Skip to main content

Government Consumption

  • Chapter
  • First Online:
Public Economics

Part of the book series: Springer Texts in Business and Economics ((STBE))

Abstract

Empirically, government expenditures represent a large share of total demand and significantly affect output, employment, and welfare. In the introductory Sect. 4.2 of this chapter, we document some selected empirical facts of government consumption. In particular, we find that government consumption is procyclical, and after an unexpected increase in consumption, output, employment, and (to a smaller extent) private consumption all increase.

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 99.00
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Hardcover Book
USD 129.99
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.

    The data are taken from IMF, OECD, and Bundesbank statistics. Please see Appendix 4.6 for the documentation of the data. The numbers for the years 2016–2018 represent estimates. The statistics are loaded and graphed with the help of the Gauss program Ch4_data.g.

  2. 2.

    The spike in government spending in Germany during the period 1990–1996 was caused by German reunification and the higher public spending in East Germany.

  3. 3.

    In April 2015, the OECD invited Costa Rica and Lithuania to open formal OECD accession talks.

  4. 4.

    OECD average public and private pension spending amounted to 9.5%.

  5. 5.

    Please take care not to equate (private and public) health expenditures with health. For example, Italy spends only half as much on health as the United States; however, in 2014, the average life expectancy in Italy was approximately 4 years longer than in the US.

  6. 6.

    Again, take care to not equate higher education spending with better education. In their article on “The Economics of International Differences in Educational Achievement” in the Handbook of Economics of Education, Hanushek and Woessmann (2011) review the literature on the determinants of educational attainment. In particular, they find that input measures such as class size or educational expenditures show little impact, while several measures of institutional structures such as school autonomy, later tracking, and the quality of the teaching force explain a significant portion of the international differences in student achievements.

  7. 7.

    See also Footnote 6 in this section for an explanation of why years of schooling might represent a better measure of educational attainment.

  8. 8.

    NATO members such as Germany and Belgium agreed to spend 2% of GDP on defense. Evidently, many NATO countries interpret this official target as a guideline.

  9. 9.

    Please see Appendix 4.6 for a description of the data on government consumption. The source of the data on real GDP, private consumption, and labor supply which is used in the computation of the results displayed in Table 4.1 is presented in Appendix 2.4.

  10. 10.

    The GAUSS computer program Ch4_data.g together with the data file Fred_data1a.txt that computes Figs. 4.7 and 4.8 is available as a download from my web page.

  11. 11.

    The HP Filter is described in Chap. 2 above. We use the parameter λ = 1600 for quarterly data.

  12. 12.

    See Brandner and Neusser (1992).

  13. 13.

    The observation that private consumption is more volatile than output does not hold for all subperiods. For example, Cooley and Prescott (1995) find that the relative volatility of personal consumption with respect to output is only 74% in the US during the period 1954–1991. In addition, these authors document that durable consumption expenditures are much more volatile than the consumption of non-durables and services. Similarly, Heer and Maußner (2009) present empirical evidence for West Germany prior to German reunification over the period 1975–1989, when consumption is only approximately half as volatile as output.

  14. 14.

    Compare this with Table 3 in Ambler and Paquet (1996). The time series are measured in logs and passed through the HP filter.

  15. 15.

    In most vector autoregression studies, the assumption that no innovation other than government spending shocks can affect government spending within a given quarter is used for identification.

  16. 16.

    On p. 2 and in Table 1 on p. 8, Murphy and Walsh (2016) summarizes the studies on the relationship between interest rates and government spending shocks.

  17. 17.

    This need not hold in all specifications of New Keynesian models. For example, Heer and Scharrer (2018) introduce a variable price of capital into these types of models, and consequently, real interest rates decline in response to higher unanticipated government consumption.

  18. 18.

    For this observation to hold, we need to assume that the change in government consumption is fully anticipated.

  19. 19.

    The special case of μ = 0 is considered by Baxter and King (1993) and Aiyagari, Christiano, and Eichenbaum (1994) and is often adapted in business cycle research.

  20. 20.

    Proportional income taxes will be introduced into the model in Chap. 5.

  21. 21.

    To derive the aggregate resource constraint (4.11), substitute (4.4), (4.8), (4.9), and (4.10) into (4.5).

  22. 22.

    Our notational convention in this book is that we use the variable x (X) as a subscript (superscript) of the parameter ρ, i.e. ρ x (ρ X), in case of a utility parameter (autoregressive parameter or parameter of a policy rule). Accordingly, ρ c denotes the substitution elasticity of private and public consumption, while ρ C denoted the autoregressive parameter of the technology shock in the consumption sector in Chap. 2.

  23. 23.

    Among others, the empirical estimates of the elasticities depend on the classification of the government expenditures, e.g., whether military spending is included.

  24. 24.

    In the Gauss computer program Ch4_subs_private_pub.g, the procedure steadystate1(.) is used in the non-linear equation problem solver to compute the solution to this problem.

  25. 25.

    The business cycle properties of the stochastic neoclassical model with government consumption will be studied subsequently. At this point, however, it may be instructive for the reader to see that one can show this result in the deterministic neoclassical growth model with the help of a simple solution technique. We will just use the non-linear equation solver that is applied abundantly in this book.

  26. 26.

    In addition to ϕ = 1, Aiyagari, Christiano, and Eichenbaum (1994) also assume that utility is additively logarithmic in total consumption and leisure.

  27. 27.

    You can also verify in the program Ch4_subs_private_pub_dyn.g that this results holds for other values of ρ c ∈ [0,  1].

  28. 28.

    For example, Schmitt-Grohé and Uribe (2007) use ρ G = 0.87 and σ G = 0.016, while Christiano and Eichenbaum (1992) apply the estimates ρ G = 0.96 and σ G = 0.020.

  29. 29.

    Some studies prefer to display impulses responses for a shock of 1% rather than one standard deviation.

  30. 30.

    In each case where we changed the value of ϕ or/and ρ c, we re-calibrated the parameter ι so that steady-state labor supply is equal to L = 0.30.

  31. 31.

    For our model and the calibration presented in Table 4.1, private consumption declines on impact for all ρ c ≥ 0.67.

  32. 32.

    In Sect. 4.5, we also consider a standard New Keynesian model with frictions in the form of sticky prices and wages. Galí and Lopez-Salido (2007) show that this model is also able to replicate the empirical fact that private consumption rises in response to an unexpected increase in government consumption if one allows for the presence of rule-of-thumb consumers who do not save.

  33. 33.

    The Frisch labor supply elasticity is derived in Appendix 4.2.

  34. 34.

    Domeij and Floden (2006) argue that these estimates are biased downward due to the omission of borrowing constraints.

  35. 35.

    The only equilibrium conditions that change in (4.38) are (4.38a) and (4.38b), which are replaced by

    $$\displaystyle \begin{aligned} \begin{aligned} \lambda_t &= \phi C_t^{-\sigma} \left(\varXi_t \right)^{\frac{1}{1-1/\rho_c}-1} \left( C^p_t\right)^{-\frac{1}{\rho_c}},\\\lambda_t w_t &= \nu_0 L_t^{\frac{1}{\nu_1}}. \end{aligned} \end{aligned}$$
  36. 36.

    You are asked to perform the numerical computation in Problem 4.3. The GAUSS program Ch4_RBC2.g that computes this problem can be downloaded from my homepage.

  37. 37.

