Skip to main content
Log in

Bayesian Local Likelihood Estimation of Time-Varying DSGE Models: Allowing for Indeterminacy

  • Published:
Computational Economics Aims and scope Submit manuscript

Abstract

This paper modifies and employs a Bayesian Local Likelihood approach to estimate time-varying parameters of a New Keynesian model and assess such time variations using US data. Our modification contributes to the expanding literature by novelly integrating indeterminacy into the time-varying estimator. Further, we implement a one-step approach based on a unified solution set obtained by the augmented linearized rational expectation model simultaneously allowing for both determinacy and indeterminacy. The evidences suggest substantial time-variations in many parameters, particularly those associated with the Fed monetary policy rule and characterized by volatilities in the economy. This study also shows that allowing time-varying parameters improves density and point forecasts in comparison to a fixed-parameter DSGE model. We investigate implications of the time-variation for monetary policy effectiveness and find that the increase in the policy response to inflation from the pre-1979 to the post-1982 alone does not suffice for explaining the U.S. economy’s shift to determinacy, unless it is accompanied by either the estimated decline in trend inflation or the estimated change in policy responses to the output growth.

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

Similar content being viewed by others

Notes

  1. Roughly speaking, indeterminacy refers to the multiplicity of rational expectations equilibria while an equilibrium that is locally isolated and uniquely determined by preferences and technologies is called determinate. See Farmer (1999) for a formal definition.

  2. Bianchi and Nicolò (2021) show that their characterization of indeterminate equilibria is equivalent to Lubik and Schorfheide (2003).

References

  • Andreasen, M. M., Fernández-Villaverde, J., & Rubio-Ramírez, J. F. (2018). The pruned state-space system for non-linear DSGE models: Theory and empirical applications. The Review of Economic Studies, 85(1), 1–49.

    Article  Google Scholar 

  • Aguiar-Conraria, L., Martins, M. M., & Soares, M. J. (2019). The Phillips Curve at 60: time for time and frequency. Bank of Finland Research Discussion Paper No. 12.

  • Bae, J., Kim, C. J., & Kim, D. H. (2012). The evolution of the monetary policy regimes in the US. Empirical Economics, 43, 617–649.

    Article  Google Scholar 

  • Benati, L., & Haroon M. (2007). U.S. evolving macroeconomic dynamics: A structural investigation. ECB Working Paper 746.

  • Bianchi, F. (2013). Regime switches, agents’ beliefs, and post-World War II US macroeconomic dynamics. Review of Economic Studies, 80(2), 463–490.

    Article  Google Scholar 

  • Bianchi, F., & Nicolò, G. (2021). A generalized approach to indeterminacy in linear rational expectations models. Quantitative Economics, 12(3), 843–868.

    Article  Google Scholar 

  • Blanchard, O. J., & Kahn, C. M. (1980). The solution of linear difference models under rational expectations. Econometrica: Journal of the Econometric Society, 1305–1311.

  • Boivin, J. (2006). Has U.S. monetary policy changed? Evidence from drifting coefficients and real-time data. Journal of Money, Credit, and Banking, 38(5), 1149–1173.

    Article  Google Scholar 

  • Canova, F. (2009). What explains the great moderation in the US? A structural analysis. Journal of the European Economic Association, 7(4), 697–721.

    Article  Google Scholar 

  • Cantore, C., Levine, P., Melina, G., & Yang, B. (2012). A fiscal stimulus with deep habits and optimal monetary policy. Economics Letters, 117(1), 348–353.

    Article  Google Scholar 

  • Castelnuovo, E. (2012). Estimating the evolution of money’s role in the US monetary business cycle. Journal of Money, Credit and Banking, 44(1), 23–52.

    Article  Google Scholar 

  • Castelnuovo, E., & Fanelli, L. (2015). Monetary policy indeterminacy and identification failures in the US: Results from a robust test. Journal of Applied Econometrics, 30(6), 924–947.

    Article  Google Scholar 

  • Christiano, L. J., Eichenbaum, M. S., & Trabandt, M. (2015). Understanding the great recession. American Economic Journal: Macroeconomics, 7(1), 110–167.

