13.11.2014 Views

T - Prism - Université Paris 1 Panthéon-Sorbonne

T - Prism - Université Paris 1 Panthéon-Sorbonne

T - Prism - Université Paris 1 Panthéon-Sorbonne

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Cahiers de Recherche PRISM-<strong>Sorbonne</strong><br />

Pôle de Recherche Interdisciplinaire en Sciences du Management<br />

The Traditional Hedging Model Revisited<br />

with a Non Observable Convenience Yield<br />

Constantin MELLIOS<br />

Professeur, Université <strong>Paris</strong> 1 Panthéon-<strong>Sorbonne</strong> PRISM<br />

Pierre SIX<br />

Rouen Business School<br />

CR-10-34<br />

PRISM-<strong>Sorbonne</strong><br />

Pôle de Recherche Interdisciplinaire en Sciences du Management<br />

UFR de Gestion et Economie d’Entreprise – Université <strong>Paris</strong> 1 Panthéon-<strong>Sorbonne</strong><br />

17, rue de la <strong>Sorbonne</strong> - 75231 <strong>Paris</strong> Cedex 05 http://prism.univ-paris1.fr/


Cahiers de Recherche PRISM-<strong>Sorbonne</strong> 10-34<br />

The Traditional Hedging Model Revisited with a Non<br />

Observable Convenience Yield<br />

Constantin Mellios a , Pierre Six b, ∗<br />

___________________________________________________________________<br />

Abstract :This article addresses the issue of hedging a constrained position in the spot<br />

(storable) commodity with futures contracts when, in particular, the convenience yield is not<br />

observable and is estimated by using the continuous-time Kalman-Bucy method. We extend<br />

the relevant literature when the investors operate under incomplete information and study its<br />

impact on optimal demands. The latter depend crucially on the investor’s initial beliefs. The<br />

speculative and mean-variance positions in the futures contract are the unique positions<br />

capturing the effect of the incomplete information and are strongly affected by the initial<br />

value of the estimation error. We achieve a decomposition allowing investors to asses the<br />

impact of both the state variables and the initial estimation error on optimal demands.<br />

Finally, a higher initial value of the estimation error exacerbates the Samuelson effect.<br />

JEL Classification: G11; G12; G13<br />

Keywords: Incomplete Information; Optimal Dynamic demand; Convenience Yield; Commodity<br />

Futures Prices; Market Prices Of Risk; Interest Rates.<br />

___________________________________________________________________<br />

__________________________<br />

a<br />

University of <strong>Paris</strong> 1 Panthéon-<strong>Sorbonne</strong>, PRISM, 17, place de la <strong>Sorbonne</strong>, 75231 <strong>Paris</strong> Cedex 05, France.<br />

Tel: + 33(0)140462807; fax: + 33(0)40463366. E-mail address: constantin.mellios@univ-paris1.fr<br />

b<br />

Rouen Busines School, Economics and Finance department, 1, rue du Maréchal Juin, 76825 Mont Saint Aignan Cedex.<br />

Tel.: +33(0)232821718. E-mail: pierre.six@rouenbs.fr.<br />

This paper received the Outstanding Derivatives Paper Award, Eastern Finance Association, Miami, 2010.<br />

0


1. Introduction<br />

In recent years, storable commodity prices’ sharp rises have been followed by abrupt decreases.<br />

Consequently, investors seek ways to hedge the risk associated with the fluctuations of these prices.<br />

Commodity futures contracts are well-adapted instruments to hedge this risk the more so as there exist<br />

numerous liquid contracts written on a wide range of commodities. This topic is therefore of great<br />

interest for academics as well as for practitioners. This paper presents a continuous-time model to<br />

determine optimal demands for risky assets by a constrained investor 1 with a fixed position in the spot<br />

commodity, who uses futures contracts as hedging instruments. It allows one to revisit the Traditional<br />

Hedging Model (THM hereafter) (Adler and Detemple, 1988b) by studying the specific case of<br />

commodities characterized by an unobservable convenience yield.<br />

Storable commodities (agricultural, metals, energy etc.) differ from other conventional assets<br />

as they are produced and consumed. In particular, holding commodity inventories provide some<br />

services by avoiding interruptions (in) and guarantying the continuity of the production process. The<br />

net flow of these services, when expressed as a rate, is called the convenience yield defined by<br />

Brennan (1991) as “the flow of services accruing to the owner of the physical inventory, but not to the<br />

owner of a contract of a future delivery”. A parallel can be drawn between the convenience yield and<br />

the dividend on a stock. However, the convenience yield is an abstract concept, non-observed in the<br />

market. For pricing and hedging purposes, therefore, the non-observability of the convenience yield,<br />

as is widely recognized in the literature (see, for instance, Schwartz, 1997), adds an extra difficulty.<br />

One way to circumvent this difficulty is the cost-of-carry model. It is at the heart of the theory of<br />

storage (Kaldor, 1939; Working, 1948; Brennan, 1958) and established a no-arbitrage relation between<br />

futures prices, spot commodity prices, the interest rate and the net convenience yield. Gibson and<br />

Schwartz (1990) suggested to compute the implied convenience yield by using the cost-of-carry<br />

formula. Unfortunately, Carmona and Ludkovski (2004 a) showed that not only the implied<br />

convenience yield is not consistent with the forward curve, but also very different values of the<br />

1 A constrained investor cannot freely trade on the underlying spot asset.<br />

1


convenience yield are obtained for different futures contracts maturities. To remedy this flaw, our<br />

objective in this paper is to take directly into account the unobservable character of the convenience<br />

yield.<br />

Many models examined the dynamic asset allocation with futures contracts (see, among<br />

others, Ho, 1984; Stulz, 1984; Adler and Detemple, 1988a, b; Duffie and Jackson, 1990; Briys et al.,<br />

1990; Duffie and Richardson, 1991; Duffie and Stanton, 1992; Lioui and Poncet, 2001). The investor’s<br />

optimal futures demand consists of three terms (see Merton, 1973; Breeden, 1979): a mean-variance<br />

speculative term, a pure hedge minimum-variance element related to the non-traded position and à la<br />

Merton-Breeden hedging components reflecting how to hedge against a stochastic opportunity set. The<br />

number of the Merton-Breeden hedging terms is equal to the number of the unspecified state variables<br />

included in the investment opportunity set.<br />

Although these models have rigorously tackled optimal demands with futures contracts, they<br />

suffer, with regard to commodities, from two drawbacks. First, the random behavior of the<br />

convenience yield, resulting in a stochastic investment set, is not modelled 2 . Yet for non-myopic<br />

investors a Merton-Breeden term is neglected. Moreover, empirical studies (see, for instance, Fama<br />

and French, 1987, 1988; Besembinder and Chan, 1992; Khan et al., 2007; Hong and Yogo, 2009) have<br />

shown that the convenience yield is a crucial variable in predicting commodities returns. Second, the<br />

models above assume that the variables describing the investment opportunity set are observable.<br />

However, the convenience yield is not observable and investors operate under incomplete information.<br />

Dothan and Feldman (1986), Detemple (1986), Gennotte (1986), Feldman (1989), Xia (2001) and<br />

Lundofte (2006) among others, investigated, in a dynamic framework, the optimal asset allocation in a<br />

partially observable economy. To the best of our knowledge, the case of hedging with futures<br />

contracts with an unobservable convenience yield has not been examined in the relevant literature yet 3 .<br />

2 A few exceptions explicitly modeling the random behavior of the convenience yield, but in a fully observable economy and<br />

in different contexts, are Hong (2001) and Bertus et al. (2009).<br />

3 Carmona and Ludkovski (2004 b) address the pricing of commodity derivatives with a partially observable convenience<br />

yield.<br />

2


Following Schwartz (1997), model 3, and Casassus and Collin-Dufresne (2005), three<br />

imperfectly correlated state variables - the spot commodity price, the instantaneous riskless rate and<br />

the convenience yield - are supposed to explain the dynamics of the futures price. In addition,<br />

according to the empirical evidence on predictability, risk-premiums in commodities are time-varying<br />

(see, for example, Liu, 2007). Market prices of risk depend on the state variables and are hence<br />

stochastic. As the convenience yield is not observable, investors estimate it by observing the spot price<br />

and the short-rate and by using the continuous-time Kalman-Bucy method. The three state variables<br />

follow a (conditionally) Gaussian distribution and the investor’s initial beliefs about the true value of<br />

the convenience yield are normally distributed. The estimation procedure has two main consequences 4 .<br />

First, the (conditionally) Gaussian filter results in a fully observable Markovian market, where the<br />

convenience yield is replaced by its estimate. It follows that futures prices and optimal demands are<br />

functions of the estimate and of the estimation error or filtering error 5 . Especially, optimal demands<br />

depend on the investor’s prior beliefs. Second, since the investor cannot trade on the spot commodity,<br />

the market is dynamically incomplete. However, in the informationally equivalent economy, as the<br />

space of the sources of uncertainty is reduced to two, the investor faces a dynamically complete<br />

market.<br />

To derive optimal demands for risky assets, an investor maximizes the expected constant<br />

relative risk aversion (CRRA) utility function of his (her) final wealth by following the no-arbitrage<br />

martingale approach (Karatzas, Lehoczky and Shreve, 1987; Cox and Huang, 1989). Inspired, for<br />

instance, by Rodriguez (2002), Stoikov and Zariphopoulou (2005), Björk et al. (2008) and Detemple<br />

and Rindisbacher (2009), we suggest an appropriate change of a martingale measure specific to the<br />

CRRA utility function - the CRRA-forward probability measure 6 . Under this measure the Merton-<br />

4 Interest readers should refer to Feldman (2007) who provides an in-depth discussion on incomplete information and on<br />

these two points especially.<br />

5 In the (conditionally) model framework, the conditional distribution of the unobserved variable given the information<br />

provided by the observations is determined by a sufficient statistic: the conditional mean. The estimation error is<br />

deterministic (Liptser and Shiryaev, 2001a,b; chapters 11,12).<br />

6 In contrast to these papers, we operate a change of a martingale measure under which asset prices are martingales and not a<br />

change of an equivalent measure.<br />

3


Breeden terms reduce to two terms. The first, as in Lioui and Poncet (2001), Munk and<br />

Sorensen (2004) and Detemple and Rindisbacher (2009), is associated with the risk of the short-rate<br />

and hedges the changes in a discount bond with a maturity that matches the investor’s horizon. The<br />

second addend, results from the stochastic prices of risk, and is couched in terms of an investor<br />

specific discount bond as it depends on the investor’s risk aversion, horizon and prior beliefs 7 . This<br />

decomposition allows us to give an economic interpretation to our results and to suggest new useful<br />

expressions of optimal demands.<br />

By isolating the orthogonal risk of the spot commodity, i.e. the part of its risk that is not<br />

correlated with the risk of the short-rate, we are able both to derive the speculative and minimumvariance<br />

proportions for each asset and to study the impact of the incomplete information. Our results<br />

show that these components related to the futures contract capture the effect of the estimation error.<br />

Those of the discount bond are affected only through the components of the futures contract. Contrary<br />

to the traditional THM, the pure hedge term is investor dependent as it involves the investor’s initial<br />

beliefs. The orthogonal risk presents another advantage. It leads to a decomposition of the investor<br />

specific bond for each and every state variable. An investor may then accurately measure the impact of<br />

the state variables on optimal demands. This impact is assessed through the sensitivity of the investor<br />

specific bond on the state variables, which depends on the initial estimation error. As a consequence,<br />

all these hedging elements are influenced by this error.<br />

An illustration applied to the copper market shows that the initial estimation error has a much<br />

stronger effect on the mean-variance and the minimum-variance demands than on the investor specific<br />

bond demand. Moreover, the futures price volatility exhibits the Samuelson effect (this volatility<br />

increases as the futures contract approaches its maturity date), which is amplified for a higher initial<br />

value of the estimation error.<br />

The remainder of the paper is organized as follows. Section 2 describes the economic<br />

framework. Section 3 analyzes optimal demands. Section 4 provides an illustration of the results of<br />

7 In the Detemple and Rindisbacher (2009) paper, this component is expressed in a less intuitive way as a function of the<br />

density of the forward probability measure.<br />

4


section 3 in the case of the copper market. Section 5 concludes and offers possible extensions. All<br />

proofs are available from the authors upon request.<br />

2. The economic framework<br />

Consider a continuous-time frictionless economy. Uncertainty in the economy is described by a<br />

complete filtered probability space ( , F , Θ,P)<br />

filtration Θ ≡ { F t<br />

: t ∈[ 0,<br />

T ]}<br />

, where [ ,T ]<br />

Ω endowed with a (right) continuous non-decreasing<br />

0 is a fixed time interval, T is the finite horizon of the<br />

economy, and<br />

F T<br />

≡ F . P is the historical probability measure. All (continuous-time) processes<br />

described below are assumed to be adapted. ( ) ′<br />

* * *<br />

correlated Brownian motions defined on ( , F , Θ,P)<br />

* *<br />

ρ ≡ dz dz dt , i ≠ j<br />

ij<br />

it<br />

motions.<br />

jt<br />

with , { S,r,δ}<br />

zSt<br />

, zrt<br />

, zδ t<br />

, stands for a 3-dimensional vector of<br />

Ω where ' denotes the transpose and<br />

i ∈ is the constant correlation coefficient of the Brownian<br />

In our model, the state variables that describe the economy are in the spirit of those of<br />