    The magnitude and the sign of the impulse responses also depend on the functional form of utility. If we employ utility function (4.40) with η L,w = 1.64 rather than Cobb-Douglas utility (4.28), private consumption also decreases in response to higher government consumption (not illustrated). As expected, the response of labor supply is much stronger than for the case with η L,w = 0.30 and amounts to + 0.11%.

  38. 38.

    Even this statement is subject to restrictions. Preferences are not completely exogenous. For example, firms’ advertising is used to influence consumer preferences or political institutions may have an effect on cultural development and, therefore, our deep utility parameters. One of the first modern economist to point out the endogeneity of preferences was Carl Christian von Weizsäcker (see, e.g., von Weizsäcker (1971)).

  39. 39.

    The expression ‘limited participation,’ as introduced by Christiano, Eichenbaum, and Evans (1997), results from the constraints that agents face in the financial market. Households can only lend funds to the firms with the help of a financial intermediary at the beginning of the period. The central bank injects money into the banking sector after the households have deposited their money at the bank. Hence, households can no longer participate in the financial market, i.e., they have limited participation. At the end of the period, the financial intermediary retrieves the loans from the firms that need to pre-finance labor costs. The different ways to introduce a motive for money demand in general equilibrium models are reviewed in Walsh (2010), among others.

  40. 40.

    Both features help to increase the cost of intertemporal substitution of consumption for the household. As a consequence, the premium on risky assets increases, and the model is also in better accordance with asset price implications. See, for example, Jermann (1998) and Uhlig (2007).

  41. 41.

    To derive that 𝜖 y is equal to the price elasticity, differentiate the demand equation (4.43) with respect to P t(j).

  42. 42.

    We used the chain rule of differentiation and the Leibniz integral rule as presented in Footnote 42 of Chap. 2.

  43. 43.

    In Problem 4.6, you are asked to study the case in which the price P Nt increases by the average inflation π rather than by the inflation in the last period. Also note that the inflation rate amounts to π − 1, while π denotes the inflation factor.

  44. 44.

    Heer and Maußner (2009) show that the (stock market) value of the firm at the beginning of period t + 1 is equal to the value of the capital at the end of period t (=beginning of period t + 1), \(V^{cd}_t= q_t K_{t+1}\). Accordingly, q t describes the (market) value of the firm relative to the replacement cost of capital and, therefore, amounts to the definition of Tobin’s q.

  45. 45.

    The structure of the model seems to be very complicated. We distinguish three production sectors (final goods, wholesale, intermediate goods) and one employment agency. If you consider the alternative case where we only postulate one production sector without an employment agency, the benefits of this fragmentation of services in the production sector are evident. If we had only one production sector that is characterized by monopolistic competition and heterogeneous firms, each firm’s labor demand and price-setting behavior would depend on its marginal costs and, hence, its capital stock. As a consequence, we would not be able to derive simple index functions in the form of (4.45) and (4.59) for the aggregate price and aggregate wage. Instead, these aggregate prices would depend on the distribution of capital in the production sector.

  46. 46.

    This specification was introduced by Constandinides (1990). As an alternative, Abel (1990) uses the ratio of consumption and habits, \(C_t/\bar C_t\), in the utility function.

  47. 47.

    The results are not sensitive to this assumption.

  48. 48.

    An inflation reaction coefficient in excess of unity prevents self-fulfilling expectations with respect to the path of inflation. See, for example, Bullard and Mitra (2002). The intuition for this behavior is quite simple. Assume that aggregate demand and prices increase and that the other reaction coefficients are equal to zero, θ R = θ Y = 0. If the reaction coefficient θ π were less than one, nominal interest rates would rise less than prices so that the real interest rate would decline. As a consequence, aggregate demand would increase further and inflation went up even more. The monetary policy clearly would become unstable. While we exclude this kind of behavior in our model, we, however, refrain from imposing a lower zero bound on the net nominal interest rate Q t − 1, as it rarely becomes binding in our simulations.

  49. 49.

    See Fig. 2.15 and Eq. (2.59) in Appendix 2.1.

  50. 50.

    The sequence does not converge for the standard secant method with κ = 1.

  51. 51.

    The reader is invited to experiment with the values of the parameters {φ y, φ w, 𝜖 y, 𝜖 w, ζ, χ} in the GAUSS program Ch4_newkeynesian.g in order to study the sensitivity of the labor impulse responses.

  52. 52.

    Farhi and Werning (2016) demonstrate in a standard New Keynesian model that the multiplier increases and exceeds unity in case of a liquidity trap (in which interest rates hit zero). Extending their model to the open economy, these authors find that the fiscal multiplier is smaller and below one for a country in a currency union. Erceg and Linde (2012) find that the fiscal multiplier is below one in an economy with fixed-exchange rates and, in accordance with the Mundell-Flemming model, above the one with flexible exchange rates. They show that their latter result is sensitive with respect to the slope of the Phillips curve and the presence of a persistent liquidity trap.

  53. 53.

    Of course, the response of hours depend on our assumption that taxes are lump-sum rather than proportional to wage income.

  54. 54.

    Linnemann and Schabert (2003) show analytically how the central bank’s rule affects the consequences of higher government consumption for labor, output, and prices. In response to higher (government) demand, labor demand increases, while the wealth effect drives up labor supply. The strength of the demand effect depends on the response of the real interest rate, which is governed by the monetary policy rule. When the rise in the real interest rate is dampened by an interest rate rule (as in our case), output and inflation can increase. They also show that, if the central bank follows a simple money-growth rule, fiscal expansions could be both deflationary and contractionary.

  55. 55.

    Ni (1995) provides empirical evidence that the estimates of the coefficient of public consumption in utility, (1 − ϕ)∕ϕ, are of small magnitude, with their signs depending on the measure of interest rates. If he uses net-of tax real taxes in his GMM estimation, he finds a negative coefficient, which corresponds to ϕ > 1 above.

  56. 56.

    The positive response of private consumption is even more pronounced if adjustment costs are smaller, e.g., with a capital adjustment cost parameter ζ = 0. In this case, firms reduce their capital stock more rapidly and investment declines more strongly, meaning that more resources are freed up for private consumption.

  57. 57.

    Heer and Scharrer (2018) present a model that is in accordance with all the empirical impulse responses of output, labor, demand components, and factor prices. For this reason, they introduce both rule-of-thumb consumers and a variable price of capital in terms of the consumption goods into an otherwise standard New Keynesian model.

  58. 58.

    Nekarda and Ramey (2013) present evidence that the price-cost markup is procyclical or at best acyclical, which causes problems for standard New Keynesian models.

  59. 59.

    Since we simulate time series of output with the help of a random number generator for the three shocks 𝜖 Z, 𝜖 G, and 𝜖 Q, the results do not lie exactly on the curves displayed in Fig. 4.22. To smooth the curve, we fitted a polynomial of order two to the data points using a simply OLS regression. The estimation is contained in the GAUSS program Ch4_new_keynes_stabil.g.

  60. 60.

    At this point, we refrain from deriving the optimal fiscal policy because it would take us too far into the field of numerical methods. Using perturbation methods of higher order, Schmitt-Grohé and Uribe (2007) derive optimal monetary and fiscal policy rules. For the fiscal policy rule, they consider a tax rule that sets total taxes as a function of government liabilities and the fiscal deficit. They find that whether the fiscal policy rule is active or passive does not significantly affect welfare.

  61. 61.

    As another sensitivity analysis, we considered the case with ϕ = 1.0, such that public consumption does not affect household utility. In this case, the output-volatility-minimizing fiscal policy is specified with ρ Y = −1.3.

