To install click the Add extension button. That's it.

The source code for the WIKI 2 extension is being checked by specialists of the Mozilla Foundation, Google, and Apple. You could also do it yourself at any point in time.

4,5
Kelly Slayton
Congratulations on this excellent venture… what a great idea!
Alexander Grigorievskiy
I use WIKI 2 every day and almost forgot how the original Wikipedia looks like.
Live Statistics
English Articles
Improved in 24 Hours
Added in 24 Hours
Languages
Recent
Show all languages
What we do. Every page goes through several hundred of perfecting techniques; in live mode. Quite the same Wikipedia. Just better.
.
Leo
Newton
Brights
Milds

Black–Scholes equation

From Wikipedia, the free encyclopedia

The surface plot of the value of a European call option over time and price of underlying stock, along with some representative stock price trajectories. If the stock price starts high above the strike price, then it is likely to end up above the strike price. If the stock price starts much below, then it has a small chance of ending up above the strike price. The average value of the trajectories' end-point is exactly equal to the height of the surface.

In mathematical finance, the Black–Scholes equation, also called the Black–Scholes–Merton equation, is a partial differential equation (PDE) governing the price evolution of derivatives under the Black–Scholes model.[1] Broadly speaking, the term may refer to a similar PDE that can be derived for a variety of options, or more generally, derivatives.

Simulated geometric Brownian motions with parameters from market data

Consider a stock paying no dividends. Now construct any derivative that has a fixed maturation time in the future, and at maturation, it has payoff that depends on the values taken by the stock at that moment (such as European call or put options). Then the price of the derivative satisfies

where is the price of the option as a function of stock price S and time t, r is the risk-free interest rate, and is the volatility of the stock.

The key financial insight behind the equation is that, under the model assumption of a frictionless market, one can perfectly hedge the option by buying and selling the underlying asset in just the right way and consequently “eliminate risk". This hedge, in turn, implies that there is only one right price for the option, as returned by the Black–Scholes formula.

YouTube Encyclopedic

  • 1/5
    Views:
    856 534
    38 788
    211 407
    1 260
    21 834
  • Introduction to the Black-Scholes formula | Finance & Capital Markets | Khan Academy
  • The Easiest Way to Derive the Black-Scholes Model
  • 19. Black-Scholes Formula, Risk-neutral Valuation
  • A simple derivation of the Black-Scholes equation
  • Black-Scholes PDE Derivation in 4 minutes

