2020 08 JRP Inflation Expectations
2020 08 JRP Inflation Expectations
Inflation Expectations,
Their Measurement and the
Estimate of Their Degree of
Anchoring
Editors:
Alexander Guarín, Luis Fernando Melo and Eliana González
Inflation Expectations,
Their Measurement
and the Estimate of
Their Degree of
Anchoring
Inflation Expectations,
Their Measurement and
the Estimate of Their
Degree of Anchoring
Eliana González
Chief of Statistics division, Banco de la República
<[email protected]>
vi
EDITORS
Alexander Guarín
Currently, Alexander is senior researcher at Macroeco-
nomics Modeling Department of Banco de la República. He
studied Economics and did a master in Economic Sciences
at Universidad Nacional de Colombia. He also obtained a
master degree in Industrial Engineering from the Univer-
sidad de los Andes (Colombia) and earned a PhD in Com-
putational Finance from the University of Essex (United
Kingdom). His current research interests include the de-
velopment and application of both numerical and econo-
metric methods for the solution of problems in finance and
macroeconomics. In particular, he is interested in meshfree
computational methods and Bayesian econometrics. Some
of his research works have been published in specialized
scientific journals (for example, Journal of Economic, Dy-
namics and Control, European Journal of Operational Re-
search, and Emerging Markets Review). < http://investiga.
banrep.gov.co/es/profile/81>.
vii
Eliana González
Statistician from Universidad Nacional of Colombia with a
master’s degree in Finance and Econometrics from Univer-
sity of London, Queen Mary College. She currently works as
a Chief of Statistics division of Banco de la República. Her
research areas of interests include Econometrics and time se-
ries applied to macroeconomic and finance. She is author of
several working papers published in Borradores de Economía
series of Banco de la República, besides she has published in
some journals as the American Journal of Agricultural Eco-
nomics and Money affairs, among others.
viii
PREFACE
In 2005, CEMLA’s Board of Governors agreed to bols-
ter economic research and collaboration among its
membership through the establishment of research
activities on topics of common interest. After a careful
analysis of the best way to implement such a program,
the heads of economic studies of the central banks on
the Steering Committee of CEMLA’s Central Bank Re-
search Network identified topics of interest and agreed
that papers on these topics should be presented at the
Network’s Annual Meetings and subsequently publis-
hed. The terms of reference for the first joint research
project were established in 2006, and the first Joint Re-
search Program book was published in 2008, entitled
Estimating and Using Unobservable Variables in the Region.
Since then, research topics have been selected
annually by the heads of economic studies at central
banks within the Research Network Steering Commit-
tee, while representatives from the participating cen-
tral banks have acted voluntarily as coordinators for
each of these projects. Additional volumes have been
published on topics such as inflationary dynamics,
persistence, and prices and wages formation; domes-
tic assets prices, global fundamentals, and financial
stability; monetary policy and financial stability in
Latin America and the Caribbean; international spi-
llovers of monetary policy, and financial decisions of
households and financial inclusion in Latin America
and the Caribbean, among others.
All of the aforementioned subjects are of particular
importance for the design and conduct of monetary
policy and the preservation of financial stability. In its
2017 Meeting, the Research Network focused on a topic
of particular interest for central banking, and whose
importance has increased over recent years: that of the
ix
measurement of inflation expectations and their an-
choring to an inflation target. One particular motive
for this renewed interest was the shocks that affected
inflation trends in the global economy in recent years,
such as commodities price fluctuations and those asso-
ciated with climate change phenomena, among others.
As argued in the literature (an overview of it is offe-
red in the Introduction to the present volume), infla-
tion expectations and, in particular, their degree of
anchoring, are fundamental for determining price
evolution and volatility developments. Therefore, an
accurate measurement of inflation expectations and
a better understanding of their determinants are fun-
damental for the design of an effective monetary poli-
cy. Nevertheless, such a measurement is a challenging
task, which has been approached through survey-based
or model-based methods, including their inference
from market prices of financial instruments. Moreo-
ver, there has been a lively debate among authorities
and researchers about the potential links between po-
licy decisions and agents’ expectations, and whether
long-term expectations may be well-anchored.
The papers included in the present volume address
these and other closely related topics (e.g. forecasts of
inflation using novel techniques). They represent an
effort by researchers of the central banks of Argenti-
na, Bolivia, Brazil, Colombia, Costa Rica, Guatemala,
Mexico, Paraguay, Peru, Spain, as well as researchers
from CEMLA and the Bank for International Settle-
ments (BIS), all of them coordinated by the Banco de
la República (Colombia) with support provided by the
Financial Stability Group of the Inter-American De-
velopment Bank.
We at CEMLA would like to thank the collaborators
in this project, and hope that these documents serve
as a showcase of the analysis carried out in the region
and contribute towards the improvement of policy de-
sign related to the core activities of central banking in
Latin America and the Caribbean.
x
Table of Contents
Editors......................................................................................... vii
Preface.......................................................................................... ix
Introduction..................................................................................1
Alexander Guarín
Luis Fernando Melo
Eliana González
1. The Formation and Measurement
of Inflation Expectations......................................................... 9
2. The Degree of Anchoring of Inflation Expectations............ 10
3. Inflation Forecasting and Its Performance Evaluation......... 11
4. Inflation Expectations and Its Relation
with Economic Policy............................................................ 12
References............................................................................. 13
xi
2.1. Inflation-linked Bonds.................................................. 24
2.2. Inflation-linked Swaps................................................... 26
2.3. Inflation-linked Options............................................... 27
2.4. Inflation Expectations from Financial
Instruments in Latin America..................................... 28
3. Modeling Interest Rates from Public Debt Markets............. 29
3.1. The Model...................................................................... 29
4. Results of Inflation Expectations from Public
Debt Markets.......................................................................... 31
4.1 Yield Curve Estimation.................................................. 31
4.2. Empirical Results........................................................... 34
5. Conclusions............................................................................. 45
References............................................................................. 46
xii
Annex B. Expected Inflation, Responses
and Dispersion by Group............................................. 75
References................................................................................... 77
xiii
5. Conclusions.......................................................................... 124
Annex ........................................................................................ 127
References................................................................................. 134
xiv
3. Empirical Analysis ................................................................ 192
3.1 Signals of Long-term Inflation Anchoring.................. 194
3.1.1 Signals from Survey Data................................... 194
3.1.2 Signals from Market Data.................................. 196
3.1.3 Selection of Signals Based on Correlation
Analysis .............................................................. 196
3.2 Factor Analysis ............................................................. 196
3.3 State-space Model ........................................................ 198
3.4 Baseline eais................................................................. 209
3.5 Robustness Analyses..................................................... 210
4. Conclusion............................................................................ 212
References........................................................................... 216
xv
4.1.2 Forecasting Exercise........................................... 235
4.2 Forecast Evaluation....................................................... 236
5. Conclusions........................................................................... 237
Annex A. Data Characteristics........................................... 238
Annex B. Results characteristics........................................ 239
References........................................................................... 242
midas-adl:........................................................................................271
midas-adl-cf:...................................................................................272
Other:.................................................................................. 272
Appendix B: Additional Tables.......................................... 273
References........................................................................... 281
xvi
Evaluation of Inflation Forecasting Models in Guatemala..........285
Juan Carlos Castañeda-Fuentes
Carlos Eduardo Castillo-Maldonado
Héctor Augusto Valle-Samayoa
Douglas Napoleón Galindo-Gonzáles
Juan Carlos Catalán-Herrera
Guisela Hurtarte-Aguilar
Juan Carlos Arriaza-Herrera
Edson Roger Ortiz-Cardona
Mariano José Gutiérrez-Morales
Abstract...................................................................................... 285
1. Introduction.......................................................................... 286
2. Forecasting Evaluation at the Central
Bank of Guatemala.............................................................. 286
2.1 Unconditional-forecasts Models.................................. 287
2.2. Conditional-forecasts Models..................................... 288
3. Data and Forecast Evaluation Methodology........................ 289
3.1 Data............................................................................... 290
3.2. Forecast Evaluation Methodology.............................. 290
4. Results.................................................................................... 293
4.1. Unconditional Forecast Evaluation............................ 293
4.1.1. Skewness and Normality................................... 293
4.1.2. rmse and mpe..................................................... 294
4.1.3. Diebold-Mariano Test....................................... 294
4.1.4. Pesaran-Timmerman Test................................. 297
4.1.5. Giacomini-Rossi Fluctuation Test.................... 298
4.1.6. Weak Efficiency Test......................................... 299
4.1.7. Strong Efficiency Test....................................... 300
4.2 Conditional Forecast Evaluation................................. 301
5. Conclusions........................................................................... 303
Annex......................................................................................... 303
xvii
Annex 1. Tables of the Unconditional
Forecast Evaluation.................................................... 303
Tables of the Conditional Forecast Evaluation................. 305
References........................................................................... 307
xviii
A.2 wes Survey Questionnaire........................................... 357
A.3 Quantitative Forecasting Inflation Expectations....... 365
A.3.1 Equations of the Statistical Analysis
Forecasting Error............................................... 365
Annex B. Self-Organizing Map Validation............................... 380
B.1 Choice of Topology...................................................... 380
B.2 Post-Training Analysis.................................................. 380
B.3 Non-Linear Auto-Regressive Neural
Networks Validation and Other Results.................... 386
B.3.1 Lag Selection..................................................... 386
B.3.2 Post-Training Analysis........................................ 390
B.3.3 MSE Evaluation................................................. 392
B.3.4 Results, Other Countries.................................. 394
B.4 arima............................................................................. 398
References........................................................................... 400
INFLATION EXPECTATIONS
AND ITS RELATION WITH ECONOMIC POLICY
xix
4.1.1 High interest rate response regimes................. 416
4.1.2 High volatility shock regimes............................ 420
4.2 Estimation results......................................................... 420
4.3 Impulse response functions......................................... 425
4.4 Counterfactuals............................................................ 442
5. Conclusions........................................................................... 450
Annex......................................................................................... 452
A. Estimated Parameters.................................................... 452
References........................................................................... 462
xx
5. Concluding Remarks............................................................ 505
6. Appendix............................................................................... 505
6.1 Parameter Estimation................................................... 505
6.2 Adaptive vs. Rational Expectations.............................. 508
6.3 Exchange Rate Regimes............................................... 510
References........................................................................... 511
xxi
Introduction
Alexander Guarín
Luis Fernando Melo
Eliana González
T
he control of inflation and its volatility are fun-
damental issues for any country. Economies with
a high level of inflation or uncertainty on its fu-
ture value can lead, for example, to high costs for eco-
nomic agents, distortions on future investment plans
and welfare implications for society. On the contrary,
economies with low levels of inflation and volatility,
for instance, can enhance their population living
conditions, access to credit sources, and confidence
indicators for international investors (e.g., Madeira
and Zafar, 2015; and Strohsal and Winkelmann, 2015).
Accordingly, keeping inflation under control be-
comes a crucial task for the monetary authority. In this
regard, a strand of the economic literature has estab-
lished an explicit relation between inflation, its long
term expectations and their anchoring to a target lev-
el. In particular, the literature has underlined the re-
lation of this anchoring to the ability of central banks
to control inflation, set up an effective monetary policy
strategy, and improve the transmission mechanisms
(e.g., Haubrich et al., 2012; Autrup and Grothe, 2014;
and Strohsal et al., 2016).
1
In this context, the appropriate measurement of inflation expec-
tations and their degree of anchoring are essential elements for mak-
ing monetary policy decisions by central banks. Nevertheless, these
variables are unobservable and, hence, their monitoring and assess-
ment are not straightforward.
In practice, inflation expectations are measured through surveys
of specific population groups (e.g., financial market agents, firms,
and consumers), or inferred from financial instruments’ market
prices (e.g., break-even inflation rates, inflation-linked bonds, swaps,
and options). However, the analyses of such expectations from these
two sources of information do not necessarily lead to the same conclu-
sions (e.g., Pierdzioch and Rülke, 2013; and Nautz and Strohsal, 2015).
These measures have different features associated with their
empirical counterparts. Survey-based expectations are a direct es-
timate of the probability distribution of inflation rates from differ-
ent economic sectors. Nonetheless, these expectations are usually
only available at low-frequencies (e.g., monthly or quarterly) and for
a limited number of short-term horizons (typically, one or two years)
(e.g., Autrup and Grothe, 2014; and Pierdzioch and Rülke, 2013).
By contrast, financial market-based expectations can be accessi-
ble in real time, at a higher-frequency (e.g., daily), and with multiple
time horizons, including the long-term ones (e.g., five or ten-year).
Nonetheless, these data are indirect measures of inflation expecta-
tions, whose measurement can be contaminated by several factors.
For instance, break-even inflation rate1 is considered a measure
of inflation compensation that, in addition to inflation expectations,
includes the inflation risk and liquidity premiums. The latter is as-
sociated with market conditions and the availability of liquid nom-
inal and inflation-linked bonds (e.g., Antunes, 2015; and Strohsal
and Winkelmann, 2015).
For authorities, another fundamental aspect is the formation pro-
cess of inflation expectations. This process is essential to understand-
ing how monetary policy decisions are transmitted to expectations
(e.g., economic channels and their speed) and, in turn, to inflation.
This enables central banks to design an effective policy strategy (e.g.,
Evans and Honkapohja, 2001; and Maertens and Rodríguez, 2013).
1
These rates are derived from the spread between nominal and inflation-
linked government bond yields.
2
The academic literature has directed its attention to two main
schemes of expectations, namely, adaptive and rational. The former
considers that inflation dynamics are based only on their own past
values, and hence agents form their expectations using the observed
price information (that is, a backward-looking rule). Under the lat-
ter scheme, each time expectations are formed, individuals consider
all available information including, for example, the learning from
previous prediction errors, the probable future actions of the central
bank as well as the agents’ beliefs (that is, a forward-looking rule)
(e.g., Taylor, 1985; Kiley, 2007; and Golden and Monks, 2009). There
have been other expectations formation mechanisms proposed
in literature. For example, Gerberding (2001) has studied a combi-
nation of both adaptive and rational schemes, while Ekeblom (2012)
has proposed some degree of learning in the formation of expecta-
tions. Other examples within this literature are Carlson and Valev
(2002), Heinemann and Ullrich (2006), and Oral et al. (2011).
As mentioned, the anchoring of inflation expectations is funda-
mental for monetary policy. In fact, the literature points out that well-
anchored expectations reduce the inflation risk premium, improve
investment decisions, enhance the valuation of long-term assets, low-
er the volatility on long-term interest rates and make them less sen-
sitive to shocks (e.g., Gürkaynak et al., 2010; Mehrotra and Yetman,
2014; and Berument and Froyen, 2015).
Inflation expectations are well-anchored (that is, central bank’s
credibility is strong) if shocks affecting current inflation and its
short-term expectation do not lead to long-run deviations from
the target level. If they are well-anchored, then long-term expecta-
tions should be insensitive to macroeconomic shocks or other sur-
prises, so that once shocks have dissipated, inflation should return
to its long-run target. On the contrary, if central bank’s credibility
is weak, economic shocks could deviate long-term inflation expecta-
tions away from its inflation target (e.g., Demertzis et al., 2009; Galati
et al., 2011; and Pagenhardt et al., 2015).
Recent literature has carried out the measurement of the degree
of anchoring through several methodologies, which capture theo-
retical aspects from two main lines of research. The first one evalu-
ates if long-term inflation expectations are moving close to a target
level, so the degree of anchoring depends on the deviation of these
expectations regarding a specific inflation target (e.g., Mehrotra
and Yetman, 2014; and Strohsal and Winkelmann, 2015).
3
The second line studies the dependence relation between short-
and long-term inf lation expectations. This literature assesses
if shocks that affect short-term inflation expectations have effects
on those of the long-term, so that the degree of anchoring depends
on how statistically significant is the joint movement between short-
and long-term inflation expectations in response to shocks (e.g.,
Gürkaynak et al., 2010; and Antunes, 2015).
Figure 1 illustrates recent research works on the anchoring of in-
flation expectations. Each of these studies is characterized accord-
ing to both the methodology considered and the source of data used
in its empirical exercises.
A broad segment of this literature has investigated mainly two is-
sues. The first one is the assessment and characterization of differ-
ences in the degree of anchoring between countries with and without
an inflation-targeting regime. For instance, Gürkaynak et al. (2005),
Gürkaynak et al. (2007), Demertzis et al. (2009), Gürkaynak et al.
(2010), and Beechey et al. (2011) examined this matter for the United
States (US) and the euro area, for sample periods between the 1990s
and the end of 2000s. These studies find that a credible inflation-
targeting strategy improves the anchoring of long-term inflation
expectations, reduces their volatility and makes them less sensitive
to inflation shocks.
The second issue is the evolution of the degree of anchoring
over time. For example, the dynamics of anchoring in the pre-
and post-Global Financial Crisis periods in the US between 2004
and 2014 is studied by Galati et al. (2011), Autrup and Grothe (2014),
Nautz and Strohsal (2015), and Strohsal et al. (2016). The first three
works state that inflation expectations have been deanchored since
the Global Financial Crisis, while the latter work points out that
the deanchoring lasted a short period in 2008, after which expecta-
tions were anchored again .
The works by Lemke and Strohsal (2013), Antunes (2015), Pagen-
hardt et al. (2015), and Scharnagl and Stapf (2015) carry out similar
research for the euro area for sample periods between 2000 and 2015.
Lemke and Strohsal (2013) and Scharnagl and Stapf (2015) stated
that although the European Sovereign Debt Crisis increased the vol-
atility of inflation expectations in 2011, these were not deanchored.
On the other hand, Antunes (2015) and Pagenhardt et al. (2015)
found that the same crisis’ events increased the joint movement
4
of short- and long-term inflation expectations, and since then the lat-
ter have been responding to economic shocks.
The variation in the degree of anchoring has also been studied
in other countries for diverse samples between 1996 and 2013.
De Pooter et al. (2014) find that inflation expectations in Brazil, Chile,
and Mexico are anchored, and that these react to us news’ surprises.
Kabundi and Schaling (2013), and Çiçek and Akar (2014) provide
evidence on the unsuccessful anchoring of inflation expectations
in South Africa and Turkey. These are due to low credibility in each
country. Mehrotra and Yetman (2014), and Berument and Froyen
(2015) show that inflation expectations are more firmly anchored
after the adoption of credible inflation-targeting regimes. Other
recent examples are the studies about the degree of anchoring
in Singapore by Ee and Supaat (2008); the US, European Monetary
Union, United Kingdom and Sweden by Strohsal and Winkelmann
(2015), and Colombia by Gamba et al. (2016).
Another topic associated directly with inflation expectations is the
continuous monitoring and forecasting of inflation. This is highly
relevant for central banks and their monetary policy strategies, par-
ticularly in economies with inflation-targeting regimes. Inflation
forecasts are computed using various types of macroeconometric
and time series methodologies. Recently, forecasting models based
on large data sets, high numbers of predictors and direct combina-
tion of different forecast models have attracted the attention of mod-
elers and practitioners. These techniques are useful considering that
central banks have inflation forecasts coming from different models.
Recent examples of these forecasting tools are the Bayesian
model averaging (BMA) (e.g., Wright, 2009), factor-augmented
vector autoregressive (FAVAR) models (e.g., Bernanke et al., 2005)
and schemes for combining forecasts, proposed by Reid (1968),
and Bates and Granger (1969). Hall and Mitchell (2007), and Geweke
and Amisano (2011) consider combinations of forecasting densities
instead of punctual predictions. Tian and Anderson (2014) pro-
posed new schemes for combining forecasts with possible structural
changes, and Kapetanios et al. (2015) extended the previous litera-
ture with weighting schemes.
A fundamental topic in forecasting is the performance evalua-
tion of prediction models and their comparison with respect to a
benchmark or other forecasts. The works by Giacomini and White
(2006), and Giacomini and Rossi (2010) are recent examples of static
5
Figure 1
RECENT LITERATURE ON THE ANCHORING
OF INFLATION EXPECTATIONS
News regressions
+ models Strohsal & Winkelmann (2015)
Time varying
Strohsal et al. (2016)
parameter models
Regression+
Autrup & Grothe (2014)
model
Kalman learning
Davis & Mack (2013)
process
6
Figure 1 (cont.)
RECENT LITERATURE ON THE ANCHORING
OF INFLATION EXPECTATIONS
Linear regression
Kabundi & Schaling (2013)
Tsenova (2011)
7
were affected by a climate phenomenon known as El Niño 2 with di-
rect effects on the food supply and its prices, as well as indirect ef-
fects on core inflation through indexation mechanisms. The impact
of these shocks on current inflation has underlined the relevance
of bringing up old and new questions about the formation and mea-
surement of inflation expectations, the estimation of their degree
of anchoring, as well as the development of more accurate forecasts
of future inflation and their relation with expectations.
This is inconsistent. S ometimes they use, for exa mple,
the empirical identification of inf lation expectation formation
processes (e.g., adaptive, rational, hybrid, or adaptive learning),
their changes over time, the statistical validation of these schemes
and the characterization of their main determinants. Likewise,
these queries relate to the measurement of an informative signal
of expectations, the choice of a suitable source of data and time
horizons as well as the theoretical and empirical implications of such
an election for monetary policy decision making.
Other questions are addressed, for instance, the measurement
of the degree of inflation expectations anchoring over time and un-
der different policy regimes, the implementation of existing meth-
odologies, the design of new methods and the comparison of their
results. A recent challenge is the prediction of variations in the de-
gree of anchoring in response to diverse shocks (e.g., climate related
shocks and commodity price’s shocks). Other discussions arise on the
evaluation of the measures of expectations as forecasts of future infla-
tion, structural changes in these predictions and how to model them.
With the aim of providing empirical and theoretical support
to the economic research and the policy decisions of central banks,
the Center for Latin American Monetary Studies (cemla) in coor-
dination with the Banco de la República (that is, the Central Bank
of Colombia) organized the 2017 Joint Research Annual Program
to study inflation expectations and other relevant topics associat-
ed with them. In the development of this program, the Financial
Stability Group of the Inter-American Development Bank and the
cemla provided academic support to the research groups through
2
This is a season of high temperatures, shortage of rains and droughts.
8
academic feedback given by professors Olivier Coibion, 3 Massimil-
iano Marcellino,4 and Andrea Tambalotti. 5
This joint program was an opportunity to deal with some of the
previous questions, learn about the current research on inflation
expectations in central banks and contribute to the burgeoning eco-
nomic literature on these issues. The results of this research agenda
are compiled in this book, which includes 13 chapters. The first one is
this Introduction. The remaining 12 chapters correspond to works
from 10 central banks (Argentina, Bolivia, Brazil, Colombia, Costa
Rica, Guatemala, Mexico, Paraguay, Peru, and Spain) and two in-
ternational institutions (Bank for International Settlements – BIS,
and cemla). These works address topics on the formation of infla-
tion expectations, their measurement through surveys and finan-
cial market data, the estimation of the degree of anchoring adopting
several methodological approaches, and the forecasts of inflation
using novel techniques. The works were divided into four main sec-
tions, as follows.
3
Associate Professor at University of Texas at Austin.
4
Full Professor of Econometrics at Bocconi University.
5
Assistant Vice President and Function Head, Macroeconomic and
Monetary Studies Function, Research and Statistics Group, Federal
Reserve Bank of New York.
9
Chapter 3 presents the work by Alonso Alfaro and Aarón Mora
from the Banco Central de Costa Rica. The authors use the model
by Mankiw and Reis (2002) to examine information rigidities in in-
flation expectations of agents from several economic sectors between
2006 and 2017. Although previous studies suggest the existence
of these rigidities in the expectations formation process in Costa
Rica, the results of this research do not support these claims. Esti-
mates show that the magnitude of the rigidities captured from data
is not large enough to validate that statement.
The work by Pablo Alonso of the Banco Central del Paraguay is pre-
sented in chapter 4. Alonso estimates a model of determinants of the
formation of inflation expectations in Paraguay since the adoption
of the inflation-targeting scheme in 2011. His results show that ex-
pectations are a function of past inflation and the credibility in the
central bank. Other variables such as the foreign exchange rate de-
preciation and the changes in oil prices do not seem to play a key role
in their determination.
6
Bolivia is under a monetary-targeting scheme, such that the main refer-
ence for future inflation are the central bank’s projections.
10
expectations were strongly anchored since 2014 due to a greater
credibility of the bcb linked with a larger intervention in the money
market, a more active communication strategy and a stable macro-
economic environment.
Chapter 7 presents the research by Fernando Nascimento
de Oliveira and Wagner Gaglianone of Banco Central do Brasil
(bcbr). They build several time-varying expectation anchoring index-
es of the bcbr from 2002 to 2017, which are based on the monetary au-
thority’s capability to anchor long-term inflation expectations. Those
indexes consider variables of fiscal and monetary policy in their es-
timation. Authors state that estimated indexes are consistent with
the central bank’s credibility perceived by economic agents in Bra-
zil over the sample period.
3. INFLATION FORECASTING
AND ITS PERFORMANCE EVALUATION
Chapters 8 and 9 present the research works developed by Lorena
Garegnani and Maximiliano Gómez, and Luis Libonatti of the
Banco Central de la República Argentina, respectively. Garegnani
and Gómez estimate Bayesian var models with Argentinian data
from 2004 to 2017, and forecast the headline inflation for several
time horizons under a rolling window scheme. In the same line
of research, Libonatti uses a mixed data sampling regression mod-
el to forecast the monthly core inflation of Argentina between 2015
and 2017 using a daily online price index captured by web scraping.
In both works, authors compare their results to forecasts from tradi-
tional benchmark models and show, in general, a good performance
of their predictions.
In chapter 10, the Economic Research Department of the Banco
de Guatemala (Banguat) presents its work. This research assesses
the performance of both unconditional and conditional inflation
forecasts for several time horizons between 2011 and 2017. These
predictions are built using time series tools and structural macro-
economic models used by the Banguat. In line with the traditional
literature, their results show that forecasts computed with time series
tools provide more accuracy in the shortest terms while structural
macroeconometric models provide better predictions for medium-
and long-term horizons.
11
In chapter 11, Héctor Zárate and Daniel Zapata from Banco de la
República (Colombia) use artificial neural networks to forecast in-
flation expectations in a set of 16 countries with inflation-targeting
regimes and a sample period between 1991 and 2016. Their predic-
tions consider different expectations patterns depending on percep-
tions about the oil shock in 2014. Authors show that their exercises
provide more accurate forecasts than the benchmark model and,
anticipate turning points of inflation in most of the cases.
12
References
13
De Pooter, Michiel, Patrice Robitaille, Ian Walker, and Michael
Zdinak (2014), Are Long-term Inflation Expectations Well-anchored
in Brazil, Chile, and Mexico?, International Finance Discussion
Papers, Board of Governors of the Federal Reserve System,
No. 1098, March.
Demertzis, Maria, Massimiliano Marcellino, and Nicola Viegi
(2009), Anchors for Inflation Expectations, dnb Working Papers,
De Nederlandsche Bank, No. 229.
Ee, Khor Hoe, and Saktiandi Supaat (2008), “The Anchoring of
Inflation Expectations in Singapore,” bis Papers, No. 35, pp.
443-449.
Ekeblom, D. (2012), Empirical Swedish Inflation Expectations. Tech-
nical report, Seminar in Macroeconomics, Department of
Economics, Lund University.
Evans, George W., and Seppo Honkapohja (2001), Learning and
Expectations in Macroeconomics, Princeton University Press.
Galati, Gabriele, Steven Poelhekke, and Chen Zhou (2011), “Did
the Crisis Affect Inflation Expectations?,” International Journal
of Central Banking, Vol. 7, No. 1, pp. 1-27.
Gamba, Santiago, Eliana Rocío González, and Luis Fernando Melo
(2016), ¿Están ancladas las expectativas de inflación en
Colombia?, Borradores de Economía, Banco de la República,
No. 940.
Gerberding, Christina (2001), The Information Content of Survey Data
on Expected Price Developments for Monetary Policy, Discussion
Paper, Economic Research Centre of the Deutsche Bundes-
bank, No. 9/01.
Geweke, John, and Gianni Amisano (2011), “Optimal Prediction
Pools,” Journal of Econometrics, Vol. 164, issue 1, pp. 130-141,
<https://doi.org/10.1016/j.jeconom.2011.02.017>.
Giacomini, Raffaella, and Barbara Rossi (2010), “Forecast Com-
parisons in Unstable Environments,” Journal of Applied Econo-
metrics, Vol. 25, No. 4, pp. 595-620.
Giacomini, Raffaella, and Halbert White (2006), “Tests of Condi-
tional Predictive Ability,” Econometrica, Vol. 74, No. 6, pp. 1545-
1578, <https://doi.org/10.1111/j.1468-0262.2006.00718.
x>.
14
Gimeno, Ricardo, and José Manuel Marqués (2009), Extraction of
Financial Market Expectations about Inflation and Interest Rates
from a Liquid Market, Documentos de Trabajo, Banco de
España, No. 906.
Golden, Brian, and Allen Monks (2009), “Measuring Inflation
Expectations in the Euro Area,” Quarterly Bulletin, No. 1,
pp. 67-84.
Gürkaynak, Refet S., Andrew T. Levin, and Eric Swanson (2010),
“Does Inflation Targeting Anchor Long-Run Inflation Expec-
tations? Evidence from the us, uk, and Sweden,” Journal of the
European Economic Association, Vol. 8, issue 6, pp. 1208-1242,
<https://doi.org/10.1111/j.1542-4774.2010.tb00553.x>
Gürkaynak, Refet S., Andrew T. Levin, Andrew N. Marder, and Eric
Swanson (2007), “Inflation Targeting and the Anchoring of
Inflation Expectations in the Western Hemisphere,” frbsf
Economic Review, Vol. 11, 25-47.
Gürkaynak, Refet S., Brian Sack, and Eric Swanson (2005), “The
Sensitivity of Long-Term Interest Rates to Economic News:
Evidence and Implications for Macroeconomic Models,”
The American Economic Review, Vol. 95, issue 1, pp. 425-436,
<doi: 10.1257/0002828053828446>.
Hall, Stephen G., and James Mitchell (2007), “Combining Den-
sity Forecasts,” International Journal of Forecasting, Vol. 23,
issue 1, pp. 1-13, <https://doi.org/10.1016/j.ijfore-
cast.2006.08.001>.
Haubrich, Joseph, George Pennacchi, and Peter Ritchken (2012),
“Inflation Expectations, Real Rates, and Risk Premia: Evi-
dence from Inflation Swaps,” The Review of Financial Studies,
Vol. 25, issue 5, pp. 1588-1629, <https://doi.org/10.1093/
rfs/hhs003>.
Heinemann, Friedrich, and Katrin Ullrich (2006), “The Impact
of emu on Inflation Expectations,” Open Economics Review,
Vol. 17, issue 2, pp. 175-195.
Kabundi, Alain, and Eric Schaling (2013), “Inflation and Inflation
Expectations in South Africa: An Attempt at Explanation,”
South African Journal of Economics, Vol. 81, issue 3, pp. 346-
355, <https://doi.org/10.1111/saje.12007>.
15
Kapetanios, George, James Mitchell, Simon Mc Kay Price, and
Nicholas W. P. Fawcett, N. (2015), “Generalised Density
Forecast Combinations,” Journal of Econometrics, Vol. 188,
issue 1, pp. 150-165, <https://doi.org/10.1016/j.jeco-
nom.2015.02.047>.
Kiley, Michael T. (2007), Monetary Policy Actions and Long-run Inflation
Expectations, feds Working Paper 2008-03, Federal Reserve.
Lemke, Wolfgang, and Till Strohsal (2013), “What Can Break-even
Inflation Rates Tell Us About the Anchoring of Inflation
Expectations in the Euro Area?,” in Annual Conference 2013
(Duesseldorf): Competition Policy and Regulation in a Global Eco-
nomic Order, No. 79794.
Madeira, Carlos, and Basit Zafar (2015), “Heterogeneous Infla-
tion Expectations and Learning,” Journal of Money, Credit
and Banking, Vol. 47, issue 5, pp. 868-996, <https://doi.
org/10.1111/jmcb.12230>.
Maertens, Luís Ricardo and Gabriel Rodríguez (2013), “Inflation
Expectations in the Presence of Policy Shifts and Structural
Breaks: An Experimental Analysis,” The Journal of Socio-Eco-
nomics, Vol. 44, June, pp. 59-67, <https://doi.org/10.1016/j.
socec.2013.02.001>.
Mankiw, N. Gregory and Ricardo Reis (2002), “Sticky Information
Versus Sticky Prices: A Proposal to Replace the New Keynes-
ian Philips Curve,” The Quarterly Journal of Economics, Vol.
117, issue 4, pp. 1295-1328, <https://doi.org/10.1162/0
03355302320935034>.
Mehrotra, Aaron, and James Yetman (2014), Decaying Expectations:
What Inflation Forecasts Tell Us about the Anchoring of Inflation
Expectations, bis Working Papers, No. 464.
Nautz, Dieter, and Till Strohsal (2015), “Are us inflation Expec-
tations Re-anchored?,” Economics Letters, Vol. 127, pp. 6-9,
<https://doi.org/10.1016/j.econlet.2014.12.023>.
Oral, Ece, Hülya Saygili, Mesut Saygili, and S. Özge Tuncel. (2011),
“Inflation Expectations in Turkey: Evidence from Panel Data,”
Journal of Business Cycle Measurement and Analysis, oecd, Vol.
2011, issue 1, pp. 5-28, <https://doi.org/10.1787/jbcma-
2011-5kgg5k53np7c>.
16
Pagenhardt, Laura, Dieter Nautz, and Till Strohsal (2015), The (De-)
Anchoring of Inflation Expectations: New Evidence from the Euro
Area, sfb 649 Discussion Paper Series, No. 2015-044, pp. 1-23.
Pierdzioch, Christian, and Jan-Christoph Rülke (2013), “Do In-
flation Targets Anchor Inflation Expectations?,” Economic
Modelling, Vol. 35, September, pp. 214-223, <https://doi.
org/10.1016/j.econmod.2013.06.042>.
Potter, Simon M., and Joshua Rosenberg (2007), Are us Inflation
Expectations Anchored, Contained or Unmoored?, Working paper,
Federal Reserve Bank of New York.
Reid, David J. (1968), “Combining Three Estimates of Gross Do-
mestic Product,” Economica, Vol. 35, No. 14, pp. 431-444,
<doi: 10.2307/2552350>
Rossi, Barbara, and Tatevik Sekhposyan (2010), “Have Economic
Models’ Forecasting Performance for us Output Growth
and Inflation Changed over Time, and When?,” International
Journal of Forecasting, Vol. 26, issue 4, pp. 808-835, <https://
doi.org/10.1016/j.ijforecast.2009.08.004>.
Sargent, Thomas, Noah Williams, and Tao Zha (2009), “The Con-
quest of South American Inflation,” Journal of Political Econ-
omy, Vol. 117, No. 2, pp. 211-256.
Scharnagl, Michael, and Jelena Stapf, J. (2015), “Inflation, De-
flation, and Uncertainty: What Drives Euro-Area Option-
Implied Inflation Expectations, and Are they Still Anchored
in the Sovereign Debt Crisis?,” Economic Modelling, Vol. 48,
August, pp. 248-269, <https://doi.org/10.1016/j.econ-
mod.2014.11.025>.
Strohsal, Till, and Lars Winkelmann (2015), “Assessing the An-
choring of Inflation Expectations,” Journal of International
Money and Finance, Vol. 50, February, pp. 33-48, <https://
doi.org/10.1016/j.jimonfin.2014.09.001>.
Strohsal, Till, Rafi Melnick, and Dieter Nautz (2016), “The Time-
Varying Degree of Inflation Expectations Anchoring,” Journal
of Macroeconomics, Vol. 48, June, pp. 62-71, <https://doi.
org/10.1016/j.jmacro.2016.02.002>.
17
Taylor, John B. (1985), “Rational Expectations Models in Macro-
economics,” in Kenneth J. Arrow and Seppo Houkupohju
(eds.), Frontiers of Economics, Basil Blackwell Publishers, pp.
391-425.
Tian, Jing, and Heather M. Anderson (2014), “Forecast Combi-
nations Under Structural Break Uncertainty,” International
Journal of Forecasting, Vol. 30, issue 1, January-March, pp. 161-
175, <https://doi.org/10.1016/j.ijforecast.2013.06.003>.
Wright, Jonathan H. (2009), “Forecasting us Inflation by Bayesian
Model Averaging,” Journal of Forecasting, Vol. 28, No. 2, pp.
131-144, <https://doi.org/10.1002/for.1088>.
18
Formation and Measurement
of Inflation Expectations
Extraction of Inflation Expectations
from Financial Instruments
Abstract
In this paper, we estimate inflation expectations for several Latin American
countries using an affine model that takes as factors the observed inflation
and the parameters generated from zero-coupon yield curves of nominal bonds.
By implementing this approach, we avoid the use of inflation-linked securi-
ties, which are scarce in many of these markets, and obtain market measures
of inflation expectations free of any risk premium, eliminating potential bi-
ases included in other measures such as breakeven rates. Our method provides
several advantages, as we can compute inflation expectations at any hori-
zon and forward rates such as the expected inflation over the five-year period
that begins five years from today. We find that inflation expectations in the
long-run are fairly anchored in Chile and Mexico, while those in Brazil and
Colombia are more volatile and less anchored. We also find that expected in-
flation increases at longer horizons in Brazil and Chile, while it is decreas-
ing in Colombia and Mexico.
Keywords: inflation expectations, affine model, real interest rate, risk
premium.
jel classification: G12, E43, E44, C54.
21
1. INTRODUCTION
A
gents’ inf lation expectations are decisive when studying
changes in many of the variables shaping households’ and
firms’ decision making. One approach to obtain inflation
expectations is based on the consensus view of specialist economic
forecasters, such as the surveys of professional forecasters by the
European Central Bank and the Federal Reserve Bank of Philadel-
phia, both of which are released quarterly. Other surveys also exist,
such as the monthly University of Michigan Survey of Consumers in
the United States, which elicits information from consumers rather
than professional economic forecasters. In Latin America, several
central banks also publish surveys about inflation expectations.1 A
drawback of these surveys is that they are released relatively infre-
quently and, thus, the information received has a time lag. Moreover,
they only cover a small range of time horizons and, as identified in
the literature (Ang et al., 2007; Chan et al., 2013), there is some bias
and inertia in their responses.