  62. 62.

    Our microfounded model has the benefit that we can quantitatively compare the welfare of different stabilization policies. In the present model, the equilibrium is not Pareto-efficient because various welfare distortions are present. First, firms in the wholesale sector operate as monopolistic competitors. Second, there is both price and wage dispersion.

  63. 63.

    We will introduce income taxes and debt in the upcoming Chaps. 5 and 7, respectively.

  64. 64.

    See, for example, Chapter 2 in Barro and Sala-i-Martin (2003) for a derivation of the transition dynamics in the continuous-time neoclassical growth model and, in particular, Section 2.6.6 for the speed of convergence.

  65. 65.

    You will be asked to compute the solution in Problem 4.1.

  66. 66.

    Appendices 4.3 and 4.4 were afforded in large parts by Alfred Maußner and are based upon the exposition in Heer, Maußner, and Ruf (2017). A more detailed description of the derivation of the microfoundations of Calvo price staggering can be found in Maußner (2000). I would like to thank Alfred for his thoughtful comments and support that have greatly helped to improve the presentation of the material in this chapter. All remaining errors are mine.

References

  • Abel, A. B. (1990). Asset pricing under habit formation and catching-up-with-the-Joneses. American Economic Review, 80, 38–42.

    Google Scholar 

  • Aiyagari, R. S., Christiano, L. J., & Eichenbaum, M. (1994). The output, employment, and interest effect of government consumption. Journal of Monetary Economics, 30, 73–86.

    Article  Google Scholar 

  • Altonij, J. G. (1986). Intertemporal substitution in labor supply: Evidence from micro data. Journal of Political Economy, 94, S176–S215.

    Article  Google Scholar 

  • Ambler, S., & Paquet, A. (1996). Fiscal spending shocks, endogenous government spending, and real business cycles. Journal of Economic Dynamics and Control, 20, 237–256.

    Article  Google Scholar 

  • Aschauer, D. A. (1985). Fiscal policy and aggregate demand. American Economic Review, 75, 117–121.

    Google Scholar 

  • Auerbach, A. J., & Gorodnichenko, Y. (2011). Fiscal multipliers in recession and expansion. NBER working paper, 17447.

    Google Scholar 

  • Barro, R. J. (1981). Output effects of government purchases. The Journal of Political Economy, 89(6), 1086–1121.

    Article  Google Scholar 

  • Barro, R. J., & Sala-i-Martin, X. (2003). Economic growth (2nd ed.). Cambridge, MA: MIT Press.

    Google Scholar 

  • Basu, S., Fernald, J. G., & Kimball, M. S. (2006). Are technology improvements contractionary?. American Economic Review, 96, 1418–1448.

    Article  Google Scholar 

  • Baxter, M., & King, R. G. (1993). Fiscal policy in general equilibrium. American Economic Review, 91, 167–185.

    Google Scholar 

  • Blanchard, O. J., & Perotti, R. (2002). An empirical characterization of the dynamic effects of changes in government spending and taxes on output. Quarterly Journal of Economics, 177, 1329–1368.

    Article  Google Scholar 

  • Brandner, P., & Neusser, K. (1992). Business cycles in open economies: Stylized facts for Austria and Germany. Review of World Economics, 128(1), 67–87.

    Article  Google Scholar 

  • Brinca, P., Holter, H. A., Krusell, P., & Malafry, L. (2016). Fiscal multipliers in the 21st century. Journal of Monetary Economics, 77, 53–69.

    Article  Google Scholar 

  • Bullard, J., & Mitra, K. (2002). Learning about monetary policy rules. Journal of Monetary Economics, 49(6), 1105–1129.

    Article  Google Scholar 

  • Calvo, G. A. (1983). Staggered prices in a utility-maximizing framework. Journal of Monetary Economics, 12, 383–398.

    Article  Google Scholar 

  • Christiano, L. J., & Eichenbaum, M. (1992). Current Real-Business-Cycle theories and aggregate labor market fluctuations. American Economic Review, 82(3), 430–450.

    Google Scholar 

  • Christiano, L. J., Eichenbaum, M., & Evans, C. L. (1997). Sticky prices and limited participation models of money: A comparison. European Economic Review, 41, 1201–1249.

    Article  Google Scholar 

  • Christiano, L. J., Eichenbaum, M., & Evans, C. L. (2005). Nominal rigidities and the dynamic effects of a shock to monetary policy. Journal of political Economy, 113(1), 1–45.

    Article  Google Scholar 

  • Clarida, R., Gali, J., & Gertler, M. (2000). Monetary policy rules and macroeconomic stability: Evidence and some theory. Quarterly Journal of Economics, 115(1), 147–180.

    Article  Google Scholar 

  • Constandinides, G. M. (1990). Habit formation: A resolution of the equity premium. Journal of Political Economy, 98, 519–543.

    Article  Google Scholar 

  • Cooley, T. F., & Prescott, E. C. (1995). Economic growth and business cycles. In T. F. Cooley (Ed.), Frontiers of business cycle research (pp. 1–64). Princeton: Princeton University Press.

    Google Scholar 

  • Davis, S. J., & Kahn, J. A. (2008). Interpreting the Great Moderation: Changes in the volatility of economic activity at the macro and micro levels. Journal of Economic Perspectives, 22(4), 155–180.

    Article  Google Scholar 

  • Devereux, M. B., Head, A. C., & Lapham, B. J. (1996). Monopolistic competition, increasing returns, and the effects of government spending. Journal of Money, Credit, and Banking, 28(2), 233–254.

    Article  Google Scholar 

  • Dixit, A., & Stiglitz, J. (1977). Monoplistic competition and optimum product diversity. American Economic Review, 67(3), 297–308.

    Google Scholar 

  • Domeij, D., & Floden, M. (2006). The labor supply elasticity and borrowing constraints: Why estimates are biased. Review of Economic Dynamics, 9, 242–262.

    Article  Google Scholar 

  • Dotsey, M., & Ireland, P. (1996). The welfare cost of inflation in general equilibrium. Journal of Monetary Economics, 37, 29–47.

    Article  Google Scholar 

  • Dynan, K. E., Elmendorf, D. W., & Sichel, D. (2006). Can financial innovation help to explain the reduced volatility of economic activity? Journal of Monetary Economics, 53(1), 123–150.

    Article  Google Scholar 

  • Erceg, C. J., Henderson, D. W., & Levin, A. T. (2000). Optimal monetary policy with staggered wage and price contracts. Journal of Monetary Economics, 46(2), 281–313.

    Article  Google Scholar 

  • Erceg, C. J., & Linde, J. (2012). Fiscal consolidation in an open economy. American Economic Review, 102(3), 186–191.

    Article  Google Scholar 

  • Farhi, E., & Werning, I. (2016). Fiscal multipliers: Liquidity traps and currency unions. In J. B. Taylor & H. Uhlig (Eds.), Handbook of the macroeconomics (pp. 2418–2492). Amsterdam: Elsevier.

    Google Scholar 

  • Fatás, A., & Mihov, I. (2001). The effects of fiscal policy on consumption and employment: Theory and evidence. CEPR Discussion Papers 2760, C.E.P.R. Discussion Papers.

    Google Scholar 

  • Fehr, H., Kallweit, M., & Kindermann, F. (2013). Should pensions be progressive? European Economic Review, 63, 94–116.

