Skip to main content

Interactions Between Fiscal and Monetary Authorities in a Three-Country New-Keynesian Model of a Monetary Union

  • Chapter
Book cover Dynamic Games in Economics

Abstract

In this paper we consider the effectiveness of various coordination arrangements between monetary and fiscal authorities within a monetary union if an economic shock has occurred. We address this problem using a multi-country New-Keynesian model of a monetary union cast in the framework of linear quadratic differential games. Using this model we study various coordination arrangements between fiscal and monetary players, including partial fiscal cooperation between only a subgroup of countries, which, to the best of our knowledge, has not been considered yet in the New-Keynesian literature. Using a simulation study we show that, in many cases and from the global point of view, partial fiscal cooperation between a subgroup of fiscal players is more efficient than non-coordination and that, in general, full cooperation without an appropriate transfer system is not a stable configuration. Furthermore, in case there is no full cooperation we show that the optimal configuration of the coordination structure depends on the type of shock that has occurred. We present a detailed analysis of the relationship between coordination structures and type of shock.

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

Access this chapter

Chapter
USD 29.95
Price excludes VAT (USA)
  • Available as PDF
  • Read on any device
  • Instant download
  • Own it forever
eBook
USD 84.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    See the next section for a selection of papers that analyse this issue in the spirit of NOEMs.

  2. 2.

    See next section for a short literature review on policy coordination in a monetary union.

  3. 3.

    It should also be noted that from the macroeconomic perspective the real cost of accessing a monetary union is reflected by the shadow cost of the abandonment of an own monetary policy so that it cannot be used as an adjustment tool in the case of an idiosyncratic shock. In other words, the cost of entering the union depends, to a large extent, on the effectiveness of the national central bank to tackle idiosyncratic shocks. It is especially important in the environment of closely integrated economies, like the EU, and in the case of small countries, such as Belgium.

  4. 4.

    See, for instance, Pierdzioch (2004) who presents a dynamic general equilibrium two-country optimising sticky-price model to analyse the consequences of international financial market integration for the propagation of asymmetric productivity shocks in the EMU.

  5. 5.

    The issue of fiscal transfers within the EMU has been studied by a number of authors such as Kletzer and von Hagen (2000), van Aarle et al. (2004), Evers (2006). The latter considers direct transfers among private sectors and indirect transfers among national fiscal authorities showing relative efficiency of such solutions.

  6. 6.

    Already Kenen (1969) emphasised the possible role which fiscal policies might play in a monetary union as potential chock-adjustment mechanisms.

  7. 7.

    One country attempts to improve its output-inflation trade-off by running a “beggar-thy-neighbour” policy. This is followed by the reaction of (an)other country(-ies) and the resulting non-cooperative outcome is a deflationary bias with all countries worse off with regard to a cooperative situation in which each country takes care of domestic inflation without attempting to affect the exchange rate (Cooper 1985).

  8. 8.

    It should be noted that prior to Beetsma et al. (2001), it has been observed in empirical studies (see e.g. Neck et al. 1999) that coordination does not necessarily lead to superior results as a result of either time-inconsistency and/or coalition formation of the fiscal policymakers against the monetary authority (see also next footnote). Furthermore, Beetsma and Bovenberg (1998) show that fiscal coordination may have a negative influence on a tax and public spending discipline, i.e. they may reduce the positive effects of monetary unification.

  9. 9.

    Rogoff (1985) already stated that there is a potential for a negative impact of coordination among a subset of actors (in this case the two fiscal authorities, leaving out the common central bank).

  10. 10.

    There are two countries in the model developed by Kirsanova et al. (2006) but this particular framework becomes computationally difficult when we add another, third country.

  11. 11.

    By F/i we denote the set of all countries except for country i.

  12. 12.

    See, among others, Fuhrer and Moore (1995), Galí and Gertler (1999), Woodford (2003), Lindé et al. (2004), Evans and McGough (2005) or Plasmans et al. (2006b).

  13. 13.

    For a similar formulation of the monetary policy rule in a model of a monetary union see van Aarle et al. (2004).

  14. 14.

    Since the seminal works of Kydland and Prescott (1977) and Barro and Gordon (1983), the quadratic loss functions are commonly used in the literature on strategic behaviour of fiscal and monetary authorities. See also Schellekens and Chadha (1999) for a more recent analysis supporting the quadratic form of the loss function.

  15. 15.

    Formula (10) applied to the 50-State US and to the 13-member state EMU would show that (ceteris paribus) the relative importance of interest rate volatility for the American economy is much lower than for the Euro-zone.

  16. 16.

    For further discussions see, for instance, Rogoff (1985), Persson and Tabellini (1993) and McCallum (1997).

  17. 17.

    This opinion was also expressed by Lars Svensson at the conference “Inflation Targeting, Central Bank Independence and Transparency,” 15–16 June 2007, Trinity College, Cambridge.

  18. 18.

    All (optimal) losses are multiplied by the factor 103.

  19. 19.

    For instance, C1’s loss \(\frac{1}{2}\int_{{ t}_{0}}^{\infty} \{\alpha_{{ i}} \hat{\pi}_{ i}^{2} (t)+\beta_{ i}\hat{y}_{ i}^{ 2}(t)+\chi_{ i}\widehat{\hat{f}}_{i}^{ 2} (t)\}e^{-\theta(t-t_{0} )} dt\) reported in the top of Table 3 is decomposed into \(\frac{1}{2}\int_{ t_{0}}^{ \infty } \{\alpha_{i}\hat{\pi}_{i}^{2}(t)\}e^{-\theta(t-t_{0})}dt\), \(\frac{1}{2}\int_{ t_{0}}^{\infty} \{\beta_{i}\hat{y}_{i}^{2}(t)\}e^{-\theta(t-t_{0} )} dt\) and \(\frac{1}{2}\int_{ t_{0}}^{\infty} \{\chi_{i}\widehat{\hat{f}}_{i}^{2}(t)\}e^{ -\theta(t-t_{0} )}dt\) in the lower part of the table.

  20. 20.

    More conventional preferences in which fiscal authorities care only two times as much about output gap as about inflation will be studied later on in this chapter.

  21. 21.

    The term “free-riding” refers here to taking advantage of others’ stabilisation policies during the stabilisation game (i.e. from time t 0, when the shock occurs, onwards). However, this term will be also used in the context of individual players breaking-up coalitional arrangements (with the same objective to take advantage of others’ cooperative stabilisation policies but themselves playing non-cooperatively and constraining own costly policies).

  22. 22.

    Deviations from a coalition are related to the coalition formation theory concept of internal stability (see Plasmans et al. 2006a, for further details).

  23. 23.

    Note that Beetsma et al. (2001) do not consider coordination in a grand coalition.

  24. 24.

    It might be argued that the above high cost of stabilisation is caused by the specific choice of policy rules which is so far away from optimum that players are forced to deviate much. In other words, it might be argued that \(\theta_{\pi}^{M}\) should be closer to 1 and \(\theta_{ \pi}^{ i}\) closer to 0. However, in other regimes, even in the fully non-cooperative regime, players are able to choose paths of stabilisation instruments close to assumed policy rules. This clearly overlures such an objection. Furthermore, the results reported in Table 3 were checked also for other parameterisations of policy rules such as: \((\theta_{ \pi}^{M} =1.5; \theta_{ \pi}^{ i} =0)\); \((\theta_{ \pi}^{M} =1; \theta_{ \pi}^{ i} =-0.5)\); \((\theta_{ \pi}^{ M}=1.25; \theta_{\pi}^{ i} =0)\); or \((\theta_{ \pi}^{ M}=1; \theta_{\pi}^{ i} =0)\) and produce similar results (under all the above assumptions).

  25. 25.

    The decrease in C3 control effort would be expected as the cooperation between C1 and C2, by increasing their activism, gives even more incentives to free-ride.

  26. 26.

    An example of such an analysis (albeit only in a two-country setting) can be found in Plasmans et al. (2006a, Chap. 3).

  27. 27.

    For the time being we focus on structural parameters of the model excluding policy rules, which together with preference parameters will be discussed in the next section.

  28. 28.

    Note that in the case of symmetric shock regimes 4, 5 and 6 denoted jointly by P are symmetric; thus, are characterised by the same social loss. However, for asymmetric shocks which always hit C1 only regimes 4 and 5 are symmetric to each other where as they are, in general, asymmetric to regime 6. Consequently, for asymmetric shocks we do not the joint P-notation for partial fiscal cooperation regimes.

  29. 29.

    In Table 8 the orderings based on J U (t 0) are in all cases but 8 exactly the same as those based on \(J_{ U}^{ \ast} (t_{0} )\), where the differences occurred only in the ordering of partial fiscal cooperation regimes 4, 5 and 6 under asymmetric shocks. Consequently, we report only orderings based on \(J_{U}^{ \ast} (t_{0})\).

  30. 30.

    Such an analysis goes far beyond the scope of this paper. It has been numerically checked that in the neighbourhood of \(\theta_{y}^{ F} =0.3\) for \(\theta_{ \pi}^{ M}=1.25\) social loss goes to the (nearly) infinite limiting value.

  31. 31.

    Naturally, there is problem with interpretation of the SGP as well as other issues related to the control variables caused by the (linear-) quadratic form of the loss functions. In reality, a negative deviation of fiscal debt from the rule, i.e. more restrictive budgetary policy, is not likely to be considered so “bad” or “undesirable” as the same positive deviation which, eventually, is going to increase public debt. It could be possible to partially take into account such issues also in the quadratic loss function but in the much complex model, which is far our of the scope of this paper.

