Derivation and Comparative Statics of the Black-Scholes Call and Put Option Pricing Formulas

Similar documents
Derivation and Comparative Statics of the Black-Scholes Call and Put Option Pricing Formulas

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

STOCHASTIC CALCULUS AND BLACK-SCHOLES MODEL

4. Black-Scholes Models and PDEs. Math6911 S08, HM Zhu

Queens College, CUNY, Department of Computer Science Computational Finance CSCI 365 / 765 Fall 2017 Instructor: Dr. Sateesh Mane.

Financial Risk Management

Black-Scholes Option Pricing

1.1 Basic Financial Derivatives: Forward Contracts and Options

25857 Interest Rate Modelling

Black-Scholes-Merton Model

From Discrete Time to Continuous Time Modeling

Lecture 8: The Black-Scholes theory

Pricing theory of financial derivatives

Greek parameters of nonlinear Black-Scholes equation

TEACHING NOTE 98-04: EXCHANGE OPTION PRICING

Option Pricing Models for European Options

Bluff Your Way Through Black-Scholes

WITH SKETCH ANSWERS. Postgraduate Certificate in Finance Postgraduate Certificate in Economics and Finance

The Black-Scholes Model

Practical Hedging: From Theory to Practice. OSU Financial Mathematics Seminar May 5, 2008

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

Mixing Di usion and Jump Processes

Arbitrage, Martingales, and Pricing Kernels

Lecture 3. Sergei Fedotov Introduction to Financial Mathematics. Sergei Fedotov (University of Manchester) / 6

Financial Derivatives Section 5

Hedging Credit Derivatives in Intensity Based Models

Reading: You should read Hull chapter 12 and perhaps the very first part of chapter 13.

Probability in Options Pricing

Definition Pricing Risk management Second generation barrier options. Barrier Options. Arfima Financial Solutions

Simulation Analysis of Option Buying

Lecture Quantitative Finance Spring Term 2015

Aspects of Financial Mathematics:

Option Pricing Formula for Fuzzy Financial Market

A Simple Approach to CAPM and Option Pricing. Riccardo Cesari and Carlo D Adda (University of Bologna)

Arbitrages and pricing of stock options

Credit Risk and Underlying Asset Risk *

FIN FINANCIAL INSTRUMENTS SPRING 2008

The Black-Scholes Equation using Heat Equation

Solving the Black-Scholes Equation

Course MFE/3F Practice Exam 1 Solutions

MASM006 UNIVERSITY OF EXETER SCHOOL OF ENGINEERING, COMPUTER SCIENCE AND MATHEMATICS MATHEMATICAL SCIENCES FINANCIAL MATHEMATICS.

1 Geometric Brownian motion

Lecture 11: Ito Calculus. Tuesday, October 23, 12

TEACHING NOTE 00-03: MODELING ASSET PRICES AS STOCHASTIC PROCESSES II. is non-stochastic and equal to dt. From these results we state the following:

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

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

Solving the Black-Scholes Equation

The Black-Scholes Equation

Stochastic Differential Equations in Finance and Monte Carlo Simulations

Math489/889 Stochastic Processes and Advanced Mathematical Finance Solutions to Practice Problems

Option Pricing in Continuous-Time: The Black Scholes Merton Theory and Its Extensions

Queens College, CUNY, Department of Computer Science Computational Finance CSCI 365 / 765 Fall 2017 Instructor: Dr. Sateesh Mane.

B. Combinations. 1. Synthetic Call (Put-Call Parity). 2. Writing a Covered Call. 3. Straddle, Strangle. 4. Spreads (Bull, Bear, Butterfly).

Pricing Barrier Options under Local Volatility

Homework Assignments

Finance: A Quantitative Introduction Chapter 8 Option Pricing in Continuous Time

Options. An Undergraduate Introduction to Financial Mathematics. J. Robert Buchanan. J. Robert Buchanan Options

AMH4 - ADVANCED OPTION PRICING. Contents

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

European call option with inflation-linked strike

Continuous Time Finance. Tomas Björk

Greek Maxima 1 by Michael B. Miller

Youngrok Lee and Jaesung Lee

Math 181 Lecture 15 Hedging and the Greeks (Chap. 14, Hull)

Option Pricing. Simple Arbitrage Relations. Payoffs to Call and Put Options. Black-Scholes Model. Put-Call Parity. Implied Volatility

Edgeworth Binomial Trees

Ṽ t (H) = e rt V t (H)

Degree project. Pricing American and European options under the binomial tree model and its Black-Scholes limit model

PAijpam.eu ANALYTIC SOLUTION OF A NONLINEAR BLACK-SCHOLES EQUATION

Dynamic Hedging and PDE Valuation

OPTION PRICE WHEN THE STOCK IS A SEMIMARTINGALE

Course MFE/3F Practice Exam 1 Solutions

EFFICIENT MONTE CARLO ALGORITHM FOR PRICING BARRIER OPTIONS

2 f. f t S 2. Delta measures the sensitivityof the portfolio value to changes in the price of the underlying