    Article  Google Scholar 

  • Féve, P., Matherod, J., & J.-G. Sahuc (2012). A pitfall with DSGE-based, estimated government spending multipliers. IDIE Working Paper, 708.

    Google Scholar 

  • Fisher, J. D. M., & Peters, R. (2010). Using stock returns to identify government spending shocks. Economic Journal, 120(544), 414–436.

    Article  Google Scholar 

  • Galí, J. (1999). Technology, employment, and the business cycle: Do technology shocks explain aggregate fluctuations? American Economic Review, 89, 249–271.

    Article  Google Scholar 

  • Galí, J., & Lopez-Salido, J. D. (2007). Understanding the effects of government spending on consumption. Journal of the European Economic Association, 5(1), 227–270.

    Article  Google Scholar 

  • Hanushek, E. A., & Woessmann, L. (2011). The economics of international differences in educational achievement. In E. A. Hanushek, S. Machin, & L. Woessmann (Eds.), Handbook of the Economics of Education (Vol. 3, pp. 89–200). Amsterdam: Elsevier.

    Google Scholar 

  • Heer, B., & Maußner, A. (2009). Dynamic general equilibrium modeling: Computational methods and applications (2nd ed.). Heidelberg: Springer.

    Book  Google Scholar 

  • Heer, B., Maußner, A., & McNelis, P. (2011). The money-age distribution: Empirical facts and the limits of three monetary models. Journal of Macroeconomics, 33, 390–405.

    Article  Google Scholar 

  • Heer, B., Maußner, A., & Ruf, H. (2017). Q-Targeting in New Keynesian models. Journal of Business Cycle Research, 13, 189–224.

    Article  Google Scholar 

  • Heer, B., Rohrbacher, S., & Scharrer, C. (2017). Aging, the Great Moderation and business-cycle volatility in a life-cycle model. Macroeconomic Dynamics, 21, 362–383.

    Article  Google Scholar 

  • Heer, B., & Scharrer, C. (2018). The age-specific burdens of short-run fluctuations in government spending. Journal of Economic Dynamics and Control, 90, 45–75.

    Article  Google Scholar 

  • Hindriks, J., & Myles, G. D. (2006). Intermediate public economics. Cambridge, MA: MIT Press.

    Google Scholar 

  • Jaimovich, N., & Siu, H. E. (2009). The young, the old, and the restless: Demographics and business cycle volatility. American Economic Review, 99, 804–826.

    Article  Google Scholar 

  • Jermann, U. J. (1998). Asset pricing in production economies. Journal of Monetary Economics, 41, 257–275.

    Article  Google Scholar 

  • Killingsworth, M. R. (1983). Labor supply. Cambridge, MA: Cambridge University Press.

    Book  Google Scholar 

  • Kwan, Y. K. (2006). The direct substitutability between government and private consumption in East Asia. NBER working paper, 12431.

    Google Scholar 

  • Linnemann, L., & Schabert, A. (2003). Fiscal policy in the New Neoclassical synthesis. Journal of Money, Credit and Banking, 35(6), 911–929.

    Article  Google Scholar 

  • Lucas, R. E. (1976). Econometric policy evaluation: A critique. Carnegie-Rochester Series on Public Policy, 1, 19–46.

    Article  Google Scholar 

  • Lugauer, S. (2012). Estimating the effect of the age distribution on cyclical output volatility across the United States. The Review of Economics and Statistics, 94(4), 896–902.

    Article  Google Scholar 

  • Lugauer, S., & Redmond, M. (2012). The age distribution and business cycle volatility: International evidence. Economics Letters, 117(3), 694–696.

    Article  Google Scholar 

  • MaCurdy, T. E. (1981). An empirical model of labor supply in a life-cycle setting. Journal of Political Economy, 89, 1059–1085.

    Article  Google Scholar 

  • Maußner, A. (2000). The analytics of New Keynesian Phillips curves. Discussion Paper Series Institute of Economics, University of Augsburg, 313.

    Google Scholar 

  • Monacelli, T., & Perotti, R. (2008). Fiscal policy, wealth effects, and markups. NBER working paper, 14584.

    Google Scholar 

  • Monacelli, T., Perotti, R., & Trigari, A. (2010). Unemployment fiscal multipliers. Journal of Monetary Economics, 57, 531–553.

    Article  Google Scholar 

  • Murphy, D., & Walsh, K. J. (2016). Spending shocks and interest rates. Working Paper.

    Google Scholar 

  • Nekarda, C., & Ramey, V. A. (2013). The cyclical behavior of the price-cost markup. Working paper, mimeo.

    Google Scholar 

  • Ni, S. (1995). An empirical analysis on the substitutability between private consumption and government purchases. Journal of Monetary Economics, 36, 593–605.

    Article  Google Scholar 

  • Ramey, V. A. (2011). Identifying government spending shocks: It’s all in the timing. The Quarterly Journal of Economics, 126, 1–50.

    Article  Google Scholar 

  • Ramey, V. A., & Shapiro, M. D. (1998). Costly capital reallocation and the effects of government spending. Carnegie-Rochester Conference Series on Public Policy, 48(1), 145–194.

    Article  Google Scholar 

  • Ravn, M. O., Schmitt-Grohé, S., & Uribe, M. (2012). Consumption, government spending, and the real exchange rate. Journal of Monetary Economics, 59, 215–234.

    Article  Google Scholar 

  • Rotemberg, J. J., & Woodford, M. (1992). Oligopolistic pricing and the effects of aggregate demand on economic activity. Journal of Political Economy, 100, 1153–1207.

    Article  Google Scholar 

  • Schmitt-Grohé, S., & Uribe, M. (2007). Optimal simple and implementable monetary and fiscal rules. Journal of Monetary Economics, 54, 1702–1725.

    Article  Google Scholar 

  • Stock, J. H., & Watson, M. W. (2003). Has the business cycle changed and why? In M. Gertler & K. Rogoff (Eds.), National Bureau of Economic Research macroeconomics annual 2002 (pp. 159–230). Cambridge, MA: MIT Press.

    Google Scholar 

  • Taylor, J. B. (1993). Discretion versus policy rules in practice. Carnegie-Rochester Conference Series on Public Policy, 33, 195–214.

    Article  Google Scholar 

  • Trabandt, M., & Uhlig, H. (2011). The Laffer curve revisited. Journal of Monetary Economics, 58, 305–327.

    Article  Google Scholar 

  • Uhlig, H. (2007). Explaining asset prices with external habits and wage rigidities in a DSGE model. American Economic Review: Papers and Proceedings, 97, 239–243.

    Article  Google Scholar 

  • von Weizsäcker, C. C. (1971). Notes on endogenous change of tastes. Journal of Economic Theory, 3, 345–372.

    Article  Google Scholar 

  • Walsh, C. E. (2005). Labor market search, sticky prices, and interest rate policy. Review of Economic Dynamics, 8, 829–849.

    Article  Google Scholar 

  • Walsh, C. E. (2010). Monetary theory and policy. Cambridge, MA: MIT Press.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Appendices

Appendix 4.1: Reverse Shooting

To compute the transition dynamics in the Numerical Example of Sect. 4.3, we first have to set the length of the transition period. When is the transition complete? We will assume that the dynamics are complete when the state variables are reasonably close to the new steady-state values. For practical reasons, we will stop searching for a better length of the transition if the divergence between the value of the state variable K t in the last period of the transition from the steady-state value is less than 0.01% or if the value of the divergence is small, e.g., less than 10−5 in absolute value. The latter number is used when the state variable is very small and 0.01% of the state variable would be close to the accuracy of the solution algorithm (of the non-linear equation solver).

