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


Deep Hedging: Learning to Remove the Drift under Trading Frictions with Minimal Equivalent Near-Martingale Measures

H. BUEHLER{\dagger} P. MURRAY{\dagger}{{\ddagger}} M. S. PAKKANEN{{\ddagger}} and B. WOOD{\dagger}
Corresponding author Email: phillip.murray@jpmorgan.com {\dagger}JP Morgan
{\ddagger}Imperial College London
Abstract

We present a machine learning approach for finding minimal equivalent martingale measures for markets simulators of tradable instruments, e.g. for a spot price and options written on the same underlying. We extend our results to markets with frictions, in which case we find “near-martingale measures” under which the prices of hedging instruments are martingales within their bid/ask spread.

By removing the drift, we are then able to learn using Deep Hedging a “clean” hedge for an exotic payoff which is not polluted by the trading strategy trying to make money from statistical arbitrage opportunities. We correspondingly highlight the robustness of this hedge vs estimation error of the original market simulator. We discuss applications to two market simulators.

1 Introduction

A long-standing challenge in quantitative finance is the development of market models for the dynamics of tradable instruments such as a spot price and options thereon. The classic approach to developing such models is to find model dynamics in a suitable parameter space under which the respective risk-neutral drift could be computed somewhat efficiently, c.f. for example [12], [15], [10] for the case of equity option markets. With this approach, realistic dynamics or estimation of statistically valid parameters are an afterthought.

This article proposes to reverse this process by starting out with training a realistic model of the market under the statistical measure – and then find an equivalent “near-martingale” measure under which the drifts of tradable instruments are constrained by their marginal costs such that there are no statistical arbitrage opportunities, that is, trading strategies which produce positive expected gains. In the absence of trading costs, this means finding an equivalent martingale measure to “remove the drift”.

Indeed, we will show that absence of statistical arbitrage under a given measure is equivalent to the conditional expectation of the returns under this measure being constrained by their marginal bid/ask prices. This result is of independent interest.

The main motivation for the present work is the application of our Deep Hedging algorithm to construct hedging strategies for contingent claims by trading in hedging instruments which include derivatives such as options. When described first in [3], we relied on markets simulated with classic quantitative finance models. In [1] we proposed a method to build market simulators of options markets under the statistical measure. Under this measure, we will usually find statistical arbitrage in the sense that an empty initial portfolio has positive value. This reflects the realities of historic data: at the time of writing the S&P 500 had moved upwards over the last ten years, giving a machine the impression that selling puts and being long the market is a winning strategy. However, naively exploiting this observation risks falling foul of the “estimation error” of the mean returns of our hedging instruments. In the context of hedging a portfolio of exotic derivatives, the presence of statistical arbitrage is undesirable as an optimal strategy will be a combination of a true hedge, and a strategy which does not depend on our portfolio, but tries to take advantage of the opportunities seen the market. It is therefore not robust against estimation error of said drifts. Hence we propose using the method presented here to generate a “clean” hedge by removing the drift of the market to increase robustness against errors in the estimation of returns of our hedging instruments.

While our examples focus on simulating equity option markets – in this case amounting to a stochastic implied volatility model – our approach is by no means limited to the equities case. In fact, it is entirely model agnostic and can be applied to any market simulator which generates paths of tradable instruments under the same numeraire which are free of classic arbitrage.

In particular, our approach can be applied to “black box” neural network based simulators such as those using Generative Adversarial Networks (GANs) described in [1] or Variational Autoencoders as in [5]. Such simulators use machine learning methods to generate realistic paths from the statistical measure, but clearly no analytic expression to describe the market dynamics can be written. Our method allows constructing an equivalent risk neutral measure through further applications of machine learning methods.

1.1 Summary of our Approach

Given instrument returns DHtDH_{t} across discrete time steps t{0=t0<<tm}t\in\{0=t_{0}<\cdots<t_{m}\}, and convex costs ctc_{t} associated with trading ata_{t} units of each instrument, we propose using our Deep Hedging algorithm introduced in [3] to find a trading strategy aa^{*} and cash amount yy^{*} which maximize the optimized certainty equivalent of a utility uu,

asupy𝔼[u(y+tatDHtct(at))y].a\longmapsto\sup_{y\in\mathbb{R}}\ {\mathbb{E}}\left[\,{u\left(y+\sum_{t}a_{t}\cdot DH_{t}-c_{t}(a_{t})\right)-y}\,\right]\ .

Let mm denote the “marginal cost” of trading in our market. We then define an equivalent measure \mathbb{Q}^{*} by setting

dd:=u(y+tatDHtmt(at)).\frac{d\mathbb{Q}^{*}}{d\mathbb{P}}:=u^{\prime}\left(y^{*}+\sum_{t}a^{*}_{t}\cdot DH_{t}-m_{t}(a^{*}_{t})\right)\ .

Under \mathbb{Q}^{*} the market has no statistical arbitrage opportunities in the sense that there is no strategy aa which has positive expected returns, i.e.

supa𝔼[tatDHtct(at)]0.\sup_{a}\ {\mathbb{E}^{*}}\left[\,{\sum_{t}a_{t}\cdot DH_{t}-c_{t}(a_{t})}\,\right]\leq 0\ . (1)

In the absence of transaction costs, \mathbb{Q}^{*} is an equivalent martingale measure. In the presence of transaction costs, we show that removing statistical arbitrage is equivalent to the measure being an equivalent near-martingale measure, in the sense that the drift of all tradable instruments must be dominated by the transaction costs. Moreover, \mathbb{Q}^{*} is minimal among all equivalent (near-) martingale measures with respect to the u~\tilde{u}-divergence from \mathbb{P} where u~\tilde{u} is the Legendre-Fenchel transform of uu.

The key insight of our utility-based risk-neutral density construction is that it relies only on solving the optimization problem of finding aa^{*} and cc^{*}, not on any particular dynamics for the market under the \mathbb{P} measure. Therefore, it lends itself to the application of modern machine learning methods. As mentioned above, this is particularly useful in the case of removing statistical arbitrage from a “black box” market simulator, such as the GAN based approach discussed in [1]. Through the choice of utility function, we are able to control the risk neutral measure we construct.

We demonstrate the power of this approach with two examples of option market simulators for spot and a number of volatilities. Specifically, we train a Vector Autoregressive (VAR) model of the form

dYt=(BA1Yt1A2Yt2)dt+ΣdWtdY_{t}=\left(B-A_{1}Y_{t-1}-A_{2}Y_{t-2}\right)dt+\Sigma dW_{t}

for a vector of log spot returns and log volatilities Yt=(logSt/St1,logσt1,,logσtK)Y_{t}=(\log{S_{t}}/{S_{t-1}},\log\sigma^{1}_{t},\ldots,\log\sigma^{K}_{t})^{\prime}, and also a neural network based GAN simulator, and then in both cases use our approach to construct the above measure such that the resulting spots and option prices are near-martingale, and free from statistical arbitrage.

1.2 Related Work

We are not aware of attempts to numerically solve for a risk-neutral density with the approach discussed here, as an application to stochastic implied volatility or otherwise. To our knowledge, ours is the first practical approach for implementing general statistically trained market models under risk-neutral measures.

The classic approach to stochastic implied volatilities is via the route of identifying analytically a risk-neutral drift given the other parameters of the specified model. The first applicable results for a term structure of implied volatilities are due to [12]. The first viable approach to a full stochastic implied volatility model for an entire fixed strike and fixed maturity option surface was presented in[15], using as parametrization also discrete local volatilities. Wissel describes the required continuous time drift adjustment for a diffusion driving a grid of such discrete local volatilities as a function of the free parameters. Unnaturally, in his approach the resulting spot diffusion takes only discrete values at the strikes of the options at each maturity date and the approach is limited to a set grid of options defined in cash strikes and fixed maturities.

More recently, a number of works have shown that when representing an option surface with a Lévy kernel we can derive suitable Heath-Jarrow-Morton conditions on the parameters of the diffusion of the Lévy kernel such that the resulting stock price is arbitrage-free, c.f. [10] and the references therein. Simulation of the respective model requires solving the respective Fourier equations for the spot price and options at each step in the path.

1.3 Outline

The rest of the article is organised as follows. In Section 2 we describe the theoretical framework and introduce the key method for constructing a risk-neutral measure in the firctionless case, which is then extended to the case with market frictions in Section 3. Then, in Section 4 we describe some of the consequences of the approach from a practical persective, and in Section 5 we provide numerical experiments demonstrating the effectiveness of the method in practice.

2 Frictionless risk-neutral case

Consider a discrete-time simulated financial market with finite time horizon where we trade over time steps 0=t0<<tm=T0=t_{0}<\cdots<t_{m}=T where TT is the maximum maturity of all tradable instruments. Fix a probability space Ω\Omega and a probability measure \mathbb{P} under which the market is simulated, which we will refer to as the “statistical” measure. For each t{t0,,tm}t\in\{t_{0},\ldots,t_{m}\}, we denote by sts_{t} the state of the market at time tt, including relevant information from the past. The state represents all information available to us, including mid-prices of all tradable instruments, trading costs, restrictions and risk limits.

The sequence of states (st)t=0,,T(s_{t})_{t=0,\ldots,T} generates a sequence 𝔽=(t)t=0,,T\mathbb{F}=({\cal F}_{t})_{t=0,\ldots,T} of σ\sigma-algebras forming a filtration. Being generative means that any t{\cal F}_{t}-measurable function f()f(\cdot) can be written as a function of sts_{t} as ff(;st)f\equiv f(\cdot;s_{t}).

