INTERNATIONAL JOURNAL OF CLIMATOLOGY
Int. J. Climatol. 24: 665–680 (2004)
Published online in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/joc.1027
ON THE ROLE OF STATISTICS IN CLIMATE RESEARCH
a
FRANCIS W. ZWIERSa, * and HANS VON STORCHb
Canadian Centre for Climate Modelling and Analysis, Meteorological Service of Canada, Victoria, BC, Canada
b Institute for Coastal Research, GKSS Research Center, Geesthacht, Germany
Received 22 March 2003
Revised 6 January 2004
Accepted 17 January 2004
ABSTRACT
We review the role of statistical analysis in the climate sciences. Special emphasis is given to attempts to construct
dynamical knowledge from limited observational evidence, and to the ongoing task of drawing detailed and reliable
information on the state, and change, of climate that is needed, for example, for short-term and seasonal forecasting. We
conclude with recommendations of how to improve the practice of statistical analysis in the climate sciences by drawing
more efficiently on relevant developments in statistical mathematics. Copyright 2004 Environment Canada. Published
by John Wiley & Sons, Ltd.
KEY WORDS:
statistical methods; statistical climatology; climate diagnostics; climate change; weather forecasting
1. INTRODUCTION
The study of the climate system is, to a large extent, the study of the statistics of weather; so, it is not
surprising that statistical reasoning, analysis and modelling are pervasive in the climatological sciences.
Statistical analysis helps to quantify the effects of uncertainty, both in terms of observation and measurement
and in terms of our understanding of the processes, that govern climate variability. Statistical analysis also
helps us to identify which of the many pieces of information derived from observations of the climate system
are worthy of synthesis and interpretation.
Statistical methods are needed for a whole gamut of activities that contribute to the ultimate synthesis
of climate knowledge, ranging from the collection of primary data, to the interpretation and analysis of
the resulting high-level data sets. Statistical procedures are fundamental components of procedures for data
retrieval from remotely sensed observations (e.g. satellite and radar), data retrieval from proxy records (e.g.
tree rings, coral records, ice and sediment cores), quality control (primarily to ensure the homogeneity
of observational data), determination of the representativity of data (e.g. urban warming effects), data
assimilation, the identification of predictive information in a given state (forecasts) and climate-change
detection. Statistical procedures are also integral to the large majority of efforts that seek physically meaningful
interpretations of observed climate variability. This includes the identification of ‘modes’ in the climate record,
such as the Arctic oscillation (AO) pattern (Thompson and Wallace, 1998), the identification of subsystems
such as the Madden–Julian oscillation (MJO; Madden and Julian, 1971) and the formulation of conceptual
models explaining the variability of climate (e.g. Hasselmann, 1976; Da Costa and Vautard, 1997; Burgers,
1999; Dobrovolski, 2000; Timmermann and Lohmann, 2000).
In the remainder of this paper we will briefly survey some of the aspects of climate science in which statistics
plays a central role. We will not, however, provide a total overview. The climate sciences, which, among
* Correspondence to: Francis W. Zwiers, Canadian Centre for Climate Modelling and Analysis, Meteorological Service of Canada,
Victoria, BC, Canada; e-mail:
[email protected]
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
666
F. W. ZWIERS AND H. VON STORCH
others, include atmospheric and oceanic physics, remote Earth observation, and palaeoclimatic reconstructions
from proxy data, are simply too broad for us to be exhaustive. Furthermore, our survey is strongly constrained
by our own research experience and interests. We hope that readers will not be offended by our selection of
topics. Although we discuss only a few areas in the climate sciences, the message that we hope to convey is
that the use of statistics is pervasive in the climate sciences, not only for the extraction and quality control of
data, but also for the synthesis of knowledge and information from that data. Thus, it is surprising that there
are only a very few comprehensive books on statistics for climatologists (e.g. Wilks, 1995; von Storch and
Zwiers, 1999) and only an additional few that deal with specialized topics (e.g. Daley, 1991; von Storch and
Navarra, 1999; Jolliffe, 2002; Jolliffe and Stephenson, 2003).
Some readers may be disappointed to see that we have not delved into fashionable topics such as neural
networks (e.g. Hsieh and Tang, 1998), cluster analysis (e.g. Mimmack et al., 2001) or wavelet analysis (e.g.
Farge et al., 1993) to any great extent. These techniques are no doubt very useful. For example, neural nets
have proven to be very useful as emulators of complex physical models. Such emulators can substantially
reduce the cost of the operational processing of high volumes of remotely sensed data (e.g. Schiller and
Doerffer, 1999). However, our perception is that the application of these kinds of techniques in diagnostic
studies has not generally improved our ability to synthesize knowledge, either from the observational record
or from the output of physically based dynamical models such as coupled global climate models. Rather,
much of the work that has had a large impact on climate research has used relatively simple techniques that
allow transparent interpretation of the underlying physics (e.g. Walker and Bliss, 1932; Madden and Julian,
1971; Wallace and Gutzler, 1981; Thompson and Wallace, 1998). A point that we will emphasize below is
that a state space approach, in which statistical models are used to describe what we have observed and
how this is related to the underlying dynamical system, is very appropriate for climate applications. Such an
approach is often used implicitly in climate research, in the sense that climate scientists do their work in a
framework that is defined by dynamical knowledge.
Some readers may also be disappointed that we have not surveyed recent developments in the statistical
sciences that may pertain to climate research. This is not to say that statistical science has nothing to offer.
Indeed, quite the opposite is true, as, for example, can be seen by reading the recent literature on extremevalue analysis (e.g. IPCC, 2002; Katz et al., 2002; Kharin and Zwiers, 2004) or forecast calibration and
verification (e.g. Jolliffe and Stevenson, 2003; Coelho et al., 2004). Rather, this choice reflects our experience
that new technology developed in the statistical sciences cannot simply be transferred to the climate sciences
by informing climatologists of its existence. Effective technology transfer across disciplines requires crossfertilization, such as has occurred in geostatistics, and as is being fostered by efforts such as the National
Center for Atmospheric Sciences (NCAR) Geophysical Statistics Project (http://www.cgd.ucar.edu/stats/).
The remainder of this paper is organized as follows. Section 2 deals with the use of statistical methods to
learn about the dynamics of the climate system. In Section 3 we discuss some of the roles of statistics in the
acquisition of data, and in Section 4 we discuss the role of statistics in forecasting. We continue in Section 5
with a discussion of climate-change detection, and complete the paper in Section 6 with some additional
discussion, a summary, and a few recommendations.
2. CONSTRUCTING KNOWLEDGE ABOUT THE DYNAMICS OF CLIMATE
When we construct knowledge, i.e. produce a structured, reasoned judgment about the functioning of the
climate system, we implicitly or explicitly assume a state space model (e.g. Honerkamp, 1994). These
statistical models clearly discriminate between a system of state variables that define a theoretical construct
and a collection of observables that contain information about the system. More formally, we describe
the system with a state vector φt that is continuous in space and time. Almost always, this state vector
cannot be observed, sometimes because of lack of suitable sensors, but also because continuous space–time
observations are generally not available. The dynamics F of this state variable are often described by a system
of differential equations:
∂t φ = F(φ, α, t) + ε
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
(1)
Int. J. Climatol. 24: 665–680 (2004)
STATISTICS IN CLIMATE RESEARCH
667
The dynamics depend on a set of free parameters α = (α1 , α2 , . . .). The ‘noise’ term ε that appears on the
right-hand side of Equation (1) arises because the dynamics can only approximately describe the behaviour of
the system regardless of the choice of the parameters, and because of the fact that seemingly random effects
act upon the system. Of course, the dynamics may generate internal noise as well, and noise may enter the
system in more subtle ways than suggested by the simple additive model in Equation (1). The dynamics F
may be derived from theoretical arguments, such as the conservation of momentum or mass, or less frequently
from an empirical fit.
Even if φt is not observable in its entirety, observations containing information about φt may be available,
for instance at some locations and at some times. Alternatively, indirect evidence may be available from proxy
information contained in media such as tree rings, ice cores, corals, and lake warves (e.g. Fritts, 1991; Cook,
1995; Bradley, 1999; Mann et al., 1999; Briffa, 2000; Crowley and Lowery, 2000; Jones et al., 2001a; Esper
et al., 2002). When these observations are combined into an observation vector ωt , an observation equation
can be used to relate the state variable to the observed variables
ωt = P(φt ) + δt
(2)
through an operator P. The noise term δt on the right-hand side of Equation (2) indicates that the observation
equation is not exactly satisfied; there may be measurement uncertainties with respect to the value, location
or timing of ωt . As with the dynamics, noise may enter into the observations more subtly (e.g. perhaps as
a multiplicative term) than indicated by Equation (2). Also, the link between φt and ωt may not be fully
known, as is the case with proxy data.
The state space approach is the core of Hasselmann’s principal interaction pattern (PIP) concept
(Hasselmann, 1988), with the PIPs being a basis of an optimally determined subspace that contains the
relevant dynamics. This concept is general and difficult to implement. Examples are provided by Achatz
et al. (1995), Achatz and Schmitz (1997), Kwasniok (1996) and Selten (1995). An important aspect of this
approach is that it provides a means for estimating (an upper limit of) the number of degrees of freedom
of the dynamical phenomenon and to characterize the system’s attractor (in the sense of a manifold within
which the system may be considered approximately closed) by specifying suitable coordinates. Selten (1995)
and Kwasniok (1996), for instance, found that less than 20 degrees of freedom were enough to describe the
Northern Hemisphere equivalent barotropic dynamics. Reduced models based on about 30–50 degrees of
freedom were found to be able to capture the essential features of the long-term behaviour of a spectral T42
quasi-geostrophic climate model.