Transcription

Voiceover: We're now gonna talk about probably the most famous formula in all of finance, and that's the Black-Scholes Formula, sometimes called the Black-Scholes-Merton Formula, and it's named after these gentlemen. This right over here is Fischer Black. This is Myron Scholes. They really laid the foundation for what led to the Black-Scholes Model and the Black-Scholes Formula and that's why it has their name. This is Bob Merton, who really took what Black-Scholes did and took it to another level to really get to our modern interpretations of the Black-Scholes Model and the Black-Scholes Formula. All three of these gentlemen would have won the Nobel Prize in Economics, except for the unfortunate fact that Fischer Black passed away before the award was given, but Myron Scholes and Bob Merton did get the Nobel Prize for their work. The reason why this is such a big deal, why it is Nobel Prize worthy, and, actually, there's many reasons. I could do a whole series of videos on that, is that people have been trading stock options, or they've been trading options for a very, very, very long time. They had been trading them, they had been buying them, they had been selling them. It was a major part of financial markets already, but there was no really good way of putting our mathematical minds around how to value an option. People had a sense of the things that they cared about, and I would assume especially options traders had a sense of the things that they cared about when they were trading options, but we really didn't have an analytical framework for it, and that's what the Black-Scholes Formula gave us. Let's just, before we dive into this seemingly hairy formula, but the more we talk about it, hopefully it'll start to seem a lot friendlier than it looks right now. Let's start to get an intuition for the things that we would care about if we were thinking about the price of a stock option. You would care about the stock price. You would care about the exercise price. You would especially care about how much higher or lower the stock price is relative to the exercise price. You would care about the risk-free interest rate. The risk-free interest rate keeps showing up when we think about taking a present value of something, If we want to discount the value of something back to today. You would, of course, think about how long do I have to actually exercise this option? Finally, this might look a little bit bizarre at first, but we'll talk about it in a second. You would care about how volatile that stock is, and we measure volatility as a standard deviation of log returns for that security. That seems very fancy, and we'll talk about that in more depth in future videos, but at just an intuitive level, just think about 2 stocks. So let's say that this is stock 1 right over here, and it jumps around, and I'll make them go flat, just so we make no judgment about whether it's a good investment. You have one stock that kind of does that, and then you have another stock. Actually, I'll draw them on the same, so let's say that is stock 1, and then you have a stock 2 that does this, it jumps around all over the place. So this green one right over here is stock 2. You could imagine stock 2 just in the way we use the word 'volatile' is more volatile. It's a wilder ride. Also, if you were looking at how dispersed the returns are away from their mean, you see it has, the returns have more dispersion. It'll have a higher standard deviation. So, stock 2 will have a higher volatility, or a higher standard deviation of logarithmic returns, and in a future video, we'll talk about why we care about log returns, Stock 1 would have a lower volatility, so you can imagine, options are more valuable when you're dealing with, or if you're dealing with a stock that has higher volatility, that has higher sigma like this, this feels like it would drive the value of an option up. You would rather have an option when you have something like this, because, look, if you're owning the stock, man, you have to go after, this is a wild ride, but if you have the option, you could ignore the wildness, and then it might actually make, and then you could exercise the option if it seems like the right time to do it. So it feels like, if you were just trading it, that the more volatile something is, the more valuable an option would be on that. Now that we've talked about this, let's actually look at the Black-Scholes Formula. The variety that I have right over here, this is for a European call option. We could do something very similar for a European put option, so this is right over here is a European call option, and remember, European call option, it's mathematically simpler than an American call option in that there's only one time at which you can exercise it on the exercise date. On an American call option, you can exercise it an any point. With that said, let's try to at least intuitively dissect the Black-Scholes Formula a little bit. So the first thing you have here, you have this term that involved the current stock price, and then you're multiplying it times this function that's taking this as an input, and this as how we define that input, and then you have minus the exercise price discounted back, this discounts back the exercise price, times that function again, and now that input is slightly different into that function. Just so that we have a little bit of background about what this function N is, N is the cumulative distribution function for a standard, normal distribution. I know that seems, might seem a little bit daunting, but you can look at the statistics playlist, and it shouldn't be that bad. This is essentially saying for a standard, normal distribution, the probability that your random variable is less than or equal to x, and another way of thinking about that, if that sounds a little, and it's all explained in our statistics play list if that was confusing, but if you want to think about it a little bit mathematically, you also know that this is going to be, it's a probability. It's always going to be greater than zero, and it is going to be less than one. With that out of the way, let's think about what these pieces are telling us. This, right over here, is dealing with, it's the current stock price, and it's being weighted by some type of a probability, and so this is, essentially, one way of thinking about it, in very rough terms, is this is what you're gonna get. You're gonna get the stock, and it's kind of being weighted by the probability that you're actually going to do this thing, and I'm speaking in very rough terms, and then this term right over here is what you pay. This is what you pay. This is your exercise price discounted back, somewhat being weighted, and I'm speaking, once again, I'm hand-weaving a lot of the mathematics, by like are we actually going to do this thing? Are we actually going to exercise our option? That makes sense right over there, and it makes sense if the stock price is worth a lot more than the exercise price, and if we're definitely going to do this, let's say that D1 and D2 are very, very large numbers, we're definitely going to do this at some point in time, that it makes sense that the value of the call option would be the value of the stock minus the exercise price discounted back to today. This right over here, this is the discounting, kind of giving us the present value of the exercise price. We have videos on discounting and present value, if you find that a little bit daunting. It also makes sense that the more, the higher the stock price is, so we see that right over here, relative to the exercise price, the more that the option would be worth, it also makes sense that the higher the stock price relative to the exercise price, the more likely that we will actually exercise the option. You see that in both of these terms right over here. You have the ratio of the stock price to the exercise price. A ratio of the stock price to the exercise price. We're taking a natural log of it, but the higher this ratio is, the larger D1 or D2 is, so that means the larger the input into the cumulative distribution function is, which means the higher probabilities we're gonna get, and so it's a higher chance we're gonna exercise this price, and it makes sense that then this is actually going to have some value. So that makes sense, the relationship between the stock price and the exercise price. The other thing I will focus on, because this tends to be a deep focus of people who operate with options, is the volatility. We already had an intuition, that the higher the volatility, the higher the option price, so let's see where this factors into this equation, here. We don't see it at this first level, but it definitely factors into D1 and D2. In D1, the higher your standard deviation of your log returns, so the higher sigma, we have a sigma in the numerator and the denominator, but in the numerator, we're squaring it. So a higher sigma will make D1 go up, so sigma goes up, D1 will go up. Let's think about what's happening here. Well, here we have a sigma. It's still squared. It's in the numerator, but we're subtracting it. This is going to grow faster than this, but we're subtracting it now, so for D2, a higher sigma is going to make D2 go down because we are subtracting it. This will actually make, can we actually say this is going to make, a higher sigma's going to make the value of our call option higher. Well, let's look at it. If the value of our sigma goes up, then D1 will go up, then this input, this input goes up. If that input goes up, our cumulative distribution function of that input is going to go up, and so this term, this whole term is gonna drive this whole term up. Now, what's going to happen here. Well, if D2 goes down, then our cumulative distribution function evaluated there is going to go down, and so this whole thing is going to be lower and so we're going to have to pay less. If we get more and pay less, and I'm speaking in very hand-wavy terms, but this is just to understand that this is as intuitively daunting as you might think, but it looks definitively, that if the standard deviation, if the standard deviation of our log returns or if our volatility goes up, the value of our call option, the value of our European call option goes up. Likewise, using the same logic, if our volatility were to be lower, then the value of our call option would go down. I'll leave you there. In future videos, we'll think about this in a little bit more depth.

