Correlations of Bond Yields

Embed Size (px)

Citation preview

  • 8/8/2019 Correlations of Bond Yields

    1/40

    Comments Welcome

    Stochastic Volatilities and Correlations of

    Bond Yields

    Bing Han

    Current Version: October 2004

    Han is with the Fisher College of Business at the Ohio State University. I thank Martin Dierker,Mark Grinblatt, Jean Helwege, Jason Hsu, Jingzhi Huang, Andrew Karolyi, Francis Longstaff, MonikaPiazzesi, Pedro Santa-Clara, Kenneth Singleton and seminar participants at the Ohio State Universityfor helpful comments and suggestions. Financial support from the Dice Center for Financial Economicsat the Ohio State University is grateful acknowledged.

    Corresponding Address: Bing Han, Department of Finance; The Ohio State University; 700D FisherHall; 2100 Neil Avenue; Columbus, OH 43210-1144. Phone: (614) 292-1875. E-mail: [email protected];

  • 8/8/2019 Correlations of Bond Yields

    2/40

    Abstract

    I develop an interest rate model with separate factors driving innovations in bondyields and their volatilities. My model features flexible and tractable affine structure

    for the covariances of bond yields. Maximum likelihood estimation of the modelwith panel data on swaptions implies pricing errors that are almost always lowerthan half of the bid-ask spread. Further, market prices of interest rate caps do notdeviate significantly from their no-arbitrage values implied by the swaptions undermy model. This supports conjectures by Collin-Dufresne and Goldstein (2003), Daiand Singleton (2003), and Jagannathan, Kaplin and Sun (2003). I also extractmarkets view on the term structure of interest rate volatility and dynamics of bondcovariances.

  • 8/8/2019 Correlations of Bond Yields

    3/40

    1 Introduction

    The turmoil in the bond markets and increased interest rate volatility since the 1980s

    have provided a boost for the rapid growth of hedging vehicles such as interest rate caps

    and swaptions. As these interest-rate derivatives become more liquid, researchers startusing their prices to evaluate term-structure models. There is a rich cross-section of

    swaptions and caps. Their prices are sensitive to both volatilities and correlations of the

    bond yields and contain valuable information regarding markets view on the evolution

    of the yield curve that is not available from interest rate data such as Treasury bonds,

    Eurodollar futures or swap rates.

    Recent studies find that it is challenging to explain market prices of swaptions and

    caps under many popular term-structure models, even for models that by construction

    fit the bond prices exactly. For example, Longstaff, Santa-Clara and Schwartz (2001)

    calibrate string-market models to the swaption data. They find that short-dated and

    long-dated swaptions tend to be priced inconsistently, and cap prices periodically deviate

    significantly from the no-arbitrage values implied by the swaptions. Other studies that

    document large and systematic pricing errors for swaptions and caps include Driessen,

    Klaassen and Melenberg (2003) and Fan, Gupta and Ritchken (2001) for multi-factor

    HJM models, and De Jong, Driessen and Pelsser (1999), as well as Hull and White (1999)

    for the Libor and swap market models.

    For the models studied in these papers, when model parameters are held constant, the

    covariances of interest rates are either deterministic or depend at most on the underlying

    interest rates. In the empirical work, models are recalibrated each date in order to capture

    time variation in the covariances. While the recalibrated models yield significantly better

    fit to the data compared to the case when model parameters are held constant (see, e.g.,

    Driessen, Klaassen and Melenberg 2003), it is not enough to continuously recalibrate

    simplistic models.

    The affine framework of Duffie and Kan (1996) accommodates stochastic covariances

    of bond yields, but may imply too strong restrictions on the covariances structure. Dai

    and Singleton (2000) find that for the affine models to be admissible, there is an important

    trade-off between flexibility in modeling the factor volatilities and correlations. In addi-

    tion, the covariances of bond yields are affine functions of the state variables governing

    bond yields in most affine models. Thus, risk factors that drive bond covariances gener-

    ally can be hedged by a portfolio consisting solely of bonds. However, Collin-Dufresne

    and Goldstein (2002) and Heidari and Wu (2001) show that interest-rate option market

    exhibits risk factors unspanned by, or independent of, the underlying yield curve.

    Empirically, Jagannathan, Kaplin and Sun (2003) find that multi-factor CIR models

    1

  • 8/8/2019 Correlations of Bond Yields

    4/40

    lead to pricing errors for caps and swaptions that are very large relative to the bid-

    ask spread, although the fit to the swap rates are very good. They suggest that the

    inability of the multi-factor CIR model to value caps and swaptions correctly may be

    due to the restrictions it imposes on the nature of stochastic volatility, and it may be

    necessary to consider models outside the affine class that are flexible in accommodatingstochastic volatility of more general forms. Collin-Dufresne and Goldstein (2003) and Dai

    and Singleton (2003) conjecture that the ultimate resolution of the swaptions and caps

    valuation puzzle may require time varying correlations and possibly factors affecting the

    volatility of yields that do not affect bond prices.

    In this paper, I develop a term structure model with properties in the conjectures

    above. My model incorporates stochastic volatilities and correlations of bond yields.

    The factors driving the covariances of bond yields are independent of the yield factors.

    Empirically, I find that market prices of interest rate caps do not deviate significantly

    from their no-arbitrage values implied from the swaptions under my model. This supportsthe conjecture by Collin-Dufresne and Goldstein (2003), Dai and Singleton (2003) and

    Jagannathan, Kaplin and Sun (2003). I am also the first to extract markets view on the

    term structure of interest rate volatility and dynamics of bond covariances from interest

    rate derivatives. Such information is valuable for risk management and valuation of exotic

    interest rate derivatives.

    The rest of this paper is organized as follows. Section 2 introduces interest rate caps

    and swaptions and the market convention for their valuations. In section 3, I develop the

    model and derive closed-form pricing formulas for European swaptions and caps. The

    main feature of the swaptions and caps data are discussed in section 4. Section 5 presents

    the econometric method used to estimate the model and section 6 presents the empirical

    results. Section 7 concludes the paper.

    2 Swaptions and Interest Rate Caps

    The swaptions studied in this paper are written on semi-annually settled interest rate

    swaps. An interest rate swap can be viewed as an agreement to exchange a fixed rate

    bond for a floating rate bond. Every six months till the maturity of the swap, one

    counterparty receives a fixed annuity and makes a floating payment tied to the six-monthLibor rate to the other counterparty of the swap contract. The coupon rate on the fixed

    leg of a swap, also known as the swap rate, is set so that the present value of the fixed

    and floating legs are equal at the start of a swap.

    Fix two dates T > . A European style by T (or into T ) receivers swaptionis a single option giving its holder the right, but not the obligation, to enter into a T

    2

  • 8/8/2019 Correlations of Bond Yields

    5/40

    year interest rate swap at date and receive semi-annual fixed payments at a pre-agreed

    coupon rate between date and T. Let D(t, T) denote the time t price of a discount

    Libor bond that matures at time T. Then at t < , the value of a forward swap that

    starts at and matures at T with a coupon rate c is given by

    V(t , , T, c) =c

    2

    2(T)i=1

    D(t, i) + D(t, T) D(t, ) (1)

    where i = +i2

    years. The sum of the first two terms is the present value of the fixed leg

    of the swap, and the third term is the present value of the floating leg. It follows that the

    payoff to the holder of a European style by T receivers swaption at its maturity date

    is given by

    Max(V( , , T , c), 0) = Max c2

    2(T)

    i=1

    D(, i) + D(, T)

    1, 0 (2)

    Thus, a swaption is an option on a portfolio of discount bonds.

    A European style by T swaption is said to be at-the-money-forward when the coupon

    rate c equals the corresponding forward swap rate F SR(0, , T )

    F SR(0, , T ) = 2

    D(0, ) D(0, T)2(T)

    i=1 D(0, i)

    (3)

    There is a second type of swaption called payers swaption where the option holder has

    the right to enter into a swap and pay fixed. For the at-the-money-forward swaptions, a by T payer swaption is worth the same as a by T receivers swaption.1

    European type at-the-money-forward swaptions are actively traded in the over-the-

    counter derivatives market, and quoted in terms of implied volatilities relative to the Black

    (1976) model as applied to the corresponding forward swap rate. The market price for a

    by T at-the-money-forward swaption is obtained by plugging the quoted Black implied

    volatility into the following formula

    (D(0, )

    D(0, T))N

    (

    2

    )

    N(

    2

    ) (4)where N() is the cumulative density function of a standard normal random variable.

    1Standard no-arbitrage arguments show that the value of a forward swap must equal the value ofthe corresponding receivers swaption minus the value of the payers swaption. By equation (1) and thedefinition of forward swap rate (3), the date-0 value of a forward starting swap whose strike equals thecorresponding forward swap rate is zero.

    3

  • 8/8/2019 Correlations of Bond Yields

    6/40

    An interest rate cap provides insurance against the rate of interest on a floating rate

    loan rising above the cap rate and gives its holder a series of European call options or

    caplets on the underlying Libor rates. Each caplet has the same strike price as the others,

    but a different expiration date. For example, a T-year cap on the six-month Libor rate

    consists of 2T1 caplets:2

    the first caplet matures in one year, and the last caplet maturesin T years. Let ti =i2

    years, ai be the actual number of days between ti and ti+1 and Lidenote the six-month Libor rate applicable to period [ti, ti+1]. Then the cash flow on the

    caplet maturing at ti+1 isai360

    max(0, K Li). Alternatively, each caplet can be viewed asa put option on a Libor discount bond. An interest rate cap is thus a portfolio of options

    on discount bonds.

    For date t < < T, let F(t , , T ) denote the Libor forward rate at time t applicable to

    the period from to T. Since Li = F(ti, ti, ti+1), the ith caplet is an option on the forward

    rate F(t, ti, ti+1). Assuming that each forward Libor rates F(ti, ti, ti+1) is lognormal with

    known volatility i, the Black model price of a cap with cap rate R is2T1i=1

    ai360

    D(0, ti+1)

    F(0, ti, ti+1)N(di) RN(di i

    ti)

    (5)

    where

    di =ln(F(0, ti, ti+1)/R) +

    2i ti/2

    i

    ti

    The market convention is to quote the price of a cap in terms of an implied volatility ,

    which is constant across caplets, so that the Black model price above at i = equals the

    market price of the cap. A T-year cap is said to be at-the-money if the cap rate R equalsthe current T-year swap rate. Only at-the-money European-type interest rate caps and

    swaptions are actively traded in the over-the-counter market. These are the data I use in

    the empirical study.

