I A I)! ”OPOIOS 7 “' -_.__, ~! 7m“ L. . ‘4, ,. ..n.;.....| . . '1’... Z...‘, .7‘ a n r . 2....1 :J I . 1:! . . .ulv. A. —. l..l t 3.4 r 4V5M4fi€§ . 5M5 - “55$, :3... 3 a ..‘_ 1.? L . s A . . .n aJ .3 Ex) . . . ...V..‘ . 11. . x. , . . av? .a; a f ‘ . . . .mfi» “gaming... ‘ $31.3» .. 'l I999 IIHNHNNHI"!”Will!!!(”MINIMUMllHUlHHHl 93 01810 3766 This is to certify that the dissertation entitled Continuous time arbitrage approached as a problem in constrained hedging presented by James A. Demopolos has been accepted towards fulfillment of the requirements for Ph.D. degree in Statistics x‘le/M Major professor Date August23, 1999 W WM“ Major Professor MS U is an Affirmative Action/Equal Opportunity Institution 0-12771 LIBRARY Michigan State University PLACE IN RETURN Box to remove this checkout from your record. To AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE CONTINUOUS TIME ARBITRAGE APPROACHED AS A PROBLEM IN CONSTRAINED HEDGING By James Andre Demopolos A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Statistics and Probability 1999 ABSTRACT CONTINUOUS TIME ARBITRAGE APPROACHED AS A PROBLEM IN CONSTRAINED HEDGING By James Andre Demopolos I characterize absence of arbitrage with tame portfolios in a model where a finite vector of stock prices is symbolized by a continuous semi-martingale with respect to the completed filtration generated by a vector-valued standard Brownian Motion. Levental and Skorohod (1995) solved this problem using probabilistic methods in the sub—case of invertible volatility matrix. They constructed an arbitrage trading strategy based upon domination at the end of the time horizon of the value of one stochastic process by that of another. This construction through domination suggests a link between the arbitrage problem and the mathematical theory of financial hedging of contingent claims. This dissertation does not assume invertible volatility. In the case of singular volatility, one faces the constraint that the dominating process constructed by Levental and Skorohod cannot always be effectively converted into a process symbolizing the accumulated capital gains of a trading strategy. Therefore, to apply the theory of hedging, one must consider hedging under constraints. This dissertation contains two primary results. First, I generalize Levental and Skorohod's characterization of arbitrage opportunities in terms of a domination relationship between stochastic processes. Second, I apply work by Cvitanic and Karatzas (1993) pertaining to hedging with constrained portfolios to this generalization to provide a new characterization of absence of arbitrage in the case of singular volatility. The proofs are probabilistic. Some examples are provided. To my wife, Emily, and my daughter, Sofia iii - ACKNOWLEDGEMENTS I would like to express my gratitude to the co-chairmen of my dissertation committee. Professor Levental's knowledge and pleasant demeanor were of great value in the completion of this research, and I am grateful for the opportunity to study probability from as accomplished a mathematician as Professor Skorokhod. I also owe sincere thanks to several other members of the faculty in the Department of Statistics and Probability. I thank Chairman Salehi for helping me in many ways during my stay at Michigan State, Professor Melfi for providing consistently excellent instruction in theoretical probability, Professor Stapleton for giving courses characterized by his diligence and insight, and Professor Page for her guidance during my experience as a consultant. Professors Erickson, Gilliland, LePage and Fabian also contributed to my successful completion of this program, and I duly thank them. I would also like to thank Cathy Sparks for her assistance with administrative matters, and Alexander White for useful discussions early in the research. iv TABLE OF CONTENTS Introduction Setting and Main Results 1.1 The Model. . . . ..................... 1.2 The Link Between Arbitrage and Constrained Hedging. ....... 1.3 The History of the Arbitrage Problem. .............. Immediate Arbitrage 2.1 Preliminaries. ........................ 2.2 Proof of the Immediate Arbitrage Theorem. ............ A Construction from the Theory of Constrained Hedging 3.1 The Construction and Closely Related Properties. .......... 3.2 An Essential Lemma in the Link to the Arbitrage Problem. ...... 3.3 Two Useful Characterizations of the Process V(-) ........... Characterizations of Arbitrage and Corollaries 4.1 Characterization of Arbitrage in Terms of Domination of Z(T; 1) ..... 4.2 Characterization of Arbitrage in Terms of the Processes V(t; o). Examples Bibliography 10 16 22 22 24 30 3O 36 4O 43 43 53 66 76 INTRODUCTION A basic problem in the construction of asset price models in mathematical finance is the determination of the conditions which are necessary and sufficient for a specified model to exhibit the absence of arbitrage, i. e., the absence of risk-free profit opportunities. In this work, a characterization of absence of arbitrage is provided in the context of a specified model for a finite vector of stock prices. The work has been motivated by results in Levental and Skorohod (1995) and Cvitanic and Karatzas (1993). The former paper characterizes absence of arbitrage in a restricted version of the model considered here. As will be explained in detail in Chapter 1, Levental and Skorohod's Corollary 3 [page 920] suggests a link between the problem of characterizing absence of arbitrage and the theory of hedging contingent claims. In the more general setting of this dissertation, fewer stochastic processes can be taken to meaningfully symbolize accumulated discounted capital gains than in Levental and Skorohod's work. This limitation motivates consideration of the problem of hedging under constraints in the course of attempting to provide a hedging based approach to the arbitrage problem. The problem of hedging contingent claims with constrained portfolios is precisely the topic of Cvitanic and Karatzas (1993). The approach of this dissertation is to generalize Corollary 3 of Levental and Skorohod to provide a characterization of arbitrage "reminiscent of hedging" in the setting of this work, and then to apply the framework of Cvitanic and Karatzas to this generalization in order to provide a new characterization of absence of arbitrage. The paper is organized in the following way: In Chapter 1, I specify the model, and motivate the link between the theory of hedging with constrained portfolios and the arbitrage problem. The dissertation's major results are stated. Chapter 1 also contains a history of research into the arbitrage problem. Chapter 2 contains the statement and proof of necessary and sufficient conditions for the absence of a special kind of arbitrage, namely, immediate arbitrage. Loosely speaking, in an immediate arbitrage, an investor does not ever let his capital gains become negative in the process of obtaining almost sure positive capital gains at the end of the time horizon. Although I did have to make modifications to the argument, the core of the proof, particularly on the necessity side, appears in Levental and Skorohod (1995) [Lemma 2, page 914.] In Chapter 3, I adapt the work of Cvitanic and Karatzas to this arbitrage problem. The details differ from their work in that in this paper, I need to impose constraints which depend upon (t,oo), whereas their constraints do not vary with (Loo). Chapter 4 accomplishes the stated objectives of this work. Theorem 3 is the promised generalization of Levental and Skorohod's Corollary 3. Theorem 4 results from application of the constructions derived from Cvitanic and Karatzas' work in Chapter 3 to the characterization given in Theorem 3. Theorem 5 extends the conclusions of Theorem 4 to address the issue of the equivalence of absence of arbitrage and the existence of an absolutely continuous local martingale measure for the stock price processes. Chapter 5 contains examples. As will be explained herein, Examples 1 and 4 show that the characterization of arbitrage in the setting of this work is meaningfully different from Levental and Skorohod's (1995) characterization. Chapter 1 Setting and Main Results 1.1 The Model. Consider a financial market in which one bond, with price process B, and d 2 1 stocks, with price processes SI, ..., Sd, are traded in the time interval 0 S t S 1. Unless otherwise specified, all processes herein will be taken to be defined for O s t S 1. Correspondingly, in definitions of stopping times interpret the infirnum of an empty time set as 1. The source of uncertainty in the market is a d-dimensional standard Brownian Motion W = (W1, W2, ..., Wd)* defined on a complete probability space (Q, F, P)l. The term “adapted” will refer throughout to the filtration {F :2 0 s t S l}, the P augmentation of the natural filtration of W, namely (1.1) ‘ Ftéo{W(s):‘OSsSt}vU where U = {A e F: P(A) = O}. The price processes of the financial instruments evolve according to the equations (1.2) I dB(t)¥B(t)r(t)dt, B(O)=1. ' * will denote matrix transpose. (1.3) dS.(t) = Smi 26.,.(t)de(t) + brow], lede Si(O)=Si E(0,00), 1:1,...,d. Here r(-) is an adapted R-valued process symbolizing the instantaneous force of interest, volatility 0(a) is an adapted d x d matrix-valued process not necessarily invertible for any (I, (D), and drift b(-) is an adapted Rd-valued process. In order that (1.2) and (1.3) have well-defined solutions, we require that (1.4) KW)! +Z|b,(t)| +Zo§j(z)}dt < 00 as. i A continuous-time trader chooses a portfolio, namely, the amount of money to invest in each of the d stocks at each time t. Formally define a portfolio by Definition 1. A portfolio is an adapted Rd-valued process it which satisfies the integrability constraint 1 1‘) u (1.5) prsmsyl‘+in'(s)a(s)[}ds < oo a.s., 0 where // // denotes the Euclidean norm in Rd_ and with Id =(1, 1, I)“ 6 Rd, the process a is defined by (1.6) a(t) = b(t) — r(t)1d. Since (1.4) implies that the paths B(-) satisfy inf{B(t) : O S t _<_ 1} > O a.s., constraint (1.5) implies that the semi-martingale X,t in Definition 2 below is a well-defined process. Definition 2. The process X7t given by (1.7) X, (z) = {3“ (s)(1t '(s)c (s))dW(s) + {13“ (5)1: ‘(s)a(s)ds o o is the discounted capital gain process associated with the portfolio 1t. To motivate Definition 2, begin from the purpose of investing in stocks, namely, the attempt to obtain capital gains in excess of those available from the less risky bond. In this light, n‘(t)o(t)dW(t) + 1t.