Kumbahar 2018

Download as pdf or txt
Download as pdf or txt
You are on page 1of 105

Centre for Efficiency and Productivity Analysis

Working Paper Series


No. WP02/2018

Stochastic Frontier Analysis: Foundations and Advances

Subal C. Kumbhakar, Christopher F. Parameter, and Valentine Zelenyuk

Date: April 2018

School of Economics
University of Queensland
St. Lucia, Qld. 4072
Australia

ISSN No. 1932 - 4398


STOCHASTIC FRONTIER ANALYSIS: FOUNDATIONS AND
ADVANCES

SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

Abstract. This chapter reviews some of the most important developments in the econo-
metric estimation of productivity and efficiency surrounding the stochastic frontier model.
We highlight endogeneity issues, recent advances in generalized panel data stochastic fron-
tier models, nonparametric estimation of the frontier, quantile estimation and distribution
free methods. An emphasis is placed on highlighting recent research and providing broad
coverage, while details are left for further reading in the abundant (although not limited to)
list of references provided.

JEL Classification: C10, C13, C14, C50

Contents

1. Introduction and Overview 3


2. The Benchmark SFM 6
3. Handling Endogeneity in the SFM 25
4. Modeling Determinants of Inefficiency 33
5. Panel Data 47
6. Nonparametric Estimation of the SFM 68
7. Quantile Estimation of the SFM 82
8. Additional Approaches/Extensions of the SFM 85

Date: March 28, 2018.


Key words and phrases. Efficiency, Productivity, Panel Data, Endogeneity, Nonparametric, Determinants of
Inefficiency, Quantile, Identification.
Subal C. Kumbhakar, Corresponding Author, Department of Economics, State University of New York at
Binghamton; e-mail: [email protected]. Christopher F. Parmeter, Department of Economics, Univer-
sity of Miami; e-mail: [email protected]. Valentin Zelenyuk School of Economics and Centre for
Efficiency and Productivity Analysis, University of Queensland; e-mail: [email protected].

1
2 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

9. Available Software to Estimate SFMs 86


10. Conclusions 87
References 89
3

1. Introduction and Overview

The primary goal of this chapter is to introduce the wide audience of this Handbook
to the range of methods, developed over the last four decades, within one of the most
popular paradigms in modern productivity analysis - the approach called Stochastic Frontier
Analysis, often abbreviated as SFA.
The first, and one of the most important, questions a reader might wonder about is why
a researcher on productivity should ever care about SFA in general and especially about the
enormous variety of different types of SFA models that have been proposed over the last
four decades. Our goal in writing this chapter was to provide the reader with a good answer
to this important question. Here, we strive to outline the essence of major types of SFA
methods, providing the minimal but most essential details, and focusing on advantages and
disadvantages of each method for dealing with various aspects that arise in practice. We
hope that upon finishing reading this chapter the reader who is barely or even unfamiliar
with SFA obtains a general understanding of the importance and relevance of different SFA
methods, along with useful/key references for further details on each method. Of course, the
reader also deserves to receive a quick answer now, to decide if it is worth it for a productivity
researcher to read this chapter further - we try to give such a quick answer in this section.
The Nobel Laureate Paul Krugman was hardly exaggerating when he once quipped that
“Productivity isn’t everything, but in the long run, it’s almost everything.” The root of
this statement can be seen when looking at various theoretical models of economic growth,
e.g., starting from Solow’s growth model, the related variations of more advanced growth
theory models or empirical growth accounting approach to productivity measurement, as
well as from the more sophisticated measurements of productivity. Regardless of how the
productivity is measured, it is inevitably tied to measuring production relationships. Such
relationships are usually modeled through the so-called production functions or, more gen-
erally, transformation functions (e.g., Shephard’s distance functions, Directional distance
4 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

functions), cost functions, etc. In the classical growth accounting approach (Solow 1957),
all the variation in growth apart from the variation of inputs is attributed to the so-called
Solow’s residual, which under certain restrictions measures what is referred to as the change
in total factor productivity (TFP).
A well known problem of simple growth accounting is that it piles up and hides many
sources for growth, the most obvious of which is the statistical error. Standard regression
methods can and are often used to, basically, estimate average relationships conditional on
various factors (inputs, demographic and geographic factors, etc.) to filter out the effect of
statistical noise. All the deviations from the estimated regression curves in such approaches
are attributed to the statistical error, and all the decision making units (DMUs) represented
in the data as observations (e.g., firms, countries, etc.), are typically assumed to be fully
efficient or on the frontier of the production relationship. Such full efficiency assumption
certainly simplifies the measurement complexity, but is it really an innocent assumption?
Indeed, while many economic models admit the assumption that all firms are efficient,
the reality that one observes in practice usually suggests there are reasonable amounts of
inefficiency in this world. Such inefficiencies could arise, for example, because of asymmet-
ric information or more generally, the problem of incomplete markets (e.g., see Stiglitz &
Greenwald 1986), which to some extent are present almost in every aspect of our lives. Dif-
ferences in inefficiencies (or in relative productivity levels)1 across firms or countries can also
arise due to different managerial practices (e.g., see Bloom et al. 2016), which could in turn
be implied by the asymmetric information problem, different cultural beliefs, traditions and
expectations (Benabou & Tirole 2016). Does accounting for such inefficiency matter for pro-
ductivity measurement? Vast literature on the subject suggests that it indeed often matters

1
Despite the variety of definitions, intuitively, production efficiency can be understood as a relative measure
of productivity. In other words, production efficiency is a productivity measure that is being normalized
(e.g., to be between 0 and 1 to reflect percentages) relative to some benchmark, such as the corresponding
frontier outcome, optimal with respect to some criteria: e.g., maximal output given certain level of input and
technology in the case of technical efficiency, or minimal cost given certain level of output and technology in
the case of cost efficiency.
5

substantially, as has been documented in thousands of articles in the last four decades. The
difference is in the approach - SFA, data envelopment analysis (DEA), free disposable hull
(FDH), etc. - and the goal of this chapter is to give a sense of a few major approaches within
the SFA paradigm.
In a nutshell, the main premise of the SFA approach is a recognition that whether all DMUs
are efficient or not is an empirical question that can and should be statistically tested against
the data, while allowing for a statistical error. To enable such testing, the SFA approach
provides a framework where the production relationship is estimated also as a conditional
average (of outputs given inputs and other factors, in the case of production function) but
the total deviation from the regression curve is decomposed into two terms - statistical noise
and inefficiency. Both of these terms are unobserved by a researcher but with relatively mild
assumptions the different approaches within SFA allow the analyst to estimate them for the
sample as a whole (e.g., representing an industry) or for each individual DMU.
Importantly, SFA approach also allows for the inefficiency term to be statistically insignif-
icant, if the data might suggest so, thus encompassing the classical approach with a naive
assumption of full efficiency as a special case and, importantly, allowing for this assumption
to be tested. Moreover, the SFA approach also contains the other extreme where one assumes
no statistical noise with all the deviations treated as inefficiency to the frontier. Thus, the
SFA approach is a natural compromise between approaches which make two extreme assump-
tions, yet also encompassing them as special cases, which can still be followed if the data
and the statistical tests from SFA would not recommend otherwise. If the tests support (or
at least cannot reject) the full efficiency hypothesis then one can proceed with the standard
regression techniques, or even with Solow’s growth accounting, but if not then accounting for
possible inefficiency could be critical for both quantitative and qualitative conclusions and,
perhaps more importantly, for the resulting policy implications. Indeed, if statistical tests
reject the hypothesis of full efficiency of DMUs, then it can be imperative to decompose the
productivity (be it Solow’s residual or any other productivity measure) further - to estimate
6 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

the inefficiency component for the sample (e.g. representing an industry) and for each indi-
vidual DMU. Moreover, SFA also provides a framework to analyze the sources of production
inefficiency and the variation of productivity levels, both of which can give important insights
into how to reduce inefficiency and increase productivity. We discuss these interesting and
important issues in this chapter. While some of the stylized facts we present here can be also
found in previous reviews (Lovell 1993, Kumbhakar & Lovell 2000, Greene 2008, Parmeter
& Kumbhakar 2014, Kumbhakar, Wang & Horncastle 2015), and it is impossible to give a
good review without following them to some degree, here we also summarize many of (what
we believe to be) the key recent developments as well as (with their help) shed some novel
perspectives onto the workhorse methods. So, all in all, our belief is that there is much value
added for the reader to complement what was done well in earlier reviews on this theme.
The rest of the chapter is structured as follows: Sections 2-4 focus on stochastic frontier
models (SFM) for cross-sectional variation in efficiency (relative productivity), where Section
2 covers the foundation laid by Aigner, Lovell & Schmidt (1977) and some closely related
research, with Section 3 discussing endogeneity issues and Section 4 focuses on modeling
the determinants of inefficiency. Section 5 focuses on SFA models for analyzing variation
of efficiency (or relative productivity) not only across firms but also over time, i.e., in the
panel data context. Section 6 reviews several prominent semi- and nonparametric approaches
to SFA. Section 7 briefly discusses a recent vein of literature focusing on quantile estima-
tion of the SFM. Section 8 presents some further extensions of the SFM, while Section 9
briefly summarizes some of the available software to estimate SFMs in practice. Section 10
concludes.

2. The Benchmark SFM

One of the main approaches to study productivity and efficiency of a cross-section of firms
is the SFM, independently proposed by Aigner et al. (1977) (ALS hereafter) and Meeusen
7

& van den Broeck (1977a) (MvB hereafter).2 Using conventional notation, let Yi be the
single-output for observation (e.g., firm) i and let yi = ln(Yi ). The SFM can be written for
a production frontier3 as

(2.1) yi = m(xi ; β) − ui + vi = m(xi ; β) + εi .

Here m(xi ; β) represents the production frontier of a firm (or more generally a DMU), with
given input vector xi . Our use of β is to clearly signify that we are parametrically specifying
our production function.4 The main difference between a standard production function setup
and the SFM is the presence of two distinct error terms in the model. The ui term captures
inefficiency, shortfall from maximal output dictated by the production technology, while the
vi term captures stochastic shocks. The standard neoclassical production function model
assumes full efficiency – so the SFM embraces it as a special case, when ui = 0, ∀i, and
allows the researcher to test this statistically.5
One shortcoming of the benchmark SFM is that the appearance of inefficiency in (2.1)
lacks any specific structural interpretation. Where is inefficiency coming from? It could
stem from inputs being used sub-optimally: workers may not put forth full effort or capital
may be improperly used, e.g., due to asymmetric information or other reasons hidden to the
researcher or even the firm. Without a specific structural link it is difficult to know just how
to treat inefficiency in (2.1). Thus, to estimate the model, several assumptions need to be
imposed. First, it is commonly assumed that inputs are independent of u and v, ui ⊥ x and
vi ⊥ x ∀x.6 Second, u and v are assumed to be independent of one another. Next, given
2
Battese & Corra (1977) and Meeusen & van den Broeck (1977b), while appearing in the same year, are
applications of the methods.
3Our discussion in this chapter will focus on a production frontier, as it is the most popular object of study,
while the framework for dual characterizations (e.g., cost, revenue, profit) or other frontiers is similar and
follows with only minor changes in notation.
4See Section 6 for a discussion on relaxing parametric restrictions on the production frontier in the SFM.
5Prior to the development of the SFM, approaches which intended to model inefficiency typically ignored v
i
leading to estimators of the SFM with less desirable statistical properties: see the work of Aigner & Chu
(1968), Timmer (1971), Afriat (1972), Dugger (1974), Richmond (1974), and Schmidt (1976).
6See Section 3 for a discussion on estimation of the SFM when some inputs are allowed to be endogenous.
8 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

that ui leads directly to a shortfall in output it must come from a one-sided distribution
implying that E[εi |x] 6= 0. This has two effects if one were to estimate the SFM using OLS.
First, the intercept of technology would not be identified, and second, without any additional
information, nothing can be said about inefficiency. Additionally, if ui is an independently
and identically distributed random variable, there is no policy implication behind it given
that nothing can directly increase or decrease inefficiency. That is, the conclusions of such
a study would be descriptive (reporting presence or absence of inefficiency) rather than
prescriptive or normative.7.
Denote E[u] as µu and ε∗i = vi − (ui − µu ), the benchmark SFM can be rewritten as

(2.2) yi = m(xi ; β) − µu − (ui − µu ) + vi ≡ m∗ (xi ; β) + ε∗i

and E[ε∗i |x] = 0. The OLS estimator could be used to recover mean inefficiency adjusted
technology m∗ (xi ; β) = m(xi , β) − µu in this case. However, rarely is the sole focus of
an analysis of productivity on the production technology. It is more likely that both the
production technology and information about inefficiency for each DMU are the targets of
interest; more structure is required on the SFM in this case.
ALS’ and MvB’s approach to extract information on inefficiency, while also estimating
technology, was to impose distributional assumptions on ui and vi , recovering the implied
distribution for εi and then estimating all of the parameters of the SFM with the maximum
likelihood estimator (MLE). vi was assumed to be distributed as a normal with mean 0
and variance σv2 by both sets of researchers, while the distribution of ui differed across the
papers; Aigner et al. (1977) assumed that ui was generated from a half-normal distribution,
N+ (0, σu2 ), whereas MvB assumed ui was distributed exponentially, with parameter σu .8
Even though the half-normal and exponential distributions are distinct, they possess sev-
eral common aspects. Both densities have modes at zero and monotonically decay (albeit at
7SeeSection 4 for models handling determinants of inefficiency
8ALS also briefly discussed the exponential distribution, but its use and development is mainly attributed
to MvB.
9

different speeds) as ui increases. The zero mode property is indicative of an industry where
there is a tendency for higher efficiency for the majority of the DMUs. Both densities would
be classified as single parameter distributions, which means that the mean and variance both
depend on the single parameter, and these distributions also possess the scaling property,
which we will discuss in section 4.3.

2.1. The Distribution of ε. Estimation of the SFM in (2.1) with maximum likelihood
requires that the density of ε, f (ε), is known. f (ε) can be determined through the distribu-
tional assumptions invoked for v and u. Not all pairs of distributional assumptions for v and
u will lead to a tractable density of f (ε), permitting estimation via maximum likelihood.
Fortunately, the half-normal specification of Aigner et al. (1977) and the exponential speci-
fication of MvB (along with the normal assumption for v), produce a density for ε that has
a closed form solution; direct application of maximum likelihood is straightforward in this
setting. For brevity we report the density of the composed error for the normal-half-normal
specification.

2
(2.3) f (ε) = φ(ε/σ)Φ(−ελ/σ),
σ

where φ(·) is the standard normal probability density function (pdf), Φ(·) is the standard
p
normal cumulative distribution function (cdf), with the parameterization σ = σu2 + σv2 and
λ = σu /σv . λ is commonly interpreted as the proportion of variation in ε due to inefficiency.
The density of ε in (2.3) can be characterized as that of a skew normal random variable
with location parameter 0, scale parameter σ and skew parameter −λ.9 This connection
has only recently appeared in the efficiency and productivity literature (Chen, Schmidt &
Wang 2014).

9The pdf of a skew normal random variable x is f (x) = 2φ(x)Φ(αx). The distribution is right skewed if α > 0
and is left skewed if α < 0. We can also place the normal, truncated-normal pair of distributional
  assumptions
  
in this class. The pdf of x with location ξ, scale ω, and skew parameter α is f (x) = ω φ ω Φ α x−ξ
2 x−ξ
ω .
See O’Hagan & Leonard (1976) and Azzalini (1985) for more details.
10 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

From f (ε) in (2.3), along with independence assumptions on ui and vi the log-likelihood
function is
n
! n n
Y X 1 X 2
(2.4) ln L = ln f (εi ) = −n ln σ + ln Φ(−εi λ/σ) − 2 ε,
i=1 i=1
2σ i=1 i

where εi = yi − m(xi ; β). The SFM can be estimated using the traditional maximum likeli-
hood estimator (MLE). The benefit of this is that under the assumption of correct distribu-
tional specification of ε, the MLE is asymptotically efficient (i.e., consistent, asymptotically
normal and its asymptotic variance reaches the Cramer-Rao lower bound). A further benefit
is that a range of testing options are available. For instance, tests related to β can easily be
undertaken using any of the classic trilogy of tests: Wald, Lagrange multiplier, or likelihood
ratio. The ability to readily and directly conduct asymptotic inference is one of the major
benefits of stochastic frontier analysis over DEA.10

2.2. Alternative Specifications. The half-normal assumption for the one-sided ineffi-
ciency term is almost without question the most common distribution for inefficiency in
practice. This stems partly from posterity, partly from the closed form solution of the
likelihood function, and partly from the availability of software to estimate the model for
applied researchers. However, none of these reasons are sufficient for blind application of the
half-normal density for inefficiency in the SFM.

2.2.1. The Exponential Distribution. The exponential assumption on inefficiency is also pop-
ular. The exponential density is

1 −u/σu
(2.5) f (u) = e , u ≥ 0.
σu

10Thisin no way suggests that inference cannot be undertaken when the DEA estimator is deployed; rather,
the DEA estimator has an asymptotic distribution which is much more complicated that the MLE for the
SFM, and so direct asymptotic inference is not available; bootstrapping techniques are required for many of
the most popular DEA estimators (Simar & Wilson 2013, Simar & Wilson 2015).
11

For the normal-exponential distributional pair, the density of ε is

1 2 2
(2.6) f (ε) = Φ(−ε/σv − σv /σu )eε/σu +σv /2σu ,
σu

with likelihood function


n n
σv2
  X 1 X
(2.7) ln L = −n ln σu + n + ln Φ(−εi /σv − σv /σu ) + εi .
2σu2 i=1
σu i=1

Like the half-normal specification for u, the exponential specification monotonically de-
creases in u, suggesting that larger levels of inefficiency are less likely to occur than small
levels of inefficiency. Both the half-normal and exponential specifications for inefficiency
stem from what are known as single parameter distributions. Single parameter distributions
are the simplest distributions and an unfortunate (yet sometimes very convenient) property
of them is that all of their moments depend on this single parameter, which can restrict the
shape that the density can potentially take.11

2.2.2. The Truncated Normal Distribution. To allow more generality into the SFM, while
guarding against distribution misspecification, a variety of one-sided distributions have been
proposed for modeling ui in the SFM. Stevenson (1980) proposed the truncated-normal
distribution as a generalization of the half-normal distribution; whereas the half-normal
distribution is the truncation of the N (0, σu2 ) at 0, the truncated-normal distribution is the
truncation of the N (µ, σu2 ) at 0. The pre-truncation mean parameter, µ, affords the SFM
more flexibility in the shape of the distribution of inefficiency.
The truncated-normal density is

(u−µ)2
1 −
(2.8) f (u) = √ e 2σu2 , u ≥ 0.
2πσu Φ(µ/σu )

This density reduces to the half-normal distribution when µ = 0 and thus provides a gener-
alization (more specifically a nesting structure), and an opportunity for inference on µ. An
11See Parmeter & Kumbhakar (2014) for a more detailed analysis of the SFM with u distributed exponentially.
12 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

intuitive appeal of deploying truncated-normal distribution in practice is that, unlike the


half-normal and exponential densities, the truncated-normal density has a mode at 0 only
when µ ≤ 0, but otherwise has a mode at µ. When µ > 0, the implication is that producers
in a given market would tend to have inefficiency ui near µ > 0 rather than near 0. This
connotation may be more realistic in some settings (e.g., the regulatory environment) than
the half-normal assumption, where the probability of being less efficient is much larger than
of being grossly inefficient.
For the normal-truncated-normal distributional pair, the density of ε is
   
1 ε+µ µ ελ .
(2.9) f (ε) = φ Φ − Φ(µ/σu ).
σ σ σλ σ

The corresponding log-likelihood function is


n  2 n  
X εi + µ X µ εi λ
(2.10) ln L = −n ln σ − − n ln Φ(µ/σu ) + ln Φ − .
i=1
σ i=1
σλ σ

2.2.3. Other Distributions. Aside from the truncated-normal specification for the distribu-
tion of u, a variety of alternatives have been proposed throughout the literature. Greene
(1980a, 1980b) and Stevenson (1980) both proposed a gamma distribution for inefficiency.
The gamma distribution generalizes the exponential distribution in much the same way that
the truncated-normal distribution nests the half-normal distribution. Ritter & Simar (1997)
advocate against the use of the gamma specification in practice noting that large samples
were required to reliably estimate the parameters of the gamma distribution due to compu-
tational identification problems with the constant of the regression. Lee (1983) proposed a
four parameter Pearson density for the specification of inefficiency; unfortunately, this distri-
bution is intractable for applied work and until now has not appeared to gain popularity. Li
(1996) proposed use of the uniform distribution for inefficiency noting an intriguing feature
13

of the subsequent composed error density: that it could be positively skewed.12 Another
specification for inefficiency appears in Carree (2002), who assumes that the distribution
of u follows a binomial specification; this allows the skewness of the composed error to be
positive or negative. Gagnepain & Ivaldi (2002) specify inefficiency as being Beta distributed
when inefficiency can be defined as a percentage (scaled between 0 and 1), while Almanidis,
Qian & Sickles (2014) further generalize Stevenson’s (1980) framework by assuming a doubly
truncated-normal distribution for inefficiency. This distributional assumption also allows the
convolved error term to be either positively or negatively skewed.
A common theme of all of the papers just mentioned is that they focus exclusively on the
distribution of inefficiency inside the SFM. Recent literature has shed light on the features
of f (ε) for the SFM when both the density of v and the density of u are changed. Horrace &
Parmeter (2014) study the behavior of the composed error when v is distributed as Laplace
and u is distributed as truncated Laplace. Nguyen (2010) considers the Laplace-Exponential
distributional pair as well as the Cauchy-Half Cauchy pair for the two error terms of the
composed error. While these alternative distributional pairs do provide different insights
into the behavior of the composed error, it remains to be seen if they will be regularly
adopted in practice and whether they provide substantially different conclusions than the
most frequently adopted distributional pairs (normal-half-normal for example); see Section
2.4 for more discussion on the perceived importance of distributional assumptions regarding
estimation of the SFM.
It is important to note that the main idea behind the SFM is that nearly any pair of
distributions can be used to model u and v. The advantage of the normal-half-normal pair
that is dominant in the literature is that the likelihood function has an easy to evaluate
expression. In general this should not be expected. More likely than not, for a range

12Prior to Li (1996) all of the previously proposed distributions always produced a composed error density
that was theoretically negatively skewed. Note that if u is distributed uniformly over the interval [0, b],
inefficiency is equally likely to be either 0 or b.
14 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

of distributional assumptions the likelihood function will contain one or more intractable
integrals, complicating estimation.13

2.2.4. Alternative Estimation Approaches of the SFM. Given the focus on inefficiency in
the SFM, and the impact that the distributional assumption on u is likely to have on the
MLE, studying the behavior of the SFM across a range of distributional assumptions is
desirable. However, outside of a few specifications (half-normal, exponential, truncated-
normal for ui and normal for vi ), the composed error density will not have a likelihood
function that lends itself for easy evaluation. In these cases it can be difficult to estimate
all of the parameters of the SFM, but several approaches exist, ranging in complexity, to
estimate the SFM when direct estimation of the likelihood function is not feasible. The
simplest approach, dubbed corrected OLS (COLS) by Olson, Schmidt & Waldman (1980),14
recognizes that OLS estimation of the SFM produces consistent estimates of the coefficients
of the frontier function aside from the intercept. The intercept is biased downward by the
p
expected level of industry inefficiency E [u] = 2/πσu . Olson et al.’s (1980) insight was that
for a given pair of distributional assumptions (normal-exponential, say), the central moments
of the OLS residuals could be used to construct consistent estimators of the parameters of
the convolved error. Once these were estimated, expected inefficiency could be estimated and
the bias in the intercept corrected. The beauty of COLS from the applied perspective is that
OLS can be used and difficult likelihood functions do not have to be derived or estimated.15
13Note that the likelihood function for the normal-half-normal pair is dependent upon the cdf of the normal
distribution, Φ(·) which contains an integral, but this can be quickly and easily evaluated across all modern
software platforms.
14See also Greene (1980a, pg. 31-32). Richmond (1974) also proposed adjusting the intercept from OLS
estimation, however, his model differs from that of Olson et al. (1980) by assuming the presence of inefficiency
(which follows a gamma distribution) but no noise.
15There exists some confusion over the terminology COLS as it relates to another method, modified OLS
(MOLS). Beginning with Winsten (1957), and discussed in Gabrielsen (1975) and Greene (1980a, pg. 32-
34), MOLS shifts the estimated OLS production function until all of the observations lie on or below the
‘frontier’. At issue is the appropriate name of these two techniques. Greene (2008) called the bounding
approach COLS, crediting Lovell (1993, pg. 21) with the initial nomenclature, and referred to MOLS as the
method in which bias corrects the intercept based on a specific set of distributional assumptions. Further,
Kumbhakar & Lovell (2000, pg. 70-71) also adopted this terminology. However, given that Olson et al. (1980,
pg. 69) explicitly used the terminology COLS, in our review we will adopt COLS to imply bias correction of
15

Several newer approaches exist as well. One that is becoming popular is maximum simu-
lated likelihood (MSL) estimation (McFadden 1989). Greene (2003) used MSL estimation to
estimate the parameters of the SFM for the normal-gamma convolution. The key to imple-
mentation of the SFM when the composed error does not produce a tractable likelihood is to
notice that the integrals that commonly remain in the density (from integrating u out of the
density) can be treated as expectations and evaluated by simulation rather than analytic op-
timization. Given that the distribution of u is assumed known (up to unknown parameters),
for a given set of parameters, draws can be taken and the subsequent expectation that is
evaluated can replace the integral. Optimization proceeds by searching over the parameter
space until a global maximum is found.
An even more recent approach to evaluating intractable likelihoods is found in Tsionas
(2012) who suggested estimation of the parameters of the SFM through the characteristic
function of the composed error. The reason that this will work is that the characteristic func-
tion is a unique representation of a distribution (whether the density does or does not exist),
and following from the convolution theorem, the characteristic function of two independent
random variables (here v and u) added together is the product of the individual charac-
teristic functions. The characteristic functions for all of the densities described above are
known, and so, using the Fast Fourier Transform, the estimated characteristic function can
be mapped to the underlying density, and subsequently, the likelihood function. Tsionas’s
(2012) method is somewhat computationally complicated, but it offers another avenue to
estimate the SFM under alternative distributional assumptions on both v and u.

