Skip to main content
Log in

An affine model for short rates when monetary policy is path dependent

  • Published:
Review of Derivatives Research Aims and scope Submit manuscript

Abstract

I propose an affine model of short rates that incorporates a random walk with stochastic drift. This framework enables my model to capture the behavior of monetary authorities in the short rate market, allowing for minor deviations while reacting strongly to deviations large enough to threaten production. Importantly, my model facilitates the derivation of closed-form bond prices, thereby providing an analytical solution for bond-option prices. I compare my model with nine standard short rate models found in the literature. Among these, five are single-factor models and four are multifactor models. Remarkably, my model outperforms all competing short rate models, including the constant elasticity of volatility, stochastic mean, and stochastic volatility models. Moreover, it yields interest rate forecasts consistent with common term structure priors and surpasses the performance of the naive random walk model. Additionally, my stochastic mean model can explain the unspanned risks documented in the literature.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6

Similar content being viewed by others

Notes

  1. Liu (2018) provides Out-of-sample evidence of predictability of excess returns for economic growth and inflation. However, Bauer and Hamilton (2017) conclude that this predictability is possibly spurious due to size distortions resulting from persistent regressors and independent variables that are not strictly exogenous. Ang et al. (2008) develop a term structure model with a time‐varying risk premium. They find that the inflation risk premium explains the upward sloping nominal term structure. Bekaert et al. (2010) also document persistent interest rates where the inflation target is time varying. Duffee (2002) finds that forecasts from a random walk model outperform the Dai and Singleton (2000) affine model.

  2. Balduzzi et al. (1998) assume a stochastic mean. However, their stochastic mean is mean-reverting. My contribution is that I introduced a model with a random walk stochastic mean.

  3. Bakshi, Gao, and Xue (2023) use of options on the 10- and 30-year Treasury bond to estimate the expected return of bond futures. These measures exhibit forecasting ability for future returns, surpassing the predictive power of the level, slope, and curvature variables typically derived from the yield curve.

  4. Ravn and Uhlig (2002) point out, the standard constant for monthly data is 1/129,600 or 1/43,200 when the fourth and third power of the number of months in a quarter are used, respectively.

  5. I compute the asymptotic p-values for ADF and WS via the MacKinnon (1994) approximation. These p-values are robust to size distortion. Results are provided for the null of driftless unit root; qualitatively similar results are obtained under the null of unit root with drift.

  6. Several papers find the nominal short rate follows a unit root process (See Perron (1989), Aït-Sahalia(1996a), and Bandi (2002)). However, Bierens (1997) and Al-Zoubi (2009) propose this conclusion is possibly flawed for two reasons: First, negative values should be realizable in a random walk with no drift; Second, given a positive drift a random walk would converge to infinity. Observed short rates do not exhibit these characteristics.

  7. See Perron (1989).

  8. I implement the one-sided Hodrick–Prescott filter introduced by Mehra (2004) to calculate the nonstationary mean. I employ a smoothing constant of 1/q = 43,200 for both the HP and bHP filters, this smoothing constant corresponds to the Ravn and Uhlig (2002) adjustments of the third power for the frequency of observations.

References

  • Ahn, D.-H., & Gao, B. (1999). A parametric nonlinear model of term structure dynamics. Review of Financial Studies, 12(4), 721–762.

    Article  Google Scholar 

  • Aït-Sahalia, Y. (1996a). Nonparametric pricing of interest rate derivatives securities. Econometrica, 64(3), 527–560.

    Article  Google Scholar 

  • Aït-Sahalia, Y. (1996b). Testing continuous-time models of the spot interest rate. Review of Financial Studies, 9(2), 385–426.

    Article  Google Scholar 

  • Aït-Sahalia, Y. and Kimmel, R., (2007). Maximum likelihood estimation of stochastic volatility models. Journal of financial economics, 83(2), 413–452.

    Article  Google Scholar 

  • Aksoy, Y., Orphanides, A., Small, D., Wiel, V., & Wilcox, D. (2006). A quantitative exploration of the opportunistic approach to disinflation. Journal of Monetary Economics, 53(8), 1877–1893.

    Article  Google Scholar 

  • Al-Zoubi, H. A. (2009). Short-term spot rate models with nonparametric deterministic drift. Quarterly Review of Economics and Finance, 49(3), 731–747.

    Article  Google Scholar 

  • Al-Zoubi, H.A., 2019. Bond and option prices with permanent shocks. Journal of Empirical Finance, 53, 272–290.

    Article  Google Scholar 

  • Andersen, T.G. and Lund, J., (1997). Estimating continuous-time stochastic volatility models of the short-term interest rate. Journal of econometrics, 77(2), 343–377.

    Article  Google Scholar 

  • Ang, A., Bekaert, G., & Wei, M. (2008). The term structure of real rates and expected inflation. The Journal of Finance, 63(2), 797–849.

    Article  Google Scholar 

  • Bakshi, G., Crosby, J., Gao, X., & Hansen, J. W. (2023a). Treasury option returns and models with unspanned risks. Journal of Financial Economics, 150(3), 103736.

    Article  Google Scholar 

  • Bakshi, G., Gao, X., & Xue, J. (2023). Recovery with applications to forecasting equity disaster probability and testing the spanning hypothesis in the treasury market. Journal of Financial and Quantitative Analysis, 58(4), 1808–1842.

    Article  Google Scholar 

  • Balduzzi, P., Das, S., & Foresi, S. (1998). The central tendency: A second factor in bond yields. The Review of Economics and Statistics, 80(1), 62–72.

    Article  Google Scholar 

  • Bandi, F. M. (2002). Short-term interest rate dynamics: A special approach. Journal of Financial Economics, 65(1), 73–110.

    Article  Google Scholar 

  • Bauer, M. D., & Hamilton, J. D. (2017). Robust bond risk premia. The Review of Financial Studies, 31(2), 399–448.

    Article  Google Scholar 

  • Bauer, M. D., & Rudebusch, G. D. (2020). Interest rates under falling stars. American Economic Review, 110(5), 1316–1354.

    Article  Google Scholar 

  • Bekaert, G., Cho, S., & Moreno, A. (2010). New Keynesian macroeconomics and the term structure. Journal of Money, Credit and Banking, 42(1), 33–62.

    Article  Google Scholar 

  • Bergstrom, A. R. (1986). The estimation of open higher-order continuous time dynamic models with mixed stock and flow data. Econometric Theory, 2(3), 350–373.

    Article  Google Scholar 

  • Bergstrom, A. R. (1989). Optimal forecasting of discrete stock and flow data generated by a higher order continuous time system. Computers & Mathematics with Applications, 17(8/9), 1203–1214.

    Article  Google Scholar 

  • Beveridge, S. and Nelson, C.R., (1981). A new approach to decomposition of economic time series into permanent and transitory components with particular attention to measurement of the ‘business cycle’. Journal of Monetary economics, 7(2), 151–174.