As we will discover in the remainder of the book, the transition in the neoclassical growth model occurs relatively fast compared to that in the other benchmark model that we consider in this book, the overlapping generations (OLG) model. Often, a transition length of fewer than 50 periods is sufficient, while we need to consider more than 100 periods in the OLG model in later chapters. We will choose 40 periods for the present case.

In essence, we have to solve for the dynamics of the three endogenous variables K t, \(C^p_t\), and L t for t = 1, …, 40 given \(K_0=\bar K\), \(L_0=\bar L\) and \(C^p_0=\bar C^p\) for the initial values and \(L_{41}=\tilde L\), \(K_{41}=\tilde K\), and \(C^p_{41}=\tilde C^p\) for the final values in case 1 and \(L_{41}=\bar L\), \(K_{41}=\bar K\), and \(C^p_{41}=\bar C^p\) in case 2. More formally, we have to solve a two-point boundary value problem.

Let us first consider case 1 with a permanent change in government consumption \(G_t=\tilde G\) for t = 1, …. For the solution, we will make use of the first-order conditions, which we restate for your convenience:

$$\displaystyle \begin{aligned} w_t & = \frac{\kappa}{\phi}\left(\frac{C_t}{1-L_t}\right)^{\frac{1}{\rho}} \left(\varXi_t \right)^{1-\frac{1}{1-1/\rho_c}} \left( C^p_t\right)^{\frac{1}{\rho_c}}, \end{aligned} $$
(4.75a)
$$\displaystyle \begin{aligned} \beta (1+r_{t+1}-\delta) &= \left(\frac{X_{t+1}}{X_t}\right)^{1-\frac{1-\sigma}{1-1/\rho_c}} \left(\frac{C_{t+1}}{C_t}\right)^{\frac{1}{\rho}} \left(\frac{\varXi_{t+1}}{\varXi_t}\right)^{1-\frac{1}{1-1/\rho_c}} \left(\frac{C^p_{t+1}}{C^p_t}\right)^{\frac{1}{\rho_c}}, \end{aligned} $$
(4.75b)
$$\displaystyle \begin{aligned} K_{t+1}&=(1-\delta) K_t + K_t^\alpha L_t^{1-\alpha}-G_t-C^p_t, \end{aligned} $$
(4.75c)

where effective consumption C t and the real interest rate r t are given by (4.21) and (4.9b), and X t is defined as follows:

$$\displaystyle \begin{aligned} X_t\equiv C_t^{1-\frac{1}{\rho}}+\kappa (1-L_t)^{1-\frac{1}{\rho}}.\end{aligned} $$
(4.76)

In the first period of the transition at t = 1, the government unexpectedly increases public consumption G t. As a consequence, the household can only adjust its behavior in period t = 1, not before. The beginning-of-period capital stock K 1, therefore, is predetermined by the decision of the household in period t = 0. The household can only choose consumption \(C^p_1\), labor L 1, and the next-period capital stock K 2. As a consequence, we have to solve for the 119 unknowns \(\{K_t\}_{t=2}^{40}\), \(\{L_t\}_{t=1}^{40}\), and \(\{C^p_t\}_{t=1}^{40}\). We require that K 40 diverges by less than 0.01% from \(\tilde K=1.2946\).

The algorithm that we will apply is called reverse shooting. We will provide a guess of K 40 that is close to the new steady-state value \(K_{41}=\tilde K\). We can compute the values L 40 and \(C^p_{40}\) with the help of the first-order conditions (4.75a) and (4.75c) for period t = 40 and using the steady-state values \(K_{41}=\tilde K\), \(L_{41}=\tilde L\), and \(C^p_{41}=\tilde C^p\) for the values of the endogenous variables in period t + 1 = 41. The problem is to solve a system of two non-linear equations in two unknowns, L 40 and \(C^p_{40}\). As the initial guess that we have to provide to the non-linear equation solver routine used in the Gauss program Ch4_subs_private_pub_dyn.g, we simply take the new steady-state values \(\tilde L\) and \(\tilde C^p\), respectively.

With the help of the values \(\{K_{40},L_{40},C^p_{40}\}\), we can compute the values \(\{K_{39},L_{39},C^p_{39}\}\) using the three non-linear equations (4.75a), (4.75b), and (4.75c). More generally, we can compute \(\{K_{t},L_{t},C^p_{t}\}\) with the help of \(\{K_{t+1},L_{t+1},C^p_{t+1}\}\) in the same way for t = 39, 38, …, 1 providing \(\{K_{t+1},L_{t+1},C^p_{t+1}\}\) as an initial value to the non-linear equation solver. In the final iteration, we have \(\{K_1,L_1,C^p_1\}\). If K 1 is close to its value in the initial steady state \(\bar K\) we are done. Otherwise, we have to adjust our guess of K 40 and retry.

How do we choose an initial guess K 40? This is not a trivial task. First, we know that K 1 is smaller than K 41 from our steady-state computation in the Gauss program Ch4_subs_private_pub.g. Since we know from the standard continuous-time Ramsey model that the transition path is saddle-point stable and that the speed of convergence declines during the transition, we would expect K 40 to lie very close to K 41.Footnote 64 For this reason, we perform a grid search of the optimal start value K 40 in the interval \([\bar K+0.9 (\tilde K-\bar K),\tilde K]\). We choose an equispaced grid of nk = 1, 000 points. For the lower values in this grid, the capital stock falls below zero in fewer than 40 periods during the recursive iteration over \(\{K_t,L_t,C^p_t\}\), and we subsequently have to exclude these values. As it turns out, the optimal point (that produces a value of K 1 closest to \(\bar K\)) is very close to \(\tilde K\), and again, we use a much tighter grid, \([\bar K+0.999 (\tilde K-\bar K),\tilde K]\), and find K 40 = 1.2945418 as our initial guess. Given this value, we simply find the solution to the non-linear equation \(f(K_1)=K_1-\bar K=0\).

How does our algorithm perform? We compute a value of K 1 that diverges by less than 10−8 from the old steady-state value \(\bar K=1.2882145\). Therefore, we have very accurately computed the transition dynamics. Let us conclude this section with some qualifying remarks:

  1. 1.

    If we had chosen a higher number for the transition periods than 40, we might have failed to compute the transition dynamics. Why? Let us consider the final values of K 39 and K 40. We have computed the values of K 39 = 1.2945418 and K 40 = 1.2945502 for an accuracy of 10−8. The values are very close to one another, and the transition is basically complete after 30–35 periods. If we had chosen an even higher number, say 100, we would not have been able to provide an initial guess for K 100 that is different from \(K_{101}=\tilde K\) given machine accuracy. If we, however, use K 100 = K 101, our algorithm will just compute the steady-state values for L 100 and \(C^p_{100}\) in the first step and for all other values \(\{K_t,L_t,C^p_t\}\) for t = 99, …, 1. If we, however, consider a smaller value, e.g., K 100 = 1.2946334 instead of \(\tilde K=1.2946335\), we will obtain a negative value for K t in our recursive iteration for t > 1 and be unable to find a solution.

  2. 2.

