17
Rigorous derivations of the Black-Scholes formula***

In Chapter 16 we provided a derivation of the Black-Scholes formula based on limiting arguments. Indeed, we have shown that when the time step in a binomial tree becomes increasingly small, the geometric random walk followed by the stock price gradually approaches that of a geometric Brownian motion. Although it provides a lot of the intuition and explains where many results come from, the approach lacks some rigor.

This chapter intends to fill these gaps by providing a more advanced treatment of the BSM model based upon advanced tools such as stochastic calculus (see Chapter 15), partial differential equations and changes of probability measures. Our main objective is to provide two rigorous derivations of the Black-Scholes formula using either partial differential equations or changes of probability measures. This chapter is therefore targeted at readers with a stronger mathematical background and who have read Chapter 15. This chapter is not mandatory to understand the upcoming chapters.

More specifically, the learning objectives are to:

  • distinguish an ordinary differential equation (ODE) from a partial differential equation (PDE);
  • understand the link between PDEs and diffusion processes as given by the Feynman-Kač formula;
  • derive and solve the Black-Scholes PDE for simple payoffs;
  • understand how to price and replicate simple derivatives with the Black-Scholes PDE;
  • apply a change of probability measure to random variables and to Brownian motions;
  • understand the role of the Fundamental Theorem of Asset Pricing to determine the no-arbitrage price of a derivative;
  • compute the price of simple and exotic derivatives using the risk-neutral probability measure;
  • using either the Black-Scholes PDE or the change of measure techniques, derive the Black-Scholes formula with the stop-loss transform.

17.1 PDE approach to option pricing and hedging

In their seminal paper [19], Fischer Black and Myron Scholes derived the dynamics of the self-financing replicating portfolio of a call option using partial differential equations. This has led to an increasing use of PDEs in mathematical finance and to an approach known as the PDE approach to option pricing.

Our objective is not to provide an introduction to the theory of PDEs; our only objective here is to lay down sufficient background material to understand the role of the Black-Scholes PDE in the replication and pricing of simple derivatives.

17.1.1 Partial differential equations

First, let us look at the following quadratic equation:

(17.1.1)numbered Display Equation

For given constants a, b and c, the goal is to obtain the value(s) of x such that ax2 + bx + c = 0 is verified. We know that, depending on the values of a, b and c, the value of the discriminant b2 − 4ac will decide whether the quadratic equation in (17.1.1) has one or two solutions, or even no (real) solution.

In physics and biology, it is common for a function F(x) describing a physical/biological quantity to be the solution of an equation of the form

(17.1.2)numbered Display Equation

with a, b and c being known constants. An equation like the one in (17.1.2) is called a differential equation and solving it means finding an analytical expression for F(x).

One one hand, solving a quadratic equation as (17.1.1) means looking for a number x whereas on the other hand, solving a differential equation as (17.1.2) means looking for a function F(x). In both cases, the existence of a solution depends on certain conditions. In what follows we will consider ordinary differential equations and then partial differential equations.

An ordinary differential equation (ODE) is a differential equation whose solution (if it exists) is a function of only one variable, say x, and which gives a relationship between the function F(x) itself and its derivatives F′(x), F′′(x), etc.

Example 17.1.1A simple ODE

Let b be a known constant and let us find a function F(x) such that:

numbered Display Equation

for all . It is easy to verify that, for this ODE,

numbered Display Equation

is one possible solution, for any value of a.

 ◼ 

A partial differential equation (PDE) is a differential equation whose function we are looking for has two variables, say t and x. The PDE is thus expressed in terms of the partial derivatives of F, namely , , , etc.

Example 17.1.2A simple PDE

Let us find a function F(t, x) such that:

(17.1.3)numbered Display Equation

for all . For this PDE, we can easily verify that the function

numbered Display Equation

is such that its partial derivatives ∂F/∂t and ∂2F/∂x2 satisfy the relationship in (17.1.3). Indeed,

numbered Display Equation

Note that the function F(t, x) = 0, for all t and x, is also a solution to this PDE.

 ◼ 

In the previous examples, we found more than one solution to the same ODE and PDE, respectively. In order to guarantee the uniqueness of a solution, ODEs and PDEs require initial or final conditions, known as boundary conditions.

For the ODE in example 17.1.1, if we add an initial condition such as F(0) = 1, then

numbered Display Equation

becomes the only solution.

