<<

AUSTRALIA’S – RESPONSES TO FIRE

Plants, and

A.M. Gill, J.C.Z. Woinarski, A. York

Biodiversity Technical Paper, No. 1 Cover photograph credits Group of 3 small photos, front cover: • Cockatiel. The Cockatiel is one of a group of highly mobile birds which track resource-rich areas. These areas fluctuate across broad landscapes in response to local rainfall or fire events. Large flocks may congregate on recently-burnt areas. /Michael Seyfort © Nature Focus • Fern regeneration post-fire, Clyde Mountain, NSW, 1988. /A. Malcolm Gill • These bull ( gulosa) are large ants which generally build small mounds and prefer open areas in which to forage for food. They are found on frequently burnt sites. Despite their fierce appearance, they feed mainly on products. /Alan York. Small photo, lower right, front cover: • Fuel reduction burning in dry . This burn is towards the “hotter” end of the desirable range. /Alan York Large photo on spine: • Forest fire, Kapalga, NT, 1990. /Malcolm Gill Small photo, back cover: • Cycad response after fire near Darwin, NT. /Malcolm Gill

ISBN 0 642 21422 0

Published by the Department of the Environment and Heritage

© Commonwealth of , 1999

Information presented in this document may be copied for personal use or pub- lished for educational purposes, provided that any extracts are acknowledged.

The views expressed in this paper are those of the authors and do not necessarily represent the views of the Department, or of the Commonwealth of Australia.

Biodiversity Convention and Strategy Section Department of the Environment and Heritage GPO Box 636 CANBERRA ACT 2601 General enquiries, telephone 1800 803772

Design: Design One Solutions, Canberra Printing: Goanna Print, Canberra Printed in Australia on recycled Australian paper AUSTRALIA’S BIODIVERSITY – RESPONSES TO FIRE , birds and invertebrates

A. Malcolm Gill CSIRO Division of Plant Industry

J.C.Z. Woinarski Parks and Wildlife Commission of the

Alan York State of

Biodiversity Technical Paper, No. 1 2 Introduction to Biodiversity Technical Paper No. 1

This is the first of the new Biodiversity Technical • Malcolm Gill, CSIRO, Canberra, who Series to be published by the Commonwealth addresses the effects of fire events on plant Department of the Environment and Heritage. biodiversity; The series has been initiated to collate and make • John Woinarski, Parks and Wildlife available information on the conservation and Commission, Northern Territory, who has sustainable use of Australia’s biodiversity to all prepared a review of the literature those involved or interested in biodiversity in concerning the impacts of fire on Australian Australian environments. birds and an annotated bibliography which The papers in this publication have been collates references to fire and Australian written by scientists with expertise in fire and its birds; and effect on Australia’s biodiversity. The papers will • Alan York, then of NSW State Forests, who add considerably to the continuing debate on fire assesses whether frequent hazard reduction in Australia. They will increase access to burning is a sustainable long-term information on how major elements of Australia’s management practice with regard to the biodiversity respond to fire and will be of conservation of terrestrial significance to land management agencies, land biodiversity. Alan’s work was made possible by managers and policy makers at all levels. The papers have been prepared for the the foresight of NSW State Forest in initiating Biodiversity Conservation Branch of Environment a long-term project in which data was collected Australia, Department of the Environment and on the same sites for some 20 years. Heritage, Canberra by: The papers were originally prepared for Environment Australia in 1996.

3 4 CONTENTS

Biodiversity and bushfires: an Australia-wide perspective on plant- changes after a fire event 9 A. Malcolm Gill Centre for Plant Biodiversity Research, CSIRO Division of Plant Industry, ACT

Summary 11

1. Preamble 13

2. Measurement of Biodiversity for Studies of ‘Disturbance’ Effects 15 2.1 What is Biodiversity? 15 2.2 Measures of biodiversity 15 2.3 Functional groups and measures of plant-species diversity 18 2.4 Declared species identities and diversity measures 20 2.5 Discussion 20

3. Models of Changes of Plant Diversity after Fire 23 3.1 Species richness (S-t models) 23 3.2 Species-area relationships (S-A models) 24 3.3 Rank-abundance and rank-cover relationships (i-mi and i-mc models) 25 3.4 Rank trajectories (i-t models) 25

4. Field Methods and Observed Changes in Plant Diversity after Fire 29 4.1 Before-and-after fire comparisons 29 4.2 Comparing adjacent burnt and unburnt stands 29 4.3 Chronological ordination using sites with different ages since fire 34 4.4 Plots, in which changes are observed in situ 35 4.5 Inferential methods 36 4.6 Discussion 37

5. Interpretation of data 39 5.1 Patterns of richness after fire: invasion-caused changes in composition 39 5.2 Patterns of richness after fire: simplification, no invasion 40 5.3 Distribution of patterns of richness change 41 5.4 A plant-community dynamics viewpoint 42

6. Fire Management of Plant Diversity 43

7. Research on Shifts in Plant Diversity after Fire 45

8. Conclusions 47

9. Acknowledgements 48

10. References cited 49

5 Fire and Australian birds: a review 55 J. C. Z. Woinarski Parks and Wildlife Commission of the Northern Territory

Summary 57

1. Introduction 59

2. History of Fire Regimes 61

3. Fire Regimes and their Impacts on Birds: A Review Across Main Environments 63 3.1 Coastal and Thickets 63 3.2 Mallee 67 3.3 Temperate Eucalypt Open Forests 69 3.4 Temperate 73 3.5 Tropical Eucalypt Open Forests and Savanna Woodlands 74 3.6 Tussock 76 3.7 Hummock Grasslands 77 3.8 and Woodlands 77 3.9 Other 78 3.10 Special Case: Islands 78

4. Research Adequacy Methods and Priorities 79

5. Conservation Overview 83

6. References 85

7. Tables 103 Table 1. Scientific Names of Birds mentioned in Text 103 Table 2. Summary Table of Incidence of Subects in Fire- References 106 Table 3. Threatened Bird Taxa for which Inappropriate Fire Regime has been listed as a Threatening Process 108

Fire and Australian birds: an annotated bibliography 113 J. C. Z. Woinarski Parks and Wildlife Commission of the Northern Territory 1. Introduction 115

2. Acknowledgements 116

3. References 117

6 Long-term effects of repeated prescribed burning on forest invertebrates: managment implications for the conservation of biodiversity 181 Alan York Centre for Biodiversity and Bioresources, School of Biological Sciences, Macquarie University

Acknowledgements 182

Executive Summary 183

1. Introduction 187

2. Methodology 189 2.1 Study Area 189 2.2 Experimental Design 189 2.3 Measurement of Environmental Parameters 192 2.3.1 Understorey Vegetation Structure 192 2.3.2 The Litter Environment 192 2.3.3 Sticks and Logs 192 2.3.4 Insolation 192 2.3.5 The Soil Environment 192 2.4 Terrestrial Invertebrate Communities 193 2.5 Analytical Procedures 193 2.5.1 Treatment, Exposure and Position Effects 193 2.5.2 Inter-relationships Between Environmental Variables 194 2.5.3 Terrestrial Invertebrate Communities 194

3. Results 199 3.1 Environmental Parameters 199 3.1.1 Understorey Vegetation Structure 199 3.1.2 Litter Biomass 202 3.1.3 Sticks & Logs 203 3.1.4 Insolation 204 3.1.5 The Soil Environment 204 3.2 Inter-relationships Between Environmental Variables 204 3.3 Terrestrial Invertebrate Communities 207 3.3.1 Ordinal Diversity 207 3.3.2 Invertebrate Abundance 207 3.3.3 Invertebrate Species Richness 213 3.3.4 Community Composition 217 3.3.5 Community Structure 234 3.3.6 Biodiversity Indicators 240

7 4. Discussion 243 4.1 Structure 243 4.2 Terrestrial Invertebrate Communities 245 4.2.1 Invertebrate Abundance 245 4.2.2 Invertebrate Species Richness 246 4.2.3 Community Composition 248 4.2.4 Community Structure 251 4.2.5 Biodiversity Indicators 253

5. Conclusions 257

References 261 BIODIVERSITY AND BUSHFIRES: An Australia-wide Perspective on Plant-species Changes after a Fire Event

A. Malcolm Gill

Centre for Plant Biodiversity Research, CSIRO Division of Plant Industry, GPO Box 1600, Canberra, ACT 2601, Australia

Biodiveristy and bushfires

SUMMARY

Fires can have major effects on plant-species specified life stage (mature) through ‘functional diversity. These effects occur through variations groups’. Major ‘functional groups’ are ‘sprouters’, in fire regimes but are most obvious in the ‘seeders’, and ‘ephemerals’. Groups of species changes that take place after a fire event, the topic which can reproduce without fire are the ‘tolerant’ of this review. Attention here is biased toward or ‘requiring’ species. Using subdivisions of these species, the source of most data. groups can assist our understanding even further. Trends observed in these data may differ from Models allow one to explore a wider range of those observed for non-vascular plants, variability than is currently known. Thus, while or other life forms. species-area, rank-performance and species- How we measure ‘diversity’ is important to richness may all be expected to change with time the way we perceive the effects of fires. Variables after fire, the way they do so will be affected by that contribute to measures of diversity include the functional groups present and their the number of species present in an area, the area contributions to the total species composition. itself, and the numbers of individual plants per Data on the changes of species numbers and species. In some cases, plant cover, height or other variables with time since fire come from a biomass for each species may be more useful range of vegetation types in Australia. Most measures than numbers of plants that are often information has come from areas burnt at very difficult to determine. different times in the past but examined at the Measurements of diversity can be expressed same time; this method assumes that the pre-fire in the form of relationships - like the numbers of species’ compositions, fire characteristics and species in areas of increasing size - or as indices. seasonalities of fires were equitable across sites. The search for a common theoretical foundation Species richness can fall, rise or remain for indices and relationships has been more-or-less constant with time after the initial unsuccessful. In Australia, the number of species, post-fire recovery phase. Declining richness is or ‘species richness’, found in a certain fixed area common in the arid and semi-arid regions. More- has been the most common and simplest measure or-less constant species richness with time after used. If data on the numbers of plants, or cover, fire is a feature of open-forests. Increasing species per species is available, the species can be ranked richness some time after fire is associated with the according to the strength of these measures and invasion of eucalypt forests by species-rich the relationships between them expressed as rank- . Because richness increases abundance or rank-cover curves; the use of such from cold to warm, and from drier to wetter curves should be explored further as a way to environments, however, richness can increase or depict the changes in plant communities that take decrease as encroachment occurs. place after fire. Species’ rank through time can be A post-fire diversity sequence from lower to plotted to indicate shifts in relative performance. higher diversity as species-rich rainforests Most indices and relationships do not encroach upon open-forests involves a change in require the explicit identity of the species present total community composition. This situation but this is fundamentally important for the contrasts with that in ephemeral-strong sequences conservation of biodiversity. Species attain where fire events strongly increase apparent plant individuality when ‘similarity’ indices are used. diversity soon after the disturbance while changes For any one site, however, the listing of species in total diversity (apparent plus that hidden in the (with a measure of performance in terms of soil in the form of and other propagules) numbers, cover, height or biomass) remains the may be subdued or absent; changes such as this basic data set. represent a simplification of an initial Interpretation of relationships and indices composition. recorded at various times after fire can be We have only a rudimentary knowledge of enhanced by the stratification of species according time sequences of plant diversity with time after to their responses to a single fire of a specified fire. Not all communities have been sampled. We ‘biological’ intensity (all foliage killed) at a do not know to what extent fire types, seasons of

11 Australia’s Biodiveristy - Responses to Fire

fire occurrence and post-fire conditions affect sequences. Is there a gradient of patterns of time sequences of species richness from arid to mesic environments as postulated here? Is any geographical gradient influenced by soil ? Managers set fires under prescribed conditions for a variety of reasons, the most common in Parks, State Forests and Reserves being to maintain biodiversity and to protect lives and property. To achieve goals managers need to manage fire regimes. Management burning can be related to times since fire or previous fire-interval distribution but the way in which this is done could have important consequences to the distribution of diversity in a landscape. A major gap in the management of biodiversity by fire would be filled by the adoption of a practical, effective, monitoring system linked with an effective response system. However, the over-arching research challenge that remains is to define ecologically appropriate fire regimes for different ecosystems in different ecological, economic and geographical contexts while identifying the need for particular shapes, sizes and spatial arrangements of burned areas.

12 1. PREAMBLE

Fires occur over the majority of the Australian conditions, of course, may be affected markedly landscape and in most vegetation types - including by fire type and severity but these are rarely rainforest - but in a wide variety of regimes (Gill recorded in studies of the changes in species et al. 1981). The ‘regime’ may include: the composition that may take place after fire. intervals between fires; whether the fires burn It is perhaps obvious that a fire has an effect peat or only the fuels above the soil surface; the that is independent of its ignition source - intensities of the fires; and, the seasons during whether this be human or natural. No distinction which the fires occur (Gill 1975, 1981). After any is made here as to the ignition source of the fire individual fire, the numbers of species present that initiated a particular sequence of change. may remain the same as the years pass but they Indeed, no such distinction is made in the may also decrease substantially or increase. How literature usually. However, there is sometimes the numbers of plant species, and other measures sensitivity among some members of society as to of diversity, change with time since fire is the the effects of fires ignited for management subject of this review. purposes – prescribed fires – compared with those The Australian landscape is not pristine. ignited by lightning. Rapid change has taken place since European There has been no attempt before now to settlement. There has been an imposition of a draw together the many data that are available on plethora of new land uses across the continent. the changes in species’ complements with time With the new land uses has come a wide variety of after fire and attempt to synthesize them. What exotic species of plants and animals. Many of these have we observed? What can we learn from a introduced species have become naturalized. collective view of these studies? Where and when Exotic species are a component of is diversity of plants markedly affected by fires? biodiversity just as native species are. However, in Can we draw conclusions for management from a conservation context, the elimination of exotic these data? What should be done next? species while maintaining all native species is a Most botanical studies of changes with time goal. Among vascular plant species there are now since fire in Australia have concerned changes in 1,952 officially recognized naturalized species in numbers of species or changes in populations of a Australia among a total of 17,590 species (Hnatiuk single species only. While the emphasis here is on 1990). The number of exotic species in Australia is temporal changes, it is being realized increasingly increasing by naturalization while the number of that the study of temporal changes cannot be known native species is increasing as the result of divorced from a consideration of spatial changes. taxonomic research. Spatial matters are considered here, albeit briefly, Among the many plant species that may be at several scales. found at any particular location, only a few - In this study the aims were to: sometimes only one - contribute significantly to (i) assess measures of biodiversity in relation to the fuel available to carry a fire but the fire may time since fire; affect all species present. The impact of any one (ii) collate all the relevant Australian literature fire depends on its characteristics (Gill 1995). on changes in plant species composition with Usually fire intensity has been the fire time since fire; characteristic to measure in ‘surface fires’ (burning (iii) identify species’, environmental and above the ground surface) in relation to vegetation characteristics that might help immediate impact while the extent of combustion explain the changes observed in vascular- may be used as a measure of the severity of plant diversity; and, ‘ground fires’ (burning peat or other material (iv) seek national patterns in diversity changes. considered as being part of the ‘ground’). In this discourse, the immediate impact of a fire will rarely be an issue as it is the events proceeding from the time of fire that are the focus. Initial

13

2. MEASUREMENT OF BIODIVERSITY FOR STUDIES OF DISTURBANCE EFFECTS

The way in which biodiversity is defined and The species-area relationship was formalized measured affects the way that changes after a fire first by Arrhenius (1921): are documented and how the results are S = cAz (1) interpreted. In this chapter a simple definition of biodiversity is given in the first section and this is or, followed by a technical section describing ways in log S = log c + z log A (2) which biodiversity may be measured. In a further section, a way in which the species contributing to where S is the number of species, A is the area the ‘biodiversity’ may be assigned to ‘functional sampled and c and z are fitted parameters. Data groups’ on the basis of their characteristics is for a set of islands undergoing primary succession described. Finally, the importance of the explicit (Rydin and Borgegard 1988) give us examples of identity of species is noted. the values for c and z. Numbers of plant species were recorded 5 times between 1886 and 1985 2.1 WHAT IS BIODIVERSITY? (Rydin and Borgegard 1988, Loehle 1990). Fitted curves to the data gave values of c from 6.3 to ‘Biodiversity’ is the variety of life. In common 10.8, z from 0.16 to 0.30 and r2 from 0.4 to 0.81 usage ‘biodiversity’ is often taken to be the variety (Rydin and Borgegard 1988, Loehle 1990). of vertebrate animals and vascular plants but in Whittaker (1965) distinguished three forms reality it also includes microbes and fungi, of diversity: “α-diversity” was the “diversity and algae, molluscs and bryophytes. In the studies within individual communities”; “β-diversity” being reviewed the focus is on vascular plants. represented the “relative extent(s) of Even with vascular plants what is observed differentiation of communities along topographic and recorded is what the observer can see directly gradients”; and, “γ-diversity” was due to - ‘apparent biodiversity’. There can be a whole “diversities of vegetation patterns, resulting from community of species that remains hidden in the both α and β diversities” (Whittaker 1965). Any soil until appropriate conditions occur - such as an suggestion that c and z reflect α and β diversities extraordinary rainfall or a fire for example - to has been dismissed by Connor and McCoy (1979). release them to become visible as above-ground There has not been universal satisfaction plants. Surveys of land for rare species or unusual with the Arrhenius (1921) equation (1) above. flora should involve, therefore, an appreciation of Gleason (1922), expanding the area of species-area the effects of rainfalls and fires on ‘hidden data from Arrhenius’ 300 dm2 to 240 m2, biodiversity’. suggested:

2.2 MEASURES OF BIODIVERSITY S = c + z log A (3)

An observer standing in a plant community may Rydin and Borgegard (1988) found that notice that as the radius of his or her gaze curves fitted to equation (3) gave values of c increases so too does the number of plant species. between -14 and -61, values of z between 13.2 and The relationship between the number of species 39.3 and values of r2 from 0.44 to 0.85 for this and the area of observation is “one of the most form of equation. studied relationships in all of ecology” (Palmer The area considered by various authors was and White 1994). Such relationships may be used further expanded by Preston (1960). His as the base from which some of the many “Arrhenius plots” (equation (2)) for nearctic and measures of diversity can be derived (i.e. neotropical birds were sigmoidal in shape with Kobayashi and Kimura 1994). three zones where the graph was more-or-less linear viz., areas up to about 100 acres (40 ha), areas from 100 acres to about 108.5 acres (ca. 1.3 million km2), and areas greater than 108.5 acres. Gitay et al. . (1991) found that neither the Arrhenius curve (equation (1)) nor the Gleason

15 Australia’s Biodiveristy - Responses to Fire

curve (equation (3)) gave good fits to their New an adjacent larger one. Replenishment of an island Zealand data. They suggested a three-parameter population would have to take place from a equation rather than the two parameter forms of relatively distant source over an inhospitable previous authors: barrier. If so, by this mechanism alone, the numbers of species occurring in a sample may be x S = c + z log A (4) expected to be larger than those on an island of where x is the additional parameter. They called the same size - as is the case (Preston 1962 a, b). this model the “general root model” which had Such observations formed the basis of island considerable success in extrapolation from the biogeographic theory (MacArthur and Wilson small areas of their plots to the whole of New 1963, 1967) which has had such a major role in Zealand with an area of about 269, 000 km2. the debates over selection of reserves for Species-area curves are usually calculated conservation of biodiversity. from data collected from a series of independent Measures of biodiversity which include areas of varying size (such as the islands reported numbers of plants could be derived from species- by Rydin and Borgegard 1988) or from a series of area curves if the numbers of plants per area were quadrats of increasing size, each larger one known. At any one site, it may be expected that incorporating all smaller ones (i.e. ‘nested’ the number of plants present would be quadrats in which the areas are not independent). proportional to area (e.g. Preston 1962 b, Both of these methods have statistical flaws. The Kobayashi and Kimura 1994): size of quadrat or sample area may have an affect N = d A (5) so using a number of separate quadrats of equal size (independent samples) seems pertinent to any where N is the number of plants in the area sampling strategy (Palmer and White 1994). and d is a constant. (Note that d is site and stand- Palmer and White (1994) found that the age specific.) Using this relationship, equations (1) geographic distances separating quadrats (see also to (4) could be modified to reflect numbers of Gleason 1922) and the number of samples taken plants rather than areas of land. For example, were important. equation 1 becomes: So far we have considered only the most -z z basic of diversity indexes, the count of species S = cd N (6) present in an area - ‘species richness’ - and its or, dependence on area. If species richness was the same in two equal areas but the numbers of plants S = c' Nz' (7) present were vastly different, would comparisons -z of diversity between the two be valid? Knight where c' equals cd (1975) examined species-rich tropical forest where Alternatively: the number of species in younger and older forests respectively were virtually the same on an log S = log c' + z log N (8) area basis - 58 and 57 per 1000 m2 - but quite different on a per-tree basis - 115 and 151 species per 500 . On these sorts of grounds, many where c' is a constant. authors have considered that numbers of plants At Myall Lakes National Park in NSW, Fox should be explicit in any diversity measure. and Fox (1986) found this relationship to hold Preston (1962 b) proposed that the chances among understorey plants in two areas with of were inversely proportional to different fire histories. The value of z was site population size and that population size was dependent. directly proportional to area so that the chances of Menhinick (1964) tried logS/logN as a extinction would be greater on smaller islands measure of diversity of insects. This is the same as than larger ones. Thus, from this reasoning alone, z in equation (8) when log c' equals 0. The values smaller areas would have fewer species than larger obtained ideally should be independent of sample areas. If the areas of concern were sample areas on size (N). Menhinick (1964) found that S /√N was the mainland, rather than islands, the chances of the most stable of the indexes (in relation to extinction may be expected to be less because a sample size) he tried across a range of sample small population there may be replenished from sizes. Kobayashi and Kimura (1994) compared six

16 Biodiveristy and bushfires

diversity indexes and found that the most stable number of theoretical models of this type of index was a new one, δ : relationship. In a plant context, numerical dominance may δ √ = S/ log N (9) be of little significance. Whittaker (1965) defined which could be regarded as a simplification of a dominant plant as one which, if removed, allows equation (3) modified by equation (5). Given x a strong positive response in one or more equal to 0.5, equation (4) becomes: remaining species. “In a given community one or a few species, the dominants, overshadow all S0.5 = c + z log A (10) others in their mass and biological activity and may strongly affect conditions of environment for If substitution for A is made (equation (5)) other species” (Whittaker 1965). Dominance also then: affects plant species richness (Specht and Specht S0.5 = c + z log (N/d) (11) 1989). Thus, biomass, cover, height or some other “importance value” (Whittaker 1965) will often be If c equals 0 and d equals 1 then: more appropriate in expressing functional relationships than number of individuals. Even so, z = S0.5/log N (12) that a species ranks first does not necessarily express its functional dominance. At very low or, z (equation (12)) equals δ (equation (9)). cover values soon after fire, for example, the species with the highest cover, biomass or In the data of Gitay et al. (1991) the best-fit numerical dominance does not necessarily exert exponent of S for vascular plants in New Zealand any functional dominance at all. The only species- was 0.46 while c was 4.5. Thus the assumption diversity relationship that allows for dominance of needed for a ratio to be derived from a species- any nature to be expressed is one involving area curve as the measure of diversity (equation species-rank. (12)) viz., that c be zero, was not met by this data. There seems to be no theory for predicting Another variant on equation (12) is given the frequency distributions of biomass or cover when x of equation (4) equals 1, i.e. z becomes values among species. Such a theory for biomass S/log N. McNaughton (1993) used this index in seems unlikely in a general sense but cover-value his studies of the effects of grazing on plant distributions (or, more specifically, foliage-cover communities while Menhinick (1964) considered distributions) among species and plant sizes may it (and rejected it) in relation to the depiction of be worth exploring because of the importance of diversity in catches. -area index to ecosystem function and its If numbers of plants per species in an area dependence on site conditions. Whittaker (1965) fitted a particular statistical distribution, such as a used net annual production as his measure of log-normal distribution (e.g. Preston 1962 a, b), functional dominance. then the abundances of individuals in each species Examination of the changes in biodiversity (m ) could be calculated from equations such as i with time since fire have mostly involved the equation (7) and ranked (by rank i) such that a simplest possible diversity index - the number of ‘rank-abundance’ curve (i-m ), could be drawn i species in a particular area. Species-area curves (Kobayashi and Kimura 1994) for a site at a express relationship at a particular time. While particular time. Such a curve would not predict species numbers present in a particular area, or which species would be of which rank, of course, species-area curves, are valuable in being able to but would indicate the shape of the curve. demonstrate some of the changes that may take Rank-abundance curves indicate the number place after fire, an increase in the number of of species in the sample (they are truncated inputs would provide further insight into the ways curves, the last entry indicating the number of in which the ecosystem may be changing. species present), the numbers of individuals per Changes in numbers of individuals, species and the relative numerical dominance of important to the expression of demographic the top-ranking species. If the rank-one species change, can be expressed in rank-abundance has 1000 individuals, its numerical dominance curves. Any numerical dominance effects on over the next may be assessed according to species numbers can be shown in the same curves. whether it has 1000, 100 or only 10 individuals In many plant communities, rank-cover curves present for example. Whittaker (1965) displays a

17 Australia’s Biodiveristy - Responses to Fire

(i-ci) will be more appropriate than rank- (ii) ‘seeders’ (species which have mature plants abundance curves. ‘Cover’ lacks, of course, the that die when exposed to the standard fire genetic and demographic implications of plant intensity), or CDR; and, number. Lamont et al. (1977) in their studies of (iii) ‘ephemerals’ (species which do not have live the plant diversity of a Western Australian ‘plants’ exposed to fires but have propagules heathland used the “importance value index” that develop into mature plants after fires (Curtis and McIntosh 1950) which included the only to senesce and die before the next one) frequency of occurrence of species among or EPH. quadrats, their cover and the numbers of Subdivision of seeders’ into those with or individuals present. without soil storage can be important to No single attribute seems ideal for the study understanding population dynamics. of diversity change with time since fire. The use To examine the sensitivities of various of constant plant numbers as the basis of an index measures of diversity, the ways in which these eliminates the expression of important measures might change after a fire is outlined for demographic change. The use of constant area the different functional groups of plants. Plant when numbers of plants are changing through numbers and cover, rather than or biomass or orders of magnitude may bias estimates of relative height, is used for convenience. species numbers present. The use of total cover, Sprouters, by definition, have a resprouting height or biomass, only, will exclude the response from mature plants after fire. A importance of dominance effects of individual percentage, usually small, may die at the time of species being recognized. Using a variety of fire, the extent of death in the population perhaps relationships will help demonstrate the dynamics depending on the life stages present and the of change in plant communities with time after characteristics of the fire. In the example (Table fire. Explicitly recognizing the species involved is 2.1), two cases in which there is no mortality of a further step in appreciating the changes that are prefire plants are considered. In the first case, taking place while classifying the species into SPRwSD, all the plants present resprout after the ‘functional groups’ (see below) may allow a more fire but there are more plants of the species general understanding. present because of substantial seed germination. It is important that the changes in species In the second case, SPRoSD, all the plants present richness detected in the field make biological as resprout as in the first case but there is no well as statistical sense. Biological sense can begin germination or other multiplication of the to be made only in the light of knowledge of population. Mature plants of the ‘seeders’ all die species identities and functional groups. at the time of the fire in our example but return in the form of seedlings. Ephemerals are not 2.3 FUNCTIONAL GROUPS AND apparent at the time of the fire but many species MEASURES OF PLANT-SPECIES with numerous seedlings emerge from soil-stored DIVERSITY seed in our example. The relationships or indexes to be There are many ways of classifying species into considered (Table 2.1) are: groups that behave similarly. A common (i) S, the number of species in a selected, fixed, classification is into life forms. Taxonomic area (i.e. species richness); groupings at a gross level e.g. cryptogams and (ii) S-A, the species-area curve; vascular plants can be instructive. In the context of (iii) S-N, the species richness versus plant this contribution, however, there are three number curve; commonly recognized functional groups (e.g. δ √ (iv) = S log N Noble and Slatyer 1980, 1981; Gill 1981, 1993): / (v) i-mi, the rank-abundance curve; and, (i) ‘sprouters’ (species which have mature plants (vi) i-ci, the rank-cover curve. that resprout after a standard fire intensity - Table 2.1 indicates the changes after fire that 100% leaf scorch) which can be divided into may be expected in the indexes or in mathematical two convenient groups - those with seedling relationships for the functional groups concerned. reproduction at the time of fire, ‘sprouters The changes that occur through a fire event may with seedlings’ or SPRwSD, and those be expected to reflect the changes that take place usually with little or no seed regeneration, over a longer time period between fires but in the ‘sprouters without seedlings’ or SPRoSD; reverse direction.

18 Biodiveristy and bushfires

Table 2.1. An indication of the likelihood of change in indexes or relationships after fire according to plant functional group. SPRwSD refers to a sprouter group of species producing seedlings as well as resprouts after fire; SPRoSD refers to a sprouter group without seedlings after fire (no mortality); CDR refers to a seeder group; and, EPH refers to species of ephemerals. S is species richness; S-A is the species-area relationship; S-N is the species-plant number relationship; δ √ = S/ log N; i-mi is the rank-abundance relationship; and, i-ci is the rank-cover relationship. δ Functional group S S-A S-N i-mi (number) i-ci(cover) SPRwSD no change no change change change change change

SPRoSD no change no change no change no change no change change

CDR no change no change change no change? change change

EPH change change change change change change

All indexes or relationships shift for the Loss in a species of ephemeral would not be ephemerals because these species were not present detected because there is no knowledge of which before the fire (by definition). By way of contrast, ephemerals were present before the fire. Where no change in any index would occur for sprouters the number of species can be separated from the without post-fire seedlings. For the two other number of plants, the measure can record the functional groups no change was expected in S or absence of a species. Where the number of plants δ S-A but could be for and would be for i-mi. alone is the basis for an open-ended relationship δ Only the rank-cover diagrams (i-ci) would record (i.e. S-N and ) then it is doubtful that the loss changes for all functional groups. Thus, the rank- would be detected when seedlings arise after the cover plot seems to be the most appropriate single fire. Thus, again, the use of the measure of relationship to use. However, any shift in an index diversity is greatly improved if the nature of the or relationship as a result of fire needs to be species responses is taken into account. considered in the light of the functional groups For the purposes of this paper the number of present. The same conclusion would follow if the species in a fixed area (i.e. S), the S-A relationship indexes or relationships were to be compared for and the i-mi or i-ci relationships seem to be the the effects of other disturbances. most useful to describe the changes taking place Fire is a natural variable but is also a major after fire. When dominance affects are considered, and often controversial management tool. Two the last two of these relationships become the major issues in the use of fire are its value in the most appropriate choices. Plant number protection of life and property and its affects on relationships are most useful if all the plants are of conservation of species (e.g. Gill and Bradstock similar size but lose their impact as far as 1994, 1995). An important role of diversity indexes dominance is concerned if sizes are varied. Cover and relationships, therefore could be to measure measurements may be essential if the numbers of the effects of fires in relation to conservation. In plants cannot be measured or if functional- Table 2, the possibility of detection of a local dominance (as opposed to numerical dominance) extinction of a species by each index and is to be represented. relationship is given for each functional group.

Table 2.2. The likelihood of detecting an extinction from various indices and relationships when applied to particular functional groups of species. Symbols are as for Table 2.1. ‘Extinction’ in this context refers to the loss of the species from the site due to fire occurrence whether or not it is ‘hidden’ or ‘apparent’.

Functional group S S-A S-N d i-mi i-ci SPRwSD yes yes ? ? yes yes

SPRoSD yes yes yes yes yes yes

CDR yes yes ? ? yes yes

EPH no no no no no no

19 Australia’s Biodiveristy - Responses to Fire

2.4 DECLARED SPECIES IDENTITIES The nested-subset analysis begins by AND DIVERSITY MEASURES examining the data for departure from ‘nestedness’ (i.e. all shorter species’ lists are A constant species richness maintained with time subsets of larger ones) then applying a statistical elapsed after fire may hide a species turnover test to assess its statistical significance. Data are which can only be detected if the identities of the examined for their departure from nestedness by: species are known and declared. All of the indexes (i) determining the shortest faunal list (in their and relationships so far discussed have an case) in which a particular species occurs; (ii) independence from any declared identity of examining all richer faunas for the same species species. Temporal variations in species diversity and recording the absences of that particular can only be properly understood when the taxa species in the richer faunas; (iii) repeating these involved are known. Even spatial comparisons are steps for all species; and, (iv) counting the somewhat hollow unless taxa are identified at numbers of absences in the data set and divide it some level - kingdom, family, species. by the number of sites multiplied by the numbers Methods of comparing species lists have of species in the richest set. The analysis is an been developed over decades. Similarity indexes improvement on species-area curves because it are based on species lists for two sites or at two requires species’ identities to be explicit but some times. Species common to both lists a, or species statistical problems remain (see Doak and Mills found in one, b, or the other list, c, may be 1994 for example). Nested subset analysis could distinguished. The grand total number of species, be relevant to analysis of temporal species changes G, in this example is then (a+b+c). An index of after fire but, as yet, has not been considered in similarity may be a/G or an index of dissimilarity this context. It has the valuable property that may be (b+c)/G. When numbers of sites or times nestedness within a time sequence of stands could are considered there is a further possibility, the be considered. absence from the two sites or times of a set of other species. Such species are often designated by 2.5 DISCUSSION d, a term increasing the variety of possible This section started with a consideration of indexes. species-area curves, a starting point familiar to Allan and Baker (1990) considered the most plant ecologists. These S-A curves can be species’ composition of a series of sites with converted to species-plant number curves and different times after fire in arid Australia. After a then, with a knowledge of the frequency plotting of species richness on a constant area distribution of numbers of plants per species, basis they graphed the dissimilarities (Bray and rank-abundance curves could be derived. Curtis 1957) between every pair of sites against Unfortunately, site specific parameters are needed the differences in times since fire. Dissimilarity to fit the basic relationships so the models lack between sites increased until the differences in universality. The reason for the lack of a solid ages between sites was about 16 years. Thereafter common theoretical base seems to be that the it remained about the same. numbers of plants or other organisms per species One class of changes after fire is that of a set may conform to a variety of model distributions of species which is gradually depleted (by (Gray 1986). Gitay et al. (1991) found either elimination) as time progresses during phase “c” equation (1) or (7) fitted data better for different as in Fig. 3.1. Elimination models have been groups of plants (cryptogams versus vascular developed in quite another context, that of plants). If the number of species in an area of evolutionary changes in species numbers on concern was small (e.g. in some arid areas of isolated mountaintops. If all mountain tops began Australia some years after fire), the use of a with the same composition but were depleted due frequency distribution of numbers of plants per to the differential reduction in habitat area species would be limited. (‘islands’) then, all species present must be subsets Hubbell (1979) proposed that frequency of a grand list of all species in all areas and would distributions of individuals per species (log- represent a sequence from relatively rich to normal or geometric) could be expected to change relatively poor. A “nested-subset” analysis has with scales and commonness of disturbances and been invented to examine such cases (Patterson the chances of local and immigrations. and Atmar 1986). The localized disturbances in rainforests - such as

20 Biodiveristy and bushfires

tree fall - could be contrasted with the major ecological sense of the relationships revealed it is disturbances in higher latitude forests of the necessary to discern ‘functional’ groups of plants, Northern Hemisphere - caused by fires for at least, as these may be expected to have different example. behaviours after fire. To make conservation sense, Noble (1989) found that ranked biomass and the individual taxa involved need to be known, if cover values in mallee communities of not finer levels of plant variation. southwestern NSW conformed to the geometric There are many other indexes (e.g. Lamont distribution soon after fire, a result he attributed et al. 1977) than those considered here and a to the dominance of Stipa, an ephemeral grass. In number of other approaches to the study of a series of stands of increasing times since fire in biodiversity. Trophic methods (e.g. Pimm 1982) jarrah ( marginata) forest, Western and mechanistic models of functional Australia (WA), rank-relative frequency curves relationships (eg. Noble et al. 1988) may be used. moved from the near log-normal type to the No approach, no index, no relationship caters geometric type (Bell and Koch 1980). Simple ideally to the measurement of the diversity of life generalizations relating the type of curve to time in its myriad forms and levels of organization. since fire in a variety of communities seem What we can do is tailor our efforts to the unlikely at present. demonstration and understanding of the diversity Diversity indices and relationships may be of nature as it relates to the purposes and scales at derived and examined without recourse to the hand and within the limits imposed by the extent identities of the organisms being known. To make of our data and knowledge.

21

Biodiveristy and bushfires

3. MODELS OF CHANGES OF PLANT DIVERSITY AFTER FIRE

In this chapter, simple graphical models are developed against which actual data can be compared. Models are useful in that: (i) the expected behaviour of various functional groups can be explored as isolated components of richness rather than as the mixtures found in the field and implicit in literature sources; (ii) the simplifications made in modelling can help focus on major phenomena about which elaborations or variations found in the field can be described.; and, (iii) the full range of variation that seems possible, but which may not yet have been detected and described, may be considered. Attention is first directed to the changes that take place in the richness of plant species (i.e. S) with various times after fire (i.e. t, in years) in a Figure 3.1. An hypothetical curve for species chosen fixed area. After fire, there may be few or richness, S, as a function of time since fire, t, in no vascular plant species obvious to the eye. years. “a”, “b” and “c” represent successive phases of change. However, it is almost inevitable that there will be seeds of some species in the soil and some of these may not have found expression in the prefire flora. Such ‘hidden’ or ‘apparent’ diversity is particularly important in the dynamics of species change after fire. Another cause of change occurs if propagules of plant species, previously unknown at the site, arrive and establish; they may be ‘apparent’ soon after arrival or remain ‘hidden’ until a fire occurs. These species may be native or exotic.

3.1 SPECIES RICHNESS (S-t models)

Curves of species richness with time are the simplest and most common way of representing the changes in communities taking place after fire. Three main periods in S-t curves may be elaborated (Fig. 3.1): “a”, the initial period after Figure 3.2. Alternatives within phases for S-t curves. fire in which species numbers increase from low (a) indicates possible variants in phase “a”. Phase “b” is considered to be either present or absent. (b) values; “b”, the period after “a” which shows a shows either increasing, fluctuating or declining plateau in richness; and “c”, a period of increased, richness in phase “c”. decreased or fluctuating richness. Fig. 3.1(a) shows the variation that could occur in the initial phase, “a”, of re-establishment re-establishment, perhaps after a particularly after fire. With a low-intensity fire, some species severe drought or a peat fire, may occur before re- of trees, for example, may not be greatly affected establishment begins. These variants may be real by the fire and remain as they were - alive and but at this stage of knowledge may be regarded as green. In this case, the curve could begin with a relatively trivial. Thus, only the curve beginning positive intercept. If all above-ground parts of at the origin is considered again here. plants are killed, the apparent number of species Phase “b” , the plateau in species richness, is begins at the origin. In the third case a delay to considered to be in one of two states - either present or absent.

23 Australia’s Biodiveristy - Responses to Fire

Figure 3.3. The five most significant, but still somewhat hypothetical, trajectories for S-t curves.

Fig. 3.2 illustrates a circumstance in which functional dominance of particular species (e.g. the species richness either increases, fluctuates or Noble 1989, Specht and Specht 1989). In some decreases in period “c”. An example of the first of arid zone communities, dominance - reflected by these could arise when a species-rich rainforest cover value - may better be depicted as a function invades a eucalypt forest of lesser richness. A of cumulative rainfall after fire (Griffin 1992). decline in richness during period “c” may occur Thus, rainfall could be a better independent by an invasion of rainforest but the rainforest in variable against which to depict changes in this case would be one that is species poor. richness after fire. Using rainfall in such cases has Fluctuating richness may occur in a community an advantage over the use of calendar years in that with many rain-responding ephemeral species. it reflects the influence of a driving variable rather Fox (1990) showed in mallee communities that than that of a surrogate. Unfortunately, such a composition and numbers of species of basis for the depiction of change does not have ephemerals change from year to year. nation-wide utility. Many species in many In Fig. 3.3 the most likely S-t curves are locations have strongly seasonal rhythms (for illustrated. Any variation in phase “a” has been flowering in particular) that are a function of the removed. Phase “b” is present in the first passage of years rather than related to quantities circumstance but absent in the second. If “b” was of antecedent rainfall. Driving variables other present in the second then it would have the than rainfall are often significant also and are general appearance of the first so is not included. collectively incorporated into ‘year’. In the third example in the Figure, “b” is absent and the richness fluctuates but not necessarily in a 3.2 SPECIES-AREA RELATIONSHIPS regular way; including phase “b” in another (S-A models) variant of this would be unlikely biologically. In the last two illustrations in Fig. 3.3, “b” is present Above, changes in species richness with time since and followed by either increasing or decreasing fire were depicted for an area of constant size. In species richness. These 5 models therefore involve this section, the area constraint is removed by the various trends in apparent richness (first 3) or consideration of the changes in species-area involve apparent richness plus that due to relationships with time after fire. S-A curves are processes of invasion and depletion of the not uncommon but have not been considered as a community (last 2). function of time since fire. In cases where species invade after fire they Species can enter a fixed area or quadrat a must have ‘tolerant’ (T) or ‘requiring’ (R) modes number of ways. ‘Entry’ can come from sources of establishment in that they can tolerate a wide external to the community, such as from range of establishment circumstances - soon after propagules of invading species, or from internal or long after fire - or “require some condition sources. In the latter category are species present in established communities” (Noble and ‘appearing’ from the soil soon after fire and Slatyer 1980). species that appear in the quadrat as the result of Changes in the time-courses of species sampling artefacts. Small plants are less likely to richness after fire are sometimes linked to the be sampled in a quadrat than large plants so, as

24 Biodiveristy and bushfires

Figure 3.4. Species-area curves for three Figure 3.5. Hypothetical rank-abundance or rank- communities representing different stages after fire. cover curves. Species richness increases, and In this example, the species-rich curve, “1”, is that numerical or functional ‘dominance’ decreases, from soon after fire; curve “3” represents a mature curves 1 to 5. community of lower richness; and, “2” is for a community at an intermediate stage. plants increase in size after a fire, they may be considerable number of other species which then more likely to appear in a sampled quadrat. A showed relatively little difference in their small population of a species is also less likely to importance values, a modified form of curve 3 or be sampled than a large one so populations 4 in Fig. 3.5. Curves from species-rich tropical building after a fire may escape sampling earlier rainforests also indicate that the highest ranked but appear later. These circumstances for a single species are considerably more numerous or fixed quadrat can be overcome if a whole series of ‘important’ than all other woody species present quadrats is used (or, less appropriately statistically, even though there is a shallow slope to curves of when nested quadrats are used). importance-value against rank for most of the Fig. 3.4 represents the way the curves might relationship (Hubbell 1979). Bell and Koch’s appear when a species-poor community (“3”) (1980) curves for all showed strong becomes a species-rich community after fire (“1”) dominance for communities at various times after (e.g. Zimmer 1940). With time after fire, then, the fire. Even soon after fire, the results of Noble community would shift from “1” to “3”. (1989) indicated strong cover dominance of an herbaceous species in a community dominated 3.3 RANK-ABUNDANCE AND RANK- structurally by woody plants when mature viz., COVER RELATIONSHIPS mallee. ( i-mi and i-mc models) Changes in the curves with time after fire may show trends because all components of the At any one time after fire, there is a range of curves may change over time viz., the numbers of possible relationship between the number of individuals (or their cover), the numbers of plants of the species present (or their cover, species, and their relationships with each other. biomass or height) and their rank. Fig. 3.5 depicts some of the range that may be possible. The 3.4 RANK TRAJECTORIES (i-t models) curves shown in Fig. 3.5 are indicative rather than definitive, representing a range of possibilities If the rank of a species was plotted against time (after Whittaker 1965 and 1972, Hubbell 1979, since fire we would be one step closer to Bell and Koch 1980, Hughes 1986 and Noble identifying species performances with time after 1989). They show increasing richness and fire rather than considering species anonymously. decreasing dominance (numerical or functional) Identifying the positions of any single species in from 1 to 5. the curves above (Fig. 3.5) is not possible. In a species-rich in , Lamont et al.’s (1977) plot of “importance values” against rank indicated a functional dominance of a few species over a

25 Australia’s Biodiveristy - Responses to Fire

Mature spinifex, Rudall River, WA. As the spinifex is in discreet clumps and discontinuous, it will not easily carry fire and has not been burnt for some time. /A Malcolm Gill

Hot fire in spinifex and acacia, , 1986. / James Eldridge

Post fire, Rudall River, WA, showing regeneration of a variety of plants including ephemerals and a Newcastelia . Immediate post-fire regeneration includes a flush of annuals which reappears in a second year at least but is ‘scanty’ in mature communities. /A Malcolm Gill

Patchy fire, Rudall River, 1994. Such fires in arid areas result in a mosaic of older and newer growth creating a variety of vegetation age classes, species number and structures. /A Malcolm Gill

26 Biodiveristy and bushfires

Late dry season fire at Kakadu, 1990, with grassy fuel and pandanus. Fires at this time of year have a high chance of getting away in most areas of Kakadu. Close to wetlands, however, management fires are lit then as it is too wet at other times. /A Malcolm Gill

Post-fire, low intensity fire, , NSW, 1985. The starting conditions and the nature of the fire can affect the type of regeneration. /A Malcolm Gill

Nothofagus forest gully, central highlands of , 1996. This area has not been burnt for a long period. /A Malcolm Gill

27 Australia’s Biodiveristy - Responses to Fire

Black Mountain, ACT, 1991. Fire behaviour will affect regeneration. /A Malcolm Gill

Resprouting of Eucalyptus maculata (smooth) and Eucalytus globulus (stringy) after the January 1994 fire, Mogo State Forest, NSW. /A Malcolm Gill

Black Mountain, 1984. Litter fuel fire. /A Malcolm Gill

Early post-fire, Nadgee, NSW, 1981, showing weed regeneration. /A Malcolm Gill

Post-January 1994 fire, Oxford Falls, NSW. Flannel flowers regenerating. /A Malcolm Gill

28 Biodiveristy and bushfires

4. FIELD METHODS AND OBSERVED CHANGES IN PLANT DIVERSITY AFTER FIRE

The detection of changes in plant diversity after 4.2 COMPARING ADJACENT BURNT fire have been made in a number of ways. In this AND UNBURNT STANDS chapter, the observations made have been grouped according to the various methods used because Comparing floras in adjacent burnt and ‘unburnt’ each method has its own values and drawbacks. areas at the same time is perhaps the simplest The sites of the studies mentioned are mapped in method of recording change with time after fire. Fig. 4.1 while selected portions of the data are Sites often are considered to be identical shown in Table 4.1. The Table shows raw data physically and biologically prefire. which can be misleading in the absence of the A classic case of this method is that of trends of data sequences. Zimmer (1940) who found that the numbers of species in a northwestern Victorian mallee 4.1 BEFORE-AND-AFTER FIRE community near Hattah had many more species COMPARISONS after the fire than before the fire (Table 4.2). The dominant mallee plants (i.e. Eucalyptus spp.) and Observations made before and after fire provide the main ground cover, of hummock grass ( one of the means of examining changes in sp. or ‘spinifex’), had all foliage removed by the diversity. Christensen and Kimber (1975) fire. Also in north-western Victoria, Cheal et al. provided data using this method in south-western (1979) reported 27 instances of greater species Australia. Species richness of vascular plant richness in burnt cf. unburnt stands of , species in tall open forests (“wet sclerophyll , shrubland, mallee and pine . forests” structurally dominated by eucalypts) In a Western Australian mallee region “620 showed that up to 3 times the number of species km ESE of ” (presumed to be near the were present after fire than before. At the same settlement of Salmon Gums) in which there was time, however, the numbers of species in the not only mallee vegetation but also woodland and ‘untreated’, or ‘control’, sites were declining shrubland, richness was greater in all burnt sites rapidly. In the drier open forests (“dry sclerophyll examined (Van der Moezel and Bell 1984). forests” - “jarrah forest” in Fig. 4.1), however, In the extensive spinifex-dominated numbers showed little change when the ‘unburnt’ grasslands of arid Australia, increases in species area was 6 years free from fire but showed a 20% richness after fire followed the same trend as that downward change for a 38 year period. observed by Zimmer (1940) in the semi-arid Heaths and heath woodlands in mallee. In northwestern WA (Warralong Station) southwestern Victoria near Anglesea revealed all immediate post-fire regeneration included a flush prefire species within 3 years of being burnt of annuals which reappeared in a second year at (Wark et al. 1987); the same was true of a least but were “scanty” in mature communities heathland near Noosa, southeastern (Burbidge 1943). Similar observations were made (Harrold 1979). In the Anglesea case, at least, the by the author and his colleagues P. H. R. Moore, species richness then declined with elapsed time. N. D. Burrows and B. Ward, in hummock grasslands near Rudall River National Park 350 km east-south-east of this area.

29 Australia’s Biodiveristy - Responses to Fire

Table 4.1 (a) Locations of sites for the studies of the Authors cited in Table 4.1 (b). Localities are shown also on the maps (Fig. 4.1 (a) and (b)). “m.a.r.” refers to mean annual rainfall noted by the cited reference or given for station in or near the study area (Bureau of Meteorology 1988). Locations of sites, when not available, have been approximated using information given in the text of the cited reference. “N.P.” means National Park. SITE Latitude Longitude m.a.r. (mm) Reference

Dark Island 36°02’ 140°29’ 450 Specht et al. 1958

Hattah 34°46’ 142°16’ 285 Zimmer 1940

Wilson’s Promontory 38°56’ 146°22’ 1043 Russell & Parsons 1978

Jervis Bay 35°08’ 150°43’ 1170 Ingwersen 1977

Cooloola N.P. 26°05’ 153°02’ 1428 McFarland 1988

Melaleuca Inlet 43°24’ 146°09’ 2400 Brown & Podger 1982

Mooliabeenee 31°18’ 116°03’ 650 Hobbs & Atkins 1990

Myall Lakes N.P. 32°41’ 152°09’ 1352 Fox 1988

N. Stradbroke Is. 27°30’ 153°30’ 1500 Specht & Specht 1989

Uluru N.P. 25°21’ 131°02’ 331 Allan & Baker 1990

Little Desert 36°35’ 141°45’ 421 McMahon 1984

Penola 37°23’ 140°50’ 690 Venning 1978

‘Jarrah forest’ 31°55’ 115°52’ 1200 Bell & Koch 1980

Sandringham 37°57’ 145°00’ 725 Molnar et al. 1989

Little Desert 36°35’ 141°45’ 421 Cheal et al. 1979

Hattah 34°46’ 142°16’ 285 Cheal et al. 1979

Sunset Country 34°39’ 141°48’ 280 Cheal et al. 1979

Wyperfeld N.P. 35°32’ 141°58’ 350 Cheal et al. 1979

Two Peoples Bay 34°57’ 118°11’ 800 Bell et al. 1984

Grampians 37°08’ 142°26’ 448 Cockburn 1978

Lake Mackay 22°13’ 128°43’ 200 Burrows & Christensen 1990

(S. of) 23°42’ 133°52’ - Allan & Griffin 1986

Herberton 17°19’ 145°26’ 1800 Unwin 1989

Anglesea 38°25 144°11’ 657 Wark et al. 1987; Wark 1997

Corinna 41°39’ 145°05’ 1952 Barker 1991

Salmon Gums 32°59’ 121°39’ 375 Van der Moezel & Bell 1984

North Head 33°50’ 151°18’ 1212 Clemens & Franklin 1980

Sorrento 31°50’ 115°45’ 711 Fox et al. 1983

30 Biodiveristy and bushfires

Table 4.1 (b) List of sites in which species richness has been examined as a function of time since fire. Refer to Table 4.1 (a) for the details of location and sources of information. Where there has been more than one investigator at a stated location, the Author of the paper is noted in brackets. “L.Desert” refers to Little Desert in Victoria (Fig. 4.1); “McM” refers to McMahon; and “C&K” refers to Christensen and Kimber. “Jarrah forest” refers to a vegetation type not a town. The fire response is noted according to the number given by Gill and Bradstock (1992). “1” refers to species in which mature plants are killed by canopy-killing fire but have seeds stored in the canopy; “8” refers to ‘seeder’ species generally; and “9” refers to ‘sprouters’. “Smax” refers to maximum species richness reported; “Smin” refers to the minimum species richness reported; “tmax” is the time in years at which the maximum richness was reported; and “tmin” is the time in years when the minimum richness was reported. “b” refers to the site being burnt relatively recently while “u” refers to an “unburnt site” ie. one with a relatively long period since fire.