    Google Scholar 

  • Cihak, M., & Nier, E. (2012). The need for special resolution regimes for financial institutions-The case of the European Union. Harvard Business Law Review, 2(2), 395–434.

    Google Scholar 

  • Clarida, R., Gali, J., & Gertler, M. (1999). The science of monetary policy: A new Keynesian perspective. Journal of Economic Literature, 37(4), 1661–1707.

    Article  Google Scholar 

  • Clarida, R., Gali, J., & Gertler, M. (2000). Monetary policy rules and Macroeco- nomic stability: Evidence and some theory. Quarterly Journal of Economics, 115, 147–180.

    Article  Google Scholar 

  • Cogley, T., & Sargent, T. J. (2005). Drifts and volatilities: Monetary policies and outcomes in the post WWII US. Review of Economic Dynamics, 8(2), 262–302.

    Article  Google Scholar 

  • Coibion, O., & Gorodnichenko, Y. (2011). Monetary policy, trend inflation, and the great moderation: An alternative interpretation. American Economic Review, 101(1), 341–370.

    Article  Google Scholar 

  • Davig, T., & Leeper, E. M. (2007). Generalizing the Taylor principle. American Economic Review, 97(3), 607–635.

    Article  Google Scholar 

  • Davig, T., & Leeper, E. M. (2010). Generalizing the Taylor principle: Reply. American Economic Review, 100(1), 618–24.

    Article  Google Scholar 

  • Davig, T., Leeper, E. M., & Walker, T. B. (2010). ‘Unfunded liabilities’ and uncertain fiscal financing. Journal of Monetary Economics, 57(5), 600–619.

    Article  Google Scholar 

  • Del Negro, M., Lenza, M., Primiceri, G. E., & Tambalotti, A. (2020). What’s up with the Phillips Curve? (No. w27003). National Bureau of Economic Research.

  • Diebold, F. X., & Mariano, R. S. (2002). Comparing predictive accuracy. Journal of Business & Economic Statistics, 20(1), 134–144.

    Article  Google Scholar 

  • Diebold, F. X., Schorfheide, F., & Shin, M. (2017). Real-time forecast evaluation of DSGE models with stochastic volatility. Journal of Econometrics, 201(2), 322–332.

    Article  Google Scholar 

  • Domit, S., Monti, F., & Sokol, A. (2019). Forecasting the UK economy with a medium-scale Bayesian VAR. International Journal of Forecasting, 35(4), 1669–1678.

    Article  Google Scholar 

  • Edge, R. M., Gürkaynak, R. S., & Kisacikoglu, B. (2013). Judging the DSGE model by its forecast. Mimeo.

  • Farmer, R. E., & Nicolò, G. (2019). Some international evidence for Keynesian economics without the Phillips curve. The Manchester School (forthcoming).

  • Farmer, L. E., & Toda, A. A. (2017). Discretizing nonlinear, non-Gaussian Markov processes with exact conditional moments. Quantitative Economics, 8(2), 651–683.

    Article  Google Scholar 

  • Farmer, R. E., Waggoner, D. F., & Zha, T. (2009). Understanding Markovswitching rational expectations models. Journal of Economic Theory, 144(5), 1849–1867.

    Article  Google Scholar 

  • Farmer, R. E., Waggoner, D. F., & Zha, T. (2011). Minimal state variable solutions to Markov-switching rational expectations models. Journal of Economic Dynamics and Control, 35(12), 2150–2166.

    Article  Google Scholar 

  • Fernández-Villaverde, J., & Rubio-Ramírez, J. F. (2007). Estimating macroeconomic models: A likelihood approach. The Review of Economic Studies, 74(4), 1059–1087.

    Article  Google Scholar 

  • Fernández-Villaverde, J., Rubio-Ramírez, J. F., Cogley, T., & Schorfheide, F. (2007). How structural are structural parameters? [with comments and discussion]. NBER Macroeconomics Annual, 22, 83–167.

    Article  Google Scholar 

  • Foerster, A., Rubio-Ramirez, J., Waggoner, D. F., & Zha, T. (2016). Perturbation methods for Markov-switching DSGE models. Quantitative Economics, 7(2), 637–669.

