a synoptic classification of inflow-generating precipitation ... · a synoptic classification of...

20
A Synoptic Classification of Inflow-Generating Precipitation in the Snowy Mountains, Australia ALISON THEOBALD AND HAMISH MCGOWAN Climate Research Group, School of Geography, Planning and Environmental Management, University of Queensland, St Lucia, Queensland, Australia JOHANNA SPEIRS Snowy Hydro, Ltd., Sydney, New South Wales, Australia NIK CALLOW Environmental Dynamics and Ecohydrology, School of Earth and Environment, University of Western Australia, Crawley, Western Australia, Australia (Manuscript received 24 October 2014, in final form 23 April 2015) ABSTRACT Precipitation falling in the Snowy Mountains region of southeastern Australia provides fuel for hydroelectric power generation and environmental flows along major river systems, as well as critical water resources for agricultural irrigation. A synoptic climatology of daily precipitation that triggers a quantifiable increase in streamflow in the headwater catchments of the Snowy Mountains region is presented for the period 1958–2012. Here, previous synoptic-meteorological studies of the region are extended by using a longer-term, year-round precipitation and reanalysis dataset combined with a novel, automated synoptic-classification technique. A three-dimensional representation of synoptic circulation is developed by effectively combining meteorological variables through the depth of the troposphere. Eleven distinct synoptic types are identified, describing key circulation features and moisture pathways that deliver precipitation to the Snowy Mountains. Synoptic types with the highest precipitation totals are commonly associated with moisture pathways originating from the northeast and northwest of Australia. These systems generate the greatest precipitation totals across the westerly and high-elevation areas of the Snowy Mountains, but precipitation is reduced in the eastern-elevation areas in the lee of the mountain ranges. In eastern regions, synoptic types with onshore transport of humid air from the Tasman Sea are the major source of precipitation. Strong seasonality in synoptic types is evident, with frontal and cutoff-low types dominating in winter and inland heat troughs prevailing in summer. Interaction between tropical and extratropical systems is evident in all seasons. 1. Introduction Inflows generated from precipitation falling in the Snowy Mountains region provide vital water sources for irrigation and environmental flows in the agriculturally important and ecologically diverse Murray–Darling ba- sin, as well as fuel for hydroelectric power generation. Located in southeastern Australia (SEA), the Snowy Mountains are one of only a few alpine regions in Aus- tralia. They form the highest part of the Great Dividing Range and include Australia’s highest peak—Mount Kosciusko—at 2228 m (Fig. 1). In contrast to younger mountain ranges such as the European Alps, the topog- raphy is not as steep, rugged, or high in elevation; instead, areas of undulating tablelands dominate the region. In response to a series of severe droughts in the region, construction began on the Snowy Mountains Hydro- Electric Scheme (‘‘Scheme’’ hereinafter) in 1949. The Scheme consists of a complex network of dams, hydro- electric power stations, tunnels, aqueducts, and pipelines that are able to divert eastward-flowing rivers under the mountain range inland to the Murray and Murrumbidgee Corresponding author address: Alison Theobald, School of Geography, Planning and Environmental Management, Level 4, Chamberlain Bldg., University of Queensland, St Lucia, Queens- land, 4072, Australia. E-mail: [email protected] AUGUST 2015 THEOBALD ET AL. 1713 DOI: 10.1175/JAMC-D-14-0278.1 Ó 2015 American Meteorological Society

Upload: others

Post on 24-Jan-2020

8 views

Category:

Documents


0 download

TRANSCRIPT

A Synoptic Classification of Inflow-Generating Precipitation in theSnowy Mountains, Australia

ALISON THEOBALD AND HAMISH MCGOWAN

Climate Research Group, School of Geography, Planning and Environmental Management, University of

Queensland, St Lucia, Queensland, Australia

JOHANNA SPEIRS

Snowy Hydro, Ltd., Sydney, New South Wales, Australia

NIK CALLOW

Environmental Dynamics and Ecohydrology, School of Earth and Environment, University of

Western Australia, Crawley, Western Australia, Australia

(Manuscript received 24 October 2014, in final form 23 April 2015)

ABSTRACT

Precipitation falling in the SnowyMountains region of southeastern Australia provides fuel for hydroelectric

power generation and environmental flows along major river systems, as well as critical water resources for

agricultural irrigation. A synoptic climatology of daily precipitation that triggers a quantifiable increase in

streamflow in the headwater catchments of the SnowyMountains region is presented for the period 1958–2012.

Here, previous synoptic-meteorological studies of the region are extended by using a longer-term, year-round

precipitation and reanalysis dataset combined with a novel, automated synoptic-classification technique. A

three-dimensional representation of synoptic circulation is developed by effectively combining meteorological

variables through the depth of the troposphere. Eleven distinct synoptic types are identified, describing key

circulation features and moisture pathways that deliver precipitation to the Snowy Mountains. Synoptic types

with the highest precipitation totals are commonly associated with moisture pathways originating from the

northeast and northwest of Australia. These systems generate the greatest precipitation totals across the

westerly and high-elevation areas of the SnowyMountains, but precipitation is reduced in the eastern-elevation

areas in the lee of the mountain ranges. In eastern regions, synoptic types with onshore transport of humid air

from the Tasman Sea are the major source of precipitation. Strong seasonality in synoptic types is evident, with

frontal and cutoff-low types dominating in winter and inland heat troughs prevailing in summer. Interaction

between tropical and extratropical systems is evident in all seasons.

1. Introduction

Inflows generated from precipitation falling in the

Snowy Mountains region provide vital water sources for

irrigation and environmental flows in the agriculturally

important and ecologically diverse Murray–Darling ba-

sin, as well as fuel for hydroelectric power generation.

Located in southeastern Australia (SEA), the Snowy

Mountains are one of only a few alpine regions in Aus-

tralia. They form the highest part of the Great Dividing

Range and include Australia’s highest peak—Mount

Kosciusko—at 2228m (Fig. 1). In contrast to younger

mountain ranges such as the European Alps, the topog-

raphy is not as steep, rugged, or high in elevation; instead,

areas of undulating tablelands dominate the region.

In response to a series of severe droughts in the region,

construction began on the Snowy Mountains Hydro-

Electric Scheme (‘‘Scheme’’ hereinafter) in 1949. The

Scheme consists of a complex network of dams, hydro-

electric power stations, tunnels, aqueducts, and pipelines

that are able to divert eastward-flowing rivers under the

mountain range inland to the Murray and Murrumbidgee

Corresponding author address: Alison Theobald, School of

Geography, Planning and Environmental Management, Level 4,

Chamberlain Bldg., University of Queensland, St Lucia, Queens-

land, 4072, Australia.

E-mail: [email protected]

AUGUST 2015 THEOBALD ET AL . 1713

DOI: 10.1175/JAMC-D-14-0278.1

� 2015 American Meteorological Society

Rivers. The Scheme provides on average 2360Gl of water

per year for irrigation, underwriting AUD $3 billion of

agricultural production in the Murray–Darling basin.

Furthermore, the water provides environmental flows

along major rivers and offers a degree of flow regulation

(Snowy Hydro Limited 2003; Ghassemi and White

2007; see also http://www.mdba.gov.au/about-basin/

how-river-runs/murrumbidgee-catchment). The Scheme’s

hydroelectric power generation meets the peak power

demand for much of eastern Australia and currently

provides 32% of all renewable annual energy produc-

tion in Australia (http://www.snowyhydro.com.au/energy/

hydro/snowy-mountains-scheme).

The SnowyMountains are typically one of Australia’s

wettest regions, but the precipitation is highly variable.

Annual precipitation from days with $10mm of

precipitation in high-elevation regions of the Snowy

Mountains varied between 2800 and 760mm between

1958 and 2012 (Fig. 2). High precipitation variability

also exists on intra-annual time scales. Precipitation in

winter and spring is heavily influenced by the prevailing

midlatitude westerly winds in conjunction with oro-

graphic enhancement, and these are commonly consid-

ered to be the wettest seasons (Snowy Hydro Limited

2003; Pook et al. 2006; Cai and Cowan 2008;

Ummenhofer et al. 2009; Chubb et al. 2011). High-

intensity and warm spring rains falling onto the snow-

pack are a major source of inflows (McGowan et al.

2009), generating as much as 50% of the total annual

inflows to the hydroelectric catchments (Snowy Hydro

Limited 2003). Although the warmer months are gen-

erally drier, heavy rainfall can result from trajectories of

warm, moist air from lower latitudes and small-scale

convective events that generate thunderstorms (Barry

1992; Basist et al. 1994; Snowy Hydro Limited 2003).

Orographic enhancement of precipitation also occurs in

summer as north and northwesterly winds flow perpen-

dicular to the Snowy Mountains (Chubb et al. 2011).

The SnowyMountains region lies toward the northern

limits of the influence of the midlatitude westerly wind

belt, where interaction between tropical and extra-

tropical weather systems is an important factor in the

generation of precipitation (Wright 1989). Accordingly,

this region is sensitive to changes in the annual cycle of

the subtropical ridge (STR) and shifts in the westerly

storm tracks (Drosdowsky 2005; Murphy and Timbal

2008; Verdon-Kidd et al. 2013; Timbal and Drosdowsky

2013), as well as hydrological-cycle changes in a warm-

ing climate that can dramatically affect seasonal pre-

cipitation and runoff (Beniston 2003; Viviroli et al.

2011). In addition, the ascent mechanisms that are re-

sponsible for providing deep convective moisture and

thus the potential for higher precipitation totals in the

Snowy Mountains vary seasonally.

Historic snow-course data have shown significant de-

clines in snow depth (Nicholls 2005; Hennessy et al.

FIG. 1. (a) Location of the Snowy Mountains region within Australia (red box) and the synoptic-analysis region

(dashed line). (b) SnowyMountains region showing the water catchment (solid black line) and precipitation gauges

used in this study (red squares indicate gauges on the western elevations, blue diamonds indicate gauges on the high

elevations, green triangles indicate gauges on the eastern slopes, and black dots indicate the gauges from the SILO

dataset). Inflow gauges at Snowy River and Yarrangobilly River are marked with asterisks.

