28.06.2014 Views

sparse grid method in the libor market model. option valuation and the

sparse grid method in the libor market model. option valuation and the

sparse grid method in the libor market model. option valuation and the

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

SPARSE GRID METHOD IN THE LIBOR MARKET<br />

MODEL. OPTION VALUATION AND THE ’GREEKS’<br />

MSc Thesis (Afstudeerscriptie)<br />

written by<br />

Vladislav Sergeev<br />

(born January 25th, 1980 <strong>in</strong> Omsk city, Russian Federation)<br />

under <strong>the</strong> supervision of Dr. Drona K<strong>and</strong>hai, <strong>and</strong> submitted to <strong>the</strong> Board of<br />

Exam<strong>in</strong>ers <strong>in</strong> partial fulfillment of <strong>the</strong> requirements for <strong>the</strong> degree of<br />

MSc <strong>in</strong> Grid Comput<strong>in</strong>g<br />

at <strong>the</strong> Universiteit van Amsterdam.<br />

Date of <strong>the</strong> public defense:<br />

November 28th, 2006<br />

Members of <strong>the</strong> Thesis Committee:<br />

Dr. Peter van Emde Boas<br />

Prof. Dr. Chris Jesshope<br />

Dr. Drona K<strong>and</strong>hai<br />

Mr. Breanndán Ó Nuallá<strong>in</strong>, MSc.


Abstract<br />

In this project we study numerical approximations of <strong>the</strong> Libor Market Model by us<strong>in</strong>g<br />

Sparse Grid techniques. This approach is a promis<strong>in</strong>g <strong>method</strong> to solve high dimensional<br />

partial differential equations. The LIBOR Market Model is a <strong>model</strong> for<br />

<strong>in</strong>terest rate movements <strong>and</strong> typically suffers from <strong>the</strong> curse of dimensionality when<br />

f<strong>in</strong>ite-difference techniques are used. As an extension of <strong>the</strong> earlier work by J. Blackham<br />

[Blackham, 2004], we target a more detailed study of <strong>the</strong> accuracy of <strong>the</strong> <strong>sparse</strong><br />

’Greeks’ (∆ <strong>and</strong> V) for 2D Chooser Option, 3D Bermudan Swaption <strong>and</strong> a 2D Digital<br />

Option. Given <strong>the</strong> results of experiments, we conclude that, despite satisfactory convergence<br />

of <strong>sparse</strong> <strong>grid</strong> values, fur<strong>the</strong>r improvements are required to enhance <strong>the</strong> quality of<br />

<strong>the</strong> ’Greeks’ <strong>in</strong> <strong>the</strong> discussed sett<strong>in</strong>g. Applicable recommendations of possible fur<strong>the</strong>r<br />

developments are <strong>in</strong>cluded accord<strong>in</strong>gly.


Acknowledgements<br />

This project would not have been possible without <strong>the</strong> support of many people. First of<br />

all, I would like to express repect <strong>and</strong> gratitude to my supervisor, dr. Drona K<strong>and</strong>hai,<br />

whose professionalism, bonded with remarkable personality, has guided <strong>and</strong> tutored me<br />

throughout <strong>the</strong> whole year. Thanks for mak<strong>in</strong>g sense out of confusion <strong>and</strong> parent<strong>in</strong>g<br />

my <strong>in</strong>terest <strong>in</strong> Derivative F<strong>in</strong>ance. Words of acknowledgment are due to Dick van<br />

Albada, Joke Blom, Walter Hoffmann, Dmitri Vasun<strong>in</strong> <strong>and</strong> Vladimir Korkhov, who at<br />

different times provided <strong>the</strong>ir <strong>in</strong>sightful comments. A large ’thank you’ goes to Kees<br />

Oosterlee <strong>and</strong> Coen Leentvaar from TU Delft, whose expertise <strong>and</strong> feedback proved to<br />

be <strong>in</strong>valuable. I am blessed to be a friend to many who had <strong>the</strong>ir h<strong>and</strong>s on <strong>the</strong> pulse<br />

of this project. Nelly’s patience, love <strong>and</strong> care secured my sanity. F<strong>in</strong>ally <strong>and</strong> most<br />

importantly, I am deeply <strong>in</strong>debted to my parents.<br />

Äîðîãèå ðîäèòåëè, ñïàñèáî çà âñå ÷òî âû äëÿ ìåíÿ ñäåëàëè. ß âàñ î÷åíü<br />

ëþáëþ.


Contents<br />

1 Introduction 1<br />

2 Asset Pric<strong>in</strong>g Overview 3<br />

2.1 Arbitrage <strong>and</strong> Risk-Neutral World . . . . . . . . . . . . . . . . . . . 3<br />

2.2 From Discrete Time to Cont<strong>in</strong>uous Time . . . . . . . . . . . . . . . . 6<br />

2.3 Pric<strong>in</strong>g with Mart<strong>in</strong>gales <strong>and</strong> Measures . . . . . . . . . . . . . . . . 13<br />

3 LIBOR Market Model 17<br />

3.1 Forward LIBOR Rates . . . . . . . . . . . . . . . . . . . . . . . . . 17<br />

3.2 LIBOR Market Model . . . . . . . . . . . . . . . . . . . . . . . . . 19<br />

3.3 Interest Rate Derivatives. Caplets, Caps <strong>and</strong> Swaptions. . . . . . . . . 22<br />

3.4 Calibration of LIBOR Market Model . . . . . . . . . . . . . . . . . . 24<br />

3.4.1 Calibration to caplets . . . . . . . . . . . . . . . . . . . . . . 25<br />

3.4.2 Correlation structure of forward rates. . . . . . . . . . . . . . 25<br />

3.5 The LIBOR Market Model PDE . . . . . . . . . . . . . . . . . . . . 28<br />

4 Computational Methods 30<br />

4.1 Monte Carlo . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30<br />

4.2 F<strong>in</strong>ite Difference Method . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

5 Sparse Grids 37<br />

5.1 Full <strong>grid</strong> approximations. . . . . . . . . . . . . . . . . . . . . . . . . 37<br />

5.2 Sparse <strong>grid</strong> approximations. . . . . . . . . . . . . . . . . . . . . . . 41<br />

5.3 A Comb<strong>in</strong>ation technique . . . . . . . . . . . . . . . . . . . . . . . . 42<br />

6 Numerical Results. Valuation <strong>and</strong> <strong>the</strong> Greeks 43<br />

6.1 Two Dimensional Case. 2D Chooser Option . . . . . . . . . . . . . . 44<br />

6.1.1 Convergence of MC, Full <strong>and</strong> Sparse Grid solutions. . . . . . 45<br />

6.1.2 Discussion: Convergence of Pric<strong>in</strong>g. . . . . . . . . . . . . . . 49<br />

6.1.3 Greeks: Delta . . . . . . . . . . . . . . . . . . . . . . . . . . 50<br />

6.1.4 Greeks: Vega . . . . . . . . . . . . . . . . . . . . . . . . . . 52<br />

6.2 Three Dimensional Case. 3D Bermudan Swaption . . . . . . . . . . . 54<br />

6.2.1 Convergence Study . . . . . . . . . . . . . . . . . . . . . . . 56<br />

6.2.2 Greeks. Delta <strong>and</strong> Vega. . . . . . . . . . . . . . . . . . . . . 58<br />

6.3 Discont<strong>in</strong>uous Payoff. 2D Digital <strong>option</strong>. . . . . . . . . . . . . . . . 61<br />

ii


6.3.1 Valuation Results. . . . . . . . . . . . . . . . . . . . . . . . 61<br />

6.3.2 Greeks. Delta <strong>and</strong> Vega. . . . . . . . . . . . . . . . . . . . . 64<br />

6.4 Discussion <strong>and</strong> Recommendations. . . . . . . . . . . . . . . . . . . . 66<br />

7 Conclusion <strong>and</strong> Future Work 67<br />

A Implementation Details 69<br />

iii


List of Figures<br />

2.1 From S<strong>in</strong>gle Period to Multi-Period . . . . . . . . . . . . . . . . . . 7<br />

3.1 Tenor Structure of Interest Rates . . . . . . . . . . . . . . . . . . . . 18<br />

4.1 Monte Carlo. Delta of <strong>the</strong> <strong>option</strong> with a digital feature. . . . . . . . . 32<br />

4.2 2D Solution mesh. . . . . . . . . . . . . . . . . . . . . . . . . . . . 33<br />

4.3 B<strong>and</strong> coefficient matrix. . . . . . . . . . . . . . . . . . . . . . . . . . 36<br />

5.1 St<strong>and</strong>ard Basis functions S 3 . . . . . . . . . . . . . . . . . . . . . . 38<br />

5.2 Hierarchical representation. . . . . . . . . . . . . . . . . . . . . . . . 39<br />

5.3 Mapp<strong>in</strong>g of hierarchical representation S ∗ 3 onto a full <strong>grid</strong>. . . . . . . 40<br />

5.4 Mapp<strong>in</strong>g of hierarchical representation Ŝ∗ 3 onto a <strong>sparse</strong> <strong>grid</strong>. . . . . . 41<br />

6.1 2D Chooser Option. MC vs. FD Convergence. . . . . . . . . . . . . . 45<br />

6.2 2D Chooser. Convergence of Full <strong>and</strong> Sparse Grid. . . . . . . . . . . 46<br />

6.3 2D Chooser. Solution Surface for level 6. . . . . . . . . . . . . . . . 47<br />

6.4 2D Chooser. Sparse vs. full solution. Slices. . . . . . . . . . . . . . . 48<br />

6.5 2D Chooser. Absolute error surface on level 8. . . . . . . . . . . . . . 49<br />

6.6 2D Chooser. Sparse delta surface on level 7. . . . . . . . . . . . . . . 50<br />

6.7 2D Chooser. Full vs. <strong>sparse</strong> delta profile on level 7 . . . . . . . . . . 51<br />

6.8 2D Chooser. Sparse vega surface on level 8. . . . . . . . . . . . . . . 52<br />

6.9 2D Chooser. Full vs. <strong>sparse</strong> vega slices on level 8. . . . . . . . . . . . 53<br />

6.10 3D Bermudan Swaption. Full <strong>and</strong> Sparse solution surfaces. . . . . . . 55<br />

6.11 3D Bermudan Swaption. Benchmark<strong>in</strong>g aga<strong>in</strong>st [Blackham, 2004] <strong>and</strong><br />

Monte Carlo. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56<br />

6.12 3D Bermudan Swaption. Solution slices correspond<strong>in</strong>g to fix<strong>in</strong>gs <strong>in</strong> L 1<br />

<strong>and</strong> L 2 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57<br />

6.13 3D Bermudan Swaption. Delta <strong>and</strong> Vega Surfaces for L 1 =5.62%. Full<br />

<strong>grid</strong>. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58<br />

6.14 3D Bermudan Swaption. L 3 -Delta. Slices correspond<strong>in</strong>g to a fix<strong>in</strong>g <strong>in</strong><br />

L 1 = 5% . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59<br />

6.15 3D Bermudan Swaption. L 3 -Vega. Slices correspond<strong>in</strong>g to a fix<strong>in</strong>g <strong>in</strong><br />

L 1 = 5.62% . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60<br />

6.16 2D Digital Option. Term<strong>in</strong>al payoff discretized on a L6 full <strong>grid</strong>. . . . 61<br />

6.17 2D Digital Option. Convergence of Monte Carlo <strong>and</strong> full <strong>grid</strong> FD. . . 61<br />

iv


6.18 2D Digital Option. Full <strong>and</strong> <strong>sparse</strong> <strong>grid</strong> solution surfaces. Level 7. . . 62<br />

6.19 2D Digital Option. Full <strong>grid</strong> Delta on L7. . . . . . . . . . . . . . . . 64<br />

6.20 2D Digital Option. Sparse Delta L7 <strong>and</strong> L9 slices. . . . . . . . . . . . 64<br />

6.21 2D Digital Option. Full <strong>grid</strong> vega on L7. . . . . . . . . . . . . . . . . 65<br />

6.22 2D Digital Option. Sparse Vega L7 <strong>and</strong> L9 slices. . . . . . . . . . . . 65<br />

v


List of Tables<br />

6.1 Setup of LMM. Input data for LMM . . . . . . . . . . . . . . . . . . 44<br />

6.2 Setup of LMM. LIBOR rates <strong>and</strong> caplet volatilies . . . . . . . . . . . 44<br />

6.3 2D Chooser. Convergence values for full <strong>and</strong> <strong>sparse</strong> <strong>grid</strong> solutions. . . 45<br />

6.4 2D Chooser. At-<strong>the</strong>-money <strong>and</strong> <strong>in</strong>-<strong>the</strong>-money <strong>option</strong> values. . . . . . 46<br />

6.5 2D Chooser. Deep out-of-<strong>the</strong>-money <strong>and</strong> at-<strong>the</strong>-money with low strike. 46<br />

6.6 3D Bermudan Swaption. Spot Rates as Table (6.2) . . . . . . . . . . 56<br />

6.7 3D Bermudan Swaption. Full <strong>and</strong> <strong>sparse</strong> on-<strong>grid</strong> solution values for<br />

L 1 =3.75%. 500000 MC simulation trials. . . . . . . . . . . . . . . . 56<br />

6.8 3D Bermudan Swaption. Full <strong>and</strong> <strong>sparse</strong> on-<strong>grid</strong> solution values for<br />

L 1 =5.62%. 500000 MC simulation trials. . . . . . . . . . . . . . . . 57<br />

6.9 2D Digital. Full <strong>and</strong> <strong>sparse</strong> <strong>grid</strong> convergence of on-<strong>grid</strong> po<strong>in</strong>ts. 500000<br />

MC simulation trials. . . . . . . . . . . . . . . . . . . . . . . . . . . 62<br />

6.10 2D Digital Cont<strong>in</strong>ued. Full <strong>and</strong> <strong>sparse</strong> <strong>grid</strong> convergence of on-<strong>grid</strong><br />

po<strong>in</strong>ts. 500000 MC simulation trials. . . . . . . . . . . . . . . . . . . 63<br />

vi


Chapter 1<br />

Introduction<br />

The modern field of Derivative F<strong>in</strong>ance is a very broad <strong>and</strong> highly scientific subject<br />

that spans several areas of knowledge. First of all, it naturally requires profound underst<strong>and</strong><strong>in</strong>g<br />

of <strong>the</strong> dynamics of f<strong>in</strong>ancial <strong>market</strong>s <strong>and</strong> f<strong>in</strong>ancial products traded <strong>in</strong> <strong>the</strong><br />

real world. Secondly, over <strong>the</strong> past several decades it has grown <strong>in</strong>to an accomplished<br />

branch of applied ma<strong>the</strong>matics that heavily employs probability <strong>the</strong>ory <strong>and</strong> <strong>the</strong> <strong>the</strong>ory<br />

of cont<strong>in</strong>uous-time stochastic processes. Thirdly, <strong>the</strong> problems of derivative pric<strong>in</strong>g,<br />

grow<strong>in</strong>g <strong>in</strong> complexity along with <strong>the</strong> dem<strong>and</strong>s of <strong>the</strong> <strong>market</strong> players, <strong>in</strong>creas<strong>in</strong>gly rely<br />

on efficient numerical approximation <strong>and</strong> simulation <strong>method</strong>s. Last but not least, high<br />

performance <strong>and</strong> parallel comput<strong>in</strong>g environments have become an <strong>in</strong>dispensable tool<br />

for <strong>the</strong> process of <strong>valuation</strong> <strong>and</strong> risk-analysis of complex multi-asset f<strong>in</strong>ancial <strong>in</strong>struments.<br />

The ma<strong>in</strong> focus of this <strong>the</strong>sis is on <strong>the</strong> family of <strong>in</strong>terest-rate derivatives, a very<br />

important class of products used to hedge exposure to <strong>in</strong>terest-rate risk. S<strong>in</strong>ce <strong>the</strong><br />

dem<strong>and</strong> for such <strong>in</strong>surance is generated by large f<strong>in</strong>ancial <strong>in</strong>stitutions (e.g. pension<br />

funds, <strong>in</strong>surance companies), banks <strong>and</strong> <strong>market</strong> makers sell<strong>in</strong>g <strong>in</strong>terest-rate derivatives<br />

have to deal with <strong>the</strong> risks of serious losses. The necessity of controll<strong>in</strong>g <strong>the</strong>se risks,<br />

<strong>and</strong> thus, f<strong>in</strong>d<strong>in</strong>g <strong>the</strong> correct value of a derivative <strong>in</strong>strument, formulates <strong>the</strong> motivation<br />

for <strong>the</strong> <strong>the</strong>ory beh<strong>in</strong>d Derivative F<strong>in</strong>ance. The second chapter of <strong>the</strong> <strong>the</strong>sis is meant to<br />

provide a brief <strong>in</strong>troduction to <strong>the</strong> ma<strong>in</strong> concepts of <strong>option</strong> pric<strong>in</strong>g <strong>in</strong> general, as well<br />

as ma<strong>the</strong>matical <strong>the</strong>ory needed while operat<strong>in</strong>g <strong>in</strong> <strong>the</strong> economy of <strong>in</strong>terest rates.<br />

Both <strong>in</strong> academia <strong>and</strong> <strong>in</strong>dustry, a great deal of research is devoted to <strong>the</strong> development<br />

of realistic <strong>model</strong>s for <strong>in</strong>terest rate dynamics 1 <strong>and</strong> frameworks for pric<strong>in</strong>g <strong>and</strong><br />

hedg<strong>in</strong>g <strong>option</strong>s. Among those discussed <strong>in</strong> <strong>the</strong> literature, one can dist<strong>in</strong>guish two fundamental<br />

approaches: spot rate <strong>model</strong>s <strong>and</strong> <strong>market</strong> rate <strong>model</strong>s. The former approach<br />

takes a fictional <strong>in</strong>stantaneous spot rate (alive for an <strong>in</strong>f<strong>in</strong>itesimal time-period dt) as a<br />

basis for <strong>model</strong>l<strong>in</strong>g of <strong>the</strong> term structure 2 . The rest of <strong>the</strong> rates are obta<strong>in</strong>ed by <strong>in</strong>tegrat<strong>in</strong>g<br />

with respect to a spot <strong>in</strong>terest rate. From a ma<strong>the</strong>matical st<strong>and</strong>po<strong>in</strong>t, formulation<br />

<strong>and</strong> pric<strong>in</strong>g products <strong>in</strong> terms of this abstraction is quite convenient. However, calibra-<br />

1 of course, <strong>the</strong>se <strong>model</strong>s are be<strong>in</strong>g designed to simulate <strong>and</strong> not to forecast <strong>the</strong> future developments <strong>in</strong><br />

<strong>the</strong> <strong>in</strong>terest rate behaviour.<br />

2 <strong>the</strong> set of consequent <strong>in</strong>terest rates with <strong>in</strong>creas<strong>in</strong>g maturity<br />

1


tion of spot-rate <strong>model</strong> to <strong>the</strong> observed <strong>market</strong> data is far from straight-forward, s<strong>in</strong>ce<br />

such rates do not exist <strong>in</strong> practice [Pelsser, 2000].<br />

In <strong>the</strong> third chapter we turn our attention to <strong>market</strong> rate <strong>model</strong>s, namely <strong>the</strong> LIBOR<br />

Market Model (LMM), where LIBOR st<strong>and</strong>s for London Interbank Offered Rate. Published<br />

as [Brace et al., 1997], it has been widely accepted <strong>and</strong> became one of <strong>the</strong> most<br />

popular <strong>model</strong>s used <strong>in</strong> practice. In contrast to spot-rate <strong>model</strong>s, it takes <strong>the</strong> process<br />

for a real <strong>market</strong> rate as <strong>the</strong> basic <strong>model</strong>l<strong>in</strong>g unit. This fact simplifies fitt<strong>in</strong>g <strong>the</strong> <strong>model</strong><br />

to <strong>market</strong> parameters to a great extent. On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, <strong>the</strong> number of <strong>in</strong>terest rates<br />

underly<strong>in</strong>g <strong>the</strong> derivative has a direct impact on <strong>the</strong> dimensionality of <strong>the</strong> problem.<br />

In fact, without sophisticated extensions, e.g. [Pietersz et al., 2005], <strong>the</strong> number of<br />

dimensions is equal to <strong>the</strong> number of underly<strong>in</strong>g forward rates.<br />

S<strong>in</strong>ce <strong>the</strong> analytical <strong>valuation</strong> formulae are rarely available <strong>in</strong> a multi-dimensional<br />

sett<strong>in</strong>g, one needs to approximate <strong>the</strong> solution with a computational <strong>method</strong> of choice.<br />

The fourth chapter conta<strong>in</strong>s <strong>the</strong> discussion of Monte Carlo simulation (MC) <strong>and</strong> F<strong>in</strong>ite<br />

Difference Method (FDM) from <strong>the</strong> perspective of LMM <strong>valuation</strong>. As of this moment,<br />

<strong>valuation</strong> of <strong>in</strong>terest-rate derivatives lies <strong>in</strong> <strong>the</strong> doma<strong>in</strong> of Monte Carlo simulation,<br />

s<strong>in</strong>ce it is weakly dependent on <strong>the</strong> dimensionality of <strong>the</strong> problem <strong>and</strong> allows to price<br />

claims contigent on a large (up to 30) number of rates. Yet, as discussed <strong>in</strong> Chapter 4,<br />

FDM has an number of important advantages over MC <strong>and</strong> <strong>the</strong> ma<strong>in</strong> reason for it to be<br />

underfavoured is its ’curse of dimensionality’. The exponential growth <strong>in</strong> <strong>the</strong> number<br />

of mesh po<strong>in</strong>ts with <strong>the</strong> <strong>in</strong>crease <strong>in</strong> <strong>the</strong> dimensionality makes it unusable <strong>in</strong> more than<br />

three dimensions. The challenge of break<strong>in</strong>g <strong>the</strong> curse motivated researchers to seek<br />

for alternative discretisation <strong>method</strong>s. One such <strong>method</strong>, known as <strong>the</strong> Sparse Grid<br />

<strong>method</strong> was <strong>in</strong>troduced <strong>in</strong> [Zenger, 1991] <strong>and</strong> has been able to show promis<strong>in</strong>g results<br />

<strong>in</strong> application to high-dimensional problems <strong>in</strong> many areas of science. It is discussed<br />

<strong>in</strong> <strong>the</strong> fifth chapter of this text.<br />

The idea of application of <strong>sparse</strong> <strong>grid</strong> technique to <strong>the</strong> field of <strong>option</strong> pric<strong>in</strong>g is not<br />

new [Reis<strong>in</strong>ger, 2004]; nor is it such with respect to <strong>the</strong> derivatives <strong>in</strong> <strong>the</strong> LIBOR Market<br />

Model [Blackham, 2004]. The author of <strong>the</strong> latter reports price convergence of full<br />

<strong>and</strong> <strong>sparse</strong> <strong>grid</strong> solutions for products cont<strong>in</strong>gent on 2,3,4 <strong>and</strong> possibly 5 forward rates.<br />

The current master <strong>the</strong>sis was conceived as an extension of <strong>the</strong> work of [Blackham,<br />

2004] <strong>in</strong> lower (2D <strong>and</strong> 3D) dimensions, with <strong>the</strong> threefold objective for <strong>the</strong> practical<br />

part:<br />

1. Independent verification of <strong>the</strong> f<strong>in</strong>d<strong>in</strong>gs regard<strong>in</strong>g <strong>the</strong> <strong>option</strong> <strong>valuation</strong> for 2D<br />

<strong>and</strong> 3D cases of [Blackham, 2004] <strong>and</strong> exploration of a larger parameter space.<br />

2. Investigation of <strong>the</strong> behaviour for a discont<strong>in</strong>uous payoff <strong>in</strong> case of a <strong>sparse</strong> <strong>grid</strong><br />

<strong>method</strong>.<br />

3. A more rigorous study of <strong>the</strong> quality of <strong>the</strong> <strong>sparse</strong> ’Greeks’. Essential to risk<br />

management, ’Greeks’ are <strong>the</strong> quantities that show sensitivity of <strong>the</strong> computed<br />

price to <strong>the</strong> changes <strong>in</strong> <strong>the</strong> <strong>in</strong>put parameters.<br />

The results <strong>and</strong> due observations are presented <strong>in</strong> <strong>the</strong> f<strong>in</strong>al chapter, followed by <strong>the</strong><br />

summary of recommendations for future work.<br />

2


Chapter 2<br />

Asset Pric<strong>in</strong>g Overview<br />

The most important goal of Derivative F<strong>in</strong>ance is f<strong>in</strong>d<strong>in</strong>g <strong>the</strong> fair price of a cont<strong>in</strong>gent<br />

claim. In this chapter, we will provide a brief <strong>in</strong>troduction to discrete <strong>and</strong> cont<strong>in</strong>uous<br />

<strong>model</strong>s used <strong>in</strong> this area of applied ma<strong>the</strong>matics, def<strong>in</strong>e <strong>the</strong> notion of no-arbitrage <strong>market</strong>s,<br />

risk-neutral world <strong>and</strong> replicat<strong>in</strong>g portfolios. The last section is go<strong>in</strong>g to prepare<br />

<strong>the</strong> ground for fur<strong>the</strong>r discussion of LIBOR <strong>market</strong> <strong>model</strong> by putt<strong>in</strong>g <strong>option</strong> <strong>valuation</strong><br />

<strong>in</strong>to a more general framework of mart<strong>in</strong>gales <strong>and</strong> measures.<br />

2.1 Arbitrage <strong>and</strong> Risk-Neutral World<br />

Simple Market In order to <strong>in</strong>troduce several important concepts, we will start with<br />

a very simple sett<strong>in</strong>g. The <strong>market</strong> should consist of only two assets: zero-coupon 1<br />

cash bond B, <strong>and</strong> a stock S. The evolution of <strong>the</strong>se two assets can be observed at<br />

two discrete time po<strong>in</strong>ts: <strong>the</strong> current moment T 0 <strong>and</strong> <strong>the</strong> time of <strong>option</strong> maturity T 1 .<br />

As cont<strong>in</strong>uously-compounded <strong>in</strong>terest rates are assumed to be constant, <strong>the</strong>re is no<br />

uncerta<strong>in</strong>ty about <strong>the</strong> value of B at T 1 . The price of <strong>the</strong> stock, on <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, is<br />

subject to uncerta<strong>in</strong>ty <strong>and</strong> evolves to one of <strong>the</strong> two possible states: ”up” - <strong>the</strong> stock<br />

price <strong>in</strong>creases <strong>and</strong> ”down” - <strong>the</strong> stock price goes down. The <strong>market</strong>, described above,<br />

is known as a one-step b<strong>in</strong>ary <strong>model</strong>.<br />

Inspired by [E<strong>the</strong>rage, 2002], this view of <strong>the</strong> economy can be presented <strong>in</strong> <strong>the</strong><br />

matrix-vector form. Prices of N assets at T 0 are denoted by vector I; values of N<br />

assets be<strong>in</strong>g <strong>in</strong> n possible states at T 1 are <strong>in</strong>corporated <strong>in</strong>to matrix D. Specifically for<br />

our case, I <strong>and</strong> D take <strong>the</strong> follow<strong>in</strong>g form:<br />

I =<br />

(<br />

S0<br />

B 0<br />

)<br />

(<br />

up down<br />

)<br />

S0 u S<br />

D =<br />

0 d<br />

B 0 e rT B 0 e rT<br />

where r - is <strong>the</strong> risk-free rate, u <strong>and</strong> d coefficients correspond to <strong>the</strong> upward <strong>and</strong> downward<br />

price movements of S 0 with u > e rT > d.<br />

1 provid<strong>in</strong>g no <strong>in</strong>terim coupon payments.<br />

3


Before start<strong>in</strong>g to operate <strong>in</strong> this economy, we need to f<strong>in</strong>ish its setup by impos<strong>in</strong>g<br />

several assumptions:<br />

• short-sell<strong>in</strong>g 2 of securities is allowed;<br />

• <strong>the</strong>re are negligible or no transaction costs;<br />

• arbitrage opportunities are absent;<br />

While <strong>the</strong> first two assumptions are easy to perceive, <strong>the</strong> assumption of no-arbitrage<br />

requires our special attention <strong>and</strong> is go<strong>in</strong>g to have a significant impact on <strong>the</strong> way we<br />

see derivative <strong>valuation</strong>. Let’s def<strong>in</strong>e arbitrage.<br />

Arbitrage Let’s set 3 θ = (θ 0 , . . . , θ N ) t to be a portfolio of N assets, where θ i is<br />

<strong>the</strong> quantity of <strong>the</strong> i th asset, held <strong>in</strong> <strong>the</strong> portfolio. The products θ t·I <strong>and</strong> D t θ, are,<br />

<strong>the</strong>refore, <strong>the</strong> price of establish<strong>in</strong>g such a portfolio at time T 0 <strong>and</strong> its value at time T 1 ,<br />

respectively. Arbitrage can be loosely def<strong>in</strong>ed as a portfolio θ, whose cost θ t·I≤ 0<br />

at T 0 , <strong>and</strong> whose vector D t θ conta<strong>in</strong>s positive values at T 1 . In o<strong>the</strong>r words, one has a<br />

chance of collect<strong>in</strong>g positive payoffs without <strong>in</strong>itial <strong>in</strong>vestment, <strong>and</strong> no chance of los<strong>in</strong>g<br />

any money. The assumption of no-arbitrage is characteristic of efficient, liquid <strong>market</strong>s,<br />

where, <strong>in</strong> <strong>the</strong>ory, <strong>the</strong> time of arbitrage existence strives to zero. This happens because<br />

<strong>the</strong>re are many <strong>market</strong> participants <strong>in</strong> search of <strong>the</strong> underpriced <strong>and</strong> overpriced assets.<br />

Arbitrage is exploited <strong>the</strong> <strong>in</strong>stant it has been found, driv<strong>in</strong>g <strong>the</strong> prices of securities to<br />

<strong>the</strong>ir fair values, elim<strong>in</strong>at<strong>in</strong>g <strong>the</strong> possibilities of riskless profits <strong>and</strong> br<strong>in</strong>g<strong>in</strong>g <strong>the</strong> <strong>market</strong><br />

to <strong>the</strong> state of equilibrium.<br />

Pric<strong>in</strong>g by Arbitrage Consider a b<strong>in</strong>ary economy where such arbitrage opportunities<br />

do not last. What would be <strong>the</strong> correct price for an <strong>in</strong>strument, whose payoff is<br />

cont<strong>in</strong>gent on <strong>the</strong> assets compris<strong>in</strong>g <strong>the</strong> economy? Obviously, this value is strongly<br />

correlated with its f<strong>in</strong>al payoff. Let’s take, for example, pla<strong>in</strong>-vanilla call <strong>option</strong> c <strong>and</strong><br />

write c(S T1 ) for a payoff of c as a function of stock price at time T 1 . If <strong>the</strong> <strong>option</strong><br />

writer could establish a portfolio θ = (ζ, γ) t , such that D t θ = (c(S 0 u), c(S 0 d)) t ,<br />

she would completely replicate <strong>the</strong> payoff of c at T 1 , <strong>and</strong>, as a result, get its value<br />

