Fixed Income Probability Measures

Interest Rate Modeling




Tutorial Contents

  1. Fixed Income Notation
  2. Fixed Income Probability Measures
  3. Multi-Currency Markets
  4. The HJM Framework
  5. Basic Fixed Income Instruments
  6. Caps, Floors and Swaptions
  7. CMS Swaps, Caps and Floors
  8. Bermudan Swaptions
  9. Structured Notes and Exotic Swaps
  10. Callable Libor Exotics and TARNs
  11. Volatility Derivatives
  12. Yield Curves and Curve Construction
  13. Managing Yield Curve Risk
  14. Vanilla Models With Local Volatility
  15. Vanilla Models With Stochastic Volatility
  16. Introduction to Term Structure Models
  17. One-Factor Short Rate Models I
  18. One-Factor Short Rate Models II
  19. Multi-Factor Short Rate Models
  20. The Libor Market Model I
  21. The Libor Market Model II
  22. Risk Management and Sensitivity Computations
  23. P&L Attribution
  24. Payoff Smoothing
  25. Vegas in the Libor Market Model


Crucial to the modeling of interest rate derivatives is the risk-neutral pricing framework. This framework is expressed in the language of stochastic processes and two particularly important concepts are that of an equivalent martingale measure and a numeraire. In this post, we will introduce the theory behind these concepts and then give some common and useful measures.

Risk-Neutral Pricing

In this section, we review the key results from stochastic calculus and modern asset pricing theory. The results we uncover will give us a framework for pricing assets in a way that is consistent with the idea of no-arbitrage, a key assumption to mathematical finance. The following is technical and it is not crucial that you understand the theory inside out. The only crucial result is that of the equivalent martingale measure and corresponding numeraire. Keep that in mind if you find this section too advanced.

Setup

Consider an economy with continuous and frictionless trading, taking place on a finite time horizon [0,T][0,T]. We assume the existence of pp traded, dividend-free assets, whose prices are characterized by the vector-valued stochastic process X(t)=(X1(t),,Xp(t))X(t) = (X_1(t),\ldots, X_p(t))^\intercal. Here, a stochastic process is simply a continuous set of vector-valued random variables, indexed by the time tt.

To model uncertainty and the flow of information, we introduce a probability space (Ω,F,P)(\Omega, \mathcal{F}, \mathbb{P}). Here, Ω\Omega is a sample space with elements ω\omega representing states of the world (or outcomes) that may materialise. F\mathcal{F} represents a specific σ\sigma-algebra on Ω\Omega (we will later see which one), effectively a family of subsets of outcomes from Ω\Omega that satisfies certain conditions. P\mathbb{P} represents a probability measure on the measure space (Ω,F)(\Omega, \mathcal{F}). The details of the above setup require an understanding of the field of measure theory. Again, it is not important you understand this thoroughly. Just know that for each AFA \in \mathcal{F} we find that AΩA \subset \Omega and that P:AF[0,1]\mathbb{P} : A \in \mathcal{F} \rightarrow [0, 1], i.e. P\mathbb{P} is a function that takes sets of outcomes and maps them to a number between 0 and 1 in such a way that each of these mappings can be considered a probability and P\mathbb{P} can be considered a probability measure.

We model the flow of information through time via a filtration F={F(t):t[0,T]}\mathbb{F} = \{\mathcal{F}(t) : t \in [0,T]\}, a family of sub-σ\sigma-algebras of F\mathcal{F} satisfying F(s)F(t)F\mathcal{F}(s) \subset \mathcal{F}(t) \subset \mathcal{F} whenever sts \le t. The best way to think about F(t)\mathcal{F}(t) is as the information available at time tt. As time passes, more and more information is revealed, so that F(t)\mathcal{F}(t) incorporates all of the previous information up to time sts \le t, plus any additional information gained from ss to tt.

We will assume that the process X(t)X(t) is adapted to F(t)\mathcal{F}(t) for all tt. Practically speaking, this just means that X(t)X(t) is fully observable at time tt. Finally, we require that the filtration F\mathbb{F} satisifes the usual conditions, which I will leave to the interested reader to explore further.

Let EP[]\mathbb{E}^\mathbb{P}[\cdot] represent the expectation operator for the measure P\mathbb{P}. When conditioning on information up to time tt, we will use the notation EtP[]:=EP[F(t)]\mathbb{E}_t^\mathbb{P}[\cdot] := \mathbb{E}^\mathbb{P}[\cdot \vert \mathcal{F}(t)].