the OLS intercept and MOLS as a procedure that shifts up (or down) the intercept to bound all of the data.
The truth is both COLS and MOLS are the same in the sense that the OLS intercept is augmented, it is just
in how each method corrects, or modifies, the intercept that is important. While we are departing from the
more mainstream use of COLS and MOLS currently deployed, given the original use of COLS, coupled with
myriad papers written by Peter Schmidt and coauthors that we discuss here, we will use the COLS acronym
to imply a bias corrected intercept.
16 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

2.3. Estimation of Individual Inefficiency. Once the parameters of the SFM have been
estimated, estimates of firm level productivity and efficiency can be recovered. Observation-
specific estimates of inefficiency are one of the main benefits of the SFM relative to neo-
classical models of production. Firms can be ranked according to estimated efficiency; the
identity of under-performing firms as well as those who are deemed best practice can also be
gleaned from the SFM. All of this information is useful in helping to design more efficient
public policy or subsidy programs aimed at improving the market, for example, insulating
consumers from the poor performance of heavily inefficient firms.
As a concrete illustration, consider firms operating electricity distribution networks that
typically possess a natural local monopoly given that the construction of competing net-
works over the same terrain is prohibitively expensive.16 It is not uncommon for national
governments to establish regulatory agencies which monitor the provision of electricity to
ensure that abuse of the inherent monopoly power is not occurring. Regulators face the task
of determining an acceptable price for the provision of electricity while having to balance
the heterogeneity that exists across the firms (in terms of size of the firm and length of
the network). Firms which are inefficient may charge too high a price to recoup a profit,
but at the expense of operating below capacity. However, given production and distribution
shocks, not all departures from the frontier represent inefficiency. Thus, measures designed
to account for noise are required to parse information from εi regarding ui .
Alternatively, further investigation could reveal what it is that makes these establishments
attain such high levels of performance. This could then be used to identify appropriate gov-
ernment policy implications and responses or identify processes and/or management prac-
tices that should be spread (or encouraged) across the less efficient, but otherwise similar,
16The current literature is fairly rich on various examples of empirical values of SFA for the estimation and
use of efficiency estimates in different fields of research. For example, in the context of electricity providers,
see Knittel (2002), Hattori (2002), and Kuosmanen (2012); for banking efficiency, see Case, Ferrari & Zhao
(2013) and references cited therein; for the analysis of the efficiency of national health care systems, see
Greene (2004) and a review by Hollingsworth (2008); for analyzing efficiency in agriculture, see Bravo-Ureta
& Rieger (1991), Battese & Coelli (1992, 1995), and Lien, Kumbhakar & Hardaker (2017), to mention just
a few.
17

units. This is the essence of the determinants of inefficiency approach which we will discuss
in Section 4. More directly, efficiency rankings are used in regulated industries such that
regulators can set tougher future cost reduction targets for the more inefficient companies,
in order to ensure that customers do not pay for the inefficiency of firms.
bu2 . This provides
The only direct estimate coming from the normal-half-normal SFM is σ
context regarding the shape of the half-normal distribution on ui and the industry average
efficiency E[u], but not on the absolute level of inefficiency for a given firm. If we are only
concerned with the average level of technical efficiency for the population, then this is all the
information that is needed. Yet, if we want to know about a specific firm, then something
else is required. The main approach to estimating firm level inefficiency is the conditional
mean estimator of Jondrow, Lovell, Materov & Schmidt (1982), commonly known as the
JLMS estimator. Their idea was to calculate the expected value of ui conditional on the
realization of composed error of the model, εi ≡ vi −ui , i.e., E[ui |εi ].17 This conditional mean
of ui given εi gives a point prediction of ui . The composed error contains individual-specific
information, and the conditional expectation is one measure of firm-specific inefficiency.
Jondrow et al. (1982) show that for the normal-half-normal specification of the SFM, the
conditional density function of ui given εi , f (ui |εi ), is N+ (µ∗i , σ∗2 ), where

−εi σu2
(2.11) µ∗i =
σ2

and

σv2 σu2
(2.12) σ∗2 = .
σ2

Given results on the mean of a truncated-normal density it follows that

σ∗ φ( µσ∗i∗ )
(2.13) E[ui |εi ] = µ∗i +   .
Φ µσ∗i∗

17Jondrow et al. (1982) also suggested an alternative estimator based on the conditional mode.
18 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

The individual estimates are then obtained by replacing the true parameters in (2.13) with
MLE estimates from the SFM.
Another measure of interest is the Afriat-type level of technical efficiency, defined as
e−ui = Yi /em(xi ) evi ∈ [0, 1]. This is useful in cases where output is measured in logarithmic
form. Further, technical efficiency is bounded between 0 and 1, making it somewhat easier
to interpret relative to a raw inefficiency score. Since e−ui is not directly observable, the
idea of Jondrow et al. (1982) can be deployed here, and E [e−ui |εi ] can be calculated (Lee &
Tyler 1978, Battese & Coelli 1988). For the normal-half-normal model, we have
 
µ∗i
1 2
Φ σ∗
− σ ∗
E e−ui |εi = e(−µ∗i + 2 σ∗ )
 
(2.14)   ,
Φ µσ∗i∗

where µ∗i and σ∗ were defined in (2.11) and (2.12), respectively. Technical efficiency estimates
are obtained by replacing the true parameters in (2.14) with MLE estimates from the SFM.
When ranking efficiency scores, one should use estimates of 1 − E [ui |εi ], which is the first
order approximation of (2.14). Similar expressions for the Jondrow et al. (1982) and Battese
& Coelli (1988) efficiency scores can be derived under the assumption that u is exponential
(Kumbhakar & Lovell 2000, p. 82), truncated-normal (Kumbhakar & Lovell 2000, p. 86),
and Gamma (Kumbhakar & Lovell 2000, p. 89); see also Kumbhakar et al. (2015).

2.3.1. Inference about the Presence of Inefficiency. Having estimated the benchmark SFM,
a natural hypothesis is whether inefficiency is even present. In this case the null hypothesis
of interest is H0 : σu2 = 0 against H1 : σu2 > 0.18 The direct way to test the H0 is through
a likelihood ratio test, keeping in mind that the unrestricted model is the assumed SFM
and the restricted model is the linear regression model (or more specifically the normal
regression model). There is a problem with implementation of this test however. Under H0

18One could test if other moments of the distribution were 0 as well, but most of the SFMs parameterize the
distribution of u with σu and so this seems the most natural.
19

σu2 is restricted to lie on the boundary of the parameter space and this precludes direct use
of a likelihood ratio test.
Coelli (1995) demonstrates that under H0 the likelihood ratio statistic in this setting is a
50:50 mixture of a χ21 distribution, the distribution of the ordinary likelihood ratio statistic
if the parameter were not on the boundary of the parameter space, and a χ20 , known as
the chi-bar-square distribution, χ̄2 , (Coelli 1995, Silvapulle & Sen 2005). This second piece
is what captures the potential presence of the σu2 parameter to lie on the boundary of the
parameter space and creates a point mass in the asymptotic distribution of the likelihood
ratio statistic.
Calculation of the test statistic itself is invariant to whether the parameter lies on the
boundary under H0 . What does change is how one goes about calculating either the p-value
or the critical value to assess the outcome of the test. In the case of the 50:50 mixture, the
critical values are determined by looking at the 2α-level critical value from a χ21 distribution.
For example, whereas the critical value for a 5% significance level is 3.841 for χ21 , it is 2.706
for the 50:50 mixture. More specifically, Table 1 presents the critical values of both the χ21
and the 50:50 mixture for a range of significance levels.

[Table 1 about here.]

An alternative type of test for the presence of inefficiency is based on the skewness of the
residuals. A variety of tests for skewness exist, notably Ahmad & Li (1997), Kuosmanen
& Fosgerau (2009) and Henderson & Parmeter (2015b). Henderson & Parmeter (2015b)
proposed a bootstrap based version of Ahmad & Li’s (1997) asymptotic test, noting that
in finite samples the bootstrap version is likely to have superior performance. This test
involves estimating the SFM using OLS and then testing whether the distribution of the
OLS residuals is symmetric. Kuosmanen & Fosgerau’s (2009) test of symmetry is also based
on the bootstrap, but rather than focus on the estimated distribution of the OLS residuals,
their test focuses exclusively on the skewness coefficient of the residuals. Both of these tests of
20 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

symmetry are appealing because they do not require parametric distributional assumptions
and can be implemented after having estimated the SFM using OLS.

2.3.2. Inference about the Distribution of Inefficiency. It is important to recognize, despite


the frequent misuse of terminology, that the JLMS or Battese-Coelli (or similar types) effi-
ciency estimators are not estimators of ui or e−ui , respectively and do not converge to them
for n → ∞. As n → ∞, the new observations represent different firms each with their own
level of inefficiency and noise (upon which JLMS conditions), rather than observations from
the same firm. Even more importantly, the JLMS estimator was not intended to estimate
unconditional inefficiency. The JLMS estimator is however, a consistent estimator for the
expected level of inefficiency conditional on the particular realizations of ε.19
The JLMS efficiency scores can be used to provide a (limited) test of the distribution
of inefficiency. The key insight to understand how a test can be constructed is that if the
distributional assumptions are correct, then the distribution of E[ui |εi ] is completely known.
b i |εi ] to the true distribution of E[ui |εi ] will
Hence a comparison of the distribution of E[u
shed light into the statistical validity of the assumed distributions for u and v. Wang &
Schmidt (2009) derived the distribution of E[ui |εi ] for the normal-half-normal SFM while
Wang, Amsler & Schmidt (2011) proposed χ2 and Komolgorov-Smirnov type test statistics
against this distribution.20
We caution readers regarding a rejection with use of this test. A rejection does not
necessarily imply that the distributional assumption on u is incorrect, it could be that the
normality distributional assumption on v or some other assumptions about the SFM (e.g.,
the parametric form of m) is violated, and this is leading to the rejection. Similarly, one
must be careful in interpreting tests on the distribution of ε (or functionals of ε) when the

19The JLMS efficiency estimator is known as a shrinkage estimator; on average, it understates the efficiency
level of a firm with small ui while it overstates efficiency for a firm with large ui .
20See also Lee (1983) for a different test based on the Pearson distributional assumption for u.
21

distribution of v is also assumed to be normal. Alternative tests similar to Wang et al. (2011)
could be formulated using the Laplace-exponential SFM of Horrace & Parmeter (2014).

2.3.3. Predicting Inefficiency. Aside from testing for the appropriate distribution of ineffi-
ciency, one should also test, or present uncertainty, as it pertains to an individual efficiency
score. Each JLMS efficiency score is a prediction of inefficiency, and it is possible to cal-
culate prediction intervals. Interestingly, few applied papers cover in depth uncertainty of
estimated efficiency scores.
A prediction interval for E[ui |εi ] was first derived by Taube (1988) and also appeared
in Hjalmarsson, Kumbhakar & Heshmati (1996), Horrace & Schmidt (1996), and Bera &
Sharma (1999) (see the discussion of this in Simar & Wilson 2010). The prediction interval
is based on f (ui |εi ). The lower (Li ) and upper (Ui ) bounds for a (1 − α)100% prediction
interval are
   
−1
 α µ∗i
(2.15) Li =µ∗i + Φ 1− 1− 1−Φ − σ∗ ,
2 σ∗
   
−1 α µ∗i
(2.16) Ui =µ∗i + Φ 1− 1−Φ − σ∗ ,
2 σ∗

where µ∗i and σ∗ are defined in (2.11) and (2.12), respectively and replacing them with their
MLE estimates will give estimated prediction intervals for E[ui |εi ].
Wheat, Greene & Smith (2014) derived minimum width prediction intervals noting that
the confidence interval studied in Horrace & Schmidt (1996) was based on a symmetric two
sided interval. Given that the distribution of ui conditional on εi is truncated (at 0) normal
and asymmetric, this form of interval is not of minimum width. Parmeter & Kumbhakar
(2014) showed that depending upon the ratio of σu to σv , the difference in relative widths
of Horrace & Schmidt’s (1996) and Wheat et al.’s (2014) prediction intervals can be quite
substantial. It is thus recommended to use the intervals provided by Wheat et al. (2014) as
these are not based on symmetry. Note that, although we could predict u and construct a
22 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

prediction interval, this information is not that useful for policy purposes unless there are
some variables that affect inefficiency and such variables can be changed by a specific policy.

2.4. Do Distributional Assumptions Even Matter? An important empirical concern


when using the SFM is the choice of distributional assumptions made for v and u. The
distribution of v has almost universally been accepted as being normal in both applied
and theoretical work (a recent exception is Horrace & Parmeter 2014); the distribution
of u is more commonly debated, but relatively little work has been devoted to discerning
the impact that alternative shapes of the distribution can have. Moreover, choice of u is
often driven through available statistical software to implement the method rather than an
underlying theoretical link between a model of productive inefficiency and the exact shape
of the corresponding distribution.
A majority of applied papers studying productivity do not rigorously check differences in
estimates, or perform inference, across different distributional assumptions. Greene (1990)
is often cited as one of the first analyses to compare average inefficiency levels across several
distributional specifications (half-normal, truncated-normal, exponential, and gamma), and
he finds little difference in average inefficiency across 123 U.S. electric generation firms.
Following Greene’s (1990) investigation into the choice of distribution, Kumbhakar & Lovell
(2000) calculated the rank correlations amongst the JLMS scores from these same four
models, producing rank correlations as low as 0.75 and as high as 0.98.21
The intuition underlying these findings is that one’s understanding of inefficiency, as mea-
sured through the JLMS score, is robust to distributional choices, at least from a ranking
perspective. The reason for this can be found in the work of (Ondrich & Ruggiero 2001,
p. 438) who have shown that the JLMS efficiency scores are monotonic in ε provided that

21In a limited Monte Carlo analysis, Ruggiero (1999) compared rank correlations of stochastic frontier
estimates assuming that inefficiency was either half-normal (which was the true distribution) or exponential (a
misspecified distribution) and found very little evidence that misspecification impacted the rank correlations
in any meaningful fashion; Horrace & Parmeter (2014) conducted a similar set of experiments and found
essentially the same results.
23

the distribution of v is log-concave (which the normal distribution is). The implication here
is that firm rankings can be obtained via the OLS residuals without the need of distribu-
tional assumptions whatsoever (Bera & Sharma 1999). Thus, in light of these insights, the
important aspect of distributional choice for u is the impact that it has on the correspond-
ing estimates of the production function; when these estimates are robust to distributional
choice, so too will be the inefficiency rankings. Thus, if interest hinges on features of the
frontier, then so long as inefficiency does not depend on conditional variables (see Section
4), one can effectively ignore the choice of distribution, as this only affects (usually but
not substantially) the level of the estimated technology, but not its shape - which is what
influences measures such as returns to scale and elasticities of substitution.

2.5. Finite Sample Identification of Inefficiency. An early analysis of the finite sample
performance of the normal-half-normal SFM by Olson et al. (1980) uncovered an interesting
phenomena, quite regularly the corrected OLS estimator would produce an estimate of σu2 ≤
0. This was deemed a ‘Type I’ failure of the SFM; further Olson et al. (1980, pg. 70) noted
that “It is also true that, in every case of Type I failure we encountered, the MLE estimate
of [σu2 ] also turned out to equal zero. (This makes some sense, though we cannot prove
analytically that it should happen.)” Waldman (1982) provided the analytic foundation
behind this result, demonstrating that a stationary point of the log-likelihood function exists,
and this stationary point is a local maximum when the sign of the skewness of residuals
stemming from OLS estimation of the SFM is positive. This is broadly viewed as a deficiency
of the SFM as an estimate of σu2 of 0 is literally interpreted as a finding of no inefficiency.
However, this is an unfortunate interpretation because it is purely a finite sample issue. If
in fact u is distributed half-normal, then as shown in Uekusa & Torii (1985), Coelli (1995),
and Simar & Wilson (2010), as n → ∞ the likelihood of drawing a random sample which
will have positive skewness decreases, and the rate of this decrease is directly related to
σu2 /σv2 ; the larger this ratio the faster the decrease in the probability of observing a random
24 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

sample with positive skew.22 The observance of OLS residuals with positive skew is, by
and of itself, of no concern. What is concerning is that for an applied researcher whose
focus is to study the efficiency level of firms, analysis of a sample where the residuals from
the SFM have positive skewness leads to the conclusion of all firms being efficient and this
finding might be incongruent with either preconceptions about the industry or perceived
publication standards when applying these methods. This has often led to various forms
of respecification: using a different data set, trying an alternative functional form for the
production function, or most likely, deploying different distributional assumptions regarding
inefficiency.
As noted by Simar & Wilson (2010), none of these respecification approaches are ap-
propriate or warranted. Again, Table 1 in Simar & Wilson (2010) evinces that even when
everything about the SFM is correctly specified, positively skewed OLS residuals are still a
regular occurrence. Their suggestion is to use special resampling techniques based on boot-
strapping to conduct inference on either overall inefficiency of the industry under study or
specific firms. The finding of OLS residuals with positive skewness is commonly denoted
the ‘wrong skew problem’, though it is not clear where this term initially originated. It is
unfortunate that this term has crept into the lexicon of productivity analysis as there really
is no problem at all, except for the problem of misinterpretation and mistreatment.
One reason why respecification is troubling is that classical statistical inference assumes
that model specification is selected independently of estimation. When specification searches
are conducted this introduces biases into the final parameter estimates. Further, there is
the concern in published research that if the researcher did encounter positive skew, that
this information is not provided to the reader. It is worth mentioning that not all SFMs are
plagued by this issue. In fact, some distributional combinations will lead to identification
of inefficiency regardless of the sign of the skewness of the OLS residuals. Examples include

22Note that the estimator of the skewness coefficient is distributed asymptotically standard normal so it is
feasible to have either negative or positive skewness in any finite sample.
25

the normal-uniform SFM of Li (1996), the normal-Weibull SFM of Tsionas (2007), the
normal-binomial SFM of Carree (2002), and the normal-doubly truncated SFM of Almanidis
et al. (2014). Even more recently, Horrace & Parmeter (2014) demonstrated, in the style
of Waldman (1982) that the log-likelihood function of the Laplace-exponential SFM is not
dependent upon the sign of the skewness of the OLS residuals. This mainly stems from the
fact that as σu2 → 0 this model converges to a regression model with error term distributed
Laplace, for which the MLE is the least absolute deviations (LAD) estimator.
Despite the history behind the impact of the sign of the skewness of the OLS residuals
on the SFM, interest still abounds surrounding this issue. Recently, Hafner, Manner &
Simar (2016) presented a generalized method which always ensures that the SFM can be
identified, and that this model will converge to the traditional SFM model as n → ∞ if the
traditional SFM is correctly specified. Bonanno, De Giovanni & Domma (2017) introduced
a generalized SFM which allows v to be distributed as a Type 1 generalized logistic which
introduces asymmetry in v, coupled with allowing dependence between u and v. These two
additional assumptions, similarly to Hafner et al. (2016), allow the parameters of the SFM
to be identified regardless of the sign of the OLS residuals. Feng, Horrace & Wu (2015)
describe a constrained MLE that uses the traditional normal-half-normal distributional pair,
but imposes a penalty in estimation to combat the potential for positive skewness of the OLS
residuals to lead to an estimate of σu2 of 0. Finally, Horrace & Wright (2016) generalize the
theory of Waldman (1982) by studying the SFM without explicit distributional assumptions.
All told, this issue is one that still generates a substantial amount of interest in the academic
community and it is one that is not likely to fade any time soon (see the discussion in
Almanidis & Sickles 2011).

3. Handling Endogeneity in the SFM

A common assumption in the SFM is that x is either exogenous or independent of both


ui and vi . If either of these conditions are violated then the MLE will be biased and most
26 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

likely inconsistent. Yet, it is not difficult to think of settings where endogeneity is likely
to exist. For example, if shocks are observed before inputs are chosen, then producers may
respond to good or bad shocks by adjusting inputs, leading to correlation between x and
v. Alternatively, if managers know they are inefficient, they may use this information to
guide their level of inputs, again, producing endogeneity. In a regression model, dealing
with endogeneity is well understood. However, in the composed error setting, these methods
cannot be simply transferred over, but require care in how they are implemented (Amsler,
Prokhorov & Schmidt 2016).
To incorporate endogeneity into the SFM in (2.1), we set m(xi ; β) = β0 + x01i β1 + x02i β2
where x1 are our exogenous inputs, and x2 are the endogenous inputs, where endogeneity
may arise through correlation of x2 with u, v or both. To deal with endogeneity we require
instruments, w, and identification necessitates that the dimension of w is at least as large
as the dimension of x2 . The natural assumption for valid instrumentation is that w is
independent of both u and v. Our following discussion here will center on the distributional
assumptions of ALS.
Why worry about endogeneity? Economic endogeneity means that the inputs in question
are choice variables and chosen to optimize some objective function such as cost minimization
or profit maximization. Statistical endogeneity arises from simultaneity, omitted variables,
and measurement errors. For example if the omitted variable is managerial ability, which
is part of inefficiency, inefficiency is likely to be correlated with inputs because managerial
ability affects inputs. This is the Mundlak argument for why omitting a management quality
variable (for us inefficiency) will cause biased parameter estimates. Endogeneity can also
be caused by simultaneity meaning that more than one variable in the model are jointly
determined.
One way to address the problem is to look at it from a purely statistical angle and use
instrumental variables. The other solution is economic, that is address the economic issue
that is causing endogeneity. We consider first the statistical solution and then the economic
27

solution. In many applied settings it is not clear what researchers mean when they attempt to
handle endogeneity inside the SFM. An excellent introduction into the myriad of influences
that endogeneity can have on the estimates stemming from the SFM can be found in Mutter,
Greene, Spector, Rosko & Mukamel (2013). Mutter et al. (2013) used simulations designed
around data based on the California nursing home industry to understand the impact of
endogeneity of nursing home quality on inefficiency measurement.

3.1. A Corrected Two Stage Least Squares Approach. The simplest approach to
accounting for endogeneity is to use a corrected two stage least squares (C2SLS) approach,
similar to the common COLS approach that has been used to estimate the SFM. This method
estimates the SFM using standard 2SLS with instruments w. This produces consistent
estimators for β1 and β2 but not β0 , as this is obscured by the presence of E[u] (to ensure
that the residuals have mean zero). The second and third moments of the 2SLS residuals
bu2 is determined, the intercept can
are then used to recover estimators of σv2 and σu2 . Once σ
q
be corrected by adding π2 σ̂u .
This represents a simple avenue to account for endogeneity, and it does not require spec-
ifying how endogeneity enters the model, i.e. through correlation with v, with u or both.
However, as with other corrected procedures based on calculations of the second and third (or
even higher) moments of the residuals, from Olson et al. (1980) and Waldman (1982), if the
initial 2SLS residuals have positive skew (instead of negative), then σu2 cannot be identified
and its estimator is 0. Further, the standard errors from this approach need to be modified
for the estimator of the intercept to account for the step-wise nature of the estimation.

3.2. A Likelihood Approach. The SFM with endogeneity has recently been studied by
Kutlu (2010), ?, Tran & Tsionas (2013), and Amsler et al. (2016). Here we describe maximum
likelihood estimation of the SFM under endogeneity. Our discussion here follows Amsler
et al. (2016) as their derivation of the likelihood relies on a simple conditioning argument as
opposed to the earlier work relying on the Cholesky decomposition. While both approaches
28 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

lead to the same likelihood function, the conditioning idea of Amsler et al. (2016) is simpler
and more intuitive.
Consider the stochastic frontier system:

(3.1) yi =xi β + εi

(3.2) x2i =wi Γ + ηi

where xi = (x1i , x2i ), β = (β1 , β2 ), wi = (x1i , qi ) is the vector of instruments, ηi is


uncorrelated with wi and endogeneity of x2i arises through cov(εi , ηi ) 6= 0. Here simultaneity
bias (and the resulting inconsistency) exists because ηi is correlated with either vi , ui or both.
The following assumptions are used by Amsler et al. (2016): ui ∼ N+ (0, σu2 ), m(x1i , x2i ; β1 , β2 ) =
β0 + x1i β1 + x2i β2 , and conditional on wi , ψi = (vi , ηi )0 ∼ N (0, Ω), where
 
σv2 Σvη
Ω= .
Σηv Σηη ηi
 
vi
Amsler et al. (2016) focused on the setting where ui is independent of ψi =  . To
ηi
derive the likelihood function, Amsler et al. (2016) condition on the instruments, w. Doing
this yields f (y, x2 |w) = f (y|x2 , w) · f (x2 |w). With the density in this form, the log-
likelihood follows suite: ln L = ln L1 + ln L2 , where ln L1 corresponds to f (y|x2 , w) and
ln L2 corresponds to f (x2 |w). These two components can be written as
n n
2 1 X 2 X
ln L1 = − (n/2) ln σ − 2 ε̃ + ln [Φ (−λc ε̃i /σ)]
2σ i=1 i i=1
n
X
ln L2 = − (n/2) ln |Σηη | − 0.5 ηi0 Σ−1
ηη ηi ,
i=1

where ε̃i = yi − β0 − xi β − µci , µci = Σvη Σ−1 2 2 2 2


ηη ηi , σ = σv + σu , λc = σu /σc and σc =

σv2 − Σvη Σ−1


ηη Σηv . The subtraction of µci in ln L1 is an endogeneity correction while it should
29

be noted that ln L2 is nothing more than the standard likelihood function of a multivariate
normal regression model (as in (3.1)). Estimates of the model parameters (β, σv2 , σu2 , Γ, Σvη )
and Σηη can be obtained by maximizing the likelihood function ln L.
While direct estimation of the likelihood function is possible, a two-step approach is also
available (Kutlu 2010). However, as pointed out by both Kutlu (2010) and Amsler et al.
(2016), this two-step approach will have incorrect standard errors. Even though the two-step
approach might be computationally simpler, it is, in general, different from full optimization
of the likelihood function of Amsler et al. (2016). This is due to the fact that the two-step
approach ignores the information provided by Γ and Σηη in ln L1 . In general full optimization
of the likelihood function is recommended as the standard errors (obtained in a usual manner
from the inverse of the Fisher information matrix) are valid.23

3.3. A Method of Moments Approach. An insightful avenue to deal with endogeneity


in the SFM that differs from the traditional corrected methods or maximum likelihood is
proposed by Amsler et al. (2016), who used the work of Hansen, McDonald & Newey (2010).
The idea is to use the first order conditions for maximization of the likelihood function under
exogeneity:

E ε22 /σ 2 − 1 = 0
 
(3.3)
 
εi φi
(3.4) E =0
1 − Ψi
 
φi
(3.5) E xi εi /σ + λxi = 0,
1 − Φi
23Typically the standard errors can be obtained either through use of the outer product of gradients (OPG)
or direct estimation of the Hessian matrix of the log-likelihood function. Given the nascency of these methods
it has yet to be determined which of these two methods is more reliable in practice, though in other settings
both tend to work well. One caveat for promoting the use of the OPG is that since this only requires
calculation of the first derivatives, it can be more stable (and more likely to be invertible) than calculation
of the Hessian. Also note that in finite samples, the different estimators of covariance of MLE estimator can
give different numerical estimates, even suggesting different implications on the inference (reject or do not
reject the null hypothesis). So, for small samples, it is often advised to check all feasible estimates whenever
there is suspicion of ambiguity in the conclusions (e.g., when a hypothesis is rejected only at say around the
10% of significance level).
30 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

where φi = φ( λεσ i ) and Φi = Φ( λεσ i ). Note that these expectations are taken over xi and yi
(and by default, εi ) and solved for the parameters of the SFM.
The key here is that these first order conditions (one for σ 2 , one for λ and the vector for
β) are valid under exogeneity and this implies that the maximum likelihood estimator is the
generalized methods of moments estimator. Under endogeneity however, this relationship
does not hold directly. But the seminal idea of Amsler et al. (2016) is that the first order
conditions (3.3) and (3.4) are based on the distributional assumptions on v and u, not on
the relationship of x with v and/or u. Thus, these moment conditions are valid whether
x contains endogenous components or not. The only moment condition that needs to be
adjusted is (3.5). In this case the first order needs to be taken with respect to w, the
exogenous variable, not x. Doing so results in the following amended first order condition:
 
φi
(3.6) E wi εi /σ + λwi = 0,
1 − Φi

where φi and Φi are identical to those in (3.5). It is important to acknowledge that this
moment condition is valid when εi and wi are independent. This is a more stringent require-
ment than the typical regression setup with E[εi |wi ] = 0. As with the C2SLS approach, the
source of endogeneity for x2 does not need to be specified (through v and/or u).