    Although swaptions and interest rate caps are traded as separate products, they are

    linked by no-arbitrage relations through the correlation structure of bond yields. The

    relative valuation of caps and swaptions is a decreasing function of bond correlations. It

    is important to note that the relative valuation of swaptions and caps can only be judged

    under a term-structure model, and not by a simple comparison of their Black implied

    volatilities quoted in the market.3

    2Note that although the cashflow of this caplet is paid at time ti+1, the applicable Libor rate Li isdetermined at ti. For this reason, the cashflow for the first caplet maturing in six months is non-stochasticand thus omitted by market convention.

    3The reason is that the Black implied volatilities apply to different underlying interest rates thatare assumed to be lognormally distributed: forward swap rates in the case of swaptions and forwardLibor rates in the case of interest rate caps. However, each forward swap rate is approximately a linearcombination of the underlying forward Libor rates. Thus, forward swap rates and forward Libor rates

    4

  • 8/8/2019 Correlations of Bond Yields

    7/40

    3 The Valuation Framework

    With empirical work in mind, I now develop a term structure model with flexible and

    intuitive specifications for the stochastic volatilities and correlations of bond yields. One

    notable feature of my model is that bond covariances are affine in a set of volatilitystate variables. My model also captures the empirical evidences of unspanned stochastic

    volatility (USV) documented in Collin-Dufresne and Goldstein (2002) and Heidari and

    Wu (2001). In my model, there are separate factors driving the innovations in bond yields

    and covariances of bond yields. Thus, my modelling framework is reminiscent of the large

    literature on stochastic volatility (e.g., Heston 1993) that specifies the joint dynamics of

    a traded asset and its volatility.

    3.1 Model

    Assume all bonds are traded. I directly model the risk-neutral dynamics of bond prices.

    The risk-neutral drifts of bond prices are determined by the no-arbitrage condition that the

    expected rate of return on all bonds equals the spot rate under the risk-neutral measure.

    The focus of my model is bond covariances. I show in section 3.2 that once bond prices

    are given, prices of swaptions and interest rate caps are determined only by the dynamics

    of bond covariances.

    By Girsanovs theorem, the instantaneous bond covariances are invariant with respect

    to an equivalent change of probability measure. Thus, I can utilize information contained

    in the historical estimates of bond covariances. It is well known that most of the observed

    variation in historical bond prices are explained by a few common factors (e.g., Litterman

    and Scheinkman 1991, and Dai and Singleton 2000). The explanatory power of these

    factors are stable and the factor loadings show a persistent pattern. However, Bliss

    (1997) and Perignon and Villa (2003) find significant time-variation in the variances of

    the common factors.

    Motivated by these findings, I assume that the yield curve is driven by N common

    factors with time-invariant weights but possibly stochastic volatilies. The risk-neutral

    dynamics of the discount bond prices are expressed as

    dD(t, T)D(t, T)

    = rtdt Nk=1

    Bk(T t)

    k(t) dZk(t) (6)

    where rt is the instantaneous short rate. For each k = 1, , N, dZk is a Brownian motionterm representing shocks to the kth factor driving the yield curve, and k is the variance

    can not be simultaneously lognormally distributed.

    5

  • 8/8/2019 Correlations of Bond Yields

    8/40

    of the kth yield factor. Without loss of generality, the yield factors are orthogonal to each

    other. Bk(T t) describes the loadings of the bond with maturity T on the kth yieldfactor at time t. It is a deterministic function of the time-to-maturity T t only. Thisensures that the term structure dynamics under my model is time-homogeneous.

    It follows from equation (6) and the Itos lemma that the date t instantaneous covari-ances of log-bond prices for a set of bonds with maturity 1, , n can be written as aproduct of three matrices

    Bt Diag(t) Bt

    where Bt is a n by N matrix whose (i, k)th element is Bk(i t), and Diag(t) is a N byN diagonal matrix whose diagonal elements are the instantaneous variances of the yield

    factors k(t), k = 1, , N. The covariance matrix above is guaranteed to be positivesemi-definite. Further, the covariance between any two discount bonds is linear in the

    variances of the N yield factors 1(t), , N(t).There are two sources for the time-variation in bond covariances under my model.

    One is predictable: as time pass by, the time-to-maturities of bonds decrease, and hence

    their loadings on the yield factors change correspondingly. Another source of movement

    in the covariances of bond yields is the stochastic volatilities of the yield factors, which I

    model next.

    Consider the general case when K of the N yield factors (K N) display stochasticvariances, and the remaining N K factors have constant variances. Let IK denote theindex of yield factors with stochastic variances. For i / IK, denote the variance of the ithfactor by a positive constant i. For each i

    IK, I model the variance of the ith factor

    as a square root process:

    di(t) = i(i i(t))dt + i

    i(t) dWit, i IK (7)

    Model parameters i and i are respectively the mean reversion speed and the long run

    mean for the variance of the ith yield factor.

    Ball and Torous (1999), Chen and Scott (2001), and Heidari and Wu (2001) find that

    innovations in interest rate levels are almost uncorrelated correlated with innovations in

    the volatility of interest rates. Motivated by this finding, I assume that each Brownian

    motion dWit driving the variances of the yield factors is uncorrelated with the dZs inequation (6). This assumption implies that bond innovations are not contemporaneously

    affected by volatility innovations. Therefore, bonds cannot be used to hedge volatility risk

    instantaneously and thus my model exhibits unspanned stochastic volatility in the sense

    Collin-Dufresne and Goldstein (2002). It also implies that dynamics for the variances of

    the yield factors are the same under the risk neutral measure and all forward measures

    6

  • 8/8/2019 Correlations of Bond Yields

    9/40

    (see Goldstein, 2000).

    In the rest of the paper, for a pair of integers K N, I denote by GAN,K the modelwith risk-neutral evolution of bond prices given by equations (6) and (7). Under the

    GAN,K, there are N factors that drive innovations in the yield curve, the first K of which

    display unspanned stochastic volatility, and the others have constant volatility.4 Theconditional covariances of bond yields are affine in the instantaneous variances of the first

    K yield factors.

    Risk-neutral bond prices dynamics are enough for deriving prices of swaptions and

    interest rate caps (see section 3.2). For the purpose of carrying out a maximum likelihood

    estimation of my model, I also need the dynamics of the volatility state variables under

    the physical measure. The volatility risk premium can be modeled as i

    i (see, e.g.,

    Heston (1993) for an equilibrium justification based on the representative agents utility).

    Under this assumption, i also follows an affine process under the physical measure:

    di(t) = i(i i(t))dt + i

    i(t) dWit (8)

    where i = ii, i = iiii , and dWit is a standard Brownian motion under the physicalmeasure.

    3.2 Model Valuation of Swaptions and Caps

    Here I derive closed-form formulas for the prices of European swaptions and caps. My

    models tractability greatly facilitates its econometric estimation with panel data of swap-

    tions to be carried out later in this study.

    Interest rate derivatives can be valued by taking expectation of their discounted payoffs

    under the spot risk-neutral measure. Unlike the equity derivatives, one can not assume

    a constant short rate, since the entire term structure is stochastic. In general, the short

    rate process is non-Markovian under the HJM type models, and thus difficult to work

    with.5 As the payoffs of swaptions and interest rate caps are homogeneous of degree 1 in

    4As in principal component analysis, I label the factors so that for i < j, the ith factor has higherunconditional variance than the j factor. Theoretically, for a given pair (N, K), IK = {1, , K} is onlyone of the possible choices for the index of yield factors which display stochastic volatility. Empirically,

    this choice fits the swaptions and interest caps data the best.5It can be shown that under my GAN,K model, the risk-neutral dynamics for the short rate r is

    rt = y(0, t) +

    t0

    c(t s)V(s)ts

    0

    c()d

    ds +

    t0

    c(t s)

    V(s) dZQs

    where B(T t) = [B1(T t) BN(T t)]; V(t) is N by N diagonal matrix whose ith diagonal elementis i(t); dZ

    Qt is a N dimensional Brownian motion stacking dZ

    Qk (t), k = 1, , N; c is a row vector of

    7

  • 8/8/2019 Correlations of Bond Yields

    10/40

    discount bond prices, it turns out that it is more convenient to use the forward risk-neutral

    measure defined by Jamshidian (1987) to value swaptions and caps.

    Let D(t , , T ) denote the date-t price of a forward contract to buy at date a bond

    that matures at T. In the absence of arbitrage, the forward bond prices are related to the

    discount bond prices by D(t , , T ) = D(t, T)/D(t, ). The forward risk-neutral measurecorresponding to date and denoted by Q uses the Libor discount bond that matures

    at as the numeraire asset. By definition, each forward bond price D(t , , T ), T > ,

    follows a martingale under Q. Thus, the dynamics of the forward bond prices under the

    corresponding forward risk-neutral measure are determined by their covariances.

    Below, I first show that given the initial bond prices, price of an European swap-

    tion is determined by the covariances of a set of discrete-tenor forward bonds under the

    corresponding forward risk-neutral measure. Then I explain how to operationalize the

    modeling assumption in section 3.1 for the covariances of forward bonds. Finally, I derive

    closed-form formulas for the prices of swaptions and caps.It is straight-forward to show that the payoff of a by T at-the-money-forward re-

    ceivers swaption, previously given in equation (2) (with coupon rate c being equal to the

    corresponding forward swap rate F SR(0, , T ) given in equation (3)), can be rewritten in

    terms of the forward bond prices as

    Max

    A() 1, 0

    (9)

    where

    A(t) =2(T)j=1

    jSj(t), Sj(t) = D(t , , j)

    D(0, , j), j = + j

    2

    j =jD(0, , j)2(T)

    k=1 kD(0, , k)(10)

    j =Kc

    2, for j = 1, . . . , 2(T ) 1, and 2(T) = (1 + c

    2)W

    W =1

    1 + (T )cNote that the js are positive constants and j j = 1.