The above setup is fairly flexible. For our purposes, we focus on the situation where information is completely governed by a dd-dimensional vector-valued Brownian Motion W(t)=(W1(t),,Wd(t))W(t) = (W_1(t),\ldots,W_d(t))^\intercal, where WiW_i is independent of WjW_j for all iji \neq j. Again, it is not important to understand the technical details of Brownian motion on any particularly deep level. For now it suffices to say: W(0)=0W(0) = 0, Wi(t)Wi(s)N(0,ts)W_i(t) - W_i(s) \sim \mathcal{N}(0, t-s) for tst \ge s and that each of these increments are independent of one another for non-overlapping times.

Above, we mentioned that F\mathcal{F} represents a specific σ\sigma-algebra on Ω\Omega. In fact, it is the one generated by WW, i.e., for all tt we have F(t)=σ({W(u),u[0,t]})\mathcal{F}(t) = \sigma \left(\{W(u), u \in [0,t] \} \right) and F=F(T)\mathcal{F} = \mathcal{F}(T).

We will assume that X(t)X(t) is described by the vector-valued Ito process

dX(t)=μ(t,ω)dt+σ(t,ω)dW(s),(2.1) \text{d}X(t) = \mu(t,\omega)\text{d}t + \sigma(t,\omega)\text{d}W(s), \tag{2.1}

where μ(s,ω)Rp\mu(s,\omega) \in \mathbb{R}^p and σ(t,ω)Rp×d\sigma(t,\omega)\in \mathbb{R}^{p\times d} are both themselves F\mathbb{F} adapted processes.

An incredibly powerful and useful concept that we require for our framework is that of a martingale. Before giving the precise definition, I will note that the concept of a martingale is really a simple one. A process is a martingale if the best guess for where the process will be in the future is where it is today. In this way, it represents a fair game, in the sense that if the martingale's current value represented the cost of entering into a bet, and it's future value your payoff from the bet, then you would on average expect to come out flat from playing the game.

Definition (Martingale)

Let Y(t)Y(t) be an adapted vector-valued process with EP[Y(t)]<\mathbb{E}^P\left[\left\vert Y(t) \right\vert\right] < \infty for all t[0,T]t \in [0,T]. We say that Y(t)Y(t) is a martingale under the measure P\mathbb{P} if, for all 0stT0 \le s \le t \le T,

EtP[Y(s)]=Y(t),a.s. \mathbb{E}_t^\mathbb{P}[Y(s)] = Y(t), \quad \text{a.s.}

The second important concept that we require is known as Ito's Lemma. Ito's lemma is essentially the stochastic version of the differential of a function. If we consider a simple scalar-valued function f(x,y)f(x,y) then the standard differential of ff is clearly just given by

df(x,y)=f(x,y)xdx+f(x,y)ydy. \text{d}f(x,y) = \frac{\partial f(x,y)}{\partial x} \text{d}x + \frac{\partial f(x,y)}{\partial y} \text{d} y.

This follows as the first order limit to the Taylor expansion

Δf(x,y)=f(x,y)xΔx+f(x,y)yΔy+122fx2Δx2+122fy2Δy2+2fxyΔxΔy+H.O.T.\begin{aligned} \Delta f(x,y) & = \frac{\partial f(x,y)}{\partial x} \Delta x + \frac{\partial f(x,y)}{\partial y} \Delta y \\ & \quad \quad \quad + \frac{1}{2} \frac{\partial^2 f}{\partial x^2} \Delta x^2 + \frac{1}{2} \frac{\partial^2 f}{\partial y^2} \Delta y^2 \\ & \quad\quad \quad + \frac{\partial^2 f}{\partial x \partial y} \Delta x \Delta y + \text{H.O.T.} \end{aligned}

where H.O.T. represents the higher-order terms in the expansion, Δf=:f(x+Δx,y+Δy)f(x,y)\Delta f = :f(x + \Delta x, y + \Delta y) - f(x,y) and Δx,Δy>0\Delta x, \Delta y > 0.

