How To Price An American Option

Similar documents
Adaptive Online Gradient Descent

Mathematical Finance

Factorization Theorems

Monte Carlo Methods in Finance

Pricing and hedging American-style options: a simple simulation-based approach

LOGNORMAL MODEL FOR STOCK PRICES

TD(0) Leads to Better Policies than Approximate Value Iteration

LECTURE 15: AMERICAN OPTIONS

Continued Fractions and the Euclidean Algorithm

CHAPTER II THE LIMIT OF A SEQUENCE OF NUMBERS DEFINITION OF THE NUMBER e.

1 Short Introduction to Time Series

Variance Reduction. Pricing American Options. Monte Carlo Option Pricing. Delta and Common Random Numbers

Lecture 3: Finding integer solutions to systems of linear equations

Numerical Analysis Lecture Notes

Algebra Unpacked Content For the new Common Core standards that will be effective in all North Carolina schools in the school year.

Taylor and Maclaurin Series

The Characteristic Polynomial

We shall turn our attention to solving linear systems of equations. Ax = b

Lecture 13: Martingales

1 The Brownian bridge construction

LINEAR ALGEBRA W W L CHEN

FACTORING POLYNOMIALS IN THE RING OF FORMAL POWER SERIES OVER Z

ECON20310 LECTURE SYNOPSIS REAL BUSINESS CYCLE

9.2 Summation Notation

Metric Spaces. Chapter Metrics

7 Gaussian Elimination and LU Factorization

4: SINGLE-PERIOD MARKET MODELS

2.1 Complexity Classes

Ideal Class Group and Units

Generating Random Numbers Variance Reduction Quasi-Monte Carlo. Simulation Methods. Leonid Kogan. MIT, Sloan , Fall 2010

Optimization under uncertainty: modeling and solution methods

Invariant Option Pricing & Minimax Duality of American and Bermudan Options

The Exponential Distribution

MATHEMATICAL METHODS OF STATISTICS

Overview of Violations of the Basic Assumptions in the Classical Normal Linear Regression Model

Solution of Linear Systems

Adaptive Search with Stochastic Acceptance Probabilities for Global Optimization

x1 x 2 x 3 y 1 y 2 y 3 x 1 y 2 x 2 y 1 0.

Recall that two vectors in are perpendicular or orthogonal provided that their dot

COMPLETE MARKETS DO NOT ALLOW FREE CASH FLOW STREAMS

U.C. Berkeley CS276: Cryptography Handout 0.1 Luca Trevisan January, Notes on Algebra

Notes on Factoring. MA 206 Kurt Bryan

Chapter 2: Binomial Methods and the Black-Scholes Formula

Modern Optimization Methods for Big Data Problems MATH11146 The University of Edinburgh

1 Sets and Set Notation.

Some Polynomial Theorems. John Kennedy Mathematics Department Santa Monica College 1900 Pico Blvd. Santa Monica, CA

3. INNER PRODUCT SPACES

CONTINUED FRACTIONS AND FACTORING. Niels Lauritzen

(Quasi-)Newton methods

Numeraire-invariant option pricing

The Monte Carlo Framework, Examples from Finance and Generating Correlated Random Variables

ELLIPTIC CURVES AND LENSTRA S FACTORIZATION ALGORITHM

Mathematics Course 111: Algebra I Part IV: Vector Spaces

CONTROLLABILITY. Chapter Reachable Set and Controllability. Suppose we have a linear system described by the state equation

THE SCHEDULING OF MAINTENANCE SERVICE

Continuity of the Perron Root

Orthogonal Projections

Master s Theory Exam Spring 2006

Thnkwell s Homeschool Precalculus Course Lesson Plan: 36 weeks

Overview of Monte Carlo Simulation, Probability Review and Introduction to Matlab

How To Prove The Dirichlet Unit Theorem

Valuation of American Options

How To Calculate Max Likelihood For A Software Reliability Model

Stochastic Inventory Control

The van Hoeij Algorithm for Factoring Polynomials

Master of Mathematical Finance: Course Descriptions

Mathematics (MAT) MAT 061 Basic Euclidean Geometry 3 Hours. MAT 051 Pre-Algebra 4 Hours

General Framework for an Iterative Solution of Ax b. Jacobi s Method

Direct Methods for Solving Linear Systems. Matrix Factorization

Faster deterministic integer factorisation

A Negative Result Concerning Explicit Matrices With The Restricted Isometry Property

The Variability of P-Values. Summary

Data Mining: Algorithms and Applications Matrix Math Review

Nonparametric adaptive age replacement with a one-cycle criterion

EC3070 FINANCIAL DERIVATIVES

Linear Algebra Notes for Marsden and Tromba Vector Calculus

Chapter 6. Cuboids. and. vol(conv(p ))

The Discrete Binomial Model for Option Pricing

NEW YORK STATE TEACHER CERTIFICATION EXAMINATIONS

Pricing Discrete Barrier Options

HOMEWORK 5 SOLUTIONS. n!f n (1) lim. ln x n! + xn x. 1 = G n 1 (x). (2) k + 1 n. (n 1)!

