Arbitrage-free market models for option prices: The multi-strike case

Similar documents
Arbitrage-free market models for interest rate options and future options: the multi-strike case

Tangent Lévy Models. Sergey Nadtochiy (joint work with René Carmona) Oxford-Man Institute of Quantitative Finance University of Oxford.

Local Volatility Dynamic Models

M5MF6. Advanced Methods in Derivatives Pricing

Local vs Non-local Forward Equations for Option Pricing

A note on the existence of unique equivalent martingale measures in a Markovian setting

Basic Arbitrage Theory KTH Tomas Björk

Advanced Topics in Derivative Pricing Models. Topic 4 - Variance products and volatility derivatives

Volatility Smiles and Yield Frowns

Hedging under Arbitrage

European call option with inflation-linked strike

Volatility Smiles and Yield Frowns

Market interest-rate models

Lecture 17. The model is parametrized by the time period, δt, and three fixed constant parameters, v, σ and the riskless rate r.

Dynamic Relative Valuation

Hedging Credit Derivatives in Intensity Based Models

Application of Stochastic Calculus to Price a Quanto Spread

1 Mathematics in a Pill 1.1 PROBABILITY SPACE AND RANDOM VARIABLES. A probability triple P consists of the following components:

Path Dependent British Options

Robust Pricing and Hedging of Options on Variance

1.1 Basic Financial Derivatives: Forward Contracts and Options

The British Russian Option

Risk Neutral Measures

1 Implied Volatility from Local Volatility

Martingale Approach to Pricing and Hedging

Youngrok Lee and Jaesung Lee

Interest rate models in continuous time

AMH4 - ADVANCED OPTION PRICING. Contents

Chapter 15: Jump Processes and Incomplete Markets. 1 Jumps as One Explanation of Incomplete Markets

Continuous Time Finance. Tomas Björk

Basic Concepts in Mathematical Finance

( ) since this is the benefit of buying the asset at the strike price rather

Pricing Barrier Options under Local Volatility

Rohini Kumar. Statistics and Applied Probability, UCSB (Joint work with J. Feng and J.-P. Fouque)

Constructing Markov models for barrier options

Introduction to Probability Theory and Stochastic Processes for Finance Lecture Notes

THE MARTINGALE METHOD DEMYSTIFIED

Hedging of Credit Derivatives in Models with Totally Unexpected Default

Pricing Dynamic Solvency Insurance and Investment Fund Protection

4: SINGLE-PERIOD MARKET MODELS

Optimal Hedging of Variance Derivatives. John Crosby. Centre for Economic and Financial Studies, Department of Economics, Glasgow University

Pricing of a European Call Option Under a Local Volatility Interbank Offered Rate Model

Stochastic Processes and Stochastic Calculus - 9 Complete and Incomplete Market Models

The Black-Scholes Model

STOCHASTIC CALCULUS AND BLACK-SCHOLES MODEL

Stochastic Volatility (Working Draft I)

1 The continuous time limit

Lecture Note 8 of Bus 41202, Spring 2017: Stochastic Diffusion Equation & Option Pricing

How do Variance Swaps Shape the Smile?

Exploring Volatility Derivatives: New Advances in Modelling. Bruno Dupire Bloomberg L.P. NY

OPTION PRICE WHEN THE STOCK IS A SEMIMARTINGALE

Optimal stopping problems for a Brownian motion with a disorder on a finite interval

SPDE and portfolio choice (joint work with M. Musiela) Princeton University. Thaleia Zariphopoulou The University of Texas at Austin

THE USE OF NUMERAIRES IN MULTI-DIMENSIONAL BLACK- SCHOLES PARTIAL DIFFERENTIAL EQUATIONS. Hyong-chol O *, Yong-hwa Ro **, Ning Wan*** 1.

KØBENHAVNS UNIVERSITET (Blok 2, 2011/2012) Naturvidenskabelig kandidateksamen Continuous time finance (FinKont) TIME ALLOWED : 3 hours

IEOR E4703: Monte-Carlo Simulation

Large Deviations and Stochastic Volatility with Jumps: Asymptotic Implied Volatility for Affine Models

MASSACHUSETTS INSTITUTE OF TECHNOLOGY 6.265/15.070J Fall 2013 Lecture 19 11/20/2013. Applications of Ito calculus to finance

The Capital Asset Pricing Model as a corollary of the Black Scholes model

The Birth of Financial Bubbles

Lecture Notes for Chapter 6. 1 Prototype model: a one-step binomial tree

Hedging under arbitrage

The Use of Importance Sampling to Speed Up Stochastic Volatility Simulations

No-Arbitrage Conditions for the Dynamics of Smiles

LIBOR models, multi-curve extensions, and the pricing of callable structured derivatives

LECTURE 2: MULTIPERIOD MODELS AND TREES

Equivalence between Semimartingales and Itô Processes

Change of Measure (Cameron-Martin-Girsanov Theorem)

Option Pricing under Delay Geometric Brownian Motion with Regime Switching

Hedging of Contingent Claims under Incomplete Information

AN ANALYTICALLY TRACTABLE UNCERTAIN VOLATILITY MODEL

NEWCASTLE UNIVERSITY SCHOOL OF MATHEMATICS, STATISTICS & PHYSICS SEMESTER 1 SPECIMEN 2 MAS3904. Stochastic Financial Modelling. Time allowed: 2 hours

American Option Pricing Formula for Uncertain Financial Market

Model-independent bounds for Asian options

On Asymptotic Power Utility-Based Pricing and Hedging

We discussed last time how the Girsanov theorem allows us to reweight probability measures to change the drift in an SDE.

Forwards and Futures. Chapter Basics of forwards and futures Forwards

Stochastic Differential equations as applied to pricing of options

An overview of some financial models using BSDE with enlarged filtrations

The Black-Scholes Model

A model for a large investor trading at market indifference prices

Probability in Options Pricing

Rough volatility models: When population processes become a new tool for trading and risk management

Optimal trading strategies under arbitrage

Practical example of an Economic Scenario Generator

Recovering portfolio default intensities implied by CDO quotes. Rama CONT & Andreea MINCA. March 1, Premia 14

Stochastic Differential Equations in Finance and Monte Carlo Simulations

Valuation of derivative assets Lecture 8

Multi-Asset Options. A Numerical Study VILHELM NIKLASSON FRIDA TIVEDAL. Master s thesis in Engineering Mathematics and Computational Science

Characterization of the Optimum

From Discrete Time to Continuous Time Modeling

An Introduction to Point Processes. from a. Martingale Point of View

PDE Approach to Credit Derivatives

The Black-Scholes Model

Numerical schemes for SDEs

Stochastic Calculus, Application of Real Analysis in Finance

Option Pricing and Calibration with Time-changed Lévy processes

MSc Financial Engineering CHRISTMAS ASSIGNMENT: MERTON S JUMP-DIFFUSION MODEL. To be handed in by monday January 28, 2013