    Article  Google Scholar 

  • Bierens, H. J. (1997). Testing the unit root with drift hypothesis against nonlinear trend stationary, with an application to the us price level and interest rate. Journal of Econometrics, 81(1), 29–64.

    Article  Google Scholar 

  • Blaskowitz, O., & Herwartz, H. (2011). On economic evaluation of directional forecasts. International Journal of Forecasting, 27(4), 1058–1065.

    Article  Google Scholar 

  • Bleich, D., Fendel, R., & Rülke, J.-C. (2012). Inflation targeting makes the difference: Novel evidence on inflation stabilization. Journal of International Money and Finance, 31(5), 1092–1105.

    Article  Google Scholar 

  • Bergmeir, C., Costantini, M., & Benítez, J. M. (2014). On the usefulness of cross-validation for directional forecast evaluation. Computational Statistics & Data Analysis, 76, 132–143.

    Article  Google Scholar 

  • Cai, L., & Swanson, N. R. (2011). In-and out-of-sample specification analysis of spot rate models: Further evidence for the period 1982–2008. Journal of Empirical Finance, 18(4), 743–764.

    Article  Google Scholar 

  • Campbell, J.Y. and Shiller, R.J., (1987). Cointegration and tests of present value models. Journal of political economy, 95(5), 1062–1088.

    Article  Google Scholar 

  • Chan, K. C., Karolyi, G. A., Longstaff, F. A., & Sanders, A. B. (1992). An empirical comparison of alternative models of the short-term interest rate. Journal of Finance, 47(3), 1209–1227.

    Google Scholar 

  • Chapman, D. A., & Pearson, N. D. (2000). Is the short rate drift actually nonlinear? Journal of Finance, 55(1), 355–388.

    Article  Google Scholar 

  • Chen, L., (1996). Stochastic mean and stochastic volatility: a three-factor model of the term structure of interest rates and its applications in derivatives pricing and risk management. Blackwell publishers

  • Cieslak, A., & Povala, P. (2015). Expected returns in treasury bonds. The Review of Financial Studies, 28(10), 2859–2901.

    Article  Google Scholar 

  • Cochrane, J. H., & Piazzesi, M. (2005). Bond risk premia. American Economic Review, 95(1), 138–160.

    Article  Google Scholar 

  • Cogley, T. and Nason, J.M., (1995). Output dynamics in real-business-cycle models. The American Economic Review, 492–511.

  • Cooper, I., & Priestley, R. (2009). Time-varying risk premiums and the output gap. The Review of Financial Studies, 22(7), 2801–2833.

    Article  Google Scholar 

  • Cox, J. C., Ingersoll, J. E., & Ross, S. A. (1985). A theory of the term structure of interest rates. Econometrica, 53(2), 385–407.

    Article  Google Scholar 

  • Dai, Q., & Singleton, K. J. (2000). Specification analysis of affine term structure models. The Journal of Finance, 55(5), 1943–1978.

    Article  Google Scholar 

  • de Jong, R. M., & Sakarya, N. (2016). The econometrics of the Hodrick–Prescott filter. Review of Economics and Statistics, 98(2), 310–317.

    Article  Google Scholar 

  • Dickey, D. A., & Fuller, W. A. (1979). Distribution of the estimators for autoregressive time series with a unit root. Journal of American Statistical Association, 74(366), 427–431.

    Article  Google Scholar 

  • Duffee, G. R. (2002). Term premia and interest rate forecasts in affine models. The Journal of Finance, 57(1), 405–443.

    Article  Google Scholar 

  • Duffee, G. R. (2011). Forecasting with the term structure: The role of no-arbitrage. Johns Hopkins University.

    Google Scholar 

  • Duffee, G. (2013). Forecasting interest rates. In Handbook of economic forecasting (Vol. 2, pp. 385–426). Elsevier.‏

  • Duffee, G. R. (2018). Expected inflation and other determinants of treasury yields. The Journal of Finance, 73(5), 2139–2180.

    Article  Google Scholar 

  • Duffie, D. and Kan, R., (1996). A yield‐factor model of interest rates. Mathematical finance, 6(4), 379–406.

    Article  Google Scholar 

  • Durham, G. B. (2003). Likelihood-based specification analysis of continuous-time models of the short-term interest rate. Journal of Financial Economics, 70(3), 463–487.

    Article  Google Scholar 

  • Fama, E. F., & Bliss, R. R. (1987). The information in long-maturity forward rates. The American Economic Review, 680–692.

  • Fama, E. (2006). The behavior of interest rates. Review of Financial Studies, 19(2), 359–379.

    Article  Google Scholar 

  • Gallant, A. R. (1987). Nonlinear statistical models. New York: Wiley.

    Book  Google Scholar 

  • Greenwood, R., & Vayanos, D. (2014). Bond supply and excess bond returns. The Review of Financial Studies, 27(3), 663–713.

    Article  Google Scholar 

  • Hamilton, J. D. (2018). Why you should never use the Hodrick–Prescott filter. Review of Economics and Statistics, 100(5), 831–843.

    Article  Google Scholar 

  • Hansen, L. P. (1982). Large sample properties of generalized method of moments estimators. Econometrica, 50(4), 1029–1054.

    Article  Google Scholar 

  • Harvey, D., Leybourne, S., & Newbold, P. (1997). Testing the equality of prediction mean squared errors. International Journal of Forecasting, 13(2), 281–291.

    Article  Google Scholar 

  • Hodrick, R. J., & Prescott, E. C. (1997). Postwar US business cycles: An empirical investigation. Journal of Money, Credit, and Banking, 29(1), 1–16.

    Article  Google Scholar 

  • Inoue, A., & Shintani, M. (2006). Bootstrapping GMM estimators for time series. Econometrics, 133(2), 531–555.

    Article  Google Scholar 

  • Jamshidian, F. (1989). An exact bond option formula. The Journal of Finance, 44(1), 205–209.

    Article  Google Scholar 

  • Joslin, S., Priebsch, M., & Singleton, K. J. (2014). Risk premiums in dynamic term structure models with unspanned macro risks. The Journal of Finance, 69(3), 1197–1233.

    Article  Google Scholar 

  • Kohn, R., & Ansley, C. F. (1987). Signal extraction for finite nonstationary time series. Biometrika, 74(2), 411–421.

    Article  Google Scholar 

  • Liu, R. (2018). Forecasting bond risk premia with unspanned macroeconomic information. Quarterly Journal of Finance, 9(1), 1940001.

    Article  Google Scholar 

  • MacKinnon, J. G. (1994). Approximate asymptotic distribution functions for unit-root and cointegration tests. Journal of Business and Economic Statistics, 12(2), 167–176.