BookTOC.txt. 1. Functions, Graphs, and Models. Algebra Toolbox. Sets. The Real Numbers. Inequalities and Intervals on the Real Number Line

South Carolina College- and Career-Ready (SCCCR) Algebra 1

α = u v. In other words, Orthogonal Projection

A linear algebraic method for pricing temporary life annuities

SECTION 10-2 Mathematical Induction

Notes 11: List Decoding Folded Reed-Solomon Codes

Lectures 5-6: Taylor Series

PUTNAM TRAINING POLYNOMIALS. Exercises 1. Find a polynomial with integral coefficients whose zeros include

Simulating Stochastic Differential Equations

Modeling and Performance Evaluation of Computer Systems Security Operation 1

Finite Differences Schemes for Pricing of European and American Options

NOTES ON LINEAR TRANSFORMATIONS

Using simulation to calculate the NPV of a project

Section 1.1. Introduction to R n

CS 522 Computational Tools and Methods in Finance Robert Jarrow Lecture 1: Equity Options

South Carolina College- and Career-Ready (SCCCR) Pre-Calculus

MATH. ALGEBRA I HONORS 9 th Grade ALGEBRA I HONORS

Some facts about polynomials modulo m (Full proof of the Fingerprinting Theorem)

Universal Algorithm for Trading in Stock Market Based on the Method of Calibration

Transcription:

The Annals of Applied Probability 2004, Vol. 4, No. 4, 2090 29 DOI: 0.24/0505604000000846 c Institute of Mathematical Statistics, 2004 arxiv:math/0503556v [math.pr] 24 Mar 2005 NUMBER OF PATHS VERSUS NUMBER OF BASIS FUNCTIONS IN AMERICAN OPTION PRICING By Paul Glasserman and Bin Yu Columbia University An American option grants the holder the right to select the time at which to exercise the option, so pricing an American option entails solving an optimal stopping problem. Difficulties in applying standard numerical methods to complex pricing problems have motivated the development of techniques that combine Monte Carlo simulation with dynamic programming. One class of methods approximates the option value at each time using a linear combination of basis functions, and combines Monte Carlo with backward induction to estimate optimal coefficients in each approximation. We analyze the convergence of such a method as both the number of basis functions and the number of simulated paths increase. We get explicit results when the basis functions are polynomials and the underlying process is either Brownian motion or geometric Brownian motion. We show that the number of paths required for worst-case convergence grows exponentially in the degree of the approximating polynomials in the case of Brownian motion and faster in the case of geometric Brownian motion.. Introduction. An American option grants the holder the right to select the time at which to exercise the option, and in this differs from a European option which may be exercised only at a fixed date. A standard result in the theory of contingent claims states that the equilibrium price of an American option is its value under an optimal exercise policy see, e.g., Chapter 8 of [6]). Pricing an American option thus entails solving an optimal stopping problem, typically with a finite horizon. Solving this optimal stopping problem and pricing an American option are relatively straightforward in low dimensions. Assuming a Markovian formulation of the problem, the relevant dimension is the dimension of the state Received June 2003; revised November 2003. Supported in part by NSF Grant DMS-00-74637. AMS 2000 subject classifications. Primary 60G40; secondary 65C05, 65C50, 60G35. Key words and phrases. Optimal stopping, Monte Carlo methods, dynamic programming, orthogonal polynomials, finance. This is an electronic reprint of the original article published by the Institute of Mathematical Statistics in The Annals of Applied Probability, 2004, Vol. 4, No. 4, 2090 29. This reprint differs from the original in pagination and typographic detail.

2 P. GLASSERMAN AND B. YU vector, and this is ordinarily at least as large as the number of underlying assets on which the payoff of the option depends. In up to about three dimensions, the problem can be solved using a variety of numerical methods, including binomial lattices, finite-difference methods and techniques based on variational inequalities. See, e.g., Chapter 5 of [0] or Chapter 9 of [9] for an introduction to these methods.) But many problems arising in practice have much higher dimensions, and these applications have motivated the development of Monte Carlo methods for pricing American options. The optimal stopping problem embedded in the valuation of an American option makes this an unconventional and challenging problem for Monte Carlo. One class of techniques, based primarily on proposals of Carrière [4], Longstaff and Schwartz [] and Tsitsiklis and Van Roy [7, 8], provides approximate solutions by combining simulation, regression and a dynamic programming formulation of the problem. Related methods have been used to solve dynamic programming problems in other contexts; Bertsekas and Tsitsiklis [] discuss several techniques and applications. In this approach to American option pricing, the value function describing the option price at each time as a function of the underlying state is approximated by a linear combination of basis functions; the coefficients in this representation are estimated by applying regression to the simulated paths. Such an approximation is computed at each step in a dynamic programming procedure that starts with the option value at expiration and works backward to find the value at the current time. Any such method clearly restricts the number of possible exercise dates to be finite; these dates may be specified in the terms of the option, or they may serve as a discrete-time approximation for a continuously-exercisable option. The convergence results available to date for these methods are based on letting the number of simulated paths increase while holding the number of basis functions fixed. Tsitsiklis and Van Roy [8] prove such a result for their method and Clément, Lamberton and Protter [5] do this for the method of Longstaff and Schwartz []. The two methods differ in the backward induction procedure they use to solve the dynamic programming problem.) The convergence established by these results is therefore convergence to the approximation that would be obtained if the calculations could be carried out exactly, without the sampling error associated with Monte Carlo. Convergence to the correct option price requires a separate passage to the limit in which the number of basis functions increases. This paper considers settings in which the number of paths and number of basis functions increase together. Our objective is to determine how quickly the number of paths must grow with the number of basis functions to ensure convergence to the correct value. The growth required turns out to be surprisingly fast in the settings we analyze. We take the underlying process to be Brownian motion or geometric Brownian motion and regress against