θ t·I= c 0 . A claim, whose payoff can be replicated by such a portfolio, is said to be<br />

atta<strong>in</strong>able. Clearly, hav<strong>in</strong>g a no-arbitrage economy <strong>in</strong> m<strong>in</strong>d, our <strong>market</strong> <strong>model</strong> should<br />

not allow for existence of several replicat<strong>in</strong>g strategies that generate <strong>the</strong> same payoff<br />

pattern <strong>and</strong> have different <strong>in</strong>itial costs. The technique of determ<strong>in</strong><strong>in</strong>g <strong>the</strong> value of cont<strong>in</strong>gent<br />

claims by f<strong>in</strong>d<strong>in</strong>g <strong>the</strong> value of <strong>the</strong>ir replicat<strong>in</strong>g portfolio is known as pric<strong>in</strong>g<br />

by arbitrage. Writ<strong>in</strong>g χ = (c(S 0 u), c(S 0 d)) t for <strong>the</strong> c’s payoff vector, <strong>the</strong> cost of a<br />

strategy θ that replicates χ <strong>and</strong> consequently <strong>the</strong> <strong>option</strong>’s value at T 0 can be calculated<br />

as follow<strong>in</strong>g:<br />

θ = (D t ) −1 χ <strong>the</strong>refore c 0 = θ t · I<br />

2 To sell an asset short means borrow<strong>in</strong>g it from o<strong>the</strong>r <strong>in</strong>vestors hav<strong>in</strong>g guaranteed to return it on a specific<br />

date <strong>in</strong> <strong>the</strong> future plus some <strong>in</strong>terest payment<br />

3 a superscript t denotes a transpose operation of a matrix or a vector<br />

4


Risk-Neutral Pric<strong>in</strong>g Ano<strong>the</strong>r important implication of no-arbitrage assumption is<br />

reflected <strong>in</strong> <strong>the</strong> Arbitrage Theorem:<br />

Theorem 2.1.1. In a <strong>market</strong> with N assets <strong>and</strong> n states at time T 1 , <strong>the</strong>re is no arbitrage<br />

if <strong>and</strong> only if <strong>the</strong>re is a vector ψ, strictly positive <strong>in</strong> all coord<strong>in</strong>ates, such that I = Dψ<br />

For <strong>the</strong> proof, an <strong>in</strong>terested reader is referred to [E<strong>the</strong>rage, 2002].<br />

The vector ψ is known as a state price vector. Let us def<strong>in</strong>e |ψ| 1 = ∑ n<br />

i=1 ψ i. If each<br />

side of <strong>the</strong> relation I = Dψ is multiplied with a scalar:<br />

1<br />

I = Dψ 1<br />

(2.1)<br />

|ψ| 1<br />

|ψ| 1<br />

<strong>the</strong>n <strong>the</strong> product ψ 1<br />

|ψ| 1<br />

can be <strong>in</strong>terpreted as a probability distribution vector. Given<br />

such a vector, <strong>the</strong> right-h<strong>and</strong> side of (2.1) can be <strong>in</strong>terpreted as <strong>the</strong> expectation operator<br />

with respect to <strong>the</strong> probability distribution ψ 1<br />

|ψ| 1<br />

. As a result, <strong>the</strong> follow<strong>in</strong>g relation<br />

holds for each asset compris<strong>in</strong>g our economy:<br />

I i T 0<br />

= |ψ| 1 E(I i T ), i = 1 . . . N.<br />

Conveniently enough, it appears 4 that <strong>the</strong> coefficient |ψ| 1 is noth<strong>in</strong>g but a risk-free<br />

discount factor e −rT .<br />

For example, we can look at <strong>the</strong> expected rate of return of <strong>the</strong> stock sett<strong>in</strong>g <strong>the</strong><br />

probabilities of up <strong>and</strong> down movements to be <strong>the</strong> p <strong>and</strong> 1 − p, <strong>and</strong> given <strong>the</strong> previous<br />

result:<br />

E(S T1 ) = pS T0 u + (1 − p)S T0 d = S T0 e rT (2.2)<br />

which implies that:<br />

p = erT − d<br />

u − d<br />

As we can see, <strong>the</strong> expectation does not rely on real-world probabilities of up or<br />

down movements of <strong>the</strong> stock, just on <strong>the</strong> magnitude of change <strong>in</strong> <strong>the</strong> price, u <strong>and</strong> d.<br />

The same price will be observed whe<strong>the</strong>r P(u) = 0.5 or, for <strong>in</strong>stance, P(u) = 0.1.<br />

Such perspective on price formation correlates with <strong>the</strong> key notion of Efficient Markets<br />

Theory - Efficient Markets Hypo<strong>the</strong>sis (EMH). The so-called ’weak form’ of EMH<br />

states that <strong>the</strong> current price of <strong>the</strong> stock already reflects <strong>the</strong> <strong>in</strong>vestors’ expectations of<br />

possible price movements [Bodie et al., 2004]. Even though probabilistic distribution<br />

of stock price movements is not present <strong>in</strong> this <strong>valuation</strong> framework, we could treat p<br />

<strong>and</strong> 1 − p as pseudo-probabilities of future payoff value. With this result, we can conclude<br />

that given a probability distribution vector, time zero price of any asset def<strong>in</strong><strong>in</strong>g<br />

<strong>the</strong> <strong>market</strong> can be found as <strong>the</strong> expectation with respect to this probability measure,<br />

discounted at a risk-free rate. The world where <strong>the</strong> stock-prices grow on average at a<br />

risk-free rate we shall call “risk-neutral” <strong>and</strong> <strong>the</strong> normalized state price vector will be<br />

addressed as risk-neutral probability measure Q.<br />

4 This can be justified by pric<strong>in</strong>g a portfolio that replicates a zero-coupon discount bond, pay<strong>in</strong>g one unit<br />

of currency at time T 1 . E.g.:<br />

B 0 = B 0 e rT (ψ 0 + · · · + ψ n ) =⇒ |ψ| 1 = (ψ 0 + · · · + ψ n ) = e −rT .<br />

5


Def<strong>in</strong>ition 2.1.2. An economy is said to be complete if every derivative security is<br />

atta<strong>in</strong>able.<br />

By a very simple argument, a <strong>market</strong> will be complete if N ≥ n. If a number of<br />

possible states exceeds a number of <strong>market</strong>ed assets <strong>the</strong>n solv<strong>in</strong>g a system of equations<br />

will not give unique values for portfolio weights <strong>and</strong> arbitrage opportunities will<br />

arise. We shall end this section with <strong>the</strong> first, light-weight edition of <strong>the</strong> Fundamental<br />

Theorem of Asset Pric<strong>in</strong>g.<br />

Theorem 2.1.3. An economy is complete <strong>and</strong> free of arbitrage if <strong>and</strong> only if <strong>the</strong>re exists<br />

a unique equivalent probability measure.<br />

We will rely on <strong>the</strong> reader’s <strong>in</strong>tuition until this <strong>the</strong>orem is restated <strong>and</strong> explicated<br />

later <strong>in</strong> <strong>the</strong> chapter to fit <strong>in</strong>to <strong>the</strong> mart<strong>in</strong>gale <strong>valuation</strong> framework. Given a risk-neutral<br />

probability measure, <strong>the</strong> T 0 price of <strong>option</strong> c 0 <strong>in</strong> a no-arbitrage complete <strong>market</strong> equals:<br />

c 0 = e −r E Q [c 1 ] (2.3)<br />

2.2 From Discrete Time to Cont<strong>in</strong>uous Time<br />

A s<strong>in</strong>gle period <strong>model</strong> is, of course, very far from practical use. Never<strong>the</strong>less, it serves<br />

well as a build<strong>in</strong>g block for a more realistic discrete multi-period <strong>model</strong> of recomb<strong>in</strong>ant<br />

b<strong>in</strong>omial trees, as suggested <strong>in</strong> [Cox et al., 1979].<br />

Recomb<strong>in</strong>ant Trees<br />

Consider a better ref<strong>in</strong>ed timel<strong>in</strong>e with N timesteps<br />

t = t 0 < t 1 < . . . < t i < . . . < t N = T<br />

This allows us to monitor <strong>the</strong> stock price time series at a f<strong>in</strong>ite number of discrete<br />

time po<strong>in</strong>ts prior to maturity of an <strong>option</strong>. The stock price at time t i+1 is now S i u<br />

or S i d, where <strong>the</strong> same growth coefficients u <strong>and</strong> d apply to every time period <strong>in</strong> <strong>the</strong><br />

partition above. Therefore, <strong>the</strong> same price can be obta<strong>in</strong>ed by simple regroup<strong>in</strong>g of <strong>the</strong><br />

coefficients, e.g. S 4 = S 0 uudu = S 0 uduu. As one can see from Figure (2.1), <strong>the</strong><br />

result of all recomb<strong>in</strong>ations is <strong>the</strong> b<strong>in</strong>omial recomb<strong>in</strong>ant tree of possible stock price<br />

movements.<br />

S<strong>in</strong>ce <strong>the</strong> s<strong>in</strong>gle period <strong>model</strong> plays a role of a construction primitive <strong>in</strong> a b<strong>in</strong>omial<br />

recomb<strong>in</strong>ant world, we already know how to f<strong>in</strong>d a fair value of an <strong>option</strong>. It is noth<strong>in</strong>g<br />

but a recursive application of no-arbitrage techniques of <strong>the</strong> previous section by means<br />

of stepp<strong>in</strong>g backwards <strong>in</strong> <strong>the</strong> tree. Aga<strong>in</strong>, <strong>in</strong> <strong>the</strong> <strong>option</strong> pric<strong>in</strong>g context <strong>the</strong> probabilistic<br />

views of <strong>the</strong> <strong>market</strong> players are disregarded <strong>and</strong> expectations of <strong>the</strong> future <strong>option</strong><br />

payoffs are found under <strong>the</strong> unique equivalent risk-neutral measure.<br />

Replication <strong>in</strong> Discrete Time This chapter has started from pric<strong>in</strong>g a derivative <strong>in</strong>strument<br />

by construction of a replicat<strong>in</strong>g portfolio <strong>in</strong> a s<strong>in</strong>gle period <strong>model</strong>. In a b<strong>in</strong>omial<br />

tree, where <strong>the</strong> evolution of a stock price <strong>and</strong> a bond is a multi-step process, <strong>the</strong><br />

replication procedure has to be performed at each time step. This means <strong>the</strong> weights of<br />

6


S 0 u<br />

d<br />

u S 0<br />

S u 2<br />

0<br />

Su<br />

3<br />

0<br />

S du 2<br />

0<br />

4<br />

S0u<br />

S<br />

du 3<br />

0<br />

S0<br />

S0<br />

ud S 0<br />

2<br />

S0ud<br />

S u d<br />

2<br />

0<br />

2<br />

T 0<br />

T<br />

1<br />

S 0<br />

d S 0<br />

2<br />

S0d<br />

S ud 3<br />

0<br />

T<br />

0<br />

T<br />

1<br />

T<br />

2<br />

3<br />

S0d<br />

T3<br />

T4<br />

S d 4<br />

0<br />

Figure 2.1: From S<strong>in</strong>gle Period to Multi-Period<br />

assets <strong>in</strong> a replicat<strong>in</strong>g portfolio have to be adjusted accord<strong>in</strong>g to what is happen<strong>in</strong>g with<br />

<strong>the</strong> prices of <strong>the</strong> assets. Intuitively, a replication strategy would be mean<strong>in</strong>gful for <strong>the</strong><br />

writer of <strong>the</strong> claim only <strong>in</strong> case this strategy does not require any ma<strong>in</strong>ta<strong>in</strong><strong>in</strong>g costs, i.e.<br />

is self-f<strong>in</strong>anc<strong>in</strong>g. In o<strong>the</strong>r words, after spend<strong>in</strong>g a certa<strong>in</strong> amount of cash to establish<br />

positions <strong>in</strong> S 0 <strong>and</strong> B 0 , <strong>the</strong> replication procedure should <strong>in</strong>volve only chang<strong>in</strong>g <strong>the</strong>ir<br />

weights <strong>and</strong> exclude any additional cash <strong>in</strong>puts. Denote a vector of portfolio weights<br />

as θ = (ζ, γ) t . A one-step example of <strong>the</strong> self-f<strong>in</strong>anc<strong>in</strong>g property for a replicat<strong>in</strong>g<br />

portfolio V would be:<br />

V i = ζ i S i + γ i B i = ζ i+1 S i + γ i+1 B i (2.4)<br />

where ζ i+1 <strong>and</strong> γ i+1 are <strong>the</strong> quantities of S <strong>and</strong> B respectively, held over <strong>the</strong> period<br />

t ∈ [ i∆t, (i + 1)∆t ) . A change <strong>in</strong> portfolio value over one-step period gives:<br />

For an N-step recomb<strong>in</strong>ant tree:<br />

V i+1 − V i = ζ i+1 (S i+1 − S i ) + γ i+1 (B i+1 − B i ) (2.5)<br />

N−1<br />

∑<br />

N−1<br />

∑<br />

V N = V 0 + ζ i+1 (S i+1 − S i ) + γ i+1 (B i+1 − B i ) (2.6)<br />

i=0<br />

If we assume <strong>in</strong>terest rates to be zero, <strong>the</strong>n it can be easily shown that for V N B −1 =<br />

Ṽ N :<br />

N−1<br />

∑<br />

Ṽ N = Ṽ0 + ζ i+1 ( ˜S i+1 − ˜S i ) or ∆Ṽi = ζ i+1 ∆ ˜S i (2.7)<br />

i=0<br />

7<br />

i=0


where ˜S i is discounted stock price 5 . Detailed account for pric<strong>in</strong>g <strong>and</strong> hedg<strong>in</strong>g of <strong>option</strong>s<br />

<strong>in</strong> a recomb<strong>in</strong>ant tree can be found <strong>in</strong> such references as [Hull, 1999] <strong>and</strong> [E<strong>the</strong>rage,<br />

2002].<br />

Log-normal distribution of stock prices Up until this moment, we haven’t been<br />

concerned with <strong>the</strong> degree of realism for <strong>the</strong> stock price development <strong>model</strong>. This<br />

simplification has been <strong>in</strong>tentional for <strong>the</strong> sake of show<strong>in</strong>g <strong>the</strong> no-arbitrage pr<strong>in</strong>ciple <strong>in</strong><br />

<strong>the</strong> most <strong>in</strong>genuous sett<strong>in</strong>g. At this po<strong>in</strong>t we shall try to remedy this <strong>and</strong> <strong>in</strong>troduce <strong>the</strong><br />

more or less realistic process for <strong>the</strong> stock price evolution. There are three important<br />

facts about <strong>the</strong> process of stock prices that can help us address this issue:<br />

S<br />

• Stock price returns, T<br />

S0<br />

= ∑ N S i<br />

i=1 S i−1<br />

are assumed to be markovian r<strong>and</strong>om<br />

variables (<strong>in</strong>dependent from history, or, equivalently, EMH)<br />

• The expectation of a future stock price for a s<strong>in</strong>gle step can be computed with<br />

respect to a certa<strong>in</strong> probability measure (Q or P). Note, that <strong>the</strong> same measure<br />

applies to <strong>the</strong> whole sequence of price changes.<br />

• In <strong>the</strong> real world, <strong>the</strong> trades on stock <strong>market</strong>s occur almost cont<strong>in</strong>uously.<br />

Unlike <strong>the</strong> first two, cont<strong>in</strong>uity of trades has not been accounted for <strong>in</strong> our b<strong>in</strong>omial<br />

view of <strong>the</strong> <strong>market</strong> so far. This can be changed if we allow N → ∞, or likewise,<br />

let dt → 0, where dt is T N<br />

. These results taken toge<strong>the</strong>r empower <strong>the</strong> Central Limit<br />

Theorem to show us that stock price returns converge <strong>in</strong> distribution to <strong>the</strong> Gaussian<br />

d<br />

−→ N (·, ·) with some mean <strong>and</strong> variance. When returns are normally<br />

process, or S T<br />

S0<br />

distributed r<strong>and</strong>om variables, <strong>the</strong> stock prices can <strong>the</strong>oretically take negative values,<br />

which is clearly unacceptable for <strong>the</strong> asset price. How can we overcome this obstacle<br />

imposed by <strong>the</strong> nature of <strong>the</strong> Gaussian distribution?<br />

It must be obvious, that expected future values of <strong>the</strong> underly<strong>in</strong>g are determ<strong>in</strong>ed by<br />

<strong>the</strong> choice of <strong>the</strong> coefficients u <strong>and</strong> d, where u > e rdt > d. Let’s set u <strong>and</strong> d such that:<br />

u = e µdt+σ√ dt<br />

d = e µdt−σ√ dt<br />

The volatility parameter σ quantifies <strong>the</strong> amount of risk characterictic of a specific<br />

stock <strong>and</strong> is calculated to be <strong>the</strong> stock’s st<strong>and</strong>ard deviation from <strong>the</strong> expected<br />

return. The drift term µ, same for both up <strong>and</strong> down movement, quantifies <strong>the</strong> constant<br />

tendency of growth <strong>in</strong> stock, or expected return. Hav<strong>in</strong>g u <strong>and</strong> d as exponential<br />

functions of dt, we conveniently work with natural logarithms of u <strong>and</strong> d <strong>and</strong> denote<br />

R i = log(S i ) − log(S i−1 ) to be stock’s s<strong>in</strong>gle period log-return. The choice of parameters<br />

as exponential functions of dt should affect <strong>the</strong> distribution of <strong>the</strong> overall process.<br />

Tak<strong>in</strong>g <strong>the</strong> up <strong>and</strong> down movements to be equally possible, i.e. P(u) = P(d) = 0.5,<br />

we arrive at <strong>the</strong> very important result for <strong>option</strong> pric<strong>in</strong>g <strong>the</strong>ory:<br />

Under a probability measure P stock returns have a lognormal distribution:<br />

( ST<br />

)<br />

log ∼ N<br />

S 0<br />

(<br />

µT, σ 2 T<br />

)<br />

(2.8)<br />

5 The discounted process ∆Ṽi can, <strong>in</strong> fact, be characterized as a discrete mart<strong>in</strong>gale process. We shall<br />

skip discussion of discrete mart<strong>in</strong>gales to <strong>in</strong>troduce <strong>the</strong>m <strong>in</strong> cont<strong>in</strong>uous time later <strong>in</strong> <strong>the</strong> text.<br />

8


or<br />

log(S T )<br />

d<br />

−→ log(S 0 ) + µT + σ √ T Z. (2.9)<br />

where Z ∼ N (0, 1) <strong>and</strong> our expected return on <strong>the</strong> stock µ be<strong>in</strong>g a mean of a s<strong>in</strong>gle<br />

period log-return 6 .<br />

Among numerous properties of lognormally distributed r<strong>and</strong>om variables, <strong>the</strong>re is<br />

one we have been look<strong>in</strong>g for: <strong>the</strong>y cannot take negative values. This type of distribution<br />

is believed to come close <strong>in</strong> attempt to mimic <strong>the</strong> real-world processes <strong>and</strong> is<br />

widely used <strong>in</strong> <strong>model</strong>l<strong>in</strong>g of different underly<strong>in</strong>g assets.<br />

Brownian Motion Due to <strong>the</strong> fact that log-returns, as an algebraic transformation<br />

of stock returns, are <strong>in</strong>dependent, identically distributed markovian r<strong>and</strong>om variables,<br />

<strong>the</strong> price change from time T 0 to time T N can be thought of as a sequence of N such<br />

variables or a stochastic process with notation {R t } N 1 . If we let dt → 0, {R t } t≥1 is a<br />

cont<strong>in</strong>uous-time stochastic process. As <strong>in</strong> [E<strong>the</strong>rage, 2002], consider a specific type of<br />

a cont<strong>in</strong>uous stochastic process:<br />

Def<strong>in</strong>ition 2.2.1. A real-valued stochastic process {W t } t≥1 is a st<strong>and</strong>ard P-Brownian<br />

motion (or a P-Wiener process) if given a probability measure P,<br />

1. for each s ≥ 0 <strong>and</strong> t > 0 <strong>the</strong> r<strong>and</strong>om variable W t+s − W s ∼ N (0, t − s)<br />

2. for each n ≥ 1 <strong>and</strong> any times 0 ≤ t 0 ≤ t 1 · · · ≤ t n , <strong>the</strong> r<strong>and</strong>om variables<br />

{W tr − W tr−1 } are <strong>in</strong>dependent.<br />

3. W 0 = 0. This property may be relaxed, as <strong>the</strong> process x + W t is a st<strong>and</strong>ard<br />

Brownian motion with W 0 = x, where x is a real constant.<br />

4. W t is cont<strong>in</strong>uous <strong>in</strong> t ≥ 0<br />

Given this def<strong>in</strong>ition, we can construct various stochastic processes Y t , that are<br />

functions of st<strong>and</strong>ard Brownian motion W <strong>and</strong> share its characteristic properties. For<br />

example:<br />

• Y t = Y 0 + σW t is a Brownian motion process with variance σ 2 t.<br />

• Y t = Y 0 + νt + σW t is a Brownian motion with drift. In o<strong>the</strong>r words, Y t ∼<br />

N (Y 0 + νt, σ 2 t).<br />

• Y t = Y 0 exp(νt + σW t ) is a geometric Brownian motion. A useful result that<br />

will be needed later is :<br />

E[e νt+σW ] = e (ν+ 1 2 σ2 )t<br />

(2.10)<br />

One can tell that our process for log-returns (2.9) is a stochastic process that is identical<br />

to Brownian motion with a drift 7 . On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, <strong>the</strong> process for stock returns is a<br />

geometric Brownian motion, tak<strong>in</strong>g strictly positive values with constant variance.<br />

)<br />

S t = S 0 exp<br />

(νt + σW t (2.11)<br />

6 Here we make use of <strong>the</strong> fact that <strong>the</strong> sum of identically distributed, <strong>in</strong>dependent r<strong>and</strong>om variables is a<br />

r<strong>and</strong>om variable for which <strong>the</strong> mean is <strong>the</strong> sum of <strong>the</strong> means <strong>and</strong> <strong>the</strong> variance is <strong>the</strong> sum of <strong>the</strong> variances.<br />

7 dW = (W t+dt − W t ) ∼ N (0, dt), accord<strong>in</strong>g to properties 1 <strong>and</strong> 2 of Brownian Motion<br />

9


A glimpse of stochastic <strong>in</strong>tegrals As we have moved to cont<strong>in</strong>uous world, <strong>the</strong> discrete<br />

sums <strong>and</strong> differences of equations like (2.6) turn <strong>in</strong>to <strong>in</strong>tegrals <strong>and</strong> differentials:<br />

∫ T ∫ T<br />

V T = V 0 + ζ t dS t + γ t dB t (2.12)<br />

0<br />

0<br />

Computation ∫ T<br />

0 γ tdB t means tak<strong>in</strong>g a Riemann-Stieltjes <strong>in</strong>tegral with respect to a<br />

smooth, differentiable function B t = e rt . On <strong>the</strong> contrary, f<strong>in</strong>d<strong>in</strong>g ∫ T<br />

0 ζ tdS t requires<br />

<strong>in</strong>tegration with respect to geometric Brownian motion <strong>and</strong> is beyond <strong>the</strong> reach of ord<strong>in</strong>ary<br />

“Newtonian” calculus. The paths of Brownian motion, despite <strong>the</strong>ir cont<strong>in</strong>uity,<br />

are not differentiable. Computation of <strong>in</strong>tegrals w.r.t. dW , also known as Itô <strong>in</strong>tegrals,<br />

lies <strong>in</strong> <strong>the</strong> doma<strong>in</strong> of Stochastic Calculus <strong>and</strong> requires some <strong>the</strong>oretical material this<br />

<strong>the</strong>sis was not meant to provide. Never<strong>the</strong>less, some most important properties <strong>and</strong><br />

results will be stated without proofs:<br />

Itô’s Processes <strong>and</strong> Itô’s Lemma We shall postpone <strong>in</strong>terpretation of (2.12) till next<br />

section. For now, let’s consider a different process:<br />

∫ T<br />

∫ T<br />

X T = X 0 + µ(t, X)dt + σ(t, X)dW t (2.13)<br />

0<br />

0<br />

or, equivaletely, <strong>in</strong> differential form:<br />

dX t = µ(t, X)dt + σ(t, X)dW t (2.14)<br />

Processes that are described by equations like (2.13) <strong>and</strong> (2.14), <strong>in</strong>volv<strong>in</strong>g Itô <strong>and</strong> Riemann<br />

<strong>in</strong>tegrals are refered to as Itô processes. Here µ(t, X) is a drift of process <strong>and</strong><br />

σ(t, X) is <strong>the</strong> volatility. The characteristic feature of an Itô process is <strong>the</strong> fact that<br />

<strong>the</strong> drift <strong>and</strong> volatility terms of (2.13) appear to be functions of time <strong>and</strong> <strong>the</strong> process<br />

itself. [Hull, 1999] provides clear <strong>in</strong>tuition why such dependancy must be <strong>the</strong> case for<br />

<strong>the</strong> process of <strong>the</strong> stock price. He argues that <strong>the</strong> expected return per unit of time, µ,<br />

should be <strong>the</strong> same regardless of <strong>the</strong> current stock value. The same argument is used<br />

for <strong>the</strong> volatility parameter - <strong>the</strong> sw<strong>in</strong>gs <strong>in</strong> <strong>the</strong> asset’s development are proportional to<br />

its price. This vision of <strong>the</strong> <strong>market</strong> transforms (2.14) <strong>in</strong>to a more specific form :<br />

dX t = µX t dt + σX t dW t (2.15)<br />

Here µ(t, X t ) = µX t <strong>and</strong> σ(t, X t ) = σX t . The Itô Process (2.15) is widely used as<br />

a <strong>model</strong> for stock price evolution <strong>and</strong> represents a Stochastic Differential Equation, or<br />

SDE. A very useful tool from Stochastic Calculus, that can sometimes be employed to<br />

obta<strong>in</strong> solutions to SDE’s such as (2.15) is Itô’s Lemma. Itô’s Lemma can be thought<br />

of as an extension of Taylor approximation formula <strong>in</strong>to stochastic world. More specifically,<br />

it states that if <strong>the</strong> sequence of r<strong>and</strong>om variables follows a Itô’s process X t with<br />

drift ν x <strong>and</strong> volatility σ t , <strong>the</strong>n <strong>the</strong> function of both time <strong>and</strong> this process f(t, X t ) is an<br />

Itô’s process as well. One dimensional version follows:<br />

df t,Xt = ∂f(t, X t)<br />

dt + ∂f(t, X t)<br />

dX t + 1 ∂t ∂X t 2<br />

10<br />

∂ 2 f(t, X t )<br />

∂Xt<br />

2 σt 2 dt (2.16)


There is an important difference between Itô’s formula <strong>and</strong> its determ<strong>in</strong>istic counterpart.<br />

The laws of stochastic calculus 8 stipulate that dWt<br />

2 = σt 2 dt, where W t is a<br />

Brownian motion. As a result, <strong>the</strong> second derivative term with respect to a stochastic<br />

factor is of <strong>the</strong> same order of accuracy as ∂f(t,X t)<br />

∂t<br />

<strong>and</strong> cannot be neglected.<br />

As a demostration of Itô’s lemma, let’s try to acquire <strong>the</strong> process for geometric<br />

Brownian motion as given by X t = W t <strong>and</strong> f(t, W t ) = S t = S 0 exp ( )<br />

νt + σW t ,<br />

which is also (2.11). The partial derivatives of S t with respect to time <strong>and</strong> Brownian<br />

motion are:<br />

<strong>and</strong> <strong>the</strong>refore:<br />

∂f(t, X t )<br />

∂t<br />

∂f(t, X t )<br />

∂ 2 f(t, X t )<br />

= νS t = σS t<br />

∂X t ∂Xt<br />

2 = σ 2 S t dt<br />

dS t = (ν + 1 2 σ2 )S t dt + σS t dW<br />

}{{} t<br />

} {{ }<br />

σ(t,S t )<br />

µ(t,S t)<br />

Thus, <strong>the</strong> process for prices given by geometric Brownian motion is an Itô process. We<br />

shall call ν + 1 2 σ2 - an <strong>in</strong>stantaneous expected return <strong>and</strong> ν - an overall expected return.<br />

Black-Scholes-Merton Model Hav<strong>in</strong>g accepted a log-normal distribution for <strong>the</strong><br />

cont<strong>in</strong>uous process of an underly<strong>in</strong>g asset, we have eventually prepared ourselves for<br />

discussion of Black-Scholes-Merton approach to <strong>option</strong> <strong>valuation</strong>. In <strong>the</strong>ir sem<strong>in</strong>al paper<br />

[Black <strong>and</strong> Scholes, 1973], Black <strong>and</strong> Scholes have suggested a <strong>model</strong> that dwells<br />

on several assumptions:<br />

• Trad<strong>in</strong>g proceeds cont<strong>in</strong>uously;<br />

• Stock prices follow a log-normal distribution, <strong>the</strong> process described by (2.15);<br />

• Absense of arbitrage opportunities;<br />

• No rebalanc<strong>in</strong>g costs;<br />

• Limitless lend<strong>in</strong>g <strong>and</strong> borrow<strong>in</strong>g is allowed;<br />

Hav<strong>in</strong>g imposed this sett<strong>in</strong>g, Black <strong>and</strong> Scholes argued that a short position <strong>in</strong> a derivative<br />

c t can be made riskless <strong>in</strong> case it is a part of a portfolio Π t such that:<br />

Π t = −c t + ∆ t S t (2.17)<br />

Both c t <strong>and</strong> S t depend on <strong>the</strong> same source of uncerta<strong>in</strong>ty, which is a Brownian Motion<br />

term W t of <strong>the</strong> stock price process. Therefore if Π t is constructed by tak<strong>in</strong>g reverse positions<br />

<strong>in</strong> a derivative <strong>and</strong> a certa<strong>in</strong> amount ∆ t of its underly<strong>in</strong>g, <strong>the</strong> Brownian motion<br />

terms cancel, elim<strong>in</strong>at<strong>in</strong>g <strong>the</strong> risk. Π t is called a Hedge portfolio. Accord<strong>in</strong>g to our<br />

no-arbitrage assumption, a position <strong>in</strong> a porfolio that has no risk must be of <strong>the</strong> same<br />

nature as a position <strong>in</strong> a risk-free cash bond B t , i.e. earn a risk-free <strong>in</strong>terest rate dur<strong>in</strong>g<br />

its lifetime:<br />

dΠ t = rΠ t dt<br />

8 Namely, <strong>the</strong> quadric variation result. See [E<strong>the</strong>rage, 2002].<br />

11


Suppose this is not true <strong>and</strong> Π t earns more or less than prescribed by no-arbitrage. Then<br />

we would have two different prices for a risk-free <strong>in</strong>vestment, driv<strong>in</strong>g <strong>market</strong> players<br />

to buy underpriced <strong>and</strong> sell overpriced positions, which naturally enforces <strong>the</strong> laws of<br />

supply <strong>and</strong> dem<strong>and</strong> br<strong>in</strong>g <strong>the</strong> <strong>market</strong> back to equilibrium. The second step towards <strong>the</strong><br />

price of an <strong>option</strong> <strong>in</strong> a cont<strong>in</strong>uous world would be f<strong>in</strong>d<strong>in</strong>g <strong>the</strong> weight ∆ t such, that <strong>the</strong><br />

value of <strong>the</strong> hedge porfolio would be exactly zero. A player with a short position <strong>in</strong><br />

an <strong>option</strong> would <strong>the</strong>n be able to compensate <strong>the</strong> claim aga<strong>in</strong>st him by sell<strong>in</strong>g his stock<br />

hold<strong>in</strong>gs. This sounds like exactly what we did <strong>in</strong> discrete time - a replicat<strong>in</strong>g strategy.<br />

