192
Macroeconomics of small open economies Manuel W¨ alti Inauguraldissertation zur Erlangung der W¨ urde eines Doctor rerum oeconomicarum der Wirtschafts- und Sozialwissenschaftlichen Fakult¨atderUniversit¨atBern Bern, October 2004

Macroeconomics of small open economies

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Macroeconomics of small open economies

Macroeconomics of

small open economies

Manuel Walti

Inauguraldissertation

zur Erlangung der Wurde eines

Doctor rerum oeconomicarum

der Wirtschafts- und Sozialwissenschaftlichen

Fakultat der Universitat Bern

Bern, October 2004

Page 2: Macroeconomics of small open economies

Wälti, Manuel: Macroeconomics of small open economies / Manuel Wälti. – Als Ms. gedr.. – Berlin : dissertation.de – Verlag im Internet GmbH, 2005 Zugl.: Bern, Univ., Diss., 2004 ISBN 3-89825-954-4 Die Fakultät hat diese Arbeit am 18. November 2004 auf Antrag der beiden Gutachter Prof. Dr. Harris Dellas und Prof. Dr. Klaus Neusser, als Dissertation angenommen, ohne damit zu den darin ausge-sprochenen Auffassungen Stellung nehmen zu wollen.

Bibliografische Information der Deutschen Bibliothek Die Deutsche Bibliothek verzeichnet diese Publikation in der Deutschen Nationalbibliografie; detaillierte bibliografische Daten sind im Internet über <http://dnb.ddb.de> abrufbar. dissertation.de – Verlag im Internet GmbH 2005 Alle Rechte, auch das des auszugsweisen Nachdruckes, der auszugsweisen oder vollständigen Wiedergabe, der Speicherung in Datenverarbeitungsanlagen, auf Datenträgern oder im Internet und der Übersetzung, vorbehalten. Es wird ausschließlich chlorfrei gebleichtes Papier (TCF) nach DIN-ISO 9706 verwendet. Printed in Germany. dissertation.de - Verlag im Internet GmbH Pestalozzistraße 9 10 625 Berlin URL: http://www.dissertation.de

Page 3: Macroeconomics of small open economies

Preface

This thesis deals with topics in applied macroeconomics. Chapter 1 sur-

veys the strand of literature in the field of open-economy macroeconomics

which applies quantitative models to the issue of transmission properties

of economic disturbances and international policy. We discuss the genesis

of quantitative open-economy models, their ability to match the data, and

their use as a laboratory for policy analysis. In doing so, we concentrate on

studies dealing with the role of exchange rate stabilization in the conduct

of monetary policy and the choice of exchange rate arrangement.

In chapter 2, we compare Taylor type interest rules which differ in the

size of the reaction coefficient on the real exchange rate. The alternative

policy regimes are evaluated in terms of macroeconomic performance and

welfare within the framework of an artificial small-open economy with opti-

mizing agents, a moderate degree of nominal price rigidity, and five kinds of

shocks (domestic and foreign). We limit the consideration to feedback rules

which require only information which could plausibly be possessed by the

central bank. The results are discussed against the background of two nat-

ural benchmark policies: strict domestic inflation targeting and a credible

and unilateral peg toward the currency of the rest of the world. We find that

introducing a moderate form of real exchange rate targeting in the original

Taylor rule induces higher welfare with respect to shocks to productivity

and foreign demand and lower welfare with respect to shocks to govern-

ment consumption and the terms of trade. However, the outcome under

rules which, unlike the original Taylor rule, allow for a considerable degree

of interest smoothing, is robust regarding the inclusion of real exchange rate

targeting.

Page 4: Macroeconomics of small open economies

iv PREFACE

In an influential paper, Jordi Gali (1999)1 studies the effects of tech-

nology shocks on employment in the G7 countries using a structural vec-

tor autoregressive model (VAR) approach. Gali finds that the response of

employment to a positive technology shock is negative and persistent. In

chapter 3, we repeat his analysis with two important modifications. First,

we add an open-economy block to Gali’s five-variable framework. Second,

the use of structural vector error correction model (VECM) methods allows

us to investigate the effect of a productivity shock on employment based

on less restrictive assumptions than Gali does. His findings, however, are

largely confirmed.

1Technology, employment, and the business cycle: do technology shocks explain ag-

gregate fluctuations?, American Economic Review 1999, 89(1), 249-271.

Page 5: Macroeconomics of small open economies

Acknowledgement

When writing this thesis, I benefited from the input of many people. Above

all I wish to thank my advisor Harris Dellas for his constant attention and

guidance. I would like to thank Klaus Neusser for kindly taking the role of

a co-examiner and for giving many valuable hints.

I am very grateful to Carlos Lenz for providing me his method for esti-

mating structural VARs and crucial sections of the RATS code; the third

chapter of this book has also greatly benefited from discussions with him.

I thank the participants of the Wednesday seminar at the Economics De-

partment – Esther Brugger, Alain Egli, Armin Hartmann, Roland Hodler,

Simon Lortscher, and Michael Manz – as well as Kurt Schmidheiny for their

valuable comments and their friendship.

Finally, I would like to thank the Swiss National Bank (SNB) for sup-

porting this project in its final stage and Werner Hermann and Ulrich Kohli

for putting the necessary time at my disposal. Thanks also to Christoph

Meyer for his assistance in obtaining the OECD MEI and IMF data at a

time when I was not yet familiar with EASY, the SNB’s economic analysis

system.

Page 6: Macroeconomics of small open economies
Page 7: Macroeconomics of small open economies

Contents

Contents . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . vii

1 Quantitative open-economy NNS models 1

1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 1

1.2 Genesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3

1.2.1 International RBC approach . . . . . . . . . . . . . . 3

1.2.2 Market incompleteness and the small country assump-

tion . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

1.2.3 New Neoclassical Synthesis . . . . . . . . . . . . . . . 7

1.3 Model evaluation . . . . . . . . . . . . . . . . . . . . . . . . 9

1.4 Laboratory for policy analysis . . . . . . . . . . . . . . . . . 13

1.5 Concluding remarks . . . . . . . . . . . . . . . . . . . . . . . 18

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2 The role of exchange rate stabilization 25

2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

2.2 The Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

2.2.1 Behavior of final good producers . . . . . . . . . . . . 29

2.2.2 Digression: Total demand for input factor i . . . . . 31

2.2.3 Behavior of intermediate good producer i . . . . . . . 32

2.2.4 Behavior of representative agent . . . . . . . . . . . . 37

2.2.5 Fiscal and monetary policy . . . . . . . . . . . . . . . 41

2.2.6 Market clearing . . . . . . . . . . . . . . . . . . . . . 42

2.2.7 Closing the model: International asset markets . . . . 43

2.3 Solution, parameterization, and diagnostic check . . . . . . . 44

2.3.1 Baseline parameterization . . . . . . . . . . . . . . . 45

2.3.2 Dynamic effects . . . . . . . . . . . . . . . . . . . . . 47

Page 8: Macroeconomics of small open economies

viii CONTENTS

2.4 The role of exchange rate stabilization . . . . . . . . . . . . 54

2.4.1 Interest-rate rules to be investigated . . . . . . . . . 56

2.4.2 Simulation results . . . . . . . . . . . . . . . . . . . . 60

2.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

2.A Stationary representation . . . . . . . . . . . . . . . . . . . . 68

2.A.1 Change in notation and useful simplifications . . . . 68

2.A.2 Equilibrium conditions . . . . . . . . . . . . . . . . . 70

2.A.3 Deflating the system . . . . . . . . . . . . . . . . . . 72

2.B Non-stochastic steady state . . . . . . . . . . . . . . . . . . 74

2.C First-order approximation . . . . . . . . . . . . . . . . . . . 77

2.C.1 Linear system . . . . . . . . . . . . . . . . . . . . . . 77

2.C.2 Digression: The New Phillips curve . . . . . . . . . . 80

2.D Exchange rate peg . . . . . . . . . . . . . . . . . . . . . . . 80

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

Tables . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

3 Technology shocks and employment in open economies 87

3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . 87

3.2 Benchmark model . . . . . . . . . . . . . . . . . . . . . . . . 92

3.2.1 Estimation method . . . . . . . . . . . . . . . . . . . 92

3.2.2 Two critical remarks . . . . . . . . . . . . . . . . . . 94

3.2.3 Replication results . . . . . . . . . . . . . . . . . . . 97

3.3 Nominal block . . . . . . . . . . . . . . . . . . . . . . . . . . 100

3.3.1 Estimation method . . . . . . . . . . . . . . . . . . . 100

3.3.2 Replication results . . . . . . . . . . . . . . . . . . . 102

3.4 Extensions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105

3.4.1 Open-economy block . . . . . . . . . . . . . . . . . . 105

3.4.2 Structural VECM approach . . . . . . . . . . . . . . 109

3.5 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . 112

3.A Univariate unit root tests . . . . . . . . . . . . . . . . . . . . 114

3.B Estimation of structural VARs and VECMs . . . . . . . . . 115

3.B.1 Non-cointegrated case . . . . . . . . . . . . . . . . . 115

3.B.2 Cointegrated case . . . . . . . . . . . . . . . . . . . . 121

3.C Bootstrap confidence intervals . . . . . . . . . . . . . . . . . 126

3.D Model specifications . . . . . . . . . . . . . . . . . . . . . . 128

Bibliography . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133

Page 9: Macroeconomics of small open economies

CONTENTS ix

Tables and figures . . . . . . . . . . . . . . . . . . . . . . . . . . . 137

Page 10: Macroeconomics of small open economies
Page 11: Macroeconomics of small open economies

Chapter 1

Quantitative

new neoclassical synthesis

models of open economies

1.1 Introduction

The traditional framework for studying policy issues in international macroe-

conomics – such as the choice of optimal exchange rate arrangements –

has long been the Mundell-Fleming-Dornbusch model. Since the Mundell-

Fleming-Dornbusch model does not provide an internal welfare criterion,

it has, for the purpose of policy evaluation, typically been supplemented

with an ad hoc loss function depending on output and inflation variabil-

ity. In recent years, the same issues have increasingly been investigated in

frameworks which integrate monopolistic competition in goods and/or labor

markets and some form of nominal rigidity into dynamic general equilib-

rium (DGE) settings. Given its solid microfoundations, the new generation

of open-economy models is far better suitable for serious policy evaluation

exercises.

Broadly speaking, there are two strands of the new open-economy macroe-

conomics literature. Models following the approach initiated by Obstfeld

and Rogoff [26], on the one hand, possess an analytical solution.1 This

advantage, however, comes at a price: It involves the imposition of many

1For a survey of small analytical open-economy models, see Lane [26] and Sarno [34].

Page 12: Macroeconomics of small open economies

2 CHAPTER 1

simplifying assumptions such as that prices are predetermined rather than

sluggish, that physical capital is exogenous rather than endogenous, that

monetary policy is passive with money stock as the instrument rather than

following a Taylor-type interest rule, that changes in monetary policy are the

only kind of shocks rather than allowing for a reasonable number of struc-

tural driving forces (including supply shocks), and that the representative

agent’s momentary utility is separable between consumption and leisure

rather than nonseparable (as is typical in the real business cycle litera-

ture). Moreover, the small -open economy version typically abstracts from

the stability problems attached to a small open economy setting (an issue

to which we will return later). All these simplifications and abstractions

make Obstfeld and Rogoff [26] and its numerous successors too stylized for

the purpose of calibration or estimation and, thus, inappropriate for the

purpose of policy analysis.

Models in the real business cycle (RBC) tradition such as Kollmann [24]

and Chari et al. [9], on the other hand, do not admit a closed-form but have

to be (approximately) solved based on numerical methods. Due to their

richer structure, however, they can be brought to data. In other words, not

only the models’ ability to explain qualitative features of observed business

fluctuations can be tested, but also their ability to account for statistical

properties like the standard deviations and cross-correlations of observed

aggregate time series. It is this capability that in our view qualifies them

for the purpose of serious policy analysis.

The present paper selectively surveys recent studies in the field of new

open-economy macroeconomics which apply realistically quantitative mod-

els to the issue of transmission properties of economic disturbances and

international policy. The remainder of the paper is organized as follows.

Section 1.2 discusses the genesis of modern quantitative open-economy busi-

ness cycle models. Section 1.3 discusses these models’ ability to fit the data

and section 1.4 their use as a laboratory for policy analysis. Section 1.5

concludes.

Page 13: Macroeconomics of small open economies

1.2. GENESIS 3

1.2 Genesis of quantitative open-economy busi-

ness cycle models

The modern quantitative open-economy business cycle models have their

roots in the methodological advances of the RBC approach and elements

of the New Keynesian models of the 1980s. We begin this section with

a brief discussion of what we consider to be the main contribution of the

RBC literature, the demonstration of how general equilibrium models can

be taken to data. Then, we point to the stability problem related to the

neoclassical model of a small open economy and the remedies suggested in

the literature. Lastly, we give a brief summary of the emergence of the

so-called New Neoclassical Synthesis.

1.2.1 International RBC approach

Nowadays, most macroeconomists would agree that the main contribution

of the RBC literature to the profession has been in terms of methodology

(see e.g. Woodford [40]): First, the RBC literature has shown how general

equilibrium models can be made quantitative; this involves realistic numer-

ical parameter values and methods to compute numerical solutions to the

equations of a model. Second, the RBC literature has emphasized the as-

sessment of a model’s validity to fit the data by comparing quantitative

features of the theoretical economy such as the standard deviations, the

cross-correlations, and the autocorrelations for key variables of the model

to those of observed aggregate time series, using some informal distance

criterion.

Let us illustrate this point by means of an example taken from the in-

ternational RBC literature. The two-country model of Backus et al. [1],

section 5, is a typical exponent of this approach. Both countries specialize

in the production of a single, imperfect substitutable intermediate good,

which implies that all trade between the countries is in intermediate goods.

Intermediate goods producers use domestic physical capital and labor as

inputs; both capital and labor, are immobile internationally. Each country

also produces a non-tradable final good, which is used for consumption,

investment, and government spending. The final good is produced by as-

Page 14: Macroeconomics of small open economies

4 CHAPTER 1

sembling domestic and imported intermediate goods. In both countries, the

representative household owns domestic firms, the capital stock, and the

time endowment; the capital stock and the time devoted to labor are rented

to the intermediate good producer. The household has access to a com-

plete contingent-claims market. Needless to say, all markets are perfectly

competitive and prices are fully flexible.

Among the notable features of international macroeconomic data and

the economic connections among countries which received much attention at

the time, is the observation that the correlations of output across countries

are larger than the analogous correlations for consumption and productivity

and that the terms of trade of the individual countries are highly variable

and persistent (compare e.g. Backus et al.). How successful is the model of

Backus et al. in mimicking these properties? A first finding is that in the

theoretical economy the consumption correlation exceeds the productivity

and output correlations. A second finding is that the fluctuations in the

terms of trade are much less variable in the theoretical economy than in the

data. Since the two findings are robust to a number of reasonable changes

in the economy (in terms of parameter values and key assumptions), Backus

et al. label them, respectively, the consumption correlations and the terms

of trade anomaly. We conclude that while for the basic version of a closed-

economy RBC model the match between theory and observations (with

respect to real economic activity) is surprisingly well,2 the open-economy

counterpart is far less successful.

1.2.2 Market incompleteness and the small country

assumption

So far, we have been talking about extensions of the basic closed-economy

RBC model, which involves the existence of a complete contingent-claims

market, to a two-country setting. An alternative setting in international

macroeconomics, however, is the case of a small open economy, often in

combination with a bonds-only structure.

From the growth literature, it is known that modifying the basic neo-

classical growth model to allow for mobility of goods and international bor-

2For a critical assessment of this statement, see e.g. King and Rebelo [23].

Page 15: Macroeconomics of small open economies

1.2. GENESIS 5

rowing and lending features two undesirable properties (compare e.g. Barro

and Sala-i-Martin [3], chapter 3): The first is that the adjustment toward

the steady state is instantaneous. The second is that if the world interest

rate is not equal the small open economy’s rate of time preference, in the

long run consumption either approaches zero and net foreign assets reach

a negative lower bound, or both consumption and net foreign assets, grow

forever and get infinite (which would imply that the small open economy

finally gets a big player). However, if the two interest rates are equal, then

the steady state depends on the economy’s initial net foreign asset position.

Translated into a stochastic framework this means that transitory shocks

to the world interest rate induce permanent changes in the net foreign asset

position. This raises a problem since standard solution techniques rely on

the existence of a stationary steady state.3

Allowing for complete contingent-claims markets would be a way to over-

come the stability problem.4 The state-contingent claims protect the small

open economy against future contingencies and, thus, work as an insurance

against the shocks’ effects. For standard specifications of the utility func-

tion, deviations of domestic consumption from their steady state values are

proportional to their foreign counterpart. In this event, since the rest of the

world behaves as a (stationary) closed-economy, the same behavior must

apply for the small open economy.

An implication of the complete markets assumption is that there is

no current account surplus or deficit – which is the reason why anyone

who intends to study current account dynamics does not assume complete

3In the words of Ghironi [21], p. 3 [emphasize in original]: ”When the model is log-

linearized, one is actually approximating its dynamics around a ’moving steady state’.

(...) De facto, one cannot perform any stochastic analysis.”4Recall that we focus here on a stochastic setting. In a deterministic setting, however,

there is no point in modelling complete markets, since there are no risks to pool.

Page 16: Macroeconomics of small open economies

6 CHAPTER 1

contingent-claims markets in the first place.5, 6 Hence, whenever we want

to make predictions regarding the dynamics of the stock of external assets

minus liabilities (the net external position) or the change in the net external

position from one period to the next (the current account), we are bound

to assume that the only asset nations trade is a one-period bond that offers

a certain one-period return, i.e., that there is free borrowing and lending in

riskless one-period bonds but no trade in contingent claims.

The literature suggests various remedies to the notorious deficiencies

in the small-country, bonds-only version of the neoclassical growth model

mentioned above. The introduction of convex capital adjustment costs can

solve the problem of instantaneous adjustment toward the steady state.7

Making the steady state level of net foreign assets unique involves one of

the following strategies. A first alternative is letting the interest rate faced

by the small open economy equal the exogenous world interest rate plus a

spread that is an increasing function of the country’s aggregate net foreign

asset position. The intuition behind this is a country-specific risk premium.

A second alternative is allowing for time varying (Uzawa-type) preferences,

which implies that the agent becomes more impatient as he becomes richer.

Allowing for effects from finite horizons is a third alternative. Within a

so-called Blanchard-Yaari-type (stochastic) overlapping generations frame-

work, a household’s life ends according to a Poisson process (which implies

that households are heterogenous with respect to date of birth and age).

5To see this, suppose there is a permanent positive productivity shock in a small

open economy with inelastic labor supply. Resources are shifted from the rest of the

world to the small open economy since agents in the world economy ”make hay where

the sun shines”, as Backus et al. [1] put it. This means that in the beginning the small

open economy is running a trade deficit to finance the additional investment. While the

resulting increase in consumption is negligible for the rest of the world, the small open

economy effect is large, resulting in a trade surplus in the subsequent periods. However,

there is no accompanying increase in foreign indebtedness to the small open economy.

In the words of Baxter [4], p. 1825, this is because ”the trade surplus may be viewed as

representing payments in fulfillment of a contingent-claims contract that specified these

payments in the event that the home country experienced an increase in productivity”.6As we will argue below, in a monetary setting there are other good reasons to deviate

from the complete asset market assumption.7Alternatively, constraints in international credit can be imposed (see Barro and Sala-

i-Martin [3].

Page 17: Macroeconomics of small open economies

1.2. GENESIS 7

However, households can buy actuarially fair annuities to hedge against un-

certainty with respect to the length of their life: As long as a household

lives, the insurance company pays a constant sum; it gets the remaining

assets when a household ends life (i.e., there is no bequest). In effect, this

leads to steady state values for consumption and foreign debt which are

positive and finite. Introducing convex portfolio adjustment costs is a final

alternative to overcome the stability problem. It implies that holding assets

in quantities which are different from a long-run level is costly.

1.2.3 New Neoclassical Synthesis

We begin this subsection with a brief overview of the empirical evidence

on the response of real and nominal variables to monetary policy impulses.

We then ask to what extent the neoclassical model is compatible with these

facts. Which finally leads us to the emergence of a new paradigm in macroe-

conomics, the New Neoclassical Synthesis (NNS).

In the last couple of years, a consensus from the empirical literature

on the short-run monetary relationships has developed. A first voluminous

body of empirical literature provides evidence on the short-run effects of

monetary policy shocks on real and nominal variables. Christiano et al. [11],

for instance, find for their benchmark model – which measures the policy

instrument by the federal funds rate – that a contractionary policy shock

leads to significant, persistent non-neutralities, to a rather slow response

of prices, and to a strong liquidity effect. These findings are robust across

alternative identification schemes; moreover, there is a long list of other

papers which find similar results. A second body of literature has produced

evidence in favor of a positively sloped short-run Phillips curve (basically the

correlation between inflation and unemployment or an alternative measure

of excess capacity in the economy).

Those pieces of evidence are hard to reconcile with the basic one-good,

one-shock RBC model. First, and most obviously, since the basic RBC

model does not refer to money, it has nothing to say about the behavior of

nominal variables. Second, the basic RBC model explains macroeconomic

fluctuations only with technology shocks; that is, demand shocks are absent

– which contrasts to the econometric evidence which identifies money as an

important source of fluctuations. A related, rather controversial property of

Page 18: Macroeconomics of small open economies

8 CHAPTER 1

the basic RBC model is that since economic fluctuations are efficient, there

is no scope for any form of government intervention, including monetary

stabilization policy.

The one-good, one-shock RBC model has subsequently been modified

to allow for money; this with the objective of learning whether monetary

forces can be an important cause of business cycle fluctuations in a neoclas-

sical setting. Cooley and Hansen [15], for instance, focus on theories of the

short-run non-neutrality of money which are in line with New Classical eco-

nomics. The first treats money as a source of confusion (the Lucas imperfect

information model); in the second, inflation acts as a distorting tax on the

holding of nominal money (the cash-in-advance approach). They find that

the quantitative importance of monetary shocks is very small and, accord-

ingly, that the two monetary models do not provide a good description of

the associations between real and nominal variables.

In the first half of the 1990s, more and more researchers began to intro-

duce New Keynesian features such as imperfect competition and sluggish or

costly nominal price adjustment into otherwise standard neoclassical busi-

ness cycle models, thereby considerably improving the models’ match to

data (see e.g. Yun [42]).8 A typical feature of NNS models is that they in-

volve substantial market failures, so that government intervention is desir-

able, and that increases in demand (e.g., due to an expansionary monetary

policy shock) stimulate aggregate activity.9

Compared to the traditional IS/LM-Phillips curve framework, the new

generation of models of the monetary transmission mechanism has a number

of advantages. First, agents are intertemporally optimizing and forward-

looking, do not in equilibrium want to change what they are doing, and

have rational expectations. Second, there is an appropriate emphasis on

real disturbances as a source of short-run variations in economic activity.

Finally, the new approach permits formal welfare evaluation of alternative

policies.

8The models’ ability to match the data crucially depends on a sufficiently large degree

of nominal price staggering or a similar mechanism. For a survey of the field in a closed-

economy context, compare e.g. Walsh [35], chapter 5.9It is probably worth to mention that with sluggish prices the effects of supply shocks

are altered, too. This is because not all firms are able to lower prices in response to an

unexpected increase in productivity and the corresponding decrease in marginal costs.

Page 19: Macroeconomics of small open economies

1.3. MODEL EVALUATION 9

The development in closed-economy macroeconomics has rapidly spilled

over into international macroeconomics. Among the first who introduced

money, imperfect competition, and price rigidities into an otherwise stan-

dard neoclassical open-economy framework were Obstfeld and Rogoff [27].

A drawback of this prototypical open-economy NNS model, however, is that

it is highly stylized and only provides qualitative results.

1.3 How well do quantitative, open-economy

NNS models match the data?

The empirical regularities in post-war data on the international business

cycle which have been used as criteria for assessing international RBC mod-

els focused on real quantities and relative prices (compare above). Open-

economy NNS models, however, should also be able to account for the fol-

lowing facts which involve the nominal exchange rate and the choice of the

exchange rate regime: (a) Nominal and measured real exchange rates are

highly volatile and very persistent; (b) nominal and real exchange rates

are strongly positively correlated; (c) the behavior of exchange rates (nom-

inal and real) varies systematically with the exchange rate regime; (d) the

behavior of the other real macroeconomic variables appears to be roughly

independent of the exchange rate regime.10

In the following, we discuss four examples of quantitative open-economy

NNS models and see how well they can account for fact (a) through (d). We

begin our study with Chari et al. [9], who aim at explaining fact (a). The

model of Chari et al. shares many features with the previously discussed

model of Backus et al. The most important differences are as follows. Since

Chari et al. consider a monetary economy, they assume a complete finan-

cial asset market rather than a complete contingent-claims market. The

intermediate goods markets are monopolistically rather than perfectly com-

petitive (i.e., competitive final good producers in each country purchase

intermediate goods from monopolistically competitive intermediate good

10Fact (b) and (c) have typically been interpreted as evidence in favor of sticky prices,

since they can only be explained within frictionless flexible-price models when most

significant shocks buffeting the economy are real – a rather unrealistic assumption, as we

have argued above (for an opposite view compare Stockman [36]).

Page 20: Macroeconomics of small open economies

10 CHAPTER 1

producers). And, intermediate good producers set nominal prices in stag-

gered contracts a la Taylor [37], rather than optimally adjusting them in

every period.

In principle, in an open-economy setting, price rigidities can take various

forms. Traditionally, prices have been assumed to be fixed in the currency

of the producer (i.e., the exporter’s country). Under this assumption the

law of one price (LOP) holds, taking for granted that trade is costless.

As a corollary, the purchasing power parity (PPP) is always satisfied and

exchange rate changes immediately feed into import prices (complete pass-

through).

However, a large body of empirical studies for many countries show that

the LOP does, more often than not, not hold and the PPP is only satisfied

in the long run. Both pieces of evidence suggest that the pass-through into

import prices might be limited. One popular story assumes that due to

high costs of arbitrage, home and foreign markets are segmented and each

individual monopolist can price-discriminate across countries; in addition

to this, producers set prices in the currency of the buyer (so-called local

currency pricing, LCP). Under this form of price stickiness, exchange rate

changes lead to proportional deviations from the LOP and the pass-through

is nil. This is exactly what Chari et al. assume.

To close the model, Chari et al. have to specify monetary policy. In

their benchmark version, they assume a (stochastic) money growth rule in

combination with a floating exchange rate.11 The first of the two countries

is calibrated to the U.S., the other to a European aggregate. Their findings,

which are based on simulations for shocks to the monetary aggregate, can be

summarized as follows. If prices are held fixed one period (i.e., one quarter),

the persistence of the real exchange rate is nil. For six period stickiness, the

persistence is considerable but still smaller than in the data. Only for an

implausibly large contract length of twelve periods they get a persistence

11Two key characteristics of modern international financial markets are the market

determination of exchange rates among the major industrialized countries and the global

financial integration. It therefore makes sense to assume that financial capital is interna-

tionally mobile. In accordance with the open-economy trilemma, the monetary authority

can then choose between an independent policy within a floating exchange rate or nom-

inal exchange rate stability at the expense of monetary independence.

Page 21: Macroeconomics of small open economies

1.3. MODEL EVALUATION 11

that roughly coincides with that found in the data.12

Duarte [17] is another example of a quantitative open-economy NNS

model; she aims at explaining fact (c) and (d). In Duarte’s model – as

opposed to Chari et al. – imports enter as final rather than as intermedi-

ate goods. There are two categories of goods, domestic and foreign, both

produced by monopolistically competitive firms in Home and Foreign, re-

spectively. Both categories of goods are aggregated to a distinct composite

consumption good; the two composites (or bundles) are then nested in a

two-input argument CES function which gives an overall consumption bas-

ket.

Further differences compared to Chari et al. are: Duarte abstracts from

capital and investment; there are two country-specific, cross-correlated pro-

ductivity shocks in addition to the two country-specific money growth shocks;

momentary utility is nonseparable between leisure and consumption; and

prices are not sluggish but set one period in advance.13

Another difference lies in the specification of the asset market. While

Chari et al. assume a complete financial asset market, Duarte assumes mar-

ket incompleteness. The assumption of asset market incompleteness is often

motivated pragmatically with a better performance regarding the dimension

consumption and output correlation. As we have seen in section 1.2, models

with complete risk sharing generate cross-country consumption correlations

that are much higher than in the data. Yet models with a bonds-only struc-

ture produce correlations which roughly coincide with those found in the

data (compare e.g. Kollmann [23]). A second argument in favor of a bonds-

only structure can be seen in the empirical evidence of non-fundamental

exchange rate fluctuations and strong deviations from the uncovered inter-

est parity (UIP). In the short to medium run, exchange rates are mostly

exogenous, that is, cannot be explained from macroeconomic fundamentals

(the Meese and Rogoff evidence). Moreover, there exists a large body of

literature that identifies (and searches to explain) deviations from the UIP.

A bonds-only structure in combination with the addition of an error term to

12Chari et al. also investigate variants of the benchmark model, in particular incomplete

financial asset markets, sticky wages, and an active rather than a passive monetary policy.

The findings of the benchmark model are largely confirmed.13As Chari et al., Duarte assumes 100% LCP.

Page 22: Macroeconomics of small open economies

12 CHAPTER 1

the UIP relation allows to account for these facts. In Duarte [17], however,

the UIP always holds.

Duarte assumes that the two countries are perfectly symmetric and cali-

brates them to U.S. data. Her model successfully generates a sharp increase

in the volatility of the real exchange rate following a switch from fixed to

flexible rates, without a similar pattern for the volatilities of output, con-

sumption, or trade flows. As we will discuss below, the LCP assumption is

crucial in generating this pattern (since it implies – in the words of Duarte

– that changes in the nominal exchange rate are dissociated from allocation

decisions).

Yet another example of a quantitative open-economy NNS model is

Monacelli [25], who aims at explaining fact (c). In Monacelli, as in Duarte,

consumption goods are directly traded, asset markets are incomplete, and

the UIP always holds. The most important differences compared to Duarte

are as follows. Monacelli considers a small country, whereby he ignores

the stability problem discussed in Subsection 1.2.2; no monetary assets are

needed to facilitate transactions (i.e., a cashless economy is considered, see

e.g. Woodford [41]); utility is separable between leisure and consumption;

prices are staggered a la Calvo [8] and fixed in the currency of the producer

(100% producer currency pricing, PCP); moreover, the model incorporates

capital and capital adjustment costs.

Finally, Monacelli assumes that monetary policy follows an interest-rate

feedback rule (which includes a term that reacts to movements of the nom-

inal exchange rate about the parity). The model is parameterized for a

hypothetical economy (i.e., the parameter values are borrowed from the

literature). Monacelli finds that the model is consistent with Mussa’s evi-

dence: The real exchange rate is between four to five times more variable

under floating than under fixed rates, independent of the underlying source

of fluctuations.

As mentioned above, another classical finding in international macroe-

conomics is that traditional exchange rate models are unable to beat a

random walk in forecasting the nominal exchange rate. Bergin [5] asks

whether open-economy NNS models can better explain the exchange rate.

His two-country model shares many features with the benchmark version of

Chari et al. The most important differences are as follows. Staggered price

Page 23: Macroeconomics of small open economies

1.4. LABORATORY FOR POLICY ANALYSIS 13

adjustment a la Calvo is assumed. Both types of price stickiness, PCP and

LCP, are allowed to coexist where the share is a parameter to be estimated.

The model has a bonds-only structure. To the UIP, a country risk premium

to the net foreign asset position plus an error term (i.e., a shock to the risk

premium) is added. Monetary policy is assumed to be active: it follows a

contemporaneous-date Taylor rule with interest rate smoothing.

A few structural parameters are pinned down using information from

prior studies. But most parameters (among others the elasticity of substi-

tution between home and foreign composite goods and the constant fraction

of firm exhibiting LCP) are estimated by maximum likelihood methods.

Quarterly data for the U.S. and an aggregate of the remaining G7 countries

are used to this aim. Bergin finds that the model performs moderately well

in that it is able to beat a random walk model for in-sample predictions.

To sum up, there is no unified modelling framework which can account

for all discussed facts. However, specific specifications match specific key

moments in the data quite well. Price staggering a la Calvo seems to be

more capable than price staggering a la Taylor. The appropriate choice of

the pricing pattern (PCP or LCP) remains an unsettled issue.

1.4 Laboratory for policy analysis

So far, we have been asking the (positive) question of how successfully

quantitative open-economy NNS models can replicate statistical properties

observed in the data. We now turn to normative aspects in international

macroeconomics such as the question whether monetary authorities should

take account of the (nominal or real) exchange rate in the conduct of mon-

etary policy and if yes, to what extent. An extreme form of taking account

of the exchange rate is stabilizing it towards a foreign currency of choice.

A step beyond fixed exchange rates is for two countries to share a common

currency. Thus, a related question asks about the ranking of alternative

international monetary arrangements: a free float, an unilateral peg, and a

currency union.

The fact that the world is interdependent with government policy pos-

sibly being a major source of the transmission of economic disturbances

across countries leads to yet another topic, namely the question wether

Page 24: Macroeconomics of small open economies

14 CHAPTER 1

there are potential benefits from international macroeconomic policy coor-

dination. The work on strategic interaction between central banks utilizing

two-country NNS models has recently been surveyed by Bowman and Doyle

[7]. Here, we concentrate on the quantitative open-economy literature deal-

ing with the role the exchange rate should play in the conduct of monetary

policy and the choice of the exchange rate regime.

In the literature concerned with the design of monetary policy, two broad

branches can be distinguished. One strand aims at deriving the globally

optimal policy (under either commitment or discretion). This is relatively

straightforward when a number of simplifying assumptions is imposed re-

garding the size of the economy, the asset market structure, the capital

accumulation, and the number and kind of distortions present in the econ-

omy (compare e.g. Clarida et al. [10] and Gali and Monecelli [15], section

5). In more general settings, however, the derivation of the globally optimal

policy is more involved and often intractable. In this event, an alternative

is to restrict the instrument rule (typically the nominal interest rate) to lie

in a given class and at the same time to assume that the central bank can

commit, once and for all, to a given policy rule for all future periods.

Hence, in rather general open-economy frameworks the question about

how responsive monetary policy should be to the exchange rate comes down

to the question whether amending or altering the original Taylor rule im-

proves economic performance and welfare.14 The followed strategy for pol-

icy evaluation is concisely described in Taylor [38], pp. 263/4: One places a

particular modification of the original rule (allowing for a direct response of

the interest rate to the exchange rate) into a parameterized stochastic DGE

model of a small open economy with sticky prices, solves the model using

a numerical solution algorithm, examines the properties of the stochastic

behavior of those macroeconomic variables which reflect potential goals of

monetary policy, and/or examines the consequences on welfare. One proviso

is in order here. It concerns the abstraction from the possibility of specu-

lative attacks when the nominal exchange rate is pegged, which is actually

an important form of cost of fixing the exchange rate.

14The original Taylor rule makes the instrument depend on current domestic inflation

and output deviations from trend with reaction coefficients 1.5 and 0.5, respectively.

Page 25: Macroeconomics of small open economies

1.4. LABORATORY FOR POLICY ANALYSIS 15

Two-country models We now come to the synopsis of some recent find-

ings. We start our study with Collard and Dellas [13] who ask how changes

in international monetary arrangements affect the properties of the busi-

ness cycle in individual countries as well as globally. Three regimes are

considered: a perfect float, an unilateral peg, and a bilateral peg. Collard

and Dellas’ [13] two-country model shares many features with the bench-

mark version of Chari et al. The most important differences are as follows.

First, Collard and Dellas [13] assume staggered nominal wage contracts a

la Calvo; furthermore, there are five shocks in the model in addition to the

two country-specific monetary shocks, namely a common and two country-

specific supply shocks and two country-specific fiscal shocks; finally, mone-

tary policy is active and follows a forward-looking Taylor rule.

The model is calibrated to Germany and France. Collard and Dellas

[13] find that in France, macroeconomic volatility under a monetary union

is comparable to that under a flexible exchange rate system but consider-

ably lower than that under an unilateral peg (where the French franc is

fixed towards the Deutsch mark). In Germany, output volatility is signifi-

cantly higher under a peg – relative to the flexible regime – and increases

even further under a currency union, whereas inflation volatility becomes

smaller. The monetary union also induces a strong negative international

transmission of country specific supply shocks.

Milton Friedman’s classical conjecture that a floating exchange rate pro-

vides the needed relative price adjustment when nominal goods prices are

sluggish, is based on the complete pass-through assumption and the notion

that monetary policy is passive. In another piece of work, Collard and Del-

las [12] investigate how much rigidity is needed in order to make a difference

for the choice of exchange rate regime and whether Friedman’s case for a

flexible regime is consistent with an activistic policy. The model they utilize

is similar to the one in Collard and Dellas [13]. The most important dif-

ferences are as follows. Collard and Dellas [12] assume that nominal goods

prices rather than wages are sluggish; prices are set in the producer’s cur-

rency (100% PCP); and, monetary policy follows a contemporaneous-date

Taylor rule with interest rate smoothing.

The model is calibrated to the postwar U.S. economy, under the assump-

tion of perfect symmetry across countries. Welfare is computed based on a

Page 26: Macroeconomics of small open economies

16 CHAPTER 1

quadratic approximation to the utility function of the representative agent

(together with a first-order approximation to the equilibrium conditions).

Collard and Dellas [12] indeed find that a high degree of sluggishness tends

to favor the flexible system, while a low degree of sluggishness favors the

fixed regime. However, the differences across exchange rate regimes in terms

of performance and welfare tend to be small.

A number of recent studies indicate that the amendment of a direct

exchange rate target (nominal or real) into an otherwise standard Taylor

rule does either not yield a greater improvement in performance or that

performance actually deteriorates (see e.g. Taylor [38]). In yet another piece

of work, Collard and Dellas [14] argue that this finding presents a challenge

for understanding actual monetary policy practices. To gain more insight,

they investigate the implication of direct exchange rate targeting within a

modelling framework which is identical to that of Collard and Dellas [12].