An alternative way of obtaining agents’ inflation expectations is
to use prices of market-traded financial instruments employed to
hedge against inflation such us inflation-linked bonds, inflation
swaps, and inflation options. One may argue that, given that inves-
tors risk their funds when taking investment decisions based on
expected future inflation and professional forecasters do not have
any vested interest, they could provide a better forecast since they
have more skin in the game. Another advantage to this approach is
that it is possible to derive the whole probability function (Gimeno
and Ibañez, 2017). This makes it possible to estimate, for example,
the probability of the occurrence of certain extreme events or the
uncertainty of future inflation. Another additional advantage in
comparison with surveys is that changes in expectations can be ob-
served almost in real time. This makes it easier to identify the effect
of specific events or decisions on inflation expectations. Unfortu-
nately, there are not many markets of inflation-linked securities avail-
able for most countries. For example, in Latin American only a few
have inflation-linked bonds, and there are no markets for inflation
1
For example, the central banks of Chile, Colombia and Mexico pub-
lish a monthly survey about inflation expectations; the Bank of Brazil
publishes a daily survey.
2
Only conventional government bonds were purchased in the Federal
Reserve Board’s first quantitative easing program. During the Federal
Reserve Board’s second quantitative easing program ( qe II), a total
of usd 600 billion-worth of government securities was purchased, of
which 26 billion was in the form of inflation-linked bonds. The fact
that more conventional bonds are being bought than inflation-linked
bonds could push down their relative yield, and therefore depress the
inflation expectations indicator in a way that is due to a mismatch in
the supply and demand for bonds used to calculate the indicator rather
than to agents’ forecasts of future consumer price trends.
3
See, for example, D’Amico et al. (2014) and Chernov and Mueller
(2012).
4
Unlike ils s, where the compensation for inflation is directly observ-
able from the price, the bond-based indicator requires a comparison
of the yields on inflation-linked bonds and conventional bonds. The
differences in the features of both types of bonds, beyond the fact that
in the case of inflation-linked bonds payments are linked to inflation
(such as, for example, their expiry), may distort the inflation expecta-
Table 1
INFLATION LINK ED SECURITIES
−1
yt ,t +k =
k
( )
Ak + BkʹX t + ut ,t +k ut � N (0,σ 2 1),
λt = λ0 + λ1X t ,
1 − e −k /t 1 − e −k /τ
yt ,t +k = Lt + St + Ct − e −k /τ + ut ,t +k ,
k /τ k / τ
Lt
S
Xt =
t
C t
π t
yt ,t +k = Ert ,t +k + Et [π t ,t +k ]+ γ t ,t +k .
Figure 1
YIELD CURVE ESTIMATES
NOMINAL (BLACK) VS. INFLATION LINKED BONDS (GRAY)
(, 2007) (, 2016)
4 8
3 7
2 6
5
1
4
0
3
−1 2
−2 1
−3 0
0 5 10 15 20 25 0 10 20 30 40
5
The break-even rate includes the spread between the liquidity premium
of the nominal and the inflation-linked bond markets. Because of that,
it decreases if the liquidity premium in the inflation-linked bond market
rises more than the premium of the nominal bond market.
Percent
6
1
0
Jun 2012
Oct 2012
Feb 2013
Jun 2013
Oct 2013
Feb 2014
Jun 2014
Oct 2014
Feb 2015
Jun 2015
Oct 2015
Feb 2016
Jun 2016
Oct 2016
Feb 2017
Jun 2017
Oct 2017
Brake-even rate Current inflation
Table 2
NOMINAL BONDS AVAILABILITY
Original bond
Number of bonds Period maturity
Et [X t +h ] = (1 + Φ + Φ 2 + + Φh −1 )µ + Φh X t .
Figure 3
35
As we previously mentioned, the model we propose allows us to com-
pute inflation expectations at different horizons. Figure 4 shows infla-
tion expectations for the one-year, five-year and ten-year horizons, as
well as the inflation targeting level established by the central bank in
each country. We can see again the different degree of anchoring by
comparing the evolution of expectations for the one-year horizon with
those for the five-years and ten-year horizons. Inflation expectations
in Brazil and Colombia show a similar pattern for all horizons while
expectations in Chile and Mexico are more volatile over the one year
horizon, showing little changes over longer horizons.
Regarding the inflation targeting levels established by the central
banks, most countries currently show inflation expectations at long
horizons within the window limits, 6 although Brazil and Colombia
have experienced recent periods where inflation expectations were
well above these limits. Both countries showed inflation expectations
above 6% before the large decreased experienced since the beginning
of 2016. On the other hand, Mexico shows long-term inflation expec-
tations slightly above the upper band of 4%, mainly due to the recent
increase in expectations after the last us presidential elections. This
effect is more apparent for the evolution of the one-year horizon, fad-
ing out at longer terms. Interestingly, it seems that the results of these
elections have barely affected inflation expectations in other countries.
For Brazil, the deep recession of 2015-2016 has affected expectations,
with a large decrease experienced since the beginning of 2016. The
path of inflation expectations changed again for Brazil at the end of
2016, with expectations turning higher at longer horizons, which sig-
nals a possible recovery. In the case of Colombia, the monetary policy
implemented by the central bank during 2016, with increases in the
policy rate from 4.5% in September 2015 to 7.75% in August 2016, have
contained inflation expectations, being now closer to the inflation tar-
get. Longer-term inflation expectations continue to show lower levels
than short-term ones for this country. Finally, Chile has experienced a
decreasing trend in short-term expectations since mid-2014 which has
been associated, first to the fall in oil prices, and since 2016 to the ap-
preciation of the Chilean peso. Although short-term inflation expecta-
tions remain below the inflation target, expected inflation at long-term
horizons is higher and have experienced little change.
6
The Bank of Brazil sets the inflation target at 4.5% with a window limit of
±1.5%. The central banks of Chile, Colombia and Mexico set the inflation
target at 3% with a window limit of ±1 percent.
1 year
Jul 2013 Jul 2013 Jul 2013 Jul 2013
5 year
Jul 2014 Jul 2014 Jul 2014 Jul 2014
Figure 4
10 year
Jan 2016 Jan 2016 Jan 2016 Jan 2016
Inflation target
9
8
7
6
5
4
3
2
0 1 2 3 4 5 6 7 8 9 10
Years
9
8
7
6
5
4
3
2
0 1 2 3 4 5 6 7 8 9 10
Years
9
8
7
6
5
4
3
2
0 1 2 3 4 5 6 7 8 9 10
Years
9
8
7
6
5
4
3
2
0 1 2 3 4 5 6 7 8 9 10
Years
7
Gutiérrez (2003) provides the values of the central bank independence
indexes for the four countries in our study. Although we should be
careful as the indexes were calculated long time ago, Mexico and Chile
show the largest values of central bank independence.
1
2
3
4
5
6
7
8
2
3
4
5
6
7
8
9
Jul 2012 Jul 2012 Jul 2012 Jul 2012
2Y2Y
Jan 2014 Jan 2014 Jan 2014 Jan 2014
5Y5Y
Jan 2015 Jan 2015 Jan 2015 Jan 2015
Figure 6
Inflation target
INFLATION EXPECTATIONS OF FORWARD RATES
8
In the case of Chile, it is 11-months horizon inflation expectations
(annual change).
Frequency Horizons
Brazil Daily Next 12 months; current year (t) and t+1, t+2,
t+3, t+4.
Mexico Monthly Next 12 months; next 1-4 years; next 5-8 years.
11
10
9
8
7
6
5
4
3
2
1
Jul 2012
Jan 2013
Jul 2013
Jan 2014
Jul 2014
Jan 2015
Jul 2015
Jan 2016
Jul 2016
Jan 2017
Jul 2017
11
10
9
8
7
6
5
4
3
2
1
Jul 2012
Jan 2013
Jul 2013
Jan 2014
Jul 2014
Jan 2015
Jul 2015
Jan 2016
Jul 2016
Jan 2017
Jul 2017
11
10
9
8
7
6
5
4
3
2
1
Jul 2012
Jan 2013
Jul 2013
Jan 2014
Jul 2014
Jan 2015
Jul 2015
Jan 2016
Jul 2016
Jan 2017
Jul 2017
11
10
9
8
7
6
5
4
3
2
1
Jul 2012
Jan 2013
Jul 2013
Jan 2014
Jul 2014
Jan 2015
Jul 2015
Jan 2016
Jul 2016
Jan 2017
Jul 2017
1
Ratio of mean square error of expected inflation from surveys, an ar(1) process
and our model with respect to a naïve prediction of expected inflation equal
to current inflation. Expected inflation in 12 months for Brazil, Colombia and
Mexico; 11 months for Chile.
5. CONCLUSIONS
References
Ang A., G. Bekaert, and M. Wei (2007), “Do Macro Variables, As-
set Markets, or Surveys Forecast Inflation Better?,” Journal of
Monetary Economics, Vol. 54, No. 4, pp. 1163-1212.
Ang A., G. Bekaert, and M. Wei (2008), “The Term Structure of
Real Rates and Expected Inflation,“ Journal of Finance, Vol.
63, pp. 797-849.
Chan, J., G. Koop, and S. Potter (2013), “A New Model of Trend
Inflation,” Journal of Business and Economic Statistics, Vol. 31,
No. 1, pp. 94-106.
Chernov, M., and P. Mueller (2012), “The Term Structure of Infla-
tion Expectations,” Journal of Financial Economics, Vol. 106,
pp. 367-394.
Carriero, A., C. A. Favero, and I. Kaminska (2006), “Financial
Factors, Macroeconomic Information and the Expectations
Theory of the Term Structure of Interest Rates,” Journal of
Econometrics, Vol. 131, pp. 339-358.
Abstract
Costa Rican inflation expectations cannot be characterized as rational un-
der any existing definition of the term. They cannot be categorized as adap-
tive either, since in addition to historical data on inflation, other macroeco-
nomic variables are important in explaining inflation expectations. Instead,
the sticky information model is considered a more sophisticated framework
to assess inflation expectations of Costa Rican agents. Results are based
on the Monthly Survey of Inflation and Exchange Rate Expectations elabo-
rated and published by the Banco Central de Costa Rica. This chapter col-
lects evidence to assess whether the expectations from this survey are subject
to information rigidities. Additionally, this chapter shows how a simulated
survey, based on a sticky information model, is capable of replicating features
from the observed survey.
Keywords: inflation expectations, sticky information, adaptive learning.
jel classification: C53, D84, E31, E58.
49
1. INTRODUCTION
C
onventional economic theory highlights the crucial influence
of expectations on changes in macroeconomic variables. Chang-
es in a variable affect the expectations related to its future move-
ment and these expectations also influence the variable’s underlying
path. This bilateral relation puts the problem of how agents form their
expectations into the front line of macroeconomic modeling.
Most central banks acknowledge the crucial role of expectations,
and argue that managing inflation expectations is paramount for at-
taining price stability and conducting monetary policy. The Banco
Central de Costa Rica (bccr) operates under an inflation targeting
regime, in order to accomplish its goal of a low and stable inflation lev-
el. It relies heavily on the inflation expectations of Costa Rican agents
aligning closely with monetary policy. It is necessary to understand
how inflation expectations are formed to anchor expectations to the
ones targeted by the bccr.
Until recently the research agenda on expectation formation
was eclipsed by the rational expectations ( re ) hypothesis started
by Muth (1961). This hypothesis revolutionized macroeconomic think-
ing during the seventies by incorporating the effect of expectations
into most economic models. As Thomas Sargent points out 1, the re
hypothesis allowed for the disappearance of any free parameters as-
sociated with expectations, so people’s beliefs became outputs of the
model in question. As a result, macroeconomists widely adopted the as-
sumption of re to arrive at tractable equilibrium solutions.
Nevertheless, a common critique for the re hypothesis is that it as-
sumes that people have much more information about the economy
than they really do, since it implies that agents construct expectations
and make decisions by gathering and conveying all available public in-
formation. This assumption is unrealistic and empirical studies often
reject the re hypothesis. There are three popular alternatives to the
re hypothesis: 1) agents use heterogeneous mechanisms to form their
expectations, as in Branch (2004) and Honkapohja and Mitra (2006); 2)
agents use different information sets, as in Angeletos and Lian (2016);
and 3) agents have different abilities to process information, see for
example Woodford (2001). A good survey of alternative approaches
to the specification of expectations is presented in Woodford (2013)
1
See Evans and Honkapohja (2005).
50 A. Alfaro, A. Mora
where the author presents how macroeconomic analysis under a new
Keynesian framework could be performed without relying on the
re hypothesis. Regardless, there are well developed theoretical alter-
natives to re, though many features observed in expectations survey
are not entirely taken into account by these alternatives. Authors like
Manski (2004) have pushed for more empirical studies that deepen
our knowledge of how people elicit and revise their expectations.
One approach to analyzing expectations formation has focused
on the role of information rigidities and has been supported by em-
pirical evidence, see Mankiw and Reis (2002), Woodford (2001),
and Sims (2003). In particular, Mankiw et al. (2003) depart from tra-
ditional empirical approaches to expectations measurement, which
have traditionally relied on measures of central tendency, such as the
mean or median; instead, they study the heterogeneity of inflation
expectations using statistics of dispersion. The idea is that the dis-
agreement among agents over inflation expectations can be explained
by information stickiness. They use the sticky information model
developed in Mankiw and Reis (2002) to explain the mean and dis-
persion of the United States’ inflation expectations. Under this frame-
work, just a fraction of the agents updates their expectations with
the most recent information available. This fraction is derived from
the bounded rationality associated with the cost of updating expec-
tations. Pfajfar and Santoro (2010) build on this line of work and in-
stead of using measures of central tendency, they perform percentile
analysis to study the heterogeneity, learning, and information sticki-
ness of inflation expectations.
Alfaro and Monge (2013) also document that Costa Rican infla-
tion expectations can neither be characterized as rational nor adap-
tive. If expectations were rational, the realized bias between expected
and realized inflation level could not be predicted: Costa Rican data
fails this test even with relaxed assumptions of rationality. On the
other hand, inflation expectations cannot be categorized as adap-
tive neither, since in addition to historical data on inflation, other
macroeconomic variables hold significant explanatory power for in-
flation expectations.
Alfaro and Monge (2013) note the need to evaluate more sophis-
ticated tools to model Costa Rican inflation expectations. This chap-
ter will evaluate the sticky information model to determine whether
this need is substantial. The main source of data for this research
comes from the Monthly Survey of Inflation and Exchange Rate
The bccr has conducted the Monthly Survey of Inflation and Ex-
change Rate Expectations since 2006. This survey gathers data on ex-
pected inflation for the next 12 months and the expected percentage
variation in the exchange rate between the Costa Rican colon (crc)
and the United States dollar (usd) for the next 3, 6, 12, 24, and 36
months2 . The questionnaire of the survey can be found in Annex A.
Responses to questions on inflation and exchange rate expectations
are point expectations that ask for a numerical expectation along with
the main factors that were considered to form these expectations.
The observation period starts on January 2006 and goes until March
2017, a total of 135 months. The individuals consulted in the survey
are categorized into four different groups depending on their profes-
sional expertise: 1) consulting, 2) stock market analyst, 3) academic,
2
Consultancy of the 24- and 36-month variation in the crc/usd exchange
rate started on December 2016.
52 A. Alfaro, A. Mora
and 4) business sector. The number of respondents to the survey
and its composition have changed during the observation period;
there were 27 respondents in January 2006, most of whom were stock
market analysts and by March 2017, there were 61 respondents pre-
dominantly from the business sector. Figure 1 presents the composi-
tion of the sample group during the observation period.
Two features of the survey responses stand out: first, the total num-
ber of responses has increased more than twofold since the survey
was first implemented, with a peak of 87 responses in June 20133 .
Second, the composition of responses has drastically changed in the
last years of the survey–the majority of responses have recently come
from individuals working in the business sector–. This compositional
shift has resulted from a change in the survey design from June 2012
to the present4 .
The bccr computes the 12-month expected inflation by averag-
ing the responses received during a particular month, expectations
coming from the business sector are dominant in the expectations
published, representing up to 80% of the responses since 2015. This
dominance of the business sector in the average expected inflation
can be observed in Figure 2 where the mean expectation is plotted
for the whole sample and by group.
The average expectation has clearly declined, staying in the single
digits since April 2009, and below 5% since April 2015. The behavior
exhibited by the inflation expectations has been in accordance with
the inflation target range of the bccr (3%-5%) since April 2015. In Jan-
uary 2016, even though the inflation target range was downgraded
to 2%-4%, expectations have continued to remain within the range
up until the last month in our sample, March 2017.
The alignment between the expected inflation rate and the tar-
get inflation range in recent years highlights the built-up credibility
of bccr towards society. For the thirty-year period preceding 2009,
Costa Rica experienced double-digit inflation rates, but the bccr
has seemingly regained credibility. Agents trust the bccr to steer
the inflation rate, which thereby anchors inflation expectations. De-
spite this tendency for inflation expectations to lie within the target
3
With 64 of them from the business sector.
4
The two samples were active for several months, but the aggregate
results did not differ.
60
40
Responses
20
0
1 4 7 10 1 4 7 10 1 4 7 10 1 4 7 10 1 4 7 10 1 4 7 10
2006 2007 2008 2009 2010 2011
Months
100
80
Responses
60
40
20
0
1 4 7 10 1 4 7 10 1 4 7 10 1 4 7 10 1 4 7 10 1
2012 2013 2014 2015 2016 2017
Months
54 A. Alfaro, A. Mora
Figure 2
EXPECTED INFLATION MONTHLY AVERAGE
16
14
12-months expected inflation
12
Monthly average
10
8
6
4
2
0
2006 2007 2008 2009 2010 2011 2012 2013 2014 2015 2016 2017
Survey date
5
Figure 9 in the Annex, shows the increase of outliers on the expecta-
tions from the business sector in recent years.
6
Since the change in the design of the survey sample involved different
nomenclature for the identifiers, the same individual can have two
identifiers, one under the former sample and another one with the
current sample.
56 A. Alfaro, A. Mora
formation by pinning down their faculty to replicate features observed
on the data. Figure 3 presents the interquartile range observed every
month by group, along with the realized inflation rate for the month
that these expectations were registered. This is done to assess whether
the dispersion tends to increase when inflation is high, as has been
suggested by Ballantyne et al. (2016) and Johannsen (2014), among
others.
For the stock market analyst and academic groups, the interquar-
tile range and inflation rate attain their maximum in the last months
of 2008. For these two groups, it may seem to be a positive correlation
between the level of inflation and interquartile range during years
near the 2008-2009 financial crisis. Nonetheless, there are periods
in which the inflation rate decreases but the dispersion of the sample
expectations does not follow the same trend; the clearest example
is the dispersion within business sector responses since 2015 the in-
terquartile range has moved around 2% despite the sharp decline
in inflation. This suggests that for the Costa Rican case there is no
clear direct relation between the dispersion in inflation expectations
and the level of inflation.
A basic regression exercise between dispersion as measured by the
interquartile range and the inflation level is shown in Table 2. Regress-
ing the interquartile range by the inflation rate does not illustrate
a significant relation between the two groups: the associated coeffi-
cients are not significant when taking into account the whole survey
or individual groups.
Elliott et al. (2008) and Engelberg et al. (2009) note that disagree-
ment among inflation expectations does not necessarily indicate
that agents face different degrees of uncertainty when forming their
expectations. This is because the survey collects point predictions
from which individual distributions or probabilistic beliefs of pos-
sible outcomes for future inflation cannot be inferred. It is possible
that two forecasters who hold identical probabilistic beliefs provide
different point predictions and it is also possible that two forecasters
with different probabilistic beliefs provide the same point forecast.
When using point forecasts, we can only interpret the phrase dis-
agreement among expectations as an acknowledgment of distinct point
forecasts; we cannot conclude anything about the uncertainty that
forecasters face.
4 20
15
Interquartile range
10
2
5
0 −5
2006
2006
2007
2008
2009
2009
2010
2011
2012
2012
2013
2014
2015
2015
2016
Inflation (right axis) Consulting
6 20
15
Interquartile range
4
10
5
2
0
0 −5
2006
2006
2007
2008
2009
2009
2010
2011
2012
2012
2013
2014
2015
2015
2016
58 A. Alfaro, A. Mora
Figure 3 (cont.)
INTERQUARTILE RANGE BY GROUP
4 20
15
Interquartile range
10
2
5
0 −5
2006
2006
2007
2008
2009
2010
2011
2011
2012
2013
2014
2015
2015
2016
Inflation (right axis) Stock market
4 20
15
Interquartile range
10
2
5
0 −5
2006
2006
2007
2008
2009
2010
2011
2011
2012
2013
2014
2015
2015
2016
Whole Stock
Coefficient survey Consulting market Academic Business
Constant 1.591c 1.209c 1.135c 1.254c 1.446c
(0.090) (0.110) (0.090) (0.138) (0.134)
Inflation −0.013 −0.007 0.003 −0.002 −0.011
(0.012) (0.015) (0.012) (0.019) (0.018)
N 135 135 135 135 135
R2 0.0087 0.0014 0.0006 0.0001 0.0029
7
The Instituto Nacional de Estadísticas y Censos (inec) of Costa Rica
publishes the inflation rate of month t until the first days of month t +1.
60 A. Alfaro, A. Mora
of the survey. The last eleven months do not yet have a realized infla-
tion level to compare to, since the last observed inflation in this paper
is March 2017. Panel A of Figure 4 compares the expected and real-
ized inflation rates, while panel B. shows the average realized bias.
Our measure of realized bias has exhibited cyclical behavior, reach-
ing its minimum at the end of 2008 and its maximum at the end of
2009. While there are months where the realized bias has been prac-
tically zero, suggesting good predictive power, it has been positive
since 2005, meaning that on average, inflation expectations have
been greater than realized inflation.
The average realized bias seems to have a general upward trend
across the entire observation period, standing above 5% during
most of 2015 and part of 2016, but decreasing since the second se-
mester of 2016. The average realized bias does not differ substantial-
ly by group–Figure 5 shows the average realized bias for each group
and also for the entire survey sample–.
As expected, the business sector has dominated recent survey re-
sults–the average bias of the business sector has largely aligned with
the average of the entire survey sample–. In addition, the average bias
has increased over the years for all four groups. Figure 5 suggests that
the differences among groups are not significant, but this can be ex-
plained as a result of using measures of central tendency such as the
average. On the other hand, valuable information can be extracted
by studying disagreement among inflation expectations via statis-
tics of dispersion. The next section explores the role of information
rigidities in explaining the heterogeneity in inflation expectations.
62
inflation
−10
−5
0
5
10
−4
−2
0
2
4
6
8
10
12
14
16
18
2006 2006
2007
2007 2007
2008
2008
A. Alfaro, A. Mora
2009 2008
2009
2009 2009
2010
2010
2010
Inflation
2011
2011
Figure 4
2011 2011
2012
.
Expected
2014
EXPECTED AND REALIZED INFLATION
2014
2014 2014
2015
2015
2015
2016
2016
2016 2016
Realized expectations bias Realized expectations bias Realized expectations bias Realized expectations bias
−10
−5
0
5
10
15
−10
−5
0
5
10
15
−10
−5
0
5
10
15
−10
−5
0
5
10
15
2006 2006 2006 2006
2007 2007 2007
2007
2008 2008 2008
Average bias
Average bias
Average bias
Figure 5
Average bias
2012 2012 2012 2012
2013 2013 2013
2013
2014 2014 2014
Business
AVERAGE REALIZED BIAS BY GROUP
2014
Consulting
Consulting
Stock market
2014 2014 2014
2015
2015 2015 2015
2015
2016 2016 2016
2016 2016 2016
2016
∞
1 Ft xt +h = (1 − λ )∑ λ j Et − j xt +h .
j =0
8
In this equation the probabilities of an update are reparametrized so
only (1 − λ ) percent of the agents update their information sets and
acquire no new information with probability λ .
64 A. Alfaro, A. Mora
λ
2 xt +h − Ft xt +h = ( Ft xt +h − Ft −1xt +h ) + vt +h,t .
1− λ
3 et +1 − Ft et +1 = β ( Ft et +1 − Ft −1et +1 ) + εt .
Estimates for Equation 3 at the survey and the group level are shown
in Table 3. These regressions can be used to assemble evidence for in-
formation rigidities present on the survey. Under a sticky information
model, the β coefficient in Equation 3 should be significant, which
is the case at the survey level. An advantage of the relation between
the ex post forecast error and the ex ante forecast revision on Equa-
tion 3 is that it enables us to map the estimated coefficient βˆ to an
estimate of the information rigidity parameter λ. In our case, this
gives an estimate of λ ( )
= β 1 + β ≈ 0.1797 1.1797 ≈ 0.15237, which
suggests that 84.76% of the agents update their information sets at a
particular period and that on average an agent updates his or her in-
formation every 1.2 months.
At the group level, the estimates of Equation 3 suggest that the ev-
idence for information rigidities is stronger among some groups
compared to others. The β coefficient for Equation 3 is significant
to various degrees among the groups, with the exception of the ac-
ademic. For consultants and stock market analysts, the coefficient
is significant at a 1% level and only at a 10% level for the businesspeo-
ple. The results imply different estimates for the rate of information
acquisition λ among groups: 82.44% of the consultants, 83.61%,
of the stock market analysts, 91.91% of academics, and 91.07% of the
Dependent variable
et +h–Ftet +h
Stock
Survey Consulting market Academic Business
Ftet +1–Ft –1et +1 0.1797c 0.213c 0.196c 0.088 0.098a
(0.056) (0.058) (0.056) (0.056) (0.056)
Observations 240 240 240 240 240
R2 0.041 0.054 0.049 0.010 0.013
Adjusted R2 0.037 0.050 0.045 0.006 0.009
Residual standard
20.290 20.940 19.937 20.367 21.007
error (df = 239)
F statistic (df = 1; 239) 10.192c 13.643c 12.409c 2.492 3.129a
9
One should keep in mind that the estimate for the academic group
is not significant and for the business group is only significant at the
10% level. The coefficients are essentially unchanged if the model is
estimated using a constant.
66 A. Alfaro, A. Mora
(var) model to generate rational forecasts10. The var model uses Cos-
ta Rican monthly data from January 1996 to March 2017 for inflation
(πt ), interest rate (it ), output gap ( yt ), an inflation index of trade
( ) ( )
partners π t C , oil prices pt oil , and annual exchange variations (et ).
The design of the var model with two lags11 is presented in 4.
4 zt = A1zt −1 + A2 zt −2 + ut ,
with
πt
i
t
yt
zt := C .
πt
oil
pt
et
10
We attempted unsuccessfully to estimate the degree of information
rigidity directly for inflation forecasts, using instrumental variables
similarly to Coibion and Gorodnichenko (2015).
11
Number of lags suggested by the Hannan-Quinn information criterion.
12
We regress imae series at date t +24 (to include a two-year period) on
the four most recent values as of date t. The residuals from this regres-
sion are set to be the cyclical component of the series.
13
Mainly composed by the inflation of the United States, the euro zone,
China and Central American countries.
68 A. Alfaro, A. Mora
4) For every t = 2, …, 135, and for each individual i = 1, …, n, a Ber-
noulli experiment with probability of success λ will be con-
ducted.
a) If the experiment is a success, individual i at time t will re-
port his or her expected 12-month forward inflation rate
π te+12 t using the 12-month forecast from the var model es-
timated with information up to time t −1:
{
5) The previous steps give for each period t a series π ie,t +12 t }
for i = 1, …, n. For each series the mean and the interquartile
range (iqr) are recorded.
14
We compute the mean of square differences between the simulated
and observed series.
70
3
4
5
0
1
2
6
7
8
0
2
4
6
8
10
12
14
18
2006 2006
2007 2007
2007 2007
2008 2008
2008 2008
2009 2009
A. Alfaro, A. Mora
2009 2009
2009 2009
2010 2010
2010 2010
2011 2011
2011 2011
Observed
Observed
2011 2011
2012 2012
Figure 6
2012 2012
2013 2013
. FL
Simulated
Simulated
2015 2015
STICKY INFORMATION MODEL SIMULATION
2015 2015
2016 2016
2016 2016
2016 2016
On the other hand, the simulated series for the interquartile range
has a correlation of only 22.55% with the series from the survey. From
panel B of Figure 6 we observe that simulated iqrs are close to the
real iqrs only in the second half of the survey. This is due to a depar-
ture from the original algorithm in Mankiw et al. (2003) where λ
is selected to maximize the correlation between the simulated series
of iqrs and the survey series. Since we are interested in the mean ex-
pectation, our simulation was modified to put more emphasis on rep-
licating the mean expectation.
The evidence of this simulated model also suggests that the sticky
information assumption may not be appropriate. The value of the
parameter λ required to match the dynamic of the mean forecast
implies dynamics of disagreement that vary significantly from those
found in the data.
4. CONCLUSIONS
72 A. Alfaro, A. Mora
ANNEX
Respondent code:
5. How do you consider that the general economic conditions for pri-
vate production activities will evolve in the next six months in con-
trast with the past six months? (Please check one box)
Will improve []
Will be the same []
Will deteriorate []
Explain why:
6. How do you label the current conditions for firms to invest in the
country? (Please check one box)
Good conditions []
Bad conditions []
Not sure []
Contact: [email protected]
Telephone: (506) 2243-3312. Fax: (506) 2243-4559
The Department of Economic Research makes readily available doc-
uments elaborated on topics related to: inflation, monetary policy,
financial stability, etc. If you want to subscribe, go to the following
address: http://www.bccr.fi.cr/suscripcion/default.aspx
74 A. Alfaro, A. Mora
Annex B. Expected Inflation, Responses
and Dispersion by Group
Figure 7
INFLATION EXPECTATIONS SURVEY, RESPONSES BY GROUP
18
14
12
10
Responses
6
4
2
0
2006
2006
2007
2007
2007
2008
2008
2008
2009
2009
2009
2010
2010
2010
2011
2011
2011
2012
2012
2012
2013
2013
2013
2014
2014
2014
2015
2015
2015
2016
2016
2016
2017
2006
Figure 8
DISPERSION OF EXPECTED INFLATION BY GROUP
20
15
10
0
Jan 1, 2006
Jan 1, 2007
Jan 1, 2008
Jan 1, 2009
Jan 1, 2010
Jan 1, 2011
Jan 1, 2012
Jan 1, 2013
Jan 1, 2014
Jan 1, 2015
Jan 1, 2016
Jan 1, 2017
76
Jan 1, 2006 Jan 1, 2006 Jan 1, 2006
A. Alfaro, A. Mora
Jan 1, 2009 Jan 1, 2009 Jan 1, 2009
Figure 8 (cont.)
78 A. Alfaro, A. Mora
Formation and Evolution of Inflation
Expectations in Paraguay
Abstract
The establishment of the inflation targeting regime in Paraguay is relatively
recent, however, the results have been satisfactory. This is because based on the
observed data, from the implementation of this framework, it has been possible
not only to reduce inflation levels, but also align inflation expectations along
the medium-term inflation target. This chapter seeks to identify the main deter-
minants of the formation of inflation expectations in Paraguay since the adop-
tion of the inflation targeting regime. This work bases the analysis on the results
obtained from the expectations surveys conducted by the country’s monetary au-
thority. The evolution of inflation should be an important factor to consider.
Moreover, for the correct functioning of the expectations channel, it is essen-
tial that the monetary authority has sufficient credibility. A credibility index
has been constructed to capture the effect of the credibility that the Banco Central
del Paraguay has acquired during the inflation targeting regime. To guarantee
the robustness of our results, the model we use has been estimated by three econo-
metric methods: ordinary least squares (ols), fully modified ols (fmols),
79
and the generalized method of moments (gmm). According to the outcomes
of all these methods, inflation expectation formation in Paraguay is deter-
mined mainly by the inflation expectation of the previous month. In addition,
the annual inflation information of the previous month is significant at the
time of forming expectations.
Furthermore, the credibility index presents an expected negative sign, as in-
flation expectations have effectively aligned around the medium-term infla-
tion target since the implementation of the inflation targeting regime. The ex-
change rate was not significant in the regressions. This could partly be due
to a relatively low pass-through of the exchange rate to total inflation, espe-
cially in the last few years.
Keywords: inflation targeting, inflation expectations, monetary policy.
jel classification: E31, E52, E58.
1. INTRODUCTION
T
he Banco Central del Paraguay (central bank of Paraguay, bcp)
officially adopted the inflation targeting regime to regulate
its monetary policy in May 2011. Prior to this policy frame-
work, Paraguay exhibited marked levels of volatility even though
there were no historical records of high inflation periods. Under
the inflation targeting regime, volatility and inflationary levels have
been reduced. These inflationary levels fostered uncertainty in eco-
nomic agents when forming their inflation expectations. All this
was reflected in the fact that these expectations showed considerable
variability, in accordance with the results obtained in the expecta-
tions surveys of economic variables carried out by the central bank
on a monthly basis.
The main purpose of this chapter is to try to identify some of the
determinants that Paraguay’s economic agents consider when form-
ing their inflationary expectations. In view of the results of the sur-
vey, a series of factors that may influence the expectations formation
of those who answered the survey have been considered. To do this,
simple econometric regressions are carried out, and the results
of these can be considered a first attempt to find the determinants
of inflation expectations in Paraguay. In addition, the regressions
highlight the importance of the establishment of the inflation target-
ing framework, not only in reducing inflation levels and their vola-
tility, but also lowering inflation expectations. Furthermore, it can
80 P. Alonso
be affirmed that the bcp has managed to gain significant credibility
with respect to the handling of the monetary policy in its attempt
to maintain a low and stable inflation. This is reflected in the cred-
ibility index, which shows the alignment of expectations around
the inflation target since the establishment of the inflation target-
ing regime.
Inflation expectations play a critical role in the process of price
formation in the market. In addition, the decisions of households
and firms depend heavily on the real return that could be expected
on the savings and investments they make. Therefore, central banks
closely monitor the development of inflation expectations in order
to implement their monetary policy in a successful manner.
The results of the empirical model of this chapter show that the es-
tablishment of the inflation targeting scheme has helped to anchor
expectations around the target, and that the dispersion of these ex-
pectations has been adjusted within the inflation range. Further-
more, this dispersion has been reduced with the decrease of the range
during the consolidation process of the inflation targeting regime.
The first part of this chapter contains a brief narrative of mon-
etary policy in Paraguay, highlighting their main characteristics,
and delineates the most important results obtained from it, espe-
cially since the implementation of the inflation targeting framework.
Next, the importance of inflation expectations in monetary policy,
in general and specifically in Paraguay, is highlighted. Subsequently,
after a description of the characteristics of the data according to the
results of the economic variables survey, an estimation model of in-
flation expectations determinants in Paraguay is shown. The main
outcomes of the model show the robustness of the results through
different methodologies of estimating. Finally, in the last section
some conclusions and final comments are presented.
50
40
30
20
10
−10
1960
1962
1964
1966
1968
1970
1972
1974
1976
1978
1980
1982
1984
1986
1988
1990
1992
1994
1996
1998
2000
2002
2004
2006
2008
2010
2012
2014
2016
Inflation GDP
unlike most countries of the region (Figure 1). Likewise, the main
problem regarding inflation has been its volatility. The macroeco-
nomic performance of Paraguay can be attributed in part to the
sound management of monetary policy. This is reflected partly in the
fact that the guarani, the local currency of Paraguay, has not been
modified since its inception, thus making it one of the oldest curren-
cies in the region. The relatively prudent management of fiscal policy
has contributed, to certain extent, to keeping inflation at a low level.
As pointed out in the document Política monetaria en Paraguay:
Metas de inflación, un nuevo esquema (bcp, 2013), the design of mon-
etary policy in Paraguay has considered the existence of a relation
between the growth of money supply and inflation. Historically,
this design has adopted a monetary policy scheme of intermediate
objectives, in this case, setting targets for the growth of a specific
monetary aggregate. Thus, the Central Bank used its instruments
to control the money supply’s growth to a level compatible with
the inflation objective, which was based on the achievement of low
82 P. Alonso
inflation, using the quantitative theory of money as a conceptual
framework reference.
Regarding economic activity, in general, the average growth of the
Paraguayan economy has been acceptable, even though it has been
characterized by its volatility. While the expansion of the economy
was quite significant in the 1970s, mainly due to the construction
of the Itaipu hydroelectric dam, there was a period of slowdown in the
1980s and 1990s. In this weakened situation and as a consequence
of a weak financial system, and the fragility of the regulatory and su-
pervisory frameworks, between 1995 and 1998, there were episodes
of large financial crises. In this period, economic authorities needed
a comprehensive reorganization of monetary and financial policy,
which was attained through the enactment of important laws that
allowed a much more stringent regulatory framework for financial
institutions.1
In 2002, the Argentine economy fell into a deep crisis, causing
the abandonment of the convertibility regime to which that country’s
exchange rate policy was subordinated. This episode also affected
the Paraguayan economy. Despite the bcp’s effort to curb capital
outflows and exchange rate depreciation through sharp increases
in the interest rates of monetary regulation instruments, the second
financial crisis occurred towards the end of 2002, although of small-
er magnitude than the first one.
Despite these episodes of crisis, the enactment of the aforemen-
tioned regulatory laws for the financial system allowed the bcp to fo-
cus more on the achievement and maintenance of low and stable
inflation, driving its monetary policy of intermediate objectives,
under a monetary aggregates framework.
As of 2004, the bcp began to lay the foundations for the establish-
ment of an inflation targeting framework, albeit in an experimental
way. Thus, the central bank modernized its monetary policy opera-
tional instruments with the establishment of a medium-term infla-
tion target with a tolerance range. Under this scheme, it was possible
to reduce the average inflation rate in the period from 2000 to 2010
to a single digit level.2
1
The Law No. 489 of the bcp and the Law No. 861 “General of Banks,
Finance, and other Credit Institutions.”
2
In that period average inflation was 8.1%, while in the 1990-2000 period
it was 15.1 percent.
84 P. Alonso
The achievement of the objectives proposed by the central bank,
its transparency and communication increase its credibility, which
contributes to that the expectations remain anchored to the tar-
get in the policy horizon. When a central bank has built a credible
and transparent reputation, a monetary policy decision aimed at con-
trolling inflation keeps inflation expectations anchored to the target.
Therefore, in the face of an expectation of controlled inflation, deci-
sions to adjust prices and wages will be made in line with the inflation
target announced by the central bank.