(t)b(t)dt — r(t)«n‘(t)1d dt = _ no) _ 11,0) - :[Sj(t)dsj(z) 80) (13(1)), which is verbally, {instantaneous gains from portfolio investment in nj(t) / SJ-(t) stock shares, j = 1, ..., d} — {opportunity cost of foregone instantaneous gains possible from the bond}. Multiplying by B’l(t) discounts these excess (or deficient) gains from stock investment to their present value at time 0. Integration across time sums the discounted instantaneous gains. Common wisdom is that a reasonable model for the processes Sj and B should not allow for risk-free profits. This-is the no arbitrage principle. Definition 3. An arbitrage is a portfolio 1: such that the associated discounted capital gain process X7t satisfies 1) There exists a C > 0 such that P{Xn(t) 2 —Cfor all 0 S t S 1} = 1. ii) P{X.(1) 2 0} = 1. iii) P{X.(1) > 0} > 0. Any portfolio 7: for which the associated Xn satisfies i) in Definition 3 is called a tame portfolio. C-tameness means that i) is satisfied with respect to a particular C. Tameness is a restriction that prevents “doubling schemes” and can be interpreted as putting a limit on borrowing. The mathematics underlying "doubling" in continuous time was set forth by Dudley (1977), who showed that in our model with d = 1, an arbitrary F 1 measurable random variable A (including, in particular, A satisfying A > O as.) can be represented as 1 l A = Ig(t)dW (t) for an adapted process g satisfying Ig2(t)dt < 00 as. In his 0 0 construction, it is possible that for each C > O, f P{min0$t$1 Ig(S)dW(S) < “C} > O. 0 The relationship between the absence of as. positive capital gains and the requirement of tarneness is treated rigorously in Dybvig and Huang (1988) [see Theorem 2, page 390.] The purpose of this dissertation is the study of conditions equivalent to the absence of arbitrage. To that end, we need define numerous objects. Let it denote Lebesgue measure on [0, 1]. As Shreve has shown, unless there is a projection-based arbitrage, then we must have (1.8) X®P{a(t, 0)) e Ran[o(t, co)]} = 1. [See Karatzas and Shreve (1998), Theorem 1.4.2, page 12.] Condition (1.8) will be assumed throughout this workz. It holds for each (t,(o) that Ran[o(t,w)] = Ran[o(t,(u)o'.(t.(o)] and that 00.0)) is injective on Ran[o.(t,co)]. Therefore, we may uniquely (up to a HEP null set) and adaptedly define a relative risk process 9 such that 9(t,(u) e Ran[o‘(t,a))] for all (1.0)) and o(-)9(o) = a(-) X®P a.s.3 Using 9, define a stopping time or: t+h (1.9) a = inft>02 ]||9(s)||2ds :00 for all he(0, 1-1]. 1 a is a legitimate stopping time because of right-continuity of the Brownian filtration, and is the key object in a characterization of the absence of a special kind of arbitrage. Definition 4. An immediate arbitrage is a portfolio 11; for which there exists a stopping time 0 S t S 1 satisfiing P{‘C < 1} > 0 such that P{Xn(t) = Ofor all t S t anan(t) > 0for all t > t}: 1. Theorem 1 (Immediate Arbitrage Theorem.) There is no immediate arbitrage if andonly ifP{0t =1} = 1. The primary contribution of this work pertains to characterization of arbitrage when immediate arbitrage does not exist. Therefore, as is consistent with Theorem 1, for the 2 See Chapter 2, Proposition 1, for the details of the necessity of ( l .8) for the absence of arbitrage. 3 To define 6 such that it is an adapted process, define 9(t) = 0:] (t) a(t), O S t S l, where 6+ is an adapted process such that for each (t,a)), o+(t,(o) is an invertible dxd matrix and for each x e Ran[o(t,a) )], o+"(t,m) x e Ran[o'(t,u))] and o(t,m) o.'l(t,w) x = x. The existence of such an adapted o. is proven in Lemma 1 of Chapter 2. remainder of this chapter all results will be given under the assumption that a = l as. In the absence of immediate arbitrage, the fundamental objects in results about existence of arbitrage are exponential local martingales. For each adapted Rd-valued process v satisfying orv = 1 a.s., where stopping time av is defined as in (1.9) with process v substituted for process 9, define for each stopping time 0 S t S 1 another stopping time C(r) by t (1.10) C(r) = inf+t>02 [{z CV (15). Z‘ (I; Q" (1)): liminf tVTC ( )Zv(‘t;t). We have that lim Zv (t ;t) exists as. The limit exists on T; “m {lfihtP{o(t,o)) is singular} > 0. Of fundamental importance in Levental and Skorohod's proof that an arbitrage exists if there exists a stopping time t such that E(Z(t; 1)) < 1 is the existence in that case of an exponential local martingale Z‘”(-) which satisfies 10 P{Z‘p(l) > Z(t; 1)} = 1. In fact, their Corollary 3 [page 920] states that in the case of ‘dt < 00 a.s., that the existence of I invertible o, with the added assumption that HP (t) 0 arbitrage is equivalent to the existence of an adapted Rd-valued process (p satisfying 1 fll@(’)ll2df < 00 as. such that 0 (1.15) P{Z“’(1)2Z(1)} =1 and P{Z“’(1)>Z(l)} >0. Equation (1.15) suggests a link between the arbitrage problem and the theory of hedging. To understand this link, begin with consideration of a "seller's objective in hedging." [See Karatzas (1996), Section 0.4 for more detail than is given here] Define a contingent claim to be a non-negative Fl-measurable random variable". One can view a contingent claim as a financial obligation at time 1 to which a seller commits himself in exchange for money at time 0. Let A be a contingent claim. For each x 2 0 such that there exists an adapted Rd-valued process it such that 1 2 (1.16) [{Haana)” + n*(r)b(z)]}dt < 00 a.s., O l 1 (1.17) x + ]n*(r)o(t)dW(z) + ]n*(t)b(z)dr 2 A a.s., 0 0 and there exists a constant C > 0 such that we have the tameness constraint ‘ Typically in work focusing on hedging, additional constraints which imply absence of arbitrage are assumed to apply with respect to process 6. A requirement related to these additional constraints is then included in the definition of a contingent claim [see Karatzas (1996), page 10]. Since mention of hedging is intended to be motivational here, and since I have not assumed absence of arbitrage, I have chosen to omit these additional details. 11 I I (1.18) P minoggl ]n*(s)o(s)dW(s) + 1r*(s)b(s)ds < —C = 0, 0 0 we have the interpretation that a seller can "hedge" his obligation to pay A at time 1 starting with the purchase price x at time 0.5 Examination of (1 .13), the integral representation of process Z‘p(-), suggests the link between (1 . 15) and this "seller's objective in hedging." Consider an "auxiliary market" in which asset price processes are characterized by the original invertible volatility 0', but the drift b is replaced by the zero vector process. Then define an adapted Rd-valued process rt by no) = [o‘(t)1"(-Z“’(t>_ X, as. (ii) If Y' is a G measurable extended random variable satisfying (1), then ‘1' 2 Y a.s Denote Y = ess sup{X, : is I }. 14 For each I, the adapted process Vo(r; .) admits a cadlag modification V(r; o), which by right-continuity is unique up to indistinguishability. Use the abbreviation V(o) to denote the process V(O; .), The processes V(r; .) and Z(t; .) are the central objects in this work's characterization of arbitrage: Theorem 4. Assume absence of immediate arbitrage. Then there is no arbitrage if and only if processes Z(r; .) and V(r; .) are indistinguishable for all constant times 0SrSl Theorem 5 below is an equivalent formulation of Theorem 4. In a sense, Theorem 4 is a characterization of absence of arbitrage in terms of a stochastic supremum, while Theorem 5 recasts this result in terms of an attained maximum. Theorem 5 is an important tool in addressing the problem of equivalence of absence of arbitrage and the existence of a probability measure Q << P such that the asset prices S;(-), i = 1, ..., d are local martingales with respect to (O, F, {Ft}o 5 1s 1, Q). Worthy of note is that the proof of Theorem 5 contrasts with approaches in the literature to the problem of existence of an absolutely continuous local martingale measure in that it does not rely upon functional analysis. Theorem 5. Assume absence of immediate arbitrage. There is no arbitrage if and only if for each constant time 0 S r S 1 there exists a 11 e D such that E (26 “Yr; 1)) = 1. 15 In the final chapter, I give examples illustrating the properties of the processes V(r; .) and their relationship to the processes Z(r; o). Example 1 demonstrates that the equivalent condition for absence of arbitrage in Theorem 4 is not equivalent to E(Z(r;1)) = 1 for all constant times r. If there is no immediate arbitrage, then the latter condition implies absence of arbitrage, but Example 1 serves as a counter-example to the reverse implication by exhibiting both E(Z(1)) < 1 and no arbitrage. Example 2 shows that we cannot simplify Theorem 4 by reducing the conditions equivalent to absence of arbitrage to the behavior of the processes V(r; .) at time 0: in Example 2, we have V(O) = 1 a.s., but V(-) and Z(-) are not indistinguishable. Although the processes V(r; .) have cadlag paths, it is not true in general that they have continuous paths. In Example 3, P{V(1/2) at 11mm aV(t)} > 0. Example 4 is due to Delbaen and Schachermayer (1998b). It is similar to Example 1, which I constructed before discovering their paper. In Example 4, there exists a v e D such that V(-) and E"[Z(l) I F.] are indistinguishable. 