Sei sulla pagina 1di 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/251527888

Forecasting the intermittent demand for slow-moving inventories: A


modelling approach

Article  in  International Journal of Forecasting · April 2012


DOI: 10.1016/j.ijforecast.2011.03.009

CITATIONS READS

39 2,093

3 authors, including:

Keith Ord
Georgetown University
239 PUBLICATIONS   19,314 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Models for composite time series (e.g. market shares) View project

All content following this page was uploaded by Keith Ord on 02 May 2016.

The user has requested enhancement of the downloaded file.


International Journal of Forecasting 28 (2012) 485–496

Contents lists available at SciVerse ScienceDirect

International Journal of Forecasting


journal homepage: www.elsevier.com/locate/ijforecast

Forecasting the intermittent demand for slow-moving inventories: A


modelling approach
Ralph D. Snyder a,∗ , J. Keith Ord b , Adrian Beaumont a
a
Department of Econometrics and Business Statistics, Monash University, Clayton, VIC 3800, Australia
b
McDonough School of Business, Georgetown University, Washington, DC 20057, USA

article info abstract


Keywords: Organizations with large-scale inventory systems typically have a large proportion of
Croston’s method
items for which demand is intermittent and low volume. We examine various different
Exponential smoothing
approaches to demand forecasting for such products, paying particular attention to the
Hurdle shifted Poisson distribution
Intermittent demand need for inventory planning over a multi-period lead-time when the underlying process
Inventory control may be non-stationary. This emphasis leads to the consideration of prediction distributions
Prediction likelihood for processes with time-dependent parameters. A wide range of possible distributions
State space models could be considered, but we focus upon the Poisson (as a widely used benchmark), the
negative binomial (as a popular extension of the Poisson), and a hurdle shifted Poisson
(which retains Croston’s notion of a Bernoulli process for the occurrence of active demand
periods). We also develop performance measures which are related to the entire prediction
distribution, rather than focusing exclusively upon point predictions. The three models are
compared using data on the monthly demand for 1046 automobile parts, provided by a
US automobile manufacturer. We conclude that inventory planning should be based upon
dynamic models using distributions that are more flexible than the traditional Poisson
scheme.
© 2011 International Institute of Forecasters. Published by Elsevier B.V. All rights reserved.

1. Introduction units (SKUs) may be of very high value, such as, for
example, spare aircraft engines. However, even when the
Modern inventory control systems may involve thou- individual units are of low value, it is not unusual for them
sands of items, many of which show very low levels of de- to represent a large percentage of the number of SKUs, so
mand. Furthermore, such items may be requested only on that they collectively represent an important element in
an occasional basis. When events corresponding to positive the planning process. Johnston and Boylan (1996a, p. 121)
demands occur only sporadically, we refer to the demand cite an example where the average number of purchases
as intermittent. When the average size of a customer order of an item by a customer was 1.32 occasions per year, and
is large, a continuous distribution is a suitable description, ‘‘For the slower movers, the average number of purchases
but when it is small, a discrete distribution is more appro- was only 1.06 per item [per] customer’’. Similarly, in the
priate. study of car parts discussed in Section 6, out of 2509 series
In this paper, our interest focuses upon intermittent with complete records for 51 months, only 1046 had (a)
demand with low volume. On occasion, such stock keeping ten or more months with positive demands, and (b) at
least some positive demands in the first 15 and the last 15
months.
∗ Demand forecasting for high volume products can be
Corresponding author. Tel.: +61 3 9905 2366; fax: +61 3 9905 5474.
E-mail addresses: ralph.snyder@monash.edu (R.D. Snyder),
handled successfully using exponential smoothing meth-
ordk@georgetown.edu (J.K. Ord), adrian.beaumont@monash.edu ods, for which a voluminous body of literature exists; see
(A. Beaumont). for example Hyndman, Koehler, Ord, and Snyder (2008)
0169-2070/$ – see front matter © 2011 International Institute of Forecasters. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.ijforecast.2011.03.009
486 R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496

and Ord, Koehler, and Snyder (1997). When volumes are • an assumption that the time gaps and the demand in
low, the exponential smoothing framework must be based active periods are statistically independent.
upon a distribution that describes count data, rather than Since the original paper by Croston, a number of extensions
the normal distribution. Further, as was recently empha- and improvements to the method have been made, notably
sized by Syntetos, Nikolopoulos, and Boylan (2010), it is not by Johnston and Boylan (1996a) and Syntetos and Boylan
sufficient to look at point forecasts when making inven- (2005). Syntetos and Boylan (2001) first showed that
tory decisions. Instead, they recommend the use of stock the original Croston estimators were biased; they then
control metrics. We accept their viewpoint completely, but (see Syntetos & Boylan, 2005) developed a new method,
since such metrics depend upon the underlying prediction which we refer to as the bias-adjusted Croston method,
distribution, we have opted to work with such distribu- and evaluated its performance in an extensive empirical
tions directly. This choice is reinforced by the observation study. Their out-of-sample comparisons indicate that the
that prediction distributions are applicable to count prob- new method provides superior point forecasts for ‘‘faster
lems beyond inventory control. Moreover, the information intermittent’’ items; that is, those with relatively short
on costs and lead times required when using inventory cri- mean times between active demand periods.
teria was not available for the data considered in Section 6. Snyder (2002) identified some logical inconsistencies
The remainder of the paper is structured as follows. in the original Croston method and examined the use of
It begins in Section 2 with a review of the literature on a time-dependent Bernoulli process. Unlike with Croston’s
forecasting intermittent demand. The focus here is on method, distinct smoothing parameters were used for
models that allow for both non-stationary and stationary the positive demands and the time gaps. Snyder went
features. For example, the demand for spare parts may on to develop a simulation procedure which provides a
increase over time as the machines age and then decline numerical determination of the predictive distribution for
as they fail completely or are withdrawn from service. the lead-time demand. Shenstone and Hyndman (2005)
In Section 3, we summarize the different models which showed that there is no possible model which will lead
will be considered in the empirical analysis and examine to the Croston forecast function unless we allow a sample
how they might be estimated and how they might be space for active period demands that can take on either
used to simulate various prediction distributions. Since negative or positive values.
our particular focus is on the ability of a model to
furnish the entire prediction distribution, not just point 2.1. Low volume, intermittent demand
forecasts, we examine suitable performance criteria in
Section 4. Issues relating to model selection are examined There is an extensive body of literature on low count
briefly in Section 5. In Section 6 we present an empirical time series models which is potentially applicable to
study using data on the monthly demand for 1046 forecasting the demand for slow moving items. Most
automobile parts. Then, in Section 7, we examine the links expositions rely on a Poisson distribution to represent the
between forecasting and management decision making, counts but introduce serial correlation through a changing
with an illustration of the use of prediction distributions in mean (and variance). Models based on lagged values of
inventory management. Finally, various conclusions from the count variable essentially have a single source of
our research are summarized briefly in Section 8. randomness (Davis, Dunsmuir, & Wang, 1999; Heinen,
2003; Jung, Kukuk, & Liesenfeld, 2006; Shephard, 1995).
2. Review of the literature on intermittent demand By contrast, models which are based upon unobservable
components have an additional source of randomness
The classic paper on this topic is that of Croston (1972), driving the evolution of the mean (Davis, Dunsmuir,
with corrections by Rao (1973). Croston’s key insight was & Wang, 2000; Durbin & Koopman, 2001; Harvey &
that: Fernandes, 1989; West & Harrison, 1997; West, Harrison,
& Migon, 1985; Zeger, 1988). In addition, there are also
When a system is being used for stock replenishment, several multiple source of error approaches based upon
or batch size ordering, the replenishment will almost integer-valued autoregressive (INAR) models (Al-Osh &
certainly be triggered by a demand which has occurred Alzaid, 1987; McCabe & Martin, 2005; McKenzie, 1988).
in the most recent interval (Croston, 1972, p. 294). Single- and dual-source of error models for count data
The net effect of this phenomenon when forecasting the were compared by Feigin, Gould, Martin, and Snyder
demand for a product which is required only intermittently (2008), who found the dual source of error model
is that the mean demand is over-estimated and the vari- to be more flexible, which is not true for Gaussian
ance is under-estimated. Thus, an inventory decision based measurements (Hyndman et al., 2008). However, their
on the application of the usual exponential smoothing for- analysis was conducted under a stationarity assumption,
while, as was noted earlier, demand series are typically
mulae will typically produce inappropriate stock levels.
non-stationary. The results in Section 6 suggest that non-
Croston then proceeded to develop an alternative approach
stationary single source of error models are competitive
based upon:
with other approaches for count time series.
• an exponential smoothing scheme for updating the
expected time gap between successive active demand 2.2. Evaluation of the Croston method
periods;
• an exponential smoothing scheme for updating the Willemain, Smart, Shockor, and DeSautels (1994) con-
expected demand in active periods; and ducted an extensive simulation study which violated some
R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496 487