    Is it possible to solve for the dynamics if we select K 2 as a starting value and iterate forward (so-called forward shooting)? Yes. The reason is as follows: Given K 1 and K 2, we can solve for \(C^p_1\) and L 1 using (4.75a) and (4.75c). In the next step, we seek to solve for \(\{K_3,C^p_2,L_2\}\). Therefore, we use (4.75a) and (4.75c) for period t = 2 and (4.75b) for period t = 1. The solution \(\{K_3,C^p_2,L_2\}\) is used in the next step. We iterate forward until we have found the solution for \(\{K_{41},C^p_{40},L_{40}\}\) and compare K 41 to the new steady-state value \(\tilde K\). If we are close, we are done. Otherwise, we have to specify a new guess for \(\tilde K\).Footnote 65

    In Chap. 6, you will encounter a problem (an OLG model with pay-as-you-go pensions) where reverse shooting is possible, while forward shooting is not. In practice, reverse shooting is often used even if forward shooting is possible. The reason is because it is often easier to provide an initial guess for the capital stock in the last period rather than the first period of the transition because the speed of convergence declines and you have to search in a smaller neighborhood around the final steady state than in the case for a guess of the capital stock in the initial period. If you use forward shooting, you have to search in a larger neighborhood of the initial steady state. This is particularly cumbersome if your state variable is not a single variable but is multi-dimensional.

  3. 3.

    In the case of a temporary government shock, we know that the new steady-state values are equal to the old steady-state values. In this case, however, the capital stock K t approaches K 41 from above. Therefore, we have to provide a guess for K 40 that is larger than \(K_{41}=\bar K\). The rest of the computational procedure is completely equivalent to the case of a permanent increase in government consumption.

Appendix 4.2: Frisch Labor Supply Elasticity for Cobb-Douglas Utility

The Frisch labor supply elasticity or intertemporal labor supply elasticity η L,w is defined as the percentage change in the labor supply in response to a 1% increase in the wage given a constant marginal utility of consumption u C:

$$\displaystyle \begin{aligned} \eta_{L,w} \equiv \frac{d L}{dw}|{}_{u_C = \mbox{const}} \frac{w}{L}. \end{aligned} $$
(4.77)

Let utility u(C, L) be a function of consumption C and labor supply L. Ignoring taxes, contributions, and pensions, the first-order condition of the household with respect to its labor supply is given by:

$$\displaystyle \begin{aligned} -u_L(C,L) = w u_C(C,L). \end{aligned} $$
(4.78)

According to (4.78), the disutility from working another time unit is equal to the utility from the additional consumption that can be afforded by the additional wage income from working an additional time unit. The total differential of (4.78) for a constant u C(C, L) is given by

$$\displaystyle \begin{aligned}- u_{LL} dL - u_{CL} dC = u_C dw.\end{aligned} $$
(4.79)

Furthermore, du C = 0 implies

$$\displaystyle \begin{aligned} u_{CC} dC + u_{CL} dL =0,\end{aligned}$$

or

$$\displaystyle \begin{aligned} dC = -\frac{u_{CL}}{u_{CC}} dL.\end{aligned}$$

Inserting the last equation into (4.79), we obtain

$$\displaystyle \begin{aligned} \eta_{L,w} \equiv \frac{d L}{dw}|{}_{u_C = \mbox{const}} \frac{w}{L} = \frac{u_C}{\frac{u_{CL}^2}{u_{CC}}-u_{LL}} \frac{w}{L}, \end{aligned} $$
(4.80)

and for the Cobb-Douglas utility function with \(u(C,L)=\frac {\left ( C^{\iota } (1-L)^{1-\iota }\right )^{1-\sigma }-1}{1-\sigma }\), the Frisch labor supply elasticity is

$$\displaystyle \begin{aligned} \eta_{L,w}= \frac{1-\iota (1-\sigma)}{\sigma} \frac{1-L}{L}.\end{aligned}$$

Appendix 4.3: Microfoundations of Calvo Price Setting

Let us consider a wholesale firm j with the relative price P t+s(j)∕P t+s.Footnote 66 In period t, the firm received the signal to choose its optimal relative price p At = P AtP t and, since then, has not received a signal to do so again up to period t + s. Between period t and t + s, the price of good j, P t(j), increases with the lagged inflation rate π t, π t+1, …, π t+s−1 in each period t + 1, t + 2, …, t + s, while the aggregate price level P t increases with the inflation rates π t+1, π t+2, …, π t+s:

$$\displaystyle \begin{aligned}\frac{P_{t+s}(j)}{P_{t+s}}=\frac{\pi_{t+s-1}\cdots\pi_t}{\pi_{t+s}\cdots\pi_{t+1}}p_{At}=\frac{\pi_t}{\pi_{t+s}}p_{At}.\end{aligned}$$

Accordingly, the firm will choose p At in period t to maximize discounted dividends (after inserting the demand function (4.43) into dividends (4.46)):

$$\displaystyle \begin{aligned} \mathbb{E}_t \ \sum_{s=0}^\infty (\beta\varphi_y)^s \frac{\varLambda_{t+s}}{\varLambda_t} \left[ \left(\frac{\pi_t}{\pi_{t+s}}p_{At}\right)^{1-\epsilon_y}Y_{t+s}-g_{t+s}\left(\frac{\pi_t}{\pi_{t+s}}p_{At}\right)^{-\epsilon_y}Y_{t+s}\right],\end{aligned}$$

where (φ y)s denotes the probability that the firm cannot adjust its price for s consecutive periods.

Differentiating this equation with respect to p At results in the first-order condition:

$$\displaystyle \begin{aligned}0= \mathbb{E}_t \sum_{s=0}^\infty (\beta\varphi_y)^s\frac{\varLambda_{t+s}}{\varLambda_t} \left[ (1-\epsilon_y)\left(\frac{\pi_t}{\pi_{t+s}}\right)^{1-\epsilon_y}Y_{t+s}p_{At}^{-\epsilon_y}+ \epsilon_y g_{t+s}\left(\frac{\pi_t}{\pi_{t+s}}\right)^{-\epsilon_y}Y_{t+s}p_{At}^{-\epsilon_y-1} \right],\end{aligned}$$

where we used the theorem that the derivative operator can be interchanged with the expectational operator and the sum operator.

The first-order equation can be re-written as

$$\displaystyle \begin{aligned} p_{At} &= \frac{\epsilon_y}{\epsilon_y-1}\frac{\varGamma_{1t}}{\varGamma_{2t}}, \end{aligned} $$
(4.81a)
$$\displaystyle \begin{aligned} \varGamma_{1t} &= \mathbb{E}_t \sum_{s=0}^\infty (\beta\varphi_y)^s \left(\frac{\pi_t}{\pi_{t+s}}\right)^{-\epsilon_y}g_{t+s}\varLambda_{t+s}Y_{t+s}= g_t \varLambda_t Y_t + (\beta\varphi_y) \mathbb{E}_t \left(\frac{\pi_t}{\pi_{t+1}}\right)^{-\epsilon_y}\varGamma_{1t+1}, \end{aligned} $$
(4.81b)
$$\displaystyle \begin{aligned} \varGamma_{2t} &= \mathbb{E}_t \sum_{s=0}^\infty (\beta\varphi_y)^s \left(\frac{\pi_t}{\pi_{t+s}}\right)^{1-\epsilon_y}\varLambda_{t+s}Y_{t+s}= \varLambda_t Y_t + (\beta\varphi_y) \mathbb{E}_t \left(\frac{\pi_t}{\pi_{t+1}}\right)^{1-\epsilon_y} \varGamma_{2t+1}. \end{aligned} $$
(4.81c)

Γ 1t and Γ 2t are simply auxiliary variables whose behaviors are described by (stochastic) first-order difference equations. Therefore, they are easily amenable to the solution with the linearization methods described in Appendix 2.3.