1.3 The History of the Arbitrage Problem. Since the late 1970's researchers have actively investigated the question of which properties of possible asset price models correspond to the absence of arbitrage opportunities. Most of the resulting articles have focused in one way or another on the notion of an equivalent martingale measure, namely, a probability measure Q equivalent to the measure P such that the discounted asset price processes (denoted Si(o)/B(o), i = 1, ..., d in this text's notation) are martingales on the original filtered probability space with measure Q replacing P. That the existence of such a martingale measure is a 16 sufficient condition for absence of arbitrage in a wide variety of circumstances was established early in research on this topic. For discrete time asset price models defined on a finite probability space with finitely many time values, Harrison and Kreps (1979) showed that the existence of such an equivalent martingale measure is necessary and sufficient for the absence of arbitrage. An early result for continuous time trading models appeared in Harrison and Pliska (1981); therein, the authors show that with a discounted price model that is a cadlag strictly positive semi-martingale, the existence of an equivalent martingale measure implies absence of arbitrage. The question of whether absence of arbitrage implies the existence of a martingale measure is complex, particularly in the case of continuous time process models. Almost all proofs of the existence of a martingale measure have employed the Hahn-Banach Theorem or one of its corollaries. In the discrete-time case, Harrison and Kreps (1979), and similarly, Harrison and Pliska (1981) employed the separating hyper-plane theorem to generate a linear fimctional symbolizing a pricing system with which one can construct an equivalent martingale measure. Both of these papers worked with finite probability spaces for the discrete-time problem. Taqqu and Willinger (1987) also established the equivalence of absence of arbitrage and existence of an equivalent martingale measure for a discrete time, finite probability space framework; their proof differed from preceding works in that it used a geometric reformulation of the no arbitrage assumption. Dalang, Morton and Willinger (1990) established the equivalence for discrete time trading in general (non—finite) probability spaces. The equivalence for general probability spaces was subsequently proved using‘somewhat simpler arguments than those in 17 Dalang, et. al. by Kabanov and Kramkov (1994) and Rogers (1995). Note that the theorems referred to above all pertained to a finite number of trading times. In discrete-time asset models, the issue for the infinite time horizon case is more complicated. Back and Pliska (1991) provide an example allowing trading in the infinite horizon which does not permit arbitrage, but for which there is no equivalent martingale measure. The notion of "no fi'ee lunch," sometimes called "no approximate arbitrage," is a stronger assertion than no arbitrage and becomes relevant here. There are several formulations of "flee lunch" in the literature. Early definitions of the existence of "free lunch" require a sequence of random variables, namely, terminal wealth levels for discounted capital gain processes, to converge topologically to a nonnegative random variable that is not as. 0. The topologies used to define the convergence vary by paper. [See, for example, Kreps (1981).] Because a sequence of trading strategies which require a trader to risk increasingly large losses, none of which produce probability one positive capital gains, seems undesirable as an approximation to arbitrage, tameness requirements were added to the definition of "free lunch." Schachermayer (1994) defines the property of "free lunch with bounded risk" as the existence of a sequence of arbitrage approxirnants which are each C-tame for a single C > 0. He proves that in the infinite time horizon discrete trading problem, "no free lunch with bounded risk" (N F LBR) is equivalent to the existence of an equivalent martingale measure. Furthermore, in the infinite horizon discrete case, the need to prevent "doubling scheme" based arbitrage becomes apparent. Harrison and Kreps (1979) provide an example of probability one positive capital gains where the minimum value of the wealth process across time is not bounded below as. 18 In continuous time trading, "doubling" based as. positive capital gains are possible in a model admitting an equivalent martingale measure even with a finite time horizon. Harrison and Pliska (1981) correct for this phenomenon by requiring tameness. Dybvig and Huang (1988) provide a rigorous analysis of the impossibility of as. positive discounted capital gains in a market admitting an equivalent martingale measure if one adds the requirement of portfolio tameness. Regarding the problem of the existence of a martingale measure, in the context of a continuous bounded semi-martingale model for the discounted asset prices on time set [0, l], Delbaen (1992) proved the equivalence of NFLBR and the existence of an equivalent martingale measure. The results of F ritelli and Lakner (1995) include that under only the assumption that the discounted asset price processes are adapted and right continuous, existence of an equivalent martingale measure is necessary and sufficient for absence of "free lunch with stopping times." In their work, the stochastic processes are defined on an arbitrary index subset of [0, co), and "free lunch with stopping times" is defined as a sequence of arbitrage approximants for which the portfolios processes lie in the linear span of {1t(t) = g{t < t S [3}; t S B are stopping times, g e L°°(P), g is FF measurable}. Duffie and Huang (1986) and Stricker (1990) study the relationship between "no free lunch" and the existence of an equivalent martingale measure under the assumption that the discounted asset price processes are in LP, 1 S p < oo. Duffie and Huang (1986) also prove some interesting results about the relationship between "no free lunch" and the relative sizes of filtrations generated by different agents' information. Delbaen and Schachermayer (1994b) establish that if the discounted asset price process 19 {S(t); 0 S t < 00} is a bounded real valued semi-martingale, then there is an equivalent martingale measure if and only if S satisfies "no free lunch with vanishing risk" (NF LVR). NFLVR is defined to hold if for any sequence of positive constants 6n satisfying limn 8n = 0, each sequence of Sn—tame portfolios an (where portfolios are defined as predictable processes it for which the stochastic integral S {In(t)dS(t); 0 S s < 00} is well-defined and converges as. to a limit as 5—) 00) must 0 co satisfy P-limn Inn(t)dS(t) = 0. As a corollary, they obtain that if S is a locally bounded 0 semi-martingale, then NFLVR is equivalent to the existence of an equivalent probability measure under which S is a local martingale. This corollary complements Delbaen and Schachermayer (1994a) in which examples are provided showing that for unbounded continuous discounted price processes, NFLBR is not equivalent to the equivalent martingale measure property. The proofs of Delbaen and Schachermayer rely heavily upon functional analysis. For the model in this thesis with invertible volatility, Levental and Skorohod (1995) prove that an equivalent martingale measure exists if and only if there is "no approximate arbitrage," a condition which means roughly the same thing as NFLVR. Their proof is more probabilistic than that of Delbaen and Schachermayer. Levental and Skorohod (1995) and Delbaen and Schachermayer (1995) both investigate the relationship between the existence of an absolutely continuous measure Q << P under which the discounted asset prices are martingales, and absence of arbitrage (as opposed to absence of "free lunch") Levental and Skorohod (1995) use the martingale representation theorem to 20 show that in their model, assuming absence of immediate arbitrage, no arbitrage is equivalent to the existence of an absolutely continuous probability measure Qr << P for each 0 S r S 1 under which (with the expression given for the one-dimensional case) {Sm/Ba), r .<. t < €90). {a} 1 , Q} is a local martingale. Delbaen and Schachermayer (1995) show, referring back to the (1994b) result proven using the Hahn-Banach theorem, that if {S(t); O S t < 00} is a locally bounded semi-martingale , then absence of arbitrage implies the existence of an absolutely continuous probability measure Q << P under which the discounted asset price process is a local martingale. Delbaen and Schachermayer (19983) consider the case of unbounded asset price processes. Assuming that {S(t); O S t < 00} is a semi-martingale, they prove that NFLVR is equivalent to the existence of a measure Q equivalent to P under which the discounted price process is a martingale transform, i. e., S Icp(t)dM (t); 0 S s < 00 , where M is an Rd valued martingale, and (p is a predictable 0 M-integrable R+-valued process. 