of the original assumptions (such as cross-correlations be- When we consider processes with low counts, the
tween active demand periods and the positive demand discrete nature of the distributions can lead to prediction
quantities) and found that substantial improvements were intervals whose actual levels may be higher than the
possible. When they tested the method on real data, the nominal level. Accordingly, we focus upon complete
benefits for one-step-ahead forecasts were modest. How- prediction distributions rather than intervals in this paper.
ever, as Johnston and Boylan (1996b) pointed out in a com-
ment, improvements are only to be expected when the 3. Models for intermittent demand and low volume
average time between active demand periods is appre-
ciably longer than the periodic review time. Willemain, The literature contains relatively little discussion of this
Smart, and Schwartz (2004) described a bootstrap-based case, although, interestingly, at the end of their paper
approach which allows for a Markov chain development of Johnston and Boylan (1996a) indicate that a simple Poisson
the probability of a non-zero demand, and indicated that process might suffice for slow movers. We move in a
their method produces better inventory decisions than ei- different direction in two respects: first, we will retain
ther exponential smoothing or the Croston method. How- the idea that the demand is measured at the ends of
ever, Gardner and Koehler (2005) point out that Willemain regular periods of time. Second, we wish to allow for lumpy
et al. (2004) did not use the correct lead-time distributions demand such that the measured demand may exceed one.
for either of these benchmark methods, nor did they ex-
amine extensions of the Croston method. Finally, from the
3.1. The basic models
perspective of prediction intervals, Willemain et al. pro-
vided an incorrect variance expression.
Our framework encompasses two possible ways of
Sani and Kingsman (1997) also conducted a sizeable
viewing time. The first assumes that time is continuous
simulation study which compared various methods. They
and that transactions occur sporadically. At the same
used multiple criteria, including overall cost criteria
time, demands are observed periodically, so that the
and the service level; they too found that the Croston
number of transactions in each period has a Poisson
method performed well, although a simple moving average
distribution. In cases where the transaction size always
provided the best overall performance. In an empirical
equals 1, the quantity demanded in each period also has
study, Eaves and Kingsman (2004) found little difference
the same Poisson distribution. To allow for the possibility
between exponential smoothing and the bias-adjusted
that transaction sizes are random, we also consider the
Croston method when using traditional point measures
possibility that they are governed by a logarithmic series
(mean absolute deviation, root mean squared error and
distribution. Then, the combined demand Y in a time period
mean absolute percentage error). They went on to argue
has a negative binomial distribution (Quenouille, 1949;
that a better measure would be to examine the average
Stuart & Ord, 1994, pp. 179–187). It belongs to the family
stock holdings for a given safety stock policy. Their
of compound Poisson distributions and is also known as
simulation results suggest that the bias-adjusted Croston
a randomly stopped sum distribution, which describes the
method works significantly better than the original
mechanism just outlined.
method in this context.
The second approach assumes that time is discrete and
Syntetos and Boylan (2005) provided a new method, in
is divided into equal (or approximately equal) periods of
the spirit of the Croston approach, which they found to be
time such as a month. When transactions take place in a
more accurate at issue points, although the results were
particular period, their combined size Y is assumed to fol-
inconclusive at other time points.
low a shifted Poisson distribution defined by Y = Z + 1,
Teunter and Duncan (2009) provided a comparative
where Z is Poisson-distributed. This shift to the right leaves
study of a number of methods. They also concluded that
the probability of a zero undefined, so it is assigned a finite
their modified Croston method is to be preferred, based
probability q. The shifted Poisson probabilities are then
upon a comparison of target and realized service levels.
weighted by p = 1−q in order to ensure that new probabil-
ities sum to one. We call the result a hurdle shifted Poisson
2.3. Point predictions versus prediction distributions process (HSP). It will be seen that one advantage of the HSP
is that it provides a link with the modified Croston method
One interesting aspect of the empirical work done thus outlined in Section 3.1.2.
far is the strong emphasis on point forecasts. Given that the Various other possibilities have been proposed over
main purpose behind forecasting intermittent demands is the years, including other compound Poisson forms such
to plan inventory levels, a more compelling analysis should as the stuttering Poisson distribution. However, we view
examine service levels, or, more generally, prediction the Poisson, negative binomial and shifted hurdle Poisson
distributions. Indeed, as Chatfield (1992) has pointed distributions as reasonable representations of the set
out, prediction intervals, which can be derived from of possible distributions. The three distributions are
prediction distributions if required, deserve a much greater summarized in Table 1 and are used in the empirical
prominence in forecasting applications. Furthermore, the comparisons in Section 6.
limited empirical evidence available, as cited above, is In Table 1, y designates the values that can be taken
consistent with the notion that Croston-type methods may by a discrete random variable Y . Its potential probability
provide more accurate prediction distributions, even if distributions are all defined over the domain y = 0, 1,
they offer little or no advantage for point forecasts. 2, . . . .
488 R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496

Table 1
Count distributions used in the empirical study.
Distribution Mass function Parameter restrictions Mean (µ)
λy
Poisson exp(−λ) λ>0 λ
y!
Γ (a + y)
 a  y
b 1 a
Negative binomial a > 0, b > 0
Γ (a)y! 1+b 1+b b

q y=0
Hurdle shifted Poisson p ≥ 0, q > 0, λ > 0, p + q = 1 p(λ + 1)
pλy−1 exp(−λ)/(y − 1)! y = 1, 2, . . .
Notes: When the iterative estimation procedure generated an estimated value of b that exceeded 99, the negative binomial was replaced by the Poisson
distribution. This condition was typically the result of under-dispersion, which would lead to having the Poisson as a limiting case. The negative binomial
form was chosen rather than the more orthodox version with p = b/(1 + b), for the sake of consistency with the Harvey-Fernandes version.