To simplify notation, we stipulate that the total number of instruments at each timestep is always nn. Let Ht(t)=(Ht(t,1),,Ht(t,n))H_{t}^{(t)}=(H_{t}^{(t,1)},\ldots,H_{t}^{(t,n)}) be the and n\mathbb{R}^{n}-valued, 𝔽\mathbb{F}-adapted stochastic process of mid-prices of the liquid instruments available to trade at tt. As above, Ht(t)H_{t}^{(t)} is a function of sts_{t}, and we also assume that HH is in L1()L^{1}(\mathbb{P}). Note that HtH_{t} can represent a wide class of instruments, including primary assets such as single equities, indices, and liquid options.

For each instrument we observe at time TT a final mark-to-market mid-value HT(t,i)H_{T}^{(t,i)} which will usually be the sum of any cashflows along the path, and which is also assumed to be a function of sTs_{T}. That means sTs_{T} must contain sufficient information from the past along the path: for example, if the iith instrument tradable at tt is a call option with relative strike kik_{i} and time-to-maturity τiTt\tau_{i}\leq T-t on a spot price process StS_{t}, then the final value of this iith instrument is the payoff on the path, HT(t,i)=(St+τi/Stki)+H_{T}^{(t,i)}=(S_{t+\tau_{i}}/S_{t}-k_{i})^{+}. Whilst for simplicity, we assume that all options mature within the time horizon, we can easily extend our method to the case where options are allowed to mature after TT by valuing them in TT at mid-prices.

We further assume that discounting rates, funding, dividends, and repo rates are zero. Extension to the case where they are non-zero and deterministic is straightforward.

At each time step tt we may chose an action atna_{t}\in\mathbb{R}^{n} to trade in the hedging instruments Ht(t)H_{t}^{(t)} based on the information available in the state sts_{t}, i.e. ata(st)a_{t}\equiv a(s_{t}). The n\mathbb{R}^{n}-valued, 𝔽\mathbb{F}-adapted stochastic process a=(a0,,am1)a=(a_{0},\ldots,a_{m-1}) defines a trading strategy over the time horizon. To ease notation, it will be useful to define at±:=max(0,±at)a^{\pm}_{t}:=\max(0,\pm a_{t}) for an action ata_{t}, where the operation is applied elementwise. We also use eie^{i} to refer to the iith unit vector.

We start with the frictionless case, where the actions are unconstrained, i.e. the set of admissible actions 𝒜t\mathcal{A}_{t} is equal to n\mathbb{R}^{n} for all tt. In this case the terminal gain of implementing a trading strategy a=(a0,,am1)a=(a_{0},\ldots,a_{m-1}) with at𝒜ta_{t}\in{\cal A}_{t} is given by

aDHT:=t=0m1atDHtwithDHt:=HT(t)Ht(t).{a\star DH_{T}}:=\sum_{t=0}^{m-1}a_{t}\cdot DH_{t}\ \ \ \mbox{with}\ \ \ DH_{t}:=H^{(t)}_{T}-H^{(t)}_{t}\ . (2)

Our slightly unusual notation of taking the performance of each instrument to maturity reflects our ambition to look at option market simulators for “floating” implied volatility surfaces where the observed financial instruments change from step to step. If the instruments tradable at each time step are, in fact, the same fixed strike and maturity instruments, then

aDHTt=0m1δtdHtwithdHt:=Ht+1Ht,{a\star DH_{T}}\equiv\sum_{t=0}^{m-1}\delta_{t}\cdot dH_{t}\ \ \ \mbox{with}\ \ \ dH_{t}:=H_{t+1}-H_{t}\ , (3)

where δt:=at+δt1\delta_{t}:=a_{t}+\delta_{t-1} starting with δ1:=0\delta_{-1}:=0.

2.1 Optimized certainty equivalents

In order to assess the performance of a trading strategy, we are looking for risk-adjusted measures of performance instead of plain expected return. We will focus on the following case: let uu be a strictly concave, strictly increasing utility function uu which is C1C^{1} and normalized to both u(0)=0u(0)=0 and u(0)=1u^{\prime}(0)=1.111 Note that the normalization is a convenience which is always achievable: If u~\tilde{u} is concave and strictly increasing, then u(x):=(u~(x)u~(0))/u~(0)u(x):=(\tilde{u}(x)-\tilde{u}(0))/\tilde{u}^{\prime}(0) satisfies these assumptions. Examples of such utility functions are the adjusted mean-volatility function u(x):=(1+λx1+λ2x2)/λu(x):=(1+\lambda x-\sqrt{1+\lambda^{2}x^{2}})/\lambda proposed in [9], or the exponential utility u(x)=(1eλx)/λu(x)=(1-e^{-\lambda x})/\lambda. We make the further assumption that u(aDHT)L1()u({a\star DH_{T}})\in L^{1}(\mathbb{P}) for all aa. (This condition is not met when uu is the exponential utility and the market contains a Black & Scholes process with negative drift.222 To see this, assume m=1m=1, n=1n=1, H0:=1H_{0}:=1, and let HT:=exp((μ12σ2)T+σTW)H_{T}:=\exp((-\mu-\frac{1}{2}\sigma^{2})T+\sigma\sqrt{T}W) for positive μ\mu and σ\sigma, and WW standard normal. Then 𝔼[u((HTH0))]=\mathbb{E}[u(-(H_{T}-H_{0}))]=-\infty. )

For a given utility function, define now the optimized certainty equivalent (OCE) of the expected utility, introduced in [2] as

U(X):=supy{𝔼[u(y+X)]y}.U(X):=\sup_{y\in\mathbb{R}}\Big{\{}{\mathbb{E}}\left[\,{u(y+X)}\,\right]-y\Big{\}}\ . (4)

The functional UU satisfies the following properties:

  1. (i)

    Monotone increasing: if XYX\geq Y then U(X)U(Y)U(X)\geq U(Y). A better payoff leads to higher expected utility.

  2. (ii)

    Concave: U(αX+(1α)Y)αU(X)+(1α)U(Y)U(\alpha X+(1-\alpha)Y)\geq\alpha U(X)+(1-\alpha)U(Y) for α[0,1]\alpha\in[0,1]. Diversification leads higher utility.

  3. (iii)

    Cash-Invariant: U(X+c)=U(X)+cU(X+c)=U(X)+c for all cc\in\mathbb{R}. Adding cash to a position increases its utility by the same amount.

This above properties mean that U(X)-U(X) is a convex risk measure. Note that the assumptions u(0)=0u(0)=0 and u(0)=1u^{\prime}(0)=1 imply that u(y)yu(y)\leq y for all yy and hence U(0)=0U(0)=0. Furthermore, UU is finite for all bounded variables XX, since by monotonicity we have U(X)U(supX)=U(0)+supX<U(X)\leq U(\sup X)=U(0)+\sup X<\infty.

Cash invariance means that U(XU(X))=0U(X-U(X))=0, i.e. U(X)-U(X) is the minimum amount of cash that needs to be added to a position in order to make it acceptable, in the sense that U(X+c)0U(X+c)\geq 0. The cash-invariance property of the OCE means in particular that optimizing UU does not depend on our initial wealth.

A classic example of such an OCE measure is the case where uu is the exponential utility with risk aversion level λ\lambda. In this case we obtain the entropy

Uλ(X)=1λlog𝔼[eλX].U_{\lambda}(X)=-\frac{1}{\lambda}\log\mathbb{E}\left[e^{-\lambda X}\right]\ .

We now consider the application of the optimized certainty equivalent to the terminal gains of a trading strategy. To this end, define

F(y,a):=𝔼[u(y+aDHT)y].F(y,a):={\mathbb{E}}\left[\,{u\big{(}\,y+{a\star DH_{T}}\,\big{)}-y}\,\right]\ . (5)
Lemma 2.1.

Suppose the market exhibits classic arbitrage. Then no finite maximizers of FF exist.

Proof.

Assume that a~\tilde{a} is a classic arbitrage opportunity with [a~DHT0]=1\mathbb{P}[\tilde{a}\star DH_{T}\geq 0]=1 and [A]=p>0\mathbb{P}[A]=p>0 for a set A={a~DHTg}A=\{\tilde{a}\star DH_{T}\geq g\} for some g>0g>0. Note that a~DHTg1A\tilde{a}\star DH_{T}\geq g\mbox{{\sl 1}}_{A}\,. Let nn\in\mathbb{N}. Then

U(na~DHT)sup:y𝔼[1Au(y+ng)y]=sup:ypu(y+ng)y=().U(n\tilde{a}\star DH_{T})\geq\sup{}_{y}:\ \mathbb{E}[\mbox{{\sl 1}}_{A}\,u(y+ng)-y]=\sup{}_{y}:\ pu(y+ng)-y=(*)\ .

The last term is optimized by y~=u(1/p)1ng\tilde{y}=u^{\prime}{}^{-1}(1/p)-ng as it solves pu(y~+ng)=1pu^{\prime}(\tilde{y}+ng)=1. Therefore, ()=pu(u(1/p)1)u(1/p)1+ng(*)=pu(u^{\prime}{}^{-1}(1/p))-u^{\prime}{}^{-1}(1/p)+ng which tends to infinity as nn\uparrow\infty. Hence, no finite maximizer of (5) exists. ∎

2.2 Utility-based risk neutral densities

In the absence of classic arbitrage, we are now able to use the optimized certainty equivalent framework to construct an equivalent martingale measure. Before moving on to our main result, we will need the following lemma.

Lemma 2.2.

