mathbf{1}_{\{b_{u}<\rho_{u}\}}\bigr{)}^{2}a_{u}^{2}(V^{1,\xi}_{u})^{2}\,dG_{u}=0,\end{split}

for all t0t\geq 0. This implies that 0ξuSudMu=0\int_{0}^{\cdot}\xi_{u}S_{u}\,\mathrm{d}M_{u}=0, since a local martingale whose quadratic variation is zero must remain constant. Using differential notation, the latter condition can be expressed as ξtStdMt=0\xi_{t}S_{t}\,\mathrm{d}M_{t}=0, for all t0t\geq 0. Hence, (2.2) gives

dVt1,ξ=r(Vt1,ξξtSt)dt+ξtStdAt=rVt1,ξdt+ξt(btρt)StdGt=rVt1,ξdt+𝟏{at=0}(𝟏{bt>ρt}𝟏{bt<ρt})(btρt)Vt1,ξdGt=rVt1,ξdt+𝟏{at=0andbtρt}|btρt|Vt1,ξdGt,\begin{split}\mathrm{d}V^{1,\xi}_{t}&=r(V^{1,\xi}_{t}-\xi_{t}S_{t})\,\mathrm{d}t+\xi_{t}S_{t}\,\mathrm{d}A_{t}\\ &=rV^{1,\xi}_{t}\,\mathrm{d}t+\xi_{t}(b_{t}-\rho_{t})S_{t}\,\mathrm{d}G_{t}\\ &=rV^{1,\xi}_{t}\,\mathrm{d}t+\mathbf{1}_{\{a_{t}=0\}}\bigl{(}\mathbf{1}_{\{b_{t}>\rho_{t}\}}-\mathbf{1}_{\{b_{t}<\rho_{t}\}}\bigr{)}(b_{t}-\rho_{t})V^{1,\xi}_{t}\,\mathrm{d}G_{t}\\ &=rV^{1,\xi}_{t}\,\mathrm{d}t+\mathbf{1}_{\{a_{t}=0\>\text{and}\>b_{t}\neq\rho_{t}\}}|b_{t}-\rho_{t}|V^{1,\xi}_{t}\,\mathrm{d}G_{t},\end{split}

for all t0t\geq 0. The solution to this equation is

Vt1,ξ=exp(rt+0t𝟏{as=0andbsρs}|bsρs|dGs),V^{1,\xi}_{t}=\exp\biggl{(}rt+\int_{0}^{t}\mathbf{1}_{\{a_{s}=0\;\text{and}\;b_{s}\neq\rho_{s}\}}|b_{s}-\rho_{s}|\,\mathrm{d}G_{s}\biggr{)},

for all t0t\geq 0. Now consider another trading strategy ξ~𝖫(S)\tilde{\xi}\in\mathsf{L}(S), with initial endowment v~=0\tilde{v}=0, comprising a long position in ξ\xi and a short position in the bank account. The value of the associated portfolio is given by

Vt0,ξ~=Vt1,ξBt=ert(exp(0t𝟏{as=0andbsρs}|bsρs|dGs)1),V^{0,\tilde{\xi}}_{t}=V^{1,\xi}_{t}-B_{t}=e^{rt}\biggl{(}\exp\biggl{(}\int_{0}^{t}\mathbf{1}_{\{a_{s}=0\;\text{and}\;b_{s}\neq\rho_{s}\}}|b_{s}-\rho_{s}|\,\mathrm{d}G_{s}\biggr{)}-1\biggr{)},

for all t0t\geq 0. Since the integrand in this expression is non-negative, it follows that V0,ξ~V^{0,\tilde{\xi}} is a non-decreasing process. This implies that V0,ξ~=0V^{0,\tilde{\xi}}_{\infty-}=0, by virtue of the assumption that the NIP condition holds, which is to say that

0𝟏{at=0andbtρt}|btρt|dGt=0.\int_{0}^{\infty}\mathbf{1}_{\{a_{t}=0\;\text{and}\;b_{t}\neq\rho_{t}\}}|b_{t}-\rho_{t}|\,\mathrm{d}G_{t}=0.

From this we conclude that bt(ω)=ρt(ω)b_{t}(\omega)=\rho_{t}(\omega), for PG-a.a(ω,t){a()=0}\textsf{P}\otimes G\text{-a.a}\;(\omega,t)\in\{a_{\cdot}(\,\cdot\,)=0\}, whence the predictable process ϑ\vartheta, defined by

ϑt𝟏{at0}btρtat2,\vartheta_{t}\coloneqq\mathbf{1}_{\{a_{t}\neq 0\}}\frac{b_{t}-\rho_{t}}{a_{t}^{2}},

for all t0t\geq 0, satisfies (2.4) and verifies the structure condition.
(\Leftarrow) Suppose the structure condition is satisfied and let ξ𝖫(S)\xi\in\mathsf{L}(S) be a trading strategy, for which the value V0,ξV^{0,\xi} of the associated portfolio with zero initial endowment is non-decreasing. It follows that V0,ξV^{0,\xi} is a finite variation process, which implies that

V0,ξt=0ξuSudMut=0tξu2Su2dMu=0tξu2Su2au2dGu=0,\langle V^{0,\xi}\rangle_{t}=\biggl{\langle}\int_{0}^{\cdot}\xi_{u}S_{u}\,\mathrm{d}M_{u}\biggr{\rangle}_{t}=\int_{0}^{t}\xi_{u}^{2}S_{u}^{2}\,\mathrm{d}\langle M\rangle_{u}=\int_{0}^{t}\xi_{u}^{2}S_{u}^{2}a_{u}^{2}\,\mathrm{d}G_{u}=0,

for all t0t\geq 0, whence at(ω)=0a_{t}(\omega)=0, for PG(ω,t){ξ()0}\textsf{P}\otimes G\;(\omega,t)\in\{\xi_{\cdot}(\,\cdot\,)\neq 0\}. Since the structure condition holds, by assumption, it follows from (2.4) that bt(ω)=ρt(ω)b_{t}(\omega)=\rho_{t}(\omega), for PG-a.a(ω,t){ξ()0}\textsf{P}\otimes G\text{-a.a}\;(\omega,t)\in\{\xi_{\cdot}(\,\cdot\,)\neq 0\}. Next, observe that 0ξuSudMu=0\langle\int_{0}^{\cdot}\xi_{u}S_{u}\,\mathrm{d}M_{u}\rangle=0 also implies that 0ξuSudMu=0\int_{0}^{\cdot}\xi_{u}S_{u}\,\mathrm{d}M_{u}=0, since a local martingale with zero quadratic variation remains constant. This can be expressed as ξtStdMt=0\xi_{t}S_{t}\,\mathrm{d}M_{t}=0, for all t0t\geq 0. Consequently, (2.2) gives

dVt0,ξ=r(Vt0,ξξtSt)dt+ξtStdAt=rVt0,ξdt+ξtSt(btρt)dGt,\mathrm{d}V^{0,\xi}_{t}=r(V^{0,\xi}_{t}-\xi_{t}S_{t})\,\mathrm{d}t+\xi_{t}S_{t}\,\mathrm{d}A_{t}=rV^{0,\xi}_{t}\,\mathrm{d}t+\xi_{t}S_{t}(b_{t}-\rho_{t})\,\mathrm{d}G_{t},

for all t0t\geq 0. Since we have established that bt(ω)=ρt(ω)b_{t}(\omega)=\rho_{t}(\omega), for PG-a.a(ω,t){ξ()0}\textsf{P}\otimes G\text{-a.a}\;(\omega,t)\in\{\xi_{\cdot}(\,\cdot\,)\neq 0\}, it follows from the expression above that dVt0,ξ=rVt0,ξdt\mathrm{d}V^{0,\xi}_{t}=rV^{0,\xi}_{t}\,\mathrm{d}t, for all t0t\geq 0, whence Vt0,ξ=V00,ξert=0V^{0,\xi}_{t}=V^{0,\xi}_{0}\mathrm{e}^{rt}=0, since V00,ξ=0V^{0,\xi}_{0}=0. Consequently, V0,ξ=0V^{0,\xi}_{\infty-}=0, implying that P(V0,ξ>0)=0\textsf{P}(V^{0,\xi}_{\infty-}>0)=0. In summary, we have demonstrated that any trading strategy that satisfies the first condition in Definition 2.3 cannot satisfy the second condition. Hence, the NIP condition holds. ∎

Inspection of (2.4) reveals that the structure condition fails if and only if there are periods during which the process aa is zero and processes bb and ρ\rho assume different values. Whenever that happens, the market effectively contains two risk-free securities with different rates of return. During such periods, the trading strategy constructed in the first part of the proof of Thoerem 2.4 generates an increasing profit by borrowing at the lower rate and investing the proceeds at the higher rate. In light of Theorem 2.4, the NIP condition may be interpreted economically as stating that the market never contains two risk-free securities with different rates of return.