The only th<strong>in</strong>g miss<strong>in</strong>g is a self-f<strong>in</strong>anc<strong>in</strong>g property of <strong>the</strong> portfolio. In discrete time, <strong>the</strong><br />

changes <strong>in</strong> ∆ t were f<strong>in</strong>anced by a ψ t amount of cash. The question is: if we add a cash<br />

bond hold<strong>in</strong>g to our hedge portfolio, will it safeguard our cont<strong>in</strong>uous-time replicat<strong>in</strong>g<br />

strategy from <strong>the</strong> need of extra money <strong>in</strong>put? Our <strong>in</strong>itial assumptions give a positive<br />

answer.<br />

Π 0 = −ψ t B 0<br />

Notation put <strong>in</strong>to words, establish<strong>in</strong>g a hedge portfolio, which replicates an <strong>option</strong> sold,<br />

requires additional <strong>in</strong>vestment <strong>in</strong> cash.<br />

Rearrang<strong>in</strong>g:<br />

dΠ 0 = d(−c 0 + ∆ t S t ) = −rψ t B 0 dt<br />

dc t = ∆ t dS t + r(c t − ∆ t S t )dt (2.18)<br />

Subscripts t should rem<strong>in</strong>d us that <strong>the</strong> hold<strong>in</strong>gs <strong>in</strong> assets ∆ t <strong>and</strong> ψ t change with time<br />

just like <strong>the</strong> prices of assets <strong>the</strong>mselves.<br />

On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, we know that a derivative c has a payoff that depends on time<br />

<strong>and</strong> a stock price S t . We can write this as c t = F (t, S t ). What we also know is that<br />

a process for <strong>the</strong> stock price evolution is be<strong>in</strong>g <strong>model</strong>led as an Itô process. Therefore<br />

we can apply Itô’s lemma to show how development <strong>in</strong> c t is expressed <strong>in</strong> terms of S t .<br />

Tak<strong>in</strong>g first derivative with respect to time <strong>and</strong> derivatives with respect to <strong>the</strong> stock<br />

price, we f<strong>in</strong>d<br />

dc t = dF (t, S t ) = ∂F (t, S t)<br />

∂t<br />

+ ∂F (t, S t)<br />

dS t + 1 ∂S t 2 σ2 t St<br />

2<br />

∂ 2 F (t, S t )<br />

∂St<br />

2 dt (2.19)<br />

Equat<strong>in</strong>g (2.18) <strong>and</strong> (2.19) <strong>and</strong> rearrang<strong>in</strong>g <strong>the</strong> terms we arrive at Black-Scholes Partial<br />

Differential Equation (PDE):<br />

∂F (t, S t )<br />

∂t<br />

+ ∂F (t, S t)<br />

∂S t<br />

rS t + 1 2<br />

∂ 2 F (t, S t )<br />

∂St<br />

2 σt 2 St 2 = rF (t, S t ) (2.20)<br />

This PDE shows us, that a derivative c t = F (t, S t ), while evolv<strong>in</strong>g <strong>in</strong> time <strong>and</strong> its<br />

underly<strong>in</strong>g asset, earns a risk-free rate. As already observed, risk preferences of <strong>in</strong>vestors<br />

do not serve as a parameter required for f<strong>in</strong>d<strong>in</strong>g <strong>the</strong> fair value of a claim. In<br />

o<strong>the</strong>r words, <strong>the</strong> assumption of a risk-neutral world, where <strong>in</strong>vestors do not require<br />

compensations for tak<strong>in</strong>g risks, is valid <strong>in</strong> cont<strong>in</strong>uous-time as well.<br />

12


Delta. In order to set up a riskfree hedge, as <strong>in</strong> (2.17), one needs to have <strong>the</strong> knowledge<br />

of <strong>the</strong> process for ∆ t . Recall, that ∆ t shows how much of <strong>the</strong> underly<strong>in</strong>g has to<br />

be held at a timepo<strong>in</strong>t t as a part of <strong>the</strong> hedge portfolio. On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, ∆ t can<br />

be <strong>in</strong>terpreted as <strong>the</strong> sensitivity of <strong>the</strong> <strong>option</strong>’s price to <strong>the</strong> movement <strong>in</strong> its underly<strong>in</strong>g<br />

<strong>and</strong>, <strong>the</strong>refore, is equivalent to <strong>the</strong> first derivative of <strong>the</strong> <strong>option</strong> value with respect to<br />

<strong>the</strong> price shifts <strong>in</strong> <strong>the</strong> asset. Note, that Black-Scholes PDE, given by (2.20), implicitly<br />

provides us with such <strong>in</strong>formation.<br />

Consider<strong>in</strong>g its utmost importance for replication of a claim, ∆ t will reappear <strong>in</strong><br />

<strong>the</strong> focus of our attention later <strong>in</strong> this text.<br />

2.3 Pric<strong>in</strong>g with Mart<strong>in</strong>gales <strong>and</strong> Measures<br />

Let’s try to fit <strong>the</strong> ma<strong>in</strong> ideas of derivative pric<strong>in</strong>g developed so far <strong>in</strong>to a couple of<br />

sentences. The risk of sell<strong>in</strong>g a derivative security can be hedged away by establish<strong>in</strong>g<br />

a replicat<strong>in</strong>g, self-f<strong>in</strong>anc<strong>in</strong>g portfolio of <strong>market</strong> assets. The cost of sett<strong>in</strong>g up a strategy<br />

that elim<strong>in</strong>ates <strong>the</strong> risk for a short side is, by no-arbitrage, <strong>the</strong> only correct price for<br />

a product be<strong>in</strong>g replicated. Accord<strong>in</strong>g to no-arbitrage, any riskless asset or a porfolio<br />

must earn a risk-free rate.<br />

We have just seen that an <strong>option</strong> value can satisfy a partial differential equation,<br />

<strong>the</strong>refore it can be approximated numerically with F<strong>in</strong>ite Difference or F<strong>in</strong>ite Volume<br />

Methods. Ano<strong>the</strong>r approach that has been discussed is risk-neutral <strong>valuation</strong>. It <strong>in</strong>volves<br />

tak<strong>in</strong>g expectations of a future payoff with respect to a unique risk-neutral probability<br />

measure, possibly result<strong>in</strong>g <strong>in</strong> a closed form expression.<br />

Complement<strong>in</strong>g each o<strong>the</strong>r, both approaches are used <strong>in</strong>tensively <strong>and</strong> ultimately<br />

arrive at <strong>the</strong> same result. In this section we will <strong>in</strong>troduce some ma<strong>the</strong>matical tools<br />

that are employed by risk-neutral <strong>valuation</strong> <strong>in</strong> cont<strong>in</strong>uous time, namely, mart<strong>in</strong>gale<br />

processes <strong>and</strong> <strong>the</strong> change of measure result.<br />

Some def<strong>in</strong>itions<br />

• Numeraire. A Numeraire is any asset B, whose price is strictly positive at any<br />

moment <strong>in</strong> time. For example, S t<br />

B t<br />

is a price of a discounted stock price S t . The<br />

price of <strong>the</strong> cash bond B t serves as a numeraire for <strong>the</strong> stock price.<br />

• Mart<strong>in</strong>gales. A mart<strong>in</strong>gale is a F t -adapted stochastic process X t such that under<br />

a certa<strong>in</strong> measure, e.g. Q:<br />

E Q = [ X t |F t<br />

]<br />

= Xs , ∀s < t<br />

In words, Q-expectation of <strong>the</strong> future value of <strong>the</strong> process is always a current<br />

value of <strong>the</strong> process. As time evolves, <strong>the</strong> process does not drift away from<br />

its start<strong>in</strong>g po<strong>in</strong>t, but fluctuates <strong>in</strong> its neighbourhood. Aga<strong>in</strong> we have already<br />

encountered such processes previously <strong>in</strong> <strong>the</strong> text. The process for discounted<br />

stock price e rt S t is a mart<strong>in</strong>gale under <strong>the</strong> risk-neutral measure, see (2.2).<br />

13


Change of Measure In order to make use of <strong>the</strong> risk-neutral <strong>valuation</strong> pr<strong>in</strong>ciple, we<br />

need to take expectations under Q. In case of one-step discrete sett<strong>in</strong>g we were able<br />

to f<strong>in</strong>d <strong>the</strong> probability weights by solv<strong>in</strong>g for a price-state vector. Needless to say, <strong>in</strong><br />

cont<strong>in</strong>uous time <strong>the</strong> same approach is hardly applicable. Yet, after we’ve taken <strong>the</strong> time<br />

step to <strong>the</strong> limit, <strong>the</strong> probability vectors become probability density functions. Without<br />

go<strong>in</strong>g much <strong>in</strong>to detail, we’ll state that chang<strong>in</strong>g <strong>the</strong> measure amounts to chang<strong>in</strong>g <strong>the</strong><br />

probability density function. Let’s sketch how expectation under one measure can be<br />

taken with respect to ano<strong>the</strong>r measure:<br />

where<br />

E P[ ∫<br />

] ∞<br />

X t |F t = X t f P (z)dz =<br />

−∞<br />

∫ ∞<br />

−∞<br />

dQ<br />

dP = f Q (z)<br />

f P (z)<br />

X t<br />

dQ<br />

dP f P (z)dz = E Q[ X t |F t<br />

]<br />

(2.21)<br />

is simply a ratio of density functions or, <strong>in</strong> f<strong>in</strong>ancial term<strong>in</strong>ology, a Radon-Nikodym<br />

derivative. It can be easily shown that:<br />

( ∫<br />

dQ<br />

t<br />

dP = exp κ s dWs P − 1 ∫ t<br />

)<br />

κ 2<br />

2<br />

sds<br />

(2.22)<br />

0<br />

is a geometric brownian motion with drift − 1 2 κ2 s <strong>and</strong> volatility κ s . Before elaborat<strong>in</strong>g<br />

on <strong>the</strong> mean<strong>in</strong>g of κ s term, we provide cont<strong>in</strong>uous time-version of Arbitrage <strong>the</strong>orem,<br />

as promised:<br />

Theorem 2.3.1. A cont<strong>in</strong>uous, complete <strong>market</strong> is free of arbitrage if <strong>and</strong> only if for<br />

every choice of numeraire <strong>the</strong>re exists a unique equivalent mart<strong>in</strong>gale measure.<br />

Restrict<strong>in</strong>g our choice of numeraire to a cash bond B t , this <strong>the</strong>orem states that<br />

<strong>the</strong>re exists only one measure under which our discounted stock process Bt<br />

−1 S t will<br />

be a mart<strong>in</strong>gale process. This measure is our risk-neutral measure Q. Ano<strong>the</strong>r very<br />

important <strong>the</strong>orem that is go<strong>in</strong>g to be crucial to discussion of <strong>the</strong> next chapter is Girsanov’s<br />

Theorem.<br />

Girsanov’s Theorem We’ll start with a simple example. Let’s denote ˜S t = Bt −1 S t ,<br />

i.e. a stock price with B t as a numeraire. By application of Itô’s lemma, <strong>the</strong> Itô process<br />

for ˜S t is:<br />

d ˜S t = (µ − r) ˜S t dt + σ ˜S t dWt<br />

P<br />

The equivalent mart<strong>in</strong>gale <strong>the</strong>orem claims that arbitrage is not feasible if <strong>the</strong>re exists a<br />

unique mart<strong>in</strong>gale measure for ˜S t is a mart<strong>in</strong>gale process. Rearrang<strong>in</strong>g <strong>the</strong> terms:<br />

( )<br />

d ˜S t = σ ˜S µ − r<br />

t<br />

σ<br />

dt + dW t<br />

P = σ ˜S t (κdt + dWt P ) = σ ˜S t dXt P (2.23)<br />

Here, <strong>the</strong> result<strong>in</strong>g process appears to be a mart<strong>in</strong>gale with respect to a new process<br />

dX P t . Of course, this appearance is only superficial as required mart<strong>in</strong>gale behaviour<br />

is ru<strong>in</strong>ed by <strong>the</strong> drift κ <strong>in</strong> <strong>the</strong> Brownian motion dX P t . So, we wouldn’t have made any<br />

progress without help from a so-called Girsanov’s Theorem:<br />

0<br />

14


Theorem 2.3.2. Let W P (t) = (W1 P (t), W2 P (t), . . . , Wd P (t)), t ≥ 0 be a d-dimensional<br />

vector of correlated Brownian motions under some measure P, with correlation values<br />

given by matrix ρ. If κ(t) = (κ 1 (t), κ 2 (t), . . . , κ d (t)), is a vector of pre-visible<br />

stochastic processes adapted to <strong>the</strong> same filtration F t , with<br />

[ ( ∫ 1 T<br />

)]<br />

E P exp κ(s) 2 ds < ∞ (2.24)<br />

2 0<br />

<strong>and</strong> Y t def<strong>in</strong>ed as<br />

( ∫ t<br />

Y t = exp κ s dWs P − 1<br />

0<br />

2<br />

∫ t<br />

0<br />

)<br />

κ 2 sds<br />

<strong>the</strong>n Y t is <strong>the</strong> Radon-Nikodym derivative dQ<br />

dP<br />

, <strong>and</strong> under <strong>the</strong> measure Q <strong>the</strong> process<br />

is also a d-dimensional Brownian motion.<br />

(2.25)<br />

dW Q (t) = dW P (t) + κ(t)ρ dt, t ≥ 0 (2.26)<br />

Shak<strong>in</strong>g off notation 9 , it simply states that by chang<strong>in</strong>g <strong>the</strong> measure to some measure<br />

Q, we can make <strong>the</strong> drift <strong>in</strong> dX P disappear. In o<strong>the</strong>r words, dX Q becomes a<br />

st<strong>and</strong>ard Brownian motion <strong>and</strong>, consequently, ˜S t is a true mart<strong>in</strong>gale. It is worth not<strong>in</strong>g,<br />

that <strong>the</strong> quantity µ−r<br />

σ<br />

has an <strong>in</strong>dependent <strong>in</strong>terpretation <strong>in</strong> CAPM. Our drift κ is<br />

commonly known as <strong>market</strong> price of risk. It shows <strong>the</strong> extent to which <strong>the</strong> volatility <strong>in</strong><br />

<strong>the</strong> asset is compensated by <strong>the</strong> excess return. From this prospective one can conclude<br />

that elim<strong>in</strong>at<strong>in</strong>g <strong>the</strong> drift, or mak<strong>in</strong>g <strong>the</strong> <strong>market</strong> price of risk κ zero, we f<strong>in</strong>d ourselves<br />

<strong>in</strong> <strong>the</strong> risk-neutral world where risk requires no compensation.<br />

Before we w<strong>in</strong>d up our discussion of prelim<strong>in</strong>aries, <strong>the</strong>re is one more result that we<br />

need to address for this chapter to be more or less complete.<br />

Brownian Mart<strong>in</strong>gale Representation Theorem Several pages ago, we have encountered<br />

a transformation of a discrete <strong>in</strong>tegral (2.6) to a cont<strong>in</strong>uous form (2.12)<br />

without provid<strong>in</strong>g any fur<strong>the</strong>r explanation. Now, after becom<strong>in</strong>g familiar with a notion<br />

of a mart<strong>in</strong>gale process, we can loosely state <strong>the</strong> Brownian Mart<strong>in</strong>gale Representation<br />

<strong>the</strong>orem, which proves to be an <strong>in</strong>dispensable tool <strong>in</strong> construction of self-f<strong>in</strong>anc<strong>in</strong>g<br />

replicat<strong>in</strong>g portfolios <strong>in</strong> cont<strong>in</strong>uous time.<br />

Theorem 2.3.3. Let W t be a Q-mart<strong>in</strong>gale process w.r.t. a filtration F t <strong>and</strong> with<br />

bounded variation E[Wt 2 ] < ∞. Then any Q-mart<strong>in</strong>gale M t w.r.t. F t <strong>and</strong> E[Mt 2 ] <<br />

∞, can be written as:<br />

where ζ s is an F t -adapted process.<br />

M t = M 0 +<br />

∫ t<br />

0<br />

ζ s dW t<br />

In o<strong>the</strong>r words, any two mart<strong>in</strong>gales with respect to <strong>the</strong> same filtration are scaled<br />

versions of each o<strong>the</strong>r. As an illustration of this result, let’s th<strong>in</strong>k that a dem<strong>and</strong> exists<br />

for a new exotic <strong>option</strong> D with some complex payoff. If <strong>the</strong>re is a way that a payoff<br />

9 Here we state <strong>the</strong> more general, multi-dimensional version, of Girsanov’s <strong>the</strong>orem.<br />

15


of D can be decomposed <strong>in</strong>to a number of less complex payoffs A,B <strong>and</strong> C that are<br />

already traded <strong>in</strong> <strong>the</strong> <strong>market</strong>, <strong>and</strong> a cash position, <strong>the</strong>n this product is atta<strong>in</strong>able. S<strong>in</strong>ce,<br />

<strong>the</strong> derivatives A,B <strong>and</strong> C are all known to have a mart<strong>in</strong>gale property 10 , <strong>the</strong>n <strong>the</strong>re<br />

is a self-f<strong>in</strong>anc<strong>in</strong>g way of replicat<strong>in</strong>g <strong>the</strong> exotic claim by cont<strong>in</strong>uously rebalanc<strong>in</strong>g a<br />

portfolio of <strong>the</strong>se less-exotic assets, speaks Theorem (2.3.3). The price of such a new<br />

product will, by no-arbitrage, be equal to <strong>the</strong> cost of establish<strong>in</strong>g such a portfolio. As<br />

for SDE, given by (2.12), by us<strong>in</strong>g a mart<strong>in</strong>gale property of d ˜S t from (2.23) <strong>and</strong> <strong>the</strong><br />

fact that at any given time t: V t = ζ t S t + γ t B t , we can show that <strong>the</strong> discounted<br />

portfolio Ṽt is a mart<strong>in</strong>gale:<br />

dṼt = d(Bt<br />

−1 V t ) = V t dBt<br />

−1<br />

+ B −1<br />

t<br />

dV t = −rṼtdt + ζ t B −1<br />

= −rṼtdt + ζ t (d ˜S t − S t dBt<br />

−1 ) + γ t rB t dtBt<br />

−1<br />

= ζ t d ˜S t − rṼtdt + r(ζ t S t + γ t B t )Bt<br />

−1 dt<br />

t<br />

dS t + γ t Bt<br />

−1 dB t<br />

= ζ t d ˜S t . (2.27)<br />

In <strong>the</strong> light of Theorem (2.3.1), it is important to note that dṼt <strong>and</strong> d ˜S t are, <strong>in</strong> fact,<br />

numeraire-rebased processes with a cash bond B t as a numeraire asset. The result<br />

of (2.27) should rem<strong>in</strong>d us of (2.7), where a discrete process ∆Ṽi appeared to be a discrete<br />

mart<strong>in</strong>gale. Theorems (2.3.3) <strong>and</strong> (2.3.1) do ensure that self-f<strong>in</strong>anc<strong>in</strong>g replication<br />

works <strong>in</strong> cont<strong>in</strong>uous time as well.<br />

After discussion of mart<strong>in</strong>gale pric<strong>in</strong>g approach, our short <strong>in</strong>troduction to <strong>the</strong> general<br />

subject of asset pric<strong>in</strong>g is more or less complete. All <strong>the</strong> tools developed with <strong>the</strong><br />

stock economy <strong>in</strong> m<strong>in</strong>d will be applied to <strong>the</strong> <strong>in</strong>terest rate <strong>market</strong>s <strong>in</strong> next chapter.<br />

10 accord<strong>in</strong>g to <strong>the</strong> risk-neutral <strong>valuation</strong> pr<strong>in</strong>ciple<br />

16


Chapter 3<br />

LIBOR Market Model<br />

Hav<strong>in</strong>g left <strong>the</strong> equity world, evolution of a stock is not relevant to <strong>the</strong> <strong>valuation</strong> process<br />

any more. From now on, we will be operat<strong>in</strong>g on a tenor structure of forward LIBOR<br />

rates, described <strong>in</strong> <strong>the</strong> first section of this chapter. While <strong>the</strong> importance of a numeraire<br />

concept has been disguised by ubiquitous discount<strong>in</strong>g so far, asset pric<strong>in</strong>g <strong>in</strong> <strong>the</strong> economy<br />

of stochastic <strong>in</strong>terest rates appeals to numeraires explicitly. The ma<strong>the</strong>matical<br />

apparatus developed <strong>in</strong> <strong>the</strong> previous chapter will prove <strong>in</strong>dispensable as we <strong>in</strong>troduce<br />

LIBOR Market Model <strong>and</strong> pric<strong>in</strong>g under term<strong>in</strong>al measure. We shall spend some time<br />

discuss<strong>in</strong>g <strong>the</strong> characteristic <strong>in</strong>terest rate products <strong>and</strong> calibration of <strong>the</strong> pric<strong>in</strong>g framework<br />

to <strong>the</strong> current <strong>market</strong> data. In <strong>the</strong> f<strong>in</strong>ale, <strong>the</strong> LIBOR Market PDE will be loosely<br />

derived.<br />

3.1 Forward LIBOR Rates<br />

Tenor Structure. In <strong>the</strong> previous chapter we have been discuss<strong>in</strong>g a sett<strong>in</strong>g that implicated<br />

a risk-free rate r which is constant regardless of <strong>the</strong> bond maturity. When<br />

pric<strong>in</strong>g claims that are cont<strong>in</strong>gent on <strong>in</strong>terest rate movements, this assumption turns<br />

<strong>in</strong>to an obstacle <strong>and</strong> has to be dismissed. If we want to acquire a more realistic <strong>model</strong><br />

of <strong>in</strong>terest rate evolution, its setup will have to be able to account for <strong>the</strong> shape of <strong>the</strong><br />

curve, that is currently observable <strong>in</strong> <strong>the</strong> real-world <strong>market</strong>s.<br />

The underly<strong>in</strong>g asset <strong>in</strong> an <strong>in</strong>terest rate economy is a zero-coupon, discount bond<br />

that pays one unit of currency at maturity. Let’s denote D(t, T ) to be <strong>the</strong> price of such<br />

a bond at time t <strong>and</strong> with maturity at T . The specific term structure, characterictic of<br />

<strong>the</strong> <strong>market</strong> be<strong>in</strong>g <strong>model</strong>led at time t = 0, is def<strong>in</strong>ed by a vector of N discount bonds<br />

ordered with respect to <strong>the</strong>ir <strong>in</strong>creas<strong>in</strong>g maturity dates, i.e.:<br />

(<br />

)<br />

D(0) = D(0, 1), . . . , D(0, 1 ≤ i < N), . . . , D(0, N) (3.1)<br />

where D(0, i) = D(t, T i ) <strong>and</strong> t < T 1 < · · · < T N−1 < T N . Note that our discount<br />

zero-coupon bond is a mere abstraction of a zero rate. Writ<strong>in</strong>g L 1 for a one-year zero<br />

17


ate <strong>and</strong> Y for a notional amount, <strong>the</strong> follow<strong>in</strong>g result must hold <strong>in</strong> a <strong>market</strong> free of<br />

arbitrage:<br />

Y ≡ Y (1 + L 1 )D(0, 1)<br />

i.e. compound<strong>in</strong>g Y with L 1 <strong>and</strong> <strong>the</strong>n discount<strong>in</strong>g with D(0, 1) produces <strong>the</strong> same<br />

notional.<br />

We shall write α i = T i − T i+1 for <strong>the</strong> time period between any two consecutive<br />

bonds, namely tenor, <strong>and</strong> for our <strong>model</strong> assume it to be constant. Therefore, <strong>the</strong> order<strong>in</strong>g<br />

(3.1) of N bonds can be referred to as <strong>in</strong>terest rate tenor structure.<br />

Forward Rates. The LIBOR <strong>market</strong> Model operates <strong>in</strong> terms of forward rates, ra<strong>the</strong>r<br />

than zero rates. We shall illustrate <strong>the</strong> notion of a forward LIBOR rate with help of<br />

a Forward Rate Agreement example. As <strong>in</strong> [Hull, 1999], a forward rate agreement<br />

(FRA) is an over-<strong>the</strong>-counter agreement that a certa<strong>in</strong> <strong>in</strong>terest rate will apply to a certa<strong>in</strong><br />

pr<strong>in</strong>cipal dur<strong>in</strong>g a specified future period of time. Given a FRA:<br />

Def<strong>in</strong>ition 3.1.1. Forward LIBOR rate L i (t) is an <strong>in</strong>terest rate one can contract for<br />

at time t to lend or borrow money for <strong>the</strong> future time period [T i , T i+1 ]. Given prices of<br />

discount bonds matur<strong>in</strong>g at T i <strong>and</strong> T i+1 , <strong>the</strong> forward rate is def<strong>in</strong>ed by:<br />

1 + α i L i (t) = D(0, T i)<br />

D(0, T i+1 )<br />

where t ≤ T i (3.2)<br />

a<br />

N-1<br />

0<br />

1<br />

2<br />

N-1<br />

N<br />

T<br />

D 0,1<br />

1<br />

L N-1<br />

D 0,2<br />

1<br />

D 0,N-1<br />

1<br />

D 0,N<br />

1<br />

Figure 3.1: Tenor Structure of Interest Rates<br />

Keep<strong>in</strong>g our assumptions of no bid-ask spread <strong>and</strong> no-arbitrage, a FRA should cost<br />

exactly noth<strong>in</strong>g if <strong>the</strong> contracted <strong>in</strong>terest rate equals a forward rate implied by a current<br />

yield curve. Instead of tak<strong>in</strong>g one of <strong>the</strong> sides <strong>in</strong> <strong>the</strong> FRA, a trader can achieve <strong>the</strong><br />

same goal by manipulat<strong>in</strong>g bonds <strong>in</strong> <strong>the</strong> money-<strong>market</strong>. Let’s illustrate this statement<br />

by rewrit<strong>in</strong>g (3.2):<br />

D(0, T i ) − D(0, T i+1 )<br />

= α i L i (0) (3.3)<br />

D(0, T i+1 )<br />

18


The right-h<strong>and</strong> side of <strong>the</strong> equality above can be perceived <strong>in</strong> <strong>the</strong> follow<strong>in</strong>g way. A<br />

<strong>market</strong> player sets up a portfolio comprised of a long position <strong>in</strong> a bond with shorter<br />

maturity T i <strong>and</strong> a short position with a T i+1 bond, borrow<strong>in</strong>g <strong>the</strong> difference between<br />

<strong>the</strong> two at <strong>the</strong> observed zero <strong>in</strong>terest rate until T i+1 . When <strong>the</strong> clock hits T i , <strong>the</strong> player<br />

receives <strong>the</strong> pr<strong>in</strong>cipal due to <strong>the</strong> short position. As α i more time passes, she has to<br />

deliver <strong>the</strong> pr<strong>in</strong>cipal on <strong>the</strong> shorted bond −D(0, T i+1 ) plus <strong>the</strong> money borrowed to<br />

1<br />

establish a portfolio, compounded at<br />

D(0,T t+1 ). The eventual extra money outflow is<br />

<strong>the</strong> right-h<strong>and</strong> side of (3.3), or <strong>the</strong> <strong>in</strong>terest payment she could have contracted for by<br />

enter<strong>in</strong>g a short side of <strong>the</strong> FRA. Tak<strong>in</strong>g reverse positions <strong>in</strong> <strong>the</strong> portfolio assets at time<br />

zero would be equivalent to tak<strong>in</strong>g a long side of <strong>the</strong> FRA <strong>and</strong> receiv<strong>in</strong>g <strong>the</strong> forward<br />

LIBOR rate for T i,i+1 .<br />

The scene for LMM <strong>and</strong> <strong>the</strong> concepts discussed so far <strong>in</strong> <strong>the</strong> chapter can be shown<br />

visually <strong>in</strong> Figure (3.1) From <strong>the</strong> figure above one could clearly see: <strong>the</strong> price of <strong>the</strong><br />

discount bond can be expressed as a product of forward LIBOR rates, implied by <strong>the</strong><br />

tenor structure:<br />

D(0, T i ) =<br />

N−1<br />

∏<br />

i=0<br />

1<br />

1 + α i L i (T i )<br />

(3.4)<br />

3.2 LIBOR Market Model<br />

LIBOR Process The foundation of this section resides on <strong>the</strong> proposition that forward<br />

LIBOR rates L i (T i ) should follow a mart<strong>in</strong>gale process. Indeed, look<strong>in</strong>g back<br />

at (3.3) we note that a discount bond D(t, T i+1 ) has a strictly positive value <strong>and</strong> <strong>the</strong>refore<br />

can be treated as a numeraire. With arbitrage-free <strong>market</strong> <strong>in</strong> m<strong>in</strong>d, <strong>the</strong> Fundamental<br />

Theorem of Asset Pric<strong>in</strong>g postulates <strong>the</strong>re must exist a unique measure Q i+1 , under<br />

which a numeraire-denom<strong>in</strong>ated portfolio is a mart<strong>in</strong>gale. This implication applies to<br />

a LIBOR rate α i L i (T i ) with α i which is a constant <strong>and</strong> leads us to <strong>the</strong> def<strong>in</strong>ition of a<br />

LIBOR process.<br />

In LMM, <strong>the</strong> process of cont<strong>in</strong>uous evolution of <strong>market</strong> rates is described by <strong>the</strong><br />

follow<strong>in</strong>g stochastic differential equation:<br />

dL i (T i ) = σ i (t)L i (T i )dW Qi+1 (3.5)<br />

where σ i is a cont<strong>in</strong>uous bounded volatility function <strong>and</strong> dW Qi+1 is a Wiener process<br />

under a mart<strong>in</strong>gale measure Q i+1 . Itô’s formula can be applied to acquire <strong>the</strong> exact<br />

solution to (3.5):<br />

(<br />

L i (t) = L i (0)exp − 1 ∫ t<br />

∫ t<br />

)<br />

σ i (s) 2 ds + σ i (s)dW Qi+1 (s) (3.6)<br />

2<br />

0<br />

With some <strong>the</strong>ory from <strong>the</strong> second chapter <strong>and</strong> one of <strong>the</strong> properties of Itô <strong>in</strong>tegrals 1 ,<br />

we can conclude that <strong>in</strong> LMM forward rates are <strong>model</strong>led to be geometric Brownian<br />

motion processes with variance κ 2 = ∫ t<br />

0 σ i(s) 2 ds <strong>and</strong> mean − 1 2 κ2 .<br />

1 If σ i (s) is a bounded, non-stochastic function, <strong>the</strong>n <strong>the</strong> Itô <strong>in</strong>tegral of <strong>the</strong> formRt<br />