Table 2
Recurrence relationships for the mean.
Relationship Recurrence relationship Restrictions

Static µt = µt −1
Damped dynamic µt = (1 − φ − α)µ + φµt −1 + α yt −1 µ > 0, φ > 0, α > 0
φ+α <1
Undamped dynamic µt = δµt −1 + α yt −1 δ > 0, α > 0
δ+α =1

Table 3 on non-negative integers: the simulated series values


Conversion of latent factors to distribution parameters. stochastically converge to zero (Grunwald, Hazma, &
Distribution Formula Hyndman, 1997), where they get trapped over moderate
Poisson λt = µt to long prediction horizons, a behavior which is examined
Negative binomial at = bµt further by Akram, Hyndman, and Ord (2009). This problem
Hurdle shifted Poisson λt = µptt − 1 does not occur with the damped stationary models.
It will be noted that in the case of the hurdle shifted
For each distribution, we allow for the possibility that Poisson distribution, we allow the probability of a non-zero
the mean of a demand distribution may change randomly demand in a period to change over time. Designated by pt ,
it may be viewed as a local or short-run probability. The
over time to reflect the effects of possible structural
damped case is governed by the recurrence relationship
change. Although the random variable is discrete, we
assume that the mean is continuous. Three possibilities pt = (1 − φ − α)p + φ pt −1 + α xt −1 , (1)
are summarized in Table 2, corresponding to a static or where xt = 0 if there is no demand and xt = 1 if there is
constant mean model, a damped dynamic mean (which demand in period t. It involves a long-run probability
may be thought of as a stationary autoregressive model designated by p. The parameters φ and α are the
for the mean), and an undamped dynamic model (which same as those used in the corresponding recurrence
corresponds to an integrated moving average model for relationship for the mean demand, consistent with the
the mean), respectively. The dynamic cases involve a local parameterization originally used by Croston. The effect is
or short-run mean µt . The damped case, being stationary, to contain the growth in the number of parameters and to
also includes a long-run mean µ. Additional lags are µt
force pt to be less than µt for all t, so that λt = −1
conceivable but are likely to be of dubious value relative pt
to the gains achieved by allowing the mean to evolve is never negative. The long-run parameters must satisfy
over time. The undamped dynamic relationship has no the condition p < µ for a similar reason. Both the seed
long-run mean because the process is non-stationary; the probability p1 and the long-run probability p are restricted
associated updating relationship corresponds to that for to the unit interval.
simple exponential smoothing. The undamped analogue of Eq. (1) with α + δ = 1 is
The distribution parameters λ and a are determined p t = δ p t −1 + α x t −1 . (2)
from the mean using the conversion formulae in Table 3. This corresponds to a simple exponential smoothing
There are two versions of the dynamic negative recurrence relationship for the probability. Again, α and δ
binomial model: the so-called unrestricted case, which are the same as the parameters used in the corresponding
uses the full parameterization as presented in these tables, recurrence relationship for the mean.
and the restricted case, where the restriction α = 1/(1 + Several other models have been considered in the
b) is applied to reduce the number of parameters. This literature on count time series, as noted earlier, and
restriction is included so that we can explore its links with those which are included in the simulation study are now
the Harvey-Fernandes method (outlined in Section 3.1.1). summarized briefly.
There is one disadvantage of using undamped models.
The simulation of prediction distributions from such 3.1.1. The Harvey-Fernandes method
models is hampered by a general problem that applies Harvey and Fernandes (1989) describe a method based
to all non-stationary count models which are defined on a local level state space model with Poisson measure-
R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496 489

ments. Their method does not allow for intermittent de- Moreover, it was envisaged that the parameters of these
mands but is a Poisson analogue of the Kalman filter based distributions would change over time, with their values
upon a negative binomial distribution, defined as a mix- being derived from the mean time gaps and mean non-
ture of Poisson distributions with a gamma mixture dis- negative demands obtained from the application of simple
tribution. The negative binomial distribution has a time exponential smoothing.
dependent mean given by the finite exponentially weighted The main advantage of the stochastic assumptions is
average that it expands Croston’s method to enable it to produce
entire prediction distributions by simulation, rather than
δ yt −1 + δ 2 yt −2 + δ 3 yt −3 · · · + δ t −1 y1 point predictions alone. It also enables the derivation and
µt = , (3)
δ + δ 2 + · · · + δ t −1 use of maximum likelihood estimates of the parameters δ
and α .
where δ is a parameter called the discount factor which
satisfies the condition 0 ≤ δ ≤ 1. The numerator and
3.2. Estimation
denominator of this expression are designated by at and bt
respectively. Used as the parameters a and b of the negative
All of the unknown model parameters were es-
binomial distribution (see Table 1) in typical period t, they
timated using the maximum likelihood method. The
are calculated recursively using the expressions
likelihood function is based on the joint distribution
at +1 = δ(at + yt ) and bt +1 = δ(bt + 1), p(y1 , . . . , yn |µ1 , θ ), where θ represents all unknown pa-
rameters other than the first mean µ1 . Using induc-
where a1 = b1 = 0. tion, in conjunction with the conditional probability law
As t increases in Eq. (3), bt converges to a constant value Pr{A, B} = Pr{B|A} Pr{A}, it can be established, for every
b = δ/(1 − δ), and the mean µt = bt /at is then governed model considered, that
by the simple exponential smoothing update relationship
n
µt +1 = δµt + α yt . The negative binomial probability 
p(y1 , . . . , yn |In−1 ) = p(yt |It −1 ), (6)
parameter is p = b/(1 + b) = δ , and consequently q = 1 −
p = α . The smoothing and discount parameters used for
t =1

