41
Views
0
CrossRef citations to date
0
Altmetric
Features

Interest rate convexity in a Gaussian framework

&
Received 14 Aug 2023, Accepted 11 May 2024, Published online: 13 Jun 2024

Abstract

The contributions of this paper are twofold: we define and investigate the properties of a short rate model driven by a general Gaussian Volterra process and, after defining precisely a notion of convexity adjustment, derive explicit formulae for it.

2010 Mathematics Subject Classification:

1. Introduction and notations

1.1. Introduction

In fixed-income markets, the different schedules of payments and the diverse currencies, margins require specific adjustments in order to price all interest-rate products consistently. This is usually referred to as convexity adjustment and has a deep impact on interest rate derivatives. Starting from Brotherton-Ratcliffe and Iben (Citation1993), Flesaker (Citation1993) and Ritchken and Sankarasubramanian (Citation1993), academics and practitioners alike have developed a series of formulae for this convexity adjustment in a variety of models, from simple stochastic rate models (Kirikos and Novak Citation1997) to some incorporating stochastic volatility features (Andersen and Piterbarg Citation2010). Recently, García-Lorite and Merino (Citation2023) used Malliavin calculus techniques to compute approximations of this convexity adjustment for various interest rate products. Motivated by the new paradigm of rough volatility in Equity markets (Bayer et al. Citation2016, El Euch et al. Citation2018, Gatheral et al. Citation2018, Fukasawa Citation2021, Jacquier et al. Citation2021, Bayer et al. Citation2023, Bonesini et al. Citation2023, Jacquier and Oumgari Citation2023), we consider here stochastic dynamics for the short rate, driven by a general Gaussian Volterra process, providing more flexibility than standard Brownian motion. In the framework of the change of measure approach in Pelsser (Citation2003), we introduce a clear definition of convexity adjustment for zero coupon bonds, in proposition 2.11, namely as the non-martingale correction of ratios of zero-coupon prices under the forward measure, for which we are able to derive closed-form expressions or asymptotic approximations. We introduce the model, derive its properties in section 2. In section 2.2, we define convexity adjustment and provide formulae for it, the main result of the paper, which we illustrate in some specific examples. Section 3 provides some further expressions for liquid interest rate products, and we highlight some numerical aspects of the results in section 4.

1.2. Model and notations

On a given filtered probability space (Ω,F,(Ft)t0,P), we are interested in short rate dynamics of the form (1) rt=θ(t)+0tφ(t,u)dWu=θ(t)+(φ(t,)W)t,(1) with θ a deterministic function and W a continuous Gaussian process adapted to the filtration (Ft)t0. Here and below, given a function ϕ and a stochastic process X, we write (ϕX)a,b:=abϕ(s)dXs, and omit a whenever a = 0. For some fixed time horizon T>0, define further, for utT, (2) ΞT(t,u):=tTφ(s,u)ds andΞT(u):=ΞT(u,u)(2) as well as Θt,T:=tTθ(s)ds. We consider a given risk-neutral probability measure Q, equivalent to P, so that the price of the zero-coupon bond at time t is given by (3) Pt,T:=EtQ[Bt,T], whereBt,T:=exp{tTrsds},(3) and we define the instantaneous forward rate process as (4) ft,T:=TlogPt,T.(4)

Remark 1.1

For modeling purposes, we shall consider kernels of convolution type, namely (5) φ(t,u)=φ(tu).(5)

1.3. Empirical motivation

The modeling framework above (and in particular the introduction of a potentially singular kernel) is motivated by empirical observations. Assume that the kernel is given by a power-law form φ(t,u)=(tu)H/12 with H(0,1), and that W is a standard Brownian motion. To estimate the Hurst exponent H, we follow the methodology devised in Gatheral et al. (Citation2018) for the instantaneous log volatility (although more refined and robust statistical estimation techniques are now available, we leave a detailed empirical analysis for future work) and compute it via the linear regression logE[|rt+Δrt|2]=2Hlog(Δ)+c, for Δ>0, for some constant c. Of course such a linear regression hinges on some assumptions in the form of (rt)t0 but a detailed analysis of short rate data is beyond the scope of the present paper, and we only provide here short insights into the potential roughness of short rates dynamics. We consider the sports interest rate data from Option Metrics.Footnote1 We consider the data from 4/1/2010 until 28/2/2023. For different dates within this period, figure  shows the available data points (circles) as well as the interpolation by splines (the extrapolation is assumed flat). In figure , we compute the time series of the yield curves, for each (interpolated) maturities and estimate the Hurst exponent for each maturity.

Figure 1. Examples of rates curves over different days from the OptionMetrics rates data.

Figure 1. Examples of rates curves over different days from the OptionMetrics rates data.

Figure 2. Top: Time series of the OptionMetrics rates for different maturities. Bottom: Estimation of the Hurst exponent for the OptionMetrics rates data.

Figure 2. Top: Time series of the OptionMetrics rates for different maturities. Bottom: Estimation of the Hurst exponent for the OptionMetrics rates data.

A similar analysis on the US Daily Treasury Par Yield Curve RatesFootnote2 yields figure .