3.4. Estimation of Individual Inefficiency. An interesting, and important finding from


Amsler et al. (2016) is that when there is endogeneity, one can potentially improve estimation
of inefficiency through the JLMS estimator. The traditional predictor of Jondrow et al.
(1982) is E(ui |εi ). However, more information is available when endogeneity is present,
namely via ηi . This calls for a modified JLMS estimator, E(ui |εi , ηi ). Note that even
though it is assumed that ui is independent from ηi (as in Amsler et al. 2016), because ηi
is correlated with vi , there is information that can be used to help predict ui even after
conditioning on εi .
31

Amsler et al. (2016) showed that ηi is independent of (ui , ε̃i ):

E(ui |εi , ηi ) = E(ui |ε̃i , ηi ) = E(ui |ε̃i ).

and that the distribution of ui conditional on ε̃i = yi − β0 − xi β − µci is N+ (µ∗ , σ∗2 ) with
µ∗ = −σu2 ε̃i /σ 2 and σ∗2 = σu2 σc2 /σ2 , which is identical to the original JLMS estimator, except
that σv2 is replaced with σc2 and ε̃i taking the place of εi . The modified JLMS estimator in
 
φ(ξi )
the presence of endogeneity becomes E(ui |εi , ηi ) = σ∗ 1−Φ(ξi ) − ξi with ξi = λε̃i /σ. Note
that E(ui |εi , ηi ) is a better predictor than E(ui |εi ) because σc2 < σv2 . The improvement in
prediction follows from the textbook identity for variances, where for any random vector
(X, Z), where X and Z are random sub-vectors, we have

var(X) = var[E(X|Z)] + E(var[X|Z]) .


| {z } | {z }
Explained U nexplained

In this case, by conditioning on both εi and ηi the conditioning set is larger than simply
conditioning on εi and so it must hold that the unexplained portion of E(ui |εi , ηi ) is smaller
than that of E(ui |εi ). It then holds that there is less variation in E(ui |εi , ηi ) as a predictor
than E(ui |εi ), which is a good thing. While it is not obvious at first glance, one benefit of
endogeneity is that researchers may be able to more accurately predict firm level inefficiency,
though it comes at the expense of having to deal with endogeneity. This improvement in
prediction may also be accompanied by narrower prediction intervals, however, this is not
known as Amsler et al. (2016) did not study the prediction intervals.

3.5. An Economic Approach to Deal with Endogeneity. An alternative to developing


valid instruments and correcting for endogeneity is to use what is known as a primal system
approach, when inputs are endogenous (Kumbhakar et al. 2015, chapt. 8). This setup
estimates the traditional SFM but appends the first order conditions stemming from cost
minimization (one could alternatively attach profit maximization or return to the outlay
conditions instead if this was a more representative behavior for the industry under study).
32 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

That is, if a producer minimizes costs24

(3.7) min p0 x, s.t. y = m(x; β) + v − u,

for input prices p, the first order conditions in this case are

mj (x; β) pj
(3.8) = , j = 2, . . . , J,
m1 (x; β) p1

where mj (x; β) is the partial derivative of m(x; β) with respect to xj . These first order
conditions are exact, which usually does not arise in practice, rather, a stochastic term is
added, which is designed to capture allocative inefficiency. That is, our empirical first order
mj (x;β) pj ξj
conditions are m1 (x;β)
= p1
e for j = 2, . . . , J where eξj captures allocative inefficiency
for the j th input relative to input 1 (the choice of input to compare to is without loss of
generality). The idea behind allocative inefficiency is that firms could be fully technically
efficient, and still have room for improvement due to over or under use of inputs, relative to
another input, given the price ratio. In general if firms are cost minimizers and one estimates
a production function, the inputs will be endogenous as these are choice variables to the firm.
Hence, a different approach is needed.
The primal system approach estimates the SFM as in (2.1) but also incorporates the
information in the J − 1 conditions in (3.8) with allocative inefficiency built in. Shephard’s
lemma in microeconomics dictates that the first order conditions are actually cost share
information, when the logarithm of the production function is taken, the first derivatives
represent the cost shares of the corresponding inputs,
∂ ln m
mj (x; β) ∂ ln xj sj /xj
(3.9) = ∂ ln m
= .
m1 (x; β) ∂ ln x1
s1 /x1

24Itis possible to treat a subset of x as endogenous; i.e., x = (x1 , x2 ), where x1 is endogenous and x2 is
exogenous.
33

sj /xj pj ξj
When these are equated to the ratio of input prices, one obtains s1 /x1
= p1
e , which can be
sj pj xj ξj
rearranged to yield s1
= p 1 x1
e . Taking logarithms produces

(3.10) ln(sj ) − ln(s1 ) − ln(pj xj ) + ln(p1 x1 ) = ξj .

If distributional assumptions are imposed on v, u and ξ, the parameters of the production


function can be estimated along with technical and allocative efficiency. An unfortunate
consequence of the primal system approach is that the input demand and cost functions are
analytically tractable only for quite specific assumptions on the production function (Cobb-
Douglas being one). In these cases a more complicated process is required to determine
the impact of technical and allocative inefficiency on costs (Kumbhakar & Wang 2006).
See Kumbhakar (2011, 2013) for more detailed discussion of these types of primal system
approaches to handle economic endogeneity across a range of settings.

4. Modeling Determinants of Inefficiency

Use of the SFM is exciting for productivity analysis because a prediction of firm level
efficiency can be obtained. However, in the benchmark SFM, ui is treated as completely
random, and so nothing connects the level of inefficiency to variables which might serve as an
explanation for the existence and the level of inefficiency. As the SFM has gained popularity
in applied productivity analysis, it has become common to introduce variables outside the
main production structure which influence output through their effect on inefficiency.25
As a concrete example, consider the study of productivity within the banking industry.
A researcher may want to know whether a bank’s level of efficiency is affected by the use of
information technology, the amount of assets the bank has access to, the type of bank, or
the type of ownership structure in place, corporate governance practices, etc. Similarly, the
25Reifschneider & Stevenson (1991) used the term ‘inefficiency explanatory variables’, while others call them
‘environmental variables’, but it is now common to refer to these variables as ‘determinants of inefficiency.’ A
variety of approaches have been proposed to model the determinants of inefficiency with the first pertaining
to panel data models (Kumbhakar 1987, Battese & Coelli 1992) (see Section 5).
34 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

government might be interested in whether regulations (such as allowing banks to merge)


improve banks’ performance. To answer these questions, the relationship between efficiency
and its potential determinants needs to be modeled and estimated.
Consider estimating what influences firm level inefficiency in the benchmark SFM. This
model assumes that both vi and ui are homoskedastic. In a traditional linear regression,
heteroskedasticity has no impact on the bias/consistency of the OLS estimator. However, if
we were to allow σu2 to depend on determinants of inefficiency, z, then ignoring this will lead
to, except in special settings, a biased and inconsistent estimator of the parameters of the
SFM. Both Kumbhakar & Lovell (2000, Section 3.4) and Wang & Schmidt (2002) provide
detailed accounts of the consequences of ignoring the presence of determinants of inefficiency
in the SFM.
p
Recall from Section 2, that E[u] = 2/πσu . Now imagine ignoring the composed structure
of ε and estimating the SFM via OLS. If it is the case that determinants of inefficiency are
present, so that σu2 = σu2 (z), this omission leads to biased parameter estimates of the SFM
given that the assumed model is

p
yi = m(xi ; β) + 2/πσu + ε∗i ,

p
with ε∗i = εi − 2/πσu , whereas the true model is

p
yi = m(xi ; β) + 2/πσu (zi ) + ε∗i ≡ m̃(xi , zi ; β, δ) + ε∗i .

The estimates of m(xi ; β) are conflated with σu (zi ), unless x and z are uncorrelated. The
reason that this issue presents itself is the fact that the mean of u, due to the truncation
at 0, must depend on the variance. Thus, it is not possible to allow u to be heteroskedastic
without the mean of u being a function of z as well. Notice here that we have specifically
separated the impacts of x and z on output, with x capturing pure production and z
capturing inefficiency. This is commonly known as the separability assumption. In some
35

settings this assumption does not have to be made, but in other settings it is a necessity
for identification. See Parmeter & Zelenyuk (2016) for a more detailed discussion of the
separability assumption. Our use of it here is more for expositional clarity.
Exactly how to model the influence of z on inefficiency is unknown and at various points
in time practitioners have deployed a simpler, two step analysis to account for the presence
of determinants of inefficiency. This approach constructs JLMS predictions in the first step,
and then regresses these inefficiency estimates on z in the second step. Pitt & Lee (1981)
were the first to implement this type of approach (in a panel data setting) and many others
followed this two-step approach blindly (Ali & Flinn 1989, Kalirajan 1990, Bravo-Ureta &
Rieger 1991). However, this route to modeling determinants of inefficiency has been met
with criticism repeatedly, and for good reason.
As explained in Battese & Coelli (1995), the first stage model is misspecified if z is ignored.
Further, Wang & Schmidt (2002) note that if x and z are correlated then an omitted variable
bias exists in the the first step rendering the second step ineffectual. Even in the special case
where x and z are uncorrelated, ignoring the dependence of u on z will lead to the estimated
JLMS predictions in the first stage to have too little variation (see also Schmidt 2011) and,
subsequently, the estimator in the second stage regression will be biased downward. Caudill
& Ford (1993) provide Monte Carlo evidence on the effects that ignoring the impact of z on
u has on the estimator of the parameters of the SFM while Wang & Schmidt (2002) provide
a detailed analysis of the bias of the second stage parameter estimators.
As should be clear, the two stage approach to account for determinants of inefficiency in
the SFM has no statistical foundation and is widely agreed upon to yield poor insights on
the actual behavior of inefficiency, as such this approach should be strictly avoided; even
with these criticisms of the two-step approach, one will occasionally happen across research
that adopts this flawed two-step methodology.
While the two stage approach has undesirable statistical properties this does not mean
that determinants of inefficiency cannot be accounted for. Quite the contrary. The preferred
36 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

approach to studying the exogenous influences on efficiency is a single-step procedure that


explicitly accounts for z.

4.1. Proper Modeling of the Determinants of Inefficiency. The first proper proposals
to model z in the SFM are Kumbhakar, Ghosh & McGuckin (1991) and Reifschneider &
Stevenson (1991), who used the normal truncated-normal SFM as the basis for estimation.26
While their focus was on the normal truncated-normal SFM, the key insights hold for the
normal-half-normal SFM, which is what we will base our discussion on here. The main idea
is to specify σu2 as a parametric function of z.27 Formally, their parameterization of σu2 is

0
(4.1) σu2 = ez δ ,

The log-likelihood function of the heteroskedastic model is the same as in (2.4), except that
we replace σu2 with (4.1).28 Here all of the model parameters are estimated simultaneously
and once they are found, technical inefficiency can be computed using (2.13) or (2.14) with
the appropriate form of σu2 substituted into the expressions.
If u follows a half-normal distribution, with the σu2 function depending upon z then the
mean of ui is

p 0 1 0
(4.2) E[ui |zi ] = 2/πezi δ = e 2 ln(2/π)+zi δ .

Note that the 1


2
ln(2/π) term can be absorbed by the constant term in zi0 δ. Therefore, by
parameterizing σu2 , we allow z to affect the expected value of inefficiency. More importantly,
however, is that the parameterization (4.1) produces maximum likelihood estimates of δ
which may not be very informative. This is because E[ui |zi ] is nonlinear in z, and therefore
26Caudill & Ford (1993), Huang & Liu (1994), Battese & Coelli (1995), Caudill, Ford & Gropper (1995),
Hadri (1999), and Wang (2002) present alternative specifications as well.
27It is also possible to model σ 2 as a function of variables, but this poses fewer problems and we omit the
v
details here. See Parmeter & Kumbhakar (2014) and Simar, Van Keilegom & Zelenyuk (2017) for more
discussion.
28Actually, given the reparameterization of the log-likelihood function, the specification for σ implies a
u
particular specification for both λ and σ.
37

the slope coefficients δ are not the marginal effects of z. For instance, assume the jth
variable in z has an estimated coefficient of 0.5. This number itself tells us very little about
the magnitude of the jth variable’s (marginal) effect on the inefficiency, though it does tell
us the direction of the effect on inefficiency. Also, the nonlinearity of the conditional mean
of u, implies that for different levels of z, there will be different expected levels of u. In these
instances the marginal effect of z may be useful for empirical purposes.
For the given parameterization of the normal-half-normal SFM, the marginal effect of the
jth variable of zi , zji on E[ui |zi ] is

∂E[ui |zi ] p
(4.3) = δj 2/πσu,i
∂zj
p
where 2/π is approximately 0.80. It is clear that (4.3) also implies
 
∂E[ui |zi ]
(4.4) sign = sign(δk )
∂zj

so that the sign of the coefficient reveals the direction of impact of zji on E[ui |zi ]. This
property does not always hold across distributional assumptions, for example in the normal-
truncated-normal SFM the sign of the coefficient cannot be interpreted directly (Parmeter &
Kumbhakar 2014). In general, only in one parameter families for the pdf of u (exponential,
half-normal, etc.) does this correspondence hold; this suggests caution in directly interpreting
the impact that a particular variable zj has on inefficiency based purely on the sign of δj .
The nonlinear nature of the relationship of E[u|z] with z implies that for a sample of n
observations we have n marginal effects for each variable. A concise statistic to present is
the average partial effect (APE) on inefficiency or the partial effect of the average (PEA):
n
!
p X
u −1
(4.5) AP E(zj ) =(δj 2/π) n σu,i
i=1
0
p
(4.6) P EA(zj ) =δju 2/πez̄u δ .
38 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

Either of these measures can be used to provide an overall sense for the impact of a given
variable on the level of inefficiency. However, these statistics should also be interpreted with
care. Neither necessarily reflects the impact of a given covariate for a given firm, but rather
on average and ceteris paribus, i.e., holding other covariates fixed; for example, it could be
that half of the sample has a very negative effect that is balanced by positive effects in the
other half of the sample, thus getting nearly zero on average, which might misrepresent the
phenomenon. It is also possible to standardize further by using elasticities, which will cancel
p
out the 2/π term, this occurs when the variables are measured in logarithms. It could also
prove useful to present the estimates of these at either quartiles or at particular points of
interest suggested by a particular empirical context (for example a specific regulation output
target).

4.2. Incorporating Determinants when u is Truncated-Normal. As we have dis-


cussed earlier, the truncated-normal distribution offers greater flexibility to model an array
of shapes of the true, but unknown, distribution of u. When determinants of inefficiency
are present and one elects to assume the truncated-normal distribution, several additional
modeling choices become available to the researcher. These additional choices are important
because, as with the choice of distributional assumption, there is typically little guidance on
how best to incorporate the determinants.
What do we mean? Consider again the truncated-normal density that would be assumed
for u, when determinants of inefficiency are present:

(u−µ(z;δ2 ))2
1 −
(4.7) f (u) = √ e 2σu (z;δ1 )2 , u ≥ 0.
2πσu (z; δ1 )Φ (µ(z; δ2 )/σu (z; δ1 ))

In this case the impact of z on u can be modeled through the pre-truncation mean, µ and
the pre-truncation standard deviation, σu . The issue with where to assume that z influences
u is that modeling either parameter as a function of z impacts all of the moments of u,
due to the truncation. Consider the conditional (on z) mean of a truncated normal random
39

variable

φ( σµ(z;δ 2)
" #
µ(z; δ2 ) u (z;δ1 )
)
(4.8) E[u|z] = σu (z; δ1 ) + .
σu (z; δ1 ) Φ( µ(z;δ2 ) )
σu (z;δ1 )

Regardless of whether σu or µ is constant, z still influences the mean of inefficiency unless


both are constant. This is what makes the choice of where to incorporate z abstruse when
using the truncated normal distribution. Parametric specification of either σu or µ will allow
for z to influence expected inefficiency, but in different manners, and, in nonlinear fashion.
Given that µ can be positive or negative, it is common to model it in a linear fashion, i.e.
0
µ(z; δ2 ) = z 0 δ2 and to model σu (z; δ1 ) as ez δ1 , to ensure positivity of the pre-truncation
standard deviation.
When we assume that u has the half-normal distribution our choice is easy because only
a single parameter exists and it is clear where z enters. However, in the truncated-normal
setup we could elect to have z enter only through the pre-truncation mean, only through the
pre-truncation standard deviation, or both. In fact, various applied papers have used any
of these three approaches. Kumbhakar et al. (1991) and Reifschneider & Stevenson (1991),
modeled the impact of determinants of inefficiency through µ,29 while Caudill & Ford (1993)
incorporated determinants through σu .30 Lastly, Wang (2002) modeled the determinants
through both µ and σu . The benefit of modeling both pre-truncation parameters jointly as
functions of z is that this leaves little room for ambiguity and makes inference of where z
belongs a viable option. The costs are that the model is more complex to estimate and may
lead to identification problems, as raised in Ritter & Simar (1997). An alternative approach,
which we discuss next, is to invoke a special assumption on the distribution that makes it
more amenable to modeling the influence of determinants of inefficiency in the SFM.

4.3. The Scaling Property. Many of the main proposals to incorporate determinants of in-
efficiency did so through the normal truncated-normal SFM. The two parameter nature of the
29See also Huang & Liu (1994) and Battese & Coelli (1995) for early approaches following this strategy.
30Other early approaches that followed this route include Caudill et al. (1995) and Hadri (1999).
40 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

truncated-normal distribution implies that determinants could influence the pre-truncation


mean, µ, the pre-truncation variance, σu2 , or both. Further still, different variables could
influence each parameter.
A popular simplification (Simar, Lovell & van den Eeckaut 1994, Wang & Schmidt 2002),
which encapsulates the normal-half-normal SFM, is to assume that inefficiency behaves as

(4.9) ui ∼ g(zi ; δ) u∗i ,

where g(·) ≥ 0 is a function of the exogenous variables while u∗i ≥ 0 is a random variable.
This behavior is known as the scaling property. Single parameter distributions, such as
the half-normal and the exponential, automatically possess this property, but more flexible
distributions, such as truncated-normal or gamma, can have this property imposed. The key
feature of the scaling property is that u∗i does not depend on zi in any fashion; u∗i is known
as base inefficiency (Wang & Schmidt 2002, Alvarez, Amsler, Orea & Schmidt 2006).
When a distribution possesses the scaling property the shape of the distribution of ui is
the same for all firms, which can be viewed as an attractive feature. The scaling function,
g(·), expands or contracts the horizontal axis so that the scale of the distribution of ui
changes while preserving the underlying shape of the distribution. In comparison, the normal
truncated-normal SFM models allow different scalings for each ui , so that for some firms the
distribution of inefficiency is close to a normal (if the pre-truncation mean is large), while
for other firms the distribution of inefficiency is the extreme right tail of a normal with a
mode of zero (if the pre-truncation mean is negative). In comparison, for a model with the
scaling property the mean and the standard deviation of u change with zi , but the shape of
the distribution is fixed.
41

Another advantage of the scaling property specification is the ease of interpretation of δ


0
when g(zi , δ) = ezi δ ,

∂ ln E[ui |z]
(4.10) = δj .
∂zj

That is, δj is the semi-elasticity (or elasticity if z is already measured on the logarithmic
scale) of expected inefficiency with respect to the jth element of z, and more importantly, this
interpretation is distinct from any distributional assumption placed on u∗ . An interpretation
of this ilk is generally not available in other model specifications. Further, the sign of the
elements of δ can be directly interpreted.
The scaling property provides an attractive economic interpretation as well. u∗ can be
interpreted as a benchmark level of inefficiency of the firm (Alvarez et al. 2006). The scaling
function then allows a firm to exploit (or fail to exploit) these talents through other variables,
z, which might include experience of the plant manager, the operating environment of the
firm, or regulatory restrictions.
The scaling property is not a fundamental feature, rather, as with the choice of distribution
on u, it is an assumption on the features of the inefficiency distribution. As such it can be
tested against models that do not possess this property for the inefficiency distribution. As it
currently stands, all tests of the scaling property hinge on a given distributional assumption,
for example, estimating the normal truncated-normal SFM and then estimating a restricted
version of the same model, but imposing the scaling property. An important avenue for future
research is the development of a test (or tests) that does not require specific distributional
assumptions.

4.4. Estimation Without Imposing Distributional Assumptions. In settings where


the researcher is comfortable with imposing the scaling property on the distribution of in-
efficiency, the SFM can be estimated without distributional assumptions. This is perhaps
42 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

the key benefit of invoking the scaling property. To understand how it is possible to esti-
mate the SFM without distributional assumptions we expound on the discussion of Simar
et al. (1994), Wang & Schmidt (2002), and Alvarez et al. (2006). The SFM with the scaling
property can be written as31

0
(4.11) yi = m(xi ; β) + vi − ezi δ u∗i .

The conditional mean of y given x and z is

0
(4.12) E[y|x, z] = x0 β − ez δ µ∗

where µ∗ = E[u∗ ] and E[v|x, z] = 0. The SFM is then

0 0 0
(4.13) yi = m(xi ; β) − ezi δ µ∗ + vi − ezi δ (ui − µ∗ ) = m(xi ; β) − ezi δ µ∗ + ε∗i ,

0
with ε∗i = vi −ezi δ (ui −µ∗ ), which, for a given parameterization of m(xi ; β), can be estimated
using nonlinear least squares (NLS) as
  n h i2
0
X
∗ −1
(4.14) β, δ, µ
b b b = min∗ n yi − m(xi ; β) + µ∗ ezi δ .
β,δ,µ
i=1

The elegance of invoking the scaling property is that the SFM can be estimated in a distri-
bution free manner via NLS; the need for NLS stems from the fact that the scaling function
must be positive and if it was specified as linear this would be inconsistent with theoretical
requirements on the variance of the distribution.
Direct NLS will produce a consistent estimator of all of the terms of the SFM. However,
the error term ε∗i is heteroskedastic,

0
var(ε∗i |xi , zi ) = σv2 + σu2∗ e2zi δ ,

31Note here that we are making the implicit assumption that z is different from x. The nonlinearity of the
scaling function does allow z and x to overlap however.
43

where σv2 = var(vi ) and σu2∗ = var(u∗i ). As such, a generalized NLS estimator would be called
for to produce an efficient estimator (as similar to the MLE). Unfortunately, a generalized
NLS algorithm hinges on distributional assumptions to appropriately separate σv2 and σu2∗ . An
alternative, which allows valid inference to be undertaken, is to compute heteroskedasticity
robust standard errors for β and δ (White 1980).
An interesting extension of this idea was recently proposed by Paul & Shankar (2017) for
the setting where ui has already been converted into technical efficiency. In this case the
level of inefficiency must be bound between 0 and 1. To account for this Paul & Shankar
(2017) model the impact of z on the level of inefficiency through a probit function. Again,
given the nonlinear nature of the probit function, this necessitates use of NLS if one wishes
to eschew distributional assumptions.
With the wide range of statistical software that can quickly implement an NLS problem, it
is perhaps surprising that this avenue has not been exploited in applied research. Certainly
the scaling property is an assumption that requires judicious justification, but not more so
than distributional assumptions imposed on the composed error structure of the SFM. It is
also possible in this nonlinear setup, that the calculation of expected firm efficiency can be
done without requiring distributional assumptions, leading to the potential for more robust
conclusions regarding observation specific inefficiency. It is also possible to estimate the SFM
in (4.13) without imposing assumptions on the scaling function, an issue we will discuss in
Section 6.
Currently no test of the scaling property exists without enforcing distributional assump-
tions. Alvarez et al. (2006) proposed standard tests of the scaling property by using the
nesting structure of the normal-truncated-normal distributional pair against the normal-
half-normal distributional pair. Unfortunately this testing facility requires distributional
assumptions on both vi and ui . A test of the statistical significance of the determinants
of inefficiency, using the NLS framework just described is available (Kim & Schmidt 2008).
44 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

Under H0 : δ = 0 it follows that

0 0
yi = m(xi ; β) − µ∗ ezi δ + ε∗i = m∗ (xi ; β) − µ∗ (1 − ezi δ ) + ε∗i ,

where m∗ (xi ; β) = m(xi ; β) + c for a constant c. That is, one can only identify µ if at least
0
one element of δ is nonzero; note that under H0 : 1 − ezi δ = 0, µ∗ cannot be separately
identified. This lack of identification creates issues for inference under the null hypothesis,
and invalidates the common asymptotic behavior of Wald and likelihood ratio tests. The
solution, which Kim & Schmidt (2008) proposed to avoid this problem, is to use the Lagrange
Multiplier (LM) test which involves estimation imposing the null hypothesis. A novel insight
of Kim & Schmidt (2008) is that the LM test they proposed has power in directions where
the scaling property does not hold. This is due to the fact that the model being tested H0 is
indifferent to ‘how’ inefficiency enters the model. Thus, while an explicit test of the scaling
property without requiring distributional assumptions would be a useful tool, the Kim &
Schmidt (2008) LM test is likely to be sufficient.
The LM test is based on the derivative of the NLS criterion function in (4.14) with respect
to δ, evaluated at the restricted estimates (δ = 0):
n
2X
(4.15) (yi − m(xi ; β) + µ∗ ) (µ∗ zi ) .
n i=1

The test statistic is designed to determine how close the derivative of the NLS objection
function (with respect to the parameters under the null hypothesis) is to 0. If the parameter
restrictions are true then this should be close to 0. The reason that distributional assump-
tions are not needed for this test to work properly is that this test is identical to an F -test,
and F -tests are invariant to the scale of the covariates (Kim & Schmidt 2008). Thus, one
can simply set µ∗ = 1 and use NLS to regress y on (x, z) and test the significance of δ.
45

4.5. Estimation When Determinants of Efficiency and Endogeneity Are Present.


Quite recently, attention has focused on estimation of the SFM when some of the deter-
minants of inefficiency may be endogenous (Amsler, Prokhorov & Schmidt 2017, Latruffe,
Bravo-Ureta, Carpentier, Desjeux & Moreira 2017). These models can be estimated using
traditional instrumental variables methods. However, given that the determinants of ineffi-
ciency enter the model nonlinearly, nonlinear methods are required. To begin, we consider
the model of Amsler, Prokhorov & Schmidt (2017),

0
(4.16) yi = x0i β + vi − ui = x0i β + vi − u∗i ezi δ ,

where the scaling property has been invoked. The covariates xi and zi are partitioned as
   
x1i z1i
xi =  , zi =  ,
x2i z2i

where x1i and z1i are exogenous and x2i and z2i are endogenous. The set of instruments
used to combat endogeneity are defined as
 
x1i
 
wi =  z1i ,
 
 
qi

where qi are the traditional outside instruments. Identification of all the parameters requires
that the dimension of q be at least as large as the dimension of x2 plus the dimension of z2
(the rank condition).
In the model of Amsler, Prokhorov & Schmidt (2017), endogeneity arises through corre-
lation between a variable in the model (x2 and/or z2 ) and noise, v. That is, both x and
z are assumed to be independent of basic inefficiency u∗ . Given that E[ui ] is not constant,
the COLS approach to deal with endogeneity proposed by Amsler et al. (2016) cannot be
used here. To develop an appropriate estimator, add and subtract the mean of inefficiency
46 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

to produce a composed error term that has mean 0,

0 0
(4.17) yi = x0i β − µ∗ ezi δ + vi − (u∗i − µ∗ )ezi δ .