    Article  Google Scholar 

  • Mehra, Y. P. (2004). The output gap expected future inflation and inflation dynamics: Another look. B.E. Journal of Macroeconomics. https://doi.org/10.2202/1534-5998.1194

    Article  Google Scholar 

  • Orphanides, A., & Wieland, V. (2000). Inflation zone targeting. European Economic Review, 44(7), 1351–1387.

    Article  Google Scholar 

  • Orphanides, A., & Wilcox, D. W. (2002). An opportunistic approach to disinflation. International Finance, 5(1), 47–71.

    Article  Google Scholar 

  • Pantula, S. G., Gonzalez-Farias, G., & Fuller, W. A. (1994). A comparison of unit-root test criteria. Journal of Business and Economic Statistics, 12(4), 449–459.

    Article  Google Scholar 

  • Perron, P. (1989). The great crash, the oil-price shock, and the unit-root hypothesis. Econometrica, 57(6), 1361–1402.

    Article  Google Scholar 

  • Phillips, P. C. B. (1987). Time series regression with a unit root. Econometrica, 55(2), 277–301.

    Article  Google Scholar 

  • Phillips, P. C. B., & Perron, P. (1988). Testing for a unit root in time series regression. Biometrika, 75(2), 335–346.

    Article  Google Scholar 

  • Phillips, P. C. B., & Sainan, J. (2021). Business cycles, trend elimination, and the HP filter. International Economic Review, 62(2), 469–520.

    Article  Google Scholar 

  • Phillips, P. C., & Shi, Z. (2021). Boosting: Why you can use the HP filter. International Economic Review, 62(2), 521–570.

    Article  Google Scholar 

  • Phillips, P. C. B., & Yu, J. (2005). Jackknifing bond option prices. Review of Financial Studies, 18(2), 707–742.

    Article  Google Scholar 

  • Phillips, P. C. B., & Yu, J. (2009). Maximum likelihood and gaussian estimation of continuous time models in finance. In T. Mikosch & J. P. Krei (Eds.), Handbook of financial time series (pp. 497–530). Berlin: Springer.

    Chapter  Google Scholar 

  • Piazzesi, M., & John, C. (2009) Decomposing the yield curve. In 2009 Meeting Papers, no. 18. Society for Economic Dynamics.

  • Ravn, M. O., & Uhlig, H. (2002). On adjusting the Hodrick–Prescott filter for the frequency of observations. Review of Economics and Statistics, 84(2), 371–376.

    Article  Google Scholar 

  • Stock, J. H., & Watson, M. W. (1988). Variable trends in economic time series. Journal of Economic Perspectives, 2(3), 147–174.

    Article  Google Scholar 

  • Tang, C. Y., & Chen, S. X. (2009). Parameter estimation and bias correction for diffusion processes. Journal of Econometrics, 149(1), 65–81.

    Article  Google Scholar 

  • Taylor, J.B., (1993), December. Discretion versus policy rules in practice. In Carnegie-Rochester conference series on public policy (Vol. 39, pp. 195–214). North-Holland.

  • Vasicek, O. (1977). An equilibrium characterization of the term structure. Journal of Financial Economics, 5(2), 177–188.

    Article  Google Scholar 

  • Yu, J. (2012). Bias in the estimation of the mean reversion parameter in continuous time models. Journal of Econometrics, 169(1), 114–122.

    Article  Google Scholar 

Download references

Acknowledgement

I would like to thank anonymous referees, Pietro Veronesi, Robert de Jong and Jun Yu for their helpful comments and suggestions. The project is partially supported by Alfaisal University research grant N. 18102.

Funding

No funding was received to assist with the preparation of this manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Haitham A. Al-Zoubi.

Ethics declarations

Conflict of interest

The authors have no relevant financial or non-financial interests to disclose.

Additional information

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Appendices

Appendix 1: Bond and option prices

This appendix provides details of the derivations related to bond and option prices under random walk mean for the short rate process.

Proof of Proposition 2.1

Consider the NSM model of the short rate introduced in Sect. 2. Under the risk-neutral measure, the short rate dynamics are given by

$$dr_{t} = - \alpha_{1} \left( {\mu_{t} - r_{t} } \right)dt + \sigma dZ_{t} ,$$
(11)
$$d\mu_{t} = \sigma_{\mu } dW_{t} .$$

It can be verified using Ito’s formula that

$$r_{t} = e^{{\alpha_{1 } }} \left[ {r_{0} - \mathop \smallint \limits_{0}^{t} \alpha \mu_{t} e^{{\alpha_{1u } }} du + \sigma \mathop \smallint \limits_{0}^{t} e^{{\alpha_{1u } }} dZ_{u} } \right]$$

is a solution to the stochastic differential equation (SDE) in (11). It can be shown using integration by parts for the second term that

$$\begin{aligned} r_{t} & = e^{{\alpha_{1} t}} \left[ {r_{0} - \mu_{t} \mathop \smallint \limits_{0}^{t} e^{{ - \alpha_{1} u}} du + \alpha_{1} \mathop \smallint \limits_{0}^{t} d\mu_{t} \left( {\mathop \smallint \limits_{0}^{t} e^{{ - \alpha_{1} t}} dt} \right)du + \sigma \mathop \smallint \limits_{0}^{t} e^{{ - \alpha_{1} u}} dZ_{u} } \right] \\ & = e^{{\alpha_{1} t}} \left[ {r_{0} + \mu_{t} \left( {e^{{ - \alpha_{1} t}} - 1} \right) - \mathop \smallint \limits_{0}^{t} \sigma_{m} dW\left( {e^{{ - \alpha_{1} t}} - 1} \right)du + \sigma \mathop \smallint \limits_{0}^{t} e^{{ - \alpha_{1} u}} dZ_{u} } \right] \\ & = e^{{\alpha_{1} t}} \left[ {r_{0} + \mu_{t} \left( {e^{{ - \alpha_{1} t}} - 1} \right) + \sigma \mathop \smallint \limits_{0}^{t} e^{{ - \alpha_{1} u}} dZ_{u} } \right]. \\ \end{aligned}$$

The expectation follows immediately from the equation given above,

$$E\left[ {r_{t} } \right] = e^{{\alpha_{1} t}} \left[ {r_{0} + E\left[ {\mu_{t} } \right]\left( {e^{{ - \alpha_{1} t}} - 1} \right) + E\left[ {\sigma \mathop \smallint \limits_{0}^{t} e^{{ - \alpha_{1} u}} dZ_{u} } \right]} \right],$$
$$E\left[ {r_{t} } \right] = e^{{\alpha_{1} t}} \left[ {r_{0} + \mu_{t} \left( {e^{{ - \alpha_{1} t}} - 1} \right)} \right].$$

Clearly, \(\underset{t\to \infty }{lim}E\left[{r}_{t}\right]=E\left[{\mu }_{t}\right]\).