SITE Vegetation Dominant Fire Smax tmax Smin tmin response (yrs) (yrs) Dark Island heath ornata 1361?2025 Hattah (Zimmer) mallee Eucalyptus spp. 9 63 1? 18 - Wilson’s Prom. heath pusilla 9 60 1 49 21 Jervis Bay heath L. laevigatum* 8311224 heath L. laevigatum 8252154 forest E. pilularis 9221184 woodland E. sieberi 9 52 1 50 4 woodland E. racemosa 9 45 1 25 4 woodland E. piperita 9482434 wet heath B. ericifolia 8422374 dry heath Casuarina distyla 8482414 Cooloola N.P. dry heath - - 44 5 33 2 wet heath - - 40 1 29 2 Inlet sere Nothofagus cunninghamii 9 25 <6 8 300 Mooliabeenee woodland B. attenuata 8 31 5 20 45 Myall Lakes N.P. open forest E. pilularis 9 37 2 35 3 N.Stradbroke Is. heath-woodland E. signata 9 25 4 18 19 Uluru N.P. hummock grass’d T. basedowii 9 60 6 22 38 hummock grassl’d T.pungens 9 52 5 29 12 L. Desert (McM) heathland B. ornata 1 42 1 25 35 Penola woodland E. baxteri 9 46 2 37 25 Jarrah forest forest E. marginata 9 31 5 20 46 Sandringham George St. scrub L. laevigatum 8 36 2 9 >70 Bay Rd. scrub L. laevigatum 8 45 1 9 >50 L. Desert(Cheal) heath B. ornata 1 35 2 18 35 Hattah (Cheal) mallee E. socialis 9 37 3 23 >40 mallee E. incrassata 9 34 3 30 >40 mallee E. incrassata 9 56 5 40 >40 mallee/w’dland E. oleosa 9 32 5 15 >40 mallee E. incrassata 9 45 5 38 >40 Sunset Country mallee E. dumosa 9 45 4 30 17 mallee E. incrassata 9 49 4 31 17 mallee E. dumosa 9 30 7 18 >40 mallee E. dumosa 9 33 7 21 >40 mallee E. incrassata 9 46 1 24 >40 mallee E. dumosa 9 34 1 25 >40 mallee E. incrassata 9 45 1 42 >30 mallee E. socialis 9 48 1 22 >40 mallee/w’dland E. dumosa 9 31 1 22 >40 mallee/w’dland E. oleosa 9 44 1 43 >40 mallee/w’dland E. oleosa 9 31 1 14 >40 mallee/w’dland E. dumosa 9 36 1 22 >40 mallee E. socialis 9 51 1 37 >40 broombush/mallee Melaleuca uncinata 9 44 1 35 >4

*Leptospermum continued over

31 Australia’s Biodiveristy - Responses to Fire

Table 4.1 (b)

SITE Vegetation Dominant Fire Smax tmax Smin tmin response (yrs) (yrs)

Wyperfeld N.P. heath B. ornata 1 26 19 22 >40 pine w’dland 8 37 19 29 >40 mallee E. dumosa 9 32 <1 30 >40 mallee E. dumosa 9 30 <1 19 >40 heath B. ornata 1 28 <1 26 >40 mallee E. incrassata 9 48 <1 42 >40 mallee E. dumosa 9 51 <1 34 >40 shrubland M. uncinata 8 26 <1 18 >40 Two Peoples Bay heath - - 78 1 67 5 Grampians heath L. myrsinoides 9 31 9 18 15 Lake Mackay grassland Triodia spp. - 26 2 5 33 S.of Alice Springs grassland T. basedowii -177510 Herberton area rainforest - - 81 - 22 - open forest E. grandis 958-24- S.W. Aus.(C&K) open forest E. marginata 9411?38>6 E. marginata 9 30 2 24 40 tall open forest - - 46(mean) 1? 14 4? - - 36(mean) 1? 7 9? - - 19(mean) 1? 9 >25 Anglesea heath L. mysinoides 9671543 open scrub E. obliqua 9583421 tall shrubland E. obliqua 9531523 woodland E. obliqua 9451393 woodland E. obliqua 9411373 woodland/scrub E. obliqua 9533431 shrub/w’dland E. baxteri 9641563 shrub/w’dland E. willisii -343241 heath L. juniperinum -423401 shrub/w’dland E. obliqua 9411363 open forest E. tricarpa 9 58 1 40 10 “gully complex” E. cypellocarpa 9 53 1 31 10 Corinna rainforest - - 117 <9 61 >78 Salmon Gums mallee - - 26 b19 u w’dland/mallee - - 20 b15 u woodland - - 18 b13 u shrubland - - 8 b 5 u North Head heath L. laevigatum 8 36 2 13 10 heath Restio fastigiatus - 36 2 26 10 Sorrento heath - - 61 b39 >9

32 Biodiveristy and bushfires

Figure 4.1. Locations of sites indicated in Table 4.1 and in the text. (a) Australia; (b) southeastern Australia.

33 Australia’s Biodiveristy - Responses to Fire

In coastal heathlands just north of Perth, Banksia ornata. The S-t curves reported by all of WA, (Sorrento) burning led to a 50% increase in these authors were similar to that of the second the number of species recorded (Fox et al. 1983). diagram in Fig. 3.3 with a short phase “a”. In The burning of scrubland invaded by the native McFarland’s (1988) study area in southeastern shrub (a ‘seeder’) during Queensland (Qld) (Cooloola National Park) there historical times in (Sandringham), was a similar short build-up time in apparent Victoria (Vic.), “dramatically increased the species richness (1 year) but this was followed by number of native species present” (Molnar et al. very little change in richness in sites 1989). A heath dominated by the same species at representative of the next 10 years. The changes North Head in Sydney, New South Wales in the first decade were similar to those noted by (NSW), also revealed more species once burnt Russell and Parsons (1978) for a wet heath in (Clemens and Franklin 1980); the same authors southern Victoria (Wilson’s Promontory) but after recorded the same phenomenon for the adjacent that period there was a slow decline in richness Restio-Leucopogon heath. from near 60 species at age 10 years to 49 species Burnt temperate rainforest areas in at 21 years after fire; the authors noted that the northwestern (near Corinna) showed trend was possibly confounded by site differences. higher richness than areas that escaped fire but Studies in the heathlands of the Grampians in there appeared to be an inconsistent effect of fire western Victoria suggested a much longer build- intensity on richness as well (Barker 1991). up period (perhaps 9 years) followed by the typical decline reported by other authors (Cockburn 4.3 CHRONOLOGICAL ORDINATION 1978). USING SITES WITH DIFFERENT AGES In the (in WA near Lake AFTER FIRE Mackay), Burrows and Christensen (1990) found only 5 to 7 species in stands more than 20 years An elaboration of the method described in the old but up to 26 species in stands burned 2 years previous section occurs when a variety of sites, before examination. Allan and Baker (1990), burnt at different times in the past, are examined. similarly, recorded a sharp drop in species richness This method has been called “chronological with time after fire in arid spinifex grassland but community ordination” (Gill 1977). While quick only after a slow rise that took about 6 years (i.e. and convenient, it has the disadvantage that it may phase “a” of Fig. 3.1). A double-humped curve be hard to find study sites which vary widely in was published by Allan and Griffin (1986) for arid fire history but are uniform in site characteristics. (south of Alice Springs), the In the best use of the technique, the effects of site second rise in richness being associated with an would be discriminated during statistical analysis. invasion by trees and beginning between 9 The method implies that there has been no affect and 27 years after fire. of season of fire occurrence or of fire intensity or In a eucalypt woodland in southeastern that these variables were uniform across all plots. (near Penola), Venning’s (1978) It assumes that all the plots had the same recent data suggested that there was a rapid restoration fire history (burnt at the same time previously) or of species richness to an average of about 43 that this feature was unimportant to the species in the first year after fire followed by a comparisons. period up to year 25 in which there was a slight Specht et al. (1958) revealed a decline in decline to an average of about 40 species; such a species richness with time since fire in a Banksia slight decrease is probably not statistically ornata heathland near Keith in South Australia significant. Presumably there was little or no (SA). “Of 36 species recorded after a fire, only 20 change in species composition during this time were found after 25 years” and “Probably only ten either. In a Banksia woodland north of Perth (at of these would persist after 50 years” (Specht et al. Mooliabeenee), WA, species richness rose after 1958). In the Little Desert in Victoria Cheal et al. fire to peak at 5 years, with 31 species per 25m2 (1979) and McMahon (1984) were able to quadrat, then declined to about 20 species ≥44 demonstrate the same trend in similar vegetation years after fire (Hobbs and Atkins 1990). types. Later Specht and Specht (1989) showed that species richness was a linear and negative function of the foliage cover of the ‘seeder’ species

34 Biodiveristy and bushfires

A curve for species richness in a “jarrah 4.4 PLOTS, IN WHICH CHANGES forest” was constructed in the Darling Range near ARE OBSERVED IN SITU Perth by Bell and Koch (1980) using a series of forest sites from different forestry ‘blocks’. There, Changes in species richness can be observed either richness peaked at 30 at 5 years then declined to in plots being monitored after fire or in those 19 in stands 50 years since last fire. However, integral to an experiment. The advantage of species composition varied a great deal between experiments is that there is a known starting point sites. from which changes may be observed. The That “There is no evidence that fire alters differences between experimental plots can be the floristic composition” of E. obliqua forests in described and accounted for in assessing the the Adelaide Hills, SA, region (Wood 1937) may results of any change (e.g. Bowman et al. 1988). mean that there has been no change detected from Effects of seasons of burning may be measured as the range of fire regimes experienced there or that well as effects of intensities. Disadvantages may be there has been no change with time after fire (or that: the range of vegetation types chosen may be both). Cochrane et al. (1962) leant towards the limited by the experimenter in order to enable the latter interpretation when considering pyric burning of the plots according to the experimental succession in the same region (near Crafers) in plan; fires in plots during critical fire-danger that within “seven to ten years the sere is periods may not be allowed by authorities; and, completed ... the composition of the sclerophyll experimental fires cannot reflect the full range of understorey is remarkably similar”. However, intensities encountered in the forest (Gill 1977, there were changes in the dominance of the Gill and Moore 1990). There can also be understorey species during this period. interactive effects between fires and grazing On sandy soils of coastal NSW in Myall animals, native or exotic (Leigh and Holgate Lakes National Park, an open forest of E. pilularis 1979, Hesp et al. 1983, Leigh et al. 1987, Noble showed a maximum richness 2 years after fire. 1989). The advantage of plots set up after fire (e.g. Richness then declined to 3 years or so whence an Ingwersen 1977, Wark et al. 1987, Wark 1997) is increase occurred again to 16 years, the longest that there is no question of any apparent change time after fire in the sequence (Fox 1988). In a being a consequence of site sampling. However, in nearby area, a greater number of species was both experimental and post-fire plot observation, recorded in a younger stand (Fox and Fox 1986). sampling problems can arise when species grow The longest time sequence examined was in into plots thereby artificially increasing numbers. wet southwest Tasmania in the vicinity of In mallee vegetation in southwestern NSW Melaleuca Inlet. There, Brown and Podger (1982) near Pooncarie, large numbers of herbaceous linked changes from sedgeland-heath to woodland species, not apparent before the fire, appeared to “mixed forest” and rainforest with time since after it (Noble 1989). Numbers of species were fire. After the initial buildup of richness to a peak boosted by post-fire rainfall conditions. A mulga of 24 species in less than 6 years, the richness shrubland (dominated by the ‘seeder’ Acacia generally declined to about 8 species in the ca. aneura) in arid Central Australia (Alice Springs 300-year sequence. There was a possible outlier in region) showed increased richness one year after the data at 19 years which indicated a temporary fire but subsequent declines in the following three increase in richness. years (Griffin and Hodgkinson 1986). This trend was evident in the control as well as the treated plots but the treated plots showed more species in years 2 to 4 than the control. A neighbouring “intergrove” community (possibly dominated by tussock grasses) showed parallel trends in richness to the mulga areas but revealed no difference between treatments. In “sclerophyll woodland” or “heathy woodland” in southeastern Queensland (Stradbroke Island), numbers of species per 10m2 declined from 26 to 18 in a 19-year period after fire (Specht and Specht 1989 p. 342).

35 Australia’s Biodiveristy - Responses to Fire

Ingwersen (1977) examined species numbers (Wilson et al. 1992). Along the putative time in plots for 4 years after a large unplanned fire in a gradient, the floristics simplify (Gilbert 1959). number of heaths, scrubs, woodlands and forests Jackson (1968) expanded the length of the sere at Jervis Bay south of Sydney. Peak richness were ending in rainforest by including a number of attained within 2 years. Numbers of species were further vegetation types beginning with sedgeland maintained or dwindled in the following two years or grassland. Brown and Podger (1982) quantified in most cases but the decline in the E.racemosa- species richness of the expanded sere in southwest E.gummifera woodland was quite marked. Tasmania (Melaleuca Inlet). Howard (1974) ‘Control’ areas were unavailable for comparison in identified a sere from grassland to rainforest Ingwersen’s study area. through Acacia stands in the northwest of the Richness in a southwestern Australian heath State. Ellis (1985), also in northwestern Tasmania, at Two Peoples Bay peaked at 78 species one year found evidence for the penetration of rainforest after fire and then gradually declined to 67 at year into eucalypt communities and into shrublands 5 (Bell et al. 1984). which, in turn, were previously grasslands. Near Canberra, Australian Capital Territory The situation in eastern Victoria in Mitchell (ACT), Purdie (1977) found that the species River National Park seems similar to that composition remained the same after burning described for Tasmania to the extent that there “dry sclerophyll vegetation” (‘open forest’ of seems to be a trend from wet sclerophyll forest (or Eucalyptus spp.). This appeared to be the case also ‘tall open forest’) to rainforest (Melick and Ashton for eucalypt forest and woodland in the northern 1991). Melick’s (1990) ordination suggested that Northern Territory (Munmarlary) where burning undisturbed rainforest (27 quadrats) had fewer treatments were as frequent as annual in species than the fire-disturbed vegetation (30 occurrence (Hoare et al. 1980, Bowman et al. quadrats). 1988). Tolhurst and Oswin (1992) studied the In tropical Australia eucalypt communities effects of a range of fire regimes in a Victorian are being replaced by rainforests in some areas. forest ( State Forest) and reported that Harrington and Sanderson (1994) using “No species was gained or lost from any photogrammetry found that wet sclerophyll forest treatment”. dominated by E. grandis was being replaced by rainforest. Unwin (1989) observed the rate of 4.5 INFERENTIAL METHODS advance of rainforests in such situations to be 1.2 m yr-1 over a 10-year study period; species Some changes in richness may take so long that richness of the eucalypt forest was much less than experiments seem impracticable and dating of that of the rainforest so the richness was stands is beyond the reach of photogrammetric or increasing during succession in his sites near recorded history. In these sorts of cases inference Herberton. may be used to piece sequences together. The Across northern Australia are ‘monsoon prime examples are for rainforests originating forests’, a type of rainforest often dominated by from eucalypt-forest or eucalypt-woodland sites. one or a few species; it may be regarded as a less- The sites in question may lie anywhere on the rich type of tropical rainforest. Fires may cause coastal and subcoastal areas of eastern and the retreat of its boundaries but, when northern Australia (as far west as The Kimberley). undisturbed, re-establishment at patch margins In Tasmania, Gilbert (1959) drew attention can occur (Bowman and Fensham 1991) over to the trend from tall open forest to rainforest in decades (Russell-Smith and Bowman 1992). his studies in the Florentine Valley. Further study Monsoon forest patches, surrounded by savanna in the same State has reinforced this: “the major (or, ‘grassy eucalypt woodland’) may contain more factor determining the nature of the rainforest- tree species (Bowman 1992) but fewer vascular eucalypt forest boundary is fire” (Neyland 1991). plant species overall (Taylor and Dunlop 1985). This is, or implies, a species-replacement argument for succession. Mount (1979), however, favoured the idea that “the pattern is one of stable fire cycles rather than one of succession”, perhaps a sort of “initial species composition model” (Egler 1954) or “complete initial floristics model”

36 Biodiveristy and bushfires

4.6 DISCUSSION A number of patterns of change in species richness may be detected from the data presented: The various methods used to discover changes (i) the numbers of vascular plant species in most taking place in species richness after fire all yield communities soon after fire rise to levels useful data. However, the data needs to be higher than those occurring immediately interpreted in the light of the methodology used. before the fire; In many cases, the actual species present are not (ii) numbers rise to reach a peak often within enumerated so whether or not the quoted richness one year but sometimes after 2 or more; for a number of sites have taxonomic and (iii) changes appear most conservative in open ecological equivalence cannot be judged. By the forests of Eucalyptus (‘dry sclerophyll forests’) choice of technique, authors may assume that (Christensen and Kimber 1975, Purdie 1977, species-richness changes occur in sites with: exact Venning 1978, Hoare et al. 1980, Bowman et ecological equivalence; identical initial species al. 1988, Tolhurst and Oswin 1992), possible composition; the same fire circumstances and exceptions being some of the forests with properties; and, equivalent post-fire heathy understoreys (presumed to be the environmental conditions. case in forests studied by Ingwersen 1977, The data reported here concern only Fox 1988 and Wark 1997); changes in species richness after fire. They do not (iv) heaths showed either no change in richness concern extinction (Gill and Bradstock 1995). In with time after fire (such as those studied by most cases there is no indication whether or not McFarland 1988 in southeastern the changes identified are a simplification of an Queensland) or decreases as in Specht et al. initial complement or whether additions and (1958), Ingwersen (1977), Cheal et al. depletions have occurred. However, most species- (1979), Clemens and Franklin (1980) and richness depletions in phase “c” of Fig. 3.1 may be Wark et al. (1987); assumed to be due, as a first approximation, to (v) rapid decreases soon after fire occurred in species deletions without any addition. Where arid and semi-arid communities (especially there has been an invasion of rainforest into other hummock grasslands and mallee) (e.g. communities, substantial or total species Zimmer (1940), Cheal et al. (1979), Burrows replacement may take place. and Christensen (1990), Noble (1989) and The range of plant communities directly Allan and Baker (1990); and, affected by fire is large. Communities affected (vi) species-replacement sequences after fire may include those in arid zones (e.g. Burbidge 1943, involve increases (species-rich rainforest - Allan and Griffin 1986, Allan and Baker 1990, Unwin 1989) or decreases (species poor Burrows and Christensen 1990) through to high rainforests - Brown and Podger 1982) in rainfall areas dominated by temperate (eg. Barker richness. 1991, Melick and Ashton 1991) and tropical (e.g. Observations of an absence of change in Ridley and Gardner 1961, Bowman and Fensham richness may be due to intervals between ‘burnt’ 1991) rainforests. Most, if not all, terrestrial plant and ‘unburnt’ being too short. Thus one of communities seem to be affected. Christensen and Kimber’s 1975 examples indicated a significant change over a 38 year period but comparisons over a relatively short period showed little change.

37

Biodiveristy and bushfires

5. INTERPRETATION OF DATA

In this chapter, the main aim is to explain the data identifications of taxa) and interactions between presented. The main hypothesis to emerge is that sites and treatments Thus, the importance of there are gradients in time sequences of species ‘appearances’ and ‘disappearances’ of species richness after fire. Thus, we may hypothesize that: cannot be accurately assessed. However, enough is in the arid zone, there is a flush of herbaceous known to hypothesize which patterns are likely to ephemeral species after fire (hidden diversity) be present and where they may occur. which increases species richness only temporally It is likely that most patterns of the form of as richness declines quickly as hummock grasses Fig.3.3 (b) - a rapid rise in numbers after fire gains cover; in open forests of relatively mesic followed by a decline almost immediately - are of areas there is either no decline in richness after the chronologically-nested type while patterns fire or else there is a slow one; in the most mesic like those in Fig. 3.3 (d and e) involve replacement sites (well drained) there is a species replacement of eucalypt-forest species by rainforests of various sequence from tall open forest to rainforest in richness. However, the long sere identified in which the richness increases or decreases Tasmania from sedgeland to rainforest (Brown according to the site suitability for rainforests. An and Podger 1982) shows a downward trend in apparent coarse-grained pattern based on rainfall species numbers during a long species- could be modified according to vegetation type, replacement series. That such a species fire history, functional groups and soil types replacement series ending in rainforest can start perhaps. Formal testing of such ideas would be from a number of communities was indicated by valuable. Howard (1974) and Ellis (1985). Furthermore, There are two main causes of the patterns in these observations do not preclude a self- richness after fire. The first is one that involves no replacement sequence within these communities addition of species during time after fire, merely with time after fire as evidenced by the changes in an absence of change or a simplification in species’ species richness in rainforest after fire described composition. The second is one that involves by Melick (1990) in eastern Victoria. species’ change (addition and/or replacement) at some time during the sequence of changes that 5.1 PATTERNS OF RICHNESS AFTER take place with time after fire. The first is a form FIRE: INVASION-CAUSED CHANGES of “complete initial floristics model” (Egler 1954, IN COMPOSITION Wilson et al. 1992) while the second conforms to Clements’ “relay floristics” model (Wilson et al. Some of the variants in compositional changes 1992). The first may be seen as having that may take place in relation to environment are chronological nestedness while the second illustrated in Fig. 5.1 (after Jackson 1968 and involves gains and losses in species with time. A Mount 1979). Compositional changes are variant of the first is due to the expression of represented by “communities”. The diagram herbaceous ephemerals as a consequence of shows four contiguous communities arranged weather variations from year to year. Examples of along an environmental gradient. “Community 4” the first of these two extremes are common in may be regarded as rainforest, “community 3” as mallee, hummock grasslands and heathlands. The eucalypt forest and the other two communities as second ‘extreme’ occurs when rainforest invades earlier stages in the possible sere or habitat eucalypt forest but may also occur when exotics sequence. In the first case, (a), the communities invade or native dominants like Leptospermum are static in species composition but may pass laevigatum (Russell and Parsons 1978) spread. through a series of phases according to the time Only broad identification of patterns can since the passage of fire. In the second case, the currently be achieved because declarations of communities can spread across the environmental species’ compositions are often not given when gradient, the only limitations to spread in the long shifts in species richness are described. The term being the fire regimes in the communities reasons for this may include the lack of present. Also, in this case, the rainforest may determination of names (despite the spread into all the other communities. There is

39 Australia’s Biodiveristy - Responses to Fire

considerable evidence for the expansion of 5.2 PATTERNS OF RICHNESS AFTER rainforest in the absence of fire into communities FIRE: SIMPLIFICATION, NO of eucalypts from southern Australia (e.g. Ellis INVASION 1985) to northeastern (eg. Harrington and Sanderson 1994) and northern Australia (e.g. Gill In communities in which simplifications in et al. 1990, Russell-Smith and Bowman 1992) richness occur with time after fire and in which thereby supporting the second type of pattern in community compositions show chronological Fig. 5.1. In the wetter parts of Tasmania, the nestedness it is tempting to suggest that all losses second case has considerable support (see in apparent species richness are due to the rise of Bowman and Jackson 1981) but if the scale of the dominant species. Specht and Specht (1989) environmental gradient is widened then aspects of showed that there was a negative correlation the first part of the diagram achieve greater between species richness in a heath community significance. with the cover of B. ornata, the community dominant. In the mallee studied by Noble (1989)

Figure 5.1 Four contiguous plant communities spread across an environmental gradient are illustrated by solid lines. (a) Communities are confined to particular fixed environmental ranges. Fires - shown by arrows - ‘recycle’ the communities (after Mount 1979). (b) Horizontal dashed lines represent habitat suited to the geographic expansion of each community in the absence of fire. Fires cause communities to change and move in composition towards that of community 1. In the absence of fire, composition moves toward that of community 4 (‘rainforest’) such that, eventually, community 4 could occupy the entire gradient (after Jackson 1968).

40 Biodiveristy and bushfires

the post-fire community was dominated by Stipa, which there are hummock grasslands, Acacia a grass; as the community matured one would shrublands, mallee etc.. The second pattern, to a expect a return to the dominance of eucalypts. large extent, seems characteristic of open eucalypt This latter example is the simplest case of what forests (‘dry sclerophyll forest’) where rainfall is may be termed ‘relay dominance’ where successive higher and the third pattern seems to be most waves of dominant species occur through time common in the highest rainfall areas where tall- even while a simplification of community open forests of eucalypts occur together with composition occurs. Such trends would be rainforests. Rainforest richness changes with obvious in rank-trajectory curves (see Section 3.4). decreasing rainfall, increasing seasonality of Zimmer (1940) ascribed diminishing species rainfall, decreasing soil fertility, decreasing richness during period “c” (Fig. 3.1) in a mallee temperatures and increasingly impeded drainage community to increasing water use by the (Kikkawa et al. 1981) so a temporal trend to dominants - the mallee eucalypts and the increasing richness in phase “c” may turn to one hummock grass Triodia - while Noble (1989) of decreasing richness relative to that of the suggested other possibilities for the effect. communities invaded along these environmental Functionally removing the dominant species could gradients. remove allelopathic suppression of other species The geographic model above is a caricature while enhanced post-fire nutrition could allow by definition. It does not reflect the detail expression of some ephemerals. In tall open perceived in some areas such as the wetter parts of forests, post-fire environments may have more Tasmania, for example, but is an hypothesis. It light, more nutrients, lower disease potential and reflects what may be expected generally across a reduced herbivory (Ashton 1981). vast continent with a plethora of species and Rank-abundance and rank-cover curves numerous vegetation and soil types. There are suggest that there is a whole hierarchy of many areas still unstudied (Fig. 5.1). dominance and that the extent of the differences Another way of examining the spatial between first and second rankings - and others - patterns in apparent richness after fire is to may vary. The shallower the slope of the curve examine what is known of the distributions of (Fig. 3.5) the greater the spread of ‘dominance’ relevant functional groups. In Section 2.3, three and the greater the potential for spatial variation functional groups were identified in relation to in dominance to be expressed. Dominance is fires. They were the ‘sprouters’, ‘seeders’ and undoubtedly important but its extent and ‘ephemerals’. Later (Section 3.1) ‘invader’ species effectiveness are important issues to resolve. Is a were identified as ‘tolerant’ (T) or ‘requiring’ (R) species which is most abundant or ‘important’ species. These T and R species may be ‘seeders’ or with a high relative density, but at a low absolute ‘sprouters’ (see Bowman 1991 for northern density (in a post-fire community for example), a Australian examples, Barker 1991 for Tasmanian functional dominant? Is a species which is examples) but presumably such ‘sprouters’ in currently first ranked but being replaced by the advancing rainforests have a relatively low second-ranked species behaving as a dominant? tolerance to repeated fires. In Victorian heathlands, invasion by 5.3 DISTRIBUTION OF PATTERNS OF Leptospermum laevigatum, a tall native ‘seeder’ RICHNESS CHANGE shrub, has been studied in relation to its effect on species richness (Molnar et al. 1989, Cheal 1996). The distribution of patterns of species richness The invasion of eucalypt forest understoreys by with time may be examined by first simplifying the bird-dispersed ‘seeder’ shrub Pittosporum the main patterns (Fig. 3.3) to three: (i) the undulatum has been noted in Victoria (Gleadow pattern of general decline in richness in phase “c”, and Ashton 1981). The exotic ‘seeder’ species a pattern in which composition through time is Pinus radiata has been invading forests in the ACT ‘nested’; (ii) the pattern in which phase “b” is and northern Victoria (see Gill and Williams present but “c” apparently is not, such that the 1996). There are other examples cited by Cheal composition of the community does not change; (1996) who also noted the commonness of and, (iii) the pattern in which species rise or ‘seeders’ among the invaders. As P. radiata and P. decline in phase “c” due to species’ addition undulatum, at least, will invade without fire and/or replacement. The first pattern seems to be disturbance, they may be regarded as ‘T’ or ‘R’ the pattern of the arid and semi-arid zones in species in some circumstances.

41 Australia’s Biodiveristy - Responses to Fire

Ephemerals constitute a large part of the ‘Sprouter’ species are very widespread in increased richness observed after fire in arid and distribution. They are found in rainforests semi-arid communities. A few ephemerals occur (Bowman 1991) and deserts (Griffin 1984). even after fire in tall open forests (Ashton 1981) Sprouters are the most common fire-response and in rainforests (McMahon 1987). They appear type in the savannas of northern Australia (Gill et to be rare in the drier eucalypt forests. al. 1990). Many species of the open eucalypt forest Ephemerals are usually herbaceous in arid areas, of southeastern Australia, including the woody in heathlands. dominants, also fall into this category. ‘Sprouters’ It would be interesting to know whether or with no observed seed regeneration are rare but not the richness of ephemeral species was related have been recorded in Victorian forests to the extent and duration that bare ground was (McMahon 1987). exposed in different plant communities during a ‘Seeders’, like ‘sprouters’, are widespread. fire rotation. Communities dominated by trees Many of the tall eucalypt forest dominants are that develop new shoots throughout their crowns intolerant ‘seeders’ (Ashton 1981) as are many of soon after fire, or which retain their canopies after the dominants of heathlands (e.g. Banksia ornata, a fire of low intensity (e.g. open forests of Specht et al. 1958). Indeed, 50-60% species in tall eucalypts), may represent the circumstance where open forests are ‘seeders’ (Ashton 1981) and full exposure of bare ground to the sun is least. In comprise over 60% of species in mallee, woodland arid areas - where fires can be quite common - and shrubland in southwestern Australia (Van der and bare ground is always exposed, the duration of Moezel and Bell 1984). In a Victorian heathland exposure of bare soil to full sun may reach a 33% species were ‘seeders’ (Wark et al. 1987), a maximum. percentage within the 20-40% found for heaths in Perhaps ephemeral richness could be related southeastern Queensland (C. Sandercoe personal to the evolutionary radiation of particular taxa and communication). Woody ‘seeders’ are not their environmental tolerances? Solanum spp. can uncommon in the arid zone (Griffin 1984) and are be ephemerals from rainforest (McMahon 1987) often dominants e.g. Acacia aneura, Thryptomene to desert but seem more common in the latter maisonneuvii and Callitris spp. ‘Seeders’ are spread environments (personal observation). C4 grasses across the landscapes of Australia, their may be expected in the north, C3 in the south; proportions varying from place to place. Aristida spp. seem more frequent in the north while Stipa spp. seem more frequent in the 5.4 A PLANT-COMMUNITY southern, temperate, half of Australia. Herbaceous DYNAMICS VIEWPOINT (e.g. Ptilotus, Gomphrena) seem more common in the northern parts of Australia The patterns observed in any one community and herbaceous Asteraceae (eg. Helichrysum, depend on the dynamics of the species in that Helipterum) in the southern. community, the species within dispersal distance A guild of ephemeral species will not that can invade, and various aspects of the physical necessarily be expressed in toto at any one time. and biological environments at those sites. In Indeed, the results of Fox (1990) suggest that the trying to understand the patterns observed, then, composition fluctuates a great deal from year to one might turn to the study of the detailed year. A number of authors have noted that in arid dynamics of the communities of interest. There is and semi-arid lands ephemeral forbs appear after no intention to attempt an analysis of such studies winter rain while grasses respond to summer rains here. Rather, it is to be pointed out that the (Griffin and Hodgkinson 1986). Will ephemerals dynamics of any of the communities studied will in northern arid lands, where rainfalls are involve the compositions and life cycles of the predominantly in summer, be largely grassy soil- propagule pool, the ‘apparent’ species pool, therefore? and, the within-dispersal-distance pool (see Fig. 6.1 in the next chapter).

42 Biodiveristy and bushfires

6. FIRE MANAGEMENT OF PLANT DIVERSITY

From earlier chapters of this Report, it is apparent Fire management involves both the that fires are a major factor affecting the plant suppression of unwanted fires and the ignition of diversity of terrestrial Australia. Fires can affect other fires under specified conditions for specified the expression of diversity and the distribution of purposes. It may also involve attempts to control diversity. Inappropriate fire regimes (Gill 1975, unwanted ignitions through education campaigns 1981) can cause the local extinction of plant and the closure of various areas at times of species (Gill and Bradstock 1995). As fire is a ‘extreme’ fire danger. management tool, as well as a natural agent of The fuels for management fires arise from community dynamics, management fires can be the extant plant community in most cases. [In used to manipulate biodiversity. Understanding some circumstances, peat derived largely from fires and their effects is important if biodiversity is communities long gone, may be the main source to be understood and properly managed. of fuel for unplanned, unwanted fires.] The fuels The fire-management aims for any parcel of for management fires come from a select group of land - implicit or explicit - are a subset of general species, not from all species making an equal management aims. General aims often include the contribution. The “fuel species” (Fig. 6.1) are maintenance of species diversity and the usually dominants but not necessarily so. Major protection of lives and property. To effectively fuels in Australian plant communities are supplied manage the ecological estate, an assessment of by grasses such as Themeda spp., Sorghum spp., diversity, especially richness, is necessary. How Heteropogon spp. and Triodia spp., by various this can be done and the results used in the sedges and shrubs, and by tree and shrub litters development of a monitoring and decision- (especially those of eucalypts). Of course, support system has been suggested by Gill and unintentional fires may be fuelled by cereal crops, Nicholls (1989). To assist in the development of pastures and plantations as well. In the depiction such systems, a data base listing the fire-responses of an ecosystem being managed by fire, the fuel of Australian vascular plant species has been species need to be distinguished (Fig. 6.1). Fuel initiated (Gill and Bradstock 1992). species can be manipulated by fire to increase

Figure 6.1. A simplified conceptual diagram illustrating inter-relationships between groups of species in a fire- prone environment.

43 Australia’s Biodiveristy - Responses to Fire

their productivity (e.g. by killing shrubby the end of phase “a”, burning may be wrongly dominants thereby allowing a grass response) or scheduled then in order to keep the intervals decreasing it (e.g. by killing the fuel species as has between fires short (to obtain maximum control of been done with Sorghum in the Northern unplanned fires) and keep the diversity at a Territory - Stocker and Sturtz 1966). maximum. This procedure, however, would be a Increasingly, exotic species are forming parts mistake because certain species of ‘seeders’ may be of Australian plant communities. Exotic plant made to become locally extinct due to species may be expected to provide fuel species and inadequacies in the time necessary for the supply members for all the functional groups identified in of seed for subsequent regeneration (Gill and earlier sections. Exotic species may enter the soil Bradstock 1995). On the other hand, if the seed pool and remain dormant until the appropriate maintenance of a diverse set of stages within the conditions occur for regeneration or they may spectrum of change after fire is seen as the invade the plant community just as other ‘requiring’ objective of management then it is possible that all species (Noble and Slatyer 1980) may do so (Fig. the species in the system could be maintained. 6.1). In conservation management, exotic species are Burning would still have to be within a domain of an unwanted part of the diversity of biota. fire regimes commensurate with the survival of all If, as asserted here, exotic species comprise species. Ways in which the variety of times since members of the same functional groups as the fire, or post-fire stages, could be achieved include: native biota, opportunities for the elimination of (i) burning at a fixed, relatively long, interval such species using fires while retaining the native but in different places each year; species may seem slim. However, if the exotic is (ii) burning at different fixed rotations in the only species in a particular functional group at different burning blocks within an a particular site or has extreme attributes making ecologically safe range of intervals; and, it the most vulnerable in a class, then the use of (iii) as for (ii) but using a stochastic or fire for its control is a distinct possibility. An probability method to allocate intervals in example of this class of circumstances occurs when each block. Pinus radiata invades perennial grassland or The last of these seems most appropriate for eucalypt open forest in the Australian Capital conservation management as it provides the most Territory (ACT) or elsewhere (Gill and Williams variation in time and space (see also Gill and 1996). In some cases fires may be used, perhaps in McCarthy 1998). Variations in such a prescription combination with other potential control agents would need to be made in some places because of (like selective herbicides). the need to accommodate other management aims Exotic species can be fuel species. Examples such as the protection of life and property. are invasive Pinus radiata (personal observation), Calibration of the probability model used (and its the shrub Lantana camara, and mission grass, form) would be essential and may be assisted by a Pennisetum polystachyon (see Gill et al. 1990). robust monitoring method (see Gill 1998). Exotic animals can affect fuel species by eating Fire management takes place at a landscape them, by transporting them or by creating or vegetation-community scale. At such scales conditions suitable for their spread. In the there are many species present, some of which Kimberley it appears that cattle are breaking may be ‘special’ in that they are rare or vulnerable down rainforest edges and either transporting to extinction. Because fire management takes seeds of grasses (fuel species) into them or place at a broad scale, any management aimed at modifying the environment such that fires can the preservation of ‘special’ species will necessarily enter more readily (McKenzie and Belbin 1991). take place at the same broad scale. The Rabbits may exert a marked effect on fuel loads in community will be managed for the species. semi-arid woodlands (Leigh et al. 1989) and no A major need for fire management today is doubt elsewhere. By affecting fuel characteristics the implementation of practical monitoring exotic species can alter fire regimes and affect the systems. Testing the suggestions already made ability of managers to maintain biodiversity. (Gill and Nicholls 1989, Gill 1998) would be a An examination of the time for the start. Having a suitable response system to the attainment of maximum species richness after fire results of monitoring is important also. Assessing may, incorrectly, suggest a suitable timing of fires where such systems are most appropriate and for the management of plant-species diversity. For implementing them is important if biodiversity is example, where maximum diversity is reached at to be managed in an accountable way.

44 Biodiveristy and bushfires

7. RESEARCH ON SHIFTS IN PLANT DIVERSITY AFTER FIRE

Research into the effects of fires on biodiversity (1980) and Gill (1981) - have been used to define has a wider compass than the focus of this review. functional groups; the attributes can often be However, the overlap is extensive. The problem is determined from observations made soon after how to assess the ecological effects of fires on the any fire. In a Register of Plant Species Responses sometimes overwhelming biodiversity of (Gill and Bradstock 1992) about 10% of the flora terrestrial Australia when this includes vascular has been listed but much of it only in the broadest and non-vascular plants, vertebrate and non- of functional groups. With the addition of time vertebrate animals, bacteria, algae and fungi, scales for key events for each species (fruiting, native and exotic. Fire effects vary with fire longevity), predictions can be made as to species’ regimes (see the Preamble) so the task is not presence or absence at particular times after a fire simply to examine the consequences of a single (Noble and Slatyer 1980); there are few data of fire on community composition over time - the this type. context of this review - but to assess all the effects Researchers simplify in order to cope with of all fire regimes on all components of the complexity. They will usually tackle one group of biodiversity across Australia, a formidable organisms such as vascular plants or vertebrate challenge. animals. They will tend to choose a single area, set Of around 18,000 vascular plant species in up plots and examine the effects of a limited range Australia (Hnatiuk 1990), a quarter (23%) are rare of fire intensities, frequencies and seasonalities for or threatened (Briggs and Leigh 1996). There are their experiments. Safety considerations usually also nearly 2,000 officially recognized naturalized limit the range of intensities used (Section 4.4). species (Hnatiuk 1990). Experimental examination Metronome-like frequencies are ideal of the responses of each of the rare and threatened experimentally but unlikely in nature. species, let alone the whole flora, to a range of fire Seasonalities adopted may be strongly influenced regimes - even if the appropriate range of regimes by restrictions imposed by fire authorities. Peat was known, and experimentally feasible to fires are ignored in experimental design because implement, and funding and land was available - the occurrence of peat fires implies prevailing seems unlikely. A variety of methods to observe drought, days, weeks or months for the effects - and not just those after a fire - needs to combustion process to be completed, and extreme be adopted. fire-suppression difficulties. Site choice may be Data for the determination of plant-species influenced by the likelihood of being able to ignite changes with time after fire is collected in a fires at the designated times. Complications such variety of ways but the variables of importance at as the interactions between burning and grazing any one stage are the number of species, the (Leigh and Holgate 1979), fires and frosts (Duff number of individuals of each species, the areas and Stocker 1989) and fires and cyclones (Stocker over which sampling occurs and the cover or and Mott 1981) are worth further examination. biomass of each species. There is a need for There are financial and practical limits to the efficient statistically-robust methods for the conduct of experiments but experiments provide collection of data on all of these variables at any controlled data. one site. Problems of methodology were State and Federal Forestry Departments addressed in Chapter 4. The more measures of have been the main instigators of fire-regime diversity that are used, and the greater the number experiments in Australia. Examples are those set of explanatory variables that are also measured, up by: the Forest Research Institute of the the greater the potential understanding. Australian Forestry and Timber Bureau (now part Interpretation of the data collected may be of CSIRO) in the Northern Territory and ACT made in terms of dominance, life cycles and (J. Hoare); the Qld Department of Forestry in functional groups. A knowledge of the attributes south-eastern Queensland (Byrne 1989); NSW of species within dispersal distance but not already State Forests (A. York and D. Binns); Victorian present may also assist. Species attributes - State government authorities (Tolhurst and Flynn defined along the lines of Noble and Slatyer 1992); the SA Woods and Forests Department

45 Australia’s Biodiveristy - Responses to Fire

(B. Gepp) and the Department of Forestry in From a management point of view, a Western Australia, now Department of research challenge is to define appropriate fire Conservation and Land Management (L. McCaw regimes for different ecosystems in different and N. D. Burrows). CSIRO Division of Wildlife social, economic, ecological, and geographic and Ecology set up the large Kapalga experiment contexts and to assess the need for, or the in the Northern Territory (Duff and Braithwaite consequences of, burned areas having particular 1989) while recently CSIRO Plant Industry sizes, shapes and arrangements. established two experiments, one in the ACT (by To assist research we can formulate J. Briggs, now with NSW National Parks and hypotheses, ask questions. For example: Wildlife Service), the other in NSW (by • Are the changes in vascular plant diversity S. Prober, now freelance). Experiments to date with time since fire simply an outcome of the have been mainly in forests and grassy woodlands, frequency of species in different functional the exception in the above examples being those groups found there as a result of particular experiments in Queensland where heath and site histories and environmental (non-fire) shrubby woodlands were chosen. There is selection? considerable scope for further experiment. • To what extent do the details within a (Experiments in which one fire only has been used functional group (such as time to flowering or in which artificial fuel has been used are not a and fruiting, longevity) affect the outcome of consideration here.) time-since-fire sequences? Because safety factors exclude the use of • Are patterns of species richness with time extreme intensities in experiments, careful after fire a function of local and national observations of the effects of unplanned severe environmental gradients? fires are an important source of information There is still a great deal to be learned about despite their occurrence being in unpredictable the changes in diversity that occur after fires locations at unpredictable times. For the same having various characteristics and occurring at reason, theoretical modelling is valuable. various times of year. Establishing a benchmark With a knowledge of the way in which fire methodology would help. regimes are distributed across the nation according to biophysical and social (including management) factors, it will be possible to predict the proportions of landscapes at particular stages after fire. Then we will be able to predict the species richness to be found in each piece of the landscape mosaic. How fire regimes are distributed in landscapes due to chance is just beginning to receive attention in Australia (Gill and McCarthy 1998). Models being developed in this context have the potential to greatly expand our understanding of the effects of fire regimes on biodiversity in Australia.

46 Biodiveristy and bushfires

8. CONCLUSIONS

Biodiversity, the variety of life, can be expressed ‘ephemerals’, ‘seeders’ and ‘sprouters’. Subdivision most simply as a list of species - ‘species richness’. of broad groups into woody or herbaceous Species richness varies with sample area (and ephemerals, ‘seeders’ with canopy-seed or soil- locations), the relationships being expressed using seed stores, and sprouters normally with or simple equations. Plant species richness can be without seedlings after fire, can be worthwhile. ‘apparent’ (in the form of plants above ground) or Identifying ‘tolerant’ and ‘requiring’ species is ‘hidden (only in the form of seeds, or other important especially where invasions occur. propagules, below ground). Apparent species Knowledge of the times to produce seeds and the richness is negatively affected by ecologically longevity of species is highly significant. dominant species, the dominance being illustrated To conserve native plant diversity, one needs by rank-abundance, rank-cover or rank-biomass to avoid extinction and prevent invasion of curves. Whether or not the first-ranked species exotics. Fire regimes affect species’ survival and are functionally dominant may need some are an essential tool of management. Both clarification. prescribed fires and unplanned fires can affect Fires occur in most of the plant communities local diversity. Management for the conservation of Australia. They can have major affects on the of biodiversity should aim at achieving suitable biota through fire regimes. Plant species richness proportions of landscape with a variety of times- changes, sometimes dramatically, with time after since-fire stages using fires within appropriate fire. In the absence of species’ identities being intensity levels at appropriate times of year and explicit, changes in richness with time are within an appropriate frequency range. A research depicted as numbers only. Where this occurs challenge is to define these ‘suitable proportions’ there can be no knowledge of species turnover. and ‘appropriate regimes’ for different ecosystems Understanding the dynamics of change with in different management, ecological, and time since fire requires a knowledge of the geographic contexts and to assess any need for any relevant functional groups of plants, such as particular spatially-explicit arrangement of burned areas.

47 Australia’s Biodiveristy - Responses to Fire

9. ACKNOWLEDGEMENTS

This report is the result of a contract with the Australian Department of the Environment and Heritage. Neal Hardy, from the Department, is to be thanked for his encouragement and facilitation of the project. Also, I would like to thank Julie McAlpine for her technical support. Penny Hohnen and Peter Moore compiled the map (Fig. 4.1) from data supplied by the author. Peter Moore provided logistic and moral support throughout the project. Ross Bradstock, Graham Griffin, Jim Noble, Kevin Tolhurst, Jann Williams and John Woinarski kindly read draft manuscripts and offered constructive suggestions for improvement.

48 Biodiveristy and bushfires

10. REFERENCES CITED

Allan, G. and Baker, L. (1990). Uluru (Ayers Bray, J. R. and Curtis, J. T. (1957). An ordination Rock-Mt. Olga) national park: an assessment of the upland forest communities of of a fire management programme. Proc. Ecol. southern Wisconsin. Ecol. Monog. 27, 325- Soc. Aust. 16, 215-220. 349. Allan, G. E. and Griffin, G. F. (1986). Fire Briggs, J. D. and Leigh, J. H. (1996). Rare or ecology of the hummock grasslands of Threatened Australian Plants. CSIRO, Central Australia. Proceedings of the 4th Melbourne. Australian Rangeland Society Conference. Brown, M. J. and Podger, F. D. (1982). Floristics Pp. 126-129. and fire regimes of a vegetation sequence Arrhenius, O. (1921). Species and area. J. Ecol. 9, from sedgeland-heath to rainforest at 95-99. Bathurst Harbour, Tasmania. Aust. J. Bot. Ashton, D. H. (1981). Fire in tall open-forests 30, 659-676. (wet sclerophyll forests). In: A. M. Gill, R. Burbidge, N. T. (1943). Ecological succession H. Groves and I. R. Noble (eds) Fire and the observed during regeneration of Triodia Australian Biota. Pp. 339-366. Australian pungens after burning. J. Roy. Soc. W.A. 28, Academy of Science, Canberra. 149-156. Barker, M. J. (1991). The Effect of Fire on West Burrows, N. D. and Christensen, P. E. S. (1990). Coast Lowland Rainforests. Tasmanian A survey of Aboriginal fire patterns in the National Rainforest Research Program western desert of Australia. USDA For. Serv. Report No.7. Gen. Tech. Rep. SE 69, 297-305. Bell, D. T. and Koch, J. M. (1980). Post-fire Bureau of Meteorology (1988). Climatic Averages. succession in the northern jarrah forest of Australia. Publishing Western Australia. Aust. J. Ecol. 5, 9-14. Service, Canberra. Bell, D.T., Hopkins, A. J. M. and Pate, J. S. Byrne, P. J. (1989). Forest fire research by the (1984). Fire in the Kwongan. In: J. S. Pate Department of Forestry in Queensland. In: and J. S. Beard. (eds) Kwongan. Plant Life of B. R. Roberts (ed) Fire Research in Rural the Sandplain. Pp.178-204. University of Queensland. Pp. 229-236. University of Western Australia Press, Nedlands, Perth. Southern Queensland, Toowoomba. Bowman, D. M. J. S. (1991). Recovery of some Cheal, D. (1996). Fire succession in heathlands northern Australian monsoon forest tree and implications for vegetation management. species following fire. Proc. Roy. Soc. Qld. Australian Department of Environment, Sport 101, 21-25. and Territories Biodiversity Series, Paper No. 8, Bowman, D. M. J. S. (1992). Monsoon forests in 67-79. north-western Australia. II. Forest-savanna Cheal, P. D., Day, J. C. and Meredith, C. W. transitions. Aust. J. Bot. 40, 89-102. (1979). Fire in the National Parks of North Bowman, D. M. J. S. and Fensham, R. J. (1991). West Victoria. National Parks Service, Response of a monsoon forest-savanna Melbourne. boundary to fire protection, Weipa, northern Christensen, P. E. and Kimber, P. C. (1975). Australia. Aust. J. Ecol. 16, 111-118. Effects of prescribed burning on the flora Bowman, D. M. J. S. and Jackson, W. D. (1981). and fauna of south-west Australian forests. Vegetation succession in southwest Proc. Ecol. Soc. Aust. 9, 85-106. Tasmania. Search 12, 358-362. Clemens, J. and Franklin, M. H. (1980). A Bowman, D. M. J. S., Wilson, B. A. and Hooper, description of coastal heath at North Head, R. J. (1988). Response of Eucalyptus forest Sydney Harbour National Park: Impact of and woodland to four fire regimes at recreation and other disturbance since 1951. Munmarlary, Northern Territory, Australia. Aust. J. Bot. 28, 463-478. J. Ecol. 76, 215-232. Cockburn, A. (1978). The distribution of Pseudomys shortridgei (Muridae: Rodentia) and its relevance to that of other heathland Pseudomys. Aust. Wildl. Res. 5, 213-219.