Taking into account that the objective of clear and transparent
communication is to give signals about the implications of monetary
policy decisions, in general terms, the expectations channel may have
a more rapid impact on the achievement of the inflation target com-
pared to others transmission mechanisms that act with a greater lag.
This makes the expectations channel an important and timely chan-
nel for the effectiveness of monetary policy.
Since the implementation of the inflation targeting regime, the Ban-
co Central del Paraguay has made a great effort to improve its cred-
ibility. As mentioned above, Paraguay’s main problem has not been
high levels of inflation, but rather high volatility. Since the formal es-
tablishment of the inflation targeting scheme by the bcp, not only have
inflation levels been reduced, but, above all, their volatility has been
reduced (Figure 2). Likewise, it has been verified in the expectations
data that there has been a decrease both in their levels and their vol-
atility given the decrease in observed inflation rates. This suggests
that the bcp has managed to increase its credibility in recent years.
As mentioned above, an interesting fact that has been observed
with the implementation of the inflation targeting regime is the re-
duction of inflation expectations (average or median) to levels closer
to the target (Figure 3 and 4). Additionally, the dispersion has been re-
duced, mainly because of the reduction of the tolerance range in 2014.
The reduction of the tolerance range can be proven through tradi-
tional statistics of variability, such as the standard deviation and the
coefficient of variation (Figure 5), which effectively show a reduction
(on average) in recent years, coinciding with the reduction of toler-
ance bands.
Finally, it was run, as an additional test, a simple model of the vola-
tility statistics with respect to a dummy variable that takes the value
of 1 if there is a reduction in the band. The variable is significant with
an expected negative sign. In summation, these results suggest that
14
12
10
0
Apr2006
Oct2006
Apr2007
Oct2007
Apr2008
Oct2008
Apr2009
Oct2009
Apr2010
Oct2010
Apr2011
Oct2011
Apr2012
Oct2012
Apr2013
Oct2013
Apr2014
Oct2014
Apr2015
Oct2015
Apr2016
Oct2016
Apr2017
Oct2017
Inflation expectation year t Inflation expectation year t+1
Inflation (annual)
Source: Banco Central del Paraguay
Figure 3
DISPERSION OF INFLATION EXPECTATIONS FOR YEAR T¹
Percentage
14
12
10
0
Jan2011
Apr2011
Jul2011
Oct2011
Jan2012
Apr2012
Jul2012
Oct2012
Jan2013
Apr2013
Jul2013
Oct2013
Jan2014
Apr2014
Jul2014
Oct2014
Jan2015
Apr2015
Jul2015
Oct2015
Jan2016
Apr2016
Jul2016
Oct2016
Jan2017
Apr2017
Jul2017
Oct2017
86 P. Alonso
Figure 4
DISPERSION OF INFLATION EXPECTATIONS FOR YEAR T+1
Percentage
12
10
0
Jan2011
Apr2011
Jul2011
Oct2011
Jan2012
Apr2012
Jul2012
Oct2012
Jan2013
Apr2013
Jul2013
Oct2013
Jan2014
Apr2014
Jul2014
Oct2014
Jan2015
Apr2015
Jul2015
Oct2015
Jan2016
Apr2016
Jul2016
Oct2016
Jan2017
Apr2017
Jul2017
Oct2017
Bands Median Target
Source: Banco Central del Paraguay
Figure 5
STANDARD DEVIATION AND COEFFICIENT OF VARIATION
Percentage
30
2.0
25
1.5 20
1.0 15
10
0.5
5
0.0 0
Jan2011
Aug2011
Mar2012
Oct2012
May2013
Dec2013
Jul2014
Feb2015
Jan2011
Aug2011
Mar2012
Oct2012
May2013
Dec2013
Jul2014
Feb2015
Apr2016
Nov2016
Jun2017
Apr2016
Nov2016
Jun2017
88 P. Alonso
In order to identify the main determinants of the process of form-
ing expectations, it is necessary to have a series of fixed horizon infla-
tion expectations. To carry out an approximation of fixed horizon
forecasts from the fixed event forecasts of the eve, we follow the work
of Dovern et al. (2012), in which this approximation is made as a
weighted average of fixed-event forecasts as follows:
12 − (m − 1) fe m − 1 fe
1 Fyf0,hm ,12 (x ) = Fy 0,m ,y 0 (x ) + F (x )
12 12 y 0,m ,y 0+1
10
0
Apr2006
Oct2006
Apr2007
Oct2007
Apr2008
Oct2008
Apr2009
Oct2009
Apr2010
Oct2010
Apr2011
Oct2011
Apr2012
Oct2012
Apr2013
Oct2013
Apr2014
Oct2014
Apr2015
Oct2015
Apr2016
Oct2016
Apr2017
Inflation expectation year t Inflation expectation year t+1
Inflation expectation 12 month-ahead
Source: Banco Central del Paraguay
90 P. Alonso
the work of Mendonça (2007), in which it is assumed that the cen-
tral bank is able to guide inflation expectations towards the target
and reaffirm its commitment to the inflation ranges. Thus, when
expectations are equal to the inflation target the credibility index
is equal to one, and decreases when expectations move away from
the target. In cases where inflation expectations are located outside
the inflation target bands, the index is equal to zero (see Annex).
Finally, it may be thought that a priori changes in the nominal ex-
change rate (guarani-dollar) should influence the formation of in-
flation expectations of economic agents on the cost side of imported
goods (and inputs), especially when considering that Paraguay is a
relatively open economy.4 A similar analysis could be made when
considering variations in oil price, since this product directly affects
the price of fuels, an important input for any production process.
6
This index was constructed according to the work of Mendonça (2007),
whose criterion is described in the Annex.
7
See Banco Central del Paraguay (2015, recuadro 1).
92 P. Alonso
Table 1
ESTIMATED EQUATIONS FOR INFLATION EXPECTATIONS
Dependent variable: inflation expectations (12 month-ahead)
Models
ols fmols gmm
Models
ols fmols gmm
94 P. Alonso
in the range occurred. In this regard, interesting results are observed
in all the estimation methodologies, as they show that the lower in-
flationary range has had an impact on getting inflation expectations
adjusted to this new range (Table 2). This also shows that the bcp
has had a significant influence on the credibility of economic agents
in achieving the inflation goal under the inflation targeting regime.8
On the other hand, an exercise was carried out that reflects the be-
havior of the inflation expectations of the group of respondents
categorized as financial entities (banks and financial companies).
The results show that the expectations of the financial agents follow
a similar pattern to the base equation (Table 3).
5. CONCLUSION
8
The introduction of the band dummy variable diminishes the signifi-
cance from the credibility index. This could be due to the fact that both
variables reflect greater credibility in the inflation targeting scheme,
so that the two variables cannot be together in the same base equation.
This proves that the bcp has achieved significant credibility in its
purpose of keeping inflation low and stable around the inflation
target. Therefore, the alignment of inflation expectations around
the target can be attributed to an increase in credibility.
It should be noted that the reduction in inflationary bands also
reflects an adjustment of inflation expectations around the target, at-
testing likewise to greater credibility of economic agents in the man-
agement of monetary policy under the inflation targeting scheme.
In addition, when the respondents are grouped in the category of fi-
nancial entities, it is observed that the expectations of these agents
follow a pattern similar to that observed in the base equation.
96 P. Alonso
ANNEX
1 if π te − π *
1
1 − lower e *
* π t − π
if π lower < π te
π − π
.
1 upper e
1 − upper e *
* π t − π
if π > πt
π −π
0 if π te ≥ π upper or π te ≤ π lower
References
Banco Central de Chile (2007), La política monetaria del Banco Central
de Chile en el marco de metas de inflación.
Banco Central de Reserva del Perú (2016), El impacto de la meta de
inflación sobre la inflación.
Banco Central del Paraguay (2003), Política monetaria en Paraguay:
metas de inflación, un nuevo esquema.
Banco Central del Paraguay (2015) Informe de política monetaria de
septiembre de 2015, <https://www.bcp.gov.py/informe-de-
politica-monetaria-i14>.
Banco Central do Brasil (2015), Inércia inflacionária e determinantes
das expectativas de inflação.
Cerisola, Martin, and R. Gaston Gelos (2005), What Drives Infla-
tion Expectations in Brazil? An Empirical Analysis, imf Working
Paper, No. wp/05/109, June.
Coibon, Olivier, and Yuriy Gorodnichenko (2012), Information Rigid-
ity and the Expectations Formations Process: A Simple Framework and
New Facts, imf Working Paper, No. wp/12/296, December.
98 P. Alonso
The Degree of Inflation
Expectation Anchoring
Anchoring of Inflation Expectations
in Latin America
Rocío Gondo
James Yetman
Abstract
We use inflation survey data from Consensus Economics to assess how firmly
inflation expectations are anchored in Latin America. Following the method-
ology proposed by Mehrotra and Yetman (2018), we model inflation forecasts
using a decay function, where forecasts monotonically diverge from an esti-
mated anchor towards recent actual inflation as the forecast horizon short-
ens. Our results suggest that most countries do have an inflation anchor,
with the estimated weight of the anchor increasing through time, indicating
more strongly anchored expectations. This is consistent with the improving
credibility of central banks’ monetary policy management over our sample
period (1993-2016). For countries with formal inflation targets, our results
indicate that inflation targeting regimes are generally credible, with estimated
anchors lying within the inflation target range for all countries in the most
recent sample that we consider.
Keywords: inflation expectations, inflation anchoring, decay function.
jel classification: E31, E58.
101
1. INTRODUCTION
M
onetary policy effectiveness, and especially the achievement
of price stability, can be greatly assisted when inflation expec-
tations are well anchored. In many models of inflation, for ex-
ample, volatile inflation expectations directly increase the volatility
of inflation outcomes. In Latin America, with a history of repeated
episodes of high inflation, many countries have adopted inflation
targeting (it) as a framework to support a move to low and stable
inflation and provide for better anchoring of inflation expectations.
Some of these countries have adopted a schedule of decreasing targets
over time with a view to gradually reducing inflation.
Challenges of inflation control for central banks in the region
remain. In 2015-2016, some countries experienced inflation rates
above the top of their target ranges, mainly commodity exporters
who experienced large currency depreciations. In the cases of Co-
lombia and Peru inflation expectations appear to have become de-
anchored to some extent, with high inflation persisting (see Figure
1). Monetary policy tightening actions were taken in response to these
developments, with their central banks raising policy rates by 3.25%
and 1%, respectively.
The goal of this paper is to assess whether or not countries have
an inflation expectations anchor and, if they do, how strongly in-
f lation expectations are anchored. For economies with formal
it frameworks, we also examine whether the anchor is consistent
with the central bank’s target. We define an inflation anchor as the
expected level of inflation in the absence of any shocks to the econ-
omy. It should be noted that the inflation anchor is not necessar-
ily equal to the inflation target for countries with an it framework.
For each country, first, we evaluate whether there is an anchor
for inflation expectations and, if so, how the anchor has evolved over
time. Second, we analyze how well identified the inflation anchor is,
using the standard deviation of the estimated anchor as an indica-
tor of the degree of anchoring. Third, we compare the anchoring
of inflation expectations between countries in the region that have
inflation targets with such anchoring in those that do not.
We model inflation forecasts using a decay function, where fore-
casts monotonically diverge from the estimated anchor towards re-
cent actual inflation as the forecast horizon shortens. We estimate
1 2
5 5
4 4
3.6 3.0
Percentages
Percentages
3 3
2.9
2 2
1 1
0 0
2014
2015
2016
2014
2015
2016
2017
2014
2015
2016
2014
2015
2016
2017
Central bank survey Expectations by:
Target Financial institutions
Economic analysts
Nonfinancial companies
1
Expectations for 12-month inflation. 2 Expectations of current year (December) 12-
month inflation.
Sources: National data.
1 f (t ,t − h ) = α (h )π * + (1 − α (h ) )π (t − h ) + ε (t ,t − h ) ,
1
We parametrize the model to separately identify the anchor and the
coefficients indicating the weight on the anchor. If there is a link
between the two (for example, adopting an inflation target leads to a
change in both the anchor and how strongly inflation is anchored), our
estimation allows for this possibility but does not impose it. As such, it
may be possible to improve the efficiency of the estimation approach
taken here.
2
Our results are conditional on the decay function. Mehrotra and Yetman
(2018) demonstrate that, provided inflation follows an autoregressive
process, a monotonically decreasing decay function should fit inflation
expectations.
3 V ( ε (t ,t − h ) ) = exp (δ 0 + δ1h + δ 2h 2 ).
The use of the exponential function here ensures that the fitted
values of the variance are positive for any values of the parameters
defining the variance ( δ 0 , δ 1 and δ 2). Note that, aside from this re-
striction, our modeling assumptions for the variance are very flexible:
It can be increasing or decreasing in the forecast horizon, or even
follow a u-shaped (or inverse u-shaped) pattern across horizons.
Forecasts made at different horizons for the inflation outcome
in a given year t are likely to be correlated, and more strongly so the
closer the two horizons are. Therefore, the correlation between the re-
sidual at two different horizons h and k is modeled as:
4 corr ( ε (t ,t − h ),ε (t ,t − k ) ) = φ0 + φ1 h − k .
Table 1
LIST OF COUNTRIES AND SAMPLE
a
Transition to an explicit it regime started in 2005 with the announcement
of an annual inflation target.
4. RESULTS
3
Consensus Forecasts also publishes average forecasts at longer horizons,
of up to ten years, for some economies in our sample, but these are
only available twice per year.
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1990-1997
1992-1999
1994-2001
1996-2003
1998-2005
2000-2007
2002-2009
2004-2011
2006-2013
2008-2015
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1.0 1.0
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1990-1997
1992-1999
1994-2001
1996-2003
1998-2005
2000-2007
2002-2009
2004-2011
2006-2013
2008-2015
1.0
0.8
0.6
0.4
0.2
0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
Notes: Horizontal axis displays the eight-year rolling sample. Periods where no line is
displayed correspond to rolling samples for which no anchor can be identified.
Source: Authors’ calculations.
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1.0 1.0
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
Notes: Horizontal axis displays the eight-year rolling sample. Periods where no line is
displayed correspond to rolling samples for which no anchor can be identified.
Source: Authors’ calculations.
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1.0 1.0
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1.0
0.8
0.6
0.4
0.2
0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
Notes: Horizontal axis displays the eight-year rolling sample. Periods where no line is
displayed correspond to rolling samples for which no anchor can be identified.
Source: Authors’ calculations.
4
In our modeling of inflation expectations, we are implicitly assuming
that changes in inflation persistence reflect changes in the anchor-
ing of inflation expectations. To the extent that declining inflation
persistence reflects changed price-setting mechanisms that results
from greater anchoring of inflation expectations, this assumption is
warranted (see Section 4.3). But there may be other, more mechani-
cal sources of changes in inflation persistence–such as changes in the
sectoral composition of the economy–that could bias our results.
b c π* s.e.( π *)
1
For Venezuela, results are for 2008-2015, since data are not available for 2016.
5
The estimated confidence intervals for the inflation anchor depend
on the functional form of the decay function. However, for a sample of
advanced and emerging countries, Mehrotra and Yetman (2018) find
that the Weibull-based decay function fits the data better than more
restrictive forms, and more general forms do not increase explanatory
power markedly.
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
10 12.5
8 10.0
6 7.5
4 5.0
2 2.5
0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1990-1997
1992-1999
1994-2001
1996-2003
1998-2005
2000-2007
2002-2009
2004-2011
2006-2013
2008-2015
8
6 Anchor
95% confidence
4
Inflation target range
2
0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1
Horizontal axis displays the eight-year rolling sample. Periods where no line is
displayed correspond to rolling samples for which no anchor can be identified.
Source: Authors’ calculations.
10 6
8 4
6 2
4 0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
20 12
15 10
8
10
6
5
4
0 2
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1990-1997
1992-1999
1994-2001
1996-2003
1998-2005
2000-2007
2002-2009
2004-2011
2006-2013
2008-2015
7
8
6
4
5
0 4
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
25 5
20 4
15 3
10 2
5 1
0 0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
40
35
30
Anchor
25 95% confidence
20
15
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1
Horizontal axis displays the eight-year rolling sample. Periods where no line is
displayed correspond to rolling samples for which no anchor can be identified.
Source: Authors’ calculations.
4.3 Effect of it
Next, we focus on the sample of countries with it and analyze wheth-
er or not the estimated anchor is consistent with the inflation tar-
get. By doing so, we are assessing whether our results are consistent
with these countries building credibility for their it monetary pol-
icy frameworks. 7 We focus on the average across all rolling samples
where a country has an it framework. Table 3 shows that the estimated
6
On the other hand, long-horizon forecasts (e.g., six-to-ten years ahead)
might relate to outcomes too far into the future to be useful for monetary
policy purposes. For monetary policy setting, the most relevant horizon is
related to the frequency with which most prices and wages are adjusted,
and hence has the greatest impact on inflation dynamics. Thus, one could
imagine wage and price-setting decisions being influenced by inflation
expectations that are anchored by a level of expected inflation that dif-
fers from expectations of long-run inflation (if, for example, forecasters
anticipated that the monetary policy framework might be adjusted in a
few years). In that case, six-to-ten year ahead inflation expectations might
not be relevant for explaining inflation dynamics, but they could still
be important for other economic decisions such as deciding to invest
in fixed assets or determining long-term savings goals.
7
The anchor of inflation expectations could become more consistent
with the inflation target, even if the central bank is not building cred-
ibility, e.g., if inflation moves towards the target for reasons unrelated
to monetary policy or the inflation target is adjusted endogenously to
track inflation. In the former case, these effects are likely to be transitory
(so are mitigated against in part by our use of rolling samples). With re-
spect to the latter case, we see limited evidence of inflation targets being
adjusted strategically in response to deviations of inflation from target
in the inflation targeters that we examine: Inflation targets are either
constant over most of the 2009-2016 period (Brazil, Chile, Colombia,
Guatemala, Mexico, Peru, and Uruguay) or follow a consistent declin-
ing path as inflation targets become more established over time (Costa
Rica, the Dominican Republic, and Paraguay).
25
20
15
10
0
0 5 10 15 20 25
Inflation anchor
25
20
15
10
0
0 5 10 15 20 25
Inflation anchor
0
0 1 2 3 4 5 6 7 8 9
Inflation anchor
1
The inflation target is the simple average of the annual inflation target for each
country in the given sample. 2 For countries that adopted it later than 2009 such
as the Dominican Republic and Paraguay, the sample starts in 2012 and 2011,
respectively.
f (t ,t − h ) = α (h )(π T (t ) + d ) + (1 − α (h ))π (t − h ) + ε (t ,t − h ).
b c d s.e.(d)
Brazil 6.37 0.38 0.38 0.028
Chile 2.58 0.56 –0.02 0.004
Colombia 11.27 0.74 0.35 0.014
Costa Rica 3.31 0.56 0.97 0.030
Dominican Republic 5.86 0.46 0.12 0.014
Guatemala 9.01 0.45 1.46 0.027
Mexico 1.29 0.29 0.54 0.005
Paraguay 1.34 1.17 0.10 0.017
Peru 0.32 0.12 0.55 0.016
Uruguay 1.45 0.52 1.67 0.025
Note: Uruguay has a target range of +/–2 percentage points; all other countries
have a target range of +/–1 percentage point.
Table 4 shows the results of these estimations, for the most recent
eight-year rolling sample. These confirm that the anchors of infla-
tion expectations are in line with the inflation target range in all
countries: within a +/−1 percentage point range in all cases except
for Guatemala and Uruguay, the latter of which has an inflation tar-
get range of +/−2 percentage points. That is, we cannot reject the hy-
pothesis that inflation expectations are anchored by the inflation
targets for most countries.
In order to complement the comparison between countries with
and without inflation targets, we further examine whether it im-
proves the anchoring of expectations. To do this, we perform a sec-
ond step panel estimation. We regress the weight of the anchor (α (h ))
for each country for each eight-year rolling sample on a set of country
characteristics. The set of regressors includes: 1) a dummy variable
that takes the value of 1 for countries with it for the full rolling sample
during the rolling sample; 2) the number of years since the adoption
of the it regime; 3) mean inflation; 4) inflation variability, measured
by the standard deviation of inflation; 5) inflation persistence, based
5. CONCLUSIONS
8
At a forecast horizon of 12 months, being under an it regime is associated
with an increase in the degree of anchoring of inflation expectations
by 0.25, and a 0.1 drop in inflation persistence is associated with an
increase in the degree of anchoring by 0.09.
Table 6
SECOND STEP ESTIMATION RESULTS
Dependent variable: standard error of the inflation anchor
10 5
8 4
6 3
4 2
2 1
23 21 19 17 15 13 11 9 7 5 3 1 23 21 19 17 15 13 11 9 7 5 3 1
8 6
6 5
4 4
2 3
0 2
23 21 19 17 15 13 11 9 7 5 3 1 23 21 19 17 15 13 11 9 7 5 3 1
5
4 2010 2014
2011 2015
3 2012 2016
2013 2017
2
1
23 21 19 17 15 13 11 9 7 5 3 1
Notes: Horizontal axis represents the forecast horizon, defined as the number of
months before the end of the calendar year being forecast. Dots represent the
realized inflation at the end of year t.
Source: Consensus Economics ©; national data.
8 7.5
6 6.0
4 4.5
2 3.0
0 1.5
−2 0.0
23 21 19 17 15 13 11 9 7 5 3 1 23 21 19 17 15 13 11 9 7 5 3 1
8 8
7 7
6 6
5 5
4 4
3 3
2 2
23 21 19 17 15 13 11 9 7 5 3 1 23 21 19 17 15 13 11 9 7 5 3 1
10
9
2010 2014
8
2011 2015
7 2012 2016
6 2013 2017
4
23 21 19 17 15 13 11 9 7 5 3 1
Notes: Horizontal axis represents the forecast horizon, defined as the number of
months before the end of the calendar year being forecast. Dots represent the
realized inflation at the end of year t.
Source: Consensus Economics ©; national data.
40 8
30 6
20 4
10 2
0 0
23 21 19 17 15 13 11 9 7 5 3 1 23 21 19 17 15 13 11 9 7 5 3 1
7.5 8
6.0 6
4.5 4
3.0 2
1.5 0
0.0 −2
23 21 19 17 15 13 11 9 7 5 3 1 23 21 19 17 15 13 11 9 7 5 3 1
2010 2011 2012 2013 2014 2015 2016 2017
Notes: Horizontal axis represents the forecast horizon, defined as the number of
months before the end of the calendar year being forecast. Dots represent the
realized inflation at the end of year t.
Source: Consensus Economics ©; national data.
8 8
6 6
4 4
2 2
23 21 19 17 15 13 11 9 7 5 3 1 23 21 19 17 15 13 11 9 7 5 3 1
8 600
6 450
4 300
2 150
0 0
23 21 19 17 15 13 11 9 7 5 3 1 23 21 19 17 15 13 11 9 7 5 3 1
2010 2011 2012 2013 2014 2015 2016 2017
Notes: Horizontal axis represents the forecast horizon, defined as the number of
months before the end of the calendar year being forecast. Dots represent the
realized inflation at the end of year t.
Source: Consensus Economics ©; national data.
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1990-1997
1992-1999
1994-2001
1996-2003
1998-2005
2000-2007
2002-2009
2004-2011
2006-2013
2008-2015
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1.0 1.0
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1990-1997
1992-1999
1994-2001
1996-2003
1998-2005
2000-2007
2002-2009
2004-2011
2006-2013
2008-2015
1.0
0.8
0.6
0.4
0.2
0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
Notes: Horizontal axis displays the eight-year rolling sample. Periods where no line is
displayed correspond to rolling samples for which no anchor can be identified.
Source: Authors’ calculations.
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1994-2001
1996-2003
1998-2005
2000-2007
2002-2009
2004-2011
2006-2013
2008-2015
1.0 1.0
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
Notes: Horizontal axis displays the eight-year rolling sample. Periods where no line is
displayed correspond to rolling samples for which no anchor can be identified.
Source: Authors’ calculations.
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1.0 1.0
0.8 0.8
0.6 0.6
0.4 0.4
0.2 0.2
0.0 0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
1.0
0.8
0.6
0.4
0.2
0.0
1993-2000
1995-2002
1997-2004
1999-2006
2001-2008
2003-2010
2005-2012
2007-2014
2009-2016
Notes: Horizontal axis displays the eight-year rolling sample. Periods where no line is
displayed correspond to rolling samples for which no anchor can be identified.
Source: Authors’ calculations.
Abstract
This chapter analyzes the time-varying degree of inflation expectations an-
choring in Bolivia and, more precisely, whether inflation expectations have
been in line with the inflation objectives announced by the Banco Central
de Bolivia (central bank of Bolivia, bcb) and if they have become better an-
chored over time. Two considerations are particularly relevant in this regard.
First, the main sources of information are the bcb survey and Focus Econom-
ics survey, which only have data for short- and medium-term inflation expec-
tations. Second, monetary policy in Bolivia is under a monetary-targeting
regime, so bcb projections represent the main references. The anchoring de-
gree analysis of short-term inflation expectations was performed considering
bcb projections, while the medium-term analysis used an implicit inflation
target. In both cases, the results indicate there is a high degree of anchoring
137
of inflation expectations in Bolivia, especially during the last four years.
This study considers information from July 2005 to June 2017, with monthly
frequency.
Keywords: inflation expectations, anchoring degree, monetary-targeting
regime, bcb projections, time-varying parameters model.
jel classification: E31, E52, E58, C32.
1. INTRODUCTION
T
he analysis of the behavior of the expectations of inflation
of economic agents has been heavily studied in the past, espe-
cially with regards to the degree of anchoring of expectations,
understood as the ability of monetary policymakers to manage infla-
tion expectations (King, 2005). Theoretical literature and monetary
policymakers agree that the anchoring of inflation expectations is of
high importance in maintaining price stability, and expectations
by private agents play an important role in macroeconomics since
they can be a determinant of macroeconomic performance. Infla-
tion expectations not only reflect private agents’ perceptions about
future inflation, but also directly impact current and future inflation.
Relatedly, a central bank should focus on the management of pri-
vate expectations through communication for two reasons (Hubert,
2015). First, the expectations channel is one of the subtlest channels
of monetary policy, because it depends on private agents’ interpreta-
tion. As King (2005) notes, “because inflation expectations matter
to the behavior of the households and firms, the critical aspect of mon-
etary policy is how decisions of the central bank affect those expecta-
tions.” Second, given the delay between policy actions and their real
effects on macroeconomic variables, central bank communication
provides policymakers with a way to promptly affect private expec-
tations to shorten the transmission lag of monetary policy.
According to Blinder et al. (2008), central bank communication
can take different forms: statements, minutes, interviews, speech-
es, or internal macroeconomic forecasts. We will focus on the latter
instrument of communication because monetary policy in Bolivia
is under a monetary-targeting regime. However, although the Ban-
co Central de Bolivia(bcb, for its acronym in Spanish) does not have
an explicit inflation target, its active communication policy and pro-
jections, announced twice per year in its Monetary Policy Report,
1
Information for previous periods is not available.
2
It is important to note that fuels are subsidized in Bolivia. In December
2010, the government decided to withdraw the subsidy which gener-
ated an environment of uncertainty, causing expectations of inflation
to increase. Although the measure was eliminated shortly, important
second-round effects were generated during the following months.
3
Although, the dispersion of expectations in a survey is a measure
of heterogeneity of beliefs rather than a measure of uncertainty ( imf,
2016), both tend to move together (Gürkaynak and Wolfers, 2007).
4
Dovern, Fritsche and Slacalek (2009), Capistran and Ramos-Francia
(2010), Siklos (2013), and Ehrmann (2015).
5
During this period, Bolivian economy went through different circum-
stances that caused strong inflationary pressures: increased international
0
5
10
15
20
25
30
35
Jun 2005 Jun 2005
Oct 2005 Oct 2005
Percentage
Percentage
Feb 2006 Feb 2006
Jun 2006 Jun 2006
Oct 2006 Oct 2006
Feb 2007 Feb 2007
Jun 2007 Jun 2007
Oct 2007 Oct 2007
Feb 2008 Feb 2008
Jun 2008 Jun 2008
Oct 2008 Oct 2008
Feb 2009 Feb 2009
Jun 2009 Jun 2009
Max-Min range
Oct 2009 Oct 2009
Feb 2010 Feb 2010
Jun 2010 Jun 2010
Oct 2010 Oct 2010
Feb 2011 Feb 2011
Jun 2017
Figure 2
CROSS-SECTIONAL STANDARD DEVIATION OF ONE YEAR-AHEAD
INFLATION EXPECTATIONS
Percentage
7
0
Jun 2005
Oct 2005
Feb 2006
Jun 2006
Oct 2006
Feb 2007
Jun 2007
Oct 2007
Feb 2008
Jun 2008
Oct 2008
Feb 2009
Jun 2009
Oct 2009
Feb 2010
Jun 2010
Oct 2010
Feb 2011
Jun 2011
Oct 2011
Feb 2012
Jun 2012
Oct 2012
Feb 2013
Jun 2013
Oct 2013
Feb 2014
Jun 2014
Oct 2014
Feb 2015
Jun 2015
Oct 2015
Feb 2016
Jun 2016
Oct 2016
Feb 2017
Jun 2017
Source: Authors’ calculations based on data.
Percentage
16
14
12
10
8
6
4
2
0
2005 2006 2007 2008 2009 2010 2011 2012 2013 2014 2015 2016 2017
Note: Inflation expectations are computed as the mean of inflation expectations for a
given year. projections and the projection range are computed as the average of
the inflation projections announced at the beginning and middle of the year.
Source: National Statistics Institute and Central Bank of Bolivia.
x for year Y0 made in month m of year Y0 and Fy 0,m ,y1 (x ) the FE forecast
fe
of variable x for year Y1 made in month m of year Y0. Then Fy 0,m ,12 (x )
fh
12 − m + 1 m −1
Fy 0,m ,12 (x ) = * Fy 0,m ,yo (x ) + * Fy 0,m ,y1 (x ) (1)
fh fe fe
1
12 12
7
Annex 1 presents the evolution of the bcb projection for the current
and next calendar year separated, and the bcb inflation projections con-
structed using the technique of equation (1) that includes the updates
at middle of each year.
0
1
2
3
4
5
6
7
8
9
Jun 2005
Jun 2005 Oct 2005
Oct 2005 Percentage
Percentage
Feb 2006
Feb 2006 Jun 2006
Jun 2006 Oct 2006
Oct 2006 Feb 2007
Feb 2007
Jun 2007 Jun 2007
Oct 2007 Oct 2007
Feb 2008 Feb 2008
Jun 2008 Jun 2008
Oct 2008 Oct 2008
Feb 2009 Feb 2009
Jun 2009 Jun 2009
Oct 2009 Oct 2009
Feb 2010 Feb 2010
Jun 2010 Jun 2010
Oct 2010 Oct 2010
Feb 2011 Feb 2011
Jun 2011 Jun 2011
Figure 4
where:
γ proj + γ π = 1
e
where πt t +n represents the inflation expectations in period t for
proj
the horizon; t +n;πt+n is the inflation projection for the horizon;
t +n;πt−1 represents observed inflation lagged one period and n
is equal to 12 months. Additionally, an error term (µt ) is included
in the equation. Note that, by construction, the sum of the coefficients
of the model must be equal to one. If the coefficient γ proj reaches
a value as close as possible to one, it would reflect a significant de-
gree of anchoring of expectations.
According to Strohsal, Melnick and Nautz (2015), the central
bank’s credibility can be gained, but it can also be lost. As a conse-
quence, the degree of inflation expectations anchoring might not be
constant over time. Meanwhile, Orphanides (2015) once pointed
out that inflation expectations are well anchored until they are not.
This means that the degree of anchoring can change over time,
so using a model with constant parameters may not be the best op-
tion. In that sense, in the present document a time-varying param-
eter model is estimated, in line with other works such as Demertzis,
Marcellino and Viegi (2012) and Strohsal, Melnick and Nautz (2015).
In the name of simplification, we assume that the state param-
eters follow a random walk process. We use the Kalman filter (Kal-
man, 1960) to compute the one-step ahead estimates of the means
and variances 8 of the states by maximum likelihood.
8
During the estimation, the variances parameters are expressed in expo-
nential form to ensure that the variances themselves are non-negative.
Table 1
RESULTS FROM TIME-VARYING PARAMETER MODEL 1
9
It is also interesting to note that the degree of anchoring of expectations
did not decline in time of the international financial crisis, something
that was analyzed in different documents such as Galati, Poelhekke
and Zhou (2011), Autrup and Grothe (2014), and Nautz and Strohsal
(2015). However, this does not imply that in that period there was a
greater degree of central bank credibility.
−0.4
−0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
1.6
1.8
Jun 2005
Jun 2005 Oct 2005
Percentage
Oct 2005
Percentage
Feb 2006
Feb 2006 Jun 2006
Jun 2006 Oct 2006
Oct 2006 Feb 2007
Feb 2007
Jun 2007 Jun 2007
Oct 2007 Oct 2007
Feb 2008 Feb 2008
Jun 2008 Jun 2008
Oct 2008 Oct 2008
Feb 2009 Feb 2009
Jun 2009 Jun 2009
Oct 2009 Oct 2009
Feb 2010 Feb 2010
3 +n + β2π t −1 + βm X t ( L ) + µt
π te|t +n = β1π tproj (3)
e
Once again πt t+n represents one year-ahead inflation expectation;
πtproj
+n is the bcb inflation projection for the horizon t + n; where n is
equal to 12 months and πt−1 represents observed inflation lagged
one period. We include X t , which represents the battery of different
10
This result does not imply that inflations expectations are rational; that
issue is not analyzed in this study.
11
The information was obtained from the Global Index of Economic
Activity (igae, for its acronym in Spanish) which represents a proxy
variable of economic activity in monthly frequency, minus its trend
value (where the trend is approximated through a Hodrick-Prescott
filter).
12
Most of the documents use movements in the nominal exchange rate.
However, in Bolivia, the exchange rate has been fixed since 2011, and it
is an important variable since it works as a nominal anchor. For this
reason, we use economic agents’ expectations of future depreciation.
13
We use expectations of fiscal balance as a proxy of the primary fiscal
balance in order to have a variable with monthly data. For this case
and the expectations of nominal depreciation we use the information
from the bcb survey employing the technique of equation (1).
14
We employ the Multivariate enso (El Niño/Southern Oscillation)
Index (mei) of the United States National Oceanic and Atmospheric
Administration (noaa) as a proxy variable to reflect the changes in the
weather condition.
15
The Food Price Index of the International Monetary Fund (imf) was con-
sidered. International food price shocks have a significant impact
on inflation in Bolivia because of the high share of food in the country’s
cpi.
16
We also use other variables like igae growth YoY, economic agents’
expectations of economic growth and the imf international energy
price index; none of these, however, showed satisfactory results.
Table 2
DETERMINANTS OF SHORT-TERM INFLATION EXPECTATIONS
We created four different models, and in each one the bcb pro-
jection remained the most important explanatory variable with
coefficients around to 0.74, close to those obtained in Section 3.1.
Also, lagged inflation was significant (at 10 percent) in all models,
with a coefficient near 0.28. The remaining variables were not sta-
tistically significant. The least relevant were the international price
17
During 2007-2008 and 2010-2011, international food prices rose expo-
nentially, so national producers decided to sell most of their production
to foreign markets, generating a shortage in local markets. This caused
an increase in the prices of some foods (like sugar) or inputs (such
as soybeans that are important for poultry farms), which translated into
an inflationary process. However, in recent years international food
prices have fallen and shown less dynamism; in addition, limits were
applied to exports in order to ensure supply to local markets. These
factors may have diminished the index’s relationship with local food
prices, so this variable turned out to be not significant in the formula-
tion of expectations.
18
The sign of the coefficient of climatic events was negative in the models.
Since the mei was used as a proxy variable, when it presents negative
values it denotes the presence of the La Niña phenomenon. This phe-
nomenon can generate heavy rains, floods and landslides, especially
in the eastern part of Bolivia, where most of the agricultural produc-
tion is located. Therefore, it can be inferred that when the La Niña
phenomenon occurs, the inflation expectations of economic agents
would increase, although not significantly. This variable’s lack of sig-
nificance is possibly explained by the fact that the effects of climatic
events generally affect food prices for no longer than three months;
prices subsequently decrease as supply normalizes in local markets.
Economic agents thus do not expect there to be a constant rise in prices
in following months.
19
It is worth mentioning that, unlike the rest of the variables, the igae
information is available to the general public with a greater lag time.
In that sense, the output gap entered the model with a lag of two
periods.
−0.2
0.0
0.2
0.4
0.6
0.8
1.0
Jun 2005
Jun 2005 Oct 2005
Oct 2005 Percentage
Percentage
Feb 2006
Feb 2006 Jun 2006
Jun 2006 Oct 2006
Oct 2006 Feb 2007
Feb 2007
Jun 2007 Jun 2007
Oct 2007 Oct 2007
Feb 2008 Feb 2008
Jun 2008 Jun 2008
Oct 2008 Oct 2008
Feb 2009 Feb 2009
Jun 2009 Jun 2009
Oct 2009 Oct 2009
Feb 2010 Feb 2010
Jun 2010 Jun 2010
Oct 2010 Oct 2010
Feb 2011 Feb 2011
Jun 2011 Jun 2011
Figure 6
Although, our main analysis has been done with the bcb survey and,
therefore, with short-term information; there are other sources
where anyone can find information on the expectations of economic
agents. Most of the research papers on this topic consider data from
international private companies that conduct surveys on different
−10
−8
−6
−4
−2
0
2
4
6
8
10
Jun 2005
Jun 2005 Oct 2005
Percentage
Oct 2005
Percentage
Feb 2006
Feb 2006 Jun 2006
Jun 2006 Oct 2006
Oct 2006 Feb 2007
Feb 2007
Jun 2007 Jun 2007
Oct 2007 Oct 2007
Feb 2008 Feb 2008
Jun 2008 Jun 2008
Oct 2008 Oct 2008
Feb 2009 Feb 2009
Jun 2009 Jun 2009
Oct 2009 Oct 2009
Feb 2010 Feb 2010
Jun 2010 Jun 2010
Oct 2010 Oct 2010
Feb 2011 Feb 2011
Jun 2011 Jun 2011
Figure 7
Feb 2015 . Feb 2015
Jun 2015 Jun 2015
Oct 2015 Oct 2015
Feb 2016 Feb 2016
Jun 2016 Jun 2016
Oct 2016 Oct 2016
Feb 2017 Feb 2017
20
The Latin Focus Consensus Forecast report is a monthly publication,
which contains macroeconomic projections from nearly 200 different
sources. It covers approximately 30 macroeconomic indicators per coun-
try for a five-year forecast horizon including economic activity (GDP),
industrial production, business confidence, consumer confidence,
inflation, monetary policy decisions and exchange rate movement.