    The no-arbitrage price of a contingent claim, which settles at time , is given by first

    taking the expectation of its payoff under the forward risk-neutral measure, and then

    multiplying it by D(0, ) (see, e.g., Lemma 13.2.3 of Musiela and Rutkowski, 1997). It

    functions defined to be the derivative of B: c() = B()

    ; and y(0, t) = log D(0, t)/t.

    8

  • 8/8/2019 Correlations of Bond Yields

    11/40

    follows that the date-0 price of a by T at-the-money-forward receivers swaption is:

    P(, T) = D(0, ) EQMax

    A() 1, 0

    (11)

    where Q is the forward risk-neutral measure. Thus, the valuation of swaption is reduced

    to computing the expectation under the forward risk-neutral measure of an arithmetic

    sum of a set of random variables Sj(), for j = 1, . . . , 2(T ). Since each Sj(t) is justa constant multiple of D(t , , j), its drift under Q

    is zero. Therefore, the covariances of

    {D(t , , j)}, j = + j2 , j = 1, . . . , 2(T ), determines their joint distributions underthe forward risk-neutral measure Q, which in turns determines the price of a by T

    European swaption by equation (11).

    Note that to value swaptions and caps, I need the covariances of forward bonds with

    fixed maturities, but it is more convienient to first model the covariances of bonds with

    fixed time to maturity (e.g., multiples of six-months).6 Let t be the date t instantaneous

    covariance matrix of changes in the logarithm of the six-month forward Libor bond prices{D(t, t + ti, t + ti+1)}19i=0, where ti = i2 years. Let H be the corresponding unconditionalcovariance matrix. Since H is a semi-definite symmetric matrix, it can be decomposed as

    H = U0U

    , where 0 is a diagonal matrix whose diagonal elements are the eigenvalues

    of H, and columns of U are the corresponding eigenvectors.7 The conditional covariance

    matrix under the GAN,K model is given by

    t = UtU (12)

    where t

    is a diagonal matrix whose first N main diagonal elements are the conditional

    variances of the N yield factors. Among them, the first K yield factors have stochastic

    variances which I model as CIR processes as in equation (7). The remaining NK factorshave constant variances. All other elements of t are zero.

    The covariances of forward bonds with fixed maturities can be obtained from t in

    two steps. First,

    log(D(t , , j)) = log(D(t , , 1)) + log(D(t, 1, 2)) + + log(D(t, j1, j))6Note that the swap rates, swaptions and caps data are all refreshed each date, in the sense that they

    have constant time-to-maturities rather than fixed maturity. The longest total maturity of the swaptions

    in my dataset is ten years. Therefore I consider bonds with time-to-maturity no more than ten years.7I normalize U to have unit length for each column. The ith column of U and the ith main diagonal

    element of 0 are the weights and the unconditional variances of the ith common factor driving the yieldcurve. Further, a rotation of these factors (obtained by multiplying a orthogonal matrix) does not changethe factor variances (is). Since all structural parameters in my model are for the dynamics of the is,they are properly identified.

    9

  • 8/8/2019 Correlations of Bond Yields

    12/40

    Second, every six months from date 0 till , j ts are multiples of six-month. On thesedates, the instantaneous covariances of the bonds on the right hand side of last equation

    can be read off from t = {cij(t)}ij. On other dates, I linearly interpolate the covariancesto preserve the continuity of the covariances as functions of time-to-maturity. 8 More

    precisely, fix integer i, j such that 1 i < j < 20; at any time t ti =i

    2 , let k be theinteger such that tk t < tk+1. Then I assume

    Cov(D(t, ti, ti+1), D(t, tj, tj+1)) = (12(ttk)) cik,jk(t)+2(ttk) cik1,jk1(t) (13)

    Now I continue with the valuation of an European by T swaption given by equa-

    tion (11). Based on the assumption that the volatility state variables are instantaneously

    uncorrelated with innovations in the yield curve, I can use an argument similar to Hull

    and White (1987) to show that each Sj() is lognormal conditional on average values of

    k (k = 1,

    , N) between date 0 and . A(), an arithmetic sum of

    {Sj()

    }j, is not con-

    ditionally lognormal. However, the corresponding geometric sum, G(t) =

    2(T)j=1 S

    jj (t),

    is. Upon replacing the difference of A() and G() by the mean of the difference,9 the

    price of an by T swaption approximately equals:

    P(, T) = D(0, ) EQMax

    G() g, 0

    (14)

    where g = 1 + EQ

    [G() A]. I apply the law of iterative expectation and derive thefollowing approximate closed-form formula for prices of at-the-money-forward European

    swaptions (see the appendix for details):

    Proposition 1 Under the GAN,K model, the price of a by T European at-the-money-

    forward swaption is given approximately by:

    P(, T) = D(0, )

    2N

    1

    2

    1)

    (15)

    =1

    21l=0

    AlUDiag

    l1U

    A

    l + Al+1UDiagl2U

    A

    l+1 AlUDiagl2U

    A

    l

    (16)

    where N() is the cumulative density function of a standard normal random variable. Theweights j, j = 1,

    , 2(T

    ), as given by (10), depend on , T as well as the current

    8I experiment with alternative interpolating schemes using functions that are exponentially decayingin time-to-maturity, and find that the interpolation scheme has a negligible influence on the prices ofswaptions and caps.

    9This approximation technique has been applied to price Asian options and basket options. It is knownto be very accurate (see, e.g., Vorst (1992)). In my case, simulations show that the typical approximationerror is less than 0.1%, even for maturity as long as = 10, and for values of the volatility state variablesthat generate bond yield volatilities that are twice as high as those observed in the data.

    10

  • 8/8/2019 Correlations of Bond Yields

    13/40

    prices of discount bonds. U is the matrix of eigenvectors of the unconditional covariance

    matrix of changes in the logarithm of the six-month forward Libor bonds. For each non-

    negative integer l 21, Diagl1 and Diagl2 are 20-by-20 matrices whose entries are zeroexcept the first N diagonal elements

    Diagl1(i, i) =1

    2i +

    eil/2 ei(l+1)/2i

    (i(0) i), i = 1, , K

    Diagl1(i, i) =1

    2i, i = K+ 1, , N

    Diagl2(i, i) =1

    4i + e

    il/2

    2

    2i e

    i2

    i 2

    2ie

    i2

    (i(0) i), i = 1, , K

    Diagl2(i, i) =1

    4i, i = K+ 1, , N

    For i = 1,

    , K, i(0) denotes the spot variance of the ith common yield factor. For each

    l, Al is a 2(T )-by-20 matrix. The 2 l + 1 to 2T l column of Al forms a lowertriangular matrix whose elements on and below diagonal are 1. All other elements of Alare zero.

    The valuation of an interest rate cap is easier. The price of a cap is the sum of the

    prices of its constituent caplets. Each caplet can be viewed as a put options on the

    corresponding forward bond.

    Proposition 2 The price of an at-the-money cap on the six-month Libor rate with time

    to maturity T is:

    2T1i=1

    D(0, ti)N(d2i) (1 + 12

    R)D(0, ti+1)N(d1i) (17)

    d1i =log(D(0, ti, ti+1)(1 +

    12

    R)) + 12

    2i ti2i ti

    and d2i = d1i

    2i ti

    where ti =i2

    year, the cap rate R equals the T year swap rate andi2 is the average expected

    variance over [0, ti] of changes in the logarithm of forward bond price D(t, ti, ti+1), given

    by equation (16) with = ti and T

    = 1

    2.

    3.3 Relation to Other Term-Structure Models

    I now discuss how my model relates to models in previous and contemporaneous studies.

    First, the dynamics of discount bond prices under my model as given by equation (6) can

    be derived from the affine model of Duffie and Kan (1996). To see this, consider an affine

    11

  • 8/8/2019 Correlations of Bond Yields

    14/40

    model with N factors driving the short rate:

    rt = 0 + 1Xt

    The risk-neutral dynamics of the risk factors Xt are given by

    dXt = ( Xt)dt +

    S(t)dZQt (18)

    where and are N by N matrices, and S is a diagonal matrix with components

    Sii(t) = i + iX(t)

    The discount bond prices are given by

    D(t, T) = eA(Tt)B(Tt)X(t) (19)

    where A() and B() can be solved from a system of ODEs with A(0) = 0, B(0) = 0.

    By equation (18), equation (19) and the Itos lemma, the risk-neutral dynamics for the

    discount bond prices under the affine model are

    dD(t, T)

    D(t, T)= rtdt B(T t)

    S(t) dZQt (20)

    which is of the same shape as bond prices dynamics under my model.

    Just like under the affine model, the covariances of bond yields in my model are

    affine in the volatility state variables. The difference is that in my model, innovations inbond yields are not contemporaneously affected by volatility innovations. In contrast, the

    stochastic volatility factors in the affine framework typically enter into bond prices, and

    can be represented as linear combinations of bond yields. In other words, the volatility

    factors in the traditional affine framework drive both the cross-sectional differences in bond

    yields and changes in conditional volatility. My model and affine models with unspanned

    stochastic volatility both break this dual role of stochastic volatility.10 Thus, my model can

    be viewed as a reduced-form representation of affine model with unspanned stochastic

    volatility.

    My model takes the current term structure of interest rates as given, and focus onunderstanding what drives stochastic covariances of bond yields and the valuation of

    10See Collin-Dufresne and Goldstein (2002) for examples of affine models that display unspannedstochastic volatility. Collin-Dufresne, Goldstein and Jones (2004) compare traditional affine models toan affine model with unspanned stochastic volatility. They show that the latter generates much morerealistic volatility estimates without sacrificing fit to the cross-sectional bond prices.