    Article  Google Scholar 

  • Galí, J. (2015). Monetary policy, inflation, and the business cycle: An introduction to the new Keynesian framework and its applications. Princeton University Press.

  • Gali, J., & Gertler, M. (1999). Inflation dynamics: A structural econometric analysis. Journal of Monetary Economics, 44(2), 195–222.

    Article  Google Scholar 

  • Galvão, A. B., Giraitis, L., Kapetanios, G., & Petrova, K. (2015). A Bayesian local likelihood method for modelling parameter time variation in DSGE models (No. 770). Working Paper.

  • Galvão, A. B., Giraitis, L., Kapetanios, G., & Petrova, K. (2016). A time varying DSGE model with financial frictions. Journal of Empirical Finance, 38, 690–716.

    Article  Google Scholar 

  • Giraitis, L., Kapetanios, G., & Yates, T. (2014). Inference on stochastic time-varying coefficient models. Journal of Econometrics, 179(1), 46–65.

    Article  Google Scholar 

  • Good, I. J. (1992). Rational decisions. Journal of the Royal Statistics Society Series B, 14, 107–114.

    Google Scholar 

  • Hoffmann, B., Peersman, G., & Straub, R. (2012). Time variation in US wage dynamics. Journal of Monetary Economics, 59(8), 769–783.

    Article  Google Scholar 

  • Ireland, P. N. (2007). Changes in the Federal Reserve’s inflation target: Causes and consequences. Journal of Money, Credit and Banking, 39(8), 1851–1882.

    Article  Google Scholar 

  • Kim, C.-J., & Nelson, C. R. (2006). Estimation of a forward-looking monetary policy rule: A time-varying parameter model using ex post data. Journal of Monetary Economics, 53, 1949–1966.

    Article  Google Scholar 

  • Justiniano, A., & Primiceri, G. E. (2008). The time-varying volatility of macroeconomic fluctuations. American Economic Review, 98(3), 604–41.

    Article  Google Scholar 

  • Kapetanios, G., Masolo, R. M., Petrova, K., & Waldron, M. (2019). A time-varying parameter structural model of the UK economy. Journal of Economic Dynamics and Control, 106, 103705.

    Article  Google Scholar 

  • Kim, J. Y. (2002). Limited information likelihood and Bayesian analysis. Journal of Econometrics, 107(1–2), 175–193.

    Article  Google Scholar 

  • King, R. G. (2000). The new IS-LM model: Language, logic, and limits. FRB Richmond Economic Quarterly, 86(3), 45–104.

    Google Scholar 

  • Kurozumi, T., & Van Zandweghe, W. (2017). Trend inflation and equilibrium stability: Firm-specific versus homogeneous labor. Macroeconomic Dynamics, 21(4), 947–981.

    Article  Google Scholar 

  • Liu, Z., Waggoner, D. F., & Zha, T. (2011). Sources of macroeconomic fluctuations: A regime-switching DSGE approach. Quantitative Economics, 2(2), 251–301.

    Article  Google Scholar 

  • Lubik, T. A., & Matthes, C. (2016). Indeterminacy and learning: An analysis of monetary policy in the Great Inflation. Journal of Monetary Economics, 82, 85–106.

    Article  Google Scholar 

  • Lubik, T. A., & Schorfheide, F. (2003). Computing sunspot equilibria in linear rational expectations models. Journal of Economic dynamics and control, 28(2), 273–285.

    Article  Google Scholar 

  • Lubik, T. A., & Schorfheide, F. (2004). Testing for indeterminacy: An application to US monetary policy. American Economic Review, 94(1), 190–217.

    Article  Google Scholar 

  • Mavroeidis, S. (2010). Monetary policy rules and macroeconomic stability: Some new evidence. American Economic Review, 100(1), 491–503.

    Article  Google Scholar 

  • Mavroeidis, S., Plagborg-Møller, M., & Stock, J. H. (2014). Empirical evidence on inflation expectations in the New Keynesian Phillips Curve. Journal of Economic Literature, 52(1), 124–88.

    Article  Google Scholar 

  • Orphanides, A. (2002). Monetary policy rules and the great inflation. American Economic Review, 92, 115–121.