21 Chapter 2 Immediate Arbitrage 2.1 Preliminaries. The following result, due to Shreve, demonstrates why we assume condition (1.8), X®P{a(t, (1)) e Ran[0'(t, 03)]} = 1. Proposition 1. If X®P{a(t, 00) e Ran[o(t, (1))]} < 1, then an immediate arbitrage exists. Proof. For each (t,(o), Rd = Ker[o"(t,0))] O Ran[o'(t,c0)], where EB denotes orthogonal sum. Define an adapted R‘Lvalued process a1 by defining a1(t,00) to be the projection of a(t,a)) on Ker[o‘(t,(o)]. Then define another adapted Rd-valued process it by (2.1) Mt): 2 21.0 ) 1+lla < )1 ‘it is a portfolio. That it meets the integrability constraint in the definition of a portfolio follows from that for each (t,co) (2.2) 0‘1: 2 0 6 Rd and 211.3 = ”a,”2 so that 0 S n‘a <1. Now define a stopping time ‘I.’ by 22 (2.3) 1: = inf{t > 0: 71({sza1(s)¢ 0} n [t, t + a] )> 0 for all e > 0}. That I is a stopping time follows from right-continuity of {F(}. That (1 .8) does not hold implies that P{t < 1} > 0. Furthermore, (2.2) yields that X, satisfies P{Xn(t) = 0 for all t S r and Xn(t) > 0 for all t > r}=1. So it is an immediate arbitrage. I Let us now attend to some technical details used in the proof of the immediate arbitrage theorem. Lemma 1. Let 0' be an adapted dxd matrix valued process. Then there exists a (non- unique) adapted process 6+ such that for each (I, (1)), 6+(t,0)) is an invertible dxd matrix and for each x e Ran[o(t,to)] we have both 0.5! (t, (o)x e Ran[o'(t, (1)) ] and oft, (9)0510, (1))x = x. Proof. Let k(t,(u) = dim(Ran[o'(t,o))]) = dim(Ran[o(t,co)]), and let {ej, fj, g], ; j = 1, ..., d}be a set of adapted Rd-valued processes such that for each (t,c0), {e1(t,0)), ..., ek(t,w)}, {f1(t,oo), ..., fd-k(t,(0)}, and {g1(t,c0), ..., gd_k(t,0))} are bases for Ran[o'(t,0))], (Ran[o*’(t,w)])i, and (Ran[o(t,0))])i, respectively. Take o.(t,c0) to be the matrix representation of the full-rank linear map on Rd defined by o+(t,(o)ej(t,a)) = o(t,c0)ej(t,0)), j = 1, ..., k(t,co), 0+(t,c0)f}(t,a)) = gj(t,o)), j = l, ..., d — k(t,oo). That Ran[o(t,c0)] = Ran[o(t,co)o'(t,u))] justifies that 0+(t,0)) is invertible. The remaining assertions made about 6+ are evident from its construction. I 23 In this chapter and the next, the Girsanov Theorem will be a useful tool. [See Karatzas and Shreve (1991), Theorem 3.5.1, page 191.] For each adapted Rd-valued process v I satisfying fl|v(t)l]2 dt < 00 as, define another adapted Rd-valued-process 0 W’ = 0. Start the construction of an immediate arbitrage by selecting a sequence of constants rk l 0 such that if stopping times ak, k = 0,1,..., are defined by (1k: (0: + n.) A 1, then (2.5) ;PHLfia,<1sa,_,}{]p(z)||2A-g;]dz gr} (1 {a<1}] < 00. Get such constants satisfying (2.5) as follows: Let r0 = 1. After selecting . {r}, i = 0, k—I} and defining {01), i = 0, k—I} as stipulated above, the divergence 1 7 1 limrw fin +r 0: {{a, 0 as t i 0} =1. Then, observing that w (t) = 0 on {t < a} and that 2({111 (5):: 0}O(a, a+a])> 0 for alls> 0' holds as. on {01 < 1}, we can conclude that I fw'(s)dW(s) (2.12) lim 0 =0 as. on {oc_ lim’k I 0 1/3 = 00. [ iliv (91%] [1020» was] Let ‘I.’ be the stopping time I l , 1/3 (2.14) 1: = inf{t>or: Iii/’(s)dW(s) + Iul'(s)9(s)ds = [my (s)“2ds] }. 0 0 26 Expression (2.13) implies that as. on {0: < 1}, both ‘I.’ > a and ifa < t S r, then (2.15) :fiu‘(s)dW(s) + :fw'(s)9(s)ds 2 “Malfdsj > 0. Now observe that for an adapted scalar process g, 111 = g0, so that for all (t,(l)), w (t,c0) e Ran[o.(t,(o)]. Therefore Lemma 1 implies that there exists an adapted process (c‘). taking values in the invertible dxd matrices such that c‘(.)[(c‘).(.)]“w (.) = w (.). Define an adapted vector process 7:0 by mm = [(ol)+(t)]"\y (t). Then define the immediate arbitrage 7t by tt(t) = B(t){t S r}7to(t). Regarding integrability constraints on a portfolio, the paths B(-) are bounded in t a.s., and (2.16) (IIB"(r)o'(r)n(r)||2+|B“'0:m'0)al}dt = ji’S'ClQlilmllz+lil'(f)9(t)|)1i s J(B2(t)+B(t)){z(t)dt _<_ i(k’“+k“) a.s. Since B“(.)n‘(.)e(.) = {. s t}ut‘(o), and x. satisfies the SDE (2.17) dx.(t) = B“1 (t)0'(t)dW(t) + B"(t)e(t)e(t)dt, x.(0) = 0, it follows from (2.15) that as. (2.18) X,(t) = 0, ift s a and x.(t) > 0, ift> a. So 71: is an immediate arbitrage with a serving as the stopping time required in the definition. 27 (Sufficiency) Suppose or = 1 as. Let 7t(-) be a portfolio such that there exists a stopping time ‘t for which P{Xn(t)= O for all 0 StSt and Xn(t)> 0 for allt> 1} =1. Define another stopping time B by B = inf{t>0: ]{t 0 as. Further observe that the Novikov Condition implies E(Z(t; B)) = 1 [See Karatzas and Shreve (1991 ), Corollary 3.5.13, page 199.] Define an adapted vector process 9 by B(t) = B(t){t < t s [3}. Then the process Z9.(-) is Z(r; . A B), so that E( Z6 (1) ) = 1. Because P{Xn(t) = 0 for t S r} = 1 and X7: satisfies the SDE (2.17), the Girsanov Theorem implies that on ((2, F1, {F t}, Pei ), the process X.(- A B) is a stochastic integral with respect to W6, a standard Brownian Motion in Rd [see (2.4)]. In particular, X.(- A B) is a (P9!) local martingale. Then, since P6 is equivalent to P, we have (2.19) P“ {X.(t) = 0 for all 0 s t St and x,(t) > 0 for allt> t} = 1. Therefore, x..(.A 13) is a (P6 ) supennartingale.‘ Then 55(X,(8)) s afar, (1)) = 0, so that P6 (x.(8) = 0) = 1 by (2.19). By probability equivalence, P(x.(8) = 0) = 1. 1 It follows from the Fatou Lemma for conditional expectation that any local martingale which is bounded below is also a superrnartingale. 28 Sor < B as. on {I <1} andP{X,,(t)> O for allt> t} = 1 imply thatP{'t = 1} =1. Since it and I were chosen arbitrarily, no immediate arbitrage exists. 29 Chapter 3 A Construction from the Theory of Constrained Hedging 3.1 The Construction and Closely Related Properties. Throughout Chapters 3 and 4, assume absence of immediate arbitrage, i. e., assume or = 1 as. The results in this chapter specialize work of Cvitanic and Karatzas (1992) and (1993) to the problem of this dissertation. This chapter contains little that Cvitanic and Karatzas did not prove. The usefulness of Corollary 1 is‘specific to this arbitrage problem; so it did not appear in their work. All proofs, barring that of the existence of the cadlag modification of V0 in Theorem 2 draw at least their key probabalistic content from Cvitanic and Karatzas. They give a different proof, which also appears in El Karoui and Quenez (1995), justifying the existence of the cadlag modification, but to the best of my knowledge the one given here has not appeared elsewhere. Their papers constructed the V process for the problem of pricing contingent claims in incomplete markets, and did not recognize its usefuleness for the arbitrage problem. Comparison with their papers will show that my approach is to'view Z(‘t’; l) as a contingent claim to be hedged in a market with d risky assets characterized by our original volatility process 0' and drift process identically zero. 30 The results and proofs of this chapter will use the notation V(-) and Z(-) otherwise reserved for ‘t = 0. All of the proofs go through without alteration for V(r;-) given any stopping time 0 S r S 1. I have chosen the simpler notation since nothing herein depends upon I. Recall that we define D to be the class of adapted Rd-valued processes v for which (1.19) A®P {v(t,00) e Ker[ o(t,co) ]} =1, 1 (1.20) P{ J]|v(s)||2ds < co} = 1, 0 (1.21) E(ZV(1)) = 1. For each v e D, Ev denotes expectation and conditional expectation with respect to the probability measure Pv with Radon-Nikodym derivative dPV/dP = Z"(1). Let V0 be an adapted Rl-valued process such that for each 0 S t S 1, Vo(t) is a version of ess supveD (EV[Z(1)| Ft]). In working with essential suprema, the following lemma will be of use. Lemma 2. F or each stopping time 0 S ‘I.’ S I, there exists a sequence {vm' n 2 1} in D for which we have the a. s. monotone convergence (3.1) ess sup..o(E"[Z(1)| F.]) =11:an EV"[Z(1)| F.]. Proof. If a family of random variables is directed upward, then the essential supremtun of the family is the as. increasing limit of a sequence in the family [see 31 Neveu (1975), Proposition VI-l-l , page 121.]1 So fix I and show that {EV[Z(1) I F.]: v e D} is directed upward. Let v] and v; be in D and define A = {EV1[Z(1)|F,] 2 .Ev2 [Z(1)|F,]} e F, . Then define an adapted process 11 by “(0 = Vl(t)[A fl {t> Tl] + V2(t)lAC I] {t > T}]- u e D: it obviously satisfies (1.19) and (1.20). For (1.21), because A 6 F., F.]) Then, for j = 1, 2, using that (1 .20) implies P{min 0 5., , z"i(t) > 0} = 1, and optionally (3.2) E(Z“(1)) = E(AE[ZV'(t;1)I~;D + E(A‘E[Z"2(t;1) stopping martingale 2‘", FT] = E 21(1)”; = 1 E[z"f(l)|F,] = 1. Z"1'(t) zvf(t) (3.3) E[z"f (1,1) So E(Z“(1)) = P(A) + P(Ac) = 1. Since for each v e D, (3.4) EV[Z(1) I F.] =E[Z(1) Z" s}. Each [in e D since, with a calculation like (3.3) justifying the third equality in (3.8), (3.8) E(Z“"(1)) = E(EIZ“"(1) I F.]) = E(Z“(s)EIZ"" E(Vo(t)) is a right continuous function [See Lipster and Shiryayev (1977), Theorem 3.1, page 55]. Fix t, and suppose tn > t, n = 1, 2, ..., satisfy t = limin t... For each v in D, it holds for all n that Vo(tn) 2 E"[Z(1)I Ftn ] a.s. Also, for each v e D, since (1 .20) implies that min05(51(Z"(1)') > 0 a.s., we have E[Z(1)Zv(1)IF.l . (3.11) ' E‘IZ<1)IF.1 = 2%) Then the paths EV[Z(1)I F.] are continuous a.s., since E[Z(1)ZV(1)I F.] is a (P) martingale with respect to the Brownian filtration {Ft}. Therefore, for each v e D, 2 In the second and final equalities in (3.10) we use the following consequence of that for each v e D, ZV is a (P) martingale. If ‘t S B are stopping times and Y is an Fig-measurable random variable, then for any v e D such that Y is (P') integrable, “E“[YI F.] = ElY Z'(I;B)| F.]. 