The Black-Scholes PDE from Scratch

Lévy models in finance

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

2.3 Mathematical Finance: Option pricing

American Option Pricing Formula for Uncertain Financial Market

25857 Interest Rate Modelling

THE BLACK-SCHOLES FORMULA AND THE GREEK PARAMETERS FOR A NONLINEAR BLACK-SCHOLES EQUATION

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

Partial differential approach for continuous models. Closed form pricing formulas for discretely monitored models

[AN INTRODUCTION TO THE BLACK-SCHOLES PDE MODEL]

Credit Risk : Firm Value Model

CHAPTER 10 OPTION PRICING - II. Derivatives and Risk Management By Rajiv Srivastava. Copyright Oxford University Press

MSC FINANCIAL ENGINEERING PRICING I, AUTUMN LECTURE 6: EXTENSIONS OF BLACK AND SCHOLES RAYMOND BRUMMELHUIS DEPARTMENT EMS BIRKBECK

Monte Carlo Simulations

Black-Scholes model: Derivation and solution

Models of Option Pricing: The Black-Scholes, Binomial and Monte Carlo Methods

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

ON AN IMPLEMENTATION OF BLACK SCHOLES MODEL FOR ESTIMATION OF CALL- AND PUT-OPTION VIA PROGRAMMING ENVIRONMENT MATHEMATICA

1 The continuous time limit

Option Pricing Model with Stepped Payoff

Forwards and Futures. Chapter Basics of forwards and futures Forwards

Hedging with Life and General Insurance Products

7.1 Volatility Simile and Defects in the Black-Scholes Model

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

Stochastic Volatility (Working Draft I)

Computational Finance. Computational Finance p. 1

Transcription:

Derivation and Comparative Statics of the Black-Scholes Call and Put Option Pricing Formulas James R. Garven Latest Revision: February 27, 2012 Abstract This paper provides an alternative derivation of the Black-Scholes call and put option pricing formulas using an integration rather than differential equations approach. The economic and mathematical structure of these formulas is discussed, and comparative statics are derived. 1 Introduction The purposes of this paper are: (1) to provide a concise overview of the economic and mathematical assumptions needed to derive the Black-Scholes (1973) call and put option pricing formulas; (2) to provide an alternative derivation and comparative statics analysis of these formulas. This paper is organized as follows. In the next section, we discuss the economic and mathematical structure of the Black-Scholes model. In the third section, we provide an alternative derivation of the Black-Scholes option pricing formulas based upon an integration rather than differential equations approach. The advantage of the integration approach lies in its simplicity, as only basic integral calculus is required. The integration approach also makes the link between the economics and mathematics of the call and put option formulas more transparent for most readers. This paper concludes with comparative statics analyses of these formulas. James R. Garven is the Frank S. Groner Memorial Chair in Finance and Professor of Finance & Insurance at Baylor University (Address: HSB 351, One Bear Place #98004, Waco, TX 76798, telephone: 254-307- 1317, e-mail: James Garven@baylor.edu). This article has benefited from the helpful comments of Soku Byoun, Richard Derrig, Gabriel Drimus, Bill Reichenstein, and J. T. Rose. Of course, the usual disclaimer applies. 1

2 Economic and Mathematical Structure of the Black-Scholes Model The most important economic insight contained in Black and Scholes seminal paper is the notion that an investor can create a riskless portfolio by dynamically hedging a long (short) position in the underlying asset with a short (long) position in a European call option. In order to prevent arbitrage, the expected return on such a portfolio must be the riskless rate of interest. Cox and Ross (1976) note that since the creation of such a hedge portfolio places no restrictions on investor preferences beyond nonsatiation, the valuation relationship between an option and its underlying asset is risk neutral in the sense that it does not depend upon investor risk preferences. Therefore, for a given price of the underlying asset, a call option written against that asset will trade for the same price in a risk neutral economy as it would in a risk averse or risk loving economy. Consequently, options can be priced as if they are traded in risk neutral economies. Black and Scholes assume that stock prices change continuously according to the Geometric Brownian Motion equation; i.e., ds = µsdt + σsdz. (1) Equation (1) is a stochastic differential equation because it contains the Wiener process dz = ε dt, where ɛ is a standard normal random variable. ds corresponds to the stock price change per dt time unit, S is the current stock price, µ is the expected return, and σ represents volatility. The Geometric Brownian Motion equation is commonly referred to as an exponential stochastic differential equation because its solution is an exponential function; specifically, S t = Se (µ σ2 /2)t+σz t, where S t represents the stock price t periods from now. Since z t is normally distributed, it follows that S t is lognormally distributed. S t s mean is µ St = Se µt, and its variance is σs 2 t = S 2 e 2µt (e σ2t 1). Ito s Lemma provides a rule for finding the differential of a function of one or more variables in which at least one of the variables follows a stochastic differential equation. At 2

