Skip to main content

Advertisement

Log in

Monetary policy and dark corners in a stylized agent-based model

  • Regular Article
  • Published:
Journal of Economic Interaction and Coordination Aims and scope Submit manuscript

Abstract

We extend in a minimal way the stylized macroeconomic Agent-Based model introduced in our previous paper (Gualdi et al. in J Econ Dyn Control 50:29–61, 2015a), with the aim of investigating the role and efficacy of monetary policy of a ‘Central Bank’ that sets the interest rate such as to steer the economy towards a prescribed inflation and employment rate. Our major finding is that provided its policy is not too aggressive (in a sense detailed in the paper) the Central Bank is successful in achieving its goals. However, the existence of different equilibrium states of the economy, separated by phase boundaries (or “dark corners”), can cause the monetary policy itself to trigger instabilities and be counter-productive. In other words, the Central Bank must navigate in a narrow window: too little is not enough, too much leads to instabilities and wildly oscillating economies. This conclusion strongly contrasts with the prediction of DSGE models.

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

Access this article

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

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7

Similar content being viewed by others

Notes

  1. The Mark I family of models was elaborated in a series of papers and books.

  2. \(\zeta _i\) is the productivity of firm i. We chose \(\zeta _i = 1\) in Gualdi et al. (2015a) and we will stick to this choice throughout the present paper as well.

  3. As a consequence of these adaptive adjustments, the economy can reach (in some regions of the parameter space) equilibrium, corresponding to the market clearing condition one would obtain in a fully representative agent framework. However, fluctuations around equilibrium persists in the limit of large system sizes giving rise to a rich phenomenology, see Gaffeo et al. (2008), Delli Gatti et al. (2008), Delli Gatti et al. (2011), Gualdi et al. (2015a).

  4. In Galì’s (2008) reference book, the quantity \(\phi _\varepsilon \) is noted \(\phi _y\).

  5. We chose \(\omega =0.2\), which corresponds to an averaging time of about \(-1/\log {(1-\omega _\tau )}\approx 4.5\) time steps.

  6. Note that the parameter f is similar to, but different from the parameter also called f in Gualdi et al. (2015a), which was used to share the cost of defaults on firms and households.

  7. To be more precise, in the default resolution described in Gualdi et al. (2015a) we introduce a bailout probability, called f there, which sets the relative impact of default costs on households and firms savings. In this sense, the present setting recovers the one in Gualdi et al. (2015a) with \(f \approx 1/2\).

  8. This value of \(\alpha _\Gamma \) has the following interpretation: when the debt of a firm equals its payroll, i.e. when \(\Phi _i = 1\), a real interest rate of \(2\,\%\) monthly leads to a freezing of all hires and a doubling of the firing rate, compared to a zero-debt situation.

  9. A higher wage translates in the availability of a larger share of unemployed workers in the hiring process.

References

  • Assenza T et al (2013) Individual expectations and aggregate macro behavior, Tinbergen Institute Discussion Paper 13–016/II

  • Adam K, Woodford M (2012) Robustly optimal monetary policy in a microfounded New Keynesian model. J Monet Econ 59(5):468–487

    Article  Google Scholar 

  • Anand K, Gai P, Marsili M (2010) The rise and fall of trust networks, Progress in Artificial Economics. Lecture Notes in Economics and Mathematical Systems 645:77

  • Anderson SP, De Palma A, Thisse JF (1992) Discrete choice theory of product differentiation. MIT Press, New York

    Google Scholar 

  • Bianchi C et al (2007) Validating and calibrating agent-based models: a case study. Comput Econ 30(3):245–264

    Article  Google Scholar 

  • Blanchard O (2014) Where danger lurks. Finance Dev 51(3):28–31

  • Bouchaud J-P (2013) Crises and collective socio-economic phenomena: simple models and challenges. J Stat Phys 151:567–606, and refs. therein

  • Buchanan M (2008) This Economy does not compute, New York Times, October 1, 2008. On this topic, see also Gualdi et al. (2015), Kirman (2010), Bouchaud J-P (2013), Farmer and Foley (2009)