Without getting into the details, in what follows we will specify final conditions to obtain unique solutions to our PDEs. More details below.

In most cases, it is difficult or even impossible to find an explicit solution to a PDE. In those cases, numerical methods are used to obtain approximations of the solution. However, for one class of PDEs, the one most commonly used in finance, there is a way to obtain the solution using stochastic calculus, as presented in Chapter 15; this is what we will describe next.

17.1.2 Feynman-Kač formula

The Feynman-Kač formula provides a solution to PDEs in the family of parabolic PDEs. This solution is expressed as the conditional expectation of a diffusion process sampled at a fixed time. Here is a simple version of the Feynman-Kač formula, which is adapted to the broader objectives of this chapter.

As always, assume that T > 0 is fixed. For a given parameter r > 0 and given functions a( · ), b( · ), h( · ), we seek to find a solution to the PDE

(17.1.4)numbered Display Equation

for all , with final condition F(T, x) = h(x), for all .

According to the Feynman-Kač formula, the solution to this PDE can be written as follows: for each (t, x) in the domain, we have

(17.1.5)numbered Display Equation

where {Xs, s ⩾ 0} is the diffusion process determined by the following SDE:

(17.1.6)numbered Display Equation

When trying to solve a PDE of the form given in (17.1.4), we can use the Feynman-Kač formula by following these steps:

  • INPUTS: parameter r > 0 and functions a( · ), b( · ), h( · ), as given by the PDE.
    1. Find the diffusion process {Xs, s ⩾ 0} given by equation (17.1.6), using stochastic calculus, as seen in Chapter 15.
    2. For each t and x in the domain, determine the conditional probability distribution of XT given that Xt = x.
    3. Compute the conditional expectation in equation (17.1.5).
  • OUTPUT: function F(t, x).

The following example illustrates this methodology.

Example 17.1.3Solving a simple PDE

We would like to find the solution to the PDE

numbered Display Equation

with boundary condition F(T, x) = x2, where σ is a given constant.

Here, our inputs are a(x) = 0 and b(x) = σ, h(x) = x2 and r = 0. Consequently, looking at (17.1.6), we consider the following SDE:

numbered Display Equation

whose solution is simply Xs = σWs. Now, using h(x) = x2 and r = 0 in (17.1.5), the solution is

numbered Display Equation

As already computed in Chapter 14 (see equation (14.3.2)), we can write

numbered Display Equation

In conclusion, the function F(t, x) = x2 + σ2(Tt) is the solution to the above PDE.

For validation purposes, let us check this. First, we can verify that the final condition F(T, x) = x2 + σ2(TT) = x2 is indeed equal to h(x) = x2. Second, we have

numbered Display Equation

Therefore,

numbered Display Equation

as expected.

 ◼ 

17.1.3 Deriving the Black-Scholes PDE

Let us recall that the Black-Scholes model can be summarized as a continuous-time financial market model composed of a risk-free asset {Bt, t ⩾ 0} and a risky asset {St, t ⩾ 0}. The dynamics of the price of the risk-free asset can be determined by the following ODE:

(17.1.7)numbered Display Equation

whose solution is simply Bt = B0ert. The dynamics of the price of the risky asset can be determined by the following SDE:

(17.1.8)numbered Display Equation

whose solution is a geometric Brownian motion: .

Let us consider a simple option with payoff VT = h(ST). We want to compute the no-arbitrage price of this derivative and find its replicating portfolio, as defined in Chapter 16.

We make the following assumptions: the time-t value Vt of this option can be written as F(t, St), where F(t, x) is a sufficiently smooth function; assume also that there exists a replicating portfolio {(Θt, Δt), 0 ⩽ tT} for this payoff, i.e. VT = ΠT, where

numbered Display Equation

and where the only cash flows of this portfolio are at time 0 and time T. Then, by the no-arbitrage principle, we have that Vt = Πt, for all 0 ⩽ tT, where

numbered Display Equation

is the value at time t of this replicating portfolio. By the above assumptions, we then have that F(t, St) = Πt, for all 0 ⩽ tT. At this stage, we do not have an expression for the function F nor for the replicating portfolio {(Θt, Δt), 0 ⩽ tT}: we only assume that they exist. Our goal is to find explicit expressions.

To meet this objective, we need to find a trading strategy that is both replicating and self-financing. This is similar to what we did in the binomial model, except that we are now in a continuous-time model.