Let f:f:\mathbb{R}\rightarrow\mathbb{R} be concave and C1C^{1}. Assume that f(ξ)L1f(\xi)\in L^{1} for all ξ\xi and that 𝔼[f(ξ)]𝔼[f(ξ)]\mathbb{E}[f(\xi^{*})]\geq\mathbb{E}[f(\xi)] for some ξ\xi^{*}. Then

ϵ|ϵ=0𝔼[f(ϵξ+ξ)]=𝔼[f(ξ)ξ].\partial_{\epsilon}|_{\epsilon=0}\mathbb{E}[f(\epsilon\xi+\xi^{*})]=\mathbb{E}[f^{\prime}(\xi^{*})\xi]. (6)
Proof.

Define Δϵ:=1ϵ(f(ϵξ+ξ)f(ξ))\Delta_{\epsilon}:=\frac{1}{\epsilon}\left(f(\epsilon\xi+\xi^{*})-f(\xi^{*})\right) such that Δϵϵ|ϵ=0f(ϵξ+ξ)=f(ξ)ξ\Delta_{\epsilon}\uparrow\partial_{\epsilon}|_{\epsilon=0}f(\epsilon\xi+\xi^{*})=f^{\prime}(\xi^{*})\xi since ff is concave. As a difference between two L1L^{1} variables ΔϵL1\Delta_{\epsilon}\in L^{1}. Since ξ\xi^{*} maximizes the expectation of ff, we also have 𝔼[Δϵ]0\mathbb{E}[\Delta_{\epsilon}]\leq 0. Using the dominated convergence shows that Δϵf(ξ)ξL1\Delta_{\epsilon}\uparrow f^{\prime}(\xi^{*})\xi\in L^{1} and therefore that taking expectations and derivatives in (6) can be exchanged. ∎

Now we give the main result allowing the construction of utility-based equivalent martingale measures.

Proposition 2.3.

Let yy^{*} and aa^{*} be finite maximizers of

y,aF(y,a):=𝔼[u(y+aDHT)y].y,a\longmapsto F(y,a):={\mathbb{E}}\left[\,{u\big{(}\,y+{a\star DH_{T}}\,\big{)}-y}\,\right]\ . (7)

Then,

D:=u(y+aDHT)D^{*}:=u^{\prime}\big{(}\,y^{*}+{a^{*}\star DH_{T}}\,\big{)} (8)

is an equivalent martingale density, i.e. the measure \mathbb{Q}^{*} defined via d:=Ddd\mathbb{Q}^{*}:=D^{*}d\mathbb{P} is an equivalent martingale measure.333 We note that if uu is not strictly increasing, then DD^{*} is an absolutely continuous, but possibly not equivalent density. An example is the CVaR “utility”u(x)=min{x,0}/(1α)u(x)=\min\{x,0\}/(1-\alpha).

Proof.

We follow broadly the discussion in Section 3.1 of [6].

Show that D(aDHT)L1D^{*}({a\star DH_{T}})\in L^{1} with zero expectation: optimality of cc^{*} and aa^{*} imply first 0=ϵ|ϵ=0F(y,ϵa+a)0=\partial_{\epsilon}\big{|}_{\epsilon=0}F(y^{*},\epsilon a+a^{*}). Secondly, lemma 2.2 shows for arbitrary aa that D(aDHTL1)D^{*}({a\star DH_{T}}\in L^{1}) with

0=ϵ|ϵ=0F(y,ϵa+a)=𝔼[DaDHT].0=\partial_{\epsilon}\big{|}_{\epsilon=0}F(y^{*},\epsilon a+a^{*})={\mathbb{E}}\left[\,{D^{*}\ {a\star DH_{T}}}\,\right]\ . (9)

Show that 𝔼t[DDHt]=0\mathbb{E}_{t}[D^{*}DH_{t}]=0: for the previous statement, set a:=(0,,1Atei,,0)a:=(0,\ldots,\mbox{{\sl 1}}_{A_{t}}\,e^{i},\ldots,0) where AtA_{t} is t{\cal F}_{t}-measurable, and where eie^{i} denotes the iith unit vector. We obtain

0=𝔼[DDHti|t].0=\mathbb{E}[D^{*}DH^{i}_{t}|{\cal F}_{t}]\ . (10)

Show that DL1D^{*}\in L^{1}: recall that aDHT=itatiDHti{a\star DH_{T}}=\mbox{$\sum_{it}$}a^{i}_{t}\cdot DH^{i}_{t}. Since uu is concave and strictly increasing, uu^{\prime} is decreasing and positive. Then,

0u(y+aDHT)\displaystyle 0\leq u^{\prime}(y^{*}+{a^{*}\star DH_{T}}) \displaystyle\leq u(y+aDHT)|DH|1|DH|>1\displaystyle u^{\prime}\left(y^{*}+{a^{*}\star DH_{T}}\right)|DH|\mbox{{\sl 1}}_{|DH|>1}\,
+u(yess supit|a|ti)1|DH|1L1\displaystyle+u^{\prime}\left(y^{*}-\mbox{ess $\sup_{it}$}|a^{*}{}^{i}_{t}|\right)\mbox{{\sl 1}}_{|DH|\leq 1}\,\in L^{1}

since yy^{*} and aa^{*} were assumed to be finite, and since the previous step with a=1a=1 implies u(y+aDHT)|DH|L1u^{\prime}\left(y^{*}+{a^{*}\star DH_{T}}\right)|DH|\in L^{1}.

Positivity of DD^{*}: since uu^{\prime} is decreasing and positive we have limnu(n)0\lim_{n\uparrow\infty}u^{\prime}(n)\geq 0 and therefore [u(aDHT)>0]=limn0[u(aDHT)>u(n)]=limn0[aDHTn]=[aDHT<]=1\mathbb{P}[u^{\prime}({a^{*}\star DH_{T}})>0]=\lim_{n\uparrow 0}\mathbb{P}[u^{\prime}({a^{*}\star DH_{T}})>u^{\prime}(n)]=\lim_{n\uparrow 0}\mathbb{P}[{a^{*}\star DH_{T}}\leq n]=\mathbb{P}[{a^{*}\star DH_{T}}<\infty]=1, since aa^{*} being almost surely finite implies that aDHTL1{a^{*}\star DH_{T}}\in L^{1}.

DD^{*} has unit expectation: optimality of yy^{*} and aa^{*} implies

0=y|y=yF(y,a)=𝔼[D]10=\partial_{y}\big{|}_{y=y^{*}}F(y,a^{*})={\mathbb{E}}\left[\,{D^{*}}\,\right]-1 (11)

and therefore that 𝔼[D]=1\mathbb{E}[D^{*}]=1.

The density DD^{*} provides an equivalent martingale density. It is minimal among all equivalent martingale densities in the following sense: the Legendre-Fenchel transform of the convex function f(x):=u(x)f(x):=-u(-x) is defined as.

u~(y):=supx{yxf(x)}.\tilde{u}(y):=\sup_{x\in\mathbb{R}}\{yx-f(x)\}\ .

The associated u~\tilde{u}-divergence between two distributions \mathbb{Q} and \mathbb{P} with \mathbb{Q}\ll\mathbb{P} is then

Df(|)=𝔼[u~(dd)].D_{f}(\mathbb{Q}|\mathbb{P})=\mathbb{E}\left[\tilde{u}\left(\frac{d\mathbb{Q}}{d\mathbb{P}}\right)\right]\ .

It is a non-symmetric measure of the similarity between two probability distributions.

Corollary 2.4.

Let u~(y)\tilde{u}(y) be the Legrendre-Fenchel transform of uu, and define DD^{*} as in (8). Then, DD^{*} is a minimizer of the u~\tilde{u}-divergence

D𝔼[u~(D))]D\longmapsto{\mathbb{E}}\left[\,{\tilde{u}\left(D)\right)}\,\right] (12)

over all equivalent martingale densities.

Proof.

The Legrende-Fenchel transform of the convex function u(x)-u(-x) is u~(y)=supx(yx+u(x))=supx(u(x)yx)\tilde{u}(y)=\sup_{x}(yx+u(-x))=\sup_{x}(u(x)-yx) which implies that for all xx,

u~(y)u(x)yx.\tilde{u}(y)\geq u(x)-yx\ . (13)

Let 𝒟e:={D>0:𝔼[D]=1,𝔼[D(aDHT)]=0for all a}{\cal D}_{e}:=\{\ D>0:\ \mathbb{E}[D]=1,\,\mathbb{E}[D\,({a\star DH_{T}})]=0\ \mbox{for all $a$}\ \} be the set of equivalent martingale densities. Equation (13) implies for yD𝒟ey\rightarrow D\in{\cal D}_{e} and xc+aDHTx\rightarrow c+{a\star DH_{T}},

infD𝒟e𝔼[u~(D)]supc,a{𝔼[u(c+aDHT)]c}=F(y,a).\inf_{D\in{\cal D}_{e}}{\mathbb{E}}\left[\,{\tilde{u}(D)}\,\right]\geq\sup_{c,a}\Big{\{}{\mathbb{E}}\left[\,{u\big{(}c+{a\star DH_{T}}\big{)}}\,\right]-c\Big{\}}=F(y^{*},a^{*})\ . (14)

Let I=u()1(0,)I=u^{\prime}{}^{-1}(\mathbb{R})\subseteq(0,\infty). For a given yIy\in I the sup in u~(y)=supx(u(x)yx)\tilde{u}(y)=\sup_{x}(u(x)-yx) is attained by x=u(y)1x=u^{\prime}{}^{-1}(y) which yields

u~(y)=u(u(y)1)yu(y)1\tilde{u}(y)=u\big{(}u^{\prime}{}^{-1}(y)\big{)}-y\ u^{\prime}{}^{-1}(y) (15)

for all yIy\in I and all xx\in\mathbb{R} as claimed above. Applying (15) to u~(D)\tilde{u}(D^{*}) yields that equality of both sides of (14), proving our claim that DD^{*} is indeed a minimizer of (12).