    Article  Google Scholar 

  • Orphanides, A. (2004). Monetary policy rules, macroeconomic stability, and inflation: A view from the trench. Journal of Money, Credit, and Banking, 36(2), 151–175.

    Article  Google Scholar 

  • Petrova, K. (2019). Quasi-Bayesian estimation of time-varying volatility in DSGE models. Journal of Time Series Analysis, 40(1), 151–157.

    Article  Google Scholar 

  • Rotemberg, J., & Woodford, M. (1997). An optimization-based econometric framework for the evaluation of monetary policy. NBER Macroeconomics Annual, 12, 246–297.

    Article  Google Scholar 

  • Romer, C. D., & Romer, D. H. (2002). The evolution of economic understanding and postwar stabilization policy. Rethinking stabilization policy (pp. 11–78). Federal Reserve Bank of Kansas City.

  • Romer, C. D., & Romer, David H. (2004). Choosing the federal reserve chair: Lessons from history. Journal of Economic Perspectives, 18(1), 129–162.

    Article  Google Scholar 

  • Sbordone, A. M. (2002). Prices and unit labor costs: A new test of price stickiness. Journal of Monetary Economics, 49, 265–292.

    Article  Google Scholar 

  • Schorfheide, F. (2000). Loss function-based evaluation of DSGE models. Journal of Applied Econometrics, 15(6), 645–670.

    Article  Google Scholar 

  • Schorfheide, F. (2005). VAR forecasting under misspecification. Journal of Econometrics, 128(1), 99–136.

    Article  Google Scholar 

  • Sims, C. (2001). Matlab Code for Solving Linear Rational Expectations Models. QM&RBC Codes.

  • Sims, C. A. (2002). Solving linear rational expectations models. Computational Economics, 20(1–2), 1–20.

    Article  Google Scholar 

  • Sims, C. A., & Zha, T. (2006). Were there regime switches in US monetary policy?. American Economic Review 96.1(2006), 54–81.

    Article  Google Scholar 

  • Woodford, M. (2003). Interest and prices: Foundation of a theory of monetary policy. Princeton University Press.

  • Wu, J. C., & Xia, F. D. (2016). Measuring the macroeconomic impact of monetary policy at the zero lower bound. Journal of Money, Credit and Banking, 48(2–3), 253–291.

    Article  Google Scholar 

Download references

Acknowledgements

We acknowledge financial support from the research project ‘Research on Parameter Identification and Application of DSGE Models Allowing for Indeterminacy’ (Approval Number: 71863008), and from the research project ‘A Research on Bias Correction of Long Memory Semi-parametric Estimators and Application of ARFIMA Models under High-Frequency Cantagion’ (Approval Number: 72263010), sponsored by National Natural Science Foundation of China.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Luyao Wu.

Ethics declarations

Competing interests

Authors have no competing interests to declare.

Additional information

Publisher's Note

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

Appendices

Appendix A

The New Keynesian model in the main body can be represented as the standard Sims (2001) form as follows.

$$\begin{aligned} \Gamma _0(\theta )\textbf{X}_t=\Gamma _1(\theta )\textbf{X}_{t-1}+\Psi (\theta )\epsilon _t+\Pi (\theta )\eta _t \end{aligned}$$

The specific matrices have relevant forms below.

$$\begin{aligned}{} & {} \Gamma _0(\theta )\equiv \begin{bmatrix} 1 &{} 0 &{} \tau &{} -1 &{} -\tau &{} -1 &{} 0 \\ \kappa &{} -1 &{} 0 &{} 0 &{} \beta &{} 0 &{} -\kappa \\ \left( 1-\rho _R\right) \psi _2 &{} \left( 1-\rho _R\right) \psi _1 &{} -1 &{} 0 &{} 0 &{} 0 &{} -\left( 1-\rho _R\right) \psi _2 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 1 &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 1 \\ 1 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 \\ 0 &{} 1 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 \end{bmatrix} \\{} & {} \Gamma _1(\theta )\equiv \begin{bmatrix} 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} -\rho _i &{} 0 &{} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} \rho _g &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} 0 &{} \rho _z \\ 0 &{} 0 &{} 0 &{} 1 &{} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 &{} 0 &{} 1 &{} 0 &{} 0 \\ \end{bmatrix},\quad \Psi (\theta )\equiv \begin{bmatrix} 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 \\ -1 &{} 0 &{} 0 \\ 0 &{} 1 &{} 0 \\ 0 &{} 0 &{} 1 \\ 0 &{} 0 &{} 0 \\ 0 &{} 0 &{} 0 \\ \end{bmatrix},\quad \Pi (\theta )\equiv \begin{bmatrix} 0 &{} 0 \\ 0 &{} 0 \\ 0 &{} 0 \\ 0 &{} 0 \\ 0 &{} 0 \\ 1 &{} 0 \\ 0 &{} 1 \\ \end{bmatrix} \end{aligned}$$

