6,105
Views
2
CrossRef citations to date
0
Altmetric
Letters

Interest-rate simulation under the real-world measure within a Gaussian HJM framework

Pages 10-16 | Received 28 Jun 2013, Accepted 14 Nov 2014, Published online: 12 Jan 2015

Abstract

This paper studies an interest-rate simulation for risk management under the real-world measure. First, this paper proposes a method to estimate the market price of risk from historical data in a Gaussian Heath, Jarrow, and Morton framework. Next, properties of the simulation are examined in connection with historical data. Finally, the market price of risk is roughly interpreted in regard to the historical change in interest rates. These results are explained through numerical examples.

1. Introduction

The term structure model of interest rates was originally used for option pricing under a risk-neutral measure; recently, it has been applied in assessment of interest-rate risk, where the model should be constructed under the real-world measure. To simulate interest-rate dynamics under the real-world measure, it is necessary to estimate the market price of risk. In practice, the market price of risk has been assumed to be zero because how to estimate the market price of risk is not obvious. Moreover, the properties of the simulation in the real world have yet to be fully elucidated, and the numerical differences between a real-world simulation and a risk-neutral simulation are not known.Footnote

In econometrics, the market price of risk has been estimated by short-rate models. For example, CitationStanton (1997) used regression analysis on historical yield data, and CitationDempster et al. (2010) used the Kalman filter. These approaches use econometric software to numerically estimate the market price of risk, and for this reason, theoretical investigation of the properties of real-world simulation remains difficult.

CitationYasuoka (2013) introduced a LIBOR market model (LMM) under the real-world measure in the framework of CitationJamshidian (1997), and proposed a procedure to estimate the market price of risk under the assumption that this price is constant. Through this approach, the following results were obtained: (1) The simulation model in the LMM is similar to an empirical term structure model with historical drift and volatility. (2) It is shown how to determine the number of factors for simulation in connection with historical data. (3) When LIBOR volatility is low, the market price of risk is roughly explained by changes in the historical forward LIBOR curve, rather than by the state of the curve.

Although the LMM is useful to simulate the flexible dynamics of the yield curve, some practitioners prefer to use a short-rate model for risk management. For example, the Hull–White model (CitationHull and White, 1990), a special case of the Heath, Jarrow, and Morton (HJM) model (CitationHeath et al., 1992), is popular. Therefore, it is reasonable to perform a study similar to CitationYasuoka (2013) on the HJM model. In this respect, the objective of this paper is to introduce a procedure for real-world simulation in a Gaussian HJM framework and to use this to study the market price of risk and properties of the simulation.

Section 2 introduces a procedure to estimate the market price of risk from historical data. Beginning with a full-factor model, it is shown that the real-world simulation model is an empirical term structure with historical drift and volatility. Section 3 studies how to decide a practical choice for the number of factors in the real-world simulation. Furthermore, the numerical meaning of the market price of risk is investigated in regard to historical trends in interest rates. Some studies have attempted to explain the market price of risk through state variables, for example the instantaneous spot rate by CitationStanton (1997). In contrast, our results explain the market price of risk through changes in the historical interest-rate curve, rather than through the curve state. Finally, Section 4 explains the obtained results through numerical examples.

2. Market price of risk and simulation

2.1. The HJM framework

This section briefly summarizes the framework of the HJM model. Let be a probability space, where T* is a time horizon, Ft is an augmented filtration, and P is the original measure. P is usually called the real-world measure (or alternatively, historical measure, etc.). f(t, T) denotes the instantaneous forward rate (hereinafter, the forward rate) with maturity TT* prevailing at time tT. We denote the inner product by · in . The dynamics of f(t, T) is assumed to be expressed by (1) where Z is a d-dimensional -Brownian motion, and α(t, T) and σ(t, T) are predictable processes satisfying some technical conditions (cf. CitationHeath et al., 1992). Usually, σ(t, T) is referred to as volatility. The instantaneous spot rate (hereinafter, the spot rate) r is given by r(t)=f(t, t). The savings account B* is defined by , and denotes the price of the zero-coupon bond at time t with maturity T. Then, it follows that (2) where and Note that is referred to as price volatility. The discounted bond price is given by