49 Australia’s Biodiveristy - Responses to Fire

Cochrane, G. R., Burnard, S., and Philpott, J. M. Gill, A. M. (1977). Management of fire-prone (1962). Land use and forest fires in the vegetation for plant species conservation in South Australia. Aust. Australia. Search (Syd.) 8, 20-26. Geographer 8(4), 143-160. Gill, A. M. (1981). Coping with fire. In: J. S. Pate Connor, E. F. and McCoy, E. D. (1979). The and A. J. McComb (eds) The Biology of statistics and biology of the species-area Australian Plants. Pp. 65-87. University of relationship. Amer. Nat. 113, 791-833. Western Australia Press, Nedlands. Curtis, J. T. and McIntosh, R. P. (1950). The Gill, A. M. (1993). Interplay of Victoria’s flora interrelationships of certain analytic and with fire. In: D. B. Foreman and N. G. synthetic phytosociological characters. Walsh (eds) Flora of Victoria. Volume 1. Ecology 31, 434-455. Introduction. Pp. 212-226. Inkata Press, Doak, D. F. and Mills, L. S. (1994). A useful role Melbourne. for theory in conservation. Ecology 73, 615- Gill, A. M. (1995). Stems and fires. In: B. L. 626. Gartner (ed) Plant Stems. Physiological and Duff, G. A. and Braithwaite, R. W. (1989). Fire Functional Morphology. Pp. 323-342. and Top End Forests - Past, Present and Academic Press, San Diego. Future Research. In: B.R.Roberts (ed) Fire Gill, A. M. (1998). Monitoring for biodiversity in Research in Rural Queensland. Pp. 84-98. fire-prone areas. In: R. Melzer (ed) Fire University of Southern Queensland, Management: Presented Papers. Queensland Toowoomba. Department of the Environment, Duff, G. A. and Stocker, G. C. (1989). The effects Rockhampton. of frosts on rainforest/open forest ecotones Gill, A. M. and Bradstock, R. A. (1992). A national in the highlands of North Queensland. Proc. register for the fire responses of plant Roy. Soc. Qld. 100, 49-54. species. Cunninghamia 2, 653-660. Egler, F. E. (1954). Vegetation science concepts. 1. Gill, A. M. and Bradstock, R. A. (1994). The Initial floristic composition, a factor in old- prescribed burning debate in temperate field vegetation development. Vegetatio 4, Australian forests: towards a resolution. Pp. 412-417. 703-712. Proceedings of the 2nd Ellis, R. C. (1985). The relationships among International Conference on Forest Fire eucalypt forest, grassland and rainforest in a Research, Coimbra, Portugal. highland area in north-east Tasmania. Aust. Gill, A. M. and Bradstock, R. A. (1995). J. Ecol. 10, 297-314. Extinctions of biota by fires. In: R. A. Fox, J. E.D., Pepper, C. and Craig, G. (1983). Bradstock, T. D. Auld, D. A. Keith, R. T. Vegetation and reaction to fire in coastal Kingsford, D. Lunney and D. P. Sivertsen heath at Sorrento, Western Australia. Mulga (eds) Conserving Biodiversity: Threats and Research Institute Annual Report No. 6. Pp. Solutions. Pp. 309-322. NSW National Parks 97-109. and Wildlife Service, Sydney. Fox, M. D. (1988). Understorey changes Gill, A. M. and McCarthy, M. C. (1998). Intervals following fire at Myall Lakes, New South between prescribed fires in Australia: what Wales. Cunninghamia 2(1), 85-95. intrinsic variation should apply? Biol.Conserv. Fox, M. D. (1990). Composition and richness of 85, 161-169. New South Wales mallee. In: J. C. Noble, P. Gill, A. M. and Moore, P. H. R. (1990). Fire J. Joss and G. K. Jones (eds) The Mallee intensities in Eucalyptus forests of Lands. A Conservation Perspective. Pp. 8-11. southeastern Australia. Proceedings 1st CSIRO, Melbourne. International Forest Fire Research Fox, M. D. and Fox, B. J. (1986). The effect of fire Conference, Coimbra, Portugal. Pp. B.24-1 frequency on the structure and floristic to 12. composition of a woodland understorey. Gill, A. M. and Nicholls, A. O. (1989). Aust. J. Ecol. 11, 77-85. Monitoring fire-prone flora in reserves for Gilbert, J. M. (1959). Forest succession in the nature conservation. In: N. Burrows, L. Florentine Valley, Tasmania. Pap. Proc. Roy. McCaw, and G. Friend (eds) Fire Soc. Tas. 93, 129-151. Management on Nature Conservation Lands. Gill, A. M. (1975). Fire and the Australian flora: a W.A. Dept. C.A.L.M. Occ. Pap. 1/89, pp review. Aust. For. 38, 4-25. 137-151.

50 Biodiveristy and bushfires

Gill, A. M. and Williams, J. E. (1996). Fire Hesp, P. A., Wells, M. R., Ward, B. H. R. and regimes and biodiversity: the effects of Riches, J. R. H. (1983). Land Resource Survey fragmentation of southeastern eucalypt of Rottnest Island. Western Australian forests by urbanisation, agriculture and pine Department of Agriculture, Perth. plantations. For. Ecol. & Managt 85, 261-278. Hnatiuk, R. J. (1990). Census of Australian Vascular Gill, A. M., Hoare, J. R. L and Cheney, N. P. Plants. Australian Flora and Fauna Series (1990). Fires and their effects in the wet-dry Number 11. Australian Government tropics of Australia. Ecol. Stud. 84, 159-178. Publishing Service, Canberra. Gill, A. M., Groves, R. H. and Noble, I. R. (eds) Hoare, J. R. L., Hooper, R. J., Cheney, N. P. and (1981). Fire and the Australian Biota. Jacobsen, K. L. S. (1980). A Report on the Australian Academy of Science, Canberra. Effects of Fire in Tall Open Forest and Gitay, H., Roxburgh, S. H. and Wilson, J. B. Woodland with particular Reference to Fire (1991). Species-area relations in New Management in Kakadu National Park in the Zealand tussock grassland, with implications Northern Territory. Unpublished Report to for nature reserve design and for community the Australian National Parks and Wildlife structure. J. Veg. Sci. 2, 113-118. Service, Canberra. Gleadow, R. M. and Ashton, D. H. (1981). Hobbs, R. J. and Atkins, L. (1990). Fire related Invasion by Pittosporum undulatum of the dynamics of a Banksia woodland in south- forests of Central Victoria. I. Invasion western Western Australia. Aust. J. Bot., 38, patterns and plant morphology. Aust. J. Bot. 97-110. 29, 705-720. Howard, T. M. (1974). Nothofagus cunninghamii Gleason, H. A. (1922). On the relation between ecotonal stages. Buried viable seed in North species and area. Ecology 3, 158-162. West Tasmania. Proc. Roy. Soc. Vic. 86, 137- Gray, J.S. (1986). Species-abundance patterns. In: 142. J. H. R. Gee and P. S. Giller (eds) Hubbell, S. P. (1979). Tree dispersion, abundance Organization of Communities: Past and Present. and diversity in a tropical dry forest. Science Symposium of the British Ecological Society. 203, 1299-1309. Pp. 53-67. Blackwell Scientific, Oxford. Hughes, R. G. (1986). Theories and models of Griffin, G. F. (1984). Vegetation patterns. In: E. species abundance. The American Naturalist C. Saxon (ed) Anticipating the Inevitable: A 128, 879-899. Patch-burn Strategy for Fire Management at Ingwersen, F. (1977). Regeneration of Vegetation Uluru (Ayers Rock - Mt Olga) National Park. after Fire at Jervis Bay and its Implications Pp. 25-37. CSIRO, Melbourne. for Management. Unpublished M.Sc. thesis, Griffin, G. F. (1992). Will it burn - should it Australian National University. 415 p. burn?: management of the spinifex Jackson, W. D. (1968). Fire, air, water, and earth - grasslands of inland Australia. In: G.P. an elemental ecology of Tasmania. Proc. Ecol. Chapman (ed) Desertified Grasslands: Their Soc. Aust. 3, 9-16. Biology and Management. Pp. 63-76. Kikkawa, J., Webb, L. J., Dale, M. B., Monteith, Academic Press, London. G. B., Tracey, J. G. and Williams, W. T. Griffin, G. F. and Hodgkinson, K. C. (1986). The (1981). Gradients and boundaries of use of fire for the management of the mulga monsoon forests in Australia. Proc. Ecol. Soc. land vegetation in Australia. In: P. Sattler Aust. 11, 39-52. (ed) The . Pp. 93-97. Royal Knight, D. H. (1975). A phytosociological analysis Society of Queensland, Brisbane. of species-rich tropical forest on Barro Harrington, G. N. and Sanderson, K. D. (1994). Colorado Island, Panama. Ecol. Monog. 45, Recent contraction of wet sclerophyll forest 259-284. in the wet tropics of Queensland due to Kobayashi, S. and Kimura, K. (1994). The invasion by rainforest. Pacific Conservation number of species occurring in a sample of a Biology 1, 319-327. biotic community and its connections with Harrold, A. (1979). Heathland regeneration after species-abundance relationship and spatial fire at Noosa. Qld. Natur. 22, 88-96. distribution. Ecol. Res. 9, 281-294.

51 Australia’s Biodiveristy - Responses to Fire

Lamont, B. B., Downes, S. and Fox, J. E. D. Melick, D. R. (1990). Ecology of rainforest and (1977). Importance-value curves and sclerophyllous communities in the Mitchell diversity indices applied to a species-rich River National Park, Gippsland, Victoria. heathland in Western Australia. Nature 265, Proc. Roy. Soc. Vic. 102, 71-87. 438-441. Melick, D. R. and Ashton, D. H. (1991). The Leigh, J. H. and Holgate, M. D. (1979). The effects of natural disturbances on warm responses of the understorey of forests and temperate rainforests in south-eastern woodlands of the to Australia. Aust. J. Bot. 39, 1-30. grazing and burning. Aust. J. Ecol. 4, 25-45. Menhinick, E. F. (1964). A comparison of some Leigh, J. H., Wimbush, D. J., Wood, D. H., species-individuals diversity indexes applied Holgate, M. D., Slee, A. V., Stanger, M. G. to samples of field insects. Ecol. 45, 859-861. and Forrester, R. I. (1987). Effects of rabbit Molnar, C. D., Fletcher, D. and Parsons, R. F. grazing and fire on a subalpine environment. (1989). Relationships between heath and I. Herbaceous and shrubby vegetation. Aust. Leptospermum laevigatum scrub at J. Bot. 35, 433-464. Sandringham, Victoria. Proc. Roy. Soc. Vic. Leigh, J. H., Wood, D. H., Holgate, M. D., Slee, 101, 77-87. A. and Stanger, M. G. (1989). Effects of Mount, A. B. (1979). Natural regeneration rabbit and grazing on two semi- processes in Tasmanian forests. Search 10, arid grassland communities in Central- 180-186. western New South Wales. Aust. J. Bot. 37, Neyland, M. G. (1991). Relict rainforest in 375-396. eastern Tasmania. Tasmanian National Loehle, C. (1990). Proper statistical treatment of Rainforest Conservation Program Technical species-area data. Oikos 57, 143-145. Report No. 6. MacArthur, R. H. and Wilson, E. O. (1963). An Noble, I. R. and Slatyer, R. O. (1980). The use of equilibrium theory of insular zoogeography. vital attributes to predict successional Evolution 17, 373-387. changes in plant communities subject to MacArthur, R. H. and Wilson, E. O. (1967). The recurrent disturbance. Vegetatio 43, 5-21. Theory of Island Biogeography. Princeton Noble, I. R. and Slatyer, R. O. (1981). Concepts University Press, Princeton, New Jersey. and models of succession. In: A. M. Gill, R. McFarland, D. C. (1988). Fire and vegetation H. Groves and I. R. Noble (eds) Fire and the composition and structure of subtropical Australian Biota. Pp. 311-335. Australian heathlands in south-east Queensland. Aust. Academy of Science, Canberra. J. Bot. 36, 533-546. Noble, I. R., Moore, A. D. and Strasser, M. J. McKenzie, N. L. and Belbin, L. (1991). (1988). Stand structure and the prediction of Kimberley rainforest communities: reserve vegetation dynamics. In: H. J. During, M. J. recommendations and management A. Werger and J. H. Willems (eds) Diversity considerations. In: N. L. McKenzie, R. B. and Pattern in Plant Communities. Pp. 263- Johnson and P. G. Kendrick (eds) Kimberley 275. SPB Academic Publishing, The Hague, Rainforests. Pp. 453-480. Surrey Beatty and The Netherlands. Sons, Chipping Norton, New South Wales. Noble, J. C. (1989). Fire studies in mallee McMahon, A. (1984). The effects of time since (Eucalyptus spp.) communities of western fire on heathlands in the Little Desert, N.W. New South Wales: The effects of fires Victoria, Australia. In: B. Dell. (ed) applied in different seasons on herbage MEDECOS IV, Proceedings 4th International productivity and their implications for Conference on Mediterranean Ecosystems. management. Aust. J. Ecol. 14, 169-187. Pp.99-100. Palmer, M. W. and White, P. S. (1994). Scale McMahon, A. R. G. (1987). The Effects of the dependence and the species-area 1982-83 Bushfires on Sites of Significance. relationship. Amer. Nat. 144, 717-740. Victorian Department of Forests & Lands, Patterson, B. D. and Atmar, W. (1986). Nested Environmental Studies Publication Series subsets and the structure of insular No. 411. mammalian faunas and archipelagos. Biol. J. McNaughton, S. J. (1993). Biodiversity and Linn. Soc. 28, 65-82. function of grazing ecosystems. Ecol. Stud. Pimm, S. L. (1982). Food Webs. Chapman and 99, 361-383. Hall, London.

52 Biodiveristy and bushfires

Preston, F. W. (1960). Time and space and the Tolhurst, K. G. and Oswin, D. A. (1992). Effects variation of species. Ecology 41, 611-627. of spring and autumn low intensity fire on Preston, F. W. (1962a,b). The canonical understorey vegetation in open eucalypt distribution of commonness and rarity. forest in west-central Victoria. Victorian Ecology 43, 185-215, 410-432. Department of Conservation and Purdie, R. W. (1977). Early stages of regeneration Environment, Research Report 349. Pp. 3.1 after burning in dry sclerophyll vegetation. I. to 3.60. Regeneration of the understorey by Unwin, G. L. (1989). Structure and composition vegetative means. Aust. J. Bot. 25, 21-34. of the abrupt rainforest boundary in the Ridley, W. F. and Gardner, A. (1961). Fires in Herberton Highland, North Queensland. rainforest. Aust. J. Sci. 23, 227-228. Aust. J. Bot. 37, 413-428. Russell, R. P. and Parsons, R. F. (1978). Effects of Van der Moezel, P. G. and Bell, D. T. (1984). Fire time since fire on heath floristics at Wilson’s in the Western Australian mallee. In: B. Dell Promontory, southern Australia. Aust. J. Bot. (ed) MEDECOS IV, Proceedings 4th 26, 53-61. International Conference on Mediterranean Russell-Smith, J. and Bowman, D. M. J. S. (1992). Ecosystems. Pp. 151-152. Conservation of monsoon rainforest isolates Venning, J. (1978). Post-fire responses of a in the Northern Territory, Australia. Biol. Eucalyptus baxteri woodland near Penola in Conserv. 59, 51-63. South Australia. Aust. For., 41(4), 192-206. Rydin, H. and Borgegard, S. O. (1988). Plant Wark, M. C. (1997). Regeneration of some forest species richness on islands over a century of and gully communities in the Angahook- plant succession: Lake Hjalmaren. Ecol. 69, Lorne State Park (north-eastern Otway 916-927. Ranges) 1-10 years after the wildfire of Specht, R. L. and Specht, A. (1989). Species February 1983. Proc. Roy. Soc. Vict. 109, 7- richness of sclerophyll (heathy) plant 36. communities in Australia - the influence of Wark, M. C., White, M. D., Robertson, D. J. and overstorey cover. Aust. J. Bot. 37, 337-350. Marriott, P. F. (1987). Regeneration of heath Specht, R. L., Rayson, P., and Jackman, M. E. and heath woodland in the north-eastern (1958). Dark Island heath (Ninety Mile Otway Ranges following the wildfire of , South Australia). VI. Pyric succession: February 1983. Proc. Roy. Soc. Vict. 99, 51-88. changes in composition, coverage, dry Whittaker, R. H. (1965). Dominance and diversity weight and mineral nutrient status. Aust. J. in land plant communities. Science 147, 250- Bot. 6 , 59-88. 260. Stocker, G. C. and Mott, J. J. (1981). Fire in the Whittaker, R. H. (1972). Evolution and tropical forests and woodlands of northern measurement of species diversity. Taxon 21, Australia. In: A. M. Gill, R. H.Groves and I. 213-251. R. Noble (eds) Fire and the Australian Biota. Wilson, J. B., Gitay, H., Roxburgh, S. H., King, Pp. 425-439. Australian Academy of Science, W. M. and Tangney, R. S. (1992). Egler’s Canberra. concept of ‘initial floristic composition’ in Stocker, G. C. and Sturtz, J. D. (1966). Use of fire succession - ecologists citing it don’t agree to establish Townsville lucerne in the what it means. Oikos 64, 591-593. Northern Territory. Aust. J. Expt. Agric. & Wood, J. G. (1937). The Vegetation of South Anim. Husb. 6, 277-279. Australia. Government Printer, Adelaide. Taylor, J.A. and Dunlop, C.R. (1985). Plant Zimmer, W. J. (1940). Plant invasions in the communities of the wet-dry tropics of mallee. Vic. Natur. 56, 143-147. Australia: the Alligator Rivers region, Northern Territory. Proc. Ecol. Soc. Aust. 13, 83-127. Tolhurst, K. and Flinn, D. (1992). (eds) Ecological Impacts of Fuel Reduction Burning in Dry Sclerophyll Forest: First Progress Report. Victorian Department of Conservation and Environment, Research Report 349.

53

Biodiveristy and bushfires

FIRE AND AUSTRALIAN BIRDS: A REVIEW

J.C.Z. Woinarski

Parks and Wildlife Commission of the Northern Territory PO Box 496 Palmerston, NT, 0831 Australia’s Biodiveristy - Responses to Fire

56 Fire and Australian Brids

SUMMARY

The literature concerning the impacts of fire on Of the threatened species whose relationships Australian birds is reviewed. Fire effects are with fire regime has been comparatively well considered for main Australian environments. documented, almost all show clear preference for The methodology of studies is reviewed. much less frequent fire than that currently Detrimental fire regimes contributed to the prevailing. The long-unburnt vegetation favoured extinction of two of the three bird species, and by these species is becoming disappearingly rare, three of the four which have and will require concerted management effort to disappeared from Australia since European maintain or increase. Most fire-sensitive colonisation. Inappropriate fire management is threatened birds have low reproductive output and now a factor in the threatened status of at least 51 limited dispersal ability. The persistence of these nationally recognised threatened bird taxa (second species is further jeopardised by habitat in importance only to habitat clearance and fragmentation, which accentuates the handicap of fragmentation, which threatens 52 taxa). In many these traits for recolonisation following fire. environments (notably heath and mallee), In temperate eucalypt forests, control inappropriate fire regime is now the main threat burning is widely used to reduce the probability of to most declining bird species. extensive wildfire. While the impact of a single Despite this recognition of the seriousness of control burn upon birds is generally less than that fire regime as a process threatening Australian of wildfire, there are few data on the long-term birds, for only a few species has detailed research impacts of a sustained regime of control burning. on the relationship between birds and fire regimes The most detailed long-term study suggests that been conducted, and only in exceptional cases has such frequent mild fires will lead to the decline this been translated into management. For many and loss of some species which are now perceived environments, there is no information on the as common and little affected by mild fires. impacts of fire regimes on birds.

57 Australia’s Biodiveristy - Responses to Fire

58 Fire and Australian Brids

INTRODUCTION

Fire has long been recognised as one of the major environmental impacts of fires, but which doesn’t factors moulding vegetation patterning across explicitly mention birds; and Australia (e.g. Mitchell 1838; Jackson 1968). (ii) review, as comprehensively as possible, Manipulation of fire, either explicitly or by the literature explicitly concerned with fire and default, is the main landscape management tool in Australian birds. many environments and most conservation An annotated bibliography of the relevant reserves. Inappropriate fire regime is the main literature is compiled as a foundation for the latter threat to many vulnerable and endangered goal. Australian birds (Garnett 1992a) and other biota Interpreting the literature is rendered (e.g. plants: Leigh et al. 1984). Yet knowledge of difficult by the variation in fires within and the history and effects of fire is negligible in many between environments. Individual fires within a environments, and there have been remarkably given environment vary in seasonal timing, extent, few long-term studies of the ecological impacts of intensity, patchiness, and the temporal pattern of a range of fire regimes, such that the long-term their (re-) occurrence. Their impacts at a consequences of fire management are usually very particular site are influenced by these diverse poorly known. factors (and also by landscape context and pre- In this essay, I seek to review the and post-fire climatic conditions), and different relationships between Australian birds and fire. bird species may respond very idiosyncratically to Previous reviews (e.g. Cowley et al. 1969; this variation. The search for general pattern in Christensen & Kimber 1975; Catling & Newsome response is further hampered by the very variable 1981; Recher 1981; Recher & Christensen 1981; and limited research effort, with few long-term Meredith 1983,1988; Suckling & MacFarlane studies and little experimentation with a range of 1983; Christensen et al. 1985; Smith 1987a; fire treatments. In this regard, it is notable that Christensen & Abbott 1989; McFarland 1993; successional responses of Australian mammals Silveira 1993; Arnold et al. 1993; Crowley 1994) have been much more thoroughly documented have summarised and interpreted studies of fire and subject to far more intensive experimental impacts on birds for broad regions or investigation (e.g. Suckling & MacFarlane 1983; environments within Australia, and this national Higgs & Fox 1993). overview owes much to these predecessors. The This document contains a general overview quantity of these reviews is surprisingly out of of fire history, an assessment of fire impacts on proportion to the very limited number of detailed birds across all main Australian environments, a primary studies, and the attempt here at a national review of the methodological approaches used in overview is the only justification for adding yet the investigation of impacts, and a concluding another review to this unbalanced literature. section reviewing fire management and bird Many of the impacts of fire on Australian conservation. environments indirectly affect birds, for example by change in nutrient availability, food resources or floristics. References on these subjects are relevant to birds even though birds may not be explicitly mentioned. However, a complete review of the environmental impacts of fire is beyond the scope of this essay, and would be redundant anyway given the excellent recent comprehensive reviews by Gill et al. (1981), Pyne (1991), Williams & Gill (1995), and Whelan (1995), and two recent broad collections of papers (McCaw et al. 1995; DEST 1996). Rather, I attempt here to: (i) interpret, in terms of consequences for birds, the literature which describes

59 Australia’s Biodiveristy - Responses to Fire

60 Fire and Australian Brids

HISTORY OF FIRE REGIMES

Evidence from pollen cores and plant fossil suggests that Aboriginal people generally burnt material has demonstrated substantial re- their lands frequently and purposefully. Most arrangement of Australian vegetation since at least resultant fires appear to have been relatively cool the Tertiary (Hill 1994), notably including a and of limited extent. Hence a fine scale mosaic of general increase in the extent of Eucalyptus open vegetation age was maintained, although frequent forests and woodlands at the expense of closed burning may have favoured early successional forests. Associated with these broad-scale species, and savannas and eucalypts generally. environmental changes, there has been a substantial Heaths may also have expanded under Aboriginal extinction of Australian fauna, notably rainforest fire regimes (Jackson 1968; Smith 1977), although and megafaunal elements, and including many shrubbiness under forests and woodlands was orders and families of birds (Rich 1991). probably much diminished (Gill 1981; Recher Environmental change accelerated during the et al. 1993; Stanton 1995). rapid climatic fluctuations of the late Pleistocene Late successional species (and fire-sensitive (Singh & Geissler 1985; Kershaw 1986), and this communities) may have been retained within the was accompanied by extinctions and range changes managed landscape by the deliberate exclusion of for much of the fauna. This period saw the end of fire from fire-sensitive areas which supported the line for the massive flightless Mihirungs particular resources (e.g. Jones 1995; Lucas & (Dromornithidae), the disappearance of flamingoes Lucas 1993) or which were otherwise of cultural (Phoenicopteridae) from Australia, the extinction of significance (Latz 1995). Areas (or environments) large species of coucal and megapodes, the which were unsuitable for habitation or supported mainland loss of the flightless Tasmanian Native relatively low densities of Aboriginal residents Hen1 and further range contractions for groups (perhaps such as much of the mallee: Harris 1990) associated with closed forests (e.g. the logrunners may have avoided the imposition of Aboriginal Orthonychidae) (Baird 1991). This change may fire regimes. have been compounded (or precipitated: Merilees Nonetheless, the flammability of much of 1968; Smith 1977; Flannery 1990, 1994) by the Australian vegetation (and recurrent drought) entry of Aboriginal people to Australia, between must have ensured that there were episodes of about 60–100,000 ybp, as Aboriginal use of fire catastrophic and extensive wildfires, even under almost certainly led to a fire regime different from Aboriginal stewardship. The existence of cohorts that previously prevalent (Jones 1969; Singh et al. of Mountain Ash Eucalyptus regnans pre-dating 1981; Nicholson 1981; Braithwaite & Estbergs European colonisation (Gill 1981) is one 1985), and consequently re-shaped the Australian demonstration of the extensive, if infrequent, landscape. The extent to which Aboriginal land occurrence of such wildfire. management was responsible for environmental The impacts upon birds of Aboriginal modification remains unclear and contested (e.g. burning regimes were probably very disparate. Williams & Gill 1995). Directly, Aboriginal people probably used fire as Any sustained change in burning regimes an aid in hunting flightless birds, as they did for will benefit some components of the ecosystem hunting macropods and other mammals but disadvantage others. Evidence from early (Braithwaite 1991). It is possible that the European accounts of firing by Aborigines (e.g. extinction or range contraction of several species Hallam 1975; Braithwaite 1991; Jones 1995), of flightless birds in the late Pleistocene was due current accounts of reasonably traditional to Aboriginal hunting (aided by fire), though Aboriginal land management (Haynes 1985, 1991; many other factors may have been involved (e.g. Lucas & Lucas 1993; Baker et al. 1993; Bradley Baird 1984). Aboriginal hunters also used smoke 1995) and interpretation of vegetation patterning to attract and trap raptors, and the remnants of (e.g. Jackson 1968; Price & Bowman 1994) such traps are still visible in parts of northern Australia (Boekel 1980). An increase in the frequency and 1 Scientific names of all birds mentioned in text or tables predictability of fires, following Aboriginal entry are given in Table 1.

61 Australia’s Biodiveristy - Responses to Fire

to Australia, probably would have led to increased forests and savanna woodlands (e.g. many abundance of birds which forage around fires (e.g. granivorous finches, doves and parrots; raptors, wood-: Braithwaite & Estbergs honeyeaters). Nonetheless, the relatively fine scale 1987; Woinarski 1990) or in recently-burnt areas of burning patterns for lands under Aboriginal (e.g. granivores). Such fire-associated species management probably maintained relatively high probably could not maintain large populations levels of heterogeneity of seral stages and under natural fire regimes of infrequent and environments generally, and hence supported high unpredictable fires. diversity of bird species. However, for birds, the most substantial Rapidly following European usurpation of impacts of Aboriginal fire regimes were probably land management across much of Australia, the experienced through resultant vegetation change. practice and purpose of burning changed. An increase in fire frequency would have led to Destructive and extensive burns were used to aid change in floristics, vegetation structure and grain clearing or to demonstrate proprietorship (Gill size of vegetation mosaics. Fire-sensitive 1981; Johnson & Purdie 1981). While some vegetation (and its associated bird fauna) would changes of incidence (and impacts) were have declined, except where this provided deliberate, others involved a complex interplay of desirable resources and could be protected. new factors. For example, the introduction and Probable losers from this re-casting of Australian rapid spread of livestock and exotic plants changed environments included species associated with fuel characteristics and understorey floristics, forests, with Casuarina or Allocasuarina thereby constraining options for fire regimes. woodlands (e.g. Glossy Black-cockatoo: Clout Other introductions, notably rabbits, interfered 1989), with closed forests, or with dense shrubby with post-fire plant recruitment processes. understoreys below Eucalyptus forests (e.g. scrub- Clearing and fragmentation of native vegetation birds, ). Probable winners included changed the scale and extent of fires. More species associated with the relatively sparse recently, fire has been used as a tool for vegetation of early seral stages (e.g. Richard’s biodiversity conservation (Good 1981; Garnett & Pipit), and those associated with eucalypt open Crowley 1994; Crowley 1995).

62 Fire and Australian Brids

FIRE REGIMES AND THEIR IMPACTS ON BIRDS: A REVIEW ACROSS MAIN ENVIRONMENTS

For all main (loosely-defined) Australian <5 years) control burning. Such frequent burning environments, I consider (where possible) current has led to the decline and/or local extinction of and pre-European fire regimes, bird species several heathland-dependent birds, most notably responses to single fires and fire regimes, Ground Parrot and Southern Emu-. Over relationships of threatened taxa to fire regimes and the last two decades, largely in response to fire management for bird conservation. The research on the requirements of such species, impacts of fire and potential regimes differ control burning in heathlands has been reduced. appreciably between these environments. In some In some areas (notably on conservation reserves), environments (notably temperate eucalypt open fire exclusion or suppression has been forests), there may be marked differences between recommended (e.g. Cooper 1974) and practised mild control fires (which usually affect only the (Meredith et al. 1984). Even in such protected understorey) and wildfire. This distinction is less heaths, occasional wildfires have proven difficult clearcut in most heaths and mallee communities, to exclude, and have burnt some reserved heaths where most vegetation occurs near the ground, and extensively (notably including the 1994 fire the response of birds to control burns may be through heathlands at Royal National Park). similar to those following wildfire (Meredith 1983). Species responses to single fires COASTAL HEATHS AND THICKETS Relative to most other Australian environments, the impacts of fire on birds in heaths has been well Fire regimes documented (Table 2). Very high mortality rates have been reported for heathland birds during In southwestern Australia, where previously wildfires (Recher et al. 1975; Fox 1978; Pescott Aboriginal people had burnt coastal heath country 1983; Wegener 1984a,b), though Main (1981) with small cool fires at intervals of 5-10 years, early suggested that many birds were able to escape colonists torched the heaths at least once every such fire. Hawking insectivores (e.g. swallows) and 2–3 years until grasslands suitable for pastoralism some raptors may be attracted to fire fronts (Main replaced the native bush (Hallam 1975; Smith 1981). 1977,1987a). Carter (1924) provided a graphic In days to months following fire in heaths, eyewitness account of these changes: the dead or exposed invertebrate and vertebrate “where there had been dense impenetrable prey attracts many predator visitors from adjacent scrub, was mostly bare sand drifts caused by habitats (e.g. Straw-necked Ibis, Torresian Crow, fire made to improve the country for cattle Laughing Kookaburra, raptors, egrets, Pied grazing”. , , Australian Raven, By contrast, in southwestern Tasmania, Magpie-lark), and seed shed by plants in response Jones (1995) noted that dense thickets and heaths to fire (e.g. Banksia spp.) attracts parrots and occurred now in areas where observers in 1829 had cockatoos (Roberts 1970; McFarland 1988,1993). recorded open sedgelands, a change attributed to While regrowth vegetation remains relatively the removal of frequent fires lit by Aboriginal open and herb and grass species are relatively people. This disparity in change reflects both the abundant (up to about 3 years post-fire), the area is difficulty of generalising about fire regimes and colonised by open-country species such as their effects, and the limited data on, and frailty of Richard’s Pipit, Brown , Little Button-quail, interpretations of, pre-European fire regimes. Elegant Parrot, swallows and martins (Smith In coastal heaths of southeastern and eastern 1987a; McFarland 1988; Brooker & Rowley 1991; Australia, European fire management was initially Hopkins & Smith 1996) (Fig. 1). In some heaths, indiscriminate, unless where directed to fire may trigger unusually abundant flowering for conversion of heaths to farmlands. More recently some plants (notably Xanthorrhoea), and some (notably since the 1950s: Luke & McArthur honeyeater and parrot species may respond with 1978), most heathlands have been managed by increased abundance (Specht 1981; McFarland high frequency (often with prescribed intervals of 1993). Insect abundance may increase rapidly after

63 Australia’s Biodiveristy - Responses to Fire

may vary substantially according to heath floristics (notably affected by moisture availability), the previous occurrence of fire, climate, heath patch size and isolation, and the patchiness of the fire (Recher et al. 1975; Meredith et al. 1984; Jordan 1987c; McFarland 1994). For example, one year following a fire that left some small unburnt patches of heath, Recher et al. (1975) reported that all but one species present before the fire were still present (although some were less common). In contrast, 2.5 years after a less patchy fire, Roberts (1970) reported that eight pre-fire resident species (Eastern Whipbird, White-cheeked Honeyeater, Little Wattlebird, Variegated Fairy-wren, Chestnut-rumped Heathwren, and Red-browed Finch) either disappeared or had declined substantially. Meredith et al. (1984) noted that Ground Parrots had not recolonised a patch of heath of suitable age, probably because it had previously been burnt at very frequent intervals. Where fire-free intervals are insufficent to allow Figure 1: Response of some heathland bird species to for maturation of plant species recruiting only fire (after McFarland 1998) through seed, the density of nectarivores may remain low, or particular nectarivore species, such fire, sometimes leading to increased breeding as Crescent Honeyeater (Recher et al. 1975; success or rapid increase in some insectivorous Recher 1981; Christensen et al. 1981) may not birds (Jordan 1987c). The total number of bird return. McFarland (1994) noted that fragmentation species may peak at one year post-fire, followed by of heaths may prevent recruitment of species with gradual decline (McFarland 1993). Resident relatively poor dispersal ability (such as Southern species may change diet, foraging behaviour or Emu-wren) to otherwise suitable heaths. nesting behaviour with vegetation and resource changes following fire (Brooker & Rowley 1991). Species responses to fire regimes As the heath structure becomes more closed In contrast to the relatively large number of (after about 3 years post-fire), the bird species studies which have considered effects of single composition changes, with loss of the opportunist fires upon birds (or relationship of species with open-country species and their replacement by time since fire), there have been remarkably few species reliant on denser vegetation such as King studies which have considered impacts of repeated Quail, Brush Bronzewing, Ground Parrot, Tawny fire or sustained fire regimes. The most detailed Grassbird, Golden-headed Cisticola, Red-backed study of repeated fire on birds is that of Rowley & Fairy-wren and Southern Emu-wren (in Brooker (1987), Brooker & Rowley (1991) and southeastern Queensland: McFarland 1988,1994) Russell & Rowley (1993) on the demography of and (in southwestern Australia: Splendid Fairy-wren over a long period (to 18 Smith 1987a; Hopkins & Smith in press). In some years) in a southwestern Australian heath. During older heaths (>10–20 years post-fire) there may be this study, the heath was subjected to at least 6 a reduction in productivity (or seed availability) mild fires (which burnt generally relatively small and some of these bird species may decline or parts of the study area) and one hot wildfire disappear (McFarland 1993,1994). For example, (which burnt 95% of the study area). This study is Smith (1985a) gives a successional pathway in remarkable for its length, the detailed monitoring heaths of southwestern Australia from Richard’s of a large number of resident birds and the precise Pipit to Striated Fieldwren to Western measurements of a wide range of life history to Western Whipbird, over a period of 50–80 years parameters. In general, the birds survived fires (although the dynamics in long-unburnt relatively well, but showed changes in behaviour vegetation may be unclear). and population structure for at least 3–5 years The pattern of return of species following fire

64 Fire and Australian Brids

post-fire. All measures of density, productivity and survival increased during fire-free periods. Fire directly affected natality and juvenile survival, and indirectly affected population density, age structure, sex ratio and group composition. The effects of wildfire may have been less than those of mild fires, because of differences in timing with respect to breeding. The longevity of Splendid Fairy- may mask critical recruitment problems post-fire. The delayed response to fire provides a warning that studies which simply report presence/absence (or even abundance changes) in the short-term after fire are likely to substantially underestimate the actual impacts of fire on populations. Repeated fires clearly threaten this population through their impact on reproductive and life history characters. Without recruitment from adjacent unburnt in old heath (or preference for older heaths) was patches (for example, if this site had been a habitat observed by Baker & Whelan (1994) in fragment), this population would have been southeastern New South Wales, Bryant (1991, eliminated by the fire regime. The parameters 1992, 1994) in Tasmania or Watkins (1985), measured in this study were used to model Burbidge et al. (1989,1990) and Cale & Burbidge population responses to a range of fire regimes (1993) in south-western Australia, and these authors and habitat fragmentation scenarios by Brooker & suggested that fire exclusion in heaths may be a Brooker (1994). more appropriate management regime for this threatened species. Threatened species Fragmentation of heaths can exacerbate fire An unusually high proportion of threatened bird management problems for Ground Parrots, as species occurs in heaths (Garnett 1992a), and the isolation may hinder their dispersal to heaths conservation of such species is mostly associated whose age would otherwise render them suitable with the provision of appropriate fire regimes. (Cale & Burbidge 1993). Ground Parrots also The relationship of the threatened Ground occur in coastal closed sedgelands in southern Parrot with fire has been subject to more studies Australia, but their abundance in these is than for any other Australian bird (Table 2). Even independent of age, as this environment is not so, appropriate fire management of the floristically- subject to the major structural and floristic rich shrubby or graminoid heaths favoured by this changes that characterise heaths of varying age species is still contested (Baker & Whelan 1994). In (Meredith et al. 1984; Gill 1996). contrast to early suggestions that fire intervals of In southwestern Victoria and far 4–5 years were desirable for Ground Parrots southeastern South Australia, the threatened (Ridpath 1972), subsequent work in coastal Victoria Rufous Bristlebird occurs in coastal heaths and and southeastern Queensland demonstrated that thickets that have been unburnt for at least population densities were low in heaths of <3 years 2 years, and it probably reaches highest densities post-fire or >18 years post-fire (Meredith et al. in far older (>25 years post-fire) heaths and 1984; Meredith 1984a,b,c; Jordan 1984b,1987b; thickets (Reilly 1991a; Belcher 1993). Fire is McFarland 1989, 1991, 1992, 1993: Fig. 2) in considered to be the major threat to this species in accord with production of seeds. Consequently, this area and Reilly (1991a) recommended that “if burning of heaths at intervals of fire is to be used as a management tool, frequent 8–10 years (in Queensland and southwestern pockets of unburnt country of sufficient size are Australia) or 10–25 years (in Victoria) was vital”. Sufficient cover to protect against predation recommended for this species (Smith 1979b; (of adults and ) may be the reason Meredith 1983,1984a; Meredith & Isles 1980; McFarland 1992; Meredith & Jaremovic 1990). However, no decrease in Ground Parrot abundance

65 Australia’s Biodiveristy - Responses to Fire

for absence from younger heaths for this in very old heath (though some birds persist in insectivorous species. Predation of nests by Foxes heath >45 years post-fire: Cale & Burbidge 1993; Vulpes vulpes is considered a main threat for some Smith 1994), probably because of lowered populations of Rufous Bristlebirds (Garnett productvity (Smith 1985a,1987a,b). 1992b). For Noisy Scrub-birds, vegetation (thickets In heathland populations of the vulnerable and low scrubby eucalypt forests) has to be at least Eastern Bristlebird (in eastern New South Wales 4–10 years post-fire (depending on floristics and and southeastern Victoria), abundance increases landscape position) before it becomes suitable: with age since fire up to at least 9 years, but may vegetation suitability also declines when very old then plateau (Jordan 1984a; Bramwell et al. 1992; (>30 years post-fire) (Smith & Robinson 1976; Pyke et al. 1995). Unburnt patches left after fire Smith 1977, 1979a, 1985a,b,c; Smith & Forrester may be critical for recolonisation (Pyke et al. 1995). 1981) (Fig. 3). The endangered Orange-bellied Parrot has a Where Western Whipbirds occur in heaths, complex relationship to seral stages of heath and these have to be 4-7 years post-fire before button-grass at its breeding grounds in south- becoming suitable (Smith 1985a), and the species western Tasmania (Brown & Wilson 1981,1984). In persists in long-unburnt thickets (up to 50 years general, it prefers to feed in relatively young heaths post-fire: Smith 1991). and sedgelands (<10 years post-fire), though it For all three species, post-fire recruitment or changes feeding preferences for successional stages recolonisation is hampered by relatively poor during different months of the breeding season (in dispersal ability and/or low reproductive rates response to variation in phenology). As its nesting (Smith 1985a) and the very fragmented nature of hollows in adjacent open forests are vulnerable to the habitat (Smith 1979a; Cale & Burbidge 1993). fire, management of fire is critical, and should Hence managers should aim to prevent entire involve the maintenance of a range of fire histories. habitat patches from being burnt (Smith 1987a). In southwestern Australia, changed fire Fire intervals of <10 years will lead to the local regimes and clearing have led to the precipitous decline and possible extinction of these three taxa. decline of four species occurring in heathlands, Fire intervals of >50 years may also be disadvan- thickets or swamp/forest margins (Ground Parrot, tageous. The fire requirements of these three taxa Western Whipbird, Western Bristlebird and Noisy are slightly different, such that an optimum Scrub-bird) and to the extinction of the Western regime for one species will not coincide exactly Australian subspecies of Rufous Bristlebird D. with that of the others (Smith 1987a). Suitability broadbenti litoralis (Milligan 1904; Ashby 1921; of heaths post-fire can be substantially delayed by Carter 1923a,b, 1924; Whittell 1936; Whitley grazing pressure by macropods (Smith 1985a; 1971; Smith 1977, 1985a, 1987a; How et al. 1987; Cale & Burbidge 1993; Hopkins & Smith 1996). Garnett 1992b; Cale & Burbidge 1993). The endangered Western Bristlebird and Noisy Scrub- bird have persisted only in very localised populations which have escaped frequent burning because of topographic protection from fire, or luck (Smith 1979a; Burbidge et al. 1986; Cale & Burbidge 1993). While the habitats of these three species differ somewhat (the Noisy Scrub-bird is not really a heathland species), it is convenient to consider them together because of their co- occurrence in a small number of sites in coastal southwestern Australia, and the detailed research conducted at these sites on relationships between all three species and fire (Smith 1987a; Cale & Burbidge 1993; Hopkins & Smith 1996). Western Bristlebirds require dry heath of at least 6–10 years post-fire or wet heath of at least 3 years post-fire before habitat is suitable (Smith 1977,1985a,1987b), and density generally declines Figure 3: Abundance of male noisy scrub-birds after fire (after Smith 1985c)

66 Fire and Australian Brids

Fire management for bird conservation With rare exceptions, heathlands have generally been burnt more frequently than is desirable for threatened heathland birds. Brooker & Rowley (1991) suggested fire-free intervals of at least 10 years for small heathland generally, and longer intervals (>20 years) are probably preferable for most threatened species (with the exception of Orange-bellied Parrot). A few currently common heathland species (e.g. Striated Fieldwren) may be disadvantaged by infrequent fires, and a mosaic of fire ages should be maintained. Many threatened heathland birds have low reproductive output and limited dispersal abilities. These characteristics suggest that protection of connecting vegetation, or unburnt patches, are critical for post-fire recolonisation (Danks 1991; Du Guesclin et al. 1995).

MALLEE Figure 4: Response of five bird species to fire in mallee-broombush (after Woinarski 1989b) Fire regimes As the vegetation becomes taller and denser In mallee and mallee-heath environments, early (10–30 years post-fire), vegetation favours the European colonists used fire as part of the mallee endemic birds (Meredith 1982) and the intrusion and clearing process (Harris 1990). Over density of Red-lored Whistler, Crested Bellbird, the course of the last century, extensive wildfires Purple-gaped Honeyeater and Southern Scrub- of high intensity have been frequent, and old (>30 robin peaks (Carpenter & Matthew 1986; years post-fire) mallee has become extremely Woinarski 1987, 1989a, b). scarce and fragmented. In response to such Older (>30 years post-fire) vegetation (with wildfires, fuel-reduction burning in mallee lands tall mallees and relatively open understorey) is has been widely used, though the practice is favoured by Malleefowl, Black-eared Miner, contentious (Cheal et al. 1979; Gill 1990). White-browed Babbler, Species responses to single fires (Carpenter & Matthew 1986; Benshemesh 1988,1990,1992; Woinarski 1989a,b; Silveira There is little information on mortality directly 1993), and some hollow-nesting birds such as due to wildfire in mallee habitats, though Striated Pardalote and Regent Parrot (Cheal et al. Benshemesh (1990) noted that at least 10 of 11 1979; Carpenter & Matthew 1986; Emison & marked Malleefowl survived a patchy but intense Bren 1989). burn, but within a few months following fire all This general successional pattern varies in but 4 had emigrated or died. Chandler (1973) response to the landscape position (e.g. dune crest suggested that mortality during mallee fires may versus swale, soil type) and spatial context of fires be very substantial. (extent and patchiness of fire, patch isolation, etc.) Species common in recently-burnt areas (Woinarski et al. 1988a). The complex floristic (<1 year post-fire) include mainly widespread mosaic typical of mallee vegetation may provide opportunists (e.g. Nankeen Kestrel, Australian habitat heterogeneity even after extensive fires, as Magpie) (Meredith 1982,1983). Subsequently different dominant plant species regrow at different (1–10 years post-fire), Chestnut Quail-, rates post-fire, thereby providing a range of Tawny-crowned Honeyeater, Hooded Robin, vegetation structure even across an area with uniform Red-capped Robin, White-fronted Honeyeater fire history. This allows some species to serially track and Shy Heathwren become abundant (Carpenter preferred vegetation structure by moving between & Matthew 1986; Woinarski 1989a,b) (Fig. 4). floristic communities (Carpenter & Matthew 1986; Woinarski et al. 1988a). Weather conditions

67 Australia’s Biodiveristy - Responses to Fire

following fire may also influence bird responses, with The vulnerable Mallee Emu-wren is closely Chandler (1973) noting that drought post-fire may associated with Triodia hummock grasses under accentuate the fire impact. mallee. Silveira (1993) noted that it was absent for 5-6 years post-fire, and then persisted in long- Species responses to fire regimes unburnt vegetation. However, Garnett (1992b) There is no information on the responses of birds reported the species bred at high densities within to repeated fire, or particular fire regimes in 5 years of fire, and Emison et al. (1987) mallee. considered that it was common in young regrowth. Low dispersal ability may limit its Threatened species recolonisation to isolated patches or following A relatively high proportion of threatened birds in very extensive non-patchy fires. The threatened Australia occurs in mallee (Garnett 1992a). Fire is Striated may also be severely one of the most serious threatening processes for disadvantaged by intense and extensive fire threatened mallee birds (Stephens 1992; Silveira (Brickhill 1980; Garnett 1992b; Silveira 1993). 1993; Smith et al. 1994, 1996). The threatened mallee-heath subspecies of The vulnerable Malleefowl requires a dense Slender-billed Thornbill iredalei hedleyi and extensive layer of litter to provide material for may be vulnerable to frequent fire (Matthew 1994). incubation. This is not available until at least Limited data on the rare Scarlet-chested 10–15 years post-fire (Frith 1962; Tarr 1965), then Parrot suggest that this species may prefer increases in older vegetation. Optimum fire interval recently-burnt (3–5 years post-fire) mallee with for this species is at least 60 years (Benshemesh hummock-grass understorey (Forshaw 1981; 1990). Fire intervals of 20 years have been Robinson et al. 1990). estimated to reduce Malleefowl densities to 6% of The vulnerable eastern subspecies of Western the maximum carrying capacity (Benshemesh Whipbird Psophodes nigrogularis leucogaster has been 1990). Young vegetation (<10 years post-fire) may recorded from a wide range of post-fire ages, but have a greater density of food resources (notably appears to be most abundant in mallee vegetation seeds) for this species, and a mosaic of fire ages may 10–25 years old in the Murray Mallee (Woinarski be desirable (Carpenter & Matthew 1986; Brickhill et al. 1988a). The subspecies, 1987; Priddell 1989,1990). Unburnt patches within P.n.oberon, has been recorded only from long- extensively-burnt vegetation appear critical for unburnt (>25 years) vegetation in southwestern recolonisation post-fire or persistence in a mainly Australia (Smith 1985a,1991; McNee 1986). burnt environment (Benshemesh 1988,1990,1992). McNee (1986) recommended that management for Another major threatening process, predation, may this species required long fire-free intervals (at least have increased impact in burnt country 30 years and preferably 50 years). (Benshemesh 1992). Suitably old mallee has The vulnerable Red-lored Whistler is most become rare (Priddell 1990), and maintenance of abundant in mallee 5-30 years post-fire, and may Malleefowl will require management to protect not persist in very old mallee (Woinarski 1987). existing old vegetation and to provide for additional Threatened populations of Major Mitchell’s long-unburnt areas in the future (Benshemesh Cockatoo and Regent Parrot may require mallee 1990, 1992, 1994). which is sufficiently old (probably >50 years post- The endangered Black-eared Miner forages fire) to form suitable nesting hollows (Cheal et al. preferentially in decorticating bark of very old 1979; Emison & Bren 1989; Silveira 1993). (>50 year post-fire) mallee eucalypts (McLaughlin Fire management for bird conservation 1992), though may occur in much younger mallee where it adjoins older vegetation (Silveira Overviews of the management of mallee birds 1993,1995). The main threat to this taxon is have recommended fire regimes which maintain a genetic swamping from the related Yellow- mosaic of vegetation of a range of ages, but with a throated Miner, a process partly brought about by bias towards retention of older ages (Cheal et al. extensive mallee clearing (Garnett 1992b), 1979; Emison & Bren 1989). Meredith (1982) however the limited area of old mallee is a factor noted that the goal of maintenance of a mosaic of in its threatened status (Starks 1987) and vegetation ages was not equivalent to, and unlikely maintenance of patches of old mallee may be an to be achieved by, unplanned random burns. important management requirement. Fire management has been complicated by

68 Fire and Australian Brids

mallee fragmentation, especially in southwestern Species responses to single fires: wildfire Australian and western New South Wales. Bird mortality may be very high in severe wildfires Individual fragments tend to each have a in temperate eucalypt forests (e.g. Hemsley 1967; homogeneous fire history, leading to reduced bird Christensen et al. 1981; Suckling & MacFarlane species diversity within fragments (Menkhorst & 1983; Wegener 1984a,b). Hood (1941) provided a Bennett 1990; Priddell 1990), though their vivid eye-witness account of birds in a forest isolation may offer some fire protection. wildfire: TEMPERATE EUCALYPT “As the terrific blast of fire swept along, OPEN FORESTS birds could be seen rising from the ground or leaving the trees in front of it only to be Fire regimes overpowered by the dense smoke and rising scorching heat. In all cases the birds fell There has been considerable debate about the fire exhausted into the flames”. regimes operating in (and understorey Some bird species, notably aerial insectivores characteristics of) temperate eucalypt open forests (e.g. swifts: Lord 1936; McCulloch 1966) and before European colonisation, and in the early raptors may follow fires, hunting disturbed insects years of settlement (Gill 1981; Nicholson 1981; or . Williams & Gill 1995). The regimes and their Where wildfire has killed the canopy trees, impacts probably showed important differences wholesale recomposition of the bird fauna may between forest types (Wakefield 1970), rendering follow. Recently burnt areas initially have very few generalisations difficult to draw. birds (Dedman 1983a,b,c,d,e), or at least a In response to Aboriginal management, the substantial decline (20-65%: Ratkowsky eucalypt forests of temperate southeastern Australia 1979,1985; Hewish 1983; Braithwaite et al. 1984; were probably generally characterised by low grassy Loyn et al. 1992a) in species richness and understoreys (maintained by either frequent cool abundance. However, even very small birds may firing or very rare fires: Benson 1981; Recher et al. survive hot wildfire (Rowley & Brooker 1987), 1993). With exclusion of Aboriginal management, especially where some patches remain unburnt these forests rapidly developed shrubby (Smith 1989). Surviving individuals may adopt understoreys, probably markedly increasing fuel unusual behaviours or diet, for example Superb levels and hence increasing the probability of hot Fairy-wrens have been recorded feeding in and extensive wildfire (Smith 1979a). While bird scorched canopy (Dedman 1983e), and the diet of species diversity at any given point probably Sooty Owls was found to be much less varied increased in response to greater structural post-fire (Loyn et al. 1986). complexity of vegetation (Recher 1969), the loss of Ground-feeding insectivores or carnivores fine-scale management probably led to increased may invade or remain in burnt areas, probably in regional homogenisation of the environment (in response to the availability of dead or injured prey, terms of floristic variation and seral ages) and hence or its greater accessibility (Hewish 1983). Species reduced regional bird species diversity. Change in such as Richard’s Pipit, Australian Magpie, ravens, understorey structure would also have Flame Robin, Scarlet Robin, Laughing disadvantaged birds reliant upon forests with grassy Kookaburra, Grey , Nankeen Kestrel understorey (Recher et al. 1993), although to an and Superb Fairy-wren colonise open burnt areas extent some of these species may have been able to until regrowth becomes too dense (up to 3 years) substitute the developing mosaic of farmlands and (Christensen 1974; Stokes 1975; Chambers 1983; forests. Braithwaite et al. 1984; Loyn 1985a,b; Reilly Frequent wildfires in temperate eucalypt 1991a,b). Granivorous birds (e.g. Red-browed forests were an inevitable consequence of limited Finch, Blue-winged Parrot, Diamond Dove, European understanding of fire management in Common Bronzewing, , Painted these systems. More recently, in response to the Button-quail) may also invade or become more human deaths and destruction of property in abundant in recently burnt areas (Recher et al. wildfire, fire suppression and/or frequent and 1985; Loyn 1985b). extensive fuel-reduction (or control) burning have In contrast, nectarivorous, frugivorous and been implemented in most temperate eucalypt foliage-gleaning insectivorous birds (e.g. Brown forests. Thornbill, Striated Thornbill, Yellow-faced