Proof of Proposition 2.2

Write

$${c}_{u}={r}_{u}-{\mu }_{u}$$
(12)

where \({c}_{u}\) is a solution of the Ornstein–Uhlenbeck equation:

$$dc\left(t\right)={\alpha }_{1}c\left(t\right)+{\sigma }_{c}d{Z}_{t}$$

Applying Ito’s lemma, the \({c}_{u}\) process is given by

$${c}_{u}={e}^{{\alpha }_{1}u}\left({c}_{0}+{\int }_{0}^{u}{\sigma }_{c}{e}^{-{\alpha }_{1}s}d{Z}_{s}\right).$$
(13)

Using Eq. (11) I obtain:

$$\begin{aligned} Cov\left[ {c_{t} ,c_{u} } \right] & = \sigma_{c}^{2} e^{{\alpha_{1} \left( {u + t} \right)}} E\left[ {\mathop \int \limits_{0}^{t} e^{{ - \alpha_{1} s}} dZ_{s} \left[ {\mathop \int \limits_{0}^{u} e^{{ - \alpha_{1} s}} dZ_{s} } \right]} \right] \\ & = \sigma_{c}^{2} e^{{\alpha_{1} \left( {u + t} \right)}} \mathop \int \limits_{0}^{u\Lambda t} e^{{ - 2\alpha_{1} s}} ds = \frac{{\sigma_{c}^{2} }}{{ - 2\alpha_{1} }}e^{{\alpha_{1} \left( {u + t} \right)}} \left( {e^{{2\alpha_{1} \left( {u\Lambda t} \right)}} - 1} \right). \\ \end{aligned}$$

Similarly,

$$\begin{aligned} Cov\left[ {\mu_{t} ,\mu_{u} } \right] & = \sigma_{m}^{2} E\left[ {\mathop \int \limits_{0}^{t} dW_{s} \mathop \int \limits_{0}^{u} dW_{s} } \right] \\ & = \sigma_{m}^{2} \left[ {\mathop \int \limits_{0}^{t\Lambda u} ds} \right] = \sigma_{m}^{2} \left( {t\Lambda u} \right). \\ \end{aligned}$$

Consequently,

$$\begin{aligned} Var\left[ {\mathop \int \limits_{0}^{t} c_{u} du} \right] & = Cov\left( {\mathop \int \limits_{0}^{t} c_{u} du,_{{}} \mathop \int \limits_{0}^{t} c_{u} ds} \right) \\ & = \mathop \int \limits_{0}^{t} \mathop \int \limits_{0}^{t} Cov\left( {c_{u} ,c_{s} } \right)duds = \mathop \int \limits_{0}^{t} \mathop \int \limits_{0}^{t} \frac{{\sigma_{c}^{2} }}{ - 2\beta }e^{{\alpha_{1} \left( {u + s} \right)}} \left( {e^{{ - 2a\left( {u\Lambda s} \right)}} - 1} \right)^{{}} duds = \frac{{\sigma_{c}^{2} }}{{2\alpha_{1}^{3} }}\left( {2\alpha_{1} t + 3 - 4e^{{\alpha_{1} t}} + e^{{2\alpha_{1} t}} } \right). \\ \end{aligned}$$

Similarly,

$$Var\left[ {\mathop \int \limits_{0}^{t} \mu \left( u \right)du} \right] = \sigma_{m}^{2} \mathop \int \limits_{0}^{t} \mathop \int \limits_{0}^{t} \frac{{\left( {s\Lambda u} \right)^{2} }}{2}^{{}} duds = \sigma_{m}^{2} \frac{{t^{3} }}{6}.$$

From Eq. (12), I have

$$E\left[ { - \mathop \int \limits_{0}^{t} r_{u} du} \right] = E\left[ { - \mathop \int \limits_{0}^{t} \left( {c_{u} + \mu_{u} } \right)du} \right]$$

Therefore,

$$E\left[ { - \mathop \int \limits_{t}^{T} r_{u} du} \right] = \frac{{r_{t} - \mu_{t} }}{\beta }\left( {1 - e^{{\alpha_{1} \left( {T - t} \right)}} } \right) - \mu_{t} \left( {T - t} \right)$$
(14)

Furthermore,

$$\begin{aligned} Var\left[ { - \mathop \int \limits_{t}^{T} r_{u} du} \right] & = Var\left[ {\mathop \int \limits_{t}^{T} \left( {c_{u} + \mu_{u} } \right)du} \right] \\ & = \frac{{\sigma_{c}^{2} }}{{2\alpha_{1}^{3} }}\left( {2\alpha_{1} \left( {T - t} \right) + 3 - 4e^{{\alpha_{1} \left( {T - t} \right)}} + e^{{2\alpha_{1} \left( {T - t} \right)}} } \right) + \sigma_{m}^{2} \frac{{\left( {T - t} \right)^{3} }}{6} \\ \end{aligned}$$
(15)

Proof of Theorem 2.1

Consider the expected value and variance of short rate specified in Propositions 2.1 and 2.2. I specify the price of a zero-coupon bond with maturity T at time t, P(t, T)using the risk neutral valuation framework:

$$P\left( {t, T} \right) = E\left[ {exp\left( { - \mathop \int \limits_{t}^{T} r_{u} du} \right)|F_{t} } \right] = E\left[ {exp\left( { - \mathop \int \limits_{t}^{T} r_{u} du} \right)|r_{t} } \right]$$

where \(\left\{ {F_{t} } \right\}\) is standard filtration.

Combining Eqs. (11) to (15), the bond price is given by

$$\begin{aligned} P\left( {t,T,r_{t} } \right) & = \exp \left( {\left[ { - \mathop \int \limits_{t}^{T} r_{u} \left( {r_{t} } \right)du} \right] + \frac{1}{2}Var\left[ { - \mathop \int \limits_{t}^{T} r_{u} \left( {r_{t} } \right)du} \right]} \right) \\ & = Exp\left( {\frac{{r_{t} - \mu_{t} }}{\beta }\left( {1 - e^{{\alpha_{1} \left( {T - t} \right)}} } \right) - \mu_{t} \left( {T - t} \right) + \frac{{\sigma_{c}^{2} }}{{4\alpha_{1}^{3} }}\left( {2\alpha_{1} \left( {T - t} \right) + 3 - 4e^{{\alpha_{1} \left( {T - t} \right)}} + e^{{2\alpha_{1} \left( {T - t} \right)}} + \frac{{\sigma_{\mu }^{2} }}{{\sigma_{c}^{2} }}\frac{{\alpha_{1}^{3} \left( {T - t} \right)^{3} }}{3}} \right)} \right) \\ & = Exp\left( {\left( {\frac{{1 - e^{{\alpha_{1} \left( {T - t} \right)}} }}{{\alpha_{1} }}} \right)r_{t} - \left( {\frac{{1 - e^{{\alpha_{1} \left( {T - t} \right)}} }}{{\alpha_{1} }} + T - t} \right)\mu_{t} + \frac{{\sigma_{c}^{2} }}{{4\alpha_{1}^{3} }}\left( {2\alpha_{1} \left( {T - t} \right) + 3 - 4e^{{\alpha_{{1\left( {T - t} \right)}} }} + e^{{2\alpha_{{1\left( {T - t} \right)}} }} } \right) + \frac{{\sigma_{\mu }^{2} }}{{\sigma_{c}^{2} }}\frac{{\alpha_{1}^{3} \left( {T - t} \right)^{3} }}{3}} \right) \\ & = Exp\left( {A\left( {t,T} \right)r_{t} - \mu_{t} \left( {A\left( {t,T} \right) + \left( {T - t} \right)} \right) + B\left( {t,T} \right) + D\left( {t,T} \right)} \right). \\ \end{aligned}$$
(16)