In a continuous-time model, the self-financing condition is given by the following stochastic equation:

(17.1.9)numbered Display Equation

It can be obtained as the limit of equation (11.2.3) in Chapter 11. In words, the variation in the portfolio value comes from changes in values from both assets within the portfolio.

Moreover, since F(t, St) = Πt, for all 0 ⩽ tT, we must also have that

numbered Display Equation

meaning that, at each instant, the variation in the portfolio value must match the variation in the derivative’s price. This is also known as the replicating condition.

We are now ready to derive the Black-Scholes PDE. First of all, F(t, St) is a transformation of a diffusion process. Therefore, using Ito’s lemma, we can write

numbered Display Equation

Substituting dSt, as given by equation (17.1.8), we further have

(17.1.10)numbered Display Equation

On the other hand, since Πt = F(t, St), the self-financing condition of (17.1.9) can be rewritten as

(17.1.11)numbered Display Equation

where we used the fact that ΘtBt = F(t, St) − ΔtSt (from the definition of the portfolio’s value) and where we have substituted the expressions for dBt and dSt as given in equations (17.1.7) and (17.1.8), respectively.

We have obtained two equivalent stochastic dynamics for F(t, St): one in equation (17.1.10) and one in equation (17.1.11). This means that for these two SDEs, the terms in front of dt must be equal and the terms in front of dWt must be equal. Consequently, we have the following system of equations:

numbered Display Equation

From the second equation, we easily find that

numbered Display Equation

Then, using this in the first equation, we deduce the so-called Black-Scholes PDE or Black-Scholes equation: if the function F(t, x) is such that

(17.1.12)numbered Display Equation

with boundary (final) condition F(T, x) = h(x), where h is the payoff function, then the above system of equations is verified.

To summarize, we have obtained that the time-t value of a simple derivative with payoff VT = h(ST) is given by F(t, St), where the function F(t, x) is the solution to the Black-Scholes PDE. The PDE comes from the model assumptions (dynamics of each asset) whereas the payoff of the option provides the terminal condition that the PDE should abide to. This boundary condition assures that we have a unique solution to this PDE.

Once we have found the price F(t, St) for the payoff VT = h(ST), then the replicating portfolio {(Θt, Δt), 0 ⩽ tT} is given by:

(17.1.13)numbered Display Equation

at every time t ∈ [0, T).

image It is also important to notice from the Black-Scholes equation (17.1.12) that the parameter μ has completely disappeared. Consequently, the option price (through function F) will not/does not depend on this parameter. This is similar to the result obtained in the binomial tree: the expected stock return under the real-world probability measure, which in that case was determined by the parameter p, is not relevant to determine the no-arbitrage price of a derivative.

In conclusion, if we can find a solution F(t, x) to the Black-Scholes PDE in (17.1.12), we will have a full solution to our pricing and hedging problem: at each time t, we have

  • the price of the derivative Vt = F(t, St);
  • the replicating portfolio (Θt, Δt) given by
    numbered Display Equation

17.1.4 Solving the Black-Scholes PDE

Luckily enough, under the assumptions of the BSM model, the Black-Scholes PDE (17.1.12) can be solved using the Feynman-Kač formula. Indeed, since F(t, x) is a solution of (17.1.12), then, from equation (17.1.5), it can be written in the following form:

numbered Display Equation

where {Xs, s ⩾ 0} is determined by the SDE

numbered Display Equation

This is simply a geometric Brownian motion with drift r whose solution we already found earlier: Xs = X0exp {(r − σ2/2)s + σWs}.

Using the standard arguments, such as formula (14.3.2), we get:

(17.1.14)numbered Display Equation

This is the pricing formula for simple derivatives with payoff VT = h(ST) obtained with the PDE approach. Note that we then have an explicit expression for the replicating portfolio as given in (17.1.13).

Let us illustrate what we have just obtained with a forward contract.

Example 17.1.4Forward contract

The payoff of a (long) forward contract with delivery price K is VT = STK. Then, using the pricing formula in (17.1.14) with h(x) = xK, we get

numbered Display Equation

From the formula in (14.1.2) for the moment generating function of a normal distribution, we know that

numbered Display Equation

and, consequently,

numbered Display Equation

This means that the time-t value of a forward contract is given by F(t, St) = St − e− r(T − t)K, as we already knew from Chapter 3. In fact, this was a model-free result, so it had to be true in any model, including the Black-Scholes-Merton model.