  32. 32.

    A change in total loss from stabilisation effort caused by a change in the value of the relevant preference parameters can be decomposed into two effects. First is the change in the use of the stabilisation instrument as it becomes more/less expensive w.r.t. to other elements of the loss. Second change is directly caused by the increased/decreased cost.

  33. 33.

    If the absolute value of \(\theta_{y}^{ i}\) is too high, the counter-cyclical output gap stabilisation effort can be overshot, i.e. output gap can be (ceteris paribus) more volatile than for lower values of \(\theta_{ y}^{i}\), and would probably require additional pro-cyclical (and costly) control effort from fiscal authorities (see previous section for more details).

  34. 34.

    For DYNARE website with the most current version of the software see: www.dynare.org.

  35. 35.

    In case x 0 does not belong to S A , vector \(y= ( S_{S}^{ T}{ S}_{S} ) ^{-1} S_{ S}^{ T}x_{ 0}\) is such that the distance between S S and x 0 is minimal (y is the least-squares solution of x 0=S S y, i.e. \(\Vert { x}_{0}{ -S}_{S}{ y}\Vert { \leq }\Vert { x}_{0}{ -S}_{S} \tilde{y}\Vert \) for all \(\tilde{y}\)).

References

  • Amato, J. D., & Laubach, T. (2003). Rule-of-thumb behaviour and monetary policy. European Economic Review, 47, 791–831.

    Article  Google Scholar 

  • Ball, L. (1999). Policy rules for open economies. In J. B. Taylor (Ed.), Monetary policy rules. Chicago: University of Chicago Press.

    Google Scholar 

  • Barro, R. J., & Gordon, D. B. (1983). Rules, discretion and reputation in a model on monetary policy. Journal of Monetary Economics, 12, 101–121.

    Article  Google Scholar 

  • Başar, T., & Olsder, G. J. (1999). SIAM series in classics in applied mathematics. Dynamic noncooperative game theory. Philadelphia: Society for Industrial and Applied Mathematics.

    Google Scholar 

  • Batini, N., & Haldane, A. G. (1999). Forward-looking rules for monetary policy. In J. B. Taylor (Ed.), Monetary policy rules. Chicago: University of Chicago Press.

    Google Scholar 

  • Beetsma, R., & Bovenberg, L. (1998). Monetary union without fiscal coordination may discipline policymakers. Journal of International Economics, 45, 239–258.

    Article  Google Scholar 

  • Beetsma, R. M. W. J., & Jensen, H. (2004). Mark-up fluctuations and fiscal policy stabilisation in a monetary union. Journal of Macroeconomics, 26(2), 357–376.

    Article  Google Scholar 

  • Beetsma, R. M. W. J., & Jensen, H. (2005). Monetary and fiscal policy interactions in a micro-founded model of a monetary union. Journal of International Economics, 67(2), 320–352.

    Article  Google Scholar 

  • Beetsma, R. M. W. J., Debrun, X., & Klaassen, F. (2001). Is fiscal policy coordination in EMU desirable? Swedish Economic Policy Review, 8, 57–98.

    Google Scholar 

  • Benhabib, J., Schmidt-Grohe, S., & Uribe, M. (2001). Monetary policy and multiple equilibria. The American Economic Review, 91, 167–185.

    Article  Google Scholar 

  • Benigno, P., & Lopez-Salido, J. D. (2002). Inflation persistence and optimal monetary policy in the euro area (International Finance Discussion Paper 749). Board of Governors of the Federal Reserve System. https://www.federalreserve.gov/pubs/ifdp/2013/default.htm.

  • Buiter, W. H. (2004). Helicopter money: irredeemable fiat money and the liquidity trap (CEPR Discussion Paper 4202).

    Google Scholar 

  • Buti, M. (2001). Comment on Beetsma, Debrun and Klaassen: is fiscal policy coordination in EMU desirable? Swedish Economic Policy Review, 8, 99–105.

    Google Scholar 

  • Buti, M. & Sapir, A. (Eds.) (1998). Economic policy in EMU: a study by the European commission services. Oxford: Oxford University Press.

    Google Scholar 

  • Buti, M., Roeger, W., & In’t Veld, J. (2001). Stabilizing output and inflation: policy conflicts and co-operation under a stability pact. Journal of Common Market Studies, 39(5), 801–828.

    Article  Google Scholar 

  • Cavallari, L., & Di Gioacchino, D. (2005). Macroeconomic stabilisation in the EMU: rules versus institutions. Review of Development Economics, 9(2), 264–276.

    Article  Google Scholar 

  • Cecchetti, S. G., McConnell, M. M., & Perez-Quiros, G. (2002). Policymakers’ revealed preferences and the output-inflation variability trade-off: implications for the European system of central banks. Manchester School, 70(4), 596–618.

    Article  Google Scholar 

  • Cooper, R. (1985). Economic interdependence and coordination of economic policies. In R. Jones & P. Kenen (Eds.), Handbook of international economics 2. Amsterdam: North-Holland.

    Google Scholar 

  • Dixit, A., & Lambertini, L. (2001). Monetary-fiscal policy interactions and commitment. European Economic Review, 45, 977–987.

    Article  Google Scholar 

  • Engwerda, J. C. (2005). LQ dynamic optimization and differential games. New York: Wiley.

    Google Scholar 

  • Engwerda, J. C., van Aarle, B., & Plasmans, J. (2002). Cooperative and non-cooperative fiscal stabilisation policies in the EMU. Journal of Economic Dynamics & Control, 26, 451–481.

    Article  Google Scholar 

  • Equipe MIMOSA (1996). La nouvelle version de MIMOSA, modéle de l’économie mondiale. Revue de L’Observatoire Français Des Conjonctures économiques, 58, 103–155.

    Google Scholar 

  • European Commission (2001). European economy: Vol. 3. Public finances in EMU-01.

    Google Scholar 

  • European Commission (2002). European economy: Vol. 3. Public finances in EMU-02.

    Google Scholar 

  • Evans, G. W., & McGough, B. (2005). Monetary policy, indeterminacy and learning. Journal of Economic Dynamics & Control, 29(11), 1809–1840.

    Article  Google Scholar 

  • Evers, M. P. (2006). Federal fiscal transfers in monetary unions: a NOEM approach. International Tax and Public Finance, 13(4), 463–488.

    Article  Google Scholar 

  • Forlati, C. (2007). Optimal monetary policy in the EMU: does fiscal policy coordination matter? (Mimeo). Universitat Pompeu Fabra.

    Google Scholar 

  • Friedman, M. (1968). The role of monetary policy. The American Economic Review, 58, 1–17.

    Google Scholar 

  • Fuhrer, J., & Moore, G. (1995). Inflation persistence. The Quarterly Journal of Economics, 110(1), 127–159.

    Article  Google Scholar 

  • Gagnon, J., & Ihrig, J. (2002). Monetary policy and exchange rate pass-through (International Finance Discussion Paper 704). Board of Governors of the Federal Reserve System. http://www.federalreserve.gov/pubs/ifdp/2013/default.htm.

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

    Article  Google Scholar 

  • Galí, J., & Monacelli, T. (2005a). Monetary policy and exchange rate volatility in a small open economy. Review of Economic Studies, 72(3), 707–734.

    Article  Google Scholar 

  • Galí, J., & Monacelli, T. (2005b). Optimal monetary and fiscal policy in a currency union (Working Paper 300). Bocconi University, IGIER (Innocenzo Gasparini Institute for Economic Research).

    Google Scholar 

  • Hughes-Hallett, A., & Ma, Y. (1996). Changing partners: the importance of coordinating fiscal and monetary policies within a monetary union. Manchester School of Economic and Social Studies, 64, 115–134.

    Article  Google Scholar 

  • Juillard, M. (1996). Dynare: a program for the resolution and simulation of dynamic models with forward variables through the use of a relaxation algorithm (CEPREMAP Working Paper 9602). http://econpapers.repec.org/paper/cpmcepmap/9602.htm.

  • Kenen, P. (1969). The theory of optimum currency areas: an eclectic view. In R. Mundell & A. Swoboda (Eds.), Monetary problems in the international economy. Chicago: University of Chicago Press.

    Google Scholar 

  • Kirsanova, T., Vines, D., & Wren-Lewis, S. (2006). Fiscal policy and macroeconomic stability within a currency union (CEPR Discussion Paper 5584).

    Google Scholar 

  • Kirsanova, T., Satchi, M., Vines, D., & Wren-Lewis, S. (2007). Optimal fiscal policy rules in a monetary union. Journal of Money, Credit, and Banking, 39(7), 1759–1784.

    Article  Google Scholar 

  • Kletzer, K., & von Hagen, J. (2000). Monetary union and fiscal federalism (CEPR Discussion Paper 2615).

    Google Scholar 

  • Kwakernaak, H. (1976). Asymptotic root loci of multivariable linear optimal regulators. IEEE Transactions on Automatic Control, 3, 378–382.

    Article  Google Scholar 

  • Kydland, F. E., & Prescott, E. C. (1977). Rules rather than discretion: the inconsistency of optimal plans. Journal of Political Economy, 85(3), 473–491.

    Article  Google Scholar 

  • Lambertini, L., & Rovelli, R. (2003). Monetary and fiscal policy coordination and macroeconomic stabilisation. A theoretical analysis (Working Paper 464). Università di Bologna, Dipartimento Scienze Economiche. http://www2.dse.unibo.it/mysql/wp_dsa.php.