This is all well and good while xx and yy are deterministic. If we were to assume that x=tx = t was our deterministic time variable, but y=X(t)y = X(t) was some stochastic variable of the form (2.1), then we have a different story. The notion of a differential breaks, since our higher order terms are no longer well-behaved. In particular, we have (dX(t))2=σ(t,ω)2(dW(t))2+H.O.T.(\text{d}X(t))^2 = \sigma(t,\omega)^2 (\text{d}W(t))^2 + \text{H.O.T.}, where it can be shown that (dW(t))2=dt(\text{d}W(t))^2 = \text{d}t in quadratic mean. Thus, we see that in the first order limit, the stochastic case leads to differential

df(t,X(t))=ftdt+fx(μ(t,ω)dt+σ(t,ω)dW(t))+12σ(t,ω)22fx2dt=(ft+μfx+12σ22fx2)dt+σfxdW(t).\begin{aligned} \text{d} f(t,X(t)) & = \frac{\partial f}{\partial t} \text{d} t + \frac{\partial f}{\partial x}(\mu(t,\omega)\text{d}t + \sigma(t,\omega)\text{d}W(t)) + \frac{1}{2} \sigma(t,\omega)^2 \frac{\partial^2 f}{\partial x^2} \text{d}t \\ & = \left(\frac{\partial f}{\partial t} + \mu \frac{\partial f}{\partial x} +\frac{1}{2}\sigma^2 \frac{\partial^2 f}{\partial x^2}\right)\text{d}t + \sigma \frac{\partial f}{\partial x} \text{d}W(t). \end{aligned}

Ito's Lemma extends this result to scalar-valued functions of multiple variables.

Theorem (Ito's Lemma)

Let f(t,x)f(t,x), x=(x1,,xp)x = (x_1,\ldots,x_p)^\intercal, denote a continuous function, let f(t,x)Rf(t,x) \in \mathbb{R} for all t[0,T]t \in [0,T], xRpx \in \mathbb{R}^p, and let fC1,2([0,T]×Rp)f \in C^{1,2}([0,T]\times \mathbb{R}^p). Let X(t)X(t) be given by (2. 1) and define a scalar process Y(t):=f(t,X(t))Y(t) := f(t,X(t)). Let xf\nabla_x f represent the pp-dimensional vector of partial derivatives of ff with respect to the non-time variables. Likewise, let HxfH_x f represent the Hessian of ff with respect to the same variables. Then Y(t)Y(t) is an Ito process with stochastic differential given by

dY(t)=f(t,X(t))tdt+xf(t,X(t))μdt+xf(t,X(t))σdW(t)+12σHxf(t,X(t))σdt.\begin{aligned} \text{d} Y(t) & = \frac{\partial f(t,X(t))}{\partial t} \text{d} t + \nabla_x f(t,X(t)) \mu \text{d} t + \nabla_x f(t,X(t)) \sigma \text{d}W(t) + \frac{1}{2} \sigma^\intercal H_x f(t,X(t)) \sigma \text{d}t. \end{aligned}

Trading Gains and Arbitrage

Consider an investor engaging in a trading strategy involving the pp assets X1,,XpX_1,\ldots,X_p. Let the trading strategy be characterized by a predictable adapted process ϕ(t,ω)=(ϕ1(t,ω),,ϕp(t,ω))\phi(t,\omega) = (\phi_1(t,\omega), \ldots, \phi_p(t,\omega))^\intercal, with ϕi(t,ω)\phi_i(t,\omega) denoting the holding at time tt in the iith asset XiX_i. The value π(t)\pi(t) of the trading strategy at time tt is thus given by

π(t)=ϕ(t)X(t). \pi(t) = \phi(t) \cdot X(t).

Clearly, over a small time interval [t,t+δ][t,t+\delta], the gain from trading is given by ϕ(t)(X(t+δ)X(t))\phi(t) \cdot (X(t+\delta) - X(t)), suggesting that

ϕ(t)dX(t)=ϕ(t)μ(t)dt+ϕ(t)σ(t)dW(t),\begin{aligned} & \phi(t)\cdot \text{d} X(t) \\ & = \phi(t) \cdot \mu(t)\text{d}t + \phi(t) \cdot \sigma(t)\text{d}W(t), \end{aligned}

is an appropriate model for trading gains in an infinitesimal interval.

A trading strategy π\pi is said to be self-financing if, for any t[0,T]t \in [0,T],

dπ(t)=ϕ(t)dX(t). \text{d}\pi(t) = \phi(t) \cdot \text{d} X(t).

This relationship expresses the concept that changes in portfolio value are solely caused by trading gains or losses, with no funds being added or withdrawn.

Definition (Arbitrage)

An arbitrage opportunity is a self-financing strategy ϕ\phi for which π(0)=0\pi(0) = 0 and, for some t[0,T]t \in [0,T],

π(t)0 a.s.andP(π(t)>0)>0. \pi(t) \ge 0 \text{ a.s.} \quad \text{and} \quad \mathbb{P}(\pi(t) > 0) > 0.