Financial interpretation

The equation has a concrete interpretation that is often used by practitioners and is the basis for the common derivation given in the next subsection. The equation can be rewritten in the form:

The left-hand side consists of a "time decay" term, the change in derivative value with respect to time, called theta, and a term involving the second spatial derivative gamma, the convexity of the derivative value with respect to the underlying value. The right-hand side is the riskless return from a long position in the derivative and a short position consisting of shares of the underlying asset.

Black and Scholes' insight was that the portfolio represented by the right-hand side is riskless: thus the equation says that the riskless return over any infinitesimal time interval can be expressed as the sum of theta and a term incorporating gamma. For an option, theta is typically negative, reflecting the loss in value due to having less time for exercising the option (for a European call on an underlying without dividends, it is always negative). Gamma is typically positive and so the gamma term reflects the gains in holding the option. The equation states that over any infinitesimal time interval the loss from theta and the gain from the gamma term must offset each other so that the result is a return at the riskless rate.

From the viewpoint of the option issuer, e.g. an investment bank, the gamma term is the cost of hedging the option. (Since gamma is the greatest when the spot price of the underlying is near the strike price of the option, the seller's hedging costs are the greatest in that circumstance.)

Derivation

The following derivation is given in Hull's Options, Futures, and Other Derivatives.[2]: 287–288  That, in turn, is based on the classic argument in the original Black–Scholes paper.

Per the model assumptions above, the price of the underlying asset (typically a stock) follows a geometric Brownian motion. That is

where W is a stochastic variable (Brownian motion). Note that W, and consequently its infinitesimal increment dW, represents the only source of uncertainty in the price history of the stock. Intuitively, W(t) is a process that "wiggles up and down" in such a random way that its expected change over any time interval is 0. (In addition, its variance over time T is equal to T; see Wiener process § Basic properties); a good discrete analogue for W is a simple random walk. Thus the above equation states that the infinitesimal rate of return on the stock has an expected value of μ dt and a variance of .

The payoff of an option (or any derivative contingent to stock S) at maturity is known. To find its value at an earlier time we need to know how evolves as a function of and . By Itô's lemma for two variables we have

Now consider a certain portfolio, called the delta-hedge portfolio, consisting of being short one option and long shares at time . The value of these holdings is

Over the time period , the total profit or loss from changes in the values of the holdings is (but see note below):

Now discretize the equations for dS/S and dV by replacing differentials with deltas:

and appropriately substitute them into the expression for :

Notice that the term has vanished. Thus uncertainty has been eliminated and the portfolio is effectively riskless. The rate of return on this portfolio must be equal to the rate of return on any other riskless instrument; otherwise, there would be opportunities for arbitrage. Now assuming the risk-free rate of return is we must have over the time period

If we now substitute our formulas for and we obtain:

Simplifying, we arrive at the celebrated Black–Scholes partial differential equation:

With the assumptions of the Black–Scholes model, this second order partial differential equation holds for any type of option as long as its price function is twice differentiable with respect to and once with respect to . Different pricing formulae for various options will arise from the choice of payoff function at expiry and appropriate boundary conditions.

