Download as pdf or txt
Download as pdf or txt
You are on page 1of 14

Global Ecology and Biogeography, (Global Ecol. Biogeogr.

) (2010) 19, 145–158

RESEARCH A biogeographic model of fire regimes


R E VIE W
in Australia: current and future
implications geb_512 145..158

R. A. Bradstock*

Centre for Environmental Risk Management of A B S T R AC T


Bushfires, Institute for Conservation Biology
and Law, University of Wollongong, NSW
Aim Patterns of fire regimes across Australia exhibit biogeographic variation in
2522, Australia response to four processes. Variations in area burned and fire frequency result from
differences in the rates of ‘switching’ of biomass growth, availability to burn, fire
weather and ignition. Therefore differing processes limit fire (i.e. the lowest rate of
switching) in differing ecosystems. Current and future trends in fire frequency were
explored on this basis.
Location Case studies of forests (cool temperate to tropical) and woodlands
(temperate to arid) were examined. These represent a broad range of Australian
biomes and current fire regimes.
Methods Information on the four processes was applied to each case study and
the potential minimum length of interfire interval was predicted and compared to
current trends. The potential effects of global change on the processes were then
assessed and future trends in fire regimes were predicted.
Results Variations in fire regimes are primarily related to fluctuations in available
moisture and dominance by either woody or herbaceous plant cover. Fire in
woodland communities (dry climates) is limited by growth of herbaceous fuels
(biomass), whereas in forests (wet climates) limitation is by fuel moisture (avail-
ability to burn) and fire weather. Increasing dryness in woodland communities will
decrease potential fire frequency, while the opposite applies in forests. In the
tropics, both forms of limitation are weak due to the annual wet/dry climate. Future
change may therefore be constrained.
Main conclusions Increasing dryness may diminish fire activity over much of
Australia (dominance of dry woodlands), though increases may occur in temperate
forests. Elevated CO2 effects may confound or reinforce these trends. The prognosis
*Correspondence: Ross Bradstock, Centre for for the future fire regime in Australia is therefore uncertain.
Environmental Risk Management of Bushfires,
Keywords
Institute for Conservation Biology and Law,
University of Wollongong, NSW 2522, Australia. Australia, climate, fire regimes, forest, fuel, global change, ignition, moisture,
E-mail: rossb@uow.edu.au woodlands.

nental extremes of latitude. By contrast, the bulk of the conti-


INTR O D U C TI O N
nent between these extremes is dry (arid or semi-arid) but still
Australia is the most fire prone of all continents, but fire regimes remarkably fire prone (Allan & Southgate, 2002; Russell-Smith
vary widely, ranging from high frequencies and relatively et al., 2007).
low intensities (e.g. 1–5-year recurrence intervals and Russell-Smith et al. (2007) divided the continent into six
< 10,000 kW m-1; Gill et al., 2000; Williams et al., 2002) in tropi- zones that reflect a gradient of rainfall seasonality, with recent,
cal savannas of the north, to low frequencies and high intensities annual fire activity a positive function of rainfall seasonality. The
(e.g. > 100-year recurrence intervals, > 10,000 kW m-1; Gill & dynamics of biomass growth (fuel) and availability to burn
Catling, 2002) in tall open forests of the cool, temperate south. (dryness) govern these trends, resulting in an immense annual
These examples represent relatively moist ecosystems at conti- area of fire in the northern wet/dry tropics. Russell-Smith et al.

© 2010 Blackwell Publishing Ltd DOI: 10.1111/j.1466-8238.2009.00512.x


www.blackwellpublishing.com/geb 145
R. A. Bradstock

(2007) also detected a strong influence of human activity on fire, may have wider applicability (e.g. exploration of global patterns
with both positive (i.e. increased ignitions) and negative effects of fire, Bowman et al., 2009). The aim is to:
(i.e. intensive pastoral and agricultural land use) evident. 1. Illustrate that variations in fire regimes across ecosystems
Broadly similar conclusions emerge from a recent analysis of reflect a systematic variation in rates of a common set of key
southern African regimes (Archibald et al., 2009). processes, rather than idiosyncratic local effects.
There is general debate about the relative importance of these 2. Predict whether climate change will have equal effects on fire
drivers (climate, fuel, ignitions), and the related influences of regimes throughout Australia, and if not, where the highest
land management, in shaping fire regimes (Westerling et al., levels of change may be anticipated.
2006; Marlon et al., 2008). Linkages between these drivers and 3. Predict the consequences of other global change effects.
underlying biophysical processes should explain variability in Will vectors of change (climate, CO2 and humans) have con-
fire regimes within and between ecosystems (Archibald et al., gruent or antagonistic effects on fire regimes? If congruent
2009). In particular, such an understanding should indicate effects are likely, only the magnitude of change will be uncertain.
which processes are likely to have the dominant influence on fire If antagonistic effects are likely, both the direction and magni-
in particular ecosystems. tude of change will be uncertain.
The drivers of fire regimes are reasonably well known for
some regions within Australia. For example, variations in litter
fuel dynamics in temperate eucalypt forests affect the season, THE B IO PHY SIC AL B ASIS F O R PR EDIC TION OF
frequency and intensity of forest fires (Walker, 1981; Raison VAR IATIO N IN F IR E R EGIMES
et al., 1983). Drought is associated with major fires in southern,
Extremes of moisture availability govern the growth/
forested regions (Ellis et al., 2004; Verdon et al., 2004), whereas
accumulation of biomass/fuel (wetness) on the one hand and its
above average rainfall leads to burning of very large areas in arid,
ability to burn (dryness) on the other (Huston, 2003; Bond &
central Australia (Allan & Southgate, 2002). Seasonal timing of
Keeley, 2005; Pausas & Bradstock, 2007; Russell-Smith et al.,
ignition can cause changes to area burned, intensity and spatial
2007; Archibald et al., 2009). Fire is therefore related to moisture
heterogeneity of fire regimes in tropical savannas (Russell-Smith
in a nonlinear way (Fig. 1). The effects of dryness are only
et al., 1997; Gill et al., 2000). Cross-regional integration of these
important following periods of high moisture and resultant
disparate effects is, however, lacking. Insights via exploration of
large-scale fire patterns in relation to key processes are also
restricted by the limited temporal scope of data (e.g. Russell-
Smith et al., 2007).
Australia is predicted to be prone to substantial shifts in
climate, including increases in temperature and evaporation,
and highly variable and uncertain changes in rainfall (CSIRO,
2007). Much of the focus on future fire under climatic change in
Australia has been on forecasting changes to fire danger indices
(e.g. Williams et al., 2001; Hennessy et al., 2005; Lucas et al.,
2007; Pitman et al., 2007). Cary (2002) used landscape-scale
simulation to show that future increases in fire danger may
translate into significant increases in area burned in forested
environments of south-eastern Australia. However, in addition
to fire weather, climate change has the potential to alter ignition
rates and the availability of fuel by, respectively, changing the
incidence of lightning and plant productivity and decomposi-
tion of dead plant material. Other influences such as the effects
of elevated CO2 on plant growth and the effects of humans on
ignitions add further complexity. Simulation using dynamic
vegetation models (e.g. Thonicke et al., 2001; Scholze et al.,
2006; Lenihan et al., 2008) offers potential for an exploration of
future Australian fire regimes via an integration of these influ-
ences, but such models are currently hampered by coarse spatial
resolution and use of inappropriate plant functional types for
Australia.
Figure 1 Potential relationships between moisture, productivity
This paper presents a conceptual model of the key drivers of and fire frequency (after Bond & Keeley, 2005; Pausas &
fire regimes, and uses it to characterise the biogeography of fire Bradstock, 2007). Corresponding trends in Australian tree cover
across Australia as a basis for predicting the effects of global and domains of strong limitation by woody (vertical hatching)
change on fire regimes. Given that the fundamentals of the and herbaceous fuels (horizontal hatching) or weak limitation
model are universal, insights from its application to Australia (unhatched) are indicated.