Under the assumption of no arbitrage, CitationHeath et al. (1992) shows that there exists ϕ(t) such that for arbitrary T. ϕ(t) is referred to as the market price of risk. Note that the use of plus/minus signs in this definition follows that in CitationJamshidian (1997) and CitationYasuoka (2013). Let be a measure equivalent to such that is a -Brownian motion. is known as a risk-neutral measure. Then, the discount price process is a -martingale for all T.

We then have (3) accordingly, (4) Although bond pricing is achieved under the interest-rate simulation for risk management is executed under .

Indeed, if we set , then it follows from equation (2) that We have (5) where the left-hand term is the excess return over r(t) of a T-maturity bond, and is a bond-price volatility. Thus, equation (5) coincides with a well-known definition of the market price of risk for the bonds market.

2.2. Market price of risk

The HJM model is Gaussian if σ(t, T) is a deterministic function in t and T. In what follows, an -valued volatility σ(t, T) is assumed to be always deterministic and continuous in t and T. Also, is deterministic and continuous.

Let be an arbitrary sequence of maturity dates with 0<Ti<Ti+1. Taking to be a d×d matrix, we assume that the rank of this matrix is equal to d. Assume that the volatility σ(t, T) is given. The market price of risk is assumed to be constant during the sample period. Setting , the forward rate process is expressed by (6)

For an integer nd, let be a sequence of lengths of time, where xi<xi+1. Let a time interval Δ t>0 be fixed, and be a sequence of observed times with t1=0 and , where J+1 is the number of observed times. Naturally, we observe the instantaneous forward rate F(tk, xi) in the interest-rate market with a fixed length of time xi from tk to the maturity date tk+xi. Taking F(tk, xi) and , respectively, as f(0, xi) and in equation (6) for each , we suppose that the following equation holds for all i. where xis is identified with a maturity date at t=s.

To simplify the notation, we denote σ(0, xi) and as σ0i and υ0i, respectively. Moreover, we denote the lth component of σ0i as . Then, where the superscript T denotes transpose. The Euler integral implies (7) for all i and k, where . Let denote the sum of the squared difference between each side of equation (7) in the time series and cross sections, neglecting the random part such that (8) where we set and We determine ϕ as the solution that minimizes . Then, ϕ is estimated by the following proposition, where EH[ ] denotes the sample mean over k=1, … , J.Footnote

Proposition 2.1

For a d-factor model with dn, let ϕ be a solution that minimizes . Then, ϕ is uniquely given as the solution of the linear equation (9) where (10)

Proof Taking as a d×n-matrix, we may assume that the rank of this matrix is d. Because it follows that Then, the partial derivative of in ϕl is (11) for l=1, … , d. We have and the Hessian of ε is given by (12) It holds that, for an arbitrary vector , (13) because is positive definite. Hence, the solution ϕ to the minimization problem is uniquely determined by solving . A constant vector is defined in equation (10). Upon substituting equation (10) into equation (11), the equation reduces to . This is equivalent to equation (9). From equation (13), is positive definite and thus rank d, so ϕ is obtained as a unique solution to equation (9). This completes the proof.

Note that for the setup, the Musiela parametrization (see CitationBrace and Musiela, 1994) allows for an alternative approach from equation (6) to equation (7).

2.3. Property of simulation

Let σ(t, T) be Rn-valued volatility. For a sequence of lengths of time we assume that is rank n. For , we obtain from the historical data. Let ϕ be the market price of risk obtained by Proposition 2.1, and let d=n.

We recall the forward rate process (6). Regarding xi as a fixed maturity date at t=0, we set Ti=xi for i=1, … , n. We consider an n-factor model as (14) for i=1, … , n. An n-factor model is called a full-factor model, and is said to be orthogonal if for all l, k with lk. The next proposition gives a single-period simulation under P by a full-factor model.