21
Focus Economics is a company that has information on economic fore-
casts for many key indicators in 127 countries. Its reports draw on many
economic and commodities price forecasts and on economic analysts
around the world.
22
There exist other institutions that provide information about economic
forecast; one of the most famous is Consensus Economics. Nevertheless,
in the case of Bolivia its report has only forecast information for the
current and next calendar year of the variables of interest for the pres-
ent document.
23
There exists forecast information for the current and next calendar
year for a longer period, but, not for the rest of the years.
Percentage
20
18
16
14
12
10
8
6
4
2
0
Jun 2005
Oct 2005
Feb 2006
Jun 2006
Oct 2006
Feb 2007
Jun 2007
Oct 2007
Feb 2008
Jun 2008
Oct 2008
Feb 2009
Jun 2009
Oct 2009
Feb 2010
Jun 2010
Oct 2010
Feb 2011
Jun 2011
Oct 2011
Feb 2012
Jun 2012
Oct 2012
Feb 2013
Jun 2013
Oct 2013
Feb 2014
Jun 2014
Oct 2014
Feb 2015
Jun 2015
Oct 2015
Feb 2016
Jun 2016
Oct 2016
Feb 2017
Jun 2017
BCB expectations FE expectations
24
Although most of the literature defines the medium term as beginning
with the fifth year ahead (see, Carrasco and Ferreiro, 2013; imf, 2016),
this document defines the medium term as beginning with the second
year ahead, like Łyziak and Paloviita, 2016.
25
There exist research papers that have used implicit inflation targets
such as Mumtaz and Theodoridis (2017).
26
Also, this level has been used as reference for the next calendar year’s
projections in the bcb Monetary Policy Report since 2015.
27
The results of Model 6 can be found in Annex 3.
0
1
2
3
4
5
6
7
8
9
Jun 2010
Jun 2005 Percentage
Oct 2005 Oct 2010
Percentage
Feb 2006
Jun 2006 Feb 2011
Oct 2006
Feb 2007 Jun 2011
Jun 2007 Oct 2011
First year
Oct 2007
Feb 2008 Feb 2012
Current year
Jun 2008
Oct 2008 Jun 2012
Feb 2009
Jun 2009 Oct 2012
Oct 2009 Feb 2013
Feb 2010
Jun 2010 Jun 2013
Oct 2010
Oct 2013
Second year
Feb 2011
Jun 2011
Feb 2014
Figure 9
Oct 2011
Feb 2012 Jun 2014
Jun 2012
Oct 2014
Table 3
DEGREE OF ANCHORING OF INFLATION EXPECTATIONS
IN THE SHORT TERM AND MEDIUM TERM
Percentage
1.1
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
Mar 2010
Jun 2010
Sep 2010
Dic 2010
Mar 2011
Jun 2011
Sep 2011
Dic 2011
Mar 2012
Jun 2012
Sep 2012
Dic 2012
Mar 2013
Jun 2013
Sep 2013
Dic 2013
Mar 2014
Jun 2014
Sep 2014
Dic 2014
Mar 2015
Jun 2015
Sep 2015
Dic 2015
Mar 2016
Jun 2016
Sep 2016
Dic 2016
Mar 2017
Jun 2017
1st year 2nd year 3rd year 4th year
banks, including the European Central Bank or the u.s. Federal Re-
serve, do not publish official inflation targets but are able to commu-
nicate the level of their inflation objective to the markets.
Although inflation expectations appear to be well anchored in the
medium term with respect to past inflation, there is a strand of lit-
erature that postulates that long-term (medium-term) expectations
should not respond to changes in short-term inflation expectations
either ( Jochmann, Koop and Potter, 2010; Łyziak and Paloviita,
2016). In that sense, we additionally create a model to study if there
is a relationship between medium-term and short-term inflation ex-
pectations using the information from Focus Economics.
If medium-term inflation expectations are well anchored, they
should not respond to changes from short-term inflation expecta-
tions. In this case, following the work of Strohsal, Melnick and Nautz
where:
α1 + α2 + α 3 = 1
Table 4
DEGREE OF ANCHORING OF INFLATION EXPECTATIONS
IN THE MEDIUM TERM, MODEL 10
28
As a reference of medium-term we choose the inflation expectations
for the fourth year of Focus Economics.
29
As a reference of short-term we choose the inflation expectations for the
first year of Focus Economics, in order to work with the same survey
sample.
5. SOME CONSIDERATIONS
REGARDING THE RESULTS
100
90
80
70
60
50
40
30
20
10
0
Feb 1998
Oct 1998
Jun 1999
Feb 2000
Oct 2000
Jun 2001
Feb 2002
Oct 2002
Jun 2003
Feb 2004
Oct 2004
Jun 2005
Feb 2006
Oct 2006
Jun 2007
Feb 2008
Oct 2008
Jun 2009
Feb 2010
Oct 2010
Jun 2011
Feb 2012
Oct 2012
Jun 2013
Feb 2014
Oct 2014
Jun 2015
Feb 2016
Oct 2016
Jun 2017
Source: Central Bank of Bolivia.
During the 1990s and the first five years of the 2000s, almost all of
the loans and deposits in the financial system were denominated
in u.s. dollars because people in Bolivia had greater confidence in the
dollar to carry out their daily transactions. This situation can be at-
tributed to the constant depreciations during this period, which
led to a loss of the value of the local currency. In 2006, when the Bo-
livian appreciated, the degree of financial dollarization in Bolivia
began to decrease. This aspect, with other measures applied by the
local authorities, allowed the de-dollarization process to accelerate.
This in turn created a more favorable environment for monetary pol-
icy and a greater role for the bcb in local economic activity. While
97 percent of loans were made in dollars at the beginning of 1998,
by mid-2017 this figure had fallen to 2.7 percent (Figure 12). In the
same period, deposits in dollars declined from 92.7 percent to 15.6
percent. These developments apparently helped to create a more
predictable environment for economic agents.
Oct 2009
Feb 2010 Feb 2010
Jun 2010 Jun 2010
Oct 2010 Oct 2010
Feb 2011 Feb 2011
Jun 2011 Jun 2011
Oct 2011 Oct 2011
Figure A.1
Feb 2012
Jun 2012
Oct 2012
Feb 2013
Jun 2013
Oct 2013
Feb 2014
Jun 2014
Oct 2014
Feb 2015
Jun 2015
Oct 2015
Feb 2016
Jun 2016
Oct 2016
Feb 2017
UPDATED BCB PROJECTION (INFLATION BY THE END OF CURRENT YEAR)
Jun 2017
−0.5
0.0
0.5
1.0
1.5
2.0
−0.8
−0.6
−0.4
−0.2
0.0
0.2
0.4
0.6
0.8
1.0
1.2
1.4
Jun 2005 Jun 2005
Oct 2005 Oct 2005
Feb 2006 Feb 2006
Annex 2
Feb 2014
Jun 2014 Jun 2014
EVOLUTION OF COEFFICIENTS IN MODEL 6
Abstract
Our objective in this paper is to build expectations anchoring indexes for in-
flation in Brazil that are fundamentally driven by the monetary authority’s
capacity to anchor long-term inflation expectations vis-à-vis short-run infla-
tion expectations. The expectations anchoring indexes are generated from a
Kalman filter, based on a state-space model that also takes into account fiscal
policy dynamics. The model’s signals are constructed using inflation expecta-
tions from the Focus survey of professional forecasters, conducted by the Banco
Central do Brasil, and from the swap and federal government bond markets,
which convey daily information of long-term inflation expectations. Although
varying across specifications, the expectations anchoring indexes that we pro-
pose tend to display a downward trajectory, more clearly in 2009, and show a
recovery starting in 2016 until the end of the sample (mid-2017).
Keywords: credibility index, inflation expectation, inflation anchoring,
Kalman filter, Banco Central do Brasil.
jel classification: E50, E52, E58.
F. Nascimento de Oliveira <[email protected]>, Research Depart-
ment, Banco Central do Brasil and ibmec/rj, and W. P. Gaglianone <wagner.gaglia-
[email protected]>, Research Department, Banco Central do Brasil. This research
was undertaken within the framework of cemla’s Joint Research Program 2017
coordinated by the Banco de la República, Colombia. The authors are especially grate-
ful for the helpful comments and suggestions given by Olivier Coibion. The authors
benefited from comments made by seminar participants in the workshop on inflation
expectations, their measurement and degree of anchoring at cemla headquarters
(Mexico City, September 14-15, 2017). We also thank Carlos Viana de Carvalho, André
Minella and Bernardus Doornik for comments made at the Second Network of Eco-
nomic Research of the Banco Central do Brasil. The views expressed in the papers
are those of the authors and do not necessarily reflect those of the Banco Central
do Brasil.
177
1. INTRODUCTION
W
ell-anchored inflation expectations are fundamental for the
conduct of monetary policy. Properly anchoring inflation ex-
pectations requires the central bank to be regarded as cred-
ible, that is, economic agents should be confident that the central
bank will react to the various shocks that affect the economy to main-
tain price stability.
Cukierman and Meltzer (1986) stressed that the future objectives
of central banks depend on inflation expectations. In this sense,
a credible commitment to an explicit inflation objective helps to an-
chor inflation expectations to the desired level. This anchoring
contributes to delivering price stability, which is the main objective
of central banks.
In turn, Blinder (2000) sent questionnaires to 127 heads of cen-
tral banks around the world asking their opinion on the importance
of central bank credibility. The answers showed clearly that credibil-
ity matters in practice. A credible central bank is one that can make
a believable commitment to low inflation policy and has complete
dedication to price stability. This will make disinflation less costly
and decrease the sacrifice ratio.
Nonetheless, building credibility is costly and takes repeated
successes to establish. Moreover, credibility evolves in asymmetric
fashion and can be lost rapidly, depending on the perception by eco-
nomic agents that the central bank is able (or not) to achieve its ob-
jectives. As famously put by Benjamin Franklin: “It takes many good
deeds to build a good reputation, and only one bad one to lose it.”1
Central banks have imperfect control over inflation in the short
run. As Gomme (2006) remarked, current inflation provides a noisy
signal of a central bank’s long-term intentions, and therefore of its
type. According to the author, a central bank is credible when the pub-
lic assigns a high probability of low inflation-type to the central bank.
In this context, a central bank will lose credibility when this prob-
ability decreases. The credibility of central banks is very much con-
cerned with people’s beliefs about what the central bank will do in
the future.
1
See Isaacson (2004).
2
The international literature on credibility indexes of central banks
is vast. They are many theoretical as well as empirical papers on the
subject. See, for example, Gomme (2006), Svenson (1993), Clarida
and Waldman (2007), Ceccheti and Krause (2002), Kaseeream (2012)
and Bordo and Siklos (2015).
3
Other papers also build credibility indexes for the Banco Central
do Brasil focusing on deviations of short-term inflation expectations
from inflation target, such as Teles and Nemoto (2005), Sicsú (2002),
Nahon and Meurer (2005), and Lowenkron and Garcia (2007).
4
Note that ci-m decreases substantially during the subprime crisis, which
like Lula’s election is also exogenous to bcb. At the end of the period,
ci-m shows a steep credibility recovery that also seems counterfactual.
12
10
0
Jan 2002
Mar 2003
May 2004
Jul 2005
Sep 2006
Nov 2007
Jan 2009
Mar 2010
May 2011
Jul 2012
Sep 2013
Nov 2014
Jan 2016
Mar 2017
Consensus inflation expectations Inflation target
(cross-sectional average)
Note: Average inflation expectations (Focus survey) with forecast horizon of one year.
Inflation targets and tolerance bands from <http://www.bcb.gov.br/pec/metas/In-
flationTargetingTable.pdf>.
Figure 2
CREDIBILITY INDEXES FROM THE LITERATURE
1.0
0.8
0.6
0.4
0.2
0.0
Jan 2002
Mar 2003
May 2004
Jul 2005
Sep 2006
Nov 2007
Jan 2009
Mar 2010
May 2011
Jul 2012
Sep 2013
Nov 2014
Jan 2016
Mar 2017
Note: CI-CK means Cecchetti and Krause (2002), CI-M denotes Medoça (2004) and
CI-MS m e a n s M e n d o ç a a n d S o u z a ( 2 0 0 9 ) . I n fl a t i o n e x p e c t a t i o n s a re t h e
survey-based cross-sectional average expectations with fixed horizon of one year.
200
100
80
60
40
20
0 Jan 2011
April 2015
Jan 2002
Jun 2004
Aug 2006
Oct 2008
Mar 2013
Jun 2017
Source: Banco Central do Brasil and authors’ calculations.
5
There are other papers in the literature that build credibility indexes
for the bcb taking different approaches from those that look at short-
term deviations of inflation expectations from the target. This is the
case of Garcia and Guillén (2011), Leal et al. (2012), Issler and Santos
(2017), and Val et al. (2017).
6
Bernanke (2007) describes inflation anchoring in the following man-
ner: “…“anchored” to mean relatively insensitive to incoming data. So,
for example, if the public experiences a spell of inflation higher than
their long-run expectation, but their long-run expectation of inflation
changes little as a result, then inflation expectations are well anchored.
If, on the other hand, the public reacts to a short period of higher-
than-expected inflation by marking up their long-run expectation
considerably, then expectations are poorly anchored”.
7
For other empirical papers with definitions of credibility, see Davis
(2012), Levieuge et al . (2015) and Dimitris et al . (2016). For theoretical
papers with definitions of central bank credibility, see Barro and Gor-
don (1983), Walsh (1995) and Blackburn and Christensen (1989).
8
See Natoli and Sigalotti (2017).
2. DATA
We have survey and market data. In the former case, we have data
from January 2002 to June 2017. In the latter case, we have data from
April 2005 to June 2017.
9
See Gaglianone (2017) for a recent survey of applied research on infla-
tion expectations in Brazil.
14
Mean forecast (1 year)
Mean forecast (2 year)
Mean forecast (3 year)
12 Mean forecast (4 year)
Mean forecast (5 year)
10
2
Jan 2002 Jun 2004 Aug 2006 Oct 2008 Jan 2011 Mar 2013 Apr 2015 Jun 2017
3.5
Std. deviation (1 year)
Std. deviation (2 year)
3.0
Std. deviation (3 year)
Std. deviation (4 year)
2.5 Std. deviation (5 year)
2.0
1.5
1.0
0.5
0.0
Jan 2002 Jun 2004 Aug 2006 Oct 2008 Jan 2011 Mar 2013 Apr 2015 Jun 2017
14
Median forecast (1 year)
Median forecast (2 year)
Median forecast (3 year)
12 Median forecast (4 year)
Median forecast (5 year)
10
2
Jan 2002 Jun 2004 Aug 2006 Oct 2008 Jan 2011 Mar 2013 Apr 2015 Jun 2017
4.5
Inter-quartile range (1 year)
4.0 Inter-quartile range (2 year)
Inter-quartile range (3 year)
3.5 Inter-quartile range (4 year)
Inter-quartile range (5 year)
3.0
2.5
2.0
1.5
1.0
0.5
0.0
Jan 2002 Jun 2004 Aug 2006 Oct 2008 Jan 2011 Mar 2013 Apr 2015 Jun 2017
10
Nowadays, the bcb releases on the internet the micro data of the Focus
survey of expectations, in a panel data with fake IDs (i.e., the identity
of the survey participants is preserved and the disclosed database only
contains anonymous participants). For more details, see the website:
http://dadosabertos.bcb.gov.br/dataset/expectativas-mercado/
resource/23f6c983-f9bd-48f8-a889-72def3ae17c8
12
Mean forecast (1 year)
11 Mean forecast (4 year)
10
3
Jan 2002 Jun 2004 Aug 2006 Oct 2008 Jan 2011 Mar 2013 Apr 2015 Jun 2017
2.5
Std. deviation (1 year)
Std. deviation (4 year)
2.0
1.5
1.0
0.5
0.0
Jan 2002 Jun 2004 Aug 2006 Oct 2008 Jan 2011 Mar 2013 Apr 2015 Jun 2017
12
Median forecast (1 year)
11 Median forecast (4 year)
10
3
Jan 2002 Jun 2004 Aug 2006 Oct 2008 Jan 2011 Mar 2013 Apr 2015 Jun 2017
4.0
Inter-quartile range (1 year)
3.5 Inter-quartile range (4 year)
3.0
2.5
2.0
1.5
1.0
0.5
0.0
Jan 2002 Jun 2004 Aug 2006 Oct 2008 Jan 2011 Mar 2013 Apr 2015 Jun 2017
10
9
8
7
6
5
4
3
2
Apr 2005
Aug 2006
Dec 2007
Apr 2009
Aug 2010
Dec 2011
Apr 2013
Aug 2014
Dec 2015
Apr 2017
swap 1y swap 4y bonds 1y bonds 4y
and tolerance bands had been set up to June of year t for the calendar
year t+2. Nowadays, the new target is announced up to June of year t
for the calendar year t+3.11 Since many signals depend on the infla-
tion target, and since our longest forecast horizon is four years, we as-
sume that the inflation target four years ahead is equal to the target
set for the calendar year t+2 (or t+3, whenever available).
In the case of market data, we have publicly available information
on federal government bonds and swaps of fixed interest rate against
inflation and a coupon from April 2005 to June 2017. The former
are obtained from Anbima (Brazilian Financial and Capital Mar-
ket Association) and the latter are registered by b3 (a Brazilian com-
pany that operates securities, commodities and futures exchange,
among others, previously known as bm&fbovespa). Federal govern-
ment bonds are nominal bonds (ltns) and inflation-indexed bonds
11
See <https://www.bcb.gov.br/pec/metas/InflationTargetingTable.
pdf>.
12
fomc means the Federal Open Market Committee of the U.S. Federal
Reserve.
13
We have yields for fixed-interest bonds with maturities of one, three
and ten years. We interpolate linearly the three- and ten-year yields
to get the four-year yields that we used to construct bei s for the swap
market.
−1
−2
−3
Jan 2002
Jul 2003
Jan 2005
Jul 2006
Jan 2007
Jul 2007
Jan 2008
Jul 2009
Jan 2010
Jul 2010
Jan 2017
Source: Banco Central do Brasil, Focus survey, cross-section average expectations.
3. EMPIRICAL ANALYSIS
6
5
4
3
2
1
0
−1
−2
−3
−4
Sep 2002
Jan 2004
May 2005
Sep 2006
Jan 2008
May 2009
Sep 2010
Jan 2012
May 2013
Sep 2014
Jan 2016
May 2017
s1 s2 s3 s4 s5 s6 s7 s8
6
4
2
0
−2
−4
−6
Sep 2002
Jan 2004
May 2005
Sep 2006
Jan 2008
May 2009
Sep 2010
Jan 2012
May 2013
Sep 2014
Jan 2016
May 2017
14
In other words, for a given sample, a weight equal to 1 is attached
to the most recent observation. After a half-life period (e.g., 1 year
=252 workdays), the weight exponentially decays to 0.5.
15
See Massey (2012).
16
In this paper, we employ the idea behind the Quandt-Andrews test,
in which a single Chow (1960) breakpoint test is performed for every
observation between two dates. The test statistics from those Chow tests
are used to build dummy variables representing the different regimes
between breakpoints.
17
Factor analysis (fa) and principal component analysis (pca) are similar
statistical techniques in the sense that both generate linear combina-
tions of the original series. However, pca is used to retain the maximum
amount of information from data in terms of total variation, whereas
fa accounts for common variance. Thus, fa is often employed to build
factors (latent variables), while pca is often used in data reduction
frameworks. See Johnson and Wichern (1992) for further details.
1 sit = Λi Ft + εit
18
We use the parsimonious number of two factors since they account
for more than half of the fraction of total variance of the set of signals.
Nonetheless, there are many alternative factor selection tools avail-
able in the literature, such as the ones proposed by Bai and Ng (2002)
or Alessi, Barigozzi and Capasso (2010).
19
These figures are computed using the eigenvalues obtained in the
solution of each factor’s linear combination, as explained in Jolliffe
(2002).
6
5
4
3
2
1
0
−1
−2
−3
−4
Sep 2002
Jan 2004
May 2005
Sep 2006
Jan 2008
May 2009
Sep 2010
Jan 2012
May 2013
Sep 2014
Jan 2016
May 2017
s1 s2 s3 s4 s5 s6 s7 s8
6
4
2
0
−2
−4
−6
Sep 2002
Jan 2004
May 2005
Sep 2006
Jan 2008
May 2009
Sep 2010
Jan 2012
May 2013
Sep 2014
Jan 2016
May 2017
s9
Sep 2002 Sep 2002
s1
Jan 2004 Jan 2004
s10
May 2005 May 2005
s2
s11
Sep 2006 Sep 2006
s3
Jan 2008 Jan 2008
s12
May 2009 May 2009
s4
s13
Figure 10
s5
s14
s6
May 2013 May 2013
s14
Rolling window weights, window of three years
s7
s15
Jan 2016 Jan 2016
s8
May 2017 May 2017
s16
Figure 11
MARKET SIGNALS
Exponential smoothing, half-life of one year
−2
−4
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
sm1 sm2 sm3 sm4 sm5 sm6 sm7
−2
−4
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
−2
−4
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
sm1 sm2 sm3 sm4 sm5 sm6 sm7
−2
−4
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
−2
−4
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
sm1 sm2 sm3 sm4 sm5 sm6 sm7
−2
−4
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
3.5
2.5
1.5
0.5
−0.5
−1.5
−2.5
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
F1,t F2,t Ft (combined factor)
Table 2
SIGNALS CONSTRUCTED FROM BREAKEVEN INFLATION (bei) MARKET DATA
Signals Description
sm1 slope from recursive ols regression, bei four years against bei one year (swaps)
sm2 recursive correlation between bei four years and one year (swaps)
sm3 Nautz and Strohsal (2015), fx-rate slope from ols (bei 4y swaps, macro shocks)
sm4 Natoli and Sigalotti (2017), slope from logit regression, ∆ bei swaps (1y, 4y)
sm5 (bei 4y swaps-inflation target)
sm6 (bei 4y swaps-inflation target)2
sm7 Demertzis et al . (2012), time-varying var, bei swaps (1y, 4y)
sm8 slope from recursive ols regression, bei four years against bei one year (bonds)
sm9 recursive correlation between bei four years and one year (bonds)
sm10 Nautz and Strohsal (2015), fx-rate slope from ols (bei 4y bonds, macro shocks)
sm11 Natoli and Sigalotti (2017), slope from logit regression, ∆ bei bonds (1y, 4y)(bei
4y bonds-inflation target)
sm12 (bei 4y bonds-inflation target)
sm13 (bei 4y bonds-inflation target)2
sm14 Demertzis et al . (2012), time-varying var bei bonds (1y, 4y)
2 xt = Axt −1 + Bt ,
3 yt = Cxt + Dvt ,
θ1 0 0 1 0 0 0 θ3 θ4 θ8 0 0
C = θ θ θ
4 A = 0 θ2 0 ; B = 0 1 0 ; 5 6 7 and D = 0 0 0
0 0 1 0 0 0 0 0 1 0 0 0
Note that the state ot = 1t plays the role of the intercept and states
ct = θ1ct −1 + ε1,t and ft = θ2 ft −1 + ε 2,t are ar(1) processes with zero mean.
On the other hand, the observable fiscal expectation (zt ) is driven
by the fiscal state ( ft ) plus an intercept and the idiosyncratic shock
v1,t . The long-term anchoring factor Ft is decomposed into two states,
ct and ft , which are designed to capture, respectively, the dynamics
of monetary and fiscal policies.
5 zt = θ3 ft + θ 4 + θ8v1,t ,
6 Ft = θ5ct + θ6 ft + θ7
S3 S9 S12 S13 S14 S15 S17 SM3 SM4 SM7 SM8 SM9 SM12 SM14
S3 1.00
S9 0.30 1.00
S12 –0.56 –0.62 1.00
S13 0.35 0.26 –0.22 1.00
S14 0.16 0.00 0.01 0.08 1.00
S15 0.00 0.26 –0.31 –0.11 –0.21 1.00
S17 0.48 0.69 –0.61 0.15 0.00 0.46 1.00
SM3 –0.07 –0.04 –0.22 0.03 –0.33 0.51 –0.15 1.00
Note: Only signals with pairwise absolute correlation below 0.7 are selected for the ES2y baseline case.
Figure 15
MONETARY POLICY CREDIBILITY STATE (ct), FISCAL POLICY STATE (ft),
EXPECTATION OF PRIMARY FISCAL BALANCE (zt) AND LONG-TERM
AND ANCHORING FACTOR (Ft)
5 0.0
4 −0.2
3 −0.4
2 −0.6
1 −0.8
0 −1.0
−1 −1.2
−2 −1.4
−3 −1.6
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
zt (left scale) ct (left scale)
Ft (left scale) ft (right scale)
20
This assumption, in principle, could be relaxed by including an error
term with zero mean and low variance (set as initial condition in the
Kalman filter estimation).
21
We limit to 1,000 the number of interactions of the maximum likeli-
hood estimations. In all estimations presented in this paper, maximum
likelihood converged before reaching the limit of interactions. For the
Kalman filter, we considered the expectation of initial state vector equal
to zero.
22
To guarantee the EAI to be inside the [0;1] interval.
23
For instance, the single fiscal expectations series, coupled with an au-
toregressive structure assumed for the fiscal state ft, might not properly
capture the core standpoint of fiscal policy. Alternative approaches
to tackle this issue could consider, for instance, a state-space model con-
taining an entire block of equations (instead of a single one) to model
the fiscal policy in a disaggregate way. On the other hand, the set
of observable variables could include data from credit default swaps
and/or real interest rates (e.g., long-maturity forwards) or even risk
premium estimates using satellite term-structure models.
24
The advantage of our approach is that the estimated risk premium
is “model-free” in the sense that it is not grounded on a specific theoreti-
cal model, but instead is solely based on survey data at the micro level.
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
Baseline (RW) Baseline (ES1y) Baseline (ES2y)
Notes: ES1y and ES2y denote the exponenially smoothed weights with half-life of one
year and 2 years, respectively, and rw means rolling window weights (window of three
years). Only signals with pairwise absolute correlation below 0.7 are selected for the
baseline case. The following signals are selected: S3, S9, S12, S13, S14, S15, S17, SM3,
SM4, SM7, SM8, SM9, SM12 and SM14.
Table 5
KALMAN FILTER ESTIMATION OF THE EXPECTATIONS
ANCHORING INDEX (BASELINE ES2Y)
4. CONCLUSION
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
Market (RW) Market (ES1y) Market (ES2y)
Notes: ES1y and ES2y denote the exponenially smoothed weights with half-life of one
year and two years, respectively, and means rolling window weights (window of
three years).
Figure 18
CREDIBILITY INDEX
Survey signals
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
Sep 2002
Jan 2004
May 2005
Sep 2006
Jan 2008
May 2009
Sep 2010
Jan 2012
May 2013
Sep 2014
Jan 2016
May 2017
Notes: ES1y and ES2y denote the exponentially smoothed weights with half-life of one
year and two years, respectively, and means rolling window weights (window of three
years).
10
−2
Apr 2005
Aug 2006
Dec 2007
Apr 2009
Aug 2010
Dec 2011
Apr 2013
Aug 2014
Dec 2015
Apr 2017
BEI swaps 1y BEI swaps 4y
BEI swaps 1y (no risk premium) BEI swaps 4y (no risk premium)
10
−2
Apr 2005
Aug 2006
Dec 2007
Apr 2009
Aug 2010
Dec 2011
Apr 2013
Aug 2014
Dec 2015
Apr 2017
BEI swaps 1y (no risk premium) BEI swaps 4y (no risk premium)
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
Market (Es2y) Market (ES2y), no risk premium
Figure 21
EXPECTATIONS ANCHORING USING A DIFFERENT METHOD
TO CONSTRUCT THE COMMON FACTOR Ft
1.0
0.9
0.8
0.7
0.6
0.5
0.4
0.3
0.2
0.1
0.0
Sep 2005
Nov 2006
Jan 2008
Mar 2009
May 2010
Jul 2011
Sep 2012
Nov 2013
Jan 2015
Mar 2016
May 2017
Note: The single-factor comes from the “minimum average partial” criterion for
selecting the number of factors.
Abstract
During the year 2016, the Banco Central de la República Argentina has be-
gun to announce inflation targets. In this context, providing the authorities
of good estimates of relevant macroeconomic variables turns out to be crucial
to make the pertinent corrections in order to reach the desired policy goals. This
paper develops a group of models to forecast inflation for Argentina, which
includes autoregressive models, and different scale Bayesian var s (bvar),
and compares their relative accuracy. The results show that the bvar model
can improve the forecast ability of the univariate autoregressive benchmark’s
model of inflation. The Giacomini-White test indicates that a bvar performs
better than the benchmark in all forecast horizons. Statistical differences
between the two bvar model specifications (small and large-scale) are not
found. However, looking at the rmses, one can see that the larger model
seems to perform better for larger forecast horizons.
Keywords: Bayesian vector autoregressive, forecasting, prior specification,
marginal likelihood, small-scale and large-scale models.
jel classification: C11, C13, C33, C53.
223
1. INTRODUCTION
S
everal long-term nominal commitments such as labor contracts,
mortgages and other debt are widespread features of modern
economies. Forecasting how the general price level will evolve
over the life of a commitment is an essential part of private sector
decision-making.
The existence of long-term nominal obligations is also among
the primary reasons economists generally believe that monetary
policy is not neutral, at least over moderate horizons.
Central banks aim is to keep inflation stable, and perhaps also
to keep output near an efficient level. With these objectives, the New
Keynesian model makes explicit that optimal policy will depend
on optimal forecasts (e.g., Svensson, 2005), and further, that policy
will be most effective when it is well understood by the public.
Under inflation targeting the central banks generally released fore-
casts in quarterly Inflation Reports in a way to be more transparent
in their actions. The costs and benefits of transparency are widely
debated, but the need for a central bank to be concerned with infla-
tion forecasting is broadly agreed. In short, inflation forecasting is of
foremost importance to households, businesses, and policymakers.
During the year 2016, the Banco Central de la República Argen-
tina (bcra) has begun to announce inflation targets. In this context,
providing the authorities of good estimates of relevant macroeco-
nomic variables turns out to be crucial to make the pertinent cor-
rections in order to reach the desired policy goals.
A standard tool in macroeconomics that is widely employed in fore-
casting is vector autoregressive (var) analysis. var s are flexible time
series models that can capture complex dynamic relationships among
macroeconomic aggregates. However, their dense parameterization
often leads to unstable inference and inaccurate out-of-sample fore-
casts, particularly for models with many variables, due to the estima-
tion uncertainty of the parameters.
Litterman (1980) and Doan, Litterman, and Sims (1984) have pro-
posed to combine the likelihood function (the data) with some infor-
mative prior distributions (the researcher’s belief about the values
of coefficients) to improve the forecasting performance of var mod-
els, introducing a Bayesian approach into var modeling.
1 yt = c + B1 yt −1 + + B p yt − p + ε t ,
2 β Ω ∼ N (β 0 ,Ω −1ξ )
where the vector β 0 is the prior mean (whose elements will represent
the coefficient in Equation 1, the matrix Ω is the known variance
of the prior, and ξ is a scalar parameter controlling the tightness
of the prior information. Even though Ω could have many shapes,
gamma and Wishart distributions are frequently used in the litera-
ture, since they ensure a normally distributed posterior.1
The conditional posterior of β can be obtained by multiplying
the prior by the likelihood function. The posterior takes the form
3 β (
, y ∼ N βˆ (ξ ),Vˆ(ξ ) , )
where
4 ( )
βˆ (ξ ) ≡ vec βˆ (ξ ) ,
and
1
If the posterior distributions are in the same family as the prior prob-
ability distribution, the prior and posterior are then called conjugate
distributions.
7 Bˆ (ξ ) ≡ ˆ[ −1
0 β 0 +(
−1
⊗ x ) y]
and
ˆ =[ −1 −1 −1 .
0 + Σ ⨂ x ʹx ]
8
9 Bˆ (ξ ) ≡ ˆ[ −1
0 β 0 ]+
ˆ[Σ −1⨂ x ʹx (x ʹx )−1 x ʹ y]
10 Bˆ (ξ ) ≡ ˆ[ −1
0 β 0 ]+
ˆ[Σ −1 x ʹx (x ʹx )β ]
ols
2
The ols estimators of a var coincide exactly with the ml estimators
conditional on the initial values.
11 p(γ y) ≈ p ( y γ )p (γ ),
12 p (γ y) = ∫ p ( y θ , γ )p (θ γ )d θ .
13 p (θ ) = ∫ p (θ , γ )π (γ )d γ .
14 p (θ , γ y) = p ( y θ , γ )p (θ , γ )p (γ ).
3
The mean of the predictive density is considered.
15 ∆Lt +τ (θˆ ) = µ + ε t ,
16 ∆Lt +τ (θˆ ) = β ʹ X t + ε t .
4
See chapter 17 of the book by Hashimzade and Thornton (2013) for a
detailed discussion about this test.
5
From December 2006 to October 2015, the index by the indec pre-
sented severe discrepancies with provincial and private price index,
and hence was discarded for that period.
3. DATA
4. RESULTS
6
In the Annex, we show the posterior estimation of the whole sample
to see the effect of this. We controlled the change in the mean due the
transition to an inflation targeting regime and indeed obtained a sig-
nificant coefficient in both models.
1 λ 2
17 σ ij2 = 2 λ13
σ j L
σ 2 = ( λ1λ4 )
2
18
The results for the hyperparameters and prior means of the small
and the big scale model are shown in Table 1. All the hyperparameters
are equal for both type of models except for the hyperparameter λ 1.
The characteristics of our hyperparameters after the optimiza-
tion procedure is as follow:
Table 1
LIST OF HYPERPARAMETER VALUES
Table 3
GIACOMINI-WHITE TEST
One-step-
0.03 0.01 0.29
ahead
Three-steps-
0.00 0.00 0.49
ahead
Six-steps-
0.09 0.05 0.41
ahead
5. CONCLUSIONS
Table A.1
LIST OF ENDOGENOUS VARIABLES
Multilateral nominal
8 bcra exchange rate
log – Unit-root
Imports of intermediate
10 indec goods
log sa Unit-root
Table B.1
SMALL bvar CHARACTERISTICS
Prior: Normal-Wishart
Autoregressive coefficient: 0
Lag decay: 2
Table B.2
SMALL bvar INFLATION EQUATION COEFFICIENT VALUES
Median sd lb ub
inf(–1) 0.468 0.066 0.338 0.598
I(–1) 0.901 0.640 –0.356 2.157
Y(–1) 2.631 3.500 –4.237 9.499
Constant 0.280 0.071 0.140 0.420
d112016 –0.197 0.144 –0.479 0.086
Number of lags 1
Prior Normal-Wishart
Autoregressive
0
coefficient
Lag decay 2
Exogenous variable
1
tightness
Median sd lb ub
inf(–1) 0.145 0.045 0.057 0.234
Luis Libonatti
Abstract
This paper presents a forecasting exercise that assesses the predictive poten-
tial of a daily price index based on online prices. Prices are compiled using
web scraping services provided by the private company PriceStats in coopera-
tion with a finance research corporation, State Street Global Markets. This
online price index is tested as a predictor of the monthly core inflation rate
in Argentina, known as “resto IPCBA” and published by the Statistics Office
of the City of Buenos Aires. Mixed frequency regression models offer a conve-
nient arrangement to accommodate variables sampled at different frequen-
cies and hence many specifications are evaluated. Different classes of these
models are found to produce a slight boost in out-of-sample predictive perfor-
mance at immediate horizons when compared to benchmark naïve models
and estimators. Additionally, an analysis of intra-period forecasts, reveals
a slight trend towards increased forecast accuracy as the daily variable ap-
proaches one full month for certain horizons.
Keywords: MIDAS, distributed lags, core inflation, forecasting.
JEL Classification: C22, C53, E37.
The author is affiliated with the Central Bank of Argentina. E-mail: luis.libonatti@
bcra.gob.ar. This research was undertaken within the framework of cemla’s Joint
Research Program 2017 coordinated by the Central Bank of Colombia. The author
gratefully acknowledges counseling and technical advisory provided by the Financial
Stability and Development (fsd) Group of the Inter-American Development Bank
in the process of writing this document. The opinions expressed in this publication
are those of the author and do not reflect the views of cemla, the fsd group,
the InterAmerican Development Bank or the Central Bank of Argentina.
245
1. INTRODUCTION
F
orecasting inf lation has become increasingly important
in Argentina as it is essential for economic agents to adjust
wages and prices—particularly in recent years—in a context
of high and volatile inflation. Having timely updates about the future
trajectory of the inflation rate is essential for conducting monetary
policy, specially, since the Central Bank is transitioning towards
an inflation targeting regime. Recent developments in the use of
“big data” have greatly facilitated tracking macroeconomic vari-
ables in real-time. A remarkable example is the construction of on-
line price indexes that are sampled daily, rather than monthly, as it
is standard for traditional price indexes from statistical offices.
The question naturally arises of whether this information can help
predict the future trajectory of traditional consumer price indexes.
Ghysels et al. (2004) introduced a regression framework that allows
for the exploitation of time series sampled at different frequencies,
known in the literature as Mixed Data Sampling (midas) regression
models. The methodology reduces to fitting a regression model to a
low-frequency variable using high-frequency data as regressors. As it
will be shown later, this technique closely resembles distributed
lag models. This paper employs this methodology to assess whether
the combination of price series sampled at different frequencies
is an effective tool for improving forecast accuracy compared to na-
ïve models, using the online price index constructed by PriceStats
in cooperation with State Street Global Markets.
The rest of the paper is organized as follows. In the next section,
a brief introduction to midas models is presented. In the third sec-
tion, existing theoretical research on midas regressions as well
as some applications in forecasting inflation are briefly reviewed.
In the fourth section, the forecasting exercise is described, and the
results are discussed. And finally, the fifth section concludes.