In words, an arbitrage opportunity is a strategy that creates something from nothing, a free lunch. In a well-functioning market, arbitrage opportunities cannot exist. If they did, they would surely be exploited until the market returned to equilibrium at a level that eliminated the opportunity. No arbitrage will form a key requirement for all of the markets that we will consider.

Equivalent Martingale Measures, Complete Markets and Arbitrage

Currently, the idea of an arbitrage opportunity is an abstract one that bears no relation to the rest of our setup. However, one of the key insights of modern asset pricing is that under certain conditions, we can encapsulate no-arbitrage via the concept of an equivalent martingale measures.

First, note that two probability measures P\mathbb{P} and P^\hat{\mathbb{P}} on the same measure space (Ω,F)(\Omega, \mathcal{F}) are equivalent if P(A)=0    P^(A)=0\mathbb{P}(A) = 0 \iff \hat{\mathbb{P}}(A) = 0 for all AFA \in \mathcal{F}. Shown via the Radon-Nikodymn Theorem, two equivalent measures are connected through a process RR with EP[R]=1\mathbb{E}^\mathbb{P}[R] = 1 such that P^(A)=EP[R1A]\hat{\mathbb{P}}(A) = \mathbb{E}^\mathbb{P}[R 1_A] for all AFA \in \mathcal{F}, known as the Radon-Nikodym derivative. RR is often denoted by dP^dP\tfrac{\text{d}\hat{\mathbb{P}}}{\text{d}\mathbb{P}}. Now, an equivalent martingale measure for the process Z(t)Z(t) is simply an equivalent measure P^\hat{\mathbb{P}} such that under P^\hat{\mathbb{P}}, Z(t)Z(t) is a martingale.

With these concept in place, we now introduce the important concept of a numeraire.

Definition (Numeraire)

A numeraire is a traded asset (a linear combination of our pp assets) with strictly positive price that pays no dividends, used to normalise the asset process. Denoting the numeraire value at time tt by N(t)N(t), the normalized asset process is given by XN(t):=X(t)N(t)X^N(t) := \tfrac{X(t)}{N(t)}.

We say that a measure QN\mathbb{Q}^N is an equivalent martingale measure induced by NN if XN(t)X^N(t) is a martingale under QN\mathbb{Q}^N. We call a trading strategy ϕ\phi permissible if

0tϕ(s)dXN(s), \int_0^t \phi(s)^\intercal \text{d} X^N(s),

is a martingale under QN\mathbb{Q}^N.

The last concept that we introduce is that of a complete market. Roughly, a complete market is one where the assets X(t)X(t) are the only ones that matter, in the sense that all other assets can be generated as a linear combination of the XiX_is.

Definition (Complete Market)

A complete market is one in which all finite-variance F(T)\mathcal{F}(T)-measurable random variables V(T)V(T) can be replicated, in the sense that there exists a permisible trading strategy ϕ\phi such that V(T)=ϕ(T)X(T)=π(T)V(T) = \phi(T) \cdot X(T) = \pi(T).

Via the fundamental theorems of asset pricing, we are now in a position to link all of these concepts, in a way that gives us a practical way of pricing derivatives.

Theorem (The Fundamental Theorems of Asset Pricing)

Restricting attention to permissible trading strategies, the following two equivalencies hold:

  1. Our economy is arbitrage-free if and only if for each numeraire NN, there exists at least one equivalent martingale measure QN\mathbb{Q}^N.
  2. An arbitrage-free market is complete if and only if for each numeraire NN, the risk neutral measure QN\mathbb{Q}^N is unique.

Using the above, if we assume that our market is both arbitrage-free and complete then for each numeraire NN, there is a corresponding equivalent martingale measure QN\mathbb{Q}^N. Hence, we can unambigously talk about the numeraire, measure pair (N,QN)(N,\mathbb{Q}^N). Under this pair, any TT-maturity derivative security paying out the random payoff V(T)V(T) at time TT can be expressed via

V(t)N(t)=EtN[V(T)N(T)], \frac{V(t)}{N(t)} = \mathbb{E}_t^N\left[\frac{V(T)}{N(T)} \right],

or

V(t)=N(t)EtN[V(T)N(T)], V(t) = N(t) \mathbb{E}_t^N \left[ \frac{V(T)}{N(T)} \right],

where EtN[]:=EQN[F(t)]\mathbb{E}_t^N[\cdot] := \mathbb{E}^{\mathbb{Q}^N}[\cdot \vert \mathcal{F}(t)].