2.4. Contingent Claim Valuation

In order for a model to be suitable for contingent claim valuation, it should admit either a numéraire portfolio or an equivalent risk-neutral probability measure. Informally, a numéraire portfolio in our modelling framework is a well-behaved trading strategy ξ\xi^{*}, whose portfolio value V1,ξV^{1,\xi^{*}}, with initial endowment v=1v=1, provides a natural benchmark against which the value of every other well-behaved portfolio can be measured. More formally, the benchmarked processes B/V1,ξ\nicefrac{{B}}{{V^{1,\xi^{*}}}} and S/V1,ξ\nicefrac{{S}}{{V^{1,\xi^{*}}}} are required to be supermartingales. On the other hand, an equivalent risk-neutral probability measure in our setting is an equivalent probability measure QP\textsf{{Q}}\sim\textsf{P}, such that the discounted stock price S^S/B\hat{S}\coloneqq\nicefrac{{S}}{{B}} is a local martingale under Q. If a model admits a numéraire portfolio, then the benchmark approach to contingent claim valuation (see Platen and Heath 2006) can be used, while traditional risk-neutral valuation can be used if it admits an equivalent risk-neutral probability measure.

Both the existence of well-behaved numéraire portfolio and the existence of an equivalent risk-neutral probability measure can be characterised in terms of no-arbitrage conditions. In more general settings than ours, Karatzas and Kardaras (2007) established that the existence of a well-behaved numéraire portfolio is equivalent to the no unbounded profit with bounded risk (NUPBR) condition, while Delbaen and Schachermayer (1994) famously demonstrated that the no free lunch with vanishing risk (NFLVR) condition is necessary and sufficient for the existence of an equivalent risk-neutral probability measure. Based on those results, a model that does not satisfy NUPBR or NFLVR cannot be used for pricing options and other contingent claims.

Crucially, both NUPBR and NFLVR are stronger conditions that NIP.222The logical dependencies between the various no-arbitrage conditions for continuous financial market models are explained in the surveys by Hulley (2010, Chapter 1) and Fontana (2015). Consequently, a model that does not satisfy the NIP condition will fail to satisfy the NUPBR and NFLVR conditions as well, which means that it will not admit a well-behaved numéraire portfolio or an equivalent risk-neutral probability measure. Such a model is unsuitable for contingent claim valuation. This is unsurprising, since we have already argued that a model cannot sustain equilibrium asset prices if it does not satisfy the NIP condition.

3. A Reflected Geometric Brownian Motion Model

Veestraeten (2008) proposed a modification of the Black and Scholes (1973) model, in which the price of a risky asset is a geometric Brownian with a reflecting lower boundary. A related model was employed by Gerber and Pafumi (2000), and Ko et al. (2010) for the value of an investment fund with a capital guarantee, while Veestraeten (2013), Neuman and Schied (2016), Hertrich (2015), and Hertrich and Zimmermann (2017) used similar models for exchange rates constrained by target zones. Finally, Thomas (2021) recently used reflected geometric Brownian motion to model house prices protected by a government guarantee. This section provides a careful analysis of the arbitrage properties of the reflected geometric Brownian motion model. We begin with a rigorous formulation of the model, before describing its basic properties. Next, we demonstrate that it violates both the no increasing profit condition and the structure condition, before analysing those failures in terms of the reflecting behaviour of the asset price in the model. Finally, we demonstrate that the option pricing formulae for the model, presented in several of the previously cited studies, exhibit numerous pathologies, which are once again related to the reflecting behaviour of the asset price in the model.

3.1. Reflected Geometric Brownian Motion

Skorokhod (1961) showed that the SDE

dSt=μStdt+σStdWt+dLt,\mathrm{d}S_{t}=\mu S_{t}\,\mathrm{d}t+\sigma S_{t}\,\mathrm{d}W_{t}+\mathrm{d}L_{t}, (3.1)

for all t0t\geq 0, with a lower reflecting boundary b>0b>0 and initial value S0>bS_{0}>b, admits a unique strong solution. This solution comprises a pair of processes SS and LL, such that

Stb;\displaystyle S_{t}\geq b; (3.2a)
L is non-decreasing with L0=0;\displaystyle\text{$L$ is non-decreasing with $L_{0}=0$}; (3.2b)
0t𝟏{Su>b}dLu=0;and\displaystyle\int_{0}^{t}\mathbf{1}_{\{S_{u}>b\}}\mathrm{d}L_{u}=0;\qquad\text{and} (3.2c)
St=S0+0tμSudu+0tσSudWu+Lt,\displaystyle S_{t}=S_{0}+\int_{0}^{t}\mu S_{u}\,\mathrm{d}u+\int_{0}^{t}\sigma S_{u}\,\mathrm{d}W_{u}+L_{t}, (3.2d)

for all t0t\geq 0, and where the integrals above are well-defined. The process SS is called a reflected geometric Brownian motion (RGBM) and the process LL is called the reflection term.

Condition (3.2c) ensures that the value of LL does not change while the value of SS exceeds bb, in which case condition (3.2d) ensures that SS behaves like a vanilla geometric Brownian motion (GBM). However, as soon as SS reaches the boundary bb, the value of LL increases, instantaneously reflecting SS back into the interval (b,)(b,\infty), after which it behaves like a vanilla GBM once again. It follows that the points of increase of LL are limited to times when SS visits bb. It is natural to interpret LtL_{t} as the cumulative amount of reflection of SS from the boundary bb, up time t0t\geq 0.

In addition to establishing the existence of a unique solution to (3.1), Skorokhod (1961) showed that the reflection term is given by

Lt=π80t𝟏{Su=b}σdu,L_{t}=\sqrt[\ ]{\frac{\pi}{8}}\int_{0}^{t}\mathbf{1}_{\{S_{u}=b\}}\sigma\,\sqrt[\ ]{\mathrm{d}u},

for all t0t\geq 0, where the integral in this expression can be defined rigorously as a limit of integral sums. There is, however, a more useful interpretation of the reflection term. Define the process b\ell^{b}, by setting

tblimε01ε0t𝟏(b,b+ε](Su)dSu=limε01ε0t𝟏{b<Sub+ε}σ2Su2du,\ell^{b}_{t}\coloneqq\lim_{\varepsilon\downarrow 0}\frac{1}{\varepsilon}\int_{0}^{t}\mathbf{1}_{(b,b+\varepsilon]}(S_{u})\,\mathrm{d}\langle S\rangle_{u}=\lim_{\varepsilon\downarrow 0}\frac{1}{\varepsilon}\int_{0}^{t}\mathbf{1}_{\{b<S_{u}\leq b+\varepsilon\}}\sigma^{2}S_{u}^{2}\,\mathrm{d}u, (3.3)

for all t0t\geq 0. This process, which is known as the local time of SS at bb, provides a non-trivial measure of the amount of time SS spends in the vicinity of bb. Pilipenko (2014, Theorem 1.3.1) showed that Lt=tb/2L_{t}=\nicefrac{{\ell^{b}_{t}}}{{2}}, for all t0t\geq 0, which is to say that the reflection term is just the scaled local time at the reflecting boundary.

3.2. Model Specification

We shall now revisit the financial market described in Section 2, comprising a risk-free bank account and a non-dividend-paying stock. As before, the value BB of the bank account is given by BtertB_{t}\coloneqq\mathrm{e}^{rt}, for all t0t\geq 0, where r0r\geq 0 is the risk-free interest rate. However, following Veestraeten (2008), we now assume that the stock price SS follows the RGBM (3.1).

We begin by noting that the RGBM model falls within the scope of the framework introduced in Section 2, which ensures that the analysis in that section is fully applicable to it. Indeed, (3.2c) allows us to rewrite (3.1) as

dSt=μStdt+σStdWt+StbdLt,\mathrm{d}S_{t}=\mu S_{t}\,\mathrm{d}t+\sigma S_{t}\,\mathrm{d}W_{t}+\frac{S_{t}}{b}\,\mathrm{d}L_{t},

for all t0t\geq 0. This equation may in turn be expressed in the form (2.1), with the continuous finite variation process AA and the continuous local martingale MM given by

Atμt+LtbandMt0tσdWs,A_{t}\coloneqq\mu t+\frac{L_{t}}{b}\qquad\text{and}\qquad M_{t}\coloneqq\int_{0}^{t}\sigma\,\mathrm{d}W_{s}, (3.4)

for all t0t\geq 0. As noted in Section 2, there exists a continuous increasing process GG satisfying (2.3), for some predictable processes ρ\rho, bb and aa. That is to say,

rdt=ρtdGt,μdt+dLtb=btdGtandσ2dt=at2dGt,r\,\mathrm{d}t=\rho_{t}\,\mathrm{d}G_{t},\qquad\mu\,\mathrm{d}t+\frac{\mathrm{d}L_{t}}{b}=b_{t}\,\mathrm{d}G_{t}\qquad\text{and}\qquad\sigma^{2}\,\mathrm{d}t=a_{t}^{2}\,\mathrm{d}G_{t}, (3.5)

for all t0t\geq 0. Setting Gtt+tbG_{t}\coloneqq t+\ell^{b}_{t}, for all t0t\geq 0, provides a convenient choice for the process GG.