246
equal weights (flat aggregation).1 Ghysels et al. (2004) suggested com-
bining yt ,a low frequency process, and , a high frequency process
that is observed a discrete and fixed number of times m each time
a new value of yt is observed, in a plain regression equation,
( m −1)
2.1 yt = ∑ θ j xt − j m + ut ,
( j = 0)
or more compactly,
2.2
1
In fact, this can be considered a special case of a midas regression.
2
This equation may also include constants, trends, seasonal terms or other
low frequency explanatory variables.
2.4
3
Foroni et al. (2015) present a detailed assessment of this strategy.
4
See for example the book by Judge et al. (1985).
248
2.5
2.6
with and
Parameterization as in equation (2.5) has proved to be quite pop-
ular and has become the standard among researchers, particularly
when q=2.
The introduction of constrained coefficients has many far-reach-
ing implications. The model turns nonlinear and lacks a closed form
solution. It is necessary to resort to nonlinear least squares and ap-
proximate the solution by numerical optimization routines. Addi-
tionally, the constraints are highly likely to introduce a bias in each
However, based on Monte Carlo simulations, when the sample
size is small relative to the number of parameters, Ghysels et al. (2016)
argue that both, parameter estimation precision and out-of-sample
forecast accuracy, gained by the increase in degrees of freedom,
far offset the effects of the bias generated by misspecified constraints.
midas models are generally intended as a direct forecasting tool
since this could prove to be more robust against misspecification
(Marcellino et al., 2006). This implies that estimation additionally
depends on the time displacement of the variables, and the
forecast horizon, 5 The direct strategy requires estimation
of as many models as per pair (d, h) is required. If T Y is the time in-
dex of latest yt available for estimation, and TX is the time index of the
latest available for both estimation and forecasting, then d can
be defined as Setting
a forecast can be computed with,
2.7
5
How many periods into the future it is necessary to forecast.
2.8
2.9
or equivalently,
2.10
2.11
θ 0, 0
x L x1−( m−1)/ m L x1− pX L x1− pX −( m−1)/ m M
y1 1 u1
y x2 L x2−( m−1)/ m L x2− pX L x2− pX −( m−1)/ m θ0,m−1 u1
2 M
M = M O M O M O M + M .
yT −1 xT −1 L xT −1−( m−1)/ m L xT − pX −1 L xT − pX −1−( m−1)/ m θ pX ,0 uT −1
yT
x L xT −( m−1)/ m L xT − pX L xT − pX −( m−1)/ m M uT
T
θ pX ,m−1
250
If different weight functions for each θ r in equation (2.9), then
the multiplicative or aggregates-based midas model is obtained (Ghy-
sels et al., 2016). On the contrary, employing a single weight function
for all m × L X coefficients vectors θ r is also possible. The first meth-
od allows for greater flexibility but at the cost of more parameters
to estimate, so this possibility will not be considered, as this may not
be convenient for a very short sample size.
Other possible extensions include constructing high frequency
factors (Marcellino and Schumacher, 2010), incorporating cointegra-
tion relations (Miller, 2013), integrating Markov switching (Guérin
and Marcellino, 2013), estimating multivariate models (Ghysels
et al., 2007), using infinite polynomials (Ghysels et al., 2007) or add-
ing low frequency autoregressive augmentations (Ghysels et al.,
2007; Clements and Galvão, 2008; Duarte, 2014), for example. Fo-
roni and Marcellino (2013) provide a comprehensive survey of pos-
sible extensions in a recent survey about mixed frequency models.
3. LITERATURE REVIEW
Clements and Galvão (2008) were among the first to study applica-
tions of midas regressions to macroeconomic variables. In their pa-
per, they forecast u.s. real quarterly output growth in combination
with three different monthly variables: i) industrial production, ii)
employment growth, and iii) capacity utilization. They find a slight
increase in out-of-sample forecast accuracy with both vintage and re-
vised data compared to two benchmarks models, an autoregression
and an adl model in particular, for short-term horizons. They also
derive and assess a model with autoregressive dynamics introduced
as a common factor shared by the low and the high-frequency lag poly-
nomials. Based on comments by Ghysels et al. (2007), they argue that
including an autoregressive term in a standard midas model, as in
the next equation,
3.1 yt = φ yt −1 + λW ( L1/ m ; γ ) xt + ut ,
3.3 (1 − φ Lh ) yt = λ (1 − φ Lh ) W ( L1/ m ; γ ) xt + ut .
6
In this instance “lead” refers to an observation of the high-frequency
predictor that corresponds to the same temporal period of the low fre-
quency variable.
252
and then they extract a common component and separate that into
a long-run and a cyclical, or short-run, component. The second step
consists in fitting the model of Clements and Galvão (2008) to cap-
ture short-term dynamics and use financial time series as high fre-
quency regressors, in addition to the long-run component previously
estimated as well as other low frequency variables. They design three
midas models, M1, M2 and M3, each with different high frequen-
cy regressors: i) M1 includes the short-term interest rate, changes
in interest rate spread and oil future prices; ii) M2 uses changes in the
wheat price, oil future quotes and the exchange rate; and finally, iii)
M3 consists of long-term rates, changes in the interest rate spreads,
and changes in the short-term rate. They contrast the out-of-sample
performance in terms of rsmfe of these models against the equations
for the inflation rate of two different low frequency vector autore-
gressions, and univariate random walks, autoregressions and autore-
gressive-moving average models. They compute all the intra-period
forecasts for the midas models and the monthly average of these daily
forecasts, and compare this average to all the low frequency models.
All the analysis is conducted for one-month-ahead and two-month-
ahead forecasts. They find on average a 20% reduction in forecast
error dispersion. The authors also provide a final empirical exercise
by using forecast combinations with the midas models and the in-
flation rate implied by financial derivatives, but this approach does
not produce any significant gains.
Duarte (2014) discusses in detail the implications of autoregressive
augmentations in midas regression models and diverse ways to incor-
porate them. She explores the out-of-sample performance of midas
models with autoregressive augmentations with no restrictions, with
an autoregressive augmentation with a common factor restriction,
and models with autoregressive augmentations with no restrictions
and a multiplicative scheme to aggregation. She then compares these
models to the same models but without the autoregressive compo-
nent, and to two low frequency benchmark models, a low frequency
autoregression and multiple regression model. She computes fore-
casts for quarterly euro area gdp growth based on three different
series: i) industrial production, ii) an economic sentiment indicator
and iii) the Dow Jones Euro Stoxx index. She disregards the seasonal
spikes impulse responses as the relevant impulse responses, as she
argues that it is not possible to single out a particularly relevant im-
pulse response for a mixed-frequency process since responses vary
3.4 S (θ ) = ( y − X θ )′ ( − X θ ) + ηθ ′D ′Dθ
1 −2 1 0 L 0
0 1 −2 1 L 0
3.5 D= , # (3.5)
M O O O O M
0 L L 1 −2 1
254
4. DATA, EXERCISE, AND RESULTS
Atkeson and Ohanian (2001), Stock and Watson (2007) and Faust
and Wright (2009) have shown that simple benchmarks are not eas-
ily beaten by more sophisticated models (at least in the case of the
US economy), and so these could serve as a good starting point
to gauge the predictive power of the daily series.
4.1 Data
The online price index is compiled by the company PriceStats in co-
operation with State Street Global Markets, a leading financial re-
search corporation. PriceStates is a spin-off company that emerged
from the Billion Prices Project at mit, founded by professors Alber-
to Cavallo and Roberto Rigobón. It is the first company, institution,
or organization to apply a big data approach to produce real-time
(daily) price indexes to track general price inflation and other re-
lated metrics. Essentially, they collect daily data of prices from on-
line retailers by “web scraping” (i.e. recording price information
contained inside specific HyperText Markup Language tags in the
retailers’ websites) and aggregate the data by replicating the meth-
odology of a traditional consumer price index, as is done by Nation-
al Statistics Offices with offline prices. Cavallo (2013) goes through
the methodology and provides comparisons between online and of-
fline price indexes for Argentina, Brazil, Chile, Colombia, and Ven-
ezuela. He concludes that online price indexes can track the dynamic
behavior of inflation rates over time fairly well with the exception
of Argentina. In fact, the construction of online price indexes was ini-
tially motivated by the desire to provide the public with an alternate
measure of the inflation rate in Argentina because from the years
7
The last one coinciding with the lifting of the majority of the capital
controls in December 2015 and a subsequent transition to a flexible
exchange rate regime and inflation targeting.
256
Figure 1
COMPARISON BETWEEN rIPCBA INFLATION AND SPSS INFLATION
AGGREGATED TO MONTHLY FREQUENCY
300 9
270 8
240 7
210 6
180 5
150 4
120 3
90 2
60 1
30 0
0 −1
2013 2014 2015 2016 2017 2018 2013 2014 2015 2016 2017 2018
Period Period
ripcba ssps
inflation rate has been fluctuating around 2.2% for ripcba and 2.1%
for the monthly aggregated ssps series, with coefficients of variation
at 35% and 49% respectively. This should pose a significant chal-
lenge for economists’ ability to formulate accurate forecasts. Figure
1 illustrates the comparison between these two indexes and provides
a quick glimpse at the potential predictive power of the high-frequen-
cy index. Overall and for the scope of this work, ripcba is available
from July 2012 to December 2017 (66 data points) while ssps ranges
from November 1, 2007 to December 31, 2017 (3,714 data points).
8
First order midas-adl-cf models include m × L X + min ( L X , h ) .
high-frequency regressors since the common factor restriction increases
the number of variates depending on the forecast horizon and the num-
ber of high frequency lags.
9
A detailed list of the models can be found in Appendix A.
258
Estimation was conducted in R with the midasr package developed
by Ghysels et al. (2016) while optimization was performed with three
routines included in optimx10 for nonlinear models or with the lm
function from the stats package for linear ones. Models that require
optimx were solved simultaneously with three optimization routines
(ucminf, nlminb and Nelder-Mead) for each model, forecast horizon
h , number of high frequency regressors LX , and out-of-sample pe-
riod. Only the best solution was kept. The algorithm was initialized
taking the hypothesis of equal weights and a null impact parameter
as starting conditions. This strategy delivered reasonable results
empirically and serves as a check on whether the high-frequency re-
gressors are actually relevant.
10
A comprehensive description about this package can be found in Nash
and Varadhan (2011).
11
Tables with rmsfe ratios are presented in Appendix B.
12
This is similar to the standard test by Diebold and Mariano (1995).
The key difference lies in that the estimation sample size is kept fixed
instead of ever expanding, as this allows to better incorporate estimation
uncertainty and to compare nested models.
260
Table 1
OUT-OF-SAMPLE PREDICTIVE PERFORMANCE, rmsfe
h =1 h =2 h =3
LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3
Almon ( q = 2 )
midas-dl 0.630 0.627 0.564 0.790 0.717 0.712 0.740 0.721 0.751
midas-adl 0.578 0.589 0.564 0.775 0.729 0.724 0.760 0.741 0.765
midas-adl-cf 0.592 0.625 0.561 0.785 0.722 0.719 0.748 0.733 0.768
Almon ( q = 3 )
midas-dl 0.660 0.623 0.571 0.819 0.731 0.720 0.755 0.724 0.757
midas-adl 0.609 0.609 0.574 0.805 0.745 0.731 0.770 0.739 0.770
midas-adl-cf 0.617 0.636 0.576 0.827 0.741 0.719 0.762 0.734 0.777
Exp. Almon ( q = 2 )
midas-dl 0.705 0.646 0.566 0.816 0.755 0.749 0.803 0.863 0.837
midas-adl 0.627 0.633 0.632 0.775 0.768 0.766 0.839 0.840 0.835
midas-adl-cf 0.639 0.629 0.557 0.765 0.745 0.885 0.875 0.831 0.860
Exp. Almon ( q = 3 )
midas-dl 0.731 0.648 0.661 0.834 0.826 0.742 0.809 0.785 0.778
262
midas-dl 0.668 0.624 0.574 0.768 0.694 0.701 0.747 0.730 0.746
midas-adl 0.571 0.622 0.568 0.728 0.707 0.716 0.732 0.750 0.748
midas-adl-cf 0.614 0.619 0.558 0.739 0.697 0.704 0.740 0.736 0.737
Flat
midas-dl 1.158 0.617 0.568 0.745 0.673 0.690 0.713 0.736 0.745
midas-adl 0.944 0.592 0.568 0.733 0.694 0.746 0.729 0.766 0.777
Nonparametric
midas-dl 0.623 0.629 0.567 0.782 0.717 0.717 0.718 0.721 0.755
EW Forecast Combination
midas-dl 0.657 0.621 0.569 0.780 0.719 0.710 0.738 0.741 0.750
midas-adl 0.585 0.595 0.570 0.765 0.729 0.728 0.763 0.760 0.764
midas-adl-cf 0.609 0.627 0.563 0.768 0.734 0.754 0.770 0.756 0.770
Autoregression
Notes: The evaluation sample comprises 36 data points, from January 2015 to December 2017. Characters in bold indicate the best
number of variables, LX , for each model and forecast horizon, h . Characters in italics indicate the best model for each number
of variables, LX , and forecast horizon, h .
Table 2
OUT-OF-SAMPLE PREDICTIVE PERFORMANCE, rmsfe
h =1 h =2 h =3
LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3
Almon ( q = 2 )
midas-dl 0.555 0.521 0.427 0.646 0.541 0.538 0.568 0.548 0.547
midas-adl 0.460 0.482 0.430 0.608 0.547 0.544 0.579 0.562 0.569
midas-adl-cf 0.508 0.528 0.422 0.625 0.548 0.543 0.586 0.574 0.573
Almon ( q = 3 )
midas-dl 0.564 0.560 0.430 0.683 0.572 0.546 0.580 0.547 0.559
midas-adl 0.465 0.523 0.435 0.642 0.581 0.551 0.578 0.561 0.584
midas-adl-cf 0.533 0.569 0.433 0.663 0.581 0.549 0.596 0.581 0.601
Exp. Almon ( q = 2 )
midas-dl 0.568 0.523 0.432 0.656 0.540 0.539 0.574 0.561 0.561
midas-adl 0.470 0.453 0.453 0.600 0.538 0.535 0.568 0.572 0.578
midas-adl-cf 0.581 0.531 0.425 0.620 0.533 0.821 0.669 0.585 0.575
Exp. Almon ( q = 3 )
midas-dl 0.620 0.549 0.430 0.566 0.551 0.533 0.555 0.559 0.559
264
midas-dl 0.598 0.491 0.444 0.592 0.506 0.523 0.587 0.570 0.583
midas-adl 0.489 0.437 0.441 0.550 0.513 0.529 0.578 0.585 0.569
midas-adl-cf 0.548 0.492 0.432 0.568 0.511 0.525 0.580 0.570 0.571
Flat
midas-dl 0.542 0.515 0.417 0.610 0.501 0.526 0.539 0.577 0.594
midas-adl 0.540 0.479 0.405 0.577 0.515 0.547 0.544 0.580 0.604
Nonparametric
midas-dl 0.551 0.539 0.432 0.633 0.543 0.540 0.554 0.548 0.557
EW Forecast
Combination
midas-dl 0.488 0.522 0.429 0.621 0.533 0.533 0.562 0.555 0.561
midas-adl 0.417 0.462 0.426 0.597 0.544 0.534 0.565 0.566 0.573
midas-adl-cf 0.541 0.527 0.426 0.624 0.551 0.592 0.595 0.573 0.575
Autoregression
Notes: The evaluation sample comprises 18 data points, from July 2016 to December 2017. Characters in bold indicate the best
number of variables, LX , for each model and forecast horizon, h . Characters in italics indicate the best model for each number
of variables, LX , and forecast horizon, h .
Figure 2
EVOLUTION OF THE RMSFE FOR HORIZON h =1 WITHIN A MONTH
FOR SELECTED MODELS WITH LX = 3
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0 5 10 15 20 25 0 5 10 15 20 25
Day within month Day within month
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0 5 10 15 20 25 0 5 10 15 20 25
Day within month Day within month
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0 5 10 15 20 25 0 5 10 15 20 25
Day within month Day within month
Almon (q=3)
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0 5 10 15 20 25 0 5 10 15 20 25
Day within month Day within month
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0 5 10 15 20 25 0 5 10 15 20 25
Day within month Day within month
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0 5 10 15 20 25 0 5 10 15 20 25
Day within month Day within month
Almon (q=3)
266
Figure 4
EVOLUTION OF THE RMSFE FOR HORIZON h =3 WITHIN A MONTH
FOR SELECTED MODELS WITH LX = 2
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0 5 10 15 20 25 0 5 10 15 20 25
Day within month Day within month
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0 5 10 15 20 25 0 5 10 15 20 25
Day within month Day within month
0.6 0.6
0.5 0.5
0.4 0.4
0.3 0.3
0 5 10 15 20 25 0 5 10 15 20 25
Day within month Day within month
Almon (q=3)
268
-- -
7
6 Observed Value
5
-- Almond (q=2)
4
3 - Nonparametric
2
1
%)
1
Mon
0
01-2015 04-2015 07_2015 10-2015 01-2016 04-2016 07-2016 10-2016 01-2017 04-2017 07-2017 10-2017 01-2018
Date
ANNEX
midas-dl:
m ×LX −1
2
A.1 πˆTrIPCBA
+h = αˆ + λˆ ∑ ∑ γ s j s πTSSPS
−d − j m
j =0 s =1
m ×LX −1
3
A.2 πˆTrIPCBA
+h
ˆ
= αˆ + λ ∑ ∑ γˆs j s πTSSPS
−d − j m
j =0 s =1
e ∑ s =1 s
2
m ×LX −1 γ js
SSPS
A.3 ˆ
πTrIPCBA
+h = αˆ + λˆ ∑ m ×L −1 2 γ j s πT −d − j m
j =0 ∑ j =0 X e ∑ s =1 s
270
e ∑ s =1 s
3
m ×LX −1 γ js
SSPS
A.4 ˆ
πTrIPCBA
+h = αˆ + λˆ ∑ m ×L −1 3 γ j s πT −d − j m
j =0 ∑ j =0 X e ∑ s =1 s
A.5
A.6
A.7
midas-adl:
A.8
A.9
A.10
A.11
A.12
midas-adl-cf:
A.14
A.15
A.16
A.17
A.18
Other:
A.19
A.20
272
Appendix B: Additional Tables
Table B.1
OUT-OF-SAMPLE PREDICTIVE PERFORMANCE, RATIO TO rmsfe OF AUTOREGRESSION ×100
h =1 h =2 h =3
LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3
Almon ( q = 2 )
midas-dl 101.8 101.4 91.2 104.2 94.7 93.9 93.1 90.8 94.5
midas-adl 93.4 95.1 91.1 102.3 96.2 95.6 95.6 93.2 96.2
midas-adl-cf 95.7 101.0 90.6 103.6 95.3 94.9 94.1 92.2 96.6
Almon ( q = 3 )
midas-dl 106.6 100.7 92.2 108.1 96.5 95.0 95.0 91.2 95.2
midas-adl 98.3 98.4 92.7 106.3 98.4 96.5 96.9 93.0 96.9
midas-adl-cf 99.8 102.7 93.0 109.2 97.8 94.9 95.8 92.3 97.7
Exp. Almon ( q = 2 )
midas-dl 113.8 104.4 91.5 107.7 99.7 98.8 101.0 108.5 105.3
midas-adl 101.3 102.2 102.1 102.4 101.4 101.1 105.5 105.7 105.1
midas-adl-cf 103.2 101.6 90.0 100.9 98.4 116.8 110.1 104.6 108.2
Exp. Almon ( q = 3 )
274
midas-adl
midas-adl-cf 107.1 106.8 90.9 107.3 114.0 114.4 103.7 100.6 103.7
Beta
midas-dl 108.0 100.8 92.7 101.4 91.6 92.6 94.0 91.8 93.8
midas-adl 92.3 100.5 91.7 96.1 93.4 94.5 92.1 94.4 94.2
midas-adl-cf 99.2 100.1 90.2 97.6 92.0 92.9 93.1 92.6 92.7
Flat
midas-dl 187.1 99.8 91.8 98.3 88.8 91.1 89.7 92.6 93.8
midas-adl 152.6 95.6 91.8 96.7 91.6 98.5 91.7 96.4 97.8
Nonparametric
midas-dl 100.6 101.6 91.7 103.3 94.6 94.6 90.4 90.7 95.0
EW Forecast
Combination
midas-dl 106.2 100.4 92.0 102.9 95.0 93.7 92.9 93.3 94.3
midas-adl 94.5 96.1 92.2 100.9 96.3 96.1 96.0 95.6 96.2
midas-adl-cf 98.4 101.4 90.9 101.4 96.9 99.5 96.9 95.1 96.9
Unconditional Mean
Notes: The evaluation sample comprises 36 data points, from January 2015 to December 2017. Characters in bold indicate the best number
of variables, LX , for each model and forecast horizon, h . Characters in italics indicate the best model for each number of variables, LX ,
and forecast horizon, h .
Table B.2
OUT-OF-SAMPLE PREDICTIVE PERFORMANCE, RATIO TO rmsfe OF AUTOREGRESSION×100
h =1 h =2 h =3
LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3
Almon ( q = 2 )
midas-dl 100.6 94.4 77.4 95.7 80.1 79.6 82.2 79.2 79.0
midas-adl 83.4 87.4 77.9 90.1 81.0 80.6 83.8 81.2 82.4
midas-adl-cf 92.0 95.7 76.5 92.6 81.2 80.5 84.8 82.9 82.8
Almon ( q = 3 )
midas-dl 102.3 101.5 78.0 101.2 84.8 80.8 83.9 79.2 80.8
midas-adl 84.2 94.7 78.9 95.1 86.0 81.6 83.6 81.1 84.4
midas-adl-cf 96.6 103.2 78.5 98.3 86.0 81.3 86.1 84.1 86.9
Exp. Almon ( q = 2 )
midas-dl 103.0 94.8 78.3 97.1 80.0 79.9 83.1 81.2 81.2
midas-adl 85.3 82.1 82.1 88.9 79.7 79.2 82.1 82.8 83.5
midas-adl-cf 105.4 96.2 77.1 91.9 79.0 121.6 96.8 84.6 83.2
Exp. Almon ( q = 3 )
midas-dl 112.4 99.5 77.9 83.8 81.6 79.0 80.3 80.8 80.8
276
midas-adl
midas-adl-cf 111.1 98.8 77.0 106.6 97.3 94.0 82.9 83.7 83.7
Beta
midas-dl 108.5 89.1 80.5 87.7 74.9 77.5 84.9 82.4 84.3
midas-adl 88.6 79.2 80.0 81.5 75.9 78.4 83.5 84.6 82.3
midas-adl-cf 99.4 89.3 78.3 84.1 75.7 77.7 83.8 82.4 82.5
Flat
midas-dl 98.3 93.4 75.6 90.3 74.2 77.9 77.9 83.4 85.9
midas-adl 98.0 86.8 73.4 85.5 76.3 81.0 78.7 83.8 87.3
Nonparametric
midas-dl 100.0 97.8 78.3 93.8 80.4 80.0 80.1 79.2 80.6
EW Forecast
Combination
midas-dl 88.4 94.7 77.7 91.9 78.9 78.9 81.2 80.2 81.2
midas-adl 75.5 83.9 77.3 88.4 80.6 79.1 81.8 81.9 82.8
midas-adl-cf 98.1 95.5 77.3 92.4 81.6 87.7 86.0 82.9 83.1
Unconditional Mean
Notes: The evaluation sample comprises 18 data points, from July 2016 to December 2017. Characters in bold indicate the best number
of variables, LX , for each model and forecast horizon, h . Characters in italics indicate the best model for each number of variables, LX ,
and forecast horizon, h .
Table B.3
OUT-OF-SAMPLE PREDICTIVE PERFORMANCE, RATIO TO rmsfe OF UNCONDITIONAL MEAN×100
h =1 h =2 h =3
LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3
Almon ( q = 2 )
midas-dl 79.7 79.4 71.4 98.7 89.6 88.9 91.8 89.5 93.2
midas-adl 73.1 74.5 71.4 96.9 91.1 90.5 94.2 91.9 94.9
midas-adl-cf 75.0 79.1 71.0 98.1 90.3 89.8 92.8 90.9 95.3
Almon ( q = 3 )
midas-dl 83.5 78.9 72.2 102.4 91.3 89.9 93.7 89.9 93.9
midas-adl 77.0 77.1 72.6 100.6 93.2 91.3 95.5 91.7 95.6
midas-adl-cf 78.1 80.4 72.9 103.4 92.6 89.8 94.5 91.0 96.3
Exp. Almon ( q = 2 )
midas-dl 89.1 81.7 71.7 101.9 94.4 93.6 99.6 107.0 103.8
midas-adl 79.4 80.1 79.9 96.9 96.0 95.7 104.0 104.2 103.6
midas-adl-cf 80.8 79.5 70.5 95.6 93.1 110.6 108.5 103.1 106.7
Exp. Almon ( q = 3 )
midas-dl 92.5 82.0 83.6 104.3 103.3 92.7 100.3 97.4 96.6
278
midas-adl-cf 71.2 101.6
Beta
midas-dl 84.5 79.0 72.6 96.0 86.7 87.6 92.6 90.5 92.5
midas-adl 72.3 78.7 71.8 91.0 88.4 89.4 90.8 93.0 92.8
midas-adl-cf 77.6 78.4 70.6 92.4 87.1 88.0 91.8 91.3 91.4
Flat
midas-dl 146.5 78.1 71.9 93.1 84.1 86.2 88.4 91.3 92.5
midas-adl 119.5 74.9 71.9 91.6 86.7 93.3 90.5 95.0 96.4
Nonparametric
midas-dl 78.8 79.6 71.8 97.8 89.6 89.6 89.1 89.4 93.7
EW Forecast Combination
midas-dl 83.2 78.6 72.1 97.5 89.9 88.7 91.6 92.0 93.0
midas-adl 74.0 75.3 72.2 95.6 91.2 90.9 94.6 94.3 94.8
midas-adl-cf 77.1 79.4 71.2 96.0 91.7 94.2 95.5 93.7 95.5
Autoregression
Notes: The evaluation sample comprises 36 data points, from January 2015 to December 2017. Characters in bold indicate the best
number of variables, LX , for each model and forecast horizon, h . Characters in italics indicate the best model for each number
of variables, LX , and forecast horizon, h .
Table B.4
OUT-OF-SAMPLE PREDICTIVE PERFORMANCE, RATIO TO rmsfe OF UNCONDITIONAL MEAN×100
h =1 h =2 h =3
LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3 LX = 1 LX = 2 LX = 3
Almon ( q = 2 )
midas-dl 79.6 74.8 61.3 93.5 78.2 77.7 83.0 80.0 79.8
midas-adl 66.1 69.2 61.7 88.0 79.1 78.7 84.6 82.0 83.2
midas-adl-cf 72.9 75.8 60.6 90.4 79.2 78.5 85.6 83.8 83.7
Almon ( q = 3 )
midas-dl 81.0 80.3 61.8 98.8 82.8 78.9 84.7 79.9 81.6
midas-adl 66.7 75.0 62.4 92.9 92.9 84.0 79.7 84.5 81.9
midas-adl-cf 76.5 81.7 62.2 95.9 84.0 79.4 87.0 84.9 87.7
Exp. Almon ( q = 2 )
midas-dl 81.5 75.0 62.0 94.8 78.1 78.0 83.9 82.0 82.0
midas-adl 67.5 65.0 65.0 86.8 77.8 77.3 82.9 83.6 84.4
midas-adl-cf 83.5 76.2 61.0 89.7 77.1 118.7 97.7 85.4 84.0
Exp. Almon ( q = 3 )
midas-dl 89.0 78.8 61.7 81.8 79.6 77.2 81.1 81.6 81.6
280
midas-adl-cf
Beta
midas-dl 85.9 70.5 63.7 85.6 73.2 75.7 85.7 83.2 85.2
midas-adl 70.2 62.7 63.3 79.6 74.1 76.5 84.3 85.5 83.1
midas-adl-cf 78.7 70.7 62.0 82.1 73.9 75.9 84.6 83.2 83.3
Flat
midas-dl 77.8 73.9 59.8 88.2 72.4 76.1 78.7 84.2 86.8
midas-adl 77.6 68.7 58.1 83.5 74.5 79.1 79.4 84.6 88.2
Nonparametric
midas-dl 79.1 77.4 62.0 91.5 78.5 78.1 80.9 80.0 81.4
EW Forecast Combination
midas-dl 70.0 75.0 61.5 89.7 77.0 77.0 82.0 81.0 82.0
midas-adl 59.8 66.4 61.2 86.3 78.7 77.2 82.6 82.7 83.6
midas-adl-cf 77.6 75.6 61.2 90.2 79.7 85.6 86.8 83.7 84.0
Autoregression
Notes: The evaluation sample comprises 18 data points, from July 2016 to December 2017. Characters in bold indicate the best number
of variables, LX , for each model and forecast horizon, h . Characters in italics indicate the best model for each number of variables,
LX , and forecast horizon, h
References
Armesto, M.T., K.M. Engemann and M.T. Owyang (2010), “Fore-
casting with Mixed Frequencies,” Federal Reserve Bank of St.
Louis Review, 92(6), pp. 521–36.
Atkeson, A., and L.E. Ohanian (2001), “Are Phillips Curves Useful
for Forecasting Inflation?,” Federal Reserve Bank of Minneapolis
Quarterly Review, 25(1), pp. 2–11.
Breitung, J., and C. Roling (2015), “Forecasting Inflation Rates Us-
ing Daily Data: A Nonparametric midas Approach,” Journal
of Forecasting, 34(7), pp. 588–603.
Cavallo, A. (2013), “Online and Official Price Indexes: Measuring
Argentina’s Inflation,” Journal of Monetary Economics, 60(2),
pp. 152–165.
Clements, M.P., and A.B. Galvão (2008), “Macroeconomic Forecast-
ing with Mixed-Frequency Data: Forecasting Output Growth
in the United States,” Journal of Business & Economic Statistics,
26(4), pp. 546–554.
Diebold, F.X., and R.S. Mariano (1995), “Comparing Predictive
Accuracy,” Journal of Business & Economic Statistics, 13(3),
pp. 134–144.
Duarte, C. (2014), Autoregressive Augmentation of MIDAS Regressions,
Banco de Portugal Working Paper 201401, Lisbon, Portugal,
Banco de Portugal.
Faust, J., and J.H. Wright (2009), “Comparing Greenbook and Reduced
Form Forecasts Using a Large Realtime Dataset,” Journal of Busi-
ness & Economic Statistics, 27(4), pp. 468–479.
Forni, M. et al. (2000), “The Generalized Dynamic-Factor Model:
Identification and Estimation,” Review of Economics and Sta-
tistics, 82(4), pp. 540–554.
Foroni, C., and M. Marcellino (2013), A Survey of Econometric Methods
for Mixed-Frequency Data, Norges Bank Working Paper 2013-
06, Oslo, Norway: Norges Bank.
Foroni, C., M. Marcellino and C. Schumacher (2015), “Unrestricted
Mixed Data Sampling (midas): midas Regressions with Unre-
stricted Lag Polynomials,” Journal of the Royal Statistical Society:
Series A (Statistics in Society), 178(1), pp. 57–82.
282
Stock, J.H., and M.W. Watson (2007), “Why Has US Inflation
Become Harder to Forecast?,” Journal of Money, Credit and
Banking, 39(1), pp. 3–33.
Abstract
Inflation forecasts are a relevant input for monetary policy decision in central
banks, particularly for those operating under inflation targeting. Therefore,
central banks must continuously evaluate the forecasting accuracy of the mod-
els used to produce inflation forecasts. In this study, we present the results
of an exhaustive evaluation of historical forecast performance for the most
important models used to forecast inflation at Banco de Guatemala. We find
evidence supporting the claim that time series models perform better for short
time horizons while conditional dsge models (both structural and semi-struc-
tural) do better in medium and long time horizons.
Keywords: economic forecasting, forecasting accuracy, forecasting
efficiency.
jel classification: C530.
285
1. INTRODUCTION
B
anco de Guatemala adopted a monetary policy framework based
on inf lation targeting (it) in 2005. Because of the forward-
looking nature of that regime, central bank authorities should
base their policy decisions on reliable inflation forecasts. In fact,
Banco de Guatemala employs an array of models to forecast inflation,
which include ols, arima , structural and semi-structural dsge type
of models, as well as forecast combinations of all, or some of these
approaches. Since each of these models provides different informa-
tion about the future path of inflation, a rigorous evaluation of their
performance is required in order to determine their reliability, so that
the central bank staff could give more weight to more reliable models,
and improve the less reliable ones or get rid of them.
This document presents the results of a thorough evaluation of the
most frequently used models by Banco de Guatemala to forecast in-
flation. Our evaluation is divided according to the type of model
employed to produce a forecast. First, we evaluate models that pro-
duce unconditional forecasts, based on four different approach-
es: 1) forecasting accuracy and bias; 2) ability to predict a change
of trend; 3) prediction similarity; and 4) forecast efficiency. Second,
we assess the performance of models that produce conditional fore-
casts, by generating in-sample projections for different scenarios
of exogenous and endogenous variables. Our main findings indicate
that time series models perform better for short time horizons, while
the dsge models are more efficient forecasting longer time horizons.
The remaining of this document is organized as follows. Section
2 presents a description of all unconditional and conditional mod-
els employed by Banco de Guatemala to generate inflation forecasts.
Section 3 describes the data and methodology employed for evalua-
tion purposes. Section 4 shows the results obtained. Finally, Section
5 concludes.
2
A first evaluation was conducted considering a wider sample (2006Q1-
2017Q2), but results from this exercise were not as expected, in par-
ticular for headline inflation forecasts. This could be due to some
periods of high volatility in headline inflation. For example, inflation
went from 14.16% in the third quarter of 2008 towards a negative value
(–0.73%) one year later (in August 2009). Therefore, in order to get
robust results, we began our evaluation from 2011Q1.
3
Anchored values of inflation are slightly different from the correspond-
ing observed values because the inflation series generated by the model
has a quarterly frequency; hence, its annualized inflation rate is the
sum of four quarterly values rather than a 12-month variation rate.
4. RESULTS
Table 1
DM TEST, QUANTITATIVE MODELS
Forecasting
horizons in dm statistic dm Statistic dm statistic
quarters (amm-iv) (efp-iv) (efp-amm)
1 1.44 (0.15) 1.71 (0.087) 1.65 (0.09)
2 1.30 (0.19) 1.97 (0.049) 2.03 (0.04)
3 0.21 (0.84) 1.79 (0.074) 1.70 (0.09)
4 –2.95 (0.00) 1.76 (0.079) 1.61 (0.11)
8 –3.35 (0.02) 2.91 (0.004) –0.87 (0.38)
Then, the dm-statistic between the efp and the iv model is pre-
sented in Column 3. The statistic is statistically significant in all
forecasting horizons, which means that the mse s of the iv model
are lower than those of the efp model. Hence, we reject the null hy-
pothesis of equal accuracy. Also, the statistic is positive for the one-
to five-quarter horizons, which means that the mse s of the iv model
are lower than those of the efp. Hence, the iv model is more accurate
in its prediction of December inflation rate in the short and interme-
diate time horizons. On the other hand, the statistic is negative from
six to seven quarters ahead; therefore, the efp model is best in the
long run to predict the inflation rate. After that, the dm-statistic be-
tween the efp and the amm model is presented in Column 4. This
is statistically significant in all forecasting horizons, which means
that we reject the null hypothesis of equal accuracy. Also, in almost
all forecasting horizons, the mse s of the amm model are lower than
those of the efp model. Therefore, the amm model is best to predict
inflation rate in December.
1 ( )
Lt θˆj −h ,R ,γˆ j −h ,R = MSE AMM ,t − MSE IV ,t
Then, we compare the forecasting accuracy between the efp and the
iv model with the use of the gr test (see Column 3). The loss func-
tion between the two models is defined by Equation 2. In this case,
the null hypothesis of equal accuracy is rejected in every forecasting
horizon since the gr-statistic is higher than the critical value. This
means that, at least in one period, one model generates more accurate
forecasts of inflation. The graphical analysis shows that the forecasts
2 ( )
Lt θˆj −h ,R ,γˆ j −h ,R = MSE EFP ,t − MSE IV ,t
3 ( )
Lt θˆj −h ,R ,γˆ j −h ,R = MSE EEP ,t − MSE DIE ,t
Table 3
COMPARISON OF THE BEST SCENARIOS BETWEEN MMS 4.0.1 AND PIGU
ANNEX
Table A1.1
PT TEST, QUANTITATIVE MODELS
Forecasting
horizons in quarters Sn statistic (iv) Sn Statistic (amm) Sn statistic (efp)
1 4.28 3.98 2.41
2 3.77 2.93 1.62
3 2.57 1.54 0.73
4 0.00 0.88 –1.01
8 0.00 –1.49 –1.53
Sources: author’s elaboration, central bank’s forecasts.
Table A1.3
gr TEST, QUANTITATIVE MODELS
Table A1.4
WEAK EFFICIENCY TEST, QUANTITATIVE MODELS
Table A1.6
ME AND RMSE, MMS 4.01, 2011Q1-2017Q2
Anchoring
Anchoring exogenous and
exogenous endogenous
Anchoring variables and variables, plus
Forecasting exogenous two periods of two periods of
horizons in Free model variables inflation inflation
quarters me rmse me rmse me rmse me rmse
Table A1.8
ME AND RMSE, MME, 2011Q1-2017Q2
Abstract
This paper has two purposes. First, it evaluates the responses to the ques-
tions on inflation expectations in the World Economic Survey (wes) for six-
teen inflation targeting countries. Second, it compares inflation expectation
forecasts across countries by using a two-step approach that selects the most
accurate linear or non-linear forecasting method for each country. Then,
Self-Organizing Maps are used to cluster inflation expectations, setting as a
benchmark June 2014, when there was a sharp decline in oil prices. Analyz-
ing inflation expectations in the context of this price change makes it pos-
sible to distinguish between countries that anticipated the oil shock smoothly
and those that had to adjust their expectations significantly. The main find-
ings from the wes in-sample comparison suggest that expert forecasts of in-
flation expectations are systematically distorted in 83 percent of the countries
in the sample. On the other hand, the out of sample forecast analysis indicates
that Non-linear Artificial Neural Networks combined with Bayesian regular-
ization outperform arima linear models for longer forecasting horizons.
This holds true for countries with both soft and brisk changes of expectations.