Appendix B

Following Petrova (2019), we adopt the BLL estimation methodology to obtain estimates of time-varying parameters in DSGE models. We start with a brief outline description of this method. Conditionally on a batch of data \(\textbf{D}^T=\left( \textbf{D}_1,\cdots ,\textbf{D}_T\right) \), Galvão et al. (2015) show that if time-varying parameters \(\theta _t\) satisfy the condition

$$\begin{aligned} \underset{j:|j-t|\le h}{\sup }\begin{Vmatrix}\theta _t-\theta _j\end{Vmatrix}^2=O_p\left( \frac{h}{t}\right) ,\text { for }t\rightarrow \infty ,1\le h \le t, \end{aligned}$$
(20)

then the conditional log-density function \({l}\left( \textbf{D}_j|F_{j-1}\right) \) is only dependent on the parameter values in period j and the information set \(F_{j-1}\) collected at the j-th time-point.

$$\begin{aligned} l\left( \textbf{D}_j|F_{j-1}\right) =l\left( \textbf{D}_j|\textbf{D}^{j-1},\theta _j\right) \end{aligned}$$

where \(\textbf{D}^{j-1}\equiv (\textbf{D}_1,\textbf{D}_2,...,\textbf{D}_{j-1})\). Observed series \(\left\{ \textbf{D}_t\right\} _{t=1}^T\) can be used to construct the sample pseudo-likelihood function:

$$\begin{aligned} L_t\left( \textbf{D}^T|\theta _t\right) =\prod _{j=1}^{T}L\left( \textbf{D}_t|\textbf{D}^{j-1},\theta _t\right) ^{w_{tj}},\text { for }t=1,...,T, \end{aligned}$$
(21)

where the weights are computed by the kernel function:

$$\begin{aligned} w_{tj}=K\left( \frac{t-j}{H}\right) , \text { for }t,j=1,...,T, \end{aligned}$$

where H is a bandwidth parameter to be selected appropriately. For estimating time-varying parameters, we choose a Gaussian kernel function in (22) to calculate the weight matrix, where the bandwidth parameter H is set equal to \(T^{0.5}\). Then, the kernel function is

$$\begin{aligned} w_{tj}=\left( \frac{1}{\sqrt{2\pi }}\right) \exp \left\{ \left( -\frac{1}{2}\right) \left( \frac{t-j}{H}\right) ^2\right\} \end{aligned}$$
(22)

The kernel approach above is extended to Bayesian inference by Petrova (2019), who shows that the resulting quasi-posterior distributions are asymptotically Gaussian and valid for constructing confidence intervals as long as the weights \(\tilde{w}_{tj}\) are modified as follows:

$$\begin{aligned} w_{tj}=\left( \sum _{j=1}^{T}\tilde{w}_{tj}^2\right) ^{-1}\left( \frac{\tilde{w}_{tj}}{\sum _{j=1}^{T}\tilde{w}_{tj}}\right) ,\text { for }j,t=1,...,T. \end{aligned}$$

One of the common methodologies for estimating \(\theta _t\) in Equation (2) and (19) is to employ the implied one-step ahead prediction distribution \(p_{\theta _t}\left( \textbf{D}_t|\textbf{D}_{t-1}\right) \) computed in (21) to form the likelihood of the observed series. Therefore, the information set \(\textbf{D}^{t-1}\) can be simplified into \(\textbf{D}_{t-1}\). Then the estimates of \(\theta _t\) are yielded by maximizing the time-varying posterior probability function which, by Bayesian rule, can be represented as the product of the local pseudo-likelihood in period t and the prior probability of parameters.