  • Leith, C., & Wren-Lewis, S. (2001). Interest rate feedback rules in an open economy with forward looking inflation. Oxford Bulletin of Economics and Statistics, 63(2), 209–231.

    Article  Google Scholar 

  • Lindé, J., Nessén, M., & Söderström, U. (2004). Monetary policy in an estimated open-economy model with imperfect pass-through (Working Paper Series 167). Sveriges Riksbank.

    Google Scholar 

  • Linnemann, L., & Schabert, A. (2002). Monetary and fiscal policy interactions when the budget deficit matters (Mimeo). University of Cologne.

    Google Scholar 

  • Lombardo, G., & Sutherland, A. (2004). Monetary and fiscal interactions in open economies. Journal of Macroeconomics, 26(2), 319–347.

    Article  Google Scholar 

  • Mankiw, N. (2001). The inexorable and mysterious tradeoff between inflation and unemployment. The Economic Journal, 111, 45–61.

    Article  Google Scholar 

  • McCallum, B., & Nelson, E. (1999). Nominal income targeting in an open-economy optimizing model. Journal of Monetary Economics, 43, 553–578.

    Article  Google Scholar 

  • McCallum, B. T. (1997). Crucial issues concerning central bank independence. Journal of Monetary Economics, 39, 99–112.

    Article  Google Scholar 

  • McCallum, B. T. (2001). Should monetary policy respond strongly to output gaps? The American Economic Review: Papers and Proceedings, 91(2), 258–262.

    Article  Google Scholar 

  • Mehra, Y. P. (2004). The output gap, expected future inflation and inflation dynamics: another look. The B.E. Journals in Macroeconomics, 4(1). doi:10.2202/1534-5998.1194.

  • Mélitz, J., & Zumer, F. (1998). Regional redistribution and stabilisation by the centre in Canada, France, the United Kingdom and the United States: new estimates based on panel data (CEPR Discussion Paper 1829).

    Google Scholar 

  • Michalak, T., Engwerda, J., & Plasmans, J. (2011). A numerical toolbox to solve N-player affine LQ open-loop differential games. Computational Economics, 37, 375–410.

    Article  Google Scholar 

  • Neck, R., Haber, G., & McKibbin, W. J. (1999). Macroeconomic policy design in the European monetary union: a numerical game approach. Empirica, 26, 319–333.

    Article  Google Scholar 

  • Persson, T., & Tabellini, G. (1993). Designing institutions for monetary stability. Carnegie-Rochester Conference Series on Public Policy, 39, 53–84.

    Article  Google Scholar 

  • Pierdzioch, Ch. (2004). Financial market integration and business cycle volatility in a monetary union (Kiel Working Paper 1115). Kiel Institute for the World Economy. http://www.ifw-kiel.de/publications/kap_e.

  • Plasmans, J., Engwerda, J. C., van Aarle, B., Di Bartolomeo, G., & Michalak, T. (2006a). Dynamic modelling of monetary and fiscal cooperation among nations. New York: Springer.

    Google Scholar 

  • Plasmans, J., Michalak, T., & Fornero, J. (2006b). Simulation, estimation and welfare implications of monetary policies in a 3-country NOEM model (NBB Working Paper 94). http://www.nbb.be/pub/06_00_00_00_00/06_03_00_00_00/06_03_05_00_00.htm.

  • Rogoff, K. (1985). Can international monetary policy coordination be counterproductive? Journal of International Economics, 18, 199–217.

    Article  Google Scholar 

  • Rotemberg, J. J., & Woodford, M. (1999). The cyclical behaviour of prices and costs (NBER Working Paper 6909). National Bureau of Economic Research.

    Google Scholar 

  • Schellekens, P., & Chadha, J. S. (1999). Monetary policy loss functions: two cheers for the quadratic (Cambridge Working Paper in Economics 9920). University of Cambridge, Faculty of Economics.

    Google Scholar 

  • Smets, F., & Wouters, R. (2002). An estimated stochastic dynamic general equilibrium model of the euro area (ECB Working Paper 171). http://www.ecb.europa.eu/pub/scientific/wps/date/html/index.en.html.

  • Svensson, L. (1997). Optimal inflation targets, “conservative” central banks, and linear inflation contracts. The American Economic Review, 87, 98–114.

    Google Scholar 

  • Taylor, J. (1993a). Discretion versus policy rules in practice. Carnegie-Rochester Conference Series on Public Policy, 39, 195–214.

    Article  Google Scholar 

  • Taylor, J. (1993b). Macroeconomic policy in a world economy: from econometric design to practical operation. New York: Norton.

    Google Scholar 

  • Taylor, J. (1995). The monetary transmission mechanism: an empirical framework. The Journal of Economic Perspectives, 9(4), 11–26.

    Article  Google Scholar 

  • Taylor, J. (2000). Reassessing discretionary fiscal policy. The Journal of Economic Perspectives, 14(3), 21–36.

    Article  Google Scholar 

  • Uhlig, H. (2003). One money, but many fiscal policies in Europe: what are the consequences? In M. Buti (Ed.), Monetary and fiscal policies in EMU: interactions and coordination. Cambridge: Cambridge University Press.

    Google Scholar 

  • van Aarle, B., Engwerda, J. C., & Plasmans, J. (2002a). Monetary and fiscal policy interaction in the EMU: a dynamic game approach. Annals of Operations Research, 109, 229–264.

    Article  Google Scholar 

  • van Aarle, B., Di Bartolomeo, G., Engwerda, J. C., & Plasmans, J. (2002b). Coalitions and dynamic interactions between fiscal and monetary authorities in the EMU. Ifo-Studien, 48(2), 207–229.

    Google Scholar 

  • van Aarle, B., Di Bartolomeo, G., Engwerda, J. C., & Plasmans, J. (2004). Policymakers’ coalitions and stabilisation policies in the EMU. Journal of Economics, 82(1), 1–24.

    Article  Google Scholar 

  • von Hagen, J. (2000). Fiscal policy and intranational risk sharing. In G. D. Hess & E. van Wincoop (Eds.), Intranational macroeconomics (pp. 272–294). Cambridge: Cambridge University Press.

    Google Scholar 

  • von Hagen, J., & Mundschenk, S. (2003). Fiscal and monetary policy coordination in EMU. International Journal of Finance and Economics, 8(4), 279–295.

    Article  Google Scholar 

  • Woodford, A. (2003). Interest and prices—foundations of a theory of monetary policy. Princeton: Princeton University Press.

    Google Scholar 

Download references

Acknowledgements

Tomasz Michalak was supported by the European Research Council, AG291528 (“RACE”) and FSF project 2003–2006.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Jacob Engwerda .

Editor information

Editors and Affiliations

Appendix

Appendix

1.1 A.1 Sensitivity Analysis with Respect to Structural Parameters

The assumed changes of one structural parameter at a time are presented in the first column of Table 5. The next columns show the social preference ordering over different regimes for four shocks.Footnote 27 The following notation is used: P—stands for regimes 4, 5 and 6; regime in bold means that players in a coalition are better off than in N; \(\hat{C}/\hat{F}\)—means that all fiscal players in grand/full fiscal coalition are better off than in N but the CB is not in this regime; \(\check{C}/\check{F}\)—vice versa; F—all players are better off than in N. In most of the cases the ordering based on \(J_{U}^{\ast}\) was the same as based on J U , with only few minor exception, therefore, only the preference ordering based on the former social loss is reported in Table 5.

Table 5 Sensitivity analysis of the benchmark model

For all combinations of parameters except for κ i,y/π =0.5, from the social point of view, regime N is preferred over F; in other words full fiscal coordination in counter-productive. Thus, it can be said that for the large set of parameters our model confirms results of Beetsma et al. (2001).

The ordering CPNF prevails for symmetric inflation shock \(v_{0S}^{\pi}\) and is robust to changes in parametrisation (except for lower value of κ i,y/π ).Footnote 28 The same ordering is valid for country-specific inflation shock \(v_{0A}^{\pi}\), however, due to asymmetry partial fiscal coalition 6 is (in general) characterised by different social loss than 4 and 5; thus, preference ordering over partial fiscal coalitions might vary. This leads to the conclusion that the grand coalition or partial fiscal coalitions should be sought as the socially efficient regimes in the case of inflation shock. The question whether such forms of coordination would be sustainable remains open. In Table 5 under \(v_{0S}^{\pi}\) and \(v_{0A}^{\pi}\) the grand coalition is preferred over non-cooperation by every player from an individual point of view. However, it does not tell us whether any player would like to deviate from this arrangement with hope that remaining fiscal players maintain cooperation and a partial fiscal regime emerges. In contrast, the fact that P-regimes are usually inferior to N for (fiscal) player(s) being in coalitions allows us to draw a conclusion that, certainly, these regimes are not sustainable in self-oriented (and myopic) environment.

The picture is less clear for output gap shocks as parameter changes have more influence on regimes’ social ordering here. Full fiscal cooperation is always least preferred (except for lower value of κ i,y/π ) so the results of Beetsma hold also in this case. However, in contrast to inflation shocks, the grand coalition scores often worse than non-cooperation or P-regimes in the case of symmetric output gap shock.