Thus, finding the u~\tilde{u}-minimal equivalent martingale measures is the dual problem of maximizing the expected utility. The key observation is that we now have a numerically efficient method to solving the primal problem via the application of machine learning methods.

In the case of the exponential utility, the u~\tilde{u}-divergence is the relative entropy of \mathbb{Q} with respect to \mathbb{P},

H(|)=𝔼[ddlogdd].H(\mathbb{Q}|\mathbb{P})=\mathbb{E}\left[\frac{d\mathbb{Q}}{d\mathbb{P}}\log\frac{d\mathbb{Q}}{d\mathbb{P}}\right]\ .

The measure  \mathbb{Q}^{*} is the minimal entropy martingale measure (MEMM) introduced by [7], given by

dd=eaDHT𝔼[eaDHT]\frac{d\mathbb{Q}^{*}}{d\mathbb{P}}=\frac{e^{-{a^{*}\star DH_{T}}}}{\mathbb{E}[e^{-{a^{*}\star DH_{T}}}]}

The measure is unique due to the strict convexity of the function u~(y)=ylogy\tilde{u}(y)=y\log y.

Remark 1.

In the case where the returns are normally distributed, and the utility is the exponential utility, then the optimization is easily shown to be equivalent to solving the classic mean-variance objective of Markowitz [11] U(X)=𝔼[X]1/2λVar[X]U(X)=\mathbb{E}[X]-1/2\lambda Var[X], and in this case the found martingale measure removes the drift while preserving the covariance of the returns.

Direct Construction of Equivalent Martingale Measures

An alternative to the above construction is described in [6] Section 3.1, as follows.

Proposition 2.5.

Define  uu as above and fix some initial wealth w0w_{0}\in\mathbb{R}. Let aa^{*} be a finite maximizer of

a𝔼[u(w0+aDHT)].a\longmapsto{\mathbb{E}}\left[\,{u\big{(}\,w_{0}+{a\star DH_{T}}\,\big{)}}\,\right]\ . (16)

Then, the measure \mathbb{Q}^{*} with density

D:=u(w0+aDHT)𝔼[u(w0+aDHT)]D^{*}:=\frac{u^{\prime}(w_{0}+{a^{*}\star DH_{T}})}{{\mathbb{E}}\left[\,{u^{\prime}(w_{0}+{a^{*}\star DH_{T}})}\,\right]} (17)

is an equivalent martingale measure.

Proof.

We prove that \mathbb{Q}^{*} is a martingale measure.444We follow broadly section 3.1 in [6].

Show 𝔼[u(w0+aDHT)aDHT]=0\mathbb{E}[u^{\prime}(w_{0}+{a^{*}\star DH_{T}})\,{a\star DH_{T}}]=0 for all aa. For an arbitrary aa we get

0=ϵ|ϵ=0𝔼[u((w0+ϵa+a)DHT)]=()𝔼[u(w0+aDHT)aDHT],0=\partial_{\epsilon}|_{\epsilon=0}\mathbb{E}[u(\,(w_{0}+\epsilon a+a^{*})\star DH_{T}\,)]\stackrel{{\scriptstyle\mbox{{\scriptsize${(*)}$}}}}{{{=}}}\mathbb{E}[u^{\prime}(w_{0}+{a^{*}\star DH_{T}})\,{a\star DH_{T}}]\ , (18)

where ()(*) follows from Lemma 2.2. Given that aa was arbitrary above also implies 𝔼[u(w0+aDHT)DHt|t]=0\mathbb{E}[u^{\prime}(w_{0}+{a^{*}\star DH_{T}})\,DH_{t}|{\cal F}_{t}]=0.

We first prove that u(aDHT)L1u^{\prime}({a^{*}\star DH_{T}})\in L^{1}: recall that aDHT=itatiDHti{a\star DH_{T}}=\mbox{$\sum_{it}$}a^{i}_{t}\cdot DH^{i}_{t}. Since uu is concave and increasing, 0u(x)u(xϵ)0\leq u^{\prime}(x)\leq u^{\prime}(x-\epsilon) for ϵ>0\epsilon>0. Then,

0u(w0+aDHT)\displaystyle 0\leq u^{\prime}(w_{0}+{a^{*}\star DH_{T}}) \displaystyle\leq u(w0+aDHT)|DH|1|DH|>1\displaystyle u^{\prime}\left(w_{0}+{a^{*}\star DH_{T}}\right)|DH|\mbox{{\sl 1}}_{|DH|>1}\,
+u(ess supit|a|ti)1|DH|1L1\displaystyle+u^{\prime}\left(-\mbox{ess $\sup_{it}$}|a^{*}{}^{i}_{t}|\right)\mbox{{\sl 1}}_{|DH|\leq 1}\,\in L^{1}

since aa^{*} was assumed to be finite.

Positivity of DD^{*}: since uu^{\prime} is decreasing and positive we have limnu(n)=0\lim_{n\uparrow\infty}u^{\prime}(n)=0 and therefore [u(w0+aDHT)>0]=limn0[u(w0+aDHT)>u(n)]=limn0[w0+aDHTn]=[w0+aDHT<]=1\mathbb{P}[u^{\prime}(w_{0}+{a^{*}\star DH_{T}})>0]=\lim_{n\uparrow 0}\mathbb{P}[u^{\prime}(w_{0}+{a^{*}\star DH_{T}})>u^{\prime}(n)]=\lim_{n\uparrow 0}\mathbb{P}[w_{0}+{a^{*}\star DH_{T}}\leq n]=\mathbb{P}[w_{0}+{a^{*}\star DH_{T}}<\infty]=1, since aa^{*} being almost surely finite implies that aDHTL1{a^{*}\star DH_{T}}\in L^{1}

Remark 2.

We note that the assumption of finiteness of aa^{*} again excludes markets with classic arbitrage opportunities.555 Assume that a~\tilde{a} is a classic arbitrage opportunity with [a~DHT0]=1\mathbb{P}[\tilde{a}\star DH_{T}\geq 0]=1 and [A]=p>0\mathbb{P}[A]=p>0 for a set A={a~DHTg}A=\{\tilde{a}\star DH_{T}\geq g\} for g>0g>0. Then 𝔼[u(w0+(na~1A)DHT)]𝔼[1Au(w0+ng)]+(1p)u(w0)pu(w0+ng)+(1p)u(w0)pu()+(1p)u(w0)\mathbb{E}[u(w_{0}+(n\tilde{a}1_{A})\star DH_{T})]\geq\mathbb{E}[1_{A}u(w_{0}+ng)]+(1-p)u(w_{0})\geq p\,u(w_{0}+ng)+(1-p)u(w_{0})\uparrow p\,u(\infty)+(1-p)u(w_{0}), e.g. no finite maximizer of 𝔼[u(aDHT)]\mathbb{E}[u({a\star DH_{T}})] exists. If uu is the exponential utility, then DD^{*} coincides with the previously defined density of the MEMM in (8).

We note that while this approach is somewhat more direct it depends on initial wealth – except in the case of the exponential utility – and lacks the interpretation of the density as a minimizer of some distance to \mathbb{P}.

We now briefly discuss some extensions of the previous results to the cases of unbounded assets, and continuous time processes.

Unbounded Assets

As pointed out in [6] Section 3.1, the requirement u(aDHT)L1u({a\star DH_{T}})\in L^{1} can be enforced at the cost of interpretability of our previous results by passing over to bounded asset prices: to this end define the random variable M:=maxi,t|DHti|M:=\max_{i,t}|DH_{t}^{i}| and set DH¯ti:=DHti/(1+M)D\bar{H}^{i}_{t}:=DH^{i}_{t}/(1+M), which are now bounded.

We can then show with the same steps as before that we can construct an equivalent martingale measure in this case as follows.

Proposition 2.6.

Let yy^{*} and aa^{*} be maximizers of the bounded problem

y,aF¯(y,a):=𝔼[u(y+aDHT1+M)y].y,a\longmapsto\bar{F}(y,a):={\mathbb{E}}\left[\,{u\left(\frac{y+{a\star DH_{T}}}{1+M}\right)-y}\,\right]\ . (19)

Then,

D:=u(y+aDHT1+M)1+MD^{*}:=\frac{u^{\prime}\!\left(\frac{y^{*}+{a^{*}\star DH_{T}}}{1+M}\right)}{1+M} (20)

is an equivalent martingale density for the unscaled problem, i.e. 𝔼[aDHT]0\mathbb{E}^{*}[{a\star DH_{T}}]\leq 0 for all aa.

Moreover, DD^{*} minimizes the scaled u~\tilde{u}-divergence

D𝔼[u~((1+M)D)]D\rightarrow{\mathbb{E}}\left[\,{\tilde{u}\big{(}(1+M)D\big{)}}\,\right] (21)

over all equivalent martingale densities.

Proof.