AMERICAN OPTION PRICING 3 polynomials in each case. We examine conditions for convergence to hold uniformly over coefficient vectors having a fixed norm, and in this sense our results provide a type of worst-case analysis. We show that for Brownian motion, the number of polynomials K = K N for which accurate estimation is possible from N paths is Olog N); for geometric Brownian motion it is O logn ). Thus, the number of paths must grow exponentially with the number of polynomials in the first case, faster in the second case. Focusing on simple models allows us to give rather precise results. Our most explicit results apply to one-dimensional problems that do not require Monte Carlo methods, but we believe they are, nevertheless, relevant to higher-dimensional problems. Many high-dimensional interest rate models have dynamics that are nearly Brownian or nearly log-brownian; see, for example, the widely used models in Chapters 4 and 5 of [3]. Our focus on polynomials helps make our results explicit and is also consistent with, for example, examples in [] and remarks in [5]. Our analysis relies on asymptotics of moments of the functions used in the regressions. To the extent that similar asymptotics could be derived for other basis functions and underlying distributions, our approach could be used in other settings. We prove two types of results, providing upper and lower bounds on K and thus corresponding to negative and positive results, respectively. For an upper bound on K, it suffices to exhibit a problem for which convergence fails. For this part of the analysis we therefore consider a single-period problem a single regression and a single step in the backward induction. The fact that an exponentially growing sample size is necessary even in a one-dimensional, single-period problem makes the result all the more compelling. For the positive results we consider an arbitrary but fixed number of steps, corresponding to a finite set of exercise opportunities. We prove a general error bound that relies on few assumptions about the underlying Markov process or basis functions, and then specialize to the case of polynomials with Brownian motion and geometric Brownian motion. Section 2 formulates the American option pricing problem, discusses approximate dynamic programming and presents the algorithm we analyze. Section 3 undertakes the single-period analysis, first in a normal setting then in a lognormal setting. Section 4 presents results for the multiperiod case. Proofs of some of the results in Sections 3 and 4 are deferred to Sections 5 and 6, respectively. 2. Problem formulation. In this section we first give a general description of the American option pricing problem, then discuss approximate dynamic programming procedures and then detail the algorithm we analyze. 2.. The optimal stopping problem. A general class of American option pricing problems can be formulated through an R d -valued Markov process

4 P. GLASSERMAN AND B. YU {St), 0 t T }, [with S0) fixed], that records all relevant financial information, including the prices of underlying assets. We restrict attention to options admitting a finite set of exercise opportunities 0 = t 0 < t < t 2 < < t m T, sometimes called Bermudan options. We preserve the continuoustime specification of S because the lengths of the intervals t i+ t i appear in some of our results.) If exercised at time t n, n = 0,,...,m, the option pays h n St n )), for some known functions h 0,h,...,h m mapping R d into [0, ). Let T n denote the set of stopping times with respect to the history of S) taking values in {t n,t n+,...,t m } and define ) V n x) = sup τ T n E[h τ Sτ)) St n ) = x], x R d, for n = 0,,...,m. Then V n x) is the value of the option at t n in state x, given that the option was not exercised prior to t n. For simplicity, we have not included explicit discounting in ). Deterministic discounting can be absorbed into the definition of the functions h n, and stochastic discounting can usually be accommodated in this formulation at the expense of increasing the dimension of S. The option values satisfy the dynamic programming equations 2) 3) V m x) = h mx), V n x) = max{h nx),e[v n+ St n+)) St n ) = x]}, n = 0,,..., m. These can be rewritten in terms of continuation values as C n x) = E[V n+ St n+)) St n ) = x], n = 0,,...,m, 4) 5) C mx) = 0, C nx) = E[max{h n+ St n+ )),C n+st n+ ))} St n ) = x], n = 0,,...,m. The option values satisfy V n x) = max{h n x),c nx)}, so these can be calculated from the continuation values. 2.2. Approximate dynamic programming. Exact calculation of 2) 3) or 4) 5) is often impractical, and even estimation by Monte Carlo is challenging because of the difficulty of estimating the conditional expectations in these equations. Approximate dynamic programming procedures replace these conditional expectations with linear combinations of known functions,

