View original document

The full text on this page is automatically extracted from the file linked above and may contain errors and inconsistencies.

History and Theory
of the NAIRU:
A Critical Review
M A R C O A . E S P I N O S A - V E G A
A N D S T E V E N R U S S E L L
Espinosa is a senior economist at the Federal Reserve
Bank of Atlanta. Russell is an assistant professor of
economics at Indiana University-Purdue University
at Indianapolis. They thank, without implicating,
Eric Leeper, Maurice Obstfeld, Mary Rosenbaum, and
Charles Whiteman for preliminary conversations on
the topic and Bob Eisenbeis and Robert Bliss for helpful
expositional comments on the final draft.

W

HAT CAUSES CHANGES IN THE RATES OF INFLATION AND UNEMPLOYMENT?
THE PRICE LEVEL AND THE LEVEL OF EMPLOYMENT RELATED?

HOW

ARE

THESE HAVE BEEN KEY

QUESTIONS FACING ECONOMISTS FOR AT LEAST FORTY YEARS.

DISCUSSIONS

ABOUT

THEM IN THE PRESS AND ELSEWHERE OFTEN CENTER ON AN APPROACH TO EXPLAINING

THE INFLATION-UNEMPLOYMENT RELATIONSHIP THAT DATES BACK TO THE

1960S

AND

1970S. ACCORDING

TO THIS APPROACH, INFLATION IS CAUSED BY AN EXCESSIVELY TIGHT LABOR MARKET THAT DRIVES UP WAGES
AND FORCES FIRMS TO RESPOND BY RAISING PRICES.

An important element of this approach is the concept of a nonaccelerating inflation rate of unemployment,
or NAIRU. As its name suggests, the NAIRU is supposed to
be an unemployment rate (or range of unemployment
rates) that produces a stable rate of inflation: if the
unemployment rate is lower than the NAIRU then the
inflation rate will tend to rise, and vice versa.
Recently, both the NAIRU and the theory of the inflationunemployment relationship on which it is based have
received a great deal of attention from the press. From
December 1995 to December 1996, for example, there were
ten articles on this subject in the Wall Street Journal, five
articles in the New York Times, and three in The
International Economy. One common feature of all these
4

articles is that they link Federal Reserve monetary policy
to the NAIRU. Most of the authors seem to assume that the
NAIRU is or should be the Fed’s principal guide for conducting monetary policy. According to this view, if the current unemployment rate is below some NAIRU estimate
(say, 6 percent) then the Fed should tighten monetary policy to head off a coming increase in the inflation rate.
Despite the extensive press coverage the NAIRU concept has received recently, the theory of the inflationunemployment relationship that it is part of is quite
controversial. Although the NAIRU is alive and well in the
media and among economic policymakers, it is no longer
very popular among academic economists. It has fallen
out of favor partly because its conceptual foundation is

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

weak and partly because its empirical track record does
not inspire confidence. Its survival is due largely to the
fact that economists have not been able to reach any consensus about alternative guides for monetary policy.
The purpose of this article is to provide some historical perspective on the “NAIRU theory” and the assumptions behind it. Most of the analysis presented in this
article is not original: it has been around for two decades
or more. However, the recent resurgence of interest in the
NAIRU indicates that there may be a need for a basic
review of its origins and a brief explanation of some of the
claims surrounding it. Readers interested in additional
details should consult the reference list.
The first section of the discussion that follows briefly
introduces the Keynesian and classical theories of macroeconomics. Keynesian theory is the macroeconomic theory on which the NAIRU is principally based while classical
theory provides the foundation for the monetarist and
neoclassical critiques of Keynesian theory that are discussed at length in this article. As we shall see, the concept of a NAIRU grew out of economists’ attempts to
reconcile the differences between Keynesian and monetarist theories on the subjects of the causes of price level
changes and the relationship between inflation and unemployment. The next section discusses the Phillips curve, a
description of the inflation-unemployment relationship
that provided the empirical and theoretical starting points
for the development of the NAIRU. The third section
reviews the monetarist critique of analysis based on the
Phillips curve and discusses a number of related questions. The next two sections explain how the NAIRU developed as a response to the monetarist critique of the
Phillips curve and raise some basic questions about the
NAIRU. The final part of the discussion reviews the concept of rational expectations, a theoretical contribution of
neoclassical theory that amplified the monetarist critique
of the Phillips curve. This section also discusses some neoclassical contributions that may offer alternatives to the
Phillips curve approach to the study of inflation, unemployment, and the effects of monetary policy.

Two Economic Traditions
lassical economic theory developed in the early
1900s, at a time when there was no formal distinction between micro- and macroeconomics. The
theory was based on the same basic assumptions that had
become widely used to study the behavior of individual
households and firms. These included the assumptions
that individuals usually act in ways that maximize their

C

self-interest, that prices are determined in the marketplace, and that markets operate efficiently. According to
classical theory, perfect competition is a good approximation of the operation of most real-life markets. The
basic assumptions of classical theory are generally understood to imply that government policies have relatively
little importance in determining economic outcomes.
Keynesian theory, which developed in the 1930s and
1940s, was the first macroeconomic theory: it was
designed specifically to study economywide phenomena,
and it was not simply an extension of the conventional
economic theory that
continued to be used to
study the behavior of
individual parts and
The NAIRU has fallen out
sectors of the economy.
of favor among academic
Keynesian theory was
based on the work of
economists partly
John Maynard Keynes,
because its conceptual
a British economist
foundation is weak and
who did most of his
work in the 1920s and
partly because its empiri1930s. One of the basic
cal track record does not
goals of Keynes’s theoinspire confidence.
ry was to explain the
persistently high rates
of unemployment that
appeared across the
world during the Great Depression. Most of this unemployment was generally believed to be “involuntary,” in the
sense that the unemployed people were willing to work at
the going wage rates but were unable to find jobs. A closely related goal of Keynes was to identify steps that the government could take to alleviate the high levels of
unemployment.
Keynesian theory assumes that some important
prices are determined or strongly influenced by forces
outside the marketplace so that many markets may not
be able to “clear” in the sense of successfully reconciling
demand with supply. It also assumes that people may not
always make the economic decisions that would be best
for them. According to Keynesian theory, perfect competition is not a good approximation of the operation of
many important real-life markets. The theory implies that
government policies can have large, important effects on
the economy and that if the policies are carefully devised
these effects can be very constructive in nature.1
Keynes’s ideas and goals placed him in direct conflict with the exponents of the reigning classical theory.

1.The monetarist and neoclassical theories developed later—monetarism in the 1950s and neoclassical theory in the 1970s. These
theories were developed as alternatives to Keynesian theory, which was then accepted by most contemporary economists. Both
theories drew heavily on the classical tradition. As we shall see, the economic theory behind the NAIRU is basically Keynesian
in nature, but it has been influenced heavily by monetarist ideas and to a lesser extent by neoclassical ones.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

5

Classical theory predicted that when unemployment was
high wages would adjust downward, stimulating more
hiring and reducing the unemployment rate. As a result,
high unemployment could not last long. It seemed obvious to Keynes (and many others) that the high, persistent levels of unemployment observed during the
Depression were inconsistent with this prediction and
that classical theory was incapable of explaining them. In
1933 prominent classical theorist A.C. Pigou published
The Theory of Unemployment; according to Keynes, this
book was “the only detailed account of the classical theory of employment” in existence at the time. In his “General Theory” article, Keynes dismisses Pigou’s book as “a
non-causative investigation into the functional relationship which determines
what level of real
wages will correspond
Classical economic theory
to any given level of
employment. . . . [It] is
was based on the assumpnot capable of telling
tions that individuals
us what determines
usually act in ways that
the actual level of
employment; and on
maximize their selfthe problem of involinterest, that prices are
untary unemployment
determined in the marketit has no direct bearing” (1964, 275).
place, and that markets
According to
operate efficiently.
K e y n e s , what prevented labor markets
from clearing, and
explained involuntary unemployment, was that when
firms’ demand for labor decreased, nominal (money)
wages did not fall as fast or as far as classical theory predicted.2 “Classical theory,” he comments, “has been accustomed to rest the supposedly self-adjusting character of
the economic system on an assumed fluidity of moneywages” (1964, 257). Keynes believed that sluggish labor
demand would not push nominal wages downward, at
least in the short run. The logic behind this belief was
that organized workers had enough market power to
resist employers’ attempts to reduce money wage rates.
As a result, Keynesian theory is often described as being
based on the assumption of “sticky wages.”3 In the classical model, unlike the Keynesian model, money wages and
prices are assumed to be perfectly flexible, so labor markets always clear. If temporary unemployment appears
because of deficient aggregate demand, then the unemployed workers will bid down nominal wages until they
have fallen far enough to eliminate the unemployment.
Keynes also criticized classical theory for failing to
provide an integrated analysis of the behavior of different
parts of the economy and for making an unwarranted
leap from analysis of individual-industry labor markets to
analysis of the determinants of aggregate employment.
6

He writes that “if the classical theory is not allowed to
extend by analogy its conclusions in respect of a particular industry to industry as a whole, it is wholly unable to
answer the question what effect on employment a reduction in money-wages will have. For it has no method of
analysis wherewith to tackle the problem” (1964, 257).
Over time, it became clear that both classical and
Keynesian theories suffered from some important deficiencies. Classical theorists needed to integrate their
microeconomic theories of individual labor markets into
a macroeconomic theory of total employment. They also
needed to explain how government policies affected the
labor market. The Keynesians needed to move in the
opposite direction, integrating their macroeconomic theory with a microeconomic theory of labor markets and
formalizing their explanation of wage-setting behavior.

The Phillips Curve
nflation and Unemployment. In 1958 British economist A.W. Phillips published the results of an empirical analysis of historical data from the U.K. labor
market. Phillips’s study was intended to help answer one
of the basic questions in macroeconomic theory, which
concerns the cause of inflation. He hoped to find empirical support for the Keynesian view that the rate of wage
inflation—that is, the rate of increase in nominal
(money) wage rates—depended on the tightness of the
labor market. Since the level of unemployment was a
readily observable indicator of the tightness of the labor
market, Phillips’s immediate goal was “to see whether
statistical evidence supports the hypothesis that the rate
of change of money wage rates in the United Kingdom can
be explained by the level of unemployment and the rate
of change of unemployment” (1958, 284).
The logic behind Phillips’s theory is very simple. If
for some reason the demand for labor were high relative
to its supply—as in Atlanta during the Olympics, to use a
modern example—then equilibrium wage rates would be
expected to rise above current wage levels, and there
would be upward pressure on nominal wages as firms bid
for additional workers. As additional workers were actually hired, moreover, the unemployment rate would fall.
The larger the discrepancy between the quantity of labor
demanded and the quantity supplied, the stronger the
upward or downward pressure on wage rates. The opposite would be true when there was excess supply of labor
and rising unemployment.
Phillips found, as he expected, that from 1861 to
1957 the growth rate of nominal wages was negatively
correlated with the rate of unemployment—that is, low
unemployment rates tended to be associated with rapidly
rising wages while high unemployment rates were associated with slowly rising wages. Phillips also found that the
strength of the unemployment versus wage-change relationship seemed to depend on the level of unemployment.

I

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

CHART 1

The Phillips Curve

Inflation Rate

When unemployment was low, decreases in unemployment tended to be associated with big increases in wage
inflation while when unemployment was high, decreases
in the unemployment rate seemed to produce small
increases in wage growth rates (see Chart 1 for a hypothetical Phillips curve). These findings appeared to confirm Keynes’s theory of the downward stickiness of
nominal wages. Tight labor markets seemed to cause
employers to bid wages up rapidly while loose markets
(high unemployment) seemed to cause workers to bid
wages down relatively slowly.
Phillips’s findings have had a profound and lasting
effect on economists’ ideas about the relationship
between inflation and unemployment. What made them
so interesting is that they seemed to establish a clear
linkage between the state of the labor market and the
rate of inflation. By the early 1960s, inflation rates in the
United States and western Europe had increased to the
point that inflation was coming to be regarded as a serious economic problem. As a result, economists and policymakers were eager for information about its possible
causes and potential cures. The Phillips curve appeared
to link the real and nominal sides of the economy.4
One possible objection to the conclusions that
Phillips (and others) drew from his findings is that standard economic theory predicts that what matters to workers is not their nominal wages but their real, or
inflation-adjusted, wages.5 Phillips did not attempt to
measure real wages or study their statistical relationship
to unemployment. Under the Keynesian assumption of
predetermined or sticky nominal prices, however,
changes in expected real and nominal wages would coincide. In addition, while Phillips’s statistical evidence
involved changes in current nominal wages, the hypothesis that he was trying to test involved changes in expected nominal wages. If workers were slow to adjust their
price expectations to actual price changes, changes in
current nominal wages could be interpreted as changes
in expected real wages.
Another problem with Phillips’s findings is that they
involve wage inflation while economists were principally
concerned about explaining price inflation. Since wages
are the biggest single component of firms’ costs, however,
most economists were willing to assume that persistent
increases in wage rates would eventually force firms to
begin increasing their prices, producing economywide

Unemployment Rate
The now-conventional Phillips curve diagram has the unemployment
rate on the horizontal axis and the inflation rate on the vertical axis.

price inflation. For this explanation for inflation to make
sense, however, it was necessary to make even more elaborate assumptions about stickiness: wages now had to be
assumed to adjust faster than goods prices, at least when
wages were rising. (In conventional Keynesian theory,
nominal wages were supposed to be slow to fall when a
decrease in aggregate demand put downward pressure on
prices; the result was a higher-than-equilibrium real
wage and involuntary unemployment.)
How was the Phillips curve related to monetary policy? Keynesian theory held that monetary policy could be
used to increase or decrease the economy’s aggregate
demand—the total nominal demand for goods and services of all types—and through it the aggregate level of
employment in the economy. The Phillips curve mechanism explained how aggregate demand management
could affect the rate of inflation. Thus, economic policymakers began to think in terms of a trade-off between
the unemployment rate and inflation rate. Although government aggregate-demand stimulus was no longer costfree, as it had been in traditional Keynesian theory
(which had viewed the price level as constant), it was
still possible for the policy authority to reduce the level
of employment if it was willing to tolerate the resulting
increase in inflation along the Phillips curve. As the next
section will show, another reason for the popularity of

2.According to Keynes, the principal source of the observed fluctuations in labor demand was the volatility of aggregate investment. Investment volatility, in turn, was caused by changes in short- and long-term business expectations and variation in
interest rates.
3.The discussion will show that the stickiness assumption was also extended to aggregate prices.
4.Phillips was not the first researcher to turn up findings of this general sort. As long ago as 1926 Irving Fisher had found a negative correlation between the rate of goods-price inflation and the level of unemployment.
5.If workers in New York City and rural Mississippi both make $2,500 per month, the worker in rural Mississippi will have a much
higher real wage because the cost of living is lower there.
Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

7

the Phillips curve is that it was seen by some prominent
economists as providing a synthesis of competing theories of inflation.
Cost-Push versus Demand-Pull Inflation. At the
time the Phillips curve analysis appeared, economists’
interest in understanding the relationship between wages,
prices, and economic activity had been growing for some
time, and there was also growing interest in studying the
effects of government policies on this relationship.
Samuelson and Solow (1960) provide a comprehensive
review of the debate on these questions that took place
after the Second World War. The debate centered on two
basic theories of the causes of inflation: demand-pull and
cost-push. Both theories can be explained using the
aggregate-demand/aggregate-supply model of output and
price level determination that was developed during the
1950s and remains popular in textbooks. Demand-pull
inflation resulted from increases in the level of aggregate
demand that occurred at or near the point of full capacity utilization—that is, at points at which the aggregate
supply curve was upward-sloping rather than flat. Costpush inflation, on the other hand, was caused by upward
shifts in the aggregate supply curve. These shifts could
allow wages and prices to rise even before full employment was reached.6
According to Samuelson and Solow, there were really no purists in this debate. Most economists believed
that inflation had both demand-pull and cost-push components, but they differed as to which component predominated. Thus, although demand-pull inflation was
associated with Keynesian theory, Keynes himself did not
dismiss the cost-push hypothesis. He was “willing to
assume that attainment of full employment would make
prices and wages flexible upward. . . . Just as wages and
prices may be sticky in the face of unemployment and
overcapacity, so may they be pushing upward beyond
what can be explained in terms of levels and shifts in
demand” (1964, 180-81).
Samuelson and Solow believed that in order to reconcile the two sides of this debate it would be necessary
for economists to improve their understanding of the
behavior of money wages with respect to the level of
employment. They saw the Phillips curve as a useful tool
for analyzing this behavior. Under some conditions, they
explained, “movements along the Phillips curve might be
dubbed standard demand-pull, and shifts of the Phillips
curve might represent the institutional changes on which
cost-push theories rest” (1960, 189).

The Monetarist Challenge to the
Keynesian Approach
he Acceleration Hypothesis. One prominent U.S.
economist who was skeptical of Keynesian theory
in general, and of Phillips curve analysis in particular, was Milton Friedman. Friedman was the champion

T
8

of monetarism, a theory that saw inflation as always and
everywhere a monetary phenomenon. He was also rather
skeptical of the Keynesian view that demand-management
policy could have significant effects on output or employment. Beginning in the mid-1960s, Friedman began to
challenge some of the conclusions about the inflationunemployment relationship that economists writing in
the 1960s and early 1970s were drawing on the basis of
Keynesian theory.
As we have seen, Keynes’s explanation for persistent
unemployment was that the prevailing level of real wages
was not compatible with labor market clearing and
instead produced excess supply of labor. This fact raised
the question of why lower, market-clearing real wages
could not be produced by reductions in nominal wages.
One explanation frequently offered was that workers
would oppose nominal wage reductions. Friedman (1976)
was very skeptical about this and other explanations that
Keynesians put forward to explain supposed nominal wage
rigidities. He was willing to concede that there might be
some situations in which wages and salaries were rigid;
the legal minimum wage, he noted, was an example of
such a rigidity. He argued, however, that situations like
these were the exception rather than the rule. In most
industries, he pointed out, relatively few workers earned
the minimum wage: what prevented workers in these
industries from reducing their wage requests in order to
avoid layoffs? And while unions could conceivably be a factor delaying wage adjustment because of their reluctance
to accept wage cuts that would benefit unemployed workers at union members’ expense, he did not believe that
unions were powerful or perverse enough to keep wages
from adjusting to full employment levels in the long run.
A second criticism Friedman raised was that
researchers had not been able to construct “decent”
empirical Phillips curves for the United States or other
countries. In later years this problem got worse, and even
ardent Keynesians were forced to acknowledge the weakness of the empirical evidence supporting the existence
of stable national Phillips curves. In 1980, for example,
prominent Keynesian Arthur Okun, commenting on the
U.S. case, wrote that “since 1970, the Phillips curve has
been an unidentified flying object and has eluded all
econometric efforts to nail it down” (1980, 166).
Friedman’s third criticism was outlined in the previous section: Phillips’s statistical evidence involved nominal wages, but standard economic theory assumes that
households and firms base their employment decisions on
real wages. Clearly, Phillips and his successors were
assuming that changes in current nominal wages were
equivalent to changes in expected future real wages. This
assumption, Friedman noted, really amounted to two
assumptions. The first was that prices, or at least price
expectations, were rigid: people did not expect the price
level to change and consequently interpreted changes in

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

their nominal wages as changes in their real wages. The
second assumption was that workers would not resist
reductions in their real wages that were caused by inflation rather than by reductions in their nominal wages.
Only if both assumptions were true could the relationship
between the rate of change in nominal wages and the
aggregate level of unemployment be stable enough to
then offer policymakers a usable menu of options.
A closely related argument made by both Friedman
(1968) and Phelps (1967) involved the long-run implications of the Phillips curve. In order to make this argument,
they imagined a situation in which a policymaker was trying to use the hypothesized inflation-unemployment
trade-off to achieve a lasting reduction in the unemployment rate. Such a policymaker, they argued, would
find that while there might indeed be an inflationunemployment trade-off in the short run, the trade-off
would disappear in the long run. In the long run, they
asserted, unemployment tended to return to a “natural
rate” (NR) that was determined by real economic forces.7
Monetary policy, in their view, could do nothing to change
the natural rate.
The analysis presented by Friedman and Phelps,
which was later summarized by Friedman (1976),
involved the relationship between real wages and unexpected inflation. The emphasis on unexpected inflation
reflected an attempt on the part of Friedman and Phelps
to reconcile the classical principle that labor supply
behavior depends on the real wage with Keynes’s observation that workers respond differently to different types
of real wage decreases: “Every trade union,” Keynes
writes, “will put up some resistance to a cut in moneywages, however small, . . . but no trade union would dream
of striking on every occasion of a rise in the cost of living”
(1964, 14-15). According to Friedman, this differential
response is due to temporary money illusion: it takes time
for workers to recognize that the price level has
increased, and until they do so they do not realize that
their real wage rates have fallen.
Friedman’s discussion can be interpreted as an
implicit description of the following hypothetical
sequence of events. Suppose the economy starts out in its
long-run equilibrium at its normal inflation rate and its
natural rate of unemployment. This equilibrium is disturbed when a monetary expansion increases households’

aggregate demand for goods and services at current
prices. Demand curves will shift to the right throughout
the economy, and the market prices of (output) goods
and services will rise. The increased market prices of
goods will cause the aggregate demand curve for labor,
plotted against the nominal wage, to shift to the right.
If workers realize that the price level has increased,
then their aggregate labor supply curve will shift to the
left, as depicted in the shift from curve D to curve D1 in
Chart 2. Equilibrium will be restored at a higher nominal
wage rate but at unchanged levels of employment, output, and real wages. If workers do not realize that
the price level has
increased—that is, if
the increase in prices
is both unperceived
Keynesian theory implies
and unexpected—
that government policies
then employment and
nominal wage rates
can have large, important
will increase along the
effects on the economy
old labor supply curve.
and that if the policies are
Workers will now be
providing more labor
carefully devised these
than they would be
effects can be very conwilling to provide at
structive in nature.
the current real wage
if they knew what that
wage really was. At
some point, however,
workers will figure out that the price level has increased,
and the aggregate labor supply curve will begin shifting to
the left, as depicted in the shift from curve S to S1 in
Chart 2. The shift in the supply curve will drive nominal
wages up further. As nominal wages rise, the supply
curves for goods and services will shift to the left, driving
the price level up further, and so on. Nominal wages will
rise faster than prices, however, as workers catch on to
the successive price increases. Eventually a new long-run
equilibrium is reached at the original unemployment rate
(the natural rate) and the original level of real wages—
the point L0 in Chart 2. Notice that once the process of
adjustment to the new long-run equilibrium gets started,
prices lead wages upward rather than the reverse.
To summarize, Friedman and Phelps argued that
unexpected inflation can drive the level of unemployment

6.Believers in cost-push inflation often identified unions as one of its main sources. Samuelson and Nordhaus point out, however, that “this view of unions as the clear-cut villain of cost-push inflation does not fit the complex historical facts. Take as an
example the depressed year of 1982, when unemployment averaged 9.7 percent of the labor force. During that year, labor costs
for union workers rose 7.2 percent, and the cost of nonunion workers rose 6 percent. Both union and nonunion wages rose
smartly in spite of high unemployment” (1989, 326).
7.Friedman defines the natural rate of unemployment as the level of unemployment “that would be ground out by the Walrasian
system of general equilibrium equations, provided there is imbedded in them the actual structural characteristics of the labor
and commodity markets, including market imperfections, stochastic variability in demands and supplies, the cost of gathering
information about job vacancies and labor availabilities, the cost of mobility, and so on” (1968, 8).

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

9

C H A R T 2 Effects of Monetary Policy on the Labor Market
S1

W

S

W2
W1

D1

W0
D

L1

L0

L

Friedman and Phelps argued that unexpected inflation can drive the level of unemployment below the natural rate, but only temporarily.

below the natural rate, but only temporarily. In the long
run, the surprise factor will disappear as workers learn
that the price level has increased; as a result, the level
of employment will go back to the natural rate. Thus, in
the long run there will be no inflation-unemployment
trade-off. Stated differently, the long-run Phillips curve
is vertical.
At this point, it is necessary to make some important
distinctions regarding the term inflation. A one-time
increase in the price level is sometimes called inflation,
but it is very different from a situation in which the price
level is increasing over time at a constant rate. Both
these situations, moreover, are different from one in
which the price level is increasing over time at a rate that
is also increasing over time (so that the price level is
accelerating upward). The inflation that the Keynesian
economists who developed Phillips curve analysis had in
mind was the type in which the price level increases at a
fixed rate. These economists believed that from the point
of view of policymakers, the cost of achieving a lower
level of unemployment was that the price level would now
increase at a higher rate. Inflation would remain constant
at this new, higher rate as long as the unemployment rate
remained at its new, lower level.
According to Friedman and Phelps, the actual relationship between inflation and unemployment was quite
different. In their minds, at least, the difference between
their view of this relationship and the Keynesian view
involved the way in which workers were assumed to form
their expectations. In describing the difference between
his view of this process and the view he attributes to
Phillips, Friedman quotes Abraham Lincoln’s famous
assertion that “you can fool all of the people some of the
10

time, you can fool some of the people all of the time, but
you can’t fool all of the people all of the time” (1976, 231).
To Friedman, Phillips’s analysis made sense only if workers could be fooled all the time—only, that is, if a given
increase in the price level (beyond some unspecified base
inflation rate) always fooled workers to exactly the same
extent, regardless of how many times they had been
fooled previously. Thus, persistent increases in the price
level could hold the labor supply curve fixed in a location
to the right of its no-surprises position, producing lower
unemployment. Higher inflation rates, moreover, shifted
the curve further than lower inflation rates and thus produced lower levels of unemployment.
Friedman and Phelps, in contrast, thought that
while it might be possible to fool all the workers some of
the time (temporarily), it was not possible to fool all of
them all of the time (permanently). Eventually, workers
would recognize that the base rate of inflation had
increased, at which point the labor supply curve would
begin to shift back and the increased inflation rate would
gradually lose its power to reduce the unemployment
rate. Further declines in unemployment could then be
achieved, if at all, only by further increases in the rate of
inflation. Thus, “the only way unemployment can be kept
below the natural rate is by an ever-accelerating inflation, which always keeps current inflation ahead of anticipated inflation” (Friedman 1976, 227).
The view underlying this “acceleration hypothesis” is
that while agents cannot be permanently fooled by inflation at a fixed rate, they can be fooled persistently, if not
permanently, by accelerating inflation. One reason to be
skeptical about this story is evidence from economies
that have experienced hyperinflations (extremely rapid

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

increases in the aggregate price level): it is not unusual
to see hyperinflation and high rates of unemployment go
hand-in-hand. It should also be emphasized that nothing
in this analysis suggests that any one-time increase in the
price level must necessarily be followed by persistent
inflation at a fixed rate that will eventually turn into
accelerating inflation. The accelerating inflation
described by Friedman and Phelps is created by design in
order to surprise economic agents. It will not result from
forces beyond the control of the policymakers, and it will
not be produced by policymakers that implement a stable
monetary policy—even if that policy involves a high
money growth rate.
The NIRU (aka NAIRU): A Response to the
Monetarists. Although the introduction to this article
focused on the NAIRU, the analysis presented so far has
concentrated on the Phillips curve. The reason for this
attention is that the Phillips curve is a key element of the
theory of the inflation-unemployment relationship that
includes the NAIRU.
As the discussion has shown, during the 1960s
Keynesian theorists came to regard the inverse (downwardsloping) empirical relationship between inflation and
unemployment—the Phillips curve relationship—as a
stable menu of options from which policymakers could
choose. The apparent concreteness of this menu helped
produce widespread confidence in the potential effectiveness of Keynesian-inspired countercyclical demand
management. To Keynesians, the job of macroeconomists
was to design demand-management policies that would
strike the right balance between the competing problems
of unemployment and inflation. Monetarists did not share
the Keynesians’ faith in the effectiveness of demand management, and during the 1960s and the 1970s there were
fierce debates between the two schools. These debates
sometimes took the form of disputes about the slope of
the Phillips curve. Keynesians believed that the Phillips
curve was quite flat, particularly at high unemployment
rates. It followed that when unemployment was high, the
unemployment rate could be reduced at little cost in
terms of increased inflation. Monetarists, on the other
hand, believed the curve was quite steep, so expansionary
demand management was likely to produce a significant
amount of inflation without providing much benefit in
terms of lower unemployment. The monetarist challenge
to Keynesian ideas about the Phillips curve culminated in
the Friedman-Phelps hypothesis that the curve was vertical in the long run.
During the severe recession of 1974-75 both the
inflation rate and the unemployment rate reached some

of the highest levels in postwar U.S. history. This experience shook public faith in Keynesianism and played a key
role in shaping the subsequent debate about inflation.
The warnings of Milton Friedman and other monetarists
that attempts to “ride the Phillips curve” might lead to
accelerating inflation began to be heeded by more and
more people, both inside and outside the ranks of professional economists. The credibility of the monetarist alternative to Keynesian theory was greatly strengthened.
Despite the credibility gains of the monetarists, however, the events of the mid-1970s did not result in the
demise of Keynesian macroeconomics or even of analysis
based on the Phillips curve. Many economists continued
to use the Phillips curve as the basis for forecasting and
policy advice. As Okun recalls, “It was hard to cast aside
a tool that had traced the United States record so well
from 1954 through the late sixties. And it was easy to
ignore the Friedman and Phelps attack on the stability of
the short-run Phillips curve, and their prophetic warning
(issued at a time when the Phillips curve was still performing admirably) that the curve would come unstuck
in a prolonged period of excess demand. Unfortunately,
most of the profession (including me) took too long to
recognize that” (1980, 166).
Some Keynesians reacted to the events of 1974-75 by
attempting to reinterpret the Phillips curve in a way that
reconciled the Keynesian and monetarist views of the
inflation-unemployment relation but preserved considerable scope for activist demand management. To do so
was necessary to acknowledge that there might indeed
be limits to the exploitability of the Phillips curve relation: in particular, attempts to use it to keep the unemployment rate below a threshold level might indeed
result in accelerating inflation. As early as 1975, for
example, Keynesians Franco Modigliani and Lucas
Papademos asserted that “the existence of NIRU [the
noninflationary rate of unemployment] is implied by both
the ‘vertical’ and the ‘nonvertical’ schools of the Phillips
curve” (1975, 142).8
What exactly was the NIRU? In the now-conventional
Phillips curve diagram, which has the unemployment rate
on the horizontal axis and the inflation rate on the vertical axis, the NIRU was the unemployment rate at which
the Keynesians’ downward-sloping Phillips curve intersected a vertical line at Friedman’s natural rate of unemployment. Thus, the NIRU was equal to the natural rate.
But while monetarists believed that the existence of a
natural rate implied that there was no useful trade-off
between inflation and unemployment, Modigliani and
Papademos interpreted the NIRU as a constraint on the

8.The NIRU was later renamed the NAIRU, or nonaccelerating inflation rate of unemployment. This name makes it clear that
sufficiently low unemployment rates are believed to be associated with accelerating inflation, not just higher fixed rates of
inflation.
Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

11

CHART 3

Inflation Rate

The Natural Rate

NR
Unemployment Rate
Keynesians believed that the economy spent most of its time in a range
of unemployment rates well to the right of the natural rate and that
unemployment rates to the left of the shaded area implied that inflation
was likely to accelerate.

ability of policymakers to exploit a trade-off that
remained both available and helpful in the short run.
Perhaps the most striking thing about the
Modigliani-Papademos argument is that while it incorporated many aspects of Friedman’s critique of Keynesian
theory, it stood Friedman’s principal policy recommendation on its head: Friedman was strongly opposed to
activist monetary policy. One of the reasons that was possible was that most expositions of the monetarist view of
the inflation-unemployment relationship—including
Friedman’s—did not seem to resolve the question of the
strength or persistence of the short-run effects of monetary policy. After all, Friedman’s inflation-acceleration
theory did seem to suggest that monetary policy could
produce temporary reductions in the level of unemployment—but these reductions could be sustained only at
the price of continually increasing inflation rates.
The remaining difference between the Keynesians
and the monetarists was actually quite fundamental: it
involved the direction of the causal relationship between
inflation and unemployment. This difference continued
to allow members of the two schools to hold contrasting
views about the sensitivity of the unemployment rate to
changes in the inflation rate (or vice versa) and thus
about whether the short-run Phillips curve trade-off was
potentially useful to policymakers.
Monetarists saw the level of unemployment as determined largely through the process of labor market clearing. The economy, in their view, was never far from the
full-employment equilibrium of the classical model.
Monetarists believed that monetary policy had a direct
and powerful influence on the price level and the inflation rate. While the channels through which it obtained
12

this influence might involve the goods and labor markets,
these markets adjusted and cleared so quickly that policy changes had little effect on them. In particular, monetary policy could affect the level of unemployment only
marginally and only by producing inflation surprises
whose impact would decrease rapidly over time. Since
unexpected changes in the inflation rate could produce
only small changes in the level of unemployment, the
Phillips curve was quite steep even in the short run. The
rate of unemployment could never stray far from the natural rate, and continued efforts to keep it below the natural rate would result mostly in accelerating inflation.
Keynesians, on the other hand, continued to believe
that the economy could and often did operate at “equilibrium” positions in which aggregate demand was deficient—positions in which there was massive excess
supply of labor and large-scale involuntary unemployment. The level of unemployment, they believed, determined the rate of inflation by determining the growth
rate of nominal wages (see above). Thus, changes in
unemployment caused changes in inflation, rather than
the reverse.
It was their belief that the level of aggregate demand
could be and often was deficient that allowed Keynesians
to believe that policies that influenced its level could play
an important role in determining the current level of
employment. As long as there was “slack” (unemployed
labor and other resources) in the economy, monetary
ease, for example, would not start a wage-price spiral
because the initial round of goods-price increases it produced (see above) would not place substantial upward
pressure on nominal wage rates. Thus, Keynesians
believed that the economy spent most of its time in a
range of unemployment rates well to the right of the natural rate/NIRU—a range within which the Phillips curve
was very flat. If demand stimulus pushed the unemployment rate too low, however, then labor market tightness
would put persistent upward pressure on the inflation
rate. This was the range where the short-run Phillips
curve was steep; it was also the range within which the
long-run increases in the inflation rate predicted by
Friedman were a serious potential problem. Thus
Modigliani and Papademos wrote that “unemployment
rates left of the shaded area [the area displaying the current range of NIRU estimates] imply a high probability
that inflation will accelerate” (1975, 147) (see Chart 3).
Despite the fundamental differences between the
monetarists and the Keynesians, the NIRU was seen by
many contemporary economists as helping build a consensus about the nature of the inflation-unemployment
relationship. According to James Tobin, the “consensus
macroeconomic framework, vintage 1970” held that “the
nonagricultural business sector plays a key role in determining the economy’s rate of inflation. . . . According to
the standard ‘augmented Phillips curve’ view, rates of

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

price and wage increase depend partly on their recent
trends, partly on expectations of their future movements,
and partly on the tightness . . . of markets for products
and labor. Variations in aggregate monetary demand,
whether the consequences of policies or other events,
affect the course of prices and output, and wages and
employment, by altering the tightness of labor and product markets, and in no other way. . . . Inflation accelerates
at high employment rates because tight markets systematically and repeatedly generate wage and price
increases. . . . At the Phelps-Friedman ‘natural rate of
unemployment,’ the degrees of resource utilization and
market tightness generate no net wage and price pressure up or down and are consistent with accustomed and
expected paths, whether stable process or any other
inflation rate. The consensus view accepted the notion of
a nonaccelerating inflation rate of unemployment
(NAIRU) as a practical constraint on policy” (1980, 23).9
Most current descriptions of the Phillips curve relationship and the NAIRU are not very different from
Tobin’s description. One difference is that most modern
descriptions see changes in monetary policy as the principal source of changes in the economy’s aggregate
demand—a view that Tobin ascribes to the monetarists.
Otherwise, the accounts are similar to Tobin’s in the
sense of asserting (1) that the current stance of [monetary] policy can be determined by looking at the unemployment rate and comparing it with its natural rate and
(2) that the current level of the unemployment rate provides a good indication of the direction and strength of
future changes in the inflation rate: low unemployment
indicates that the rate of inflation will increase in the
short run and accelerate in the long run.
As Tobin pointed out, the macroeconomic consensus
about the nature of the inflation-unemployment relationship did not extend to the question of whether policymakers could or should exploit that relationship. In a recent
column in the Wall Street Journal, Friedman recounts: “I
introduced the concept of the natural rate in 1968 as part
of an article on ‘The Role of Monetary Policy.’ . . . The natural rate is a concept that does have a numerical counterpart—but that counterpart is not easy to estimate and
will depend on particular circumstances of time and
place. More important, an accurate estimate is not necessary for proper monetary policy. I introduced the concept
in a section titled ‘What Monetary Policy Cannot Do.’ It
was part of an explanation of why, in my opinion, the monetary authority cannot adopt ‘a target for employment or
unemployment . . . ; be tight when unemployment is less
than the target; be easy when unemployment is higher
than the target’” (WSJ, September 24, 1996).

To summarize, the NAIRU was born out of an attempt
by proponents of the Phillips curve to address the monetarist critique of policy prescriptions based on the curve.
In the minds of many Keynesians, the NAIRU theory successfully reformulated the natural rate hypothesis as a relatively minor qualification of Keynesian theories about
the usefulness of the Phillips curve as a guide to monetary
(or fiscal) policy. From the monetarist perspective, however, the NAIRU was simply another name for the natural
rate. The NAIRU theory, moreover, was based on a fundamental misunderstanding of the natural rate hypothesis—a hypothesis that demonstrated the ineffectiveness
of government demand-management policy.
There is a sense in which it is hard to blame the
NAIRU proponents for ignoring monetarist assertions that
monetary policy was inherently neutral. After all, monetarists such as Friedman had long argued that activist
monetary policy was in fact the principle source of shortrun economic fluctuations. The monetarists gave further
ground to the Keynesians by maintaining a distinction
between the short run and the long run and by speaking of
money illusion as a channel that gave policymakers access
to a short-run inflation-unemployment trade-off. To the
dismay of the leading monetarists, proponents of the
NAIRU were quite successful in capitalizing on its appeal
as a simple, intuitive guide for giving policy advice.

Some Questions about the NAIRU
here are additional problems with using the NAIRU
concept to formulate policy rules that are not
directly connected to the Keynesian-monetarist
debate. One of these involves the relationship between
changes in relative prices and changes in the aggregate
price level. Relative price changes signal degrees of relative scarcity in the economy: they reveal how highly the
economy values different goods and services and are
often associated with changes in the quantities of those
goods and services produced or employed. One very
important relative price is the real (or relative) wage,
which is the purchasing power of the nominal wage in
terms of goods and services and can be loosely defined as
the average nominal wage divided by the average price
level. Changes in real wages reflect changes in the value
of labor services relative to the value of other goods and
services. They are often associated with changes in the
level of employment.
Both conventional monetarist theory and the
Keynesian/monetarist synthesis of the 1970s predict
that the mechanism by which monetary policy creates
inflation involves repeated increases in both nominal
and real wages and temporary decreases in the rate of

T

9.Tobin seems to have been the first writer to actually use the term NAIRU; recall that Modigliani and Papademos used the
acronym NIRU (noninflationary rate of unemployment).
Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

13

unemployment. According to these theories, the changes
in nominal wages caused by monetary policy do not
result in permanent changes in the level of real wages
because the price level eventually adjusts to offset the
nominal wage changes.
These macroeconomic theories are often “inverted”
to produce rules for conducting monetary policy that are
based on current levels of unemployment or current rates
of change in nominal wages. The simplest rule of this type
is that when the unemployment rate is lower than the
NAIRU, monetary policy has become too “easy” and
should be tightened to head off the coming inflationary
spiral. Unfortunately (or perhaps fortunately), the fact
that there is little agreement on the precise value of the
NAIRU makes this rule hard to implement. An alternative
rule that does not suffer from this problem is to tighten
policy whenever nominal wages begin growing more rapidly than
prices
so that real
Friedman and Phelps
wages begin to rise.
argued that unexpected
As the introduction to this article
inflation can drive the level
noted, one of the reaof unemployment below the
sons that the NAIRU
natural rate, but only temhas attracted so much
attention recently is
porarily; in the long run
that the level of unemthere will be no inflationployment has been
unemployment trade-off.
low—lower than many
NAIRU estimates. As a
result, some economists have called on
monetary policymakers to move to tighten policy, and
others have suggested that they begin watching nominal
wage changes closely and tighten policy as soon as there
is any sign that real wages are rising.
Non-Policy-Induced Changes in Real Wages. Do
policy rules of this sort make sense, even if we accept the
underlying theory of the effects of monetary policy? One
fundamental problem with these rules is that they implicitly assume that any change in labor market conditions
that produces lower unemployment or higher real wage
rates must have resulted from monetary policy.10 Of
course, virtually every economist acknowledges that
changes in labor productivity produce persistent increases in the relative price of labor, thus causing real wage
rates to change independently of monetary policy
changes. Consequently, it is often suggested that policymakers should respond only to increases in nominal
wages that result in real wage increases that cannot be
attributed to gains in productivity.
Is it reasonable to assume that every increase in real
wages that cannot be directly linked to an increase in
productivity has been caused by a change in monetary
14

policy and will eventually be followed by an increase in
inflation? Two big problems with this assumption are that
labor productivity is notoriously difficult to measure and
that productivity data become available only after a considerable time lag. However, even if labor productivity
could be measured in a timely and accurate manner, it
would not follow that increases in wages that were not
associated with productivity gains were necessarily
caused by monetary policy. Not all changes in the demand
for goods and services come from changes in monetary
policy, or even government policy, and some of these
changes may affect both the relative price of U.S. labor
(that is, the real wage) and the level of U.S. employment.
Examples include changes in foreign demand for U.S.
exports—particularly exports of goods that are labor- or
human capital-intensive—or changes in domestic tastes
favoring goods of the same type.
As we shall see, if wages and prices are perfectly
flexible then changes in relative prices—including relative wages—should have no effect on the aggregate price
level. If wages or prices are sticky, however, then relative
wage or price changes may appear to produce aggregate
price level changes and may even appear to produce persistent inflation.
Why is the possibility of non-policy-induced changes
in the relative price of labor important? Most economists
would agree that it makes sense to use monetary policy to
resist real wage or unemployment rate changes if these
changes are simply a lane on the road to a permanent
increase in the rate of inflation. Most economists would
also agree that policymakers should not resist real wage
or unemployment rate changes that are associated with
permanent (or persistent) increases in the relative price
of labor—even if these changes appear to produce temporary increases in the inflation rate. Resisting changes
of this sort would risk letting monetary policy interfere
with the important job of relative price changes, which is
to ensure that inputs and outputs continue to be used
and produced efficiently.
Unfortunately, it is not always easy to distinguish
temporary changes in the inflation rate from permanent
ones. As a result, the fact that there may have been many
occasions in the past when increases in the relative price
of labor produced temporary increases in the inflation
rate may reinforce some economists’ present tendency to
advocate tightening in response to current increases in
nominal and real wages. Thus, real wage changes that are
not caused by policy-induced changes in aggregate
demand can create a great deal of confusion for policymakers who are trying to use wage growth rates or unemployment rates as guides to monetary policy.
Relative Price Changes and the Aggregate Price
Level. The most common method for measuring changes
in the aggregate price level involves taking a fixed basket
of goods and determining how the money cost of that bas-

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

ket has changed over time. The price index produced by
this method is equal to the market value of the basket at
a particular point in time divided by the market value of
the same basket in a fixed base year—typically, the year
in which the basket was chosen. The consumer price
index (CPI), which is the most closely watched price
index, is constructed in this manner. The following example illustrates the impact of a relative price change—a
change in the price of a single good relative to the prices
of other goods—on a price index like the CPI.
Imagine a household that consumes (1) directly
provided labor services (for example, cleaning services),
(2) the services of durable goods (such as personal computers), and (3) food (bread). Now suppose that the
demand for directly provided labor services increases—
perhaps because foreign tourism in the United States has
increased and hotels and condo owners are hiring people
to clean the rooms and condos foreign tourists have rented. Standard microeconomic theory predicts that this
increase in demand will lead to an increase in the price
of these services. Assume, for the moment, that the prices
of the two other classes of consumption goods do not
change (an assumption that will have to be abandoned
later). Thus, both the absolute and relative prices of
direct labor services have increased.
How will a change in the price of direct labor services—a relative price change—affect the CPI, which is
a measure of the overall price level? The fixed-marketbasket method for constructing the CPI amounts to
assigning different fixed weights to the prices of the different items in the basket. For the purposes of this example, assume that cleaning services, personal computer
services, and bread are the only items in the basket and
that their initial prices are $10 per hour for cleaning services, $10 for computer services, and $10 per loaf of
bread. Also assume that a typical individual allocates 10
percent of his or her spending to cleaning services, another 10 percent to computer services, and the remaining 80
percent to buying bread. Finally, assume that the initial
prices of these items are the same as the ones from the
base year. We can then construct the initial value of our
hypothetical price index:

CPI initial =

0.1($10) + 0.1($10) + 0.8($10) $10
=
= 1.00.
0.1($10) + 0.1($10) + 0.8($10) $10

Now suppose that the price of cleaning services doubles
but the other two prices remain unchanged. The CPI
would then be

CPI =

0.1($20) + 0.1($10) + 0.8($10) $11
=
= 1.10.
0.1($10) + 0.1($10) + 0.8($10) $10

In this case, the reported inflation rate would be 10 percent.
An important question, however, is whether it is really reasonable to hold the weights of the three goods/services fixed in light of the large increase in the price of one
of them. From elementary microeconomics, we know that
the price increase is likely to produce a “substitution
effect” on spending: people will respond to the relative
price increase by substituting out of market-delivered
cleaning services, either by accepting slightly messier
homes or doing more cleaning themselves. They may also
buy additional durable goods (such as carpet-cleaning
machines) to help them do their own cleaning. With this
likelihood in mind, and ignoring for the moment the possibility of further adjustment in relative prices, let’s imagine the effects of allowing the quantity weights to adjust.
Assume that U.S. households change their spending patterns so that they purchase fewer hours of cleaning services (labor), whose weight falls from 0.1 to 0.05, and
more durables services, whose weight rises from 0.1 to
0.15). (Note that the government agency that constructs
the actual CPI does not make these kinds of adjustments,
except quite infrequently—see below.) Our “revised”
June CPI would look like this:

0.05($20) + 0.15($10) + 0.8($10)
0.1($10) + 0.1($10) + 0.8($10)
$10.50
=
= 1.05,
$10

CPI( rev) =

in which case the rate of inflation would now be only be
only 5 percent.
Clearly, the increase in the value of the price
index—that is, in the aggregate price level—is smaller
when the quantity weights are allowed to adjust to
changes in expenditure patterns. In other words, the substitution effect acts to restrain the “inflationary” effects
of relative price increases.

10. One noteworthy aspect of Friedman’s explanation of the Phillips curve mechanism was that he was as willing as most other
economists to accept the notion that increases in wage rates were essentially equivalent to increases in the price level. In
Friedman’s words: “Fisher talked about price changes, Phillips about wage changes, but I believe that for our purposes that is
not an important distinction. Both Fisher and Phillips took for granted that wages are a major component of total cost and
that prices and wages would tend to move together. So both of them tended to go very readily from rates of wage change to
rates of price change, and I shall do as well” (1976, 218). Of course, Friedman may have taken this approach not because he
agreed with the assumption that all wage-rate changes necessarily produce proportional price level changes but because he
was able to make his point about the natural rate of unemployment without worrying about this distinction.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

15

A second effect of relative price changes on the price
level is the often-overlooked “income effect.” An increase
in the absolute (dollar) price of cleaning services
reduces households’ purchasing power: they are no longer
able to afford the quantities of the three goods that they
were purchasing initially. This loss of purchasing power
will typically cause them to reduce their purchases of all
goods—even goods that are not closely related to the
goods whose prices have changed. In our example, households are slightly poorer because of the increased price of
cleaning services, and they react by reducing their purchases of bread. Bakers may be forced to respond by
reducing the price they charge for bread, which we will
assume falls to $9.38 per loaf.
What does the newly revised CPI look like after
accounting for the income effect?

0.05($20) + 0.15($10) + 0.8($9.38)
0.1($10) + 0.1($10) + 0.8($10)
$10
=
= 1.00.
$10

CPI( rev2 ) =

Thus, after the substitution and income effects have
worked their way through the economy, the increase in
the price level caused by an increase in a relative price—
in this case, something like a relative wage—is zero.
Unfortunately, the CPI as currently calculated does
not capture the substitution effect in a timely fashion:
while the quantity weights are periodically changed to
reflect changes in spending patterns, this revision happens only once every five years. Income-effect-induced
price changes will be captured as soon as they occur, but
these often take a long time to work their way through
the economy. It may take households some time to realize
that their real income has decreased and some additional time to adjust to the decrease; until they do adjust,
they may dig into their savings to finance higher-thannormal expenditures. Consequently, relative price or
wage increases may produce increases in the measured
price level in both the short run and the medium run,
even though they may have no long-run price level effects
once the income and substitution effects work their way
through the economy.
Menu Costs. Ball and Mankiw (1995) have developed a theory that provides a more detailed and specific
explanation of the process by which increases in relative
prices produce temporary increases in the aggregate
price level. Their key postulate is that there are “menu
costs”—costs of changing prices—that prevent nominal
prices from being fully flexible. Suppose, for example,
that veal is a key ingredient in many of the items on a
restaurant’s menu and that its market price has gone up
by a small amount. The restaurant owner is consequently
faced with an uncomfortable choice: increase the prices
16

of veal-based dishes to reflect the new veal price, which
will require an expensive reprinting of all the menus in
the restaurant, or simply absorb the price increase.
Changing announced prices may be costly for many
firms other than restaurants. The Ball-Mankiw theory
predicts that these costs will produce a “range of inaction”—a range of input-price increases small enough that
they will not cause producers to increase the prices of
outputs. They explain that “when a firm experiences a
shock to its desired relative price, it changes its actual
price only if the desired adjustment is large enough to
warrant paying the menu cost. . . . In this setting, shifts in
relative prices can affect the price level” (1995, 162). To
understand the latter point, imagine a no-menu-cost situation in which the prices of a small number of goods rise
substantially but the aggregate price level does not rise
because the income effect of these price increases
reduces the demand for a large number of other goods
and causes their prices to decline slightly. When there
are menu costs, however, it may not pay the producers of
these other goods to cut their prices in response to small
demand decreases. As a result, there may not be a large
number of small price decreases to offset the small number of large price increases, and the aggregate price level
may rise.
The Ball-Mankiw theory can help explain how a onetime increase in real wage rates or other relative prices
can produce a temporary increase in the aggregate price
level (as measured by a price index) and how repeated
increases in real wages or relative prices can produce a
temporary increase in the inflation rate.11 As the authors
note, this explanation presumes that the relative price
increases are concentrated in particular industries, and
thus require large price adjustments, while the resulting
income-effect-driven demand decreases are spread across
many different industries and consequently require relatively small adjustments. As applied to wage rates, the
theory predicts that the increases in real wages that are
most likely to result in temporary increases in inflation
are increases that are concentrated, at least initially,
among workers in particular industries. These wage
increases will produce cost increases in these industries
that exceed their ranges of inaction and will consequently impel the industries to increase their product prices
substantially.
Price Stickiness: The Empirical Evidence. Menu
costs are one possible example of a “nominal rigidity”—a
source of friction that prevents money prices from adjusting in the perfectly flexible manner assumed by classical
theory. Much of Keynesian theory, including the theory
behind the NAIRU, is based on the assumption that the
economy is afflicted by many other price rigidities of this
general type. As a result, one natural strategy for convincing skeptics of the validity of the theory would be to
describe the nature and source of these rigidities as pre-

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

cisely as possible. It would be helpful, for example, to be
able to identify the rigidities that are severe enough to
prevent nominal wages from adjusting to eliminate a persistent excess supply of labor—the rigidities, that is, that
allegedly permit persistent involuntary unemployment.
Similarly, it would be helpful to be able to identify the
frictions that allegedly make firms slow to adjust their
prices to increases in demand and workers slow to adjust
their wage demands to increases in prices. This information would make it much easier for skeptics to understand how aggregate demand stimulus could produce
significant (if temporary) increases in output and
employment.
Given the wealth of NAIRU-based advice that is currently being offered to policymakers, it may seem reasonable to infer that there is plenty of good evidence
supporting the claim that nominal rigidities are widespread and substantial. In reality this is not at all the
case. In a recent paper, Wynne (1995) reports the results
of a systematic search for empirical studies documenting
price stickiness. Despite the widespread acceptance of
theories based on sticky prices, he was able to find only a
small number of studies, including only three that used
data from the post-World War II period. Wynne also points
out that these studies would not stand up well against
some elementary objections to their methodology. For
example, the goods and services whose prices are examined in these studies account for a very small fraction of
GDP; they also include, in many cases, goods whose prices
are known a priori to be relatively inflexible or which
“exhibit little or no quality changes over time.” Wynne
goes on to point out that “many hi-tech products have
remarkably flexible prices” (1995, 7).
What about the assumption that is widely considered
absolutely fundamental to Keynesianism—the assumption that nominal wages are sticky downward? Zarnowitz
notes that “the average annual money earnings from
wages declined in about half of the business contractions
of 1860-1914 and in all of those of 1920-38, according to
the data compiled in Phelps Brown 1968. . . . In contrast,
they kept rising through the period 1945-60, which
witnessed four moderate or mild recessions. . . . Data for
1889-1914 from Rees 1961 show that peaks and troughs in
annual earnings matched nearly two thirds of the like
business cycle turns of the period, but those in hourly
earnings fewer than half. . . . The conclusion is that most
of the major business downturns and some of the minor
ones have historically been associated with declines in
nominal wage earnings” (1992, 146).

The NAIRU’s Empirical Record. As Okun (1980)
explains in a passage quoted above, for roughly fifteen
years ending in the late 1960s U.S. inflation and unemployment data seemed to line up along a stable Phillips
curve. The stagflation of the 1970s destroyed this empirical relationship. During the last twenty years, econometricians have not had much success identifying a stable,
reliable relationship between inflation and unemployment.
Of course, econometricians’ inability to construct an
empirically reliable Phillips curve makes it impossible for
them to produce a reliable estimate of the
NAIRU. Recently this
problem has become a
Despite the credibility
serious one for econogains of the monetarists,
mists who think monetary policy should be
the events of the midbased on the NAIRU.
1970s did not result in
During the past two
the demise of Keynesian
years, for example, the
U.S. unemployment
macroeconomics or even
rate has been quite
of analysis based on the
low—lower than
Phillips curve.
many widely public i z e d NAIRU estimates. However, the
inflation rate has
shown no signs of increasing (to say nothing of accelerating) in the way the NAIRU theory predicts. As Fred
Bleakley reports in a recent article in the Wall Street
Journal (February 1996), the failure of relatively low
unemployment rates to produce higher inflation rates has
led several prominent economists to revise their estimates of the NAIRU downward. Ex post revisions of this
sort are probably very frustrating for policymakers who
are seeking a reliable guidepost for monetary policy.
As frustrating as the current situation may be, economists and policymakers definitely prefer it to the 1970s,
when inflation rates and unemployment rates were high
simultaneously rather than low simultaneously. By the
end of the decade even inveterate Keynesians had begun
to lose faith in the usefulness of the NAIRU concept. At
the close of the 1970s, Tobin warned that “as for the
shape of the short-run trade-off [between inflation and
unemployment], Murphy’s Law of macroeconomics
assures us that it is an L with the corner wherever it happens to be. . . . It is possible that there is no NAIRU, no
natural rate, except one that floats around with actual

11. The theory does not imply that changes in relative prices can produce permanent price level increases. If the restaurant owner
believes that the change in the price level is permanent then it will make sense for him to revise his menu immediately since
he will have to revise the menu eventually and the longer he waits the greater his losses will be. Similarly, if relative prices rise
gradually over a period of time then the theory predicts that the inflation rate may increase during the same period of time
but not that the inflation rate will increase permanently.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

17

history. It is just as possible that the direction the economy is moving is at least as important a determinant of
acceleration and deceleration as its level. These possibilities should give policymakers pause as they embark on
yet another application of the orthodox demand-management cure for inflation” (1980, 61-62).12

Neoclassical Macroeconomics
he 1970s: Theory and Evidence Collide.
Economically, the decade of the 1970s was dominated by major “supply shocks”—principally, the
OPEC oil embargo and the resulting increases in world oil
prices. Supply shocks were not easily incorporated into
Keynesian theory. Historically, Keynesian theorists had
concentrated on studying the effects of
changes in aggregate
From the monetarist perdemand and had
implicitly assumed the
spective, the NAIRU theory
existence of a stawas based on a fundamenble aggregate suptal misunderstanding of the
ply schedule. As a
result, the supply
natural rate hypothesis—
shocks of the 1970s
a hypothesis that demoncaused forecasts based
strated the ineffectiveness
on Keynesian predictions to generate huge
of government demanderrors. As Tobin pointmanagement policy.
ed out, “the inflationary components of the
expansions, 1971-73
and 1975-79, were unexpectedly and distressingly large.
The disinflationary consequence of the first contraction,
1969-71, was distressingly small. Indeed, money wages
‘exploded’ while unemployment was rising. . . . The major
economic events of the decade were the extraordinary
changes in world supplies and prices of specific commodities. Their interaction with macroeconomic indicators and events confronted both policymakers and analysts
with problems for which they were unprepared. . . . No one
foresaw in 1970 the main economic events of the decade
or the formidable challenges those surprises would pose
for macroeconomics and stabilization policy. We macroeconomists were caught unawares. It was not simply that
our models, theoretical and econometric, now had to be
applied to novel situations. Worse than that, the shocks of
the 1970s required some fundamental rethinking and
rebuilding” (1980, 21-23).
Although Tobin acknowledged that Keynesian theory
faced problems, he was not at all ready to abandon the
Keynesian ship. In his view, the “consensus model” of the
early 1970s was in need of extension and refinement
rather than replacement. As noted earlier, however, the
high inflation rates of 1974-75 pushed many other economists in the direction of the monetarists.

T

18

In retrospect, it is clear that the record of 1974-75
posed big problems for both Keynesians and monetarists.
While Keynesians could try to explain the high unemployment as a consequence of insufficiently aggressive
management of aggregate demand, they could not explain
how the inflation rate had become so high when the labor
market was clearly the opposite of tight. Monetarists, on
the other hand, could blame accelerating inflation on
overly aggressive demand management but could not
explain how a too-expansionary policy could have produced such high unemployment. To make matters worse,
monetarism held that recessions were almost always
caused by monetary tightening (see Friedman and
Schwartz 1963), but if a major tightening had occurred
then the inflation rate should have fallen.
A New (and Old) Approach to Macroeconomics.
The inability of Keynesian and monetarists theories to
explain the key macroeconomic events of the 1970s
caused these theories to become discredited in the minds
of many economists. This widespread disenchantment
with traditional macroeconomic theory left the field open
for a group of young economists who were attempting to
develop a new approach to macroeconomics on the foundation provided by the classical paradigm. The research
program of these economists came to be known as neoclassical economics.13
The neoclassical attempt to build on classical principles involves formalizing many of the concepts that
have been used informally by classical and monetarist
economists. Neoclassical economics is based on the classical assumption that individual households and firms
make the decisions that maximize their well-being subject to their budget and technological constraints.
Neoclassical economists extend this assumption to
intertemporal decisions—an extension that forces them
to study the interaction between current choices and
future choices and to attempt to trace out the consequences of these choices over time. They prefer to conduct
these investigations in general equilibrium settings—
that is, in formal models that try to take into account the
complex and often simultaneous interactions among different economic variables in both the short run and the
long run.
A key principle of neoclassical economics is that in
order to determine the economic impact of a hypothetical
change in government policy—a tax cut or an increase in
the money supply growth rate, for example—it is necessary to consider the possibility that individual households
and firms may react to government policy changes by
changing the ways in which they make their own economic decisions. Neoclassical economists’ effort to
describe the nature of these changes in individual “decision rules” focuses on the manner in which the individuals formulate their economic expectations. More
specifically, a fundamental and formative assumption of

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

neoclassical economic theory is that the economic expectations of households and firms are formulated in the
most accurate possible manner, given the information
available to them—including information about changes
in government policy. This assumption is known as rational expectations.
As we have seen, the question of how workers formed
their expectations about future prices became a key issue
in the debate between the Keynesians and monetarists
over the inflation-unemployment relationship. The analysis used by the original Keynesians did not include any
formal description of the way expectations of this sort
were formed. The expectational assumption behind the
Friedman-Phelps natural rate hypothesis—a hypothesis
that was (as we have also seen) partially incorporated
into early-1970s Keynesianism—was “adaptive expectations.” Adaptive expectations is the assumption that people base their expectations about the future values of
economic variables on the past values of these variables,
emphasizing values from the recent past. In the case of
inflation, one specific adaptive expectations assumption
that was commonly used in econometric studies was that
next year’s rate of inflation was expected to be equal to a
weighted average of the values of past inflation rates,
with the weight of a particular past inflation rate declining as it receded further into history. As we note below,
because adaptive expectational assumptions do not take
into account the systematic changes in ways the public
forms its expectations that may occur when the government changes policy, results obtained using them will be
very different from those obtained using the assumption
of rational expectations.
The following two examples illustrate the potential
impact of rational expectations on the effects of government policy. First, imagine that the Smith family is considering buying a house in a particular neighborhood. The
family wants to make sure the house will bring a good
price if they have to sell it in the future. The Smiths will
probably use the price information from recent sales of
comparable homes to estimate the future resale price of
the home they are looking at. Suppose, however, that the
Smiths learn that the government has decided to build an
interstate highway extension that will, when completed,
come within a thousand feet of their prospective home.
Will they take this change in government policy into consideration when estimating the future sale price of the
home, or will they continue to concentrate exclusively on
past sale price information?
For a second example, imagine that during a mild
recession the government decides to try to stimulate the

economy by giving temporary tax breaks to families who
buy new homes. Suppose, for the sake of argument, that
this policy really does succeed in influencing potential
home buyers and that the economy actually improves as
a result. Now suppose that the government, emboldened
by the apparent success of its new policy, makes the decision to use it to combat future recessions. What will happen the next time the economy begins to slow down? Will
people remember the tax break that was offered during
the previous recession and decide to hold back on their
new-home purchases until the government decides to
offer another tax break? If they do, then the recession
may come sooner and be more severe than it would have
been otherwise, and the effects of the tax break policy
will be almost exactly
the opposite of what
the government inReal wage changes that
tended.
These examples
are not caused by policyillustrate two imporinduced changes in aggretant things about the
gate demand can confuse
ways in which rational, forward-looking
policymakers who are
individuals are likely
trying to use wage growth
to respond to changes
rates or unemployment
in government policy.
First, in projecting the
rates as guides to moneconsequences of their
tary policy.
economic decisions
individuals are likely
to consider not only
the consequences of similar past decisions but also all
the other relevant information that may be available—
including information about the effects of government
policies. When it comes to predicting inflation, for
example, people will not look exclusively at inflation
rates from the recent past, as adaptive expectations
assumed. Instead, they will also try to make use of any
information available to them about the motives and
behavior of monetary policymakers. Second, just as people will try to learn from the results of their own past
decisions, they will also try to learn from their past
observations about the effects of government policy. In
particular, people will try to distinguish unsystematic
variation in government behavior from systematic
changes in government policy. Suppose, for example,
that people discover that every time the unemployment
rate is above a certain percentage, monetary policymakers react by increasing the money supply in an effort to
reduce the rate of unemployment. It will not be long

12. For a closer look at the question of the estimation and empirical usefulness of the NAIRU, see Staiger, Stock, and Watson (1997)
and Chang (1997).
13. For summaries of some of the innovations this research program produced, see Lucas and Sargent (1979) and Miller (1995).

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

19

before both employers and employees begin to take into
consideration the effects of this policy in their wage and
salary negotiations. If the unemployment rate is above
the threshold percentage at the time of the negotiations, then the wage and salary levels that emerge from
the negotiations may include upward adjustments for
expected price increases. As a result, the final negotiated salary may be the same, in real terms, as it would
have been if the government had not acted, and the government’s actions may not end up having any effect on
the level of employment.
How did neoclassical theory view the Phillips curve?
To neoclassical economists the Friedman-Phelps critique
of Keynesian notions
about the effects of
monetary policy was a
Neoclassical economics
step in the right direcis based on the classical
tion, but only a rather
tentative step. As we
assumption that individual
have seen, Friedman
households and firms
and Phelps forced
make the decisions that
Keynesians to accept
the natural rate as a
maximize their well-being
long-run constraint on
subject to their budget
demand-management
and technological conpolicy but did not succeed in suppressing
straints.
their belief in the existence and exploitability of a short-run
Phillips curve relationship. Neoclassical economists,
however, argued that even if a statistical relationship
between inflation and unemployment did exist in the
short run, it might be impossible for the government to
exploit the relationship because people might respond to
government demand-management policy in ways that
would frustrate the goals of the policy.
The first economist to make this point was Robert
Lucas, who is generally regarded as the founder of the
neoclassical school. The formal model Lucas (1972)
developed and analyzed had three basic features that
have become characteristic of neoclassical macroeconomic theory. First, the model integrated microeconomics and macroeconomics by studying the impact of the
decisions of individual households and firms on the values of economic aggregates. Second, the model was
dynamic—that is, it took intertemporal considerations
into account, including the expectations of households
and firms. Third, the model was stochastic—that is, it
accounted for the fact that many decisions had to be
made under uncertain circumstances and that the decisions of the households and firms played a role in determining the nature of this uncertainty.
In Lucas’s model, individuals are “farmers” who
simultaneously provide labor, produce goods, and con20

sume goods. These individuals face fluctuations in prices
that are caused partly by changes in “real” economic conditions—good or bad crops—and partly by unsystematic
changes in monetary policy. The latter take the form of
random deviations from a systematic path of the money
supply. Each period, the change in the price of any particular good is caused partly by a change in real economic conditions and partly by a change in monetary policy.
Individuals would like to respond differently to price
fluctuations that come from different sources. If the relative prices of the particular goods they produce increase,
then they want to work harder and increase their production of these goods, for standard microeconomic reasons. If, however, the increase in the price of the good a
particular individual produces is simply part of an
increase in the overall price level (that is, in absolute
prices)—so that the relative price of this good has not
changed—then there is no reason for that individual to
increase his production or work effort. Thus, if individuals could distinguish relative price changes from absolute
price changes with 100 percent accuracy, then they would
never increase their work effort in response to absolute
price changes. As a result the Phillips curve for this economy would be vertical, even in the short run.
In Lucas’s model, as in most real-life situations, individuals do not possess complete information about the
current state of the economy. In particular, individuals
are assumed to be unable to observe the current prices of
any goods other than the goods they produce.
Consequently they cannot tell for certain whether
changes in the prices of “their goods” represent absolute
or relative price changes. However, individuals do know
the statistical properties of the two different types of
price fluctuations. They can use this information to calculate the average part of each price change that represents a relative price movement and then respond only to
that part of the price change.14 This is the key place
where the assumption of “rational expectations” is used
in the model.
Now suppose that, during a particular period, relative prices happen to remain entirely unchanged because
there have been no changes in real economic conditions.
At the same time, the absolute price level rises by a largerthan-normal amount because there has been a largerthan-normal increase in the money supply. Individuals
will have no way of knowing that this particular price
change is all absolute; consequently, they will proceed
under the assumption that some part of it represents a
relative price change. As a result, they will increase their
work effort in response to the price increase. The larger
the absolute price increase, moreover, the larger their
work-effort increase will be. Thus, monetary-policyinduced changes in the price level will have real effects
of a type consistent with a Keynesian-looking short-run
Phillips curve.15

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

Can monetary policymakers use this short-run
Phillips curve to increase the levels of employment and
output? Suppose that in an effort to do so they increase
the average money growth rate by some fixed percentage. If individuals are aware of this change in policy they
will realize that prices are now going to increase at a
higher average rate. As a result, the fact that the price
level increases at a higher rate next period or in subsequent periods will not surprise or confuse them, and the
policy-induced increase in the inflation rate will have no
effect on work effort. People will still respond to unsystematic price level changes in the same way they did
previously, but they will now expect a higher average
rate of inflation. The statistical Phillips curve will shift
up by the amount of the increase in the average inflation
rate, but the Phillips curve facing policymakers will be
vertical.
Lucas’s 1972 paper had a tremendous impact on the
economics profession: it is arguably the most influential
single contribution by a macroeconomist in the last fifty
years. There are two basic reasons for its significance.
The first reason, already noted, is that the paper represented a huge methodological advance in macroeconomic theory, combining as it did general equilibrium theory,
dynamic analysis, and rational expectations.16 The second reason is that he provided a qualitative explanation
for two phenomena that were both puzzling and troubling
to macroeconomists—the fact that the seemingly reliable
Phillips curve of the 1950s and 1960s had begun shifting
upward erratically at just about the time that policymakers began to try to use it to guide monetary and fiscal policy and the (closely related) fact that deliberate changes
in monetary and fiscal policy did not seem to be having
the effects on employment and output that were predicted by Keynesian theory.
What does Lucas’s theory predict about the natural
rate and the NAIRU? In his model, systematic changes in
monetary policy have no effect on the level of employment, and the labor market does not play any special role

in the mechanism by which a monetary expansion produces inflation. As a result, in the context of the model it
would not make sense for the government to focus on
unemployment rates or wage changes as guides for monetary policy.
Lucas’s paper also makes two broader points whose
potential applicability extends far beyond the specific features of his model. The first point, discussed earlier, is
that theories of the effects of government policy that are
based on the assumption that people make systematic
forecasting errors are not very sensible: since people have
strong economic incentives to correct such errors, the
changes in their behavior induced by changes in policy
are likely to disappear
very quickly as they
revise their forecastA fundamental assumption
ing schemes. This
of neoclassical economic
point had already been
made by Friedman and
theory is that the economPhelps, but Lucas’s
ic expectations of houseanalysis reinforced it
holds and firms are
in an exceptionally
stark and rigorous way.
formulated in the most
The second point,
accurate possible manner,
which was an entirely
given the information
new contribution, is
that the existence of
available to them.
statistical relationships between variables of interest to
policymakers is no guarantee that these relationships
can be exploited by policymakers, regardless of how reliable the relationships may seem to be. In Lucas’s model,
the Phillips curve is, by construction, a very reliable statistical relationship—a relationship in which the levels of
employment and output fluctuate around long-run averages that can be thought of as the analogues of the natural rate or NAIRU. However, the short-run component
of the Phillips curve relationship, which is the only

14. For example, individuals may know that, on average, one-third of the increase in the price of a good represents an increase
in the relative price of that good while two-thirds represents an increase in the absolute price level. In this case, if individuals
observe that the price of their good has increased by, say, 3 percent, then they will estimate that the relative price of the good
has increased by 1 percent and will increase their work effort accordingly.
15. Workers in Lucas’s model can be viewed as displaying a type of “money illusion”: they supply additional labor in response to
expansionary monetary policy because for a time after the policy is implemented they believe, incorrectly, that the purchasing power of their income is higher than it will actually turn out to be. Unlike the analysts who preceded him, however, Lucas
provided a rigorous explanation for the source of workers’ money illusion. This extra step was crucial because it enabled him
to ask (and answer) the question of whether the mechanism generating the money illusion would allow it to be exploited by
policymakers. As we shall see, he concluded that it would not.
16. The rational expectations assumption was developed and first used by Muth (1961). However, Lucas (1972) was the first economist to accomplish the conceptually and mathematically challenging task of including rational expectations in a dynamic
stochastic general equilibrium model. Sargent and Wallace (1975) illustrated the central importance of rational expectations
by inserting this expectational assumption into a simple macroeconomic model of an otherwise-conventional (that is, nonneoclassical) type. The results were similar to those reported by Lucas: the model generated a Phillips curve-type relationship
that government policy was powerless to exploit.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

21

component that involves changes in the levels of employment and output, is generated by forces that have nothing
to do with the systematic (policy-determined) component
of monetary policy. As a result, deliberate, policy-induced
changes in the inflation rate have no power to influence
the unemployment rate in Lucas’s model.17
Neoclassical Economics in Perspective. In the
quarter-century since Lucas published this seminal
paper, neoclassical theory has become the dominant
school of thought among academic macroeconomists. To
be sure, the neoclassical school has not escaped criticism. The rational expectations assumption, in particular, has been criticized as requiring unrealistically high
levels of economic knowledge and forecasting ability on
the part of households and firms and also because the
econometric restrictions it implies are regularly rejected by the
The significance of neodata. As a result, in
recent years there has
classical macroeconomics
been a renewed interis not that it has provided
est in the implications
anything like a definitive
of adaptive expectations, especially relamacroeconomic model but
tively sophisticated
that it has imposed more
adaptive mechanisms
rigorous scientific discisuch as least squares
learning, Bayesian
pline on macroeconomic
updating, and genetic
theorizing.
algorithms (see, for
example, Marcet and
Sargent 1989 and
Arifovic 1995). The goal of this research program is to try
to better replicate the way in which real-world individuals learn from their mistakes and adjust their expectations to changes in the economic environment.
Neoclassical use of general equilibrium models has
been criticized on the grounds that the existing versions
of these models are too simplistic and restrictive to capture the complex and diverse behavior of real-world
households and firms. A closely related criticism is that
neoclassical models simply cannot explain important
macroeconomic phenomena. For example, although the
“policy ineffectiveness” prediction of the original Lucas
article has remained a fundamental part of the neoclassical message, a great many economists continue to believe
that monetary policy has substantial real effects, and
there is a good deal of empirical evidence supporting this
position.18
In hindsight, it is clear that the significance of neoclassical macroeconomics is not that it has provided anything like a definitive macroeconomic model but instead
that it has imposed more rigorous scientific discipline on
macroeconomic theorizing. Stated differently, neoclassical macroeconomic theory is at an early stage of develop22

ment, and there are many basic questions to which it has
not yet been able to provide definitive answers. However,
it has been very successful at identifying the logical and
conceptual problems with the Keynesian and monetarist
theories that preceded it.
Neoclassical macroeconomics has made a second
major contribution to macroeconomic thought—a contribution that is less direct but perhaps equally important.
By creating skepticism among economists that monetary
or fiscal policy is responsible for business cycle fluctuations, it has forced them to recognize the possibility that
the fluctuations may be caused by real forces—that is, by
changes in technology, tastes, or resource costs of the
type that cause supply and demand curves to shift in conventional microeconomic theory. In recent years, one of
the fastest-growing branches of neoclassical macroeconomics has been real business cycle theory, which tries to
attribute cyclical fluctuations to random changes in technological productivity. Kydland and Prescott (1982) pioneered in the development of this theory, and Nelson and
Plosser (1982) provided empirical evidence that is widely viewed as indicating the importance of real as opposed
to nominal factors in driving the business cycle.19
One basic prediction of real business cycle theory is
that the observed changes in real wages and hours
worked represent fluctuations in the relative value of
labor—a prediction that has been emphasized in real
business cycle studies by Hansen (1985) and Prescott
(1986). As we have seen, this implication of the theory
provides another argument against conducting monetary
policy using rules of thumb based on the unemployment
rate or the rate of wage inflation. Another interesting
implication of neoclassical theory (though not necessarily of real business cycle theory) is that monetary policy
and fiscal policy interact so that the effects of changes in
monetary policy may depend partly or wholly on the
response of fiscal policy. The first neoclassical economists to make this point forcefully were Sargent and
Wallace (1981), who constructed a simple model in
which the inflationary implications of a change in monetary policy depended critically (and dramatically) on how
the government managed its debt. Again, this implication
of the theory suggests that any reasonable set of rules for
monetary policy guidance must be multidimensional in
nature.20

Conclusion
conomic commentators regularly urge the Fed to
use the level of unemployment or the rate of
change in wages as leading indicators of inflation
and as guides to whether they should ease or tighten
monetary policy.21
The logic behind this approach is based on modern
(post-1970s) Keynesian macroeconomics and, more
specifically, on the Phillips curve and the NAIRU.

E

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

According to this view, inflation is caused by excessive
“aggregate demand,” and changes in aggregate demand
show up first in the labor markets. Low levels of unemployment—levels below the natural rate/NAIRU—
reflect the fact that excessive aggregate demand has
produced a tight labor market. A tight labor market will
put upward pressure on wages. Increases in wages will
force firms to increase their prices and will consequently produce a higher rate of inflation. Since modern
Keynesianism sees the state of monetary policy as the
principal determinant of the level of aggregate demand,
a tight labor market also reflects excessively expansionary monetary policy and indicates the need for corrective
Fed tightening.
This article has attempted to provide some basic
information about this NAIRU theory of the causes of
inflation and the role of monetary policy. We began by
describing the Phillips curve, an apparent empirical
relationship between wage increases and unemployment
that Keynesian economists used as the basis for a theory
of the inflation-unemployment relationship. The theory
implies that policymakers could use demand stimulus or
restraint to produce lower or higher unemployment at the
cost of higher or lower inflation. Monetarist economists,
who were deeply skeptical of Keynesian views about the
effectiveness of demand management, developed a critique of the Phillips curve that was based on the concept
of a “natural” rate of unemployment. According to the
monetarists, attempts to use monetary or fiscal policy to
keep the unemployment rate below the natural rate
might have limited success in the short run but in the
long run would produce continually increasing inflation.
The stagflation (simultaneous high inflation and
high unemployment) that afflicted the U.S. economy dur-

ing the 1970s shook economists’ faith in the existence of
a stable Phillips curve and greatly increased the credibility of the monetarist “acceleration hypothesis.” The proponents of Keynesian theory weathered the monetarist
critique by accepting the natural rate—which they
rechristened the NAIRU—as a long-run constraint on
demand-management policies that, in their view,
remained effective in the short run. Although the monetarists were not satisfied with the Keynesians’ response to
their critique, the fact that the two schools of macroeconomic thought were working with a common set of theoretical weapons prevented the monetarists from
overwhelming the Keynesians’ defenses. As a result, the
modified Keynesian theory of the 1970s became the standard theory taught to economics students and used by
policymakers. A basic feature of this theory was a simple
rule of thumb for monetary policy: tighten policy when
the unemployment rate is below the NAIRU or when real
wages are rising, and ease policy when the reverse is true.
The low unemployment rates observed in the mid-1990s
have caused many commentators to urge the Fed to consider using this rule of thumb as a justification for preemptive monetary tightening.
After describing the historical development of the
NAIRU theory, the discussion raises some practical questions about the validity of the theory and its usefulness
as the basis for policy advice. Perhaps the most important question involved the difficulty of distinguishing
policy-induced changes in absolute wages from changes
in relative wages associated with real changes in the economy—changes that it would not make sense for monetary
policymakers to attempt to oppose. A second question
focused on the fact that there is very little empirical evidence supporting the notion of sticky prices on which

17. In his paper, Lucas imagines a researcher who tries to use a statistical analysis of data from his model to provide advice to policymakers. The researcher runs a linear regression with output or employment as the dependent variable and the inflation
rate as the independent variable. He finds that the coefficient estimate for the inflation rate is positive and consequently advises policymakers that using monetary policy to increase the inflation rate is likely to succeed in increasing the levels of employment and output. As we have seen, however, this policy advice is incorrect.
18. Strictly speaking, neoclassical theory does not preclude monetary policy from having real effects: it simply rules out real
effects that rely on frictionless markets not clearing or on the public being systematically fooled. Thus, although it is arguably
fair to describe monetary policy ineffectiveness as a characteristic feature of neoclassical theory, there are an increasing number of neoclassical models in which monetary policy has short-run real effects. Leeper and Gordon (1992) and references therein are examples of key contributors to the rapidly growing “liquidity effects” literature, which uses real business cycle models
(see below) to study the short-run effects of changes in monetary policy. There are also a few neoclassical models in which monetary policy has long-run real effects. Examples of the latter type include Wallace (1984), Bhattacharya, Guzman, and Smith
(1996), Espinosa and Russell (1997a, 1997b), and Bullard and Russell (1997).
19. For a more detailed description of real business cycle theory and a review of the formative developments in the theory see
Prescott (1986).
20. The following statement by former Federal Reserve Governor Lawrence Lindsey provides a good example of unidimensional
reliance on the NAIRU: “The NAIRU is a useful theoretical construct . . . sufficient for making quick ‘on your feet’ estimates of
likely economic performance. . . . If I knew with certainty that the NAIRU was 5.837. . . I would have the information I needed
to know with certainty that I should tighten” (1996, 10).
21. Ironically, experts who specialize in studying the properties of the business cycle classify wages as lagging rather than leading indicators. See Moore (1961) and Zarnowitz (1992).

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

23

Keynesian theory is based, and a third involved the
empirical weakness of the Phillips curve relationship that
provides the basis for the NAIRU.
The discussion also includes neoclassical economics, a relatively new school of macroeconomic thought
that has provided a second, more fundamental challenge
to Keynesian thought. We described the fundamental
principles of neoclassical theory and went on to explain
how Robert Lucas, one of the theory’s founders, used
these principles to construct a groundbreaking theoretical model whose properties cast doubt on the short-run
effectiveness of monetary policy and thus on the usefulness of monetary policy rules based on the NAIRU.
Neoclassical theory still has a large number of basic
macroeconomic questions to answer. However, it has produced huge logical and methodological improvements in
macroeconomic analysis, and it has left the Keynesian
and monetarist theories that preceded it largely discred-

ited—including the modern form of Keynesian theory
that provides the basis for the NAIRU. Recent developments in neoclassical theory indicate that business cycle
fluctuations in employment and output may be caused
primarily by real forces—a situation that, if true,
increases the danger that monetary policy based on the
NAIRU may interfere with the proper functioning of the
price system.
Our own view is that proponents of the NAIRU have
never provided anything like a satisfactory answer to the
neoclassical critique, or even to the questions raised in
this article. Given that this is the case, it is hard to give
much credence to the commentators who urge the Fed to
base its monetary policy on the NAIRU. Unfortunately,
neoclassical economists have yet to provide monetary
policymakers with reliable policy rules to replace NAIRUbased rules. Until they do, monetary policy decision making will remain a difficult task.

R E F E R E N C E S
ARIFOVIC, JASMINA. 1995. “Genetic Algorithms and Inflationary
Economies.” Journal of Monetary Economics 36 (December):
219-43.

FRIEDMAN, MILTON, AND ANNA J. SCHWARTZ. 1963. A Monetary
History of the United States, 1867-1960. Princeton, N.J.:
Princeton University Press.

BALL, LAURENCE, AND N. GREGORY MANKIW. 1995. “Relative-Price
Changes as Aggregate Supply Shocks.” Quarterly Journal of
Economics (February): 161-93.

HANSEN, GARY. 1985. “Indivisible Labor and the Business Cycle.”
Journal of Monetary Economics 16 (November): 309-27.

BHATTACHARYA, JOYDEEP, MARK G. GUZMAN, AND BRUCE D. SMITH.
1996. “Some Even More Unpleasant Monetarist Arithmetic.”
Cornell University, Center for Analytic Economics, CAE
Working Paper 95-04, April.
BULLARD, JAMES, AND STEVEN RUSSELL. 1997. “How Costly Is
Sustained Low Inflation for the U.S. Economy?” Federal
Reserve Bank of St. Louis and IUPUI Working Paper.
CHANG, ROBERTO. 1997. “Is Low Unemployment Inflationary?”
Federal Reserve Bank of Atlanta Economic Review 82 (First
Quarter): 4-13.
ESPINOSA, MARCO A., AND STEVEN RUSSELL. 1997a. “Can Higher
Inflation Reduce Real Interest Rates in the Long Run?”
Canadian Journal of Economics (forthcoming).
———. 1997b. “Conventional Monetary Policy Wisdom in the
Diamond Model.” Federal Reserve Bank of Atlanta, working
paper, forthcoming.
FISHER, IRVING. 1926. “A Statistical Relationship between
Unemployment and Price Changes.” International Labor
Review 13 (June): 785-92.
FRIEDMAN, MILTON. 1968. “The Role of Monetary Policy.”
American Economic Review 68 (March): 1-17.
———. 1976. “Wage Determination and Unemployment.”
Chap. 12 in Price Theory. Chicago: Aldine Publishing Company.

24

KEYNES, JOHN M. 1964. Reprint. The General Theory of
Employment, Interest, and Money. San Diego: Harcourt Brace
and Company. Originally published, 1953.
KYDLAND, FINN, AND EDWARD PRESCOTT. 1982. “Time to Build and
Aggregate Fluctuations.” Econometrica 50 (November): 134570.
LEEPER, ERIC M., AND DAVID B. GORDON. 1992. “In Search of the
Liquidity Effect.” Journal of Monetary Economics 29 (June):
341-69.
LINDSEY, B. LAWRENCE. 1996. “NAIRU Disrobed.” International
Economy (March/April): 8-13.
LUCAS, ROBERT E., JR. 1972. “Expectations and the Neutrality of
Money.” Journal of Economic Theory 4 (April): 103-24.
LUCAS, ROBERT E., JR., AND THOMAS SARGENT. 1979. “After
Keynesian Macroeconomics.” Federal Reserve Bank of
Minneapolis Quarterly Review 3 (Spring): 1-16.
MARCET, ALBERT, AND THOMAS SARGENT. 1989. “Convergence of
Least-Square Learning Mechanisms in Self-Referential Linear
Stochastic Models.” Journal of Economic Theory 48 (August):
337-68.
MODIGLIANI, FRANCO, AND LUCAS PAPADEMOS. 1975. “Targets for
Monetary Policy in the Coming Year.” Brookings Papers on
Economic Activity, no. 1:141-63.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

MOORE, GEOFFREY. 1961. Business Cycle Indicators. Vol. 1 of
Contributions to the Analysis of Current Business Conditions,
National Bureau of Economic Research. Princeton, N.J.:
Princeton University Press.
MUTH, J.F. 1961. “Rational Expectations and the Theory of
Price Movements.” Econometrica 29 (July): 315-35.
NELSON, CHARLES R., AND CHARLES I. PLOSSER. 1982. “Trends and
Random Walks in Macroeconomic Time Series: Some Evidence
and Implications.” Journal of Monetary Economics 10:139-62.
OKUN, ARTHUR M. 1980. “Postwar Macroeconomic Performance.”
In The American Economy in Transition, edited by Martin
Feldstein, 162-69. Chicago: University of Chicago Press.
PHELPS, EDMUND. 1967. “Phillips Curves, Expectations of
Inflation, and Optimal Unemployment over Time.” Economica
34 (August): 254-81.

SAMUELSON, PAUL A., AND ROBERT M. SOLOW. 1960. “Analytical
Aspects of Anti-Inflation Policy.” American Economic Review
50 (May): 177-194.
SARGENT, THOMAS, AND NEIL WALLACE. 1975. “Rational
Expectations, the Optimal Monetary Instrument, and the
Optimal Money Supply Rule.” Journal of Political Economy 83
(April): 241-54.
———. 1981. “Some Unpleasant Monetarist Arithmetic.”
Federal Reserve Bank of Minneapolis Quarterly Review 5
(Fall): 1-17.
STAIGER, DOUGLAS, JAMES STOCK, AND MARK WATSON. 1997. “The
NAIRU, Unemployment and Monetary Policy.” Journal of
Economic Perspectives 11 (Winter): 33-50.
TOBIN, JAMES. 1980. “Stabilization Policy Ten Years After.”
Brookings Papers on Economic Activity, no. 1:19-71.

PHILLIPS, A.W. 1958. “The Relationship between Unemployment
and the Rate of Change of Money Wage Rates in the United
Kingdom, 1861-1957.” Economica 25 (November): 283-99.

WALLACE, NEIL. 1984. “Some of the Choices for Monetary Policy.”
Federal Reserve Bank of Minneapolis Quarterly Review 8
(Winter): 15-24.

PIGOU, A.C. 1933. The Theory of Unemployment. London:
Macmillan.

WYNNE, MARK. 1995. “Sticky Prices: What Is the Evidence?”
Federal Reserve Bank of Dallas Economic Review (First
Quarter): 1-12.

PRESCOTT, EDWARD. 1986. “Theory Ahead of the Business Cycle.”
Carnegie-Rochester Conference Series on Public Policy 25
(Autumn): 11-44.
SAMUELSON, PAUL A., AND WILLIAM NORDHAUS. 1989. Economics.
13th ed. New York: McGraw-Hill.

ZARNOWITZ, VICTOR. 1992. Business Cycles: Theory, History,
Indicators, and Forecasting. Vol. 27 of Studies in Business
Cycles, National Bureau of Economic Research. Chicago:
University of Chicago Press.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

25

Identifying Monetary
Policy: A Primer
TA O Z H A
The author is an economist in the macropolicy section
of the Atlanta Fed’s research department. He is grateful
to Roberto Chang, Frank King, Eric Leeper, Larry Wall,
and especially Jerry Dwyer and Mary Rosenbaum for
valuable comments.

T

HE POPULAR PRESS AND UNDERGRADUATE ECONOMICS TEXTBOOKS HAVE LONG CONCLUDED
THAT AN INCREASE IN THE FEDERAL FUNDS RATE TARGET BY THE

COMMITTEE (FOMC)
TIONARY PRESSURES.

FEDERAL OPEN MARKET

TENDS TO SLOW GROWTH OF NATIONAL OUTPUT AND REDUCE INFLA-

ECONOMISTS

GENERALLY AGREE ON THIS POINT, BUT THEY DISAGREE

CONSIDERABLY ABOUT THE QUANTITATIVE IMPACT OF MONETARY POLICY.

FOR EXAMPLE, A GROUP OF ECON-

OMISTS CALLED MONETARISTS ARGUE THAT “IN THE SHORT RUN, WHICH MAY BE AS LONG AS THREE TO TEN
YEARS, MONETARY CHANGES AFFECT PRIMARILY OUTPUT” BUT NOT PRICES

(FRIEDMAN 1992, 48)

WHILE

OTHER ECONOMISTS SUCH AS REAL BUSINESS CYCLE THEORISTS POSTULATE THAT MONETARY CHANGES
AFFECT ONLY PRICES BUT HAVE LITTLE OR NO EFFECT ON OUTPUT

As it stands, economists’ beliefs about the quantitative importance of monetary policy stem largely from theoretical models through which the policy effects of
changing monetary policy are inferred. It is no surprise,
then, that different conclusions arise from different
experiments or theories. The actual economy, however,
is not the result of any such controlled experiment.
Obviously, a central bank cannot change policy for the
sake of examining its effect on the economy. In the real
world, inferences about the quantitative effect of monetary policy must rely on observations of actual economic
activity in which many variables are changing simultaneously. What can be observed is the equilibrium outcome
of interaction among all players in the economy—the
central bank, financial market participants, producers,
and consumers. On this playing field, sorting out the
central bank’s behavior from that of the many other
participants is the first and critical step in attempting
26

(COOLEY AND HANSEN 1995).

to estimate the actual impact of monetary policy. This
sorting-out process is known in technical parlance as
identification.
Identification of monetary policy is partly a conceptual (economic) issue and partly an empirical (technical) one. Conceptually, the process requires that one
understand the economics of the demand and supply of
money, or, in other words, the interaction between the
central bank’s reaction to economic conditions and the
private sector’s response to policy actions. Empirically,
one needs sophisticated mathematical tools to isolate the
central bank’s behavior from all other behaviors in the
observed data and examine its consequences.
The purpose of this article is to explain these two
issues: the conceptual one of why identification of monetary policy is important and the empirical issue of how
difficult it is in practice. The article focuses on these two
issues exclusively because of how vital careful identifica-

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

CHART 1

Demand for and Supply of Wine
p
S

P r i c e

tion is for an accurate assessment of policy effects. Given
this purpose, the article refrains from discussing how to
resolve the disparate views about the actual quantitative
effect of monetary policy on a given country’s economy.
The discussion first explores identification of monetary
policy as having much in common with issues familiar to
us from basic economics principles. The article then discusses the identification issue special to the analysis of
monetary policy and illustrates the process with a few
examples of identifying monetary policy in different
countries.

D

Demand and Supply
he abstract concept of money is clear: money is
something the public accepts in exchange for
goods and services. In reality, however, the measure of money is not so well defined: money can be currency in circulation, reserves, the monetary base (the
sum of currency in circulation and reserves), M1 (currency plus checkable deposits), or M2 (M1 plus other
assets). Whichever monetary aggregate is used, the
analysis of monetary policy inevitably encounters the two
blades of the monetary scissors: demand for and supply of
money. Thus it is appropriate to begin exploring the
importance of identifying monetary policy with an analysis of the demand and supply of money.
A simple, familiar example is instructive: the
demand-supply relationship in the market of goods and
services, in this case the wine market. If one has the data
on the price of wine (p) and the quantity bought and sold
(q), the bivariate demand-supply relationship can be
described by the following two equations:

T

q = a1p + a2 X + ed (demand),

(1)

q = b1p + b2 Y + es (supply),

(2)

where X is a set of variables (such as the government’s
excise tax and consumers’ income) that affect the
demand for wine, Y is a set of variables (such as the government’s excise tax and weather condition) that affect
the supply of wine, the a coefficients describe the behavior of consumers, and b, the behavior of producers.
Before proceeding, explaining a few common notations and notions will lay the groundwork for discussion
of these and additional equations. The notations q, p, X,
and Y in equations (1) and (2) are variables while a1, a2,
b1, and b2 are coefficients. The sharp distinction between
the “coefficient” and “variable” is an important one. A
variable has a quantitative value observed in the data so
that, for example, the variable q represents the price of
wine bought and sold. A coefficient does not come directly from the data; rather, its quantitative value must be
obtained by statistical methods. The process of obtaining
the value of a coefficient is called estimation, a concept

q
Q u a n t i t y

B o u g h t

a n d

S o l d

important throughout the article. For instance, coefficients such as a1 are to be estimated through equations
(1) and (2). Finally, the notation ed represents a random
change that cannot be described by normal demand
behavior, and this article calls ed a demand shock. The
word shock has its familiar meaning of referring to something unpredictable. Similarly, the supply shock es in
equation (2) indicates an unpredicted change in the supply of wine.
Economic theory implies that price is inversely related to quantity demanded (that is, a1 < 0). It also tells us
that when the price is higher, the firm is willing to produce more wine (b1 > 0). These relationships can be
depicted in a two-dimensional figure like Chart 1, where
the downward-sloping curve represents demand behavior
and the upward-sloping curve represents supply behavior.
Chart 1 is drawn under the assumption that variables
other than p and q are held fixed. Therefore, if there are
any changes in X or Y, the curves in Chart 1 will shift from
the original equilibrium position. For example, when the
government raises the tax on wine, both demand and supply will fall, their curves will shift to the left, and the
equilibrium will change from E1 to E2 (Chart 2). How
much the quantity of wine will be reduced (from q1 to q2)
in the market depends on the behavior of consumers and
producers, which is described in equations (1) and (2).
In other words, it depends on how far the demand and
supply curves in Chart 2 will shift. The policy analyst, to
assess the tax’s effect on the behavior of consumers and
producers, must understand (correctly identify) both the
demand function (1) and the supply function (2).
The argument about the importance of identifying
the demand for and supply of wine can be carried over to
the money market, although monetary analysis is of
course far more complicated (see, for example, Leeper
1992). To begin with, one can think of the quantity of

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

27

C H A R T 2 Effect of the Government’s Tax on Wine
S2

p

P r i c e

S1

E2

p2

E1

p1

D1
D2

q1

q2
Q u a n t i t y

B o u g h t

money as resembling the quantity of wine and the opportunity cost of holding money (the interest rate) as the
price of wine. Let M represent money and R, the nominal
interest rate. Thus, analogous to the wine market, where
q and p are jointly determined, M and R are determined
by both demand and supply in the money market. Assume
that all deposits in M do not bear interest.1 The demand
function for money derived in standard textbooks can be
expressed as
M – P = a1y + a2R + eMD (money demand),

(3)

where y represents national output; P, the general price
level; a1, the coefficient of y; a2, the coefficient of R; and
eMD, the money demand shock.2 The coefficient a1 measures the percent change in demand for money in
response to a percent change in output y, and the measure is known as the income elasticity of money demand.3
As consumer income rises, the demand for goods and services will increase, and in turn their demand for money
will rise so that they can purchase goods and services.
Thus, the coefficient a1 is expected to be positive. The
coefficient a2, known as the interest elasticity of money
demand, measures the percent change in money demand
in response to a percent change in the interest rate. Since
the public is willing to hold less (real) money (M – P)
when the cost (R) of holding money increases, the interest elasticity a2 is expected to be negative. If one depicts
the demand curve in the (M, R) plane (see Chart 3), the
curve of demand for money has a negative slope (analogous to the downward-sloping curve in the demand for
wine in Chart 1).
When broad monetary aggregates such as M1 or M2
are used, the term money supply in general involves not
merely the behavior of the central bank but also the
28

a n d

q
S o l d

behavior of banks and other financial institutions whose
liabilities (such as checking deposits) serve as part of the
medium of exchange as well as the behavior of depositors
who decide how much currency to hold in relation to
deposits. A central bank, through its open market operations or discount window lending, can affect monetary
aggregates through the banking system. To see how a monetary aggregate such as M2 is affected, suppose that the
Federal Reserve decides to increase the monetary base
(the sum of the currency in circulation and reserves) by
buying Treasury securities worth, say, $x from a seller.
Suppose the seller, now becoming a depositor, decides to
deposit the full amount of $x in Bank A, creating $x in
deposits in the bank. After meeting the reserve requirement (that is, the certain percentage of $x that must be
kept in the bank), Bank A lends part of the deposit to
households who, now becoming depositors, decide to
deposit the loans in, say, Bank B. The process can continue. Eventually, such a chain of deposit expansions
through bank loans makes an increase in M2 a multiple of
the initial increase in the monetary base. Thus the term
money multiplier, defined by the ratio of the monetary
aggregate (like M2) to the monetary base, is used to indicate the extent to which money is created or multiplied
through the participation of both banks and depositors.
The incentive to increase deposits in the banking
system lies in the prospect of making profitable loans. If
the prospect is dim or the demand for loans falls off,
banks may not create deposits up to the limit the reserve
requirements allow. Thus they may, from time to time,
have excess reserves in addition to required reserves.
Furthermore, because of the uncertainty about deposit
flows and transaction clearing within a day and from day
to day, banks typically hold some excess reserves
although there are incentives to minimize them. Clearly,

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

M = a3R + a4 Xs + eMS (money supply),

(4)

where the coefficient a3 is the interest elasticity of money
supply, the coefficient a4 is the elasticity in relation to Xs,
which is a set of variables (such as reserves and output)
that can influence the supply of money, and eMS is the
money supply shock, which is uncorrelated with the
money demand shock eMD. The a coefficients are to be
estimated, and the sign of a3 is expected to be positive.
One interpretation for the positive sign of a3 follows the
logic in Box 1: since each dollar of excess reserves is costly to hold because of forgone interest, the amount of
excess reserves tends to decline as the rate of interest
rises. Through the money multiplier effect, deposits and
monetary aggregates tend to increase. Thus, one would
expect an upward-sloping curve for the money supply
function as depicted in Chart 3.
From the viewpoint of the policymaker the question
is, Why is it important to separate the demand for money
from the supply of money? Remember that in the example of the wine market, distinguishing the demand behavior of consumers and the supply behavior of producers is
important for assessing the effect of the government’s tax
on the behavior of consumers and producers (recall
Chart 2). Here, the central bank needs to assess policy
effects in order to attain its objective. When policy
actions shift the money supply curve (for example, from
MS1 to MS2 in Chart 4), the change in the equilibrium
quantity of money and the equilibrium rate of interest
(from E1 to E2 in Chart 4) depends on two factors: the
slope of the money demand curve as well as the slope of
the money supply curve. Thus this section has shown that
understanding the demand and supply of money is crucial
for assessing (identifying) policy effects on, say, the
quantity of money (M) and the rate of interest (R).

CHART 3

R
MS

I n t e r e s t

R a t e

Money Demand and Money Supply

N o m i n a l

banks’ decisions about how much needs to be held as
excess reserves, combined with depositors’ decisions
about how their portfolio should be allocated, can cause
the supply of money to change. The central bank is not
the only player whose behavior influences the money
supply or the money multiplier. Taking all these behaviors
into account, the supply function for money can be
derived from the money multiplier (see Box 1 on page 32)
and usually has the following form:

MD

M
M o n e y

Central Banks’ Behavior
he distinction between the demand for and supply
of money, explained in the monetary model (3)-(4),
is analogous to the intuitive demand-supply analysis in the wine market. What the model does not show is a
problem unique to policy analysis: separating the central
bank’s behavior from the behavior of the banking system
and depositors. It is the purpose of this section to discuss
in detail the behavior of the central bank.
For simplicity of analysis, textbooks usually use the
money supply curve in Chart 3 to represent the central
bank’s behavior under either of the following two assumptions. One is that the central bank is in complete control of
a broad monetary aggregate like M2, in which case the
money supply curve is vertical (Chart 5). The other
assumption is that unpredicted changes in reserves are
caused solely by unpredicted monetary policy shifts (eMS)
and that these changes are the only random sources affecting (that is, shifting) the money supply curve in Chart 3.
Two caveats are in order. First, these two assumptions are generally not a good description of what actually happens. The central bank can heavily influence
broad monetary aggregates such as M2 but cannot control them completely. Moreover, unpredicted changes in
reserves can be caused by shifts in banks’ demand for
reserves or by depositors’ portfolio adjustment and thus

T

1. This assumption is made for the convenience of analysis. Some deposits such as savings accounts in M2 do pay interest on their
balances. Then, the cost of holding these assets is reflected by the difference between the interest rate on these assets and the
interest rate on other assets such as government bonds. This feature would complicate the analysis but not alter the basic
conclusions.
2. All variables discussed in this article except for interest rates are logarithmic. Thus, M – P is the log of real balances (the money
stock deflated by the general price level).
3. Note the adherence to the convention of using the term income instead of output; both names denote the variable y. The concept
of elasticity, the percent change in one variable in response to a percent change in another variable, is frequently used in the
article.
Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

29

C H A R T 4 Shift of the Money Supply
R

E1

R1

MS2

E2

R2

N o m i n a l

I n t e r e s t

R a t e

MS1

MD

M1

M2

M

M o n e y

do not necessarily indicate policy shocks. Second, even if
these two assumptions are reasonable, the money supply
function discussed does not to this point describe the real
world of the central bank’s behavior.
What is the real world behavior of the central bank,
after all? More important, to what extent can a policy
analyst write down a function (functions) or equation
that gives a good approximation of that behavior? The
macroeconomic policy aspect of many central banks’
behavior reflects both their responsibility for controlling
inflation and their attention to policy’s effect on overall
economic activity. In the day-to-day implementation of
U.S. monetary policy, for example, the Federal Reserve
sets a target for the federal funds rate according to its
objective. Its attempts to meet the target require tracking
the amount of reserves and subsequently of deposit flows
and monetary aggregates. In choosing its target, the
Federal Reserve regularly examines economic forecasts
prepared by its staff. The staff frequently explore the
historical relationships between key macroeconomic
variables (such as inflation and output) and policy
instruments (such as reserves and the federal funds
rate) and provide alternative economic outlooks under
different assumptions about policy instruments such as
different levels of the federal funds rate. Policymakers
then decide what actions to take in order to attain their
objectives.
The process of such policymaking is common across
different industrial countries. For example, for the Bank
of France, senior management “assesses the reserve position of the banking system and evaluates whether current
market interest rates, especially the interbank rate, are
consistent with the current stance of monetary policy and
foreign rates. Instructions are then given to the money
market trading room at the Bank of France to intervene
30

in the interbank market on the basis of the evaluations of
money market and general macroeconomic conditions”
(Batten and others 1990, 78). The Bank of Canada “uses
economic projections to translate the Bank’s objectives
into suggested paths for the instruments of policy, and
uses various economic and financial indicators, notably
monetary aggregates, to monitor progress and help the
Bank to act in a timely fashion when necessary” (Duguay
and Poloz 1994, 197).
In short, a central bank tries to achieve its objective
subject to the constraints imposed by the private sector’s
activity. As a result, the central bank comes out with a
strategy or plan by reviewing the state of the economy.4
This article refers to this strategy as the policy reaction
function(s) and henceforth uses it to mean monetary policy or the central bank’s behavior throughout. The reaction function is therefore composed of two components:
the systematic reaction of policy to economic conditions
and unexpected shifts in policy (policy shocks).
The discussion first turns to the systematic component of monetary policy because it is the essence of the
specification (that is, description) of a reaction function.
For illustration, suppose the Federal Reserve’s objective
is to stabilize inflation at some low level with the federal
funds rate as a policy instrument. Since the Federal
Reserve has no direct control over the general price level,
it uses the federal funds rate to influence intermediate
targets such as the three-month Treasury bill rate and
M2. Unfortunately, there is no simple linkage of the federal funds rate with M2 and of M2 with the general price
level, at least in the short run. The price level today is
affected not only by current and past movements in other
variables such as the federal funds rate, M2, and output
but also by previous changes in the price level itself. The
changes in all the variables reflect the interaction

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

where FR stands for the federal funds rate; M, for M2; R,
the three-month Treasury bill rate; and X, a set of other
crucial variables that are used by the Federal Reserve to
predict fluctuations in the general price level.5 To give an
example of which variables are contained in X, suppose
monthly data are used to estimate the b coefficients in
the Federal Reserve’s reaction function (5). The set X
should include all the crucial variables the Federal
Reserve uses in making its policy decisions. The variables
that may be excluded in X are the price level and output
in the present month, on the grounds that the data on
these variables are available only after the end of the
month. Similarly, one should include in the reaction function not only variables (such as the federal funds rate,
commodity prices, M2, exchange rates, output, and the
price level) in the previous months but also variables
such as commodity prices, M2, and exchange rates in the
present month. Current data on commodity prices and
financial variables provide the Federal Reserve with
information about the market’s expectation of future
inflation while the data from the previous months help
predict future economic activity.
Given the systematic component b1M + b2R + b3X in
the policy reaction function (5), the sign of the coefficient b1 is particularly interesting because it indicates
how the federal funds rate responds to a change in the
monetary aggregate. Suppose there is an increase in the
monetary aggregate. If the central bank believes that
such an increase will lead to a rise in future inflation, it
will tend to increase the federal funds rate in order to offset the rising monetary aggregate. The sign of b1 is therefore expected to be positive.
The second component of equation (5) is the random shock eMS, which reflects an unpredicted shift in
monetary policy. The notion of randomness here is the
same as when newspapers use the term shock to refer to
an oil shock, which appears random because it is unpredictable. Likewise, policy shocks occur when the central
bank’s instrument changes unpredictably. To explain fur-

R a t e

R
MS

I n t e r e s t

FR = b1M + b2R + b3X + eMS (policy reaction), (5)

CHART 5

Perfectly Inelastic Money Supply (a3=0)

N o m i n a l

between policy actions and economic activity in the current and previous periods. To attain price stability, the
Federal Reserve will adjust its federal funds target in
response to changes in all crucial economic variables
such as M2, the price level, and output. The reaction
function can therefore be summarized as

M
M o n e y

ther, consider the Federal Reserve’s policy. Suppose the
Federal Reserve’s objective is to keep inflation low in the
long run and its policy instrument is the federal funds
rate. In the short run (say, three to ten years), however,
the dynamic relationships between output, unemployment, inflation, and the federal funds rate are complicated, and the trade-off between inflation and output may be
substantial and is uncertain. A policy decision reached by
sifting through such uncertain relationships can be as
unpredictable as any other economic condition.6 Such an
unpredicted movement in the federal funds rate is called
a policy shock—eMS in equation (5)—while the predicted movement (systematic reaction) is characterized by
b1M + b2R + b3X.
Note the close connection between the functional
forms (4) and (5): the reaction function (5) can be
rearranged to have the same functional form as the
money supply function (4), and Xs in (4) can then be
thought of as including both X and the federal funds rate
(FR) in (5). Is the sign of a3 in the newly derived function (4) positive as in the original money supply function? Recall the argument for the positive sign of b1 in
equation (5): the Federal Reserve tends to increase the
federal funds rate in order to offset the rising monetary
aggregate (leaning against the wind, so to speak). When
the federal funds rate (FR) is expected to rise, the threemonth Treasury bill rate (R) will tend to rise because
there is a strong positive relationship between R and
expected FR.7 Thus one should expect the positive sign of

4. Formally, one can think of the strategy coming from the first-order conditions in the central bank’s maximization problem in
a theoretical model.
5. The term eMS will be discussed later.
6. See Leeper, Sims, and Zha (1996) for further discussions.
7. Such a relationship is known as the expectation theory of the term structure in the economic literature. For details of the theory, the reader can consult any standard textbook of monetary economics or finance—for example, Mishkin (1992).

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

31

B O X

1

Traditional Approach to the Supply of Money
n standard textbooks, “money supply” typically refers to
the joint behavior of the central bank, commercial banks,
and depositors. The derivation of the money supply function
or equation in this box draws directly from McCallum (1989,
55-73). Suppose that M1 is the definition of money and the
central bank resembles the Federal Reserve. By definition,

I

M = C + D,

where the ratios cr, rr, and k are assumed to be constant (or
are not influenced by other variables if changing over time).
The assumption that e(R) decreases with R is based on the
belief that the banks will hold less excess reserves when the
interest rate R rises. A rise in R indicates the opportunity
cost of holding the excess reserves.
Using (A4), one can rewrite equation (A1) as

(A1)
M = (cr + 1)D.

where C stands for the currency in circulation (held by the
nonbank private sector) and D the checkable deposits. The
monetary base MB that is heavily influenced by the central
bank is
MB = C + TR.

(A2)

Note that TR stands for the total reserves, which can be further broken into two components:
TR = RR + ER,

Combining (A2), (A3), (A4), (A5), (A6), and (A7) leads to
MB = [cr + k + e(R)]D.

M
cr + 1
=
.
MB cr + K + e( R)

(A10)

Using the expression m(R; k, cr) to summarize the righthand term of equation (A10) yields the simple money supply
function:
M = m(R; k, cr)MB.

C/D = cr,

(A9)

The money multiplier, defined by the ratio of money to the
monetary base, can be derived from equations (A8) and
(A9):

(A3)

where RR is the required reserves and ER is the excess
reserves.1 The key relationships between the deposits and
the other variables are

(A8)

(A11)

(A4)

TR/D = rr,

(A5)

RR/D = k,

(A6)

ER/D = e(R), e´(R) < 0,

(A7)

It is obvious from (A7) and (A10) that the money supply
function (A11) implies the upward-sloping curve of money
supply. Thus, the function (A11) can be written in the form
of equation (4), where the condition a3 > 0 reflects the
upward-sloping curve of the money supply.

1. Some central banks, like Canada’s, no longer have the legal reserve requirement. But such banks still hold reserves in response to the
withdrawal of deposits. In the case of Canada, RR can be thought of as the desired reserves—the amount the banks desire to hold (see
Barro and Lucas 1994).

32

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

a3 and the upward-sloping curve of the newly derived
function (4) when depicted in the two-dimensional (M,
R) chart. Now, however, the new function (4) has a different interpretation: it describes the policy behavior, not
a joint behavior of the central bank, commercial banks,
and depositors. Despite the nuances in interpreting the
same functional form (4), the practice of calling the reaction function “money supply” is very common because it
is always intuitive to think of demand and supply.
Accordingly, this article shall continue to interchange the
terms.
In summary, this section discusses the behavior of
the central bank, emphasizes the significance of understanding the policy’s systematic response to economic
conditions, and shows how such behavior can be modeled
or approximated by the policy reaction function (5).
Moreover, it reinterprets the traditional money supply
function discussed in the previous section as the policy
reaction function but does so without changing the characteristics of the money supply function (for example,
the upward-sloping curve of the supply function). Given
this reinterpretation, one is able to analyze the effect of
monetary policy in the intuitive framework of demand
and supply, as will be shown in later sections.

Other Points about Identifying Monetary Policy
he antecedent sections establish the importance of
identifying monetary policy (separating money
demand from money supply) and describe how the
money supply function (4) can be used to approximate a
central bank’s behavior. Even so, the point about the
importance of identifying monetary policy is still often
misunderstood. Two popular contentions merit further
discussion.
One position is that central banks know exactly what
their monetary policy or behavior is and from their viewpoint there is no need to separate the money demand and
money supply. For example, if a central bank’s objective
is a commitment to price stability, monetary policy is to
make the general price level stable, and thus the private
sector’s behavior (such as the demand for money) is not
the policymaker’s concern. While this notion seems prima
facie sensible, it is simply incorrect. It confuses the central bank’s objective with its policy, which, as the discussion in the last section argues, is a strategy designed to
achieve the objective. The real issue is not whether the
central bank knows its objective; the real issue is whether
it knows how to form a strategy (monetary policy) to
attain its objective. The formation is difficult and
requires a thorough understanding of the interaction
between the central bank’s behavior (money supply) and
the private sector’s activity.
Consider, for example, Canadian monetary policy.
Analysis of Canadian monetary policy is instructive not
only because most countries resemble Canada in the

T

sense that they are small and open relative to the U.S.
economy but also because it is of interest to U.S. policymakers as the U.S. economy has become increasingly
integrated with the rest of the world, especially with
other major industrialized countries. Suppose the price
of world commodities suddenly drops while other conditions or variables do not change. Since Canada is an
exporter of raw materials and commodities, Canadian
residents’ income will decline. Falling Canadian income
means a decrease in the demand for Canadian money,
which by itself would lower the exchange value of
Canadian currency. If a falling Canadian currency has a
direct positive impact on Canadian prices in the short run
(see Dornbusch and Krugman 1976), the Bank of Canada
will try to stabilize
the exchange rate in
hopes of stabilizing
the price level. In the
process of formulating
To assess the effect of
such a monetary polimonetary policy requires
cy, the Bank of Canada
must have a fairly
understanding the interacaccurate idea of the
tion between the central
demand for Canadian
bank’s behavior and the
money. It also must
have a strong sense
private sector’s activity.
of its own behavior
(money supply) in
order to predict the
equilibrium quantity
of money and the
equilibrium rate of interest (the intersection of demand
and supply in Chart 3), as changes in both the money
stock and the interest rate will affect the exchange rate
and the price level. This example is important because
monetary policy in most countries, unlike in the United
States, resembles Canadian monetary policy in the sense
that the domestic economy is heavily influenced by foreign economies and the exchange rate plays a considerable role in policy formation.
In the case of the United States, some would say that
monetary policy is easy to formulate: it calls for simple
adherence to the federal funds rate target the Federal
Reserve itself chooses. Again, this argument is a sophism.
The federal funds rate target is not set arbitrarily; it
reflects the Federal Reserve’s concern about its own
objective of, say, price stability. When fluctuations in economic activity or the repercussions of past policy choices
threaten such an objective under the current rate of federal funds, a new target for the federal funds rate will be
chosen. Indeed, as shown in Chart 6, the federal funds
rate has changed over time, sometimes frequently. How
the target is set reflects how the Federal Reserve reacts
to the changing state of the economy, which is described
by the reaction function (5) or (4). The Federal Reserve’s

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

33

CHART 6

Time Series Pattern of Federal Funds Rate and Three-Month Treasury Bill Rate
20
Federal Funds

Percent

16

12

8

4

Three-Month
Treasury Bill

0
1960

1965

1970

1980

1975

1985

1990

1995

July 1959–March 1996

reaction function can be complicated because there is no
simple relationship between the federal funds rate and
the general price level, at least within three to ten years.
The other popular contention that questions the role
of separating money demand and money supply argues
that although a central bank’s decision is based on its
staff’s forecasts of a wide range of macroeconomic variables, such forecasts do not identify distinctive behaviors
of the central bank and the private sector. For example,
in most models that forecast real gross domestic product
(GDP), monetary aggregates, interest rates, and prices,
the central bank’s reaction function is not explicitly specified or sorted out. But it does not follow that the policymaker has no idea of money demand and money supply.
In fact, during the process of decision making, the central
bank’s behavior and the private sector’s behavior are
closely examined by looking into the past movements of
various key macroeconomic variables (such as M2 and
the general price level). When conducting monetary policy, the policymaker always wants to know how much
changes in M2 are influenced by present monetary policy
(the money supply side) and how much those are merely
caused by portfolio shifts in the private sector (the money
demand side). If, say, the money demand curve shifts to
the right from MD1 to MD2 and if the central bank desires
to have the money stock at M* (Chart 7) in order to keep
inflation in check, an economic model that explicitly separates the central bank’s behavior and the private sector’s
behavior can undoubtedly aid the policymaker in deciding how the money supply curve needs to be shifted
accordingly (from MS1 to MS2 in Chart 7). Moreover, such
a model allows one to forecast different paths of macroeconomic variables conditional on different policy
34

actions in the future. For example, the Federal Reserve
may be interested in deciding whether the federal funds
rate in the next two years should be 5 percent or 6 percent or 4 percent. If the economic model distinguishes
policy behavior and the private sector’s behavior, it can be
used to examine how policy actions in the future would
lead to different forecasts of the price level, M2, the
unemployment rate, and other variables.
The discussion in this section replies to some prevalent naive thinking about the issue of identifying monetary
policy. It reinforces the point that the actual formation of
monetary policy in any country is a complicated process.
To assess the effect of monetary policy requires understanding the interaction between the central bank’s
behavior and the private sector’s activity.

More on Demand and Supply
o far, the discussion has been concerned with why
identification of monetary policy is important in
policy analysis. It has not yet answered the question
of how we identify monetary policy in practice. How difficult is it to estimate the money demand function (3) and
the money supply function (4) or to obtain from the
observed data the values of a coefficients in both equation (3) and (4) so that the actual curves of money
demand and supply can be plotted? This section turns to
this “how” question, which is the essence of identification
integral to all empirical study in economics. It begins
with the familiar wine market example and then discusses how to estimate the demand and supply of money.
As shown in the dots in Chart 8, the data on the
quantity and price of wine are the equilibrium outcome
from movements in both demand and supply. These

S

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

C H A R T 7 Monetary Policy Reaction

R a t e

R

I n t e r e s t

MS2

E2

MS1

N o m i n a l

E1
MD2

MD1
M1(M*)

M

M o n e y

movements, caused by either the e shocks or other factors such as the government’s tax and the consumers’
preference for wine over beer, or by both, make the estimation of the demand and supply curves a challenge. A
popular approach uses the data to estimate the relationship of quantity to price with the equation
q = g1p + e.

(6)

The obvious problem with this approach is that one cannot be sure whether equation (6), after being estimated,
is a demand function, a supply function, or a combination
of the two. Suppose g1 is estimated to be –3 as indicated
by the curve Edata in Chart 8, implying that the quantity of
wine increases by 3 percent when the price falls by 1 percent. This estimated relationship between q and p does
not mean that the actual demand for wine (indicated by
the curve D in Chart 8) is as elastic as –3. The reason g1
and a1 are not equal is that g1 represents the coefficient
in the relationship (6) that is directly observed in the
data while a1 represents the coefficient in the wine
demand function (1) that is not directly observable.8
Suppose that the policy analyst mistakenly took the estimated function represented by Edata in Chart 8 as the
demand function. The analyst would anticipate that the
quantity demanded will rise substantially when the price
of wine drops. But since the actual demand curve represented by D in the chart is much steeper than the curve
Edata, the actual demand is less elastic than the one estimated and the quantity actually demanded will not rise
so substantially. Then, any conclusions based on this estimate of consumer behavior can be misleading.

The wine market example illustrates that even
when demand and supply have simple, uncomplicated
relationships, identifying (that is, estimating) each of
them is difficult. The same identification problem exists
if one tries to estimate both money demand and money
supply from the observed data because the data themselves are not sufficient to distinguish supply from
demand. Suppose one wants to identify the money supply function. The question is, How can one distinguish
the specific behavior of the central bank from the
observed data, or how can one figure out the money supply curve in Chart 3? Clearly, one cannot follow the practice of estimating equation (6) in the wine market
example and use the data on M and R to estimate such
a relationship. Some factor or factors are needed that
will shift the money demand curve but not the supply
curve. Then, if one traces out the observed data as illustrated by the dots in Chart 9, the time series pattern will
precisely reveal the money supply curve that describes
the central bank’s behavior. Thus, the basic idea of
achieving identification is to isolate factors that are in
one of the relationships, such as the money demand
function (3), but not shared with the others, such as the
money supply function (4). An assumption about which
factor influences which equation is called an identifying
assumption.
Are there such factors? As discussed above, since
the central bank is unable to observe the data on output
(y) and the general price level (P) within the present
month, the set of variables Xs in (4) does not contain y
and P. Thus, changes in current output and the current
price level serve as the shifters that move the money

8. For similar reasons, g1 and b1 are not equal.
Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

35

C H A R T 8 Equilibrium Quantity and Price of Wine

P r i c e

p

S2

1

S1

S4

S3

3
Edata

D2

D1

D4

D3

q
Q u a n t i t y

B o u g h t

demand curve but not the money supply curve (Chart 9)
and help estimate the money supply curve from the data.
The money demand function can also be estimated
(identified) in similar spirit. Recall that the information
set Xs in the money supply function (4) contains the
exchange rate or the commodity price index or both,
which are excluded in the money demand function (3).
Movements in the exchange rate or commodity price
index will shift the money supply curve but not the money
demand curve (Chart 10). The money demand curve can
then be traced out when there are enough changes in the
exchange rate or commodity price index.
Charts 9 and 10 demonstrate the basic idea of
achieving identification, but the actual estimation of both
the money demand function (3) and the money supply
function (4) is far more complicated. When both output
(shifting the money demand curve) and the exchange
rate (shifting the money supply curve) change at once,
the money demand and money supply curves will shift
simultaneously (Chart 11).9 Therefore, one cannot estimate the demand function (3) or the supply function (4)
in isolation, as Charts 9 and 10 seem to suggest. Indeed,
both functions must be estimated jointly, not in isolation.
See Box 2 (page 38) for a discussion of the technical difficulties involved. Nonetheless, the basic idea of achieving identification is clear: conceptually, to do so one
needs factors that shift the demand curve independent of
the supply curve or vice versa; technically, the demand
and supply functions must be estimated simultaneously.

Different Empirical Approaches
conomic research has historically used a variety of
empirical approaches to uncover (identify) policy
effects from the observed data. To circumvent both
the conceptual and technical difficulties in identifying

E
36

a n d

S o l d

monetary policy, many approaches invoke implausible
identifying assumptions. This section considers a few
examples or approaches taken from economics journals
and argues that assumptions that are convenient for statistical purposes but are not sensible in economics terms
are likely to generate misleading results.
Example 1. One traditional approach, which has
been exploited at least since Friedman and Schwartz
(1963), is to use a single variable (such as a monetary
aggregate, an interest rate, or an exchange rate) as an
indicator of monetary policy. For example, unpredicted
changes in a monetary aggregate—be it reserves or an M
variable—are often attributed mainly to monetary policy
shocks. The common practice is to estimate the relationship of the monetary aggregate to the other variables and
then interpret the residuals calculated from such an estimation as policy shocks (for example, Barro 1977 for the
United States and Wogin 1980 for Canada).
The assumption underlying this practice is that this
relationship represents the central bank’s behavior.
Unfortunately, this single-equation approach, analogous
to the estimation exercise along the line Edata in Chart 8,
fails to take account of the fact that the data on monetary
aggregates are what is observed at equilibrium (the point
intersected by the demand and supply curves). Indeed,
monetary aggregates such as reserves or M2 are influenced by not only the central bank’s behavior but also the
demand behavior of other sectors in the economy. Suppose
that the equilibrium outcome is such that the monetary
aggregate and the interest rate are positively correlated as
indicated by the single estimation curve Edata in Chart 11.
Apparently, it is a mistake to interpret the curve Edata as
the actual money supply curve MS.
The point that the policy behavior inferred through
the time series patterns of a single variable can be mis-

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

R

MS

N o m i n a l

I n t e r e s t

R a t e

C H A R T 9 Tracing Out the Money Supply Function

MD3

MD1

MD2

MD4

M
M o n e y

R
MS4

MS1

MS2

MS3

N o m i n a l

I n t e r e s t

R a t e

C H A R T 1 0 Tracing Out the Money Demand Function

MD

M
M o n e y

leading was made long ago by Tobin (1970). Tobin presented a dynamic general equilibrium model to show that
the same time series evidence used by Milton Friedman
and other monetarists could lead to a completely different
interpretation of monetary policy effects. Despite Tobin’s
warning, however, researchers oftentimes continue to use
the single-equation approach to modeling monetary policy, at least in part because identifying monetary policy is
conceptually difficult and mathematical tools are only
now being developed to address the identification issue
seriously.10

Example 2. In recognition of both the inadequacy of
single-equation approaches and the nature of a central
bank’s reaction to the state of the economy, recent
research on identification of monetary policy has developed ways of handling the complex relationships of multiple economic variables (see Box 2). One approach is to
include both policy instruments (such as an interest rate)
and other macroeconomic variables (such as the general
price level) in the same framework (as in Sims 1992, Grilli
and Roubini 1995, Eichenbaum and Evans 1995, and
Dungey and Pagan 1997). This approach is certainly a

9. The situation is analogous to that depicted in Chart 3.
10. Romer and Romer (1989, 1990), following the spirit of Friedman and Schwartz, invent a single dummy variable indicating
the changes in U.S. monetary policy. But as Leeper (1997) forcibly argues, the Romers’ dummy variable does not identify the
Federal Reserve’s behavior.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

37

B O X

2

Empirical Methods
his box, focusing on the case of a small open economy, is
largely drawn from Zha (1996). Assume the structural
model is of a linear, dynamic form called vector autoregression (VAR):

T

A(L)y(t) = e(t),

(B1)

where A(L) is an m 3 m matrix polynomial in lag operator
L, y(t) is an m 3 1 vector of observations of m variables, and
e(t) is an m 3 1 vector of i.i.d. structural shocks so that
Ee(t) = 0, Ee(t)e(t)´ = I.

(B2)

The reduced form of (B1) can be obtained by multiplying A–1
0
through (B1).
A natural way of estimating the model is to explore the
shape of a likelihood function (which describes how likely
the model parameters are to lie within a certain range of values) and to obtain the values of parameters that are most
likely to occur (the values so obtained are called maximum
likelihood [ML] estimates). If the likelihood function is
complicated, finding ML estimates may become problematic. For the reduced form of (B1), however, the ML estimates
turn out to be simply the ordinary least squares (OLS) estimates in each equation. The OLS estimation is straightforward and can be easily computed using any statistical
software package.
To see how the system (B1) can be used to model a
small open economy, break (B1) into two blocks—the first
block concerns the home (small) economy, and the second
block concerns the foreign (the rest of the world) economy.
To be specific, let

 A11 ( L) A12 ( L)
A( L) = 
,
 A21 ( L) A22 ( L)


 y (t)
y(t) =  1  ,
 y2 (t)
 e (t)
e(t) =  1  .
 e 2 (t)

The matrix As is the coefficient matrix of Ls in A(L), where
Ls is the lag operator L raised to s power. In most works of the
identified VAR literature, the restrictions are imposed only
on A0—the contemporaneous coefficient matrix. The ML
estimates of A0 depend only on the estimated covariance
∑) of reduced-form residuals; this can be easily seen
matrix (^
by writing out the concentrated likelihood function of A0
(see Sims and Zha 1995 for details):

(

T
 T
A0 exp− trace Sˆ A0′ A0
 2

).

(B3)

Because of (B3), obtaining the estimates of and inference
about model parameters is in principle straightforward
(Sims and Zha 1995), and when A0 follows a successive representation, the estimation is straightforward even in practice (Sims 1980).
If the small open economy framework is taken seriously, one will impose the restriction that A21(L) = 0, meaning
that the small country takes changes in foreign economic
conditions as given or exogenous. This small-economy
restriction makes the easily implemented procedure developed by Sims and Zha (1995) invalid, mainly because the
concentrated likelihood (B3) no longer holds. In principle,
various iterative procedures can be used. For example, one
∑ to solve the ML estimate of
begins with the unrestricted ^
A0 with the restriction A0 21 = 0 imposed. The estimates of
other structural parameters (As, s ≥ 1) can then be recovered, and a new reduced form covariance matrix is accord∑
ingly formed.1 Use this new matrix to replace the previous ^
and repeat the procedure until ^
∑ converges.
Since the size of a small open economy model is typically large relative to closed economy models, this iterative
procedure not only is cumbersome but can be computationally prohibitive as well, especially when one computes the
inference of the ML estimates. Consequently, previous
researchers have not accounted for the small open economy
features in their models. The method developed in Zha
(1996), which allows for more general cases than the small
open economy example here, provides a practically feasible
way of obtaining the ML estimates as well as their inference.

1. The details of how the estimates are recovered are discussed in Zha (1996). The idea of this iterative procedure is also mentioned in Dias,
Machado, and Pinheiro (1996).
38

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

R

MS2

MS1

MS3

MS4

I n t e r e s t

R a t e

C H A R T 1 1 Simultaneous Changes in Money Demand and Money Supply

N o m i n a l

Edata

MD1

MD2

MD3

MD4
M

M o n e y

C H A R T 1 2 Effect of Contractionary Shock (EMS)
R

MS1
R2

E2

E1

R1

N o m i n a l

I n t e r e s t

R a t e

MS2

MD

M2

M1

M

M o n e y

considerable improvement over that outlined in Example
1. Nonetheless, all these works suffer from a common problem: they make identifying assumptions that seem implausible in the description of a central bank’s behavior.
As explained above, identifying assumptions are
those that help distinguish different behaviors (for example, demand and supply) in the actual economy. They are
necessary because, analogous to the example of demand
and supply, one needs some factors that shift only the
supply curve in order to identify the demand function
(Chart 10) and other factors that shift only the demand
curve in order to identify the supply function (Chart 9).
While the aforementioned works use assumptions that are
convenient for statistical computations, the important dif-

ference in the approach called for in the previous sections
is that it argues for economically sensible assumptions.
Specifically, the common assumption used in the
works cited is that different behaviors follow successive
relationships. Although the successive assumption makes
it convenient to estimate the model (see Box 2), it seldom
represents the structure of the actual economy. For example, in Eichenbaum and Evans (1995) the money stock M
influences the interest rate R contemporaneously but not
vice versa, an assumption that essentially takes a3 in the
money supply equation (4) to be zero, implying that the
money supply is perfectly inelastic (Chart 5).11 In their
study of the U.S. economy, for example, Eichenbaum and
Evans use nonborrowed reserves as M and the federal

11. The money demand function is not explicitly specified in their papers.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

39

CHART 13

Perverse Dynamic Responses of Price after
Monetary Contraction: The Price Puzzle

Money

M

0

1

2
Ye a r

General Price Level

P

1

Ye a r

2

funds rate as R. Thus, an inelastic money supply (a3 = 0)
means that the Federal Reserve does not respond, within
the month, to fluctuations in the federal funds rate. In
fact, the Federal Reserve frequently influences the federal funds rate in pursuit of its objective, so the assumption
that money supply is perfectly inelastic seems at odds
with the Federal Reserve’s targeting of the federal funds
rate. It is therefore not surprising that this extreme
assumption would lead to results that are inconsistent
with views widely held by both policymakers and economists (Gordon and Leeper 1994 and Leeper 1995).
Before reviewing an example of these inconsistent
results, it is important to explain the concept of contractionary monetary policy shock that is often used in economics journals. Recall that this article uses the phrase
policy shocks—eMS in the money supply equation (4)—to
describe unpredicted shifts in monetary policy. Thus, the
shock eMS in equation (4) is said to be contractionary if it
shifts the money supply curve to the left (from MS1 to MS2
in Chart 12), moving the equilibrium outcome from point
E 1 to point E 2. The word contractionary is adopted
because, subsequent to this shock, the money stock M
40

contracts from M1 to M2 while the interest rate R rises
from R1 to R2.
Now, to present an example, consider one of the
firmly established views in policy analysis: the price level
falls after an unpredicted contraction in monetary policy.
When one uses Eichenbaum and Evans’s successive
assumption to model several industrial countries such as
the United States, Japan, and Germany, the model generates the inconsistent result (often termed the price puzzle) that the price level would rise, not fall, in response
to a contractionary monetary policy shock (Chart 13).12 If
the model is intended to be useful for policy decisions,
such a puzzle is indeed troublesome because it implies
that monetary policy must expand the money stock (or
lower the federal funds rate) in order to lower inflation.
Would one recommend such a policy? Does anyone really
believe that inflation will fall if the central bank increases the supply of money (or lowers interest rates)?
When a model produces inconsistent results such as
the price puzzle, one needs to examine carefully the
underlying (identifying) assumptions to see if they make
good economic sense. If a central bank reacts quickly to
changes in the interest rate, it makes no economic sense
to assume that the interest elasticity a3 in the money supply equation (4) is zero. If one insists on a successive representation by letting a3 be zero, equation (4) is then no
longer the policy reaction function (or the money supply
function).
Example 3. The above example suggests that a reasonable identification of the central bank’s behavior
inevitably leads to a breakdown of the successive representation commonly used in economics journals. A recent
work of Cushman and Zha (1997) argues for a better representation of policy’s systematic behavior and makes
progress in the specification and estimation of behavioral
relationships. In that study, both the money demand
equation and the money supply equation are in the same
form as equations (3) and (4) in this article. Using
Canada as a study case, the paper devotes special attention to Canada’s relationship with the U.S. economy and
the systematic component (a3R + a4 Xs) in the policy
reaction function (4). In particular, the set Xs contains a
wide variety of macroeconomic variables to which the
Bank of Canada would react. Some information, such as
output and the general price level in both Canada and the
United States, is not readily available to the Bank of
Canada in a timely fashion (because data such as industrial output and the consumer price index for a given
month are not released until after the end of the month).
These pieces of information are therefore excluded from
Xs in the money supply function (4). The Bank of Canada,
however, can react quickly to changes in other key macroeconomic conditions—the exchange rate, the U.S. interest rate, and commodity prices—for which data are
available daily. These economic conditions convey infor-

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

Conclusion
he monetary policy reaction in any actual economy
is complicated, and “the policy framework is a
pragmatic one. There are no simple rules” (Duguay

T

CHART 14

R

MS

1
0.13

N o m i n a l

I n t e r e s t

R a t e

Estimated Money Demand and
Money Supply Functions

MD
M
M o n e y

Source: Adapted from Cushman and Zha (1997).

CHART 15

Relationship between Money Supply and
Exchange Rate

E x c h a n g e R a t e
U S $ / C $

mation about the current state of both the Canadian
economy and the U.S. economy, about possible actions in
U.S. monetary policy, and about future inflation.
The estimated money demand and money supply curves
by Cushman and Zha (1997) are depicted in Chart 14.
The money supply is almost inelastic to the domestic
interest rate but, as shown in Chart 15, very elastic to the
exchange rate. This condition implies that the Bank of
Canada, unlike the Federal Reserve, responds mainly to
changes in the exchange rate rather than to the domestic
interest rate. Evidently, systematic behavior of central
banks may differ across countries (such as the United
States and Canada). For the U.S. economy, it is a mistake
to assume the interest elasticity a3 in the money supply
function (4) to be zero because the Federal Reserve targets
the federal funds rate. In other words, assuming a3 = 0
may be expedient statistically, but it yields results that are
not sensible. For the Canadian economy, it is a mistake to
assume the exchange elasticity in the money supply function to be zero because, as shown in Chart 15, the Bank of
Canada responds to changes in the exchange rate. Indeed,
if the exchange rate elasticity were assumed to be zero,
the price puzzle, which does not exist in Cushman and
Zha’s original model, would be present.
The above example of identifying Canadian monetary
policy shows the importance of using sensible identifying
assumptions even though such assumptions may raise
empirical difficulties. Identifying the U.S. monetary policy
involves a similar task of separating the Federal Reserve’s
behavior from the private sector’s behavior. Leeper, Sims,
and Zha (1996) discuss how difficult it is to achieve
reasonable identification of U.S. monetary policy.
Understanding each country’s relationship with the rest of
the world and each central bank’s systematic behavior is a
necessary step when one makes identifying assumptions.
This section reviews several identification approaches
used in policy analysis. Some, such as the single-equation
approach, fail to separate policy’s systematic response to
the state of the economy from the response of the economy
to policy (supply from demand). Those that attempt such
separation have often imposed extreme assumptions that
would lead to inconsistent or puzzling results. All the
examples discussed echo the same message: avoiding
extreme or unreasonable identifying assumptions when
modeling monetary policy in a given country is crucial for
eliminating puzzling results, achieving correct identification, and producing sensible monetary policy analysis.

Policy Reaction

1
1.45

Q u a n t i t y

o f

M o n e y

and Poloz 1994, 197). The discussion here establishes the
importance of identifying monetary policy, explains the
difficulties involved in identification, cautions about the
potential danger of making extreme assumptions about
the behavioral relationships, and sheds some light on the
progress there has been toward adequately identifying
monetary policy (as in Cushman and Zha 1997). In particular, the article uses the simple examples of demand
and supply to illustrate how quickly the difficulty in identification of monetary policy can become overwhelming if
one wishes to separate the central bank’s behavior from

12. The author thanks Roberto Chang for suggesting such an exercise.
Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

41

others’ behavior in the economy. The difficulty contributes to the disparity and uncertainty in economists’
views on the effects of monetary policy.
The essential point is that because of the complexity inherent in monetary policy reaction unique to different countries, an economic model usable for policy
analysis in a given country requires both cogitable reasoning conceptually and serious effort empirically.
Notwithstanding significant progress in both theory and

econometrics, the gap between economic theory and
empirical observations is still wide because theoretical
models have not yet produced the same time series pattern of macroeconomic variables as those that characterize the actual economy. The challenge in future research
is to narrow the gap and move toward a good economic
model usable for a discussion of actual policy effects in
different countries.

R E F E R E N C E S
BARRO, ROBERT J. 1977. “Unanticipated Money Growth and
Unemployment in the United States.” American Economic
Review 67 (March): 101-15.

FRIEDMAN, MILTON, AND ANNE J. SCHWARTZ. 1963. A Monetary
History of the United States, 1867-1960. Princeton, N.J.:
Princeton University Press.

BARRO, ROBERT J., AND ROBERT F. LUCAS. 1994. Macroeconomics.
Burr Ridge, Ill.: Richard D. Irwin, Inc.

GORDON, DAVID B., AND ERIC M. LEEPER. 1994. “The Dynamic
Impacts of Monetary Policy: An Exercise in Tentative
Identification.” Journal of Political Economy 102
(December): 1228-47.

BATTEN, DALLAS S., MICHAEL P. BLACKWELL, IN-SU KIM, SIMON E.
NOCERA, AND YUZUVU OZEKI. 1990. “The Conduct of Monetary
Policy in the Major Industrial Countries: Instruments and
Operating Procedures.” International Monetary Fund
Occasional Paper 70, July.
COOLEY, THOMAS F., AND GARY D. HANSEN. 1995. “Money and the
Business Cycle.” In Frontiers of Business Cycle Research, edited by Thomas F. Cooley, 175-216. Princeton, N.J.: Princeton
University Press.
CUSHMAN, DAVID O., AND TAO ZHA. 1997. “Identifying Monetary
Policy in a Small Open Economy under Flexible Exchange
Rates.” Journal of Monetary Economics 39 (June, forthcoming).
DIAS, FRANCISCO C., JOSE A.F. MACHADO, AND MAXIMIANO R.
PINHEIRO. 1996. “Structural VAR Estimation with Exogeneity
Restrictions.” Oxford Bulletin of Economics and Statistics 58
(May): 417-22.
DORNBUSCH, RUDIGER, AND PAUL KRUGMAN. 1976. “Flexible
Exchange Rates in the Short Run.” Brookings Papers on
Economic Activity, no. 3:537-75.
DUGUAY, PIERRE, AND STEPHEN POLOZ. 1994. “The Role of
Economic Projections in Canadian Monetary Policy
Formulation.” Canadian Public Policy 20, no. 2:189-99.
DUNGEY, MARDI, AND ADRIAN PAGAN. 1997. “Towards a Structural
VAR Model of the Australian Economy.” Australian National
University, unpublished manuscript.
EICHENBAUM, MARTIN, AND CHARLES EVANS. 1995. “Some Empirical
Evidence on the Effects of Shocks to Monetary Policy on
Exchange Rates.” Quarterly Journal of Economics 110
(November): 975-1009.
FRIEDMAN, MILTON. 1992. Money Mischief: Episodes in Monetary
History. New York: Harcourt Brace Jovanovich.

42

GRILLI, VITTORIO, AND NOUVIEL ROUBINI. 1995. “Liquidity and
Exchange Rates: Puzzling Evidence from the G-7 countries.”
Yale University, unpublished manuscript.
LEEPER, ERIC M. 1992. “Facing Up to Our Ignorance about
Measuring Monetary Policy Effects.” Federal Reserve Bank of
Atlanta Economic Review 77 (May/June): 1-16.
———. 1995. “Reducing Our Ignorance about Monetary Policy
Effects.” Federal Reserve Bank of Atlanta Economic Review 88
(July/August): 1-38.
———. 1997. “Narrative and VAR Approaches to Monetary
Policy: Common Identification Problems.” Journal of
Monetary Economics 40 (forthcoming).
LEEPER, ERIC M., CHRISTOPHER A. SIMS, AND TAO ZHA. 1996. “What
Does Monetary Policy Do?” Brookings Papers on Economic
Activity, no 2:1-78.
MCCALLUM, BENNETT T. 1989. Monetary Economics: Theory and
Policy. New York: Macmillan.
MISHKIN, FREDERIC S. 1992. The Economics of Money, Banking,
and Financial Markets. 3d ed. New York: Harper Collins.
ROMER, CHRISTINA D., AND DAVID H. ROMER. 1989. “Does Monetary
Policy Matter? A New Test in the Spirit of Friedman and
Schwartz.” In NBER Macroeconomics Annual 1989, edited by
Olivier J. Blanchard and Stanley Fischer, 121-70. Cambridge,
Mass.: MIT Press.
———. 1990. “New Evidence on the Monetary Transmission
Mechanism.” Brookings Papers on Economic Activity, no.
1:149-98.
SIMS, CHRISTOPHER A. 1980. “Macroeconomics and Reality.”
Econometrica 48 (January): 1-48.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

———. 1992. “Interpreting the Macroeconomic Time Series
Facts: The Effects of Monetary Policy.” European Economic
Review 36 (June): 975-1000.

WOGIN, GILLIAN. 1980. “Unemployment and Monetary Policy
under Rational Expectations: Some Canadian Evidence.”
Journal of Monetary Economics 6 (January): 59-68.

SIMS, CHRISTOPHER A., AND TAO ZHA. 1995. “Error Bands for
Impulse Responses.” Federal Reserve Bank of Atlanta Working
Paper 95-6, September.

ZHA, TAO. 1996. “Identification, Vector Autoregression, and
Block Recursion.” Federal Reserve Bank of Atlanta Working
Paper 96-8, August.

TOBIN, JAMES. 1970. “Money and Income: Post Hoc Ergo Propter
Hoc?” Quarterly Journal of Economics 84 (May): 301-17.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

43

International
Settlements:
A New Source of
Systemic Risk?
R O B E R T A . E I S E N B E I S
The author is senior vice president and director of
research at the Atlanta Fed. He thanks Clyde
Farnsworth, Craig Furfine, Diana Hancock, Pat
Parkinson, and Alice P. White for helpful comments.

T

HE VERY REAL SIGNIFICANT SOCIAL COSTS OF SYSTEMIC RISK HAVE LONG SERVED AS AN IMPORTANT RATIONALE FOR A FEDERAL PRESENCE IN THE DOMESTIC PAYMENTS SYSTEM.1

RECENT

MARKET DEVELOPMENTS HAVE HEIGHTENED CONCERNS ABOUT THE POTENTIAL FOR SYSTEMIC
RISK IN THE PAYMENTS SYSTEM.

FIRST, THE SHEER GROWTH IN LARGE-VOLUME PAYMENTS HAS

RAISED THE POTENTIAL COSTS SHOULD A NUMBER OF INSTITUTIONS FAIL.

SECOND, TECHNOLOGY AND TECH-

NOLOGICAL CHANGE SEEM TO BE REDEFINING THE KINDS OF TRANSACTIONS TAKING PLACE AS WELL AS
INCREASING THE SPEED WITH WHICH THESE TRANSACTIONS CAN BE COMPLETED AND FUNDS TRANSFERRED.

FOR

EXAMPLE, BOTH COMPUTER AND OPTIONS PRICING TECHNOLOGIES NOW PERMIT THE UNBUNDLING,

RESTRUCTURING, AND CREATION OF TRANSACTIONS (SUCH AS SWAPS AND DERIVATIVES) WHOSE RISKS, LEGAL
STATUS, AND RELATED CHARACTERISTICS ARE JUST NOW BEGINNING TO BE UNDERSTOOD.

A third dynamic behind the increased concern about
systemic risk in the payments system is the globalization
of financial markets, which is tying economies and markets together in ways that introduce additional issues
about the mechanisms by which traditional clearing (the
notification and transfer of documents and orders to purchase and sell assets) and settlement (the transfer of
final payment) take place. Finally, the fears about the
supposed potential for systemic risk associated with

44

clearing and settlement loss have been given greater credence by the lack of internal controls within major institutions, which have been exposed by the actions of rogue
traders in Kidder, Bankers Trust, and Barings (see
Edwards 1996). As the problems in these institutions
have been unwound, greater appreciation has emerged of
just how complex and segmented the institutional
arrangements for clearing and settling transactions have
become.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

Despite the fact that securities, futures and options,
and derivatives are increasingly cleared under a variety of
institutional arrangements, final settlement usually
takes place in the interbank market. In general, clearing
of transactions—be they securities, derivatives, or other
assets—is almost exclusively done by the private sector
while settlement can take place in the wholesale banking sector or through central banks (see BIS 1997b).2
Markets have become tiered as more and more transactions are cleared through several layers of institutions
before they are ultimately settled (see Corrigan 1990).
Equally important, the introduction of new instruments,
such as swaps, collateralized mortgage obligations, and
off-exchange derivatives, and their associated methods
for transferring cash flows and settlement relationships
have resulted in seemingly unrelated markets and institutions being linked together in ways that both create
and may de facto transfer risks from one market to another. Increasingly, these transactions and markets are
assuming an international dimension that can also have
significant domestic market implications (see BIS
1997b).
This article examines whether internationalization
has changed the nature of and potential vulnerability of
the financial system to systemic risks and looks at a
method to mitigate them. The Lamfalussy Report, which
examined and proposed standards for payments systems
settlement and risk control features, indicates that system vulnerability is critically linked to the length of time
that participants are exposed to credit and liquidity risks
(see BIS 1993b). The analysis presented here suggests
that regulatory and legal structures can also have systemic risk dimensions. As to the fundamental question of
whether new risks are being introduced, the answer
seems to be no. Moreover, recent institutional and regulatory developments may act to reduce the potential
scope and the size of these risks and limit implicit taxpayer liabilities should these risks be realized.

Risks in Payments
egardless of the institutional arrangements, there
are four generally accepted generic types of payments system risks that have been identified and
have been the focus of much attention. These include
operational, legal, credit, and liquidity risks (see
Eisenbeis 1995 or BIS 1997b). While it is easy to differentiate these risks conceptually, in reality they tend to be
interrelated. The realization of one can lead to occurrences of the others, and this dynamic has not changed
with the evolution of the new instruments and markets
just described. These interrelationships among risks can

R

be illustrated by considering credit risk, which arises
when the purchaser of an asset defaults by failing to settle any or all of its obligations. Credit risk arises as a logical by-product of separating the clearing and settlement
functions, which under current institutional arrangements nearly always involves an extension of temporary
credit.
Credit risk is a function of the potential loss exposure when a buyer initiates a transaction ordering its
bank to transfer funds but then cannot make payment
without going into an overdraft situation. The buyer’s
bank, which is attempting to settle on behalf of the buyer,
is faced with essentially three alternatives. First, it can
provide credit to the customer until funds are received.
Second, the transaction can be canceled,
or the bank can comRecent market developplete the transaction
itself. If the buyer’s
ments have heightened
bank takes the place of
concerns about the potenthe customer and comtial for systemic risk in the
pletes the transaction,
it may then take pospayments system. . . . As
session of the goods or
to the question of whether
asset (or any other of
new risks are being introthe customer’s available collateral) and
duced, the answer seems
proceed to unwind the
to be no.
transaction. Finally,
in the extreme, the
buyer’s bank can default on its own obligation to settle if the time for settlement has not yet occurred.
If the buyer has good collateral and a sound credit
rating, then extension of credit may be the best alternative. Canceling the transaction may not be an option,
especially when delivery of the good or service has
already taken place and there is no available collateral.
Settlement failure in this example could be controlled if the buyer’s bank were to put a hold on the
buyer’s funds at the time payment is initiated, collateralizing the transaction. Organized futures markets effectively accomplish this control through the use of margins,
mark-to-market accounting, and settlement requirements. For good customers, however, collateralization
may not be necessary, practical, or efficient, especially if
both the probability of default and the expected loss are
small relative to the bank’s resources. The lack of a hold
or similar type of collateral policy illustrates that an institution’s vulnerability and exposure to credit risk often
results from the underlying conventions, practices, and

1. See Benston and Kaufman (1995) for a review of the evidence on fragility and systemic risk.
2. For a discussion of the risks and recent developments in exchange-traded derivatives markets, see BIS (1997a).

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

45

structures of the markets involved rather than from the
realization of performance risks associated with the
underlying projects and investments.
As markets have become increasingly global, differences in timing and clearing and settlement conventions
and differences in bankruptcy laws can add important
temporal and other dimensions to credit risks not always
found in domestic markets. This consideration was clearly demonstrated in 1974 when Herstatt Bank failed and
was closed by German authorities. Herstatt had entered
into agreements to exchange deutsche marks for dollars. The mark leg of
the transaction was
settled, but the dollar
As markets have become
portion was not setincreasingly global, differtled in New York at
t h e time Herstatt
ences in timing and
was closed since the
clearing and settlement
deadline on CHIPS
conventions and differences
(Clearinghouse I n terbank Payments
in bankruptcy laws can
System) for final setadd important temporal
tlement was approxand other dimensions to
i m a t e l y 4 : 3 0 P. M .
eastern
standard time.
credit risks.
This difference in settlement times for the
two sides of the transaction left the counterparties to the foreign exchange
transaction thinking that they had more funds than they
did. When the dollar transactions failed to settle, the
result was large losses to the U.S. counterparties. This
temporal dimension to credit/systemic risk has come to
be known as Herstatt risk and can be very large.3
A more recent example of this type of event is the
closing of the Bank of Credit and Commerce International (BCCI) in 1991. The Industrial Bank of Japan
had paid 44 billion yen into BCCI’s branch in Tokyo, for
which payment was to be received in New York from
BCCI’s New York branch. When BCCI was closed, the dollar portion of the transaction was never completed,
and Industrial Bank of Japan became a creditor for
$30 million.
These examples may at first look like ordinary credit risk in that loss exposure resulted from the inability of
Herstatt and BCCI to pay. But the incidence of the losses
and ultimate position of the banks’ creditors was determined by both home country laws and the intervention
policies of their regulatory authorities, whose actions
usually cannot be easily predicted or priced.4 The losses
to dollar counterparties in the Herstatt case were the
consequence of the timing of the closure of the institution rather than the realization of estimable default risk.
Had the German authorities waited until the U.S. dollar
markets had settled, then the losses to those expecting
46

dollar transfers would not have occurred and the risks
would not have been realized. Such exposure is better
characterized as settlement uncertainty rather than settlement or credit risk since it is not possible to estimate
reliably and cost out the implications associated with the
vagaries of sometimes untested statutes governing transactions and of regulatory actions and policies. Note, too,
that although the size of the losses may not have been
affected by the closure timing, the distribution of the
losses was significantly affected by legal structures and
governmental action. At the same time, numerous initiatives by governmental bodies such as the Federal Reserve
and the Bank for International Settlements (BIS) are
continually seeking to identify and institute policies to
limit these problems (see Bank of England 1994 and BIS
1989, 1990, 1993a-c, 1997a, b).
Herstatt-type risk can also be involved solely in dollar clearing systems. In Asia the Chase Manhattan Bank
operates a dollar clearing and settlement service through
its Tokyo branch. The system provides a limited overdraft
facility and promises finality of settlement guaranteed by
Chase Manhattan. Participants are permitted to settle
overdrafts in New York across the Tokyo/New York business day. Furthermore, Tokyo balances at the end of the
day may be transferred to New York through the New York
offices of Chase or Tokyo banks or through CHIPS. In this
system any problems that may arise in this satellite settlement and clearing system quickly have the potential to
transmit liquidity and credit risk from Asia to New York,
and ultimately to the Federal Reserve, if it affects CHIPS,
Chase, or significant New York correspondents. A failure
to settle in New York on payments guaranteed in Japan by
Chase creates a form of Herstatt risk that would end up
having to be resolved in New York. At present, concern
about such clearing and settlement systems stems from
the sheer size of the potential losses rather than from a
true understanding of well-articulated scenarios on how
the risks would be played out.5

Sources of Payments Uncertainty
henever clearing and settlement of financial
assets are separated in the international arena,
a given country’s rules usually establish the
exact point in time that a transaction has been completed and the obligation satisfied. The issue centers on
transaction finality and the legal criteria for when debts
are discharged and who bears the losses in the event of
default. Finality usually occurs when the party selling the
asset actually has “good funds” and the transaction is
both irrevocable and unconditional. Importantly, since
many central bank settlement systems can involve the
extension of intraday credit, finality may or may not correspond to the time that the buyer actually settled. For
example, because Fedwire provides finality as a matter of
Federal Reserve policy, acceptance of a payment order

W

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

carries with it the “guarantee” of good funds to the
receiver and also discharges the debt, since the sender’s
reserve account is debited and the receiver’s bank
account is credited, even though the sender’s bank may
default on the settlement of its reserve account with the
Fed at the end of the day. When the settling institutions
are located in two separate countries, the specifics of the
transactions in terms of settlement, discharge of debt,
and so forth may sometimes be governed by the laws of
two separate countries and, if transactions involve clearinghouses, the laws where they are located as well.
The legal status of claims can quickly become very
murky when the problems involved in settlement failures
in cross-border bilateral and multilateral netting
arrangements are examined, especially those transactions involving forward-dated contracts in foreign
exchange, derivatives, and other cross-border markets
(see BIS 1997b). Under netting systems, debt and credit
orders are cumulated, and only the net difference is
transferred at an agreed-upon time. This procedure contrasts with real-time gross settlement systems (RTGS),
which continually process and settle transactions as the
orders are received. Final disposition of the liability
under netting systems depends critically on the legal
rules governing the disposition of debts and transactions
in the event of a default or bankruptcy.
As an example, if two institutions have entered into
a bilateral netting arrangement, then completion of all
the transactions subject to the arrangement is contingent on settlement of the net position. Should one of the
parties fail to settle because of a bankruptcy, all the
gross transactions subject to netting may have to be
undone. The determining factor here depends upon the
legal rules affecting the markets in which the transaction was settled. Since the legal rules may differ according to where settlement takes place, and this location
may be beyond the receiver’s control, settlement uncertainty may exist.
The exact status of cross-border transactions, therefore, is determined by several sets of laws. These include

the laws governing bilateral netting arrangements and
those governing the particular settlement market
involved as well as the bankruptcy provisions and other
related laws of the country of the failed institution (or
the laws of the resident country if the transaction is
recorded on the books of a branch of the failed bank). For
example, netted transactions may or may not be regarded
as discharged. The bankruptcy court with jurisdiction
over the transactions may decide to unbundle netted
transactions, demanding payment for debts owed and disavowing liabilities to creditors. In addition, country bankruptcy law may give
creditors the right
to offset their liabilities to a failed entity
Final disposition of the
against their claims
liability under netting
on that entity. Thus,
systems depends critically
debts owed on foreign
exchange may be dison the legal rules governcharged with debts on
ing the disposition of
securities, loans, or any
debts and transactions
other assets. Not only
do the bankruptcy laws
in the event of a default
affect the size of the
or bankruptcy.
losses but also the way
in which the losses may
be apportioned across
various creditors.
The legal situation in multilateral netting arrangements introduces complexities several orders of magnitude greater than those affecting bilateral arrangements.
There is considerable variation across countries in treatment of transactions, and thus uncertainty exists about
how particular bankruptcies will be treated. The key
point is that this legal uncertainty often can undermine
the efficiency of bilateral and multilateral netting
arrangements and creates the very real possibility that
systemic risks could be heightened rather than reduced
when the laws governing netting are not uniform across
countries. Because these legal uncertainties complicate

3. Notice, however, that it may be a misnomer to call this type of event risk, at least in the Frank Knight ([1921] 1971) tradition;
see also Hu (1994). The incidence of loss resulted from the German governmental action, which seems almost impossible to
assign a probability to, and hence may be better characterized as regulatory uncertainty. See BIS (1996) for a comprehensive
discussion of risks in foreign exchange markets and efforts that both private- and public-sector entities have made to identify,
monitor, and control these risks.
4. Bankruptcy statutes can clearly affect the distribution of claims as well. For example, some countries have what is known as a
zero-hour rule, which means that transactions taking place after the time the institution is legally closed are regarded as
invalid and will be unwound.
5. See, for example, General Accounting Office (1994). An exception is Edwards (1996), who describes the possible paths of a breakdown in derivatives markets. He describes a scenario in which an end-user fails to meet its obligations as a counterparty. This
failure in turn brings down a major dealer, thereby spilling over to both other counterparties and dealers. These disruptions
are then transmitted to other markets as uncertainty both raises contract prices and leads to reluctance to enter into contracts.
There are then price breaks, credit disruptions, falling asset prices, and, ultimately, real effects. Edwards analyzes the likelihood that such a scenario would be realized and concludes that true dealer credit exposures are small and substantially smaller than their exposure on loans and other assets.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

47

assessment of the likely outcome of a default scenario for
many transactions, authorities have paid great attention
to putting transactions on a common legal basis and, as
discussed in the next section, some nations have moved
to establish real-time gross settlement as the basis for
clearing and settlement.

Responses to Uncertainty
oth private- and public-sector entities have
responded to the increased uncertainties, market
risks, and evolving market technologies in many
interesting ways. The responses affect contract design
and the micromarket structure of exchanges and their
rules governing transactions. They have given rise to proposals to change laws governing transactions and suggestions to increase
governmental crossborder cooperation in
financial rules, regulaSystems are evolving
tion, and supervision
toward real-time gross
as well as changes in
the structural design
settlement, which contains
of transfer systems.
inherent incentives for
Given the cominstitutions engaged in
plexity of financial
transactions and their
offering payments services
interrelationships,
to price and monitor their
measuring, monitorexposures.
ing, and pricing what
institutions’ true risk
exposures to each
other are and how
these risks flow directly and indirectly through relationships with related customer groups is difficult. For example, Customer X may have several relationships with its
primary bank (Bank A). These might include a loan, a
swap, a deposit account, and several foreign exchange
transactions. Customer X may also have similar relationships and transactions outstanding with Bank B. In addition, Bank A may also have made loans in the form of
advancing federal funds to Bank B. If Customer X fails,
the entirety of its net position with Bank A across all the
relationships and transactions represents its net direct
risk exposure. Bank A may also be indirectly exposed
through Bank B if the customer’s default causes Bank B
to default on its federal funds obligations to A’s primary
bank.
Measuring and monitoring these interrelated exposures across the world, across different markets and time
zones, is a truly daunting modeling and monitoring problem. It is made even more so by the dynamic and continual evolution of new instruments and markets.
Central bank and market responses to these challenges have been to substitute rules and other mechanisms to control customer risk-taking incentives. A

B

48

number of control mechanisms have been designed to
limit uncertainty and to provide incentives for member
institutions to control their own risk exposures. These
include maintenance of adequate capitalization, reliance
upon contract design to allocate risk and losses, collateralization of transactions, use of outside guarantees and
bonding, pricing, imposition of system membership
requirements, and self-imposed (and system-mandated)
caps and other limits on risk exposure to individual and
related parties. For example, in the United States, the
Federal Reserve imposed limits in 1986 on participating
banks’ net exposures across Fedwire and CHIPS as well
as bilateral limits on exposures to individual participants.
Collateralization of certain positions is also required, and
the system charges for intraday credit that is extended.
Contracting activities also have focused on apportioning risks, defining performance, and allocating losses
among participants in a payments system or exchange in
the event that a default occurs. Because of the difficulties in continuously measuring and monitoring total risk
exposure to individual system members, caps on the
amount of exposure with any member have been imposed,
and the system imposes a similar total cap across all system members. In the case of the U.S. CHIPS system
(which is not a real-time gross settlement system), participants require same-day settlement, engage in realtime monitoring, have established limits on exposures,
have required collateral to cover the largest two exposures, and have instituted a loss-sharing arrangement.6
System members also impose various types of membership and participation requirements, such as the maintenance of minimum capital requirements.
It has also been recognized that accounting rules—
such as mark-to-market requirements—can affect the
ease of information transfer and reduce monitoring costs.
Such rules have been especially widely used in the case
of futures, options, and commodities exchanges.
Finally, systems are evolving toward real-time gross
settlement despite the supposed efficiency advantages of
netting arrangements. Real-time gross settlement systems require those engaging in payments activities to collateralize payments fully as they are initiated. The
benefits of doing so are weighed against the costs of
uncertainty and credit risks. Such systems contain inherent incentives for institutions engaged in offering payments services to price and monitor their exposures.
Furthermore, real-time gross settlement reduces risk
exposure by limiting the duration of both credit and liquidity risk.
The first real-time gross settlement system was the
Federal Reserve’s Fedwire (see BIS 1997b). By the end
of the 1980s, six of the Group of 10 countries had instituted RTGS systems. As the European Union proceeds,
Lamfalussy Standards (BIS 1993b) specify that RTGS
systems must be in place, and the union’s umbrella set-

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

TA B L E 1

Features of Selected Funds Transfer Systems

Country

System
(Planned)

Type

Date

Central Bank
Daylight Credit
Yes

Belgium

ELLIPS

RTGS

1996

Canada

IIPS
(LVTS)

Net
Net

1976
1997

France

SAGITTAIRE
(TBF)

Net
RTGS

1984
1997

Germany

EIL-ZV
EAF2

RTGS
Net

1987
1996

Italy

BISS
(BI-REAL)
ME
SIPS

RTGS
RTGS
Net
Net

1989
1997
1989
1989

Japan

BOJ-NET
FEYCS

Net+RTGS
Net

1988
1989

Netherlands

FA
(TOP)

RTGS+Net
(RTGS)

1985
1997

Yes

Sweden

RIX

RTGS

1986

Yes

Switzerland

SIC

RTGS

1987

No

United Kingdom

CHAPS

RTGS

1984

Yes

United States

CHIPS
Fedwire

Net
RTGS

1970
1918

No
Yes

Yes
Yes

Yes

No

Source: BIS (1997b).

tlement system, Target, which will link settlement systems within the union, is also designed as a real-time
gross settlement system. The progress of the European
Union and the European Monetary Union have also contributed to the conversion of netting systems such as the
U.K. CHAPS system to real-time gross settlement even
though the United Kingdom is not projected to join the
European Monetary Union initially. Table 1 briefly summarizes some of the salient characteristics of settlement
systems in selected developed countries and illustrates
the extent to which they are evolving toward real-time
gross settlement.

Conclusions and Implications
he present path on which payments systems are
moving involves a seeming contradiction. On the
one hand, markets are becoming more integrated

T

and global in scope. At the same time they are becoming
more segmented in the sense that there is a growing separation evolving between the clearing and settlement of
transactions. This increasing separation raises the
prospect that there may be a need to invoke the safety net
and introduces a possible distortion into the international payments system. As a consequence, both publicsector and private markets have given great attention
to attempting to identify and control risk exposures.
Perhaps one of the more interesting developments in this
evolution of regional and globalized payments markets in
both the public and private sectors has been the push
toward real-time gross settlement systems with collateralization. Nowhere are these efforts more apparent than
in Europe, where the struggle to create a single financial marketplace has focused attention and generated
analyses of the underlying issues, with the Bank for

6. Real-time gross settlement may also improve risk management. In the case of derivatives clearinghouses, real-time gross settlement facilitates the use of intraday margin calls and the receipt of final funds before the end of the day.

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997

49

International Settlements, the Group of Ten, and central
banks spearheading much of this work.
Casual empiricism suggests several reasons why the
systems are evolving in this direction despite considerable analysis suggesting that netting arrangements are
more operationally efficient. The first reason is that systems, instruments, and markets are evolving faster than
the political entities can bring their various rules and
regulations into harmony despite the many initiatives
that have been undertaken. Second, harmonizing systems
to control effectively the systemic risks (such as Herstatt
risk) inherent in nonsynchronized clearing and settlement systems, such as foreign exchange markets, even if
all the legal rules are in place requires extensive interna-

tional coordination and cooperation. Third, central banks
realize that, regardless of the explicit rules governing
exchanges and settlement arrangements, they still may
be thrust into the role of the lender of last resort should
major participants get into financial difficulties that
threaten to bring down settlement and clearing systems.
In the United States, the decline in member bank reserve
balances reduces payments system participants’ liquidity
positions and increases the likelihood that intraday credit may have to be extended. Finally, the movement toward
expanding the overlapping hours that exchanges are
open will increasingly make the operation of net settlement systems more difficult.

R E F E R E N C E S
BANK OF ENGLAND AND APACS. 1994. “The Development of RealTime Gross Settlement (RTGS) in the United Kingdom.” Bank
for International Settlements information release, April.
BANK FOR INTERNATIONAL SETTLEMENTS (BIS). 1989. “Report on
Netting Schemes.” Prepared by the Group of Experts on
Payments Systems of Central Banks of the Group of Ten
Countries, February.
———. 1990. “Report of the Committee on Interbank Netting
Schemes of the Central Banks of the Group of Ten Countries.”
November.
———. 1993a. “Central Bank Payment and Settlement
Services with Respect to Cross-Border and Multi-Currency
Transactions.” Report prepared by the Committee on Payment
and Settlement Systems of the Central Banks of the Group of
Ten Countries, September.
———. 1993b. “Minimum Common Features for Domestic
Payments Systems.” Report of the Committee of Governors of
the Central Banks of the Member States of the European
Economic Community. Action 2 of the report on issues of common concern to EC central banks in the field of payments systems, by the Working Group on EC Payment Systems,
November.
———. 1993c. Payment Systems in the Group of Ten Countries. Report prepared by the Committee on Payment and
Settlement Systems of the Central Banks of the Group of Ten
Countries. Basle, December.
———. 1996. Settlement Risk in Foreign Exchange Transactions. Report prepared by the Committee on Payment and
Settlement Systems of the Central Banks of the Group of Ten
Countries. Basle, March.

and Settlement Systems of the Central Banks of the Group of
Ten Countries. Basle, March.
———. 1997b. Real-Time Gross Settlement Systems. Report
prepared by the Committee on Payment and Settlement
Systems of the Central Banks of the Group of Ten Countries.
Basle, March.
BENSTON, GEORGE J., AND GEORGE G. KAUFMAN. 1995. “Is the
Banking and Payments System Fragile?” Journal of Financial
Services Research 9 (December): 209-40.
CORRIGAN, E. GERALD. 1990. “Perspectives on Payments System
Risk Reduction.” In The U.S. Payments System: Efficiency,
Risk, and the Role of the Federal Reserve, edited by David B.
Humphrey. Proceedings of a Symposium on the U.S. Payments
System sponsored by the Federal Reserve Bank of Richmond.
Boston, Mass.: Kluwer Academic Publishers.
EDWARDS, FRANKLIN R. 1996. The New Finance: Regulation and
Financial Stability. Washington, D.C.: AEI Press.
EISENBEIS, ROBERT A. 1995. “Private-Sector Solutions to
Payments System Fragility.” Journal of Financial Services
Research 9 (December): 327-49.
GENERAL ACCOUNTING OFFICE. 1994. Financial Derivatives:
Actions Needed to Protect the Financial System. Report to
Congressional Requestors, GAO/GGD-94-133. Washington, D.C.:
Government Printing Office, May.
HU, JIE. 1994. “Information Ambiguity: Recognizing Its Role in
Financial Markets.” Federal Reserve Bank of Atlanta
Economic Review 79 (July/August): 11-21.
KNIGHT, FRANK H. [1921] 1971. Risk, Uncertainty, and Profit.
Reprint, Chicago: University of Chicago Press.

———. 1997a. Clearing Arrangements for Exchange-Traded
Derivatives. Report prepared by the Committee on Payment

50

Federal Reserve Bank of Atlanta E C O N O M I C R E V I E W Second Quarter 1997