Since we have

numbered Display Equation

the replicating portfolio is given by

numbered Display Equation

and

numbered Display Equation

Note that it is a static strategy, i.e. it is constant with respect to time. In other words, to replicate a long forward contract, we must hold (at any time t) one unit of the underlying asset and be short e− rTK units of the risk-free asset. This is also a result we had obtained in Chapter 3.

 ◼ 

Note that we have found our first solution to the Black-Scholes PDE (17.1.12): it is given by the function F(t, x) = x − e− r(T − t)K. We will soon find other solutions.

17.1.5 Black-Scholes formula

Let us now specialize the above results to the case of a call option. Similar to what we did before, we will substitute the notation F(t, x) by C(t, x). In this case, the final condition become C(T, x) = (xK)+ and by (17.1.14) we have

numbered Display Equation

Note that this expectation is the stop-loss transform of a lognormally distributed random variable. Indeed, we have

numbered Display Equation

Consequently, using once again the stop-loss formula in (14.1.7) of Chapter 14, we have

numbered Display Equation

where the functions d1(t, x) and d2(t, x) have been defined in (16.3.3) of Chapter 16.

Again, we can verify (with tedious calculations) that C(t, x) is indeed a solution to the Black-Scholes PDE (see the Greek letters and their relationship to the Black-Scholes equation (PDE) in equation (20.5.7)). We could also show that as we approach maturity, then

numbered Display Equation

As a byproduct of the above verification, we get

numbered Display Equation

from which we deduce the replicating portfolio of a call option:

numbered Display Equation

This is consistent with the results previously derived in Chapter 16.

Finally, using the put-call parity, we can deduce the analytic expression of the put option price P(t, St). This can also be achieved by exploiting the linearity of the Black-Scholes PDE (see the exercises).

17.2 Risk-neutral approach to option pricing

In the binomial tree model, we obtained the price of a derivative as the value of its replicating portfolio and then we reorganized the expression of this price using weights q and 1 − q. In other words, the price of the derivative was also given by a discounted risk-neutral expectation, i.e. an expectation taken with respect to another probability measure as opposed to the actuarial/real-world probability measure . See, for example, the expectation in (9.3.3).

Thanks to the FTAP, as discussed in Section 13.3.2 of Chapter 13, we can hope to find such a risk-neutral probability in the Black-Scholes-Merton model since this model is assumed to be free of arbitrage opportunities. This is the foundation of a probabilistic approach1 called the risk-neutral approach or the martingale approach to option pricing, as opposed to the PDE approach.

Therefore, the objective of this section is to obtain directly the value of both a call and a put option in the Black-Scholes-Merton model, without using the binomial limiting argument. Our goal is to formalize some of the statements made in Section 16.2.4 about risk-neutral probabilities. In that section, we said that there exists a risk-neutral probability such that, for any time 0 < tT, we have

numbered Display Equation

or, equivalently,

numbered Display Equation

We will see how to rigorously obtain this artificial probability measure and the corresponding risk-neutral dynamics of the risky asset price.

Before going any further, we will recall the definition of a probability measure, as seen in an introductory probability course, and then see how to introduce another probability measure in the same probability model using a technique known as a change of probability measure.

17.2.1 Probability measure

In elementary and more advanced probability theory, for a given sample space Ω, we define a probability measure as a real-valued mapping defined on the set of all events, i.e. subsets of Ω, with the following properties:

  1. for any event E⊆Ω, we have ;
  2. ;
  3. for any sequence of disjoint events E1, E2, …⊆Ω, we have
    numbered Display Equation
    This last property is known as infinite additivity and it implies finite additivity: for any n disjoint events E1, E2, …, En⊆Ω, we also have
    numbered Display Equation

Example 17.2.1Throwing a regular die

We throw a well-balanced die, as we did in Chapter 1. Therefore, the sample space Ω is

numbered Display Equation

For this experiment, we should define by:

numbered Display Equation

The mapping assigns the value 1/6 to each event. It is easy to see that this meets the three criteria necessary to be called a probability measure.

 ◼ 

In an introductory probability course, the modelling is usually done with a fundamental space Ω and only one probability measure (rarely defined in an explicit way). Unfortunately, it creates the false impression that is just a symbol with no particular meaning and/or specific definition.