Proper estimation through instrumental variables requires that the moment condition
h 0
i
(4.18) E vi − (u∗i − µ∗ )ezi δ |wi = 0.

The nonlinearity of these moment conditions would necessitate use of nonlinear two stage
least squares (NL2SLS) (Amemiya 1974).
Latruffe et al. (2017) have a similar setup as Amsler, Prokhorov & Schmidt (2017), using
the model in (4.16), but develop a four step estimator for the parameters; additionally, only
x2 is treated as endogenous. Latruffe et al.’s (2017) approach is based on Chamberlain (1987)
using the construction of efficient moment conditions. The vector of instruments proposed
in Latruffe et al. (2017) is defined as
 
x1i
 
(4.19) wi (γ, δ) =  qi0 γ ,
 
 
0
zi ezi δ

where qi0 γ captures the linear projection of x2 on the external instruments q. The four-stage
estimator is defined as

Step 1: Regress x2 on q to estimate γ. Denote the OLS estimator of γ as γ


b.
Step 2: Use NLS to estimate the SFM in (4.16). Denote the NLS estimates of (β, δ) as
(β̈, δ̈). Use the NLS estimate of δ and the OLS estimate of γ in Step 1 to construct
the instruments wi (b
γ , δ̈).
Step 3: Using the estimated instrument vector wi (b
γ , δ̈), calculate the NL2SLS esti-
mator of (β, δ) as (β,
e δ).
e Use the NL2SLS estimate of δ and the OLS estimate of γ

in Step 1 to construct the instruments wi (b


γ , δ).
e
47

Step 4: Using the estimated instrument vector wi (b


γ , δ),
e calculate the NL2SLS esti-

mator of (β, δ) as (β,


b δ).
b

This multi-step estimator is necessary in the context of efficient moments because the
actual set of instruments is not used directly, rather wi (γ, δ) is used, and this instrument
vector requires estimates of γ and δ. The first two steps of the algorithm are designed to
construct estimates of these two unknown parameter vectors. The third step then is designed
to construct a consistent estimator of wi (γ, δ), which is not done in Step 2 given that the
endogeneity of x2 is ignored (note that NLS is used as opposed to NL2SLS). The iteration
from Step 2 to Step 3 does produce a consistent estimator of wi (γ, δ), and as such, Step
4 produces consistent estimators for β and δ). While Latruffe et al. (2017) proposed a set
of efficient moment conditions to handle endogeneity, the model of Amsler, Prokhorov &
Schmidt (2017) is more general because it can handle endogeneity in the determinants of
inefficiency as well.

5. Panel Data

Our current discussion of the SFM has focused on having access to cross-sectional data.
When repeated observations of firms are available, then more useful information about ineffi-
ciency (and often with more flexibility) can be extracted and a range of panel data SFMs are
available to the applied researcher. Here we highlight some of the most prominent models.
The advantage of panel data is that more information on inefficiency and productivity can
be parsed, and in particular, shed light on changes in efficiency or productivity, which differs
from a cross-sectional setting, which can only provide a static portrayal of inefficiency.
While Pitt & Lee (1981) were the first to consider extending the cross sectional SFM to
the panel data setting, it was Schmidt & Sickles (1984) who brought prominence to the use
of models tailored exclusively to panel data. They raise three problems with cross-sectional
models that are used to measure inefficiency and productivity: First, if the MLE is used to
estimate the parameters of the SFM and inefficiency through JLMS, everything is contingent
48 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

on distributional assumptions for both noise and inefficiency; Second, technical inefficiency
is assumed to be independent of the regressor(s).32; Third, the JLMS estimator is not a
consistent estimator of u, as E[u|ε] never approaches u as the number of cross-sectional
units approach infinity (n → ∞). Access to panel data can, to varying degrees, mitigate all
of these issues. However, with panel data comes a range of additional assumptions that the
researcher needs to carefully consider before proceeding.
To begin, consider the benchmark linear panel data regression model:

(5.1) yit = m(xit ; β) + ci + vit .

Aside from the indexing of our data by individual, i, and time, t, we have the presence
of firm specific heterogeneity, ci . The common dilemma facing application of the linear
panel data regression model is how to treat the relationship between ci and xit . Under the
fixed effects (FE) framework (Wooldridge 2010), xit is allowed to be correlated with ci and
the parameters of the model can be estimated consistently using the within transformation
(Baltagi 2013). Under the random effects (RE) framework, xit and ci are required to be
uncorrelated, leading to OLS being a consistent estimator, but is ultimately inefficient given
that the variance-covariance matrix of the composed error term c + v is no longer diagonal.
A feasible generalized least squares approach is available to obtain asymptotically efficient
estimators of the parameters of the regression model in this case.
Now, to think about where inefficiency enters the model in (5.1), we must characterize
the nature of inefficiency. If inefficiency is assumed to be constant over time, then it is likely
that ci might be augmented to also capture inefficiency. If inefficiency is time-varying then
we could include a second, one-sided error term to be convolved with vit in (5.1), in much
the same way we did in the benchmark SFM. Or, it could be that inefficiency is composed of
both a time-invariant component and a time-varying component. All told, the general panel

32If firms maximize profit, and inefficiency is known to the firm, then this assumption is unlikely to be true
as firms may adjust their inputs to account for inefficiency (e.g., see Mundlak 1961).
49

data SFM is

(5.2) yit = m(xit ; β) + ci − ηi + vit − uit = m(xit ; β) + αi + εit ,

where αi = ci − ηi with ci capturing time-invariant heterogeneity and ηi encapsulating time-


invariant inefficiency while εit = vit − uit with uit representing time-varying inefficiency. The
panel data SFM looks identical to the panel data regression model in (5.1), except that,
due to uit > 0, εit no longer has mean zero, and αi no longer solely captures individual
specific heterogeneity. Early approaches that studied inefficiency in panel data settings
placed restrictions on how inefficiency entered the panel data SFM. As time progressed,
fewer assumptions were made, especially as more advanced econometric techniques were
exploited.

5.1. Time-invariant Technical Inefficiency Models. When inefficiency in the panel


data SFM is assumed to be time-invariant, it is possible to estimate the model without
the need for distributional assumptions. To begin, we assume that uit does not exist in
(5.2) and all time-invariant unobserved heterogeneity is inefficiency, αi = ηi . With these
restrictions, the panel data SFM is written as

(5.3) yit = m(xit ; β) − ηi + vit ; i = 1, . . . , n; t = 1, . . . , T.

This model is termed the time-invariant SFM. Aside from the one-sided nature of ηi , this
model can be estimated with standard panel data regression techniques, once an assumption
on the underlying statistical relationship (either the FE or RE framework) between xit and ηi
is made. Which framework to deploy depends upon the relationship that one assumes exists
between the covariates of the model and firm level inefficiency. Under the FE framework
correlation is allowed between xit and ηi , whereas under the RE framework no correlation is
permitted between xit and ηi . Regardless of which framework is deemed appropriate, neither
requires distributional assumptions for η or v. This freedom from imposing a parametric
50 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

assumption on the distribution of ηi (i.e., we have some statistical requirements on the


distribution but do not require a precise parametric form) has led to the time-invariant SFM
being referred to as a distribution free approach (Schmidt & Sickles 1984). To estimate
the time-invariant SFM respecting the one-sided nature of ηi , a simple transformation is
needed to interpret the individual effect as time-invariant inefficiency as opposed to pure firm
heterogeneity. One major limitation of the time-invariant SFM is that separate identification
of inefficiency and individual heterogeneity is not considered. Additionally, the production
technology is assumed to be time constant, which may be a further limitation depending
upon the time dimension one has access to.
We briefly discuss how to estimate the time-invariant SFM under the FE framework, which
was first proposed by Schmidt & Sickles (1984). For ease of exposition, we assume m(·) is
linear in xit . The time-invariant SFM is

(5.4) yit =β0 + x0it β + vit − ηi

=(β0 − ηi ) + x0it β + vit

(5.5) =ci + x0it β + vit

where ci ≡ β0 − ηi . Under the FE framework, ηi and thus αi , i = 1, . . . , n are allowed to


have arbitrary correlation with xit .
Given the similarity of the time-invariant SFM and a traditional panel data regression
model, Schmidt & Sickles (1984) used standard estimation methods to estimate the pa-
rameters of the model, namely, within estimation. The within transformation subtracts
cross-sectional means of the data from each cross section (e.g., replacing yit by yit − ȳi· and
P
xit by xit − x̄i· , where ȳi· = (1/T ) t yit , etc.), thereby eliminating ci . OLS can then be
used to estimate the transformed model, essentially regressing transformed y on transformed
x. The OLS estimator with the transformed data, β,
b is a consistent estimator for β. An

estimator of ci , ĉi , is constructed from the mean of the residuals for each cross sectional
51

unit, i.e., ĉi = ȳi· − x0i· β,


b but it is biased, because ηi > 0 ∀i. A simple transformation will

produce a consistent estimator of ηi . Once ĉi is determined, η̂i is estimated as (Schmidt &
Sickles 1984):

(5.6) η̂i = max{ĉi } − ĉi ≥ 0, i = 1, . . . , n.


i

This formulation implicitly assumes that the most efficient firm/DMU in the sample is 100%
efficient. In other words, estimated inefficiency in the fixed-effects model is relative to the
best firm/DMU in the sample. If one is interested in estimating firm-specific technical
efficiency, it can be obtained from

(5.7) T
d E i = e−η̂i , i = 1, . . . , n.

Operating under the FE framework may be more appropriate for empirical applications in
which inefficiency is believed to be correlated with the inputs used. However, a disadvantage
of using the time-invariant SFM under the FE framework is that no other time-invariant
variables can be included. For example, the gender of a plant manager, or ownership status
of the firm (which may not change over a short time). Effectively, the influence of time
constant variables will be accumulated in (and distort) the estimates of inefficiency.
In settings where time-invariant variables are expected to be relevant regressors in the
production model, an alternative is to operate under the RE framework. Estimation of
the model still does not require distributional assumptions on v or η, but OLS on the
transformed model no longer represents an efficient estimator given that the composed error
term, vit − ηi no longer has a diagonal variance-covariance matrix [besides the requirement of
no correlation between inefficiency and inputs]. Schmidt & Sickles (1984) discuss estimation
under the RE framework through generalized least squares as well. Another alternative, if
one was uncomfortable with the implications stemming from RE framework would be to make
distributional assumptions, and estimate the model via maximum likelihood. This avenue
52 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

was suggested by Pitt & Lee (1981) and can allow time-invariant covariates to enter the
model while still identifying time-invariant inefficiency. The cost is the use of distributional
assumptions so that the likelihood function can be constructed. Following Aigner et al.
(1977), Pitt & Lee (1981) assume that ηi follows a half-normal distribution and vit follows a
normal distribution. Kumbhakar (1987) discussed estimation of inefficiency in such a model
by extending the JLMS formulation.

5.2. Time-varying Technical Inefficiency Models. The time-invariant SFM allows in-
efficiency to differ across individuals, but restricts any change over time. The implication
of this is that an inefficient firm could not improve productivity over time by lessening in-
efficiency. This may be unrealistic in a variety of applied settings, or where T is large. We
must consider models that allow both technology and inefficiency to change over time to
accommodate the idea of productivity and efficiency improvement at the firm level.
A nice feature of time-varying SFMs is that the time-invariant SFM is a special case
and, correspondingly, the time-invariant specification can be tested, opening up a variety of
inferential opportunities for empirical analyses. To introduce the time-varying SFM, recall
the model in (5.5):

(5.8) yit = ci + x0it β + vit .

To allow ci to be time-varying, one may impose some reasonable and tractable structure,
e.g., Cornwell, Schmidt & Sickles (1990) suggested replacing ci by cit where

(5.9) cit = c0i + c1i t + c2i t2 ,

where t is the time trend variable. The parameterization in (5.9) allows the parameters to
be firm-specific. If the number of cross-sectional units (n) is not large, one can define n firm
dummies and interact these dummies with time and time squared. These variables along
with the regressors (i.e., the x variables) are then used in a standard OLS regression. The
53

coefficients associated with the firm dummies and their interactions are the estimates of c0i ,
c1i , and c2i . These estimated coefficients can be used to obtain estimates of cit , c̃it . Again, the
within estimator can be used to consistently estimate β along with the 3n parameters from
the parameterization of cit . Finally, ĉit (the estimator of relative inefficiency) is obtained
from

(5.10) ĉit = ĉt − c̃it where ĉt = max(c̃jt ) ∀t.


j

In this model, efficiency is calculated relative to the best firm in each year. Since the firm
with the maximum c̃jt is likely to change over time, different firms may be fully efficient
(or inefficient at different levels) in different years. An alternative would be to calculate
ĉjt = maxjt (c̃jt ), the maximum over all j and t and replace ĉt with this definition in (5.10),
then efficiency is relative to the firm that was the most efficient over the entire sample period.
The Cornwell et al. (1990) estimation procedure is easy to implement. It relies on the
standard panel data estimator with the FE framework. Note that since t appears in the
inefficiency function, it cannot also appear as a regressor in xit , which would be required
if one were to capture technical change, i.e., a shift in the production frontier, m(x). In
other words, the above model cannot separate inefficiency from technical change, which is
an obvious drawback of this approach. In general, if one wants to have both time-varying
inefficiency and technical change, then the distribution free route of Cornwell et al. (1990)
will not work. In this case distributional assumptions will be necessary to allow time (and
higher powers of it) to enter the model in various places.
In a model with large n and small T the model will have too many parameters (3n
parameters in the cit function alone). A somewhat parsimonious time-varying inefficiency
model was proposed by Lee & Schmidt (1993):

(5.11) yit = m(xit ; β) + vit − uit = m(xit ; β) + εit .


54 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

where uit = ui `t and `t represent time specific effects to be estimated. This model is quite
flexible in its ability to model time-varying inefficiency. However, the temporal pattern of
inefficiency is assumed to be exactly the same for all firms (`t ). Under the FE framework,
this specification can be viewed as an interactive effects panel data model and estimation
can be undertaken by introducing both firm and time dummies. Though no distributional
assumptions are required by Lee & Schmidt (1993), the structure of inefficiency is similar to
that assuming the scaling property discussed above. Again, given that inefficiency depends
directly upon time it is difficult to model both time-varying inefficiency and technical change
in 5.11.
A similar idea was used prior to Lee & Schmidt (1993) in Kumbhakar (1990) and Battese
& Coelli (1992), who proposed time-varying SFMs, but made distributional assumptions
on both vit and uit and estimated the corresponding likelihood functions. Lee & Schmidt’s
(1993) model is more general than either the Kumbhakar (1990) or Battese & Coelli (1992)
models as both can be derived as special cases with appropriate parametric restrictions on
`t . Further still, the time-invariant SFM is also a special case: `t = 1 ∀ t. Once `t and ui
are estimated, inefficiency can be estimated from

(5.12) ûit = max{ûj `ˆt } − ûi `ˆt .


j

So far, the time-varying models that we have discussed treat inefficiency in a fully de-
terministic fashion, i.e. no distributional assumptions are required. In the Lee & Schmidt
(1993) time-varying SFM, both ui and `t are deterministic. This model can also be estimated
treating the time component as deterministic, but the individual component as stochastic
(through a distributional assumption). The deviation from the Lee & Schmidt (1993) time-
varying SFM in (5.11) is that uit = G(t)ui with G(t) being a deterministic function of time
and ui ∼ N+ (µ, σu2 ) (Kumbhakar 1990, Battese & Coelli 1992). The ideas discussed pertain-
ing to the scaling property appear here, where firms have a base level of inefficiency, and
55

then, through time, become more or less efficient. The stochastic component, ui , utilizes the
panel structure of the data in this model. The G(t) component is common across individuals
(as in, but not limited to, Lee & Schmidt 1993).
Given ui ≥ 0, uit ≥ 0 is ensured by having G(t) > 0. Undoubtedly, the most popular form
of G(t) is that proposed by Battese & Coelli (1992)

(5.13) G(t) = exp [γ(t − T )] ,

where T is the terminal period of the sample. The specification for G(t) is a simplification of
the first attempt to introduce stochasticity into the time-varying SFM by Kumbhakar (1990)
that assumes a more general specification of G(t) given by

−1
G(t) = 1 + exp(γ1 t + γ2 t2 )

(5.14) .

The Battese & Coelli (1992) specification essentially enforces γ2 = 0 in the Kumbhakar
(1990) time-varying SFM. The popularity of the Battese & Coelli (1992) time-varying SFM
has been aided by the freely available statistical package Frontier V4.1 which implements
this model at the push of a button (see Section 9 as well). Other specifications for G(t)
have also been proposed, see Cuesta (2000) and Kumbhakar & Wang (2005) for more recent
examples. Little research has been done on comparing a variety of forms of G(t). Lastly,
modeling technical change in the Kumbhakar (1990) or Battese & Coelli (1992) framework
is trivial because the imposition of distributional assumptions allows inclusion of t (as a
deterministic time-trend, e.g., linear, quadratic, etc.) as a component of xit .

5.3. Models that Separate Firm Heterogeneity from Inefficiency. While the time-
invariant SFM is a standard panel data model where ci is the unobservable individual effect,
a notable drawback of this approach is that inefficiency is indistinguishable from individual
heterogeneity. All time-invariant heterogeneity is confounded with inefficiency, and there-
fore ĉi will capture heterogeneity in addition to, or even instead of, inefficiency (Greene
56 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

2005b). An important question for practitioners using the time-invariant SFM is how to
view the time-invariant component. Should it be thought of as persistent inefficiency (as per
Kumbhakar 1991, Kumbhakar & Hjalmarsson 1993, Kumbhakar & Heshmati 1995, Kumb-
hakar & Hjalmarsson 1998) or is it more appropriate to think of it as individual heterogeneity,
capturing the effects of unobserved time-invariant covariates? If it is the latter, then the in-
sights from the time-invariant panel data SFMs are incorrect. A less rigid perspective is that
the truth lies somewhere in the middle; inefficiency may be decomposed into two components:
one that is persistent over time and one that varies over time.
Unless persistent inefficiency is disentangled from time-invariant individual heterogeneity,
practitioners need to choose between either the case in which ci represents persistent ineffi-
ciency or ci represents an individual-specific effect (heterogeneity). Here, we will discuss both
specifications. In particular, we will consider models in which inefficiency is time-varying
irrespective of whether the time-invariant component is treated as inefficiency or not. Thus,
the model we will describe is

(5.15) yit = ci + x0it β + vit − uit .

Compared to a standard panel data model, we have the additional time-varying inefficiency
term, −uit , in (5.15). If one treats ci , i = 1, . . . , n as a random variable that may be
correlated with xit but does not capture inefficiency, then the above model becomes what
has been termed the ‘true fixed-effects’ panel SFM (Greene 2005a). The model is labeled
as the ‘true random-effects’ SFM when ci is treated as uncorrelated with xit . Note that
these specifications are of the same nature as the models proposed by Kumbhakar (1991),
Kumbhakar & Hjalmarsson (1993), Kumbhakar & Heshmati (1995), and Kumbhakar &
Hjalmarsson (1998). The difference is in the interpretation of the ‘time-invariant term’, ci .
Estimation of the model in (5.15) is not straightforward. When ci , i = 1, . . . , n, are
embedded in the FE framework, the model encounters the incidental parameters problem
57

(Neyman & Scott 1948). The incidental parameters problem arises when the number of
parameters to be estimated increases with the number of cross-sectional units in the data,
which is the case with the ci in (5.15). In this situation, consistency of the parameter
estimates is not guaranteed even if n → ∞ because the number of ci increases with n.
Therefore, usual asymptotic results may not apply. In addition to this specific statistical
problem, another technical issue in estimating (5.15) is that the number of parameters to be
estimated can be prohibitively large for large nT .
For a standard linear panel data model (i.e., one that does not have −uit in (5.15)), the
literature has developed estimation methods to deal with this problem. These methods
involve transforming the model so that ci is removed before estimation. Without ci in the
transformed model, the incidental parameters problem no longer exists and the number of
parameters to be estimated no longer increases with the number of individuals. Methods
of transformation include conditioning the model on ci ’s sufficient statistic33 to obtain the
conditional MLE, and the within-transformation model or the first-difference transformation
model to construct the marginal MLE (e.g., Cornwell & Schmidt 1992). For the basic panel
data SFM, this could be done by transforming the error term if assumptions on vit and
uit are such that the composed error term’s distribution is closed-skew normal (i.e., the
normal-half-normal distributional pair).
These standard methods, however, are usually not applicable to (5.15). For the conditional
MLE of (5.15), Greene (2005b) showed that there is no sufficient statistic for ci . For the
marginal MLE, the resulting model after the within or first-difference transformation usually
does not have a closed form likelihood function, if one uses standard procedures.34 In general
this would not pose an issue as regression methods can be easily applied. However, given
the precise interest in recovering estimates of the parameters of the distribution of ineffi-
ciency, maximum likelihood or specific moments of the distribution of the transformed error
33Asufficient statistic contains all the information needed to compute any estimate of the parameter.
34Colombi, Martini & Vittadini (2011) showed that the likelihood function has a closed form expression.
Chen et al. (2014) considered a special case of Colombi et al. (2011) and derived a closed form expression.
58 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

component are needed. This precipitates methods that can recover information regarding
uit .
Greene (2005b) proposed a tentative solution. He assumed uit follows a simple i.i.d. half-
normal distribution and suggested including n dummy variables in the model for ci , i =
1, . . . , n and then estimating the model by MLE without any transformation. He found that
the incidental parameters problem does not cause significant bias to the model parameters
when T is relatively large (e.g., T ≥ 10). The problem of having to estimate more than n
parameters is dealt with by employing an advanced numerical algorithm.
There are some recent econometric developments on this issue. First, Chen et al. (2014)
proposed a solution in the FE framework. They showed that the likelihood function of
the within transformed and the first-difference model have closed form expressions using
results in Domı́nguez-Molina, González-Farı́as & Ramos-Quiroga (2003). The same theorem
in Domı́nguez-Molina et al. (2003) is used by Colombi, Kumbhakar, Martini & Vittadini
(2014) to derive the log-likelihood function in the RE framework.
Using a different approach, Wang & Ho (2010) solve the problem classified in Greene
(2005b) by proposing a class of SFMs in which the within and first-difference transformations
on the model can be carried out while also providing a closed form likelihood function. The
main advantage of such a model is that because the ci s are removed from the model in (5.15),
the incidental parameters problem is avoided entirely. As such, consistency of the estimates
is obtained for either n → ∞ or T → ∞, which is invaluable for applied settings. A further
computational benefit is that the elimination of ci s reduces the number of parameters to be
estimated to a manageable number. The catch is in the specification of inefficiency which is
the product of an i.i.d non-negative random component and a deterministic function of zit
(determinants of inefficiency). Formally, the Wang & Ho (2010) model is:

(5.16) yit = ci + x0it β + εit ,


59

where εit = vit − uit with vit ∼ N (0, σv2 ) and uit = git u∗i with u∗i ∼ N+ (µ, σu2 ), which is
the now familiar scaling property model with a truncated-normal distribution for the basic
distribution of inefficiency.
For the scaling function Wang & Ho (2010) set git = g(zit0 δ). What allows the model
transformation to be applied is the scaling property; the within and first-difference transfor-
mations leave this stochastic term intact as u∗i does not change with time. As Wang & Ho
(2010) showed that the within-transformed and the first-differenced models are algebraically
identical we have only provided discussion on the first-differenced model. However, a lim-
itation of their model is that it does not completely separate persistent and time-varying
inefficiency, a subject which we now turn our attention to. Lastly, as with the models of
Kumbhakar (1990) or Battese & Coelli (1992), the use of distributional assumptions allows
both time-varying inefficiency and technical change to be modeled in (5.16).