However, when forecasting one step ahead, the performance between the two
methods is similar.
309
jel classifications: C02, C222, C45, C63, E27.
Keywords: Inflation expectations, Machine learning, Self-organizing
maps, Nonlinear auto-regressive neural network, Expectation surveys, Time
series models.
1. INTRODUCTION
C
ross-country data from economic expectations surveys have
recently highlighted the importance of analyzing and forecast-
ing public expectations to gain insight into crucial empirical is-
sues in macroeconomics. Expectations can influence the future path
of real economic variables and help guide policy decision-makers,
and inflation expectations are particularly important for countries
that utilize inflation targeting as their primary monetary policy
framework. The usefulness of inflation expectations is manifested
in various realms of economic analysis. They are critical for i) testing
theories of informational inflation rigidity (Coibion et al., 2012); ii)
estimating key structural parameters, such as the intertemporal
substitution elasticity (Crump et al., 2015); iii) testing public un-
derstanding of monetary policy, such as the Taylor rule (Carvalho
and Nechio, 2014); and iv) assessing how well inflation expectations
may be anchored among economic agents, which is key in assessing
the effectiveness of central bank communication. Lastly, New Keynes-
ian macroeconomic models have successfully used inflation expecta-
tions to predict real inflation (Henzel and Wollmershäuserab, 2008).
Expectation surveys have featured a wide range of respondents,
including economic experts, central bankers, financial agents, con-
sumers, and firms. Those surveyed often have to make important
decisions that take into account inflation and survey data, and their
responses provide information on the effectiveness of economic poli-
cies and institutional confidence. The World Economic Survey (wes)
collects data on inflation expectations across countries and surveys
more than 1,000 economic experts in approximately 120 countries.
The respondents evaluate present economic conditions and predict
the economic outlook of the country in which they reside, giving
special attention to price trends in their answers to both qualitative
and quantitative questions.
Thus, we must assess the suitability of wes data surveys and select
the appropriate methods to accurately forecast inflation expectations.
311
According to the wes forecast error analysis, we observe that even
though the forecasts meet at least the minimum standard when com-
pared to a random walk, economic experts have made systematic
errors in their predictions. That is, inflation was under- predicted
while increasing and over-predicted while declining in most of the
countries. Moreover, the mean squared error decomposition illus-
trated that there were systematic distortions in the inflation forecasts
in around 83 percent of the countries. The evidence suggests that al-
though the accuracy of the forecasts increases as the forecasting hori-
zon decreases, this relationship is not monotonic. This finding does
not support the hypothesis that forecasts have improved over time,
which may signal that there is a non-linear data- generating process.
Second, turning to a much more complex analysis, the som rep-
resentation allows us to cluster countries based on the evolution
of inflation expectations before the oil price shock. It is important
to note that the low inflation expectations cluster is relatively small
compared to the high and neutral clusters for inflation-targeting
countries. We find that in the one step- forward forecasts, the neural
network only slightly improves on forecasts of the arima , but that
it outperforms the arima model in the two step-forward forecasts
for Canada, Colombia, Chile, Poland, Hungary, and Sweden. There-
fore, using a non-linear neural network along with Bayesian regular-
ization leads to an improvement in expectations forecasts.
This paper contains five sections apart from this introduction
and proceeds as follows. In Section 2 we describe the wes data
and evaluate the responses to both qualitative and quantitative
inflation questions. In Section 3, we provide the methodologies
for clustering and forecasting, emphasizing the merits of the artifi-
cial neural network approach. In Section 4, we summarize the main
results, including the cluster analysis and forecasting accuracy. Fi-
nally, in Section 5 we present our conclusions and propose future
lines of research.
1
A survey form of the World Economic Survey, the wes questionnaire,
is included in Appendix A, see Figure 14.
313
Poland, Sweden, Thailand, and South Africa. 2 The relationship be-
tween the indicator of wes inflation expectations and the observed
annual inflation rate is illustrated through a simple exploratory analy-
sis that uses time plots and correlation statistics.3 The observed infla-
tion rate and the corresponding inflation expectations are depicted
in Figure 1 for some selected countries. For each country, inflation
was measured by annual changes in the Consumer Price Index. Ac-
cording to Figure 1, wes expectations move in tandem with actual
inflation for most of the period under study except during idiosyn-
cratic and global shocks that affected specific national economies.4
Figure 2 displays the correlation coefficient over time and the
coherence as a function of the frequency between the wes inflation
expectations and real annual inflation. The plot of the correlation
coefficient shows the existence of different patterns of linear asso-
ciation. For example, while the correlation in Mexico has increased
over time, it has decreased in Canada. On the other hand, Colom-
bia has experienced an inverted u-shaped correlation pattern that
peaks in the middle of 2002. According to frequency domain analy-
sis, higher coherence was found in higher frequencies of the spectral
distribution in most of the countries, which suggests that the relation-
ship between inflation expectations and observed inflation is strong
predominantly during short cycles. It is important to note that Asian
countries have higher coherence in lower frequencies, which points
to a different trend between expectation and observed inflation. 5
2
Figure 11 in Appendix A contains the full-time series length.
3
To see the other countries’ inflation expectations, see Figure 15 in
the Appendix.
4
In addition, we include a summary of the data, their histograms and cor-
relations which are relevant to the som analysis: Figure 12 in the Ap-
pendix reveals the heterogeneity of the variables, and Figure 13 displays
the correlation between them. Table 9 in the Appendix shows a brief
summary of the wes expectations data.
5
To see the spectral decomposition of the other countries, see Figure
16 in the Appendix.
0
2
4
6
8
10
12
Mar 2000 Dec 1991
Jun 1992
Sep 2000 Dec 1992
Mar 2001 Jun 1993
Sep 2001 Dec 1993
Jun 1994
Mar 2002 Dec 1994
Sep 2002 Jun 1995
Dec 1995
Mar 2003 Jun 1996
Sep 2003 Dec 1996
Jun 1997
Mar 2004 Dec 1997
Sep 2004 Jun 1998
Dec 1998
Mar 2005 Jun 1999
Sep 2005 Dec 1999
Jun 2000
Annual Inflation
Mar 2006 Dec 2000
Sep 2006 Jun 2001
Dec 2001
Mar 2007 Jun 2002
Sep 2007 Dec 2002
Jun 2003
Mar 2008 Dec 2003
Sep 2008 Jun 2004
Dec 2004
Mar 2009 Jun 2005
Figure 1
()
()
Jun 2006
Mar 2010 Dec 2006
Sep 2010 Jun 2007
Dec 2007
Mar 2011 Jun 2008
Sep 2011 Dec 2008
Jun 2009
Mar 2012 Dec 2009
Sep 2012 Jun 2010
Dec 2010
Jun 2012
Mar 2014 Dec 2012
Jun 2013
COMPARISON OF THE INFLATION EXPECTATIONS
Sep 2014
0
1
2
3
4
5
6
7
8
9
315
−1
0
1
2
3
4
5
6
4
8
−1
0
1
2
3
4
5
6
Dec 1991 Dec 1991
Jun 1992 Jun 1992
Dec 1992 Dec 1992
Jun 1993 Jun 1993
Dec 1993 Dec 1993
Jun 1994 Jun 1994
Dec 1994 Dec 1994
Jun 1995 Jun 1995
Dec 1995 Dec 1995
Jun 1996 Jun 1996
Dec 1996 Dec 1996
Jun 1997 Jun 1997
Dec 1997 Dec 1997
Jun 1998 Jun 1998
Dec 1998 Dec 1998
Jun 1999 Jun 1999
Dec 1999 Dec 1999
Jun 2000 Jun 2000
Annual Inflation
Dec 2000 Dec 2000
Jun 2001 Jun 2001
Dec 2001 Dec 2001
Jun 2002 Jun 2002
Dec 2002 Dec 2002
Jun 2003 Jun 2003
Dec 2003 Dec 2003
Jun 2004 Jun 2004
Dec 2004 Dec 2004
Jun 2005 Jun 2005
()
0
1
2
3
4
5
6
7
8
9
0
1
2
3
4
5
6
7
8
9
Figure 2
CORRELATION AND COHERENCE COEFFICIENTS OF QUALITATIVE WES
INFLATION EXPECTATION WITH OBSERVED ANNUAL INFLATION
0.8
0.4
0.6
Coherence
0.3
0.4
0.2 0.2
0.1 0.0
1990 1995 2000 2005 2010 2015 2020 0.00 0.62 1.24 1.86 2.48 3.10
Date Frequency from 0 to
0.2
0.6
Coherence
0.0
0.4
−0.2
0.2
−0.4
0.0
1990 1995 2000 2005 2010 2015 2020 0.00 0.62 1.24 1.86 2.48 3.10
Date Frequency from 0 to
0.4 1.0
Correlation coefficient
0.8
0.2
0.6
Coherence
0.0 0.4
0.2
−0.2
0.0
1990 1995 2000 2005 2010 2015 2020 0.00 0.62 1.24 1.86 2.48 3.10
Date Frequency from 0 to
317
Figure 2 (cont.)
CORRELATION AND COHERENCE COEFFICIENTS OF QUALITATIVE WES
INFLATION EXPECTATION WITH OBSERVED ANNUAL INFLATION
0.2 0.8
Correlation coefficient
0.0 0.6
Coherence
−0.2 0.4
−0.4
0.2
−0.6
0.0
1990 1995 2000 2005 2010 2015 2020 0.00 0.62 1.24 1.86 2.48 3.10
Date Frequency from 0 to
6
The respective statistics equations are presented in Appendix A.3,
and mae and U-statistic results are in Tables 10 and 11, respectively.
See Appendix.
7
To see the other countries quantitative inflation expectation, see Figure
17 in Appendix A.3.
8
The quarter-specific forecasting error by country is plotted in Figure
18, Appendix A.3.
319
The evidence for Colombia suggests that actual annual inflation
was overestimated during the period from 2000 to 2003, and from
2003 to 2007 the expectations were close to the observed inflation
rate. The 2008 financial crises led expectations to undershoot ob-
served inflation for a short period of time, but soon after, expecta-
tions began to overshoot observed inflation until 2014. Eventually,
the 2014 oil shock induced a period of undershooting. There are dif-
ferent patterns across the countries. For example, in Mexico expec-
tations were close to actual inflation until the oil shock, but after
the shock, they overestimated observed inflation rates. In Tables
3 and 4 we count the number of years in which inflation was over-
estimated and underestimated respectively by respondents, to the
quarterly wes survey. For instance, the results indicate that annual
inflation in Colombia was overestimated, on average, in 14 of 25 years
and for Mexico in 17 of 26 years. There is evidence that systematic
overestimation was greater than underestimation. The exception
occurs in the case of Brazil in which, on average, in 15 of 26 years in-
flation was underestimated by economic experts.
Finally, a cross-country comparison using the U-statistic confirms
that the wes -forecasts in every country at least meet the minimum
standard when compared with the random walk alternative.
321
Table 2
THEIL ERROR DECOMPOSITION OF THE WES
FORECAST ERRORS Q1 1991 TO Q2 2016
323
Table 3
OVERESTIMATION OF WES FORECASTS QI 1991 TO Q2 2016
325
Table 4
UNDERESTIMATION OF WES FORECASTS Q1 1991 TO Q2 2016
327
0.0
2.0
4.0
6.0
8.0
10.0
12.0
14.0
0.00
0.50
1.00
1.50
2.00
2.50
3.00
3.50
Ene 2000 Ene 2000
Mar 2000 Mar 2000
Ene 2001 Ene 2001
Mar 2001 Mar 2001
Ene 2002 Ene 2002
Mar 2002 Mar 2002
Ene 2003 Ene 2003
Mar 2003 Mar 2003
Ene 2004 Ene 2004
Mar 2004 Mar 2004
Ene 2005 Ene 2005
Mar 2005 Mar 2005
Ene 2006 Ene 2006
wes expectations
Mar 2006 Mar 2006
Ene 2007 Ene 2007
Mar 2007 Mar 2007
Ene 2008 Ene 2008
Figure 3
()
()
Ene 2009 Ene 2009
wes expectations
Mar 2006 Mar 2006
Ene 2007 Ene 2007
Mar 2007 Mar 2007
Ene 2008 Ene 2008
Figure 3
329
3. METHODOLOGY
the outputs and inputs.9 To see more details about anns see Hagan
et al. (2014).
9
In the nar-nn Model, to perform multi-step forecasts, the network
is transformed into a recurrent network after their parameters were
trained as a feed-forward network.
331
patterns should be represented on the map by nearby output units;
see Kohonen (2001).
The som architecture consists of a two-layer network: in the first lay-
er the inputs are multiplied with weights that were initialized as small
numbers. Then the results are evaluated by a competitive function
that produces a wining neuron (Best Matching unit). The weights
are updated according to the learning rule, equation (2), and the
neuron’s neighborhood is updated as well. See Figure 5 below.
2 wi (q ) = (1 − α )w i (q − 1) + α ( p (q ) )
Figure 6
WEIGHT SOM VECTORS OF WES EXPECTATIONS FOR THE NEXT 6 MONTHS
333
3 ( )
Yt = g Yt −1 + Yt −2 + ... + Yt − p + et
4 ( ( ) )
Yt +1 = f 2 W 2 f 1 W 1 Yt ,Yt −1 ,...,Yt − p + b1 + b2 + et +1
( ) (Yt − Yt ) +α ∑ni =1xi2
T
F (x ) = β ∑ t =1 Yt − Yt
T
5
P ( D |α , β , M ) P (α , β | M )
6 P (α , β | D , M ) =
P ( D |M )
The prior joint density for α and β is assumed from the uniform
distribution. Consequently, the posterior can be obtained by com-
puting the following probabilities:
10
Overfitting is a characteristic that should be avoided and occurs when
the neural network fit the data closely in the training set, but in the test-
ing set and out of sample, the fitting is poor.
P ( D |X , β , M ) P ( X |α , M )
8 P ( X |D ,α , β , M ) =
P ( D |α , β , M )
9 e n − e −n
a=
e n + e −n
335
10 a = n
For the output layer the linear function is used.11 The final archi-
tecture in matrix notations and scalar is:
11
( ( )
Yt +1 = f 2 W 2 f 1 W 1 Yt ,Yt −1 ,...,Yt − p + b1 + b2 )
10 p
yt +1 = ∑w 2j fj1 ∑wi1+ pYt1+i − p + b 1 + b 2
j =1 i =0
n
p =
(
2 p − p min ) −1
(p max
−p min
)
where wi1+ p , i =1, ... , p, w 2j , i =1, ... , p are the weights of the output layer,
b 1 is the biases of the first layer, and b 2 the biases of the second layer.
Figure 7 displays the observed data (black line), the fit in the train-
ing set (blue line), the forecasts in horizons 1 and 2 (green and orange
lines, respectively), and the out- of-sample forecasts eight steps ahead
(yellow line). Also, the figure is divided into three blocks. The block
on the left corresponds to the training set from Q31991 to Q2 2014;
the center block corresponds to the testing set from Q3 2014 to Q2
2016, which occurs after the oil shock period, and the right block
is the forecasting period.
3.4 arima
Box and Jenkins proposed the arima model in 1970 (Box et al., 2016).
The general expression of an arima model is the following:
Θs ( Ls )θ ( L )
Yt = εt
12
( )
Φs Ls ϕ ( L ) ∆sD ∆d
11
Notice that before training the network, data normalization, which
transforms the data in the interval between [-1, 1], is required to make
the training algorithm faster.
Jul 1991
Mar 1992
Nov 1992
Jul 1993
Mar 1994
Observed
Nov 1994
Jul 1995
Mar 1996
Nov 1996
Jul 1997
Mar 1998
Nov 1998
Jul 1999
Mar 2000
One-step-ahead Training Nov 2000
Jul 2001
Training set
Mar 2002
Nov 2002
Jul 2003
Mar 2004
Nov 2004
Figure 7
Testing set
Jul 2005
Mar 2006
Nov 2006
One-step-ahead Testing
Jul 2007
Mar 2008
Nov 2008
Out of sample
Jul 2009
DATA BLOCK DIVISION AND OUT OF SAMPLE SETS
Mar 2010
Nov 2010
Jul 2011
Mar 2012
Two-step-ahead
Nov 2012
Jul 2013
Mar 2014
Nov 2014
Jul 2015
Mar 2016
Out of sample
Nov 2016
Jul 2017
( ) ( )
where Θs Ls = 1 − Θs Ls − Θ2s L2s − Θ3s L3s − ... − ΘQs LQs is a seasonal
( ) (
moving average polynomial, Φs Ls = 1 − Φs Ls − Φ2s L2s − ... − Φ 3s L3s )
is the sea sona l auto -reg ressive poly nomia l, θ ( L ) = (1 − θ L
1 1 − θ2 L2 − ... −
( )
θ ( L ) = 1 − θ1L1 − θ2 L2 − ... − θq Lq is the regular moving average polynomi-
( )
al, and ϕ ( L ) = 1 − ϕ1L − ϕ2 L2 − ... − ϕ p Lp is a regular auto-regressive
1
4. RESULTS
In this section, we present the main results of the clustering and fore-
casting for inflation expectations across countries. First, we present
the som analysis that includes three sequential steps: the choice of the
map topology based on data, the training and validation stages of the
som neural network, and the elaboration of the clustering map of
agent expectations (in Appendix B we include a detailed explana-
tion of these steps). Then we overlap agents’ inflation expectations
on the resulting som map. Finally, the nar-nn results are provided.
12
The arima models chosen are described in Appendix D.
13
Appendix B explains the choice of topology as well as the post-training
analysis of the results.
A key tool in this analysis is the feature map or heat map that is the
representation of a single variable across the map (Figure 6). In this
application, the colors identify the intensity of the indicator. For ex-
ample: while the blue color is associated with low expectations, the red
is associated with high expectations. Clustering can be performed
by using hierarchical clustering on the weight learned vectors of the
variable. This procedure requires one to set the number of clusters.
339
Thus, given the nature of the expectations, we choose three clusters
to represent low, neutral, and high expectations.
341
Table 5
CLASSIFICATION OF INFLATION EXPECTATIONS
AND LAG SELECTED IN THE nar-nn MODEL
Brazil Brisk 1
Canada Brisk 8
Chile Smooth 4
Colombia Brisk 5
Czech R. Smooth 6
Korea R. Smooth 2
Mexico Smooth 6
Norway Smooth 1
Switzerland Brisk 8
United K. Smooth 6
Hungary Smooth 10
Philippines Brisk 1
Poland Smooth 7
Sweden Smooth 1
Thailand Brisk 4
South A. Brisk 1
14
In most of the cases mean and median, of the lag chosen, are both
the smallest. However, in Colombia, Czech Republic and Switzerland
this is not the case, even though the lag’s mean is closer to the small-
est mean.
15
These results for all datasets and training sets are presented in Tables
12 and 13, respectively, in Appendix 3.
16
To see the other countries, see Figure 24 in Appendix C.
17
A summary of results of the neural networks parameters is presented
in Table 14 in Appendix 3.
18
A simulation of 1000 networks was performed to ensure that the mse
presented belongs to the average neural network find after specifying
the model previously described. See Table 15 and Appendix 3.
343
Table 6
LAG STATISTICS TEST DATA - ONE-STEP AHEAD FORECAST. SAMPLE = 30
Countries Lags 1 2 3 4 5 6 7 8 9 10
Brazil mean 1.65 2.08 1.85 1.86 1.89 1.85 1.83 1.84 1.92 2.04
median 1.57 2.07 1.85 1.86 1.83 1.85 1.83 1.84 1.92 2.04
Canada mean 2.04 1.84 1.59 1.63 1.62 1.54 1.56 1.52 1.75 1.73
median 2.04 1.74 1.59 1.63 1.62 1.54 1.56 1.52 1.75 1.73
Switzerland mean 1.32 1.21 1.04 1.04 1.02 0.98 1.06 0.79 0.93 0.78
median 1.30 1.22 1.04 1.04 1.02 0.98 1.06 0.78 0.94 0.83
Chile mean 4.34 2.76 2.77 2.70 2.74 2.80 2.94 3.13 3.13 2.93
median 4.38 2.76 2.79 2.68 2.76 2.81 3.00 3.06 3.13 2.91
Colombia mean 4.82 2.88 2.91 2.90 2.83 2.88 2.86 3.27 3.21 3.27
345
1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
Jul 1991 Jul 1991
Mar 1992 Mar 1992
Nov 1992 Nov 1992
Jul 1993 Jul 1993
Mar 1994 Mar 1994
Nov 1994 Nov 1994
Jul 1995 Jul 1995
Mar 1996 Mar 1996
Nov 1996 Nov 1996
Jul 1997 Jul 1997
Mar 1998 Mar 1998
Nov 1998 Nov 1998
Jul 1999 Jul 1999
Mar 2000 Mar 2000
Nov 2000 Nov 2000
Observed
One-step-ahead Testing
Jul 2001 Jul 2001
Mar 2002 Mar 2002
Nov 2002 Nov 2002
Jul 2003 Jul 2003
Mar 2004 Mar 2004
Nov 2004 Nov 2004
Jul 2005 Jul 2005
()
Figure 10
Mar 2006
()
Mar 2006
Nov 2006 Nov 2006
Jul 2007 Jul 2007
Two-step-ahead
Nov 2008
Jul 2009 Jul 2009
Mar 2010 Mar 2010
Nov 2010 Nov 2010
One-step-ahead Training
Jul 2011 Jul 2011
Mar 2012 Mar 2012
Nov 2012 Nov 2012
Jul 2013 Jul 2013
Mar 2014 Mar 2014
Nov 2014 Nov 2014
Jul 2015 Jul 2015
Out of sample
Mar 2016 Mar 2016
Nov 2016 Nov 2016
FORECASTS OF INFLATION EXPECTATIONS USING THE NAR-NN MODEL
Observed
One-step-ahead Testing
Jul 2001 Jul 2001
Mar 2002 Mar 2002
Nov 2002 Nov 2002
Jul 2003 Jul 2003
Mar 2004 Mar 2004
Nov 2004 Nov 2004
Jul 2005 Jul 2005
()
Two-step-ahead
Nov 2008 Nov 2008
Jul 2009 Jul 2009
Mar 2010 Mar 2010
Nov 2010 Nov 2010
One-step-ahead Training
Jul 2011 Jul 2011
Mar 2012 Mar 2012
Nov 2012 Nov 2012
Jul 2013 Jul 2013
Mar 2014 Mar 2014
Nov 2014 Nov 2014
Jul 2015 Jul 2015
Out of sample
Mar 2016 Mar 2016
Nov 2016 Nov 2016
FORECASTS OF INFLATION EXPECTATIONS USING THE NAR-NN MODEL
347
4.4 Forecast Accuracy
Table 7
MSE COMPARISON AT TESTING DATA SETS FOR COUNTRIES
WITH BRISK INFLATION EXPECTATIONS
Table 8
MSE COMPARISON AT TESTING DATA SETS FOR COUNTRIES
WITH SOFT INFLATION EXPECTATIONS
349
the closer the expert is to the end of the year, the smaller the predic-
tion bias. This group includes Colombia and Switzerland among
others. The other group of countries exhibit increasing bias in the
last quarter of the prediction period and include Brazil, Canada,
and Chile.
Additionally, the quality of the quantitative question is judged
by standard measures of forecast evaluation at different horizons:
rmse, mae, and U-Theil. Thus, we concluded that the forecasts
meet a minimum standard compared to the random walk reference
and that economic experts have made systematic errors in their pre-
dictions. Inflation was under- predicted when it was rising and over-
predicted when it was declining in most of the countries. The Theil
decomposition of the mae illustrated that 83 percent of the countries
experienced systematic distortion in their forecasts, which means
that the increase in accuracy with shorter forecast horizons is not
monotonic. The evidence does not support the claim that forecasts
have improved over time due to a non-linear generating data process.
The evidence also suggests that turning points of observed average
inflation were mostly anticipated in most cases. This issue may be
an interesting area for further research.
On the other hand, a Self-Organizing Map analysis of surveys ex-
pectations before the impending oil shock allows us to classify infla-
tion expectations as either brisk or soft based on the speed with which
expectations shift. Using this classification, we can select the most
appropriate forecasting method. We notice that the low-inflation ex-
pectations cluster is relatively small compared to high and neutral
clusters for inflation targeting countries. The Nonlinear auto-re-
gressive neural network and arima methods were used as competing
candidates to forecast inflation expectations. The results indicate
that in the one step ahead forecasts the neural network is slightly bet-
ter, but in two step-ahead forecasts, it outperforms the arima mod-
el significantly. For Canada, Colombia, Chile, Poland, Hungary,
and Sweden in particular, the neural network produces significant
improvement in the two-step ahead forecasts.
Further research is required to provide theoretical economic ex-
planations for the results of each country. Moreover, this combina-
tion between machine learning and statistics can be implemented
in a follow-up paper to forecast actual inflation.
Figure 11
EXPECTED INFLATION RATE FOR THE NEXT SIX MONTHS
Wes Qualitative Question
2 3 4 5 6 7 8
Inflation exp
8
Inflation exp
6
4
2
7 8
Inflation exp
2 3 4 5 6
351
Figure 11 (cont.)
EXPECTED INFLATION RATE FOR THE NEXT SIX MONTHS
Wes Qualitative Question
8 6
Inflation exp
4 2
8
Inflation exp
4 2 6
8
Inflation exp
4 2 6
6
Inflation exp
4
2
8
Inflation exp
6
4
2
8
6
Inflation exp
4
2
353
Figure 11 (cont.)
EXPECTED INFLATION RATE FOR THE NEXT SIX MONTHS
Wes Qualitative Question
Inflation exp
2 3 4 5 6 7
8
Inflation exp
6
4
2
8
Inflation exp
6
4
2
8
Inflation exp
6
4
2
355
Table 9
DATA SUMMARY OF wes EXPECTATIONS FROM Q3 1991 TO Q2 2016
Selected countries
Capital Private
Overall economy expenditures consumption Inflation rate
Min 1 1 1 1
1stQ 4.8 4.7 4.57 4
Median 5.8 5.7 5.5 5.5
Mean 5.79 5.59 5.44 5.32
3rdQ 6.8 6.6 6.5 6.8
Max 9 9 9 9
Figure 13
SCATTER PLOT OF AGENTS’ EXPECTATIONS OF ECONOMIC SITUATION
FOR NEXT SIX MONTHS
Figure 14
EXAMPLE OF WORLD ECONOMIC SURVEY (WES) QUESTIONNAIRE
357
0
2
4
6
8
-2
0
1
2
3
4
5
6
7
8
-1
10
12
0
5
10
15
20
25
Mar-2000 Mar-2000 Dec-1995
Sep-2000 Sep-2000 Jun-1996
Mar-2001 Mar-2001 Dec-1996
Jun-1997
Sep-2001 Sep-2001
Dec-1997
Mar-2002 Mar-2002 Jun-1998
Sep-2002 Sep-2002 Dec-1998
Mar-2003 Mar-2003 Jun-1999
Dec-1999
Sep-2003 Sep-2003 Jun-2000
Mar-2004 Mar-2004 Dec-2000
Sep-2004 Sep-2004 Jun-2001
Mar-2005 Mar-2005 Dec-2001
Jun-2002
Sep-2005 Sep-2005 Dec-2002
Annual Inflation
Sep-2007 Sep-2007 Jun-2005
Mar-2008 Mar-2008 Dec-2005
Sep-2008 Sep-2008 Jun-2006
Dec-2006
Mar-2010 Mar-2010 Jun-2008
Sep-2010 Dec-2008
Sep-2010
Jun-2009
Mar-2016 Jun-2016
Sep-2016 Sep-2016
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
0
1
2
3
4
5
6
0
2
4
6
8
0
2
4
6
8
10
12
10
12
Mar-2000 Mar-2000 Mar-2000
Sep-2000 Sep-2000 Sep-2000
Mar-2001 Mar-2001 Mar-2001
Sep-2001 Sep-2001 Sep-2001
Mar-2002 Mar-2002 Mar-2002
Sep-2002 Sep-2002 Sep-2002
Mar-2003 Mar-2003 Mar-2003
Sep-2003 Sep-2003 Sep-2003
Mar-2004 Mar-2004 Mar-2004
Sep-2004 Sep-2004 Sep-2004
Mar-2005 Mar-2005 Mar-2005
Sep-2005 Sep-2005 Sep-2005
Annual Inflation
Sep-2007 Sep-2007 Sep-2007
Mar-2008 Mar-2008 Mar-2008
Sep-2008 Sep-2008 Sep-2008
Sep-2009 Sep-2009 Sep-2009
Mar-2010 Mar-2010 Mar-2010
Figure 15 (cont.)
Mar-2016 Mar-2016
Sep-2016 Sep-2016 Sep-2016
359
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
10.00
0
1
2
3
4
5
-3
-2
-1
0
2
4
6
8
0
2
4
6
8
-4
-2
10
12
10
12
14
16
Mar-2000 Mar-2000 Mar-2000
Sep-2000 Sep-2000 Sep-2000
Mar-2001 Mar-2001 Mar-2001
Sep-2001 Sep-2001 Sep-2001
Mar-2002 Mar-2002 Mar-2002
Sep-2002 Sep-2002 Sep-2002
Mar-2003 Mar-2003 Mar-2003
Sep-2003 Sep-2003 Sep-2003
Mar-2004 Mar-2004 Mar-2004
Sep-2004 Sep-2004 Sep-2004
Mar-2005 Mar-2005 Mar-2005
Sep-2005 Sep-2005 Sep-2005
Annual Inflation
Sep-2007 Sep-2007 Sep-2007
Mar-2008 Mar-2008 Mar-2008
Sep-2008 Sep-2008 Sep-2008
Mar-2009
Mar-2009 Mar-2009
Sep-2009 Sep-2009
Sep-2009
Mar-2010 Mar-2010 Mar-2010
Figure 15 (cont.)
Sep-2010
Mar-2016 Mar-2016
Sep-2016 Sep-2016
Sep-2016
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
10.00
10.00
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
-2
0
2
4
6
8
10
12
-6
-4
-2
0
2
4
6
8
10
-2
-1.5
-1
-0.5
0
0.5
1
1.5
2
2.5
3
3.5
Mar-2000 Mar-2000 Mar-1995
Sep-2000 Sep-2000 Sep-1995
Mar-2001 Mar-2001 Mar-1996
Sep-1996
Sep-2001 Sep-2001 Mar-1997
Mar-2002 Mar-2002 Sep-1997
Sep-2002 Sep-2002 Mar-1998
Mar-2003 Sep-1998
Mar-2003 Mar-1999
Sep-2003 Sep-2003 Sep-1999
Mar-2004 Mar-2004 Mar-2000
Sep-2004 Sep-2000
Sep-2004 Mar-2001
Mar-2005 Mar-2005 Sep-2001
Sep-2005 Sep-2005 Mar-2002
Annual Inflation
Sep-2007 Sep-2007
Mar-2008 Mar-2005
Mar-2008 Sep-2005
Sep-2008 Sep-2008 Mar-2006
Mar-2009 Mar-2009 Sep-2006
Sep-2009 Mar-2007
Sep-2009 Sep-2007
Mar-2010 Mar-2010 Mar-2008
Figure 15 (cont.)
Mar-2016
Sep-2016 Sep-2016 Sep-2016
361
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
0.00
1.00
2.00
3.00
4.00
5.00
6.00
7.00
8.00
9.00
10.00
10.00
Figure 16
CORRELATION COEFFICIENTS BETWEEN WES QUALITATIVE
INFLATION EXPECTATION AND ANNUAL INFLATION
. .: . .:
363
Figure 16 (cont.)
CORRELATION COEFFICIENTS BETWEEN WES QUALITATIVE
INFLATION EXPECTATION AND ANNUAL INFLATION
. : . :
1
∑ e ( L,Q (h ),t )
2016
13
26 1991
1 2
∑ e ( L,Q (h ),t )
2016
14 1991
26
Theil U.statistic:
1 2
∑ e ( L,Q (h ),t )
2016
26 1991
15
1 2 1 2
∑ q ( L,Q (h ),t ) ∑
2016 2016
1991 1991
p ( L,t )
26 26
Bias share:
1 2 1 2
∑
2016
26 1991 q ( L ,Q ( h ) ,t ) − ∑
2016
26 1991
p ( L ,t )
16 V(h)=
1 2
∑ e ( L,Q (h ),t )
2016
26 1991
The spread share:
2
Sq (h ) − Sq (h )
S (h ) =
17 1 2
∑ e ( L,Q (h ),t )
2016
26 1991
K (h ) =
( )
2 1 − rq ,p (h ) Sq (h ) − S (h )
18 1 2
∑ e ( L,Q (h ),t )
2016
26 1991
365
Table 10
MAE OF WES SURVEY QUANTITATIVE INFLATION QUESTION
367
0
1
2
3
4
5
6
7
0
1
2
3
4
5
6
7
8
9
10
0
2
4
6
8
10
12
14
16
wes expectations
Mar 2006 Mar 2006 Mar 2006
Ene 2007 Ene 2007 Ene 2007
Mar 2007 Mar 2007 Mar 2007
Ene 2008 Ene 2008 Ene 2008
()
()
0
1
2
3
4
5
6
7
8
9
10
0
2
4
6
8
10
12
14
Ene 2000 Ene 2000 Ene 2000
Mar 2000 Mar 2000 Mar 2000
Ene 2001 Ene 2001 Ene 2001
Mar 2001 Mar 2001 Mar 2001
Ene 2002 Ene 2002 Ene 2002
Mar 2002 Mar 2002 Mar 2002
Ene 2003 Ene 2003 Ene 2003
Mar 2003 Mar 2003 Mar 2003
Ene 2004 Ene 2004 Ene 2004
Mar 2004 Mar 2004 Mar 2004
Ene 2005 Ene 2005 Ene 2005
Mar 2005 Mar 2005 Mar 2005
Ene 2006 Ene 2006 Ene 2006
Mar 2006 Mar 2006 Mar 2006
Ene 2007 Ene 2007 Ene 2007
Mar 2007 Mar 2007 Mar 2007
Ene 2008 Ene 2008 Ene 2008
()
Mar 2008
()
()
Ene 2009 Ene 2009 Ene 2009
Figure 17 (cont.)
369
COUNTRIES’ QUANTITATIVE INFLATION EXPECTATIONS AND ANNUAL INFLATION
−2
0
2
4
6
8
10
12
14
−2
0
2
4
6
8
10
12
−1.0
−0.5
0.0
0.5
1.0
1.50
2.0
2.5
3.0
3.5
4.0
Ene 2000 Ene 2000
Ene 2000
Mar 2000 Mar 2000
Mar 2000
Ene 2001 Ene 2001
Ene 2001
Mar 2001 Mar 2001
Mar 2001
Ene 2002 Ene 2002
Ene 2002
Mar 2002 Mar 2002
Mar 2002
Ene 2003 Ene 2003
Ene 2003
Mar 2003 Mar 2003
Mar 2003
Ene 2004 Ene 2004
Ene 2004
Mar 2004 Mar 2004
Mar 2004
Ene 2005 Ene 2005
Ene 2005
Mar 2005 Mar 2005
Mar 2005
Ene 2006 Ene 2006
Ene 2006
Mar 2006 Mar 2006
Mar 2006
Ene 2007 Ene 2007
Ene 2007
Mar 2007 Mar 2007
Mar 2007
Ene 2008 Ene 2008
()
Ene 2008
()
Ene 2009
Figure 17 (cont.)
371
COUNTRIES’ QUANTITATIVE INFLATION EXPECTATIONS AND ANNUAL INFLATION
Figure 18
QUARTER-SPECIFIC FORECASTING ERROR BY COUNTRY
373
Figure 18 (cont.)
QUARTER-SPECIFIC FORECASTING ERROR BY COUNTRY
375
Figure 18 (cont.)
QUARTER-SPECIFIC FORECASTING ERROR BY COUNTRY
377
Figure 18 (cont.)
QUARTER-SPECIFIC FORECASTING ERROR BY COUNTRY
379
ANNEX B. SELF-ORGANIZING MAP VALIDATION
19
The quantization error is not comparable between maps because it is
susceptible to map size. To see more about topographic errors see the
Post-training analysis section.
1.9
Best matching unit error 1.8
1.7
1.6
1.5
1.4
5 10 15
1.2
Quantization error
1.0
0.8
0.6
0.4
5 10 15
1.8
Error node distant
1.6
1.4
1.2
5 10 15
50
Average sample per node
40
30
20
10
5 10 15
381
Figure 20
POST-TRAINING ANALYSIS
. /
.
383
Figure 21
THE U-MATRIX
. 45
.