3.3. Failure of the No Increasing Profit Condition

The behaviour of the stock price at the reflecting boundary raises concerns about the arbitrage properties of the RGBM model. Indeed, it seems plausible that an arbitrageur could exploit this behaviour by purchasing the stock when its price reaches the boundary and unwinding the position immediately afterwards, as the stock price is reflected off the boundary. The guiding intuition is that, upon reaching the boundary, the behaviour of the stock price is completely predictable over the next instant of time, and therefore arbitrageable. The following proposition confirms this intuition.333The trading strategy constructed in the proof of Proposition 3.1 is similar in spirit to the trading strategy in Fontana (2015, Example 7.1).

Proposition 3.1.

The RGBM model violates the NIP condition.

Proof.

Define the process ξ\xi, by setting ξt𝟏{St=b}\xi_{t}\coloneqq\mathbf{1}_{\{S_{t}=b\}}, for all t0t\geq 0. Then ξ\xi is predictable, since it is obtained by applying the measurable function 𝟏{=b}:\mathbf{1}_{\{\,\cdot\,=b\}}:\mathbb{R}\rightarrow\mathbb{R} to the continuous (and hence predictable) process SS. Moreover, since every bounded predictable process is integrable with respect to any semimartingale, it follows that ξ𝖫(S)\xi\in\mathsf{L}(S), which establishes that ξ\xi is a valid trading strategy. Now, since leb{t+|St(ω)=b}=0\operatorname{leb}\{t\in\mathbb{R}_{+}\,|\,S_{t}(\omega)=b\}=0, for P-a.a.ωΩ\textsf{P}\text{-a.a.}\;\omega\in\Omega, the properties of the Lebesgue and Itô integrals and the local time process ensure that

ξtStBtdBt\displaystyle\frac{\xi_{t}S_{t}}{B_{t}}\,\mathrm{d}B_{t} =𝟏{St=b}rStdt=0\displaystyle=\mathbf{1}_{\{S_{t}=b\}}rS_{t}\,\mathrm{d}t=0
and
ξtdSt\displaystyle\xi_{t}\,\mathrm{d}S_{t} =𝟏{St=b}μStdt+𝟏{St=b}σStdWt+𝟏{St=b}dLt=dLt,\displaystyle=\mathbf{1}_{\{S_{t}=b\}}\mu S_{t}\,\mathrm{d}t+\mathbf{1}_{\{S_{t}=b\}}\sigma S_{t}\,\mathrm{d}W_{t}+\mathbf{1}_{\{S_{t}=b\}}\,\mathrm{d}L_{t}=\mathrm{d}L_{t},

for all t0t\geq 0. By substituting these identities into (2.2), we see that the value V0,ξV^{0,\xi} of a portfolio that implements the strategy ξ\xi, with zero initial endowment, satisfies the SDE

dVt0,ξ=Vt0,ξξtStBtdBt+ξtdSt=rVt0,ξdt+dLt,\mathrm{d}V^{0,\xi}_{t}=\frac{V^{0,\xi}_{t}-\xi_{t}S_{t}}{B_{t}}\,\mathrm{d}B_{t}+\xi_{t}\,\mathrm{d}S_{t}=rV^{0,\xi}_{t}\,\mathrm{d}t+\mathrm{d}L_{t}, (3.6)

for all t0t\geq 0. First, since LL is non-decreasing, it follows that V0,ξV^{0,\xi} is non-decreasing as well. Second, since dLt(ω)>0\mathrm{d}L_{t}(\omega)>0, for PG-a.a.(ω,t){S()=b}\textsf{P}\otimes G\text{-a.a.}\>(\omega,t)\in\{S_{\cdot}(\,\cdot\,)=b\}, it follows that 𝟏{τb<}Vt0,ξ>0\mathbf{1}_{\{\tau_{b}<\infty\}}V^{0,\xi}_{t}>0, for all tτbt\geq\tau_{b}, where τbinf{t0|St=b}\tau_{b}\coloneqq\inf\{t\geq 0\,|\,S_{t}=b\} denotes the first-passage time of SS to bb. Since the properties of geometric Brownian motion ensure that P(τb<)>0\textsf{P}(\tau_{b}<\infty)>0, this implies that P(V0,ξ>0)>0\textsf{P}(V^{0,\xi}_{\infty-}>0)>0. Together, these observations confirm that ξ\xi generates an increasing profit. ∎

Figure 3.1 plots a simulated path for the stock price and the corresponding path for the reflection term, together with the path for the stock position in the arbitrage portfolio described in the proof of Proposition 3.1. Figure 3.11(a) illustrates the reflection of the stock price off the boundary, while Figure 3.11(b) illustrates the behaviour of the the reflection term (which is effectively the local time of the stock price at the boundary). Each time the stock price visits the boundary, we see an instantaneous increase in the refection term, which nudges the stock price away from the boundary. Figure 3.11(c) shows that the stock holding in the arbitrage portfolio switches from zero to one in that instant, before immediately switching back to zero as the stock price is instantaneously reflected off the boundary. In effect, the stock is purchased and immediately sold when its price reaches the boundary, realising a profit equal to the increment in the reflection term in that instant. This profit is deposited in the bank account, so that the value of the arbitrage portfolio at any time is the cumulative sum of the instantaneous profits realised up to that time, plus interest.

Refer to caption
(a) Sample path of the stock price.
Refer to caption
(b) Sample path of the reflection term.
Refer to caption
(c) Sample path of the arbitrage strategy.
Figure 3.1. Sample path for the stock price and the reflection term determined by (3.1), with μ=0\mu=0, σ=0.5\sigma=0.5, b=1b=1 and S0=2S_{0}=2, as well as the corresponding sample path for the immediate arbitrage strategy in the proof of Proposition 3.1.

In the case when the risk-free interest rate is zero, it follows from (3.6) that the value of the arbitrage portfolio is simply the value of the reflection term (or the scaled local time process). In that case, Figure 3.11(b) also illustrates the behaviour of the value of the arbitrage portfolio. We see that the portfolio requires no initial investment of capital, since its initial value is zero. We also see that its value is non-decreasing, which reflects the riskless nature of the strategy as well as the fact that it requires no trading capital to fund margin calls. Moreover, the portfolio value increases instantaneously from zero at the first-passage time of the stock price to the reflecting boundary, and increases again each subsequent time the stock price visits the boundary. Effectively, the arbitrage strategy harvests the local time of the stock price at the boundary as a riskless profit.

Veestraeten (2008) claimed that the RGBM model does not admit arbitrage opportunities. He justified this assertion by citing two properties of the stock price in the model. First, since reflection off the boundary is instantaneous, the Lebesgue measure of the time spent there by the stock price is zero. Second, the stock price follows a continuous process. Based on those two observations (both of which are correct), he argued that the model is arbitrage-free and therefore admits an equivalent risk-neutral probability measure. Although Proposition 3.1 refutes this argument, we shall nevertheless dwell on it, because it has been repeated by several other authors.

The main problem is that Veestraeten’s (2008) argument focuses on the wrong measure of the time spent by the stock price at the reflecting boundary. Although the Lebesgue measure of the time spent there is indeed zero, that is irrelevant. Instead, we should focus on the local time of the stock price at the boundary. Being a continuous process whose paths are of unbounded total variation on compact intervals means that the sample paths of the stock price are extremely irregular. As a result of this irregularity, the stock price spends a non-zero amount of time infinitesimally close to the boundary, each time it visits, even though it spends no time actually on the boundary. The time spent arbitrarily close to the boundary is what the local time process measures. Each time the stock price reaches the boundary, the clock measuring its local time there ticks over and the local time increment is added to the stock price, nudging it away from the boundary. The arbitrage strategy constructed in the proof of Proposition 3.1 exploits this behaviour, by purchasing and immediately selling the stock each time it reaches the boundary, risklessly harvesting the local time increments in the process.

Another problem with Veestraeten’s (2008) argument is its apparent reliance on the idea that it is practically impossible for an arbitrageur to time their trades to coincide with the stock price reaching and leaving the reflecting boundary. Mathematically, an arbitrage is merely predictable process that satisfies certain technical conditions, and a model is arbitrage-free (in the appropriate sense) if no such process exists. The question of whether the strategy could be implemented in practice has no bearing on its existence as a well-defined mathematical object, and is thus irrelevant to the arbitrage properties of the model. Even though no trader could implement the strategy described in the proof of Proposition 3.1, it nevertheless exists as a well-defined predictable process satisfying the conditions in Definition 2.3.

