ArticlePDF Available

A Statistical Downscaling Model for Southern Australia Winter Rainfall

Authors:
  • Western Power

Abstract and Figures

A technique for obtaining downscaled rainfall projections from climate model simulations is described. This technique makes use of the close association between mean sea level pressure (MSLP) patterns and rainfall over southern Australia during winter. Principal components of seasonal mean MSLP anomalies are linked to observed rainfall anomalies at regional, gridpoint, and point scales. A maximum of four components is sufficient to capture a relatively large fraction of the observed variance in rainfall at most locations. These are used to interpret the MSLP patterns from a single climate model, which has been used to simulate both present-day and future climate. The resulting downscaled values provide 1) a closer representation of the observed present-day rainfall than the raw climate model values and 2) alternative estimates of future changes to rainfall that arise owing to changes in mean MSLP. While decreases are simulated for later this century (under a single emissions scenario), the downscaled values, in percentage terms, tend to be less.
Content may be subject to copyright.
A Statistical Downscaling Model for Southern Austr alia Winter Rainfall
YUN LI
CSIRO Mathematical and Information Sciences, CSIRO Climate Adaptation Flagship, Wembley, Western Australia, Australia
IAN SMITH
CSIRO Marine and Atmospheric Research, CSIRO Water for a Healthy Country Flagship, Aspendale, Victoria, Australia
(Manuscript received 24 July 2007, in final form 29 July 2008)
ABSTRACT
A technique for obtaining downscaled rainfal l projections from climate model simulations is described.
This technique makes use of the close assoc iation between mean sea level pressure (M SLP) patterns and
rainfall over southern Australia during winter. Principal compo nents of seasonal mean MSLP anomalies
are linked to observed rainfall anomalies at regional, gridpoint, and point scales. A maximum of four
components is sufficient to capture a relatively large fraction of the observed variance in rainfall at most
locations. These are used to interpret the MSLP patterns from a single climate model, which has been used
to simulate both present-day and future climate. The resulting downscaled values provide 1) a closer
repres entation of the observed present-day rainfall than the raw climate model values and 2) alternative
estimates of fut ure changes to rainfall that arise owing to changes in mean MSLP. While decreases are
simulated for later this century (under a single emiss ions scenario), the downsc aled values, in p ercent age
terms, tend to be less.
1. Introduction
There is already evidence that the global pattern of
annual streamflow trends is partly due to the effects of
anthropogenic climate change, mainly because of a
poleward expansion of the subtropical dry zone (Lu
et al. 2007). However, uncertainties in projected rainfall
changes for later this century plague estimates of im-
pacts on future runoff and water storages (Milly et al.
2008). In particular, there are several difficulties asso-
ciated with interpreting changes in variables simulated
at a resolution of 100–200 km in terms of changes to be
expected at smaller catchment scales. This affects the
ability to simulate, for example, expected changes in
surface runoff and streamflows into water catchments.
There is an increasing demand for more reliable esti -
mates of these changes by water resource managers
who need to make long-term decisions about future
infrastructure demands (e.g., new reservoirs, pipelines,
drainage, etc.). Southwest Western Australia has al-
ready been severely affected by a downturn in rainfall
and even more serious reductions in runoff (Bates et al.
2008), and there is a growing concern that this pattern of
reduced rainfall may be extending to the eastern states.
Southern Australia (the region southward of 308S)
receives the bulk of its annual rainfall during the winter
half of the year (i.e., May–October). Rainfall during the
austral winter months from June to August (JJA) is
largely controlled by large-scale atmospheric circulation
patterns, as opposed to convectively dominated sys-
tems that dominate regions to the north and in summer
(Smith et al. 2008). The winter patterns include incur-
sions of cold air outbreaks from higher latitudes asso-
ciated with depressions originating in the Indian Ocean
or Southern Ocean. On the other hand, the persistence
of anticyclones is associated with relatively dry condi-
tions. Variability of rainfall is strongly linked to the
relative strength and location of these patterns which
affect wind strength and direction (Sturman and Tapper
1996). A number of studies (e.g., Smith et al. 2000) have
documented the strong inverse relationship between
mean sea level pressure (MSLP) and southwest Western
Australia (SWWA) regional winter rainfall on monthly
Corresponding author address: Dr. Yun Li, CSIRO Mathemat-
ical and Information Sciences, Wembley, Western Australia 6913,
Australia.
E-mail: yun.li@csiro.au
1142 JOURNAL OF CLIMATE VOLUME 22
DOI: 10.1175/2008JCLI2160.1
Ó 2009 American Meteorological Society
and seasonal time scales. Allan and Haylock (1993)
suggested that southern Australian winter rainfall is mod-
ulated by a long-term MSLP signal with a pronounced
trend in recent decades. This large-scale pressure pat-
tern is referred to as the southern annular mode (SAM)
and influences rainfall across southern parts of the con-
tinent in some seasons (Meneghini et al. 2006; Hendon
et al. 2007). Li et al. (2005) modeled winter extreme
rainfall over SWWA and its associated changes with the
SAM. They showed that the upward trend of the SAM
is consistent with the decrease in extreme winter rainfall
observed over SWWA that has been apparent since the
mid-1960s.
Climate models sometimes have difficulty in realisti-
cally simulating rainfall at regional and smaller scales
(Hewitson and Crane 2006), which can affect the con-
fidence that can be placed in their attempts to simulate
future changes. This arises because of a number of fac-
tors, including the fact that different models use different
parameterization schemes for various physical processes
and that the results not only reflect the effects of forced
changes but also internal variability resulting from the
chaotic nature of the system (Ra
¨
isa
¨
nen 2007). In par-
ticular, ensemble rainfall projections for Australia are
characterized by relatively large uncertainties (Whetton
et al. 2005). For example, results from increased CO
2
experiments (Cai and Watterson 2002; Hope 2006)
show that SWWA becomes drier as midlatitude MSLP
increases. Although these results are useful, it also needs
to be borne in mind that the degree of drying suggested
by these GCM results is accompanied by considerable
uncertainty—particularly as the simulated amounts for
present-day conditions tend to underestimate the ob-
served amounts. For example, of seven climate models
analyzed by Hope (2006), the percentage errors asso-
ciated with estimates for present-day (1961–2000) mean
June and July rainfall for SWWA range from 280% to
0%. These same models yield percentage changes for
later this century (2081–2100) that range from 23% to
225%, the average reduction being 213%.
Various approaches to overcoming the uncertainties
accompanying future climate change projections are
being developed, including the assessment of the per-
formance of individual models as a guide to the reliability
of their predicted changes (Maxino et al. 2008; Perkins
et al. 2007; Suppiah et al. 2007; Whetton et al. 2007;
Perkins and Pitman 2008; Smith and Chandler 2008,
manuscript submitted to Climatic Change). Downscaling
is another method that can potentially assist in the as-
sessment of climate models. A simple test for a model is
that it not only provide an accurate estimate for present-
day regional rainfall, but that it should also simulate the
observed relationship between regional rainfall and
other key variables (e.g., MSLP). If these criteria can be
satisfied, then it is arguable that any simulated changes
in rainfall are more likely to be reliable than otherwise.
Downscaling cannot only provide an indication of any
such relationships, it can also potentially provide al-
ternative estimates for rainfall changes if the model-
simulated changes in the key variables are believed to
be more reliable than the rainfall estimates themselves
(e.g., Benestad 2001).
There are two main approaches to downscaling: dy-
namical (Christensen et al. 1998; Murphy 1999; Schmidli
et al. 2006) and statistical (Zorita and von Storch 1997,
1999; von Storch et al. 2000). The former involves
nesting regional models with relatively high horizontal
resolution with a coarser-resolution global climate model.
The latter approach is based on historical (empirical)
studies of the relationship between the l arge-scale cli-
mate anomalies and local climate fluctuations. There
are numerous ways t o develop statistical downscaling
models (Zorita and von Storch 1997; Fowler et al. 2007),
but it is important to note that a statistical downscaling
approach to climate change results implicitly assumes
that any derived historical relationships also hold for the
future (Wilby 1997).
A number of studies have involved statistical down-
scaling as a tool in seasonal rainfall predictions for
Australia (Hughes et al. 1999; Charles et al. 2004).
Hughes et al. (1999) developed a nonhomogeneous
hidden Markov model (NHMM) for SWWA using three
atmospheric predictors: the mean large-scale value for
MSLP, the north–south MSLP gradient, and east–west
gradient in geopotential height at 850 hPa. They found
that the model could provide credible reproductions of
at-site precipitation and their spatial association and
dry- and wet-spell length statistics for a range of site
locations. In a subsequent study, Charles et al. (2004)
also identified dewpoint temperature depression at 850
hPa (RH850) as an important factor and demonstrated
some predictability of rainfall occurrence probabilities
but less so with actual rainfall amounts. Another tech-
nique, referred to as analog-based downscaling, has been
used in an attribution study of the rainfall decline over
SWWA (Timbal et al. 2006) and in interpreting climate
change results for Australia (Timbal and McAvaney
2001; Timbal 2004; Timbal and Jones 2008). Hope et al.
(2006) used a self-organizing map (SOM) to identify
the typical synoptic patterns that govern daily rain-
fall events over SWWA and showed how a ch ange
in the frequency of rainfall-bearing events could ex-
plain the observed decline in winter rainfall since the
mid-1970s.
Rather than referring to just north–south or east–west
gradients of MSLP independently for each region or
1MARCH 2009 L I A N D S M I T H 1143
point of interest (as is typically done in the models de-
scribed above), or dealing with daily synoptic patterns, we
are more interested in the seasonal time scale and diag-
nosing a limited number of large-scale MSLP patterns
that may be relevant to rainfall over a range of regions
across southern Australia. This is done by deriving em-
pirical orthogonal functions (EOFs) and their associated
principal components (PCs) of observed MSLP fields
over the wider Australian region. These are then linked to
observed winter rainfall totals (from various regions) via
multiple linear regression, referred to as principal com-
ponent regression (Draper and Smith 1981; Jolliffe 2002).
The resultant predictor equations (or statistical models)
can then be applied to any subsequent MSLP field (either
observed or predicted) to provide an estimate for rainfall.
This can be particularly convenient where climate models
provide simulations for present-day and future MSLP
fields. An example of such an application is presented
using the results from a single model, the Commonwealth
Scientific and Industrial Research Organisation Mark
version 3.5 (CSIRO Mk3.5), set of results from an A2
emissions scenario climate change experiment and the
implications for model assessment discussed.
2. Data
Mean sea level pressure data were extracted from
the National Centers for Environmental Prediction–
National Center for Atmospheric Research (NCEP–
NCAR) reanalysis globally archived d ataset with a
horizontal resolution of 2.5 832.58 (additional infor-
mation is available online at http://www.cdc.noaa.gov/
index.html; Kalnay et al. 1996). The data analyzed here
comprise the gridded JJA average values for each year
covering the region 608S–08,908E–1808 (Fig. 1).
The rainfall data analyzed here were provided by
the National Climate Centre (NCC) of the Bureau of
Meteorology (BoM) and consist of gridded data on a
0.25830.258 grid. These are described by Lo et al.
(2007), who indicate their confidence in the validity of
the data after 1948. In addition, four target regions were
selected: 1) southwest Western Autralia—the region
southwest of the line joining 308S, 1158E and 358S,
1208E; 2) the south Australian region (SA)—the box
area bounded by 328–368S, 1358–1408E; 3) the Victorian
region (VIC)—the box area bounded by 378–398S, 1438
1478E; and 4) the Tasmanian region (TAS)—the box
FIG. 1. Study area used for analyzing MSLP data and the four regional-scale study areas: 1)
SWWA (southwest of the line joining 308S, 1158E and 358S, 1208E); 2) SA (the box area
bounded by 328–368S, 1358–1408E); 3) VIC (the box area bounded by 378–398S, 1438–1478E);
4) TAS (the box area bounded by 418–438S, 1458–1488E).
1144 JOURNAL OF CLIMATE VOLUME 22
area bounded by 418–438S, 1458–1488E (Fig. 1). For each
region, the average winter (June–August) rainfall totals
over the 58-yr period from 1948 to 2005 were generated
(from the BoM Web site http://www.bom.gov.au/cgi-
bin/silo/cli_var/area_timeseries.pl). The driest region is
SA (111 mm) followed by VIC (249 mm), SWWA
(300 mm), and then TAS (373 mm). Finally, rainfall
data were extracted for sites corresponding to the lo-
cations of the capital cities within each region: Perth
Airport ( 31.928 38S, 115.97398E), Adelaide Airport
(34.95248S, 138.52048E), Melbourne Regional Office
(37.80758S, 144.978E), and Hobart Airport (42.8389 8S,
147.49928E).
The model data used to demonstrate the application
of the downscaling methodology were generated as part
of CSIRO’s participation in the World Climate Re-
search Programme (WCRP) Coupled Model Inter-
comparison Project phase 3 (CMIP3) (see http://www-
pcmdi.llnl.gov/ipcc/about_ipcc.php). The model used is
a version of the CSIRO Mk3 model (Gordon et al.
2002), referred to as Mk3.5 (Hirst 2007; Smith 2007),
and the climate change experiment is based on the A2
emissions scenario. Rotstayn et al. (2008, manuscript
submitted to Int. J. Climatol.) have assessed the Mk3.5
model in terms of its ability to simulate El Nin
˜
o–
Southern Oscillation (ENSO) teleconnection patterns
over Australia and found that it performs as well as, if
not better than, other international models.
3. Principal component analysis of MSLP
Let X(t) 5 [X
1
(t), X
2
(t), ..., X
p
(t)] be an n 3 p data
matrix, X
i
(t):i 5 1, ..., p; t 5 1, ..., n
fg
is a vector
containing n (yearly) values of the ith centered predic-
tor (MSLP anomalies related to the average over n
years at the ith grid point), and p is the number of
predictors (i.e., p 5 925 grid points over covering the
region 608S–08,908E–1808 in our study). The PC com-
ponent is obtained by
Z
m
(t) 5 å
p
k51
e
km
X
k
(t), m 5 1, ..., M; t 5 1, ..., n,
(1)
where e
km
are the elements (loadings) of the mth ei-
genvector of the covariance matrix
S 5
1
n 1
X
T
X.
The number M of components retained in the analysis is
based on a Scree test (Wilks 1995). Further details can
be found in Jolliffe (2002) and Wilks (1995).
The analysis of the MSLP data yields eight modes that
explain 95% of the total variance. The spatial patterns
are shown in Fig. 2, while the associated time series are
shown in Fig. 3. The trend correlation coefficient and
the correlations with the JJA Southern Oscillation index
(SOI) and the SAM index (Mo and White 1985; Karoly
1990; Gong and Wang 1999; Thompson and Wallace
2000) for each of the modes are shown in Table 1.
The first mode (Fig. 2a) represents an almost zonal
north–south gradient pattern but has a center over the
New Zealand region extending back over Australia,
contrasting with anomalies farther south. It explains the
majority (38%) of the variance in MSLP, and its time
series (Fig. 3a) exhibits a very significant increase (trend
correlation r 510.73) over the period from 1948
to 2005. It is also strongly linked to both the SOI
(r 5 0.44) and the SAM index (r 510.77), which
has also increased in strength over time. Note that the
link between these indices remains after the data are
detrended. The second mode (Fig. 2b) explains about
FIG. 2. (a)–(h) The first eight leading principal component pat-
terns of winter MSLP over the study area. Variance explained is
indicated in parentheses.
1M
ARCH 2009 L I A N D S M I T H 1145
28% variance and represents contrasting MSLP anom-
alies between the tropics and the Southern Ocean re-
gion directly to the south of Australia where it could be
expected to be associated with anomalous westerly
winds. Its time series (Fig. 3b) also exhibits a significant
increase (r 510.31) over time. After detrending, it can
be seen to be significantly linked to both the SOI
(r 510.36) and the SAM (r 5 0.41). The link to the
SOI corresponds to the fact that this mode explains
some of the MSLP variability over northern Australia.
It is similar to the pattern noted by Allan and Haylock
(1993, Fig. 7a) that is associated with relatively wet
winters in SWWA.
The third mode (Fig. 2c) explains about 13% of
the total variance and represents an east–west pattern
over the Southern Ocean. This pattern could be ex-
pected to be associated with anomalous southeasterly
winds over SWWA and anomalous southerly winds
over Tasmania. The fourth mode (Fig. 2d) explains 8%
of the total variance and is associated with negative
MSLP anomalies over the entire continent. It is linked
to interannual variations in the SAM (r 510.29).
Mode 6 (Fig. 2f) explains only 3% of the total variance
and is associated with positive MSLP anomalies cen-
tered on southeastern Australia. However, it is strongly
associated with the detrended SOI (r 5 0.54). The
remaining modes (5, 7, and 8) explain smaller-scale
features of MSLP variability and do not exhibit any
significant links to either ENSO or the SAM. Only the
first two modes exhibit significant long-term trends over
time.
The links between each of these modes and Austra-
lian winter rainfall is illustrated in Fig. 4, which shows
the correlations between individual gridpoint rainfall
FIG. 3. (a)–(h) The standardized time series (Z
1
Z
8
) of the leading eight MSLP modes.
1146 JOURNAL OF CLIMATE VOLUME 22
time series and each mode, while Table 2 summarizes
the links with regional rainfall. First, all but modes 5 and
8 are significantly linked to one or more of the regional
rainfall time series. Of these, mode 7 is probably in-
consequential since it is only weakly linked to SWWA
(r 510.27), which is much more strongly linked to the
higher modes 1–4. Mode 6 has a negative impact on SA
rainfall (r 5 0.26) and reflects the fact that this mode
corresponds to positive MSLP anomalies and suppressed
rainfall over the entire continent (Figs. 2f and 4f), re-
flecting the influence of ENSO. However, SA rainfall is
more strongly linked to higher modes 1, 2, and 4. To
check the robust relationship between the rainfall and
MSLP modes, we use a bootstrap technique to assess
the correlations between the eight leading PCs and JJA
rainfall. This is done by resampling, 1000 times, each PC
series and the JJA rainfall with replacement, and then
determining if the resulting correlation is significant
(i.e., if the MSLP mode score series has strong link to
rainfall in a specific region then the 95% resampled
correlations should be significant greater than zero at
the 0.05 level). Figure 5 shows regions (marked in red)
where 95% of the 1000 resampled series resulted in
correlations significant at the 0.05 level. It is evident that
modes 1–4 still yield significant correlation patterns
(Figs. 5a–d). Mode 6, however, when resampled, appears
TABLE 1. Links between the leading eight JJA MSLP modes, the
SOI, and the SAM index (1948–2005). The numbers within pa-
rentheses in the fourth and fifth columns represent correlations
between detrended time series. Values significant at p , 0.05 are in
bold.
Time
series
Variance
(%)
Trend
correlation r SOI r SAM
Z
1
37.5 10.73 20.44 (20.36) 0.77 (0.53)
Z
2
27.9 10.31 0.24 (0.36) 20.08 (20.41)
Z
3
12.5 10.01 20.14 (20.15) 0.12 (0.25)
Z
4
8.2 10.15 0.04 (0.08) 0.29 (0.29)
Z
5
3.7 20.03 0.21 (0.21) 0.01 (0.21)
Z
6
2.6 20.12 20.49 (20.54) 20.38 (20.38)
Z
7
1.5 10.11 20.13 (20.10) 0.10 (0.01)
Z
8
1.3 10.15 20.12 (20.08) 0.05 (20.05)
SOI 20.25 1.00 20.17 (20.003)
SAM 10.70 1.00
FIG. 4. (a)–(h) Correlation patterns between JJA rainfall and the leading eight time MSLP modes.
1M
ARCH 2009 L I A N D S M I T H 1147
to only yield significant correlations over the eastern
half of the continent, outside of the regions of interest
here. This suggests that the pattern evident in Fig. 4f
represents an overestimate of the importance of this
mode over most of the continent. Likewise, none of the
remaining modes (5, 7, and 8) yield any significant
correlations, again highlighting the importance of re-
sampling when looking for robust links.
In summary, it appears that the most valuable links
between MSLP and regional rainfall are contained in
modes 1–4. Of these, mode 4 has a negative impact on
SWWA rainfall since it appears to be associated with
positive MSLP anomalies in this region (Fig. 2d). On the
other hand, it has a positive impact on SA rainfall (Fig.
4d), consistent with enhanced southwesterly winds over
this region. A similar relationship is seen with mode 3
(Fig. 2c) where it suppresses rainfall in SWWA owing
to positive MSLP anomalies but enhances rainfall in
southwest TAS (Fig. 4c) owing to enhanced southerly/
southwesterly winds there.
As expected, the leading modes of MSLP variability
have the strongest links to regional winter rainfall. We
TABLE 2. Correlations between regional rainfall totals and various indices (1948–2005). Values significant at the 0.05 level are in bold.
Region SOI SAM Z
1
Z
2
Z
3
Z
4
Z
5
Z
6
Z
7
Z
8
SWWA Raw 0.34 20.37 20.40 0.37 20.34 20.28 0.25 20.17 0.24 0.11
Detrended 0.30 20.32 20.37 0.47 20.35 20.26 0.25 20.2 0.27 0.15
SA Raw 0.36 20.03 20.25 0.47 0.16 0.44 0.08 20.28 0.19 20.03
Detrended 0.41 20.13 20.51 0.45 0.16 0.43 0.08 20.26 0.17 20.05
VIC Raw 0.43 20.19 20.36 0.59 0.20 0.18 0.15 20.18 0.02 0.11
Detrended 0.44 20.21 20.52 0.63 0.20 0.18 0.15 20.19 0.02 0.12
TAS Raw 0.31 20.12 20.15 0.69 0.31 0.06 0.21 0.07 20.17 0.12
Detrended 0.36 20.2 20.33 0.70 0.31 0.04 0.22 0.08 20.19 0.11
FIG. 5. (a)–(h) Bootstrap assessment of significant correlations between the eight leading PC score series and JJA
rainfall. This is done by resampling, 1000 times, each PC series and the JJA rainfall with replacement, and then
determining if the resulting correlation is significant. The regions with 95% of the 1000 resampled correlations
significant at the 0.05 level are marked in red.
1148 JOURNAL OF CLIMATE VOLUME 22
note that modes 1 and 2 (detrended) are oppositely
correlated with both the SOI and the SAM (Table 1).
While both modes correspond to positive MSLP anom-
alies over northern Australia, they correspond to con-
trasting anomalies over southern Australia (Figs. 2a
and 2b). As a consequence, mode 1 (mode 2) reflects
increases (decreases) in MSLP and suppressed (en-
hanced) rainfall over much of southern Australia (Figs.
4a and 4b).
4. Principal component regression models
for rainfall
We use principal component regression (PCR) (e.g.,
Draper and Smith 1981; Jolliffe 2002) to arrive at a
predictive model for rainfall [ Y(t)] as follows:
Y(t) 5
å
M
m51
a
m
Z
m
(t) 1 e
t
, (2)
where e
t
is the residual variability not described by the
(M
5 8) MSLP modes that we have identified. For this
analysis, we could retain all eight modes but it is ap-
parent that not all the modes are necessary depending
on the region (see section 3). The method for develop-
ing or ‘‘calibrating’’ a predictive model is to use data
from a ‘‘training’’ period comprising the 43 years, 1948–
90, and to verify using data for the 15 years, 1991–2005.
The aim is to develop a robust model that provides a
downscaled prediction for rainfall given a predicted
large-scale MSLP field. Note that the rainfall data can
comprise regional-scale averages, gridpoint values, or
specific station values.
To decide how many modes are necessary to help
describe any particular rainfall time series, we use cross-
validation (Stone 1974). That is, using data from the
training period, we monitor the additional variance
explained by progressively adding in each of the modes.
We find that, in most cases, the first four (Z
1
to Z
4
)
modes are sufficient to help explain the variance. In the
FIG. 6. (a)–(d) The relative contribution in mm (thick lines) of each of the first four MSLP modes to
regional winter rainfall totals. The boxes (thin horizontal lines) represent the 50% (95%) confidence
intervals.
1M
ARCH 2009 L I A N D S M I T H 1149
case of Tasmania, only three of the modes add any
significant information.
Figure 6 provides an illustration of the relative con-
tribution (in mm) of each of the four components to the
regional winter rainfall totals in terms of the regression
coefficients associated with each of the first four stan-
dardized MSLP modes. The 95% confidence levels are
estimated using 1000 bootstrap replications as described
in the appendix. The total percentage variance explained
for each region is SWWA: 52%, SA: 55%, VIC: 61%,
and TAS: 56%. It is evident that Z
1
represents a negative
influence on rainfall for all four regions, whereas Z
2
represents a positive influence, particularly for TAS.
This is consistent with Figs. 4a and 4b and the fact that
the SAM has increased over recent decades (Cai et al.
2005; Li et al. 2005; Meneghini et al. 2006). As expected,
Z
3
and Z
4
have a negative influence on SWWA, but
positive influences on the eastern regions.
To quantify the uncertainty associated with any pre-
dictions with the model, we apply a bootstrap sampling
approach proposed by Stine (1985, see appendix A).
The estimates of the 95% confidence intervals of pre-
dicted rainfall for 1000 bootstrapping samples with re-
placement using independent testing MSLP data over
1991–2005. The uncertainty of the model is then indi-
cated by the spread of bootstrapping prediction inter-
vals for a given input MSLP field, shown as dashed curves
in Figs. 7 and 8.
5. Results
Figure 7 compares the predicted and observed rainfall
for all four regions, while Fig. 8 compares the results for
individual capital city stations within each of the regions
(Perth, Adelaide, Melbourne, and Hobart). In all cases,
a separate model was derived using the observed data
FIG. 7. (a)–(d) Predicted (red dash curve) vs observed JJA rainfall amounts (black solid curve) for
each of the four regions. The results for 1948–90 correspond to the training period, while the results for
1991–2005 correspond to the verification period. The blue dash curves are upper and lower bands of
the 95% confidence interval for the verified predictions.
1150 JOURNAL OF CLIMATE VOLUME 22
over the training period 1948–90. The predicted values
after 1990 therefore indicate the true skill of this ap-
proach. The uncertainty in terms of bootstrapping 95%
confidence intervals associated with the verified pre-
dictions after 1990 are also indicated in Figs. 7 and 8 by
the dashed curves.
Table 3 summarizes this skill by showing the associ-
ated correlation coefficients and the ratio of rms errors
(RMSE) to the climatology of JJA rainfall.
In general, the performance of the models evident in
the training period is maintained during the subsequent
verification period. All models perform reasonably well
at the regional scales since the percentage variance
explained in each case is relatively high (about 50% in
both periods) and the correlation between predicted
and observed values are all highly significant at the 0.01
level. In the case of the results for the four capital city
sites, it can be seen that the downscaling performance is
generally less, which is to be expected if local (unpre-
dictable) factors are contributing to the observed rainfall
amounts in each gauge. These factors are less relevant at
the larger regional scales since they tend to cancel out in
the averaging process.
To detect where this downscaling technique is most
useful, we have also used high-resolution (0.25830.258)
gridded rainfall data within each of the four selected
regions. Figure 9 indicates how the performance of the
models varies across the grid points within each region
by showing the correlation between observed and
predicted rainfall over the verification period 1991–
2005. In SWWA (Fig. 9a) the skill increases from the
south where it is moderate (r ’10.5) to high (r ’10.9)
in the north. In SA (Fig. 9b) the reverse holds true
where the skill is highest along the coastal regions but is
poor (r ’10.3) in the far north. Over VIC (Fig. 9c), the
skill is highly variable, being lowest (r ’10.5) in the
northwest and southeast of the box. Finally, the skill
over TAS (Fig. 9d) is highest over the northwest and
FIG. 8. (a)–(d) As in Fig. 6 but for the four station sites.
1M
ARCH 2009 L I A N D S M I T H 1151
central regions and lowest toward the east and south,
where Hobart is located.
6. Application to climate change simulations
We apply the downscaling models to MSLP data de-
rived from climate model (GCM) simulations for both
the present-day and future climate. The climate model
(CSIRO Mk3.5) has been used in an experiment in
which greenhouse gases are prescribed to increase ac-
cording to the A2 emissions scenario (Solomon et al.
2007). Mean winter (JJA) MSLP fields were calculated
for both present-day conditions (1971–2000) and for
later this century (2071–2100). Figure 10 compares the
observed (NCEP) long-term average JJA MSLP with
those simulated by the GCM for present-day and future
conditions. The pattern correlation between the ob-
served and the GCM-simulated long-term average JJA
MSLP is 0.969 for the present day and 0.967 for the
future. The MSLP values simulated for the future are
higher than present over most of the continent, with the
largest increases found just to the south. Decreases are a
feature of the differences toward the very high latitudes.
To downscale the GCM-generated values for MSLP
to regional JJA rainfall, we calibrated PCR model using
the observed rainfall and NCEP data from 1948 to 2005
based on the first four MSLP modes. The predictive
model linking rainfall (Y) to MSLP (via X) can be
written as Eq. (A2) in the appendix, which enables us to
generate downscaled rainfall estimates using GCM-
generated values for MSLP.
The relationship between observed, simulated, and
downscaled rainfall totals is summarized in Fig. 11. The
mean downscaled value is shown for each region ac-
companied by the 95% confidence intervals (thin hori-
zontal lines in Fig. 11). These confidence intervals re-
flect the uncertainty associated with the downscaling
model as estimated by the bootstrap procedure in the
appendix. In the SWWA case, the downscaled value
represents a significant underestimate, in the case of SA
and VIC they represent slight overestimates, while, in
the case of TAS, the downscaled value is a significant
overestimate. In all cases, the raw GCM values repre-
sent underestimates. Table 4 summarizes the results for
the four regions and indicates that, in the case of SWWA,
the GCM underestimate is 260%. [This is comparable
to the average underestimate (256%) from seven GCMs
analyzed by Hope 2006.] In all cases, the downscaled
values are higher than the GCM values but the per-
centage errors are much less: SWWA (232% cf.
260%), SA (114% cf. 220%), and VIC (113% cf.
251%). In the case of TAS, the downscaled error of
120% is similar in magnitude to the GCM error (217%)
although different in sign. On average, the GCM un-
derestimates winter rainfall totals by 237% whereas the
downscaled values represent an average overestimate of
only 14%. Note that the average of the absolute error in
the downscaled values is 20%, much less than the 37%
associated with the GCM values.
The GCM results for later this century all indicate
decreases compared to the present-day values. The
largest decrease (236%) is simulated for SA and
the smallest (27%) for TAS. In the case of SWWA, the
simulated decrease of 217% is comparable to the av-
erage (218%) from the seven GCMs analyzed by Hope
(2006). The downscaled values also indicate decreases,
the largest being 223% for SA and the smallest being
25% for SWWA. In the case of TAS, the downscaled
value (212%) is slightly larger than the GCM estimate
(27%). The average decrease according to the GCM
values is 222%, much higher than the average down-
scaled estimate of 213%. Moreover, according to the
Wilcoxon–Mann–Whitney rank sum test (Iman 1994),
the mean values for present and future downscaled
rainfall are significantly different (with p value 0), and
the difference and the percentage difference (%diff) be-
tween the present and future periods are also significantly
different from zero (with p- value 0). The associated
95% confidence intervals are given in brackets in the final
two columns of Table 4. Note that the Wilcoxon–Mann–
Whitney rank sum test (Iman 1994) is used because the
distribution of the downscaled rainfall series is un-
known. By taking into account the uncertainty of the
downscaled rainfall, it is evident that winter rainfall
over all southern Australian regions displays a highly
significant decrease with the smallest change evident
over SWWA (2
5%). These decreases are consistent
with the increases in simulated MSLP.
TABLE 3. Performance of the downscaling models at reproduc-
ing regional- and station-scale rainfall over the training period
(1948–90) and the verification period (1991–2005). Here r is the
correlation between the predictions and observations, r(%) is the
ratio of RMSE (mm) to the climatology of JJA rainfall.
Training period
(1948–90)
Verification period
(1991–2005)
r r (%) r r (%)
Regions
SWWA 0.72 12 0.69 13
SA 0.74 22 0.61 22
VIC 0.78 15 0.64 19
TAS 0.75 16 0.80 12
Stations
Perth 0.69 15 0.68 13
Adelaide 0.66 24 0.66 20
Melbourne 0.36 25 0.54 26
Hobart 0.51 31 0.55 25
1152 JOURNAL OF CLIMATE VOLUME 22
7. Discussion
We have analyzed winter (JJA) average MSLP data
for the Australian region (608S–08,908E–1808) and
found that the four leading modes can be used to ex-
plain a large fraction of the variability in winter rainfall
over much (but not all) of southern Australia. PCR
models have been developed that allow downscaled
estimates of rainfall to be made given predicted MSLP
fields. Rainfall predictions can be made at either re-
gional, grid square, and point scales. The results indicate
that robust models can be derived with relatively high
predictive capability, particularly at large regional scales,
but this performance is not uniform within regions. For
example, the variability of rainfall at capital city sites
Melbourne (within VIC) and Hobart (within TAS) ap-
pear to be less strongly linked to the large-scale MSLP
patterns than Perth (within SWWA) or Adelaide (within
SA). Even so, MSLP explains about 30% of the total
variance in the worst case and up to ;45% in the best
case.
GCM estimates for winter rainfall totals represent
relatively large underestimates (on average, 237%).
These may be caused by a number of factors including
poor physical representation of precipitation processes,
the absence of significant topography due to the coarse
horizontal resolution, or that the model underestimates
the moisture content of the atmosphere because of
problems elsewhere. Whatever the causes, the results
indicate some caution should be attached to the climate
change results for winter rainfall when such large dif-
ferences exist. However, the downscaled values provide
much more realistic estimates for present-day rainfall,
the average error being only 14% (average absolute
error being 20%).
The downscaled estimates for future changes to rain-
fall suggest a relatively small decrease of 29 mm for
SWWA compared to the GCM value of 221 mm. This
FIG. 9. The correlation between predicted and observed JJA rainfall over the verification
period 1991 to 2005 at grid points within each region: (a) SWWA, (b) SA, (c) VIC, and (d) TAS.
1M
ARCH 2009 L I A N D S M I T H 1153
translates to a percentage change of 25% compared to
the GCM value of 217%. In the case of SA, the abso-
lute decreases in rainfall are very similar (232 and 229
mm) but the percentage differences are different (236%
and 223%). In the case of VIC (233 and 245 mm) the
downscaled changes are much larger, but the GCM se-
verely underestimates present-day rainfall so that the
downscaled percentage change (216%) is, in fact, much
less than the GCM value (227%). Finally, the down-
scaled changes for TAS (254 mm) are much larger in
magnitude than the GCM changes (223) but the per-
centage changes are not so different (212% compared
to 27%).
We further show that the climate change signals, in
terms of the downscaled rainfall over these regions, are
highly significant (with p values 0) based on the
Wilcoxon–Mann–Whitney rank sum test (Iman 1994).
The average of the downscaled percentage changes is
213%, which is much less than that of the GCM (222%).
The question here is which set of values is likely to be
more accurate. A commonsense assumption is that that
the better the model representation of present-day con-
ditions, the more reliable an y subseq uent p redicted
changes. Whetton et al. (2007) demonstrate that, at least
in the world of climate mode results, this is appropriate
for rainfall at midlatitude regions. Importantly, they
showed that the degree of matching of current patterns
of MSLP was highly relevant to reliability of predicted
changes in the future. On this basis, downscaled values
are more likely to represent the changes in precipitation
that may arise owing to increased CO
2
concentrations in
the atmosphere than the raw model values themselves.
In this one specific case, the decreases simulated by
the GCM for later this century may be overestimates.
In applying this technique to GCM results, it will be
important to assess how well each GCM simulates both
the present-day mean MSLP patterns and also MSLP
variability. In this particular case, the GCM is know to
FIG. 10. Long-term average winter MSLP (hPa): (a) from NCEP data in 1970–2000; (b) simulated by
the CSIRO Mk3.5 for the present period 1970–2000; (c) simulated by the CSIRO Mk3.5 for the future
period 2071–2100. (d) MSLP difference between the future and present periods.
1154 JOURNAL OF CLIMATE VOLUME 22
perform well at simulating observed rainfall–ENSO
teleconnection patterns over Australia (Rotstayn et al.
2008, manuscript submitted to Int. J. Climatol.), but we
have not performed a detailed MSLP assessment. This
needs to be borne in mind when considering this single
result since a proper projection of rainfall changes will
need to be supplemented by similar analyses of other
GCM results.
It should be noted that the downscaling method in
this paper can only represent the changes in rainfall
linked to changes in the atmospheric circulation as we
only used one predictor field (MSLP) in the downscal-
ing model. Changes in rainfall in the future may be
brought about by changes in other factors not consid-
ered in the downscaling methods, for example, changes
in atmospheric humidity. To some extent, some changes
in humidity may be accounted for by the changes in the
MSLP patterns insofar as they can affect the direction of
the prevailing winds. They will not, however, be able to
account for large-scale changes in humidity associated
with global warming, but this effect is difficult to in-
corporate using present-day observations (Charles et al.
1999).
Like all other statistical downscaling methods, the
assumption of stationarity may be questionable (e.g.,
Huth 1997; Slonosky et al. 2001; Fowler and Kilsby
2002). The degree of nonstationarity in projected climate
change was assessed by Hewitson and Crane (2006) who
suggested that circulation dynamics may be robust to
nonstationarities.
Finally, in addition to the assumptions of linearity and
stationarity, downscaling results are constrained by the
quality of the predictors (e.g., MSLP) simulated by the
GCMs. In this case, a greater consensus about the ex-
pected changes to the MSLP patterns may lead to an
improved consensus of the expected changes to rainfall
by downscaling techniques. Further work will investi-
gate the application of these techniques to an ensemble
of GCM results.
Acknowledgments. We thank two anonymous ref-
erees whose comments greatly improved the paper. We
FIG. 11. Present (1971–2000) and future (2071–2100) JJA rainfall totals (thick lines) for (a) SWWA,
(b) SA, (c) VIC, and (d) TAS. The symbol ‘‘o’’ represents the observed value; the ‘‘1’’ and ‘‘x’’
symbols represent the GCM-simulated values for the present and future periods, respectively. The
boxes (thin horizontal lines) represent the 50% (95%) confidence intervals.
1M
ARCH 2009 L I A N D S M I T H 1155
thank Mark Collier for providing the CSIRO Mk3.5
data. This work was supported by a Western Australian
state government Indian Ocean Climate Initiative, the
CSIRO Climate Adaptation Flagship project and the
CSIRO Water for a Healthy Country Flagship project,
and the Australian Department of Climate Change
via the Australian–China Climate Change Partnership
Program.
APPENDIX
Bootstrapping Prediction Intervals
In general, a fitted PCR model (2) based on the least
squares fit to the n year training data (X(t), Y(t)):f
t 5 1, ..., ng is
^
Y(t) 5
å
M
m51
^a
m
Z
m
(t), (A1)
which can be rewritten as
^
Y(t) 5
å
M
m51
^a
m
å
p
k51
e
km
X
k
(t)
0
@
1
A
5 å
p
k51
^
b
k
X
k
(t), (A2)
where
^
b
k
5 å
M
m51
^a
m
e
km
. Now, assume we have N . n
year observed datasets. A calibrated or ‘‘trained’’ model
based on the dataset (X(t), Y(t)): t 5 1, ..., nfgcan be
tested for its robustness by using independent data
( X(t 1 h), Y(t 1 h)): h 5 1, ..., N n
fg
.
To quantify the uncertainty of downscaled rainfall
estimates using Eq. (A2) related to a ‘‘future’’ X
i
(t 1 h)
(i 5 1, ..., p; h 5 1, ..., N n), we need to establish
the cumulative distribution function G for the confi-
dence interval of the prediction error Y(t 1 h)
^
Y(t 1 h), where
^
Y(t 1 h) 5
å
p
k51
^
b
k
X
k
(t 1 h) by using
(A2). A 100(1 a)% predication interval for a future
rainfall Y(t 1 h) based on a ‘‘future’’ X
i
(t 1 h)
(i 5 1, ..., p; h 5 1, ..., N n) is given by
[
^
Y(t 1 h) 1 G
1
(a/2),
^
Y(t 1 h) 1 G
1
(1 a/2)].
(A3)
However, as the distribution F of the residual variability
e
t
in the PCR model (2) is unknown, we cannot obtain
the distribution G analytically. So we apply a bootstrap
sampling approach proposed by Stine (1985) to estimate
the distribution G. Briefly the adapted approach is as
follows.
First, the parameter estimate
^
b
k
are calculated using
the observed training data (X(t), Y(t)), t 5 1, ..., n
fg
.
These estimates are then used to calculated fitted values
^
Y(t) using Eq. (A2) and residuals e
t
5 Y(t)
^
Y(t). For
an independent testing data X(t 1 h), 1 # h #
f
N ng, predicted value
^
Y(t 1 h) 5
å
p
k 5 1
^
b
k
X
k
(t 1 h).
The error distribution F is estimated by the empirical
distribution of residuals, which we denote F
n
. This is
then used to construct B bootstrapped samples of the
form
TABLE 4. Regional JJA mean rainfall for both the present (1971–2000) and future (2071–2100). A comparison between observed,
GCM simulated, and downscaled GCM values. The 95% confidence levels for the downscaled (DS) amounts are based on 1000 bootstrap
samples with replacement and are shown in parentheses in columns 3 and 6. The confidence levels for the difference and the percentage
differences (between the future and present periods) are estimated based on the Wilcoxon–Mann–Whitney rank sum test and are shown
in parentheses in the final two columns.
Region
Present (1971–2000) Future (2071–2100)
Source Amount (mm)
Diff
with obs (mm) Diff (%) Amount (mm)
Diff with
(1971–2000) (mm) Diff (%)
SWWA Obs 300
GCM 121 2179 260 100 221 217
DS 205 (154, 261) 295 232 196 (142, 254) 29(212, 27) 25(24.99, 22.5)
SA Obs 111
GCM 88 223 220 56 232 236
DS 126 (94, 158) 115 14 97 (66, 130) 229 (230, 227) 223 (223.5, 221.5)
VIC Obs 249
GCM 122 2127 251 89 233 227
DS 282 (230, 337) 152 13 237 (182, 293) 245 (246, 241) 216 (215.9, 214)
TAS Obs 373
GCM 307 266 217 284 223 27
DS 448 (387, 508) 175 20 394 (335, 453) 2
54 (261, 253) 212 (213.5, 211.9)
Average GCM 237 222
DS 14 213
1156 JOURNAL OF CLIMATE VOLUME 22
(X(t), Y
(t)), t 5 1, ..., n
fg
,(X(t 1 h), Y
(t 1 h)),
with
^
Y
(t) 5
^
Y(t) 1 e
t
and Y
(t 1 h) 5
^
Y(t 1 h) 1 e
9
t
,
where e
t
and e
9
t
are independently sampled from F
n
;
that is, they are randomly sampled with replacement
from the set of residuals e
1
, ..., e
n
fg. The asterisk su-
perscript denotes a value constructed for a particular
bootstrap sample.
Each bootstrapped sample is used to calculated a
simulated estimate
^
b
k
, predicted value
^
Y
(t 1 h), and
predicted error e
9
t
5 Y
(t 1 h)
^
Y
(t 1 h). The em-
pirical distribution of e
9
t
, which we denote
~
G, is then an
estimate of the distribution of the bootstrap prediction
errors. It can be used as the distribution function G in
Eq. (A3). Therefore, a 100(1 a)% predication interval
for a future rainfall Y(t 1 h) can be estimated as
[
^
Y(t 1 h) 1
~
G
1
(a/2),
^
Y(t 1 h) 1
~
G
1
(1 a/2)]. (A4)
In our rainfall analysis, estimates of the 95% confidence
interval of predicted rainfall for 1000 bootstrapping
samples using independent testing MSLP data over
1991–2005 are shown as dashed blue curves in Figs. 6
and 7. Further, bootstrapping estimates of the uncer-
tainty (50% and 95% confidence intervals for 1000
bootstrap replications) of downscaled GCM-simulated
present and future MSLP to regional mean rainfall to-
tals based on the model calibrated using observed
rainfall and NCEP MSLP are shown in Fig. 11.
REFERENCES
Allan, R. J., and M. R. Haylock, 1993: Circulation features asso-
ciated with the winter rainfall decrease in southwestern
Australia. J. Climate, 6, 1356–1367.
Bates, B. C., P. Hope, B. Ryan, I. Smith, and S. Charles, 2008: Key
findings from the Indian Ocean Climate Initiative and their
impact on policy development in Australia. Climatic Change,
89, 339–354.
Benestad, R. E., 2001: A comparison between two empirical
downscaling strategies. Int. J. Climatol., 21, 1645–1668.
Cai, W., and I. G. Watterson, 2002: Modes of interannual varia-
bility of the Southern Hemisphere circulation simulated by
the CSIRO climate model. J. Climate, 15, 1159–1174.
——, G. Shi, and Y. Li, 2005: Multidecadal fluctuations of winter
rainfall over southwest Western Australia simulated in the
CSIRO Mark 3 coupled model. Geophys. Res. Lett., 32,
L12701, doi:10.1029/2005GL022712.
Charles, S. P., B. C. Bates, P. H. Whetton, and J. P. Hughes, 1999:
Validation of downscaling models for changed climate con-
ditions: Case study of southwestern Australia. Climate Res.,
12, 1–14.
——, ——, I. N. Smith, and J. P. Hughes, 2004: Statistical down-
scaling of daily precipitation from observed and modelled
atmospheric fields. Hydrol. Processes, 18, 1373–1394.
Christensen, O. B., J. H. Christensen, B. Machenhauer, and M.
Botzet, 1998: Very high-resolution climate simulations over
Scandinavia—Present climate. J. Climate, 11, 3204–3229.
Draper, N. R., and H. Smith, 1981: Applied Regression Analysis.
2nd ed. John Wiley and Sons, 709 pp.
Fowler, H. J., and C. G. Kilsby, 2002: Precipitation and the North
Atlantic Oscillation: A study of climatic variability in north-
ern England. Int. J. Climatol., 22, 843–866.
——, S. Blenkinsop, and C. Tebaldi, 2007: Review linking climate
change modelling to impacts studies: Recent advances in
downscaling techniques for hydrological modelling. Int. J.
Climatol., 27, 1547–1578.
Gong, D., and S. Wang, 1999: Definition of Antarctic oscillation
index. Geophys. Res. Lett., 26, 459–462.
Gordon, H. B., and Coauthors, 2002: The CSIRO Mk3 Climate System
Model. CSIRO Atmospheric Research Tech. Paper 60, 130 pp.
Hendon, H. H., D. W. J. Thompson, and M. C. Wheeler, 2007:
Australian rainfall and surface temperature variations asso-
ciated with the Southern Hemisphere annular mode. J. Cli-
mate, 20, 2452–2467.
Hewitson, B. C., and R. G. Crane, 2006: Consensus between GCM
climate change projections with empirical downscaling: Pre-
cipitation downscaling over South Africa. Int. J. Climatol., 26,
1315–1337.
Hirst, A. C., 2007: The CSIRO Mk3.5 Model—Simulations and
evaluation. Extended Abstracts, 14th National AMOS Conf.,
AMOS Publication 22, Adelaide, South Australia, Australia,
AMOS, 50 pp.
Hope, P. K., 2006: Projected future changes in the synoptic systems
influencing southwest Western Australia. Climate Dyn., 26,
765–780.
——, W. Drosdowsky, and N. Nicholls, 2006: Shifts in the synoptic
systems affecting southwest Western Australia. Climate Dyn.,
26, 751–764.
Hughes, J. P., P. Guttorp, and S. P. Charles, 1999: A non-
homogeneous hidden Markov model for precipitation occur-
rence. Appl. Stat., 48, 15–30.
Huth, R., 1997: Potential of continental-scale circulation for the
determination of local daily surface variables. Theor. Appl.
Climatol., 56, 165–186.
Iman, R. L., 1994: A Data-Based Approach to Statistics. Duxbury
Press, 348 pp.
Jolliffe, I. T., 2002: Principal Component Analysis.2nded.Springer,
487 pp.
Kalnay, E., and Coauthors, 1996: The NCEP/NCAR 40-Year
Reanalysis Project. Bull. Amer. Meteor. Soc., 77, 437–471.
Karoly, D., 1990: The role of transient eddies in low-frequency
zonal variations of the Southern Hemisphere circulation.
Tellus, 42, 41–50.
Li, Y., W. Cai, and E. P. Campbell, 2005: Statistical modelling of
extreme rainfall in southwest Western Australia. J. Climate,
18, 852–863.
Lo, F., M. C. Wheeler, H. Meinke, and A. Donald, 2007: Proba-
bilistic forecasts of the onset of the north Australian wet
season. Mon. Wea. Rev., 135, 3506–3530.
Lu, J., G. A. Vecchi, and T. Reichler, 2007: Expansion of the
Hadley cell under global warming. Geophys. Res. Lett., 34,
L06805, doi:10.1029/2006GL028443.
Maxino, C. C., B. J. McAvaney, A. J. Pitman, and S. E. Perkins,
2008: Ranking the AR4 climate models over the Murray-
Darling Basin using simulated maximum temperature, mini-
mum temperature and precipitation. Int. J. Climatol., 28,
1097–1112.
Meneghini, B., I. Simmonds, and I. Smith, 2006: Association be-
tween Australian rainfall and the southern annular mode. Int.
J. Climatol., 27, 109–121.
1M
ARCH 2009 L I A N D S M I T H 1157
Milly, P. C. D., J. Betancourt, M. Falkenmark, R. M. Hirsch, Z. W.
Kundzewicz, D. P. Lettenmaier, and R. J. Stouffer, 2008:
Stationarity is dead: Whither water management? Science,
319, 573–574.
Mo, K. C., and C. H. White, 1985: Teleconnections in the Southern
Hemisphere. Mon. Wea. Rev., 113, 22–37.
Murphy, J., 1999: An evaluation of statistical and dynamical
techniques for downscaling local climate. J. Climate, 12, 2256–
2284.
Perkins, S. E., and A. J. Pitman, 2008: Do weak AR4 models bias
projections of future climate changes over Australia? Climatic
Change, in press.
——, ——, N. J. Holbrook, and J. McAveney, 2007: Evaluation of
the AR4 climate models’ simulated daily maximum temper-
ature, minimum temperature, and precipitation over Aus-
tralia using probability density functions. J. Climate, 20, 4356–
4376.
Ra
¨
isa
¨
nen, J., 2007: How reliable are climate models? Tellus, 59A,
2–29.
Schmidli, J., F. Cheristoph, and P. L. Vidale, 2006: Downscaling
from GCM precipitation: A benchmark for dynamical and
statistical downscaling methods. Int. J. Climatol., 26, 679–689.
Slonosky, V. C., P. D. Jones, and T. D. Davies, 2001: Atmospheric
circulation and surface temperature in Europe from the 18th
century to 1995. Int. J. Climatol., 21, 63–75.
Smith, I. N., 2007: Global climate modelling within CSIRO: 1981
to 2006. Aust. Meteor. Mag., 56, 131–152.
——, P. McIntosh, T. J. Ansell, C. J. C. Reason, and K. McInnes,
2000: Southwest Western Australian winter rainfall and its
association with Indian Ocean climate variability. Int. J. Cli-
matol., 20, 1913–1930.
——, L. Wilson, and R. Suppiah, 2008: Characteristics of the
northern Australian rainy season. J. Climate, 21, 4298–4311.
Solomon, S., D. Qin, M. Manning, Z. Chen, M. Marquis, K. B.
Averyt, M. Tignor, and H. L. Miller, Eds., 2007: Climate
Change 2007: The Physical Science Basis. Cambridge Univer-
sity Press, 987 pp.
Stine, R. A., 1985: Bootstrap prediction intervals for regression.
J. Amer. Stat. Assoc., 80, 1026–1031.
Stone, M., 1974: Cross-validatory choice and assessment of statistical
predictions (with discussion). J. Roy. Stat. Soc., 36, 111–147.
Sturman, A., and N. Tapper, 1996: The Weather and Climate of
Australia and New Zealand. Oxford University Press, 476 pp.
Suppiah, R., K. J. Hennessy , P. H. Whetton, K. McInnes, I.
Macadam, J. Bathols, J. Ricketts, and C. M. Page, 2007:
Australian climate chang e projections derived from simula-
tions performed f or the IPCC 4th Assessment Report. Aust.
Meteor. Mag., 56, 131–152.
Thompson, D. W. J., and J. M. Wallace, 2000: Annular modes in
the extratropical circulation. Part I: Month-to-month varia-
bility. J. Climate, 13, 1000–1016.
Timbal, B., 2004: Southwest Australia past and future rainfall
trends. Climate Res., 26, 233–249.
——, and B. J. McAvaney, 2001: An analogue-based method to
downscale surface air temperature: Application for Australia.
Climate Dyn.,
17, 947–963.
——, and D. Jones, 2008: Future projections of winter rainfall in
Southeast Australia using a statistical downscaling technique.
Climatic Change, 86, 165–187.
——, J. M. Arblaster, and S. Power, 2006: Attribution of the late-
twentieth-century rainfall decline in southwest Australia. J.
Climate, 19, 2046–2062.
von Storch, H., B. Hewitson, and L. Mearns, 2000: Review of
empirical downscaling techniques. Regional climate devel-
opment under global warming. Proc. RegClim Spring Meeting,
General Tech. Rep. 4, Jevnaker, Torbjørnrud, Norway, Nor-
wegian Institute for Air Research, 29–46.
Whetton, P. H., K. L. McInnes, R. N. Jones, K. J. Hennessy, R.
Suppiah, C. M. Page, J. Bathols, and P. J. Durack, 2005:
Australian climate change projections for impact assessment
and policy application: A review. CSIRO Marine and At-
mospheric Research Paper 001, 34 pp.
——, I. Macadam, B. Bathols, and J. O’Grady, 2007: Assessment
of the use of current climate patterns to evaluate regional
enhanced greenhouse response patterns of climate models.
Geophys. Res. Lett., 34, L14701, doi:10.1029/2007GL030025.
Wilby, R. L., 1997: Non-stationarity in daily precipitation series:
Implications for GCM down-scaling using atmospheric cir-
culation indices. Int. J. Climatol., 17, 439–454.
Wilks, D. S., 1995: Statistical Methods in the Atmospheric Science.
Academic Press, 467 pp.
Zorita, E., and H. von Storch, 1997: A survey of statistical down-
scaling results. Tech. Rep. 97/E/20, GKSS, 42 pp.
——, and ——, 1999: The analog method as a simple statistical
downscaling technique: Comparison with more complicated
methods. J. Climate, 12, 2474–2489.
1158 JOURNAL OF CLIMATE VOLUME 22
... One of the most commonly employed methods in SD includes principal component regression (PCR, Kistler et al 2001, Sankarasubramanian et al 2008, Palomino-Lemus et al 2015, applied to the predictor field, in which the leading principal components (PCs) are fed into a multiple linear regression model to generate the SD model, overcoming multicollinearity problems. There are several methodologies to obtain the leading PCs: by projecting the EOFs from calibration period using observational data onto the GCMs outputs (Li and Smith 2009, Palomino-Lemus et al 2015 or using common EOFs (Benestad 2001(Benestad 2002. The fitted SD model is applied to the results from the GCMs outputs, providing high-resolution projections (Wilks 2006, Benestad et al 2008, Li and Smith 2009, Eden and Widmann 2014, Song et al 2014, Palomino-Lemus et al 2015, Mtongori et al 2016. ...
... There are several methodologies to obtain the leading PCs: by projecting the EOFs from calibration period using observational data onto the GCMs outputs (Li and Smith 2009, Palomino-Lemus et al 2015 or using common EOFs (Benestad 2001(Benestad 2002. The fitted SD model is applied to the results from the GCMs outputs, providing high-resolution projections (Wilks 2006, Benestad et al 2008, Li and Smith 2009, Eden and Widmann 2014, Song et al 2014, Palomino-Lemus et al 2015, Mtongori et al 2016. An effective technique to reduce potential uncertainties in the projections is the use of the multi-model ensemble (Benestad 2002, Thomson et al 2006, Semenov and Stratonovitch 2010, Wilby and Harris 2006, Maurer 2007. ...
... Following the scheme proposed by Smith (2009) andPalomino-Lemus et al (2015), a statistical downscaling based on the Principal Component Regression (PCR) method is employed to achieve statistical relationships between the large-scale predictor (NCEP summer SLP) and regional rainfall. The PCs accounting for a high percentage of explained reanalysis SLP variance, and presenting significant correlations with precipitation (see text S1 and figure S1 in the supporting information for details) were considered as predictors in a multiple regression model, for each precipitation grid point, to establish the statistical downscaling (SD) model, using 1950-1993 and 1994-2010 as the calibration and validation periods, respectively. ...
Article
Full-text available
Climate change projections for the last 30 years of the 21st century, for boreal summer precipitation in tropical America, have been made by developing a statistical downscaling (SD) model applied to the SLP outputs of 20 GCMs of CMIP5, for present climate (1970–2000), and for future (2071–2100) under the RCP2.6, RCP4.5 and RCP8.5 scenarios. For present climate, many SD GCMs faithfully reproduce the precipitation field in many regions of the study area. For future climate, as the radiative forcing increases, the projected changes intensify and the regions affected expand, with higher coherence between models. The zone between central and southeastern Brazil registered the most pronounced precipitation changes by a large number of SD models, even for the RCP2.6. Except for this region in Brazil, in general, the changes in rainfall range from moderate (± 25%) to intense (from ±70% to ±100%) as the radiative forcing increases from the RCP2.6−RCP8.5. For this latter scenario, all SD models present significant precipitation changes for more than 50% of the area, in some cases reaching 75% of area with significant changes. For the ensemble mean, the results show three extensive regions with significant changes under the three scenarios, the most highlighted changes being for the RCP8.5: a northwest-southeast band that extends from northern Mexico to eastern Brazil, crossing through northern Colombia, along with the regions in the south of the study area, with generally moderate precipitation increases; and a band that extends from eastern Ecuador to southeastern Brazil, with major decreasing changes. This pattern of change could be related with a possible strengthening in frequency in terms of La Niña events for the end of the century.
... This comparison study further reinforces the findings of [Das, 1955, Fealy and Sweeney, 2007, Ailliot et al., 2015, Holsclaw et al., 2017, Bertolacci et al., 2019, Xie et al., 2023, all reporting success with marginal models based on Gamma densities. PCA regression was used in Li and Smith [2009] for precipitation downscaling using CM variables together with mean sea level pressure (MSLP) measurements, alleviating the underprediction problems faced by CMs. Our JGNM can be seen as a generalisation, where the refinement of variables used in a regression model is made in a data-driven way. ...
... Frost et al. [2011], a review of six downscaling methods (namely Chandler [2002], Timbal [2004], Mc-Gregor [2005], Mehrotra and Sharma [2007], Chiew et al. [2009] and a HMMs-based implementation of Kirshner [2007]), found that HMMs displayed weaknesses in capturing spatial correlation. Applying Bayesian HMMs to rainfall, Song et al. [2014] (expanding on the approach of Li and Smith [2009]) uses a covariance matrix construction with spatial distances used only in error correlation. Holsclaw et al. [2017] assume a more complicated time-dependency in transition probabilities, which calls for a Polya-Gamma-based MCMC sampling scheme. ...
Preprint
Full-text available
This work introduces a novel approach for generating conditional probabilistic rainfall forecasts with temporal and spatial dependence. A two-step procedure is employed. Firstly, marginal location-specific distributions are jointly modelled. Secondly, a spatial dependency structure is learned to ensure spatial coherence among these distributions. To learn marginal distributions over rainfall values, we introduce joint generalised neural models which expand generalised linear models with a deep neural network to parameterise a distribution over the outcome space. To understand the spatial dependency structure of the data, a censored latent Gaussian copula model is presented and trained via scoring rules. Leveraging the underlying spatial structure, we construct a distance matrix between locations, transformed into a covariance matrix by a Gaussian Process Kernel depending on a small set of parameters. To estimate these parameters, we propose a general framework for the estimation of Gaussian copulas employing scoring rules as a measure of divergence between distributions. Uniting our two contributions, namely the joint generalised neural model and the censored latent Gaussian copulas into a single model, our probabilistic approach generates forecasts on short to long-term durations, suitable for locations outside the training set. We demonstrate its efficacy using a large UK rainfall data set, outperforming existing methods.
... Climate models are the primary tools available to simulate future climate impacts under different emission scenarios . But there is an agreement among the scientific community that global climate model (GCM)-simulated climate data cannot be directly used as input to hydrological models (Li and Smith, 2009;Liu et al., 2014). In order to use the output of a GCM for conducting hydrological impact studies, downscaling is used in this study, which is a process of converting the coarse spatial resolution of the GCM output into a fine resolution. ...
... Climate models sometimes have difficulty in realistically simulating climatic variables on smaller scales, which in turn affects the confidence that can be placed in their attempts to simulate future changes (Hewitson and Crane, 2006;Li and Smith, 2009). In order to use the output of a GCM for conducting hydrological impact studies, downscaling is used, which is a process of converting the coarse spatial resolution of the GCM output into a fine resolution. ...
Thesis
The study was conducted in the Bilate River Watershed. Bilate River is one of the inland rivers of Ethiopia that drains in to the northern watershed of the Lake Abaya-Chamo Drainage Basin which forms part of the Main Ethiopian Rift and in turn is part of an active rift system of the Great Rift Valley in Africa. This study examined the extent and nature of rainfall variability from recorded data while estimation of evapotranspiration was derived from recorded weather data. Future climate scenarios of precipitation and temperature for the Bilate Watershed were also generated. Analysis of rainfall variability was made by the rainfall anomaly index, coefficient of variance and Precipitation Concentration Index. The FAO-56 reference ET (ETo) approach was used to determine the amount of evapotranspiration. Estimation of the onset and the end of the growing season, and the length of the growing period was done using Instat software. The results show that mean annual rainfall of the upper (2307 m.a.s.l), middle (1772 m.a.s.l) and lower (1361 m.a.s.l) altitude zones of the watershed are in the order of 1100 mm, 1070 mm and 785 mm with CV of 12%, 15% and 17% respectively. Based on the rainfall data record of the latest 30 years, there was a high temporal anomaly in rainfall between 1980 and 2013. The wettest years recorded a Rainfall Anomaly Index of +5, +6 and +8 for stations in the upper, middle and lower altitude zones respectively, where the driest year recorded value is -5 in all the stations. The average onset date of rainfall for the upper zone is April 3+ 8 days, for the middle zone April 10 + 10 days and for the lower zone April 11+ 11 days with CV of 23%, 26% and 29% respectively. The average end dates of the rainy season in the upper and middle zones are October 3+ 5 days and September 25+ 7 days with CV 5% and 7%. The main rainy season ends earlier in the lower zone; it is on July 12 + 10 days with CV of 14%. Climate change scenarios were generated for two Representative Concentration Pathways (RCPs): RCP 4.5 and RCP 8.5 using 20 GCMs from CMIP5 bias-corrected under three future time slices, near-term (2010-2039), mid-century (2040-2069) and end-century (2071-2099). Rainfall is projected to increase in total amount under all-time slices and emissions pathways but with pronounced inter and intra-variability. Minimum temperature will significantly increase during mid-century by 1.810C (RCP 4.5) and xiii 2.550C (RCP 8.5) and by 2.10C (RCP 4.5) and 4.270C (RCP8.5) during end-century. The projected increase in maximum temperature during mid-century is 1.430C under RCP 4.5 and 1.99 0C under RCP 8.5 and during end-century by 1.650C under RCP 4.5 and 3.50C under RCP8.5 during end-century. The Soil and Water Assessment Tool (SWAT) model was selected to simulate stream flow of the watershed. The Alaba Kulito gauging station monthly stream flows from 1990 to 1996 and 1997 to 2002 were used for stream flow calibration and validation respectively. The respective statistical results of the coefficient of determination (R2), Nash–Sutcliffe coefficient (NSE) and percent bias (PB) are 0.79, 0.78 and 0.56 for the calibration period and 0.64, 0.60 and -21.7 for the validation period which show that the model predicted the stream flow at the Alaba Kulito gauging station reasonably. The annual stream flow increased progressively throughout the century for all time periods under both RCP scenarios. The increases under RCP 8.5 scenario are the larger compared to RCP 4.5 scenarios, approximately 42.42% during the 2080s period. The six GCMs selected to see the uncertainties related to GCMs suggest that the river flow will change by small amounts of −6.18 to 7.83% change compared with the baseline. The simulated runoff in the Bilate River depends on the projected amount of rainfall embedded in the GCM structures selected to simulate the future climate and is less dependent on the local temperature increment. The study also assessed the farmers‘ perceptions of the changes on climatic variables and their adaptation options to the impacts of climate variability and change. The determinant factors that influence the choice of farmers to climate change adaptation were also investigated. Above 92% of the surveyed farm households perceived variability and change in climatic variables but 59% of the households participated in one or other of the six major adaptation strategies which most prevailed inside farmers of the watershed. Changing crop variety, using water harvesting scheme, intensifying irrigation, using cover crop or/and mulching, reducing the number of livestock owned and getting offfarm jobs are the main adaptation strategies used by the farming households. The results from the binary logistic model further showed that age and educational level of the household head, farm size and the income level of the household are household characteristics that significantly affect the choice of adaptation options, while access to climate information in the form of seasonal forecasts and local agro ecology are other factors that determined the selection of adaptation methods by the farming households in the study area. The main constraints to adaptation to climate change in the study area were seen to be the knowledge gap in the form of lack of information, shortage of labour and minimal land size. These were the three most explained constraints to climate change as explained by responding household heads.
... Currently, most predictions use global circulation models (GCMs), but GCMs are not suitable for surface parameters in specific regions and at sub-grid scales (Risbey and Stone, 1996). To solve this problem, many statistical models have been developed based on physical connections between regional factors and other variables (e.g., Hughes and Guttorp, 1994;Wilby and Wigley, 1997;Goodess and Palutikof, 1998;Salathe, 2003;Hewitson and Crane, 2006;Fowler et al., 2007;Chu et al., 2008;Li and Smith, 2009;Huang et al., 2011;Ruan et al., 2015). For example, Li and Smith (2009) linked the mean sea-level pressure with rainfall over southern Australia during winter and developed a statistical model to predict future regional precipitation. ...
... To solve this problem, many statistical models have been developed based on physical connections between regional factors and other variables (e.g., Hughes and Guttorp, 1994;Wilby and Wigley, 1997;Goodess and Palutikof, 1998;Salathe, 2003;Hewitson and Crane, 2006;Fowler et al., 2007;Chu et al., 2008;Li and Smith, 2009;Huang et al., 2011;Ruan et al., 2015). For example, Li and Smith (2009) linked the mean sea-level pressure with rainfall over southern Australia during winter and developed a statistical model to predict future regional precipitation. Ruan et al. (2015) successfully developed a model to forecast late-winter precipitation over southwest China based on sea-level pressure in Western Europe and sea-surface temperature in the Western Pacific. ...
Article
Full-text available
Using observations and reanalysis, we develop a linear regression model to predict April precipitation in the northwestern United States (PNWUS) with a 1-month lead. Not only does this model reproduce April PNWUS for the training period, but also its predictions are robust and reliable for the independent test period. Two independent factors, Arctic stratospheric ozone (ASO) and geopotential height in western North America at 700 hPa (H700WNA), are used to construct the linear regression model. Based on observations, the possible mechanism by which ASO and H700WNA affect PNWUS is as follows. April circulation anomalies over the North Pacific related to March ASO changes can extend eastward to the western United States, causing April PNWUS anomalies. When the April H700WNA is abnormally high, there will be an anti-cyclonic circulation anomaly over the northwestern United States, which not only inhibits water-vapor transport from the North Pacific to the northwestern United States but also suppresses convective activity. This process also influences the April PNWUS. A transient experiment using a climate model with a longer period also agrees with the above result from observations. Using the predicted April H700WNA from the climate forecast system (CFS) of the National Centers for Environmental Prediction (NCEP) based on March data and the observed March ASO, the April PNWUS predicted by this statistical model is closer to observations than the April PNWUS predicted directly by the CFS using March data.
... При получении статистических соотношений используются различные методы: множественная линейная регрессия, корреляционный анализ и многие другие, в том числе искусственные нейронные сети (см. ссылки в работах [Kidson and Tompson, 1998;Li and Smith, 2009]). ...
... При этом может привлекаться крупномасштабная информация не только из того региона, для которого требуется получить детализированные поля. Так, в работе [Li and Smith, 2009] рассматривается процедура восстановления зимних осадков в Южной Австралии по полям давления на уровне моря. При этом привлекается информация из региона, значительно превышающего Австралию. ...
Article
Full-text available
Рассматриваются современные методы повышения детализации метеорологических полей с недостаточным пространственным разрешением. Кратко приводятся основные достоинства и недостатки статистических, физических и динамико-статистических методов. Большее внимание уделяется динамическим методам, основанным на интегрировании численных моделей атмосферы с высоким пространственным разрешением. Анализируются различные способы использования данных с грубым разрешением для коррекции состояния численной модели при подобном интегрировании. Кратко рассматривается методология циклического инкрементного усвоения данных, пригодная для повышения детализации метеорологических полей.
... where T is the time coefficient and V is the spatial eigenvector. In this process, the main information of the variable field is represented by several typical eigenvectors [36,37]. ...
Article
Full-text available
Based on data from the Gurbantunggut Desert, the largest fixed/semi-fixed desert in China, and ERA5-Land reanalysis data, the long-term variations and spatial surface heat source (SHS) differences in the Gurbantunggut Desert are discussed herein. The results show the following: (1) The hourly SHS at the Kelameili station during the 2013–2021 period was a weak heat source at night; contrastingly, it was a strong heat source during the day. The duration of the hourly SHS increased gradually from January to July, but it decreased gradually from July to December. The daily SHS showed obvious seasonal variation, reaching the maximum in summer and the minimum in winter. The ERA5-Land reanalysis can reproduce all the variation characteristics of the SHS well. (2) The climatology (i.e., multi-year mean) of the monthly SHS intensity was lower than 50 W/m2 during the January–March and September–December periods in the Gurbantunggut Desert, indicating a weak heat source. On the other hand, the climatology recorded in April–August was higher than 50 W/m2, with a strong heat source. From the perspective of spatial distribution, the eastern and western regions of the Gurbantunggut Desert show strong heat sources, while the central region shows weak heat sources. The spatial distribution of the first and second modes of the empirical orthogonal function (EOF) decomposition reflected the consistent spatial variability and a north–south (or east–west) polarity variation of the monthly SHS in the Gurbantunggut Desert, respectively. (3) The yearly SHS showed negative anomalies during the 1950–1954, 1964–1982 and 2004–2015 periods, and positive anomalies during the 1955–1963, 1983–2003 and 2016–2021 periods in the Gurbantunggut Desert. Additionally, the time series of the SHS anomalies was positively correlated with the Interdecadal Pacific Oscillation (IPO) index. During the negative IPO phase, the yearly SHS showed a negative anomaly in the Gurbantunggut Desert, while the yearly SHS showed a positive anomaly during the positive IPO phase in most regions of the Gurbantunggut Desert.
... There have been many statistical downscaling methods proposed by past researchers. The simplest form is linear regression, which estimates target predictand using an optimized linear combination of local circulation features [19][20][21]. The features are usually represented as the leading Principal Components (PC) of moisture, pressure, and wind field. ...
Article
Full-text available
Precipitation downscaling is widely employed for enhancing the resolution and accuracy of precipitation products from general circulation models (GCMs). In this study, we propose a novel statistical downscaling method to foster GCMs' precipitation prediction resolution and accuracy for the monsoon region. We develop a deep neural network composed of a convolution and Long Short Term Memory (LSTM) recurrent module to estimate precipitation based on well-resolved atmospheric dynamical fields. The proposed model is compared against the GCM precipitation product and classical downscaling methods in the Xiangjiang River Basin in South China. Results show considerable improvement compared to the European Centre for Medium-Range Weather Forecasts (ECMWF)-Interim reanalysis precipitation. Also, the model outperforms benchmark downscaling approaches, including (1) quantile mapping, (2) the support vector machine, and (3) the convolutional neural network. To test the robustness of the model and its applicability in practical forecasting, we apply the trained network for precipitation prediction forced by retrospective forecasts from the ECMWF model. Compared to the ECMWF precipitation forecast, our model makes better use of the resolved dynamical field for more accurate precipitation prediction at lead times from 1 day up to 2 weeks. This superiority decreases with the forecast lead time, as the GCM's skill in predicting atmospheric dynamics is diminished by the chaotic effect. Finally, we build a distributed hydrological model and force it with different sources of precipitation inputs. Hydrological simulation forced with the neural network precipitation estimation shows significant advantage over simulation forced with the original ERA-Interim precipitation (with NSE value increases from 0.06 to 0.64), and the performance is only slightly worse than the observed precipitation forced simulation (NSE = 0.82). This further proves the value of the proposed downscaling method, and suggests its potential for hydrological forecasts.
... We select the following SD models as baselines : (1) linear regression, (2) nearest neighbor, and (3) random forest, all of which have been extensively applied and verified for SD tasks (Gangopadhyay et al., 2005;Hutengs & Vohland, 2016;Li & Smith, 2009). For each of the model, we adopt same input variables as for CNN, with optional feature extractions before feeding the input to the model. ...
Article
Full-text available
Precipitation process is generally considered to be poorly represented in numerical weather/climate models. Statistical Downscaling (SD) methods, which relate precipitation with model resolved dynamics, often provide more accurate precipitation estimates compared to model’s raw precipitation products. We introduce the convolutional neural network (CNN) model to foster this aspect of SD for daily precipitation prediction. Specifically, we restrict the predictors to the variables that are directly resolved by discretizing the atmospheric dynamics equations. In this sense, our model works as an alternative to the existing precipitation-related parameterization schemes for numerical precipitation estimation. We train the model to learn precipitation-related dynamical features from the surrounding dynamical fields by optimizing a hierarchical set of spatial convolution kernels. We test the model at 14 geogrid points across the contiguous United States. Results show that provided with enough data, precipitation estimates from the CNN model outperform the reanalysis precipitation products, as well as SD products using linear regression, nearest neighbor, random forest, or fully-connected deep neural network. Evaluation for the test set suggests that the improvements can be seamlessly transferred to numerical weather modeling for improving precipitation prediction. Base on the default network, we examine the impact of the network architectures on model performance. Also, we offer simple visualization and analyzing approaches to interpret the models and their results. Our study contributes to the following two aspects: Firstly, we offer a novel approach to enhance numerical precipitation estimation; secondly, the proposed model provides important implications for improving precipitation-related parameterization schemes using a data-driven approach.
Article
We propose a statistical downscaling model based on multi-way functional principal component analysis (FPCA) for rainfall prediction. The model mainly explains the relationship between the winter mean sea level pressure (MSLP) and rainfall in southern Australia from the perspective of functional data. Compared with the traditional approach of feature extraction based on principal component analysis, the multi-way FPCA needs not only fewer principal components to capture most variance in MSLP, bus also greatly avoid the loss of spatial information. A functional principal component (FPC) regression is further developed to simulate both current and future rainfall. The main results show that the first five leading FPCs are sufficient to capture the spatial characteristics of winter MSLP, achieving the purpose of efficient dimensionality reduction. Specifically, no more than three FPCs are required to develop the functional dowscaling models for the winter rainfall over four studied regions. The functional downscaling model provides a good skill in terms of the correlation higher than 0.7 between the predictions and observations, and the ratio of root mean square error to the climatology of winter rainfall below 20% over four regions. The developed downscaling models are further used to interpret the MSLP patterns from four CMIP5 climate models (ACCESS1.3, BCC-CSM1.1-m, CESM1-CAM5 and MPI-ESM-MR), which have been used to simulate both present-day and future climate. The resulting downscaled values based on ensemble MSLP provides (1) a closer representation of observed present-day rainfall than the raw climate model values; (2) alternative estimates of future changes in rainfall that arises from changes in MSLP.
Article
Full-text available
The stepwise regression model (SRM) is a widely used statistical downscaling method that could be used to establish a statistical relationship between observed precipitation and predictors. However, the SRM cannot reflect the contributions of predictors to precipitation reasonably, which may not be the best model based on several possible competing predictors. Bayesian model averaging (BMA) is a standard inferencing approach that considers multiple competing statistical models. The BMA infers precipitation predictions by weighing individual predictors based on their probabilistic likelihood measures over the training period, with the better-performing predictions receiving higher weights than the worse-performing ones. Furthermore, the BMA provides a more reliable description of all the predictors than the SRM, leading to a sharper and better calibrated probability density function (PDF) for the probabilistic predictions. In this study, monthly precipitation at fifteen meteorological stations over the period of 1971–2012 in the Heihe River basin (HRB), which is located in an arid area of Northwest China, was simulated using the Bayesian model averaging (BMA) and the stepwise regression model (SRM), which was then compared with the observed datasets (OBS). The results showed that the BMA produced more accurate results than the SRM when used to statistically downscale large-scale variables. The multiyear mean precipitation results for twelve of the fifteen meteorological stations that were simulated by the BMA were better than those simulated by the SRM. The RMSE and MAE of the BMA for each station were lower than those of the SRM. The BMA had a lower mean RMSE (− 13.93%) and mean MAE (− 14.37%) compared with the SRM. The BMA could reduce the RMSEs and MAEs of precipitation and improve the correlation coefficient effectively. This indicates that the monthly precipitation simulated by the BMA has better consistency with the observed values.
Article
Full-text available
How much can we trust model-based projections of future anthropogenic climate change? This review attempts to give an overview of this important but difficult topic by using three main lines of evidence: the skill of models in simulating present-day climate, intermodel agreement on future climate changes, and the ability of models to simulate climate changes that have already occurred. A comparison of simulated and observed present-day climates shows good agreement for many basic variables, particularly at large horizontal scales, and a tendency for biases to vary in sign between different models, but there is a risk that these features might be partly a result of tuning. Overall, the connection between model skill in simulating present-day climate and the skill in simulating future climate changes is poorly known. An intercomparison of future climate changes between models shows a better agreement for changes in temperature than that for precipitation and sea level pressure, but some aspects of change in the latter two variables are also quite consistent between models. A comparison of simulations with observed climate changes is, in principle, a good test for the models, but there are several complications. Nonetheless, models have skilfully simulated many large-scale aspects of observed climate changes, including but not limited to the evolution of the global mean surface air temperature in the 20th century. Furthermore, although there is no detailed agreement between the simulated and observed geographical patterns of change, the grid box scale temperature, precipitation and pressure changes observed during the past half-century generally fall within the range of model results. Considering the difficulties associated with other sources of information, the variation of climate changes between different models is probably the most meaningful measure of uncertainty that is presently available. In general, however, this measure is more likely to underestimate than overestimate the actual uncertainty.
Article
How much can we trust model-based projections of future anthropogenic climate change? This review attempts to give an overview of this important but difficult topic by using three main lines of evidence: the skill of models in simulating present-day climate, intermodel agreement on future climate changes, and the ability of models to simulate climate changes that have already occurred. A comparison of simulated and observed present-day climates shows good agreement for many basic variables, particularly at large horizontal scales, and a tendency for biases to vary in sign between different models, but there is a risk that these features might be partly a result of tuning. Overall, the connection between model skill in simulating present-day climate and the skill in simulating future climate changes is poorly known. An intercomparison of future climate changes between models shows a better agreement for changes in temperature than that for precipitation and sea level pressure, but some aspects of change in the latter two variables are also quite consistent between models. A comparison of simulations with observed climate changes is, in principle, a good test for the models, but there are several complications. Nonetheless, models have skilfully simulated many large-scale aspects of observed climate changes, including but not limited to the evolution of the global mean surface air temperature in the 20th century. Furthermore, although there is no detailed agreement between the simulated and observed geographical patterns of change, the grid box scale temperature, precipitation and pressure changes observed during the past half-century generally fall within the range of model results. Considering the difficulties associated with other sources of information, the variation of climate changes between different models is probably the most meaningful measure of uncertainty that is presently available. In general, however, this measure is more likely to underestimate than overestimate the actual uncertainty.
Article
The dominant modes of low-frequency variation of the Southern Hemisphere (SH) circulation in winter have been identified using 15 years of monthly mean analyses of the SH troposphere. The two leading modes are primarily zonally-symmetric, representing out-of-phase variations of geopotential height between middle and high latitudes in one case and between the tropics and middle latitudes in the other. There are accompanying variations of zonal wind and temperature associated with these modes, which have an equivalent barotropic vertical structure in the extratropics. The two modes represent 25% of the variance of the monthly mean height in the SH. The high-latitude mode is also the dominant pattern of variation of 5-day average height anomalies, indicating that this is an important mode on time scales from about a week to a season. Using composites of monthly mean fields and transient eddy statistics for opposite extremes of this high latitude mode, the interaction between transient eddies and the anomalous mean flow is described. During one extreme of the variation, there is an increased height gradient, a stronger zonal jet and more baroclinicity at high latitudes. Although there may be some doubt about the reliability of the transient eddy statistics, they show an enhanced storm track at high latitudes, with transient eddy fluxes which help to maintain the anomalous mean flow. DOI: 10.1034/j.1600-0870.1990.00005.x
Article
The leading modes of variability of the extratropical circulation in both hemispheres are characterized by deep, zonally symmetric or "annular" structures, with geopotential height perturbations of opposing signs in the polar cap region and in the surrounding zonal ring centered near 45 latitude. The structure and dynamics of the Southern Hemisphere (SH) annular mode have been extensively documented, whereas the existence of a Northern Hemisphere (NH) mode, herein referred to as the Arctic Oscillation (AO), has only recently been recognized. Like the SH mode, the AO can be defined as the leading empirical orthogonal function of the sea level pressure field or of the zonally symmetric geopotential height or zonal wind fields. In this paper the structure and seasonality of the NH and SH modes are compared based on data from the National Centers for Environmental Prediction-National Center for Atmospheric Research reanalysis and supplementary datasets. The structures of the NH and SH annular modes are shown to be remarkably similar, not only in the zonally averaged geopotential height and zonal wind fields, but in the mean meridional circulations as well. Both exist year-round in the troposphere, but they amplify with height upward into the stratosphere during those seasons in which the strength of the zonal flow is conducive to strong planetary wave-mean flow interaction: midwinter in the NH and late spring in the SH. During these "active seasons," the annular modes modulate the strength of the Lagrangian mean circulation in the lower stratosphere, total column ozone and tropopause height over mid- and high latitudes, and the strength of the trade winds of their respective hemispheres. The NH mode also contains an embedded planetary wave signature with expressions in surface air temperature, precipitation, total column ozone, and tropopause height. It is argued that the horizontal temperature advcction by the perturbed zonal-mean zonal wind field in the lower troposphere is instrumental in forcing this pattern. A companion paper documents the striking resemblance between the structure of the annular modes and observed climate trends over the past few decades.
Chapter
Principal component analysis has often been dealt with in textbooks as a special case of factor analysis, and this tendency has been continued by many computer packages which treat PCA as one option in a program for factor analysis—see Appendix A2. This view is misguided since PCA and factor analysis, as usually defined, are really quite distinct techniques. The confusion may have arisen, in part, because of Hotelling’s (1933) original paper, in which principal components were introduced in the context of providing a small number of ‘more fundamental’ variables which determine the values of the p original variables. This is very much in the spirit of the factor model introduced in Section 7.1, although Girschick (1936) indicates that there were soon criticisms of Hotelling’s method of PCs, as being inappropriate for factor analysis. Further confusion results from the fact that practitioners of ‘factor analysis’ do not always have the same definition of the technique (see Jackson, 1981). The definition adopted in this chapter is, however, fairly standard.