34 lim infn Vo(tn) 2 EV[Z(1) I F t], implying lim infn Vo(tn) _>_ Vo(t) a.s. Then by the F atou Lemma, (3.12) lim inf,E(vo(t,)) 2 E(lim inano(tn)) 2 E(Vo(t)). Since V0 is a (P) supermartingale by (i), E(V0(tn)) S E(Vo(t)) for each 11. Therefore (3.12) implies that limnE(Vo(tn)) = E(Vo(t)). so we have the equivalent condition, and the process denoted by V in (iii) exists. For (iii), suppose \7 is an adapted process with (as) cadlag paths such that {7(1) = 2(1) a.s. Suppose that there exists a q e Q = {rational q: 0 S q S 1} such that P{ V(q) < V(q)} > 0. (By right continuity, { V(t) — V(t) 2 0 for all 0 S t S 1}c equals I quQ{ V(q) — V(q) < 0} modulo null sets.) Apply Lermna 2 to obtain the monotone convergence V(q) = lianEv"[Z(1) I Fq]. Then, for n sufficiently large, we see that \7 cannot be a (Pvn) supermartingale from equivalence of P and P"n and both {7(1) = 2(1) (P) as. and P{ v(q) < EV“[Z(1) | Fq]} > 0. I Corollary 1. P{V(t) S Z(t)f0r all 0 S t S 1} =1. Proof. The result will follow from (iii) of Theorem 2 once we show that 2(a) is a (PV) supermartingale for each v e D. For each v e D, any adapted process Y(-) is a (P‘) supermartingale if and only if Y(-)Z"(-) is a (P) supermartingale. If v e D, we have A®P{0(t,m) e Ran[o'(t,m)], v(t,a)) e (Ran[o‘(t,o)])i} =1, so that Z(-)ZV(-) is (P) supermartingale 29*V(.). I 35 3.2 An Essential Lemma in the Link to the Arbitrage Problem. Lemma 3 below will be essential in constructing an arbitrage in the proof of the final theorem in Chapter 4. In the proof of Lemma 3, it is necessary to use the Doob~Meyer Decomposition of a cadlag supermartingale. Below is the formulation of the Doob- Meyer Decomposition Theorem given in KOpp (1984) [Theorem 3.8.10, page 122]. Theorem (Boob-Meyer Decomposition.) LetX be a right-continuous supermartingale. Then X has a unique decomposition X = M - A, where M is a local martingale and A is a predictable increasing process. The conclusion of Theorem 2 that for each v e D, V is a (P‘) supermartingale with cadlag paths implies that for each v e D we have the unique (PV) Doob-Meyer Decomposition V = Lv - A". (Uniqueness is up to (PV) indistinguishability, which is the same as (P) indistinguishability by equivalence of P and PV.) Each Lv is a (P‘) local martingale, and each AV is an adapted process with cadlag and non-decreasing paths satisfying A"(O) = 0. Lemma 3. Let L denote the (P) local martingale in the (P) Doob-Meyer decomposition of V. Then L(-) = V(0)Z°(o) for an adapted Rd-valued process (p satisfying (3.13) A®P{ 0: L(t) = 0}, then for each (t, (1)), t 2 Q(L).,, implies 36 (p(t,o)) = 0. [See Lipster and Shiryayev (1977), Lemma 6.2, page 208.] (3.13) will follow once we prove (3.14) For each v e D, 71®P{(p’(t,a))v(t,m) S 0} = 1. To show that (3.14) implies (3.13), suppose that (3. 13) does not hold. Then define the process I10 by putting uo(t_.o)) the projection of (p(t,0.)) on Ker[0'(t,(t))] = (Ran[o"(t,a)])i. Then we have that A®P{ uo(t,03) ¢ 0} > 0. Let process It be given by 110) = #1100). if 11.0) a o; 110) = 0 if too) = 0. ”110(1)“ The Novikov Criterion implies that us D because u is bounded uniformly in (t,c0). (3.14) does not hold for this It, since (p'u = ”up”. To prove (3.14), consider that for each v e D, where (PV) Brownian Motion W" is as defined in (2.4), it is possible to write the (PV) Doob-Meyer Decomposition of V as I II: (3.15) V(t): V(O) + Irv (s)dw"(s) —AV(t), OStSl, 0 1 where f’ is an adapted Rd-valued process such that Illf” (t)II2dt < 00 as. With respect to 0 (3.15) note that for each v e D, Lv is adapted to {F1}, and not necessarily adapted the P-augmentation of the natural filtration of W". If this latter filtration is denoted { Ftv ; 0 S t. S 1}, then it is possible that there exists an adapted v such that there exist t such that FtV are strict subsets of the corresponding Ft. Therefore, it is not possible to directly apply the Martingale Representation Theorem to obtain the stochastic integral 37 representation of L"(-) in (3.15). If V(O) = 0, then because V(-) is a nonnegative (P) supermartingale, V(-) is indistinguishable from the zero process, and clearly we may take F to be the zero vector process for each v e D. 3 If V(O) > 0, then for each fixed v e D, to construct the process f’ start from that Lv(o) is a (PV) local martingale is equivalent to that LV(-)Z"(o) is a (P) local martingale. Then since LV(.)Z"(.) is a nonnegative process, there exists an adapted Rd-valued process g = g" satisfying org = 1 as. such that L"(-)ZV(-) = V(0)Zg(.), Then the K0 formula gives that dZ" (t) (3.16) dD(t) = (1le = W» ng, V(0)Z‘(t) 2:0) 7(7) 9 ‘ (an) — V(O) ),,a28(.)ZV(.) t) + V(O—)—-Zx(’)sz(),Z"(.) t) (2%) [ I (2%)? [ I V(O)Zg(t) =_Z.V_()_Ig*(t)dwo) +v*(t)dW(t) —g*(t)v(t)dt +IIVII2(t)dt I = Lv(t)(v*(t) — g* (t)XdW(t) + v(t)dt) So we have equation (3.15) with f’(-) = LV(-)[v(o) — g(o)]. Then because dL(t) = —L(t)(p(t)dW(t), 0 S t S 1, it follows from uniqueness of the (P) Doob-Meyer Decomposition and equivalence of the (P‘) that for each v e D (3.17) 7L®P{f’(t,0)) = —L(t,0))(p(t,(o)} = 1. and consequently, up to indistinguishability, 3 If Y(t), 0 St S 1, a process with cadlag and non-negative paths. is an ((2, F, {F.}, P) supermartingale, and S = inf{t > 0: Y(t) = 0}, then P{Y(t) = 0 for all S < t S 1} = 1. [See Elliott (1982), Theorem 4.16, page 38.] 38 t (3.18) A"(t)= A(t) — ]L(s)tp*(s)v(s)ds; OStSl. 0 Equation (3.18) motivates the completion of the argument. F ix v e D. Then define for each n a process un e D, with E(Z""(1)) = 1 following from uniform boundedness of un in (t,m): Io“(t)v(t) > 0} Ia*(t)v(t) s OI 3.19 n t = t t . Then, by (3.18), 1 at (3.20) A”n (l) = A(l) — n I {‘8 (t)V(t) > 0jL(t)l| _ ‘anmsoI . OI 1+I|v(t)|I L(t)

O}> 0}the right-hand side of (3.20) tends to —oo as n —> oo. (Recall that for each (t,o)), L(t,o)) = 0 implies cp(t,ct)) = 0 6 Rd.) Such divergence would contradict that for each n the paths A1L1 n(-) are non-decreasing. 39 3.3 Two Useful Characterizations of the Process V(o). It will be useful that the following result holds when we stop process V. Karatzas and Shreve (1998) give a different presentation of this proof which is based upon a similar underlying approach [see Remark 5.6.7, page 215.] Lemma 4. For each stopping time 0 S t S 1, (3.21) V(t) = ess szrpveDKEv[Z(I) /F,]). Proof. For each 1, denote the right-hand side of (3.21) by Y(t). First assume that 1: is a simple stopping time, ‘t = 21515,, {t = tk}tk. Then for each v e D, (3.22) EVIz(1)|F,] = §{e=tk}EV[Z(1)|F,k]. k=l (The right-hand side of (3.22) is FT-measurable because {I = tk} O A e F I for any A e Ftk , Equality of integrals on F: sets follows from that {t = tk} O A e Ftk for any A e F.) Evident from (3.22) is that Y(T) 3 2,5,5,“ = amt.) = V(T). For the reverse inequality, choose by Lemma 2 sequences {vk,m; m 2 1} c_: D for k = l, ..., n such that for each k, limmIEvk’m[Z(1) IFtk ] = V(tk) a.s. Then, in light of (3.22), it holds as. on {t = tk} that Y(I) 2 liran Evkm[2(1) | F.] = V(t). Now let I be an arbitrary stopping time. Stopping times Tk, k = 1, 2, ..., given by . ‘-l . . -J—on J——_ 1 a sequence {vk,m; m 2 1} g; D such that (3.24) V(tk) = luanE"“~'“[2(1) |I~‘Tk ] a.s. Define adapted processes ukm, k 2 l, m 2 1, by “k,m(t) = vk,m(t){t > 11,}. Then each um e D. (Calculation (3.3) shows that for any v in D and stopping time B, v(-){- > B} is also in D.) For each k, we have (3.24) with the processes um replacing the processes vbm. Furthermore, for each k,m, with the last equality in (3.25) holding because ZN"m (1; H) = 1 as, (3.25) Y(t) 2 E“"’"‘[Z(1) I F.1= E““’"‘[E“"“‘[Z(1) IF.,I I F.] = E[E“"‘"‘[Z(1) Ink] | F.] as. (Refer to footnote 2 on page 33 for more detail on this calculation.) Letting m —> 00 for fixed k in (3.25) shows that for each k, Y(t) 2 E[ V(rk) I F.] as. Therefore, the Fatou Lemma implies (326) Bore» 211m sup. E(Emtl) I F.]) = 1m. E(V(rt)) 2 Ewe». Then Y(t) S V(t) as. and (3.26) imply Y(t) = V(t) as. 41 Although properties of the processes Z"(o) dependent upon v e D were used in proving properties of the process V(o) above, for each t, V(t) is in fact a version of an essential supremum taken over a much wider class than D. Define the class N by N = {adapted Rd-valued processes v: or" = 1 as. and l®P{v(t,m)e Ker[o(t,o))]} = 1}. Lemma 5. For each stopping time 0 S t S 1, (3.27) V(t) = ess supv.E)(r(E[Z(1)Zv (1:;1) /F.]). Proof. D _c_ N implies "S". For "2", fix a stopping time t and v e N. Then define two sequences, one of stopping times BH and the other of adapted processes v", by t (3.28) Bn = inf {t > O: I{t < s}IIv(s)II:2 ds 2 n}, 0 (3.29) vn(t) = v(t){1: < t S Bn}. Each vn e D. Clearly A®P{vn(t,m) e Ker[o(t,0))]} = 1, and we have for each n that l JIIv,(t)I|2dt S n as. The Novikov Criterion implies that E(ZV“(1)) = 1. 0 The paths Zv(r; .) are continuous a.s., and 2""(13 1) = ZV(T; B"). Therefore, limn Zv"(t; 1) = Zv(r; C(13)) = Zv(‘t; 1) as. Conclude through the Fatou Lemma: (3.30) V(t) 2 lim supn EV"[Z(1) | F.] = lim supn E[Z(l)Zvn(t; 1) | F.] 2 E[Z(1)ZV(1:; 1) | F.] as. I 42 Chapter 4 Characterizations of Arbitrage and Corollaries 4.1 Characterization of Arbitrage in Terms of Domination of Z(t; l). The following simple fact is presented first to avoid repetitive justification. Lemma 6. Let (p be an adapted Rd-valued process satisfying 01‘p = 1 a. s., and let 0 S t S 1 be a stopping time satisfiing P{r < I} > 0 such that processes Z‘”(o) and Z(t; .) are not indistinguishable. Then there exists a constant 0 < 5 < 1 such that the stopping time B defined by B = inf{t > t : Z”(t) S 6Z(t; t) and Z(t; t) > 0} satisfies P{B < I} > 0. Proof. Suppose that no such 8 exists. Then since the paths Z‘p(o) and Z(t; .) are nonnegative and continuous, (4.1) P{Z"(t)2Z(t; t) forall‘tStSl} =1. In particular, Z“’(1:) 2 Z(t; r) = l as. Then that Z‘”(-) is a continuous supermartingale with Z"(O) = 1 as. implies P{Z‘p(t) = Z(t; t) for all 0 S t S t} = 1. 43 Furthermore, (4.1) implies that that CWO) 2 §9(r) a.s. Then there exist localization stopping times 1],, with limklnk = §°(r) as. such that for each k the process Z"(- A m) — Z(r; . A m.) is a nonnegative martingale. Then 23(0) — Z(‘t; O) = 0 as. implies that each of these localized martingales is indistinguishable from the zero process. Then path continuity implies both that (4.2) P{Z‘p(t) = Z(t; t) for all 0 s t _<_ 6(a)} = 1 and that Z‘°(§9(t)) = 0 as. on {§9(t) < 1}. Because Z‘”(-) is a nonnegative supermartingale, we can replace §9(t) by 1 in (4.2). So Z‘p(-) and Z(‘L’; .) are indistinguishable. I The following theorem is an extension of Corollary 3 of Levental and Skorohod (1995) [page 920]. It serves as the link between the V process detailed in the last section and our arbitrage problem. Theorem 3. Assume absence of immediate arbitrage. Arbitrage exists if and only if there exists both an adapted Rd-valued process (p satisfying A®P{cp(t,(1)) e Ran[o'(t,w)]} = 1 and P{a‘” = I} = I and a stopping time 0 S ‘I.’ S 1 such that the processes Z‘p(t; .) and Z(t; .) are not indistinguishable and P{Z‘” (1:; 1) 2 Z(t; 1)} = 1. Proof. (Sufficiency) Lemma 6, applied to (p(-){- > 1}, implies that there exists a constant 0 < 5 < 1 such that the stopping time B defined by (3 = inf{t > a; 2““ (r; t) s 52(1; t) and Z(T; t) > 0} satisfies P{B < l} > 0. Assume first that Z(r; 1)> 0 as. on {B < 1}. Then it also holds as. on {B < 1} that (,3) 1 2 2304) : [Zimmjvwnj . 52mm Z(I;l) Z(nB) Z(B;l) Z(Bd) So zt(8;1)/z(p;1) 2 8‘1 > 1 as. on {B < 1). It follows from the Itc‘) Formula that on {B < t < §9(B)} Zion) '_ Z‘Past) (4.4) d{ 2033)] _ wt) ———(9* (t)— (p *(t)IdW(t)+9(t)dt). (The computational details underlying (4.4) are precisely the same as those given explicitily in calculation (3.16).) Let ((3'),. denote an adapted process such that for each (t,w), (o')+(t,(o) is an invertible dxd matrix and (4.5) o‘(t,to) (6);} (t, u) x = x for all x e Ran[o‘(t,m)]. Lemma 1 guarantees that such an (0'). exists. Then because of the range hypothesis on (p and the construction of 0, (4.6) l®P{o (on) (0— (p)= 0— (p}=1. Choose the portfolio 71 as follows: (47) n0) = Bari—LII," ;){B< 130)(90)— 4.0)) We have assumed Z(‘t;l) > 0 as. on {B < 1}. So Z‘p(t;1)2Z(t;l)> 0 as. on {B < 1}. Then since B 2 t, 45 (4.8) [{0 <1}Ip (1)132 +|Itp(t)|flit < 00 as. The paths B(-){Z‘° (B; .) / Z (B; o)} are bounded in t as. because it holds as. that nonnegative supermartingale Z(B; .) cannot hit 0 on {Z(t;l) > 0}. So, in light of (4.8), 7: satisfies the integrability constraint required of a portfolio. Where random variable M is given by M = 5111305151 B(t)%(-IiB—;;t_t)2’ 1 2 (4.9) IUIo*(t)n(t)“ + In*(t)a(t)|]dt s o (M2 +1) I11) «41910-00112 + I(e‘(t)- 1 as. on {(3 < 1}, where P{B < 1} > 0, completes justification of that 1: is an arbitrage. Now drop the assumption that Z(r; l) > 0 as. on {B < 1}, but treat the situation where P({Z(T; 1) = O}O{B < 1}) > 0 under the added assumption that P{C"(B) 2 §6(B)} = 1. The It?) Formula gives that on {B < t < Q9(B)}, 46 1 1 . (4.11) ((2033)) _ Z(B;t)9 (t)(dW(t)+9(t)dt) With constant 5 and matrix-valued process (6’). as they were earlier in the proof, define the portfolio rt' by Z¢(B;’){ Z(B;!) 1) < r s v}(c*);‘(t)(e(t) — 40)) (4.12) rt'(t) = B(t) K :1: -1 B(t)—__Z(13;t){B < r s vim )+ (090) , where K = (8"1 — 1) / 2 > 0, and st0pping time y is defined 7 = inf{t > 0: Z(B; t) = K /(2 + 2K)}. Path-continuity on (B; .) and K /(2 + 2K) < 1 imply that a.s., Z (B; l) > 0 on {y = 1} and (4.13) v < 49(1)) on {Z(B; 4903)) = 0} = I I{B < 911910124 = 00}. Paths Z‘p (B; . A y) / Z(B; . A y) and K / Z(B; . A y) are bounded in t as. S0 inequality (4.14) below shows that (4.13) and P{§¢(B) 2 §9(B)} = 1 together imply that n' satisfies the portfolio integrability constraint. With random variable M' defined by Z"(fl;t Ar)+ K Z(fl;t) M, = 3111305151 3(1) 9 47 (4.14) I{IIo*(t)a'(t)II2 + Ia'*(t)a(t)l}dt s O (M'2 + 1) I {p < t s y}I|o(t) — a(t)|I"' + |(e‘(t) — a‘(t))e(t)|}lt + l 2(M'2 + 1) I {,6 < z s y}IIt9(t)H2dt. 0 (Since Z(t; B) > 0 on {B < 1} and Z(B; 1) > 0 as. on {y=1}, P{Z‘p (r; 1) 2 Z(r;1)} =1 1 implies that IIB < t}II(p(t)I|2 dt < 00 as. on {y = 1}. ) 0 The associated discounted capital gain process satisfies _ 2¢(/3s) K (4.15) X20) - IMGQI d[Z(,B;s) + Z(fl;s)I : Z"(B;t/\T) _ 1 + K Z(B;t A T) Z(B;t A 1’) So 7t' is (1 +K)-tame. On {B =1}, an(1) =0 a.s. On {B <1}fl{y< 1}, X..'(1)21+K a.s. With regard to Xnv(l) on {B <1}fl{y =1}, recall that we have Z(T; B) > O on {B <1} and Z(B; 1) > 0 as. on {y = 1}. Then calculation (4.3) shows that 28(13;1)/z(p; 1) 2 8“ as. on {B <1}n{y=1}, so that an(1) 2 6'1 — 1 — K > 0 as. on this set. So if is an arbitrage. 48 All that remains in the sufficiency argtunent is to produce an arbitrage given P{C"(B) < §9(B)} > 0. In this case, define the portfolio it" by B(t) —————- I:“’(B) 0: Z(Q“’(B); t) = 1 /2}. If 48(9) < 49(0), then 116 «119101124 = 00, so that 2‘90); 1) = o as. on {43(0) < §°(B)}- Since I S B, we also have Z“’(r; l) = O as. on this set. Therefore, Z‘p(t; l) 2 Z (t; 1) as. implies (4.17) z (r; 1) = 0 as. on {$10) < 49113)}. On 13(1)) 0 on {9115) < 49(0)) c; {13 < 1). and Z(B; 43(0)) > 0 as. on {Q‘p(B) < §6(B)}. Therefore, path continuity. (4.17), and (4.18) imply that p < §°(B) as. on {§“’(B) < §°(B)}. Then a calculation like (4.14) shows that 7:" is a suitably integrable portfolio. it" is an arbitrage since 1 Z(C“’(fl);t/\p) (4-19) Xfl-(t) = I - 1]I§”(fl)<§9(fl)l 49 Observe that 1t" is l-tame, Xn~(1) = O as. on {C(13) _>_ 3(6)}, and Xn»(1) = 1 as on {:m»<§%m) (Necessity) Suppose that no is a C—tame arbitrage portfolio. Then defining 1: = (20’an produces an arbitrage for which P{Xn(t) + 1 > O for all 0 S t S 1} = 1. Define the adapted vector process (p as follows: 1 . (4.20) (0(1) _ a(z) - magma). Then O: Xn(t) ¢ 0}. It follows from the definition of r and that dX,,(t) = B"(t)1t'(t)o(t)[dW(t) + 6(t)dt] that up to a P-null set, (4.22) {r<1}= {Masts1:o‘(t)n(t){rst<(r+g)A1}¢0) >0 foralls>0}. P{a = 1} = 1 implies 6(1) > r as. on (I < 1}. That ‘11: is an arbitrage implies P{r < 1} > 0. So (4.22) implies that x®P(cp(t,m)(r,, < t < a(t),} i 6(t,w){rw < t < 6(1),») > 0. 50 Then to see that Z“’(r; .) and Z(r; .) are not indistinguishable, examine the explicit formula for process Z"(1:; o), (1.11), and conclude that the map [v] —-) Z"('c; .) defined on the equivalence classes of adapted Rd-valued processes v satisfying 01V = 1 as. partitioned by the relation V15 V2 if A®P(Vl(tam){‘tm < t < €Vl(T)w} = V2(t,(t)){‘tw < t < Cvz(1)w}) = 1 is an injective map. Now calculate that on {O s t < §9(r)}, Z‘p(1:;t) _ ’ {t 0} ; {gem = 1} that x, is an arbitrage capital gain, and on {Z(m) = 0} that Z‘” (I; .) is a nonnegative process, we obtain that P{Z“’(t; 1) Z Z(‘L‘; 1)} = 1. I 51 Remark. Under the assumption that P{a = 1} = 1, the proof of necessity in the previous theorem characterizes the capital gain process Xn(-) associated with an arbitrary arbitrage 1t(-) as C({Z"°(r; .) / Z(r; o)} — 1), for a constant C > 0, on the stochastic interval {(t,(t)): 0 S t < §9(r)m}. Here I = inf {t > O: Xn(t) at O}, and (p is an adapted process satisfying a range requirement. Therefore, it is interesting to question if it is ever necessary in producing arbitrage to hold asset shares at times t 2 gen) when §9(r) < 1, i. e., at times t when the ratio Z“’('c; t) / Z(r; t) is undefined because Z(r; t) =10. The answer is no. Assume absence of immediate arbitrage. If arbitrage exists, then there is an arbitrage 1: such that, where T = inf {t > O: Xn(t) ¢ 0}, there exists a stopping time 7 such that y < 6(1) a.s. on {Z(t; 6(1)) = O} and X,‘(1) = Xn(y) a.s. To justify the assertion in the preceding paragraph, suppose it is an arbitrage. Put stopping time i = inf {t > O: X {I (t) :t 0}. By the proof of necessity in Theorem 3, there exists an adapted Rd-valued (p satisfying a” = 1 a.s., A®P{ %:Z‘P(i ;t)SBZ(%;t)andZ(% ;t)>0} satisfiesP{B<1}>O, K=(5"—1)/2 > o, and y = inf{t > B: Z(B; t) = K /(2 + 2K)}. 1: is an arbitrage, and Xn(1) = Xn(y) a.s. Let T = inf{t > 0: X,,(t) at 0}. Note that i <5 s r s 7 holds a.s. on {i < 1}. We have 7 < cam) a.s. on {Z(B; 6(5)) = 0} [c. f. (4.13)]. Therefore, 13 s 1: < 6(1)) a.s. on {Z(B; @903» = 0}, and consequently, 4%) = @945) a.s. on {Z(B; c911)» = 0) and {211; do» = 0} = {Z(B; 4113)) = 0} a.s. So 7 < 6(1) a.s. on {Z(r; 33(1)) = O}, and we have proven the assertions in the remark. 4.2 Characterization of Arbitrage in Terms of the Processes V(r; .). Lemma 7. Assume P{a = 1} = 1. Then Z(r; o) is indistinguishable fi-om V(r; .) for all constants 0 _< r S 1 If and only if Z( r; .) is indistinguishable fiom V(r; .) for all stopping times 0 S 2'5 1 . Proof. Sufficiency is obvious. (N ecessity.) First prove that for any stopping time T, in order to prove Z(‘t; .) and V(r; .) are indistinguishable it suffices to show (4.25) For each 0 S t S 1, V(r; t v r) 2 Z(‘L‘; t v t) a.s. Corollary 1 states that P{V(T; t) S Z(‘L’; t) for all 0 S t S 1} = 1, so that we may replace "_>_" in (4.25) by "=". The paths V(r; .) are right-continuous, and the paths Z(‘t; .) are continuous. So (4.25) implies 53 P{V(‘C;t)=Z(T;t) foralltStS l} = 1. In particular, V(T;‘r) = Z(t;r) = 1 a.s. Since V(r; .) is a (P) supermartingale by Theorem 2 and V(r; 0) S Z(‘C; O) = 1 a.s., we then have For each 0 St S 1, V(r; tA'c) = 1: Z(T; t /\ T) a.s. Then by right-continuity, Z(‘C; .) and V(r; .) are indistinguishable given (4.25). Now let 1' be a simple stopping time, T = Zfir = tj}tj. Fix 0 S t S 1. Because of indistinguishability of Z(tj; .) and V(tj; .) for each j, Lemmas 2 and 4 imply that for each j there exists a sequence {V1.16 k 2 1} g D such that a.s., (4.26) lika EV” [Z(t j;1) Fm] = V(tj; t v r) = Z(tj; t v r). Since for each v e D, (4.27) E"[Z(t;1) FM] = 2,-{r = tj } EV[Z(t j;1) PW] it follows that (4.28) lika E"1-*[Z(r;1)l=,,,] = Z(‘C; t v 1:) a.s. on {r = tj}. Since supJ-,kEV"k [Z(‘C;1)Ftvt] S V(r; t v T) a.s., (4.25) holds for this simple I . Now let 0 S 1: S 1 be any stopping time. There exist simple stopping times In 2 1' with limnlrtn = r a.s. Fix 0 S t S 1. By the preceding paragraph, for each n V(t v In; .) and Z(t V In; .) are indistinguishable processes. So for each 11, there exists a sequence {vm k; k 2 1} g D such that a.s., 54 (4.29) lika Ev” [Z(t v Tn;l) Ftwn] = V(t v In; t v 1,.) = Z(t v tn; t v In) = 1. For each n, k, the adapted process link, defined by pn,k(s) = vn‘k(s){s > t v In} lies in D [see (3.3)], and we have Ew‘ [Z(t v 7:51le, = E“ [Z(t v 7,51) EV," ]. For each n, with the last equality below holding because Z"""‘ (t v r;t v In ) = 1 a.s. for each n, k [see footnote 2, page 33], (4.30) V(t;tv1) 2 likaE“"~*[Z(r;l)F,,,] = Z(r;tvr) lirnkTE““"‘[Z(tv1:;1)FM] = Z(T; t v ‘1') lika E ”""‘ [Z (t v r;t v 2'")E”"'* [Z(t v r";1) EWJEVJ = Z(‘t; t v T) lika E[Z(t v 1:;t v tn)E"“'k [Z(t v In ;1) PW," 117nm]. Then monotone convergence (4.29) yields that for each n, (4.31) V(t; t v 1') _>. Z(‘C; t v r) E[Z(t v t;t v en)|F,,,]. We have lirnn Z(t v 1.