1714 JOURNAL OF APPL IED METEOROLOGY AND CL IMATOLOGY VOLUME 54

2008) in association with rising alpine temperatures.

Climate-change modeling of snow depth by Hennessy

et al. (2003, 2008) and Hendrikx et al. (2013) shows a

continuation of these declines. Regional-scale hydro-

climaticmodeling predicts rainfall and runoff to decrease

on average and to increase in extremity over coming

decades (Chiew et al. 2011). Meanwhile, demand for

both water and energy are forecast to increase into the

future (Christensen et al. 2007) resulting in increased

vulnerability of water resources that are already under

stress (Viviroli et al. 2011). Despite the importance of

inflow-generating precipitation in the SnowyMountains,

there remains a knowledge gap regarding the long-term,

historic climatological behavior of the synoptic weather

systems that deliver precipitation to the region.

The cool-season synoptic circulation over various re-

gions of SEA has been well studied (e.g., Wright 1989;

Pook et al. 2006; Landvogt et al. 2008; Risbey et al. 2009;

Gallant et al. 2012), although few studies relate specifi-

cally to the Snowy Mountains area (Colquhoun 1978;

Chubb et al. 2011; Fiddes et al. 2015). These studies

commonly report cool-season precipitation declines.

From these cool-season-focused studies, it is widely re-

ported that cold fronts and closed lows, including cutoff

lows, are responsible for the majority of wintertime

precipitation. These types of weather systems can occur

year-round (Landvogt et al. 2008; Wright 1989), how-

ever, and are also important for inflow generation out-

side the winter season, particularly when they interact

with tropical systems. For instance, the highest 7-day

accumulated rainfall total in the Snowy Mountains re-

gion fell between 27 February and 4March 2012, causing

widespread flooding (Bureau of Meteorology 2012).

Strengthening of the East Australia Current and asso-

ciated Tasman Sea warming due to climate change

suggest increasing summer precipitation in SEA (Cai

et al. 2005; Shi et al. 2008; Gallant et al. 2012).

Previous studies of synoptic circulation over SEA

have predominantly used manual-classification schemes,

which are, by their nature, subjective and limited in the

number of meteorological variables on which they are

based. The majority consider only the cool-season period.

Manual approaches are considered to be time-consuming,

and, with the exception of Pook et al. (2014), the majority

of SEA studies have covered relatively short time periods

of a few decades at most. Several SEA studies have fo-

cused particularly on the period of extended drought that

persisted for much of the 1990s and 2000s in an attempt to

understand the significant precipitation decline that oc-

curred during this period (e.g., Risbey et al. 2013). Studies

confined to shorter periodsmay not be fully representative

of synoptic circulation over a multidecadal period.

McGowan et al. (2009) identified the role that multi-

decadal ocean–atmosphere teleconnections may play in

the hydroclimate of the Snowy Mountains region. Ac-

cordingly, there is a need to extend current understanding

of synoptic circulation beyond the last few decades and to

encompass year-round synoptic systems.

Previous manual classifications of precipitation sys-

tems worldwide and inAustralia focusmainly on surface

pressure fields with limited analyses of mid- and upper-

level atmospheric properties. Between three and five

synoptic types were identified for SEA (Wright 1989;

Pook et al. 2006; Landvogt et al. 2008; Chubb et al. 2011;

Risbey et al. 2013). Studies that have applied automated

techniques have been based on a single atmospheric

FIG. 2. (top) Annual precipitation across the western (plus signs), high (squares), and eastern

(asterisks) elevations and (bottom) annual number of precipitation days $ 10mm for the pe-

riod 1958–2012.

AUGUST 2015 THEOBALD ET AL . 1715

variable, usually mean sea level pressure (MSLP;

Whetton 1988; Kidson 2000; Jiang 2011; Renwick 2011),

because of difficulties encountered in combining multi-

level data (Kidson 2000). Multivariable classifications

either classify each variable separately (e.g., Bettolli

et al. 2010) or use data-reduction techniques (e.g., Stahl

et al. 2006; Moron et al. 2008). Self-organizing maps that

are based on MSLP (Cassano and Cassano 2010) or

500-hPa geopotential height (Newton et al. 2014a,b)

were employed in Canadian studies to assess catchment-

scale hydroclimatic variability. These approaches, as

noted by Stahl et al. (2006), conceal the complex three-

dimensional nature and internal variability of synoptic

types. Consequently, causes of variability in seasonal

rainfall that depend not only on the annual cycle of

surface pressure, but also on seasonal changes to the

atmospheric circulation in the mid- and upper tropo-

sphere (Pook et al. 2006), may not be identified.

Here we present a seasonal, 55-yr (1958–2012) synoptic

climatology of inflow-generating precipitation days for

the Snowy Mountains region. A novel, objective method

is developed that describes synoptic types on the basis of a

suite of 21 meteorological variables throughout the depth

of the troposphere. Use of reanalysis data and an auto-

mated approach allows a multidecadal time period to be

investigated and removes much of the subjectivity of

manual classifications (Yarnal 1993). Important is that

this approach allows robust conclusions to be drawn with

regard to patterns of synoptic circulation that are re-

sponsible for inflow-generating precipitation, enabling

ongoing research to investigate trends in synoptic circu-

lation variability and their significance for local to re-

gional hydroclimate. Such knowledge is essential to better

understand the drivers of variability in historical records

of precipitation so as to make better-informed water-

management decisions (Viviroli et al. 2011).

A description of the data andmethods used is outlined

in section 2. Results of the synoptic climatology and the

variability of the synoptic weather types are presented in

section 3. A discussion and conclusions are presented in

section 4.

2. Data and methods

a. Data

A network of private, tipping- and weighing-bucket

precipitation gauges operated by Snowy Hydro Limited

(SHL) and the Queensland government’s Scientific In-

formation for Land Owners (SILO) ‘‘Patched Point

Dataset’’ (https://www.longpaddock.qld.gov.au/silo/ppd/

index.php?reset5reset, accessed January 2014; Jeffrey

et al. 2001) from the Australian Bureau of Meteorology

(BoM) recording stations provide the daily precipitation

observations used in this study (Fig. 1b). Some records in

the SHL dataset begin in the 1950s, but many of the early

data records (before 1975) are discontinuous. Installation

of an automatic weather station within the study area in

1996 and a substantial increase in precipitation gauge

network size after 2004 vastly improved data quality.

Newer, heated and weighing gauges at higher elevations,

some with wind shields, more accurately record snowfall.

Records from 56 SHL active gauges within the Snowy

Mountains have been used in this study, which covers the

period of 1958–2012. Data are recorded instantaneously

and were aggregated to daily totals to 0900 local time (in

line with the BoM convention for daily precipitation ob-

servations). SHL data from two inflow-recording stations,

Yarrangobilly River at Ravine station and the Snowy

River above Guthega station (Fig. 1b), were also aggre-

gated to daily totals.

Data from 10 gauges within the Snowy Mountains

were selected from the SILO dataset (Fig. 1b) and used

to ensure a continuous daily precipitation record. The

SILOdataset consists of observations fromBoMgauges,

filled in with interpolated values where observations are

unavailable. Rainfall is interpolated using the geo-

statistical method of ordinary kriging (Jeffrey et al.

2001). An interpolated surface (0.058 3 0.058 grid reso-

lution) is produced for each day, from which missing

data values for point locations are extracted from the

nearest grid cell. The provision of daily data for in-

dividual stations rather than a gridded product allows

direct insertion of data (Pook et al. 2010) into disconti-

nuities in the SHL record. For any point location, the

nearest 30 stations and all within 100 km are used for the

interpolation, whichever is greater. It has been shown

that normalized precipitation removes the component

of precipitation variability that is due to topographic

influences and can be reliably interpolated at time scales

from hourly to monthly (Jeffrey et al. 2001). Topo-

graphic effects are subsequently accounted for by de-

normalizing the interpolated surface to derive the final

rainfall surface. The SILO dataset is available for the

period from 1889 to present. Precipitation data from

all seasons were considered. The standard climatologi-

cal seasons have been used throughout this study:

December–February (DJF), March–May (MAM),

June–August (JJA), and September–November (SON).

For the purpose of producing a synoptic climatology,

the European Centre for Medium-Range Weather

Forecasts (ECMWF) ERA-Interim and ERA-40 re-

analysis products (Dee et al. 2011) were used as input to a

clustering algorithm and to construct composite maps of

the synoptic atmospheric circulation associated with

precipitation days. ERA-Interim spans the period from

1979 to near–real time and is available at a default 0.758

1716 JOURNAL OF APPL IED METEOROLOGY AND CL IMATOLOGY VOLUME 54

latitude3 0.758 longitude grid resolution. ERA-40 spans

the period 1958–2002 and has a default resolution of

2.58 3 2.58, but a 0.758 3 0.758 resolution can also be

obtained and is used here. Reanalysis products at

coarser scales have been shown in previous studies to be

unsuitable for regional climate assessments (Eichler and

Gottschalck 2013) and to have difficulty accurately de-

tecting features such as surface pressure fronts and

troughs. As a result, past synoptic studies have needed to

supplement coarse-resolution reanalysis data with sat-

ellite imagery and manual-analysis charts (e.g., Pook

et al. 2006). Furthermore, ECMWF reanalyses have

improved representation of Southern Hemisphere high-

latitude atmospheric circulation when compared with

other reanalysis products (Marshall 2003).

ERA-Interim represents the latest reanalysis product

(at the time of writing) fromECMWFand has addressed

several data-assimilation issues that were encountered

in ERA-40 (Dee et al. 2011). Comparison of clustering

results from the two products for an overlapping 22-yr

period (1979–2001) demonstrated no significant differ-

ence in the output (not shown). This result is in agree-

ment with Hoskins and Hodges (2005) who evaluated

the impact of changes to observing systems in their

analysis of Southern Hemisphere storm tracks. They

concluded that their climatological description re-

mained robust between different reanalyses and the pre-

and postsatellite eras.