5.4. Models that Separate Persistent and Time-varying Inefficiency. Although sev-
eral models discussed earlier can separate firm-heterogeneity from time-varying inefficiency
(which is either modeled as the product of a time-invariant random variable and a deter-
ministic function of covariates or distributed i.i.d. across firms and over time), none of these
models consider persistent technical inefficiency. It is important to quantify persistent ineffi-
ciency, especially in short panels, as it captures the effects of inputs like management quality
(Mundlak 1961). Unless there is a change in something that affects management practices
at the firm (for example new government regulations or a change in ownership), it is unlikely
that persistent inefficiency will change. The importance of persistent inefficiency contrasts
with time-varying as this can change over time without requiring structural changes which
impact the firm.
This distinction between the time-varying and persistent components is important from a
policy perspective as each yields different implications. Colombi et al. (2014) refer to time-
varying inefficiency as short-run inefficiency and mention that it can arise due to failure in
60 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

allocating resources properly in the short run. They argued that, for example, a hospital
with excess capacity may increase its efficiency in the short-run by reallocating the work
force across different activities. Thus, some of the physicians’ and nurses’ daily working
hours might be changed to include other hospital activities such as acute discharges. This
is a short-run improvement in efficiency that may be independent of short-run inefficiency
levels in the previous period, which can justify the assumption that uit is i.i.d. However,
this does not impact the overall management of the hospital and so is independent from
time-invariant inefficiency.
To help formalize this issue more clearly we consider the model35

(5.17) yit = β0 + x0it β + εit = β0 + x0it β + vit − (ηi + uit )

Technical inefficiency is represented as ηi + uit where ηi is the persistent, firm-specific com-


ponent (for example, time-invariant ownership or geographic location) and uit is the time-
varying component of technical inefficiency which is firm and time specific. Model (5.17)
generalizes the previously discussed models because it allows for firm heterogeneity, time-
invariant and time-varying inefficiency all at once.
Such a decomposition is desirable because, since ηi does not change over time, for a firm
to improve efficiency a structural change in policy or management would need to arise.
Additionally, ηi does not fully capture firm level inefficiency because it does not account for
learning over time since it is time-invariant; the time-varying component, uit can capture
this aspect. In (5.17) the level of overall firm inefficiency, as well as the components, are
important to know because they convey different types of information. Thus, for example,
it may be argued that if residual inefficiency for a firm is relatively large in a particular year
this is due to an event which is unlikely to occur in the following next year. Alternatively,
if persistent inefficiency is large, then a firm is expected to operate with a relatively high

35This
is the model proposed by Kumbhakar & Hjalmarsson (1993), Kumbhakar & Heshmati (1995), and
Kumbhakar & Hjalmarsson (1998), among others.
61

level of inefficiency over time, unless some changes in policy and/or management occur.
Therefore, a large value of ηi is more concerning in the long run given its persistent nature
than is a high value of uit .
The specification in (5.17) offers that advantage of testing for the presence of the persistent
nature of technical inefficiency without the imposition of a specific parametric form of time-
dependence. Furthermore, by including time in the xit vector, (5.17) has the ability to
separate exogenous technical change from technical inefficiency.
To estimate the model we rewrite (5.17) as

(5.18) yit = αi + x0it β + ωit = (β0 − ηi − E[uit ]) + x0it β + vit − (uit − E[uit ]).

The error, ωit , has zero mean and constant variance. Model (5.18) is a standard panel
data model with firm-specific heterogeneity (one-way error component model), and can be
estimated either by the within transformation (under the FE framework) or by generalized
least-squares (under the RE framework).
The SFM in (5.18) can be estimated under the FE framework using a step-wise procedure.

Step 1: The standard within transformation can be performed on (5.18) to remove αi


before estimation. Since both the components of ωit are zero mean and constant
variance random variables, the within transformed ωit will generate a random vari-
able that has zero mean and constant variance. OLS can be used on the within
transformed version of (5.18) to obtain consistent estimates of β.
Step 2: Given the estimate of β, β̂, from Step 1, construct the pseudo-residuals rit =
yit −x0it β̂, which contain information on αi +ωit . Using these, we first estimate αi from
the mean of rit for each i. Then, we can estimate αi from maxi α̂i − α̂i = maxi {r̄i }− r̄i
where r̄i is the mean (over time) of rit for firm i. Note that the intercept, β0 , and
ωit are eliminated by taking the mean of rit over time for a firm. The above formula
gives an estimate of αi relative to the best firm in the sample.
62 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

Step 3: With our estimates of β and ηi , we calculate residuals eit = yit − x0it β̂ + η̂i ,
which contains information on β0 + vit − uit . At this stage additional distributional
assumptions are required to separate vit from uit . Here we follow convention and
assume vit ∼ i.i.d. N (0, σv2 ) and uit ∼ i.i.d. N+ (0, στ2 ). MLE can be deployed here,
treating eit as the dependent variable, to estimate β0 and the parameters associated
with vit and uit . The log-likelihood for this setup is, letting N = nT ,
n X
T n T
X 1 XX 2
(5.19) ln L = −N ln σ + ln Φ(−eit λ/σ) − 2 e
i=1 t=1
2σ i=1 t=1 it

Note that the parameters to be estimated here are β0 , σν2 and στ2 . Once these pa-
rameters have been estimated a JLMS conditional mean or median technique can be
used to estimate uit for each observation.

To summarize estimation under the FE framework, we estimate (5.18) using standard FE


panel data tools to obtain consistent estimates of β in Step 1. Step 2 estimates persistent
technical inefficiency, ηi . Lastly, Step 3 involves estimation of β0 and the parameters as-
sociated with the distributional assumptions imposed on the random components, vit and
uit . One can then use the JLMS formula to estimate the time-varying (residual) component
of inefficiency, uit . Note that no distributional assumptions are used in the first two steps.
Without further assumptions, residual inefficiency cannot be identified and hence, distribu-
tional assumptions are needed in the last step. This model can also be estimated under the
RE framework (see also Colombi et al. 2014).

5.5. Models that Separate Firm Effects, Persistent Inefficiency and Time-varying
Inefficiency. All of the panel data SFMs introduced so far have departed from the general
model introduced in (5.2) in some aspect pertaining to the four separate error components.
This is due to the fact, that until recently, it was not clear how to estimate the full panel
data SFM represented by (5.2). The models of Kumbhakar, Lien & Hardaker (2014) and
Colombi et al. (2014) overcome the limitations of the previous models by embracing the
63

nature of the four component structure inherent in the general panel data SFM. In the
SFM represented in (5.2) the four components take into account different factors affecting
output, given the inputs. As in Greene (2005b, 2005a) the first component captures firms’
latent heterogeneity, which needs to be extricated from inefficiency; the second component
captures time-varying inefficiency, the third component captures time-invariant inefficiency
as in Kumbhakar & Hjalmarsson (1993), Kumbhakar & Heshmati (1995), and Kumbhakar &
Hjalmarsson (1998) while the fourth component captures stochastic shocks beyond control
of the firm.
The ability to estimate model (5.2) allows improvement over the previous models in several
ways. To begin, while some of the time-varying inefficiency models just described can ac-
commodate firm effects, these models fail to acknowledge the potential for factors that might
have time-invariant effects on firm inefficiency. Second, SFMs which allow time-varying in-
efficiency commonly assume that the inefficiency level of the firm at time t is independent of
its previous level of inefficiency; it is more reasonable to assume that a firm may eliminate
some of its inefficiency by mitigating short-run rigidities, while other sources of inefficiency
may remain over time. The former is captured by the time-invariant component, ηi , and
the latter by the time-varying component, uit . Finally, many panel SFMs do consider time-
invariant inefficiency, but do not simultaneously account for the presence of unobserved firm
heterogeneity. In doing so, these models confound time-invariant inefficiency with firm ef-
fects (heterogeneity). The models proposed by Greene (2005b, 2005a), Kumbhakar & Wang
(2005), Wang & Ho (2010) and Chen et al. (2014) decompose the error term in the production
function into three components: a firm-specific time-varying inefficiency term; a firm-specific
effect capturing latent heterogeneity; and a time- and firm-varying random error term. How-
ever, these models consider any producer-specific, time-invariant component as unobserved
heterogeneity. Thus, although firm heterogeneity is now accounted for, it comes at the cost of
ignoring long-term inefficiency. As before, latent heterogeneity is confounded with long-run
inefficiency.
64 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

Estimation of the panel data SFM in (5.2) can be undertaken in a single stage MLE
method based on distributional assumptions on the four components (Colombi et al. 2011).
We first describe a simpler, multi-step procedure (Kumbhakar et al. 2014). For this, we
rewrite the model in (5.2) as

(5.20) yit = β0∗ + x0it β + αi + εit ,

where β0∗ = β0 − E[ηi ] − E[uit ]; αi = ci − ηi + E[ηi ]; and εit = vit − uit + E[uit ]. With this
specification both αi and εit are zero mean and constant variance random variables. (5.20)
is estimated in three steps.

Step 1: Standard random effect panel regression is used to estimate β̂ (since (5.20) is
a common panel data model). Predicted values of αi and εit , denoted by α̂i and ε̂it
are also available after estimating (5.20).
Step 2: Time-varying technical inefficiency, uit , is estimated using ε̂it from Step 1.
Since

(5.21) εit = vit − uit + E[uit ],

by assuming vit is i.i.d. N (0, σv2 ) and uit is i.i.d. N+ (0, σu2 ), which yields E[uit ] =
p
2/π σu , and ignoring the difference between the true and predicted values36 of εit ,
we can estimate (5.21) using standard SFA techniques. Doing so produces predictions
of the time-varying technical inefficiency component uit , E [e−uit |εit ], (i.e., Battese &
Coelli 1988), which we call relenting technical efficiency (RTE).
Step 3: Estimate ηi following a similar strategy as in Step 2. For this we use α̂i from
Step 1. Since

(5.22) αi = ci − ηi + E[ηi ],

36Which is the standard practice in any two- or multi-step procedure.


65
p
by assuming ci is i.i.d. N (0, σµ2 ), ηi is i.i.d. N+ (0, ση2 ), where E[ηi ] = 2/π ση , es-
timate (5.22) using the standard normal-half-normal cross-section SFM and obtain
estimates of the persistent technical inefficiency component, ηi following JLMS. Per-
sistent technical efficiency (PTE) can then be estimated as PTE = e−ηi , where η̂i is
the JLMS estimator of ηi . Overall technical efficiency (OTE) is then constructed as
the product of PTE and RTE, i.e., OTE = PTE×RTE.

It is possible to extend this model (in steps 2 and 3) to include PTE and RTE that is
distributed as truncated-normal or exponential as opposed to half-normal.
While the multi-step approach of Kumbhakar et al. (2014) is straightforward to implement,
it is inefficient relative to full MLE. However, given the structure of the four separate errors,
deriving the likelihood function was previously seen as infeasible. However, using insights
related to the closed-skew normal distribution, as in Colombi et al. (2014), a tractable
likelihood function turned out to be easily obtainable.
Colombi et al. (2014) made skew normal distributional assumptions for both ci − ηi and
vit −uit in (5.20).37 Assuming vit is i.i.d normal and uit is i.i.d half-normal, the composed error
vit −uit has a skew normal distribution. The same set of assumptions can be used for ci and ηi .
Thus, model (5.2)’s likelihood, can be derived. Even though the log-likelihood for (5.2) can
be determined based on skew normal assumptions for the time-varying and time-invariant
error components, it can be daunting to implement. Greene & Fillipini (2014) recently
proposed a simulation based optimization routine which circumvents many of the challenges
associated with direct optimization. They used a trick suggested by Butler & Moffitt (1982),
conditioning on ci and ηi . This conditioning eliminates many of the computational hurdles
that direct optimization of the likelihood function presents.

5.6. The Four Component Panel Data SFM with Determinants of Inefficiency.
A further generalization of the four component model in (5.2) involves the inclusion of
37Theskew normal distribution is a more general distribution than the normal distribution, allowing for
asymmetry (Azzalini 1985).
66 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

determinants of inefficiency, either for the time-varying or the time invariant components.
An estimator for this model was recently proposed in Badunenko & Kumbhakar (2017),

(5.23) yit = m(xit ; β) + ci − ηi + vit − uit ,

2 2 2 2
where ηi ∼ N+ (0, ση,i ), uit ∼ N+ (0, σu,it ), ci ∼ N (0, σc,i ), and vit ∼ N (0, σv,it ). These
distributional assumptions are imposed so that the time invariant composed error and the
time-varying composed error both follow the closed skew normal distribution. Each of the
variance parameters of the four components is dependent upon a set of covariates and spec-
0 0 0
2
ified as an exponential function: ση,i = ση2 ezη,i δη , σc,i
2
= σc2 ezc,i δc , σu,it
2
= σu2 ezu,it δu , and
0
2
σv,it = σv2 ezv,it δv . The time-constant and time-varying z vectors can overlap due to the as-
sumed distributional assumptions, that is zc,i can share elements with zη,i and zu,it can share
elements with zv,it .
To estimate this four component model, Badunenko & Kumbhakar (2017) used the insights
of Greene & Fillipini (2014) and deployed simulated maximum likelihood techniques. The
benefit of this approach is that rather than having T integrals to evaluate, by conditioning
on ci − ηi , the likelihood function can be written as the product of T univariate integrals.
Simulation methods are required to construct draws of ci − ηi inside the convolution density.
The final log-likelihood function is
n R
"T #!
X X Y 2  εitr   εitr λit 
(5.24) L= log R−1 φ Φ ,
i=1
σ
r=1 t=1 it
σit σit
p 0 0
p 0 0
p 0 p 0 
where σit = ezu,it δu + ezv,it δv , λit = ezu,it δu −zv,it δv , εitr = it − ezc,i δc Vir − ezη,i δη |Uir |
and it = yit − m(xit ; β). R is the number of draws over which to numerically evaluate the
integral (larger R increases accuracy but slows down the routine, smaller R leads to faster
computation but decreases accuracy). Lastly, both Vir and Uir are random draws from a
standard normal distribution. Implementation of this routine is straightforward if one has
access to a standard normal random number generator (which is typically available in any
67

general statistical software). Once draws for Vir and Uir have been constructed, the likelihood
is evaluated for the current set of parameters (β, δu , δv , δη , δc ). This process is then iterated
over different sets of parameter values. Naturally, one can impose constancy at various parts
of the error components by restricting δ` = 0 for ` ∈ {u, v, c, η}.

5.7. Inference Across the Panel Data SFM. The most general SFM in the panel con-
text is the model which allows for firm specific heterogeneity, persistent technical efficiency,
relenting technical inefficiency and individual-time specific idiosyncratic shocks. Colombi
et al. (2014) denote this model as TTT (True for having firm specific heterogeneity, True
for having time constant inefficiency and True for having time-varying inefficiency). The
majority of all panel data models that have appeared in the literature are special cases of
TTT. For example, the widely used true RE model of Greene (2005b) is a special case of
the TTT model. The same holds for all of the models we have discussed above. Naturally,
inference is necessary to determine the model which best fits the data at hand. One ben-
efit of nearly all of the panel data SFM discussed here is that standard panel data type
tests (coefficient significance, fixed versus random effects framework, serial correlation, etc.)
are easily implemented. This is similar to the benefits of the cross-sectional SFM that we
discussed earlier.
What is less straightforward is to test the most general TTT model against more restricted
versions. Testing any of the previous models against the most general TTT model is a non-
standard problem because, under the null hypothesis, one or more of the parameters of
interest lie on the boundary of the parameter space. Under reasonable assumptions the
asymptotic distribution of the log-likelihood ratio test statistic is χ̄2 , as discussed in Section
2.3.1. For example, the model of Pitt & Lee (1981) could be tested against the TTT model
with the log-likelihood ratio test statistic but using χ̄2 to determine the p-value, see Table 1.
Future research focusing on adapting testing procedures to the TTT framework is impor-
tant moving forward. As discussed earlier, the presence of both time-varying and invariant
68 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

efficiency yields different policy recommendations and so working with models that document
their presence, or lack of one, are important for proper analysis.

6. Nonparametric Estimation of the SFM

6.1. Early Attempts. In a nutshell, the semiparametric and nonparametric approaches to


SFA typically use the benchmark SFM of Aigner et al. (1977) as the stepping-stone, general-
izing it in different ways by relaxing all or some parametric assumptions by utilizing existing
semiparametric and nonparametric statistical methods, such as the Nadaraya-Watson esti-
mator, the local polynomial estimator or the likelihood (pseudo or local) estimators.
To facilitate further and more precise discussion, recall that the benchmark SFM for a
sample of n DMUs is given by:

(6.1) yi = m(xi ) + vi − ui = m(xi ) + εi , i = 1, . . . , n,

where m(·) is the frontier of the production technology that can be used to transform vector
of inputs x ∈ Rq+ into scalar output yi , perturbed by some statistical noise vi and adjusted
by technical inefficiency ui . As we discussed in Section 2, traditional parametric estimation
of the model begins by assuming a particular functional form for the production technology,
most commonly a Cobb-Douglas or a Translog, besides making distributional assumptions
on both vi and ui , which help to identify and estimate the unknown parameters via, say, the
maximum likelihood approach. All the asymptotic results (consistency, asymptotic normal-
ity) are conditional on these assumptions and if they happen to be incorrect then, strictly
speaking, all these results may be invalid. In such cases, the parametric MLE will be incon-
sistent or converging in probability not to the truth (e.g., true elasticities) but to some other
numbers, which can even be very far from the truth if the parametric assumption made on
a function is far from the true one.
The early attempts to estimating SFM nonparametrically or semiparametrically go back to
at least Banker & Maindiratta (1992), Fan, Li & Weersink (1996) and Kneip & Simar (1996).
69

Specifically, Banker & Maindiratta (1992) proposed a nonparametric approach in the spirit of
the DEA estimator but embedded in a maximum likelihood framework, similar to parametric
SFA, and thus allow for modeling both the noise and the inefficiency. A few years later, Fan
et al. (1996) proposed estimating the production frontier in another flexible manner, using
nonparametric kernel regression methods embedded into the parametric maximum likelihood.
About the same time, Kneip & Simar (1996) suggested using the kernel regression estimator
(Nadaraya-Watson in particular) for the panel data SFM.
Importantly, note that the estimated conditional mean E[yi |x] of the production frontier
is a biased estimator when ignoring the inefficiency term. Indeed, a critical assumption for
consistent estimation of the production frontier in a regression setting is E[εi |x] = 0 and due
to the one-sided nature of ui , this assumption is not satisfied, because E[εi |x] = µu 6= 0 in
the simplest case when inefficiency is independent of the inputs, or more generally, E[εi |x] =
µu (x) 6= 0 ∀x. Therefore, the production frontier cannot be identified in the regression
setup, where one would estimate

(6.2) yi = m(xi ) + εi = m(xi ) + µu + (εi − µu ) ≡ m∗ (xi ) + ε∗i .

Realizing this, Fan et al. (1996) proposed correcting the estimation bias of m(x) via a
three-stage semiparametric pseudo-likelihood estimation of the SFM. In this approach, at
the first stage, one estimates (6.2) non-parametrically.38 Results from this first stage are
then fed into the second stage, involving parametric MLE with particular assumptions on
the distribution of the noise and inefficiency that help identifying and disentangling the
two.39 Once the parameters of this symbiosis of MLE and kernel-regression are estimated,
the estimated conditional mean can then, in the third stage, be corrected for the bias by

38They used a local constant (Nadaraya-Watson) regression, although other consistent nonparametric esti-
mators can be used there too.
39In their work, the normal-half-normal assumption was used, but other assumptions as discussed above can
be used there too. Note, however, that for some alternative distributional assumptions on u, for example
exponential or truncated-normal, a concentrated version of the log-likelihood function may not exist, causing
identification problems.
70 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

the estimated mean of inefficiency (as in COLS), µ̂u (xi ) to get a consistent estimator m(xi )
given by

(6.3) m(x b ∗ (xi ) − µ̂u (xi ),


b i) = m

Kneip & Simar (1996) also proposed a similar strategy for correcting the bias occurring
in estimating (6.2) nonparametrically, but avoided using MLE due to the possibility of
disentangling the noise from inefficiency without distributional assumptions, by utilizing the
panel-data SFA framework.
The approaches of Fan et al. (1996) and Kneip & Simar (1996) provided a useful framework
and formed a foundation on which many other approaches have been built.40 For example,
more recent approaches of Kuosmanen & Kortelainen (2012) and Parmeter & Racine (2012)
share some essence of Fan et al. (1996) except that they required the estimated produc-
tion frontier to obey traditional axioms of production, such as monotonicity and concavity,
something that Fan et al. (1996) did not accommodate in their approach. Specifically,
Parmeter & Racine (2012) employ the framework of Fan et al. (1996) but combine it with
constraint weighted bootstrapping (Hall & Huang 2001, Du, Parmeter & Racine 2013) to
ensure that monotonicity and concavity are enforced during estimation. More recently, Noh
(2014) made improvements to the approach of Parmeter & Racine (2012), which resulted in
small sample performance gains. On the other hand, Kuosmanen & Kortelainen (2012) used
an entirely different estimation approach, concave nonparametric least-squares (CNLS), to
impose monotonicity and concavity. Lastly, Martins-Filho & Yao (2015) showed that while
the estimator of Fan et al. (1996) is consistent, the parametric estimator for the parameters

of the density of the convolved error yields an asymptotic bias (when normalized by n)
and proposed an alternative estimator that estimates the distributional parameters and the
unknown frontier jointly.

40See Parmeter & Zelenyuk (2016) for a more comprehensive review of this topic.
71

6.2. Local Likelihood Methods. The local likelihood approach (Tibshirani & Hastie
1987) is known to be a natural alternative to the semi-parametric pseudo-likelihood, and
it was first proposed in the SFA context by Kumbhakar, Park, Simar & Tsionas (2007).
This approach closely resembles the parametric likelihood approach with the only (yet fun-
damental) difference being the kernel-based weights (instead of the equal weights) used to
weigh each individual contribution to the likelihood, which help in localizing the estimation
in the direction of each continuous variable through the bandwidths. Specifically, for a given
regression error density, fε (ε, θ), we have the local log-likelihood function
n
X
−1
(6.4) Ľn (θ(x), mx ) = (n|h|) ln fε (yi − m(xi ); θ(x))Kix ,
i=1

where mx captures the conditional mean of y given x (a q × 1 vector of covariates) and θ is


q  
xis −xs
h−1
Q
the vector of remaining parameters of fε , Kix = s k hs
is the standard product
s=1
kernel where k(·) is any second order univariate kernel (Epanechnikov, Gaussian, e.g.), hs
is the smoothing parameter for the sth covariate (and is the sth element of vector h), while
|h| = h1 h2 · · · hq .
Kumbhakar et al. (2007) used a local-linear approximation for the unknown production
function m(xi ) combined with the assumption of a normal, half-normal convolved error term,
where parameters are also modeled as unknown functions of the covariates,
n h  i
2 (x ) 2 (x )
X
−1 2 2 −σ̈x λ̈x (xi )−0.5σ̈x
(6.5) Ľn =(n|h|) −0.5σ̈x (xi ) − 0.5ε̈i e i
+ ln Φ −ε̈i e i
Kix
i=1

where ε̈i = yi − m̈x (xi ), m̈x (xi ) = m̈0 − m̈01 (xi − x), σ̈x2 (xi ) = σ̈02 + σ̈120 (xi − x), and
λ̈x (xi ) = λ̈0 + λ̈01 (xi − x).41
Noting that often the main focus of interest is related to σu , Park, Simar & Zelenyuk (2015)
suggested directly parameterizing the local likelihood function in terms of ln σv2 and ln σu2

41One could also use a quadratic approximation, but note that even in this local-linear case, there are already
3 + 3q parameters to estimate (i.e., optimize over) at each point of interest x: these are the three functional
estimates, m̈0 , σ̈02 and λ̈0 and the 3q derivative estimates of the functions, m̈1 , σ̈12 and λ̈1 .
72 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

which also impose positivity of σv2 and σu2 throughout the estimation, making it more stable
computationally. Park et al. (2015) also outlined an asymptotic theory for modeling discrete
variables in the context of the local-likelihood approach, which can be imperative for many
applications, which many covariates that researchers have access to are categorical in nature
(regulated vs. non-regulated firms or industries, private vs. publicly owned companies, male
vs. female managers, etc.). The local-likelihood function in this case would be
n h  2 c d   2 c d 
2 c d 2 c d
X
c d −1
Ľ(θ(x , x ), mxc ,xd ) =(n|h|) −0.5 ln eσ̈v (xi ,xi ) + eσ̈u (xi ,xi ) − 0.5ε̈2i / eσ̈v (xi ,xi ) + eσ̈u (xi ,xi )
i=1
 2 c d 2 c d
p i
c d c d
+ ln Φ −ε̈i eσ̈u (xi ,xi )/2−σ̈v (xi ,xi )/2 / eσ̈v (xi ,xi ) + eσ̈u (xi ,xi ) Kixc W i (xdi ).
2 2
(6.6)

where xci is a vector of continuous regressors while xdi is a vector of discrete regressors
and W i (xdi ) is an appropriate discrete kernel, e.g., the one proposed by Aitchison & Aitken
(1976) or its variations. The theory in Park et al. (2015) is derived for the case of kernel from
I(xd 6=xd )
Racine & Li (2004), given by W i (xd ) = kj=1 ωj ij j , which is a standardized version of
Q

the Aitchison-Aitken kernel, standardized so that the bandwidths for a jth discrete variable,
here denoted as ωj , are always between 0 and 1, regardless of the number of categories.
However, this theory also extends (with some modifications) to cases with other discrete
kernels. For example, one might prefer the so-called discrete Epanechnikov kernels, which
are particularly useful and can be superior to others in case of sparse data (e.g., see Chu,
Henderson & Parmeter (2017) and the references cited therein). One can also use more
adaptive bandwidths, e.g., allow for bandwidths of some or all continuous regressors to differ
across categories of some or all discrete variables (e.g., see Li, Simar & Zelenyuk (2016) for
related discussion).
Standard optimization algorithms can be used here, but as with any nonlinear optimiza-
tion, careful choice of starting values is imperative, especially in selecting the bandwidths.
For example, Kumbhakar et al. (2007) suggested starting with the local-linear least-squares
73

estimates for m̈0 and m̈1 and the global, parametric maximum likelihood estimates for σ 2
and λ (from Aigner et al. 1977) so that m̈0 is properly corrected (as in Fan et al. 1996).
Selection of the bandwidths is a very important step here (as is true in general for kernel-
based methods) and many interesting general selection methods can be adapted to the cur-
rent context. One of the most popular approaches is cross-validation.42 Kumbhakar et al.
(2007) outlined how to use least-squares cross-validation (LSCV) for their approach. Mean-
while, Park et al. (2015) suggested using maximum likelihood cross-validation (MLCV),
which is more natural for the local-likelihood approach, although it may be more demanding
in computation. For the starting values in numerical optimization of LSCV or MLCV for
selecting optimal bandwidths, one could use the so-called rules-of-thumb bandwidths that
reflect the rates of convergence required for the asymptotic theory, e.g., for a continuous
variable xcs , use h0 (xcs ) = 1.06 × n−1/(4+q) σ̂xcs , where σ̂xcs is estimated standard deviation of
xcs , and ω0 = n−2/(q+4) for the discrete bandwidths.
Kneip, Simar & Van Keilegom (2015) provide an update of the Kumbhakar et al. (2007)
estimator whereby the distributional assumption on the inefficiency term can be dropped.
The only parametric assumption required in Kneip et al. (2015) is that the two-sided error
term is normal, which allows them to rely on penalized likelihood, where the unknown density
is constructed non-parametrically via a histogram over the support of the covariate space
and the penalty term is included to ensure appropriate smoothness of the resulting density.
Both the theory and simulated evidence appearing in Kneip et al. (2015) suggest that this
estimator works quite well in a variety of settings. To date, no application of this method
has appeared to our knowledge and so it represents an exciting opportunity moving forward.