Structural Models of Credit Risk and Some Applications

Lecture 4. Finite difference and finite element methods

Transcription:

Working Paper Series National Centre of Competence in Research Financial Valuation and Risk Management Working Paper No. 38 Arbitrage-free market models for option prices: The multi-strike case Martin Schweizer Johannes Wissel First version: May 7 Current version: February 8 This research has been carried out within the NCCR FINRIS project on Mathematical Methods in Financial Risk Management

Arbitrage-free market models for option prices: The multi-strike case Martin Schweizer Johannes Wissel Department of Mathematics, ETH Zurich Rämistrasse CH-89 Zurich Switzerland {martin.schweizer,johannes.wissel}@math.ethz.ch 5th February 8 Abstract This paper studies modelling and existence issues for market models of option prices in a continuoustime framework with one stock, one bond and a family of European call options for one fixed maturity and all strikes. After arguing that classical implied volatilities are ill-suited for constructing such models, we introduce the new concepts of local implied volatilities and price level. We show that these new quantities provide a natural and simple parametrization of all option price models satisfying the natural static arbitrage bounds across strikes. We next characterize absence of dynamic arbitrage for such models in terms of drift restrictions on the model coefficients. For the resulting infinite system of SDEs for the price level and all local implied volatilities, we then study the question of solvability and provide sufficient conditions for existence and uniqueness of a solution. We give explicit examples of volatility coefficients satisfying the required assumptions, and hence of arbitrage-free multi-strike market models of option prices. ey words option prices, market model, implied volatility, static arbitrage, dynamic arbitrage, drift restrictions, existence result MSC Classification Numbers 6H, 9B8 JEL Classification Numbers C6, G3 Introduction Consider a financial market where the following assets are all traded liquidly: a bank account bond paying no interest, a stock S, and a collection of European call options C, T on S with various strikes and maturities T T. Our ultimate goal is to establish a framework for pricing and hedging possibly exotic derivatives in an arbitrage-free way, using all the liquid tradables as potential hedging instruments. The present paper takes a step in that direction. In order to achieve our goal, we want to construct a class of models for bond, stock and options having at least the following features: Of course, the model should be arbitrage-free. Any initial option price data from the market can be reproduced by the model; this could be called perfect calibration or smile-consistency. Empirically observed stylized facts from market time series, i.e., characteristic features of the joint dynamics of stock and options, can be incorporated in the model. This requires that explicit expressions for option price processes and their dynamics should be available. The overwhelming majority of the literature uses the martingale approach, where one specifies the dynamics of the underlying S under some pricing i.e., martingale measure Q and defines option prices

by C t, T := E Q [S T + F t ]. This obviously satisfies, and a perfect fit of the entire initial option surface as in is for instance possible with the so-called smile-consistent models. However, is usually not feasible, or if it is to some extent, this often comes at the cost that it entails a loss in. We discuss this in more detail in the next section. An alternative approach is the use of market models where one specifies the joint dynamics of all tradable assets here, stock and options. This gives and by construction, and the remaining issue is to ensure absence of arbitrage to have as well. In interest rate modelling, this leads to the well-known HJM drift conditions; but the case of options is more complicated. In fact, absence of dynamic arbitrage again corresponds to drift conditions for the joint dynamics of S and the C, T. But in addition, absence of static arbitrage enforces a number of relations between the various C, T and S, and this means that the state space of these processes is constrained as well. To obtain a tractable model, one must therefore reparametrize the tradables in such a way that the parametrizing processes have a simple state space and yet capture all the static arbitrage constraints. We explain this in more detail in the next section, but the point here is that this modelling task is quite difficult. The literature with actual results on arbitrage-free market models for option prices is quite small and most compactly summarized in terms of the families and T. Again, a more thorough discussion is postponed to the next section. For the case = {}, T = {T } of one single call option available for trade, there are both an existence result and some explicit examples for models. For models with = {}, T =, one fixed strike, all maturities, the drift restrictions are well-known, but existence of models has been proved only very recently. The other extreme =,, T = {T } all strikes, one fixed maturity is the focus of this paper; it is more difficult and has to the best of our knowledge no precursors in terms of parametrization or results. Finally, the case =,, T =, of the full surface of strikes and maturities is still open despite some recent work by Carmona and Nadtochiy [3]; see Section 6 for more details. The paper is structured as follows. In Section, we give an overview of the literature that is most closely related to the problem studied here, and we explain in more detail the nature of our contribution. Section 3 reviews market models for stochastic implied volatilities. We characterize absence of arbitrage in terms of drift restrictions and provide a general existence result for the case of a single option C, T. This is done in order to illustrate where one meets difficulties with classical implied volatilities when passing to models with multiple strikes. Our main contribution is contained in Section 4. Instead of modelling stock price and classical implied volatilities, we introduce for a set of maturity-t call prices with strikes price level and local implied volatilities which parametrize in a natural and simple way all possible arbitrage-free option prices. We provide explicit formulas for these new quantities as functions of stock price and classical implied volatilities, and vice versa. In analogy to classical implied volatilities, we then characterize their arbitrage-free dynamics in terms of drift restrictions. In Section 5 we provide explicit and fairly general examples of arbitrage-free dynamic models for the price level and the local implied volatilities. To prove existence and uniqueness of a solution to the corresponding infinite system of SDEs, we adapt results from [44] to our setting. Section 6 concludes and points out a number of open questions. Background, motivation, and literature This section discusses in more detail what we want to and what we can achieve with our approach. Moreover, it also gives an overview of related literature, and for this, a slightly broader perspective is useful. So let us look at models that exploit or produce information about an underlying stock as well as options written on S.. Martingale models In the martingale approach, one writes down a dynamic model usually an SDE for a stock price martingale S under a probability measure Q and defines C t, T := E Q [ ST + Ft ], t T