    12

  • 8/8/2019 Correlations of Bond Yields

    15/40

    interest rate derivatives. This aspect of my model is similar to the HJM (1992) framework

    or its generalization such as the string or random field framework (see, e.g., Goldstein 2000,

    and Santa-Clara and Sornette, 2001). If a model fits the bond prices with errors, such

    errors will get carried over to the interest rate derivatives and confound the inferences

    about the factors driving the bond covariances.My model is closely related to the string-market model of Longstaff, Santa-Clara

    and Schwartz (2001) who also directly model the risk-neutral dynamics of discount bond

    prices. The two models share the property that the instantaneous covariance matrix of

    bond yields and the unconditional historical covariance matrix are diagonalized by the

    same matrix. The difference is that Longstaff, Santa-Clara and Schwartz (2001) consider

    only constant covariances while I explicitly model the dynamics of the risk factors driving

    the stochastic bond covariances.

    Two contemporaneous theoretical papers contain models similar to mine. First, Kim-

    mel (2004) develops a class of random field models in which the volatility of forward ratesdepends on a set of latent variables, not the level or shape of the forward curve. The latent

    volatility variables themselves follow general diffusion process. He focuses on conditions

    necessary for the existence and uniqueness of the forward rate process, characterizing

    the behavior of the latent variable random field model and deriving theoretic results for

    derivative pricing. For tractability of empirical work, I choose bond prices rather than

    forward rates as model fundamentals, and focus on affine process for the state variables

    that drive the stochastic covariances of bond yields. These state variables also have eco-

    nomic meaning in my model: they are the instantaneous variance of common factors that

    drive the yield curve innovations. Second, my model is closely related to the generalized

    affine framework proposed by Collin-Dufresne and Goldstein (2003), which also allows

    unspanned stochastic volatility but maintains the tractability of the affine models.

    4 Data

    A panel dataset of interest rates and implied volatility for at-the-money-forward European

    type swaptions is used to estimate my model. The interest rates include six-month and

    one-year Libor rates as well as two, three, four, five, seven, ten and fifteen-year swap rates.

    The total maturity of the swaptions I use is no greater than ten years and there are 34such swaptions. The maturities of the swaptions range from six months to five years and

    the tenors of the underlying swaps are between one and seven years. Implied volatilities

    for Libor interest-rate caps of maturity two, three, four, five, seven and ten years are used

    in the study of relative valuation of swaptions and caps. All data are collected from the

    Bloomberg system and represent the average of best bid and ask among many large swap

    13

  • 8/8/2019 Correlations of Bond Yields

    16/40

    and swap derivatives brokers. There are 220 weekly observations for each series, sampled

    every Friday from January 24, 1997 to April 6, 2001.

    Discount bond prices (with maturity that are multiples of six-months) in the Libor-

    swap market are needed in order to compute the market prices and model prices for

    swaptions and caps. I apply a least-square cubic spline approximation to the Libor andswap rates to get the par yield curve, and then bootstrap the discount bond prices from

    the par yield curve.11 The unconditional covariance matrix of forward bond yields is

    computed using the discount bond prices between January 17, 1992 and January 17,

    1997. Its eigenvector matrix U is used in forming instantaneous covariances as specified

    in (12).

    Table 1 reports the mean and standard deviation of swaption implied volatilities.

    On average, the implied volatility is humped as a function of swaption maturity, with a

    maximum at two years. On high volatility dates, the swaption implied volatility tends

    to decrease monotonically with option maturity. Consistent with mean reversion in theinterest rates, the implied volatility of swaption is usually monotonically decreasing as a

    function of the tenor of the underlying swap. Further, the standard deviation of swaption

    implied volatility decreases with option maturity as well as the tenor of the underlying

    swap. For example, although the implied volatilities of 0.5 into 1 and 5 into 5 swaption

    have about the same sample mean, the first is almost three times as volatile as the latter.

    Figure 1 confirms that there is a fair amount of time-series variation in the swaption

    implied volatilities, especially for short-dated swaptions. My sample period includes the

    Asian crisis, the Russian moratorium and the Long Term Capital Management crisis,

    the crash of technology stocks, as well as quiet periods of low interest rate volatility.These different volatility environments help us to estimate the dynamics of bond-yield

    covariances.

    Figure 1 also shows that implied volatilities for short-dated and long-dated swaptions

    do not always track each other closely. For example, the sample correlation between the

    implied volatility for the 0.5 into 1 swaption and that for the 5 into 5 swaption is only

    0.51. A principal component analysis also suggests there are at least two factors driving

    the innovations in the swaption implied volatilities.

    11The least-squares cubic spline approximation fits the swap rates very well, with average absolute

    fitted error about 0.76 basis points. Interpolation schemes that exactly fit observed Libor-swap ratestend to lead to unreasonably high estimates for the volatilities of long term bonds and somewhat ruggedcorrelations. An alternative method to generate reasonable estimates for the bond covariances is to putsome smoothness condition on the shape of the forward rate curve (see, e.g., Driessen, Klaassen andMelenberg 2003).

    14

  • 8/8/2019 Correlations of Bond Yields

    17/40

    5 Econometric Method

    The econometric exercise in my paper parallels many previous studies that evaluate the

    performance of short-rate based term structure models and infer dynamics of the risk

    factors driving the short rate using bond prices or swap rates. The focus of my studyis the stochastic covariances of bond yields. I infer dynamics of the risk factors driving

    bond covariances and obtain conditional estimates of bond covariances from the swaption

    implied volatilities data.12

    I adopt the maximum likelihood approach used by Chen and Scott (1993), Pearson

    and Sun (1994), Duffie and Singleton (1997), Jagannathan, Kaplin and Sun (2003) and

    others. The volatility state variables are unobservable. Thus, they need to be filtered

    at the same time as the model is estimated. For a given set of model parameters, the

    volatility state variables are inverted from the prices of swaptions that are assumed to

    be priced exactly. The percent pricing errors of the other swaptions are assumed to be

    i.i.d and jointly normal N(0, ). The estimated parameters maximize the sum of thelog-likelihoods of the exactly fitted swaptions and the percent pricing errors of the other

    swaptions.

    The estimated parameters for the volatility dynamics under the GAN,K model maxi-

    mize a log-likelihood function function log L = log L1 + log L2, where

    log L1 =T1t=1

    Ki=1

    lnf(i,t+1|i,t) T

    t=2

    ln(J,t)

    log L2 = (34

    K)(T

    1)

    2 ln(2) T

    1

    2 ln|| 1

    2

    Tt=2

    e

    t1et

    where et is a vector of the percent pricing errors of the non-exactly-fitted swaptions,

    J,t is the determinant of the matrix of first order partial derivatives of K exactly-fitted

    swaptions with respect to each of the Kvolatility factors, and f(i,t+1|i,t) is the transitiondensity of the ith volatility factor (assumed to follow independent CIR process). Omitting

    the subscript i,

    f(t+1|t) = 2ceuw

    w

    u

    q/2Iq(2

    uw)

    withc =

    2

    2(1 e(Tt)) , u = cte(Tt), w = cT, q =

    2

    2 1,

    and Iq is a qth order modified Bessel function of the first kind. Since the data are sampled

    12Collin-Dufresne, Goldstein and Jones (2004) find that interest rate volatility can not be extractedfrom the cross-section of bond prices.

    15

  • 8/8/2019 Correlations of Bond Yields

    18/40

    weekly, the above formula is applied with T t = 1/52 year.To implement the maximum likelihood approach above, I have to choose the swaptions

    that are to be fitted exactly. Each swaption provides information about the covariances

    of a segment of the yield curve and there are no good reasons to choose one swaption

    over another to be fitted exactly.13 For all results reported in this paper, I invert theK volatility state variables under the GAN,K model by assuming that K portfolios of

    swaptions (corresponding to the first K principal components of the swaption implied

    volatilities) are fitted exactly, and measurement errors apply to other principal compo-

    nents.14 This approach not only circumvents the arbitrariness of standard approach of

    fitting specific instruments exactly, it also approximately orthogonalizes the matrix of

    measurement errors.

    6 Empirical Results

    The empirical results presented below answer, among other things, the following questions:

    1) how many yield factors and how many covariance factors are needed to explain the

    swaptions data? 2) can my model explain the relative valuation of swaptions and interest

    rate caps? 3) how do the covariances of bond yields implied from the swaptions compare

    to the historical estimates based on bond prices alone?

    6.1 Number of Factors Underlying the Swaptions Data

    Recall that my GAN,K model have N yield factors, and the first K yield factors have

    stochastic volatilities. When K = 0, my model specifies a constant covariance matrix

    for bonds with fixed time-to-maturity. There are only N parameters under the GAN,0model, which are the constant variances of the N yield factors. They can be estimated

    via the General Methods of Moments approach. The pricing error of each swaption

    provides a moment condition. The test statistic for over-identifying moment restrictions

    is asymptotically distributed as 234N.

    The constant covariance model could not fit the swaptions data well. Even when there

    13For my model with only one volatility factor, I find that the parameter estimates and the impliedvolatility state variable can differ significantly depending on the choice of the swaption to be fitted exactly.This reflects the fact that my model with only one volatility factor is unable to fit simultaneously bothshort-maturity options on short-tenor swaps and long-maturity options on long-tenor swaps. My modelwith multiple factors driving the volatilities and correlations of bond yields fits the data much better andthe estimation results are not sensitive to the choices of swaptions to be fitted exactly.

    14Collin-Dufresne, Goldstein and Jones (2004) has applied a similar approach in estimating variousaffine model using bond prices.

    16

  • 8/8/2019 Correlations of Bond Yields

    19/40

    are 4 factors driving the innovations in the yield curve, the mean absolute pricing error

    is 9.86%, and the average RMSE for the swaptions is 10.21%. The constant covariance

    model tends to underprice swaptions. The test statistics for over-identifying moment

    restrictions is 129.18. Since the 99% critical value for 230 is 50.89, I can reject the

    GAN,0 model. Constant covariance model with fewer yield factors performs even worse.Increasing the number of yield factors to N > 4 does not significantly improve model fit.