where

$$\begin{aligned} & A\left( {t,T} \right) = \left( {\frac{{1 - e^{{\alpha_{1} \left( {T - t} \right)}} }}{{ - \alpha_{1} }}} \right) \\ & B\left( {t,T} \right) = \left[ {\frac{{A\left( {t,T} \right) + \left( {T - t} \right)}}{{2\alpha_{1}^{2} }} + \frac{{A\left( {t,T} \right)^{2} }}{{4\alpha_{1} }}} \right]\sigma_{c}^{2} \\ \end{aligned}$$

and

$$D\left(t,T\right)=\frac{{\sigma }_{\mu }^{2}{\left(T-t\right)}^{3}}{12}.$$

If HP and bHP trends are used, the bond price will be estimated using the signal to noise ratio (q). The bond price is given by

$$\begin{gathered} = Exp\left( {\frac{{r_{t} - \mu _{t} }}{{\alpha _{1} }}\left( {1 - e^{{\alpha _{1} \left( {T - t} \right)}} } \right) - \mu _{t} \left( {T - t} \right) + \frac{{\sigma ^{2} }}{{4\alpha _{1}^{3} \left( {1 + q} \right)}}\left( {2\alpha _{1} \left( {T - t} \right) + 3 - 4e^{{\alpha _{1} \left( {T - t} \right)}} + e^{{2\alpha _{1} \left( {T - t} \right)}} + q\frac{{\alpha _{1}^{3} \left( {T - t} \right)^{3} }}{3}} \right)} \right) \hfill \\ = Exp\left( {\left( {\frac{{1 - e^{{\alpha _{1} \left( {T - t} \right)}} }}{{\alpha _{1} }}} \right)r_{t} - \left( {\frac{{1 - e^{{\alpha _{1} \left( {T - t} \right)}} }}{{\alpha _{1} }} + \left( {T - t} \right)} \right)\mu _{t} + \frac{{\sigma ^{2} }}{{2\alpha _{1}^{2} \left( {1 + q} \right)}}\left( {\frac{{1 - e^{{\alpha _{1} \left( {T - t} \right)}} }}{{\alpha _{1} }}} \right) + \frac{{\sigma ^{2} }}{{2\alpha _{1}^{2} \left( {1 + q} \right)}}\left( {T - t} \right)} \right. \hfill \\ \left. {\quad + \frac{{\sigma ^{2} }}{{2\alpha _{1} \left( {1 + q} \right)}}\left( {\frac{{1 - 2e^{{\alpha _{1} \left( {T - t} \right)}} + e^{{2\beta \left( {T - t} \right)}} }}{{2\alpha _{1}^{2} }}} \right) + \left( {\frac{{q\sigma ^{2} \left( {T - t} \right)^{3} }}{{12\left( {1 + q} \right)}}} \right)} \right) \hfill \\ = Exp\left( {A\left( {t,T} \right)r_{t} - \mu _{t} \left( {A\left( {t,T} \right) + \left( {T - t} \right)} \right) + \frac{{\sigma ^{2} }}{{2\alpha _{1}^{2} \left( {1 + q} \right)}}A\left( {t,T} \right) + \frac{{\sigma ^{2} }}{{2\alpha _{1}^{2} \left( {1 + q} \right)}}\left( {T - t} \right)} \right. \hfill \\ \left. {\quad + \frac{{\sigma ^{2} }}{{2\alpha _{1} \left( {1 + q} \right)}}A\left( {t,T} \right)^{2} + \left( {\frac{{q\sigma ^{2} \left( {T - t} \right)^{3} }}{{12\left( {1 + q} \right)}}} \right)} \right) \hfill \\ = Exp\left( {A\left( {t,T} \right)r_{t} - \mu _{t} \left( {A\left( {t,T} \right) + \left( {T - t} \right)} \right) + B\left( {t,T} \right) + D\left( {t,T} \right)} \right) \hfill \\ \end{gathered}$$

where

$$A\left( {t,T} \right)\left( {\frac{{1 - e^{{\alpha_{1} \left( {T - t} \right)}} }}{{\alpha_{1} }}} \right),$$
$$B\left( {t,T} \right) \equiv \frac{{\sigma^{2} }}{{2\alpha_{1}^{2} \left( {1 + q} \right)}}\left( {A\left( {t,T} \right) + \left( {T - t} \right)} \right) + \frac{{\sigma^{2} A\left( {t,T} \right)^{2} }}{{2\beta \left( {1 + q} \right)}},$$

and

$$D\left( {t, T} \right) = \left( {\frac{{q\sigma^{2} \left( {T - t} \right)^{3} }}{{12\left( {1 + q} \right)}}} \right).$$

Appendix 2: Signal–noise ratio

This appendix provides details of the derivations related to the signal–noise ratio,

$$q = \sigma_{\eta }^{2} /\sigma_{v}^{2} = \sigma_{\eta }^{2} /\left( {1 - \rho^{2} - \mathop \sum \limits_{j = 1}^{p} \rho_{j}^{2} } \right)\sigma_{c}^{2} .$$

The resulting transitory component from (3), ct, possesses weak dependence properties with mean zero. Thus, ct is an AR(p) process:

$$c_{t + 1} = \rho c_{t} + \mathop \sum \limits_{j = 1}^{p} \rho_{j} c_{t - j} + v_{t + 1} , v_{t} \sim NID\left( {0,\sigma_{v}^{2} } \right).$$

The variance is

$$= E\left( {c_{t + 1}^{2} } \right) - \left( {Ec_{t + 1} } \right)^{2} .$$

It follows immediately that

$$\sigma_{c}^{2} = \frac{{\sigma_{v}^{2} }}{{1 - \rho^{2} - \mathop \sum \nolimits_{j = 1}^{p} \rho_{j}^{2} }}.$$
(17)

Now, consider the case in which \(\mu_{t}\) is a random walk process:

$$\Delta \mu_{t + 1} = \eta_{t + 1} \sim NID\left( {0,\sigma_{\eta }^{2} } \right).$$