Figure 3. Top: Time series of the US Treasury rates for different maturities. Bottom: Estimation of the Hurst exponent for the US Treasury rates.

Figure 3. Top: Time series of the US Treasury rates for different maturities. Bottom: Estimation of the Hurst exponent for the US Treasury rates.

2. Gaussian martingale driver

2.1. Dynamics of the zero-coupon bond price

We assume first that W is a continuous Gaussian martingale with γW(t):=E[Wt2] finite for all t0. In this case, the (predictable) quadratic variation process γW() is clearly deterministic, but also continuous and increasing, and therefore its derivative γW exists almost everywhere. In order to ensure existence of the rate process in (Equation1), we assume the following (we write dλ for the Lebesgue measure on R+):

Assumption 2.1

For each t[0,T], φ(t,)L1(dλ)L2(γW), and φ is of convolution type (Equation5).

Lemma 2.2

Under assumption 2.1, (ΞT(t,)W)t is an (Ft)t[0,T] Gaussian semimartingale.

Proof.

From (Equation2), ΞT is in general not in convolution form (Equation5). However, since φ is, we can write ΞT(t,u):=tTφ(s,u)ds=tTφ(su)ds=Φ(Tu)Φ(tu), where the function Φ is defined as Φ(z):=zφ(u)du. The stochastic integral then reads (ΞT(t,)W)t=0tΞT(t,u)dWu=0t[Φ(tu)Φ(Tu)]dWu, which corresponds to a two-sided moving average process in the sense of Basse-O'Connor and Graversen (Citation2010, Section 5.2). Assumption 2.1 then implies that for each t[0,T], the function ΞT(t,) is absolutely continuous on [0,t] and tΞT(t,)L2(γW) and the statement follows from Basse-O'Connor and Graversen (Citation2010, Theorem 5.5).

Remark 2.3

  • The L2 property ensures that the stochastic integral (φ(t)W)t is well defined.

  • The assumption does not imply that the short rate itself, while Gaussian, is a semimartingale.

Proposition 2.4

The price of the zero-coupon bond at time t reads Pt,T=exp{Θt,T+12tTΞT(u)2du+(ΞT(t,)W)t}, and the discounted bond price P~t,T:=Pt,Texp{0trsds} is a Q-martingale satisfying dP~t,TP~t,T=ΞT(t) dWt.

Corollary 2.5

The instantaneous forward rate satisfies fTT=rT and, for all t[0,T), ft,T=θ(T)+0tφ(T,u)dWu+tTφ(T,u)ΞT(u)du.

In differential form, for any fixed T>0, for t[0,T], this is equivalent to dft,T=φ(Tt)dWtφ(Tt)ΞT(t)dt.

Algorithm 2.6

For simulation purposes, we consider a time grid T:={0=t0<t1<<tN=T} and discretize the stochastic integral along this grid with left-point approximations as (ΞT(ti,)W)ti=0tiΞT(ti,u)dWuk=0i1ΞT(ti,tk)(Wtk+1Wtk), for each i=1,,N. The vector (ΞT(ti,)W)tiT of stochastic integrals can then be simulated along the grid directly as ((ΞT(t1,)W)t1(ΞT(tN,)W)tN)(ΞT(t1,t0)ΞT(t2,t0)ΞT(t2,t1)ΞT(tN1,t0)ΞT(tN1,t1)ΞT(tN1,tN2)ΞT(tN,t0)ΞT(tN,t1)ΞT(tN,tN1))×(Wt1Wt0WtNWtN1), where the middle matrix is lower triangular (we omit the null terms everywhere for clarity).

Example 2.7

With φ(t)=σeκt, for σ>0, θ(t):=r0eκt+μ(1eκt) and W=W a Brownian motion, we recover exactly the Vasicek model (Vasicek Citation1977), namely rt=r0+κ0t(μrs)ds+σWt.

Example 2.8

Consider the extension of the Vasicek model proposed by Hull and White (Citation1990), where drt=(ζ(t)a(t)rt)dt+σ(t)dWt, where ζ(),a() and σ() are sufficiently smooth deterministic functions of time. Direct computations yield the solution, with A(t):=0ta(s)ds, rt=r0+0te(A(t)A(s))ζ(s)ds+0te(A(t)A(s))σ(s)dWs. Letting θ(t):=r0+0te(A(t)A(s))ζ(s)ds,φ(t,s):=e(A(t)A(s)) anddWt=σ(t)dWt makes it coincide exactly with our setup in (Equation1). Now assumption 2.1 holds if and only if A(t)A(s)=A(ts) for all 0st, namely when the function a is linear or constant. Note that, as mentioned in Brigo and Mercurio (Citation2006, Section 3.3), the function a is often assumed constant in practice.

Proof of proposition 2.4.