Daily mean values of the following variables were

used in this study: MSLP; 500-hPa geopotential height;

wind vectors at the surface, 850, 700, 500, and 250 hPa;

relative humidity and temperature at 850, 700, 500, and

250hPa; and (the computed) 1000–500-hPa atmospheric

thickness. The subtropical jet stream (STJ), the strength

and position of which over Australia is known to in-

fluence steering and development of synoptic systems,

was calculated as the magnitude of the wind vector at

250hPa (Risbey et al. 2009). For our purposes, tem-

perature and humidity profiles are standard airmass in-

dicators of atmospheric stability and available moisture,

respectively (Davis and Kalkstein 1990). Wind vectors

provide critical information on direction of moisture

transport and steering of synoptic systems (Pook et al.

2006). Mean sea level pressure provides an indicator of

the large-scale current atmospheric state (Eder et al.

1994) while thickness provides information on advection

and frontal position (e.g., Pook et al. 2006). Variables

were obtained for the synoptic analysis area bounded by

latitudes 208–468S and longitudes 1208–1608E (Fig. 1a).

This region is considered to be extensive enough to

capture all synoptic weather systems affecting the

Snowy Mountains, including those originating in, and

interacting with, tropical latitudes.

b. Methods

1) QUALITY CONTROL

For the purposes of this study, all data were subject to

quality control and any data flagged as unsuitable for

climatological purposes were automatically removed.

Data were checked on a gauge-by-gauge basis for

anomalous values by calculating the maximum, mini-

mum, and range. Furthermore, any data with a quality

flag indicating potentially bad data were subject to ad-

ditional quality control and in a few cases were removed

because of anomalously high half-hourly values. In ad-

dition, SHL data coded as ‘‘good’’ but with a zero

amount where the corresponding SILO dataset showed a

value greater than zero were disregarded when calculat-

ing the daily mean values.

2) PRECIPITATION THRESHOLD

This study investigates days on which precipitation

generates inflow to reservoirs within the Snowy Moun-

tains water catchment (Fig. 1b). To define a threshold

amount for a precipitation day, a Lyne and Hollick filter

(Lyne and Hollick 1979) was applied to separate

‘‘quickflow’’ (surface runoff) from base flow in the in-

flow dataset (Nathan and McMahon 1990) at the Snowy

River and Yarrangobilly River stations. As alpine and

subalpine sites, respectively, these two stations provide

indicative conditions across the region. Precipitation

from the most relevant gauges to the inflow stations was

then correlated with quickflow at a lag time of 11 day

(based on prior knowledge of the behavior of pre-

cipitation and runoff in these catchments). The appli-

cation of the Lyne and Hollick filter to inflow data

resulted in a threshold precipitation amount of 10mm,

above which quickflow and therefore inflow was en-

hanced (not shown).

Only precipitation during the period from December

to April was considered for the determination of this

threshold. During the cool season, precipitation can be

stored as snowpack or can enhance inflow during spring

snowmelt, and, particularly during this period, a static

precipitation–runoff threshold does not apply. It is ac-

knowledged that such a precipitation–runoff threshold

is typically dynamic and depends upon numerous

catchment conditions but is necessary for the purpose of

synoptic classification of precipitation days in this study.

3) GROUPING DATA

Different precipitation regimes are experienced

across the Snowy Mountains region, in part because of

topographic interaction with the prevailing flow (Chubb

et al. 2011; Fiddes et al. 2015). Following Chubb et al.

AUGUST 2015 THEOBALD ET AL . 1717

(2011), the 56 SHL gauges were divided into western-,

high-, and eastern-elevation groups (Fig. 1b) on the basis

of the similarity of the statistics shown in Table 1. Al-

though the high-elevation group contains a greater

number of gauges, most of these began operating after

2004. Prior to 2004, the number of gauges was more

similar across all groups (Table 1). Such regionalization

of data has been used successfully in previous studies

in a variety of locations worldwide, helping to reduce

noise and spatial autocorrelation in the data (Whetton

1988; Widmann and Schär 1997; Kidson 2000; Chubb

et al. 2011; Plavcova et al. 2014).

A continuous daily precipitation record back to 1958

was not possible using only the SHL data; therefore,

comparisons between the SHL and SILO datasets were

carried out to determine the suitability, and associated

uncertainty, in combining the two datasets. Correla-

tions, contingency tables, and their associated skill-score

statistics [bias, probability of detection (POD), and

mean absolute error (MAE)] were calculated (Wilks

2006; Beesley et al. 2009; Tozer et al. 2012). Data from

all 10 SILO gauges within the watershed boundary

provided the best comparison with the most recent SHL

data (which include high-resolution, heated, and fenced

gauges), suggesting that using all gauges in the SILO

dataset back to 1958 provides the most robust estimate

of precipitation.

Contingency-table statistics were better overall for

gauges within the high-elevation group, with POD

scores indicating that at least 78%–84% of precipitation

days over 10mm are accurately detected by SILO. The

MAE between the SILO and SHL datasets varies be-

tween 0.03mm (high elevations) and 3.66mm (eastern

elevations), in good agreement with Jeffrey et al. (2001)

for the study region. SILO estimates of precipitation do

not demonstrate a forecast bias for the high elevations,

but there is a tendency to underforecast (overforecast)

precipitation days after (before) 2006 in the western and

eastern elevations. Lower skill scores in the western and

eastern elevations are likely due to fewer gauges in these

areas. Particularly in the east, the only two gauges within

the SILO dataset are located in the southern part of the

catchment. Lack of representation in the northern

catchment may explain the lower skill there. Despite the

lower skill scores for the eastern and western elevations,

only 4.5%–4.9% of days are detected solely in

these groups.

A daily precipitation amount for each group was cal-

culated as the mean daily precipitation from each gauge

within the group (Chubb et al. 2011; Dai et al. 2014;

Fiddes et al. 2015). Following WMO guidelines and

Chubb et al. (2011), a minimum of four data values were

used to calculate a mean (WMO 2011). When this was

unachievable using only SHL data, the SILO data were

included in the calculation. The change in the number

of SHL gauges, particularly within the high-elevation

group, was shown to have minimal effect on the calcula-

tion of the mean, with an average difference of 0.11mm

between mean values. Prior to the commencement of

data recording in the western (eastern) elevations by the

Snowy Hydro Scheme in 1961 (1960), daily totals were

calculated solely from the SILO dataset.

Each day with a precipitation total greater than the

threshold amount of 10mmwas extracted from the daily

record. The corresponding ECMWF reanalysis vari-

ables for each of those days were collated into a data

matrix and standardized on a monthly basis (over the

full study period), using z scores, to account for the in-

consistent units (Yarnal 1993; Hart et al. 2006; Wilson

et al. 2013; Gao et al. 2014). Standardizing climate data

has been demonstrated to have the additional benefit of

removing seasonal variations in intensity of synoptic

systems, allowing comparison between data with dif-

ferent variabilities and means (Gao et al. 2014; Jiang

2011;Wilks 2006; Kidson 2000; Yarnal 1993). The strong

influence of the seasonal cycle of the STR on weather

patterns in SEA means clustering applied to standard-

ized data better reflects the daily changes in atmospheric

circulation (Yarnal 1993). For example, Pook et al.

(2006, their Fig. 3) show the variability in monthly mean

MSLP across the cool season. In our study, the monthly

standardized MSLP data reveal that precipitation days

are always associated with negative anomalies, which

become stronger during the cooler months and are at a

maximum in May (not shown). Similar month-to-month

variations are apparent in the other variables used in this

study (not shown for brevity). The standardized data

matrix was used as input into the cluster analysis.

TABLE 1. Statistics of grouped precipitation gauges.

Group

Avg

elev (m)

Avg annual

precipitation (mm)

Correlation

between gauges

No. gauges

prior to 2004

No. gauges

after 2004

Western 391 1246 .0.78 2 5

High 1484 1602 .0.74 8 45

Eastern 1107 857 .0.70 3 6

1718 JOURNAL OF APPL IED METEOROLOGY AND CL IMATOLOGY VOLUME 54

4) CLUSTER ANALYSIS

To generate synoptic types for the period 1958–2012,

cluster analysis was applied to the standardized meteo-

rological variables for each daily precipitation total

greater than the threshold amount of 10mm. The cluster

analysis was performed across all seasons simulta-

neously, given the prior removal of the seasonal cycle by

data standardization. Equal weighting was given to each

variable, given the importance of each in influencing the

synoptic circulation and precipitation received, as out-

lined in section 2a. Hierarchical average-linkage clus-

tering gave an initial indication of the number of groups

contained in the data (Wilks 2006; Hart et al. 2006;

Trauth 2007). The k-means clustering method of Wilson

et al. (2013), with a city-block distance measure, was

then applied to the standardized variables to assign each

precipitation day to a synoptic type (Hart et al. 2006;

Wilson et al. 2013). The iterative nature of the k-means

technique refines the clusters by reclassifying days until

the smallest within-cluster solution is found and days

with similar meteorological characteristics are classified

in the same cluster (Hart et al. 2006). The algorithm was

tested for a range of clusters between 2 and 20. Exami-

nation of a plot of the distance measure against number

of clusters for the point at which the line flattens out, and

after which distance increases again, commonly gives an

indication of the optimum number of clusters (Wilks

2006; Tan et al. 2006; Wilson et al. 2013). This was used

in conjunction with physical interpretation of composite

maps (generated from the mean value of all days as-

signed to each cluster; Kalkstein et al. 1987). Thek-means

technique was initialized several times using random

subsets of the data as cluster seed values. The iteration for

which the sum of distances was smallest was then used as

the cluster seeds for the full dataset.

Comparison of clustering results between the ERA-

Interim andERA-40 reanalyses for an overlapping 22-yr

period produced no significant differences in the

resulting synoptic types, with a hit rate of .80% in the

assignment of individual days to the same synoptic type.

This is in agreement with Hoskins and Hodges (2005)

whose comprehensive comparison of synoptic clima-

tologies remained robust between different reanalyses

products and pre- and postsatellite eras.