AMERICAN OPTION PRICING 5 sometimes called features but more commonly referred to as basis functions. Thus, for each n =,...,m, let ψ nk, k = 0,...,K, be functions from R d to R and consider approximations of the form K Cn x) β nk ψ nk x), for some constants β nk, or the corresponding approximation for V n. Working with approximations of this type reduces the problem of finding the functions C n to one of finding the coefficients β nk. The methods of Longstaff and Schwartz [] and Tsitsiklis and Van Roy [7, 8] select coefficients through least-squares projection onto the span of the basis functions. Other methods applying Monte Carlo to solve 2) 3) include Broadie and Glasserman [2, 3], Haugh and Kogan [9] and Rogers [6]; for an overview, see Glasserman [8]. To simplify notation, we write S n for St n ). We write ψ n for the vector of functions ψ n0,...,ψ nk ). The following basic assumptions will be in force throughout: A0) ψ n0 for n =,...,m; E[ψ n S n )] = 0, for n =,...,m; and Ψ n = E[ψ n S n )ψ n S n ) ] is finite and nonsingular, n =,...,m. For any square-integrable random variable Y define the projection Thus, Π n Y = ψ n S n)ψ n E[Y ψ ns n )]. 6) Π n Y = a k ψ nk S n ) with 7) a 0,...,a K ) = Ψ n E[Y ψ n S n )]. We also write Π n Y )x) = a k ψ nk x) for the function defined by the coefficients 7). Define an approximation to 4) 5) as follows: C m x) 0, 8) C n x) = Π n max{h n+ S n+ ),C n+ S n+ )})x).

6 P. GLASSERMAN AND B. YU As in 6), the application of the projection Π n results in a linear combination of the basis functions, so 9) C n x) = Π n max{h n+,c n+ })x) = β nk ψ nk x) with β n = β n0,...,β nk ) defined as in 7) but with Y replaced by 0) V n+ S n+ ) max{h n+ S n+ ),C n+ S n+ )}. With the payoff functions h n fixed, we can rewrite 9) using the operator ) L n C n+ = Π n max{h n+,c n+ }). Exact calculation of the projection in 8) is usually infeasible, but it is relatively easy to evaluate a sample counterpart of this recursion defined from a finite set of simulated paths of the process S. We consider the following procedure to approximate the coefficient vectors β n and the continuation values C n. Step. Set Ĉm = 0 and ˆV m = max{h m,ĉm} = h m. Step 2. For each n =,...,m, repeat the following steps: Generate N paths {S i),...,si) n+ }, i =,...,N, up to time t n+, independent of each other and of all previously generated paths. Calculate ˆγ n = N N i= ˆV n+ S i) n+ )ψ ns i) n ), calculate the coefficients ˆβ n = Ψ n ˆγ n and set 2) 3) Ĉ n = ˆβ n ψ n ˆL n Ĉ n+ ˆΠ n max{h n+,ĉn+}, ˆV n = max{h n,ĉn}. Step 3. Set Ĉ0S 0 ) = N N i= ˆV S i) ) and ˆV 0 S 0 ) = max{h 0 S 0 ),Ĉ0S 0 )}. A few aspects of this algorithm require comment. In Step 3 we simply average the estimated values at t to get the continuation value at time 0 because S0) is fixed. The operators ˆL n and ˆΠ n implicitly defined in 2) are the sample counterparts of those in 6) and ), using estimated rather than exact coefficients. The coefficient estimates in Step 2 use the matrices Ψ n. In ordinary least-squares regression, each Ψ n would be replaced with its sample counterpart, N N i= ψ n S i) n )ψ n S i) n ),