3.4. Failure of the Structure Condition

Theorem 2.4 showed that a continuous financial market model satisfies the NIP condition if and only if it satisfies the structure condition. Proposition 3.1 proved that the RGBM model does not satisfy the NIP condition, by explicitly constructing an immediate profit strategy. Hence, the structure condition must fail as well. The next proposition proves this directly.

Proposition 3.2.

The RGBM model does not satisfy the structure condition.

Proof.

Suppose that the structure condition does in fact hold. In that case, there is a predictable process ϑ\vartheta satisfying (2.4). Using (3.5), that equation can be manipulated as follows:

bt(ω)ρt(ω)=ϑt(ω)at2(ω)\displaystyle b_{t}(\omega)-\rho_{t}(\omega)=\vartheta_{t}(\omega)a_{t}^{2}(\omega)
\displaystyle\Rightarrow\qquad bt(ω)dGt(ω)ρt(ω)dGt(ω)=ϑt(ω)at2(ω)dGt(ω)\displaystyle b_{t}(\omega)\,\mathrm{d}G_{t}(\omega)-\rho_{t}(\omega)\,\mathrm{d}G_{t}(\omega)=\vartheta_{t}(\omega)a_{t}^{2}(\omega)\,\mathrm{d}G_{t}(\omega)
\displaystyle\Rightarrow\qquad μdt+dLt(ω)brdt=ϑt(ω)σ2dt\displaystyle\mu\,\mathrm{d}t+\frac{\mathrm{d}L_{t}(\omega)}{b}-r\,\mathrm{d}t=\vartheta_{t}(\omega)\sigma^{2}\,\mathrm{d}t
\displaystyle\Rightarrow\qquad dtb(ω)=2dLt(ω)=2b(ϑt(ω)σ2(μr))dt,\displaystyle\mathrm{d}\ell^{b}_{t}(\omega)=2\,\mathrm{d}L_{t}(\omega)=2b\bigl{(}\vartheta_{t}(\omega)\sigma^{2}-(\mu-r)\bigr{)}\,\mathrm{d}t,

for PG-a.a.(ω,t)Ω×+\textsf{P}\otimes G\text{-a.a.}\;(\omega,t)\in\Omega\times\mathbb{R}_{+}. This gives rise to a contradiction, since

0<τb(ω)b(ω)=τb(ω)b(ω)limε0τb(ω)εb(ω)=limε0τb(ω)ετb(ω)dtb(ω)=limε0τb(ω)ετb(ω)2b(ϑt(ω)σ2(μr))dt=0,\begin{split}0<\ell^{b}_{\tau_{b}(\omega)}(\omega)&=\ell^{b}_{\tau_{b}(\omega)}(\omega)-\lim_{\varepsilon\downarrow 0}\ell^{b}_{\tau_{b}(\omega)-\varepsilon}(\omega)\\ &=\lim_{\varepsilon\downarrow 0}\int_{\tau_{b}(\omega)-\varepsilon}^{\tau_{b}(\omega)}\mathrm{d}\ell^{b}_{t}(\omega)\\ &=\lim_{\varepsilon\downarrow 0}\int_{\tau_{b}(\omega)-\varepsilon}^{\tau_{b}(\omega)}2b\bigl{(}\vartheta_{t}(\omega)\sigma^{2}-(\mu-r)\bigr{)}\,\mathrm{d}t=0,\end{split}

for P-a.aω{τb<}\textsf{P}\text{-a.a}\;\omega\in\{\tau_{b}<\infty\}, and P(τb<)>0\textsf{P}(\tau_{b}<\infty)>0. Hence, the structure condition cannot hold. ∎

We can approach Proposition 3.2 from a slightly more intuitive angle. Using (3.4), we obtain the following expressions for the continuous finite variation process A^\hat{A} and the continuous local martingale M^\hat{M}, defined by (2.5):

A^t=0t(μr)S^udu+0tdLuBuandM^t=0tσS^udWu,\hat{A}_{t}=\int_{0}^{t}(\mu-r)\hat{S}_{u}\,\mathrm{d}u+\int_{0}^{t}\frac{\mathrm{d}L_{u}}{B_{u}}\qquad\text{and}\qquad\hat{M}_{t}=\int_{0}^{t}\sigma\hat{S}_{u}\,\mathrm{d}W_{u},

for all t0t\geq 0, where S^S/B\hat{S}\coloneqq\nicefrac{{S}}{{B}} is the discounted stock price. Since the local time of the stock price at the reflecting boundary increases instantaneously each time gets there, while the Lebesgue measure of the time spent by the stock price at the boundary is zero, we have

0𝟏{St(ω)=b}dtb(ω)>0and0𝟏{St(ω)=b}dt=0,\int_{0}^{\infty}\mathbf{1}_{\{S_{t}(\omega)=b\}}\,\mathrm{d}\ell^{b}_{t}(\omega)>0\qquad\text{and}\qquad\int_{0}^{\infty}\mathbf{1}_{\{S_{t}(\omega)=b\}}\,\mathrm{d}t=0,

for P-a.a.ω{τb<}\textsf{P}\text{-a.a.}\;\omega\in\{\tau_{b}<\infty\}. Consequently,

0𝟏{St(ω)=b}dA^t(ω)=0𝟏{St(ω)=b}(μr)bdt+0𝟏{St(ω)=b}dLt(ω)Bt=0𝟏{St(ω)=b}dtb(ω)2Bt>0\displaystyle\begin{split}\int_{0}^{\infty}\mathbf{1}_{\{S_{t}(\omega)=b\}}\,\mathrm{d}\hat{A}_{t}(\omega)&=\int_{0}^{\infty}\mathbf{1}_{\{S_{t}(\omega)=b\}}(\mu-r)b\,\mathrm{d}t+\int_{0}^{\infty}\mathbf{1}_{\{S_{t}(\omega)=b\}}\,\frac{\mathrm{d}L_{t}(\omega)}{B_{t}}\\ &=\int_{0}^{\infty}\mathbf{1}_{\{S_{t}(\omega)=b\}}\,\frac{\mathrm{d}\ell^{b}_{t}(\omega)}{2B_{t}}>0\end{split}
and
0𝟏{St(ω)=b}dM^t(ω)\displaystyle\int_{0}^{\infty}\mathbf{1}_{\{S_{t}(\omega)=b\}}\,\mathrm{d}\langle\hat{M}\rangle_{t}(\omega) =0𝟏{St(ω)=b}σ2b2Bt2dt=0,\displaystyle=\int_{0}^{\infty}\mathbf{1}_{\{S_{t}(\omega)=b\}}\sigma^{2}\frac{b^{2}}{B_{t}^{2}}\,\mathrm{d}t=0,

for P-a.a.ω{τb<}\textsf{P}\text{-a.a.}\;\omega\in\{\tau_{b}<\infty\}. Since P(τb<)>0\textsf{P}(\tau_{b}<\infty)>0, it follows that A^\centernotM^\hat{A}\centernot\ll\langle\hat{M}\rangle, in violation of the structure condition.444See the interpetation of the structure condition following Definition 2.2. In other words, the structure condition fails because a non-trivial set of paths of A^\hat{A} (those for which the first-passage time of the stock price to the reflecting boundary is finite) increase on sets of Lebesgue-measure zero (corresponding to the times when the stock price is at the reflecting boundary), while the paths of M^\langle\hat{M}\rangle cannot change on sets of Lebesgue-measure zero.

3.5. Inconsistent Option Pricing Formulae

Proposition 3.1 established that the RGBM model does not satisfy the NIP condition. As a result, the NFLVR condition also fails, which rules out the existence of an equivalent risk-neutral probability measure. Operating under the mistaken belief that risk-neutral valuation is possible for the RGBM model, Veestraeten (2008) derived pricing formulae for European options written on the stock. Subject to an amendment to the pricing formula for European put options by Hertrich and Veerstraeten (2013), these formulae have also been used by Hertrich (2015) and Hertrich and Zimmermann (2017). Since their derivations are predicated on an incorrect assumption, we expect them to exhibit some inconsistencies.