The automated clustering procedure was validated by

comparison with amanual classification for a 5-yr period

(2008–12) and nonparametric hypothesis testing of the

precipitation assigned to each cluster. Surface and 500-hPa

height charts, readily available from BoM (http://

www.bom.gov.au/australia/charts/archive/), and NOAA

satellite imagery (http://www.ncdc.noaa.gov/gibbs/year)

were used to classify each day on the basis of identification

of key surface and upper-air features (Davis andKalkstein

1990; Yarnal 1993), and the presence of cloud bands.

Classification on the basis of these variables revealed

several days that could have been placed into more than

one cluster. Further inspection of additional reanalysis-

generated variables (temperature, humidity, and wind

vectors) showed that most days could belong to only one

particular cluster. Overall, only 8% of days were moved

into a different cluster on the basis of the manual anal-

ysis, from that generated by the k-means algorithm.

3. Results and analysis

a. Synoptic classification of precipitation-bearingsystems

The application of the threshold precipitation

amount to daily precipitation for the period of 1958–

2012 resulted in 3443 days being identified with inflow-

generating precipitation and requiring synoptic clas-

sification. A day was classified if at least 10mm of

precipitation was recorded in the western, high, or

eastern group. These specific days account for almost

40% of all precipitation days, that is, those for which

precipitation $ 1mm is recorded. It is acknowledged

that this definition of a precipitation day does not ac-

count for multiday precipitation events, in which a

series of synoptic types may traverse the region as a

precipitation-bearing weather system evolves through

time. Instead, each individual day is assigned to a

specific synoptic type, following Pook et al. (2006) and

Risbey et al. (2009). Figure 2 shows the annual pre-

cipitation across the different elevations (Fig. 2a) and

the annual number of precipitation days of at least

10mm (Fig. 2b) between 1958 and 2012, demonstrat-

ing the high degree of interannual variability in the

precipitation of the Snowy Mountains region. A sta-

tistically significant trend in precipitation of 138mm

(10 yr)21 in the eastern elevations is apparent, but

western and high elevations and annual precipitation

days exhibit nonsignificant decreases.

A two-sided Wilcoxon–Mann–Whitney test demon-

strated that the median rainfall amount in correspond-

ing clusters of the manual and automated classification

was equal at a 95% confidence level (p , 0.05). Fur-

thermore, the manual classification confirmed that the

automated scheme was capable of detecting expected

synoptic patterns. The types were not as clearly defined

as in previous manual-classification studies (e.g., Pook

et al. 2006; Chubb et al. 2011), however, likely because

of the greater number of variables being considered and

thus the greater possible combinations of variables as

well as the multitype nature of some synoptic systems.

AUGUST 2015 THEOBALD ET AL . 1719

The k-means clustering method, applied to the daily,

standardized reanalysis data for a range of cluster

numbers k, suggested the optimum number of clusters to

be 10 or 11 (Fig. 3). Comparison of composite maps for

each of these solutions further suggested that 11 syn-

optic types better represented known synoptic systems

and reinforced initial hierarchical clustering results. A

Wilcoxon–Mann–Whitney test determined that median

precipitation between types was significantly different,

rejecting the null hypothesis that all medians were equal

(p , 0.05), for 80% of all possible combinations (con-

sidering precipitation across all areas of the catchment).

Given that only those days experiencing precipitation$

10mm have been classified, rather than rain versus no-

rain days, and the natural variability in precipitation

amount from individual occurrences of the same type, it

is more likely that the null hypothesis will not be re-

jected in all cases. This result, along with physical in-

terpretation of composite maps for each reanalysis

parameter (Kalkstein et al. 1987; Wilson et al. 2013),

demonstrates that each of the 11 clusters represents

specific synoptic types.

Composite charts showing average meteorological

conditions for a number of key parameters for each of

the 11 synoptic types are presented in Figs. 4–9. In ad-

dition to the clustering parameters, an analysis of co-

lumnar precipitable water (PW) and relative vorticity

FIG. 3. Within-cluster sum of distances. Similar to a scree plot,

the point at which the plot flattens out (shown by the circled area)

indicates that there are 11 clusters in the data.

FIG. 4. Compositemaps for each synoptic type showingMSLP (colored contours; hPa) and 500-hPa height (contour lines; m). Longitude

and latitude are displayed on the x and y axes, respectively. The area represented in these compositemaps is the synoptic-analysis region as

shown in Fig. 1a.

1720 JOURNAL OF APPL IED METEOROLOGY AND CL IMATOLOGY VOLUME 54

for each synoptic type was conducted, and it is shown in

Fig. 10. These additional parameters give further in-

formation on available moisture and system develop-

ment, respectively. Some between-type similarities exist

in individual variables for a given level (e.g., MSLP), but

when all variables are considered together each type

has distinct characteristics and distinguishing features.

Table 2 summarizes the three-dimensional characteris-

tics, ascent mechanisms, andmoisture pathways for each

synoptic type. The frequency of occurrence of each

synoptic type and the resulting precipitation contribu-

tions in each elevation group are presented in Table 3.

The synoptic classification reveals that 8 of the 11

types (all except T1, T5, and T9) represent atmospheric

circulation patterns with a connection to tropical lati-

tudes, in particular where a north or northwesterly air-

flow (specifically between 700 and 500 hPa; Figs. 7–9)

advects a conveyor of warm, moist air originating from

the warm oceans surrounding tropical Australia toward

the Snowy Mountains (Figs. 6–8). These tropical-

connected systems deliver over 70% of total precipita-

tion greater than 10mm across the whole catchment

(Table 3). Three of the tropical-connected synoptic

types—T8, T11, and T4—display synoptic circulation

conducive to northwest cloud bands (NWCBs). For T4,

however, NWCBs are detected on only ;10% of oc-

currences, when conditionsmatch those inWright (1989,

their Fig. 3). NWCBs form over the warm surface waters

to the northwest of Australia, where deep convection

feeds moisture from the tropical Indian Ocean along the

cloud band to southeastern Australia (Tapp and Barrell

1984; Sturman and Tapper 2006). This circulation, evi-

dent in Figs. 4 and 5 for T8 (and to a lesser extent T4 and

T11), features moisture aligned with the core of the STJ

and its region of maximum intensity [as shown in Tapp

and Barrell (1984)]. Together these three synoptic types

account for 26% of all days over 10mm. In a similar

way, and as confirmed by the manual classification,

T10-classified synoptic types have circulation that is

conducive to cloud bands extending northward along

the east coast, commonly seen as an easterly dip and

cloud-band pattern (Gallant et al. 2012; Fiddes et al.

2015). Downstream anticyclones and ridging—apparent

in T3, T4, T6, T7, T8, T10, and T11—contribute to

tropical moisture transport and enhancement of warm

air advection (WAA; Fig. 6), which, combined with an

anticlockwise rotation of winds with height (‘‘backing’’),

signifies forced ascent (Figs. 7–9). With the exception of

FIG. 5. As in Fig. 4, but showing the subtropical jet stream strength at the 250-hPa level (colored contours; m s21) and direction

(black arrows).

AUGUST 2015 THEOBALD ET AL . 1721

T11, all tropical-connected systems demonstrate upper-

level divergence in the exit quadrant of the STJ (Fig. 5).

All synoptic types exhibit airflow directions that are

conducive to orographic enhancement of precipitation.

Synoptic types T1, T2, T5, T6, and T8 can be grouped

as cold-cored frontal-type days [including contributions

from closed and cutoff lows (T2 and T6) and prefrontal

troughs (T5 and T8)], and together account for 53% of

all classified precipitation days (Table 3). Cutoff lows

here follow the definition in the SEA studies of Pook

et al. (2006), Risbey et al. (2009), and Chubb et al.

(2011), among others, in which closed circulation can be

apparent at the surface or midlevels with a trough above

or below, respectively. Themanual analysis showed that

T2 fulfils the traditional criteria of cutoff lows on ap-

proximately 60% (50%) of occurrences and T6 fulfils

them on 72% (44%) of occurrences when considering

MSLP (500-hPa geopotential height). In addition,

closed-low types have associated fronts on 75% (T2)

and 45% (T6) of occurrences. Cold-frontal types have

in common differential cyclonic vorticity advection

(CVA) at 500 hPa as an ascent mechanism (Fig. 10), but

closed-low types demonstrate stronger cyclonic vortic-

ity maxima than do embedded cold fronts and pre-

frontal types. The types T4 and T7 show the signature of

heat lows and troughs interacting with cold-cored

extratropical fronts to the south of Australia (Sturman

and Tapper 2006; Gallant et al. 2012). Localized accel-

eration of the STJ core along the enhanced thickness

gradient, jet-stream divergence in the poleward exit

quadrant, and strong WAA (associated with strong

downstream anticyclones) are consistent with enhanced

system development and higher precipitation totals

(Risbey et al. 2009). Cold fronts (T1) occur at a fre-

quency that is similar to those of other frontal types,

although they deliver smaller amounts of precipitation

across all elevations, consistent with lower humidity and

precipitable water. Strong cold-air advection (CAA),

clockwise rotation of winds with height, and jet-streak

divergence downstream of the front indicate ascent is

provided primarily by frontal lift. Development of

closed and cutoff lows is associated with a localization

and concentration of the STJ, often forming on the

poleward side of the jet. This relationship is represented

in Fig. 5, which shows a strong jet core located to the

north of the closed low for types 2 and 6. Similarly, for

classifications that include the passage of a cold front

(T1 and, to some extent, T5), meridional excursions of

the polar jet that cause it to merge with the subtropical

jet near the location of the front are a known feature

(Sturman and Tapper 2006; Risbey et al. 2009). Fur-

thermore, Risbey et al. (2009) associate cyclonic

FIG. 6. As in Fig. 4, but showing 1000–500-hPa thickness (colored contours; m) and MSLP (line contours; hPa).