for, and T T,. These models by construction satisfy the requirement of being arbitrage-free. Calibration as in to given market option prices is more or less feasible for instance in stochastic volatility models e.g. Hull and White [3], Heston [3], Davis [] or in models with jumps e.g. Merton [37], Barndorff-Nielsen and Shephard [3], Carr et al. [4], and several of the models also match some of the stylized features for S alone. But of course, calibration is limited by the fact that one has only a finite number of parameters to be fitted. A perfect fit of the entire option surface C, T for =,, T =, is achieved by the so-called smile-consistent models, most prominent among which are the local volatility model of Dupire [4] and discrete-time implied tree models like Derman and ani []. A good overview on smile-consistent pricing is given by Skiadopoulos [43], and some recent papers like Carr et al. [6] or Rousseau [39] also produce in addition fairly realistic dynamic behaviour for S alone. nowing at time all call option prices C, T for, is equivalent to knowing the marginal distribution of S T under Q; this observation goes back to Breeden and Litzenberger []. Hence perfect fitting of all C, T with,, T T can be achieved by constructing a martingale with the corresponding marginals for S T, T T, under Q, and this can be done in many ways and situations; see for instance Madan and Yor [36], Carr et al. [5], Bibby et al. [6], Atlan [], Hamza and lebaner [8]. There are also many papers on calibration or empirical analyses of various models, even if we do not quote any of this work here. But from our perspective, all these models suffer from the same fundamental drawback: In general, there are no explicit expressions for the processes C, T, and so their joint dynamics with S are not really available. Another question of interest in the context of models for stocks and options is the link between the implied and the instantaneous volatility. This has been studied both for local and for stochastic volatility models, and the typical results are asymptotic relationships close to maturity and for at-the-money options; see for instance Berestycki et al. [4], [5] or Durrleman [6]. But again, these papers neither provide nor study the joint dynamics of S and C, T.. Market models As already explained in the introduction, a natural way to construct a model satisfying the requirements of perfect calibration and of joint dynamics is to use a market model, where one specifies the dynamics of all liquid tradables simultaneously. This goes back to ideas from interest rate modelling, and absence of dynamic arbitrage there leads to the well-known drift conditions of Heath et al. [9]. The same type of conditions also appears in option price models. But in addition, static arbitrage bounds lead to restrictions on the state space of the quantities used to describe the model, and so the choice of a suitable parametrization becomes a crucial issue. As a matter of fact, the same problem arises in the interest rate context if one insists on modelling zero coupon bond prices; but it is easily resolved there by passing to forward rates instead. In the literature, some work has been done in special cases. If the option collection consists of a single call C = C, T, one has the static arbitrage bounds S t + C t S t as well as the terminal condition C T = S T +. Specifying directly for the pair S, C dynamics which obey these state space constraints is quite delicate. It is much easier to reparametrize the option price C by its implied volatility ˆσ via C t = c S t,, T tˆσ t, where c is the well-known Black-Scholes [7] function given in 3. below. Then the pair S, ˆσ may take any value in,, the static arbitrage bounds and terminal condition are satisfied, and one can proceed to specify and study models for the joint dynamics of S, ˆσ. Such market models of implied volatilities for a single option have first been proposed in Lyons [35] and Schönbucher [4], and arbitrage-free examples have been constructed in Babbar []. Even in this apparently simple situation, the construction is not entirely straightforward: in an Itô process framework over a Brownian filtration, the drifts are essentially determined by the volatilities of S and ˆσ, and if one takes these nonlinear drift restrictions into account, the question whether the resulting two-dimensional SDE system for S and ˆσ admits a solution becomes nontrivial. The situation becomes much more complicated if our option collection contains more than one single call. In the literature, one can find several variants of necessary conditions on the implied volatility dynamics for the resulting model to be arbitrage-free; see for instance Schönbucher [4], Brace et al. [9], 3

[] and Ledoit et al. [33]. However, none of these works provide any explicit example of a multi-option market model; in other words, no sufficient conditions are given, and the existence of such models with specified dynamics remains an open issue. The key difficulty is that the well-known static no-arbitrage conditions for calls with different strikes and maturities see e.g. Carr and Madan [7] or Davis and Hobson [] entail rather complicated relations between the implied volatilities of these options; this is illustrated in some more detail in Section 3. We believe that there is a fundamental reason for this problem: despite their importance as a market standard to quote option prices, classical implied volatilities are unsuited for modelling call prices in a multi-option model. Put bluntly, they give the wrong parametrization. Of course, the idea of replacing implied volatilities by another parametrization of call prices in option market models is not entirely new. For the case = {}, T =, of a family with one fixed strike and all maturities T >, Schönbucher [4] has introduced the forward implied volatilities ˆσ fw T := T tˆσ T,. T and we have recently used in [4] new techniques from [44] for infinite-dimensional SDE systems to prove existence results for this class of models. The main contribution in [4] is to show how one can handle the complicated SDE systems that arise via the drift restrictions coming from absence of dynamic arbitrage. The choice of the parametrization. is taken from Schönbucher [4] and has its roots in the obvious analogy to the well-known forward rates for interest rate modelling. As a matter of fact, the results in [4] are more generally given for a maturity term structure of options with one fixed convex or concave payoff function h and all maturities T >. The special case h = log corresponds to a market model for variance swaps, where the drift conditions take a particularly simple form; the resulting model has been explicitly analyzed in Bühler []. Jacod and Protter [3] also study models for options with one fixed payoff function and all maturities and parametrize via the maturity derivatives T C tt. However, they do not specify CT by joint dynamics with S, and so their work falls into the realm of the martingale approach discussed in Section.. In this paper, we consider the other extreme of the spectrum. We want to construct arbitrage-free market models for call option prices in the case =,, T = {T } of a family with one fixed maturity T and all strikes >. This is substantially more difficult than the case of all maturities with one fixed strike because it requires new ideas already at the modelling level. Our main achievement is to introduce a new parametrization of option prices for the multi-strike case in such a way that arbitrage-free dynamic modelling becomes tractable. We define these new quantities, called local implied volatilities, in Section 4. They have no comparable precursor or analogue in interest rate theory because the traded assets in interest rate market models, the zero coupon bonds, simply do not have any strike structure. The key feature of these new parameters is that they have a simple state space and yet capture precisely all the static arbitrage restrictions. Once they have been constructed, dynamic arbitrage conditions and existence results for the corresponding dynamic option models still need to be dealt with, but this can be achieved by using our techniques developed in [44] and [4]. 3 Market models for implied volatilities In this section, we review market models for implied volatilities and explain why their usefulness in arbitrage-free modelling is mostly limited to the case of a single traded option C, T. The general setup along with the concept of implied volatilities is introduced in Section 3.. In Section 3. we characterize absence of dynamic arbitrage in such models in terms of drift restrictions. This recovers the results in Section 3.3 of Schönbucher [4]. Our Section 3.3 provides sufficient conditions for existence and uniqueness of the corresponding dynamics of S and ˆσ, T in the case of only one pair, T and discusses the difficulties that arise if one tries to generalize this approach to a model of implied volatilities for more than one strike. 3. Implied volatilities of call options Throughout this paper, we work with the following setup. Let Ω, F, P be a probability space and T > a fixed maturity. Let S t t T be a positive process modelling a stock price, and B t t T a positive 4