The price of the zero-coupon bond at time t then reads (6) Pt,T:=EtQ[exp{tTrsds}]=EtQ[exp{tT(θ(s)+0sφ(s,u)dWu)ds}]=eΘt,TEtQ[exp{tT(0sφ(s,u)dWu)ds}].(6) Using Fubini, we can write (7) tT(0sφ(s,u)dWu)ds=0t(tTφ(s,u)ds)dWutT(uTφ(s,u)ds)dWu=0tΞT(t,u)dWu+tTΞT(u)dWu,(7) using (Equation2). Plugging this into (Equation6), the zero-coupon bond then reads Pt,T=eΘt,Texp{0tΞT(t,u)dWu}EtQ×[exp{tTΞT(u)dWu}]=eΘt,Texp{(ΞT(t,)W)t}EtQ[e(ΞTW)t,T]. Conditional on Ft, (ΞTW)t,T is centered Gaussian with Vt[(ΞTW)t,T]=tTΞT(u)2du, hence Pt,T=eΘt,Texp{(ΞT(t,)W)t+12tTΞT(u)2du}. By Fubini and assumption 2.1, (ΞT(t,)W)t=0tΞT(t,u)dWu=0t(ΞT(u)+utsΞT(s,u)ds)dWu=0tΞT(u)dWu+0tutsΞT(s,u)dsdWu=0tΞT(u)dWu+0t0ssΞT(s,u)dWuds=0tΞT(u)dWu+0t0sφ(s,u)dWuds. This is an L1-Dirichlet process (Russo and Tudor Citation2006, Definition 2), written as a decomposition of a local martingale and a term with zero quadratic variation. Therefore log(P,T),log(P,T)t=0tΞT(u)2du and (8) dlog(P,T)=(θ(t)+(tΞT(t,)W)t12ΞT(t)2)dt+ΞT(t)dWt.(8) Now, Itô's formula with Xt:=log(Pt,T), using (Equation8) yields PT,T=Pt,T+tTPs,TdXs+12tTPs,TdX,Xs, hence, for each T>0, dPT,T=dPt,TPt,TdXt12Pt,TdX,Xt, and therefore, since PT,T=1, dPt,TPt,T=dXt+12dX,Xt=(θ(t)+(tΞT(t,)W)trt12ΞT(t)2)dt+ΞT(t)dWt+12d(0tΞT(u)2du)=rtdt+ΞT(t)dWt12ΞT(t)2dt+12ΞT(t)2dt=rtdt+ΞT(t)dWt. The dynamics of the discounted zero-coupon bond price in the lemma follows immediately.

Proof of corollary 2.5.

It follows by direct computation starting from the instantaneous forward rate (Equation4): ft,T=TΘt,TT0tΞT(t,u)dWu12TtTΞT(u)2du=TΘt,TT0t(tTφ(s,u)ds)dWu12TtT(uTφ(s,u)ds)2du=θ(T)+0tT(tTφ(s,u)ds)dWu12TtT(uTφ(s,u)ds)2du=θ(T)+0tφ(T,u)dWu12(TTφ(s,T)2ds+tTT[(uTφ(s,u)ds)2]du)=θ(T)+0tφ(T,u)dWutTφ(T,u)(uTφ(s,u)ds)du=θ(T)+0tφ(T,u)dWu+tTφ(T,u)ΞT(u)du.

Remark 2.9

The two lemmas above correspond to the two sides of the Heath–Jarrow–Morton framework. From the expression of the instantaneous forward rate, let αt,T:=φ(Tt)ΞT(t) and βt,T:=φ(Tt), so that dft,T=βt,TdWtαt,Tdt, and consider the discounted bond price P~t,T:=Pt,Texp{0trsds}=exp{0trsdstTft,sds}=:exp{Zt}. Itôs' formula then yields (9) dP~t,TP~t,T=dZt+12dZ,Zt.(9) From the differential form of ft,T, we can write, for any t[0,T), ft,T=f0,T+0tdfs,T=f0,T+0t(φ(T,u)dWuφ(T,u)ΞT(u)du)=f0,T+0tβu,TdWu+0tαu,Tdu, so that, using stochastic Fubini, we obtain Ft,T:=tTft,sds=tT(f0,s+0tβu,sdWu+0tαu,sdu)ds=tTf0,sds+0ttTβu,sdsdWu+0ttTαu,sdsdu. Now, tTf0,sds=tT(fs,s0sufu,sdu)ds=tTrsds0ttTufu,sdsdutTuTufu,sdsdu=tTrsds0t(tTufu,sdsuTufu,sds)du0TuTufu,sdsdu=tTrsds+0tutufu,sdsdu0TuTufu,sdsdu, using Fubini, so that Ft,T=tTrsds+0tutufu,sdsdu0TuTufu,sdsdutTf0,sds+0ttTβu,sdsdWu+0ttTαu,sdsdu, and dFt,T=(tTαt,sdsrt)dt+(tTβt,sds)dWt. Therefore, dZt=d(0trsdstTft,sds)=rtdtdFt,T=rtdtdFt,T=(tTαt,sds)dt(tTβt,sds)dWt, and (Equation9) gives dP~t,TP~t,T=(tTαt,sds12(tTβt,sds)2)dt(tTβt,sds)dWt. The discounted process (P~t,T)t[0,T] is a local martingale if and only if its drift is null: for t(0,T), T{tTαt,sds12[tTβt,sds]2}=αt,Tβt,TtTβt,sds=φ(Tt)×[ΞT(t)tTφ(s,t)ds], which is equal to zero by definition of the functions. Therefore the drift (as a function of T) is constant. Since it is trivially equal to zero at T = t, it is null everywhere and (P~t,T)t[0,T] is a Q-local martingale.