Schwartz (1997) and Casassus and Collin-Dufresne (2005). There are three imperfectly correlated<br />

variables: the spot commodity price,<br />

yield, δ<br />

t<br />

. They are governed by the following stochastic processes:<br />

S<br />

t<br />

, the instantaneous riskless rate, r<br />

t<br />

, and the (net) convenience<br />

dSt<br />

S<br />

dr<br />

t<br />

t<br />

*<br />

[ r + σ λ ] dt + σ dz X ≡ ln( S ),<br />

S ≡ > 0<br />

+ δ dt =<br />

S<br />

(1a)<br />

t<br />

t<br />

S<br />

[ ( t)<br />

− αr<br />

] dt + σ dz<br />

*<br />

, r ≡ r<br />

=<br />

r<br />

t<br />

r rt 0<br />

St<br />

S<br />

St, t<br />

t 0<br />

θ (1b)<br />

[ θ −δ<br />

] dt + σ dz<br />

*<br />

, δ ≡ δ<br />

d δ<br />

t<br />

= κ<br />

δ t<br />

δ δt<br />

0<br />

(1c)<br />

σ i<br />

and λ i<br />

(.) represent the constant strictly positive volatility of the state variables and the market price<br />

of risk associated with the state variables, respectively.<br />

λ =<br />

0<br />

+ + , λ<br />

rt<br />

= λr<br />

0<br />

+ λrrrt<br />

and<br />

St<br />

λS<br />

λSX<br />

X<br />

t<br />

λS<br />

δδt<br />

λδ t<br />

= λδ<br />

0<br />

+ λδδδ<br />

( t)<br />

. λ<br />

S 0, λSX<br />

, λSδ<br />

, λr<br />

0,<br />

λrr<br />

, λδ<br />

0<br />

and λδδ<br />

are constants. The convenience yield and the spot<br />

price are related through inventory decisions (Routledge, Seppi, and Spatt 2000). To take into account<br />

this relation, (see, for instance, Brennan, 1958; Dincerler et al. 2005), the price of risk associated with<br />

the (log) of the spot price process is an affine function of the level of both the (log) of the spot price<br />

5


and the convenience yield. The prices of risk associated with the short-rate and the convenience yield<br />

are affine functions of their own level.<br />

The short rate follows a process such that the model incorporates the initial term structure of<br />

interest rates (see Hull and White, 1990; Heath et al., 1992). In particular, its drift depends on a<br />

2<br />

deterministic function θ t)<br />

{ αf<br />

( t)<br />

+ ∂ f ( t)<br />

+ σ D<br />

α<br />

( t + σ λ }<br />

describes the initial forward yield curve,<br />

with respect to x and ( )<br />

r<br />

(<br />

0 t 0 r 2<br />

)<br />

r r 0<br />

2<br />

y<br />

= , where α is a constant, f ( )<br />

∂<br />

x<br />

stands for the partial derivative (column gradient vector)<br />

D u<br />

denotes the deterministic function D u<br />

( y)<br />

( 1−<br />

exp( 2uy)<br />

)<br />

0 t<br />

1<br />

= for two<br />

2u<br />

2<br />

−<br />

scalars u and y. Empirical studies (see Fama and French, 1988 ; Brennan, 1991) reported that the<br />

convenience yield should be specified by a mean-reverting process. It follows then a mean-reverting<br />

process with a constant long term mean, θ<br />

δ<br />

and a constant speed of mean reversion k.<br />

The investor does not observe the convenience yield but draws inferences about it from his (her)<br />

observations of the spot price of the commodity and the interest rate. It is assumed that (s)he views the<br />

prior distribution of δ as Gaussian with given mean<br />

m( 0) ≡ m and variance e( 0) ≡ e . The agent<br />

estimates the unobserved stochastic process of the convenience yield {δ(t): t ∈ [0,T]} given noisy<br />

observations of the joint processes {S(t), r(t): t ∈ [0,T]}. The agent utilizes the information available,<br />

F<br />

S , r<br />

t<br />

( S , r , u ≤ t)<br />

= σ 0 ≤<br />

u<br />

u<br />

, to derive the optimal estimate of the state variable, δ(t). The conditional<br />

, r<br />

mean is defined as follows: m(<br />

t)<br />

E[ δ ( t)<br />

F ( t)<br />

] S<br />

≡ E [ δ ( t)<br />

]<br />

error) is given by ( e,<br />

t)<br />

= E ( δ ( t)<br />

− m(<br />

t)<br />

)<br />

= , and the estimation error (the mean square<br />

2 , r<br />

[ F ( t)<br />

]<br />

S<br />

t<br />

ε . m<br />

t<br />

and ε ( e,<br />

t)<br />

obey the following system of<br />

differential equations (Liptser and Shiryayev, 2001 a, b, chapters 11-12) – detailed computation is<br />

available from the authors upon request:<br />

dm<br />

t<br />

⎡ κ ⎤ ⎡<br />

⎤<br />

Sδ<br />

κ<br />

Sδ<br />

= κ[ θ − mt<br />

] dt + ⎢σ<br />

ρS<br />

0<br />

+ ρS<br />

ε ( e,<br />

t)<br />

⎥dzSt<br />

+ ⎢σ<br />

ρr<br />

0<br />

+ ρr<br />

ε ( e,<br />

t)<br />

⎥dz<br />

(2b)<br />

δ δ<br />

ε<br />

δ<br />

ε<br />

rt<br />

⎣ σ<br />

S ⎦ ⎣ σ<br />

S ⎦<br />

ε<br />

∞<br />

− ε<br />

2<br />

ε ( e,<br />

t)<br />

= ε<br />

∞<br />

−<br />

(2c)<br />

e −ε<br />

2<br />

1−<br />

exp( ∆t)<br />

e −ε<br />

∞<br />

6


where κ ≡ σ λ −1<br />

∆ ≡ b 2 − 4ac ,<br />

ρ − ρ ρ ρrδ<br />

− ρSδ<br />

ρSr<br />

1<br />

ρSr<br />

ρ ≡ , ρr<br />

0<br />

≡ , ρ<br />

2<br />

S ε<br />

≡ , ρ<br />

2 r ε<br />

≡ − and<br />

2<br />

− ρ<br />

1− ρ<br />

− ρ<br />

Sδ<br />

rδ<br />

Sr<br />

S δ S Sδ<br />

,<br />

S 0<br />

2<br />

1−<br />

ρSr<br />

− b + ∆<br />

ε<br />

2<br />

≡ ,<br />

2a<br />

2 2<br />

[ − ρ − ρ − 2ρ<br />

ρ ρ ]<br />

2<br />

σ δ<br />

1<br />

S 0 r 0 Sr r 0 S 0<br />

− b − ∆<br />

ε<br />

∞<br />

≡<br />

with<br />

2a<br />

1<br />

Sr<br />

c ≡ . The innovation processes, dz<br />

St<br />

and<br />

2<br />

Sr<br />

1<br />

Sr<br />

⎡κ ⎤<br />

S<br />

a ≡ − δ<br />

⎡ σ ⎤<br />

⎢ ⎥ ,<br />

δ<br />

b ≡ −2 ⎢κ<br />

+ κ<br />

Sδ ρS<br />

0 ⎥ and<br />

⎣ σ<br />

S ⎦ ⎣ σ<br />

S ⎦<br />

dz<br />

rt<br />

, are given by<br />

*<br />

dz<br />

rt<br />

= dz rt<br />

and<br />

dz<br />

St<br />

1<br />

=<br />

σ<br />

S<br />

⎡dSt<br />

⎢<br />

⎢⎣<br />

St<br />

⎡dSt<br />

− E⎢<br />

⎢⎣<br />

St<br />

F<br />

S , r<br />

t<br />

⎤⎤<br />

⎥⎥<br />

⎥⎦<br />

⎥⎦<br />

and are observable Brownian motions (see Liptser and Shiryaev; 2001<br />

a,b ). ε ≡ ε ( ∞,<br />

e)<br />

denotes the steady state value of the estimation error which does not depend on e.<br />

∞<br />

The system of equations (1a) to (1c) can be written under the equivalent fully observable<br />

economy as follows:<br />

dS t<br />

+ mtdt<br />

=<br />

2<br />

(3a)<br />

S<br />

dr<br />

t<br />

t<br />

dm<br />

[ rt<br />

+ σ<br />

S<br />

( λS<br />

0<br />

+ λSX<br />

X<br />

t<br />

+ λSδ mt<br />

)] dt + σ<br />

S<br />

[ ρSrdzrt<br />

+ 1−<br />

ρSr<br />

dzwt<br />

]<br />

[ θr<br />

t − rt<br />

] dt σ<br />

rdzrt<br />

= β ( ) +<br />

(3b)<br />

t<br />

⎡<br />

⎤<br />

2 κ<br />

Sδ<br />

= κ[ θ − mt<br />

] dt + σ ρr<br />

dzrt<br />

+ ⎢σ<br />

1−<br />

ρSr<br />

ρS<br />

+<br />

( e,<br />

t)<br />

⎥<br />

0<br />

ε dz<br />

(3c)<br />

δ δ δ<br />

δ<br />

2<br />

wt<br />

⎢<br />

⎣<br />

σ<br />

⎥<br />

S<br />

1−<br />

ρSr<br />

⎦<br />

Liptser and Shiryaev (2001 a, b) shown that the information generated by the state variable<br />

process in the original economy, {S(t), r(t), δ(t): t ∈ [0,T]}, is equivalent to the information generated<br />

by the observable process {S(t), r(t), m(t): t ∈ [0,T]}. Thus, the investor makes his (her) financial<br />

decisions as if (s)he faces the fully observable Markovian market given by (3a-c) (see Feldman, 2007).<br />

This has an impact on the covariances between the state variables. Although the conditional<br />

covariance between the short rate and the conditional mean is not modified, that between the (log) spot<br />

price and the conditional mean is a function of the estimation error:<br />

Cov<br />

( dX ; dm ) σ σ ρ dt κ ε ( e,<br />

t dt<br />

= . This covariance has two addends. The first is equal to the<br />

t t t S δ Sδ<br />

+<br />

Sδ<br />

)<br />

covariance in the original economy. To this addend, as a consequence of the estimation procedure, a<br />

second component is added involving the estimation error. The impact of the latter on the covariance<br />

depends on the sign of κ<br />

S δ<br />

≡ σ<br />

SλSδ<br />

−1. When κ S δ<br />

> 0 , the filtering error has a positive effect and vice<br />

versa.<br />

7


It is worth pointing out that in the original economy three Brownian motions govern the<br />

dynamics of the state variables, while in the informationally equivalent economy two sources of<br />

uncertainty, the innovation processes, z<br />

St<br />

and z<br />

rt<br />

, determine the evolution of the three state variables.<br />

For reasons that will become clear later, since the innovation processes are correlated,<br />

dz<br />

St<br />

can be<br />

written in the following manner:<br />

dz<br />

St<br />

= ρ dz + 1−<br />

ρ dz , where z<br />

wt<br />

is a Brownian motion<br />

Sr<br />

rt<br />

2<br />

Sr<br />

wt<br />

uncorrelated with z<br />

rt<br />

. It follows that, as can be seen in equation (3c), the effect of the estimation error<br />

on the conditional mean is isolated and appears through z<br />

wt<br />

.<br />

In order to obtain compact formulae, the equations above can be expressed in a matrix form.<br />

Let Y [ X r m ] ′ , λ [ λ λ ] ′ and [ dz dz ] ′<br />

t<br />

=<br />

t t t<br />

t<br />

=<br />

rt wt<br />

dz be the vectors of the state variables, the<br />

t<br />

=<br />

rt wt<br />

prices of risk and the Wiener processes respectively. Then:<br />

dY<br />

t<br />

′<br />

= µ , , (4a)<br />

[ ( t)<br />

− µ<br />

YYt<br />

] dt + σ<br />

Y<br />

( t e) dzt<br />

λ = λ 0<br />

+ λ Y , (4b)<br />

t<br />

Y<br />

t<br />

where<br />

′<br />

σ =<br />

S<br />

⎡ 1 2⎤<br />

⎢σ<br />

S<br />

λs<br />

0<br />

− σ ⎥<br />

µ = ⎢ βθ ( 2 S<br />

( t)<br />

) ⎥<br />

r<br />

t<br />

⎢<br />

⎥<br />

⎢ κθ δ<br />

⎥<br />

⎢⎣<br />

⎥⎦<br />

2<br />

[ σ<br />

S<br />

ρSr<br />

σ<br />

S<br />

1−<br />

ρSr<br />

]<br />

′<br />

,<br />

⎡−<br />

σ λ −1 − κ<br />

⎡ ′<br />

⎤<br />

σ ⎤<br />

S SX<br />

Sδ<br />

S<br />

⎢ ⎥<br />

µ =<br />

⎢<br />

⎥ ′<br />

Y<br />

⎢<br />

0 β 0<br />

⎥<br />

, ( ) ⎢<br />

′<br />

σ ⎥<br />

Y<br />

t,<br />

e = σ<br />

r<br />

with<br />

⎢ ⎥<br />

⎢⎣<br />

0 0 κ ⎥<br />

′<br />

⎦<br />

⎢σ<br />

m(<br />

t,<br />

e)<br />

⎥<br />

⎣ ⎦<br />

, σ = [ 0]<br />

, σ ( e,<br />

t)<br />

′<br />

= [ σ ρ σ ( t,<br />

e)<br />

]<br />

r<br />

σ r<br />

m<br />

δ rδ<br />

mw<br />

⎡λr<br />

0 ⎤<br />

. λ<br />

0<br />

= ⎢ ⎥ and<br />

⎣λw0<br />

⎦<br />

⎡ 0<br />

λY<br />

= ⎢<br />

⎣λ<br />

wX<br />

λ<br />

λ<br />

rr<br />

wr<br />

0 ⎤<br />

λ<br />

⎥<br />

wm ⎦<br />

are such that<br />

λ<br />

wX<br />

λ<br />

ρSr<br />

≡ , λwr<br />

≡ − λ<br />

2<br />

rr<br />

,<br />

1− ρ<br />

SX<br />

2<br />

1−<br />

ρSr<br />

Sr<br />

λ<br />

wm<br />

λ<br />

≡ and<br />

Sδ<br />

2<br />

1−<br />

ρSr<br />

λ<br />

w0<br />

≡<br />

λ<br />

S 0<br />

1−<br />

ρ<br />

2<br />

Sr<br />

−<br />

ρ<br />

Sr<br />

1−<br />

ρ<br />

2<br />

Sr<br />

λ . We denote by<br />

r 0<br />

⎡<br />

⎤<br />

2 κ<br />

Sδ<br />

σ<br />

mw(<br />

t,<br />

e)<br />

≡ ⎢σ<br />

δ<br />

1−<br />

ρSr<br />

ρS<br />

0<br />

+ ε ( t,<br />

e)<br />

⎥ the<br />

2<br />

⎢⎣<br />

σ<br />

S<br />

1−<br />

ρSr<br />

⎥⎦<br />

sensitivity of the estimate of the convenience yield to<br />

dz<br />

wt<br />

.<br />

There are in the economy a locally riskless asset, the savings account, such that:<br />