Proposition 2.2

If is rank n and orthogonal, then, for an arbitrary f(0, Ti) and Δ s>0 with Δ s<T1, is simulated by (15)

Proof The Euler integral of equation (14) implies (16) for i=1, … , n. The inner product between and vanishes from equation (9) for all l. Since is rank n and orthogonal, the family of n-dimensional vectors spans . Hence, for all i. Substituting this and equation (10) into the second term of equation (16), we have equation (15).

For example, if the principal components of historical covariance are imposed on the volatility, then such volatility is orthogonal. Furthermore, represents a kind of historical change in the sample forward rate, Therefore, equation (15) is interpreted as a term structure model with historical drift and volatility.

The following corollary shows the averages of interest rates after the single-period simulation. Comparing equation (17) with equation (18), we see the difference between a real-world simulation and a risk-neutral simulation.

Corollary 2.3

Under the assumption of Proposition 2.2, is normally distributed under both and with the following expectation. (17) (18) The standard deviations of in both simulations are equal to .

2.4. Volatility associated with principal component analysis

Returning to the setup in Section 2.2, we estimate the market price of risk more explicitly. We denote a sample covariance matrix by V such that (19) for i, j. We assume that V has rank dn. By the usual arguments pertaining to principal component analysis, the covariance matrix is decomposed into where and are the lth eigenvalue and the lth eigenvector, respectively. We may assume that all eigenvectors are chosen such that and ρl>0. el is also known as the lth principal component of the covariance, so it holds that For a positive integer kd, the accumulated contribution rate Ck is defined as .

Let σ(t, T) be a d-dimensional volatility satisfying then the market price of risk is obtained as follows.

Proposition 2.4

For a d-factor model with dn, if σ0i is given as the principal components, then it follows that where .

Proof The left-hand side of equation (9) becomes . Similarly, the right-hand side of equation (9) becomes We have . From , we have . Since ρl≠0, we have for all l. This completes the proof.

We call ζl and the lth market price of risk (MPR) score and the lth volatility, respectively. The norm of is Thus, ρl represents the magnitude of the lth volatility, we call ρl the lth volatility risk. Consequently, the market price of risk is explained as the MPR score per unit risk of the volatility, this is analogous to CitationYasuoka (2013) on the LMM.

Note that positive and negative values of ϕ coincide with positive and negative MPR scores. Hence, MPR score is affected not only by the definition of ϕ, but also by the definition of el. To investigate positive and negative values of ϕ, we must fix their definitions. This is the reason that we assume for all l.

3. Numerical analysis of real-world simulation

To reduce the dimensions of the simulation to practical bounds, there are at least two approaches. One approach is to approximate the dynamics of the forward rate while retaining the arbitrage-free framework. This was already studied in CitationYasuoka (2013). Another approach is to approximate the expression (15) in Proposition 2.2. The point of the latter approach is that a historical and simple structure is retained. For the sake of convenience, we adopt the latter approach. In the following, we denote the d-factor simulated forward rate as f˜ to distinguish it from the full-factor model.

Proposition 3.1

For a full-factor model, we assume that σ0i are given as the principal components, and that is rank n. If Cd is sufficiently close to 1 for some d<n, then the instantaneous forward rate at time Δ s is approximated by a d-factor model. (20)

Proof From the assumption, equation (15) holds. The second term in equation (15) does not depend on the number of factors. If Cd is sufficiently close to 1, it holds in the sense of probability, that Therefore, equation (20) is obtained.

Recall that Proposition 2.2 is introduced as a discrete-time version of an arbitrage-free model. In contrast, the above model is no more than an approximation of an arbitrage-free model. Indeed, in a d-factor model (d<n), does not vanish as it does in equation (20); hence, it is not trivial to obtain equation (20). However, the above approximation is mostly accurate if the number of factors is chosen such that Cd≈ 1.