146 Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd
A biogeographic model of fire in Australia

its relationship to differing ecosystem characteristics is the key


B Plant functional types to understanding these variations.
The rate of switching in each case will be a function of inter-
Soils actions between climate (mean and variability) and soils (fertil-
A Habitats ity, texture, depth). This will shape habitats and the selection of
Climate available plant functional types (Fig. 2). Therefore, both the lim-
iting switch and its inherent switching rate will vary among
S Australian ecosystems as a result of patterns of biogeographic
CO2 variation in these factors (Hutchinson et al., 2005).
Humans
I Biomass (B)
Figure 2 Influences of biogeographic factors (climate, soils, Surface or near-surface fuels (primarily dead leaf material) are
habitats, plant functional types) on fire regimes via four ‘switches’ critical for the spread of fire in most Australian vegetation types
(biomass growth, B; availability for burning, A; ambient fire (Catchpole, 2002). Fuel quantity is therefore affected by foliar
weather, S; ignitions, I). Potential effects of changing climate,
cover (Walker, 1981) and leaf attributes. A primary distinction
human activity and atmospheric CO2 are indicated by dashed
can be made between herbaceous and woody plants because of
lines.
differences in the characteristics, rate of supply and arrangement
of material derived from these life-forms.
plant growth. Temporal fluctuations in moisture availability will Tree cover will be a primary determinant of the type of
therefore influence fire frequency. Characteristic patterns of surface fuel. Tree cover is positively correlated with patterns of
mean and variance in available moisture are therefore likely to available moisture that arise from climate and interactions with
have a fundamental influence on variation of the fire regime soil depth/texture (Beadle, 1981; Williams et al., 1996; Specht &
(Gill et al., 2002; Russell-Smith et al., 2007). In particular, mois- Specht, 1999; Fensham et al., 2005). Fire regimes may also
ture may influence fire regimes through determination of fuel reduce tree cover below edaphic/climatic potential (e.g. Liedloff
types, via selection of key plant species that are highly influential & Cook, 2007).
(‘fuel species’; Gill et al., 2002). Eucalypt open forests (i.e. canopy projective foliage cover
A small number of large fires account for the bulk of area > 30%; Gill, 1994; Specht & Specht, 1999) occupy areas with
burned over time in many ecosystems (Boer et al., 2008; Cui & relatively high moisture availability (Fig. 1). Litter from trees
Perera, 2008). Significant area burned, via large fires, can only forms the principal surface fuel (Walker, 1981; Raison et al.,
occur when available fuel is spread across large areas. Connec- 1983), augmented by variable contributions from shrubs and
tivity of available fuel is important due to the contagious nature herbs. Fuel quantity and extent is a therefore a function of the
of fire spread (Peters et al., 2004; Pueyo, 2007). Landscape- and turnover of perennial foliage with major variations in cover and
regional-scale patterns of moisture fluctuation will influence the load of surface fuel driven primarily by the degree of tree and
probability of large fires and thus area burned via effects on the shrub cover (a function of moisture availability) and time since
production and drying of contiguous fuel. Temperature, wind the last fire (Walker, 1981; Specht & Specht, 1999; Berry &
speed and humidity directly influence the rate of spread of fires Roderick, 2002).
and the degree to which fires can bridge fuel discontinuities Woodlands (10–30% tree cover; Specht & Specht, 1999) are
(Catchpole, 2002; Peters et al., 2004; Boer et al., 2008). As a common in drier environments (Fig. 1), and cover vast areas of
consequence, area burned will also be influenced by these ‘fire the interior (Beadle, 1981; Groves, 1994). There is interplay
weather’ variables, as recognized in models of fire behaviour and between woody litter (trees) and herbaceous fuels in woodlands.
fire danger rating systems (Catchpole, 2002). Climate will also Litter fuel from trees is typically discontinuous because of rela-
determine the frequency of lightning and the ensuing probabil- tively low cover, so that overall fuel availability is strongly influ-
ity of ignition (e.g. Price & Rind, 1994). enced by fluctuations in herbaceous biomass. Connectivity of
Biomass production (B), its availability to burn (A), fire surface fuels is provided by herbaceous material (e.g. grasses and
weather (S) and ignition (I) represent a hierarchy (Fig. 2) of herbs; e.g. Hodgkinson, 2002; Prober et al., 2007, 2008).
conditional processes governing fire (Archibald et al., 2009).
They can be regarded as four hypothetical ‘switches’ that must
Availability to burn (A)
be simultaneously activated for fire to occur. Area burned will be
a direct function of the spatial scale (i.e. the contiguous area) In forested systems, drought alters the moisture status of com-
over which they jointly prevail. Should any switch be ‘off’, fire pacted litter beds derived from woody plants and thus availabil-
will not occur. Fire can therefore be constrained in differing ity to burn in horizontal (e.g. differing aspects) and vertical (e.g.
ways, through the effect of turning different switches ‘off’. Some within litter bed) planes. This changes the propensity for large
switches are ‘off’ more often than others in different ecosystems, fires to develop by altering the connectivity of fuels across land-
resulting in fundamentally different fire regimes. Identification scapes. Droughts may also temporarily increase litter fall from
of the ‘limiting’ switch (i.e. the switch activated least often) and trees and shrubs in both temperate and tropical environments

Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd 147
R. A. Bradstock

(e.g. Pook et al., 1997; Cook, 2003). By contrast, in grass and a) Effect of biomass growth (B)
herb fuels, dry periods initiate rapid leaf death (curing) of
WSF
perennial and ephemeral species (e.g. Catchpole, 2002; Spessa DSF
et al., 2005). TF TGW AW
Ecosystems in the northern, monsoonal tropics experience
prolonged annual wet and dry seasons, whereas those in south- 1 5 10 100
ern, temperate regions may experience severe drought on a mul- Potential IFI (years)
Moisture
tidecadal cycle, influenced by El Niño oscillations (e.g. Cullen &
Grierson, 2009). Pronounced water deficits are a regular sea- b) Effect of fire weather (A+S)
sonal feature of arid and semi-arid environments (Hutchinson
et al., 2005). AW
TF
TGW DSF WSF
Fire spread (S)
1 5 10 100
Severe ambient weather conditions (high temperatures, wind Potential IFI (years)
Moisture
speeds and low humidity) result in the rapid spread of fires,
irrespective of fuel type (Catchpole, 2002). Ease of ignition and Figure 3 Potential minimum length of the interfire interval (IFI)
flame transfer are increased by high temperatures and low as determined by: (a) biomass growth (B); (b) fire weather (i.e.
humidity. Wind directly affects flame length and depth, and availability of dry biomass plus influence of wind, temperature
propagation of fire via embers. Area burned will therefore be and humidity on rate of spread of fire, A + S) in relation to trends
directly influenced by these variables (Gill et al., 2002; Boer in available moisture, as indicated by direction of arrows (see
et al., 2008). Fire weather is a function of latitude and rainfall, Figure 1). Different forest and woodland types are indicated
with the annual incidence of severe fire danger conditions (tropical forest, TF; temperate dry sclerophyll forest, DSF;
cool temperate wet sclerophyll forest, WSF; temperate grassy
declining with increasing rainfall and latitude (Gill & Moore,
woodland, TGW; arid woodland, AW). Strong limitation by B and
1990; Williams et al., 2001; McCaw & Hanstrum, 2003; Lucas
A + S is indicated by vertical and horizontal hatching, respectively,
et al., 2007). while the unhatched area indicates weak limitation (see Fig. 1).

Ignition (I)
result, fire regimes in woodlands dominated by grass/
Definitive data on ignitions from lightning versus anthropo- herbaceous fuels will be governed by fluctuations in biomass
genic sources are elusive, due to the remoteness of many regions growth. Moisture variation will therefore strongly influence fire
and difficulty of attribution to sources (e.g. McCaw & Han- regimes.
strum, 2003). Many official records indicate large proportions of By contrast, in forested systems where woody litter fuels are
ignitions of unknown origin. Lightning ignition rates (I) gener- prominent, regular patterns of fuel dynamics occur that are
ally vary from being common annually in the tropics to rare in governed by rapid accumulation after fire (Walker, 1981; Raison
the cool temperate zone (Kuleshov et al., 2006). Anthropogenic et al., 1983). The amount of fuel is therefore less tightly coupled
ignitions are a function of population density and land use, with to short-term moisture variations. Sufficient litter is usually
rates high near urban development (Gill & Williams, 1996; present for propagation of fire at most times except for very
Bradstock & Gill, 2001) and low in dry pastoral/agricultural soon after fire (i.e. fuel is non-limiting). Fire will therefore be
landscapes (Noble & Grice, 2002; Russell-Smith et al., 2007). limited primarily by fluctuations in availability to burn and
propagation potential, governed, respectively, by drought and
ambient weather at the time of ignition.
C U R R EN T F I R E R E G I M E S I N A U S T R A LI A :
Ecosystems can be arranged (Figs 1 & 3) along axes represent-
PATTER N S O F LI M I T I N G ‘ S W I T C H E S ’
ing either biomass/fuel (switch B) limitation, or a combination
Potential influences of fuel, fire weather and ignitions on fire of fuel availability (switch A) plus fire spread potential (a func-
regimes are coupled (Fig. 2). Grass/herbaceous fuels will be tion of ambient weather, switch S). This latter combination is
important where moisture deficits (low to moderate tree cover) consistent with fire danger rating systems [e.g. the McArthur
and high levels of fire danger are severe and regular (Fig. 1). Forest Fire Danger Index (FFDI); Catchpole, 2002]. Thus high
Woody/litter fuels (moderate to high tree cover) occur where values of fire danger indices, conducive to a rapid rate of spread
such effects are less severe and regular (Fig. 1) or where trees can and large area burned, are improbable without both severe
access other sources of water (e.g. tropical savannas; Bowman & drought and ambient weather (e.g. Bradstock et al., 2009).
Prior, 2005). Growth and curing of grasses and herbs (perennial This approach contrasts the sensitivity of fire regimes to rela-
and ephemeral) respond rapidly to fluctuations in available tive effects of biomass growth (B) on the one hand and fire
moisture. C4 grasses may be highly influential in this regard, due danger (A + S) on the other (Fig. 3). These represent, respec-
to their productivity under pronounced conditions of seasonal tively, the indirect (‘bottom-up’) or direct (‘top-down’) influ-
variability (Murphy & Bowman, 2007; Osborne, 2008). As a ence of climate (Fig. 2). Predictions of potential fire regimes (i.e.