69 Australia’s Biodiveristy - Responses to Fire

absent until trees become sufficiently old to form suitable hollows (Milledge & Palmer 1990; Kavanagh 1990,1991). Depending upon the bird and tree species, this may be between 100 and >250 years post-fire (Loyn 1985a,b; Milledge et al. 1991; Nelson & Morris 1994), and hollow availability may continue to increase with tree ages perhaps up to 1000 years (Mawson & Long 1994). The relationship of fire with hollow formation is complex: fires may kill canopy trees but these (and their hollows) may persist as dead stags; fires may lead to hollow formation (or change in dimensions of existing hollows) in surviving trees or may destroy hollow-bearing trees; and frequent fires may alter recruitment processes and hence dictate future availability of hollows (Cowley 1971; Ashton 1975; Saunders 1979; Saunders et al. 1982; Ambrose 1982; Calder et al. 1983; Inions 1985; Smith & Lindenmayer 1988; Inions et al. 1989; Gibbons 1994; Mawson & Long 1994). Figure 5: Response of 6 bird species to wildfire in Species responses to single fires: temperate eucalypt open forest (after Reilly 1991a) control burning Honeyeater, Yellow-tufted Honeyeater, Environmental changes, and hence changes in bird Golden Whistler, Grey Fantail) disappear or are species composition, are less pronounced with generally much less abundant for at least two years single control burns (Christensen & Kimber after hot wildfire (Ratkowsky 1979,1985; Hewish 1975). Where this is managed successfully, fires are 1983; Braithwaite et al. 1984; Recher et al. relatively limited in extent, leave unburnt patches, 1985,1987a; Loyn 1985b; Loyn et al. 1992a; Reilly and do not consume the canopy or kill trees. 1991a,b) (Fig. 5). However, in some cases, wildfires Bird survival during control burns is may stimulate flowering of eucalypts and hence relatively high (Christensen et al. 1981; Abbott & attract greater abundance of some nectarivores Christensen 1994). For example, following a (Recher et al. 1985). Declines may be delayed for control fire in an open forest in southeastern 2-5 years for some species, such as Eastern Yellow Australia, Cowley found that at least 18 of 27 Robin (Marchant 1985; Jordan 1988) and Rufous banded birds living in the understoey remained in Whistler (Reilly 1991b), possibly because of pre-fire territories. However, subsequent structural characteristics of regrowth at this period mortality is unknown and may be high (Suckling & MacFarlane 1983). (Christensen et al. 1981). Pre-fire total bird abundance may be Short-term changes in bird species reached by 2 years post-fire and exceeded 3 years composition are relatively minor (Christensen & post-fire (Christensen 1974; Loyn et al. 1992a). Kimber 1975; Christensen et al. 1985; Tolhurst Rapid regrowth of woody vegetation within 2–3 1996) and largely restricted to understorey species years post-fire allows the return of some foliage- (Christensen et al. 1985; Christensen & Abbott gleaning birds, and the loss of the open-country 1989; Nichols & Muir 1989; Wardell-Johnson & invaders. Most of the original bird assemblage is Nichols 1991; Wardell-Johnson & Christensen then gradually regained (Catling & Newsome 1992). Species which feed from the open ground 1981). By about 50 years post-fire (Loyn 1985a) (such as Flame Robin, Scarlet Robin, Grey only species typical of old-growth forest—hollow- Shrike-thrush, Laughing Kookaburra, Superb nesting birds, honeyeaters, some raptors, trunk- Fairy-wren, Red-browed Finch, Pied Currawong, gleaining insectivores, frugivores and some White-winged Chough, Australian Magpie, Buff- insectivores of the canopy foliage (Scotts 1991; rumped Thornbill and Painted Button-quail in Taylor 1991)—are missing. southeastern Australia and Grey Shrike-thrush, Hollow-dependent species are scarce or

70 Fire and Australian Brids

Western Yellow Robin, Scarlet Robin, Australian Magpie, White-winged Triller, Australian Raven in southwestern Australia: Cowley 1974; Christensen et al. 1985; Christensen & Abbott 1989; Loyn et al. 1992b) generally increase post- fire. Species that feed from relatively dense shrubs initially decrease (e.g. Brown Thornbill, White- browed Scrubwren and Olive Whistler southeastern Australia and White-browed Scrubwren and Inland Thornbill in southwestern Australia) or disappear (Red-winged Fairy-wren, White-breasted Robin and Golden Whistler in southwestern Australia: Christensen & Kimber 1975; Christensen et al. 1985), but generally return (and sometimes exceed pre-fire abundance) within 2–3 years post-fire (Christensen & Kimber 1975; Christensen et al. 1985) (Fig. 6). Birds persisting in burnt areas may undergo change in behaviour and diet (Wooller & Calver 1988). Control fires generally have little impact on canopy birds (Christensen & Kimber 1975; Christensen et al. 1981,1985), but increased Figure 6: Response of a ground-feeding bird (Scarlet numbers of lorikeets and honeyeaters may be Robin) and shrub-feeding bird (Golden Whistler) to attracted to fire-induced flowering of eucalypts control burn in temperate eucalypt open forest. (after (Christensen et al. 1985; Recher et al. 1985; Christensen et al. 1985) Christensen & Abbott 1989; Loyn et al. 1992b). to sustained changes in understorey floristics and The season of burning has some impact on structure, including decline in litter and increase responses (Loyn et al. 1992b), but this may be very in weeds (Cowley 1971; Recher & Serventy 1991), minor (Christensen et al. 1985). At least some and ultimately to changes in recruitment patterns birds raised young post-fire from nests started just for canopy trees. pre-fire (Cowley 1974). Where control fires are Where fire regimes lead to relatively unusually hot, bird responses may be more permanent, rather than short-term, changes in substantial and recovery more gradual, possibly understorey, the bird species composition of the because of greater depletion of invertebrates in understorey will also be relatively permanently hot fires (Christensen et al. 1985). changed. Where frequent control burning is used Total bird abundance may decline to reduce fuel loads over sustained periods, birds immediately after fire, then exceed pre-fire levels which favour shrubby undergrowth (e.g. Golden by 5 months post-fire (Kimber 1974; Christensen Whistler) or dense leaf litter (e.g. Pilot-bird) will & Kimber 1975). However, there may be much be disadvantaged (Rowley et al. 1988; Recher & variability in such responses: for example, Wooller Lim 1990; Taylor 1991; Recher & Serventy 1991; & Calver (1988) reported a decline in bird Recher et al. 1993). Birds which require relatively abundance that was sustained for at least 3 years open understoreys may be favoured by very post-fire. frequent fires or very long intervals between fires, Species responses to fire regimes but not by intermediate frequencies. For example, Superb may decline or disappear from While relatively minor impacts of control fires forests without frequent control burning because have been demonstrated in many studies, these are of the spread of wiregrass several years after fires essentially short-term responses to single (or few) (Cowley et al. 1969; Catling & Newsome 1981; fire events, and may belie more substantial or Suckling & McFarlane 1983; Smith 1994), but insidious long-term impacts of a fire regime of occur also where long-unburnt forests develop a frequent control burns (Recher et al. 1985). dense tall shrubby understorey which shades out Consistent and frequent burning is likely to lead grass (Loyn et al. 1985a,b).

71 Australia’s Biodiveristy - Responses to Fire

Frequent burning may also lead to change in Threatened species the spatial patterning of eucalypt open forests A small proportion of the bird species in relative to vegetation dominated by more fire- temperate open forests is threatened relative to sensitive plant species (notably rainforest). Such those in other habitats (Garnett 1992a). Forestry change may be to the disadvantage of birds of and clearing, rather than fire, are the primary rainforests or their margins (Ferrier 1985; Holmes threatening process for most of these taxa. Most 1988,1989). of the threatened bird species occurring in While milder fires generally have less impact temperate eucalypt open forests (e.g. Sooty Owl, than hot wildfires on invertebrates (Majer Masked Owl, , southern subspecies 1984,1985; Christensen et al. 1985) and other of Red-tailed Black-cockatoo and Long-billed resources for birds, the cumulative impacts of Black-cockatoo) are associated with old repeated mild fires upon these resources, and vegetation, usually because of their requirement hence upon birds, may be more severe (Friend for hollows. The endangered Orange-bellied 1995), although the limited evidence is not Parrot nests in hollows in eucalypt open forest consistent (Springettt 1976; Abbott et al. 1984; fringing heaths and button-grass in Friend 1995). southwestern Tasmania. Fire management is Forestry critical for these species, to ensure protection of existing hollows and to allow the development of Much of the management of temperate eucalypt sufficient areas of suitable old trees for the future. forests is associated with forestry. Fire The endangered Helmeted Honeyeater is management is a significant component of forestry restricted to riparian eucalypt forests near activities. The impacts upon birds of interactions Melbourne. Two small isolated populations of fire and forestry practice (or comparisons disappeared after recent wildfire (Backhouse 1987). between the two) have been considered by Cowley While widlfire is a continuing threat to the only (1971), McIlroy (1978), Wilson (1981), Green remaining population, fire prevention works may (1982), Tingay & Tingay (1984), Loyn (1985a), create more disturbance than fire (Backhouse 1987). Recher et al. (1985,1987b), Dickinson et al. (1986), The endangered Forty-spotted Pardalote Recher (1991), Curry (1991), Milledge et al. occurs in coastal eucalypt open forests in (1991), Taylor (1991), Wardell-Johnson & Nichols Tasmania. Isolated populations may have been (1991), Wardell-Johnson & Christensen (1992) eliminated by wildfire, but mild fires are probably and Abbott & Christensen (1994). In some not detrimental to this canopy-feeding bird eucalypt forests (notably E.regnans and (Rounsevell & Woinarski 1983), although its E.delegatensis), hot fire following harvesting is requirement for hollows for breeding suggests essential for tree regrowth. Such fires may that fire management may be required for its exacerbate harvesting effects for species such as maintenance. Spotted Quail-thrush (Wilson 1981), though The rare Rufous Scrub-bird and northern McIlroy (1978) claimed that post-harvest fires had populations of the vulnerable Eastern Bristlebird little impact on bird species. Piles of slashed inhabit ecotones between rainforests and eucalypt vegetation left after harvesting may attract more open forests, and are threatened by too frequent birds if left unburnt (Dickinson et al. 1986; Curry fires (Ferrier 1985; Holmes 1988,1989), though 1991). Recher et al. (1987b) noted that the effects both may also be disadvantaged by rainforest of wildfire upon birds were more severe in small expansion associated with fire exclusion. retained (unlogged) strips than in larger strips or in larger unlogged areas. Regrowth forests Fire management for bird conservation following logging showed greater impact of A number of studies have recommended that fire wildfire upon birds than did burnt unlogged management for bird conservation in eucalypt forests (Recher et al. 1985), and the combined open forests should involve the flexible use of a effects of logging and fire were more serious than broad range of fire regimes, with specific attention either individually. Hollow availability may be to threatened bird species (e.g. Recher 1981,1991; more affected by forestry activities than by fire Wardell-Johnson et al. 1989). Until more is regime (Lindenmayer et al. 1990). known about the long-term effects of particular regimes, this is probably the most prudent strategy.

72 Fire and Australian Brids

3.4 TEMPERATE WOODLANDS Bee-eater, Black-faced Cuckoo-shrike, White- winged Triller, Superb Fairy-wren, , Rufous Songlark, Grey Fantail, Brown Fire regimes Treecreeper, Varied , Grey-fronted Little information is available on the pre- Honeyeater, White-naped Honeyeater, Noisy European fire regimes in temperate woodlands, or Friarbird and Dusky Wood-. Golden the impacts of current regimes on birds of this Whistler and Speckled Warbler were more habitat. Hopkins (1985a) considered that fires common in the unburnt area. were probably infrequent in temperate woodlands Open-country bird species invaded Banksia of southwestern Australia under Aboriginal land woodlands in southwestern Australia soon after management. He also noted that single fires in fire, but most species present before fire were woodlands could have dramatic effects on again recorded within a few months after fire vegetation structure, that repeated burning could (Bamford 1985a,b). No species showed a cause permanent structural and floristic changes preference for the long unburnt (22 years post- and that regeneration rates were generally very fire) sites. slow. In contrast, Prober & Thiele (1993) considered that woodlands in a slightly higher Species responses to fire regimes rainfall area in southeastern Australia were burnt There is very limited information on response of every year under Aboriginal land management, birds to repeated fire, or particular fire regimes, in and this maintained a floristically rich grass/herb temperate woodlands. Adam & Robinson (1996) understorey dominated by Themeda. found greatly reduced density of Grey-crowned Subsequent to European settlement, in both Babblers in roadside remnants that were annually southeastern and southwestern Australia, burnt compared to unburnt strips, in association woodlands have been much affected by clearing, with greatly reduced shrub cover. fragmentation, grazing, forestry operations and alteration of ecological processes (including Threatened species changed fire regimes) (Saunders 1989; Saunders & Woodland birds are declining across much of Curry 1990; Bennett 1993; Robinson 1994; Yates southern Australia (Recher & Lim 1990; et al. 1994). The age structure of most woodlands Robinson 1991). In many woodlands, hollow- has probably markedly changed since European nesting birds (and other fauna) are threatened by colonisation, with removal of most old trees on current (and/or projected) limited supply of public lands and absence of regrowth on private hollows (Saunders et al. 1982; Traill 1993; Bennett (grazed) lands (Bennett 1993). 1993), which may be influenced by fire regimes (Webster & Ahern 1992; Quin & Baker-Gabb Species responses to single fires 1993). Birds which or forage on the ground Turner (1987,1992) considered short-term are threatened by introduced predators and (8 months post-fire) and longer-term (8 years floristic or vegetation structural changes post-fire) responses of birds to a single wildfire in associated with grazing or changed fire regimes a mixed Callitris-Eucalyptus woodland in (Bennett 1993). A recent review of threatened southeastern Australia. In the short-term, bird woodland birds (Robinson 1994) did not list abundance and richness was much reduced. inappropriate fire regimes as a major threatening Crested Pigeon, cuckoos, Rainbow Bee-eater, process, but little is known of the fire regimes Speckled Warbler, White-browed Babbler, preferred or required by such species, nor of White-plumed Honeyeater, Little Friarbird, impacts of the interaction between fire and Yellow-faced Honeyeater, Chestnut-breasted grazing. Management of fire may be a major Mannikin, White-browed Wood-swallow, Superb concern if grazing is excluded from woodland Fairy-wren and declined, while conservation reserves. Emu, White-winged Triller and Dusky Wood- Casuarina or Allocasuarina may be a locally swallow were more common in burnt vegetation. prominent component of some woodland At 8 years post-fire, richness and abundance of communities and provides focal resources for birds was greater in the regrowth vegetation. some bird species, notably the endangered Glossy Species more common in the burnt area included Black-cockatoo. A high frequency of intense fires Galah, Cockatiel, , Rainbow since European settlement has led to decline in

73 Australia’s Biodiveristy - Responses to Fire

mature Casuarina, and hence of this dependent patterns comparable to those in temperate open bird (for example, its extinction from forests (Bowman 1988). is associated with extensive hot wildfires around Species responses to single fires 1920: Green & McGarvie 1971). Its persistence requires management to ensure fire exclusion, or There may be very marked short-term response by long intervals between intense fire (Joseph 1982; birds to fire in tropical eucalypt open forests and Clout 1989). For example, Joseph (1982) found savanna woodlands. Typically, the slow mild fires that woodland areas 22 years post-fire were still lead to little direct bird mortality (Braithwaite unsuitable for Glossy Black-cockatoos. 1985). Many hawking insectivores (e.g. wood- The extinction of the Paradise Parrot from swallows, swifts) and raptors are attracted to fires. grassy woodlands of inland northeastern New From hours to months after fire, a wide South Wales and southeastern Queensland has range of ground-feeding birds (e.g. Torresian been linked to change in the fire regime following Crow, raptors, Pied Butcherbird, Straw-necked European settlement (Chisholm 1922,1945), Ibis, Red-tailed Black-cockatoo, Little Corella, although pastoralism, drought or weeds (singly, or Galah, Blue-winged Kookaburra, Red-backed in combination) may also have been involved Kingfisher, Magpie-lark) is attracted to burnt (Forshaw 1981; Garnett 1992b). areas to feed on resources made more accessible by the removal of the dense grass layer (Crawford Fire management for bird conservation 1972,1979; Beeton 1985; Braithwaite 1985; There is an urgent need for data to guide fire Braithwaite & Estbergs 1987; Press 1987; management for the conservation of temperate Woinarski 1990) (Fig. 7). Many of these species woodland birds. Some fire management actions may track fires or recently-burnt areas across the for Superb Parrot, Turquoise Parrot, Grey- landscape (Crawford 1972; Woinarski & crowned Babbler and Glossy Black-Cockatoo have Tidemann 1991,1992; Woinarski et al. 1992). been described (e.g. Davidson & Chambers 1991; This attraction may be reduced in intense late Dry Davidson & Robinson 1992; Webster & Ahern season fires, as the hot fires may destroy a far 1992; Quin & Baker-Gabb 1993). higher proportion of seeds than do the cooler burns of the early dry season (Woinarski 1990). 3.5 TROPICAL EUCALYPT OPEN Individual bird species may also differ in their FORESTS AND SAVANNA short-term responses to late and early Dry season WOODLANDS fires (Woinarski 1990; Tidemann 1993b). Few bird species of tropical open forests and Fire regimes woodlands seem to be clearly disadvantaged in the short-term by fire. Those that do, include the Fire management by Europeans has varied insectivorous Golden-headed Cisticola and Red- substantially across the extensive tropical eucalypt backed Fairy-wren, which feed and shelter in open forests and savanna woodlands of northern dense grass (Crawford 1972,1979; Woinarski 1990) Australia. In the Top End and Kimberley, the and Purple-crowned Fairy-wren which requires incidence of hot extensive late Dry season fires has dense grass and pandanus (Rowley 1987,1993; increased (Braithwaite & Estbergs 1985; Lewis Rowley & Russell 1993). Early dry season fires 1985). This increase has largely been due to may also destroy the nests and/or young of birds pastoral goals, but a high frequency of extensive which nest on the ground or in grass tussocks late Dry season fires occurs across all land tenures, (notably Partridge Pigeon and Masked Finch: including conservation reserves (Russell-Smith & Woinarski 1990; Lucas & Lucas 1993). Bowman 1992; Russell-Smith & Ryan 1994). In tropical eucalypt open forests and savanna Fires in tropical open forests and savanna woodlands, hot fires late in the Dry season may woodlands are generally relatively mild and destroy hollow-bearing trees, to the detriment of immediate impacts are often restricted to the hollow-nesting birds, including Gouldian Finch understorey (Braithwaite & Estbergs 1985; and Palm Cockatoo (Young 1991; Tidemann Bowman 1988), superficially more similar to 1992; Stanton 1992,1995; Crowley 1995). While control burning than wildfire in temperate this may be a serious threat for Palm Cockatoos, systems. Rapid build up of fuel and long dry the smaller hollows suitable for Gouldian Finch seasons almost inevitably lead to a high frequency appear to be relatively abundant and not limiting of fires, hence generally denying successional (Tidemann et al. 1992).

74 Fire and Australian Brids

fuel levels are kept low by grazing), fire-sensitive vegetation can expand into the open forests and savanna (Russell-Smith & Dunlop 1987; Stanton 1992,1995; Crowley 1995; Harrington & Sanderson 1994). In the Wet tropics of , rainforest expansion has been at the expense of the narrow margin of wet sclerophyll forest, and bird taxa associated with this vegetation (including the vulnerable Southern Cassowary, and isolated populations of Eastern Yellow Robin, Yellow Thornbill, Buff-rumped Thornbill, Crested Shrike-tit, White-naped Honeyeater and Yellow- faced Honeyeater: Crowley 1995; Harrington & Sanderson 1994) are threatened by fire suppression. In contrast, too frequent burning of the rainforest/eucalypt open forest margins is considered threatening to the vulnerable Black- breasted Button-quail in southeastern Queensland, and fire exclusion has been recommended to maintain the dense leaf litter it requires (Hughes & Hughes 1991). Figure 7: Short-term response of birds to fire: Presumably as a result of a complex tropical savanna woodland (after Woinarski 1990) interaction between grazing pressure, burning regime and periods of unusually high rainfall, wet grassland flats on Cape York Peninsula are being invaded by dense Melaleuca forests (Stanton 1995). Species responses to fire regimes, and This has severely disadvantaged the endangered threatened species Golden-shouldered Parrot (Garnett & Crowley The high frequency and intensity of fires late in 1994; Crowley 1995; Stanton 1995). the Dry season has led to the decline of fire- For the tropical eucalypt open forests and sensitive vegetation within the tropical eucalypt savanna woodlands themselves, frequent fires landscape (notably monsoon rainforests, Callitris (annual or biennial) decrease structural complexity and some Acacia woodlands: Bowman 1988; of the vegetation (Bowman et al. 1988) and the McKenzie & Belbin 1991; Russell-Smith & incidence or productivity of shrubs bearing fleshy Bowman 1992; Bowman & Panton 1993; Price & fruits. Hence, tropical open forests and savanna Bowman 1994) since European settlement, and woodlands protected from fire have different bird hence disadvantaged the birds associated with species composition to those burnt frequently these (Woinarski 1993b; Woinarski & Fisher (Porter & Henderson 1983; Woinarski 1990). The 1995a,b). Inappropriate fire regimes may have led more structurally diverse forests resulting from to the extinction of the subspecies of Northern infrequent fires generally having richer bird Scrub-robin from monsoon rainforests and Acacia assemblages, especially of frugivorous birds and thickets of the Northern Territory, although those which forage or nest in shrubs (Porter & whether this population ever existed is Henderson 1983; Woinarski et al. 1988; questionable (Bennett 1983). Evidence from Woinarski 1990). abandoned nesting mounds of the rainforest- Grass species composition may also be dependent Orange-footed Scrubfowl located now determined by the season and frequency of fires in tropical eucalypt open forests suggests that (e.g. Smith 1960; Tothill 1992), which may affect contraction of rainforests, and hence decline of granivorous birds, though the evidence is limited their bird species, occurred also in the period of (Tidemann 1990; Garnett & Crowley 1995b). Aboriginal land management (Stocker 1971; Fitzherbert & Baker-Gabb (1988) considered six Russell-Smith 1985; Bowman et al. 1994b). granivorous bird species (Gouldian Finch, Where fire is excluded (generally only Yellow-rumped Mannikin, Chestnut-backed possible in areas without long dry seasons, or where Button-quail, Partridge Pigeon,

75 Australia’s Biodiveristy - Responses to Fire

Golden-shouldered Parrot and Hooded Parrot) to 3.6 TUSSOCK GRASSLANDS be threatened by too frequent fires in the tropical savannas. Garnett & Bredl (1985) noted that the There is surprisingly little information on the threatened Star Finch was largely restricted to impact upon birds of fire in tussock grasslands long-unburnt patches of grass on Cape York (Fitzherbert & Baker-Gabb 1988; Woinarski Peninsula. On Cape York Peninsula, pastoral 1993a). Research on effects of fires in savanna management often involves “storm-burning” after woodlands is at least partly transferable to the first rains of the Wet season (Crowley 1995). grasslands. Burning patterns in tussock grasslands This may have removed much of the annual across Australia are very variable, ranging from sorghum, possibly to the detriment of the attempted exclusion to frequent firing to promote endangered Gouldian Finch (Tidemann 1993a; “green pick” (Walker & Tothill 1992; Pressland et Tidemann et al. 1993), though possibly to the al. 1992). Their impact is almost everywhere advantage of Golden-shouldered Parrots (Garnett regulated or confounded by grazing pressure. & Crowley 1994). Burning patterns in the floodplain grasslands The decline of two trunk-gleaning of northern Australia have now changed from those insectivores, the Black Treecreeper of open forests employed by Aboriginal people, who lit mild early on Cape York Peninsula and the northern dry season fires to decrease the likelihood of subspecies of Crested Shrike-tit of open forests subsequent extensive destructive fire and to across northern Australia, has been linked to manipulate graminoid species composition to frequent intense and extensive late dry season fires favour Magpie Goose, a pivotal food resource (Robinson & Woinarski 1992; Garnett & Crowley (Jones 1980; Lewis 1989; Lucas & Lucas 1993; 1995a). Lucas & Russell-Smith 1993). Burning patterns The vulnerable Red Goshawk occurs widely now tend to be more erratic, with a higher across open forests of northern Australia. frequency of wildfire (Press 1988). Although there are reports of nestlings being In semi-arid Mitchell grasslands, land killed by fire and nest trees being destroyed by fire management usually attempts fire exclusion (Orr (Aumann & Baker-Gabb 1991) and suggestions & Holmes 1984; Anderson et al. 1988). This leads that high fire frequency is disadvantageous (Debus to change in plant species composition and & Czechura 1988), the evidence to evaluate fire structure of the grasslands, and, as fire promotes effects is insufficient. seeding (Scanlan 1980), to decrease in seed Exclusion of fire has been recommended for availability. The consequences of such habitat and the riparian habitat of the threatened Purple- resource changes to birds are unknown crowned Fairy-wren (Rowley & Russell 1993), (Woinarski 1993a). and less frequent fire (especially of extensive hot Fire is also used extensively to maintain fires) for the threatened Black Treecreeper and semi-arid grasslands under invasion by “woody Crested Shrike-tit (Robinson & Woinarski 1992; weeds” (Walker & Tothill 1992; Sullivan et al. Garnett & Crowley 1995a). 1992), and to aid in the clearance of woody species to create grasslands (Johnson & Purdie 1981; Fire management for bird conservation Anderson & Back 1992). Given the correlation of Fire management for the conservation of birds in bird species richness with vegetation structural the tropical eucalypt open forests and savanna complexity (Recher 1969), this habitat alteration woodlands has emphasized the need for probably reduces local species richness. maintenance of a range of burning practices In temperate grasslands, burning regimes preferably creating a fine-scale mosaic of fire have also changed appreciably from relatively histories (Press 1987; Brooker et al. 1990; Stanton frequent firing by Aboriginal people (Nicholson 1992; Crowley 1995; Garnett & Crowley 1995a), 1981) usually to a contemporary regime of with emphasis on burning early in the dry season attempted fire exclusion with occasional hot to prevent destructive late Dry season fires wildfire (Fitzherbert & Baker-Gabb 1988). (Tidemann 1992; Woinarski 1993a) or “storm- Occasional wildfires in temperate grasslands of burning” in the early Wet season (Garnett & the may have severe impacts on grassland Crowley 1994; Crowley 1995) for the birds, notably the threatened Plains-wanderer management of particular bird species. (Baker-Gabb et al. 1990). Many grasses and other grassland plant species depend upon specific

76 Biodiveristy and bushfires

burning regimes (Lunt 1991). Plant species Longer-term impacts are less clearcut. The diversity is increased by regular burning at about 5 availability of hummock grass seeds generally year intervals (McDougall 1989). While these increases with time since fire up to about 10-20 floristic (and vegetation structural) responses to years but may then decline (Westoby et al. 1988). fire regime probably influence grassland bird Invertebrate numbers may be low in the first year species composition, there is no substantial after fire (Masters 1993). Several bird species information on bird-fire relationships in show preference for, or are restricted to, long- temperate grasslands. unburnt hummock grasslands. These include Early seral grasslands occurring soon after Rufous-crowned Emu-wren, Spinifexbird and the hot fires in heaths and forests attract a wide range threatened , all insectivores of typical grassland birds, such as Richard’s Pipit, which shelter or nest in large clumps of hummock Blue-winged Parrot, quail and Australian Magpies grass (Pedler 1991; Garnett et al. 1993; Reid et al. (Loyn 1985b; Dickinson et al. 1986; Brooker & 1993a,b). The threatened Carpentarian Rowley 1991). Grasswren, of sandstone ranges in the Gulf of Carpentaria hinterland, also requires long- 3.7 HUMMOCK GRASSLANDS unburnt hummock grasslands or, at least, is severely disadvantaged by the current regime of Hummock grasslands form a distinctive and frequent extensive hot fires (CSIRO 1976; extensive environment acoss much of arid and Schodde 1982; McKean & Martin 1985,1989). semi-arid Australia and in sandstone ranges of This may also be the case for the northern Australia. Suijdendorp (1981) (in the Kimberley) and White-throated summarises prevailing fire regimes in hummock Grasswren (of the sandstone grasslands and their broad ecological impacts. massif) (Fitzherbert & Baker-Gabb 1988), Current fire regimes generally differ from the however, although extensive hot fires are now patch mosaic burns used by Aboriginal people to frequent in this environment (Russell-Smith & changed scale of fires with fewer small mild fires Ryan 1994), the evidence of their impact on but more frequent extensive wildfire (Kimber White-throated is unclear (Noske 1982; Griffin et al. 1983; Griffin 1984; Burbidge 1988,1992a,b: Woinarski 1992). 1985; Reid & Fleming 1992; Baker The threatened Night Parrot formerly et al. 1993; Russell-Smith & Ryan 1994). occurred in hummock grasslands and other arid Hummock grasslands may be expanding under habitats across much of inland Australia. Its rapid some current fire regimes at the expense of other decline may have been due to increased frequency vegetation types, such as Acacia woodlands and of extensive fires (Ashby 1924a). Recent reports shrublands (Bowman et al. 1994a; Latz 1995). have been from an area where small control burns Environmental responses to fire in hummock had led to a mosaic of vegetation ages and had grasslands (and other arid communities) may be prevented extensive fire (Garnett et al. 1993). particularly influenced by season of fire and Fine-scale mosaic burning (largely to reduce rainfall following fire (Noble et al. 1984; Reid the likelihood of extensive hot wildfire) has been et al. 1993a). recommended for the conservation of birds In contrast to some detailed studies generally in hummock grasslands (Gibson 1986; demonstrating successional patterns in mammals Pedler 1991; Reid et al. 1993a,b; Baker et al. 1993). and of hummock grasslands (e.g. Masters 1993), there has been little research on the 3.8 ACACIA SHRUBLANDS AND relationships of birds to fire regimes in hummock WOODLANDS grasslands. Reid et al. (1993a,b) found that short- term effects of fire in hummock grasslands of There is relatively little information on the birds central Australia include invasion of recently- of Acacia woodlands and shrublands, fire regimes burnt areas by nomadic open-country birds operating in them, or the effects of fire regimes on (including White-winged Triller, Budgerigar, their birds. Aboriginal use of fire in semi-arid and Black Honeyeater, Crimson Chat, Zebra Finch, arid Acacia communities may have been generally Banded Whiteface and Masked Wood-swallow). restricted to infrequent burning of many small Pioneer plants may provide a rich seed resource patches, creating a fine-scale mosaic of soon after fire (Latz 1995), and many granivores successional states (Griffin & Hodgkinson 1986). are hence attracted to recently-burnt areas (Baker European colonisers in these environments et al. 1993). 77 Australia’s Biodiveristy - Responses to Fire

initially had no fire strategy, or used fire to clear 3.9 OTHER HABITATS country. The incidence of extensive wildfire has accordingly increased markedly (Griffin & In chenopod shrublands, Brooker et al. (1979) Hodgkinson 1986). noted that White-winged Fairy-wrens Some Acacia formations, such as Mulga disappeared from a bluebush site for at least A. aneura, are notably fire-sensitive, and are 5 years post-fire. Wildfire may be a substantial declining under current management (Reid et al. threat to the Nullabor Quail-thrush, especially as 1993b; Bowman et al. 1994a). Reid et al. (1993b) regrowth of chenopods is often prevented (or noted that: delayed) by rabbit grazing (Burbidge & Pedler “given the importance of this species to the 1993,1996; Pedler & Burbidge 1995). associated rich bird community, it would There is little information on the impact of seem that the destruction of extensive areas fire regimes on birds in other Australian of mulga by fire would decimate bird environments. In some cases (such as mangroves, populations, at least until the mulga had saltpans, rainforests), this is because fire may not regenerated to a certain age”. be a major management issue, at least in the short- Most bird species will recolonise mulga by term. about 10–15 years post-fire (Reid et al. 1993a,b), though older mulga may support more mistletoe, 3.10 SPECIAL CASE: ISLANDS and hence the bird species associated with this Increased use and extent of fire following (e.g. White-fronted Honeyeater: Reid et al. European colonisation contributed to the 1993a). Recently-burnt mulga contains many extinction of the Kangaroo Island Emu (Ashby generalist nomadic species, such as White-winged 1924a) and the Glossy Black-Cockatoo on King Triller and Zebra Finch (Reid et al. 1993a). Island (Green & McGarvie 1971), although the The loss, through burning and mechanical evidence for its role in the former case is limited. clearing, of Brigalow Acacia harpohylla forests and Birds which nest in colonies on, or close to, other trees in inland southeastern Queensland, the ground may be particularly vulnerable to fire. may have widespread ripple effects, as this area is The most obvious examples of this susceptibility probably an important wintering base for many are seabirds, whose island (or, occasionally bird species migrating from southeastern Australia mainland) colonies have been frequently (Nix 1993). devastated by fire, often killing many thousands Increased incidence of intense wildfire is (and/or a large proportion) of the nesting adults, regarded as the major conservation threat to birds young and/or (MacGillivray 1910; Hull of Lancewood Acacia shirleyi woodlands 1922; Whitley 1944; Fowler 1945; Abbott 1981; (Woinarski & Fisher 1995 a,b). Lane 1976; White 1979a,b; Garnett 1987; Chatto Conservation managers have recommended 1995), and/or rendering vegetation subsequently that extensive wildfires in Acacia communities less suitable for breeding (Pescott 1976; Paton & should be avoided, and hence biodiversity Paton 1977; Brothers 1983; Brothers & Skira maintained, by mosaic burning, with small mild 1987,1988; Skira & Brothers 1988a,b; Walker & fires (Griffin & Hodgkinson 1986). Hulsman 1993). In many cases, these fires have been deliberately lit by fishermen.

78 Biodiveristy and bushfires

4. RESEARCH ADEQUACY, METHODS AND PRIORITIES

Recher et al. (1985) noted that: (e.g. Meredith et al. 1984; McFarland 1988), this “Considering the frequency with which approach may also provide too few data to enable fires occur in eucalypt forests and woodlands statistical analysis of responses of such species. and their dramatic impact on the In most cases, this approach also fails to landscape, there is remarkably little consider fire regime, concentrating instead on information about the effects of fire on time since last fire. Where more detailed fauna or the long-term consequences of information on fire histories of sites are known burning on forest ecosystems”. (e.g. Russell-Smith & Ryan 1994), systematic This is a striking feature of this review of the sampling of sites should be able to indicate the effects of fire on Australian birds. In general, there impacts of a range of fire regime parameters is insufficient information available to direct rather than simply period since last occurrence. management or to evaluate a range of Repeated sampling of individual sites conservation options. Most information is anecdotal or serendipitous. There have been very There have been a small number of valuable few long-term studies. There have been few studies which have monitored bird assemblages at comparisons of different fire regimes, or of single sites over an extended time following treatments with controls. There have been wildfire (Hewish 1983; Smith 1985a; Reilly relatively few studies with individually-marked 1991a,b; Turner 1987,1992; Brooker & Rowley birds. For some environments, and many 1991; Cale & Burbidge 1993). In some cases these biogeographic regions, there is effectively no have included comparisons with before-fire information on responses of birds to current, or abundances (Hewish 1983; Recher et al. 1987b), alternative, fire regimes (Table 2). or with nearby unburnt vegetation (Hewish 1983; The disparate approaches, and limitations, of Recher et al. 1987b; Turner 1992). More studies research on birds has hampered comparisons, have considered short-term (0–2 years post-fire) synthesis and overview. A similar problem has responses (e.g. Roberts 1970; Ratkowsky recently been described for research on the 1979,1985; Dedman 1983a–e,1984; Recher et al. impacts of fire upon invertebrates (Friend 1995). 1985; Loyn et al. 1992a), during which changes may be most rapid. In general, the interpretation Sampling sites across a range of fire ages (and extrapolation) of these studies is constrained A relatively high proportion of studies have by lack of replication, such that it may be difficult considered succession (or change in bird species to distinguish idiosyncratic site (or fire) factors composition with post-fire age) by comparing bird and responses from more general responses. assemblages at separate sites across a range of ages Because control burning is, by definition, (e.g. Cheal et al. 1979; Meredith et al. 1984; Loyn more tractable than wildfire, there has been 1985a; Bamford 1985a,b; Carpenter & Matthew somewhat more systematic sampling of the short- 1986; McFarland 1988; Woinarski 1989b). Such term responses of birds to control fires (Hodgson research provides a broad-brush response to fire & Heislers 1972; Cowley 1974; Kimber 1974; ages, and is attractive in that sampling of a range Christensen & Kimber 1975; Ratkowsky of ages can be undertaken over a short time 1978,1979; Tingay & Tingay 1984; Wardell- period. However, the ages sampled are Johnson & Christensen 1992; Woinarski 1990; constrained by availability, rendering prediction Loyn et al. 1992b), usually with before-after from beyond the sampled range uncertain (Baker comparisons, and occasionally with comparisons & Whelan 1994). Because different sites are with unburnt controls (Kimber 1974; Christensen sampled to represent different post-fire ages, it is et al. 1985; Woinarski 1990; Loyn et al. 1992b; also likely that fire effects are compounded or Tolhurst 1996) or between a range of control confounded by other environmental factors or burning approaches (notably season of burn: local site effects (Burbidge et al. 1989). Except Woinarski 1990; Loyn et al. 1992b; Tolhurst 1996). where specific searches are made for rare species However, these studies cover a remarkably small proportion of the environments in which

79 Australia’s Biodiveristy - Responses to Fire

control burning is regularly and extensively As noted above, a special case may be birds which implemented. Most also report changes in the nest in isolated colonies, where a single fire may more common species only (for which more data destroy many individuals and/or a large are compiled), and hence may overlook impacts of proportion of the population. fire upon rarer birds (Christensen et al. 1985; An unusual relationship between birds and Meredith 1988). Again, the extrapolation from fire is the consumption of ash and charcoal by a these studies is usually hampered by inadequate range of bird species (Baldwin 1965; Coate 1985; replication; though generally consistent findings Pescott 1985; Hutchins 1988), which may be across a range of studies at separate locations related to mineral requirements for breeding, but suggest that the main conclusions are probably limited analysis has failed to establish any pattern robust. or compelling explanation of this feature. Sampling sites of known long-term Autecological studies fire regimes Some detailed long-term autecological studies There have been extremely few studies which have revealed important subtleties of fire impacts, have examined longer-term effects upon birds of such as delayed responses due to reduced breeding specified fire regimes (Meredith 1988). success in the years after fire (Marchant 1985; Exceptional have been studies in tropical eucalypt Rowley & Brooker 1987; Brooker & Rowley open forests, where Porter & Henderson (1983) 1991; Russell & Rowley 1993), the fate of compared bird assemblages in plots of three dispersing birds, role of unburnt patches and different fire regimes (unburnt, annually burnt complex patterns of use of mosaics of different fire and burnt at 2–5 year intervals) maintained over age (Smith 1979a,1985a,1987a; Brown & Wilson nearly 30 years; and Woinarski (1990) compared 1981,1984; Benshemesh 1992), change in bird assemblages in plots of four regimes population age structure, sex ratio and (unburnt, burnt annually in the early dry season, survivorship (Rowley & Brooker 1987; Brooker & burnt annually in the late dry season and burnt Rowley 1991; Russell & Rowley 1993), and biennially) maintained over 14 years. The critical requirements for apparently trivial difficulty of maintaining consistent fire treatments resources (e.g. cobwebs) which may be over such long period (and the unreality of particularly affected by fire (Brooker & Rowley assuming that such strictly-enforced regimes 1991; Recher 1991). The subtlety but importance could occur beyond experimental sites) has of these factors suggests that brief monitoring provided a strong disincentive for such studies. post-fire may provide misleading assessments of longer-term impacts. It also demonstrates that Fire season and breeding management of fire for the conservation of There have also been very few studies which have individual bird species must be preceded by compared the effects of fires occurring at different detailed long-term research (Smith 1987a). seasons. Several authors have advocated that fires should avoid the breeding season, as fires then Translation of research to management may destroy a high proportion of the season’s Such long-term autecological studies are reproductive output for birds nesting on the particularly valuable in that they can provide ground or in low vegetation (Bedggood 1980; specific information on which to base carefully Reilly 1991a,b; Baker et al. 1993), and this impact targetted fire management (e.g. Hopkins 1985b; may be compounded when fires are frequent Brooker & Brooker 1994). There has been (Lucas & Lucas 1993; Brooker & Brooker 1994). remarkably little effective translation of research For example, Rowley & Brooker (1987) suggested results to evaluated management options - though that a single hot wildfire had less impact on the some exceptions include Meredith (1982), Smith survival of Splendid Fairy-wrens than did cool (1987a), Benshemesh (1990), McFarland (1992), control fires, because the wildfire occurred after Cale & Burbidge (1993) and Hopkins & Smith the breeding season whereas the control fires were (1996) - possibly because of the paucity of earlier. In contrast, Christensen et al. (1985) quantitative data. considered that occasional fires during the Brooker & Brooker (1994) provide easily the breeding season probably have limited and most detailed assessment of a range of fire insignificant impact on populations of forest birds. management options for any Australian bird

80 Biodiveristy and bushfires

species, using modelling to predict survival/ reported that a forest bird assemblage proved extinction probabilities for a range of fire regimes remarkably resilient to a wildfire largely because across a range of habitat patch sizes. However, this of small areas (in gullies) that were left unburnt. case is exceptional, as it is based on data from one Rowley & Brooker (1987) noted that Splendid of the very few long-term studies of the responses Fairy-wrens that retained some unburnt patches of individual species to fire (Rowley & Brooker in their territories persisted after fire, whereas 1987; Russell & Rowley 1993). Comparable data those whose territories were completely burnt are available for very few other species (probably disappeared. However, Brooker & Rowley (1991) being limited to Noisy Scrub-bird, Western found no apparent preference for unburnt patches Bristlebird and possibly Western Whipbird, of heath among a range of bird species after fire. Rufous Scrub-bird and Ground Parrot). Yet such Recolonisation of burnt areas is generally modelling may be vital for management of fire- quicker if smaller areas are burnt and are susceptible threatened species, to decide whether surrounded by similar unburnt vegetation (e.g. to translocate populations, how to protect Reilly 1991b), but again the evidence is more populations from fire, whether to provide a range anecdotal than analytical. Small burnt areas may of fire ages in different patches, etc. These studies attract more concentrated populations of post-fire provide good examples of how research should scavengers than extensively burnt areas precede and guide management. (Woinarski 1990). In the absence of a sufficient research Small distant isolates which are unlike their history, fire management (especially in surrounds are less quickly recolonised post-fire conservation reserves) should conservatively than are large isolates close to unburnt source include a broad range of regimes, so as not to areas (Reilly 1991a,b; McFarland 1991), and foreclose options (Recher 1981,1991; Christensen unburnt corridors may be important in this et al. 1985; Brooker et al. 1990). Fire management recolonisation (Danks 1991; DuGuesclin et al. should be accompanied by ongoing monitoring of 1995). Small fragments may also suffer more impacts upon the fauna, and particularly on the severe fire impacts than large fragments (Recher known fire-sensitive species, and assessment of et al. 1987b), possibly because there is a higher these impacts should then be used to refine, or probability of some unburnt areas (or some more precisely target, that management (Wardell- survivors) in larger fragments. Increased Johnson et al. 1989). Acceptance of the fragmentation of suitable habitat, either through importance of fire management for conservation clearing or widespread application of of fauna is relatively recent. Only 20 years ago, inappropriate fire regimes, may exacerbate local Newsome et al. (1975) considered that: fire effects, especially for species with relatively “it is too early to think of utilising fire as a poor dispersal ability. tool for management of fauna in National In some cases, fragmentation may also offer Parks”. some protection from landscape-wide fires (for example, much of the limited long-unburnt mallee Landscape context exists as small patches isolated by farmland), and Relatively little research has been directed at possibly more flexibility in manipulating spatial analysis of fire effects, such as the role of vegetation ages (e.g. because it is relatively easy to unburnt patches, the relationship between fire impose contrasting regimes on different isolates). extent and direct impacts and recolonisation, and However, in general, fragments are probably more the process of recolonisation of burnt isolates. prone to homogenising fires than are extensive Many experienced field workers have reported habitats (Menkhorst & Bennett 1990), support that fires which comprehensively burn areas have relatively small populations and are less likely to be greater impacts than patchy burns which leave recolonised post-fire (Brooker & Brooker 1994), some unburnt areas (e.g. Recher 1981; Smith with the result that they are less likely to retain 1989; Reilly 1991a); however there is little environmental diversity (Williams et al. 1994) and quantitative documentation of the role, or sustain, over the long-term, populations of species required size, of unburnt patches. Using radio- which are associated with particular seral stages. telemetry, Benshemesh (1990) demonstrated the Fragments, particularly smaller ones, may also be importance of small unburnt patches in the post- subjected to fire regimes which are fire persistence of Malleefowl. Smith (1989)

81 Australia’s Biodiveristy - Responses to Fire

unlike those occurring in extensive habitats, Grazing by other herbivores may delay or typically being burnt much more frequently or stall recovery of vegetation after fire, extending or much less frequently (depending upon their magnifying fire impact. The interaction of rabbits, surrounds). This disruption of ecological processes fire and chenopod shrublands is threatening is recognised as a major management problem in Nullabor Quail-thrush (Burbidge & Pedler 1996). fragmented habitats (e.g. Abensperg-Traun & Grazing by macropods may stall regrowth of Smith 1993; Yates et al. 1994), and is likely to be heath, to the detriment of Western Bristlebirds critically important to fire-susceptible threatened (Cale & Burbidge 1993). bird species in fragmented habitats. Fire management, even in conservation reserves, serves many purposes other than Fire control optimising environments for particular bird species. There has been little research on the impacts In order to include bird conservation in a multi- upon birds of fire control mechanisms other than purpose fire management plan, it is critical that fuel reduction burning. Adam & Robinson (1996) research be directed at the assessment the and Davidson & Robinson (1992) suggested that consequences for birds of fire/land-use interactions. land managers who must protect property from There has also been little study of the wildfire should preferably use judicious slashing interaction between fire and predation, although rather than control burning for roadside remnants the compounding effects of these two factors has used by Grey-crowned Babblers. However, Schulz been recognised (e.g. Braithwaite & Estbergs 1987; (1991) noted that slashing also had problems for Benshemesh 1990; Recher 1991). While birds may the conservation of this species. The provision of survive fire, lack of vegetation cover in burnt areas water sources for fire fighting may locally alter may make them far more susceptible to predation, bird community dynamics in arid or semi-arid and predator density may increase in recently- areas (Cheal et al. 1979; Meredith 1982). burnt areas. Firebreaks and associated road networks are likely Multidisciplinary studies to lead to increased penetration of predators, and increased predation post-fire (May 1994). With notable exceptions (e.g. Recher et al. 1985; Tolhurst & Flinn 1992), almost all research on the Interactive effects between fire and other effects of fire upon birds has been carried out in land uses or threatening processes an ecological vacuum, without synchronous There has been limited research directed at the complementary studies of the responses of other impacts upon birds of fire/forestry interactions components of the ecosystem. However (e.g. Recher et al. 1985,1987b; Dickinson et al. multidisciplinary approaches may provide the 1986), but many aspects of fire management in mechanisms for explaining why individual bird forestry operations have not been assessed in species respond in particular ways to a given fire terms of impacts upon birds. or fire regime, and hence provide powerful Other than preliminary studies by Brooker capability to better guide management. (1988) and Tidemann (1990), there has been no research investigating the impacts upon birds of Representation of biogeographic regions and environments interactions between fire and grazing by livestock. Grazing by stock leads to changes in fuel loads, The research effort directed at the relationships floristic composition, vegetation structure, between Australian birds and fire has been very recruitment of tree species, successional process inequitably distributed across habitats and and, hence, to different fire regimes to those geographic regions (Table 2). While the known operating in ungrazed areas. Pastoralists may also occurrence of fire-sensitive threatened birds in impose particular fire regimes with the single aim heaths and mallee is reasonable grounds for giving to provide short-term benefit to stock (Head et al. these environments high priority attention, other 1992). Removal of stock from woodland environments where fire is a major management fragments is advocated as a conservation measure issue have been neglected. Critical gaps occur in the for woodland birds (Robinson 1994), but the long-term effects of control burning in temperate impact of this removal on fire regimes is unclear. eucalypt forests, and in the impacts of fire regimes Removal of the immediate perceived problem, in woodlands (including those dominated by Acacia, grazing, may lead to its replacement with another Callitris and Casuarina species), tussock grasslands, management issue, fire. hummock grasslands and rainforest margins.