Next we consider the numerical meaning of the market price of risk. The ith observable trend of the historical forward rate is defined by . Intuitively, this means the change of the forward rate curve during the period. The ith rolled trend of the historical forward rate is defined as . It follows that (21)

The first term is the ith observable trend, and the second term is almost equal to the sample mean of the forward rate slope.Footnote

For example, if the forward rate curve is upward sloping and remains unchanged, then equation (21) implies . This is known as the roll-down of the forward rate curve. Furthermore, when the upward sloping forward rate curve is rising fast, the rolled trend, that is, the roll-up of the forward rate curve, might be positive. Accordingly, the rolled trend corresponds to the roll-down or roll-up of the forward rate curve, which we are familiar with from financial experience. Proposition 2.2 shows that the real-world simulation reproduces the rolled trend of the entire forward rate curve. We next explain the meaning of the market price of risk in connection with the above trends.

Proposition 3.2

For a d-dimensional model with dn, we assume that the volatility σ0i is given by principal components. If it holds that for all i, then we have the following approximations. (22)

Proof From Proposition 2.4, the MPR score is given by The assumption implies equation (22). Since , we have the second approximation. This completes the proof.

Proposition 3.2 shows that ϕ is largely determined by the rolled trend in the historical forward rate. Therefore, the numerical meaning of the market price of risk is roughly interpreted with respect to the observable trend. From equation (22), approximately represents the coordinates of the rolled trend in the principal component space. In other words, the lth MPR score ζl is approximated by the lth projection of the rolled trend to the space. The principal components e1, e2, and e3 are known to explain the movement of the forward rate curve in terms of level, slope, and curvature, respectively. Similarly, , and ζ3 approximately represent the level, slope, and curvature factors of the rolled trend. Hence, ϕl roughly means the lth projection of the rolled trend per lth volatility risk.

Consider an investor who holds bonds for all maturities and expects the roll-down of the entire forward rate curve as an excess return. Then, the first market price of risk ϕ1 is a risk-adjusted measure for the roll-down return of the entire forward rate curve. The values are interpreted in a similar way. shows rough estimates of ϕ1 in each case. shows rough estimates of ϕ1 and ϕ2 in association with the two factors of the observable trend, where we assume that most of the forward rate curves in the sample data are upward sloping. These tables are exactly same as those under the LMM ( and in CitationYasuoka, 2013).

Table 1. First market price of risk ϕ1 with respect to the observable trend and average shape of the initial forward rate curve.

Table 2. Market prices of risks ϕ1 and ϕ2 with respect to the observable trend.

4. Numerical examples

We have data for Japanese LIBOR swap rates from 2 April 2007 to 31 August 2009, which is the same data set used in CitationYasuoka (2013). Setting δ=0.5 (year) and xii for i=1, 2, … , the 6-month forward LIBOR is obtained for this period. For convenience, we regard the forward LIBOR as the instantaneous forward rate in the following example. According to CitationYasuoka (2013), we divide the sample period into two periods. Period A is earlier and covers 2 April 2007 to 16 June 2008, period B is later and covers 16 June 2008 to 31 August 2009. presents the time series of the forward rates for xi=0, 2, 5, 10 years and illustrates periods A and B. presents the forward rate curves at three days (2 April 2007, 16 June 2008, and 31 August 2009). These curves show that the observable trend is bear-flattening in period A, and bull-steepening in period B.

Figure 1. Forward LIBOR in Japanese LIBOR/swap market from 2 April 2007 to 31 August 2009, where xi = 0, 2, 5, 10 years. Period A is from 2 April 2007 to 16 June 2008, period B is from 16 June 2008 to 31 August 2009. The swap data are provided by Mizuho Information and Research Institute.

Figure 1. Forward LIBOR in Japanese LIBOR/swap market from 2 April 2007 to 31 August 2009, where xi = 0, 2, 5, 10 years. Period A is from 2 April 2007 to 16 June 2008, period B is from 16 June 2008 to 31 August 2009. The swap data are provided by Mizuho Information and Research Institute.