    Allowing stochastic covariances greatly improves models fitting performance for swap-

    tions. Table 2 shows that under the GA4,1 model, the overall mean absolute swaption

    pricing error is reduced to 3.28%, and the average (median) RMSE for the swaptions

    is 3.97% (3.53%). The swaption RMSE under GA4,1 is always smaller than that under

    GA4,0. The difference is especially big in the months leading up to the Long Term Capital

    Management (LTCM) crisis in the fall of 1998. Between January 1998 to the end of Au-

    gust 1998, the swaption RMSE under the constant covariance GA4,0 model raises steadily

    from around 10% to over 30%, but it stays around 2.5% under the GA4,1 model.However, the GA4,1 model, with only one factor driving the stochastic covariances

    of bond yields, has difficulty fitting simultaneously both short and long dated swaptions,

    especially during the LTCM crisis. As can be seen from Figure 1, the implied volatility for

    the 0.5 into 1 swaption jumps from just above 10% before the LTCM crisis to more than

    24% during the crisis, while the implied volatility for the 5 into 1 swaption only increases

    from about 13% to 17%. In order to generate this feature of the data using models with

    one volatility factor, the mean reversion speed 1 must be high. A large 1 implies fast

    reversion to the mean. Thus, a volatility shock is expected to die out quickly, which

    implies much smaller changes in the implied volatilities for the long-dated swaptions than

    short-dated swaptions. Yet a high 1 would produce large swaption pricing errors during

    normal market when implied volatilities for short dated and long dated swaptions tend

    to move by about the same amount.

    My model with multiple covariance factors solves the problem above. The mean

    absolute pricing error for the swaptions is 2.17% under the GA4,2 model. The median

    swaption RMSE is 2.61% and the maximum is 8.00%. The GA4,3 model performs even

    better. It implies a mean absolute pricing error of 2.06%. Panel A of Figure 2 shows that

    the swaption RMSE under GA4,3 is lower than 4% except during the LTCM crisis. On

    161 of the 220 weeks, the RMSE is lower than 3% (about half of the bid-ask spread).15

    In addition to having multiple factors driving the bond covariances, it is also important

    to have a sufficient number of yield factors. For example, although GA2,2 has more

    15Note that the bid-ask spread for the swaption implied volatility is usually 1 volatility point (i.e., 1%)during my sample period. For a typical implied volatility of 16% for at-the-money-forward swaption, thistranslates into a bid-ask spread of about 6% bid-ask spread for the swaption price, since price of swaptionat-the-money-forward is almost linear in its Black implied volatility.

    17

  • 8/8/2019 Correlations of Bond Yields

    20/40

    volatility factors than GA4,1, it often produces larger pricing errors. The GA4,2 model

    significantly dominates both GA4,1 and GA2,2, as formal statistical tests presented below

    also shows.

    The swaption pricing errors obtained in the previous studies are substantially higher

    than ours, even when the models are frequently recalibrated in these studies in order tofit the data better. For example, De Jong, Driessen and Pelsser (1999) find that for a

    variety of Libor and swap market model specifications, the mean absolute pricing error

    ranges from 0.87 to 2.28 volatility points, or about 5% to 14% assuming an average im-

    plied volatility of 16%. Driessen, Klaassen and Melenberg (2003) report average absolute

    swaption pricing error between 7.23% and 9.50% for several HJM type models. Jagan-

    nathan, Kaplin and Sun (2003) report swaption pricing errors that are larger than the

    bid-ask spread. For example, under the two-factor CIR model, the mean absolute pricing

    error is about 8% for 3-month into 2-year swaption, and 31 .39% for 2-year into 5-year

    swaption.

    16

    The best prior fitting performance for the swaptions data comes from the four-factor

    string-market model of Longstaff, Santa-Clara and Schwartz (2001). The median swaption

    RMSE is 3.10%, but it becomes over 15% during the LTCM crisis. Just like my GA4,0model, their model assumes constant covariances of bond yields (with constant time-to-

    maturities). The difference is that they recalibrate their model each date which implicitly

    allows four factors to drive the covariances of bond yields. In contrast, I explicitly model

    the dynamics of the risk factors that drive the covariances of bond yields. As a result,

    my GA4,2 model, with only two factors driving the covariances of bond yields, is able

    to fit the swaption data better, both under normal market condition and during market

    crises.17

    Table 3 reports the Diebold and Mariano (1995) test statistics for pairwise comparisons

    of alternative specifications of my model. Each model specification implies a time series

    of mean squared swaption pricing errors SS E(t) corresponding to the optimized model

    parameters: SS E(t) = 134

    34i=1

    2i,t, where i,t is the percent pricing error for the ith

    swaption on date t. Consider two models that give rise to {SS E1(t)}Tt=1 and {SS E2(t)}Tt=1respectively (T = 220 in my case). The null hypothesis that the two models have the

    same pricing accuracy is equivalent to the null hypothesis that the population mean ()

    16

    Umantsev (2001) shows that the three-factor affine model A1(3) of Dai and Singleton (2000) fitsthe swaption volatilities better than the multi-factor CIR model, especially when the swaptions data areincluded in the estimation procedure. However, the pricing errors for swaptions with long option or swapmaturity are still quite large.

    17As reported earlier, over the entire sample of 220 weekly observations, the median swaption RMSEunder the GA4,2 model is 2.61% and the maximum is 8.00%. The sample in Longstaff, Santa-Clara andSchwartz (2001) is the first 128 weeks of mine. Over this subsample, the median RMSE of my GA4,2model is 2.43%.

    18

  • 8/8/2019 Correlations of Bond Yields

    21/40

    of the difference in pricing errors (dt) is zero, where dt = SS E1(t) SS E2(t). Dieboldand Mariano (1995) show that if{dt}Tt=1 is covariance stationary and short memory, thenthe asymptotic distribution of the sample mean of pricing error difference d = 1

    T

    Tt=1 dt

    is normally distributed as T(

    d ) N(0, 2fd(0))

    where fd(0) is the spectral density of the pricing error difference at frequency:

    fd(0) =1

    2

    =

    rd() where rd() = E[(dt )(dt )]

    The formula for fd(0) shows correction for serial correlation in dt. Diebold and Mariano

    (1995) obtain a consistent estimate of 2fd(0) by the sum of available autocovariances up

    to a certain truncation lag q:18

    2 fd(0) =q

    =q

    rd(), where rd() = 1T

    Tt=||+1

    (dt d)(dt|| d)

    Thus, under the null hypothesis of equal pricing accuracy across the two models, the

    following statistic

    S =

    T d

    2 fd(0)

    is distributed asymptotically as standard normal N(0, 1).Table 3 confirms my earlier results on the comparison of model fitting performance.

    Models with more yield factors and volatility factors usually fit the swaptions data signifi-

    cantly better. The largest improvement comes from increasing the number of yield factors

    from one to two, or from two to three, and from increasing the number of volatility factors

    from one to two. In each pair of such comparisons, models with more yield or volatility

    factors imply at least 1% lower swaption RMSE on at least 200 weeks out of a total of 220

    weeks. These cases also give the largest Diebold and Mariano (1995) statistics. Overall,

    the GA4,3 model with four yield factors and three separate volatility factors best fit the

    swaptions data. The GA4,2 model and the GA3,3 model also provides excellent fit to the

    swaptions data.

    Results in Table 3 are also consistent with Heidari and Wu (2001) who find by princi-pal component analysis that one needs three volatility factors that are independent of the

    underlying yield curve to capture the movement of the swaption implied volatility surface.

    However, the principal component analysis is unconditional, and completely ignores the

    18In Table 3, lag order q = 52. I verify that my results are robust to alternative lag order q such as 5,10, 26, and 102.

    19

  • 8/8/2019 Correlations of Bond Yields

    22/40

    rich information contained in the cross-section of swaptions that my stochastic-covariance

    model tries to capture. Through rigorous modeling and econometric estimation, I obtain

    estimates of conditional covariances of bond yields and infer information about the dy-

    namics of the factors that drive bond covariances. I can also address issues such as the

    relative valuation of swaptions and interest rate caps. These are impossible to achievewith principal component analysis.

    6.2 Relative Valuation of Swaptions and Caps

    The relative valuation of swaptions and interest rate caps crucially depends on the co-

    variances of bond yields, and thus provides a diagnostic test for my model. This is an

    out-of-sample test in the sense that I first estimate my model with the swaptions data,

    then compare the model implied no-arbitrage values for the interest rate caps given by

    equation (17) to their market prices. The differences are referred to as cap pricing errors.Table 4 reports the mean and mean absolute pricing errors of interest rate caps. The

    constant-covariance model, even with as many as four yield factors, still implies large

    mean absolute pricing error (10.36%). Just like the case for the swaptions, introducing

    stochastic covariances substantially improves model fit to the caps. The mean absolute

    cap pricing error is 5.41% under GA4,1, 4.88% under GA4,2, and 4.54% under the GA4,3model. The average pricing error is 0.62% under the GA4,2 model and only 0.03% under

    the GA4,3 model. Only the 2-year cap shows some systematic overvaluation by the GA4,2model. The GA4,3 model, which is favored by the statistical tests based on models fit

    to the swaptions, also implies the smallest pricing errors for the interest rate caps. Its

    improvement over the GA4,2 model is mainly reducing the pricing error of the 2-year cap.

    Panel B of Figure 2 shows that differences between the market prices of interest rate

    caps and their no-arbitrage values implied from the swaptions according to my GA4,3model are typically lower than 6% (about the bid-ask spread). There are only two periods

    when the cap RMSE is higher than 6%. The first occurs during the LTCM crisis. For

    example, the cap RMSE jumps to 19.9% on August 28, 1998 and 30.4% on September

    4, 1998, right after The Times headlined Meriwether fund plummets by 44%. For the

    next 4 weeks, the average cap pricing error stays below 6% until it shoots up again to

    25.8% on October 2, 1998. During that week, many banks revealed huge exposure to or

    investment in the hedge-fund which led to a 3.5 billion bail out of LTCM.

    The second period of relatively large cap pricing errors starts at the beginning of

    August 1999, when, according to my GA4,3 model, the no-arbitrage prices of individual

    interest rate caps implied from the swaptions are lower than the market prices of caps by

    about 8%. Supporting the view that caps are overvalued relative to swaptions during this

    20

  • 8/8/2019 Correlations of Bond Yields

    23/40

    period, various news reports indicate that proprietary desks are short cap volatility and

    long swaption volatility.19 The cap pricing errors revert to around 3% (about half of the

    bid-ask spread) by the end of 1999.