6.3. Local Least-Squares Approaches. In spite of the appealing theoretical advantages


of the likelihood-based approaches, they involve numerical optimization of the local likelihood
function over many parameters at each point of interest, which can be computationally
42Formore discussions on the pros and cons, as well as references on this approach in general, see Henderson
& Parmeter (2015a).
74 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

complex, especially if bootstrap methods are needed to conduct inference. An attractive


alternative that is much simpler to compute is provided by adopting the local least-squares
methods; because these methods do not require nonlinear optimization (given closed form
solutions), only basic matrix operations are required, marking dramatic improvements in
computation time.
Recently, Simar et al. (2017) (SVKZ hereafter) proposed what can be viewed as a semi- or
nonparametric generalization of COLS (Olson et al. 1980)43, which also allows for modeling
determinants of inefficiency. Specifically, they considered a generalization of (6.1) given by

(6.7) yi = m(xi , zi ) + vi − ui = m(xi , zi ) + εi .

where m(xi , zi ) is the production frontier evaluated at xi , the realizations of inputs for
observation i, and at zi , the realization of the so-called environmental factors faced by the
observation i, and disturbed by the realizations of statistical noise vi and inefficiency ui . In
general, they required fairly general and mild conditions on the model, e.g., (ui |xi = x, zi =
z) ∼ D+ (µu (x, z), σu2 (x, z)) with D+ (·, ·) being a non-negative random variable with mean
µu (·, ·) and finite positive variance σu2 (·, ·), while (vi |xi = x, zi = z) ∼ D(0, σv2 (x, z)) with
D(0, ·) being a random variable with mean zero and finite positive variance σv2 (·, ·). They also
assumed that, conditional on (xi , zi ), ui and vi are independent random variables. Further,
given that vi has a symmetric distribution around zero, while ui is a positive random variable
from a skewed distribution E[εi |x, z] = −E[ui |x, z] 6= 0. Therefore, after recentering, we
have

(6.8) yi = m(xi , zi ) + vi − ui + E[ui |x, z] − E[ui |x, z] = m∗ (xi , zi ) + ε∗i

where m∗ (xi , zi ) = m(xi , zi ) − E[ui |x, z] and ε∗i = εi + E[ui |x, z]. Adapting the strategy of
COLS from Olson et al. (1980), SVKZ proposed in the first stage the estimator of m∗ (x, z),

43As with our earlier discussion, SVKZ referred to this approach as nonparametric MOLS, but cite Olson
et al. (1980), who used the term COLS and so we refer to it as COLS here.
75

b ∗ (x, z) using local-polynomial least-squares, noting that under mild regularity conditions
m
and appropriate choice of bandwidths, such estimators have desirable statistical properties
(consistency, asymptotic normality, etc.; see Fan & Gijbels 1996, Li & Racine 2007, Hen-
derson & Parmeter 2015a). Then, in the second stage, they utilized the moment conditions
implied by the assumptions on ui and vi , namely

E[ε∗ |x, z] =0,

E[(ε∗ )2 |x, z] =σu2 (x, z) + σv2 (x, z),

E[(ε∗ )3 |x, z] = − E (u − E[u|x, z])3 |x, z ,


 

and estimate the second and third moments of ε∗ using local-polynomial methods with the
residuals εb∗i = yi − m
b ∗ (xi , zi ) from the first stage, i.e.,
n
X
(6.9) m
b 2 (x, z) = ε2i
Ai (x, z)b
i=1

and
n
X
(6.10) m
b 3 (x, z) = ε3i ,
Ai (x, z)b
i=1

where Aj (x, z) would vary depending upon the local smoothing method used. If one desires
to estimate the level of the frontier in SVKZ’s setup, then a (local) parametric distributional
assumption for ui is needed, although the ranking of output would be independent of this
distributional choice. Importantly, note that if the moments of ui depend on x or z, then the
frontier correction will also depend on x and z implying that any features of the production
frontiers, such as returns to scale, may depend on the distribution of ui . One therefore needs
to either make some type of distributional assumption or to assume a type of separability
assumption, such as E[u|x, z] = E[u|z]. With the normal-half-normal framework, SVKZ
76 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

showed (adapting Olson et al. 1980), that


( r   1/3 )
π π
(6.11) σ
bu (x, z) = max 0, m
b 3 (x, z)
2 π−4
 
2 2 π−2
(6.12) σ
bv (x, z) =mb 2 (x, z) − σ
bu (x, z) ,
π

These estimates can then be used to obtain the estimates of the efficiency scores for each
observation, in the spirit of Jondrow et al. (1982), generalized to the heteroskedastic case
involving E[ui |εi , xi , zi ] instead of E[ui |εi ]. However, as mentioned in the parametric con-
text above, one should be careful interpreting these estimates of efficiency scores, as they
are “predicted values” conditional on unobserved εi , replaced with its estimate for the spe-
cific realization i, and as such the prediction intervals tend to be quite wide (see Simar &
Wilson (2010), for related discussion). In turn, the conditional mean of inefficiency can be
consistently estimated as
r
2
(6.13) µ
bu (x, z) = σ
bu (x, z).
π

and then use it at any point of interest (x, z) to form a consistent estimate of the level of
frontier, m(x, z), using

(6.14) m(x,
b b ∗ (x, z) + µ
z) = m bu (x, z).

SVKZ also derived the asymptotic properties of these estimators, generalizing earlier results
from Fan & Yao (1998) and Chen, Cheng & Peng (2009).
Finally, and perhaps most interestingly, SVKZ pointed out that if one is only interested
in the influence of z or x on the (conditional mean) efficiency, or as a special case to test if
E[u|x, z] is a constant, then no parametric distributional specification is required for ui , only
a condition that it belongs to the one parameter scale family of distributions. Specifically,
77

they showed that an elasticity of E[u|x, z], ψ` , defined as

∂µu (x, z) ψ`
(6.15) ξψ` (x, z) =
∂ψ` µu (x, z)

assuming that µu (x, z) 6= 0, can be estimated as

1 ∂m
c 3 (x, z) ψ`
(6.16) ξbψ` (x, z) =
3 ∂ψ` m
b 3 (x, z)

where m
b 3 (x, z) and ∂m
c 3 (x, z)/∂ψ` , are the estimates from the local polynomial estimator

b 3 (x, z) 6= 0 for the particular combination of interest (x, z). Impor-


and provided that m
tantly, SVKZ also derived the asymptotic law for this elasticity estimator, showing that

(nhp+d+2 )1/2 ξbψ` (x, z) − ξψ` (x, z) −→ N (0, s2ξ` (x, z)),

(6.17)

In turn, these asymptotic results can be used for statistical testing about influence of
elements in (x, z) onto expected inefficiency.
A practical limitation of SVKZ is that the estimated production technology may not satisfy
axioms of production. One might be tempted to follow Kuosmanen & Kortelainen (2012)
or Parmeter & Racine (2012), imposing the desired constraints first, and then recovering
E[u|x,
b z]. However, as we noted earlier, the methods of Kuosmanen & Kortelainen (2012)
and Parmeter & Racine (2012) work when the distribution of inefficiency is independent of
x and z, i.e. when u is homoskedastic. The issue the applied researcher faces here is much
more subtle. When heteroskedasticity is present in u, one must recognize that what is being
estimated in the first stage is a conditional mean, and not a production frontier. Thus, it
is not necessarily the case that the axioms of production should be expected to hold when
estimating the conditional mean.
Consider the case of a monotonic production function. The conditional mean of output
could be non-monotonic if E[u|x,
b z] was non-monotonic, even though the production function
78 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

is monotonic. Further, it is well known that adding two concave functions might not produce
a concave function, so even if E[u|x,
b z] was concave, adding it to the production frontier
may not produce a concave production function. And therein lies the danger of imposing
constraints when estimating the conditional mean, it is not necessarily the case that they
should be satisfied. This might seem innocuous except for the fact that imposing constraints
on a conditional mean that are incorrect will not produce a consistent estimator and typically,
consistent estimates in the first stage are needed for the second stage (recovering inefficiency)
to produce valid estimates.
Take for example the discussion in Kuosmanen, Johnson & Saastamoinen (2015, pg. 233),
who consider estimation of a production frontier nonparametrically, while also allowing u to
depend on x. In this case they stated (in our notation) “. . . Note that the shape of function
g can differ from that of frontier m because E(ui |xi ) is a function of inputs x . . . It is also
worth noting that function g is not necessarily monotonic increasing and concave even if
the production function m satisfies these axioms because −E(ui |xi ) can be a non-monotonic
and non-concave function of inputs . . . To apply CNLS in step 1, we need to assume that
the curvature of the production function m dominates and that function g is monotonic
increasing and concave (at least by approximation).” Unless the conditional mean of
output satisfies the axioms of production, it is recommended the axiomatic restrictions be
enforced after consistent, unrestricted estimation of the conditional mean as this will ensure
that the first stage estimator of the conditional mean is consistent. How exactly to do this is
a relatively unexplored area in stochastic frontier analysis and is a fruitful avenue for future
research.
Figures 10.1-10.3 illustrate the pitfalls of enforcing constraints ex ante on the conditional
mean of y (given x). We have a single input, x, and our production frontier is logarithmic,
which is naturally monotonic and concave. When inefficiency is homoskedastic we see that
the conditional mean is just a shift down of the production frontier, and remains both
monotone and concave. However, if we allow heteroskedasticity of inefficiency, e.g. through
79

a quadratic relationship, then, depending on the nature of heteroskedasticity, we can violate


monotonicity, Figure 10.2, or concavity, Figure 10.3 of E[y|x]. This quadratic relationship
is not beyond the pale, even in the parametric setting.44

[Figure 1 about here.]

[Figure 2 about here.]

[Figure 3 about here.]

6.4. Avoiding Distributional and (Some) Parametric Assumptions When Deter-


minants of Inefficiency Are Present. Here we discuss the approach of Tran & Tsionas
(2009) and Parmeter, Wang & Kumbhakar (2017). Let the SFM be:

(6.18) yi = m(xi ) + vi − ui = m(xi ) + vi − ui + E[ui |zi ] − E[ui |zi ] = m∗ (xi , zi ) + ε∗i .

where m∗ (xi , zi ) = m(xi ) + g(zi ), (ui |zi = z) ∼ D+ (µu (x, z), σu2 (x, z)), while (vi |xi , zi ) ∼
D(0, σv2 ). This model is a special case of SVKZ’s model. Now, if we specify our production
technology as m(xi ) = x0i β and E[ui |zi ] = g(zi ), then if β were known, g(zi ) could be
identified as the conditional mean of ε̃i = yi − x0i β given zi .
However, β is unknown and must be estimated. It can be estimated as follows. Condi-
tioning only on zi in equation (6.18) we have

(6.19) E[yi |zi ] = E[xi |zi ]0 β − g(zi ).

Subtracting (6.19) from (6.18) yields

(6.20) yi − E[yi |zi ] = (xi − E[xi |zi ])0 β + εi .

If E[yi |zi ] and E[xi |zi ] were known, β could be estimated via OLS from (6.20). The idea is to
replace the unknown conditional means with their nonparametric estimates (Robinson 1988).
44Wang (2002) documents non-monotonic efficiency effects in a panel of Philippine rice farmers based on the
age of the farmer.
80 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

To estimate both β and g(zi ) we replace E[yi |zi ] and E[xi |zi ] in (6.20) with
n
X
Ê[y|zi ] = Aj (zi )yj
j=1
n
X
Ê[xs |zi ] = Aj (zi )xsj ,
j=1

For a given bandwidth, the conditional expectations for y and each element of x can be
estimated and OLS can then be used to obtain a consistent estimator of β. That is, instead
of the usual regression of y on x, one performs the modified OLS regression of ỹ on x̃, where
we have used the notation w̃ = w − Ê[w|z] to denote a random variable that has been con-
ditionally demeaned. The estimates for β can then be used to obtain a consistent estimator
of the conditional mean of inefficiency via standard nonparametric regression techniques.
Let ε̌i = yi − x0i β,
b where βb is our estimate from the OLS regression of ỹ on x̃. We then

estimate g(zi ) nonparametrically via local-polynomial least-squares as


n
X
(6.21) gb(zi ) = Aj (zi )ε̌j .
j=1

In the cross-sectional regression setting, without assuming some structure on the distribu-
tions of the error components, it is not possible to identify the impact that any given variable
has on output directly, i.e. through the frontier, indirectly through inefficiency or both.45
One way to achieve identification is through invocation of the separability assumption. This
assumption, described in exceeding detail in Simar & Wilson (2007), essentially requires two
distinct sets of variables: those which influence the frontier and those which solely influence
inefficiency. In the context of a model for which two-sided noise does not exist (the standard
DEA framework), when this assumption is satisfied, a two-step approach is available which

45Hall & Simar (2002) discussed nonparametric identification of the mean of inefficiency subject to the
variance of the noise distribution diminishing as n → ∞. Horrace & Parmeter (2011) showed how to
nonparametrically identify the full distribution of inefficiency if one assumes that v is distributed normal.
81

can produce consistent estimators of both the frontier function and the inefficiency of a firm
(Simar & Wilson 2007, Banker & Natarajan 2008, Simar & Wilson 2011).
In general it is recommended that if variables which influence inefficiency exist, that this
information should be used directly, with a single stage estimator, such as maximum like-
lihood. When the separability assumption holds, then the partly linear model of Tran &
Tsionas (2009) and Parmeter et al. (2017) could be deployed (albeit with some parametric
assumptions imposed) or the additive model previously described can be used.46
Importantly, the separability assumption can be tested in the stochastic frontier context,
including the fully nonparametric or semiparametric frameworks. We can compare the es-
timates from the additively separable SFM, with that from a fully nonparametric model to
determine if there are statistical differences. Fortunately, this type of setup is conducive to
inference through either a residual sum of squares test or a conditional moment test. See
the discussion in chapter 6 of Henderson & Parmeter (2015a).

6.5. Future Directions in Semi- and Nonparametric Estimation and Inference


of the SFM. One of the future directions of research within non- and semiparametric SFA
is, naturally, related to statistical inference. The asymptotic results developed in the above
mentioned papers as well as various testing procedures developed in the general statistics
community make a solid foundation for this to happen, with careful adaptation and extensive
Monte Carlo evidence supporting the theory. Additionally, few of the methods discussed here
have been fully developed in the panel data setting.

46
The approach of SVKZ allows for both x and z to influence both the frontier and inefficiency and as such
the separability assumption is not required. Yet, one may say that there is also a kind of ‘separability’
structure involved implicitly: (x, z) is assumed to influence the frontier via the first moment, while for the
inefficiency term, u, the same (x, z) is modeled through the skedastic function defining the second moment.
Besides helping with statistical identification, such structure can be viewed as quite natural to the context
of measurement. Indeed, one often thinks of the frontier as the level, and so using the (conditional) first
moment, measuring the (conditional) average level of outputs, would be very natural. Meanwhile, the
inefficiency is often understood as the deviation from the frontier, so it would be a more natural way to
model it with the second moment. In addition, one could also think of the inefficiency as a reflection of
the uncertainty and related ‘risk’ to produce less than the potential and beyond the usual (and symmetric)
noise, and it is very common to model risk through the second moment.
82 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

It is worth noting that neither Kumbhakar et al. (2007), nor Martins-Filho & Yao (2015),
nor Park et al. (2015), nor SVKZ imposed any axioms of production on the frontier, e.g.,
monotonicity (i.e., require ∇mx ≥ 0 ∀x), although some of them have brief discussions
about possible extensions to do so. Specifically, to impose the desired constraints, one could
adapt ideas from Daouia & Simar (2005) and Daouia & Park (2013), or use DEA or FDH
on the fitted values from these methods (thus using the stochastic DEA or stochastic FDH
approaches of Simar & Zelenyuk 2011), or to employ the constraint weighted bootstrapping
(Hall & Huang 2001, Du et al. 2013), as was adapted to the baseline SFM by Parmeter &
Racine (2012).

7. Quantile Estimation of the SFM

A recent development in the estimation of the SFM has been to embrace the use of quantile
methods (Bernini, Freo & Gardini 2004, Know, Blankmeyer & Stutzman 2007, Liu, Laporte
& Ferguson 2008, Behr 2010). Quantile regression is known to provide a more complete
picture of a conditional distribution (Koenker & Hallock 2001, Koenker 2005) and provides
a robust alternative to ordinary least squares. Whereas the ordinary least squares estimator
stems from minimization of the sum of squared errors, the conditional quantile estimator is
determined through minimization of the “check” function (Koenker & Bassett 1978) defined
for a particular quantile, the median say.
The conditional quantile function Qy (τ |x) for a random variable y with conditional CDF
F (y|x) is defined as F −1 (τ |x) = inf {y : F (y|x) ≥ τ } where τ is the τ th conditional quantile
of the random variable y. Rather than directly inverting of the conditional distribution
function, the conditional quantile can be determined through the loss function

(7.1) ρτ () =  (τ − 1{ < 0}) .


83

ρτ () is known as the check function. For a traditional linear in parameters framework,
Qy (τ |x) = x0i β(τ ), the quantile estimator is found by minimizing
n
X
(7.2) min ρτ (yi − x0i β(τ )) ,
β
i=1

for a given τ . When the error terms are i.i.d., the conditional quantiles represent vertical
shifts of the conditional median function by the appropriate quantile of the error distribution.
However, when heteroskedasticity is present, the conditional quantiles are no longer vertical
shifts of the conditional median, but will have varying slopes; moreover, the quantiles will
become nonlinear.
The use of conditional quantile estimation to recover the frontier is appealing because in
general a frontier can be thought of as a quantile in the distribution of output. At issue is
the appropriate quantile, τ . For example, Bernini et al. (2004, pg. 379) estimate the frontier
with the conditional quantile estimator using τ = 0.5, 0.9 and 0.975. τ = 0.5 corresponds
to the median and is equivalent to the conditional mean in the case that σu2 = 0 (see the
discussion in Horrace & Parmeter 2014). Know et al. (2007, pg. 79) estimate conditional
quantiles for τ = 0.85, 0.9 and 0.95, while Liu et al. (2008, pg. 1080) consider τ = 0.5 and
0.8. Lastly, Behr (2010, pg. 572) recommended use of τ = 0.95 for estimation of production
frontiers and τ = 0.05 for estimation of cost frontiers.
What is lost in the recommendations of this earlier research is how one estimates (or
predicts) individual efficiency once the frontier has been estimated. Currently the standard
practice is to treat any firm whose output lies above the frontier as fully efficient, and
any firm whose output is below the frontier as inefficient, with inefficiency defined as the
difference between the estimated frontier and observed output. However, both of these
recommendations ignore the fact that the composed error term represents inefficiency and
noise. There does not exist at present an approach that separates inefficiency from noise in
a manner similar to Jondrow et al. (1982). One idea could be to use the conditional mode
84 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

as proposed in Materov (1981). This estimator can be interpreted as a maximum likelihood


estimator for the distribution of the joint density of v and u, and more importantly, for
positive residuals, it is always 0, which is akin to how inefficiency is currently calculated
using conditional quantile estimation. Unfortunately, as with the conditional mean, the
conditional mode estimator requires distributional assumptions for it to be operational.
Lastly, we mention two important caveats with quantile estimation of frontiers. First,
heteroskedasticity in either v or u has, until now, not been accounted for. This is a severe
limitation as heteroskedasticity is commonly seen as present in v in applied efficiency studies,
and researchers typically have access to an array of determinants of inefficiency, which induce
heteroskedasticity in the inefficiency term. Moreover, unlike estimation of a conditional
mean, when conditional heteroskedasticity is present, this can affect consistent estimation of
the conditional quantile. Second, estimation of the conditional quantile for a specific value
of τ is an implicit assumption on the ratio of signal to noise between σu2 and σv2 . To see this,
more clearly, Figures 10.4-10.7 present the results of quantile estimation for τ = 0.5, 0.8,
0.85, 0.9, and 0.95 for 1,000 observations drawn from the model

vi −ui
(7.3) yi = x0.4
i e ,

with vi ∼ N (0, 1) and ui ∼ N+ (0, σu2 ). In Figure 10.4 the inefficiency draws are taken with
σu2 = 0.01, in Figure 10.5 we have σu2 = 0.25, in Figure 10.6 σu2 = 1, and in Figure 10.7 σu2 = 4.
In the case where σu2 = 4, this corresponds to a λ = σu /σv = 2 which is of a decent size for
an applied efficiency study. In this case the true frontier is approximately equal to the 85th
quantile. It is clear that interpreting the frontier for a given quantile as the benchmark for a
firm being efficient or inefficient is implicitly a statement on the ratio between the variance
of the noise and the inefficiency for the sample. In Figure 10.4, where λ = 0.01, the situation
where there is almost no inefficiency, the frontier is nearly equivalent to the median, which
is the least absolute deviation estimator that Horrace & Parmeter (2014) discussed.
85

[Figure 4 about here.]

[Figure 5 about here.]

[Figure 6 about here.]

[Figure 7 about here.]

While the quantile estimator marks an interesting and robust alternative to traditional
stochastic frontier analysis, it should be clear that more work needs to be done. We direct the
reader to the earlier referenced papers for more details and additional insights on how best to
use conditional quantile methods at present for conducting efficiency analysis. Furthermore,
panel estimation of quantiles, as well as semi and non-parametric estimation of quantiles,
is still in its infancy in this area and extensions to the SFM have as yet to appear in the
literature.

8. Additional Approaches/Extensions of the SFM

As with any review or summary article, there is never enough space to cover all top-
ics equally or broadly enough. The SFM has been studied and used for 40 years now
and even though we have covered a range of approaches and insights, there are still many
topics which we did not cover. These include finite mixture models (Caudill 2003, Orea
& Kumbhakar 2004, Greene 2005b), the zero-inefficiency SFM (Kumbhakar, Parmeter &
Tsionas 2013), the meta-frontier (Battese & Rao 2002, Battese, Rao & O’Donnell 2004),
total factor productivity change and its individual components (Hulten 2001), the two-tier
frontier (Polachek & Yoon 1987, Polachek & Yoon 1996), sample selection in the SFM
(Kumbhakar, Tsionas & Sipiläinen 2009, Greene 2010), and directional distance function
estimation (Atkinson & Tsionas 2016). Parmeter & Kumbhakar (2014) cover broadly es-
timation and inference of finite mixture models, the zero-inefficiency SFM and issues per-
taining to sample selection. Full details on the measurement of total factor productiv-
ity and separation into distinct components can be found in Kumbhakar et al. (2015,
86 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

chapt. 11). Both the two-tier frontier (Kumbhakar & Parmeter 2009, Kumbhakar &
Parmeter 2010, Papadopoulos 2015) and meta-frontier (O’Donnell, Rao & Battese 2008, Am-
sler, O’Donnell & Schmidt 2017) have started to receive more attention recently, but as of
yet, no broad review of either exists. Regarding the estimation of directional distance func-
tions, we refer interested readers to Färe, Martins-Filho & Vardanyan (2010) for a thorough
treatment.

9. Available Software to Estimate SFMs

Despite the popularity of the SFM, only the most basic implementations of it are available
across a wide array of statistical platforms. For example, in the R programming environment
the frontier (Coelli & Henningsen 2013) package allows for cross-sectional estimation of
the SFM assuming either the half-normal or truncated-normal distribution for ui and the
Battese & Coelli (1992) and Battese & Coelli (1995) panel data estimators of the SFM are
implemented.47 There are similar estimators available in LIMDEP through the NLOGIT
module but these also include the normal-gamma specification as well as the true fixed and
true random effects estimators along with the latent class stochastic frontier estimator. There
are also several modules in the STATA software as described in Kumbhakar et al. (2015)
which implement several other panel data estimators as described earlier. Additionally, many
authors provide their own personal codes. For example, Young Hoon Lee provides GAUSS
code for a variety of cross-sectional and panel data stochastic frontier estimators on his
webpage complete with several datasets (https://sites.google.com/site/yhnlee3/SFM-code).
Federico Belotti provides integrated STATA code which works with the basic frontier

47The frontier package accesses the Frontier V4.1 Fortran codes originally developed by Tim Coelli, which
is also freely available (at http://www.uq.edu.au/economics/cepa/frontier.php), although fairly outdated by
now.
87

capabilities. These new codes are sfcross and sfpanel and can be obtained through his
48
blog http://www.econometrics.it.
However, there does not yet exist a singular software that implements all of the available
estimators described here. This should not be surprising. As with any applied field, as
statistical improvements are made, there is a lag with available software and the array of
options makes it infeasible to include all discussed models in a singular package. Researchers
interested in the newest methods can invest in programming these methods and disseminating
them to the field, or can collaborate with the authors of the original models to develop
software that can be made widely available, and we strongly encourage researchers to do so.