any given point in time, the value of the call option (C ) depends upon the value of the underlying asset; i.e., C = C (S,t). 1 Since the value of the call option is a function of the stock price which evolves over time according to equation (1), Ito s Lemma justifies the use of a Taylor-series-like expansion for the differential dc : dc = C C dt + t S ds + 1 2 C 2 S 2 ds2. (2) Since ds 2 = S 2 σ 2 dt, substituting for ds 2 in equation (2) yields equation (3): dc = C C dt + t S ds + 1 2 σ2 S 2 2 C dt. (3) S2 Suppose we wish to construct a hedge portfolio consisting of one long call option position worth C (S, t) and a short position in some quantity t of the underlying asset worth S per share. We express the hedge ratio t as a function of t because the portfolio will be dynamically hedged; i.e., as the price of the underlying asset changes through time, so will t. Then the value of this hedge portfolio is V = C (S, t) - t S, which implies dv = dc t ds = ( C t + 1 ) ( ) 2 σ2 S 2 2 C C S 2 dt + S t ds. (4) Note that there are stochastic as well as deterministic terms on the right-hand side of equation (4). The deterministic terms are represented by the first product, whereas the stochastic terms are represented by the second product. However, by setting t equal to C, the second product equals zero and we have: S dv = dc t ds = ( C t + 1 ) 2 σ2 S 2 2 C S 2 dt. (5) Since this is a perfectly hedged portfolio, it has no risk. In order to prevent arbitrage, the 1 The call option value also depends upon the exercise price of the option (X ), the interest rate (r), and the volatility of the underlying asset (σ 2 S). However, since these parameters are constants, the differential equation for C only depends upon S and t. Since t is a deterministic time trend whereas instantaneous rates of return on S are assumed to be normally distributed, it follows that C is once differentiable in C and twice differentiable in S. 3

hedge portfolio must earn the riskless rate of interest r; i.e., dv = rv dt. (6) We will assume that t = C C, so V = C (S, t) - S. Substituting this into the right-hand S S side of equation (6) and equating the result with the right-hand side of equation (5), we obtain: ( r C S C ) ( C dt = S t + 1 ) 2 σ2 S 2 2 C S 2 dt. (7) Dividing both sides of equation (7) by dt and rearranging results in the Black-Scholes (nonstochastic) partial differential equation 2 rc = C t + 1 2 σ2 S 2 2 C C + rs S2 S. (8) Equation (8) shows that the valuation relationship between a call option and its underlying asset is deterministic because dynamic hedging enables the investor to be perfectly hedged over infinitesimally small units of time. Since risk preferences play no role in this equation, this implies that the price of a call option can be calculated as if investors are risk neutral. Suppose there are τ = T t periods until option expiration occurs, where T is the (fixed) expiration date and t is today. Then the value of a call option (C ) at time t must satisfy equation (8), subject to the boundary condition on the option s payoff at expiration; i.e., C T = Max[S T X, 0], where X represents the option s exercise price and S T represents the stock price at expiration. Black and Scholes transform their version of equation (8) into the heat transfer equation of physics, 3 which allows them to (quite literally) employ a textbook solution procedure 2 Note that equation (8) in this paper and equation (7) in the original Black-Scholes paper (cf. Black and Scholes (1973), available on the web at http://www.jstor.org/stable/1831029) are the same equation (although the notation is marginally different). 3 Wilmott (2001, p. 156) notes that heat transfer equations date back to the beginning of the 19th century, and have been used to model a diverse set of physical and social phenomena, such as the flow of heat from one part of an object to another, electrical activity in the membranes of living organisms, the dispersion of populations, the formation of zebra stripes, and the dispersion of air and water pollution, 4

taken from Churchill s (1963) classic introduction to Fourier series and their applications to boundary value problems in partial differential equations of engineering and physics. This solution procedure results in the following equation for the value of a European call option: C = SN(d 1 ) Xe rt N(d 2 ), (9) where d 1 = ln(s/x) + (r +.5σ2 )t σ ; t d 2 = d 1 σ t; σ 2 = variance of underlying asset s rate of return; and N (z) = standard normal distribution function evaluated at z. Equation (9) implies that the price of the call option depends upon five parameters; specifically, the current stock price S, the exercise price X, the riskless interest rate r, the time to expiration t, and the volatility of the underlying asset σ. If it is not possible to form a riskless hedge portfolio, then other more restrictive assumptions are needed in order to obtain a risk neutral valuation relationship between an option and its underlying asset. Rubinstein (1976) has shown that the Black-Scholes option pricing formulas also obtain under the following set of assumptions: 1. The conditions for aggregation are met so that securities are priced as if all investors have the same characteristics as a representative investor; 2. The utility function of the representative investor exhibits constant relative risk aversion (CRRA); 3. The return on the underlying asset and the return on aggregate wealth are bivariate lognormally distributed; 4. The return on the underlying asset follows a stationary random walk through time; and 5. The riskless rate of interest is constant through time. Collectively, assumptions 1, 2, and 3 constitute sufficient conditions for valuing the option as if both it and the underlying asset are traded in a risk neutral economy. Assumptions among other things. 5