  • Caballero RJ (2010) Macroeconomics after the crisis: time to deal with the pretense-of-knowledge syndrome. J Econ Perspect 24:85–102

    Article  Google Scholar 

  • Chakraborti A, Challet D, Chatterjee A, Marsili M, Zhang Y-C, Chakrabarti BK (2015) Statistical mechanics of competitive resource allocation using agent-based models. Phys Rep 552:1–25

    Article  Google Scholar 

  • Christiano Lawrence J, Eichenbaum Martin, Evans Charles L (2005) Nominal rigidities and the dynamic effects of a shock to monetary policy. J Polit Econ 113:1–45

    Article  Google Scholar 

  • da Gama Batista J, Bouchaud J-P, Challet D (2015) Sudden trust collapse in networked societies. Eur Phys J B 88(3):1–11. doi:10.1140/epjb/e2015-50645-1

  • Dawid H, Fagiolo G (2008) Agent-based models for economic policy design: introduction to the special issue. J Econ Behav Organ 67:351–354

  • Dawid H, Gemkow S, Harting P, van der Hoog S, Neugarty M (2011) The Eurace@Unibi Model: an agent-based macroeconomic model for economic policy analysis, working paper, Bielefeld University

  • De Nardi M, French E, Benson D (2012) Consumption and the great recession. Mimeo, Federal Reserve Bank of Chicago

  • Deissenberg C, van der Hoog S, Dawid H (2008) Eurace: a massively parallel agent-based model of the european economy. Appl Math Comput 204:541–552

    Google Scholar 

  • Delli Gatti D, Gaffeo E, Gallegati M, Giulioni G, Palestrini A (2008) Emergent macroeconomics: an agent-based approach to business fluctuations. Springer, Berlin

    Google Scholar 

  • Delli Gatti D, Desiderio S, Gaffeo E, Cirillo P, Gallegati M (2011) Macroeconomics from the bottom-up. Springer, Berlin

    Book  Google Scholar 

  • Dosi G, Fagiolo G, Roventini A (2005) An evolutionary model of endogenous business cycles. Comput Econ 27:3–24 and Fagiolo and Roventini (2012)

    Article  Google Scholar 

  • Dosi G, Fagiolo G, Roventini A (2010) Schumpeter meeting Keynes: a policyfriendly model of endogenous growth and business cycles. J Econ Dyn Control 34:1748–1767 and Fagiolo and Roventini (2012)

    Article  Google Scholar 

  • Dosi G et al (2013) Income distribution, credit and fiscal policies in an agent-based Keynesian model. J Econ Dyn Control 37(8):1598–1625

    Article  Google Scholar 

  • Dosi G et al (2015) Fiscal and monetary policies in complex evolving economies. J Econ Dyn Control 52:166–189

    Article  Google Scholar 

  • Ehrhardt G, Marsili M, Vega-Redondo F (2006) Phenomenological models of socioeconomic network dynamics. Phys Rev E 74:036106

    Article  Google Scholar 

  • Fagiolo G et al (2006) Empirical validation of agent-based models: a critical survey, No. 2006/14. LEM Working Paper Series

  • Fagiolo G, Roventini A (2012) Macroeconomic policy in DSGE and agent-based models. Working Paper

  • Farmer D, Foley D (2009) The economy needs agent-based modelling. Nature 460:685–686

    Article  Google Scholar 

  • Gaffeo E, Delli Gatti D, Desiderio S, Gallegati M (2008) Adaptive microfoundations for emergent macroeconomics. East Econ J 34:441–463

    Article  Google Scholar 

  • Gai P, Haldane A, Kapadia S (2011) Complexity, concentration and contagion. J Monet Econ 58:453–470

    Article  Google Scholar 

  • Galí J (2008) Monetary policy, inflation and the business cycle. Princeton University Press, Princeton

    Google Scholar 

  • Gross DB, Souleles NS (2002) Do liquidity constraints and interest rates matter for consumer behavior? Evidence from credit card data. Q J Econ 107:149–185

    Article  Google Scholar 

  • Gualdi S, Tarzia M, Zamponi F, Bouchaud J-P (2015a) Tipping points in macroeconomic agent-based models. J Econ Dyn Control 50:29–61