Suppose we make the counterfactual assumption that an equivalent risk-neutral probability measure QP\textsf{{Q}}\sim\textsf{P} does in fact exist for the RGBM model. In that case, the discounted stock price process S^\hat{S} is a Q-local martingale. Given a European contingent claim written on the stock, with maturity T0T\geq 0 and payoff function h:[b,)h:[b,\infty)\rightarrow\mathbb{R}, we may (by hypothesis) apply a standard risk-neutral valuation approach to obtain its pricing function Vh:[0,T]×[b,)V^{h}:[0,T]\times[b,\infty)\rightarrow\mathbb{R}, as follows:

Vh(t,S)Et,SQ(BtBTh(ST))=Et,SQ(er(Tt)h(ST)),V^{h}(t,S)\coloneqq\textsf{E}^{\textsf{{Q}}}_{t,S}\biggl{(}\frac{B_{t}}{B_{T}}h(S_{T})\biggr{)}=\textsf{E}^{\textsf{{Q}}}_{t,S}\bigl{(}\mathrm{e}^{-r(T-t)}h(S_{T})\bigr{)},

for all (t,S)[0,T]×[b,)(t,S)\in[0,T]\times[b,\infty). Here Et,SQ()\textsf{E}^{\textsf{{Q}}}_{t,S}(\,\cdot\,) is the expected value operator with respect to the risk-neutral probability measure Qt,S\textsf{{Q}}_{t,S}, under which St=SbS_{t}=S\geq b.

Consider European call and put options on the stock, with a common strike price of KbK\geq b and a common maturity date T0T\geq 0. The pricing functions for these instruments are given by

C(t,S)\displaystyle C(t,S) Et,SQ(er(Tt)(STK)+)\displaystyle\coloneqq\textsf{E}^{\textsf{{Q}}}_{t,S}\bigl{(}\mathrm{e}^{-r(T-t)}(S_{T}-K)^{+}\bigr{)} (3.7)
and
P(t,S)\displaystyle P(t,S) Et,SQ(er(Tt)(KST)+),\displaystyle\coloneqq\textsf{E}^{\textsf{{Q}}}_{t,S}\bigl{(}\mathrm{e}^{-r(T-t)}(K-S_{T})^{+}\bigr{)}, (3.8)

for all (t,S)[0,T]×[b,)(t,S)\in[0,T]\times[b,\infty). Veestraeten (2008) derived a putative risk-neutral transition density for the stock price, which Veestraeten (2008), Hertrich and Veerstraeten (2013) and Hertrich (2015) used to evaluate (3.7) and (3.8). In so-doing, they obtained the following pricing functions for European calls and puts:555The option pricing formulae are formulated differently (but equivalently) in the cited papers. We have chosen the formulations in Hertrich (2015, Equations (11) and (12)), which produce compact expressions.

C(t,S)=SΦ(z1)Ker(Tt)Φ(z1σTt)+1θ(S(bS)1+θΦ(z3)Ker(Tt)(Kb)θ1Φ(z3θσTt))\begin{split}C(t,S)&=S\Phi(z_{1})-K\mathrm{e}^{-r(T-t)}\Phi\bigl{(}z_{1}-\sigma\sqrt[\ ]{T-t}\bigr{)}\\ &\hskip 50.00008pt+\frac{1}{\theta}\biggl{(}S\Bigl{(}\frac{b}{S}\Bigr{)}^{1+\theta}\Phi(z_{3})-K\mathrm{e}^{-r(T-t)}\Bigl{(}\frac{K}{b}\Bigr{)}^{\theta-1}\Phi\bigl{(}z_{3}-\theta\sigma\sqrt[\ ]{T-t}\bigr{)}\biggr{)}\end{split} (3.9)

and

P(t,S)=Ker(Tt)Φ(z1+σTt)ber(Tt)Φ(z4)S(Φ(z4+σTt)Φ(z1))1θ(S(bS)1+θ(Φ(z4+θσTt)Φ(z3))ber(Tt)Φ(z4)+Ker(Tt)(Kb)θ1Φ(z3θσTt)),\begin{split}P(t,S)&=K\mathrm{e}^{-r(T-t)}\Phi\bigl{(}-z_{1}+\sigma\sqrt[\ ]{T-t}\bigr{)}-b\mathrm{e}^{-r(T-t)}\Phi(z_{4})\\ &\hskip 50.00008pt-S\bigl{(}\Phi\bigl{(}-z_{4}+\sigma\sqrt[\ ]{T-t}\bigr{)}-\Phi(z_{1})\bigr{)}\\ &\hskip 50.00008pt-\frac{1}{\theta}\biggl{(}S\Bigl{(}\frac{b}{S}\Bigr{)}^{1+\theta}\Bigl{(}\Phi(z_{4}+\theta\sigma\sqrt[\ ]{T-t}\bigr{)}-\Phi(z_{3})\Bigr{)}-b\mathrm{e}^{-r(T-t)}\Phi(z_{4})\\ &\hskip 100.00015pt+K\mathrm{e}^{-r(T-t)}\Bigl{(}\frac{K}{b}\Bigr{)}^{\theta-1}\Phi\bigl{(}z_{3}-\theta\sigma\sqrt[\ ]{T-t}\bigr{)}\biggr{)},\end{split} (3.10)

for all (t,S)[0,T]×[b,)(t,S)\in[0,T]\times[b,\infty), where

z1lnSK+(r+12σ2)(Tt)σTt,\displaystyle z_{1}\coloneqq\frac{\ln\frac{S}{K}+\bigl{(}r+\frac{1}{2}\sigma^{2}\bigr{)}(T-t)}{\sigma\sqrt[\ ]{T-t}}, z3lnb2KS+(r+12σ2)(Tt)σTt,\displaystyle z_{3}\coloneqq\frac{\ln\frac{b^{2}}{KS}+\bigl{(}r+\frac{1}{2}\sigma^{2}\bigr{)}(T-t)}{\sigma\sqrt[\ ]{T-t}},
z4lnbS(r12σ2)(Tt)σTt,and\displaystyle z_{4}\coloneqq\frac{\ln\frac{b}{S}-\bigl{(}r-\frac{1}{2}\sigma^{2}\bigr{)}(T-t)}{\sigma\sqrt[\ ]{T-t}},\qquad\text{and} θ2rσ2.\displaystyle\theta\coloneqq\frac{2r}{\sigma^{2}}.

Here Φ()\Phi(\,\cdot\,) denotes the cumulative distribution function for a standard normal random variable.

As an immediate consequence of (3.7), we deduce that

C(t,S)Et,SQ(er(Tt)ST)S,C(t,S)\leq\textsf{E}^{\textsf{{Q}}}_{t,S}\bigl{(}\mathrm{e}^{-r(T-t)}S_{T}\bigr{)}\leq S, (3.11)

for all (t,S)[0,T]×[b,)(t,S)\in[0,T]\times[b,\infty), where the second inequality follows because S^\hat{S} is a supermartingale under Q, by virtue of Fatou’s lemma.666Since S^\hat{S} is (by assumption) a local martingale under Q, which is bounded from below, an application of Fatou’s lemma reveals that it is also a Q-supermartingale. Consequently, EQ(er(Tt)ST|t)=ertEQ(S^T|t)ertS^t=St,\textsf{E}^{\textsf{{Q}}}\bigl{(}\mathrm{e}^{-r(T-t)}S_{T}\,|\,\mathscr{F}_{t}\bigr{)}=\mathrm{e}^{rt}\textsf{E}^{\textsf{{Q}}}(\hat{S}_{T}\,|\,\mathscr{F}_{t})\leq\mathrm{e}^{rt}\hat{S}_{t}=S_{t}, for all t0t\geq 0. Note that (3.11) is a model-independent upper bound on call prices, whose violation contradicts the assumption that Q is an equivalent risk-neutral probability measure. The next lemma reveals that this bound can be violated if call prices are determined by (3.9).

Proposition 3.3.

The price of a call option obtained from (3.9) violates the upper bound (3.11) under certain parameter regimes, when the stock price is close to the reflecting boundary.

Proof.

Suppose the model parameters satisfy θ=1\theta=1, which is to say that r=12σ2r=\frac{1}{2}\sigma^{2}. We consider the price of a call option at time t[0,T)t\in[0,T) and assume that the stock price is located on the boundary bb at that time. In that case, (3.9) gives the following expression for the call price:

C(t,b)=2bΦ(lnbKσTt+σTt)2Ker(Tt)Φ(lnbKσTt).C(t,b)=2b\Phi\biggl{(}\frac{\ln\frac{b}{K}}{\sigma\sqrt[\ ]{T-t}}+\sigma\sqrt[\ ]{T-t}\biggr{)}-2K\mathrm{e}^{-r(T-t)}\Phi\biggl{(}\frac{\ln\frac{b}{K}}{\sigma\sqrt[\ ]{T-t}}\biggr{)}. (3.12)

Now, since

limKbΦ(lnbKσTt+σTt)=Φ(σTt)>12,\lim_{K\downarrow b}\Phi\biggl{(}\frac{\ln\frac{b}{K}}{\sigma\sqrt[\ ]{T-t}}+\sigma\sqrt[\ ]{T-t}\biggr{)}=\Phi\bigl{(}\sigma\sqrt[\ ]{T-t}\bigr{)}>\frac{1}{2},

we can choose a small enough strike price K>bK^{\prime}>b and a large enough risk-free interest rate r>0r^{\prime}>0 to ensure that

Φ(lnbKσTt+σTt)>12+Kber(Tt)Φ(lnbKσTt).\Phi\biggl{(}\frac{\ln\frac{b}{K^{\prime}}}{\sigma\sqrt[\ ]{T-t}}+\sigma\sqrt[\ ]{T-t}\biggr{)}>\frac{1}{2}+\frac{K^{\prime}}{b}\mathrm{e}^{-r^{\prime}(T-t)}\Phi\biggl{(}\frac{\ln\frac{b}{K^{\prime}}}{\sigma\sqrt[\ ]{T-t}}\biggr{)}.