“; t v In) = 1 a.s., so that (4.31) and the Fatou Lemma imply V(‘t; t v 1:) 2 Z(‘L'; t v I) a.s. I Theorem 4. Assume absence of immediate arbitrage. There is no arbitrage if and only if processes Z(r; .) and V(r; .) are indistinguishable for all constant times 0 _< r S1. Proof. (Necessity) Suppose that Z(r; .) and V(r; .) are not indistinguishable. Consider the (P) Doob-Meyer Decomposition, cadlag V(r; .) = L(-) — A(-). Lemma 3 implies that L(o) = V(r; O)Z"’(o) for an adapted Rd-valued process (p satisfying ()t‘p = 1 a.s. 55 and X®P{cp(t,o)) e Ran[o‘(t,m)]} == 1. A(-) is an adapted process with non-decreasing and cadlag paths starting at A(O) = O. From V(r; O) S Z(r; O) = 1 a.s. and V(r; 1) = V(r; O)Z“’(1) — A(1) = Z(r; 1) a.s., we obtain that Z‘”(1) 2 Z(r; 1) a.s. It cannot be that Z“’(-) and Z(r; .) are indistinguishable: otherwise, in order for V(r; 1) = Z(r; 1) a.s. to hold, we would need both V(r; O) = 1 and A(1) = O a.s. These values would contradict that Z(r; .) and V(r; .) are not indistinguishable because the paths A(-) are non-decreasing. By Lemma 6 then, there exists a stopping time B 2 t for which (4.32) P{p<1}>0 and O_ Z(B; 1) a.s. If P{Z‘p(B; 1) > Z(B; 1)} > 0, then Z‘p(B; .) and Z(B; .) are not indistinguishable, so that Theorem 3 implies that an arbitrage exists. If Z‘”(B; l) = Z(B; 1) a.s., then Z‘°(1) 2 Z(r; 1) a.s. and (4.32) imply that Z(B; 1) = O a.s. on {[3 < 1}. In this case, apply Theorem 3 with o, the zero vector process, in the role of (p. We have o(t,(n) e Ran[o‘(t,co)] for each (Leo) and (4.33) 2°43; 1)=1 2{B=1}=Z(B;1). Z°(B; .) and Z(B; .) are not indistinguishable, and so an arbitrage exists. (Sufficiency) Suppose that there exists an arbitrage given a = 1 a.s. By Theorem 3, there exist both an adapted Rd-valued process (p satisfying or‘p = 1 a.s. and X®P { O. The existence of such a V(-) implies through (iii) of Theorem 2 that Z(t; .) and V(‘t; .) are not indistinguishable. Then Lemma 7 implies that there exists a constant time r such that Z(r; .) and V(r; .) are not indistinguishable. Lemma 6 implies that there exists a stopping time B > T such that (4.32)’ P{B<1}>Oand0 [0, 1] be a continuous and strictly decreasing deterministic function satisfying g(O) = 1. Define stopping times 7 and n as follows: (434) Y = inf{t > BI Z‘p(T; t) = g(t — B)Z(t; 0} n = inf{t > y: Z“°(t; t) = Z(t; t)}. Path-continuity of Z‘p(t; .) and Z(‘t; o), (4.32)’, and that Z‘p(t; 1) 2 Z(‘L’; 1) a.s. together imply that a.s. on {p < 1} (4.35) Z“’(t; 7) = g(11 - B)Z(t; )1) < Z(t; Y) and Z‘”(t; n) = Z(t; It). Now define 17(.)by (4.36) V(t) = a((lt A 11 - B)”)Z“ 1 — c1. Define a sequence of stopping times {[31,}. ; k 2 1} by 51.1: = inf{t>r: Ze+p1(r;t) v Z“1(r;t) =k} 59 Then 81,1. = 1 for k sufficiently large holds as., so that the monotone convergence theorem implies that there exists a k; such that (4.42) E(29+“1 1 — a. (111 e D implies that Z”1 (.) is a martingale process.) Denote 1:1 = 131, kl . Now consider that the absence of arbitrage, Theorem 4 and Lemma 7 together imply that V(‘L’l; .) and Z(t]; .) are indistinguishable. So there exists a sequence (v?) ; n 2 1} in D such that (4.43) limnT 13129"le(1:51))?Tl ] = 1 a.s. Then (4.44) 1im.E(ZG*“I maze”? (151)) = (2) limnTE(Ze+“1(r;tl)E[Z9+V" (t1;1)F,l]) = E(Ze+“1(r;rl)). Because 29+“) (r;o /\ TI) is a uniformly bounded local martingale, and therefore a martingale, E( Z 6+” (r;t])) = 1. Therefore, there exists a 112 e D such that E(Z9+“1 (r;tl)29+“2 (t1;1)) > 1 — s2. By considering the stopping times (132,1. ; k _>. k; + 1} defined by 9 Br... = inf{t>‘tt : ze+“l(rn.)z Wont) v z“1(r;tt)z“2 1 — 82. Then by induction, there exist a sequence of processes {un ; n 2 1} in D and a sequence of stopping times to S 121 S 12 S (put to = r) such that the processes Gn(-) and Hn(o) defined by ' n' 6 . n _ (4.46) 6.1-) = 1’12 +“j(Tj—1;'/\Tj)s H.(-) = HZ“J(t,--i;-Ar,-) . j=l j=1 are uniformly bounded martingales satisfying (4-47) E(Gn(1){tn =1}) AE(Hn(1){Tn =1}) > 1-8n. Define the adapted Rd-valued process 1.1 by (4.48) 11(1) = 2141mm «5 tn}. n=l Then A®P{ 110,19) 6 Ker[0'(t,ot))} = l. on“ = l a.s. so that the processes Z“(-) and 29+“(.) are well defined and continuous. Our method of selecting the stopping times tn imply that there exist integers N1 < N2 < such that for each n, r, = inf{t > o : zero; t) v Z“(r; t) = N}. 61 Therefore P(Un{‘tn = 1}) = 1. So the summation in (4.48) has finitely many non-zero I ., summands as, and consequently, fl|p(t)“"dt < 00 a.s. By monotone convergence, 0 E(Z“(r;1)) = limnT 15(H,.(1){tn = 1}) = 1. So 1.1 e D. Similarly, to complete the proof, E(Z°*“ 0: Xn(t) ¢ 0}. The folloWing result is interesting in comparison with that conclusion. Note that the assumption of Corollary 2 below implies that for all stopping times 0 S ‘t S 1, 6(1) = 1 a.s. Loosely speaking, we need check for indistinguishability of V(r;o) and Z(r;o) for r > 0 only if an explosion of 9 occurs before time 1 and "kills the market initiated at time 0.” Corollary 2. Assume P{Qem) = I} = I. Then there is no arbitrage if and only if processes V(-) and Z(-) are indistinguishable. Equivalently, there is no arbitrage if and only if there exists a 11(o) e D such that E(Z9 +11(I )) = 1. Proof. The proof of Theorem 5 establishes the equivalence of the two characterizations. For necessity, apply Theorem 4 with r = O. 62 (Sufficiency) Assume V(.) and Z(-) are indistinguishable. Since P{gem) = 1} = 1 implies P{a = l} = 1, Theorem 4 applies. Since Z(l; .) is indistinguishable from the constant 1 process, V(l; .) is indistinguishable from Z(l; o). Now fix 0 S r < 1. As in the proof of Lemma 7, to prove that V(r;-) and Z(r;o) are indistinguishable it suffices to show (4.25)’ For each t with r S t S 1, V(r; t) 2 Z(r; t) a.s. If r S t S 1, then by Lemma 2 there exists a sequence vn in D such that (4.49) limnTEV"12(1)1F.] = V(t) = Z(t) a.s. Equivalently, since r S t, (4.50) Z(t) = Z(r)(lim,,TEV"[Z(r; l)11=.] ) a.s. P{§9(O) = l} = 1 and r < 1 imply that Z(r) > O a.s. so that a.s. (4.51) V(r; t) 2 limnT Ev"[Z(r; l)|Ft] = Z(t) / Z(r) = Z(r; t). In the statements of the remaining corollaries, recall that we assume absence of immediate arbitrage, i. e., we assume a = l a.s. The following result is generally understood; Levental and Skorohod (1995) prove it without reference to invertibility of 0' [see page 917]. Here it is an easy corollary. Corollary 3. If E(Z(t; 1)) = I for all stopping times 0 S r S I, then no arbitrage exists. 63 Proof. The zero vector process lies in D. Apply Theorem 5. I The following Corollary explicitly formulates the link between Theorem 5 and the notion of an equivalent local martingale measure for the stock price processes. 1 . Corollary 4. Assume j H9(t)”’dt < oo a.s. Then there is no arbitrage if and only if for 0 eachi= 1, d, 6151(1) = 5,0) Brandi/1.0), lSde where there exists a probability measure Q ~ P such that W(-) is a d-dimensional standard Brownian Motion with respect to (Q, F, {Ft}o 5 t S 1, Q). Proof. No arbitrage is equivalent to that there exists a v(-) e D such that 1~:(z9 * v(1)) = 1. Let W.) be given by w, (t) = Wi (t) + 1(e,(s) + v,(s))ds; 1 s i s d. 0 Apply Girsanov's Theorem with probability measure Q equivalent to P defined by dQ/dP =z"*V(1) . I l 7 Under the assumptions Hp (t)||‘ dt < 00 a.s., Z(l) is a.s. bounded, and E(Z(1)) < 1, 0 Levental and Skorohod construct an arbitrage which does not require invertible volatility [(1995), Example 5, page 924]. The following corollary accomplishes the same result under a slightly weaker integrability assumption. 64 Corollary 5. If 6(0) = 1 a. s. and there exists a constant C for which Z(I) S C a. s., then there is no arbitrage if and only ifE(Z(])) = 1. Proof. (Sufficiency) If E(Z(1)) = 1, then the proof of Corollary 3 implies that Z(-) and V(-) are indistinguishable. Then Corollary 2 implies absence of arbitrage. (Necessity) For each v e D and O S t S 1, EV[ Z(1)| Ft] S C. Then it follows from right-continuity of paths V(o) that P{V(t) S C for all 0 S t S 1} = 1. Theorem 4 states that absence of arbitrage implies that Z(-) and V(-) are indistinguishable. Therefore, Z(-) is a uniformly bounded local martingale. (3(0) = l a.s. implies that we can take 1 as the limit of the localization stopping times.) A uniformly bounded local martingale is a martingale. [See Revuz and Yor (1991), Proposition IV.1.7, page 118.] Therefore, E(Z(1))= 1. I 65 Chapter 5 Examples In all examples, assume that r is the zero process. Example 1. This is an example where E(Z(1)) < l and Z(l) > O a.s., but V(-) and 2(0) are indistinguishable. This example shows through Corollary 2 that when 0(a) can be singular, the result that E(Z(l )) < 1 implies existence of arbitrage is not true. The example can be written in one dimension, but is easier to read in a construction with d=2.La l 0 (Y(t) E [O 0]. To define 9, let 0 = to < t; < t; < with limnTtn = 1. Where W = (W1, Wz)‘, define for k 2 1 sigma algebra G. = a(wza) — W2(tk-1); tk_1 S t < tk). Take for k 2 1, AR from Gk satisfying both P(Ak) < 1 and P(flkAk) = 1/2. Define a deterministic process f(o) satisfying af= 1 a.s. by f(t) = ((1 — t)", O) ', O S t < 1, and a stopping time t by ‘L' = inf{t > t]: Zf(t1; t) = 1/2}. Then define 0(a) by 66 (0,0)' for tStl (5.1) 9(1): [hAjfl{th}]f(z) for tk 0 a.s. To see that E(Z(1)) < 1, consider separately integrals over the cells of the partition Q: (flAk)U(UBk), where Bk: A§\(UA°.) for k21. k_>_1 k21 j tk on UJ-(kBj. (5.5.b) E(Z(l); B1) = P( A: ){1 — E(Zf(t1; tk A t); (ma/3.93} = P(Afi ){1 — EjakE(Z(l); 31)} = P(Afi ){1 - 2i t, and 68 consider the Gk- measurable random variable Y = Afi / P( Afi ), which satisfies E(Y) = 1. There exists an adapted Rl-valued process (1)0 satisfying 01‘p0 = 1 a.s. such that if the Rz-valued process (p is defined by (1)(t) = (O, (po(t))', then the processes E[YI 17.2] and Z“’(-) are indistinguishable. (p e N, since the definition of 6 results in that N = {adapted v = (v1, v2)‘ : onv =1 a.s., A®P{v1(t,m) = 0}=1}. To prove Z(t) = E[Z(1)Z“’(t; 1) I F1], show that Z(-)Z‘p(t; .) is a (P) martingale. Since Y is Gk-measurable with pH > t, and therefore independent of F3 , X®P{s St implies (p(s,(o) = O} = 1. So Z‘p(t; .) and Z‘p(-) are indistinguishable. In particular Z‘p(t; 1) = Y a.s. For any v e N, Z(o)Z"(-) is the (P) supermartingale 29%). Then that 2(.)z¢(t; .) is a (P) martingale follows from that (5.6) E(Z(I)Z‘° 0 a.s. Let d = 1, and a(t) = {t >_ 1/2}, 0 S t S 1. Let deterministic f(t) = (l—t)"l, for O S t < 1, and define a stopping time t by t = inf{t > in; z‘(1/2, t) = 1/2}. Choose A from Fla with P(A) = 1/2; then define 6 as follows: 0 (t) '= f(t)(A n {1/2 < t st }) Take b = 09 = 9. This model is valid. 