On Identification Issues in Business Cycle Accounting Models

Pedro Brinca (Nova School of Business and Economics, Universidade Nova de Lisboa, Lisboa, Portugal)
Nikolay Iskrev (Banco de Portugal, Lisboa, Portugal)
Francesca Loria (Board of Governors of the Federal Reserve System, Washington, DC, USA)

Essays in Honour of Fabio Canova

ISBN: 978-1-80382-636-3, eISBN: 978-1-80382-635-6

ISSN: 0731-9053

Publication date: 16 September 2022

Abstract

Since its introduction by Chari, Kehoe, and McGrattan (2007), Business Cycle Accounting (BCA) exercises have become widespread. Much attention has been devoted to the results of such exercises and to methodological departures from the baseline methodology. Little attention has been paid to identification issues within these classes of models. In this chapter, the authors investigate whether such issues are of concern in the original methodology and in an extension proposed by Šustek (2011) called Monetary Business Cycle Accounting. The authors resort to two types of identification tests in population. One concerns strict identification as theorized by Komunjer and Ng (2011) while the other deals both with strict and weak identification as in Iskrev (2010). Most importantly, the authors explore the extent to which these weak identification problems affect the main economic takeaways and find that the identification deficiencies are not relevant for the standard BCA model. Finally, the authors compute some statistics of interest to practitioners of the BCA methodology.

Keywords

Citation

Brinca, P., Iskrev, N. and Loria, F. (2022), "On Identification Issues in Business Cycle Accounting Models", Dolado, J.J., Gambetti, L. and Matthes, C. (Ed.) Essays in Honour of Fabio Canova (Advances in Econometrics, Vol. 44A), Emerald Publishing Limited, Leeds, pp. 55-138. https://doi.org/10.1108/S0731-90532022000044A004

Publisher

:

Emerald Publishing Limited

Copyright © 2022 This is a work completed under the United States Government and therefore is in the public domain.


1 Introduction

The business cycle accounting (BCA) procedure developed by Chari et al. (2007) and recently revived by Brinca, Chari, Kehoe, and McGrattan (2016) has sparked great interest among quantitative and theoretical economists. This tool allows to look the data through the lens of a standard business cycle model and, most importantly, to detect and quantitatively assess to which extent and in which equilibrium conditions the model performs better or worse. The so-called ‘wedges’ – which in practical terms are whatever makes the equilibrium conditions not hold – can be mapped into frictions of richer models and vice versa. In this context, the BCA exercise can thus be thought of as some ex ante diagnosis tool for researchers to know in which broad classes of models it is worth or not worth investing time if one wants to explain fluctuations in macroeconomic aggregates such as GDP, investment and working hours during a particular economic episode.

BCA has become a standard tool of business cycle analysis and, since its inception, literally hundreds of applications of the original methodology, have been performed.

Examples can be found in Kobayashi and Inaba (2006) for Japan, Simonovska and Söderling (2008) for Chile and Lama (2009) for Argentina, Mexico and Brazil. The results seem to conclude, much in line with Chari et al. (2007), that total factor productivity and distortions to the labour choice are relevant, whereas distortions to the savings decision are considerably less important. Some authors focus their analysis to one type of distortions as in Restrepo-Echavarria and Cheremukhin (2010) or Cociuba and Ueberfedt (2010), where the focus is on the labour-leisure margin, or other numerous studies which deal with total factor productivity, such as Islam, Dai, and Sakamoto (2006). Another line of work looks into a selected sample of countries and a specific periods of fluctuations such as output drops (see, e.g., Dooyeon & Doblas-Madrid, 2012). Brinca (2014) instead provides a comprehensive exercise for 22 OECD countries covering the period from 1970 to 2012. It looks at the quantitative relevance of each distortion over the whole business cycle and not just booms or busts. The results confirm the findings of Chari et al. (2007) regarding the Great Depression and the 1981 recession in the United States, while stressing the relevance of the international channels of transmission of these distortions.

In terms of methodological departures, Otsu (2009) extends the methodology to a two-country setting. Šustek (2011) includes a Taylor-type nominal interest rate setting rule and an extra asset, government bonds, to study the behaviour of nominal variables such as the nominal interest rate on bonds and the inflation rate, and gives it the name of monetary business cycle accounting (MBCA). These departures have also been explored. For instance, Brinca (2013) applies Šustek (2011) model to perform a MBCA exercise for Sweden, comparing the 1990s crisis with the period of the Great Recession.

While much attention has been devoted to the BCA methodology, no efforts have been made in investigating whether the parameter estimation procedure associated with this tool is affected by identification deficiencies. The question of identifiability in dynamic stochastic general equilibrium (DSGE) is an important one as it might jeopardize the consistency and adversely affect the precision of parameter estimates. This issue becomes even more important in light of the fact that, in the past years, DSGE models have become a standard and important asset within the toolkit of economic policy-makers to make quantitative statements about real and nominal variables. When these models are brought to the data researchers should be cautious in taking for granted the empirical credibility of their estimated parameters and, thus, of the economic implications that the latter entail. Indeed, due to identifiability issues inherent to DSGE models, it is far from obvious that parameters can be inferred successfully even when one has an infinite sample of observed data and when full-information methods such as maximum likelihood (ML) are employed in estimation. Most importantly, as pointed out by Canova and Sala (2009), in some cases these identification deficiencies can result in significantly different economic inference from the theoretical models of interest. To the extent that the BCA exercises by Chari et al. (2007) and Šustek (2011) draw quantitative conclusions from their respective DSGE models of reference it is of outmost importance that their identification potential is carefully analyzed.

This chapter builds up on the literature which studies local sample and population identification issues specific to linearized DSGE models. Our methodological approach to the analysis of such issues is closely related to the work by Canova and Sala (2009) whose contribution was to provide: (i) a working language which allows researchers to classify identification problems; (ii) formal graphical inspection tools to detect those problems; and (iii) possible ways to obviate them. Due to the high number of parameters to be estimated in both models graphical inspection quickly becomes unmanageable and dispersive. This leads us to bring into our toolkit the formal identification tests developed by Komunjer and Ng (2011) and Iskrev (2010).

The results presented in this chapter suggest that the standard and MBCA model do not suffer from strict identification failures when estimation is restricted to the parameters governing the law of motion of the latent variables and that this is not true anymore once one extends the estimation to the deep parameters of the model. We show that restricting estimation of some deep parameters can obviate these strict identification failures. We also find that both models are affected by weak identification deficiencies and that these are induced by several parameters of the model not exerting a distinct effect on our objective function of interest, namely the likelihood function of the model. Finally, we explore to which extent this type of identification failure affects the main conclusions to be drawn from (M)BCA exercises. We find that the main takeaways from a standard BCA exercise are not overturned when one explicitly takes into account the weak identifiability of the model’s parameters. We are still investigating whether this result holds through in the MBCA framework.

Finally, we analyze how overall identification strength of the estimated parameter vector varies across sample sizes. To do so, we compute empirical distance measures as in Qu and Tkachenko (2017). We find that in both models the parameter set is overall well identified even if a practitioner is faced with a data set of only 20 data points per observable.

This chapter is organized as follows. Section 2 presents the prototype MBCA economy. The methodology to run the identification tests is illustrated in Section 3. Their results are reported in Section 4 while Section 5 discusses their economic relevance. In Section 6, we present statistics of interest to practitioners and then conclude in Section 7.

2 The Prototype (M)BCA Economy

We start by introducing the MBCA prototype economy as in Šustek (2011), which is an extension of the prototype economy in Chari et al. (2007), a neoclassical growth model with labour-leisure choice. The extension consists in allowing for an extra asset (government bonds) and introducing a nominal interest rate setting rule.

2.1 Description of the Economy

There is an infinitely lived representative agent that maximizes expected discounted utility and a representative firm, both price-takers in all markets. The economy experiences one of finitely many events st, where st=(s0,,st) is the history of events up to period t which occur with probability πt(st). There are six exogenous stochastic variables which are all function of the random variable st. Four of them are the same as in Chari et al. (2007). These are the efficiency wedge Zt(st), the labour wedge 1τl,t(st), the investment wedge 1/[1+τx,t(st)] and the government wedge gt(st). With the Šustek (2011) extension, two more stochastic variables are added: an asset market wedge 1/[1+τb,t(st)] and a monetary policy wedge R˜t(st).

The representative household chooses how much to consume ct(st) and how much labour to supply lt(st). Given the discount factor β it solves the following maximization problem:

(2.1)max{ct(st),lt(st),kt+1(st),bt(st)}E0Σt=0βtUct(st),1l(st)(1+gn)t,

subject to the budget constraint:

(2.2)ct(st)+[1+τx(st)]xt(st)+[1+τb(st)](1+gn)bt(st)[1+Rt(st)]pt(st)bt1(st1)pt(st)=[1τl(st)]wt(st)lt(st)+rt(st)kt(st1)+Tt(st)

where xt(st) is investment, gn the population growth rate, bt(st) are bond holdings paying a net nominal rate of return Rt(st) in all states of the world, pt(st) is the nominal price of goods in terms of a numeraire, wt(st) the wage rate, rt(st) the real rental rate of return on capital kt(st) held at the beginning of period and Tt(st) lump-sum transfers from the government. Capital accumulation follows:

(2.3)(1+gn)kt+1(st)=(1δ)kt(st)+xt(st)

where δ is capital’s depreciation rate. The production function for the representative firm is given by

(2.4)yt(st)=Fkt(st1),Zt(st)lt(st),

which is assumed to exhibit constant returns to scale (CRS). The aggregate resource constraint is then given by

(2.5)yt(st)=ct(st)+gt(st)+xt(st).

There is a monetary authority who reacts to deviations from steady-state output y and inflation π by setting the nominal interest rate Rt(st) according to:

(2.6)Rt(st)=(1ρR)R+ωy(lnyt(st)lny)+ωπ(πt(st)π)+ρRRt1(st1)+R˜t(st)

where ρR[0,1) and πt(st)lnpt(st)lnpt1(st1) is the inflation rate. In addition, it is assumed that ωπ>1, thus eliminating explosive paths for inflation.

Just like in Chari et al. (2007) it is assumed that the state st follows a Makov process of the type

(2.7)st+1=P0+Pst+Qεs,t+1,

where εs,t+1N(0,I). Moreover, the mapping between this process of the underlying event st=(sZt,slt,sxt,sgt,sbt,sR˜t) and the wedges Zt,1τl,t,11+τx,t,gt,11+τb,t,Rt˜ is one-to-one and onto. This setup is equivalent to assuming that agents use only past realizations of wedges to forecast future ones. Note that in the case where the matrix P is diagonal then, irrespectively of whether the covariance matrix QQQQis diagonal or not (i.e. whether the shocks are allowed to be correlated or not), the MBCA model is block-recursive in the sense that shocks to the wedges of the standard BCA setup only affect real variables while leaving the model’s nominal variables – interest rate Rt and inflation rate πt– unaffected. In this sense, we can think of the MBCA theoretical framework as nesting the plain-vanilla BCA one. This is why we avoid a detailed exposition of the latter.

2.2 Equilibrium Conditions

Equilibrium allocations are pinned down by the production function in (2.4), the aggregate resource constraint in (2.5) and the first-order conditions with respect to labour, capital and bond holdings below:

(2.8)Ul,t(st)Uc,t(st)=[1τl,t(st)]wt(st),
(2.9)1=βEt{Uc,t+1(st+1)Uc,t(st)[[1+τx,t+1(st+1)](1δ)+rt+1(st+1)1+τx,t(st)]},
(2.10)1=βEt{Uc,t+1(st+1)Uc,t(st)1+τb,t+1(st+1)1+τb,t(st)pt(st)pt+1(st+1)[1+Rt(st)]}.

In Appendix B, we derive these conditions and present the full-fledged model.

The notation v^vt(1+gz)tVtNt(1+gz)t refers to model variables Vt which are not only expressed in per-capita terms vt but also detrended v^t.

2.3 Operational Model

In the operational version of the model which we bring to the data we consider quantities which are not only expressed in per-capita terms but also detrended (see Appendix B.5 for derivations). To highlight the differences between this model’s and the previous model’s variables we introduce the notation v^vt(1+gz)tVtNt(1+gz)t.

The model is given by the CRS Production Function:

(2.11)y^t(st)=k^t(st1)αztlt(st)1α,

the aggregate resource constraint:

(2.12)y^t(st)=c^t(st)+g^t+x^t(st),

the capital accumulation law:

(2.13)(1+gn)(1+gz)k^t+1(zt)=(1δ)k^t(zt1)+x^t(zt),

the Taylor rule:

(2.14)Rt(st)=(1ρR)R+ωy(lny^t(st)lny^)+ωπ(πt(st)π)+ρRRt1(st1)+R˜t,

the F.O.C. for labour:

(2.15)ψc^t(st)1lt(st)=(1τl,t)(1α)k^t(st1)αzt1αlt(st)α,

the F.O.C. for capital:

(2.16)1=β˜Etc^t+1c^tσ1lt+11ltψ(1σ)(1+τx,t+1)(1δ)+αk^t+1(st)α1zt+1lt+1(st+1)1α1+τx,t

and the F.O.C. Bonds:

(2.17)1=β˜Etc^t(st)c^t+1(st+1)1+τb,t+11+τb,tpt(st)pt+1(st+1)[1+Rt(st)],

where β˜=β/(1+gz).

Notice that the operational model the efficiency wedge is thus given by zt=Zt(1+gn)t and the government wedge by g^t=gt(1+gn)t.

The structural parameters of the BCA model are: (i) gn, population growth rate of labour-augmenting technological process; (ii) gz, growth rate of labour-augmenting technological process; (iii) α, parameter which determines the share of (and weight on) net capital stock in the Cobb-Douglas CRS production function; (iv) β, subjective discount factor, reflecting the time preference of the household; (v) δ, depreciation rate of net capital stock; and (vi) ψ, Frisch elasticity of labour supply. In the MBCA model the deep parameters also include (vii) ρR, weight on lagged nominal interest rate in the Taylor rule (extent of ‘interest rate smoothing’); (viii) ωπ, coefficient on deviations of inflation from its steady-state value in Taylor rule; (ix) ωy, coefficient on deviations of output from its steady-state value in Taylor rule; and (x) πss, steady-state inflation.

3 Methodology

The problem of identification emerges when a researcher seeks to infer the parameters of his theoretical model from a sample of observed data. In general and loosely speaking, the requirements for ‘successful estimation’ are that (i) the objective function (e.g. the log-likelihood function) has a unique maximum; (ii) the Hessian at the mode is negative definite and has full rank; and that (iii) the objective function exhibits ‘sufficient’ curvature. Local identifiability is crucial as it guarantees consistent parameter estimation and that the estimator has the usual asymptotic properties, where by ‘local’ we mean that the maximum in condition (i) above is at least locally unique. It is thus of outmost importance to investigate thoroughly whether the conditions for local identifiability are satisfied within the standard and MBCA framework.

In this analysis, we employ strict and weak identification tests. Both are performed in population, that is, given a theoretical sample of infinite length. The former provides a yes or no answer to the question whether a parameter is identifiable or not and, thus, whether condition (ii) above is satisfied or not. The latter investigates condition (iii) above and is thus informative about whether the likelihood exhibits small curvature, where small is in relation to an economically relevant parameter range. The weak identification test also instructs about whether the small curvature is due to the fact that the parameters have no effect on the objective function or to the fact that the variation which they induce on the objective function is not distinct enough from other parameters which are being estimated.

As to the strict identification tests in population, consider the case where the researcher has a sample of length T generated by (3.4) with θ=θ0. In this context, one can ask the following question: If the sample was infinitely large, that is, T, under which conditions would it be possible to uncover the value θ0 and the model that generated the data? Problems specific to the dynamic nature of DSGE models make it difficult to test whether the conditions for local identifiability mentioned above hold given a sample of data. Indeed, as pointed out by Canova and Sala (2009), the mapping from the structural parameters to the solution coefficients is typically unknown and the latter, in turn, usually appear in a non-linear way in the objective function. These are some of the reasons why the rank and order conditions of Rothenberg (1971) derived for simultaneous system of equations cannot be applied. As pointed out by Komunjer and Ng (2011) other reasons are that classical conditions: (i) are derived for static models whose reduced-form errors are orthogonal to the regressors, an assumption which is implausible for DSGE models given that they are dynamic; (ii) do not recognize the fact that reduced-form parameters themselves can be not identifiable unless all state variables are observed: and (iii) are for simultaneous system of equations which is not the form of DSGE model solutions.

Alternative rank and order conditions for strict identification are derived by Komunjer and Ng (2011) using the spectral density matrix and by Iskrev (2010) using the likelihood function of the DSGE model. Both tests treat parameter identification as a property of the underlying structural model. This is motivated by the fact that DSGE models completely characterize the data generating process. This is in contrast with other types of models where the mapping from the model to the data is only partially known. Therefore, the economic model is the origin of identification problems which appear in a particular data set. It is then straightforward to see that identification problems may occur as an intrinsic property of the model when, for instance, the restrictions that the model imposes on the joint distribution of the observed variables do not contain sufficient information about some parameters of interest. It is important to recognize the fact that, in general, these restrictions are a function of the parameters. Hence, also the data crucially contributes to identify those parameter values for which the model can account well for the movements in the data.

An advantage of the Iskrev (2010) approach over Komunjer and Ng (2011) is that key objects of interest are obtained analytically rather than numerically. Furthermore, it can detect not only strict but also weak identification problems. A central tool in his analysis is the expected Fisher information matrix, as first suggested by Rothenberg (1971). It is intuitive to understand why this matrix comes handy to study identification problems. The information matrix measures the curvature of the expected log-likelihood surface and, as pointed out by Rothenberg (1971), it is informative about the (degree of) informational content available in the sample about the unknown parameters. For instance, one should expect identification deficiencies to arise when the log-likelihood surface is flat or nearly flat with respect to the parameters to be estimated. The degree of ‘flatness’ can be detected and quantified via the information matrix.

There are two main reasons why parameters might be unidentifiable or just weakly identifiable; they can be broadly classified as a ‘sensitivity’ and a ‘collinearity’ factor. They originate from the fact that the economic features which operate via the problematic parameters may be nearly or completely irrelevant with respect to the variables of the model used in estimation. The ‘sensitivity’ factor signals that the identification problem occurs because the features are inherently unimportant while the ‘collinearity’ factor attributes it to the nearly redundant nature of these features given others present in the model. The information matrix can be used to assess the importance of the two factors via a simple decomposition. This analysis thus allows not only to flag problematic parameters but also to quantify the strength and discern the nature of their identification deficiencies.

3.1 State Space Form

Following Chari et al. (2007) the state space form of the model is given by:

(3.1)Xt+1=A(θ)Xt+B(θ)εt+1,Yt=C^(θ)Xt+ωt,ωt=D^(θ)ωt1+ηt,

where, in the standard BCA model, the vectors of state variables and the vector of observables are, respectively, given by Xt=[log(k^t)log(k),log(zt)log(z),τltτl,τxtτx,log(g^t)log(g^),1] Yt=[logy^tlog(y^),logx^tlog(x^),logltlog(l),logg^tlog(g^)], whereas in the MBCA model Xt=[log(k^t)log(k),log(zt)log(z),τltτl,τxtτx, log(g^t)log(g^),τbtτb,R˜tR˜,1],Yt=[logy^tlog(y^),logx^tlog(x^),logltlog(l),logg^tlog(y^),RtR,πtπ] .

The parameters of the model are collected in a vector θ and belong to a set Θnθ. The matrices A,B,C^ and D^ are, respectively, the ones describing: (i) the transition of the states; (ii) the variance covariance matrix of the shocks to the wedges utB(θ)εt given by ΩBB since it is assumed that E[εtεt]=I; (iii) the mapping from the states to the observables; and (iv) the serial correlation of the measurement errors (set to 0 here).

For both models, the state space matrices A(θ) and B(θ) are obtained by solving the rational expectations systems using the Gensys algorithm developed by Sims (2002) (see Appendix C for more details).

Let us assume that Eηtηt=R and Eεtηs=0 for all periods t and s. Next, define Y¯tYt+1D^Yt. Then we can rewrite (3.1) as

(3.2)Xt+1=A(θ)Xt+B(θ)εt+1,Y¯t=C¯(θ)Xt+C^(θ)B(θ)εt+1+ηt+1,

where C¯(θ)=C^(θ)A(θ)D^(θ)C^(θ).

Stacking the vector of innovations and measurement errors into a nϵ×1 vector ϵt=εt,ηt yields the following representation:

(3.3)Xt+1=A(θ)Xt+B(θ)εt+1,Y¯t=C¯(θ)Xt+D¯(θ)ϵt+1.

In all periods and all identification tests we set the measurement errors equal to 0 so that D^=R=04×4 and

(3.4)Xt+1nX×1=A(θ)nX×nXXtnX×1+B(θ)nX×nεεt+1nε×1,Yt+1nY×1=C(θ)nY×nXXtnX×1+D(θ)nY×nεεt+1nε×1,

where C(θ)=C^(θ)A(θ) and D(θ)=C^(θ)B(θ).

3.2 Estimated Parameters

The estimated parameters are those governing the stochastic process:

(3.5)st+1=P0+Pst+Qεs,t+1

underlying the wedges and are thus the ones appearing in the matrices P0, P and Q. More specifically, for the standard BCA model the stochastic process of the wedge shocks takes the form:

(3.6)logzt+1τlt+1τxt+1logg^t+1st+1=z¯τ¯lτ¯xg¯P0+ρzρz,τlρz,τxρz,gρτl,zρτlρτl,τxρτl,gρτx,zρτx,τlρτxρτx,gρg,zρg,τlρg,τxρgPlogztτltτxtlogg^tst+q11000q21q2200q31q32q330q41q42q43q44Qεz,t+1ετl,t+1ετx,t+1εg,t+1εs,t+1,

The estimated steady-state vector of wedge shocks [logzt+1,τlt+1,τxt+1,logg^t+1] are given by (I4×4P)1P0=[log(z)τlτxlog(g^)] which we define as [zssτlssτxssgss] for convenience.

In the MBCA model the stochastic process of the wedge shocks takes the form:

(3.7)logzt+1τlt+1τxt+1logg^t+1τbt+1R˜t+1st+1=z¯τ¯lτ¯xg¯τ¯bR¯P0+ρzρz,τlρz,τxρz,gρz,τbρz,R˜ρτl,zρτlρτl,τxρτl,gρτl,τbρτl,R˜ρτx,zρτx,τlρτxρτx,gρτx,τbρτx,R˜ρg,zρg,τlρg,τxρgρg,τbρg,R˜ρτb,zρτb,τlρτb,τxρτb,gρτbρτb,R˜ρR˜,zρR˜,τlρR˜,τxρR˜,gρR˜,τbρR˜Plogztτltτxtlogg^tτbtR˜tst+q1100000q21q220000q31q32q33000q41q42q43q4400q51q52q53q54q550q61q62q63q64q65q66Qεz,t+1ετl,t+1ετx,t+1εg,t+1ετb,t+1εR˜,t+1εs,t+1.

The estimated steady-state vector of wedge shocks [logzt+1,τlt+1,τxt+1,logg^t+1,τbt+1,R˜t+1] is given by (I6×6P)1P0=[log(z)τlτxlog(g^)τbR˜] which, again for convenience, we define as [zssτlssτxssgssτbssR˜ss]. Šustek (2011) imposes τbss=0 and R˜ss=0 in estimation. In the identification tests we explore, inter alia, the case where these two parameters are estimated as well.

In Section 4, we also present results for the case where the structural parameters of the respective models are included in the identification analysis.

3.3 Komunjer and Ng (2011) Test for Strict Identification

In a DSGE model two parameter vectors θ0 and θ1 are observationally equivalent if the spectral density matrix evaluated at θ0 is equal to the spectral density matrix evaluated at θ1. A parameter set θ0Θ is defined to be locally identifiable from the autocovariances of Yt if there exists an open neighboorhood of θ0 such that θ0 and θ1 being observationally equivalent necessarily implies θ1=θ0.

Formally, Komunjer and Ng (2011) state that two triples (θ0,InX,Inε) and (θ1,T,U) are observationally equivalent if

(3.8)A(θ1)=TA(θ0)T1
(3.9)B(θ1)=TB(θ0)U
(3.10)C(θ1)=C(θ0)T1
(3.11)D(θ1)=D(θ0)U
(3.12)Σ(θ1)=UΣ(θ0)U1

where A(·),B(·),C(·),D(·) and Σ(·) are the matrices of the state space model described in (3.4) with T and U being full rank matrices.