$$\begin{aligned} \begin{aligned} \pi (\theta _t)&\equiv p_t\left( \theta _t|\textbf{D}^T\right) \\&=\frac{L_t\left( \textbf{D}^T|\theta _t\right) p\left( \theta _t\right) }{p\left( \textbf{D}^T\right) } \propto \prod _{j=1}^{T}L \left( \textbf{D}_j|\textbf{D}_{j-1},\theta _t\right) ^{w_{tj}}p\left( \theta _t\right) \end{aligned} \end{aligned}$$

For convenience, the formula above usually is taken in logarithms:

$$\begin{aligned} \log p_t\left( \theta _t|\textbf{D}^T\right) =-\sum _{j=1}^{T} w_{tj}\log L\left( \textbf{D}_t|\textbf{D}_{j-1},\theta _t\right) -\log p\left( \theta _t\right) . \end{aligned}$$
(23)

The standard Kalman filter methodology is adequate to compute \(L_j\left( \textbf{D}_j|\textbf{D}_{j-1},\theta _t\right) \) for the purpose of evaluating the log-likelihood function \(L_t\left( \textbf{D}^T|\theta _t\right) \). However, this posterior is intractable as the likelihood cannot be computed in closed form. Hence, we use the Random Walk Metropolis-Hastings (RWMH) algorithm, which is one of the most widely employed standard approaches among Markov Chain Monte Carlo (MCMC) techniques to evaluate (23) numerically. The RWMH proceeds by selecting an arbitrary proposal Markov Chain \(q\left( \theta _t^i|\theta _t^{i-1}\right) \) and moving it forward by a way of randomly accepted draws from the chain at the last iterate. The acceptance probability is dependent on the posterior distribution. In the majority of Bayesian contexts, a Gaussian proposal distribution is used,

$$\begin{aligned} q\left( \theta _t^{(i)}|\theta _t^{(i-1)}\right) =\mathcal {N}\left[ \theta _t';\mu \left( \theta _t^{(i-1)}\right) ,\mathbf {\Sigma }\left( \theta _t^{(i-1)}\right) \right] , \end{aligned}$$

where \(\theta _t'\) are the candidate parameter estimates (represented as a swarm of particles) with \(\mu \) and \(\mathbf {\Sigma }\) denoting the mean and covariance matrix of the joint Gaussian distribution, respectively. The Gaussian mean, obtained by a maximization routine, is the mode of local posterior distribution, which is different during every period.

Once the inverse Hessian matrix of log likelihood is evaluated at posterior modes \(\mu \left( \theta _t\right) =\hat{\theta }_t\), the variance \(\mathbf {\Sigma }\left( \hat{\theta }_t \right) \) of the proposal densities of parameters can be obtained. Afterwards, parameter initial values \(\theta _t^0\) can be drawn from the joint Gaussian distribution,

$$\begin{aligned} \mathcal {N}\left[ \theta _t^0;\hat{\theta _t},c_0^2\mathbf {\Sigma }\left( \hat{\theta _t}\right) \right] . \end{aligned}$$

After the initial step, the MH recursive steps proceed. For \(i=1,...,n_{sim}\), we draw particles \(\theta _t^{(i)}\) from the proposal Gaussian distribution,

$$\begin{aligned} \mathcal {N}\left[ \theta _t';\theta _t^{(i-1)},c^2\mathbf {\Sigma }\left( \hat{\theta _t}\right) \right] \end{aligned}$$
(24)

With respect to each draw, the covariance matrix of the Gaussian distribution is invariant. After generating a candidate parameter set via (24), it is accepted in the acceptance probability as follows,

$$\begin{aligned} \alpha \left( \theta _t',\theta _t^{(i-1)}\right) =\textbf{1}\wedge \frac{\pi (\theta _t')q\left( \theta _t^{(i-1)}|\theta _t'\right) }{\pi \left( \theta _t^{(i-1)}\right) q\left( \theta _t'|\theta _t^{(i-1)}\right) } \end{aligned}$$

where \(a\wedge b=\min (a,b)\). We set \(\theta _t^{(i)}\leftarrow \theta _t'\) if the candidate parameter particle set is accepted and \(\theta _t^{(i)}\leftarrow \theta _t^{(i-1)}\) if it is rejected. \(c_0^2\) and \(c^2\) are scaling parameters that tune the step length of the RWMH algorithm. Tuning them, we aim to keep acceptance probability within the range 25–35%.