1722 JOURNAL OF APPL IED METEOROLOGY AND CL IMATOLOGY VOLUME 54

curvature of the jet-stream core, the exit region of

which is divergent, with the largest amounts of pre-

cipitation (.5mm and, in particular, for synoptic systems

that produce $ 15mm of precipitation) from frontal

systems—evident here in Fig. 6 for frontal-type days

over 10mm.

Synoptic type T3 is representative of inland heat

troughs extending from a low pressure center in north-

ern Australia—known locally as the Cloncurry low

(Sturman and Tapper 2006; Gallant et al. 2012). Strong

WAA, into the divergent region of the jet stream, is

apparent in the vicinity of the trough (Fig. 6).

Topographic interaction of the prevailing airflow in

each synoptic type creates differences in precipitation

contributions between elevation areas. Synoptic types

generating the greatest precipitation totals across

westerly elevations are the result of approaching cold

fronts or troughs, closed lows and troughs that extend

toward northwest Western Australia (T2, T4, T6, T7,

and T8; Table 3). This is similar for the high elevations,

where NWCBs associated with days that are classified

as type 8 and with closed lows (T6) bring the highest

percentage of precipitation totals per day. A common

feature in each of these types (T2, T4, T6, T7, and T8)

is a downstream anticyclone, with WAA and relatively

high PW to the northeast and northwest of the study

region (Figs. 6 and 10), along with divergence in the

poleward exit quadrant of the jet stream (Fig. 5) and

orographic enhancement. As a result, thermal wind

is enhanced and directed southeastward along the

thickness gradient, causing acceleration of the STJ

(Fig. 5)—apparent here for those synoptic systems

classified as T2, T6, and T8. Classifications that include

closed lows alone account for approximately 20% of

days and contribute 26%, 23%, and 16% of total pre-

cipitation to the western, high, and eastern groups,

respectively.

The contribution of precipitation from frontal systems

and the midlatitude westerly airflow is reduced in east-

ern elevations, in the lee of the mountain range. Instead,

onshore easterlies in the subtropics associated with

downstream anticyclones or ridging that advect warm,

humid air from a moisture corridor along the east coast

via inland, meridional troughs (T3, T7, and T10), and

offshore lows (T9), are the major sources of precipita-

tion (Table 3). WAA is a common ascent mechanism for

these types. Differences in the spatial distribution of

precipitation between synoptic types shown here are

consistent with investigations by Chubb et al. (2011) and

Fiddes et al. (2015).

FIG. 7. As in Fig. 4, but for relative humidity (colored contours; %) and wind direction (black arrows) at 850 hPa.

AUGUST 2015 THEOBALD ET AL . 1723

b. Seasonality of synoptic types

Clear seasonality in the frequency of synoptic types is

evident in Fig. 11, which reflects seasonal movement of

the STR. The mean contribution of each type to sea-

sonal precipitation accumulations (Fig. 12) demon-

strates the high degree of intra-annual variability. The

greatest between-type variability, in terms of both fre-

quency and precipitation, occurs in winter and summer.

Orographic enhancement of precipitation for all types is

evident, with highest elevations consistently receiving

the largest precipitation totals in all seasons (Fig. 12).

The majority of types can produce seasonal precipita-

tion accumulations beyond the 95th percentile (.2 stan-

dard deviations) and could be considered as extreme

(Pook et al. 2012). Notable large falls of precipitation,

exceeding 300mm in thewestern and high elevations, have

resulted from the dominant types T4 and T7 in summer

(not shown). As noted in section 3a, these types display

properties that are consistent with enhanced system de-

velopment, strong WAA, and higher precipitation totals.

Table 4 summarizes the mean precipitation from all

days per season and reinforces previous studies that this

region is dominated by cool-season precipitation (nearly

60%), often associated with frontal and closed-low

systems in the lower-to-midtroposphere (surface–500hPa;

Figs. 11 and 12). However, this study highlights that a sig-

nificant proportion of inflow-generating precipitation days

of $10mm (approximately 20%) are recorded during

summer months, often related to the occurrence of inland

heat troughs (i.e., T4 and T7) and increased convection. A

further 20% of days occurred in the transitional autumn

season. Mean daily precipitation from all synoptic types is

similar in all seasons, but the fewer number of precipitation

days occurring in summer and autumn and the higher PW

indicate a higher intensity for warm-season precipitation

days (Table 4).

In austral summer (DJF), precipitation is dominated

by warm-cored heat-trough types T3, T4, and T7. Sum-

mer types are associated with weaker midlevel troughs

and weaker cyclonic vorticity, often displaced upstream

of the SnowyMountains. Instead, stronger ridges, higher

moisture, enhanced WAA, and low-level (upper level)

convergence (divergence) dominate and act as ascent

mechanisms. Moist air, with high PW, is entrained from

tropical latitudes toward the Snowy Mountains region

(Figs. 6–8, 10). These systems deliver relatively consis-

tent falls across all elevations (Fig. 12).

The occurrence of each synoptic type and their asso-

ciated precipitation are more consistent in autumn

FIG. 8. As in Fig. 7, but for 700 hPa.

1724 JOURNAL OF APPL IED METEOROLOGY AND CL IMATOLOGY VOLUME 54

(MAM), although T1, T3, T4, and T5 dominate slightly,

representing a mix of warm- and cold-core synoptic

types. This is a transition season with precipitation still

generated from a moist airflow and WAA from tropical

latitudes (T3 and T4). However, the northward move-

ment of the STR and midlatitude westerly wind belt is

apparent as prefrontal troughs and cold fronts associ-

ated with embedded lows in the Southern Ocean (T1

and T5) begin to cross the region more frequently.

Cold-cored frontal types T1, T2, T5, T6, and T8 and

offshore low pressure centers (T9) dominate in winter

(JJA). All demonstrate CVA with a maximum either

over SEA (T1, T2, and T9) or upstream (T5, T6, and T8).

Winter types generally display higher maximum cy-

clonic vorticity than do summer types. Only one-half of

these types have a tropical moisture corridor at 700 hPa,

and the influence of themidlatitude westerly wind belt is

clear, with a moisture source in the lower atmosphere

(850 hPa) over the Southern Ocean or Tasman Sea.

Low-level convergence, a northwest moisture corridor,

and WAA are common features for T6 and T8, which

deliver the highest winter precipitation totals. In eastern

elevations the influence of orography is apparent,

with much lower precipitation resulting from the pre-

dominantly westerly flow. Instead, low pressure centers

situated off the east coast (T9) are the dominant source

of precipitation.

Similar to autumn, spring (SON) observes an even

spread in the percentage of total seasonal precipitation

days, although cold-core types T2, T6, and T8 along with

T4 occur slightly more frequently. Prefrontal and closed-

low systems (T2, T6, and T8) make the highest contri-

butions to precipitation totals, but, given that this is a

transitional season, systems connected to tropical lati-

tudes are also common (e.g., T4 and T7). Occurrence of

these warm systems, on top of a late-season isothermic

(‘‘ripe’’) snowpack, is considered to generate the greatest

snowmelt and inflow in the Snowy Mountains.

4. Discussion and conclusions

Presented here is a 55-yr (1958–2012) synoptic cli-

matology of daily synoptic circulation systems that de-

liver greater than 10mm of precipitation for the Snowy

Mountains region of southeastern Australia. This is the

first study to link synoptic circulation throughout the

tropospheric column to precipitation at the surface in

the Snowy Mountains. The use of a suite of variables

throughout the depth of the troposphere, applied to a

large gridded analysis area, expands on previous studies

FIG. 9. As in Fig. 7, but for 500 hPa.

AUGUST 2015 THEOBALD ET AL . 1725

(e.g., Wilson et al. 2013) and results in 11 synoptic types.

The clustering method reveals subtle differences in, for

example, the position and orientation of surface troughs

and the location of moisture corridors, which directly

affect the amount of precipitation received at different

sites in the SnowyMountains region. Some types display

similar attributes at certain levels, but each represents a

particular synoptic atmospheric circulation for this re-

gion. The method used in this study has demonstrated

that difficulties in combining variables from different

atmospheric levels (Kidson 2000) can be overcome,

providing a vertical profile of atmospheric conditions

during specific precipitation days. It offers an automated

approach to the traditional map classification (Yarnal

1993) and generates synoptic types with no a priori

forcing of the clustering algorithm, minimizing sub-

jectivity. The importance of moisture source regions and

ascent mechanisms in delivering precipitation to the

region of interest is demonstrated.

The use of a daily precipitation threshold of $10mm

differs from previous southeastern Australia studies in

which all precipitation days have been considered. Im-

portant is that it allows classification of synoptic circu-

lation associated with precipitation days that trigger a

quantifiable increase in streamflow in the headwater

catchments of Australia’s iconic and economically im-

portant river system: the Murray–Darling basin.

Seasonal variations in the frequency of occurrence

of each synoptic type highlight the variability of

atmospheric circulation affecting the Snowy Mountains

region. Although certain synoptic systems are pre-

dominant during the cool or warm seasons, they may

occur at any time (Wright 1989), and a clear seasonal

signal in their incidence is apparent (Fig. 11).

The seasonal movement of the STR is critical to the

synoptic systems experienced in SEA. In winter the STR

is in its most northerly position over central Australia,

allowing the passage of frontal systems associated with

the midlatitude westerly wind belt to traverse southern

Australia. Accordingly, winter synoptic types frequently

relate to the passage of cold fronts and closed lows and

are typically associated with CVA. Areas of descent

coincide with CAA while CVA and ascending motion

are enhanced downwind of the front or trough (Risbey

et al. 2009). Jet streaks display greater intensity in win-

ter. Divergence in their poleward exit quadrant, along

with CVA in midlevels, results in rising motion (Sturman

and Tapper 2006) and is a common ascent mechanism in

winter. In addition, the highest mean winter precipitation

totals result from types demonstrating convergence at

FIG. 10. As in Fig. 4, but for columnar precipitablewater (colored contours;mm) and relative vorticity at 500 hPa31025 s21. Solid vorticity

contours indicate positive (anticyclonic) vorticity, and dotted contours indicate negative (cyclonic) vorticity.