2.2. Convexity adjustments

We now enter the core of the paper, investigating the influence of the Gaussian driver on the convexity of bond prices. We first start with the following simple proposition:

Proposition 2.10

For any T,τ0, d(1Pt,τ)=(Ξτ(t,t)2γW(t)rt)dtPt,τΞτ(t,t)Pt,τdWt,d(Pt,TPt,τ)=Pt,TPt,τ(ΞT(t,t)Ξτ(t,t))×{Ξτ(t,t)γW(t)dt+dWt}, and there exists a probability measure Qτ such that WtQτ is a Qτ-Gaussian martingale and (10) d(Pt,TPt,τ)=Pt,TPt,τΣtT,τdWtQτ,(10) under Qτ, where ΣtT,τ:=ΞT(t)Ξτ(t)=Φ(Tt)Φ(τt).

Note that, from the definition of ΞT in (Equation2), ΣtT,τ is non-negative whenever τT. In standard Fixed Income literature, the probability measure Qτ corresponds to the τ-forward measure.

Proof.

From the definition of the zero-coupon price (Equation3) and proposition 2.4, Pt,T>0 is strictly positive almost surely and dPt,TPt,T=rtdt+ΞT(t,t)dWt, and therefore Itô's formula implies that, for any 0tτ, d(1Pt,τ)=dPt,τPt,τ2+dPt,τ,Pt,τPt,τ3=(Ξτ(t,t)2γW(t)rt)dtPt,τΞτ(t,t)dWtPt,τ. Therefore d(Pt,TPt,τ)=Pt,Td(1Pt,τ)+dPt,TPt,τ+dPt,Td(1Pt,τ)=Pt,TPt,τ{(Ξτ(t,t)2γW(t)rt)dtΞτ(t,t)dWt+(rtdt+ΞT(t,t)dWt)ΞT(t,t)Ξτ(t,t)γW(t)dt}=Pt,TPt,τ(ΞT(t,t)Ξτ(t,t))×{Ξτ(t,t)γW(t)dt+dWt}. Define now the Doléans–Dade exponential Mt:=exp{0tΞτ(s,s)γW(s)dWs120t[Ξτ(s,s)γW(s)]2ds}, and the Radon–Nikodym derivative dQτdP:=M. Girsanov's Theorem (Øksendal Citation2003, Theorem 8.6.4) implies that WtQτ:=Wt0tΞτ(s,s)γW(s)ds is a Gaussian martingale and Pt,TPt,τ satisfies (Equation10) under Qτ.

The following proposition is key and provides a closed-form expression for the convexity adjustments:

Proposition 2.11

For any τ0 let t1,t20. We then have EQτ[Pt,t1Pt,t2]=P0,t1P0,t2Ctτ(t1,t2), for any t[0,t1t2], where Ctτ(t1,t2):=exp{0t(Σst2,τΣst1,τ)Σst2,τγW(s)ds} is the convexity adjustment factor.

Remark 2.12

  • When t = 0 or t1=t2 or Pt,t1Pt,t2 is constant, there is no convexity adjustment, i.e. Ctτ(t1,t2)=1.

  • More interestingly, if t2=τ, then Σtt2,τ=Σtt2,t2=Ξt2(t,t)Ξt2(t,t)=0 and Ctτ(t1,t2)=Ctt2(t1,t2)=exp{0t(Σst2,t2Σst1,t2)Σst2,t2γW(s)ds}=1, and the process (Pt,t1Pt,t2)t0 is a Qτ-martingale on [0,t1t2].

  • Regarding the sign of the convexity adjustment, we have Σst2,τΣst1,τ=(Ξt2(s,s)Ξτ(s,s))(Ξt1(s,s)Ξτ(s,s))=Ξt2(s,s)Ξt1(s,s)=st2φ(z,s)dz+st2φ(z,s)dz=t1t2φ(z,s)dz. Since φ() is strictly positive, then sgn(Σst2,τΣst1,τ)=sgn(t1t2). Furthermore, since Σst2,τ=Ξt2(s,s)Ξτ(s,s)=st2φ(z,s)dz+sτφ(z,s)dz=t2τφ(z,s)dz, then sgn(Σst2,τ)=sgn(τt2), and therefore, assuming γW strictly positive (as will be the case in all the examples considered here),

    Table 1. aaaa

    Considering without generality t1<t2, the convexity adjustment is therefore greater than 1 for τ<t2 and less than 1 above.

Proof of proposition 2.11.