⎧ ⎫<br />

= ⎨∫ t β ( t)<br />

exp r(<br />

s)<br />

ds⎬<br />

, with initial condition β ( 0) = 1, and three risky traded assets. The spot<br />

⎩ 0 ⎭<br />

commodity whose dynamics are given in equation (3a), a discount bond with maturity<br />

T<br />

B<br />

, whose<br />

8


price, at time t,<br />

0 ≤ t ≤ TB<br />

, is Bt<br />

( r,<br />

TB<br />

) ≡ Bt<br />

( TB<br />

) , and a futures contract written on a commodity with<br />

maturity<br />

T<br />

H<br />

, whose price, at date t,<br />

0 ≤ t ≤ T H<br />

≤ T , is denoted H ( Y , e,<br />

T ) ≡ H ( e,<br />

T ). The price<br />

B<br />

t<br />

t<br />

H<br />

t<br />

H<br />

dynamics of the last two securities can be written as follows:<br />

dBt<br />

( TB<br />

)<br />

= [ rt<br />

−σ<br />

B(<br />

t,<br />

TB<br />

) λrt<br />

] dt −σ<br />

B(<br />

t,<br />

TB<br />

) dzrt<br />

, (5a)<br />

B ( T )<br />

t<br />

dH<br />

H<br />

B<br />

t<br />

( e,<br />

TH<br />

)<br />

( e,<br />

T )<br />

t<br />

H<br />

[ σ<br />

Hw( t,<br />

e,<br />

TH<br />

) λwt<br />

+ σ<br />

Hr<br />

( t,<br />

e,<br />

TH<br />

) λrt<br />

] dt + σ<br />

Hr<br />

( t,<br />

e,<br />

TH<br />

) dzrt<br />

+ σ<br />

Hw( t,<br />

e,<br />

TH<br />

) dzwt<br />

= , (5b)<br />

where ( t,<br />

T ) ≡ σ D ( T − t)<br />

Hr<br />

2<br />

σ σ ( t , e,<br />

T ) σ h ( t,<br />

e,<br />

T ) 1−<br />

ρ + σ ( t,<br />

e) h ( t,<br />

e,<br />

T )<br />

B<br />

B<br />

r<br />

α<br />

B<br />

( t e,<br />

T ) σ h ( t,<br />

e,<br />

T ) ρ + σ h ( t,<br />

e,<br />

T ) σ δ<br />

ρ h ( t,<br />

e,<br />

T )<br />

σ ≡ +<br />

are such that:<br />

,<br />

H S X H Sr r r H<br />

rδ<br />

m<br />

Hw<br />

H<br />

≡ and<br />

S<br />

X<br />

H<br />

H<br />

Sr<br />

mw<br />

m<br />

H<br />

⎡∂thX<br />

⎢<br />

⎢<br />

∂thr<br />

⎢⎣<br />

∂thm<br />

( t,<br />

e,<br />

TH<br />

)<br />

( t,<br />

e,<br />

TH<br />

)<br />

( t,<br />

e,<br />

T )<br />

H<br />

⎤ ⎡ 0<br />

⎥ ⎢<br />

⎥<br />

=<br />

⎢<br />

−1<br />

⎥⎦<br />

⎢⎣<br />

1<br />

0<br />

α<br />

0<br />

( t,<br />

e,<br />

TH<br />

)<br />

( t,<br />

e,<br />

TH<br />

)<br />

( t,<br />

e,<br />

T )<br />

( TH<br />

, e,<br />

TH<br />

) ⎤ ⎡1⎤<br />

( T )<br />

⎥<br />

=<br />

⎢<br />

H<br />

, e,<br />

TH<br />

⎥ ⎢<br />

0<br />

( ) ⎥ ⎢ ⎥ ⎥⎥ TH<br />

, e,<br />

TH<br />

⎦ ⎣0⎦<br />

σ<br />

mw(<br />

t,<br />

e)<br />

λwX<br />

⎤⎡hX<br />

⎤ ⎡hX<br />

σ ( ) +<br />

⎥⎢<br />

⎥ ⎢<br />

mw<br />

t,<br />

e λwr<br />

σ ρr<br />

λrr<br />

⎥⎢<br />

hr<br />

⎥<br />

,<br />

⎢<br />

h<br />

. (5c)<br />

δ δ<br />

r<br />

κ + σ ( , ) ⎥⎦<br />

⎢⎣<br />

⎥⎦<br />

⎢<br />

mw<br />

t e λwm<br />

hm<br />

H ⎣hm<br />

Moreover, we denote by σ ( t e,<br />

T ) σ ( t,<br />

e,<br />

T ) 2 + σ ( t,<br />

e,<br />

T ) 2<br />

contract.<br />

H<br />

,<br />

H<br />

Hw H Hr H<br />

≡ the volatility of the futures<br />

We consider the case of a constrained investor who is endowed with a fixed position θ<br />

S<br />

in the<br />

spot commodity, which means that the investor cannot freely trade the spot commodity. In contrast,<br />

(s)he can freely invest in the discount bond and the futures contract, and in the risk-free asset. We<br />

denote by θ<br />

βt<br />

, θ<br />

Bt<br />

and θ<br />

Ht<br />

the units invested in the risk-free asset, the zero-coupon bond and the<br />

futures contract, respectively. Investor’s time t wealth is then given by:<br />

where<br />

t<br />

t<br />

t<br />

S<br />

t<br />

Bt<br />

t<br />

( TB<br />

) M<br />

t<br />

W = θ β<br />

β + θ S + θ B + . (6)<br />

M<br />

t<br />

is the margin account, since the investor trades futures contracts. The expression of<br />

given by (Duffie and Stanton, 1992):<br />

M<br />

t<br />

=<br />

t<br />

∫<br />

0<br />

M<br />

t<br />

is<br />

t<br />

⎧ ⎫<br />

exp⎨∫<br />

rv<br />

dv⎬θ HudHu<br />

( TH<br />

)<br />

(7)<br />

⎩ u ⎭<br />

The investor, endowed with an initial wealth W(0), has an investment horizon<br />

T I<br />

< T<br />

, a<br />

constant relative risk aversion γ > 0 and seeks to maximize his (her) CRRA utility function of<br />

terminal wealth. Under the complete information economy, there are two sources of uncertainty and,<br />

9


as a consequence of the investor’s fixed position in the spot commodity, two risky traded securities.<br />

Therefore, the constrained investor faces a dynamically complete financial market. It is worth pointing<br />

out that in the original economy, there are three sources of uncertainty and two traded assets. As a<br />

result, the market is not complete for the constrained investor. The market completeness is directly<br />

related to incomplete information. The martingale approach for complete markets, developed by<br />

Karatzas et al. (1987) and Cox and Huang (1989), can then be applied to determine the constrained<br />

investor’s optimal asset allocation by solving the following static program:<br />

1−γ<br />

⎡W<br />

⎤<br />

TI<br />

max Εt<br />

⎢ ⎥<br />

WTI<br />

⎢⎣<br />

1−<br />

γ ⎥⎦<br />

(8a)<br />

s.t.<br />

Wt<br />

G<br />

t<br />

⎡W<br />

= Et<br />

⎢<br />

⎢⎣<br />

G<br />

T<br />

T<br />

I<br />

I<br />

⎤<br />

⎥<br />

⎥⎦<br />

(8b)<br />

where [ ⋅ F ] ≡ [] .<br />

Ε denotes the expectation, under P, conditional on the information, F t , available at<br />

t<br />

E t<br />

time t and<br />

t<br />

t<br />

= ⎧<br />

= ⎨∫<br />

+ ∫ + ∫<br />

t<br />

β ( t)<br />

1 2 '<br />

G( t)<br />

exp ru<br />

ds λu<br />

λu<br />

dz<br />

ξ ( t)<br />

⎩ 2<br />

0 0<br />

0<br />

u<br />

⎫<br />

⎬, with G(0) = 1, represents the numéraire or<br />

⎭<br />

optimal growth portfolio such that the value of any admissible portfolio relative to this numéraire is a<br />

martingale under P (see Long, 1990; Merton, 1990; Bajeux-Besnainou and Portait, 1997).<br />

stands<br />

for the norm in R 2 and ξ (t)<br />

is the Radon-Nikodym derivative of the so-called, unique, risk-neutral<br />

probability measure Q equivalent to the historical probability P , such that the relative price (with<br />

respect to the savings account chosen as numéraire), of any risky security is a Q-martingale (see<br />

Harrison and Pliska, 1981).<br />

3 Optimal demands under incomplete information<br />

Within the financial market described above, the solution to the static problem (8), which is a standard<br />

Lagrangian optimization problem, allows us to derive the optimal wealth at time t:<br />

where<br />

W<br />

t<br />

* −1 γ<br />

t t t<br />

t<br />

gγ<br />

( T ) B ( T )<br />

= k G B<br />

, (9)<br />

I<br />

γt<br />

I<br />

gγ<br />

[{ { G<br />

} ]<br />

k is a Lagrange multiplier, ( T ) E G B ( T )<br />

B = is and ≡1− γ<br />

γt<br />

I<br />

t<br />

t<br />

t<br />

I<br />

TI<br />

g . ( )<br />

γ<br />

1<br />

B γ<br />

can be<br />

t<br />

T I<br />

written as follows:<br />

10


B<br />

γt<br />

gγ<br />

[{ { G<br />

} ]<br />

( T ) = E G B ( T )<br />

I<br />

t<br />

t<br />

t<br />

2<br />

⎧ gγ<br />

exp⎨−<br />

⎩ 2<br />

I<br />

TI<br />

where [ ] ′<br />

B σ<br />

t<br />

∫<br />

0<br />

= B(0,<br />

T )<br />

λ −σ<br />

( u,<br />

T )<br />

σ t,<br />

T ) ≡ ( t,<br />

T ) 0 .<br />

(<br />

I B I<br />

u<br />

B<br />

gγ<br />

I<br />

I<br />

2<br />

2<br />

⎡ ⎧ gγ<br />

Et<br />

⎢exp⎨−<br />

⎢⎣<br />

⎩ 2<br />

du − g<br />

γ<br />

t<br />

t<br />

∫<br />

λ −σ<br />

( u,<br />

T )<br />

'<br />

∫[ λu<br />

−σ<br />

B<br />

( u,<br />

TI<br />

)]<br />

0<br />

0<br />

u<br />

B<br />

I<br />

2<br />

⎫⎤<br />

dz(<br />

u)<br />

⎬⎥<br />

⎭⎥⎦<br />

⎫<br />

du⎬<br />

⎭<br />

The computation of the optimal wealth reduces to that of B γ t<br />

( T I<br />

). Its computation may be<br />

simplified by making an appropriate change of numéraire. Jamishidian (1987, 1989) was the first to<br />

use a discount bond as a numéraire and introduced the forward martingale measure equivalent to P.<br />

This measure was applied to asset allocation by Lioui and Poncet (2001), Munk and Sorensen (2004)<br />

and Detemple and Rindisbacher (2009). In the spirit of Rodriguez (2002), Stoikov and Zariphopoulou<br />

(2005) and Björk et al. (2008), we operate a change of probability measure specific to CRRA utility<br />

functions. The Radom-Nikodym derivative defines the probability measure<br />

Geman et al., 1995):<br />

( g γ )<br />

P equivalent to P (see<br />

( gγ<br />

)<br />

dP<br />

dP<br />

S , t<br />

t<br />

F<br />

[ λu<br />

−σ<br />

B(<br />

u,<br />

TI<br />

)]<br />

2 t<br />

t<br />

⎪⎧ g<br />

2<br />

γ<br />

′<br />

≡ exp⎨−<br />

∫ λu<br />

−σ<br />

B(<br />

u,<br />

TI<br />

) du − gγ<br />

∫<br />

dz<br />

⎪⎩<br />

2<br />

0<br />

0<br />

u<br />

⎪⎫<br />

⎬<br />

⎪⎭<br />

By using Baye’s rule for conditional expectations:<br />

B<br />

γt<br />

( T )<br />

I<br />

T<br />

( ) ⎡ ⎪⎧<br />

I<br />

⎪⎫<br />

⎤<br />

g g<br />

2<br />

γ γ<br />

= Et<br />

⎢exp⎨−<br />

∫ λu<br />

−σ<br />

B ( u,<br />

TI<br />

) du⎬⎥<br />

, (10)<br />

⎢⎣<br />

⎪⎩ 2γ<br />

t<br />

⎪⎭ ⎥⎦<br />

where<br />

( g<br />

E ) γ<br />

[] . stand for the expectation under<br />

t<br />

( g γ )<br />

P conditional on<br />

S r<br />

F , t<br />

.<br />

B γ t<br />

( T I<br />

) is a function of the investor’s risk aversion coefficient, horizon, and initial beliefs. As<br />

it has the functional form of a discount bond, it will be called the investor specific discount bond. Its<br />

dependence on the investor’s initial beliefs distinguishes the expression of B γ t<br />

( T I<br />

) from that in a fully<br />

observable economy. It is stochastic because of the stochastic character of the market prices of risk. It<br />

is worth pointing out that B γ t<br />

( T I<br />

) is not a traded financial asset but it can be duplicated, in our<br />

complete market, by existing traded securities. As the expectation in equation (10) involves a<br />

quadratic function, following the literature of the term structure of interest rates (see Dai and<br />