process with B T = P-a.s., modelling the price of a non-defaultable zero coupon bond with maturity T. Moreover, for > let C t t T be a nonnegative process modelling the price of a European call option on S paying S T B T + = S T + at time T. Finally, let cs,, Υ be the Black-Scholes function logs/ + cs,, Υ = SN Υ logs/ N Υ Υ >, Υ Υ 3. cs,, = S +, where N denotes the standard normal distribution function. Clearly, c is strictly increasing in Υ with lim Υ cs,, Υ = S for all S,. If the model consisting of B, S and C for a fixed > does not admit an elementary arbitrage opportunity, then it is well known that we have for all t T S t B t + C t S t see e.g. [4, Proposition.]. This allows us to give Definition 3.. The implied volatility of the price C t is the unique parameter ˆσ t satisfying c S t, B t, T tˆσ t = C t. 3. Since the Black-Scholes function c has the homogeneity property cs,, Υ = S c S, Υ for a suitable function c, the implied volatility is invariant under a change of numeraire; in other words, the defining condition 3. can be rewritten, for every positive numeraire process M, as c S t /M t, B t /M t, T tˆσ t = C t /M t. Therefore, we always use from now on the bond B as numeraire, and in the sequel, all price processes B, S, C denote B-discounted price processes, so that B. 3. Drift restrictions for implied volatilities Let W be an m-dimensional Brownian motion on Ω, F, P, F = F t t T the P-augmented filtration generated by W, and F = F T. We denote for d N and p by L p loc Rd the space of all R d -valued, progressively measurable, locally p-integrable in t, P-a.s. processes on [, T]. We model a stock price process S t t T and, for some set of strikes,, a family of price processes C t t T of call options paying S T + at time T by with dynamics C t = c S t,, T tx t 3.3 ds t = µ t S t dt + σ t S t dw t t T, S = s, 3.4 dx t = u t X t dt + v t X t dw t t T, X = x. 3.5 Here c is the Black-Scholes function from 3., µ, u are in L loc R, and σ, v are in L loc Rm. Each X = ˆσ is a positive process modelling the square of the implied volatility of C. It is now natural to ask under which conditions there exists a common equivalent local martingale measure for the discounted price processes S, C for all. The existence of such a measure is essentially equivalent to the drift restrictions µ t = σ t b t, 3.6 u t = T t X t σ t + log S t v t + 6 T tx t + 4 v t σ t + b t vt 3.7 for all and a market price of risk process b L loc Rm. More precisely, we have the following result. 5

Theorem 3.. a If there exists a common equivalent local martingale measure Q for S and C for all, then there exists a market price of risk process b L loc Rm such that 3.6 and 3.7 hold for a.e. t [, T], P-a.s. b Conversely, suppose that the coefficients µ, σ, u and v satisfy, as functions of S and X, the relations 3.6, 3.7 for a.e. t [, T], P-a.s., for some bounded uniformly in t, ω process b L loc Rm. Also suppose that there exists a family of positive continuous adapted processes S, X on [, T] satisfying the system 3.4, 3.5. Then there exists a common equivalent local martingale measure Q on F T for S t t T, C t t T for all. One such measure is given by dq dp := E b dw T, where E denotes the stochastic exponential. Moreover, if σ is bounded, then S and C for all are martingales under Q. This result is essentially proved in Section 3.3 of Schönbucher [4]. Note that the free input parameters in the model are the stock volatility σ and the family of processes v for all, i.e., the volatilities of the implied volatilities X; they determine together with the market price of risk b the drifts µ and u via 3.6 and 3.7. The v are often called volvols. 3.3 Existence problems in arbitrage-free implied volatility models We now turn to the question of existence and uniqueness of solutions for arbitrage-free implied volatility models. This is an important issue; without an existence result, it is not possible to specify a concrete model, and uniqueness is the basis for any convergence result of an eventual numerical implementation. For the case of one single call option, a positive result can be found in Babbar []. In order to point out the major difficulties arising in the general case of several calls, we review here the case of a single option by discussing a slightly more general version of the basic result of Babbar []. Consider the model 3.4, 3.5, where µ and u are given by 3.6, 3.7, and take the case where the coefficients v are nonzero constants and σ is in L loc Rm. In general, two problems will arise:. Because of the nonlinear dependence on X of the drifts u in 3.7, a solution of 3.5 will in general only exist up to an explosion time which may be strictly less than T with positive probability.. Due to the factor T t, the drifts u will typically not be in L loc R. The solution of 3.5 will explode at maturity, i.e., for t ր T, and C will no longer be a local martingale on [, T]. So for a general specification of the coefficients σ, v, we must expect that the system 3.4, 3.5 does not have a non-exploding solution on [, T], and that there does not exist an arbitrage-free model with these coefficients. See also Schönbucher [4], Section 3.5, for a discussion of the second problem. For a positive result, we need to make a choice for σ, v which avoids the above difficulties. As in Babbar [], we restrict to the case = {} of a single call option, and to simplify notation we drop the dependence on in the quantities X, u, v. We choose the processes v and σ as functions of the state variables t, S, X, writing with a slight abuse of notation v t = vt, S, X, σ t = σt, S, X, and define for a fixed market price of risk process b the processes µ t = µt, S, X, u t = ut, S, X by the drift restrictions 3.6, 3.7. In order to obtain a unique strong solution to our SDEs, we have to impose some sort of Lipschitz condition on the coefficients. To that end, let U R d and Θ be a possibly 6

empty set. We say that a function f : Θ U R k is locally Lipschitz on U if f, x is bounded for fixed x U and if there exists a continuous function C, on U such that fθ, x fθ, x Cx, x x x x, x U, θ Θ. One easily checks that if f, g are locally Lipschitz on U and h : fu R k is locally Lipschitz on fu, then f + g, fg and h f are locally Lipschitz on U. In the following const denotes a generic positive constant whose value can change from one line to the next. We now have the following result. Proposition 3.3. Let b be a progressively measurable process which is uniformly bounded and σt, s, x := log/svt, s, x + x ft, s, x + T tgt, s, x, 3.8 where f, g, v : [, T], R m are locally Lipschitz on, and satisfy ft, s, x =, gt, s, x const, vt, s, x const + x + log/s. 3.9 Then the system 3.4, 3.5 with 3.6, 3.7 has a unique positive non-exploding solution S, X on [, T]. For the proof, we need an existence and uniqueness result for strong solutions of SDEs with locally Lipschitz coefficients. Proposition 3.4. Let x R d and β : [, R d R d, γ : [, R d R d m be functions which are locally Lipschitz on R d. Suppose that for f {β, γ} ft, x const + x t, x R d. 3. Then there exists a unique strong solution on [, to the SDE dx t = β t, X t dt + γ t, Xt dwt, X = x. Proof. This follows from Chapter 5, Theorems 3. and 3. in Durrett [5] for the time-homogeneous case. The time-inhomogeneous case is proved completely analogously. We now come to the Proof of Proposition 3.3. In order to deal with the positivity requirement for S and X, we work with a suitable transformation of the state variables. Let a be sufficiently large. Then there exists a convex smooth strictly increasing function ψ : R, such that { ψz = z for z a, z for z a. 3. Let ϕ :, R be the inverse of ψ. We apply Proposition 3.4 with d = to the SDE system for the processes Y = log S, Z = ϕx, that is, the system dy = µ t, e Y, ψz σ t, e Y, ψz dt + σ t, e Y, ψz dw, Y = log s, 3. dz = ūt, Y, Zdt + vt, Y, ZdW, Z = ϕx, 3.3 where ūt, y, z = u t, e y, ψz ψzϕ ψz + v t, e y, ψz ψz ϕ ψz, vt, y, z = v t, e y, ψz ψzϕ ψz. If we have a solution Y, Z to 3., 3.3, then by Itô s lemma S, X := e Y, ψz is a solution to 3.4, 3.5, and vice versa. 7