AMERICAN OPTION PRICING 7 calculated from the simulated values themselves. Owen [4] calls the use of the exact matrix quasi-regression.) In our examples, the Ψ n are indeed available explicitly and using this formulation simplifies the analysis. In Step 2 we have used an independent set of paths to estimate coefficients at each date, though the algorithms of Longstaff and Schwartz [] and Tsitsiklis and Van Roy [7, 8] use a single set of paths for all dates. This modification is theoretically convenient because it makes the coefficients of Ĉn+ independent of the points at which Ĉn+ is evaluated in the calculation of ˆγ n. This distinction is relevant only to the multiperiod analysis of Section 4 and disappears in the single-period analysis of Section 3. The worst case over all multiperiod problems is at least as bad as the worst single-period problem. The results in Section 3 thus provide lower bounds on the worst-case convergence rate for multiperiod problems whether one uses independent paths at each date or a single set of paths for all dates. 3. Single-period problem. For the single-period problem, we fix dates t < t 2 and consider the estimation of coefficients β 0,...,β K in the projection of a function of S 2 onto the span of ψ k S ), k = 0,...,K. Thus, 4) β = β 0,...,β K ) = Ψ γ with Ψ = Ψ and γ = E[Y ψ S )] for some Y. In a simplified instance of the algorithm of the previous section, we simulate N independent copies S i),y i) ), i =,...,N, and compute the estimate β = Ψ γ, 5) where γ is the unbiased estimator of γ with components 6) γ k = N N i= Y i) ψ k S i) ), k = 0,,...,K. We analyze the convergence of β and γ) as both N and K increase. We denote by x the Euclidean norm of the vector x. For a matrix A, we denote by A the Euclidean matrix norm, meaning the square root of the sum of squared elements of A. It follows that Ax A x and then from 4) and 5), 7) Ψ γ γ β β Ψ γ γ. The Euclidean norm on vectors is a measure of the proximity of the functions determined by vectors of coefficients. To make this more explicit, let b and c be coefficient vectors and let S n have density g n. Then K 2 b k ψ nk x) c k ψ nk x)) g n x)dx = b c) Ψ n b c)

8 P. GLASSERMAN AND B. YU and Ψ n b c 2 b c) Ψ n b c) Ψ n b c 2. Thus, the Euclidean norm on vectors gives the L 2 norm with respect to g n ) for the functions determined by the vectors, up to factors of Ψ n and Ψ n that will prove to be negligible in the settings we consider. We therefore investigate the convergence of the expected squared difference E[ β β 2 ]. Because this is the mean square error of β, we also denote it by MSE β). Thus, 7) implies 8) Ψ 2E[ γ γ 2 ] MSE β) Ψ 2 E[ γ γ 2 ]. For a given number of replications N and basis functions K, MSE β) can be made arbitrarily large or small by multiplying β by a constant. To get meaningful results, we therefore adopt the following normalization: A) β =. We investigate the convergence of the supremum of the MSE β) over all β satisfying this condition. In order to investigate how N must grow with K, we assume that the regression representation is, in fact, valid, in a sense implied by the following two conditions: A2) Y has the form for some constants a k. Y = a k ψ 2k S 2 ), A3) There exist functions f k : R + R +, k = 0,...,K, such that E[f k t 2 )ψ 2k S 2 ) S ] = f k t )ψ k S ), t 2 t. Condition A3) states that the ψ nk S n ) are martingales, up to a deterministic function of time. Condition A2), though a strong assumption in practice, makes Theorems and 2 more compelling: the rapid growth in the number of paths implied by the theorems holds even though we have chosen the correct basis functions, in the sense of A2). The results of Section 4 give sufficient conditions for convergence without such an assumption. Under assumptions A2) and A3), we have 9) γ k = E[Y ψ k S )] [ K ] = E a l ψ 2l S 2 )ψ k S ) l=0

AMERICAN OPTION PRICING 9 = l=0 a l f l t ) f l t 2 ) E[ψ ls )ψ k S )]. The restriction on β in A) then restricts a. Returning to the analysis of MSE β), 8) indicates that we need to analyze the mean square error of γ, for which since E[ γ] = γ) we get 20) 2) E[ γ γ 2 ] = = Var[ γ k ] N E[Y 2 ψ 2 ks )] N γk. 2 Thus, using 8), A2) and the Cauchy Schwarz inequality, 22) MSE β) Ψ 2 E[ γ γ 2 ] Ψ 2 K N E[Y 2 ψ 2 ks )] Ψ 2 a 2 l E[ψ2j 2 N S 2)ψk 2 S )]. l=0 k,j=0 To get a lower bound, we may define Y = a K ψ 2KS 2 ), with a K chosen such that the corresponding β satisfies β =. Using 8) and 20), we then get 23) sup MSE β) β = Ψ 2 = Ψ 2 K a 2 K N E[Y 2 ψ 2 k S )] N ) γk 2 N E[ψ2 2K S 2)ψ 2 k S )] N ) γk 2. From 22) and 23) we see that the key to the analysis of the uniform convergence of MSE β) lies in the growth of fourth-order moments of the form E[ψ 2 2j S 2)ψ 2 k S )]. This, in turn, depends on the choice of basis functions and on the law of the underlying process S. We analyze the case of polynomials with Brownian motion and geometric Brownian motion. 3.. Normal setting. For this section, let {St),0 t T } be a standard Brownian motion. We define the basis functions through the Hermite polynomials n/2 ) i n!x n 2i H en x) = n 2i)!i!2 i=0 i, n = 0,,...,