They find that real exchange rate targeting is indeed never a good idea

while nominal exchange rate targeting is likely to be irrelevant.

Small open economy models In a rather influential paper, Gali and

Monacelli [15] analyze within a small open economy setting the macroeco-

nomic implications of the three alternative policy regimes strict domestic

inflation targeting, strict consumer price index (CPI) inflation targeting,

and an unilateral peg. As in Duarte, consumption goods are directly traded.

The most important differences compared to Duarte are as follows. In Gali

and Monacelli, the home economy is small (while the rest of the world is

treated as a closed economy following an optimal policy); the related sta-

bility problem is overcome by assuming complete financial asset markets;

no monetary assets are needed to facilitate transactions (cashless economy);

utility of the representative household is assumed to be separable between

leisure and consumption; finally, staggered price adjustment a la Calvo and

100% PCP is assumed.

Gali and Monacelli report quantitative results for a parameterized ver-

sion of their model; parameter values are borrowed from the literature. They

find that a peg amplifies both output gap and inflation volatility, relative to

a strict domestic inflation targeting, with the strict CPI inflation targeting

regime lying somewhere in between.

Page 27: Macroeconomics of small open economies

1.4. LABORATORY FOR POLICY ANALYSIS 17

Ghironi [16] compares the performance of alternative monetary policy

rules for the (approximately) small open economy of Canada. The model

shares key features with Monacelli [25]. The most important differences

are as follows. Rather than ignoring the stability problem, Ghironi deals

with it by assuming Blanchard-Yaari-type overlapping generations; money

is incorporated; and, the household’s momentary utility is nonseparable be-

tween leisure and consumption. Ghironi estimates an extended Taylor-type

reaction function which is backward looking and comprises an exchange rate

target, a CPI inflation target (in addition to a domestic inflation target),

and interest rate smoothing.

The structural (or non-policy) parameters are estimated using quarterly

data from Canada and the U.S. Despite the simple structure of the exoge-

nous processes, the model matches several key moments in the data quite

well. According to the simulation results, the benchmark policy (the es-

timated extended Taylor-type rule) dominates the considered alternatives

(strict CPI targeting, an unilateral peg, and the original Taylor rule) in

terms of welfare, whereby risk diversification (the covariance between con-

sumption and leisure) is playing a crucial role.

A final example of a quantitative NNS model used for policy analysis is

Kollmann [24], who optimizes reaction coefficients for different variants of

Taylor-type interest rules. His model shares many features with Chari et al.

The most important differences are as follows. Kollmann considers a small

country within a bonds-only structure; he overcomes the stability problem

by adding to the UIP a risk premium to the net foreign asset position;

he assumes staggered price adjustment a la Calvo; and, he ignores direct

services provided by money.

In the benchmark case, in which the policy instrument only depends on

domestic inflation and the output gap, the optimized policy rule has infla-

tion and output coefficients of 3.01 and -0.01, respectively. Kollmann also

experiments with extended Taylor rules. Independent whether the nominal

exchange rate in growth rates or in levels is appended to the benchmark

rule, the optimized reaction coefficient is close to zero and the welfare gains

are minuscule. CPI targeting yields essentially the same welfare as domes-

tic inflation targeting. And, a peg greatly raises the variability of both

consumption and output.

Page 28: Macroeconomics of small open economies

18 CHAPTER 1

1.5 Concluding remarks

This paper selectively surveyed the strand of literature within open-economy

macroeconomics which applies quantitative business cycle models to the

issue of transmission properties of economic disturbances and international

policy. The discussion revealed that there is one major point of disagreement

over the appropriate specification of open-economy models: it concerns the

nature of price stickiness.

The assumption of LCP has been motivated by a number of well doc-

umented facts in the data: pervasive deviations from the LOP, persistent

deviations from the PPP, and a close to zero correlation between nominal

exchange-rate changes and inflation (at the business cycle frequency). As

indicated above, one interpretation of these facts is that exporters price-

discriminate across markets and in addition post prices in the buyer’s cur-

rency. Direct evidence on the choice of invoice currencies in international

trade, however, seems to contradict this interpretation.15 Moreover, there

are other reasons thinkable why consumer prices do not respond much to

exchange rates. There might be transportation or distribution costs; im-

ports may incorporate a substantial nontradable marketing input and/or

may be distributed through an imperfectly competitive retailing network;

finally, there might be true pricing to market.16 Each one of these reasons

might weaken the link between consumer and original price without assum-

ing LCP. We conclude that until additional (microeconometric) studies are

available we should be cautious in interpreting the existing evidence in favor

of one or the other policy regime (see Engel [18]).

To conclude, we would like to point to some recent developments in

closed-economy macroeconomics that might in the future spill over into in-

ternational finance. Most models discussed in this study assume that the

central bank can perfectly observe and react to current shocks. In reality,

however, disturbances (and possibly also steady state values) are not di-

rectly observable and policy makers as well as the private sector have to

solve complicated signal-extraction problems instead. In particular, central

15For an overview, see Engel [18] and Obstfeld [30].16True pricing to market means optimal price discrimination in the sense of monopo-

listic firms intentionally setting different prices in different markets because of different

market conditions (see Bergin and Feenstra [6]).

Page 29: Macroeconomics of small open economies

1.5. CONCLUDING REMARKS 19

bankers have only incomplete information regarding current values of the

output gap and the natural rate of interest. Some authors include devia-

tions from trend, lagged values of the gap or previous period’s conditional

expectations of the current gap, in lieu of the current gap itself. Yet the

lack of a learning process on the part of the monetary authority makes

these short-cuts unsatisfactory. A related issue is uncertainty about the

true structure of the economy: The decision makers at central banks have

only vague notions about the workings of the economy. It would be inter-

esting to explore both types of practical complications in an open-economy

setting and examine their implications for monetary policy.

Literally all models discussed in this study assume that the nominal

price or wage setting mechanism is time dependent. The reason for the per-

vasiveness of staggered price adjustment a la Calvo in particular, is a very

pragmatic one, namely tractability. However, the lack of a formal optimiza-

tion underpinning is a potential disadvantage. Recently, some research has

been devoted to price settings in which the nominal price rigidity is derived

as an endogenous result from microeconomic optimality conditions. Bakhshi

et al. [2], for instance, derive a closed-form solution for short-term inflation

in the NNS modelling framework of Dotsey et al. [16].17 The resulting

state-dependent New Phillips curve relates inflation not only to expected

future inflation and current and expected future real marginal costs (as the

New Phillips curve in standard Calvo models does) but – most interestingly

– also to lagged inflation, with fast decreasing weights. The number of leads

and the size of the coefficients are endogenous and depend on the level of

steady-state inflation and on firms’ beliefs about future adjustment costs.

Again, it would be interesting to investigate this alternative pricing scheme

in an open-economy context and examine implications for monetary policy.

17Dotsey et al. [16] assume that firms face stochastic menu costs which are i.i.d. across

firms and across time.

Page 30: Macroeconomics of small open economies

Bibliography

[1] Backus, David K., Patrick J. Kehoe, and Finn E. Kydland (1995),

International business cycles: theory and evidence, in: T. F. Cooley,

ed., Frontiers of Business Cycle Research, Princeton University Press

[2] Bakhshi, Hasan, Hashmat Khan, and Barbara Rudolf (2004), The

Phillips curve under state-dependent pricing, Working Paper, Swiss

National Bank

[3] Barro, Robert J., and Xavier Sala-i-Martin (1995), Economic Growth,

McGraw-Hill

[4] Baxter, Marianne (1995), International trade and business cycles, in:

G. Grossman and K. Rogoff, eds., Handbook of International Eco-

nomics, Vol. 3, Elsevier Science

[5] Bergin, Paul R. (2004), How well can the new open economy macroeco-

nomics explain the exchange rate and current account?, NBER Work-

ing Paper No. 10356

[6] Bergin, Paul R., and Robert C. Feenstra (2001), Pricing-to-market,

staggered contracts, and real exchange rate persistence, Journal of In-

ternational Economics 54(2), 333-359

[7] Bowman, David, and Brian Doyle (2003), New Keynesian, open-

economy models and their implications for monetary policy, Board of

Governors of the Federal Reserve System, International Finance Dis-

cussion Papers No. 762

[8] Calvo, Guillermo A. (1983), Staggered prices in a utility-maximizing

framework, Journal of Monetary Economics 12(3), 383-398

Page 31: Macroeconomics of small open economies

BIBLIOGRAPHY 21

[9] Chari, V.V., Patrick J. Kehoe, and Ellen R. McGrattan (2000), Can

sticky price models generate volatile and persistent real exchange

rates?, Federal Reserve Bank of Minneapolis, Research Department

Staff Report 277

[10] Clarida, Richard, Jordi Gali, and Mark Gertler (2001), Optimal mon-

etary policy in open versus closed economies: an integral approach,

American Economic Review, AEA Papers and Proceedings 91(2), 248-

252

[11] Christiano, Lawrence J., Martin Eichenbaum, and Charles L. Evans

(1999), Monetary policy shocks: what have we learned and to what

end?, in: J. B. Taylor and M. Woodford, eds., Handbook of Macroeco-

nomics, Vol. 1A, Elsevier Science

[12] Collard, Fabrice, and Harris Dellas (2001), Price rigidity and the selec-

tion of the exchange rate regime, Mimeo

[13] Collard, Fabrice, and Harris Dellas (2002), Exchange rate systems and

macroeconomic stability, Journal of Monetary Economics 49(3), 571-

599

[14] Collard, Fabrice, and Harris Dellas (2002), Monetary policy rules and

exchange rate targeting, Mimeo

[15] Cooley, Thomas F., and Gary D. Hansen, (1995), Money and the busi-

ness cycle, in: T. Colley (ed.), Frontiers of business cycle research,

Princeton University Press

[16] Dotsey, Michael, Robert G. King, and Alexander L. Wolman (1999),

State-dependent pricing and the general equilibrium dynamics of

money and output, The Quarterly Journal of Economics 114(2), 655-

690

[17] Duarte, Margarida (2003), Why don’t macroeconomic quantities re-

spond to exchange rate volatility?, Journal of Monetary Economics

50(4), 889-913

Page 32: Macroeconomics of small open economies

22 CHAPTER 1

[18] Engel, Charles (2002), The responsiveness of consumer prices to ex-

change rates and the implications for exchange-rate policy: a survey of

a few recent new open-economy macro models, NBER Working Paper

No. 8725

[19] Gali, Jordi, and Tommaso Monacelli (2002), Monetary policy and ex-

change rate volatility in a small open economy, NBER Working Paper

No. 8905

[20] Ghironi, Fabio (2000), Alternative monetary rules for a small open

economy: the case of Canada, Working Paper, Boston College

[21] Ghironi, Fabio (2000), Towards new open economy macroeconomics,

Working Paper, Boston College

[22] King, Robert G., and Sergio T. Rebelo (1999), Rescuscitating real

business cycles, in: J. B. Taylor and M. Woodford, eds., Handbook

of Macroeconomics Vol. 1, Elsevier Science

[23] Kollmann, Robert (1996), Incomplete asset markets and the cross-

country consumption correlation puzzle, Journal of Economic Dynam-

ics and Control 20(5), 945-961

[24] Kollmann, Robert (1997), The exchange rate in a dynamic-optimizing

current account model with nominal rigidities: a quantitative investi-

gation, IMF Working Paper WP/97/7

[25] Kollmann, Robert (2002), Monetary policy rules in the open economy:

effects on welfare and busincess cycles, Journal of Monetary Economics

49(5), 989-1015

[26] Lane, Philip R. (2001), The new open economomy macroeconomics: a

survey, Journal of International Economics 54(2), 235-266

[27] McCallum, Bennett T., and Edward Nelson (2004), Timeless perspec-

tive vs. discretionary policy in forward-looking models, Federal Reserve

Bank of St. Louis Review 86(2), 43-56

[28] Mendoza, Enrique G. (1991), Real business cycles in a small open econ-

omy, American Economic Review 81(4), 797-818

Page 33: Macroeconomics of small open economies

BIBLIOGRAPHY 23

[29] Monacelli, Tommaso (2000), Into the Mussa puzzle: monetary policy

regimes and real exchange rate in a small open economy, Working Pa-

per, Boston College

[30] Obstfeld, Maurice (2002), Exchange rates and adjustment: perspec-

tives from the new open economy macroeconomics, CEPR Discussion

Paper Series No. 3533

[31] Obstfeld, Maurice, and Kenneth Rogoff (1995), Exchange rate dynam-

ics redux, Journal of Political Economy 103(3), 624-660

[32] Obstfeld, Maurice, and Kenneth Rogoff (1995), The mirage of fixed

exchange rates, Journal of Economic Perspectives 9(4), 73-96

[33] Obstfeld, Maurice, and Kenneth Rogoff (1996), Foundations of inter-

national macroeconomics, The MIT Press

[34] Sarno, Lucio (2001), Toward a new paradigm in open economy model-

ing: where do we stand?, Federal Reserve Bank of St. Louis

[35] Schmitt-Grohe, Stephanie, and Martin Uribe (2002), Closing small

open economy models, CEPR Discussion Paper No. 3096

[36] Stockman, Alan C. (1987), The equilibrium approach to ex-

change rates, Federal Reserve Bank of Richmond, Economic Review

March/April

[37] Taylor, John B. (1980), Aggregate dynamics and staggered contracts,

Journal of Political Economy 88(1), 1-23

[38] Taylor, John B. (2001), The role of the exchange-rate in monetary-

policy rules, American Economic Review, AEA Papers and Proceedings

91(2), 263-267

[39] Walsh, Carl E. (2003), Monetary theory and policy, second edition,

The MIT Press

[40] Woodford, Michael (forthcoming), Revolution and evolution in

twentieth-century macroeconomics, in: P. Gifford, ed., Frontiers of the

mind in the twenty-first century, Harvard University Press

Page 34: Macroeconomics of small open economies

24 CHAPTER 1

[41] Woodford, Michael (2003), Interest and prices: foundations of a theory

of monetary policy, Princeton University Press

[42] Yun, Tack (1996), Nominal price rigidity, money supply endogeneity,

and busincess cycles, Journal of Monetary Economics 37(2-3), 345-370

Page 35: Macroeconomics of small open economies

Chapter 2

On the role of exchange rate

stabilization in the conduct of

monetary policy of a small

open economy

2.1 Introduction

In the case of a small open economy (with free financial capital mobility)

whose policy is oriented towards domestic goals and for which shocks from

the rest of the world are important: What role should the stabilization of

the exchange rate play in the conduct of monetary policy? In the new open

economy macroeconomics literature concerned with the design of monetary

policy, two broad branches can be distinguished. One strand aims at de-

riving the globally optimal policy (under either commitment or discretion).

This is relatively straightforward when a number of simplifying assump-

tions are imposed regarding the size of the economy, the specification of the

preferences, the capital accumulation, and the number and kind of distor-

tions present in the economy (compare e.g. Clarida et al. [10] and Gali and

Monecelli [15], section 5).

In more general settings, however, the derivation of the globally optimal

policy is more involved and often intractable. In this event, an alternative

is to restrict the instrument rule (typically the nominal interest rate) to lie

Page 36: Macroeconomics of small open economies

26 CHAPTER 2

in a given class and at the same time to assume that the central bank can

commit, once and for all, to a given policy rule for all future periods. In

the rather general framework of quantitative open economy New Neoclassi-

cal Synthesis (NNS) models, the question about how responsive monetary

policy should be to the exchange rate, thus, comes down to the question

whether amending or altering the original Taylor rule improves economic

performance and welfare.1 A number of studies have recently appeared on

this subject (compare e.g. Ghironi [16], Monacelli [25], Collard and Dellas

[11], Gali and Monacelli [15], and Kollmann [22]).2 The typical result is

either that directly reacting to the exchange rate does not yield a greater

improvement in performance or welfare or that performance actually dete-

riorates (Taylor [34], p. 266). In Kollmann [22], for instance, the optimized

reaction coefficient is close to zero; the involved welfare gains are minuscule.

This finding, however, contrasts to actual monetary policy practices.

Casual observation reveals that central banks of small and semi-small open

economies often follow the rule of thumb – expressed in Obstfeld and Rogoff

[27], p. 93 – saying that a substantial appreciation of the real exchange rate

accompanied by slow output growth and low inflation furnishes a case for

cutting interest rates. To gain potential insights into the underpinnings

of actual policy practices, we add a terms-of-trade target to Taylor-type

interest rules as they have been proposed in the literature and evaluate them

in terms of their implications on the business cycles and welfare within a

variant of Kollmann’s [22] model of a small open economy.3

1The original Taylor [32] rule makes the instrument depend on current domestic infla-

tion and output deviations from trend with reaction coefficients 1.5 and 0.5, respectively.2Monacelli [25] presents qualitative results only, while the other studies conduct per-

formance and/or welfare analyzes as well; Collard and Dellas [11] investigate a two-

country model, while the other studies consider small open economies. – Two other

frequently cited examples which do not belong in the class of NNS models (in that they

are not fully microfounded but directly specify aggregate supply and demand curves) are

Ball [2] and Svensson [31]. Ball derives the optimal instrument rule assuming an ad hoc

loss function within a backward-looking framework; Svensson explores optimal reaction

functions for different variants of the loss function and two versions of the Taylor rule

within a forward-looking framework.3Collard and Dellas [11] do a similar exercise in a two-country setting. They find that

real exchange rate targeting is never a good idea while nominal exchange rate targeting

is likely to be irrelevant.

Page 37: Macroeconomics of small open economies

2.1. INTRODUCTION 27

We limit the consideration to operational feedback rules, that is, to pol-

icy rules which require only information which could plausibly be possessed

by the central bank. The results are discussed against the background of

two natural benchmark policies, namely strict (domestic) inflation target-

ing and a credible and unilateral peg of the domestic currency toward the

currency of the rest of the world. We find that introducing a moderate form

of terms-of-trade targeting in the original Taylor rule induces higher wel-

fare with respect to shocks to productivity and foreign demand and lower

welfare with respect to shocks to government consumption and the terms of

trade. The outcome under rules that allow for a considerable degree of in-

terest smoothing and a high relative weight on inflation, however, is robust

regarding the inclusion of terms-of-trade targeting.

In the remainder of this section, we briefly describe Kollmann’s [22]

original setting and then discuss our modifications. In Kollmann’s model,

the small open economy specializes in a continuum of tradable intermediate

goods and imports a continuum of foreign intermediate goods. Each differ-

entiated good, whether domestic or foreign, is produced by a monopolist.

Domestic intermediate goods producers use domestic (physical) capital and

labor as inputs; both inputs are immobile internationally.

The small open economy also produces a non-tradable final good, which

is used for consumption, investment, and government spending. The fi-

nal good is produced by assembling imperfectly substitutable domestic and

imported intermediate goods. While nominal prices in the perfectly com-

petitive final good and input markets are fully flexible, intermediate goods

producers set their prices for a stochastic number of periods as in Calvo

[5]. The representative household owns the firms of both sectors as well as

the capital stock and the time endowment; the capital stock and the time

devoted to labor are rented to the domestic intermediate goods producers.

Other distinguishing features of Kollmann’s model are: International

risk sharing is imperfect due to incomplete international financial markets;

to the uncovered interest parity, a country risk premium to the net foreign

asset position is added; the exchange rate pass-through is limited (100%

local-currency pricing); and, direct services provided by money are ignored

(a cashless economy). The model is calibrated to quarterly data for Japan,

Germany, and the U.K.

Page 38: Macroeconomics of small open economies
Page 39: Macroeconomics of small open economies

2.2. THE MODEL 29

a stationary representation of the system of equilibrium conditions. Ap-

pendix 2.B derives the non-stochastic steady state. Appendix 2.C presents

a first-order approximation to the equilibrium conditions, given that the

monetary policy obeys a Taylor-type feedback rule. Appendix 2.D describes

the implementation of an exchange rate peg.

2.2 The Model

2.2.1 Behavior of final good producers

There is a representative, competitive firm which produces a non-tradable,

homogenous consumption/investment good. To produce this good, the firm

has to purchase intermediate goods from domestic and foreign intermediate

goods producers. The country-specific, final good is assembled according to

a CRTS production technology

Y(st)

= F[Xd(st), Xf

(st)]

where st is one out of finitely many states the economy experiences in period

t (i.e., all variables in the model follow a discrete state stochastic process)

and the two input factors are CES-aggregators given by

Xd(st)

=

[∫ 1

0

Xd(i, st)θ

di

] 1

θ

Xf(st)

=

[∫ 1

0

Xf(i, st)θ

di

] 1

θ

where i stands for a differentiated intermediate good and θ ∈ (0, 1) is the

lower-level substitution parameter. In what follows, F [•] is specified to

F[Xd(st), Xf

(st)]

=[ω1−ρXd

(st)ρ

+ (1 − ω)1−ρ Xf(st)ρ]1/ρ

where ρ ∈ (−∞, 1) is the upper-level substitution parameter and ω ∈ (0, 1)

is a distribution parameter (in fact, one minus the import share).

The final good producer simultaneously solves two problems. First, the

firm minimizes costs of producing a given level of Xd (st); this yields condi-

tional factor demand functions Xd (i, st) for all i ∈ [0, 1]. Similarly, the firm

minimizes costs of producing a given level of Xf (st). Second, the firm mini-

mizes costs of producing a given level of Y (st); this yields conditional factor

demand functions Xd (st) and Xf (st). These are both static problems; to

simplify notation, we therefore skip the state-in-time-t label.

Page 40: Macroeconomics of small open economies

30 CHAPTER 2

Conditional factor demand functions Xd (i) and Xf (i) The problem

of minimizing costs of producing a given level of Xd can be stated as

minXd(i)

∫ 1

0

Px (i) Xd (i) di

subject to

Xd =

[∫ 1

0

Xd (i)θ di

] 1

θ

where Px (i) is the price of the domestic intermediate good i denoted in

domestic currency. From the FOC to this problem, we can derive the final

good producer’s conditional factor demand function

Xd (i) =

[Px (i)

Px

] 1

θ−1

Xd for all i ∈ [0, 1]

and an expression for the price index of the domestic intermediate goods,

the producer price index (PPI) in brief,

Px =

[∫ 1

0

Px (i)θ

θ−1 di

] θ−1

θ

.

There is an analogous problem for the foreign intermediate good i, from

which we get

Xf (i) =

[P ∗

x (i)

P ∗x

] 1

θ−1

Xf for all i ∈ [0, 1]

with

P ∗

x =

[∫ 1

0

P ∗

x (i)θ

θ−1 di

] θ−1

θ

.

Note that P ∗

x (i) (and, thus, the intermediate goods price index, P ∗

x ) is

denoted in terms of the currency of the seller, i.e., in foreign currency.

Conditional factor demand functions Xd and Xf The problem of

minimizing the costs of producing a given level of Y can be stated as

minXd, Xf

PxXd + eP ∗

xXf

Page 41: Macroeconomics of small open economies

2.2. THE MODEL 31

subject to

Y =[ω1−ρ

(Xd)ρ

+ (1 − ω)1−ρ (Xf)ρ]1/ρ

where e is the price of foreign money in units of domestic money (i.e., the

nominal exchange rate). From the FOCs to this problem, the following two

conditional factor demand functions can be derived

Xd = ω

(Px

P

) 1

ρ−1

Y Xf = (1 − ω)

(eP ∗

x

P

) 1

ρ−1

Y.

Assuming that the representative final good producer efficiently produces

one unit of Y provides us with an expression for the CPI as a function of

the domestic and foreign PPI

P =[ωP

ρρ−1

x + (1 − ω) (eP ∗

x )ρ

ρ−1

] ρ−1

ρ

Recall that all conditional factor demand functions derived in this subsec-

tion need to be satisfied in any state of the world.

2.2.2 Digression: Total demand for input factor i

To derive a function for total demand for input i, we start by noting that

intermediate good i is produced for the domestic final good sector and for

the export market

X(i, st)

= Xd(i, st)

+ Xd∗(i, st)

where Xd∗ (i, st) denotes the quantity of intermediate good i used in the rest

of the world. In step one of the domestic final good producer’s optimization

problem, we have derived a function for the domestic demand for good i,

Xd (i, st); we would like to come up with a similar expression for the export

demand for good i, Xd∗ (i, st).

Following Kollmann [22], we assume that the export demand function

resembles the domestic demand function

Xd∗(i, st)

=

[Px (i, st)

Px (st)

] 1

θ−1

Xd∗(st)

for all i ∈ [0, 1]

Page 42: Macroeconomics of small open economies

32 CHAPTER 2

where Xd∗ (st) is exogenous.5 We end up with the following total conditional

factor demand function

X(i, st)

=

[Px (i, st)

Px (st)

] 1

θ−1 [Xd(st)

+ Xd∗(st)]

︸ ︷︷ ︸X(st)

. (2.1)

Note that the elasticity of total demand for input factor i with respect to

input price Px (i, st), defined as

ε ≡ −∂X (i, st)

∂Px (i, st)

Px (i, st)

X (i, st),

is given by1

1 − θ.

In words: If the lower-level substitution parameter, θ, is close to 1, the price

elasticity of demand is large (in the limit +∞); if θ is much smaller than 1,

the price elasticity of demand is small (in the limit 0).

2.2.3 Behavior of intermediate good producer i

Each intermediate goods producer supplies a differentiated intermediate

good i and demands inputs in a competitive fashion. Thus, firm i has to

make two simultaneous decisions: how much capital and labor to lease in

each period (thereby acting as a price taker) and what output price to charge

for the differentiated intermediate good (thereby acting as a monopolist).

Input demand

Firm i combines capital K (i, st) and labor h (i, st) to produce intermediate

good i according to a CRTS production technology

X(i, st)

= F[K(i, st), h(i, st), A(st), Γt

]

where A (st) is an exogenous stationary stochastic technological shock and

Γt is deterministic technical progress. F [•], A (st), and Γt are identical for

5Recall that in contrast to Kollmann [22] we assume producer-currency pricing. This

explains the difference between his and our specification.

Page 43: Macroeconomics of small open economies

2.2. THE MODEL 33

all i. In what follows, F [•] is specified to as Cobb-Douglas

F[K(i, st), h(i, st), A(st), Γt

]= A

(st)K(i, st)α [

Γth(i, st)]1−α

(2.2)

where α ∈ [0, 1] is a positive constant.

In order to decide how much capital and labor to lease in each period,

intermediate good producer i solves the static cost minimizing problem

(again, we skip the state-in-time-t label)

minK(i),h(i)

PzK (i) + PWh (i)

subject to

X (i) = AK (i)α [Γh (i)]1−α

where z is the real rental rate and W is real wage. From the FOC to this

problem, we can derive the conditional demand functions for capital

K [X (i) ,W, z, ...] =

1 − α

W

z

)1−α

A−1Γα−1X (i)

and for labor

h [X (i) ,W, z, ...] =

1 − α

W

z

)−α

A−1Γα−1X (i) .

The value function to the cost minimizing problem is given by (in real

terms)

C [X (i) ,W, z, ...] = A−1χ−1zαW 1−αΓα−1X (i)

where χ−1 ≡ α−α (1 − α)α−1. Note that in the presence of a CRTS Cobb-

Douglas technology, (real) average costs, Ca (i) ≡ C (i) /X (i), equal (real)

marginal costs, Cm (i) ≡ ∂C (i) /∂X (i), i.e.,

Ca (i) = Cm (i) .

Moreover, note that Cm (i) and Ca (i) are the same for all i

Cm (i) = Cm = Ca (i) = Ca = A−1χ−1zαW 1−αΓα−1. (2.3)

Page 44: Macroeconomics of small open economies

34 CHAPTER 2

To get more economic insight – and for later reference – we derive the

inverse of the two conditional factor demands, thereby substituting Cm for

A−1χ−1zαW 1−αΓα−1. We end up with

z(st)

= Cm

(st)α

X (st)

K (st)︸ ︷︷ ︸MPK(st)

(2.4)

and

W(st)

= Cm

(st)(1 − α)

X (st)

h (st)︸ ︷︷ ︸MPh(st)

. (2.5)

In words: In the optimum, the real rental return is equated to the marginal

product of capital, MPK, times real marginal costs; similarly, the real wage

is equated to the marginal product of labor, MPh, times real marginal costs.

We will comment on this peculiarity further below.

Output supply: The flexible price case

To have a suitable reference point, we start by considering the case where

intermediate goods producers can adjust prices in every period. In this

instance, firm i solves the following static profit maximization problem

maxPx(i)

Πx (i) = [Px (i) − P Cm ] X (i)

such that

X (i) =

[Px (i)

Px

] 1

θ−1

X.

The FOC to this problem can be rearranged to the optimality condition

Px (i) =1

θP Cm.

In words: Monopolist i sells at a price which is greater than the socially

optimal price, which is its marginal costs (in nominal terms). As a conse-

quence, the monopolist’s optimal output must be below the socially optimal

(or competitive) output level. The price distortion in form of the markup,

1/θ, is larger when the representative final good producer, facing a price in-

crease, reduces its demand only slightly (that is, when the price elasticity of

demand is small). Note that all firms sell at the same price, i.e., Px(i) = Px.

Page 45: Macroeconomics of small open economies

2.2. THE MODEL 35

The inefficiency in the final good market has its mirror image in the two

input markets. To make this clear, consider the long-run equilibrium where

the relative price Px/P is assumed to be 1 and, thus, Cm = θ. From the two

conditions which implicitly define the factor demands, equation (2.4) and

(2.5), follows that on average there is a wedge between the marginal product

of labor and the real wage on the one hand and the marginal product of

capital and the real rental price on the other hand. Generally speaking, the

average firm pays a rental price which is smaller than the socially optimal

rental price, given by the marginal product. Again, the wedge (in the form

of θ) is larger when the price elasticity of demand is small.

Output supply: The sticky price case

We now turn to the situation where prices are sluggish. Apart from price

sluggishness we also allow for sustained (or steady state) inflation; that is,

the central bank manipulates its policy instrument such that the money

supply and the domestic price level are growing at a constant rate. This

assumption has to be seen as a means of preventing that in a calibrated

version of the model, large shocks affecting the economy lead to a nominal

interest rate which hits the zero lower bound.

Price change signal Let q be the probability that a firm gets a price-

change signal in a given period. Note that q does not depend on the duration

of the interval of price fixity. By visualizing an event tree, we immediately

see that (1 − q)j is the probability of being stuck in period t + j with the

price which was set at t and that (1 − q)j−1 q is the probability of adjusting

in period t + j (or, alternatively, of adjusting in j periods).

Imagine a monopolist who gets a price-change signal in period t. The

probability that firm i can adjust its price in period 1 is q; the probability

that firm i can adjust in period 2 is (1 − q) q; the probability that firm i

can adjust in period 3 is (1 − q)2 q, and so on. Thus, the time over which

a price is fixed can be considered as a discrete random variable with p.d.f.

(1 − q)t−1 q. The average time over which a price is fixed, thus, is given by

∞∑

t=1

t (1 − q)t−1 q = q[1 + 2 (1 − q) + 3 (1 − q)2 + ...

]=

1

q.

Page 46: Macroeconomics of small open economies

36 CHAPTER 2

Sustained inflation Monopolist i anticipates that on average the do-

mestic price level is growing at a constant rate. Therefore, the monopolist

chooses a ”deflated” price of the good as the choice variable and then multi-

plies the ”deflated” price by the steady-state growth factor of the domestic

price level to get the optimal nominal price of good i in time t.6 Formally,

Px

(i, st)

= Ξ · px

(i, st)

where px (i, st) is the ”deflated” price of good i and Ξ is the steady-state

growth factor of the PPI.

Firm i’s profit maximization behavior We are now in the position to

state firm i’s profit maximization problem in the presence of sticky prices.

A firm resetting the price of its good in period t chooses a price px (i, st) in

order to maximize

Πx

(i, st)+

∞∑

τ=1

st+τ

P b(st+τ

∣∣ st)(1 − q)τ−1

[qΠx

(i, st+τ

)+ (1 − q) Πx

(i, st+τ

)]

(2.6)

where P b (st+τ | st) denotes the τ -step pricing kernel which is used to value

date t + τ profits,7 Πx (i, st+τ ) is the profit attained when the firm gets a

price-change signal in period t + τ (an event which occurs with probability

(1 − q)τ−1 q), given by

Πx

(i, st+τ

)=[Ξτpx

(i, st+τ

)− P

(st+τ

)Cm

(st+τ

)]X(i, st+τ

),

and Πx (i, st+τ ) is the profit attained when the firm gets no price-change

signal in period t + τ (an event which occurs with probability (1 − q)τ ),

given by

Πx

(i, st+τ

)=[Ξτ px

(i, st)− P

(st+τ

)Cm

(st+τ

)]X(i, st+τ

).

The maximization takes place subject to the sequence of total conditional

factor demand constraints (2.1) (expressed in terms of px (i, st)),

X(i, st+τ

)=

[Ξτ px (i, st)

Px (st)

] 1

θ−1

X(st+τ

)for all τ = 0, 1, ...

6Deflated is put in quotation marks because it only refers to the deterministic part of

PPI inflation.7 P b (st+τ | st) gives the value of an asset which pays exactly one unit of money in

state sj in period t + τ and zero otherwise.

Page 47: Macroeconomics of small open economies

2.2. THE MODEL 37

The expected profit is maximized at8

px

(st)

=1

θ

∑∞

τ=0

∑st+τ P b (st+τ | st) (1 − q)

τ(Ξτ )

1θ−1 Px (st)

11−θ P (st+τ ) Cm (st+τ ) X (st+τ )

∑∞

τ=0

∑st+τ P b (st+τ | st) (1 − q)

τ[Ξτ ]

θ

θ−1 Px (st)1

1−θ X (st+τ ).

(2.7)

PPI evolution under a Calvo price setting structure Given that

prices are set a la Calvo, how does the aggregate intermediate price index

evolve over time?

Recall that the PPI is given by (expressed in terms of px (i, st))

Px

(st)

=

∫ 1

0

[Ξ · px

(i, st)] θ

θ−1 di

θ−1

θ

.

The Calvo price setting structure allows us to make statements about the

price of every individual producer i. We know that in any period t some

producers are stuck with a price set t− j periods ago, where j goes toward

infinity (recall that all producers which set their price in the same period

will choose the same price). By visualizing ones more an event tree, it

becomes evident that in every period t a fraction of (1 − q)j−1 q of prices

from period t − j survives. The price index, thus, becomes

Px

(st)

=

q

∞∑

j=1

(1 − q)j−1 [Ξ · px

(st+1−j

)] θθ−1

θ−1

θ

.

This expression can be turned into the following non-linear difference equa-

tion

Px

(st)

=

qpx

(st) θ

θ−1 + (1 − q) Ξθ

θ−1 Px

(st−1

) θθ−1

θ−1

θ

. (2.8)

2.2.4 Behavior of representative agent

The representative household maximizes lifetime utility

∞∑

τ=0

st+τ

βτπ(st+τ

∣∣ st)U

[C(st+τ

),M (st+τ )

P (st+τ ), l(st+τ

)]

8Due to the symmetry of producers, all firms which get a price-change signal in period

t will set the same price such that px (i, st) = px (st).

Page 48: Macroeconomics of small open economies

38 CHAPTER 2

where β ∈ (0, 1) is the constant discount factor and π (st+τ | st) is the (ob-

jectively known) probability of state st+τ conditional on being in state st in

period t.9 The arguments in the momentary (or period-by-period) utility

function are consumption, C (st), real balances, M (st) /P (st), and leisure,

l (st). In what follows, the momentary utility is specified to

U

[C(st),M (st)

P (st), l(st)]

=1

1 − σ

[Ψ(st)ν

l(st)1−ν

]1−σ

− 1

where

Ψ(st)≡

C(st)η

+ ζ

[M (st)

P (st)

]η 1

η

.

The parameters satisfy the conditions σ, η > 0 and ν, ζ ∈ (0, 1). For a

discussion of this kind of nonseparable preferences compare e.g. Chari et al.

[6].

Constraints

Budget constraint The agent’s maximization is subject to three con-

straints which are discussed in turn. The first is the budget constraint.

We assume complete financial markets, that is, there is a world market in

one-period claims which completely spans the relevant uncertainty faced by

the households in the small open economy and the rest of the world about

future income, prices, etc. The payoff of these claims is assumed to be in

terms of domestic paper money.10 The period-by-period budget constraint

in nominal terms is given by∑

st+1

P b(st+1

∣∣ st)B(st+1

)+ M

(st)

+ P(st) [

C(st)

+ I(st)]

+ T(st)

≤ B(st)

+ M(st−1

)+ N

(st)

+ Π(st)

+ P(st) [

z(st)K(st−1

)+ W

(st)h(st)]

. (2.9)

Let us start the discussion of inequality (2.9) by considering the source of

funds. There are three assets in the economy. Of each asset, the household

9Given that the stochastic processes in the model are Markovian, π (st+τ | st) can be

derived via recursion (see e.g. Ljungqvist and Sargent [23]).10This is just for convenience; the claims could equally well be assumed to be denoted

in the world currency.

Page 49: Macroeconomics of small open economies

2.2. THE MODEL 39

brings a certain amount into period t. B (st) is the share of a claim which

is contingent on the state at t being st and pays out one unit of paper

money; hence, B (st) · 1 is the total amount of paper money paid out in

period t. M (st−1) is the agent’s holdings of nominal money balances and

K (st−1) is the amount of physical capital which is leased to the intermediate

goods producers at the nominal rental price P (st) z (st). Note that K (st)

is somehow split up in a continuum of differentiated capital supplies, such

that

K(st)

=

∫ 1

0

K(i, st)di.

Apart from the three assets carried over into period t, there are three

additional sources of funds. First, the agent sacrifices some leisure-time to

labor. The amount of hours worked, h (st), is leased to the intermediate

goods producers at the nominal wage rate P (st) W (st). Again, h (st) is

somehow split up in a continuum of differentiated labor supplies, such that

h(st)

=

∫ 1

0

h(i, st)di.

Second, the agent receives a nominal lump-sum transfer from the govern-

ment, N (st). Finally, the agent receives the nominal profits of the monop-

olistically competitive intermediate goods producers,

Π(st)

=

∫ 1

0

Π(i, st)di.