Under Qτ, the process defined as Xt:=Pt,T/Pt,τ satisfies dXt=XtΣtT,τdWtQτ, is clearly lognormal and hence Itô's formula implies dlog(Xt)=dXtXt12dX,XtXt2=ΣtT,τdWtQτ12(ΣtT,τ)2γW(t)dt, so that Xt=X0exp{0tΣsT,τdWs120t(ΣsT,τ)2γW(s)ds}, and therefore Pt,TPt,τ=P0,TP0,τexp{0tΣsT,τdWs120t(ΣsT,τ)2γW(s)ds}. With successively T=t1 and T=t2, we can then write Pt,t1Pt,τ=P0,t1P0,τexp{0tΣst1,τdWs120t(Σst1,τ)2γW(s)ds},Pt,t2Pt,τ=P0,t2P0,τexp{0tΣst2,τdWs120t(Σst2,τ)2γW(s)ds}, so that Pt,t1Pt,t2=P0,t1P0,t2exp{0tΣst1,τdWs120t(Σst1,τ)2γW(s)ds0tΣst2,τdWs+120t(Σst2,τ)2γW(s)ds}=P0,t1P0,t2exp{0t(Σst1,τΣst2,τ)dWs+120t[(Σst2,τ)2(Σst1,τ)2]γW(s)ds}=P0,t1P0,t2exp{0t(Σst1,τΣst2,τ)dWs+120t(Σst1,τΣst2,τ)2γW(s)ds}exp{120t[(Σst1,τ)2+(Σst2,τ)22Σst1,τΣst2,τ]γW(s)ds+120t[(Σst2,τ)2(Σst1,τ)2]γW(s)ds}=P0,t1P0,t2exp{0t(Σst1,τΣst2,τ)dWs+120t(Σst1,τΣst2,τ)2γW(s)ds}exp{0t[(Σst2,τ)2Σst1,τΣst2,τ]γW(s)ds}. The first exponential is a Doléans-Dade exponential martingale under Qτ, thus has Qτ-expectation equal to one, and the proposition follows.

2.3. Examples

Let W=W be a standard Brownian motion, so that γW(t)=t and γW(t)=1.

2.3.1. Exponential kernels

Assume that φ(t)=eαt for some α>0, then the short rate process is of Ornstein-Uhlenbeck type and ΞT(t,u)=Φ(Tu)Φ(tu) withΦ(z):=1αeαz. We can further compute Ξτ(t,t)=Φ(τ,t)Φ(t,t), and ΣtT,τ=ΞT(t,t)Ξτ(t,t)=Φ(T,t)Φ(t,t)Φ(τ,t)+Φ(t,t)=Φ(T,t)Φ(τ,t). Therefore the diffusion coefficient ΣtT,τ and the Girsanov drift Ξτ(t,t) read Ξτ(t,t)=1α(eα(τt)1) andΣtT,τ=1α(eα(Tt)eα(τt)). Finally, regarding the convexity adjustment, logCtτ(t1,t2)=e2αt12α3{(eαt1eαt2)eατ+e2αt2eα(t1+t2)}. Note that, as α tends to zero, namely rt=θ(t)+Wt (in the limit), we obtain Ctτ(t1,t2)=exp{(t2t1)(t2τ)t}.

2.3.2. Riemann–Liouville kernels

Let H(0,1) and H±:=H±12. If φ(t)=tH, with, the short rate process (Equation1) is driven by a Riemann–Liouville fractional Brownian motion with Hurst exponent H. Furthermore, with H+:=H+12, ΞT(t,u)=Φ(Tu)Φ(tu) withΦ(z):=zH+H+. Therefore the diffusion coefficient ΣtT,τ and Girsanov drift Ξτ(t,t) read Ξτ(t,t)=(τt)H+H+ andΣtT,τ=(τt)H+(Tt)H+H+. Regarding the convexity adjustment, we instead have Ctτ(t1,t2)=exp{0t(Σst2,τΣst1,τ)Σst2,τds} Unfortunately, there does not seem to be a closed-form simplification here. We can however provide the following approximations:

Lemma 2.13

The following asymptotic expansions are straightforward and provide some closed-form expressions that may help the reader grasp a flavor on the roles of the parameters:

  • As t tends to zero, logCtτ(t1,t2)=tH+2(t2H+t1H+)×(t2H+τH+)+O(t2).

  • For any η>0, as ε tend to zero, logCtt1ϵ(t1,t1+ϵ)=1+η2H(t12H(t1t)2H)ϵ2+O(ϵ3).

Proof.

From the explicit computation of ΣtT,τ above, we can write, as s tends to zero, ΣsT,τ=(τs)H+(Ts)H+H+=τH+TH+H++O(s). As a function of s, Σst2,τ is continuously differentiable. Because we are integrating over the compact [0,t], we can integrate term by term, so that logCtτ(t1,t2)=0t(Σst2,τΣst1,τ)Σst2,τds=0t{(τH+t2H+H+τH+t1H+H++O(s))×(τH+t2H+H++O(s))}ds=0t{(t1H+t2H+H++O(s))×(τH+t2H+H++O(s))}ds=t1H+t2H+H+τH+t2H+H+t+O(t2), where we can check by direct computations that the term O(t2) is indeed non null.