A necessary sufficient condition for identification is thus checking that the mapping

(3.13)δS(θ,T,U)=vec(TA(θ)T1)vec(TB(θ)U)vec(C(θ)T1)vec(D(θ)U),vec(UΣ(θ)U1)

has full rank. The rank condition for local identification at θ0 when the state space system is square (i.e. nε=nY) is thus given by

(3.14)rankΔS(θ0)=nθ+nX2+nε2,

where

(3.15)ΔS(θ0)ΔΛS(θ0),ΔTS(θ0),ΔUS(θ0)δ(θ0,InX,Inε)θ,δ(θ0,InX,Inε)vecT,δ(θ0,InX,Inε)vecU.

They also establish the following necessary order condition for identification:

(3.16)nθ+nX2+nε2nΛS=(nX+nY)(nX+nε)+nε(nε+1)/2.

It requires the number of equations defined by δS to be at least as large as the number of unknowns in those equations and can be rewritten as:

(3.17)nθnYnX+nε(nX+nYnε)+nε(nε+1)2nδ.

Minimality and left-invertibility of the state space system are maintained assumptions of these conditions and we thus verify that they hold in our analysis. The first condition gets fulfilled by rewriting (3.4) in the particular form:

(3.18)X˜t+1=X1,t+1X2,t+1=A˜1(θ)0A˜2(θ)0X1,tX2,t+B˜1(θ)B˜2(θ)εt+1,
(3.19)Yt+1=C˜1(θ)C˜2(θ)X1,t+1X2,t+1,

so that

(3.20)X1,t+1=A˜1(θ)A(θ)X1,t+B˜1(θ)B(θ)εt+1,
(3.21)Yt+1=C˜1(θ)A˜1(θ)+C˜2(θ)C˜2(θ)C(θ)X1,t+C˜1(θ)B˜1(θ)+C˜2(θ)B˜2(θ)D(θ)εt+1.

Left-invertibility is ensured by the following assumption: For every θΘ, rank P(z;θ)=nX+nε in |z|>1, where P(z;θ)zInXA(θ)B(θ)C(θ)D(θ),z.

3.4 Iskrev (2010) Test for Strict and Weak Identification

In the following exposition, we closely follow Iskrev (2010).

3.4.1 Preliminaries

The log-likelihood function of a sample YT of data can be obtained using the sequence of one-step ahead prediction errors et|t1=YtC^(θ)X^t|t1. The latter can be easily constructed using the one-step ahead forecasts of the state vector X^t|t1 returned by the Kalman filter. Assuming that the structural shocks are Gaussian implies that the conditional distribution of et|t1 is Gaussian as well, with mean 0 and covariance matrix St|t1=C^(θ)Pt|t1C^(θ), where Pt|t1=EXtX^t|t1XtX^t|t1 can also be obtained from the Kalman filter recursions and is the covariance matrix of the one-step ahead forecasts conditional on information up to time t − 1. The log-likelihood of the sample is then given by

(3.22)lT(θ)=const.12Σt=1Tlog|St|t1|12Σt=1Tet|t1St|t11et|t1,

Under some regularity conditions the ML estimator θ˜T is consistent, asymptotically efficient and asymptotically normally distributed with

(3.23)T(θ˜Tθ0)d(0,I01),

where I0 is the asymptotic Fisher information matrix evaluated at the true value of θ. More formally,

(3.24)I0limT1TIT,

where IT is the finite sample Fisher information matrix given by

(3.25)ITElT(θ)θlT(θ)θ.

3.4.2 General Principles of Identification Analysis

Suppose that inference about the parameters of the model collected in the vector θ is made using a sample with T observations of a random vector Y with a known probability density function p(YT;θ), where YT=[Y1,,YT]. The latter, when considered as a function of θ, contains all available sample information about θ associated with the observed data. It is then straightforward to see that a prerequisite for successful inference about θ is that its values imply distinct values of the density function p(YT;θ). More formally, a point θ0Θ is said to be identified if:

(3.26)Prp(YT;θ)=p(YT;θ0)=1θ=θ0.

That is, if the density function yields the same value when evaluated at θ and at θ0 this implies that θ is equal to θ0. It is possible to rewrite this condition in terms of the log-likelihood function lT(θ)logP(YT;θ):

(3.27)E0lT(θ0)E0lT(θ),θ.

This follows from Jensen’s inequality, see Rao (1971), and the logarithmic function being concave. It implies that the function H(θ0,θ)E0lT(θ)lT(θ0) achieves a maximum at θ=θ0, and that θ0 is identified if and only if the maximum is unique. The conditions for local uniqueness of a maximum at θ0 are that (i) H(θ0,θ)θ|θ=θ0=0 and (ii) 2H(θ0,θ)θθ|θ=θ0 is negative definite. If the maximum at θ0 is locally unique then θ0 is locally identified, that is, there exists an open neighborhood of θ0 where (3.26) holds θ.1 One can show that, see Bowden (1973), the condition in (i) is always fulfilled and that the Hessian matrix in (ii) is equal to the negative of the Fisher information matrix. This leads to the following result by Rothenberg (1971): Let θ0 be a regular point2 of the information matrix IT(θ). Then θ0 is locally identifiable if and only if IT(θ0) is non-singular.

In general, non-singularity of the Fisher information matrix is both necessary and sufficient for local identification.3 The information matrix is singular whenever the expected log-likelihood function is flat at θ0. In this case, due to the lack of the variability induced by the parameters on the log-likelihood function, it is impossible to make inference about the parameters even with an infinite sample of data. There are two reasons why this might occur, following Iskrev (2010) we call them the ‘sensitivity’ and ‘collinearity’ factors. Either the parameters have no effect on the expected log-likelihood (‘lack of sensitivity’), or different parameter values induce the same changes in the expected log-likelihood (‘perfect collinearity’). It is thus useful to formalize ideas in order to investigate to which extent the two channels are at work. This can be done by using the fact that the information matrix is equal to the covariance matrix of the scores and can thus be expressed as:

(3.28)IT(θ0)=Δ1/2RT(θ0)Δ1/2,

where Δ=diagIT(θ0) is a diagonal matrix containing the variances of the elements of the score vector, and RT(θ0) is the correlation matrix of the score vector. Thus a parameter θiθ is locally unidentifiable if:

  • (I)

    Lack of sensitivity: The expected log-likelihood is not affected by small changes in θi, that is,

    (3.29)ΔiElT(θ0)θi2=E2lT(θ0)θi2=0

  • (II)

    Perfect collinearity: The effect of small changes in θi on the expected log-likelihood can be offset by varying other parameters, that is,

    (3.30)ϱi11/RTii=1,

    where RTii is the ith diagonal element of the inverse of RT. As Iskrev (2010) puts its

[t]he intution about the meaning of ϱi comes from a well-known property of the correlation matrix Tucker, Cooper, and Meredith (1972), which implies that ϱi is the coefficient of multiple correlation between the partial derivative of the log-likelihood with respect to θi and the partial derivatives of the log-likelihood with respect to the other elements of θ.

Conditions (I) and (II) characterize the case in which the expected log-likelihood is completely flat and the parameters are thus not identifiable in a strict sense. The case of weak identification, on the other hand, arises when the expected log-likelihood features little curvature with respect to some parameters. We delve further into this issue next.

3.4.3 Identification Strength

Local identifiability guarantees, in general, consistent estimation of θ. The precision with which θ is estimated is governed by the curvature of the expected log-likelihood function in the neighborhood of θ0, of which the rank conditions above are not informative. Identification is weak whenever small changes in θ do not induce sufficiently large changes in lT(θ) or, equivalently, when small changes in lT(θ) are associated with large changes in θ. By weak we mean that the estimates are prone to be imprecisely estimated even in the presence of an infinitely large sample of data. The degree of ‘weakness’ is thus related to the degree of precision. The latter is not an absolute but rather a relative concept which varies according to the application at hand.

We already saw that a central tool when investigating whether a parameter is locally identifiable or not is the Fisher information matrix. This is because, as shown in Rothenberg (1971), the latter is indicative of the degree of curvature of the expected log-likelihood function. To understand the next logical step in the analysis, namely the relationship between the curvature and the precision of the ML estimator θ^T it is useful to recall its asymptotic distribution described in (3.23). It is then straightforward to see that IT1(θ0)/T is the sample counterpart of the covariance matrix of θ^T and, analogously, ITii1(θ0)/T is the sample counterpart of the variance of θ^i.

Asymptotic efficiency of ML estimation implies that θ^T has the smallest asymptotic covariance matrix within the class of consistent estimators. This follows directly from the fact that, according to the Cramér-Rao theorem, the lower bound of the asymptotic covariance of any consistent estimator θ is given by the inverse of its asymptotic information matrix I0. Concurrently, the covariance matrix of any unbiased estimator is bounded below by the inverse of the sample information matrix IT so that biITii1 represents the lower bound on the variance of any unbiased estimator θi. To measure identification strength we can thus construct bounds on 1 standard deviation intervals for the individual parameters.

As shown in Iskrev (2010), it is possible to relate the size of the bounds to the potential roots of identification deficiencies. Indeed, using the decomposition of IT(θ) in (3.25) and the properties of the correlation matrix one obtains:

(3.31)bi=1Δi(1ρi)2.

When can one ascribe the identification problem to the ‘sensitivity’ or ‘correlation’ factor? In the first case, we know that the parameter does exert an irrelevant or weak effect on the likelihood, in which case Δi0. In the second case, the effects induced on the likelihood by one parameter are compensated, and thus made redundant, by changes in other parameters, in which case ρi1. In both cases, the sources of weak identification lead to a large bi and make inference about a parameter value challenging at best.

Notice that the sensitivity factor alone cannot guarantee successful parameter identification. Indeed, even if Δi is large, nearly perfect collinear effects of a parameter θi with respect to the other parameters θi lead to values of ρi close to 1, in which case identification remains weak. This example illustrates well the difference between the information about θi contained in the likelihood when the other parameters are known, see Δi, and when they are not known, see bi. This second source of information is smaller and the difference is increasing in the ‘correlation’ factor, see ϱi.

4 Results

In this section, we report results for the tests by Komunjer and Ng (2011) and Iskrev (2010). Results for the case where investment adjustment costs are introduced can be found in Appendix A. In the BCA and MBCA models, strict and local identifications are checked at the parameter set estimated (or fixed, as it is the case for the deep parameters) in Chari et al. (2007) and Šustek (2011), respectively.

4.1 Komunjer and Ng (2011)

We explore whether a parameter is identifiable in a strict sense. To answer this question in population we show results of the test by Komunjer and Ng (2011) first. Since this test requires computing the rank of the matrix ΔS(θ0) in (3.15) to check whether the rank condition for strict identification holds we report results for different tolerance levels. Indeed, the rank of a matrix is equal to the number of its non-zero eigenvalues which are found by numerical routines using a cut-off to establish whether they are sufficiently small. Matlab, for instance, uses the tolerance Tol = max(size(M))EPS ||M||, where EPS is the float point precision of M. As pointed out by Komunjer and Ng (2011) this default tolerance does not take into account the fact that the matrix ΔS(θ0) is often sparse and can thus be misleading. Results are thus reported for 11 tolerance values ranging from a maximum tolerance of 1e−2 to a minimum of 1e−11, along with the Matlab default one.

To isolate the parameters which are not strictly identifiable even with an infinite sample of data combinations of parameters which cause full rank failures in the ΔS(θ0) are searched by inspecting its change in rank and its null space. This is first done at a high tolerance level of 1e−3 to flag the most difficult parameters to identify and then at the lowest tolerance level for which identification fails in order to find additional problematic parameters. Komunjer and Ng (2011) choose the highest tolerance level of 1e−3 ‘on the grounds that the numerical derivatives are computed using a step size of 1e−3’. In Appendix D, we report results for the case where a Matlab selected measure of the step size is being used when computing numerical derivatives. When this other measure is used the results suggest more identification power in both models.

In general, the lower the tolerance level the higher the rank of the matrix since more of the smallest eigenvalues are considered to be numerically different from 0. Then, according to the rank-nullity theorem – which states that the sum of the rank and the nullity of a matrix is equal to its number of columns – the null space will be smaller as well. This would lead to think that the set of parameters which are flagged as troublesome should be larger the lower the tolerance level. However, the way the set of problematic parameters is found by identifying the number of columns of the orthonormal basis for the null space of the matrix ΔTS – obtained via singular value decomposition – whose (absolute) sum of elements is greater than the tolerance value (i.e. numerically larger than 0). This is because the vectors contained in a basis must be linearly independent and a null vector would always be linearly dependent with the other vectors. Thus, the lower the tolerance level, the more columns (and associated parameters) will fall within this set, the lower is the null space of the matrix ΔTS and, hence, the smaller is the set of parameters which are not strictly identifiable.

Finally, we perform conditional identification tests. More specifically, when the rank condition for strict identification fails, we check which parameters, when restricted, can enable identification of the remaining parameters.

4.1.1 Chari et al. (2007) BCA Model

Table 1 reveals that the model’s parameters are strictly identifiable at Tol = 1e−11 and at the Matlab default tolerance level. When inspecting the null space of ΔS(θ0) no problematic parameters are found. This result might be explained by the fact that the lower is the tolerance level, the smaller is the set of problematic parameters, as explained above.

Table 1.

Komunjer and Ng Test Results BCA Model.

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 29 25 15 51 40 62 0
e−03 29 25 16 54 45 69 0
e−04 29 25 16 a 45 69 0
e−05 29 25 16 54 45 69 0
e−06 29 25 16 54 45 69 0
e−07 29 25 16 54 45 69 0
e−08 30 25 16 54 46 70 0
e−09 30 25 16 54 46 70 0
e−10 30 25 16 55 46 70 0
e−11 30 25 16 55 46 71 1
Default = 2.756906e−12 30 25 16 55 46 71 1
Required 30 25 16 55 46 71 1

Summary: nθ=30,nX=5,nε=4.

Order condition: nθ=30,nδ=50.

Moving to the case where also the deep parameters of the model are included in the identification test reveals a different picture, as reported in Table 2. Indeed, the extended parameter set does not pass the test at any tolerance value and several problematic parameters are found. At Tol = 1e−3 the latter mainly concern some steady-state values of the wedge shocks, off-diagonal elements of the matrix P and all deep parameters. Once the tolerance is further lowered to Tol = Default also several off-diagonal elements of the Q matrix are flagged as troublesome. At the default tolerance level, the model’s parameters might be strictly identifiable if at least two restrictions are imposed. Indeed, all matrices are full rank with the exception of ΛS which is short rank by 2. We thus check whether fixing some parameters alleviates the strict identification deficiencies. As emerges from Table 3 there are several sets of restricted parameter combinations which allow to do so.

Table 2.

Komunjer and Ng Test Results BCA Model (Deep Parameters Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 33 25 15 52 44 64 0
e−03 35 25 16 57 51 71 0
e−04 35 25 16 57 51 71 0
e−05 35 25 16 58 51 71 0
e−06 35 25 16 58 51 71 0
e−07 36 25 16 58 52 72 0
e−08 36 25 16 59 52 74 0
e−09 37 25 16 60 53 75 0
e−10 37 25 16 60 53 76 0
e−11 37 25 16 60 53 76 0
Default = 5.513812e−12 37 25 16 61 53 76 0
Required 37 25 16 62 53 78 1

Summary: nθ=37,nX=5,nε=4.

Order condition: nθ=37,nδ=50.

Problematic parameters at Tol = 1e−3: zss, τlss, gss, ρτl,z, ρz,τl, ρτx,τl, ρg,τl, ρτl,τx, ρτl,g, q22, gn, gz, β, ψ, σ.

Problematic parameters at Tol = 5.513812e−12: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρz,g, ρτl,g, ρτx,g, ρg, q21, q31, q22, q32, q42, q33, q43, q44, gn, gz, β, δ, ψ, σ, α.

Table 3.

Komunjer and Ng Conditional Test Results BCA Model, Tol = 5.513812e−12.

Fixed ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
τlss zss 37 25 16 62 53 78 1
ρτl,z zss 37 25 16 62 53 78 1
ρz,τl zss 37 25 16 62 53 78 1
ρτx,τl zss 37 25 16 62 53 78 1
ρg,τl zss 37 25 16 62 53 78 1
ρτl,τx zss 37 25 16 62 53 78 1
ρτl,g zss 37 25 16 62 53 78 1
q21 zss 37 25 16 62 53 78 1
q22 zss 37 25 16 62 53 78 1
ψzss 37 25 16 62 53 78 1
gss τlss 37 25 16 62 53 78 1
gn τlss 37 25 16 62 53 78 1
gz τlss 37 25 16 62 53 78 1
β τlss 37 25 16 62 53 78 1
ρτl,z gss 37 25 16 62 53 78 1
ρz,τl gss 37 25 16 62 53 78 1
ρτx,τl gss 37 25 16 62 53 78 1
ρg,τl gss 37 25 16 62 53 78 1
ρτl,τx gss 37 25 16 62 53 78 1
ρτl,g gss 37 25 16 62 53 78 1
q21 gss 37 25 16 62 53 78 1
q22 gss 37 25 16 62 53 78 1
ψgss 37 25 16 62 53 78 1
gn ρτl,z 37 25 16 62 53 78 1
gz ρτl,z 37 25 16 62 53 78 1
β ρτl,z 37 25 16 62 53 78 1
gn ρz,τl 37 25 16 62 53 78 1
gz ρz,τl 37 25 16 62 53 78 1
β ρz,τl 37 25 16 62 53 78 1
gn ρτx,τl 37 25 16 62 53 78 1
gz ρτx,τl 37 25 16 62 53 78 1
β ρτx,τl 37 25 16 62 53 78 1
gn ρg,τl 37 25 16 62 53 78 1
gz ρg,τl 37 25 16 62 53 78 1
β ρg,τl 37 25 16 62 53 78 1
ρτl,g ρτl,τx 37 25 16 62 53 78 1
gn ρτl,τx 37 25 16 62 53 78 1
gz ρτl,τx 37 25 16 62 53 78 1
β ρτl,τx 37 25 16 62 53 78 1
gn ρτl,g 37 25 16 62 53 78 1
gz ρτl,g 37 25 16 62 53 78 1
β ρτl,g 37 25 16 62 53 78 1
gn q21 37 25 16 62 53 78 1
gz q21 37 25 16 62 53 78 1
β q21 37 25 16 62 53 78 1
gn q22 37 25 16 62 53 78 1
gz q22 37 25 16 62 53 78 1
β q22 37 25 16 62 53 78 1
ψ gn 37 25 16 62 53 78 1
ψ gz 37 25 16 62 53 78 1
ψ β 37 25 16 62 53 78 1
Required 37 25 16 62 53 78 1

Summary: nθ=37,nX=5,nε=4.

Order condition: nθ=37,nδ=50.

4.1.2 Šustek (2011) MBCA Model

The results for the MBCA model by Šustek (2011) resemble the ones just reported for the standard BCA model. Indeed the MBCA model fulfils the rank condition for strict identification established by Komunjer and Ng (2011) at Tol = 1e−11 and at the Matlab default tolerance when the baseline set of estimated parameters is considered (Table 4) or when the latter contains also [τbss,R˜ss] (Table 5). This is no longer true once the deep parameters are included in estimation. Indeed, also in this case, several steady-state wedge shocks and off-diagonal elements of the P and Q matrix are not strictly identifiable, though this is only true when both [τbss,R˜ss] and the deep parameters of the model are included in estimation (Table 6). Indeed, when only the deep parameters of the model are included in estimation on top of the baseline set of parameters considered in Šustek (2011) then this is only true at a low Tol = Default since at Tol = 1e−3 only a few steady-state wedge shocks are found to not meet the requirements for strict identifiability (Table 7). A similar pattern emerges once investment adjustment costs are introduced in the model (see Appendix A).

Table 4.

Komunjer and Ng Test Results MBCA Model.

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 60 49 33 101 89 130 0
e−03 60 49 36 108 96 143 0
e−04 60 49 36 109 96 144 0
e−05 60 49 36 109 96 144 0
e−06 60 49 36 109 96 144 0
e−07 60 49 36 109 96 145 0
e−08 60 49 36 109 96 145 0
e−09 61 49 36 109 97 145 0
e−10 61 49 36 109 97 145 0
Default = 1.165290e−11 61 49 36 110 97 146 1
e−11 61 49 36 110 97 146 1
Required 61 49 36 110 97 146 1

Summary: nθ=61,nX=7,nε=6.

Order condition: nθ=61,nδ=105.

Table 5.

Komunjer and Ng Test Results MBCA Model (τbss and R˜ss Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 62 49 33 103 91 131 0
e−03 62 49 36 110 98 144 0
e−04 62 49 36 111 98 146 0
e−05 62 49 36 111 98 146 0
e−06 62 49 36 111 98 146 0
e−07 62 49 36 111 98 147 0
Default = 1.193257e−08 62 49 36 111 98 147 0
e−08 62 49 36 111 98 147 0
e−09 63 49 36 111 99 147 0
e−10 63 49 36 112 99 147 0
e−11 63 49 36 112 99 148 1
Required 63 49 36 112 99 148 1

Summary: nθ=63,nX=7,nε=6.

Order condition: nθ=63,nδ=105.

Problematic parameters at Tol = 1e−3: zss, gss.

Problematic parameters at Tol = 1.000000e−10: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q21, q31, q51, q22, q32, q42, q52, q33, q53, q63, q44.

Table 6.

Komunjer and Ng Test Results MBCA Model (τbss,R˜ss and Deep Parameters Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 70 49 35 106 100 134 0
e−03 71 49 36 114 107 147 0
e−04 71 49 36 118 107 149 0
e−05 71 49 36 118 107 150 0
e−06 71 49 36 118 107 150 0
e−07 72 49 36 119 108 151 0
Default = 2.386514e−08 72 49 36 120 108 153 0
e−08 72 49 36 120 108 154 0
e−09 73 49 36 120 109 154 0
e−10 73 49 36 121 109 156 0
e−11 74 49 36 122 110 157 0
Required 74 49 36 123 110 159 1

Summary: nθ=74,nX=7,nε=6.

Order condition: nθ=74,nδ=105.

Problematic parameters at Tol = 1e−3: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q21, q31, q51, q22, q32, q42, q52, q33, q43, q53, q63, q44, q54, q64, q55, gn, gz, β, δ, ψ, σ, ρR, ωπ, ωy, πss.

Problematic parameters at Tol = 1.000000e−11: zss, τlss, τxss, gss, τbss, R˜ss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q11, q21, q31, q41, q51, q61, q22, q32, q42, q52, q62, q33, q43, q53, q63, q44, q54, q64, q55, q65, q66, gn, gz, β, δ, ψ, σ, α, ρR, ωπ, ωy, πss.

Table 7.

Komunjer and Ng Test Results MBCA Model (Deep Parameters Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 68 49 35 104 98 133 0
e−03 69 49 36 113 105 146 0
e−04 69 49 36 116 105 148 0
e−05 69 49 36 116 105 148 0
e−06 69 49 36 116 105 148 0
e−07 70 49 36 117 106 149 0
e−08 70 49 36 118 106 152 0
e−09 71 49 36 118 107 153 0
e−10 71 49 36 119 107 154 0
Default = 2.330580e−11 72 49 36 120 108 155 0
e−11 72 49 36 120 108 155 0 Required 72 49 36 121 108 157 1

Summary: nθ=72,nX=7,nε=6.

Order condition: nθ=72,nδ=105.

Problematic parameters at Tol = 1e−3: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q22, q53, gn, gz, β, ψ, σ, ρR, ωπ, ωy, πss.

Problematic parameters at Tol = 1.000000e−11: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q11, q21, q31, q51, q61, q22, q32, q42, q52, q62, q33, q43, q53, q63, q44, q54, q64, q55, q65, q66, gn, gz, β, δ, ψ, σ, α, ρR, ωπ, ωy, πss.

As to the conditional identification tests, we find that when the parameter set also includes the deep parameters of the model it is possible to fix some model parameters so as to make the rest identifiable (see Table 8) only when τbss and R˜SS are not included in estimation. The sets which restrict the least number of parameters are two and are found at Tol = 1e−11, namely (i) {πSS,zSS} and (ii) {πSS,gSS}

Table 8.

Komunjer and Ng Conditional Test Results MBCA Model, Tol = 2.330580e−11.

Fixed ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
πss zss 72 49 36 121 108 157 1
πss gss 72 49 36 121 108 157 1
Required 72 49 36 121 108 157 1

Summary: nθ=72,nX=7,nε=6.

Order condition: nθ=72,nδ=105.

4.2 Iskrev (2010)

In this subsection, we show the results of the parameter identification analysis of the baseline BCA and MBCA models. We only summarize the results on the strict identification tests and then focus on the weak identification analysis.