148 Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd
A biogeographic model of fire in Australia

the limits of average interfire interval) can be made, with real- ephemeral herbs fuels are common (Allan & Southgate, 2002;
ization of potential being dependent on the rate of ignition (i.e. Hodgkinson, 2002). The tree layer, often dominated by Acacia
the fourth ‘switch’: Fig. 2). spp. (Hodgkinson, 2002), is sparse or clumped. Herbaceous
Fire regime potential was explored for five case studies cover (perennial and ephemeral) can be sparse in dry times, but
(Table 1) covering a gradient of tree cover spanning high to low also grows vigorously in response to rain (Griffin et al., 1983;
moisture (Figs 1 & 3) and thus degree of limitation by either Allan & Southgate, 2002; Southgate & Carthew, 2007). Time
woody litter or herbaceous fuels. The case studies are: tropical since fire, in interaction with rainfall, also affects herbage cover,
open forest (TF); arid woodlands (AW); temperate grassy wood- though these effects can be negated after high rainfall (Allan &
lands (TGW); dry sclerophyll shrubby forests (DSF); wet scle- Southgate, 2002; Southgate & Carthew, 2007). High-rainfall
rophyll forests (WSF). These occur across gradients of latitude, events therefore irregularly provide fuel connectivity by stimu-
climate (monsoonal tropics to cool temperate) and available lating herbage which is otherwise absent or slow to develop (a
moisture. Potential fire regimes in forested communities (tree ‘fuel irruption’; Southgate & Carthew, 2007). Such events occur
cover > 30%) tend to be limited primarily by occurrence of at decadal to multidecadal recurrence intervals with consequent
severe fire weather. The degree of limitation increases with effects on fire regime potential (Allan & Southgate, 2002;
increasing moisture (Fig. 1 & 3). By contrast, potential fire Hodgkinson, 2002).
regimes in woodland communities tend to be limited primarily TGW is intermediate between TF and AW, reflecting moisture
by biomass growth. Limitation increases with decreasing mois- status in deep, well-drained soils on the margins of inland,
ture (Figs 1 & 3). Communities with ‘intermediate’ tree cover semi-arid environments (Beadle, 1981; Table 1). Perennial and
(i.e. c. 30%) may be particularly susceptible to factors that alter ephemeral grasses/herbs are prominent, and Eucalyptus, Callitris
the balance between woody overstorey and grass/herb understo- and Allocasuarina spp. are common as trees. Herbage growth
rey (e.g. a change from limitation by B to A or S or vice versa). may fluctuate with level of stocking of native and domestic
The case studies also represent the influence of particular grazing animals and occasional prolonged drought (Noble &
plant species/functional types and land-use context (e.g. pasto- Grice, 2002; Prober et al., 2007). Recovery after fire is relatively
ral to urban hinterlands). Examples are discussed using relevant rapid (e.g. 1–5 years), depending on fluctuations in rainfall
data from a typical region in each instance (Table 1). (Prober et al., 2007, 2008). The potential for fire is therefore high
(e.g. 1–5-year frequency; Fig. 3), because temporal biomass
limitation is weak (i.e. biomass fluctuations are rapid and
Potential fire regimes
regular). The balance between C3 (e.g. Poa) and C4 (e.g.
TF represents grassy, open forests (i.e. tree foliar cover > 30%; Themeda) grasses in these systems may be important (Prober
Table 1) that occur in high-rainfall, coastal regions of northern et al., 2007), with biomass and spatial cover possibly declining if
Australia. These are the mesic variant of tropical savannah C4 abundance declines.
forests and woodlands (Williams et al., 2002) where moisture In comparison to TF (biomass non-limiting), B in the tem-
availability is sufficient to support relatively high tree cover, perate forest examples (DSF, WSF) is weakly limiting due to the
consisting of a mix of evergreen, deciduous and semi-deciduous rate of post-fire accumulation of litter. There is sufficient surface
species. TF represents an extreme where the influences of both litter for propagation of fire at 1–3 years in severe weather
available biomass (B) and fire danger (A + S) converge so that (Raison et al., 1983; Morrison et al., 1996; Huston, 2003). Quasi-
neither influence is strongly limiting (Fig. 3). Frequency of equilibrium surface fuel loads are reached after about 10 years
occurrence of conditions conducive to major fires is close to post-fire (Walker, 1981; Raison et al., 1983). Fire potential there-
annual, due to the effects of both regular annual prolonged wet fore tends to be limited by fire weather (Fig. 3b).
and dry seasons on grass/herb fuels under a monsoonal climate. In DSF, days of extreme fire danger (i.e. FFDI > 49) are rela-
The wet season provides a concentrated moist period for rapid tively rare (e.g. an average of 1 day per year) in mainland south-
growth of grasses, while the dry season provides prolonged eastern and south-western regions (Gill & Moore, 1990; McCaw
annual drought (A) and lengthy sequences of days with fire & Hanstrum, 2003; Lucas et al., 2007). In south-eastern DSF,
weather (S) suitable for the rapid spread of large, relatively characterized by non-seasonal rainfall, regular El Niño–
intense fires (Table 1; Williams et al., 2002). Annual grasses such Southern Oscillation (ENSO)-related droughts (e.g. 5-year fre-
as Saga spp. (= Sorghum; Williams et al., 2002) are suited to these quency) may provide severe fire danger conditions suitable for
conditions and therefore have the potential to control fire major fires (Verdon et al., 2004; Nicholls & Lucas, 2007; Table 1).
regimes and effectively usurp the influence of litter accumula- South-western DSF occurs in a mediterranean climate, with a
tion from trees. long annual summer dry period. Severe fire danger episodes are
In contrast to TF, AW and TGW represent differing degrees of regularly generated (e.g. 5-year frequency) by strong cyclonic
temporal limitation in herbaceous biomass growth (Table 1, activity to the north (McCaw & Hanstrum, 2003; Table 1).
Fig. 3a). As in TF, severe fire weather in AW and TGW is non- Extreme fire danger is rarer in southern Tasmania (Lucas et al.,
limiting or weakly limiting (Table 1, Fig. 3b). Variants of AW 2007; Table 1), but a positive association between summer
cover a vast expanse of the continent. The ground layer consists dryness and area burned has been shown in western WSF
of perennial hummock grasses (e.g. Triodia and Plectrachne regions (Nicholls & Lucas, 2007). The potential for major fires is
spp.) or tussock grasses (e.g. Aristida and Eragrostis spp.) and therefore regular in mainland DSF (5–10-year intervals, Fig. 3b)

Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd 149
150
Table 1 Attributes of case studies that affect fire regimes in differing Australian ecosystems.