It now only remains to check the conditions of Proposition 3.4 for the coefficients in the system 3., 3.3. The local Lipschitz condition is clearly satisfied since f, g, v, ψ, ϕ, ϕ are locally Lipschitz. To check 3., first note that we have } ψzϕ ψz const + z z R, ψz ϕ ψz 3.4 const + z z R. This follows for z / [ a, a] by direct computation from 3. and for z [ a, a] by continuity of the functions on the left-hand sides of 3.4. Next, note that from 3.7 and 3.8 we obtain ut, s, x = ft, s, x gt, s, x T t gt, s, x + 6 T tx + 4 vt, s, x σt, s, x + b vt, s, x, and now 3.8 and 3.9 imply that ut, s, x and vt, s, x are bounded. Together with 3.4, this yields 3. for the coefficients of 3.3. Finally, 3.8 and 3.9 imply σ t, e y, ψz const + ψz const + a + z. Together with 3.6, this yields 3. for the coefficients of 3., and the proof is complete. The main ideas for the volatility specifications in Proposition 3.3 are the following. First, we choose v in a form which ensures that u from 3.7 is bounded in X; this is why we need the asymptotic behaviour vt, s, x x for x in 3.9. Once v is given, the choice of σ in 3.8 is then necessary to remove the singularity and ensure boundedness of u from 3.7 near maturity, i.e., for T t ց. We can now also illustrate one of the main difficulties in extending this result to >. In a market model 3.4, 3.5 for a stock S and several call options and squared implied volatilities X satisfying the no-arbitrage conditions 3.6, 3.7, it is not clear how to choose the coefficients σ, v to ensure non-explosion of the drifts u and thus absence of arbitrage for T t ց. Already for =, the above method would force us to choose σ for given v and v in such a way that we keep both u and u under control, and it is not clear if or how this could be achieved. It does not help either if one tries to first specify σ and then find suitable v and v ; getting simultaneous control over both u and u looks equally hard. The true reason behind these difficulties is the fact that implied volatilities cannot take arbitrary values across a spectrum of strikes: absence of arbitrage between different options enforces awkward constraints and relations between the corresponding implied volatilities. For an easy example, take strikes < < and suppose that the call price C is given by some implied volatility ˆσ >. If the implied volatility ˆσ for the strike now exceeds a certain finite bound depending on S, ˆσ,,, the price curve c S,, T tˆσ is no longer decreasing on [, ], which leads to an immediate arbitrage opportunity. This is the same effect we have already seen in Section where modelling S, C was very delicate whereas modelling S, ˆσ was straightforward. In the present two-strike situation, we now find that modelling S, ˆσ, ˆσ is unpleasant, and the reason is again that the state space of this process is very complicated due to static arbitrage constraints. We conclude again that we need another, better suited parametrization and this is why we think that classical implied volatilities are not the right choice for constructing market models of option prices. To overcome these problems, we introduce in the next section a transformation of the variables S, X,, to a new parametrization which has a very simple state space, and for which a non-singular specification of the arbitrage-free dynamics consequently becomes straightforward. 4 Local implied volatilities and price level of option price curves It has been a market standard for a long time to use implied volatilities for quoting option prices, and they have been extensively analyzed in many respects; a good overview can be found in Lee [34]. In particular, the statistical behaviour of ˆσ is well known from many empirical studies on the dynamics of the surface ˆσ, T see for example Cont and da Fonseca [8] for a list of references in this area. But for the purpose of a theoretical analysis, implied volatilities in an arbitrage-free setting suffer from the 8

serious drawback that they cannot take arbitrary positive values across different strikes, and we have just seen in Section 3.3 that this makes the construction of option market models via implied volatilities a tremendous if not impossible challenge. In this section, we therefore propose to parametrize call prices by a new set of quantities which do not suffer from the above problems. In Section 4., we introduce local implied volatilities and the price level which model an arbitrage-free set of call prices in a natural way. We provide an interpretation for these quantities and show how they are related to classical implied volatilities and the stock price. Examples are given in Section 4.. In Section 4.3, we then derive the arbitrage-free dynamics of local implied volatilities and the price level. The resulting infinite system of SDEs is studied in more detail in Section 5; it is still complicated but tractable. 4. The new parametrization: Definitions and basic properties We resume the setup of Section 3.. On some probability space Ω, F, P, we have a discounted bond price process B and positive processes S t t T and C t t T, >, modelling the discounted prices of a stock S and European call options on S with one fixed maturity T > and all strikes >. In the notation of Section 3, we have here =,. By setting C t := S t, the model is specified through the processes C,, on the interval [, T]. Definition 4.. A function Γ : [, [, is called a price curve. A price curve is called statically arbitrage-free if it is convex and satisfies Γ + for all. This definition is motivated by the following Proposition 4.. If the above market C, does not admit an elementary arbitrage opportunity, then for each t [, T the price curve Γ := C t is statically arbitrage-free. Proof. See Davis and Hobson [], Theorem 3.. Since C t = S t, one easily checks that any statically arbitrage-free price curve satisfies the elementary static arbitrage bounds S t + C t S t. Besides absence of arbitrage, a further economically reasonable requirement for a European call option model is that lim C t = for all t [, T, and under some non-degeneracy conditions on S one can also show that the inequalities in Definition 4. must be strict. This motivates that we restrict ourselves to the following class of models. Definition 4.3. An option model C, is called admissible if the price curve Γ := C t has an absolutely continuous derivative with Γ > for a.e. > and < Γ < for all >, and lim Γ = for each t [, T, and if we have C T = S T + for all, P-a.s. We now introduce a new set of fundamental quantities which allow a straightforward parametrization of admissible option models. Let N denote the quantile function and n = N the density function of the standard normal distribution. To motivate our subsequent definition, recall the Black- Scholes function cs,, Υ in 3. and note that its first and second partial derivatives with respect to the strike are given by c S,, T tσ = Nd and c S,, T tσ = nd T t σ with d = logs/ T tσ / T t σ This motivates the following. Hence we have the identity n N c S,, T tσ σ =. 4. T t c S,, T tσ Definition 4.4. Let C t t T be admissible. The local implied volatility of the price curve at time t [, T is the measurable function X t given by X t := T tc t n N C t for a.e. >. 4. Next, for a fixed constant,, we define the price level of the price curve at time t [, T as Y t := T tn C t. 4.3 9