We start with the standard BCA model in the case where the deep parameters of the model are estimated on top of the ones governing the VAR(1) process of the innovations to the wedges, for a total of 37 parameters. The rank of the information matrix is 33. Using population data it is possible to obtain gn and gz can assumed to be known since in our detrending it set to a value such that average detrended output is equal to 0.4 With 35 parameters, the rank of the information matrix is still 33. By brute force, we check the rank of the information matrix for all possible combinations of 33 parameters out of the 35 (595 such combinations) being restricted and find that: (i) 21 combinations deliver a rank of 31; (ii) 355 combinations give rank of 32; and (iii) 219 combinations result in a rank of 33. Thus, we can fix 219 possible pairs of parameters and identify the ones left unrestricted in the respective cases.

As to the MBCA model, we start with 74 parameters which corresponds to the case where the deep parameter of the model as well as the steady-state innovations to the asset market and monetary policy wedge (τbssandR˜ss) are estimated on top of the ones governing the VAR(1) process. Fixing gn and gz as discussed above leaves 72 parameters and the rank of the information matrix is 68. There are 1,028,790 possible combinations of 68 out of 72 parameters which can be fixed, for 16,652 of them the resulting rank of the information matrix is 68. When all deep parameters are excluded from the analysis, and thus only the wedge parameters are considered 63 parameters are left. In this case, the rank of the information matrix is 62 and fixing any one of the following set of parameters gives a full rank of 62: {τbss,ρτb,z,ρτb,τl,ρτb,τx,ρτb,g,ρz,τb,ρτl,τb,ρτx,τb,ρg,τb,ρR˜,τb,ρτb,R˜,q51,q52,q53,q54,q55}. For instance, fixing τbss as done in Šustek (2011) gives a full rank of 62. If the non-wedge parameters are added 71 parameters are estimated and the rank of the information matrix is 68. There are 57,155 combinations of 68 out of 71 parameters and for 59 of them fixing the relevant parameters gives a rank of 68.

Next, we study whether the parameters are affected by weak identification problems. We report Cramér-Rao lower bounds (CRLBs) as a measure of parameter estimation uncertainty in absolute terms, and relative CRLBs, rCRLB(θi)=CRLB(θi)/abs(θi), as a measure of relative uncertainty. The CRLBs reflect the uncertainty which arises from both the ‘sensitivity’ and the ‘collinearity’ factors. This is because, as discussed in Section 3.4.2, the CRLB is the product of these two components. The first component concerns the sensitivity of the log-likelihood function with respect to θi, whereas the second relates to the degree of collinearity between the derivative of the likelihood with respect to θi and the derivatives of the likelihood with respect to all other free parameters θi.

In order to interpret them it is useful to recall the following facts. The sensitivity factor for a parameter reports the value of the conditional CRLB, that is, the lower bound of uncertainty given that all other parameters are known and no collinearity is present. As to the collinearity factor, it is indicative of the increase in the CRLB when the other parameters are unknown as well. The magnitude of the sensitivity component, unlike the collinearity one, depends on the scale of the parameters and is thus divided by the respective parameter values.

4.2.1 Chari et al. (2007) BCA Model

Table 9 deals with the standard BCA model and the case where the structural parameters are assumed to be known. The first column shows the values of the parameters governing the stochastic process of the wedges, the second one the CRLBs and the third one the relative CRLBs. The latter measure can be used to compare the relative identification strength across parameters which take different values. For instance, zss is much worse identified than gss. This is because the CRLB of zss relative to its value is 6.3701 while only 0.3617 for gss. Further examination of Table 9 suggests that the worst identified parameters, using (arbitrarily) a cut-off value of 1, are zss, ρτl,z,ρτx,z,ρg,z,ρz,τl,ρτx,τl,ρg,τl,ρz,τx,ρτl,τx,ρg,τx,ρz,g,ρτl,g, q21, q31, q41, q32, q42 and q44.

Table 9.

BCA, Parameter Identification (All Deep Parameters Fixed).

Value CRLB rCRLB
zss −0.0239 0.1524 6.3701
τlss 0.3279 0.2513 0.7662
τxss 0.4834 0.3763 0.7783
gss −1.5344 0.5550 0.3617
ρz 0.9800 0.0484 0.0494
ρτl,z −0.0330 0.0631 1.9123
ρτx,z −0.0702 0.1135 1.6161
ρg,z 0.0048 0.0627 13.0322
ρz,τl −0.0138 0.0353 2.5588
ρτl 0.9564 0.0675 0.0706
ρτx,τl −0.0460 0.1346 2.9252
ρg,τl −0.0081 0.0578 7.1346
ρz,τx −0.0117 0.0825 7.0314
ρτl,τx −0.0451 0.0768 1.7024
ρτx 0.8962 0.0994 0.1109
ρg,τx 0.0488 0.0964 1.9744
ρz,g 0.0192 0.0791 4.1117
ρτl,g 0.0569 0.0760 1.3357
ρτx,g 0.1041 0.0866 0.8321
ρg 0.9711 0.0907 0.0934
q11 0.0116 0.0007 0.0578
q21 0.0014 0.0028 2.0187
q31 −0.0105 0.0111 1.0555
q41 −0.0006 0.0013 2.2905
q22 0.0064 0.0004 0.0633
q32 0.0010 0.0099 9.6266
q42 0.0061 0.0085 1.3946
q33 0.0158 0.0110 0.6912
q43 0.0142 0.0034 0.2412
q44 0.0046 0.0047 1.0269

Table 10 considers the case where all parameters in the BCA model are estimated with the exception of four out of the seven structural parameters in the BCA model being held fixed so as to guarantee strict identification of the remaining parameters. As described above, there are many such combinations which would have guaranteed a full rank information matrix. On top of leaving out gn and gz which we can calculate using additional data we leave out β and ψ from the analysis. Not surprisingly, the relative uncertainty measures increase. Moreover, the set of worst identified parameters includes also τlss, τxss, ρτx,g and q33.

Table 10.

BCA, Parameter Identification (Some Deep Parameters Free).

Value CRLB rCRLB
zss −0.0239 1.7607 73.6039
τlss 0.3279 0.4714 1.4376
τxss 0.4834 1.5796 3.2674
gss −1.5344 0.7628 0.4972
ρz 0.9800 0.0609 0.0622
ρτl,z −0.0330 0.0683 2.0712
ρτx,z −0.0702 0.1159 1.6503
ρg,z 0.0048 0.0695 14.4480
ρz,τl −0.0138 0.0571 4.1391
ρτl 0.9564 0.0750 0.0784
ρτx,τl −0.0460 0.1441 3.1325
ρg,τl −0.0081 0.0830 10.2360
ρz,τx −0.0117 0.1214 10.3566
ρτl,τx −0.0451 0.0798 1.7692
ρτx 0.8962 0.1288 0.1437
ρg,τx 0.0488 0.0970 1.9863
ρz,g 0.0192 0.1046 5.4386
ρτl,g 0.0569 0.1020 1.7917
ρτx,g 0.1041 0.1243 1.1947
ρg 0.9711 0.1141 0.1175
q11 0.0116 0.0057 0.4902
q21 0.0014 0.0040 2.8614
q31 −0.0105 0.0149 1.4217
q41 −0.0006 0.0013 2.3349
q22 0.0064 0.0035 0.5488
q32 0.0010 0.0123 11.8844
q42 0.0061 0.0117 1.9119
q33 0.0158 0.0215 1.3562
q43 0.0142 0.0046 0.3233
q44 0.0046 0.0068 1.4834
δ 0.0118 0.0014 0.1215
σ 1.0000 0.5176 0.5176
α 0.3500 0.3171 0.9059

The ‘sensitivity’ and ‘collinearity’ components of the parameters’ CRLBs for the case where the deep parameters are fixed are reported in Table 11 and labelled as ‘sens.’ and ‘coll.’. It is immediate to see that the channel which drives weak identification is the collinearity one. Indeed, while all parameters exert a strong effect on the likelihood, some of them induce a variation in the likelihood which is very similar to other parameters. For some parameters, the sensitivity factor is so strong that it outweighs the negative effect of the collinearity factor on their overall uncertainty. For others, however, the collinearity factor predominates and leads their relative uncertainty to be high. It is also interesting to take a look at the largest multiple correlation coefficients between lT(θ)/θi and lT(θ)/θi, namely ϱi(n). We report the single and pairwise correlation coefficients in the columns labelled by ϱi(1) and ϱi(2), respectively. The problematic parameters exert an effect on the likelihood which is mostly collinear with the elements of the matrices they belong to in the VAR(1) process of the innovations to the wedges (i.e. P0, P and Q). The fact that the steady-state innovations to the wedges are strongly correlated with elements of both the P0 and P matrix is due to the fact that they are obtained as (IP)1P0.

Table 11.

BCA, Information Matrix Decomposition (Observing 4 Standard Variables).

CRLB/para. sens/para. coll. ϱi ϱi(1) ϱi(2)
zss 6.370 0.575 11.082 0.995920 0.886 (gss) 0.931 (gss, ρg,z)
τlss 0.766 0.018 41.833 0.999714 0.739 (τxss) 0.905 (τxss,ρz,τx)
τxss 0.778 0.009 82.492 0.999927 0.886 (gss) 0.934 (τlss, gss)
gss 0.362 0.006 59.573 0.999859 0.886 (τxss) 0.971 (zss, τxss)
ρz 0.049 0.000 153.529 0.999979 0.988 (ρg,z) 0.991 (ρτl,z,ρg,z)
ρτl,z 1.912 0.030 64.792 0.999881 0.907 (ρz) 0.982 (ρτl,τx,ρτl,g)
ρτx,z 1.616 0.026 62.859 0.999873 0.908 (ρg,z) 0.974 (ρτx,ρτx,g)
ρg,z 13.032 0.097 134.773 0.999972 0.988 (ρz) 0.990 (ρz, ρτx,z)
ρz,τl 2.559 0.014 176.574 0.999984 0.988 (ρg,τl) 0.993 (ρg,τl,ρz,g)
ρτl 0.071 0.001 108.028 0.999957 0.978 (ρτl,g) 0.990 (ρτl,τx,ρτl,g)
ρτx,τl 2.925 0.026 112.667 0.999961 0.969 (ρτx,g) 0.986 (ρτx,ρτx,g)
ρg,τl 7.135 0.036 196.423 0.999987 0.988 (ρz,τl) 0.993 (ρg,τx, ρg)
ρz,τx 7.031 0.012 600.752 0.999999 0.988 (ρg,τx) 0.999 (ρz, ρz,g)
ρτl,τx 1.702 0.009 181.791 0.999985 0.983 (ρτl,g) 0.998 (ρτl,z,ρτl,g)
ρτx 0.111 0.001 126.383 0.999969 0.978 (ρτx,g) 0.997 (ρτx,z,ρτx,g)
ρg,τx 1.974 0.004 474.827 0.999998 0.988 (ρz,τx) 0.999 (ρg,z, ρg)
ρz,g 4.112 0.005 881.973 0.999999 0.988 (ρg) 0.999 (ρz, ρz,τx)
ρτl,g 1.336 0.005 272.329 0.999993 0.983 (ρτl,τx) 0.999 (ρτl,z,ρτl,τx)
ρτx,g 0.832 0.005 163.595 0.999981 0.978 (ρτx) 0.998 (ρτx,z,ρτx)
ρg 0.093 0.000 684.987 0.999999 0.988 (ρz,g) 0.999 (ρg,z,ρg,τx)
q11 0.058 0.036 1.628 0.788970 0.780 (q31) 0.788 (q21, q31)
q21 2.019 0.247 8.160 0.992462 0.747 (q41) 0.756 (q11, q41)
q31 1.056 0.038 27.708 0.999349 0.951 (q41) 0.960 (q11, q41)
q41 2.291 0.654 3.504 0.958421 0.951 (q31) 0.958 (q21, q31)
q22 0.063 0.045 1.399 0.699484 0.622 (q42) 0.632 (τlss, q42)
q32 9.627 0.388 24.829 0.999189 0.951 (q42) 0.951 (τlss, q42)
q42 1.395 0.062 22.658 0.999026 0.951 (q32) 0.955 (q22, q32)
q33 0.691 0.024 28.643 0.999390 0.909 (q43) 0.911 (ρτx, q43)
q43 0.241 0.027 9.087 0.993927 0.909 (q33) 0.909 (zss, q33)
q44 1.027 0.058 17.694 0.998402 0.088 (τlss) 0.173 (τlss,τxss)

Table 12 presents the decomposition of parameter uncertainty into a sensitivity and collinearity component for the case where also some deep parameters are estimated. The conclusions drawn about identification patterns drawn above are largely unaffected. The only key difference is that steady-state parameters are mainly collinear with the deep parameters of the model and vice versa. This is due to the fact that the latter are directly informative about each other via the steady-state equations of the model.

Table 12.

BCA (Some Deep Parameters Estimated), Information Matrix Decomposition.

CRLB/para. sens/para. coll. ϱi ϱi(1) ϱi(2)
zss 73.604 0.575 128.045 0.999970 0.886 (gss) 0.955 (gss, α)
τlss 1.438 0.018 78.490 0.999919 0.887 (α) 0.909 (zss, α)
τxss 3.267 0.009 346.308 0.999996 0.900 (δ) 0.970 (ρg, α)
gss 0.497 0.006 81.884 0.999925 0.886 (τxss) 0.971 (zss, τxss)
ρz 0.062 0.000 193.365 0.999987 0.988 (ρg,z) 0.991 (ρτl,z,ρg,z)
ρτl,z 2.071 0.030 70.178 0.999898 0.907 (ρz) 0.982 (ρτl,τx,ρτl,g)
ρτx,z 1.650 0.026 64.190 0.999879 0.908 (ρg,z) 0.974 (ρτx,ρτx,g)
ρg,z 14.448 0.097 149.415 0.999978 0.988 (ρz) 0.990 (ρz, ρτx,z)
ρz,τl 4.139 0.014 285.622 0.999994 0.988 (ρg,τl) 0.993 (ρg,τl,ρz,g)
ρτl 0.078 0.001 120.051 0.999965 0.978 (ρτl,g) 0.990 (ρτl,τx,ρτl,g)
ρτx,τl 3.132 0.026 120.650 0.999966 0.969 (ρτx,g) 0.986 (ρτx,ρτx,g)
ρg,τl 10.236 0.036 281.808 0.999994 0.988 (ρz,τl) 0.993 (ρg,τx, ρg)
ρz,τx 10.357 0.012 884.848 0.999999 0.988 (ρg,τx) 0.999 (ρz, ρz,g)
ρτl,τx 1.769 0.009 188.922 0.999986 0.983 (ρτl,g) 0.998 (ρτl,z,ρτl,g)
ρτx 0.144 0.001 163.762 0.999981 0.978 (ρτx,g) 0.997 (ρτx,z,ρτx,g)
ρg,τx 1.986 0.004 477.688 0.999998 0.988 (ρz,τx) 0.999 (ρg,z, ρg)
ρz,g 5.439 0.005 1166.579 1.000000 0.988 (ρg) 0.999 (ρz, ρz,τx)
ρτl,g 1.792 0.005 365.308 0.999996 0.983 (ρτl,τx) 0.999 (ρτl,z,ρτl,τx)
ρτx,g 1.195 0.005 234.875 0.999991 0.978 (ρτx) 0.998 (ρτx,z,ρτx)
ρg 0.118 0.000 862.241 0.999999 0.988 (ρz,g) 0.999 (ρg,z,ρg,τx)
q11 0.490 0.036 13.799 0.997371 0.780 (q31) 0.788 (q21, q31)
q21 2.861 0.247 11.566 0.996255 0.747 (q41) 0.756 (q11, q41)
q31 1.422 0.038 37.321 0.999641 0.951 (q41) 0.960 (q11, q41)
q41 2.335 0.654 3.572 0.960019 0.951 (q31) 0.958 (q21, q31)
q22 0.549 0.045 12.138 0.996600 0.622 (q42) 0.632 (τlss, q42)
q32 11.884 0.388 30.653 0.999468 0.951 (q42) 0.951 (τlss, q42)
q42 1.912 0.062 31.063 0.999482 0.951 (q32) 0.955 (q22, q32)
q33 1.356 0.024 56.198 0.999842 0.909 (q43) 0.911 (ρτx, q43)
q43 0.323 0.027 12.183 0.996626 0.909 (q33) 0.909 (zss, q33)
q44 1.483 0.058 25.559 0.999234 0.408 (σ) 0.660 (τxss, σ)
δ 0.122 0.022 5.444 0.982983 0.900 (τxss) 0.949 (zss, α)
σ 0.518 0.015 35.581 0.999605 0.752 (τxss) 0.868 (τxss, q44)
α 0.906 0.003 292.762 0.999994 0.887 (τlss) 0.973 (τxss, ρg)

We thus conclude that weak identification can be attributed to the fact that parameters in the steady-state vector, in the autoregressive matrix and in standard deviations of the fundamental innovations have a similar effect on the likelihood as other parameters appearing within the matrices they belong to. In order to understand how this pattern comes about we perform the following experiment. First, we assume that the observed variables are not output, investment, hours and government consumption (i.e. logy^t,logx^t,loglt,logg^t) but rather the four innovations to the wedges (i.e.log(zt),τlt,τxt,log(g^t)). Second, we assume that the parameters governing the VAR(1) process of the innovations to the wedges are estimated in isolation of the BCA model. In other words, we perform identification analysis on the parameters assuming that a VAR is run directly on the wedges, which are treated as observable in this experiment. This allows us to shed light on the question whether the weak identification patterns described above are an intrinsic property of the model or of the VAR(1) process itself. Table 13 presents the information matrix decomposition for the case just described.

Table 13.

BCA Model, Information Matrix Decomposition (Observing 4 Wedges).

CRLB/para. sens/para. coll. ϱi ϱi(1) ϱi(2)
zss 8.528 0.137 62.163 0.999871 0.997 (gss) 0.998 (τlss,gss)
τlss 1.062 0.013 79.708 0.999921 0.985 (gss) 0.998 (τxss,gss)
τxss 1.007 0.004 251.762 0.999992 0.999 (gss) 1.000 (τlss,gss)
gss 0.501 0.001 390.297 0.999997 0.999 (τxss) 1.000 (τlss,τxss)
ρz 0.053 0.002 30.412 0.999459 0.937 (ρz,τl) 0.994 (ρz,τx,ρz,g)
ρτl,z 0.893 0.042 21.480 0.998916 0.938 (ρτl) 0.993 (ρτl,τx,ρτl,g)
ρτx,z 1.206 0.023 51.881 0.999814 0.935 (ρτx,τl) 0.993 (ρτx,ρτx,g)
ρg,z 15.015 0.323 46.481 0.999769 0.934 (ρg,τl) 0.993 (ρg,τx, ρg)
ρz,τl 2.922 0.074 39.262 0.999676 0.991 (ρz,g) 0.996 (ρz,τx,ρz,g)
ρτl 0.024 0.001 27.282 0.999328 0.990 (ρτl,g) 0.996 (ρτl,τx,ρτl,g)
ρτx,τl 1.433 0.022 65.457 0.999883 0.990 (ρτx,g) 0.996 (ρτx,ρτx,g)
ρg,τl 6.950 0.118 58.772 0.999855 0.989 (ρg) 0.995 (ρg,τx, ρg)
ρz,τx 7.522 0.061 123.966 0.999967 0.992 (ρz,g) 0.999 (ρz, ρz,g)
ρτl,τx 1.110 0.013 84.669 0.999930 0.992 (ρτl,g) 0.999 (ρτl,z,ρτl,g)
ρτx 0.161 0.001 204.424 0.999988 0.991 (ρτx,g) 0.999 (ρτx,z,ρτx,g)
ρg,τx 2.512 0.014 182.688 0.999985 0.991 (ρg) 0.999 (ρg,z, ρg)
ρz,g 4.432 0.024 185.122 0.999985 0.992 (ρz,τx) 1.000 (ρz, ρz,τx)
ρτl,g 0.851 0.007 127.227 0.999969 0.992 (ρτl,τx) 1.000 (ρτl,z,ρτl,τx)
ρτx,g 1.341 0.004 306.941 0.999995 0.991 (ρτx) 1.000 (ρτx,z,ρτx)
ρg 0.122 0.000 274.011 0.999993 0.991 (ρg,τx) 1.000 (ρg,z,ρg,τx)
q11 0.058 0.035 1.630 0.789720 0.781 (q31) 0.789 (q21, q31)
q21 0.378 0.247 1.530 0.756825 0.748 (q41) 0.757 (q11, q41)
q31 0.137 0.038 3.593 0.960488 0.951 (q41) 0.960 (q11, q41)
q41 2.292 0.652 3.515 0.958676 0.951 (q31) 0.959 (q21, q31)
q22 0.058 0.045 1.282 0.625594 0.624 (q42) 0.625 (q32, q42)
q32 1.258 0.387 3.253 0.951570 0.951 (q42) 0.952 (q22, q42)
q42 0.208 0.061 3.384 0.955345 0.951 (q32) 0.955 (q22, q32)
q33 0.058 0.024 2.404 0.909374 0.909 (q43) 0.909 (ρτl,τx, q43)
q43 0.064 0.026 2.404 0.909367 0.909 (q33) 0.909 (ρτl,τx, q33)
q44 0.058 0.058 1.000 0.008064 0.002 (q31) 0.003 (ρg,τl, ρg)

The main takeaways are that (i) the collinearity patterns are preserved and (ii) for almost all steady-state innovations to the wedges, collinearity is higher. The intuition for result (ii) is that by fixing deep parameters the model imposes additional restrictions on the steady states of the innovations to the wedges via the steady-state equations. This results in additional information and, therefore, improved identification strength.

To further develop intuition and visualize our findings, we plot pairwise correlations coefficients between lT(θ)/θi and lT(θ)/θi across the entire parameter space in Figs. 1 and 2. Clearly, correlation coefficients between the derivatives with respect to the same parameter are equal to 1, marked as dark red squares on the main diagonal. The following results emerge. First, the figures confirm the main finding that the steady-state innovations to the wedges (IP)1P0, the elements of the autoregressive matrix P and the Choleski decomposition of the variance covariance matrix Q induce variation in the likelihood which is mainly similar to that generated by other parameters in the matrix they belong to. Indeed, in Fig. 2 the strongest collinearity is among elements of the same matrix and only some elements of the P and Q have a weakly collinear effect. Second, collinearity is less pervasive when innovations to the wedges are assumed to be observed and the parameters are estimated using the VAR in isolation from the model. The additional links between parameters and their likelihood introduced by the model equations are thus not sufficiently rich to let each parameter in the VAR(1) matrices P0,P and Q raise its own, distinct voice. In other words, estimating the wedge parameters within a model does not obviate the collinearity patterns which emerge in isolation from it.

Fig. 1. BCA, Pairwise Correlations (Observing 4 Standard Variables).

Fig. 1.

BCA, Pairwise Correlations (Observing 4 Standard Variables).

Fig. 2. BCA, Pairwise Correlations (Observing 4 Wedges).

Fig. 2.

BCA, Pairwise Correlations (Observing 4 Wedges).

4.2.2 Šustek (2011) MBCA Model

Tables 14 and 15 contain information about the relative CRLBs as well as about the sensitivity and collinearity components for the MBCA model wedge parameters. Using again a cut-off value of 1 the set of worst identified parameters is given by zss, ρτl,z,ρτx,z,ρg,z,ρτb,z,ρR˜,z,ρz,τl,ρτx,τl,ρg,τl,ρτb,τl,ρR˜,τl,ρz,τx,ρτl,τx,ρg,τx,ρτb,τx,ρR˜,τx,ρz,g,ρτl,g,ρτx,g,ρτb,g,ρR˜,g,ρz,τb,ρτl,τb,ρτx,τb,ρg,τb,ρR˜,τb,ρτx,R˜,ρg,R˜,ρτb,R˜, q41, q51, q61, q32, q52, q53, q63, q54, q64 and q66. This amount to 39 out of 61 parameters, that is, roughly two thirds of the estimated parameters are weakly identified. As in the standard BCA model these parameters concern the steady-state innovation to the efficiency wedge zss as well as the off-diagonal elements of the P and Q matrices. Identification deficiencies can again be attributed to the strong collinearity between the effects that parameters within the same matrix exert on the likelihood. We focus on the case where some deep parameters are included in the identification analysis in Tables 16 and 17. Also in this case, we keep fixed the other deep parameters so as to guarantee a full rank information matrix. Among the many available choices we choose to leave out β, ψ as well as steady-state inflation πss and the weight on inflation in the Taylor rule ωπ. Again, relative CRLBs increase as expected. Additional parameters which are weakly identified are τlss,τxss,ρz,R˜, q31, q33, q43, δ, σ, α, ωπ. The collinearity patterns are similar to the ones in the BCA model.

Table 14.

MBCA, Parameter Identification (All Deep Parameters, τbss and R˜ss Fixed).