The price index (4.45) implies

$$\displaystyle \begin{aligned}P_t^{1-\epsilon_y}=(1-\varphi_y)P_{At}^{1-\epsilon_y}+\varphi_y P_{Nt}^{1-\epsilon_y}=(1-\varphi_y)P_{At}^{1-\epsilon_y}+\varphi_y (\pi_{t-1}P_{t-1})^{1-\epsilon_y}.\end{aligned}$$

The second equality follows from the updating rule (4.47) and the fact that the non-optimizers are a random sample of optimizers and non-optimizers. Dividing by P t on both sides yields:

$$\displaystyle \begin{aligned} 1 = (1-\varphi_y)p_{At}^{1-\epsilon_y} + \varphi_y (\pi_{t-1}/\pi_t)^{1-\epsilon_y}. \end{aligned} $$
(4.81d)

Finally, consider the definition of \(\tilde Y_t\) given in (4.50):

$$\displaystyle \begin{aligned}\tilde Y_t = \int_{0}^1 Y_{t}(j)\; dj = \int_0^1 \left(\frac{P_{t}(j)}{P_t}\right)^{-\epsilon_y}Y_t\; dj=\left(\frac{\tilde P_t}{P_t}\right)^{-\epsilon_y}Y_t,\end{aligned}$$

with the definitions

$$\displaystyle \begin{aligned} \tilde P_t^{-\epsilon_t}\equiv \int_0^1 P_{t}(j)^{-\epsilon_y}\; dj,\end{aligned}$$

and

$$\displaystyle \begin{aligned} s_t^y \equiv \left(\frac{\tilde P_t}{P_t}\right)^{-\epsilon_y}.\end{aligned}$$

Therefore,

$$\displaystyle \begin{aligned} \tilde Y_t&= s_t^y Y_t. \end{aligned} $$
(4.81e)

Using the same reasoning for \(\tilde P_t\) as for the price index P t above results in the following first-order difference equation for the dispersion of individual prices P t(j) in the wholesale sector:

$$\displaystyle \begin{aligned} s_t^y = (1-\varphi_y)p_{At}^{-\epsilon_y} + \varphi_y (\pi_{t-1}/\pi_t)^{-\epsilon_y}s_{t-1}^y. \end{aligned} $$
(4.81f)

Appendix 4.4: Microfoundations of Wage Setting

Consider the real wage W t(h)∕P t of a household member h who has set his wage optimally in period t to \(\tilde w_t=W_{At}/P_t\) and who has not been able to do so again until period s. Between period t and t + s, the nominal wage of the household member h increases with the lagged inflation rate π t, π t+1, …, π t+s−1 in each period t + 1, t + 2, …, t + s, while the aggregate price level P t increases with the inflation rates π t+1, π t+2, …, π t+s:

$$\displaystyle \begin{aligned}\frac{W_{Nt+s}}{P_{t+s}}=\frac{\prod_{i=1}^s \pi_{t+i-1}W_{At}}{\prod_{i=1}^s \pi_{t+i}P_t}=\frac{\pi_t}{\pi_{t+s}}\tilde w_t.\end{aligned}$$

The demand for his type of labor service follows from (4.58)

$$\displaystyle \begin{aligned}L_{t+s}(h)=\left(\frac{(\pi_{t}/\pi_{t+s})\tilde w_t}{w_{t+s}}\right)^{-\epsilon_w} L_{t+s},\end{aligned} $$

where w t+s = W t+sP t+s denotes the real wage prevailing in period t + s. Accordingly, the Lagrangian for the optimal real wage is represented by:

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} \mathscr{L} &=& \mathbb{E}_t \sum_{s=0}^\infty (\beta\varphi_w)^s&\Bigg\{ \frac{(C_{t+s}(h)-\chi \bar C_{t+s})^{1-\sigma}-1}{1-\sigma}\\ &&& - \frac{\nu_0}{1+\frac{1}{\nu_1}}\left[\left(\frac{(\pi_t/\pi_{t+s})\tilde w_{t}}{w_{t+s}}\right)^{-\epsilon_w} L_{t+s}\right]^{1+\frac{1}{\nu_1}}\\ &&&+ \gamma^M_0 \frac{\left(\frac{M_{t+1}(h)}{P_t}\right)^{1-\gamma^M_1}-1}{1-\gamma^M_1}\\ &&&+\varLambda_{ht+s}\left[ \frac{\pi_t}{\pi_{t+s}}\tilde w_t\left(\frac{(\pi_t/\pi_{t+s})\tilde w_{t}}{w_{t+s}}\right)^{-\epsilon_w} L_{t+s} + RMT \right]\Bigg\},\vspace{-2pt}\end{array}\end{aligned} $$

where (φ w)s denotes the probability that the household cannot adjust its wage optimally for s periods and RMT t is a placeholder for the remaining terms of the household budget constraint (4.61).

The first-order condition with respect to \(\tilde w_t\) is represented by

$$\displaystyle \begin{aligned}\begin{array}{r*{20}l} 0 &=\mathbb{E}_t \sum_{s=0}^\infty (\beta\varphi_w)^s\Bigg\{\epsilon_w \nu_0 \tilde w_t^{-\epsilon_w(1+\frac{1}{\nu_1})-1}\left(\frac{(\pi_t/\pi_{t+s})}{w_{t+s}}\right)^{-\epsilon_{w}(1+\frac{1}{\nu_1})} L_{t+s}^{1+\frac{1}{\nu_1}}\\ &\quad \quad +(1-\epsilon_w)\varLambda_{ht+s}\tilde w_t^{-\epsilon_w}w_{t+s}^{\epsilon_w}\left(\frac{\pi_t}{\pi_{t+s}}\right)^{1-\epsilon_w} L_{t+s}\Bigg\}. \vspace{-2pt}\end{array}\end{aligned} $$

Using Λ ht+s = Λ t+s, this equation can be arranged to read as

$$\displaystyle \begin{aligned} \tilde w_t &=\frac{\epsilon_w}{\epsilon_w-1}\frac{\varDelta_{1t}}{\varDelta_{2t}}, \end{aligned} $$
(4.82a)

where

$$\displaystyle \begin{aligned} \notag\varDelta_{1t} &=\nu_0 \mathbb{E}_t \sum_{s=0}^\infty (\beta\varphi_w)^s \left(\frac{\pi_t \tilde w_t}{\pi_{t+s}w_{t+s}}\right)^{-\epsilon_w(1+\frac{1}{\nu_1})} L_{t+s}^{1+\frac{1}{\nu_1}}, \end{aligned} $$
$$\displaystyle \begin{aligned} &=\nu_0\left(\frac{\tilde w_t}{w_t}\right)^{-\epsilon_w(1+\frac{1}{\nu_1})}L_t^{1+\frac{1}{\nu_1}} + (\beta\varphi_w)\mathbb{E}_t \left(\frac{\pi_t\tilde w_t}{ \pi_{t+1}\tilde w_{t+1}}\right)^{-\epsilon_w(1+\frac{1}{\nu_1})}\varDelta_{1t+1}, \end{aligned} $$
(4.82b)
$$\displaystyle \begin{aligned} \notag\varDelta_{2t} &=\mathbb{E}_t \sum_{s=0}^\infty (\beta\varphi_w)^s \varLambda_{t+s}\left(\frac{\tilde w_t}{w_{t+s}}\right)^{-\epsilon_w}\left(\frac{\pi_t}{\pi_{t+s}}\right)^{1-\epsilon_w} L_{t+s}, \end{aligned} $$
$$\displaystyle \begin{aligned} &=\varLambda_t \left(\frac{\tilde w_t}{w_t}\right)^{-\epsilon_w}L_t + (\beta\varphi_w)\mathbb{E}_t \left(\frac{\tilde w_t}{\tilde w_{t+1}}\right)^{-\epsilon_w}\left(\frac{ \pi_t}{\pi_{t+1}}\right)^{1-\epsilon_w}\varDelta_{2t+1}. \end{aligned} $$
(4.82c)

Again, Δ 1t and Δ 2t are simply auxiliary variables whose behaviors are described by (stochastic) first-order difference equations. Therefore, they are easily amenable to the solution with the linearization methods described in Appendix 2.3.