0 P. GLASSERMAN AND B. YU where n/2 denotes the integer part of n/2. The Hermite polynomials have the following useful properties: They are orthogonal with respect to the standard normal density φ, in the sense that H e0 and { 0, i j, H ei x)h ej x)φx)dx = i!, i = j. They define martingales, in the sense that see, e.g., [5], page 5) [ ) E t i/2 2 H St2 ) St ] e i ) = t i/2 H St ) e i ), t2 t for t 2 t. And their squares admit the expansion 24) The functions 25) H en x)) 2 = n!) 2 n i=0 H e2i x) i!) 2 n i)!. ψ nk x) = k! H ek x/ t n ) satisfy A3) with f k t) = t k/2. They are also orthogonal and their Ψ matrix is the identity. Thus, β = γ and β = γ. We can now state the main result of this section. Let ρ = t 2 /t, and for ρ define c ρ = 2log2 + ρ). Table Estimates of MSE β) for various combinations of K basis functions and N paths. The critical values K = log N/c ρ are displayed by in the bottom row and also indicated by the horizontal line through the table N K 500 000 2000 4000 8000 6000 32000 64000 28000 0.0 0.00 0.00 0.00 0.00 0.00 0.00 0.00 0.00 2 0.08 0.04 0.02 0.0 0.00 0.00 0.00 0.00 0.00 3 0.67 0.3 0.7 0.08 0.04 0.02 0.0 0.00 0.00 4 5.6 3.0.6 0.73 0.36 0.8 0.09 0.05 0.02 5 52.7 23.4 3.5 6.0 3..5 0.8 0.40 0.20 6 427.2 55.7 93.3 38.4 24.0 0.8 6.2 3..5 7 2403 202 600.8 300.4 50.2 75. 37.5 8.8 9.4 8 447 5723 2862 43 75.4 357.7 78.9 89.4 44.7 9 9856 4928 2464 232 66 308 54 0 609 3054 527 764 38 280 405 702 2 023 Bound 2.5 2.8 3. 3.4 3.7 3.9 4.2 4.5 4.8

AMERICAN OPTION PRICING Theorem. Let ψ nk be as in 25) and suppose A2) holds. If K = δ) logn/c ρ for some δ > 0, then 26) lim sup MSE β) = 0. N β = If K = + δ)log N/c ρ for some δ > 0, then 27) lim sup MSE β) =. N β = This result shows rather precisely that, from a sample size of N, the highest K for which coefficients of polynomials of order K can be estimated uniformly well is Olog N). Equivalently, the sample size required to achieve convergence grows exponentially in K. This is illustrated numerically in Table, which shows estimates of MSE β) for various combinations of N and K. The results shown are for Y = ρ K/2 H ek S 2 / t 2 )/ K!, with t = and t 2 = 2, a special case of the Y we use to prove 27). The estimates are computed as follows. For each entry of the table, we generate 5000 batches, each consisting of N paths. From each batch we compute β and then take the average of β β 2 over the 5000 batches. This average provides our estimate of MSE β) in each case with K 6. For K 7 this produced unacceptably high variability, so for those cases we calculated MSE β) from 5000 replications of N = 500,000 and then scaled the estimate by N. The bottom row of the table displays the critical values K = logn/c ρ provided by Theorem ; these values are also indicated by the horizontal line through the table. As indicated by the theorem, MSE β) explodes along any diagonal line through the table steeper than the critical line, and remains small above the critical line. The proof of the theorem uses the following two lemmas, proved in Section 5. Lemma. For the ψ nk in 25) and ρ = t 2 /t, 28) 29) E[ψ 2k2 S 2 )ψ k S )] = { 0, k k 2, ρ k/2, k = k 2, k E[ψ 2k2 S 2 ) 2 ψ k S ) 2 k 2 ] = ρ k 2k k ) k k ) ) k2, k with k k 2 the minimum of k and k 2. Equation 29) is strictly increasing in k and k 2.

2 P. GLASSERMAN AND B. YU 30) For the special case k = k 2 = K, 29) yields ) ) 2 E[ψ 2K S 2 ) 2 ψ K S ) 2 ] = ρ k 2k 2K. k k As a step toward bounding this expression, let k denote the index of the largest summand so that 2k ) ) 2 ) ) 2 K 2k K 3) ρ k k k = max 0 k K ρ k. k k For k, we have the following lemma. Lemma 2. As K, k = 2 2 + K + o)). ρ Proof of Theorem. We bound MSE β) from above based on 22). Combining the fact that β = γ because Ψ = I) with 9) and 28) we get β k = a l E[ψ 2l S 2 )ψ k S )] = a k ρ k/2. l=0 Thus, β = implies a 2 k ρk. From 30) and 3), we get 32) ρ k 2k k ) K k ) 2 < E[ψ 2KS 2 ) 2 ψ KS ) 2 ] 2k ) ) 2 K < K + )ρ k k k. Recalling that Ψ = I and applying the inequality a 2 k ρk to 22) we get 33) 34) sup MSE β) sup Ψ 2 β = β = N K + ) N < < ρ k a 2 l E[ψ2lS 2 2 )ψks 2 )] l=0 k,l=0 K k,l=0 E[ψ 2 2k S 2)ψ 2 l S )] K + )2 ρ K K + ) 2 E[ψ2K 2 N S 2)ψK 2 S )] K + )5 N 2k ) ) 2 K ρ K k k k, where 33) follows from Lemma and the last inequality follows from 32).