11


Singleton, 2000, 2002; Ahn et al., 2002), B γ t<br />

( T I<br />

) can be expressed as an exponential quadratic function<br />

of the state variables:<br />

⎧<br />

' 1 ' ⎫<br />

B<br />

t<br />

( TI<br />

) = exp⎨b<br />

( t,<br />

e,<br />

TI<br />

) + b ( t,<br />

e,<br />

TI<br />

) Yt<br />

+ Yt<br />

b ( t,<br />

e,<br />

TI<br />

) Yt<br />

⎬ , (11a)<br />

γ 0γ<br />

1γ<br />

2γ<br />

⎩<br />

2<br />

⎭<br />

where b γ<br />

( t,<br />

e,<br />

T ) is a deterministic function unnecessary to the hedging analysis and b γ<br />

( t,<br />

e,<br />

T ) ,<br />

0 I<br />

b γ<br />

( t,<br />

e,<br />

T ) are functions solving the following system of ordinary differential equations:<br />

2 I<br />

∂ b<br />

2γ<br />

( t,<br />

e,<br />

T ) − b ( t,<br />

e,<br />

T ) µ ( t,<br />

e) − ( t,<br />

e) b ( t,<br />

e,<br />

T ) + b ( t,<br />

e,<br />

T ) Σ ( t,<br />

e) b ( t,<br />

e,<br />

T )<br />

t 2γ I 2γ<br />

I Yγ<br />

µ<br />

Yγ<br />

− a<br />

∂ b<br />

− a<br />

1γ<br />

= 0<br />

2x2<br />

′<br />

2γ<br />

'<br />

( t,<br />

e,<br />

TI<br />

) − µ<br />

Yγ<br />

( t,<br />

e) b1<br />

γ<br />

( t,<br />

e,<br />

TI<br />

) + b2γ<br />

( t,<br />

e,<br />

TI<br />

)<br />

γ<br />

( t,<br />

e,<br />

TI<br />

) + b2γ<br />

( t,<br />

e,<br />

TI<br />

) ΣY<br />

( t,<br />

e) b1<br />

γ<br />

( t,<br />

e,<br />

TI<br />

)<br />

( t,<br />

TI<br />

) = 02<br />

t 1γ µ<br />

with the terminal conditions b<br />

2γ ( T e,<br />

) = 0 2 2<br />

and<br />

1<br />

( T , e,<br />

) = 02<br />

I<br />

, T I x<br />

I<br />

2γ<br />

I<br />

T I<br />

I<br />

Y<br />

2γ<br />

I<br />

1 I<br />

, (11b)<br />

, (11c)<br />

b γ<br />

. 02<br />

x 2<br />

and 02<br />

are a 2-dimensional<br />

matrix and a 2-dimensional vector of zeroes respectively. ( t, e) ≡ σ ( t,<br />

e) σ ( t e)<br />

Y Y<br />

Y<br />

,<br />

′<br />

Σ is the variancecovariance<br />

matrix of<br />

Y , ( t e,<br />

T ) µ ( t)<br />

− g σ ( t,<br />

e)<br />

′[ λ − σ ( T )]<br />

t<br />

'<br />

µ<br />

γ<br />

,<br />

I<br />

≡<br />

γ Y<br />

0 Bt I<br />

and µ<br />

Yγ ( t,<br />

e)<br />

≡ µ<br />

Y<br />

+ gγσ<br />

Y<br />

( e,<br />

t)<br />

λY<br />

are the two terms of the expected changes in<br />

Y<br />

t<br />

under<br />

( g γ )<br />

gγ<br />

′<br />

P . a t,<br />

T ) ≡ λ [ λ −σ<br />

( t,<br />

T )]<br />

and<br />

1γ<br />

(<br />

I<br />

Y 0 B I<br />

γ<br />

a<br />

g<br />

γ<br />

2 γ<br />

≡ λY<br />

λY<br />

. Applying Itô’s lemma to expression (6) and by the self-financing property, we obtain<br />

γ<br />

′<br />

the dynamics of the investor’s wealth 8 :<br />

dW t<br />

rt<br />

t<br />

(<br />

St S t<br />

( e TH<br />

TB<br />

)<br />

t<br />

) ⎤dt<br />

+ (<br />

St S<br />

+<br />

t<br />

( e TH<br />

TB<br />

)<br />

t<br />

) ′ dzt<br />

W ⎢⎣<br />

⎡ ′<br />

= + λ π σ + σ , , π π σ σ , , π , (12)<br />

⎥⎦<br />

t<br />

⎡σ<br />

B( t,<br />

TB<br />

) σ<br />

Hr<br />

( t,<br />

e,<br />

TH<br />

)<br />

with the initial condition W(0), ( )<br />

( ) ⎥ ⎤<br />

σ<br />

t<br />

e,<br />

TH<br />

, TB<br />

= ⎢<br />

and π<br />

t<br />

≡ [ π<br />

Bt<br />

π<br />

Ht<br />

]<br />

⎣ 0 σ<br />

t<br />

te,<br />

TH<br />

⎦<br />

' .<br />

S<br />

W<br />

π ≡ B T W and<br />

π<br />

St<br />

≡ θS<br />

t t<br />

is the fixed proportion invested in the spot commodity,<br />

Bt<br />

θBt<br />

t<br />

(<br />

B<br />

)<br />

t<br />

Ht<br />

Ht<br />

t<br />

( e TH<br />

) Wt<br />

π ≡ θ H , represent the proportions invested in the two traded risky securities respectively.<br />

In what follows, we shall distinguish the speculative or mean-variance proportion,<br />

MV<br />

π<br />

t<br />

, the minimum-<br />

8 Investor’s utility function satisfies the Inada conditions. Consequently, optimal wealth is positive and (12) is a well defined<br />

equation.<br />

12


variance proportion,<br />

that,<br />

MPR<br />

t<br />

mV<br />

π<br />

t<br />

, from the hedging proportion,<br />

B γ<br />

.<br />

π , related to the investor specific bond, ( )<br />

t<br />

T I<br />

π , related to the discount bond ( )<br />

IR<br />

t<br />

B and<br />

The dynamics of optimal wealth are obtained by applying Itô’s lemma to equation (9). It is<br />

well-known that the diffusion vector of optimal wealth, σ<br />

W*<br />

, is sufficient to compute the optimal asset<br />

t<br />

allocation:<br />

1<br />

σW* = λt<br />

+ g σ<br />

B(<br />

t,<br />

TI<br />

) σ<br />

B<br />

( e,<br />

TI<br />

)<br />

t<br />

γ<br />

+<br />

(13)<br />

γt<br />

γ<br />

Optimal proportions invested in risky assets are derived by equating the diffusion part of the selffinanced<br />

portfolio (12) to that of optimal wealth (13) (Karatzas et al., 1987). By rearranging terms, we<br />

obtain the following decomposition of the traded risky asset proportions (a detailed calculation is<br />

available from the authors upon request):<br />

t<br />

T I<br />

π = π + π + π + π , (14a)<br />

t<br />

MV<br />

t<br />

mV<br />

t<br />

IR<br />

t<br />

MPR<br />

t<br />

( TB<br />

) − rt<br />

( ) ⎥ ⎤<br />

e,<br />

TH<br />

⎦<br />

MV<br />

⎡ ⎤<br />

MV<br />

π<br />

⎡<br />

≡ ⎢ ⎥ =<br />

1 −<br />

µ<br />

Bt<br />

Bt<br />

π<br />

t<br />

Σ( t,<br />

e,<br />

TH<br />

, TB<br />

)<br />

1<br />

MV<br />

⎢ , (14b)<br />

⎣π<br />

Ht ⎦ γ<br />

⎣ µ<br />

Ht<br />

π<br />

mV<br />

t<br />

⎡π<br />

≡ ⎢<br />

⎣π<br />

mV<br />

Bt<br />

mV<br />

Ht<br />

⎤<br />

⎥ = −Σ<br />

⎦<br />

−1<br />

( t,<br />

e,<br />

TH<br />

, TB<br />

) σ ( t,<br />

e,<br />

TH<br />

, TB<br />

) σ Sπ<br />

St<br />

′<br />

, (14c)<br />

π<br />

π<br />

IR<br />

t<br />

MPR<br />

t<br />

IR<br />

⎡π<br />

⎤<br />

Bt<br />

≡ ⎢ = g<br />

IR ⎥ γ<br />

Σ<br />

⎣π<br />

Ht ⎦<br />

⎡π<br />

≡ ⎢<br />

⎣π<br />

MPR<br />

Bt<br />

MPR<br />

Ht<br />

⎤<br />

⎥ = Σ<br />

⎦<br />

where µ<br />

Bt( T B<br />

), Ht( e, T H<br />

)<br />

−1<br />

( t,<br />

e,<br />

T , T ) σ ( t,<br />

e,<br />

T , T ) σ B(<br />

t,<br />

T )<br />

H<br />

B<br />

−1<br />

( t,<br />

e,<br />

T , T ) σ ( t,<br />

e,<br />

T , T ) σ ( e,<br />

T )<br />

H<br />

B<br />

H<br />

H<br />

B<br />

B<br />

′<br />

′<br />

Bγt<br />

I<br />

I<br />

, (14d)<br />

, (14e)<br />

µ are the instantaneous expected returns of the two traded risky assets and<br />

′<br />

Σ denotes the variance-covariance matrix of the traded assets.<br />

( t , e,<br />

T , T ) ≡ σ ( t,<br />

e,<br />

T , T ) σ ( t,<br />

e,<br />

T , T )<br />

H<br />

B<br />

H<br />

B<br />

H<br />

B<br />

σ e,<br />

T ) is the diffusion vector of e,<br />

T ) and is obtained by applying Itô’s lemma to<br />

Bγt<br />

(<br />

I<br />

(11a): σ ( e,<br />

T ) σ ( t,<br />

e)<br />

Bγt<br />

I<br />

B γ t<br />

(<br />

I<br />

∂Y<br />

Bγ<br />

t<br />

( e,<br />

TI<br />

)<br />

=<br />

Y<br />

, with ∂<br />

Y<br />

B<br />

γ t<br />

( e,<br />

TI<br />

) Bγ<br />

t<br />

( e,<br />

TI<br />

) = b1 γ<br />

( t,<br />

e,<br />

TI<br />

) + b2<br />

γ<br />

( t,<br />

e,<br />

TI<br />

) Yt<br />

.<br />

B ( e,<br />

T )<br />

γt<br />

I<br />

Optimal demand (equation 14a) admits the traditional decomposition. The first component is<br />

the speculative fund evolving stochastically over time because of the random character of the market<br />

13


prices of risk. As expected, it is a decreasing function of the risk aversion coefficient. The other three<br />

terms are hedging terms. The first is usually qualified as the preference-free minimum-variance term,<br />

since it does not depend on the investor’s risk aversion. It serves to hedge the risk associated with the<br />

fixed position in the spot commodity. The other two hedging addends involve the investor’s attitude<br />

IR<br />

vis-à-vis the risk. The role of π<br />

t<br />

is to hedge against random fluctuations in the instantaneous return<br />

of the discount bond B<br />

t<br />

( T I<br />

), which is a function of the instantaneously riskless rate. The second<br />

MPR<br />

component, π<br />

t<br />

, offsets the risk generated by the investor specific bond. It arises because the prices<br />

of risk are stochastic.<br />

Two main differences distinguish our model from other existing models (see, among others,<br />

Ho, 1984; Stulz, 1984; Adler and Detemple, 1988a, b; Duffie and Jackson, 1990; Briys et al., 1990;<br />

Duffie and Richardson, 1991; Lioui et al., 1996). First, concerning the last two hedging components,<br />

in these models there are as many hedging terms, called Merton-Breeden terms, as state variables. In<br />

our model, only two hedging ingredients emerge. This result is in the spirit of Lioui and Poncet (2001)<br />

and Detemple and Rindisbacher (2009). However, compared to these papers,<br />

MPR<br />

π<br />

t<br />

, in particular, is<br />

expressed in a more intuitive and convenient way. Contrary to Lioui and Poncet paper, we provide a<br />

solution to B γ<br />

e,<br />

T ) and we specify its volatility σ e,<br />

T ) . With regard to Detemple and<br />

t<br />

(<br />

I<br />

Bγt<br />

(<br />

I<br />

MPR<br />

Rindisbacher model, we express π<br />

t<br />

in terms of the investor specific bond and not in terms of the<br />

more abstract density of the forward measure. Second, in contrast to the above (complete information)<br />

models, the intertemporal demand for risky assets depends on the investor’s prior beliefs. This means<br />

that the minimum-variance ingredient especially cannot be implemented by regression analysis.<br />

Indeed, under complete information, this element is common to all investors who agree on the second<br />

moments of the asset price distribution (see, for instance, Anderson and Danthine 1981; Adler and<br />

Detemple 1988b).<br />

The decomposition given by equations (14) is in line with standard results. However, it does<br />

not allow one to determine the demand for each risky asset, to isolate the effect of the estimation error<br />

on these demands and to assess the impact of the state variables as well. Let us recall that it is possible<br />

to disentangle the effect of the estimation error on the conditional mean (see equation 3c) by<br />

14


decomposing the Brownian motion z<br />

St<br />

in two uncorrelated Wiener processes ( z rt<br />

and z<br />

wt<br />

). It follows<br />

that the futures price dynamics contain a random part involving z<br />

wt<br />

and a diffusion term as a function<br />

of the estimation error. Moreover, optimal demands (14a-e) depend on the variance-covariance matrix<br />

of the traded assets Σ ( t , e,<br />

T , T ) ≡ σ ( t,<br />

e,<br />

T , T ) σ ( t,<br />

e,<br />

T , T )<br />

H<br />

B<br />

H<br />

B<br />

′<br />

H<br />

B<br />

, which can be partitioned in a adequate<br />

way. Indeed, Anderson and Danthine (1981) and Adler and Detemple (1988b) showed that the<br />

variance-covariance matrix can be partitioned so that the mean-variance demand of some assets can be<br />

expressed as a function of the mean-variance demand of other assets. Liu (2007) generalized this<br />

separation result to the total demand in traded assets. We suggest a partition of the diffusion matrix<br />

that is suited to our framework. To this effect, we introduce into the financial market a synthetic asset 9<br />

perfectly correlated with the orthogonal source of risk z<br />

wt<br />

:<br />

dH<br />

H<br />

wt<br />

wt<br />

( e,<br />

TH<br />

)<br />

= µ<br />

Hwt<br />

( e,<br />

TH<br />

) dt + σ<br />

Hw( t,<br />

e,<br />

TH<br />

) dzwt<br />

, (15)<br />

( e,<br />

T )<br />

H<br />

where, by absence of arbitrage opportunities µ Hwt<br />

( e, T H<br />

) r t<br />

+ σ Hw<br />

( t,<br />

e,<br />

T H<br />

) λ wt<br />

= , and<br />

λ<br />

wt<br />

2<br />

[ λ − ρ λ ] − ρ<br />

= is the price of risk related to z<br />

wt<br />

.<br />

St<br />

Sr<br />

rt<br />

1<br />

Sr<br />

In Propositions 1 and 2 below, we take advantage of the synthetic asset to provide more<br />

insightful expressions of the minimum-variance, the speculative and the hedging components.<br />