  • Gualdi S, Bouchaud J-P, Cencetti G, Tarzia M, Zamponi F (2015b) Endogenous crisis waves: a stochastic model with synchronized collective behavior. Phys Rev Lett 114:088701. arXiv:1409.3296v1

  • Haber G (2008) Monetary and fiscal policy analysis with an agent-based macroeconomic model. Jahrbücher für Nationalökonomie und Statistik, pp 276–295

  • Hommes C (2013) Behavioral rationality and heterogeneous expectations in complex economic systems. Cambridge University Press, for a review on behavioral experiments, and Assenza et al. (2013) for experiments related to DSGE modelling

  • Kirman A (2010) The economic crisis is a crisis for economic theory. CESifo Econ Stud 56:498–535

    Article  Google Scholar 

  • Kothari SP, Lewellen J, Warner JB (2014) The behavior of aggregate corporate investment. http://ssrn.com/abstract=2511268

  • LeBaron B, Tesfatsion L (2008) Modeling macroeconomies as open-ended dynamic systems of interacting agents. Am Econ Rev 98(2):246–250

  • Lein SM (2010) When do firms adjust prices? Evidence from micro panel data. J Monet Econ 57:696–715

    Article  Google Scholar 

  • Lengnick M, Wohltmann H-W (2013) Agent-based financial markets and New Keynesian macroeconomics: a synthesis. J Econ Interact Coord 8(1):1–32

    Article  Google Scholar 

  • Ludvigson SC (2004) Consumer confidence and consumer spending. J Econ Perspect 18:29–50

    Article  Google Scholar 

  • Mandel A, Furst S, Lass W, Meissner F, Jaeger C (2009) Lagom generiC: an agent-based model of growing economies, ECF Working Paper 1 /2009

  • Mankiw NG (2014) Principles of macroeconomics, Cengage Learning, see in particular Ch. 14

  • Marsili M (2014) Complexity and financial stability in a large random economy. Quant Finance 14:1663–1675

    Article  Google Scholar 

  • Patzelt F, Pawelzik K (2011) Criticality of adaptive control dynamics. Phys Rev Lett 107:238103

    Article  Google Scholar 

  • Schelling T (1978) Micromotives and macrobehaviour. W.W. Norton & Co Ltd, New York

  • Souleles NS (2004) Expectations, heterogeneous forecast errors, and consumption: micro evidence from the Michigan consumer sentiment surveys. J Money Credit Bank 36:39–72

    Article  Google Scholar 

  • Taleb NN (2007) The Black Swan. Random House, London

    Google Scholar 

  • Tesfatsion L (2006) Agent-based computational modeling and macroeconomics. Post Walrasian macroeconomics: beyond the dynamic stochastic general equilibrium model

  • van der Hoog S, Dawid H (2015) Bubbles, crashes and the financial cycle: insights from a stock-flow consistent agent-based macroeconomic model, University of Bielefeld working paper, http://ssrn.com/abstract=2616662

Download references

Acknowledgments

This work was partially financed by the CRISIS project. We want to thank all the members of CRISIS for most useful discussions, in particular during the CRISIS meetings. The input and comments of T. Assenza, J. Batista, E. Beinhocker, D. Challet, D. Delli Gatti, D. Farmer, J. Grazzini, C. Hommes, F. Lillo, G. Tedeschi; S. Battiston, A. Kirman, A. Mandel, M. Marsili and A. Roventini are warmly acknowledged. JPB wants to sincerely thank J.-C. Trichet and A. Haldane for very encouraging comments on this endeavor. Finally, we thank our two referees for insightful and constructive remarks that helped improving the manuscript.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Stanislao Gualdi.

Appendices

Appendix 1: Price, production and wage updates

1.1 The households timeline

  • At the beginning of the time step households are characterized by a certain amount of savings \(S(t)\ge 0\).