Temperate grassy woodlands Temperate dry sclerophyll Cool temperate wet


Tropical open forest (TS) (TGW) Arid woodlands (AW) forests (DSF) sclerophyll forests (WSF)
R. A. Bradstock

Indicative bioregion and Arnhem Coast; Central New South Wales–south-western slopes and MacDonnell Ranges Sydney basin – warm, wet (F3). Long hot Tasmanian west, cold (B2).
agro-climatic zone Arnhem. Hot, seasonally Brigalow Belt south. Warm, seasonally wet/ Desert – high water summers and mild winters. Very cold winters with short
(Hutchinson et al., 2005) wet/dry (I1) dry. Long hot summers and mild winters limitation (G) Jarrah Forest, mediterranean climate warm summers
with significant moisture limits on growth (summer dry season, e.g. Perth E1)
including mediterranean climates (summer
dry season, e.g. Adelaide E1) and
mid-latitude eastern continental climates
with wetter summers and drier winters (e.g.
central New South Wales E3)
Present land uses Pastoralism, conservation Winter cereals and summer crops, grazing Pastoralism, Urban, conservation reserves, water Forestry, hydroelectricity,
reserves, indigenous conservation, catchments, forestry (south-west), mining conservation reserves
indigenous (south-west)
Mean annual rainfall (mm) 1539 (Darwin Post 585 (Dubbo Darling St), 279 (Alice Springs 1214 (Sydney Observatory Hill), 1212 (Maydena)
(Weather station) Office) 529 (Adelaide West Terrace) Post Office) 868 (Perth Regional)
Rainfall seasonality. Predominantly summer Even seasonal distribution (63 January, 43 Summer peak (43 Late summer–early winter peak (131 June, Winter–early summer peak
Months of highest and lowest (393 January, 1 July). September, Dubbo). January, 9 September) 69 September, Sydney). (128 August, 60 February)
mean rainfall (mm) Predominantly winter (72 June, 20 January, Predominantly winter (182 June, 9 January,
Adelaide) Perth)
Canopy tree cover/ > 30%, Eucalyptus, 10–30% Eucalyptus, Casuarina, Callitris < 10–30% Acacia 30–70% Eucalyptus, Angophora, Eucalyptus, Nothofagus,
composition Terminalia Allocasuarina Atherosperma
Understorey structure/ Annual (e.g. Saga) and Perennial (e.g. Themeda) and annual Hummock grasses Diverse shrubs (e.g. Proteaceae, Fabaceae, Shrubs (sclerophyll and
composition perennial grasses grasses, herbs and shrubs (e.g. Triodia), tussock Myrtaceae), sedges (e.g. Cyperaceae), mesic)
grasses (Astrebla), graminoids
herbs
Average number of Very > 30 (Jabiru) (Williams 23 (Dubbo), 18.3 (Adelaide) (Lucas et al., > 60 (Alice Springs) 7.6 (Sydney) (Lucas et al., 2007) 2 (Hobart) (Lucas et al.,
High to Extreme fire weather et al., 2002) 2007) (Williams et al., 2001) > 10 (Perth) (Gill & Moore, 1990) 2007)
days per annum (weather
station)
Fire season Winter–spring Summer–autumn Spring–summer Spring–early summer (south-east), Late summer–autumn
summer–autumn (south-west)
Present ignition sources Predominantly human Predominantly lightning Human and lightning Predominantly human, variable lightning Predominantly lightning
Present range of interfire 2–5 (Russell-Smith et al., > 10, often multidecadal in length (Hobbs, 10–80 (Allan & 5–15 near urban areas, 10–25 in remote > 20 (Gill & Catling, 2002)
intervals (years) 1997, 2007; Williams 2002; Lunt et al. 2006) Southgate, 2002) areas (south-east) (Gill & Moore, 1997;
et al., 2002) Bradstock & Gill, 2001)

Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd
A biogeographic model of fire in Australia

and less regular (> 10 years – Fig. 3b) in western Tasmania fuel characteristics (i.e. flammability) differ among available
(WSF). functional types (fuel species). Bowman (2000) postulated that
alternative states of flammability (potential dominance by dif-
fering fuel species) exist in some Australian ecosystems. Present
Realization of fire regime potential
patterns of fire regimes could reflect such ‘flammable feedback’,
Ignition rates in TF are currently high, due mainly to anthropo- in interaction with the influences outlined above.
genic sources (Russell-Smith et al., 2007), resulting in high- For this to occur, alternative fuel species must have the poten-
frequency fire regimes (Table 1). Fire regimes are therefore close tial to occupy large expanses of habitat in order to influence fuel
to their potential limit (Fig. 3). Variations in summer rain may connectivity and fire regimes. Such potential is poorly under-
affect the quantity of grass fuels and the length of the dry season, stood; hence the degree to which ‘flammable feedback’ is cur-
with high rainfall leading to greater area burned (Harris et al., rently driving selection between alternative fuel species is
2008). In AW, ignition rates (anthropogenic and lightning) also unknown. The availability of species in differing ecosystems not
appear to be sufficient to saturate opportunities where herba- only reflects current conditions (soils and climate) but also phy-
ceous fuel, after curing, is non-limiting, leading to episodes of logenetic legacies and the varying Quaternary influences of
major fire over vast areas (Heydon et al., 2000; Allan & South- people (ignitions and species introductions). Nonetheless, land-
gate, 2002; Russell-Smith et al., 2007). scape flammability may be altered by the introduction of new
By contrast, in TGW fire is often constrained through low functional types in interaction with future changes to habitat
anthropogenic ignitions and a higher level of availability of fire (Fig. 2).
suppression due to higher population density (e.g. agricultural
holdings and small towns), leading to a long-term absence of
burning in natural fragments (Table 1). In larger non- GL O B AL C HANGE EF F EC TS
fragmented tracts of temperate woodlands, multidecadal fire Influences of global change on fire regimes are potentially varied
regimes predominate under the influence of lightning (Table 1). (Figs 2, 4 & 5), due to the complexity of underlying processes.
Current fire recurrence intervals are therefore considerably The influence of climatic change can be distinguished from
longer than their potential minimum (cf. Fig. 3). High domestic other influences such as elevated CO2 and land use (Figs 2, 4 &
stocking in many regions has degraded perennial grass popu- 5). Climate change has the potential to affect B, A, S and I
lations (e.g. Noble, 1997; Hobbs, 2002). Replacement by (Fig. 2), with changes in available moisture (via rainfall, tem-
ephemeral grasses, herbage and woody plants has resulted in perature, evaporation) directly affecting B and A, while changes
diminished fuel continuity and quantity, enhancing the sensi- to temperature and wind will affect S. Elevated CO2 may mostly
tivity of fire potential to rainfall variations and contributing to a affect B and possibly A, through alteration of the growth and
decrease in ignitions and area burned (Noble & Grice, 2002). competitive performance of plant functional types. Human
In DSF ignitions are highly varied (Table 1). In some areas of
eastern New South Wales (NSW), for example, anthropogenic
ignition rates (sometimes complemented by lightning) are suf-
a) Effect of biomass growth (B)
ficient to exploit most opportunities for major fires, resulting in
5–10-year fire recurrence intervals at landscape scales (e.g. Brad- e TGW AW cg
stock & Kenny, 2003), whereas in more remote regions with low
rates of anthropogenic ignition, average interfire intervals may
be > 20 years under the predominant influence of lightning 1 5 10 100
(Bradstock, 2008). Similar trends apply in other regions such as Potential IFI (years)
south-western Australia (Table 1). Fulfilment of fire regime
potential in DSF is therefore highly variable (Table 1, Fig. 3). b) Effect of fire weather (A+S)
In WSF, lightning ignition rates and anthropogenic ignition
rates are relatively low (e.g. King et al., 2006; Table 1). These may AW
be insufficient to saturate opportunities for major fires created f TGW
by irregular summer drought, resulting in fire regimes of mul-
tidecadal to century-scale frequency (Table 1, Fig. 3). Limitation
of fire regimes by ignition (I) may be generally more common 1 5 10 100
with increasing moisture in forested communities. The degree of Potential IFI (years)
limitation will also depend on rates of human ignition as a
Figure 4 Effects of global change on the interfire interval (IFI) in
function of context (e.g. proximity to towns/cities; Bradstock &
woodlands (temperate grassy woodland, TGW; arid woodland,
Gill, 2001). AW) via changes to (a) biomass growth, B, and (b) fire weather,
Fire regimes resulting from any particular combination of B, A+S. Arrows indicate the direction of effects on potential and
A, S and I will affect the relative proportions of plant functional current fire regimes. Symbols indicate the effects of climate
types in landscapes, given inherent life-history and regeneration change (fire danger, f; grass/herb growth, g), elevated CO2 (c) and
attributes. In turn, this may further influence the size of fires if exotic species (e).

Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd 151
R. A. Bradstock

a) Effect of biomass growth (B) weather non-limiting). Thus there is potential for area burned to
WSF either increase or decrease in woodlands due to effects on B but
c li the balance of these effects suggests a decline (Fig. 4).
DSF
By contrast, forests are expected to be sensitive to changes in
e TF cg
fire danger and frequency of severe drought under climate
change (A + S; Fig. 5b). For example, the recent succession of
1 5 10 100 major fire seasons in southern Australia (Ellis et al., 2004;
Potential IFI (years) Bradstock, 2008) is coincident with a prolonged period of severe
drought and elevated fire danger (Lucas et al., 2007). General
increases in future drought and fire danger indices are predicted
in most forested regions (Table 2) but the magnitude of pre-
b) Effect of fire weather (A+S)
dicted change is variable. Hennessy et al. (2005) and Lucas et al.
(2007) predicted substantial increases in drought frequency and
TF f DSF WSF
days of very high to extreme fire danger across broad regions of
the south-eastern mainland but very little change for southern
1 5 10 100 Tasmania (Hobart; Table 2). These studies predict an increase in
Potential IFI (years) severity of drought and fire danger, rather than a change in
drought frequency, as they utilize the current pattern of daily
Figure 5 Effects of global change on the interfire interval (IFI) in
forests (tropical forest, TF; temperate dry sclerophyll forest, DSF; weather as their basis. Changes to ENSO severity are predicted
cool temperate wet sclerophyll forest, WSF) via changes to (a) (e.g. CSIRO, 2007) but details of potential changes in ENSO
biomass growth, B, and (b) fire weather, A+S. Arrows indicate the frequency are lacking. These could be crucial to changes in fire
direction of effects on potential and current fire regimes. Symbols activity in many forested regions. Current estimates of change to
indicate the effects of climate change (fire danger, f; grass/herb fire danger may therefore underestimate the potential for
growth, g; litter fuel accumulation, li), elevated CO2 (c) and exotic climate change to alter fire regimes.
species (e). Declining moisture may result in reduced rates of litter fuel
accumulation in DSF and WSF (Fig. 5a, Table 2) as demon-
strated by comparative trends in litter fuel accumulation under
activity can affect B and A via the introduction of new plant differing present-day moisture scenarios (Walker, 1981; Raison
functional types. Activities such as grazing and clearing may et al., 1983; Huston, 2003). Such effects may tend to reduce area
directly affect B, while there is ever-present potential for human burned. Notwithstanding this possibility, the prognosis for
activity to affect ignition (I). forests is for increasing area burned due to effects of elevated fire
Responses of differing ecosystems can therefore be expected danger. The TF case study may be an exception, due to the weak
to vary according to the balance of these influences and the limiting effects of both fuel and fire weather. The potential for
nature of fire regime limitation in each instance. Woodlands climate change to affect either of these influences is more con-
should inherently be sensitive to factors which affect B whereas strained; however, a reduction in grass growth due to drying
forests should be primarily sensitive to factors influencing A (Table 2, Fig. 5a) has the potential to decrease area burned
and S. (Harris et al., 2008).
There are few detailed predictions of changes to ignition rates
from lightning under climate change (e.g. Hennessy et al., 2005).
Climate change effects
Earlier work suggests that there may be a trend for increasing
In woodlands, changes in available moisture have the potential lightning ignitions (Price & Rind, 1994; Goldammer & Price,
to affect B. Rainfall at stations relevant to the AW and TGW case 1998) due greater atmospheric instability under global
studies is predicted to either increase or decrease under a range warming.
of 2070 scenarios (CSIRO, 2007). The median predictions for
the case-study stations (Table 2) are for a decrease in all cases.
Other global change effects
The consequent decline in available moisture will be exacer-
bated by predicted increases in evaporation in all cases (Table 2). Changes to fire regimes stemming from new plant functional
Increases in summer rainfall and various levels of decrease in types, altered plant growth, land use and ignition rates range
cooler season rainfall are generally predicted. Consequent from those that are well known (e.g. the effects of exotic grasses)
changes to fuel (B; Table 2, Fig. 4b) may include decreased grass/ to speculative and highly uncertain effects (e.g. the conse-
herbage in AW and TGW. Increases in summer rain could favour quences of elevated CO2).
the growth of the important C4 grass component in these cases, In TF and AW there is evidence that new plant species (exotic
particularly in non-tropical environments (Murphy & Bowman, grasses) may affect fire regimes. In TF, the introduction of
2007). While fire danger may tend to increase in these regions gamba grass (Andropogon gayanus), an African exotic, has
due to climate change (Table 2), such effects may be less signifi- resulted in a dramatic elevation of fuel loads and consequent fire
cant due to inherent fuel limitation in woodlands (i.e. fire intensity (Rossiter et al., 2003). The rate of expansion of gamba

152 Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd
Table 2 Global change scenarios in case studies in differing Australian ecosystems.

Temperate grassy
woodlands Temperate dry sclerophyll Cool temperate wet
Global change attribute Tropical open forest (TF) Arid woodlands (AW) (TGW) forests (DSF) sclerophyll forests (WSF)

Rainfall (% change) -1 -9 to -17 -4 to -7 -4 to -8 -3 to -6


-7 to -13* -11 to -19†
Temperature (°C change) +1.7 to +3.2 +1.9 to +3.7 +1.7 to +3.3 +1.6 to +3.0 +1.1 to +2.1
+1.5 to +2.8* +1.4 to +2.7†
Evaporation (% change) +5 to +10 +4 to +7% + 4 to +9 +5 to +9 +5 to +10
+3 to +6* +4 to +7†
Fire danger (very high to extreme days Increase‡ Increase‡ +4.4 to +20.8§ + 0.4 to 6.6 (Sydney)§ 0 to +0.2 (Hobart)§
year–1) +1.6 to +11.5*§ Increase likely‡ (Perth)
Sensitivity (direction of change in mass) of Annual grasses Perennial grasses and Perennial grasses and Woody plant litter and shrub crowns Woody plant litter
main fuel types to: (1) climate change and (1) decrease annual herbs/grasses annual herbs/grasses (1) decrease (1) no change
(2) elevated CO2 (2) decrease (1) decrease (1) decrease (2) increase (2) decrease

Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd
(2) decrease (2) decrease
Woody plant litter
(1) decrease
(2) increase
New plant functional types Gamba grass Buffel grass Tree plantations Exotic grasses – mediterranean areas
Trend in ignitions +Anthropogenic -Anthropogenic +Anthropogenic +Anthropogenic

Climatic predictions are 2070 90p scenarios from CSIRO (2007) for Darwin (TF), Alice Springs (AW), Dubbo (TGW), Adelaide* (TGW mediterranean), Sydney (DSF), Perth† (DSF mediterranean) and Hobart
(WSF).
Fire danger scenarios are based on Williams et al. (2001)‡ and Lucas et al. (2007)§ for 2050.
A biogeographic model of fire in Australia