To summarise, sensitivity analysis of the benchmark model confirms the result of Beetsma et al. (2001) as in all the cases but one full fiscal coordination is worse than non-cooperation. Furthermore, this result can be extended further, as it comes out that F is the worst of all regimes, including partial fiscal cooperation. For inflation shocks, the grand coalition is the socially optimal outcome, and this regime is better than non-cooperation from the individual point of view. However, whether C is sustainable remains still an open question. Partial fiscal cooperation is suboptimal w.r.t. the grand coalition but gives better results than non-cooperation from the social perspective. Unfortunately, in many cases these regimes are suboptimal from individual point of view, thus, possibly unsustainable in the non-cooperative environment, especially if players are myopic. The ineffectiveness results about full fiscal coordination hold also for output gap shocks, but it more difficult to draw some definite conclusions about the preference ordering over the other regimes as the differences in social loss between them are small and, therefore, sensitive, to changes in parametrisation.

It is apparent from Table 5 that the influence of forward-lookingness on our model calls for more attention, as lower values of this parameter may have an important impact on the result obtained above. Table 6 shows the optimal losses for symmetric inflation shock \(v_{0S}^{\pi}\) and benchmark parametrisation but with κ i,y/π =0.5. What is the reason for the improved effectiveness of F regime w.r.t. N when the economies in a monetary union are characterised by lower forward-lookingness?

Table 6 Optimal losses for (\(v_{0S}^{\pi}\), κ i,y =0.5)
Table 7 Symmetric inflation shock, the number of equilibria in LQDGs

Change of an important model parameter certainly influenced the reduced form matrices \(\tilde{B}_{4}\) and \(\tilde{B}_{5}\) which show the influence of control instruments on output gaps and inflations, respectively.

$$\begin{aligned} \tilde{B}_{4} =&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} 0.7007 & 0.1468 & 0.1468 & -0.6631 \\ 0.1468 & 0.7007 & 0.1468 & -0.6631 \\ 0.1468 & 0.1468 & 0.7007 & -0.6631\end{array} \right ], \quad \mbox{and} \\ \tilde{B}_{5} =&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} 0.0134 & 0.0026 & 0.0026 & -0.0125 \\ 0.0026 & 0.0134 & 0.0026 & -0.0125 \\ 0.0026 & 0.0026 & 0.0134 & -0.0125\end{array} \right ]. \end{aligned}$$

The pattern is closely comparable to the benchmark (all the values are now more or less 15 % lower than before and all the signs are preserved). With a bigger backward-looking component, economies are more persistent and converge slower to the equilibrium. This is reflected by nearly two times as more losses in regimes N, C and P for κ i,y/π =0.5 compared to benchmark with \(\kappa_{i,y/\pi}=\frac{2}{3}\) in Table 3. As far as the cost of the control instruments is concerned, we observe the following pattern. For regimes where only fiscal players fully or partially cooperate (i.e. F and P), (total) cost of control of those players involved in the coalition gets lower when backward-looking component becomes more eminent. In the case of full fiscal cooperation. This reduction is drastic, from 7.2424 to 0.1974. In contrast, the loss of the fiscal players from the control effort (in P regimes) increases more than 10 times, from 0.0336 to 0.3682. The same holds for the non-cooperative regime, where the (non-cooperative) fiscal players have their control effort increased from 0.0285 to 0.6618. In spite of the somewhat increased volatility of inflation and output gap, the above decrease in control cost in the F regime accompanied by the increase in control cost in N regime (see Fig. 3) makes full fiscal cooperation more attractive than previously.

Fig. 3
figure 3

Benchmark vs. model with lower forward-lookingness, \(v_{0S}^{\pi}\), regime N

The reason for these more moderate actions should be sought in the fact that more backward-lookingness in the model means that economies are more persistent and more slowly converging to equilibrium. This means that it is more costly to control them as the use of instruments is less efficient, which successfully diminishes the conflict between authorities, which choose more reasonable policies.

To improve our understanding of the meaning of forward-lookingness in the model similar sensitivity check to the previous one was conducted but now assuming that κ i,y/π is set to 0.5 as the benchmark. The results are presented in Table 8 which was constructed in the same way as Table 5.Footnote 29

Table 8 Sensitivity analysis, benchmark model but with lower forward-lookingness

In most of the cases, \(J_{U}^{\ast}(t_{0},F)\) is preferred over \(J_{U}^{\ast}(t_{0},N)\) which confirms that forward-lookingness is decisive for the social profitability of full fiscal cooperation in our model. Similarly to Table 6 the social ordering is especially robust for both symmetric and asymmetric inflation shocks. In all these cases full fiscal coordination is the second best regime just after the grand coalition. Regimes P with smaller fiscal coalitions score worse, where as the worst result is obtained for non-cooperation. Consequently, in contrast to previous findings and Beetsma et al. (2001), the above results strongly advocate the need for coordination in the case of inflation shocks. Furthermore, any form of coordination is better than non-cooperation, not only from the social point of view, but also from the perspective of individual authorities.

The results are a little less clear for output gap shocks as, for a few parameter combinations, non-cooperation comes back to the position it took in Table 5, i.e. just after C but before N. It happens when the CB gets more powerful w.r.t. fiscal players in its influence on the output gap (either higher values of γ i or lower value of η i ).

1.2 A.2 Sensitivity Analysis w.r.t. Preference Parameters

Thus far, our sensitivity analysis was performed only w.r.t. structural parameters of the economies (excluding policy rules). It is interesting to take a closer look into preference parameters in players’ loss functions. In the benchmark, fiscal players cared five times more about development of output gaps than inflation whereas the CB cared five times more about inflation than output gap. We argued that such a preferences were in line with other parameters in the model as they guarantee that no variables are overrepresented in the total loss of players (as decompositions in previous sections confirm). In this section we vary relative preferences of fiscal and monetary authorities regarding output gap and inflation. In particular, let \(r_{F}^{\pi/y}\) and \(r_{CB}^{\pi/y}\) denote the ratio between α i and β i and between β CB,i and α CB,i , respectively, i.e. \(r_{F}^{\pi/y}:=\frac{\alpha_{i}}{\beta_{i}}\) and \(r_{CB}^{\pi/y}:=\frac{\beta_{CB,i}}{\alpha_{CB,i}}\). The sensitivity analysis will be performed for \(r_{F}^{\pi/y}\) and \(r_{CB}^{\pi/y}\) simultaneously changing from 0 to 1 with step 0.1. For \(r_{F}^{\pi /y}=r_{CB}^{\pi/y}=0\) we have a situation where governments are concerned only about output gaps (i.e. are very liberal in stabilisation sense) while the CB is concerned only about inflation (i.e. is very conservative). In other words, this is the situation when preferences are totally opposite. In contrast, when \(r_{F}^{\pi/y}=r_{CB}^{\pi/y}=1\) fiscal authorities as well as the monetary ones are equally interested in deviations of both variables, which, taking into account that the weight of the control instrument does not change and is equal between all the players, means that under symmetric shocks their objectives are the same in this extreme case. The results of the sensitivity check of the benchmark model with preferences amended in the above way are presented in Table 9. The ordering in this table, similarly to previous tables, is based on \(J_{U}^{\ast}(t_{0})\).

Table 9 Sensitivity analysis, benchmark model with altered preference

It is evident from Table 9 that the grand coalition is the socially optimal regime for lower values of \(r_{F/CB}^{\pi/y}\), when preferences of various authorities are opposite. The second best choice are partial fiscal cooperation regimes, where as fiscal cooperation scores worst, even worse than non-cooperative regime N. This pattern is observed in the neighbourhood of the benchmark for all shocks except for \(v_{0S}^{y}\). In contrast, when \(r_{F/CB}^{\pi/y}\) becomes larger, first partial fiscal cooperation becomes more socially profitable than C, then F more profitable than N and, finally, when preferences of governments and the CB coincide, F becomes the most profitable outcome of all. This last result is interesting as previously, in the majority of situations, C was the most socially desirable outcome. However, when \(r_{F/CB}^{\pi /y}\approx r_{F/CB}^{\pi/y}\) this regime is not so efficient any more because under equal bargaining power assumption the loss of the CB is under-represented in the joint loss of the grand coalition. This leads to the situation in which the interest rate is less important in the joint loss than fiscal debts of individual countries and, therefore, is used more extensively than under F, where free-riding between fiscal coalition and the CB prevents both groups from an overuse of their control instruments. It is easily visible in Table 10 which shows the decomposed players losses for symmetric price shock and symmetric preferences, i.e. \(r_{F}^{\pi /y}=r_{CB}^{\pi/y}=1\). Loss from \(\chi_{F,C3}\widehat{\hat{f}}^{2}_{C3}\) under F is bigger than under C and, at the same time, \(\chi_{CB}\widehat{\hat{\imath}}_{U}^{2}\) is lower under F than under C as fiscal players in F cannot rely so much as in C on interest rate to stabilise economies due to free-riding of the CB. Since, in social loss based on \(J_{U}^{\ast}\) interest rate has relatively bigger share than in the joint loss of the grand coalition, full fiscal coordination scores better than the grand coalition where interest rate is relatively overused. Another important observation from Table 9 is that partial fiscal cooperation, as in most of the cases analysed before, is very often the second best choice.