0 σ i(s)dW (s) converges<br />

<strong>in</strong> distribution to N(0,Rt<br />

0 σ i(s) 2 ds)<br />

19<br />

0


Change of Numeraire Before we can discuss <strong>the</strong> key result of LMM - <strong>model</strong>l<strong>in</strong>g<br />

forward rates under term<strong>in</strong>al measure, it is necessary to address such issue as change<br />

of numeraire. This result should rem<strong>in</strong>d us of <strong>the</strong> change of measure result from <strong>the</strong><br />

first chapter, as chang<strong>in</strong>g <strong>the</strong> numeraire implicates change of measure, accord<strong>in</strong>g to<br />

Theorem 2.3.1. Similar to narrative of [Pelsser, 2000], consider some derivative product<br />

with payoff V (T ) <strong>and</strong> two numeraire assets D <strong>and</strong> B. Then, <strong>in</strong> <strong>the</strong> Risk-Neutral<br />

world, value of such product at time zero is def<strong>in</strong>ed as:<br />

[ ] V (T )<br />

E D D(T, T ) ∣ F t = V (t)<br />

[ ] V (T )<br />

or E B D(t, T )<br />

B(T, T ) ∣ F t = V (t) (3.7)<br />

B(t, T )<br />

where D <strong>and</strong> B are equivalent mart<strong>in</strong>gale measures imposed by <strong>the</strong> choice of numeraire<br />

D <strong>and</strong> B. With no arbitrage assumption:<br />

[ ]<br />

[ ]<br />

V (T )<br />

V (t) = E D V (T )<br />

D(T, T ) ∣ F t D(t, T ) = E B B(T, T ) ∣ F t B(t, T ) (3.8)<br />

Let V D (T ) be D-denom<strong>in</strong>ated payoff of <strong>the</strong> product. Now, if we want to express e.g.<br />

<strong>the</strong> D-denom<strong>in</strong>ated value of our product as <strong>the</strong> B-expectation, <strong>the</strong> equation (3.8) can<br />

be rewritten as<br />

E D[ [<br />

] [<br />

]<br />

V D (T )|F t = E<br />

B D(T, T )/D(t, T )<br />

V D (T ) B(T, T )/B(t, T ) ∣ F t = E B V D (T ) dD<br />

]<br />

dB ∣ F t (3.9)<br />

The last expectation operator should look familiar. It has <strong>the</strong> same mean<strong>in</strong>g as <strong>the</strong><br />

D(T,T )/D(t,T )<br />

change of measure result (2.21) for stochastic processes. Variable<br />

B(T,T )/B(t,T ) is<br />

noth<strong>in</strong>g more than a “numeraire” notation for a Radon-Nikodym derivative dD/dB,<br />

previously def<strong>in</strong>ed as (2.22). To summarize, we have derived a recipe of how <strong>the</strong><br />

expectation of a numeraire discounted value can be expressed <strong>in</strong> terms of expectation<br />

under a measure imposed by a different choice of numeraire. This result will be needed<br />

<strong>in</strong> <strong>the</strong> discussion of pric<strong>in</strong>g under term<strong>in</strong>al measure.<br />

Term<strong>in</strong>al Measure Still keep<strong>in</strong>g <strong>the</strong> discussion quite abstract from derivatives with<br />

any specific payoff pattern, we shall address <strong>the</strong> issue of pric<strong>in</strong>g <strong>in</strong>terest rate products<br />

that are cont<strong>in</strong>gent on several forward rates. While simple, one-rate derivatives allow<br />

for closed-form expressions 2 , pric<strong>in</strong>g <strong>option</strong>s on several rates requires some extra effort.<br />

The complications <strong>in</strong> valu<strong>in</strong>g such derivatives is <strong>the</strong> fact that every LIBOR rate<br />

is a mart<strong>in</strong>gale process under its own measure. These measures are imposed by <strong>the</strong><br />

discount bonds with associated maturity dates, or numeraire assets.<br />

The key idea of LMM is <strong>model</strong>l<strong>in</strong>g a process for an arbitrary valid forward rate<br />

<strong>in</strong> terms of <strong>the</strong> probability measure rul<strong>in</strong>g <strong>the</strong> Brownian motion of <strong>the</strong> rate with most<br />

distant payoff. In o<strong>the</strong>r words, all <strong>the</strong> uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> development of <strong>the</strong> price<br />

for derivative should be def<strong>in</strong>ed solely by <strong>the</strong> mart<strong>in</strong>gale measure of <strong>the</strong> last forward<br />

rate, or “term<strong>in</strong>al” measure. We shall assume Brownian Motions driv<strong>in</strong>g <strong>the</strong> <strong>in</strong>dividual<br />

rates to be <strong>in</strong>dependent processes <strong>and</strong>, for now, no correlation exists between any two<br />

Brownian motion terms.<br />

2 as we shall see fur<strong>the</strong>r on <strong>in</strong> this text<br />

20


At this po<strong>in</strong>t of <strong>the</strong> text all necessary ma<strong>the</strong>matical tools have been covered to<br />

proceed with derivation. It is <strong>the</strong> recursive application of multi-dimensional Girsanov’s<br />

<strong>the</strong>orem <strong>and</strong> Change of Measure results that will help collect all <strong>the</strong> LIBOR rates under<br />

<strong>the</strong> same measure. First, recall that accord<strong>in</strong>g to Theorem (2.3.2), we can go from<br />

one measure to ano<strong>the</strong>r by <strong>in</strong>troduc<strong>in</strong>g <strong>the</strong> drift <strong>in</strong>to <strong>the</strong> orig<strong>in</strong>al LIBOR process, as<br />

<strong>in</strong> (2.26). In order to do so, we need to obta<strong>in</strong> a drift-correction term κ, which is at <strong>the</strong><br />

same time <strong>the</strong> volatility of a Radon-Nikodym derivative.<br />

1. Work<strong>in</strong>g out drift-correction term<br />

It was agreed that discount bonds D(t, T i ) serve as numeraire assets that impose<br />

mart<strong>in</strong>gale measures on separate forward rates. Let Q i denote such numeraireborn<br />

measure, <strong>the</strong>refore mak<strong>in</strong>g dL i (t) a mart<strong>in</strong>gale. So, from (3.9):<br />

Y t =<br />

dQi<br />

dQ i+1 = D(t, T i)/D(0, T i )<br />

D(t, T i+1 )/D(0, T i+1 ) = D(0, T i+1)<br />

D(0, T i ) (1 + α iL i (t)) (3.10)<br />

At <strong>the</strong> same time, from (2.25):<br />

( ∫ t<br />

Y t =<br />

dQi<br />

dQ i+1 = exp κ(s)dW Qi+1 (s) − 1 2<br />

0<br />

∫ t<br />

Note that Y t is a mart<strong>in</strong>gale process described by <strong>the</strong> SDE:<br />

0<br />

)<br />

κ(s) 2 ds<br />

(3.11)<br />

dY t = κ(t)Y t dW i+1 (3.12)<br />

The f<strong>in</strong>al expression for κ(t) now requires application of Itô’s Lemma to (3.10):<br />

dY t = D(0, T i+1)<br />

D(0, T i ) α idL i (t) = D(0, T i+1)<br />

D(0, T i ) α iσ i (t)L i (t)dW i+1 (3.13)<br />

The last equality comes directly from def<strong>in</strong>ition of a LIBOR process. Equat<strong>in</strong>g<br />

(3.12) <strong>and</strong> (3.13):<br />

κ(t)Y t dW i+1 = D(0, T i+1)<br />

D(0, T i ) α iσ i (t)L i (t)dW i+1<br />

D(0,<br />

κ(t)<br />

✟ ✟✟✟ T i+1 )<br />

✟ D(0, T i ) (1 + α iL i (t))✘✘ dW i+1 ✘ D(0,<br />

=<br />

✟ ✟✟✟ T i+1 )<br />

✟ D(0, T i ) α iσ i (t)L i (t)✘✘ dW i+1 ✘<br />

κ(t) = α iσ i (t)L i (t)<br />

1 + α i L i (t)<br />

(3.14)<br />

2. Adjust<strong>in</strong>g <strong>the</strong> drift.<br />

Now that we have a drift correction term κ(t) we can apply Girsanov’s <strong>the</strong>orem<br />

to get a change of measure relation 3 :<br />

dW i = dW i+1 − κ(t)ρ i,i+1 dt = dW i+1 − α iρ i,i+1 σ i (t)L i (t)<br />

dt<br />

1 + α i L i (t)<br />

3 In case of a s<strong>in</strong>gle underly<strong>in</strong>g Brownian Motion, correlation term ρ i,j = 1<br />

21


For <strong>in</strong>stance, let’s set κ T i = αiσi(t)Li(t)<br />

1+α i L i (t)<br />

to be <strong>the</strong> drift-correction term for <strong>the</strong> term<strong>in</strong>al<br />

measure <strong>and</strong> f<strong>in</strong>d <strong>the</strong> L i−3 (t) forward rate under <strong>the</strong> term<strong>in</strong>al measure.<br />

dL i−3 (t) = σ i−3 (t)L i−3 (t)dW i−2<br />

[<br />

]<br />

= σ i−3 (t)L i−3 (t) dW i−1 − κ T i−2(t)ρ i−3,i−2 dt<br />

( )<br />

]<br />

= σ i−3 (t)L i−3 (t)[<br />

dW i − κ T i−1(t)ρ i−3,i−1 dt − κ T i−2(t)ρ i−3,i−2 dt<br />

We proceed with open<strong>in</strong>g <strong>the</strong> brackets <strong>and</strong> rearrang<strong>in</strong>g <strong>the</strong> terms:<br />

(3.15)<br />

dL i−3 (t)<br />

[<br />

]<br />

L i−3 (t) = −σ i−3(t)κ T i−1(t)ρ i−3,i−1 dt + σ i−3 (t) dW i − κ T i−2(t)ρ i−3,i−2 dt<br />

]<br />

= −σ i−3 (t)dt<br />

[κ T i−2(t)ρ i−3,i−2 + κ T i−1(t)ρ i−3,i−1 + σ i−3 (t)dW i (3.16)<br />

The general form:<br />

where<br />

µ i (t) =<br />

dL i (t)<br />

L i (t) = µ i(t)dt + σ i (t)dW N (3.17)<br />

{<br />

−σi (t) ∑ N−1<br />

j=i+1 κT j (t)ρ i,j i < N − 1<br />

0 i = N − 1<br />

(3.18)<br />

3.3 Interest Rate Derivatives. Caplets, Caps <strong>and</strong> Swaptions.<br />

So far <strong>in</strong> this chapter we have been mostly busy with <strong>the</strong> description of <strong>the</strong> <strong>model</strong> that<br />

tries to immitate forward rates. However, our idea of a claim cont<strong>in</strong>gent on such rates<br />

has been quite an abstract one. Now it’s good time to dismiss <strong>the</strong> assumption of “some”<br />

payoff pattern <strong>and</strong> discuss several more specific <strong>in</strong>terest rate products, that are traded<br />

<strong>in</strong> <strong>the</strong> <strong>market</strong>s.<br />

A Caplet. A simpliest of all <strong>in</strong>terest-rate derivatives. It is analogous to a European<br />

call <strong>option</strong> from <strong>the</strong> equity derivatives family. It’s payoff is as follows:<br />

c i (T i+1 ) = α i (L i (T i ) − K i ) + (3.19)<br />

One can’t help but notice an important difference between a caplet <strong>and</strong> it’s equity counterpart.<br />

At maturity, <strong>the</strong> holder of a caplet has <strong>the</strong> <strong>option</strong> to take a long side of <strong>the</strong><br />

forward contract to borrow money at an <strong>in</strong>terest rate K, also known as a strike rate for<br />

α i units of time. This is <strong>in</strong> contrast to an <strong>in</strong>-<strong>the</strong>-money call <strong>option</strong> that immediately<br />

22


pays off at expiry. Conveniently, <strong>the</strong> exact solution for such product becomes available<br />

via analytical risk-neutral <strong>valuation</strong> along <strong>the</strong> follow<strong>in</strong>g l<strong>in</strong>es:<br />

[<br />

] [ ( ) ]<br />

c i (t)<br />

D(t, T i+1 ) = c i (T i+1 )<br />

L i (T i ) − K i EQi+1 = α i E Qi+1 +<br />

(3.20)<br />

D(T i+1 , T i+1 )<br />

1<br />

In <strong>market</strong> practice, <strong>the</strong> process for dL i is commonly assumed to be geometric Brownian<br />

Motion (3.6), <strong>and</strong> s<strong>in</strong>ce it’s payoff is pla<strong>in</strong>-vanilla, we compute this expectation by<br />

<strong>in</strong>tegrat<strong>in</strong>g with respect to log-normal distribution of dL i , that results <strong>in</strong>:<br />

c i (t) = D(t, T i+1 ) [ L i (t)Φ(d 1 ) − K i Φ(d 2 ) ] (3.21)<br />

where Φ(·) is cumulative normal distribution function <strong>and</strong><br />

d 1 = log(L i(t)/K i ) + 0.5σ 2 i (T i − t)<br />

σ i<br />

√<br />

Ti − t<br />

d 2 = d 1 − σ i<br />

√<br />

Ti − t<br />

The equation (3.21) is commonly refered to as Black’s formula <strong>and</strong> is widely used <strong>in</strong><br />

<strong>the</strong> <strong>in</strong>dustry. The popularity of this formula has enabled <strong>the</strong> traders to quote caplet<br />

prices <strong>in</strong> terms of Black implied volatilities.<br />

A Cap. It is worth mention<strong>in</strong>g that <strong>in</strong>dividual caplets are <strong>the</strong> <strong>in</strong>struments that are not<br />

traded <strong>in</strong> <strong>the</strong> <strong>market</strong>s. Instead, caps, i.e. portfolios consist<strong>in</strong>g of consecutive caplets<br />

are bought <strong>and</strong> sold. The holder of a cap has an <strong>option</strong> to enter <strong>in</strong>to a forward contract<br />

at maturity of each caplet. It’s value at time t is <strong>the</strong>refore:<br />

C(t) =<br />

N−1<br />

∑<br />

i=1<br />

c i (t) (3.22)<br />

Individual caplet prices <strong>and</strong> volatilities can be extracted from quoted cap prices via<br />

boot-strapp<strong>in</strong>g algorithm.<br />

Swaps <strong>and</strong> Swaptions.<br />

• A Swap is a contract between two <strong>market</strong> players to exchange a series of float<strong>in</strong>grate<br />

payments for a series of fixed-rate payments. This <strong>in</strong>strument can help hedge<br />

aga<strong>in</strong>st high <strong>in</strong>terest rates via enter<strong>in</strong>g a payer swap, <strong>in</strong> which <strong>the</strong> party has to<br />

pay a fixed rate <strong>in</strong> exchange for <strong>the</strong> float<strong>in</strong>g. So, <strong>the</strong> cash outflow for <strong>the</strong> fixed<br />

side <strong>and</strong> an arbitrary payment date is:<br />

V fix<br />

i (t) = D(t, T i+1 )α i K (3.23)<br />

On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, engagement <strong>in</strong> a reciever swap, where <strong>the</strong> fixed rate is received<br />

<strong>and</strong> float<strong>in</strong>g is payed, would mean hav<strong>in</strong>g strong reasons to believe that<br />

23


<strong>the</strong> <strong>in</strong>terest rates are go<strong>in</strong>g to fall. The expected cash outflow for one payment at<br />

T i <strong>in</strong> this case would be:<br />

Vi flo (t) = D(t, T i+1 )α i E Qi+1 [L i (T )] = D(t, T i ) − D(t, T i+1 ) (3.24)<br />

} {{ }<br />

L i (t)<br />

Given a set of payment dates {T i } N t=0, <strong>the</strong> value of, e.g. payer swap, is:<br />

V pswap<br />

t =<br />

N∑<br />

i=t<br />

V flo<br />

i<br />

−<br />

N∑<br />

i=t<br />

V fix<br />

i (3.25)<br />

The strike rate is set such that <strong>the</strong> current price of <strong>the</strong> swap is zero. Such rate K<br />

is called par swap rate:<br />

K par<br />

t = D(t, T n) − D(t, T N )<br />

∑ N<br />

i=n+1 a i−1D(t, T i )<br />

• European Swaption gives a holder <strong>the</strong> right to take a side <strong>in</strong> a swap contract at<br />

maturity of <strong>the</strong> contract. It’s payoff <strong>in</strong> case of a payer swap <strong>option</strong>:<br />

V ESw<br />

T<br />

= max(V pswap<br />

T<br />

, 0) (3.26)<br />

• Bermudan Swaption gives a holder <strong>the</strong> right to enter a swap on a predef<strong>in</strong>ed<br />

number of dates <strong>in</strong> <strong>the</strong> future. We shall assume <strong>the</strong>se dates are <strong>the</strong> fix<strong>in</strong>g dates<br />

of <strong>the</strong> tenor structure. It’s payoff is as follows:<br />

V BSw<br />

t = max(H t (t), E t (t)) (3.27)<br />

where H N−1 (T N−1 ) = 0 <strong>and</strong> E t (t) = V swap<br />

t . H t (t) is a common notation<br />

for <strong>the</strong> hold value (a contract has not been exercised up to <strong>and</strong> <strong>in</strong>clud<strong>in</strong>g time<br />

t. E t (t) is a common notation for immediate exercise value, which <strong>in</strong> case of<br />

Bermudan swaption equals value of <strong>the</strong> swap [Peterbarg, 2003].<br />

Chooser Option. This product gives a holder <strong>the</strong> right to enter <strong>in</strong>to a forward contract<br />

at <strong>the</strong> rate equal to <strong>the</strong> difference between <strong>the</strong> maximum of <strong>the</strong> previous forward rates<br />

<strong>and</strong> a strike rate K. The payoff follows:<br />

V Cho<br />

T<br />

= α T −1 ( max<br />

0


1. Tenor structure of N <strong>in</strong>terest rates<br />

2. Instantaneous volatility function σ i (t), i ∈ [1..N]<br />

3. Correlation structure of LIBOR rates.<br />

Tenor structure of <strong>in</strong>terest rates is available to us immediately but volatility <strong>and</strong> correlation<br />

require some extra work.<br />

3.4.1 Calibration to caplets<br />

LIBOR Market Model assumes that forward LIBOR rates follow a lognormal distribution,<br />

i.e L i (T i ) = L i (t) exp(Z i+1 ), where Z i+1 ∼ N (− 1 2 κ2 , κ 2 ) under Q T i+1<br />

,<br />

<strong>and</strong>:<br />

κ 2 =<br />

∫ Ti<br />

t<br />

||σ(s)|| 2 ds<br />

One can see that, <strong>in</strong> order to get a value for variance of a LIBOR process, we need to<br />

<strong>in</strong>tegrate an <strong>in</strong>stantaneous volatility function, that is unfortunately is not at our disposal<br />

by default. What we do have is Black implied volatilites that are available to public.<br />

Calibration to caplets implies that we should take quoted caplet volatility values as a<br />

template/skeleton <strong>and</strong> fit <strong>the</strong> curve that matches <strong>the</strong> view of <strong>the</strong> <strong>market</strong>. Discussion on<br />

approximat<strong>in</strong>g <strong>the</strong> <strong>in</strong>stantaneous volatility function can be found <strong>in</strong> [Rebonato, 1998].<br />

In this <strong>the</strong>sis, a simplified approach of a flat <strong>in</strong>stantaneous volatility function will<br />

be adopted. To be more specific, we shall assume LIBOR rate volatility to be constant<br />

for a given tenor. Calibration to caplets <strong>in</strong> this case is particularly easy. By sett<strong>in</strong>g:<br />

or<br />

σ 2 BLACK(T i − t) = κ 2<br />

√<br />

1<br />

σ BLACK =<br />

T t − t<br />

∫ Ti<br />

t<br />

||σ(s)|| 2 ds<br />

we can accept Black volatilities as an average of an <strong>in</strong>stantaneous volatility function<br />

throughout <strong>the</strong> given tenor <strong>and</strong> use <strong>the</strong>m directly.<br />

3.4.2 Correlation structure of forward rates.<br />

Hav<strong>in</strong>g calibrated to caplet volatilities, <strong>the</strong> setup of <strong>the</strong> <strong>model</strong> is still not complete. In<br />

our discussion of pric<strong>in</strong>g under term<strong>in</strong>al measure we have neglected any correlation<br />

that might exist between <strong>the</strong> LIBOR processes. As for this moment, any given process<br />

for a LIBOR rate is driven solely by its Brownian Motion with drift, imposed by<br />

change of measure. If we are <strong>model</strong>l<strong>in</strong>g such tenor structure of N LIBOR rates, <strong>the</strong><br />

result<strong>in</strong>g processes can be shown <strong>in</strong> <strong>the</strong> follow<strong>in</strong>g matrix-vector form with matrix A as<br />

25


an identity matrix 4 :<br />

⎡<br />

⎢<br />

⎣<br />

dL 1 (t)<br />

L 1 (t)<br />

.<br />

dL N (t)<br />

L N (t)<br />

⎤<br />

⎡<br />

⎥<br />

⎦ = ⎢<br />

⎣<br />

µ 1 (t)<br />

.<br />

µ N−1 (t)<br />

⎤<br />

⎡<br />

⎥ ⎢<br />

⎦ dt+ ⎣<br />

σ 1 (t)<br />

.<br />

σ N (t)<br />

⎤ ⎛<br />

⎥ ⎜<br />

⎦ ⎝<br />

a 11 . . . a 1N<br />

. . . . . .<br />

a N1 . . . a NN<br />

⎞<br />

⎟<br />

⎠<br />

A<br />

⎡<br />

⎢<br />

⎣<br />

dW 1 (t)<br />

.<br />

dW N (t)<br />

One might f<strong>in</strong>d it hard to believe that all forward rates <strong>in</strong> <strong>the</strong> <strong>model</strong>led term structure<br />

are completely <strong>in</strong>dependent. For this specific reason, <strong>the</strong> <strong>model</strong> needs to be calibrated<br />

to account for multiple sources of uncerta<strong>in</strong>ty <strong>in</strong> <strong>the</strong> behaviour of an <strong>in</strong>terest rate.<br />

It is worth try<strong>in</strong>g to underst<strong>and</strong> <strong>the</strong> idea by look<strong>in</strong>g at only 2 forward rates dL 1 (t)<br />

<strong>and</strong> dL 2 (t), where dW 1 (t) <strong>and</strong> dW 2 (t) are two <strong>in</strong>dependent Brownian motions. The<br />

Itô process that is driven by N-factor Brownian Motion takes <strong>the</strong> follow<strong>in</strong>g form:<br />

dL i (t)<br />

L i (t) = µ i(t)dt +<br />

⎤<br />

⎥<br />

⎦<br />

Q N+1<br />

N∑<br />

α i dW i (t) (3.29)<br />

As we are consider<strong>in</strong>g correlation that is characteristic of Brown<strong>in</strong>an Motions, we can<br />

neglect <strong>the</strong> drift terms s<strong>in</strong>ce <strong>the</strong>y are determ<strong>in</strong>istic <strong>and</strong> do not affect r<strong>and</strong>omness or<br />

stochastic factors <strong>in</strong> any way. The coefficients α <strong>and</strong> β will be used for notational<br />

convenience, <strong>in</strong>stead of just α i as <strong>in</strong> (3.29):<br />

i=1<br />

dL 1 (t) = · · · + α 1 dW 1 (t) + β 1 dW 2 (t)<br />

dL 2 (t) = · · · + α 2 dW 1 (t) + β 2 dW 2 (t)<br />

(3.30)<br />

We postulate that dW 1 (t) <strong>and</strong> dW 2 (t) are two <strong>in</strong>dependent Wiener processes that<br />

follow N (0, dt). Covariance of dL 1 (t) with itself is its variance:<br />

[ (α1<br />

Cov(dL 1 (t), dL 1 (t)) = E dW 1 (t) + β 1 dW 2 (t) )( α 1 dW 1 (t) + β 1 dW 2 (t) )]<br />

= E<br />

[α1dW 2 1 (t) 2 + 2α 1 β 1 dW 1 (t)dW 2 (t) + β1dW 2 2 (t) 2]<br />

= α1E[dW 2 1 (t) 2 ] + 2α 1 β 1 E[dW 1 (t)dW 2 (t)] +β 2<br />

} {{ }<br />

1E[dW 2 (t) 2 ]<br />

(<br />

= α1<br />

2 Var ( dW 1 (t) )<br />

) )<br />

− E 2 [dW 1 (t)] + β<br />

} {{ } } {{ }<br />

1dt<br />

2<br />

dt<br />

0<br />

= α1dt 2 + β1dt 2 = (α1 2 + β1)dt 2 = Var ( dL 1 (t) )<br />

With calibration to caplets approach, this means that (α 2 1 + β 2 1)dt = ||σ i || 2 dt where<br />

||σ i || is Black implied volatility. With this result at h<strong>and</strong>, we can extract <strong>the</strong> volatility<br />

4 Matrix A conta<strong>in</strong>s coefficients that determ<strong>in</strong>e to what extent <strong>the</strong> <strong>in</strong>dividual Brownian Motion term contributes<br />

to <strong>the</strong> composite Wiener process that drives <strong>the</strong> stochastic process dL i (t)/L i (t)<br />

0<br />

26


<strong>and</strong> weight factor parts from <strong>the</strong> coefficients via scal<strong>in</strong>g:<br />

dL 1 (t) = . . . + ||σ 1 ||( α 1<br />

||σ 1 || dW 1(t) + β 1<br />

||σ 1 || dW 2(t) = ||σ 1 || ( a 11 dW 1 (t) + a 12 dW 2 (t) )<br />

dL 2 (t) = . . . + ||σ 2 ||( α 2<br />

||σ 2 || dW 1(t) + β 2<br />

||σ 2 || dW 2(t) = ||σ 2 || ( a 21 dW 1 (t) + a 22 dW 2 (t) )<br />

For Variance, we have:<br />

Var(dL 1 (t)) = ||σ 1 || 2 (a 2 11 + a 2 12)dt = ||σ 1 || 2 BLACK (3.31)<br />

For <strong>the</strong> equality above to hold, <strong>the</strong> correlation components <strong>in</strong> a matrix row should add<br />

up to one : a 2 11 + a 2 12 = 1. So it follows:<br />

√<br />

a 22 = 1 − a 2 11<br />

With all said above <strong>and</strong> for present moment, our coefficient matrix takes <strong>the</strong> follow<strong>in</strong>g<br />

form:<br />

[ ] [ ] [ √ ] [ ]<br />

a11 a 12 dW1 (t) a11 1 − a<br />

2<br />

≡ √ 11 dW1 (t)<br />

a 21 a 22 dW 2 (t) a 22 1 − a<br />

2<br />

22<br />

dW 2 (t)<br />

A<br />

Now, let’s look at covariance between two rates:<br />

(<br />

)<br />

Cov(dL 1 (t), dL 2 (t)) = ||σ 1 || ||σ 2 || Cov a 11 dW 1 (t) + a 12 dW 2 (t), a 21 dW 1 (t) + a 22 dW 2 (t)<br />

[ (a11<br />

= ||σ 1 || ||σ 2 || E dW 1 (t) + a 12 dW 2 (t) )( a 21 dW 1 (t) + a 22 dW 2 (t) )]<br />

= ||σ 1 || ||σ 2 || E<br />

[a 11 a 21 dW 1 (t) 2 + a 12 a 22 dW 2 (t) 2]<br />

= ||σ 1 || ||σ 2 || (a 11 a 21 + a 12 a 22 ) dt<br />

Cor(dL 1 (t), dL 2 (t)) = ✟✟ ||σ 1<br />

✟|| ✟✟ ✟ ||σ 2 ||(a 11 a 21 + a 12 a 22 )✚dt<br />

✟||σ ✟ 1 ||✟ √ dt ✟ ✟ ||σ✟✟<br />

2 ||✟ √ dt ✟ = ρ<br />

F<strong>in</strong>ally, lets denote: (<br />

1 ρ<br />

ρ 1<br />

A<br />

)<br />

= ρ<br />

Matrix ρ is symmetric <strong>and</strong> positive-def<strong>in</strong>ite, threfore we can apply Cholesky Decomposition<br />

to f<strong>in</strong>d <strong>the</strong> matrix coefficient matrix:<br />

( ) ( ) (<br />

AA T a11 a<br />

=<br />

12 a11 a 21<br />

a 2 11 + a 2 12 a 11 a 21 + a 12 a 22<br />

a 21 a 22 a 12 a 22 a 21 a 11 + a 22 a 12 a 2 21 + a 2 22<br />

A T =<br />

)<br />

= ρ<br />

27


Correlation. Now one th<strong>in</strong>g that is left to specify is <strong>the</strong> correlation values between<br />

any pair of rates. We go along with <strong>the</strong> approach from [Rebonato, 1998] 5 , where it is<br />

suggested that:<br />

ρ ij = e −β|T i−T j |<br />

(3.32)<br />

Beta is estimated from <strong>market</strong> data. This form makes sense as <strong>the</strong> correlation between<br />

two rates is more likely to decrease as <strong>the</strong>y are located far<strong>the</strong>r apart on <strong>the</strong> tenor structure.<br />

3.5 The LIBOR Market Model PDE<br />

Let’s consider an <strong>in</strong>terest rate economy whose population is described by a set of forward<br />

rates:<br />

L(t) = (L 1 (t), L 2 (t), . . . , L N−1 (t))<br />

Suppose we need to price an <strong>in</strong>terest rate product U(t, L) ≡ U(t, L(t)). Accord<strong>in</strong>g<br />

to risk-neutral <strong>valuation</strong> pr<strong>in</strong>ciple, <strong>the</strong> value of such a derivative at time t < T N <strong>in</strong><br />

LIBOR Market Model will be:<br />

U(t, L)<br />

D(t, T N ) = EQT N<br />

[ ]<br />

U(TN , L)<br />

D(T N , T N ) ∣ F t<br />

(3.33)<br />

where Q T N<br />

is <strong>the</strong> term<strong>in</strong>al mart<strong>in</strong>gale measure imposed by <strong>the</strong> forward rate with latest<br />

maturity. Values of discount bonds D(t, T N ) <strong>and</strong> D(T N , T N ) are known at time t <strong>and</strong><br />

<strong>the</strong>refore:<br />

U(t, L) = D(t, T N )E [ QT N<br />

U(T N , L) ∣ ]<br />

∣F t (3.34)<br />

However, <strong>the</strong> closed-form solutions based on (3.34) are rarely available due to <strong>the</strong><br />

multi-asset feature of most LIBOR derivatives. One approach that readily offers itself<br />