By 4., the quantity X t can be interpreted as an implied volatility in the sense of a functional of the call price curve that yields back the volatility parameter for option prices given by Black-Scholes prices. The terminology local implied volatility is justified by the following result. Proposition 4.5. Let X and Y be the local implied volatilities and price level of an admissible model C,. Suppose that for a small interval I = [a, b], and a fixed t < T we have X t = X t a for all I. Then there exists a unique pair x t, z t, such that holds for all, I. It is given by c z t,, T tx t c zt,, T tx t = Ct C t 4.4 x t = X t a, z t = exp X t ay t X t a a dh X + log a + thh T tx ta. The proof is given at the end of this section. By Proposition 4.5, the local implied volatility X t a at the strike a is that unique implied volatility parameter x t for which the Black-Scholes formula prices all call option differences C t C t with strikes in the small interval I = [a, b] consistently, i.e., with the same implied volatility parameter x t and the same implied stock price z t. Note that this interpretation only holds locally, in the sense that the assumption X t = X t a for [a, b] can only hold approximately if the interval [a, b] is very small. The implied stock price z t in general depends on I and differs from the underlying stock price C t = S t, unless the price curve C t is generated by a Black-Scholes model; see Example 4. below. As we shall see later, the price level Y t which does not depend on or I by construction serves as a convenient substitute for the implied stock price z t in our option model framework. For an admissible option model, we clearly have by definition X t >, Y t R. The main motivation for Definition 4.4 is that the set of positive local implied volatility curves and real-valued price levels is up to some integrability conditions in a one-to-one relation to admissible option price models, as is shown in the following Theorem 4.6. Let X, Y be the local implied volatilities and price level of an admissible model C. Then C t = N Y t k dh X thh dk, [,, 4.5 T t C t = N Y t dh X thh,,, 4.6 T t C t = n Y t dh X thh T t X t, a.e.,. 4.7 T t Conversely, for continuous adapted processes X >, Y on [, T] for which the right-hand side of 4.5 is finite P-a.s., define C t via 4.5. Then C, is an admissible model having local implied volatilities X and price level Y. Proof. We start with the second assertion. 4.5 implies 4.6 and 4.7, so the price curves Γ := C t are strongly arbitrage-free. Finiteness of the integral in 4.5 ensures lim Γ =, and using the behaviour of the integrand N C T = Yt k dh X t hh T t I { Y T > k in 4.5 for t ր T gives } dh kdk I { Y X T hh T > k } dh kdk. X T hh Because Xh >, the integrand k gk := I { Y T > } k dh is decreasing, has values and, and X T hh its integral from to is C T = S T. So either g on [, ], in which case C T = S T, or g

drops to before, in which case S T = gkdk <. This shows that C T = S T +. Finally, solving 4.6 at = for Y t gives 4.3, and solving 4.7 for X t via plugging in 4.6 yields 4.. For the first assertion, note that C t defined by 4.6 solves the first order ODE 4. for C t with initial condition 4.3. Since x n N x is locally Lipschitz on,, the solution is unique and thus must be given by 4.6. Finally 4.5 follows by integrating 4.6 and using lim C t =. Remark. Definition 4.4 and Theorem 4.6 can be extended to the following setting. Let C, be an option model such that the price curve Γ := C t is statically arbitrage-free, has absolutely continuous derivative, and satisfies lim Γ = for each t [, T. We also suppose that there exists a constant, with I t := { > C t, } for all t [, T. Note that I t is an interval since C t is an increasing function. Then define for t [, T X t := T tc N C t if C t >, t n if C t =, / I t, if C t =, I t and Y t by 4.3. Then in analogy to Theorem 4.6, the option prices can be written as in 4.5 4.7 with the obvious interpretation of the integrals k dh X thh for [, ]-valued functions X th. Conversely, for a process X satisfying X t, ] for I t and X t = for / I t with intervals I t containing, and a real-valued process Y, define C t again via 4.5. Then we obtain a model with statically arbitrage-free price curves having local implied volatilities X and price level Y. This extended definition of the local implied volatilities can for example be used to construct models in which the stock price S only takes values in some interval I [, ; in this case, absence of arbitrage requires that C t = for / I. Here is a sufficient criterion for the finiteness of the right-hand side of 4.5. Note that this is a condition on X. log Proposition 4.7. If there exists > such that X t T t for a.e., then the outer integral in 4.5 is finite. Proof. We may assume >. Then for sufficiently large k depending on and Y t dh X k thh 8T t = log k log 8T t, N k Y k dh t X t hh dh log h h T t log k 8 + log 8 + T Y t t 4 logk, Yt k dh X t hh T t N 4 log k n 4 logk = π e log k = So the integral is finite P-a.s., for each t. π k. Remarks. If there exists an equivalent martingale measure Q P for the model C,, then we have C t = Q[S T > F t ], and hence 4.6 implies the relation Yt Q[S T > F t ] = N T t between the price level and the risk-neutral probability of the call C being in the money at maturity. In particular, this relation shows that we have to exclude the choice = in the definition 4.3 of the price level unless we are dealing with a defaultable stock model. Under the assumption in the default probability P[S T = F t ] is zero if and only if we have C t =. This condition can be ensured in a local implied volatility model for instance by demanding that for some ǫ > and x > we have X t x log for a.e. ǫ; this implies dh X for and hence thh C t = by 4.6.