4 and 5 constitute necessary conditions for the Black-Scholes option pricing formulas, since they guarantee that the riskless rate of interest, the rate of return on the underlying asset, and the variance of the underlying asset will be proportional to time. 4 3 Black-Scholes Option Pricing Formula Derivations Rubinstein (1987) notes that the Black-Scholes call option pricing formula can be derived by integration if one assumes that the probability distribution for the underlying asset is lognormal. Fortunately, as noted earlier, this assumption is implied by the Geometric Brownian Motion equation given by equation (1). Therefore, our next task is to derive the Black-Scholes call option pricing formula by integration. Once we obtain the Black-Scholes call option pricing formula, the put option pricing formula follows directly from the put-call parity theorem. The value today (C ) of a European call option that pays C t = Max[S t X, 0] at date t is given by the following equation: C = V (C t ) = V (Max[S t X, 0]), (10) where V represents the valuation operator. Having established in equation (8) that a risk neutral valuation relationship exists between the call option and its underlying asset, it follows that we can price the option as if investors are risk neutral. Therefore, the valuation operator V determines the current price of a call option by discounting the risk neutral expected value of the option s payoff at expiration (Ê (C t)) at the riskless rate of interest, as indicated in equation (11): C = e rt Ê (C t ) = e rt (S t X)ĥ(S t)ds t, (11) X 4 Brennan (1979) shows that if the return on the underlying asset and the return on aggregate wealth are bivariate lognormally (normally) distributed, then the representative investor must have CRRA (CARA) preferences. Stapleton and Subrahmanyam (1984) extend Rubinstein s and Brennan s results to a world where payoffs on options depend on the outcomes of two or more stochastic variables. 6

where ĥ(s t) represents the risk neutral lognormal density function of S T. We begin our analysis by computing the expected value of C t : E(C t ) = E[Max(S t X, 0)] = X (S t X)h(S t )ds t, (12) where h(s t ) represents S t s lognormal density function. 5 To evaluate this integral, we rewrite it as the difference between two integrals: E(C t ) = X S t h(s t )ds t X X h(s t )ds t = E X (S t ) Xe rt [1 H(X)], (13) where E X (S t ) = the partial expected value of S T, truncated from below at X ; 6 H(X) = the probability that S t X. By rewriting the terminal stock price S T as the product of the current stock price S and the t-period lognormally distributed price ratio S t /S, S t = S (S t /S), equation (13) can be rewritten as E(C t ) = S (S t /S)g(S t /S)dS t /S X g(s t /S)dS t /S X/S X/S = SE X/S (S t /S) X[1 G(X/S)], (14) where g(s t /S) = lognormal density function of S T /S; 7 5 The difference between the ĥ(st) and h(st) density functions is easily explained. Since ST is a lognormally distributed random variable, E(S t) = 0 S th(s t)ds t =Se (µ+.5σ2 )t. Since all assets in a risk neutral economy are expected to return the riskless rate of interest, this implies that in a risk neutral economy, Ê(S t) = Se rt, which in turn implies that (µ +.5σ 2 )t = rt. The ĥ(st) density function is risk neutral in the sense that its location parameter µt is replaced by (r.5σ 2 )t. 6 The partial expected value of X for values of X ranging from y to z is Ey(X) z = z Xf(X)dX. If we y replace the y and z limits of integration with - and, then we have a complete expected value, or first moment. See Winkler, Roodman and Britney (1972) for more details concerning the computation of partial moments for a wide variety of probability distributions, including the normal distribution. 7 Note that g(s τ /S) = h(s T )/S. 7

and E X/S (S t /S) = the partial expected value of S T /S, truncated from below at X /S; G(X/S) = the probability that S t /S X /S. Next, we evaluate the right-hand side of equation (14) by considering its two integrals separately. First, consider the integral that corresponds to the partial expected value of the terminal stock price, SE X/S (S t /S). Let S t /S = e kt, where k is the rate of return on the underlying asset per unit of time. Since S t /S is lognormally distributed, ln (S t /S) = kt is normally distributed with density f (kt), mean µ k t and variance σk 2t.8 Furthermore, g(s τ /S) = (S/S t ) f (kt). 9 Also, since S t /S = e kt, ds t /S = e kt tdk. Consequently, (S t /S)g(S t /S)dS t /S = e kt f(kt)tdk. Thus, SE X/S (S t /S) = S ln(x/s) e kt 1 f(kt)tdk = S 2πσ 2 t ln(x/s) e kt e {.5[(kt µt)2 /σ 2 t]} tdk. (15) Next, we simplify equation (15) s integrand by adding the terms in the two exponents, multiplying and dividing the resulting expression by e.5σ2t, and rearranging: e kt e {.5[(kt µt)2 /σ 2 t]} = e {.5t[(k2 2µk+µ 2 2σ 2 k)/σ 2 ]} = e {.5t[(k2 2µk+µ 2 2σ 2 k+σ 4 σ 4 )/σ 2 ]} = e {.5t[((k µ σ2 ) 2 σ 4 2µσ 2 )/σ 2 ]} = e (µ+.5σ2 )t e {.5[(kt (µ+σ2 )t) 2 /σ 2 t]}. (16) In equation (16), the term e (µ+.5σ2 )t = E(S T /S), the mean of the t-period lognormally distributed price ratio S T /S. Therefore, we can rewrite equation (15) with this term appearing 8 For the sake of simplifying notation, the k subscripts are dropped in the subsequent analysis. 9 The relationship between f (kt) and g(s t/s) is apparent from differentiating the cumulative distribution function F (kt) with respect to S t/s : df (kt)/d(s t/s) = (df (kt)/d(kt))(dkt/d(s t/s)) = f(kt)(s/s t). 8