Conceptually simpler analyses formulate a model, in a given space, with a few unknown parameters. The
task of the statistician is then to estimate these parameters and to determine the skill of the fitted model
in reproducing the variability of the observables. Examples include the delayed action oscillator model of
the El Niño–southern oscillation (ENSO) phenomenon (with local wind and sea-surface temperature (SST)
observations as observables; Suarez and Schopf, 1988; Battisti and Hirst, 1989; Mechoso et al., 2003) and
the stochastic climate model (with the spectra of, for instance, SST as observables; Hasselmann, 1976;
Frankignoul, 1985). The analysis of observed data in the framework of a spatial wave-like phenomena, as
in the case of principal oscillation patterns (POPs; e.g. analysis of the MJO by von Storch and Xu (1990))
or a frequency wavenumber analysis (analysis of extratropical storm track by Hayashi (1982) and Speth
and Madden (1983)) are other examples. Questions regarding air–sea interaction (Does the ocean drive the
atmosphere or vice versa?) can also be addressed successfully in this format (von Storch, 2000; Zorita et al.,
1992; Goodman and Marshall, 2003).
A widely pursued line of research is the identification of ‘modes’ in the observational record or in extended
quasi-realistic model simulations. A large variety of methods have been developed in the past several decades,
with empirical orthogonal functions (EOFs) as the most frequently used. Other techniques include variants
of EOFs, such as rotated EOFs (Richman, 1986), extended EOFs (Weare and Nasstrom, 1982), Hilbert EOFs
(or, as they are often named, complex EOFs; Wallace and Dickinson, 1972), singular spectrum analysis (SSA;
Vautard et al., 1992) and multi-channel SSA (MSSA; Vautard, 1995). Other techniques used to identify modes
include the analysis of teleconnections (Wallace and Gutzler, 1981), empirical orthogonal teleconnections (van
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
668
F. W. ZWIERS AND H. VON STORCH
den Dool et al., 2000), composite analysis, canonical correlation analysis (Barnett and Preisendorfer, 1987),
singular value decomposition (Bretherton et al., 1992) and redundancy analysis (WASA, 1998; Wang and
Zwiers, 2001). See von Storch and Zwiers (1999) for an overview.
In all cases, the quest is not for information about the state of the system, but rather about the dynamics of
the system. Initially, the goal will be modest, i.e. simply to gather information that will eventually allow one
to articulate the dynamics encompassed in Equation (1). Ultimately, however, the question becomes more
specific: Does Equation (1) successfully describe the observed variability?
The danger with ‘blind’ techniques, such as canonical correlation, EOF, POP and other types of analysis, is
that they deliver dynamically interesting looking numbers and patterns, but reliable problem-independent tests
are unavailable. Confirmatory analysis, whereby one tests a specific hypothesis developed from exploration
with such techniques on independent data, is hardly possible in this context. This is not to say that descriptive,
exploratory analysis with these tools is not useful. The systems considered are complex and non-linear, and
often do not yield to analysis with first principles. In that case, exploratory analysis is a valuable aid that may
lead to recognition and identification of dynamical mechanisms of variability. One should, however, keep in
mind that the patterns obtained from methods such as EOF analysis are not a priori constructed to extract
information about dynamical behaviour: they simply provide an efficient representation of variability. There is
no guarantee that the often tacitly made assumption that EOFs represent dynamical ‘modes’ of variability will
hold. The recent discussion about the dynamical interpretation of the AO and the North Atlantic oscillation is
an indication of this problem (Baldwin and Dunkerton, 1999; Deser, 2000; Monahan et al., 2000, 2001, 2003;
Wallace, 2000; Ambaum et al., 2001; Thompson and Wallace, 2001). Also the discussion about the ‘Indian
Ocean dipole’ (Saji et al., 1999) is sometimes drawn in this direction (Dommenget and Latif, 2002, 2003;
Behera et al., 2003; Jolliffe, 2003). Another such example concerns the interpretation of the coupled modes
that are obtained by performing a singular vector decomposition on a cross-covariance matrix (Bretherton
et al., 1992; for subsequent caveats on the usage of this technique see Newman and Sardeshmukh (1995) and
Cherry (1996)).
A symptom of the difficulty with confirmatory analysis is the often heard, ill-posed, question about the
significance of EOFs. If ‘significance’ is meant to stand for ‘non-degenerate’ (in the sense that the eigenvectors
are not primarily mixed combinations of the true underlying eigenvectors), then tools such as the North et al.
(1982) ‘rule of thumb’ appear to be useful, even though the inferences made with these rules may not be
precise. For example, whereas the North et al. (1982) rule has its basis in asymptotic statistical theory, we
inevitably apply it in circumstances when samples are very far from being asymptotically large. But simple
‘rules’ are not available if the question concerns the uncertainty of the pattern itself. Here, it is necessary to
build confidence and understanding through the analysis of simplified examples, to articulate the particular
question carefully regarding the pattern uncertainty that we wish to assess, and then to apply resampling
tools carefully to the full pattern estimation procedure (e.g. Michaelson, 1987; Barnston and van den Dool,
1993; Efron and Tibshirani, 1993). In most cases, the only credible way in which to discriminate between the
effects of the sampling properties of the observables and truly dynamical features of the system is dynamical
plausibility and reproducibility in detailed dynamical models.
3. ACQUIRING HIGH-QUALITY, LOW-LEVEL INFORMATION
Traditionally, from the 19th to the middle of the 20th century, climatology was bookkeeping for meteorology.
The problem was to record in an objective and representative way typical climatic variables such as nearsurface minimum, maximum and mean temperature, precipitation amounts, surface pressure, wind speeds,
wave heights and the like. These data were, and are, used for planning and safety standards of houses and,
for instance, offshore operations. As such, the data need to be comparable — in space and time.
In many applications the data are assumed to fit a stationary (if we disregard the annual cycle) probability
distribution for describing the ‘normal’ range of variability, or to fit an extreme value distribution for assessing
risks of rare events. This task does not really pose a challenge for modern climatology as long as the assumption
of stationarity is satisfied. The challenge is with temporally changing statistics. This is not a new challenge:
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
STATISTICS IN CLIMATE RESEARCH
669
Brückner (1890) described a systematic approach for making data comparable in space and time in order to
describe large-scale climatic variations objectively over centuries. Brückner (1890), among other variables,
attempted to homogenize water-level records from the Caspian Sea by comparing simultaneous observations
from different locations.
Often, records exhibit creeping or sudden changes. Sometimes these changes are not due to changes in the
climatic variables, but to changes in the instruments used to measure the variables, the physical characteristics
of the instrument’s immediate environment (its exposure), and recording and observing procedures. Easy-tounderstand effects relate to a change in the exposure of instruments, as described by Karl et al. (1993),
Peterson et al. (1998) and Vincent and Gullett (1999) among others. Also, observation practices, such as
changes in the observation time (already noted by Ellis (1890) and Donnell (1912); a modern reference is
Karl et al. (1986)) or changing ship routing affect the climate record. More examples are listed in Folland et al.
(2001). Less obvious effects are found in remotely sensed data, e.g. related to the changing flight altitude
(orbital drift), which played an important role in the recent controversy about the apparent inconsistency
between the warming at the surface and the much smaller rate of warming in the lower troposphere
(e.g. Christy et al, 2000, 2001, 2003; NRC, 2000; Santer et al., 2000, 2003; Folland et al., 2001; Mears
et al., 2003).
Other unwanted effects appear if the data are not representative of what they are assumed to be. Growing
populated areas produce changes in the local micro-climate (urban warming) that are real and, when studying
the climate of cities, very relevant. However, such local changes cannot be interpreted as being representative
of climate change on the continental or even global scale. Climatologists understand how to avoid these
problems (Jones et al., 1990; Gallo and Owen, 1999; Peterson and Vose, 1997; Peterson et al., 1999), but
misunderstanding about this effect has been a cause for continuing doubt about global warming among
lay people.
Statistics helps in these cases in first discriminating such changes as being either ‘inhomogeneities’,
i.e. related to changes in the instrument, observing procedure, the local environment, or a true climatic
variation, and to correct for inhomogeneities if needed. Sudden changes are sometimes detected with
classical ‘change-point’ analysis (e.g. Busuioc and von Storch, 1996). Sudden and creeping changes
are found by systematically comparing neighbouring stations, a technique that was already in use by
Brückner in 1890. Peterson et al. (1998) give an extensive overview about the state of the art. Also,
for indirect climate data, like tree rings, such techniques to correct for non-climatic influences have
been developed (e.g. Cook, 1995; Fritts, 1976, 1991). In addition, dendro-climatologists have developed
several data-adjustment techniques that better preserve low-frequency and secular variability when several
records are joined together to form a very long temperature time series (e.g. Briffa et al., 2001; Esper
et al., 2002).
Another modern challenge is the global, dynamically consistent analysis of the synoptic state of the
atmosphere or ocean. Geostatistical techniques, which include ‘optimal interpolation’ in meteorology, have
been used successfully; also, Kalman-filter-type techniques have been developed, tested and applied (Evensen,
1992, 1994; Burgers et al., 1998; Houtekamer and Mitchell, 2001). Such combined statistical–dynamical
analysis is in routine use at the weather services. In case of the European Centre for Mediumrange Weather Forecasting (ECMWF) and the National Centers for Environmental Prediction (NCEP),
these schemes have been used to analyse retrospectively weather observations for periods of 15 years
(ECMWF ERA-15, 1979–94; Gibson et al., 1997) and 50 years (NCEP, 1958–2001; Kalnay et al.,
1996). The ECMWF has recently completed an extended, high-resolution global 40 year reanalysis (see
http://www.ecmwf.int/research/era). These retrospective products have been immensely useful to climate
research because they provide data with complete global coverage that are as consistent as possible
with our archives of in situ and remotely sensed observations, and that are not affected by the frequent
changes that occur in operational analysis systems that assimilate data into weather forecasting models. However, there remains an urgent need for very high resolution (<50 km) regional retrospective
analyses for the study of climate on scales where the impacts of climate variability and change can
be assessed.