The household allocates its resources between investment in the three

assets, in consumption, and in taxes (the use of funds). P b (st+1| st) is

the one-step pricing kernel, B (st+1) is the amount of claims regarding a

particular state sj, I (st) is real investment expenditures, and T (st) is a

nominal lump-sum tax.

Time endowment and capital evolution Apart from the period-by-

period budget constraint in nominal terms, there are two additional con-

straints to the households maximization problem. First, time devoted to

labor and leisure is equal to total time endowment (normalized to 1)

l(st)

+ h(st)

= 1. (2.10)

Page 50: Macroeconomics of small open economies

40 CHAPTER 2

Second, the evolution of the capital stock follows

K(st)

= I(st)− Φ

[I(st)

+

, K(st−1

)

]+ (1 − δ) K

(st−1

)(2.11)

where δ ∈ (0, 1) is the rate of depreciation. In words: I (st) units of gross

investment involve adjustment costs which depend positively on the amount

of gross investment, I (st), and negatively on the amount of the capital

already in place, K (st−1). The Uzawa-type adjustment cost function in

equation (2.11) is specified to

Φ[I(st), K(st−1

)]≡

φ

2

[I (st)

K (st−1)− δ

]2

K(st−1

)

where φ > 0 is a cost parameter.

Digression: Real demand for money

To get some economic insight, we derive a function which implicitly defines

the real demand for money. Combining the FOC for C (st) and for M (st)

yields

ζ

[M (st)

P (st) C (st)

]η−1

= 1 − β

∑st+1 π (st+1| st) Λ1 (st+1)

Λ1 (st)(2.12)

where Λ1 (st) is the nominal shadow price on the first constraint to the

agent’s maximization problem. From the FOC for B (st+1), we get an ex-

pression for the one-step pricing kernel

P b(st+1

∣∣ st)

= βπ(st+1

∣∣ st) Λ1 (st+1)

Λ1 (st). (2.13)

The nominal gross return of an asset that yields one unit of money in state

st+1 with certainty is given by

R(st)

=1

P b (st+1| st).

Substituting for P b (st+1| st) from equation (2.13), provides us with

1

R (st)= β

st+1

π(st+1

∣∣ st) Λ1 (st+1)

Λ1 (st)

Page 51: Macroeconomics of small open economies

2.2. THE MODEL 41

which can alternatively be written as

1 = β∑

st+1

π(st+1

∣∣ st) UC (st+1)

UC (st)

R (st) P (st)

P (st+1)

where UC (st) stands for ∂U(st)/∂C(st). The ratio R (st) P (st) /P (st+1) is,

of course, the gross real interest rate and the resulting condition the famous

Lucas asset pricing equation.

Finally, substituting Λ1 (st) /R (st) for β∑

st+1 π (st+1| st) Λ1 (st+1) in

equation (2.12) leads to the following condition which characterizes the de-

mand for real money balances as a function of the nominal rate of interest

and real consumption

ζ

[M (st)

P (st) C (st)

]η−1

=[1 − R

(st)−1].

2.2.5 Fiscal and monetary policy

The output of the final good production is either used as a consumption

good, a capital good, or it is absorbed by the public sector. It is assumed

that public services do neither provide utility to households nor are they

an input to (private) production. We further assume that the government

spends amount G (st) of the final good and that the cyclical component of

G (st) is determined exogenously. Since Ricardian equivalence holds in this

model, we can - without loss of generality - assume that the government

runs a balanced budget each period

P(st)G(st)

= T(st).

Monetary policy is assumed to be either active or passive. Active mon-

etary policy is implemented by means of a simple instrument rule. Granted

that the exchange rate is floating, the nominal interest rate, R(st), is a lin-

ear function of the lagged nominal rate, the long-run equilibrium nominal

rate and deviations of the actual values of output, inflation, and the terms

of trade from their respective target values

ln R(st) = ρr ln R(st−1) + (1 − ρr)ln R + φy

[ln Y (st) − ln Y

]

+ φπ

[ln πx(s

t) − ln πx

]+ φe

[ln etot(st) − ln etot

](2.14)

Page 52: Macroeconomics of small open economies

42 CHAPTER 2

where πx(st) is PPI inflation and etot(st) is the small open economy’s terms

of trade (the relative price of imports and exports), defined by

etot(st) ≡e(st)P ∗

x (st)

Px(st).

The policy parameters (or reaction coefficients) ρr, φy, φπ, and φe determine

how aggressively policy responds to the lagged interest rate and to devia-

tions of the target variables from their respective target values. Like the

target values Y , πx, and etot, they are chosen by the monetary authority. If

the (gross) nominal interest rate follows (2.14), money holdings are demand

determined, that is, the money supply is set so as to satisfy any money

demand that prevails at the ongoing interest rate.

A passive monetary policy can have two forms (within the present set-

ting). One alternative is that the growth process of the nominal money

supply is specified as follows

M(st+1

)= M

(st)µ(st)

(2.15)

where the variable µ (st) (one plus the money growth rate) is assumed to

follow a stochastic AR(1) process (to be specified) and the nominal exchange

rate is assumed to be completely flexible.11 A second alternative is that the

nominal exchange rate is fixed toward the currency of the rest of the world

at an arbitrary value. In this event, the nominal money supply is chosen

such that the exchange rate target is achieved to the full extent in every

period.

Under all three regimes, new money is introduced through lump-sum

transfers from the monetary authority to households:

M(st)− M

(st−1

)= N

(st).

2.2.6 Market clearing

The following market clearing conditions hold in equilibrium. Final good

market:

Y(st)

= C(st)

+ I(st)

+ G(st).

11Obviously, in the steady state µ = Ξ.

Page 53: Macroeconomics of small open economies

2.2. THE MODEL 43

Intermediate goods market i:12

X(i, st)

= Xd(i, st)

+ Xd∗(i, st)

from which follows that

X(st)

= Xd(st)

+ Xd∗(st).

Labor market:

h(st)

=

∫ 1

0

h(i, st)di.

Capital market:

K(st)

=

∫ 1

0

K(i, st)di.

Money market:13

Md(st)

= M s(st)

Since all these markets clear in equilibrium, by Walras law the market for

financial securities clears, too.

2.2.7 Closing the model: International asset markets

In order to close the model, the Euler equation of the representative agent

of the small open economy has to be linked to the international financial

asset market. From the expression for the one-step pricing kernel (2.13) and

the FOC for C (st), we get the following optimality condition

P b(st+1

∣∣ st)

= βπ(st+1

∣∣ st) UC (st+1) /P (st+1)

UC (st) /P (st).

In the rest of the world, agents have access to the same array of financial

assets as in the domestic economy. Thus, there is a similar condition for

12Note that since intermediate goods prices are eroded over time as inflation is above

average, intermediate goods producers do not charge the same output prices and, thus,

do not produce the same amount of output (this phenomenon is called price dispersion).

Since indirect demand functions depend on output (among others), conditional factor

demands are asymmetric, too.13In fact, we do not distinct between money demand and supply but just write M (st).

Page 54: Macroeconomics of small open economies

44 CHAPTER 2

the rest of the world14

P b(st+1

∣∣ st)

= βπ(st+1

∣∣ st) U∗

C (st+1) / [e (st+1) P ∗ (st+1)]

U∗

C (st) / [e (st) P ∗ (st)]

where we assume that domestic and foreign households share the same sub-

jective discount factor. Arbitrage implies that

UC (st+1)

UC (st)

P (st)

P (st+1)=

U∗

C (st+1)

U∗

C (st)

P ∗ (st)

P ∗ (st+1)

e (st)

e (st+1).

Iteration yields (compare e.g. Chari et al. [7], p. 14)

UC (st)

U∗

C (st)

P ∗ (st) e (st)

P (st)=

UC (s0)

U∗

C (s0)

P ∗ (s0) e (s0)

P (s0).

We end up with the following expression for the nominal exchange rate

e(st)

= κU∗

C (st)

UC (st)

P (st)

P ∗ (st)(2.16)

where κ ≡e(s0)UC(s0)P ∗(s0)

P (s0)U∗

C(s0)

and the ratio U∗

C (st) /P ∗ (st) is exogenous.

2.3 Solution, parameterization, and diagnos-

tic check

The equilibrium of this economy is a sequence of prices and quantities for

which the representative household’s problem and the firms’ problems are

solved and markets are cleared. In order to compute those sequences, the

optimality conditions for the representative household and the firms, the

government’s budget constraint, the central bank’s feedback rule, and the

market clearing conditions must be put together. The resulting system of

equations we call the (competitive) equilibrium conditions. We then ap-

ply the standard linear approximation method from the RBC literature

(compare e.g. King et al. [18] and King and Rebelo [20]), which involves

14Recall that the nominal price of a one-period claim that pays out one unit of paper

money if state st occurs and nothing otherwise, P b(st+1

∣∣ st), is denominated in home

currency; thus, the foreign price level has to be converted.

Page 55: Macroeconomics of small open economies

2.3. SOLUTION, PARAMETERIZATION, AND DIAGNOSTIC CHECK 45

determining the properties of the non-stochastic steady state (of the system

which describes the equilibrium) and taking a first-order Taylor expansion

of the equilibrium conditions around the steady state. The resulting linear

system of structural relations determine the prices and quantities (in terms

of percentage deviations from the steady state) in period t = 0, 1, 2, ...; it

can be solved by means of any computational algorithm for solving and

simulating linear rational expectations models.15

But before we can solve and simulate the model we have to assign values

to the structural parameters. Moreover, we undertake the following change

of notation: We write Yt for Y (st) etc. and let Et• be a function which

takes the expected value of the term inside the curly bracket, based on

information available in t.16

2.3.1 Baseline parameterization

Rather than assigning values to the structural parameters such that the

behavior of the model economy matches features of measured data for a

real-life economy, we borrow a set of plausible values from other studies, a

strategy which is not uncommon in the literature. Following Collard and

Dellas [10], we set the parameters to:17

15The one we make use of is a version of the classical solution algorithm developed by

King, Plosser, and Rebelo [18].16One may wonder why we did not do this from the beginning on. The reason is that

in our view, solving the representative household’s and the firms’ optimization problems

in terms of the states of the economy makes the underlying economics more transparent.17Collard and Dellas [10] investigate the role played by price sluggishness in the perfor-

mance of alternative exchange rate regimes in a two-country NNS model. Their model is

calibrated on the postwar US economy. For parameter values they heavily rely on Cooley

and Prescott [14] and Chari et al. [7]. Collard and Dellas find that a high degree of price

sluggishness tends to favor the flexible regime while a low degree favors the fixed regime.

Page 56: Macroeconomics of small open economies

46 CHAPTER 2

Category: Parameter: Value

Final good producer: ω (one minus import share) 0.8000

ρ (substitution parameter - upper level) 0.3333

θ (substitution parameter - lower level) 0.8000

Intermediate goods prod.: α (capital elasticity) 0.2813

Discount factor: β 0.9880

Momentary utility: σ (relative risk aversion) 1.5000

η (parameter in Ψ(•)) 1.5600

ζ (weight of money - liquidity service) 0.0649

Rate of depreciation: δ 0.0250

Adjustment costs: φ 10.000

Time devoted to labor: h (= 1 − l) 0.3100

Degree of price stickiness: q 0.2500

GDP devoted to govt. exp.: g 0.2200

Steady state inflation: Ξ 1.0260

Note that the discount factor is set so as to imply (approximately) a 5%

annual subjective discount rate, following the formula

β =

(1

1 + 0.05

)0.25

.

The degree of price stickiness is chosen such that the average duration prices

remain fixed is four years. The implied steady state values are given by:

C 0.3944

M/P 0.4910

l 0.6900

Y (= X) 0.6274

Xf 0.1255

Xd 0.5020

I 0.0950

G 0.1380

K 3.8012

R 1.0385

Cm 0.8000

z 0.0371

ν 0.3394

Page 57: Macroeconomics of small open economies

2.3. SOLUTION, PARAMETERIZATION, AND DIAGNOSTIC CHECK 47

The autoregressive parameters of the shocks are set to:

Parameter: Value

ρa (productivity) 0.9500

ρg (government expenditures) 0.9700

ρm (money growth) 0.4900

ρl (foreign lambda) 0.9500

ρp (foreign intermediate goods prices) 0.9500

ρx (foreign intermediate goods demand) 0.9500

Standard deviations are normalized to 0.01.

2.3.2 Dynamic effects

We are now in the position to compute the model’s equilibrium dynamics for

alternative specifications of monetary policy. For the time being, we focus on

three policy regimes: (i) constant money growth, (ii) an unilateral exchange

rate peg, and (iii) strict domestic inflation targeting. For each regime, the

impulse responses will be analyzed and the role of key parameters (like

the degree of price stickiness) investigated. The purpose of this exercise is

twofold. First, we want to gain some intuition on how the model works.

Second, given that we have no measure of fit, comparing the dynamics of

our model with the qualitative predictions of models of the same class and

with comparable parameterizations is a way to make sure that there are no

coding or other errors.18 The results are summarized in table 1.

Domestic technology shock

Fixed money supply The effect of sticky prices is best understood in

terms of variations in the markup of price over marginal costs. In the style

of King and Wolman [19] and Goodfriend and King [17] we define the time

varying, endogenous average markup as the ratio of the intermediate price

level to nominal marginal costs of production

υt ≡Px,t

PtCm,t

18We say ”to make sure” even though we are well aware of the fact that one can never

be virtually sure that ones code is faultless.

Page 58: Macroeconomics of small open economies

48 CHAPTER 2

or, alternatively, in terms of deflated variables (see appendix 2.A):

υt = (ptCm,t)−1

where pt = Pt/Px,t is the deflated PPI. In terms of percentage deviations

from the steady state we have

υt = −(pt + Cm,t

).

Suppose monetary policy follows a constant money growth rule. More-

over, suppose there are no capital adjustment costs (i.e., φ = 1). If prices

are fully flexible (which is implemented by letting q almost but not exactly

1), the markup is not affected by a positive technology shock; the elasticity

of υ with respect to a supply shock is literally zero. We conclude that the

responses of the real variables to the shocks are efficient.19 Not surprisingly,

the dynamics are in line with the predictions of the baseline RBC model

of a closed one-good, one-shock economy (compare e.g. King and Rebelo,

[20]).

The marginal product of labor increases above its steady state value and

stays there for a protracted period. This results in a rise of the real wage;

the elasticity of Wt is close to 0.8.20 Accordingly, there is a great incentive

to substitute intertemporally and to take less leisure now and in the near

future than in the far future. However, there is also an offsetting income

effect. Initially, the substitution effect outweighs the income effect and,

thus, work effort responses positively. The positive labor response amplifies

the productivity shock; the elasticity of Xdt is about 1.4. As long as the

domestic productivity is the only source of shocks, total output of domestic

intermediate goods production is given by Xt = ωXdt . And so, Xt moves

along with Xdt .

The rise in At together with the positive labor response lead to a marginal

product of capital which is higher than normal. The impact effect of the

real rental price, zt, is about 0.9. Most of the additional input therefore

19In the words of Rotemberg and Woodford [29], the average markup represents a

measure of if and how fluctuations in real variables are inefficient.20The impulse effect of Wt is not as large as the impulse effect on MPht. This is

because Wt = Cm,t + MPht and a gain in productivity leads to a reduction of real

marginal costs below the long-run level (the elasticity of Cm,t is about −0.2).

Page 59: Macroeconomics of small open economies

2.3. SOLUTION, PARAMETERIZATION, AND DIAGNOSTIC CHECK 49

is invested, which leads to a higher capital stock. As productivity decays

geometrically, the intertemporal substitution effect is outbalanced by the

offsetting income effect: work effort drops below its steady-state level. Also,

investment becomes lower than normal as the capital stock is reduced to its

stationary level.

How do international trade aspects come into play? An increase in

the supply of domestic intermediate goods, Xt, lowers the relative price

of domestic intermediate goods, i.e., induces a deterioration of the terms

of trade. As a consequence, the amount of intermediate inputs imported

by the representative final good producer, Xft , decreases while the amount

of inputs purchased from domestic intermediate goods producers, Xdt , in-

creases (expenditure-switching).

What about the nominal variables? The rise in consumption induced by

a positive supply shock tends to raise the demand for money. The fact that

nominal money growth remains constant gives rise to a decrease of the CPI,

Pt, (and, thus, an increase of (M/P )t) and an increase of the opportunity

costs of holding money, Rt.21 The nominal exchange rate, finally, increases

(i.e., the domestic currency depreciates).

The picture changes dramatically when prices are assumed to be sticky

(q = 0.25). To begin with, the dynamics now exhibit the hump-shaped

pattern typical for sticky-price economies. In addition to this, the elasticity

of the markup amounts to 2.3.22 Why this massive rise in the markup in the

presence of sticky prices? A sudden gain in productivity produces a shift in

the marginal cost schedule. But because prices do not fall immediately in

proportion to the decline in costs, markups rise. This is inevitable: Firms

would like to lower prices (thereby expanding output); but a large fraction

of firms does not receive a price adjustment signal and, thus, remains stuck

with the price from the previous period. Some prices, however, do fall

and thus output increases. The transitory rise in the markup – relative

to what would happen under flexible prices – lessens the output effect of

the technology shock; the elasticity of Xt is −0.5, compared to 1.1 in the

21When the weight for money in the utility function, ζ, is non-zero, changes in real

balances affect the marginal utility of consumption and labor and thereby Ct and ht; this

effect, however, is small.22Notice: υt falls back to normal rather quickly and approaches its long-run level just

after a few quarters.

Page 60: Macroeconomics of small open economies

50 CHAPTER 2

flex-price case.23

Since the increase of the markup is temporary, we observe a strong

substitution effect (see e.g. King and Wolman [19], p. 20): the supply of

labor is massively reduced; the same is true for investment. As the gap

returns to normal, the supply of labor expands and investment increases.

The response of the input factors is mirrored by the response of rental prices:

the impact effect of both Wt and zt now is negative (the elasticities are −0.9

and −2.9, respectively).

This result is standard in models of this type. For an illustration, con-

sider the baseline sticky price model of a closed economy as put forward by

Clarida et al. [9], with a comparable parameterization. In this model, the

output gap can only increase if the marginal costs increase or, alternatively,

if the markup falls. Moreover, there is a link between markup and inflation.

With fully flexible prices and a fixed money supply, the impact effect of the

markup to a positive technology shock is zero; the elasticity of the real wage

with respect to the supply shock is about 1.0. If prices are sticky, however,

the impact effect of the markup to a positive technology shock is significant

and the elasticity of the real wage with respect to the supply shock is about

−0.7.

What happens when we introduce capital adjustment costs (i.e., when

we set φ to its baseline value of 10) in the flex price environment? Recall

that some of the windfall associated with a productivity shock is consumed,

some is invested. (The third component of aggregate demand, government

consumption, is unaffected by a productivity shock.) In the presence of high

adjustment costs, investment will not increase by as much as in the previ-

ously discussed no-adjustment cost scenario. Given that the representative

agent has an incentive to smooth consumption (i.e., that the agent does

not just want to expand consumption by the same amount by which the

response of investment turns out to be smaller), the only way for aggregate

demand and aggregate supply to be equalized is that hours worked drop be-

low their long-run level. Hence, the most significant difference compared to

the no-adjustment cost scenario (besides the significantly smaller response

of investment) is the negative impact effect of work effort.24

23Notice: Xt lies above normal from period 2 onward.24Starting from the baseline parameterization, a way to provoke a positive impact effect

Page 61: Macroeconomics of small open economies

2.3. SOLUTION, PARAMETERIZATION, AND DIAGNOSTIC CHECK 51

Exchange rate peg Under the previously discussed floating regime with

constant money growth, et increases in response to a positive technology

shock, that is, the domestic currency depreciates. Accordingly, to keep

et = 0 for all t, monetary policy needs to be contractionary: Mt decreases

and returns to normal only slowly.25 The cost of stabilizing the nominal

exchange rate (in terms of a contractionary monetary policy), however, are

negligible if prices are flexible and still small if prices are sticky.

Strict domestic inflation targeting If prices are sticky and monetary

policy actively seeks to stabilize domestic inflation, the response of the real

variables to a positive domestic technology shock is pretty much the same

as under the previously discussed floating regime under flexible prices and

with a fixed money supply. The picture differs, however, with regard to the

nominal variables: Since monetary policy succeeds in stabilizing domestic

inflation around its long-run equilibrium level, the effect on πx,t (and, thus,

the markup) is zero. Given that prices are sticky, a deterioration of the

terms of trade (i.e., a rise in etott ) leads to a rise of the nominal exchange

rate; the domestic CPI becomes negative on impact and returns to normal

only gradually.

Domestic fiscal policy shock

With a constant money growth and flexible prices, a shock to government

expenditures produces the standard results (compare e.g. Baxter and King,

[3]). A persistent (but not permanent) shift in government consumption

financed by a lump-sum tax means higher future taxes, which induces a

moderate negative wealth effect. To this negative wealth effect the repre-

sentative household responds by decreasing consumption and leisure. As a

corollary, ht and Xt jump to positive values. Given the slowly adjusting

capital stock, the shift in ht leads to a marginal productivity of labor below

average which in turn leads to a negative percentage deviation of the real

wage from its steady state value.

of work effort other than setting φ to 1.0 is to let domestic and foreign inputs become

closer substitutes, i.e., to increase the parameter ρ (compare Collard and Dellas [10]).25Compare appendix 2.D to learn how we implement an exchange rate peg.

Page 62: Macroeconomics of small open economies

52 CHAPTER 2

The behavior of investment is influenced by two offsetting effects. On the

one hand, due to the increased government absorption of resources, there

are reduced opportunities for private uses of output, i.e., consumption and

investment.26 On the other hand, the increase in labor input shifts up the

marginal product for capital, a force which works in the direction of more

investment. Given the baseline parameterization, the first effect outweighs

the second and, hence, I becomes slightly negative for a protracted period.

In the long run, however, investment is above normal as the economy works

to rebuild the capital stock.

If intermediate prices are assumed to be sticky, the reaction of domestic

inflation is weakened and slowed down. The markup now is significantly

positive on impact and outweighs the negative marginal productivity of la-

bor, which brings about a positive impact effect of the real wage. Otherwise,

in qualitative terms the picture largely remains the same as under flexible

prices.

Foreign shocks

How does the economy respond to a shock to the nominal foreign shadow

price, Λ∗

1,t? Suppose money growth is constant, the exchange rate is flexible,

and intermediate prices are sticky (q = 0.25). Moreover, suppose the inter-

mediate goods produced in the small open economy represent a negligible

input into world final good production. Finally, suppose for a moment that

the world monetary authority succeeds in fully stabilizing the foreign inter-

mediate goods price level. The third together with the second assumption

imply that the foreign CPI coincides with the foreign PPI, which in turn is

constant.

To derive an expression for the terms of trade, note that the equilibrium

condition et = κΛ∗

1,t/Λ1,t can be extended to

etott

(=

etP∗

x,t

Px,t

)= κ

Λ∗

1,tP∗

x,t

Λ1,tPx,t

.

Taking into account our previously made assumptions (and the definition

of Λ∗

1,t = U∗

C,t/P∗

x,t), we end up with

etott = κU∗

C,t/λ1,t,

26Keep in mind that the negative wealth effect induces consumption to decrease.

Page 63: Macroeconomics of small open economies

2.3. SOLUTION, PARAMETERIZATION, AND DIAGNOSTIC CHECK 53

where κ = κ/P ∗

x and λ1,t = Λ1,tPx,t. In terms of percentage deviations from

the steady state, we have

etott = U∗

C,t − λ1,t.

Note that λ1,t is an endogenous variable which responds to the exogenous

forces in the model in much the same way as any other endogenous variable,

while U∗

C,t (which under the given assumptions equals Λ∗

1,t) is exogenous.

Now, consider a positive shock to Λ∗

1,t(= U∗

C,t). For the baseline (and in

fact for any reasonable alternative) parameterization, although λt is affected

too, by far the biggest part of the sudden increase in etott stems from the

change in Λ∗

1,t. In response to this unexpected deterioration of the terms of

trade, the final good producers substitute away from foreign inputs toward

domestic inputs. At the same time, the terms of trade deterioration induces

a negative wealth effect (the small open economy needs to export more in

order to purchase a given bundle of imports) and, thus, a drop of consump-

tion and leisure (which in turn leads to an increase in ht). The decrease

in consumption is smoothed by a cut in investment. The two price levels

(domestic and consumption) rise and the final good production declines.

The picture remains the same when we allow for changes in the foreign

PPI: A positive shock to P ∗

x,t has identical effects on the nominal and real

variables in the model as a positive shock to Λ∗

1,t – with the prominent

exception of the nominal exchange rate (compare table 1). Since within the

policy simulation exercise variations in et are of no concern for us, we will

further below refer to a so-called ”terms-of-trade shock”, where we mean a

temporary change in either Λ∗

1,t or P ∗

x,t.

Under the same scenario (constant money growth, flexible exchange rate,

sticky intermediate prices) and under the additional assumption that capital

is costless to adjust, the dynamics implied by a foreign demand shock in

favor of domestic intermediate inputs (that is, a unexpected increase in Xd∗t )

are consistent with the traditional Mundell-Fleming-Dornbusch model (as

described e.g. in Clarida and Gali [8]): the terms of trades improve and the

price levels and the final good production rise.

Page 64: Macroeconomics of small open economies

54 CHAPTER 2

Domestic money growth shock

Suppose money growth follows a stochastic process: What are the effects of

a shock to the money growth process, given that prices and exchange rates

are fully flexible? If ζ is non-zero, the model’s predictions are in line with

the standard money-in-the-utility-model of a closed one-sector economy as

described e.g. in Walsh [35], chapter 2. This is to say, the effects on real

variables (except real balances) are extremely small.

However small these effects are, where do they come from? Suppose

the growth rate of money follows the stochastic process µt = ρmµt−1 + εm,t

where ρm ∈ (0, 1) and εm,t is i.i.d. with zero mean. Now consider the

effect of a positive shock to µt. Since future money growth will be above

average for a protracted period, expectations of future inflation instantly

rise; both price levels jump to higher levels. How does this affect real

balances? Given the baseline parameterization, the CPI rises more than in

proportion to the rise in the nominal money stock and, thus, real money

balances decrease on impact. As both the nominal money stock and the CPI

gradually climb to their new steady state values, real money balances returns

to normal. The reduction of real money balances lowers the marginal utility

of consumption and - by affecting the ratio of the marginal utility of leisure

to the marginal utility of consumption - causes the agent to substitute away

from consumption towards leisure. As a consequence, work effort as well as

intermediate goods output fall.

If intermediate prices are sticky, the same mechanism is at work. Again,

the adjustment process of prices is now weakened and slowed down. Over-

all, the implied dynamics are in line with the traditional Mundell-Fleming-

Dornbusch model: A shock to the money growth rate results in a perma-

nent nominal depreciation, a permanent rise in the price level (both PPI

and CPI), a temporary rise of intermediate and final good production, and

a temporary deterioration of the terms of trades.

2.4 The role of exchange rate stabilization

The purpose of this section is to evaluate selected members of the family

of simple policy rules represented by equation (2.14). The chosen strategy

for policy evaluation follows the steps proposed by Taylor [34], pp. 263/4:

Page 65: Macroeconomics of small open economies

2.4. THE ROLE OF EXCHANGE RATE STABILIZATION 55

First, we place a particular specification of equation (2.14) into the pa-

rameterized model. Second, we solve the model using a numerical solution

algorithm. Third, we examine the properties of the stochastic behavior of

those macroeconomic variables which reflect potential goals of monetary

policy, in particular the variability of Yt, πx,t, πt, etott , and Rt.

27 Finally, we

examine the consequences on welfare, measured by means of a quadratic

approximation of the representative household’s lifetime utility, given by

W ≈ (1 − β)−1

[U +

C · UC

2(1 + ξC) σCC +

m · Um

2(1 + ξm) σmm

+l · Ul

2(1 + ξl) σll + C · m · UCm · σCm + C · l · UCl · σCl + m · l · Uml · σml

]

where U ≡ U [C, m, l] with m ≡ M/P , UC ≡ ∂U/∂C, ξC ≡ C ·UCC/UC , and

σCC = Corr(Ct, Ct) are functions of the structural parameters of the model.

Both performance and welfare are conditioned on the kind of disturbance

which hits the economy; this is because unless the model is calibrated we

cannot make predictions regarding the relative size of each kind of shock.

Why focusing on instrument rules such as (2.14)? A first argument given

in the literature refers to their simplicity : Taylor rules may serve as an in-

formative guideline for policy or as an aid in promoting policy transparency

(compare e.g. Walsh [35], p. 549). A second argument has been pointed out

by McCallum (compare e.g. [24]): the optimal policy rule determined by

solving an explicit policy design problem crucially depends on the form of

the model under consideration.28 The recommended research strategy is to

search for a policy rule which possesses robustness in the sense of yielding

a desirable outcome in policy simulation experiments in a wide variety of

models. A third argument can be seen in the fact that the derivation of the

(globally) optimal policy would be relatively straightforward if a number

of simplifying assumptions were imposed, but is much more involved and

in fact intractable in the rather general setting at hand. In this event, an

alternative is to restrict the instrument rule to lie in a given class and at

the same time to assume that the central bank can commit, once and for

all, to a given policy rule for all future periods.

27On the choice of Yt compare below.28In the words of Taylor [33], p. 11, ”[t]he optimal rule exploits properties of a model

which are specific to that model, and when the optimal rule is then simulated in another

model those properties are likely to be different and the optimal rule works poorly.”

Page 66: Macroeconomics of small open economies

56 CHAPTER 2

2.4.1 Interest-rate rules to be investigated

Given that we do not intend to systematically search the whole parameter

space of ρr, φy, φπ, φe as e.g. Kollmann [22] does (for a reasoning compare

above), the next question to answer is which configurations we want to look

at. We begin by defining two policies which represent suitable reference

points. These two policies are strict domestic inflation targeting and an

unilateral exchange rate peg. When monetary policy seeks to completely

stabilize PPI inflation, it does – obviously – not make allowance for variables

other than domestic inflation. Conversely, when the nominal interest rate

and the nominal money supply are set such that the nominal exchange rate

is fixed at an arbitrary value, monetary policy does not make allowance

for variables other than the exchange rate. In this sense the two reference

points are two (feasible) extreme cases; we call them benchmark policies.

Next, we consider three examples of simple instrument rules as they

have been proposed in the literature. Recall that when monetary policy is

assumed to be active, it follows the generalized Taylor type interest rule

(2.14). Suppose, the target values Y , πx, and etot coincide with the respec-

tive long-run equilibrium values. Rule (2.14) can then be written as

Rt = ρrRt+1 + (1 − ρr)[φyYt + φππx,t + φee

tott

]. (2.17)

In terms of equation (2.17), the baseline specifications of the three rules

are characterized by the parameter configurations ρr, φy, φπ, φe equal

to 0, 0.5, 1.5, 0, 1/3, 0.4, 1.5, 0, and 2/3, 0.1, 2, 0; we label them,

respectively, Rule 1, Rule 2, and Rule 3.29 They roughly correspond to,

respectively, Rule III, I and V in table 1 of Taylor [33].30 We compare the

29This kind of policies are often referred to as flexible inflation targeting. Flexible

inflation targeting gives, on the one hand, leeway to the central bank to respond to

economic shocks and, on the other hand, represents a strong commitment to keeping

inflation low and stable.30To see this, note that it is the relative size of the parameters that matter. Rule I in

table 1 of Taylor [33] e.g. is given by (in Taylor’s terminology)

3.0 · π + 0.8 · y + 1.0 · i−1

where i is the nominal interest rate. Multiplying by 1/3 yields

1.0 · π + 0.26 · y + 0.3 · i−1

Page 67: Macroeconomics of small open economies

2.4. THE ROLE OF EXCHANGE RATE STABILIZATION 57

achievement of these three rules with that of the two benchmark policies and

investigate the robustness of the findings. Finally, we modify the baseline

specifications of the three rules to allow for exchange rate targeting and

explore the consequences of this modification on welfare and performance;

and again, we perform robustness checks.

Output targeting Specification (2.17) assumes that the target values of

output, inflation etc. are equivalent to their respective steady state values.

Note that in the absence of secular growth, the long-run equilibrium output

coincides with the deterministic trend in output. The general rule (2.17),

thus, can be interpreted as describing a reaction to deviations of actual

output from trend.

In the recent literature on policy rules, however, output is typically not

stabilized around trend but around the contemporaneous potential output,

defined as the period t level of output which would obtain under flexible

prices. If we want this to be the case, Rule 1 (in log form) is modified to

ln Rt = ln R + 0.5 · ln Yt + 1.5 · (ln πx,t − ln πx)

where ln Yt ≡ ln Yt − ln Y pott denotes the (log) output gap. In terms of

percentage deviations from the steady-state, we have

Rt = 0.5 · Yt + 1.5 · πx,t.

Note that independent of whether prices are flexible or not, long-run output

remains the same. Hence,

Yt = (ln Yt − ln Y ) −(ln Y pot

t − ln Y)

= Yt − Y pott

where Y pott is the percentage deviation of actual potential output from its

steady state value.31

which is equivalent to

0.6 (1.5 · π + 0.4 · y) + 0.3 · i−1.

31Given a particular state of the economy, what output would obtain under flexible

prices? Suppose, in terms of the linear rational expectations modelling framework utilized

to solve the model, Y pott is a function of exactly the same state variables as Yt. Then,

to compute potential output, we need to import the elasticity values of Yt for q ' 1 and

the policy regime under consideration. Moreover, we set φy = 0 (if prices are flexible,

the output gap is zero for all t anyway).

Page 68: Macroeconomics of small open economies

58 CHAPTER 2

In our policy simulation experiments, we will consider interest rate rules

which involve deviations of actual output from trend as realistic and oper-

ational specifications:32 The central bank takes the fitted trend as a proxy

for the potential output not because it does not know better, but because it

lacks an appropriate current-period measure for the potential output.33 In

this sense, our choice represents a short-cut for allowing for imperfect ob-

servability. The central bank does not have all relevant information available

about the state of the economy and, thus, potential output is measured (or

estimated) with an error. Conversely, we consider it plausible that data on

inflation is readily available and is measured with sufficient accuracy.

Interest smoothing The general rule (2.17) has a dynamic form, that is,

the current interest rate is a weighted average of some desired value which

depends on the state of the economy and the lagged interest rate, where the

relative weights depend on the parameter ρr. The baseline specifications of

Rule 2 and 3 have interest smoothing parameters of 1/3 and 2/3, respec-

tively, whereas there is no interest smoothing in the baseline specification

of Rule 1. One rationale for interest smoothing can be seen in the fact that

whenever the steady state net growth factor of the domestic price level is

larger than minus the steady state net real interest rate, there is a monetary

distortion present in the model.34 This distortion could be eliminated by

following the Friedman rule, which in turn calls for a perfect stabilization

of Rt at 1 (i.e., Rt = 0 for all t). Another rationale can be seen in the

32Another example of a realistic and operational specification would be the prede-

termination of Yt and πx,t, i.e., the assumption that the current inflation and output

deviation from trend cannot be influence by current monetary policy decisions (compare

e.g. Woodford [37], chapter 5).33Rotemberg and Woodford [28], p. 93: ”There are two reasons why such variables [like

the current potential output] may simply be unobservable by the central bank. These

are that some important economic data are collected retrospectively and that even the

data that are collected concurrently need to be processed before their message about the

economy as a whole can be distilled.”34In terms of our model, the condition reads: whenever Ξ > β there is a monetary

distortion present in the model. To make the link to the informal statement, let us define

β ≡ 1/(1 + r) where r ∈ (0,∞] is the agent’s rate of time preference and Ξ ≡ (1 + x)

where x is the net rate of steady state PPI inflation. The inequality Ξ > β approximately

implies x > −r.

Page 69: Macroeconomics of small open economies

2.4. THE ROLE OF EXCHANGE RATE STABILIZATION 59

argument, that in a rather general NNS framework an inertial (or history-

dependent) dynamic response of interest-rate policy to disturbances would

be preferable to a purely forward-looking approach (see e.g. Woodford [37],

chapter 7).

Exchange rate targeting The rule of thumb – expressed in Obstfeld and

Rogoff [27], p. 93, and taken up in Taylor [34] – saying that a substantial

appreciation of the real exchange rate,

erealt ≡

etP∗

t

Pt

,

accompanied by slow output growth and low inflation furnishes a case for

cutting interest rates, represents a concise description of actual monetary

policy behavior of many small and semi-small open economies.

Since only intermediate goods are traded in our model, it might be a

sensible starting point to assume that the central bank targets the terms

of trade, rather than the real exchange rate. In this instance, Rule 1 is

modified to

Rt = 0.5 · Yt + 1.5 · πx,t + φeetott (2.18)

where the parameter on the terms of trade, φe, is negative. Rule 2 and 3

are amended accordingly.35

CPI inflation targeting In the baseline specifications of Rule 1 to 3, the

nominal interest rate reacts to deviations of PPI inflation from its steady

state value. Since it is the intermediate sector where price rigidities and the

implied distortions occur, this might be a sensible starting point. However,

35Given our assumption that P ∗

t = P ∗

x,t, the real exchange rate can be rewritten as

erealt =

etP∗

x,t

pt

=etott

pt

where et ≡ et/Px,t and pt ≡ Pt/Px,t. In terms of percentage deviations from the steady-

state, we have

erealt = et − pt + P ∗

x,t = etott − pt.

For examples of studies which make allowance for the real exchange rate instead of (or

as an alternative to) the terms of trade, compare e.g. Taylor [34] and Kollmann [22].

Page 70: Macroeconomics of small open economies

60 CHAPTER 2

the central bank might prefer to make allowance for CPI inflation rather

than PPI inflation. Why this? Recall that the CPI, defined as

Pt =[ωP

ρρ−1

x,t + (1 − ω)[etP

x,t

] ρρ−1

] ρ−1

ρ

,

includes the price for foreign intermediate goods expressed in the domestic

currency and, consequently, the nominal exchange rate, et. Therefore, tar-

geting CPI inflation can be considered as a way for monetary policy to take

changes in the nominal exchange rate into account. Rule 1 is modified to

Rt = 0.5 · Yt + 1.5 · πt;

Rule 2 and 3 are amended accordingly.