Proposition 1.<br />

a) The mean-variance component can be written as follows:<br />

π<br />

MV<br />

Ht<br />

( e,<br />

TH<br />

) −<br />

( t,<br />

e,<br />

T ) 2<br />

1 µ<br />

Hwt<br />

rt<br />

= , (16a)<br />

γ σ<br />

Hw<br />

H<br />

( t,<br />

e,<br />

TH<br />

, TB<br />

)<br />

( t,<br />

T )<br />

1 µ ( T ) − r Σ<br />

= π . (16b)<br />

MV Bt B t HB<br />

π<br />

Bt<br />

−<br />

2<br />

2<br />

γ σ<br />

B( t,<br />

TB<br />

) σ<br />

B B<br />

b) The minimum-variance component can be written as follows:<br />

( t,<br />

e,<br />

TH<br />

)<br />

( t,<br />

e,<br />

T )<br />

mV HwS<br />

π<br />

Ht<br />

2<br />

σ<br />

Hw H<br />

St<br />

MV<br />

Ht<br />

Σ<br />

= −<br />

π , (17a)<br />

9 This asset can be duplicated, in our complete market, by a portfolio of the risk-free asset and the two traded risky assets.<br />

15


( t,<br />

e,<br />

TH<br />

, TB<br />

)<br />

( t,<br />

T )<br />

Σ ( t,<br />

T ) Σ<br />

= −<br />

−<br />

π . (17b)<br />

mV BS B<br />

HB<br />

π<br />

Bt<br />

π<br />

2 St<br />

2<br />

σ<br />

B<br />

( t,<br />

TB<br />

) σ<br />

B B<br />

{ S;<br />

B;<br />

H }<br />

Σ , k , j ∈ ; stands for the covariance between assets k, j k ≠ j .<br />

kj<br />

H w<br />

mV<br />

Ht<br />

Σ<br />

σ<br />

HwS<br />

Hw<br />

2<br />

( t,<br />

e,<br />

TH<br />

) 1−<br />

ρSrσ<br />

S<br />

=<br />

2<br />

( e,<br />

t;<br />

T ) σ ( e,<br />

t;<br />

T )<br />

H<br />

Hw<br />

H<br />

,<br />

Σ<br />

HB<br />

σ<br />

( t,<br />

e,<br />

TH<br />

, TB<br />

)<br />

2<br />

( t,<br />

T )<br />

B<br />

B<br />

σ<br />

Hr<br />

= −<br />

σ<br />

( t,<br />

e,<br />

TH<br />

)<br />

( t,<br />

T )<br />

B<br />

B<br />

ΣBS<br />

( t,<br />

TB<br />

)<br />

and =<br />

σ<br />

B<br />

2<br />

( t,<br />

T ) σ ( t,<br />

T )<br />

B<br />

ρ σ<br />

B<br />

Sr<br />

S<br />

B<br />

.<br />

Proof. Available from the authors upon request.<br />

Despite their difference in interpretation and use, the mean-variance and the minimumvariance<br />

demands have two common features. First, they are expressed in a simple recursive manner<br />

and consist of a fund specific to the futures contract and a fund related to the discount bond.<br />

MV<br />

π<br />

Ht<br />

and<br />

mV<br />

π<br />

Ht<br />

depend on the first two moments of the synthetic asset and are thus directly related to<br />

zwt<br />

underscoring the importance of the decomposition of z<br />

St<br />

. In other terms, these elements are<br />

MV<br />

directly related to the orthogonal source of risk of the estimate of the convenience yield. π<br />

Ht<br />

reflects<br />

the investor’s expectations about the orthogonal source of risk, while<br />

mV<br />

π<br />

Ht<br />

hedges against the<br />

orthogonal risk affecting the fixed position in the spot commodity. The speculative and minimumvariance<br />

proportions of the discount bond are associated with the risk of the short rate. However, since<br />

MV<br />

the futures contract is correlated with the short rate, these proportions are adjusted by π<br />

Ht<br />

and<br />

respectively, weighted by the usual variance/covariance ratios. Second, Proposition 1 underlines the<br />

relation between the investor’s prior beliefs and the demand in the futures contract. Indeed, as<br />

mV<br />

π<br />

Ht<br />

explained above, the estimation error influences the estimate of the convenience yield through<br />

z<br />

wt<br />

.<br />

MV<br />

π<br />

Ht<br />

and<br />

mV<br />

π<br />

Ht<br />

are devoted to this source of uncertainty. It follows that the mean-variance and<br />

minimum-variance demands of the futures contract are functions of the estimation error and thus of the<br />

MV<br />

mV<br />

investor’s initial prior beliefs. π<br />

Bt<br />

and π<br />

Bt<br />

are impacted by the estimation procedure only through<br />

MV<br />

mV<br />

π<br />

Ht<br />

and π<br />

Ht<br />

.<br />

The mean-variance and minimum variance proportions are functions of volatilities and of<br />

covariances. Moreover, the investor’s position, short or long, depend on the sign of the fixed position<br />

MV<br />

and of that of the covariances. Proposition 1 shows that π<br />

Ht<br />

and<br />

π are functions of σ ( t , e,<br />

)<br />

mV<br />

Ht<br />

Hw<br />

T H<br />

16


and are likely to exhibit the same pattern as that of this volatility, which dominates the futures price<br />

volatility. Especially, as will be seen later, futures price volatility evolves according to the Samuelson<br />

effect, which suggests that the volatility of futures prices for contracts close to maturity exceeds that of<br />

MV<br />

mV<br />

more distant contracts. This effect implies that π<br />

Ht<br />

and π<br />

Ht<br />

are increasing functions, in absolute<br />

values, of the futures contract maturity. The mean-variance and minimum variance demands in the<br />

MV<br />

mV<br />

bond are modified by π<br />

Ht<br />

and π<br />

Ht<br />

weighted by<br />

Σ<br />

−<br />

HB<br />

σ<br />

( t,<br />

e,<br />

TH<br />

, TB<br />

)<br />

( t,<br />

T ) 2<br />

have a positive impact on the futures contract, the covariance ( t , e,<br />

T , T )<br />

B<br />

B<br />

. As interest rates are expected to<br />

Σ should be negative<br />

MV<br />

mV<br />

implying that this ratio should be positive. Therefore, the sign of π<br />

Ht<br />

and π<br />

Ht<br />

will determine their<br />

MV<br />

mV<br />

impact (negative or positive) on π<br />

Bt<br />

and π<br />

Bt<br />

.<br />

We turn now to the study of the last two hedging elements in equation (14). A direct<br />

calculation of (14d) gives:<br />

π = 0 , (18a)<br />

π<br />

IR<br />

Ht<br />

σ<br />

( t,<br />

TI<br />

)<br />

( t,<br />

T )<br />

IR<br />

B<br />

Bt<br />

= gγ<br />

, (18b)<br />

σ<br />

B B<br />

The hedging demand stemming from the stochastic behavior of the return of the discount bond with<br />

HB<br />

H<br />

B<br />

maturity<br />

T<br />

I<br />

reduces to the proportion,<br />

IR<br />

π<br />

Bt<br />

, invested in the bond that provides insurance against<br />

interest rate risk. It is independent of the estimation error. Its expression is given in (18b) confirming a<br />

well-known result, that is it is proportional to the ratio of the volatilities of two bonds with maturity T<br />

I<br />

and T<br />

B<br />

respectively (see Lioui and Poncet, 2001; Munk and Sorensen, 2004). Given that investors are,<br />

in general, more risk averse than the Bernoulli investor and that<br />

T < T , this proportion is positive and<br />

I<br />

B<br />

less than one.<br />

MPR<br />

The component π<br />

t<br />

hedges against the risk generated by the investor specific discount bond.<br />

It arises from the need to hedge against the stochastic prices of risk, which they are functions of the<br />

state variables. Equation (14e) involves the diffusion term of ( )<br />

t<br />

T I<br />

B γ<br />

. By definition, σ e,<br />

T )<br />

depends on the volatilities of the state variables. It can then be written in the following manner:<br />

Bγt<br />

(<br />

I<br />

17


σ<br />

( e,<br />

T ) + σ rΨ<br />

( e,<br />

T ) + σ m(<br />

t,<br />

e)<br />

Ψ ( e T )<br />

Bγt<br />

( , TI<br />

) σ X Ψγ<br />

Xt I<br />

γrt<br />

I<br />

γmt<br />

e = , , (19)<br />

where Ψ ( e, T ) = ∂ B ( e,<br />

T ) B ( e,<br />

T ), i ∈{ X , r m}<br />

B γ<br />

e,<br />

T ) .<br />

t<br />

(<br />

I<br />

it I i γt<br />

I γt<br />

I<br />

,<br />

γ<br />

are the components of the gradient vector of<br />

MPR<br />

π<br />

t<br />

can be decomposed into three terms for each and every state variables:<br />

I<br />

π = π + π + π , (20)<br />

MPR<br />

t<br />

MPRX<br />

t<br />

MPRr<br />

t<br />

MPRm<br />

t<br />

The following proposition is devoted to these terms.<br />

Proposition 2.<br />

The optimal hedging proportions generated by the investor specific bond may be decomposed for each<br />

and every state variable.<br />

a) The risk associated with the (log) spot commodity price is hedged by the futures contract and the<br />

discount bond.<br />

π<br />

MPRX<br />

Ht<br />

Σ<br />

( t,<br />

e,<br />

T )<br />

( e,<br />

T )<br />

HwS H<br />

= Ψ<br />

2 γXt<br />

I<br />

, (21a)<br />

σ<br />

Hw( t,<br />

e,<br />

TH<br />

)<br />

( e,<br />

T )<br />

( t,<br />

e,<br />

TH<br />

, TB<br />

)<br />

( t,<br />

T )<br />

Σ ( t,<br />

T ) Σ<br />

= Ψ −<br />

π . (21b)<br />

π<br />

MPRX BS B<br />

HB<br />

Bt<br />

2 Xt I<br />

σ<br />

B( t,<br />

TB<br />

)<br />

γ σ<br />

2<br />

B B<br />

MPRX<br />

Ht<br />

MPRr<br />

b) The risk associated with the interest rate is hedged by the discount bond ( π = 0 ).<br />

π<br />

Σ<br />

( t,<br />

T )<br />

( e,<br />

T )<br />

MPRr Br B<br />

Bt<br />

= Ψ . (22)<br />

2 γrt<br />

I<br />

σ<br />

B( t;<br />

TB<br />

)<br />

c) The risk associated with the estimate of the convenience yield is hedged by the futures contract and<br />

the discount bond.<br />

π<br />

Σ<br />

( t,<br />

e,<br />

T )<br />

( e,<br />

T )<br />

MPRm Hwm H<br />

Ht<br />

= Ψ<br />

2 γmt<br />

I<br />

, (23a)<br />

σ<br />

Hw( t,<br />

e,<br />

TH<br />

)<br />

( e,<br />

T )<br />

( t,<br />

e,<br />

TH<br />

, TB<br />

)<br />

( t,<br />

T )<br />

Σ ( t,<br />

T ) Σ<br />

= Ψ −<br />

π . (23b)<br />

π<br />

MPRm Bm B<br />

HB<br />

Bt<br />

2 mt I<br />

σ<br />

B( t;<br />

TB<br />

)<br />

γ σ<br />

2<br />

B B<br />

MPRm<br />

Ht<br />

Σ , i ∈{ X ; r;<br />

m} , j { B;<br />

H } stands for the covariance between the assets { }<br />

i , j<br />

∈<br />

w<br />

Ht<br />

B; and the state<br />

H w<br />

Σ<br />

; .<br />

σ<br />

variables { X r;<br />

m}<br />

Br<br />

B<br />

( t,<br />

TB<br />

)<br />

= −<br />

2<br />

( t;<br />

T ) D ( t T )<br />

B<br />

α<br />

1<br />

,<br />

B<br />

,<br />

Σ<br />

σ<br />

Hwm<br />

Hw<br />

( t,<br />

e,<br />

TH<br />

)<br />

=<br />

2<br />

( t,<br />

e,<br />

T ) σ ( t,<br />

e,<br />

T )<br />

H<br />

σ ( t,<br />

e)<br />

Hw<br />

mw<br />

H<br />

,<br />

ΣBm(<br />

t,<br />

TB<br />

)<br />

= −<br />

σ<br />

B<br />

σ δ<br />

ρr<br />

δ<br />

2<br />

( t;<br />

T ) D ( t,<br />

T )<br />

B<br />

α<br />

B<br />

.<br />

Proof. Available from the authors upon request.<br />

18


Proposition 2 seems to suggest a decomposition à la Merton-Breeden as the optimal hedging<br />

proportions are derived for each and every state variable. At a first sight, this result may be in<br />

contradiction with the decomposition given in equations (14) in terms of two bonds. However, a<br />

preliminary remark is in order. Thanks to the change of probability measure mentioned above, the<br />

Merton-Breeden hedging terms reduce to two addends involving two bonds whatever the number of<br />

the state variables. In order to study how the state variables and incomplete information affect optimal<br />

demands, the latter are couched in terms of each variable. An important difference with the familiar<br />