1726 JOURNAL OF APPL IED METEOROLOGY AND CL IMATOLOGY VOLUME 54

850hPa and trajectories of warm, moist air from tropical

latitudes (T6 and T8).

In summer, the southward movement of the STR sees

southern Australia under the influence of a band of high

pressure, associated with the descending branch of the

Hadley cell. Accordingly, frontal systems associated

with themidlatitude westerly wind belt are pushed south

of Australia. Instead, downstream anticyclones and

TABLE 2. Description of key characteristics for each synoptic type. Directional information is referred to withstandard compass-point

notation: north (N), west (W), south (S), east (E) and corresponding points between.

Synoptic type Synoptic summary Ascent mechanism Moisture pathway

T1: embedded cold

fronts

Cold fronts over or rapidly approaching

region; cold core; midlevel trough; no

connection to tropics

Frontal lift; CVA; convergence of polar

jet and STJ over Tasmania; weak

downstream divergence in poleward

exit quadrant of jet streak

Southern Ocean; no

tropical moisture

connection at any

tropospheric level

T2: closed lows Closed lows, including cutoff lows

centered over Tasmania; cold core;

weak downstream ridge; strong

trough at 500 hPa, with cutoff

circulation on ;50% of occasions

WAA over Coral Sea enhanced by

downstream ridge; CVA; divergence

in poleward exit quadrant of STJ jet

streak over Snowy Mountains region

Southern Ocean 850 hPa;

NW tropical 700 hPa

T3: inland heat

troughs

Heat troughs extend equatorward from

low pressure center in northwest

Queensland; warm core; downstream

anticyclone; midlevel trough

Convergent airflow at 850 hPa; winds

back strongly; strong WAA, enhanced

by downstream ridging; weak

divergence in poleward exit quadrant

of STJ

NE tropical 850–500 hPa;

high PW content

(;40mm) along NE

pathway

T4: narrow,

interacting,

inland heat

troughs

Narrow inland heat troughs aligned

NW–SE, interacting with cold-core

low; strong downstream anticyclone;

upstream midlevel trough; NWCBs

on ;10% of occasions

Convergent airflow at 850 hPa; WAA

into divergent poleward exit

quadrant of STJ

Southern Ocean and NE

tropical 850 hPa; NW

tropical 700–500 hPa;

high (;40mm) PW

along NE and NW

pathways

T5: prefrontal

troughs;

approaching

cold fronts

Prefrontal troughs and approaching cold

fronts (west of Snowy Mountains);

cold core; upstream midlevel trough

CVA; divergence in poleward exit

quadrant of jet streak, south of

Snowy Mountains region

Southern Ocean; no

tropical moisture

connection

T6: upstream

closed lows

Approaching closed lows, including

cutoff lows over bight; cold core;

downstream anticyclone; upstream,

strong midlevel trough, with cutoff

circulation on ;45% of occasions

Downstream anticyclone enhances

WAA into area of divergent

poleward exit quadrant of jet streak;

CVA; winds back strongly

Southern Ocean 850 hPa;

NW 700–500 hPa;

moderate PW

(;25mm) along

NW pathway

T7: broad,

interacting

inland heat

troughs

Broad inland heat troughs west of study

region, aligned N–S; interaction with

low pressure system SW of region;

strong downstream anticyclone

Convergent airflow at 850 hPa; WAA

enhanced by downstream anticyclone;

divergence in poleward exit quadrant

of STJ SW of Snowy Mountains

NE tropical 850 hPa;

NW tropical

700–500 hPa; extensive

area of high PW

(;40mm) along NE

and NW pathways

T8: approaching

prefrontal

troughs and

cold fronts

NWCBs

Approaching cold fronts and prefrontal

troughs aligned NW–SE; warm core;

strong downstream anticyclone;

NWCBs on ;40% of occasions

WAA, enhanced by downstream

anticyclone; divergence in poleward

exit quadrant of jet streak upstream

of Snowy Mountains; CVA; winds

back strongly

Southern Ocean 850 hPa;

NW tropical

700–500 hPa; moderate

(;20mm) PW along

NW pathway

T9: offshore low

pressure

systems

Offshore low pressure centers and

troughs; includes east coast lows; cold

core; midlevel closed circulation; no

interaction with tropics

WAA on poleward side of cyclonic

circulation; CVA

Tasman Sea; no tropical

moisture connection

at any level; moderate

PW (;25mm) over

Tasman Sea

T10: easterly

dips

Closed low pressure centers and troughs

aligned N–S, N, or NW of Snowy

Mountains and extending to tropics;

cold core; midlevel closed circulation

Convergent airflow at 850 hPa;

downstream ridging and associated

WAA; divergence in poleward exit

quadrant of STJ

NE tropical 850–500 hPa;

high (.30mm) PW

along NE pathway

T11: noninteracting

inland heat

troughs

Inland troughs; warm-core lows over

central Australia; lack of upstream

trough; NWCBs on ;35% of

occasions

Convergent airflow at 850 hPa;

downstream anticyclone enhances

WAA; winds back strongly

NW tropical

850–500 hPa; high PW

(;30mm) along

NW pathway

AUGUST 2015 THEOBALD ET AL . 1727

ridging enhance WAA and entrainment of tropical

moisture. The warmer seasons are indicative of higher

precipitable water and weaker ascent (Milrad et al.

2014). Areas of low-level convergence that are present

in the dominant summer types indicate positive vertical

velocity and, in addition to strong WAA, provide an

ascent mechanism, in the absence of the CVA present in

winter (Risbey et al. 2009).

In all seasons, interaction between tropical and

extratropical systems (Wright 1989), tropical moisture

pathways, and vertical ascent profiles with low-level

(upper level) convergence (divergence) (Gyakum

2008) are highlighted as important factors generating

precipitation of$10mm, confirming trajectory analyses

conducted by McIntosh et al. (2007) and Chubb et al.

(2011). In a similar way, event precipitation isotope

analyses conducted by Callow et al. (2014) identified

significant synoptic-type variability in the isotopic sig-

nature of precipitation in the Snowy Mountains. The

results of our study further support the dominance of

moisture-source pathways in controlling isotopic vari-

ability in this region.

It is acknowledged that using daily variables may

mask some of the contributions of subdaily systems such

as pre- and postfrontal flow to precipitation over the

region (Gallant et al. 2012) and excludes multiday and

TABLE 3. Percentage occurrence of each synoptic type across all elevation groups (first row) and percentage of total precipitation$ 10mm

received from each synoptic type across each elevation group and in total (last four rows) for the period of 1958–2012.

T1 T2 T3 T4 T5 T6 T7 T8 T9 T10 T11

% occurrence 10.4 11.3 8.6 9.7 10.5 9.6 10.3 10.8 7.3 6.2 5.3

% precipitation

West 6.8 10.6 6.0 10.9 9.2 15.2 12.6 12.4 4.4 6.8 5.1

High 8.8 10.9 6.0 10.4 10.9 12.4 11.0 12.1 6.1 6.1 5.3

East 4.1 6.7 13.9 9.6 7.1 9.0 11.4 8.5 10.0 12.3 7.3

All elev 7.3 10.0 7.5 10.4 9.7 12.6 11.6 11.5 6.3 7.5 5.6

FIG. 11. Intra-annual variability and relative contributions of each synoptic type to the total number of seasonal

precipitation days $ 10mm for (a) DJF, (b) MAM, (c) JJA, and (d) SON.

1728 JOURNAL OF APPL IED METEOROLOGY AND CL IMATOLOGY VOLUME 54

multitype events. Accordingly, slow-moving synoptic

systems may be overrepresented. However, when con-

sidering synoptic-scale circulation, which generally op-

erates on time scales of a few days, the use of daily

variables is considered to be appropriate (Barry and

Carleton 2001; Sturman and Tapper 2006; Pook et al.

2006). In addition, the nature of the composite plots may

result in some features in the synoptic-scale circulation

becoming masked (Milrad et al. 2014). Highly variable

precipitation, complex terrain, and a relatively low

density gauge network in the SILO dataset may all

contribute to interpolation errors. Extensive testing and

comparison of the SHL and SILO datasets have been

used to quantify the uncertainty in using these in-

terpolated precipitation data. This has permitted con-

struction of a continuous daily precipitation record for

the Snowy Mountains region; such a record is consid-

ered to be essential for compiling a long-term synoptic

climatology (Chappell and Agnew 2001).

In summary, this novel method proposes that synoptic

typing can be successfully based on atmospheric vari-

ables throughout the depth of the troposphere. The

higher number of types in this study compares well to

other automated typing schemes, which have generally

produced more types than manual methods, with more

subtle differences between types (e.g., Newton et al.

FIG. 12. Intra-annual distribution and variability of mean precipitation (per precipitation day) associated with each

synoptic type for (a) DJF, (b) MAM, (c) JJA, and (d) SON. Wider pale-gray bars represent high elevations, narrow

black bars represent western elevations, and narrow dark-gray bars represent eastern elevations.

TABLE 4. Mean number of days and seasonal precipitation received from all synoptic types across all elevations for the period of

1958–2012.

Season

Mean No. of precipitation

days $ 10mm

Percentage of total

precipitation $ 10mm

Mean seasonal

precipitation $ 10mm

Mean daily precipitation

from all types (mm)

DJF 12 19.6 175 43.6

MAM 13 20.8 185 43.1

JJA 20 29.6 264 39.6

SON 18 30.0 268 44.7

AUGUST 2015 THEOBALD ET AL . 1729

2014a,b; Plavcova et al. 2014; Kidson 2000). Although

the k-means technique has been widely used in synoptic

classifications, this is the first time, to our knowledge,

that it has been applied to multilevel and multiparam-

eter gridded meteorological data. It has revealed the

complex three-dimensional nature of synoptic-scale

circulation, including the importance of the influence

of moisture pathways from tropical latitudes in the

generation of high precipitation totals. It provides a

method for linking regional-scale precipitation data to

synoptic-scale atmospheric circulation. The spatial dis-

tribution of precipitation associated with synoptic types

has implications for water resources management (Frei

and Schär 1998; Newton et al. 2014a,b) in this region.