The wage index (4.59) implies

$$\displaystyle \begin{aligned}W_t^{1-\epsilon_w}=(1-\varphi_w)W_{At}^{1-\epsilon_w}+\varphi_w (\pi_{t-1}W_{t- 1})^{1-\epsilon_w}\end{aligned}$$

and thus, the real wage equals

$$\displaystyle \begin{aligned} w_t^{1-\epsilon_w} &= (1-\varphi_w)\tilde w_t^{1-\epsilon_w}+\varphi_w\left(\frac{\pi_{t-1}}{\pi_t}w_{t-1}\right)^{1-\epsilon_w}. \end{aligned} $$
(4.82d)

Finally, consider the index

$$\displaystyle \begin{aligned}\tilde L_t^{1+\frac{1}{\nu_1}} = \int_0^1 L_{t}(h)^{1+\frac{1}{\nu_1}}dh,\end{aligned}$$

in the families of current-period utility functions. Using labor demand function (4.58), this index can be re-written as

$$\displaystyle \begin{aligned}\tilde L_t^{1+\frac{1}{\nu_1}}=L_t^{1+\frac{1}{\nu_1}}\int_0^1 \left(\frac{W_{t}(h)}{W_t}\right)^{-\epsilon_w(1+\frac{1}{\nu_1})}dh.\end{aligned}$$

Therefore,

$$\displaystyle \begin{aligned}\tilde L_t &= s^w_t L_t.\end{aligned} $$
(4.82e)

implies the definition of the wage dispersion measure (4.63).

Next, we need to derive the dynamics of the wage dispersion measure \(s^w_t\). For this reason, consider

$$\displaystyle \begin{aligned} \begin{aligned} \bar W_t^{-\epsilon_w(1+\frac{1}{\nu_1})}&=\int_0^1 W_{t}(h)^{-\epsilon_w(1+\frac{1}{\nu_1})}dh\\ &=(1-\varphi_w)(W_{At})^{-\epsilon_w(1+\frac{1}{\nu_1})}+\varphi_w (\pi_{t-1}W_{Nt-1})^{-\epsilon_w(1+\frac{1}{\nu_1})}. \end{aligned} \end{aligned}$$

Accordingly, wage dispersion \(s_t^w\) is described by the following equation:

$$\displaystyle \begin{aligned}(s^w_t)^{1+\frac{1}{\nu_1}}=\left(\frac{\bar W_t}{W_t}\right)^{-\epsilon_w(1+\frac{1}{\nu_1})}=\left(\frac{\bar W_t/P_t}{W_t/P_t}\right)^{-\epsilon_w(1+\frac{1}{\nu_1})} =\left(\frac{\bar w_t}{w_t}\right)^{-\epsilon_w(1+\frac{1}{\nu_1})}.\end{aligned}$$

Using the same line of argument employed to derive (4.81f) yields the dynamic equation for the measure of wage dispersion \(s^w_t\):

$$\displaystyle \begin{aligned} (s_t^w)^{1+\frac{1}{\nu_1}} = (1-\varphi_w)\left(\frac{\tilde w_{t}}{w_t}\right)^{-\epsilon_w(1+\frac{1}{\nu_1})} + \varphi_w\left(\frac{\pi_{t-1}w_{t-1}}{\pi_t w_t}\right)^{-\epsilon_w(1+\frac{1}{\nu_1})}(s^w_{t-1})^{1+\frac{1}{\nu_1}}.\end{aligned} $$
(4.82f)

Appendix 4.5: Monetary Policy Analysis in the New Keynesian Model

Figure 4.25 illustrates the impulse response functions to an interest rate shock of one percentage point in the benchmark New Keynesian model with ϕ = 1.0. In accordance with empirical observations, restrictive monetary policy decreases all private demand, investment and private consumption; moreover, output and labor decline. As expected, inflation also decreases under a restrictive monetary policy.

Fig. 4.25
figure 25

Impulse responses to an interest rate shock

The reasons are as follows: Following an increase in the nominal interest rate Q t, prices only adjust slowly so that also the real interest rate in the economy increases. As a consequence, firms reduce their investment, and Tobin’s q decreases. In addition, households postpone consumption to later periods in accordance with their Euler condition (4.66c). Therefore, demand decreases, and prices fall. Since wholesale producers are slow to adjust their prices, the mark-up increases.

Appendix 4.6: Data Sources

The time series on government expenditures that we use in this chapter are attached as separate Excel files to my Matlab/Gauss programs.

Problems

4.1

Solve the transition dynamics for the Numerical Problem in Sect. 4.3 by forward shooting as described in Appendix 4.1.

4.2

Derive the Frisch labor supply elasticity (4.39) of the Cobb-Douglas utility function (4.28).

4.3

In applied work, researchers often select model parameters that are not empirically observable or only estimated with a high degree of uncertainty by optimizing the behavior of the RBC model. Using the model in Sect. 4.4, use a grid search over ϕ ∈ [0, 1.5] and ρ C ∈ [0.3, 1.3] to find the minimum distance (i.e., the minimum of the squared deviations) of the theoretical second moments (as implied by the model) from the empirical second moments (as presented in Table 4.1).

4.4

Use the preferences

$$\displaystyle \begin{aligned} u(C^p,L)=\frac{(C^p)^{1-\sigma} (1-L)^{1+\vartheta}}{1-\sigma},\;\; \sigma>1, \vartheta>0,\end{aligned}$$

and recompute the RBC model with stochastic government. Set σ = 2.0, and calibrate 𝜗 such that steady-state labor supply is equal to L = 0.3. Does private consumption increase after an increase in government consumption?

4.5

Derive (4.72) using the individual budget constraint, the government budget, and the firms’ first-order equations. Apply Euler’s theorem according to which the aggregate output is equal to the sum of all factor payments for a constant-returns-to-scale technology under perfect competition.

4.6

Derive the equilibrium dynamics (4.73 ) for the New Keynesian model in Sect. 4.5. Assume, however, that firms in the wholesale sector who cannot optimally choose their price adjust their nominal price P NT according to the average inflation rate:

$$\displaystyle \begin{aligned} P_{Nt}= \pi P_{Nt-1}.\end{aligned}$$

Rights and permissions

Reprints and permissions

Copyright information

© 2019 Springer Nature Switzerland AG

About this chapter

Check for updates. Verify currency and authenticity via CrossMark

Cite this chapter

Heer, B. (2019). Government Consumption. In: Public Economics. Springer Texts in Business and Economics. Springer, Cham. https://doi.org/10.1007/978-3-030-00989-2_4

Download citation

Publish with us

Policies and ethics