    Under my GA4,3 model, the market prices of interest rate caps do not deviate signif-

    icantly from their no-arbitrage values implied from the swaptions. In contrast, using afour-factor constant-covariance model, Longstaff, Santa-Clara and Schwartz (2001) find

    periodically large and somewhat bimodally distributed pricing errors for interest rate

    caps. The average cap pricing error is 10.38%, and the mean absolute cap pricing error is

    14.19%. Note that their model and my model both have four factors driving the innova-

    tions in the yield curve. The difference is that my model also has separate factors driving

    the covariances of bond yields. My model seems to realistically capture the dynamics

    of bond covariances and largely eliminates the large and systematic mispricing between

    swaptions and caps found by Longstaff, Santa-Clara and Schwartz (2001) and several

    other studies.

    20

    As an additional test of goodness-of-fit for my stochastic covariance model, I regress

    changes in the market prices of interest rate derivatives on the changes in the fitted prices

    under the GA4,3 model. The regression is run individually for each swaption and interest

    rate cap. If my model describes the data well, then the estimated intercept should be

    very close to 0, and the slope coefficient to be indistinguishable from 1, as well as a high

    R2 for the regression. This is what I find. None of the intercept is even close to being

    statistically different from 0. The average estimated slope coefficient is not significantly

    different from 1. The average R2 is 0.92 (0.78) across all the regressions corresponding to

    the swaptions (interest rate caps).

    6.3 The Volatility Factors and Implied Covariances

    Through the model estimation, I also extract markets view about the dynamics of the

    volatility factors and obtain implied bond covariances. Here I compare the implied volatil-

    ities and correlations of bond yields to their historical estimates (both unconditional and

    conditional). All results below are based on the estimated GA4,3 model.

    19However, this position is risky. On August 5, 1999, Dow Jones Newswire reports that two major USinvestment banks lose millions of dollars when they had to unwind big swaption positions and cover shortcap positions. Interestingly, this correlation play on the forward rates was reported to be put back onjust three weeks later.

    20For example, the smallest mean absolute pricing error for caplets among all the Libor and swapmarket models tested in De Jong, Driessen and Pelsser (1999) is about 20%. Driessen, Klaassen andMelenberg (2003) find that the average absolute pricing errors for several multi-factor HJM models areall bigger than 10%. Jagannathan, Kaplin and Sun(2003) find smallest mean absolute pricing error forcaps under one, two, or three-factor CIR models are still much larger than the bid-ask spreads.

    21

  • 8/8/2019 Correlations of Bond Yields

    24/40

    Recall that instantaneous covariances of bond yields under the GA4,3 model are linear

    in the variances of four yield factors. Plots of the factor loadings obtained from the

    principal component analysis confirm that the first three yield factors approximately

    represent the level, slope and curvature of the term structure. The fourth factor only

    affects the short end of the yield curve (less than 2-year maturity).The first volatility factor, or the variance of the level yield factor, is quite persistent.

    On the other hand, the second and the third volatility factor, or the variance of the slope

    and the curvature yield factor mean revert very fast. The half-life, defined as log2

    ,

    where is the mean reversion speed, measures the time when a mean-reverting variable is

    expected to reach a value halfway between the current level and its long-run mean. Based

    on the model parameter estimates presented in Table 5 Panel A, the half life of the first

    volatility factor is about 6 years under the physical measure, while it is only about 3 to

    4 months for the second and the third volatility factor.

    Several implications follow from the weights of the yield factors and the mean reversionspeeds of their stochastic volatility. First, the long-dated swaptions are mainly influenced

    by the first volatility factor. Second, swaptions of different maturities and tenors are about

    equally sensitive to innovations in the first volatility factor. Therefore, the first volatility

    factor proxies for the average level of the surface of swaption implied volatilities. Third,

    the second volatility factor proxies for the spread between the implied volatilities of short

    and long dated swaptions of the same tenor.21 Finally, only short-dated swaptions on

    one-year and two-year swaps are sensitive to the third volatility factor. I have confirmed

    these implications with numerical analysis.

    Panel B of Table 5 reports the sample correlations between the yield factors and thevolatility factors under the GA4,3 model. It strongly supports my models assumption

    that shocks to the yield factors are uncorrelated with their stochastic volatilities. Panel C

    of Table 5 shows that the time series average of the implied variances of the yield factors

    are very close to the unconditional variances of the yield factors. Figure 3 shows that

    the average implied covariances of bond yields are also very close to their unconditional

    historical estimates.22

    Finally, Figure 4 compares the implied covariances of forward bonds to the conditional

    covariances obtained from the Dynamic Conditional Correlation (DCC) model, a new class

    of multivariate GARCH models proposed by Engle (2002). He shows that this model

    provides a very good approximation to a variety of time varying correlation process, yet it

    21The sample correlations between implied 2 and the difference between the Black implied volatilitiesfor 0.5 into T and 5 into T year swaptions are all about 0.95, for T = 1, . . . , 5.

    22The only noticeable difference occurs for the correlations of bonds with maturity below 3 years wherethe implied correlations are lower than the historical correlations. But most of the differences are smallerthan 0.05. The maximum difference is 0.08.

    22

  • 8/8/2019 Correlations of Bond Yields

    25/40

    still has the flexibility of univariate GARCH model. I apply this model to changes in the

    logarithm of six-month forward bonds. Conditional covariances are estimated following a

    two-step procedure as described in detail by Engle and Sheppard (2001).

    Several interesting results can be observed from Figure 4. The instantaneous volatility

    of six-month forward rate implied from the swaptions closely tracks the GARCH(1,1)estimate. For example, in the case of six-month forward rate with two year maturity

    plotted in the top panel of Figure 4, the correlation between the two conditional volatility

    estimates is 0.62. There is no systematic difference between the two conditional volatility

    estimate. The bottom panel in Figure 4 plots estimates for the correlation between the

    two and five year maturity six-month forward rates. The correlation implied from the

    swaptions is more volatile than that estimated from the bond prices under the DCC

    model. The sample correlation between the two correlation series is only 0.15. During

    periods of market crises, the implied correlation drops precipitously but this is not reflected

    in the DCC estimate. The comparison of implied correlations and historical correlationsdeserves further investigation in future studies.

    7 Conclusion

    Recent studies find that it is difficult for many popular term-structure models to explain

    the valuation of swaptions and interest rate caps, although these models can price the

    underlying bonds very well or even fit them exactly by design. A potential limitation of

    these models is that the covariances of bond yields are either deterministic, or can be com-

    pletely hedged by the underlying bonds. Collin-Dufresne and Goldstein (2003), Dai andSingleton (2003) and Jagannathan, Kaplin and Sun (2003) conjecture that the resolution

    of the swaptions and caps valuation puzzle may require models that can accommodate

    more general forms of stochastic volatility and time varying correlations, such as having

    volatility factors that do not affect bond prices.

    In this paper, I develop and evaluate a term-structure model that has separate factors

    driving innovations in bond yields and their stochastic volatilities and correlations. The

    covariances of bond yields are affine in a set of volatility state variables that are not

    spanned by bonds. Empirically, my model explains the cross-section and the time-series

    variation of the swaption implied volatilities very well, and reconciles the relative valuationof swaptions and interest rate caps. These results support the conjectures by Collin-

    Dufresne and Goldstein (2003), Dai and Singleton (2003) and Jagannathan, Kaplin and

    Sun (2003).

    The modelling of factors driving stochastic covariances of bond yields and the econo-

    metric estimation using interest rate derivative prices in my study are parallel to a large

    23

  • 8/8/2019 Correlations of Bond Yields

    26/40

    literature that studies short-rate based term-structure models using bond prices. More

    empirical research is needed to better understand the rich dynamics of interest rates

    volatilities and correlations as well as the valuation of interest rate derivatives, and to

    evaluate the class of models proposed in this paper as well as in Collin-Dufresne and

    Goldstein (2003) and Kimmel (2004).

    24

  • 8/8/2019 Correlations of Bond Yields

    27/40

    Table 1: Descriptive Statistics of European Swaption Volatilities

    This table presents descriptive statistics for the mid-market Black implied volatilities for 34 at-the-money-forward European swaptions analyzed in the paper. The data are Friday close collected from theBloomberg, from January 24, 1997 to April 6, 2001. Expiration refers to the number of years till optionexpiration and Tenor refers to the maturity of the underlying swap. The last column reports the firstorder serial correlation of each swaption implied volatility series. The swaption implied volatilities areannualized and expressed in percentage points.