9’ is adapted, because A e Fm. 9 e Ran[o"] for all (t,00). Moreover, P{‘C S l} = 1, so that 0' and b satisfy (1.4), the market parameter integrability constraint. P {t < 1} = 1 also implies that Z(l) = Ac + 1/2 A a.s. That Z(-) and V(-) arenot indistinguishable follows from F1 )2 measurability of Z(l ). We have V(1/2) = 2(1) a.s., and Z(1)< Z(1/2) a.s. on A. To see that V(O) = 1, consider the nonnegative {Pd-martingale Y(o), with Y(t) = E[2A°I Ft]. Since Y(O) = E(2A°) = 1, 7O Y(-) = Z”(-) for an adapted process (1) satisfying d” = 1 a.s. (Take (p(t) = O for t > Q‘p(0) for each 0).) Because Ac 6 F 1/2, A®P{tp(t,0)) = 0 for all t > 1/2} = l, i.e., 7t®P{(p(t,00) e Ker[o(t,0))]} = 1. So (1) e N. Then by Lemma 5, V(0) 2 E(Z(1)Z“’(1)) = E(2Ac) =1. Example 3. This example shows that we do not have in general that V(-) is a continuous process. Here P{V(1/2) at lim ma V(t)} > 0. Let d = 1, and let 0 = to <11 < with limn tn = 1/2. Define Gk = c(W(t) - W(tk-1),tk_1 < t S t.) for each k 2 1. Choose a sequence of sets Ak, such that Ag 6 Gk and P(Ak) < 1 for each k, and P(flkAk) = 1/2. Define A = flkAk 6 F1 /2, and let 6(0), 9(-) and b(-) be as in Example 2, using this particular A when defining 6(a). Fix 0 S t < 1/2. Take k so that tk_1 > t and define a martingale process Y(o) by Y(t) = P( Afi )‘1E[Afi I Ft]. As argued for similar processes in earlier examples, Y(-) = Z‘p(o) for an adapted process (1) satisfying 06" = 1 a.s. and such that (p(t) = O for t > §°(O) holds for each 0). Ak e Gk c Fm. So 7L®P{s > 1/2 implies (p(s,o)) = 0} = l; equivalently, (p e N. Ag 6 Gk also implies through independence of G. and Flt—1 that A®P{s < tk-1 implies (p(s,00) = O}= 1. Then t < lg-) implies Z‘p(1) = Z“’(t; l) a.s. Now, using in the second equality below that Afi ; AC , where Z(l) = 1 a.s. on Ac, and using independence in the final equality, 71 (5.8) V(t) 2 E[Z(1)Z‘°(t;1)lFt] = 1512(1))111)! F11=P(A:). Since I < 1/2 implies Z(t) = l as, and V(t) S Z(t) a.s., we have V(t) = 1 a.s. for each t < 1/2. So lirn ma V(t) = l a.s. From Example 2, we know that V(1/2) = 2(1) = Ac +1/2 A a.s. Then P{V(1/2) at limm/z V(t)} = 1/2. Example 4. This example is the central object of Delbaen and Schachermayer (1998b). The Z process of this example satisfies EZ(1) < 1 and Z( 1) > 0 a.s. Furthermore, there exists a v e D such that Z(-)Z"(-) is a martingale. Given such Z and v,foreachOStSl E[Z(1)|E]= 53519121020115] =20). so that the processes V(o), Z(o), and E"[Z(1) I F.] are mutually indistinguishable. Because Z(1)> 0 a.s., Corollary 2 implies absence of arbitrage. One can construct a similar example in a market with d = 1, but the expression below using (1 = 2 is nice in terms of readability. Define 1 0 0(1) a [0 0} so that for all (t,(D), Ker[o] = {x 5 R2: x1 = 0}. Define the deterministic processes f and gfortimesOSt0:Zf(t)=l/2}, B=inf{t>0:Zg(t)=2}. Since Zf(1) = O a.s., we have t < 1 a.s. Define the process Z by stopping Zf: put Z(t) = Zf(‘t A 11 At). So defining Z is equivalent to defining process 9 by 9(t) = [(1/(1—t)){ts t A B}, 01‘. Observe that 9 satisfies 60,00) 6 Ran[o’(t,o))] for all (L(t)). If b = 66, then b = 9, and the market parameters integrability constraint (1.4) holds because 1' < 1 as: 1 1 #201,120) + Z|br(t)!}dt =1 + I{t$1/\B}Il—tdt. 0 LI 1' 0 — Check first that that E(Z(1)) < 1: (5.9) E(Zf(t/\B)) = 12f (t)dP + jzf(rA13)dP. {B=1} {13<1} P{B = l} = 1/2 because bounded local martingale Zg(B A t) is a martingale with Zg(B) = 2 on {1} < 1} and Zg(B) = 0 a.s. on {13 = 1}. Therefore, the first summand in (5.9) is 1/2 P{B = 1} = 1/4. To calculate the integral over {B < 1}, use independence of B and the process Zf(- At) in the first equality below, and in the second, that for any t e [0, 1), {Zf(S/\T), F5; OS 5 S t} is a martingale by boundedness of f(-){- S t}. 73 (5.10) jz/(ti At)dP = [E(Z/(tAt))P{13edz} = P{B<1}=1/2. {13d} 10.!) So E(Z(1))=1/4 +1/2 = 3/4. Now consider the adapted process v given by V(t) = g(t){t S 1: A B} From I < 1 a.s., it 1 holds that [va (OHZdt < 00 a.s. We have v(t,0)) e Ker[o(t,0))] for all (t,0)). Furthermore, 0 ZV(-) is the martingale zg(r A 13 A .). So v e D. Finish the example by showing E(Z(1)Z"(1)) = 1. We use the martingale property of the process Zg(B A .) in the second equality below, and refer to (5.10) in the fourth. (5.11) E(Z(1)2"<1))= E(Z‘tt A 1)) 2% A 1))) = E(Zf(t A 1)) Zg(B)) =2 jzf(tA13)dP = 1. {15d} 74 BIBLIOGRAPHY 75 Bibliography [1] Back, K., Pliska, S. (1991) On the fimdamental theorem of asset pricing with an infinite state space, J. Math. Economics, 20, 1—18. [2] Cvitanic, J ., Karatzas, I. (1992) Convex duality in convex portfolio optimization, Ann. Appl. Probability, 2, 767-818. [3] Cvitanic, J ., Karatzas, I. (1993) Hedging contingent claims with constrained portfolios, Ann. Appl. Probability, 3, 652-681. [4] Dalang, R.C., Morton, A., Willinger W. (1990) Equivalent martingale measures and no-arbitrage in stochastic security market models, Stochastics, 29, 185—201 . [5] Delbaen, F. (1992) Representing martingale measures when asset prices are continuous and bounded, Mathematical Finance, 2, 107-130. [6] Delbaen, F., Schachermayer, W. (1994a) Arbitrage and free-lunch with bounded risk, for unbounded continuous processes, Mathematical Finance, 4, 343-348. [7] Delbaen, F., Schachermayer, W. (1994b) A general version of the fimdamental theorem of arbitrage pricing, Mathematische Annalen, 300, 463—520. [8] Delbaen, F ., Schachermayer, W. (1995) The existence of absolutely continuous local martingale measures, Ann. Appl. Probability, 5, 926—945. [9] Delbaen, F., Schachermayer, W. (1998a) T he fimdamental theorem of asset pricing for unbounded stochastic processes, Mathematische Annalen, 312, 215—250. [10] Delbaen, F., Schachermayer, W. (1998b) A simple counterexample to several problems in the theory of asset pricing, Mathematical Finance, 8, 1—1 1. [11] Dudley, R.M., (1977) Wiener fimctionals as Ito integrals, Ann. Probability, 5, 140—141. [12] Duffie, D., Huang, CF. (1986) Multi-period security markets with dzjfirential information, J. Math. Econom., 15, 283——303. [13] Dybvig, P.H., Huang CF. (1988) Nonnegative wealth, absence of arbitrage, and feasible consumption plans, Rev. Financial Studies, 1, 377—401. [14] El Karoui, N., Quenez, MC. (1995) Dynamic programming and pricing of contingent claims in an incomplete market, SIAM J. Control and Optimization, 33, 29-66. 76 [15] Elliott, R.J. (1982) Stochastic Calculus and Applications, Springer-Verlag, New York. [16] Fritelli, M., Lakner, P. (1995) Arbitrage and flee-lunch in a general financial market model: the fundamental theorem of asset pricing, In IMA Vol. 65: Mathematical Finance 89-92, M. Davis, D. Duffie, W. Fleming and S. Shreve eds., Springer-Verlag, New York. [17] Harrison, J .M, Kreps, D.M. (1979) Martingales and arbitrage in multiperiod security markets, J. Econom. Theory, 20, 381—408. [18] Harrison, J .M., Pliska, SR. (1981) Martingales and stochastic integrals in the theory of continuous trading, Stochastic Processes and Appl., 11, 215—260. [19] Kabanov, Yu., Kramkov, DO. (1994) No arbitrage and equivalent martingale measures: an equivalent proof of the Harrison-Pliska theorem, Theory Probab. Appl., 39, 523—527. [20] Karatzas, I. (1996) Lectures on the Mathematics of Finance, CRM Monographs 8, American Mathematical Society. [21] Karatzas, I., Shreve SE. (1991) Brownian Motion and Stochastic Calculus, Second Edition, Springer-Verlag, New York. [22] Karatzas, I., Shreve, SE. (1998) Methods of Mathematical Finance, Springer- Verlag, New York. [23] KOpp, RE. (1984) Martingales and Stochastic Integrals, Cambridge University Press, Cambridge. [24] Kreps, D. (1981) Arbitrage and equilibrium in economies with infinitely-many commodities, J. Math. Economics, 8, 15-35. [25] Levental S., Skorohod A.V. (1995) A necessary and sufiicient condition for absence of arbitrage with tame portfolios, Ann. Appl. Probability 5, 906-925. [26] Lipster, R.S., Shiryayev, A.N. (1977) Statistics of Random Processes 1: general theory, Springer-Verlag, New York. [27] Neveu, J. (1975) Discrete-Parameter Martingales, English translation, North- Holland, Amsterdam and American Elsevier, New York. [28] Revuz, D., Yor, M (1991) Continuous Martingales and Brownian Motion, Springer, Berlin. 77 [29] Rogers, L.C.G. (1995) Equivalent martingale measures and no arbitrage, Stochastics, 51, 41-50. [30] Schachermayer, W. (1994) Martingale measures for discrete-time processes with infinite horizon, Mathematical Finance, 4, 25—56. [31] Stricker, C. (1990) Arbitrage et lois de martingale, Ann. Inst Henri Poincare, 26, 451—460. [32] Taqqu M., Willinger W. (1987) The analysis of finite security markets using martingales, Adv. Appl. Probab., 19, 1-25. 78