Example 17.2.2Probability measures for economic states

Assume the sample space of economic states for the upcoming year is

numbered Display Equation

According to actuary A, the probability of observing these events is:

numbered Display Equation

But actuary B is more pessimistic. According to her, the probability of observing these events is:

numbered Display Equation

Verifying the three conditions in the definition of a probability measure presented above, we have thus defined two probability measures and for the same experiment.

 ◼ 

We say that and are equivalent probability measures if, for any event E, we have

numbered Display Equation

or, equivalently,

numbered Display Equation

This is often written as . More or less, two probability measures are equivalent if they agree on all impossible events, which means then that they also agree on all events of probability equal to one. For any other event, they can differ.

Example 17.2.3Probability measures for economic states (continued)

Because actuaries A and B agree that all four events have a non-zero probability, we have that .

A third actuary (C) assesses the likelihood of each event of Ω and concludes that an economic boom is impossible. Mathematically, . Therefore is not equivalent to and is not equivalent to .

 ◼ 

17.2.2 Changes of probability measure

Let us now present a methodology leading to a change of probability measure in a given probability model.

Let be a probability measure on a sample space Ω, both designed for a given experiment. Now, choose a random variable Z such that:

  1. Z ⩾ 0;
  2. .

For such a given random variable Z, we can define another probability measure as follows: for any event E⊆Ω, set

(17.2.1)numbered Display Equation

where we should stress that the expectation is taken with respect to . The notation emphasizes that for each such random variable Z, there is a new probability measure attached to it. In probability theory, Z is known as the Radon-Nikodym derivative of with respect to . This is just terminology as no derivatives are really involved.

We need to make sure that meets the three conditions to be a probability measure. It is rather easy to verify that for any event E⊆Ω, we have and . A little bit more work is needed to verify that for any sequence of disjoint events E1, E2, …⊆Ω, we have

numbered Display Equation

In conclusion, for each such random variable Z, we now have a methodology to construct a new probability measure , usually different from the existing .

Let us now compute expectations of the form with respect to the probability , i.e. compute weighted averages using the weights given by . First, by definition (17.2.1), we have

numbered Display Equation

Indeed, since the random variable is a Bernoulli random variable, i.e. it takes only the values 0 and 1, then

numbered Display Equation

The result follows by the definition in (17.2.1).

More generally, one can show that2 for any random variable X, we have a similar relationship:

(17.2.2)numbered Display Equation

The expectation of an arbitrary random variable X computed with the probability measure is equal to the expectation with respect to where X is distorted by the random variable Z.

To successfully use the above methodology, we first need to choose a positive random variable Z with unit mean. Then, we can determine the distribution of another random variable X with respect to or, in other words, the impact of the change of measure on the distribution of X.

Example 17.2.4Change of measure for a normal distribution

Let us consider the case of a normally distributed random variable where the notation emphasizes that the distribution is affected by the choice of the probability measure ( in this case). This statement means that: for any , we have

numbered Display Equation

An equivalent characterization of the normal distribution is given by its m.g.f.: for any , we have

numbered Display Equation

See Chapter 14.

Fix a real number α. Let us now find the distribution of X with respect to a probability measure defined with the random variable

numbered Display Equation

For simplicity, we have chosen to write instead of .

First of all, it is easy to verify (see exercise 17.4) that Zα is indeed such that

  1. Zα ⩾ 0;
  2. .

Now, let us show that . Using the characterization of the normal distribution by its m.g.f., it suffices to verify that

numbered Display Equation

for all , where stands for the expectation with respect to . Indeed, using (17.2.2), we have

numbered Display Equation

where, in the last step, we used the fact that or, equivalently, that . The result follows.

 ◼ 

This last example shows that it is possible to apply a well-chosen change of measure to shift the mean of a normally distributed random variable, i.e. to move from a mean of 0 under to a mean of α under . Note that the variance is not affected by this change of measure. This will be of paramount importance for linear Brownian motions and the Black-Scholes-Merton model.

Example 17.2.5Change of measure for an exponential distribution

Let us now consider an exponentially distributed random variable , with α > 0. This means that, with respect to , the random variable X admits the following probability density function:

numbered Display Equation

Again, an equivalent characterization of the exponential distribution is given by its m.g.f. Indeed, for any λ < α, we have

(17.2.3)numbered Display Equation

It is possible to change the probability measure from to to make X an exponentially distributed random variable with a larger parameter β > α with respect to . Indeed, it suffices to use

numbered Display Equation

to define .