385
B.3 Non-Linear Auto-Regressive Neural Networks Validation and Other Results
B.3.1 Lag Selection
Table 12
LAG STATISTICS ON ALL DATA, ONE STEP-AHEAD FORECASTS, SAMPLE OF 30
Countries Lags 1 2 3 4 5 6 7 8 9 10
Brazil mean 2.01 1.99 1.88 1.90 1.93 1.86 1.87 1.88 1.89 1.88
median 1.99 1.99 1.88 1.90 1.90 1.86 1.87 1.88 1.89 1.88
Canada mean 1.17 1.38 1.30 1.29 1.30 1.30 1.31 1.32 1.29 1.30
median 1.16 1.38 1.31 1.29 1.30 1.30 1.31 1.32 1.29 1.30
Switzerland mean 1.03 1.08 1.04 1.04 1.03 1.03 0.83 0.33 0.50 0.65
median 1.03 1.09 1.04 1.04 1.03 1.03 0.87 0.31 0.50 0.83
387
Table 13
LAG STATISTICS ON TRAIN DATA, ONE STEP-AHEAD FORECAST, SAMPLE OF 30
Countries Lags 1 2 3 4 5 6 7 8 9 10
Brazil mean 2.04 1.98 1.88 1.90 1.94 1.86 1.87 1.89 1.89 1.86
median 2.03 1.98 1.88 1.90 1.90 1.86 1.87 1.89 1.89 1.86
Canada mean 1.10 1.34 1.28 1.26 1.27 1.27 1.28 1.30 1.24 1.26
median 1.07 1.34 1.28 1.26 1.27 1.27 1.28 1.30 1.24 1.26
Switzerland mean 1.01 1.07 1.05 1.04 1.03 1.03 0.80 0.29 0.46 0.64
median 1.01 1.07 1.05 1.04 1.03 1.03 0.85 0.27 0.46 0.83
Chile mean 2.04 2.03 2.02 2.15 2.04 1.96 1.99 2.00 1.94 2.01
median 2.03 2.03 1.98 2.18 2.11 1.87 2.11 2.07 1.94 2.02
389
B.3.2 Post-Training Analysis
Table 14
NEURAL NETWORKS RESULTS OF TRAINING PHASE
Total Effective Maximum Sum
number number of sum squared squared of Total
Countries of parameters of parameters parameters epoch
391
B.3.3 MSE Evaluation
Table 15
NEURAL NETWORK SIMULATIONS STATISTICS BY DATASETS, SAMPLE OF 1,000
Brazil Korea
Training Testing Training Testing
All data set set All data set set
mean 2.01 2.05 1.65 1.47 1.44 1.86
median 2.00 2.04 1.61 1.47 1.44 1.86
std 0.04 0.03 0.18 0.01 0.01 0.03
maximum 2.09 2.09 2.11 1.52 1.44 2.51
minimum 1.92 1.97 1.24 1.36 1.34 1.62
Canada Mexico
Training Testing Training Testing
All data set set All data set set
mean 1.33 1.31 1.52 0.97 1.03 0.34
median 1.32 1.30 1.52 0.90 0.95 0.30
std 0.06 0.06 0.01 0.24 0.24 0.17
maximum 2.09 2.09 2.11 3.91 4.00 2.94
minimum 1.92 1.97 1.24 0.90 0.95 0.30
Chile Norway
Training Testing Training Testing
All data set set All data set set
mean 2.21 2.17 2.69 1.87 1.91 1.41
median 2.23 2.18 2.68 1.86 1.90 1.42
std 0.06 0.07 0.03 0.04 0.04 0.03
maximum 1.33 1.36 1.01 2.06 2.12 1.51
minimum 0.55 0.54 0.67 1.82 1.86 1.25
Colombia Switzerland
Training Testing Training Testing
All data set set All data set set
mean 1.59 1.48 2.83 0.31 0.27 0.78
median 1.57 1.46 2.78 0.31 0.27 0.78
std 0.08 0.07 0.21 0.00 0.00 0.00
maximum 1.93 1.76 3.69 0.31 0.27 0.78
minimum 1.57 1.46 2.77 0.31 0.27 0.78
Hungary Philippines
Training Testing Training Testing
All data set set All data set set
mean 1.59 1.47 2.85 2.66 2.60 3.42
median 1.59 1.48 2.75 2.66 2.59 3.43
std 0.11 0.17 0.54 0.04 0.05 0.07
maximun 1.73 1.58 7.60 2.81 2.74 3.88
minimum 0.68 0.00 2.74 2.59 2.51 2.98
Poland Sweden
Training Testing Training Testing
All data set set All data set set
mean 0.89 0.90 0.72 1.25 1.26 1.14
median 0.89 0.90 0.72 1.24 1.25 1.15
std 0.01 0.01 0.01 0.03 0.04 0.09
maximun 1.02 1.04 0.80 1.35 1.37 1.31
minimum 0.88 0.90 0.66 1.21 1.21 0.86
393
1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
01/07/91 01/07/91 01/07/91
01/02/92 01/02/92 01/02/92
01/09/92 01/09/92 01/09/92
01/04/93 01/04/93 01/04/93
Observed
01/11/93 01/11/93 01/11/93
01/06/94 01/06/94 01/06/94
01/01/95 01/01/95 01/01/95
01/08/95 01/08/95 01/08/95
01/03/96 01/03/96 01/03/96
01/10/96 01/10/96 01/10/96
01/05/97 01/05/97 01/05/97
01/12/97 01/12/97 01/12/97
01/07/98 01/07/98 01/07/98
01/02/99 01/02/99 01/02/99
01/09/99 01/09/99 01/09/99
01/04/00 01/04/00 01/04/00
01/11/00 01/11/00 01/11/00
One-step-ahead Training
01/06/01 01/06/01 01/06/01
01/01/02 01/01/02 01/01/02
01/08/02 01/08/02 01/08/02
01/03/03 01/03/03 01/03/03
01/10/03 01/10/03 01/10/03
B.3.4 Results, Other Countries
.
01/07/05 01/07/05
.
01/07/05
Figure 23
.
01/04/07 01/04/07 01/04/07
One-step-ahead Testing
01/11/07
01/06/08 01/06/08 01/06/08
01/01/09 01/01/09 01/01/09
01/08/09 01/08/09 01/08/09
01/03/10 01/03/10 01/03/10
01/10/10 01/10/10 01/10/10
01/05/11 01/05/11 01/05/11
01/12/11 01/12/11 01/12/11
01/07/12 01/07/12 01/07/12
01/02/13 01/02/13
Two-step-ahead
01/02/13
01/09/13 01/09/13 01/09/13
01/04/14 01/04/14 01/04/14
01/11/14 01/11/14 01/11/14
01/06/15 01/06/15 01/06/15
01/01/16 01/01/16 01/01/16
01/08/16 01/08/16 01/08/16
01/03/17 01/03/17 01/03/17
COUNTRIES’ INFLATION RATE FORECAST FOR THE NEXT SIX MONTHS BY NAR-NN
01/10/17 01/10/17
Out of sample
01/10/17
1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
01/07/91 01/07/91 01/07/91
01/02/92 01/02/92 01/02/92
01/09/92 01/09/92 01/09/92
01/04/93 01/04/93 01/04/93
Observed
01/11/93 01/11/93 01/11/93
01/06/94 01/06/94 01/06/94
01/01/95 01/01/95 01/01/95
01/08/95 01/08/95 01/08/95
01/03/96 01/03/96 01/03/96
01/10/96 01/10/96 01/10/96
01/05/97 01/05/97 01/05/97
01/12/97 01/12/97 01/12/97
01/07/98 01/07/98 01/07/98
01/02/99 01/02/99 01/02/99
01/09/99 01/09/99 01/09/99
01/04/00 01/04/00 01/04/00
01/11/00
One-step-ahead Training
01/11/00 01/11/00
01/06/01 01/06/01 01/06/01
01/01/02 01/01/02 01/01/02
01/08/02 01/08/02 01/08/02
01/03/03 01/03/03 01/03/03
01/10/03 01/10/03 01/10/03
01/05/04 01/05/04 01/05/04
01/12/04 01/12/04 01/12/04
.
01/07/05 01/07/05
.
01/07/05
.
One-step-ahead Testing
01/11/07 01/11/07 01/11/07
01/06/08 01/06/08 01/06/08
01/01/09 01/01/09 01/01/09
01/08/09 01/08/09 01/08/09
01/03/10 01/03/10 01/03/10
01/10/10 01/10/10 01/10/10
01/05/11 01/05/11 01/05/11
01/12/11 01/12/11 01/12/11
01/07/12 01/07/12 01/07/12
Two-step-ahead
01/02/13 01/02/13 01/02/13
01/09/13 01/09/13 01/09/13
01/04/14 01/04/14 01/04/14
01/11/14 01/11/14 01/11/14
01/06/15 01/06/15 01/06/15
01/01/16 01/01/16 01/01/16
395
01/08/16 01/08/16 01/08/16
01/03/17 01/03/17 01/03/17
COUNTRIES’ INFLATION RATE FORECAST FOR THE NEXT SIX MONTHS BY NAR-NN
Out of sample
01/10/17 01/10/17 01/10/17
1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
01/07/91 01/07/91 01/07/91
01/02/92 01/02/92 01/02/92
01/09/92 01/09/92 01/09/92
01/04/93 01/04/93 01/04/93
Observed
01/11/93 01/11/93 01/11/93
01/06/94 01/06/94 01/06/94
01/01/95 01/01/95 01/01/95
01/08/95 01/08/95 01/08/95
01/03/96 01/03/96 01/03/96
01/10/96 01/10/96 01/10/96
01/05/97 01/05/97 01/05/97
01/12/97 01/12/97 01/12/97
01/07/98 01/07/98 01/07/98
01/02/99 01/02/99 01/02/99
01/09/99 01/09/99 01/09/99
01/04/00 01/04/00 01/04/00
One-step-ahead Training
01/11/00 01/11/00 01/11/00
01/06/01 01/06/01 01/06/01
01/01/02 01/01/02 01/01/02
01/08/02 01/08/02 01/08/02
01/03/03 01/03/03 01/03/03
01/10/03 01/10/03 01/10/03
01/05/04 01/05/04 01/05/04
01/12/04 01/12/04 01/12/04
01/07/05 01/07/05 01/07/05
.
01/02/06 01/02/06 01/02/06
.
.
01/09/06 01/09/06 01/09/06
Figure 23 (cont.)
Two-step-ahead
01/02/13 01/02/13 01/02/13
01/09/13 01/09/13 01/09/13
01/04/14 01/04/14 01/04/14
01/11/14 01/11/14 01/11/14
01/06/15 01/06/15 01/06/15
01/01/16 01/01/16 01/01/16
01/08/16 01/08/16 01/08/16
01/03/17 01/03/17 01/03/17
COUNTRIES’ INFLATION RATE FORECAST FOR THE NEXT SIX MONTHS BY NAR-NN
Out of sample
01/10/17 01/10/17 01/10/17
1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
1
2
3
4
5
6
7
8
9
01/07/91 01/07/91 01/07/91
01/02/92 01/02/92 01/02/92
01/09/92 01/09/92 01/09/92
01/04/93 01/04/93 01/04/93
Observed
01/11/93 01/11/93 01/11/93
01/06/94 01/06/94 01/06/94
01/01/95 01/01/95 01/01/95
01/08/95 01/08/95 01/08/95
01/03/96 01/03/96 01/03/96
01/10/96 01/10/96 01/10/96
01/05/97 01/05/97 01/05/97
01/12/97 01/12/97 01/12/97
01/07/98 01/07/98 01/07/98
01/02/99 01/02/99 01/02/99
01/09/99 01/09/99 01/09/99
01/04/00 01/04/00 01/04/00
01/11/00
One-step-ahead Training
01/11/00 01/11/00
01/06/01 01/06/01 01/06/01
01/01/02 01/01/02 01/01/02
01/08/02 01/08/02 01/08/02
01/03/03 01/03/03 01/03/03
01/10/03 01/10/03 01/10/03
01/05/04 01/05/04 01/05/04
01/12/04 01/12/04 01/12/04
01/07/05 01/07/05
.
01/07/05
.
01/02/06 01/02/06 . 01/02/06
01/09/06 01/09/06 01/09/06
Figure 23 (cont.)
One-step-ahead Testing
01/11/07 01/11/07 01/11/07
01/06/08 01/06/08 01/06/08
01/01/09 01/01/09 01/01/09
01/08/09 01/08/09 01/08/09
01/03/10 01/03/10 01/03/10
01/10/10 01/10/10 01/10/10
01/05/11 01/05/11 01/05/11
01/12/11 01/12/11 01/12/11
01/07/12 01/07/12 01/07/12
Two-step-ahead
01/02/13 01/02/13 01/02/13
01/09/13 01/09/13 01/09/13
01/04/14 01/04/14 01/04/14
01/11/14 01/11/14 01/11/14
01/06/15 01/06/15 01/06/15
01/01/16 01/01/16 01/01/16
397
01/08/16 01/08/16 01/08/16
01/03/17 01/03/17 01/03/17
COUNTRIES’ INFLATION RATE FORECAST FOR THE NEXT SIX MONTHS BY NAR-NN
Out of sample
01/10/17 01/10/17 01/10/17
B.4 arima
In the arima modeling, various tests were performed before mod-
eling the series in order to understand the generating data process
and find the best (p,d,q)(P,D,Q) order suit to the series. We began
to perform the Augmented Dickey-Fuller (ADF) test (see Dickey
and Fuller, 1981) and the Kwiatkowski, Phillips, Schmidt, and Shin
(KPSS) test (see Kwiatkowski et al., 1992 to find the differentiation
order (Table 16). In the Dickey-Fuller test, we started including
the trend and constant over the regression for which all the series
rejected the null hypothesis of the unit root. For the KPSS test, like
the ADF test, we included the trend and constant terms and almost
all the series did not reject the null hypothesis of stationary except
for Switzerland and Norway, where the Switzerland series became
stationary after the first 8 observations were excluded from the tests.
To find the seasonal difference order, the Canova-Hansen test (see
Canova and Hansen, 1995) was implemented, which has a null hy-
pothesis of no unit roots at seasonal frequencies. This test comple-
ments the HEGGY test of seasonal unit roots.
Once the difference orders were determined and the respective
transformations were applied, such as applying logarithms if nec-
essary, we proceed to explore the autocorrelation function, partial
autocorrelation, extended autocorrelation function, and informa-
tion criterion AIC and BIC. We used these factors to find the autore-
gressive and moving average coefficients. A group of possible models
were tested on each country, for which the most suitable model had to
accomplish five conditions:
• Low BIC, AICc, and rmse
• coefficients statistically different to zero.
• the residuals should be uncorrelated through time.
• the cross-correlation function between the predicted errors
and the observed time series should be close to zero.
KPSS
ADF t-Stat Stat (p,d,q)(P,D,Q) order
Brazil 5.871 0.089 (1,0,0)
*Log transformation
399
References
Box, G.E.P. et al. 2016. Time Series Analysis: Forecasting and Control.
New York, United States: Wiley.
Canova, F., and B.E. Hansen. 1995. “Are Seasonal Patterns Constant
over Time? A Test for Seasonal Stability.” Journal of Business
Economic Statistics 13(3): 237-252. doi:10.1080/07350015
.1995.10524598.
Carvalho, C., and F. Nechio. 2014. “Do People Understand Monetary
Policy?” Journal of Monetary Economics 66: 108-123.
Claveria, O., E. Monte and S. Torra. 2016. “A Self-Organizing Map
Analysis of Survey-Based Agents’ Expectations before Im-
pending Shocks for Model Selection: The Case of the 2008
Financial Crisis.” International Economics 146: 40-58. doi:
10.1016/j.inteco.2015.11.003.
Coibion, O. 2012. “Are the Effects of Monetary Policy Shocks Big or
Small?” American Economic Journal: Macroeconomics 4(2): 1-32.
Crump, R.K. et al. 2015. “Subjective Intertemporal Substitution.”
Federal Reserve Bank of New York Staff Reports 734. New
York, United States: Federal Reserve Bank of New York.
Dickey, D.A, and W.A. Fuller. 1981. “Likelihood Ratio Statistics for
Autoregressive Time Series with a Unit Root.” Econometrica
49(4): 10-57. doi:10.2307/1912517.
Fildes, R., and H. Stekler. 2002. “The State of Macroeconom-
ic Forecasting.” Journal of Macroeconomics 24(4): 435-468.
doi:10.1016/s0164-0704(02)00055-1.
Hagan, M.T., H.B. Demuth and M.H. Beale. 2002. Neural Network
Toolbox: User’s Guide. Natick, United States: The MathWorks,
Inc.
Hagan, M.T. et al. 2014. Neural Network Design. Second edition.
Boulder, United States: University of Colorado.
Hamella, S., and H. Haupt. 2007. “Suitability of wes Data for
Forecasting Inflation.” In: G. Goldrian, editor. Handbook
of Survey-Based Business Cycle Analysis. Cheltenham, United
Kingdom: Edward Elgar.
401
Titterington, D.M. 2004. “Bayesian Methods for Neu-
ral Networks.” Statistical Science 19(1): 128–139.
doi:10.1214/088342304000000099.
Wehrens, R., and L. Buydens. 2007. “Self- and Super-Organising
Maps in R: The Kohonen Package.” Journal of Statistical Software
21 (5). URL http://www.jstatsoft.org/v21/i05
White, H. 1989. “Learning in Artificial Neural Networks: A Statistical
Perspective.” Neural Computation 1 (4): 425-464. doi:10.1162/
neco.1989.1.4.425.
Yuriy, Y., and C. Olivier. 2015. “Is the Phillips Curve Alive and
Well after All? Inflation Expectations and the Missing Dis-
inflation.” American Economic Journal: Macroeconomics 7(1):
197-232. https://ideas.repec.org/a/aea/aejmac/
v7y2015i1p197- 232.html
Zhang, G., B.E. Patuwo and M.Y. Hu. 1998. “Forecasting with Arti-
ficial Neural Networks: “The State of the Art.” International
Journal of Forecasting 14(1): 35–62. doi: http://doi.org/10.
1016/S0169-2070(97)00044-7. URL http://www.sciencedi-
rect.com/science/article/pii/S0169207097000447.
Abstract
In the 1990s, after experiencing high levels of inflation, several countries
in Latin America passed constitutional amendments providing greater
autonomy to their central banks. A few years later, many central banks
increased their exchange rate flexibility and later adopted inflation targeting
frameworks. These institutional changes coincided with sharp reductions
in inflation and its variability. In this paper, we ask if the observed reduction
of inflation is possibly related to changes in monetary policy. To answer this
question, we build and estimate a Markov-Switching Dsge model for an
open economy with monetary factors for Brazil, Chile, Colombia, Mexico,
and Peru, all of whom formally adopted inflation targeting regimes between
1999 and 2002. Regimes are classified according to their relative weights
of inflation in an interest rate reaction function. Although ex-ante these
regimes need not be associated with the introduction of the inflation targeting
framework, the coincidence of a regime switch with a more responsive interest
rate - inflation relationship is striking. Furthermore, the Markov-Switching
Dsge model allows us to generate counterfactuals of what could have
happened if the observed change towards a more aggressive fight against
inflation had not taken place. In general, we observe that if monetary
policy had remained dovish, these countries would have experienced higher
and more variable levels of inflation and more pronounced variations in Gdp
with small gains in average economic growth. Therefore, we conclude that
The authors thank Junior Maih for making his rise toolbox for the solution and esti-
mation of Markov Switching Rational Expectations models available and for patiently
answering all of our questions. The views expressed in this presentation are those
of the author, and not necessarily those of cemla or egade Business School of Tec-
nológico de Monterrey.
405
the introduction of inflation targeting represented a favorable regime switch
in the implementation of monetary policy in Latin America.
Keywords: Monetary policy, inflation, Markov-switching Dsge, Bayesian
Maximum Likelihood methods.
jel: E31, E37, E52, E58, C11.
1. INTRODUCTION
B
eginning in the late 1980s, many countries around the world en-
acted new central banking legislation to grant more autonomy
to their monetary authorities. For example, see Figure 1,
which uses a sample of indexes of central bank independence from
182 countries since 1970, produced by Garriga (2016). Figure 1 shows
a sharp increase in the number of reforms toward increased central
bank independence in the 1990s. This shift came in response to the
traumatic inflationary and hyper-inflationary episodes experienced
in the previous decades, and it was reinforced by evidence showing
that “central bank independence promotes price stability” without
“measurable impact on real economic performance” (e.g., Alesina
and Summers (1993)).
In Latin America, starting with Venezuela in 1974, several coun-
tries had reforms to strengthen the independence of their central
banks1. In some countries, and for different reasons (from depletion
of reserves to the desire to gain greater control of monetary policy),
many central banks increased their exchange rate flexibility. The pro-
cess continued with the adoption of inflation targeting frameworks
to direct monetary policy. These institutional changes coincided with
1
According to Garriga (2016), since 1970, countries that took positive
reforms towards independence were the following: Venezuela in 1974;
Chile in 1975; Haiti in 1979; Mexico in 1985; Brazil in 1988; Chile
in 1989; El Salvador in 1991; Argentina, Colombia, Ecuador, Nica-
ragua, Peru, and Venezuela in 1992; Mexico in 1993; Bolivia, Costa
Rica, Paraguay, and Uruguay in 1995; Honduras in 1996; Cuba in 1997;
Nicaragua and Venezuela in 1999; El Salvador in 2000; Guatemala
and the Dominican Republic in 2002; and Uruguay in 2008 and 2010.
Meanwhile, negative reforms hindering Central Bank independence
include the following: Argentina and El Salvador in 1973, Panama
in 1975, El Salvador in 1982, Uruguay in 1997, Venezuela in 2001,
Argentina in 2003, Ecuador in 2008, Venezuela in 2009, Nicaragua
in 2010, and Argentina in 2012.
15
10
Number of reforms
−5
Year of
Positive Inflation
reforms Exchange Targeting
Average 1980- 1990- 2000- 2010- towards rate introduc-
inflation 1989 1999 2009 2015 independence flexibility tion
Brazil 121.7 147.1 6.6 6.2 1988 1999 1999
Chile 19.9 11.8 3.5 3 1975 and 1999 1999
1989
Colombia 20.8 19.9 6.1 3.1 1992 1999 1999
Mexico 53.1 18.3 5.1 3.6 1985 and 1995 2001
1993
Peru 111 78.5 2.6 3 1992 2002 2002
2. MODEL
2.1
2.2
2.3
2
The equation reduces to the closed economy variant when α = 0.
2.4
where i s a wo r l d i n f l a t io n s h o c k w h i c h i s t r e a t e d
as an unobservable and is assumed to follow an exogenous process:
Terms of trade, in turn, are as-
sumed to follow a law of motion for their growth rate:
2.5
The dsge system with constant parameters has the following ma-
trix form:
3.1
′
3
with X t = ⎡⎢ yt πt Rt Δqt Δet πt* yt*at ⎥⎤ .
⎣ ⎦
3.2
3.3
where stands for the parameters of the model, Ytobs are the ob-
served variables,4 and M provides the policy function for the observ-
ables. Following Bianchi and Ilut (2017), we introduce the possibility
of regime change for the structural parameters and the volatilities
through two Markov chains, ξ sp and ξvo . The former denotes the un-
observed regime associated with the monetary parameters subject
to regime shifts and takes on discrete values 5
and the latter
stands for the shock volatilities, assumes discrete values, 6
3.4
4
gdp growth, inflation rate, interest rate, change in the terms of trade
and nominal depreciation.
5
Where 1 and 2 are the high and low response to inflation regimes
⎛ ⎞
⎜⎜i.e ψ sp=1 > ψ sp=2 ⎟⎟, respectively.
⎝ π,ξt π,ξt ⎟⎠
6
Where 1 and 2 are the low and high volatility regimes. In order to define
the high volatility regime, we included into the model the following
restriction: σa,ξ vol=1 < σa,ξ vol=2 .
3.5
3.6
7
( ) ⎛
⎝ q,ξt a,ξt R,ξt y ,ξt π ,ξt ⎟
⎞
Where: Σ ξ vo = diag ⎜⎜σ vo ,σ vo ,σ vo ,σ * vo ,σ * vo ⎟⎟.
⎠
8
The routines used for the computations were implemented using rise,
an object-oriented Matlab toolbox for solving and estimating Markov
switching rational expectation models, developed by Junior Maih.
3.8
where:
• We use the posterior mode as the initial value for the Metrop-
olis Hasting algorithm, with 100.000 iterations.
9
rise toolbox optimization routine.
4. RESULTS
100
50
0
1996 1998 2001 2004 2007 2009 2012 2016
50
0
1996 1998 2001 2004 2007 2009 2012 2016
Figure 3
SMOOTHED PROBABILITIES FOR CHILE
100
50
0
1998 2000 2002 2004 2006 2008 2010 2012 2014 2016
50
0
1998 2000 2002 2004 2006 2008 2010 2012 2014 2016
100
50
0
1995 1998 2001 2004 2007 2009 2012 2015
50
0
1995 1998 2001 2004 2007 2009 2012 2015
Figure 5
SMOOTHED PROBABILITIES FOR MEXICO
100
50
0
1981 1984 1988 1992 1996 2000 2003 2007 2011 2016
50
0
1981 1984 1988 1992 1996 2000 2003 2007 2011 2016
100
50
0
1995 1998 2001 2004 2007 2009 2012 2015
50
0
1995 1998 2001 2004 2007 2009 2012 2015
Country
Parameter Distribution Brazil Chile Colombia Mexico Peru
Beta 0.1738 0.2053 0.7092 0.8564 0.1318
2 0.5
Output growth
0 0
−2 −0.5
−4 −1
2 4 6 8 10 12 2 4 6 8 10 12
5 0.5
0 0
Inflation
−5 −0.5
−10 −1
2 4 6 8 10 12 2 4 6 8 10 12
4 0.5
Interest rate
2
0
0
−2 −0.5
2 4 6 8 10 12 2 4 6 8 10 12
10 1
Real interest rate
5
0.5
0
−5 0
2 4 6 8 10 12 2 4 6 8 10 12
5 0.5
∆ Exchange rate
0 0
−5 −0.5
−10 −1
2 4 6 8 10 12 2 4 6 8 10 12
0.2 1
Output growth
0 0
−0.2 −1
−0.4 −2
2 4 6 8 10 12 2 4 6 8 10 12
1 5
0 0
Inflation
−1 −5
−2 −10
2 4 6 8 10 12 2 4 6 8 10 12
0.4 2
Interest rate
0.2 1
0 0
−0.2 −1
2 4 6 8 10 12 2 4 6 8 10 12
2 10
Real interest rate
1 5
0 0
−1 −5
2 4 6 8 10 12 2 4 6 8 10 12
1 5
∆ Exchange rate
0 0
−1 −5
−2 −10
2 4 6 8 10 12 2 4 6 8 10 12
0.5
Output growth
0
−0.5
−1
2 4 6 8 10 12
0
Inflation
−1
−2
2 4 6 8 10 12
2
Interest rate
0
2 4 6 8 10 12
4
Real interest rate
0
2 4 6 8 10 12
0
∆ Exchange rate
−1
−2
2 4 6 8 10 12
1 0.2
Output growth
0.5 0.1
0 0
−0.5 −0.1
2 4 6 8 10 12 2 4 6 8 10 12
2 0.1
1 0.05
Inflation
0 0
−1 −0.05
2 4 6 8 10 12 2 4 6 8 10 12
1 0.1
Interest rate
0.5 0.05
0 0
2 4 6 8 10 12 2 4 6 8 10 12
1 0.05
Real interest rate
0.5
0
0
−0.5 −0.05
2 4 6 8 10 12 2 4 6 8 10 12
2 0.1
∆ Exchange rate
1 0.05
0 0
−1 −0.05
2 4 6 8 10 12 2 4 6 8 10 12
0.2 0.2
Output growth
0.1 0.1
0 0
−0.1 −0.1
2 4 6 8 10 12 2 4 6 8 10 12
1 1
0.5 0.5
Inflation
0 0
−0.5 −0.5
2 4 6 8 10 12 2 4 6 8 10 12
0.4 0.4
Interest rate
0.2 0.2
0 0
−0.2 −0.2
2 4 6 8 10 12 2 4 6 8 10 12
0.5 0.2
Real interest rate
0
0
−0.2
−0.5 −0.4
2 4 6 8 10 12 2 4 6 8 10 12
1 1
∆ Exchange rate
0.5 0.5
0 0
−0.5 −0.5
2 4 6 8 10 12 2 4 6 8 10 12
0.4
Output growth
0.2
−0.2
2 4 6 8 10 12
0.5
Inflation
−0.5
2 4 6 8 10 12
0.3
Interest rate
0.2
0.1
0
2 4 6 8 10 12
0.2
Real interest rate
−0.2
2 4 6 8 10 12
0.5
∆ Exchange rate
−0.5
2 4 6 8 10 12
0.1 0.2
Output growth
0.05 0.1
0 0
2 4 6 8 10 12 2 4 6 8 10 12
0.1 0.2
0 0
Inflation
−0.1 −0.2
−0.2 −0.4
2 4 6 8 10 12 2 4 6 8 10 12
0.1 0.2
Interest rate
0 0
−0.1 −0.2
−0.2 −0.4
2 4 6 8 10 12 2 4 6 8 10 12
0.1 0.2
Real interest rate
0
0
−0.1
−0.2 −0.2
2 4 6 8 10 12 2 4 6 8 10 12
0 5
∆ Exchange rate
−2 0
−4 −5
−6 −10
2 4 6 8 10 12 2 4 6 8 10 12
0.4 0.15
Output growth
0.2 0.1
0 0.05
−0.2 0
2 4 6 8 10 12 2 4 6 8 10 12
0.2 0.4
0.2
Inflation
0
0
−0.2 −0.2
2 4 6 8 10 12 2 4 6 8 10 12
0.2 0
Interest rate
0
−0.2
−0.2
−0.4 −0.4
2 4 6 8 10 12 2 4 6 8 10 12
0.2 0
Real interest rate
0 −0.2
−0.2 −0.4
−0.4 −0.6
2 4 6 8 10 12 2 4 6 8 10 12
5 0
∆ Exchange rate
−1
0
−2
−5 −3
2 4 6 8 10 12 2 4 6 8 10 12
0.15
Output growth
0.1
0.05
0
2 4 6 8 10 12
0.2
0.1
Inflation
−0.1
2 4 6 8 10 12
0
Interest rate
−0.1
−0.2
−0.3
2 4 6 8 10 12
0
Real interest rate
−0.1
−0.2
−0.3
2 4 6 8 10 12
5
∆ Exchange rate
−5
2 4 6 8 10 12
0 0
Output growth
−0.5
−1
−1
−1.5 −2
2 4 6 8 10 12 2 4 6 8 10 12
3 3
2 2
Inflation
1 1
0 0
2 4 6 8 10 12 2 4 6 8 10 12
1.5 2
Interest rate
1
1
0.5
0 0
2 4 6 8 10 12 2 4 6 8 10 12
1 1
Real interest rate
0 0
−1 −1
−2 −2
2 4 6 8 10 12 2 4 6 8 10 12
3 3
∆ Exchange rate
−2 2
1 1
0 0
2 4 6 8 10 12 2 4 6 8 10 12
0 0
Output growth
−0.5
−5
−1
−1.5 −10
2 4 6 8 10 12 2 4 6 8 10 12
6 30
4 20
Inflation
2 10
0 0
2 4 6 8 10 12 2 4 6 8 10 12
1.5 10
Interest rate
1
5
0.5
0 0
2 4 6 8 10 12 2 4 6 8 10 12
2 10
Real interest rate
0 0
−2 −10
−4 −20
2 4 6 8 10 12 2 4 6 8 10 12
6 30
∆ Exchange rate
4 20
2 10
0 0
2 4 6 8 10 12 2 4 6 8 10 12
0
Output growth
−0.5
−1
−1.5
2 4 6 8 10 12
1.5
1
Inflation
0.5
0
2 4 6 8 10 12
1
Interest rate
0.5
0
2 4 6 8 10 12
0.5
Real interest rate
−0.5
−1
2 4 6 8 10 12
1.5
∆ Exchange rate
0.5
0
2 4 6 8 10 12
0.05 0.1
Output growth
0.05
0
0
−0.05 −0.05
2 4 6 8 10 12 2 4 6 8 10 12
0.15 0.1
0.1
Inflation
0.05
0.05
0 0
2 4 6 8 10 12 2 4 6 8 10 12
0.01 0.1
Interest rate
0 0
−0.01 −0.1
−0.02 −0.2
2 4 6 8 10 12 2 4 6 8 10 12
0 0.1
Real interest rate
−0.05 0
−0.1 −0.1
−0.15 −0.2
2 4 6 8 10 12 2 4 6 8 10 12
0 5
∆ Exchange rate
−2
0
−4
−6 −5
2 4 6 8 10 12 2 4 6 8 10 12
0.04 1
Output growth
0.02 0.5
0 0
−0.02 −0.5
2 4 6 8 10 12 2 4 6 8 10 12
0.1 6
0.05 4
Inflation
0 2
−0.05 0
2 4 6 8 10 12 2 4 6 8 10 12
0.02 1.5
Interest rate
0 1
−0.02 0.5
−0.04 0
2 4 6 8 10 12 2 4 6 8 10 12
0.05 2
Real interest rate
0 0
−0.05 −2
−0.1 −4
2 4 6 8 10 12 2 4 6 8 10 12
2 0
∆ Exchange rate
0 −2
−2 −4
−4 −6
2 4 6 8 10 12 2 4 6 8 10 12
0.2
Output growth
0.1
0
2 4 6 8 10 12
0.4
Inflation
0.2
0
2 4 6 8 10 12
0.1
Interest rate
−0.1
−0.2
2 4 6 8 10 12
0
Real interest rate
−0.2
−0.4
2 4 6 8 10 12
0
∆ Exchange rate
−2
−4
−6
2 4 6 8 10 12
Observed 0.64 1.26 1.97 3.85 4.21 1.09 3.44 4.27 1.11 2.26 5.73 2.53 4.65 3.31 0.71
Output High 0.99 3.28 3.30 3.75 2.78 0.74 3.42 4.11 1.22 1.77 4.84 2.73 4.97 2.74 0.55
growth response
Low 1.00 3.63 3.62 3.65 4.75 1.30 3.37 4.47 1.31 3.46 8.85 2.56 5.38 5.71 1.06
response
Observed 6.31 3.72 0.59 3.06 2.51 0.82 9.84 7.43 1.12 20.15 24.78 1.23 3.62 3.18 0.88
Inflation High 3.89 2.73 0.70 2.93 1.90 0.65 7.13 4.93 1.04 11.08 7.65 0.69 3.34 2.12 0.64
response
Low 15.73 5.80 0.37 3.14 3.88 1.23 17.13 12.26 0.89 26.83 23.69 0.88 6.89 8.68 1.26
response
Observed 16.49 7.00 0.42 4.59 2.04 0.44 12.56 10.50 2.33 25.38 26.36 1.04 6.91 6.03 0.87
Interest High 10.69 3.84 0.36 4.76 1.44 0.30 9.01 6.41 1.61 18.90 13.06 0.69 5.43 2.61 0.48
rate response
Low 12.59 4.17 0.33 4.76 3.02 0.63 12.58 8.48 1.57 30.81 27.84 0.90 4.42 5.26 1.19
response
Real High 6.80 3.49 0.51 1.83 2.04 1.11 1.89 2.80 3.64 7.82 7.94 1.01 2.09 2.61 1.25
interest response
rate
Low −3.14 7.36 2.34 1.63 1.48 0.91 −4.56 8.94 1.07 3.98 15.61 3.92 −2.47 13.33 5.40
response
Observed 1.64 9.13 5.58 0.67 4.87 7.25 1.36 6.60 0.89 −0.59 4.81 8.10 0.45 2.70 6.02
Nominal High −0.60 8.61 14.26 0.72 2.86 4.00 −3.15 8.44 1.34 −5.59 9.92 1.77 0.25 3.29 13.40
depreciation response
Low 11.24 9.23 0.82 1.28 7.50 5.88 6.33 9.15 0.77 5.74 14.88 2.59 3.77 8.15 2.16
response
Figure 12
COUNTERFACTUAL FOR HIGH AND LOW RESPONSE REGIMES FOR BRAZIL
10
−10
2000 2005 2010 2015
10
30
20
10
0
2000 2005 2010 2015
30
20
10
0
−10
10
−10
10
5
0
0.2
0.1
0
2000 2005 2010 2015
20
10
0
−10
2000 2005 2010 2015
10
5
0
−5
10
30
20
10
0
1995 2000 2005 2010
−20
20
−20
20
−20
100
50
0
1985 1990 1995 2000 2005 2010 2015
100
50
0
1985 1990 1995 2000 2005 2010 2015
60
40
20
0
−20
−40
15
10
5
0
−5
1990 2005 2010 2015
100
50
50
0
1990 2005 2010 2015
5
0
−5
A. Estimated Parameters
Table 4
ESTIMATED PARAMETERS OF BRAZIL
462
Kim, C.-J., Nelson, C. R., et al. (1999). State-Space Models with Regime
Switching: Classical and Gibbs-Sampling Approaches with Applica-
tions. mit Press Books, 1.
Liu, P. and Mumtaz, H. (2011). Evolving Macroeconomic Dynamics
in a Small Open Economy: An estimated Markov Switching
dsge Model for the uk. Journal of Money Credit and Banking,
43(7):1443-1474.
Lubik, T. A. and Schorfheide, F. (2007). Do Central Banks Respond
to Exchange Rate Movements? A Structural Investigation.
Journal of Monetary Economics, 54(4):1069-1087.
Maih, J. (2015). Efficient Perturbation Methods for Solving Regime-Switch-
ing dsge Models. Working Paper 2015/01, Norges Bank.
Ortiz, A. and Sturzenegger, F. (2007). Estimating Sarb’s Policy
Reaction Rule. South African Journal of Economics, 75(4):659-
680.
Schmitt-Grohé, S. and Uribe, M. (2003). Closing Small Open
Economy Models. Journal of International Economics, 61(1):163-
185.
Sims, C. A. (2002). Solving Linear Rational Expectations Models.
Computational Economics, 20(1):1-20.
Sims, C. A. and Zha, T. (2006). Were There Regime Switches in us
Monetary Policy?, The American Economic Review, 96(1):54-81.
Abstract
We estimate a hidden Markov model where inflation is determined by gov-
ernment deficits financed through money creation and/or by destabilizing
expectations dynamics (expectations can potentially divorce inflation from
fundamentals). The baseline model, proposed by Sargent et al. (2009), is used
to analyze the interaction between fiscal deficits, inflation expectations,
and inflation in Mexico. The model is able to distinguish between causes
and remedies of hyperinflation, such as persistent or transitory shocks to sei-
gniorage-financed fiscal deficits, de-anchoring of inflation expectations from
fiscal fundamentals, and cosmetic (non-fundamental) monetary reforms.
The behavior of monetized deficits provides an adequate account of high in-
flation episodes and stabilizations for the period 1969-1994. We then extend
This research was developed within the framework of cemla’s Joint Research Program
2017 coordinated by the Central Bank of Colombia. We thank counselling and techni-
cal advisory provided by the Financial Stability and Development (FSD) Group of the
Inter-American Development Bank in the process of writing this document. We are
especially grateful to Daniel Chiquiar and our academic advisor Andrea Tambalotti,
for their feedback during this project. We would like to thank, for thoughtful com-
ments and suggestions, Nicolás Amoroso, Oscar Budar Mejía, Julio Carrillo, Alfonso
Cebreros, Gerardo Hernández del Valle, Juan Ramón Hernández, Raul Ibarra, Emil-
iano Luttini, André Martínez-Fritscher, Felipe Meza, Alberto Ortiz-Bolaños, Fernando
Pérez-Cervantes, Manuel Ramos-Francia, Juan Sherwell, our discussant Christian
Velásquez, and the participants of the cemla Research Network and the XXII An-
nual Meeting of the Central Bank Research Network (Bogota, 2017). The opinions
expressed in this publication are those of the authors and do not reflect the views
of cemla, the FSD Group, the Inter-American Development Bank, or those of Banco
de México. E-mails: [email protected], [email protected],
and [email protected].