Figure 2. Implied forward LIBOR curves at three days (2 April 2007, 16 June 2008, and 31 August 2009). The observable trend is bear flat in period A and bull steep in period B.

Figure 2. Implied forward LIBOR curves at three days (2 April 2007, 16 June 2008, and 31 August 2009). The observable trend is bear flat in period A and bull steep in period B.

In the data analysis, are directly implied from the LIBOR/swap rates, but is not since usually . So, we approximate by a linear interpolation such that shows the eigenvalues, contribution rates, and the market prices of risk for each period. Since the observable trend is almost bear-flattening in period A, the ‘bear flat’ column of shows that ϕ1 is near zero and ϕ2 is positive. shows that and in this period, which are roughly explained in . The observable trend of period B is bull-steepening. shows that ϕ1 and ϕ2 are both negative in the ‘bull steep’ column. The results are that and . These values are also consistent with .

Table 3. Eigenvalues and the market prices of risk for periods A and B.

For comparison, the rightmost column presents the market price of risk that is implied by the LMM framework on the same sample (table 7.1 in [7]). As mentioned at the end of Section 3, the numerical interpretation of the market price of risk in this study is largely similar to that in the LMM. This commonality is reflected in that the first and the second marker prices of risk show similar tendency between two models.

5. Conclusion

The interest-rate simulation model has been introduced under the real-world measure in the Gaussian HJM framework. The procedure to estimate the market price of risk was presented under the assumption that market price of risk is constant. The market price of risk is interpreted by the change in the historical forward rate curve rather than the change in the state of the curve.

In particular, when the principal components are imposed on the volatilities, the simulation model is almost equivalent to a term structure model with historical drift and volatility. Accordingly, the mean of the forward rate obtained from the simulation is strongly affected by sample period selection.

Although the possibility of a negative interest rate is a drawback of the Gaussian model, this paper's approach is practical for interest-rate risk management in actual businesses. As an example of present use, the Hull–White model, which is a special case of the HJM model, is successfully used for risk management.

Additional information

Funding

This work was partially supported by JSPS KAKENHI [Grant Number 26380399].

Notes

In this paper, an interest-rate simulation under the real-world measure is referred to as a real-world simulation for short. Similarly, the simulation under a risk-neutral measure is referred to as a risk-neutral simulation.

For arbitrary historical data {Ak}k=1, … , J+1, the sample mean is usually evaluated by using . In this sense, EH[ ] is not the sample mean of Ak because .

For a more precise explanation, see note 2.

References

  • A. Brace and M. Musiela, A multifactor Gauss Markov implementation of Heath, Jarrow, and Morton. Math. Finance, 1994, 4, 259–283. doi: 10.1111/j.1467-9965.1994.tb00095.x
  • M.A.H. Dempster, E.A. Medova and M. Villaverde, Long-term interest rates and consol bond valuation. J. Asset Manag., 2010, 11, 113–135. doi: 10.1057/jam.2010.7
  • D. Heath, R. Jarrow and A. Morton, Bond pricing and the term structure of interest rates: A new methodology. Econometrica, 1992, 61, 77–105. doi: 10.2307/2951677
  • J. Hull and A. White, Pricing interest-rate derivative securities. Rev. Financ. Stud., 1990, 3, 573–592. doi: 10.1093/rfs/3.4.573
  • F. Jamshidian, A LIBOR and swap market models and measures. Finance Stoch., 1997, 1, 293–330. doi: 10.1007/s007800050026
  • R. Stanton, A nonparametric model of term structure dynamics and the market price of interest rate. J. Finance, 1997, 52, 1973–2002. doi: 10.1111/j.1540-6261.1997.tb02748.x
  • T. Yasuoka, LIBOR market model under the real-world measure. Int. J. Theor. Appl. Finance, 2013, 16(4):1350024 (18 pages). doi: 10.1142/S0219024913500246