Again, it is easy to verify (see exercise 17.5) that Zβ is such that

  1. Zβ ⩾ 0;
  2. .

Let us now compute the m.g.f. of X with respect to :

numbered Display Equation

where in the second last step we used the formula in (17.2.3). Since

numbered Display Equation

for all λ < β, we can conclude that .

 ◼ 

Likelihood ratio

One might wonder how to choose the random variable Z in order to change the probability distribution of X (with respect to the original probability ) to a prescribed distribution with respect to the probability measure .

For a continuous random variable, this is given by a ratio of probability density functions (p.d.f.). Assume that X has a p.d.f. fX(x) with respect to and that we are looking for the change of probability measure such that we would obtain the p.d.f. fZX(x) with respect to . In particular, after the change of measure, we should have the following:

numbered Display Equation

This is the expectation of X with respect to .

First, we know that the -expectation of the random variable given by

numbered Display Equation

is given by

numbered Display Equation

In other words, we must choose:

numbered Display Equation

We easily verify that Z ⩾ 0, as the p.d.f.s are positive functions and

numbered Display Equation

Similarly, for discrete random variables, a similar change of probability measures, given by the ratio of probability mass functions, will allow a prescribed distribution to be obtained.

17.2.3 Girsanov theorem

The last two examples have shown the effect of a given change of measure on the distribution of a random variable X. In the Black-Scholes-Merton model, we will need to change the distribution of a whole stochastic process, namely that of a linear Brownian motion.

This situation will be taken care of by the so-called Girsanov theorem, which is essentially the analog for Brownian motions of the change of measure discussed in example 17.2.4 for normal random variables. The Girsanov theorem is the last building block needed before we can finally determine the no-arbitrage price of a derivative and obtain risk-neutral pricing formulas. In this section we present a simple version of the latter.

Consider a -standard Brownian motion W = {Wt, 0 ⩽ tT}, i.e. a standard Brownian motion under , and for a fixed number θ define the process Wθ = {Wθt, 0 ⩽ tT} by

numbered Display Equation

The Girsanov theorem states that if we use the random variable

numbered Display Equation

to obtain the probability measure , then Wθ = {Wθt, 0 ⩽ tT} will be a -standard Brownian motion.

At first, this result can be rather confusing. Indeed, by definition W is a -SBM and since

numbered Display Equation

then Wθ is a linear Brownian motion with drift θ with respect to . On the other hand, since by the Girsanov theorem Wθ is a -SBM and

numbered Display Equation

then W is a linear Brownian motion with drift of − θ with respect to . We have summarized this discussion in Table 17.1.

Table 17.1 Impact of the Girsanov theorem on (linear) Brownian motions

with respect to
W is a standard BM linear BM
Wθ is a linear BM standard BM

Whereas example 17.2.4 showed how we can change the probability measure to shift the mean of a normally distributed random variable, the Girsanov theorem tells us how to change the probability measure to shift the drift of a linear Brownian motion.

17.2.4 Risk-neutral probability measures

As in the general n-period binomial model and in the trinomial tree model (see e.g. Section 13.3.2), the concept of risk-neutral probability measure is fundamental in the theory of continuous-time derivatives pricing. In any continuous-time model with a stock price process S = {St, 0 ⩽ tT} and a risk-free asset price process B = {Bt, 0 ⩽ tT}, a probability measure is called a risk-neutral probability measure if it is such that:

  1. is equivalent to ;
  2. the stochastic process is a -martingale, i.e. a martingale with respect to the probability measure , which means that, for all 0 ⩽ t1 < t2T,
    numbered Display Equation

The probability measure is also known as an equivalent martingale measure (EMM).

As discussed earlier in this chapter, the equivalence condition means that and have the same zero-probability events or, equivalently, the same probability-one events. Intuitively, if an event is impossible or certain with respect to , then it is still the case under the risk-neutral probability measure .

The martingale condition says that if the risky asset price S is discounted with the risk-free asset B, then, with respect to a risk-neutral probability , this discounted price does not have any remaining trend as it behaves as a martingale.

Using now a more formal definition of probability measure, recall from Section 13.3.2 that the FTAP states that:

  • a market model is free of arbitrage opportunities if and only if there exists at least one risk-neutral probability measure ;
  • an arbitrage-free market model is complete if and only if there exists a unique risk-neutral probability measure . Otherwise the market is said to be incomplete.