  • After each firm chooses its production \(Y_i(t)\), price \(P_i(t)\) and wage \(W_i(t)\) for the current time step (see later) wages are paid. Since firms use a one-to-one linear technology taking only labor as input and productivity is set to 1, the production equal the workforce of the firm. Hence, the total amount of wages paid is given by

    $$\begin{aligned} W_T(t) = \sum _i W_i(t)Y_i(t) \end{aligned}$$
    (14)
  • Once the total payroll of the economy is determined, interests on deposits are paid and households set a consumption budget as

    $$\begin{aligned} C_B(t) = c(t) [ S(t) + W_T(t) + \rho ^{\text {d}}(t)S(t)] \end{aligned}$$
    (15)

    where \(c(t)\in [0,1]\) is the propensity to consume and may depend on inflation/interests on deposits, see Eq. (9).

  • The consumption budget is distributed among firms using an intensity of choice model (Anderson et al. 1992). The demand of goods for firm i is therefore:

    $$\begin{aligned} D_i(t) = \frac{C_B(t)}{p_i(t)} \frac{e^{- \beta p_i(t) / \overline{p}(t)}}{\sum \nolimits _i e^{-\beta p_i(t) / \overline{p}(t)}} \ , \end{aligned}$$
    (16)

    where \(\beta \) is the price sensitivity parameter determining an exponential dependence of households demand to the price offered by the firm. Indeed, \(\beta = 0\) corresponds to complete price insensitivity and \(\beta \rightarrow \infty \) means that households select only the firm with the lowest price. In this sense, as long as \(\beta >0\) firms compete on prices.

  • The actual consumption C(t) (limited by production) is given by

    $$\begin{aligned} C(t){:=} \sum _{i=1}^{N_\mathrm{F}}p_i(t)\min {\{Y_i(t),D_i(t)\}}\le C_B(t) = \sum _{i=1}^{N_\mathrm{F}}p_i(t)D_i(t) \end{aligned}$$
    (17)

    and households accounting therefore reads

    $$\begin{aligned} S(t+1) = [1+\rho ^{\text {d}}(t)]S(t) + W_T(t) - \sum _i p_i(t)\min {\{D_i,Y_i\}} + \Delta (t) \end{aligned}$$
    (18)

    where \(\Delta (t)\) are dividends paid (see below for a definition of this last term).

Fig. 8
figure 8

Scatter plot in one time step of firm profits versus the price offered. The red line correspond to a moving flat average of 100 consecutive points. Parameters are: \(R=2\) (with \(\eta ^0_-=0.1\)), \(\Theta =2\), \(\gamma _p=\gamma _w=0.05\), \(\alpha _\Gamma =50\), \(\Gamma _0=0\), \(\beta =2\), \(\alpha _c=4\), \(N=5000\) and \(\rho ^{*}=0.5\,\%\)

1.2 The firms timeline

  • At the beginning of the time step t firms with \(\Phi _i(t) \ge \Theta \) become inactive and are removed from the simulation. Default costs are computed as

    $$\begin{aligned} {\mathcal {D}}(t) = -\sum _{i\ \mathrm{bankrupt}}{\mathcal {E}}_i(t) \ . \end{aligned}$$
    (19)

    Firms with \(\Phi _i(t) < \Theta \) are instead allowed to continue their activity and contribute to total loans \({\mathcal {E}}^-(t)\) and total firms savings \({\mathcal {E}}^+(t)\) as

    $$\begin{aligned} {\mathcal {E}}^-= & {} -\sum _{i\ \mathrm{not \ bankrupt}}\min {\{{\mathcal {E}}_i(t),0\}}\nonumber \\ {\mathcal {E}}^+= & {} \sum _{i\ \mathrm{not \ bankrupt}}\max {\{{\mathcal {E}}_i(t),0\}}. \end{aligned}$$
    (20)
  • Active firms set production, price and wage for the current time step following simple adaptive rules which are meant to represent an heuristic adjustment. In particular:

    • Price: Prices are updated through a random multiplicative process which takes into account the production-demand gap experienced in the previous time step and if the price offered is competitive (with respect to the average price). The update rule for prices reads:

      $$\begin{aligned} \text {If } Y_i(t)< D_i(t)\Rightarrow & {} {\left\{ \begin{array}{ll} &{} \text {If } p_i(t) < \overline{p}(t) \Rightarrow p_i(t+1) = p_i(t) (1 + \gamma _p \xi _i(t) ) \\ &{} \text {If } p_i(t) \ge \overline{p}(t) \Rightarrow p_i(t+1) = p_i(t) \\ \end{array}\right. }\nonumber \\ \text {If } Y_i(t)> D_i(t)\Rightarrow & {} {\left\{ \begin{array}{ll} &{} \text {If } p_i(t) > \overline{p}(t) \Rightarrow p_i(t+1) = p_i(t) (1 - \gamma _p \xi _i(t) ) \\ &{} \text {If } p_i(t) \le \overline{p}(t) \Rightarrow p_i(t+1) = p_i(t)\\ \end{array}\right. }\nonumber \\ \end{aligned}$$
      (21)

      where \(\xi _i(t)\) are independent uniform U[0, 1] random variables and \(\gamma _p\) is a parameter setting the relative magnitude of the price adjustment (we set it to \(5\,\%\) unless stated otherwise). Fig. 8, which plots the average profit of firms as a function of the offered price, shows that these rules lead to a reasonable emergent “optimizing” behavior of firms. As expected, the profit reaches a maximum for prices slightly above the average price \(\overline{p}(t)\). Higher prices are not competitive and firms lose clients, while lower prices do not cover production costs.

    • Production: Independently of their price level, firms try to adjust their production to the observed demand. When firms want to hire, they open positions on the job market; we assume that the total number of unemployed workers, which is \(N_\mathrm{F} u(t)\), is distributed among firms according to an intensity of choice model which depends on both the wage offered by the firmFootnote 9 and on the same parameter \(\beta \) as it is for Eq. (16); therefore the maximum number of available workers to each firm is:

      $$\begin{aligned} u^*_i(t) = \frac{e^{\beta W_i(t) / \overline{w}(t)}}{\sum \nolimits _i e^{\beta W_i(t) / \overline{w}(t)}} N_\mathrm{F} u(t) \ . \end{aligned}$$
      (22)

      The production update is then defined as:

      $$\begin{aligned} \begin{aligned} \text {If } Y_i(t) < D_i(t)&\Rightarrow Y_i(t+1)=Y_i(t)+ \min \{ \eta ^+_i ( D_i(t)-Y_i(t)), u^*_i(t) \} \\ \text {If } Y_i(t) > D_i(t)&\Rightarrow Y_i(t+1)= Y_i(t) - \eta ^-_i [Y_i(t)-D_i(t)] \\ \end{aligned} \end{aligned}$$
      (23)

      where \(\eta ^\pm \in [0,1]\) are what we denote as the hiring/firing propensity of the firms. According to this mechanism, the change in output responds to excess demand (there is an increase in output if excess demand is positive, a decrease in output if excess demand is negative, i.e. if there is excess supply). The propensities to hire/fire \(\eta _\pm \) are the sensitivity of the output change to excess demand/supply.

    • Wage: The wage update rule we chose follows (in spirit) the choices made for price and production. We propose that at each time step firm i updates its wage as:

      $$\begin{aligned} \begin{aligned} W^T_i(t+1)=W_i(t)[1+\gamma _w (1 - \Gamma \Phi _i) \varepsilon \xi ^\prime _i(t)] \quad \hbox {if}\quad {\left\{ \begin{array}{ll} Y_i(t) &{}< D_i(t)\\ {\mathcal {P}}_i(t) &{}> 0 \end{array}\right. }\\ W_i(t+1)=W_i(t)[1-\gamma _w (1 + \Gamma \Phi _i) u \xi ^\prime _i(t)] \quad \hbox {if}\quad {\left\{ \begin{array}{ll} Y_i(t) &{}> D_i(t)\\ {\mathcal {P}}_i(t) &{}< 0 \end{array}\right. } \end{aligned} \end{aligned}$$
      (24)

      where \(\gamma _w\) is a certain parameter, \({\mathcal {P}}_i(t)\) is the profit of the firm at time t and \(\xi ^\prime _i(t)\) an independent U[0, 1] random variable. If \(W^T_i(t+1)\) is such that the profit of firm i at time t with this amount of wages would have been negative, \(W_i(t+1)\) is chosen to be exactly at the equilibrium point where \({\mathcal {P}}_i(t)=0\) otherwise \(W_i(t+1) = W^T_i(t+1)\).