outside the integral: 1 SE X/S (S t /S) = SE(S t /S) 2πσ 2 t 1 = E(S t ) 2πσ 2 t ln(x/s) ln(x/s) e {.5[(kt (µ+σ2 )t) 2 /σ 2 t]} tdk e {.5[(kt (µ+σ2 )t) 2 /σ 2 t]} tdk. (17) Next, we let y = [kt (µ + σ 2 )t]/σ t, which implies that kt = (µ + σ 2 )t + σ ty and tdk = σ tdy. With this change in variables, [ln(x /S) - (µ + σ 2 )t]/σ t= δ 1 becomes the lower limit of integration. Thus equation (17) may be rewritten as: SE X/S (S t /S) = E(S t ) = E(S t ) [e.5y2 δ 1 δ1 n(y)dy / 2π]dy = E(S t )N(δ 1 ), (18) where n(y) is the standard normal density function evaluated at y and N(δ 1 ) is the standard normal distribution function evaluated at y = δ 1. In order to complete our computation of the expected value of C t, we must evaluate the integral X X/S g(s t /S)dS t /S. As noted earlier, since S t /S = e kt is lognormally distributed, ln (S t /S) = kt is normally distributed with density f (kt), mean µt and variance σ 2 t. Furthermore, this change in variable implies that f(kt)tdk must be substituted in place of g(s t /S)dS t /S. Therefore, X X/S g(s t /S)dS t /S = X f(kt)tdk ln(x/s) = X 1 2πσ 2 t ln(x/s) e {.5[(kt µt)2 /σ 2 t]} tdk. (19) Next, we let z = [kt µt]/σ t, which implies that kt = µt + σ tz and tdk = σ tdz. With this change in variables, the lower limit of integration becomes [ln(x/s) µt]/σ t = -(δ 1 - σ t) = δ 2. Substituting these results into the right-hand side of equation (19) 9

yields equation (20): X X/S g(s t /S)dS t /S = X = X [e.5z2 δ 2 δ2 n(z)dz / 2π]dz = XN(δ 2 ), (20) where n(z) is the standard normal density function evaluated at z and N(δ 2 ) is the standard normal distribution function evaluated at z = δ 2. Substituting the right-hand sides of equations (18) and (20) into the right-hand side of equation (14), we obtain E(C t ) = E(S t )N(δ 1 ) XN(δ 1 σ t). (21) Since equation (11) defines the price of a call option as the discounted, risk neutral expected value of the option s payoff at expiration; i.e., C = e rt Ê (C t ), our next task is to compute the risk neutral expected value of the stock price on the expiration date (Ê(S t)) and the risk neutral value for δ 1 (which we will refer to as d 1 ). In footnote (3), we noted that (µ +.5σ 2 )t = rt in a risk neutral economy. Since E(S t ) = Se (µ+.5σ2 )t, this implies that Ê(S t) = Se rt. In the expression for δ 1, the term (µ + σ 2 )t appears. In a risk neutral economy, (µ + σ 2 )t = (r +.5σ 2 )t; therefore,d 1 = [ln(s/x) +(r +.5σ 2 )t]/σ t. Substituting the right-hand side of equation (21) into the right-hand side of equation (11) and simplifying yields the Black-Scholes call option pricing formula: [ C = e rt Ê (C t ) = e rt Se rt N(d 1 ) XN(d 1 σ ] t) = SN(d 1 ) Xe rt N(d 2 ). (22) Now that we have the Black-Scholes pricing formula for a European call option, the put option pricing formula follows directly from the put-call parity theorem. Suppose two portfolios exist, one consisting of a European call option and a riskless discount bond, 10