Value CRLB rCRLB
zss −0.0246 0.1333 5.4109
τlss 0.1267 0.0345 0.2720
τxss 0.4649 0.1310 0.2818
gss −1.5389 0.0592 0.0385
R˜ss 0.0000 0.0023 0.0023
ρz 0.8541 0.2833 0.3317
ρτl,z −0.0437 0.1504 3.4424
ρτx,z −0.0557 0.1393 2.5014
ρg,z 0.0533 0.1774 3.3285
ρτb,z 0.0633 0.4210 6.6544
ρR˜,z −0.0141 0.0229 1.6232
ρz,τl −0.1482 0.4083 2.7553
ρτl 1.0580 0.2108 0.1993
ρτx,τl −0.0335 0.1629 4.8637
ρg,τl 0.0587 0.2139 3.6444
ρτb,τl −0.2979 0.6935 2.3279
ρR˜,τl 0.0146 0.0295 2.0178
ρz,τx 0.2654 0.3643 1.3724
ρτl,τx −0.0014 0.2048 146.3072
ρτx 1.0877 0.1840 0.1692
ρg,τx −0.0974 0.2229 2.2881
ρτb,τx 0.0850 0.6202 7.2964
ρR˜,τx 0.0005 0.0358 71.6947
ρz,g −0.0094 0.1084 11.4917
ρτl,g 0.0097 0.0619 6.3770
ρτx,g 0.0026 0.0326 12.5353
ρg 1.0053 0.0479 0.0477
ρτb,g −0.0076 0.2253 29.6442
ρR˜,g 0.0004 0.0088 21.8878
ρz,τb −0.0654 0.1679 2.5674
ρτl,τb 0.0465 0.0865 1.8607
ρτx,τb −0.0116 0.0763 6.5772
ρg,τb 0.0241 0.0940 3.9020
ρτb 0.8263 0.2712 0.3282
ρR˜,τb 0.0063 0.0137 2.1719
ρz,R˜ 0.7994 0.7968 0.9967
ρτl,R˜ −0.7219 0.4557 0.6313
ρτx,R˜ 0.4016 0.5299 1.3194
ρg,R˜ 0.3411 0.5499 1.6123
ρτb,R˜ 0.1200 1.2617 10.5144
ρR˜ 0.4412 0.1101 0.2495
q11 0.0110 0.0006 0.0579
q21 0.0037 0.0009 0.2430
q31 0.0058 0.0024 0.4065
q41 0.0009 0.0013 1.4043
q51 0.0005 0.0100 19.9724
q61 0.0003 0.0003 1.1280
q22 0.0092 0.0007 0.0717
q32 −0.0008 0.0028 3.5125
q42 0.0050 0.0014 0.2832
q52 −0.0175 0.0189 1.0819
q62 0.0000 0.0003 0.0003
q33 0.0029 0.0022 0.7422
q43 0.0117 0.0050 0.4255
q53 −0.0013 0.0237 18.2111
q63 0.0001 0.0033 32.6605
q44 0.0087 0.0067 0.7715
q54 0.0014 0.0292 20.8730
q64 −0.0002 0.0045 22.4300
q55 0.0219 0.0116 0.5293
q65 0.0040 0.0005 0.1246
q66 0.0010 0.0012 1.1923
Table 15.

MBCA, Information Matrix Decomposition (Observing 6 Standard Variables).

CRLB/para. sens/para. coll. ϱi ϱi(1) ϱi(2)
zss 5.411 0.437 12.388 0.996736 0.923 (τxss) 0.945 (τxss,R˜ss)
τlss 0.272 0.079 3.436 0.956717 0.451 (gss) 0.630 (zss, gss)
τxss 0.282 0.017 16.867 0.998241 0.923 (zss) 0.955 (zss, R˜ss)
gss 0.038 0.009 4.104 0.969858 0.736 (τxss) 0.850 (zss, τlss)
R˜ss 0.002 0.000 5.086 0.980482 0.496 (τxss) 0.676 (zss, τxss)
ρz 0.332 0.001 420.889 0.999997 0.993 (ρz,τx) 0.999 (ρz,τx,ρz,g)
ρτl,z 3.442 0.018 187.343 0.999986 0.989 (ρτl,τx) 0.997 (ρτl,τx,ρτl,g)
ρτx,z 2.501 0.005 456.118 0.999998 0.992 (ρτx) 0.998 (ρτx,ρτx,g)
ρg,z 3.329 0.021 159.247 0.999980 0.990 (ρg,τx) 0.997 (ρg,τx, ρg)
ρτb,z 6.654 0.035 191.861 0.999986 0.979 (ρτb,τx) 0.989 (ρτb,τx,ρτb,g)
ρR˜,z 1.623 0.017 94.732 0.999944 0.990 (ρR˜,τx) 0.996 (ρR˜,τx,ρR˜,g)
ρz,τl 2.755 0.014 194.002 0.999987 0.917 (ρτx,τl) 0.982 (ρτx,τl,ρτb,τl)
ρτl 0.199 0.002 118.928 0.999965 0.790 (ρτb,τl) 0.965 (ρτl,τx,ρτl,τb)
ρτx,τl 4.864 0.024 201.703 0.999988 0.949 (ρg,τl) 0.986 (ρz,τl,ρτb,τl)
ρg,τl 3.644 0.045 80.899 0.999924 0.949 (ρτx,τl) 0.975 (ρg,τx,ρg,τb)
ρτb,τl 2.328 0.013 185.382 0.999985 0.790 (ρτl) 0.939 (ρτb,τx,ρτb)
ρR˜,τl 2.018 0.045 44.361 0.999746 0.675 (ρR˜,g) 0.931 (ρR˜,τx,ρR˜,τb)
ρz,τx 1.372 0.003 435.327 0.999997 0.993 (ρz) 0.999 (ρz, ρz,g)
ρτl,τx 146.307 0.696 210.110 0.999989 0.989 (ρτl,z) 0.997 (ρτl,z,ρτl,g)
ρτx 0.169 0.000 488.457 0.999998 0.992 (ρτx,z) 0.998 (ρτx,z,ρτx,g)
ρg,τx 2.288 0.014 163.487 0.999981 0.990 (ρg,z) 0.998 (ρg,z, ρg)
ρτb,τx 7.296 0.032 231.052 0.999991 0.979 (ρτb,z) 0.990 (ρτb,z,ρτb,g)
ρR˜,τx 71.695 0.605 118.481 0.999964 0.990 (ρR˜,z) 0.996 (ρR˜,z,ρR˜,g)
ρz,g 11.492 0.081 141.658 0.999975 0.929 (ρτx,g) 0.988 (ρτx,g,ρτb,g)
ρτl,g 6.377 0.072 88.102 0.999936 0.824 (ρτb,g) 0.921 (ρτb,g,ρR˜,g)
ρτx,g 12.535 0.120 104.264 0.999954 0.937 (ρg) 0.989 (ρz,g,ρτb,g)
ρg 0.048 0.001 47.038 0.999774 0.937 (ρτx,g) 0.952 (ρz,g,ρτb,g)
ρτb,g 29.644 0.222 133.747 0.999972 0.824 (ρτl,g) 0.920 (ρz,g,ρτx,g)
ρR˜,g 21.888 0.659 33.218 0.999547 0.675 (ρR˜,τl) 0.820 (ρR˜,z,ρR˜,τx)
ρz,τb 2.567 0.014 180.589 0.999985 0.918 (ρτx,τb) 0.982 (ρτx,τb,ρτb)
ρτl,τb 1.861 0.017 107.845 0.999957 0.767 (ρτb) 0.967 (ρτl,ρτl,τx)
ρτx,τb 6.577 0.031 212.675 0.999989 0.966 (ρg,τb) 0.986 (ρz,τb,ρg,τb)
ρg,τb 3.902 0.049 78.892 0.999920 0.966 (ρτx,τb) 0.977 (ρg,τl,ρg,τx)
ρτb 0.328 0.002 165.624 0.999982 0.767 (ρτl,τb) 0.947 (ρτb,τl,ρτb,τx)
ρR˜,τb 2.172 0.046 47.425 0.999778 0.528 (ρR˜,z) 0.932 (ρR˜,τl,ρR˜,τx)
ρz,R˜ 0.997 0.038 25.915 0.999255 0.889 (ρτx,R˜) 0.953 (ρτl,R˜,ρτx,R˜)
ρτl,R˜ 0.631 0.039 16.073 0.998063 0.500 (ρR˜) 0.871 (ρτb,R˜,ρR˜)
ρτx,R˜ 1.319 0.038 34.906 0.999590 0.899 (ρg,R˜) 0.969 (ρz,R˜,ρg,R˜)
ρg,R˜ 1.612 0.164 9.860 0.994844 0.899 (ρτx,R˜) 0.924 (ρz,R˜,ρτx,R˜)
ρτb,R˜ 10.514 0.367 28.685 0.999392 0.481 (ρR˜) 0.868 (ρτl,R˜,ρR˜)
ρR˜ 0.249 0.016 15.685 0.997965 0.846 (q55) 0.941 (ρR˜,z, q55)
q11 0.058 0.016 3.584 0.960297 0.937 (q31) 0.960 (q21, q31)
q21 0.243 0.050 4.858 0.978583 0.939 (q51) 0.968 (q11, q51)
q31 0.407 0.021 19.169 0.998638 0.937 (q11) 0.950 (q11, q41)
q41 1.404 0.720 1.951 0.858718 0.840 (q31) 0.842 (q21, q31)
q51 19.972 0.870 22.955 0.999051 0.970 (q61) 0.979 (q21, q61)
q61 1.128 0.273 4.131 0.970260 0.970 (q51) 0.970 (q31, q51)
q22 0.072 0.020 3.675 0.962262 0.912 (q52) 0.939 (q42, q52)
q32 3.512 0.154 22.840 0.999041 0.840 (q42) 0.884 (ρτb,τl, q42)
q42 0.283 0.129 2.187 0.889370 0.840 (q32) 0.842 (q22, q32)
q52 1.082 0.025 43.524 0.999736 0.970 (q62) 0.976 (q22, q62)
q62 0.000 0.000 4.237 0.971743 0.970 (q52) 0.970 (q32, q52)
q33 0.742 0.038 19.698 0.998711 0.745 (q43) 0.786 (ρτl,g, q43)
q43 0.426 0.055 7.692 0.991514 0.745 (q33) 0.756 (ρτl,g, q33)
q53 18.211 0.335 54.425 0.999831 0.970 (q63) 0.973 (ρg,R˜, q63)
q63 32.661 0.819 39.874 0.999685 0.970 (q53) 0.973 (ρg,R˜, q53)
q44 0.772 0.055 13.997 0.997445 0.515 (ρτl,g) 0.580 (ρz,g,ρτx,g)
q54 20.873 0.311 67.113 0.999889 0.970 (q64) 0.976 (ρg,R˜, q64)
q64 22.430 0.410 54.717 0.999833 0.970 (q54) 0.976 (ρg,R˜, q54)
q55 0.529 0.019 27.447 0.999336 0.943 (q65) 0.956 (ρR˜, q65)
q65 0.125 0.020 6.091 0.986429 0.943 (q55) 0.943 (ρτl,R˜, q55)
q66 1.192 0.058 20.582 0.998819 0.242 (ρR˜) 0.454 (ρR˜, q55)
Table 16.

MBCA, Parameter Identification (Some Deep Parameters Free).

Value CRLB rCRLB
zss −0.0246 2.1912 88.9226
τlss 0.1267 0.4744 3.7431
τxss 0.4649 6.2711 13.4892
gss −1.5389 0.0694 0.0451
R˜ss 0.0000 0.0023 0.0023
ρz 0.8541 0.5182 0.6068
ρτl,z −0.0437 0.3059 6.9991
ρτx,z −0.0557 0.1738 3.1195
ρg,z 0.0533 0.3181 5.9690
ρτb,z 0.0633 0.7185 11.3585
ρR˜,z −0.0141 0.0880 6.2419
ρz,τl −0.1482 0.4659 3.1440
ρτl 1.0580 0.3470 0.3280
ρτx,τl −0.0335 0.9381 28.0039
ρg,τl 0.0587 0.2177 3.7081
ρτb,τl −0.2979 1.2796 4.2953
ρR˜,τl 0.0146 0.1146 7.8510
ρz,τx 0.2654 0.4307 1.6228
ρτl,τx −0.0014 0.4689 334.9210
ρτx 1.0877 0.5609 0.5157
ρg,τx −0.0974 0.2500 2.5668
ρτb,τx 0.0850 1.1771 13.8482
ρR˜,τx 0.0005 0.0920 184.0737
ρz,g −0.0094 0.5517 58.5039
ρτl,g 0.0097 0.1280 13.1959
ρτx,g 0.0026 0.2396 92.1551
ρg 1.0053 0.2162 0.2151
ρτb,g −0.0076 0.2403 31.6135
ρR˜,g 0.0004 0.0235 58.7368
ρz,τb −0.0654 0.5120 7.8303
ρτl,τb 0.0465 0.0947 2.0375
ρτx,τb −0.0116 0.6453 55.6283
ρg,τb 0.0241 0.1851 7.6786
ρτb 0.8263 0.5197 0.6289
ρR˜,τb 0.0063 0.0362 5.7450
ρz,R˜ 0.7994 2.0086 2.5127
ρτl,R˜ −0.7219 0.5524 0.7652
ρτx,R˜ 0.4016 7.3077 18.1964
ρg,R˜ 0.3411 0.7627 2.2359
ρτb,R˜ 0.1200 3.9863 33.2192
ρR˜ 0.4412 0.1855 0.4204
q11 0.0110 0.0046 0.4202
q21 0.0037 0.0012 0.3225
q31 0.0058 0.0188 3.2388
q41 0.0009 0.0014 1.5813
q51 0.0005 0.0187 37.4840
q61 0.0003 0.0005 1.6802
q22 0.0092 0.0061 0.6675
q32 −0.0008 0.0189 23.6018
q42 0.0050 0.0015 0.2959
q52 −0.0175 0.0225 1.2844
q62 0.0000 0.0004 0.0004
q33 0.0029 0.0202 6.9657
q43 0.0117 0.0815 6.9617
q53 −0.0013 0.1007 77.4485
q63 0.0001 0.0112 111.8444
q44 0.0087 0.1096 12.5977
q54 0.0014 0.0854 61.0234
q64 −0.0002 0.0184 92.0608
q55 0.0219 0.0165 0.7539
q65 0.0040 0.0029 0.7225
q66 0.0010 0.0022 2.2075
gn 0.0037 0.0923 24.9454
gz 0.0040 0.0112 2.8109
δ 0.0118 0.1014 8.5917
σ 1.0000 2.7391 2.7391
α 0.3500 0.3514 1.0039
ρR 0.7500 0.2378 0.3170
ωπ 1.5000 1.8906 1.2604
Table 17.

MBCA (Some Deep Parameters Estimated), Information Matrix Decomposition.

CRLB/para. sens/para. coll. ϱi ϱi(1) ϱi(2)
zss 88.923 0.437 203.582 0.999988 0.967 (α) 0.985 (δ, α)
τlss 3.743 0.079 47.283 0.999776 0.451 (gss) 0.675 (gss, α)
τxss 13.489 0.017 807.496 0.999999 0.981 (gn) 0.990 (gn, δ)
gss 0.045 0.009 4.807 0.978125 0.736 (τxss) 0.853 (τlss, α)
R˜ss 0.002 0.000 5.282 0.981918 0.566 (gn) 0.835 (gn, α)
ρz 0.607 0.001 769.983 0.999999 0.993 (ρz,τx) 0.999 (ρz,τx,ρz,g)
ρτl,z 6.999 0.018 380.905 0.999997 0.989 (ρτl,τx) 0.997 (ρτl,τx,ρτl,g)
ρτx,z 3.120 0.005 568.821 0.999998 0.992 (ρτx) 0.998 (ρτx,ρτx,g)
ρg,z 5.969 0.021 285.575 0.999994 0.990 (ρg,τx) 0.997 (ρg,τx, ρg)
ρτb,z 11.358 0.035 327.492 0.999995 0.979 (ρτb,τx) 0.989 (ρτb,τx,ρτb,g)
ρR˜,z 6.242 0.017 364.272 0.999996 0.990 (ρR˜,τx) 0.996 (ρR˜,τx,ρR˜,g)
ρz,τl 3.144 0.014 221.371 0.999990 0.917 (ρτx,τl) 0.982 (ρτx,τl,ρτb,τl)
ρτl 0.328 0.002 195.729 0.999987 0.790 (ρτb,τl) 0.965 (ρτl,τx,ρτl,τb)
ρτx,τl 28.004 0.024 1,161.364 1.000000 0.949 (ρg,τl) 0.986 (ρz,τl,ρτb,τl)
ρg,τl 3.708 0.045 82.311 0.999926 0.949 (ρτx,τl) 0.975 (ρg,τx,ρg,τb)
ρτb,τl 4.295 0.013 342.062 0.999996 0.790 (ρτl) 0.939 (ρτb,τx,ρτb)
ρR˜,τl 7.851 0.045 172.608 0.999983 0.675 (ρR˜,g) 0.931 (ρR˜,τx,ρR˜,τb)
ρz,τx 1.623 0.003 514.749 0.999998 0.993 (ρz) 0.999 (ρz, ρz,g)
ρτl,τx 334.921 0.696 480.977 0.999998 0.989 (ρτl,z) 0.997 (ρτl,z,ρτl,g)
ρτx 0.516 0.000 1,489.071 1.000000 0.992 (ρτx,z) 0.998 (ρτx,z,ρτx,g)
ρg,τx 2.567 0.014 183.404 0.999985 0.990 (ρg,z) 0.998 (ρg,z, ρg)
ρτb,τx 13.848 0.032 438.524 0.999997 0.979 (ρτb,z) 0.990 (ρτb,z,ρτb,g)
ρR˜,τx 184.074 0.605 304.196 0.999995 0.990 (ρR˜,z) 0.996 (ρR˜,z,ρR˜,g)
ρz,g 58.504 0.081 721.174 0.999999 0.929 (ρτx,g) 0.988 (ρτx,g,ρτb,g)
ρτl,g 13.196 0.072 182.309 0.999985 0.824 (ρτb,g) 0.921 (ρτb,g,ρR˜,g)
ρτx,g 92.155 0.120 766.509 0.999999 0.937 (ρg) 0.989 (ρz,g,ρτb,g)
ρg 0.215 0.001 212.213 0.999989 0.937 (ρτx,g) 0.952 (ρz,g,ρτb,g)
ρτb,g 31.613 0.222 142.631 0.999975 0.824 (ρτl,g) 0.920 (ρz,g,ρτx,g)
ρR˜,g 58.737 0.659 89.143 0.999937 0.675 (ρR˜,τl) 0.820 (ρR˜,z,ρR˜,τx)
ρz,τb 7.830 0.014 550.781 0.999998 0.918 (ρτx,τb) 0.982 (ρτx,τb,ρτb)
ρτl,τb 2.038 0.017 118.095 0.999964 0.767 (ρτb) 0.967 (ρτl,ρτl,τx)
ρτx,τb 55.628 0.031 1,798.754 1.000000 0.966 (ρg,τb) 0.986 (ρz,τb,ρg,τb)
ρg,τb 7.679 0.049 155.247 0.999979 0.966 (ρτx,τb) 0.977 (ρg,τl,ρg,τx)
ρτb 0.629 0.002 317.345 0.999995 0.767 (ρτl,τb) 0.947 (ρτb,τl,ρτb,τx)
ρR˜,τb 5.745 0.046 125.447 0.999968 0.528 (ρR˜,z) 0.932 (ρR˜,τl,ρR˜,τx)
ρz,R˜ 2.513 0.038 65.329 0.999883 0.889 (ρτx,R˜) 0.953 (ρτl,R˜,ρτx,R˜)
ρτl,R˜ 0.765 0.039 19.483 0.998682 0.500 (ρR˜) 0.871 (ρτb,R˜,ρR˜)
ρτx,R˜ 18.196 0.038 481.414 0.999998 0.899 (ρg,R˜) 0.969 (ρz,R˜,ρg,R˜)
ρg,R˜ 2.236 0.164 13.674 0.997322 0.899 (ρτx,R˜) 0.924 (ρz,R˜,ρτx,R˜)
ρτb,R˜ 33.219 0.367 90.628 0.999939 0.593 (ρR) 0.868 (ρτl,R˜,ρR˜)
ρR˜ 0.420 0.016 26.429 0.999284 0.850 (ρR) 0.950 (q55, ωπ)
q11 0.420 0.016 26.007 0.999260 0.937 (q31) 0.960 (q21, q31)
q21 0.323 0.050 6.449 0.987903 0.939 (q51) 0.968 (q11, q51)
q31 3.239 0.021 152.717 0.999979 0.937 (q11) 0.950 (q11, q41)
q41 1.581 0.720 2.197 0.890440 0.840 (q31) 0.842 (q21, q31)
q51 37.484 0.870 43.082 0.999731 0.970 (q61) 0.979 (q21, q61)
q61 1.680 0.273 6.153 0.986706 0.970 (q51) 0.970 (q51, ωπ)
q22 0.668 0.020 34.190 0.999572 0.912 (q52) 0.939 (q42, q52)
q32 23.602 0.154 153.473 0.999979 0.840 (q42) 0.884 (ρτb,τl, q42)
q42 0.296 0.129 2.285 0.899189 0.840 (q32) 0.842 (q22, q32)
q52 1.284 0.025 51.670 0.999813 0.970 (q62) 0.976 (q22, q62)
q62 0.000 0.000 4.512 0.975134 0.970 (q52) 0.970 (q32, q52)
q33 6.966 0.038 184.862 0.999985 0.745 (q43) 0.786 (ρτl,g, q43)
q43 6.962 0.055 125.846 0.999968 0.745 (q33) 0.756 (ρτl,g, q33)
q53 77.448 0.335 231.461 0.999991 0.970 (q63) 0.973 (ρg,R˜, q63)
q63 111.844 0.819 136.547 0.999973 0.970 (q53) 0.973 (ρg,R˜, q53)
q44 12.598 0.055 228.544 0.999990 0.515 (ρτl,g) 0.627 (ρτl,g, σ)
q54 61.023 0.311 196.210 0.999987 0.970 (q64) 0.976 (ρg,R˜, q64)
q64 92.061 0.410 224.580 0.999990 0.970 (q54) 0.976 (ρg,R˜, q54)
q55 0.754 0.019 39.094 0.999673 0.943 (q65) 0.956 (ρR˜, q65)
q65 0.722 0.020 35.319 0.999599 0.943 (q55) 0.952 (q55, ρR)
q66 2.208 0.058 38.106 0.999656 0.535 (ωπ) 0.729 (ρR, ωπ)
gn 24.945 0.037 666.842 0.999999 0.981 (τxss) 0.988 (τxss,ρg,z)
gz 2.811 0.120 23.506 0.999095 0.961 (δ) 0.967 (gn, δ)
δ 8.592 0.039 222.782 0.999990 0.961 (gz) 0.979 (zss, gz)
σ 2.739 0.016 174.691 0.999984 0.557 (zss) 0.723 (zss, q44)
α 1.004 0.003 370.406 0.999996 0.974 (τxss) 0.990 (zss, τxss)
ρR 0.317 0.006 52.055 0.999815 0.850 (ρR˜) 0.902 (ρτl,R˜,ρR˜)
ωπ 1.260 0.020 63.537 0.999876 0.738 (ρR) 0.886 (q66, ρR)

Finally, we carry out the identification analysis on the wedge parameters with the twist of treating the innovations to the wedges as observables for the MBCA model as well. We again find that (i) parameters of the same matrices having a collinear effect on the likelihood is a property of the VAR process itself (Table 18) and (ii) collinearity is less pervasive when the parameters are estimated in isolation to the model (Figs. 3 and 4).

Fig. 3. MBCA, Pairwise Correlations (Observing 6 Standard Variables).

Fig. 3.

MBCA, Pairwise Correlations (Observing 6 Standard Variables).

Fig. 4. MBCA, Pairwise Correlations (Observing 6 Wedges).

Fig. 4.

MBCA, Pairwise Correlations (Observing 6 Wedges).

Table 18.

MBCA Model, Information Matrix Decomposition (Observing 6 Wedges).