Using those parameter values, (3.12) gives C(t,b)>bC(t,b)>b, in violation of (3.11). The continuity of the call pricing function (3.9) with respect to the stock price ensures that the bound will also be violated for stock prices above the reflecting boundary, but sufficiently close to it. ∎

Proposition 3.3 shows that call prices obtained from (3.9) can exceed the upper bound (3.11) when the stock price is close to the reflecting boundary. This is illustrated in Figure 3.2, which also illustrates that Black and Scholes (1973) call prices are consistent with the upper bound. While the prices obtained from the two formulae converge as the stock price increases, we see that the RGBM call pricing formula (3.9) produces inflated prices that are inconsistent with risk-neutral valuation when the stock price approaches the reflecting boundary. This echoes our analysis of the failures of NIP and the structure condition, which revealed that the arbitrage pathologies of the RGBM model are intimately related to its reflecting behaviour.

Refer to caption
Figure 3.2. The dependence of the RGBM call price (3.9) (solid red curve) and the Black and Scholes (1973) call price (solid blue curve) on the stock price. The vertical dashed line is the reflecting boundary for the RGBM model and the sloped dashed line is the upper bound (3.11) for the call price. The parameter values are Tt=10T-t=10, K=2K=2, r=0.125r=0.125, σ=0.5\sigma=0.5 and b=1b=1.

We turn our attention now to put options. From (3.8), we obtain

P(t,S)Et,SQ(er(Tt)(KST))Ker(Tt)S,P(t,S)\geq\textsf{E}^{\textsf{{Q}}}_{t,S}\bigl{(}\mathrm{e}^{-r(T-t)}(K-S_{T})\bigr{)}\geq K\mathrm{e}^{-r(T-t)}-S, (3.13)

for all (t,S)[0,T]×[b,)(t,S)\in[0,T]\times[b,\infty), where the second inequality follows because S^\hat{S} is a supermartingale under Q, as previously noted. In this case, (3.13) imposes a model-independent lower bound on European put prices, whose violation once again contradicts the assumption that Q is an equivalent risk-neutral probability measure. The next lemma reveals that this bound can be violated if put prices are determined by (3.10).

Proposition 3.4.

The price of a put option obtained from (3.10) violates the lower bound (3.13) under certain parameter regimes, when the stock price is close to the reflecting boundary.

Proof.

Suppose the model parameters satisfy θ=1\theta=1, which is to say that r=12σ2r=\frac{1}{2}\sigma^{2}. We consider the price of a put option at time t[0,T)t\in[0,T) and assume that the stock price is located on the boundary bb at that time. In that case, (3.10) gives the following expression for the put price:

P(t,b)=Ker(Tt)(12Φ(lnbKσTt))2b(Φ(σTt)Φ(lnbKσTt+σTt)).\begin{split}P(t,b)&=K\mathrm{e}^{-r(T-t)}\biggl{(}1-2\Phi\biggl{(}\frac{\ln\frac{b}{K}}{\sigma\sqrt[\ ]{T-t}}\biggr{)}\biggr{)}\\ &\hskip 50.00008pt-2b\biggl{(}\Phi\bigl{(}\sigma\sqrt[\ ]{T-t}\bigr{)}-\Phi\biggl{(}\frac{\ln\frac{b}{K}}{\sigma\sqrt[\ ]{T-t}}+\sigma\sqrt[\ ]{T-t}\biggr{)}\biggr{)}.\end{split} (3.14)

Now, since

Φ(σTt)>12andlimKΦ(lnbKσTt+σTt)=0,\Phi\bigl{(}\sigma\sqrt[\ ]{T-t}\bigr{)}>\frac{1}{2}\qquad\text{and}\qquad\lim_{K\uparrow\infty}\Phi\biggl{(}\frac{\ln\frac{b}{K}}{\sigma\sqrt[\ ]{T-t}}+\sigma\sqrt[\ ]{T-t}\biggr{)}=0,

we can choose a large enough strike price K>bK^{\prime}>b to ensure that

Φ(σTt)Φ(lnbKσTt+σTt)>12.\Phi\bigl{(}\sigma\sqrt[\ ]{T-t}\bigr{)}-\Phi\biggl{(}\frac{\ln\frac{b}{K^{\prime}}}{\sigma\sqrt[\ ]{T-t}}+\sigma\sqrt[\ ]{T-t}\biggr{)}>\frac{1}{2}.

Combining this with the fact that

12Φ(lnbKσTt)<11-2\Phi\biggl{(}\frac{\ln\frac{b}{K^{\prime}}}{\sigma\sqrt[\ ]{T-t}}\biggr{)}<1

allows us to deduce from (3.14) that P(t,b)<Ker(Tt)bP(t,b)<K^{\prime}\mathrm{e}^{-r(T-t)}-b, in violation of (3.13). The continuity of the put pricing function (3.10) with respect to the stock price ensures that the bound will also be violated for stock prices above the reflecting boundary, but sufficiently close to it. ∎

Proposition 3.4 shows that put prices obtained from (3.9) can violate the lower bound (3.13) when the stock price is close to the reflecting boundary. This is illustrated in Figure 3.3, which also illustrates that Black and Scholes (1973) put prices are consistent with the lower bound. While the prices obtained from the two formulae converge as the stock price increases, we see that the RGBM put pricing formula (3.10) produces diminished prices that are inconsistent with risk-neutral valuation when the stock price approaches the reflecting boundary. In other words, the pathologies of the RGBM model are once again evident near the reflecting boundary.

Refer to caption
Figure 3.3. The dependence of the RGBM put price (3.10) (solid red curve) and the Black and Scholes (1973) put price (solid blue curve) on the stock price. The vertical dashed line is the reflecting boundary for the RGBM model and the sloped dashed line is the lower bound (3.13) for the put price. The parameter values are Tt=10T-t=10, K=2K=2, r=0.02r=0.02, σ=0.2\sigma=0.2 and b=1b=1.

We conclude by emphasising that any violation of the bounds (3.11) and (3.13) is inconsistent with the assumption that an equivalent risk-neutral probability measure exists. Consequently, Propositions 3.3 and 3.4 imply that either the option pricing formulae (3.9) and (3.10) are incorrect or the assumption that an equivalent risk-neutral probability measure exists is false.

3.6. Problems with No-Negative-Equity Guarantees

An equity release mortgage (ERM) is a loan made to an elderly property-owning borrower that is collateralised by their property.777ERMS are typically known as reverse mortgages outside the U.K. In the U.K., ERMs typically embody a no-negative equity guarantee (NNEG) stipulating that the amount due for repayment is no more than the minimum of the rolled-up loan amount and the property value at the time of repayment, which would be the time of the borrower’s death or their entry into a care home. This obligation to repay the minimum of two future values implies that a NNEG involves a put option issued to the borrower.