2.4. Extension to smooth Gaussian Volterra semimartingale drivers

Let now W in (Equation1) be a Gaussian Volterra process with a smooth kernel of the form Wt=0tK(t,u)dWu, for some standard Brownian motion W. Assuming that K is a convolution kernel absolutely continuous with square integrable derivative, it follows by Basse-O'Connor and Graversen (Citation2010) that W is a Gaussian semimartingale (yet not necessarily a martingale) with the decomposition Wt=0tK(u,u)dWu+0t(0u1K(u,s)dWs)du=:0tK(u,u)dWu+A(t), where A is a process of bounded variation satisfying dA(t)=A(t)dt=(0t1K(t,s)dWs)dt and hence the Itô differential of Wt reads dWt=K(t,t)dWt+A(t)dt, and its quadratic variation is dW,Wt=0tK(u,u)2du. The short rate process (Equation1) therefore reads rt=θ(t)+0tφ(tu)dWu=θ(t)+0tφ(tu)(K(u,u)dWu+A(u)du)=θ~t+0tφ~(t,u)K(u,u)dWu, where θ~t:=θ+0tφ(tu)A(u)du and φ~(t,u):=φ(tu)K(u,u). If φ~ satisfies assumption 2.1, then the analysis above still holds.

2.4.1. Comments on the Bond process

Let Rt,T:=tTrsds be the integrated short rate process and Bt,T:=eRt,T the bond price process on [0,T].

Lemma 2.14

The process (Bt,T)t[0,T] satisfies BT,T=1 and, for t[0,T), dBt,TBt,T=rtdt=(θ(t)+0tφ(tu)A(u)du+0tφ(tu)K(u,u)dWu)dt.

Proof.

For any t[0,T), we can write rt=θ(t)+0tφ(tu)d(0uK(s,s)dWs+A(u))=θ(t)+0tφ(tu)A(u)du+0tφ(tu)K(u,u)dWu. and therefore (11) dRt,T=rtdt=(θ(t)+0tφ(tu)A(u)du+0tφ(tu)K(u,u)dWu)dt.(11) Itô's formula (Alòs et al. Citation2001, Theorem 4) then yields BT,T=Bt,TtTBs,TdRs,T+12tTBs,TdR,Rs,T=Bt,T+tTBs,T{(θ(s)+0sφ(s,u)A(u)du)+0sφ(s,u)K(u,u)dWu}ds. so that, since BT,T=1, the lemma follows from dBt,T=d(tTBs,T{(θ(s)+0sφ(s,u)A(u)du)+0sφ(s,u)K(u,u)dWu}ds)=Bt,T{(θ(t)+0tφ(tu)A(u)du)+0tφ(tu)K(u,u)dWu}dt.

Remark 2.15

We can also write Rt,T in integral form as follows, using stochastic Fubini: Rt,T=tT[θ(s)+0sφ(s,u)A(u)du+0sφ(s,u)K(u,u)dWu]ds=Θt,T+tT(0sφ(s,u)A(u)du)ds+tT(0sφ(s,u)K(u,u)dWu)ds=Θt,T+0t(tTφ(s,u)ds)A(u)du+0t(tTφ(s,u)ds)K(u,u)dWu+tT(uTφ(s,u)ds)A(u)du+tT(uTφ(s,u)ds)K(u,u)dWu=Θt,T+0tΦt(u)A(u)du+0tΦtK(u)dWu+tTΦu(u)A(u)du+tTΦuK(u)dWu, with Φt(u):=tTφ(s,u)ds and ΦtK(u):=Φt(u)K(u,u). As a consistency check, we have dRt,T=θ(t)dt+Φt(t)A(t)dt+ΦtK(t)dWtΦt(t)A(t)dtΦtK(t)dWt+0ttΦt(u)A(u)dudt+0ttΦtK(u)dWudt=(θ(t)+Φt(t)A(t)Φt(t)A(t)+0ttΦt(u)A(u)du+0ttΦtK(u)dWu)dt+(ΦtK(t)ΦtK(t))dWt=(θ(t)+Φt(t)A(t)Φt(t)A(t)+0ttΦt(u)A(u)du)dt+0ttΦtK(u)dWudt=(θ(t)+0ttΦt(u)A(u)du)dt+0ttΦtK(u)dWudt=(θ(t)+0tφ(tu)A(u)du)dt0tφ(tu)K(u,u)dWudt, which corresponds precisely to (Equation11).