2.4.2 Simulation results

The simulation results are summarized in table 2. In what follows, the

trend output specification for a policy rule is called TREND, the potential

output specification is called GAP, PPI (or domestic) inflation targeting is

abbreviated by DIT, the exchange rate peg is called PEG, CPI inflation

targeting is called CIT, the standard deviation of a variable is denoted by

Std(•), shocks to technology, government consumption, terms of trade, and

foreign demand, respectively, are labelled A-, G-, TOT-, and FD-shocks.

Throughout, the parameter Ξ is set to 1.0.36 Moreover, we focus on reac-

tion coefficient configurations for which a rational expectations equilibrium

exists.37

Benchmark policies

Performance We start by discussing the findings for the two benchmark

policies strict DIT and PEG for the flexible price case. Compared to strict

DIT, the PEG raises output and inflation volatility with respect to all four

36We are safe to make this assumption since at this stage of the work the model is not

calibrated and we may restrict it to small enough shocks.37Determinacy of equilibrium cannot be taken for granted in NNS models; this holds

particularly true when monetary policy is defined in terms of interest-rate rules (see e.g.

Woodford [36]).

Page 71: Macroeconomics of small open economies

2.4. THE ROLE OF EXCHANGE RATE STABILIZATION 61

shocks. Regarding the terms of trade variability, the relative performance is

rather mixed: under the PEG, Std(etot) is clearly raised in the presence of

A-shocks, whereas it is slightly reduced in the presence of G-shocks. With

respect to TOT- and FD-shocks, the differences are negligible. Note that in

the presence of TOT-shocks strict DIT tends to stabilize the interest rate

almost as well as the PEG (the active policy actually performs better). This

is because under a PEG, the gross domestic interest rate is directly propor-

tional to Λ∗

1-shocks and in this sense Λ∗

1-disturbances can be interpreted as

shocks to the world nominal interest rate – to which the domestic nominal

interest rate is equated under the PEG.

Changing the degree of price flexibility from q ' 1 to 0.25 has the

following consequences on the outcome under the PEG. With respect to

all four shocks the volatility of inflation and the terms of trade improves

whereas the volatility of output deteriorates. The increase in Std(Y ) is

particularly significant with respect to TOT-shocks. In contrast, under

strict DIT there are hardly any differences distinguishable as the degree of

price flexibility varies. Note that despite the lower inflation volatility under

sticky prices, Std(π) is still higher under the PEG.

Welfare Given that prices are flexible, moving from strict DIT to the PEG

affects the variances and the respective covariances of C, m, and l, which in

turn affect welfare. However, the differences in terms of welfare are rather

small – from which we conclude that the welfare function is quite flat in

the neighborhood of the set of points we are evaluating it.38 If anything,

the PEG is dominated by strict DIT. The dominance of strict DIT becomes

somewhat more pronounced when prices are sticky: While under strict DIT

welfare is hardly affected as the degree of price flexibility varies, we observe

an albeit small deterioration in welfare under the PEG. This is particularly

true with respect to TOT-shocks.

Baseline rules

Flexible prices Next, we compare the three baseline specifications of

Rule 1, 2, and 3 with each other. If prices are flexible, then, in terms

38As a matter of fact, the differences in terms of welfare are minuscule for all policy

alternatives considered here; this finding is notorious for that kind of studies.

Page 72: Macroeconomics of small open economies

62 CHAPTER 2

of welfare, Rule 1 is invariably dominated by Rule 2 and 3 whereas the

differences between Rule 2 and 3 are small.39 This pronounced discrepancy

in the outcome for Rule 1 on the one hand and Rule 2 and 3 on the other

hand disappears for the GAP-specification. Only regarding A-shocks, Rule

1 yields minimally lower welfare than the two other rules.

A similar pattern emerges from comparing the three simple policy rules

with the two benchmark policies strict DIT and PEG. For the TREND-

specification, Rule 3 delivers virtually the same level of welfare as strict DIT,

Rule 2 yields somewhat higher welfare with respect to A- and G-shocks, and

Rule 1 clearly and invariably involves welfare losses. This pattern practically

disappears for the GAP-specification.

The differences between the TREND-specification of Rule 1 and strict

DIT in terms of volatilities are invariably significant, while the performances

under Rule 2 and 3 (again, the TREND-specifications) come relatively close

to the one under strict DIT. Rule 3 e.g. differs from strict DIT only insofar

as (and this finding does not surprise) the variability of πx is higher; the

variability of π, on the other hand, is slightly smaller with respect to A- and

FD-shocks and slightly higher for G- and TOT-shocks. From the differences

in terms of performance between Rule 1 on the one side and Rule 2 and 3

on the other side, we conclude that under flexible prices interest smoothing

in combination with a more aggressive reaction to inflation pays off in terms

of both inflation and output variability.

Sticky prices If prices are sticky, the differences across the rules in terms

of welfare decrease by about one half to two thirds. But still, Rule 1 is

dominated by Rule 2 and 3. Interestingly, the ranking between Rule 2 and

3 gets now reversed: Rule 2 yields slightly more welfare with respect to all

shocks but FD-shocks.

In terms of volatilities, Rule 1 performs better in the sticky price case

compared to the flex price case, and this with respect to all shocks;40 the

only exception is Std(etot) for which we observe a move in the opposite

39Rule 3 yields a little bit more welfare with respect to A-shocks and insignificantly

less welfare with respect to TOT-shocks.40This does not come at a surprise to us since the original Taylor [32] rule has been

proposed for a sticky price environment.

Page 73: Macroeconomics of small open economies

2.4. THE ROLE OF EXCHANGE RATE STABILIZATION 63

direction in the case of G- and TOT-shocks. Under all three rules output

variability with respect to A-shocks is smaller than under strict DIT. Rule

2 invariably performs better than strict DIT in terms of Std(Yt). The same

is true for Rule 3 in terms of Std(πt). In comparison to Rule 2 and 3, an

adoption of Rule 1 induces higher inflation variability (both PPI and CPI

inflation). With respect to output variability, however, Rule 1 outperforms

Rule 3.

Varying the parameter φy A weaker response to output deviations from

trend in Rule 1 relative to the respective baseline value invariably leads to

higher welfare, whereby the biggest benefits arise with respect to A-shocks.

Rule 2 and 3 are less sensitive to lowering φy; the observed welfare gains

or losses are negligible. In terms of performance, a smaller φy-parameter in

Rule 1 brings about somewhat more variability in output deviations from

trend and significantly less volatility in inflation as well as in the interest

rate.

A stronger response to output from trend in Rule 1 leads to welfare losses

with respect to all four shocks (the biggest losses can be observed regarding

A-shocks). In the case of Rule 2, a higher φy leads to welfare losses with

respect to A-shocks while the outcome regarding the remaining three shocks

is mainly unaffected. In the case of Rule 3, raising φy seems to have no

effect whatsoever. If output is measured with respect to deviations from

its potential level, Rule 1 gets fairly robust in that varying the parameter

φy has very little consequences for welfare and performance. If anything, a

higher φy leads to modest welfare gains.

Varying the parameter ρr A modest rise in the smoothness parame-

ter ρr of Rule 1 leads to substantial gains in welfare. In fact, introducing

smoothing in Rule 1 leads to the largest welfare gains we have found among

all investigated modifications. In terms of performance, a smoothing pa-

rameter of 1/3 brings about less variability not only for inflation but also

for output – and this with respect to all types of shocks – with the only

exception that with respect to TOT-shocks, Std(Y ) increases. Again, this

pattern disappears when it is assumed that the current potential output

level is observable or can be measured accurately. In this event, the only

Page 74: Macroeconomics of small open economies

64 CHAPTER 2

gains in terms of welfare from interest smoothing can be observed with

respect to A-shocks. However, these gains are much smaller than for the

TREND-specification.

We conclude that the differences in terms of welfare and performance

between the baseline specifications of Rule 1 on the one hand and Rule 2

and 3 on the other hand mainly stem from the absence of interest smoothing

in Rule 1.

Terms of trade targeting

Suppose, the central bank targets the terms of trade (in addition to devi-

ations of PPI inflation and output from trend). Lowering φe to −0.1 and

further to −0.2 and finally to −0.3 in Rule 1 yields increasingly higher wel-

fare compared to the baseline specification with respect to A-shocks and

(to a much lower extent) FD-shocks. The same experiment leads to welfare

losses with respect to G- and TOT-shocks. Without knowing more about

the relative size of each type of shock we cannot decide how much terms of

trade targeting the welfare maximizing central bank should opt for.

In the case of Rule 2 and 3, lowering φe (relative to the baseline value)

leads to shifts in welfare of much smaller magnitudes. In particular, going

from φe = 0 to -0.1 has barely any effects on welfare. Only with respect

to A-shocks we can detect a minor shift (for Rule 2 a small gain and for

Rule 3 a small loss; note that the small gain turns into a loss for the GAP-

specification).

In the case of Rule 1, a moderate form of terms of trade targeting leads

to lower inflation and interest rate volatilities with respect to A- and FD-

shocks, while the same volatilities are higher with respect to G- and TOT-

shocks. Output variability practically remains unaffected with respect to

A-, G-, and FD-shocks and improves somewhat with respect to TOT-shocks.

In the case of Rule 2, the same experiment leads to a similar pattern – with

the difference that output variability now is clearly negatively affected and

that the effects on Std(R) are much smaller. Letting Rule 3 respond to

the terms of trade leads to moderate effects on output volatility (Std(Y )

improves with respect to A- and FD-shocks and deteriorates with respect to

G- and TOT-shocks) while the effects on inflation volatility are negligible.

Page 75: Macroeconomics of small open economies

2.4. THE ROLE OF EXCHANGE RATE STABILIZATION 65

Varying the parameter φy How are the numerical results under terms of

trade targeting affected by decreasing the coefficient on deviations of output

from trend? When we compare Rule 1 for the parameter configuration

ρr, φy, φπ, φe = 0, 0.1, 1.5, − 0.3 with 0, 0.5, 1.5, − 0.3 we observe

a deterioration in terms of welfare with respect to A-shocks (considerably)

and FD-shocks (marginally) and an improvement with respect to G-shocks

(considerably) and TOT-shocks (marginally). Output volatility is hardly

affected by lowering φy (we only observe a small increase with regard to

A-shocks). However, we observe a pronounced effect on inflation volatility

with respect to A-shocks (deterioration) and G-shocks (improvement). The

effects of lowering φy on Rule 2 and 3 in terms of welfare are small.41 If

anything, we observe a moderate loss in the case of Rule 2 and with respect

to A-shocks.

How are the numerical results affected by increasing the coefficient on

deviations of output from trend? In the case of Rule 1, the parameter con-

figuration 0, 1.5, 1.5, − 0.3 leads to indeterminacy. One way to get rid

of indeterminacy is to raise the smoothness parameter from 0 to 0.1 (which

admittedly biases the outcome towards welfare gains). We observe welfare

gains with respect to A-shocks (considerably) and FD-shocks (marginally).

The outcome for Rule 2, again, seems to be fairly stable regarding shifts

in φy, while the outcome for Rule 3 is slightly improved with respect to A-

and G-shocks and barely affected with respect to TOT- and FD-shocks.42

Increasing the φy-parameter in Rule 3 in conjunction with a very mod-

est terms of trade targeting also pays off in terms of output and inflation

volatility.

CPI inflation targeting

Suppose the central bank targets CPI inflation rather than PPI inflation.

For all three rules, the CIT outcome is going to be compared to the outcome

41For Rule 2, we compare the parameter configuration 1/3, 0.1, 1.5, − 0.1

with 1/3, 0.4, 1.5, − 0.1. For Rule 3, we compare the parameter configuration

2/3, 0, 2, − 0.1 with 2/3, 0.1, 2, − 0.1.42For Rule 2, we compare the parameter configuration 1/3, 1.5, 1.5, − 0.1

with 1/3, 0.4, 1.5, − 0.1. For Rule 3, we compare the parameter configuration

2/3, 0.5, 2, − 0.1 with 2/3, 0.1, 2, − 0.1.

Page 76: Macroeconomics of small open economies

66 CHAPTER 2

for the respective baseline specification (which involves DIT). In the case

of Rule 1, CIT leads to modest welfare gains with respect to A- and FD-

shocks and brings about losses (modest as well) with respect to G- and

TOT-shocks. Welfare is invariably lower in the case of Rule 2 and 3, albeit

these differences are minuscule. Lowering φy in Rule 1 invariably improves

welfare in almost the same manner as allowing for modest interest smoothing

does, while nothing is gained by raising φy. As for the case of PPI targeting,

the results for Rule 2 and 3 are fairly robust regarding moderate shifts in

φy. If anything, we can observe welfare gains with respect to A-shocks when

lowering the parameter on output in Rule 2 and when increasing the same

parameter in Rule 3.

In the case of Rule 1, CIT leads to a modest improvement of output

and inflation volatility with respect to A- and FD-shocks and an equally

modest deterioration with respect to G- and TOT-shocks. For Rule 2 and

3 the differences in terms of performance are even smaller (and in fact

negligible).

Strict CTI Finally, we undertake an assessment of strict CTI. If prices

are flexible, there is hardly any difference in terms of welfare between strict

DIT and strict CIT, respectively; both versions of active policies dominate

the PEG. The only difference we are aware of is that with respect to A-

shocks, strict CIT achieves slightly more welfare. However, if prices are

sticky, strict CIT invariably yields less welfare than strict DIT targeting.

The biggest differences can be observed with respect to TOT-shocks.

2.5 Conclusions

In this paper, we have compared simple monetary policy rules in the frame-

work of a hypothetical small open economy with optimizing agents and

monopolistic competition in intermediate product markets. The role the

exchange rate is playing in manipulating the policy instrument has been

explored in the presence of a moderate degree of price stickiness, perfect

exchange rate pass-through into import prices, and a small number of do-

mestic and foreign structural shocks. We found that a moderate form of

exchange rate targeting in the original Taylor rule induces higher perfor-

Page 77: Macroeconomics of small open economies

2.5. CONCLUSIONS 67

mance with respect to productivity and foreign demand shocks and lower

performance with respect to government and terms-of-trade shocks; without

knowing more about the relative size of each type of shock we cannot decide

whether directly reacting to the exchange rate is preferable. For rules with

a moderate to considerable degree of persistence and a larger relative weight

on inflation, the outcome was practically unaffected by the inclusion of an

exchange rate target.

There are some obvious limitations to the analysis conducted in this

paper which may indicate directions for future work. Some investigation

of the robustness of the findings was presented here, but more needs to be

done. One may want to vary key parameters of the model (such as the

trade elasticity and the degree of openness) and/or key assumptions such

as the degree of exchange rate pass-through – and assess the numerical

sensitivity of the results. One may also want to calibrate the model (or

an extended version of it) to quarterly data of a real-world, approximately

small open economy like the one of Switzerland – an exercise which would

involve, among others, getting an idea of the relative size of the structural

shocks. Together with the adoption of Sims’ [30] non-linear solution method

this would, after all, allow one to evaluate unconditional welfare and make

policy recommendations.

Page 78: Macroeconomics of small open economies

68 CHAPTER 2

2.A Stationary representation

2.A.1 Change in notation and useful simplifications

Before presenting the system of equations which describes the equilibrium

of the economy, we undertake a change in notation and some useful sim-

plifications. In what follows, we write Yt for Y (st) etc. Moreover, we let

Et • be a function which takes the expected value of the term inside the

curly bracket, based on information available in t. The optimal pricing rule

can then be rewritten as follows43

px,t =1

θ

∑∞

τ=0 βτ (1 − q)τ (Ξτ )1

θ−1 Et

Λ1,t+τ

Λ1,tP

1

1−θ

x,t+τPt+τCm,t+τXt+τ

∑∞

τ=0 βτ (1 − q)τ (Ξτ )θ

θ−1 Et

Λ1,t+τ

Λ1,tP

1

1−θ

x,t+τXt+τ

.

Both the nominator and the denominator of the pricing rule can be ex-

pressed in terms of expectational difference equations. We end up with the

following rather compact expression for the optimal price

px,t =1

θ

S1,t

S2,t

where

S1,t ≡ P1

1−θ

x,t PtCm,tXt + β (1 − q) Ξ1

θ−1 Et

Λ1,t+1

Λ1,t

S1,t+1

and

S2,t ≡ P1

1−θ

x,t Xt + β (1 − q) Ξθ

θ−1 Et

Λ1,t+1

Λ1,t

S2,t+1

.

Next, the FOC for lt from the household’s maximization problem44

(1 − ν) Ψν(1−σ)t l

(1−ν)(1−σ)−1t = Λ2,t,

43Before changing notation we multiply the nominator and the denominator of equa-

tion (2.7) by π(st+1

∣∣ st)/π(st+1

∣∣ st)

and then substitute P b(st+1

∣∣ st)/π(st+1

∣∣ st)

for

βΛ1

(st+1

)/Λ1 (st).

44The variables Λ1,t, Λ2,t and Λ3,t are the (nominal) shadow prices on the first, sec-

ond, and third constraint to the household’s maximization problem (compare Subsection

2.2.4).

Page 79: Macroeconomics of small open economies

2.A. STATIONARY REPRESENTATION 69

and for ht,

Λ1,tPtWt = Λ2,t,

are combined to

Λ1,tPtWt = (1 − ν) Ψν(1−σ)t l

(1−ν)(1−σ)−1t .

Also, the FOC for It,

Λ3,t = Λ1,tPt

[1 − φ

(It

Kt−1

− δ

)]−1

,

and for Kt,45

Λ3,t = βEt Λ1,t+1Pt+1zt+1

+Λ3,t+1

2

(It+1

Kt

)2

−φ

2δ2 + 1 − δ

],

are combined to

Λ1,tPtΦ−1t = βEt

Λ1,t+1Pt+1

(zt+1 + Φ−1

t+1

2

(It+1

Kt

)2

−φ

2δ2 + 1 − δ

])

where

Φt = 1 − φ

(It

Kt−1

− δ

).

Finally, we take notice of the fact that variable Bt shows up in exactly one

optimality condition, namely the FOC for the shadow price Λ1,t. Since this

would have been the equation which determined Bt, that particular FOC

(i.e., the budget constraint) can be dropped from the system of equilibrium

conditions (compare Subsection 2.2.6).

45Observe that

φ

(It+1

Kt

− δ

)It+1

Kt

−φ

2

(It+1

Kt

− δ

)2

+ (1 − δ) =φ

2

(It+1

Kt

)2

−φ

2δ2 + 1 − δ.

Page 80: Macroeconomics of small open economies

70 CHAPTER 2

2.A.2 Equilibrium conditions

We end up with a system of 22 equations plus policy rule (2.14) which simul-

taneously hold at all points in time, in a total of 23 endogenous variables:46

Final good producers:

Y : Yt =[ω1−ρ

(Xd

t

)ρ+ (1 − ω)1−ρ

(Xf

t

)ρ]1/ρ

(2.19)

Xd : Xdt = ω

(Px,t

Pt

) 1

ρ−1

Yt (2.20)

Xf : Xft = (1 − ω)

(etP

x,t

Pt

) 1

ρ−1

Yt (2.21)

Intermediate goods producers:

h : Xt = AtKαt l1−α

t (2.22)

Cm : ztKt = αCm,tXt (2.23)

W : Wtht = (1 − α) Cm,tXt (2.24)

p : px,t =1

θ

S1,t

S2,t

(2.25)

S1 : S1,t = P1

1−θ

x,t PtCm,tXt + β (1 − q) Ξ1

θ−1 Et

Λ1,t+1

Λ1,t

S1,t+1

(2.26)

S2 : S2,t = P1

1−θ

x,t Xt + β (1 − q) Ξθ

θ−1 Et

Λ1,t+1

Λ1,t

S2,t+1

(2.27)

Evolution of PPI:

Px : Pθ

θ−1

x,t = qpθ

θ−1

x,t + (1 − q) Ξθ

θ−1

t−1 Pθ

θ−1

x,t−1 (2.28)

46There are fifteen endogenous quantities (C , l, h, I, K, M , Y , X, Xd, Xf , Cm, Ψ,

Φ, S1, S2) and eight endogenous prices – including the policy variable and the shadow

price in the household’s budget constraint (z, W , P , Px, px, e, R, Λ1).

Page 81: Macroeconomics of small open economies

2.A. STATIONARY REPRESENTATION 71

Representative household:

C : Λ1,tPt = νCη−1t Ψ

ν(1−σ)−ηt l

(1−ν)(1−σ)t (2.29)

Ψ : Ψt =

[Cη

t + ζ

(Mt

Pt

)η] 1

η

(2.30)

M : ζ

(Mt

PtCt

)η−1

=(1 − R−1

t

)(2.31)

P : Λ1,tPtWt = (1 − ν) Ψν(1−σ)t l

(1−ν)(1−σ)−1t (2.32)

Λ :Λ1,t

Rt

= βEt Λ1,t+1 (2.33)

z : Λ1,tPtΦ−1t = (2.34)

βEt

Λ1,t+1Pt+1

(zt+1 + Φ−1

t+1

2

(It+1

Kt

)2

−φ

2δ2 + 1 − δ

])

Φ : Φt = 1 − φ

(It

Kt−1

− δ

)(2.35)

K : Kt = It −φ

2

(It

Kt−1

− δ

)2

Kt−1 + (1 − δ) Kt−1 (2.36)

l : 1 = lt + ht (2.37)

Explicit market clearing conditions:

X : Xt = Xdt + Xd∗

t (2.38)

I : Yt = Ct + It + Gt (2.39)

International asset market:

e : et = κΛ∗

1,t

Λ1,t

. (2.40)

Notice: At, Gt, U∗

C,t, P ∗

x,t, and Xd∗t are exogenous variables. Moreover,

equations (2.19), (2.20), and (2.21) together ensure that

Pt =[ωP

ρρ−1

x,t + (1 − ω)(etP

x,t

) ρρ−1

] ρ−1

ρ

.

Page 82: Macroeconomics of small open economies

72 CHAPTER 2

2.A.3 Deflating the system

In the presence of sustained inflation (Ξ > 1), the nominal variables in

the system of equilibrium conditions are non-stationary. We can get rid of

this non-stationarity by deflating the system appropriately.47 The choice of

the deflator is arbitrary; we prefer Px. We end up with the following 23

equations plus policy rule (2.14) in 24 endogenous variables:

Final good producers:

Y : Yt =[ω1−ρ

(Xd

t

)ρ+ (1 − ω)1−ρ

(Xf

t

)ρ]1/ρ

Xd : Xdt = ω

(1

pt

) 1

ρ−1

Yt

Xf : Xft = (1 − ω)

(etP

x,t

pt

) 1

ρ−1

Yt

where pt ≡ Pt/Px,t and et ≡ et/Px,t.

Intermediate goods producers:

h : Xt = AtKαt h1−α

t

Cm : ztKt = αCm,tXt

W : Wtht = (1 − α) Cm,tXt

px : px,t =1

θ

s1,t

s2,t

s1 : s1,t = ptCm,tXt + β (1 − q) Ξ1

θ−1 Et

λ1,t+1

λ1,t

π1

1−θ

x,t+1s1,t+1

s2 : s2,t = Xt + β (1 − q) Ξθ

θ−1 Et

λ1,t+1

λ1,t

πθ

1−θ

x,t+1s2,t+1

where px,t ≡ px,t/Px,t, s1,t ≡ S1,tP−

2−θ1−θ

x,t , s2,t ≡ S2,t/P1

1−θ

x,t , and λ1,t ≡ Λ1,tPx,t.

47Note that equations (2.35), (2.37), (2.36), ( 2.19), (2.20), (2.38), (2.23), (2.24), and

( 2.39) are stated in terms of either (stationary) real variables or (likewise stationary)

relative prices.

Page 83: Macroeconomics of small open economies

2.A. STATIONARY REPRESENTATION 73

Representative household:

C : λ1,tpt = νCη−1t Ψ

ν(1−σ)−ηt l

(1−ν)(1−σ)t

Ψ : Ψt =

[Cη

t + ζ

(mt

pt

)η] 1

η

m : ζ

(mt

ptCt

)η−1

=(1 − R−1

t

)

p : λ1,tptWt = (1 − ν) Ψν(1−σ)t l

(1−ν)(1−σ)−1t

λ1 : λ1,t = βRtEt

λ1,t+1

πx,t+1

z : λ1,tpt1

Φt

= βEt

λ1,t+1pt+1

(zt+1 +

1

Φt+1

2

(It+1

Kt

)2

−φ

2δ2 + 1 − δ

])

Φ : Φt = 1 − φ

(It

Kt−1

− δ

)

K : Kt = It −φ

2

(It

Kt−1

− δ

)2

Kt−1 + (1 − δ) Kt−1

l : lt = 1 − ht

where mt ≡ Mt/Px,t.

Explicit market clearing conditions:

I : Yt = Ct + It + Gt

X : Xt = Xdt + Xd∗

t

International asset market:

e : κΛ∗

1,t

λ1,t

= et

Page 84: Macroeconomics of small open economies

74 CHAPTER 2

PPI and CPI inflation:

πx : 1 = qpθ

θ−1

x,t + (1 − q) Ξθ

θ−1

(1

πx,t

) θθ−1

π : πx,t+1 =pt

pt+1

πt+1

where πx,t+1 ≡ Px,t+1/Px,t and πt+1 ≡ Pt+1/Pt.

Note that equation (2.28) (the evolution of the PPI) now is expressed in

terms of PPI inflation, πx,t. Moreover, πx,t now shows up in the equations

which determine s1,t, s2,t, and λi,t. Finally, PPI inflation and CPI inflation

are related by πx,t+1 = (pt/pt+1)πt+1.

2.B Non-stochastic steady state

In the non-stochastic steady state we have sustained inflation (π = πx = Ξ)

and, thus, all prices grow at the same rate. It follows that (P/Px =) p =

(p/Px =) px = 1.

From λ1 = βRλ1π−1x , we get

R =Ξ

β.

From pρ

1−ρ = ω + (1 − ω) eρ

ρ−1

x P∗

ρρ−1

x , we get

e =1

P ∗x

.

From s1 = pCmX + β (1 − q) Ξ1

θ−1λ1

λ1π

1

1−θx s1, we get

CmX

s1

= 1 − β (1 − q) .

Similarly, from s2 = X + β (1 − q) Ξθ

θ−1λ1

λ1π

θ1−θx s2, we get

X

s2

= 1 − β (1 − q) .

Page 85: Macroeconomics of small open economies

2.B. NON-STOCHASTIC STEADY STATE 75

It follows thats1

s2

= Cm.

From px = 1θ

s1

s2together with the previous results, we get

Cm = θ

which is the expected result.

Transforming K = I− φ2

(IK− δ)2

K+(1 − δ) K yields a quadratic equation

in I/K, given by

0 = −1

(I

K

)2

+ (1 + δφ)I

K−

1

2δ2φ − δ.

The two solutions for the ratio I/K are δ and δφ+2φ

, respectively. Here, we

focus on the first oneI

K= δ.

It follows that

z =1 − β (1 − δ)

β.

From z = CmαXK

, we get

K

h=( z

αθ

) 1

α−1

=

(1 − β (1 − δ)

αβθ

) 1

α−1

.

Let us assume that Y = X and that A = 1. It follows that

Y

K=

1 − β (1 − δ)

αβθ

and, consequently, that

Y =

(Y

K

) αα−1

h.

We are now in the position to derive an expression for C/Y

C

Y= 1 −

I

Y−

G

Y

Page 86: Macroeconomics of small open economies

76 CHAPTER 2

where G/Y is steady state government expenditures in percent of steady

state output and whereI

Y=

I

K

K

Y.

From Xf = (1 − ω) e1

ρ−1

(P ∗

x

) 1

ρ−1 p−1

ρ−1 Y and previous results, we get

Xf = (1 − ω) Y.

Also, from Xd = ω(

1p

) 1

ρ−1

Y , we get

Xd = ωY.

From the assumption that Y = X and previous results, we get

Xd∗

Y

(=

Xd∗

X

)= 1 − ω.

Also,ωY

Y

(=

Xd

X

)= ω.

From ζ(

mC

)η−1= 1 − R−1, we get

m

Y=

C

Y

(R − 1

ζR

) 1

η−1

.

And, from Ψη = Cη + ζmη we get

Ψ

Y=

C

Y

[1 + ζ

(m/Y

C/Y

)η] 1

η

.

Suppose the time devoted to labor as a fraction of total endowment is given.

The implied CES weight ν can be computed as follows. Dividing the FOC

for C,

νCη−1Ψν(1−σ)−η (1 − h)(1−ν)(1−σ) = λ1,

by the FOC for h,

(1 − ν) Ψν(1−σ) (1 − h)(1−ν)(1−σ)−1 = λ1W,

Page 87: Macroeconomics of small open economies

2.C. FIRST-ORDER APPROXIMATION 77

provides us withνCη−1Ψ−η

(1 − ν) (1 − h)−1 =1

W.

Substituting out for Ψ = (Cη + ζmη)1

η yields

ν

1 − ν

Cη(1 − h)

C(Cη + ζmη)=

1

W.

An expression for W can be found from the condition

W =(1 − α) θY

h.

Substituting out for W yields

Y

C

1

h

(1 − h)θ (1 − α)[1 + ζ

(mC

)η] =1 − ν

ν.

Substituting out for m/C taking the inverse yields

ν =

(1 − h)θ (1 − α)[

1 + ζ1

1−η

(R−1

R

) ηη−1

]CY

h+ 1

−1

.

2.C First-order approximation

2.C.1 Linear system

Sofar, we have presented the system of equations which describes the equilib-

rium and have found the non-stochastic steady state of that system. Next,

we consider the first-order approximation to the equilibrium conditions (in

terms of percentage deviations from the steady state):

Final good producers:

Y : Yt = ωXdt + (1 − ω) Xf

t

Xd : Xdt = −

(1

ρ − 1

)pt + Yt

Xf : Xft =

(1

ρ − 1

)et +

(1

ρ − 1

)P ∗

x,t −

(1

ρ − 1

)pt + Yt

Page 88: Macroeconomics of small open economies

78 CHAPTER 2

Intermediate goods producers:

h : Xt = At + αKt + (1 − α) ht

Cm : zt = Cm,t + Xt − Kt

W : Wt = Cm,t + Xt − ht

px : px,t = s1,t − s2,t

s1 : s1,t =CmX

s1

[Cm,t + Xt

]

+β (1 − q)

[Et

λ1,t+1

− λ1,t +

(1

1 − θ

)Et πx,t+1 + Et s1,t+1

]

s2 : s2,t =X

s2

Xt

+β (1 − q)

[Et

λ1,t+1

− λ1,t +

1 − θ

)Et πx,t+1 + Et s2,t+1

]

Representative household:

C : λ1,t + pt = (η − 1) Ct + [ν (1 − σ) − η] Ψt + (1 − ν) (1 − σ) lt

Ψ :

Y

Ψt =

(C

Y

Ct + ζ(m

Y

mt − ζ(m

Y

pt

m : (R − 1) (η − 1)[mt − pt − Ct

]= Rt

p : λ1,t + pt + Wt = ν (1 − σ) Ψt + [(1 − ν) (1 − σ) − 1] lt

λ1 : λ1,t − Et

λ1,t+1

= Rt − Et πx,t+1

z : λ1,t + pt − Φt = Et

λ1,t+1

+ Et pt+1 + [1 − β (1 − δ)] Et zt+1

−β (1 − δ) Et

Φt+1

+ βφδ2Et

It+1

− βφδ2Kt

Φ : Φt = −φδIt + φδKt−1

K : Kt = δIt + (1 − δ) Kt−1

l : lt = −h

lht

Page 89: Macroeconomics of small open economies

2.C. FIRST-ORDER APPROXIMATION 79

Explicit market clearing conditions:

I : Yt =C

YCt +

I

YIt +

G

YGt

X : Xt = ωXdt + (1 − ω) Xd∗

t

International capital markets:

e : Λ∗

1,t − λ1,t = et

PPI and CPI inflation:

πx : πx,t =q

(1 − q)px,t

π : πx,t = pt−1 − pt + πt.

Policy rule: If monetary policy is assumed to be active, it follows the

generalized Taylor type interest rule (2.14). Suppose, the target values Y ,

πx, and etot coincide with the respective long-run equilibrium values. Rule

(2.14) can then be rewritten as

Rt = ρrRt+1 + (1 − ρr)[φyYt + φππx,t + φee

tott

]

where

etott = et + P ∗

x,t.

Exogenous variables: The variable At is assumed to follow the stochas-

tic AR(1) process

ln At = ρa ln At−1 + (1 − ρa) ln A + εa,t

where ρa ∈ [0, 1), A denotes the unconditional mean of At, and εa,t is

i.i.d.(0, σ2

εa

). This process can be rearranged as follows

(ln At − ln A

)= ρa

(ln At−1 − ln A

)+ εa,t

Page 90: Macroeconomics of small open economies

80 CHAPTER 2

from which we get

A : At = ρaAt−1 + εa,t.

Similarly, we have

G : Gt = ρgGt−1 + εg,t

Λ∗ : Λ∗

1,t = ρlΛ∗

1,t−1 + εl,t

Xd∗ : Xd∗t = ρxX

d∗t−1 + εx,t

P ∗

x : P ∗

x,t = ρpP∗

x,t−1 + εp,t

where ρg, ρµ, ρl, ρp, ρx ∈ [0, 1). All driving forces are i.i.d. with zero mean

and constant variance (and uncorrelated among each others, of course).

2.C.2 Digression: The New Phillips curve

Substituting out for s1 and s2 in the equation for px yields the following

expectational difference-equation

px,t = [1 − β (1 − q)] Cm,t + β (1 − q) Et πx,t+1 + β (1 − q) Et px,t+1 .

Replacing px,t by 1−qq

πx,t provides us with the familiar New Phillips curve

πx : πx,t = [1 − β (1 − q)]q

1 − qCm,t + βEt πx,t+1 .

2.D Exchange rate peg

Suppose the central bank pegs the domestic currency unilaterally and cred-

ibly to the world currency. There is just one way for the central bank of

the small open economy to effectively fix the nominal exchange rate: it has

to choose the same steady state inflation rate as the monetary authority of

the rest of the world does, that is, it has to choose Ξ = Ξ∗.

For Ξ = Ξ∗ > 1, the asset market equilibrium condition can be rewritten as

et = κΛ∗

t

Λ1,t

= κλ∗

t

λ1,t

Px,t

P ∗x,t

Page 91: Macroeconomics of small open economies

2.D. EXCHANGE RATE PEG 81

where the ratio Px,t/P∗

x,t is stationary; the original system of equations has

then to be deflated in a meaningful way.

However, things get simpler when we assume that Ξ = Ξ∗ = 1 (which

is the reason why we prefer this alternative). In this event, the original

(non-deflated) system of equilibrium conditions becomes stationary. The

non-deflated linearized system can then be modified as follows: We skip the

Taylor rule (since monetary policy becomes passive) and let et = 0 for all t

(since et is fixed at an arbitrary value). It follows that Λ1 becomes a control

variable, given by Λ1,t = Λ∗

1,t. R and M adjust endogenously. Moreover,

equation

Λ1,t − Et

Λ1,t+1

= (1 − ρl) Λ∗

1,t = Rt

determines R while M is determined via the money demand equation. Note

that when the Λ∗

1-shock is silent, Rt = et = 0.

Page 92: Macroeconomics of small open economies

Bibliography

[1] Backus, David K., Patrick J. Kehoe, and Finn E. Kydland (1995),

International business cycles: Theory and evidence, in: Thomas F.

Cooley, ed., Frontiers of Business Cycle Research, Princeton University

Press

[2] Ball, Laurance (1999), Policy rules for open economies, in: John B.

Taylor, ed., Monetary policy rules, NBER Studies in Business Cycles

[3] Baxter, Marianne, and Robert G. King (1993), Fiscal policy in general

equilibrium, American Economic Review 83(3), 315-334

[4] Blanchard, Olivier J., and Charles M. Kahn (1980), The solution of lin-

ear difference models under rational expectations, Econometrica 48(5),

1305-1312

[5] Calvo, Guillermo A. (1983), Staggered prices in a utility-maximizing

framework, Journal of Monetary Economics 12(3), 383-398

[6] Chari, V.V., Patrick J. Kehoe, and Ellen R. McGrattan (2000), Sticky

price models of the business cycle: can the contract multiplier solve

the persistence problem?, Econometrica 68(5), 1151-1179

[7] Chari, V.V., Patrick J. Kehoe, and Ellen R. McGrattan (2000), Can

sticky price models generate volatile and persistent real exchange

rates?, NBER Working Paper No. w7869

[8] Clarida, Richard, and Jordi Gali (1994), Sources of real exchange-rate

fluctuations: how important are nominal shocks?, Carnegie-Rochester

Conference Series on Public Policy 41, 1-65

Page 93: Macroeconomics of small open economies

BIBLIOGRAPHY 83

[9] Clarida, Richard, Jordi Gali, and Mark Gertler (1999), The science of

monetary policy: a New Keynesian perspective, Journal of Economic

Literature 37(4), 1661-1707

[10] Collard, Fabrice, and Harris Dellas (2001), Price rigidity and the selec-

tion of the exchange rate regime, Mimeo

[11] Collard, Fabrice, and Harris Dellas (2002), Monetary policy rules and

exchange rate targeting, Mimeo

[12] Collard, Fabrice, and Harris Dellas (2002), Technology shocks and em-

ployment, CEPR Discussion Paper No. 3680

[13] Collard, Fabrice, and Harris Dellas (2003), Inflation targeting, Mimeo

[14] Cooley, Thomas F., and Edward C. Prescott (1995), Economic growth

and business cycles, in: Thomas F. Cooley, ed., Frontiers of Business

Cycle Research, Princeton University Press

[15] Gali, Jordi, and Tommaso Monacelli (2002), Monetary policy and ex-

change rate volatility in a small open economy, NBER Working Paper

No. 8905

[16] Ghironi, Fabio (2000), Alternative monetary rules for a small open

economy: the case of Canada, Working Paper, Boston College

[17] Goodfriend, Marvin, and Robert G. King (1997), The new neoclassical

synthesis and the role of monetary policy, in: Ben S. Bernanke and

Julio Rotemberg, NBER Macroeconomics Annual 1997

[18] King, Robert G., Charles I. Plosser, and Sergio T. Rebelo (1988), Pro-

duction, growth, and business cycles, Journal of Monetary Economics

21, 195-232 and 309-341

[19] King, Robert G., and Alexander L. Wolman (1996), Inflation targeting

in a St. Louis model of the 21st century, NBER Working Paper No.