Clearly, since NN is arbitrary in the above, if have another numeraire pair (M,QM)(M, \mathbb{Q}^M) then

N(t)EtN[V(T)N(T)]=M(t)EtM[V(T)M(T)]. N(t) \mathbb{E}_t^N \left[ \frac{V(T)}{N(T)} \right] = M(t) \mathbb{E}_t^M \left[ \frac{V(T)}{M(T)} \right].

If we consider the time TT payoff V(T)=Y(T)M(T)V(T) = Y(T)M(T) then this shows that

EtN[N(t)M(T)N(T)M(t)Y(T)]=EtM[Y(T)], \mathbb{E}_t^N \left[ \frac{N(t) M(T)}{N(T) M(t)} Y(T) \right] = \mathbb{E}_t^M \left[Y(T) \right],

By comparison, this shows that the Radon-Nikodym derivative for the measure change between measures QN\mathbb{Q}^N and QM\mathbb{Q}^M is characterised by the process

η(t)=EtN[dQMdQN]=M(t)/M(0)N(t)/N(0).(2.2)\begin{aligned} \eta(t) & = \mathbb{E}_t^N \left[ \frac{\text{d}\mathbb{Q}^M}{\text{d}\mathbb{Q}^N}\right] = \frac{M(t)/M(0)}{N(t)/N(0)}. \tag{2.2} \end{aligned}

Girsanov's Theorem

We have established that there is a close link between the concept of arbitrage and the existence and uniqueness of an equivalent martingale measure. We saw that there is even an easy way to characterize the Radon-Nikodym derivative linking two equivalent martingale measures via their corresponding numeraires. From a practical perspective, this does not tell us a huge amount about how we convert our asset price dynamics when acting under different numeraires. However, in the setup that we have introduced above, there is a practical way of switching between measures.

Consider two measures, P\mathbb{P} and P(θ)\mathbb{P}(\theta) related by the density ηθ(t)=EtP[dP(θ)dP]\eta^\theta(t) = \mathbb{E}_t^{\mathbb{P}}\left[\tfrac{\text{d}\mathbb{P}(\theta)}{\text{d}\mathbb{P}} \right], where ηθ(t)\eta^\theta(t) is an exponential martingale given by the Ito process

dηθ(t)=ηθ(t)dW(t). \text{d}\eta^\theta(t) = -\eta^\theta(t)^\intercal \text{d}W(t).

The dd-dimensional process θ\theta is known as the market price of risk. Assuming that ηθ\eta^\theta defines a martingale, we have the following important result:

Theorem (Girsanov's Theorem)

Suppose that ηθ(t)\eta^\theta(t) defined above is a martingale, then for all t[0,T]t \in [0, T]

Wθ(t)=W(t)+0tθ(s)ds W^\theta(t) = W(t) + \int_0^t \theta(s) \text{d}s

is a Brownian motion under the measure P(θ)\mathbb{P}(\theta).

Fixed Income Probability Measures

As we have seen, selecting an equivalent martingale measure is largely a matter of choosing a numeraire. Here, we go through a number of important numeraires used in fixed income pricing. Throughout, we assume that our market precludes arbitrage and is complete, and we use V(t)V(t) to denote the time tt price of a derivative security making an F(T)\mathcal{F}(T)-measurable payment of V(T)V(T).

The Risk Neutral Measure

The most fundamental numeraire in the fixed-income world (and even more generally in derivatives pricing) is the continuously compounded money market account β(t)\beta(t), satisfying the locally deterministic SDE

dβ(t)=r(t)β(t)dt,β(0)=1, \text{d}\beta(t) = r(t)\beta(t)\text{d}t, \quad \beta(0) = 1,

where r(t)r(t) is the same short rate that we saw in Section 1. Solving this SDE shows that

β(t)=exp(0tr(s)ds).(2.3) \beta(t) = \exp\left(\int_0^t r(s) \text{d} s\right). \tag{2.3}

We denote the equivalent martingale measure associated with β\beta by Q\mathbb{Q}. Under this measure, we find that

V(t)=β(t)EtQ[V(T)β(T)],tT. V(t) = \beta(t)\mathbb{E}_t^\mathbb{Q}\left[\frac{V(T)}{\beta(T)} \right], \quad t \le T.