the update of the mean also become an integral part of the where It −1 = {µt , θ}, t = 1, 2, . . . , n. The univariate dis-
negative binomial distribution formula. This corresponds tributions in this decomposition are a succession of one-
to what we earlier called the restricted undamped step-ahead prediction distributions. In the case of the static
negative binomial model. Given that this is effectively the models, the maximum likelihood estimate of the common
asymptotic form of the Harvey-Fernandes method, the two mean is just a simple average. In the other cases, the appro-
approaches should give similar predictions. priate dynamic relationship is applied to obtain the means
of the univariate distributions.
The same basic approach is used for the Harvey-
3.1.2. The modified Croston model
Fernandes (HF) model. In this case, the initial mean µ1 is
The original Croston method examined a series of the
not needed and the likelihood is calculated from the period
time gaps between periods with demand occurrences
with the first demand. Successive means are calculated
and a series of the non-zero demand quantities. Simple
using Eq. (3), and the term corresponding to t = 1 in Eq. (6)
exponential smoothing is then applied to the two derived is dropped. In the case of the modified Croston method,
series to obtain estimates of their means, and a point estimates of the parameters were also obtained using the
prediction is established from the two results. The same prediction decomposition of the likelihood. The details are
parameters δ and α are used in both of the simple provided by Hyndman et al. (2008, pp. 281–283).
exponential smoothing recurrence relationships.
Building on earlier ideas from the work of Snyder (2002)
3.3. Prediction distributions
and Shenstone and Hyndman (2005), the Croston method
was modified by Hyndman et al. (2008, pp. 281–283) to
A simulation approach is used to obtain all of the pre-
incorporate probabilistic assumptions. It was envisaged
diction distributions, although analytical methods could be
that:
developed for static models — see Ord, Snyder, and Beau-
1. The time gaps between active demand periods will be mont (2010). Given that the models involve first-order
governed by a shifted geometric distribution recurrence relationships, the joint prediction distribution
can be decomposed into a product of univariate one-step-
Pr{τ = t } = qt −1 p t = 1, 2, . . . , (4) ahead prediction distributions, as follows:
where p is the probability of a positive demand in a n +h

given period and q = 1 − p. p(yn+1 , . . . , yn+h |In ) = p(yt |It −1 ). (7)
2. The positive demands Y + will be governed by a shifted t =n+1
Poisson distribution
In the simulation approach, successive future series
(λ)y−1 values are then generated from each future one-step-
+
Pr{Y = y|µ } = +
exp(−λ),
(y − 1)! ahead distribution. This process is repeated 100,000 times,
to give a sample from the joint distribution. Marginal and
y = 1, 2, . . . , (5)
lead time relative frequency distributions are then used as
where µ +
= λ+ 1 is the mean of the positive demands. approximations for the prediction distributions.
490 R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496

4. Prediction performance measures 4.2. Distribution based scores

Many different measures can be used to evaluate Although the MASE and similar measures are useful
predictive performance, but we focused primarily on in determining the performances of point forecasting
three: the mean absolute scaled error, the prediction methods, they do not provide any information on other
likelihood score and the discrete ranked probability score. characteristics of the predictive distributions. In the
Each of these will be defined in the following sub-sections. following sections we describe two criteria that can be
For the moment, we consider these measures from a used to measure the forecasting performance relative to
general perspective. the predictive distribution. Typically, such measures could
Let M denote a measure of the prediction performance be used with fairly small numbers of hold-out observations
that can be calculated for all models under consideration. for a single series, but would be averaged over a number of
If this measure is defined so that an increase means an series to determine the overall performance for a group of
improved prediction, we write S = 1; otherwise, when series, as in Section 6.
a decrease signifies an improvement, we let S = −1.
It is convenient to benchmark all models against the 4.2.1. Prediction likelihood score (PLS)
static Poisson model. We use Mp and Mi to represent the The joint prediction distribution p(yn+1 , . . . , yn+h |In )
measure for a static Poisson distribution and for another summarizes all of the characteristics of a future series,
model i, respectively. Moreover, it makes sense to use a including the central tendency, variability, autocorrela-
scale-free summary statistic to facilitate comparisons. We tion, skewness and kurtosis. Since we are interested in
therefore recommend the use of statistics of the form prediction distributions rather than point forecasts, this
joint distribution is a natural criterion to use. Here,
Rip = 100S (log Mi − log Mp ). (8)
In consists of all quantities that inform the calculation
Rip may be interpreted as the percentage change in of these probabilities, including the estimation sample
the measure for model i relative to the static Poisson y1 , . . . , yn , the parameters, and the states of the pro-
model. Rip > 0 indicates that model i is a better predictor cess at the end of period n. Assuming that we withhold
of yn+1 , . . . , yn+h than the static Poisson model. the series values yn+1 , . . . , yn+h for evaluation pur-
Once the statistic in Eq. (8) is in place, it is straightfor- poses, p(yn+1 , . . . , yn+h |In ) is the likelihood that these
ward to compare any two models. R12 = R1p − R2p mea- values come from the model under consideration. We call
sures the percentage difference between any two models this the prediction likelihood score (PLS), although its log-
1 and 2. Hence, R12 > 0, or equivalently R1p > R2p , indi- arithm is more commonly known as the logarithmic score
cates that model 1 is a better predictor than model 2. (Gneiting & Raftery, 2007). Note that Czado, Gneiting, and
Held (2009) use this measure in a study of cross-sectional
4.1. Mean absolute scaled error Poisson and negative binomial regression models, but with
constant coefficients.
The PLS could be evaluated in several ways. We
One performance measure in common use is the mean
consider the joint distribution of {yn+1 , . . . , yn+h } given
absolute percentage error (MAPE), defined as
the information up to and including period n, namely In .
Applying the same logic as in the derivation of Eq. (7), it
100   ŷn (j) − yn+j 
h  
MAPE =  , (9) can be established that
h j =1
 yn+j 
n +h

where ŷn (j) designates the prediction of the series value p(yn+1 , . . . , yn+h |In ) = p(yt |It∗−1 ), (11)
yn+j made at origin n, and h is the prediction horizon. t =n+1

However, it fails for low count data whenever the value where It∗ = {In , yn+1 , . . . , yt }; the change in notation
of zero is encountered in the series. We therefore used serves to indicate that the parameters are estimated using
the mean absolute scaled error (Hyndman & Koehler, 2006) only the first n observations, whereas the means are
instead: updated each time. Each univariate distribution describes
h  n the uncertainty in the typical ‘future’ period t, as seen from
1 1 
the beginning of this period, using the ‘past’ information
MASE = |yn+j − ŷn (j)| |yt − yt −1 |. (10)
h j =1 n − 1 t =2 contained in It∗−1 . Thus, the joint prediction mass function
can be found from the product of h one-step-ahead
In Section 6, the empirical study is based upon h = 6; univariate prediction distributions. This is the forecasting
that is, lead times of 1–6 are employed in the calculations. analogue of the prediction error decomposition of the
Other measures such as the geometric mean absolute error likelihood function used in estimation; see Hyndman
(GMAE) or geometric root mean square error (GRMSE) et al. (2008). The means of these one-step prediction
could be employed, but would be expected to give similar distributions are calculated using the various forms
results. In all cases, extreme values may result if the of exponential smoothing implied by the damped and
number of active demand periods in the estimation sample undamped transition relationships given in Table 2. Since
is very small (or even zero), which is why we required interest may focus upon forecasts for the demand over the
a minimum number of active demand periods in the lead-time as well as one-step-ahead, we also examine the
empirical work reported in Section 6. PLS for the sum over the lead-time Sn (h) = yn+1 +· · ·+yn+h
R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496 491