We cover the main differences to the previous case: first, we see that

0=yF(y,a)=𝔼[u(y+aDHT1+M)1+M1].0=\partial_{y}F(y^{*},a^{*})={\mathbb{E}}\left[\,{\frac{u^{\prime}\left(\frac{y^{*}+{a^{*}\star DH_{T}}}{1+M}\right)}{1+M}-1}\,\right]\ . (22)

Then,

0=ϵF(y,ϵa+a)=𝔼[u(y+aDHT1+M)1+MaDHT]0=\partial_{\epsilon}F(y^{*},\epsilon a+a^{*})={\mathbb{E}}\left[\,{\frac{u^{\prime}\left(\frac{y^{*}+{a^{*}\star DH_{T}}}{1+M}\right)}{1+M}{a\star DH_{T}}}\,\right] (23)

showing that DD^{*} is an equivalent martingale density. Using (13) with y(1+M)Dy\rightarrow(1+M)D for D𝒟eD\in{\cal D}_{e} and z(c+aDHT)/(1+M)z\rightarrow(c+{a\star DH_{T}})/(1+M) yields as before

infD𝒟e𝔼[g~((1+M)D)]supc,a{𝔼[u(c+aDHT1+M)]c}.\inf_{D\in{\cal D}_{e}}{\mathbb{E}}\left[\,{\tilde{g}\big{(}\,(1+M)D\,\big{)}}\,\right]\geq\sup_{c,a}\Big{\{}{\mathbb{E}}\left[\,{u\left(\frac{c+{a\star DH_{T}}}{1+M}\right)}\,\right]-c\Big{\}}\ . (24)

Equality in DD^{*} follows as before. ∎

Continuous Time

We note that our method of proof also works in a continuous time: let Gt(a):=0tar𝑑HrG_{t}(a):=\int_{0}^{t}\!\!\,a_{r}\,dH_{r} where dHt=μtdt+σtdWtdH_{t}=\mu_{t}\,dt+\sigma_{t}\,dW_{t} (i.e. the classic setup with fixed instruments). Assume that y,ay^{*},a^{*} maximize y,a𝔼[u(y+GT(a))y]y,a\mapsto\mathbb{E}[u(y+G_{T}(a))-y] and that 𝔼[u(GT(a))]<\mathbb{E}[u(G_{T}(a^{*}))]<\infty which again excludes markets with classic arbitrage. Then, then same statement as above is true with virtually the same proof.

Let Gt:=Gt(a)G^{*}_{t}:=G_{t}(a^{*}) and notice that if u3u\in\mathbb{C}^{3} then f(x):=logu(x)f(x):=\log u^{\prime}(x) has derivative f(x)=u′′(x)u(x)f^{\prime}(x)=\frac{u^{\prime\prime}(x)}{u^{\prime}(x)} which is the Arrow-Pratt coefficient of absolute risk aversion of uu, c.f. [6] section 2.3. Standard calculus shows that

Dt=exp(0tu′′(Gr)u(Gr)σr𝑑Wr120t(u′′(Gr)u(Gr)σr)2𝑑r)D^{*}_{t}=\exp\left(\int_{0}^{t}\!\!\frac{u^{\prime\prime}(G^{*}_{r})}{u^{\prime}(G^{*}_{r})}\sigma_{r}\,dW_{r}-\frac{1}{2}\int_{0}^{t}\!\!\left(\frac{u^{\prime\prime}(G^{*}_{r})}{u^{\prime}(G^{*}_{r})}\sigma_{r}\right)^{2}\!\!dr\right) (25)

Under \mathbb{Q}^{*} our assets are driftless and satisfy dHt=σtdWtdH_{t}=\sigma_{t}\,dW^{*}_{t} for a \mathbb{Q}^{*}-Brownian motion WW^{*}. This implies the well-known result

u′′(Gt)u(Gt)=μtσt2.\frac{u^{\prime\prime}(G^{*}_{t})}{u^{\prime}(G^{*}_{t})}=\frac{\mu_{t}}{\sigma^{2}_{t}}\ . (26)

3 Transaction costs and trading constraints

The previous section enables us to simulate markets from a martingale measure in the absence of trading frictions. In practise, trading strategies will be subject to trading cost and constraints such as liquidity and risk limits. Our use-case is training a Deep Hedging agent. We therefore now extend the previous results to the case of generalized cost functions which will cover both trading cost and most trading constraints.

generalized cost function is a non-negative, t\mathcal{F}_{t}-measurable function ct(at)c(at;st)c_{t}(a_{t})\equiv c(a_{t};s_{t}) with values in [0,][0,\infty], which is convex in ata_{t}, lower semi-continuous, and normalized to ct(0)=0c_{t}(0)=0. To impose convex restrictions on our trading activity, we set transaction cost to infinity outside the admissible set. Indeed, let 𝒜t{\cal A}_{t} be be a convex set of admissible trading actions, and c¯t\bar{c}_{t} an initial const function. We then use ct(at):=c¯t(at)+1at𝒜tc_{t}(a_{t}):=\bar{c}_{t}(a_{t})+\infty\mbox{{\sl 1}}_{a_{t}\not\in{\cal A}_{t}}\,. (We note that this construction is lower semi-continuous.) As example, let us assume the iith instrument is not tradable in tt. We then impose ct(at)=c_{t}(a_{t})=\infty whenever |ati|>0|a^{i}_{t}|>0.

In reverse, if ctc_{t} is a generalized cost function, we may call 𝒜t:={an:ct(a)<}{\cal A}_{t}:=\{a\in\mathbb{R}^{n}:\,c_{t}(a)<\infty\} the convex set of admissible actions. Note also that by construction 0𝒜t0\in{\cal A}_{t}.

Example 3.1.

The simplest trading costs are proportional. Assume that Δt\Delta_{t} and Vt\mathrm{V}_{t} are observable Black & Scholes delta and vega of the mid-prices Ht(t)H^{(t)}_{t} for the trading instruments available at tt, and that the cost of trading aia^{i} units of Htt,iH^{t,i}_{t} is proportional to its delta and vega with cost factors gΔ±g^{\pm}_{\Delta} and gV±g^{\pm}_{\mathrm{V}} for buying and selling, respectively. We also impose that we may trade at most Vmax\mathrm{V}_{\mathrm{max}} units of vega per time step. The corresponding cost function is given by

ct(a):={a+(gΔ+Δt+gV+Vt)aVtVmax+a(gΔΔt+gVVt)aVt>Vmaxc_{t}(a):=\left\{\begin{array}[]{ll}a^{+}\cdot\left(g^{+}_{\Delta}\Delta_{t}+g^{+}_{\mathrm{V}}\mathrm{V}_{t}\right)&a\cdot\mathrm{V}_{t}\leq\mathrm{V}_{\mathrm{max}}\\ \ \ \ \ \ +a^{-}\cdot\left(g^{-}_{\Delta}\Delta_{t}+g^{-}_{\mathrm{V}}\mathrm{V}_{t}\right)&\\ \infty&a\cdot\mathrm{V}_{t}>\mathrm{V}_{\mathrm{max}}\end{array}\right. (27)
Example 3.2.

Consider trading cost which apply only to net delta and vega traded, e.g.

ct(a):={gΔ+(aΔt)++gV+(aVt)+aVtVmax+gΔ(aΔt)+gV(aVt)aVt>Vmaxc_{t}(a):=\left\{\begin{array}[]{ll}g^{+}_{\Delta}\left(a\cdot\Delta_{t}\right)^{+}+g^{+}_{\mathrm{V}}\left(a\cdot\mathrm{V}_{t}\right)^{+}&a\cdot\mathrm{V}_{t}\leq\mathrm{V}_{\mathrm{max}}\\ \ \ \ \ \ +g^{-}_{\Delta}\left(a\cdot\Delta_{t}\right)^{-}+g^{-}_{\mathrm{V}}\left(a\cdot\mathrm{V}_{t}\right)^{-}&\\ \infty&a\cdot\mathrm{V}_{t}>\mathrm{V}_{\mathrm{max}}\end{array}\right. (28)

The terminal gain of implementing a trading policy aa with cost function cc is given by

aDHTCT(a)whereCT(a):=t=0T1ct(at).{a\star DH_{T}}-C_{T}(a)\ \ \ \mbox{where}\ \ \ C_{T}(a):=\sum_{t=0}^{T-1}c_{t}(a_{t})\ . (29)

The marginal cost of trading small quantities of the iith asset in tt are given as

γti+:=+ϵ>0ct(ϵei)andγti:=ϵ>0ct(ϵei).\gamma^{i+}_{t}:=+\partial_{\epsilon>0}c_{t}(\epsilon e^{i})\ \ \ \mbox{and}\ \ \ \gamma^{i-}_{t}:=-\partial_{\epsilon>0}c_{t}(-\epsilon e^{i})\ . (30)

They define the marginal cost function

mt(a):=a+γt+aγt,m_{t}(a):=a^{+}\cdot\gamma^{+}_{t}-a^{-}\cdot\gamma^{-}_{t}\ , (31)

3.1 Statistical arbitrage and near-martingale measures

Under the statistical measure we expect there to be statistical arbitrage opportunities, i.e. trading strategies aa such that we expect to make money:

𝔼[aDHT]>0.\mathbb{E}[{a\star DH_{T}}]>0\ . (32)

In the absence of transaction costs, the market will be free from statistical arbitrage if and only if we are under a martingale measure.666 Assume that there is a tt such that ft:=𝔼[DHt|t]0f_{t}:=\mathbb{E}[DH_{t}|{\cal F}_{t}]\not=0. Set at:=signfta_{t}:=\mathrm{sign}\,f_{t}. Then the strategy a=(0,,at,,0)a=(0,\ldots,a_{t},\ldots,0) is a statistical arbitrage strategy. Since the gains of trading with transaction costs are almost surely never greater than the gains in the absence of transaction costs, it is clear that if \mathbb{Q} is an equivalent martingale measure for the market, then there are no statistical arbitrage opportunities under transaction cost, either, i.e. 𝔼[aDHTCT(a)]0\mathbb{E}_{\mathbb{Q}}\left[\,{a\star DH_{T}}-C_{T}(a)\right]\leq 0 for all policies aa (equality is acheived with a0a\equiv 0). Taking the limit to small transaction cost, it becomes inutitively clear that 𝔼[aDHTMT(a)]0\mathbb{E}_{\mathbb{Q}}\left[\,{a\star DH_{T}}-M_{T}(a)\right]\leq 0 as well for marginal cost. In fact, inuitively it makes sense that the market is free of statistical arbitrage with full cost cc if and only if it is free of statistical arbitrage with marginal cost mm.