<strong>in</strong> such circumstances is <strong>the</strong> appplication of <strong>the</strong> Feynman-Kac <strong>the</strong>orem to l<strong>in</strong>k <strong>the</strong><br />

expectation (3.34) of <strong>the</strong> functionals of SDE solutions to <strong>the</strong> solution of <strong>the</strong> partial differential<br />

equation. A detailed account of this <strong>method</strong>ology with application to LIBOR<br />

Market Model can be found <strong>in</strong> [Pietersz, 2002]. Without proof, we now sketch <strong>the</strong><br />

derivation of <strong>the</strong> LIBOR Market PDE, proceed<strong>in</strong>g along <strong>the</strong> follow<strong>in</strong>g l<strong>in</strong>es:<br />

First, let’s denote Ũ = U(t, L)/D(t, T N ) at any given time t. The process dŨ(t, L)<br />

can be obta<strong>in</strong>ed by application of multi-dimensional Itô’s Lemma:<br />

N−1<br />

∂Ũ<br />

dŨ(t, L) =<br />

∂t dt + ∑<br />

1<br />

2<br />

i,j=1<br />

N−1<br />

∂ 2 Ũ<br />

∂L i (t)∂L j (t) dL ∑<br />

i(t)dL j (t) +<br />

i=1<br />

∂Ũ<br />

∂L i (t) dL i(t)<br />

The LIBOR processes dL i (t) under <strong>the</strong> term<strong>in</strong>al measure are known from (3.17), so<br />

5 also used <strong>in</strong> [Blackham, 2004]<br />

28


we can substitute <strong>in</strong>to <strong>the</strong> equation above to get:<br />

⎛<br />

dŨ(t, L) = ⎝ ∂Ũ<br />

N−1<br />

∂t + ∑<br />

1<br />

2<br />

ρ ij σ i (t)σ j (t)L i L j<br />

i,j=1<br />

N−1<br />

∑<br />

∂Ũ<br />

+ µ i (t)L i ∂L i<br />

dW QT N<br />

i=1<br />

∂ 2 Ũ<br />

∂L i ∂L j<br />

+<br />

N−1<br />

∑<br />

i=1<br />

⎞<br />

∂Ũ<br />

µ i (t)L i<br />

⎠<br />

∂L i<br />

dt<br />

(3.35)<br />

It is important to remember, that <strong>in</strong> order to comply with no-arbitrage conditions<br />

<strong>and</strong> (3.34), <strong>the</strong> numeraire-rebased process dŨ(t, L) has to be a mart<strong>in</strong>gale under term<strong>in</strong>al<br />

measure Q T N<br />

. To satisfy this requirement, <strong>the</strong> drift term dt <strong>in</strong> (3.35) must be<br />

equal to zero. The f<strong>in</strong>al PDE takes <strong>the</strong> follow<strong>in</strong>g form:<br />

∂Ũ<br />

∂t + 1 N−1<br />

∑<br />

ρ i,j σ i (t)σ j (t)L i L j<br />

2<br />

i,j=1<br />

∂ 2 Ũ<br />

+<br />

∂L i ∂L j<br />

N−1<br />

∑<br />

i=1<br />

µ i (t)L i<br />

∂Ũ<br />

∂L i<br />

= 0 (3.36)<br />

where<br />

µ i (t) =<br />

{<br />

−σ i (t) ∑ N−1 α k L k (t)<br />

k=i+1 1+α k L k (t) ρ ikσ k (t) if i < N − 1<br />

0 if i = N − 1<br />

<strong>and</strong> term<strong>in</strong>al Dirichlet boundary conditions given by <strong>the</strong> <strong>option</strong> payoff function p(·) at<br />

time T N :<br />

Ũ(T N , L) = p(L(T N )) (3.37)<br />

As a result, <strong>the</strong> derivative price at time t is given by (3.34), where Ũ satisfies (3.36).<br />

This LMM PDE imposes a relation between <strong>the</strong> time derivative (<strong>the</strong> manner <strong>in</strong> which<br />

our price changes <strong>in</strong> time) <strong>and</strong> ”space”-derivatives (that describe how numeraire-rebased<br />

price is affected by <strong>the</strong> change <strong>in</strong> one of <strong>the</strong> underly<strong>in</strong>g rates).<br />

29


Chapter 4<br />

Computational Methods<br />

In order to price derivatives, whose values are not accessible analytically, we need to<br />

employ numerical <strong>method</strong>s to approximate <strong>the</strong> solution. First, we look at Monte Carlo<br />

(MC) simulation <strong>and</strong> provide a general overview of its dist<strong>in</strong>ctive features. Second, we<br />

address <strong>the</strong> F<strong>in</strong>ite Difference (FD) approach <strong>and</strong> derive a discrete form of <strong>the</strong> LMM<br />

PDE. The ma<strong>in</strong> focus of this <strong>the</strong>sis is on <strong>the</strong> FD approach, while Monte Carlo simulation<br />

is used as a benchmark.<br />

4.1 Monte Carlo<br />

Monte Carlo Simulation approach takes advantage of <strong>the</strong> Law of Large Numbers from<br />

Probability <strong>the</strong>ory. This result states that, given n <strong>in</strong>dependent, identically distributed<br />

r<strong>and</strong>om variables with mean µ, <strong>the</strong> average of <strong>the</strong>se variables will converge to µ, as<br />

n → ∞. Let’s see how this can be applied to derivative pric<strong>in</strong>g <strong>in</strong> <strong>the</strong> framework of<br />

LMM.<br />

Simulation The risk-neutral <strong>valuation</strong> ansatz def<strong>in</strong>es <strong>the</strong> price of a derivative to be<br />

<strong>the</strong> discounted expectation of a future payoff taken under <strong>the</strong> equivalent mart<strong>in</strong>gale<br />

measure. As before, <strong>the</strong> process for a simulated forward rate is assumed to be given<br />

by (3.5). If <strong>the</strong> lifetime of a derivative is divided <strong>in</strong>to N small <strong>in</strong>tervals of length ∆t, a<br />

cont<strong>in</strong>uous path of (3.5) can be approximated more accurately 1 if we operate <strong>in</strong> terms<br />

of log L i (t). The LIBOR process for log L i (t) can be obta<strong>in</strong>ed by application of Itô’s<br />

lemma to (3.5) <strong>and</strong> <strong>the</strong> discretization follows:<br />

log L i (t + ∆t) − log L i (t) = − 1 2 σ(t)2 ∆t + σ(t) ( W (t + ∆t) − W (t) ) . (4.1)<br />

From <strong>the</strong> property 1 of Brownian Motion <strong>in</strong> section (2.2), we know that W (t + ∆t) −<br />

W (t) ∼ N (0, ∆t). Then, given <strong>the</strong> discrete process (4.1), we can easily go back to<br />

operat<strong>in</strong>g <strong>in</strong> terms of <strong>the</strong> underly<strong>in</strong>g rate:<br />

{<br />

L i (t + ∆t) = L i (t) exp − 1 2 σ i(t) 2 ∆t + σ i (t) √ }<br />

∆t ɛ i . (4.2)<br />

1 s<strong>in</strong>ce <strong>the</strong> <strong>in</strong>crement <strong>in</strong> <strong>the</strong> simulated process does not depend on <strong>the</strong> orig<strong>in</strong>al process that we are try<strong>in</strong>g<br />

to approximate.<br />

30


with ɛ i be<strong>in</strong>g <strong>in</strong>dependent st<strong>and</strong>ard normal r<strong>and</strong>om variables. In practice, we need to<br />

<strong>model</strong> <strong>the</strong> dynamics of a tenor structure of N forward rates. Pric<strong>in</strong>g under term<strong>in</strong>al<br />

measure <strong>in</strong> LMM <strong>in</strong>troduces a drift-correction term <strong>in</strong>to an <strong>in</strong>dividual LIBOR process.<br />

So, discretization of (3.17) yields:<br />

L i (T n+1 )<br />

L i (T n )<br />

= exp<br />

{(<br />

−σ i (T n )<br />

N∑<br />

k=i+1<br />

α k ρ ik σ k (T n )L k (T n )<br />

− σ i(T n ) 2 )∆t+σ i (T n ) √ }<br />

∆tɛ i<br />

1 + α k L k (T n ) 2<br />

(4.3)<br />

Hav<strong>in</strong>g simulated <strong>the</strong> relevant LIBOR paths for <strong>the</strong> underly<strong>in</strong>g rates with (4.3), we<br />

can compute sample payoff value for a derivative at maturity. In o<strong>the</strong>r words, we get a<br />

sample value of an <strong>in</strong>dependent r<strong>and</strong>om variable V (T ) that represents an <strong>option</strong> payoff.<br />

Next step, by recipe, is to repeat <strong>the</strong> sampl<strong>in</strong>g procedure n number of times <strong>and</strong> take <strong>the</strong><br />

average over <strong>the</strong> all sampled payoff values. In <strong>the</strong> limit, by <strong>the</strong> Law of Large Numbers,<br />

this average will converge to <strong>the</strong> expectation value.<br />

Recall that, <strong>in</strong> order for <strong>the</strong> risk-neutral <strong>valuation</strong> pr<strong>in</strong>ciple to hold, we must operate<br />

<strong>in</strong> terms of numeraire-rebased quantities:<br />

[ ]<br />

V (t) V (T )<br />

N(t) = EQ N(T ) ∣ F t<br />

where N is <strong>the</strong> numeraire asset <strong>and</strong> Q is <strong>the</strong> unique equivalent mart<strong>in</strong>gale measure<br />

imposed by <strong>the</strong> choice of N. Thus, we have to make sure that every sample payoff<br />

V (T ), is denom<strong>in</strong>ated <strong>in</strong> terms of <strong>the</strong> term<strong>in</strong>al bond D, represent<strong>in</strong>g our choice of<br />

numeraire <strong>in</strong> LMM. Then, by averag<strong>in</strong>g over all D-rebased sample payoff values, we<br />

shall obta<strong>in</strong> <strong>the</strong> value of <strong>the</strong> Q-expectation <strong>and</strong>, equivalently, V (t)/D(t, T ).<br />

St<strong>and</strong>ard Error If we are able to obta<strong>in</strong> <strong>the</strong> approximate value for expectation, it is<br />

easy to get a number for st<strong>and</strong>ard deviation ω. As <strong>in</strong> [Hull, 1999], <strong>the</strong> expression for a<br />

st<strong>and</strong>ard error of Monte Carlo is:<br />

ω<br />

√ n<br />

A convenient fact from [Hull, 1999] also states, that a 95% <strong>in</strong>terval of trust for <strong>the</strong><br />

derivative lies with<strong>in</strong> approximately 2 st<strong>and</strong>ard errors:<br />

µ − 1.96ω √ n<br />

< c t < µ + 1.96ω √ n<br />

Sometimes <strong>the</strong> st<strong>and</strong>ard error of MC simulation can be reduced by means of variance<br />

reduction techniques, such as anti<strong>the</strong>tic variables or control variates. Yet, this subject is<br />

outside <strong>the</strong> scope of present work. A useful reference to such techniques is [Rebonato,<br />

1998].<br />

Correlated Brownian Motions If we recall <strong>the</strong> discussion from <strong>the</strong> previous chapter<br />

on calibration to correlation between <strong>the</strong> rates, an <strong>in</strong>dividual LIBOR process can be<br />

driven by a composite Brownian motion ˆɛ i such, that:<br />

ˆɛ i =<br />

i∑<br />

α ik W k<br />

k=1<br />

31


Values for α ik are obta<strong>in</strong>ed by application of <strong>the</strong> Cholesky decomposition to a correlation<br />

matrix 2 ρ.<br />

Early Exercise Options with early-exercise feature are quite a hard nut for Monte<br />

Carlo simulation due to so-called “Monte Carlo on Monte Carlo” effect. Indeed, at<br />

every po<strong>in</strong>t of time T i , where early-exercise is allowed, one has to f<strong>in</strong>d:<br />

V (T i ) = (max[E(T i ), H(T i )]) +<br />

with E(T i ) <strong>and</strong> H(T i ) denot<strong>in</strong>g immediate exercise value <strong>and</strong> hold value at time T i ,<br />

respectively. In order to f<strong>in</strong>d H(T i ) (i.e. <strong>the</strong> expectation of <strong>the</strong> future payoff), canonical<br />

Monte Carlo <strong>method</strong> would have to run an entire simulation run at every such timepo<strong>in</strong>t<br />

for all sample paths. In contrast, recomb<strong>in</strong>ant tree <strong>and</strong> PDE <strong>method</strong>s have <strong>the</strong> value of<br />

H(T i ) available at each time-step, work<strong>in</strong>g by backward <strong>in</strong>duction.<br />

The authors of [Longstaff <strong>and</strong> Schwartz, 2001] have proposed a <strong>method</strong> which<br />

approximates <strong>the</strong> conditional expectation value H(T i ) on <strong>the</strong> basis of cross-sectional<br />

<strong>in</strong>formation stored for each simulated path. The idea of such approximation is to use a<br />

least-square fit <strong>method</strong> to regress <strong>the</strong> estimated value with a polynomial basis function<br />

of a chosen type <strong>and</strong> order.<br />

Greeks The weakest po<strong>in</strong>t of MC simulation<br />

appears to be <strong>the</strong> computation of<br />

<strong>the</strong> ’Greeks’, <strong>the</strong> quantities that relate to<br />

parameter-specific risk <strong>in</strong> an <strong>option</strong> position.<br />

We have already encountered one<br />

such sensitivity, ∆, earlier <strong>in</strong> <strong>the</strong> text. In<br />

order to compute a ∆ value with MC,<br />

<strong>the</strong> entire simulation procedure has to be<br />

repeated for a small shift ɛ <strong>in</strong> <strong>the</strong> respected<br />

underly<strong>in</strong>g asset value. The same<br />

<strong>method</strong>ology is to be applied with respect<br />

to o<strong>the</strong>r possible factors of risk, for <strong>in</strong>stance,<br />

volatility. The po<strong>in</strong>twise <strong>valuation</strong><br />

nature of Monte Carlo drastically <strong>in</strong>creases<br />

<strong>the</strong> computational costs <strong>in</strong> case of<br />

<strong>the</strong> overall sensitivity profile, characteristic<br />

of a derivative <strong>in</strong>strument.<br />

Delta value<br />

0.24<br />

0.22<br />

0.2<br />

0.18<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

10<br />

FD solution. Mesh 256x256<br />

100000 MC trials<br />

400000 MC trials<br />

1<br />

0.1<br />

0.01<br />

Shift <strong>in</strong> <strong>the</strong> underly<strong>in</strong>g rate <strong>in</strong> bp. (1bp = 0.01%)<br />

Figure 4.1: Monte Carlo. Delta of <strong>the</strong> <strong>option</strong><br />

with a digital feature.<br />

Even greater difficulties arise when <strong>the</strong> ’Greeks’ of <strong>the</strong> products with discont<strong>in</strong>uous<br />

payoff are considered, e.g. digital <strong>option</strong>s. By a simple argument, if ɛ is not large<br />

enough, <strong>the</strong> proportion of sample paths end<strong>in</strong>g up <strong>in</strong>-<strong>the</strong>-money vs. out-of-<strong>the</strong>-money<br />

is likely to rema<strong>in</strong> unchanged. Even though <strong>the</strong>re will more paths approach<strong>in</strong>g <strong>the</strong><br />

digital barrier, a sensitivity (e.g. ∆ or V) will show zero. On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, <strong>the</strong> value<br />

of <strong>the</strong> ’Greek’ letter could explode if <strong>the</strong> small change <strong>in</strong> price is divided by an even<br />

smaller ɛ, as <strong>in</strong> Figure (4.1). In general, a very large number of simulation trials <strong>and</strong><br />

2 which, for example, could be given by (3.32).<br />

0.001<br />

32


a good compromise on <strong>the</strong> size of ɛ are required to capture <strong>the</strong> reaction of MC <strong>option</strong><br />

value to small parameter shifts.<br />

4.2 F<strong>in</strong>ite Difference Method<br />

While Monte Carlo simulation rema<strong>in</strong>s <strong>the</strong> <strong>in</strong>dustry’s tool of choice for pric<strong>in</strong>g <strong>in</strong>terest<br />

rate derivatives with<strong>in</strong> LMM framework, <strong>the</strong> difficulties mentioned above motivate<br />

researchers appeal to alternatives approaches.<br />

d=2<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0<br />

1<br />

2<br />

h<br />

1<br />

}<br />

3<br />

4<br />

} h 2<br />

with:<br />

d=1<br />

h=<br />

L=<br />

m<br />

U h , L<br />

( h 1,h<br />

) t<br />

2<br />

( 2,3) t<br />

Figure 4.2: 2D Solution mesh.<br />

A numerical technique where <strong>the</strong> issues with<br />

early exercise <strong>and</strong> <strong>the</strong> greeks are solved quite<br />

naturally is <strong>the</strong> F<strong>in</strong>ite Differences Method. This<br />

approach works by approximat<strong>in</strong>g a PDE solution<br />

on a discrete mesh of po<strong>in</strong>ts. An orig<strong>in</strong>al<br />

PDE is rewritten <strong>in</strong>to a difference equation (or<br />

a system of such) to be solved ei<strong>the</strong>r directly or<br />

iteratively. Discretization can be done accord<strong>in</strong>g<br />

to one of <strong>the</strong> several discretization schemes<br />

that vary <strong>in</strong> <strong>the</strong> degree of accuracy, stability<br />

<strong>and</strong> speed of convergence [Travella <strong>and</strong> R<strong>and</strong>al,<br />

2000].<br />

Solution space. The number of subsequent <strong>in</strong>terest<br />

rates, that underlie a derivative product, determ<strong>in</strong>es <strong>the</strong> “Space”-dimensionality<br />

of our LIBOR Market PDE. In o<strong>the</strong>r words, an i-th dimension <strong>in</strong> our solution mesh<br />

is assigned to an [L m<strong>in</strong><br />

i . . . L max<br />

i ] range <strong>in</strong> an i-th forward rate. The step size <strong>in</strong> an<br />

underly<strong>in</strong>g rate L i is def<strong>in</strong>ed by a number of mesh po<strong>in</strong>ts M i <strong>in</strong> <strong>the</strong> correspond<strong>in</strong>g<br />

dimension. Let’s denote: 3<br />

• h = (h 1 , h 2 , . . . , h d ) t to be a mesh width vector of a d-dimensional mesh, where<br />

h i = Lmax i<br />

−L m<strong>in</strong><br />

i<br />

M i −1<br />

.<br />

• L i±k = (L 1 , . . . , L i ± kh i , . . . , L d ) t to be a vector of coord<strong>in</strong>ates of <strong>the</strong> po<strong>in</strong>t<br />

on <strong>the</strong> mesh.<br />

• Uh,L m to be <strong>the</strong> solution at time level m at coord<strong>in</strong>ate L on a <strong>grid</strong> with space<br />

discretization h.<br />

Time discretisation. Relation (3.37) stipulates that <strong>the</strong> boundary condition for our<br />

equation should correspond to <strong>the</strong> value of an <strong>option</strong> payoff at T N ≡ T d+1 . In reality,<br />

an eventual cash <strong>in</strong>flow from such product fixes at T d <strong>and</strong> stays unchanged for [T d , T d +<br />

α d ]. This fact allows us to restrict <strong>the</strong> time horizon to that part of <strong>the</strong> tenor structure<br />

where <strong>the</strong> dynamics of work<strong>in</strong>g forward rates do affect <strong>the</strong> outcome. Discount<strong>in</strong>g with<br />

D(t, T d+1 ), as <strong>in</strong> (3.34), ensures that extra α d time of position existence is accounted<br />

3 In this section we shall take advantage of convenient notation used <strong>in</strong>[Blackham, 2004]<br />

33


for. With all said above, it is natural for <strong>the</strong> LMM PDE to be solved backwards <strong>in</strong> time.<br />

We set τ = T d − t, ∆τ = <strong>and</strong> discretize <strong>the</strong> time derivative as:<br />

T d<br />

M−1<br />

∂u<br />

∂t = −∂u ∂τ = −Um+1 h,L<br />

− Uh,L<br />

m<br />

∆τ<br />

+ O(∆τ) (4.4)<br />

Space discretisation. For <strong>the</strong> time be<strong>in</strong>g, we have chosen second-order approximations<br />

of spatial derivatives.<br />

• First derivative (by central differences):<br />

• For cross-derivative term with i ≠ j:<br />

∂U<br />

∂L i<br />

≈ U m h,L i+1<br />

− U m h,L i−1<br />

2h i<br />

+ O(h 2 ) (4.5)<br />

∂ 2 U<br />

∂L i ∂L j<br />

≈ U m h,L i+1,j+1<br />

+ U m h,L i−1,j−1<br />

− U m h,L i+1,j−1<br />

− U m h,L i−1,j+1<br />

4h i h j<br />

+ O(h 2 )<br />

• For a second derivative term:<br />

∂ 2 U<br />

= ∂2 U<br />

∂L i ∂L i ∂L 2 i<br />

≈ U m h,L i+1<br />

− 2U m h,L + U m h,L i−1<br />

h 2 i<br />

(4.6)<br />

+ O(h 2 ) (4.7)<br />

• The discretized µ ∆ i drift-correction term is as follows:<br />

{<br />

µ ∆ −σi (m∆τ) ∑ N−1 α k ρ ik L k<br />

i (m∆τ) =<br />

k=i+1 1+α k L k<br />

σ k (m∆τ) if i < N − 1<br />

0 if i = N − 1<br />

(4.8)<br />

LMM PDE discretized. Note, that <strong>the</strong> time difference equation (4.4) is first-order<br />

accurate, while approximations of spatial derivatives (4.5), (4.6) <strong>and</strong> (4.7) are secondorder<br />

accurate. An important aspect of rewrit<strong>in</strong>g a PDE <strong>in</strong>to a difference form is <strong>the</strong><br />

choice of an accurate discretization scheme, such as Crank-Nicolson scheme, which is<br />

second-order accurate both <strong>in</strong> space <strong>and</strong> time [Crank <strong>and</strong> Nicolson, 1949].<br />

The general, so-called θ-scheme of PDE discretization <strong>in</strong>to F<strong>in</strong>ite Difference form<br />

is:<br />

U m+1<br />

h,L<br />

− Uh,L<br />

m = θW m+1<br />

h,L<br />

+ (1 − θ)Wh,L m (4.9)<br />

∆τ<br />

where W m+1<br />

h,L<br />

<strong>and</strong> Wh,L m are implicit-<strong>in</strong>-time <strong>and</strong> explicit-<strong>in</strong>-time discretizations of rema<strong>in</strong><strong>in</strong>g<br />

PDE terms.<br />

W m h,L = 1 2<br />

N−1<br />

∑<br />

i,j=1<br />

i≠j<br />

N−1<br />

1 ∑<br />

σi 2 (m∆τ)L 2 i<br />

2<br />

i=1<br />

ρ ij σ i (m∆τ)σ j (m∆τ)L i L j<br />

U m h,L i+1,j+1<br />

+U m h,L i−1,j−1<br />

−U m h,L i+1,j−1<br />

−U m h,L i+1,j−1<br />

4h ih j<br />

U m h,L i+1<br />

−2U m h,L +U m h,L i−1<br />

h 2 i<br />

+ ∑ N−1<br />

i=1 µ i(m∆τ)L i<br />

U m h,L i+1<br />

−U m h,L i−1<br />

2h i<br />

(4.10)<br />

34


Three different θ values represent three canonical discretization schemes. By sett<strong>in</strong>g<br />

θ = 0.5 for Crank-Nicolson, we shall first average <strong>and</strong> <strong>the</strong>n discrim<strong>in</strong>ate explicit <strong>and</strong><br />

implicit parts as follows:<br />

U m+1<br />

h,L<br />

− ∆τ<br />

2 W m+1<br />

h,L<br />

= Uh,L m + ∆τ<br />

2 W h,L m (4.11)<br />

As a result of such discretization we should arrive at Ax = b system of discrete stencil<br />

equations, where A is a M i ×· · ·×M d b<strong>and</strong> matrix of known coefficients, x is a vector<br />

of unknown solutions U m+1 <strong>and</strong> b is a vector of known values that correpond to <strong>the</strong><br />

right h<strong>and</strong> side of (4.11). The necessity to simultaneously solve a large system of<br />

discrete equations poses <strong>the</strong> dilemma of <strong>the</strong> choice of an efficient iterative solver, such<br />

as BiCGSTAB [van der Vorst, 1992] or Multi<strong>grid</strong> [Wessel<strong>in</strong>g, 2004].<br />

Stencil equation for 2 forward rates. Given <strong>the</strong> discrete equation for 2 forward<br />

rates, we can rearrange <strong>the</strong> terms of (4.11) to arrive at <strong>the</strong> follow<strong>in</strong>g stencil equality:<br />

−ψU m+1<br />

i−1,j−1 −βU m+1<br />

i−1,j ψU m+1<br />

i−1,j+1<br />

−ζU m+1<br />

i,j−1 (γ + 2)U m+1<br />

i,j −ηU m+1<br />

i,j+1<br />

ψU m+1<br />

i+1,j−1 −αU m+1<br />

i+1,j −ψU m+1<br />

i+1,j+1<br />

} {{ }<br />

unknowns<br />

where <strong>the</strong> known coefficients are:<br />

=<br />

ψUi−1,j−1 m βUi−1,j m −ψUi−1,j+1<br />

m<br />

ζUi,j−1 m −γUi,j m ηUi,j+1<br />

m<br />

−ψUi+1,j−1 m }<br />

αUi+1,j m {{<br />

ψUi+1,j+1<br />

m }<br />

knowns<br />

α = ∆τσ2 i (m∆τ)L2 i<br />

4h 2 i<br />

β = ∆τσ2 i (m∆τ)L2 i<br />

4h 2 i<br />

γ = ∆τσ2 i (m∆τ)L2 i<br />

2h 2 i<br />

+ ∆τµi(m∆τ)Li<br />

4h i<br />

− ∆τµ i(m∆τ)L i<br />

4h i<br />

+ ∆τσ2 j (m∆τ)L2 j<br />

2h 2 j<br />

ψ = ∆τρ ijσ i (m∆τ)σ j (m∆τ)L i L j<br />

8h ih j<br />

ζ = ∆τσ2 j (m∆τ)L2 j<br />

4h 2 j<br />

η = ∆τσ2 j (m∆τ)L2 j<br />

4h 2 j<br />

− ∆τµ j(m∆τ)L j<br />

4h j<br />

+ ∆τµ j(m∆τ)L j<br />

4h j<br />

− 1<br />

<strong>and</strong> <strong>the</strong> b<strong>and</strong> coefficient matrix A hav<strong>in</strong>g <strong>the</strong> same pattern as Figure (4.3).<br />

Forward rate expiry. While solv<strong>in</strong>g (4.9), we should keep <strong>in</strong> m<strong>in</strong>d that <strong>the</strong> lifetime<br />

of any given forward rate is bounded by its fix<strong>in</strong>g date on <strong>the</strong> tenor structure. Has such<br />

maturity of an underly<strong>in</strong>g rate been reached, <strong>the</strong> dimensionality of orig<strong>in</strong>al problem<br />

reduces by one. The simpliest <strong>and</strong> most natural way to <strong>in</strong>troduce such behavior <strong>in</strong>to<br />

<strong>valuation</strong> process is to reset forward rate’s volatility value to zero after it has matured.<br />

Therefore, by solv<strong>in</strong>g LMM PDE backwards <strong>in</strong> time, we gradually <strong>in</strong>crease <strong>the</strong> number<br />

of rates that are sett<strong>in</strong>g <strong>in</strong> motion <strong>the</strong> numerical mechanism of <strong>option</strong> <strong>valuation</strong>.<br />

Notation σ(m∆τ) <strong>and</strong> µ(m∆τ) is meant to serve as a rem<strong>in</strong>der of this feature.<br />

35


Greeks. Opposed to MC simulation, comput<strong>in</strong>g<br />

solutions for <strong>the</strong> Greeks with F<strong>in</strong>ite Difference<br />

Method comes at much lesser cost <strong>in</strong> terms<br />

of both time <strong>and</strong> accuracy. Approximations to<br />

’spatial’ derivatives are implicitly provided with<br />

<strong>the</strong> discrete PDE solution. The profiles for ∆<br />

<strong>and</strong> Γ can be extracted from <strong>the</strong> solution mesh<br />

<strong>in</strong> <strong>the</strong> form of discrete first <strong>and</strong> second derivatives,<br />

i.e (4.5) <strong>and</strong> (4.6). In order to acquire <strong>the</strong><br />

solution for a V, <strong>the</strong> numerical procedure needs<br />

to be repeated with a small shift <strong>in</strong> volatility of<br />

an underly<strong>in</strong>g rate.<br />

Figure 4.3: B<strong>and</strong> coefficient matrix.<br />

36


Chapter 5<br />

Sparse Grids<br />

In <strong>the</strong> previous chapter we discussed two computational <strong>method</strong>s that are available<br />

for pric<strong>in</strong>g of <strong>the</strong> claims with<strong>in</strong> <strong>the</strong> LMM framework. In practice, Monte Carlo is almost<br />

always <strong>in</strong> favour, s<strong>in</strong>ce its computational costs are only l<strong>in</strong>early dependant on<br />

<strong>the</strong> dimensionality of <strong>the</strong> problem. F<strong>in</strong>ite Difference Method suffers from <strong>the</strong> ’curse<br />

of dimensionality’, <strong>the</strong> number of po<strong>in</strong>ts grows exponentially with <strong>the</strong> <strong>in</strong>crease <strong>in</strong> <strong>the</strong><br />

number of underly<strong>in</strong>g variables. This only fact underm<strong>in</strong>es all <strong>the</strong> nice features FD has<br />

to offer, s<strong>in</strong>ce a great number of realistic products are cont<strong>in</strong>gent on up to 30 LIBOR<br />

rates. Sparse Grid <strong>method</strong>s, suggested by Zenger [Zenger, 1991], operate on spatial<br />

discretizations that rely on far fewer po<strong>in</strong>ts <strong>and</strong>, at <strong>the</strong> same time, offer reasonable<br />

accuracy. In this chapter we shall look at <strong>the</strong> common issues related to function approximation,<br />

<strong>in</strong>troduce Sparse Grid technique <strong>and</strong> postulate results for error analysis.<br />

At <strong>the</strong> end we shall discuss <strong>the</strong> comb<strong>in</strong>ation technique used for function <strong>in</strong>terpolation<br />

on a <strong>sparse</strong> <strong>grid</strong>. Presentation of <strong>the</strong> subject <strong>in</strong> this chapter relies on exposition <strong>and</strong><br />

notation of [Garcke, 2005].<br />

5.1 Full <strong>grid</strong> approximations.<br />

Consider a PDE of <strong>the</strong> form Lu = f, subject to certa<strong>in</strong> boundary conditions, def<strong>in</strong>ed<br />

on a d-dimensional solution space [0, 1] d . An equidistant d-dimensional <strong>grid</strong> Ω l could<br />