Table 10 Optimal losses for \(v_{0S}^{\pi}\), \(\alpha_{F,i}=\alpha _{CB}=\frac{1}{2}\alpha_{CB}=\frac{1}{2}\alpha_{F,i}\)

Next to every column with social preference ordering we show average social loss obtained for different levels of \(r_{F/CB}^{\pi/y}\). Obviously, the less conflicting preferences are the lower average common loss suffered by the union is. Thus, the percentage difference between the average losses for \(r_{F/CB}^{\pi/y}=0\) and \(r_{F/CB}^{\pi/y}=1\) is the highest for symmetric and asymmetric inflation shocks (61.5 % and 59 %, respectively), and much more moderate for both output shock (13.5 % and around 0), which confirms are previous results that the biggest gains from choosing an appropriate regime is to be expected in the former case.

A number of interesting conclusions can be drawn the above analysis. First of all, the relative antagonism between the CB and governments in the monetary union is an important factor which strongly determines the profitability of full fiscal coordination. In contrast to other various findings from the literature, in our model, strongly independent bank of a monetary union is not so profitable from the common perspective and more intermediate arrangements are advisable. Secondly, if bargaining power in the grand coalition do not coincide with socially optimal preferences, this regime might be counter-productive w.r.t. full-fiscal coalition, which in turn, can turn out to be optimal due to free-riding.

The analyses in Table 9 is made under the assumption that both types of authorities simultaneously change their preferences from the most conflicting to the same ones. This was rather theoretical simulation as having little chances to be realised in (the European) practice as the ECB independence is strongly safeguarded by relevant treaties. It is also interesting to study the more realistic situation in which the strong CB’s focus on inflation remains unchanged while governments, at the beginning fixed only at inflation, become gradually interested in inflation. More formally, we consider the case where \(r_{CB}^{\pi/y}\) is kept constant at 0 where as \(r_{F}^{\pi/y}\) changes from 0 to 1. One possible interpretation of such simulations in Table 11, which one can think of, are more and more stringent provisions of the SGP, which additionally to fiscal debt issues regulates also inflation in the EMU Member States.

Table 11 Sensitivity analysis, benchmark model with altered preference

In general, the outcomes in Table 11 are reasonably similar to those from Table 9 as far as main trends are considered, i.e. the less intensive conflict between authorities makes partial fiscal cooperation regimes as well as full fiscal cooperation more interesting from the social point of view. Of course, always restrictive CB makes it impossible to reach the same outcome as in the previous case. For \(r_{CB}^{\pi/y}=0\) and \(r_{F}^{\pi/y}=1\) (i.e. the last row of Table 11) the social orderings are similar \(r_{F/CB}^{\pi /y}=0.5\) previously (the middle of the Table 9). Accordingly, minimal average loss obtained for the last case is now higher than when authorities’ preferences were more alike. However, what is important, social losses at the end of both tables are not much different which shows that similar low social welfare can be obtained either by making preferences of fiscal and monetary authorities more parallel, or by safeguarding the CB independence and making government to be more equally oriented about inflations and output gaps. The first proposition seems to be rather unacceptable by the modern economic school, but the second one seems not only to be acceptable from this point of view, but actually implemented in the current European practice (in the form of the strongly independent ECB and the SGP, which makes governments more “inflation-aware”).

The issue related to the SGP will be discussed further in this paper but first we will consider (nearly) optimal policy rules.

1.3 A.3 Nearly Optimal Policy Rules

In the LQDG framework it is not possible to analytically optimise certain parameters of the model, however, an approximate analyses can be performed numerically. We will use this method to study how various combinations of policy rules’ parameters influence output gap and inflation volatility and which of them are likely to bring (nearly) optimal outcome from the social point of view. Due to the space constraints we will focus mainly on the symmetric inflation shock. Tables 12, 13, 14 show (optimal) losses together with their decomposition in the non-cooperative regime for different values of \(\theta_{y}^{i}\) and \(\theta _{\pi}^{U}\). More specifically, Table 12 shows cases in which \(\theta_{\pi}^{U}=1.25\) and \(\theta_{y}^{i}\) changes from 0 to −1; Table 13 cases in which \(\theta_{\pi }^{U}=1.5\) and \(\theta_{y}^{i}\) as before; and, finally, Table 14 shows cases in which \(\theta_{\pi}^{U}=1.75\) and as before. Such an analysis, albeit approximate, may give us an important insight into efficiency of different policy rules’ combinations.

Table 12 Optimal losses for \(\theta^{M}_{\pi}=1.25\) and \(\theta^{F}_{y}=0, 0.1,0.3, 0.5, 0.8, 1.0\)
Table 13 Optimal losses for \(\theta^{M}_{\pi}=1.50\) and \(\theta^{F}_{y}=0, 0.1, 0.3, 0.5, 0.8, 1.0\)
Table 14 Optimal losses for \(\theta^{M}_{\pi}=1.75\) and \(\theta^{F}_{y}=0, 0.1, 0.3, 0.5, 0.8, 1.0\)

Figure 4 compares \(J_{U}^{\ast}\) losses for different combinations of \(\theta_{\pi}^{U}\) and \(\theta_{y}^{i}\). In general, from the monetary authority perspective, a rule less focused on inflation (i.e. \(\theta_{\pi}^{U}=1.25\), Table 12) results in higher losses than for the benchmark value (i.e. \(\theta _{\pi }^{U}=1.50\), Table 13), whereas a rule more focused on inflation (i.e. \(\theta_{\pi}^{U}=1.75\), Table 14) generates lower losses. The only exception from this pattern is a combination of coefficients \(\theta_{\pi}^{U}=1.50\) and \(\theta_{y}^{i}=0\) which produces the lowest social loss, i.e. is an optimal Taylor rules parameters’ combination (ceteris paribus) for the non-cooperative regime. From the fiscal authority perspective the stronger reaction to output, the higher loss and vice versa. Finally, it should be mentioned that for combinations \((\theta_{\pi}^{U}=1.25, \theta_{y}^{i}=-0.3)\) and \((\theta_{\pi}^{U}=1.25, \theta _{y}^{i}=-0.5) \) strong irregularities emerge, explanation of which should be sought in mathematical properties of the model.Footnote 30

Fig. 4
figure 4

Social loss for various combinations of policy rules parameters, regime N

Let us now focus on the individual players’ perspective. Due to the irregularities mentioned above we will exclude from our analysis Table 12. For \(\theta_{\pi}^{U}=1.50\), the loss of the fiscal players is not monotonic and reaches its minimum for the benchmark parametrisation (i.e. \(\theta_{y}^{i}=0.50\)). However, for \(\theta _{\pi}^{U}=1.75\) the results are more clear-cut as the stronger reaction of governments to output always leads to the lower loss. This is totally at odds with the CB losses which behave in the exactly opposite way. The reasons of this difference can be found in the decomposition of losses. Stronger reaction of fiscal debt to output gap leads to its lower volatility (in Table 14 \(\beta_{F,C1}\hat{y}_{C1}^{2}\), and, consequently, \(\beta_{CB}\hat{y}_{CB}^{2}\) decrease with \(\theta _{y}^{i}\)), however, this positive effect is reached at the very expense of inflation volatility which grows accordingly. This is detrimental for the CB as this authority is mainly concerned about inflation under benchmark parametrisation. Overall, the conflict of interest between both types of authorities is clearly visible here. Highest value of \(\theta_{\pi}^{U}\) without a counter-response in fiscal rule is damaging to loss of governments as the CB’s strong reaction to inflation makes output gap very volatile. Thus, governments use the most of their control effort to improve the situation, however, only higher (absolute) values of \(\theta_{y}^{i}\) make them increasingly better off. This pattern is robust even for \(\theta _{y}^{i}\) reaching minus one. In contrast, as mentioned above, for the more moderate CB’s policy rule (i.e. \(\theta_{\pi}^{U}=1.5\)) and for \(\theta_{y}^{i}\) higher than half, fiscal loss start to increase. This means that, if fiscal rule responses to tempered monetary rule too strongly, there is an effect of overshooting the fiscal policy rule. As a result, governments are pushed to deviate from such a rule much stronger than before because the (overshot) rule must be discretionary corrected. Consequently, \(\chi_{F,C1}\widehat{\hat{f}}^{2}_{C1}\) grows from 0.0269 in benchmark to 0.6282 for \(\theta_{y}^{i}=-1\). Comparing the CB’s losses between Tables 12, 13, 14 it is evident that more reactionary stance is in the interest of the CB as its loss decreases with increasing \(\theta_{\pi}^{U}\) due to the lower inflation deviation in the union.

To sum up, from the individual point of view, we have a situation where CB has incentives to increase \(\theta_{\pi}^{U}\) and, at the same time, for high values of \(\theta_{\pi}^{U}\), governments have incentives to increase \(\theta_{y}^{i}\). When both types of authorities do it at the same time, the economies end up in a position which is not only suboptimal from the social point of view (right-down corner in Fig. 4), but also from individual ones. For \(\theta_{\pi}^{U}=1.75\) and \(\theta _{y}^{i}=-1.0\) governments obtain loss of 2.4264 and the CB of 6.5807, which is the outcome Pareto-dominated by other combinations, e.g. the benchmark. Unfortunately, the socially optimal combination \((\theta _{\pi}^{U}=1.5, \theta_{y}^{i}=0)\) does not Pareto-dominate combination \((\theta_{\pi}^{U}=1.75, \theta_{y}^{i}=-1.0)\) so the mutual agreement to move toward the optimum seems unlikely to be obtained.