NNEG valuation plays a crucial role in the design and management of ERMs, and several valuation approaches have been considered.888See Buckner and Dowd (2020) for a survey of NNEG valuation. Thomas (2021) recently proposed a new approach, built on the assumption that the government will intervene in the residential real estate market if property prices fall by more than a certain proportion, thereby establishing a de facto lower bound for house prices. Based on this idea, he proposed an RGBM model for house prices, where the lower reflecting boundary b>0b>0 corresponds to the level at which the government will enter the property market to support prices. By rearranging the pricing formula for a European put on a dividend-paying security in Hertrich (2015), Thomas (2021) obtained the following formula for the price of a NNEG with maturity date T>0T>0 and loan principal KbK\geq b, written on a house whose current price is SbS\geq b:

P(t,S)=Ker(Tt)Φ(z1+σTt)Seq(Tt)Φ(z1)ber(Tt)Φ(z3+σTt)+Seq(Tt)Φ(z3)+1θ(ber(Tt)Φ(z3+σTt)Seq(Tt)(bS)1+θ(Φ(z4)Φ(z2))Ker(Tt)(Kb)θ1Φ(z2θσTt)),\begin{split}P(t,S)&=K\mathrm{e}^{-r(T-t)}\Phi\bigl{(}-z_{1}+\sigma\sqrt[\ ]{T-t}\bigr{)}-S\mathrm{e}^{-q(T-t)}\Phi(-z_{1})\\ &\hskip 50.00008pt-b\mathrm{e}^{-r(T-t)}\Phi\bigl{(}-z_{3}+\sigma\sqrt[\ ]{T-t}\bigr{)}+S\mathrm{e}^{-q(T-t)}\Phi(-z_{3})\\ &\hskip 50.00008pt+\frac{1}{\theta}\biggl{(}b\mathrm{e}^{-r(T-t)}\Phi\bigl{(}-z_{3}+\sigma\sqrt[\ ]{T-t}\bigr{)}\\ &\hskip 100.00015pt-S\mathrm{e}^{-q(T-t)}\Bigl{(}\frac{b}{S}\Bigr{)}^{1+\theta}\bigl{(}\Phi(z_{4})-\Phi(z_{2})\bigr{)}\\ &\hskip 100.00015pt-K\mathrm{e}^{-r(T-t)}\Bigl{(}\frac{K}{b}\Bigr{)}^{\theta-1}\Phi\bigl{(}z_{2}-\theta\sigma\sqrt[\ ]{T-t}\bigr{)}\biggr{)},\end{split} (3.15)

for all (t,S)[0,T]×[b,)(t,S)\in[0,T]\times[b,\infty), where

z1\displaystyle z_{1} lnSK+(rq+12σ2)(Tt)σTt,\displaystyle\coloneqq\frac{\ln\frac{S}{K}+\bigl{(}r-q+\frac{1}{2}\sigma^{2}\bigr{)}(T-t)}{\sigma\sqrt[\ ]{T-t}}, z2\displaystyle z_{2} lnb2KS+(rq+12σ2)(Tt)σTt,\displaystyle\coloneqq\frac{\ln\frac{b^{2}}{KS}+\bigl{(}r-q+\frac{1}{2}\sigma^{2}\bigr{)}(T-t)}{\sigma\sqrt[\ ]{T-t}},
z3\displaystyle z_{3} lnSb+(rq+12σ2)(Tt)σTt,\displaystyle\coloneqq\frac{\ln\frac{S}{b}+\bigl{(}r-q+\frac{1}{2}\sigma^{2}\bigr{)}(T-t)}{\sigma\sqrt[\ ]{T-t}}, z4\displaystyle z_{4} lnbS+(rq+12σ2)(Tt)σTt,\displaystyle\coloneqq\frac{\ln\frac{b}{S}+\bigl{(}r-q+\frac{1}{2}\sigma^{2}\bigr{)}(T-t)}{\sigma\sqrt[\ ]{T-t}},
and
θ\displaystyle\theta 2(rq)σ2.\displaystyle\coloneqq\frac{2(r-q)}{\sigma^{2}}.

As before, r0r\geq 0 is the continuously compounding risk-free interest rate, while σ>0\sigma>0 is the volatility of the price of the house. The parameter q0q\geq 0, which is called the deferment rate, is the continuously compounding discount rate that yields the deferment price of the house, when applied to its current price.999The deferment price of a house is the price payable now, for possession at some future date. Hence, the deferment rate qq may be regarded as a type of convenience yield.

Since (3.15) is a straightforward extension of the European put pricing formula (3.10), in order to accomodate a dividend-paying stock (with the deferment rate playing the role of a dividend yield), it shares the defects of that formula. First, the non-existence of an equivalent risk-neutral probability measure (or even a numéraire portfolio) for the RGBM model means that there is no economic justification for (3.15) as a pricing formula for NNEGs. Second, an easy modification of Proposition 3.4 shows that (3.15) can violate the model-independent lower bound

P(t,S)Ker(Tt)Seq(Tt),P(t,S)\geq K\mathrm{e}^{-r(T-t)}-S\mathrm{e}^{-q(T-t)}, (3.16)

for all (t,S)[0,T]×[b,)(t,S)\in[0,T]\times[b,\infty), when the house price is close to the reflecting boundary.101010The lower bound (3.16) is a direct analogue of (3.13), for the case of a dividend-paying stock. It can be derived in exactly the same way, if one assumes the existence of an equivalent risk-neutral probability measure. As before, any violation of that bound contradicts the existence of an equivalent risk-neutral probability measure. To make matters worse, the next lemma highlights an additional defect of (3.15) that makes it especially ill-suited to NNEG valuation.

Proposition 3.5.

The price of a sufficiently long-dated NNEG obtained from (3.15) violates the lower bound (3.16), if r=0r=0 and 0<q<12σ20<q<\frac{1}{2}\sigma^{2}.

Proof.

Suppose r=0r=0 and 0<q<12σ20<q<\frac{1}{2}\sigma^{2}, in which case 1<θ<0-1<\theta<0. Fix a current time t0t\geq 0 and a house price SbS\geq b, and observe that

z1+σTt\displaystyle-z_{1}+\sigma\sqrt[\ ]{T-t} =O(1Tt)+12σ(1θ)Tt,\displaystyle=\operatorname{O}\Bigl{(}\frac{1}{\sqrt[\ ]{T-t}}\Bigr{)}+\frac{1}{2}\sigma(1-\theta)\sqrt[\ ]{T-t}, z1\displaystyle-z_{1} =O(1Tt)12σ(1+θ)Tt,\displaystyle=\operatorname{O}\Bigl{(}\frac{1}{\sqrt[\ ]{T-t}}\Bigr{)}-\frac{1}{2}\sigma(1+\theta)\sqrt[\ ]{T-t},
z3+σTt\displaystyle-z_{3}+\sigma\sqrt[\ ]{T-t} =O(1Tt)+12σ(1θ)Tt,\displaystyle=\operatorname{O}\Bigl{(}\frac{1}{\sqrt[\ ]{T-t}}\Bigr{)}+\frac{1}{2}\sigma(1-\theta)\sqrt[\ ]{T-t}, z3\displaystyle-z_{3} =O(1Tt)12σ(1+θ)Tt,\displaystyle=\operatorname{O}\Bigl{(}\frac{1}{\sqrt[\ ]{T-t}}\Bigr{)}-\frac{1}{2}\sigma(1+\theta)\sqrt[\ ]{T-t},
z4\displaystyle z_{4} =O(1Tt)+12σ(1+θ)Tt,\displaystyle=\operatorname{O}\Bigl{(}\frac{1}{\sqrt[\ ]{T-t}}\Bigr{)}+\frac{1}{2}\sigma(1+\theta)\sqrt[\ ]{T-t}, z2\displaystyle z_{2} =O(1Tt)+12σ(1+θ)Tt,\displaystyle=\operatorname{O}\Bigl{(}\frac{1}{\sqrt[\ ]{T-t}}\Bigr{)}+\frac{1}{2}\sigma(1+\theta)\sqrt[\ ]{T-t},
and
z2θσTt\displaystyle z_{2}-\theta\sigma\sqrt[\ ]{T-t} =O(1Tt)+12σ(1θ)Tt,\displaystyle=\operatorname{O}\Bigl{(}\frac{1}{\sqrt[\ ]{T-t}}\Bigr{)}+\frac{1}{2}\sigma(1-\theta)\sqrt[\ ]{T-t},

for all T>tT>t. From this it follows that

limTΦ(z1+σTt)\displaystyle\lim_{T\uparrow\infty}\Phi\bigl{(}-z_{1}+\sigma\sqrt[\ ]{T-t}\bigr{)} =1,\displaystyle=1, limTΦ(z1)\displaystyle\lim_{T\uparrow\infty}\Phi(-z_{1}) =0,\displaystyle=0,
limTΦ(z3+σTt)\displaystyle\lim_{T\uparrow\infty}\Phi\bigl{(}-z_{3}+\sigma\sqrt[\ ]{T-t}\bigr{)} =1,\displaystyle=1, limTΦ(z3)\displaystyle\lim_{T\uparrow\infty}\Phi(-z_{3}) =0,\displaystyle=0,
limTΦ(z4)\displaystyle\lim_{T\uparrow\infty}\Phi(z_{4}) =1,\displaystyle=1, limTΦ(z2)\displaystyle\lim_{T\uparrow\infty}\Phi(z_{2}) =1,\displaystyle=1,
and
limTΦ(z2θσTt)\displaystyle\lim_{T\uparrow\infty}\Phi\bigl{(}z_{2}-\theta\sigma\sqrt[\ ]{T-t}\bigr{)} =1,\displaystyle=1,

by virtue of 1<θ<0-1<\theta<0. Consequently, (3.15) gives

limTP(t,S)=Kb+1θ(bK(Kb)θ1)=K+bθ(1θ(Kb)θ).\lim_{T\uparrow\infty}P(t,S)=K-b+\frac{1}{\theta}\biggl{(}b-K\Bigl{(}\frac{K}{b}\Bigr{)}^{\theta-1}\biggr{)}=K+\frac{b}{\theta}\biggl{(}1-\theta-\Bigl{(}\frac{K}{b}\Bigr{)}^{\theta}\biggr{)}. (3.17)

The second term in this expression is negative, since K>b>0K>b>0 and θ<0\theta<0, which implies that limTP(t,S)<K\lim_{T\uparrow\infty}P(t,S)<K. We may therefore choose T>tT^{\prime}>t large enough, such that

P(t,S)<KSeq(Tt),P(t,S)<K-S\mathrm{e}^{-q(T-t)},

for all T>TT>T^{\prime}, in violation of (3.16). ∎

Proposition 3.5 shows that NNEG prices obtained from (3.15) violate the lower bound (3.16) for sufficiently long-dated contracts, when the risk-free interest rate is zero and the deferment rate is not too large.111111We could generalise the result by proving that long-dated NNEG prices obtained from (3.15) do not obey (3.16) for any parameter regimes with 1<θ<0-1<\theta<0, but the proof would be messier. Since NNEGs are typically long-dated instruments, the upshot is that they are dramatically undervalued by (3.15) in low interest rate environments.