      The above rules are intuitive: if a firm makes a profit and it has a large demand for its good, it will increase the pay of its workers. The pay rise is expected to be large if the firm is financially healthy and/or if unemployment is low (i.e. if \(\varepsilon \) is large) because pressure on salaries is high. Conversely, if the firm makes a loss and has a low demand for its good, it will reduce the wages. This reduction is drastic if the company is close to bankruptcy, and/or if unemployment is high, because the pressure on salaries is then low. In all other cases, wages are not updated.

      The parameters \(\gamma _{p,w}\) allow us to simulate different price/wage update timescales, i.e. the aggressivity with which firms react to a change of their economic conditions. In the following we set \(\gamma _p=0.05\) and \(\gamma _w= \gamma _p\). The case \(\gamma _w=0\) corresponds to removing completely the wage update rule, such that the version of the model with constant wage is recovered.

  • After prices, productions and wages are set and interests paid, consumption and accounting take place. Since each firm has total sales \(p_i\min {\{Y_i,D_i\}}\) firms profits are

    $$\begin{aligned} {\mathcal {P}}_i(t)= & {} p_i(t)\min {\{Y_i(t),D_i(t)\}} - W_i(t)Y_i(t)\nonumber \\&+\, \rho ^{\text {d}}\max {\{{\mathcal {E}}_i(t),0\}}+ \rho ^{\ell }\min {\{{\mathcal {E}}_i(t),0\}} \ . \end{aligned}$$
    (25)

    When firms have both positive \({\mathcal {E}}_i\) and \({\mathcal {P}}_i\) dividends are paid as a fraction \(\delta \) of the firm cash balance \({\mathcal {E}}_i\). The update rule for firms cash balance is therefore

    $$\begin{aligned} {\mathcal {E}}_i(t+1) = {\mathcal {E}}_i(t) + {\mathcal {P}}_i(t) - \delta {\mathcal {E}}_i(t)\theta ({\mathcal {P}}_i(t))\theta ({\mathcal {E}}_i(t)) \end{aligned}$$
    (26)

    where \(\theta (x)=1\) if \(x>0\) and 0 otherwise. Correspondingly, households savings are updated as

    $$\begin{aligned} S(t+1)= & {} S(t) + \sum _iW_i(t)Y_i(t) - \sum _i p_i(t)\min {\{Y_i(t),D_i(t)\}}\nonumber \\&+\, \delta \sum _i{\mathcal {E}}_i(t)\theta ({\mathcal {P}}_i(t))\theta ({\mathcal {E}}_i(t)). \end{aligned}$$
    (27)

    The dividends share \(\delta \) is set to \(2\,\%\) unless stated otherwise and the \(\Delta (t)\) term in Eq. (18) is given by

    $$\begin{aligned} \Delta (t) = \delta \sum _i{\mathcal {E}}_i(t)\theta ({\mathcal {P}}_i(t))\theta ({\mathcal {E}}_i(t)) \end{aligned}$$
    (28)
  • Finally, an inactive firm has a finite probability \(\varphi \) (which we set to 0.1) per unit time to get revived; when it does so its price is fixed to \(p_i(t) = \overline{p}(t)\), its wage to \(W_i(t) = \overline{w}(t)\), its workforce is the available workforce \(Y_i(t) = u(t)\) and its cash-balance is the amount needed to pay the wage bill \({\mathcal {E}}_i(t) = W_i(t) Y_i(t)\). This small “liquidity” is provided by firms with positive \({\mathcal {E}}_i\) in shares proportional to their wealth \({\mathcal {E}}_i\).