<strong>the</strong>n be taken as a discrete approximation of such solution space, with :<br />

• l = (l 1 , . . . , l d ) ∈ N d as a multi-<strong>in</strong>dex, i.e. discretization resolution of Ω l<br />

• h l := (h l1 , . . . , h ld ) := (2 −l 1<br />

, . . . , 2 −l d<br />

) as a mesh width<br />

<strong>in</strong> each coord<strong>in</strong>ate direction t ∈ [1 . . . d].<br />

Given Ω l , <strong>the</strong>n <strong>the</strong> discretized PDE, denoted as L l u l = f l , can be approximated<br />

with ei<strong>the</strong>r st<strong>and</strong>ard basis or hierarchical basis functions.<br />

37


(j-1)h (j+1)h<br />

0<br />

1<br />

Figure 5.1: St<strong>and</strong>ard Basis functions S 3<br />

St<strong>and</strong>ard basis functions. Let’s def<strong>in</strong>e a correspond<strong>in</strong>g space S l of piecewise d-<br />

l<strong>in</strong>ear basis functions as:<br />

S l := span{f l,j | j t = 0, . . . , 2 l t<br />

, t = 1, . . . , d} (5.1)<br />

where <strong>the</strong> d-l<strong>in</strong>ear piecewise function f l,j is obta<strong>in</strong>ed by a tensor product operation<br />

over simple l<strong>in</strong>ear ’pagoda’ functions:<br />

f l,j :=<br />

d⊗<br />

f l,j (5.2)<br />

Such one-dimensional functions have support [x j − h l , x j + h l ] ∩ [0, 1], as illustrated<br />

<strong>in</strong> Figure (5.1) above, <strong>and</strong> have <strong>the</strong> follow<strong>in</strong>g def<strong>in</strong>ition:<br />

t=1<br />

{<br />

1 − |x/hl − j| x ∈ [(j − 1)h<br />

f l,j (x) =<br />

l , (j + 1)h l ] ∩ [0, 1]<br />

0 o<strong>the</strong>rwise<br />

(5.3)<br />

Hierarchical basis functions. One of <strong>the</strong> ma<strong>in</strong> advantages of hierarchical bases over<br />

<strong>the</strong> st<strong>and</strong>ard nodal po<strong>in</strong>t bases described above is <strong>the</strong> fact that <strong>the</strong> approximat<strong>in</strong>g function<br />

space acquires a multi-level structure. By us<strong>in</strong>g <strong>the</strong> hierarchy of basis functions<br />

we can dist<strong>in</strong>guish between high-level <strong>and</strong> low-level representations, i.e., bases that<br />

conta<strong>in</strong> a significant part of <strong>the</strong> <strong>in</strong>formation <strong>and</strong> those where approximation is quite<br />

rough [Bungartz <strong>and</strong> Dornseifer, 1998], respectively.<br />

Let’s denote :<br />

• |l| 1 := ∑ d<br />

t=1 l t to be a discrete multi-<strong>in</strong>dex L 1 -norm.<br />

• |l| ∞ := max 1≤t≤d l t to be a discrete multi-<strong>in</strong>dex L ∞ -norm.<br />

The first step on <strong>the</strong> way to hierachical representation is obta<strong>in</strong><strong>in</strong>g a hierarchical difference<br />

space T l from <strong>the</strong> stardard base of <strong>the</strong> same level:<br />

T l = S l \<br />

d⊕<br />

S l−et (5.4)<br />

where e t is <strong>the</strong> unit vector <strong>in</strong> t-th dimension. The result<strong>in</strong>g function space consists<br />

of basis functions that are characteristic of <strong>the</strong> current level l <strong>and</strong> are not a part of <strong>the</strong><br />

t=1<br />

38


T<br />

0 , 0<br />

T<br />

0 , 1<br />

T<br />

0 , 2<br />

T<br />

0 , 3<br />

|i|<br />

1 = 0<br />

T<br />

1 ,0<br />

T<br />

1 , 1<br />

T<br />

1 ,2<br />

· · ·<br />

|i|<br />

1<br />

= 1<br />

T<br />

1 ,3<br />

T<br />

2 ,0<br />

T<br />

2 , 1<br />

·<br />

·<br />

T<br />

2 ,2<br />

T<br />

2 , 3<br />

|i|<br />

1 = 2<br />

T<br />

3 ,0<br />

|i|<br />

1 = 3<br />

|i|<br />

1 = 4<br />

T<br />

3 , 1<br />

|i|<br />

1<br />

= 5<br />

T<br />

3 ,2<br />

|i|<br />

1<br />

= 6<br />

T<br />

3 , 3<br />

Figure 5.2: Hierarchical representation.<br />

spaces with a smaller |l| 1 value. For example, apply<strong>in</strong>g (5.4) to S 3 should yield l<strong>in</strong>ear<br />

images of spaces labeled T 3,0 or, T 0,3 on Figure (5.2). Note that, unlike S 3 , <strong>the</strong> support<br />

of <strong>the</strong> basis functions of T 3 is not overlapp<strong>in</strong>g.<br />

Now we can use (5.4) to obta<strong>in</strong> <strong>the</strong> f<strong>in</strong>al n-level hierarchy of subspaces S ∗ n:<br />

S ∗ n =<br />

n⊕<br />

· · ·<br />

l 1=0<br />

n⊕<br />

T l =<br />

l d =0<br />

⊕<br />

|l| ∞ ≤n<br />

T l (5.5)<br />

’≤’ corresponds to <strong>in</strong>dex-wise relation <strong>in</strong> <strong>the</strong> subscript of <strong>the</strong> direct sum operator. Figure<br />

(5.2) illustrates such multi-level subspace decompostion S3 ∗ <strong>in</strong> two dimensions.<br />

Figure (5.3) shows how hierarchical subspaces T l1,l 2<br />

reconstruct Ω 3,3 , where each tuple<br />

relates to <strong>the</strong> subspace <strong>in</strong>dex. As stated <strong>in</strong> <strong>the</strong> beg<strong>in</strong>n<strong>in</strong>g of this chapter, our orig<strong>in</strong>al<br />

<strong>in</strong>tent is to approximate a PDE solution space. The equation of our <strong>in</strong>terest, given<br />

by (3.36), is subject to Dirichlet boundary condition, which implies no degrees of freedom<br />

exist on <strong>the</strong> boundaries of solution doma<strong>in</strong>. For this specific reason, subspaces<br />

with l = 0 <strong>in</strong>dex components have not been <strong>in</strong>cluded <strong>in</strong> <strong>the</strong> hiearchical representation<br />

of (5.3).<br />

39


33 23 33 13 33 23 33<br />

32 22 32 12 32 22 32<br />

33 23 33 13 33 23 33<br />

31 21 31 11 31 21 31<br />

33 23 33 13 33 23 33<br />

32 22 32 12 32 22 32<br />

33 23 33 13 33 23 33<br />

Figure 5.3: Mapp<strong>in</strong>g of hierarchical representation S ∗ 3 onto a full <strong>grid</strong>.<br />

At this po<strong>in</strong>t, a reasonable question to ask <strong>in</strong> <strong>the</strong> context of our PDE problem would<br />

be: how accurate is <strong>the</strong> Sl ∗ -approximation of a solution u, given that u is a smooth,<br />

cont<strong>in</strong>uous function on Ω?<br />

Here we shall provide <strong>the</strong> result for such error analysis <strong>in</strong> case of a two dimensional<br />

discretized PDE L i,j u i,j = f i,j . Results for d-dimensional case can be easily derived.<br />

The follow<strong>in</strong>g def<strong>in</strong>itions will apply:<br />

• The approximate solution of u on Si,j ∗ , denoted with u∗ i,j , can be decomposed as:<br />

u ∗ i,j =<br />

i⊕<br />

j⊕<br />

s=1 t=1<br />

u s,t<br />

• As <strong>in</strong> [Zenger, 1991], we shall def<strong>in</strong>e a semi-norm:<br />

∂4 u<br />

|u| = ∥ ∥ ∥∞<br />

∂x 2 ∂y 2<br />

∂<br />

assum<strong>in</strong>g that a mixed derivative<br />

4 u<br />

∂x 2 ∂y<br />

of u exists <strong>and</strong> is a cont<strong>in</strong>uous function<br />

2<br />

itself.<br />

• In <strong>the</strong> orig<strong>in</strong>al paper [Zenger, 1991], <strong>the</strong> follow<strong>in</strong>g a property of <strong>the</strong> representation<br />

<strong>in</strong> <strong>the</strong> product-type basis has been shown:<br />

∥<br />

∥u i,j<br />

∥<br />

∥∞ ≤ 4 −i−j−1 |u|<br />

Now, by us<strong>in</strong>g properties of <strong>in</strong>f<strong>in</strong>ite series <strong>and</strong> matrix norms, it can be shown that <strong>the</strong><br />

error of approximation is<br />

∥<br />

∥u − u ∗ ∥<br />

i,j ∞<br />

= · · · = 1 (<br />

)<br />

36 |u| h 2 i + h 2 j + h 2 i h 2 j<br />

(5.6)<br />

giv<strong>in</strong>g <strong>the</strong> convergence order of <strong>in</strong>terpolation of <strong>the</strong> function with hierachical basis<br />

function space as O(h 2 i + h2 j ), which is second-order convergence. For <strong>the</strong> derivation<br />

of this result an <strong>in</strong>terested reader is referred to [Griebel et al., 1992] or [Zenger, 1991].<br />

However, representation of a function <strong>in</strong> a hiearchical basis still suffers from <strong>the</strong><br />

’curse of dimensionality’. As argued <strong>in</strong> [Garcke, 2005], <strong>the</strong> number of basis functions,<br />

which describe a u ∈ Sl ∗, <strong>in</strong> hierarchical basis is (2l + 1) d , where l is <strong>the</strong> level of<br />

discretization <strong>in</strong> each of <strong>the</strong> d dimensions. For <strong>in</strong>stance, tak<strong>in</strong>g l = 4 <strong>and</strong> d = 10, <strong>the</strong><br />

number of basis functions is 2 × 10 12 !<br />

40


00 03 02 03 01 03 02 03 00<br />

30 30<br />

20 12 20<br />

30 30<br />

10 21 11 21 10<br />

30 30<br />

20 12 20<br />

30 30<br />

00 03 02 03 01 03 02 03 00<br />

Figure 5.4: Mapp<strong>in</strong>g of hierarchical representation Ŝ∗ 3 onto a <strong>sparse</strong> <strong>grid</strong>.<br />

5.2 Sparse <strong>grid</strong> approximations.<br />

In order to decrease <strong>the</strong> amount of po<strong>in</strong>ts <strong>in</strong>volved <strong>in</strong> <strong>the</strong> approximation, we turn to <strong>the</strong><br />

<strong>sparse</strong> discretization suggested <strong>in</strong> [Zenger, 1991]. Our solution space is chosen such<br />

that <strong>the</strong> hierarchical basis functions with <strong>the</strong> small support <strong>and</strong>, <strong>the</strong>refore, a small contribution<br />

to <strong>the</strong> f<strong>in</strong>al approximation, are not <strong>in</strong>cluded <strong>in</strong> <strong>the</strong> result<strong>in</strong>g discrete space. The<br />

authors of <strong>the</strong> orig<strong>in</strong>al paper suggested <strong>the</strong> follow<strong>in</strong>g comb<strong>in</strong>ation of basis functions:<br />

Ŝn ∗ = ⊕<br />

T l (5.7)<br />

|l| 1≤n<br />

The required set of difference spaces is acquired by substitution of L ∞ -norm of (5.5)<br />

for L 1 -norm <strong>in</strong> <strong>the</strong> direct sum condition of (5.7). Figure (5.2) shows difference spaces<br />

T i,j compris<strong>in</strong>g Ŝ∗ 3 outl<strong>in</strong>ed with <strong>the</strong> bold frame. The lower right triangular section<br />

provides discretization details smaller than required tolerance <strong>and</strong>, thus, is excluded.<br />

The set of spaces T i,j which meet <strong>the</strong> |l| 1 ≤ n condition is called a <strong>sparse</strong> <strong>grid</strong>. Accord<strong>in</strong>g<br />

to [Garcke, 2005], its number of degrees of freedom equals:<br />

|Ŝ∗ n| = O(h −1<br />

n<br />

log(h −1<br />

n ) d−1 ) (5.8)<br />

while <strong>the</strong> <strong>in</strong>terpolation error, by similar analysis as <strong>in</strong> (5.6), shows<br />

∥<br />

∥u − û ∗ ∥<br />

n,n ∞<br />

= 1 (<br />

48 |u| h2 n log(h −1<br />

n ) + 4 )<br />

3<br />

(5.9)<br />

For <strong>the</strong> proof, an <strong>in</strong>terested reader is referenced to [Zenger, 1991].<br />

As one can tell from (5.9) <strong>and</strong> (5.8), <strong>the</strong> drastic reduction <strong>in</strong> <strong>the</strong> number of degrees<br />

of freedom is complemented by an <strong>in</strong>significant <strong>in</strong>crease <strong>in</strong> <strong>the</strong> <strong>in</strong>terpolation error. For<br />

<strong>the</strong> 2D case, <strong>the</strong> use of Ŝ∗ n <strong>in</strong>stead of Sn ∗ decreases <strong>the</strong> number of po<strong>in</strong>ts from O(h −2<br />

n ) to<br />

O(h −1<br />

n log(h −1<br />

n )). At <strong>the</strong> same time, <strong>the</strong> accuracy of representation stays at reasonable<br />

O(h 2 nlog(h −1<br />

n )), as compared to O(h 2 n).<br />

Figure (5.4) shows <strong>the</strong> result<strong>in</strong>g <strong>sparse</strong> <strong>grid</strong> space, obta<strong>in</strong>ed by application of (5.7)<br />

to Sn ∗ of Figure (5.2). Boundary spaces have been kept <strong>in</strong> order to populate a very<br />

<strong>sparse</strong> space.<br />

41


5.3 A Comb<strong>in</strong>ation technique<br />

So far, we’ve had a d-l<strong>in</strong>ear basis function as <strong>the</strong> unit of approximation on a discrete<br />

function space. The <strong>sparse</strong> <strong>grid</strong> comb<strong>in</strong>ation technique, <strong>in</strong>troduced <strong>in</strong> [Griebel et al.,<br />

1992], employes a different technique lead<strong>in</strong>g to <strong>the</strong> function representation on a <strong>sparse</strong><br />

<strong>grid</strong>. Similar to Richardson extrapolation, it is based on <strong>the</strong> pr<strong>in</strong>ciple of multi-variate<br />

extrapolation <strong>and</strong> relies on <strong>the</strong> fact that lead<strong>in</strong>g-order discretization errors cancel out as<br />

<strong>the</strong> result of l<strong>in</strong>ear comb<strong>in</strong>ation of <strong>grid</strong>s. The comb<strong>in</strong>ation technique uses a node-based<br />

discretization, as opposed to function-based one.<br />

The <strong>grid</strong>s are comb<strong>in</strong>ed as follows 1 :<br />

• 2D case :<br />

• 3D case :<br />

• dD case:<br />

û c n,n =<br />

∑<br />

u i,j − ∑<br />

u i,j (5.10)<br />

i+j=n+1 i+j=n<br />

û c n,n =<br />

∑<br />

∑<br />

u i,j − 2 u i,j + ∑<br />

u i,j (5.11)<br />

i+j=n+2<br />

i+j=n+1 i+j=n<br />

∑d−1<br />

( ) d − 1 ∑<br />

û c n,n = (−1) k k<br />

k=0<br />

|l| 1=n−k<br />

u |l| (5.12)<br />

First, PDE is solved on <strong>the</strong> <strong>in</strong>dividual <strong>grid</strong>s by means of a numerical <strong>method</strong> (e.g.,<br />

f<strong>in</strong>ite difference <strong>method</strong>, as <strong>in</strong> Chapter 4). Second, solutions are l<strong>in</strong>early comb<strong>in</strong>ed to<br />

produce a <strong>in</strong>terpolated approximation to u.<br />

Essencially, <strong>the</strong> comb<strong>in</strong>ed approximation of a solution to <strong>the</strong> PDE u c l<br />

is not <strong>the</strong><br />

same as <strong>the</strong> u ∗ l<br />

. However, it can be shown [Griebel et al., 1992] via asymptotic error<br />

expansion for 2D, that <strong>the</strong> estimation is of <strong>the</strong> same order of accuracy:<br />

∥<br />

∥ u − û<br />

c ∥∞ n,n ≤ Kh 2 l (1 + 5 4 log 2(h −1<br />

l<br />

)) = O(h −1<br />

n log(h −1<br />

n )) (5.13)<br />

where K is some constant <strong>and</strong> h l = 2 −l . Consider referr<strong>in</strong>g to [Blackham, 2004] for a<br />

d-dimensional error analysis.<br />

Recently, advances have been made <strong>in</strong> <strong>the</strong> subject of comb<strong>in</strong>ation techniques for<br />

<strong>sparse</strong> <strong>grid</strong>s. Novel comb<strong>in</strong>ation approaches have been <strong>in</strong>troduced <strong>in</strong> [Leentvaar <strong>and</strong><br />

Oosterlee, 2006] or [Reis<strong>in</strong>ger, 2004].<br />

1 here we usePnotation for a direct sum operator to emphasize <strong>the</strong> node-based nature of discretization,<br />

used <strong>in</strong> comb<strong>in</strong>ation technique, <strong>in</strong>stead of <strong>the</strong> commonLdirect sum notation<br />

42


Chapter 6<br />

Numerical Results. Valuation<br />

<strong>and</strong> <strong>the</strong> Greeks<br />

In Chapter 1 we have outl<strong>in</strong>ed <strong>the</strong> list of objectives for <strong>the</strong> practical part of this <strong>the</strong>sis.<br />

Accord<strong>in</strong>g to <strong>the</strong> first objective, our <strong>in</strong>tention is to revisit <strong>valuation</strong> results of [Blackham,<br />

2004]. Therefore, we choose to price a chooser <strong>option</strong> <strong>and</strong> a bermudan swaption<br />

products <strong>in</strong> <strong>the</strong> sett<strong>in</strong>g specified <strong>in</strong> [Blackham, 2004]. Even though real-world LMM<br />

derivatives are cont<strong>in</strong>gent on a larger number of forward rates, it has been decided to<br />

constra<strong>in</strong> <strong>the</strong> scope of work by look<strong>in</strong>g <strong>in</strong>to two <strong>and</strong> three rate cases.<br />

As a start<strong>in</strong>g po<strong>in</strong>t for each case, <strong>the</strong> convergence behaviour of <strong>the</strong> price is studied<br />

by means of Monte Carlo simulation. The follow<strong>in</strong>g step <strong>in</strong>volves implementation<br />

of a f<strong>in</strong>ite difference solver to obta<strong>in</strong> a PDE solution on a full mesh of po<strong>in</strong>ts, while<br />

earlier Monte Carlo results serve as a useful benchmark. Convergence of a <strong>sparse</strong> <strong>grid</strong><br />

solution, obta<strong>in</strong>ed with <strong>the</strong> comb<strong>in</strong>ation technique of Chapter 5, is <strong>the</strong>n compared with<br />

a full <strong>grid</strong> approximation. As required by our orig<strong>in</strong>al motivation, fur<strong>the</strong>r tests <strong>in</strong>clude<br />

solv<strong>in</strong>g for different strike <strong>and</strong> spot rates <strong>and</strong> tak<strong>in</strong>g a closer look at <strong>the</strong> ’Greeks’ (∆<br />

<strong>and</strong> V).<br />

F<strong>in</strong>ally, we address <strong>the</strong> case of a discont<strong>in</strong>uous payoff <strong>and</strong> present results for <strong>valuation</strong><br />

<strong>and</strong> <strong>the</strong> ’Greeks’ of a 2D digital <strong>option</strong>.<br />

Setup of Pric<strong>in</strong>g Framework Before any <strong>valuation</strong> can take place, one needs to calibrate<br />

<strong>the</strong> pricer code to <strong>the</strong> real <strong>market</strong> data. The same calibration dataset is used<br />

as <strong>in</strong> [Blackham, 2004]. Movements <strong>in</strong> <strong>the</strong> underly<strong>in</strong>g rates are restricted to <strong>the</strong> doma<strong>in</strong><br />

1 of [0.0, 0.15] d . Volatility term structure is assumed flat, mean<strong>in</strong>g <strong>the</strong> value of an<br />

<strong>in</strong>stantaneous volatility function rema<strong>in</strong>s constant throughout <strong>the</strong> tenor period. Oneyear<br />

LIBOR forward rates <strong>and</strong> correspond<strong>in</strong>g caplet volatilities are provided <strong>in</strong> (6.2).<br />

The amount of cash that underlies ga<strong>in</strong>s or losses <strong>in</strong> <strong>in</strong>terest rate movements (a notional<br />

amount) is chosen to be 10000. The correlation matrix is given by formula (3.32). We<br />

assume strong correlation between <strong>the</strong> neighbour<strong>in</strong>g rates, which can be achieved by<br />

sett<strong>in</strong>g β = 0.1. The follow<strong>in</strong>g two tables give a summary of our setup:<br />

1 unless stated o<strong>the</strong>rwise<br />

43


Currency EUR<br />

Tenor size α i 1 year<br />

Strike K 5.5%<br />

Volatility Constant<br />

Correlation parameter β 0.1<br />

Forward rate range L m<strong>in</strong> − L max 0%-15%<br />

Notional amount 10000 e<br />

Table 6.1: Setup of LMM. Input data for LMM<br />

Start date End date Forward LIBOR Rate Volatility %<br />

T 0 29.07.04 29.07.05 2.423306 0<br />

T 1 29.07.05 29.07.06 3.281384 24.73<br />

T 2 29.07.06 29.07.07 3.931690 22.45<br />

T 3 29.07.07 29.07.08 4.364818 19.36<br />

T 4 29.07.08 29.07.09 4.680236 17.43<br />

T 5 29.07.09 29.07.10 4.933085 16.15<br />

T 6 29.07.10 29.07.11 5.135066 15.02<br />

T 7 29.07.11 29.07.12 5.273314 14.24<br />

Table 6.2: Setup of LMM. LIBOR rates <strong>and</strong> caplet volatilies<br />

6.1 Two Dimensional Case. 2D Chooser Option<br />

A chooser <strong>option</strong> is a European-style derivative product, whose payoff is determ<strong>in</strong>ed<br />

by (3.28). A two-rate chooser payoff follows:<br />

V Cho2D<br />

3 = α 2 (max(L 1 , L 2 ) − K) +<br />

All action takes place on a T 0 , ..., T 3 subset of a tenor structure. The relevant discount<br />

bond prices can be calculated accord<strong>in</strong>g to (3.4).<br />

A few words need to be said about impos<strong>in</strong>g boundary conditions on 2D solution<br />

space for FD. Two of <strong>the</strong> four boundaries, correspond<strong>in</strong>g to maximum values <strong>in</strong> underly<strong>in</strong>g<br />

rates, are always fixed to <strong>the</strong> maximum possible payoff value: L max − K. As<br />

of <strong>the</strong> rema<strong>in</strong><strong>in</strong>g boundaries, with one of <strong>the</strong> rates be<strong>in</strong>g at its m<strong>in</strong>imum L m<strong>in</strong> , <strong>the</strong>re<br />

are two possible scenarios. In <strong>the</strong> first case, solution can be considered fixed to <strong>the</strong><br />

<strong>in</strong>tr<strong>in</strong>sic value of <strong>the</strong> <strong>option</strong>. The second case proceeds by solv<strong>in</strong>g a lower dimensional<br />

problem with respect to <strong>the</strong> non-L m<strong>in</strong> rate <strong>and</strong> current time step. This solution is conveniently<br />

given by <strong>the</strong> analytical Black formula. Even though <strong>the</strong> second approach<br />

seems more logical <strong>in</strong>tuitively, experience so far shows that accuracy of f<strong>in</strong>al solutions<br />

wasn’t significantly affected by preference of one <strong>method</strong> over ano<strong>the</strong>r. In order to<br />

rema<strong>in</strong> confident <strong>in</strong> this fact, we cont<strong>in</strong>ued to use <strong>the</strong>m <strong>in</strong>terchangably.<br />

It was convenient to start with an out-of-<strong>the</strong>-money sett<strong>in</strong>g (strike of 5.5% <strong>and</strong><br />

time-zero forward rates as provided <strong>in</strong> Table (6.2)), s<strong>in</strong>ce <strong>the</strong>re are results available for<br />

this specific case <strong>in</strong> [Blackham, 2004].<br />

44


6.1.1 Convergence of MC, Full <strong>and</strong> Sparse Grid solutions.<br />

MC convergence <strong>and</strong> its st<strong>and</strong>ard error are shown <strong>in</strong> Figure (6.1). Simulation was<br />

performed for <strong>the</strong> maximum of 4, 000, 000 trials, reach<strong>in</strong>g m<strong>in</strong>imum st<strong>and</strong>ard error of<br />

0.022. FD solution for full <strong>grid</strong> discretization was computed for up to discretization<br />

level L = 9. Figure (6.1) shows conv<strong>in</strong>c<strong>in</strong>g convergence behavior of simulated solution<br />

Level of Discretization<br />

6 6.5 7 7.5 8 8.5 9<br />

10.2<br />

MC Simulation Value+Std.Error<br />

Full-Grid FD value<br />

10.15<br />

10.1<br />

Option price<br />

10.05<br />

10<br />

9.95<br />

9.9<br />

9.85<br />

500000 1e+006 1.5e+006 2e+006 2.5e+006 3e+006 3.5e+006 4e+006<br />

Number of MC simulations<br />

Figure 6.1: 2D Chooser Option. MC vs. FD Convergence.<br />

<strong>and</strong> full-<strong>grid</strong> FD, provid<strong>in</strong>g accurate reference values for fur<strong>the</strong>r tests. For our <strong>sparse</strong><br />

approximation experiments we have used a classical <strong>sparse</strong> <strong>grid</strong> (a.k.a ”maximumnorm-based”<br />

type) <strong>and</strong> a basic comb<strong>in</strong>ation technique, both described <strong>in</strong> <strong>the</strong> previous<br />

chapter. Table (6.3) <strong>and</strong> Figure (6.2) give idea of FD convergence on full <strong>and</strong> <strong>sparse</strong><br />

<strong>grid</strong>s 2 :<br />

Level<br />

Full <strong>grid</strong> Execution Sparse Grid Execution Full <strong>grid</strong> Sparse <strong>grid</strong><br />

solution time(sec.) solution time(sec.) po<strong>in</strong>ts count po<strong>in</strong>ts count<br />

5 10.5969 1.5 11.351 0.36 1089 257<br />

6 10.1755 6.6 11.477 0.88 4225 577<br />

7 10.0964 30.1 9.8731 2.16 16641 1281<br />

8 10.0762 176.4 10.005 5.18 66049 2817<br />

9 10.0762 651.4 10.165 12.89 263169 6145<br />

10 10.153 31.64 1050625 13313<br />

11 9.998 74.86 4198401 28673<br />

Table 6.3: 2D Chooser. Convergence values for full <strong>and</strong> <strong>sparse</strong> <strong>grid</strong> solutions.<br />

2 The execution time for <strong>sparse</strong> <strong>grid</strong> technique does not <strong>in</strong>clude <strong>in</strong>terpolation procedure<br />

45


11.8<br />

11.6<br />

11.4<br />

Black-Scholes Boundary<br />

Fixed Boundary<br />

Full Grid, Level 9 Value<br />

11.2<br />

Option price<br />

11<br />

10.8<br />

10.6<br />

10.4<br />

10.2<br />

10<br />

9.8<br />

6 7 8 9 10 11<br />

Discretization Level<br />

Figure 6.2: 2D Chooser. Convergence of Full <strong>and</strong> Sparse Grid.<br />

Level<br />

L 1 =L 2 =5.5%, K=5.5% K = 3.9% L 1 =L 2 =7.5%, K=5.5%<br />

Full Sparse Full Sparse Full Sparse<br />

5 83.669 71.9602 49.325 37.3126 251.755 262.1945<br />

6 82.947 83.0537 48.869 37.9287 251.535 258.5324<br />

7 82.547 81.1169 48.818 41.3625 251.449 253.7469<br />

8 82.551 79.0752 48.790 46.9370 251.438 253.1529<br />

9 82.4895 48.0288 251.9835<br />

10 82.5755 48.0717 251.8426<br />

MC<br />

Price Std.error Price Std.error Price Std.error<br />

82.399 0.01298 48.653 0.0089 251.468 0.0218<br />

Table 6.4: 2D Chooser. At-<strong>the</strong>-money <strong>and</strong> <strong>in</strong>-<strong>the</strong>-money <strong>option</strong> values.<br />

Level<br />

L 1 =L 2 =2%, K=5.5% L 1 =L 2 =2%, K=2%<br />

Full Sparse Full Sparse<br />

5 0.0572 0.5587 32.7686 23.4880<br />

6 0.0269 0.0239 31.2000 28.7051<br />

7 0.0202 0.0701 30.1360 31.3129<br />

8 0.0193 0.1163 30.1523 28.5730<br />

9 0.0383 28.5550<br />

10 0.0246 29.8075<br />

MC<br />

Price Std.error Price Std.error<br />

0.0194 0.000153 30.0825 0.00472<br />

Table 6.5: 2D Chooser. Deep out-of-<strong>the</strong>-money <strong>and</strong> at-<strong>the</strong>-money with low strike.<br />

In addition to <strong>the</strong> out-of-<strong>the</strong> money case from [Blackham, 2004], we were <strong>in</strong>terested<br />

<strong>in</strong> <strong>the</strong> convergence of <strong>sparse</strong> solution <strong>in</strong> case of different strike prices <strong>and</strong> different<br />

values of spot forward rates. The results are presented <strong>in</strong> Tables (6.4) <strong>and</strong> (6.5), where<br />

MC prices have been calculated for 500000 simulation trials.<br />

46


Full Solution. Level 6<br />

Option Price<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

0 100<br />

200<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

Forward Rate L2 0.04<br />

0.02<br />

0<br />

0<br />

0.120.140.16<br />

Forward Rate L1<br />

0.020.040.060.080.1<br />

(a) Full Grid Surface.<br />

Sparse Grid. Level 6<br />

Option Price<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

900<br />

800<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

Forward Rate L2 0.04<br />

0.02<br />

0<br />

0<br />

0.120.140.16<br />

Forward Rate L1<br />

0.020.040.060.080.1<br />

(b) Sparse Grid Surface.<br />

Figure 6.3: 2D Chooser. Solution Surface for level 6.<br />

47


900<br />

800<br />

Full Solution<br />

Sparse Solution<br />

450<br />

400<br />

Full Solution<br />

Sparse Solution<br />

700<br />

350<br />

600<br />

300<br />

Option value<br />

500<br />

400<br />

Option value<br />

250<br />

200<br />

300<br />

150<br />

200<br />

100<br />

100<br />

50<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

(a) Option Values for L 2 =3.75%<br />

(b) Plot (a) zoomed <strong>in</strong>.<br />

900<br />

800<br />

Full Solution<br />

Sparse Solution<br />

450<br />

400<br />

Full Solution<br />

Sparse Solution<br />

700<br />

350<br />

600<br />

300<br />

Option value<br />

500<br />

400<br />

Option value<br />

250<br />

200<br />

300<br />

150<br />

200<br />

100<br />

100<br />

50<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

(c) Option Values for L 2 =7.5%<br />

(d) Plot (c) zoomed <strong>in</strong>.<br />

Figure 6.4: 2D Chooser. Sparse vs. full solution. Slices.<br />

The Figure (6.4) above shows how Level 6 <strong>sparse</strong>-<strong>grid</strong> approximated solution corresponds to<br />