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
670
F. W. ZWIERS AND H. VON STORCH
4. FORECASTING
Statistics, and statistical reasoning, has also played key roles in the forecasting of weather and climate on
time scales ranging from hours to several seasons. Statistical methods have been used to assess the potential
predictability of climate and weather, to develop schemes for initializing dynamical forecasting models, for
post-processing dynamical forecasts (both to remove biases and to add additional skill), and to forecast future
weather and climate states empirically without the aid of dynamical models. Statistical concepts have also
been key to the evaluation of forecasting systems and have provided objective forecast skill scores that cannot
be manipulated by the forecaster.
4.1. Weather forecasting
The multiply chaotic evolution of weather, with rapidly growing differences between any two rather similar
states (analogues), makes weather forecasting a field for statistical analysis. First, the problem of forecasting
is cast in a statistical framework: the future weather state is understood as a random variable conditioned
upon the initial state and, in certain cases, external parameters such as SST or soil moisture. Of course, the
forecast lead, i.e. the time in the future that the forecast is supposed to predict, is an important parameter as
well. In the following we will briefly touch upon four problems associated with weather forecasting that have
been dealt with via statistical methods: initial conditions, predictability, forecast improvement via the perfect
prognosis (PP) or model output statistics (MOS) approach, and the assessment of forecast skill.
Early weather forecasts, whether subjective or objective, were essentially statistical. Forecasts were often
based on a process called weather typing (e.g. Köppen, 1923), which is similar to modern-day cluster analysis
of weather states. In contrast, modern weather forecasts are produced by numerical weather prediction (NWP)
systems, with sophisticated data assimilation systems for determining the initial state plus advanced highresolution dynamical global or regional models of the atmosphere. (Note that ‘frozen’ versions of such systems
are used to produce the retrospective analyses (reanalysis) discussed in Section 3). Statistics continues to play
an important role in NWP. At the ‘front end’ of the forecasting system, data assimilation systems are used to
specify the initial conditions for NWP models. More broadly, data assimilation is closely related to objective
analysis, the spatial (and sometimes temporal) interpolation of data that are scattered irregularly in space
(and time) to a regular grid in space (and a fixed point in time). The methods used range from techniques
that are primarily statistical to methods that are fully integrated with the dynamics of the system that is
being analysed.
The primarily statistical techniques, which are often termed objective analysis, include ad hoc methods
such as Cressman (1959) successive correction (see Given and Ray (1994) for an application; see Trapp
and Doswell (2000) and Askelson et al. (2000) for discussion of the use of this technique in the analysis
of radar data), to optimal interpolation techniques (Gandin, 1963; Thiébaux and Pedder, 1987; see Sokolov
and Rintoul (1999) for a recent inter-comparison of techniques). In the latter, care is often taken to model
the spatial correlation structure of forecast errors so as to maintain the dynamical balance between forecast
parameters, such as geopotential and winds (e.g. Thiébaux, 1976). At the opposite end of the spectrum are
modern three-dimensional (space only) and four-dimensional (space and time) variational techniques (e.g.
Daley, 1991) that incorporate so-called adjoint models (e.g. Le Dimet and Talagrand, 1986; Thépaut and
Coultier, 1991; Tanguay et al., 1995) of the atmospheric dynamics. These techniques are now also being
coupled with state space (Kalman filtering) approaches that include explicit representation of the dynamics of
the system that is to be predicted and the statistical nature of the observables (e.g. Daley, 1991; Rabier et al.,
1998; Houtekamer and Mitchell, 2001). An overview of methods used in oceanography is given by Robinson
et al. (1998).
Statistics also plays an important role at the output end of NWP systems. The performance of these NWP
systems is far better than that of any statistical forecast system on scales of a few hundred kilometres and
greater, but they often fail to deliver skilful forecasts of local, impact-relevant variables, such as the danger of
freezing or sunshine duration in a physiographically structured landscape. Statistical methods have long been
used to correct for systematic errors, particularly biases, in NWP products and to derive transfer functions
that relate well the forecast tropospheric variables to impact-relevant surface variables.
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
STATISTICS IN CLIMATE RESEARCH
671
Two classes of forecast-improvement methods have been developed, namely PP (Klein et al., 1959) and
MOS (Glahn and Lowry, 1972; Klein and Glahn, 1974). In the former, a statistical model, often a multiple
regression, is derived from simultaneous observations of the tropospheric variables and the variable of interest,
whereas MOS ‘consists of determining a statistical relationship between a predictand and variables forecast
by a numerical model at some projection time’ (Glahn and Lowry, 1972). The latter is more efficient, and
is able to adjust for forecast model biases and other kinds of systematic error, but it needs to be updated
whenever a component of the forecast system is modified. The PP approach has received renewed interest
in the past 10 years for ‘downscaling’ large-scale climate-change information to the local level (e.g. Wilby
et al., 1998; Widman et al., 2003). The MOS approach has continued to be used for weather forecasting, and
considerable research has been conducted on updatable MOS systems that more easily adapt to the frequent
changes that are made in weather forecasting systems (Ross, 1987; Wilson and Vallée, 2002; Yuval and Hsieh,
2003). MOS approaches are now also being used extensively for seasonal forecasting (see below).
In addition, statistics plays important roles in the study of predictability (Lorenz, 1982) and in the assessment
of the skill of forecasting systems (e.g. Mason and Graham, 1999). The predictability question relates to the
forecast lead for which all forecast skill is lost, i.e. that lead at which the distribution of the verifying
observations, conditional upon the forecast, is as wide as the unconditional distribution. Related to this
is the question of whether the uncertainty of the forecast itself may be predicted. Advancement in all of
these areas (the specification of initial conditions, predictability studies, and dynamical and statistical forecast
improvements) is measured by means of forecast skill scores (Murphy and Epstein, 1989; Gandin and Murphy,
1992; Murphy and Wilks, 1998; Mason and Graham, 1999; Wilks, 2000). Overviews are given by Stanski
et al. (1989), Livezey (1999) and von Storch and Zwiers (1999).
4.2. Seasonal forecasting
The instantaneous state of the atmosphere is not generally predictable beyond about 2 weeks because of its
inherent chaotic nature. Nonetheless, in many parts of the world there has been gradual progress in forecasting
seasonal mean conditions with leads of up to about a year. This has been achieved by conditioning on parts
of the climate system that evolve more slowly than the atmosphere and communicate with the atmosphere
(e.g. see Shukla (1998) and references cited therein). Most effort to date has focused on the tropical ocean
as the source of predictable climate signals on time scales of months to seasons (e.g. Livezey et al., 1997;
Hoerling and Kumar, 2002, 2003) although it is recognized that there may also be other signal sources, such
as soil moisture (Fennesy and Shukla, 1999).
Considerable work has been done to assess the potential predictability of climate on seasonal to interannual time scales (e.g. Zwiers et al., 2000 and references cited therein), and decadal time scales (e.g. Latif
and Barnett, 1996; Rowell and Zwiers, 1999; Saravanan et al., 2000; Collins and Allen, 2002). Much of this
work has used the potential predictability concept that was first outlined and applied by Madden (1976). This
is essentially a classical analysis of variance (ANOVA; e.g., Scheffé, 1959) conducted in the time domain so
that one looks for evidence of excess variability on time scales longer than the scale of interest (e.g. seasonal;
see Zwiers (1996)). This evidence is interpreted as an indication of potential predictive skill.
Observational studies (e.g. Madden, 1976) have been supplemented with perfect model studies in which
the time evolution of a slow part of the climate system (typically SST) is prescribed. Two-way ANOVA is
often applied to ensembles of such simulations to quantify the proportion of inter-annual or decadal variability
that is ascribable to lower boundary forcing of the climate system (e.g. Zwiers, 1996; Rowell, 1998; Zwiers
et al., 2000).
The most reliable seasonal forecasting tools are presently statistical, prevalently canonical correlation
analysis (e.g. Barnston, 1994; Shabbar and Barnston, 1996; Hwang et al., 2001). This approach is moderately
successful in parts of the world where there are teleconnections between the ocean surface and the atmosphere.
Many other approaches have been used, including POP analysis (Tang, 1995), SSA, MSSA, neural networks
(Tangang et al., 1998).
While statistical methods are prominent, the widely accepted goal is to be able to produce reliable dynamical
forecasts. Most dynamical systems are presently two tier (Bengtsson et al., 1993), meaning that lower
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
672
F. W. ZWIERS AND H. VON STORCH
boundary conditions (usually SST) are first forecast, followed by a calculation to determine the atmospheric
response to the boundary forcing forecast. Lower boundary forecasts may be statistical or dynamical in nature.
Most systems presently require MOS to adjust for biases in the forecast system (e.g. Derome et al., 2001) and
to link surface parameters to the generally more skilfully forecast large-scale circulation. This is an area of
active research (e.g. Krishnamurti et al., 1999, 2000; Doblas-Reyas et al., 2000; Kharin et al., 2001; Kharin
and Zwiers, 2002, 2003a,b; Yun et al., 2003). Anderson et al. (1999) assess statistical and dynamical seasonal
forecasting methods. Kharin and Zwiers (2003a) describe several methods for producing probability forecasts
of seasonal conditions, and Kharin and Zwiers (2003b) discuss some properties of skill scores that are often
used to evaluate seasonal forecasts. In connection with seasonal probability forecasts, see also Buzzia and
Palmer (1998), Mason and Graham (1999) and Wilks (2001).
Skill is considerably more difficult to assess in climate forecasting than in NWP because of the increased
time scale of the forecasts, and hence the reduced frequency with which forecasts can be made and evaluated.
This substantially slows progress in this area. Kharin and Zwiers (2002) discuss the limitations of some
seasonal forecasting improvement techniques. Wilks (2000) demonstrates an evaluation of the skill of seasonal
forecasts produced at the NCEP and discusses diagnostics of the consistency of seasonal forecasts that are
issued monthly.
5. CLIMATE-CHANGE ASSESSMENT
Human activity is altering the composition of the Earth’s atmosphere through the addition of greenhouse gases
such as carbon dioxide and chlorofluorocarbons, and various species of aerosol. Science has been aware of
the potential for global warming by greenhouse gases at least since Arrhenius (1896). This prospect has
been the focus of much physical research and data analysis during the past 20 years. See, for example, the
assessments of the Intergovernmental Panel on Climate Change (e.g. Houghton et al., 2001), and the United
States National Assessment (NAST, 2001; MacCracken et al., 2003).
Several external forcing factors are thought to affect the mean state of the climate on decadal and longer
time scales (Houghton et al., 2001). These include anthropogenically caused changes in radiative forcing
that result from the emission of greenhouse gases and the creation of aerosols from fossil fuel and biomass
burning. In addition, there are natural external forcing factors, such as changes in solar irradiance and volcanic
activity. However, the climate system does not vary because of external factors alone. The climate system,
even when not perturbed by external factors, produces substantial amounts of natural internal variability (von
Storch et al., 2001) on large spatial scales and long temporal scales. Hence, detection and attribution of the
effects of external forcing are statistical signal-to-noise problems (Hasselmann, 1979, 1993).
The detection part of this problem is the process of demonstrating that an observed change is not likely
to have been entirely the result of natural internal variability. The attribution aspects of the problem are
more difficult because it is not possible to conduct controlled experiments with the climate system. The
practical approach that has been taken in the climate research community involves statistical analysis and
the assessment of multiple lines of evidence to demonstrate that: (a) observed changes are consistent with
forcing of the climate by a combination of anthropogenic and natural external factors; and (b) the changes
are inconsistent with alternative, physically plausible explanations.
The detection technique that has been used in most studies performed to date has several equivalent
representations (Hegerl and North, 1997; Zwiers, 1999; Hegerl and Allen, 2002). It can be cast as a multiple
regression problem (Hasselmann, 1993, 1997; Allen and Tett, 1999) in which a field of n ‘observations’ y is
represented as a linear combination of signal patterns g1 , . . . , gm plus residual climate noise n:
y=
m
ai gi + n = Ga + n
(3)
i=1
where G = (g1 | . . . |gm ) is the matrix composed of the signal patterns and a is the vector composed of the
unknown amplitudes.
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
STATISTICS IN CLIMATE RESEARCH
673
The signal patterns are derived by using physically based models to simulate the response to estimated
changes in external forcing that are thought to have occurred during the past two centuries. These physical
models generally also simulate internal variability. Therefore, the signals (responses to external forcing) are
usually estimated by averaging a small ensemble of runs with the same model, each run being started from
different initial conditions. This averaging reduces contamination of the signal by the internal noise that is
simulated by the physical model. The amplitudes a are estimated either by means of a standard least-squares
method, or by the total least-squares method (Ripley and Thompson, 1987). The latter method is generally
preferred because it takes account of the fact that the signal estimates are not completely free of the effects
of internal noise.
The detection issue is dealt with by testing the null hypothesis HD : a = 0 that the signals have zero amplitude
in the observations. The test that is used in most cases is a variant of the Hotelling T 2 -test, although Bayesian
approaches to assess the presence of the signals in the data have also been used (e.g. Berliner et al., 2000;
Schnur and Hasselmann, 2004).
Research on climate-change detection and attribution over the last 10–20 years has searched for evidence
of anthropogenic and natural external signals primarily in the observed global temperature record. Some recent
studies have also considered some other climate elements, such as the global ocean heat content (Barnett et al.,
2001; Reichert et al., 2002) or even the propagation of ocean waves (Pfizenmayer and von Storch, 2001).
Statistical evidence typically contributes to an attribution assessment by testing the null hypothesis
HA : a = 1 that the model simulated signals all have the correct amplitude. If this can be demonstrated,
then there is evidence that the physical model has responded to the historical changes in external forcing in
the same way as the observed system, which in turn contributes to a comprehensive attribution assessment, as
described in Mitchell et al. (2001). A difficulty with this approach is that the analyst must interpret a failure
to reject HA as evidence in support of an attribution assessment. This situation is not entirely acceptable,
because it means that the likelihood of attribution cannot be controlled by acquiring more information or
better delineating the signals G. In fact, when more data become available, the power of the test increases,
making the rejection of HA more likely. Levine and Berliner (1999) have pointed out that a more satisfactory
approach would be to conduct a test of inconsistency in which consistency is the alternative hypothesis. Such
a test has, apparently, not yet been applied in the climate literature. Bayesian methods (e.g. Hasselmann,
1998; Berliner et al., 2000; Schnur and Hasselmann, 2004) approach the problem in a more satisfactory way
by using evidence from the observations to estimate the a posteriori probability of attribution for a suitable
defined attribution criterion.
A difficulty with the detection and attribution analysis is that an estimate of the covariance matrix Cnn of the
residual noise field is required to make statistical inferences about the amplitudes a. However, the instrumental
record gathered during the past 150 years (e.g. Jones, et al., 2001b; Jones and Moberg, 2003) cannot provide
a reliable estimate of residual noise covariability. This is because the length of the observed record is not
sufficient to estimate variability on the decadal and longer time scales that are important for detection and
attribution. Also, natural internal variability is confounded with the effect of anthropogenic and natural
forcing during the instrumental period. Palaeo-reconstructions of past climate are a possible future source of
information for this purpose. There is growing confidence that these records provide realistic representations
of hemispheric-scale decadal variability (e.g. Hegerl et al., 2003), but they presently do not adequately resolve
spatial variability on the scales needed for detection and attribution. Thus, the covariance matrix is generally
estimated from long control simulations performed with a climate model in which concentrations of greenhouse
gases and aerosols are fixed at present or pre-industrial levels.
The climate models used in detection and attribution studies are likely not to be able to simulate climate
variability correctly on all spatial and temporal scales. This problem is circumvented, in part, by performing
the detection and attribution analysis in a reduced dimension space that is spanned by a small number of
EOFs of the estimated internal variability. Thus, a constraint on the choice of the number of EOFs is that
the variability of the residual noise should be consistent with the variability of the control simulation in the
dimensions that are retained. An approximate chi-squared test can be used for this purpose (Allen and Tett,
1999). Detection and attribution studies typically use approximately 10 EOFs, although some studies (e.g.
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
674
F. W. ZWIERS AND H. VON STORCH
North and Wu, 2001) have used a much higher level of truncation. Experience has shown that results are
typically not very sensitive to the level of truncation (e.g. see Zwiers and Zhang (2003)).
Multiple linear regression is not the only technique that has been used in detection and attribution studies.
There is also a substantial body of work (see Santer et al. (1996) and Zwiers (1999)) that relies on pattern
correlation methods. These methods are closely related to optimal detection with one signal pattern.
6. SUMMARY AND DISCUSSION
We have tried to survey some of the ways in which statistics pervades the climate sciences. Obviously, a total
overview is impossible: the climate sciences, which include atmospheric and oceanic physics, remote Earth
observation, and palaeoclimatic reconstructions from proxy data, are simply too broad for us to be exhaustive.
We have covered only a limited part of the field, and have not used this paper to attempt to transfer new
technology to climate science from the statistical sciences. Instead, our purpose has been to demonstrate that
statistical concepts and methods are necessary in all facets of the climate science enterprise, ranging from the
gathering of data to the derivation of knowledge from carefully synthesized data products.
Statistical analysis is needed to interpret observations, from either the real world or the artificial world of
a climate model, in a proper framework. Statistical reasoning allows analysts to extract information about
some part of the climate system by making a number of simplifying assumptions about the way in which
the system generates information (e.g. an observed process is stationary and ergodic) and about the way in
which the data analysed have been observed. Within the context of these assumptions, statistical reasoning
imposes an important element of rigour when extracting information from data. Again, within the context of
the explicit, and implicit, approximating assumptions, statistical methods allow one to deal explicitly with the
effects of uncertainty on inferences and to quantify its effects on forecasts, projections, etc.
This need for statistical thinking stems foremost from two particularities of the climate system, which are
that the climate system has a large number of components (degrees of freedom) and that it is impossible
to conduct laboratory experiments with the Earth system. Consequently, there is considerable scope for the
misinterpretation of statistical evidence. Flawed results can be avoided, not through the use of a ‘silver bullet’,
but by clearly articulating all assumptions required to apply a given analysis technique. This suggestion applies
to both the concrete analysis of specific data and to the conceptual framework within which we build our
theories and knowledge (Petersen, 1999; Sarewitz and Pielke, 1999; von Storch, 2001a,b). For the latter, the
concept of state space models is particularly useful, as it helps us to discriminate between our hypothetical
construction of dynamics and the observational process. Flawed results, when they occur, are not easy to
identify, and significant work is sometimes required to uproot them. Examples of the successful disclosure
of such problems are given by Allen and Smith (1996) and Nitsche et al. (1994).
When dealing with environmental systems, different types of uncertainty arise (Sarewitz and Pielke, 1999;
Risbey et al., 2000). One type of uncertainty is inherent, like the weather in 50 days or greenhouse gas
emissions in 50 years. No improved understanding of atmospheric or social dynamics will provide us with
information; the only way to reduce the uncertainty is to wait for 40 days, or 40 years, until we enter the
period prior to the future date of interest in which the weather or greenhouse gas emissions are actually
predictable. However, there is another type of uncertainty, which is temporary or malleable. By collecting
additional data and by improving our conceptual understanding of the phenomenon under consideration, we
may in some cases reduce uncertainty. One example is the additional information provided by the Tropical
Atmosphere Ocean buoy array in the equatorial Pacific, which allows a more complete analysis of the state
of the tropical Pacific and has thus provided knowledge to improve ENSO forecasting (see the extensive
bibliography available at http://www.pmel.noaa.gov/tao/). A second example is the improved understanding
of the climate system that helps us to constrain our estimates of the climate sensitivity to increased greenhouse
gases (e.g. Gregory et al., 2002).
The role of statistics is, again, twofold. It helps to quantify the degree of ‘inherent’ uncertainty, and it helps
to assess the value of learning. The Bayesian approach seems to be more successful in dealing explicitly
with both types of uncertainty, as is demonstrated for instance by Risbey et al. (2000) in an analysis of the
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
STATISTICS IN CLIMATE RESEARCH
675
inherent and malleable uncertainties of global warming. Another example is Hasselmann’s (1998) study of
the role of preconceptions in detecting climate change (also see Schnur and Hasselmann (2004)). An account,
mainly seen from the standpoint of theoretical statisticians, on the potential of Bayesian methods in climate
analysis is provided by Berliner et al. (1998).
In frequentist analysis, the approach that most readers are familiar with, assumptions about uncertainty are
often hidden. In general, we condition our inferences on many things without being explicit. An example
is climate-change detection: the methodology used in most studies (see Section 5) implicitly assumes that
our current assessment of potentially important external forcings (greenhouse-gas emissions, sulphur dioxide
emissions, solar variability, and volcanic variability) is exhaustive. To be fair, however, we should hasten to
add that most authors of climate-change-detection papers do clearly state the limitations of their analyses.
Clearly, inferences will only be as good as our assumptions about uncertainty.
We have the impression that the discussion about statistical methodology in the climate sciences is generally
not very deep and that straightforward craftsmanship is pursued in many cases. As a consequence, much of
the statistical practice in climate science is of a home-grown nature. It is often ad hoc, and is sometimes
not well informed or supported by statistical theory. Clearly, the link between climate and the statistical
sciences should continue to be improved with additional efforts such as the Geophysical Statistics Project
(http://www.cgd.ucar.edu/stats/) at the NCAR. However, such links cannot consist of simple unidirectional
transfers of knowledge from the statistical sciences into the climate sciences: statisticians need to collaborate
closely with climate scientists to ensure that the advanced methods that they develop will yield new information
that can provide new, clear, insights about the climate system and its dynamics.
We feel that the cooperation between the statistical and climate sciences does not function as well as
that between, for example, statistical and biomedical science. Among other reasons, this is due to the two
particularities mentioned above: confounded dynamics with a large number of degrees of freedom, so that
it is not always easy to identify isolated problems amenable to statistical analysis, and the impossibility of
generating additional independent data in experiments (apart of numerical experiments with physically based
models). Thus, better communication between statisticians and climatologists requires a better understanding
by statisticians of the specifics of climate science, and a greater effort by climatologists to communicate the
specifics of open problems to statisticians.
One way to overcome these communication problems between the different scientific cultures of statistical
and climate sciences is to arrange many more occasions where statisticians can meet climatologists, including
meteorologists, oceanographers and other geo-scientists, in a constructive environment. Successful activities
along these lines include the International Meetings on Statistical Climatology (Murphy and Zwiers, 1993;
also see http://imsc.seos.uvic.ca/), whose ninth meeting will take place in May 2004, in Capetown, South
Africa, or the NCAR Geophysical Statistics Project cited above. Useful activities in this respect in the past
include the 1993 and 1997 ‘Aha Huliko’a Hawaiian Winter Workshops on ‘Statistical Methods in Physical
Oceanography’ and ‘Monte Carlo Simulations in Oceanography’ (Müller and Henderson, 1993, 1997), the
1993 Autumn School ‘Analysis of Climate Variability — Applications of Statistical Techniques’ organized
by the Commission of the European Community (von Storch and Navarra, 1999) and the ‘Statistics and
Physical Oceanography Report’ (Chelton, 1994; Panel on Statistics and Oceanography, 1994). Many more such
activities are required. Greater opportunities for joint research, which means a greater emphasis by funding
agencies on cross-disciplinary research that specifically links climatologists and statisticians, is also required.
REFERENCES
Achatz U, Schmitz G. 1997. On the closure problem in the reduction of complex atmospheric models by PIPs and EOFs: a comparison
for the case of a two-layer model with zonally symmetric forcing. Journal of the Atmospheric Sciences 54: 2452–2474.
Achatz U, Schmitz G, Greisiger K-M. 1995. Principal interaction patterns in baroclinic wave life cycles. Journal of the Atmospheric
Sciences 52: 3201–3213.
Allen MR, Smith LA. 1996. Monte Carlo SSA: detecting irregular oscillations in the presence of colored noise. Journal of Climate 9:
3373–3404.
Allen MR, Tett SFB. 1999. Checking for model consistency in optimal fingerprinting. Climate Dynamics 15: 419–434.
Ambaum MHP, Hoskins BJ, Stephenson DB. 2001. Arctic oscillation or North Atlantic oscillation? Journal of Climate 14: 3495–3507.
Anderson J, van den Dool H, Barnston A, Chen W, Stern W, Ploshay J. 1999. Present day capabilities of numerical and statistical model
for atmospheric extratropical seasonal simulation and prediction. Bulletin of the American Meteorological Society 80: 1349–1362.
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
676
F. W. ZWIERS AND H. VON STORCH
Arrhenius SA. 1896. On the influence of carbonic acid in the air upon the temperature of the ground. Philosophical Magazine and
Journal of Science 41: 237–276.
Askelson MA, Aubagnac J-P, Straka JM. 2000. An adaptation of the Barnes filter applied to the objective analysis of radar data. Monthly
Weather Review 128: 3050–3082.
Baldwin MP, Dunkerton TJ. 1999. Propagation of the Arctic oscillation from the stratopshere to the troposphere. Journal of Geophysical
Research 104: 30 937–30 946.
Barnett TP, Preisendorfer R. 1987. Origins and levels of monthly and seasonal forecast skill for United States surface air temperature
determined by canonical correlation analysis. Monthly Weather Review 115: 1825–1850.
Barnett TP, Pierce DW, Schnur R. 2001. Detection of anthropogenic change in the world’s oceans. Science 292: 270–274.
Barnston AG. 1994. Linear statistical short-term climate predictive skill in the Northern Hemisphere. Journal of Climate 7: 1513–1564.
Barnston AG, van den Dool H. 1993. A degeneracy in cross-validated skill in regression-based forecasts. Journal of Climate 6: 963–977.
Battisti DS, Hirst AC. 1989. Interannual variability in a tropical atmosphere–ocean model: influences of the basic state, ocean geometry
and nonlinearity. Journal of the Atmospheric Sciences 46: 1687–1712.
Behera SK, Rao SA, Saji HN, Yamagata T. 2003. Comments on ‘A cautionary note on the interpretation of EOFs’. Journal of Climate
16: 1087–1093.
Bengtsson L, Schlese U, Roeckner E, Latif M, Barnett TP, Graham N. 1993. A two-tiered approach to long-range climate forecasting.
Science 261: 1026–1029.
Berliner ML, Royle JA, Wikle CK, Millif RF. 1998. Bayesian methods in atmospheric sciences. Bayesian Statistics 6: 83–100.
Berliner LM, Levine RA, Shea DJ. 2000. Bayesian climate change assessment. Journal of Climate 13: 3805–3820.
Bradley RS. 1999. Paleoclimatology: Reconstructing Climates of the Quaternary. Academic Press: San Diego.
Bretherton CS, Smith C, Wallace JM. 1992. An intercomparison of methods for finding coupled patterns in climate data. Journal of
Climate 5: 541–560.
Briffa KR. 2000. Annual climate variability in the Holocene: interpreting the message of ancient trees. Quaternary Science Reviews 19:
87–105.
Briffa KR, Osborn TJ, Schweingruber FH, Harris IC, Jones PD, Shiyatov SG, Vaganov EA. 2001. Low-frequency temperature variations
from a northern tree ring density network. Journal of Geophysical Research 106: 2929–2941.
Brückner E. 1890. Klimaschwankungen seit 1700 nebst Bemerkungen über die Klimaschwankungen der Diluvialzeit. In Geographische
Abhandlungen, Al-Penck A (ed.). E.D. Hölzel: Wien and Olmütz.
Burgers G. 1999. The El Niño stochastic oscillator. Climate Dynamics 15: 521–531.
Burgers G, van Leeuwen PJ, Evensen G. 1998. On the analysis scheme in the ensemble Kalman filter. Monthly Weather Review 126:
1719–1724.
Busuioc A, von Storch H. 1996. Changes in the winter precipitation in Romania and its relation to the large-scale circulation. Tellus
Series A: Dynamic Meteorology and Oceanography 48: 538–552.
Buzzia R, Palmer TN. 1998. Impact of ensemble size on ensemble prediction. Monthly Weather Review 126: 2503–2518.
Chelton DB. 1994. Physical oceanography: a brief overview for statisticians. Statistical Science 9: 150–166.
Cherry S. 1996. Singular value decomposition analysis and canonical correlation analysis. Journal of Climate 9: 2003–2009.
Christy JR, Parker DE, Brown SJ, Macadam I, Stendel M, Norris WB. 2001. Differential trends in tropical sea surface and atmospheric
temperatures since 1979. Geophysical Research Letters 28: 183–186.
Christy JR, Spencer RW, Braswell WD. 2000. MSU tropospheric temperatures: dataset construction and radiosonde comparison. Journal
of Atmospheric and Oceanic Technology 17: 1153–1170.
Christy JR, Spencer RW, Norris WB, Braswell WD, Parker DE. 2003. Error estimates of Version 5.0 of MSU–AMSU bulk atmospheric
temperatures. Journal of Atmospheric and Oceanic Technology 20: 613–629.
Coelho CA, Pezzulli S, Balmaseda M, Doblas-Reyes FJ, Stephenson DB. 2004. Forecast calibration and combination: a simple Bayesian
approach for ENSO. Journal of Climate in press.
Collins M, Allen MR. 2002. Assessing the relative roles of initial and boundary conditions in interannual to decadal climate predictability.
Journal of Climate 15: 3104–3109.
Cook ER. 1995. Temperature histories in tree rings and corals. Climate Dynamics 11: 211–222.
Cressman GP. 1959. An operational objective analysis system. Monthly Weather Review 87: 367–374.
Crowley TJ, Lowery T. 2000. How warm was the Medieval Warm Period? Ambio 29: 51–54.
Da Costa E, Vautard R. 1997. A qualitative realistic low-order model of the extratropical low-frequency variability built from long
records of potential vorticity. Journal of the Atmospheric Sciences 54: 1064–1084.
Daley R. 1991. Atmospheric Data Analysis. Cambridge University Press: Cambridge, UK.
Derome J, Brunet G, Plante A, Gagnon N, Boer GJ, Zwiers FW, Lambert S, Ritchie H. 2001. Seasonal predictions based on two
dynamical models. Atmosphere–Ocean 39: 485–501.
Deser C. 2000. On the teleconnectivity of the ‘Arctic oscillation’. Geophysical Research Letters 27: 779–782.
Doblas-Reyes FJ, Déqué M, Piedelievre J-P. 2000. Multi-model spread and probabilistic forecasts in PROVOST. Quarterly Journal of
the Royal Meteorological Society 126: 2069–2087.
Dobrovolski SG. 2000. Stochastic Climate Theory and its Application. Springer Verlag: Berlin.
Dommenget D, Latif M. 2002. A cautionary note on the interpretation of EOFs. Journal of Climate 15: 216–225.
Dommenget D, Latif M. 2003. Reply to Comments of Behera et al. Journal of Climate 16: 1094–1097.
Donnell CA. 1912. The effect of the time of observation on mean temperatures. Monthly Weather Review 40: 708.
Efron B, Tibshirani R. 1993. An Introduction to the Bootstrap. Chapman and Hall: New York.
Ellis W. 1890. On the difference produced in the mean temperature derived from daily maxima and minima as dependent on the time
at which the thermometers are read. Quarterly Journal of the Royal Meteorological Society 16: 213–218.
Esper J, Cook ER, Schweingruber RH. 2002. Low-frequency signals in long tree-ring chronologies for reconstructing past temperature
variability. Science 295: 2250–2253.
Evensen G. 1992. Using the extended Kalman filter with a multilayer quasi-geostrophic ocean-model. Journal of Geophysical Research
97: 17 905–17 924.
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
STATISTICS IN CLIMATE RESEARCH
677
Evensen G. 1994. Sequential data assimilation with a nonlinear quasi-geostrophic model using Monte Carlo methods to forecast error
statistics. Journal of Geophysical Research 99: 10 143–10 162.
Farge M, Goirand E, Philipovitch T. 1993. Wavelets and wavelet packets to analyse, filter and compress two-dimensional turbulent
flows. In Statistical Methods in Physical Oceanography, Müller P, Henderson D (eds). Proceedings of the ‘Aha Huliko’a Hawaiian
Winter Workshop. University of Hawaii at Manoa, 12–15 January.
Fennesy MJ, Shukla J. 1999. Impact of initial soil wetness on seasonal atmospheric prediction. Journal of Climate 12: 3167–3180.
Folland CK, Karl TR, Christy JR, Clarke RA, Gruza GV, Jouzel J, Mann ME, Oerlemans J, Salinger MJ, Wang S-W. 2001. Observed
climate variability and change. In Climate Change 2001: The Scientific Basis. Contribution of Working Group I to the Third Assessment
Report of the Intergovernmental Panel on Climate Change, Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ, Dai X,
Maskell K, Johnston CA (eds). Cambridge University Press: Cambridge, UK, and New York, USA.
Frankignoul C. 1985. Sea surface temperature anomalies, planetary waves, and air–sea feedback in the middle latitudes. Reviews in
Geophysics 23: 357–390.
Fritts HC. 1976. Tree Rings and Climate. Academic Press: London.
Fritts HC. 1991. Reconstructing Large-scale Climatic Patterns from Tree Ring Data. The University of Arizona Press: Tucson.
Gallo KP, Owen TW. 1999. Satellite-based adjustments for the urban heat island temperature bias. Journal of Applied Meteorology 38:
806–813.
Gandin L. 1963. Objective analysis of meteorological fields. Gridromet: Leningrad. English translation, 1965. Israel Program for Scientific
Translation: Jerusalem.
Gandin LS, Murphy AH. 1992. Equitable skill scores for categorical forecasts. Monthly Weather Review 120: 361–370.
Gibson JK, Kalberg P, Uppala S, Hernandes A, Nomura A, Serrano E. 1997. ECMWF reanalysis report series 1 — ERA description.
ECMWF, Reading, UK.
Given T, Ray PS. 1994. Response of a two-dimensional dual-Doppler radar wind synthesis. Journal of Atmospheric and Oceanic
Technology 11: 239–255.
Glahn HR, Lowry DA. 1972. The use of model output statistics (MOS) in objective weather forecasting. Journal of Applied Meteorology
11: 1203–1211.
Goodman JC, Marshall J. 2003. The role of neutral singular vectors in middle-latitude air–sea coupling. Journal of Climate 16: 88–102.
Gregory JM, Stouffer RJ, Raper SCB, Stott PA, Rayner NA. 2002. An observationally based estimate of the climate sensitivity. Journal
of Climate 15: 3117–3121.
Hasselmann KF. 1976. Stochastic climate models. Part I. Theory. Tellus 28: 473–485.
Hasselmann KF. 1979. On the signal-to-noise problem in atmospheric response studies. In Meteorology of the Tropical Ocean, Shaw DB
(ed.). Royal Meteorological Society: London; 251–259.
Hasselmann KF. 1988. PIPs and POPs: the reduction of complex dynamical systems using principal interaction and principal oscillation
patterns. Journal of Geophysical Research 93: 11 015–11 021.
Hasselmann K. 1993. Optimal fingerprints for the detection of time dependent climate change. Journal of Climate 6: 1957–1971.
Hasselmann KF. 1997. Multi-pattern fingerprint method for detection and attribution of climate change. Climate Dynamics 13: 601–612.
Hasselmann K. 1998. Conventional and Bayesian approach to climate change detection and attribution. Quarterly Journal of the Royal
Meteorological Society 124: 2541–2565.
Hayashi Y. 1982. Space–time spectral analysis and its applications to atmospheric waves. Journal of the Meteorological Society of
Japan 60: 156–171.
Hegerl GC, Allen MR. 2002. Origins of model–data discrepancies in optimal fingerprinting. Journal of Climate 15: 1348–1356.
Hegerl GC, North GR. 1997. Statistically optimal methods for detecting anthropogenic climate change. Journal of Climate 10:
1125–1133.
Hegerl GC, Crowley TJ, Baum SK, Kim K-W, Hyde WT. 2003. Detection of volcanic, solar and greenhouse gas signals in paleoreconstructions of Northern-Hemispere temperature. Geophysical Research Letters 30(5): 1242. DOI: 10.1029/2002GL016635.
Hoerling MP, Kumar A. 2002. Atmospheric response patterns associated with tropical forcing. Journal of Climate 15: 2184–2203.
Hoerling MP, Kumar A. 2003. The perfect ocean for drought. Science 299: 691–694.
Honerkamp J. 1994. Stochastic Dynamical Systems: Concepts, Numerical Methods, Data. VCH Publishers.
Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ, Dai X, Maskell K, Johnston CA (eds). 2001. Climate Change 2001:
The Scientific Basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on Climate
Change. Cambridge University Press: Cambridge, UK, and New York, USA.
Houtekamer PL, Mitchell HL. 2001. A sequential ensemble Kalman filter for atmospheric data assimilation. Monthly Weather Review
129: 123–137.
Hsieh WH, Tang B. 1998. Applying neural network models to prediction and analysis in meteorology and oceanography. Bulletin of
the American Meteorological Society 79: 1855–1870.
Hwang S-O, Schemm J-KE, Barnston AG, Kwon W-T. 2001. Long-lead seasonal forecast skill in far eastern Asia using canonical
correlation analysis. Journal of Climate 14: 3005–3016.
IPCC. 2002. IPCC Workshop on changes in extreme weather and climate events, 11–13 June, Beijing, China. http://www.ipcc.ch/pub/
extremes.pdf. [Accessed January 2004].
Jolliffe IT. 2002. Principal Component Analysis, 2nd edition. Springer-Verlag.
Jolliffe I. 2003. A cautionary note on artificial examples of EOFs. Journal of Climate 16: 1084–1086.
Jolliffe I, Stephenson DB (eds). 2003. Forecast Verification: A Practitioner’s Guide in Atmospheric Science. Wiley.
Jones PD, Moberg A. 2003. Hemispheric and large-scale surface air temperature variations: an extensive review and an update to 2001.
Journal of Climate 16: 206–213.
Jones PD, Groisman PY, Coughlan M, Plummer N, Wang W-C, Karl TR. 1990. Assessment of urbanization effects in time series of
surface air temperatures over land. Nature 347: 169–172.
Jones PD, Osborn TJ, Briffa KR. 2001a. The evolution of climate over the last millennium. Science 292: 662–667.
Jones PD, Osborn TJ, Briffa KR, Folland CK, Horton EB, Alexander LV, Parker DE, Rayner NA. 2001b. Adjusting for sampling
density in grid box land and ocean surface temperature time series. Journal of Geophysical Research 106: 3371–3380.
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
678
F. W. ZWIERS AND H. VON STORCH
Kalnay E, Kanamitsu M, Kistler R, Collins W, Deaven D, Gandin L, Iredell M, Saha S, White G, Woollen J, Zhu Y, Chelliah M,
Ebisuzaki W, Higgins W, Janowiak J, Mo KC, Ropelewski C, Wang J, Leetmaa A, Reynolds R, Jenne R, Joseph D. 1996. The
NCEP/NCAR 40-year reanalysis project. Bulletin of the American Meteorological Society 77: 437–472.
Karl TR, Williams Jr CN, Young PJ. 1986. A model to estimate the time of observation bias associated with monthly mean maximum,
minimum and mean temperatures for the United States. Journal of Climate and Applied Meteorology 23: 1489–1504.
Karl TR, Quayle RG, Groisman PY. 1993. Detecting climate variations and change: new challenges for observing and data management
systems. Journal of Climate 6: 1481–1494.
Katz RW, Parlange MB, Naveau P. 2002. Statistics of extremes in hydrology. Advances in Water Resources 25: 1287–1304.
Kharin VV, Zwiers FW. 2002. Climate predictions with multi-model ensembles. Journal of Climate 15: 793–799.
Kharin VV, Zwiers FW. 2003a. Improved seasonal probability forecasts. Journal of Climate 16: 1684–1701.
Kharin VV, Zwiers FW. 2003b. On the ROC score of probability forecasts. Journal of Climate 16: 4145–4150.
Kharin VV, Zwiers FW. 2004. Estimating extremes in transient climate change simulations. Journal of Climate submitted for publication.
Kharin VV, Zwiers FW, Gagnon N. 2001. Skill of seasonal hindcasts as a function of the ensemble size. Climate Dynamics 17: 835–843.
Klein WH, Glahn HR. 1974. Forcasting local weather by means of model output statistics. Bulletin of the American Meteorological
Society 55: 1217–1227.
Klein WH, Lewis BM, Enger I. 1959. Objective predictions of five-day mean temperature during winter. Journal of Meteorology 16:
672–682.
Köppen W. 1923. Die Klimate der Erde. Wlater de Gruyter: Berlin.
Krishnamurti TN, Kishtawal CM, LaRow TE, Bachiochi DR, Zhang Z, Willifor CE, Gadgil S, Surendran S. 1999. Improved weather
and seasonal climate forecasts from multimodel superensemble. Science 285: 1548–1550.
Krishnamurti TN, Kishtawal CM, Zhang Z, LaRow TE, Bachiochi DR, Willifor CE, Gadgil S, Surendran S. 2000. Multi-model
ensemble forecasts for weather and seasonal climate. Journal of Climate 13: 4196–4216.
Kwasniok F. 1996. The reduction of complex dynamical systems using principal interaction patterns. Physica D 92: 28–60.
Latif M, Barnett TP. 1996. Decadal climate variability over the North Pacific and North Atlantic: dynamics and predictability. Journal
of Climate 9: 2407–2423.
Le Dimet FX, Talagrand O. 1986. Variational algorithms for analysis and assimilation of meteorological observations: theoretical aspects.
Tellus Series A: Dynamic Meteorology and Oceanography 38: 97–110.
Levine RA, Berliner LM. 1999. Statistical principles for climate change studies. Journal of Climate 12: 564–574.
Livezey RE. 1999. The evaluation of forecasts. In Analysis of Climate Variability: Applications of Statistical Techniques, 2nd edition,
von Storch H, Navarra A (eds). Springer Verlag: 177–196.
Livezey RE, Masutani M, Leetma A, Rui H, Ji M, Kumar A. 1997. Teleconnective response of the Pacific–North American regions
atmosphere to large central equatorial Pacific SST anomalies. Journal of Climate 10: 1787–1820.
Lorenz EN. 1982. Atmospheric predictability with a large numerical model. Tellus 34: 505–513.
MacCracken MC, Barron EJ, Easterling DR, Felzer BS, Karl TR. 2003. Climate change scenarios for the U.S. National Assessment.
Bulletin of the American Meteorological Society 84: 1711–1723.
Madden RA. 1976. Estimates of the natural variability of time-averaged sea-level pressure. Monthly Weather Review 104: 942–952.
Madden RA, Julian PR. 1971. Detection of a 40–50 day oscillation in the zonal wind in the tropical Pacific. Journal of the Atmospheric
Sciences 28: 702–708.
Mann M, Bradley RS, Hughes MK. 1999. Northern Hemisphere temperatures during the past millenium: inferences, uncertainties, and
limitations. Geophysical Research Letters 26: 759–762.
Mason SJ, Graham NE. 1999. Conditional probabilities, relative operating characteristics, and relative operating levels. Weather
Forecasting 14: 713–725.
Mears CA, Schabel MC, Wentz FJ. 2003. A reanalysis of the MSU channel 2 tropospheric temperature record. Journal of Climate 16:
3650–3664.
Mechoso CR, Neelin JD, Yu J-Y. 2003. Testing simple models of ENSO. Journal of the Atmospheric Sciences 60: 305–318.
Michaelson J. 1987. Cross-validation in statistical climate forecast models. Journal of Climate and Applied Meteorology 26: 1589–1600.
Mimmack GM, Mason SJ, Galpin JS. 2001. Distance metrics in hierarchical cluster analysis: defining regions. Journal of Climate 14:
2790–2797.
Mitchell JFB, Karoly DJ, Hegerl GC, Zwiers FW, Marengo J. 2001. Detection of climate change and attribution of causes. In Climate
Change 2001. The Scientific Basis. Contribution of Working Group I to the Third Assessment Report of the Intergovernmental Panel on
Climate Change, Houghton JT, Ding Y, Griggs DJ, Noguer M, van der Linden PJ, Dai X, Maskell K, Johnston CA (eds). Cambridge
University Press: Cambridge, UK, and New York, USA.
Monahan AH, Fyfe J, Flato G. 2000. A regime view of Northern Hemisphere atmospheric variability and change under global warming.
Geophysical Research Letters 27: 1139–1142.
Monahan AH, Pandolfo L, Fyfe JC. 2001. The preferred structure of variability of the Northern Hemisphere atmospheric circulation.
Geophysical Research Letters 28: 1019–1022.
Monahan AH, Fyfe JC, Pandolfo L. 2003. The vertical structure of wintertime climate regimes of the Northern Hemisphere extratropical
atmosphere. Journal of Climate 16: 2005–2021.
Müller P, Henderson D (eds). 1993. Statistical Methods in Physical Oceanography. Proceedings ‘Aha Huliko’a Hawaiian Winter
Workshop. University of Hawaii at Manoa, 12–15 January.
Müller P, Henderson D (eds). 1997. Monte Carlo Simulations in Oceanography. Proceedings ‘Aha Huliko’a Hawaiian Winter Workshop.
University of Hawaii at Manoa, 14–17 January.
Murphy AH, Epstein ES. 1989. Skill scores and correlation coefficients in model verification. Monthly Weather Review 117: 572–581.
Murphy AH, Wilks DS. 1998. A case study of the use of statistical models in forecast verification: precipitation probability forecasts.
Weather Forecasting 13: 795–810.
Murphy AH, Zwiers FW. 1993. International Meetings on Statistical Climatology. Bulletin of the American Meteorological Society 9:
1721–1727.
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
STATISTICS IN CLIMATE RESEARCH
679
NAST (National Assessment Synthesis Team). 2001. Climate Change Impacts on the United States: The Potential Consequences of
Climate Variability and Change. Cambridge University Press: Cambridge, UK.
Newman M, Sardeshmukh PD. 1995. A caveat concerning singular value decomposition. Journal of Climate 8: 352–360.
Nitsche G, Wallace JM, Kooperberg C. 1994. Is there evidence of multiple equilibria in the planetary-wave amplitude? Journal of the
Atmospheric Sciences 51: 314–322.
North GR, Wu Q. 2001. Detecting climate signals using space time EOFs. Journal of Climate 14: 1839–1863.
North R, Bell TL, Cahalan RF, Moeng FJ. 1982. Sampling errors in the estimation of empirical orthogonal functions. Monthly Weather
Review 10: 699–706.
NRC. 2000. Reconciling Observations of Global Temperature Change. National Research Council: Washington, DC.
Panel on Statistics and Oceanography. 1994. Report on statistics and physical oceanography. Statistical Science 9: 167–221.
Petersen AC. 1999. Philosophy of climate science. Bulletin of the American Meteorological Society 81: 265–271.
Peterson TC, Vose RS. 1997. An overview of the Global Historical Climatology Network temperature data base. Bulletin of the American
Meteorological Society 78: 2837–2849.
Peterson TC, Easterling DR, Karl TR, Groisman P, Nicholls N, Plummer N, Torok S, Auer I, Boehm R, Gullett D, Vincent L,
Heino R, Tuomenvirta H, Mestre O, Szentimrey T, Saliner J, Førland E, Hanssen-Bauer I, Alexandersson H, Jones P, Parker D. 1998.
Homogeneity adjustments of in situ atmospheric climate data: a review. International Journal of Climatology 18: 1493–1517.
Peterson TC, Gallo KP, Livermore J, Owen TW, Huang A, McKittrick DA. 1999. Global rural temperature trends. Geophysical
Research Letters 26: 329–332.
Pfizenmayer A, von Storch H. 2001. Anthropogenic climate change shown by local wave conditions in the North Sea. Climate Research
19: 15–23.
Rabier F, Thépaut J-N, Courtier P. 1998. Extended assimilation and forecast experiments with a four-dimensional variational assimilation
system. Quarterly Journal of the Royal Meteorological Society 124: 1–39.
Reichert BK, Schnur R, Bengtsson L. 2002. Global ocean warming tied to anthropogenic forcing. Geophysical Research Letters 29(11):
20-1, 20-4. DOI: 10.1029/2001GL013954.
Richman MB. 1986. Rotation of principal components. International Journal of Climatology 6: 293–335.
Ripley BD, Thompson M. 1987. Regression techniques for the detection of analytical bias. Analyst 112: 377–383.
Risbey JS, Kandlikar M, Karoly DJ. 2000. A protocol to articulate and quantify uncertainties in climate change detection and attribution.
Climate Research 16: 61–78.
Robinson AR, Lermusiaux PFJ, Sloan III NQ. 1998. Data assimilation. In The Global Coastal Ocean. Processes and Methods, Brink KH,
Robinson AR (eds). The Sea, vol. 10. John Wiley: New York; 541–593.
Ross GH. 1987. An updateable model output statistics scheme. Programme on short- and medium range weather prediction, PSMP
Report Series, No. 25. World Meteorological Organization; 25–28.
Rowell DP. 1998. Assessing potential seasonal predictability with an ensemble of multi-decadal GCM simulations. Journal of Climate
11: 109–120.
Rowell DP, Zwiers FW. 1999. The global distribution of the sources of decadal variability and mechanisms over the tropical Pacific
and southern North America. Climate Dynamics 15: 751–772.
Saji NH, Goswami BN, Vinayachandran PN, Yamagata T. 1999. A dipole mode in the tropical Indian Ocean. Nature 401: 360–363.
Santer BD, Wigley TML, Barnett TP, Anyamba E. 1996. In Climate Change 1995. The Science of Climate Change, Houghton JT, Meira
Filho LG, Callandar BA, Harris N, Kattenberg A, Maskell K (eds). Cambridge University Press: Cambridge; 407–444.
Santer BD, Wigley TML, Boyle JS, Gaffen DJ, Hnilo JJ, Nychka D, Parker DE, Taylor KE. 2000. Statistical significance of trend
differences in layer-average temperature time series. Journal of Geophysical Research 105: 7337–7356.
Santer BD, Wigley TML, Meehl GA, Wehner MF, Mears C, Schabel M, Wentz FJ, Ammann C, Arblaster J, Bettge T, Washington WM,
Taylor KE, Boyle JS, Brggemann W, Doutriaux C. 2003. Influence of satellite data uncertainties on the detection of externally-forced
climate change. Science 300: 1280–1284.
Saravanan R, Danabasoglu G, Doney SC, McWilliams JC. 2000. Decadal variability and predictability in the midlatitude
ocean–atmosphere system. Journal of Climate 13: 1073–1097.
Sarewitz D, Pielke Jr R. 1999. Prediction in science and policy. Technology in Science 21: 121–133.
Scheffé H. 1959. The Analysis of Variance. Wiley: New York.
Schiller H, Doerffer R. 1999. Neural network for emulation of an inverse model — operational derivation of case II water properties
from MERIS data. International Journal of Remote Sensing 20: 1735–1746.
Schnur R, Hasselmann K. 2004. Optimal filtering for Bayesian detection of climate change. Climate Dynamics submitted for publication.
Selten FN. 1995. An efficient description of the dynamics of barotropic flow. Journal of the Atmospheric Sciences 52: 915–936.
Shabbar A, Barnston AG. 1996. Skill of seasonal climate forecasts in Canada using canonical correlation analysis. Monthly Weather
Review 124: 2370–2385.
Shukla J. 1998. Predictability in the midst of chaos: a scientific basis for climate forecasting. Science 282: 728–731.
Sokolov S, Rintoul SR. 1999. Some remarks on interpolation of nonstationary oceanographic fields. Journal of Atmospheric and Oceanic
Technology 16: 1434–1449.
Speth P, Madden RA. 1983. Space–time spectral analyses of Northern Hemisphere geopotential heights. Journal of the Atmospheric
Sciences 40: 1086–1100.
Stanski HR, Wilson LJ, Burrows WR. 1989. Survey of common verification methods in meteorology. WMO WWW Technical Report
8, WMO TD 358.
Suarez MJ, Schopf PS. 1988. A delayed action oscillator for ENSO. Journal of the Atmospheric Sciences 45: 3283–3287.
Tang B. 1995. Periods of linear development of the ENSO cycle and POP forecast experiments. Journal of Climate 8: 682–691.
Tangang FT, Tang B, Monahan AH, Hsieh WW. 1998. Forecasting ENSO events: a neural network-extended EOF approach. Journal
of Climate 11: 29–41.
Tanguay M, Bartello P, Gauthier P. 1995. Four-dimensional data assimilation with a wide range of scales. Tellus Series A: Dynamic
Meteorology and Oceanography 47: 974–997.
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)
680
F. W. ZWIERS AND H. VON STORCH
Thépaut J-N, Coultier P. 1991. Four-dimensional data assimilation using the adjoint of a multilevel primitive equation model. Quarterly
Journal of the Royal Meteorological Society 117: 1225–1254.
Thiébaux HJ. 1976. Anisotropic correlation functions for objective analysis. Monthly Weather Review 104: 994–1002.
Thiébaux HJ, Pedder MA. 1987. Spatial Objective Analysis: With Applications in Atmopsheric Science. Academic Press: London.
Thompson DWJ, Wallace JM. 1998. The Arctic oscillation signature in the wintertime geopotential height and temperature fields.
Geophysical Research Letters 25: 1297–1300.
Thompson DWJ, Wallace JM. 2001. Regional climate impacts of the Northern Hemisphere annular mode. Science 293: 85–89.
Timmermann A, Lohmann G. 2000. Noise-induced transitions in a simplified model of the thermohaline circulation. Journal of Physical
Oceanography 30: 1891–1900.
Trapp RJ, Doswell CA. 2000. Radar data objective analysis. Journal of Atmospheric and Oceanic Technology 17: 105–120.
Van den Dool HM, Saha S, Johansson A. 2000. Empirical orthogonal teleconnections. Journal of Climate 13: 1421–1435.
Vautard R. 1995. Patterns in time: SSA and MSSA. In Analysis of Climate Variability: Applications of Statistical Techniques, von
Storch H, Navarra A (eds). Springer Verlag: 259–280.
Vautard R, Yiou P, Ghil M. 1992. Singular spectrum analysis: a toolkit for short, noisy chaotic signals. Physica D 58: 95–126.
Vincent LA, Gullett DW. 1999. Canadian historical and homogeneous temperature datasets for climate change analyses. International
Journal of Climatology 19: 1375–1388.
Von Storch H. 2001a. Statistics — an indispensable tool in dynamical modelling. In Models in Environmental Research, von Storch H,
Flöser G (eds). Springer Verlag.
Von Storch H. 2001b. Models between academia and applications. In Models in Environmental Research, von Storch H, Flöser G (eds).
Springer Verlag.
Von Storch H, Navarra A (eds). 1999. Analysis of Climate Variability: Applications of Statistical Techniques, 2nd updated extended
edition. Springer Verlag.
Von Storch H, Xu JS. 1990. Principal oscillation pattern analysis of the tropical 30- to 60-day oscillation: part I: definition of an index
and its prediction. Climate Dynamics 4: 175–190.
Von Storch H, Zwiers FW. 1999. Statistical Analysis in Climate Research. Cambridge University Press.
Von Storch H, von Storch J-S, Müller P. 2001. Noise in the climate system — ubiquitous, constitutive and concealing. In Mathematics
Unlimited — 2001 and Beyond. Part II, Engquist B, Schmid W (eds). Springer Verlag: 1179–1194.
Von Storch JS. 2000. Signatures of air–sea interactions in a coupled atmosphere–ocean GCM. Journal of Climate 13: 3361–3379.
Walker GT, Bliss EW. 1932. World weather V. Memoirs of the Royal Meteorological Society 4(36): 53–84.
Wallace JM. 2000. North Atlantic oscillation/annual mode: two paradigms–one phenomenon. Quarterly Journal of the Royal
Meteorological Society 126: 791–805.
Wallace JM, Dickinson RE. 1972. Empirical orthogonal representation of time series in the frequency domain. Part I: theoretical
considerations. Journal of Applied Meteorology 11: 887–892.
Wallace JM, Gutzler DS. 1981. Teleconnections in the geopotential height field during the Northern Hemisphere winter. Monthly Weather
Review 109: 784–812.
Wang XL, Zwiers FW. 2001. Using redundancy analysis to improve dynamical seasonal mean 500 hPa geopotential forecasts.
International Journal of Climatology 21: 637–654.
WASA. 1998. Changing waves and storms in the Northeast Atlantic? Bulletin of the American Meteorological Society 79: 741–760.
Weare BC, Nasstrom JN. 1982. Examples of extended empirical orthogonal function analyses. Monthly Weather Review 110: 481–485.
Widman M, Bretherton CS, Salathé EP. 2003. Statistical precipitation downscaling over the northwestern United States using numerically
simulated precipitation as a predictor. Journal of Climate 16: 799–816.
Wilby RL, Wigley TML, Conway D, Jones PD, Hewitson BC, Main J, Wilks DS. 1998. Statistical downscaling of general circulation
model output: a comparison of methods. Water Resources Research 34: 2995–3008.
Wilks DS. 1995. Statistical Methods in the Atmospheric Sciences: An Introduction. International Geophysics Series, Vol. 59. Academic
Press.
Wilks DS. 2000. Diagnostic verification of Climate Prediction Center long-lead outlooks, 1995–8. Journal of Climate 13: 2389–2403.
Wilks DS. 2001. A skill score based on economic value for probability forecasts. Meteorological Applications 8: 209–219.
Wilson LJ, Vallée M. 2002. The Canadian Updateable Model Output Statistics (UMOS) system: design and development test. Weather
Forecasting 17: 206–222.
Yun WT, Stephanova L, Krishnamurti TN. 2003. Improvement of the multi-model superensemble technique for seasonal forecasts.
Journal of Climate 16: 3834–3840.
Yuval, Hsieh WW. 2003. An adaptive nonlinear MOS scheme for precipitation forecasts using neural networks. Weather Forecasting
18: 303–310.
Zorita E, Kharin V, von Storch H. 1992. The atmospheric circulation and sea surface temperature in the North Atlantic in winter. Their
interaction and relevance for Iberian rainfall. Journal of Climate 5: 1097–1108.
Zwiers FW. 1996. Interannual variability and predictability in an ensemble of AMIP climate simulations conducted with the CCC
GCM2. Climate Dynamics 12: 825–848.
Zwiers FW. 1999. The detection of climate change. In Anthropogenic Climate Change, von Storch H, Flöser G (eds). Springer: Berlin;
161–206.
Zwiers FW, Zhang X. 2003. Toward regional scale climate change detection. Journal of Climate 16: 793–797.
Zwiers FW, Wang XL, Sheng J. 2000. The effects of specifying bottom boundary conditions in an ensemble of GCM simulations.
Journal of Geophysical Research 105: 7295–7315.
Copyright 2004 Environment Canada. Published by John Wiley & Sons, Ltd.
Int. J. Climatol. 24: 665–680 (2004)