Merton-Breeden terms is that the decomposition in Proposition 2 preserves the advantages of<br />

equations (14) since it depends on the characteristics of the two bonds.<br />

In the same manner as for the minimum-variance and mean-variance elements, these hedging<br />

terms are expressed in a convenient recursive way. The futures contract serves to hedge the orthogonal<br />

risk of the estimate of the convenience yield and directly captures the effect of the estimation error.<br />

The latter is also captured through Ψ ( e, ), which, given that ( )<br />

γit T I<br />

B γ<br />

has the functional form of a<br />

discount bond, admit a simple economic interpretation. It represents the sensitivity of the investor<br />

discount bond on the three state variables (see also Wachter, 2002). In particular, ( e, )<br />

t<br />

T I<br />

Ψ measures<br />

γmt T I<br />

the sensitivity of B γ t<br />

( T I<br />

) to changes in the estimate of the convenience yield and thus the impact of the<br />

incomplete information.<br />

Proposition 2 has another two advantages. First, it allows an investor to assess the impact of<br />

each state variable on optimal hedging demands and to decide which of the state variables are worth to<br />

be comprised in the investment opportunity set. This assessment is based on ( e, )<br />

Ψ weighted by the<br />

γit T I<br />

covariance/variance ratios. Second, Proposition 2 shows which risky assets should be used and in what<br />

proportions in order to hedge the risk of the investor specific bond related to the state variables. To<br />

gain intuition on how state variables affect Ψ ( e, )<br />

γit T I<br />

, let us rewrite expression (10) in a convenient<br />

way separating the price of risk related to the orthogonal risk from that associated with the short rate<br />

risk:<br />

B<br />

γt<br />

( ) ⎡ ⎪⎧<br />

g gγ<br />

2<br />

2<br />

( TI<br />

) Et<br />

⎢exp⎨−<br />

[ λwu<br />

+ λru<br />

−σ<br />

B<br />

( u,<br />

TI<br />

) ]<br />

T I<br />

⎪⎫<br />

⎤<br />

γ<br />

= ∫<br />

du⎬⎥<br />

(10b)<br />

⎢⎣<br />

⎪⎩ 2γ<br />

t<br />

⎪⎭ ⎥⎦<br />

19


The sign of Ψ ( e, ) and ( e, )<br />

γXt T I<br />

Ψ can be examined through the effect of X(t) and m(t) on λ<br />

wt<br />

,<br />

γmt T I<br />

while the analysis of the sign of ( e, )<br />

Ψ is more complicated since r(t) impacts both λ<br />

rt<br />

and<br />

γrt T I<br />

Note that for more risk-averse investors than the logarithmic investor, γ ≥1<br />

and g ≥ 0<br />

γ<br />

λ<br />

wt<br />

.<br />

. Therefore, as<br />

the exponential involves a quadratic function, when<br />

λ is positive (negative), the sign of Ψ ( e, )<br />

wt<br />

γXt T I<br />

should be the opposite (same) to that of X(t ) and m(t) on<br />

λ<br />

wt<br />

. The same reasoning applies to<br />

Ψ ( e, ) . Unfortunately, it is not possible to directly deduce the sign of Ψ ( e, )<br />

γmt T I<br />

γrt T I<br />

. However, as the<br />

correlation of the short rate and the risky assets, the spot commodity in particular, is negative, we<br />

guess that its sign is negative.<br />

4. Illustration<br />

This section provides an illustration of our model in the case of the copper market. We simulate how<br />

speculative and hedging demands react to the investor’s horizon, to the state variables changes and to<br />

the investor’s initial beliefs. The parameters values are based on those estimated by Schwartz (1997)<br />

and Casassus and Collin-Dufresne (2005). To focus on the influence of the convenience yield and the<br />

spot price on optimal demands, we run our simulations for a flat (forward) initial term structure of the<br />

risk-free rate. The values of α and the initial forward yield curve, f ( ) , are computed so that θ (∞ )<br />

and β are respectively equal to the values of the long term mean (3%) and speed of mean-reversion<br />

(0.2) of the risk-free rate provided in Casassus and Collin-Dufresne (2005). Table 1a gathers the<br />

values of the parameters used in our simulations.<br />

[INSERT TABLE 1a ABOUT HERE]<br />

The parameters values are set so that we can reproduce some stylized facts characterizing<br />

commodities. Spot commodity prices and the convenience yield (see Bessembinder et al., 1995) as<br />

well as the short rate exhibit mean reversion, so that α > 0 and k > 0. As Casassus and Collin-Dufresne<br />

(2005) pointed out, mean-reversion in the state variables is reinforced by that of prices of risk. Thus,<br />

0 t<br />

r<br />

λ<br />

SX<br />

<<br />

rr<br />

0 and λ < 0 . To assess the impact of the state variables, we consider three scenarios allowing<br />

us to take into account backwardation (more frequently observed in commodity markets) and contango<br />

20


situations. For each scenario, the short-term rate is equal to its long-term mean in order to focus our<br />

analysis on the specific impact of the spot price and the convenience yield. The first scenario describes<br />

a situation where the (log) spot price and the estimate of the convenience yield are equal to their longterm<br />

mean. For the second (third) scenario, called backwardation (contango), the spot price is 10%<br />

above (below) its long term mean and the estimate of the convenience yield is equal to 9.5% (3.5%).<br />

The three scenarios are shown inTable 1b.<br />

[INSERT TABLE 1b ABOUT HERE]<br />

The filtering error steady state value can be computed using the values in Table 1a: ε ∞ =0.59%.<br />

We then study the effect of the initial value of the estimation error for high values above (below) the<br />

steady state, e = 1% > ε ∞ ( e = 0.1% < ε ∞ ). When examining optimal demands as a function of the<br />

futures contract maturity, we let this maturity vary between 0 and 18 months, since copper contracts<br />

are usually available for maturities up to 24 months (less liquid contracts). Investor’s horizon also<br />

varies in the same interval. Throughout the analysis, we set the bond maturity equal to seven years,<br />

T B =7, and, for simplicity, the constrained position equal to 1, π = 1. Finally, we consider the case of<br />

a more-risk averse investor than the log-utility (myopic) investor, and we retain a value of γ = 3 . For<br />

compactness of notation, we drop out the parentheses from the quantities we study.<br />

[INSERT FIGURES 1a AND 1b ABOUT HERE]<br />

Figures 1a,b display σ H and σ Hw as a function of T H , respectively. Figure 1a shows the<br />

decreasing pattern of the futures contract volatility as a function of its maturity, i.e. the Samuelson<br />

effect. This effect is more pronounced for a high initial estimation error. Moreover, the higher the<br />

initial error, the lower the volatility of the futures price. This can be explained as follows. For the<br />

values used in this illustration, κ S<br />

> 0 , which implies that e has a positive impact on σ<br />

mw<br />

and thus on<br />

δ<br />

σ H . However, mean reversion in the market prices of risk outweighs this positive effect resulting in a<br />

decreasing σ H as a function of e. As the maturity of the futures contract increases, this phenomenon is<br />

amplified. Figure 1b reveals that the behavior of σ H is essentially dictated by that of σ Hw . Indeed,<br />

additional results, not reproduced here, show that, as expected, σ Hr is positive, but its value is low and<br />

decreases from around 3.2% to around 2.4%.<br />

St<br />

21


[INSERT TABLE 2 ABOUT HERE]<br />

Table 2 displays the covariance/variance ratios in propositions 1 and 2 as a function of T H . As<br />

these ratios determine the investor’s position (short or long) as well as its magnitude, it is interesting to<br />

examine their evolution over time. As expected, Σ HwS /σ Hw 2<br />

is positive and since it is inversely<br />

proportional to σ Hw , from the analysis above, it follows that it is an increasing function of both the<br />

futures contract maturity and the initial value of the estimation error. Their influence is substantial.<br />

Indeed, when e = 0.1%, this ratio increases by 78.33% between T H = 3 months and T H = 18 months.<br />

When e = 1%, the rise in this ratio, for the same period, is 91.4%. Moreover, the longer the maturity<br />

the higher the impact of e on this ratio. In contrast, Σ HB /σ 2 B is negative and, as discussed above, much<br />

less sensitive than Σ HwS /σ 2 Hw to T H and e.<br />

[INSERT TABLE 3 ABOUT HERE]<br />

Let us now analyze how futures contracts maturity affects the minimum-variance and meanvariance<br />

proportions (see tables 3 and 4 respectively). For future reference, we denote by<br />

Σ<br />

≡ π , l ∈ { mV , MV , MPR,<br />

MPRX , MPRm}<br />

the part of the bond demand related to the<br />

l<br />

HB<br />

π<br />

B,<br />

H<br />

−<br />

2<br />

σ<br />

B<br />

l<br />

H<br />

futures contract demand. Since π = 1, the minimum-variance futures proportion is equal to -Σ HwS<br />

St<br />

/σ 2 Hw and mimics its behavior. The minimum-variance futures hedge demand is an opposite position to<br />

the spot commitment. The pure hedge of the discount bond is a little more complicated. The spot<br />

ΣBS<br />

commodity is negatively correlated with the interest rate so that<br />

2<br />

σ<br />

B<br />

is negative and<br />

Σ<br />

−<br />

BS<br />

π<br />

S<br />

σ 2<br />

B<br />

is<br />

positive. However, as seen above,<br />

mV<br />

π<br />

B<br />

exacerbates the effect of the first term.<br />

mV<br />

is adjusted by second term π<br />

B,H<br />

, which is also negative and<br />

[INSERT TABLE 4 ABOUT HERE]<br />

Table 4 displays the mean-variance proportions for the three scenarios considered in Table 1b.<br />

The impact of T H and the initial estimation error on<br />

MV<br />

π<br />

H<br />

is similar to that on<br />

mV<br />

π<br />

H<br />

. The term<br />

1 µ<br />

Bt(<br />

TB<br />

) − rt<br />

γ σ<br />

B<br />

( t,<br />

T ) 2<br />

B<br />

is independent on T H and e. It is the same across all scenarios and is equal to 128.3%.<br />

Since Σ HB /σ 2 MV<br />

B is negative, π<br />

B,H<br />

adds to the above term resulting in a substantial speculative demand,<br />

22


MV<br />

π<br />

B<br />

, for the discount bond. When the spot price is above its long-term mean, speculative demands are<br />

higher than those when the spot price is below its long-term mean. This result, may, at a first sight, be<br />

seen as counter-intuitive. However, it can be explained by mean-reversion in the market prices of risk.<br />

The higher the spot price, the lower λ wt , and the higher the estimate of the convenience yield, the<br />

higher λ wt . However, the spot price effect dominates that of the estimate.<br />

[INSERT TABLE 5 ABOUT HERE]<br />

MV<br />

Table 5 performs a similar analysis as for π<br />

H<br />

and<br />

the investor’s horizon.<br />

π of , i ∈{ X , r m}<br />

mV<br />

H<br />

Ψ γ it<br />

,<br />

as a function of<br />

Ψ<br />

γit<br />

are, in absolute value, increasing functions of T I and the results confirm the<br />

intuition about their sign explained below Proposition 2. In addition, a similar explanation to the case<br />

of<br />

MV<br />

π<br />

H<br />

and<br />

mV<br />

π<br />

H<br />

can be put forward to understand the behavior of the sensitivities in the three<br />

MV<br />

scenarios. Finally, in contrast to π<br />

H<br />

and<br />

a negligible effect on<br />

value of e.<br />

mV<br />

π<br />

H<br />

, a change in the value of the initial estimation error has<br />

Ψ<br />

γit<br />

. The sensitivities depend more on the level of the state variables than on the<br />

[INSERT TABLE 6 ABOUT HERE]<br />

Table 6 exhibits the hedging components related to the investor discount bond and generated<br />

by the random behavior of the prices of risk. These demands depend on both T H and T I . However, to<br />

simplify their understanding, without loss of generality, we set T H =T I . As is demonstrated in<br />

Proposition 2, π is a weighted average of the sensitivities Ψ<br />

Xt<br />

and Ψ<br />

mt<br />

. The weighs are equal to<br />

MPR<br />

Ht<br />

the variance covariance ratios Σ HwS /σ Hw<br />

2<br />

and Σ Hwm /σ Hw 2 , which, in our case, are positive. In contrast,<br />

γ<br />

γ<br />

Ψ<br />

γXt<br />

and Ψγ<br />

mt<br />

MPR<br />

have an opposite sign and therefore the two terms in π<br />

Ht<br />

partially offset each other.<br />

MPR<br />

This explains the low values of π<br />

Ht<br />

for short maturities especially and the change of the position<br />

(long or short) for e =0 .01%. π is a function of Ψ<br />

Xt<br />

, Ψ<br />

mt<br />

and Ψ<br />

rt<br />

. To study this term turns out to<br />

be more complex than<br />

π<br />

MPR<br />

Bt<br />

B<br />

( t;<br />

T )<br />

MPR<br />

π<br />

Ht<br />

Σ<br />

BX<br />

( t,<br />

T )<br />

=<br />

σ<br />

MPR<br />

Bt<br />

γ<br />

MPR<br />

. However, write π<br />

Bt<br />

as follows facilitates its interpretation:<br />

( e,<br />

T )<br />

Σ<br />

( t,<br />

T )<br />

( e,<br />

T )<br />

B<br />

Br B<br />

Bm B<br />

Ψ<br />

2 γXt<br />

I<br />

+ Ψ<br />

2 γrt<br />

I<br />

+ Ψ<br />

2 γmt<br />

,<br />

B<br />

σ<br />

B<br />

( t;<br />

TB<br />

)<br />

σ<br />

B<br />

( t;<br />

TB<br />

)<br />

Σ<br />

γ<br />

( t,<br />

T )<br />

γ<br />

MPR<br />

( e T ) −π<br />

I<br />

B,<br />

Ht<br />

23


As Σ BX /σ B<br />

2<br />

=-55.8%, Σ Br /σ B<br />

2<br />

=-17.5% and Σ Bm /σ B<br />

2<br />

=-56.0%, the first term, in the equation above, is<br />

negative, while the second and third components are positive. Moreover, the first term partially<br />