Appendix 2: Firms’ adaptive behavior leads to a second order phase transition

We start by analysing the effect of adaptation of firms. In order to get a first insight it is useful to consider a simplified setting where \(\Gamma =\Gamma _0\) (i.e. \(\alpha _\Gamma =0\)), \(\rho ^{\ell }(t)=\rho _0(t)=0\), \(f=1\), \(c(t)=c_0=0.5\) (hence \(\alpha _c=0\)) and wages are constant and equal to 1 (\(\gamma _w=0\)). In this case the basic model described in Gualdi et al. (2015a) (with constant wages) is recovered.

Intuitively, the coupling between financial fragility and hiring/firing propensity should have a stabilizing effect on the economy. Moreover, the full unemployment phase at \(R< R_c\) is deeply affected by the presence of \(\alpha _\Gamma \): for \(\alpha _\Gamma \ne 0\) the unemployment rate in this phase is no longer one, but becomes smaller than one and continuously changing with R. In order to derive an estimate of these continuous values we use an intuitive argument (at \(\Theta =\infty \)) which is justified a posteriori by the good match with numerical results. Given the critical ratio \(R = \eta ^0_+/\eta ^0_- = R_c\) separating the high/low unemployment phases when there is no adaptation (i.e. \(\Gamma _0=0\)) one can expect that equilibrium values of the unemployment rate different from 0 and 1 can only be stable if \(\eta _+^{i}/\eta _{-}^{i}\) remains around the critical value \(R_c\) at \(\Gamma _0=0\). Near criticality therefore we enforce that:

$$\begin{aligned} \frac{\eta _+^i}{\eta _-^i} = \frac{\eta ^0_+ (1 - \Gamma _0 \Phi _i)}{\eta ^0_- (1 + \Gamma _0 \Phi _i) } = R_c\Rightarrow -\Gamma _0 \Phi \approx \frac{R_c\eta _-^0-\eta _+^0}{R_c\eta _-^0+\eta _+^0}. \end{aligned}$$
(29)

An explicit form of \(\Phi \) in terms of the employment rate \(\varepsilon ={\overline{Y}}\) can be obtained with the additional assumption that the system is always close to equilibrium (i.e. \(p\approx 1\) and \(D \approx Y\), at least when \(\eta ^i_+/\eta ^i_-\sim R_c\)), which allows one to express households savings in terms of the firms’ production. Indeed (see the discussion in Gualdi et al. 2015a) at equilibrium \(W = C_B = N_\mathrm{F} Y = c(W + S)\), from which it follows that \(N_\mathrm{F} Y = W = S c/(1-c)\). For \(c=0.5\) as in our simulations one thus has \(S=N_\mathrm{F} Y\). Since the total amount of money is conserved (in our simulations \(N_\mathrm{F} \overline{{\mathcal {E}}} + S = N_\mathrm{F}\overline{{\mathcal {E}}} + N_\mathrm{F}Y= N_\mathrm{F}\), see Appendix 3) one finally obtains that \(\overline{{\mathcal {L}}} = 1 - Y\) and \(\Phi = (Y-1)/Y = (\varepsilon -1)/\varepsilon \), hence

$$\begin{aligned} \frac{\Gamma _0}{\varepsilon } = \frac{R_c\eta _-^0-\eta _+^0}{R_c\eta _-^0+\eta _+^0} + \Gamma _0 =\frac{R_c- R}{R_c+ R} + \Gamma _0\ . \end{aligned}$$
(30)

Note that according to this formula the employment goes to \(\varepsilon =1\) at the critical point \(R = R_c\). Above this value, the economy is in the “good” state and employment sticks to \(\varepsilon =1\) (this is because in the argument the effect of \(\Theta \) has been neglected). Moreover, when \(R < R_c\), \(\varepsilon \) is proportional to \(\Gamma _0\) and therefore in the limit \(\Gamma _0\rightarrow 0\) one has \(\varepsilon =0\) for all \(R < R_c\). This is the “bad” phase of full unemployment at \(\Gamma _0=0\), which becomes in this case a phase where employment grows steadily but remains of order \(\Gamma _0\) except very close to the critical point.