153
R. A. Bradstock

grass is rapid in places (e.g. the Darwin hinterland), though the ecosystems. Instead, global change factors are more likely to have
bulk of TF remains unaffected. The potential for further spread antagonistic effects (Figs 4 & 5). Both the direction and magni-
and consequent changes to fire regimes across northern Austra- tude of changes to Australian fire regimes are therefore uncer-
lia is high. This situation is paralleled in arid areas of northern tain, given the difficulty in quantifying the magnitude of the
and central Australia (AW) by buffel grass (Pennisetum ciliare; vectors.
Clarke et al., 2005). This species has high drought tolerance, thus The outcome for future fire regimes in woodlands (TGW and
providing levels of biomass and spatial connectivity that may AW) will depend largely on the interplay between moisture and
exceed that contributed by native grasses and herbage (Clarke elevated CO2 effects on B. If a decline in moisture occurs as
et al., 2005). predicted under the median 2070 scenarios (Table 2) then a
Reafforestation of agricultural land in the south (TGW and decline in grass/herb biomass will result. This may complement
some DSF regions) for timber, paper or fuel, carbon credits, any negative effect of elevated CO2 on area burned and fire
salinity control and habitat restoration (e.g. exotic conifers – frequency (Fig. 4). The spread of exotic functional types such as
Pinus spp.; native and exotic Eucalyptus spp.) may enhance land- buffel grass could play a contrary role over significant areas of
scape surface fuel connectivity, particularly in heavily frag- AW due to their effect on landscape connectivity (Fig. 4a).
mented landscapes, where cleared land may otherwise be Enhancement of woody plant growth in woodlands may also
seasonally fuel free due to cropping and/or grazing. Further be context specific. In arid and semi-arid regions, woody cover is
fragmentation of natural vegetation through urbanization and dominated by Acacia, Casuarina and Callitris species, which
agriculture may counteract such trends. form litter beds of low flammability (Bradstock & Gill, 1993;
Elevated atmospheric concentrations of carbon dioxide may Bradstock & Cohn, 2002). In higher-rainfall TGW areas, euca-
differentially affect plant growth, via enhanced effects on woody lypts with litter beds of high flammability are more prominent.
species relative to herbs – particularly C4 grasses (Wang, 2007; Negative effects of elevated CO2 on fire, through stimulation of
Osborne, 2008). Such effects may be dependent on water and woody growth, are therefore likely to be more pronounced in
nutrient availability, which may be particularly important in semi-arid and arid woodlands. Resolution of the outcome of
Australia, given the predominant low-fertility soils and aridity opposing effects of increasing summer rainfall (favourable for
(Hughes, 2003; Steffen & Canadell, 2005). C4 grasses) and elevated CO2 (favourable for woody plants) may
Elevated CO2 could affect woodlands and TF by altering the be particularly significant in TGW. In moist TF regions (grassy
balance between grasses, herbs, shrubs and trees (e.g. Banfai & woodland/open forest), encroachment of rainforest from fire-
Bowman, 2005; Berry & Roderick, 2006). Major changes in protected enclaves, due to CO2 effects, may also diminish the
grass/tree ratios in southern African savannas, with resultant area burned.
effects on fire regimes (i.e. area burned positively related to grass In contrast to woodlands, the balance between fire weather/
cover), are postulated to be driven by fluxes in atmospheric CO2 danger effects (A + S) and fuel effects (B) will determine the
concentration (Bond et al., 2003). Increases in tree and shrub nature of future fire regimes in forests (Fig. 5). Because of the
density at the expense of grass/herb cover are implicated in inherent limitation of A + S in forests, predicted increases in fire
historical declines in area burned in TGW (e.g. Noble, 1997; danger make an increase in average area burned plausible. Fore-
Noble & Grice, 2002). Elevated CO2 may increase accumulation cast changes in fire danger are conservative and do not account
of forest litter fuels via an increase in growth and accession of for crucial changes in major drought frequency that may deter-
litter. Increases in the C : N ratio of leaves (reduced palatability mine the magnitude of any forcing of fire regimes by climate
for vertebrate and invertebrate consumers) could slow decom- change. Potential increases in ignitions (I) from urban expan-
position (e.g. Gleadow et al., 1998; Wang, 2007). sion may be synergistic with fire weather effects in many forested
Rates of fire incidence are positively related to population regions.
density and proximity to the urban interface (Bradstock & Gill,
2001; Keeley & Fotheringham, 2001). Continuing expansion of
C O NC L USIO NS
existing major urban centres and population drift from cities to
rural areas may alter ignition patterns in south-eastern and Climate change and elevated CO2 may lead to a decrease in area
south-western forested regions in particular (DSF, WSF). burned and fire frequency in Australian systems dominated by
herbaceous fuels, due to effects on the limiting ‘switch’ (B).
Factors such as spread of exotic grasses have the potential to
Future trends in fire regimes
cause contrary effects in these ecosystems. In forests dominated
Factors that affect the performance of herbaceous/grass func- by woody, litter fuels, alterations to B through CO2 or other
tional types (B) will have major effects on Australian fire effects may be less critical than changes to fire danger (A + S).
regimes, because systems dominated by herbaceous fuels (e.g. Climate change therefore has a strong potential to increase area
woodlands) cover the bulk of the continent. Given that a range burned in temperate forested regions through an increase in the
of effects on herbaceous growth and spatial continuity (B) are severity of fire danger, including the contribution of drought.
plausible, the potential for major changes to Australian fire These examples do not represent fire regime trends in all
regimes is high. Nonetheless, synergistic effects of all differing Australian ecosystems, or all the possible effects of global
global change factors on fire regimes appear unlikely in many change. Further extension and refinement of this approach may

154 Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd
A biogeographic model of fire in Australia

be appropriate, though the adequacy of understanding of fun- Archibald, S., Roy, D.P., van Wilgen, B.W. & Scholes, R.J. (2009)
damental processes (e.g. fuel dynamics, fire behaviour, ignition What limits fire? An examination of drivers of burnt area in
patterns and sources) may impose constraints. For example, Southern Africa. Global Change Biology, 15, 613–630.
shrub-dominated communities (e.g. various shrublands, heaths Banfai, D.S. & Bowman, D.M.J. (2005) Dynamics of a savanna-
and shrubby woodlands) are of great significance in terms of forest mosaic in the Australian monsoon tropics inferred from
biodiversity and some (e.g. shrubby woodlands in the south- stand structures an d historical aerial photography. Australian
west) cover an extensive area. Such communities occur across a Journal of Botany, 53, 185–194.
range of environments comparable to TGW through to WSF. Beadle, N.C.W. (1981) The vegetation of Australia. Cambridge
Fire regimes in some cases may be governed not only by in situ University Press, Cambridge.
processes but also by characteristics of adjacent communities Berry, S.L. & Roderick, M.L. (2002) CO2 and land use effects on
(e.g. heath patches within tropical savannah or temperate DSF). Australian vegetation over the last two centuries. Australian
Differing approaches may be needed to predict future fire Journal of Botany, 50, 511–531.
regimes at this scale. Ultimately, given the complexity of Berry, S.L. & Roderick, M.L. (2006) Changing Australian veg-
responses and drivers of fire and potential feedback processes, etation from 1788 to 1988: effects of CO2 and land-use
appropriate numerical simulation models may be required to change. Australian Journal of Botany, 54, 325–338.
fully explore the problem. Boer, M.M., Sadler, R.J., Bradstock, R.A., Gill, A.M. & Grierson,
This overview has also focused on factors governing area P.F. (2008) Spatial scale invariance of southern Australian
burnt by large fires and resultant effects on fire frequency, rather forest fires mirrors the scaling behaviour of weather events.
than effects on fire intensity and season and the detail of spatial Landscape Ecology, 23, 899–913.
heterogeneity of regimes. Terrain diversity will influence fire size Bond, W.J. & Keeley, J.E. (2005) Fire as a global ‘herbivore’: the
and spatial patterns of intensity, for example, because of effects ecology and evolution of flammable ecosystems. Trends in
of moisture on fuel availability (A) and rate of spread (S). The Ecology and Evolution, 20, 387–394.
approach could be extended to explore these components of fire Bond, W.J., Midgley, G.F. & Woodward, F.I. (2003) The impor-
regimes, particularly as information on spatial variation in fire tance of low atmospheric CO2 and fire in promoting the spread
severity accumulates. While these deficiencies are considerable, of grasslands and savannas. Global Change Biology, 9, 973–982.
there is sufficient evidence to indicate that variation in the fire Bowman, D.M.J.S. (2000) Australian rainforests: islands of green
regime among Australian ecosystems can be explained by varia- in a land of fire. Cambridge University Press, Cambridge.
tions in the rates of four basic drivers. This provides a basis to Bowman, D.M.J.S. & Prior, L.D. (2005) Why do evergreen trees
guide the development of quantitative fire regime models in the dominate the Australian seasonal tropics? Australian Journal
future, as well as systematic data collection to document the of Botany, 53, 379–399.
nature of the processes that shape these drivers. Bowman, D.M.J.S., Balch, J.K., Artaxo, P. et al. (2009) Fire in the
The concept of spatial connectivity of the drivers of fire earth system. Science, 324, 481–484.
underpins the ‘four switch’ concept presented here. Further Bradstock, R.A. (2008) Effects of large fires on biodiversity in
exploration of the way that landscapes develop ‘flammable con- south-eastern Australia: disaster or template for diversity?
nections’ through interactions between fuel, weather and terrain International Journal of Wildland Fire, 17, 809–822.
(Falk et al., 2007) will provide an incisive basis for predicting Bradstock, R.A. & Cohn, J.S. (2002) Demographic characteris-
and managing future fire regimes. tics of mallee pine (Callitris verrucosa) in fire-prone mallee
communities of central New South Wales. Australian Journal
of Botany, 50, 653–665.
A C K N O W L ED G E M E N T S Bradstock, R.A. & Gill, A.M. (1993) Fire in semi-arid, mallee
shrublands: size of flames from discrete fuel arrays and their
This work was completed with assistance from the Common-
role in the spread of fire. International Journal of Wildland
wealth Department of Climate Change as part of a national
Fire, 3, 3–12.
assessment of the consequences of climate change for fire and
Bradstock, R.A. & Gill, A.M. (2001) Living with fire and biodi-
biodiversity. Input from colleagues engaged on this project is
versity at the urban edge: in search of a sustainable solution to
gratefully acknowledged, with particular thanks to Dick Will-
the human protection problem in southern Australia. Journal
iams. Penny Watson, Alan Andersen and two referees provided
of Mediterranean Ecology, 2, 179–195.
invaluable comments on the manuscript.
Bradstock, R.A. & Kenny, B.J. (2003) An application of plant
functional types to fire management in a conservation reserve
in southeastern Australia. Journal of Vegetation Science, 14,
REF ER EN C ES
345–354.
Allan, G.E. & Southgate, R.I. (2002) Fire regimes in the spinifex Bradstock, R.A., Cohn, J.S., Gill, A.M., Bedward, M. & Lucas, C.
landscapes of Australia. Flammable Australia: the fire regimes (2009) Prediction of the probability of large fires in the
and biodiversity of a continent (ed. by R.A. Bradstock, J.E. Sydney region of south-eastern Australia using components of
Williams and A.M. Gill), pp. 145–176. Cambridge University fire weather. International Journal of Wildland Fire, 18, 932–
Press, Cambridge. 943.

Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd 155
R. A. Bradstock

Cary, G.J. (2002) Importance of a changing climate for fire Gill, A.M. & Williams, J.E. (1996) Fire regimes and biodiversity:
regimes in Australia. Flammable Australia: the fire regimes and the effects of fragmentation of southeastern Australian euca-
biodiversity of a continent (ed. by R.A. Bradstock, J.E. Williams lypt forests by urbanization, agriculture and pine plantations.
and A.M. Gill), pp. 26–46. Cambridge University Press, Forest Ecology and Management, 85, 261–278.
Cambridge. Gill, A.M., Ryan, P.G., Moore, P.H.R. & Gibson, M. (2000) Fire
Catchpole, W. (2002) Fire properties and burn patterns in het- regimes of World Heritage Kakadu National Park, Australia.
erogeneous landscapes. Flammable Australia: the fire regimes Austral Ecology, 25, 616–625.
and biodiversity of a continent (ed. by R.A. Bradstock, J.E. Gill, A.M., Bradstock, R.A. & Williams, J.E. (2002) Fire regimes
Williams and A.M. Gill), pp. 50–75. Cambridge University and biodiversity: legacy and vision. Flammable Australia: the
Press, Cambridge. fire regimes and biodiversity of a continent (ed. by R.A.
Clarke, P.J., Latz, P.K. & Albrecht, D.E. (2005) Long-term Bradstock, J.E. Williams and A.M. Gill), pp. 429–446. Cam-
changes in semi-arid vegetation: invasion of an exotic peren- bridge University Press, Cambridge.
nial grass has larger effects than rainfall variability. Journal of Gleadow, R.M., Foley, W.J. & Woodrow, I.E. (1998) Enhanced
Vegetation Science, 16, 237–248. CO2 alters the relationship between photosynthesis and
Commonwealth Scientific & Industrial Research Organisation defence in cyanogenic Eucalyptus cladocalyx F.Muell. Plant,
& Australian Bureau of Meteorology (2007) Climate Cell and Environment, 21, 12–22.
change in Australia. CSIRO and Bureau of Meteorology, Goldammer, J.G. & Price, C. (1998) Potential impacts of climate
Canberra. http://www.climatechangeinaustralia.com.au/ change on fire regimes in the tropics based on MAGICC and
technical_report.php (accessed 17th December 2009). a GISS GCM-derived lightning model. Climatic Change, 39,
Cook, G.D. (2003) Fuel dynamics, nutrients, and atmospheric 273–296.
chemistry. Fire in tropical savannas: the Kapalga fire experi- Griffin, G., Price, N.F. & Portlock, H.F. (1983) Wildfires in the
ment (ed. by A.N. Andersen, G.D. Cook and R.J. Williams), central Australian rangelands 1970–1980. Journal of Environ-
pp. 47–58. Springer-Verlag, New York. mental Management, 17, 311–323.
Cui, W. & Perera, A.H. (2008) What do we know about forest fire Groves, R.H. (1994) Australian vegetation. Cambridge Univer-
size distribution, and why is this knowledge useful for forest sity Press, Cambridge.
management? International Journal of Wildland Fire, 17, 234– Harris, S., Tapper, N., Packham, D., Orlove, B. & Nicholls, N.
244. (2008) The relationship between the monsoonal summer rain
Cullen, L.E. & Grierson, P.F. (2009) Multi-decadal scale variabil- and dry-season fire activity of northern Australia. Interna-
ity in autumn-winter rainfall in south-western Australia since tional Journal of Wildland Fire, 17, 674–684.
1655 AD as reconstructed from tree rings of Callitris columel- Hennessy, K.J., Lucas, C., Nicholls, N., Bathols, J.M., Suppiah, R.
laris. Climate Dynamics, 33, 433–444. & Ricketts, J.H. (2005) Climate change impacts on fire-weather
Ellis, S., Kanowski, P. & Whelan, R.J. (2004) National inquiry in south-east Australia. Report C/1061. CSIRO, Aspendale,
on bushfire mitigation and management. Council of Austra- Victoria.
lian governments report. Commonwealth of Australia, Heydon, D.T., Friar, J.K. & Pianka, E.R. (2000) Fire driven
Canberra. dynamic mosaics in the Great Victoria Desert Australia. I. Fire
Falk, D., Miller, C., McKenzie, D. & Black, A. (2007) Cross-scale geometry. Landscape Ecology, 15, 373–381.
analysis of fire regimes. Ecosystems, 10, 809–823. Hobbs, R. (2002) Fire regimes and their effects in Australian
Fensham, R.J., Fairfax, R.J. & Archer, S. (2005) Rainfall, land-use temperate woodlands. Flammable Australia: the fire regimes
and woody vegetation cover change in semi-arid Australian and biodiversity of a continent (ed. by R.A. Bradstock, J.E.
savanna. Journal of Ecology, 93, 596–606. Williams and A.M. Gill), pp. 305–326. Cambridge University
Gill, A.M. (1994) Patterns and processes in open forests of Press, Cambridge.
Eucalyptus in southern Australia. Australian vegetation (ed. Hodgkinson, K. (2002) Fire regimes in Acacia wooded land-
by R.H. Groves), pp. 197–226. Cambridge University Press, scapes: effects on functional processes and biological diversity.
Cambridge. Flammable Australia: the fire regimes and biodiversity of
Gill, A.M. & Catling, P.C. (2002) Fire regimes and biodiversity of a continent (ed. by R.A., Bradstock, J.E. Williams and
forested landscapes on southern Australia. Flammable Austra- A.M. Gill), pp. 351–372. Cambridge University Press,
lia: the fire regimes and biodiversity of a continent (ed. by R.A. Cambridge.
Bradstock, J.E. Williams and A.M. Gill), pp. 351–369. Cam- Hughes, L. (2003) Climate change and Australia: trends, projec-
bridge University Press, Cambridge. tions and impacts. Austral Ecology, 28, 423–443.
Gill, A.M. & Moore, P.H.R. (1990) Fire intensities in Eucalyptus Huston, M. (2003) Understanding the effects of fire and other
forests of south-eastern Australia. Proceedings of the 1st Inter- mortality-causing disturbances on species diversity. Fire in
national Conference on Forest Fire Research, Coimbra Portugal. ecosystems of south-west Western Australia: impacts and man-
Paper B24. agement. (ed. by I. Abbott and N. Burrows), pp. 37–70. Back-
Gill, A.M. & Moore, P.H.R. (1997) Contemporary fire regimes in huys, Leiden.
the forests of southwestern Australia. Contract Report for Envi- Hutchinson, M.F., McIntyre, S., Hobbs, R.J., Stein, J.L., Garnett,
ronment Australia, Canberra. S. & Kinloch, J. (2005) Integrating a global agro-climatic

156 Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd
A biogeographic model of fire in Australia