10. Conclusions

This review was meant to highlight some of the most important econometric developments
over the past 40 years which improve the estimation of measurements of productivity and
efficiency. We covered the workhorse SFM, and discussed avenues to include determinants
of inefficiency and productivity, dealing with endogeneity, what to do when one has panel
data, quantile estimation, and robust methods involving nonparametric regression and local-
likelihood to place as few restrictions as possible on the frontier or the behavior of inefficiency.
All told, a variety of methods and models exist for the practitioner and our hope is that
this review will encourage applied researchers to move away from some of the basic SFMs in
search of more robust and insightful conclusions.
While much has been covered, much remains unsaid. Important areas that are still being
developed include modeling dependence between statistical noise and inefficiency, selection
of firm technology, handling heterogeneous technology in a sample of firms, and how to allow
a subset of firms to be fully efficient. While our discussion was couched in terms of the single
equation stochastic production frontier, system based approaches surrounding cost, profit,
48One can install these commands via
net install sfcross, all from(http://www.econometrics.it/stata)
net install sfpanel, all from(http://www.econometrics.it/stata)
88 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

or revenue frontiers are also available and, similar to the other methods mentioned earlier
without any details, they deserve attention and separate reviews.
89

References
Afriat, S. N. (1972), ‘Efficiency estimation of production functions’, International Economic Review
13(3), 568–598.
Ahmad, I. A. & Li, Q. (1997), ‘Testing symmetry of an unknown density function by kernel method’, Journal
of Nonparametric Statistics 7, 279–293.
Aigner, D. & Chu, S. (1968), ‘On estimating the industry production function’, American Economic Review
58, 826–839.
Aigner, D. J., Lovell, C. A. K. & Schmidt, P. (1977), ‘Formulation and estimation of stochastic frontier
production functions’, Journal of Econometrics 6(1), 21–37.
Aitchison, J. & Aitken, C. (1976), ‘Multivariate binary discrimination by the kernel method’, Biometrika
63, 413–420.
Ali, M. & Flinn, J. C. (1989), ‘Profit efficiency among Basmati rice producers in Pakistan Punjab’, American
Journal of Agricultural Economics 71(2), 303–310.
Almanidis, P., Qian, J. & Sickles, R. C. (2014), Stochastic frontier models with bounded inefficiency, in
R. C. Sickles & W. C. Horrace, eds, ‘Festschrift in Honor of Peter Schmidt Econometric Methods and
Applications’, Springer: New York, pp. 47–82.
Almanidis, P. & Sickles, R. C. (2011), The skewness issue in stochastic frontier models: Fact or fiction?, in
I. van Keilegom & P. W. Wilson, eds, ‘Exploring Research Frontiers in Contemporary Statistics and
Econometrics’, Springer Verlag, Berlin.
Alvarez, A., Amsler, C., Orea, L. & Schmidt, P. (2006), ‘Interpreting and testing the scaling property in
models where inefficiency depends on firm characteristics’, Journal of Productivity Analysis 25(2), 201–
212.
Amemiya, T. (1974), ‘The nonlinear two-stage least-squares estimator’, Journal of Econometrics 2, 105–111.
Amsler, C., O’Donnell, C. J. & Schmidt, P. (2017), ‘Stochastic metafrontiers’, Econometric Reviews 36, 1007–
1020.
Amsler, C., Prokhorov, A. & Schmidt, P. (2016), ‘Endogeneity in stochastic frontier models’, Journal of
Econometrics 190, 280–288.
Amsler, C., Prokhorov, A. & Schmidt, P. (2017), ‘Endogeneity environmental variables in stochastic frontier
models’, Journal of Econometrics 199, 131–140.
Atkinson, S. E. & Tsionas, E. G. (2016), ‘Direcitonal disance functions: Optimal endogenous directions’,
Journal of Econometrics 190, 301–314.
Azzalini, A. (1985), ‘A class of distributions which includes the normal ones’, Scandinavian Journal of
Statistics 12(2), 171–178.
Badunenko, O. & Kumbhakar, S. C. (2017), ‘Economies of scale, technical change and persistent and time-
varying cost efficiency in Indian banking: Do ownership, regulation and heterogeneity matter?’, Euro-
pean Journal of Operational Research 260, 789–803.
Baltagi, B. H. (2013), Econometric Analysis of Panel Data, 5th edn, John Wiley & Sons, Great Britain.
Banker, R. D. & Maindiratta, A. (1992), ‘Maximum likelihood estimation of monotone and concave produc-
tion frontiers’, Journal of Productivity Analysis 3(4), 401–415.
Banker, R. D. & Natarajan, R. (2008), ‘Evaluating contextual variables affecting productivity using data
envelopment analysis’, Operations Research 56(1), 48–58.
Battese, G. E. & Coelli, T. J. (1988), ‘Prediction of firm-level technical efficiencies with a generalized frontier
production function and panel data’, Journal of Econometrics 38, 387–399.
Battese, G. E. & Coelli, T. J. (1992), ‘Frontier production functions, technical efficiency and panel data:
With application to paddy farmers in India’, Journal of Productivity Analysis 3, 153–169.
Battese, G. E. & Coelli, T. J. (1995), ‘A model for technical inefficiency effects in a stochastic frontier
production function for panel data’, Empirical Economics 20(1), 325–332.
Battese, G. E. & Corra, G. S. (1977), ‘Estimation of a production frontier model: With application to the
pastoral zone off Eastern Australia’, Australian Journal of Agricultural Economics 21(3), 169–179.
90 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

Battese, G. E. & Rao, D. S. P. (2002), ‘Technology gap, efficiency and a stochastic metafrontier function’,
International Journal of Business and Economics 1, 1–7.
Battese, G. E., Rao, D. S. P. & O’Donnell, C. J. (2004), ‘A metafrontier production function for estimation
of technical efficiencies and technology gaps for firms operating under different technologies’, Journal
of Productivity Analysis 21, 91–103.
Behr, A. (2010), ‘Quantile regression for robust bank efficiency score estimation’, European Journal of
Operational Research 200, 568–581.
Benabou, R. & Tirole, J. (2016), ‘Mindful economics: The production, consumption, and value of beliefs’,
Journal of Economic Perspectives 30(3), 141–164.
Bera, A. K. & Sharma, S. C. (1999), ‘Estimating production uncertainty in stochastic frontier production
function models’, Journal of Productivity Analysis 12(2), 187–210.
Bernini, C., Freo, M. & Gardini, A. (2004), ‘Quantile estimation of frontier production function’, Empirical
Economics 29, 373–381.
Bloom, N., Lemos, R., Sadun, R., Scur, D. & Van Reenen, J. (2016), ‘International data on measuring
management practices’, American Economic Review 106(5), 152–156.
Bonanno, G., De Giovanni, D. & Domma, F. (2017), ‘The ‘wrong skewness’ problem: a re-specification of
stochastic frontiers’, Journal of Productivity Analysis 47(1), 49–64.
Bravo-Ureta, B. E. & Rieger, L. (1991), ‘Dairy farm efficiency measurement using stochastic frontiers and
neoclassical duality’, American Journal of Agricultural Economics 73(2), 421–428.
Butler, J. & Moffitt, R. (1982), ‘A computationally efficient quadrature procedure for the one factor multi-
nomial probit model’, Econometrica 50, 761–764.
Carree, M. A. (2002), ‘Technological inefficiency and the skewness of the error component in stochastic
frontier analysis’, Economics Letters 77(1), 101–107.
Case, B., Ferrari, A. & Zhao, T. (2013), ‘Regulatory reform and productivity change in indian banking’, The
Review of Economics and Statistics 95(3), 1066–1077.
Caudill, S. B. (2003), ‘Estimating a mixture of stochastic frontier regression models via the EM algorithm:
A multiproduct cost function application’, Empirical Economics 28(1), 581–598.
Caudill, S. B. & Ford, J. M. (1993), ‘Biases in frontier estimation due to heteroskedasticity’, Economics
Letters 41(1), 17–20.
Caudill, S. B., Ford, J. M. & Gropper, D. M. (1995), ‘Frontier estimation and firm-specific inefficiency
measures in the presence of heteroskedasticity’, Journal of Business & Economic Statistics 13(1), 105–
111.
Chamberlain, G. (1987), ‘Asymptotic efficiency in estimation with conditional moment restrictions’, Journal
of Econometrics 34(2), 305–334.
Chen, L.-H., Cheng, M.-Y. & Peng, L. (2009), ‘Conditional variance estimation in heteroscedastic regression
models’, Journal of Statistical Planning and Inference 139(2), 236–245.
Chen, Y.-Y., Schmidt, P. & Wang, H.-J. (2014), ‘Consistent estimation of the fixed effects stochastic frontier
model’, Journal of Econometrics 181(2), 65–76.
Chu, C.-Y., Henderson, D. J. & Parmeter, C. F. (2017), ‘On discrete Epanechnikov kernels’, Computational
Statistics and Data Analysis . forthcoming.
Coelli, T. & Henningsen, A. (2013), frontier: Stochastic Frontier Analysis. R package version 1.1-0.
URL: http://CRAN.R-Project.org/package=frontier
Coelli, T. J. (1995), ‘Estimators and hypothesis tests for a stochastic frontier function: A Monte Carlo
analysis’, Journal of Productivity Analysis 6(4), 247–268.
Colombi, R., Kumbhakar, S., Martini, G. & Vittadini, G. (2014), ‘Closed-skew normality in stochastic fron-
tiers with individual effects and long/short-run efficiency’, Journal of Productivity Analysis 42(2), 123–
136.
Colombi, R., Martini, G. & Vittadini, G. (2011), A stochastic frontier model with short-run and long-
run inefficiency random effects. Department of Economics and Technology Management, University of
Bergamo, Working Paper Series.
91

Cornwell, C. & Schmidt, P. (1992), ‘Models for which the MLE and the conditional MLE coincide’, Empirical
Economics 17(2), 67–75.
Cornwell, C., Schmidt, P. & Sickles, R. C. (1990), ‘Production frontiers with cross-sectional and time-series
variation in efficiency levels’, Journal of Econometrics 46(2), 185–200.
Cuesta, R. A. (2000), ‘A production model with firm-specific temporal variation in technical inefficiency:
With application to Spanish dairy farms’, Journal of Productivity Analysis 13, 139–152.
Daouia, A. & Park, B. U. (2013), ‘On projection-type estimators of multivariate isotonic functions’, Scandi-
navian Journal of Statistics 40, 363–386.
Daouia, A. & Simar, L. (2005), ‘Robust nonparametric estimators of monotone boundaries’, Journal of
Multivariate Analysis 96(2), 311–331.
Domı́nguez-Molina, J. A., González-Farı́as, G. & Ramos-Quiroga, R. (2003), Skew normality in stochastic
frontier analysis. Comunicación Técnica No I-03-18/06-10-2003 (PE/CIMAT).
Du, P., Parmeter, C. F. & Racine, J. S. (2013), ‘Nonparametric kernel regression with multiple predictors
and multiple shape constraints’, Statistica Sinica 23(3), 1347–1371.
Dugger, R. (1974), An application of bounded nonparametric estimating functions to the analysis of bank
cost and production functions, PhD thesis, University of North Carolina , Chapel Hill.
Fan, J. & Gijbels, I. (1996), Local Polynomial Modelling and its Application, Chapman and Hall.
Fan, J. & Yao, Q. (1998), ‘Efficient estimation of conditional variance functions in stochastic regression’,
Biometrika 85, 645–660.
Fan, Y., Li, Q. & Weersink, A. (1996), ‘Semiparametric estimation of stochastic production frontier models’,
Journal of Business & Economic Statistics 14(4), 460–468.
Färe, R., Martins-Filho, C. & Vardanyan, M. (2010), ‘On functional form representation of multi-output
production technologies’, Journal of Productivity Analysis 33(1), 81–96.
Feng, Q., Horrace, W. C. & Wu, G. L. (2015), Wrong skewness and finite sample correction in parametric
stochastic frontier models. Center for Policy Research - The Maxwell School, working paper N. 154.
Gabrielsen, A. (1975), On estimating efficient production functions. Working Paper No. A-85, Chr. Michelsen
Institute, Department of Humanities and Social Sciences, Bergen, Norway.
Gagnepain, P. & Ivaldi, M. (2002), ‘Stochastic frontiers and asymmetric information models’, Journal of
Productivity Analysis 18(2), 145–159.
Greene, W. (2004), ‘Distinguishing between heterogeneity and inefficiency: stochastic frontier analysis of the
World Health Organization’s panel data on national health care systems’, Health Economics 13(9), 959–
980.
Greene, W. H. (1980a), ‘Maximum likelihood estimation of econometric frontier functions’, Journal of Econo-
metrics 13(1), 27–56.
Greene, W. H. (1980b), ‘On the estimation of a flexible frontier production model’, Journal of Econometrics
13(1), 101–115.
Greene, W. H. (1990), ‘A gamma-distributed stochastic frontier model’, Journal of Econometrics 46(1-
2), 141–164.
Greene, W. H. (2003), ‘Simulated likelihood estimation of the normal-gamma stochastic frontier function’,
Journal of Productivity Analysis 19(2), 179–190.
Greene, W. H. (2005a), ‘Fixed and random effects in stochastic frontier models’, Journal of Productivity
Analysis 23(1), 7–32.
Greene, W. H. (2005b), ‘Reconsidering heterogeneity in panel data estimators of the stochastic frontier
model’, Journal of Econometrics 126(2), 269–303.
Greene, W. H. (2008), The econometric approach to efficiency analysis, in C. A. K. L. H. O. Fried & S. S.
Schmidt, eds, ‘The Measurement of Productive Efficiency and Productivity Change’, Oxford University
Press, Oxford, United Kingdom, chapter 2.
Greene, W. H. (2010), ‘A stochastic frontier model with correction for sample selection’, Journal of Produc-
tivity Analysis 34(1), 15–24.
Greene, W. H. & Fillipini, M. (2014), Persistent and transient productive inefficiency: A maximum simulated
likelihood approach. CER-ETH - Center of Economic Research at ETH Zurich, Working Paper 14/197.
92 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

Hadri, K. (1999), ‘Estimation of a doubly heteroscedastic stochastic frontier cost function’, Journal of Busi-
ness & Economic Statistics 17(4), 359–363.
Hafner, C., Manner, H. & Simar, L. (2016), ‘The “wrong skewness” problem in stochastic frontier model: a
new approach’, Econometric Reviews . forthcoming.
Hall, P. & Huang, H. (2001), ‘Nonparametric kernel regression subject to monotonicity constraints’, The
Annals of Statistics 29(3), 624–647.
Hall, P. & Simar, L. (2002), ‘Estimating a changepoint, boundary or frontier in the presence of observation
error’, Journal of the American Statistical Association 97, 523–534.
Hansen, C., McDonald, J. B. & Newey, W. K. (2010), ‘Instrumental variables estimation with flexible
distributions’, Journal of Business and Economic Statistics 28, 13–25.
Hattori, T. (2002), ‘Relative performance of U.S. and Japanese electricity distribution: An application of
stochastic frontier analysis’, Journal of Productivity Analysis 18(3), 269–284.
Henderson, D. J. & Parmeter, C. F. (2015a), Applied Nonparametric Econometrics, Cambridge University
Press, Cambridge, Great Britain.
Henderson, D. J. & Parmeter, C. F. (2015b), ‘A consistent bootstrap procedure for nonparametric symmetry
tests’, Economics Letters 131, 78–82.
Hjalmarsson, L., Kumbhakar, S. C. & Heshmati, A. (1996), ‘DEA, DFA, and SFA: A comparison’, Journal
of Productivity Analysis 7(2), 303–327.
Hollingsworth, B. (2008), ‘The measurement of efficiency and productivity of health care delivery’, Health
Economics 17(10), 1107–1128.
Horrace, W. C. & Parmeter, C. F. (2011), ‘Semiparametric deconvolution with unknown error variance’,
Journal of Productivity Analysis 35(2), 129–141.
Horrace, W. C. & Parmeter, C. F. (2014), A Laplace stochastic frontier model. University of Miami Working
Paper.
Horrace, W. C. & Schmidt, P. (1996), ‘Confidence statements for efficiency estimates from stochastic frontier
models’, Journal of Productivity Analysis 7, 257–282.
Horrace, W. C. & Wright, I. A. (2016), Stationary points for parametric stochastic frontier models. Center
for Policy Research - The Maxwell School, working paper N. 196.
Huang, C. J. & Liu, J.-T. (1994), ‘Estimation of a non-neutral stochastic frontier production function’,
Journal of Productivity Analysis 5(1), 171–180.
Hulten, C. R. (2001), Total factor productivity. a short biography, in C. R. Hulten, E. R. Dean & M. J.
Harper, eds, ‘New Developments in Productivity Analysis’, University of Chicago Press, Chicago, IL,
pp. 1–54.
Jondrow, J., Lovell, C. A. K., Materov, I. S. & Schmidt, P. (1982), ‘On the estimation of technical efficiency
in the stochastic frontier production function model’, Journal of Econometrics 19(2/3), 233–238.
Kalirajan, K. P. (1990), ‘On measuring economic efficiency’, Journal of Applied Econometrics 5(1), 75–85.
Kim, M. & Schmidt, P. (2008), ‘Valid test of whether technical inefficiency depends on firm characteristics’,
Journal of Econometrics 144(2), 409–427.
Kneip, A. & Simar, L. (1996), ‘A general framework for frontier estimation with panel data’, Journal of
Productivity Analysis 7(2), 187–212.
Kneip, A., Simar, L. & Van Keilegom, I. (2015), ‘Frontier estimation in the presence of measurement error
with unknown variance’, Journal of Econometrics 184, 379–393.
Knittel, C. R. (2002), ‘Alternative regulatory methods and firm efficiency: Stochastic frontier evidence form
the U.S. electricity industry’, The Review of Economics and Statistics 84(3), 530–540.
Know, K. J., Blankmeyer, E. C. & Stutzman, J. R. (2007), ‘Technical efficiency in Texan nursing facilities:
a stochastic production frontier approach’, Journal of Economics and Finance 31(1), 75–86.
Koenker, R. (2005), Quantile Regression, Cambridge University Press.
Koenker, R. & Bassett, G. (1978), ‘Regression quantiles’, Econometrica 46(1), 33–50.
Koenker, R. & Hallock, K. (2001), ‘Quantile regression’, Journal of Economic Perspectives 15, 143–156.
Kumbhakar, S. C. (1987), ‘The specification of technical and allocative inefficiency in stochastic production
and profit frontiers’, Journal of Econometrics 34(1), 335–348.
93

Kumbhakar, S. C. (1990), ‘Production frontiers, panel data, and time-varying technical inefficiency’, Journal
of Econometrics 46(1), 201–211.
Kumbhakar, S. C. (1991), ‘The measurement and decomposition of cost-inefficiency: The translog cost
system’, Oxford Economic Papers 43(6), 667–683.
Kumbhakar, S. C. (2011), ‘Estimation of production technology when the objective is to maximize return to
the outlay’, European Journal of Operational Research 208, 170–176.
Kumbhakar, S. C. (2013), ‘Specification and estimation of multiple output technologies: A primal approach’,
European Journal of Operational Research 231, 465–473.
Kumbhakar, S. C., Ghosh, S. & McGuckin, J. T. (1991), ‘A generalized production frontier approach for
estimating determinants of inefficiency in US diary farms’, Journal of Business & Economic Statistics
9(1), 279–286.
Kumbhakar, S. C. & Heshmati, A. (1995), ‘Efficiency measurement in Swedish dairy farms: An application
of rotating panel data, 1976-88’, American Journal of Agricultural Economics 77(3), 660–674.
Kumbhakar, S. C. & Hjalmarsson, L. (1993), Technical efficiency and technical progress in Swedish dairy
farms, in K. L. H. Fried & S. Schmidt, eds, ‘The Measurement of Productive Efficiency’, Oxford Uni-
versity Press, Oxford, United Kingdom.
Kumbhakar, S. C. & Hjalmarsson, L. (1998), ‘Relative performance of public and private ownership under
yardstick competition: Electricity retail distribution’, European Economic Review 42(1), 97–122.
Kumbhakar, S. C., Lien, G. & Hardaker, J. B. (2014), ‘Technical efficiency in competing panel data models:
A study of Norwegian grain farming’, Journal of Productivity Analysis 41(2), 321–337.
Kumbhakar, S. C. & Lovell, C. A. K. (2000), Stochastic Frontier Analysis, Cambridge University Press.
Kumbhakar, S. C., Park, B. U., Simar, L. & Tsionas, E. G. (2007), ‘Nonparametric stochastic frontiers: A
local maximum likelihood approach’, Journal of Econometrics 137(1), 1–27.
Kumbhakar, S. C. & Parmeter, C. F. (2009), ‘The effects of match uncertainty and bargaining on labor
market outcomes: evidence from firm and worker specific estimates’, Journal of Productivity Analysis
31(1), 1–14.
Kumbhakar, S. C. & Parmeter, C. F. (2010), ‘Estimation of hedonic price functions with incomplete infor-
mation’, Empirical Economics 39(1), 1–25.
Kumbhakar, S. C., Parmeter, C. F. & Tsionas, E. (2013), ‘A zero inefficiency stochastic frontier estimator’,
Journal of Econometrics 172(1), 66–76.
Kumbhakar, S. C., Tsionas, E. G. & Sipiläinen, T. (2009), ‘Joint estimation of technology choice and technical
efficiency: an application to organic and conventional dairy farming’, Journal of Productivity Analysis
31(2), 151–161.
Kumbhakar, S. C. & Wang, H.-J. (2005), ‘Production frontiers, panel data, and time-varying technical
inefficiency’, Journal of Econometrics 46(1), 201–211.
Kumbhakar, S. C. & Wang, H.-J. (2006), ‘Estimation of technical and allocative inefficiency: A primal
system approach’, Journal of Econometrics 134(3), 419–440.
Kumbhakar, S. c., Wang, H.-J. & Horncastle, A. (2015), A Practitioners Guide to Stochastic Frontier Analysis
Using Stata, Cambridge University Press, Cambridge, United Kingdom.
Kuosmanen, T. (2012), ‘Stochastic semi-nonparametric frontier estimation of electricity distribution net-
works: Application of the StoNED method in the Finnish regulatory model’, Energy Economics
34, 2189–2199.
Kuosmanen, T. & Fosgerau, M. (2009), ‘Neoclassical versus frontier production models? Testing for the
skewness of regression residuals’, The Scandinavian Journal of Economics 111(2), 351–367.
Kuosmanen, T., Johnson, A. & Saastamoinen, A. (2015), Stochastic nonparametric approach to efficiency
analysis: A unified framework, in J. Zhu, ed., ‘Data Envelopment Analysis’, International Series in
Operations Research & Management Science, Springer Science, New York., chapter 7, pp. 191–244.
Kuosmanen, T. & Kortelainen, M. (2012), ‘Stochastic non-smooth envelopment of data: semi-parametric
frontier estimation subject to shape constraints’, Journal of Productivity Analysis 38(1), 11–28.
Kutlu, L. (2010), ‘Battese-Coelli estimator with endogenous regressors’, Economics Letters 109, 79–81.
94 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

Latruffe, L., Bravo-Ureta, B. E., Carpentier, A., Desjeux, Y. & Moreira, V. H. (2017), ‘Subsidies and tech-
nical efficiency in agriculture: Evidence from European dairy farms’, American Journal of Agricultural
Economics 99, 783–799.
Lee, L. (1983), ‘A test for distributional assumptions for the stochastic frontier function’, Journal of Econo-
metrics 22(2), 245–267.
Lee, L.-F. & Tyler, W. G. (1978), ‘The stochastic frontier production function and average efficiency: An
empirical analysis’, Journal of Econometrics 7, 385–389.
Lee, Y. & Schmidt, P. (1993), A production frontier model with flexible temporal variation in technical
efficiency, in K. L. H. Fried & S. Schmidt, eds, ‘The Measurement of Productive Efficiency’, Oxford
University Press, Oxford, United Kingdom.
Li, D., Simar, L. & Zelenyuk, V. (2016), ‘Generalized nonparametric smoothing with mixed discrete and
continuous data’, Computational Statistics and Data Analysis 100, 424–444.
Li, Q. (1996), ‘Estimating a stochastic production frontier when the adjusted error is symmetric’, Economics
Letters 52(3), 221–228.
Li, Q. & Racine, J. (2007), Nonparametric Econometrics: Theory and Practice, Princeton University Press.
Lien, G., Kumbhakar, S. C. & Hardaker, J. B. (2017), ‘Accounting for risk in productivity analysis: an
application to Norwegian dairy farming’, Journal of Productivity Analysis 47(3), 247–257.
Liu, C., Laporte, A. & Ferguson, B. S. (2008), ‘The quantile regression approach to efficiency measurement:
Insights from Monte Carlo simulations’, Health Economics 17, 1073–1087.
Lovell, C. A. K. (1993), Production frontiers and productive efficiency, in C. A. K. L. H. O. Fried & S. S.
Schmidt, eds, ‘The Measurement of Productive Efficiency’, Oxford University Press, Oxford, United
Kingdom, chapter 1.
Martins-Filho, C. B. & Yao, F. (2015), ‘Semiparametric stochastic frontier estimation via profile likelihood’,
Econometric Reviews 34(4), 413–451.
Materov, I. S. (1981), ‘On full identification of the stochastic production frontier model (in Russian)’,
Ekonomika i Matematicheskie Metody 17, 784–788.
McFadden, D. (1989), ‘A method of simulated moments for estimation of discrete response models without
numerical integration’, Econometrica 57(5), 995–1026.
Meeusen, W. & van den Broeck, J. (1977a), ‘Efficiency estimation from Cobb-Douglas production functions
with composed error’, International Economic Review 18(2), 435–444.
Meeusen, W. & van den Broeck, J. (1977b), ‘Technical efficiency and dimension of the firm: Some results on
the use of frontier production functions’, Empirical Economics 2(2), 109–122.
Mundlak, Y. (1961), ‘Empirical production function free of management bias’, Journal of Farm Economics
43(1), 44–56.
Mutter, R. L., Greene, W. H., Spector, W., Rosko, M. D. & Mukamel, D. B. (2013), ‘Investigating the
impact of endogeneity on inefficiency estimates in the application of stochastic frontier analysis to
nursing homes’, Journal of Productivity Analysis 39(1), 101–110.
Neyman, J. & Scott, E. L. (1948), ‘Consistent estimation from partially consistent observations’, Economet-
rica 16, 1–32.
Nguyen, N. B. (2010), Estimation of technical efficiency in stochastic frontier analysis, PhD thesis, Bowling
Green State University.
Noh, H. (2014), ‘Frontier estimation using kernel smoothing estimators with data transformation’, Journal
of the Korean Statistical Society 43, 503–512.
O’Donnell, C. J., Rao, D. S. P. & Battese, G. E. (2008), ‘Metafrontier frameworks for the study of firm-level
efficiencies and technology ratios’, Empirical Economics 34, 231–255.
O’Hagan, A. & Leonard, T. (1976), ‘Bayes estimation subject to uncertainty about parameter constraints’,
Biometrika 63(1), 201–203.
Olson, J. A., Schmidt, P. & Waldman, D. A. (1980), ‘A Monte Carlo study of estimators of stochastic frontier
production functions’, Journal of Econometrics 13, 67–82.
Ondrich, J. & Ruggiero, J. (2001), ‘Efficiency measurement in the stochastic frontier model’, European
Journal of Operational Research 129(3), 434–442.
95