and the other consisting of a European put option and a share of stock against which both options are written. The call and put both have exercise price X and t periods to expiration, and the riskless bond pays off X dollars at date t. Then these portfolios date t payoffs are identical, since the payoff on the first portfolio is Max(S t X, 0) + X = Max(S t, X) and the payoff on the second portfolio is Max(X S t, 0) + S T = Max(S t, X). Consequently, the current value of these portfolios must also be the same; otherwise there would be a riskless arbitrage opportunity. Therefore, the price of a European put option, P, can be determined as follows: P = C + Xe rt S = SN(d 1 ) Xe rt N(d 2 ) + Xe rt S = Xe rt [1 N(d 2 )] S [1 N(d 1 )] = Xe rt N( d 2 ) SN( d 1 ). (23) 4 Comparative Statics As indicated by equations (22) and (23), the prices of European call and put options depend upon five parameters: S, X, t, r, and σ. Next, we provide comparative statics analysis of call and put option prices. 4.1 Relationship between the call and put option prices and the price of the underlying asset Consider the relationship between the price of the call option and the price of the underlying asset, C/ S: C/ S = N(d 1 ) + S( N(d 1 )/ d 1 )( d 1 / S) Xe rt ( N(d 2 )/ d 2 )( d 2 / S) = N(d 1 ) + Sn(d 1 )( d 1 / S) Xe rt n(d 2 )( d 2 / S). (24) Substituting d 2 = d 1 -σ t, d 2 / S = d 1 / S, and n(d 2 ) = n(d 1 -σ t) into equation (24), 11

we obtain: C/ S = N(d 1 ) + ( d 1 / S)[Sn(d 1 ) Xe rt n(d 1 σ t)] 1 = N(d 1 ) + ( d 1 / S) [Se.5d2 1 Xe rt e.5(d 1 σ t) 2 ]. (25) 2π Recall that d 1 = [ln(s/x)+(r+.5σ 2 )t]/σ t. Solving for S, we find that S = Xe d 1σ t (r+.5σ 2 )t. Substituting this expression for S into the bracketed term on the right-hand side of equation (25) s yields C/ S = 1 N(d 1 ) + ( d 1 / S) [Xe.5d2 1 e d 1 σ t (r+.5σ 2 )t Xe rt e.5(d 1 σ t) 2 ] 2π = 1 N(d 1 ) + ( d 1 / S) [X(e (r+.5σ2 )t+d 1 σ t.5d 2 1 e (r+.5σ 2 )t+d 1 σ t.5d 2 1 )]. (26) 2π Since the bracketed term on the right-hand side of equation (26) equals zero, C/ S = N(d 1 ) > 0; i.e., the price of a call option is positively related to the price of its underlying asset. Furthermore, from our earlier analysis of the Black-Scholes partial differential equation, we know that a dynamically hedged portfolio consisting of one long call position per t = C/ S short shares of the underlying asset is riskless (cf. equation (5)). Since C/ S = N (d 1 ), N (d 1 ) is commonly referred to as the option s delta, and the dynamic hedging strategy described here is commonly referred to as delta hedging. Intuitively, since the put option provides its holder with the right to sell rather than buy, one would expect an inverse relationship between the price of a put option and the price of its underlying asset. Next, we analyze this relationship by finding P/ S: P/ S = N( d 1 ) + Xe rt ( N( d 2 )/ d 2 )( d 2 / S) S( N( d 1 )/ d 1 )( d 1 / S) = N( d 1 ) + Xe rt n( d 2 )( d 1 / S) Sn( d 1 )( d 1 / S) = N( d 1 ) + ( d 1 / S)[Xe rt n(σ t d 1 ) Sn( d 1 )] (27) Given the symmetry of the standard normal distribution about its mean of zero, Sn( d 1 ) = 12

Sn(d 1 ) and Xe rt n( d 2 ) = Xe rt n(d 2 ). Since we know from equation (26) that Sn(d 1 ) Xe rt n(d 1 σ t) = 0, it follows that the bracketed term on the right-hand side of equation (27), Xe rt n(σ t d 1 ) Sn( d 1 ) = 0. Therefore, P/ S = N( d 1 ) < 0. Thus, our intuition is confirmed; i.e., the price of a put option is inversely related to the price of its underlying asset. 4.2 Relationship between call and put option prices and the exercise price Consider next the relationship between the price of the call option and the exercise price, C/ X : C/ X = e rt N(d 2 ) + S( N(d 1 )/ d 1 )( d 1 / X) Xe rt ( N(d 2 )/ d 2 )( d 2 / X) = e rt N(d 2 ) + Sn(d 1 )( d 1 / X) Xe rt n(d 2 )( d 2 / X) (28) Substituting d 2 = d 1 -σ t, d 2 / X = d 1 / X and n(d 2 ) = n(d 1 -σ t) into equation (28) gives us C/ X = e rt N(d 2 ) + ( d 1 / X)[Sn(d 1 ) Xe rt n(d 1 σ t)] = e rt N(d 2 ) < 0. (29) Intuition suggests that since the price of a call option is inversely related to its exercise price, one would expect to find that a positive relationship exists between the price of a put option and its exercise price, which we confirm in equation (30): P/ X = e rt N( d 2 ) + Xe rt ( N( d 2 )/ d 2 )( d 2 / X) S( N( d 1 )/ d 1 )( d 1 / X) = e rt N( d 2 ) + Xe rt n( d 2 )( d 1 / X) Sn( d 1 )( d 1 / X) = e rt N( d 2 ) + ( d 1 / X)[Xe rt n(σ t d 1 ) Sn( d 1 )] = e rt N( d 2 ) > 0. (30) 13