CRLB/para. sens/para. coll. ϱi ϱi(1) ϱi(2)
zss 5.486 0.099 55.563 0.999838 0.861 (τxss) 0.998 (τxss,τbss)
τlss 0.308 0.016 18.980 0.998611 0.977 (τbss) 0.995 (gss, τbss)
τxss 0.233 0.004 64.877 0.999881 0.904 (gss) 0.998 (zss, τbss)
gss 0.045 0.007 6.672 0.988704 0.904 (τxss) 0.914 (zss, τbss)
τbss 0.080 0.004 20.423 0.998801 0.977 (τlss) 0.996 (τlss, gss)
R˜ss 0.002 0.000 6.817 0.989181 0.838 (τlss) 0.908 (zss, τxss)
ρz 0.095 0.001 113.826 0.999961 0.994 (ρz,τx) 0.999 (ρz,τx,ρz,g)
ρτl,z 1.691 0.017 102.092 0.999952 0.994 (ρτl,τx) 0.999 (ρτl,τx,ρτl,g)
ρτx,z 0.875 0.009 102.811 0.999953 0.994 (ρτx) 0.999 (ρτx,ρτx,g)
ρg,z 2.179 0.048 45.056 0.999754 0.991 (ρg,τx) 0.999 (ρg,τx, ρg)
ρτb,z 3.307 0.027 122.830 0.999967 0.994 (ρτb,τx) 0.999 (ρτb,τx,ρτb,g)
ρR˜,z 2.189 0.023 95.026 0.999945 0.994 (ρR˜,τx) 0.999 (ρR˜,τx,ρR˜,g)
ρz,τl 0.575 0.016 35.237 0.999597 0.952 (ρτx,τl) 0.977 (ρτl,ρτx,τl)
ρτl 0.072 0.002 31.555 0.999498 0.937 (ρτb,τl) 0.977 (ρτl,τx,ρτl,τb)
ρτx,τl 1.511 0.049 30.748 0.999471 0.952 (ρz,τl) 0.976 (ρτx,ρτx,τb)
ρg,τl 2.055 0.127 16.157 0.998083 0.858 (ρg) 0.982 (ρg,τx,ρg,τb)
ρτb,τl 0.726 0.019 37.927 0.999652 0.963 (ρR˜,τl) 0.977 (ρτb,τx,ρτb)
ρR˜,τl 2.191 0.074 29.561 0.999428 0.963 (ρτb,τl) 0.976 (ρR˜,τx,ρR˜,τb)
ρz,τx 0.399 0.003 118.786 0.999965 0.994 (ρz) 0.999 (ρz, ρz,g)
ρτl,τx 68.331 0.642 106.404 0.999956 0.994 (ρτl,z) 0.999 (ρτl,z, ρτl,g)
ρτx 0.058 0.001 106.998 0.999956 0.994 (ρτx,z) 0.999 (ρτx,z, ρτx,g)
ρg,τx 1.542 0.033 46.985 0.999773 0.991 (ρg,z) 0.999 (ρg,z, ρg)
ρτb,τx 3.188 0.025 128.102 0.999970 0.994 (ρτb,z) 0.999 (ρτb,z, ρτb,g)
ρR˜,τx 79.866 0.807 98.976 0.999949 0.994 (ρR˜,z) 0.999 (ρR˜,z,ρR˜,g)
ρz,g 1.514 0.089 16.932 0.998255 0.953 (ρτx,z) 0.977 (ρτl,g, ρτx,g)
ρτl,g 1.310 0.087 15.082 0.997800 0.937 (ρτb,g) 0.964 (ρz,g, ρτb,g)
ρτx,g 3.269 0.221 14.792 0.997712 0.953 (ρz,g) 0.967 (ρz,g, ρτb,g)
ρg 0.020 0.002 8.356 0.992814 0.858 (ρg,τl) 0.947 (ρg,z, ρg,τb)
ρτb,g 4.759 0.262 18.154 0.998482 0.963 (ρR˜,g) 0.974 (ρτl,g, ρR˜,g)
ρR˜,g 13.415 0.945 14.195 0.997515 0.963 (ρτb,g) 0.967 (ρR˜,τl, ρτb,g)
ρz,τb 0.628 0.016 38.536 0.999663 0.952 (ρτx,τb) 0.979 (ρτl,τb,ρτx,τb)
ρτl,τb 0.792 0.023 33.771 0.999561 0.935 (ρτb) 0.977 (ρτl,z,ρτl)
ρτx,τb 2.103 0.061 34.475 0.999579 0.952 (ρz,τb) 0.979 (ρτx,z,ρτx,τl)
ρg,τb 2.409 0.144 16.731 0.998212 0.845 (ρτx,τb) 0.981 (ρg,z,ρg,τl)
ρτb 0.126 0.003 40.518 0.999695 0.968 (ρR˜,τb) 0.977 (ρτb,z,ρτb,τl)
ρR˜,τb 2.442 0.077 31.533 0.999497 0.968 (ρτb) 0.977 (ρR˜,z,ρR˜,τl)
ρz,R˜ 0.380 0.037 10.192 0.995175 0.953 (ρτx,R˜) 0.979 (ρτl,R˜,ρτx,R˜)
ρτl,R˜ 0.379 0.042 9.006 0.993817 0.937 (ρτb,R˜) 0.966 (ρz,R˜,ρτb,R˜)
ρτx,R˜ 0.450 0.050 9.012 0.993824 0.953 (ρz,R˜) 0.968 (ρz,R˜,ρτl,R˜)
ρg,R˜ 1.255 0.312 4.017 0.968518 0.836 (ρτx,R˜) 0.865 (ρg,z,ρτx,R˜)
ρτb,R˜ 6.461 0.595 10.849 0.995743 0.968 (ρR˜) 0.976 (ρτl,R˜,ρR˜)
ρR˜ 0.259 0.031 8.453 0.992978 0.968 (ρτb,R˜) 0.969 (ρR˜,z,ρτb,R˜)
q11 0.058 0.016 3.585 0.960317 0.937 (q31) 0.960 (q21, q31)
q21 0.212 0.050 4.231 0.971666 0.939 (q51) 0.968 (q11, q51)
q31 0.072 0.021 3.389 0.955473 0.937 (q11) 0.950 (q11, q41)
q41 1.404 0.719 1.952 0.858840 0.840 (q31) 0.842 (q21, q31)
q51 4.598 0.870 5.286 0.981945 0.970 (q61) 0.979 (q21, q61)
q61 1.128 0.273 4.131 0.970265 0.970 (q51) 0.970 (q31, q51)
q22 0.058 0.020 2.964 0.941350 0.912 (q52) 0.939 (q42, q52)
q32 0.302 0.154 1.967 0.861055 0.840 (q42) 0.861 (q22, q42)
q42 0.246 0.129 1.899 0.850068 0.840 (q32) 0.842 (q22, q32)
q52 0.118 0.025 4.746 0.977547 0.970 (q62) 0.976 (q22, q62)
q62 0.000 0.000 4.127 0.970205 0.970 (q52) 0.970 (q32, q52)
q33 0.058 0.038 1.538 0.759661 0.746 (q43) 0.759 (q43, q53)
q43 0.084 0.055 1.520 0.752922 0.746 (q33) 0.752 (q33, q63)
q53 1.383 0.334 4.133 0.970290 0.970 (q63) 0.970 (q33, q63)
q63 3.379 0.819 4.127 0.970198 0.970 (q53) 0.970 (q43, q53)
q44 0.058 0.055 1.052 0.310266 0.307 (q54) 0.309 (q54, q64)
q54 1.282 0.311 4.126 0.970180 0.970 (q64) 0.970 (q44, q64)
q64 1.689 0.410 4.124 0.970155 0.970 (q54) 0.970 (q44, q54)
q55 0.058 0.019 2.993 0.942520 0.943 (q65) 0.943 (ρg, q65)
q65 0.061 0.020 2.993 0.942515 0.943 (q55) 0.943 (ρg, q55)
q66 0.058 0.058 1.000 0.007684 0.001 (ρτl) 0.002 (ρτl,ρτl,g)

5 Economic Relevance

To assess which wedge is most important in accounting for cyclical fluctuations in the data Brinca et al. (2016) use the statistic fiY given by:

fiY=1/Σt(YtYit)2Σt,j(1/(YtYjt)2),

where j={z,τl,τx,g} in the BCA and j={z,τl,τx,g,τb,R˜} in the MBCA model. Yt is actual data of a given observable Y={y,l,x} (output, labour or investment)5 and Yit is the component of observable Y due to wedge i and with fiY[0,1], Σ fiY=1. For instance, fix roughly measures the fraction of movement and level in actual investment explained by wedge i. The statistic fiY ranges from 0 to 1 and is increasing in the explanatory power of a wedge. Indeed, when Yt = Yit, then fiY=1 since fjY=0 ji, whereas, in the limit, fiY=0 when the deviation of actual data from its component due to a given wedge goes to infinity. Notice that since the MSE statistic is used a model can be penalized along both the variance and the bias dimension. In other words both missing the data by a constant and missing the variation in the data is penalized.

When simulating the observables, we follow Chari et al. (2007) and consider two classes of counterfactual economies, namely the ‘one-wedge-on’ and ‘one-wedge-off’ economies. These economies are constructed by feeding the estimated wedges back into the model either one at a time (‘one-wedge-on’) or all but one (‘one-wedge-off’). It is important to point out the fact that the wedges have both a direct and a forecasting effect on the model. In the experiments, we seek to retain the forecasting effect only. This is done by setting the inactive wedges equal to some constant value (typically their intercept) at time t while preserving the estimated stochastic process and the realization of the wedges at time t − 1 to forecast their future realizations. Notice that this procedure would not be necessary if the matrices P and Q in the VAR(1) law of motion of the wedge shocks were diagonal.

The intuition behind the ‘one-wedge-on’ and ‘one-wedge-off’ counterfactual economies is the following. On one hand, the first seeks to understand how far a single wedge channel can bring the model to replicate the movements in the data while turning off the other margins. On the other hand, the second asks how badly the model performs when freezing the same channel and while keeping the others active.

5.1 Chari et al. (2007) BCA Model

Table 19 shows the fiY statistic for different observables and counterfactual economies as well as its one standard error bands given by fi + /− Sd(fi), where the standard deviations are computed using the delta method and the parameter covariance matrix (inverse Fisher information matrix) for the case when only the wedges parameters are assumed unknown (and thus the deep parameters of the model are fixed). To calculate the statistics we focus on the 1982 recession episode as in Chari et al. (2007) and use their original data.

Table 19.

BCA Model, Uncertainty Around fi Statistics.

Observable Counterfactual Economy fi-Sd(fi) fi fi-Sd(fi)
Output Efficiency wedge on 0.63 0.78 0.93
Labour wedge on 0.05 0.13 0.20
Investment wedge on 0.00 0.04 0.09
Government wedge on 0.01 0.05 0.10
Hours Efficiency wedge on −0.07 0.06 0.20
Labour wedge on 0.69 0.89 1.09
Investment wedge on −0.02 0.02 0.06
Government wedge on −0.01 0.02 0.06
Investment Efficiency wedge on 0.57 0.68 0.79
Labour wedge on 0.07 0.19 0.31
Investment wedge on −0.02 0.06 0.13
Government wedge on 0.04 0.07 0.10
Output Efficiency wedge off −0.01 0.04 0.09
Labour wedge off −0.07 0.20 0.47
Investment wedge off 0.11 0.46 0.80
Government wedge off 0.00 0.31 0.62
Hours Efficiency wedge off 0.43 0.90 1.36
Labour wedge off −0.06 0.02 0.10
Investment wedge off −0.18 0.05 0.28
Government wedge off −0.13 0.03 0.20
Investment Efficiency wedge off −0.02 0.08 0.19
Labour wedge off −0.06 0.28 0.62
Investment wedge off 0.05 0.13 0.21
Government wedge off 0.07 0.50 0.93

The one standard deviation bands around fi statistics for the one-wedge-on counterfactual economies cover an economically non-negligible range. However, these bands never overlap and are always quite far from each other. This means that if one used this statistic to measure and rank the relative importance of the wedges in replicating movements in the data, the main conclusions would not be overturned.6 Thus, the main result of Chari et al. (2007) still holds trough: the labour and the efficiency wedge play primary roles in explaining business cycle fluctuations during the period covering 1982 recession, whereas the investment and government wedge are negligible.

This is no longer true if one considers the one-wedge-off economies since the statistics exhibit such a higher level of uncertainty that the bands around them overlap. Our analysis thus discourages from using this statistic only to evaluate the relative importance of the wedges in a BCA model.

The intuition behind this result is that while the one-wedge-on fi statistics are a function of only one object subject to uncertainty, the one-wedge-off fi statistics are a function of three such objects. Indeed, in the one-wedge-on counterfactual economies only one wedge is active whereas in the one-wedge-off experiments this is the case for three wedges.

5.2 Brinca et al. (2016) Multi-country BCA Analysis

We now explore whether the results above for the 1982 recession in the United States, focus of the Chari et al. (2007), change if one considers multiple countries and focuses on the 2008 recession, as in Brinca et al. (2016). Tables 20 and 21, respectively, show the fi statistics for the one-wedge-on and one-wegde-off economies. Specifically, each table reports the fi statistics for each country, observable and wedge. Numbers are marked in bold only if, for a given observable and country, the most important wedge’s fi statistic is significantly different from the others (i.e. the bands do not overlap).7

Table 20.

ϕ Statistics for Observable‘s Components – Great Recession – One-Wedge-On.

Countries Output Labour Investment
Australia 0.69 0.27 0.02 0.02 0.64 0.12 0.12 0.10 0.53 0.25 0.03 0.17
Austria 0.71 0.07 0.11 0.12 0.28 0.07 0.19 0.43 0.62 0.05 0.20 0.10
Belgium 0.87 0.05 0.05 0.03 0.19 0.56 0.15 0.06 0.57 0.18 0.15 0.08
Canada 0.49 0.13 0.16 0.20 0.17 0.16 0.27 0.35 0.40 0.09 0.39 0.05
Denmark 0.58 0.06 0.29 0.06 0.26 0.13 0.49 0.11 0.19 0.05 0.69 0.07
Finland 0.93 0.01 0.03 0.02 0.38 0.02 0.09 0.49 0.60 0.03 0.29 0.06
France 0.93 0.01 0.04 0.02 0.71 0.02 0.19 0.08 0.69 0.03 0.21 0.05
Germany 0.79 0.03 0.12 0.07 0.26 0.16 0.33 0.23 0.43 0.04 0.47 0.06
Iceland 0.25 0.14 0.51 0.09 0.36 0.25 0.26 0.12 0.02 0.02 0.94 0.03
Ireland 0.21 0.24 0.46 0.08 0.06 0.30 0.58 0.04 0.07 0.07 0.80 0.06
Israel 0.77 0.03 0.16 0.04 0.38 0.26 0.08 0.27 0.21 0.08 0.60 0.11
Italy 0.68 0.08 0.17 0.07 0.18 0.13 0.59 0.09 0.33 0.07 0.56 0.04
Japan 0.62 0.10 0.15 0.12 0.18 0.14 0.44 0.22 0.36 0.15 0.32 0.16
Korea 0.51 0.18 0.18 0.14 0.37 0.24 0.16 0.23 0.44 0.09 0.33 0.14
Luxembourg 0.96 0.01 0.01 0.02 0.61 0.16 0.14 0.07 0.38 0.12 0.07 0.42
Mexico 0.55 0.11 0.26 0.07 0.22 0.23 0.43 0.10 0.24 0.14 0.48 0.12
Netherlands 0.89 0.02 0.05 0.03 0.38 0.10 0.24 0.26 0.70 0.03 0.22 0.04
New Zealand 0.44 0.09 0.40 0.08 0.22 0.15 0.53 0.10 0.08 0.03 0.83 0.05
Norway 0.76 0.04 0.05 0.15 0.27 0.08 0.22 0.37 0.77 0.03 0.06 0.12
Spain 0.11 0.05 0.81 0.03 0.15 0.14 0.63 0.07 0.02 0.01 0.96 0.01
Sweden 0.98 0.00 0.01 0.01 0.63 0.02 0.19 0.15 0.74 0.02 0.22 0.02
Switzerland 0.89 0.02 0.06 0.02 0.86 0.03 0.03 0.07 0.05 0.02 0.91 0.02
Great Britain 0.65 0.12 0.14 0.09 0.15 0.26 0.39 0.10 0.33 0.14 0.38 0.12
USA 0.15 0.50 0.26 0.06 0.03 0.80 0.15 0.01 0.08 0.08 0.78 0.04
Table 21.

ϕ Statistics for Observable‘s Components – Great Recession – One-Wedge-Off.

Countries Output Labour Investment
Australia 0.06 0.10 0.44 0.39 0.77 0.02 0.11 0.09 0.11 0.21 0.15 0.50
Austria 0.03 0.35 0.17 0.44 0.52 0.16 0.08 0.20 0.02 0.14 0.02 0.82
Belgium 0.01 0.16 0.36 0.40 0.64 0.05 0.12 0.14 0.07 0.16 0.25 0.43
Canada 0.04 0.58 0.21 0.13 0.59 0.22 0.08 0.06 0.03 0.71 0.03 0.18
Denmark 0.02 0.25 0.03 0.69 0.69 0.06 0.01 0.20 0.02 0.17 0.01 0.80
Finland 0.02 0.15 0.16 0.64 0.79 0.02 0.02 0.12 0.01 0.09 0.02 0.86
France 0.01 0.07 0.04 0.87 0.16 0.06 0.04 0.73 0.00 0.04 0.01 0.95
Germany 0.01 0.04 0.10 0.84 0.26 0.03 0.07 0.60 0.03 0.15 0.07 0.74
Iceland 0.54 0.31 0.02 0.14 0.83 0.10 0.01 0.05 0.49 0.24 0.00 0.23
Ireland 0.19 0.12 0.06 0.61 0.95 0.01 0.00 0.04 0.19 0.09 0.00 0.68
Israel 0.05 0.15 0.09 0.71 0.94 0.01 0.01 0.05 0.06 0.16 0.01 0.75
Italy 0.01 0.20 0.03 0.76 0.41 0.09 0.01 0.43 0.04 0.59 0.02 0.33
Japan 0.01 0.47 0.07 0.40 0.49 0.21 0.03 0.18 0.02 0.35 0.02 0.53
Korea 0.01 0.01 0.02 0.96 0.16 0.01 0.02 0.81 0.04 0.03 0.01 0.92
Luxembourg 0.02 0.35 0.23 0.38 0.90 0.04 0.02 0.04 0.02 0.29 0.03 0.65
Mexico 0.03 0.12 0.10 0.72 0.57 0.05 0.04 0.31 0.04 0.10 0.01 0.83
Netherlands 0.00 0.06 0.02 0.91 0.25 0.04 0.01 0.63 0.02 0.27 0.04 0.65
New Zealand 0.04 0.37 0.04 0.56 0.65 0.13 0.01 0.20 0.04 0.25 0.01 0.70
Norway 0.03 0.39 0.42 0.11 0.78 0.08 0.09 0.02 0.05 0.41 0.16 0.32
Spain 0.08 0.80 0.01 0.11 0.62 0.32 0.00 0.05 0.15 0.29 0.00 0.52
Sweden 0.01 0.12 0.21 0.65 0.48 0.06 0.10 0.33 0.02 0.19 0.09 0.69
Switzerland 0.07 0.15 0.06 0.73 0.71 0.05 0.02 0.22 0.19 0.49 0.03 0.28
Great Britain 0.00 0.04 0.03 0.90 0.66 0.01 0.01 0.29 0.02 0.14 0.01 0.77
United States 0.43 0.07 0.15 0.27 0.89 0.01 0.03 0.05 0.22 0.18 0.01 0.35

Two results stand out from Table 20. First, across the vast majority of countries, the fi statistics for output point to the efficiency wedge being the most relevant wedge. This result can be established with precision, as the uncertainty bands around the fi statistic for φAY do not overlap with those of other wedges. In other countries, the investment wedge takes a predominant role in explaining fluctuations in output (see Iceland, Ireland and Spain) and can also be distinguished from others. Second, the fi statistics for the United States are no longer statistically different from each other, as it was the case for the 1982 recession. These two results are largely confirmed by the one-wedge-off results in Table 21, for which we again find less discriminatory power due to lower precision of the fi statistics in the case where multiple wedges are active.

Notice that Chari et al. (2007) use a longer sample than Brinca et al. (2016) (specifically, 1959–2004 vs. 1980–2014) as the latter uses a balanced panel for 24 OECD countries for which the data availability is more limited. The larger uncertainty in the fi statistics thus also points to the crucial role of sample size in BCA applications, which we discuss in the next section.

6 Statistics for Practitioners

In this section, we present some statistics that might be of special interest to (monetary) BCA practitioners. We show how the size of the sample available affects the overall strength of parameter identification.

6.1 Empirical Distance Measures

While the analysis in the previous sections has been carried out in the time domain using the state space representation of the DSGE model solution the methodology proposed by Qu and Tkachenko (2012) and Qu and Tkachenko (2017) takes a frequency domain perspective to study identification issues. In particular, rank conditions for local identification of dynamic parameters are established by studying the spectral density matrix which maps the structural parameters to functions in the Banach space. Local identifiability occurs if and only if a local change in the parameter values leads to a different image. This analysis thus requires to study the Jacobian of the spectral density matrix w.r.t. the deep parameters of the model. If steady-state-related parameters are estimated as well, like in our case, it is possible to incorporate in the analysis also the first-order properties of the model by considering an extra term involving the steady-state parameters. In both cases, under some regularity conditions, the rank conditions are necessary and sufficient for local identification also in the time domain. More specifically, it can be shown that when the model is non-singular, like in our case, analyzing the rank conditions in the frequency domain is equivalent to inspecting the rank of the information matrix in the time domain. This analysis is thus expected to deliver analogous results to Iskrev (2010) and Komunjer and Ng (2011). For a more detailed comparison of these tests we refer the reader to Qu and Tkachenko (2012).

In general, the solution of a DSGE model log-linearized around its steady state can be expressed in state space form. Assuming no measurement errors, as it is the case in the BCA and MBCA model, it is given by:

(6.1)Xt+1=A(θ)Xt+B(θ)εt+1,Yt=C^(θ)Xt

and can be rewritten as a vector moving average (VMA) representation using:

(6.2)Yt(θ)=C^(θ)A(θ)Xt1+B(θ)εt=C^(θ)A(θ)A(θ)Xt2+B(θ)εt1+B(θ)εt=C^(θ)A(θ)2Xt2+C^(θ)A(θ)B(θ)εt1+B(θ)εt=

where, following the exposition in Qu and Tkachenko (2012) we made explicit the dependence of Yt on θ. More generally the process can be expressed as:

(6.3)Yt(θ)=C^(θ)A(θ)kXtk+C^(θ)A(θ)k1B(θ)εtk+1++C^(θ)A(θ)B(θ)εt1+B(θ)εt.

If all eigenvalues of A(θ) then limkA(θ)kXtk=0. It thus follows that

(6.4)Yt(θ)=Σj=0C^(θ)A(θ)jB(θ)εtj
(6.5)=Σj=0hj(θ)εtj
(6.6)=H(L;θ)εt,

where hj(θ)(j=0,,) are nY×nε matrices and H(L;θ)=Σj=0hj(θ)Lj the matrix of lagged polynomials.

Equation 6.5 along with some standard assumption about the process of the error term ϵ implies that Yt(θ) is covariance stationary and has a spectral density matrix fθ(ω) that can be represented as:

(6.7)fθ(ω)=12πH(exp(iω);θ)Σ(θ)H(exp(iω);θ),

where X denotes the conjugate transpose of a generic complex matrix X.

In some cases, like in the BCA and MBCA methodological framework, not only dynamic but also steady-state parameters are estimated. In this case, it is useful to define the augmented parameter vector as θ¯=(θ,α) and make explicit the relationship between data observables Yto, model observables Yt(θ) and steady states μ(θ¯)

(6.8)Yto=μ(θ¯)+Yt(θ),

where it is important to recall the fact that the model observables are expressed as log-deviations to form their respective steady-state values. This representation is useful in the case where steady-state parameters are estimated since considering only the second-order properties would omit the full identification potential of these parameters. This is why, in this case, the identification of θ¯ will be examined on the properties of μ(θ¯) and fθ¯(ω) jointly.

Definition 6.1. The parameter vector θ¯ is said to be locally identifiable from the first and second order properties of {Yt} at θ¯=θ¯0 if there exists an open neighborhood of θ¯0 in which μ(θ¯1)=μ(θ¯0) and fθ¯1(ω)=fθ¯0(ω) for all ω[π,π] necessarily implies θ¯1=θ¯0.

Now consider the following object:

(6.9)G¯(θ¯)=ππvec(fθ(ω))θ¯vec(fθ(ω))θ¯dω+μ(θ¯)θ¯μ(θ¯)θ¯.

Under some mild assumptions Qu and Tkachenko (2012) show that:

Theorem 6.1. θ¯ is locally identifiable from the first and second order properties of Yto at a point θ¯0 if and only if G¯(θ¯0) is nonsingular. Establishing whether the parameter set θ is identifiable thus reduces to checking that the matrix G¯(θ¯0) is full rank.