465
the model to analyze the possibility that fiscal policy can affect inflation expec-
tations in a context of Central Bank independence, as is the case of Mexico
after 1994. We find evidence that the exchange rate and sovereign interest
rate spreads influence the evolution of aggregate prices.
Keywords: inflation, inflation expectations, fiscal policy.
jel: E31, E42, E52, E63.
1.INTRODUCTION
1
Fischer et al. (2002), Catao and Terrones (2005), and Lin and Chu
(2013), among others, document international evidence regarding
the relationship between ination rates, scal decits, and money supply.
Rogers and Wang (1994) estimate that between 1977 and 1990, scal
and monetary shocks accounted for 60 percent of the variance of ina-
tion in Mexico.
2
Agents have adaptive expectations or backward-looking expectations
when these are formed by extrapolating past values of the variable be-
ing predicted.
467
inflation and inflation expectations even in the context of Central
Bank independence?
A strand of the macroeconomic literature proposes that fiscal
policy is relevant to achieving price stability even in an environ-
ment where monetary policy is conducted by an independent Cen-
tral Bank. 3 We extend the baseline model along several dimensions
with the objective of documenting evidence, perhaps indirect, or re-
butting the possibility that fiscal policy is relevant in determining
inflation and inflation expectations in a context of Central Bank in-
dependence. A variable of interest we consider is the spread in the
sovereign interest rate embi. This variable, which can be considered
forward-looking, reflects the fiscal situation of the government.
To the extent that economic agents perceive potential risks in terms
of the ability of the government to make debt payments, it may also
affect the credibility of the Central Bank. The perception of this
type of risk is incorporated in the prices of sovereign debt. The state
of public finances is often considered to affect the exchange rate;
this is the second variable we assess in the model. The results indi-
cate that both variables are relevant in determining inflation expec-
tations and inflation.4
We proceed as follows: Section 2 presents the baseline model
and describes the mechanisms that drive the behavior of the different
variables. Section 3 presents the main results for the baseline model:
(1) the parameter values of the model and their implications in terms
3
There exists a vast literature studying the relevance of fiscal policy
and its interaction with monetary policy for the determination of infla-
tion, a seminal paper is Sargent and Wallace (1981). Though we will
not attempt to provide an exhaustive set of references, some addi-
tional examples are provided by Sims (2016), Leeper (1991), Davig
et al. (2011), Sargent and Zeira (2011), Woodford (2001), and Bianchi
and Ilut (2017). For an introductory treatment of the fiscal theory
of the price level, see Christiano and Fitzgerald (2000). Central Banks
frequently express concern related to how fiscal imbalances may affect
the effectiveness of monetary policy (e.g., Carstens and Jácome (2005)
and Ramos-Francia and Torres-Garcia (2005)).
4
There are different mechanisms through which these variables could
potentially be relevant; we explore the impact through expectations
and the demand for real money balances. We discuss the evidence
of the extent to which these variables are influenced by international
and exogenous factors, with a focus on the case of Mexico, such as prices
of commodities in global markets.
The baseline model is the one featured in Sargent et al. (2009), con-
structed to study the relationship between inflation, fiscal deficits,
and inflation expectations. An ad- vantage of this model is its sim-
ple structure, which allows for the estimation of its parameters us-
ing only the historic series of one of the main variables, in our case
the monthly inflation series (the estimation algorithm is described
briefly in the next section and in the Appendix). With these param-
eters, the model accounts for an observed sequence of inflation as a
result of fiscal deficits and a particular process for the formation
of inflation expectations. The framework consists of three main
components: a money demand function, the budget constraint of the
government, a process that models the formation of expectations,
and the (exogenous and stochastic) evolution of deficits. We now de-
scribe each of these components.
5
In a seminal paper, Cagan (1956) specifies a demand for real balances
and backward-looking expectations to explain several European hyper-
inflation episodes.
469
where represents the weight that the expected price level
has on the current price level P t , and is the weight that
the nominal balances relative to output have on the price level at time
t.6 Thus, if the public expects a higher price level in t + 1, their real
balances demand Mt/P t will fall.
The next equation represents the budget constraint of the gov-
ernment, where dt (a stochastic variable) is the part of the real defi-
cit of the government that is monetized (net of debt emissions, so it
must be covered by printing money). Thus, the growth of nominal
balances per unit of output is determined according to the follow-
ing equation:
9
This is obtained with λ∈(0,1), θ∈(0,1), and γ>0.
10
For example, Branch (2004) develops a micro-founded model where
agents optimally choose not to update their beliefs according to a
rational expectations algorithm because the information it requires
is too costly (rational expectations algorithms usually require a lot
of information). In the type of models we are considering, adaptive
471
Assuming constant-gain expectations (cge) is key in determining
the dynamics of the model. Panel (a) of Figure 1 shows the change
in gross inflation πt+1−π t as a function of expectations βt , with a con-
stant real fiscal deficit. As shown in the Figure, there are two values
of β that imply a constant inflation equilibrium: β1 and β 2 . In the
adaptive expectations literature, β1 and β 2 are known as self-confirm-
ing equilibria. As implied by the Figure, β1 is a locally stable equi-
librium, thus, if the beliefs of the public regarding future inflation
are not sufficiently high then πt+1−πt will converge to zero and βt+1
to β1. Additionally, equation (4) implies that πt will also converge
to β1. However, if βt > β 2, then πt+1−πt will increase, with unbounded
dynamics. Therefore, βt>β 2 implies that the model will eventually
generate a hyperinflation episode. This phenomenon is called es-
cape dynamics by Sargent et al. (2009).11
Panel (b) of Figure 1 presents another result of cge: assuming
β t induces escape dynamics, a hyperinflation episode can be pre-
vented if the deficit is reduced. This Panel shows two dynamic paths
for πt+1−π t as a function of βt . The only difference between these paths
is the level of fiscal deficit. The dynamics shown in blue correspond
to a high fiscal deficit, while the dynamics in green correspond to a
low fiscal deficit. Assuming a high deficit and βt = βˆ , if the deficit
is not reduced then it will provoke an escape dynamics of inflation
and expectations as shown with blue arrows in Panel (b) of Figure 1.
However, if the government reduces its fiscal deficit to a sufficiently
low level then, even when βt = βˆ , it will be able to prevent an escape
dynamics. Furthermore, πt+1−π t will converge to a low and stable in-
flation equilibrium as shown by the green arrows in the Figure.
Finally, cge implies a non-trivial computational advantage: given
the complexity of the function that will be used to estimate all the pa-
rameters involved in the model, assuming this type of expectations
80
60
40
πt+1−πt
20
−20
0 β1 5 βt 10 15 β2
200 High dt
150
100
πt+1−πt
Low dt
50
β2
0
β1
−50
0 β15 βt 10 15 β̂
Note: these figures considers βt−1 =1.02 and the estimated parameters shown in Table 1.
473
allows us to reduce the computational burden.12 We discuss the im-
plications of using rational expectations in the Appendix.
12
The next section explains some of the details involved in estimating
the parameters of model.
13
A stochastic process xt is said to be a discrete Markov process if xt takes
values in a set I with and for all the Markov property
is satisfied: This property states
that past realizations of the process do not affect future
values, only the present state xt affects xt+1.
14
Formally, a hidden Markov process is a pair {x t, y t} such that x t is
a (standard) Markov process and there exists a function f such that
for all and:
475
Figure 2
NON-LINEAR EFFECT OF FISCAL DEFICITS
.
—
d1
—
d2
60
40
—
πt
d3
20
0 5 10 15 20 25
βt
.
40
— —
d1 d2
βt+1
20
—
d3
0
0 5 10 15 20 25
βt
Note: these figures consider βt−1 =1.02 and the estimated parameters as described in
the next section.
16
17
Certainty in the model implies In equilibrium, Using
477
generate an escape dynamics, the government can implement a re-
form to prevent this from happening. Sargent et al. (2009) define
two types of reforms: a reform is said to be cosmetic if the govern-
ment is able to (temporarily) control inflation but the median level
of fiscal deficit is not altered. Following Panel (a) of Figure 1, a cos-
metic reform can fail if the expected inflation rate associated with
inflation βt+1 is such that βt+1>β 2. However, a cosmetic reform can be
successful if 18
A structural reform, on the other hand, oc-
curs when the government is able to control the inflation rate by re-
ducing the median level of fiscal deficit, Panel (b) of Figure 1 is
an example of a structural reform where the government succeeded
in controlling an escape dynamics.
An important contribution of the model is its ability to identi-
fy whether a reform is cosmetic or structural. Previous literature
had only studied structural reforms, even though the notion of a
cosmetic reform was part of academic and economic policy discus-
sions. The inclusion of cosmetic reforms in the model represents
a reduced form approach to consider different episodes in Latin
America, when governments attempted to control inflation without
tackling fiscal deficits. Discussions of eco- nomic events often point
to the role of the exchange rate, which is not explicitly included in the
baseline model, and we explore below through different extensions
of the baseline model.
18
Sargent et al. (2009) argue that in Peru a cosmetic reform was enough
to control the inflationary crisis this country experienced in 1985.
19
In the Appendix we provide further details regarding the estimation
of the model. Ramirez de Aguilar (2017) describes the computational
procedure.
20
The sic is a Bayesian selection criterion between two models, A and B.
Let Lx, Px, nx be the log-likelihood, the number of parameters, and the
sample size in model x ∈ {A, B}, respectively. Then, the Schwarz crite-
rion for model x is computed as SICx=log(nx)Px−2Lx. If SICA<SICB, then
model A is preferred.
21
The estimation of ν=0.1147 implies that the weight agents give to their
past expectations is 0.8853. Hence, if inflation is stable for only
479
Table 1
PARAMETER ESTIMATION
Note: the numbers shown in parentheses represent the standard deviation of each parameter,
computed using the Hessian matrix of the maximum likelihood problem (see MacDonald
and Zuccini (2009)).
one month, this will not be enough to reduce β because past beliefs
have more weight on expectations. Only if the inflation rate is stable
for several consecutive months will β also become stable.
481
of median deficit This is consistent with the economic his-
tory of Mexico; during the decade of the 1960s, the inflation
rate in Mexico achieved its lowest value during the second half
of the twentieth century: an average of 2.8%, which is repli-
cated by the model.22
22
In this section we draw from Cardenas (2015), who provides an exhaus-
tive narrative of the economic history of Mexico during the period
of our analysis. Historical series for output and the inflation rate
data presented in this section were obtained in the Historic Statistics
of Mexico published by inegi.
23
Cardenas (2015) argues that the crisis presented during 1987 is a
direct consequence of the unwillingness of the government to reduce
its deficit during 1982-1987.
483
Tesobonos), which required a stable exchange rate in order
to keep this debt sustainable. However, political events led to
a significant depreciation of the domestic currency in 1994
accompanied by capital outflows (Calvo and Mendoza (1996),
Cole and Kehoe (1996) analyze these events). The government
faced a debt crisis, the private financial sector found itself
in bankruptcy, and the inflation rate reached 51% in 1995.
The government negotiated loans with the International
Monetary Fund (imf) and with the United States in order to fi-
nance its debt.
24
Some of the policies adopted by the Central Bank after 1994 were: (i)
restoration of the level of international reserves to gain credibility, (ii)
the use of an objective of cumulative current account balances that
private banks held at the Central Bank as the primary monetary policy
instrument, (iii) adoption of an inflation-targeting policy, and (iv)
to improve transparency, the Central Bank began to publish reports
communicating monetary policy decisions as well as quarterly reports
of the economy. For a more detailed description of these policies
see Ramos-Francia and Torres- Garcia (2005).
25
Furthermore, as documented by Chiquiar et al. (2010), the inflation
rate after 2000-2001 became a stationary process and initiated its con-
vergence towards the inflation target.
• During the last sub-period (Region (6) of Figure 3), the model
predicts that fiscal deficits were at the lowest median and vari-
ance hidden states. The model also shows that the expected
inflation rate has fluctuated within the range of the target
of Banco de México: an inf lation rate of 3% that can vary
between 2% and 4%. The model proposes that a necessary
condition to anchor inflation and its expectations is a low mon-
etization of fiscal deficit. The only year in which the fiscal
deficit had a slight probability of being at a higher median
state was in 2009, in the course of the global financial crisis.
However, since the inflation rate remained low after 2009,
the baseline model predicts that Mexico has remained in a
low fiscal deficit regime.
Considering the inflation history previously described, we ob-
serve that the model predicts a deficit distribution with an elevated
mean and variance during those years in which the inflation rate
was elevated, as in 1987 (a year characterized by the highest infla-
tion rate presented in Mexico during the second half of the twentieth
century). In those years in which the inflation rate was moderately
high, as in 1975, the model predicts a fiscal deficit with a moderate
mean and lower variance than in 1987. Finally, in those years where
the inflation rate is low, the fiscal deficit density is characterized by a
low mean and variance.
26
In 2008 there was a methodological modification in bpt, it became
a wider measure of fiscal deficits: after 2008 the bpt considers part
of the investments made by two important state- owned firms (pemex
and cfe) that before were considered as long-term debt (this type
of investments are called pidiregas).
27
The bpt does not consider the revenue and expenditures of Banco
de México or the public financial sector. The financial sector of the
485
Figure 3
DYNAMICS OF THE MODEL
(3)
10
0
1970
1973
1976
1979
1982
1985
1988
1991
1994
1997
2000
2003
2006
2009
2012
2015
Year
150
Data
100
50 Model
0
1969
1972
1975
1978
1981
1984
1987
1990
1993
1996
1999
2002
2005
2008
2011
2014
2017
Year
Notes: Panel (a) plots the median real scal deficit relative to output together with the
10th and 90th percentile of the annual deficit distribution. Panel (b) shows the annual
inflation rate predicted by the model given the real scal deficit, and the data. Panel
(c) shows the expected inflation rate according to the algorithm (4). Panel (d)
plots P [dt = d2|πt, φ] + P[dt = d3|πt, φ] where d2 and d3 are the moderate and low levels
of mean fiscal deficit.
Source: and model results.
150
100
50
0
1969
1972
1975
1978
1981
1984
1987
1990
1993
1996
1999
2002
2005
2008
2011
2014
2017
Year
.
%
100
60
40
20
(2)
0
1969
1972
1975
1978
1981
1984
1987
1990
1993
1996
1999
2002
2005
2008
2011
2014
2017
Year
Notes: Panel (a) plots the median real scal deficit relative to output together with the
10th and 90th percentile of the annual deficit distribution. Panel (b) shows the annual
inflation rate predicted by the model given the real scal deficit, and the data. Panel
(c) shows the expected inflation rate according to the algorithm (4). Panel (d)
plots P [dt = d2|πt, φ] + P[dt = d3|πt, φ] where d2 and d3 are the moderate and low levels
of mean fiscal deficit.
Source: and model results.
487
computes an alternative fiscal deficit measure called Requerimien-
tos Financieros del Sector Público (rfsp), which incorporates the fi-
nancial requirements of the government at the federal level. This is a
broader measure of fiscal deficit since it includes the bpt in addition
to all revenues and expenditures of the public financial sector that
provide funds for public policy.28
Panel (a) of Figure 4 displays the estimated sequence of fiscal defi-
cits from the model, as well as the bpt and the rfsp relative to gdp be-
tween 1977 and 2016. As shown in the Figure, there is an adequate
approximation of the model to the bpt data before 1991 and to the rfsp
after 1993. During 1991 and 1992, both series show a fiscal surplus.
The model cannot match this feature of the data given the assumption
of a log-normal distribution, and deficits cannot be negative. Addi-
tion- ally, the model predicts a higher deficit during 1994-1996 rela-
tive to those observed in the data; in 1995 the model predicts a fiscal
deficit relative to output of 6.1% of gdp, while the Rfsp exhibits a fiscal
deficit of 2.5% of gdp. The baseline model can only attribute the spike
in inflation of that year to fiscal deficits. We will see that the exten-
sions of this model can better account for the rates of inflation during
this episode. During 1977-2016, the model’s median deficit variance
is 53.7% of the variance presented in the fiscal deficit data.29
Panel (b) of Figure4 displays the model’s implied monetary base
growth rate compared with Banco de México’s data between 1969
and 1970. 30 The Figure shows that the model approximates the data’s
31
As explained by Meza (2017), the Central Bank transfers resources
to the Ministry of Finance (equivalent to the Treasury in the U.s.),
after determining its earnings and following legally specified rules.
This is called the Remanente de Operación de Banco de México. In the
United States, the Federal Reserve transfers to the Treasury most of its
interest earnings from government debt. As further discussed below,
this can be perfectly consistent with a regime of monetary dominance.
32
In this sense, the approach is complementary to models that consider
regime-switching environments, e.g. Chung et al. (2007), Cadavid-
Sanchez et al. (2017), and Bianchi and Ilut (2017).
33
For the period, Meza (2017) estimates seigniorage at an average of 0.66
p.p. of gdp for the period 1995-2016.
489
Figure 4
DATA AND MODEL COMPARISON
6 Model RFSP
3
P10
0
−3
BPT
−6
1977
1980
1983
1986
1989
1992
1995
1998
2001
2004
2007
2010
2013
2016
Year
Data
100
50
Model
−50
1969
1972
1975
1978
1981
1984
1987
1990
1993
1996
1999
2002
2005
2008
2011
2014
2016
Year
Notes: the series presented are - Panel (a): in blue the estimated scal deficit with the
10th/ 90th percentiles of the estimated deficit distribution. In red/orange the
or the relative to . Panel (b): In blue/red the model/data monetary base
annual growth rate, respectively.
Source: Banco de México and .
34
We have explored additional extensions of the model. For example,
incorporating the cetes interest rate, and another specification that
includes the target for the inflation rate of Banco de México. However,
the fit of these alternative specifications is less favorable (results avail-
able upon request). Further exploration of alternative specifications
would certainly be an interesting topic for future research.
491
Empirical evidence shows that sovereign interest rate spreads
are, to a large extent, driven by international factors such as risk
appetite, market volatility, terms of trade, global liquidity, conta-
gion from events such as the Russian crisis or the LTCM collapse
in 1998, and even U.s. macroeconomic news. 35 In the same fashion,
exchange rate fluctuations are linked to global financial factors (to
give some recent examples, Gabaix and Maggiori (2015) and Itskhoki
and Mukhin (2017)), and the Mexican peso is sometimes considered
a commodity currency (see Kohlscheen (2010)). The state of public ac-
counts can make the economy vulnerable to these external shocks. 36
Our model allows us to explore empirically the possibility that
fiscal policy can make the evolution of inflation sensitive to events
in international financial markets. The results motivate the need
for further theoretical developments in this area, in particular for de-
veloping economies, where sovereign interest rate spreads and ex-
change rates seem to be of primary relevance. The historical narrative
of events in Mexico for the period 1969-1994 supports this interpre-
tation; events such as significant drops in the price of oil or sudden-
stops make the economy vulnerable when fiscal accounts are in a dire
situation and the government may be forced to turn to the Central
Bank to cover its financial needs. Even in a context of de jure mon-
etary dominance, economic agents may consider that these risks
are still present, and thus we aim to capture this possibility in the
estimation of our model. 37
35
There is an extensive literature that documents these facts, including
Longstaff et al. (2011), González-Rozada and Levy-Yeyati (2008), Bunda
et al. (2009), Ciarlone et al. (2009), Hilscher and Nosbusch (2010),
and Ozatay et al. (2009).
36
The issue of endogeneity is addressed by exploiting alternative method-
ologies in Cortés-Espada (2013) and Lopez-Villavicencio and Mignon
(2016).
37
These channels have been considered by Zoli (2005) in the case of Bra-
zil, by assessing the impact of news concerning fiscal variables and fis-
cal policy on sovereign interest rate spreads and the exchange rate
and discussing the potential implications for monetary policy. Cerisola
and Gelos (2005) find that the stance of fiscal policy (proxied by the
ratio of the consolidated primary surplus to gdp) is important to deter-
mine inflation expectations in the case of Brazil and argue that fiscal
policy is instrumental in anchoring inflation expectations.
100
Floating Exchange Rate
50
−50
1977
1980
1983
1986
1989
1992
1995
1998
2001
2004
2007
2010
2013
2016
Source: Banco de México and .
38
In the Appendix, we describe the different exchange rate regimes
in Mexico.
493
that this variation has a weight ξ on expectations. Hence, for each pe-
riod t, the expected inflation rate is determined as follows:
Given that Mexico had a fixed ner during 1969-1994 and after 1995
the ner is in a floating regime, we estimate the model allowing ξ to
change during 1969-1994 and 1995-2016. Hence, the model allows agents
to give a weight ξ 1 to the ner variation during a fixed exchange rate re-
gime and a weight ξ 2 when the ner is in a floating regime. To estimate
this model, again we consider the monthly inflation sequence accord-
ing to the inpc between January of 1969 and December of 2016, and the
sequence of the monthly variation in the peso-dollar ner documented
by Banco de México for that period. Table 2 presents the estimated pa-
rameters of this version compared with the baseline model estimation.
Considering the exchange rate as a variable that can influence inflation
expectations (and hence, inflation), the model can account for 75.8%
of the variance observed in the inflation data, while the baseline model
can explain 61.6% of this variance. Also, as suggested by the Diebold-
Mariano test, during 2000-2016 the ner and baseline models produce
different in-sample forecasts of observed inflation (at a 1% significance
level) and the modified model has a higher correlation with the infla-
tion data. 39 This result emphasizes the relevance of the exchange rate
for the determination of the inflation rate in Mexico.40
39
The hypothesis test proposed in Diebold and Mariano (1995) allows to assess
Defining ekt=ykt−yt for k ∈ {i, j} and considering a loss-function g(e), the null
of the forecasts and show that, if the time series considered are covariance
stationary and short memory, it has a t-Student distribution. Then, they
construct a statistic that, under the same assumptions, is asymptotically
N (0, 1).
40
More formally, according to the sic comparison, the ordering of the models
is the following: the model with the embi spread and the ner in the for-
mation of expectations, the model with the ner in the real balances
demand function (presented in the following section), the model with
only the ner in the formation of expectations and, finally, the base-
line model.
495
Figure 6
IMPULSE-RESPONSE FUNCTIONS OF INFLATION
1
Difference with Equilibrium Inflation (%)
0.6
0.4
0.2
−
0.026 p.p. High d (1982-1987)
0
−
Low d (1995-2016)
−0.2
−0.4
0 10 20 30
Time
section, a higher fiscal deficit magnifies the effect that βt has on infla-
tion (in fact this effect is nonlinear). Hence, if fiscal deficit increas-
es, the effect that the ner variation has on πt will grow because this
variation affects βt . It can be shown that:
∂πt ∂π ∂βt λξ
9 = t = πt .
∂∆NER ∂βt ∂∆NER 1 − λβt − dt
This equation highlights two important results: (i) the erpt is in-
creasing in dt; (ii) a higher inflation rate implies a higher erpt. Figure
6 shows the impulse-response function of inflation given a 1% depre-
ciation in the ner. As this Figure suggests, when fiscal deficit is high
(e.g., during 1982-1987) the erpt to inflation is 0.821 p.p. However,
if fiscal deficit is low the erpt of a 1% depreciation is 0.026 p.p. Hence,
a low fiscal deficit financed by the Central Bank not only translates
into low inflation, but also into a limited erpt. A low pass-through
41
The low level of pass-through is consistent with estimates in the lit-
erature for Mexico, see Albagli et al. (2015), Capistrán et al. (2011),
Cortés-Espada (2013), and Kochen and Samano (2016). Furthermore,
there is evidence of a declining erpt in environments with more stable
inflation and with the adoption of inflation targets (see Baqueiro
et al. (2003), Choudhri and Hakura (2006) and Lopez-Villavicencio
and Mignon (2016)). Capistrán et al. (2011) and Cortés-Espada (2013)
document a lower erpt for Mexico under the inflation targeting regime.
42
The perception of economic agents of the fiscal responsibility of the
government may depend on the particular historical context. For ex-
ample, Sargent and Zeira (2011) describe how the anticipation of a
future government bailout of banks caused a jump in inflation in Israel
in 1983. They argue that the public anticipated that this bailout would
eventually be financed by monetary expansion. Alternatively, Chung
et al. (2007) explore an environment where monetary and fiscal re-
gimes evolve according to a Markov process, this possibility can change
the impact of policy shocks. These authors argue that, to the extent that
there has been a history of changes in policy regimes, private agents
can ascribe a probability distribution over the different regimes.
497
Figure 7
INFLATION AND EXPECTATIONS IN THE NER MODEL
Inflation Data
150
Average Inflation
100 2006-2016:
Data: 3.91%
Baseline Model Model: 3.56%
ner Model
50 Baseline Model: 3.15%
0
1969
1972
1975
1978
1981
1984
1987
1990
1993
1996
1999
2002
2005
2008
2011
2014
2017
Year
.
%
200
150
Average Expected
100 ner Model Inflation
2006-2016:
Model: 3.37%
50 Baseline Model: 3.15%
Baseline Model
0
1969
1972
1975
1978
1981
1984
1987
1990
1993
1996
1999
2002
2005
2008
2011
2014
2017
Year
Source: .
10
11
We allow the parameters {ν, ξ} to vary because the ner had a change
in its regime.
If parameters ξ and σ are positive and statistically significant,
it would imply that the embi spread and the ner influence inflation.
In fact, these variables can generate the escape dynamics that in the
baseline model could only be ignited by the behavior of fiscal defi-
cits.43 Figure 9 exemplifies how an escape dynamics, that leads to high
or hyperinflation, can occur in this scenario: suppose that initially
βt =β ∗ and that ∆NER t , ∆EMBIt are limited. This implies that infla-
tion and its expectations will converge to a low inflation equilibrium
as the blue arrows show. However, if the fiscal authority starts to con-
siderably increase its deficit (which is no longer financed with money
creation and is therefore translated into debt) this would be reflected
43
In the baseline model, an escape dynamics can only occur if fiscal deficit
increases for a considerable period, because it is the only way to raise
inflation expectations.
499
Figure 8
EMBI, NER AND INFLATION
20
380 embi ner
18
330
16
280
14
230
12
180
130 10
80 8
1999
2001
2003
2005
2007
2009
2011
2013
2015
2017
Year
2001
2003
2005
2007
2009
2011
2013
2015
2017
Year
0.8
0.6
0.4
πt+1−πt
∆NER>>0
0.2
∆EMBI>>0
β∗
0
−0.2
β2
−0.4
0 5 10 15
βt
Note: these figures considers βt−1 =1.02 and the estimated parameters of the
extension.
in the embi spread and influence the ner. In our model, the incre-
ment in these variables will affect inflation expectations. Further-
more, if this effect is large enough, as shown with an orange arrow
in the Figure, it will cause that βt>β 2 , which will lead to high infla-
tion (as shown with red arrows). Consequently, if σ and ξ are signifi-
cant and positive then, even in a context of monetary dominance,
our model suggests the possibility of high inflation caused by the fis-
cal authority via expectations.
To estimate this model, once again we consider the inflation se-
quence according to the INPC during 1969-2016, the ner varia-
tion registered by Banco de México, and the embi spread reported
by Bloomberg after 1994. The main results of this extension are:
501
• The estimation for σ suggests that, everything else constant,
if the embi spread increases 100 basis points, the rate of infla-
tion rises by 0.24 p.p.44
12
44
To find the impact that the embi spread has on inflation, we again have
to consider an impulse-response function as in Figure 6.
45
Alternatively, this expression can be rewritten as a demand for real
balances that depends on the exchange rate.
8
Data Average: 3.91%
Data and Model Average: 3.80%
Baseline Model Average: 3.15%
Annual Inflation Rate (%)
Baseline Model
2 embi and ner Model
bm Inflation Target
0
2006
2007
2008
2009
2010
2011
2012
2013
2014
2015
2016
2017
Year
503
Figure 11
INFLATION IN THE EXTENDED MODEL
. 1970-2016
200
Data
150
Annual Inflation (%)
Extension
100
50
Baseline Model
0
1969
1972
1975
1978
1981
1984
1987
1990
1993
1996
1999
2002
2005
2008
2011
2014
2017
Year
. 2006-2016
Data
Data Average: 3.91%
6 Extension Average: 4.01%
Annual Inflation (%)
4
ner in rbd Extension
3%
2
Banco de México’s Extension Target
0
2006
2007
2008
2009
2010
2011
2012
2013
2014
2015
2016
2017
Year
The baseline model and the extensions that we have presented allow
us to assess the role of fiscal policy in the determination of inflation
and its expectations. Even in a context of Central Bank indepen-
dence, a large literature has explored the role of fiscal policy in de-
termining inflation. We exploit a simple model and provide evidence
of the relevance of fiscal policy in determining the behavior of ag-
gregate prices in Mexico as well as the importance of expectations.
Admittedly, the theoretical framework we utilize is relatively sim-
ple and models with more structure, perhaps in the inter-temporal
dimension, would increase our understanding of the relationship
between fiscal policy and inflation in emerging economies. Further-
more, it is sometimes argued that Central Bank independence acts
as a mechanism that increases fiscal responsibility of the govern-
ment in developing countries (Bodea and Higashijima (2015), Minea
and Tapsoba (2014)). We believe that further research is necessary
to understand the institutional arrangements that govern the rela-
tionship between a central bank and the fiscal authority in the pres-
ence of competing objectives and constraints.
6. APPENDIX
505
where Xt is a constant equal to 1 if {πt, βt.dt} satisfy 1−λβt−1>0 and δ(1−
λβt−γdt)>θ(1−λβt−1).46 Assuming β0 =π 0 and given a sequence of fiscal
deficits the model can generate a sequence for the expected
inflation rate and for the actual inflation rate How-
ever, the hidden Markov states among other parameters, must
be estimated to generate a sequence of fiscal deficits. Table 3 shows
the parameters that need to be estimated.
Table 3
MODEL PARAMETERS
46
These constraints guarantee that the model’s inflation rate is bounded
and that the real balances demand is positive.
13
where Ω is the set of all the vectors φ that satisfy the constraints rel-
evant for each parameter. Because there is no analytical solution
to this maximization problem, has to be approximated numeri-
cally. To do this, we used a constrained optimization algorithm based
on the bfgs (Broyden-Fletcher-Goldfarb-Shanno) method of Nocedal
and Wright (2006) and the block-wise method of Sims et al. (2006).
Given the computational burden of the maximum-likelihood
optimization problem, Sargent et al. (2009) fix three parameters
to reduce the complexity on the estimation. These parameters are:
θ = 0.99, δ = 100, and γ = 1. The value assigned to θ is consistent with
the behavior of nominal balances in the five countries these authors
studied. Fixing δ = 100 implies that, in every period, inflation cannot
surpass 10,000%. Finally, γ was fixed because the maximum-likeli-
hood algorithm cannot identify γ and dt separately. Once dt is esti-
mated for each period, γ is re-normalized so that the mean of fiscal
507
deficits estimated by the model matches the mean observed in the
data (in our case, for Mexico for the period 1977-2016).
14
β1
0
πt+1−πt
−5
−10
−15
−5 0 5 10 15
βt
–
10 High dt
β2
0
β1
πt+1−πt
−10
–
Low dt
β
−20
−5 0 5 10 15
Note: These figures considers βt−1 =1.02 and the estimated parameters shown in Table 1.
509
and its expectations to β 2. However, if the government reduces its fis-
cal deficits, it will change the dynamics on inflation and its expecta-
tions inducing a convergence to β1.
Figure 12 points out an important difference between rational
expectations and cge: when agents use the cge algorithm, if the in-
flation rate induces a high βt then this could provoke an escape dy-
namics and eventually a hyperinflation episode, where the dynamics
between inflation and its expectations are unbounded. However,
with rational expectations, even with an extremely high fiscal deficit,
agents always adapt their expectations to prevent a hyperinflation
spiral. If fiscal deficit is high, rational expectations imply a stable
equilibrium with a high inflation rate and no escapes.
Even though cge and rational expectations induce different dy-
namics on the variables involved in the model, the inflation equi-
libria they predict are similar. Sargent et al. (2009) argue that, in the
context of hyperinflation models, “an adaptive expectations version
of the model shares steady states with the rational expectations ver-
sion, but has more plausible out-of-steady state dynamics.” Besides,
rational expectations may induce multiple equilibria that are hard
to compute. Given the computational problem rational expectations
may induce and the fact that some Latin American countries have
experienced hyperinflation episodes with escape dynamics which
a strictly rational expectations model cannot account for, cge are
necessary for the purposes of this study.
Date ner
Begining End Regime Begining End
April 1954 August 1976 Fixed 12.50 12.50
September 1976 August 1982 Controlled 20.50 48.80
Variation
September 1982 December 1982 Generalized 50.00 70.00
Controlled
System
December 1982 August 1985 Controlled 95.00 281.00
System
August 1985 November 1991 Controlled 282.30 3,073.00
Flotation
November 1991 December 1994 Floating 3,074.10 N 3.99
Intervals
with
Controlled
Variation
December 1994 December 2016 Floating N 4.88 N 20.51
References
Albagli, E., A. Naudon, and R. Vergara (2015): “Inflation Dynamics
in LATAM: A Comparison with Global Trends and Implica-
tions for Monetary Policy,” Economic Policy Papers Central
Bank of Chile 58, Central Bank of Chile.
Baqueiro, A., A. Díaz De León, and A. Torres-García (2003): “Temor
a La Flotación o a La Inflación? La Importancia del “Traspaso
del Tipo De Cambio A Los Precios,” Working Papers 2003-
02, Banco de México.
Bianchi, F. and C. Ilut (2017): “Monetary/Fiscal Policy Mix and
Agent’s Beliefs,” Review of Economic Dynamics, 26, 113-139.
511
Bodea, C. and M. Higashijima (2015): “Central Bank Independence
and Fiscal Policy: Can the Central Bank Restrain Deficit
Spending?” British Journal of Political Science, 1, 1-24.
Branch, W. A. (2004): “The Theory of Rationally Heterogeneous
Expectations: Evidence from Survey Data on Inflation Ex-
pectations,” Economic Journal, 114, 592-621.
Bunda, L., A. J. Hamann, and S. Lall (2009): “Correlations in
emerging market bonds: The role of local and global fac-
tors,” Emerging Markets Review, 10, 67-96.
Cadavid-Sanchez, S., A. Martinez-Fritscher, and A. Ürtiz-Bolaños
(2017): “Monetary and Fiscal Policies Interactions in Mexico:
1981-2016,” Working paper, cemla.
Cagan, P. (1956): “The Monetary Dynamics of Hyperinflation,”
in Studies in the Quantity Theory of Money, Chicago University
Press, Milton Friedman ed.
Calvo, G. A. and E. G. Mendoza (1996): “Mexico’s balance-of-
payments crisis: a chronicle of death foretold,” Journal of
International Economics, 41, 235-264.
Capistrán, C., R. Ibarra-Ramírez, and M. Ramos-Francia (2011): “Ex-
change Rate Pass-Through to Prices: Evidence from Mexico,”
Working Papers 2011-12, Banco de México.
Cardenas, E. (2015): El Largo Curso de la Economía Mexicana: De
1780 a Nuestros Días, Ciudad de Mexico: Fondo de Cultura
Economica.
Carrasco, C. A. and J. Ferreiro (2013): “Inflation targeting and
inflation expectations in Mexico,” Applied Economics, 45,
3295-3304.
Carstens, A. and L. L. Jácome (2005): “Latin American Central
Bank Reform; Progress and Challenges,” IMF Working Papers
05/114, International Monetary Fund.
Catao, L. A. and M. E. Terrones (2005): “Fiscal deficits and infla-
tion,” Journal of Monetary Economics, 52, 529-554.
Cerisola, M. and G. Gelos (2005): “What Drives Inflation Expecta-
tions in Brazil? An Empirical Analysis,” imf Working Papers,
05, 1.
Chiquiar, D., A. Noriega, and M. Ramos-Francia (2010): “A time-
series approach to test a change in inflation persistence:
the Mexican experience,” Applied Economics, 42, 3067-3075.
513
Kochen, F. and D. Samano (2016): “Price-Setting and Exchange
Rate Pass- Through in the Mexican Economy: Evidence from
CPI Micro Data,” Working Papers 2016-13, Banco de México.
Kocherlakota, N. (2012): “Central Bank Independence and Sover-
eign Default,” Bank du France, Financial Stability Review, 16.
Kohlscheen, E. (2010): “Emerging floaters: Pass-throughs and
(some) new commodity currencies,” Journal of International
Money and Finance, 29, 1580-1595.
Leeper, E. M. (1991): “Equilibria under “active” and “passive”
monetary and fiscal policies,” Journal of Monetary Economics,
27, 129 - 147.
Lin, H.-Y. and H.-P. Chu (2013): “Are fiscal deficits inflationary?”
Journal of International Money and Finance, 32, 214-233.
Longstaff, F. A., J. Pan, L. H. Pedersen, and K. J. Singleton (2011):
“How Sovereign Is Sovereign Credit Risk?” American Economic
Journal: Macroeconomics, 3, 75-103.
Lopez-Villavicencio, A. and V. Mignon (2016): “Exchange Rate
Pass-through in Emerging Countries: Do the Inflation Envi-
ronment, Monetary Policy Regime and Institutional Quality
Matter?” Working Papers 2016-07, CEPII Research Center.
Macdonald, L. and W. Zuccini (2009): Hidden Markov Models for
Time Series, New York: CRC Press, 2nd ed.
Marcet, A. and J. P. Nicolini (2003): “Recurrent Hyperinflations
and Learning,” American Economic Review, 93, 1476-1498.
Meza, F. (2017): “Mexico from the 1960s to the 21st Century: From
Fiscal Dominance to Debt Crisis to Low Inflation,” Working
papers, itam.
Minea, A. and R. Tapsoba (2014): “Does inflation targeting improve
fiscal discipline?” Journal of International Money and Finance,
40, 185-203.
Nocedal, J. and S. J. Wright (2006): Numerical Optimization, New
York: Springer, 2nd ed.
Ozatay, F., E. Ozmen, and G. Sahinbeyoglu (2009): “Emerging
market sovereign spreads, global financial conditions and
U.s. macroeconomic news,” Economic Modelling, 26, 526 - 531.
Ramirez De Aguilar, A. (2017): “Mexico’s Recent Inflation History
as a Result of Fiscal Deficit and its Expectations,” B. A. thesis,
Banco de México and itam.
515
CENTER FOR LATIN AMERICAN MONETARY STUDIES
www.cemla.org