1.4 A.4 SGP Analysis

Within our model we can also investigate the effects of the major policy-surveillance institution of the EMU, namely the SGP. The SGP imposes a framework of fiscal stringency and coordination measures that aim at securing the implementation of the BEPGs. In our model the effects of various levels of the SGP stringency can be studied by considering (i) different levels of the countercyclical parameter \(\theta_{y}^{i}\) in the fiscal rule; and (ii) different weights associated with the domestic fiscal deficit, χ i , in the objective functions of the fiscal players. We compare the following three cases, each characterised by stricter SGP provisions than the other:Footnote 31

  1. I.

    In the fiscal rule the coefficient measuring a countercyclical reaction of fiscal debt to deviation of output gap is two times smaller than in the benchmark, i.e.: \(\theta_{y}^{i,new}=-0.25\);

  2. II.

    As above, but, additionally, deviations from the rule are more costly (i.e. are more severely punished by the SGP provisions), \(\chi_{i}^{new,II}=1.5\chi_{i}\);

  3. III.

    As in point I, but, additionally, \(\chi_{i}^{new,III}=3\chi _{i}\).

It is expected that smaller countercyclical reaction of the fiscal rule is going to force fiscal authorities to deviate stronger from the rule than in the benchmark. On the other hand, more costly deviations from the rule in cases II and III are likely to diminish the use of fiscal instrument w.r.t. case I. As far as individual losses of players are concerned, it is possible to directly compare new cases to the benchmark, however, it is not exactly obvious whether we can do so with the social loss. Whereas governments, as public authorities, might be bound by tougher SGP provisions, it does not have to lead to an automatic increase of the social loss. In the benchmark we assumed that χ U =χ i . Now, we are going to compute “adjusted” social loss of the entire union \(J_{U}^{\ast}(t_{0})\), denoted \(J_{U}^{\ast A}(t_{0})\), by assuming that cost of the deviation of the fiscal instrument from the rule is unchanged, i.e. equals to χ U =χ i as in the benchmark, instead of \(\chi_{U}=\chi _{i}^{new,II} \) or \(\chi_{U}=\chi_{i}^{new,III}\). By doing so we will see what is the contribution of a change in use of a fiscal instrument in the total change of the loss from stabilisation effort.Footnote 32

Players’ losses in the first three regimes in the case of a symmetric inflation shock and benchmark parametrisation are shown in Tables 15 and 16. In spite of vast differences between cases a few general conclusions can be drawn. First of all, lower counter-cyclical reaction of fiscal debt (case I) always makes the losses of fiscal players higher than in the benchmark, which means that the assumed value \(\theta _{y}^{i}=-0.5\) was chosen relatively well for the initial simulations and which confirms our findings from the previous section.Footnote 33 Secondly, in all the new cases higher SGP stringency leads to increasing losses of fiscal players from output gap volatility. This is natural as fiscal authorities refrain from using the more expensive control instruments. Interestingly, in different regimes we obtain different relationships between SGP stringency and the amount of the control instrument used. Whereas in all regimes with any form of cooperation (i.e. C, F, and P-regimes) the higher χ i , the less control instrument is used (compare cases II to I and III to II), then under non-cooperation this relationship is highly non-linear. For \(\chi _{i}^{new,II}\) governments decide to use \(\frac{\chi_{C1}\widehat{\hat {f}}^{2}_{i}}{\chi_{C1}}=\widehat{\hat{f}}^{2}_{i}=1190\) which is nearly 40 % more (not less) than for χ i in case I, however when χ i increases to \(\chi_{i}^{new,III}\) they contract substantially control action to \(\frac{\chi_{C1}\widehat{\hat{f}}^{2}_{i}}{\chi_{C1}}= \widehat{\hat{f}}^{2}_{i}=300\). The SGP regulating the use of control instrument influences also the use of interest rate by the CB of the union. In many cases when control action of fiscal authorities is diminished the response of the CB also fades out, i.e. conflict between both types of authorities is hampered. In relative terms, the biggest reductions in the control effort of the monetary authority is obtained under F III where cost of the control effort is lowered from 25.83 to 3.00. On the other hand, under N III we also witness quite a reduction in the fiscal control effort w.r.t. N I, but the main driving force in this case is a free-riding of fiscal players, which forces the CB to increase its engagement in the union economy not stabilised enough by national governments. As the loss from the CB’s control instrument is an important part of \(J_{U}^{\ast}(t_{0})\) and \(J_{U}^{A\ast}(t_{0})\), this leads to higher social loss under N III than under F III.

Table 15 Regimes N, C and F for different levels of the SGP stringency
Table 16 Regimes N, C and F for different levels of the SGP stringency

To summarise, we established the third factor (next to the degree of backward-lookingness and loss functions’ preferences) which heavily determined the results obtained for the benchmark parametrisation of the model. The increased SGP stringency reduces incentives of fiscal players to use control instruments, therefore, in situations where high social losses where driven by the conflict between authorities (notably regime F), such a firmer stance is beneficial to the union-wide economic interest. However, in situations in which free-riding is present (notably regime N under benchmark) increased SGP stringency may lead to more extensive free-riding of governments as undertaking any actions become more costly. This, in turn, makes the CB to intervene and increases social loss of the union. In other words, the stringent SGP has both positive and negative effects in the context of this paper and is able to make unprofitable regime to become profitable.

Similar analysis has been performed for 3 other shocks. Since the conflict under \(v_{0A}^{\pi}\) is less eminent also the social gains from higher SGP stringency are lower. As before, both output shocks are characterised by the lower variability of losses between different regimes, however, still SGP stringency is able to make non-cooperation inferior to fiscal cooperation, at least, in the case of the symmetric shock.

1.5 A.5 Model Derivations

1.5.1 A.5.1 Reduced Form of the Model

Defining

$$\begin{aligned} K_{y} :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \kappa_{1,y} & 0 & \cdots& 0 \\ 0 & \kappa_{2,y} & \cdots& 0 \\ \cdots& \cdots& \cdots& \cdots\\ 0 & 0 & \cdots& \kappa_{n,y}\end{array} \right ] , \qquad K_{\pi}:=\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \kappa_{1,\pi} & 0 & \cdots& 0 \\ 0 & \kappa_{2,\pi} & \cdots& 0 \\ \cdots& \cdots& \cdots& \cdots\\ 0 & 0 & \cdots& \kappa_{n,\pi}\end{array} \right ] , \\ G :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \gamma_{1} & 0 & \cdots& 0 \\ 0 & \gamma_{2} & \cdots& 0 \\ \cdots& \cdots& \cdots& \cdots\\ 0 & 0 & \cdots& \gamma_{n}\end{array} \right ] , \qquad E:=\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \eta_{1} & 0 & \cdots& 0 \\ 0 & \eta_{2} & \cdots& 0 \\ \cdots& \cdots& \cdots& \cdots\\ 0 & 0 & \cdots& \eta_{n}\end{array} \right ] , \\ \varXi :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \xi_{1} & 0 & \cdots& 0 \\ 0 & \xi_{2} & \cdots& 0 \\ \cdots& \cdots& \cdots& \cdots\\ 0 & 0 & \cdots& \xi_{n}\end{array} \right ] , \qquad B:=\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \beta_{1} & 0 & \cdots& 0 \\ 0 & \beta_{2} & \cdots& 0 \\ \cdots& \cdots& \cdots& \cdots\\ 0 & 0 & \cdots& \beta_{n}\end{array} \right ] , \\ R :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} 0 & \rho_{12} & \cdots& \rho_{1n} \\ \rho_{21} & 0 & \cdots& \rho_{2n} \\ \cdots& \cdots& \cdots& \cdots\\ \rho_{n1} & \rho_{n2} & \cdots& 0\end{array} \right ] , \\ D :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \delta_{12} & \delta_{13} & \cdots& \delta_{1n} \\ -\sum_{j\in F/2}\delta_{2j} & \delta_{23} & \cdots& \delta_{2n} \\ \cdots& \cdots& \cdots& \cdots\\ \delta_{n2} & \delta_{n3} & \cdots& -\sum_{j\in F/n}\delta_{nj}\end{array} \right ] , \\ \varPsi_{y} :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \psi_{1,y} & 0 & \cdots& 0 \\ 0 & \psi_{2,y} & \cdots& 0 \\ \cdots& \cdots& \cdots& \cdots\\ 0 & 0 & \cdots& \psi_{n,y}\end{array} \right ] \\ V :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \varsigma_{12} & \varsigma_{13} & \cdots& \varsigma_{1n} \\ -\sum_{j\in F/2}\varsigma_{2j} & \varsigma_{23} & \cdots& \varsigma _{2n} \\ \cdots& \cdots& \cdots& \cdots\\ \varsigma_{n2} & \varsigma_{n3} & \cdots& -\sum_{j\in F/n}\varsigma _{nj}\end{array} \right ] , \\ \varPsi_{\pi} :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \psi_{1,y} & 0 & \cdots& 0 \\ 0 & \psi_{2,y} & \cdots& 0 \\ \cdots& \cdots& \cdots& \cdots\\ 0 & 0 & \cdots& \psi_{n,y}\end{array} \right ] \\ \varPsi :=&\left [ \begin{array}{@{}c@{\quad}c@{}} \varPsi_{y} & 0 \\ 0 & \varPsi_{\pi} \end{array} \right ] , \quad \mbox{and} \quad v_{t}:=\left [ \begin{array}{@{}c@{}} v_{t}^{y} \\ v_{t}^{\pi}\end{array} \right ] , \\ \iota_{n} :=&\left [ \begin{array}{@{}c@{}} 1 \\ 1 \\ 1 \\ \cdots\\ 1\end{array} \right ] _{n}, \qquad S:=\left [ \begin{array}{@{}c@{\quad}c@{}} -\iota_{ ( n-1 ) } & I_{n-1} \end{array} \right ] , \\ \varTheta_{\pi}^{F} :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \theta_{\pi}^{1} & \theta_{\pi}^{2} & \cdots& \theta_{\pi}^{n}\end{array} \right ] ^{T} \quad\mbox{and}\quad \varTheta_{y}^{F}:= \left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{}} \theta_{y}^{1} & \theta_{y}^{2} & \cdots& \theta_{y}^{n}\end{array} \right ] ^{T} \end{aligned}$$