4.3 Relationship between call and put option prices and the interest rate Consider the relationship between the price of the call option and the rate of interest, C/ r: C/ r = txe rt N(d 2 ) + S( N(d 1 )/ d 1 )( d 1 / r) Xe rt ( N(d 2 )/ d 2 )( d 2 / r) = txe rt N(d 2 ) + Sn(d 1 )( d 1 / r) Xe rt n(d 2 )( d 2 / r) = txe rt N(d 2 ) + ( d 1 / r)[sn(d 1 ) Xe rt n(d 1 σ t)] = txe rt N(d 2 ) > 0. (31) Since the right-hand side of equation (31) is positive, this implies that the price of a call option is positively related to the rate of interest. This comparative static relationship is commonly referred to as the option s rho. Intuition suggests that since the price of a call option is positively related to the rate of interest, one would expect to find that an inverse relationship exists between the price of a put option and the rate of interest, which we confirm in equation (32): P/ r = txe rt N( d 2 ) + Xe rt ( N( d 2 )/ d 2 )( d 2 / r) + S( N( d 1 )/ d 1 )( d 1 / r) = txe rt N( d 2 ) + Xe rt n( d 2 )( d 2 / r) Sn( d 1 )( d 1 / r) = txe rt N( d 2 ) + ( d 1 / r)[xe rt n(σ t d 1 ) Sn( d 1 )] = txe rt N( d 2 ) < 0. (32) 14

4.4 Relationship between call and put option prices and the time to expiration Consider next the relationship between the price of the call option and the time to expiration, C/ t: C/ t = rxe rt N(d 2 ) + S( N(d 1 )/ d 1 )( d 1 / t) Xe rt ( N(d 2 )/ d 2 )( d 2 / t). (33) Substituting X = Se d 1σ t+(r+.5σ 2 )t into equation (33) and simplifying further, we obtain C/ t = rxe rt N(d 2 ) + S[n(d 1 )/ d 1 )( d 1 / t) e d 1σ t+(r+.5σ 2 )t rt n(d 1 σ t)( d 2 / t) = rxe rt N(d 2 ) + S 1 2π [( d 1 / t)e.5d2 1 ( d2 / t)e d 1σ t+(r+.5σ 2 )t rt.5(d 1 σ t) 2 ] = rxe rt N(d 2 ) + S 1 2π [( d 1 / t)e.5d2 1 ( d2 / t)e d 1σ t+d 1 σ t+rt rt+.5σ 2 t.5σ 2 t.5d 2 1 ] = rxe rt N(d 2 ) + Sn(d 1 )[( d 1 / t) ( d 2 / t)] = rxe rt N(d 2 ) + Sn(d 1 ).5σ t. (34) Since both terms on the right-hand side of equation (34) are positive, this implies that the price of a call option is positively related to the time to expiration; i.e., longer lived European call options are worth more than otherwise identical shorter lived European call options. Next, we consider the relationship between the price of a put option and the time to expiration: P/ t = rxe rt N( d 2 ) + Xe rt ( N( d 2 )/ d 2 )( d 2 / t) S( N( d 1 )/ d 1 )( d 1 / t). (35) 15

Substituting X = Se d 1σ t+(r+.5σ 2 )t into equation (35) and simplifying further, we obtain P/ t = rxe rt N( d 2 ) + Se d 1σ t+(r+.5σ 2 )t rt n(σ t d 1 )( d 2 / t) Sn(d 1 )( d 1 / t) = rxe rt N( d 2 ) + S 2π e d 1σ t+(r+.5σ 2 )t rt.5(σ t d 1 ) 2 ( d 2 / t) Sn( d 1 )( d 1 / t) = rxe rt N( d 2 ) + S 2π [( d 2 / t)e d 1σ t+d 1 σ t+rt rt+.5σ 2 t.5σ 2 t.5d 2 1 ( d1 / t)e.5d2 1 ] = rxe rt N( d 2 ) + Sn(d 1 )[( d 2 / t) ( d 1 / t)] = rxe rt N( d 2 ) + Sn(d 1 ).5σ t. (36) It is not possible to sign P/ t, because the first term is negative and the second term is positive. Therefore, the relationship between put value and time to expiration is ambiguous. 10 4.5 Relationship between call and put option prices and volatility of the underlying asset Finally, consider the relationship between the price of the call option and the volatility of the underlying asset, C/ σ: C/ σ = S N(d 1) d 1 d 1 σ Xe rt N(d 2) d 2 d 2 σ. (37) 10 The theta for an option represents the rate at which its value changes with the passage of time, given a fixed time to expiration. Therefore, the first term on the left-hand side of equation (8) corresponds to the theta for a call option which expires in T periods, where T represents a fixed unit of time. Thus theta for a call option, Θ C, is equal to 1 C/ t, where C/ t corresponds to the right-hand side of equation (34); i.e., Θ C = rxe rt N(d 2) Sn(d 1).5σ t < 0. Since the theta of a call option is negative, this implies that with the passage of time, call options become less valuable. However, since, Θ P = rxe rt N( d 2) Sn(d 1).5σ k t, this implies that the theta for a put option may be positive or negative depending upon the relative magnitudes of the two terms with opposing signs. 16