I specify \(\mu_{t}\) as a driftless random walk process:

$$\mu_{t + 1} = \mu_{t} + \eta_{t} ,$$

which can be written as

$$\mu_{t + 1} = \mu_{0} + \mathop \sum \limits_{i = 1}^{t} \eta_{t} .$$

The variance follows immediately

$$\sigma_{{\mu_{{}} }}^{2} = \left( {t + 1} \right)\sigma_{\eta }^{2} ,$$

and

$$\sigma_{{\Delta \mu_{t + 1} }}^{2} = \left( {t + 1} \right)\sigma_{\eta }^{2} + \left( t \right)\sigma_{\eta }^{2} - 2cov\left( {\mu_{t + 1} , \mu_{t} } \right) = \left( {t + 1} \right)\sigma_{\eta }^{2} + \left( t \right)\sigma_{\eta }^{2} - 2\left( {t + 1 - 1} \right)\sigma_{\eta }^{2} = \sigma_{\eta }^{2}$$

Because the HP filter assume that \(W_{t}\) and \(Z_{t}\) are two independent Brownian motions I have,

$$\sigma^{2} = \sigma_{\mu }^{2} + \sigma_{c}^{2} ,$$
(18)

Because \(\mu_{t}\) is driftless a random walk process, it follows that

$$\sigma_{\mu }^{2} = \left( {t + 1 - t} \right)\sigma_{\eta }^{2} ,$$

and

$$\sigma_{\Delta \mu }^{2} = \sigma_{\eta }^{2}$$

.

Therefore,

$$q = \frac{{\sigma_{\Delta \mu }^{2} }}{{\sigma_{c}^{2} }} = \frac{{\sigma_{\eta }^{2} }}{{\sigma_{c}^{2} }} = \frac{{\sigma_{\mu }^{2} }}{{\sigma_{c}^{2} }} .$$
(19)

Appendix 3: Autocorrelated error term

This appendix provides details of derivation related to the variance of the autocorrelated error term, \(\pi_{t + 1} ,\), and the variance of the White noise error term, \(\varepsilon_{t + 1} ,\) of the stochastic mean NSM model.

3.1 The HP filter case

From Eq. (4a) and Eq. (5a) I have

$$\mathop \sum \limits_{j = 1}^{p} \rho_{j} c_{t - j} + v_{t + 1} = c_{t + 1} - \rho c_{t}$$
$$\pi_{t + 1} = \mathop \sum \limits_{j = 1}^{p} \rho_{j} c_{t - j} + v_{t + 1} + \eta_{t + 1} = \mathop \sum \limits_{j = 1}^{p} \rho_{j} c_{t - j} + \varepsilon_{t + 1} .$$

It follows that

$$\pi_{t + 1} = c_{t + 1} - \rho c_{t} + \eta_{t + 1} ,$$

Following Hodrick and Prescott (1997) and Ravn and Uhlig (2002), I make the assumption of independence between the permanent and transitory shocks, such that \(\sigma_{c,\eta }^{2} = 0\). (See Kohn & Ansley, 1987). Hence, the variance follows immediately

$$\sigma_{\pi }^{2} = \sigma_{c}^{2} + \rho^{2} \sigma_{c}^{2} - 2\rho {\text{cov}} \left( {c_{t + 1} ,c_{t} } \right) + \sigma_{\eta }^{2} .$$
$$\sigma_{\pi }^{2} = \sigma_{c}^{2} + \rho^{2} \sigma_{c}^{2} - 2\rho^{2} \sigma_{c}^{2} + \sigma_{\eta }^{2} .$$
$$\sigma_{\pi }^{2} = \left( {1 - \rho^{2} } \right)\sigma_{c}^{2} + \sigma_{\eta }^{2} .$$

Using (19) I obtain,

$$\sigma_{\pi }^{2} = \frac{{\left( {1 - \rho^{2} } \right)}}{{\left( {1 - \rho^{2} - \mathop \sum \nolimits_{j = 1}^{p} \rho_{j}^{2} } \right)}}\sigma_{v}^{2} + \sigma_{\eta }^{2} .$$

Defining \(\Omega = 1 - \rho^{2}\) and \(\phi = - \mathop \sum \limits_{j = 1}^{p} \rho_{j}^{2}\) we can write,

$$\sigma_{\pi }^{2} = \frac{\Omega }{\Omega + \phi }\sigma_{v}^{2} + \sigma_{\eta }^{2} ,$$

which can be written as,

$$\sigma_{\pi }^{2} = \frac{\Omega }{\Omega + \phi }\sigma_{v}^{2} + \sigma_{\eta }^{2} + \sigma_{v}^{2} - \sigma_{v}^{2}$$
$$= \left( {\frac{\Omega }{\Omega + \phi } - 1} \right)\sigma_{v}^{2} + \sigma_{\eta }^{2} + \sigma_{v}^{2} ,$$
$$= \left( {\frac{\Omega }{\Omega + \phi } - 1} \right)\sigma_{v}^{2} + \sigma_{{}}^{2} ,$$
$$= \left( {\frac{\Omega }{\Omega + \phi } - 1} \right)\sigma_{v}^{2} + \sigma_{{}}^{2} ,$$
$$\sigma_{\pi }^{2} = - \frac{\phi }{\Omega + \phi }\sigma_{v}^{2} + \sigma_{{}}^{2} ,$$

Because \(\Omega = 1 - \rho^{2}\) and \(\alpha_{1} = \rho - 1,\) I can write: \(\Omega = \alpha_{1}^{2} - 2\alpha_{1}\), then I have,

\(\sigma_{\pi }^{2} = \left( { - \frac{\varphi }{{\left( {\varphi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}}} \right)\sigma_{v}^{2} + \sigma^{2}\).