The synoptic-typing method that was applied here

allows long-term and climatologically significant periods

to be examined, enabling a robust investigation of the

impacts of synoptic circulation on the hydroclimate of

the Snowy Mountains region. Future research will in-

vestigate temporal variability of the synoptic types in

relation to interannual drivers of climate variability.

Increased understanding gained from this synoptic cli-

matology has implications for water resource manage-

ment in regional areas, and this method could be readily

applied to other hydroclimate studies and to other re-

gions worldwide.

Acknowledgments. We thank Snowy Hydro Limited

scientific staff for helpful and constructive discussions,

and we thank three anonymous reviewers whose com-

ments have greatly improved this manuscript. We also

thank Joshua Soderholm for initial assistance with pro-

gramming. Alison Theobald was supported by an

Australian Postgraduate Award and by Snowy Hydro

Limited. We thank Snowy Hydro Limited and the

Queensland Government Department of Science, In-

formation Technology, Innovation and the Arts for pro-

viding precipitation data. ERA-Interim and ERA-40 data

were provided by ECMWF (Reading, United Kingdom;

http://apps.ecmwf.int/datasets/). Satellite imagery was pro-

vided by NOAA (http://www.ncdc.noaa.gov/gibbs/year).

REFERENCES

Barry, R. G., 1992: Mountain Weather and Climate. 2nd ed.

Routledge, 402 pp.

——, and A. M. Carleton, 2001: Synoptic and Dynamic Climatol-

ogy. Routledge, 620 pp.

Basist, A., G. D. Bell, and V. Meentemeyer, 1994: Statistical re-

lationships between topography and precipitation patterns.

J. Climate, 7, 1305–1315, doi:10.1175/1520-0442(1994)007,1305:

SRBTAP.2.0.CO;2.

Beesley, C. A., A. J. Frost, and J. Zajaczkowski, 2009: A compar-

ison of the BAWAP and SILO spatially interpolated daily

rainfall datasets. Proc. 18th World IMACS/MODSIM Con-

gress, Cairns, Australia, Modelling and Simulation Society of

Australia and New Zealand and International Association for

Mathematics and Computers in Simulation, 3886–3892.

[Available online at http://www.mssanz.org.au/modsim09/I13/

beesley.pdf.]

Beniston, M., 2003: Climatic change in mountain regions: A review

of possible impacts. Climatic Change, 59, 5–31, doi:10.1023/

A:1024458411589.

Bettolli, M. L., O. C. Penalba, and W. M. Vargas, 2010: Synoptic

weather types in the south of South America and their re-

lationship to daily rainfall in the core crop-producing region in

Argentina. Aust. Meteor. Oceanogr. J., 60, 37–48.

Bureau of Meteorology, 2012: Exceptional heavy rainfall across

southeast Australia. Bureau of Meteorology Special Climate

Statement 39, 19 pp. [Available online at http://www.bom.gov.

au/climate/current/statements/scs39.pdf.]

Cai, W., and T. Cowan, 2008: Evidence of impacts from rising

temperature on inflows to the Murray-Darling Basin. Geo-

phys. Res. Lett., 35, L07701, doi:10.1029/2008GL033390.

——, G. Shi, T. Cowan, D. Bi, and J. Ribbe, 2005: The response of

the southern annular mode, the East Australian Current, and

the southernmid-latitude ocean circulation to global warming.

Geophys. Res. Lett., 32, L23706, doi:10.1029/2005GL024701.

Callow,N., H.McGowan, L.Warren, and J. Speirs, 2014:Drivers of

precipitation stable isotope variability in an alpine setting,

Snowy Mountains, Australia. J. Geophys. Res. Atmos., 119,

3016–3031, doi:10.1002/2013JD020710.

Cassano, E. N., and J. J. Cassano, 2010: Synoptic forcing of pre-

cipitation in the Mackenzie and Yukon River basins. Int.

J. Climatol., 30, 658–674, doi:10.1002/joc.1926.

Chappell, A., and C. T. Agnew, 2001: Geostatistical analysis and

numerical simulation of West African Sahel rainfall. Land

Degradation, A. J. Conacher, Ed., GeoJournal Library, Vol.

58, Springer, 19–35.

Chiew, F. H. S., W. J. Young, andW. Cai, 2011: Current drought and

future hydroclimate projections in southeast Australia and im-

plications for water resources management. Stochastic Environ.

Res. Risk Assess., 25, 601–612, doi:10.1007/s00477-010-0424-x.

Christensen, J. H., and Coauthors, 2007: Regional climate pro-

jections. Climate Change 2007: The Physical Science Basis,

S. Solomon et al., Eds., Cambridge University Press, 847–940.

Chubb, T. H., S. T. Siems, and M. J. Manton, 2011: On the decline

of wintertime precipitation in the Snowy Mountains of

southeastern Australia. J. Hydrometeor., 12, 1483–1497,

doi:10.1175/JHM-D-10-05021.1.

Colquhoun, J. R., 1978: Snowfall on the New South Wales Snowy

Mountains. Department of Science and the Environment

Bureau of Meteorology Tech. Rep. 25, 33 pp. [Available online

at http://www.australianweathernews.com/snow/Colquhoun_

Snowfall_NSW_Snowy_Mountains.pdf.]

Dai, J., M. Manton, S. Siems, and E. Ebert, 2014: Estimation of

daily winter precipitation in the Snowy Mountains of south-

eastern Australia. J. Hydrometeor., 15, 909–920, doi:10.1175/

JHM-D-13-081.1.

Davis, R. E., and L. S. Kalkstein, 1990: Development of an auto-

mated spatial synoptic climatological classification. Int.

J. Climatol., 10, 769–794, doi:10.1002/joc.3370100802.

Dee, D. P., and Coauthors, 2011: The ERA-Interim reanalysis:

Configuration and performance of the data assimilation sys-

tem. Quart. J. Roy. Meteor. Soc., 137, 553–597, doi:10.1002/

qj.828.

Drosdowsky, W., 2005: The latitude of the subtropical ridge over

eastern Australia: The L index revisited. Int. J. Climatol., 25,

1291–1299, doi:10.1002/joc.1196.

1730 JOURNAL OF APPL IED METEOROLOGY AND CL IMATOLOGY VOLUME 54

Eder, B. K., J. M. Davis, and P. Bloomfield, 1994: An automated

classification scheme designed to better elucidate the dependence

of ozone on meteorology. J. Appl. Meteor., 33, 1182–1199,

doi:10.1175/1520-0450(1994)033,1182:AACSDT.2.0.CO;2.

Eichler, T. P., and J. Gottschalck, 2013: A comparison of Southern

Hemisphere cyclone track climatology and interannual vari-

ability in coarse-gridded reanalysis datasets. Adv. Meteor.,

891260, doi:10.1155/2013/891260.

Fiddes, S., A. B. Pezza, and V. Barras, 2015: Synoptic climatology

of extreme precipitation in alpine Australia. Int. J. Climatol.,

35, 172–188, doi:10.1002/joc.3970.

Frei, C., and C. Schär, 1998: A precipitation climatology of the

Alps from high-resolution rain-gauge observations. Int. J. Clima-

tol., 18, 873–900, doi:10.1002/(SICI)1097-0088(19980630)18:

8,873::AID-JOC255.3.0.CO;2-9.

Gallant, A. J. E., A. S. Kiem, D. C. Verdon-Kidd, R. C. Stone, and

D. J. Karoly, 2012: Understanding hydroclimate processes

in the Murray-Darling Basin for natural resources manage-

ment. Hydrol. Earth Syst. Sci., 16, 2049–2068, doi:10.5194/

hess-16-2049-2012.

Gao, X., P. Schlosser, E. Monier, and D. Entekhabi, 2014: An

analogue approach to identify heavy precipitation events:

Evaluation and application to CMIP5 climate models in the

United States. J. Climate, 27, 5941–5963, doi:10.1175/

JCLI-D-13-00598.1.

Ghassemi, F., and I. White, 2007: Inter-Basin Water Transfer: Case

Studies from Australia, United States, Canada, China and In-

dia. 1st ed. International Hydrology Series, Cambridge Uni-

versity Press, 462 pp.

Gyakum, J. R., 2008: The application of Fred Sanders’ teaching to

current research on extreme cold-season precipitation events

in the Saint Lawrence River valley region. Synoptic–Dynamic

Meteorology and Weather Analysis and Forecasting, Meteor.

Monogr., No. 55, Amer. Meteor. Soc., 241–250.

Hart,M., R. deDear, and R.Hyde, 2006: A synoptic climatology of

tropospheric ozone episodes in Sydney, Australia. Int.

J. Climatol., 26, 1635–1649, doi:10.1002/joc.1332.

Hendrikx, J., C. Zammit, E. Ö. Hreinsson, and S. Becken, 2013:

A comparative assessment of the potential impact of cli-

mate change on the ski industry in New Zealand and

Australia. Climatic Change, 119, 965–978, doi:10.1007/

s10584-013-0741-4.

Hennessy, K., P. Whetton, I. Smith, J. Bathols, M. Hutchinson, and

J. Sharples, 2003: The impact of climate change on snow

conditions in mainland Australia. CSIRO Atmospheric

Research Rep., 47 pp. [Available online at http://www.

climatechange.vic.gov.au/__data/assets/pdf_file/0015/73212/

TheImpactofClimateChangeonSnowConditions2003.pdf.]

——, ——, K. Walsh, I. N. Smith, J. M. Bathols, M. Hutchinson,

and J. Sharples, 2008: Climate change effects on snow condi-

tions in mainland Australia and adaptation at ski resorts

through snowmaking. Climate Res., 35, 255–270, doi:10.3354/

cr00706.

Hoskins, B. J., and K. I. Hodges, 2005: A new perspective on

Southern Hemisphere storm tracks. J. Climate, 18, 4108–4129,

doi:10.1175/JCLI3570.1.

Jeffrey, S. J., J. O. Carter, K. B. Moodie, and A. R. Beswick, 2001:

Using spatial interpolation to construct a comprehensive ar-

chive of Australian climate data. Environ. Modell. Software,

16, 309–330, doi:10.1016/S1364-8152(01)00008-1.

Jiang, N., 2011: A new objective procedure for classifying New

Zealand synoptic weather types during 1958–2008. Int.

J. Climatol., 31, 863–879, doi:10.1002/joc.2126.

Kalkstein, L. S., G. Tan, and J. A. Skindlov, 1987: An evaluation of

three clustering procedures for use in synoptic climatological

classification. J. Climate Appl. Meteor., 26, 717–730,

doi:10.1175/1520-0450(1987)026,0717:AEOTCP.2.0.CO;2.

Kidson, J. W., 2000: An analysis of New Zealand synoptic types

and their use in defining weather regimes. Int. J. Climatol.,

20, 299–316, doi:10.1002/(SICI)1097-0088(20000315)20:

3,299::AID-JOC474.3.0.CO;2-B.

Landvogt, P. K., J. A. T. Bye, and T. P. Lane, 2008: An in-

vestigation of recent orographic precipitation events in

northeast Victoria. Aust. Meteor. Mag., 57, 235–247.

Lyne, V., and M. Hollick, 1979: Stochastic time-variable rainfall–

runoff modelling. Proc. Hydrology and Water Resources

Symposium, Perth, Australia, Institution of Engineers, 89–92.

Marshall, G. J., 2003: Trends in the southern annular mode from

observations and reanalyses. J. Climate, 16, 4134–4143,

doi:10.1175/1520-0442(2003)016,4134:TITSAM.2.0.CO;2.

McGowan, H. A., S. K. Marx, J. Denholm, J. Soderholm, and B. S.

Kamber, 2009: Reconstructing annual inflows to the headwa-

ter catchments of the Murray River, Australia, using the Pa-

cific decadal oscillation. Geophys. Res. Lett., 36, L06707,

doi:10.1029/2008GL037049.

McIntosh, P. C., M. J. Pook, J. S. Risbey, S. N. Lisson, and

M. Rebbeck, 2007: Seasonal climate forecasts for agriculture:

Towards better understanding and value. Field Crops Res.,

104, 130–138, doi:10.1016/j.fcr.2007.03.019.

Milrad, S. M., E. H. Atallah, J. G. Gyakum, and G. Dookhie, 2014:

Synoptic typing and precursors of heavy warm-season pre-

cipitation events at Montreal, Quebec. Wea. Forecasting, 29,

419–444, doi:10.1175/WAF-D-13-00030.1.

Moron, V., A. W. Robertson, M. N. Ward, and O. Ndiaye, 2008:

Weather types and rainfall over Senegal. Part I: Observa-

tional analysis. J. Climate, 21, 266–287, doi:10.1175/

2007JCLI1601.1.

Murphy, B. F., and B. Timbal, 2008: A review of recent climate

variability and climate change in southeastern Australia. Int.

J. Climatol., 28, 859–879, doi:10.1002/joc.1627.

Nathan, R. J., and T. A. McMahon, 1990: Evaluation of automated

techniques for base flow and recession analyses.Water Resour.

Res., 26, 1465–1473, doi:10.1029/WR026i007p01465.

Newton, B. W., T. D. Prowse, and B. R. Bonsal, 2014a: Evaluating

the distribution of water resources in western Canada using

synoptic climatology and selected teleconnections. Part 1:

Winter season.Hydrol. Processes, 28, 4219–4234, doi:10.1002/

hyp.10233.

——, ——, and ——, 2014b: Evaluating the distribution of water

resources in western Canada using synoptic climatology and

selected teleconnections. Part 2: Summer season. Hydrol.

Processes, 28, 4235–4249, doi:10.1002/hyp.10235.

Nicholls, N., 2005: Climate variability, climate change and the

Australian snow season. Aust. Meteor. Mag., 54, 177–185.

Plavcova, E., J. Kysely, and P. Stepanek, 2014: Links between

circulation types and precipitation in central Europe in the

observed data and regional climate model simulations. Int.

J. Climatol., 34, 2885–2898, doi:10.1002/joc.3882.Pook, M. J., P. C. McIntosh, and G. A. Meyers, 2006: The synoptic

decomposition of cool-season rainfall in the southeastern

Australian cropping region. J. Appl. Meteor. Climatol., 45,

1156–1170, doi:10.1175/JAM2394.1.

——, J. S. Risbey, and P. McIntosh, 2010: East coast lows, atmo-

spheric blocking and rainfall: A Tasmanian perspective. IOP

Conf. Ser.: Earth Environ. Sci., 11, 012011, doi:10.1088/

1755-1315/11/1/012011.

AUGUST 2015 THEOBALD ET AL . 1731

——,——, and——, 2012: The synoptic climatology of cool-season

rainfall in the Central Wheatbelt of Western Australia. Mon.

Wea. Rev., 140, 28–43, doi:10.1175/MWR-D-11-00048.1.

——, ——, and ——, 2014: A comparative synoptic climatology of

cool-season rainfall in major grain-growing regions of south-

ern Australia. Theor. Appl. Climatol., 117, 521–533,

doi:10.1007/s00704-013-1021-y.

Renwick, J. A., 2011: Kidson’s synoptic weather types and surface

climate variability over New Zealand.Wea. Climate, 31, 3–23.

Risbey, J. S., M. J. Pook, P. C. Mcintosh, C. C. Ummenhofer, and

G. Meyers, 2009: Characteristics and variability of synoptic

features associated with cool season rainfall in southeastern

Australia. Int. J. Climatol., 29, 1595–1613, doi:10.1002/

joc.1775.

——, ——, and ——, 2013: Spatial trends in synoptic rainfall in

southern Australia. Geophys. Res. Lett., 40, 3781–3785,

doi:10.1002/grl.50739.

Shi, G., J. Ribbe, W. Cai, and T. Cowan, 2008: An interpretation of

Australian rainfall projections. Geophys. Res. Lett., 35,L02702, doi:10.1029/2007GL032436.

Snowy Hydro Limited, 2003: Engineering Features of the Snowy

Mountains Scheme. 4th ed. Snowy Hydro Ltd., 20 pp.

Stahl, K., R. D. Moore, and I. G. McKendry, 2006: The role of

synoptic-scale circulation in the linkage between large-scale

ocean–atmosphere indices and winter surface climate in

British Columbia, Canada. Int. J. Climatol., 26, 541–560,

doi:10.1002/joc.1268.

Sturman, A. P., andN. J. Tapper, 2006:TheWeather and Climate of

Australia and New Zealand. 2nd ed. Oxford University Press,

541 pp.

Tan, P., M. Steinbach, and V. Kumar, 2006: Cluster analysis: Basic

concepts and algorithms. Introduction to Data Mining, Pear-

son Addison-Wesley, 487–568.

Tapp, R. G., and S. L. Barrell, 1984: The northwest Australian

cloud band: Climatology, characteristics and factors associ-

ated with development. J. Climatol., 4, 411–424, doi:10.1002/

joc.3370040406.

Timbal, B., and W. Drosdowsky, 2013: The relationship between

the decline of southeastern Australian rainfall and the

strengthening of the subtropical ridge. Int. J. Climatol., 33,

1021–1034, doi:10.1002/joc.3492.

Tozer, C. R., A. S. Kiem, and D. C. Verdon-Kidd, 2012: On the

uncertainties associated with using gridded rainfall data as a

proxy for observed. Hydrol. Earth Syst. Sci., 16, 1481–1499,

doi:10.5194/hess-16-1481-2012.

Trauth, M., 2007: MATLAB Recipes for Earth Sciences. 3rd ed.

Springer-Verlag, 336 pp.

Ummenhofer, C. C., A. S. Gupta, A. S. Taschetto, and M. H.

England, 2009: Modulation of Australian precipitation by

meridional gradients in east Indian Ocean sea surface temper-

ature. J. Climate, 22, 5597–5610, doi:10.1175/2009JCLI3021.1.

Verdon-Kidd, D. C., A. S. Kiem, and R. Moran, 2013: Links be-

tween the Big Dry in Australia and hemispheric multi-decadal

climate variability—Implications for water resource manage-

ment. Hydrol. Earth Syst. Sci., 10, 13 539–13 593, doi:10.5194/

hessd-10-13539-2013.

Viviroli, D., and Coauthors, 2011: Climate change and mountain

water resources: Overview and recommendations for re-

search, management and policy. Hydrol. Earth Syst. Sci., 15,

471–504, doi:10.5194/hess-15-471-2011.

Whetton, P. H., 1988: A synoptic climatological analysis of rainfall

variability in south-eastern Australia. J. Climatol., 8, 155–177,

doi:10.1002/joc.3370080204.

Widmann, M., and C. Schär, 1997: A principal component

and long-term trend analysis of daily precipitation in

Switzerland. Int. J. Climatol., 17, 1333–1356, doi:10.1002/

(SICI)1097-0088(199710)17:12,1333::AID-JOC108.3.0.CO;2-Q.

Wilks, D. S., 2006: Statistical Methods in the Atmospheric Sciences.

2nd ed. Academic Press, 627 pp.

Wilson, L., M. J. Manton, and S. T. Siems, 2013: Relationship be-

tween rainfall and weather regimes in south-eastern Queens-

land, Australia. Int. J. Climatol., 33, 979–991, doi:10.1002/

joc.3484.

WMO, 2011: Guide to climatological practices—2011 edition.

WMO Publ. WMO-No. 100, 117 pp. [Available online at

http://www.wmo.int/pages/prog/wcp/ccl/guide/documents/

WMO_100_en.pdf.]

Wright, W. J., 1989: A synoptic climatological classification of

winter precipitation in Victoria. Aust. Meteor. Mag., 37, 217–

229.

Yarnal, B., 1993: Synoptic Climatology in Environmental Analysis:

A Primer. John Wiley and Sons, 256 pp.

1732 JOURNAL OF APPL IED METEOROLOGY AND CL IMATOLOGY VOLUME 54