Utilising the explicit formula (2.3), we find that the above is equivalent to

V(t)=EtQ[exp(tTr(s)ds)V(T)].(2.4) V(t) = \mathbb{E}_t^\mathbb{Q}\left[\exp\left(-\int_t^T r(s)\text{d}s\right)V(T) \right]. \tag{2.4}

As we saw in the Section 1, a ZCB pays out 1 unit at its maturity TT. Hence, if we set V(T)=1V(T) = 1 then V(t)=P(t,T)V(t) = P(t,T) for tTt \le T and we arrive at the bond pricing formula

P(t,T)=EtQ[exp(tTr(s)ds)].(2.5) P(t,T) = \mathbb{E}_t^\mathbb{Q}\left[\exp\left(-\int_t^T r(s)\text{d}s\right)\right]. \tag{2.5}

This result is important because it shows that specification of the dynamics of the short-rate r(t)r(t) under Q\mathbb{Q} is enough to determine ZCB prices at all times and maturities. Models based on such a direct specification are known as short rate models. Note the similarity between (2.5) and (1.1). Clearly, if r(t)r(t) is deterministic then the two expressions will agree and r(u)=f(t,u)r(u) = f(t,u)for all utu \ge t. However, if r(t)r(t) is random, these two results diverge, and we find that f(t,u)EtQ[r(u)]f(t,u) \neq \mathbb{E}_t^\mathbb{Q}[r(u)]. However, if we were to replace f(t,u)f(t,u) by the equivalent futures rate q(t,u)q(t,u) then the above does hold. The exact result goes as follows.

Lemma

Assuming that mark-to-market takes place continuously, under regularity conditions on rr, futures rate F(,T,T+τ)F(\cdot, T, T+\tau) is a Q\mathbb{Q}-martingale, and

F(t,T,T+τ)=EtQ[L(T,T,T+τ)]. F(t,T,T+\tau) = \mathbb{E}_t^\mathbb{Q}[L(T,T,T+\tau)].

Proving this result is relatively simple. We exploit the fact that the mark-to-market process leads to daily cash proceeds proportional to dF(t,T,T+τ)=F(t+dt,T,T+τ)F(t,T,T+τ)\text{d}F(t,T,T+\tau) = F(t+\text{d}t, T, T+\tau) - F(t,T,T+\tau). Deflating each of these back to the current time tt and noting that the futures value is always zero shows that holding the futures contract until an arbitrary time s[t,T]s \in [t,T] leads to the expression V(t)=0=EtQ[tsβ(s)1dF(s,T,T+τ)]V(t) = 0 = \mathbb{E}_t^\mathbb{Q}\left[\int_t^s \beta(s)^{-1} \text{d}F(s,T,T+\tau) \right]. Since this holds for arbitrary time horizons, we deduce that

EtQ[dF(s,T,T+τ)]=0,s[t,T], \mathbb{E}_t^\mathbb{Q}\left[\text{d}F(s, T, T+\tau) \right] = 0, \quad s \in [t,T],

which implies that FF is a Q\mathbb{Q}-martingale. The result follows immediately.

Taking the limit as τ0\tau \searrow 0 on both sides of the above then gives our result for the instantaneous futures rate, i.e.

q(t,u)=EtQ[r(u)]. q(t,u) = \mathbb{E}_t^\mathbb{Q}[r(u)].

The T-Forward Measure

The TT-forward measure QT\mathbb{Q}^T uses the TT-bond as numeraire. We will denote the conditional expectation operator under the TT-forward measure by EtT[]\mathbb{E}_t^T[\cdot], so that

V(t)=P(t,T)EtT[V(T)P(T,T)]=P(t,T)EtT[V(T)].(2.6)\begin{aligned} V(t) & = P(t,T)\mathbb{E}_t^T\left[ \frac{V(T)}{P(T,T)} \right] \\ & = P(t,T)\mathbb{E}_t^T\left[ V(T) \right]. \tag{2.6} \end{aligned}

By comparing (2.4) and (2.6), it should be clear that switching from the risk-neutral to the TT-forward measure decouples the expectation of the discount factor D(t,T):=β(t)β(T)D(t,T) := \tfrac{\beta(t)}{\beta(T)} from the expectation of the payoff V(T)V(T). This is often convenient when pricing IR derivatives. From equation (2.2), we see that the change of measure process from the risk-neutral to the TT-forward measure is given by density

EtQ[dQTdQ]=P(t,T)/P(0,T)β(t). \mathbb{E}_t^\mathbb{Q}\left[\frac{\text{d}\mathbb{Q}^T}{\text{d}\mathbb{Q}} \right] = \frac{P(t,T)/P(0,T)}{\beta(t)}.