Appendix C

1.1 The Point Forecasts are Computed as Posterior Means

$$\begin{aligned} \hat{\textbf{D}}_{T+h|T}=\int _{\textbf{D}_{T+h}}\textbf{D}_{T+h}\times p\left( \textbf{D}_{T+h}|\textbf{D}_{1:T}\right) \textbf{d}\textbf{D}_{T+h}\approx \frac{1}{n_{sim}}\sum _{j=1}^{n_{sim}}\textbf{D}_{T+h}^{(j)} \end{aligned}$$
(25)

where the draws \(\textbf{D}_{T+h}^{(j)}\) are obtained using Algorithm 1.

1.2 Mean Forecast Bias

Based on forecast estimates of Eq. (25), the expectation biases can be computed as follows.

$$\begin{aligned} \textbf{Bias}\left( n|h\right) =\frac{1}{p-h}\sum _{T=E}^{E+p-h}\left( \textbf{D}_{n,T+h}-\hat{\textbf{D}}_{n,T+h|T}\right) \end{aligned}$$
(26)

where \(\textbf{D}_{n,T+h}\) is actually observed values at time \(T+h\); \(\hat{\textbf{D}}_{n,T+h|T}\) denote h-steps forward expectations; p is the number of forecasting rounds, and n represents variables y, \(\pi \) or i, respectively.

The root mean squared error (RMSE) is computed as follows,

$$\begin{aligned} \textbf{RMSE}(n|h)=\sqrt{\frac{1}{p-h}\sum _{T=E}^{E+p-h}\left( \textbf{D}_{n,T+h}-\hat{\textbf{D}}_{n,T+h|T}\right) ^2} \end{aligned}$$
(27)

where E is the first forecast origin and p stands for the number of h-step forecasting rounds.

1.3 Log Predictive Scores

The approach of log predictive scores suggested by Good (1992) is popularly used in the literature. Scoring rules can be used to compare the ability of forecasts between a collection of models by assigning the predictive distribution a specified value that materializes at observations. Here, we report the differences between the log predictive scores computed from the time-varying DSGE model and those computed from the fixed-parameter DSGE model, and also report the differences between the log predictive scores from the TV-DSGE model and those from the TV-BVAR model. If relative log predictive scores are larger than zero, it means that the time-varying parameter DSGE model is comparatively better for prediction. The log predictive scores can be computed as the summation of log marginal predictive likelihood:

$$\begin{aligned} \textbf{S}(n|h)=\frac{1}{p-h}\sum _{T=E}^{E+p-h}\log p\left( \textbf{D}_{n,T+h}|\textbf{D}_{n,1:T}\right) \end{aligned}$$
(28)

1.4 Probability Integral Transform (PIT)

The log predictive scores rank predictive accuracies according to estimated likelihoods of the realized observables. As is well-known, a well-calibrated density forecast should satisfy two criteria: unbiasedness and efficiency. Equipped with this idea, we turn to another tool for comparing the density forecasts, that is, the probability integral transform, which can be used to assess the extent to which these two criteria are satisfied. According to Diebold et al. (2017), the probability integral transform is computed as the cumulative probability of the random variable \(\textbf{D}_{n,T+h}\) evaluated at the ex-post realizations of the variables collected in \(\textbf{D}_{n,T+h}\),

$$\begin{aligned} \textbf{PIT}_{n,h,T}=\int _{-\infty }^{\textbf{D}_{n,T+h}}p\left( \textbf{D}_{n,T+h}|\textbf{D}_{n,1:T}\right) \textbf{d}\textbf{D}_{n,T+h} \end{aligned}$$

The formula above is based on the predictive density function evaluated at the observed values until the \(T^{th}\) time point. In practice, the probability integral transform can be obtained by the Monte Carlo average approach as follows:

$$\begin{aligned} \textbf{PIT}_{n,h,T}\approx \frac{1}{n_{sim}}\sum _{j=1}^{n_{sim}}\textbf{I}\left( \textbf{D}_{n,T+h}^{(j)}\le \textbf{D}_{n,T+h}\right) \end{aligned}$$
(29)