Here is our formal result:

Proposition 3.3.

We call \mathbb{Q} a near martingale measure if any of the following equivalent conditions hold:

  • the measure \mathbb{Q} is free from statistical arbitrage with full cost cc;

  • the measure \mathbb{Q} is free from statistical arbitrage with marginal cost mm; and

  • the expected return from any hedging instrument is within its marginal bid/ask spread in the sense that

    Ht(t,i)γtiMarginalbid price𝔼[HT(t,i)|t]Expected gainsHt(t,i)+γti+Marginalask price,\underbrace{H^{(t,i)}_{t}-\gamma^{i-}_{t}}_{\begin{array}[]{c}\mbox{Marginal}\\ \mbox{bid price}\end{array}}\leq\underbrace{\mathbb{E}_{\mathbb{Q}}\big{[}H^{(t,i)}_{T}\big{|}\mathcal{F}_{t}\big{]}}_{\mbox{Expected gains}}\leq\underbrace{H^{(t,i)}_{t}+\gamma^{i+}_{t}}_{\begin{array}[]{c}\mbox{Marginal}\\ \mbox{ask price}\end{array}}\ , (33)

    with γti±\gamma^{i\pm}_{t} defined in (30).

Proof.

Assume first there are no statistical arbitrage opportunities with full cost cc. We will show (33). Let AtA\in{\cal F}_{t} arbitrary and let etie^{i}_{t} the policy with unit vector eie^{i} at tt and zero elsewhere; for ease of notation we will also write etie^{i}_{t} for simply the unit vector, seen at a time tt.

Absence of statistical arbitrage implies that 01ϵ𝔼[(±ϵeti1A)DHTCT(±ϵeti1A)]0\geq\frac{1}{\epsilon}\mathbb{E}_{\mathbb{Q}}\big{[}{(\pm\epsilon e^{i}_{t}1_{A})\star DH_{T}}-C_{T}(\pm\epsilon e^{i}_{t}1_{A})\big{]} for all ϵ>0\epsilon>0, and therefore

0ϵ>0𝔼[1A{±ϵetiDHTCT(±ϵeti)}]=𝔼[1A{DHtiγti±}]0\geq\partial_{\epsilon>0}\mathbb{E}_{\mathbb{Q}}\big{[}1_{A}\left\{\pm\epsilon{e^{i}_{t}\star DH_{T}}-C_{T}(\pm\epsilon e^{i}_{t})\right\}\big{]}=\mathbb{E}_{\mathbb{Q}}\big{[}1_{A}\ \big{\{}DH_{t}^{i}\mp\gamma^{i\pm}_{t}\big{\}}\big{]}

which yields (33).

Assume now that (33) holds, and let aa be arbitrary. Then 𝔼[attDHtmt(at)]0\mathbb{E}_{\mathbb{Q}}[\sum{}_{t}a_{t}\cdot DH_{t}-m_{t}(a_{t})]\leq 0 by construction of our marginal cost (31) and (33). Hence, there is no statistical arbitrage with cost mm. Since ctmtc_{t}\geq m_{t} it is also clear that if there is no statistical arbitrage with marginal cost mm, then there is also no statistical arbitrage with full cost cc. ∎

Remark 3.

Under the conditions of the above theorem the conditional expectation 𝔼[HT(t,i)|t]\mathbb{E}_{\mathbb{Q}}\big{[}H^{(t,i)}_{T}\big{|}\mathcal{F}_{t}\big{]} defines a martingale “micro-price” [14] within the bid–ask spread.

In the absence of trading costs or trading constraints then equality is acheived. That is, the market is free from statistical arbitrage if and only if

𝔼[HT(t,i)|t]=Ht(t,i).\mathbb{E}_{\mathbb{Q}}\big{[}H^{(t,i)}_{T}\big{|}\mathcal{F}_{t}\big{]}=H^{(t,i)}_{t}\ .

resulting in the classic formulation of the price process being a martingale under \mathbb{Q}.

3.2 Utility-based near-martingale measures under trading frictions

We now proceed with constructing a near-martingale measure \mathbb{Q}^{*} via the same duality as in the zero transaction cost case. Define again the function

F(y,a):=𝔼[u(y+aDHTMT(a))y]F(y,a):={\mathbb{E}}\left[\,{u\big{(}y+{a\star DH_{T}}-M_{T}(a)\big{)}-y}\,\right] (34)

just as in (7), but this time with marginal transaction costs.

Proposition 3.4.

Let yy^{*} and aa^{*} be finite maximizers of y,aF(y,a)y,a\mapsto F(y,a). Then

D:=u(y+aDHTMT(a))D^{*}:=u^{\prime}\!\left(y^{*}+{a^{*}\star DH_{T}}-M_{T}(a^{*})\right) (35)

is an equivalent density, and the measure \mathbb{Q}^{*} defined by d:=Ddd\mathbb{Q}^{*}:=D^{*}d\mathbb{P} is a near-martingale measure. Moreover, the density DD^{*} minimizes the u~\tilde{u}-divergence among all equivalent near-martingale densities.

Proof.

To show that DD^{*} is a equivalent near- martingale density most of the previous proof applies as before, except of course (9) since DD^{*} is not an equivalent martingale measure. Instead, we will show that there is no statistical arbitrage under \mathbb{Q}^{*}. Let therefore AtA\in{\cal F}_{t} be arbitrary, and denote by etie^{i}_{t} the strategy with unit vector eie^{i} in tt and zero otherwise; for notational simplicity we will also use etie^{i}_{t} to refer simply the unit vector, in tt.

Define F±(ϵ):=±𝔼[u(y+(±ϵ1Aeti+a)DHTMT(±ϵ1Aeti+a))y]F_{\pm}(\epsilon):=\pm{\mathbb{E}}\left[\,{u\big{(}y^{*}+{(\pm\epsilon{\mbox{{\sl 1}}_{A}\,}e^{i}_{t}+a^{*})\star DH_{T}}-M_{T}(\pm\epsilon\mbox{{\sl 1}}_{A}\,e^{i}_{t}+a^{*})\big{)}-y^{*}}\,\right]. Consider the derivative ϵF±(0)=𝔼[1A{DHti±ϵ>0mt(±ϵeti+at)}]\partial_{\epsilon}F_{\pm}(0)={\mathbb{E}^{*}}\left[\,{\mbox{{\sl 1}}_{A}\,\left\{DH^{i}_{t}\pm\partial_{\epsilon>0}m_{t}(\pm\epsilon e^{i}_{t}+a^{*}_{t})\right\}}\,\right]: we recall that mt(a)=at+γt+atγtm_{t}(a)=a^{+}_{t}\cdot\gamma^{+}_{t}-a^{-}_{t}\gamma^{-}_{t}. Therefore

()=±ϵ>0mt(±ϵeti+at)={γti+if ati>0,+γtiif ati<0,γti±if ati=0.(*)=\pm\partial_{\epsilon>0}m_{t}(\pm\epsilon e^{i}_{t}+a^{*}_{t})=\left\{\begin{array}[]{ll}-\gamma^{i+}_{t}&\mbox{if $a^{*i}_{t}>0$,}\\ +\gamma^{i-}_{t}&\mbox{if $a^{*i}_{t}<0$,}\\ \mp\gamma^{i\pm}_{t}&\mbox{if $a^{*i}_{t}=0$.}\\ \end{array}\right.

Since (a,y)(a^{*},y^{*}) are optimal we must have 0[min(),max()]0\in[\min(*),\max(*)]. Given that AtA\in{\cal F}_{t} was arbitrary we obtain

{𝔼[DHt|t]=+γti+if ati>0,γti±𝔼[DHt|t]+γti+if ati=0, and γti=𝔼[DHt|t]if ati<0.\left\{\begin{array}[]{rlll}&\mathbb{E}^{*}[DH_{t}|{\cal F}_{t}]&=+\gamma^{i+}_{t}&\mbox{if $a^{*i}_{t}>0$,}\\ -\gamma^{i\pm}_{t}\leq&\mathbb{E}^{*}[DH_{t}|{\cal F}_{t}]&\leq+\gamma^{i+}_{t}&\mbox{if $a^{*i}_{t}=0$, and }\\ -\gamma^{i-}_{t}=&\mathbb{E}^{*}[DH_{t}|{\cal F}_{t}]&&\mbox{if $a^{*i}_{t}<0$.}\\ \end{array}\right. (36)

This is in fact a more precise statement than (33).