AMERICAN OPTION PRICING 3 To get a lower bound on the supremum of MSE β) we use 23) with Y = ρ K/2 ψ 2K S 2 ) a K ψ 2KS 2 ), for which β K = and β k = 0, k K. By applying Lemma and the lower bound in 32), 23) becomes sup MSE β) ) β = Ψ 2 a 2 K E[ψ 2 N 2KS 2 )ψks 2 )] 35) Ψ 2 a 2 K N E[ψ2 2K S 2)ψ 2 K S )] K + N ρk k 2k ) ) 2 K k. By Stirling s approximation n! 2nπ n e )n and Lemma 2 we get K k ) = = K! k!k k )! 2KπK/e) K + o)) 2k π 2K k )πk /e) k K k /e) K k + o)) k 36) = + o)), 2abKπa ak bbk with a = 2/2 + ρ) and b = a. Also, 2k ) k = 2k! k!k! 4k = π2k /e) 2k + o)) 2k πk /e) 2k + o)) 37) = 22aK akπ + o)). By substituting 36) and 37) into 34) and 35) we get ρ bk 2 2aK 2NK + ) + o)) akπabkπa 2aK b2bk sup MSE β) β = K + ) 5 ρ bk 2 2aK 2N + o)). akπabkπa 2aK b2bk

4 P. GLASSERMAN AND B. YU Simple algebra verifies that c ρ = 2alog2) 2aloga) 2blogb) + blogρ), so we can rewrite these bounds as e cρk 2NK + ) akπabkπ + o)) sup MSE β) β = K + )5 e cρk 2N + o)). akπabkπ If K = δ)log N/c ρ for some δ > 0, then as N, { K + ) 5 e cρk } log 2N + o)) = δ logn + olog N), akπabkπ so 26) holds. If K = + δ)log N/c ρ for some δ > 0, then as N, { e cρk } log 2NK + ) + o)) = δ logn + ologn), akπabkπ and 27) holds. 3.2. Lognormal setting. We now take S to be geometric Brownian motion, St) = expwt) t/2), with W a standard Brownian motion. For the basis functions ψ nk ψ k, we use multiples of the powers x k to get the martingales 38) ψ k St)) = e kwt) k2 t/2. These functions satisfy A0). The main result of this section is the following: Theorem 2. Let the ψ k be as in 38) and suppose A2) holds. If δ)log N K = 5t + t 2 for some δ > 0, then If for some δ > 0, then lim sup MSE β) = 0. N β = K = + δ)log N 3t + t 2 lim sup MSE β) =. N β =

AMERICAN OPTION PRICING 5 Compared with the normal case in Theorem, we see that here K must be much smaller of the order of logn. Accordingly, N must be much larger of the order of expk 2 ). The analysis in this setting is somewhat more complicated than in the normal case because the ψ k are no longer orthogonal. To prove the theorem we state some lemmas that are proved in Section 5. Lemma 3. For t 2 t and k,k 2 = 0,...,K, E[ψ k S )ψ k2 S 2 )] = e k k 2 t, E[ψ 2 k S )ψ 2 k 2 S 2 )] = e k2 t +k 2 2 t 2+4k k 2 t, and E[ψ k S ) 2 ψ k2 S 2 ) 2 ] is strictly increasing in k and k 2. Using the first statement in the lemma, we find that the matrix Ψt) with ijth entry E[ψ i St))ψ j St))] is given by e t e 2t e Kt Ψt) = e 2t e 4t e 2Kt........ e Kt e 2Kt e K2 t We write Ψ for Ψt ). Lemma 4. We have Ψt) K+) 2 e 2K2t and, with Ct) = exp 2e/e t ) 2 ), e Ψt) C t ) K t)kk + ) e t. Proof of Theorem 2. Condition A2) and the martingale property of the ψ k St)) imply that E[Y S ] = a k ψ k S ), and, thus, that β k = a k, k = 0,,...,K. In this case, the normalization β = is equivalent to a 0,...,a K ) =. Applying this in 22) and then applying Lemmas 3 and 4 we get ] sup MSE β) sup β = β = Ψ 2 [ K N E ψk 2 S 2)ψK 2 S )