with prediction distribution p(Sn (h)|In ). In Section 6, the 5. Model selection


PLS is averaged over all of the series we consider in order
to provide an overall assessment of each model, rather than There are two principal approaches to model selection.
a selection criterion for individual series. The first uses an information criterion such as the AIC or
The following simple example illustrates why it is BIC (see, for example, Hyndman et al., 2008, pp. 105–108),
important to use a measure such as the PLS rather than one and relies upon the fit of the data to the estimation
which focuses exclusively on point forecasts. sample, with suitable penalties for extra parameters.
The second method, known as prediction validation,
Example 1: Model choice using PLS
uses an estimation sample to specify the parameter
Consider two competing static Gaussian forecasting values, and then selects a procedure based upon the
models, indexed as 1 and 2, with a common mean µ and out-of-sample forecasting performances of the competing
variances, from the estimation sample, estimated as V1 < models. Despite the popularity of prediction validation
V2 . (For example, the two models might correspond to (e.g. Makridakis & Hibon, 2000), Billah, King, Snyder,
estimation with and without the removal of outliers.) From and Koehler (2006) found it to be generally inferior to
Eq. (11), the comparative form of the PLS reduces to other methods for point forecasting, particularly to those
  based upon information criteria. This conclusion is not
V V
R12 = 100h log V2 − log V1 + − , (12) affected by the particular out-of-sample point forecasting
V2 V1 criterion selected, such as those described in the previous
section. We ran several small simulation experiments for
where V denotes the one-step-ahead forecast mean
the distributions currently under consideration, which
squared error evaluated over the hold-out sample for
confirmed the conclusions of Billah et al. (2006). This
periods n + 1, . . . , n + h.
conclusion is especially true in the present case, when the
If an in-sample criterion, based upon forecast variances,
hold-out sample for a single series is based upon only six
were used to select a model, it is clear that model 1
observations.
would be selected. However, if we are interested in the
Although the prediction validation methods are not
validity of the prediction distribution (or, more specifically, recommended for model selection for individual series,
in making a safety stock decision), we need to ensure they are useful for assessing the overall performance across
that the model selected represents the uncertainty in the multiple series when we use criteria such as PLS and DRPS
forecasts properly. In this simple case, the two procedures to evaluate the prediction distributions. Such comparisons
would give rise to the same value of V . It can be shown are common in forecasting competitions, and are useful
that expression (12) is positive at V = V1 and negative when a decision as to a general approach to forecasting
at V = V2 , so that the choice of model depends upon a group of series such as a set of SKUs must be made.
the prediction distribution. Thus, expression (12) indicates Accordingly, the comparisons in the next section are made
that we should only choose model 1 if V is sufficiently close using the criteria discussed in Section 4 for comparing the
to V1 ; that is, if R12 > 0. In other words, the method with overall model performances.
the variance closest to that of the mean squared prediction
error is chosen. 6. An empirical study of auto parts demand

4.2.2. Discrete ranked probability score (DRPS) The study used data on slow-moving parts for a US
Another possible measure is the ranked probability automobile company; these data were also discussed
score (Epstein, 1969; Murphy, 1971). It uses the L2 -norm by Hyndman et al. (2008, pp. 283–286). The data set
to measure the distance between two distributions: consists of 2674 monthly series, of which 2509 had

complete records. The data cover a period of 51 months;
45 observations were used for estimation and 6 were

L2 (x, F ) = (F̂ (y) − F (y))2 ,
withheld for comparing the forecasting performances one
y=0
to six steps ahead. Considering only those series with at
where F̂ (y) is the sample-based approximation of the least two active periods, the average time lapse or gap
distribution function F . When F is discrete and there is only between positive demands is 4.8 months. The average
a single observation x, positive demand is 2.1, with an average variance-to-mean
 ratio of 2.3, meaning that most of the series are over-
0 if y ≤ x dispersed. As was noted earlier, only 1046 of these series
F̂ (y) =
1 if y > x. had (a) ten or more months with positive demands, and (b)
at least some positive demands in the first 15 and the last
In our calculations, the infinite sum was truncated at 15 months; our forecasting study was restricted to these
y = 100; the numerical error caused by this truncation 1046 series, for which the average time lapse was 2.5, with
is negligible. an average variance-to-mean ratio of 1.9. The purpose of
In Section 6, we calculate the DRPS for each one-step- these restrictions was to ensure that each part had at least
ahead forecast relating to the sample withheld, and then some inventory activity during the estimation periods.
average over the h(= 6) values; the procedure follows the We examined the performance of the models defined
same logic as for the PLS. The DRPS for the total lead-time in Tables 1–3 for one-step, multiple-step and lead time
demand is also considered. demand predictions. For the purpose of comparison, we
492 R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496

Table 4
Comparison of the forecasting performances of different methods for 1046 US automobile parts series (best cases in bold).
Predictions One-step Multiple-step Lead time demand
Criterion PLS DRPS MASEb DRPS MASE PLSa DRPS MASE

Static models
Poisson 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0
Hurdle shifted Poisson 12.0 9.5 0.0 9.5 0.0 6.6 1.7 0.0
Negative binomial 14.5 13.7 0.0 13.7 0.0 9.2 11.1 0.0
Damped dynamic models
Poisson 10.9 16.7 15.4 15.0 13.0 10.8 33.0 24.4
Hurdle shifted Poisson 16.9 21.8 12.8 20.1 10.9 13.4 34.1 22.2
Negative binomial – unrestricted 20.5 25.7 15.9 24.3 13.5 15.3 41.6 25.3
Negative binomial – restricted 18.3 21.4 11.1 10.7 –9.2 14.4 22.0 –13.9
Undamped dynamic models
Poisson 10.2 18.4 19.4 18.2 18.8 10.9 38.2 32.4
Hurdle shifted Poisson 17.2 22.7 15.8 22.2 15.1 14.1 37.1 26.8
Negative binomial – unrestricted 20.1 26.9 18.9 26.3 18.2 15.2 44.2 31.7
Negative binomial – restricted 15.1 23.2 23.3 22.0 21.6 13.4 41.8 34.5
Others
Croston 15.8 18.5 8.9 18.5 8.8 12.2 28.2 17.4
Harvey and Fernandes 16.0 24.9 25.7 22.9 23.0 14.3 43.9 35.7
Zeros − inf 10.0 68.4 10.0 68.4 − inf –2.8 26.8
Info criteria based on the best choice of 9 models
AIC 18.6 23.9 15.0 23.3 14.3 13.5 37.0 25.6
BIC 16.4 20.7 11.7 20.8 11.7 11.7 30.0 20.2
a
The PLS values are averaged using a 2% trimmed mean in order to avoid a small number of extreme values.
b
The MASEs for the static models are the same, as the best predictor for each of them is the mean.