the SNKM model can be rewritten as:

$$\begin{aligned} y_{t} = {}& K_{y}E_{t}y_{t+1}+ ( I_{n}-K_{y} ) y_{t-1}-G ( \mathbf{\iota}_{n}i_{t}-E\mathbf{\pi}_{t+1} ) +Ef_{t} \\ &{}- K_{y}RE_{t}y_{t+1}+Ry_{t}- ( I_{n}-K_{y} ) Ry_{t-1}-K_{y}DE_{t}s_{t+1} \\ &{}+ Ds_{t}- ( I_{n}-K_{y} ) Ds_{t-1}+v_{t}^{y}, \end{aligned}$$
(12)
$$\begin{aligned} \pi_{t}={} & K_{\pi}BE_{t} \pi_{t+1}+ ( I_{n}-K_{\pi} ) B\pi _{t-1}+ \varXi y_{t}+\varXi Vs_{t}+v_{t}^{\pi}, \end{aligned}$$
(13)
$$\begin{aligned} s_{t+1}={} & s_{t}+S\pi_{t+1}, \end{aligned}$$
(14)
$$\begin{aligned} v_{t+1}={} & \varPsi v_{t}+\varepsilon _{t+1}, \end{aligned}$$
(15)

where I m is m×m identity matrix (m=n−1,n), s t :=[s 12,t s 1n,t ] and y t , π t , \(v_{t}^{y}\) and \(v_{t}^{\pi}\) are appropriately defined vectors of size n each. In particular, it can be shown that every s ij,t :=p j,t p i,t (ji) can be expressed in terms of s 12,t ,…,s 1n,t . For example, in a three-country monetary union we have six bilateral real exchange rates: s 12,t =p 2,t p 1,t , s 13,t =p 3,t p 1,t , s 21,t =p 1,t p 2,t , s 23,t =p 3,t p 2,t , s 31,t =p 1,t p 3,t , and s 32,t =p 2,t p 3,t . Clearly, the last four variables can be expressed as a combination of the first two, i.e. s 21,t =−s 12,t , s 23,t =s 13,t s 12,t , s 31,t =−s 13,t and s 32,t =s 12,t s 13,t .

Defining fiscal and monetary policy rule vectors as:

$$\begin{aligned} f_{t} &:=\varTheta_{\pi}^{F} \pi_{t}+\varTheta_{y}^{F}y_{t},\quad \mbox{ and} \end{aligned}$$
(16)
$$\begin{aligned} i_{t} &:=\theta_{\pi}^{U} \omega^{T}\pi_{t}+\theta_{y}^{U}\omega ^{T}y_{t}, \end{aligned}$$
(17)

substituting them into system (12)–(15) and rearranging we get:

$$\begin{aligned} &{-}K_{y} ( I_{n}-R ) E_{t}y_{t+1}-GE_{t} \pi_{t+1}+K_{y}DE_{t}s_{t+1} \\ &\quad=- \bigl( I_{n}-E\varTheta_{y}^{F}+G \iota_{n}\theta_{y}^{U}\omega ^{T}-R \bigr) y_{t}- \bigl( G\iota_{n}\theta_{\pi}^{U} \omega ^{T}-E\varTheta_{\pi}^{F} \bigr) \pi_{t} \\ &\qquad{}+ ( I_{n}-K_{y} ) ( I_{n}-R ) y_{t-1}+Ds_{t}- ( I_{n}-K_{y} ) Ds_{t-1}+I_{n}v_{t}^{y}, \end{aligned}$$
(18)
$$\begin{aligned} &{-}K_{\pi}BE_{t}\pi_{t+1}=\varXi y_{t}-\pi_{t}+ ( I_{n}-K_{\pi} ) B \pi_{t-1}+\varXi Vs_{t}+I_{n}v_{t}^{\pi}, \end{aligned}$$
(19)
$$\begin{aligned} &s_{t+1}-SE_{t}\pi_{t+1}=s_{t}, \end{aligned}$$
(20)
$$\begin{aligned} &v_{t+1}=\varPsi v_{t}+\varepsilon _{t+1}. \end{aligned}$$
(21)

Introducing three additional vectors of variables a t+1:=y t , b t+1:=π t and c t+1:=s t we may rewrite system (18)–(21) as:

$$\begin{aligned} &{-}K_{y} ( I_{n}-R ) E_{t}y_{t+1}-GE \pi_{t+1}+K_{y}Ds_{t+1} \\ &\quad=- \bigl( I_{n}-E\varTheta_{y}^{F}+G \iota_{n}\theta_{y}^{U}\omega ^{T}-R \bigr) y_{t}- \bigl( G\iota_{n}\theta_{\pi}^{U} \omega ^{T}-E\varTheta _{\pi}^{F} \bigr) \pi_{t} \\ &\qquad{}+Ds_{t}+ ( I_{n}-K_{y} ) ( I_{n}-R ) a_{t}- ( I_{n}-K_{y} ) Dc_{t}+I_{n}v_{t}^{y} \end{aligned}$$
(22)
$$\begin{aligned} &{-}K_{\pi}BE\pi_{t+1}=\varXi y_{t}- \pi_{t}+\varXi Vs_{t}+ ( I_{n}-K_{\pi } ) Bb_{t}+I_{n}v_{t}^{\pi}, \end{aligned}$$
(23)
$$\begin{aligned} &{-}S\pi_{t+1}+s_{t+1}=s_{t}, \end{aligned}$$
(24)
$$\begin{aligned} &a_{t+1}=y_{t}, \end{aligned}$$
(25)
$$\begin{aligned} &b_{t+1}=\pi_{t}, \end{aligned}$$
(26)
$$\begin{aligned} &c_{t+1}=s_{t}, \end{aligned}$$
(27)
$$\begin{aligned} &v_{t+1}=\varPsi v_{t}+\varepsilon _{t+1}. \end{aligned}$$
(28)

Defining: A 11=−K y (IR), A 12=−G, A 13=K y D, \(B_{11}=- ( I-E\varTheta _{y}^{F}+G\iota _{n}\theta_{y}^{U}\omega^{T}-R ) \), \(B_{12}=- ( G\iota _{n}\theta _{\pi}^{U}\omega^{T}-E\varTheta_{\pi}^{F} ) \), B 13=D, B 14=(I n K y )(I n R), B 16=−(I n K y )D, B 17=[0 n×n I n ], A 22=−K π B, B 21=Ξ, B 22=−I n , B 23=ΞV, B 25=(I n K π )B, B 27=[I n 0 n×n ], A 32=−S, B 77=Ψ, the system (22)–(28) in state-space form as:

$$ E_{t}z_{t+1}=A^{-1}Bz_{t}+F \upsilon_{t}, $$
(29)

where \(z_{t}:= [ y_{t}^{T} \ \pi_{t}^{T} \ s_{t}^{T} \ a_{t}^{T} \ b_{t}^{T} \ c_{t}^{T} \ v_{t}^{T}] ^{T}\) or \(z_{t}:= [ z_{1,t}^{T}\ z_{2,t}^{T}\ v_{t}^{T} ] ^{T}\) with \(z_{1,t}:= [ a_{t}^{T}\ b_{t}^{T}\ c_{t}^{T} ] ^{T}\), \(z_{2,t}:= [ y_{t}^{T}\ \pi_{t}^{T}\ s_{t}^{T} ] ^{T}\),

$$\begin{aligned} \upsilon_{t} :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{}} 0_{1\times n} & 0_{1\times n} & 0_{1\times ( n-1 ) } & 0_{1\times n} & 0_{1\times n} & 0_{1\times ( n-1 ) } & \varepsilon _{t}^{T}\end{array} \right ] ^{T}, \\ A :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{}} A_{11} & A_{12} & A_{13} & 0_{1} & 0_{1} & 0_{2} & 0_{3} \\ 0_{1} & A_{22} & 0_{2} & 0_{1} & 0_{1} & 0_{2} & 0_{3} \\ 0_{4} & A_{32} & I_{ ( n-1 ) } & 0_{4} & 0_{4} & 0_{5} & 0_{6} \\ 0_{1} & 0_{1} & 0_{1} & I_{n} & 0_{1} & 0_{2} & 0_{3} \\ 0_{1} & 0_{1} & 0_{1} & 0_{1} & I_{n} & 0_{2} & 0_{3} \\ 0_{4} & 0_{4} & 0_{5} & 0_{4} & 0_{4} & I_{ ( n-1 ) } & 0_{6} \\ 0_{7} & 0_{7} & 0_{8} & 0_{7} & 0_{7} & 0_{8} & I_{2n}\end{array} \right ], \\ B :=&\left [ \begin{array}{@{}c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{\quad}c@{}} B_{11} & B_{12} & B_{13} & B_{14} & 0_{1} & B_{16} & B_{17} \\ B_{21} & B_{22} & B_{23} & 0_{1} & B_{25} & 0_{2} & B_{27} \\ 0_{4} & 0_{4} & I_{ ( n-1 ) } & 0_{4} & 0_{4} & 0_{5} & 0_{6} \\ 0_{1} & 0_{1} & 0_{1} & I_{n} & 0_{1} & 0_{2} & 0_{3} \\ 0_{1} & 0_{1} & 0_{1} & 0_{1} & I_{n} & 0_{2} & 0_{3} \\ 0_{4} & 0_{4} & 0_{5} & 0_{4} & 0_{4} & I_{ ( n-1 ) } & 0_{6} \\ 0_{7} & 0_{7} & 0_{8} & 0_{7} & 0_{7} & 0_{8} & \varPsi \end{array} \right ], \\ F :=&\left [ \begin{array}{@{}c@{\quad}c@{}} 0_{1} & 0_{1} \\ 0_{1} & 0_{1} \\ 0_{4} & 0_{4} \\ 0_{1} & 0_{1} \\ 0_{1} & 0_{1} \\ 0_{4} & 0_{4} \\ I_{n} & 0_{1} \\ 0_{1} & I_{n}\end{array} \right ], \end{aligned}$$

where 01, 02, 03, 04, 05, 06, 07, 08, 09 are zero matrices of dimensions n×n, n×(n−1), n×2n, (n−1)×n, (n−1)×(n−1), n×2n, 2n×n, 2n×(n−1) and 2n×1.