classification with bioregional boundaries in Australia. Global Noble, J.C. & Grice, A.C. (2002) Fire regimes in semi-arid and
Ecology and Biogeography, 14, 197–212. tropical pastoral lands: managing biological diversity and eco-
Keeley, J.E. & Fotheringham, C.J. (2001) Historic fire regime in system function. Flammable Australia: the fire regimes and
southern California shrublands. Conservation Biology, 15, biodiversity of a continent (ed. by R.A. Bradstock, J.E. Williams
1536–1548. and A.M. Gill), pp. 373–400. Cambridge University Press,
King, K.J., Cary, G.J., Bradstock, R.A., Chapman, J., Pyrke, A. & Cambridge.
Marsden-Smedley, J.B. (2006) Simulation of prescribed Osborne, C.P. (2008) Atmosphere, ecology and evolution: what
burning strategies in south-western Tasmania, Australia: drove the Miocene expansion of C(4) grasslands? Journal of
effects on unplanned fires, fire regimes, and ecological man- Ecology, 96, 35–45.
agement values. International Journal of Wildland Fire, 15, Pausas, J. & Bradstock, R.A. (2007) Plant persistence fire traits
526–540. along a productivity and disturbance gradient in Mediterra-
Kuleshov, Y., Mackerras, D. & Darveniza, M. (2006) Spatial dis- nean shrublands of SE Australia. Global Ecology and Biogeog-
tribution and frequency of lightning activity and lightning raphy, 16, 330–340.
flash density maps for Australia. Journal of Geophysical Peters, D.P., Pielke, R.A., Bestelmeyer, B.T., Allen, C.D., Munson-
Research, 111, D19105. doi:10.1029/2005JD006982. McGee, S. & Havstad, K.M. (2004) Cross-scale interactions,
Lenihan, J.M., Bachelet, D., Neilson, R.P. & Drapek, R. (2008) nonlinearities, and forecasting catastrophic events. Proceed-
Response of vegetation distribution, ecosystem productivity, ings of the National Academy of Sciences USA, 101, 15130–
and fire to climate change scenarios for California. Climatic 15135.
Change, 87 (Suppl. 1), S215–S230. Pitman, A.J., Narisma, G.T. & McAneney, J. (2007) The impact
Liedloff, A.C. & Cook, G.D. (2007) Modelling the effects of of climate change on the risk of forest and grassland fires in
rainfall variability and fire on tree populations in an Austra- Australia. Climatic Change, 84, 383–401.
lian tropical savanna with the FLAMES simulation model. Pook, E.W., Gill, A.M. & Moore, P.H.R. (1997) Long-term varia-
Ecological Modelling, 201, 269–282. tion of litter fall, canopy leaf area and flowering in a Eucalyp-
Lucas, C., Hennessy, K.J., Mills, G.A. & Bathols, J.M. (2007) tus maculata forest on the south coast of New South Wales.
Bushfire weather in southeast Australia: recent trends and pro- Australian Journal of Botany, 45, 737–755.
jected climate change impacts. Report to the Climate Institute Price, C. & Rind, D. (1994) Possible implications of global
of Australia Bushfire Cooperative Research Centre, Australian climate change on global lightning distributions and frequen-
Bureau of Meteorology and CSIRO Marine and Atmospheric cies. Journal of Geophysical Research, 99, 10823–10832.
Research, Canberra. Prober, S.M., Thiele, K.R. & Lunt, I.D. (2007) Fire frequency
Lunt, I.D., Jones, N., Spooner, P.G. & Petrow, M. (2006) Effects regulates tussock grass composition, structure and resilience
of European colonization on indigenous ecosystems: post- in endangered temperate woodlands. Austral Ecology, 32, 808–
settlement changes in tree stand structures in Eucalyptus- 824.
Callitris woodlands in central New South Wales, Australia. Prober, S.M., Lunt, I.D. & Thiele, K.R. (2008) Effects of fire
Journal of Biogeography, 33, 1102–1115. frequency and mowing on a temperate, derived grassland soil
McCaw, L. & Hanstrum, B. (2003) Fire environment of Medi- in south-eastern Australia. International Journal of Wildland
terranean south-west Western Australia. Fire in ecosystems of Fire, 17, 586–594.
south-west Western Australia: impacts and management (ed. by Pueyo, S. (2007) Self-organised criticality and the response of
I. Abbott and N. Burrows), pp. 87–106. Backhuys, Leiden. wildland fires to climate change. Climatic Change, 82, 131–
Marlon, J.R., Bartlein, P.J., Carcaillet, C., Gavin, D.G., Harrison, 161.
S.P., Higuera, P.E., Joos, F., Power, M.J. & Prentice, I.C. (2008) Raison, R.J., Woods, P.V. & Khanna, P.K. (1983) Dynamics of
Climate and human influences on global biomass burning fine fuels in recurrently burnt eucalypt forests. Australian For-
over the past two millennia. Nature Geoscience, 1, 697–702. estry, 46, 294–302.
Morrison, D.A., Buckney, R.T., Bewick, B.J. & Cary, C.J. (1996) Rossiter, N.A., Setterfield, S.A., Douglas, M.M. & Hutley, L.B.
Conservation conflicts over burning bush in south-eastern (2003) Testing the grass-fire cycle: alien grass invasion in the
Australia. Biological Conservation, 76, 167–175. tropical savannas of northern Australia. Diversity and Distri-
Murphy, B.P. & Bowman, D.M.J.S. (2007) Seasonal water avail- butions, 9, 169–176.
ability predicts the relative abundance of C3 and C4 grasses in Russell-Smith, J., Ryan, P.G. & DuRieu, R. (1997) A LANDSAT
Australia. Global Ecology and Biogeography, 16, 160–169. MSS-derived fire history of Kakadu National Park, mon-
Nicholls, N. & Lucas, C. (2007) Inter-annual variations of area soonal northern Australia, 1980–1994: seasonal extent, fre-
burnt in Tasmanian bushfires: relationships with climate and quency and patchiness. Journal of Applied Ecology, 34, 748–
predictability. International Journal of Wildland Fire, 16, 540– 766.
546. Russell-Smith, J., Yates, C.P., Whitehead, P.J., Smith, R., Craig, R.,
Noble, J.C. (1997) The delicate and noxious scrub: CSIRO studies Allan, G.E., Thackway, R., Frakes, R., Cridland, S., Meyer,
on native tree and shrub proliferation in the semi-arid wood- M.C.P. & Gill, A.M. (2007) Bushfires ‘down under’: patterns
lands of eastern Australia. CSIRO Wildlife and Ecology, and implications of contemporary Australian landscape
Lyneham, Australian Capital Territory. burning. International Journal of Wildland Fire, 16, 361–377.

Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd 157
R. A. Bradstock

Scholze, M., Knorr, W., Arnell, N.W. & Prentice, I.C. (2006) A Wang, X. (2007) Effects of species richness and elevated carbon
climate-change risk analysis for world ecosystems. Proceedings dioxide on biomass accumulation: a synthesis using met-
of the National Academy of Sciences USA, 103, 13116– analysis. Oecologia, 152, 595–605.
13120. Westerling, A.L., Hidalgo, H.G., Cayan, D.R. & Swetnam, T.W.
Southgate, R. & Carthew, S. (2007) Post-fire ephemerals and (2006) Warming and earlier spring increase western U.S.
spinifex-fuelled fires: a decision model for bilby habitat man- forest wildfire activity. Science, 313, 940–943.
agement in the Tanami Desert, Australia. International Journal Williams, A.A., Karoly, D.J. & Tapper, N. (2001) The sensitivity
of Wildland Fire, 16, 741–754. of Australian fire danger to climate change. Climatic Change,
Specht, R.L. & Specht, A. (1999) Australian plant communities: 49, 171–191.
dynamics of structure, growth and biodiversity. Oxford Univer- Williams, R.J., Duff, G.A., Bowman, D.M.J. & Cook, G.D. (1996)
sity Press, Melbourne. Variation in the composition and structure of tropical savan-
Spessa, A., McBeth, B. & Prentice, I.C. (2005) Relationships nas as a function of rainfall and soil texture along a large scale
among fire frequency, rainfall and vegetation patterns in the climatic gradient in the Northern Territory, Australia. Journal
wet–dry tropics of northern Australia: an analysis based on of Biogeography, 23, 747–756.
NOAA-AVHRR data. Global Ecology and Biogeography, 14, Williams, R.J., Griffiths, A.D. & Allan, G. (2002) Fire regimes
439–454. and biodiversity in the wet-dry tropical savanna landscapes of
Steffen, W. & Canadell, J.G. (2005) Carbon dioxide fertilisation northern Australia. Flammable Australia: the fire regimes and
and climate change policy. Australian Greenhouse Office, biodiversity of a continent (ed. by R.A. Bradstock, J.E. Williams
Commonwealth Department of Environment and Heritage, and A.M. Gill), pp. 281–304. Cambridge University Press,
Canberra. Cambridge.
Thonicke, K., Venevski, S., Sitch, S. & Cramer, W. (2001) The
role of fire disturbance for global vegetation dynamics: cou- B IO SK ETC H
pling fire into a dynamic global vegetation model. Global
Ecology and Biogeography, 10, 661–678. Ross Bradstock is Director of the Centre for Environ-
Verdon, D.C., Kiem, A.S. & Franks, S.W. (2004) Multi-decadal mental Risk Management of Bushfires at the University
variability of forestfire risk – eastern Australia. International of Wollongong. He has research interests in fire ecology,
Journal of Wildland Fire, 13, 165–171. conservation biology, fire regimes and consequences of
Walker, J. (1981) Fuel dynamics in Australian vegetation. Fire global change. He is an adviser on fire management
and the Australian biota (ed. by A.M. Gill, R.H. Groves and policies and practices to land and fire managers.
I.R. Noble), pp. 101–127. Australian Academy of Science,
Canberra. Editor: Brad Murray

158 Global Ecology and Biogeography, 19, 145–158, © 2010 Blackwell Publishing Ltd

You might also like