<strong>the</strong> full <strong>grid</strong> counterpart as we freeze one of <strong>the</strong> forward rates.<br />

48


6.1.2 Discussion: Convergence of Pric<strong>in</strong>g.<br />

Results from Tables (6.3), (6.4) <strong>and</strong> (6.5) prove that <strong>the</strong> use of BiCGSTAB+ILU iterative<br />

solver with second order FD <strong>method</strong> manages to produce promis<strong>in</strong>g results <strong>in</strong><br />

case of <strong>sparse</strong> discretization. In addition to impressive improvement <strong>in</strong> speed, <strong>the</strong> “per<br />

<strong>grid</strong> po<strong>in</strong>t“ convergence of <strong>sparse</strong> <strong>grid</strong> PDE solution is by far better than that of a full<br />

<strong>grid</strong>. This feature of <strong>sparse</strong> <strong>grid</strong> technique proves to be crucial when approximat<strong>in</strong>g<br />

functions <strong>in</strong> higher dimensional spaces.<br />

In order to ga<strong>in</strong> a better <strong>in</strong>sight on accuracy of <strong>the</strong> solution we have plotted <strong>the</strong><br />

slices of solution surface, where full <strong>and</strong> <strong>sparse</strong> approximations were compared aga<strong>in</strong>st<br />

each o<strong>the</strong>r. Figure (6.4) presents two of such slices, correspond<strong>in</strong>g to L 2 = 7.5% <strong>and</strong><br />

L 2 = 3.75% fix<strong>in</strong>gs. On a broader range both plots show relatively strong agreement<br />

of both full <strong>and</strong> <strong>sparse</strong> approximations. Yet, as we zoom closer, <strong>the</strong> plot(d) exhibits<br />

<strong>in</strong>accuracies that form a dist<strong>in</strong>guishable “wave” pattern, <strong>and</strong> less so plot(b).<br />

The orig<strong>in</strong> of this pattern is better illustrated by Figure (6.5), that provides an <strong>in</strong>sight<br />

<strong>in</strong>to <strong>the</strong> absolute error distribution, characteristic of a <strong>sparse</strong> <strong>grid</strong> solution on level 8.<br />

It shows that <strong>the</strong> largest differences between solutions appear along <strong>the</strong> center of <strong>the</strong><br />

doma<strong>in</strong> <strong>and</strong> <strong>in</strong> <strong>the</strong> region of discont<strong>in</strong>uities of <strong>in</strong>itial boundary conditions.<br />

Error Profile. Level 8<br />

25<br />

20<br />

15<br />

10<br />

-10<br />

-505<br />

-15<br />

Error<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

Forward Rate L2 0.04<br />

0.02<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04 Forward Rate L1<br />

0.02<br />

Figure 6.5: 2D Chooser. Absolute error surface on level 8.<br />

S<strong>in</strong>ce all <strong>grid</strong>s of <strong>sparse</strong> discretization are treated <strong>in</strong>dependently by our iterative<br />

solver, a discretization error is to be <strong>in</strong>troduced while solv<strong>in</strong>g subproblems on ”narrower”<br />

<strong>and</strong> borderl<strong>in</strong>e <strong>grid</strong>s of <strong>sparse</strong> space (as <strong>in</strong> Figure (5.2)), where spatial step h i<br />

<strong>in</strong> one dimension is significantly larger than <strong>in</strong> <strong>the</strong> o<strong>the</strong>r. In <strong>the</strong> process of comb<strong>in</strong>ation,<br />

some po<strong>in</strong>ts rema<strong>in</strong> less-<strong>in</strong>fluenced or unaffected at all by addition <strong>and</strong> subtraction<br />

of “<strong>sparse</strong>r” <strong>grid</strong>s, <strong>and</strong> <strong>the</strong>refore <strong>in</strong>troduc<strong>in</strong>g “spikes” <strong>in</strong>to solution. For <strong>in</strong>stance, <strong>the</strong><br />

volatile center of <strong>the</strong> solution is such due to <strong>the</strong> po<strong>in</strong>ts contributed by <strong>the</strong> “narrowest”<br />

49


<strong>grid</strong>s with <strong>the</strong> largest multi-<strong>in</strong>dex value.<br />

6.1.3 Greeks: Delta<br />

In Chapter 2, we have discussed sett<strong>in</strong>g up a hedge portfolio Π t . It has been made<br />

clear that f<strong>in</strong>d<strong>in</strong>g a process ∆ t for hold<strong>in</strong>gs <strong>in</strong> an underly<strong>in</strong>g is vital for be<strong>in</strong>g able to<br />

replicate an <strong>option</strong> payoff. Derivatives <strong>in</strong> LIBOR Market Model, <strong>in</strong> this sense, belong<br />

to a “multi-asset” type of products. Thus, with LIBOR rates as underly<strong>in</strong>g assets, we<br />

can denote:<br />

∆ i = ∂U<br />

∂L i<br />

(6.1)<br />

where ∆ i is <strong>option</strong>’s sensitivity to <strong>the</strong> change <strong>in</strong> a i-th forward rate. Has <strong>the</strong> f<strong>in</strong>al solution<br />

been obta<strong>in</strong>ed, f<strong>in</strong>d<strong>in</strong>g deltas with respect to one of <strong>the</strong> forward rates is straightforward.<br />

With full-<strong>grid</strong> F<strong>in</strong>ite Differences, delta values can be extracted directly from<br />

<strong>the</strong> f<strong>in</strong>al mesh of po<strong>in</strong>ts <strong>in</strong> <strong>the</strong> form of first derivatives. In case of <strong>sparse</strong> <strong>grid</strong>s, it<br />

is not possible to apply <strong>the</strong> same procedure to <strong>the</strong> comb<strong>in</strong>ation result, due to <strong>the</strong> its<br />

non-uniformity. Instead, we can approximate delta surface by recomput<strong>in</strong>g <strong>the</strong> solution<br />

with a small shift <strong>in</strong> one of <strong>the</strong> underly<strong>in</strong>gs. The Figure (6.6) below presents <strong>the</strong><br />

surface of delta solution with respect to <strong>the</strong> forward rate L 1 .<br />

Interpolated Delta surface<br />

Delta Value<br />

1.2<br />

0.8 1<br />

0.6<br />

0.4<br />

0.2<br />

-0.2 0<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

Forward Rate L1<br />

0.04<br />

0.02<br />

0.02<br />

0.04<br />

0.06<br />

0.08<br />

0.1<br />

0.12<br />

0.14<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.2<br />

Forward Rate L2<br />

Figure 6.6: 2D Chooser. Sparse delta surface on level 7.<br />

50


1<br />

Full Delta Value<br />

Sparse Delta Value<br />

1<br />

Full Delta Value<br />

Sparse Delta Value<br />

0.8<br />

0.8<br />

Delta Value for L2 = 3.2%<br />

0.6<br />

0.4<br />

Delta Value for L2 = 4.6%<br />

0.6<br />

0.4<br />

0.2<br />

0.2<br />

0<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

(a) Delta values for L2=3.2%<br />

(b) Delta values for L2=4.6%<br />

1<br />

Full Delta Value<br />

Sparse Delta Value<br />

1<br />

Full Delta Value<br />

Sparse Delta Value<br />

0.8<br />

0.8<br />

Delta Value for L2 = 5.6%<br />

0.6<br />

0.4<br />

Delta Value for L2 = 6.7%<br />

0.6<br />

0.4<br />

0.2<br />

0.2<br />

0<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

(c) Delta values for L2=5.6%<br />

(d) Delta values for L2=6.7%<br />

1<br />

0.9<br />

Full Delta Value<br />

Sparse Delta Value<br />

1<br />

Full Delta Value<br />

Sparse Delta Value<br />

Delta Value for L2 = 7.5%<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

Delta Value for L2 = 11.2%<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

-0.1<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

(e) Delta values for L2=7.5%<br />

(f) Delta values for L2=11.2%<br />

Figure 6.7: 2D Chooser. Full vs. <strong>sparse</strong> delta profile on level 7<br />

51


Remark. The way we impose boundary conditions, that is keep<strong>in</strong>g <strong>the</strong>m fixed or<br />

sett<strong>in</strong>g <strong>the</strong>m to Black-Scholes value, does make a difference when we look at extracted<br />

Delta profile. Yet, <strong>in</strong> both cases, <strong>the</strong> approximation is still not accurate enough to<br />

provide smooth surfaces on <strong>the</strong> Delta plot. Tak<strong>in</strong>g advantage of <strong>the</strong> fact that people<br />

are mostly <strong>in</strong>terested <strong>in</strong> <strong>the</strong> ∆’s <strong>in</strong> at-<strong>the</strong>-money or close to at-<strong>the</strong>-money states, <strong>the</strong><br />

disturbed boundaries were excluded from <strong>the</strong> plots above.<br />

6.1.4 Greeks: Vega<br />

∆ is just one of <strong>the</strong> several sensitivities that help explore exposure to risk <strong>in</strong> <strong>the</strong> short<br />

position as a function of different <strong>model</strong> parameters. O<strong>the</strong>r ’Greek letters’ <strong>in</strong>clude Θ<br />

with respect to time changes <strong>and</strong> Γ as <strong>the</strong> second derivative with respect to changes <strong>in</strong><br />

underly<strong>in</strong>g. In this part of <strong>the</strong> text we turn to vega, V, a sensitivity of a product with<br />

respect to <strong>the</strong> volatility <strong>in</strong> one of <strong>the</strong> underly<strong>in</strong>g forward rates:<br />

V = ∂U<br />

∂σ i<br />

(6.2)<br />

Information about V is not encapsulated <strong>in</strong> <strong>the</strong> solution explicitly. Here, whe<strong>the</strong>r full<br />

<strong>and</strong> <strong>sparse</strong>, <strong>the</strong> solution has to be recomputed with a small shift <strong>in</strong> volatility value of<br />

one of <strong>the</strong> underly<strong>in</strong>g forward rates. Figure (6.8) below shows <strong>the</strong> <strong>sparse</strong> solutions<br />

for vega with respect to <strong>the</strong> σ 1 on Level 8. Plots <strong>in</strong> Figure (6.9) illustrate slices of a<br />

result<strong>in</strong>g vega profile with fixed L 2 rate.<br />

Sparse Vega po<strong>in</strong>ts. Level 8<br />

Vega Value<br />

0.03<br />

0.025<br />

0.015<br />

0.02<br />

0.005<br />

0.01<br />

-0.005 0<br />

-0.01<br />

0<br />

0.02 0.04 0.06 0.08<br />

0.1 0.12 0.14 0.16<br />

0<br />

Forward Rate L1<br />

0.04<br />

0.02<br />

0.16<br />

0.14<br />

0.12<br />

0.03<br />

0.025<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

-0.005<br />

-0.01<br />

0.1<br />

0.08<br />

Forward Rate L2<br />

0.06<br />

Figure 6.8: 2D Chooser. Sparse vega surface on level 8.<br />

52


0.03<br />

0.025<br />

Full vega Value<br />

Sparse vega Value<br />

0.03<br />

0.025<br />

Full vega Value<br />

Sparse vega Value<br />

0.02<br />

0.02<br />

vega Value for L2 = 3.2%<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

vega Value for L2 = 4.6%<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

-0.005<br />

-0.005<br />

-0.01<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

-0.01<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

(a) Vega values for L2=3.2%<br />

(b) Vega values for L2=4.6%<br />

0.03<br />

0.025<br />

Full vega Value<br />

Sparse vega Value<br />

0.03<br />

0.025<br />

Full vega Value<br />

Sparse vega Value<br />

0.02<br />

0.02<br />

vega Value for L2 = 5.6%<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

vega Value for L2 = 6.7%<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

-0.005<br />

-0.005<br />

-0.01<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

-0.01<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

(c) Vega values for L2=5.6%<br />

(d) Vega values for L2=6.7%<br />

0.03<br />

0.025<br />

Full vega Value<br />

Sparse vega Value<br />

0.03<br />

0.025<br />

Full vega Value<br />

Sparse vega Value<br />

0.02<br />

0.02<br />

vega Value for L2 = 7.5%<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

vega Value for L2 = 11.2%<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

-0.005<br />

-0.005<br />

-0.01<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

-0.01<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L1<br />

(e) Vega values for L2=7.5%<br />

(f) Vega values for L2=11.2%<br />

Figure 6.9: 2D Chooser. Full vs. <strong>sparse</strong> vega slices on level 8.<br />

53


Naturally, <strong>the</strong> characteristic feature of ’spikes’ of a <strong>sparse</strong> comb<strong>in</strong>ation solution is<br />

to be expected <strong>in</strong> <strong>the</strong> higher dimensional problems, where <strong>the</strong> contribution of coarse<br />

<strong>grid</strong>s is bigger. The follow<strong>in</strong>g section summarizes results for 3D Bermudan Swaption<br />

product.<br />

6.2 Three Dimensional Case. 3D Bermudan Swaption<br />

As discussed <strong>in</strong> Section (3.3), a Bermudan swaption is an <strong>in</strong>terest rate derivative with<br />

<strong>the</strong> ’early-exercise’ feature. It allows <strong>the</strong> holder to enter <strong>in</strong>to a swap contract on a set<br />

of prescribed dates <strong>and</strong> its general payoff is given by (3.27). S<strong>in</strong>ce we are comput<strong>in</strong>g<br />

<strong>the</strong> price of a swaption <strong>in</strong> <strong>the</strong> framework of a LIBOR Market Model, <strong>the</strong> term<strong>in</strong>al bond<br />

price D(t, T N+1 ) will play a role of a numeraire asset <strong>and</strong> Q will be its associated<br />

mart<strong>in</strong>gale measure. In case of 3 underly<strong>in</strong>g rates, one can engage <strong>in</strong> a swap at <strong>the</strong><br />

follow<strong>in</strong>g set of fix<strong>in</strong>g dates (T 0 ,T 1 ,T 2 ,T 3 ). Denote Vt<br />

BSw (T t , . . . , T N : T N+1 ) to be<br />

a Bermudan Swaption value at time t with expiry time T N+1 <strong>and</strong> set of fix<strong>in</strong>g dates<br />

(T t , . . . , T N ). Then, its value at T 0 follows:<br />

V BSw<br />

0 (T 0 , T 1 , T 2 , T 3 : T 4 ) = max ( H 0 (T 0 ), E 0 (T 0 ) ) (6.3)<br />

with a hold value H 0 (T 0 ) recursively def<strong>in</strong>ed from T 3 :<br />

[ (<br />

max<br />

H 2 (T 2 ) = D(2, 4) E Q H3 (T 3 ), E 3 (T 3 ) )<br />

]<br />

D(3, 4) ∣ F 2 = E Q[ ]<br />

(L 3 − K) + |F 2<br />

[ (<br />

max<br />

H 1 (T 1 ) = D(1, 4) E Q H2 (T 2 ), E 2 (T 2 ) )<br />

]<br />

D(2, 4) ∣ F 1<br />

[ (<br />

max<br />

H 0 (T 0 ) = D(0, 4) E Q H1 (T 1 ), E 1 (T 1 ) )<br />

]<br />

D(1, 4) ∣ F 0<br />

where <strong>the</strong> first relation holds s<strong>in</strong>ce:<br />

E 3 (T 3 ) = D(3, 4) (L 3 − K) +<br />

<strong>and</strong> H 3 (T 3 ) = 0<br />

(6.4)<br />

Figure (6.10) shows solution surfaces for 3D Bermudan Swaption with L 1 =5.62% fix<strong>in</strong>g:<br />

54


L1=5.62% Sparse Solution<br />

Swaption Price<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

0 200<br />

400<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

Forward Rate L3 0.06<br />

0.04<br />

0.02<br />

0<br />

0.02<br />

0.04<br />

0.06<br />

0.08<br />

0.1<br />

0.12<br />

0.14<br />

0.16<br />

Forward Rate L2<br />

(a) Full Solution Surface for L 1 =5.62%<br />

L1=5.62% Sparse Solution<br />

Swaption Price<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

1800<br />

1600<br />

1400<br />

1200<br />

1000<br />

800<br />

600<br />

400<br />

200<br />

0<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

Forward Rate L3 0.06<br />

0.04<br />

0.02<br />

0<br />

0.02<br />

0.04<br />

0.06<br />

0.08<br />

0.1<br />

0.12<br />

0.14<br />

0.16<br />

Forward Rate L2<br />

(b) Sparse Solution Surface for L 1 =5.62%<br />

Figure 6.10: 3D Bermudan Swaption. Full <strong>and</strong> Sparse solution surfaces.<br />

55


6.2.1 Convergence Study<br />

Table (6.6) shows steady convergence of full <strong>grid</strong> values. On <strong>the</strong> o<strong>the</strong>r h<strong>and</strong>, convergence<br />

of a <strong>sparse</strong> <strong>grid</strong> solution is not very smooth. However, it soon becomes clear<br />

that such convergence behaviour should be attributed to a large <strong>in</strong>terpolation error.<br />

Level<br />

Full Exec. Time Sparse Exec. Time Full po<strong>in</strong>t Sparse po<strong>in</strong>t<br />

solution (<strong>in</strong> sec.) solution (<strong>in</strong> sec.) count count<br />

4 31.639 11.94 4913 593<br />

5 29.802 121.11 39.691 1.51 35937 1505<br />

6 29.212 1107.16 38.494 5.27 274625 3713<br />

7 33.341 17.48 2146689 8961<br />

8 36.259 54.39 16974593 21249<br />

9 32.532 158.14 135005697 49665<br />

Table 6.6: 3D Bermudan Swaption. Spot Rates as Table (6.2)<br />

Figure (6.11) shows <strong>the</strong> discrepancy<br />

between <strong>the</strong> calculated price <strong>and</strong><br />

<strong>the</strong> benchmark provided <strong>in</strong> [Blackham,<br />

2004]. 500000 simulation<br />

runs of MC showed <strong>the</strong> value of<br />

29.064 with st<strong>and</strong>ard error 0.133.<br />

As suggested <strong>in</strong> [Andersen, 1999],<br />

numeraire-rebased swap value has<br />

been chosen as <strong>the</strong> explanatory variable<br />

for regression. Conditional expectations<br />

were fitted with 4th-degree<br />

Legendre polynomial basis function.<br />

Tables (6.7) <strong>and</strong> (6.8) below show<br />

good convergence of on-<strong>grid</strong> po<strong>in</strong>ts<br />

for a set of different at-<strong>the</strong>-money <strong>and</strong><br />

out-of-<strong>the</strong>-money forward rates.<br />

Option price<br />

38<br />

37<br />

36<br />

35<br />

34<br />

33<br />

32<br />

31<br />

30<br />

29<br />

Level of Discretization<br />

3 3.5 4 4.5 5 5.5 6<br />

MC Simulation Value+Std.Error<br />

Full-Grid FD value<br />

Full-Grid J.Blackham value<br />

28<br />

100000 150000 200000 250000 300000 350000 400000 450000 500000<br />

Number of MC simulations<br />

Figure 6.11: 3D Bermudan Swaption. Benchmark<strong>in</strong>g<br />

aga<strong>in</strong>st [Blackham, 2004] <strong>and</strong> Monte<br />

Carlo.<br />

Level<br />

L 2 =5.62% L 2 =1.875% L 2 =3.75%<br />

L 3 =3.75% L 3 =3.75% L 3 =2.81%<br />

Full Sparse Full Sparse Full Sparse<br />

4 51.335 8.189 4.684<br />

5 51.828 50.361 8.243 10.985 4.337<br />

6 51.897 51.694 8.232 9.472 4.246 9.546<br />

7 52.605 8.772 4.823<br />

8 51.499 8.272 4.348<br />

9 51.711 8.107 4.245<br />

MC<br />

Price Std.error Price Std.error Price Std.error<br />

52.2593 0.2053 8.3734 0.0601 4.294 0.0493<br />

Table 6.7: 3D Bermudan Swaption. Full <strong>and</strong> <strong>sparse</strong> on-<strong>grid</strong> solution values for<br />

L 1 =3.75%. 500000 MC simulation trials.<br />

56


Level<br />

L 2 =1.87% L 2 =1.875% L 2 =5.62%<br />

L 3 =7.5% L 3 =3.75% L 3 =5.62%<br />

Full Sparse Full Sparse Full Sparse<br />

4 196.542 10.945 176.422<br />

5 196.766 204.006 10.903 178.124<br />

6 196.628 207.901 10.875 13.080 178.474<br />

7 194.337 11.263 165.244<br />

8 196.054 11.067 178.232<br />

9 196.378 10.796 177.958<br />

MC<br />

Price Std.error Price Std.error Price Std.error<br />

196.514 0.337 11.062 0.075 177.208 0.383<br />

Table 6.8: 3D Bermudan Swaption. Full <strong>and</strong> <strong>sparse</strong> on-<strong>grid</strong> solution values for<br />

L 1 =5.62%. 500000 MC simulation trials.<br />

400<br />

350<br />

Full Solution L6<br />

Sparse Solution L6<br />

Sparse Solution L8<br />

L1=3.75% L2=3.75%<br />

200<br />

L1=3.75% L2=3.75%<br />

Full Solution L6<br />

Sparse Solution L6<br />

Sparse Solution L8<br />

300<br />

150<br />

Option value<br />

250<br />

200<br />

150<br />

Option value<br />

100<br />

100<br />

50<br />

50<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

(a) L 1 <strong>and</strong> L 2 out-of-<strong>the</strong>-money<br />

(b) Plot (a) Zoomed.<br />

400<br />

350<br />

Full Solution L6<br />

Sparse Solution L6<br />

Sparse Solution L8<br />

L1=5.62% L2=5.62%<br />

200<br />

L1=5.62% L2=5.62%<br />

Full Solution L6<br />

Sparse Solution L6<br />

Sparse Solution L8<br />

300<br />

150<br />

Option value<br />

250<br />

200<br />

150<br />

Option value<br />

100<br />

100<br />

50<br />

50<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

(c) L 1 <strong>and</strong> L 2 <strong>in</strong>-<strong>the</strong>-money<br />

(d) Plot (b) Zoomed.<br />

Figure 6.12: 3D Bermudan Swaption. Solution slices correspond<strong>in</strong>g to fix<strong>in</strong>gs <strong>in</strong> L 1<br />

<strong>and</strong> L 2 .<br />

57


6.2.2 Greeks. Delta <strong>and</strong> Vega.<br />

In this section we provide some results <strong>and</strong> comments on <strong>the</strong> results for <strong>the</strong> delta <strong>and</strong><br />

vega of <strong>the</strong> 3D Bermudan Swaption. Figure (6.13) shows <strong>the</strong> correspond<strong>in</strong>g surface<br />

plots <strong>in</strong> case of a full <strong>grid</strong> approximation.<br />

L1=5% Full Delta L6<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

Forward Rate L3 0.06<br />

0.04<br />

0.02<br />

0.02<br />

0.04<br />

0.06<br />

0.08<br />

0.1<br />

0.12<br />

0.14<br />

0.16<br />

Forward Rate L2<br />

(a) Delta<br />

L1=5.62% Full Vega Solution<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

-0.01 0<br />

-0.02<br />

-0.03<br />

0.14<br />

0.12 0.1 0.08 0.06<br />

Forward Rate L3<br />

0.04<br />

0.02<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

0.02<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

-0.01<br />

-0.02<br />

-0.03<br />

Forward Rate L2<br />

(b) Vega<br />

Figure 6.13: 3D Bermudan Swaption. Delta <strong>and</strong> Vega Surfaces for L 1 =5.62%. Full<br />

<strong>grid</strong>.<br />

58


Figures (6.14) <strong>and</strong> (6.15) show <strong>the</strong> slices of a <strong>sparse</strong> <strong>grid</strong> delta <strong>and</strong> vega profiles, respectively,<br />

plotted aga<strong>in</strong>st a full <strong>grid</strong> solution. Basis po<strong>in</strong>t (bp) is equal to 0.01%.<br />

1.4<br />

L1=5% L2=1.25%<br />

1.4<br />

L1=5% L2=1.25%<br />

1.2<br />

1.2<br />

1<br />

1<br />

Delta value<br />

0.8<br />

0.6<br />

Delta value<br />

0.8<br />

0.6<br />

0.4<br />

0.4<br />

0.2<br />

0.2<br />

Full Delta L6<br />

Sparse Delta L8. dL3 = 7bp<br />

Sparse Delta L8. dL3 =2bp<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

0<br />

Full Delta L6<br />

Sparse Delta L10. dL = 2bp<br />

-0.2<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

(a) Full L6 vs. Sparse L8. L 2 = 1.25%<br />

(b) Full L6 vs. Sparse L10. L 2 = 1.25%<br />

1.4<br />

L1=L2=5%<br />

1.4<br />

L1=5% L2=5%<br />

1.2<br />

1.2<br />

1<br />

1<br />

Delta value<br />

0.8<br />

0.6<br />

Delta value<br />

0.8<br />

0.6<br />

0.4<br />

0.4<br />

0.2<br />

Full Delta L6<br />

Sparse Delta L6. dL3 = 30bp<br />

Sparse Delta L6. dL3 = 7bp<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

0.2<br />

Full Delta L6<br />

Sparse Delta L8. dL8 = 7bp<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

(c) Full L6 vs. Sparse L6. L 2 = 5%<br />

(d) Full L6 vs. Sparse L8. L 2 = 5%<br />

1.4<br />

L1=5% L2=5.62%<br />

1.4<br />

L1=5% L2=5.62%<br />

1.2<br />

1.2<br />

1<br />

1<br />

Delta value<br />

0.8<br />

0.6<br />

Delta value<br />

0.8<br />

0.6<br />

0.4<br />

0.4<br />

0.2<br />

Full Delta L6<br />

Sparse Delta L8. dL3 = 7bp<br />

Sparse Delta L8. dL3 = 2bp<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

0.2<br />

Full Delta L6<br />

Sparse Delta L10. dL10 = 2bp<br />

0<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

(e) Full L6 vs. Sparse L8. L 2 = 5.62%<br />

(f) Full L6 vs. Sparse L10. L 2 = 5.62%<br />

Figure 6.14: 3D Bermudan Swaption. L 3 -Delta. Slices correspond<strong>in</strong>g to a fix<strong>in</strong>g <strong>in</strong> L 1 = 5%<br />

59


0.035<br />

0.03<br />

L1=3.75% L2=3.75% Shift <strong>in</strong> L3 = 10bp<br />

Full Vega L6<br />

Sparse Vega L7<br />

Sparse Vega L9<br />

0.035<br />

0.03<br />

L1=3.75% L2=3.75% Shift <strong>in</strong> L3 = 0.1bp<br />

Full Vega L6<br />

Sparse Vega L7<br />

Sparse Vega L9<br />

0.025<br />

0.025<br />

Vega value<br />

0.02<br />

0.015<br />

0.01<br />

Vega value<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0.005<br />

0<br />

0<br />

-0.005<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

-0.005<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

(a) L 2 = 3.75% Step <strong>in</strong> σ 3 = 10bp<br />

(b) L 2 = 3.75% Step <strong>in</strong> σ 3 = 0.1bp<br />

0.04<br />

0.03<br />

L1=5.62% L2=5.62% Shift <strong>in</strong> L3 = 10bp<br />

Full Vega L6<br />

Sparse Vega L7<br />

Sparse Vega L9<br />

0.04<br />

0.03<br />

L1=5.62% L2=5.62% Shift <strong>in</strong> L3 = 0.1bp<br />

Full Vega L6<br />

Sparse Vega L7<br />

Sparse Vega L9<br />

0.02<br />

0.02<br />

Vega value<br />

0.01<br />

Vega value<br />

0.01<br />

0<br />

0<br />

-0.01<br />

-0.01<br />

-0.02<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

-0.02<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

(c) L 2 = 5.62% Step <strong>in</strong> σ 3 = 10bp<br />

(d) L 2 = 5.62% Step <strong>in</strong> σ 3 = 0.1bp<br />

Figure 6.15: 3D Bermudan Swaption. L 3 -Vega. Slices correspond<strong>in</strong>g to a fix<strong>in</strong>g <strong>in</strong> L 1 = 5.62%<br />

Remarks. One can make several observations when look<strong>in</strong>g at <strong>the</strong> plots above. First,<br />

<strong>sparse</strong> solution for <strong>the</strong> greeks has proved to be stable with respect to <strong>the</strong> size of <strong>the</strong> shift<br />

<strong>in</strong> <strong>the</strong> underly<strong>in</strong>g <strong>and</strong> volatility step. Second, despite <strong>the</strong> clear presence of well-def<strong>in</strong>ed<br />

’wiggles’, <strong>sparse</strong> <strong>grid</strong> approximation is capable of outl<strong>in</strong><strong>in</strong>g <strong>the</strong> shape of <strong>the</strong> delta <strong>and</strong> vega<br />

slopes.<br />

60


6.3 Discont<strong>in</strong>uous Payoff. 2D Digital <strong>option</strong>.<br />

The third <strong>and</strong> f<strong>in</strong>al task case that we have considered is <strong>the</strong> case of a discont<strong>in</strong>uous<br />

payoff, characterictic of so-called digital <strong>option</strong>s. Figures (6.16) <strong>and</strong> (6.5) show <strong>the</strong><br />

digital payoff that describes our specific product:<br />

{<br />

V Digital 0.02 L1 > K OR L<br />

(T ) =<br />

2 > K<br />

0 o<strong>the</strong>rwise<br />

(6.5)<br />

With full <strong>grid</strong> of level 6 <strong>and</strong> K=5.5%, such payoff can be demonstrated visually:<br />

Discont<strong>in</strong>uous Payoff<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

0.16<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

Forward Rate L2 0.06<br />

0.04<br />

0.02<br />

0<br />

0<br />

0.02<br />

0.04<br />

0.06<br />

0.08<br />

0.1<br />

0.12<br />

0.14<br />

0.16<br />

Forward Rate L1<br />

0.02<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

Figure 6.16: 2D Digital Option. Term<strong>in</strong>al payoff discretized on a L6 full <strong>grid</strong>.<br />