5507

[20] King, Robert G., and Sergio T. Rebelo (1999), Rescuscitating real busi-

ness cycles, in: John B. Taylor and Michael Woodford, eds., Handbook

of Macroeconomics Vol. 1, Elsevier Science

Page 94: Macroeconomics of small open economies

84 CHAPTER 2

[21] Kollmann, Robert (1997), The exchange rate in a dynamic-optimizing

current account model with nominal rigidities: aA quantitative inves-

tigation, IMF Working Paper WP/97/7

[22] Kollmann, Robert (2002), Monetary policy rules in the open economy:

effects on welfare and business cycles, Journal of Monetary Economics

49(5), 989-1015

[23] Ljunqvist, Lars, and Thomas J. Sargent (2000), Recursive macroeco-

nomic theory, The MIT Press

[24] McCallum, Bennett T. (1999), Issues in the design of monetary policy

rules, in: John B. Taylor and Michael Woodford, eds., Handbook of

Macroeconomics Vol. 1, Elsevier Science

[25] Monacelli, Tommaso (2000), Into the Mussa puzzle: monetary policy

regimes and real exchange rate in a small open economy, Working Pa-

per, Boston College

[26] Obstfeld, Maurice, and Kenneth Rogoff (1995), Exchange rate dynam-

ics redux, Journal of Political Economy 103(3), 624-660

[27] Obstfeld, Maurice, and Kenneth Rogoff (1995), The mirage of fixed

exchange rates, Journal of Economic Perspectives 9(4), 73-96

[28] Rotemberg, Julio J., and Michael Woodford (1999), Interest rate rules

in an estimated sticky price model, in: John B. Taylor, ed., Monetary

policy rules, NBER Studies in Business Cycles

[29] Rotemberg, Julio J., and Michael Woodford (1999), The cyclical be-

havior of prices and costs, in: John B. Taylor and Michael Woodford,

eds., Handbook of Macroeconomics Vol. 1, Elsevier Science

[30] Sims, Christopher A. (2003), Second order accurate solution of discrete

time dynamic equilibrium models, Working paper

[31] Svensson, Lars E.O. (2000), Open-economy inflation targeting, Journal

of International Economics 50(1), 155-183

Page 95: Macroeconomics of small open economies

BIBLIOGRAPHY 85

[32] Taylor, John B. (1993), Discretion versus policy rules in practice,

Carnegie-Rochester Conference Series on Public Policy 39, 195-214

[33] Taylor, John B. (1999), Introduction, in: John B. Taylor, ed., Monetary

policy rules, NBER Studies in Business Cycles

[34] Taylor, John B. (2001), The role of the exchange-rate in monetary-

policy rules, American Economic Review, AEA Papers and Proceedings

91(2), 263-267

[35] Walsh, Carl E. (2003), Monetary theory and policy, second edition,

The MIT Press

[36] Woodford, Michael (2001), The Taylor rule and optimal monetary pol-

icy, American Economic Review, AEA Papers and Proceedings 91(2),

232-237

[37] Woodford, Michael (2003), Interest and prices: foundations of a theory

of monetary policy, Princeton University Press

Page 96: Macroeconomics of small open economies

86 TABLES CHAPTER 2

Tables Chapter 2

Table 1a+b: Elasticities of key variables

Table 2a-e: Performance and welfare of alternative policy regimes

Page 97: Macroeconomics of small open economies

Policy regime: Constant money growth rate

Flexible prices (q =1), no capital adjustment costs (φ =1)

A G µ Λ* Xd* Px*Y 1.133 0.182 -0.036 -0.299 -0.064 -0.299C 0.673 -0.083 -0.082 -0.189 -0.085 -0.189 I 4.687 0.096 0.100 -1.190 -0.072 -1.190h 0.187 0.153 -0.038 -0.008 0.156 -0.008πx -0.848 0.122 1.650 -0.006 0.127 -0.006π -0.658 0.092 1.652 0.189 0.096 0.189e 0.102 -0.027 1.658 0.968 -0.025 -0.032etot 0.950 -0.149 0.008 0.975 -0.152 0.975

Sticky prices (q =0.25), no capital adjustment costs (φ =1)

A G µ Λ* Xd* Px*Y -0.729 0.464 3.425 -0.341 0.228 -0.341C 0.079 0.007 1.020 -0.203 0.009 -0.203 I -5.142 1.582 18.377 -1.412 1.471 -1.412h -2.042 0.490 4.106 -0.058 0.506 -0.058πx -0.366 0.049 0.754 0.005 0.051 0.005π -0.270 0.034 0.930 0.197 0.035 0.197e 0.114 -0.028 1.634 0.968 -0.027 -0.032etot 0.480 -0.078 0.880 0.964 -0.078 0.964

Flexible prices (q =1), capital adjustment costs (φ =10)

A G µ Λ* Xd* Px*Y 0.831 0.177 -0.044 -0.223 -0.059 -0.223C 0.812 -0.080 -0.078 -0.225 -0.087 -0.225 I 2.120 0.046 0.034 -0.538 -0.033 -0.538h -0.067 0.148 -0.044 0.057 0.160 0.057πx -1.083 0.118 1.644 0.053 0.130 0.053π -0.843 0.089 1.647 0.236 0.099 0.236e 0.114 -0.026 1.658 0.965 -0.025 -0.035etot 1.196 -0.144 0.014 0.912 -0.156 0.912

Table 1a: Elasticities of key variables

Page 98: Macroeconomics of small open economies

Policy regime: Exchange rate peg

Flexible prices (q =1), no capital adjustment costs (φ =1)

A G Λ* Xd* Px*Y 1.133 0.183 -0.316 -0.064 -0.299C 0.673 -0.082 -0.213 -0.084 -0.189 I 4.684 0.097 -1.202 -0.072 -1.189h 0.187 0.154 -0.024 0.157 -0.008πx -0.950 0.149 -0.981 0.152 0.025π -0.760 0.119 -0.785 0.122 0.220e 0.000 0.000 0.000 0.000 0.000etot 0.950 -0.149 0.981 -0.152 0.975

Sticky prices (q =0.25), no capital adjustment costs (φ =1)

A G Λ* Xd* Px*Y -1.018 0.536 -2.695 0.298 -0.269C -0.023 0.032 -0.984 0.033 -0.180 I -6.628 1.954 -13.705 1.828 -1.030h -2.388 0.577 -2.872 0.589 0.028πx -0.409 0.060 -0.382 0.061 0.018π -0.327 0.048 -0.306 0.049 0.215e 0.000 0.000 0.000 0.000 0.000etot 0.409 -0.060 0.382 -0.061 0.982

Policy regime: Strict domestic inflation targeting

Sticky prices (q =0.25), no capital adjustment costs (φ =1)

A G Λ* Xd* Px*Y 1.131 0.181 -0.295 -0.066 -0.295C 0.664 -0.085 -0.182 -0.087 -0.182 I 4.711 0.095 -1.193 -0.074 -1.193h 0.184 0.152 -0.003 0.154 -0.003πx 0.000 0.000 0.000 0.000 0.000π 0.190 -0.030 0.195 -0.030 0.195e 0.950 -0.148 0.973 -0.151 -0.027etot 0.949 -0.148 0.973 -0.151 0.973

Table 1b: Elasticities of key variables (Cont.)

Page 99: Macroeconomics of small open economies

Benchmark policies:

Strict domestic inflation targeting (strict DIT) - Reaction coeff.: ρr =0.0, φy =0.0, φπ =1000, φe =0.0

Sticky prices (q =0.25) Flexibel prices (q =1)Shocks to: Shocks to:

A G TOT FD A G TOT FDstd (Y ) 0.2915 0.0729 0.0766 0.0200 0.2905 0.0728 0.0765 0.0198std (πx ) 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000std (π ) 0.0240 0.0029 0.0185 0.0031 0.0240 0.0029 0.0185 0.0031std (etot ) 0.4758 0.0561 0.2721 0.0509 0.4751 0.0561 0.2721 0.0508std (R ) 0.0152 0.0019 0.0010 0.0024 0.0152 0.0019 0.0010 0.0024

welfare -114.9880 -114.9854 -114.9852 -114.9853 -114.9880 -114.9854 -114.9852 -114.9853

PegSticky prices Flexibel pricesShocks to: Shocks to:

A G TOT FD A G TOT FDstd (Y ) 0.3147 0.0951 0.3129 0.0354 0.3531 0.0749 0.0988 0.0203std (πx ) 0.0468 0.0065 0.0415 0.0066 0.0958 0.0150 0.0998 0.0154std (π ) 0.0374 0.0052 0.0332 0.0053 0.0766 0.0120 0.0799 0.0123std (etot ) 0.4209 0.0424 0.1748 0.0369 0.5048 0.0558 0.2726 0.0514std (R ) 0.0000 0.0000 0.0160 0.0000 0.0000 0.0000 0.0160 0.0000

welfare -114.9967 -114.9861 -115.0068 -114.9860 -114.9887 -114.9854 -114.9858 -114.9853

Rule 1:

Baseline specification: flexible DIT - Reaction coeff.: ρr =0.0, φy =0.5, φπ =1.5, φe =0.0

Sticky prices Flexibel pricesShocks to: Shocks to:

A G TOT FD A G TOT FDstd (Y ) 0.2690 0.0735 0.0714 0.0173 0.4243 0.1044 0.1086 0.0298std (πx ) 0.2684 0.0647 0.0648 0.0195 0.4159 0.0945 0.1003 0.0312std (π ) 0.2661 0.0654 0.0704 0.0193 0.4129 0.0949 0.1042 0.0309std (etot ) 0.3247 0.0866 0.3056 0.0393 0.4356 0.0652 0.2824 0.0477std (R ) 0.2683 0.0604 0.0616 0.0206 0.4117 0.0895 0.0961 0.0320

welfare -115.1059 -114.9916 -114.9914 -114.9860 -115.2905 -115.0002 -115.0015 -114.9868

Flexible domestic inflation plus terms of trade targeting - Reaction coeff.: ρr =0.0, φy =0.5, φπ =1.5, φe =-0.1

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2715 0.0743 0.0630 0.0184std (πx ) 0.2060 0.0831 0.1163 0.0129std (π ) 0.2035 0.0838 0.1216 0.0128std (etot ) 0.3570 0.0947 0.3373 0.0434std (R ) 0.2087 0.0782 0.1094 0.0145

welfare -115.0592 -114.9957 -115.0046 -114.9856

Table 2a: Performance and welfare of alternative policy regimes

Page 100: Macroeconomics of small open economies

Rule 1 (Cont.):

Flexible domestic inflation plus terms of trade targeting - Reaction coeff.: ρr =0.0, φy =0.5, φπ =1.5, φe =-0.2

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2721 0.0757 0.0555 0.0195std (πx ) 0.1283 0.1058 0.1830 0.0044std (π ) 0.1257 0.1065 0.1882 0.0049std (etot ) 0.3955 0.1043 0.3754 0.0483std (R ) 0.1341 0.1002 0.1719 0.0065

welfare -115.0169 -115.0023 -115.0332 -114.9854

Flexible domestic inflation plus terms of trade targeting - Reaction coeff.: ρr =0.0, φy =0.5, φπ =1.5, φe =-0.3

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2699 0.0781 0.0503 0.0203std (πx ) 0.0424 0.1340 0.2703 0.0068std (π ) 0.0406 0.1348 0.2754 0.0080std (etot ) 0.4416 0.1158 0.4217 0.0543std (R ) 0.0468 0.1277 0.2545 0.0041

welfare -114.9902 -115.0127 -115.0905 -114.9854

Flexible domestic inflation plus terms of trade targeting - Reaction coeff.: ρr =0.0, φy =0.5, φπ =1.5, φe =-0.4

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2632 0.0818 0.0493 0.0207std (πx ) 0.1172 0.1697 0.3863 0.0218std (π ) 0.1196 0.1705 0.3915 0.0226std (etot ) 0.4977 0.1295 0.4787 0.0617std (R ) 0.1029 0.1626 0.3653 0.0184

welfare -115.0028 -115.0296 -115.2023 -114.9859

Flexible CPI inflation targeting (flexible CIT) - Reaction coeff.: ρr =0.0, φy =0.5, φπ =1.5, φe =0.0

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2635 0.0749 0.0795 0.0167std (πx ) 0.2624 0.0660 0.0721 0.0186std (π ) 0.2597 0.0666 0.0761 0.0182std (etot ) 0.3227 0.0861 0.3035 0.0391std (R ) 0.2580 0.0626 0.0745 0.0191

welfare -115.0966 -114.9921 -114.9946 -114.9859

Table 2b: Performance and welfare of alternative policy regimes (Cont.)

Page 101: Macroeconomics of small open economies

Rule 2:

Baseline specification: flexible DIT - Reaction coeff.: ρr =0.3333, φy =0.4, φπ =1.5, φe =0.0

Sticky prices Flexibel pricesShocks to: Shocks to:

A G TOT FD A G TOT FDstd (Y ) 0.2520 0.0691 0.0759 0.0162 0.2905 0.0727 0.0764 0.0199std (πx ) 0.0337 0.0034 0.0007 0.0032 0.0577 0.0063 0.0042 0.0056std (π ) 0.0232 0.0066 0.0195 0.0022 0.0386 0.0091 0.0219 0.0034std (etot ) 0.4303 0.0613 0.2729 0.0459 0.4758 0.0561 0.2720 0.0509std (R ) 0.0157 0.0026 0.0045 0.0023 0.0292 0.0029 0.0029 0.0035

welfare -114.9869 -114.9853 -114.9852 -114.9854 -114.9890 -114.9854 -114.9852 -114.9853

Flexible domestic inflation plus terms of trade targeting - R. coeff.: ρr =0.3333, φy =0.4, φπ =1.5, φe =-0.1

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2595 0.0680 0.0688 0.0171std (πx ) 0.0264 0.0045 0.0059 0.0023std (π ) 0.0170 0.0079 0.0256 0.0019std (etot ) 0.4395 0.0632 0.2831 0.0473std (R ) 0.0142 0.0030 0.0046 0.0023

welfare -114.9868 -114.9853 -114.9852 -114.9854

Flexible domestic inflation plus terms of trade targeting - R. coeff.: ρr =0.3333, φy =0.4, φπ =1.5, φe =-0.2

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2681 0.0668 0.0631 0.0183std (πx ) 0.0187 0.0057 0.0121 0.0014std (π ) 0.0134 0.0094 0.0327 0.0022std (etot ) 0.4497 0.0653 0.2947 0.0488std (R ) 0.0137 0.0033 0.0061 0.0024

welfare -114.9868 -114.9853 -114.9855 -114.9854

Flexible domestic inflation plus terms of trade targeting - R. coeff.: ρr =0.3333, φy =0.4, φπ =1.5, φe =-0.3

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2783 0.0656 0.0606 0.0197std (πx ) 0.0104 0.0070 0.0188 0.0003std (π ) 0.0159 0.0110 0.0409 0.0030std (etot ) 0.4613 0.0677 0.3079 0.0505std (R ) 0.0144 0.0038 0.0086 0.0027

welfare -114.9872 -114.9853 -114.9861 -114.9853

Table 2c: Performance and welfare of alternative policy regimes (Cont.)

Page 102: Macroeconomics of small open economies

Rule 2 (Cont.):

Flexible CIT - Reaction coeff.: ρr = 0.3333, φy=0.4, φπ = 1.5, φe = 0.0

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2500 0.0695 0.0791 0.0161std (πx ) 0.0345 0.0034 0.0014 0.0033std (π ) 0.0230 0.0062 0.0164 0.0019std (etot ) 0.4280 0.0609 0.2706 0.0456std (R ) 0.0089 0.0031 0.0078 0.0015

welfare -114.9870 -114.9854 -114.9854 -114.9854

Rule 3:

Baseline specification: flexible DIT - Reaction coeff.: ρr =0.6666, φy =0.1, φπ =2.0, φe =0.0

Sticky prices Flexibel pricesShocks to: Shocks to:

A G TOT FD A G TOT FDstd (Y ) 0.2703 0.0744 0.0885 0.0173 0.2908 0.0729 0.0770 0.0199std (πx ) 0.0154 0.0012 0.0074 0.0020 0.0246 0.0019 0.0117 0.0033std (π ) 0.0103 0.0018 0.0118 0.0013 0.0099 0.0019 0.0122 0.0013std (etot ) 0.4526 0.0543 0.2589 0.0475 0.4759 0.0561 0.2722 0.0509std (R ) 0.0114 0.0018 0.0070 0.0017 0.0161 0.0021 0.0073 0.0024

welfare -114.9869 -114.9854 -114.9855 -114.9854 -114.9880 -114.9854 -114.9853 -114.9853

Flexible domestic inflation plus terms of trade targeting - R. coeff.: ρr = 0.6666, φy =0.1, φπ =2.0, φe =-0.1

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2739 0.0739 0.0848 0.0177std (πx ) 0.0128 0.0009 0.0055 0.0017std (π ) 0.0122 0.0022 0.0139 0.0016std (etot ) 0.4566 0.0548 0.2626 0.0481std (R ) 0.0129 0.0021 0.0077 0.0019

welfare -114.9870 -114.9854 -114.9854 -114.9854

Flexible domestic inflation plus terms of trade targeting - R. coeff.: ρr =0.6666, φy =0.1, φπ =2.0, φe =-0.2

Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2778 0.0734 0.0812 0.0181std (πx ) 0.0101 0.0006 0.0036 0.0013std (π ) 0.0148 0.0026 0.0164 0.0019std (etot ) 0.4608 0.0554 0.2665 0.0486std (R ) 0.0148 0.0023 0.0090 0.0022

welfare -114.9872 -114.9854 -114.9854 -114.9854

Table 2d: Performance and welfare of alternative policy regimes (Cont.)

Page 103: Macroeconomics of small open economies

Rule 3 (Cont.):

Flexible domestic inflation plus terms of trade targeting - R. coeff.: ρr =0.6666, φy =0.1, φπ =2.0, φe =-0.3Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2820 0.0730 0.0779 0.0187std (πx ) 0.0073 0.0003 0.0018 0.0010std (π ) 0.0180 0.0030 0.0191 0.0023std (etot ) 0.4652 0.0559 0.2706 0.0493std (R ) 0.0170 0.0027 0.0108 0.0025

welfare -114.9875 -114.9854 -114.9854 -114.9853

Flexible CIT - Reaction coeff.: ρr =0.6666, φy =0.1, φπ =2.0, φe =0.0Sticky pricesShocks to:

A G TOT FDstd (Y ) 0.2684 0.0747 0.0915 0.0172std (πx ) 0.0160 0.0013 0.0076 0.0020std (π ) 0.0067 0.0013 0.0089 0.0009std (etot ) 0.4509 0.0541 0.2574 0.0473std (R ) 0.0102 0.0013 0.0062 0.0015

welfare -114.9871 -114.9855 -114.9857 -114.9854

Strict CPI inflation targeting (strict CIT):

Strict CIT - Reaction coeff.: ρr =0.0, φy =0.0, φπ =1000, φe =0.0

Sticky prices Flexibel pricesShocks to: Shocks to:

A G TOT FD A G TOT FDstd (Y ) 0.2688 0.0761 0.1078 0.0177 0.2905 0.0730 0.0776 0.0199std (πx ) 0.0153 0.0018 0.0115 0.0020 0.0241 0.0029 0.0185 0.0031std (π ) 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000 0.0000std (etot ) 0.4510 0.0531 0.2496 0.0473 0.4761 0.0562 0.2725 0.0510std (R ) 0.0232 0.0028 0.0210 0.0029 0.0122 0.0015 0.0035 0.0019

welfare -114.9887 -114.9855 -114.9876 -114.9854 -114.988 -114.985 -114.985 -114.985

Table 2e: Performance and welfare of alternative policy regimes (Cont.)

Page 104: Macroeconomics of small open economies
Page 105: Macroeconomics of small open economies

Chapter 3

Technology shocks and

employment in open economies

3.1 Introduction

The one-good, one-shock RBC model as described e.g. in King and Rebelo

[23] features a strongly positive correlation between labor input and mea-

sured labor productivity, and this regardless whether technology shocks are

considered to be highly persistent or permanent.1 This prediction contra-

dicts the facts. Hansen and Wright [18] e.g. report estimates of the correla-

tion between total hours worked and measured labor productivity (output

divided by total hours) for quarterly U.S. data. They consider two measures

of total hours (and, thus, productivity) and several sample periods. Their

point estimates are near zero or slightly negative while the individual series

for total hours and productivity are, respectively, strongly and moderately

procyclical.

This anomaly has spurred a vast literature. To illustrate the problem,

consider a labor-supply/labor-demand graph with labor on the abscissa and

real wage on the ordinate. Under perfect competition the real wage equals

the marginal product of labor which – under standard assumptions – is di-

1King and Rebelo [23] discuss the effects of both shocks that have a highly persistent

but not permanent and shocks that have a permanent effect on total factor productivity.

In the case of a permanent positive productivity shock, the long-run effect on labor is

nil. In the short and medium run, however, the effect on labor is strongly positive.

Page 106: Macroeconomics of small open economies

88 CHAPTER 3

rectly proportional to average labor productivity.2 Productivity shocks –

i.e., shifts of the labor demand curve – imply a positive correlation between

labor and the real wage. To get a roughly zero correlation between labor

and real wage, shifts of the supply curve have somehow to be introduced. If

both labor demand and labor supply shocks are active, then the predicted

correlation might be close to zero. This can be achieved, among others, by

augmenting the basic one-good, one-shock RBC model to allow for govern-

ment consumption (compare e.g. Christiano and Eichenbaum [5]).

Modifying the basic one-good, one-shock RBC model to allow for gov-

ernment consumption may improve the model’s ability to produce uncon-

ditional correlations between labor input and average labor productivity

similar to those found in the data. Yet the predicted correlation condi-

tional on an exogenous productivity shock – regardless whether assumed to

be highly persistent or permanent – remains positive and high.

In an influential paper, Gali [13] studies the effects of technology shocks

on labor input in the G7 countries using a bivariate structural vector au-

toregressive model (VAR) approach. Gali’s results can be summarized as

follows: (i) The estimated conditional correlations between the proxy for

labor input and measured labor productivity are negative for technology

shocks and positive for non-technology shocks.3 (ii) The labor proxy shows a

persistent decline in response to a positive technology shock. (iii) Measured

productivity increases temporarily in response to a positive non-technology

shock. Given that standard labor proxies are (strongly) procyclical, Gali

concludes that shocks other than technology shocks must play the dominant

role in business cycles.

In the sense of a robustness check, Gali also estimates a higher dimen-

sional extension of the bivariate VAR including – in addition to labor input

and productivity – real money balances, inflation, and a real interest rate;

the findings for the bivariate VAR are largely confirmed.

The impact of Gali’s work on the profession was rather big, all the more

as independent studies published at about the same time found similar ev-

2For the CRTS Cobb-Douglas production function Yt = ZtNχt K1−χ

t where Yt is out-

put, Zt is total factor productivity, Nt is labor, Kt is capital, and χ ∈ [0, 1] is a parameter,

we have ∂Yt/∂Nt = χYt/Nt.3Examples of non-technology shocks are tax rate changes, demographic changes in

the labor force, and shocks to government spending or the nominal money supply.

Page 107: Macroeconomics of small open economies

3.1. INTRODUCTION 89

idence.4 The subsequent work went into two directions. One route was to

check whether the measured response of labor input on a productivity shock

depends on (or can be explained by) an incorrect specification of the utilized

structural VAR framework. Gali’s identification strategy – which goes back

to Blanchard and Quah [4] – relies on the assumption that only technology

shocks have a permanent effect on measured labor productivity. Uhlig [32]

challenges the theoretical foundations for this assumption and presents a

neoclassical business cycle model in which there are two shocks which may

influence labor productivity in the long run apart from technology shocks,

namely changes to the dividend tax rate and shifts in the preferences. A

second critical point within Gali’s econometric framework is the choice of

the proxy for labor input and its treatment as either a level-stationary or

a difference-stationary variable. Altig et al. [1] estimate a full-fledged New

Neoclassical Synthesis (NNS) model with eight structural shocks; the mon-

etary shock is identified by strategies which are standard in the literature,

the productivity shock is identified as in Gali, and the remaining six shocks

are identified by means of what they call model-based strategies. They

employ average weekly hours worked per person rather than total hours or

employment series as Gali does. If average hours enter the VAR in levels

rather than in first differences, Altig et al. find that labor input rises in

response to a positive technology shock.5

A second route was to accept the Gali evidence but to question the con-

clusion that exogenous variations in technology play a very limited role, if

any, as sources of the business cycle. Collard and Dellas [7], for instance,

argue that the predictions of an RBC model are fairly sensitive to the de-

gree of openness to trade and to the trade elasticity. They underpin their

conjecture by presenting an international RBC model that produces – for

plausible parameter values – conditional correlations between labor input

and productivity of the same sign and magnitude as those estimated by Gali.

Francis and Ramey [12] attain the same goal by modifying a standard RBC

model of a closed economy to allow for habit formation in consumption.

4Basu et al. [2] and Shea [31] investigate Solow-residual based measures of technolog-

ical shocks.5Christiano et al. [6] take up the issue of whether the specification of average hours

worked in levels or in first-differences is more adequate and provide additional evidence.

Page 108: Macroeconomics of small open economies

90 CHAPTER 3

Adding to this body of literature, we extend the discussion in two direc-

tions. First, we add an open-economy block to Gali’s five-variable VAR and

then repeat his analysis. Second, we use – alongside with standard struc-

tural VAR methods – structural vector error correction model (VECM)

methods and then, again, repeat his analysis.

Let us briefly elaborate on the two modifications. As mentioned above,

Uhlig [32] and others argue that there might be sources of long-run stochas-

tic movements in labor productivity other than technology shocks; perma-

nent changes in the capital income tax rate and shifts in preferences are

frequently cited candidates.6 We suggest that within an international con-

text there is yet another potential source of shifts in labor productivity,

which has (to our knowledge) not received much attention in the literature

sofar, namely permanent terms-of-trade shocks. Whatever the empirical

importance of each of these sources might be,7 we cannot exclude that the

shocks identified by Gali are contaminated in that they capture disturbances

other than genuine technological changes. A structural modelling frame-

work which allowed one to disentangle the alternative potential origins of

long-run shifts in labor productivity would therefore be highly desirable.

As a first step towards that end we suggest to repeat Gali’s exercise with

the difference that we include a set of openness variables and then to inves-

tigate whether this modification has an effect on the estimated dynamics

and conditional correlations. In principle, the inclusion of additional vari-

ables allows a more precise identification of the shocks that do not have a

permanent effect on measured productivity (we call them ”non-technology

shocks”). This in turn leads to a more precise identification of the conglom-

erate which – the methodological flaws of the utilized method notwithstand-

ing – we continue to call ”technology shocks”. While Gali in his robustness

6As usual in a business cycle context, we abstract from the possibility of endogenous

technological progress.7In a recent attempt to provide evidence in support of his identification scheme, Gali

[14] compares the co-movement between ”his” shocks and measures of dividend tax rates

(both for the U.S.). He finds that the two series are uncorrelated whereas the very same

shocks are significantly positively correlated with independent measures of technological

change (in particular, those provided by Basu et al. [2]). Gali takes this as evidence in

favor of the view that technology changes are the only empirically relevant kind of shocks

which plausibly induce a permanent shift in measured labor productivity.

Page 109: Macroeconomics of small open economies

3.1. INTRODUCTION 91

check adds variables which make allowance for monetary aspects, we focus

here on variables that capture aspects of international trade and finance

such as net exports and the terms of trade.

In the specification of his higher dimensional VAR, Gali incorporates

two cointegrating relations, one between money and the price level (both in

first differences) and another between the interest rate and changes in the

price level. Although from a theoretical point of view such a proceeding

certainly is justifiable, one can argue that by specifying the number and the

form of the cointegrating relations present in the VAR, more restrictions

are imposed than absolutely necessary.8 Any estimation method which gets

by with less restrictive assumptions should be favored. This is particularly

true when we plan to estimate models for which no priors regarding the

number and the form of the cointegrating relations exist (e.g. the models

including an open-economy block). Therefore, our second modification is

the use of structural VECM rather than standard structural VAR methods.

The estimations are conducted with quarterly data for the G7 countries

(minus Germany) plus Australia, Canada, Switzerland, Spain, and New

Zealand. We utilize employment in manufacturing as a proxy for labor in-

put; all variables are treated as integrated of order one. In a few cases,

the inclusion of openness variables indeed lead to an estimated correlation

between labor input growth and measured labor productivity growth con-

ditional on a positive ”technology shock” which is either close to zero or

negative but not significantly different from zero. However, we cannot dis-

cern a systematic pattern that would indicate a relationship between the

incorporation of international trade aspects and the size of the conditional

correlation coefficient. Moreover, we find that extending the standard struc-

tural VAR framework to the cointegrated case with an arbitrary number of

cointegrating relations and general linear restrictions on the cointegration

space does not alter the Gali evidence.

The remainder of this paper is organized as follows. In section 3.2, we

discuss Gali’s bivariate structural VAR and present our replication results.

In section 3.3, we discuss Gali’s robustness check and present replication

8In Gali [13], the relations are embedded despite the fact that in most cases for-

mal tests indicate that the presumed relations do not belong to the cointegrating space

(compare the technical appendix to Gali’s article).

Page 110: Macroeconomics of small open economies

92 CHAPTER 3

results. In section 3.4, we repeat Gali’s analysis with the two modifications

discussed above and report the main findings. Section 3.5 concludes and

gives directions for future work. Most technicalities are postponed to the

appendices. Appendix 3.A sketches our estimation strategy in connection

with univariate unit root tests. Appendix 3.B contains the applied method

for estimating structural VARs and VECMs and for computing dynamics

and conditional correlations. Appendix 3.C describes the computation of

percentile confidence intervals for the estimated impulse-response functions

and conditional correlation coefficients and appendix 3.D, finally, contains

the model specifications (VAR and VECM).

3.2 Benchmark model

3.2.1 Estimation method

Suppose the economic variables of interest can be expressed as a distributed

lag of some unobserved exogenous shocks, whereby the number of shocks

equals the number of endogenous variables and the shocks are uncorrelated

at all leads and lags. Let xt be a n × 1 covariance-stationary vector of

economic variables. The structural vector moving average (VMA) represen-

tation for xt is given by

xt = A (L) ut (3.1)

where A (L) is a matrix infinite-order lag polynomial and ut is a n×1 vector

of white noise disturbances

E (ut) = 0

E (utu′

τ ) =

Σu for t = τ

0 otherwise

where Σu is a symmetric positive definite matrix.

In Gali’s [13] bivariate model, the first variable is average labor pro-

ductivity in first differences, denoted by ∆(yt − nt) where yt is the natural

logarithm of real GDP and nt is the log of labor input; the second vari-

able is a detrended measure of the log of labor input. The first structural

disturbance is labelled ”technology shock” and the second ”non-technology

shock”. Gali continues by imposing two identifying restrictions. First, the

Page 111: Macroeconomics of small open economies

3.2. BENCHMARK MODEL 93

structural shocks are mutually uncorrelated and their variances are nor-

malized to one (i.e., Σu = In). Second, only technology shocks have a

permanent effect on the (log) level of labor productivity. The latter restric-

tion corresponds to the assumption that the cumulative impulse-response

of the non-technology shock on ∆(yt − nt) must equal to zero, that is, in

[x1,t

x2,t

]=

[A11(L) A12(L)

A21(L) A22(L)

] [u1,t

u2,t

](3.2)

A12(1) is set to 0.9 This sort of long-run identifying restriction requires that

the level of the endogenous variable on which the restriction is imposed

(here: (yt − nt)) is non-stationary but not cointegrated with any of the

other non-stationary endogenous variables in the system (in case the proxy

for labor input turns out to be I(1): with nt). In the case at hand, there

is every reason to believe that (yt − nt) is non-stationary and that it is not

cointegrated with nt.

Gali employs the following series. For the U.S. (sample period: 1948:1-

1994:4), he makes use of real GDP and either total employee-hours in nona-

gricultural establishments or employed civilian labor force. All three series

9To see this, consider the following univariate process (in first-differences)

∆xt = θ (L) εt

where θ (L) is an infinite-order lag polynomial and εt is a white noise variable. The effect

of εt on ∆xt+j is given by∂∆xt+j

∂εt

= θj .

Now, note that

xt+j = xt+j − xt+j−1 + xt+j−1 − xt+j−2 + xt+j−2 − ...

= ∆xt+j + ∆xt+j−1 + ...

Thus, the effect of εt on xt+j (the level) is given by

∂xt+j

∂εt

=∂∆xt+j

∂εt

+∂∆xt+j−1

∂εt

+ ... +∂∆xt

∂εt

+ ...

= θj + θj−1 + ... + θ0

=∑j

i=0θi.

For ε to have no effect on x in the long run, we must have that∑j

i=0θi = 0.

Page 112: Macroeconomics of small open economies

94 CHAPTER 3

are drawn from Citibase. For the remaining G7 OECD countries Canada,

United Kingdom, Germany, France, Italy and Japan (sample periods are

country specific and depend on data availability), he makes use of GDP

(drawn from the OECD Quarterly National Accounts) and employed civil-

ian labor force (drawn from the OECD Quarterly Labor Force statistics).

Gali detrends the measure of labor input in two alternative ways: either

he takes first differences or he removes a fitted linear time-trend from the

original series.

Estimating the bivariate VAR by means of a method similar to the

one described in appendix 3.B provides Gali with estimates of the MA-

coefficients of model (3.1). Based on the estimated MA-coefficients, point

estimates for the impulse response functions (IRF) and the conditional co-

variances and correlations can be computed. The standard errors for the

conditional correlations and the confidence intervals for the IRF are gener-

ated using a Monte Carlo method.

3.2.2 Two critical remarks

Before proceeding by replicating Gali’s results for the bivariate VAR, we

critically assess two key ingredients of the econometric framework.

Sources of long-run stochastic movements in labor productivity

In the basic one-good, one-shock RBC model, there is (as the name says)

only one source of shocks, namely changes in the production function.10 In

richer frameworks, however, there might be sources of long-run stochastic

movements in labor productivity other than technology shocks, from which

we conclude that the assumption upon which Gali’s identification strategy

relies may not be valid and that the identified shocks may contain things

other than genuine technological changes.

Two candidates suggested in the literature are permanent changes in the

dividend tax rate and shifts in preferences (see e.g. Uhlig [32]).11 A further

10While productivity is typically assumed to follow a highly persistent but not per-

manent autoregressive process, the model can readily be extended to the case where

productivity follows a random walk with a positive drift - a typical homework assign-

ment for graduate students.11Note that in the presence of non-distorting taxation, permanent shocks to govern-

Page 113: Macroeconomics of small open economies

3.2. BENCHMARK MODEL 95

source which has not (to our knowledge) received much attention so far

are permanent terms-of-trade shocks. Consider a standard NNS model of a

small open economy such as Kollmann [24]. To simplify matters, suppose

money growth is constant, the exchange rate is flexible, and asset markets

are complete. Moreover, domestic intermediate goods prices are supposed

to be moderately sticky. In response to an unexpected, permanent im-

provement of the terms of trade, final good producers substitute away from

domestic inputs toward foreign inputs, a force, which works in the direction

of a contraction of the domestic intermediate goods production. At the same

time, the terms-of-trade improvement induces a positive permanent wealth

effect (the small open economy needs to export less in order to purchase a

given bundle of imports) and, thus, a permanent rise in consumption and

leisure. The rise in consumption is smoothed by an increase in investment,

which leads in the long run to a higher capital stock. Altogether, the model

predicts a lasting shift of measured labor productivity.

Removing trends from macroeconomic time series Good economet-

ric practice requires that the specification of an econometric model is based

on theory whenever this is feasible. This is particularly true for the choice of

the number and kind of endogenous (an possibly exogenous) variables and

the identification scheme. When theory offers no clear guidelines as is the

case for the choice of the number of lags to be included in a VAR, then our

decision is usually based on statistical inference. Finally, for some aspects

of the empirical model we have no choice other than accepting what the

real world offers us; this is typically true for the choice of the data range.

A related issue is the question about the appropriate way of removing

trends from macroeconomic time series. In principle, macroeconomic time

series may be level stationary, difference stationary, or trend stationary. In

the last two cases, the time trend may be either common or series specific

and it may have structural breaks. It is a well-known fact that the amount

of macroeconomic data is not large enough to get precise information about

the true data generating process; in other words, there is no reliable method

for distinguishing among the alternatives listed above. As a way to cope

with this ambiguity, we follow a rather pragmatic approach, partly based

ment consumption do not belong in this category; see e.g. Baxter and King [3].

Page 114: Macroeconomics of small open economies

96 CHAPTER 3

on empirical evidence, partly based on common sense.

To illustrate our approach, let us have a closer look at the labor proxy

we choose. In the basic RBC model, all of the variation in aggregate hours

arises due to movements in hours per worker (the ”intensive margin”). Yet

movements of individuals in and out of employment (the ”extensive mar-

gin”) seems to be important for understanding aggregate labor supply (see

e.g. Hansen’s [17] seminal study). We are therefore convinced that an ac-

curate proxy for labor input should incorporate the extensive margin.

What we actually would like to have is a measure of total hours worked

in the non-farm business sector.12 However, if we had the choice between a

measure of people with jobs in an economy (the extensive margin only) and a

measure of the individual choice of the number of hours worked for the same

economy (the intensive margin only), we would go for the employment series.

In the case of the OECD MEI data base – the widely accessible data base

we are making use of – this is exactly the situation: There is either a series

of employment or a series of average weekly hours, both in manufacturing.13

We, thus, decide to restrict our investigation on employment.

It remains to be decided how the trend should be removed from the

chosen labor proxy. In principle, there is no theoretical reason against a

time trend in employment in manufacturing. This trend component may

stem from long-run developments on the demand and the supply side of the

labor market, such as changes in the demographic composition (immigration

or a long-run drop in the birth-rate), a rise in real income (and the related

substitution and income effects), shifts in the production structure, etc.