Appendix D

We consider the following VAR(2) model with time-varying parameters

$$\begin{aligned} \begin{aligned} \textbf{Y}_t&=c_t + \sum _{j=1}^2B_{j,t}\textbf{Y}_{t-j}+\nu _t,\quad \textbf{VAR}(\nu _t)=R_t\\ \beta _t&=\{c_t,B_{1,t},B_{2,t}\} \\ \beta _t&=\beta _{t-1}+e_t,\quad \textbf{VAR}(e_t)=Q \end{aligned} \end{aligned}$$

The covariance matrix of the error term \(\nu _t\), i.e. \(R_t\) has time-varying elements. The observable vector \(\textbf{Y}_t=(y_t,\pi _t,i_t)\), every element being output growth, inflation and interest rate, respectively. For simplicity we consider the following structure for \(R_t\),

$$\begin{aligned} R_t=A_t^{-1}H_tA_t^{-1'} \end{aligned}$$

where \(A_t\) is a lower triangular matrix and \(H_t\) is a diagonal matrix. For a three variable VAR above,

$$\begin{aligned} A_t= \begin{bmatrix} 1 &{} 0 &{} 0 \\ \alpha _{12,t} &{} 1 &{} 0 \\ \alpha _{13,t}&{} \alpha _{23,t} &{} 1 \end{bmatrix},\quad H_t= \begin{bmatrix} h_{1,t} &{} 0 &{} 0 \\ 0 &{} h_{2,t} &{} 0 \\ 0&{} 0 &{} h_{3,t} \end{bmatrix} \end{aligned}$$

where

$$\begin{aligned} \alpha _{ij,t}=\alpha _{ij,t-1}+V_t,\quad \textbf{VAR}(V_t)=D \end{aligned}$$

and

$$\begin{aligned} \ln h_{i,t}=\ln h_{i,t-1}+z_{i,t},\quad \textbf{VAR}(z_{i,t})=g_i \end{aligned}$$

for \(i=1,2,3\).

The Gibbs and MH algorithm for estimating this three variable time-varying VAR model. The sample is the same as what has been used in estimating the New Keynesian model. We choose the first 40 quarterly data as warm-up sample to estimate a three- variable VAR(2) model. The estimates function as initial values when estimating the TV-BVAR model. We implement MH algorithm with a number of draws of 370,000, from which we drop the first about 10% draws.

Appendix E

The paper employs three U.S. quarterly time series: GDP growth rate \(100\Delta \log GDP_t\), quarterly growth rate of the urban CPI index \(100\Delta \log CPI_t\) and the Federal Funds rate \(100R_t/4\). To compare the monetary policy effects to the economy, it divides the sample over two subsample periods. The first subsample, 1948:Q1–1979:Q4, corresponds to the Great Inflation period. The second one, 1980:Q1–2019:Q4, corresponds to the Great Moderation period characterized by dramatically milder macroeconomic volatilities. The measurement equations relating the relevant elements of to the three observables are given by (9) in the main body. The raw data should be processed such that we estimate the model to match them.

Firstly, Growth rate of real output per capita.

$$\begin{aligned} YGR_t=100\left[ \ln \left( GDP_t\right) -\ln \left( GDP_{t-1}\right) \right] \end{aligned}$$

Secondly, Annualized inflation rate.

$$\begin{aligned} INFL_t=400\ln \left( \frac{CPI_t}{CPI_{t-1}}\right) \end{aligned}$$

Thirdly the Fed funds rate

$$\begin{aligned} INT_t=FFR_t/4 \end{aligned}$$

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

Wu, J., Wu, L. Bayesian Local Likelihood Estimation of Time-Varying DSGE Models: Allowing for Indeterminacy. Comput Econ (2023). https://doi.org/10.1007/s10614-023-10478-0

Download citation

  • Accepted:

  • Published:

  • DOI: https://doi.org/10.1007/s10614-023-10478-0

Keywords

JEL Classification

Navigation