Substituting N(d 2) d 2 = n(d 2 ) = n(d 1 σ k t), d 2 σ = d 1 σ t, and X = Se d 1σ t+(r+.5σ 2 )t into the right-hand side of equation (37) and simplifying further, we obtain [ C/ σ = S n(d 1 ) d 1 = S ] σ e d 1σ t+(r+.5σ 2 )t rt d 2 n(d 1 σ k t) σ t+rt+.5σ 2 t rt.5(d 1 σ t) 2 [ e.5d2 1 2π d 1 σ e d1σ 2π ( d1 σ t ) ]. (38) Since e d 1σ t+rt+.5σ 2 t rt.5(d 1 σ t) 2 = e.5d2 1 and n(d1 ) = e.5d2 1 2π, equation (38) can be rewritten as [ ( d1 C/ σ = Sn(d 1 ) σ d1 σ )] t = Sn(d 1 ) t > 0. (39) As equation (39) indicates, the price of a call option is positively related to the volatility of the underlying asset. This comparative static relationship is commonly referred to as the option s vega. Next, we consider the relationship between the price of a put option and the volatility of the underlying asset, P/ σ: P/ σ = Xe rt N( d 2) d 2 d 2 σ S N( d 1) d 1 d 1 σ = Xe rt n( d 2 ) d 2 σ + Sn( d 1) d 1 σ. (40) Substituting d 2 = σ t d 1, d 2 σ = d 1 σ t, and X = Se d 1σ t+(r+.5σ 2 )t into the right hand side of equation (40) and simplifying further yields: P/ σ = S[n( d 1 ) d t+rt+.5σ 1 σ e d1σ 2 t rt.5(σ t d 1) 2 ( d 1 2π σ t)] = Sn( d 1 )[ d 1 σ ( d 1 σ t)] = Sn( d 1 ) t > 0. (41) 17

As equation (41) indicates, the price of a put option is positively related to the volatility of the underlying asset. Table 1 summarizes the comparative statics of the Black-Scholes call and put option pricing formulas. 18

Table 1. Comparative Statics of the Black-Scholes Model Derivative Call Option Put Option C P and C S S S = N (d 1) > 0 (delta) C P C and X X X = -e rt N (d 2 ) < 0 C P and C r r r = txe rt N (d 2 ) > 0 (rho) C P C and t t t =rxe rt N(d 2 ) + Sn(d 1 ).5σ t > 0 theta C P and σ σ (vega) - C t = rxe rt N(d 2 ) Sn(d 1 ).5σ t < 0 C σ = Sn(d 1) t > 0 P S = -N (-d 1) < 0 P X = e rt N (-d 2 ) > 0 P r = -txe rt N (-d 2 ) < 0 P t = rxe rt N( d 2 )+Sn(d 1 ).5σ? k t <> 0 - P t =rxe rt N( d 2 ) Sn(d 1 ).5σ? k t <> 0 P σ = Sn( d 1) t > 0 19

5 References Black, F., and M. Scholes (1973), The Pricing of Options and Corporate Liabilities, Journal of Political Economy 81:637-659. Brennan M. J., 1979, The Pricing of Contingent Claims in Discrete Time Models, Journal of Finance, Vol. 34, No. 1 (March), pp. 53-68. Churchill, R. V. (1963), Fourier Series and Boundary Value Problems, 2 nd edition, New York: McGraw-Hill. Cox, J. C., and S. A. Ross (1976), The Valuation of Options for Alternative Stochastic Processes, Journal of Financial Economics 3:145-166. Rubinstein, M. (1976), The Valuation of Uncertain Income Streams and the Pricing of Options, Bell Journal of Economics 7:407-425. Rubinstein, M. (1987), Derivative Assets Analysis, The Journal of Economic Perspectives, 1(Autumn): 73-93. Stapleton, R. C. and M. G. Subrahmanyam, 1984, The Valuation of Multivariate Contingent Claims in Discrete Time Models, Journal of Finance, Vol. 39, No. 1 (March), pp. 207-228. Wilmott, P. (2001), Paul Wilmott Introduces Quantitative Finance, Chichester, England: John Wiley & Sons Ltd. Winkler, R. L., G. M. Roodman and R. R. Britney (1972), The Determination of Partial Moments, Management Science 19(3):290-296. 20