However, in what follows we assume – and here the common sense comes in

– that all time series employed in this study are I(1), including the nominal

12To be more precise, we would like to have a measure of labor services in the non-farm

business sector. But since such a measure is not available, we prefer the proxy total hours

worked. Suppose we knew of each individual in the labor force how many hours he or

she works during a given period. In the aggregate, this gives total hours worked.13Average weekly hours series are available only for a subset of the countries included

in our study; moreover, only the series for the U.S. is of an acceptable quality. When we

experimented with a version of the benchmark model for the USA which involves average

hours series in levels, we got the same result as reported in Altig et al. [1] and others:

The response of hours to a technology shock is positive at all dates; also, the correlation

between measured productivity growth and hours is positive (and significant).

Page 115: Macroeconomics of small open economies

3.2. BENCHMARK MODEL 97

and real variables that might be added to the bivariate VAR in an attempt

to get a more precise measure of what we term technology shocks . It goes

without saying that we conduct standard (univariate) unit root tests, and

in most cases the analysis indeed indicates that the series at hand contain a

random walk component. But we would like to stress that we consider our

basic assumption to hold independent of those test results.

3.2.3 Replication results

In order to replicate Gali’s results for the bivariate VAR, we proceed as

follows:

• Step 1: Data. We obtain data on gross domestic output and employ-

ment in manufacturing for the G7 countries plus Australia, Canada,

Switzerland, Spain, and New Zealand (in what follows abbreviated by

AUS for Australia etc.). The data bases are OECD MEI (mostly) and

IMF (partly). All series (except interest rates) are seasonally adjusted

(where necessary, we apply Census X12).

• Step 2: Univariate analysis. The series (yt − nt) and nt are pretested

in order to assess their order of integration. A description of the test-

ing strategy we follow is given in appendix 3.A. There is only one

country for which the evidence regarding (yt − nt) is mixed, namely

AUS. There is only one country for which the hypothesis that nt is

I(1) has to be rejected, namely ESP.

• Step 3: Model specification. Regardless of the findings of step

2 and as a matter of principle we consider both (yt − nt) and nt as

I(1)-variables. Accordingly, our benchmark model is a VAR in first

differences given by

∆xt = D1∆xt−1 + D2∆xt−2 + ... + Dp∆xt−p + εt

with

xt =[

(yt − nt) nt

]′

where p is the number of lags in the model, and εt is an 2 × 1 vector

containing the reduced form residuals which are assumed to be white

Page 116: Macroeconomics of small open economies

98 CHAPTER 3

noise. In order to determine the number of lags, we follow Enders [10],

p. 396, and estimate the reduced form VAR using the undifferenced

(non-stationary) data.14 We then apply a classical LR test (or pairing

down lag length strategy).15 The sample period, finally, is country

specific and depends on data availability. In some cases, data avail-

ability is limited in such a way that we have to do without a particular

series or a particular country. For instance, the time series for GER

only start with the reunification and in our view are too short for the

purpose of model estimation.

• Step 4a: Estimating impulse response functions (IRF) and

the conditional correlations (CC) – the former in terms of levels,

the latter in terms of growth rates. For a description of the applied

estimation method see appendix 3.B.

• Step 4b: Computing bootstrap confidence intervals (CI) for

the IRF and the CC. In our view it is not meaningful to display

standard errors for the CC. A correlation has a lower and an upper

bound (i.e., a truncated distribution). For this reason we prefer to

display percentile confidence bands (for a description of the applied

bootstrap method see appendix 3.C). However, to ensure compara-

bleness between Gali’s results and ours we will also display standard

errors.

Let us now turn to the replication results for the benchmark specifica-

tion (figure 1 and table 1). For each country investigated, we report the

dynamic response of employment to a technology shock (in terms of per-

centage deviations from average) with 5 and 95% confidence bands over

almost 5 years (beginning with t = 0) and the correlation between labor

productivity growth and employment growth conditional on a technology

shock with 5 and 95% percentile points. For each model specification, we

start with a rather detailed discussion of the results for the USA. We then

go on by discussing the rest of the G7 OECD countries minus Germany

14This procedure is standard in the univariate case (compare e.g. Hamilton [16], p. 553

or table 17.3). The argumentation carries over to the vector case.15We are careful and use a rather small size of tests (since the type one error cumulates).

Page 117: Macroeconomics of small open economies

3.2. BENCHMARK MODEL 99

(CAN, FRA, GBR, ITA, JPN) and, finally, the additionally investigated

OECD countries (AUS, CHE, ESP, NZL).

USA For the USA, both the estimated dynamics and the value for the

correlation between measured labor productivity growth and employment

growth conditional on a technology shock (CC), are pretty much in line with

the corresponding results reported in Gali [13] and in the technical appendix

that comes along with the article. A positive technology shock leads to a

negative and significant response in period 0 and 1. The response remains

negative but gets insignificant from period 2 after the shock onwards. The

estimated CC is negative and significant (-.82 compared to -.84 in Gali).

Note that the CI does not cover the point estimate; for an explanation and

a possible remedy compare appendix 3.C.

FRA, GBR, and ITA For FRA, we observe a j-shaped response of em-

ployment to a technology shock.16 After a weakly negative impact effect,

employment decreases for a couple of periods, then begins to rise, returns

to its average, and eventually gets positive (after about 3 years). For GBR

and ITA, on the other hand, the response of employment remains nega-

tive over the entire horizon. This general picture corresponds to the results

reported in Gali. Since our bootstrapped confidence bands are somewhat

narrower than those computed by Gali, the responses for FRA and GBR

are significant not only on impact and for the subsequent two periods but

for the first three to four periods. For ITA, the negative effect is significant

even over the entire horizon.

For GBR, the estimated CC is close to the corresponding value reported

by Gali (-0.87 compared to -0.91). For FRA, our point estimate is higher

(-0.31 compared to -0.81) while it is lower for ITA (-0.96 compared to -0.41).

In all three cases, the point estimates are significantly different from zero

(in the case of FRA where the 0.95 percentile point is -0.01 rather narrowly,

though). Note that the bootstrapped standard deviations are significantly

lower than those reported by Gali.

16FRA is the only country for which Gali utilizes detrended employment rather than

employment in first differences. Preliminary experiments lead us to follow suit; however,

unlike Gali we use the Hodrick-Prescott filter to remove the trend.

Page 118: Macroeconomics of small open economies

100 CHAPTER 3

CAN and JPN For all the investigated countries except CAN and JPN,

Gali finds a significantly negative response, at least on impact. We therefore

prefer to discuss the two countries CAN and JPN separately within the G7-

block. For both countries, the impact effect is negative but insignificant.

Employment rises quickly and turns positive in period 1 or 2; it remains

so in the medium and long run. In the case of JPN, the positive long-run

effect is actually significant.

For CAN, the estimated CC is higher than the corresponding value re-

ported by Gali (-0.27 compared to -0.59) and insignificant. For JPN, the

CC is somewhat lower than the value reported by Gali (-0.12 compared to

-0.07) but still insignificant. Note that for JPN the bootstrapped standard

deviation is clearly larger than the value reported by Gali (0.28 compared

to 0.08).

AUS, CHE, ESP, and NZL The response of employment for AUS and

CHE is similar to the one for FRA and GBR; the dynamics for ESP and

NZL are comparable to the one for ITA. In all four cases, the estimated

CCs are significantly negative. The variability is rather small for AUS and

ESP (about the same size as for GBR) and somewhat larger for CHE and

NZL (about the same size as for the USA).

Summary The replication results for the benchmark specification largely

confirm Gali’s findings. For the USA and GBR, the estimated dynamics

and the CC are in line with the corresponding point estimates reported by

Gali. For CAN, FRA, ITA, and JPN, the results coincide in qualitative

terms.

3.3 Nominal block

3.3.1 Estimation method

In order to check robustness, Gali augments the basic VAR by a nominal

block consisting of log nominal money (e.g. M1), the log price index (e.g. the

GDP deflator), and a nominal interest rate (e.g. the 90-day money market

rate) – denoted by mt, pt, and rt, respectively.

Page 119: Macroeconomics of small open economies

3.3. NOMINAL BLOCK 101

The extension of model (3.2) to the case of more than two variables is

straightforward. In this event, x2,t is no longer a scalar but an (n − 1) × 1

vector of variables which enter the model for theoretical or empirical rea-

sons. The n×1 vector xt of endogenous variables is still driven by the same

two structural disturbances. Matrices A21(L) and A22(L) have the appro-

priate dimensions. To identify the extended VAR, Gali makes the following

assumptions: (i) The conglomerate we term technology shock is orthogonal

to each of the n − 1 non-technology shocks; (ii) only the technology shock

has a long-run effect on (yt − nt); (iii) the n − 1 non-technology shocks

are orthogonal to each other. Given this set of assumptions, the structural

representation (3.2) can be recovered from the reduced form VAR to be

described below.

Note that assumption (iii) is completely arbitrary. This poses no prob-

lem as long as the (squared) MA-coefficients are finally added up – which

is the case in the process of computing the correlation conditional on an

aggregated non-technology shock (the sum of the components driven by

the n − 1 individual non-technology shocks). The response function to the

aggregated non-technology shock, however, is sensitive regarding alterna-

tive identification schemes. For instance, it makes a difference whether the

individual non-technology shocks are identified by means of an arbitrary

short-run identifying assumption (as is the case here) or an equally arbi-

trary long-run identifying assumption. But since in this study we entirely

focus on the IRF to and the correlation conditional on a technology shock,

this issue is not a concern for us.

According to the technical appendix that comes with Gali’s paper, uni-

variate analysis of the data on ∆pt, ∆mt, and rt indicates that all three series

can be characterized as I(1) variables. Further univariate analysis lead Gali

to the conclusion that ∆mt and ∆pt as well as rt and ∆pt+1 are cointe-

grated with cointegrating vectors [1,−1] (implying stationary processes for

∆ (mt − pt) and (rt − ∆pt+1), respectively).

As further reported in the technical appendix to Gali, cointegration tests

(Johansen procedure based on the trace statistic) on the five variable vector

xt = [(yt − nt) , nt, ∆mt, ∆pt, rt]′

point to the presence of a cointegration rank equal to 2 (which is consistent

Page 120: Macroeconomics of small open economies

102 CHAPTER 3

with the previous conjecture).17 However, tests of the joint hypothesis that

the vectors [0, 0, 1,−1, 0] and [0, 0, 0, 1,−1] belong to the cointegration space

are rejected.18 Nevertheless, Gali continues by estimating a reduced form

VAR with

xt =[

∆ (yt − nt) ∆nt ∆ (mt − pt) ∆2pt rt − ∆pt+1

]′.

3.3.2 Replication results

In order to replicate Gali’s results for the higher-dimensional structural

VAR, we proceed as follows.

• Step 1: Data. We come up with series for the GDP deflator, narrow

nominal money (mostly M1, in some cases M0), and the nominal

interest rate (3 month money market rate and 10-year government

bonds) for the OECD countries included in our study.

• Step 2: Univariate analysis. The series ∆pt, (rt − ∆pt+1), and

(mt − pt) are pretested. There is only one country for which the hy-

pothesis that (mt − pt) has a unit root and a drift has to be rejected,

namely ESP. The hypothesis that ∆pt has a unit root has to be re-

jected (or at least the evidence is mixed) for CAN, ITA, and NZL. And

finally, the hypothesis that (rt − ∆pt+1) is I(1) cannot be rejected for

AUS, FRA, JPN, and USA.

• Step 3: Model specification. Regardless of the findings of step 2

and as a matter of principle we consider ∆pt, (mt − pt), and rt as

17The reported test results refer to U.S. data; labor input is measured by total hours

worked. Note that (rt − ∆pt) is not precisely equal to (rt − ∆pt+1).18In terms of the VECM specification

∆xt = Πxt−1 +∑p−1

i=1Γi∆xt−i + εt

Gali imposes the restriction

Π =

[0 0 1 −1 0

0 0 0 −1 1

].

Page 121: Macroeconomics of small open economies

3.3. NOMINAL BLOCK 103

I(1)-variables. The reduced form VAR(p) is given by

xt = D1xt−1 + D2xt−2 + ... + Dpxt−p + εt

with

xt =[

∆ (yt − nt) ∆nt ∆ (mt − pt) ∆2pt rt − ∆pt+1

]′

where εt is assumed to be white noise. Lag-length is tested for

xt =[

(yt − nt) nt (mt − pt) ∆pt rt

]′.

The sample period is country specific and depends on data availability.

• Step 4: Estimating IRF and CC and computing bootstrap

CI. For a description of the applied estimation method see appendix

3.B. A description of the applied bootstrap method can be found in

appendix 3.C.

The results for the original five-variable VAR (hereafter ”VAR 0”) are

contained in figure 2 and table 1.

USA Again, we start by discussing the results for the USA. When looking

at the response of employment to a technology shock for ”VAR 0” reported

by Gali, the first thing which catches ones eye is that the confidence band is

clearly narrower compared to the benchmark specification. Moreover, the

response has a pronounced j-shaped form: After a modest negative impact

effect, employment continues to drop in period 1, then returns to its average,

and eventually turns positive (from period 7 onwards).

We observe a similar pattern. After a negative impact effect, employ-

ment further declines in period 1 and 2 and then rises and eventually turns

positive, but not as much as in Gali.19 In the long run (from 2 to 3 years

after the shock onwards), employment falls back below average on a level

that is less negative than in the case of the benchmark model. (We suppose

this is also the case for Gali’s estimation but we cannot tell since he chooses

19The negative effect on employment is significant for periods 0 to 2, as in the bench-

mark model.

Page 122: Macroeconomics of small open economies

104 CHAPTER 3

a horizon of just 3 years.) Also, the confidence band is somewhat narrower

than for the benchmark specification.

According to Gali, there is barely a difference between the CC estimates

for the two model specifications (-.82 for ”VAR 0” compared to -.84 for the

benchmark model). As far as our replication results are concerned, the point

estimate for CC is less negative compared to the benchmark specification

(-.66 compared to -.82), but still significant. The width of the bootstrapped

CI for the two model specifications differ only slightly. Note that the boot-

strapped standard deviation of .21 is rather large compared to the (very

small) .08 reported by Gali.

FRA, GBR, and ITA In the case of FRA, going from the bivariate to

the five-variable VAR gives rise to a widening of the confidence band, in

particular in the short run. As a consequence, the negative impact effect is

not significant any more. Moreover, the j-shaped response which we have

observed for the benchmark specification disappears. The medium- and

long-run effect on employment now is negative, albeit insignificantly. In the

case of GBR, the inclusion of additional variables gives rise to a j-shaped

response. While in the benchmark specification employment was below

average over the entire horizon, it now begins to rise in the medium run

and becomes positive in the long run. The positive long-run effect is not

significant, though. In the case of ITA, the inclusion of additional variables

seems to have no effect on the response of employment.

In the cases of GBR and ITA, the estimated CCs are close to the respec-

tive point estimates for the benchmark model and highly significant. The

variability gets somewhat bigger for GBR whereas for ITA it remains rather

small. For FRA, the point estimate is lower (i.e., more negative) compared

to the benchmark model; at the same time, the CI becomes wider such that

the CC is not significantly different from zero any more.

CAN and JPN For CAN, the impact effect is still insignificant. The

subsequent rise is less pronounced; so is the (insignificant) positive long-run

effect. The estimated CC is somewhat lower than for the benchmark model

but remains insignificant. For JPN, the only effect from augmenting the

VAR which we can discern is that the estimated CC now is positive (while

Page 123: Macroeconomics of small open economies

3.4. EXTENSIONS 105

it was weakly negative for the benchmark specification).

AUS, CHE, ESP, and NZL In the case of the remaining OECD coun-

tries investigated, the CC estimates are close to those for the respective

benchmark models. In the case of CHE, the confidence bands tend to be-

come narrower compared to the benchmark specification, while in the cases

of AUS and NZL they tend to become wider. (In the case of ESP, there is

no effect observable with respect to the bootstrapped confidence bands.)

Summary The findings for the bivariate VAR are largely confirmed. In

the case of the USA, the estimated CC is somewhat less negative than both

the value we got for the benchmark model and the corresponding value

reported by Gali, but still significant.

3.4 Extensions

3.4.1 Open-economy block

In the introduction, we argued that – despite the methodological flaws of

the utilized identification strategy and in the sense of an explorative study

– we want to check the robustness of the Gali evidence to the inclusion

of variables that capture aspects of international trade and finance. Our

choice of openness variables is partly based on theoretical models of small

open economies such as Kollmann [24]; it involves the current account bal-

ance (cat) and the terms of trade (tott).20 Unfortunately, for a number of

countries the series cat is too short for the purpose of model estimation

(this is the case for ESP, FRA, and ITA). Apart from that, for some of the

countries for which the series cat is available, its quality is questionable.

Therefore, we utilize the main component of the current account, the trade

balance (or net exports, nxt), in addition or as a proxy to the series cat.

This, and the fact that we are interested in the effect of the inclusion of

each individual international trade and finance variable lead us to estimating

a number of different model specifications (for a detailed description see

20We say ”partly” since a pure model-based approach would include the world interest

rate and possibly the world price level as an exogenous variable.

Page 124: Macroeconomics of small open economies

106 CHAPTER 3

appendix 3.D). Note that in some of these specifications we omit money from

the model. In our view, this is justifiable as most central banks nowadays

use a short-term nominal interest rate as their policy instrument. Therefore,

monetary aggregates such as M1 have lost some of their importance for the

conduct of (and the measurement of exogenous shocks to) monetary policy.

Adding an open economy block to Gali’s extended VAR involves the

following steps:

• Step 1: Data. We come up with series for the real effective exchange

rate,21 net exports, and the current account balance for the OECD

countries included in the set.

• Step 2: Univariate analysis. The series tott, nxt, and cat are

pretested. There is no country for which the hypothesis that tott fol-

lows a random walk (with or without drift) has to be rejected. There

are a few countries for which the hypothesis that nxt follows a ran-

dom walk without drift can be rejected (or at least the evidence is

mixed), namely AUS, GBR, JPN, and NZL. There are a few coun-

tries for which the hypothesis that cat follows a random walk without

drift can be rejected (or at least the evidence is mixed), namely AUS,

GBR, and NZL.

• Step 3: Model specification. Regardless of the findings of step 2 and

as a matter of principle we consider tott, nxt, and cat as I(1)-variables.

Moreover, we assume (for the time being) that ∆mt and ∆pt on the

one hand and rt and ∆pt+1 on the other hand are cointegrated with

cointegrating vectors [1,−1].

• Step 4: Estimating IRF and CC and computing bootstrap CI.

Same as in subsection 3.3.2.

The results for the structural VAR specifications including international

trade and finance variables (”VAR 1” to ”VAR 7”) are contained in figure

3 and 4 and in table 1. For details on the specifications compare appendix

3.D.

21As a matter of fact, the ratio of the import deflator to the export deflator would

have been the better proxy.

Page 125: Macroeconomics of small open economies

3.4. EXTENSIONS 107

USA As usual we start with reporting the results for the USA. Above,

we have pointed to the j-shaped response of employment which we observed

for Gali’s five-variable specification (”VAR 0”). This general pattern is not

preserved for the VARs including international trade and finance variables,

at least not for the larger ones. We make the following observations. First,

the inclusion of more variables leads to an upward shift of the impact effect.

Since the width of the confidence bands remains almost unaltered across

the alternative VAR specifications, we conclude that the more variables

are incorporated in the model the more likely it is that the impact effect

is insignificant. Second, while for ”VAR 0” the response of employment

to a technology shock gets positive after about 3 years, employment does

never hit the zero-line in the case of the seven specifications incorporating

openness variables. Finally, the negative (and insignificant) long-run effect

is more pronounced than it was for ”VAR 0”.

What about the CC estimates? For the models which do not include the

variable tott (i.e., ”VAR 1”, and ”VAR 2”), the CC remains clearly negative.

Since the CI is only moderately wider compared to the benchmark case, the

point estimates are significant, albeit narrowly. The inclusion of the variable

tott, however, leads to a significant widening of the CI such that for ”VAR

3” through ”VAR 7” the point estimates get insignificant. Moreover, for

the larger models, the CC is less negative and reaches values of about -0.2

to -0.3.

FRA, GBR, and ITA In the case of FRA, the inclusion of tott brings the

j-shaped response back (which we observed for the benchmark specification

and which got lost for ”VAR 0”). At the same time, the confidence bands

tend to become wider the larger the VAR is. For GBR, we observe no such

widening of the confidence bands. Interestingly, the inclusion of either cat

or nxt leads to a significant (negative) long-run effect, while in the medium

run the (negative) response is insignificant; incorporating tott brings about

a reversion of this pattern. In the case of ITA, the permanent (negative)

effect is a little bit less pronounced for the extended VARs and in some

cases the effect is barely significant (the width of the confidence band is

stable across the specifications).

In the case of FRA, the estimated CC for ”VAR 4” and ”VAR 6” is about

Page 126: Macroeconomics of small open economies

108 CHAPTER 3

as weakly negative as for ”VAR 0”; the point estimates are insignificant for

all considered specifications. In the cases of GBR and ITA, the findings for

CC are robust to the inclusions of openness variables.

CAN and JPN In the case of CAN, the inclusion of cat (or nxt) together

with tott in ”VAR 0” gives rise to a positive (albeit insignificant) impact

effect. Moreover, we observe a positive and significant medium to long-run

effect for ”VAR 1” and for the model versions which include tott but omit

∆mt. The weakly negative CC is insignificant for all seven specifications.

In the case of JPN, we observe no effect whatsoever.

AUS, CHE, ESP, and NLZ In the case of AUS, CHE, and ESP the

inclusion of international trade and finance variables does not affect the

response of employment, whereas in the case of NZL the inclusion of both

tott and cat lifts the impulse response considerably over the entire horizon.

While we have observed a pronounced and significant negative long-run

effect for the benchmark model and ”VAR 0”, the long-run effect now is

much weaker and not different from zero any more.

In the cases of AUS, CHE, and ESP, we hardly observe any effect on the

CC estimates whereas in the case of NZL the same value becomes narrowly

significant when tott is included (”VAR 3”) and – due to a pronounced

widening of the CI – insignificant for specifications which include cat (”VAR

2”, ”VAR 5”, and ”VAR 7”).

Summary In case of the USA, we observe for models which do not include

the variable tott a CC which is clearly negative and (narrowly) significant.

The inclusion of the variable tott leads to a pronounced widening of the

CI; the point estimates (which are less negative for the larger models) get

insignificant. For most of the remaining OECD countries included in our

set, the CC is negative and significant across all alternative specifications.

The only exceptions are CAN, JPN and FRA, for which the estimated CC

is insignificant across all alternative specifications and NZL, for which we

get an insignificant point estimate for models which include cat.

Page 127: Macroeconomics of small open economies

3.4. EXTENSIONS 109

3.4.2 Structural VECM approach

In a last step, we extend the previous two robustness checks (which involved

the addition of a nominal and an open-economy block) to the cointegrated

case with an arbitrary number of cointegrating relations and general linear

restrictions on the cointegration space. So far, we have (for all specifica-

tions) assumed that there are exactly two cointegrating relations with a

very specific form (see subsection 3.3.1). This restriction is now going to

be relaxed. To this aim, we apply a two stage procedure which involves the

following modifications:

• Step 3: Model specification. Consider the VECM

∆xt = Πxt−1 +∑p−1

i=1Γi∆xt−i + εt. (3.3)

where εt is assumed to be Gaussian white noise. All variables that

enter xt are assumed to be I(1). For lag-lenth determination we esti-

mate a VAR using the non-stationary xt (compare above). Then, in

a first stage, we perform a rank test (for details regarding the testing

strategy compare appendix 3.D).

• Step 4: Estimating IRF and CC and computing bootstrap CI.

In a second stage, we set the rank of Π in model (3.3) according to

the test results of the modified step 3 (the vectors are normalized

reasonably) and estimate the VECM with the estimation method for

structural VARs extended to the cointegrated case (compare appendix

3.B). This provides us with estimates of the MA-coefficients in model

(3.2), on the base of which we can compute point estimates of the IRF

and the CC and, finally, a measure for the estimation variability.

The findings for the VECM version of Gali’s five-variable specification

(henceforth ”VECM 0”) and for the various extensions (”VECM 1” through

”VECM 7”) can be found in figure 5 to 6 and in table 2. For half of the

countries investigated, the test results do not point out to the presence of

a cointegration rank equal to 2; these countries include AUS, ESP, FRA,

GBR, NZL, and the USA. For the countries for which the hypothesis of

rank 2 cannot be rejected, there is virtually no case for which the joint

hypothesis that the vectors [0, 0, 1,−1, 0] and [0, 0, 0,−1, 1] belong to the

cointegration space cannot be rejected.

Page 128: Macroeconomics of small open economies

110 CHAPTER 3

USA In the case of the USA, the response of employment to a technology

shock for ”VECM 0” can be summarized as follows. Compared to both the

benchmark model and the restricted counterpart ”VAR 0”, the decrease

after the (moderately) negative impact effect is more pronounced and more

prolonged. In the medium to long run, employment reverses as in ”VAR

0”, but then remains negative. The confidence band is rather wide. For

”VECM 1” the response is similar to the one for ”VECM 0”, while for

”VECM 2” (which involves cat) we observe a long-run response which is

somewhat more negative and a confidence band which is wider than for

”VECM 0”.

For ”VECM 3” and ”VECM 5” (which include, respectively, tott and

tott in combination with cat) we observe something like an inverted j-shaped

response in that the effect is positive on impact, then rises over the medium

run, and eventually gets negative. However, for both specifications the

dynamic response features a weak negative trend, which suggests that the

growth rate rather than the level of the series is affected by the shock. This

anomaly makes us suspicious and in what follows we disregard the evidence

for the concerned specifications.22

For ”VECM 4” and ”VECM 6” (which contain tott in combination with

nxt) the negative impact effect is less pronounced than for ”VECM 0”.

Moreover, we observe no subsequent decline in employment. For both spec-

ifications, only the impact effect is (narrowly) significant. For ”VECM 7”

(which involves mt in addition to tott and cat), we observe the same inverted

j-shaped response as for ”VECM 3” and ”VECM 5”; but now, the dynamic

response on employment converges in the very long-run to an admittedly

rather low value (-0.9).

What about the CC estimates? As can be read off the relevant table,

relaxing the number and kind of cointegrating relationships alone (”VECM

0”) leads to a modest widening of the CI. At the same time, the point

estimate rises to -.42 (compared to -.66 for the unrestricted version ”VAR

0”). Together this implies that the CC is not significantly different from

zero any more. Adding either nxt or cat and omitting mt (”VECM 1” and

”VECM 2”) brings about a further increase in the point estimate, while

22We would like to stress here that within the entire set of countries and model speci-

fications these are the only two cases where such a problem arises.

Page 129: Macroeconomics of small open economies

3.4. EXTENSIONS 111

the width of the CI is comparable to the one for ”VECM 0”. Adding tott(”VECM 4” and ”VECM 6”), however, leads to a significant decrease of the

point estimate to a value in the neighborhood of the one observed for the

benchmark model, while the CI still remains unaltered. Accordingly, the

negative correlations for ”VECM 4” and ”VECM 6” are only marginally

insignificant.

Remaining OECD countries in the set In the case of FRA, the impact

effect for ”VECM 0” is zero; thereafter, the response gets positive and sig-

nificant (even in the long run). Skipping mt and adding openness variables

leads to a moderately negative impact effect. The estimated CC is positive

for ”VECM 0” and weakly negative for the other VECM specifications; none

of the point estimates is significant. In the case of GBR, the response of

employment for ”VECM 0” is similar to the one for the benchmark model.

Within the specifications which include tott, employment gradually returns

to the long-run equilibrium. The CCs are significantly negative for all spec-

ifications.

In the case of CAN, the employment response estimated for ”VECM

0” is moderately negative on impact – as for ”VAR 0” –, but then quickly

gets positive and remains so in the long run. The inclusion of tott together

with either cat or nxt leads to a pronounced upward shift of the short-

run response. For the remaining countries (AUS, CHE, ESP, ITA, JPN,

NZL), the effect of relaxing the rank assumption and introducing additional

variables is modest (if anything).23

What about the estimated CCs? In the case of AUS, CHE and ITA, the

CCs are significantly negative for all specifications, as for CAN and JPN

the CCs (some of which are positive) are always insignificant. In the case

of NZL, the inclusion of nxt leads to a moderate widening of the CI; for

those specifications, the point estimates are not significantly different from

zero. Note that the pronounced widening of the CI which we observed for

the VAR specifications involving cat is not apparent anymore.

23Within this subset of countries, there is exactly one specification for which we observe

a j-shaped response, namely ESP ”VECM 1”, rank 3.

Page 130: Macroeconomics of small open economies

112 CHAPTER 3

Summary In the case of the USA, relaxing the number and kind of cointe-

grating relationships and including international trade and finance variables

leads for the smaller models (which do not include tott) to an estimated CC

which is close to zero. For the larger models (which include tott) the point

estimate is in the neighborhood of the one reported for the benchmark

specification, while the CI is wider compared to the benchmark case such

that the estimated negative correlation are marginally significant. For the

majority of the investigated countries, however, the finding that the CC is

negative and significant is robust across the alternative specifications. Only

for CAN, FRA, and JPN we find nonsignificant point estimates; but this

finding too is robust across the alternative specifications.

3.5 Conclusions

In this paper, we raised the question whether the negative response of em-

ployment to productivity shocks estimated by Gali [13] is robust to the

inclusion of international trade and finance variables and to relaxing the

restrictions on the number and kind of cointegrating relationships imposed

in Gali’s higher-dimensional VAR framework. Results based on quarterly

data for a set of 10 OECD countries suggest that there is no systematic

relationship between the provision for variables that catch openness aspects

and the size of the conditional correlation coefficient. Also, we find that

going from a standard structural VAR approach to a two stage procedure

within a VECM framework does not alter Gali’s evidence.

On several occasions we have pointed to the fact that, since the shock we

identify is in principle a conglomerate of disturbances which permanently

affect measured labor productivity, any inference referring to an accurately

measured, genuine technology shock is likely to be distorted. In a follow-

up project to this study, we plan to disentangle the potential sources of

long-term shifts in average productivity, with particular emphasis on the

discrimination between technological and terms-of-trade disturbances. This

will require incorporating openness variables (as we did in this study) and

making additional identifying assumptions.

We have also pointed to the fact that employment in manufacturing, the

series which we made use of, is not an ideal proxy for overall labor input.

Page 131: Macroeconomics of small open economies

3.5. CONCLUSIONS 113

Drawing on more specific data bases should allow us to repeat the exercise

with total hours in non-farm business activities, at least for the subset of

the countries investigated in our study for which such a series exist. In the

case of Switzerland, for instance, we are aware of two (inofficial) estimates

of total hours, whereas in the case of the U.S. we could utilize the same

series as Gali [13] does (drawn from Citibase).

Page 132: Macroeconomics of small open economies

114 CHAPTER 3

3.A Univariate unit root tests

In practice, the specification of a time series process is carried out empir-

ically by means of unit root tests. As it is well known, the asymptotic

distributions of statistics from regressions involving unit roots are sensitive

to the presence of deterministic regressors. It matters whether we include

an intercept, an intercept plus a time trend, or neither an intercept nor

a time trend (see e.g. Hamilton [16]. As a consequence, one has to fol-

low a full-fledged estimation strategy rather than computing a single test

statistic. The testing strategy we apply is a pragmatic one and provides an

alternative to more involved (but in our view more error prone) strategies

as proposed e.g. by Enders [10]; it is largely based on Elder and Kennedy

[9].24

A crucial element of Elder and Kennedy’s testing strategy is making use of

prior knowledge regarding the growth status of a series based on theory.25 In

the case at hand, we presume that (yt −nt) is growing, that the status of nt

(measured by employment) is unknown, and that HP-filtered employment

is by definition not growing. Our priors for the other series are: (i) mt,

(mt − pt): growing, (ii) nxt, cat, ∆pt, (rt − ∆pt+1): not growing, (iii) tott:

unknown status (that is – as with employment – we cannot rule out that

there is some kind of growth, either positive or negative).

Another well-known fact is that in the presence of structural breaks, the

ADF and PP test statistics are biased toward the nonrejection of a unit

root (compare e.g. Enders [10], pp. 243-251). We suspect that some of the

series might indeed exhibit trend breaks. However, unit root tests that allow

for the presence of structural breaks assume that we know something about

when the break occurs (either on theoretical or on empirical grounds), but

in the case at hand we have no clue. Similar reasoning holds in the presence

of heteroscedasticity. We conclude that our test results might be biased

24We have to extend Elder and Kennedy [9] in some respects, among others in order

to permit the case where a series is integrated of higher order. A detailed description of

the applied testing strategy is available upon request.25Elder and Kennedy [9] also mention inspecting the data as a source of prior knowl-

edge. However, we trust more in theory than in visual checks.

Page 133: Macroeconomics of small open economies

3.B. ESTIMATION OF STRUCTURAL VARS AND VECMS 115

toward the nonrejection of a unit root.

3.B Estimation of structural VARs and

VECMs

Lenz [26] proposes a rather flexible method for estimating structural VARs

with a wide range of identification schemes. In subsection 3.B.1, we present

Lenz’s original estimation method for the non-cointegrated case. In subsec-

tion 3.B.2, we extend the method to the cointegrated case, thereby drawing

on results provided by Hoffmann [20] and Johansen [21].

3.B.1 Non-cointegrated case

Notation

Consider the covariance-stationary reduced form VAR(p)

D (L) xt = εt (3.4)(I − D1L − D2L

2 − ... − DpLp)xt = εt

xt = D1xt−1 + D2xt−2 + ... + Dpxt−p + εt

where xt is a n × 1 vector of economic variables, D (L) is a matrix finite-

order lag polynomial, and εt is a n × 1 vector of reduced form disturbances

which is characterized as follows

E (εt) = 0

E (εtε′

τ ) =

Σ for t = τ

0 otherwise

where Σ is a symmetric positive definite matrix.26 If |D (L) | has all its

characteristic roots greater than one in modulus, it is invertible and there

26In practice, of course, we allow for deterministic variables.

Page 134: Macroeconomics of small open economies

116 CHAPTER 3

exists a reduced form VMA representation

xt = C (L) εt

xt =(I + C1L + C2L

2 + ...)εt

where C (L) = D (L)−1.

Also, consider the structural VAR (p)

B (L) xt = ut(B0 − B1L − B2L

2 − ... − BpLp)xt = ut

B0xt = B1xt−1 + B2xt−2 + ... + Bpxt−p + ut

where E (utu′

t) = In. If B (L) is invertible there exists a structural VMA

representation

xt = A (L) ut

xt =(A0 + A1L + A2L

2 + ...)ut

where A (L) = B (L)−1.

Some useful corollaries

Substituting A (L) ut for xt in B (L) xt = ut yields ut = B (L) A (L) ut. It

follows that27

B0A0 = I.

Moreover, from A (L) = B (L)−1 follows that

A (1) = B (1)−1 .

Combining xt = C (L) εt and xt = A (L) ut yields

A (L) ut = C (L) εt. (3.5)

27From the fact that there are no lags on the right hand side of

ut =(B0 − B1L − B2L

2 − ... − BpLp) (

A0 + A1L + A2L2 + ...

)ut,

we conclude that B0A0 = I while the other terms equal zero.

Page 135: Macroeconomics of small open economies

3.B. ESTIMATION OF STRUCTURAL VARS AND VECMS 117

As (3.5) must hold for all t (and C0 = I), we have

A0ut = εt.

Substituting A0ut for εt in (3.5) yields

A (L) ut = C (L) A0ut

from which follows that A (1) = C (1) A0. Moreover, as A (L) ut = C (L) A0ut

must hold for all t, we have

Ai = CiA0 for i = 1, 2, ...

Key idea

The key idea of Lenz’s estimation method is to perform an orthogonal de-

composition of the observed residuals based on a total of n (n − 1) /2 re-

strictions on A0, B0, A (1), and B (1). This decomposition is accomplished

in two steps.

Step 1: Defining and computing S and defining Q

Let S be the lower triangular Cholesky decomposition of Σ, i.e., a lower

triangular matrix satisfying

Σ = SS ′.

Moreover, let Q be an arbitrary orthogonal matrix, i.e., a matrix satisfying28

Q′Q (= QQ′) = I and Q−1 = Q′.

From A0ut = εt follows that

E (εtε′

t) = E (A0utu′

tA′

0)

Σ = A0A′

0.

28In step 2, it will be shown that Q is uniquely determined – up to the sign of the

diagonal elements.

Page 136: Macroeconomics of small open economies

118 CHAPTER 3

Making use of the two matrices S and Q, we can derive a new expression

for A0 as follows

A0A′

0 = SQ (SQ)′ .

We end up with

A0 = SQ.

Step 2: Computing Q

In general, restrictions can be imposed on each of the four matrices A0,

B0, A (1), and B (1). We begin by expressing each of these four matrices

in terms of S, Q, and C (1). First, recall that A0 = SQ. Second, from

B0A0 = I we obtain

B′

0 = (S ′)−1

Q.

Third, from A (1) = C (1) A0 we immediately get

A (1) = C (1) SQ.

Finally, from B (1) = A (1)−1 we get

B (1)′ =[[C (1) S]′

]−1Q.

Let H be an n× 4n selection matrix whose function is to choose the appro-

priate restrictions on A0, B′

0, A (1), and B (1)′ (what is meant by ”choosing

the appropriate restrictions” will become clear further below), that is

H

S

(S ′)−1

C (1) S[[C (1) S]′

]−1

Q

︸ ︷︷ ︸matrix of restrictions

.

Then, we define

Z ≡ H

S

(S ′)−1

C (1) S[[C (1) S]′

]−1

.

Page 137: Macroeconomics of small open economies

3.B. ESTIMATION OF STRUCTURAL VARS AND VECMS 119

In order to solve for Q, all we need to do is re-ordering the endogenous

variables (and the associated shocks) of the VAR such that H times the

matrix of restrictions can be written in a lower triangular form.29 Let us

denote this lower triangular matrix by R

R ≡ H

S

(S ′)−1

C (1) S[[C (1) S]′

]−1

Q.