82 Biodiveristy and bushfires

5. CONSERVATION OVERVIEW

Of the three species and four subspecies of birds It is striking that, across a broad range of which have become extinct since European environments, most fire-sensitive threatened settlement, inappropriate fire regimes caused, or species require fire intervals longer than those contributed to, the loss of two species (Kangaroo which have been imposed since European Island Emu and Paradise Parrot) and three settlement. Too frequent burning has endangered subspecies (Northern Territory subspecies of species such as Noisy Scrub-bird, Western Northern Scrub-robin, southwestern Australian Bristlebird, Malleefowl and Ground Parrot. The subspecies of Rufous Bristlebird and southwestern old-growth (or mid to late seral) vegetation that Australian subspecies of Lewin’s Rail). Fire is now these species require, or are most abundant in, is recognised as a main threatening process for many now becoming disappearingly rare. The rare, vulnerable and endangered Australian birds. maintenance of suitably-aged vegetation is Brouwer & Garnett (1990) listed inappropriate fire required to retain such species. The endangerment regime as a threat for 22 of 52 threatened of so many species reliant on relatively old Australian bird species, and Garnett 1992b vegetation is a clear indication that land managers considered it threatened 51 taxa (second only to are now generally burning far more extensively or habitat clearance and fragmentation, which frequently than prior to European settlement, or threatened 52 taxa) (Table 3). that fires now are generally more destructive. This recognition of the significant impact of The very low fire frequency, or fire fire is not recent. For example, in reviewing the exclusion, required by many of these species (e.g. status of threatened Australian birds, Ashby preferred intervals of at least 20 years for most (1924b) noted: threatened heathland birds (Smith 1985a) or at “The most serious factor of all is the least 60 years for Malleefowl (Benshemesh 1990)) destruction of both food supply and shelter, will pose serious management problems, and, of course, breeding haunts, by bush- especially where many of these species somewhat fires. I consider that these … factors, paradoxically live in very fire-prone environments especially …. bush-fires, account for more (mallee and heath), where potentially competing than nine-tenths of the disappearance of fire management goals exist, and where adjacent certain forms … I am confident that the populations of humans lead almost inevitably to indiscriminate burning of bush, which is uncontrollable increased ignition of fires. the concomitant of all farming and grazing The long fire-free periods that many of operations, is by a long way the major cause these fire-sensitive birds require, and the very old of the disappearance of many of our rarer trees (generally centuries to millenia old) required birds”. by most of the many hollow-nesting birds, Despite this long-standing appreciation of the suggests that much of the bird fauna has adapted problem, there has been very limited research to an environment which has a long history of directed at the impacts of fire regimes on birds and little or mild disturbance. The association with, or little informed management of fire for bird reliance upon, long-unburnt vegetation by a conservation. To a large extent, myth, anecdote and significant component of the bird fauna is in casual observations substitute for detailed marked contrast to that shown by other vertebrate knowledge in management advice (Meredith 1982). groups, where the successional sequence is often However, recent research has served to focus played out within 10 to 20 years (e.g. Fox 1982; more sharply on the intricacies of fire regime, and Masters 1993), other than for hollow-dependent away from the perception that occasional species. This disparity may be because most of the catastrophic wildfires are inevitably the most mammal, and frog fauna is terrestrial, and serious fire threats to bird conservation (Recher hence responds to the relatively rapid changes in 1981). Rather, minor changes in fire regime may understorey, ground and litter characteristics be critical (Saunders 1985) and lead to almost rather than to the slower changes in taller shrubs imperceptibly gradual, but inexorable, bird and trees. Alternatively, the lesser mobility of decline (Brooker & Brooker 1994). mammals, reptiles and frogs may have precluded

83 Australia’s Biodiveristy - Responses to Fire

the development of species tightly associated with (such as Richard’s Pipit, Little Button-quail, Blue- old seral stages. winged Parrot, Scarlet Robin, Flame Robin) are However, two features generally common to generally widespread and common, and typically most of the fire-sensitive threatened bird species occur after fire across a broad range of are low reproductive output and relatively limited environments. In the first few years post-fire, the dispersal abilities (Smith 1977,1985a,c,1987a,b; presence of such invading generalist species Reilly 1991b): indeed, these features probably alongside recolonisers may lead to local species partly define why these species are threatened by richness which is greater than that pre-fire (e.g. frequent fire. While these are useful adaptive Christensen et al. 1985). However, on a landscape traits in stable (and often resource-poor) scale this is not equivalent to increased environments, they are profound handicaps where biodiversity, as habitat characteristics (and hence that environment is disturbed. Such species are bird species composition) of early seral stages of particularly vulnerable to habitat fragmentation disparate environments tend to be more similar to and fire, and their conservation must involve each other than are later stages. Hence, frequent landscape-scale perspective and management, else regular fires across a range of environments will recovery from fire may be fatally compromised. result in a convergence of their bird fauna, and a Friend (1993) considered the responses to loss of the late seral species which contribute fire of mallee mammals, reptiles and frogs, and much of the distinctiveness to different sought to compare the ecological characteristics of environments. species according to their preferred seral stages. Such homogenisation may be avoided by Following this approach, although close to planned fine-scale mosaic burning with a range of tautological, it is clear that most fire-sensitive bird fire regimes, but with priority protection of species are mainly insectivorous and forage in current long-unburnt areas (and ensuring future dense shrubs, shrubby understorey or thick leaf provision of such areas). Such management has litter. been recommended across a very broad range of The broad requirements of some of these Australian environments (Recher 1981,1991; threatened fire-sensitive bird species are now Christensen et al. 1985; Braithwaite 1985; Emison reasonably obvious. However, many species now & Bren 1989; Brooker et al. 1990). The approach considered reasonably abundant may be more has the advantage of conservatively covering a subtly disadvantaged by current fire regimes. The broad spectrum of fire regime options, until better detailed study of the common Splendid Fairy- knowledge of the requirements of individual wren by Rowley & Brooker (1987), Brooker & species is available. Such knowledge will be Rowley (1991), Russell & Rowley (1993) and acquired only when a much more comprehensive, Brooker & Brooker (1994) has illustrated the strategic and concerted research effort is danger of small but repeated impacts associated established. with a regime of frequent control burning. Where such regimes are sustained over long periods (decades to centuries), the bird assemblage within habitats will be gradually re-sorted and species now considered secure may be lost. Over longer periods (decades to millenia), sustained fire regimes will re-arrange spatial relationships between habitats, resulting in decline or extinction of fire-sensitive habitats (and their associated bird fauna). The long period of such change relative to that of most studies on fire impacts suggests great caution should be exercised in interpretation of studies which suggest minor or no changes associated with fire. In general, currently threatened bird species will benefit from longer intervals between fires. Such change will disadvantage species associated with early seral stages. However, these species

84 Fire and Australian Brids

REFERENCES

Abensperg-Traun, M. and Smith, G.T. (1993). birds of the south-west corner of Western Predictable effects of agricultural Australia. Emu 20, 123–124. development on the long-term availability of Ashby, E. (1924a). Notes on extinct or rare hollows for animals: observations from the Australian birds, with suggestions as to some Western Australian wheatbelt. Pacific of the causes of their disappearance. Part I. Conservation Biology 1, 78–79. Emu 23, 178–183. Abbott, I. (1981). Seabird Islands: No. 106. Ashby, E. (1924b). Notes on extinct or rare Mondrain Island, Archipelago of the Australian birds, with suggestions as to some Recherche, Western Australia. Corella 5, of the causes of their disappearance. Part II. 60–61. Emu 23, 294–298. Abbott, I., & Christensen, P. (1994). Application Ashton, D.H. (1975). The root and shoot of ecological and evolutionary principles to development of Eucalyptus regnans F.Muell. forest management in Western Australia. Australian Journal of Botany 23, 867–887. Australian Forestry, 57, 109–122. Aumann, T., & Baker-Gabb, D. J. (1991). The Abbott, I., Van Heurck, P. and Wong, L. (1984). ecology and status of the Red Goshawk in Responses to long-term fire exclusion: northern Australia. RAOU Report No. 75. physical, chemical and faunal features of Royal Australasian Ornithologists Union. litter and soil in a Western Australian forest. Backhouse, G. N. (1987). Management of Australian Forestry 47, 237–242. remnant habitat for conservation of the Adam, P., & Robinson, D. (1996). Negative effects Helmeted Honeyeater Lichenostomus of fuel-reduction burning on the habitat of melanops cassidix. In D. A. Saunders, G. W. the Grey-crowned Babbler Pomatostomus Arnold, A. A. Burbidge, & A. J. M. Hopkins temporalis. Victorian Naturalist 113, 4–9. (Eds.), Nature conservation: the role of Ambrose, G.J. (1982). An ecological and remnants of native vegetation (pp. 287–294). behavioural study of vertebrates using Chipping Norton: Surrey Beatty. hollows in eucalypt branches. Ph.D. thesis. Baird, R.F. (1984). The Pleistocene distribution of La Trobe University: Bundoora. the Tasmanian Native-hen Gallinula mortierii Anderson, E. and Back, P. (1992). Fire in brigalow mortierii. Emu 84, 119–123. lands. In B.R. Roberts (Ed.) Fire research in Baird, R.F. (1991). Avian fossils from the rural Queensland. (pp. 324–339). of Australia. In P. Vickers-Rich, Toowoomba: University of Southern J.M. Monaghan, R.F. Baird and T.H. Rich Queensland. (Eds.) Vertebrate plaeontology of Australasia Anderson, E.R., Pressland, A.J., McLennan, S.R., pp.809–870. Lilydale: Pioneers Design Clem, R.L. and Rickert, K.G. (1988). The Studios. role of fire in native pasture management. In Baker, J., & Whelan, R. J. (1994). Ground Parrots W.H. Burrows, J.C. Scanlan and M.T. and fire at Barren Grounds, New South Rutherford (Eds.) Native pastures in Wales: a long–term study and an assessment Queensland - the resources and their management of management implications. Emu 94, (pp. 112–124). Brisbane: Queensland 300–304. Department of Primary Industries. Baker, L., Woenne–Green, S., & Mutitjulu Arnold, G. W., Smith, G. T., Rowley, I. C. R., & Community (1993). Anangu knowledge of Brooker, M. G. (1993). The effects of fire on vertebrates and the environment. In J. R. W. the abundance and distribution of animals in Reid, J. A. Kerle, & S. R. Morton (Eds.), Australian ecosystems, with emphasis on Uluru fauna. The distribution and abundance of Mediterranean ecosystems. In L. Trabaud & vertebrate fauna of Uluru (Ayers Rock-Mount R. Prodon (Eds.), Fire in Mediterranean Olga) National Park, N.T. (pp. 79–132). ecosystems (pp. 237–257). Brussels: Canberra: Australian National Parks and Commission of the European Communities. Wildlife Service. Ashby, E. (1921). Notes on the supposed “extinct” Baker-Gabb, D. J., Benshemesh, J. S., & Maher, P.

85 Australia’s Biodiveristy - Responses to Fire

N. (1990). A revision of the distribution, Bowman, D.M.J.S. (1988). Stability amid status and management of the turmoil?: towards an ecology of north Plains-wanderer Pedionomus torquatus. Emu, Australian eucalypt forests. Proceedings of the 90 161–168. Ecological Society of Australia 15, 149–158. Baldwin, M. (1965). Bird eating charcoal. Emu, 64 Bowman, D.M.J.S. and Panton, W.J. (1993). 208. Decline of Callitris intratropica in the Bamford, M. J. (1985a) The dynamics of small Northern Territory: implications for pre- vertebrates in relation to fire in banksia and post-European colonisation fire regimes. woodland near Perth, Western Australia. PhD, Journal of Biogeography 20, 373–381. Murdoch University. Bowman, D.M.J.S., Wilson, B.A. and Hooper, Bamford, M. J. (1985b). The fire-related dynamics R.J. (1988). Response of Eucalyptus forest and of small vertebrates in Banksia woodland: a woodland to four fire regimes, Munmarlary, summary of research in progress. In J. R. Northern Territory, Australia. Journal of Ford (Ed.), Fire ecology and management in Ecology 76, 215–232. Western Australian ecosystems (pp. 107–110). Bowman, D.M.J.S. Latz, P.K. and Panton, W.J. Perth: Western Australian Institute of (1994a). Pattern and change in an Acacia Technology. aneura shrubland and Triodia hummock Bedggood, G.W. (1980). Birdlife between Lake grassland mosaic on rolling hills in central Tyers and Marlo, Victoria. Australian Bird Australia. Australian Journal of Botany 43, Watcher 8, 147–162 25–37. Beeton, R.J.S. (1985). The Little Corella - a Bowman, D. M. J. S., Woinarski, J. C. Z., & seasonally adapted species. Proceedings of the Russell-Smith, J. (1994b). Environmental Ecological Society of Australia 13, 53–63. relationships of Orange-footed Scrubfowl Belcher, C. (1993). Rufous Bristlebird survey and Megapodius reinwardt nests in the Northern habitat analysis, Port Campbell National Park Territory. Emu 94, 181–185. 1992. Department of Conservation and Bradley, J. J. (1995). Fire: emotion and politics. In Natural Resources. D.B. Rose (Ed.) Country in flames. Proceedings Bennett, A.F. (1993). Fauna conservation in box of the 1994 symposium on biodiversity and fire in and ironbark forests. Victorian Naturalist North Australia. (pp. 25–32). Canberra: 110, 15–23. Department of the Environment, Sport and Bennett, S. (1983). The Northern Scrub-robin Territories. superciliaris in the Northern Braithwaite, L. W., Clayton, M., MacLean, L., & Territory. Emu 83, 105–107. Parker, B. S. (1984). Vertebrate fauna of a Benshemesh, J. (1988). Report on a study of 144-ha water catchment within eucalypt forest malleefowl ecology. Department of being harvested for woodpulp at Eden, south- Conservation, Forests and Lands. eastern New South Wales. CSIRO Wildlife Benshemesh, J. (1990). Management of and Rangelands Research. Malleefowl - with regard to fire. In J. C. Braithwaite, R. W. (1985). Fire and fauna. In R. Noble, P. J. Joss, & G. K. Jones (Eds.), The W. Braithwaite (Ed.), Kakadu Fauna Survey. mallee lands: a conservation perspective (pp. Final report to Australian National Parks and 206–211). Melbourne: CSIRO. Wildlife Service. (pp. 634–650). Darwin: Benshemesh, J. S. (1992) The conservation biology of CSIRO. Malleefowl, with particular regard to fire. PhD, Braithwaite, R.W. (1991). Aboriginal fire regimes Monash University. of monsoonal Australia in the 19th century. Benshemesh, J. S. (1994). Malleefowl Leipoa Search 22, 247–249. ocellata. Department of Conservation and Braithwaite, R.W. and Estbergs, J.A. (1985). Fire Natural Resources. patterns and woody vegetation trends in the Benson, D.H. (1981). Vegetation of Upper Alligator Rivers region of northern Australia. Mangrove Creek, Wyong, New South In J.C. Tothill and J.J. Mott (Eds.) Ecology Wales. Cunninghamia 1, 7–22. and management of the world’s savannas (pp. Boekel, C. (1980). Birds of Victoria River Downs 359–364). Canberra: Australian Academy of Station and of Yarralin, Northern Territory. Science. Part 1. Australian Bird Watcher 8, 171–193.

86 Biodiveristy and bushfires

Braithwaite, R., & Estbergs, J. (1987). Fire-birds Service, Tasmania. of the Top End. Australian Natural History Brown, P. B., & Wilson, R. I. (1984). The 22, 299–302. Orange-bellied Parrot. In R. H. Groves & Bramwell, M., Pyke, G., Adams, C., & Coontz, P. W. D. L. Ride (Eds.), Species at risk (pp. (1992). Habitat use by Eastern Bristlebirds 106–116). Canberra: Australian Academy of in Barren Grounds Nature Reserve. Emu 92, Science. 117–121. Bryant, S. L. (1991). The Ground Parrot, Pezoporus Brickhill, J. (1980). Striated Grasswren Amytornis wallicus, in Tasmania: distribution, density and striatus. In C. Haigh (Ed.), Endangered . Scientific report no. 91/1. animals of New South Wales (p. 68). Sydney: Parks, Wildlife & Heritage, Tasmania. NSW National Parks and Wildlife Service. Bryant, S. L. (1992). The Ground Parrot and age Brickhill, J. (1987) The conservation status of of vegetation in Tasmania. In L. Joseph malleefowl in New South Wales. (Ed.), Issues in the conservation of parrots in M.Nat.Res.Sc., University of New England. Australasia and : challenges to Brooker, L. C., & Brooker, M. G. (1994). A model conservation biology. (pp. 42–45). Melbourne: for the effects of fire and fragmentation on Royal Australasian Ornithologists Union. the population viability of the Splendid Bryant, S. L. (1994). Habitat and potential diet of Fairy–wren. Pacific Conservation Biology 1, the Ground Parrot in Tasmania. Emu 94, 344–358. 166–171. Brooker, M. G. (1988). Some aspects of the Burbidge, A. (1985). Fire and mammals in biology and conservation of the Thick-billed hummock grasslands of the arid zone. In J. Grasswren Amytornis textilis in the Shark Bay R. Ford (Ed.), Fire ecology and management in area, Western Australia. Corella 12, 101–108. Western Australian ecosystems (pp. 91–94). Brooker, M. G., & Rowley, I. (1991). Impact of Perth: Western Australian Institute of wildfire on the nesting behaviour of birds in Technology. heathland. Wildlife Research 18, 249–263. Burbidge, A. H., & Pedler, L. (1993). Conservation Brooker, M. G., Ridpath, M. G., Estbergs, A. J., status of the Nullabor Quail-thrush. World Bywater, J., Hart, D. S., & Jones, M. S. (1979). Wide Fund for Nature. Bird observations on the north-western Burbidge, A. H., & Pedler, L. (1996). Habitat of Nullabor Plain and neighbouring regions, the threatened Nullabor Quail-thrush. 1967–1978. Emu 79, 176–190. Wildlife Research, 23. Brooker, M. G., Braithwaite, R. W., & Estbergs, J. Burbidge, A. A., Folley, G. L., & Smith, G. T. A. (1990). Foraging ecology of some (1986). The Noisy Scrub-bird. Western insectivorous and nectarivorous species of Australian Wildlife Management Program 2. birds in forests and woodlands of the Department of Conservation and Land Wet-Dry tropics of Australia. Emu 90, Management. 215–230. Burbidge, A. H., Watkins, D., & McNee, S. Brothers, N. P. (1983). Seabird Islands: No. 136. (1989). Conservation of the Ground Parrot in Actaeon Island, Tasmania. Corella 7, 89–90. Western Australia. Final report (Project 118). Brothers, N. P., & Skira, I. J. (1987). Seabird World Wildlife Fund. Islands: No. 173. Chappell Island, Furneaux Burbidge, A. H., McNee, S., Newbey, B., & Rolfe, Group, Tasmania. Corella 11, 81–82. J. (1990). Conservation of the Ground Parrot in Brothers, N. P., & Skira, I. J. (1988). Seabird Western Australia. Supplementary report: Islands: No. 185. , Furneaux project 118. World Wildlife Fund. Group, Tasmania. Corella 12, 85–86. Calder, T.G., Golding, B.G. and Manderson, A.D. Brouwer, J., and Garnett, S. (Eds) (1990). (1983). Management for arboreal species in Threatened birds of Australia: an annotated list. the Wombat State Forest. Graduate School Royal Australasian Ornithologists Union: of Environmental Science. Environmental Melbourne. Report no. 16. Clayton: Monash University. Brown, P. B., & Wilson, R. I. (1981). A survey of Cale, P. G., & Burbidge, A. H. (1993). Research the Orange-bellied Parrot Neophema plan for the Western Ground Parrot, Western chrysogaster in Tasmania, Victoria and South Whipbird and Western Bristlebird. Australian Australia. National Parks and Wildlife National Parks and Wildlife Service.

87 Australia’s Biodiveristy - Responses to Fire

Campbell, A. G. (1937). Birds of Wilson’s Christensen, P. E. S., Wardell-Johnson, G., & Promontory. Emu 37, 157. Kimber, P. (1985). Birds and fire in Carpenter, G., & Matthew, J. (1986). The birds of southwestern forests. In A. Keast, H. F. Billiatt Conservation Park. South Australian Recher, H. Ford, & D. Saunders (Eds.), Birds Ornithologist 30, 29–37. of eucalypt forests and woodlands: ecology, Carter, T. (1923a). Birds of Broome Hill district. conservation, management (pp. 291–299). Part II. Emu 23, 223–235. Chipping Norton: Surrey Beatty. Carter, T. (1923b). Birds of the Broome Hill Christidis, L. and Boles, W.E. (1994), The district. Part I. Emu 23, 125–142. and species of birds of Australia and its Carter, T. (1924). Birds of the Broome Hill territories., Melbourne: Royal Australasian district. Part III. Emu 23, 306–318. Ornithologists Union. Catling, P. C., & Newsome, A. E. (1981). Clout, M. N. (1989). Foraging behaviour of Responses of the Australian vertebrate fauna Glossy Black-cockatoos. Australian Wildlife to fire. In A. M. Gill, R. H. Groves, & I. R. Research 16, 467–473. Noble (Eds.), Fire and the Australian biota Coate, K. (1985). Black Honeyeaters feeding on (pp. 273–310). Canberra: Australian ash. Western Australian Naturalist 16, 51–52. Academy of Science. Cooper, R. P. (1974). The avifauna of Wilson’s Chambers, W. (1983). Birds at Urquhart’s Bluff: Promontory. Part 1. Australian Bird Watcher before and after the fire. Geelong Naturalist 5, 137–174. 20, 50–51. Cowley, R.D. (1971). Birds and forest Chandler, L. G. (1973). In the wake of a bushfire. management. Australian Forestry 3, 234–250. Wildlife in Australia 10, 140–141. Cowley, R. D. (1974). Effects of prescribed Chatto, R. (1995). The effects of fire on a burning on birds of the mixed species forests breeding colony of Australian Pelicans. of West Central Victoria. In Third Fire Corella 19, 70. Ecology Symposium (pp. 58–65). Melbourne: Cheal, P. D., Day, J. C., & Meredith, C. W. Forests Commission, Victoria. (1979). Fire in the national parks of north-west Cowley, R. D., Heislers, A., & Ealey, E. H. H. Victoria. National Parks Service. (1969). Effects of fire on wildlife. Victoria’s Chisholm, A. H. (1922). The “lost” Paradise Resources 11, 18–22. Parrot. Emu 22, 4–17. Crawford, D. N. (1972). Birds of Darwin area, Chisholm, A. H. (1945). Birds of the Gilbert with some records from other parts of diary, Part 2. Emu 44, 183–200. Northern Territory. Emu 72, 131–148. Christensen, P. (1974). The concept of fauna Crawford, D. N. (1979). Effects of grass and fires priority areas. In Third Fire Ecology on birds in the Darwin area, Northern Symposium (pp. 66–73). Melbourne: Forests Territory. Emu 79, 150–152. Commission, Victoria. Crowley, G. M. (1995). Fire on Cape York Christensen, P., & Abbott, I. (1989). Impact of fire Peninsula. Cape York Peninsula Land Use in the eucalypt forest ecosystem of southern Study. Western Australia: a critical review. CSIRO (1976). A survey of the fauna of the lower Australian Forestry 52, 103–121. McArthur River Region, Northern Territory. Christensen, P. E., & Kimber, P. C. (1975). Effect Report to Mimets Pty Ltd. (CSIRO Division of prescribed burning on the flora and fauna of Wildlife Research; Canberra) of south-west Australian forests. Proceedings Curry, G. N. (1991). The influence of proximity of the Ecological Society of Australia 7, 85–107. to plantation edge on diversity and Christensen, P., Recher, H., & Hoare, J. (1981). abundance of bird species in an exotic pine Responses of open forests (dry sclerophyll plantation in north-eastern New South forests) to fire regimes. In A. M. Gill, R. H. Wales. Wildlife Research 18, 299–314. Groves, & I. R. Noble (Eds.), Fire and the Danks, A. (1991). The role of corridors in the Australian biota (pp. 367–393). Canberra: management of an endangered . In Australian Academy of Science. D. A. Saunders & R. J. Hobbs (Eds.), Nature conservation 2: the role of corridors (pp. 291–296). Chipping Norton: Surrey Beatty.

88 Biodiveristy and bushfires

Davidson, I., & Chambers, L. (1991). Vegetation Emison, W. B., Beardsell, C. M., Norman, F. I., & management for Superb Parrot foraging habitat Loyn, R. H. (1987). Atlas of Victorian birds. in Victoria. Victorian Department of Melbourne: Department of Conservation, Conservation and Environment. Forests and Lands, and Royal Australasian Davidson, I., & Robinson, D. (1992). Ornithologists Union. Grey-crowned Babbler Pomatostomus temporalis. Ferrier, S. (1985). Habitat requirements of a rare Department of Conservation and Natural species, the Rufous Scrub-bird. In A. Keast, Resources. H. F. Recher, H. Ford, & D. Saunders Debus, S. J. S., & Czechura, G. V. (1988). The (Eds.), Birds of eucalypt forests and woodlands: Red Goshawk Erythrotriorchis radiatus: a ecology, conservation, management (pp. review. Australian Bird Watcher 12, 175–199. 241–248). Chipping Norton: Surrey Beatty. Dedman, V. (1983a). G. F. N. C. Otway Fitzherbert, J. C., & Baker-Gabb, D. J. (1988). Regeneration Survey – Progress report. Australasian grasslands and their threatened Geelong Naturalist 20, 52–56. avifauna. In P. D. Goriup (Ed.), Ecology and Dedman, V. (1983b). G.F.N.C. Otways conservation of grassland birds (pp. 227–250). regeneration survey – 2nd progress report. Cambridge: International Council for Bird Geelong Naturalist 20, 98–100. Preservation. Dedman, V. (1983c). Mammals and birds after the Flannery, T. F. (1990). Pleistocene faunal loss: fires: Forest Road, Anglesea. Geelong implications of the aftershock for Australia’s Naturalist 20, 27. past and future. Archaeology in Oceania 25, Dedman, V. (1983d). Moggs Creek: eight days 45–67. later. Geelong Naturalist 20, 22–25. Flannery, T. F. (1994). The future-eaters: an Dedman, V. (1983e). Notes on the effects of the ecological history of the Australasian lands and Ash Wednesday fire on the Ironbark Basin, people. Sydney: Reed. Point Addis. Geelong Naturalist 20, 25–26. Forshaw, J. M. (1981). Australian Parrots. Dedman, V. (1984). Otways regeneration survey. Melbourne: Lansdowne Press. Third progress report. Geelong Naturalist 21, Fowler, S. (1945). The fire on Mondrain Island. 89–92. Emu 44, 334–335. Department of Environment, Sport and Territories Fox, A. (1978). The ’72 fire of Nadgee Nature (1996). Fire and biodiversity: the effects and Reserve. Parks & Wildlife 2, 5–24. effectiveness of fire management. Proceedings Fox, B.J. (1982). Fire and mammalian secondary of the Conference held 8–9 October 1994, succession in an Australian coastal heath. Footscray, Melbourne. Biodiversity Series, Ecology 63, 1332–1341. Paper No. 8. Canberra: Department of Friend, G.R. (1993). Impact of fire on small Environment, Sport and Territories vertebrates in mallee woodlands and Dickinson, K. J. M., Wall, L. E., & Wilson, R. I. heathlands of temperate Australia: a review. (1986). Birds in a partly clearfelled dry Biological Conservation 65, 99–114. eucalypt forest on dolerite in southeastern Friend, G.R. (1995). Fire and invertebrates - a Tasmania. Papers and Proceedings of the Royal review of research methodology and the Society of Tasmania 120, 39–49. predictability of post-fire response patterns. Du Guesclin, P., Smith, S., O’Shea, B., & Debbis, CALMScience Supplement 4, 165–174. C. (1995). “Brushing for bristles”: habitat Frith, H. J. (1962). Conservation of the corridors for the Rufous Bristlebird. In A. Mallee-fowl (Leipoa ocellata, Gould). CSIRO Bennett, G. Backhouse, & T. Clark (Eds.), Wildlife Research 7, 33–49. People and nature conservation: perspectives on Garnett, S. (1987). Seabird Islands: No. 171. private land use and endangered species recovery Kusamet Island, Torrest Strait, Queensland. (pp. 163–165). Chipping Norton: Surrey Corella 11, 77–78. Beatty. Garnett, S. (1992a). The action plan for Australian Emison, W. B., & Bren, W. M. (1989). Common birds. Canberra: Australian National Parks birds of the mallee, northwestern Victoria. In and Wildlife Service. J. C. Noble & R. A. Bradstock (Eds.), Garnett, S. (1992b). Threatened and extinct birds of Mediterranean landscapes in Australia: mallee Australia. Melbourne: Royal Australasian ecosystems and their management (pp. 221–242). Ornithologists Union. Melbourne: CSIRO.

89 Australia’s Biodiveristy - Responses to Fire

Garnett, S., & Bredl, R. (1985). Birds in the Good, R.B. (1981). The role of fire in vicinity of Edward River Settlement. Part II. conservation reserves. In A. M. Gill, R. H. Discussion, references, list of passerines. Groves, & I. R. Noble (Eds.), Fire and the Sunbird 15, 25–40. Australian biota (pp. 529–550). Canberra: Garnett, S. T., & Crowley, G. M. (1994). The Australian Academy of Science. ecology and conservation of the Green, R. H. (1982). The activity and movement Golden-shouldered Parrot. Cape York of fauna in compartment 2, Maggs Peninsula Land Use Strategy. Mountain, Tasmania, in the first five years of Garnett, S., & Crowley, G. (1995a). The decline forest regeneration. Records of the Queen of the Black Treecreeper Climacteris picumnus Victoria Museum 75, 1–31. melanota on Cape York Peninsula. Emu 95, Green, R. H., & McGarvie, A. M. (1971). The 66–68. birds of King Island. Records of the Queen Garnett, S., & Crowley, G. (1995b). Feeding Victoria Museum 40, 1–42. ecology of Hooded Parrots Psephotus Griffin, G.F. (1984). Fire, vegetation, and man. In dissimilis during the early wet season. Emu E.C. Saxon (Ed.) Anticipating the inevitable: a 95, 54–61. patch-burn strategy for fire management at Garnett, S., Crowley, G., Duncan, R., Baker, N., Uluru (Ayers Rock - Mt Olga) National Park & Doherty, P. (1993). Notes on live Night (pp. 7–12). Melbourne: CSIRO. Parrot sightings in north-western Griffin, G.F. and Hodgkinson, K.C. (1986). The Queensland. Emu 93, 292–296. use of fire for the management of mulga land Gibbons, P. (1994). Sustaining key old-growth vegetation in Australia. In P. Sattler (Ed.) characteristics in native forests used for The mulga lands. (pp.93–97). Brisbane: Royal wood production: retention of trees with Society of Queensland. hollows. In T.W. Norton and S.R. Dovers Griffin, G.F., Price, N.F. and Portlock, H.F. (Eds.) Ecology and sustainability of southern (1983). Wildfires in the central Australian temperate ecosystems (pp. 59–84). Canberra: rangelands, 1970–1980. Journal of CSIRO. Environmental Management 17, 311–323. Gibson, D. F. (1986). A biological survey of the Hallam, S. J. (1975). Fire and hearth: a study of in the Northern Territory. Aboriginal usage and European usurpation in Conservation Commission of the Northern south-western Australia. Canberra: Australian Territory. Institute of Aboriginal Studies. Gill, A. M. (1981). Post-settlement fire history in Harrington, G. N., & Sanderson, K. D. (1994). Victorian landscapes. In A. M. Gill, R. H. Recent contraction of wet sclerophyll forest Groves, & I. R. Noble (Eds.), Fire and the in the wet tropics of Queensland due to Australian biota (pp. 77–98). Canberra: invasion by rainforest. Pacific Conservation Australian Academy of Science. Biology 1, 319–327. Gill, A. M. (1990). Fire management of mallee Harris, C. R. (1990). The history of mallee land lands for species conservation. In J. C. use: Aboriginal and European. In J. C. Noble, P. J. Joss, & G. K. Jones (Eds.), The Noble, P. J. Joss, & G. K. Jones (Eds.), The mallee lands: a conservation perspective (pp. mallee lands: a conservation perspective (pp. 202–205). Melbourne: CSIRO. 147–151). Melbourne: CSIRO. Gill, A. M. (1996). How fires affect biodiversity. Haynes, C. D. (1985). The pattern and ecology of In DEST (Eds.), Fire and biodiversity: the munwag: traditional Aboriginal fire regimes effects and effectiveness of fire management. in north-central Arnhemland. Proceedings of Proceedings of the conference held 8–9 October the Ecological Society of Australia 13, 203–214. 1994, Footscray, Melbourne (pp. 47–55 Haynes, C. D. (1991). Use and impact of fire. In (&123–124)). Canberra: Department of the C. D. Haynes, M. G. Ridpath and M. A. J. Environment, Sport and Territories. Williams (Eds.), Monsoonal Australia: Gill, A. M., Groves, R. H. and Noble, I. R. landscape, ecology and man in the northern (1981). Fire and the Australian biota. lowlands. (pp. 61–71). Balkema: Rotterdam. Canberra: Australian Academy of Science. Head, L. (1989). Prehistoric Aboriginal impacts on Australian vegetation: an assessment of the evidence. Australian Geographer 20, 37–46.

90 Biodiveristy and bushfires

Head, L. (1993). Unearthing prehistoric cultural Hopkins, A. J. M., & Smith, G. T. (1996). Fire: landscapes: a view from Australia. effects and management implications. In A. Transactions of the Institute of British J. M. Hopkins & G. T. Smith (Eds.), The Geography 18, 481–499. natural history of Two Peoples Bay Nature Head, L., O’Neill, A.L., Martick, J.K. and Reserve Perth: Department of Conservation Fullager, R.L.K. (1992). A comparison of and Land Management (CALMScience Aboriginal and pastoral fires in the north- Supplement no. 5). west Northern Territory. In I. Moffatt and A. How, R. A., Dell, J., & Humphreys, W. F. (1987). Webb (Eds.), Conservation and development The ground vertebrate fauna of coastal areas issues in northern Australia. Darwin: North between and Albany, Western Australia Research Unit. Australia. Records of the Western Australian Hemsley, J. H. (1967). Bushfire – S.E. Tasmania 7th Museum 13, 553–574. February, 1967. Some aspects of the fire in Hughes, P., & Hughes, B. (1991). Notes on the relation to and plant life. Tasmanian Black-breasted Button-quail at Widgee, Forestry Commission. Queensland. Australian Bird Watcher 14, Hewish, M. (1983). The effect of a wildfire on 113–118. birdlife in a eucalypt forest: a preliminary Hull, A. F. B. (1922). A visit to the Archipelago of report on the Lerdederg Gorge seven weeks the Recherche S.W. Australia. Emu 21, after the Wombat State Forest fire. Geelong 277–289. Naturalist 20, 3–16. Hutchins, B. R. (1988). Black Honeyeaters Higgs, P. and Fox, B.J. (1993). Interspecific feeding among charcoal and ash. South competition: a mechanism for rodent Australian Ornithologist 30, 160. succession after fire in wet heathland. Inions, G. (1985). The interactions between Australian Journal of Ecology 18, 193–201. possums, habitat trees and fire. B.Sc. (Hons) Hill, R.S. (Ed.) (1994). History of Australian thesis. Australian National University, vegetation: to Recent. Cambridge: Canberra. Cambridge University Press. Inions, G.B., Tanton, M.T. and Davey, S.M. Hodgson, A., & Heislers, A. (1972). Some aspects of (1989). The effect of fire on the availability the role of forest fire in South-eastern Australia. of hollows in trees used by the Common Forests Commission, Victoria. Brushtail Possum, Trichosurus vulpecula Kerr Holmes, G. (1988). Eastern Bristlebird: summary 1872, and Ringtail Possum, Pseudocheirus conservation statement for northern populations. peregrinus Boddarts 1785. Australian Wildlife QNPWS. Research 16, 449–458. Holmes, G. (1989). Eastern Bristlebird. Species Jackson, W.D. (1968). Fire, air, water, earth - an management plan for northern populations. elemental ecology of Tasmania. Proceedings of Queensland NPWS and NSW NPWS. the Ecological Society of Australia 3, 9–16. Hood, J. B. (1941). Birds and bushfires. South Johnson, R.W. and Purdie, R.W. (1981). The role Australian Ornithologist 15, 125–127. of fire in the establishment and management Hopkins, A.J.M. (1985a). Fire in the woodlands of agricultural systems. In A. M. Gill, R. H. and associated formations of the semi-arid Groves, & I. R. Noble (Eds.), Fire and the region of south-western Australia. In J. R. Australian biota (pp. 497–528). Canberra: Ford (Eds.), Fire ecology and management in Australian Academy of Science. Western Australian ecosystems (pp. 83–90). Jones, R. (1969). Fire-stick farming. Australian Perth: Western Australian Institute of Natural History 16, 224–228. Technology. Jones, R. (1980). Hunters in the Australian coastal Hopkins, A. J. M. (1985b). Planning the use of savanna. In D. R. Harris (Ed.), Human fire on conservation lands in south-western Ecology in Savanna Environments London: Australia. In J. R. Ford (Ed.), Fire ecology and Academic Press. management in Western Australian ecosystems Jones, R. (1995). Mindjongrok: legacy of the (pp. 203–208). Perth: Western Australian firestick. In D.B. Rose (Ed.) Country in Institute of Technology. flames. Proceedings of the 1994 symposium on biodiversity and fire in North Australia. (pp. 11–18). Canberra: Department of the Environment, Sport and Territories.

91 Australia’s Biodiveristy - Responses to Fire

Jordan, R. (1984a). The Eastern Bristlebird. Leigh, J., Boden, R. and Briggs, J. (1984). Extinct Effects of fire on a population. RAOU Report and endangered plants of Australia. Melbourne: 11, 30. MacMillan. Jordan, R. (1984b). The Ground Parrot – Effect Lewis, H.T. (1985). Burning the “Top End”: of fire on a population. RAOU Report 11, and cattle. In J.R. Ford (Ed.). Fire 28–29. ecology and management in Western Australian Jordan, R. (1987a). The Barren Grounds rolling ecosystems (pp. 21–31). Perth: Western bird survey. RAOU Report 27, 15–18. Australian Institute of Technology. Jordan, R. (1987b). The Ground Parrot in Barren Lewis, H. T. (1989). Ecological and technological Grounds Nature Reserve. RAOU Report 27, knowledge of fire: Aborigines versus park 19–23. rangers in northern Australia. American Jordan, R. (1987c). The Southern Emu–wren in Anthropologist 91, 940–961. Barren Grounds. RAOU Report 27, 24. Lindenmayer, D.B., Norton, T.W., and Tanton Jordan, R. (1988). Population changes of some M.T. (1990). Differences between wildfire common insectivore species in woodland and clearfelling on the structure of montane near the wardens house. RAOU Report 51, ash forests of Victoria and their implications 16–18. for fauna dependant on tree hollows. Joseph, L. (1982). The Glossy Black–cockatoo on Australian Forestry 53, 61–68. Kangaroo Island. Emu 82, 46–49. Lindenmayer, D.B., Cunningham, R.B., Donnelly, Kavanagh, R. P. (1990). Survey of Powerful and C.F., Tanton, M.T. and Nix, H.A. (1993). Sooty Owls in south–eastern New South Wales. The abundance and development of cavities Final report (Project 120). World Wildlife in Eucalyptus trees: a case study in the Fund (Australia). montane forests of Victoria, southeastern Kavanagh, R. P. (1991). The target species Australia. Forest Ecology and Management 60, approach to wildlife management: gliders 77–104. and owls in the forests of southeastern New Lord, E. A. R. (1936). Notes on swifts. Emu 35, South Wales. In D. Lunney (Ed.), 216–218. Conservation of Australia’s forest fauna (pp. Loyn, R. H. (1985a). Bird populations in 377–383). Mosman: Royal Zoological successional forests of Mountain Ash Society of NSW. Eucalyptus regnans in central Victoria. Emu Kershaw, A.P. (1986). Climatic change and 85, 213–230. Aboriginal burning in north-east Australia Loyn, R. H. (1985b). Ecology, distribution and during the last two glacials. Nature 322, density of birds in Victorian forests. In A. 47–49. Keast, H. F. Recher, H. Ford, & D. Saunders Kimber, P. C. (1974). Some effects of prescribed (Eds.), Birds of eucalypt forests and woodlands: burning on Jarrah Forest birds. In Third Fire ecology, conservation, management (pp. 33–46). Ecology Symposium (pp. 49–57). Melbourne: Chipping Norton: Surrey Beatty. Forests Commission, Victoria. Loyn, R. H., Traill, B. J., & Triggs, B. E. (1986). Kimber, R.G. (1982). Black lightning: Aborigines Prey of Sooty Owls in East Gippsland before and fire in central Australia and the Western and after fire. Victorian Naturalist 103, desert. Archaeology in Oceania 18, 38–45. 147–149. Lane, S. G. (1976). Seabird islands: No. 18. Loyn, R. H., Cameron, D. G., Traill, B. J., Sloan, Broughton Island, New South Wales. J. F., Malone, B. S., Schulz, M., Earl, G. E., Australian Bird Bander 14, 10–13. & Triggs, B. E. (1992a). Flora and fauna of the Latz, P. (1995). Fire in the desert: increasing Cooaggalah Forest Block, East Gippsland, biodiversity in the short-term, decreasing it Victoria. Ecological Survey Report 20. in the long term. In D.B. Rose (Ed.) Country Department of Conservation and in flames. Proceedings of the 1994 symposium on Environment. biodiversity and fire in North Australia. (pp. 77–86). Canberra: Department of the Environment, Sport and Territories.

92 Biodiveristy and bushfires

Loyn, R. H., Hewish, M. J., & Considine, M. Masters, P. (1993). The effects of fire-driven (1992b). Short-term effects of fuel reduction succession and rainfall on small mammals in burning on bird populations in Wombat spinifex grassland at Uluru National Park, State Forest. In K. Tolhurst & D. Flinn Northern Territory. Wildlife Research 20, (Eds.), Ecological impact of fuel reduction 803–813. burning in dry sclerophyll forest: first progress Matthew, J. (1994). The status, distribution and report (pp. 5.1–5.11). Melbourne: habitat of the Slender-billed Thornbill Department of Conservation and Acanthiza iredalei in South Australia. South Environment (Research Report no. 349). Australian Ornithologist 32, 1–19. Lucas, D., & Russell-Smith, J. (1993). Traditional Mawson, P. R., & Long, J. L. (1994). Size and age resources of the South Alligator floodplain: parameters of nest trees used by four species utilisation and management. Australian Nature of parrot and one species of cockatoo in Conservation Agency. south-west Australia. Emu 94, 149–155. Lucas, K. and Lucas, D. (1993). Aboriginal fire May, S. A. (1994). Sustaining key old-growth management of the Woolwonga wetlands in characteristics in native forests used for Kakadu National Park. Report to Australian wood production: the significance of exotic Nature Conservation Agency. vertebrate predators. In T.W. Norton and Luke, R. H. and McArthur, A. G. (1978). Bushfires S. R. Dovers (Eds.), Ecology and sustainability in Australia. Canberra: Australian of southern temperate ecosystems (pp. 51–58). Government Publishing Service. Canberra: CSIRO. Lunt, I.D. (1991). Management of remnant McCaw, W. L., Burrows, N. D., Friend, G. R., & lowland grasslands and grassy woodlands for Gill, A. M. (1995). Landscape fires ’93: nature conservation: a review. Victorian proceedings of an Australian bushfire Naturalist 108, 56–66. conference, Perth, Western Australia, 27–29 MacGillivray, W. D. K. (1910). Along the Great September 1993. CALMScience Supplement 4 Barrier Reef. Emu 10, 216–233. McCulloch, E. M. (1966). Swifts and bushfires. Mackowski, C. M. (1984). The ontogeny of Emu 65, 290. hollows in Blackbutt, Eucalyptus nilularis and McDougall, K.L. (1989). The re-establishment of its relevance to the management of forests Themeda triandra (Kangaroo Grass): for possums, gliders and timber. In A.P. implications for the restoration of grassland. Smith and I. D. Hume (Eds ). (pp. 553–567) Melbourne: Department of Conservation, Chipping Norton: Surrey Beatty. Forests and Lands. Main, A. R. (1981). Fire tolerance of heathland McFarland, D. C. (1988). The composition, animals. In R. L. Specht (Ed.), Ecosystems of microhabitat use and response to fire of the the World. 9B. Heathlands and related avifauna of subtropical heathlands in Coloola shrublands. Analytical studies. Amsterdam: National Park, Queensland. Emu 88, Elsevier. 249–257. Majer, J. D. (1984). Short-term responses of soil McFarland, D. C. (1989). The Ground Parrot and litter invertebrates to a cool autumn Pezoporus wallicus (Kerr) in Queensland: burn in jarrah () forest habitat, biology and conservation. Department in Western Australia. Pedobiologia 26, of Conservation, Parks & Wildlife, 299–247. Queensland. Majer, J. D. (1985). Fire effects on invertebrate McFarland, D. C. (1991). The biology of the fauna of forest and woodland. In J. R. Ford Ground Parrot, Pezoporus wallicus, in (Ed.), Fire ecology and management in Western Queensland. III. Distribution and Australian ecosystems (pp. 103–106). Perth: abundance. Wildlife Research 18, 199–213. Western Australian Institute of Technology. McFarland, D. C. (1992). Fire and the Marchant, S. (1985). Breeding of the Eastern management of ground parrot habitat. In Yellow Robin. In A. Keast, H. F. Recher, H. B. R. Roberts (Ed.), Fire Research in rural Ford, & D. Saunders (Eds.), Birds of eucalypt Queensland (pp. 483–495). Toowomba: forests and woodlands: ecology, conservation, University of Southern Queensland. management (pp. 231–240). Chipping Norton: Surrey Beatty.

93 Australia’s Biodiveristy - Responses to Fire

McFarland, D. C. (1993). Fire and bird Meredith, C. W. (1983). Fire and birds. The conservation. In C. P. Catterall, P. V. result of two studies and their relevance to Driscoll, K. Hulsman, D. Muir, & A. Taplin fuel reduction burning. In E. H. M. Ealey (Eds.), Birds and their habitats: status and (Eds.), Fighting fire with fire (pp. 193–202). conservation in Queensland (pp. 41–44). Melbourne: Graduate School of St Lucia: Queensland Ornithological Society Environmental Science Monash University. Inc. Meredith, C. W. (1984a). The Ground Parrot. McFarland, D. C. (1994). Notes on the Brush RAOU Conservation Statement, 1. Bronzewing elegans and Southern Meredith, C. W. (1984b). Management of the Emu-wren Stipiturus malachurus in Cooloola Ground Parrot Pezoporus wallicus in Victoria. National Park. Sunbird 24, 14–17. Fisheries & Wildlife Division, Victoria. McIlroy, J.C. (1978). The effects of forestry Meredith, C. W. (1984c). Recent records of the practices on wildlife in Australia: a review. Ground Parrot Pezoporus wallicus in western Australian Forestry 41, 78–94. Victoria. Geelong Naturalist 21, 3–4. McKean, J. L., & Martin, K. C. (1985). Meredith, C. W. (1988). Fire in the Victorian Distribution and status of the Carpentarian environemnt - a discussion paper. Conservation Grasswren Amytornis dorotheae. Conservation Council of Victoria. Commission of the Northern Territory. Meredith, C. W., & Isles, A. C. (1980). A study of McKean, J. L., & Martin, K. C. (1989). the Ground Parrot (Pezoporus wallicus) in Distribution and status of the Carpentarian Victoria. report No. 304. Environmental Grass-wren Amytornis dorotheae. Northern Studies Division of the Ministry for Territory Naturalist 11, 12–19. Conservation, Victoria. McKenzie, N. L. and Belbin, L. (1991). Meredith, C. W., & Jaremovic, R. (1990). Current Kimberley rainforest communities: reserve status and management of the Ground Parrot in recommendations and management Victoria. Arthur Rylah Institute for considerations. In N. L. McKenzie, R. B. Environmental Research. Johnson and P. G. Kendrick (Eds.) Kimberley Meredith, C. W., Gilmore, A. M., & Isles, A. C. rainforests of Australia. (pp. 453–468). (1984). The Ground Parrot (Pezoporus Chipping Norton: Surrey Beatty. wallicus Kerr) in south-eastern Australia: a McLaughlin, J. (1992). The floristic and structural fire-adapted species? Australian Journal of features of Black-eared Miner Manorina Ecology 9, 367–380. melanotis habitat. RAOU Report No. 84. Royal Merilees, D. (1968). Man the destroyer: late Australasian Ornithologists Union. Quaternary changes in the Australian McLaughlin, J. (1994). Searches for the Black-eared fauna. Journal of the Royal Society of Miner Manorina melanotis in the Victorian Western Australia 51, 1–24. Murray Mallee. RAOU Report No. 93. Royal Milledge, D. R., & Palmer, C. L. (1990). The Sooty Australasian Ornithologists Union. Owl in Mountain Ash forests in the Victorian McNee, S. (1986). Surveys of the Western Whipbird Central Highlands. Department of and Western Bristlebird in Western Australia. Conservation and Environment. RAOU Report No. 18. Royal Australasian Milledge, D. R., Palmer, C. L., & Nelson, J. L. Ornithologists Union. (1991). “Barometers of change”: the Menkhorst, P. W., & Bennett, A. F. (1990). distribution of large owls and gliders in Vertebrate fauna of mallee vegetation in Mountain Ash forests of teh Victorian southern Victoria. In J. C. Noble, P. J. Joss, Central Highlands and their potential as & G. K. Jones (Eds.), The mallee lands: a management indicators. In D. Lunney (Ed.), conservation perspective (pp. 39–53). Conservation of Australia’s Forest Fauna (pp. Melbourne: CSIRO. 53–65). Sydney: Royal Zoological Society of Meredith, C. W. (1982). The research input to NSW. fire management in natural areas with Milligan, A. W. (1904). Notes on a trip to the particular reference to wildlife research in Wongan Hills, Western Australia. Emu 4, the semi-arid lands of Victoria and New 2–11. South Wales. In A. Heislers, P. Lynch, & B. Walters (Eds.), Fire ecology in semi-arid lands (pp. 10pp). Deniliquin: CSIRO.

94 Biodiveristy and bushfires

Mitchell, T.M. (1838). Three expeditions into the Paton, J. B., & Paton, D. C. (1977). Seabird interior of eastern Australia, with descriptions of Islands: No. 52. Wright Island, South recently explored regions of Australia Felix, and Australia. Corella 1, 68-69. the present colony of New South Wales London: Pedler, L. (1991). Rare bird survey: implications for T. & W. Boone. fire management for Uluru National Park. Nelson, J. L., & Morris, B. J. (1994). Nesting Australian National Parks and Wildlife requirements of the Yellow-tailed Service. Black-cockatoo, Calyptorhynchus funereus, in Pedler, L., & Burbidge, A. H. (1995). The range Eucalyptus regnans forest, and implications and status of the Nullabor Quail-thrush. for forest management. Wildlife Research 21, South Australian Ornithologist 32, 45–52. 267–278. Pescott, T. W. (1976). Seabird Islands: No. 27. Newsome, A. E., McIlroy, J., & Catling, P. (1975). Lady Julia Percy Island, Victoria. Australian The effects of extensive wildfire on Bird Bander 14, 29–31. populations of twenty ground vertebrates in Pescott, T. (1983). Beach-washed birds after the south-east Australia. Proceedings of the Ash Wednesday fire. Geelong Naturalist 20, Ecological Society of Australia 9, 107–123. 17–19. Nichols, O. G., & Muir, B. (1989). Vertebrates of Pescott, T. (1985). Black Honeyeaters and the jarrah forest. In B. Dell, J. J. Havel, & N. charcoal. Geelong Naturalist 22, 37–39. Malajczuk (Eds.), The Jarrah Forest (pp. Porter, J. W., & Henderson, R. (1983). Birds and 133–153). Dordrecht: Kluwer. burning histories of open forest at Gundiah, Nicholson, P. H. (1981). Fire and the Australian southeastern Queensland. Sunbird 13, 61–69. Aboriginal - an enigma. In A. M. Gill, R. H. Press, A. J. (1987). Fire management in Kakadu Groves, & I. R. Noble (Eds.), Fire and the National Park: the ecological basis for the Australian biota (pp. 55–76). Canberra: active use of fire. Search 18, 244–248. Australian Academy of Science. Press, A. J. (1988). Comparisons of the extent of Nix, H.A. (1993). Bird distributions in relation to fire in different land management systems in imperatives for habitat conservation in the Top End of the Northern Territory. Queensland. In C. P. Catterall, P. V. Driscoll, Proceedings of the Ecological Society of Australia K. Hulsman, D. Muir, & A. Taplin (Eds.), 15, 167–175. Birds and their habitats: status and conservation Pressland, A. J., Scanlan, J. C. and McLennan, S. in Queensland (pp. 12–21). St Lucia: (1992). The role of fire in the grazing lands Queensland Ornithological Society Inc. of Queensland. In B.R. Roberts (Ed.) Fire Noble, J.C., Harrington, G.N. and Hodgkinson, research in rural Queensland. (pp. 237–250). K.C. (1984). The ecological significance of Toowoomba: University of Southern irregular fire in Australian rangelands. In Queensland. P. J. Joss, P. W. Lynch and O. B. Williams Price, O. and Bowman, D.M.J.S. (1994). Fire- (Eds.), Rangelands: a resource under siege. stick forestry: a matrix model in support of (pp.577–580) Canberra: Australian Academy skilful fire management of Callitris of Science. intratropica R.T. Baker. Journal of Noske, R. (1988). The status and biology of the Biogeography 21, 573–580. White-throated Grass-wren. Australian Priddel, D. (1989). Conservation of rare fauna: National Parks & Wildlife Service. the Regent Parrot and the Malleefowl. In J. Noske, R. A. (1992a). Do Grasswrens have the C. Noble & R. A. Bradstock (Eds.), numbers? Reply to Woinarski (1992). Mediterranean landscapes in Australia: mallee Northern Territory Naturalist 13, 5–8. ecosystems and their management (pp. Noske, R. A. (1992b). The status and ecology of 243–249). Melbourne: CSIRO. the White-throated Grasswren Amytornis Priddel, D. (1990). Conservation of the woodwardi. Emu 92, 39–51. Malleefowl in New South Wales: an Orr, D.M. and Holmes, W.E. (1984). Mitchell experimental management study. In J. C. grasslands. In G.N. Harrington, A.D. Noble, P. J. Joss, & G. K. Jones (Eds.), The Wilson and M.D. Young (Eds.) Management mallee lands: a conservation perspective (pp. of Australia’s rangelands. Melbourne: CSIRO. 71–74). Melbourne: CSIRO.