Orea, L. & Kumbhakar, S. C. (2004), ‘Efficiency measurement using a latent class stochastic frontier model’,
Empirical Economics 29(1), 169–183.
Papadopoulos, A. (2015), ‘The half-normal specification for the two-tier stochastic frontier model’, Journal
of Productivity Analysis 43(2), 225–230.
Park, B. U., Simar, L. & Zelenyuk, V. (2015), ‘Categorical data in local maximum likelihood: theory and
applications to productivity analysis’, Journal of Productivity Analysis 43(1), 199–214.
Parmeter, C. F. & Kumbhakar, S. C. (2014), ‘Efficiency Analysis: A Primer on Recent Advances’, Founda-
tions and Trends in Econometrics 7(3-4), 191–385.
Parmeter, C. F. & Racine, J. S. (2012), Smooth constrained frontier analysis, in X. Chen & N. Swanson, eds,
‘Recent Advances and Future Directions in Causality, Prediction, and Specification Analysis: Essays in
Honor of Halbert L. White Jr.’, Springer-Verlag, New York, New York, chapter 18, pp. 463–489.
Parmeter, C. F., Wang, H.-J. & Kumbhakar, S. C. (2017), ‘Nonparametric estimation of the determinants
of inefficiency’, Journal of Productivity Analysis 47(3), 205–221.
Parmeter, C. F. & Zelenyuk, V. (2016), A bridge too far? the state of the art in combining the virtues of
stochastic frontier analysis and data envelopment analysis. University of Miami Working Paper 2016-10.
Paul, S. & Shankar, S. (2017), An alternative specification for technical efficiency effects in a stochastic
frontier production function. Crawford School Working Paper 1703.
Pitt, M. M. & Lee, L.-F. (1981), ‘The measurement and sources of technical inefficiency in the Indonesian
weaving industry’, Journal of Development Economics 9(1), 43–64.
Polachek, S. W. & Yoon, B. J. (1987), ‘A two-tiered earnings frontier estimation of employer and employee
information in the labor market’, The Review of Economics and Statistics 69(2), 296–302.
Polachek, S. W. & Yoon, B. J. (1996), ‘Panel estimates of a two-tiered earnings frontier’, Journal of Applied
Econometrics 11(2), 169–178.
Racine, J. S. & Li, Q. (2004), ‘Nonparametric estimation of regression functions with both categorical and
continuous data’, Journal of Econometrics 119(1), 99–130.
Reifschneider, D. & Stevenson, R. (1991), ‘Systematic departures from the frontier: A framework for the
analysis of firm inefficiency’, International Economic Review 32(1), 715–723.
Richmond, J. (1974), ‘Estimating the efficiency of production’, International Economic Review 15(2), 515–
521.
Ritter, C. & Simar, L. (1997), ‘Pitfalls of normal-gamma stochastic frontier models’, Journal of Productivity
Analysis 8(2), 167–182.
Robinson, P. M. (1988), ‘Root-n consistent semiparametric regression’, Econometrica 56, 931–954.
Ruggiero, J. (1999), ‘Efficiency estimation and error decomposition in the stochastic frontier model: A Monte
Carlo analysis’, European Journal of Operational Research 115(6), 555–563.
Schmidt, P. (1976), ‘On the statistical estimation of parametric frontier production functions’, The Review
of Economics and Statistics 58(2), 238–239.
Schmidt, P. (2011), ‘One-step and two-step estimation in SFA models’, Journal of Productivity Analysis
36(2), 201–203.
Schmidt, P. & Sickles, R. C. (1984), ‘Production frontiers and panel data’, Journal of Business & Economic
Statistics 2(2), 367–374.
Silvapulle, M. & Sen, P. (2005), Constrained Statistical Inference, Wiley, Hoboken, New Jersey.
Simar, L., Lovell, C. A. K. & van den Eeckaut, P. (1994), Stochastic frontiers incorporating exogenous
influences on efficiency. Discussion Papers No. 9403, Institut de Statistique, Universite de Louvain.
Simar, L., Van Keilegom, I. & Zelenyuk, V. (2017), ‘Nonparametric least squares methods for stochastic
frontier models’, Journal of Productivity Analysis 47(3), 189–204.
Simar, L. & Wilson, P. W. (2007), ‘Estimation and inference in two-stage, semi-parametric models of pro-
duction processes’, Journal of Econometrics 136(1), 31–64.
Simar, L. & Wilson, P. W. (2010), ‘Inferences from cross-sectional, stochastic frontier models’, Econometric
Reviews 29(1), 62–98.
Simar, L. & Wilson, P. W. (2011), ‘Two-stage DEA: Caveat emptor’, Journal of Productivity Analysis
36(2), 205–218.
96 SUBAL C. KUMBHAKAR, CHRISTOPHER F. PARMETER, AND VALENTIN ZELENYUK

Simar, L. & Wilson, P. W. (2013), ‘Estimation and inference in nonparametric frontier models: Recent
developments and perspectives’, Foundations and Trends in Econometrics 5(2), 183–337.
Simar, L. & Wilson, P. W. (2015), ‘Statistical approaches for nonparametric frontier models: A guided tour’,
International Statistical Review 83(1), 77–110.
Simar, L. & Zelenyuk, V. (2011), ‘Stochastic FDH/DEA estimators for frontier analysis’, Journal of Pro-
ductivity Analysis 36(1), 1–20.
Solow, R. (1957), ‘Technical change and the aggregate production function’, The Review of Economics and
Statistics 39(3), 312–320.
Stevenson, R. (1980), ‘Likelihood functions for generalized stochastic frontier estimation’, Journal of Econo-
metrics 13(1), 58–66.
Stiglitz, J. E. & Greenwald, B. C. (1986), ‘Externalities in economies with imperfect information and incom-
plete markets’, Quarterly Journal of Economics 101(2), 229–264.
Taube, R. (1988), Möglichkeiten der effizienzmess ung von öffentlichen verwaltungen. Duncker & Humbolt
GmbH, Berlin.
Tibshirani, R. & Hastie, T. (1987), ‘Local likelihood estimation’, Journal of the American Statistical Asso-
ciation 82, 559–568.
Timmer, C. P. (1971), ‘Using a probabilistic frontier production function to measure technical efficiency’,
The Journal of Political Economy 79(4), 776–794.
Tran, K. C. & Tsionas, E. G. (2009), ‘Estimation of nonparametric inefficiency effects stochastic frontier
models with an application to British manufacturing’, Economic Modelling 26, 904–909.
Tran, K. C. & Tsionas, E. G. (2013), ‘GMM estimation of stochastic frontier models with endogenous
regressors’, Economics Letters 118, 233–236.
Tsionas, E. G. (2007), ‘Efficiency measurement with the Weibull stochastic frontier’, Oxford Bulletin of
Economics and Statistics 69(5), 693–706.
Tsionas, E. G. (2012), ‘Maximum likelihood estimation of stochastic frontier models by the Fourier trans-
form’, Journal of Econometrics 170(2), 234–248.
Uekusa, M. & Torii, A. (1985), ‘Stochastic production functions: An application to Japanese manufacturing
industry (in Japanese)’, Keizaigaku Ronsyu (Journal of Economics) 51(1), 2–23.
Waldman, D. M. (1982), ‘A stationary point for the stochastic frontier likelihood’, Journal of Econometrics
18(1), 275–279.
Wang, H.-J. (2002), ‘Heteroscedasticity and non-monotonic efficiency effects of a stochastic frontier model’,
Journal of Productivity Analysis 18(2), 241–253.
Wang, H.-J. & Ho, C.-W. (2010), ‘Estimating fixed-effect panel stochastic frontier models by model trans-
formation’, Journal of Econometrics 157(2), 286–296.
Wang, H.-J. & Schmidt, P. (2002), ‘One-step and two-step estimation of the effects of exogenous variables
on technical efficiency levels’, Journal of Productivity Analysis 18, 129–144.
Wang, W. S., Amsler, C. & Schmidt, P. (2011), ‘Goodness of fit tests in stochastic frontier models’, Journal
of Productivity Analysis 35(1), 95–118.
Wang, W. S. & Schmidt, P. (2009), ‘On the distribution of estimated technical efficiency in stochastic frontier
models’, Journal of Econometrics 148(1), 36–45.
Wheat, P., Greene, B. & Smith, A. (2014), ‘Understanding prediction intervals for firm specific inefficiency
scores from parametric stochastic frontier models’, Journal of Productivity Analysis 42, 55–65.
White, H. (1980), ‘A heteroskedasticity-consistent covariance matrix estimator and a direct test for het-
eroskedasticity’, Econometrica 48, 817–838.
Winsten, C. B. (1957), ‘Discussion on Mr. Farrell’s paper’, Journal of the Royal Statistical Society Series A,
General 120(3), 282–284.
Wooldridge, J. M. (2010), Econometric Analysis of Cross Section and Panel Data, 2nd edn, MIT Press,
Cambridge, Massachusetts.
Figures 97

Figure 10.1. Concave and Monotonic Conditional Mean and Production


Frontier Under Homoskedastic Inefficiency. The solid line is the production
frontier while the dashed line is the conditional mean of output.
Conditional Mean is
Concave and Monotonic


0.55

2.5

0.50


● ●
● ●

2.0

0.45
● ●
●● ●

● ●


● ● ●
1.5

● ● µu

0.40
● ●
y

●●


●● ● ● ● ● ●

●●

0.35

1.0

● ● ● ● ●
● ● ●
● ●
● ●
● ●
● ● ●
● ●
0.30


●●
0.5



● ● ●


0.25



0.0

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0

x x
98 Figures

Figure 10.2. Concave but Non Monotonic Conditional Mean and Production
Frontier Under Heteroskedastic Inefficiency. The solid line is the production
frontier while the dashed line is the conditional mean of output.
Conditional Mean is
Non−Monotonic
3.0


● ●
2.5

0.4


●● ●●
● ●
● ●
2.0

● ●
● ●

● ● ● ● ●
● ●

0.3
● ●

● ●● ●
µu

y

●● ●
1.5

● ●


● ● ●
● ●
● ● ●● ● ●

● ● ● ●
0.2
1.0

● ●
● ● ●

● ●
● ●
● ● ●
● ●

0.5



0.1

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0

x x
Figures 99

Figure 10.3. Monotonic but Non Concave Conditional Mean and Production
Frontier Under Heteroskedastic Inefficiency. The solid line is the production
frontier while the dashed line is the conditional mean of output.
Conditional Mean is
Non−Concave

● ●
2.5

0.7
● ●

● ● ●
● ●
2.0



● ●
● ●

0.6

● ● ●
1.5

● ●
● ●

● ●●
● ● ●●
● ● µu
● ● ●
y

1.0

● ●
● ●● ● ●
● ●

● ● ●● ● 0.5
● ●
● ●●
0.5

● ● ●


●● ●
● ● ●●

● ●

0.0

0.4



−0.5

0.0 0.2 0.4 0.6 0.8 1.0 0.0 0.2 0.4 0.6 0.8 1.0

x x
100 Figures

Figure 10.4. Conditional quantile estimation of a univariate SFM with σu2 =


0.01. ●
● ● ●

● ●
● ●


15


● ●

True Frontier ●
τ = 0.5
τ = 0.8 ●

τ = 0.85
τ = 0.9 ● ● ●

τ = 0.95 ●
● ●

● ● ●

● ● ●
10

● ●
● ●
● ●
● ●

● ●●
● ●
● ●

● ●● ●

● ● ●
● ●● ●
y

● ● ●
● ●
● ● ● ● ●
● ● ● ● ● ● ●●
● ●
● ● ● ● ● ●●
● ●
● ● ● ●
● ●●
● ● ● ● ● ●
● ● ● ●
● ● ● ● ● ●

● ●● ● ● ● ●
● ● ● ● ● ●
● ●● ● ● ●● ●● ●
●●
5

● ● ●● ● ●
● ● ● ● ● ● ●● ● ●
● ●● ●● ●● ● ● ● ● ● ● ●●

● ●● ●
● ● ● ● ●
● ● ●

● ● ● ● ●● ● ●
● ● ● ● ● ● ● ● ● ●●
● ● ●
● ●● ● ● ●● ● ● ●● ● ●● ● ●● ●
● ● ●● ● ● ● ●● ● ● ●●●● ●
●● ● ● ● ● ● ● ● ●
● ●● ● ● ●● ● ●
●● ● ● ● ● ● ●
● ● ● ●● ● ● ● ● ● ●
●● ● ● ● ●
● ● ● ●

● ●
● ●● ●
● ● ● ●● ●●
●●●
● ● ● ● ●
● ●● ● ● ● ● ●● ● ●● ● ● ●
● ●●
●●● ●● ● ● ● ●
●● ● ● ●●
● ● ● ● ● ● ● ●●● ● ●
● ● ●
●● ● ●

●● ●● ● ● ● ●●

● ●● ●● ●
● ●● ● ●
●● ●

●● ● ●
●● ● ● ●● ● ● ●● ●
●●● ● ● ●●● ● ●
●●●● ● ● ●● ● ●
● ● ●
●●
● ●●●
● ● ●
● ●● ● ● ● ●
● ● ● ●
●● ● ●
● ●
●●●●● ●● ● ● ●● ● ● ●● ●●
● ●● ●
● ●● ● ●
● ● ● ● ● ● ● ● ● ● ●
● ●● ●●●● ●● ● ● ●● ● ● ●●●● ●
● ● ● ●● ● ●● ●
● ● ●● ● ● ● ● ●● ● ● ●●● ●
●● ● ● ●● ● ●● ● ●●●● ● ● ● ● ●●●
●●
●● ● ●●●●● ●●
●●
● ●●● ●●● ● ●●● ●● ●●●● ● ● ● ● ● ●●●● ● ●●
●● ●●
●●
● ● ●

● ● ● ● ● ●● ● ●

● ●● ● ●● ● ● ● ● ● ● ●

●● ●
●● ● ●● ● ● ● ●● ● ●● ● ● ● ● ● ●● ● ● ● ●● ●● ● ●
●●● ●● ● ● ● ● ● ● ● ●●●●●●●● ●●
● ●● ●● ●●●● ● ●●● ●
●● ●
● ●●●●
●●
●●



● ●


● ●●
● ●
● ● ●● ● ●●●●●●
● ● ●● ●●●
● ●
●● ●● ● ● ● ●
● ●● ● ●●
● ●
●● ● ●●●●● ● ●● ● ● ●● ●●●● ● ●

● ●●●
●●
●● ●
● ● ● ●● ● ● ●● ● ● ●●●● ● ● ● ●
●●● ● ●●● ●●●●●
●●●●
● ● ● ●● ●●● ● ●●● ● ●●●●●● ●● ●●●●● ●●●● ● ● ● ●● ●● ●● ●●●● ●●●● ● ●●
●●●
● ●● ● ●
● ●●● ●● ●● ● ● ●● ● ●● ● ● ●● ●
● ●● ●● ● ● ● ●●
● ●● ●● ● ● ●●


●●●
●● ● ●●
● ●●● ●●●● ● ●● ●● ●● ● ●● ● ● ●●
● ●● ● ●● ●
0

2 4 6 8 10

x
Figures 101

Figure 10.5. Conditional quantile estimation of a univariate SFM with σu2 =


0.25.
● ●


15




True Frontier
τ = 0.5 ●
τ = 0.8 ● ●

τ = 0.85

τ = 0.9 ●

τ =●0.95 ●



10



● ●

● ●
● ●
● ● ● ●

● ●
● ● ●
y


● ●

● ●
● ● ● ●
● ● ● ● ● ●
● ● ●●
● ● ● ● ● ●
● ● ●

● ●● ● ●
● ● ● ● ● ●
● ● ●● ●● ●
5

● ● ●
● ●● ●

● ●
● ●● ●
● ● ●
● ●● ● ● ●● ● ●● ● ●
●● ● ●
●●● ● ● ● ● ●● ● ● ● ● ●●
● ● ● ● ●●●
● ● ● ● ● ● ● ● ●
● ● ● ● ● ● ●●● ●
●● ● ●
● ● ● ●● ● ● ● ●●● ●
● ● ●● ● ●● ● ● ●

● ●● ● ● ●● ● ●
● ● ● ● ●
● ●
● ● ● ● ● ● ● ● ●
● ● ● ● ●● ●
● ●● ●
● ●
●● ● ● ●● ● ● ● ●
● ●● ●●●
● ●
●● ●
● ● ● ●●●
● ● ● ●
● ● ●●●● ●
● ●● ●● ● ● ● ●● ●
●● ● ● ● ●
● ● ● ● ●● ● ● ●●

● ● ●● ● ● ● ●● ● ● ●● ●
●●● ●● ●
● ●
● ● ●● ● ● ●● ●● ●● ● ● ● ● ●
● ● ● ●
●● ●● ●
●● ● ● ●● ● ● ● ● ●
●● ●●● ● ● ●● ●● ● ● ● ● ● ● ●●
● ●●
● ●● ●● ● ● ● ●● ● ● ●●
● ● ●● ● ●● ● ●
● ●
● ● ●●●●● ●
●●●●●●● ●● ● ●● ● ●●● ● ●● ● ●●●● ●

●● ● ●●●● ●● ● ●● ● ●● ● ●●●
● ● ● ●● ● ● ● ● ●

● ●● ●● ●
●● ● ●● ● ●● ●●●● ● ● ● ● ● ● ● ● ● ●
● ● ● ● ●● ●
● ● ● ●

● ●
●●● ●● ●
●● ●●●●●
●●
●● ●● ● ● ● ●●
● ●● ●● ● ● ●● ● ● ●● ●
●● ●●
●● ● ●
● ●● ● ● ●● ● ●●●●● ●●●●
● ●●

●●●
●● ●● ● ● ●
● ● ●
●●●● ●● ●


●●●
● ●
●● ●● ● ●● ●●
●●●● ●●●● ● ● ● ●●●●●
●● ●●● ● ● ●
● ●●●●
●● ●●●●●● ●●● ●●●●●
● ●
● ● ●
●●●● ●●●● ●● ● ●
● ●
● ●● ● ● ●
●● ●
● ●●●● ●
● ● ●●● ● ● ● ● ●
●● ●●
●● ● ● ● ●●●●● ● ●
● ●● ●● ● ● ● ●● ● ●● ●
●● ●
●● ● ●● ●●●● ● ● ●● ● ● ●● ●● ● ●● ● ● ● ● ●● ●
●● ● ● ●●●● ●●●● ●●●●● ●● ●● ● ● ●●
●●● ●●
● ● ●● ● ● ● ●● ● ●●● ●●● ●
●●●● ● ● ● ●● ●●●


●●● ● ●
● ●● ●●
●●●

●●●● ●● ● ●● ● ●●● ●
● ●●● ●
●●● ●
● ●

●●●

●● ●●●●●
● ● ●

●●●

● ●●

●●● ● ●● ●●● ●● ●●● ● ● ●
●● ● ●●●●



●●
● ● ● ●

●●●●● ● ●
●● ●●● ●● ● ● ●●● ●
●● ● ● ● ● ●● ● ●● ●●

0

2 4 6 8 10

x
102 Figures

Figure 10.6. Conditional quantile estimation of a univariate SFM with σu2 =


1.
● ●


15

True Frontier
τ = 0.5
τ = 0.8 ●
● ●

τ = 0.85
τ = 0.9


τ = 0.95 ●




10




●● ●
● ●

y

● ●●
● ●
● ●
● ● ●

● ● ●
● ● ● ●
● ● ●
●● ● ● ●

5

● ●
● ● ●
● ● ●● ●
● ● ●
● ● ● ●
● ● ● ●
● ● ● ●
● ●● ● ● ● ● ●
●● ● ● ● ●
●● ● ● ● ●
● ●●
● ● ● ● ● ● ● ●
●● ●● ●● ●

● ●
● ● ●● ● ● ●● ●●
● ● ●
●●
●● ●
● ● ● ● ●
● ● ●

● ● ●● ● ● ●●● ● ●● ● ● ● ●
●● ● ●
●● ●● ● ● ● ● ●● ● ● ●● ● ●
● ● ● ●● ● ●
●● ● ●● ● ● ● ● ●● ● ●●● ● ● ●● ● ●
● ● ● ● ● ● ● ● ● ●●● ● ●● ● ●●
● ●● ● ●
● ● ● ●●● ● ●
● ● ● ● ● ● ● ●
●●

●● ● ● ● ● ●
● ●●
●●
●● ●
● ●
● ●
● ● ● ● ● ● ● ● ●● ● ● ● ● ●●
●●
●●● ● ● ● ● ●
● ●●● ● ● ● ● ● ●● ●● ●
● ● ●● ● ●● ●
● ●●● ● ● ● ●● ●● ● ●●●
● ●● ●●●● ● ●● ●● ●●● ● ●●● ● ● ●●
●● ●
●●
●● ●● ● ●●●●●●
●● ●● ●● ●● ● ● ● ●● ● ● ● ●● ● ●●● ●● ●●●● ●● ●
●●●●●

●●●●

● ● ● ●
●● ● ● ●●● ●●●● ● ● ● ● ●● ● ● ●● ●● ●●
● ●●●
● ● ● ●● ●● ●● ●
● ●●●
● ●●●● ● ●
● ● ●
●● ● ●
● ●● ● ●● ● ● ● ● ● ● ● ● ●● ●
● ● ● ●● ● ●● ● ● ●● ● ●
● ● ● ●●● ●●● ● ● ●●

● ●

● ● ● ● ●●● ●●
●●●● ●●●● ● ●●●●
● ● ●
● ● ●● ●● ● ●● ● ●● ● ●● ●●● ● ●● ● ●●
●●
●●●●●
●● ● ●●● ● ●●●
●●● ●●●●●●
● ● ●●
●● ● ● ●●●
● ●●●
● ●●●●
●● ●●● ● ●●● ● ● ● ●● ●●●● ●●
● ● ●● ●●●● ●●
● ●● ● ●●● ●
●● ●
●● ●●
●●
● ●● ● ●
● ● ●●●●● ●
●●●●● ●●●●●●●●●
● ●●●●●●●●●●●


●●● ●

●●
● ●●● ●●
● ●●●
●● ●●●●●
●● ●
●●●●
●●● ●●●●

●●
● ●
●● ● ●

●●
● ●●
●● ●●●● ●● ● ● ●●●

● ●● ●●● ●
●●
●●●●● ● ●
● ●●●


● ● ● ● ● ●● ● ● ●
● ●
● ● ●
● ●● ●●
● ●
● ● ● ●● ● ●● ● ●● ● ●
●● ●
● ●● ●● ●●●
● ●● ●● ●

● ●
●●
●●●●●●●●●
● ●●
●● ●● ●
●●
● ●●●
●● ●●●● ●● ● ● ●● ●● ● ● ● ●●●●

●● ●● ●●
●● ●
●●● ● ●
● ●●●●● ● ●● ●
●●
● ●●● ● ●
● ●●●●
● ●●
0

2 4 6 8 10

x
Figures 103

Figure 10.7. Conditional quantile estimation of a univariate SFM with σu2 =


4.

15

True Frontier ●
τ = 0.5
τ = 0.8

τ = 0.85 ●

τ = 0.9
τ = 0.95

10



● ●


y

● ●

● ●
● ● ●
● ●

● ●
● ● ● ● ●
5

● ● ●
● ● ● ●
● ●
● ●
● ● ●

● ● ●
●● ● ●● ●
● ● ●
● ● ● ● ●
● ● ● ●● ●● ●
● ● ●

● ● ● ● ●
● ● ●●● ● ● ● ●
● ● ● ● ● ●●● ● ● ● ●
● ● ●
● ● ● ●
● ● ●● ● ● ● ● ● ● ● ● ● ●●
● ●●● ● ● ● ●● ●● ● ●●
● ● ● ● ● ● ● ● ● ● ● ●● ● ●●● ●●●● ● ● ●● ●
● ● ● ● ● ● ● ● ●●● ● ● ● ●● ●
● ● ● ●● ● ● ● ● ● ● ●●● ●● ● ●● ● ● ● ●●

●● ● ●
●●●● ● ● ●● ● ● ● ● ●●●

● ● ●● ●● ● ● ● ● ● ●●●● ● ●● ●


● ● ●●● ● ●● ●● ●
●● ● ● ● ● ● ●
● ● ●●●● ●● ● ●
● ●● ● ● ●● ● ●● ● ● ● ●
● ● ● ● ● ● ● ●● ●●● ●

● ●● ● ● ●● ● ● ● ●● ●● ●● ● ●

●● ● ● ●● ● ● ●● ● ●● ● ● ●● ● ● ● ●●● ●●●●●●●●● ● ● ●
● ●● ● ● ●●
●● ● ●● ● ●● ● ●
● ●● ●●● ● ● ● ● ●●
● ●●●

● ● ●● ●●● ●
● ● ● ● ●●
● ●● ● ● ● ●●
●●●●●●●●● ●

●● ●●●
●● ●
●●●●●● ● ●●
● ●●●●●●●●● ●● ●●●
●●●●
●● ●● ●● ●
●● ● ● ● ● ●●● ●● ●●●●●●● ● ●●●● ●●●
● ●● ● ●●●●● ●●● ● ●
● ● ● ●●
● ●●● ●●●● ●● ● ● ●
● ●
●● ● ●● ● ●● ●● ●●● ●
● ● ●●●●●

● ●
●●● ● ●

●●


● ●
●●●
●●●●● ●●●
●● ●●
● ● ●●● ●

●● ●
●●●● ● ● ●



● ●● ●
● ●
●●●●●●●

● ● ●●
● ●
●●●
● ●●


● ●
●● ●●●●


●●●●


●● ● ● ●● ●● ● ●● ● ●●
●● ● ●
●●


●●
● ●●●
●●●

● ● ●
●●●● ●● ●● ●●● ●●●●●● ● ●●● ●● ●● ●● ● ●● ●●●
●●●●● ● ●●
●●

● ●
●●●
●●
●●●
● ●
●●●
● ●●●●
● ● ●●
●●
● ●

●●
● ●●
●●●●●●● ●●
● ●●●●●●

● ●●●
● ●
●●●
● ●
● ●●
●●●
● ● ●●●
● ● ●●●
●● ●
●●●●
●●●
● ●●
●●● ●●● ●●

●●●● ●
●● ● ● ●●●●●●● ●
●● ●● ●●
●●
●●

● ●●
● ●●●●
●●

●●●
●●●● ● ●●● ●●●●●
●●
●●●●●
●● ●
0

● ● ●● ●● ● ● ● ●●● ● ● ●

2 4 6 8 10

x
104 Tables

Table 1. Right tail critical values for both a χ21 and a 50:50 mixture of a χ20
and a χ21 , denoted as χ̄2 .

Significance Level 0.01 0.05 0.1 0.15 0.2 0.25


2
χ1 6.634 3.841 2.706 2.072 1.642 1.323
χ̄2 5.412 2.706 1.642 1.074 0.708 0.455

You might also like