MPR<br />

MPR<br />

compensates the third one and given that π<br />

B,Ht<br />

takes low values, π<br />

Bt<br />

is, for a large part, driven by<br />

the second element. With regard to the impact of e, an inspection of Table 5 reveals, however, that<br />

Ψγrt<br />

is insensible to e. Indeed, a quick calculation shows that the difference<br />

MPR<br />

MPR<br />

MPR<br />

MPR<br />

π ( e = %) − π ( e 0.1% ) is similar to the difference ( e = 1 %) − π ( e = 0.1% )<br />

B, Ht<br />

1<br />

B,<br />

Ht<br />

=<br />

π .<br />

Bt<br />

Bt<br />

5. Concluding remarks<br />

In this article we have studied the Traditional Hedging Model (Adler and Detemple, 1988 a,b), i.e. the<br />

issue of hedging a fixed cash asset position with correlated futures contracts, for commodity markets<br />

when the convenience yield is not observable and is estimated given the information conveyed by the<br />

spot commodity and the short-rate by using the continuous-time Kalman filter. The estimate of the<br />

convenience yield and the estimation error directly affect the mean-variance and minimum-variance<br />

demands in futures contracts. Moreover, the initial value of the estimation error has a heavy impact on<br />

these demands. The positions in the discount bond are impacted by the estimation procedure only<br />

through the demand in futures contracts. We achieve a decomposition of the hedging component<br />

related to the stochastic prices of risk by distinguishing the effect of each and every state variable.<br />

Contrary to the other two demands, these hedging elements are not very sensitive to the initial<br />

estimation error.<br />

Our framework can be extended in order to take into account richer dynamics of the state<br />

variables including, for example, jumps in the state variables (see Jeanblanc et al. 2009) and/or other<br />

utility functions (recursive utility, loss aversion, for instance). In the paper, we assume that investors<br />

use the information coming from interest rates and spot commodity prices to infer the convenience<br />

yield. However, investors can have access to private information regarding the value of the<br />

convenience yield through their business activity. This information would result in an additional<br />

source of uncertainty. It would then be optimal for the investor to use two futures contracts as hedging<br />

instruments.<br />

24


References<br />

Adler M. and J. Detemple (1988a). On the Optimal Hedge of a Non-Traded Cash Position,<br />

Journal of Finance, 43, 143-153.<br />

Adler M. and J. Detemple (1988b). Hedging with Futures in an Intertemporal Portfolio Context,<br />

Journal of Futures Markets, 8, 249-269.<br />

Ahn, D-H., Dittmar, R., Gallant, R., 2002. Quadratic Term Structure Models: Theory and Empirical<br />

Evidence, Review of Financial Studies 15, 243-288.<br />

Anderson R. And J-P Danthine (1981). Cross Hedging, Journal of Political Economy, 89, 1182- 1196.<br />

Bajeux-Besnamou, I., Portait, R., 1997. The Numeraire Portfolio: A New Perspective on Financial<br />

Theory, The European Journal of Finance 3, 291-309.<br />

Bessembinder H. and K. Chan (1992). Time-Varying Risk Premia and Forecastable Returns in<br />

Futures Markets, Journal of Financial Economics, 32, 169-193.<br />

Bessembinder H., J. Coughenour, P. Seguin and S.M. Smoller (1995). Mean Reversion in<br />

Equilibrium Asset Prices: Evidence from the Futures Term Structure, Journal of Finance, 50,<br />

205-238.<br />

Bertus M., J. Godbey and J. Hilliard (2009). Minimum Variance Cross-Hedging Under Mean<br />

Reverting Spreads, Stochastic Convenience Yields and Jumps: An Application to the<br />

Airline Industry, Journal of Futures Markets, 29, 736-756.<br />

Björk, T., Davis, M.H., Landén, C., 2008. Optimal Investment under Partial Information. Unpublished<br />

working paper, Stockholm School of Economics.<br />

Breeden D. (1979). An Intertemporal Asset Pricing Model With Stochastic Consumption and<br />

Investment Opportunities, Journal of Financial Economics, 7, 263-296.<br />

Brennan, M., 1958. The Supply of Storage, American Economic Review 48, 50-72.<br />

25


Brennan M. (1991). The Pricing Of Convenience And The Valuation Of Commodity Contingent<br />

Claims, in: Lund, D., Oskendal, B. (Eds), Stochastic models and option values. Elsevier Science<br />

Publishers, North Holland, Amsterdam, The Netherlands, 33-71.<br />

Briys E., M. Crouhy and H. Schlesinger (1990). Optimal Hedging Under Intertemporally<br />

Dependent Preferences, Journal of Finance, 45, 1315-1324.<br />

Carmona R. and M. Ludkovski (2004a). Spot Convenience Yield Models for the Energy Market,<br />

Unpublished Working Paper.<br />

Carmona R. and M. Ludkovski (2004b), Convenience Yield Model With Partial Observations and<br />

Exponential Utility, Unpublished Working Paper.<br />

Casassus J. and P. Collin-Dufresne (2005). Stochastic Convenience Yield Implied from Commodity<br />

Futures and Interest Rates, Journal of Finance, 5, 2283-2331.<br />

Cox J. and C. Huang (1989). Optimum Consumption and Portfolio Policies When Asset Prices<br />

Follow a Diffusion Process, Journal of Economic Theory, 49, 33-83.<br />

Dai, Q., Singleton K., 2000. Specification Analysis of Affine Term Structure Models, The Journal of<br />

Finance 55, 1943-1978.<br />

Dai Q. and K. Singleton (2002). Expectation Puzzles, Time Varying Risk Premia, and Affine Models<br />

of the Term Structure, Journal of Financial Economics, 63, 415-441.<br />

Detemple J. (1986). Asset Pricing in a Production Economy with Incomplete Information,<br />

Journal of Finance, 41, 383-391.<br />

Detemple J. and M. Rindisbacher (2009). Dynamic Asset Allocation: Portfolio Decomposition<br />

Formula and Applications, Review of Financial Studies, forthcoming.<br />

Dincerler, C., Z. Khokher, Simin T., 2005. An Empirical Analysis of Commodity Convenience Yields,<br />

Unpublished Working Paper, The Pennsylvania State University.<br />

Dothan M. and D. Feldman (1986). Equilibrium Interest Rates and Multiperiod Bonds in a<br />

Partially Observable Economy, Journal of Finance, 41, 369-382.<br />

26


Duffie D. and O. Jackson (1990). Optimal Hedging and Equilibrium in A Dynamic Futures Market,<br />

Journal of Economic Dynamics and Control, 14, 21-33.<br />

Duffie D. and H. Richardson (1991). Mean-Variance Hedging in Continuous Time, Annals of Applied<br />

Probability, 1, 1-15.<br />

Duffie D. and R. Stanton (1992). Pricing Continuously Resettled Contingent Claim, Journal of<br />

Economic Dynamics and Control, 16, 561-573.<br />

Fama E. and K. French (1987). Commodity Futures Prices: Some Evidence on Forecast Power,<br />

Premium and The theory of Storage, Journal of Business, 60, 55–74.<br />

Fama, E., French, K., 1988. Business Cycles and the Behaviour of Metals Prices, Journal of Finance<br />

43, 1075-1093.<br />

Feldman D. (2007). Incomplete Information Equilibria: Separation Theorems and Other Myths, Annals<br />

of Operations Research, 161, 119-149.<br />

Geman H., N. El Karoui and J-C. Rochet (1995). Changes of Numeraire, Changes of<br />

Probability<br />

Measure and Option Pricing, Journal of Applied Probability, 32, 443-458.<br />

Gennotte G. (1986). Optimal Portfolio Choice Under Incomplete Information, Journal of<br />

Finance, 41, 733-746.<br />

Gibson, R., Schwartz, E., 1990. Stochastic Convenience Yield and the Pricing of Oil Contingent<br />

Claims, Journal of Finance 45, 959-976.<br />

Heath D., R. Jarrow R. and R. Morton (1992). Bond Pricing and the Term Structure of Interest Rates:<br />

A New Methodology for Contingent Claims Valuation, Econometrica, 60, 77-105.<br />

Ho T. (1984). Intertemporal Commodity Futures Hedging and the Production Decisions,<br />

Journal of Finance, 39, 351-376.<br />

Hong H. (2001). Stochastic Convenience Yield, Optimal Hedging and The Term Structure of Open<br />

Interest and Futures Prices, Unpublished Working Paper, Graduate School of Business,<br />

Stanford University.<br />

Hong H. and M. Yogo (2009). Digging Into Commodities, Unpublished Working Paper, Princeton<br />

University and Wharton of University of Pennsylvania.<br />

27


Hull J. and A. White (1990). Pricing Interest-Rate-Derivative Securities, Review of Financial Studies,<br />

4, 573-592.<br />

Jamshidian, F., 1987. Pricing of Contingent Claims in the One Factor Term Structure Model, Merrill<br />

Lynch Capital Markets.<br />

Jamshidian F. (1989). An Exact Bond Option Formula, Journal of Finance, 44, 205-209.<br />

Jeanblanc M., V. Lacoste and S. Roland (2009). Utility Maximization Under a Partially Information<br />

Observed Jump-Diffusion Model, Unpublished Working Paper, Evry University and Rouen<br />

Business School.<br />

Kaldor, N., 1939. Speculation and Economic Stability, Review of Economic Studies 7, 1-27.<br />

Karatzas I., J. Lehoczky and S. Shreve (1987): “Optimal Portfolio and Consumption Decisions or a<br />

Small Investor on A Finite Horizon.” Siam Journal on Control and Optimization, 25, 1557-1586.<br />

Lioui, A., Nguyen Duc Trong, P., Poncet, P., 1996. Optimal dynamic hedging in incomplete futures<br />

markets. The Geneva Papers on Risk and Insurance Theory 21-1, 103-122.<br />

Lioui A. and P. Poncet (2001). On The Optimal Portfolio Choice Under Stochastic Interest Rates,<br />

Journal of Economic Dynamics and Control, 25, 1841-1865.<br />

Liptser R. S. and A. N. Shiryaev (2001a), Statistics of Random Processes I, 2nd Edition, Springer,<br />

New-York.<br />

Liptser R. S. and A. N. Shiryaev (2001b). Statistics of Random Processes II, 2nd Edition,<br />

Springer, New-York.<br />

Litzenberger, R., Rabinowitz, N. 1995. Backwardation in Oil Futures Markets: Theory and<br />

Empirical Evidence, Journal of Finance 50, 1517-1545.<br />

Liu J. (2007). Portfolio Selection in Stochastic Environments, Review of Financial Studies,20, 1-39.<br />

Long, J. (1990). The Numeraire Portfolio, Journal of Financial Economics, 26, 29-69.<br />

Lundtofte F. (2006). The Effect of Information Quality on Optimal Portfolio Choice, Financial<br />

Review, 41, 157-185.<br />

28


Merton R. (1973). An Intertemporal Capital Asset Pricing Model, Econometrica, 41, 867-887.<br />

Merton R. (1980). On Estimating the Expected Return on The Market: An Exploratory Investigation,<br />

Journal of Financial Economics, 8, 323-362.<br />

Munk, C. Sorensen, C., 2004. Optimal Consumption and Investment Strategies with Stochastic<br />

Interest Rates, Journal of Banking and Finance 28, 1987-2013.<br />

Rodriguez, J. 2002. Hedging Demands, Incomplete Markets, and Imperfect Information, Unpublished<br />

Working Paper, Sloan School of Management, Massachusetts Institute of Technology.<br />

Routledge, B., Seppi, D., Spatt, C., 2000. Equilibrium Forward Curves for Commodities, Journal of<br />

Finance 55, 1297-1338.<br />

Schwartz E. (1997). The Stochastic Behavior of Commodity Prices: Implications for Valuation and<br />

Hedging, Journal of Finance, 52, 923-973.<br />

Stoikov, S., Zariphopoulou T., 2005. Dynamic Asset Allocation and Consumption Choice in<br />

Incomplete Markets, Australian Economic Papers, 44, 414-454.<br />

Stulz, R., 1984. Optimum Hedging Policies, Journal of Financial and Quantitative Analysis 19, 127-<br />

140.<br />

Wachter J. (2002). Portfolio and Consumption Decisions Under Mean-Reverting Returns: An Exact<br />

Solution for Complete Markets, Journal of Financial and Quantitative Analysis, 37, 63-91.<br />

Working H. (1953). Futures Trading and Hedging, American Economic Review, 43, 1314-1343.<br />

Xia Y. (2001). Learning About Predictability: The Effect of Parameter Uncertainty on Optimal<br />

Dynamic Asset Allocation, Journal of Finance, 56, 205-246.<br />

29


Figures<br />

0.24<br />

0.22<br />

ε=0.01<br />

ε=0.001<br />

0.24<br />

0.22<br />

ε=0.01<br />

ε=0.001<br />

0.2<br />

0.2<br />

0.18<br />

0.18<br />

σ H<br />

0.16<br />

σ Hw<br />

0.16<br />

0.14<br />

0.14<br />

0.12<br />

0.12<br />

0.1<br />

0.1<br />

0 0.5 1 1.5<br />

Futures contract time to maturity (T H<br />

in years)<br />

Fig. 1a. Futures price volatility as a function of the<br />

futures contract time to maturity. This figure plots σ H as a<br />

function of T H for maturities varying from 0 to 1.5 years,<br />

and for e=0.1% (dash line) and e=1% (solid line).<br />

Parameters values are from Table 1a.<br />

0.08<br />

0 0.5 1 1.5<br />

Futures contract time to maturity (T H<br />

in years)<br />

Fig. 1b. Futures contract sensitivity to the spot price<br />

orthogonal risk as a function of the futures contract time to<br />

maturity. This figure plots σ Hw as a function of T H for<br />

maturities varying from 0 to 1.5 years for e=0.1% (dash<br />

line) and e=1% (solid line). Parameters values are from<br />

Table 1a.<br />

30


Tables<br />

Table 1a: State variables dynamics parameters for the copper market<br />

σ S λ S0 λ SX λ Sδ σ r α f 0 (t) λ r0 λ rr θ δ Κ σ δ ρ rδ ρ Sδ ρ Sr<br />

2 2 . 8 % 2 2 . 3 4 - 5 . 0 3 8 . 4 2 1 % 6 % 5 . 2 8 % 0 . 2 - 1 4 6 . 4 6 % 1 . 1 2 0 % 0 . 1 6 0 . 7 0 . 1 4<br />