Analogously to Komunjer and Ng (2011) the eigenvalues of G(θ¯0) are obtained numerically, with the sole exception of the default Matlab tolerance level being used to determine its rank. This is due to the fact that the matrix is not as sparse as the one considered by Komunjer and Ng (2011). Also, the derivatives fθ(ωj)θk¯ are computed numerically as [fθ0+ekhk(ωj)fθ0(ωj)]/ where ek is a unit vector whose kth element is equal to 1 and hk is the step size. Unlike Komunjer and Ng (2011), however, the step size is not set to 1e−3 but to 1e−7.

To estimate fθ(ωj)θk¯ and to compute the integral in 6.9 we make use of the symmetric difference quotient and Gaussian quadrature, respectively. We find that both models are locally identifiable only when the first- and second-order properties of the data are used since we find the matrix G¯(θ¯0) in (6.9) to be full rank but the matrix G¯(θ¯)=ππvec(fθ(ω))θ¯vec(fθ(ω))θ¯dω) to be rank deficient.

If θ¯ is shown to be locally unidentifiable, as it is the case here for both the BCA and MBCA models, it is possible to trace out a curve of points χ which are observationally equivalent to θ¯0 in a local neighborhood of the latter. Indeed Qu and Tkachenko (2012), following Rothenberg (1971), show that this curve can be defined using the function θ(v) such that:

(6.10)θ¯(v)v=c(θ¯),θ¯(0)=θ¯0,

where c(θ¯0) is the eigenvector which corresponds to the smallest eigenvalue of G¯(θ¯), v is a scalar which varies in a neighborhood of 0 such that θ¯(v)δ(θ¯0).8

As shown by Qu and Tkachenko (2012), along the non-identification curve χ the parameter set θ¯ is not identified at θ¯0 since θ¯(v)v=0 and thus c(θ¯)=0, which implies:

(6.11)vec(fθ¯(v)(ω))v=vec(fθ¯(v)(ω))θ¯(v)c(θ¯)=0

for all ω[π,π].

The curve can be traced out recursively using the Euler method such that θ¯(vj+1)=θ¯(vj)+c(θ¯(vj))h, where h is the step size (fixed at 1e−04).

We apply two methods (the first due to Qu & Tkachenko, 2012 and the latter due to Qu & Tkachenko, 2017) can be used to robustify conclusions on both local and global identifications success or failure.

The first method recognizes the fact that it is important to make sure that the points on the curve result in identical spectral densities. It thus computes the maximum absolute and relative deviations of fθ¯1(ω) from fθ¯0(ω) across the frequencies: maxω[0,π]||fθ¯(ω)fθ¯0(ω)||9 and maxω[0,π]||fθ¯(ω)fθ¯0(ω)||\fθ¯0,hl(ω), where fθ¯0,hl(ω) denotes the (h, l) element of the spectral density matrix given the parameter set θ and is evaluated at the same frequency and element of fθ¯0(ω) that maximizes the numerator. In the BCA model we find that at norm (θ¯0θ¯) = 0.1276 the maximum absolute deviation on the curve is 0.0015 whereas the maximum relative deviation in relative form is 1.3811e−04. When norm (θ¯0θ¯) = is 1.30 the two measures are 0.0083 and 7.5130e−04, respectively. As to the MBCA model, when norm (θ¯0θ¯) = 0.0251 the maximum absolute deviation on the curve is 1.0226e−04 and the maximum relative deviation in relative form is 3.2755e−05. When norm (θ¯0θ¯) = 2.1728 the two measures are 2.3858e−04 and 7.4645e−05, respectively. In both models, the points on the curve θ¯0 for the smallest norm are chosen such that at least the absolute measure is bigger than the numerical errors associated with the Euler method, which Qu and Tkachenko (2017) say should remain near or below 1.0E−04.

The second method computes a so-called ‘empirical distance between DSGE models’ using a range of sample sizes. It is equal to the testing power of the likelihood ratio test of the null hypothesis that fθ¯(ω) is the true spectral density against the alternative hypothesis that hφ0(ω) is the true spectral density. In the context of our analysis, hφ0(ω) is taken to be fθ¯0(ω) but the test is general enough to accommodate different parameter vectors as well as different model structures. The empirical distance measure has the following properties. First, its value is between 0 and 1 for any sample size T and significance level α. A higher value thus means that it is easier to distinguish between the spectral densities implied by the two parameter vectors θ¯ and θ¯0 and thus argues in favour of stronger identification. Second, consider what Qu and Tkachenko (2017) refer to as the ‘Kullback–Leibler (KL) distance between two DSGE models (and, more generally, between two vector linear processes) with spectral densities fθ¯0(ω) and fθ¯0(ω) ’. It is given by

(6.12)KL(θ¯0,θ¯1)=KL(θ0,θ1)+14πμ(θ¯0)μ(θ¯1)fθ¯11(0)μ(θ¯0)μ(θ¯1)

Where

(6.13)KL(θ0,θ1)=14πππtr(fθ¯11(ω)fθ¯0(ω))logdet(fθ¯11(ω)fθ¯0(ω))nYdω

To provide intuition, let l(θ¯) denote the frequency domain approximate likelihood (or the time domain Gaussian likelihood) based on Yt(θ¯). Then, T1EYt(θ¯),θ¯=θ¯0L(θ¯0)L(θ¯1)KL(θ¯0,θ¯1). If the Kullback–Leibler distance KL(θ¯,θ¯0) is non-zero its value increase consistently with T and approaches 1 as T. As becomes evident from Tables 22 and 23 all empirical distance measures (p-values of the likelihood ratio tests) are well above the 5% significance level for all sample sizes considered, even for a small sample size of just T = 20, that is, five years of observations, and for the smallest normed differences between the points on the identification curve and the points at which local identification is checked in the respective models. This is reassuring evidence for practitioners since even with a relatively small sample size the overall identification strength is sufficiently high.

Table 22.

BCA Model, Wedges Parameters, Empirical Distance Measures.

Empirical Distance Measures on the Curve When Norm (θ¯0θ¯) = 0.13
Sample T = 20 T = 40 T = 80 T = 120 T = 160 T = 200 T = 1,000
Size 0.12 0.17 0.24 0.32 0.38 0.44 0.96
Empirical Distance Measures on the Curve When Norm (θ¯0θ¯) = 1.30
Sample T = 20 T = 40 T = 80 T = 120 T = 160 T = 200 T = 1,000
Size 1.00 1.00 1.00 1.00 1.00 1.00 1.00
Table 23.

MBCA Model, Wedges Parameters (no τbss,R˜ss), Empirical Distance Measures.

Empirical Distance Measures on the Curve When Norm (θ¯0θ¯) = 0.03
Sample T = 20 T = 40 T = 80 T = 120 T = 160 T = 200 T = 1,000
Size 0.25 0.40 0.6204 0.77 0.87 0.93 1.00
Empirical Distance Measures on the Curve When Norm (θ¯0θ¯) = 2.17
Sample T = 20 T = 40 T = 80 T = 120 T = 160 T = 200 T = 1,000
Size 1.00 1.00 1.00 1.00 1.00 1.00 1.00

7 Conclusion

In the past years, BCA exercises have sparked great interest among theoretical and applied macroeconomists insofar as they can shed light on which classes of models are able to explain fluctuations in macroeconomic aggregates during a particular economic episode. These exercises involve ML estimation of the stochastic process governing the latent variables. Even though they have been extensively performed, the methodology has yet to be properly scrutinized in terms of identification deficiencies. Given that quantitative recommendations are made the statistical and economic relevance of potential identification issues of the methodology is of the utmost importance.

In this chapter, we take seriously the notes of caution raised by the literature which investigates identification issues in DSGE models. Indeed, we first perform strict and weak local identification tests as theorized by Komunjer and Ng (2011) and Iskrev (2010) on the parameters vectors estimated in Chari et al. (2007) and Šustek (2011). We find that in both the standard and MBCA frameworks the model parameters are strictly identifiable. This is no longer true once one extends the estimation to the deep parameters of the model. In these cases, we show how to obviate such failures by imposing restrictions on the space of estimated parameters. Our analysis also points out that both models are affected by weak identification problems and thus suffer from a low degree of estimation precision. In particular, we find that the elements which suffer from weak identification are the off-diagonal elements of the VAR(1) law of motion of the latent variables. This is due to the fact that while these parameters do affect the likelihood, the effect which they exert on the latter is strongly collinear. We find that this is an inherent property of the VAR(1) process evaluated at the estimated parameter vectors. Indeed, we show that when innovations to the wedges are assumed to be observed and the parameters are estimated using the VAR in isolation from the model it is the off-diagonal elements which have a strongly collinear effect on the likelihood. Second, we investigate the economic severity of these identification problems by computing the uncertainty around a statistic which is used to rank which classes of models best explain business cycle fluctuations. We find that the main conclusion is not overturned for the standard BCA model. We are currently investigating whether this result also holds for the MBCA framework. Finally, we investigate a question of interest to practitioners, namely how identification strength varies across sample sizes. We show that even in small samples the parameter sets in both frameworks are overall well identified by computing empirical distance measures as in Qu and Tkachenko (2017).

A. Appendix – BCA and MBCA Model with Investment Adjustment Costs

A.1. Komunjer and Ng (2011)

A.1.1 Chari et al. (2007) BCA Model

We introduce the version of the BCA model by Chari et al. (2007) which allows for investment adjustment costs calibrated at the ‘normal’ intensity used by Bernanke et al. (1999), see Appendix C.2. The results are similar to the baseline case with the default parameters being strictly identifiable and several steady-state wedge shocks, off-diagonal elements of the P and Q matrix as well as deep parameters not being identifiable once the set of estimated parameters is extended to the latter (Tables A-1 and A-2).

Table A-1.

Komunjer and Ng Test Results BCA Model With Normal Adjustment Costs.

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 29 25 15 51 40 62 0
e−03 29 25 16 53 45 68 0
e−04 29 25 16 54 45 69 0
e−05 29 25 16 54 45 69 0
e−06 29 25 16 54 45 69 0
e−07 29 25 16 54 45 69 0
e−08 30 25 16 54 46 70 0
e−09 30 25 16 54 46 70 0
e−10 30 25 16 55 46 70 0
e−11 30 25 16 55 46 71 1
Default = 1.378453e−12 30 25 16 55 46 71 1
Required 30 25 16 55 46 71 1

Summary: nθ=30,nX=5,nε=4.

Order condition: nθ=30,nδ=50.

Table A-2.

Komunjer and Ng Test Results BCA Model With Normal Adjustment Costs (Deep Parameters Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 34 25 15 53 45 64 0
e−03 36 25 16 57 52 71 0
e−04 36 25 16 57 52 71 0
e−05 36 25 16 58 52 71 0
e−06 36 25 16 58 52 71 0
e−07 36 25 16 58 52 71 0
e−08 37 25 16 60 53 74 0
e−09 37 25 16 61 53 75 0
e−10 38 25 16 61 54 76 0
e−11 38 25 16 63 54 77 0
Default = 1.378453e−12 38 25 16 63 54 78 0
Required 39 25 16 64 55 80 1

Summary: nθ=39,nX=5,nε=4.

Order condition: nθ=39,nδ=50.

Problematic parameters at Tol = 1e−3: zss, τlss, τxss, gss, ρτl,z, ρτx,z, ρg,z, ρz,τl, ρτx,τl, ρg,τl, ρz,τx, ρτl,τx, ρg,τx, ρτl,g, ρτx,g, q21, q22, q32, q42, q33, q43, β, ψ, σ, a, b.

Problematic parameters at Tol = 1.378453e−12: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρz,g, ρτl,g, ρτx,g, ρg, q21, q31, q22, q32, q42, q33, q43, q44, gn, gz, β, δ, ψ, σ, α, a, b.

In line with the analysis carried out for the baseline model, we check which restricted parameter combinations would allow the other parameters to be strictly identifiable. As reported in Table A-3 we find three such sets at the lowest tolerance level for which identification fails in A-2, mostly consisting of three parameters each. These sets are (i) {a,ρg}, (ii) {b,ρg} and (iii) {a,ψ} and thus involve the parameters governing the degree of investment adjustment costs (a and b), the autocorrelation of the government wedge and the inverse of the constant Frisch elasticity of labour supply ψ.

Table A-3.

Komunjer and Ng Conditional Test Results BCA Model With Normal Adjustment Costs, Tol = 1.378453e−12.

Fixed ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
a ρg 39 25 16 64 55 80 1
b ρg 39 25 16 64 55 80 1
a ψ 39 25 16 64 55 80 1
Required 39 25 16 64 55 80 1

Summary: nθ=39,nX=5,nε=4.

Order condition: nθ=39,nδ=50.

A.1.2 Šustek (2011) MBCA Model

An analogous pattern to the one found in the baseline MBCA model and reported in Section 4 emerges once investment adjustment costs are introduced in the model (see Tables A-4 and A-7).

Table A-4.

Komunjer and Ng Test Results MBCA Model With Normal Adjustment Costs.

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 60 49 33 101 89 130 0
e−03 60 49 36 107 96 142 0
e−04 60 49 36 109 96 144 0
e−05 60 49 36 109 96 144 0
e−06 60 49 36 109 96 144 0
e−07 60 49 36 109 96 145 0
e−08 60 49 36 109 96 145 0
e−09 61 49 36 109 97 145 0
e−10 61 49 36 110 97 145 0
e−11 61 49 36 110 97 146 1
Default = 5.826450e−12 61 49 36 110 97 146 1
Required 61 49 36 110 97 146 1

Summary: nθ=61,nX=7,nε=6.

Order condition: nθ=61,nδ=105.

Table A-5.

Komunjer and Ng Test Results MBCA Model with Normal Adjustment Costs (τbss and R˜ss Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 62 49 33 103 91 132 0
e−03 62 49 36 109 98 143 0
e−04 62 49 36 111 98 146 0
e−05 62 49 36 111 98 146 0
e−06 62 49 36 111 98 146 0
e−07 62 49 36 111 98 147 0
e−08 62 49 36 111 98 147 0
Default = 2.983143e−09 62 49 36 111 98 147 0
e−09 63 49 36 111 99 147 0
e−10 63 49 36 111 99 147 0
e−11 63 49 36 112 99 148 1
Required 63 49 36 112 99 148 1

Summary: nθ=63,nX=7,nε=6.

Order condition: nθ=63,nδ=105.

Problematic parameters at Tol = 1e−3: zss, gss.

Problematic parameters at Tol = 1.000000e−10: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl,ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q21, q31, q51, q22, q32, q42, q33, q43, q53, q44, q54, q64, q55.

Table A-6.

Komunjer and Ng Test Results MBCA Model With Normal Adjustment Costs (Deep Parameters Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 69 49 35 104 99 133 0
e−03 70 49 36 113 106 146 0
e−04 70 49 36 116 106 148 0
e−05 70 49 36 117 106 148 0
e−06 70 49 36 117 106 148 0
e−07 71 49 36 118 107 149 0
e−08 71 49 36 119 107 152 0
e−09 72 49 36 119 108 153 0
e−10 73 49 36 120 109 154 0
Default = 2.330580e−11 73 49 36 121 109 155 0
e−11 73 49 36 122 109 157 0
Required 74 49 36 123 110 159 1

Summary: nθ=74,nX=7,nε=6.

Order condition: nθ=74,nδ=105.

Problematic parameters at Tol = 1e−3: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q22, q42, q43, q53, gn, gz, β, ψ, σ, ρR, ωπ, ωy, πss, a, b.

Problematic parameters at Tol = 1.000000e−11: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q11, q21, q31, q51, q61, q22, q32, q42, q52, q62, q33, q43, q53, q63, q44, q54, q64, q55, q65, q66, gn, gz, β, δ, ψ, σ, α, ρR, ωπ, ωy, πss, a, b.

Table A-7.

Komunjer and Ng Test Results MBCA Model with Normal Adjustment Costs (τbss,R˜ss and Deep Parameters Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 71 49 35 105 101 134 0
e−03 72 49 36 114 108 146 0
e−04 72 49 36 118 108 149 0
e−05 72 49 36 119 108 150 0
e−06 72 49 36 119 108 150 0
e−07 73 49 36 120 109 151 0
Default = 9.546056e−08 73 49 36 120 109 151 0
e−08 73 49 36 121 109 154 0
e−09 74 49 36 121 110 154 0
e−10 74 49 36 122 110 156 0
e−11 75 49 36 123 111 158 0
Required 76 49 36 125 112 161 1

Summary: nθ=76,nX=7,nε=6.

Order condition: nθ=76,nδ=105.

Problematic parameters at Tol = 1e−3: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q21, q31, q51, q22, q32, q42, q52, q33, q43, q53, q63, q44, q54, q64, q55, gn, gz, β, δ, ψ, σ, ρR, ωπ, ωy, πss, a, b.

Problematic parameters at Tol = 1.000000e−11: zss, τlss, τxss, gss, τbss, R˜ss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q11, q21, q31, q41, q51, q61, q22, q32, q42, q52, q62, q33, q43, q53, q63, q44, q54, q64, q55, q65, q66, gn, gz, β, δ, ψ, σ, α, ρR, ωπ, ωy, πss, a, b.

As to the conditional identification tests, we find that only for the MBCA model with normal adjustment costs where τbss and R˜SS not included in estimation it is possible to reparameterize the model in a way which makes the other parameters identifiable. We find 27 parameter combinations which restrict only four parameters at the lowest tolerance level for which identification fails in A-6 (see Table A-8). These sets mainly feature the steady-state innovations to the efficiency and government wedge (zss and gss) as well as steady-state inflation πss and some deep parameters. The parameter b which governs, with a, the degree of investment adjustment costs also shows up prominently in these sets. This is because this parameter is a function of other deep parameters and fixing it thus provides an additional restriction which can help to identify the latter.

Table A-8.

Komunjer and Ng Conditional Test Results MBCA Model with Normal Adjustment Costs, Tol = 2.330580e−11.

Fixed ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
zss τlss πss b 74 49 36 123 110 159 1
zss ρz πss b 74 49 36 123 110 159 1
zss ρτl,τx πss b 74 49 36 123 110 159 1
zss ρτb πss b 74 49 36 123 110 159 1
zss ρτb,R˜ πss b 74 49 36 123 110 159 1
zss q31 πss b 74 49 36 123 110 159 1
zss q51 πss b 74 49 36 123 110 159 1
zss q54 πss b 74 49 36 123 110 159 1
zss ψ πss b 74 49 36 123 110 159 1
zss σ πss b 74 49 36 123 110 159 1
zss πss a b 74 49 36 123 110 159 1
τlss gss πss b 74 49 36 123 110 159 1
gss ρz πss b 74 49 36 123 110 159 1
gss ρτl,τx πss b 74 49 36 123 110 159 1
gss ρτb πss b 74 49 36 123 110 159 1
gss ρτb,R˜ πss b 74 49 36 123 110 159 1
gss q31 πss b 74 49 36 123 110 159 1
gss q51 πss b 74 49 36 123 110 159 1
gss q54 πss b 74 49 36 123 110 159 1
gss ψ πss b 74 49 36 123 110 159 1
gss σ πss b 74 49 36 123 110 159 1
gss πss a b 74 49 36 123 110 159 1
ρz α πss b 74 49 36 123 110 158 0
ρτb,R˜ α πss b 74 49 36 123 110 158 0
q54 α πss b 74 49 36 123 110 158 0
ψ α πss b 74 49 36 123 110 158 0
σ α πss b 74 49 36 123 110 158 0
Required 74 49 36 123 110 159 1

Summary: nθ=74,nX=7,nε=6.

Order condition: nθ=74,nδ=105.

A.2 Iskrev (2010)

A.2.1 Chari et al. (2007) BCA Model

We first study strict identification in the standard BCA model featuring investment adjustment costs of normal intensity. We start with the 38 parameters case (only a is considered in estimation since b is de facto a function of other structural parameters). Fixing gn and gz in the standard BCA case discussed above leaves 36 parameters and the rank of the information matrix is found to be 33. There are 7,140 combinations of 33 parameters out of 36 and as many as 2,406 combinations of parameters which, when fixed, deliver a rank of 33 and thus enable identification of the ones left unrestricted.

Table A-9 is instructive about weak identification problems. Interestingly, relative uncertainty decreases such that q31 and q44 no longer feature among the set of worst identified parameters. This is no longer true for q31 once also the deep parameters of the model are considered in the identification analysis (see Tables A-10 and A-11). The same results of the no adjustment costs counterpart hold through, with the exception of α also being one of the most badly identified parameters.

Table A-9.

BCA Model with Normal Adjustment Costs, Parameter Identification (All Deep Parameters Fixed).

Value CRLB rCRLB
zss −0.0239 0.0955 3.9926
τlss 0.3279 0.1450 0.4423
τxss 0.4834 0.2142 0.4431
gss −1.5344 0.3212 0.2093
ρz 0.9800 0.0488 0.0498
ρτl,z −0.0330 0.0514 1.5588
ρτx,z −0.0702 0.1022 1.4548
ρz,g 0.0048 0.0596 12.3854
ρz,τl −0.0138 0.0351 2.5492
ρτl 0.9564 0.0528 0.0552
ρτx,τl −0.0460 0.1147 2.4941
ρg,τl −0.0081 0.0470 5.7929
ρz,τx −0.0117 0.0836 7.1323
ρτl,τx −0.0451 0.0699 1.5497
ρτx 0.8962 0.0941 0.1050
ρg,τx 0.0488 0.0995 2.0369
ρz,g 0.0192 0.0802 4.1698
ρτl,g 0.0569 0.0663 1.1650
ρτx,g 0.1041 0.0949 0.9114
ρg 0.9711 0.0926 0.0953
q11 0.0116 0.0007 0.0578
q21 0.0014 0.0015 1.0792
q31 −0.0105 0.0078 0.7389
q41 −0.0006 0.0013 2.2903
q22 0.0064 0.0004 0.0636
q32 0.0010 0.0067 6.5323
q42 0.0061 0.0050 0.8153
q33 0.0158 0.0075 0.4712
q43 0.0142 0.0023 0.1619
q44 0.0046 0.0036 0.7903
Table A-10.

BCA Model with Normal Adjustment Costs, Wedges Parameter Identification.

Value CRLB rCRLB
zss −0.0239 2.3406 97.8465
τlss 0.3279 0.5711 1.7414
τxss 0.4834 1.9206 3.9729
gss −1.5344 0.7576 0.4937
ρz 0.9800 0.0554 0.0565
ρτl,z −0.0330 0.0545 1.6526
ρτx,z −0.0702 0.1190 1.6937
ρz,g 0.0048 0.0654 13.5906
ρz,τl −0.0138 0.0548 3.9733
ρτl 0.9564 0.0586 0.0612
ρτx,τl −0.0460 0.1305 2.8372
ρg,τl −0.0081 0.0660 8.1379
ρz,τx −0.0117 0.1096 9.3467
ρτl,τx −0.0451 0.0876 1.9431
ρτx 0.8962 0.1101 0.1229
ρg,τx 0.0488 0.1026 2.1003
ρz,g 0.0192 0.0934 4.8530
ρτl,g 0.0569 0.0795 1.3964
ρτx,g 0.1041 0.1715 1.6475
ρg 0.9711 0.1042 0.1073
q11 0.0116 0.0064 0.5536
q21 0.0014 0.0035 2.4901
q31 −0.0105 0.0163 1.5549
q41 −0.0006 0.0013 2.2904
q22 0.0064 0.0046 0.7156
q32 0.0010 0.0113 10.9783
q42 0.0061 0.0116 1.8940
q33 0.0158 0.0244 1.5433
q43 0.0142 0.0045 0.3185
q44 0.0046 0.0045 0.9858
δ 0.0118 0.0015 0.1285
σ 1.0000 0.4395 0.4395
α 0.3500 0.4199 1.1996
Table A-11.

BCA Model with Normal Adjustment Costs (Deep Parameters Estimated), Information Matrix Decomposition.