17.2.5 Risk-neutral dynamics

Let us now put all the pieces together and determine how to price derivatives in the BSM model using the risk-neutral approach. In the BSM model, the risky asset S is a geometric Brownian motion of the form

numbered Display Equation

and the risk-free asset is such that

numbered Display Equation

with B0 = 1. The resulting financial market is free of arbitrage opportunities.3 Thus, according to the FTAP, if the market is arbitrage-free then there exists at least one risk-neutral probability measure under which the process given by is a martingale.

Dividing St by Bt we get

numbered Display Equation

In other words, the discounted stock price process S/B is still a geometric Brownian motion. We know that this GBM is a -martingale if and only if (see Section 14.5.3). Therefore, if we need to find an EMM, the probability measure is not going to be our candidate. That is why we called the real-world probability measure to mark the difference.

In our quest for an EMM , let us define (using a very suggestive notation) by

(17.2.4)numbered Display Equation

and by

(17.2.5)numbered Display Equation

One has to be careful: recall from Table 17.1 that, with respect to , the new process is not a standard Brownian motion, it is a linear Brownian motion.

Using the Girsanov theorem, if we set with , i.e. if we use the following change of measure

(17.2.6)numbered Display Equation

then is a -standard Brownian motion and, consequently, is a -martingale. This last statement about the martingale property of is backed up by the work we did in Section 14.5.3.

So, the probability measure defined with the random variable in (17.2.6) is the (unique) risk-neutral probability measure in the BSM model. Indeed, it is now easy to verify that is a -martingale. By definition of the geometric Brownian motion S in the BSM model, we have

numbered Display Equation

where in the second last step we used the definition of given in (17.2.4) and in the last step we used the definition of given in (17.2.5).

Since , then we can summarize our findings as: for all 0 ⩽ tT,

numbered Display Equation

We can also say that the continuous-time stochastic process {St, t ⩾ 0} is

  • a GBM with drift coefficient under the real-world/actuarial probability measure ;
  • a GBM with drift coefficient under the risk-neutral probability measure .

Note that the volatility coefficient is σ with respect to both probability measures.

17.2.6 Risk-neutral pricing formulas

Imagine a financial derivative with payoff VT = ST, that is a derivative providing one unit of the underlying risky asset at maturity T. To avoid arbitrage opportunities, this derivative must be worth exactly St, i.e. we must have Vt = St, at any time 0 ⩽ tT. From the definition of a risk-neutral probability measure, we already have that

numbered Display Equation

In other words, for this specific derivative, we have

numbered Display Equation

for all 0 ⩽ tT. This should be reminiscent of the risk-neutral pricing formulas obtained in the general binomial tree model (see e.g. equation (11.3.1)) and in the BSM model (see e.g. Section 16.4).

It turns out that4 for any European payoff VT, simple or exotic, we have a similar risk-neutral pricing formula: for all 0 ⩽ tT,

numbered Display Equation

or, written differently,

(17.2.7)numbered Display Equation

Note that as a particular case of this pricing formula, the initial no-arbitrage price is given by

numbered Display Equation

This is why the risk-neutral probability , under which we can find the no-arbitrage price of any derivative, is also called the pricing probability measure.

Note that, for a simple payoff VT = h(ST), the pricing formula in (17.2.7) coincides with the pricing formula in (17.1.14), obtained with the PDE approach

inline Again, let us emphasize that is an artificial probability measure meant to find the price of a derivative. In other words, it is a byproduct of no-arbitrage pricing. The measure should not be used for risk management purposes such as computing likelihoods of events and scenarios.

Let us put the risk-neutral pricing formula of (17.2.7) to the test.

Example 17.2.6Forward contract

The payoff of a forward contract with fixed delivery price K is given by VT = STK. Then by equation (17.2.7) we have

numbered Display Equation

for any 0 ⩽ tT. Since {e− rtSt, 0 ⩽ t ⩽ T} is a -martingale (by definition of ), we have

numbered Display Equation

and then

numbered Display Equation

as obtained a few times already in this book.