also included the Harvey-Fernandes and modified Croston and DRPS results suggest that the traditional static
models described in Sections 3.1.1 and 3.1.2, as well as Poisson distribution can be too restrictive for intermittent
the simplistic all-zeros forecast which ignores the data inventories. They confirm that better predictions may
and simply forecasts a zero demand for all periods. The be obtained from distributions which allow for over-
required prediction distributions were obtained using the dispersion, with the negative binomial distribution being
simulation approach outlined in Section 3.3. the best option. This outcome was to be expected, because
The measures PLS, DRPS and MASE were calculated the negative binomial distribution has been used widely
using withheld data from periods 46 to 51. These measures in inventory control for slow moving items, presumably
were computed for the 1046 series and benchmarked because it has been found to work well in practice.
against the static Poisson case using Eq. (8), but the Interestingly, the MASE, and similar measures based on
computational process was different for each. In the case point predictions, fail to reflect this important conclusion.
of the PLS, Eq. (8) was applied to each series and then Instead, they imply that the zero prediction is best for both
averaged. Given the form of the log-likelihood ratio, this is one-step and multi-step predictions, which would result in
equivalent to averaging the log-likelihood across all series zero stocks for all items!
and then taking the ratios for different models. We would The main aim of this study was to assess methods that
have liked to have applied a series-by-series approach to allow for serial correlation through the use of dynamic
the DRPS and MASE, but the single series measures could specifications. The performance measures all indicate that
take on zero values. Therefore, the DRPS and MASE were there is a significant improvement, and the PLS and
averaged across series before Eq. (8) was applied to the DRPS are reasonably consistent in their rankings. Each
resulting averages. measure indicates that the unrestricted negative binomial
A summary of the results, in terms of percentage distribution is best, although there is no clear indication of
improvements, is reported in Table 4. In all cases, larger whether damped or undamped dynamics should be used.
values are better. A summary using medians instead of The difference between the two forms of the dynamics
averages was also produced, but they were found to lead is small according to both criteria. Undamped dynamics
to similar conclusions and so are not reported here. The gets our vote here, because it has strong links with the
summary for the PLS of lead-time demand was based on widely used simple exponential smoothing forecasting
a trimmed mean, in order to avoid distortions from a few method, and because it has been found that most business
extreme cases. The zeros prediction method has Pr(Y = and economic time series are not stationary. Furthermore,
0) = 1 and Pr(Y ̸= 0) = 0, so its PLS results are reported fewer parameters need to be estimated, which is important
as minus infinity. when the series are short. However, as is indicated in
A number of interesting observations can be made. Section 3.1, this choice has one inconvenient aspect:
First, the results in Table 4 are reasonably consistent simulated series values eventually converge to a fixed
across the different types of forecasts. Second, the PLS point of zero.
R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496 493

Table 5 Table 6
Percentage breakdown of AIC model selections. Percentage breakdown of BIC model selections.
Distribution Dynamics Distribution Dynamics
Static Damped Undamped Total Static Damped Undamped Total

Poisson 15 18 1 34 Poisson 31 10 5 46
Negative binomial 21 11 4 36 Negative binomial 24 2 5 31
Hurdle Poisson 24 5 0 30 Hurdle Poisson 22 0 0 23
Total 60 34 6 100 Total 77 13 10 100

more appropriate. Moreover, no particular distribution


One interesting finding is that the Harvey-Fernandes
dominated. It is not clear how much emphasis should
method out-performed the widely-used Croston method
be placed on these particular outcomes when the multi-
of forecasting, or at least our adaptation of it, which allows
model approach did not work as well as the encompassing
for maximum likelihood estimation of its parameters. It
approach. Nevertheless, at first sight, the prominence of
was shown in Section 3.1.1 that the Harvey-Fernandes
static and Poisson selections may be surprising. For those
method has a limiting form corresponding to our restricted
series, it is likely that all of the models (except ‘zeros’)
version of the negative binomial model with undamped
would perform well, and the AIC is simply guiding the
dynamics. Not surprisingly, the two models had similar
modeler to the simpler schemes. In the remaining cases,
forecasting performances, with the Harvey-Fernandes
the static and Poisson versions are inadequate and more
method having a slight edge. However, the unrestricted
complex models are needed. With a different mix of series,
version of the negative binomial model was markedly
the AIC may well outperform the encompassing approach,
better than either of these approaches. What clearly
but it needs to be kept in mind that non-stationary models
emerges from this study is the fact that, based on the
are better able to adapt to structural changes in a series,
PLS and DRPS criteria, our undamped negative binomial
and thus may be preferable as an ‘insurance policy’.
model significantly out-performs both the adapted Croston
There may be concern that our initial culling of the
method and the Harvey-Fernandes method.
series has had a distorting effect on the results. If all of
The assumptions underlying the hurdle shifted Poisson
the series had been included, the results for the MASE and
distribution, as outlined in Section 3.1, appear to make it
the DRPS would have shifted in the direction of the zeros
a strong candidate for intermittent demand forecasting.
However, the results from Table 4 indicate that its method, with the latter still having the value minus infinity
performance lags behind that of the damped negative for the PLS. In practice, items with such low levels of
binomial model, according to all of the measures. In demand would often be covered, not by carrying stock, but
earlier numerical studies using the same data set, a similar simply by placing special replenishment orders as needed.
outcome (not reported here) was obtained for the zero-
inflated Poisson distribution. 7. Use of simulated demands in inventory control
In our framework, the undamped hurdle shifted Poisson
model is the closest to the modified Croston model. Instead Once forecasts of the demand have been obtained, they
of smoothing the time gaps, we smooth the demand can be fed into the inventory control decision process.
occurrence indicator variable using Eq. (2). Interestingly, In the Gaussian case (Harrison, 1967; Snyder, Koehler,
the undamped hurdle shifted Poisson model does better & Ord, 1999), this can be done analytically, but in non-
than the adapted Croston method. Gaussian cases it is often necessary to resort to simulation.
Multi-model approaches using information criteria for We will provide an example in this section, but a more
model selection were also explored in the study. Both comprehensive exposition is given by Hyndman et al.
the methods designated as Others and the restricted (2008, Chapter 18).
versions of our negative binomial models in Table 4 Our focus is on a part, the demand for which is governed
were excluded from the set of possible models for this by a Poisson distribution, the mean of which changes
part of the project. Both the Akaike (AIC) and Bayesian according to the undamped dynamic equation underlying
(BIC) information criteria were considered. As might be simple exponential smoothing, as defined in Tables 1 and
expected, the AIC approach yielded the better forecasts on 2. We assume that the maximum likelihood estimate of the
average. Nevertheless, the unrestricted damped negative smoothing parameter α has been found using 55 months
binomial case employed as an encompassing model out- of data and is 0.1. It is now the beginning of month 56;
performed the multi-model approaches. a simple exponential smoothing forecasting routine has
The percentage breakdowns of the models selected yielded a point prediction for this month of 0.75, and
by the AIC and the BIC are shown in Tables 5 and 6. a replenishment order must be placed with a supplier.
Static models proved to be quite adequate for about 60% There is a delivery lead time of 2 months, meaning that
of the series according to the AIC and 77% according a replenishment order placed now is not delivered until
to the more stringent BIC. We need to keep in mind the beginning of month 58. The size of the replenishment
the fact that the asymptotic justification for the AIC is order is found by comparing the current total supply (stock
based upon the forecasting performance, whereas the minus backlog plus outstanding replenishment orders)
BIC is a consistent criterion for selecting the true model. with a pre-determined order-up-to level (OUL). The size of
Given the forecasting focus of this study, the AIC seems the OUL determines the size of the replenishment order,
494 R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496