6 P. GLASSERMAN AND B. YU Ψ 2 N K + )E[ψ2 KS 2 )ψ 2 KS )] e C 2 t )K 2 K + ) 2 t ) 2K K + t +K 2 t 2 e t N e5k2. If we now take K = δ)log N 5t +t 2, then as N, e log {Ct ) 2 K 2 K + ) 3 t ) 2K } t +K 2 t 2 e t N e5k2 + o)) = δ logn + olog N), which proves the first assertion in the theorem. For the second part of the theorem, define Y = e KWt 2) K 2 t 2 /2, for which β is 0,...,0,). The corresponding vector γ is Ψβ, the last column of Ψ. Applying this in 23) and using Lemmas 3 and 4 we get sup MSE β) K β = Ψ 2 N E[ψ2 KS 2 )ψks 2 )] ) γ N k) 2 Ψ 2 N E[ψ2 K S 2)ψ 2 K S )] γ K )2 ) NK + ) 2 e 2K2 t e 5K2 t +K 2 t 2 e 2K2 t ) t = +K 2 t 2 + o)). NK + ) 2e3K2 If we now take K = +δ)log N 3t +t 2, then as N, { } t log +K 2 t 2 NK + ) 2 e3k2 + o)) = δ logn + ologn), proving the second assertion in the theorem. The analysis of this section differs from the normal setting of Section 3. in that the polynomials 38) are not orthogonal. In the Brownian case, the Hermite polynomials are orthogonal and after appropriate scaling) martingales. In using 38), we have chosen to preserve the martingale property rather than orthogonality. As a consequence Ψ and / Ψ appear in our bounds on MSE β). From Lemma 4 we see that Ψ has an asymptotically negligible effect on the upper bound for MSE β), and with or without the factor of / Ψ, the lower bound on MSE β) is exponential in a multiple of K 2. The slower convergence rate in the lognormal setting therefore does not appear to result from the lack of orthogonality.

AMERICAN OPTION PRICING 7 4. Multiperiod problem. We now turn to conditions that ensure convergence of the multiperiod algorithm in Section 2.2 as both the number of basis functions K and the number of paths N increase. We first formulate a general result bounding the error in the estimated continuation values, then specialize to the normal and lognormal settings. 4.. General bound. We use the following conditions. B) E[ψ 2 nk S n)] and E[ψ 4 nk S n)] are increasing in n and k. As explained in the discussion of the single-period problem, we need some normalization on the regression coefficients in order to make meaningful statements about worst-case convergence. For a problem with m exercise opportunities, we impose B2) β m =. This condition is analogous to the one we used in the single-period problem, where β was a vector of coefficients at time t and Y was a linear combination of functions evaluated at St 2 ). We also need a condition on the functions h n that determine the payoff upon exercise at time t n. The following condition turns out to be convenient: B3) E[h 4 ns n )] tn t n ) 2K E[ψ 4 nk S n)], for n = 0,,...,m. Suppose S n has density g n and define the weighted L 2 norm on functions G:R R, G n = Gx) 2 g n x)dx. With Ĉn the estimated continuation value defined by 2), we analyze the error E[ Ĉn C n 2 n ]. We need some additional notation. Let t n+ c = max, B K = max n=,...,m t n n=,...,m Ψ n, H K = max{c K,BK 2 K + )}, A K = K + )H K E[ψmK 4 S m)]. Under A0), B K is well defined. We can now state the main result of this section. 39) Theorem 3. If assumptions A0) and B) B3) hold, then E[ Ĉn C n 2 n] 2 m n ) K + )2 N B K A m n K E[ψ2 mks m )]) 2 + o)). This result is proved in Section 6. Its consequences will be clearer once we illustrate it in the normal and lognormal settings.

8 P. GLASSERMAN AND B. YU 4.2. Multiperiod examples. 4.2.. Normal setting. As in Section 3., let S be a standard Brownian motion and let the ψ nk be as in 25). Each Ψ n is then the identity matrix, n =,...,m. It follows that 40) Also, B K = max n Ψ n = K +. H K = max{c K,BK 2 K + )} = ck for all sufficiently large K. To bound E[ψ 4 mk S m)] which appears in A K ), we use 29) with t = t 2 and k = k 2 = K) and then Stirling s formula and Lemma 2 to get ) ) 2 E[ψmKS 4 2k K 9 3 m )] = K + ) k k 4 + o)). 2K 3 π 332K The expression on the right follows from 32), 36) and 37) upon noting that with ρ = we get a = 2/3 and b = /3. Substituting this expression and 40) into 39) yields E[ Ĉn C n 2 ] < 2 m n ) It now follows that if for some δ > 0, then 4) ) K + )2m 2n+5/2 9 3 m n N 4 c m n)k + o)). 2K 3 π 332K K = lim N β m = δ)log N m n)2log3 + logc) sup E[ Ĉn C n 2 n] = 0. In other words, we have convergence of the estimated continuation values at all exercise opportunities, as both N and K increase. If the basis functions eventually span the true optimum, in the sense that C n C n n 0 as K, then by the triangle inequality, 4) holds with C n replaced by C n. On the other hand, from Theorem we know that if K = +δ)log N/c ρ for any δ > 0, with ρ = t m /t m, then lim sup N β m = E[ Ĉ m C m 2 m ] =. Thus, the crititcal rate of K for the multiperiod problem is OlogN), just as in the single-period problem.