Because \(q = \frac{{\sigma_{\eta }^{2} }}{{\sigma_{v}^{2} }}\) we obtain,

$$\sigma_{\eta }^{2} = q\sigma_{v}^{2} ,$$
$$\sigma_{{}}^{2} = \left( {1 + q} \right)\sigma_{v}^{2} ,$$
$$\sigma_{v}^{2} = \left( {\frac{1}{1 + q}} \right)\sigma_{{}}^{2} .$$

Plug this in (19) I get,

$$\sigma_{\pi }^{2} = \left( { - \frac{\phi }{\Omega + \phi }} \right)\left( {\frac{1}{1 + q}} \right)\sigma_{{}}^{2} + \sigma_{{}}^{2} ,$$
$$\sigma_{\pi }^{2} = \left( {1 - \frac{\phi }{{\left( {\Omega + \phi } \right)\left( {1 + q} \right)}}} \right)\sigma_{{}}^{2} .$$

Because \(\Omega = 1 - \rho^{2}\) and \(\alpha_{1}^{{}} = \rho - 1,\) we can write: \(\Omega = \alpha_{1}^{2} - 2\beta\), then we have,

$$\sigma_{\pi }^{2} = \left( {1 - \frac{\phi }{{\left( {\phi + \alpha_{1}^{2} - 2\beta } \right)\left( {1 + q} \right)}}} \right)\sigma_{{}}^{2} ,$$
$$\sigma_{\pi }^{2} = \left( {\frac{{\left( {\alpha_{1}^{2} - 2\alpha_{1} } \right)\left( {1 + q} \right) + q\phi }}{{\left( {\phi + \alpha_{1}^{2} - 2\alpha_{1} } \right)\left( {1 + q} \right)}}} \right)\sigma_{{}}^{2} ,$$
$$\sigma_{\pi }^{2} = \left( {\frac{{\left( {\alpha_{1}^{2} - 2\alpha_{1} } \right) + \frac{q}{{\left( {1 + q} \right)}}\phi }}{{\left( {\phi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}}} \right)\sigma^{2}$$

Let

$$\partial = \left( {\frac{{\left( {\alpha_{1}^{2} - 2\alpha_{1} } \right) + \frac{q}{{\left( {1 + q} \right)}}\phi }}{{\left( {\phi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}}} \right)$$

Then

$$\sigma_{\pi }^{2} = \partial \sigma^{2} .$$

3.2 The BN filter case

Because the BN filter assumes that \(Corr\left({\eta }_{t},{\nu }_{t}\right)={\rho }_{\eta \upsilon }=1\), I have

$$\sigma_{\pi }^{2} = \left( { - \frac{\varphi }{{\left( {\varphi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}}} \right)\sigma_{v}^{2} + \sigma_{v}^{2} + \sigma_{\mu }^{2} + 2\sigma_{v} \sigma_{\mu }$$
$$\sigma_{\pi }^{2} = \left( {1 - \frac{\varphi }{{\left( {\varphi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}}} \right){ }\sigma_{v}^{2} + \sigma_{\mu }^{2} + 2\sigma_{v} \sigma_{\mu }$$
$$\sigma_{\pi }^{2} = \left( {1 - \frac{\varphi }{{\left( {\varphi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}}} \right){ }\sigma_{v}^{2} + \sigma^{2} - \sigma_{v}^{2}$$
$$\sigma_{\pi }^{2} = \left( { - \frac{\varphi }{{\left( {\varphi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}}} \right)\sigma_{v}^{2} + \sigma^{2}$$
$$\sigma_{\pi }^{2} = \left( {1 - \frac{\varphi }{{\left( {\varphi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}} + 2\frac{{\sigma_{\mu } }}{{\sigma_{v} }}} \right){ }\sigma_{v}^{2} + \sigma_{\mu }^{2}$$
$$\sigma_{\pi }^{2} = \left( { - \frac{\varphi }{{\left( {\varphi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}} + 2\frac{{\sigma_{\mu } }}{{\sigma_{v} }}} \right){ }\sigma_{v}^{2} + \sigma_{v}^{2} + \sigma_{\mu }^{2}$$

Define \(\lambda =\frac{{\sigma }_{v}^{2}}{{\sigma }^{2}}\) I obtain,

$$\sigma_{\pi }^{2} = \left( { - \frac{\varphi }{{\left( {\varphi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}}} \right)\lambda \sigma^{2} + \sigma^{2}$$
$$\sigma_{\pi }^{2} = \left( {1 - \frac{\varphi }{{\left( {\varphi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}}\lambda } \right)\sigma^{2}$$

Appendix 4: GMM estimation

Define λ as the entire parameter vector. I have the following orthogonality conditions:

  1. 1.

    The Aït-Sahalia Model: \(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} - \alpha_{0} - \alpha_{1} r_{t} - \alpha_{2} r_{t}^{2} - \alpha_{3} r_{t}^{ - 1} } \right]\). The moment conditions are given by:

    $$h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\varepsilon_{t + 1} , \varepsilon_{t + 1} r_{t } , \varepsilon_{t + 1} r_{t}^{2} , \varepsilon_{t + 1} r_{t}^{ - 1} , \varepsilon_{t + 1}^{2} - \sigma^{2} r_{t}^{2\gamma } , \left( {\varepsilon_{t + 1}^{2} - \sigma^{2} r_{t}^{2\gamma } } \right)r_{t } } \right]$$
  2. 2.

    The CKLS model: \(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} - \alpha_{0} - \alpha_{1} r_{t} } \right]\). The moment conditions are given by:

    $$h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\varepsilon_{t + 1} , \varepsilon_{t + 1} r_{t, } \varepsilon_{t + 1}^{2} - \sigma^{2} r_{t}^{2\gamma } , \left( {\varepsilon_{t + 1}^{2} - \sigma^{2} r_{t}^{2\gamma } } \right)r_{t } } \right]$$
  3. 3.

    The AG model: \(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} - \alpha_{0} - \alpha_{1} r_{t} } \right]\). The moment conditions are given by:

    $$h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\varepsilon_{t + 1} , \varepsilon_{t + 1} r_{t, } \varepsilon_{t + 1} r_{t}^{2} , \varepsilon_{t + 1}^{2} - \sigma^{2} r_{t}^{2\gamma } , \left( {\varepsilon_{t + 1}^{2} - \sigma^{2} r_{t}^{3} } \right)r_{t } , \left( {\varepsilon_{t + 1}^{2} - \sigma^{2} r_{t}^{3} } \right)r_{t}^{3} } \right]$$
  4. 4.

    The CIR Model: \(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} - \alpha_{0} - \alpha_{1} r_{t} } \right]\). The moment conditions are given by:

    $$h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\varepsilon_{t + 1} , \varepsilon_{t + 1} r_{t, } \varepsilon_{t + 1}^{2} - \sigma^{2} r_{t} , \left( {\varepsilon_{t + 1}^{2} - \sigma^{2} r_{t } } \right)r_{t } } \right]$$
  5. 5.

    The Vasicek Model: \(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} - \alpha_{0} - \alpha_{1} r_{t} } \right]\). The moment conditions are given by:

    $$h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\varepsilon_{t + 1} , \varepsilon_{t + 1} r_{t, } \varepsilon_{t + 1}^{2} - \sigma^{2} , \left( {\varepsilon_{t + 1}^{2} - \sigma^{2} } \right)r_{t } } \right]$$
  6. 6.