We now show that DD^{*} minimizes the u~\tilde{u}-divergence among all measures D𝒟e:={D>0:𝔼[D]=1,𝔼[D(aDHTMT(a))]0for all a}D\in{\cal D}_{e}:=\{\ D>0:\ \mathbb{E}[D]=1,\mathbb{E}[D\,({a\star DH_{T}}-M_{T}(a))]\leq 0\ \mbox{for all~$a$}\ \}. We apply (13) again with yD𝒟ey\rightarrow D\in{\cal D}_{e} and xy+aDHTMT(a)x\rightarrow y+{a\star DH_{T}}-M_{T}(a). This yields

𝔼[u~(D)]𝔼[u(y+aDHTMT(a))]𝔼[D(y+aDHTMT(a))]𝔼[u(y+aDHTMT(a))y],\mathbb{E}[\tilde{u}(D)]\geq\mathbb{E}[u(y+{a\star DH_{T}}-M_{T}(a))]-\mathbb{E}[D\,(y+{a\star DH_{T}}-M_{T}(a))]\geq\mathbb{E}[u(y+{a\star DH_{T}}-M_{T}(a))-y]\ , (37)

where the last inequality holds since DD does not admit statistical arbitrage. The right hand side is maximized in (a,y(a^{*},y^{*}). For the left hand side, apply again (15) which yields

𝔼[u~(D)]=𝔼[u(y+aDHTMT(a))y]𝔼[aDHTMT(a)]=0.\mathbb{E}[\tilde{u}(D^{*})]=\mathbb{E}[u(y^{*}+{a^{*}\star DH_{T}}-M_{T}(a^{*}))-y^{*}]-\underbrace{\mathbb{E}^{*}[{a^{*}\star DH_{T}}-M_{T}(a^{*})]}_{=0}\ . (38)

This proves that DD^{*} is u~\tilde{u}-minimal among all near-martingale measures. ∎

Considering that any equivalent true martingale measure is also a near-martingale measure, this result is a formalization of the intuitive notion that in order to avoid statistical arbitrage we do not truly have to find a full martingale measure, but that we only have to “bend” the drifts of our trading instruments enough to be dominated by prevailing trading cost.

4 Learning to Simulate Risk-Neutral Dynamics

The key insight of our utility-based risk-neutral density construction is that it relies only on solving the optimization problem of find aa^{*} and yy^{*}, not on specifying any particular dynamics for the market under the \mathbb{P} measure. Therefore, it can be done in a data-driven, model agnostic way, lending itself to the application of modern machine learning methods. Specifically, given a set of NN samples from a \mathbb{P} market simulator, we may make the sample set risk neutral by numerically solving the optimization problem on the NN paths, and then using our formulation to reweight the paths so that the resulting weighted sample is a (near-)martingale. As mentioned above, this is particularly useful in the case of removing statistical arbitrage from a “black box” market simulator, such as the GAN based approach discussed in [1].

Our approach enables the adaptation of GAN and other advanced machine learning approaches so that they can not only simulate realistic samples from the statistical measure, but also from an equivalent risk neutral measure. Moreover, through the choice of utility function, we are able to control the risk neutral measure we construct.

We solve the stochastic control problem (7) through an application of the ‘Deep Hedging’ methods of [3]: we can pose (7) as a reinforcement learning problem and use a neural network to represent our trading policy aa, and since the function FF is fully differentiable, use stochastic gradient methods to find a,ya^{*},y^{*}, and hence DD^{*}.

4.1 Deep Hedging under Risk-Neutral Dynamics

Our primary application is in the pricing and hedging of exotic options via utility-based Deep Hedging. With a portfolio of derivatives represented by the random variable ZZ to hedge, the Deep Hedging problem under the statistical measure \mathbb{P} is to maximize the optimized certainty equivalent

𝕌(Z):=supa,y𝔼[u(y+Z+aDHTCT(a))y]\mathbb{U}_{\mathbb{P}}(Z):=\sup_{a,y}\ {\mathbb{E}_{\mathbb{P}}}\left[\,{u\big{(}y+Z+{a\star DH_{T}}-C_{T}(a)\big{)}-y}\,\right] (39)

over strategies aa and yy\in\mathbb{R}. An optimal solution aa^{*}_{\mathbb{P}} is called an optimal hedge for ZZ. We note that in the presence of statistical arbitrage 𝕌(0)>0\mathbb{U}^{*}(0)>0. Deep Hedging under a near-martingale measure \mathbb{Q}^{*} then is

𝕌(Z):=supa,y𝔼[u(y+Z+aDHTCT(a))y].\mathbb{U}^{*}(Z):=\sup_{a,y}\ \mathbb{E}^{*}\left[u\big{(}y+Z+{a\star DH_{T}}-C_{T}(a)\big{)}-y\right]. (40)

Since under \mathbb{Q}^{*} we have 𝕌(0)=0\mathbb{U}^{*}(0)=0 we note that 𝕌\mathbb{U}^{*} represents an indifference price for ZZ in the sense of [3] section 3.

In the case of hedging under exponential utility with zero transaction costs, it is straightforward to show that the optimal hedge for the derivative ZZ under the statistical measure can be written as

a=a+a0,a^{*}_{\mathbb{P}}=a_{\mathbb{Q}}^{*}+a_{0}^{*}\ , (41)

where aa_{\mathbb{Q}}^{*} is an optimal hedge for ZZ under the minimal entropy martingale measure (MEMM), and where a0a_{0}^{*} is an optimal statistical arbitrage strategy, i.e. an optimal “hedge” for an empty initial portfolio. In this sense we may regard aa^{*}_{\mathbb{Q}} as a net hedging strategy for ZZ.

A consequence of this is that the hedge found by solving the Deep Hedging problem under the statistical measure will be a sum of a true hedge, and a component which does not depend on ZZ and is simply seeking profitable opportunities in the market. Solving the optimization problem under the risk neutral measure will then directly remove the statistical arbitrage component of the strategy, leaving a “clean” hedge for the derivative, which is not sensitive to the estimation of the mean returns of our hedging instruments.

5 Numerical implementations

To demonstrate our approach we apply it to two market simulators. First, we discuss a simple, but usable multivariate “PCA” vector autoregressive model for spot and a form of implied volatilities. Secondly, we also present results for a Generative Adversarial Network based simulator based on the ideas presented in [1].

5.1 Vector Autoregressive market simulator

For the first numerical experiment, we build a VAR market simulator as follows. For the simulation we use discrete local volatilies (DLVs) as arbitrage-free parametrization of option prices. We do not use the underlying model dynamics; the only use of DLVs is arbitrage-free parametrization of the option surface. We briefly recap the relevant notation: assume thaty 0=τ0<τ1<<τm0=\tau_{0}<\tau_{1}<\cdots<\tau_{m} are time-to-maturities and 0<x1<<1<<xn0<x_{1}<\ldots<1<\ldots<x_{n} relative strikes.777See [4] for the use of inhomogeneous strike grids. Also define the additonal boundary strikes 0x0x10\leq x_{0}\ll x_{1} and xn+1:=1+2xnxnx_{n+1}:=1+2x_{n}\gg x_{n}.

For i=1,,ni=1,\ldots,n and j=1,,mj=1,\ldots,m we denote by Cj,iC^{j,i} the price of the call option with payoff (Sτj/S0xi)+(S_{\tau_{j}}/S_{0}-x_{i})^{+} at maturity τj\tau_{j}. Define Δj,i:=Cj,i+1Cj,ixi+1xi\Delta^{j,i}:=\frac{C^{j,i+1}-C^{j,i}}{x_{i+1}-x_{i}}, Γj,i:=Δj,iΔj,i1\Gamma^{j,i}:=\Delta^{j,i}-\Delta^{j,{i-1}} and Θj,i:=Cj,iCj,i1τjτj1\Theta^{j,i}:=\frac{C^{j,i}-C^{j,i-1}}{\tau_{j}-\tau_{j-1}}. The discrete local volatility surface (σj,i)j,i(\sigma^{j,i})_{j,i} for j=1,,mj=1,\ldots,m and i=1,,ni=1,\ldots,n is defined by

σj,i:=2Θj,ixj,iΓj,i2,\sigma^{j,i}:=\sqrt{\frac{2\,\Theta^{j,i}}{x^{j,i}{}^{2}\Gamma^{j,i}}}\ , (42)

where we set σj,i=\sigma^{j,i}=\infty whenever the square root is imaginary. We also set σj,i=0\sigma^{j,i}=0 if 0/00/0 occurs. We recall that a given surface of option prices is free of static arbitrage if and only if σ<\sigma<\infty. Moreover, given a surface of finite discrete local volatilities, we can reconstruct the surface of arbitrage-free call prices by solving the implicit finite difference scheme implied by (42). This involves inverting sequentially mm tridiagonal matrices, an operation which is “on graph” in modern automatic adjoint differentiation (AAD) machine learning packages such as TensorFlow, see [4] for further details.

For a given time series of vectors of historical log spot returns and log DLVs

Yr=(logSrSr1,logσr1,1,,logσrm,n),Y_{r}=\left(\log\frac{S_{r}}{S_{r-1}},\log\sigma^{1,1}_{r},\ldots,\log\sigma^{m,n}_{r}\right)^{\prime}\ , (43)

we estimate a vector autoregression model of the form

Yr=(BA1Yr1A2Yr2)dt+dtZr,Zr𝒩(0,Σ)Y_{r}=\left(B-A_{1}Y_{r-1}-A_{2}Y_{r-2}\right)dt+\sqrt{dt}Z_{r}\ ,\ \ \ Z_{r}\sim{\cal N}(0,\Sigma) (44)

where each A1,2A_{1,2} is a mn+1×mn+1mn+1\times mn+1 coefficient matrix, B=(B0,B1,,Bmn)B=(B_{0},B_{1},\ldots,B_{mn})^{\prime} is an intercept, and Σ\Sigma is a volatility matrix.

Constructing \mathbb{P} we train the model to historical data from EURO STOXX 50, using standard regression techniques from the Statsmodels Python package [13]. Once the model has been trained, we can simulate new sample paths of log spot returns and discrete local volatilities by sampling new noise variables ZrZ_{r} and stepping the model forward. We then convert the DLVs to option prices using the methods detailed above, so that we can simulate market states of spot and option prices.

We generate 10510^{5} paths, of length 30 days from a VAR model, where each path consists of spot and both put and call option prices on a grid of maturities {20,40,60}\{20,40,60\} and relative strikes {0.85,,1.15}\{0.85,\ldots,1.15\}.