2.4.2. Specific example

Consider the kernel K(t,s)=eβ(ts) with β>0, so that K(t,t)=1 and tK(t,s)=βK(t,s). In this case, A(t)=0ttK(t,s)dWs=βWt, so that dWt=dWtβWtdt, which is an Ornstein-Uhlenbeck process, with covariance, for all s,t0, E[WsWt]=E[0sK(s,u)dWu0tK(t,u)dWu]=0sK(s,u)K(t,u)du=eβ|ts|eβ(s+t)2β. The short rate dynamics in (Equation1) then reads rt=θ(t)+0tφ(t,u)dWu=θ(t)+0tφ(t,u)(dWuβWudu)=θ~(t)+0tφ(t,u)dWu, with θ~(t)=θ(t)β0tφ(tu)Wudu, and the zero-coupon bond dynamics ( proposition 2.4) reads Pt,T=exp{Θ~t,T+12tTΞT(u)2du+0tΞT(t,u)dWu}, with Θ~t,T:=tTθ~sds=tT(θ(s)β0sφ(su)Wudu)ds=Θt,TβtT0sφ(su)Wududs. Applying stochastic Fubini, we then obtain Θ~t,T=Θt,Tβ0Tt+u(1t/T)Tφ(su)dsWudu=Θt,Tβ0T(Φ(Tu)Φ(ttT))Wudu. We note that the convexity adjustment in proposition 2.11 is only affected by a different weighting scheme in the integral given by the function γW. In our case, from the covariance computation above, γW(t)=12β(1e2βt), and therefore γW(t)=e2βt.

3. Pricing OIS products and options

3.1. Simple compounded rate

Using proposition 2.4, we can compute several OIS products and options Consider the simple compounded rate (12) rS(t0,T):=1D(t0,T)(i=0n11Pti,ti+11),(12) where D(t0,T) is the day count fraction and n the number of business days in the period [t0,tn]. The following then holds directly: rS(t0R,T)=1D(t0,T)(i=0n1exp{ΘtiR,ti+1R12tiRti+1RΞ(u,u)2du(Ξ(tiR,)W)tiR}1), where the superscript R refers to reset dates; we use the superscript A to refer to accrual dates below.

3.2. Compounded rate cashflows with payment delay

The present value at time zero of a compounded rate cashflow is given by PVflow=P0,TpD(t0A,tnA)EQTp[rS]=P0,TpD(t0A,tnA)EQTp[1D(t0A,tnA)×{i=0n1(1+D(tiA,ti+1A)D(tiR,ti+1R)(Pt,tiRPt,ti+1R1))1}], where rS denotes the compounded RFR rate. In the case where there is no reset delays, namely tiR=tiA for all i=0,,n, then PVflow=P0,TpEQTp[i=0n1(Pt,tiRPt,ti+1R)1]=P0,TpEQTp[Pt,t0RPt,tnR1]=P0,Tp(P0,t0RP0,tnRCtTp(t0R,tnR)1)=P0,Tp(P0,TRSP0,TRECtTp(TRS,TRE)1), where t0R=TRS and tnR=TRE, using the convexity adjustment formula given in proposition 2.11.

3.3. Compounded rate cashflows with reset delay

Assuming now that tiRtiA, we can write rtS=r~tS+rtS,adj, from (Equation12), where r~tS:=1D(t0R,tnR)(Pt,TRSPt,TRE1), and rtS,adj is implied from the decomposition above. Therefore PVflow=P0,TpD(t0A,tnA)EQTp[rtS]=P0,TpD(t0A,tnA)EQTp[r~tS+rtS,adj]=P0,TpD(t0A,tnA)EQTp×[1D(t0R,tnR)(Pt,TRSPt,TRE1)+rtS,adj]=P0,TpD(t0A,tnA)×{1D(t0R,tnR)(P0,TRSP0,TRECtTp(TRS,TRE)1)+EQTp[rtS,adj]}=P0,TpD(t0A,tnA)D(t0R,tnR)×{P0,TRSP0,TRECtTp(TRS,TRE)1+D(t0R,tnR)EQTp[rtS,adj]}. Assume now that EQTp[rtS,adj]=r0S,adj, so that we can simplify the above as PVflow=P0,TpD(t0A,tnA)D(t0R,tnR)×{P0,TRSP0,TRECtTp(TRS,TRE)1+D(t0R,tnR)r0S,adj}=P0,TpD(t0A,tnA)D(t0R,tnR)×{P0,TRSP0,TRECtTp(TRS,TRE)1+D(t0R,tnR)(r0Sr~0S)}=P0,TpD(t0A,tnA)D(t0R,tnR)×{P0,TRSP0,TRECtTp(TRS,TRE)1+D(t0R,tnR)×(r0S1D(t0R,tnR)(P0,TRSP0,TRE1))}=P0,TpD(t0A,tnA)D(t0R,tnR)×{P0,TRSP0,TRE(CtTp(TRS,TRE)1)+D(t0R,tnR)r0S}.

4. Numerics

4.1. Zero-coupon dynamics

In figure , we analyze the impact of the parameter, α in the Exponential kernel case (section 2.3.1) and H in the Riemann-Liouville case (section 2.3.2), on the dynamics of the zero-coupon bond over a time span [0,1] and considering a constant curve θ()=6%. In order to compare them properly, the underlying Brownian path is the same for all kernels. Unsurprisingly, we observe that the Riemann-Liouville case creates a lot more variance in the dynamics.

Figure 4. Dynamics of the zero-coupon bond in the Exponential (left) and the Riemann-Liouville (right) kernel case.

Figure 4. Dynamics of the zero-coupon bond in the Exponential (left) and the Riemann-Liouville (right) kernel case.