95 Australia’s Biodiveristy - Responses to Fire

Prober, S.M. and Thiele, K.R. (1993). The Recher, H. F., & Serventy, D. L. (1991). Long ecology and genetics of remnant grassy term changes in the relative abundance of White Box woodlands in relation to their birds in Kings Park, Perth, Western conservation. Victorian Naturalist 110, Australia. Conservation Biology 5, 90–102. 30–36. Recher, H. F., Lunney, D., & Posamentier, H. Pyke, G.H., Saillard, R. and Smith, J. (1995). (1975). A grand natural experiment – the Abundance of Eastern Bristlebirds in Nadgee wildfire. Australian Natural History 18, relation to habitat and fire history. Emu 95, 150–163. 106–110. Recher, H. F., Allen, D., & Gowing, G. (1985). Pyne, S.J. (1991). Burning bush: a fire history of The impact of wildfire on birds in an Australia. New York: Henry Holt & intensively logged forest. In A. Keast, H. F. Company. Recher, H. Ford, & D. Saunders (Eds.), Birds Quin, B. R., & Baker-Gabb, D. (1993). of eucalypt forests and woodlands: ecology, Conservation and management of the Turquoise conservation, management (pp. 283–290). Parrot Neophema pulchella in north-east Chipping Norton: Surrey Beatty. Victoria. Department of Conservation and Recher, H. F., Davis, W. E., & Holmes, R. T. Environment. (1987a). Ecology of Brown and Striated Ratkowsky, A. V. (1978). The effect of a spring Thornbills in forests of south-eastern New fire on the number of bird species. South Wales, with comments on forest Tasmanian Naturalist (53), 11–12. management. Emu 87, 1–13. Ratkowsky, A. V. (1979). The bird species of Mt. Recher, H. F., Shields, J., Kavanagh, R., & Webb, Nelson in relation to microhabitat and G. (1987b). Retaining remnant mature forest recent bushfires. Tasmanian Naturalist (57), for nature conservation at Eden, New South 12–18. Wales: a review of theory and practice. In D. Ratkowsky, A. V. (1985). The effect of a severe fire A. Saunders, G. W. Arnold, A. A. Burbidge, on the number of bird species in a wet & A. J. M. Hopkins (Eds.), Nature sclerophyll environment. Tasmanian conservation: the role of remnants of native Naturalist, 8–9. vegetation (pp. 177–194). Chipping Norton: Recher, H.F. (1969). Bird species diversity and Surrey Beatty. habitat diversity in Australia and North Recher, H. F., Hutchings, P. A. and Rosen, S. America. American Naturalist 103, 75–80. (1993). The biota of the Hawkesbury- Recher, H. F. (1981). Death of an Australian Nepean catchment: reconstruction and myth: fire and its effects on wildlife. In P. J. restoration. Australian Zoologist 29, 3–41. Stanbury (Ed.), Bushfires: their effect on Reid, J., & Fleming, M. (1992). The conservation Australian life and landscape. (pp. 39–48). status of birds in arid Australia. Rangelands Sydney: Macleay Museum, University of Journal 14, 65–91. Sydney. Reid, J. R. W., Kerle, J. A., & Morton, S. R. Recher, H. F. (1991). The conservation and (1993a). Birds. In J. R. W. Reid, J. A. Kerle, management of eucalypt forest birds: & S. R. Morton (Eds.), Uluru fauna: the resource requirements for nesting and distribution and abundance of vertebrate fauna foraging. In D. Lunney (Ed.), Conservation of of Uluru (Ayers Rock – Mount Olga) National Australia’s forest fauna. (pp. 25–34). Sydney: Park, N.T. (pp. 36–57). Canberra: Australian Royal Zoological Society of NSW. National Parks and Wildlife Service. Recher, H. F., & Christensen, P. E. (1981). Fire Reid, J. R. W., Kerle, J. A., & Morton, S. R. and the evolution of the Australian biota. In (1993b). Ecological patterns and processes of A. Keast (Ed.), Ecological biogeography in importance. In J. R. W. Reid, J. A. Kerle, & Australia (pp. 135–162). The Hague: Junk. S. R. Morton (Eds.), Uluru fauna: the Recher, H. F., & Lim, L. (1990). A review of distribution and abundance of vertebrate fauna current ideas of the extinction, conservation of Uluru (Ayers Rock – Mount Olga) National and management of Australia’s terrestrial Park, N.T. (pp. 133–148). Canberra: vertebrate fauna. Proceedings of the Ecological Australian National Parks and Wildlife Society of Australia 16, 287–301. Service.

96 Biodiveristy and bushfires

Reilly, P. (1991a). The effect of wildfire on bird Chipping Norton: Surrey Beatty. populations in a Victorian coastal habitat. Rowley, I., & Russell, E. (1993). The Emu 91, 100–106. Purple-crowned Fairy-wren Reilly, P. (1991b). The effect of wildfire on bush coronatus. II. Breeding biology, social bird populations in six Victorian coastal organisation, demography and management. habitats. Corella 15, 134–142. Emu 93, 235–251. Ridpath, M. G. (1972). The effects of fire on Rowley, I., Russell, E., Brown, R., & Brown, M. fauna. In R. J. Hooper & M. Rowell (Eds.), (1988). The ecology and breeding biology of Proceedings of the 1971 Tropical and Arid Fire the Red-winged Fairy-wren Malurus elegans. Symposium. (pp. 64–66). Darwin: NT Emu 88, 161–176. Government Printer. Russell, E. M., & Rowley, I. (1993). Demography Roberts, P. E. (1970). Some effects of a bushfire of the cooperatively breeding Splendid on heathland birdlife. Proceedings of the Royal Fairy-wren, Malurus splendens (Maluridae). Zoological Society of New South Wales 89, Australian Journal of Zoology 41, 475–505. 40–43. Russell-Smith, J. (1985). A record of change: Robinson, A. C., Casperson, K. D., & Copley, P. studies of vegetation history in the B. (1990). Breeding records of the South Alligator Region, Northern Territory. Malleefowl (Leipoa ocellata) and Proceedings of the Ecological Society of Australia Scarlet-chested Parrot (Neophema splendida) 13, 191–202. within the Yellabinna Wilderness area, South Russell-Smith, J. and Bowman, D.M.J.S. (1992). Australia. South Australian Ornithologist 31, Conservation of monsoon rainforest isolates 8–12. in the Northern Territory, Australia. Robinson, D. (1991). Threatened birds in Biological Conservation 59, 51–63. Victoria: their distributions, ecology and Russell-Smith, J. and Dunlop, C. (1987). The future. Victorian Naturalist 108, 67–77. status of monsoon vine forests in the Robinson, D. (1994). Research plan for Northern Territory. In The rainforest legacy threatened woodland birds of southeastern (pp.227–288). Canberra: Australian Australia. Melbourne: Department of Government Publishing Service. Conservation and Natural Resources. Russell-Smith, J. and Ryan, P. (1994). Long-term (Technical Report series no. 133). monitoring of the effects of management-imposed Robinson, D., & Woinarski, J. C. Z. (1992). A fire regimes on old-growth vegetation in Kakadu review of records of the Northern Shrike-tit National Park: fire history 1980–1993. Report Falcunculus frontatus whitei in northwestern to Department of Environment, Sport and Australia. South Australian Ornithologist 31, Territories. 111–117. Saunders, D. A. (1985). Human impact: the Rounsevell, D. E., & Woinarski, J. C. Z. (1983). response of forest and woodland bird Status and conservation of the Forty-spotted communities. Whither the future? A Pardalote, Pardalotus quadragintus (Aves: synthesis. In A. Keast, H. F. Recher, H. Pardalotidae). Australian Wildlife Research 10, Ford, & D. Saunders (Eds.), Birds of eucalypt 343–349. forests and woodlands: ecology, conservation, Rowley, I. (1987). Conservation of the Purple-crowned management (pp. 355–357). Chipping Fairy-wren Malurus coronatus in northern Norton: Surrey Beatty. Australia. World Wildlife Fund (Australia). Saunders, D.A. (1989). Changes in the avifauna of Rowley, I. (1993). The Purple-crowned a region, district and remnant as a result of Fairy-wren Malurus coronatus. I. History, fragmentation of native vegtetation: the distribution and present status. Emu 93, Wheatbelt of Western Australia. A case 220–234. study. Biological Conservation 50, 99–135. Rowley, I., & Brooker, M. (1987). The response of Saunders, D.A. and Curry, P.J. (1990). The impact a small insectivorous bird to fire in of agricultural and pastoral industries on heathlands. In D. A. Saunders, G. W. birds in the southern half of Western Arnold, A. A. Burbidge, & A. J. M. Hopkins Australia – past, present and future. (Eds.), Nature Conservation: the role of Proceedings of the Ecological Society of Australia remnants of native vegetation (pp. 211–218). 16, 303–321.

97 Australia’s Biodiveristy - Responses to Fire

Saunders, D. A., Smith, G. T., & Rowley, I. Smith, E.L. (1960). Effects of burning and (1982). The availability and dimensions of clipping at various times during the wet tree hollows that provide nest sites for season on tropical tall grass range in cockatoos (Psittaciformes) in Western northern Australia. Journal of Rangeland Australia. Australian Wildlife Research 9, Management 13, 197–203. 541–556. Smith, G. T. (1977). The effect of environmental Scanlan, J.C. (1980). Effects of spring wildfires on change on six rare birds. Emu 77, 173–179. Astrebla (mitchell grass) grasslands in north- Smith, G. T. (1979a). The Noisy Scrub-bird. In west Queensland under varying levels of M. J. Tyler (Ed.), The status of endangered growing season rainfall. Australian Australasian wildlife (pp. 117–121). Adelaide: Rangelands Journal 2, 162–168. Royal Zoological Society of South Australia. Schodde, R. (1982). The Fairy-wrens. A monograph Smith, G. T. (1979b). The status of Australian of the Maluridae. Melbourne: Lansdowne. parrots. In M. J. Tyler (Ed.), The status of Schulz, M. (1991). The Grey-crowned Babbler endangered Australasian wildlife (pp. Pomatostomus temporalis - a cause for concern 101–108). Adelaide: Royal Zoological in Victoria. Australian Bird Watcher 14, Society of South Australia. 37–43. Smith, G. T. (1985a). Fire effects on populations Scotts, D. J. (1991). Old-growth forests: their of the Noisy Scrub–bird (Atrichornis ecological characteristics and value to clamosus), Western Bristle-bird (Dasyornis forest-dependent vertebrate fauna of longirostris) and Western Whip-bird south-east Australia. In D. Lunney (Ed.), (Psophodes nigrogularis). In J. R. Ford (Ed.), Conservation of Australia’s forest fauna (pp. Fire ecology and management in Western 147-159). Sydney: Royal Zoological Society Australian ecosystems (pp. 95–102). Perth: of NSW. Western Australian Institute of Technology. Silveira, C. E. (1993). Recovery plan for threatened Smith, G. T. (1985b). The Noisy Scrub-bird mallee birds - addressing fire regimes. Royal Atrichornis clamosus. Does it’s past suggest a Australasian Ornithologists Union. future? In A. Keast, H. F. Recher, H. Ford, & Silveira, C. E. (1995). The Black–eared Miner. D. Saunders (Eds.), Birds of eucalypt forests and Australian Bird Watcher 16, 96–109. woodlands: ecology, conservation, management (pp. Singh, G., Kershaw, A.P. and Clark, R. (1981). 301–308). Chipping Norton: Surrey Beatty. Quaternary vegetation and fire history in Smith, G. T. (1985c). Population and habitat Australia. In A. M. Gill, R. H. Groves, & I. selection of the Noisy Scrub-bird, Atrichornis R. Noble (Eds.), Fire and the Australian biota clamosus, 1962–83. Australian Wildlife (pp. 23–54). Canberra: Australian Academy Research 12, 479–485. of Science. Smith, G. T. (1987a). The changing environment Singh, G. and Geissler, E.A. (1985). Late for birds in the south-west of Western Cainozoic history of vegetation, fire, lake Australia; some management implications. In levels and climate at Lake George, New D. A. Saunders, G. W. Arnold, A. A. South Wales, Australia. Philosophical Burbidge, & A. J. M. Hopkins (Eds.), Nature Transactions of the Royal Society, London. B311, Conservation: The role of remnants of native 379–447. vegetation (pp. 269–277). Chipping Norton: Skira, I. J., & Brothers, N. P. (1988a). Seabird Surrey Beatty. Islands: No. 183. Little , Smith, G. T. (1987b). Observations on the biology , Tasmania. Corella 12, of the Western Bristlebird Dasyornis 80–81. longirostris. Emu 87, 111–118. Skira, I. J., & Brothers, N. P. (1988b). Seabird Smith, G. T. (1991). Ecology of the Western Islands: No. 184. Great Dog Island, Furneaux Whipbird Psophodes nigrogularis in Western Group, Tasmania. Corella 12, 82–84. Australia. Emu 91, 145–157. Smith, A. P., and Lindenmayer, D. (1988). Tree Smith, G. T., & Forrester, R. I. (1981). The status hollow requirements of Leadbeater’s Possum of the Noisy Scrub-bird Atrichornis clamosus. and other possums and gliders in timber Biological Conservation 19, 239–254. production ash forests of the Victorian Smith, G. T., & Robinson, F. N. (1976). The central highlands. Australian Wildlife Research Noisy Scrub-bird: an interim report. Emu 15, 347–362. 76, 37–42.

98 Biodiveristy and bushfires

Smith, L. H. (1994). A critical analysis of the Suckling, G. C., & MacFarlane, M. A. (1983). factors responsible for the decline of the The effects of fire on fauna – a review. In E. Superb Menura novaehollandiae in Ealey (Ed.), Fighting fire with fire. (pp. Sherbrooke Forest, Victoria. Australian Bird 107–128). Melbourne: Monash University. Watcher 15, 238–249. Suijdendorp, H. (1981). Responses of the Smith, P. (1989). Changes in a forest bird hummock grasslands of northwestern community during a period of fire and Australia to fire. In A. M. Gill, R. H. Groves, drought near Bega, New South Wales. & I. R. Noble (Eds.), Fire and the Australian Australian Journal of Ecology 14, 41–54. biota (pp. 417–424). Canberra: Australian Smith, P. J., Pressey, R. L., & Smith, J. E. (1994). Academy of Science. Birds of particular conservation concern in Sullivan, M., Cowan, D. and Pressland, A. (1992). the Western Division of New South Wales. The effects of controlled burns in the mulga Biological Conservation 69, 315–338. lands of southwest Queensland. In B.R. Smith, P. J., Smith, J. E., & Pressey, R. L. (1996). Roberts (Ed.) Fire research in rural Birds of particular conservation concern in the Queensland. (pp. 300–307). Toowoomba: Western Division of New South Wales: University of Southern Queensland. distribution, habitats and threats. NSW Tarr, H. E. (1965). The Mallee-Fowl in National Parks and Wildlife Service. Wyperfeld National Park. Australian Bird Specht, R. L. (1981). Responses to fires in Watcher 2, 140–144. heathlands and related shrublands. In A. M. Taylor, R. J. (1991). Fauna conservation in Gill, R. H. Groves, & I. R. Noble (Eds.), production forests in Tasmania. Hobart: Fire and the Australian biota (pp. 395–415). Forestry Commission Tasmania. Canberra: Australian Academy of Science. Thackway, R. and Creswell, I.D. (1995). An Springett, J.A. (1976). The effect of prescribed interim biogeographic regionalisation for burning on the soil fauna and on litter Australia. Canberra: Australian Nature decomposition in Western Australian forests. Conservation Agency. Australian Journal of Ecology 1, 72–82. Tidemann, S. C. (1990). Relationships between Stanton, J. P. (1992). J.P. Thomson oration. The finches and pastoral practices in northern neglected lands: recent changes in the Australia. In J. Pinowski & J. D. ecosystems of Cape York Peninsula and the Summers–Smith (Eds.), Granivorous birds and challenge of their management. Journal of the agriculture (pp. 305–315). Warsaw: Queensland Geographical Society 7, 1–18. PWN-Polish Scientific Publishers. Stanton, P. (1995). A tropical Queensland Tidemann, S. C. (1992). Conservation of the perspective. In D. B. Rose (Ed.), Country in Gouldian Finch, N.T. World Wide Fund for flames. Proceedings of the 1994 symposium on Nature. biodiversity and fire in North Australia (pp. Tidemann, S. C. (1993a). Management of a 71–76). Canberra: Department of the threatened species: the Gouldian Finch Environment, Sport and Territories, and example. In C. P. Catterall, P. V. Driscoll, K. North Australia Research Unit. Hulsman, D. Muir, & A. Taplin (Eds.), Birds Starks, J. (1987). The status and distribution of the and their habitats: status and conservation in Black-eared Miner (Manorina melanotis) in Queensland (pp. 123–131). St Lucia: Victoria. Technical report no.49. Arthur Rylah Queensland Ornithological Society Inc. Insitute of Environmental Research. Tidemann, S. C. (1993b). Where are Gouldian Stephens, S. (1992). Endangered species and Finches after the breeding season? Victorian communities and threatening processes in the Naturalist 110, 238–243. Murray Mallee. Australian National Parks Tidemann, S.C., Boyden, J., Elvish, R., Elvish, J. and Wildlife Service. and O’Gorman, B. (1992). Characteristics of Stocker, G. C. (1971). The age of charcoal from nest sites and breeding habitat of the old jungle fowl nests and vegetation change endangered Gouldian Finch Erythrura on Melville Island. Search 2, 28–30. gouldiae and abundant, co-occurring Long- Stokes, T. (1975). The effect of a bushfire on the tailed Finch Poephila acuticauda at two sites in banding of Flame Robins in the Brindabella the Northern Territory. Journal of Tropical Ranges. Australian Bird Bander 13, 75–76. Ecology 8, 373–388.

99 Australia’s Biodiveristy - Responses to Fire

Tidemann, S. C., McArtney, J., & Smith, I. Walker, T. A., & Hulsman, K. (1993). Seabird (1993). Queensland Gouldian Finches Islands: No. 221. Wilson Island, Great Barrier Erythrura gouldiae and air-sac mite Reef, Queensland. Corella 17, 155–157. Sternostoma tracheacolum. Sunbird 23, 36–40. Wardell-Johnson, G., & Christensen, P. (1992). A Tingay, A., & Tingay, S. R. (1984). Bird communities review of the effects of disturbance on in the of Western Australia. wildlife of the karri forest. In Research on the Australian Conservation Foundation. impact of forest management in south-west Tolhurst, K. (1996). Effects of fuel reduction Western Australia (pp. 33–57). Perth: burning on fauna in a dry sclerophyll forest. Department of Conservation and Land In DEST (Ed.), Fire and biodiversity: the Management (Occasional Paper 2/92). effects and effectiveness of fire management. Wardell-Johnson, G., & Nichols, O. (1991). Proceedings of the conference held 8–9 October Forest wildlife and habitat management in 1994, Footscray, Melbourne (pp. 113–121 southwestern Australia: knowledge, research (&129)). Canberra: Department of the and direction. In D. Lunney (Ed.), Environment, Sports and Territories. Conservation of Australia’s forest fauna (pp. Tolhurst, K. and Flinn, D. (eds.) (1992). Ecological 161–192). Mosman: Royal Zoological impact of fuel reduction burning in dry Society of NSW. sclerophyll forest: first progress report. Wardell-Johnson, G., McCaw, W. L., & Masey, Melbourne: Department of Conservation K. G. (1989). Critical data requirements for and Environment (Research Report no. 349). the effective management of fire on nature Tothill, J.C. (1992). Fire in black speargrass conservation lands in south Western (Heteropogon contortus) pastures – a ten year Australia. In N. Burrows, L. McCaw, & G. comparison of burnt and unburnt Friend (Eds.), Fire management on nature treatments. In B.R. Roberts (Ed.) Fire conservation lands Perth: Department of research in rural Queensland. (pp. 274–279). Conservation and Land Management. Toowoomba: University of Southern Watkins, D. (1985). Report of the R.A.O.U. Ground Queensland. Parrot survey in Western Australia. Royal Traill, B.J. (1993). Forestry, birds, mammals and Australasian Ornithologists Union. management in box and ironbark forests. Webster, R., & Ahern, L. (1992). Management for Victorian Naturalist 110, 11–14. conservation of the Superb Parrot (Polytelis Turner, R. J. (1987). Effect of fire on birds – swainsonii) in New South Wales and Victoria. Weddin Mountain. In Disappearing islands. Department of Conservation and Natural Proceedings of a seminar on conservation and Resources. co-operation in the Central West (pp. 66–86). Wegener, A. (1984a). Animals killed by bushfires Bathurst: National Parks and Wildlife in coastal Victoria, February 1983. Australian Service, NSW. Bird Watcher 10, 248–250. Turner, R. J. (1992). Effect of wildfire on birds at Wegener, A. (1984b). Survey of animal species killed Weddin Mountain, New South Wales. by the Ash Wednesday fires along the Great Corella 16, 65–74. Ocean Road. Geelong Naturalist 21, 13–16. Vickers-Rich, P. (1991). The and Westoby, M., Rice, B., Griffin, G. and Friedel, M. Tertiary history of birds on the Australian (1988). The soil seed banks of Triodia plate. In (Eds.) P. Vickers-Rich, J.M. basedowii in relation to time since fire. Monaghan, R.F. Baird and T.H. Rich Australian Journal of Ecology 13, 161–169. Vertebrate plaeontology of Australasia. (pp. Whelan, R. J. (1995). The ecology of fire. 721–808). Lilydale: Pioneers Design Studios. Cambridge: Cambridge University Press. Wakefield, N.A. (1970). Bushfire frequency and White, G. (1979a). Seabird islands: No. 75. vegetational change in south-eastern Louisa Island, Tasmania. Corella 3, 61–62. Australian forests. Victorian Naturalist 87, White, G. (1979b). Seabird Islands: No. 76. De 152–158. Witt Island, Tasmania. Corella 3, 63–65. Walker, J. and Tothill, J.C. (1992). Fires in the Whitley, G. P. (1944). Fire and Petrels: the grasslands of Queensland. In B.R. Roberts mystery of Mondrain Island. Emu 44, 6–7. (Ed.) Fire research in rural Queensland. (pp. Whitley, G. P. (1971). Field notes on birds by 251–262). Toowoomba: University of Thomas Carter. Western Australian Naturalist Southern Queensland. 12, 41–44.

100 Fire and Australian Birds

Whittell, H. M. (1936). The Bristle-birds of Woinarski, J. C. Z., & Fisher, A. (1995b). Wildlife Western Australia. Emu 35, 197–201. of lancewood (Acacia shirleyi) thickets and Williams, J. E., & Gill, A. M. (1995). The impact of woodlands in northern Australia: 2. fire regimes on native forests in eastern New comparisons with other environments of the South Wales. NSW National Parks and region (Acacia woodlands, Eucalyptus savanna Wildlife Service. woodlands and monsoon rainforests). Williams, J.E., Whelan, R.J. and Gill, A.M. Wildlife Research 22, 413–443. (1994). Fire and environmental Woinarski, J. C. Z., & Tidemann, S. C. (1991). heterogeneity in southern temperate forest The bird fauna of a deciduous woodland in ecosystems: implications for management. the wet–dry tropics of northern Australia. Australian Journal of Botany 42, 125–137. Wildlife Research 18, 479–500. Wilson, R. I. (1981). The woodchip industry and Woinarski, J. C. Z., & Tidemann, S. C. (1992). Tasmanian birds. Tasmanian Bird Report 11, Survivorship and some population 11–14. parameters for the endangered Gouldian Woinarski, J. C. Z. (1987). Notes on the status Finch Erythrura gouldiae and two other finch and ecology of the Red–lored Whistler species at two sites in tropical northern Pachycephala rufogularis. Emu 87, 224–231. Australia. Emu 92, 33–38. Woinarski, J. C. Z. (1989a). Broombush harvesting Woinarski, J. C. Z., Eckert, H. J., & Menkhorst, in southeastern Australia. In J. C. Noble & R. P. W. (1988a). A review of the distribution, A. Bradstock (Eds.), Mediterranean landscapes in habitat and conservation status of the Australia: mallee ecosystems and their management Western Whipbird Psophodes nigrogularis (pp. 362–378). Melbourne: CSIRO. leucogaster in the Murray mallee. South Woinarski, J. C. Z. (1989b). The vertebrate fauna Australian Ornithologist 30, 146–153. of broombush Melaleuca uncinata vegetation Woinarski, J. C. Z., Tidemann, S. C., & Kerin, S. in north-western Victoria, with reference to (1988b). Birds in a tropical mosaic: the effects of broombush harvesting. Australian distribution of bird species in relation to Wildlife Research 16, 217–238. vegetation patterns. Australian Wildlife Woinarski, J. C. Z. (1990). Effects of fire on the Research 15, 171–196. bird communities of tropical woodlands and Woinarski, J. C. Z., Whitehead, P. J., Bowman, D. open forests in northern Australia. M. J .S. & Russell-Smith, J. (1992). Australian Journal of Ecology 15, 1–22. Conservation of mobile species in a variable Woinarski, J. C. Z. (1992). The conservation environment: the problem of reserve design status of the White-throated Grasswren in the Northern Territory, Australia. Global Amytornis woodwardi, an example of Ecology and Biogeography Letters 2, 1–10. problems in status designation. Northern Wooller, R. D., & Brooker, K. S. (1980). The Territory Naturalist 13, 1–5. effects of controlled burning on some birds Woinarski, J. C. Z. (1993a). Australian tropical of the understorey in Karri forest. Emu 80, savannas, their avifauna, conservation status 165–166. and threats. In C. P. Catterall, P. V. Driscoll, Wooller, R. D., & Calver, M. C. (1988). Changes K. Hulsman, D. Muir, & A. Taplin (Eds.), in an assemblage of small birds in the Birds and their habitats: status and conservation understorey of dry sclerophyll forest in in Queensland (pp. 45–63). St Lucia: south-western Australia after fire. Australian Queensland Ornithological Society Inc. Wildlife Research 15, 331–338. Woinarski, J. C. Z. (1993b). A cut-and-paste Yates, C. J., Hobbs, R.J. and Bell, R.W. (1994). community: birds of monsoon rainforests in Landscape-scale disturbance and Kakadu National Park, Northern Territory. regeneration in semi-arid woodlands of Emu 93, 100–120. southwestern Australia. Pacific Conservation Woinarski, J. C. Z., & Fisher, A. (1995a). Wildlife Biology 1, 214–221. of lancewood (Acacia shirleyi) thickets and Young, J. (1991). Dealing with fire in parks and woodlands in northern Australia: 1. variation protected areas. In Tropics under fire: fire in vertebrate species composition across the management on Cape York Peninsula (pp. environmental range occupied by lancewood 43–45). Cairns: Cairns and Far North vegetation in the Northern Territory. Environment Centre. Wildlife Research 22, 379–411.

101

Fire and Australian Birds

7. TABLES

Table 1. Scientific names of birds mentioned in Table 1. continued text (following Christidis & Boles 1994). ORDER CHARADRIIFORMES ORDER STRUTHIONIFORMES Pedionomidae Casuariidae Pedionomus torquatus Plains-wanderer Casuarius casuarius Southern Cassowary Burhinidae Dromaius novaehollandiaeEmu Burhinus grallarius Bush Stone-curlew Dromaius baudinianus Kangaroo Island Emu ORDER COLUMBIFORMES ORDER Megapodiidae Phaps chalcoptera Common Bronzewing Leipoa ocellata Malleefowl Phaps elegans Brush Bronzewing Megapodius reinwardt Orange-footed Scrubfowl Ocyphaps lophotes Crested Pigeon Geophaps smithii Partridge Pigeon ypsilophora Brown Quail Geopelia cuneata Diamond Dove Coturnix chinensis King Quail

ORDER ANSERIFORMES ORDER PSITTACIFORMES Cacatuidae Anaseranatidae Cacatuidae Anseranas semipalmata Magpie Goose ProboscigerProbosciger aterrimusaterrimus PalmPalm Cockatoo Cockatoo CalyptorhynchusCalyptorhynchus banskiibanskii Red-tailedRed-tailed ORDER PELECANIFORMES Black-CockatooBlack-Cockatoo Pelecanidae CalyptorhynchusCalyptorhynchus lathamilathami GlossyGlossy Black-Cockatoo Black-Cockatoo Pelecanus conspicillatus Australian Pelican CalyptorhynchusCalyptorhynchus funereusfunereus Yellow-tailedYellow-tailed Black-CockatooBlack-Cockatoo ORDER CICONIIFORMES Calyptorhynchus baudinii CalyptorhynchusLong-billed Black- baudinii Threskionithidae Long-billedCockatoo Black- Threskiornis spinicollis Straw-necked Ibis Cacatua roseicapilla CockatooGalah CacatuaCacatua tenuirostrisroseicapilla GalahLong-billed Corella ORDER FALCONIFORMES CacatuaCacatua pastinatortenuirostris Long-billedWestern Corella Corella Accipitridae CacatuaCacatua sanguineapastinator WesternLittle Corella Corella Erythrotriorchis radiatus Red Goshawk Cacatua leadbeateri Major Mitchell’s Cacatua sanguinea Little Corella Aquila audax Wedge-tailed Eagle Cockatoo Cacatua leadbeateri Major Mitchell’s Falconidae Nymphicus hollandicus Cockatiel Cockatoo Falco cenchroides Nankeen Kestrel Psittacidae Nymphicus hollandicus Cockatiel Polytelis anthopeplus Regent Parrot Psittacidae ORDER GRUIFORMES Platycercus icterotis Western Rosella Polytelis swainsonii Superb Parrot Rallidae Barnardius zonarius Australian Ringneck Polytelis anthopeplus Regent Parrot Rallus pectoralis Lewin’s Rail Purpureicephalus spurius Red-capped Parrot Platycercus icterotis Western Rosella Gallinula mortierii Tasmanian Native-hen Psephotus chrysopterygius Golden-shouldered Barnardius zonarius Australian Ringneck Otididae Parrot Purpureicephalus spurius Red-capped Parrot Ardeotis australis Australian Bustard Psephotus dissimilis Hooded Parrot Psephotus chrysopterygius Golden-shouldered Psephotus pulcherrimus Paradise Parrot ORDER TURNICIFORMES Parrot Melopsittacus undulatus Budgerigar Turnicidae Psephotus dissimilis Hooded Parrot Neophema chrysostoma Blue-winged Parrot Turnix velox Little Button-quail Psephotus pulcherrimus Paradise Parrot Neophema elegans Elegant Parrot Turnix varia Painted Button-quail Melopsittacus undulatus Budgerigar Neophema chrysogaster Orange-bellied Parrot Turnix melanogaster Black-breasted Button- Neophema chrysostoma Blue-winged Parrot Neophema pulchella Turquoise Parrot quail Neophema elegans Elegant Parrot Neophema splendida Scarlet-chested Parrot Neophema chrysogaster Orange-bellied Parrot Pezoporus wallicus Ground Parrot continued over Neophema pulchella Turquoise Parrot Pezoporus occidentalis Night Parrot

continued over 103 Australia’s Biodiveristy - Responses to Fire

Table 1. continued Table 1. continued Dasyornis broadbenti Rufous Bristlebird ORDER STRIGIFORMES Pycnoptilus floccosus Pilotbird Strigidae Sericornis frontalis White-browed Ninox strenua Powerful Owl Scrubwren Tytonidae Hylacola pyrrhopygia Chestnut-rumped Tyto tenebricosa Sooty Owl Heathwren Tyto novaehollandiae Masked Owl Hylacola cauta Shy Heathwren ORDER APODIFORMES Calamanthus fuliginosus Striated Fieldwren Apodidae Chthonicola sagittata Speckled Warbler Hirundapus caudacutus White-throated Needletail Gerygone fusca Western Gerygone Apus pacificus Fork-tailed Swift Acanthiza pusilla Brown Thornbill Acanthiza apicalis Inland Thornbill ORDER CORACIIFORMES Acanthiza inornata Western Thornbill Halcyonidae Acanthiza reguloides Buff-rumped Thornbill Dacelo novaeguineae Laughing Kookaburra Acanthiza iredalei Slender-billed Thornbill Dacelo leachii Blue-winged Kookaburra Acanthiza chrysorrhoa Yellow-rumped Thornbill Todiramphus pyrrhopygia Red-backed Kingfisher Acanthiza nana Yellow Thornbill Meropidae Acanthiza lineata Striated Thornbill Merops ornatus Rainbow Bee-eater nigricincta Banded Whiteface Meliphagidae ORDER PASSERIFORMES Anthochaera chrysoptera Little Wattlebird Menuridae Philemon corniculatus Noisy Friarbird Menura novaehollandiae Philemon citreogularis Little Friarbird Atrichornithidae Manorina flavigula Yellow-throated Miner Atrichornis rufescens Rufous Scrub-bird Manorina melanotis Black-eared Miner Atrichornis clamosus Noisy Scrub-bird Lichenostomus chrysops Yellow-faced Honeyeater Climacteridae Lichenostomus melanops Yellow-tufted Climacteris picumnus Brown Treecreeper Honeyeater [C.p. melanota Lichenostomus Maluridae melanops cassidix Helmeted Honeyeater Malurus coronatus Purple-crowned Lichenostomus cratitius Purple-gaped Fairy-wren Honeyeater Malurus cyaneus Superb Fairy-wren Lichenostomus plumulus Grey-fronted Honeyeater Malurus splendens Splendid Fairy-wren Lichenostomus penicillatus White-plumed Malurus lamberti Variegated Fairy-wren Honeyeater Malurus elegans Red-winged Fairy-wren lunatus White-naped Honeyeater Malurus leucopterus White-winged Fairy-wren Phylidonyris pyrrhoptera Crescent Honeyeater Malurus melanocephalus Red-backed Fairy-wren Phylidonyris nigra White-cheeked Stipiturus malachurus Southern Emu-wren Honeyeater Stipiturus mallee Mallee Emu-wren Phylidonyris albifrons White-fronted Stipiturus ruficeps Rufous-crowned Honeyeater Emu-wren Phylidonyris melanops Tawny-crowned Amytornis woodwardi White-throated Grasswren Honeyeater Amytornis dorotheae Certhionyx niger Black Honeyeater Amytornis striatus Striated Grasswren Epthianura tricolor Crimson Chat Amytornis textilis Thick-billed Grasswren Petroicidae Pardalotidae multicolor Scarlet Robin Pardalotus punctatus Spotted Pardalote Petroica goodenovii Red-capped Robin Pardalotus quadragintus Forty-spotted Pardalote Petroica phoenicea Flame Robin Pardalotus striatus Striated Pardalote cucullata Hooded Robin Dasyornis brachypterus Eastern Bristlebird australis Eastern Yellow Robin Dasyornis longirostris Western Bristlebird Eopsaltria griseogularis continued over continued over

104 Fire and Australian Birds

Table 1. continued Table 1. continued Eopsaltria georgiana White-breasted Robin Lonchura castaneothorax Chestnut-breasted Drymodes superciliaris Northern Scrub-robin Mannikin Drymodes brunneopygia Southern Scrub-robin Erythrura gouldiae Gouldian Finch Pomatostomidae Dicaeidae Pomatostomus superciliosus White-browed Babbler Dicaeum hirundinaceum Mistletoebird Cinclosomatidae Hirundinidae Psophodes olivaceus Eastern Whipbird Hirundo nigricans Tree Martin Psophodes nigrogularis Western Whipbird Sylviidae Cinclosoma punctatum Spotted Quail-thrush Megalurus timoriensis Tawny Grassbird Cinclosoma castanotus Chestnut Quail-thrush Eremiornis carteri Spinifexbird Cinclosoma cinnamomeumCinnamon Quail-thrush Cincloramphus mathewsi Rufous Songlark (includes Nullabor Quail- Cisticola exilis Golden-headed Cisticola thrush)

Neosittidae Daphoenositta chrysopteraVaried Sittella Pachycephalidae Falcunculus frontatus Crested Shrike-tit Oreoica gutturalis Crested Bellbird Pachycephala olivacea Olive Whistler Pachycephala rufogularis Red-lored Whistler Pachycephala pectoralis Golden Whistler Pachycephala rufiventris Rufous Whistler Colluricincla harmonica Grey Shrike-thrush Dicruridae Grallina cyanoleuca Magpie-lark Rhipidura fuliginosa Grey Fantail Campephagidae Coracina novaehollandiae Black-faced Cuckoo-shrike Lalage sueurii White-winged Triller Aatamidae Artamus personatus Masked Artamus superciliosus White-browed Woodswallow Artamus cyanopterus Dusky Woodswallow Cracticus torquatus Grey Butcherbird Cracticus nigrogularis Pied Butcherbird Gymnorhina tibicen Australian Magpie Strepera graculina Pied Currawong Corvidae Corvus coronoides Australian Raven Corvus orru Torresian Crow Corcoracidae Corcorax melanorhamphosWhite-winged Chough Motacillidae Anthus novaeseelandiae Richard’s Pipit Passeridae Taeniopygia guttata Zebra Finch Poephila personata Masked Finch Neochmia ruficauda Star Finch Neochmia temporalis Red-browed Finch continued over

105 Australia’s Biodiveristy - Responses to Fire

Table 2. Summary table of incidence of Table 2. continued subjects in fire-bird references. 5 , , , This tally is based on references included within the Furneaux, , annotated bibliography. Where possible, I have tried Naracoorte Coastal Plain to restrict inclusion in this list to subjects which form 4 Great Sandy Desert, Mulga a substantial component or focus of the paper Lands, NSW South western considered. Note that the same material may be Slopes, Riverina included in several papers (notably in reviews), such 3 , Central that a large tally doesn’t necessarily equate to a Kimberley, Mt Isa Inlier, substantial research effort. A single paper may , consider several subjects, so the tallies do not sum to Nullabor, the number of papers included in the annotated 2 Carnarvon, D’Entrecasteaux, bibliography (361). Bioregions are the biogeographic Lofty Block, Mitchell Grass regions defined in Thackway & Cresswell (1995). Downs, Simpson-Strezlecki Dunefields, , Wet BROAD SUBJECTS Tropics, Woolnorth No. of references subject 1 , Ben Lomond, 141 threatened species 127management North, Brigalow 78 wildfire Belt South, , 60 review , 46 community Darling Riverine Plains, 41 control burning , Einasleigh 26 historic change; seabirds Uplands, Eyre & Yorke 24 hollows Blocks, Finke, Flinders and 21 forestry Olary Ranges, , 20 mortality Gawler, Sandplains, 17succession Gibson Desert, Hampton, 13 isolate , 10 Aboriginal burning MacDonnell Ranges, 9 old-growth Murchison, , 6 charcoal; Aboriginal knowledge , , Yalgoo 2 corridors; slash-burn ENVIRONMENTS BIOREGIONS No. of references habitat No. of references Bioregions 116 (temperate) eucalypt open 37Murray-Darling Depression forest 33 South East Corner 103 heath 29 South East Coastal Plain 48 tussock grassland 28 Jarrah Forest 47mallee 22 25 tropical eucalypt open forest 19 , 24 tropical eucalypt savanna Top End Coastal, Warren woodland 14 Pine Creek-Arnhem 21 hummock grassland 19 rainforest 12 South Eastern Queensland 17(temperate) eucalypt woodland 11 Swan Coastal Plain 15 thicket 10 Cape York Peninsula 10 sedgeland 9 West and South West 8 Acacia woodland/shrubland 8 Ord-Victoria Plains 6 wetlands 7Gulf Falls and 5Uplands Callitris woodland, pandanus, (Allo)Casuarina woodland, 6 Freycinet, NSW North Coast, chenopod shrubland Victorian Midlands 2 Banksia woodland continued over 1 riparian vegetation continued over Fire and Australian Birds

Table 2. continued Table 2. continued 1 Kangaroo Island Emu INDIVIDUAL BIRD SPECIES Australian Pelican, Magpie No. of references bird species Goose, Australian Bustard, 33 Ground Parrot Black-breasted Button-quail, 15 Western Whipbird Chestnut-backed Button-quail, 14 Malleefowl Little Button-quail, Bush 13 Noisy Scrub-bird Stone-curlew, Wedge-tailed 8 Eastern Bristlebird, Western Eagle, Plains-wanderer, Palm Bristlebird Cockatoo, Pink Cockatoo, 7Gouldian Finch Yellow-tailed Black-cockatoo, 6 Rufous Bristlebird, White-tailed Black-cockatoo, Black-eared Miner Long-billed Corella Little 5 Glossy Black-cockatoo, Corella Red-capped Parrot, Golden-shouldered Parrot, Brush Bronzewing, Flock Carpentarian Grass-wren Bronzewing, Partridge Pigeon, White-throated Grass-wren Fork-tailed Swift, Tree Martin, 4 Orange-footed Scrubfowl, Spinifex Bird, Richard’s Pipit, Sooty Owl, Superb Lyrebird, Pilot Bird, Spotted Quail- Splendid Fairy-wren thrush, Northern Scrub-robin, 3 Cassowary, Hooded Parrot, Eastern Yellow Robin, Scarlet Nullabor Quail-thrush, Robin, Crested Shrike-tit, Striated Grass-wren, Purple- Black Grass-wren, White- crowned Fairy-wren, Black winged Fairy-wren, Red- Honeyeater winged Fairy-wren, Western 2 Red Goshawk, Powerful Owl, Thornbill, Striated Thornbill, Red-tailed Black-Cockatoo, Brown Thornbill, Yellow- Superb Parrot, Regent Parrot, rumped Thornbill, Black Turquoise Parrot, Scarlet- Treecreeper, Helmeted chested Parrot, Orange-bellied Honeyeater, Mistletoebird, Parrot, Paradise Parrot, Western Yellow-rumped Pardalote, Rosella, Ringneck Parrot, Night Forty-spotted Pardalote, Star Parrot, White-throated Finch, Yellow-rumped Needletail, Rufous Scrub-bird, Mannikin Grey-crowned Babbler, Flame Robin, Red-lored Whistler, Thick-billed Grass-wren, Rufous-crowned Emu-wren, Southern Emu-wren, Mallee Emu-wren, Brown Thornbill, Slender-billed Thornbill, Forty- spotted Pardalote continued over

107 Australia’s Biodiveristy - Responses to Fire

Table 3. Threatened bird taxa for which Table 3. continued inappropriate fire regime has been listed as a threatening process by Garnett (1992b). Malurus coronatus Purple-crowned Fairy- Extinct taxa are excluded. Conservation status: coronatus wren (western subsp.) (V) R=rare, K=insufficiently known, V=vulnerable, Stipiturus malachurus Southern Emu-wren (Mt E=endangered. intermedius Lofty Ranges subsp.) (E) Stipiturus malachurus Southern Emu-wren Casuarius casuarius Southern Cassowary (V) parimeda (Eyre Peninsula subsp.) Leipoa ocellata Malleefowl (V) (V) Botaurus poiciloptilus Australasian Bittern (K) Stipiturus mallee Mallee Emu-wren (V) Erythrotriorchis radiatus Red Goshawk (V) Amytornis dorotheae Carpentarian Grasswren Turnix castanota Chestnut-backed Button- (K) quail (K) Amytornis striatus Striated Grasswren Turnix olivii Buff-breasted Button- striatus (sandplain subsp.) (K) quail (K) Amytornis striatus Striated Grasswren Turnix varia scintillans Painted Button-quail merrotsyi (Flinders Ranges subsp.) (Houtmans Abrolhos (K) subsp) (V) Pardalotus quadragintus Forty-spotted Pardalote Turnix melanogaster Black-breasted Button- (E) quail (V) Dasyornis brachypterus Eastern Bristlebird (V) Geophaps smithii smithii Partridge Pigeon (eastern Dasyornis longirostris Western Bristlebird (E) subsp.) (V) Dasyornis broadbenti Rufous Bristlebird Geophaps smithii blaauwi Partridge Pigeon (western broadbenti (Victorain subsp.) (R) subsp.) (K) Dasyornis broadbenti Rufous Bristlebird Probosciger aterrimus Palm Cockatoo (K) whitei South Australian subsp.) Calyptorhynchus Red-tailed Black- (K) Cockatoo Acanthiza pusilla Brown Thornbill (King banskii graptogyne (southeastern subsp.) (E) arcibaldi Island subsp.) (K) Calyptorhynchus Glossy Black-Cockatoo Acanthiza iredalei Slender-billed Thornbill lathami halmaturinus (Kangaroo Island subsp.) hedleyi eastern subsp.) (K) (E) Manorina melanotis Black-eared Miner (E) Eclectus roratus Eclectus Parrot (R) Lichenostomus melanops Psephotus chrysopterygius Golden-shouldered Parrot cassidix Helmeted Honeyeater (E) (E) Psophodes nigrogularis Western Whipbird Neophema chrysogaster Orange-bellied Parrot (E) nigrogularis (western heath subsp.) (E) Neophema splendida Scarlet-chested Parrot (R) Psophodes nigrogularis Western Whipbird Pezoporus wallicus Ground Parrot (western oberon (western mallee subsp.) flaviventris subsp.) (E) (R) Pezoporus occidentalis Night Parrot (K) Psophodes nigrogularis Western Whipbird Ninox rufa queenslandica Rufous Owl (eastern lashmari (Kangaroo Island subsp.) subsp.) (R) (R) Ninox rufa meesi Rufous Owl (Cape York Psophodes nigrogularis Western Whipbird Peninsula subsp.) (K) leucogaster (eastern subsp.) (V) Tyto tenebricosa Sooty Owl (R) Falcunculus frontatus Crested Shrike-tit Tyto novaehollandiae Masked Owl whitei (northern subsp.) (K) melvillensis (Melville Island subsp.) Pachycephala rufogularis Red-lored Whistler (V) (K) Erythrura gouldiae Gouldian Finch (E) Atrichornis rufescens Rufous Scrub-bird (R) Cisticola juncidis Zitting Cisticola Atrichornis clamosus Noisy Scrub-bird (E) normani (Normanton subsp.) (K) continued over Zoothera lunulata Bassian Thrush (South halmaturina Australian subsp.) (R)

108 Fire and Australian Birds

Mallee Fowl. Fires in the mallee country may create flushes of food resources for the Mallee Fowl, but generally long-unburnt country is most favourable, especially for breeding. A complex fire mosaic is probably optimal for this species, with special management effort directed to maintaining the most vulnerable long-unburnt patches. /K Thaler © ANBG. Inset photo /Joe Benshemesh

Noisy Scrub-bird. The range and abundance of this Splendid Fairy-Wren. The Splendid Fairy-wren is a unusual bird has declined catastrophically over the small bird found in heathlands in southwestern last 150 years. It is now restricted to a very few Australia. If heathland patches are totally burnt by populations in south-western Australia. Much of this fire, local extinctions may occur, and the Fairy-wren’s decline has been due to the imposition of an poor dispersal ability renders recruitment and inappropriate fire regime. Despite living in highly recolonisation difficult. Cooler fires may also affect flammable heath and thicket vegetation, the Noisy habitat suitability through impacts upon prey Scrub-bird is dependent upon vegetation which has abundance, nest site availability, nest parasitism and remained unburnt for long periods, typically greater predation. /Tom & Pam Gardner © Nature Focus than 40 years. /Ray Smith

109 Australia’s Biodiveristy - Responses to Fire

Gouldian Finch. The endangered Gouldian Finch Red-backed Fairy-wren. The Red-backed Fairy-wren has complex relationships with fire. In its tropical occurs in dense ground layer vegetation across savanna woodland habitat, dry season fires remove northern Australia. Fires remove most of this the dense grass layer, allowing it easier access to vegetation, and may knock out local populations. grass seeds on the ground. Hence, it seeks out A consistent regime of frequent fires over extensive recently burnt areas and in the short term is favoured areas will greatly suppress populations of this by frequent (annual) fires. However, fires may species. /Tom & Pam Gardner © Nature Focus change the seed productivity and timing in the subsequent seeding period, and little is known of the longer-term consequences of a frequent fire regime upon grass species composition and hence food resources. Fire management for this species should aim to maximise environmental variability. /Ian Morris

Magpie Goose. Aboriginal people have a long tradition of burning the northern floodplains once the waters have receded during the dry season. The purpose of burning is to increase the suitability of the area for Magpie Goose feeding and nesting. Where Aboriginal fire management of the floodplains has been removed, Magpie Geese have become disadvantaged. /G O’Neill © ANPWS

110 Fire and Australian Birds

Eastern Bristlebird. The Eastern Bristlebird is one of a Red-tailed Black Cockatoo. Like many parrots and a group of species restricted to heathlands and coastal range of other birds, Red-tailed Black Cockatoos thickets in temperate southern Australia. Recently- require hollows for nest sites, typically formed in very burnt areas are unsuitable, with population densities old trees. Intense fires may destroy such important increasing as vegetation density builds up, for at trees, although fire may have a role in hollow least 15 years post-fire. Its persistence in an area formation. In northern Australia, Red-tailed Black- may require careful management, aimed principally cockatoos also track fires, as food is most readily toward fire suppression. However, some fire is available in recently-burnt areas. almost inevitable in this habitat, and remaining / Babs & Bert Wells © Nature Focus unburnt patches are then critical for re-establishment of populations. / Norman Chaffer Estate © Nature Focus

Ground Parrot. The Ground Parrot is generally severely disadvantaged by frequent fires in heathlands. In some habitats and areas it may also decline in long-unburnt (>15 years) vegetation, but elsewhere fire exclusion is recommended. The impacts of fire may be compounded by habitat fragmentation, which is probably appreciably greater now for heathlands than it was before European colonisation. Such fragmentation may decrease the probability of recolonisation of isolates where fire (and unsuitable post-fire habitat) has destroyed local populations. / John Gray © Nature Focus

111 FIRE AND AUSTRALIAN BIRDS An annotated bibliography

J.C.Z. Woinarski

Parks and Wildlife Commission of the Northern Territory PO Box 496 Palmerston, NT, 0831

INTRODUCTION

This annotated bibliography compiles all (361) bird community, whether it deals with references I could locate which relate to fire and management issues, etc. Bioregions follow Australian birds. It is based on systematic searches Thackway and Cresswell (1995). No bioregions of the journals: are listed for most review articles or where the Emu (1960–1995) observations reported are relatively trivial (e.g. Corella (1977–1995) birds seen feeding on charcoal). A map of the Australian Bird Watcher (1970–1994) locations of primary sources reported is presented South Australian Ornithologist (1968–1994) as an appendix. Australian Bird Bander (1969–1976) The annotations given are my own Sunbird (1992–1995) summaries of relevant sections of the articles, Australian Journal of Ecology (1980–1995) except where material is presented in quotation (Australian) Wildlife Research (1984–1995) marks. Proceedings of the Ecological Society of Australia Much of the material presented here is (1980–1995) discussed in the accompanying review. Northern Territory Naturalist (1978–1994) Although I have made every attempt to be Australian Forestry (1977–1995) comprehensive, many studies of the relationship Australian Forest Research (1981–1987) between birds and fire are presented in the grey Victorian Naturalist (1975–1992) literature. I would be grateful for notification of and less comprehensive searches through any papers I may have missed, or any comment on other journals, reports, theses and other my interpretation of any articles. publications, notably including the general Gill, A.M., Moore, P.H.R., and Martin, W.K. bibliography of Gill et al. (1994) on fire ecology in (1994). Bibliography of fire ecology in Australia, as well as publications indicated to me Australia (including fire science and fire by a range of authorities in most States and management). Edition 4. NSW National Territories. Parks and Wildlife Service, Hurstville. All entries are stored on the bibliographic Thackway, R. and Creswell, I.D. (1995). An data base Endnote Plus. For each reference, I list interim biogeographic regionalisation for keywords and bioregion(s) in italics. Keywords Australia: a framework for setting priorities include the (broad) vegetation formation to which in the National Reserves System the article refers, the main bird taxa considered, Cooperative Program. Version 4.0. whether the article deals with threatened taxa, Australian Nature Conservation Agency: whether the article describes responses of a whole Canberra.