    Expiration Tenor Mean Median Standard Min Max SerialDeviation Correlation

    0.5 1 14.77 13.85 3.29 9.70 27.00 0.931 1 16.12 15.45 2.79 11.80 26.00 0.952 1 16.85 16.50 2.11 13.10 23.20 0.943 1 16.66 16.40 1.87 13.00 21.80 0.934 1 16.35 16.10 1.69 13.00 20.70 0.925 1 16.05 15.90 1.50 12.90 20.25 0.93

    0.5 2 15.50 14.55 3.02 10.40 27.00 0.941 2 16.10 15.50 2.48 12.20 24.00 0.942 2 16.37 16.05 1.95 13.00 22.50 0.933 2 16.12 15.90 1.70 12.90 21.00 0.934 2 15.82 15.73 1.51 12.80 20.10 0.925 2 15.51 15.50 1.37 12.70 19.75 0.92

    0.5 3 15.48 14.58 2.89 10.60 27.00 0.931 3 15.87 15.30 2.28 12.20 23.00 0.942 3 16.04 15.70 1.81 12.90 21.50 0.933 3 15.79 15.70 1.57 12.80 20.20 0.934 3 15.50 15.50 1.41 12.70 19.85 0.92

    5 3 15.18 15.20 1.29 12.60 19.50 0.91

    0.5 4 15.42 14.60 2.73 10.80 26.50 0.931 4 15.65 15.30 2.12 12.10 22.50 0.932 4 15.74 15.55 1.67 12.80 20.30 0.933 4 15.50 15.40 1.48 12.70 19.95 0.934 4 15.20 15.20 1.33 12.60 19.50 0.915 4 14.85 14.90 1.22 12.40 19.05 0.91

    0.5 5 15.36 14.60 2.60 11.00 26.50 0.921 5 15.40 15.10 2.00 11.30 21.50 0.932 5 15.46 15.30 1.57 12.70 19.80 0.923 5 15.22 15.18 1.40 12.60 19.55 0.92

    4 5 14.89 14.90 1.27 12.50 19.10 0.915 5 14.50 14.53 1.15 12.10 18.55 0.90

    0.5 7 15.18 14.50 2.45 11.00 26.00 0.921 7 15.12 14.80 1.80 12.00 20.50 0.922 7 15.07 14.90 1.49 12.40 19.45 0.923 7 14.82 14.70 1.35 12.30 19.20 0.91

    25

  • 8/8/2019 Correlations of Bond Yields

    28/40

    Table 2: Swaption Pricing Errors

    This table reports the mean absolute value of the percent pricing errors of 34 at-the-money-forwardEuropean swaptions under various model specifications. For the model corresponding to the columnlabelled by (N, K), there are N factors driving the evolution of yield curve, and the first K yield factorshave unspanned stochastic volatility. All models are estimated using a panel dataset consisting of 220weekly observations on 34 swaptions from January 24, 1997 to April 6, 2001. The percent pricing error fora swaption is the difference between its fitted model price and its market price expressed as a percentageof market price.

    Expiration Tenor Model(4,0) (4,1) (2,2) (4,2) (3,3) (4,3)

    0.5 1 13.36 7.78 12.45 4.23 3.43 3.311 1 10.56 4.22 7.46 3.32 2.74 2.842 1 9.78 3.13 3.57 3.08 3.14 3.143 1 9.63 3.14 3.39 2.93 2.91 2.794 1 9.58 3.84 4.11 3.37 3.15 3.185 1 9.44 4.48 3.80 3.03 2.94 3.01

    0.5 2 13.19 5.16 2.69 2.12 3.28 1.851 2 10.76 2.41 3.34 2.59 3.15 2.572 2 9.71 1.50 2.48 1.50 1.49 1.613 2 9.44 1.77 1.49 1.41 1.36 1.304 2 9.11 2.42 1.39 1.47 1.19 1.305 2 9.22 3.18 1.18 1.27 1.35 1.31

    0.5 3 13.26 4.20 6.65 2.85 2.64 2.281 3 10.62 1.70 4.91 2.17 1.87 1.992 3 9.33 1.56 2.12 0.99 1.13 1.003 3 8.97 1.97 0.91 1.17 1.08 1.024 3 8.82 2.59 0.79 1.05 0.95 1.005 3 9.08 3.24 1.26 1.36 1.46 1.35

    0.5 4 12.55 3.66 5.51 2.50 2.13 2.251 4 10.11 1.76 3.46 1.55 1.43 1.512 4 8.84 2.28 0.94 1.20 1.43 1.183 4 8.58 2.70 1.02 1.33 1.32 1.294 4 8.59 3.13 1.15 1.29 1.33 1.285 4 9.18 3.78 2.22 2.26 2.23 2.14

    0.5 5 11.71 3.72 3.49 2.84 2.86 2.931 5 9.97 2.31 2.15 1.76 1.81 1.762 5 8.54 3.06 1.27 1.44 1.55 1.47

    3 5 8.42 3.47 1.48 1.45 1.54 1.514 5 8.66 3.92 1.89 2.01 2.10 1.975 5 9.76 4.37 3.52 3.42 3.15 3.13

    0.5 7 11.04 4.26 4.69 4.52 4.54 4.651 7 9.03 2.99 2.72 2.38 2.37 2.402 7 8.24 3.80 2.16 1.81 1.86 1.863 7 8.28 4.15 2.00 1.97 2.06 1.96

    Average 9.86 3.28 3.05 2.17 2.15 2.0626

  • 8/8/2019 Correlations of Bond Yields

    29/40

    Table 3: Tests of Alternative Model Specifications

    This table reports pairwise comparisons of pricing accuracy of alternative specifications of my model. Allmodels are estimated using a panel dataset consisting of 220 weekly observations on 34 swaptions fromJanuary 24, 1997 to April 6, 2001. For the model labelled by (N, K), there are N factors driving theevolution of the term structure of bond yields. The first K factors have unspanned stochastic volatility,and the other NK factors have constant volatility. The DM column reports the Diebold and Mariano(1995) statistic that tests the null hypothesis that the models in the first two columns have the samepricing accuracy for swaptions. Under the null hypothesis, the Diebold and Mariano statistic is distributedasymptotically as standard normal N(0, 1). Column four reports the number of weeks (out of a totalof 220 weeks) that Model 1 implies a higher root-mean-squared percent pricing error (RMSE) for thecross-section of swaptions, and column five reports the average reduction in swaption RMSE (reportedin percent) implied by Model 2 during these weeks. Similarly, the sixth column reports the number ofweeks that Model 2 implies a higher swaption RMSE, and the last column reports the average reductionin swaption RMSE (reported in percent) implied by Model 1 during these weeks.

    Model 1 Model 2 DM RM SE 2 < RMSE1 RMSE2 > RMSE1Statistic Frequency Mean Difference Frequency Mean Difference

    (1,1) (2,1) 6.46 211 1.88 9 0.47

    (2,1) (3,1) 9.59 213 1.29 7 0.31

    (3,1) (4,1) 0.26 119 0.17 101 0.16

    (2,1) (2,2) 2.87 218 1.02 2 0.11

    (2,2) (3,2) 6.68 211 1.36 9 0.48

    (3,2) (4,2) 2.18 178 0.19 42 0.08

    (3,1) (3,2) 3.61 200 1.17 20 0.21

    (3,2) (3,3) 3.55 166 0.32 54 0.21

    (3,3) (4,3) 2.50 161 0.16 59 0.10

    (4,0) (4,1) 2.38 209 7.17 11 0.62

    (4,1) (4,2) 2.83 214 1.21 6 0.18

    (4,2) (4,3) 2.10 149 0.31 71 0.21

    (4,3) (4,4) 0.52 127 0.21 93 0.20

    27

  • 8/8/2019 Correlations of Bond Yields

    30/40

    Table 4: Interest Rate Caps Pricing Errors

    This table reports the average and mean absolute pricing errors for the interest rate caps of maturitytwo, three, four, five, seven and ten years under various model specifications. For the model labelled by(N, K), there are N factors driving the evolution of the term structure of bond yields. The first K factorshave unspanned stochastic volatility, and the other N K factors have constant volatility. The pricingerror for an interest rate cap is defined as the difference between its no-arbitrage model price impliedfrom the swaptions and its market price expressed as a percentage of the market price. The model price is

    obtained from equation (17) using the model parameters (and implied volatility state variables) estimatedfrom the swaptions data. The dataset consists of 220 weekly observations on each series from January24, 1997 to April 6, 2001.

    Maturity Mean Percent Pricing Errors Mean Absolute Percent Errors(4,0) (4,1) (4,2) (3,3) (4,3) (4,0) (4,1) (4,2) (3,3) (4,3)

    2 -3.04 0.21 6.07 -2.82 4.52 10.96 6.27 7.26 4.52 5.943 -3.71 -3.26 0.91 -4.57 0.04 10.45 5.12 4.72 5.25 4.254 -3.37 -3.10 0.04 -3.92 -0.47 10.67 5.06 4.58 5.12 4.365 -3.35 -3.92 -1.51 -4.43 -1.77 10.53 5.35 4.51 5.33 4.457 -1.69 -3.33 -1.51 -3.44 -1.64 10.16 5.44 4.34 4.86 4.35

    10 0.81 -2.07 - 0.28 - 1.88 - 0.49 9.43 5.24 3.86 4.14 3.92

    Average -2.39 -2.58 0.62 -3.51 0.03 10.36 5.41 4.88 4.87 4.54

    28

  • 8/8/2019 Correlations of Bond Yields

    31/40

    Table 5: Maximum Likelihood Estimation of the GA4,3 Model

    This table reports the results of maximum likelihood estimation of the GA4,3 model. There are four factorsdriving the innovation of bond yields. The first three factors display unspanned stochastic volatility, andthe fourth factor has constant volatility. The model is estimated with a panel dataset of 220 weeklyobservations on 34 at-the-money-forward European swaptions from January 24, 1997 to April 6, 2001.Panel A reports the estimates for i, i and i (i = 1, 2, 3) that govern the risk-neutral dynamics ofthe volatility state variables. The is are the market prices of volatility risk. The standard errors ofparameter estimates reported in the parentheses are obtained by calculating the inverse of the informationmatrix based on the Hessian of the likelihood function. Panel B reports the sample correlations betweenthe yield factors and each volatility state variable, as well as the correlations between the implied volatilitystate variables. Panel C reports the sample mean of the implied variances of the first three yield factorsand their unconditional variances.

    Panel A: Parameter Estimates

    100 1 0.0509 0.0354 0.0081 -0.0582

    (0.0016) (0.0091) (0.0033) (0.1811)2 1.9703 0.0041 0.0138 -0.1479

    (0.0329) (0.0001) (0.0056) (0.4186)3 2.6144 0.0018 0.0079 -0.2007

    (0.3945) (0.0003) (0.0032) (0.2936)4 - 0.0026 - -

    - (0.0002) - -

    Panel B: Correlations between Yield Factors and Volatility Factors

    Factor Yield Factors Volatility Factors

    z1 z2 z3 z4 1 2 3z1 1z2 -0.0000 1z3 -0.0000 0.0000 1z4 -0.0000 -0.0000 0.0000 11 -0.0302 0.0084 -0.0162 -0.0579 12 0.0028 -0.0362 0.0613 0.0714 0.0955 13 -0.0291 -0.0434 -0.0791 -0.1288 0.3488 0.1473 1

    Panel C: Unconditional and Average Implied Factor Variance

    1031 1032 1033Historical 0.4584 0.0617 0.0159

    Avergae Implied 0.4428 0.0761 0.0178

    29

  • 8/8/2019 Correlations of Bond Yields

    32/40

    Figure 1: Time Series of Swaption Implied Volatilities

    This figure plots the time series of implied volatilities for 4 at-the-money-forward European-style swap-

    tions. The data are mid-market quotes of annualized implied volatilities measured in percentage. The

    sample is weekly from January 24, 1997 to April 6, 2001.