    The BDF Model: \(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} + \alpha_{1} \left( {\theta_{t} - r_{t} } \right)} \right]\). The moment conditions are given by:

    $$h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\varepsilon_{t + 1} , \varepsilon_{t + 1} r_{t, } \varepsilon_{t + 1} r_{1, t } , \varepsilon_{t + 1} r_{2, t ,} \varepsilon_{t + 1}^{2} - \sigma^{2} , \left( {\varepsilon_{t + 1}^{2} - \sigma^{2} } \right)r_{t } } \right]$$

    where

    $$\theta_{t} = \alpha_{\theta ,1} \left[ {B\left( {T_{2} } \right)T_{1} r_{1,t} - B\left( {T_{1} } \right)T_{2} r_{2,T} } \right]$$
    $$B\left( T \right) = \frac{{2\left( {e^{\delta T} - 1} \right)}}{{\left( {\delta + k} \right)\left( {e^{\delta T} - 1} \right) + 2\delta }}$$
    $$\sigma^{2} = \sigma_{0}^{2} + \sigma_{1}^{2} r_{t}$$

    and

    $$\delta = \sqrt {\left( {\delta^{2} + \sigma_{1}^{2} } \right)}$$

    where \(\sigma_{1} = 0\) if BDF-VAS is considered and \(\sigma_{0} = 0\) if BDF-CIR is considered.

  7. 7.

    The Heston Model: \(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} - \alpha_{0} - \alpha_{1} r_{t} } \right]\) and \(\varepsilon_{v, t + 1} = \left[ {v_{t + 1} - v_{t} - \alpha_{v} - \alpha_{v} v_{t} } \right]\). The moment conditions are given by:

    $$h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\varepsilon_{t + 1} , \varepsilon_{v, t + 1} , \varepsilon_{t + 1} r_{t} , \varepsilon_{v, t + 1} v_{t } , \varepsilon_{v, t + 1}^{2} - \sigma_{v}^{2} , \left( {\varepsilon_{v, t + 1}^{2} - \sigma_{v}^{2} } \right)v_{t } } \right]$$

    I follow Andersen, and Lund (1997) and specify \(v_{t + 1}\) as a GARCH(1,1) model.

  8. 8.

    The Chen Model: \(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} + \alpha_{1} \left( {\theta_{t} - r_{t} } \right)} \right],\) \(\varepsilon_{v, t + 1} = \left[ {v_{t + 1} - v_{t} - \alpha_{vo} - \alpha_{v1} v_{t} } \right],\) and \(\varepsilon_{\theta , t + 1} = \left[ {\theta_{t + 1} - \theta_{t} - \alpha_{\theta o} - \alpha_{\theta 1} \theta_{t} } \right].\) The moment conditions are given by: \(h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\varepsilon_{t + 1} , \varepsilon_{v, t + 1} ,\varepsilon_{\theta , t + 1} , \varepsilon_{t + 1} r_{t, } \varepsilon_{v, t + 1} v_{t, } , \varepsilon_{\theta , t + 1} \theta_{t, } , \varepsilon_{v, t + 1}^{2} - \sigma_{v}^{2} , \left( {\varepsilon_{v, t + 1}^{2} - \sigma_{v}^{2} ,} \right)v_{t } , \varepsilon_{\theta , t + 1}^{2} - \sigma_{\theta }^{2} , \left( {\varepsilon_{\theta , t + 1}^{2} - \sigma_{\theta }^{2} ,} \right)\theta_{t } } \right]\). I follow Andersen, and Lund (1997) and specify \(v_{t + 1}\) as a GARCH(1,1) model.

  9. 9.

    The Al-Zoubi (2019) I(2) model: \(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} - \alpha_{1} c_{t} } \right]\).

    $$h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\pi_{t + 1} , \varepsilon_{t + 1} c_{t} , \varepsilon_{t + 1}^{2} - \sigma^{2} , \left( {\varepsilon_{t + 1}^{2} - \sigma^{2} } \right)c_{t} } \right]$$
  10. 10.

    The NSM-HP and NSM-bHP Models:\(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} - \alpha_{1} c_{t} } \right]\). Letting:

    $$\partial = \left( {\frac{{\left( {\alpha_{1}^{2} - 2\alpha_{1} } \right) + \frac{q}{{\left( {1 + q} \right)}}\phi }}{{\left( {\phi + \alpha_{1}^{2} - 2\alpha_{1} } \right)}}} \right)$$

    then

    $$h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\pi_{t + 1} , \pi_{t + 1} c_{t} , \pi_{t + 1}^{2} - \partial \sigma^{2} , \left( {\pi_{t + 1}^{2} - \partial \sigma^{2} } \right)c_{t} } \right]$$

    .

  11. 11.

    The NSM-HAM Model:

    \(\varepsilon_{t + 1} = \left[ {r_{t + 1} - r_{t} - \alpha_{1} \left( {r_{t} - b_{0} - b_{1} r_{t - 7} - b_{2} r_{t - 8} - b_{3} r_{t - 9} - b_{4} r_{t - 10} } \right) } \right]\). The moment conditions are given by:

    \(h\left( {r_{t + 1} ,\lambda } \right) = \left[ {\varepsilon_{t + 1} , \varepsilon_{t + 1} r_{t} , \varepsilon_{t + 1}^{2} - \sigma^{2} , \left( {\varepsilon_{t + 1}^{2} - \sigma^{2} } \right)r_{t } } \right]\).

To test the validity of my model, I minimize the GMM criterion of the form,

$$\frac{1}{T}\left[ {f\left( {x_{t + 1} ,Y_{t} ,\lambda } \right)} \right]{\prime} W_{T} \frac{1}{T}\left[ {f\left( {x_{t + 1} ,Y_{t} ,\lambda } \right)} \right],$$
(18)

where WT is a consistent estimate of \(\left( {{\text{var}} \left[ {\left( {1/T} \right)\left( {f\left( {x_{t + 1} ,Y_{t} ,\lambda } \right)_{t} } \right)} \right]} \right)^{ - 1}\) and Yt is a K-dimensional vector of instrumental variables.

Under the null hypothesis that GMM restrictions are valid, I have that:

$$\frac{1}{T}\left[ {f\left( {x_{t + 1} ,Y_{t} ,\lambda } \right)} \right]{\prime} W_{T} \frac{1}{T}\left[ {f\left( {x_{t + 1} ,Y_{t} ,\lambda } \right)} \right]\mathop \sim \limits^{a} \chi_{L - k}^{2} .$$
(19)

For my model to be robust with respect to heteroskedasticity and autocorrelation variance, I follow Inoue and Shintani (2006) and use the Parzen kernel of Gallant (1987) with two lags to calculate the moments weighting matrix.

Rights and permissions

Springer Nature or its licensor (e.g. a society or other partner) holds exclusive rights to this article under a publishing agreement with the author(s) or other rightsholder(s); author self-archiving of the accepted manuscript version of this article is solely governed by the terms of such publishing agreement and applicable law.

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Al-Zoubi, H.A. An affine model for short rates when monetary policy is path dependent. Rev Deriv Res (2024). https://doi.org/10.1007/s11147-024-09202-3

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s11147-024-09202-3

Keywords

JEL Classification

Navigation