In addition to providing an interpretation for the local implied volatilities and price level, we can also express these new quantities in terms of the classical implied volatilities ˆσ t of the admissible option prices C t and the stock price S t. These formulas are a bit lengthy but explicit. Proposition 4.8. Define Υ t := T tˆσ t and d t, := logst/ Υt. Then the local Υt implied volatilities and price level are given by X t = ˆσ t n [ + + 4 N N d t, n d t, Υ t ˆσ t d d ˆσ t log S t + Υ t d ˆσ t d ˆσ t log S t Υt 4 Υ t d ˆσ t Y t = T t N N d t, n d t, Υ t Proof. From 3. we obtain d ˆσ t + d ˆσ t C t = c St,, Υ t + c Υ St,, Υ t T t d d ˆσ t, C t = c St,, Υ t + c Υ St,, Υ t T t d d ˆσ t n d t, Υ t ˆσ t d 4.8 d ˆσ t ], d ˆσ t. 4.9 + c ΥΥ St,, Υ t T t d d ˆσ t + cυ St,, Υ t T t d d ˆσ t. Computing the partial derivatives of the Black-Scholes function cs,, Υ yields C t = N d t, + n d t, d Υ t ˆσ t d ˆσ t, 4. C t = n d t, [ + log S t Υt + Υ d t ˆσ t d ˆσ t 4. + 4 log ] S t Υt 4 Υ t d ˆσ t d ˆσ t + Υ t d ˆσ t d ˆσ t. Now the result follows by inserting 4., 4. into 4., 4.3. Note that by using 4., 4., we could express the strong static no-arbitrage restrictions C t, and C t > of Proposition 4. in terms of the implied volatilities ˆσ. However, it is not at all obvious how one could parametrize those implied volatility curves ˆσ t which satisfy the resulting conditions. In contrast, Theorem 4.6 says that any positive-valued local implied volatility curve X satisfying the simple bound in Proposition 4.7 is compatible with the static no-arbitrage restrictions in an admissible model. In other words, absence of dynamic arbitrage can be characterized and ensured by imposing only conditions on the coefficients in the dynamics; there are no restrictions on the state space of the X and Y, which can take arbitrary values in, and R, respectively. We therefore argue that this new parametrization is very natural and convenient for constructing models. Remark. In [4], we have shown how one can parametrize an infinite family of option price processes for one fixed payoff function e.g., calls with one fixed strike and all maturities T >. The parametrization chosen in [4] is via the forward implied volatilities over the remaining time to maturity, and we have shown that this is very convenient to analyze in terms of absence of dynamic arbitrage. The local implied volatilities and price level introduced here play the analogous role for models of calls with one fixed maturity and all strikes >. It remains to give the Proof of Proposition 4.5. By 4.5 and the assumption in Proposition 4.5, we have for any, I C t C t = Y t k dh X Yt thh a dh dk = X + log a thh X log k ta X ta dk. T t T t N N

For any x t >, z t > we have d d c z t,, T tx logzt/ t = N T tx t T t xt for all, and therefore c z t,, T tx t c zt,, T tx log zt t = x N t So clearly x t = X t a, z t = exp X t a Y t a dh X + log a thh X + ta T tx ta satisfy 4.4. To see the uniqueness of x t, z t, note that by using 4., then 4.4, and finally 4., we obtain for I X t a = X t = T tc t n N C t = T tx t log k x t T t dk. 4. n N c zt,, T tx t = x T t c zt,, T txt t. Moreover, 4. shows that c z t,, T tx t c zt,, T tx t is strictly increasing in zt, so the uniqueness of z t follows. 4. Examples We now illustrate the concept of local implied volatilities in several stock models. By Theorem 4.6, local implied volatilities and price level could have been equivalently defined via 4.5. This equation and along with it also the new parametrization can be motivated quite naturally as an option pricing formula in a certain stock martingale model driven by a one-dimensional Brownian motion, as explained in our first example. Example 4.9. Suppose that the stock price process S = S t t T satisfies S T = F W T 4.3 for a Brownian motion W under a pricing measure Q and a strictly increasing bijective differentiable function F : R I,. Suppose further that both S and W generate the same filtration F t t T. Our aim is now to fit the function F to a given admissible price curve C of call option prices at time. Under absence of arbitrage we obtain for the call option prices C t with [ ] [ C t = E Q ST + ] Ft = EQ F WT T W t + t T t + Wt F t = E Q [ F ] + W T t + y [ = E Q I {F W T t+y k} dk] = = = y= W t y= Wt [ ] Q F W T t + y k dk Q [ W F k y T t ]dk N y= Wt y= Wt Wt F k T t dk 4.4 for t [, T]. In particular, for = we obtain the underlying stock price model S t = C t. The local implied volatilities are now introduced as a straightforward parametrization of the function F. Since F : R I is strictly increasing, bijective, and differentiable, for a fixed > there exists a unique integrable function f : I [, ] with F k = k fhdh. If we define and Y t := W t, then F k = k dh Xhh Xk := fkk k 4.5 and 4.4 becomes equation 4.5. Solving this for Xk and Y then leads to Theorem 4.6 and shows how one can fit local implied volatilities, and hence the function F, to the initial price curve C. 3

Theorem 4.6 says that the admissible price curves C at time are in a one-to-one relation to positive local implied volatilities Xk and real-valued price levels Y. The above calculations show that the admissible price curves C are also in a one-to-one relation to strictly increasing bijective differentiable functions F : R, as in 4.3. This reflects the well-known result that modelling statically arbitrage-free price curves of T-maturity calls is equivalent to specifying the distribution of the stock price S T under the pricing measure. Our parametrization of the arbitrage-free option prices then arises quite naturally via the nonnegative quantities fk parametrizing the set of functions F k as above. Once the representation 4.4 is found, the scaling in the definition 4.5 of the local implied volatilities Xk is chosen to establish compatibility with the Black-Scholes model, in the sense that Xk is constant and yields the volatility parameter if S follows a geometric Brownian motion. Example 4.. For the Black-Scholes model with volatility σ > we have S t = S exp σ W t σ t, where W is a Brownian motion under the risk-neutral measure, and ˆσ t = σ for all >. So 4.8 simplifies to X t = σ, recovering the identity 4. again, and 4.9 yields Y t = T t d t, = σ log S t σ log T tσ = W t + σ logs / Tσ. Plugging this into the expression for z t in Proposition 4.5 readily shows that z t = S t in the Black-Scholes model. Example 4.. In Heston s [3] stochastic volatility model, the stock and instantaneous variance are modelled by the -dimensional diffusion S, ζ given by ds t = S t ζt d W t, dζ t = κ θ ζ t dt + ν ζ t ρd W t + ρ d W t for a -dimensional Brownian motion W, W under the risk-neutral measure, with constants κ, θ, ν > and ρ,. Stochastic volatility models can reproduce implied volatility smiles and skews, and Heston s model is a popular choice in practice since there exists a semi-closed formula for call prices see [3] which allows fast calibration of the model parameters to market prices. Because of the complicated structure of the price formula, however, no simple expression for local implied volatilities in Heston s model seems to be available. Numerical calculations show that for typical parameter values, local implied volatilities exhibit a similar but more pronounced smile and skew structure than classical Black-Scholes implied volatilities; see Figure..35 T t. local implied vol class implied vol.35 T t.5 local implied vol class implied vol.3.3.5.5...5.7.8.9...3 S.5.7.8.9...3 S Figure : Classical versus local implied volatility as function of moneyness S t in Heston model for different times to maturity. Parameter values are κ = 3, θ =., ν =.45, ρ =.5 and ζ t =.. 4.3 Arbitrage-free dynamics of the local implied volatilities In this section, we derive the dynamics of the local implied volatilities under absence of arbitrage. Let W be an m-dimensional Brownian motion on Ω, F, P, F = F t t T the P-augmented filtration generated 4