Table 7 month 58. The ratio of the ensemble average of sales to the
Simulation of future demands from the undamped dynamic Poisson
ensemble average demand in month 58 can be viewed as
model.
an estimate of the fill-rate. This is unlikely to exactly equal
Period Point prediction (mean) Simulated demand
the target fill-rate with the initial trial value for the OUL,
56 0.75 0 and therefore, keeping the demand scenarios unchanged,
57 0.675 2
the OUL is adjusted by a solver until the estimated fill
58 0.8075 1
rate equals the target value of 90%. The deviation of OUL
from its initial value of the mean extended lead time
which in turn determines the level of service provided to demand represents the safety stock. For practical reasons,
customers in month 58. The problem is to find a value for any resulting non-integer values of the OUL are rounded to
the OUL which ensures that the expected fill rate in month the next highest integer, in which case the target fill rate is
58 is at least 90%. Demands occurring during shortages are usually exceeded.
backlogged. Regardless of which dynamic demand model is used,
The analysis begins by assigning a trial value to the a simulation approach is needed because the analytical
OUL and determining the consequent expected fill rate in form of future distributions of demands, as seen from the
month 58. We then adjust the OUL until the expected fill forecast origin, are unknown, with the exception of the
rate equals the target value of 90%. Initially, the analysis distribution for the first future period. Because of the as-
ignores the integer property of the stock and treats the OUL sociated computational loads, an approach like this would
as a continuous quantity. This search procedure is typically have been completely impractical in the era when most
automated by the use of a solver such as Goal-Seeker in of the basic ideas underpinning inventory control theory
Microsoft Excel. were originally developed, ideas which still form the main-
Since the predicted demand is 0.75 for each future stay of most texts on the subject. Analytical approaches
month, as seen at the beginning of month 56, and the were once necessary for tractable computations, but now
extended lead time of interest is 2+1 = 3 months, we set the the raw computational power provided by modern com-
initial trial value of the OUL equal to the mean extended puters enables us to find ordering parameters such as the
lead time demand (three times the monthly demand, or OUL using approaches like the one described here in a frac-
2.25). There is no safety stock for this initial situation. tion of a second. Approaches like this are therefore now
We can then simulate the future demands for months feasible even for large numbers of parts.
56, 57 and 58, as shown in Table 7. The first value of 0
was simulated from a Poisson distribution with a mean of
0.75. The mean was then revised in the light of this new 8. Conclusions
simulated demand using simple exponential smoothing
(with α = 0.1), to give a new mean of 0.675, with this In this paper we have introduced some new models for
change being a reflection of presumed permanent changes forecasting intermittent demand time series based on a
in the market for the inventory. The second value of variety of count probability distributions, coupled with a
2 was simulated from a Poisson distribution with a variety of dynamic specifications to account for potential
mean of 0.675. This procedure was then run again (new serial correlation. These models were then compared with
mean = 0.8075), resulting in a simulated demand of 1 for established forecasting procedures using a database of
month 58. car parts demands. Particular emphasis was placed on
Given these three simulated future demands, it is then prediction distributions from these models rather than
possible to find the corresponding sales in month 58, as point forecasts, because the latter ignore features such as
shown in Table 8. The OUL represents the total stock which variability and skewness, which can be important for safety
is available to meet demand over the extended lead time of stock determination.
3 months. After the OUL of 2.25 is used to meet demands The empirical results suggest that although many series
of 0 and 2 in months 56 and 57, respectively, month 58 may be able to be modeled adequately using traditional
begins with a stock of 0.25. This is not sufficient to meet static schemes, substantial gains may be achievable by
the simulated demand of 1 in month 58. Simulated sales using dynamic versions for many of the others. A similar
immediately from stock can only be 0.25. The unsatisfied argument favors the use of richer models than the Poisson.
demand must be deferred until the next delivery. Thus, an effective forecasting framework for SKUs that
This experiment can be repeated to create, say, 100 have low volume intermittent demands must look beyond
hypothetical future ensembles of the demand and sales in the traditional static Poisson format.

Table 8
Sales calculation from simulated demands.
Period Variable Value

56 Order-up-to level 2.25 Trial value specified by user (OUL)


56 Demand 0 Simulated demand (D56)
57 Demand 2 Simulated demand (D57)
58 Open stock 0.25 max (OUL–D56–D57, 0)
58 Demand 1 Simulated demand (D58)
58 Sales 0.25 min (Open stock, D58)
R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496 495