 ◼ 

17.2.6.1 Black-Scholes formula

Now, let us consider the no-arbitrage price of a call and a put option. In particular, if we consider CT = (STK)+, then from (17.2.7) we have

numbered Display Equation

for any 0 ⩽ tT. Since S is a -geometric Brownian motion, it possesses the Markovian property as discussed in Chapter 14. Consequently, as we did in Chapter 16, we can further write

numbered Display Equation

where

numbered Display Equation

Applying once more the stop-loss formula in equation (14.1.7), we obtain a dynamic version of the Black-Scholes formula: for 0 ⩽ t < T,

numbered Display Equation

where

numbered Display Equation

By the put-call parity, we have

numbered Display Equation

and we can recover the Black-Scholes formula for the price of a put option: for all 0 ⩽ tT,

numbered Display Equation

It is important to emphasize that the risk-neutral pricing formula in equation (17.2.7) can be applied to any European payoff VT, not only simple payoffs of the form VT = h(ST). Consequently, we now have a powerful methodology to price exotic options such as Asian, barrier and lookback options, as seen in Chapter 7, using the risk-neutral dynamics of {St, t ⩾ 0} and the corresponding path-dependent payoff functional.

17.3 Summary

Black-Scholes-Merton model

  • Black-Scholes-Merton model  Risk-free asset: {Bt, t ⩾ 0}.
  • Risky asset: {St, t ⩾ 0}.
  • BSM model:
    numbered Display Equation
    or, equivalently,
    numbered Display Equation

PDE approach to option pricing and hedging

  • Assumptions: for a payoff VT = h(ST),
    • Vt = F(t, St), where F(t, x) is a sufficiently smooth function;
    • – exists {(Θt, Δt), 0 ⩽ tT}, with time-t value Πt = ΔtSt + ΘtBt such that:
      numbered Display Equation
  • Objective: find explicit expressions for F(t, x), Θt and Δt.
  • First step toward a solution:
    • F(t, x) is the solution of the Black-Scholes PDE:
      numbered Display Equation
      with final condition F(T, x) = h(x);
    • – the replicating portfolio is given by:
      numbered Display Equation
  • Using the Feynman-Kač formula, we can solve the Black-Scholes PDE:

    numbered Display Equation

Risk-neutral approach to option pricing

  • Change of probability measure: for a positive random variable Z with unit mean, define
    numbered Display Equation
  • Girsanov theorem: if Wθt = θt + Wt, then the probability measure defined with
    numbered Display Equation
    is such that {Wθt, 0 ⩽ tT} is a -standard Brownian motion.
    with respect to
    W is a standard BM linear BM
    Wθ is a linear BM standard BM
  • In the BSM model, if we choose θ = (μ − r)/σ, then is an EMM.
  • With respect to :
    • – {St/Bt, 0 ⩽ tT} is a martingale;
    • – the (risk-neutral) distribution of St is given by
      numbered Display Equation
  • Risk-neutral pricing formula: for any European payoff VT, simple or exotic, we have:

    numbered Display Equation

17.4 Exercises

  1. Verify that the following functions are solutions to the Black-Scholes PDE:

    1. V(t, x) = 0;
    2. V(t, x) = x;
    3. V(t, x) = ert;
    4. V(t, x) = xKe− r(T − t);
    5. .

    In each case, identify the corresponding financial derivative.

  2. Use the Feynman-Kač formula to find the solution to the following PDE problem:

    numbered Display Equation

    where μ, σ are given constants.

  3. Let us consider a financial market such that St = Wt (a SBM) and Bt = 1 for t ⩾ 0. Determine whether the trading strategy (Δt, Θt) given by Δt = 2Wt and Θt = −tW2t is self-financing.

  4. Verify that Zα in example 17.2.4 is such that Zα ⩾ 0 and .

  5. Verify that Zβ in example 17.2.5 is such that Zβ ⩾ 0 and .

  6. Verify that if and if the change of measure is given by

    numbered Display Equation

    then .

  7. Find the explicit change of measure behind the conditional probability associated with a fixed event F such that : .

  8. Suppose that in a financial market, three assets are traded: a risk-free asset such that

    numbered Display Equation

    and two risky assets, such that

    numbered Display Equation

    Note that both risky assets depend on the same Brownian motion W = {Wt, t ⩾ 0}. We would like to determine the conditions such that this market is free of arbitrage opportunities. You will find that

    numbered Display Equation

    must hold to prevent arbitrage opportunities. This is known as the Sharpe ratio or the market price of risk. There are two approaches to reach such a conclusion:

    1. Build a self-financing strategy with both risky assets that replicates the risk-free asset. Show that the Sharpe ratio must hold to prevent arbitrage opportunities.
    2. Invoke the Fundamental Theorem of Asset Pricing and let to show that λ must correspond to the Sharpe ratio to prevent arbitrage opportunities.

Notes

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset