03.10.2014 Views

Potential Output: Concepts and Measurement - Department of Labour

Potential Output: Concepts and Measurement - Department of Labour

Potential Output: Concepts and Measurement - Department of Labour

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Labour</strong> Market Bulletin 1995:1 Pages 72–115<br />

<strong>Potential</strong> <strong>Output</strong>: <strong>Concepts</strong> <strong>and</strong> <strong>Measurement</strong><br />

DARREN GIBBS 1<br />

This paper is concerned with the concept <strong>of</strong> potential output—the economy’s sustainable<br />

level <strong>of</strong> economic activity. The paper outlines the significance <strong>of</strong> potential output<br />

for policymakers, <strong>and</strong> surveys methods which have been devised to estimate potential<br />

output empirically. Preliminary analysis for New Zeal<strong>and</strong> suggests that the sustainable<br />

rate <strong>of</strong> economic growth currently lies in the range <strong>of</strong> 3–4 percent per annum. However,<br />

for many reasons, considerable uncertainty is attached to these estimates. This limits<br />

the usefulness <strong>of</strong> these measures in policymaking. 2<br />

1 Introduction<br />

THIS PAPER IS CONCERNED with the practical measurement <strong>of</strong> ‘potential<br />

output’. The concept is a familiar one to economists. Put simply, potential<br />

output is the level <strong>of</strong> economic activity at which aggregate dem<strong>and</strong> <strong>and</strong><br />

aggregate supply is consistent with a stable inflation rate. In a dynamic context,<br />

the economy’s ‘potential growth rate’ can be similarly defined. (The terms<br />

‘potential growth rate’ <strong>and</strong> ‘sustainable growth rate’ are <strong>of</strong>ten used interchangeably.)<br />

By calculating the level <strong>and</strong> rate <strong>of</strong> growth <strong>of</strong> potential output, <strong>and</strong><br />

comparing the results with observed output trends, one can obtain a measure <strong>of</strong><br />

the degree <strong>of</strong> spare capacity in the economy, <strong>and</strong> the rate at which capacity is<br />

exp<strong>and</strong>ing.<br />

The development <strong>of</strong> empirical measures <strong>of</strong> the degree <strong>of</strong> spare capacity in the<br />

economy, <strong>and</strong> the economy’s potential growth rate, is important for a number <strong>of</strong><br />

reasons. The most common use <strong>of</strong> such measures is as a guide for monetary <strong>and</strong><br />

fiscal policy, <strong>and</strong> as a source <strong>of</strong> information for use in developing forecasts <strong>of</strong><br />

economic growth. If reliable, they would provide an important guide for<br />

policymakers in determining whether developments in the real economy are<br />

consistent with the maintenance <strong>of</strong> price stability. In the New Zeal<strong>and</strong> context,<br />

they could also assist fiscal policymakers in determining whether spending<br />

decisions <strong>and</strong> tax settings are consistent with the Fiscal Responsibility Act 1994. 3<br />

1<br />

Darren Gibbs is an economist at the Reserve Bank <strong>of</strong> New Zeal<strong>and</strong>. He has worked<br />

previously as an adviser in the <strong>Labour</strong> Market Policy Group <strong>of</strong> the <strong>Department</strong> <strong>of</strong> <strong>Labour</strong>.<br />

2<br />

The paper has benfited from comments made by colleagues at the Reserve Bank <strong>and</strong> the<br />

<strong>Department</strong> <strong>of</strong> <strong>Labour</strong>. However, all views expressed are those <strong>of</strong> the author, <strong>and</strong> not<br />

necessarily those <strong>of</strong> the Reserve Bank or the <strong>Department</strong> <strong>of</strong> <strong>Labour</strong>.<br />

3<br />

Although most economists would share this view the consensus regarding the usefulness<br />

<strong>of</strong> empirical measures <strong>of</strong> potential output is not unanimous. For example, Plosser <strong>and</strong><br />

Schwert (1979) invoke the Lucas (1976) critique to argue that ‘these models cannot be used<br />

to predict the future path <strong>of</strong> output under different policy regimes. It would seem, therefore,<br />

that modelling “potential output” is an exercise with little merit, serving only to perpetuate<br />

the idea that its use as a policy guide can be justified through economic theory’ (p 185).


Darren Gibbs<br />

73<br />

These measures are also relevant for labour market analysts <strong>and</strong> commentators.<br />

In addition to their use in developing employment <strong>and</strong> wage projections,<br />

which rely on forecasts <strong>of</strong> the overall level <strong>of</strong> activity, they also play a part in<br />

informed decision-making regarding the feasibility <strong>of</strong> various labour market<br />

policy settings. For instance, a knowledge <strong>of</strong> how close to capacity the economy<br />

is can tell us about the extent to which efforts to reduce unemployment, or<br />

trends in employment growth, are likely to result in upward pressure on wages<br />

or prices.<br />

The aim <strong>of</strong> this paper is to present a brief survey <strong>of</strong> the techniques that have<br />

been used overseas to measure potential output, <strong>and</strong> to apply several <strong>of</strong> these<br />

techniques in order to estimate preliminary potential output series for New<br />

Zeal<strong>and</strong>. Section 2 defines what is meant by the term ‘potential output’, <strong>and</strong><br />

outlines more fully why the concept is <strong>of</strong> interest to central banks, governments<br />

<strong>and</strong> private sector businesses. The main techniques used by economists to<br />

empirically measure potential output are then discussed in section 3. Section 4<br />

presents the results obtained when applying a number <strong>of</strong> these techniques to<br />

New Zeal<strong>and</strong> data, <strong>and</strong> section 5 uses econometric techniques in an effort to<br />

select which <strong>of</strong> the estimates is preferable. The paper ends with a brief<br />

conclusion, which summarises the main ideas <strong>and</strong> results <strong>and</strong> suggests possible<br />

directions for future research.<br />

2 Economic growth <strong>and</strong> potential output<br />

Following World War II, a stable economic environment saw st<strong>and</strong>ards <strong>of</strong> living<br />

in most OECD countries grow rapidly. 4<br />

For many years New Zeal<strong>and</strong> also<br />

enjoyed strong economic growth <strong>and</strong> almost everybody who wanted a job was<br />

able to find work. However by the late 1960s economic performance had begun<br />

to deteriorate in most OECD countries, <strong>and</strong> the extent <strong>of</strong> deterioration was<br />

particularly marked in the case <strong>of</strong> New Zeal<strong>and</strong>. Whereas New Zeal<strong>and</strong>’s<br />

st<strong>and</strong>ard <strong>of</strong> living was ranked third in the developed world in 1950 (behind the<br />

United States <strong>and</strong> Canada), by 1975 it had dropped to fifteenth. During the<br />

decade following the first oil shock in 1973, New Zeal<strong>and</strong>’s economic<br />

performance slumped significantly. Growth during 1975 to 1984 averaged little<br />

more than 1.5 percent; <strong>and</strong> our relative st<strong>and</strong>ard <strong>of</strong> living dropped further to<br />

twenty-first place on the OECD ladder.<br />

The deterioration in economic performance prompted a serious rethink <strong>of</strong><br />

economic policy. As a consequence, since the mid 1980s, successive governments<br />

have implemented a comprehensive programme <strong>of</strong> economic <strong>and</strong> financial<br />

reform. These reforms have included: the deregulation <strong>of</strong> the banking industry;<br />

the freeing <strong>of</strong> controls on international capital flows (including the floating <strong>of</strong> the<br />

4<br />

The term ‘st<strong>and</strong>ard <strong>of</strong> living’ as used here refers only to measured pecuniary benefits as<br />

proxied by GDP.


74<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

exchange rate); the removal <strong>of</strong> a complex system <strong>of</strong> export subsidies, import<br />

licences <strong>and</strong> quotas; a programme <strong>of</strong> tariff reductions; the commercialisation<br />

(<strong>and</strong> in most cases privatisation) <strong>of</strong> State-owned market industries; the granting<br />

<strong>of</strong> operational independence to the Reserve Bank to pursue the sole target <strong>of</strong><br />

price stability; legislation designed to improve the ability <strong>of</strong> employers <strong>and</strong><br />

employees to negotiate more flexible <strong>and</strong> mutually beneficial employment<br />

relationships; <strong>and</strong>, latterly, legislation binding the Government to conduct fiscal<br />

policy in a manner which is both transparent <strong>and</strong> prudent. All together these<br />

‘supply-side’ policies are designed to improve the efficiency with which scarce<br />

resources are allocated within the economy, thus raising productivity growth <strong>and</strong><br />

increasing New Zeal<strong>and</strong>’s long-term sustainable rate <strong>of</strong> output growth.<br />

In undergoing these reforms New Zeal<strong>and</strong> has incurred substantial<br />

adjustment costs. However with economic growth reaching 6.0 percent in the<br />

year to March 1995, employment recovering strongly, <strong>and</strong> inflation <strong>and</strong> current<br />

account pressures broadly in check, many economists now agree that the fruits<br />

<strong>of</strong> these reforms are beginning to be seen. Consequently, the focus <strong>of</strong> attention<br />

has now turned to the following question: what is the economy’s sustainable rate<br />

<strong>of</strong> growth?<br />

Although the concept <strong>of</strong> potential output is widely used in economic analysis,<br />

there has been a considerable divergence <strong>of</strong> opinion as to its precise definition<br />

<strong>and</strong>, as we shall see in section 3, as to the best method <strong>of</strong> measuring it. Taken<br />

literally, the term means the maximum possible output <strong>of</strong> an economy if all <strong>of</strong><br />

its resources are ‘fully employed’. For example, at one extreme we could define<br />

potential output as the level <strong>of</strong> GDP that could be produced if everybody <strong>of</strong><br />

working age worked 24 hours per day, every day <strong>of</strong> the year. Alternatively, the<br />

term could be defined as some ‘normal’ level <strong>of</strong> production given ‘average’<br />

factor utilisation rates.<br />

The definition considered in most recent studies, however, is somewhat more<br />

restrictive. The contemporary approach is to define the potential output <strong>of</strong> an<br />

economy as the maximum level <strong>of</strong> output obtainable without generating an increase<br />

in inflation. Thus, calculations <strong>of</strong> potential output are based either explicitly or<br />

implicitly on estimates <strong>of</strong> the ‘natural rate <strong>of</strong> unemployment’—the rate <strong>of</strong><br />

unemployment which prevails when expectations <strong>of</strong> inflation are realised, <strong>and</strong><br />

toward which the economy will tend to converge following a disturbance<br />

(Friedman, 1968).<br />

Clearly, this concept is somewhat different from one that represents the<br />

maximum attainable level <strong>of</strong> output with a given set <strong>of</strong> inputs. The importance<br />

<strong>of</strong> defining potential output in a manner that is consistent with stable inflation<br />

was recognised by Okun (1962) in his seminal paper on potential output. Okun<br />

argued that:<br />

. . . potential GNP . . . is not a measure <strong>of</strong> how much output could be generated by<br />

unlimited amounts <strong>of</strong> aggregate dem<strong>and</strong>. The nation would probably be most


Darren Gibbs 75<br />

productive in the short run with inflationary pressures pushing the economy. But the<br />

social target <strong>of</strong> maximum production <strong>and</strong> employment is constrained by a social<br />

desire for price stability <strong>and</strong> free markets. The full employment goal must be<br />

understood as striving for maximum production without inflationary pressure . . .<br />

(p 99).<br />

There are a number <strong>of</strong> reasons why central banks <strong>and</strong> governments, <strong>and</strong> more<br />

generally the private sector, require accurate measures <strong>of</strong> the degree <strong>of</strong> spare<br />

capacity in the economy <strong>and</strong> the economy’s long-term sustainable growth rate.<br />

The Reserve Bank’s need for economy-wide estimates <strong>of</strong> potential output reflects<br />

its statutory responsibility to maintain stability in the general level <strong>of</strong> prices.<br />

Although previous Reserve Bank research has been quite successful in providing<br />

estimates <strong>of</strong> the inflationary consequences <strong>of</strong> an increase in, say, unit labour costs<br />

or the world price <strong>of</strong> tradeable goods, reliable estimates <strong>of</strong> the impact <strong>of</strong><br />

domestic dem<strong>and</strong> have proved more elusive, partly because researchers have<br />

been unsure <strong>of</strong> how to capture the influence <strong>of</strong> ‘excessive’ dem<strong>and</strong> pressures.<br />

Simple economic theory tells us that if dem<strong>and</strong> for goods <strong>and</strong> services in an<br />

economy outstrips supply at a given price level, then the price level will rise. But<br />

what level <strong>of</strong> dem<strong>and</strong> is consistent with stable prices? If the Reserve Bank<br />

mismeasures the level <strong>of</strong> potential output <strong>and</strong> calculates a positive output gap<br />

(dem<strong>and</strong> exceeds sustainable supply) when in fact the gap is negative<br />

(sustainable supply exceeds dem<strong>and</strong>), the consequent (ex-post procyclical)<br />

monetary policy tightening will tend instead to amplify the business cycle (<strong>and</strong><br />

the perturbations in inflation that stem from it), <strong>of</strong>fsetting some <strong>of</strong> the efficiency<br />

<strong>and</strong> allocative gains that a stable price environment may be expected to deliver.<br />

Similarly, if the Reserve Bank observes a rise in the rate <strong>of</strong> output growth, it<br />

needs to distinguish whether this is purely a positive dem<strong>and</strong> shock (<strong>and</strong><br />

therefore possibly inflationary) or simply the result <strong>of</strong> the economy responding<br />

to faster growth in the capacity <strong>of</strong> the economy to supply (<strong>and</strong> therefore not<br />

inflationary). Likewise, if the Reserve Bank observes a fall in actual output<br />

growth, it needs to distinguish whether this is the result <strong>of</strong> a negative dem<strong>and</strong><br />

shock (<strong>and</strong> therefore possibly deflationary) or a negative supply shock (<strong>and</strong><br />

therefore possibly inflationary). If policymakers wish to operate monetary policy<br />

in the least-cost manner possible (thus maximising the benefits), it is imperative<br />

that they have tools which will enable them to attempt to make such distinctions.<br />

The concept <strong>of</strong> potential output is also important from the perspective <strong>of</strong> the<br />

Government. In the short term, an assessment <strong>of</strong> the degree <strong>of</strong> excess dem<strong>and</strong><br />

or excess capacity in the economy will influence the fiscal policy stance. As with<br />

monetary policy, an over-expansionary stance at a time when excess dem<strong>and</strong><br />

conditions are evident or an over-contractionary stance when the spare resources<br />

are plentiful will tend to amplify the business cycle. In the medium term, views<br />

regarding the economy’s sustainable growth potential (<strong>and</strong> thus tax base) are<br />

also required to guide the formulation <strong>of</strong> the Government’s fiscal strategy. This


76<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

is particularly important in the case <strong>of</strong> New Zeal<strong>and</strong> where the Fiscal<br />

Responsibility Act requires the Government to conduct fiscal policy in a<br />

transparent <strong>and</strong> prudent manner, <strong>and</strong> to achieve specific reductions in the net<br />

public debt to GDP ratio. The calculation <strong>of</strong> ‘cyclically adjusted’ fiscal balances<br />

based on measures <strong>of</strong> potential output is st<strong>and</strong>ard practice in most other OECD<br />

countries (see, for example, Giorno et al, 1995). With regard to labour market<br />

policy, measures <strong>of</strong> potential output may also be <strong>of</strong> assistance in identifying the<br />

factors underlying unemployment <strong>and</strong> thus guide the appropriate policy<br />

responses. That part <strong>of</strong> unemployment which is identified as being unrelated to<br />

cyclical variation in economic activity will clearly require a different policy<br />

response from that which merely reflects the impact <strong>of</strong> a temporary shock to the<br />

economy.<br />

The importance <strong>of</strong> accurately estimating the economy’s sustainable growth<br />

rate is not restricted to the public sector. To the extent that dem<strong>and</strong> for some<br />

types <strong>of</strong> goods <strong>and</strong> services is related to the size <strong>of</strong> the aggregate economy,<br />

private sector investors must also form views on the economy’s sustainable<br />

growth rate to guide them in cost-benefit analysis <strong>and</strong> in making decisions<br />

regarding the timing <strong>of</strong> investments. This is particularly important for those<br />

investment projects in which there is a considerable time lag between the project<br />

planning <strong>and</strong> approval phase <strong>and</strong> project completion. For example, the planning<br />

<strong>and</strong> construction <strong>of</strong> hydroelectric power stations can take seven to eight years.<br />

Over this length <strong>of</strong> time, relatively small errors in the projection <strong>of</strong> annual<br />

economic growth can accumulative into large deviations in the expected size <strong>of</strong><br />

the economy, with the result that the project may be completed either years<br />

before or after the capacity is required.<br />

3 How is potential output measured?<br />

A number <strong>of</strong> alternative techniques for measuring potential output have been<br />

applied in the literature. These techniques range from the simple univariate<br />

smoothing techniques commonly used in the 1960s <strong>and</strong> early 1970s, to more<br />

complex structural techniques based on the ‘production function’ which became<br />

dominant in the late 1980s. More recently, techniques have been developed<br />

which provide a middle ground to these two extremes <strong>and</strong> which allow one to<br />

exploit relevant empirical information suggested by economic theory without<br />

imposing a rigid (<strong>and</strong> quite possibly incorrect) structure. This section briefly<br />

discusses the theory <strong>and</strong> methodology underlying the main techniques used, <strong>and</strong><br />

summarises the main advantages <strong>and</strong> disadvantages associated with each.<br />

3.1 Simple univariate techniques<br />

Univariate techniques use information on actual output to estimate potential<br />

output—no other information (such as trends in factor inputs) is used. The most


Darren Gibbs 77<br />

common procedure for estimating potential output is to fit a trend either to<br />

actual output or through its peaks. The choice between these two methods<br />

implies different views <strong>of</strong> the business cycle. The essential point is whether the<br />

gap between potential output <strong>and</strong> actual output over some cyclically neutral<br />

historical period is, on average, zero or positive. In the latter case there is<br />

presumably scope for appropriate macroeconomic policies to increase the mean<br />

level <strong>of</strong> output; but in the former case macroeconomic policies may be able to<br />

decrease the variability <strong>of</strong> output, but not increase its mean (assuming that initial<br />

macroeconomic policy settings are structurally optimal). 5<br />

The most attractive feature <strong>of</strong> these essentially theory-free estimates is their<br />

simplicity. The main univariate techniques are the trends through peaks<br />

approach, the linear time trend approach, <strong>and</strong> the Hodrick-Prescott filter<br />

approach. Discussion <strong>of</strong> the latter is reserved until later in this section where a<br />

more general application <strong>of</strong> filtering techniques is outlined.<br />

Trends through peaks approach<br />

The trends through peaks approach is one <strong>of</strong> the most popular techniques<br />

considered in the literature, <strong>and</strong> was the accepted methodology in the 1960s <strong>and</strong><br />

early 1970s. To apply this approach, seasonally adjusted output is first graphed.<br />

It is then assumed that ‘major’ peaks in the series represent output where<br />

resources in the economy are utilised at 100 percent <strong>of</strong> capacity. A straight line is<br />

drawn between each <strong>of</strong> the major peaks, <strong>and</strong> is then extrapolated (using the<br />

same slope as the line connecting the previous two peaks) beyond the last one.<br />

This line is taken to be potential output. The output gap is represented by the<br />

difference between the line representing potential output, <strong>and</strong> actual output.<br />

Because the approach results in a succession <strong>of</strong> discrete changes in a linear time<br />

trend, it is also sometimes referred to as the ‘split time trend’ method.<br />

This approach reflected the prevailing economic thinking at that time. The<br />

supply side <strong>of</strong> the economy was seen as deterministic, with changes in dem<strong>and</strong><br />

seen as the prime factor underlying observed business cycles. By focusing on the<br />

peaks <strong>of</strong> cycles, potential output was defined as the maximum possible output<br />

in a physical capacity or engineering sense; <strong>and</strong> so the calculated output gaps<br />

were almost always negative (ie actual output was below potential). As Laxton<br />

<strong>and</strong> Tetlow (1992) note, this reflected a belief that the unbridled economy tends<br />

towards inefficient outcomes, <strong>and</strong> that the goal <strong>of</strong> macroeconomic policy should<br />

be to <strong>of</strong>fset this tendency.<br />

As growth through the 1960s was relatively constant, <strong>and</strong> inflation was low<br />

<strong>and</strong> stable, the conceptual link between the output gap <strong>and</strong> inflation which had<br />

5<br />

As Laxton et al (1994) note, if the short-run Phillips curve is asymmetric, trend output<br />

will lie below potential output. In this circumstance, a policy aimed at decreasing the<br />

variability <strong>of</strong> output will also increase mean output.


78<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

been identified by Okun was ignored. However with the advent <strong>of</strong> greater<br />

variability in growth <strong>and</strong> inflation in the early 1970s, the practical shortcomings<br />

<strong>of</strong> this approach were exposed.<br />

Christiano (1981) identifies three such shortcomings. First, deciding what<br />

constitutes a major peak is purely a matter <strong>of</strong> judgment, <strong>and</strong> the assumption that<br />

each major peak represents the same intensity <strong>of</strong> resource utilisation is bold.<br />

Secondly, it is unreasonable to assume that potential output grows at a constant<br />

arithmetic rate between peaks, <strong>and</strong> that structural breaks only occur at the peaks.<br />

Thirdly, the potential output line which follows the most recent peak, a period<br />

particularly important for policy purposes, is subject to substantial revision<br />

when a new peak occurs. Hence, potential output estimates created using this<br />

technique will be unreliable when the economy is suspected to have undergone<br />

a structural shift, or when the rate <strong>of</strong> capital accumulation <strong>and</strong>/or growth in the<br />

labour force have changed. This latter shortcoming seems particularly applicable<br />

to New Zeal<strong>and</strong>.<br />

As the ‘natural rate’ concept became more widely accepted in the economics<br />

pr<strong>of</strong>ession in the mid 1970s, <strong>and</strong> the predictions made by the trends through<br />

peaks estimates proved increasingly misleading, the use <strong>of</strong> this method became<br />

less widespread. However, up until recently, the trends through peaks method<br />

was still used by the OECD to decompose actual output into trend <strong>and</strong> cycle<br />

when constructing fiscal policy indicators (see Giorno et al, 1995).<br />

Linear time trend approach<br />

The initial response <strong>of</strong> researchers to the demise <strong>of</strong> the trends through peaks<br />

approach was to estimate linear time trends that were calculated to run through<br />

the centre <strong>of</strong> the business cycle. While this was an advance, in that this approach<br />

utilised all the information contained in the actual output series (rather than just<br />

the peaks) <strong>and</strong> allowed for the possibility <strong>of</strong> both negative <strong>and</strong> positive output<br />

gaps, this technique generally assumed that actual output was equal to potential<br />

output on average over the business cycle, for which there was no guarantee<br />

(particularly over small samples). Not surprisingly, therefore, this approach also<br />

yielded implausible estimates <strong>of</strong> potential output.<br />

As world economic growth began to slow down in the mid 1970s, in response<br />

to the oil price shocks <strong>and</strong> lower growth rates in measured productivity,<br />

researchers began to turn their attention to the impact <strong>of</strong> supply-side influences<br />

on economic activity. Initially however most work continued to assume that the<br />

supply side <strong>of</strong> the economy evolved only gradually. Rather than seek to explain<br />

these developments explicitly, researchers tended to resort to the use <strong>of</strong> such<br />

tools as moving averages, spline functions <strong>and</strong> shift dummies. Indeed, Laxton<br />

<strong>and</strong> Tetlow quote a Bank <strong>of</strong> Canada research memor<strong>and</strong>um describing the<br />

method used for specifying total factor productivity:


Darren Gibbs 79<br />

To put it starkly, given the data <strong>and</strong> research, both a sharp eye <strong>and</strong> a flexible 18-inch<br />

ruler were applied to the data in a manner consistent with existing priors. (From<br />

Gosselin et al, 1981.)<br />

It soon became clear that more sophisticated techniques were required. New<br />

structural approaches began to be developed which sought to explain growth in<br />

terms <strong>of</strong> factor inputs. In retrospect, it seems surprising that this advance was<br />

so long in coming given that the theoretical growth models based on such<br />

relationships had been developed as far back as the late 1940s by Harrod (1959)<br />

<strong>and</strong> Domar (1947).<br />

3.2 Multivariate structural techniques<br />

Unlike the approaches discussed above, multivariate techniques make use <strong>of</strong><br />

variables other than just actual output to derive estimates <strong>of</strong> potential output.<br />

Typically these techniques involve relating potential output to trends in the<br />

quantity (<strong>and</strong> in some cases quality) <strong>of</strong> labour <strong>and</strong>/or capital inputs. In this<br />

section, approaches are outlined based on the seminal work <strong>of</strong> Okun (1962); the<br />

output/capital approach; the inversion <strong>of</strong> input dem<strong>and</strong> functions; <strong>and</strong> the<br />

production function approach.<br />

Okun’s approach<br />

A simple, widely used <strong>and</strong> long st<strong>and</strong>ing method <strong>of</strong> calculating potential output<br />

is based on the seminal work <strong>of</strong> Arthur Okun (1962). Okun’s method relies less<br />

on economic theory than it does on an empirical regularity, <strong>and</strong> is based on a<br />

simple relationship linking the gap between actual <strong>and</strong> potential output with the<br />

gap between actual <strong>and</strong> natural rates <strong>of</strong> unemployment. This approach,<br />

therefore, directly links the estimation <strong>of</strong> potential output to a judgment or<br />

estimate <strong>of</strong> the natural rate <strong>of</strong> unemployment.<br />

In Okun’s original work, using US data from 1947Q2 to 1960Q1, the natural<br />

rate <strong>of</strong> unemployment was assumed (relatively uncontroversially) to be 4<br />

percent. Therefore, using Okun’s method, potential output equals actual output<br />

when aggregate dem<strong>and</strong> is exactly at the level that yields an unemployment rate<br />

<strong>of</strong> 4 percent. If aggregate dem<strong>and</strong> is lower, part <strong>of</strong> the potential output is not<br />

produced—there is unrealised potential or a gap between actual <strong>and</strong> potential<br />

output. 6<br />

6<br />

Okun argued that to the extent that low utilisation rates <strong>and</strong> accompanying low pr<strong>of</strong>its<br />

<strong>and</strong> personal incomes hold down investment in plant, equipment, research, housing <strong>and</strong><br />

education, the failure to use one year’s potential fully can influence future potential<br />

output. Because today’s actual output influences tomorrow’s productive capacity, success<br />

in the stabilisation objective promotes more rapid economic growth. This recognition that<br />

history can affect the future constitutes an early expression <strong>of</strong> the hysteresis ideas which<br />

are prevalent in recent literature.


80<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

Okun employed three different methods <strong>of</strong> relating output to the unemployment<br />

rate. The most widely known method was based on a regression linking<br />

quarterly changes in the unemployment rate (U t<br />

, expressed in percentage points)<br />

<strong>and</strong> quarterly percentage changes in GNP (Q t<br />

). That is:<br />

( U U )<br />

t<br />

− = α + β<br />

t−1<br />

( Q − Q )<br />

t<br />

Q<br />

t−1<br />

t−1<br />

According to the estimates from this equation, the unemployment rate will rise<br />

by α percentage points from one quarter to the next if real GNP is unchanged,<br />

as secular gains in productivity <strong>and</strong> growth in the labour force push up the<br />

unemployment rate. For each extra 1 percent <strong>of</strong> GNP, unemployment is β<br />

percentage points lower. Furthermore, from a position where actual <strong>and</strong><br />

potential output are equal, these estimates imply that output must grow by α β<br />

percent per quarter in order to stabilise the unemployment rate. Thus, given the<br />

assumption that the unemployment rate will not change when the economy<br />

grows at its potential rate, α β provides an estimate <strong>of</strong> the potential growth rate. 7<br />

As Okun himself acknowledged, his methodology is very simplistic. A<br />

number <strong>of</strong> weaknesses have been identified in the literature. First, in making a<br />

direct connection between unemployment <strong>and</strong> output Okun assumed that the<br />

influence <strong>of</strong> slack economic activity on average hours, labour force participation<br />

<strong>and</strong> productivity are all proxied by the unemployment rate.<br />

Secondly, the method relies on the assumption that the same statistical<br />

relations prevailing in the historical data would have applied if the economy had<br />

been operating at full employment. No correction is made for factors such as<br />

working age population growth or increased immigration, which would imply<br />

more unemployment for any given level <strong>of</strong> output (we correct for this in our<br />

empirical estimates later in the paper by defining output in per capita terms).<br />

Thirdly, the natural rate <strong>of</strong> unemployment is determined by assumption, <strong>and</strong><br />

assumed to be constant over the relevant time period. 8 This led Gordon (1979)<br />

to argue that:<br />

The original Okun/Heller concept . . . should have been called not potential output,<br />

but arbitrarily defined full employment output (p 187).<br />

7<br />

Christiano (1981) slightly extends Okun’s original approach by allowing unemployment<br />

rate changes to depend on various lagged values <strong>of</strong> GNP growth. Assuming that the<br />

estimated lag coefficients would have held if the economy had been operating at full<br />

employment, Christiano derives estimates <strong>of</strong> potential output <strong>and</strong> output gaps.<br />

8<br />

This problem can be eliminated by separately estimating a series for the natural rate <strong>of</strong><br />

unemployment which varies through time. Techniques to achieve this will be discussed<br />

later when we outline the more complex multivariate approaches to the measurement <strong>of</strong><br />

potential output which routinely estimate a time-varying natural rate <strong>of</strong> unemployment as<br />

part <strong>of</strong> the general approach.


Darren Gibbs 81<br />

Fourthly, Plosser <strong>and</strong> Schwert (1979) demonstrate both theoretically <strong>and</strong><br />

empirically that Okun’s procedure <strong>of</strong> inverting a regression <strong>of</strong> unemployment on<br />

output is valid only if the correlation coefficient is equal to ±1.<br />

Finally, a more significant problem with the technique is that it can result in<br />

potential output series which display unreasonable volatility. Indeed, in a recent<br />

study by the OECD (1994) using an Okun’s Law based technique, they found<br />

that the potential output series derived for some countries was more volatile<br />

than the actual output series on which they were based.<br />

Because <strong>of</strong> these shortcomings, the reliability <strong>of</strong> potential output estimates<br />

obtained using these techniques also suffered as output <strong>and</strong> inflation became<br />

more variable in the 1970s. As noted by Adams et al (1987):<br />

. . . because these approaches tend to deal somewhat mechanically with any change<br />

in economic conditions, they are not well suited to separating permanent from<br />

transitory influences. These approaches tend to extrapolate any change in growth<br />

automatically, provided it is maintained long enough, <strong>and</strong>, even when they produce<br />

correct judgments about historical developments in potential output, they do not<br />

identify the underlying determinants. Such methods are, therefore, unreliable for<br />

forecasting, particularly when the economic environment is undergoing substantial<br />

change.<br />

Adams’ last comment is particularly pertinent in the case <strong>of</strong> New Zeal<strong>and</strong>.<br />

The output-capital ratio approach<br />

Whereas Okun’s measure relies solely on the labour market as an indicator <strong>of</strong><br />

the output gap, this approach hinges on the existence <strong>of</strong> a stable proportional<br />

relation between the stock <strong>of</strong> capital <strong>and</strong> potential output. The method assumes<br />

that fluctuations in the observed output/capital ratio are largely due to<br />

deviations in output from its potential level.<br />

An example <strong>of</strong> the use <strong>of</strong> this method is contained in Panic (1978). First he<br />

constructs an output/capital ratio series ( Q t<br />

K ), <strong>and</strong> then he fits a least squares<br />

t<br />

linear trend to the actual output/capital series, to yield a ‘capacity’ output/<br />

capital series, as follows:<br />

⎛ Qt<br />

⎞<br />

⎜ ⎟ = α0 + α1t+<br />

e<br />

⎝ K ⎠<br />

t<br />

t<br />

where Q t<br />

<strong>and</strong> K t<br />

represent output <strong>and</strong> the capital stock at time t <strong>and</strong> e t<br />

is a<br />

r<strong>and</strong>om error. The capacity output/capital ratio is taken to be the points on a line<br />

with the time derivative α 1<br />

raised just enough so that it touches only one <strong>of</strong> the<br />

observed Q t<br />

Q<br />

c<br />

t<br />

K data points. The adjusted trend ratio ( )<br />

t<br />

K is the assumed capacity<br />

t<br />

output/capital ratio. From here, a series measuring potential output is simply<br />

Q<br />

c<br />

t<br />

.<br />

given by K t ( K )<br />

t


82<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

As with the other approaches that we have described so far, this approach is<br />

computationally relatively simple to apply. However a major weakness in this<br />

approach is that it relies exclusively on capital stock series, which are among the<br />

least reliable series available. 9 Estimates <strong>of</strong> the capital stock generally rely on an<br />

initial starting value for the capital stock, updated using gross investment data,<br />

allowing for an assumed rate <strong>of</strong> depreciation. There is a variety <strong>of</strong> problems<br />

associated with such estimates including: the reliability <strong>of</strong> initial capital stock<br />

estimates; the use <strong>of</strong> accounting rather than economic rates <strong>of</strong> depreciation;<br />

capital vintage effects; scrapping rates; <strong>and</strong> the use <strong>of</strong> capital definitions that<br />

exclude components such as human capital.<br />

Inverted input dem<strong>and</strong> approach<br />

Approaches have also been derived which rely on the inversion <strong>of</strong> estimated<br />

input dem<strong>and</strong> functions. One such approach is based on first estimating the<br />

desired quantity <strong>of</strong> labour or capital input as a function <strong>of</strong> output <strong>and</strong> other<br />

variables. Next, given the estimated dem<strong>and</strong> function, the desired input level is<br />

replaced with the actual input level in the equation, <strong>and</strong> the equation is solved<br />

for output. On the assumption that the relationship between desired input <strong>and</strong><br />

output holds at efficient points in production, the output series computed in the<br />

second step <strong>of</strong> the procedure can be taken as an estimate <strong>of</strong> potential output.<br />

Christiano (1981) provides two examples <strong>of</strong> the use <strong>of</strong> this approach. The first,<br />

calculated by the US Council <strong>of</strong> Economic Advisors (CEA, 1979), relies on a<br />

labour dem<strong>and</strong> function. The following labour dem<strong>and</strong> function relating the<br />

labour input gap to current <strong>and</strong> lagged output gaps is the basis for this approach:<br />

⎛ L ⎞<br />

t<br />

log<br />

⎜<br />

p ⎟<br />

⎝ L ⎠<br />

=<br />

where:<br />

t<br />

n<br />

∑<br />

s=<br />

0<br />

⎛ Q<br />

bs<br />

log<br />

⎜<br />

⎝ Q<br />

t−s<br />

p<br />

t−s<br />

⎞<br />

⎟<br />

⎠<br />

L t<br />

= hours dem<strong>and</strong>ed measured in efficiency units 10<br />

L p t<br />

= full employment (or potential) hours dem<strong>and</strong>ed in efficiency units<br />

Q t<br />

= actual output<br />

Q p t<br />

= capacity (or potential) output<br />

<strong>and</strong> where t = 1, . . ., T <strong>and</strong> n <strong>and</strong> T are finite numbers. It is assumed that the<br />

sum <strong>of</strong> the b coefficients approximately equals unity.<br />

10<br />

Efficiency units are derived by weighting hours for different sub-groups according to<br />

their relative wage.<br />

9<br />

In the case <strong>of</strong> New Zeal<strong>and</strong> there are no <strong>of</strong>ficial capital stock data. The only estimates<br />

available are those estimated by the Reserve Bank <strong>of</strong> New Zeal<strong>and</strong> <strong>and</strong> the Research Project<br />

on Economic Planning at Victoria University. Both are estimated using the perpetual<br />

inventory approach.


Darren Gibbs 83<br />

p<br />

⎛ Qt<br />

Assuming that labour productivity ⎜ ⎞<br />

p<br />

⎟<br />

⎝ Lt<br />

⎠ follows a simple time pattern,<br />

Christiano rearranges the above equation to express actual labour input as a<br />

function <strong>of</strong> actual output <strong>and</strong> that time pattern, <strong>and</strong> obtains estimates <strong>of</strong> the b s<br />

coefficients. Using these coefficients <strong>and</strong> an independent measure <strong>of</strong> L p t<br />

, 11 the<br />

dem<strong>and</strong> equation can be inverted to yield an estimate <strong>of</strong> the output gap as a<br />

function <strong>of</strong> lagged values <strong>of</strong> the input gap.<br />

A second approach which makes use <strong>of</strong> a capital (or investment) dem<strong>and</strong><br />

function was proposed by Hickman (1964). In this approach the dem<strong>and</strong> for<br />

capital, K * t<br />

, is expressed as:<br />

* * *<br />

t 0 1 t 2 t 3<br />

log K = α + α logQ + α log P + α t<br />

where Q t * <strong>and</strong> P t * represent permanent income <strong>and</strong> relative output prices at time<br />

t. (The model assumes annual data.)<br />

Assuming that the capital stock follows a partial adjustment process, <strong>and</strong> that<br />

both permanent income <strong>and</strong> relative output prices depend linearly on current<br />

<strong>and</strong> lagged values <strong>of</strong> actual output <strong>and</strong> actual prices respectively, it is possible<br />

to estimate Q t * as a function <strong>of</strong> capital <strong>and</strong> output price series.<br />

The relative advantage <strong>of</strong> the inverted dem<strong>and</strong> function approaches discussed<br />

above is that although, computationally, they are slightly more burdensome than<br />

the other approaches discussed so far, they do go some way towards addressing<br />

some <strong>of</strong> the weaknesses in the simpler approaches. They are also simpler to<br />

apply than the next technique we will discuss, which requires us to estimate the<br />

underlying production function. However, as is usually the case, the search for<br />

computational simplicity requires a trade-<strong>of</strong>f to be made with theoretical<br />

desirability. Strong assumptions need to be made about functional forms <strong>and</strong> the<br />

time behaviour <strong>of</strong> various variables.<br />

There are a number <strong>of</strong> difficulties with applying these approaches to New<br />

Zeal<strong>and</strong> data. In the case <strong>of</strong> the labour dem<strong>and</strong> approach, sufficiently disaggregated<br />

data over a long enough time period is either extremely unreliable or<br />

simply unavailable. Without this disaggregation, the approach differs little from<br />

that suggested by Okun; <strong>and</strong> the capital dem<strong>and</strong> approach has all the limitations<br />

that were discussed earlier in the context <strong>of</strong> the output/capital ratio approach.<br />

The production function approach<br />

The final approach that we shall discuss in this section is the production function<br />

approach which, at least on theoretical grounds, is by far the most desirable. The<br />

essence <strong>of</strong> the production function approach is an explicit modelling <strong>of</strong> output<br />

in terms <strong>of</strong> the underlying factor inputs, which involves the specification <strong>and</strong><br />

11<br />

Christiano derives a measure <strong>of</strong> potential hours dem<strong>and</strong> by applying cyclically adjusted<br />

participation <strong>and</strong> unemployment rates to eight age-gender groups in the population.<br />

Hours are converted to efficiency units by valuing them using relative wages. This is <strong>of</strong>ten<br />

known as ‘Perry weighting’, based on Perry (1971).


84<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

estimation <strong>of</strong> production functions linking output to factor inputs, as well as the<br />

determination <strong>of</strong> the quantity <strong>and</strong> quality <strong>of</strong> inputs. The non-accelerating<br />

inflation constraint is usually introduced by relating the potential level <strong>of</strong> factor<br />

inputs to the natural level <strong>of</strong> unemployment.<br />

Adams et al (1987) identify four components to the production function<br />

approach. First, a two factor production function, relating the output <strong>of</strong> the<br />

private business sector to the inputs <strong>of</strong> labour <strong>and</strong> capital <strong>and</strong> to total factor<br />

productivity. Secondly, a pair <strong>of</strong> equations that relate the intensities with which<br />

labour <strong>and</strong> capital are used over the business cycle to the ratio <strong>of</strong> actual to<br />

normal output. Thirdly, a pair <strong>of</strong> equations which are used to determine the<br />

potential levels <strong>of</strong> the factor inputs. 12 And fourthly, an equation or wage-price<br />

model that is used to determine the natural rate <strong>of</strong> unemployment.<br />

This approach relies heavily on the production function, but does not require<br />

the explicit modelling <strong>of</strong> the dem<strong>and</strong> for <strong>and</strong> supply <strong>of</strong> factor inputs <strong>and</strong> total<br />

factor productivity. The approach assumes that, in the short term, the potential<br />

inputs <strong>of</strong> factors are determined principally by the behaviour <strong>of</strong> unemployment<br />

relative to its natural rate, <strong>and</strong> output relative to its normal level. It is further<br />

assumed that the growth in factor inputs <strong>and</strong> total factor productivity<br />

underlying the projections <strong>of</strong> potential output can be based on judgments about<br />

the macroeconomic environment supplemented by explicit views on key<br />

relationships.<br />

Variants <strong>of</strong> the production function approach are used by both the International<br />

Monetary Fund (IMF) (Adams et al, 1987; Adams <strong>and</strong> Coe, 1990) <strong>and</strong><br />

the Organisation for Economic Cooperation <strong>and</strong> Development (OECD) (Torres<br />

<strong>and</strong> Martin, 1990).<br />

Adams et al (1987) identified four major advantages that the production<br />

approach has over the more traditional approaches discussed so far. First, it<br />

allows for an explicit accounting for growth in terms <strong>of</strong> the contributions <strong>of</strong><br />

factor inputs <strong>and</strong> total factor productivity. Secondly, it explicitly accounts for the<br />

links between the product <strong>and</strong> factor markets that underlie relationships such as<br />

Okun’s Law, without imposing a constant Okun coefficient. Thirdly, it allows one<br />

to analyse the impact <strong>of</strong> various disturbances including, for example, an increase<br />

in energy prices. Finally, it can be adapted for forecasting purposes to determine<br />

the underlying rates <strong>of</strong> economic growth that may be envisaged, <strong>and</strong> the effects<br />

<strong>of</strong> such factors as changes in the natural rate <strong>of</strong> unemployment.<br />

However, as with all the approaches discussed so far, this approach also has<br />

disadvantages. First, its theoretical supremacy is at the cost <strong>of</strong> computational<br />

12<br />

These equations relate the inputs <strong>of</strong> factors to the deviation <strong>of</strong> the unemployment rate<br />

from its natural rate, to the ratio <strong>of</strong> actual to normal output, <strong>and</strong> to a number <strong>of</strong> factor<br />

specific circumstances. <strong>Potential</strong> inputs are found by determining the levels <strong>of</strong> input when<br />

unemployment is at its natural rate <strong>and</strong> output is at its normal level (hence introducing the<br />

inflation constraint).


Darren Gibbs 85<br />

simplicity. The approach requires that an assumption be made regarding the<br />

mathematical form <strong>of</strong> the production function. Typically, the Cobb-Douglas <strong>and</strong><br />

CES specifications are used as well as a variety <strong>of</strong> more flexible forms, though<br />

economists have had a great deal <strong>of</strong> difficulty agreeing on which form is best.<br />

Furthermore, in addition to the need to measure the quantity <strong>of</strong> inputs used<br />

in the process (which, as already outlined, is a feature <strong>of</strong> structural approaches),<br />

the approach also requires that some estimate is made <strong>of</strong> the quality <strong>of</strong> factor<br />

inputs. Studies as early as that <strong>of</strong> Solow (1957) have found that the physical<br />

quantity <strong>of</strong> inputs provides a poor explanation <strong>of</strong> actual movements in output.<br />

Indeed, this study found that factor inputs could explain only one-third <strong>of</strong><br />

growth with the ‘Solow residual’ or total factor productivity accounting for the<br />

remainder. It is this residual that captures changes in the quality <strong>of</strong> inputs <strong>and</strong><br />

production processes. As Laxton <strong>and</strong> Tetlow note:<br />

One important advantage <strong>of</strong> the production function approach over the time trend<br />

approach is that it is easy to keep track <strong>of</strong> the major contributing factors to potential<br />

output. That is, variations in its underlying determinants such as labour force<br />

growth, capital formation, <strong>and</strong> trend total factor productivity. For this reason the<br />

production function approach was seen as an attractive framework for organising the<br />

data. However, there remained a problem. Since economists had no useful model <strong>of</strong><br />

the determinants <strong>of</strong> productivity, the resulting estimates were still essentially<br />

exogenous time trends (p 7).<br />

The use <strong>of</strong> ad hoc polynomial time trends as regressors in production functions<br />

to explain productivity (see, for example, Perl<strong>of</strong>f <strong>and</strong> Wachter, 1979), together<br />

with their quantitative importance to the results, constituted a very significant<br />

weakness in the empirical application <strong>of</strong> the production function approach.<br />

Moreover estimates derived from these techniques proved to be very sensitive<br />

to the type <strong>of</strong> approach used. Together these factors severely limited the practical<br />

usefulness <strong>of</strong> the estimates.<br />

Although the recent theoretical development <strong>of</strong> the endogenous growth<br />

literature has provided some insights <strong>and</strong> stimulated debate regarding the<br />

determinants <strong>of</strong> long-run growth, endogenous growth theory has received little<br />

attention in empirical research to date.<br />

Partly in response to the above problems, a new literature emerged in the<br />

early 1980s in which changes in potential output <strong>and</strong> the natural rate <strong>of</strong><br />

unemployment were treated as stochastic (rather than deterministic) phenomena.<br />

3.3 Stochastic filtering techniques<br />

Laxton <strong>and</strong> Tetlow (1992) argue that there is insufficient knowledge about the<br />

true structure <strong>of</strong> the economy to make the structural approach practical.<br />

Moreover, researchers using structural techniques have tended to concentrate on<br />

highly stylised models focusing on specific, easily identifiable shocks, <strong>and</strong> have


86<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

tended to cling too long to obsolete estimates <strong>of</strong> potential output or to overinterpret<br />

apparent changes in potential output.<br />

Laxton <strong>and</strong> Tetlow argue that potential output is best characterised as being<br />

driven by a stochastic process so that in addition to the myriad <strong>of</strong> policy, demographic,<br />

<strong>and</strong> commodity price shocks there are regular unidentifiable shocks,<br />

some <strong>of</strong> which have lasting effects on potential output. Given the extensive<br />

programme <strong>of</strong> economic <strong>and</strong> financial reform in New Zeal<strong>and</strong> over the last ten<br />

years, <strong>and</strong> the change in behaviour <strong>and</strong> decision-making patterns (<strong>and</strong> thus<br />

structural relationships) that are likely to have taken place, it seems particularly<br />

reasonable to reach this conclusion in the case <strong>of</strong> New Zeal<strong>and</strong>.<br />

They also note a tendency to use much more judgment <strong>and</strong> more structured,<br />

but less detailed, macroeconomic models. This they attribute to:<br />

. . . a by-product <strong>of</strong> the realisation that, although economists have a broad grasp <strong>of</strong><br />

the relationships between macroeconomic activity <strong>and</strong> such observables as potential<br />

output, the structural form <strong>of</strong> these relationships remains elusive. This would seem<br />

to call for a method <strong>of</strong> measuring potential output that reflects general beliefs<br />

concerning the nature <strong>of</strong> these relationships, without imposing too much formal<br />

structure (p 12).<br />

It is the difficulty <strong>of</strong> finding sufficient evidence to arrive at a consensus on<br />

structure that has spawned astructural methods <strong>of</strong> measurement such as the<br />

Hodrick-Prescott filter, at least as complementary tools in much the same way<br />

as VAR models can be seen as complementary to structural macroeconomic<br />

models. The Hodrick-Prescott filter is a popular technique for smoothing time<br />

series data <strong>and</strong> for this reason it worth discussing it in its own right. However,<br />

it is also interesting because it encompasses the deterministic time trend as a<br />

special case <strong>and</strong> it provides the basis for the multivariate filter technique<br />

suggested by Laxton <strong>and</strong> Tetlow which is discussed later in this section.<br />

Hodrick-Prescott filter approach<br />

The rationale for using the Hodrick-Prescott (HP) filter (Hodrick <strong>and</strong> Prescott,<br />

1980) is that it can help to decompose an observed shock into a supply <strong>and</strong> a<br />

dem<strong>and</strong> component. This is done by making the identifying assumption that<br />

supply shocks have permanent or lasting effects on output while dem<strong>and</strong> shocks<br />

tend to be temporary. Of course, if the temporary dem<strong>and</strong> component contains<br />

a great deal <strong>of</strong> persistence it is very difficult to distinguish between the two,<br />

particularly at the end <strong>of</strong> the sample. Moreover, there is also the possibility <strong>of</strong><br />

supply shocks which are temporary (eg due to climatic factors such as the 1992<br />

electricity crisis). Furthermore, the two sources <strong>of</strong> shock need not be independent.<br />

In endogenous growth models, cycles <strong>and</strong> growth are part <strong>of</strong> the same<br />

process. Nevertheless, as noted earlier, since supply shocks <strong>and</strong> dem<strong>and</strong> shocks<br />

have inherently different implications for inflation we need some way <strong>of</strong><br />

disentangling dem<strong>and</strong> shocks from supply shocks.


Darren Gibbs 87<br />

The HP filter is derived by minimising the sum <strong>of</strong> the squared deviations <strong>of</strong><br />

a variable (in this case, output), Q t<br />

, from its trend τ, subject to a smoothness<br />

constraint, λ, that penalises squared variations in the growth <strong>of</strong> the trend series.<br />

That is, the HP filter is calculated to minimise:<br />

T<br />

T−1<br />

QHP =<br />

∑ t − τ t + λ ∑<br />

τ t − τ − τ t − τ t<br />

t=<br />

1<br />

t=<br />

2<br />

[ + 1 1 −1<br />

]<br />

2<br />

( Q ) ( ) ( )<br />

Thus, the calculated HP trend series is a function <strong>of</strong> the smoothness constraint,<br />

<strong>and</strong> <strong>of</strong> both past <strong>and</strong> future values <strong>of</strong> output. The user can determine the<br />

smoothness in the trend series by choosing an appropriate value for the<br />

smoothness parameter. Higher values <strong>of</strong> λ imply a larger weight on the<br />

smoothness <strong>of</strong> the series, so that in the limit, as λ becomes arbitrarily large, the<br />

trend series will converge on a linear time trend. As Laxton <strong>and</strong> Tetlow note, this<br />

would be consistent with an extreme Keynesian model in which supply shocks<br />

are deterministic, <strong>and</strong> variations in output come almost entirely from dem<strong>and</strong><br />

shocks. Conversely, a very small value <strong>of</strong> λ will effectively eliminate the penalty<br />

function, <strong>and</strong> thus the HP trend series will be equal to the actual series. This<br />

would be consistent with an extreme real-business-cycle model in which most<br />

variations in output are also variations in potential or trend output, <strong>and</strong> hence<br />

are driven by supply shocks.<br />

King <strong>and</strong> Rebelo (1989) show that the optimal value <strong>of</strong> λ is a function <strong>of</strong> the<br />

ratio <strong>of</strong> the variances <strong>of</strong> these shocks <strong>and</strong> that, for some data generating process<br />

<strong>of</strong> permanent <strong>and</strong> temporary shocks, the HP filter is an optimal (minimumvariance)<br />

linear inverse filter. However, there is considerable debate concerning<br />

the relative variance <strong>of</strong> supply <strong>and</strong> dem<strong>and</strong> shocks. In practice, users <strong>of</strong> the HP<br />

filter have tended to set λ = 1600 when smoothing quarterly data following the<br />

initial approach <strong>of</strong> Hodrick <strong>and</strong> Prescott. An alternative approach is to choose λ<br />

so as to generate an output gap series which is consistent with the results <strong>of</strong><br />

other business cycle dating techniques.<br />

Relative to the trends through peaks or linear time trend approach, the HP<br />

filter approach has the advantage that it allows for structural change to occur at<br />

any point in the series (although the rate <strong>of</strong> change is constrained by the choice<br />

<strong>of</strong> λ). However, this approach still has a number <strong>of</strong> weaknesses.<br />

First, the choice <strong>of</strong> λ can have a large effect on the conclusion reached, as can<br />

the choice <strong>of</strong> sample period chosen. If an ‘average’ level <strong>of</strong> λ is chosen, any<br />

dramatic change in the actual series will tend to be attributed to both dem<strong>and</strong><br />

<strong>and</strong> supply disturbances. Even if this attribution is correct on average, in the case<br />

<strong>of</strong> pure positive dem<strong>and</strong> shocks the technique will tend to understate the<br />

amount <strong>of</strong> excess dem<strong>and</strong> in the economy, while understating the amount <strong>of</strong><br />

excess supply in the case <strong>of</strong> pure positive supply shocks. For example, during a<br />

period <strong>of</strong> disinflation when monetary conditions are being tightened, the HP<br />

2


88<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

filter will tend to understate the amount <strong>of</strong> excess supply in the economy.<br />

Secondly, as with the trend through peaks approach, the level <strong>of</strong> uncertainty<br />

in the estimates will be greater at the end <strong>of</strong> the sample, precisely where it<br />

matters the most. This is because, if the effects <strong>of</strong> dem<strong>and</strong> shocks are persistent,<br />

it will be difficult to distinguish between permanent <strong>and</strong> temporary shocks; <strong>and</strong><br />

at the end <strong>of</strong> the sample there is almost no information available about the effects<br />

<strong>of</strong> the latest shocks.<br />

Finally, the level <strong>of</strong> potential output produced by this simple detrending<br />

technique is clearly not consistent with the traditional equilibrium notion <strong>of</strong><br />

potential output. There is no guarantee that the sample period chosen is<br />

cyclically neutral, particularly when dealing with small samples. For these<br />

reasons, Laxton <strong>and</strong> Tetlow suggest that very large confidence intervals should<br />

be placed around potential output estimates that are derived solely through the<br />

use <strong>of</strong> univariate techniques. Despite these problems, the HP filter is a very<br />

popular tool for obtaining estimates <strong>of</strong> potential output.<br />

The multivariate filter approach<br />

Laxton <strong>and</strong> Tetlow (1992) have suggested the use <strong>of</strong> a multivariate filter to<br />

overcome some <strong>of</strong> the problems with the univariate HP procedure.<br />

This approach attempts to retain the computational simplicity <strong>of</strong> the Hodrick-<br />

Prescott filter approach, but supplements that approach with additional<br />

information based on established economic relationships. Therefore the approach<br />

can be thought <strong>of</strong> as quasi-structural in that it makes use <strong>of</strong> theoretical<br />

relationships <strong>and</strong>/or empirical regularities, without needing to impose<br />

restrictions based on imperfect representations <strong>of</strong> the true structure.<br />

The specific multivariate filter approach suggested by Laxton <strong>and</strong> Tetlow<br />

relies on the short-term output-inflation relationship (the Phillips curve) <strong>and</strong> the<br />

output-unemployment relationship (Okun’s Law) to attribute the proportion <strong>of</strong><br />

a given shock that can be deemed as originating from a supply disturbance.<br />

These relationships are specified as follows:<br />

( −1 , −1)<br />

π ,<br />

( −1 −1) ( −1 −1)<br />

π = π + BL ( ) Q − τ + ε<br />

t et t Q t t<br />

U − τ = C( L) U − τ + D( L)<br />

Q − τ + ε<br />

t U, t t U, t t Q, t U,<br />

t<br />

where π is the inflation rate, π e is the expected inflation rate, Q is the log <strong>of</strong><br />

output, U is unemployment, τ i,t<br />

is the trend value for variable I, <strong>and</strong> J(L) are<br />

polynomial lag operators. The generalised problem is, therefore, to minimise:<br />

T<br />

T<br />

T<br />

T−1<br />

Q MVF = ∑ ηε 2<br />

t Q t + ∑ θε 2<br />

t t + ∑ γ ε 2<br />

, π , t U , t + λ ∑ τ t − τ t − τ t − τ t<br />

t=<br />

1<br />

t=<br />

1<br />

t=<br />

1<br />

t=<br />

2<br />

[( + 1 ) ( −1)<br />

]<br />

subject to the Phillips curve <strong>and</strong> Okun’s Law equations. That is, estimates <strong>of</strong><br />

potential output are then chosen in a way that helps to improve the fit <strong>of</strong> the<br />

2


Darren Gibbs 89<br />

inflation rate equation <strong>and</strong> the unemployment rate equation. If there is useful<br />

explanatory power in these equations, then the data on inflation <strong>and</strong> unemployment<br />

will help in the identification <strong>of</strong> potential output. Of course, the procedure<br />

can be developed to take account <strong>of</strong> other known relationships such as changes<br />

in the yield curve, the import penetration rate, real unit labour costs, or any<br />

other relationship that may <strong>of</strong>fer insights which help to separate out the trend<br />

from the cycle.<br />

Laxton <strong>and</strong> Tetlow further generalise the HP optimisation by allowing for<br />

time varying weights on the various terms in the optimisation process (these are<br />

given by {η,θ,γ}). This means that users have the option <strong>of</strong> exercising judgment<br />

to increase or decrease the weight attached to a particular piece <strong>of</strong> information<br />

at a particular time. For example, one may wish to attach very little weight to<br />

the inflation equation during a period covered by wage <strong>and</strong> price controls, or<br />

attach a very high weight on actual output when there is good reason to believe<br />

that actual <strong>and</strong> potential output are close to coinciding. (In this way, benchmarks<br />

are established for the potential output series.) There are a number <strong>of</strong> ways <strong>of</strong><br />

choosing the relative weights. Laxton <strong>and</strong> Tetlow suggest that a sensible<br />

approach is to base them on the relative uncertainty in the relationships.<br />

Laxton <strong>and</strong> Tetlow conducted Monte Carlo tests under various assumptions<br />

about the relative importance <strong>of</strong> dem<strong>and</strong> <strong>and</strong> supply shocks to compare the<br />

performances <strong>of</strong> the HP <strong>and</strong> multivariate filters. Their results indicated that the<br />

multivariate filter improves on the estimates <strong>of</strong> the HP filter regardless <strong>of</strong> the<br />

proportion <strong>of</strong> output stemming from supply disturbances, but the extent <strong>of</strong> the<br />

improvement varies negatively with the proportion <strong>of</strong> shocks coming from<br />

supply sources. The HP filter does best when supply shocks are prevalent; but<br />

since the multivariate filter uses information on inflation, it improves the<br />

estimates most when supply shocks are less dominant. However, their results<br />

also suggest that although the multivariate filter performs better than the HP<br />

filter, there is still considerable uncertainty in the estimates.<br />

3.3 Summary<br />

As we have seen, there are number <strong>of</strong> empirical methods that have been used to<br />

calculate potential output, <strong>and</strong> the policy conclusions reached can be sensitive<br />

to the methods chosen. Simple time trend techniques appear to be too inflexible<br />

to generate plausible estimates while, in practice, structural techniques are too<br />

rigid <strong>and</strong> rely on unreliable data <strong>and</strong> unjustifiable assumptions. The multivariate<br />

filter approach, which represents something <strong>of</strong> a middle ground, has several<br />

desirable characteristics which attempt to overcome the problems. Nevertheless,<br />

none <strong>of</strong> the techniques surveyed is free from serious deficiencies. Therefore, the<br />

relative desirability <strong>of</strong> each approach is ultimately an empirical question. This<br />

also suggests that policymakers should not place too great a reliance on one<br />

single measure.


90<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

These conclusions are shared by Laxton <strong>and</strong> Tetlow. They note that regardless<br />

<strong>of</strong> the method used estimates <strong>of</strong> potential output should be interpreted with<br />

caution, since the confidence b<strong>and</strong>s around such estimates are quite wide. This<br />

is seen as especially important when these estimates are used to guide<br />

macroeconomic policy decisions. This view is emphasised by the results <strong>of</strong> the<br />

preliminary empirical work discussed in the remainder <strong>of</strong> this paper.<br />

4 Empirical estimates <strong>of</strong> potential output for New Zeal<strong>and</strong><br />

In this section we apply a selection <strong>of</strong> the techniques discussed in section 3 to<br />

estimate preliminary potential output series for New Zeal<strong>and</strong>. The section<br />

describes how each series was constructed <strong>and</strong> uses graphical techniques to first<br />

compare the results <strong>of</strong> those techniques which are similar in nature, <strong>and</strong> then<br />

compare the results obtained across completely different techniques. A statistical<br />

comparison <strong>of</strong> the derived output gaps is reserved for section 5.<br />

4.1 Data<br />

A variety <strong>of</strong> data has been used in constructing the potential output series<br />

described in this section. The measure <strong>of</strong> actual output used was the seasonally<br />

adjusted quarterly production-based real GDP statistics produced by Statistics<br />

New Zeal<strong>and</strong> (QGDP). These data are <strong>of</strong>ficially available for the period 1977Q2<br />

to 1994Q1, though the National Bank has backdated this data to 1965Q4. For<br />

those techniques based solely on QGDP (ie the univariate techniques) a potential<br />

output series covering the whole <strong>of</strong> the period 1965Q4 to 1994Q1 was calculated.<br />

As noted above, the multivariate estimates <strong>of</strong> potential output require a<br />

variety <strong>of</strong> additional data series as inputs into the estimation process. The<br />

construction <strong>of</strong> these additional series is discussed as they arise. The start dates<br />

for the potential output series derived from these approaches were dependent<br />

on the availability <strong>of</strong> these additional series. The estimated multivariate potential<br />

output series typically start in the early 1970s.<br />

For the purposes <strong>of</strong> the discussion in this section <strong>and</strong> the statistical analysis<br />

that follows we focus on the single period 1977Q2–1994Q1—the period during<br />

which <strong>of</strong>ficial data are available on movements in actual output. In the interest<br />

<strong>of</strong> brevity we concentrate mainly on reporting the final output from these<br />

techniques (ie the estimated potential output series <strong>and</strong> the associated output<br />

gaps), <strong>and</strong> not on the intermediate estimation results from which they were<br />

constructed.<br />

4.2 Simple univariate estimates<br />

Given the dubious theoretical quality <strong>of</strong> these measures, the only technique<br />

utilised under this heading is the trend through peaks approach. As noted earlier,<br />

a particular difficulty with this approach is establishing what constitutes a<br />

significant peak; <strong>and</strong> how to project out the potential output line past the last


Darren Gibbs 91<br />

significant peak. We use two variants to derive two alternative series. In the first,<br />

the potential output line is extrapolated past the last peak (the December quarter<br />

1986 GST induced boom) at the same slope as that connecting the previous two<br />

peaks (ie the st<strong>and</strong>ard approach). In the second, we extrapolate on the basis <strong>of</strong><br />

the average slope during the period 1965Q4–1982Q2.<br />

Figure 1 illustrates the estimated series potential output series (tp1 <strong>and</strong> tp2)<br />

<strong>and</strong> the corresponding output gaps. Note that these gaps are by definition<br />

always negative.<br />

While both these series indicate the development <strong>of</strong> a substantial output gap<br />

over the early 1990s, only tp2 suggests that this gap closed significantly by<br />

March 1994 (to less than 2 percent <strong>of</strong> GDP).<br />

4.3 Structural multivariate estimates<br />

Multivariate estimates <strong>of</strong> potential output (<strong>and</strong> thus the output gap) were<br />

constructed using three <strong>of</strong> the techniques described in section 3. The techniques<br />

used were the labour input related Okun approach, the capital input related<br />

output-capital ratio approach, <strong>and</strong> the combined input related general<br />

production function approach. These are discussed in turn.<br />

FIGURE 1: Estimated potential output <strong>and</strong> output gap series—trends through<br />

peaks approach<br />

20<br />

<strong>Output</strong> gap (%)<br />

15<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

-20<br />

10000<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

<strong>Output</strong> ($m 82/83)<br />

June years<br />

tp1: output gap (left axis)<br />

gdp (right axis)<br />

tp2: potential output (right axis)<br />

tp2: output gap (left axis)<br />

tp1: potential output (right axis)


92<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

<strong>Labour</strong> input related approaches<br />

Two potential output series were estimated using solely labour market data. The<br />

first approach relied on relating the size <strong>of</strong> the output gap to the deviation <strong>of</strong><br />

the actual employment rate (one minus the actual unemployment rate) from the<br />

estimated ‘natural rate’ <strong>of</strong> employment (one minus the natural rate <strong>of</strong><br />

unemployment), where the natural rate <strong>of</strong> unemployment is that rate at which<br />

the labour market is in equilibrium <strong>and</strong> there is no pressure for nominal wages<br />

to rise. The natural rate <strong>of</strong> unemployment is usually taken to represent that part<br />

<strong>of</strong> unemployment not explained by purely cyclical factors (in other words, the<br />

part <strong>of</strong> unemployment due to frictional, structural, <strong>and</strong> classical unemployment).<br />

This technique relies on Okun’s assumption that the influence <strong>of</strong> slack economic<br />

activity on average hours, labour force participation <strong>and</strong> productivity are all<br />

proxied by the unemployment rate.<br />

Obtaining an accurate measure <strong>of</strong> the unobserved natural rate <strong>of</strong> unemployment<br />

is no easier than estimating the unknown rate <strong>of</strong> growth in potential<br />

output. Two approaches have been commonly applied in the literature. The first<br />

involves directly estimating a reduced form model for the unemployment rate,<br />

including both ‘natural’ <strong>and</strong> cyclical factors as determinants. For example,<br />

studies have related the actual unemployment rate to actual output; the real<br />

exchange rate; real consumer wages; the marginal tax on wages; the unionisation<br />

FIGURE 2: Estimated natural unemployment rate<br />

12<br />

10<br />

Unemployment rate (%)<br />

8<br />

6<br />

4<br />

2<br />

0<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

June years<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

Estimated natural rate<br />

Actual unemployment rate


Darren Gibbs 93<br />

FIGURE 3: Estimated potential output <strong>and</strong> output gap series—labour input<br />

methods<br />

20<br />

15<br />

10000<br />

<strong>Output</strong> gap (%)<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

-20<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

<strong>Output</strong> ($m 82/83)<br />

June years<br />

yl1: output gap (left axis)<br />

gdp (right axis)<br />

yl1: output gap (left axis)<br />

yl2: output gap (left axis)<br />

yl1: potential output (right axis)<br />

rate; the real unemployment benefit; the minimum wage rate; <strong>and</strong> the volume <strong>of</strong><br />

world trade. The natural rate <strong>of</strong> unemployment is then obtained by solving out<br />

the equation with the cyclical factors eliminated.<br />

An alternative approach involves the estimation <strong>of</strong> a Phillips curve or twoequation<br />

wage-price model. This model is subsequently solved for an<br />

unemployment rate that is consistent with stable inflation.<br />

It was decided to adopt the Phillips curve approach in this paper. However,<br />

the estimated natural unemployment rate series also displayed unreasonable<br />

volatility. This was removed by passing the series through the Hodrick-Prescott<br />

filter, producing the natural unemployment rate series depicted in Figure 2.<br />

Clearly this methodology is deficient. However, its use was necessitated by<br />

the absence <strong>of</strong> better estimates <strong>of</strong> the natural rate <strong>of</strong> unemployment (which is a<br />

major task in its own right). Although the true natural rate series is likely to be<br />

less smooth, the estimated natural rate <strong>of</strong> 7.75 percent as <strong>of</strong> March 1994 seems<br />

plausible; however recent falls in unemployment (to 6.3 percent in June 1995)<br />

combined with only moderate wage growth suggest that it was likely to have<br />

been lower than this by mid 1995.<br />

The second labour input related potential output series estimate is based<br />

directly on Okun’s model relating the levels <strong>of</strong> output <strong>and</strong> unemployment as


94<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

FIGURE 4: Estimated potential output <strong>and</strong> output gap series—capital input<br />

methods<br />

20<br />

15<br />

10000<br />

<strong>Output</strong> gap (%)<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

8000<br />

6000<br />

4000<br />

2000<br />

<strong>Output</strong> ($m 82/83)<br />

-20<br />

0<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

June years<br />

yk1: output gap (left axis)<br />

gdp (right axis)<br />

yk2: potential output (right axis)<br />

yk2: output gap (left axis)<br />

yk1: potential output (right axis)<br />

discussed in section 3. The technique involves regressing the employment rate<br />

on output per working age person, <strong>and</strong> a time trend. This yields an estimate <strong>of</strong><br />

the potential per capita growth rate which, when combined with the rate <strong>of</strong><br />

growth in the working age population, results in an estimate <strong>of</strong> the total potential<br />

growth rate. Using this approach, the estimated output elasticity <strong>of</strong> the<br />

employment rate was 0.35 (which compares well with Okun’s original estimate<br />

<strong>of</strong> 0.35–0.40), while the estimated potential per capita growth rate was around 2<br />

percent per annum.<br />

Having obtained this estimate (which is expressed as a growth rate) the level<br />

<strong>of</strong> potential output is then calculated by choosing one time period at which the<br />

level <strong>of</strong> actual output <strong>and</strong> potential output are assumed to coincide (ie a<br />

benchmark). In this case, March 1985 was chosen as the benchmark (almost all<br />

<strong>of</strong> the more sophisticated estimates which are discussed later in this section<br />

support this choice).<br />

Figure 3 illustrates the estimated series potential output series derived from<br />

applying the above techniques (yl1 <strong>and</strong> yl2), <strong>and</strong> the corresponding output gaps.<br />

While both techniques produce output gaps which display a similar trend, the<br />

variance <strong>of</strong> yl2 is considerably greater than that associated with yl1. Both


Darren Gibbs 95<br />

estimates suggest that a negative output gap existed in March 1992, with the<br />

more intuitively plausible yl1 measure suggesting that a negative output gap <strong>of</strong><br />

around 1.5 percent <strong>of</strong> GDP remained in the 1994 March quarter.<br />

Capital input related approaches<br />

Two techniques <strong>of</strong> relating solely capital inputs to output were applied. The first<br />

approach, similar in nature to that used to derive yl1 above, was to relate<br />

deviations in actual from potential output to deviations in capacity utilisation<br />

from historical norms. The capacity utilisation series used was that collected in<br />

the Quarterly Survey <strong>of</strong> Business Opinion (produced by the New Zeal<strong>and</strong><br />

Institute <strong>of</strong> Economic Research).<br />

The second approach used was the output/capital ratio approach discussed<br />

in section 4 which relies on the assumption <strong>of</strong> a stable proportional relationship<br />

between the stock <strong>of</strong> capital <strong>and</strong> potential output. Annual capital stock estimates<br />

for the non-government sector (excluding dwellings) produced by the Research<br />

Project on Economic Planning (Philpott, 1994) were interpolated to produce<br />

quarterly capital stock estimates through to March 1993; <strong>and</strong> capital stock data<br />

for the remaining year <strong>of</strong> the sample were estimated from SNA data with the<br />

rate <strong>of</strong> depreciation set at 1.8% per quarter. 13<br />

Figure 4 illustrates the estimated potential output series derived from applying<br />

the above techniques (yk1 <strong>and</strong> yk2), <strong>and</strong> the corresponding output gaps.<br />

As the figures illustrate, the two estimates are highly correlated. The yk1<br />

measure suggests that the economy has been operating at above potential since<br />

1993. Also note that yk2 is similar to the trend through peaks estimates in that<br />

the estimated output gap is by construction always negative. Both series suggest<br />

that the degree <strong>of</strong> spare capacity has declined sharply over the last two years.<br />

The production function approach<br />

The previous two approaches related potential output to either some measure <strong>of</strong><br />

labour input or some measure <strong>of</strong> capital input. By contrast, the production<br />

function approach relates potential output to both labour <strong>and</strong> capital inputs, <strong>and</strong><br />

by so doing should yield better estimates <strong>of</strong> potential output.<br />

Only one potential output series was estimated using this technique. The<br />

functional form chosen for the estimated production function was the constant<br />

elasticity <strong>of</strong> substitution (CES) function. The estimated function was <strong>of</strong> the<br />

following form:<br />

⎡<br />

LF E H<br />

Q= PWA• ⎛ ⎝ ⎜ ⎞<br />

⎟ • ⎛ PWA⎠<br />

⎝ ⎜ ⎞<br />

⎟ • ⎛ −ρ<br />

⎛<br />

LF⎠<br />

⎝ ⎜ ⎞⎞<br />

⎢α⎜<br />

⎟⎟ + − •<br />

⎝<br />

E ⎠⎠<br />

⎣⎢<br />

( 1 α)( QCU K)<br />

1<br />

ρ<br />

−ρ<br />

⎤<br />

⎥<br />

⎦⎥<br />

13<br />

This rate was chosen based on previous work at the Reserve Bank (See Brooks <strong>and</strong><br />

Gibbs, 1991).


96<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

where<br />

Q = actual output<br />

PWA = population <strong>of</strong> working age<br />

LF = labour force<br />

E = employment<br />

H = hours worked<br />

QCU = capacity utilisation<br />

K = capital stock<br />

The function implies that output is a function <strong>of</strong>: the population <strong>of</strong> working age;<br />

the labour force participation rate; the employment rate; hours per employee;<br />

<strong>and</strong> the utilised capital stock. All labour force data were taken from the<br />

Household <strong>Labour</strong> Force Survey, while the capacity utilisation <strong>and</strong> capital stock<br />

series used were the same as noted earlier in this section.<br />

A potential output series based on the above function was estimated. First,<br />

the production function was estimated so as to yield estimates <strong>of</strong> the parameters<br />

α <strong>and</strong> ρ. Secondly, these parameters were used to create the predicted values<br />

from the regression with the following adjustments: the actual series <strong>of</strong> labour<br />

FIGURE 5: Estimated potential output <strong>and</strong> output gap series—production<br />

function method<br />

20<br />

15<br />

10000<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

-20<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

<strong>Output</strong> gap (%)<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

<strong>Output</strong> ($m 82/83)<br />

June years<br />

pf1: output gap (left axis)<br />

gdp (right axis)<br />

pf1: potential output (right axis)


Darren Gibbs 97<br />

force participation rates was replaced with a series in which all observations<br />

were equal to the average rate <strong>of</strong> participation over the sample period; the actual<br />

series <strong>of</strong> employment rates was replaced with the estimated ‘natural’ employment<br />

rates derived earlier in this section; the actual series <strong>of</strong> hours worked per<br />

employee was replaced with a series in which all observations were equal to the<br />

average hours per employee over the sample period; <strong>and</strong> the actual series <strong>of</strong><br />

capacity utilisation was replaced by a series in which all observations were equal<br />

to the average rate <strong>of</strong> capacity utilisation over the sample period. 14 The influence<br />

<strong>of</strong> technology was proxied as is st<strong>and</strong>ard in these techniques by adding back the<br />

residuals to the above predicted values.<br />

Figure 5 illustrates the estimated potential output series <strong>and</strong> the estimated<br />

output gaps obtained by following the above approach.<br />

As Figure 5 illustrates, the output gap derived from the production function<br />

based estimate <strong>of</strong> potential output is similar to the others derived so far. The mid<br />

1980s is identified as a period in which output was above potential, while the<br />

early 1990s is identified as a period during which output was well below potential.<br />

<strong>Output</strong> is estimated to have moved above potential in late 1993/early 1994.<br />

4.4 Stochastic filtering approaches<br />

Four filtering approaches were applied following the methodology discussed in<br />

section 3. The first relies on the conventional Hodrick-Prescott filter, while the<br />

remaining three involve various forms <strong>of</strong> the multivariate filter which include<br />

progressively greater degrees <strong>of</strong> ‘outside’ information.<br />

Hodrick-Prescott filter<br />

The first approach tried was a straightforward application <strong>of</strong> the Hodrick-<br />

Prescott filter. As noted in section 3, the only parameter that must be chosen is<br />

the smoothness constraint (λ). Three variations <strong>of</strong> the constraint were applied<br />

yielding three different potential output series. The first variation involved<br />

setting λ to equal 1600—the value recommended by Hodrick <strong>and</strong> Prescott. The<br />

second variation involved setting λ to equal 200. This results in greater nonlinearity<br />

<strong>of</strong> the potential output series so that supply shocks are assumed to be<br />

more dominant than in the case when λ is equal to 1600. The third variation<br />

involved setting λ to equal 100,000. This results in greater linearity <strong>of</strong> the<br />

potential output series so that dem<strong>and</strong> shocks are assumed to be more dominant<br />

than in the case where λ is equal to 1600.<br />

Figure 6 illustrates the three estimated potential output series (hp1600, hp200,<br />

hp100,000) obtained by following the above approach, <strong>and</strong> the estimated output<br />

gaps.<br />

14<br />

As noted earlier, these assumptions require that the period covered is cyclically neutral.<br />

Moreover, these assumptions could be improved by modelling the trend <strong>of</strong> each <strong>of</strong> these<br />

series individually.


98<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

FIGURE 6: Estimated potential output <strong>and</strong> output gap series—<br />

Hodrick-Prescott method<br />

20<br />

<strong>Output</strong> gap (%)<br />

15<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

-20<br />

10000<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

<strong>Output</strong> ($m 82/83)<br />

June years<br />

hp200: output gap (left axis)<br />

hp100,000: output gap (left axis)<br />

hp200: potential output (right axis)<br />

hp100,000: potential output (right axis)<br />

hp1600: output gap (left axis)<br />

gdp: (right axis)<br />

hp1600: potential output (right axis)<br />

The output gaps illustrated in Figure 6 once again display the same broad<br />

trends as those estimated earlier. The impact <strong>of</strong> variations in the smoothness<br />

constraint can be seen clearly. When l is equal to 200, the estimated potential<br />

output series is more closely correlated with the actual output series so that the<br />

absolute size <strong>of</strong> the estimated output gap tends to be smaller. By contrast, when<br />

l is equal to 100,000, the estimated potential output series is a near linear trend<br />

<strong>and</strong> the absolute size <strong>of</strong> the estimated output gaps tends to be larger. All series<br />

suggest that actual output rose above potential output during 1993, but this is<br />

less pronounced in the case <strong>of</strong> the hp200 filter.<br />

Multivariate filter approach—the bivariate filter<br />

We now supplement the information generated by the pure Hodrick-Prescott<br />

filter with information derived from the evolution <strong>of</strong> inflation using the<br />

multivariate filter approach discussed in section 3.<br />

The procedure was applied as follows. First, an initial ‘Phillips curve’ was<br />

estimated whereby today’s inflation is described as a function <strong>of</strong> expected<br />

inflation (this was estimated assuming adaptive expectations using a fourth<br />

order autoregressive model) <strong>and</strong> the history <strong>of</strong> output gaps (the Hodrick-Prescott<br />

output gap derived above was used in the first iteration). The information


Darren Gibbs 99<br />

FIGURE 7: Estimated potential output <strong>and</strong> output gap series—bivariate filter<br />

method<br />

20<br />

15<br />

10000<br />

<strong>Output</strong> gap (%)<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

-20<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

<strong>Output</strong> ($m 82/83)<br />

June years<br />

bf200: output gap (left axis)<br />

bf100,000: output gap (left axis)<br />

bf200: potential output (right axis)<br />

bf100,000: potential output (right axis)<br />

bf1600: output gap (left axis)<br />

gdp (right axis)<br />

bf1600: potential output (right axis)<br />

contained in this regression was than embedded within the Hodrick-Prescott<br />

minimisation formula to yield a bivariate estimate <strong>of</strong> potential output which was<br />

based on both the observed behaviour <strong>of</strong> output <strong>and</strong> the observed behaviour <strong>of</strong><br />

inflation. Next, the Phillips curve was re-estimated with the initial Hodrick-<br />

Prescott derived output gap being replaced by the new estimated bivariate<br />

output gap. The bivariate filter was then re-estimated using the information from<br />

this updated regression. This iterative procedure was repeated until the<br />

coefficients in the re-estimated Phillips curve converged.<br />

As noted in section 3, the procedure can be generalised to allow independently<br />

determined weights to be placed on the various terms in the minimisation<br />

problem. This facility was used so as to effectively give zero weight to developments<br />

in inflation during the 1982–84 wage <strong>and</strong> price freeze, <strong>and</strong> to the 1986 <strong>and</strong><br />

1989 GST induced increases in indirect tax.<br />

As with the potential output estimates derived using the pure Hodrick-<br />

Prescott approach three variations <strong>of</strong> the smoothness parameter were used, so<br />

that three different bivariate filters were derived. Figure 7 illustrates the three<br />

estimated potential output series (bf1600, bf200, bf100,000) obtained by following<br />

the above approach, <strong>and</strong> the estimated output gaps.


100<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

FIGURE 8: Comparison <strong>of</strong> the HP1600 <strong>and</strong> BF1600 filters<br />

4<br />

6<br />

2<br />

5<br />

<strong>Output</strong> gap (%)<br />

0<br />

-2<br />

-4<br />

-6<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Inflation rate (%)<br />

-8<br />

-1<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

June years<br />

bf1600: output gap (left axis)<br />

hp1600: output gap (left axis)<br />

Inflation rate (right axis)<br />

The general pattern <strong>of</strong> the bivariate output gaps is similar to the Hodrick-<br />

Prescott derived estimates. However, unlike the Hodrick-Prescott estimates, the<br />

bivariate estimates suggest that the output gap remained negative (ie output<br />

below potential) at the end <strong>of</strong> the sample period. The influence <strong>of</strong> adding the<br />

inflation process to the filtering problem is best illustrated in Figure 8 which<br />

graphs the hp1600 <strong>and</strong> bf1600 filters against the quarterly inflation rate.<br />

Compared with the hp1600 filter, the bf1600 filter tends to result in a more<br />

positive (or less negative) output gap when inflation is rising <strong>and</strong> a more<br />

negative (or less positive) output gap when inflation is falling. In other words,<br />

in the latter case declining inflation is interpreted as additional evidence that<br />

actual output is below potential, with the result that the estimated level <strong>of</strong><br />

potential output is higher <strong>and</strong> thus the estimated negative output gap is larger.<br />

Multivariate filter approach: Adding information regarding unemployment<br />

The multivariate filter is simply a further generalisation <strong>of</strong> the bivariate filter in<br />

which we include, in this case, additional information on the evolution <strong>of</strong><br />

unemployment. To be specific, we make use <strong>of</strong> an estimated Okun relationship<br />

to map ‘unemployment gaps’ into output gaps. The unemployment gap was<br />

calculated using the natural rate <strong>of</strong> unemployment series derived earlier in this<br />

section. As with the bivariate filter, the initial output gap used in the Okun


Darren Gibbs 101<br />

FIGURE 9: Estimated potential output <strong>and</strong> output gap series—multivariate<br />

filter method<br />

20<br />

15<br />

10000<br />

<strong>Output</strong> gap (%)<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

-20<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

<strong>Output</strong> ($m 82/83)<br />

1992<br />

1993<br />

June years<br />

mfa200: output gap (left axis)<br />

mfa100,000: output gap (left axis)<br />

mfa200: potential output (right axis)<br />

mfa100,000: potential output (right axis)<br />

mfa1600: output gap (left axis)<br />

gdp (right axis)<br />

mfa1600: potential output (right axis)<br />

equation was that derived using the Hodrick-Prescott method. The information<br />

contained in the Okun regression was than embedded within the Hodrick-<br />

Prescott minimisation formula (along with the Phillips curve regression) to yield<br />

a bivariate estimate <strong>of</strong> potential output which is based on both the observed<br />

behaviour <strong>of</strong> output <strong>and</strong> the observed behaviour <strong>of</strong> inflation <strong>and</strong> unemployment.<br />

As with the bivariate filter, an iteractive technique was used to generate<br />

successive replacement output gaps in the Phillips curve <strong>and</strong> Okun equation,<br />

until the estimated coefficients converged.<br />

As with the potential output estimates derived using the previous filtering<br />

techniques three variations <strong>of</strong> the smoothness parameter were used, so that three<br />

different multivariate filters were derived. Figure 9 illustrates the three estimated<br />

potential output series (mfa1600, mfa200, mfa100,000) obtained by following the<br />

above approach, <strong>and</strong> the estimated output gaps.<br />

The general pattern <strong>of</strong> the multivariate output gaps is similar to the bivariate<br />

filter estimates. However, the size <strong>of</strong> the negative output gap is more pronounced<br />

in the late 1980s <strong>and</strong> early 1990s than that estimate derived from the<br />

bivariate filter. Note also that the mfa200 measure suggests that potential output<br />

fell in absolute terms in 1991. This is consistent with accelerated capital<br />

scrapping <strong>and</strong> a rising natural unemployment rate.


102<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

FIGURE 10: Comparison <strong>of</strong> the BF1600 <strong>and</strong> MFA1600 filters<br />

4<br />

2<br />

<strong>Output</strong> gap (%)<br />

0<br />

-2<br />

-4<br />

-6<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

June years<br />

bf1600: output gap (left axis)<br />

mfa1600: output gap (left axis)<br />

unemployment gap (left axis)<br />

The influence <strong>of</strong> adding the information regarding unemployment to the<br />

filtering problem is best illustrated in Figure 10 which graphs the bf1600 <strong>and</strong><br />

mfa1600 filters against the estimated unemployment gap (a negative unemployment<br />

gap implies that actual unemployment is above the estimated natural rate).<br />

Compared with the bf1600 filter, the mfa1600 filter results in a more negative (or<br />

less positive) output gap when the actual unemployment rate is assessed as<br />

above the natural rate. This is because the presence <strong>of</strong> a negative unemployment<br />

gap is interpreted as additional evidence that actual output is below potential,<br />

with the result that the estimated level <strong>of</strong> potential output is higher <strong>and</strong> thus the<br />

estimated negative output gap is larger.<br />

The multivariate filter: Adding information regarding the yield curve<br />

The final estimates that we shall consider involve a further extension <strong>of</strong> the<br />

above approach so as to include additional supplementary information on the<br />

evolution <strong>of</strong> the yield curve. Specifically, we seek to relate the yield gap between<br />

90-day bank bills <strong>and</strong> five-year government bonds (expressed as a deviation<br />

from a moving mean) to the output gap. The estimated ‘IS curve’ implies that<br />

high short-term rates relative to long-term rates (a negatively sloped yield curve)<br />

are associated with a negative output gap. This may reflect the fact that a


Darren Gibbs 103<br />

FIGURE 11: Estimated potential output <strong>and</strong> output gap series—multivariate<br />

filter method<br />

20<br />

15<br />

10000<br />

<strong>Output</strong> gap (%)<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-15<br />

-20<br />

8000<br />

6000<br />

4000<br />

2000<br />

0<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

<strong>Output</strong> ($m 82/83)<br />

1992<br />

1993<br />

June years<br />

mfb200: output gap (left axis)<br />

mfb100,000: output gap (left axis)<br />

mfb200: potential output (right axis)<br />

mfb100,000: potential output (right axis)<br />

mfb1600: output gap (left axis)<br />

gdp (right axis)<br />

mfb1600: potential output (right axis)<br />

tightening in monetary policy (resulting in a positive yield gap: short rates<br />

exceed long rates) tends to temporarily depress economic growth during the<br />

transition to low inflation; or the fact that a positive yield gap may cause some<br />

people to expect interest rates to fall in the near future with the result that some<br />

investment projects are delayed (see Ebert, 1994).<br />

The information contained in the IS regression was than embedded within the<br />

Hodrick-Prescott minimisation formula (along with the Phillips curve <strong>and</strong> Okun<br />

regression) to yield a multivariate estimate <strong>of</strong> potential output which is based on<br />

both the observed behaviour <strong>of</strong> output <strong>and</strong> the observed behaviour <strong>of</strong> inflation,<br />

unemployment <strong>and</strong> the yield curve. Again, an iterative technique was used to<br />

generate successive replacement output gaps in the IS regression until the<br />

estimated coefficients converged, <strong>and</strong> three variations <strong>of</strong> the smoothness parameter<br />

were used so that three different multivariate filters were derived. Figure<br />

11 illustrates the three estimated potential output series (mfb1600, mfb200,<br />

mfb100,000) obtained by following the above approach, <strong>and</strong> the estimated output<br />

gaps.<br />

The general pattern <strong>of</strong> the multivariate output gaps is similar to the previous<br />

estimates. However, the size <strong>of</strong> the negative output gap is less pronounced in the


104<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

FIGURE 12: Comparison <strong>of</strong> the MFA1600 <strong>and</strong> MFB1600 filters<br />

4<br />

6<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1977<br />

1978<br />

1979<br />

1980<br />

1981<br />

1982<br />

1983<br />

1984<br />

1985<br />

1986<br />

1987<br />

1988<br />

1989<br />

1990<br />

1991<br />

1992<br />

1993<br />

<strong>Output</strong> gap (%)<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

-1<br />

-2<br />

Yield gap (%)<br />

June years<br />

mfa1600: output gap (left axis)<br />

mfb1600: output gap (left axis)<br />

yield gap (right axis)<br />

late 1980s <strong>and</strong> early 1990s than the estimates using the previous multivariate<br />

filter. The influence <strong>of</strong> adding the information regarding the yield gap to the<br />

filtering problem is illustrated in Figure 12 which graphs the mfa1600 <strong>and</strong><br />

mfb1600 filters against the yield gap.<br />

Compared with the mfa1600 filter, the mfb1600 filter tends to result in a less<br />

positive output gap the more positive the yield gap. This is because the presence<br />

<strong>of</strong> a positive yield gap is interpreted as additional evidence that actual output is<br />

below potential, with the result that the estimated level <strong>of</strong> potential output is<br />

higher <strong>and</strong> thus the negative output gap is larger.<br />

<strong>Potential</strong> growth rates implied by each measure<br />

Table 1 presents the estimated potential growth rate over the year to March 1994<br />

(the actual growth rate was 5.4 percent in the numbers used in this study though<br />

this has since been revised up to 6.0 percent) <strong>and</strong> the estimated output gap (as a<br />

proportion <strong>of</strong> GDP) as at March 1994 for each <strong>of</strong> the 19 potential output series<br />

generated in this section <strong>of</strong> the paper. 15<br />

15<br />

The size <strong>of</strong> the output gap is only reported for those measures that are broadly<br />

consistent with the ‘natural rate’ as opposed to the maximum physical capacity definition<br />

(eg output gaps for the trends through peaks approach can only be negative so the sign on<br />

the gap term is <strong>of</strong> no significance).


Darren Gibbs 105<br />

TABLE 1: Estimated potential growth rates <strong>and</strong> magnitude <strong>of</strong> output gap<br />

Method used Growth rate <strong>Output</strong> gap<br />

tp1 2.21<br />

tp2 1.06<br />

yl1 4.05 -1.52<br />

yl2 3.54<br />

yk1 3.16 3.43<br />

yk2 1.52<br />

pf1 1.39 1.34<br />

hp1600 1.73 3.41<br />

hp200 3.55 1.52<br />

hp100,000 1.10 3.53<br />

bf1600 3.55 -0.25<br />

bf200 5.42 -1.10<br />

bf100,000 1.50 -0.20<br />

mfa1600 2.79 -0.90<br />

mfa200 4.19 -1.95<br />

mfa100,000 1.62 -1.83<br />

mfb1600 3.76 -0.76<br />

mfb200 4.94 -1.75<br />

mfb100,000 1.66 -0.35<br />

As Table 1 illustrates the various measures produce vastly different<br />

conclusions regarding the potential growth rate over the year to March 1994, <strong>and</strong><br />

the size <strong>and</strong> sign <strong>of</strong> the output gap. If, for example, mfa200 is the correct estimate<br />

<strong>of</strong> potential output then, assuming the potential growth rate over the year to<br />

March 1995 is the same as in 1994, this would imply that the economy could<br />

sustain a further 6.1 percent before the output gap termed positive <strong>and</strong> pressures<br />

on inflation developed. 16 On the other h<strong>and</strong>, if pf1 is the correct estimate <strong>of</strong><br />

potential output, then the estimates suggest that inflationary pressures are likely<br />

to have been developing since the beginning <strong>of</strong> last year. The question is, how<br />

does one select which (if any) is the best estimate?<br />

5 Econometric properties <strong>of</strong> estimated output gaps<br />

In section 4 we estimated 19 different output gaps. Given that the different<br />

estimates led to different conclusions, the usefulness <strong>of</strong> these measures will be<br />

limited; unless one estimate can be isolated as being superior to the rest, or the<br />

range <strong>of</strong> valid alternatives can at least be narrowed down. In this section we<br />

make use <strong>of</strong> both correlation <strong>and</strong> regression analysis in an attempt to do<br />

just that.<br />

16<br />

The estimate <strong>of</strong> 6.1 percent is calculated by summing last period’s output gap <strong>and</strong> the<br />

negative <strong>of</strong> the estimated potential growth rate.


106<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

TABLE 2: Correlations between estimated output gaps <strong>and</strong> the change <strong>and</strong><br />

level <strong>of</strong> inflation<br />

Change in<br />

Gap <strong>and</strong><br />

gap <strong>and</strong><br />

change in<br />

change in<br />

Change in import import<br />

Method inflation Inflation penetration penetration<br />

used 77Q2–94Q1 77Q2–94Q1 77Q2–94Q1 77Q2–94Q1<br />

tp1 0.1064 0.5815 -0.0013 0.2144<br />

tp2 0.1577 0.1230 0.0740 0.2272<br />

yl1 -0.0451 0.7760 -0.0934 0.1426<br />

yl2 0.0380 0.8218 -0.0545 0.2302<br />

yk1 0.2735 0.2811 0.0806 0.2046<br />

yk2 0.1696 0.3681 0.0617 0.2102<br />

pf1 0.1768 0.4695 0.0612 0.2126<br />

hp1600 0.2415 0.0272 0.1406 0.2231<br />

hp200 0.3372 0.1461 0.1932 0.2225<br />

hp100,000 0.1439 0.1655 0.1052 0.2078<br />

bf1600 0.3643 0.4530 0.1639 0.2252<br />

bf200 0.3930 0.3891 0.1650 0.1744<br />

bf100,000 0.1645 0.2100 0.0790 0.2108<br />

mfa1600 0.2910 0.6542 0.0986 0.2236<br />

mfa200 0.2624 0.6744 0.0643 0.1933<br />

mfa100,000 0.1614 0.3821 0.0590 0.2121<br />

mfb1600 0.3223 0.5442 0.1377 0.2179<br />

mfb200 0.2712 0.6226 0.0891 0.1783<br />

mfb100,000 0.1595 0.2484 0.0701 0.2099<br />

5.1 Correlation analysis<br />

From the st<strong>and</strong>point <strong>of</strong> a central bank that is required to maintain price stability,<br />

the usefulness <strong>of</strong> the estimated output gap series depends on the strength <strong>and</strong><br />

reliability <strong>of</strong> its link with inflation. Keeping in mind the theory that the output<br />

gap should be related to the rate <strong>of</strong> acceleration in inflation (so that positive<br />

output gaps are associated with an increasing inflation rate), one way <strong>of</strong><br />

assessing this linkage is to analyse the simple correlation between the output gap<br />

<strong>and</strong> the change in inflation. Given that previous Reserve Bank research has found<br />

stronger correlations between the output gap (or some measure <strong>of</strong> excess<br />

dem<strong>and</strong>) <strong>and</strong> the level <strong>of</strong> inflation, these correlations were also computed. The<br />

results are presented in the first two columns <strong>of</strong> Table 2, which lead to the<br />

following conclusions. First, in all but one case, the estimated correlation coefficient<br />

is positive. This is to be expected, <strong>and</strong> implies that more positive values<br />

<strong>of</strong> the output gap are associated with larger changes in the rate <strong>of</strong> inflation or<br />

higher levels <strong>of</strong> inflation itself.


Darren Gibbs 107<br />

Secondly, focusing on the relationship between the output gaps <strong>and</strong> the<br />

change in inflation, the strongest correlations were obtained with those output<br />

gaps derived using the filtering techniques. Again this is to be expected, given<br />

that information about the evolution <strong>of</strong> inflation was used in the construction <strong>of</strong><br />

the underlying output series; though the Hodrick-Prescott derived estimates<br />

(which did not use this information) produced similar correlations to the<br />

multivariate filters. The magnitude <strong>of</strong> these correlation coefficients is comparable<br />

to those estimated for the G7 economies by Torres <strong>and</strong> Martin (1990) which<br />

ranged from 0.27 (for Germany) to 0.53 (for the United States), with an average<br />

correlation coefficient <strong>of</strong> 0.41.<br />

Thirdly, focusing on the relationship between the output gaps <strong>and</strong> the level<br />

<strong>of</strong> inflation, the strongest correlations were obtained with the labour input<br />

derived output gaps, although the multivariate filter based estimates were not<br />

far behind. In almost all cases, the correlation coefficient between the estimated<br />

output gap <strong>and</strong> the level <strong>of</strong> inflation was substantially higher than that between<br />

the output gap <strong>and</strong> the change in inflation. The relative strength <strong>of</strong> these<br />

correlations was further emphasised when they were recalculated over a more<br />

recent data period (1986Q1–1994Q1).<br />

Finally, among the filtered output gaps, those estimates based on smaller<br />

values <strong>of</strong> λ—the smoothness constraint—typically out-performed those based on<br />

higher values, with λ = 200, tending to provide the best correlation among each<br />

type <strong>of</strong> filter. This result suggests that supply shocks have been more dominant<br />

than dem<strong>and</strong> shocks over the sample period, which is perhaps not surprising<br />

given the extent <strong>of</strong> financial <strong>and</strong> economic reform.<br />

Another feature that we might expect to see in a good estimate <strong>of</strong> the output<br />

gap is a positive correlation between the output gap <strong>and</strong> the change in the<br />

import penetration rate. This reflects the idea that conditions <strong>of</strong> excess dem<strong>and</strong><br />

will typically result in a greater proportion <strong>of</strong> domestic expenditure on imported<br />

goods (together with a running down <strong>of</strong> domestic stocks). Indeed, this type <strong>of</strong><br />

relationship could also have been added to the multivariate filter, in the same<br />

way that we added data on inflation, unemployment <strong>and</strong> the yield gap. As<br />

Torres <strong>and</strong> Martin found a stronger correlation between the change in the output<br />

gap <strong>and</strong> import penetration growth for the G7 economies, these correlation<br />

coefficients were also calculated. The last two columns <strong>of</strong> Table 2 present the<br />

estimated correlation coefficients which lead to the following conclusions. First,<br />

in almost all cases, the correlation is positive. This is as one would expect, <strong>and</strong><br />

implies that more positive values <strong>of</strong> the output gap are associated with increased<br />

import penetration.<br />

Secondly, focusing on the relationship between the level <strong>of</strong> the output gap <strong>and</strong><br />

the change in import penetration, the strongest correlations were obtained with<br />

those output gaps derived using the Hodrick-Prescott <strong>and</strong> bivariate filtering<br />

techniques, though in all cases the correlations were relatively weak. The


108<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

magnitude <strong>of</strong> these correlation coefficients is comparable with those estimated<br />

for the G7 economies by Torres <strong>and</strong> Martin which ranged from 0.02 (for Canada)<br />

to 0.33 (for France), with an average correlation coefficient <strong>of</strong> 0.15.<br />

Finally, focusing on the relationship between the change in the output gap <strong>and</strong><br />

the change in import penetration, the estimated correlations were in all cases<br />

higher than those noted above (as in Torres <strong>and</strong> Martin), <strong>and</strong> were <strong>of</strong> a very<br />

similar magnitude. However, in absolute terms, the strength <strong>of</strong> the correlation is<br />

still weak.<br />

5.2 Regression analysis<br />

Given the finding that the estimated output gaps are more highly correlated with<br />

the level, rather than the change, in inflation (particularly over the last decade),<br />

it was decided to pursue this relationship further using regression analysis.<br />

In order to compare the explanatory power <strong>of</strong> the estimated output gaps, the<br />

following approach was employed. First, a simple reduced form econometric<br />

equation for quarterly consumer price inflation was estimated, including data on<br />

unit labour costs, the nominal exchange rate, <strong>and</strong> world export <strong>and</strong> import<br />

TABLE 3: Regression results—the output gap <strong>and</strong> quarterly consumer<br />

price inflation (1977Q2–1994Q1)<br />

Coefficient<br />

Regression<br />

<strong>Output</strong> gap on gap adjusted<br />

used term t statistic R-squared<br />

tp1 0.0380 1.602 0.8682<br />

tp2 0.0650 1.853 0.8702<br />

yl1 0.1053 1.222 0.8656<br />

yl2 0.0207 1.186 0.8654<br />

yk1 0.0845 2.504 0.8764<br />

yk2 0.0780 2.236 0.8737<br />

pf1 0.0842 2.337 0.8747<br />

hp1600 0.1085 2.414 0.8755<br />

hp200 0.1725 2.826 0.8799<br />

hp100,000 0.0451 1.689 0.8689<br />

bf1600 0.1801 3.497 0.8877<br />

bf200 0.1849 3.353 0.8860<br />

bf100,000 0.0580 1.965 0.8712<br />

mfa1600 0.1591 3.527 0.8881<br />

mfa200 0.1485 3.633 0.8894<br />

mfa100,000 0.0622 2.061 0.8721<br />

mfb1600 0.1788 3.362 0.8861<br />

mfb200 0.1957 3.352 0.8860<br />

mfb100,000 0.0540 1.916 0.8708


Darren Gibbs 109<br />

prices. The equation was then re-estimated with the addition <strong>of</strong> one <strong>of</strong> the<br />

estimated output gaps. If the output gap is to be useful for explaining <strong>and</strong><br />

forecasting the inflation process then we would want the estimated coefficient<br />

on the gap term to be <strong>of</strong> the correct sign, statistically significant <strong>and</strong> <strong>of</strong> a<br />

plausible magnitude. 17<br />

The relationship between the estimated output gaps <strong>and</strong> inflation was<br />

assumed to be linear in nature (ie the Phillips curve is linear). Recent research<br />

by Laxton et al (1994) suggests that this assumption may not be valid. Their<br />

research using pooled data from the G7 countries indicated that the relationship<br />

may in fact be non-linear (so that larger output gaps have a proportionately<br />

larger impact on inflation) <strong>and</strong> quite possibly asymmetric (positive output gaps<br />

have a different impact on inflation compared with negative gaps). Assuming<br />

symmetric non-linearity (<strong>of</strong> the cubic form) tended to worsen the equations’ fit.<br />

A more extensive testing for other forms <strong>of</strong> non-linearity <strong>and</strong> asymmetry was not<br />

attempted in this paper. However, given the significant policy implications <strong>of</strong><br />

non-linearity (<strong>and</strong> especially asymmetry) in the output-inflation process, further<br />

research along these lines appears warranted.<br />

This procedure gives us 19 possible models, each using a different output gap.<br />

Table 3 presents the estimated coefficients on the output gap terms, their t<br />

statistics, <strong>and</strong> the regression adjusted R-squared for each <strong>of</strong> the 19 alternative<br />

models.<br />

An examination <strong>of</strong> Table 3 leads to the following conclusions. First, in all<br />

cases the estimated coefficient on the output gap is positive as expected, <strong>and</strong> this<br />

is statistically significant in all regressions except those including the labour<br />

input derived output gaps. This latter result is interesting since the simple<br />

correlation between these output gaps <strong>and</strong> the inflation rate was very strong.<br />

This suggests that the unit labour cost term in the regression is already capturing<br />

the same information as these output gap measures.<br />

Secondly, the estimated coefficients range from 0.0207 to 0.1957, though the<br />

coefficients <strong>of</strong> the most significant gaps range from 0.1485 to 0.1957. These<br />

coefficients imply that a positive output gap equal to 1 percent <strong>of</strong> potential GDP<br />

will add around 0.15–0.195 to the quarterly inflation rate or 0.60–0.78 over the<br />

full year to the annual inflation rate, while a positive output gap equal to 3<br />

percent <strong>of</strong> potential GDP will add 0.45–0.59 to the quarterly inflation rate or<br />

1.80–2.34 over a full year to the annual inflation rate. Bearing in mind the period<br />

over which these equations have been estimated, these figures seem reasonably<br />

plausible. Given the Reserve Bank’s target <strong>of</strong> 0–2 percent annual movements in<br />

17<br />

One possibility, not investigated here, is that inflation may respond to the rate at which<br />

actual output is growing relative to potential in addition to the size <strong>of</strong> the output gap itself.<br />

Thus, there may be a distinction between short-run potential output <strong>and</strong> long-run potential<br />

output. Both have been found to be important by the Bank <strong>of</strong> Engl<strong>and</strong> (1994).


110<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

the Consumers Price Index these estimates imply that the Bank would, in the<br />

absence <strong>of</strong> inflationary pressures from other sources, need to ensure that actual<br />

output remained within around 2 percent either side <strong>of</strong> the estimated potential<br />

level <strong>of</strong> output. 18<br />

Finally, as in the correlation analysis, those estimates based on smaller values<br />

<strong>of</strong> λ—the smoothness constraint—typically outperformed those based on higher<br />

values, with λ = 200, tending to provide the best explanatory power among each<br />

type <strong>of</strong> filter. It is interesting to note that the choice <strong>of</strong> λ = 200 is the preferred<br />

parameter for Canada—another ‘small’ open economy. Again this result suggests<br />

that supply shocks have been more dominant than dem<strong>and</strong> shocks over the<br />

sample period.<br />

The task now is to assess whether any one model or group <strong>of</strong> models<br />

performs statistically better than the rest. This was done by applying the nonnested<br />

model selection tests developed by Davidson <strong>and</strong> Mackinnon (1981) <strong>and</strong><br />

Cox (1961 <strong>and</strong> 1962), known as the J test <strong>and</strong> Cox test respectively. The basic<br />

hypothesis that is tested is that Model A is the true model <strong>and</strong> that Model B is<br />

not. The testing procedure requires that the reverse hypothesis also be tested, as<br />

in small samples the procedure can generate conflicting or inconclusive results<br />

(ie both models are accepted or both models are rejected). Table 4 presents the<br />

results <strong>of</strong> the J test while Table 5 presents the results <strong>of</strong> the Cox test. In both<br />

tables, the statistics are presented in the form <strong>of</strong> p (or probability) values so that<br />

low values <strong>of</strong> p (say, less than 0.10) indicate the null hypothesis (that Model A is<br />

the true model) is rejected. Each row <strong>of</strong> the tables test the null hypothesis that<br />

Model A is superior to the alternative hypothesis (given by the columns). Cells<br />

with a p value <strong>of</strong> less than 0.10 are shaded so as to highlight the key results.<br />

An examination <strong>of</strong> Table 4 <strong>and</strong> Table 5 clearly indicates that the multivariate<br />

filters (including the bivariate filter) are to be preferred to those estimated using<br />

other techniques. Indeed, for those multivariate techniques which allow for<br />

greater non-linearity in the estimated potential output series (ie those where λ =<br />

1600 or λ = 200) the null hypothesis is accepted against all alternative techniques<br />

including the more linear forms <strong>of</strong> the multivariate filter.<br />

Overall, the tests suggest that there is little difference between the bf1600,<br />

mfa1600, <strong>and</strong> mfa200 output gaps.<br />

5.3 Summary<br />

The analysis in the previous section suggests that the bf1600, mfa1600, <strong>and</strong><br />

mfa200 output gaps are in a statistical sense equally good at explaining trends<br />

in inflation. The potential growth rates suggested by these series range from 2.79<br />

18<br />

It should be noted that the other variables in the regression, particularly unit labour<br />

costs, may also capture part <strong>of</strong> the total cyclical influence. Therefore, the estimated<br />

coefficient on the output gap term should not be interpreted as the total elasticity with<br />

respect to cyclical pressures.


TABLE 4: Non-nested model selection: J test<br />

tp1 tp2 yl1 yl2 yk1 yk2 pf1 hp1600 hp200 hp100 bf1600 bf200 bf100 mfa1600 mfa200 mfa100 mfb1600 mfb200 mfb100<br />

tp1 0.366 0.530 0.752 0.045 0.120 0.102 0.072 0.016 0.372 0.003 0.002 0.271 0.004 0.002 0.212 0.006 0.006 0.313<br />

tp2 0.887 0.553 0.372 0.091 0.173 0.161 0.139 0.041 0.823 0.006 0.006 0.505 0.006 0.003 0.388 0.009 0.010 0.590<br />

yl1 0.237 0.142 0.486 0.031 0.073 0.052 0.046 0.015 0.227 0.002 0.003 0.138 0.001 0.001 0.111 0.003 0.003 0.151<br />

yl2 0.276 0.102 0.451 0.022 0.058 0.049 0.016 0.007 0.080 0.002 0.002 0.085 0.002 0.001 0.086 0.003 0.003 0.010<br />

yk1 0.405 0.559 0.557 0.356 0.332 0.274 0.240 0.072 0.542 0.020 0.020 0.379 0.021 0.013 0.325 0.025 0.018 0.377<br />

yk2 0.718 0.504 0.956 0.582 0.152 0.476 0.352 0.080 0.487 0.014 0.016 0.457 0.011 0.009 0.604 0.017 0.021 0.363<br />

pf1 0.818 0.699 0.687 0.689 0.168 0.754 0.444 0.108 0.479 0.017 0.022 0.425 0.011 0.011 0.488 0.017 0.026 0.466<br />

hp1600 0.579 0.827 0.816 0.180 0.184 0.714 0.608 0.173 0.395 0.016 0.030 0.962 0.016 0.014 0.787 0.029 0.029 0.991<br />

hp200 0.307 0.669 0.688 0.234 0.186 0.489 0.523 0.967 0.970 0.022 0.101 0.678 0.052 0.039 0.574 0.098 0.076 0.615<br />

hp100 0.466 0.445 0.664 0.190 0.064 0.120 0.091 0.065 0.033 0.004 0.007 0.239 0.004 0.003 0.225 0.005 0.007 0.359<br />

bf1600 0.593 0.974 0.800 0.603 0.500 0.946 0.765 0.494 0.184 0.574 0.697 0.760 0.588 0.340 0.870 0.899 0.418 0.884<br />

bf200 0.223 0.368 0.684 0.429 0.254 0.444 0.535 0.666 0.953 0.862 0.334 0.636 0.225 0.213 0.543 0.328 0.239 0.542<br />

bf100 0.870 0.811 0.928 0.407 0.088 0.210 0.151 0.185 0.052 0.495 0.007 0.011 0.006 0.005 0.431 0.008 0.011 0.763<br />

mfa1600 0.852 0.878 0.334 0.976 0.613 0.654 0.403 0.566 0.714 0.561 0.726 0.472 0.595 0.391 0.623 0.960 0.490 0.716<br />

mfa200 0.348 0.541 0.554 0.699 0.492 0.799 0.983 0.971 0.903 0.848 0.704 0.938 0.976 0.706 0.895 0.635 0.402 0.818<br />

mfa100 0.879 0.931 0.952 0.570 0.097 0.332 0.211 0.244 0.056 0.652 0.009 0.012 0.603 0.008 0.006 0.010 0.013 0.380<br />

mfb1600 0.800 0.693 0.575 0.968 0.371 0.543 0.379 0.683 0.894 0.420 0.386 0.340 0.381 0.341 0.189 0.393 0.571 0.442<br />

mfb200 0.643 0.993 0.617 0.506 0.222 0.843 0.723 0.614 0.493 0.848 0.230 0.238 0.675 0.231 0.133 0.597 0.541 0.689<br />

mfb100 0.954 0.784 0.840 0.436 0.079 0.155 0.144 0.165 0.045 0.761 0.007 0.009 0.608 0.006 0.005 0.256 0.008 0.010<br />

Darren Gibbs 111


112<br />

TABLE 5: Non-nested model selection: Cox test<br />

tp1 tp2 yl1 yl2 yk1 yk2 pf1 hp1600 hp200 hp100 bf1600 bf200 bf100 mfa1600 mfa200 mfa100 mfb1600 mfb200 mfb100<br />

tp1 0.224 0.308 0.688 0.000 0.021 0.008 0.000 0.000 0.166 0.000 0.000 0.109 0.000 0.000 0.075 0.000 0.000 0.166<br />

tp2 0.871 0.349 0.017 0.001 0.083 0.057 0.033 0.000 0.795 0.000 0.000 0.412 0.000 0.000 0.275 0.000 0.000 0.578<br />

yl1 0.005 0.000 0.222 0.000 0.000 0.000 0.000 0.000 0.012 0.000 0.000 0.002 0.000 0.000 0.001 0.000 0.000 0.002<br />

yl2 0.043 0.000 0.169 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000 0.000<br />

yk1 0.169 0.440 0.354 0.010 0.160 0.102 0.074 0.001 0.399 0.000 0.000 0.189 0.000 0.000 0.130 0.000 0.000 0.179<br />

yk2 0.706 0.486 0.952 0.393 0.017 0.412 0.234 0.002 0.479 0.000 0.000 0.432 0.000 0.000 0.575 0.000 0.000 0.343<br />

pf1 0.801 0.682 0.687 0.585 0.025 0.732 0.343 0.011 0.475 0.000 0.000 0.406 0.000 0.000 0.462 0.000 0.000 0.449<br />

hp1600 0.448 0.814 0.774 0.000 0.035 0.666 0.540 0.073 0.398 0.000 0.000 0.958 0.000 0.000 0.752 0.001 0.001 0.990<br />

hp200 0.044 0.591 0.586 0.000 0.030 0.364 0.416 0.963 0.966 0.004 0.027 0.608 0.008 0.003 0.419 0.028 0.007 0.519<br />

hp100 0.290 0.327 0.553 0.000 0.000 0.032 0.016 0.007 0.000 0.000 0.000 0.147 0.000 0.000 0.114 0.000 0.000 0.259<br />

bf1600 0.468 0.971 0.793 0.438 0.386 0.940 0.751 0.485 0.178 0.577 0.654 0.750 0.535 0.254 0.860 0.886 0.321 0.874<br />

bf200 0.003 0.158 0.579 0.11 0.085 0.297 0.426 0.600 0.947 0.838 0.245 0.549 0.121 0.127 0.423 0.222 0.120 0.413<br />

bf100 0.850 0.783 0.918 0.055 0.001 0.129 0.072 0.066 0.000 0.473 0.000 0.000 0.000 0.000 0.367 0.000 0.000 0.739<br />

mfa1600 0.826 0.869 0.392 0.974 0.534 0.642 0.402 0.554 0.694 0.566 0.691 0.384 0.590 0.310 0.616 0.955 0.406 0.706<br />

mfa200 0.098 0.408 0.575 0.602 0.373 0.765 0.982 0.967 0.894 0.838 0.663 0.930 0.974 0.666 0.879 0.579 0.279 0.786<br />

mfa100 0.869 0.922 0.947 0.371 0.001 0.254 0.130 0.090 0.000 0.635 0.000 0.000 0.590 0.000 0.000 0.000 0.000 0.346<br />

mfb1600 0.791 0.684 0.592 0.965 0.218 0.534 0.378 0.667 0.883 0.440 0.310 0.235 0.394 0.262 0.094 0.401 0.510 0.451<br />

mfb200 0.642 0.993 0.629 0.536 0.056 0.831 0.707 0.533 0.392 0.838 0.123 0.120 0.665 0.128 0.041 0.589 0.477 0.679<br />

mfb100 0.949 0.752 0.808 0.102 0.000 0.82 0.061 0.045 0.000 0.742 0.000 0.000 0.561 0.000 0.000 0.188 0.000 0.000<br />

<strong>Labour</strong> Market Bulletin 1995: 1


Darren Gibbs 113<br />

to 4.19 percent (averaging 3.51 percent); while all the series suggest that a small<br />

negative output gap still existed in March 1994, ranging from between 0.25 to<br />

1.95 percent <strong>of</strong> GDP (averaging 1 percent <strong>of</strong> GDP). It is comforting that the<br />

preferred estimates span a range which would be considered plausible by most<br />

economists. However the lack <strong>of</strong> precision, which characterises empirical studies<br />

<strong>of</strong> this type, severely hinders the usefulness <strong>of</strong> these measures for policy analysis.<br />

Nevertheless, if we combine the information about potential growth rates, the<br />

output gap, <strong>and</strong> the elasticities <strong>of</strong> the output gaps in the estimated inflation<br />

equations, <strong>and</strong> assume (somewhat heroically) that growth in potential output<br />

over the year to March 1995 is the same as over the year to March 1994, the<br />

estimates imply that actual output could grow by between 3.8 percent <strong>and</strong> 6.1<br />

percent in the year to March 1995 (in point to point terms) before the output gap<br />

becomes positive <strong>and</strong> inflationary pressures emerge. Beyond that, every extra 1<br />

percent <strong>of</strong> growth would contribute between 0.59 percent to 0.72 percent to the<br />

annual inflation rate.<br />

6 Conclusion<br />

Empirical measures <strong>of</strong> the degree <strong>of</strong> spare capacity in the economy <strong>and</strong> <strong>of</strong> the<br />

economy’s potential (non-inflationary) growth rate are important for a number<br />

<strong>of</strong> reasons. Such measures, if robust, are particularly important for monetary <strong>and</strong><br />

fiscal policy, for labour market monitoring, forecasting <strong>and</strong> policy <strong>and</strong> for private<br />

sector planning.<br />

This paper surveyed a variety <strong>of</strong> techniques used to measure potential output.<br />

In the presence <strong>of</strong> uncertainties surrounding the true structural determinants <strong>of</strong><br />

the supply side <strong>of</strong> the economy together with poor quality data, the paper argues<br />

that multivariate filtering techniques are likely to be preferable to more rigid<br />

structural approaches such as those based on the estimation <strong>of</strong> production<br />

functions (especially in the short term).<br />

This view is supported by preliminary estimates <strong>of</strong> output gaps with New<br />

Zeal<strong>and</strong> data. Nevertheless, although statistical techniques identify some<br />

measures as better than others, these estimates are still imprecise. The current<br />

potential growth rates preferred by the study range from 2.79 to 4.19 percent<br />

(averaging 3.51 percent); while all series suggest that a small negative output gap<br />

still existed in March 1994, ranging from between 0.25 to 1.95 percent <strong>of</strong> GDP<br />

(averaging 1 percent <strong>of</strong> GDP).<br />

Although the theory hypothesises a relationship between the measured<br />

output gaps <strong>and</strong> the change in inflation, the empirical measures derived in this<br />

paper display a stronger relationship with the level <strong>of</strong> inflation. Regression<br />

analysis suggests that between 0.59 percent to 0.72 percent is added to the annual<br />

inflation rate for each percentage point <strong>of</strong> GDP in which actual output exceeds<br />

potential. However, further research aimed at testing for a non-linear or


114<br />

<strong>Labour</strong> Market Bulletin 1995: 1<br />

asymmetric inflationary response to both the change <strong>and</strong> the level <strong>of</strong> the output<br />

gap is required.<br />

One characteristic <strong>of</strong> the preferred estimates is that they suggest, perhaps not<br />

surprisingly, that supply shocks have dominated over the last two decades.<br />

Given the different inflationary dynamics implied by dem<strong>and</strong> versus supply<br />

shocks, policymakers would be wise to keep this in mind when seeking to draw<br />

conclusions regarding recently observed sharp increases in actual output.<br />

Given the wide range <strong>of</strong> estimates <strong>of</strong> potential output which appear<br />

consistent with the data, estimates must be treated with considerable caution.<br />

Although further research is likely to yield better estimates than those presented<br />

here, empirical measures <strong>of</strong> potential output are unlikely ever to be sufficiently<br />

robust as to be useful for fine-tuning macroeconomic policy. As Giorno et al<br />

(1995) argue:<br />

. . . the OECD has revised its estimation methods to provide a single measure <strong>of</strong><br />

potential output. . . . Nonetheless, it is clear from this work <strong>and</strong> the wide range <strong>of</strong><br />

analytic <strong>and</strong> survey based indicators which are available that significant margins <strong>of</strong><br />

error are involved in their estimation <strong>and</strong> use. Reliance therefore cannot wholly be<br />

placed on a single measure <strong>of</strong> potential or trend output, <strong>and</strong> related indicators must<br />

therefore be treated with due caution (p 6).


Darren Gibbs 115<br />

References<br />

Adams, C; Fenton, P R; <strong>and</strong> Larsen, F (1987) <strong>Potential</strong> output in major industrial countries, Staff Studies<br />

for the World Economic Outlook, World Economic <strong>and</strong> Financial Surveys, IMF, pp 1–38.<br />

Adams, C <strong>and</strong> Coe D T (1990) A systems approach to estimating the natural rate <strong>of</strong> unemployment <strong>and</strong><br />

potential output for the United States, IMF Staff Papers, Vol 37, No 2, pp 232–293.<br />

Christiano, L J (1981) A survey <strong>of</strong> measures <strong>of</strong> capacity utilisation, IMF Staff Papers, March, pp 144–198.<br />

Domar, E D (1947) ‘Expansion <strong>and</strong> employment’, American Economic Review, March 1947.<br />

Giorno, C; Richardson, P; Roseveare, D; <strong>and</strong> van den Noord, P (1995) Estimating potential output,<br />

output gaps, <strong>and</strong> structural budget balances, OECD Working Papers, No 152.<br />

Gordon, R J (1979) A comment on the Perl<strong>of</strong>f <strong>and</strong> Wachter paper, Carnegie Rochester Conference Series<br />

10, eds K Brunner <strong>and</strong> A H Meltzer.<br />

Harrod, R F (1957) ‘Domar <strong>and</strong> dynamic economics’, Economic Journal, September 1957.<br />

Hodrick, R J <strong>and</strong> Prescott, E C (1980) Post-war US business cycles: An empirical investigation, Discussion<br />

Paper 451, Evanston, Illinois: Centre for Mathematical Studies in Economics <strong>and</strong> Management<br />

Science, Northwestern University.<br />

King, R G <strong>and</strong> Rebelo, S T (1989) Low frequency filtering <strong>and</strong> real business cycles, Working Paper No 205,<br />

Rochester Center for Economic Research, University <strong>of</strong> Rochester.<br />

Laxton, D; Meredith, G; <strong>and</strong> Rose, D (1994) Asymmetric effects <strong>of</strong> economic activity on inflation: Evidence<br />

<strong>and</strong> policy implications, IMF Working Paper WP/94/139.<br />

Laxton, D <strong>and</strong> Tetlow, R (1992) A simple multivariate filter for the measurement <strong>of</strong> potential output, Bank<br />

<strong>of</strong> Canada Working Paper.<br />

Minford, P (1994) Deregulation <strong>and</strong> unemployment—The UK experience <strong>and</strong> lessons for New Zeal<strong>and</strong>,<br />

Liverpool Macroeconomic Research Ltd.<br />

OECD (1994) Cyclical adjustment <strong>of</strong> general government balances, Draft Working Paper.<br />

Okun, A M (1962) <strong>Potential</strong> GDP: Its measurement <strong>and</strong> significance, Proceedings <strong>of</strong> the Business <strong>and</strong><br />

Economic Statistics Section, American Statistical Association, pp 98–104.<br />

Panic, M (1978) Capacity utilisation in UK manufacturing industry, Discussion Paper No 5, National<br />

Economic Development Office.<br />

Perl<strong>of</strong>f, J M <strong>and</strong> Wachter, M L (1979) A production function nonaccelerating inflation approach to potential<br />

output: Is measured potential output too high?, Carnegie-Rochester Conference Series 10, eds K Brunner<br />

<strong>and</strong> A H Meltzer.<br />

Solow, R M (1957) ‘Technical change <strong>and</strong> the aggregate production function’, Review <strong>of</strong> Economics <strong>and</strong><br />

Statistics 39, pp 312–320.<br />

Torres, R <strong>and</strong> Martin, J P (1990) ‘Measuring potential output in the seven major OECD countries’,<br />

OECD Economic Studies, No 14, Spring, pp 127–150.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!