Our study has indicated that simple exponential Feigin, P. D., Gould, P., Martin, G. M., & Snyder, R. D. (2008). Feasible
smoothing can work well in conjunction with an unre- parameter regions for alternative discrete state space models.
Statistics and Probability Letters, 78, 2963–2970.
stricted negative binomial distribution. It has also indi- Gardner, E. S., & Koehler, A. B. (2005). Comment on a patented bootstrap-
cated that there is little advantage in using a multi-model ping method for forecasting intermittent demand. International Jour-
approach with information criteria for model selection. nal of Forecasting, 21, 617–618.
The usual caveat that such results are potentially data de- Gneiting, T., & Raftery, A. E. (2007). Strictly proper scoring rules, prediction
and estimation. Journal of the American Statistical Association, 102,
pendent must be made. Nevertheless, we expect that sim- 359–378.
ilar results would emerge for other datasets. This being Grunwald, G. K., Hazma, K., & Hyndman, R. J. (1997). Some properties and
the case, it appears that the Croston method, even in an generalizations of Bayesian time series models. Journal of the Royal
Statistical Society, Series B, 59, 615–626.
adapted form, should possibly be replaced by exponential Harrison, P. J. (1967). Exponential smoothing and short-term sales
smoothing, coupled with a negative binomial distribution. forecasting. Management Science, 13, 821–842.
There are a number of series, most of which were Harvey, A. C., & Fernandes, C. (1989). Time series models for count or
excluded from the sample of 1046 SKUs used here, for qualitative observations. Journal of Business and Economic Statistics, 7,
407–422.
which the demand is very low, perhaps in the order of one Heinen, A. (2003). Modelling time series count data: an autoregressive
or two units per year. In such cases, a static model might be conditional Poisson model. Working paper. University of California. San
preferable, although from the stock control perspective the Diego.
Hyndman, R. J., & Koehler, A. B. (2006). Another look at measures of
decision will often lie between holding one unit of stock
forecast accuracy. International Journal of Forecasting, 22, 679–688.
and holding zero stock and submitting replenishment Hyndman, R. J., Koehler, A. B., Ord, J. K., & Snyder, R. D. (2008). Forecasting
orders as needed. with exponential smoothing: the state space approach. Berlin and
There still remains the important issue of what to do Heidelberg: Springer-Verlag.
Johnston, F. R., & Boylan, J. E. (1996a). Forecasting for items with
with new parts which have no or only limited demand data. intermittent demand. Journal of the Operational Research Society, 47,
In such cases, maximum likelihood methods applied to 113–121.
single series are going to be ineffective. This question is the Johnston, F. R., & Boylan, J. E. (1996b). Forecasting intermittent demand: a
comparative evaluation of Croston’s method. Comment. International
subject of an on-going investigation into the forecasting of
Journal of Forecasting, 12, 297–298.
demand for slow moving inventories, which explores the Jung, R. C., Kukuk, M., & Liesenfeld, R. (2006). Time series of count data:
possibility of extending the maximum likelihood principle modeling, estimation and diagnostics. Computational Statistics and
to multiple time series in a quest to overcome the paucity Data Analysis, 51, 2350–2364.
Makridakis, S., & Hibon, M. (2000). The M3-competition: results,
of data. conclusions and implications. International Journal of Forecasting, 16,
451–476.
McCabe, B. P. M., & Martin, G. M. (2005). Bayesian predictions of low count
Acknowledgement time series. International Journal of Forecasting, 21, 315–330.
McKenzie, E. (1988). Some ARMA models for dependent sequences of
Poisson counts. Advances in Applied Probability, 20, 822–835.
This research was supported by Australian Research Murphy, A. H. (1971). A note on the ranked probability score. Journal of
Council Discovery Grant DPO877424, and by the Sebes Applied Meteorology, 10, 155–156.
Fellowship funded by Mr Alvin J Delaire. Ord, J. K., Koehler, A. B., & Snyder, R. D. (1997). Estimation and prediction of
a class of dynamic nonlinear statistical models. Journal of the American
Statistical Association, 92, 1621–1629.
Ord, J. K., Snyder, R. D., & Beaumont, A. (2010). Forecasting for intermittent
References demand for slow-moving items. Department working paper 12/10.
Quenouille, M. H. (1949). A relation between the logarithmic, Poisson, and
Akram, M., Hyndman, R. J., & Ord, J. K. (2009). Exponential smoothing and negative binomial series. Biometrics, 5, 162–164.
non-negative data. Australian and New Zealand Journal of Statistics, 51, Rao, A. V. (1973). A comment on: Forecasting and stock control for
415–432. intermittent demands. Operational Research Quarterly, 24, 639–640.
Al-Osh, M. A., & Alzaid, A. A. (1987). First-order integer valued Sani, B., & Kingsman, B. G. (1997). Selecting the best periodic inventory
autoregressive (INAR(1)) process. Journal of Time Series Analysis, 8, control and demand forecasting methods for low demand items.
261–275. Journal of the Operational Research Society, 48, 700–713.
Shenstone, L., & Hyndman, R. J. (2005). Stochastic models underlying
Billah, B., King, M. L., Snyder, R. D., & Koehler, A. B. (2006). Exponential
Croston’s method for intermittent demand forecasting. Journal of
smoothing model selection for forecasting. International Journal of
Forecasting, 24, 389–402.
Forecasting, 22, 239–247.
Shephard, N. (1995). Generalized linear autoregressions. Working paper.
Chatfield, C. (1992). Calculating interval forecasts. Journal of Business and
Nuffield College. Oxford.
Economic Statistics, 11, 121–144.
Snyder, R. D. (2002). Forecasting sales of slow and fast moving inventories.
Croston, J. D. (1972). Forecasting and stock control for intermittent
European Journal of Operational Research, 140, 684–699.
demands. Operational Research Quarterly, 23, 289–303.
Snyder, R. D., Koehler, A. B., & Ord, J. K. (1999). Lead time demand for
Czado, C., Gneiting, T., & Held, L. (2009). Predictive model assessment for
simple exponential smoothing: an adjustment factor for the standard
count data. Biometrics, 65, 1254–1261.
deviation. Journal of the Operational Research Society, 50, 1079–1082.
Davis, R. A., Dunsmuir, W. T., & Wang, Y. (1999). Modelling time series Stuart, A., & Ord, J. K. (1994). Kendall’s advanced theory of statistics (6th ed.)
of count data. In S. Ghosh (Ed.), Asymptotics, nonparametrics and time London: Edward Arnold.
series (pp. 63–114). New York: Marcel Dekker. Syntetos, A. A., & Boylan, J. E. (2001). On the bias of intermittent demand
Davis, R. A., Dunsmuir, W. T. M., & Wang, Y. (2000). On autocorrelation in estimates. International Journal of Production Economics, 71, 457–466.
a Poisson regression model. Biometrika, 87, 491–505. Syntetos, A. A., & Boylan, J. E. (2005). The accuracy of intermittent demand
Durbin, J., & Koopman, S. J. (2001). Time series analysis by state space estimates. International Journal of Forecasting, 21, 303–314.
methods. Oxford: Oxford University Press. Syntetos, A. A., Nikolopoulos, K., & Boylan, J. E. (2010). Judging the
Eaves, A. H. C., & Kingsman, B. G. (2004). Forecasting for the ordering judges through accuracy implication metrics: The case of inventory
and stock-holding of spare parts. Journal of the Operational Research forecasting. International Journal of Forecasting, 26, 134–143.
Society, 55, 431–437. Teunter, R. H., & Duncan, L. (2009). Forecasting intermittent demand:
Epstein, E. S. (1969). A scoring system for probability forecasts of ranked a comparative study. Journal of the Operational Research Society, 60,
categories. Journal of Applied Meteorology, 8, 985–987. 321–329.
496 R.D. Snyder et al. / International Journal of Forecasting 28 (2012) 485–496

West, M., & Harrison, J. (1997). Bayesian forecasting and dynamic models Ralph D. Snyder is an associate professor in the Department of
(2nd ed.) New York: Springer-Verlag. Econometrics and Business Statistics at Monash University, Australia.
West, M., Harrison, J., & Migon, H. S. (1985). Dynamic generalized models Having published extensively on business forecasting and inventory
and Bayesian forecasting (with discussion). Journal of the American management, he has played a leading role in the establishment of the class
Statistical Association, 80, 73–97. of innovations state space models for exponential smoothing.
Willemain, T. R., Smart, C. N., & Schwartz, H. F. (2004). A new approach
to intermittent forecasting demand for service parts inventories. J. Keith Ord is a professor in the McDonough School of Business,
International Journal of Forecasting, 20, 375–387. Georgetown University, Washington, DC. He has authored over 100
Willemain, T. R., Smart, C. N., Shockor, J. H., & DeSautels, P. A. (1994). research papers on statistics and its applications, and ten books, including
Forecasting intermittent demand in manufacturing: a comparative Kendall’s Advanced Theory of Statistics. He is a fellow of the American
evaluation of Croston’s method. International Journal of Forecasting, Statistical Association and the International Institute of Forecasters.
10, 529–538.
Zeger, S. L. (1988). A regression model for time series counts. Biometrika, Adrian Beaumont, a statistician, is a research assistant at Monash
75, 621–629. University.

View publication stats

Potrebbero piacerti anche