Since R = ZQ is lower triangular, it follows that

R′ = Q′Z ′

is upper triangular.30 Premultiplying the above equation by Q yields an

expression for Z ′

Z ′ = QR′.

Under the assumption that Z has full rank, the QR-decomposition of Z ′

yields Q. As the QR-decomposition is unique only up to the sign of the

diagonal elements of R′, the columns of R′ and Q can now be appropriately

normalized so that the IRF of the ith variable to the ith shock has the

desired sign.

Computing the structural VMA coefficients in practice

In practice, we start by re-ordering the variables in the system and defining

H to make sure that R equals a lower triangular matrix. Next, estimating

the reduced form VAR provides us with estimates of the reduced form VAR

coefficients Di for i = 1, 2, ..., p and Σ. Once we have estimates of Di

for i = 1, 2, ..., p, we can compute estimates of the VMA representation

coefficients Ci for i = 1, 2, ... (the necessary computations are carried out

automatically in econometric packages like RATS), and, thus, of C (1) (in

29We do not claim that the proposed method can handle all existing combinations of

short-run and long-run restrictions on the structural form of a model.30In general, if U is lower triangular, then U−1 is lower triangular, too, while U ′ is

upper triangular.

Page 138: Macroeconomics of small open economies

120 CHAPTER 3

fact a truncated version of it). Also, the Cholesky decomposition of Σ,

labelled S, can be computed. From H, S, and C (1), we can compute

Z. The QR-decomposition of Z ′ yields Q. A0 = SQ and Ai = CiA0 for

i = 1, 2, ... can be computed, in turn.

Illustrative example

Consider Gali’s [13] five-variable VAR. Suppose ∆(yt−nt) is the first variable

in vector xt. We assume that the technology shock is orthogonal to the

four non-technology shocks and only the technology shock has a long-run

effect on the level of yt − nt. We also assume that non-technology shocks

are orthogonal to each other and that they have an arbitrary recursive

structure. We end up with 10 additional restrictions (in addition to the

restriction that all shocks are orthogonal to each other). In terms of the

previously discussed framework, the matrix of restrictions is given by

A0

B′

0

A (1)

B (1)′

=

· · · · ·

· · 0 0 0

· · · 0 0

· · · · 0

· · · · ·

· · · · ·

· · · · ·

· · · · ·

· · · · ·

· · · · ·

· 0 0 0 0

· · · · ·

· · · · ·

· · · · ·

· · · · ·

· · · · ·

· · · · ·

· · · · ·

· · · · ·

· · · · ·

.

Page 139: Macroeconomics of small open economies

3.B. ESTIMATION OF STRUCTURAL VARS AND VECMS 121

And, the matrix which picks out the appropriate rows of the matrix ofrestrictions is equal to

H =

0 0 0 0 0

0 1 0 0 0

0 0 1 0 0

0 0 0 1 0

0 0 0 0 1

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

1 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

0 0 0 0 0

.

3.B.2 Cointegrated case

Notation and key idea

So far, the vector process xt has been assumed to be covariance-stationary.

Now, suppose all components of xt are I (1). If all components of xt are

I (1), the D (L)-polynomial in the reduced form VAR(p) given in equation

(3.4) has roots equal to one and is therefore not invertible. Model (3.4) can

be re-parameterized as follows

∆xt =∑p−1

i=1Γi∆xt−i + Πxt−1 + εt (3.6)

where Π = D (1) has rank 0 ≤ r ≤ n and εt is Gaussian white noise.

Suppose we know r (in practice, the rank of Π is determined empirically by

means of rank tests). Moreover, suppose 0 < r < n (the interesting case).

In this event, Π can be factorized as Π ≡ αβ′ where β and α are n × r

matrices, each with rank r. In addition, xt is cointegrated such that β′xt is

I (0). System (3.6) is then called a Vector Error Correction Model (VECM),

that is, a VAR that incorporates cointegrating restrictions.

From the VECM, it is possible to derive a VMA representation

∆xt = C (L) εt

where C (1) has rank (n − r). Since D (L) is not invertible, the derivation

of the C (L)-polynomial is not completely straightforward (compare e.g.

Watson [33], subsection 3.2). In practice, however, the computation of Ci

for i = 1, 2, ... can readily be implemented in standard software packages like

RATS. Note that Ci is restricted by the assumption regarding the number

of cointegrating relations.

Page 140: Macroeconomics of small open economies

122 CHAPTER 3

The extension of Lenz’s [26] estimation method to the non-stationary, coin-

tegrated case involves an additional step. The purpose of this step is to dis-

tinguish innovations which have permanent effects from those which have

transitory effects. This goal is accomplished by a transformation of the

residuals using information that are available from the VECM estimation

(i.e., from the Johansen procedure; compare Hoffmann [20]).

Step 1: Distinguishing innovations that have permanent effects

from those that have transitory effects

Johansen [21] has shown that in the cointegrated case, the long-run impact

matrix C (1) can be given by the following representation

C (1) = β⊥ (α′

⊥Γ (1) β⊥)

−1α′

where α⊥, β⊥ are the orthogonal complements of α and β. Notice: α′

⊥α = 0

and α′

⊥is (n − r) × n. Moreover, Johansen has shown that the vector of

permanent and transitory disturbances is given by

ηt =

[η1,t

η2,t

]=

[α′

α′Ω−1

]

n×n

εt

where subvector η1t contains the disturbances that have permanent effects on

the components of xt (the levels!), subvector η2,t contains the disturbances

that have temporary effects, and Ω is the variance-covariance matrix of the

reduced form disturbances.

Let us define

P−1 ≡

[α′

α′Ω−1

]

from which we get ηt = P−1εt. For any initial choice of α⊥ and α, we have

P =[

Ωα⊥ (α′

⊥Ωα⊥)−1 α (α′Ωα)−1

].

Moreover, we have

Σ ≡ E (ηtη′

t) = P−1Ω(P−1

)′=

[α′

⊥Ωα⊥ 0(n−r)×r

0r×(n−r) α′Ω−1α

].

Page 141: Macroeconomics of small open economies

3.B. ESTIMATION OF STRUCTURAL VARS AND VECMS 123

Note that Σ is positive definite.

This completes step 1, the orthogonalization of permanent and transitory

disturbances. However, permanent shocks among themselves and transitory

shocks among themselves are not yet orthogonal. This will be accomplished

within step 2 and 3.

Step 2 + 3: Orthogonalizing the previously distinguished shocks

among themselves

We are now in the position to re-formulate our problem such that it becomes

almost equivalent to the non-cointegrated case. We are looking at two

models. The first model is a reduced form VMA representation given by

∆xt = C (L) Pηt

∆xt =(I + C1L + C2L

2 + ...)Pηt.

It is computationally straightforward to get the matrix of long-run multi-

pliers to this model

C (1) P =[

β⊥ (α⊥Γ (1) β⊥)−1 0n×r

].

The first block with the dimension n × (n − r) corresponds to the long-run

multipliers for η1,t; the second block with the dimension of n× r corresponds

to the long-run multipliers for η2,t (a matrix of zeros). Let us define

Ψn×(n−r) ≡ β⊥ (α⊥Γ (1) β⊥)−1

from which we get

C (1) P =[

Ψn×(n−r) 0n×r

].

Note that Ψ can be partitioned into two blocks

Ψn×(n−r) =

[Ψ1 (n−r)×(n−r)

Ψ2 r×r

]

where Ψ1 corresponds to the long-run effects on the first n − r components

of xt (in levels).

Page 142: Macroeconomics of small open economies

124 CHAPTER 3

The second model is a structural VMA representation given by

∆xt = A (L) ut

∆xt =(A0 + A1L + A2L

2 + ...)ut

where E (utu′

t) = In.

Combining ∆xt = C (L) Pηt and ∆xt = A (L) ut yields

A (L) ut = C (L) Pηt. (3.7)

As (3.7) must hold for all t (and C0 = I), we have

A0ut = Pηt.

Substituting P−1A0ut for ηt in (3.7) yields

A (L) ut = C (L) A0ut

from which follows that A (1) = C (1) A0. Moreover, as A (L) ut = C (L) A0ut

must hold for all t, we have

Ai = CiA0 for i = 1, 2, ...

As in the stationary case, we are going to perform an orthogonal decompo-

sition of the η-residuals based on a total of n (n − 1) /2 restrictions on A0

and A (1). Let S be the Cholesky decomposition of Σ

Σ = SS ′.

Note that because Σ is block diagonal, its Cholesky decomposition is block

diagonal too

SS ′ =

[SuS

u (n−r)×(n−r) 0(n−r)×r

0r×(n−r) SlS′

l r×r

]

whereby each block consists of a lower triangular matrix. Moreover, let Q

be an arbitrary orthogonal matrix.

From A0ut = Pηt follows that

E (A0utu′

tA′

0) = E (Pηtη′

tP′)

A0A′

0 = PΣP ′.

Page 143: Macroeconomics of small open economies

3.B. ESTIMATION OF STRUCTURAL VARS AND VECMS 125

Making use of the two matrices S and Q, we can derive a new expression

for A0 as follows

P−1A0A′

0(P′)−1 = SQ (SQ)′

A0 = PSQ.

This represents the first of two conditions relating the two models. The

second one follows directly from A (1) = C (1) A0 and is given by

A(1) = C(1)PSQ.

In analogy to the non-cointegrated case, let H be an n×2n selection matrix

whose function is to choose the appropriate restrictions on A0 and A (1),

that is

H

[PS

C(1)PS

]Q.

We define

Z ≡ H

[PS

C (1) PS

].

In order to solve for Q, all we need to do is re-ordering the endogenous

variables (and the associated shocks) of the VAR such that H times the

matrix of restrictions can be written in a lower triangular form; let us

denote this matrix by R

R ≡ H

[PS

C (1) PS

]Q = ZQ.

Note that the range of specifications of H is restricted by the fact that

C (1) P =[

Ψn×(n−r)

∣∣ 0n×r

]; as a consequence, long-run restrictions on

A(1) can only be imposed on the first n − r variables. Moreover, note that

since S is block diagonal, Z is block diagonal, too

Z =

[Zu (n−r)×(n−r) 0(n−r)×r

0r×(n−r) Zl r×r

].

Under the assumption that both Zu and Zl, have full rank, the QR-decomposition

of Z ′

u yields Qu and similarly Z ′

l yields Ql, from which we finally get

Q =

[Qu (n−r)×(n−r) 0(n−r)×r

0r×(n−r) Ql r×r

].

Page 144: Macroeconomics of small open economies

126 CHAPTER 3

Computing the structural VMA coefficients in practice

In practice, we start by re-ordering the variables in the system and defining

H to make sure that R equals a lower triangular matrix. Next, estimating

the VECM (based on a rank assumption) provides us with estimates of

Π, Γi for i = 1, 2, ..., p − 1, and Ω. We can then compute P . From this

we get estimates of the VMA representation coefficients Ci for i = 1, 2, ...

and ηt, from which we get estimates of Σ in turn. Then, the Cholesky

decomposition of Σ, labelled S, can be computed. From H, S, P , and

C (1), we can compute Z. With the QR-decompositions of Zu′

and Zl′

at

our disposal, we get Q. A0 = P SQ and Ai = CiA0 for i = 1, 2, ... can be

computed, in turn.

3.C Bootstrap confidence intervals

Estimating a structural VAR or VECM by means of the estimation method

presented in appendix 3.B provides us with point estimates of the impulse

response functions (IRF) and conditional correlations (CC). Next, we would

like to come up with a measure for the estimation variability. We find

that percentile confidence intervals (CI) are particularly well suited for this

purpose. The chosen strategy is based on bootstrapping the VAR and was

originally proposed by Runkle [30].

Consider a covariance-stationary, reduced form VAR(p)

xt = γ + D1xt−1 + D2xt−2 + ... + Dpxt−p + εt (3.8)

where γ is a vector of constant terms. By estimating this VAR, we get the

coefficient estimates γ and the stacking matrix

D ≡[D′

1, D′

2, ..., D′

p

]

as well as the series of the estimated residuals, εtTt=1.

In the words of Runkle [30], pp. 438/9, ”[t]he basic insight behind the

bootstrap is that since the estimated residuals of the model are a represen-

tative sample of the true disturbances [i.e., they are i.i.d.], it should not

Page 145: Macroeconomics of small open economies

3.C. BOOTSTRAP CONFIDENCE INTERVALS 127

matter in what order the disturbances occur. This means that the distri-

bution of the estimator can be determined by generating large numbers of

artificial observations from the actual data and the estimated residuals.”

To guarantee that the (empirical) distribution function of the ε’s has mean

zero, we subtract ε = 1T

∑Tt=1 εt from every single εt. With draws (without

replacement) from the shaken, mean adjusted estimated residuals and the

coefficient estimates γ and D we generate 1000 artificial series

x∗

t = γ + D1x∗

t−1 + D2x∗

t−2 + ... + Dpx∗

t−p + (ε∗t − ε)

where x∗

t is the simulated value (the first p observations are taken as initial

conditions) and ε∗t is a draw from the shaken estimated residuals, ε.

We end up with 1000 simulated series x∗

tTt=1. For each individual series, we

estimate D∗ =[D∗′

1 , D∗′

2 , ..., D∗′

p

]; in each round, we compute – based on D∗

and the previously discussed identifying assumptions – the structural VMA

coefficients,[A∗′

1 , A∗′

2 , ..., A∗′

k

]where k is the chosen horizon. Finally, we can

compute the (empirical) relative distribution function for each element in

A∗

k.

We use the α/2 and (1 − α) /2 percentile points of the distribution functions

as confidence intervals. An example should make this point clearer. Note

that the empirical relative distribution function is computed for every single

element in A, that is every

aij,k =∂xi,t+k

∂uj,t

for i, j = 1, 2, ..., n and k = 1, 2, ..., k. In the case of a 0.90 confidence, the

interval is given by element 1000 ∗ 0.10/2 = 50 and element 1000 ∗ (1 −

0.10/2) = 950.

Runkle’s method has been criticized for being affected by a small sample

bias. The small sample bias distorts the initial VAR coefficient estimates;

the bootstrapped estimates are then biased again - usually towards the sta-

tionary region (compare Kursteiner [25]). As a consequence, it can happen

that the confidence bands constructed in this way do not contain the orig-

inal parameter estimates. To overcome this problem, Kilian [22] suggests

Page 146: Macroeconomics of small open economies

128 CHAPTER 3

a small-sample bias correction. The planned follow-up study will contain

Kilian’s algorithm.31

3.D Model specifications

VAR models

Gali’s bivariate VAR in first differences is given by

∆xt = D1∆xt−1 + D2∆xt−2 + ... + Dp∆xt−p + εt

where

xt =[

(yt − nt) nt

]′.

We term it the ”benchmark model”. Lag-length is tested for the VAR in

levels.

Gali’s five-variable VAR has the form

xt = D1xt−1 + D2xt−2 + ... + Dpxt−p + εt

where

xt =[

∆ (yt − nt) ∆nt ∆ (mt − pt) ∆2pt rt − ∆pt+1

]′.

We term it ”VAR 0”. Lag-length is tested for

xt =[

(yt − nt) nt mt − pt ∆pt rt

]′.

We consider a number of modifications of ”VAR 0”. The data vectors of

the estimated VARs are listed below. The models are numbered from 1 to

7. Lag length determination is analogous to that for ”VAR 0”.

1.

xt =[

∆ (yt − nt) ∆nt ∆2pt rt − ∆pt+1 ∆nxt

]′

31A detailed description of the strategy to be applied and its implementation is available

upon request.

Page 147: Macroeconomics of small open economies

3.D. MODEL SPECIFICATIONS 129

2.

xt =[

∆ (yt − nt) ∆nt ∆2pt rt − ∆pt+1 ∆cat

]′

3.

xt =[

∆ (yt − nt) ∆nt ∆2pt rt − ∆pt+1 ∆tott]′

4.

xt =[

∆ (yt − nt) ∆nt ∆2pt rt − ∆pt+1 ∆nxt ∆tott]′

5.

xt =[

∆ (yt − nt) ∆nt ∆2pt rt − ∆pt+1 ∆cat ∆tott]′

6.

xt =[

∆ (yt − nt) ∆nt ∆ (mt − pt) ∆2pt rt − ∆pt+1 ∆nxt ∆tott]′

7.

xt =[

∆ (yt − nt) ∆nt ∆ (mt − pt) ∆2pt rt − ∆pt+1 ∆cat ∆tott]′

VECM models

Data vectors Consider the following VECM

∆xt = Πxt−1 +∑p−1

i=1Γi∆xt−i + εt

where all variables entering xt are supposed to be I(1). The data vectors

of the estimated VECMs are listed below. The models are again numbered

from 0 to 7. Lag length determination is analogous to that for ”VAR 0”.

0.

xt =[

(yt − nt) nt ∆mt ∆pt rt

]′

1.

xt =[

(yt − nt) nt ∆pt rt nxt

]′

Page 148: Macroeconomics of small open economies

130 CHAPTER 3

2.

xt =[

(yt − nt) nt ∆pt rt cat

]′

3.

xt =[

(yt − nt) nt ∆pt rt tott]′

4.

xt =[

(yt − nt) nt ∆pt rt nxt tott]′

5.

xt =[

(yt − nt) nt ∆pt rt cat tott]′

6.

xt =[

(yt − nt) nt ∆mt ∆pt rt nxt tott]′

7.

xt =[

(yt − nt) nt ∆mt ∆pt rt cat tott]′

Rank tests For each VECM, rank tests are performed. Thereby, we

proceed as follows. We begin by running CATS in RATS. Then, we check

the adequacy of the model regarding time independence of the residuals,

etc. (for a motivation compare Johansen [21], pp. 20/1). This includes:

• A visual check of the autocorrelation and the cross-correlation func-

tions for the individual residual series.

• A visual check of the standardized residuals and the histograms.

• A formal analysis of the residuals: (i) LM test for residual auto-

correlation of order 1 and 4, respectively. (ii) For each individual

residual series descriptive statistics are computed (like standard de-

viation, skewness, excess kurtosis) and a normality test is performed

(Shenton-Brownman test, known as Jarque-Bera test). (iii) More-

over, a modified version of the Shenton-Brownman test for normality

of the individual series is performed (for a description see Hansen and

Juselius [19], p. 27 and p. 73).

• A visual check of the eigenvalues of the companion matrix.

Page 149: Macroeconomics of small open economies

3.D. MODEL SPECIFICATIONS 131

The correlogram together with the (uni- and multivariate) normality test

statistics are our main criteria for assessing the adequacy of the model. If

the residuals are far from white noise, the number of lags is augmented.

However, we do not do this mechanically but look for a compromise be-

tween white noise criterium and degrees-of-freedom considerations.32 For

the smaller models, we consider 8 lags as being a rather large number; in

the case of ”VECM 6” and ”VECM 7” we try not to go beyond 4 lags.

If there is no hint of misspecification, we perform rank tests (table 1 in

Osterwald-Lenum [29]). We work with 90% critical values. Then, we set the

rank of Π according to the test result and normalize the vectors reasonably.

Finally, we perform some checks to see whether the supposed number of

cointegrating relationships is valid. Criteria: same as above; in addition to

this: visual check of the cointegrating relationships.

Testing restrictions on the β-vector An informal visual check allows

us to see whether a specified structural relation is contained in the space

spanned by β. Moreover, the α coefficient indicate which cointegrating

relationship is important for a variable. To make these statements more

precise we perform a series of formal tests. The testing strategy depends

on the specification of a model (number of endogenous variables and of

cointegrating relationships):

• Model 0: If the test points out to the presence of a cointegration

rank > 2, we directly go on by estimating a structural VECM. If

r = 2, we test the joint hypothesis that the vectors [0, 0, 1,−1, 0] and

[0, 0, 0,−1, 1] belong to the cointegration space. If r = 1, we test the

two hypothesis individually. If the matrix Π is the null matrix, the

model can be estimated as a VAR in first-differences.

• Model 1-3: If r > 1, we directly go on by estimating a structural

VECM. If r = 1, we test the hypothesis that the vector [0, 0,−1, 1, 0]

32Johansen [21], p. 21: ”It is our experience that if a long lag length is required to get

white noise residuals then it often pays to reconsider the choice of variables, and look

around for another important explanatory variable to include in the information set.”

Page 150: Macroeconomics of small open economies

132 CHAPTER 3

belongs to the cointegration space. If the matrix Π is the null matrix,

the model can be estimated as a VAR in first-differences.

• Model 4+5: If r > 1 we directly go on by estimating a structural

VECM. If r = 1, we test the hypothesis that the vector [0, 0,−1, 1, 0, 0]

belongs to the cointegration space. If the matrix Π is the null matrix,

the model can be estimated as a VAR in first-differences.

• Model 6+7: If r > 2, we directly go on by estimating a structural

VECM. If r = 2, we test the joint hypothesis that the vectors

[0, 0, 1,−1, 0, 0, 0] and [0, 0, 0,−1, 1, 0, 0]

belong to the cointegration space. If r = 1, we test the two hypothesis

individually. If the matrix Π is the null matrix, the model can be

estimated as a VAR in first-differences.

Page 151: Macroeconomics of small open economies

Bibliography

[1] Altig, David, Laurence J. Christiano, Martin Eichenbaum, and Jesper

Linde (2002), Technology shocks and aggregate fluctuations, Working

paper (preliminary and incomplete), Federal Reserve Bank of Cleveland

[2] Basu, Susanto, John G. Fernald, and Miles Kimball (1998), Are tech-

nology improvements contractionary?, Board of Governors of the Fed-

eral Reserve System, International Finance Discussion Paper 625

[3] Baxter, Marianne, and Rober G. King (1993), Fiscal policy in general

equilibrium, The American Economic Review 83(3), 315-34

[4] Blanchard, Olivier J., and Danny Quah (1989), The dynamic effects of

aggregate demand and supply disturbances, American Economic Re-

view 79(4), 655-673

[5] Christiano, Lawrence J., and Martin Eichenbaum (1992), Current real

business cycle theories and aggregate labor market fluctuations, Amer-

ican Economic Review 82(3), 430-450

[6] Christiano, Laurance J., Martin Eichenbaum, and Robert Vigfusson

(2003), What happens after a technology shock?, NBER Working Pa-

per No. 9819

[7] Collard, Fabrice, and Harris Dellas (2002), Technology shocks and em-

ployment, CEPR Discussion Paper No. 3680

[8] Cooley, Thomas F., and Edward C. Prescott (1995), Economic growth

and business cycles, in: Thomas F. Cooley, ed., Frontiers of Business

Cycle Research, Princeton University Press

Page 152: Macroeconomics of small open economies

134 CHAPTER 3

[9] Elder, John, and Peter E. Kennedy (2001), Testing for unit roots: what

should students be taught?, Journal of Economic Education 32(2), 137-

146

[10] Enders, Walter (1995), Applied econometric times series, Wiley Text

Books

[11] Fisher, Jonas D.M. (2003), Technology shocks matter, Federal Reserve

Bank of Chicago, Working Paper No. 2002-14

[12] Francis, Neville, and Valerie A. Ramey (2002), Is the technology-driven

real business cycle hypothesis dead? Shocks and aggregate fluctuations

revisited, NBER Working Paper No. 8726

[13] Gali, Jordi (1999), Technology, employment, and the business cycle: do

technology shocks explain aggregate fluctuations?, American Economic

Review 89(1), 249-271

[14] Gali, Jordi (2003), On the role of technology shocks as a source of

business cycles: Some new evidence, Mimeo

[15] Gonzalo, Jesus, and Serena Ng (2001), A systematic framework for ana-

lyzing the dynamic effects of permanent and transitory shocks, Journal

of Economic Dynamics and Control 25, 1527-1546

[16] Hamilton, James D. (1994), Time series analysis, Princeton University

Press

[17] Hansen, Gary D. (1985), Indivisible labor and the business cycle, Jour-

nal of Monetary Econonomics 16(3), 309-327

[18] Hansen, Gary D., and Randall Wright (1992), The labor market in real

business cycle theory, Federal Reserve Bank of Minneapolis Quarterly

Review (Spring), 2-12

[19] Hansen, Henrik, and Katarina Juselius (1995), CATS in RATS: Coin-

tegration analysis of time series

[20] Hoffmann, Mathias (2001), Long run recursive VAR models and QR

decompositions, Economics Letters 73(1), 15-20

Page 153: Macroeconomics of small open economies

BIBLIOGRAPHY 135

[21] Johansen, Soren (1995), Likelihood-based inference in cointegrated vec-

tor autoregressive models, Oxford University Press

[22] Kilian, Lutz (1998), Small-sample confidence intervals for impulse re-

sponse functions, Review of Economics and Statistics 80(2), 218-230

[23] King, Robert G., and Sergio T. Rebelo (1999), Rescuscitating real busi-

ness cycles, in: John B. Taylor and Michael Woodford, eds., Handbook

of Macroeconomics Vol. 1, Elsevier Science

[24] Kollmann, Robert (2002), Monetary policy rules in the open economy:

effects on welfare and busincess cycles, Journal of Monetary Economics

49(5), 989-1015

[25] Kuersteiner, Guido (undated manuscript), Lecture 6: Additional re-

sults for VAR’s, Department of Economics, MIT

[26] Lenz, Carlos (forthcoming), A simple method for identifying structural

VAR models using combinations of short- and long-run restrictions,

Mimeo, University of Basel.

[27] Lutkepohl, Helmut (1991), Introduction to multiple time series analy-

sis, Springer-Verlag

[28] McCallum, Bennett T. (1993), Unit roots in macroeconomic time se-

ries: some critical issues, Federal Reserve Bank of Richmond Economic

Quarterly 79(2), 13-43

[29] Osterwald-Lenum, Michael (1992), A note with quantiles of the asymp-

totic distribution of the maximum likelihood cointegration rank test

statistics, Oxford Bulletin of Economics and Statistics 54(3), 461-472

[30] Runkle, David E. (1987), Vector autoregressions and reality, Journal

of Business and Economics Statistics 5(4), 437-454

[31] Shea, John (1998), What do technology shocks do?, NBER Working

Paper No. 6632

[32] Uhlig, Harald (2003), Do technology shocks lead to a fall in total hours

worked?, Mimeo

Page 154: Macroeconomics of small open economies

136 CHAPTER 3

[33] Watson, Mark W. (1994), Vector autoregressions and cointegration, in:

Robert F. Engle and Daniel L. McFadden, Handbook of Econometrics

Vol. 4, Elsevier Science

Page 155: Macroeconomics of small open economies

TABLES AND FIGURES CHAPTER 3 137

Tables and figures chapter 3

Table 1a-e: Structural VAR

Table 2a-j: Structural VECM

Figure 1a-c: Benchmark model

Figure 2a-c: VAR 0

Figure 3a+b: USA, VAR 1-7

Figure 4a-c: VECM 0

Figure 5a+b: USA, VECM 1-7

Figure 6a-i: Remaining countries in the set, VECM 1-7

Page 156: Macroeconomics of small open economies

FRA

Nbe

nchm

ark

VA

R 0

VA

R 1

VA

R 2

VA

R 3

VA

R 4

VA

R 5

VA

R 6

VA

R 7

T88

8179

-79

79-

79-

ES:

FR

OM

1980

.219

78.4

1979

.2-

1979

.219

79.2

-19

79.2

-

TO20

02.1

1998

.419

98.4

-19

98.4

1998

.4-

1998

.4-

L8

24

44

4

CC

-0.3

05-0

.528

-0.3

91-

-0.4

14-0

.285

--0

.263

-St

d0.

141

0.31

30.

281

-0.

279

0.26

7-

0.28

7-

UpB

-0.0

120.

159

0.22

3-

0.18

80.

234

-0.

228

-Lo

B-0

.467

-0.8

37-0

.630

--0

.637

-0.5

80-

-0.6

42-

¦ Lo

B - U

pB ¦

0.45

50.

996

0.85

3-

0.82

50.

814

-0.

870

-

Gal

i (19

99)

CC

-0.8

1-

Std

0.27

-

GB

Rbe

nchm

ark

VA

R 0

VA

R 1

VA

R 2

VA

R 3

VA

R 4

VA

R 5

VA

R 6

VA

R 7

T16

373

158

158

115

117

115

7575

ES:

FR

OM

1961

.219

83.4

1962

.319

62.3

1973

.219

72.4

1973

.219

83.2

1983

.2

TO20

01.4

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

L4

48

84

24

22

CC

-0.8

71-0

.799

-0.7

96-0

.813

-0.7

31-0

.738

-0.7

32-0

.911

-0.8

94St

d0.

092

0.20

90.

173

0.17

40.

140

0.22

30.

142

0.14

70.

113

UpB

-0.6

71-0

.415

-0.2

44-0

.250

-0.4

20-0

.162

-0.3

97-0

.599

-0.6

18Lo

B-0

.930

-0.9

00-0

.783

-0.7

95-0

.840

-0.8

76-0

.826

-0.9

58-0

.958

¦ Lo

B - U

pB ¦

0.25

90.

485

0.53

90.

545

0.42

00.

714

0.42

90.

359

0.34

0

Gal

i (19

99)

CC

-0.9

1-

(UK

)St

d.0.

16-

Not

es: I

n th

e ca

se o

f FR

Ai,

n(t

) is p

roxi

ed b

y H

P fil

tere

d em

ploy

men

t and

r(t

) by

a sh

ort-t

erm

inte

rest

rate

(3-m

onth

PIB

OR

).In

the

case

of G

BR

, r(t)

is p

roxi

ed b

y a

long

-term

rate

(10-

year

gov

ernm

ent b

onds

).Th

e #

of o

bser

vatio

ns (T

) dep

ends

on

the

chos

en #

of l

ags (

L) a

nd o

n da

ta a

vaila

bilit

y. E

S st

ands

for e

ffec

tive

sam

ple,

CC

for t

he c

orre

latio

n be

twee

n

aver

age

labo

r pro

duct

ivity

gro

wth

and

em

ploy

men

t gro

wth

con

ditio

nal o

n a

posi

tive

tech

nolo

gy sh

ock,

UpB

for u

pper

bou

nd (0

.95

perc

entil

e po

int),

and

LoB

for l

ower

bou

nd (0

.05

perc

entil

e po

int).

Sta

ndar

d de

viat

ions

(Std

) are

incl

uded

in o

rder

to m

ake

our f

indi

ngs c

ompa

rabl

e to

thos

e of

Gal

i (19

99).

Tabl

e 1a

: Stru

ctur

al V

AR

Page 157: Macroeconomics of small open economies

ITA

benc

hmar

kV

AR

0V

AR

1V

AR

2V

AR

3V

AR

4V

AR

5V

AR

6V

AR

7T

120

7777

-79

79-

79-

ES:

FR

OM

1972

.219

79.4

1979

.4-

1979

.219

79.2

-19

79.2

-

TO20

02.1

1998

.419

98.4

-19

98.4

1998

.4-

1998

.4-

L8

44

-2

2-

2-

CC

-0.9

61-0

.955

-0.9

39-

-0.9

75-0

.972

--0

.977

-St

d0.

082

0.09

70.

148

-0.

097

0.15

2-

0.13

5-

UpB

-0.7

40-0

.732

-0.5

95-

-0.7

54-0

.670

--0

.685

-Lo

B-0

.969

-0.9

65-0

.966

--0

.989

-0.9

85-

-0.9

86-

¦ Lo

B - U

pB ¦

0.22

90.

233

0.37

1-

0.23

50.

315

-0.

301

-

Gal

i (19

99)

CC

-0.4

7-

Std.

0.12

-

USA

benc

hmar

kV

AR

0V

AR

1V

AR

2V

AR

3V

AR

4V

AR

5V

AR

6V

AR

7T

156

162

162

162

115

115

115

115

115

ES:

FR

OM

1963

.219

61.3

1961

.319

61.3

1973

.219

73.2

1973

.219

73.2

1973

.2

TO20

02.1

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

L12

44

44

44

44

CC

-0.8

20-0

.658

-0.7

70-0

.763

-0.6

82-0

.622

-0.5

41-0

.288

-0.1

87St

d0.

161

0.21

20.

260

0.25

60.

319

0.29

40.

297

0.32

60.

342

UpB

-0.2

71-0

.105

-0.0

46-0

.055

0.30

60.

242

0.26

10.

500

0.57

6Lo

B-0

.789

-0.7

94-0

.891

-0.8

87-0

.735

-0.7

27-0

.697

-0.5

89-0

.577

¦ Lo

B - U

pB ¦

0.51

80.

689

0.84

50.

832

1.04

10.

969

0.95

81.

089

1.15

3

Gal

i (19

99)

CC

-0.8

4-0

.82

Std.

0.26

0.08

Not

es: I

n th

e ca

se o

f ITA

, r(t

) is p

roxi

ed b

y a

shor

t-ter

m in

tere

st ra

te, i

n th

e ca

se o

f the

USA

by

a lo

ng-te

rm ra

te.

The

# of

obs

erva

tions

(T) d

epen

ds o

n th

e ch

osen

# o

f lag

s (L)

and

on

data

ava

ilabi

lity.

ES

stan

ds fo

r eff

ectiv

e sa

mpl

e, C

C fo

r the

cor

rela

tion

betw

een

av

erag

e la

bor p

rodu

ctiv

ity g

row

th a

nd e

mpl

oym

ent g

row

th c

ondi

tiona

l on

a po

sitiv

e te

chno

logy

shoc

k, U

pB fo

r upp

er b

ound

(0.9

5 pe

rcen

tile

poin

t),an

d Lo

B fo

r low

er b

ound

(0.0

5 pe

rcen

tile

poin

t). S

tand

ard

devi

atio

ns (S

td) a

re in

clud

ed in

ord

er to

mak

e ou

r fin

ding

s com

para

ble

to th

ose

of G

ali (

1999

).

Tabl

e 1b

: Stru

ctur

al V

AR

(Con

t.)

Page 158: Macroeconomics of small open economies

CA

Nbe

nchm

ark

VA

R 0

VA

R 1

VA

R 2

VA

R 3

VA

R 4

VA

R 5

VA

R 6

VA

R 7

T80

7981

7981

8181

8181

ES:

FR

OM

1982

.219

82.2

1981

.419

82.2

1981

.419

81.4

1981

.419

81.4

1981

.4

TO20

02.1

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

L4

42

42

22

22

CC

-0.2

65-0

.537

-0.0

08-0

.042

-0.0

36-0

.055

-0.0

670.

330

0.22

3St

d0.

277

0.35

20.

315

0.41

60.

329

0.32

20.

327

0.34

80.

340

UpB

0.19

80.

419

0.46

90.

679

0.47

30.

467

0.43

90.

736

0.65

7Lo

B-0

.720

-0.7

30-0

.582

-0.7

02-0

.628

-0.6

01-0

.671

-0.4

48-0

.525

¦ Lo

B - U

pB ¦

0.91

81.

149

1.05

11.

381

1.10

11.

068

1.11

01.

184

1.18

2

Gal

i (19

99)

CC

-0.5

9-

Std.

0.32

-

JPN

benc

hmar

kV

AR

0V

AR

1V

AR

2V

AR

3V

AR

4V

AR

5V

AR

6V

AR

7T

8483

8363

8385

6585

65ES

: F

RO

M19

81.2

1981

.219

81.2

1986

.219

81.2

1980

.419

85.4

1980

.419

85.4

TO

2002

.120

01.4

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

2001

.4L

44

44

42

22

2

CC

-0.1

220.

355

0.28

60.

190

0.02

40.

285

0.53

50.

498

0.61

6St

d0.

281

0.39

00.

404

0.33

60.

372

0.43

30.

381

0.47

20.

426

UpB

0.22

60.

563

0.79

10.

619

0.52

20.

735

0.80

10.

781

0.84

2Lo

B-0

.696

-0.6

85-0

.628

-0.4

98-0

.717

-0.7

08-0

.449

-0.7

66-0

.515

¦ Lo

B - U

pB ¦

0.92

21.

248

1.41

91.

117

1.23

91.

443

1.25

01.

547

1.35

7

Gal

i (19

99)

CC

-0.0

7-

Std.

0.08

-N

otes

: In

the

case

of b

oth

CA

N a

nd JP

N, r

(t) i

s pro

xied

by

a sh

ort-t

erm

inte

rest

rate

.Th

e #

of o

bser

vatio

ns (T

) dep

ends

on

the

chos

en #

of l

ags (

L) a

nd o

n da

ta a

vaila

bilit

y. E

S st

ands

for e

ffec

tive

sam

ple,

CC

for t

he c

orre

latio

n be

twee

n

aver

age

labo

r pro

duct

ivity

gro

wth

and

em

ploy

men

t gro

wth

con

ditio

nal o

n a

posi

tive

tech

nolo

gy sh

ock,

UpB

for u

pper

bou

nd (0

.95

perc

entil

e po

int),

and

LoB

for l

ower

bou

nd (0

.05

perc

entil

e po

int).

Sta

ndar

d de

viat

ions

(Std

) are

incl

uded

in o

rder

to m

ake

our f

indi

ngs c

ompa

rabl

e to

thos

e of

Gal

i (19

99).

Tabl

e 1c

: Stru

ctur

al V

AR

(Con

t.)

Page 159: Macroeconomics of small open economies

AU

Sbe

nchm

ark

VA

R 0

VA

R 1

VA

R 2

VA

R 3

VA

R 4

VA

R 5

VA

R 6

VA

R 7

T15

010

312

812

811

111

511

710

310

3ES

: F

RO

M19

64.4

1976

.219

70.1

1970

.119

74.2

1973

.219

72.4

1976

.219

76.2

TO

2002

.120

01.4

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

2001

.4L

24

88

84

24

4

CC

-0.9

53-0

.900

-0.7

74-0

.763

-0.8

23-0

.929

-0.9

27-0

.898

-0.8

79St

d0.

095

0.10

50.

123

0.12

00.

136

0.13

80.

121

0.11

30.

132

UpB

-0.7

40-0

.614

-0.4

20-0

.412

-0.4

08-0

.577

-0.7

25-0

.576

-0.5

38Lo

B-0

.991

-0.9

31-0

.791

-0.7

85-0

.831

-0.9

33-0

.979

-0.9

13-0

.912

¦ Lo

B - U

pB ¦

0.25

10.

317

0.37

10.