Constructing \mathbb{Q}^{*} we construct the risk neutral measure \mathbb{Q}^{*} by solving 34 with the exponential utility, using proportional transaction costs for all instruments set to γ=0.001\gamma=0.001. To parametrize our policy action, we use a two layer feedforward neural network, with 64 units in each layer and ReLU activation functions. We train for 2000 epochs on the training set of 10510^{5} paths.

Assessing Performance – Figure 1 compares out-of-sample the expected value of the option payoffs vs. their prices for the full grid of calls and puts under both the statistical and the risk-free measure in relation to trading cost.

Refer to caption
Figure 1: Average realised drift for call options (top) and put options (bottom) under the \mathbb{P} market simulator (left) and \mathbb{Q}^{*} simulator (right), by strike and maturity.

The expected payoff under the changed measure has been flattened to zero, and now lies within the transaction cost level, so that the tradable drift has been removed.

To further confirm that statistical arbitrage has indeed been eliminated from the market simulator under \mathbb{Q}^{*}, we train a second strategy under the new measure, with identical neural network architecture and the same utility function but incresed transaction cost γ=0.002\gamma=0.002. Figure 2 shows the distributions of terminal gains of respective estimated optimal strategies under \mathbb{P} and \mathbb{Q}. We compare the method using the exponential utility, and the adjusted mean-volatility utility u(x):=(1+λx1+λ2x2)/λu(x):=(1+\lambda x-\sqrt{1+\lambda^{2}x^{2}})/\lambda. In both cases, the distribution of gains is now tightly centred at zero confirming that statistical arbitrage has been removed.

Refer to caption
Figure 2: Gains distribution of estimated optimal policy under \mathbb{P} and \mathbb{Q}^{*} for exponential utility (left) and adjusted mean-volatility (right).

5.2 GAN market simulator

To demonstrate the flexibility of our approach, we now apply it to a more data driven simulator for spot and option prices based on Generative Adversarial Networks (GANs) [8] as in[1].

We illustrate the effect on Deep Hedging of changing measure with the following numerical experiment. We hedge a short position in a digital call option, with market instruments being the spot and at the money call options with maturities 2020 and 4040 days. We first train a network under the zero portfolio to find a maximal statistical arbitrage strategy, then use this to construct the risk neutral density. We then train two Deep Hedging networks to hedge the digital, one under the original, unweighted, market and one under the risk neutral market. All networks are trained to maximize exponential utility. Figure 3 compares the final hedged PNL of the two strategies on the left, and the PNL of the strategies, with the statistical arbitrage component subtracted, on the right (i.e. aa_{\mathbb{Q}}^{*} vs. aa0a_{\mathbb{P}}^{*}-a_{0}^{*}). The distribution of PNL from the risk neutral hedge is clearly less wide tailed, and the righthand plot demonstrates that we have removed the statistical arbitrage element as the distributions now align.

Refer to caption
Figure 3: Comparison of the distribution of hedged PNL from a Deep Hedging strategy trained under \mathbb{P} vs one trained under \mathbb{Q}^{*} (left) and with statistical arbitrage removed (right). Both strategies are evaluated under \mathbb{P}.

Robustness of \mathbb{Q}^{*} by removing the statistical arbitrage component of the strategy, the risk neutral hedge aa_{\mathbb{Q}}^{*} represents a more robust hedge with respect to uncertainty in the market simulator, i.e. when the future distribution of the market at model deployment differs slightly from the distribution of the training data generated by the market simulator. Consider the case where the future market returns follow a distribution ~\tilde{\mathbb{P}}, which is similar to \mathbb{P} in the sense that H(~|)cH(\tilde{\mathbb{P}}|\mathbb{P})\leq c for some small cc where HH is the relative entropy. For illustration, we can construct such a measure by simply perturbing the weights of the simulated paths slightly.

In particular, we consider perturbations which are unfavourable for the original strategy, where the strategy was over-reliant on the perceived drift in the market, which is no longer present under ~\tilde{}\mathbb{P}. Figure 4 shows the new PNL distributions under measures ~\tilde{\mathbb{P}} with H(~|)=cH(\tilde{\mathbb{P}}|\mathbb{P})=c for c=0.05,0.5.c=0.05,0.5. What is striking is that a relatively small shift in the measure can significantly worsen the distribution of hedged PNL of the original Deep Hedging model, but that the distribution of PNL of the model trained on the risk neutral measure remains practically invariant, indicating that the model performances more consistently with respect to uncertainty. Naturally, the method provides robustness against estimation errors for mean returns of the underlying assets.

Refer to caption
Figure 4: Comparison of hedged PNL from both strategies with respect to uncertainty, evaluated under ~\tilde{\mathbb{P}} with H(~|)=cH(\tilde{\mathbb{P}}|\mathbb{P})=c for c=0.05,0.5c=0.05,0.5.

Conclusion

We have presented a numerically efficient method for computing a risk-neutral density for a set of paths over a number of time steps. Our method is applicable to paths of derivatives and option prices in particular, hence we effectively provide a framework for statistically learned stochastic implied volatility via the application of machine learning tools. Our method is generic and does not depend on the market simulator itself, except that it requires that the simulator does not produce classic arbitrage opportunities. It also caters naturally for transaction costs and trading constraints, and is easily extended to multiple assets.

The method is particularly useful to introduce robustness to a utility-based machine learning approach to the hedging of derivatives, where the use of simulated data is essential to train a ‘Deep Hedging’ neural network model. If trained directly on data from the statistical measure, in addition to risk management of the derivative portfolio, the Deep Hedging agent will pursue statistical arbitrage opportunities that appear in the data, thus the hedge action will be polluted by drifts present in the simulated data. By applying our method, we remove any statistical arbitrage opportunities from the simulated data, resulting in a policy from the Deep Hedging agent that seeks to only manage the risk of the derivative portfolio, without exploiting any drifts. This in turn makes the suggested hedge more robust to any uncertainty inherent in the simulated data.

References

  • [1] Lanjun Bai, Hans Buehler, Mangnus Wiese and Ben Wood “Deep Hedging: Learning to Simulate Equity Option Markets” In SSRN, 2019 URL: https://ssrn.com/abstract=3470756
  • [2] Aharon Ben-Tal and Marc Teboulle “An old-new concept of convex risk measures: The optimized certainty equivalent” In Mathematical Finance 17.3 Wiley Online Library, 2007, pp. 449–476
  • [3] H. Buehler, L. Gonon, J. Teichmann and B. Wood “Deep Hedging” In Quantitative Finance 0.0 Routledge, 2019, pp. 1–21 URL: https://ssrn.com/abstract=3120710
  • [4] Hans Buehler and Evgeny Ryskin “Discrete Local Volatility for Large Time Steps (Short Version)”, 2016 URL: https://ssrn.com/abstract=2783409
  • [5] Hans Buehler et al. “A data-driven market simulator for small data environments” In Available at SSRN 3632431, 2020
  • [6] Hans Föllmer and Alexander Schied “Stochastic Finance” De Gruyter, 2008 DOI: doi:10.1515/9783110212075
  • [7] Marco Frittelli “The minimal entropy martingale measure and the valuation problem in incomplete markets” In Mathematical finance 10.1 Wiley Online Library, 2000, pp. 39–52
  • [8] Ian Goodfellow et al. “Generative adversarial nets” In Advances in neural information processing systems 27, 2014
  • [9] V. Henderson and D. Hobson “Utility indifference pricing : an overview”, 2009 URL: https://warwick.ac.uk/fac/sci/statistics/staff/academic-research/henderson/publications/indifference_survey.pdf
  • [10] Jan Kallsen and Paul Krühner “On a Heath-Jarrow-Morten approach for Stock Options” In Finance and Stochastics 19, 2015, pp. 583–615 URL: https://arxiv.org/pdf/1305.5621.pdf
  • [11] Harry Markowitz “Portfolio Selection” In The Journal of Finance 7.1, 1952, pp. 77–91 DOI: 10.2307/2975974
  • [12] Philipp Schönbucher “A market model for stochastic implied volatility” In Phil. Trans. R. Soc. A., 1999, pp. 2071–2092 URL: https://ssrn.com/abstract=182775
  • [13] Skipper Seabold and Josef Perktold “statsmodels: Econometric and statistical modeling with python” In 9th Python in Science Conference, 2010
  • [14] Sasha Stoikov “The micro-price: a high-frequency estimator of future prices” In Quantitative Finance 18.12 Routledge, 2018, pp. 1959–1966
  • [15] Johannes Wissel “Arbitrage-free market models for option prices”, 2007 URL: http://www.nccr-finrisk.uzh.ch/media/pdf/wp/WP428_D1.pdf

Disclaimer

Opinions and estimates constitute our judgement as of the date of this Material, are for informational purposes only and are subject to change without notice. It is not a research report and is not intended as such. Past performance is not indicative of future results. This Material is not the product of J.P. Morgan’s Research Department and therefore, has not been prepared in accordance with legal requirements to promote the independence of research, including but not limited to, the prohibition on the dealing ahead of the dissemination of investment research. This Material is not intended as research, a recommendation, advice, offer or solicitation for the purchase or sale of any financial product or service, or to be used in any way for evaluating the merits of participating in any transaction. Please consult your own advisors regarding legal, tax, accounting or any other aspects including suitability implications for your particular circumstances. J.P. Morgan disclaims any responsibility or liability whatsoever for the quality, accuracy or completeness of the information herein, and for any reliance on, or use of this material in any way.
Important disclosures at: www.jpmorgan.com/disclosures