Figure 3.4 illustrates the issue described by Proposition 3.5. For the chosen parameter values, we see that (3.15) produces NNEG prices below the lower bound (3.16), for maturities in excess of 10 years. We also see that the prices obtained from that formula converge asymptotically to the limit (3.17). By contrast, Black (1976) NNEG prices are consistent with the lower bound (3.16), for all maturities.121212The Black (1976) put pricing formula is proposed by advocates of a market-consistent approach to NNEG valuation (see Buckner and Dowd 2020). To appreciate the severity of the undervaluation issue for long-dated contracts, note that the Thomas (2021) price of a 20-year NNEG in Figure 3.4 is only 46%46\% of the value of the lower bound and a mere 29%29\% of the value of the corresponding Black (1976) NNEG price.

Refer to caption
Figure 3.4. The dependence of the RGBM NNEG price (3.15) (solid red curve) and the Black (1976) NNEG price (solid blue curve) on time to expiry. The horizontal dashed line is the long maturity asymptote (3.17) for the RGBM model and the dashed curve is the lower bound (3.16) for the price of the NNEG. The parameter values are S=1S=1, K=0.9K=0.9, r=0r=0, q=0.03q=0.03, σ=0.3\sigma=0.3 and b=0.5b=0.5.

4. Conclusions

Intuitively, an arbitrage is a trading strategy that realises a riskless profit without requiring an upfront investment of capital. But in the context of continuous-time finance, it is best to think of an arbitrage simply as a predictable process that satisfies certain technical conditions, with only a tenuous link to feasible real-world trading strategies. Moreover, there are several notions of arbitrage in continuous-time finance, with different implications for the mathematical features of a model. These observations warn us that intuition and heuristic reasoning are an unreliable substitute for mathematical rigour, when analysing the arbitrage properties of a continuous-time financial market model.

The salience of this warning is illustrated by the recent literature advocating the use of reflected geometric Brownian motion (RGBM) as a security price model. Several authors have argued that the RGBM model does not offer any arbitrage opportunities, but their argument is informal and heuristic. In particular, they do not specify what type of arbitrage opportunities are precluded and they do not formally verify the associated no-arbitrage condition. Instead, they appeal to an intuitive idea that arbitrageurs cannot profit from the reflecting behaviour of the security price in the model because the price process is continuous and the time spent on the reflecting boundary has Lebesgue measure zero. Based on these observations, they wrongly conclude that the RGBM model admits an equivalent risk-neutral probability measure, which leads them to derive invalid option pricing formulae.

A careful mathematical analysis of the RGBM model shows that it violates even the weakest no-arbitrage condition considered in the literature. Consequently, it does not offer any of the structural features required for contingent claim valuation. In particular, it does not admit a numéraire portfolio or an equivalent risk-neutral probability measure. As a result, there is no theoretical justification for the published RGBM option pricing formulae. Indeed, a close examination of those formulae reveals that they behave pathologically under certain conditions.

References

  • Black (1976) Black, F. (1976). The pricing of commodity contracts. J. Finan. Econ. 3(1–2), 167–179.
  • Black and Scholes (1973) Black, F. and M. Scholes (1973). The pricing of options and corporate liabilities. J. Pol. Econ. 81(3), 637–654.
  • Buckner and Dowd (2020) Buckner, D. and K. Dowd (2020). The Eumaeus guide to equity release mortgage valuation: Restating the case for a market consistent approach. Technical report, The Eumaeus Project. Second edition.
  • Delbaen and Schachermayer (1994) Delbaen, F. and W. Schachermayer (1994). A general version of the fundamental theorem of asset pricing. Math. Ann. 300(3), 463–520.
  • Delbaen and Schachermayer (1995) Delbaen, F. and W. Schachermayer (1995). The existence of absolutely continuous local martingale measures. Ann. Appl. Probab. 5(4), 926–945.
  • Feng et al. (2020) Feng, L., P. Jiang, and Y. Wang (2020). Constant elasticity of variance models with target zones. Physica A 537, 1–22.
  • Fontana (2015) Fontana, C. (2015). Weak and strong no-arbitrage conditions for continuous financial markets. Int. J. Theor. Appl. Finance 18(1), 34.
  • Gerber and Pafumi (2000) Gerber, H. U. and G. Pafumi (2000). Pricing dynamic investment fund protection. N. Amer. Actuarial J. 4(2), 28–37.
  • Hertrich (2015) Hertrich, M. (2015). A cautionary note on the put-call parity under an asset pricing model with a lower reflecting barrier. Swiss J. Econ. Statist. 151(3), 227–260.
  • Hertrich and Veerstraeten (2013) Hertrich, M. and D. Veerstraeten (2013). Valuing stock options when prices are subject to a lower boundary: A correction. J. Futures Markets 33(9), 889–890.
  • Hertrich and Zimmermann (2017) Hertrich, M. and H. Zimmermann (2017). On the credibility of the Euro/Swiss Franc floor: A financial market perspective. J. Money, Credit, Banking 49(2–3), 567–578.
  • Hulley (2010) Hulley, H. (2010). Strict Local Martingales in Continuous Financial Market Models. Ph. D. thesis, University of Technology, Sydney.
  • Jacod and Shiryaev (2003) Jacod, J. and A. N. Shiryaev (2003). Limit Theorems for Stochastic Processes (Second ed.). Berlin: Springer.
  • Karatzas and Kardaras (2007) Karatzas, I. and C. Kardaras (2007). The numéraire portfolio in semimartingale financial models. Finance Stoch. 11(4), 447–493.
  • Karatzas and Shreve (1991) Karatzas, I. and S. E. Shreve (1991). Brownian Motion and Stochastic Calculus (Second ed.). New York: Springer.
  • Ko et al. (2010) Ko, B., E. S. W. Shiu, and L. Wei (2010). Pricing maturity guarantee with dynamic withdrawal benefit. Ins.: Mathematics Econ. 47(2), 216–223.
  • Melnikov and Wan (2021) Melnikov, A. and H. Wan (2021). On modifications of the Bachelier model. Ann. Finance 17(2), 187–214.
  • Neuman and Schied (2016) Neuman, E. and A. Schied (2016). Optimal portfolio liquidation in target zone models and catalytic superprocesses. Finance Stoch. 20(2), 495–509.
  • Pilipenko (2014) Pilipenko, A. (2014). An Introduction to Stochastic Differential Equations with Reflection. Potsdam: Potsdam University Press.
  • Platen and Heath (2006) Platen, E. and D. Heath (2006). A Benchmark Approach to Quantitative Finance. Berlin: Springer.
  • Revuz and Yor (1999) Revuz, D. and M. Yor (1999). Continuous Martingales and Brownian Motion (Third ed.). Berlin: Springer.
  • Schweizer (1995) Schweizer, M. (1995). On the minimal martingale measure and the Föllmer-Schweizer decomposition. Stoch. Anal. Appl. 13(5), 573–599.
  • Skorokhod (1961) Skorokhod, A. V. (1961). Stochastic equations for diffusion processes in a bounded region. Theory Probab. Appl. 6(3), 264–274.
  • Thomas (2021) Thomas, R. G. (2021). Valuation of no-negative-equity guarantees with a lower reflecting barrier. Ann. Actuar. Sci. 15(1), 115–143.
  • Veestraeten (2008) Veestraeten, D. (2008). Valuing stock options when prices are subject to a lower boundary. J. Futures Markets 28(3), 231–247.
  • Veestraeten (2013) Veestraeten, D. (2013). Currency option pricing in a credible exchange rate target zone. Appl. Finan. Econ. 23(11), 951–962.