Since the (T+τ)(T+\tau)-bond is another traded asset and has price P(t,T+τ)P(t,T+\tau), we find that by definition, the time tt forward bond price P(t,T,T+τ)=P(t,T+τ)P(t,T)P(t,T, T+\tau) = \tfrac{P(t,T+\tau)}{P(t,T)} is a martingale under the TT-forward measure. This is, in fact, where the name TT-forward measure comes from.

While the forward bond price P(t,T,T+τ)P(t, T, T+\tau) is a martingale under the TT forward measure, we find that the forward Libor rate L(t,T,T+τ)L(t,T,T+\tau) is a martingale under the (T+τ)(T+\tau)-forward measure QT+τ\mathbb{Q}^{T+\tau}. This follows immediately from the relation

L(t,T,T+τ)=1τ(P(t,T)P(t,T+τ)1). L(t,T,T+\tau)=\frac{1}{\tau}\left(\frac{P(t,T)}{P(t,T+\tau)} - 1\right).

Using this fact, we find that

f(t,u)=limτ0L(t,u,u+τ)=limτ0Etu+τ[L(u,u,u+τ)]=Etu[f(u,u)]=Etu[r(u)],\begin{aligned} f(t,u) & = \lim_{\tau \searrow 0} L(t, u, u + \tau) \\ & = \lim_{\tau \searrow 0} \mathbb{E}_t^{u + \tau}[L(u, u, u + \tau)] \\ & = \mathbb{E}_t^u [f(u,u)] \\ & = \mathbb{E}_t^u[r(u)], \end{aligned}

a result analogous to that of the futures rate result under the risk-neutral measure.

The Spot Measure

When working on a full tenor structure 0=T0<<TN0 = T_0 < \cdots < T_N of forward rates, it can be inconvenient to work with a particular forward measure TnT_n. Evolving through time, once t>Tnt > T_n, our numeraire is invalidated and we must switch measure to one that is still live. Another way to avoid this issue is by introducing a numeraire that extends to arbitrary horizons by construction. While the money market account β\beta accomplishes this, it can be awkward to work with an inherently continuous construct when dealing with a discrete tenor structure. An alternative is the discrete-time equivalent to β\beta.

At time 00, we invest £11 in 1P(0,T1)\tfrac{1}{P(0,T_1)} T1T_1-bonds. Clearly, at T1T_1, this investment returns

1P(0,T1)=1+τ0L(0,0,T1). \frac{1}{P(0,T_1)} = 1 + \tau_0 L(0,0,T_1).

Reinvesting (rolling) this amount at T1T_1 into T2T_2-bonds then returns

1P(0,T1)1P(T1,T2)=(1+τ0L(0,0,T1))(1+τ1L(T1,T1,T2))\begin{aligned} \frac{1}{P(0,T_1)}\cdot\frac{1}{P(T_1,T_2)} & = (1 + \tau_0 L(0,0,T_1))(1 + \tau_1 L(T_1,T_1,T_2)) \end{aligned}

at T2T_2. While t(T1,T2]t \in (T_1, T_2], this investment clearly has value 1P(0,T1)P(t,T2)P(T1,T2)\frac{1}{P(0,T_1)}\cdot\frac{P(t,T_2)}{P(T_1,T_2)}. By repeating this strategy at each date in the tenor structure, we arrive at an investment with time tt value B(t)B(t) given by

B(t)=i=0n(1+τ0Li(Ti))P(t,Tn+1),t(Tn,Tn+1],\begin{aligned} B(t) & = \prod_{i = 0}^n (1 + \tau_0 L_i(T_i))P(t,T_{n+1}), \quad\quad t \in (T_n, T_{n+1}], \end{aligned}

where we have again used the shorthand Ln(t)=L(t,Tn,Tn+1)L_n(t) = L(t,T_n,T_{n+1}).