    01/24/97 07/25/97 01/23/98 07/24/98 01/22/99 07/23/99 01/21/00 07/22/00 01/20/01

    8

    10

    12

    14

    16

    18

    20

    22

    24

    26

    28

    Date

    Swaptio

    nImpliedVolatility

    0.5 into 1

    5 into 1

    0.5 into 5

    5 into 5

    30

  • 8/8/2019 Correlations of Bond Yields

    33/40

    Figure 2: Time Series of Pricing Errors for Swaptions and Interest Rate Caps

    This graph plots the time series of root mean squared pricing errors (RMSE) for swaptions and interest

    rate caps under the GA4,3 model. The model is estimated via maximum likelihood using 34 at-the-

    money-forward European swaptions. Panel A plots the RMSE of these 34 swaptions. Panel B plots the

    RMSE for interest rate caps with maturity two, three, four, five, seven and ten years. For each interest

    rate cap, the pricing error is calculated as the difference between its no-arbitrage value implied from the

    swaptions according to the GA4,3 model and its market price, expressed as a percentage of the market

    price. The dataset consists of 220 weekly observations on each series from January 24, 1997 to April 6,

    2001.

    01/24/97 11/24/97 09/24/98 07/25/99 05/24/00 03/24/01

    0

    2

    4

    6

    8

    Date

    RootMeanSquaredPercentErr

    or

    Swaptions

    01/24/97 11/24/97 09/24/98 07/25/99 05/24/00 03/24/010

    10

    20

    30

    40

    Date

    RootMe

    anSquaredPercentError

    Interest Rate Caps

    31

  • 8/8/2019 Correlations of Bond Yields

    34/40

    Figure 3: Average Implied Covariances versus Unconditional Covariances

    This graph compares the time series average of the option-implied conditional covariances of the six-month

    forward rates (up to 10 years in maturity) under the GA4,3 model to the their unconditional estimates.Panel A plots the average implied volatility and the sample standard deviation of the six-month forward

    rates (both annualized). Panel B plots the difference between the average implied correlations and the

    sample correlations based on weekly data from January 24, 1997 to April 6, 2001. The implied volatilities

    and implied correlations are calculated according to the covariance structure specified in equation (12)

    under the estimated GA4,3 model.

    1 2 3 4 5 6 7 8 9 100.008

    0.009

    0.01

    0.011

    0.012

    0.013

    0.014

    Maturity

    Volatility of SixMonth Forward Rates

    Average ImpliedUnconditional

    1 23 4

    5 67 8

    9

    1234

    56789

    0.1

    0.05

    0

    0.05

    0.1

    Maturity

    Difference Between Average Implied and Unconditional Correlations

    Maturity

    32

  • 8/8/2019 Correlations of Bond Yields

    35/40

    Figure 4: Conditional Covariances: GARCH versus Option-Implied

    This graph compares the conditional volatilities and correlations of the six-month Libor forward ratesimplied from the swaptions according to my GA4,3 model to those estimated from the historical bond

    prices according to Engle (2002)s GARCH model with dynamic conditional correlations. Panel A illus-

    trates two estimates of the volatility of the six-month forward rate with two-year maturity. Panel B plots

    two estimates of the correlation between the two-year maturity and five-year maturity six-month forward

    rates. In both panels, the horizonal line marks the unconditional estimate. The data sample is weekly

    from January 24, 1997 to April 6, 2001.

    01/24/97 11/24/97 09/24/98 07/25/99 05/24/00 03/24/010.005

    0.01

    0.015

    0.02

    Date

    Volatility

    GARCH(1,1)OptionImplied

    01/24/97 11/24/97 09/24/98 07/25/99 05/24/00 03/24/010.4

    0.5

    0.6

    0.7

    0.8

    0.9

    1

    Date

    Correlation

    Bivariate GARCH

    OptionImplied

    33

  • 8/8/2019 Correlations of Bond Yields

    36/40

    Appendix

    Proof of Proposition 1:

    Under the forward risk-neutral measure Q, Sj(t) is a martingale, since it differs from

    D(t , , j) only by a constant, for all j = 1, , 2(T). Moreover, the covariances among{d log D(t , , j)}j and {d log Sj}j are the same. Let S(t) be a vector of length 2(T )that stacks all Sj(t). Denote t as the instantaneous covariance matrix of d log S and Rtas the Cholesky decomposition of t so that RtR

    t = . Using equations (12) and (13),t can be expressed explicitly in terms of U and k(t), k = 1, , n, as follows: for anytime t < = i/2, let integer m be such that m/2 t < (m + 1)/2, then

    t = (1 2(t m/2)) A1UtUA1 + 2(t m/2) A2UtU

    A

    2

    where A1 and A2 are both 2(T ) by 20 matrix. All elements ofA1 (resp. A2) are zero,except that the 2(T

    ) by 2(T

    ) submatrix consisting of i

    m + 1th to i

    m +jth

    column (resp. from i mth to i +j 1 mth column) is a lower triangular matrix whoseelements on and below diagonal are 1. It follows each variance and covariance term in tis linear in the volatility state variables k(t), k = 1, , N.

    Let dZt be 2(T ) dimension standard Brownian motions, then under Q,dS(t)

    S(t)= RtdZt

    By Itos lemma and definition of G, this implies that

    d log G =

    2(T)j=1

    jd log Sj = 1

    2

    diag()dt +

    RtdZt

    where diag is an operator that takes the diagonal of a matrix. Note that if k(t) isdeterministic instead of following process (7), then G() is lognormal.

    Now I calculate the expectation in equation (14) via the law of iterated expectationby first conditioning on a vector V which stacks the average values of k(t) over timeinterval [0, ], for k = 1, , n. Based on the assumption that the volatility state vari-ables are instantaneously uncorrelated with innovations in the yield curve, and followingan argument of Hull-White (1987) (see their lemma on P284-285), G() is lognormallydistributed conditional on V:

    G() = exp( ) (21)where = 1

    2

    diag() and is a normal mean zero random variable whose variance

    is (

    ). is the average covariance matrix of S over [0, ] corresponding to V. Tocalculate the expectation in (14), I use the following simple lemma:

    Lemma 1 Let be a Gaussian random variable with zero mean and the variance 2. For

    34

  • 8/8/2019 Correlations of Bond Yields

    37/40

    any strictly positive numbers a and b, I have

    E

    max(ae0.52 b, 0)

    = aN(h) bN(h )

    where h = 1ln(a/b) + 0.5.

    Note that EQ

    [A] = 1 since each Sj is a martingale under Q and Sj(0) = 1. Hence

    g = EQ

    [G()]. Apply lemma 1 to G() as in (21), it follows that the price of a by Tat-the-money-forward swaption is approximately

    P = D(0, ) EQ

    V

    EMax

    G g, 0

    |V

    = D(0, ) EQ

    V

    N(1

    2

    ) N(1

    2

    )

    = D(0, )

    2N

    1

    2

    EQ

    V[]

    1)

    Here I use the well known fact at-the-money option is almost linear in the Black volatilitywhich follows from the property of the normal cumulative density function N() and thefact that expectation operator is linear.

    To finish the evaluation of the expectation above, I just need to compute the averageexpected value of the s from time 0 to and hence the average expected covariancematrix of the forward bonds involved in pricing by T swaption. This can be done withsimple but tedious algebra, using the conditional expectation of a mean reverting process:

    E0[j(t)] = ejtj(0) + (1 ejt)j

    Proof of Proposition 2: A cap on six-month Libor rate of final maturity T year withcap rate R consists of 2 T 1 caplets Ci, i = 1, . . . , 2T 1. Let i = i2 and D(t, i, i+1be the forward Libor bond price. Each caplet Ci protects the six-month Libor rate Liapplicable between time [i, i+1] year from rising above R. Its pays off

    12

    max(Li R, 0)per unit dollar of principal at time i+1. Since Li = 2(

    1D(i,i,i+1)

    1), the payoff of capletCi discounted to time i can be rewritten as

    1

    1 + 0.5Li

    1

    2max(Li R, 0) = (1 + 0.5R)max( 1

    1 + 0.5R 1

    1 + 0.5Li, 0)

    = (1 + 0.5R)max(

    1

    1 + 0.5R D(i, i, i+1), 0)Hence each caplet Ci is a European put option on forward Libor bond D(t, i, i), witha strike price 1

    1+0.5Rand maturing at i. A cap can be also viewed as a portfolio of put

    options on forward Libor bond. I can price each caplet Ci using the corresponding forwardmeasure Qi. The key quantity is the average expected variance of each forward Liborbond (of fixed maturity), which can be calculated from equations (12), (7) and (13).

    35

  • 8/8/2019 Correlations of Bond Yields

    38/40

    References

    [1] Ball, C. and W. Torous (1999), The Stochastic Volatility of Short-Term InterestRates: Some International Evidence, Journal of Finance, 54, no.6, 2339-2359.

    [2] Black, F. (1976), The Pricing of Commodity Contracts, Journal of Financial Eco-nomics 3, 167 179

    [3] Bliss, R., (1997), Movements in the Term Structure of Interest Rates, FederalReserve Bank of Atlanta Economic Review, Fourth Quarter, 1997, 16-33.

    [4] Chen, R. and L. Scott (1993), Maximum Likelihood Estimation For a MultifactorEquilibrium Model of the Term Structure of Interest Rates, Journal of Fixed Income,3, 14-31.

    [5] Chen, R. and L. Scott (2001), Stochastic Volatility and Jumps in Interest Rates:

    An Empirical Analysis, Working Paper, Rutgers University.[6] Collin-Dufresne, P. and R. S. Goldstein (2002), Do Bonds Span the Fixed Income

    Markets? Theory and Evidence fo