by W, and F = F T. We suppose that we have positive processes X t for a.e. >, satisfying the condition in Proposition 4.7, and a real valued process Y t with P-dynamics dx t = u t X t dt + v t X t dw t t T, 4.6 dy t = β t dt + γ t dw t t T, 4.7 where β, u L loc R, and γ, v L loc Rm for a.e.. We also suppose that u, v are uniformly bounded in ω, t, and that the initial local implied volatility curve satisfies dh X h < for all >. Now define processes C t, by 4.5, so that X t, Y t are by construction and Theorem 4.6 the local implied volatilities and price level of the option prices C t,. Remember that S t = C t, and note that for defining C t via 4.5, the values X t are not needed. Our aim is now to show that the existence of a common equivalent local martingale measure for C for all is essentially equivalent to the drift restrictions β t = Y t T t u t = T t [ γt γ t b t, 4.8 γ t + v th X dh thh + Y t dh X thh γ t + ] v th X dh thh v t + v t v t b t 4.9 for a market price of risk process b L loc Rm. More precisely, we have the following result. Theorem 4.. a If there exists a common equivalent local martingale measure Q for all C, then there exists a market price of risk process b L loc Rm such that 4.8, 4.9 for a.e. > hold for a.e. t [, T], P-a.s. b Conversely, suppose that the coefficients β, γ, u and v satisfy, as functions of Y t and X t, the relations 4.8, 4.9 for a.e. > for a.e. t [, T], P-a.s. for some bounded uniformly in t, ω process b L loc Rm. Also suppose that there exists a family of continuous adapted processes X >, Y satisfying the system 4.6 for a.e. > and 4.7. Then there exists a common equivalent local martingale measure Q on F T for C. One such measure is given by dq dp := E b dw, 4. T where E is again the stochastic exponential. c In the situation of a or b, the dynamics of C under Q are given by dc t = n Y t k dh X thh k v t h γ t + T t T t X t hh dh dk d W t 4. for and a Q-Brownian motion W = W b s ds. The equations 4.8, 4.9 for the local implied volatility setting are the analogues to the drift restrictions.3,.4 in [4] for the forward implied volatility modelling. Note that the free input parameters are the market price of risk process b as well as γ and the family of processes v for all, i.e., the volatilities of the state variables Y and X; they determine the drifts β and u via 4.8, 4.9. Note also that since S t = C t, the volatility σ t of the stock price process ds t = σ t S t d W t can easily be derived from 4. and 4.5 as σ t = n Yt k dh X t hh T t T t γ t + k v th X thh dh dk / N Yt k dh X t hh T t dk. This implies that if γ or the volvols v are random, we obtain for the stock price S a model with a certain quite specific stochastic volatility. Whether or not this is Markovian depends on γ, v. Example 4.3. Let X be a positive measurable function on, satisfying the condition in Proposition 4.7 and Y R. Take m =, γ, and v for all. Then Theorem 4. yields β = b, u = for all, and thus Y t = W t t b sds and X t = X for all. Hence we recover the arbitrage-free one-factor model with constant strike-dependent local implied volatility of Example 4.9. In Section 5, we construct more generally arbitrage-free option price models with stochastic and thus potentially more realistic local implied volatility processes. 5

The remainder of this section is devoted to the proof of Theorem 4.. We use Proposition 4.4. Let Z t k := Y t k dh X. Under P, the dynamics of C thh t for each fixed are then given by dc t = + n Ztk T t n Ztk T t [ T t Z tk T t T t γ t + k γ t + k v th X thh dh dk dw t. v th X dh thh + β t k v t h uth X thh dh ] dk dt Proof. Formally this follows from applying Itô s lemma under the integral in 4.5 and then using 4.6, 4.7. Using the condition in Proposition 4.7 one can show that we may apply Fubini s theorem for stochastic integrals see Protter [38], Chap. IV, Theorem 65 to justify interchanging the dk-integral and the stochastic integral. A detailed proof can be found in [46], Section 4.7.3. Proof [ of Theorem ] 4.. a Since F is generated by W, Itô s representation theorem implies that we have E dq F t = E bdw for some process b L t loc Rm, and dp W := W b t dt is a Q-Brownian motion by Girsanov s theorem. Now Proposition 4.4 yields dc t = where µ t k := Z tk T t n Ztk T t T t µ t kdk dt + γ t + k n Ztk T t T t γ t + k v th X thh dh dk d W t, 4. v th X dh thh + β t k v t h uth X thh dh + γ t + k v th X dh thh b t for k >. Since the C are local Q-martingales for all, by Fubini s theorem we have P-a.s. for a.e. t µ t k = for a.e. k, 4.3 and then for all k by continuity of µ t in k. Letting k in 4.3, we obtain 4.8. Finally, 4.9 follows after a straightforward calculation if we differentiate 4.3 with respect to k. b Define dq dp := E bdw T on F T; then W := W b t dt is a Q-Brownian motion on [, T] by Girsanov s theorem. Another lengthy but straightforward calculation shows that 4.8 and 4.9 imply µ t k = for all k. Plugging this and dw t = d W t + b t dt into Proposition 4.4, we obtain c under b. It now follows easily from c that C for all are Q-local martingales on [, T]. c The assertion under b has been proved together with b above. Under a the assertion follows from 4. and 4.3. 5 A class of arbitrage-free local implied volatility models In this section, we apply the existence and uniqueness results of [44] to the infinite system 4.6, 4.7 of SDEs arising in Section 4, providing explicit examples of arbitrage-free local implied volatility models. This requires some additional work: An existence result for general SDEs like in [44] uses assumptions on both the drift and the volatility coefficients, but in the case of our system 4.6, 4.7 we may only choose the volatility coefficients γ, v. Our aim is therefore to find conditions on the coefficients γ, v such that the drift coefficients β, u given by 4.8, 4.9 behave nicely and the results of [44] can be applied. We first adapt the framework for infinite systems of SDEs developed in [44] to the present setup in Section 5.. Then we provide an existence result in Section 5.. We generalize Example 4.3 to nonzero v and hence to stochastic local implied volatilities. Our approach is similar in spirit to the existence results in [44, Section 5] for interest rate term structure models or in [4, Section 3] for forward implied volatility term structure models. 6