115 ACKNOWLEDGEMENTS

I am very grateful to Allan Burbidge, Grant Wardell-Johnson, Graeme Smith, David Paton, Richard Loyn, Doug Robinson, David Baker- Gabb, Charlie Silveira, David Rounsevell, Robert Taylor, Dan Lunney, Jann Williams, Harry Recher, Hugh Ford, Dick Turner, Mike Fleming, Gay Crowley and Stephen Garnett for comments on an earlier draft and/or pointers to obscure references. I thank the Royal Australasian Ornithologists Union for library access, and Tony Norton for helping arrange library access at ANU. Belinda Oliver constructed the accompanying map. I thank Neal Hardy of the Department of the Environment and Heritage for his patience and assistance in this project.

116 REFERENCES

Abbott, I. (1981). Seabird Islands: No. 106. Arnold, G. W., Smith, G. T., Rowley, I. C. R., Mondrain Island, Archipelago of the & Brooker, M. G. (1993). The effects of fire Recherche, Western Australia. Corella, 5, on the abundance and distribution of animals 60–61. in Australian ecosystems, with emphasis on seabirds; heath; tussock grassland; mortality Mediterranean ecosystems. In L. Trabaud & Bioregion: Esperance Plains R. Prodon (Eds.), Fire in Mediterranean Fires since 1801 have repeatedly consumed much ecosystems (pp. 237–257). Brussels: of the vegetation, destroying many breeding Commission of the European Communities. seabirds. review; heath; mallee; eucalypt open forest; isolate Reviews studies of effects of fire on birds (and other fauna) in Mediterrean ecosystems (heath, Abbott, I., & Christensen, P. (1994). mallee and open forest) in southwestern and Application of ecological and evolutionary southeastern Australia. More studies are required principles to forest management in Western on the influence of fire mosaic or patchiness. Australia. Australian Forestry, 57, 109–122. Mammals (at least in eastern Australia) show management; control burning; eucalypt open forest pronounced seral responses to fire. Lizards and Bioregion: Jarrah Forest frogs are generally not greatly affected by fire. Reviews some studies of the impacts of fuel Fauna of mediterranean systems shows less clear reduction burning on birds in Jarrah forests, seral responses to fire than those in temperate suggesting limited effect. ecosystems, as the fauna of mediterranean ecosystems has adapted to substantial climatic variability. There are substantial differences Adam, P., & Robinson, D. (1996). Negative between bird species in response to fire, for effects of fuel-reduction burning on the example the Ground Parrot is fire dependent habitat of the Grey-crowned Babbler whereas Noisy Scrub-bird is fire sensitive. Pomatostomus temporalis. Victorian Naturalist, Long-term demographic studies are needed to 113, 4–9. comprehend and predict responses to fire. Fire Grey-crowned Babbler; threatened species; management is an essential component of the management; control burning; eucalypt woodland management of conservation reserves: this may be Bioregions: Victorian Midlands; South Eastern complicated by fragmentation. Highlands Frequent (annual) fuel-reduction burning (in remnant vegetation along roadside verges) is Ashby, E. (1921). Notes on the supposed having a considerable impact on the habitat (and “extinct” birds of the south-west corner of hence numbers) of the threatened Grey-crowned Western Australia. Emu, 20, 123–124. Babbler. Warns that findings of little impact of Ground Parrot; Western Whipbird; succession; heath; fuel-reduction burning on birds should be treated threatened species. with caution. Alternative mechanisms of fire Bioregion: Swan Coastal Plain protection (e.g. slashing) should be considered in Frequent fires (lit by farmers) had changed order to safeguard these populations. heathlands to grasslands at a site in southwestern Australia eliminating Ground Parrots and Western Whipbirds.

117 Australia’s Biodiveristy - Responses to Fire

Ashby, E. (1924a). Notes on extinct or rare Backhouse, G. N. (1987). Management of Australian birds, with suggestions as to some remnant habitat for conservation of the of the causes of their disappearance. Part I. Helmeted Honeyeater Lichenostomus melanops Emu, 23, 178–183. cassidix. In D. A. Saunders, G. W. Arnold, review; threatened species; Night Parrot; hummock A. A. Burbidge, & A. J. M. Hopkins (Eds.), grassland; Ground Parrot; heath; wetlands; Kangaroo Nature conservation: the role of remnants of Island Emu. native vegetation (pp. 287–294). Chipping Speculates that Night Parrots persist only where Norton: Surrey Beatty. hummock grasses are “unswept by fires”. Helmeted Honeyeater; eucalypt open forest; wildfire; Disappearance of Ground Parrot due to frequent management; threatened species; isolate fires. Extinction of Kangaroo Island Emu was due Bioregion: South Eastern Highlands to hunting and excessive burning. The sites of two small isolated populations of Helmeted Honeyeaters were severely burnt by bushfire, destroying much of the habitat. Both Ashby, E. (1924b). Notes on extinct or rare populations disappeared after fire. At the site of Australian birds, with suggestions as to some the remaining population (a narrow riparian strip) of the causes of their disappearance. Part II. fire prevention works may create more Emu, 23, 294–298. disturbance than fire. review; threatened species Review of decline of Australian birds. “The most Baird, R. F. (1986). Historical records of the serious factor of all is the destruction of both food Glossy Black Cockatoo Calyptorhynchus supply and shelter, and, of course, breeding haunts, lathami and Red-tailed Black Cockatoo by bush-fires. I consider that these … factors, C. magnificus in south-eastern Australia. South especially … bush-fires, account for more than Australian Ornithologist, 30, 38–45. nine-tenths of the disappearance of certain forms Glossy Black-Cockatoo; historic change; threatened … I am confident that the indiscriminate burning species; review; Casuarina woodland of bush, which is the concomitant of all farming Bioregion: Woolnorth and grazing operations in bush country, is by a long Considers historic change in the distributions of way the major cause of the disappearance of many Glossy Black-Cockatoo and Red-tailed of our rarer birds.” Black-Cockatoo. Declines initially as a result of climate change, more recently exacerbated by Aumann, T., & Baker-Gabb, D. J. (1991). The clearing and fire (at least for Glossy ecology and status of the Red Goshawk in Black-Cockatoo on King Island) northern Australia. RAOU Report No. 75. Royal Australasian Ornithologists Union. Baker, J., & Whelan, R. J. (1994). Ground Red Goshawk; tropical eucalypt open forest; threatened Parrots and fire at Barren Grounds, New species; management. South Wales: a long-term study and an Fires have been reported to burn out the nests of assessment of management implications. Emu, Red Goshawk, killing young. In the long term, a 94, 300-304. combination of grazing and frequent fires may Ground Parrot; threatened species; heath. result in reduction of prey and loss of nest trees, Bioregion: South East Corner however short-term results show little clear Some previous studies have suggested that indication of differences in density or breeding suitability of heaths for Ground Parrots (and hence success related to burning history. Ground Parrot abundance) peaks at around 10 years post-fire, and hence have recommended fire intervals of 8-10 years. In contrast, this study Australian Biological Research Group Pty. suggests that there is little evidence for a decline in Ltd. (1987). Management Plan for the Ground suitability, but rather that abundance plateaus (at c Parrot in the Orbost Region. Australian 0.25birds/ha) after around 5 years. This was based Biological Research Group Pty Ltd. on a series of censuses of the same locality from Ground Parrot; heath; sedgelands; threatened species; immediately after fire to 10.7 years post-fire. Few management. studies have examined abundance in heaths >15 Bioregion: South east Corner years. No single fire regime is typical of all areas Fire regimes proposed for heathlands to maintain where Ground Parrots occur. suitability for Ground Parrots. 118 Fire and Australian birds

Baker, L., Woenne-Green, S., & Mutitjulu Bamford, M. J. (1985a) The dynamics of small Community (1993). Anangu knowledge of vertebrates in relation to fire in banksia vertebrates and the environment. In J. R. W. woodland near Perth, Western Australia. PhD, Reid, J. A. Kerle, & S. R. Morton (Eds.), Murdoch University. Uluru fauna. The distribution and abundance of Banksia woodland; succession; community vertebrate fauna of Uluru (Ayers Rock-Mount Bioregion: Swan Coastal Plain Olga) National Park, N.T. (pp. 79-132). Bird communities were assessed at six sites across Canberra: Australian National Parks and a range of ages (0-22 years) since fire. Soon after Wildlife Service. fire in Banksia woodland, bird richness decreased, management; Aboriginal knowledge; mallee; hummock but the site was colonised by open-country birds grassland; Acacia woodland. from surrounding farmlands. Bioregion: Describes Aboriginal knowledge of wildlife. Many granivores are favoured by fire, but Emu dislike Bamford, M. J. (1985b). The fire-related freshly-burnt country. Burning in spring can dynamics of small vertebrates in Banksia disrupt their breeding and cause them to move woodland: a summary of research in progress. away. Fires lit in strong winds can kill many small In J. R. Ford (Ed.), Fire ecology and animals. Discusses traditional Aboriginal burning management in Western Australian ecosystems regimes. (pp. 107-110). Perth: Western Australian Institute of Technology. Banksia woodland; community. Baker-Gabb, D. J., Benshemesh, J. S., & Bioregion: Swan Coastal Plain Maher, P. N. (1990). A revision of the Six Banksia sites with a range of fire histories were distribution, status and management of the monitored for vertebrates. Birds were more Plains-wanderer Pedionomus torquatus. Emu, affected by fire than reptiles but less than 90, 161-168. mammals. Most bird species were recorded within Plains-wanderer; tussock grassland; wildfire; a few months of fire. At one site, the number of management; threatened species. bird species was dramatically lower soon after a Bioregions: Murray Darling Depression; Channel fire, but returned to pre-fire levels within 1 year, Country; Riverina; Victorian Volcanic Plain although total number of individuals remained Plains-wanderers are sedentary in sparse low for at least 2 years. No species showed a grasslands unless these are overgrazed, burnt or preference for the long unburnt sites. A few cultivated. It is likely that most young and many species were more common in the most recently adults at one site perished when an extensive fire burnt sites. These were mainly species typical of swept through the area, and wildfire may have the surrounding farmlands. severe impact on this species. This was the first such fire recorded from this area for 70 years. Bates, R. (1980). After the fire: some observations on the effect of the February 20, Baldwin, M. (1965). Bird eating charcoal. 1980 (Ash Wednesday) bushfires on wildlife in Emu, 64, 208. the Adelaide Hills. South Australian Naturalist, charcoal 54, 77-79. Four species (Fairy Martin, Dusky Wood-swallow, wildfire; eucalypt open forest Double-barred Finch and Zebra Finch) recorded Bioregion: Lofty Block. eating charcoal.

119 Australia’s Biodiveristy - Responses to Fire

Bedggood, G. W. (1980). Birdlife between Benshemesh, J. (1988). Report on a study of Lake Tyers and Marlo, Victoria. Australian malleefowl ecology. Department of Bird Watcher, 8, 147-162. Conservation, Forests and Lands. heath; eucalypt open forest; control burning. mallee; Malleefowl; threatened species; wildfire Bioregion: South east Corner Bioregion: Murray Darling Depression Suggests that the widespread spring fuel reduction A detailed case study of the response of burns result in heavy toll of young birds and eggs, Malleefowl to an intense but patchy fire. Results excessive loss of hollows suitable for breeding, loss are summarised in Benshemesh (1990). Small of protective undergrowth, loss of ground-nesting unburnt patches were vital for the persistence of birds, and reduced germination for the species. Long fire-free intervals (>60 years) are spring-flowering plants. optimal.

Beeton, R. J. S. (1985). The little corella: a Benshemesh, J. (1990). Management of seasonally adapted species. Proceedings of the Malleefowl - with regard to fire. In Ecological Society of Australia, 13, 53-63. J. C. Noble, P. J. Joss, & G. K. Jones (Eds.), Little Corella; tussock grassland; tropical eucalypt The mallee lands: a conservation perspective savanna woodland (pp. 206-211). Melbourne: CSIRO. Bioregions: Ord-Victoria Plains; mallee; Malleefowl; threatened species; wildfire Victoria-Bonaparte Bioregion: Murray Darling Depression At a season when food is generally limited Optimal fire frequency for Malleefowls is >60 (March-April), Little Corellas flock to small areas years. Broad-scale fires eliminate Malleefowl in recently burnt, presumably because such fires the short-term, and even 20-30 years post-fire, increase accessibility to seeds. breeding densities are only about one third of those in long-unburnt mallee (based on four matched pairs of 20-30 year old and >40 year old Belcher, C. (1993). Rufous Bristlebird survey sites). Patchy burns provide some refuges from and habitat analysis, Port Campbell National which recolonisation can occur. Sufficient litter Park 1992. Department of Conservation and for nesting is generally unavailable until 10-15 Natural Resources. years post-fire. Previous authors had suggested Rufous Bristlebird; heath; thicket; threatened species; that food resources were more abundant in isolate younger mallee and that periodic fire was needed Bioregion: South East Coastal Plain for their maintenance. Modelled densities were At Port Campbell National Park optimal habitat 6% of maximum carrying capacity for 20 year fire for Rufous Bristlebird is climax heathland more interval, 30% at 40 year intervals, and 54% at 60 than 25 years post-fire. Isolated populations are year intervals. susceptible to elimination by fire, and the poor At least 10 of 11 radio-marked and banded dispersal ability of the species may limit birds survived a patchy but intense fire, but in the subsequent recolonisation months following the fire all but 4 had emigrated or died. Dispersing birds used corridors of unburnt vegetation rather than traversing Bennett, S. (1983). The Northern Scrub-robin extensive burnt areas. The remaining birds bred Drymodes superciliaris in the Northern in the season following fire in small unburnt Territory. Emu, 83, 105-107. patches. These birds foraged extensively in burnt Northern Scrub-robin; rainforest; threatened species areas, using the rich pulse of herbs growing after Bioregion: Gulf Coastal fire. If records from the Roper River earlier this century are valid, the Northern Scrub-robin has become extinct in the NT, most likely because of change in fire regimes leading to decline in the extent of rainforest thickets.

120 Fire and Australian birds

Benshemesh, J. S. (1992) The conservation Boekel, C. (1980). Birds of Victoria River biology of Malleefowl, with particular regard to Downs Station and of Yarralin, Northern fire. PhD, Monash University. Territory. Part 1. Australian Bird Watcher, 8, Malleefowl; mallee; Callitris woodland; wildfire; 171-193. old-growth; threatened species. tropical eucalypt savanna woodland; hunting Bioregion: Murray Darling Depression Bioregion: Ord-Victoria Plains Detailed study of habitat requirements and Describes the use of smoke and traps by biology of the Malleefowl, particularly in response Aboriginal people for catching eagles and kites. to a wildfire. Malleefowl requires long-unburnt mallee to provide the extensive litter needed for breeding. Such old-growth has become extremely Bowman, D. M. J. S., Woinarski, J. C. Z., & limited. Malleefowls may survive fires if unburnt Russell-Smith, J. (1994). Environmental patches occur. relationships of Orange-footed Scrubfowl Megapodius reinwardt nests in the Northern Territory. Emu, 94, 181-185. Benshemesh, J. S. (1994). Malleefowl Leipoa Orange-footed Scrubfowl; historic change; rainforest. ocellata. Department of Conservation and Bioregions: Top End Coastal; Pine Creek Natural Resources. Arnhem; Central Arnhem Malleefowl; mallee; threatened species; wildfire; Environmental changes, probably including management historic changes in fire regime, have led to some Bioregion: Murray-Darling Depression contraction of rainforests, as evidenced by the Reviews information on Malleefowl (in Victoria). occurrence of old scrubfowl mounds in areas “The extent and frequency of fires pose a serious which are now eucalypt open forests. threat to the conservation of Malleefowl as remaining populations may be destroyed and habitat quality reduced for 40 years or more … the Braithwaite, L. W., Clayton, M., MacLean, L., effect of fire is exacerbated by the fragmentation & Parker, B. S. (1984). Vertebrate fauna of a due to clearing, as isolated reserves that are 144-ha water catchment within eucalypt forest entirely burnt are unlikely to be recolonised … being harvested for woodpulp at Eden, more effective fire control may be the single most south-eastern New South Wales. CSIRO important factor in improving the conservation Wildlife and Rangelands Research. status of the species.” The scale of fire in mallee eucalypt open forest; wildfire; community lands is a problem, as it may be comparable to that Bioregion: South East Corner of the largest reserves. Birds may survive and breed Birds were surveyed at a site for 15 months before after patchy fires. and 2 months after January wildfire. Parts of the site were logged during this period. Species Bill, M. E. (1932). Lyre-birds and bushfires. richness declined after fire. Laughing Victorian Naturalist, 49, 24. Kookaburra, Grey Butcherbird, Flame and Scarlet Superb Lyrebird; wildfire Robins increased after fire. Striated Thornbill, Red Wattlebird, Crescent Honeyeater and Pied Currawong decreased. Blakers, M., Davies, S. J. J. F., & Reilly, P. N. (1984). The atlas of Australian birds. Melbourne: Melbourne University Press. review Distributional information on all Australian birds, but includes information on threatening processes. Notes Orange-bellied Parrot prefers heaths and button-grass plains <15 years post-fire for feeding. Increase in fire frequencies or intensities may have led to decline in Gouldian Finch and Pictorella Mannikins. Notes references to fire for other species (e.g. Emerald Dove, Partridge Pigeon, Paradise Parrot, Scarlet-chested Parrot, Golden-headed Cisticola, White-throated

Grass-wren, Eastern Bristlebird). 121 Australia’s Biodiveristy - Responses to Fire

Braithwaite, R. W. (1985). Fire and fauna. In country around waterholes. The short-term R. W. Braithwaite (Ed.), Kakadu Fauna Survey. attraction of many mobile species to burnt areas is Final report to Australian National Parks and in contrast to the succession described in Wildlife Service. (pp. 634-650). Darwin: temperate areas. CSIRO. management; tropical eucalypt open forest; tropical eucalypt savanna woodland; rainforest Bramwell, M., Pyke, G., Adams, C., & Coontz, Bioregions: Pine Creek Arnhem; Top End Coastal P. (1992). Habitat use by Eastern Bristlebirds A wide range of carnivorous and hawking birds are in Barren Grounds Nature Reserve. Emu, 92, attracted to fire. Soon after fire, many carnivorous 117-121. and granivorous birds move into burnt areas to Eastern Bistlebird; threatened species; heath; eucalypt take advantage of greater abundance or woodland. accessibility of resources. Some nests (of Brown Bioregion: South east Corner Honeyeater, Mistletoebird, White-throated Eastern Bristlebird population density was greater Honeyeater) were found to be destroyed by fire. in heath/woodland 9 years after fire than 6 years after fire.

Braithwaite, R. W. (1996). Biodiversity and fire in savanna landscapes. In O. Solbrig, Brickhill, J. (1980). Striated Grasswren E. Medina, & J. F. Silva (Eds.), Biodiversity and Amytornis striatus. In C. Haigh (Ed.), savanna ecosystem processes: a global perspective Endangered animals of New South Wales (pp. 121-140). Berlin: Springer-Verlag. (pp. 68). Sydney: NSW National Parks and tussock grasslands; tropical eucalypt savanna woodland; Wildlife Service. tropical eucalypt open forest; review threatened species; mallee; hummock grassland; Reviews some previous studies from northern Striated Grass-wren Australia. In general, there is little succession, but Extensive wildfires in mallee have severe and increase in bird abundance soon after fires. long-term impacts on Striated Grass-wren.

Braithwaite, R. W., & Estbergs, J. (1987). Brickhill, J. (1987) The conservation status of Fire-birds of the Top End. Australian Natural malleefowl in New South Wales. M.Nat.Res.Sc., History, 22, 299-302. University of New England. community; tropical eucalypt open forest; tropical mallee; Malleefowl; threatened species eucalypt savanna woodland. Bioregion: Murray Darling Depression Bioregions: Top End Coastal; Pine Creek Arnhem Periodic fire may be needed to maintain the food Several species (notably Brown Falcon, Black resources on which the Malleefowl depends. A Falcon, Whistling Kite, Black Kite, fine-scale fire-age mosaic benefits Malleefowl , Tree Martin) are attracted to fires, through provision of a diversity of food types. often in large aggregations. Torresian Crow, raptors, , Straw-necked Ibis, Brooker, L. C., & Brooker, M. G. (1994). A Black-faced Cuckoo-shrike, nightjars, Red-tailed model for the effects of fire and fragmentation Black-cockatoo, Little Corella, Galah, quail, on the population viability of the Splendid , Blue-winged Kookaburra, Fairy-wren. Pacific Conservation Biology, 1, Red-backed Kingfisher, Forest Kingfisher, 344-358. Partridge Pigeon and Magpie-lark forage in the Splendid Fairy-wren; isolate; management; wildfire; burnt area immediately (to several months) after control burning; heath; long-term study fire, consuming carrion, more accessible animal Bioregion: Swan Coastal Plain food or fallen seeds. In the months following fire, Based on 17 years breeding, parasitism and vegetation regrowth attracts herbivorous insects survival data, a model of the effects of fire (both and birds feeding on them (e.g. Straw-necked wildfire and controlled burning) on populations of Ibis). In the wet season, many of these species Splendid Fairy-wrens is developed and explored. migrate into more arid areas where fires continue The frequency and extent of fire can readily drive to occur, though most are then no longer isolated populations to extinction. This associated with burnt areas but rather open

122 Fire and Australian birds

probability increases with smaller population size material (e.g. cobwebs) and/or inadequate food for (=smaller area of isolate), and increase in production. Yellow-rumped Thornbills used probability of fire. In this case, the main effect of novel nest sites post-fire, but used only 2 species fire is manifested through increase in subsequent of plant as nest sites, compared to 11 pre-fire. nest predation. Of 26 species which bred in the site in the year preceding the fire, 21 nested in burnt areas in the year following fire. White-browed Scrubwren Brooker, M. G. (1988). Some aspects of the vacated the area and did not nest for 2 years biology and conservation of the Thick-billed post-fire (possibly due to limited food resources Grasswren Amytornis textilis in the Shark Bay through lack of litter). White-cheeked area, Western Australia. Corella, 12, 101-108. Honeyeater did not nest until 4 years post-fire. Thick-billed Grass-wren; succession Inland Thornbill became rare and had failed to Bioregion: Carnarvon nest by 5 years post-fire. Fantailed Cuckoo Increased tourism and de-stocking could lead to remained abundant at the site but did not breed changes in fire regime. Evidence of past severe until its hosts (White-browed Scrubwren and fires, but current frequency of fires is low. Inland Thornbill) returned to breed. Little Thick-billed Grasswrens can occur in vegetation Button-quail and Elegant Parrot were recorded within a few years post-fire, but fire is likely to be nesting only after the fire (attracted to the a main management consideration. temporary conversion of heath to grassland), and White-winged Triller and Dusky Wood-swallow were more numerous breeders post-fire. Brooker, M. G., & Rowley, I. (1991). Impact Although most species appeared to have of wildfire on the nesting behaviour of birds in some individuals which survived fire and showed heathland. Wildlife Research, 18, 249-263. adaptable behaviour in subsequent nesting, less heath; eucalypt open forest; wildfire; long-term study; obvious effects (such as increased mortality, Western Thornbill; Splendid Fairy-wren; lowered productivity and altered age structure) Yellow-rumped Thornbill; breeding; control burning may be more critical in determining the ultimate Bioregion: Swan Coastal Plain viability of populations. An intense wildfire burnt almost all of the study “For small heathland passerines, control site half-way through a nine-year study: three burns at any time of year and even five years apart wildfires burnt the rest of the study area over the could make an area uninhabitable.” Suggests next 3 years. Birds changed their nest siting after intervals of perhaps as long as 10 years are needed fires. In contrast to pre-fire locations, Splendid to maintain these populations. Fairy-wrens nested only in resprouter plant species in the first year post-fire. By the second year post-fire, they used some seeding shrubs. Brooker, M. G., Ridpath, M. G., Estbergs, A. J., They had difficulty attaching nests to post-fire Bywater, J., Hart, D. S., & Jones, M. S. (1979). substrate, with several nests falling to the ground. Bird observations on the north-western Western Thornbills were unable to nest in Nullabor Plain and neighbouring regions, preferred Hakea in the year post-fire but nested 1967-1978. Emu, 79, 176-190. instead in holes of eucalypts and under shedding chenopod shrubland; wildfire bark. They changed their placement of nests in Bioregion: Nullabor Xanthorrhea. Very small unburnt patches were A widespread fire followed by drought may have not used selectively by any of the three main bird led to local loss (for at least 5 years) of species considered. The height distribution of White-winged Fairy-wren from a bluebush site. nests generally changed in burnt vegetation. Breeding of Splendid Fairy-wren was delayed by 3-5 weeks in the year after fire, and the number of nests built per group had almost doubled by the second year (probably due to high rate of nest failure). Western Thornbills also delayed breeding by up to 5 weeks in burnt areas (cf unburnt) in the year after fire, and only 59% of females attempted to breed. The delay was due to shortage of nesting

123 Australia’s Biodiveristy - Responses to Fire

Brooker, M. G., Braithwaite, R. W., & Brouwer, J., & Garnett, S. (Ed.). (1990). Estbergs, J. A. (1990). Foraging ecology of Threatened birds of Australia: an annotated list. some insectivorous and nectarivorous species Melbourne: Royal Australasian Ornithologists of birds in forests and woodlands of the Union. Wet-Dry tropics of Australia. Emu, 90, review; threatened species 215-230. Of 52 threatened bird species, inappropriate fire tropical eucalypt open forest; tropical eucalypt savanna regime is a factor affecting status for 22 species. woodland; community Bioregions: Pine Creek Arnhem; Top End Coastal Study of foraging behaviours of insectivorous and Brown, P. B., & Wilson, R. I. (1981). A survey nectarivorous birds across a range of habitats. of the Orange-bellied Parrot Neophema “… controlled burning and the prevention of chrysogaster in Tasmania, Victoria and South previously frequent wildfires are but a few examples Australia. National Parks and Wildlife Service, of disturbances which may change … the abundance Tasmania. and diversity of birds.” More than half of the Orange-bellied Parrot; threatened species; insectivorous species forage from the ground or management; eucalypt open forest; sedgelands; hollows from shrubs and grass. “The lower strata are the Bioregion: West and South West most affected by fire … (and) the needs of this large Uncontrolled burning of button-grass plains in ground and shrub-foraging guild should therefore southwestern Tasmania extend to the breeding sites be considered in the formulation of fire of Orange-bellied Parrots in fringing open forest, management plans. A range of fire types throughout to the extent that, when nesting, they “are the year might increase the habitat diversity … with constantly at risk from wildfires”. However, the a positive effect on the diversity of bird species.” Parrots feed in the button grass plains, sedgelands and heaths, and prefer relatively young vegetation (albeit in a seasonally rotating order: in October Brothers, N. P. (1983). Seabird Islands: No. and November they fed mainly in 7-8 year old 136. Actaeon Island, Tasmania. Corella, 7, regrowth, in December they mainly fed in 1-4 year 89-90. old regrowth, in January and February they mainly seabirds; heath; tussock grassland fed in regrowth >8 years post-fire). A fire Bioregion: D’Entrecasteaux management plan is required. “The vegetation on the island has been frequently burnt for many years and … this may have significantly affected the habitat suitable for Brown, P. B., & Wilson, R. I. (1984). The burrowing” for nesting seabirds. Orange-bellied Parrot. In R. H. Groves & W. D. L. Ride (Eds.), Species at risk (pp. 106-116). Canberra: Australian Academy of Brothers, N. P., & Skira, I. J. (1987). Seabird Science. Islands: No. 173. Chappell Island, Furneaux Orange-bellied Parrot; heath; sedgelands; threatened Group, Tasmania. Corella, 11, 81-82. species; management. seabirds; tussock grassland Bioregion: West and South West Bioregion: Furneaux Fire management is critical for the maintenance of Firing and grazing has substantially modified the Orange- bellied Parrot at their breeding grounds. vegetation, considerably reducing the distribution Relatively young regrowth heath (7 years post-fire) and abundance of breeding shearwaters. is preferred for feeding, and is probably unsuitable by 10-12 years. In its breeding area it roosts in thick regrowth of 7-10 years post-fire. Brothers, N. P., & Skira, I. J. (1988). Seabird Islands: No. 185. Little Dog Island, Furneaux Group, Tasmania. Corella, 12, 85-86. seabirds; tussock grassland Bioregion: Furneaux Repeated fires have changed vegetation patterning, probably affecting suitability for shearwaters.

124 Fire and Australian birds

Bryant, S. L. (1991). The Ground Parrot, Burbidge, A. H., & Pedler, L. (1993). Pezoporus wallicus, in Tasmania: distribution, Conservation status of the Nullabor density and conservation status. Scientific report Quail-thrush. World Wide Fund for Nature. no. 91/1. Parks, Wildlife & Heritage, Nullabor Quail-thrush; threatened species; chenopod Tasmania. shrubland; management Ground Parrot; threatened species; heath; sedgelands. Bioregion: Nullabor Bioregion: West and South West Hot fires may make areas unsuitable for Nullabor Provided density estimates of Ground Parrots at Quail-thrush, and this effect may be exacerbated 185 sites of varying ages post-fire. No clear peak by grazing preventing regrowth. in abundance at a particular age. Did not recommend prescribed burning on current information. Burbidge, A. H., & Pedler, L. (1997). Habitat of the threatened Nullabor Quail-thrush. Wildlife Research, 24. Bryant, S. L. (1992). The Ground Parrot and Nullabor Quail-thrush; chenopod shrublands; age of vegetation in Tasmania. In L. Joseph threatened species; management (Ed.), Issues in the conservation of parrots in Bioregion: Nullabor Australasia and Oceania: challenges to Interacting with grazing by rabbits and livestock, conservation biology. (pp. 42-45). Melbourne: and weed invasion, fire is a main threat to the Royal Australasian Ornithologists Union. Nullabor Quail-thrush. Habitat should be Ground Parrot; threatened species; heath. managed to decrease risks (or frequency) of Bioregion: West and South West extensive fire (especially hot wildfire). Ground Parrots occur in heath across a very broad range of ages after fire, though are least abundant in early regrowth. Burbidge, A. H., Watkins, D., & McNee, S. (1989). Conservation of the Ground Parrot in Western Australia. Final report (Project 118). Bryant, S. L. (1994). Habitat and potential diet World Wildlife Fund. of the Ground Parrot in Tasmania. Emu, 94, Ground Parrot; threatened species; heath; 166-171. management; control burning threatened species; heath; sedgelands; Ground Parrot. Bioregion: Esperance Plains Bioregion: West and South West Ground Parrots in WA prefer long-unburnt Recorded in vegetation from 1 to 90 years heath, but it is not clear at what age heaths post-fire. Minimum vegetation cover required is become suitable. “The major approach for the achieved in most situations in Tasmania within management of Ground Parrots in these areas at 12 months post-fire. Growth rates (and phenology) this time must be one of total fire exclusion”. If of moorland plant species are very variable, leading fuel reduction burning must be done (for buffer to wide range of vegetation ages used. strips) this should be done in autumn, to avoid disruption of breeding). Describes floristics and results of radio-tracking. Burbidge, A. A., Folley, G. L., & Smith, G. T. (1986). The Noisy Scrub-bird. Western Australian Wildlife Management Program 2. Burbidge, A. H., McNee, S., Newbey, B., & Department of Conservation and Land Rolfe, J. (1990). Conservation of the Ground Management. Parrot in Western Australia. Supplementary Noisy Scrub-bird; eucalypt open forest; heath; report: project 118. World Wildlife Fund. threatened species; management. Ground Parrot; heath; management; threatened Bioregion: Esperance Plains species. The former frequent burning and drainage of Bioregion: Esperance Plains swamps would have led to decline of this species. Results of some censuses and floristic associations Management requires fire control. for Ground Parrot in . The western subspecies appears to have very different habitat and management requirements to the eastern subspecies. Monitoring of sites following wildfire is outlined.

125 Australia’s Biodiveristy - Responses to Fire

Cale, P. G., & Burbidge, A. H. (1993). Research Calver, M. C., Hobbs, R. J., Horwitz, P., & plan for the Western Ground Parrot, Western Main, A. R. (1996). Science, principles and Whipbird and Western Bristlebird. Australian forest management: a response to Abbott and National Parks and Wildlife Service. Christensen. Australian Forestry, 59, 1-6. Ground Parrot, Western Whipbird, Western management; control burning; eucalypt open forest Bristlebird; threatened species; management; heath; Bioregion: Jarrah Forest mallee. Disputes claims (made by Abbott and Bioregions: Esperance Plains; Jarrah Forest; Christensen) of limited impacts of fuel reduction Warren burning in jarrah forests; no new data are Reviews conservation status, research priorities presented. and responses to fire for three threatened species (Western Whipbird, Western Bristlebird and Ground Parrot). Fire management of these three Campbell, A. G. (1937). Birds of Wilson’s species (and also Noisy Scrub-bird) should be Promontory. Emu, 37, 157. integrated at sites of co-occurrence, although this eucalypt open forest may be complicated by somewhat different Bioregion: South east Coastal Plain responses to fire. For Ground Parrot, cautions Birds in the Wilson’s Promontory area have been against correlative studies of abundance and time much less common because of fire, and are likely since fire (other factors may be involved). Limited to be further affected. “The forests of the west data from WA suggest response may differ to that coast are gone and the forests of the east are recorded from eastern Australia: birds in WA use doomed owing to the ravages of fire … the utterly long-unburnt vegetation (>20, and maybe >30 senseless waste of the natural assets of this years post-fire), and are more abundant in a national park can only be termed a satire upon the long-unburnt area than an adjacent 6-year methods of those in control”. post-fire area. This difference may reflect slower growth rates of heath in WA. “Determining the Carpenter, G., & Matthew, J. (1986). The age after fire at which vegetation becomes suitable birds of Billiatt Conservation Park. South for Ground Parrots and the length of time for Australian Ornithologist, 30, 29-37. which it remains suitable is essential for the mallee; Callitris woodlands; community; hollows proper long-term management of the species”: Bioregion: Murray Darling Depression this can only be done by monitoring existing Vegetation of several ages since fire was sampled. populations and/or determining the time after fire “Fire has an important influence on bird at which colonisation occurs. “Fire has been populations within the Park due to its effect on identified as the greatest threat to P.n. vegetation. For example, hollows suitable for nigrogularis … (but) little is known about the nesting are occasionally produced by fires. responses to fire of P.n. oberon”. Management Elsewhere, suitable hollows exist only in and research for the former subspecies should long-unburnt eucalypts and Callitris.” Areas burnt consider whether there is an upper limit on the <10 years before were unsuitable for Gilbert’s and age of vegetation used; for the latter subspecies Red-lored Whistlers, Southern Scrub-robin and research should examine more comprehensively Western Whipbird, but suitable for Hooded and long-term responses of known populations to Red-capped Robins. Relationship of birds with vegetation age and hence derive fire management age since fire may be changed with varying plans. For Western Bristlebirds, research has floristics: Callitris is slower-growing than indicated that heaths older than c45 years become eucalypts and maintains shrubbiness longer, less suitable (due to decrease in productivity and therefore species requiring shrubby vegetation floristic/structural changes). Research and may persist in Callitris longer than in eucalypts. management should continue long-term Southern Scrub-robin, Shy Heath-wren, Crested monitoring of population size with heaths of Bellbird and Purple-gaped Honeyeater were increasing age, and examination of the impacts of common in areas burnt 23 years previously, or fire breaks. older areas that had shrubby understorey. Mallee-fowl mounds were most common in areas older than 23 years post-fire with open understorey, but recently burnt areas may offer

126 Fire and Australian birds

rich food supplies for this species. Several species other features of the Australian environment. showed no apparent relationship with fire (Grey Primitive birds are mainly in least fire-prone Shrike-thrush, Golden Whistler, Splendid and environments (rainforests). Fire-prone Variegated Fairy-wrens, White-browed Babbler, environments tend to be species-poor overall. Weebill, Inland Thornbill, Yellow-rumped Ground Parrots may be fire-specialists, requiring Pardalote, Grey Butcherbird, and Brown-headed, heath of a certain range of ages post-fire. Superb Yellow-plumed, White-eared and Spiny-cheeked Lyrebird and Common Bronzewing may require Honeyeaters). fire to eliminate wiregrass and enhance production of Acacia seeds respectively. White-browed Scrubwren decreased after fire in a eucalypt forest. Carter, T. (1923a). Birds of the Broome Hill In eucalypt forests, bird numbers in the district. Part I. Emu, 23, 125-142. understorey and ground layers decline with long Malleefowl; mallee; heath; threatened species periods after fire, and more frequent burning may Bioregion: Avon Wheatbelt be required for these lower layers. In 1902 the Malleefowl was common in coastal vegetation, but its population had diminished by 1920 because of burning of the coastal vegetation Chaffer, N. (1954). The Eastern Bristle-bird. to improve grazing for cattle. Emu, 54, 153-162. Eastern Bristlebird; threatened species; heath; wildfire Bioregion: South East Corner. Carter, T. (1923b). Birds of Broome Hill Repeated bushfires may have reduced the district. Part II. Emu, 23, 223-235. numbers of Eastern Bristlebirds, although they Western Whipbird; heath; threatened species have survived in areas of frequent fire. Bioregion: Avon Wheatbelt Repeated fires have destroyed much of the heath and thickets that the Western Whipbird requires, Chambers, W. (1983). Birds at Urquhart’s and it is becoming rarer. Bluff: before and after the fire. Geelong Naturalist, 20, 50-51. eucalypt open forest; wildfire Carter, T. (1924). Birds of the Broome Hill Bioregion: South east Coastal Plain District. Part III. Emu, 23, 306-318. Anecdotal notes of species before fire and then Rufous Bristlebird; heath; threatened species gradually recolonising after fire. By six weeks Bioregions: Warren; Avon Wheatbelt post-fire, Australian Magpie, Grey Shrike-thrush, Repeated burning of heaths had severe impact on Crimson Rosella, Pied Currawong, Grey Rufous Bristlebird. Currawong, Australian Raven, Welcome Swallow, Masked Lapwing, Sulphur-crested Cockatoo and Yellow-tailed Black Cockatoo had been recorded. Catling, P. C., & Newsome, A. E. (1981). Within weeks after that, first returns of Responses of the Australian vertebrate fauna to White-eared Honeyeater, Eastern Spinebill and fire. In A. M. Gill, R. H. Groves, & I. R. Noble Flame Robin. (Eds.), Fire and the Australian biota (pp. 273-310). Canberra: Australian Academy of Science. Chandler, L. G. (1973). In the wake of a review bushfire. Wildlife in Australia, 10, 140-141. Considers a series of propositions that the wildfire; mallee Australian vertebrate fauna is fire-adapted. In Bioregion: Murray Darling Depression eucalypt forests, the greatest bird diversity occurs General notes, suggesting high mortality of small in forests 5-6 years post-fire then mature forests birds during mallee wildfire. Malleefowl may be declines as habitat. There is little evidence of particularly disadvantaged, with eggs lost, young distinct seral stages, but simply gradual changes in killed and regrowth vegetation unsuitable. Effects abundance. All species in eucalypt forests have the may be accentuated if drought follows fire. capacity to survive fires and recover quickly. There are very few fire specialists. It is difficult to ascribe life history traits to adaptation to fire as opposed to

127 Australia’s Biodiveristy - Responses to Fire

Chatto, R. (1995). The effects of fire on a Chisholm, A. H. (1922). The “lost” Paradise breeding colony of Australian Pelicans. Parrot. Emu, 22, 4-17. Corella, 19, 70. Paradise Parrot; threatened species; eucalypt woodland. Australian Pelican; mortality The Paradise Parrot may have declined (now to Bioregion: Top End Coastal extinction) over the last century because grazing In two of the last four years, fire burnt through an and frequent firing led to loss of its seed resource. island colony of Australian Pelicans killing at least 1000 young birds (=most of the population). Fires were probably deliberately lit. Chisholm, A. H. (1945). Birds of the Gilbert diary, Part 2. Emu, 44, 183-200. Paradise Parrot; eucalypt woodland; threatened species. Cheal, P. D., Day, J. C., & Meredith, C. W. Altered fire regimes since European settlement (1979). Fire in the national parks of north-west may have contributed to the extinction of the Victoria. National Parks Service. Paradise Parrot. threatened species; mallee; heath; management; hollows Bioregion: Murray-Darling Depression Bird communities were censused in a range of Christensen, P. (1974). The concept of fauna vegetation types (mallee and heath) across a range priority areas. In Third Fire Ecology Symposium of regrowth ages. “After a fire in mallee vegetation, (pp. 66-73). Melbourne: Forests Commission, virtually no birds will breed in the burnt area in Victoria. the first three to four years”. Some birds (mainly eucalypt open forest; wildfire granivores) will forage in such burnt areas, but Bioregion: Warren generally at low abundance. When regrowth trees Bird abundance and richness increases sharply in begin to emerge, bird diversity and densities Karri forest up to at least 2 years after a very hot increase rapidly (to peak at about 15 years burn. Some species of open formations (e.g. Scarlet post-fire, though richness may continue to Robin) can invade wetter forests after hot fires. increase). No bird species appears to be restricted Other species (e.g. Red-winged Fairy-wren) decline to young regrowth, but several are restricted to but then increase within 2 years. mallee of >15-20 years post-fire. These include hollow-nesting species (e.g. Striated Pardalote, Christensen, P., & Abbott, I. (1989). Impact of parrots), some canopy species and many fire in the eucalypt forest ecosystem of ground-foraging species. Most of the “typical” southern Western Australia: a critical review. mallee species favour old vegetation. However Australian Forestry, 52, 103-121. very old mallee may become unsuitable, even for review; community; management; eucalypt open forest; these species. Malleefowl may prefer vegetation succession around 15-25 years post-fire, as food resources Bioregions: Jarrah Forest; Warren. probably decline with increasing time since fire. Current periods between fuel-reduction burns are In heaths, few species occur in very young 5-6 years for Jarrah forest and 7-9 years for Karri regrowth, richness probably peaks at c15-20 years forest, other than in conservation reserves. Limited post-fire, and older heaths tend to be evidence suggests that Aboriginal burning in these species-poor. forests may have been more frequent (3-5 year Predation on birds may be high in the cycles) or much less frequent. Reviews research on immediate aftermath of fire. In general, old mallee effects of fire on soil nutrients, floristics, structure, had more stable bird assemblages (greater invertebrates and vertebrates. Most studies report similarity in species composition between an initial short-term decrease in understorey birds geographically separated sites compared to followed by an increase above pre-fire levels younger regrowth), and their bird assemblages following fuel-reduction burns in Jarrah and Karri included more insectivores and greater forests, with little change in canopy species. Some stratification of foraging zones. The restriction of species (e.g. White-winged Chough, Western most mallee specialists to old vegetation (and the Yellow Robin, Scarlet Robin) invaded or increased generalist nature of transient species occurring in in abundance in recently-burnt areas. Honeyeaters young regrowth) argues for a relative stability of also became more common in association with old mallee and the adaptation of most mallee birds to long intervals between fire.

128 Fire and Australian birds

fire-induced flowering. Fire intensity is the major Christensen, P., Recher, H., & Hoare, J. factor influencing the impact of fire on bird (1981). Responses of open forests (dry populations. sclerophyll forests) to fire regimes. In A. M. Gill, R. H. Groves, & I. R. Noble (Eds.), Fire and the Australian biota (pp. 367-393). Christensen, P. E., & Kimber, P. C. (1975). Canberra: Australian Academy of Science. Effect of prescribed burning on the flora and review; eucalypt open forest; widlfire; mortality fauna of south-west Australian forests. In very hot wildfires, bird mortality in fires may be Proceedings of the Ecological Society of Australia, high. With cooler fires, mortality is usually low, 7, 85-107. however post-fire mortality may be substantial. review; eucalypt open forest; control burning; Noisy Scrub-bird requires a late seral stage and community; succession cannot survive frequent fires. The nectarivorous Bioregions: Jarrah Forest; Warren Crescent Honeyeater may disappear from heaths Birds showed surprisingly small changes following for several years after fire. Post-fire changes in cool fire. Birds were sampled at two sites (one forests are most pronounced for birds of ground unburnt for 40 years and one subject to an intense and understorey layers. In general these birds prescribed fire) in dry sclerophyll forest. There was decline in abundance after fire, but then increase to little difference in the bird communities before fire. at least pre-fire levels within 2-3 years post-fire From one month to one year post-fire there was a Recolonisation and increased populations may be very slight decline of ground and understorey due to greater insect numbers on regrowth species. By 2 years post-fire the population of vegetation. understorey and ground-dwelling species (e.g. Rufous Treecreeper, Inland Thornbill, Western Thornbill, Western Yellow Robin, Golden Christensen, P. E. S., Wardell-Johnson, G., & Whistler) had surpassed pre-fire levels, and that of Kimber, P. (1985). Birds and fire in canopy species was unchanged. Following a fire in southwestern forests. In A. Keast, wet sclerophyll forest, bird numbers decreased H. F. Recher, H. Ford, & D. Saunders (Eds.), initially, but by 5 months post-fire abundance was Birds of eucalypt forests and woodlands: ecology, higher than pre-fire (due especially to more conservation, management (pp. 291-299). Western Yellow Robin, Inland Thornbill, Scarlet Chipping Norton: Surrey Beatty. Robin, Rufous Treecreeper and Grey eucalypt open forest; management; community; control Shrike-thrush). In the second year after fire there burning; wildfire; succession was an increase in birds of the shrub layer Bioregion: Jarrah Forest (particularly of Inland Thornbill and White-browed Reports two studies of birds and fire in Jarrah Scrubwren), and of White-naped Honeyeater forests. A hot prescribed fire burnt the site in the (which foraged in epicormic growth). Some species third year of a 5 year study (a control site remained (e.g. Red-winged Fairy-wren and Golden Whistler) unburnt). Bird species richness increased in the disappeared for the first year post-fire, but returned burnt site. 13 species (including White-winged in 2-3 years post-fire. White-browed Scrubwren Triller, Australian Magpie, Australian Raven) changed foraging behaviour after fire. Extensive hot appeared after fire that had not been recorded fires or frequent cool fires would produce a before, and others (e.g. Dusky Wood-swallow, Tree homogenisation of the landscape to the detriment Martin, Scarlet Robin, Western Yellow Robin, of bird species diversity. Inland Thornbill, Western Thornbill) became more abundant (for varying periods post-fire). Some nectarivorous species (Purple-crowned Lorikeet, Silvereye, Red Wattlebird, ) increased post-fire, because of fire-induced flowering in the eucalypt overstorey. Canopy-feeding insectivores showed no or little effect of fire. Golden Whistler, White-breasted Robin and White-browed Scrub-wren declined after fire (the latter two did not reappear until 2 years post-fire).

129 Australia’s Biodiveristy - Responses to Fire

A second study considered long-term impacts Coate, K. (1985). Black Honeyeaters feeding of fire regimes, comparing bird assemblages in on ash. Western Australian Naturalist, 16, 3 areas with different fire regimes (unburnt for 51-52. 14 years, burnt in a severe wildfire 1 year before charcoal; Black Honeyeater study, and burnt in a mild fire 6 months before Up to 4 Black Honeyeaters were feeding study). The unburnt and mildly burnt sites had simultaneously in old fireplaces, consuming ash. more species and individuals than the site which Analysis showed this had high levels of calcium, had been burnt by wildfire. Golden Whistler, and the behaviour may be explained by female Inland Thornbill, Western Yellow Robin and birds seeking calcium before egg-laying. White-breasted Robin were most common in the unburnt forest (and least common in the wildfire Cooper, R. P. (1972). The occurrence of the site). Scarlet Robin was least common in the Pilot-bird on Wilson’s Promontory. Australian wildfire site. Western Thornbill was least common Bird Watcher, 4, 137-143. in the unburnt site. The abundance of small Pilot-bird; eucalypt open forest; wildfire insectivores of the canopy showed little variation. Bioregion: South east Coastal Plain Rapid response of birds to mild fire From being common last century in densely compared to more substantial impact of hot vegetated gullies of the Strzelcki Ranges, the wildfire may be due to a delayed recovery of Pilot-bird has now been largely wiped out of the invertebrates in the latter. As a general rule, the area by clearing and wildfire, particularly the effect of mild or moderate fires on birds is disastrous fires of 1939. inversely proportional to their main foraging height. However responses of species are somewhat idiosyncratic, and the individual Cooper, R. P. (1974). The avifauna of Wilson’s responses of rare species in particular need to be Promontory. Part 1. Australian Bird Watcher, considered. 5, 137-174. Fire intensity is the major factor determining heath; eucalypt open forest; management impact of fire on birds. Season of burn may be Bioregion: South east Coastal Plain relatively unimportant, other than through its Substantial changes to vegetation (increased impact on intensity: this view is contrary to scrubbiness) of Wilson’s Promontory over the last popular belief that spring burning is especially century were due to changes in fire regime. This detrimental due to its alleged disruption of must have changed bird species composition. A breeding. A diverse burning pattern may be the very large wildfire in 1951 (during the breeding most appropriate management. season) probably resulted in massive mortality of birds. Argues against fuel reduction burning and burning of heathlands. “Fire has a most Clout, M. N. (1989). Foraging behaviour of detrimental effect on the avifauna of any area.” Glossy Black-cockatoos. Australian Wildlife Research, 16, 467-473. Glossy Black-cockatoo; Allocasuarina woodland; Cooper, R. P. (1975). The avifauna of Wilson’s hollows; eucalypt open forest; threatened species; Promontory. Part 4. Australian Bird-watcher, management 6, 17-34. Bioregion: South eastern Corner heath; Ground Parrot; threatened species Glossy Black-cockatoos are reliant on seeds of Bioregion: South east Coastal Plain Allocasuarina. These species are fire-sensitive, Ground Parrots reported to flee just ahead of fire requiring heat to open cones, but are easily killed front. The species disappeared from an area for at by intense fire. Post-European changes in fire least four years post-fire. regime have reduced abundance of Allocasuarina and the large dead trees in which Glossy Black-cockatoos nest. Conservation of this species depends upon appropriate fire management.