CRLB/para. sens/para. coll. ϱi ϱi(1) ϱi(2)
zss 97.846 0.410 238.923 0.999991 0.926 (gss) 0.980 (gss, α)
τlss 1.741 0.019 90.135 0.999938 0.801 (α) 0.861 (τxss,ρτx)
τxss 3.973 0.006 681.977 0.999999 0.946 (gss) 0.983 (ρg,τl, α)
gss 0.494 0.003 151.290 0.999978 0.946 (τxss) 0.989 (zss, α)
ρz 0.057 0.000 211.820 0.999989 0.992 (ρz,g) 0.994 (ρτl,z,ρz,g)
ρτl,z 1.653 0.038 43.825 0.999740 0.900 (ρτl) 0.987 (ρτl,τx,ρτl,g)
ρτx,z 1.694 0.019 89.376 0.999937 0.948 (ρz,g) 0.985 (ρτx,ρτx,g)
ρz,g 13.591 0.074 184.275 0.999985 0.992 (ρz) 0.993 (ρz, ρτx,z)
ρz,τl 3.973 0.012 336.677 0.999996 0.991 (ρg,τl) 0.995 (ρg,τl,ρz,g)
ρτl 0.061 0.001 75.356 0.999912 0.983 (ρτl,g) 0.992 (ρτl,τx,ρτl,g)
ρτx,τl 2.837 0.018 155.079 0.999979 0.980 (ρτx,g) 0.991 (ρτx,ρτx,g)
ρg,τl 8.138 0.027 298.583 0.999994 0.991 (ρz,τl) 0.994 (ρz,τl, ρg)
ρz,τx 9.347 0.010 979.814 0.999999 0.992 (ρg,τx) 0.999 (ρz, ρz,g)
ρτl,τx 1.943 0.012 166.317 0.999982 0.987 (ρτl,g) 0.999 (ρτl,z,ρτl,g)
ρτx 0.123 0.001 195.368 0.999987 0.986 (ρτx,g) 0.998 (ρτx,z,ρτx,g)
ρg,τx 2.100 0.003 673.543 0.999999 0.992 (ρz,τx) 0.999 (ρz,g, ρg)
ρz,g 4.853 0.004 1,280.821 1.000000 0.992 (ρg) 0.999 (ρz, ρz,τx)
ρτl,g 1.396 0.006 229.345 0.999990 0.987 (ρτl,τx) 0.999 (ρτl,z,ρτl,τx)
ρτx,g 1.648 0.004 459.104 0.999998 0.986 (ρτx) 0.999 (ρτx,z,ρτx)
ρg 0.107 0.000 1,051.673 1.000000 0.992 (ρz,g) 0.999 (ρz,g,ρg,τx)
q11 0.554 0.036 15.585 0.997939 0.780 (q31) 0.788 (q21, q31)
q21 2.490 0.247 10.064 0.995051 0.747 (q41) 0.755 (q11, q41)
q31 1.555 0.038 40.815 0.999700 0.951 (q41) 0.960 (q11, q41)
q41 2.290 0.654 3.504 0.958405 0.951 (q31) 0.958 (q21, q31)
q22 0.716 0.045 15.827 0.998002 0.622 (q42) 0.632 (τlss, q42)
q32 10.978 0.388 28.313 0.999376 0.951 (q42) 0.951 (τlss, q42)
q42 1.894 0.062 30.769 0.999472 0.951 (q32) 0.955 (q22, q32)
q33 1.543 0.024 63.950 0.999878 0.909 (q43) 0.910 (ρτx, q43)
q43 0.319 0.027 12.003 0.996524 0.909 (q33) 0.909 (zss, q33)
q44 0.986 0.058 16.985 0.998265 0.113 (σ) 0.389 (δ, σ)
δ 0.129 0.013 9.575 0.994531 0.934 (gss) 0.966 (zss, α)
σ 0.440 0.011 40.138 0.999690 0.940 (α) 0.980 (τxss,ρg,τl)
α 1.200 0.002 504.842 0.999998 0.940 (σ) 0.983 (τxss,ρg,τl)

B. Appendix – Model Derivations

B.1. Representative Consumer

B.1.1. Optimization Problem of the Household

Suppose households own the capital stock and rent it out at rate rt. They also work for wages at rate wt per unit of labour input and pay takes on labour, investment and bond holdings. Then, the optimization problem for the household looks as follows:

Objective function:

max{ct(st),lt(st),kt+1(st),bt(st)}Σt=0ΣstβtUct(st),1l(st)Nt(st)=(1+gn)tsettingN0=1,

where ct(st),xt(st)0,st,t.

Let xt(st) be any random variable. The expectation over the discounted sum of future possible realizations can be written as:

E0Σt=0βtxt(st)Σt=0Σstβtxt(st)μt(st)

The second sum expresses that the expectation is a probability weighted average of the different possible realizations of the variable. We can thus rewrite the objective function as follows:

max{ct(st),lt(st),kt+1(st),bt(st)}E0Σt=0βtUct(st),1l(st)(1+gn)t,

Budget constraint:

ct(st)+[1+τx(st)]xt(st)+[1+τb(st)](1+gn)bt(st)[1+Rt(st)]pt(st)bt1(st1)pt(st)=[1τl(st)]wt(st)lt(st)+rt(st)kt(st1)+Tt(st)

Capital accumulation law:

(B-1)Nt+1(st+1)kt+1(st)=[(1δ)kt(st)+xt(st)]Nt(st)(1+gn)kt+1(st)=(1δ)kt(st)+xt(st)

B.1.2. Lagrangian Function

Given that there is one budget constraint for each realization of st in each time period t one can write the Lagrangian function in the following way:

L=E0Σt=0βtUct(st),1l(st)Nt(st)Σt=0Σstλ˜t{ct(st)+[1+τx(st)]xt(st)+[1+τb(st)](1+gn)bt(st)[1+Rt(st)]pt(st)bt1(st1)pt(st)[1τl(st)]wt(st)lt(st)rt(st)kt(st1)Tt(st)}

Making use of expression (B-1) for investment and recalling that the expectation is a probability weighted sum of the possible realizations, one can integrate the budget constraints into the first part of the function one obtains:

L=E0{Σt=0βt[Uct(st),1l(st)Nt(st)λt[ct(st)+[1+τx(st)][(1+gn)kt+1(st)(1δ)kt(st)]+[1+τb(st)][(1+gn)bt(st)[1+Rt(st)]pt(st)bt1(st1)pt(st)][1τl(st)]wt(st)lt(st)rt(st)kt(st1)Tt(st)]]}

with λt=λ˜tβtμt(st).

B.1.3. First-order Necessary Conditions

Differentiating the Lagrangian with respect to the choice variables of the household, one obtains the following first-order necessary conditions which hold st,t:

For each state of the world, st, and each point of time t it holds that:

(B-2)Lct(st)=βtμt(st)[Uc,t(st)(1+gn)tλt]=0λt=Uc,t(st)(1+gn)t
(B-3)Lnt(st)=βtμt(st)[Ul,t(st)(1+gn)t+λt[1τl,t(st)]wt(st)]=0λt=Ul,t(st)(1+gn)t[1τl,t(st)]wt(st)

The intratemporal optimality condition is then given by:

(B-4)Ul,t(st)Uc,t(st)=[1τl,t(st)]wt(st).

When differentiating the Lagrangian with respect to capital kt+1(st), one has to note that at time t + 1 the net capital stock of the previous period, kt(st1), becomes kt+1(st), which leads to an additional component in the derivative.

(B-5)Lkt+1(st)=βtμt(st)[λt[1+τx,t(st)](1+gn)]+Σst+1>stβt+1μt+1(st+1)[λt+1[1+τx,t+1(st+1)](1δ)+rt+1(st+1)]=01=β(1+gn)Σst+1>stμt+1(st+1)μt(st)λt+1λt[1+τx,t+1(st+1)](1δ)+rt+1(st+1)1+τx,t(st)1=β(1+gn)Σst+1>stμt+1(st+1|st)Uc,t+1(st+1)(1+gn)t+1Uc,t(st)(1+gn)t×[[1+τx,t+1(st+1)](1δ)+rt+1(st+1)1+τx,t(st)]1=βΣst+1>stμt+1(st+1|st)Uc,t+1(st+1)Uc,t(st)[[1+τx,t+1(st+1)](1δ)+rt+1(st+1)1+τx,t(st)]1=βEtUc,t+1(st+1)Uc,t(st)[1+τx,t+1(st+1)](1δ)+rt+1(st+1)1+τx,t(st),

where in the third step we used μt+1(st+1)μt(st)μt+1(st+1|st) and equation (B-2).

When differentiating the Lagrangian with respect to bonds, one has to note that at time t + 1 the bonds of the previous period, bt1(st1), become bt(st), which leads to an additional component in the derivative.

(B-6)Lbt(st)=βtμt(st)λt[1+τb,t(st)](1+gn)[1+Rt(st)]pt(st)+Σst+1>stβt+1μt+1(st+1)λt+1[1+τb,t+1(st+1)]1pt+1(st+1)=01=β(1+gn)Σst+1>stμt+1(st+1)μt(st)λt+1λt1+τb,t+1(st+1)1+τb,t(st)pt(st)pt+1(st+1)[1+Rt(st)]1=β(1+gn)Σst+1>stμt+1(st+1)μt(st)Uc,t+1(st+1)(1+gn)t+1Uc,t(st)(1+gn)t1+τb,t+1(st+1)1+τb,t(st)pt(st)pt+1(st+1)[1+Rt(st)]1=βΣst+1>stμt+1(st+1|st)Uc,t+1(st+1)Uc,t(st)1+τb,t+1(st+1)1+τb,t(st)pt(st)pt+1(st+1)[1+Rt(st)]1=βEtUc,t+1(st+1)Uc,t(st)1+τb,t+1(st+1)1+τb,t(st)pt(st)pt+1(st+1)[1+Rt(st)]1=βEtUc,t+1(st+1)Uc,t(st)1+τb,t+1(st+1)1+τb,t(st)pt(st)pt+1(st+1)[1+Rt(st)]

where in the fourth step we used μt+1(st+1)μt(st)μt+1(st+1|st) and equation (B-2).

B.2. Representative Producer

The representative producer operates an aggregate CRS production function:

(B-7)yt(st)=Fkt(st1),Zt(st)lt(st),

where F(.,.) has the standard properties and Zt(st)=zt(st)1+gzt.

B.2.1. Optimization Problem of the Firm

The producer maximizes profits:

(B-8)yt(st)wt(st)lt(st)rt(st)kt(st1)

by setting:

(B-9)wt(st)=Fl,tkt(st1),zt(st)(1+gz)tlt(st)
(B-10)•••
(B-11)•••

B.3. Additional Model Equations

The aggregate resource constraint is given by:

(B-12)yt(st)=ct(st)+gt(st)+xt(st).

Monetary policy is assumed to set the interest rate according to a Taylor rule of the following type:

(B-13)Rt(st)=(1ρR)R+ωy(lnyt(st)lny)+ωπ(πt(st)π)+ρRRt1(st1)+R˜t(st),

where ρR[0,1),πt(st)lnpt(st)lnpt1(st1) is the inflation rate and a variable’s symbol without a time subscript denotes the variable’s steady-state (or balanced growth path) value. In addition, it is assumed that ωπ>1, thus eliminating explosive paths.

B.4. Functional Forms and Auxiliary Assumptions

From here on, we make the following functional form assumptions:

and F(k,Zl)=kα(Zl)1α and U(c,1l)=c1lψ1σ1σ

It is assumed that the state st stfollows a Markov process of the form μ(st|st1) and that the wedges in period t can be used to uncover the event st uniquely, in the sense that the mapping from the event st to the wedges zt,1τl,t,11+τl,t,gt,11+τb,t,Rt˜ is one-to-one and onto. Given this assumption, without loss of generality, let the underlying event st=(szt,slt,sxt,sgt,sbt,sR˜t), and let logzt(st)=szt,τl,t(st)=slt,τx,t(st)=sxt and loggt(st)=sgt. Given the unique mapping between st and the wedges we make following auxiliary choices:

logzt=logz(st), logg^t=logg^t(st),τl,t=τl(st),τx,t=τx(st),τb,t=τb(st),R˜t=R˜(st)

Note that we have effectively assumed that agents use only past wedges to forecast future wedges and that the wedges in period t are sufficient statistics for the event in period t. More precisely, the VAR representation of the underlying state st is modelled as follows

st+1=P0+Pst+Qεs,t+1,

where εs,t+1N(0,I).

B.5. Operational Model

In the operational model, we consider quantities which are not only expressed in per-capita terms but also detrended. To highlight the differences between this model’s and the previous model’s variables we introduce the notation v^vt(1+gz)tVtNt(1+gz)t. The model is then given by:

  • CRS production function:

    (B-14)y^t(st)=k^t(st1)αztlt(st)1α

  • Aggregate resource constraint:

    (B-15)y^t(st)=c^t(st)+g^t+x^t(st)

  • Capital accumulation law:

    (B-16)(1+gn)(1+gz)t+1k^t+1(zt)=(1δ)(1+gz)tk^t(zt1)+(1+gz)tx^t(zt)(1+gn)(1+gz)k^t+1(zt)=(1δ)k^t(zt1)+x^t(zt)

  • Taylor rule:

    (B-17)Rt(st)=(1ρR)R+ωy(lny^t(st)lny^)+ωπ(πt(st)π)+ρRRt1(st1)+R˜t,

  • F.O.C. labour:

    (B-18)ψc^t(st)1lt(st)=(1τl,t)(1α)k^t(st1)αzt1αlt(st)α

  • F.O.C. capital:

    (B-19)1=βEt{c^t+1(1+gz)t+1σ(1lt+1)ψ(1σ)c^t(1+gz)tσ(1lt)ψ(1σ)×[(1+τx,t+1)(1δ)+αk^t+1(st)α1zt+1lt+1(st+1)1α1+τx,t]}=β˜Et{c^t+1c^tσ1lt+11ltψ(1σ)[(1+τx,t+1)(1δ)+αk^t+1(st)α1zt+1lt+1(st+1)1α1+τx,t]},

    where β˜=β/(1+gz)σ.

  • F.O.C. bonds:

    (B-20)1=βEt{c^t(st)(1+gz)tc^t+1(st+1)(1+gz)t+11+τb,t+11+τb,tpt(st)pt+1(st+1)[1+Rt(st)]}1=β˜Et{c^t(st)c^t+1(st+1)1+τb,t+11+τb,tpt(st)pt+1(st+1)[1+Rt(st)]}

    where β˜=β/(1+gz).

B.6. Steady State

The model in steady state is given by:

  • CRS production function:

    (B-21)y^=k^αzl1α

  • Aggregate resource constraint:

    (B-22)y^=c^+g^+x^

  • Capital accumulation law:

    (B-23)(1+gn)(1+gz)k^=(1δ)k^+x^

  • Taylor rule:

    (B-24)π=R˜(1ρR)ωπ+πSS,

  • F.O.C. labour:

    (B-25)ψc^1l=(1τl)(1α)k^αz1αlα

  • F.O.C. capital:

    (B-26)1=β˜(1+τx)(1δ)+αk^α1zl1α1+τx

  • F.O.C. bonds:

    (B-27)1=β˜exp(π)(1+R)R=1β˜exp(π)1,

    where we used that πt=logptpt1.

To solve for the steady state of real variables, we start by solving (B-26) w.r.t. k^:

(B-28)k^=[αβ˜(1+τx)[1β˜(1δ)]]11αzlΛzl

Plugging this expression for k^ in equation (B-25) yields:

(B-29)ψc^1l=(1τl)(1α)(Λzl)αz1αlαψc^1l=(1τl)(1α)Λαzc^=1ψ(1τl)(1α)Λαz(1l)

Inserting (B-28), (B-29) and (B-23) into the aggregate resource constraint leads to:

(B-30)k^α(zl)1α=c^+g^+(1+gn)(1+gz)k^(1δ)k^Λzlα(zl)1α=1ψ(1τl)(1α)Λαz(1l)+g^+(1+gn)(1+gz)(1δ)ΛzlΛαzl=1ψ(1τl)(1α)Λαz1ψ(1τl)(1α)Λαzl+g^+(1+gn)(1+gz)(1δ)Λzl
(B-31)Λαzl+1ψ(1τl)(1α)Λαzl(1+gn)(1+gz)(1δ)Λzl=g^+1ψ(1τl)(1α)Λαzl=g^+1ψ(1τl)(1α)ΛαzΛαz1+1ψ(1τl)(1α)Λz(1+gn)(1+gz)(1δ)

Using (B-31) in (B-28) one obtains:

(B-32)k^=Λzl=Λzg^+1ψ(1τl)(1α)ΛαzΛαz1+1ψ(1τl)(1α)Λz(1+gn)(1+gz)(1δ)=g^+1ψ(1τl)(1α)ΛαzΛα11+1ψ(1τl)(1α)(1+gn)(1+gz)(1δ)

B.7. Definitions

Below we define the notation used for the model’s variables and parameters.

B.7.1. Variables

Lower-case variables define quantities in per-capita terms as follows:

  • (a)

    Z: Labour-augmenting technical change Z=z(1+gz)t.

  • (b)

    b: One period, nominal risk-free bonds; purchased in period t, pay off in period t + 1.

  • (c)

    c: Consumption.

  • (d)

    g: Government consumption.

  • (e)

    k: Net capital stock.

  • (f)

    l: Labour.

  • (g)

    N: Population Nt=N0(1+gn)t.

  • (h)

    r: Rental rate on capital.

  • (i)

    R: Nominal interest rate.

  • (j)

    t: Time period.

  • (k)

    w: Wage rate.

  • (l)

    x: Investment.

  • (m)

    y: Output.

  • (n)

    s: State of the world at time t.

  • (o)

    π: Inflation rate.

  • (p)

    τl: Labour tax.

  • (q)

    τx: Tax on investment.

  • (r)

    τb: Tax on bond holdings.

B.7.2. Parameters

The following parameters appear in the model:

  • (a)

    gn: Population growth rate of labour-augmenting technological process.

  • (b)

    gz: Growth rate of labour-augmenting technological process.

  • (c)

    α: Parameter which determines the share of (and weight on) net capital stock in the Cobb-Douglas CRS production function.

  • (d)

    β: Subjective discount factor, reflecting the time preference of the household.

  • (e)

    δ: Depreciation rate of net capital stock.

  • (f)

    ψ: Frisch elasticity of labour supply.

  • (g)

    ρR: Weight on lagges nominal interest rate in Taylor rule (extent of ‘interest rate smoothing’).

  • (h)

    ωπ: Coefficient on deviations of inflation from its steady-state value in Taylor’s rule.

  • (i)

    ωy: Coefficient on deviations of output from its steady-state value in Taylor’s rule.

C. Appendix – Gensys State Space

C.1. Log-linearized Equilibrium Conditions

We start by writing the system of equations in terms of k and s. This is done by replacing r, w, c^ and x^ in the first-order conditions with functions of the states. Thus, we start with:

(C-1)c^t+g^t+(1+gz)(1+gn)k^t+1(1δ)k^t=y^t=k^tα(ztlt)1α
(C-2)ψc^t1lt=(1τlt)(1α)k^tαltαzt1α
(C-3)(1+τxt)c^tσ(1lt)ψ(1σ)  =β^Etc^t+1σ(1lt+1)ψ(1σ)[αk^t+1α1(zt+1lt+1)1α+(1δ)(1+τxt+1)],

where ψ=λ/(1λ) and which can be reduced to the following:

ψ[k^tα(ztlt)1α(1+gn)(1+gz)k^t+1+(1δ)k^tg^t]    =(1τlt)(1α)k^tαltαzt1α(1lt)(1+τxt)[k^tα(ztlt)1α(1+gn)(1+gz)k^t+1+(1δ)k^tg^t]σ(1lt)ψ(1σ)    =β^Et[k^t+1α(zt+1lt+1)1α(1+gn)(1+gz)k^t+2        +(1δ)k^t+1g^t+1]σ(1lt+1)ψ(1σ)        [αk^t+1α1(zt+1lt+1)1α+(1δ)(1+τxt+1)].

Next, we compute the steady state of the system for constant values for z, the taxes and government spending:

k^/l=(1+τx)(1β^(1δ))β^αz1α1/(α1)c^=(k^/l)α1z1α(1+gz)(1+gn)+1δk^g^=ξ1k^g^c^=(1τl)(1α)(k^/l)αz1α/ψ(11/(k^/l)k^)=ξ2ξ3k^,

where the last two equations imply k^=(ξ2+g^)/(ξ1+ξ3),c^=ξ1k^g^, l=(1/(k^/l))k^.

The log-linearization is done around these steady-state values. Detrended consumption is obtained via (C-1) and given approximately by:

c^tc^logc^tk^α(zl)1α[αlogk^t+(1α)(logzt+loglt)](1+gz)(1+gn)k^logk^t+1+(1δ)k^logk^tg^logg^t.

The labour input is then derived from the static first-order condition (C-2):

0ψ{k^α(zl)1α[αlogk^t+(1α)(logzt+loglt)](1+gz)(1+gn)k^logk^t+1+(1δ)k^logk^tg^logg^t}+(1α)(1τl)k^αlαz1α(1l){1/(1τl)τltαlogk^t+αloglt(1α)logzt+l/(1l)loglt},

which we write succinctly as

loglt=φlklogk^t+φlzlogzt+φllτlt+φlglogg^t+φlklogk^t+1.

Using this equation for logl, we use the other static first-order conditions to write logy^, logx^ and logc^ as follows:

logy^t=φyklogk^t+φyzlogzt+φylτlt+φyglogg^t+φyklogk^t+1=(α+(1α)φlk)logk^t+(1α)(1+φlz)logzt+(1α)[φllτlt+φlklogk^t+1]logx^t=(1+gz)(1+gn)k^/x^logk^t+1(1δ)k^/x^logk^tlogc^t=φcklogk^t+φczlogzt+φclτlt+φcglogg^t+φcklogk^t+1=[y^logytx^logxtg^logg^t]/c^,

where the φ ’s are known functions of the parameters.

Capital is derived from the dynamic first-order condition (C-4):

0(1+τx)c^σ(1l)ψ(1σ)ψ(1σ)l/(1l)logltσlogc^t+c^σ(1l)ψ(1σ)τxtβ^Et{[αk^α1(zl)1α+(1δ)(1+τx)]·[c^σ(1l)ψ(1σ)ψ(1σ)l/(1l)loglt+1σlogc^t+1]+c^σ(1l)ψ(1σ)[αk^α1(zl)1α(1α)·(loglt+1+logzt+1logk^t+1)+(1δ)τxt+1]},

which simplifies to:

0(1+τx)ψ(1σ)l/(1l)logltσlogc^t+τxt  Et{(1+τx)ψ(1σ)l/(1l)loglt+1σlogc^t+1  +β^[r(1α)(loglt+1+logzt+1logk^t+1)+(1δ)τxt+1]},

where r=αy^/k^ and can be rewritten as:

(C-4)0φklloglt+φkclogc^t+τxt+φkElEtloglt+1+φkEcEtlogc^t+1+φkEzEtlogzt+1+φkEkEtlogk^t+1+φkEtxEtτxt+1.

C.2. Allowing for Adjustment Costs

To do log-linear computation (as in the baseline economy) in the case with adjustment costs and τct=τkt=0, we start with:

c^t+g^t+(1+gz)(1+gn)k^t+1(1δ)k^t+φ(x^t/k^t)k^t=y^t=k^tα(ztlt)1αψc^t1lt=(1τlt)(1α)k^tαltαzt1α(1+τxt)c^tσ(1lt)ψ(1σ)/(1φ(x^t/k^t))  =β^Etc^t+1σ(1lt+1)ψ(1σ)[αk^t+1α1(zt+1lt+1)1α+(1δ        φ(x^t+1/k^t+1)+φ(x^t+1/k^t+1)x^t+1/k^t+1)        (1+τxt+1)/1φ(x^t+1/k^t+1)],

where,

φ(x/k)=a2xkb2.

and b is set equal to the investment-capital trend rate (i.e. b=(1+gz)(1+gn)1+δ). To allow for different intensities of adjustment costs, a is raised from 0 (no adjustment costs) to 12.88 (the level used by Bernanke, Gertler, and Gilchrist (1999), the normal adjustment costs BGG level) to 4*12.88 (extreme adjustment costs).

Assuming φ(x^/k^)=φ(x^/k^)=0, the log-linearization of these equations yields the same results as in the benchmark with the exception of the intertemporal condition:

0(1+τx){ψ(1σ)l/(1l)logltσlogc^t+η(logx^tlogk^t)}+τxt  Et{(1+τx)ψ(1σ)l/(1l)loglt+1σlogc^t+1      +β^[r(1α)(loglt+1+logzt+1logk^t+1)        +(1+τx)(1+gz)(1+gn)η(logx^t+1logk^t+1)        +(1δ)τxt+1]},

where r=αy^/k^,η=φ(x^/k^)(x^/k^)=ab and the term in red is what is added to the baseline log-linearized dynamic equilibrium condition due to the presence of adjustment costs. This system can be rewritten as

(C-5)0φklloglt+φkclogc^t+τxt+φkxlogx^t+φkklogk^t+φkElEtloglt+1+φkEcEtlogc^t+1+φkEzEtlogzt+1+φkEkadjEtlogk^t+1+φkEtxEtτxt+1+φkExadjEtxt+1.

and differs from its baseline counterpart (C-4) by the terms in red.