4.2. Impact of the roughness on convexity

We compare in figure  the impact of the (roughness of the) kernel on the convexity adjustment. We consider a constant curve θ()=6% and (t,t1,t2,τ)=(1,2,3,2). As α tends to zero (exponential kernel case) and as H tends to 12 (Riemann-Liouville case), the convexity adjustments converge to the same value (as expected), approximately equal to 2.718. In figure , we consider example 2.4.2, shifting away from a standard Brownian driver.

Figure 5. Left: Impact of the exponential factor α on the convexity for the Exponential kernel from section 2.3.1. Right: Impact of the Hurst exponent H on the convexity for the power-law kernel from section 2.3.2.

Figure 5. Left: Impact of the exponential factor α on the convexity for the Exponential kernel from section 2.3.1. Right: Impact of the Hurst exponent H on the convexity for the power-law kernel from section 2.3.2.

Figure 6. Left: Impact of the exponential factor α on the convexity for the Exponential kernel with standard Brownian motion (black dashed) and with OU driver with different β parameters. Right: Same but with the power-law kernel.

Figure 6. Left: Impact of the exponential factor α on the convexity for the Exponential kernel with standard Brownian motion (black dashed) and with OU driver with different β parameters. Right: Same but with the power-law kernel.

Acknowledgments

The authors would like to thank Damiano Brigo for helpful comments. ‘For the purpose of open access, the author(s) has applied a Creative Commons Attribution (CC BY) licence (where permitted by UKRI, “Open Government Licence” or “Creative Commons Attribution No-derivatives (CC BY-ND) licence” may be stated instead) to any Author Accepted Manuscript version arising’.

Disclosure statement

No potential conflict of interest was reported by the author(s).

Additional information

Funding

AJ is supported by the EPSRC [grant numbers EP/W032643/1 and EP/T032146/1].

Notes

References

  • Alòs, E., Mazet, O. and Nualart, D., Stochastic calculus with respect to Gaussian processes. Ann. Probab., 2001, 29, 766–801.
  • Andersen, L. and Piterbarg, V., Interest Rate Modeling, 2010 (Atlantic Financial Press).
  • Basse-O'Connor, A. and Graversen, S.-E., Path and semimartingale properties of chaos processes. Stoch. Process. Appl., 2010, 120, 522–540.
  • Bayer, C., Friz, P. and Gatheral, J., Pricing under rough volatility. Quant. Finance, 2016, 16, 887–904.
  • Bayer, C., Friz, P.K., Fukasawa, M., Gatheral, J., Jacquier, A. and Rosenbaum, M., Rough Volatility, 2023 (SIAM).
  • Bonesini, O., Jacquier, A. and Pannier, A., Rough Volatility, Path-Dependent PDEs and Weak Rates of Convergence, 2023. arXiv:2304.03042.
  • Brigo, D. and Mercurio, F., Interest Rate Models-Theory and Practice: With Smile, Inflation and Credit, Vol. 2, 2006 (Springer).
  • Brotherton-Ratcliffe, R. and Iben, B., Yield curve applications of swap products. In Advanced Strategies in Financial Risk Management, edited by R.J. Schwartz and C.W. Smith, pp. 400–450, 1993 (New York Institute of Finance: New York).
  • El Euch, O., Fukasawa, M. and Rosenbaum, M., The microstructural foundations of leverage effect and rough volatility. Finance Stoch., 2018, 22, 241–280.
  • Flesaker, B., Arbitrage free pricing of interest rate futures and forward contracts. J. Futures Mark., 1993, 13, 77–91.
  • Fukasawa, M., Volatility has to be rough. Quant. Finance, 2021, 21, 1–8.
  • García-Lorite, D. and Merino, R., Convexity Adjustments à la Malliavin, 2023. arXiv:2304.13402.
  • Gatheral, J., Jaisson, T. and Rosenbaum, M., Volatility is rough. Quant. Finance, 2018, 18, 933–949.
  • Hull, J. and White, A., Pricing interest-rate-derivative securities. Rev. Financ. Stud., 1990, 3, 573–592.
  • Jacquier, A., Muguruza, A. and Pannier, A., Rough Multifactor Volatility for SPX and VIX Options, 2021. arXiv:2112.14310.
  • Jacquier, A. and Oumgari, M., Deep curve-dependent PDEs for affine rough volatility. SIAM J. Financ. Math., 2023, 14, 353–382.
  • Kirikos, G. and Novak, D., Convexity conundrums: Presenting a treatment of swap convexity in the hall-white framework. Risk Mag., 1997, 10, 60–61.
  • Øksendal, B., Stochastic Differential Equations, 2003 (Springer).
  • Pelsser, A., Mathematical foundation of convexity correction. Quant. Finance, 2003, 3, 59–65.
  • Ritchken, P. and Sankarasubramanian, L., Averaging and deferred payment yield agreements. J. Futures Mark., 1993, 13, 23–41.
  • Russo, F. and Tudor, C.A., On bifractional Brownian motion. Stoch. Process. Appl., 2006, 116, 830–856.
  • Vasicek, O., An equilibrium characterization of the term structure. J. Financ. Econ., 1977, 5, 177–188.