Table(1a) displays the parameters values for the dynamics (1 a-c) in the case of the copper market.<br />

Table 1b: State variables values for the three market scenarios<br />

Long Term Mean Scenario Backwardation Scenario Contango Scenario<br />

X 4.498 4.59 4.39<br />

r 3% 3% 3%<br />

m 6.46% 9.5% 3.5%<br />

Table (1b) displays the state variables values for the long term mean,<br />

backwardation and contango scenarios<br />

Table 2. Covariance/variance ratios as a function of the futures contract maturity<br />

e<br />

Σ HwS /σ 2 Hw (%) Σ HB /σ 2 B (%) Σ Hwm /σ 2 Hw (%)<br />

T H (years) T H (years) T H (years)<br />

0.25 0.5 1 1.5 0.25 0.5 1 1.5 0.25 0.5 1 1.5<br />

0.1 % 115.8 132.0 166.8 206.5 -48.5 -44.5 -41.4 -41.0 72.3 82.4 104.1 128.9<br />

1 % 119.9 140.3 182.8 229.5 -48.6 -44.8 -41.8 -41.5 94.3 110.4 143.8 180.6<br />

Table 2 displays the covariance/variance ratios, Σ HwS /σ Hw 2 , Σ HB /σ B 2 and Σ Hwm /σ Hw 2 , as functions of the<br />

futures contract time to maturity varying from 0.25 to 1.5 years, and for e=0.1 % and e=1%. Parameters values<br />

are given in table 1a.<br />

Table 3. Minimum-variance proportions as a function of the futures contract maturity<br />

e<br />

mV<br />

mV<br />

π (%)<br />

π (%)<br />

π (%)<br />

mV<br />

Ht<br />

T H (years) T H (years) T H (years)<br />

0.25 0.5 1 1.5 0.25 0.5 1 1.5 0.25 0.5 1 1.5<br />

0.1% -115.8 -132.0 -166.8 -206.5 -74.3 -77.7 -91.4 -112.1 -0.4 -3.8 -17.5 -38.2<br />

1 % -119.9 -140.3 -182.8 -229.5 -77.2 -83.1 -101.2 -125.9 -3.3 -9.2 -27.3 -52.0<br />

B,<br />

Ht<br />

Bt<br />

Table 3 displays the minimum variance proportions π , π<br />

mV<br />

H<br />

mV<br />

B,H<br />

andπ as functions of the futures contract time to maturity<br />

varying from 0.25 to 1.5 years, and for e=0.1% and e=1%. Parameters values are given in table 1a.<br />

mV<br />

B<br />

31


Table 4. Mean-variance proportions as a function of the futures contract maturity<br />

e<br />

e<br />

e<br />

MV<br />

MV<br />

π (%)<br />

π (%)<br />

π (%)<br />

MV<br />

Ht<br />

T H (years) T H (years) T H (years)<br />

0.25 0.5 1 1.5 0.25 0.5 1 1.5 0.25 0.5 1 1.5<br />

Long Term Mean Market<br />

0.1 % 49.5 56.4 71.0 87.2 24.0 25.1 29.4 35.8 152.3 153.4 157.7 164.1<br />

1 % 51.3 59.9 77.5 96.4 24.9 26.8 32.4 40.0 153.2 155.1 160.7 168.3<br />

Backwardation Market<br />

0.1 % 11.3 12.9 16.3 20.0 5.5 5.7 6.7 8.2 133.8 134.0 135.0 136.5<br />

1 % 11.7 13.7 17.8 22.1 5.7 6.1 7.4 9.2 134.0 134.4 135.7 137.4<br />

Contango Market<br />

0. 1 % 97.5 111.1 139.8 171.7 47.3 49.4 57.9 70.5 175.6 177.7 186.2 198.8<br />

1 % 100.9 117.9 152.7 189.9 49.1 52.8 63.9 78.7 177.4 181.0 192.2 207.0<br />

B,<br />

Ht<br />

Bt<br />

Table 4 displays the mean-variance proportions π , π and<br />

Table 5.Investor specific bond sensitivities to state variables as a function of<br />

the investor’s horizon<br />

MV<br />

H<br />

MV<br />

B<br />

MV<br />

π<br />

B,H<br />

as a function of the futures contract<br />

time to maturity varying from 0.25 to 1.5 years for the three scenarios identified in table 1b. For each<br />

scenario e=0.1% and e=1%. Parameters values are given in table 1a and γ=3.<br />

e<br />

e<br />

e<br />

(%)<br />

γXt<br />

Ψ Ψ (%)<br />

Ψ (%)<br />

T I (years) T I (years) T I (years)<br />

0.25 0.5 1 1.5 0.25 0.5 1 1.5 0.25 0.5 1 1.5<br />

Long Term Mean Market<br />

0.1 % 7.0 12.1 19.0 23.4 -19.2 -36.6 -67.1 -93.2 -11.1 -18.0 -25.4 -28.7<br />

1 % 6.8 11.6 18.0 22.3 -19.2 -36.4 -66.9 -93.0 -10.7 -17.2 -24.0 -27.1<br />

Backwardation Market<br />

0.1 % 1.9 3.7 7.0 9.7 -17.5 -34.2 -65.1 -92.5 -3.0 -5.4 -8.9 -10.9<br />

1 % 1.8 3.5 6.6 9.2 -17.5 -34.2 -65.0 -92.4 -2.9 -5.2 -8.4 -10.3<br />

Contango Market<br />

0. 1 % 13.4 22.5 33.6 40.0 -21.4 -39.5 -69.6 -94.3 -21.1 -33.5 -45.5 -50.3<br />

1 % 12.9 21.4 31.8 38.0 -21.2 -39.2 -69.2 -93.9 -20.4 -31.9 -43.0 -47.5<br />

γrt<br />

γmt<br />

Table 5 displays the sensitivities Ψ , Ψ and<br />

X<br />

r<br />

Ψ<br />

m<br />

as a function of the futures contract time to maturity<br />

varying from 0.25 to 1.5 years for the three scenarios identified in table 1b. For each scenario e=0.1% and<br />

e=1%. Parameters values are given in table 1a and γ=3<br />

Table 6. The investor specific discount bond proportions as a function of investment<br />

horizon and futures contract maturity<br />

MPR<br />

MPR<br />

π (%)<br />

π (%)<br />

π (%)<br />

e<br />

e<br />

e<br />

MPR<br />

Ht<br />

B,<br />

Ht<br />

T H =T I (years) T H =T I (years) T H =T I (years)<br />

0.25 0.5 1 1.5 0.25 0.5 1 1.5 0.25 0.5 1 1.5<br />

Long Term Mean Market<br />

0.001 0.1 1.2 5.2 11.3 0.1 0.5 2.2 4.6 5.7 10.2 17.5 24.0<br />

0.01 -1.9 -2.7 -1.6 2.1 -0.9 -1.2 -0.7 0.9 4.6 8.3 14.4 19.9<br />

Backwardation Market<br />

0.001 0.0 0.4 2.3 5.9 0.0 0.2 1.0 2.4 3.7 7.1 13.5 19.3<br />

0.01 -0.5 -0.8 0.0 2.6 -0.2 -0.3 0.0 1.1 3.4 6.6 12.4 17.8<br />

Contango Market<br />

0.001 0.3 2.1 8.6 17.7 0.1 0.9 3.6 7.3 8.2 14.0 22.5 29.6<br />

0.01 -3.7 -5.1 -3.6 1.4 -1.8 -2.3 -1.5 0.6 6.1 10.4 16.9 22.4<br />

Bt<br />

Table 6 displays the market price of risk proportions π , π and<br />

MV<br />

H<br />

MV<br />

B<br />

MV<br />

π<br />

B,H<br />

as a function of the<br />

futures contract time to maturity and investment horizon varying from 0.25 to 1.5 years for the<br />

three scenarios identified in table 1b.T H =T I . For each scenario e=0.1% and e=1%. Parameters<br />

values are given in table 1a and γ=3.<br />

32


Liste des cahiers de recherche du PRISM – ANNEE 2010<br />

N° Titre Auteurs<br />

CR-10-01 Vers une approche constructionniste de la valorisation de<br />

l’entreprise<br />

J.-J. Pluchart,<br />

L. Barbara<br />

CR-10-02 « Tu pousses le bouchon un peu trop loin, Maurice ! » Vers un<br />

repérage des leviers publicitaires influençant les enfants.<br />

Application au domaine alimentaire<br />

P. Ezan, M. Gollety,<br />

N. Guichard,<br />

V. Nicolas-Hémar<br />

CR-10-03 De la nécessité de prendre en considération simultanément les<br />

différents contextes sociaux des enfants pour comprendre leur<br />

comportement alimentaire<br />

P. Ezan, M. Gollety,<br />

N. Guichard,<br />

V. Nicolas-Hémar<br />

CR-10-04 La valeur économique du brevet « bloquant » A. Bami, G. A. Shiri<br />

CR-10-05 Etude exploratoire sur l’effet des dimensions de la couleur des<br />

sites de marque sur les préférences des enfants<br />

H. Ben Miled-Chérif,<br />

N. Bezaz-Zeghache<br />

CR-10-06 Gestion tribale de la marque et distribution spécialisée : le cas<br />

Abercrombie & Fitch<br />

J.-F. Lemoine,<br />

O. Badot<br />

CR-10-07 Le knowledge management et la communication financière de<br />

l’entreprise : principes, leviers et mise en œuvre<br />

J.-J. Pluchart,<br />

S. Ayoub<br />

CR-10-08 The impact of social and environmental variables on S.-E Gadioux<br />

performance: a panel data application to international banks<br />

CR-10-09 Les relations entre la performance sociétale et la performance S.-E Gadioux<br />

financière des organisations : une étude empirique comparée<br />

des banques européennes et des banques non européennes<br />

CR-10-10 Le changement organisationnel des Entreprises Socialement J.-J. Pluchart,<br />

CR-10-11<br />

CR-10-12<br />

CR-10-13<br />

Responsables (ESR)<br />

L’influence de la familiarité et de l’intérêt pour la catégorie de<br />

produit sur l’innovativité et le comportement innovateur du<br />

consommateur : une approche modérée par le bouche à oreille<br />

L'impact conjoint des variables situationnelles et individuelles<br />

sur les réactions du consommateur face à un nouveau produit:<br />

d'une étude exploratoire qualitative à un essai de modélisation<br />

Coordination, engagement et RSE au cœur de la quête<br />

managériale du changement perpétuel<br />

D. Gnanzou<br />

A. Sellami<br />

A. Sellami<br />

O. Uzan,<br />

B. Condomines<br />

CR-10-14 De l’homo métis à l’homo academicus J.-J. Pluchart<br />

CR-10-15<br />

L’opérationnalisation du succès de carrière : intérêts et limites<br />

E. Hennequin<br />

des méthodologies actuelle<br />

CR-10-16 Quelle réussite professionnelle pour les enseignantschercheurs<br />

E. Hennequin<br />

?<br />

CR-10-17 La confiance institutionnelle en question J.-J. Pluchart<br />

CR-10-18 L’impact des mécanismes internes de gouvernement de H. Ben Ayedl’entreprise<br />

sur la qualité de l’information comptable<br />

Koubaa<br />

CR-10-19 La gouvernance des entreprises socialement responsables J.-J. Pluchart<br />

CR-10-20 Les mécanismes de régulation du marché du coaching : les H. Cloet<br />

conventions<br />

CR-10-21 How R&D Competition affects Investment Choices T. Lafay, C. Maximin<br />

CR-10-22 Les stratégies d’innovation dans le commerce indépendant de J.-F. Lemoine<br />

proximité<br />

CR-10-23 Optimal dynamic demands in commodity futures markets with a<br />

stochastic convenience yield<br />

C.Mellios<br />

P. Six<br />

CR-10-24 Comment le site internet d’une enseigne modifie le R. Vanheems<br />

comportement de ses clients en magasin<br />

CR-10-25 Les violences familiales : une problématique organisationnelle ? E. Hennequin<br />

N. Wielhorski<br />

CR-10-26 Etat de l’art sur l’harmonisation vie privée – vie professionnelle<br />

des salariés<br />

S . Kilic


CR-10-27<br />

CR-10-28<br />

Les mécanismes de régularisation du marché du coaching : les<br />

conventions<br />

La diffusion de la fraude en entreprise, le cas de la collusion<br />

tacite<br />

H. Cloet<br />

P. Jacquinot<br />

A. Pellissier-Tanon<br />

S. Strtak<br />

JJ. Lemoine<br />

CR-10-29 Agent virtuel et confiance des internautes vis-à-vis d’un site<br />

Web<br />

CR-10-30 Peut-on encore lancer un programme de recherche en<br />

management stratégique ?<br />

CR-10-31 Equivalent Risky Allocation S. Plunus<br />

R. Gillet<br />

G. Hubner<br />

CR-10-32<br />

CR-10-33<br />

CR-10-34<br />

Voluntary financial disclosure, introduction of IFRS and the<br />

setting of a communication policy: An empirical test on SBF<br />

French firms using a publication score<br />

RSE et développement des populations pauvres du sud : Une<br />

analyse relative aux OMD comme cadre d’analyse<br />

macroéconomique du développement.<br />

The Traditional Hedging Model Revisited with a Non Observable<br />

Convenience Yied<br />

JJ. Notebaert<br />

S. Edouard<br />

A. Gratacap<br />

H. De La Bruslerie<br />

H. Gabteni<br />

D. Gnanzou<br />

C.Mellios<br />

P. Six

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!