In order to obtain LQDG NKM, we assume that E t z 1,t+1=z 1,t+1, i.e. that economic agents in the deterministic NKM make neither systematic nor random errors when predicting the future. Furthermore, substituting monetary and fiscal rules (6)–(7) in which deviation is possible into system (12)–(15) in the way presented above and performing similar transformations we obtain the system:

$$\begin{aligned} &{-}K_{y} ( I_{n}-R ) E_{t}y_{t+1}-GE \pi_{t+1}+K_{y}Ds_{t+1} \\ &\quad=- \bigl( I_{n}-E\varTheta_{y}^{F}+G \iota_{n}\theta_{y}^{U}\omega ^{T}-R \bigr) y_{t}- \bigl( G\iota_{n}\theta_{\pi}^{U} \omega ^{T}-E\varTheta _{\pi}^{F} \bigr) \pi_{t} \\ &\qquad{}+ Ds_{t}+ ( I_{n}-K_{y} ) ( I_{n}-R ) a_{t}- ( I_{n}-K_{y} ) Dc_{t}+I_{n}v_{t}^{y}+E \hat{f}_{t}-G\hat{\imath}_{t} \end{aligned}$$
(30)
$$\begin{aligned} &{-}K_{\pi}BE\pi_{t+1}=\varXi y_{t}- \pi_{t}+\varXi Vs_{t}+ ( I_{n}-K_{\pi } ) Bb_{t}+I_{n}v_{t}^{\pi} \end{aligned}$$
(31)
$$\begin{aligned} &{-}S\pi_{t+1}+s_{t+1}=s_{t} \end{aligned}$$
(32)
$$\begin{aligned} &a_{t+1}=y_{t} \end{aligned}$$
(33)
$$\begin{aligned} &b_{t+1}=\pi_{t} \end{aligned}$$
(34)
$$\begin{aligned} &c_{t+1}=s_{t} \end{aligned}$$
(35)
$$\begin{aligned} &v_{t+1}=\varPsi v_{t}, \end{aligned}$$
(36)

which, compared to the system (22)–(28), has two additional vectors of control variables \(\hat{f}_{t}\) and \(\hat {\imath}_{t}\). System (30)–(36) in state-space form can we written as:

$$ z_{t+1}=A^{-1}Bz_{t}+A^{-1}Cu_{t}, $$
(37)

where \(u_{t}:= [ \hat{f}_{t}^{T}\ \hat{\imath}_{t} ] ^{T}\) and

$$ C:=\left [ \begin{array}{@{}c@{\quad}c@{}} E & -G\iota_{n} \\ 0_{1} & 0_{1} \\ 0_{1} & 0_{1} \\ 0_{1} & 0_{1} \\ 0_{4} & 0_{4} \\ 0_{7} & 0_{9}\end{array} \right ]. $$

1.5.2 A.5.2 Initial Condition Derivation

Initial condition z 0 should position the system on the saddle-path so that the model would converge to the equilibrium. We propose two alternative ways of deriving this initial condition:

  1. 1.

    One way to obtain z 0 which positions the system on the saddle-path is to solve the RE version of the model and then use the initial state obtained. This initial state, by definition (if RE-model is stable), meets the required condition because it positions the system on the saddle path. In particular, at t=0 vector of endogenous non-predetermined variables z 1,t will “jump” to a saddle path whereas vector of endogenous state (predetermined) variables z 2,t will have a value of 0. The initial value of shock vector v t should follow the same assumptions made while solving the RE SNKM, i.e. its initial value should equal to standard deviation of ε t . A number of freeware applications is available to solve RE model with DYNARE by Juillard (1996) being probably the most famous.Footnote 34

  2. 2.

    Another method to position the system on the saddle-path is to calculate the orthogonal projection of the shock v t onto the stable subspace at time t=1. This method will be described below in more details.

Let

$$ z_{t+1}=\bar{A}z_{t},\quad z(0)=z_{0}, $$
(38)

be the deterministic NKM, where \(\bar{A} :=A^{-1}B\).

Now, let \(\bar{A}=SJS^{-1}\) be a Jordan decomposition of \(\bar{A}\) such that \(J=\operatorname{diag} ( \varLambda_{S}, \varLambda_{U} ) \) and S=[S S S U ], where Λ S contains all stable eigenvalues of \(\bar{A}\) and Λ U all unstable eigenvalues of \(\bar{A}\) and S S (S U ) is the with Λ S (Λ U ) corresponding stable (unstable) subspace of \(\bar{A}\). Then, if z 0 belongs to S S we have z(0)=S S y for some y (\(y= ( S_{S}^{T}S_{S} ) ^{-1}S_{S}^{T}z_{0}\)).Footnote 35 In that case we may write:

$$\begin{aligned} z(t)& =e^{\bar{A}t}z(0)=Se^{Jt}S^{-1}S_{\bar{S}}y \\ &=Se^{Jt} \left [ \begin{array}{c} I \\ 0\end{array} \right ] y=S\left [ \begin{array}{c} e^{\varLambda_{S}t} \\ 0\end{array} \right ] y =S\left [ \begin{array}{c} e^{\varLambda_{S}t} ( S_{S}^{T}S_{S} ) ^{-1}S_{S}^{T}z_{0} \\ 0\end{array} \right ], \end{aligned}$$

which is the solution for t≥0. In our simulations we always consider such orthogonal projection of z onto the stable subspace at time t=1 as the initial condition \(\tilde{z}_{0}\).

1.5.3 A.5.3 Change from a Discrete- to a Continuous-Time Model

Following Kwakernaak (1976) let the continuous time system be:

$$\begin{aligned} \dot{x}& =Ax+Bu, \\ y& =Cx+Du. \end{aligned}$$

Under the assumptions that u(t)=u(t i ), t i tt i+1 and Δ=t i+1t i the equivalent discrete-time system is:

$$\begin{aligned} x(i+1)& =A_{Cl}x(i)+B_{d}u(i) \end{aligned}$$
(39)
$$\begin{aligned} y(i)& =C_{d}x(i)+D_{d}u(i), \end{aligned}$$
(40)

where

$$\begin{aligned} A_{Cl} =&e^{A\varDelta }, \\ B_{d} =& \biggl( \int_{0}^{\varDelta }e^{A\tau}d\tau \biggr) B, \\ C_{d} =&Ce^{A\varDelta } \quad\mbox{and} \\ D_{d} =&C \biggl( \int_{0}^{\varDelta }e^{A\tau}d\tau \biggr) B+D. \end{aligned}$$

Assuming Δ=1 we may rewrite the continuous time system in terms of discrete time system matrices as:

$$\begin{aligned} A& =\log A_{Cl}=\log ( I+A_{Cl}-I ) \approx A_{Cl}-I, \end{aligned}$$
(41)
$$\begin{aligned} B& = \biggl[ \int_{0}^{1}e^{A\tau}d \tau \biggr] ^{-1}B_{d}= \bigl( e^{A}-I \bigr) ^{-1}AB_{d}, \end{aligned}$$
(42)
$$\begin{aligned} C& =C_{d}e^{-A}, \end{aligned}$$
(43)
$$\begin{aligned} D& =D_{d}-C \biggl( \int_{0}^{\varDelta }e^{A\tau}d \tau \biggr) B \\ & =D_{d}-C \bigl[ A^{-1} \bigl( e^{A}-I \bigr) \bigr] B. \end{aligned}$$
(44)

Rights and permissions

Reprints and permissions

Copyright information

© 2014 Springer-Verlag Berlin Heidelberg

About this chapter

Cite this chapter

Michalak, T., Engwerda, J., Plasmans, J. (2014). Interactions Between Fiscal and Monetary Authorities in a Three-Country New-Keynesian Model of a Monetary Union. In: Haunschmied, J., Veliov, V., Wrzaczek, S. (eds) Dynamic Games in Economics. Dynamic Modeling and Econometrics in Economics and Finance, vol 16. Springer, Berlin, Heidelberg. https://doi.org/10.1007/978-3-642-54248-0_12

Download citation

Publish with us

Policies and ethics