C.3. Extension to MBCA – Šustek (2011)

The MBCA model features also bonds and prices. We thus have two additional equations which describe the dynamics of the nominal interest rate and inflation. The Taylor rule takes the form:

(C-6)Rt=(1ρR)R+ωy(logy^tlogy^)+ωπ(πtπ)+ρRRt1+R˜tRt=φπ0+φπylogy^t(st)+φππt+φπRRt1+R˜t.

whereas the first-order condition for bonds is given by

(C-7)(1+τbt)c^tσ(1lt)ψ(1σ)=β^Etc^t+1σ(1lt+1)ψ(1σ)[(1+τbt+1)exp(πt+1)(1+Rt)],

We follow the lines of Chari, Kehoe and McGrattan (2007) for the log-linearization of the F.O.C. for bonds and obtain

0(1+τb)c^σ(1l)ψ(1σ)ψ(1σ)l/(1l)logltσlogc^t      +c^σ(1l)ψ(1σ)τbt  β^Et{[(1+τb)exp(π)(1+R)]      ·[c^σ(1l)ψ(1σ)ψ(1σ)l/(1l)loglt+1σlogc^t+1]      +c^σ(1l)ψ(1σ)[(1+τb)exp(π)(1+R)          ·((1+τb)1τbt+1πt+1+Rt)]},

which simplifies to:

0ψ(1σ)l/(1l)logltσlogc^t+(1+τb)1τbt  Et{[ψ(1σ)l/(1l)loglt+1σlogc^t+1]      +[(1+τb)1τbt+1πt+1+Rt]},

and can be rewritten as:

0φRlloglt+φRclogc^t+φRτbτbt+φRElEtloglt+1+φREcEtlogc^t+1+φREτbEtτbt+1+Etπt+1Rt.

C.4. Gensys State Space Representation

C.4.1. BCA – Chari et al. (2007)

The models we are interested in can be cast in the form:

(C-8)Γ0yt=Γ1yt1+C+Ψϵt+Πηt

t=1,,T, where C is a vector of constants, t is an exogenously evolving, possibly serially correlated, random disturbance and ηt is an expectational error, satisfying Etηt+1=0, all t. The ηt terms are not given exogenously, but instead are treated as determined as part of the model solution.

Within the context of the BCA model, the matrices Γ0yt,Γ1yt1, C, Ψϵt,Πηt are given by:

Γ0yt=φkEk0010000φkl0φkcφkElφkEcφkEzφkEτx010000000000000001000000000000000100000000000000010000000000000001000000000φykφyzφyτl0φyg0100000000φxk00000010000000φlkφlzφlτl0φlg00010000000000100001000000000φcg0φcyφcx0010000000000001000000000000000010000010000000000000000100000000000logk^t+1logztτltτxtlogg^t1logy^tlogx^tlogltlogg^tobslogc^tEt{loglt+1}Et{logc^t+1}Et{logzt+1}Et{τx,t+1}
Γ1yt1=0000000000000000ρzρz,τlρz,τxρz,gz¯0000000000ρτl,zρτlρτl,τxρτl,gτ¯l0000000000ρτx,zρτx,τlρτxρτx,gτ¯x0000000000ρg,zρg,τlρg,τxρgg¯000000000000001000000000φyk00000000000000φxk00000000000000φlk00000000000000000000000000000000000000000000000000000001000000000000000100000000000000010000000000000001logk^tlogzt1τlt1τxt1logg^t11logy^t1logx^t1loglt1logg^t1obslogc^t1Et1{loglt}Et1{logc^t}Et1{logzt}Et1{τx,t}
C=000000000000000
Ψϵt=0000q11000q21q2200q31q32q330q41q42q43q440000000000000000000000000000000000000000εz,tετl,tετx,tεg,t
Πηt=000000000000000000000000000000000000000000001000010000100001ηEl,tηEc,tηEz,tηEτx,t

C.4.2. MBCA – Šustek (2011)

Within the context of the MBCA model, the matrices Γ0yt,Γ1yt1, C, Ψϵt,Πηt are given by:

Γ0yt=φkEk001000000φkl000φkcφkElφkEcφkEzφkEτx00010000000000000000000001000000000000000000000100000000000000000000010000000000000000000001000000000000000000000100000000000000000000010000000000000φykφyzφyτl0φyg0001000000000000φxk00000000100000000000φlkφlzφlτl0φlg00000100000000000000100000010000000000000001φπ0φπy000φπ1000000000000φRτb0000φRl001φRcφRElφREc00φREτb10000φcg000φcyφcx00001000000000000000010000000000000000000000001000000010000000000000000000000100000000000000000000001000000000000000000000000000100000000logk^t+1logztτltτxtlogg^tτbtR˜t1logy^tlogx^tlogltlogg^tobsπtRtlogc^tEt{loglt+1}Et{logc^t+1}Et{logzt+1}Et{τx,t+1}Et{τb,t+1}Et{πt+1}
Γ1yt1=0000000000000000000000ρzρz,τlρz,τxρz,gρz,τbρz,R˜z¯00000000000000ρτl,zρτlρτl,τxρτl,gρτl,τbρτl,R˜τ¯l00000000000000ρτx,zρτx,τlρτxρτx,gρτx,τbρτx,R˜τ¯x00000000000000ρg,zρg,τlρg,τxρgρg,τbρg,R˜g¯00000000000000ρτb,zρτb,τlρτb,τxρτb,gρτbρτb,R˜τ¯b00000000000000ρR˜,zρR˜,τlρR˜,τxρR˜,gρR˜,τbρR˜R˜¯0000000000000000000010000000000000φyk00000000000000000000φxk00000000000000000000φlk000000000000000000000000000000000000000000000000000000φπR0000000000000000000000000000000000000000000000000000000000000000100000000000000000000010000000000000000000001000000000000000000000100000000000000000000010000000000000000000001logk^tlogzt1τlt1τxt1logg^t1τbt1R˜t11logy^t1logx^t1loglt1logg^t1obsπt1Rt1logc^t1Et1{loglt}Et1{logc^t}Et1{logzt}Et1{τx,t}Et1{τb,t}Et1{πt}
C=000000000000000000000
Ψϵt=000000q1100000q21q220000q31q32q33000q41q42q43q4400q51q52q53q54q550q61q62q63q64q65q66000000000000000000000000000000000000000000000000000000000000000000000000000000000000000000εz,tετl,tετx,tεg,tετb,tεR˜,t
Πηt=000000000000000000000000000000000000000000000000000000000000000000000000000000000000000000100000010000001000000100000010000001ηEl,tηEc,tηEz,tηEτx,tηEτb,tηEπ,t

C.4.3. BCA – Chari et al. (2007) With Adjustment Costs

Allowing for adjustment costs in the standard BCA model requires modifying the state vector y(t) and the matrices Γ0yt,Γ1yt1, C, Ψϵt,Πηt in the following way:

Γ0yt=φkEkadj001000φkxadjφkl0φkcφkElφkEcφkEzφkEτxφkExadj01000000000000000010000000000000000100000000000000001000000000000000010000000000φykφyzφyτl0φyg01000000000φxk000000100000000φlkφlzφlτl0φlg0001000000000001000010000000000φcg0φcyφcx0010000000000000100000000000000000100000010000000000000000010000000000000000000100000000logk^t+1logztτltτxtlogg^t1logy^tlogx^tlogltlogg^tobslogc^tEt{loglt+1}Et{logc^t+1}Et{logzt+1}Et{τx,t+1}Et{logxt+1}
Γ1yt1=φkkadj0000000000000000ρzρz,τlρz,τxρz,gz¯00000000000ρτl,zρτlρτl,τxρτl,gτ¯l00000000000ρτx,zρτx,τlρτxρτx,gτ¯x00000000000ρg,zρg,τlρg,τxρgg¯00000000000000010000000000φyk000000000000000φxk000000000000000φlk0000000000000000000000000000000000000000000000000000000000100000000000000001000000000000000010000000000000000100000000000000001logk^tlogzt1τlt1τxt1logg^t11logy^t1logx^t1loglt1logg^t1obslogc^t1Et1{loglt}Et1{logc^t}Et1{logzt}Et1{τx,t}Et1{logxt}
C=0000000000000000
Ψϵt=0000q11000q21q2200q31q32q330q41q42q43q44000000000000000000000000000000000000000000000000εz,tετl,tετx,tεg,t
Πηt=00000000000000000000000000000000000000000000000000000001000001000001000001000001ηEl,tηEc,tηEz,tηEτx,tηExt

C.4.4. MBCA – Šustek (2011) with Adjustment Costs

Allowing for adjustment costs in the standard BCA model requires modifying the state vector y(t) and the matrices Γ0yt,Γ1yt1, C, Ψϵt,Πηt in the following way:

Γ0yt=φkEkadj00100000φkxadjφkl000φkcφkElφkEcφkEzφkEτx00φkEτxadj0100000000000000000000001000000000000000000000010000000000000000000000100000000000000000000001000000000000000000000010000000000000000000000100000000000000φykφyzφyτl0φyg00010000000000000φxk000000001000000000000φlkφlzφlτl0φlg0000010000000000000001000000100000000000000001φπ0φπy000φπ10000000000000φRτb0000φRl001φRcφRElφREc00φREτb100000φcg000φcyφcx0000100000000000000000100000000000000000000000001000000001000000000000000000000001000000000000000000000001000000000000000000000000000010000000000000000001000000000000logk^t+1logztτltτxtlogg^tτbtR˜t1logy^tlogx^tlogltlogg^tobsπtRtlogc^tEt{loglt+1}Et{logc^t+1}Et{logzt+1}Et{τx,t+1}Et{τb,t+1}Et{πt+1}Et{logxt+1}
Γ1yt1=φkkadj0000000000000000000000ρzρz,τlρz,τxρz,gρz,τbρz,R˜z¯000000000000000ρτl,zρτlρτl,τxρτl,gρτl,τbρτl,R˜τ¯l000000000000000ρτx,zρτx,τlρτxρτx,gρτx,τbρτx,R˜τ¯x000000000000000ρg,zρg,τlρg,τxρgρg,τbρg,R˜g¯000000000000000ρτb,zρτb,τlρτb,τxρτb,gρτbρτb,R˜τ¯b000000000000000ρR˜,zρR˜,τlρR˜,τxρR˜,gρR˜,τbρR˜R˜¯000000000000000000000100000000000000φyk000000000000000000000φxk000000000000000000000φlk00000000000000000000000000000000000000000000000000000000φπR00000000000000000000000000000000000000000000000000000000000000000001000000000000000000000010000000000000000000000100000000000000000000001000000000000000000000010000000000000000000000100000000000000000000001logk^tlogzt1τlt1τxt1logg^t1τbt1R˜t11logy^t1logx^t1loglt1logg^t1obsπt1Rt1logc^t1Et1{loglt}Et1{logc^t}Et1{logzt}Et1{τx,t}Et1{τb,t}Et1{πt}Et1{logxt}
C=0000000000000000000000
Ψϵt=000000q1100000q21q220000q31q32q33000q41q42q43q4400q51q52q53q54q550q61q62q63q64q65q66000000000000000000000000000000000000000000000000000000000000000000000000000000000000000000εz,tετl,tετx,tεg,tετb,tεR˜,t
Πηt=0000000000000000000000000000000000000000000000000000000000000000000000000000000000000000000000000000000001000000010000000100000001000000010000000100000001ηEl,tηEc,tηEz,tηEτx,tηEτb,tηEπ,tηEx,t

D. Appendix – Derivatives with Alternative Stepsize

We here report the results for the case where the stepsize used to compute the numerical derivatives is not set to 1e−3 like in Komunjer and Ng (2011) but rather automatically selected by Matlab using the function nuderst. The returned step size is the maximum of 1e−4 times the absolute value of the current parameter and 1e−7.

The main results of the previous analysis hold through with two notable exceptions. Fist, as becomes evident from Tables D-1 and D-2, both the baseline BCA and MBCA model are strictly identifiable already at a tolerance level of 1e−9 (vs. 1e−11). Second, the BCA model is strictly identifiable even when the deep parameters are estimated at a tolerance level of 1e−10, as reported in Table D-3. This is not the case for the MBCA model (see Table D-5). A few things:

Table D-1.

Komunjer and Ng Test Results BCA Model.

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 29 25 15 51 40 62 0
e−03 29 25 16 54 45 69 0
e−04 29 25 16 54 45 69 0
e−05 29 25 16 54 45 69 0
e−06 29 25 16 54 45 69 0
e−07 30 25 16 54 46 69 0
e−08 30 25 16 54 46 69 0
e−09 30 25 16 55 46 71 1
e−10 30 25 16 55 46 71 1
e−11 30 25 16 55 46 71 1
Default = 2.756906e−12 30 25 16 55 46 71 1
Required 30 25 16 55 46 71 1

Summary: nθ=30,nX=5,nε=4.

Order condition: nθ=30,nδ=50.

Table D-2.

Komunjer and Ng Test Results BCA Model (Deep Parameters Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 33 25 15 52 44 64 0
e−03 34 25 16 57 50 71 0
e−04 34 25 16 57 50 71 0
e−05 35 25 16 57 51 71 0
e−06 36 25 16 57 52 71 0
e−07 36 25 16 59 52 72 0
e−08 37 25 16 59 53 73 0
e−09 37 25 16 60 53 74 0
e−10 37 25 16 62 53 76 0
e−11 37 25 16 62 53 78 1
Default = 5.513812e−12 37 25 16 62 53 78 1
Required 37 25 16 62 53 78 1

Summary: nθ=37,nX=5,nε=4.

Order condition: nθ=37,nδ=50.

Table D-3.

Komunjer and Ng Test Results MBCA Model.

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 60 49 33 101 89 130 0
e−03 60 49 36 108 96 143 0
e−04 60 49 36 109 96 144 0
e−05 60 49 36 109 96 144 0
e−06 60 49 36 109 96 144 0
e−07 61 49 36 109 97 145 0
e−08 61 49 36 110 97 145 0
e−09 61 49 36 110 97 146 1
e−10 61 49 36 110 97 146 1
Default = 1.165290e−11 61 49 36 110 97 146 1
e−11 61 49 36 110 97 146 1
Required 61 49 36 110 97 146 1

Summary: nθ=61,nX=7,nε=6.

Order condition: nθ=61,nδ=105.

Table D-4 (D-6) reports results for when τbss and R˜ss (and deep parameters) are estimated.

Table D-4.

Komunjer and Ng Test Results MBCA Model (τbss and R˜ss Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 60 49 33 101 89 130 0
e−03 60 49 36 108 96 143 0
e−04 60 49 36 109 96 144 0
e−05 60 49 36 109 96 144 0
e−06 60 49 36 109 96 144 0
e−07 61 49 36 109 97 145 0
Default = 1.165290e−11 61 49 36 110 97 146 0
e−08 61 49 36 110 97 146 0
e−09 61 49 36 110 97 146 0
e−10 61 49 36 110 97 146 0
e−11 61 49 36 110 97 146 0
Required 63 49 36 112 99 148 1

Summary: nθ=63,nX=7,nε=6.

Order condition: nθ=63,nδ=105.

Problematic parameters at Tol = 1e−3: τbss, R˜ss.

Problematic parameters at Tol = 1.000000e−11: τlss, τxss, gss, τbss, R˜ss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q21, q31, q51, q22, q32, q42, q52, q33, q53, q63, q44, q54.

Table D-5.

Komunjer and Ng Test Results MBCA Model (Deep Parameters Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 67 49 35 104 97 133 0
e−03 68 49 36 113 104 146 0
e−04 69 49 36 115 105 148 0
e−05 70 49 36 115 106 148 0
e−06 70 49 36 116 106 149 0
e−07 71 49 36 118 107 149 0
e−08 72 49 36 118 108 151 0
e−09 72 49 36 120 108 154 0
e−10 72 49 36 121 108 156 0
Default = 2.330580e−11 72 49 36 121 108 156 0
e−11 72 49 36 121 108 157 1
Required 72 49 36 121 108 157 1

Summary: nθ=72,nX=7,nε=6.

Order condition: nθ=72,nδ=105.

Problematic parameters at Tol = 1e−3: zss, τlss, τxss, gss, ρτl,z, ρz,τl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρτl,g, ρτl,τb, ρτl,R˜, ρτx,R˜, q21, q22, ψ.

Problematic parameters at Tol = 2.330580e−11: zss, τlss, τxss, gss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q11, q21, q31, q41, q51, q61, q22, q32, q42, q52, q62, q33, q43, q53, q63, q44, q54, q64, q55, q65, q66, gn, gz, β, δ, ψ, σ, α, ρR, ωπ, ωy, πss.

Table D-6.

Komunjer and Ng Test Results MBCA Model (τbss,R˜ss and Deep Parameters Estimated).

Tol ΔΛS ΔTS ΔUS ΔΛTS ΔΛUS ΔS Pass
e−02 67 49 35 104 97 133 0
e−03 68 49 36 113 104 146 0
e−04 69 49 36 115 105 148 0
e−05 69 49 36 115 105 148 0
e−06 70 49 36 116 106 148 0
e−07 71 49 36 118 107 150 0
Default = 2.330580e−11 72 49 36 121 108 157 0
e−08 72 49 36 119 108 152 0
e−09 72 49 36 121 108 154 0
e−10 72 49 36 121 108 156 0
e−11 72 49 36 121 108 157 0
Required 74 49 36 123 110 159 1

Summary: nθ=74,nX=7,nε=6.

Order condition: nθ=74,nδ=105.

Problematic parameters at Tol = 1e−3: zss, τlss, τxss, gss, τbss, R˜ss, ρτl,z, ρz,τl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρτl,g, ρτl,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρτb,R˜, q21, q22, ψ, σ, ωπ, πss.

Problematic parameters at Tol = 1.000000e−11: zss, τlss, τxss, gss, τbss, R˜ss, ρz, ρτl,z, ρτx,z, ρg,z, ρτb,z, ρR˜,z, ρz,τl, ρτl, ρτx,τl, ρg,τl, ρτb,τl, ρR˜,τl, ρz,τx, ρτl,τx, ρτx, ρg,τx, ρτb,τx, ρR˜,τx, ρz,g, ρτl,g, ρτx,g, ρg, ρτb,g, ρR˜,g, ρz,τb, ρτl,τb, ρτx,τb, ρg,τb, ρτb, ρR˜,τb, ρz,R˜, ρτl,R˜, ρτx,R˜, ρg,R˜, ρτb,R˜, ρR˜, q11, q21, q31, q41, q51, q61, q22, q32, q42, q52, q62, q33, q43, q53, q63, q44, q54, q64, q55, q65, q66, gn, gz, β, δ, ψ, σ, α, ρR, ωπ, ωy, πss.

Notes

1

Global identification would extend the uniqueness requirement to the whole parameter space.

2

A point is regular if it belongs to an open neighborhood where the rank of the matrix does not vary.

3

Please refer to Iskrev (2010) for further details.

4

This detrending method is used in Brinca et al. (2016) and is found to make the ML estimation procedure very robust across a large set of OECD countries.

5

We use Chari et al. (2007) actual data for the study of the 1982 recession.

6

Testing that the fi statistics are not significantly different from each other would be the most rigorous way of answering this question.

7

In cases where the upper bound and the lower bound only overlapped at a given number (e.g. fi statistic of wedge 1 UB = 0.02 and fi statistic of wedge 2 LB = 0.02), we still counted them as the fi statistic overlapping and so the number would not be reported in bold.

8

As pointed out by Qu and Tkachenko (2012) it is usually the case that δ(θ¯0) is unknown and, thus, so is the domain of the curve. This is why in constructing the non-identification curve first a wide support is considered, the model solved and the spectrum computed. The resulting curve is then truncated so as to exclude points which (i) are associated with indeterminacy, (ii) violate the natural bounds of the parameters and (iii) yield fθ¯(ω) different from fθ¯0(ω).

9

There is no need to consider ω[π,0] because fθ¯(ω) is equal to the conjugate of fθ¯(ω).

References

Bernanke, B., Gertler, M., & Gilchrist, S. (1999, March). The financial accelerator in a quantitative business cycle framework. In J. B. Taylor & M. Woodford (Eds.), Handbook of Macroeconomics (pp. 13411393). Amsterdam: Elsevier.

Bowden, R. J. (1973). The theory of parametric identification. Econometrica, 41, 10691074.

Brinca, P. (2013). Monetary business cycle accounting for Sweden. The B.E. Journal of Macroeconomics, 13(1), 21946116.

Brinca, P. (2014). Distortions in the Neoclassical Growth Model: A cross-country analysis. Journal of Economic Dynamics and Control, 47(C), 119.

Brinca, P., Chari, V. V., Kehoe, P. J., & McGrattan, E. R. (2016). Accounting for business cycles with wedges. In J. B. Taylor & H. Uhlig (Eds.), Handbook of macroeconomics (2nd ed., pp. 10131063). Amsterdam: Elsevier.

Canova, F., & Sala, L. (2009, May). Back to square one: Identification issues in DSGE models. Journal of Monetary Economics, 56(4), 431449.

Chari, V. V., Kehoe, P. J., & McGrattan, E. R. (2007). Business cycle accounting. Econometrica, 75(3), 781836.

Cociuba, S. E., & Ueberfedt, A. A. (2010). Trends in U.S. hours and the labor wedge. Bank of Canada Working Paper Series No. 28. https://ideas.repec.org/p/bca/bocawp/10-28.html

Dooyeon, C., & Doblas-Madrid, A. (2012). Business cycle accounting East and West: Asian Finance and the Investment Wedge. Review of Economic Dynamics. doi:10.1016/j.red.2012.10.003

Iskrev, N. (2010). Evaluating the strength of identification in DSGE models. An a priori approach. Technical report.

Islam, N., Dai, E., & Sakamoto, H. (2006). Role of TFP in China’s growth. Asian Economic Journal, 20(2), 127156.

Kobayashi, K., & Inaba, M. (2006). Business cycle accounting for the Japanese economy. Japan and the World Economy, 18(4), 418440.

Komunjer, I., & Ng, S. (2011, November). Dynamic identification of dynamic stochastic general equilibrium models. Econometrica, 79(6), 19952032.

Lama, R. (2009). Accounting for output drops in Latin America. IMF Working Paper Series 67. International Monetary Fund. https://ideas.repec.org/p/imf/imfwpa/2009-067.html

Otsu, K. (2009). International business cycle accounting. IMES Discussion Paper Series 29. Institute for Monetary and Economic Studies, Bank of Japan. https://ideas.repec.org/p/ime/imedps/09-e-29.html

Qu, Z., & Tkachenko, D. (2012, March). Identification and frequency domain quasi-maximum likelihood estimation of linearized dynamic stochastic general equilibrium models. Quantitative Economics, 3(1), 95132.

Qu, Z., & Tkachenko, D. (2017). Global identification in DSGE models allowing for indeterminacy. Review of Economic Studies, 84(3), 13061345.

Rao, C. R. (1971). Linear statistical inference and its applications (Vol. 39, 577591). New York, NY: Wiley.

Restrepo-Echavarria, P., & Cheremukhin, A. A. (2010). The labor wedge as a matching friction. European Economic Review, 68(C), 7192.

Rothenberg, T. (1971). Identification in parametric models. Econometrica, 39, 577591.

Simonovska, I., & Söderling, L. (2008). Business cycle accounting for Chile. IMF Working Paper Series No. 61. International Monetary Fund. https://ideas.repec.org/p/imf/imfwpa/2008-061.html

Sims, C. (2002, October). Solving linear rational expectations models. Computational Economics, 20(1–2), 120.

Šustek, R. (2011, October). Monetary business cycle accounting. Review of Economic Dynamics, 14(4), 592612.

Tucker, L., Cooper, L., & Meredith, W. (1972). Obtaining squared multiple correlations from a correlation matrix which may be singular. Psychometrika, 37, 143148.

Acknowledgements

The views expressed are those of the authors and do not necessarily reflect those of the Bank of Portugal, the Federal Reserve Board, the Federal Reserve System or their staff. We thank Stéphane Bonhomme, Fabio Canova, Ellen McGrattan, Ricardo Reis, Barbara Rossi, Roman Šustek and Harald Uhlig for stimulating comments. We are also grateful to seminar participants at Norges Bank Brown Bag seminar, 2nd HenU/INFER Workshop on Applied Macroeconomics, The University of Chicago Econometrics Reading Group, 3rd Annual Conference of the International Association for Applied Econometrics, 22nd International Conference of the Society for Computational Economics on Computing in Economics and Finance, 10th Annual Meeting of the Portuguese Economic Journal and 31st Annual Congress of the European Economic Association. Pedro Brinca acknowledges funding from Fundação para a Ciência e a Tecnologia (UID/ECO/00124/2013, UID/ECO/00124/2019, PTDC/EGE-ECO/7620/2020; and Social Sciences DataLab, LISBOA-01-0145-FEDER-022209), POR Lisboa (LISBOA-01-0145-FEDER-007722 and LISBOA-01-0145-FEDER-022209), POR Norte (LISBOA-01-0145-FEDER-022209) and CEECIND/02747/2018.