Since we constructed B(t)B(t) through a permissible trading strategy and B(t)>0B(t) > 0 for all tt, B(t)B(t) is a valid numeraire. The measure generated by B(t)B(t) is called the spot (Libor) measure, denoted by QB\mathbb{Q}^B with associated conditional expectation EtB[]\mathbb{E}_t^B[\cdot]. Then

V(t)=EtB[V(T)B(t)B(T)], V(t) = \mathbb{E}_t^B\left[ V(T) \frac{B(t)}{B(T)}\right],

where

B(t)B(T)=P(t,Tn+1)P(T,Tk+1)i=n+1k11+τkLk(Tk),t(Tn,Tn+1],T(Tk,Tk+1].\begin{aligned} \frac{B(t)}{B(T)} & = \frac{P(t,T_{n+1})}{P(T,T_{k+1})} \prod_{i = n+1}^k \frac{1}{1 + \tau_k L_k(T_k)}, \quad\quad t \in (T_n, T_{n+1}], T \in (T_k,T_{k+1}]. \end{aligned}

It is clear that the spot measure resembles the risk neutral measure, since the risk-neutral measure is in some sense its continuous counterpart. In fact, if the futures rate is marked-to-market discretely, on dates T0,,TNT_0,\ldots,T_N, it turns out to be a martingale under the spot measure.

Terminal and Hybrid Measures

The big advantage of the spot measure that we discussed was that it remains alive throughout the span of the tenor structure. This property is key for valuing securities which may mature at any date in the tenor structure (e.g. barriers or range accruals).

Clearly, picking the TT-forward measure corresponding to the final maturity in the tenor structure, i.e. T=TNT = T_N also satisfies this neat property. The measure induced by P(t,TN)P(t,T_N), QN\mathbb{Q}^N is called the terminal measure. As usual,

V(t)=P(t,TN)EtN[V(T)P(T,TN)]. V(t) = P(t,T_N)\mathbb{E}_t^N\left[\frac{V(T)}{P(T,T_N)} \right].

One interpretation of the term V(T)P(T,TN)\tfrac{V(T)}{P(T,T_N)} is as the time TNT_N proceeds of rolling the payoff V(T)V(T) into the TNT_N-bond. Alternatively, we could roll V(T)V(T) into the spot numeraire asset B(T)B(T), leading to TNT_N payoff V(T)B(TN)B(T)V(T) \cdot \tfrac{B(T_N)}{B(T)}. This gives rise to the equivalent formula

V(t)=P(t,TN)EtN[V(T)B(TN)B(T)]. V(t) = P(t,T_N)\mathbb{E}_t^N\left[V(T) \frac{B(T_N)}{B(T)} \right].

Clearly, this idea generalises. For instance, we can define a numeraire such as

P^(t,T):={P(t,T),tT,B(t)B(T),t>T. \hat{P}(t,T) := \begin{cases} P(t,T), & t \le T, \\ \tfrac{B(t)}{B(T)}, & t > T. \end{cases}

This asset corresponds to the strategy of initially investing into the TT-bond and then rolling our proceeds into the spot measure at TT. Letting Q^T\hat{\mathbb{Q}}^T denote the corresponding measure with conditional expectation operator E^tT\hat{\mathbb{E}}_t^T, we can write

V(t)=P^(t,T)E^tT[V(TN)P^(TN,T)]=P^(t,T)E^tT[V(TN)B(T)B(TN)],T<TN,t<TN. \begin{aligned} V(t) & = \hat{P}(t,T) \hat{\mathbb{E}}_t^T\left[ \frac{V(T_N)}{\hat{P}(T_N, T)} \right] \\ & = \hat{P}(t,T)\hat{\mathbb{E}}_t^T \left[V(T_N) \frac{B(T)}{B(T_N)}\right], \quad T < T_N, t < T_N. \end{aligned}

The associated measure to such numeraire is known as a hybrid measure. Clearly, many such hybrid measures exist and can be useful for different situations.

Swap Measures

Since the annuity factor Ak,m(t)A_{k,m}(t) of Section 1 is just a linear combination of ZCBs, it qualifies as a numeraire asset. The associated measure Qk,m\mathbb{Q}^{k,m} is known as a swap measure or annuity measure. We have

V(t)=Ak,m(t)Etk,m[V(T)Ak,m(T)], V(t) = A_{k,m}(t) \mathbb{E}_t^{k,m}\left[\frac{V(T)}{A_{k,m}(T)} \right],

where Etk,m[]\mathbb{E}_t^{k,m}[\cdot] denotes the expectation operator under Qk,m\mathbb{Q}^{k,m}.

In the absence of arbitrage, the forward swap rate Sk,m(t)S_{k,m}(t), which we have seen is defined by

Sk,m(t)=P(t,Tk)P(t,Tk+m)Ak,m(t) S_{k,m}(t) = \frac{P(t,T_k) - P(t,T_{k+m})}{A_{k,m}(t)}

is clearly a martingale under Qk,m\mathbb{Q}^{k,m}, giving the measure its name.

© 2022 Will Douglas, All Rights Reserved.