Eq. (30) is plotted in Fig. 9 together with numerical results. Note that in this case the representative firm approximation (\(N_\mathrm{F}=1\)) is in good agreement with numerical results also for \(N_\mathrm{F}=10,000\), as it was for the discontinuous transition obtained for \(\Gamma =0\). In the inset of Fig. 9 one can see that the variance of the fluctuations of employment rate is diverging as long as the critical value of R is approached. This is confirmed by a spectral analysis of the unemployment time series (see Fig. 10). In order to obtain the power spectrum we apply the GSL Fast Fourier Transform algorithm to the time series \(\varepsilon (t)-\langle \varepsilon \rangle \). As one can see in Fig. 10 the power spectrum is well approximated by an Ornstein-Uhlenbeck form:

$$\begin{aligned} I(\omega )=I_0\frac{\omega _0^2}{\omega _0^2+\omega ^2} \end{aligned}$$
(31)

with \(\omega _0\) going linearly to 0 when \(\eta ^0_+\) approaches its critical value, meaning that the relaxation time \(\omega _0^{-1}\) diverges as one approaches the critical point. Note that this is not the case for the Mark 0 model with \(\Gamma _0=0\) which instead has a white noise power spectrum even in proximity of the transition line. The first order (discontinuous) transition for \(\Gamma _0=0, \Theta =\infty \) is thus replaced by a second order (continuous) transition when the firms adapt their behavior as a function of their financial fragility.

Fig. 9
figure 9

Inverse of the average employment rate \(\Gamma _0/\overline{\varepsilon }\) as a function of the ratio \(R=\eta ^0_+/\eta ^0_-\) with \(\eta ^0_-=0.1\) and \(\gamma =0.01\) when \(\Gamma _0>0\). When the employment rate is rescaled with the parameter \(\Gamma _0\) (here \(\Gamma _0=10^{-3},\ 10^{-4}\)) the different lines collapse and Eq. (30) agrees with numerical simulations. In the inset we also plot the rescaled variance, still as a function of \(\eta ^0_+\). Approaching the critical point the variance of the unemployment fluctuations diverges, together with their relaxation time going to infinity. The other parameters are: \(\delta =0.02\), \(\Theta =5\), \(\gamma _w=0\), \(c=0.5\), \(\beta =0\) and \(\varphi =0.1\)

Fig. 10
figure 10

Logarithm of the normalized power spectrum for Mark 0 with adaptive firms (\(\Gamma _0=10^{-3}\)), \(\gamma _p=0.05\) and \(N_\mathrm{F}=1\) (left) and \(N_\mathrm{F}=1000\) (right). The other parameters are set as in Fig. 9. The main plot shows two examples of the spectrum for \(\eta _-^0=0.1\) and \(\eta _+^0=0.05\) (black line) and \(\eta ^0_+=0.09\) (red line) in the left plot, \(\eta _+^0=0.05\) (black line) and \(\eta ^0_+=0.08\) (red line) in the right plot. The time series is made of \(2^{28}\) time steps after \(T_{eq}=500\, 000\) and the logarithm of the spectrum is averaged over a moving window of 100 points for a better visualization. With both system sizes the fit with Eq. (31) (blue dashed lines) is good with the only difference that when \(N_\mathrm{F}>1\) a clear oscillatory pattern appears at high frequencies, that becomes sharper and sharper as \(N_\mathrm{F}\) increases. In the inset of each figure we plot the value of \(\omega _0\) in Eq. (31) obtained from the fit as a function of the ratio \(R=\eta _+^0/\eta _-^0\). In both cases \(\omega _0\) goes linearly to 0 as the critical value is approached

Finally, note that the presence of a continuum of states for the unemployment rate whenever \(\Gamma _0>0\) and \(R<R_c\) holds also with \(\gamma _w>0\) (when wages are not constant). It was however simpler to perform analytical computations with constant wages.

Appendix 3: Pseudo-code of Mark 0

We present here the pseudo-code for the Mark 0 code described in Sect. 2.2 and Appendix 1. The source code of the baseline Mark-0 is available on demand.

figure a
figure b

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

Gualdi, S., Tarzia, M., Zamponi, F. et al. Monetary policy and dark corners in a stylized agent-based model. J Econ Interact Coord 12, 507–537 (2017). https://doi.org/10.1007/s11403-016-0174-z

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s11403-016-0174-z

Keywords

Navigation