373

0.42

30.

356

0.25

40.

337

0.37

4

CH

Ebe

nchm

ark

VA

R 0

VA

R 1

VA

R 2

VA

R 3

VA

R 4

VA

R 5

VA

R 6

VA

R 7

T80

8484

8484

8484

8484

ES:

FR

OM

1982

.219

81.2

1981

.219

81.2

1981

.219

81.2

1981

.219

81.2

1981

.2

TO20

02.1

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

L8

44

44

44

44

CC

-0.9

51-0

.874

-0.8

79-0

.855

-0.8

46-0

.857

-0.8

03-0

.845

-0.7

89St

d0.

238

0.20

40.

170

0.18

40.

202

0.16

90.

194

0.19

40.

233

UpB

-0.1

94-0

.317

-0.4

20-0

.401

-0.3

30-0

.421

-0.2

90-0

.327

-0.1

97Lo

B-0

.960

-0.9

44-0

.950

-0.9

51-0

.935

-0.9

30-0

.909

-0.9

20-0

.909

¦ Lo

B - U

pB ¦

0.76

60.

627

0.53

00.

550

0.60

50.

509

0.61

90.

593

0.71

2N

otes

: In

the

case

of b

oth

AU

S an

d C

HE,

r(t

) is p

roxi

ed b

y a

shor

t-ter

m in

tere

st ra

te.

The

# of

obs

erva

tions

(T) d

epen

ds o

n th

e ch

osen

# o

f lag

s (L)

and

on

data

ava

ilabi

lity.

ES

stan

ds fo

r eff

ectiv

e sa

mpl

e, C

C fo

r the

cor

rela

tion

betw

een

av

erag

e la

bor p

rodu

ctiv

ity g

row

th a

nd e

mpl

oym

ent g

row

th c

ondi

tiona

l on

a po

sitiv

e te

chno

logy

shoc

k, U

pB fo

r upp

er b

ound

(0.9

5 pe

rcen

tile

poin

t),an

d Lo

B fo

r low

er b

ound

(0.0

5 pe

rcen

tile

poin

t). S

tand

ard

devi

atio

ns (S

td) a

re in

clud

ed in

ord

er to

mak

e ou

r fin

ding

s com

para

ble

to th

ose

of G

ali (

1999

).

Tabl

e 1d

: Stru

ctur

al V

AR

(Con

t.)

Page 160: Macroeconomics of small open economies

ESP

benc

hmar

kV

AR

0V

AR

1V

AR

2V

AR

3V

AR

4V

AR

5V

AR

6V

AR

7T

8672

73-

7373

-72

-ES

: F

RO

M19

80.4

1980

.419

80.4

-19

80.4

1980

.4-

1980

.4-

TO

2002

.119

98.3

1998

.4-

1998

.419

98.4

-19

98.3

-L

22

2-

22

-2

-

CC

-0.9

13-0

.827

-0.8

78-

-0.8

85-0

.883

--0

.823

-St

d0.

082

0.17

40.

184

-0.

144

0.18

9-

0.18

2-

UpB

-0.7

19-0

.442

-0.4

19-0

.487

-0.4

25-0

.429

LoB

-0.9

84-0

.939

-0.9

31-0

.937

-0.9

43-0

.926

¦ Lo

B - U

pB ¦

0.26

50.

497

0.51

2-

0.45

00.

518

-0.

497

-

NZ

Lbe

nchm

ark

VA

R 0

VA

R 1

VA

R 2

VA

R 3

VA

R 4

VA

R 5

VA

R 6

VA

R 7

T75

7474

5776

7657

7657

ES:

FR

OM

1983

.319

83.3

1983

.319

87.4

1983

.119

83.1

1987

.419

83.1

1987

.4

TO20

02.1

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

L4

44

22

22

22

CC

-0.7

30-0

.836

-0.6

29-0

.522

-0.5

97-0

.744

-0.4

03-0

.864

-0.5

46St

d0.

185

0.19

20.

201

0.38

00.

253

0.22

00.

430

0.18

70.

382

UpB

-0.3

66-0

.237

-0.1

600.

311

-0.0

16-0

.214

0.51

5-0

.424

0.39

6Lo

B-0

.886

-0.8

39-0

.818

-0.8

95-0

.856

-0.8

91-0

.862

-0.9

15-0

.857

¦ Lo

B - U

pB ¦

0.52

00.

602

0.65

81.

206

0.84

00.

677

1.37

70.

491

1.25

3N

otes

: In

the

case

of b

oth

ESP

and

NZL

, r(t

) is p

roxi

ed b

y a

shor

t-ter

m in

tere

st ra

te.

The

# of

obs

erva

tions

(T) d

epen

ds o

n th

e ch

osen

# o

f lag

s (L)

and

on

data

ava

ilabi

lity.

ES

stan

ds fo

r eff

ectiv

e sa

mpl

e, C

C fo

r the

cor

rela

tion

betw

een

av

erag

e la

bor p

rodu

ctiv

ity g

row

th a

nd e

mpl

oym

ent g

row

th c

ondi

tiona

l on

a po

sitiv

e te

chno

logy

shoc

k, U

pB fo

r upp

er b

ound

(0.9

5 pe

rcen

tile

poin

t),an

d Lo

B fo

r low

er b

ound

(0.0

5 pe

rcen

tile

poin

t). S

tand

ard

devi

atio

ns (S

td) a

re in

clud

ed in

ord

er to

mak

e ou

r fin

ding

s com

para

ble

to th

ose

of G

ali (

1999

).

Tabl

e 1e

: Stru

ctur

al V

AR

(Con

t.)

Page 161: Macroeconomics of small open economies

FRA

NV

EC

M 0

VEC

M 1

VEC

M 2

VEC

M 3

VEC

M 4

VEC

M 5

VEC

M 6

VEC

M 7

T79

77-

7979

-79

ES:

FR

OM

1979

.219

79.4

-19

79.2

1979

.2-

1979

.2

TO19

98.4

1998

.4-

1998

.419

98.4

-19

98.4

T - N

OV

7946

-58

54-

50L

46

-4

4-

4

Ran

k1

2-

24

-4

CC

0.37

8-0

.033

--0

.138

-0.3

64-

-0.0

30U

pB0.

727

0.47

0-

0.67

90.

245

-0.

442

LoB

-0.2

45-0

.483

--0

.576

0.55

5-

-0.4

23 ¦

LoB

- UpB

¦0.

972

0.95

3-

1.25

50.

310

-0.

865

Ran

k2

CC

0.31

1U

pB0.

699

LoB

-0.4

49¦ L

oB -

UpB

¦1.

148

Not

es: I

n th

e ca

se o

f FR

Ai,

n(t

) is p

roxi

ed b

y H

P fil

tere

d em

ploy

men

t and

r(t

) by

a sh

ort t

erm

inte

rest

rate

(3-m

onth

PIB

OR

).Th

e nu

mbe

r of o

bser

vatio

ns (T

) dep

ends

on

the

chos

en n

umbe

r of l

ags (

L) a

nd o

n da

ta a

vaila

bilit

y. E

S st

ands

for e

ffec

tive

sam

ple,

NO

V fo

r num

ber o

f var

iabl

es, C

C fo

r cor

rela

tion

betw

een

aver

age

labo

r pro

duct

ivity

gro

wth

and

em

ploy

men

t gro

wth

con

ditio

nal o

na

posi

tive

tech

nolo

gy sh

ock,

UpB

for u

pper

bou

nd (0

.95

perc

entil

e po

int),

and

LoB

for l

ower

bou

nd (0

.05

perc

entil

e po

int).

Ran

k 0

is e

quiv

alen

t to

a V

AR

in fi

rst-d

iffer

ence

s.

Tabl

e 2a

: Stru

ctur

al V

ECM

Page 162: Macroeconomics of small open economies

GB

RV

EC

M 0

VEC

M 1

VEC

M 2

VEC

M 3

VEC

M 4

VEC

M 5

VEC

M 6

VEC

M 7

T73

158

158

113

111

111

7373

ES:

FR

OM

1983

.419

62.3

1962

.319

73.4

1974

.219

74.2

1983

.419

83.4

TO20

01.4

2001

.420

01.4

2001

.420

01.4

2001

.420

01.4

2001

.4T

- NO

V52

117

117

8262

6244

44L

48

86

88

44

Ran

k1

10

10

14

4C

C-0

.930

-0.7

45-0

.849

-0.7

67-0

.717

-0.7

13-0

.727

-0.7

98U

pB-0

.560

-0.3

19-0

.560

-0.1

77-0

.201

-0.2

39-0

.267

-0.2

30Lo

B-0

.959

-0.7

80-0

.862

-0.7

95-0

.689

-0.7

09-0

.840

-0.8

92 ¦

LoB

- UpB

¦0.

399

0.46

10.

302

0.61

80.

488

0.47

00.

573

0.66

2

Ran

k2

11

3C

C-0

.792

-0.7

87-0

.611

-0.8

00U

pB-0

.402

-0.3

64-0

.205

-0.2

38Lo

B-0

.905

-0.8

16-0

.640

-0.8

86¦ L

oB -

UpB

¦0.

503

0.45

20.

435

0.64

8

Not

es: I

n th

e ca

se o

f GB

R, r

(t) is

pro

xied

by

a lo

ng te

rm ra

te (1

0-ye

ar g

over

nmen

t bon

ds).

The

num

ber o

f obs

erva

tions

(T) d

epen

ds o

n th

e ch

osen

num

ber o

f lag

s (L)

and

on

data

ava

ilabi

lity.

ES

stan

ds fo

r eff

ectiv

e sa

mpl

e,N

OV

for n

umbe

r of v

aria

bles

, CC

for c

orre

latio

n be

twee

n av

erag

e la

bor p

rodu

ctiv

ity g

row

th a

nd e

mpl

oym

ent g

row

th c

ondi

tiona

l on

a po

sitiv

e te

chno

logy

shoc

k, U

pB fo

r upp

er b

ound

(0.9

5 pe

rcen

tile

poin

t), a

nd L

oB fo

r low

er b

ound

(0.0

5 pe

rcen

tile

poin

t).R

ank

0 is

equ

ival

ent t

o a

VA

R in

firs

t-diff

eren

ces.

Gre

y st

ands

for:

num

ber o

f ran

ks n

ot b

ased

on

test

resu

lts (r

obus

tnes

s che

ck).

Tabl

e 2b

: Stru

ctur

al V

ECM

(Con

t.)

Page 163: Macroeconomics of small open economies

ITA

VE

CM

0V

ECM

1V

ECM

2V

ECM

3V

ECM

4V

ECM

5V

ECM

6V

ECM

7T

7676

-76

76-

76ES

: F

RO

M19

80.1

1980

.1-

1980

.119

80.1

-19

80.1

TO19

98.4

1998

.4-

1998

.419

98.4

-19

98.4

T - N

OV

5555

-55

51-

47L

44

-4

4-

4

Ran

k1

2-

33

-2

CC

-0.9

48-0

.958

--0

.920

-0.9

25-

-0.9

53U

pB-0

.525

-0.6

49-

-0.5

06-0

.522

--0

.704

LoB

-0.9

76-0

.970

--0

.941

-0.9

42-

-0.9

46 ¦

LoB

- UpB

¦0.

451

0.32

1-

0.43

50.

420

-0.

242

Ran

k2

CC

-0.9

52U

pB-0

.634

LoB

-0.9

68¦ L

oB -

UpB

¦0.

334

Not

es: I

n th

e ca

se o

f ITA

, r(t

) is p

roxi

ed b

y a

shor

t ter

m in

tere

st ra

te.

The

num

ber o

f obs

erva

tions

(T) d

epen

ds o

n th

e ch

osen

num

ber o

f lag

s (L)

and

on

data

ava

ilabi

lity.

ES

stan

ds fo

r eff

ectiv

e sa

mpl

e,N

OV

for n

umbe

r of v

aria

bles

, CC

for c

orre

latio

n be

twee

n av

erag

e la

bor p

rodu

ctiv

ity g

row

th a

nd e

mpl

oym

ent g

row

th c

ondi

tiona

l on

a po

sitiv

e te

chno

logy

shoc

k, U

pB fo

r upp

er b

ound

(0.9

5 pe

rcen

tile

poin

t), a

nd L

oB fo

r low

er b

ound

(0.0

5 pe

rcen

tile

poin

t).R

ank

0 is

equ

ival

ent t

o a

VA

R in

firs

t-diff

eren

ces.

Tabl

e 2c

: Stru

ctur

al V

ECM

(Con

t.)

Page 164: Macroeconomics of small open economies

USA

VE

CM

0V

ECM

1V

ECM

2V

ECM

3V

ECM

4V

ECM

5V

ECM

6V

ECM

7T

159

159

159

114

116

114

116

116

ES:

FR

OM

1962

.319

62.3

1962

.319

73.4

1973

.219

73.4

1973

.219

73.2

TO20

02.1

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

2002

.1T

- NO

V11

811

811

883

9177

8787

L8

88

64

64

4

Ran

k1

12

21

32

3C

C-0

.417

0.06

2-0

.053

-0.5

08-0

.844

0.19

6-0

.749

-0.0

32U

pB0.

173

0.35

20.

309

0.42

30.

001

0.53

90.

047

0.49

6Lo

B-0

.664

-0.5

21-0

.519

-1.0

00-0

.849

-0.9

87-0

.745

-0.6

15 ¦

LoB

- UpB

¦0.

837

0.87

30.

828

1.42

30.

850

1.52

60.

792

1.11

1

Ran

k2

CC

-0.6

02U

pB0.

246

LoB

-0.6

42¦ L

oB -

UpB

¦0.

888

Not

es: I

n th

e ca

se o

f the

USA

, r(t)

is p

roxi

ed b

y a

long

term

inte

rest

rate

.Th

e nu

mbe

r of o

bser

vatio

ns (T

) dep

ends

on

the

chos

en n

umbe

r of l

ags (

L) a

nd o

n da

ta a

vaila

bilit

y. E

S st

ands

for e

ffec

tive

sam

ple,

NO

V fo

r num

ber o

f var

iabl

es, C

C fo

r cor

rela

tion

betw

een

aver

age

labo

r pro

duct

ivity

gro

wth

and

em

ploy

men

t gro

wth

con

ditio

nal o

na

posi

tive

tech

nolo

gy sh

ock,

UpB

for u

pper

bou

nd (0

.95

perc

entil

e po

int),

and

LoB

for l

ower

bou

nd (0

.05

perc

entil

e po

int).

Ran

k 0

is e

quiv

alen

t to

a V

AR

in fi

rst-d

iffer

ence

s.G

rey

stan

ds fo

r: nu

mbe

r of r

anks

not

bas

ed o

n te

st re

sults

(rob

ustn

ess c

heck

).Li

ght g

rey

stan

ds fo

r: m

odel

spec

ifica

tion

unst

able

(gro

wth

rate

rath

er th

an le

vel o

f ser

ies s

eem

s to

be sh

ocke

d).

Tabl

e 2d

: Stru

ctur

al V

ECM

(Con

t.)

Page 165: Macroeconomics of small open economies

CA

NV

EC

M 0

VEC

M 1

VEC

M 2

VEC

M 3

VEC

M 4

VEC

M 5

VEC

M 6

VEC

M 7

T80

8080

8082

8282

82ES

: F

RO

M19

82.2

1982

.219

82.2

1982

.219

81.4

1981

.419

81.4

1981

.4TO

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

T - N

OV

5959

5959

6969

6767

L4

44

42

22

2

Ran

k0

01

23

33

3C

C-0

.381

-0.3

17-0

.214

-0.3

24-0

.372

0.04

9-0

.101

0.20

9U

pB0.

185

0.24

10.

266

0.30

60.

521

0.64

80.

592

0.69

2Lo

B-0

.734

-0.7

3-0

.638

-0.5

87-0

.851

-0.7

62-0

.754

-0.5

82 ¦

LoB

- UpB

¦0.

919

0.97

10.

904

0.89

31.

372

1.41

01.

346

1.27

4

Ran

k2

32

34

CC

-0.0

06-0

.266

-0.2

34-0

.580

-0.2

48U

pB0.

463

0.43

60.

544

0.36

30.

447

LoB

-0.5

50-0

.582

-0.6

71-0

.665

-0.6

35¦ L

oB -

UpB

¦1.

013

1.01

81.

215

1.02

81.

082

Not

es: I

n th

e ca

se o

f CA

N, r

(t) i

s pro

xied

by

a sh

ort t

erm

inte

rest

rate

.Th

e nu

mbe

r of o

bser

vatio

ns (T

) dep

ends

on

the

chos

en n

umbe

r of l

ags (

L) a

nd o

n da

ta a

vaila

bilit

y. E

S st

ands

for e

ffec

tive

sam

ple,

NO

V fo

r num

ber o

f var

iabl

es, C

C fo

r cor

rela

tion

betw

een

aver

age

labo

r pro

duct

ivity

gro

wth

and

em

ploy

men

t gro

wth

con

ditio

nal o

na

posi

tive

tech

nolo

gy sh

ock,

UpB

for u

pper

bou

nd (0

.95

perc

entil

e po

int),

and

LoB

for l

ower

bou

nd (0

.05

perc

entil

e po

int).

Ran

k 0

is e

quiv

alen

t to

a V

AR

in fi

rst-d

iffer

ence

s.

Tabl

e 2e

: Stru

ctur

al V

ECM

(Con

t.)

Page 166: Macroeconomics of small open economies

JPN

VE

CM

0V

ECM

1V

ECM

2V

ECM

3V

ECM

4V

ECM

5V

ECM

6V

ECM

7T

8484

6484

8464

8666

ES:

FR

OM

1981

.219

81.2

1986

.219

81.2

1981

.219

86.2

1980

.419

85.4

TO20

02.1

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

2002

.1T

- NO

V63

6343

6359

3971

51L

44

44

44

22

Ran

k0

12

12

44

2C

C-0

.154

0.18

10.

268

0.20

20.

222

0.38

50.

583

0.24

5U

pB0.

230

0.55

60.

591

0.58

60.

626

0.78

70.

857

0.72

8Lo

B-0

.703

-0.6

41-0

.446

-0.6

39-0

.638

-0.5

96-0

.870

-0.7

26 ¦

LoB

- UpB

¦0.

933

1.19

71.

037

1.22

51.

264

1.38

31.

727

1.45

4

Ran

k2

CC

0.65

0U

pB0.

775

LoB

-0.5

90¦ L

oB -

UpB

¦1.

365

Not

es: I

n th

e ca

se o

f JPN

, r(t

) is p

roxi

ed b

y a

shor

t ter

m in

tere

st ra

te.

The

num

ber o

f obs

erva

tions

(T) d

epen

ds o

n th

e ch

osen

num

ber o

f lag

s (L)

and

on

data

ava

ilabi

lity.

ES

stan

ds fo

r eff

ectiv

e sa

mpl

e,N

OV

for n

umbe

r of v

aria

bles

, CC

for c

orre

latio

n be

twee

n av

erag

e la

bor p

rodu

ctiv

ity g

row

th a

nd e

mpl

oym

ent g

row

th c

ondi

tiona

l on

a po

sitiv

e te

chno

logy

shoc

k, U

pB fo

r upp

er b

ound

(0.9

5 pe

rcen

tile

poin

t), a

nd L

oB fo

r low

er b

ound

(0.0

5 pe

rcen

tile

poin

t).R

ank

0 is

equ

ival

ent t

o a

VA

R in

firs

t-diff

eren

ces.

Tabl

e 2f

: Stru

ctur

al V

ECM

(Con

t.)

Page 167: Macroeconomics of small open economies

AU

SV

EC

M 0

VEC

M 1

VEC

M 2

VEC

M 3

VEC

M 4

VEC

M 5

VEC

M 6

VEC

M 7

T10

113

113

111

411

411

410

310

3ES

: F

RO

M19

77.1

1969

.319

69.3

1973

.419

73.4

1973

.419

76.3

1976

.3TO

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

T - N

OV

7010

010

083

7777

7474

L6

66

66

64

4

Ran

k1

11

11

12

2C

C-0

.817

-0.8

78-0

.869

-0.7

81-0

.785

-0.7

85-0

.924

-0.9

03U

pB-0

.454

-0.5

89-0

.615

-0.5

66-0

.518

-0.5

42-0

.598

-0.5

38Lo

B-0

.864

-0.9

04-0

.906

-0.8

45-0

.848

-0.8

47-0

.934

-0.9

13 ¦

LoB

- UpB

¦0.

410

0.31

50.

291

0.27

90.

330

0.30

50.

336

0.37

5

Ran

k2

3C

C-0

.772

-0.6

21U

pB-0

.252

-0.0

59Lo

B-0

.817

-0.8

79¦ L

oB -

UpB

¦0.

565

0.82

0

Not

es: I

n th

e ca

se o

f AU

S, r

(t) i

s pro

xied

by

a sh

ort t

erm

inte

rest

rate

.Th

e nu

mbe

r of o

bser

vatio

ns (T

) dep

ends

on

the

chos

en n

umbe

r of l

ags (

L) a

nd o

n da

ta a

vaila

bilit

y. E

S st

ands

for e

ffec

tive

sam

ple,

NO

V fo

r num

ber o

f var

iabl

es, C

C fo

r cor

rela

tion

betw

een

aver

age

labo

r pro

duct

ivity

gro

wth

and

em

ploy

men

t gro

wth

con

ditio

nal o

na

posi

tive

tech

nolo

gy sh

ock,

UpB

for u

pper

bou

nd (0

.95

perc

entil

e po

int),

and

LoB

for l

ower

bou

nd (0

.05

perc

entil

e po

int).

Ran

k 0

is e

quiv

alen

t to

a V

AR

in fi

rst-d

iffer

ence

s.G

rey

stan

ds fo

r: nu

mbe

r of r

anks

not

bas

ed o

n te

st re

sults

(rob

ustn

ess c

heck

).

Tabl

e 2g

: Stru

ctur

al V

ECM

(Con

t.)

Page 168: Macroeconomics of small open economies

CH

EV

EC

M 0

VEC

M 1

VEC

M 2

VEC

M 3

VEC

M 4

VEC

M 5

VEC

M 6

VEC

M 7

T82

8484

8484

8484

84ES

: F

RO

M19

81.4

1981

.219

81.2

1981

.219

81.2

1981

.219

81.2

1981

.2TO

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

T - N

OV

5163

6363

5959

5555

L6

44

44

44

4

Ran

k1

22

12

23

4C

C-0

.915

-0.9

50-0

.886

-0.7

42-0

.727

-0.8

58-0

.770

-0.9

15U

pB-0

.400

-0.1

60-0

.275

-0.1

45-0

.021

-0.1

90-0

.113

-0.3

53Lo

B-0

.943

-0.9

56-0

.959

-0.8

92-0

.886

-0.9

42-0

.888

-0.9

22 ¦

LoB

- UpB

¦0.

543

0.79

60.

684

0.74

70.

865

0.75

20.

775

0.56

9

Ran

k2

33

4C

C-0

.763

-0.9

02-0

.890

-0.9

19U

pB-0

.366

-0.4

46-0

.314

-0.4

82Lo

B-0

.908

-0.9

43-0

.924

-0.9

32¦ L

oB -

UpB

¦0.

542

0.49

70.

610

0.45

0

Not

es: I

n th

e ca

se o

f CH

E, r

(t) i

s pro

xied

by

a sh

ort t

erm

inte

rest

rate

.Th

e nu

mbe

r of o

bser

vatio

ns (T

) dep

ends

on

the

chos

en n

umbe

r of l

ags (

L) a

nd o

n da

ta a

vaila

bilit

y. E

S st

ands

for e

ffec

tive

sam

ple,

NO

V fo

r num

ber o

f var

iabl

es, C

C fo

r cor

rela

tion

betw

een

aver

age

labo

r pro

duct

ivity

gro

wth

and

em

ploy

men

t gro

wth

con

ditio

nal o

na

posi

tive

tech

nolo

gy sh

ock,

UpB

for u

pper

bou

nd (0

.95

perc

entil

e po

int),

and

LoB

for l

ower

bou

nd (0

.05

perc

entil

e po

int).

Ran

k 0

is e

quiv

alen

t to

a V

AR

in fi

rst-d

iffer

ence

s.

Tabl

e 2h

: Stru

ctur

al V

ECM

(Con

t.)

Page 169: Macroeconomics of small open economies

ESP

VE

CM

0V

ECM

1V

ECM

2V

ECM

3V

ECM

4V

ECM

5V

ECM

6V

ECM

7T

6971

-71

71-

71ES

: F

RO

M19

81.3

1981

.2-

1981

.219

81.2

-19

81.1

TO19

98.3

1998

.4-

1998

.419

98.4

-19

98.3

T - N

OV

4850

-50

46-

56L

44

-4

4-

2

Ran

k1

0-

00

-3

CC

-0.7

84-0

.760

--0

.765

-0.7

62-

-0.7

67U

pB-0

.488

-0.4

69-

-0.4

77-0

.475

--0

.544

LoB

-0.9

01-0

.896

--0

.900

-0.8

98-

-0.9

33 ¦

LoB

- UpB

¦0.

413

0.42

7-

0.42

30.

423

-0.

389

Ran

k2

31

1C

C-0

.736

-0.5

31-0

.923

-0.9

69U

pB-0

.418

0.14

4-0

.512

-0.5

41Lo

B-0

.890

-0.8

27-0

.928

-0.9

39 ¦

LoB

- UpB

¦0.

472

0.97

10.

416

0.39

8

Ran

k3

CC

-0.5

80U

pB0.

140

LoB

-0.8

55¦ L

oB -

UpB

¦0.

995

Not

es: I

n th

e ca

se o

f ESP

, r(t

) is p

roxi

ed b

y a

shor

t ter

m in

tere

st ra

te.

The

num

ber o

f obs

erva

tions

(T) d

epen

ds o

n th

e ch

osen

num

ber o

f lag

s (L)

and

on

data

ava

ilabi

lity.

ES

stan

ds fo

r eff

ectiv

e sa

mpl

e,N

OV

for n

umbe

r of v

aria

bles

, CC

for c

orre

latio

n be

twee

n av

erag

e la

bor p

rodu

ctiv

ity g

row

th a

nd e

mpl

oym

ent g

row

th c

ondi

tiona

l on

a po

sitiv

e te

chno

logy

shoc

k, U

pB fo

r upp

er b

ound

(0.9

5 pe

rcen

tile

poin

t), a

nd L

oB fo

r low

er b

ound

(0.0

5 pe

rcen

tile

poin

t).R

ank

0 is

equ

ival

ent t

o a

VA

R in

firs

t-diff

eren

ces.

Gre

y st

ands

for:

num

ber o

f ran

ks n

ot b

ased

on

test

resu

lts (r

obus

tnes

s che

ck).

Tabl

e 2i

: Stru

ctur

al V

ECM

(Con

t.)

Page 170: Macroeconomics of small open economies

NZ

LV

EC

M 0

VEC

M 1

VEC

M 2

VEC

M 3

VEC

M 4

VEC

M 5

VEC

M 6

VEC

M 7

T75

7558

7375

5875

58ES

: F

RO

M19

83.3

1983

.319

87.4

1984

.119

83.3

1987

.419

83.3

1987

.4TO

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

2002

.120

02.1

T - N

OV

5454

4742

5045

4643

L4

42

64

24

2

Ran

k2

31

22

14

1C

C-0

.861

-0.7

57-0

.783

-0.5

30-0

.689

-0.8

01-0

.408

-0.7

88U

pB-0

.447

-0.1

65-0

.521

-0.1

33-0

.095

-0.4

620.

127

-0.3

32Lo

B-0

.896

-0.8

76-0

.957

-0.6

83-0

.823

-0.9

64-0

.738

-0.9

65 ¦

LoB

- UpB

¦0.

449

0.71

10.

436

0.55

00.

728

0.50

20.

865

0.63

3

Ran

k3

2C

C-0

.410

-0.6

85U

pB0.

018

0.13

6Lo

B-0

.754

-0.9

07¦ L

oB -

UpB

¦0.

772

1.04

3

Not

es: I

n th

e ca

se o

f NZL

, r(t

) is p

roxi

ed b

y a

shor

t ter

m in

tere

st ra

te.

The

num

ber o

f obs

erva

tions

(T) d

epen

ds o

n th

e ch

osen

num

ber o

f lag

s (L)

and

on

data

ava

ilabi

lity.

ES

stan

ds fo

r eff

ectiv

e sa

mpl

e,N

OV

for n

umbe

r of v

aria

bles

, CC

for c

orre

latio

n be

twee

n av

erag

e la

bor p

rodu

ctiv

ity g

row

th a

nd e

mpl

oym

ent g

row

th c

ondi

tiona

l on

a po

sitiv

e te

chno

logy

shoc

k, U

pB fo

r upp

er b

ound

(0.9

5 pe

rcen

tile

poin

t), a

nd L

oB fo

r low

er b

ound

(0.0

5 pe

rcen

tile

poin

t).R

ank

0 is

equ

ival

ent t

o a

VA

R in

firs

t-diff

eren

ces.

Tabl

e 2j

: Stru

ctur

al V

ECM

(Con

t.)

Page 171: Macroeconomics of small open economies

Figure 1a: Benchmark model FRA:

response of n to T-shock

5 10 15 20-2

-1

0

1

GBR:

response of n to T-shock

5 10 15 20-2

-1

0

1

ITA:

response of n to T-shock

5 10 15 20-1.5

-1.0

-0.5

0.0

0.5

USA:

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

Page 172: Macroeconomics of small open economies

Figure 1b: Benchmark model (Cont.) CAN:

response of n to T-shock

5 10 15 20-1

0

1

2

JPN:

response of n to T-shock

5 10 15 20-0.5

0.0

0.5

1.0

1.5

AUS:

response of n to T-shock

5 10 15 20-1.2

-0.8

-0.4

-0.0

0.4

CHE:

response of n to T-shock

5 10 15 20-2

-1

0

1

Page 173: Macroeconomics of small open economies

Figure 1c: Benchmark model (Cont.) ESP:

response of n to T-shock

5 10 15 20-4

-2

0

2

NLZ:

response of n to T-shock

5 10 15 20-2.7

-1.8

-0.9

-0.0

0.9

Page 174: Macroeconomics of small open economies

Figure 2a: VAR 0 FRA:

response of n to T-shock

5 10 15 20-2

-1

0

1

GBR:

response of n to T-shock

5 10 15 20-2

-1

0

1

ITA:

response of n to T-shock

5 10 15 20-1.5

-1.0

-0.5

0.0

0.5

USA:

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

Page 175: Macroeconomics of small open economies

Figure 2b: VAR 0 (Cont.) CAN:

response of n to T-shock

5 10 15 20-1

0

1

2

JPN:

response of n to T-shock

5 10 15 20-0.5

0.0

0.5

1.0

1.5

AUS:

response of n to T-shock

5 10 15 20-1.2

-0.8

-0.4

-0.0

0.4

CHE:

response of n to T-shock

5 10 15 20-2

-1

0

1

Page 176: Macroeconomics of small open economies

Figure 2c: VAR 0 (Cont.) ESP:

response of n to T-shock

5 10 15 20-4

-2

0

2

NLZ:

response of n to T-shock

5 10 15 20-2.7

-1.8

-0.9

-0.0

0.9

Page 177: Macroeconomics of small open economies

Figure 3a: USA, VAR 1-7 VAR 1:

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

VAR 2:

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

VAR 3:

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

VAR 4:

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

Page 178: Macroeconomics of small open economies

Figure 3b: USA, VAR 1-7 (Cont.) VAR 5:

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

VAR 6:

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

VAR 7:

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

Page 179: Macroeconomics of small open economies

Figure 4a: VECM 0 FRA: Rank 1

response of n to T-shock

5 10 15 20-0.50

-0.25

0.00

0.25

0.50

GBR: Rank 1

response of n to T-shock

5 10 15 20-1.6

-0.8

0.0

0.8

ITA: Rank 2

response of n to T-shock

5 10 15 20-1.5

-1.0

-0.5

0.0

0.5

USA: Rank 1

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

Page 180: Macroeconomics of small open economies

Figure 4b: VECM 0 (Cont.) CAN: Rank 2

response of n to T-shock

5 10 15 20-0.8

0.0

0.8

1.6

JPN: Rank 2

response of n to T-shock

5 10 15 20-0.5

0.0

0.5

1.0

1.5

AUS: Rank 1

response of n to T-shock

5 10 15 20-1.2

-0.8

-0.4

-0.0

0.4

CHE: Rank 1

response of n to T-shock

5 10 15 20-2

-1

0

1

Page 181: Macroeconomics of small open economies

Figure 4c: VECM 0 (Cont.) ESP: Rank 1

response of n to T-shock

5 10 15 20-4

-2

0

2

NLZ: Rank 2

response of n to T-shock

5 10 15 20-2.7

-1.8

-0.9

-0.0

0.9

Page 182: Macroeconomics of small open economies

Figure 5a: USA, VECM 1-7 VECM 1: Rank 1

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

VECM 2: Rank 2

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

VECM 3: Rank 2

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

1.0

VECM 4: Rank 1

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

1.0

Page 183: Macroeconomics of small open economies

Figure 5b: USA, VECM 1-7 (Cont.) VECM 5: Rank 3

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

1.0

VECM 6: Rank 2

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

VECM 7: Rank 3

response of n to T-shock

5 10 15 20-1.0

-0.5

0.0

0.5

1.0

Page 184: Macroeconomics of small open economies

Figure 6a: FRA, VECM 1-7 VECM 1: Rank 2

response of n to T-shock

5 10 15 20-0.50

-0.25

0.00

0.25

0.50

VECM 3: Rank 2

response of n to T-shock

5 10 15 20-0.50

-0.25

0.00

0.25

0.50

VECM 4: Rank 4

response of n to T-shock

5 10 15 20-0.50

-0.25

0.00

0.25

0.50

VECM 6: Rank 4

response of n to T-shock

5 10 15 20-0.50

-0.25

0.00

0.25

0.50

Page 185: Macroeconomics of small open economies

Figure 6b: GBR, VECM 1-7 VECM 1: Rank 1

response of n to T-shock

5 10 15 20-1.6

-0.8

0.0

0.8

VECM 3: Rank 1

response of n to T-shock

5 10 15 20-1.6

-0.8

0.0

0.8

VECM 4: Rank 1

response of n to T-shock

5 10 15 20-1.6

-0.8

0.0

0.8

VECM 6: Rank 3

response of n to T-shock

5 10 15 20-2

-1

0

1

Page 186: Macroeconomics of small open economies

Figure 6c: ITA, VECM 1-7 VECM 1: Rank 2

response of n to T-shock

5 10 15 20-1.5

-1.0

-0.5

0.0

0.5

VECM 3: Rank 3

response of n to T-shock

5 10 15 20-1.5

-1.0

-0.5

0.0

0.5

VECM 4: Rank 3

response of n to T-shock

5 10 15 20-1.5

-1.0

-0.5

0.0

0.5

VECM 6: Rank 2

response of n to T-shock

5 10 15 20-1.5

-1.0

-0.5

0.0

0.5

Page 187: Macroeconomics of small open economies

Figure 6d: CAN, VECM 1-7 VECM 2: Rank 1

response of n to T-shock

5 10 15 20-0.8

0.0

0.8

1.6

VECM 2: Rank 2

response of n to T-shock

5 10 15 20-0.8

0.0

0.8

1.6

VECM 6: Rank 3

response of n to T-shock

5 10 15 20-0.8

0.0

0.8

1.6

VECM 7: Rank 3

response of n to T-shock

5 10 15 20-0.8

0.0

0.8

1.6

Page 188: Macroeconomics of small open economies

Figure 6e: JPN, VECM 1-7 VECM 1: Rank 1

response of n to T-shock

5 10 15 20-0.5

0.0

0.5

1.0

1.5

VECM 2: Rank 2

response of n to T-shock

5 10 15 20-0.5

0.0

0.5

1.0

1.5

VECM 4: Rank 2

response of n to T-shock

5 10 15 20-0.5

0.0

0.5

1.0

1.5

VECM 7: Rank 2

response of n to T-shock

5 10 15 20-0.5

0.0

0.5

1.0

1.5

Page 189: Macroeconomics of small open economies

Figure 6f: AUS, VECM 1-7 VECM 1: Rank 1

response of n to T-shock

5 10 15 20-1.2

-0.8

-0.4

-0.0

0.4

VECM 2: Rank 1

response of n to T-shock

5 10 15 20-1.2

-0.8

-0.4

-0.0

0.4

VECM 6: Rank 2

response of n to T-shock

5 10 15 20-1.5

-1.0

-0.5

0.0

0.5

VECM 7: Rank 2

response of n to T-shock

5 10 15 20-1.5

-1.0

-0.5

0.0

0.5

Page 190: Macroeconomics of small open economies

Figure 6g: CHE, VECM 1-7 VECM 1: Rank 2

response of n to T-shock

5 10 15 20-2

-1

0

1

VECM 2: Rank 2

response of n to T-shock

5 10 15 20-2

-1

0

1

VECM 6: Rank 3

response of n to T-shock

5 10 15 20-2

-1

0

1

VECM 7: Rank 4

response of n to T-shock

5 10 15 20-2

-1

0

1

Page 191: Macroeconomics of small open economies

Figure 6h: ESP, VECM 1-7 VECM 1: Rank 3

response of n to T-shock

5 10 15 20-4

-2

0

2

VECM 3: Rank 1

response of n to T-shock

5 10 15 20-4

-2

0

2

VECM 4: Rank 1

response of n to T-shock

5 10 15 20-5.0

-2.5

0.0

2.5

VECM 6: Rank 3

response of n to T-shock

5 10 15 20-4

-2

0

2

Page 192: Macroeconomics of small open economies

Figure 6i: NZL, VECM 1-7 VECM 1: Rank 3

response of n to T-shock

5 10 15 20-2.7

-1.8

-0.9

-0.0

0.9

VECM 2: Rank 1

response of n to T-shock

5 10 15 20-2.7

-1.8

-0.9

-0.0

0.9

VECM 3: Rank 2

response of n to T-shock

5 10 15 20-2.7

-1.8

-0.9

-0.0

0.9

VECM 7: Rank 2

response of n to T-shock

5 10 15 20-2.7

-1.8

-0.9

-0.0

0.9