Technical note: A subtlety obscured by the discretization approach above is that the infinitesimal change in the portfolio value was due to only the infinitesimal changes in the values of the assets being held, not changes in the positions in the assets. In other words, the portfolio was assumed to be self-financing.[citation needed]

Alternative derivation

Here is an alternative derivation that can be utilized in situations where it is initially unclear what the hedging portfolio should be. (For a reference, see 6.4 of Shreve vol II).[3]

In the Black–Scholes model, assuming we have picked the risk-neutral probability measure, the underlying stock price S(t) is assumed to evolve as a geometric Brownian motion:

Since this stochastic differential equation (SDE) shows the stock price evolution is Markovian, any derivative on this underlying is a function of time t and the stock price at the current time, S(t). Then an application of Itô's lemma gives an SDE for the discounted derivative process , which should be a martingale. In order for that to hold, the drift term must be zero, which implies the Black—Scholes PDE.

This derivation is basically an application of the Feynman–Kac formula and can be attempted whenever the underlying asset(s) evolve according to given SDE(s).

Solving methods

Once the Black–Scholes PDE, with boundary and terminal conditions, is derived for a derivative, the PDE can be solved numerically using standard methods of numerical analysis, such as a type of finite difference method.[4] In certain cases, it is possible to solve for an exact formula, such as in the case of a European call, which was done by Black and Scholes.

The solution is conceptually simple. Since in the Black–Scholes model, the underlying stock price follows a geometric Brownian motion, the distribution of , conditional on its price at time , is a log-normal distribution. Then the price of the derivative is just discounted expected payoff , which may be computed analytically when the payoff function is analytically tractable, or numerically if not.

To do this for a call option, recall the PDE above has boundary conditions [5]

The last condition gives the value of the option at the time that the option matures. Other conditions are possible as S goes to 0 or infinity. For example, common conditions utilized in other situations are to choose delta to vanish as S goes to 0 and gamma to vanish as S goes to infinity; these will give the same formula as the conditions above (in general, differing boundary conditions will give different solutions, so some financial insight should be utilized to pick suitable conditions for the situation at hand).

The solution of the PDE gives the value of the option at any earlier time, . To solve the PDE we recognize that it is a Cauchy–Euler equation which can be transformed into a diffusion equation by introducing the change-of-variable transformation

Then the Black–Scholes PDE becomes a diffusion equation

The terminal condition now becomes an initial condition

where H(x) is the Heaviside step function. The Heaviside function corresponds to enforcement of the boundary data in the S, t coordinate system that requires when t = T,

assuming both S, K > 0. With this assumption, it is equivalent to the max function over all x in the real numbers, with the exception of x = 0. The equality above between the max function and the Heaviside function is in the sense of distributions because it does not hold for x = 0. Though subtle, this is important because the Heaviside function need not be finite at x = 0, or even defined for that matter. For more on the value of the Heaviside function at x = 0, see the section "Zero Argument" in the article Heaviside step function.

Using the standard convolution method for solving a diffusion equation given an initial value function, u(x, 0), we have

which, after some manipulation, yields

where is the standard normal cumulative distribution function and

These are the same solutions (up to time translation) that were obtained by Fischer Black in 1976.[6]

Reverting to the original set of variables yields the above stated solution to the Black–Scholes equation.

The asymptotic condition can now be realized.

which gives simply S when reverting to the original coordinates.

References

  1. ^ Øksendal, Bernt (1998). "Option Pricing". Stochastic Differential Equations : An Introduction with Applications (5th ed.). Berlin: Springer. pp. 266–283. ISBN 3-540-63720-6.
  2. ^ Hull, John C. (2008). Options, Futures and Other Derivatives (7 ed.). Prentice Hall. ISBN 978-0-13-505283-9.
  3. ^ Shreve, Steven (2004). Stochastic Calculus for Finance II (1st ed.). Springer. pp. 268–272. ISBN 0-387-40101-6.
  4. ^ Wilmott, Paul; Howison, Sam; Dewynne, Jeff (1995). "Finite-difference Methods". The Mathematics of Financial Derivatives. Cambridge University Press. pp. 135–164. ISBN 0-521-49789-2.
  5. ^ Chan, Raymond (2021-07-03), Black-Scholes Equations (PDF)
  6. ^ See equation (16) in Black, Fischer S. (1976). "The Pricing of Commodity Contracts". Journal of Financial Economics. 3 (1–2): 167–179. doi:10.1016/0304-405X(76)90024-6.
This page was last edited on 24 March 2024, at 00:44
Basis of this page is in Wikipedia. Text is available under the CC BY-SA 3.0 Unported License. Non-text media are available under their specified licenses. Wikipedia® is a registered trademark of the Wikimedia Foundation, Inc. WIKI 2 is an independent company and has no affiliation with Wikimedia Foundation.