The stripe of different color <strong>in</strong> <strong>the</strong> corner h<strong>in</strong>ts that <strong>in</strong> such sett<strong>in</strong>g <strong>the</strong> strike rate K<br />

is an off-<strong>grid</strong> po<strong>in</strong>t. Indeed, <strong>in</strong> case of level 6 discretization, K falls roughly between<br />

values 0.05391 <strong>and</strong> 0.05625.<br />

6.3.1 Valuation Results.<br />

Level of Discretization<br />

7 7.5 8 8.5 9 9.5 10<br />

21.2<br />

MC Simulation Value+Std.Error<br />

Full-Grid FD value<br />

21<br />

20.8<br />

Option price<br />

20.6<br />

20.4<br />

20.2<br />

20<br />

19.8<br />

19.6<br />

500000 1e+006 1.5e+006 2e+006 2.5e+006 3e+006<br />

Number of MC simulations<br />

Figure 6.17: 2D Digital Option. Convergence of Monte Carlo <strong>and</strong> full <strong>grid</strong> FD.<br />

61


Figure (6.17) above shows Monte Carlo <strong>and</strong> Full <strong>grid</strong> FD solutions for <strong>in</strong>creas<strong>in</strong>g number<br />

of simulation trials <strong>and</strong> different levels of discretization, respectively. The FD convergence<br />

requires discretisation level 10.<br />

Plots <strong>in</strong> Figure (6.18) illustrate <strong>the</strong> solution surfaces for full <strong>and</strong> <strong>sparse</strong> <strong>grid</strong> discretizations:<br />

2D Digital Solution.<br />

2D Digital Solution. Sparse Grid.<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

200<br />

180<br />

160<br />

140<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

Forward Rate L2 0.04<br />

0.02<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06<br />

0.04<br />

Forward Rate L1<br />

0.02<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

Forward Rate L2<br />

0.06<br />

0.04<br />

0.02<br />

0.14<br />

0.12<br />

0.1<br />

0.08<br />

0.06Forward Rate L1<br />

0.04<br />

0.02<br />

(a) Full Solution<br />

(b) Sparse Solution<br />

Figure 6.18: 2D Digital Option. Full <strong>and</strong> <strong>sparse</strong> <strong>grid</strong> solution surfaces. Level 7.<br />

The bumps on <strong>the</strong> <strong>sparse</strong> solution surface <strong>and</strong> oscillations on <strong>the</strong> boundaries reveal <strong>the</strong><br />

presence of ’spikes’. Still, Tables (6.9) <strong>and</strong> (6.10) show relatively satisfactory convergence<br />

<strong>in</strong> case of on-<strong>grid</strong> po<strong>in</strong>ts:<br />

Level<br />

L 1 =1.87% L 1 =1.87% L 1 =3.75% L 1 =3.75%<br />

L 2 =3.75% L 2 =5.62% L 2 =3.75% L 2 =4.68%<br />

Full Sparse Full Sparse Full Sparse Full Sparse<br />

5 17.092 23.628 89.300 130.156 22.065 29.124 51.215 51.094<br />

6 15.436 22.188 84.271 97.389 19.716 23.805 47.414 54.958<br />

7 16.490 17.813 86.700 85.402 21.138 20.356 49.646 50.124<br />

8 16.015 15.403 85.473 82.088 20.481 21.968 48.617 49.894<br />

9 15.776 16.443 84.863 87.704 20.153 20.470 48.101 48.572<br />

10 16.031 84.883 19.991 20.163 48.341<br />

11 15.769 84.665 20.042 48.044<br />

12 15.653 84.462 20.160 48.049<br />

MC<br />

Price Std.error Price Std.error Price Std.error Price Std.error<br />

15.661 0.0793 84.758 0.141 20.013 0.0886 47.735 0.124<br />

Table 6.9: 2D Digital. Full <strong>and</strong> <strong>sparse</strong> <strong>grid</strong> convergence of on-<strong>grid</strong> po<strong>in</strong>ts. 500000 MC simulation trials.<br />

62


Level<br />

L 1 =3.75% L 1 =4.68% L 1 =5.62% L 1 =5.62% L 1 =5.62%<br />

L 2 =5.62% L 2 =3.75% L 2 =1.87% L 2 =3.75% L 2 =5.62%<br />

Full Sparse Full Sparse Full Sparse Full Sparse Full Sparse<br />

5 89.456 93.811 48.683 48.707 94.433 144.257 95.101 105.22 117.35 158.660<br />

6 84.664 85.297 44.155 53.857 87.891 108.154 88.791 92.075 111.236 137.341<br />

7 87.192 82.942 46.758 47.944 91.012 91.672 91.977 87.894 114.009 123.701<br />

8 85.965 88.521 45.535 47.775 89.430 86.233 90.408 93.333 112.555 115.810<br />

9 85.352 85.157 44.925 45.498 88.644 89.739 89.625 91.783 111.835 113.733<br />

10 85.229 45.319 88.988 89.533 113.023<br />

11 85.055 44.913 88.480 89.292 112.396<br />

12 85.340 44.921 88.171 89.592 111.965<br />

MC<br />

Price Std.error Price Std.error Price Std.error Price Std.error Price Std.error<br />

84.601 0.141 44.492 0.122 87.453 0.1413 88.572 0.141 111.124 0.137<br />

Table 6.10: 2D Digital Cont<strong>in</strong>ued. Full <strong>and</strong> <strong>sparse</strong> <strong>grid</strong> convergence of on-<strong>grid</strong> po<strong>in</strong>ts. 500000 MC simulation trials.<br />

Remarks. From results presented above, it can be concluded that both 2D Digital full<br />

<strong>grid</strong> <strong>and</strong>, more so, <strong>sparse</strong> <strong>grid</strong> have less dist<strong>in</strong>ct on-<strong>grid</strong> convergence than <strong>the</strong> cases of 2D<br />

Chooser <strong>and</strong> 3D Bermudan. As mentioned before, our strike price happens to lie between<br />

<strong>grid</strong> po<strong>in</strong>ts, which is, <strong>in</strong> fact, a typical problem of a f<strong>in</strong>ite difference space discretization<br />

relative to pric<strong>in</strong>g of <strong>the</strong> digital payoff. Consequently, <strong>the</strong> discont<strong>in</strong>uity of <strong>the</strong> <strong>option</strong> value<br />

is approximated with a large quantification error <strong>and</strong> <strong>the</strong> accuracy of a FD solution is deteriorated.<br />

This issue, specific of a canonical spacial discretization, shows itself to a much<br />

larger extent <strong>in</strong> case of <strong>sparse</strong> <strong>grid</strong> delta calculation.<br />

63


6.3.2 Greeks. Delta <strong>and</strong> Vega.<br />

Delta Figures (6.19) <strong>and</strong> (6.20) show <strong>the</strong> delta profile characteristic of a 2D digital<br />

<strong>option</strong> <strong>and</strong> how full approximation compares with <strong>sparse</strong> <strong>grid</strong> representation.<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

-0.1<br />

0.02 0.04 0.06 0.08 0.1<br />

Forward Rate L1 0.12 0.14<br />

Forward Rate L2<br />

0.02 0.040.060.080.10.120.14<br />

Figure 6.19: 2D Digital Option. Full <strong>grid</strong> Delta on L7.<br />

L1=4.68%<br />

L1=4.68%<br />

0.8<br />

0.7<br />

Full Delta L7<br />

Sparse Delta L7<br />

Sparse Delta L9<br />

0.8<br />

0.7<br />

Full Delta L7<br />

Sparse Delta L7<br />

Sparse Delta L9<br />

0.6<br />

0.6<br />

0.5<br />

0.5<br />

Delta value<br />

0.4<br />

0.3<br />

Delta value<br />

0.4<br />

0.3<br />

0.2<br />

0.2<br />

0.1<br />

0.1<br />

0<br />

0<br />

-0.1<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

-0.1<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L2<br />

Forward Rate L2<br />

(a) L 1 = 4.68%. Shift <strong>in</strong> L 1 = 10bp<br />

(b) L 1 = 4.68%. Shift <strong>in</strong> L 1 = 1bp<br />

L1=5.62%<br />

L1=5.62%<br />

0.6<br />

0.5<br />

Full Delta L7<br />

Sparse Delta L7<br />

Sparse Delta L9<br />

0.6<br />

0.5<br />

Full Vega L7<br />

Sparse Vega L7<br />

Sparse Vega L9<br />

0.4<br />

0.4<br />

Delta value<br />

0.3<br />

0.2<br />

Vega value<br />

0.3<br />

0.2<br />

0.1<br />

0.1<br />

0<br />

0<br />

-0.1<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

-0.1<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L3<br />

Forward Rate L2<br />

(c) L 1 = 5.62%. Shift <strong>in</strong> L 1 = 10bp<br />

(d) L 1 = 5.62%. Shift <strong>in</strong> L 1 = 1bp<br />

Figure 6.20: 2D Digital Option. Sparse Delta L7 <strong>and</strong> L9 slices.<br />

64


Figures (6.21) <strong>and</strong> (6.22) show <strong>the</strong> vega profile characteristic of a 2D digital <strong>option</strong><br />

<strong>and</strong> how full approximation compares with <strong>sparse</strong> <strong>grid</strong> representation.<br />

0.015<br />

0.01<br />

0.005<br />

-0.005 0<br />

-0.01<br />

-0.015<br />

-0.02<br />

-0.025<br />

0 0.02 0.04 0.06 0.08 0.1 0.12<br />

Forward Rate L1 0.14 0.16<br />

0.14<br />

0.015<br />

0.01<br />

0.005<br />

0<br />

-0.005<br />

-0.01<br />

-0.015<br />

-0.02<br />

-0.025<br />

0.02<br />

0.04<br />

0.06<br />

0.08<br />

0.10.12<br />

Forward Rate L2<br />

Figure 6.21: 2D Digital Option. Full <strong>grid</strong> vega on L7.<br />

L1=4.68%<br />

L1=4.68%<br />

0.014<br />

0.012<br />

Full Vega L7<br />

Sparse Vega L7<br />

Sparse Vega L9<br />

0.014<br />

0.012<br />

Full Vega L7<br />

Sparse Vega L7<br />

Sparse Vega L9<br />

0.01<br />

0.01<br />

0.008<br />

0.008<br />

Vega value<br />

0.006<br />

0.004<br />

Vega value<br />

0.006<br />

0.004<br />

0.002<br />

0.002<br />

0<br />

0<br />

-0.002<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

-0.002<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L2<br />

Forward Rate L2<br />

(a) L 1 = 4.68%. Shift <strong>in</strong> σ 1 = 10bp<br />

(b) L 1 = 4.68%. Shift <strong>in</strong> σ 1 = 0.1bp<br />

L1=5.62%<br />

L1=5.62%<br />

0.005<br />

0<br />

Full Vega L7<br />

Sparse Vega L7<br />

Sparse Vega L9<br />

0.005<br />

0<br />

Full Vega L7<br />

Sparse Vega L7<br />

Sparse Vega L9<br />

-0.005<br />

-0.005<br />

Vega value<br />

-0.01<br />

Vega value<br />

-0.01<br />

-0.015<br />

-0.015<br />

-0.02<br />

-0.02<br />

-0.025<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

-0.025<br />

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16<br />

Forward Rate L2<br />

Forward Rate L2<br />

(c) L 1 = 5.62%. Shift <strong>in</strong> σ 1 = 10bp<br />

(d) L 1 = 5.62%. Shift <strong>in</strong> σ 1 = 0.1bp<br />

Figure 6.22: 2D Digital Option. Sparse Vega L7 <strong>and</strong> L9 slices.<br />

Remarks. As anticipated, <strong>sparse</strong> <strong>grid</strong> delta approximation appeared to be very dem<strong>and</strong><strong>in</strong>g<br />

to <strong>the</strong> size of <strong>the</strong> shift. It becomes obvious from Figure (6.20), plot(b), that<br />

many of <strong>the</strong> less accurate solution po<strong>in</strong>ts become irresponsive to <strong>the</strong> shift as large as<br />

1%. When <strong>the</strong> step size is reduced to one-tenth of a percent, <strong>the</strong> differential of <strong>the</strong> value<br />

with respect to step <strong>in</strong> <strong>the</strong> underly<strong>in</strong>g becomes zero. Such behavior of <strong>the</strong> delta solution<br />

is, aga<strong>in</strong>, <strong>the</strong> magnified effect of quantification error mentioned earlier. Discont<strong>in</strong>uity<br />

65


of <strong>the</strong> payoff is also accountable for <strong>the</strong> severe spikes <strong>in</strong> <strong>sparse</strong> vega solution near <strong>the</strong><br />

at-<strong>the</strong>-money position.<br />

6.4 Discussion <strong>and</strong> Recommendations.<br />

In f<strong>in</strong>ancial community, <strong>the</strong> quality of a computational <strong>method</strong> is, to a large extent,<br />

determ<strong>in</strong>ed by its ability to show accurate approximations of risk sensitivities, such as<br />

∆ <strong>and</strong> V. We can conclude from ∆ <strong>and</strong> V slice plots of each test case that <strong>the</strong> impact of<br />

<strong>the</strong> spikes effect is be<strong>in</strong>g magnified as we start operat<strong>in</strong>g on smaller ranges of values.<br />

What didn’t show <strong>in</strong> quality of price estimates, shows dist<strong>in</strong>ctly <strong>in</strong> <strong>the</strong> values for <strong>the</strong><br />

Greeks. This fact makes immediate use of a <strong>sparse</strong> <strong>grid</strong> <strong>method</strong> <strong>in</strong> a discussed sett<strong>in</strong>g<br />

quite problematic. Steps need to be taken to improve <strong>the</strong> accuracy of approximation.<br />

There are several possible recipes one could follow:<br />

• Iterative Solver. When benchmark<strong>in</strong>g aga<strong>in</strong>st <strong>the</strong> results from[Blackham, 2004],<br />

<strong>the</strong> latter shows smoo<strong>the</strong>r convergence <strong>in</strong> both discretizations, <strong>and</strong>, moreover,<br />

surpris<strong>in</strong>gly close results on earlier levels. Most likely, such behaviour is <strong>in</strong>flicted<br />

by <strong>the</strong> choice of an iterative solver. The author of [Blackham, 2004]<br />

reports to have used second-order f<strong>in</strong>ite differences with a multi<strong>grid</strong> iterative<br />

solver. The multi<strong>grid</strong> technique is designed to operate on a hierarchy of <strong>in</strong>creas<strong>in</strong>gly<br />

“coarser” meshes where, after desired accuracy has been reached, every<br />

<strong>in</strong>dividual mesh represents a part of <strong>the</strong> common solution.<br />

However, despite smoo<strong>the</strong>r convergence, <strong>the</strong> author of [Blackham, 2004] reports<br />

<strong>the</strong> presence of spikes <strong>in</strong> <strong>the</strong> solution. Thus, similar <strong>in</strong>accuracies are to be expected<br />

<strong>in</strong> <strong>the</strong> results for <strong>the</strong> Greeks.<br />

• Order of FD. A possible way of with deal<strong>in</strong>g discretization errors could be ref<strong>in</strong><strong>in</strong>g<br />

spatial discretization to a fourth-order f<strong>in</strong>ite differences, as explored <strong>in</strong><br />

[Leentvaar <strong>and</strong> Oosterlee, 2006]. Aga<strong>in</strong>, such discretization poses difficulties<br />

for ”narrow” <strong>grid</strong>s <strong>and</strong> solutions adjacent to boundary conditions, that appear to<br />

rely on ”ghost” outlier po<strong>in</strong>ts. These issues can be addressed with application of<br />

extrapolation or rewrit<strong>in</strong>g of PDE with one-sided differences on <strong>the</strong> borders of<br />

solution doma<strong>in</strong>.<br />

• Comb<strong>in</strong>ation technique <strong>and</strong> <strong>sparse</strong> <strong>grid</strong> type. It would be worthwhile <strong>in</strong>vestigat<strong>in</strong>g<br />

<strong>the</strong> effect of <strong>the</strong> choice of an alternative <strong>sparse</strong> <strong>grid</strong> comb<strong>in</strong>ation technique.<br />

Some of <strong>the</strong> possible alternatives to <strong>the</strong> “classical” <strong>sparse</strong> <strong>grid</strong> <strong>and</strong> comb<strong>in</strong>ation<br />

technique are discussed <strong>in</strong> [Klimke, 2003] <strong>and</strong> [Leentvaar <strong>and</strong> Oosterlee,<br />

2006], respectively.<br />

• Quantification Error. In practice, <strong>the</strong> quantification error of discont<strong>in</strong>uous payoff<br />

can be significantly reduced by peform<strong>in</strong>g coord<strong>in</strong>ate transformation. As described<br />

<strong>in</strong> [Travella <strong>and</strong> R<strong>and</strong>al, 2000], by rewrit<strong>in</strong>g <strong>the</strong> PDE <strong>in</strong> different terms,<br />

one can transform <strong>the</strong> <strong>grid</strong> <strong>in</strong> such a way that a larger amount of po<strong>in</strong>ts is concentrated<br />

<strong>in</strong> <strong>the</strong> discont<strong>in</strong>uity region. S<strong>in</strong>ce this procedure does not degrade <strong>the</strong><br />

accuracy of discretization, it could be an employed by <strong>the</strong> <strong>sparse</strong> <strong>grid</strong> <strong>method</strong> to<br />

battle <strong>the</strong> quantification error <strong>in</strong> case of digital <strong>option</strong>.<br />

66


Chapter 7<br />

Conclusion <strong>and</strong> Future Work<br />

In this <strong>the</strong>sis, we have <strong>in</strong>vestigated <strong>the</strong> application of <strong>sparse</strong> <strong>grid</strong> comb<strong>in</strong>ation technique<br />

to <strong>the</strong> 2D <strong>and</strong> 3D <strong>in</strong>terest rate products <strong>in</strong> <strong>the</strong> LIBOR Market Model framework.<br />

The selected test cases have been solved with second-order f<strong>in</strong>ite differences<br />

<strong>and</strong> BiCGSTAB iterative solver.<br />

In two dimensions, a chooser <strong>option</strong> <strong>and</strong> a digital <strong>option</strong> show good agreement<br />

of MC, full <strong>and</strong> <strong>sparse</strong> solutions <strong>in</strong> case of on-<strong>grid</strong> po<strong>in</strong>ts for higher levels of discretization.<br />

For <strong>the</strong> chooser <strong>option</strong>, <strong>the</strong> price has converged to <strong>the</strong> value provided <strong>in</strong><br />

[Blackham, 2004]. In addition, we have <strong>in</strong>cluded results for different sets of strike <strong>and</strong><br />

spot rates <strong>in</strong> case of on-<strong>grid</strong> solution po<strong>in</strong>ts, where good convergence <strong>in</strong> both full <strong>and</strong><br />

<strong>sparse</strong> discretizations was recorded.<br />

In three dimensions, <strong>the</strong> <strong>valuation</strong> of a bermudan swaption demonstrated convergence<br />

behaviour on a full <strong>grid</strong>. However, <strong>the</strong> obta<strong>in</strong>ed FD <strong>and</strong> MC values were different<br />

from <strong>the</strong> benchmark provided <strong>in</strong> [Blackham, 2004]. While convergence of a benchmark<br />

problem solution on a <strong>sparse</strong> <strong>grid</strong> was not smooth due to a large <strong>in</strong>terpolation<br />

error, results for on-<strong>grid</strong> values were by far more conv<strong>in</strong>c<strong>in</strong>g.<br />

The accuracy of approximations of <strong>the</strong> ’Greeks’ has been shown to suffer seriously<br />

from <strong>the</strong> ’spikes’, <strong>in</strong>troduced <strong>in</strong>to <strong>the</strong> f<strong>in</strong>al comb<strong>in</strong>ation by approximations on<br />

<strong>the</strong> coarser <strong>grid</strong>s <strong>in</strong> both 2D <strong>and</strong> 3D. Moreover, <strong>the</strong> delta of <strong>the</strong> digital <strong>option</strong> becomes<br />

a victim of a large quantification error.<br />

Our suggestions for <strong>the</strong> future work are <strong>the</strong> follow<strong>in</strong>g:<br />

• ref<strong>in</strong><strong>in</strong>g discretization scheme of F<strong>in</strong>ite Differences to <strong>the</strong> fourth order could help<br />

to obta<strong>in</strong> more accurate solutions on <strong>the</strong> coarse <strong>grid</strong>s of <strong>sparse</strong> discretization.<br />

• it would be <strong>in</strong>terest<strong>in</strong>g to see how results of BiCGSTAB compare with those<br />

of a different iterative solver. In our view, a study of different comb<strong>in</strong>ations<br />

of discretization order <strong>and</strong> iterative solver could make a positive change on <strong>the</strong><br />

quality of <strong>the</strong> ’Greeks’.<br />

• a study of <strong>the</strong> effect of <strong>grid</strong> stretch<strong>in</strong>g <strong>and</strong> coord<strong>in</strong>ate transformation on <strong>the</strong> quality<br />

of digital ’Greeks’ with <strong>sparse</strong> <strong>grid</strong> approach.<br />

• comb<strong>in</strong>ation technique is <strong>in</strong>herently well-fit for parallel comput<strong>in</strong>g. When solv<strong>in</strong>g<br />

problems <strong>in</strong> higher than 3D, parallel computation of solutions on <strong>in</strong>dividual<br />

<strong>grid</strong>s will be required.<br />

67


• reduction <strong>in</strong> <strong>the</strong> dimension of <strong>the</strong> PDE could be attempted by fast-drift approximation,<br />

as <strong>in</strong> [Pietersz et al., 2005].<br />

At <strong>the</strong> end we conclude that <strong>sparse</strong> <strong>grid</strong> comb<strong>in</strong>ation technique is a promis<strong>in</strong>g<br />

<strong>method</strong> to help revive F<strong>in</strong>ite Difference approach <strong>in</strong> <strong>the</strong> <strong>in</strong>terest-rate <strong>option</strong> pric<strong>in</strong>g<br />

field <strong>in</strong> higher dimensions.<br />

68


Appendix A<br />

Implementation Details<br />

In this appendix, we present a brief list of software tools we employed throughout <strong>the</strong><br />

project:<br />

1. C++, as <strong>the</strong> language of choice, allowed us to take advantage of <strong>the</strong> power of<br />

templates, operator overload<strong>in</strong>g <strong>and</strong> generic STL conta<strong>in</strong>ers <strong>and</strong> algorithms.<br />

2. Sub<strong>grid</strong>s have been implemented on top of Blitz++ generic array library, described<br />

<strong>in</strong> [Veldhuizen].<br />

3. LMM PDE solver made use of <strong>the</strong> BiCGSTAB rout<strong>in</strong>e <strong>and</strong> supplimentary data<br />

structures of <strong>the</strong> free-source iterative <strong>method</strong>s library, described <strong>in</strong> [Dongarra<br />

et al.].<br />

4. As <strong>in</strong> case of full <strong>grid</strong>, off-<strong>grid</strong> po<strong>in</strong>ts of <strong>sparse</strong> <strong>grid</strong> solution need to be found<br />

with <strong>in</strong>terpolation. The current implementation of <strong>sparse</strong> <strong>grid</strong> solver doesn’t<br />

provide a rout<strong>in</strong>e to <strong>in</strong>terpolat<strong>in</strong>g non-uniform meshes, so this procedure has<br />

been outsourced to Matlab. Interpolation of off-<strong>grid</strong> values has been peformed<br />

as a post-process<strong>in</strong>g step with <strong>grid</strong>data <strong>and</strong> <strong>grid</strong>data3 rout<strong>in</strong>es.<br />

5. GnuPlot software came very h<strong>and</strong>y <strong>in</strong> draw<strong>in</strong>g pretty graphs <strong>and</strong> color-mapped<br />

surfaces.<br />

6. The text of <strong>the</strong> <strong>the</strong>sis has been typeset <strong>in</strong> L A TEX.<br />

69


Bibliography<br />

L. Andersen. A simple approach to <strong>the</strong> pric<strong>in</strong>g of bermudan swaptions <strong>in</strong> <strong>the</strong> multifactor<br />

LIBOR Market Model. Journal of Computational F<strong>in</strong>ance, 3:5–32, 1999.<br />

F. Black <strong>and</strong> M. Scholes. The pric<strong>in</strong>g of <strong>option</strong>s <strong>and</strong> corporate liabilities. Journal of<br />

Political Economy, 81(3):637–659, 1973.<br />

J. Blackham. Sparse <strong>grid</strong> solutions to <strong>the</strong> LIBOR Market Model. Master’s <strong>the</strong>sis,<br />

Magdalen College, University of Oxford, 2004.<br />

Z. Bodie, A. Kane, <strong>and</strong> A.J. Marcus. Investments. 6th edition edition, 2004.<br />

A. Brace, D. Gatarek, <strong>and</strong> M. Musiela. The <strong>market</strong> <strong>model</strong> of <strong>in</strong>terest rate dynamics.<br />

Ma<strong>the</strong>matical F<strong>in</strong>ance, 7(2):127–155, 1997.<br />

H.J. Bungartz <strong>and</strong> T. Dornseifer. Sparse <strong>grid</strong>s: Recent developments for elliptic partial<br />

differential equations. In W. Hackbusch <strong>and</strong> G. Wittum, editors, Multi<strong>grid</strong> Methods<br />

V, Lecture Notes <strong>in</strong> Computational Science <strong>and</strong> Eng<strong>in</strong>eer<strong>in</strong>g, pages 45–70, Berl<strong>in</strong><br />

Heidelberg, 1998. Spr<strong>in</strong>ger.<br />

J. Cox, S. Ross, <strong>and</strong> M. Rub<strong>in</strong>ste<strong>in</strong>. Option pric<strong>in</strong>g, a simplified approach. J F<strong>in</strong>ancial<br />

Econ, 7:229–263, 1979.<br />

J. Crank <strong>and</strong> P. Nicolson. A practical <strong>method</strong> for numerical e<strong>valuation</strong> of solutions of<br />

partial differential equations of <strong>the</strong> heat-condution type. Ma<strong>the</strong>matical Proceed<strong>in</strong>gs<br />

of <strong>the</strong> Cambridge Philosophical Society, 43:50–67, 1949.<br />

J. Dongarra, A. Lumsda<strong>in</strong>e, R. Pozo, <strong>and</strong> K. Rem<strong>in</strong>gton. IML++ v. 1.2 Iterative Methods<br />

Library. Available at: http://math.nist.gov/iml++/.<br />

A. E<strong>the</strong>rage. F<strong>in</strong>ancial Calculus. Cambridge University Press, 2002.<br />

J. Garcke. Sparse <strong>grid</strong> tutorial. http://wwwmaths.anu.edu.au/˜garcke/<br />

paper/<strong>sparse</strong>GridTutorial.pdf, 2005.<br />

M. Griebel, M. Schneider, <strong>and</strong> C. Zenger. A comb<strong>in</strong>ation technique for <strong>the</strong> solution<br />

of <strong>sparse</strong> <strong>grid</strong> problems. In Proceed<strong>in</strong>gs of <strong>the</strong> IMACS International Symposium on<br />

Iterative Methods <strong>in</strong> L<strong>in</strong>ear Algebra, pages 263–281, Amsterdam, 1992. Elsevier.<br />

URL citeseer.ist.psu.edu/griebel92comb<strong>in</strong>ation.html.<br />

J. Hull. Options, Futures <strong>and</strong> o<strong>the</strong>r Derivatives. Prentice Hall, 1999.<br />

70


A. Klimke. Piecewise multil<strong>in</strong>ear <strong>sparse</strong> <strong>grid</strong> <strong>in</strong>terpolation <strong>in</strong> MATLAB, IANS report<br />

2003/019. Technical report, University of Stuttgart, 2003. URL citeseer.ist.<br />

psu.edu/klimke03piecewise.html.<br />

C.C.W. Leentvaar <strong>and</strong> C.W. Oosterlee. Pric<strong>in</strong>g multi-asset <strong>option</strong>s with <strong>sparse</strong> <strong>grid</strong>s<br />

<strong>and</strong> fourth-order f<strong>in</strong>ite differences. Delft University of Technology, 2006.<br />

F. Longstaff <strong>and</strong> E. Schwartz. Valu<strong>in</strong>g american <strong>option</strong>s by simulation: A simple leastsquares<br />

approach. The review of F<strong>in</strong>ancial Studies, 14(1):133–148, 2001.<br />

A. Pelsser. Efficient Methods for Valu<strong>in</strong>g Interest Rate Derivatives. Spr<strong>in</strong>ger F<strong>in</strong>ance,<br />

2000.<br />

V. Peterbarg. A practitioner’s guide to pric<strong>in</strong>g <strong>and</strong> hedg<strong>in</strong>g callable <strong>libor</strong> exotics <strong>in</strong><br />

forward <strong>libor</strong> <strong>model</strong>s. Available at: http://ssrn.com/abstract=427084,<br />

2003.<br />

R. Pietersz. PDE pric<strong>in</strong>g for BGM. Available at: http://ssrn.com/abstract=<br />

302266, 2002.<br />

R. Pietersz, A. Pelsser, <strong>and</strong> M. van Regenmortel. Fast drift approximated pric<strong>in</strong>g <strong>in</strong> <strong>the</strong><br />

BGM <strong>model</strong>. Available at: http://ssrn.com/abstract=427084, 2005.<br />

R. Rebonato. Interest-Rate Option Models: Underst<strong>and</strong><strong>in</strong>g, Analyz<strong>in</strong>g <strong>and</strong> Us<strong>in</strong>g Models<br />

for Exotic Interest Rate <strong>option</strong>s. (Wiley Series <strong>in</strong> F<strong>in</strong>ancial Eng<strong>in</strong>eer<strong>in</strong>g). John<br />

Wiley & Son Ltd., 2nd edition edition, 1998.<br />

C. Reis<strong>in</strong>ger. Numerische Methoden fur hochdimensionale parobolische Gleichungen<br />

an Beispiel von Optionspreisaufgaben. PhD <strong>the</strong>sis, Ruprecht-Karls-Universitat, Heidelberg,<br />

Germany, 2004.<br />

D. Travella <strong>and</strong> C. R<strong>and</strong>al. Pric<strong>in</strong>g F<strong>in</strong>ancial Instruments. The F<strong>in</strong>ite Difference<br />

Method. Wiley, 2000.<br />

H.A. van der Vorst. BiCGSTAB: A fast <strong>and</strong> smoothly converg<strong>in</strong>g variant of Bi-CG<br />

for <strong>the</strong> solution of nonsymmetric l<strong>in</strong>ear systems. SIAM J. Sci. Stat. Comput., 13:<br />

631–644, 1992.<br />

T. Veldhuizen. Blitz++ User’s Guide. A C++ class library for scientific comput<strong>in</strong>g.<br />

Available at: http://www.oonumerics.org/blitz/.<br />

P. Wessel<strong>in</strong>g. An Introduction to Multi<strong>grid</strong> Methods. R.T. Edwards, Inc., 2004.<br />

C. Zenger. Sparse <strong>grid</strong>s. In W. Hackbusch, editor, Parallel Algorithms for Partial<br />

Differential Equations, Notes on Numerical Fluid Mechanics; 31, pages 241–251.<br />

Vieweg, Braunschweig, 1991.<br />

71

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!