130 Fire and Australian birds

Cowley, R. D. (1971). Birds and forest Cowley, R. D., Heislers, A., & Ealey, E. H. H. management. Australian Forestry, 35, 234-250. (1969). Effects of fire on wildlife. Victoria’s review; eucalypt open forest; management; forestry Resources, 11, 18-22. In eucalypt forests, birds which feed or nest on the mallee; Malleefowl; Superb Lyrebird; eucalypt open ground (e.g. Superb Fairy-wren, White-browed forest; review Scrubwren, Spotted Quail-thrush, Brown Malleefowl require abundant litter for nesting. Thornbill and White-eared Honeyeater) may be This may not be available for at least 10 years those most affected by fire. Birds which nest within post-fire. In contrast, Superb Lyrebird may 1.5m of the ground may be vulnerable to require periodic fires to remove wire grass. fuel-reduction fire during the breeding season. Food (e.g. seed) is generally more available after a fire. Cool burns generally leave unburnt patches, Crawford, D. N. (1972). Birds of Darwin area, typically in gullies, and regrowth of grasses and with some records from other parts of shrubs is usually rapid after fire. Frequent fires can Northern Territory. Emu, 72, 131-148. change shrubby understorey to grass, to the tropical eucalypt open forest; tussock grassland; advantage of species such as Spotted Quail-thrush hummock grassland. and Buff-rumped Thornbill but to the disadvantage Bioregions: Top End Coastal; Pine Creek Arnhem of species associated with shrubs (e.g. Burning of grasslands during the dry season leads White-browed Scrubwren, Brown Thornbill, to local shifts in habitat use by Golden-headed Common Bronzewing). Frequent fires in wetter Cisticola and Red-backed Fairy-wren (both may forest may result in a dense cover of wire grass, to occur in monsoon rainforest fringes when most of the detriment of species foraging in litter (such as the open forests and grasslands have been burnt). Superb Lyrebird, Bassian Thrush, Eastern Yellow Barn Owl concentrates in cleared areas early in Robin, Superb Fairy-wren). Some honeyeaters are the dry but disperses to extensive burnt areas later advantaged by frequent light fires, because these in the season (presumably because hunting over promote profuse flowering. unburnt grasslands is difficult). In contrast, there does not appear to be a concentration of Pheasant Coucals in unburnt patches of grassland later in Cowley, R. D. (1974). Effects of prescribed the season (possibly suggesting seasonal burning on birds of the mixed species forests movement). White-throated Grass-wren recorded of West Central Victoria. In Third Fire Ecology only in long-unburnt hummock grassland. Symposium (pp. 58-65). Melbourne: Forests Commission, Victoria. eucalypt open forest; control burning; community Crawford, D. N. (1979). Effects of grass and Bioregion: Victorian Midlands fires on birds in the Darwin area, Northern Short-term effects of a (fairly hot) prescribed burn Territory. Emu, 79, 150-152. in open forest were examined, partly through tropical eucalypt open forest. observation of changed distribution of banded Bioregion: Top End Coastal birds. Of 27 banded birds (of 7 species: Monitored bird populations in 24 sites (1.6ha) in White-browed Scrubwren, Superb Fairy-wren, wet and dry seasons over 28 months. All sites were Straited Thornbill, Brown Thornbill, burnt at some stage over the study period. White-throated Treecreeper, Eastern Yellow Red-backed Fairy-wren was disadvantaged by fire, Robin, White-eared Honeyeater) regularly using and its persistence appears to depend on some the site pre-fire, 18 were known to have survived cover remaining after fire. Within 2 weeks of (all species other than White-eared Honeyeater). burning, there is an increase in migratory ground- Home ranges changed little. Birds appeared more and mid-level foragers, and Pied Butcherbird, wary after fire (and logs were used as refuge). Black-faced Cuckoo-shrike, Magpie-lark and Several species with nests started before the fire Red-tailed Black-cockatoo. Their numbers successfully raised young after the fire. Only one declined slightly at >2 months post-fire. species (Olive Whistler, which prefers dense understorey) apparently disappeared. Several species characteristic of open areas (Australian Magpie, Buff-rumped Thornbill, Painted Quail) invaded after fire.

131 Australia’s Biodiveristy - Responses to Fire

Crowley, G. M. (1995). Fire on Cape York The relative importance of fire as a factor Peninsula. Cape York Peninsula Land Use affecting bird conservation is discussed for every Study. Queensland bioregion. Existing fire regimes may review; tropical eucalypt open forest; rainforest; be a management problem for birds in Mitchell hollows; Cassowary; Golden-shouldered Parrot; Grass Downs, , Einasleigh threatened species; management Uplands, Cape York Peninsula, Wet Tropics, but Bioregion: Cape York Peninsula are not a problem (or there is insufficient Birds that are advantaged by fire in the information) for other bioregions. short-term, or by a high frequency regime, tend to be nomadic granivores, carnivores or omnivores. Of particular concern is the effect of Curry, G. N. (1991). The influence of fires on tree hollows used for nesting by birds. proximity to plantation edge on diversity and Fires may lead to hollow formation, or to abundance of bird species in an exotic pine destruction of hollows and nests. Ecotone plantation in north-eastern New South Wales. between rainforest and wet eucalypt forest may be Wildlife Research, 18, 299-314. the preferred habitat for Cassowary, and this management; slash-burn; forestry species may require fire for its maintenance. Windrows in plantations allow the entry or Widespread late dry season fires have probably led persistence of many bird species. These are to the decline of the Brown Treecreeper. Lack of usually burnt to reduce fire risk to the plantation, burning after the first storms of the wet season is but it would be preferable (for bird conservation) considered a threat to the Golden-shouldered not to burn these windrows, but rather to reduce Parrot. A wide variety of burning histories is likely fire hazard by careful positioning of windrows. to maximise bird diversity. Curry, P. J. (1986). Habitat characteristics of CSIRO Wildlife Research (1976). A survey of the Thick-billed Grasswren Amytornis textilis the fauna of the Lower McArthur River Region, in grazed shrublands in Western Australia. In Northern Territory. Mimets Development Pty P. J. Joss, P. W. Lynch, & O. B. Williams Ltd. (Eds.), Rangelands: a resource under siege (pp. tropical eucalypt open forest; tropical eucalypt savanna 566). Canberra: Australian Academy of woodland; hummock grassland; tussock grassland; Science. Carpentarian Grass-wren; Bush Stone-curlew; Thick-billed Grasswren; chenopod shrubland; Acacia Australian Bustard. shrubland Bioregions: Gulf Falls and Uplands; Gulf Coastal Bioregion: Carnarvon Increased frequency of fire, particularly in The western subspecies of Thick-billed hummock grasslands, must have a deleterious Grasswren had previously been considered to be effect on the survival of such terrestrial species as possibly on the verge of extinction and reliant on Carpentarian Grass-wren, quail, Bush saltbush and bluebush. Instead, the subspecies was Stone-curlew and perhaps Bustard. found to be widespread in seral shrublands that replace Acacia shrublands for at least 40 years after wildfire. Cummings, B., McDonald, B., & Taplin, A. (1993). Knowledge of birds in Queensland biogeographic regions and threats to their Danks, A. (1991). The role of corridors in the conservation. In C. P. Catterall, P. V. Driscoll, management of an endangered passerine. In K. Hulsman, D. Muir, & A. Taplin (Eds.), Birds D. A. Saunders & R. J. Hobbs (Eds.), Nature and their habitats: status and conservation in conservation 2: the role of corridors (pp. Queensland (pp. 178-186). St Lucia: 291-296). Chipping Norton: Surrey Beatty. Queensland Ornithological Society Inc. Noisy Scrub-bird; eucalypt open forest; thicket; review corridors; management; threatened species. Bioregions: Mitchell Grass Downs, South East Bioregion: Jarrah Forest Queensland, , Cape York Dispersal of Noisy Scrub-bird is assisted by Peninsula, Wet Tropics, Channel Country, Mulga corridors which are unburnt (and protected from Lands, Brigalow Belt, , Gulf other disturbance). Plains, Desert Uplands, Central Mackay Coast. 132 Fire and Australian birds

Davidson, I., & Chambers, L. (1991). Dedman, V. (1983a). G.F.N.C. Otway Vegetation management for Superb Parrot Regeneration Survey - Progress report. foraging habitat in Victoria. Victorian Geelong Naturalist, 20, 52-56. Department of Conservation and eucalypt open forest; heath; wildfire Environment. Bioregion: South east Coastal Plain Superb Parrot; threatened species; management By six months after hot fire in heath and eucalypt Bioregion: Riverina open forest, birds were returning slowly. Larger Some corridors used by Superb Parrots may be species (Australian Raven, , Laughing enhanced by the cessation of burning; landowners Kookaburra, Sulphur-crested Cockatoo) are may be reluctant to increase or maintain suitable conspicuous. Smaller birds were not seen habitat patches because of their perception that regularly until bracken was high enough to this increases the risk of wildfire. provide cover, but by 6 months post-fire Brown and Striated Thornbills were being seen regularly. Davidson, I., & Robinson, D. (1992). Welcome Swallows were common over the burnt Grey-crowned Babbler Pomatostomus temporalis. heathland. Department of Conservation and Natural Resources. Dedman, V. (1983b). G.F.N.C. Otways Grey-crowned Babbler; eucalypt woodland; threatened regeneration survey - 2nd progress report. species; management; control burning Geelong Naturalist, 20, 98-100. Bioregions: Victorian Midlands; South Eastern eucalypt open forest; wildfire Highlands Bioregion: South east Coastal Plain Reviews information on this species. Threats to Almost one year after hot fire, bird numbers and declining Grey-crowned Babblers include fire species were increasing. Honeyeaters were protection works (e.g. firebreaks and fuel attracted to flowering Xanthorrhoea. Superb reduction burning) and collection of firewood. Fairy-wrens were foraging in regrowth bracken. Some species were probably breeding. Pied Davies, S. J. J. F., Smith, G. T., & Robinson, F. Currawongs, Sulphur-crested Cockatoos and Red N. (1982). The Noisy Scrubbird in Western Wattlebirds were noted frequently. Australia. In R. H. Groves & W. D. L. Ride (Eds.), Species at risk: research in Australia (pp. Dedman, V. (1983c). Mammals and birds after 117-127). Canberra: Australian Academy of the fires: Forest Road, Anglesea. Geelong Science. Naturalist, 20, 27. Noisy Scrub-bird; threatened species; heath; eucalypt eucalypt open forest; wildfire open forest; management. Bioregion: South east Coastal Plain Bioregion: Jarrah Forest List of 13 bird species observed in severely burnt The Noisy Scrub-bird disappeared from areas eucalypt open forest two weeks after fire. where eucalypt forests fringing swamps had been burnt. With strict fire control populations have built up. Dedman, V. (1983d). Moggs Creek: eight days later. Geelong Naturalist, 20, 22-25. eucalypt open forest; wildfire Debus, S. J. S., & Czechura, G. V. (1988). The Bioregion: South east Coastal Plain Red Goshawk Erythrotriorchis radiatus: a Eight days following severe fire in eucalypt open review. Australian Bird Watcher, 12, 175-199. forest, few birds were seen. These included Red Goshawk; historical change; threatened species; Crimson Rosella, Red Wattlebird, tropical eucalypt open forest White-throated Treecreeper and Scarlet Robin. Since the , Australia’s rich raptor community has declined, possibly because of climate change and perhaps Aboriginal fire regimes. Currently, the most serious threats to the Red Goshawk are mismanagement of tropical rangelands (overstocking and too frequent burning), vegetation destruction and wetland drainage. 133 Australia’s Biodiveristy - Responses to Fire

Dedman, V. (1983e). Notes on the effects of Disney, H. J. d. S. (1968). Bushfires and their the Ash Wednesday fire on the Ironbark effect on fauna and flora. Australian Natural Basin, Point Addis. Geelong Naturalist, 20, History, 16, 87-89. 25-26. review eucalypt open forest; wildfire Banded birds returned to their pre-burn Bioregion: South east Coastal Plain territories following fire. Backburns may be Bird lists from 3 weeks and 6 weeks after hot fire particularly dangerous to small birds. in eucalypt open forest. Superb Fairy-wrens were observed in scorched tree-tops. Commonest birds in the burnt area were Eastern Yellow Robin and Du Guesclin, P., Smith, S., O’Shea, B., & White-throated Treecreeper. Other species Debbis, C. (1995). “Brushing for bristles”: present included White-browed Scrubwren, habitat corridors for the Rufous Bristlebird. In Crimson Rosella, Brown Thornbill and Striated A. Bennett, G. Backhouse, & T. Clark (Eds.), Thornbill. People and nature conservation: perspectives on private land use and endangered species recovery (pp. 163-165). Chipping Norton: Surrey Dedman, V. (1984). Otways regeneration Beatty. survey. Third progress report. Geelong Rufous Bristlebird; heath; corridors; isolate Naturalist, 21, 89-92. Bioregion: South East Coastal Plain wildfire; eucalypt open forest Recolonisation after fire is a problem for Rufous Bioregion: South east Coastal Plain Bristlebirds because of their poor dispersal ability. One year after a hot wildfire, large birds Corridors may provide effective means for (Yellow-tailed Black-cockatoo, Crimson Rosella, recolonisation. Red Wattlebird, Pied Currawong) appear more obvious in burnt forest. Emison, W. B., & Bren, W. M. (1989). Common birds of the mallee, northwestern Dickinson, K. J. M., Wall, L. E., & Wilson, R. Victoria. In J. C. Noble & R. A. Bradstock I. (1986). Birds in a partly clearfelled dry (Eds.), Mediterranean landscapes in Australia: eucalypt forest on dolerite in southeastern mallee ecosystems and their management (pp. Tasmania. Papers and Proceedings of the Royal 221-242). Melbourne: CSIRO. Society of Tasmania, 120, 39-49. mallee; heath; eucalypt woodland; Callitris woodland; eucalypt open forest; slash-burn; forestry; community; Casuarina woodland; hollows; management management Bioregion: Murray-Darling Depression Bioregion: Freycinet Many species are dependent on hollows in Bird assemblages were sampled in unlogged woodlands, and the maintenance of these requires forest, forest clearfelled and slash burnt, and active management (including of fire regimes). In forests clearfelled without subsequent burning. the previous decade, wildfires have burnt much of Species typical of open areas (e.g. Blue-winged the mallee shrublands such that long-unburnt Parrot, Superb Fairy-wren) invaded the cut and patches are now very rare. Species associated with burnt sites. Flame Robin, Dusky Robin, Superb such habitat include those nesting in hollows, and Fairy-wren and Grey Shrike-thrush were more those which require dense litter. The management common in the slash-burnt area than the unburnt of mallee for fauna probably requires maintenance logged forest: Crescent Honeyeater and Eastern of a mosaic of vegetation of a range of ages, but Spinebill were more common in the unburnt cut with a bias towards retention of older ages. forest. Prescribed burning under regenerating forests may disadvantage some species which depend on ground habitats (e.g. Spotted Quail-thrush).

134 Fire and Australian birds

Emison, W. B., Beardsell, C. M., Norman, F. Fitzherbert, J. C., & Baker-Gabb, D. J. (1988). I., & Loyn, R. H. (1987). Atlas of Victorian Australasian grasslands and their threatened birds. Melbourne: Department of avifauna. In P. D. Goriup (Ed.), Ecology and Conservation, Forests and Lands, and Royal conservation of grassland birds (pp. 227-250). Australasian Ornithologists Union. Cambridge: International Council for Bird review Preservation. Describes habitat and distribution for Victorian review; tussock grasslands; hummock grasslands; birds. The Mallee Emu-wren attains highest tropical eucalypt savanna woodland; management; densities in early years after fire. The Southern threatened species; Gouldian Finch; Yellow-rumped Emu-wren is “able temporarily to colonise Mannikin; Chestnut-backed Button-quail; Partridge ephemeral habitats such as … scrubs regenerating Pigeon; Golden-shouldered Parrot; Hooded Parrot; after fire.” Carpentarian Grass-wren; White-throated Grass-wren; Black Grass-wren Carpentarian, White-throated and Black Ferrier, S. (1985). Habitat requirements of a Grass-wrens are threatened by fire in hummock rare species, the Rufous Scrub-bird. In A. grasslands, and require aging (>6 years) Triodia. Keast, H. F. Recher, H. Ford, & D. Saunders In tropical grasslands, fire frequency has increased (Eds.), Birds of eucalypt forests and woodlands: since European settlement. “Large areas are burnt ecology, conservation, management (pp. each year … The decline of six bird species - the 241-248). Chipping Norton: Surrey Beatty. Gouldian Finch, Yellow-rumped Mannikin, Rufous Scrub-bird; management; eucalypt open forest; Chestnut-backed Button-quail, Partridge Pigeon, rainforest; threatened species. Golden-shouldered Parrot and Hooded Parrot - is Bioregion: NSW North Coast probably related to the changed fire regime and Fire has played a major role in the Rufous associated grazing pressure.” After habitat Scrub-bird’s current distribution and, in destruction (by over-grazing and conversion of particular, its close association with rainforest. grasslands to agriculture) “fire is undoubtedly the High intensity wildfires are known to have next most significant threat”. In tropical savannas, destroyed territories. Prescribed burning probably fire effects are poorly understood, but changed has little direct effect (though habitat suitability phenology or floristics have probably affected was low 10 months after a prescribed fire), and some birds. In some central and southern may prevent destruction through wildfire (though grasslands, a decrease in fire frequency (but most are naturally buffered by their dampness or increase in intensity) “has had an equally proximity to rainforests). However long-term destructive effect”. influence of prescribed burning may be complex. Suitable habitat in eucalypt forests is probably associated with a particular post-fire seral stage. Fitzherbert, K., McLaughlin, J., & With long absence of fire, the eucalypt forests Baker-Gabb, D. (1992). Black-eared Miner may develop into rainforests, to the detriment of Manorina melanotis. Department of Rufous Scrub-birds (due to reduction in ground Conservation and Environment. cover density). Black-eared Miner; threatened species; mallee; management; wildfire. Bioregion: Murray-Darling Depression Reviews information on the species. Black-eared Miners inhabit areas that have not been burnt for more than 60 years, a limited and declining habitat. Suppression of wildfires and reduction in fire frequency will assist this species.

135 Australia’s Biodiveristy - Responses to Fire

Ford, H. A. (1989). Ecology of birds: an Garnett, S. (1987). Seabird Islands: No. 171. Australian perspective. Chipping Norton: Kusamet Island, Torres Strait, Queensland. Surrey Beatty. Corella, 11, 77-78. review; eucalypt open forest; Ground Parrot; heath seabirds; tussock grassland; thicket; mortality Includes a brief review of the effect of fires in Bioregion: Cape York Peninsula eucalypt forest. Changes are not dramatic and, The grass on the island is often burnt during except after severe or extensive fire, the bird October and November which is likely to destroy assemblage will have recovered its original nests and chicks of the breeding Bridled and composition within 1-2 years post-fire. Ground Black-naped Terns. Parrot cannot survive in habitat where fires occur at intervals of less than every few years or greater than 20 years. Garnett, S. (1992a). The action plan for Australian birds. Canberra: Australian National Parks and Wildlife Service. Forshaw, J. M. (1981). Australian Parrots. review; threatened species; management Melbourne: Lansdowne Press. Describes the former and current threats to all review; Golden-shouldered Parrot; Ground Parrot; threatened Australian birds, and research and Scarlet-chested Parrot; tropical eucalypt open forest; management requirements. Altered fire regimes are heath; mallee; hummock grassland; threatened species currently threatening 51 taxa (16 confirmed, 35 Fire regime is contributing to the decline of speculative). Only one other threatening process is Golden-shouldered Parrot. The Ground Parrot is affecting more taxa (habitat clearance and dependent on a regime of mosaic burning at 8-10 fragmentation which affects 54 taxa). year intervals. Scarlet-chested Parrot is associated with recently-burnt mallee with hummock grass. Garnett, S. (1992b). Threatened and extinct birds of Australia. Melbourne: Royal Fowler, S. (1945). The fire on Mondrain Australasian Ornithologists Union. Island. Emu, 44, 334-335. review; threatened species; management seabirds; tussock grassland; heath; mortality Reviews (including previously unpublished Bioregion: Esperance Plains information) the status of Australian threatened Deliberately lit fire burnt 60% of Mondrain birds. Altered fire regimes are a current confirmed Island, leading to seabird deaths. or speculated threat to 51 Australian bird taxa, second only to clearing and fragmentation (affecting 52 taxa). The most serious fire effects Fox, A. (1978). The ’72 fire of Nadgee Nature are increase in frequency of late dry season fires in Reserve. Parks & Wildlife, 2, 5-24. tropical woodlands and all fires in the mallee. heath; eucalypt open forest; mortality; wildfire Discusses the threat of fire (or relationship with Bioregion: South East Corner fire) for Kangaroo Island Emu, Australasian After fire, 609 birds were found dead on beach. Bittern (reports desertion of a swamp for 2 years Most (89%) came from a backburn area rather after burning), Red Goshawk, MalleeFowl than from the original fire. 49 bird species were (habitat should be protected against too frequent recorded as killed: the most common dead birds burning), Buff-breasted Button-quail (where late were Little Wattlebird and dry season fires during the nesting season are Honeyeater. regarded as the major threat), Chestnut-backed Button-quail, Painted Button-quail, Black-breasted Button-quail (control burns on Frith, H. J. (1962). Conservation of the rainforest margins at intervals of less than 4 years Mallee-fowl (Leipoa ocellata, Gould). CSIRO render the habitat unsuitable), Lewin’s Rail, Wildlife Research, 7, 33-49. Partridge Pigeon (early dry season fires may Malleefowl; threatened species; management; mallee. destroy the ground nests and eggs), Red-cheeked Bioregion: Murray Darling Depression Parrot (main threat is fire in the late dry season Malleefowl requires adequate litter for breeding. which may destroy nest trees), Eclectus Parrot This is not available until at least 10-15 years (nests have been destroyed by management fires), post-fire. Golden-shouldered Parrot, Partridge Pigeon,

136 Fire and Australian birds

Orange-bellied Parrot, Scarlet-chested Parrot, Bioregion: Cape York Peninsula Ground Parrot (fire continues to be the main Fire regimes determine boundaries (and relative threat), Night Parrot, Palm Cockatoo (the hollow extent) of grassland flats and Melaleuca trees used for nesting are particularly vulnerable woodlands. Use of fire to control Melaleuca and to fire, and many have been destroyed by recent early wet season burning (to increase food fires, including destruction of eggs and young), resources) are required to conserve Red-tailed Black-cockatoo, Masked Owl, Sooty Golden-shouldered Parrots. Owl, Rufous Owl (the hollow trees used for nesting are particularly vulnerable to fire), Rufous Scrub-bird, Noisy Scrub-bird, Purple-crowned Garnett, S., & Crowley, G. (1995a). The Fairy-wren, Southern Emu-wren (where decline of the Black Treecreeper Climacteris fragmentation exacerbates the threat of fire due to picumnus melanota on Cape York Peninsula. decreased possibility of recolonisation), Mallee Emu, 95, 66-68. Emu-wren, Striated Grass-wren (the major threat historical change; tropical eucalypt open forest; Black is extensive fire), Carpentarian Grass-wren, Treecreeper; management Black-eared Miner, Helmeted Honeyeater, Bioregion: Cape York Peninsula Forty-spotted Pardalote, Rufous Bristlebird, Marked decline in Black Treecreeper on Cape Eastern Bristlebird (the main threat to the species York Peninsula over the period 1920 to 1995. appears to be change in fire regime), Western Principal reason thought to be change in fire Bristlebird (fire is the main threat), Brown regime, with larger hotter fires now. Remnant Thornbill, Slender-billed Thornbill, Northern populations are now mainly in areas with natural Scrub-robin (possible extinction of the NT fire breaks (rivers etc). Conservation of the species subspecies has been attributed to changed fire depends upon reintroduction of fine-grained regime), Western Whipbird (fire appears to be the mosaic burning. main threat), Crested Shrike-tit, Red-lored Whistler (in remnant habitat the effects of fire Garnett, S., & Crowley, G. (1995b). Feeding may be critical), Bassian Thrush and Zitting ecology of Hooded Parrots Psephotus dissimilis Cisticola. during the early wet season. Emu, 95, 54-61. Hooded Parrot; tropical eucalypt open forest; tropical Garnett, S., & Bredl, R. (1985). Birds in the eucalypt savanna woodland vicinity of Edward River Settlement. Part II. Bioregion: Pine Creek Arnhem Discussion, references, list of passerines. Detailed foraging study of Hooded Parrots. Fire Sunbird, 15, 25-40. and grazing may affect plant species composition tropical eucalypt open forest; tussock grassland; historic and hence food availability for this species. change; Star Finch; community; threatened species Bioregion: Cape York Peninsula Garnett, S., Crowley, G., Duncan, R., Baker, “Hunting, introduction of plants and animals, and N., & Doherty, P. (1993). Notes on live Night burning are the effects of people most likely to be Parrot sightings in north-western detrimental to the birds of the Edward River Queensland. Emu, 93, 292-296. area.” Grass around the breeding lagoon Night Parrot; tussock grassland; hummock grassland; is never burnt, and Star Finch is now restricted to Spinifex-bird; Rufous-crowned Emu-wren; threatened this area, suggesting that the prevailing fire species regime in the region has become unsuitable for Bioregion: Mt Isa Inlier this species. Night Parrot may be present in this area due to favourable fire management (small control burns Garnett, S. T., & Crowley, G. M. (1994). The leading to seral mosaic and preventing extensive ecology and conservation of the wildfire). Rufous-crowned Emu-wrens and Golden-shouldered Parrot. Cape York Peninsula Spinifex-birds are also favoured by this Land Use Strategy. management, as both prefer spinifex which has Golden-shouldered Parrot; threatened species; remained unburnt for long periods. management; tropical eucalypt savanna woodlands; tussock grasslands; historic change; control burning

137 Australia’s Biodiveristy - Responses to Fire

Gibbons, P. (1994). Sustaining key old-growth different responses of shrubs and sedges to fire. characteristics in native forests used for wood Differences between Queensland and Victoria in production: retention of trees with hollows. In preferred ages may be due to faster processes in T. W. Norton & S. R. Dovers (Eds.), Ecology the north. Hence, it is misleading to extrapolate and sustainability of southern temperate results to different areas or habitats. ecosystems (pp. 59-84). Canberra: CSIRO. forestry; old-growth; hollows; management; review; eucalypt open forest Green, R. H. (1982). The activity and Reviews use of hollows by birds (and other fauna); movement of fauna in compartment 2, Maggs hollow-formation; relationship between fire (and Mountain, Tasmania, in the first five years of forestry operations) and hollow formation. forest regeneration. Records of the Queen Victoria Museum, 75, 1-31. forestry; eucalypt open forest; succession; control Gibson, D. F. (1986). A biological survey of the burning Tanami Desert in the Northern Territory. Bioregion: Ben Lomond Conservation Commission of the Northern Mainly concerned with recovery of bird Territory. populations after intensive logging, but considers hummock grassland; Little Button-quail; Richard’s interaction with burning. Pipit; Rufous-crowned Emu-wren Bioregion: Tanami Little Button-quail and Richard’s Pipit occur Green, R. H., & McGarvie, A. M. (1971). The mostly in recently burnt country. Rufous-crowned birds of King Island. Records of the Queen Emu-wren is particularly common in Victoria Museum, 40, 1-42. long-unburnt hummock grass. historic change; wildfire; Glossy Black-Cockatoo; Forty-spotted Pardalote; eucalypt open forest; Casuarina woodland; threatened species Gill, A. M. (1990). Fire management of mallee Bioregion: Woolnorth lands for species conservation. In J. C. Noble, Clearing for agriculture and wildfire (especially P. J. Joss, & G. K. Jones (Eds.), The mallee major fires around 1920) led to the extinction on lands: a conservation perspective (pp. 202-205). King Island of Glossy Black-Cockatoo and Melbourne: CSIRO. Forty-spotted Pardalote, as well as decline in mallee other species. Bioregion: Murray Darling Depression Describes characteristics of mallee fires. Species such as Malleefowl and Mallee Emu-wren are Hadlington, P., & Hoschke, F. (1959). threatened by high frequency of fires. Observations on the ecology of the phasmatid Ctenomorphodes tessulata (Gray). Proceedings of the Linnaean Society of New South Wales, 84, Gill, A. M. (1996). How fires affect 146-159. biodiversity. In DEST (Ed.), Fire and eucalypt open forest; management biodiversity: the effects and effectiveness of fire Bioregion: South Eastern Highlands management. Proceedings of the conference held While some fire regimes may decrease the 8-9 October 1994, Footscray, Melbourne (pp. abundance of phasmatids, others may lead to 47-55 (&123-124)). Canberra: Department of increases, presumably because they decrease the the Environment, Sports and Territories. number of egg parasites or the abundance (or Ground Parrot; Mistletoebird; threatened species; predation pressure) of predatory birds. heath; eucalypt open forest Description of some impacts upon biodiversity of a range of fire regimes. Hot fires in open forests Harrington, G. N., & Sanderson, K. D. (1994). may eliminate mistletoe (and hence lead to decline Recent contraction of wet sclerophyll forest in in ). Briefly reviews Queensland and the wet tropics of Queensland due to invasion Victorian studies of Ground Parrot response to by rainforest. Pacific Conservation Biology, 1, fire, noting that associations with vegetation age 319-327. of graminoid heaths seem to be related to tropical eucalypt open forest; management; rainforest. Bioregion: Wet Tropics

138 Fire and Australian birds

Fire suppression has led to contraction of ecotonal White-throated Treecreeper) were more common wet sclerophyll forests at the expense of expanding (or more obvious) in burnt areas. Four of these rainforest. The eucalypt forests are important for species are insectivorous ground-feeders. isolated populations of Eastern Yellow Robin, Recolonisation of burnt areas occurred very quickly, Yellow Thornbill, Buff-rumped Thornbill, even before the smoke had cleared. Previously Crested Shrike-tit, White-naped Honeyeater and common species which were rarer or absent in Yellow-faced Honeyeater, and these populations burnt areas included Golden Whistler, Rufous may require the establishment of a more frequent Whistler, Grey Fantail, Yellow-faced Honeyeater, burning regime. Yellow-tufted Honeyeater, Spotted Pardalote and Striated Pardalote, though Rufous Whistler and Yelow-faced Honeyeater may have migrated from Haynes, C. D. (1985). The pattern and the area. Most of these species forage in shrubs or ecology of munwag: traditional Aboriginal fire canopy, most of which was destroyed in the fire. regimes in north-central Arnhemland. Proceedings of the Ecological Society of Australia, 13, 203-214. Hodgson, A., & Heislers, A. (1972). Some Hooded Parrot; tropical eucalypt open forest aspects of the role of forest fire in South-eastern Bioregions: Top End Coastal, Pine Creek Australia. Forests Commission, Victoria. Arnhem. eucalypt open forest; review Changes from Aboriginal fire regimes may be Few birds are killed directly in fuel reduction fires. implicated in the decline of Hooded Parrot and Lyrebirds recolonised forest within 2 years of an some grass finches. extensive wildfire. Frequent burning of forests may favour species which forage in open ground (e.g. Spotted Quail-thrush, Buff-rumped Thornbill) but Hemsley, J. H. (1967). Bushfire - S.E. disadvantage those using shrubs (e.g. White-browed Tasmania 7th February, 1967. Some aspects of Scrubwren, Brown Thornbill) or eating Acacia the fire in relation to animal and plant life. seeds (Common Bronzewing). Tasmanian Forestry Commission. wildfire; mortality Bioregions: Freycinet, D’Entrecasteaux, Holmes, G. (1988). Eastern Bristlebird: Tasmanian Midlands summary conservation statement for northern 931 birds of 60 species were found dead on populations. QNPWS. beaches adjacent to areas burnt by wildfire. 25 Eastern Bristlebird; threatened species species were recorded from burnt area during the Bioregion: NSW North Coast first two months after the fire. Inappropriate fire regimes threaten Eastern Bristlebird.

Hewish, M. (1983). The effect of a wildfire on birdlife in a eucalypt forest: a preliminary Holmes, G. (1989). Eastern Bristlebird. Species report on the Lerdederg Gorge seven weeks management plan for northern populations. after the Wombat State Forest fire. Geelong Queensland NPWS and NSW NPWS. Naturalist, 20, 3-16. Eastern Bristlebird; rainforest; eucalypt open forest; eucalypt open forest; wildfire; community; succession threatened species; management. Bioregion: Victorian Midlands Bioregions: South Eastern Queensland; NSW Censuses in open forest before, 3 weeks and 7 weeks North Coast. after severe bushfire, including some census points Eastern Bristlebird in northern NSW and that were unburnt and others that were burnt only southeastern Queensland uses ecotones between mildly. The fire caused an immediate reduction in rainforest and open forest. Suitability is bird species diversity (25-26 species pre-fire falling determined by time since fire. Fire management is to 19-20 species post-fire). Unburnt sites required. maintained species richness. Mildly burnt sites were intermediate in richness. Some species (Eastern Yellow Robin, Grey Shrike-thrush, Superb Fairy-wren, White-browed Scrubwren,

139 Australia’s Biodiveristy - Responses to Fire

Hood, J. B. (1941). Birds and bushfires. South by Western Bristlebird (this change may be Australian Ornithologist, 15, 125-127. delayed by macropod grazing). Three congeneric mortality; wildfire. honeyeaters showed contrasting post-fire Bioregion: Naracoorte Coastal Plain responses. Tawny-crowned Honeyeater was Effects of wildfire on birds depend on weather recorded 1 year post-fire and peaked at 3 years conditions. In relatively mild fires, some species post-fire then declined to reach the unburnt heath (e.g. Fork-tailed Swift, Australian Raven) are abundance at 6 years post-fire. White-cheeked attracted to fire fronts to feed on disturbed or Honeyeater was first recorded 3 years post-fire, dead insects. In very hot fires (such as the and remained at low levels for the next 3 years. extensive wildfire of 1939), few birds were New Holland Honeyaeter was not recorded until observed to survive. Species noted to have been 6 years post-fire. killed included Emu, Stubble Quail, Masked Lapwing, Bush Stone-curlew, Sulphur-crested Cockatoo, Long-billed Corella, Eastern Rosella, How, R. A., Dell, J., & Humphreys, W. F. Crimson Rosella, Red-rumped Parrot, (1987). The ground vertebrate fauna of coastal Wedge-tailed Eagle, Australian Magpie, ravens areas between Busselton and Albany, Western and Laughing Kookaburra. “As the terrific blast of Australia. Records of the Western Australian fire swept along, birds could be seen rising from Museum, 13, 553-574. the ground or leaving the trees in front of it only heath; historic change; eucalypt open forest; Malleefowl; to be overpowered by the dense smoke and rising Western Whipbird; Rufous Bristlebird; Noisy scorching heat. In all cases the birds fell exhausted Scrub-bird; threatened species into the flames”. Bioregion: Warren Decline over this century in several birds in the Warren district was due to changes in fire regime, Hopkins, A. J. M. (1985). Planning the use of particularly affecting heath. Species which fire on conservation lands in south-western declined as a result included Malleefowl, Western Australia. In J. R. Ford (Ed.), Fire ecology and Whipbird, Rufous Bristlebird and Noisy management in Western Australian ecosystems Scrub-bird. (pp. 203-208). Perth: Western Australian Institute of Technology. review; Noisy Scrub-bird; management; Western Hughes, P., & Hughes, B. (1991). Notes on Whipbird; Western Bristlebird; heath; thicket; eucalypt the Black-breasted Button-quail at Widgee, open forest; threatened species Queensland. Australian Bird Watcher, 14, Bioregion: Jarrah Forest 113-118. Outlines a fire management plan for Two Peoples Black-breasted Button-quail; eucalypt open forest; Bay Nature Reserve, with major aims to conserve rainforest Noisy Scrub-bird, Western Whipbird and Bioregion: South eastern Queensland Western Bristlebird The Black-breasted Button-quail is dependent on deep leaf-litter, and occurs in vine thickets and adjacent scrubby eucalypt open forest. The Hopkins, A. J. M., & Smith, G. T. (1996). Fire: widespread practice of frequent burning may be effects and management implications. In A. J. detrimental to it. It can recover with absence of fire, M. Hopkins & G. T. Smith (Eds.), The natural and fire exclusion is recommended. history of Two Peoples Bay Nature Reserve Perth: Department of Conservation and Land Management (CALMScience Supplement). Hull, A. F. B. (1922). A visit to the Archipelago heath; threatened species; thicket; eucalypt open forest; of the Recherche S.W. Australia. Emu, 21, succession 277-289. Bioregion: Jarrah Forest seabirds; mortality; tussock grassland; heath Following fire in dense closed heath, Richard’s Bioregion: Esperance Plains Pipit initially colonised (to maximum density at 2 Noted repeated intense deliberate burns: “so years post-fire) then declined to absence at 6 years many times have fires been put through the scrub post-fire. Striated Fieldwren increased post-fire to that only possibly birds such as the Mutton Bird replace Pipits. As the regrowth heath proceeded could survive”. from open to closed the Heath-wren was replaced 140 Fire and Australian birds

Hunt, T. J., & Kenyon, R. F. (1970). The Jordan, R. (1984b). The Ground Parrot - rediscovery of the Mallee Whipbird in Effect of fire on a population. RAOU Report, Victoria. Australian Bird Watcher, 3, 222-226. 11, 28-29. Western Whipbird; mallee; heath; threatened species. Ground Parrot; threatened species; heath; wildfire Bioregion: Murray Darling Depression Bioregion: South East Corner Western Whipbirds may have left parts of the Big Populations in 5yr. post-fire coastal heath were c. Desert and Sunset Country this century because 0.2birds/ha. For c10 months after wildfire, no the vegetation was too old (and hence understorey Ground Parrots were recorded. Then birds too open). Recorded here in 9 year old regrowth. became resident: these were most probably Fires about a decade previously suggest that many immatures. areas would now be suitable. A previous record of this species was in “recently burnt” vegetation, though there are also records from more mature Jordan, R. (1987a). The Barren Grounds regrowth. rolling bird survey. RAOU Report, 27, 15-18. heath; eucalypt woodland; wildfire; community Bioregion: South East Corner Hutchins, B. R. (1988). Black Honeyeaters Four year survey of heathland with some feeding among charcoal and ash. South woodland following fire. Four species (Grey Australian Ornithologist, 30, 160. Shrike-thrush, Crimson Rosella, Brown charcoal; Black Honeyeater Thornbill, Rufous Whistler) showed no change. A number of Black Honeyeaters visited old Four species (Golden Whistler, Flame Robin, campfires and fed on charcoal and ash. As the Australian Magpie, Australian Raven) showed birds included males, the explanation that ash was immediate post-fire increase, followed by decline. eaten to provide calcium for egg-laying is Four species (White-browed Scrubwren, Beautiful insufficient. Firetail, White-eared Honeyeater, Eastern Spinebill) showed post-fire decline followed by increase. Two species (Southern Emu-wren, Jones, R. (1980). Hunters in the Australian Tawny-crowned Honeyeater) showed post-fire coastal savanna. In D. R. Harris (Ed.), Human population explosion (1-2 years post-fire) then Ecology in Savanna Environments London: decline. Academic Press. Aboriginal knowledge; Aboriginal burning; tussock grasslands; management; wetlands Jordan, R. (1987b). The Ground Parrot in Bioregion: Top End Coastal Barren Grounds Nature Reserve. RAOU Describes the traditional use of fire for hunting Report, 27, 19-23. and other resource management, including Ground Parrot; heath; threatened species burning of floodplains to maintain waterfowl Bioregion: South East Corner populations. Population density of Ground Parrots reaches a peak at 4-8 years post-fire in heathland, then declines to virtually absent by 12 years post-fire. Jordan, R. (1984a). The Eastern Bristlebird. Effects of fire on a population. RAOU Report, 11, 30. Jordan, R. (1987c). The Southern Emu-wren Eastern Bristlebird; threatened species; heath; wildfire in Barren Grounds. RAOU Report, 27, 24. Bioregion: South East Corner Southern Emu-wren; heath; wildfire Before a wildfire in coastal heathlands, Eastern Bioregion: South East Corner Bristlebirds were common. Following fire, no Following fire in coastal heathland, a few birds were detected for 2 months. Population emu-wrens moved in during the first year, but numbers then built up to approach pre-fire levels recolonisation became rapid after about at 2 years post-fire. All were occupying patches of 12 months. Birds were presumably moving in relatively dense regrowth. from unburnt patches (mostly within 100m.). Then recruitment was rapid, in response to large post-fire increases in insect numbers.

141 Australia’s Biodiveristy - Responses to Fire

Jordan, R. (1988). Population changes of some Kavanagh, R. P. (1990). Survey of Powerful and common insectivore species in woodland near Sooty Owls in south-eastern New South Wales. the wardens house. RAOU Report, 51, 16-18. Final report (Project 120). World Wildlife eucalypt woodland; wildfire; community Fund (Australia). Bioregion: South East Corner Sooty Owl; Powerful Owl; eucalypt open forest; Relatively small changes in abundance that could rainforest; old-growth; forestry; hollows; management; be attributable to wildfire (based on banding threatened species results before and up to five years post-fire) in Bioregions: South East Corner; South Eastern coastal woodland. Brown Thornbill and Highlands. White-bowed Scrubwren showed short-term Powerful Owl requires old-growth forests (across increases, possibly due to increased numbers of a broad floristic range); Sooty Owl requires insects. No change for Striated Thornbill, which old-growth forest, especially where there are foraged in the canopy (which remained relatively rainforest elements in the understorey. Too unaffected by fire). Eastern Yellow Robin declined frequent fire or logging will seriously for several years post-fire possibly because of the disadvantage these species. increased density of ground-layer vegetation. Eastern Spinebill increased for several years Kavanagh, R. P. (1991). The target species post-fire, possibly because of increased nectar approach to wildlife management: gliders and availability. owls in the forests of southeastern New South Wales. In D. Lunney (Ed.), Conservation of Australia’s forest fauna (pp. 377-383). Mosman: Joseph, L. (1982). The Glossy Black-cockatoo Royal Zoological Society of NSW. on Kangaroo Island. Emu, 82, 46-49. hollows; Sooty Owl; Powerful Owl; eucalypt open Glossy Black-cockatoo; Casuarina woodland; eucalypt forest; forestry; management; old-growth; threatened open forest; threatened species species Bioregion: Lofty Block Bioregions: South East Corner; South Eastern Glossy Black-cockatoo is dependent on Casuarina Highlands seeds and prefers to feed in taller, mature trees. Sooty Owls and Powerful Owls are associated Areas burnt 22 and 11 years prior to this survey with old-growth forests, and provide good were still unsuitable for Glossy Black-cockatoos. indicators for conservation planning.

Joseph, L., Emison, W. B., & Bren, W. M. Kimber, P. C. (1974). Some effects of (1991). Critical assessment of the conservation prescribed burning on Jarrah Forest birds. In status of Red-tailed Black-Cockatoos in Third Fire Ecology Symposium (pp. 49-57). south-eastern Australia with special reference Melbourne: Forests Commission, Victoria. to nesting requirements. Emu, 91, 46-50. eucalypt open forest; control burning; community; Red-tailed Black-Cockatoo; hollows; eucalypt open succession forest; management Bioregion: Jarrah Forest Bioregions: Murray Darling-Depression; Short-term effects of (hot) control fire were limited, Victorian Midlands with slight reduction in the number of birds Feeding habitat (Eucalyptus baxteri open forests) immediately following fire and extending to the first of Red-tailed Black-cockatoos in this area is spring after fire. After 2 years numbers increased to threatened by fuel-reduction burning. Nesting 25% above pre-fire levels. The greatest increase was trees are being lost by lack of regeneration and for species occurring in low to mid canopy levels clearing (including for firewood). (Western Gerygone, Brown-headed Honeyeater, White-naped Honeyeater, Grey Shrike-thrush, Inland Thornbill, Western Thornbill, Golden Whistler). Bird populations in a forest which had been unburnt for 40 years were appreciably less than in regularly burnt forests. This is so particularly for species occurring in understorey and lower canopies (e.g. Rufous Treecreeper, Western Yellow Robin, Western Spinebill, Golden Whistler, Brown Thornbill). 142 Fire and Australian birds

King, B., & King, D. (1983). Moggs Creek: Lewis, H. T. (1989). Ecological and Ash Wednesday plus nineteen days. Geelong technological knowledge of fire: Aborigines Naturalist, 20, 19-21. versus park rangers in northern Australia. eucalypt open forest; wildfire American Anthropologist, 91, 940-961. Bioregion: South east Coastal Plain Aboriginal knowledge; Aboriginal burning; General notes on a brief visit to an management; tussock grasslands; review; historic intensively-burnt open forest at 19 days following change; Magpie Goose; wetlands fire. The only birds seen were a single Black-faced Bioregion: Top End Coastal Cuckoo-shrike, a woodswallow, Australian Raven Reviews burning practices of Aboriginal people and Australian Magpie. and park rangers. Aboriginal people note that burning is important for Magpie Geese (and other waterfowl), as they prefer to feed in burnt areas King, B. R., Limpus, C. J., & Walker, T. A. around their nest sites. Last fires of the year in (1991). Seabird islands: No. 210. Fife Island, floodplains are always set before waterfowl begin Great Barrier Reef, Queensland. Corella, 15, to nest. 59-61. seabirds; tussock grassland Bioregion: Cape York Peninsula Lindenmayer, D. B., Norton, T. W., & Tanton, Several fires have been recorded from the island, M. T. (1990). Differences between wildfire affecting nesting seabirds. and clearfelling on the structure of montane ash forests of Victoria and their implications for fauna dependent on tree hollows. King, D. F. (1987). Further fauna surveys at Australian Forestry, 53, 61-68. Moggs Creek - post Ash Wednesday wildfire. hollows; forestry; review Geelong Naturalist, 23, 81-84. Bioregion: South Eastern Highlands eucalypt open forest; wildfire; community Fire and clearfelling affect hollow availability very Bioregion: South east Coastal Plain differently (age range of trees, spatial patterning, List of bird species recorded during one visit three abundance and longevity of hollow-bearing trees), years post-fire. Part of ongoing survey (see and hollow-nesting birds (such as owls and Dedman 1984). cockatoos) are more affected by clearfelling than by fire. Lane, S. G. (1976). Seabird islands: No. 18. Broughton Island, New South Wales. Lord, E. A. R. (1936). Notes on swifts. Emu, Australian Bird Bander, 14, 10-13. 35, 216-218. seabirds; tussock grassland; mortality White-throated Needletail; wildfire Bioregion: NSW North Coast White-throated Needletail (=Spine-tailed Swift) Fishermen frequently set fire to the grasslands on feeds on insects in smoke over bushfires. Flocks of this island. Burning may have interfered with the swifts followed fires for several days. breeding storm-petrels, but apparently not with shearwaters (which nest in deep burrows). Burnt areas may assist the shearwaters taking off or Loyn, R. H. (1985a). Bird populations in landing. successional forests of Mountain Ash Eucalyptus regnans in central Victoria. Emu, 85, 213-230. Lane, S. G. (1982). Seabird Islands: No. 119. eucalypt open forest; wildfire; hollows; forestry; Frederick Island, Archipelago of the succession; management Recherche, Western Australia. Corella, 6, Bioregion: South Eastern Highlands 61-62. Mountain Ash forests are rarely dry enough to seabirds; heath; tussock grassland burn, but when they do, trees are killed over Bioregion: Esperance Plains extensive areas. Little regeneration occurs Much of the vegetation was extensively burnt by naturally in absence of fire. Bird abundance was fire, probably in the year preceding this report. If estimated at a series of sites of varying age since this fire occurred during the breeding season it fire (including 39, 73, 90 and 223 years post-fire) would probably have destroyed many nesting birds.

143 Australia’s Biodiveristy - Responses to Fire

or logging. Young regrowth (after clear-felling) mature than in regrowth forests. In mixed foothill had a very different bird fauna to mature forests, forests, birds that feed from bare open ground including many species typical of open areas (e.g. (e.g. Scarlet Robin, Buff-rumped Thornbill, Nankeen Kestrel, Richard’s Pipit, Australian Spotted Quail-thrush, White-throated Nightjar, Magpie, Blue-winged Parrot, Flame Robin) or Painted Button-quail) are more common on low shrubbery (Superb Fairy-wren, Red-browed ridges than in gullies and in more recently burnt Finch). The bird species composition changed areas. In Mountain Ash forests, a few little between 50 to about 200 years post-fire. The open-country birds (Richard’s Pipit, Nankeen abundance of some birds that were associated with Kestrel, Blue-winged Parrot) occur for about particular plants (e.g. Correa lawrenciana, Acacia 3 years post-fire, but regeneration is rapid and dealbata) changed with the successional increase regrowth is soon colonised by bird species typical and then decrease of these. Hole-nesting birds of the understorey of mature forests. were more abundant in forests >100 years than 39-80 years, although dead trees left after the 1939 fire offered a reasonable number of hollows. Loyn, R. H. (1987). Effects of patch area and A few species were largely restricted to forests habitat on bird abundances, species numbers >200 years. These included Powerful Owl, Sooty and tree health in fragmented Victorian Owl, Australian Owlet-Nightjar and forests. In D. A. Saunders, G. W. Arnold, Sulphur-crested Cockatoo (all hollow-nesters), A. A. Burbidge, & A. J. M. Hopkins (Eds.), Mistletoebird, Australian King-Parrot, Lewin’s Nature conservation: the role of remnants of Honeyeater and Satin (all at least native vegetation (pp. 65-77). Chipping partial fruit-eaters) and Varied Sittella. The Norton: Surrey Beatty. immediate effects on birds of wildfire in Mountain eucalypt open forest; community; isolate Ash forests must be profound. After the 1939 Bioregion: South east Coastal Plain fires, flocks of Gang-gang Cockatoos were Bird assemblages were recorded from 56 forest reported from many suburban areas from which fragments, and the distribution of individual they were normally absent. species, and assemblage parameters, were related to a number of environmental factors (including fire history). Patch size, grazing history and Loyn, R. H. (1985b). Ecology, distribution and presence of Noisy Miners accounted for most density of birds in Victorian forests. In variation: fire history was not strongly correlated A. Keast, H. F. Recher, H. Ford, & D. with assemblage parameters, but this was possibly Saunders (Eds.), Birds of eucalypt forests and because fire history showed little variation woodlands: ecology, conservation, management between patches. (pp. 33-46). Chipping Norton: Surrey Beatty. eucalypt open forest; community; hollows Bioregions: South east Coastal Plain; South Loyn, R. H., Traill, B. J., & Triggs, B. E. Eastern Highlands; Australian Alps; Victorian (1986). Prey of Sooty Owls in East Gippsland Midlands; Riverina before and after fire. Victorian Naturalist, 103, Discusses effects of fire in a range of forests in 147-149. southeastern Australia. Compares densities of Sooty Owl; threatened species; rainforest; eucalypt open hole-nesting birds in a range of mature and forest regrowth forests. Highest proportion of Bioregion: South East Corner hollow-nesters is in mature River Red Gum Diet of Sooty Owl was restricted 4 months after forests. Mistletoe is more common in older forests fire to one terrestrial rodent species, in contrast to (and hence density of Mistletoebirds is greater). varied diet pre-fire. Owls were observed in burnt Some honeyeaters, foliage-gleaning insectivores forests, and contraction in diet may have been due and bark-foraging birds are more common in to greater ease of catching rats in burnt areas.

144 Fire and Australian birds

Loyn, R. H., Cameron, D. G., Traill, B. J., understorey (Brown Thornbill, White-browed Sloan, J. F., Malone, B. S., Schulz, M., Earl, G. Scrubwren) generally decreased. Nectarivorous E., & Triggs, B. E. (1992a). Flora and fauna of birds (lorikeets and honeyeaters) were more the Cooaggalah Forest Block, East Gippsland, common in burnt areas, attracted to what Victoria. Ecological Survey Report 20. appeared to be fire-induced high-quality Department of Conservation and flowering in some eucalypts. Three species which Environment. were rarely recorded before fires, became wildfire; eucalypt open forest; rainforest; community. abundant post-fire (Red-browed Finch, Bioregion: South East Corner White-winged Chough, Pied Currawong). In A very intense bushfire burnt the study area general, there was little change in total bird during survey. The fire had burnt especially richness or abundance, and most individual fiercely in some rainforest gully vegetation. species showed no significant change. Notes that Sampling continued 1-4 months, 1 year and 2 the study areas were small, and no burnt site was years post-fire. In the 1-4 months post-fire, bird >300m from unburnt forest. Also, the area had numbers were reduced to 57% of pre-fire levels. relatively few shrubs before burning, so impacts Honeyeaters departed, and little food remained may be more profound in forests of different for them. Numbers of granivorous and structure. The study deals with changes over 2 frugivorous birds were also greatly reduced. years only, and monitoring is needed to detect Bark-foraging and ground-foraging (e.g. Superb longer-term changes. Lyrebird, which were observed to forage in ash) birds showed little change, and some carnivorus birds and Flame and Scarlet Robins increased. Lucas, D., & Russell-Smith, J. (1993). Treecreepers and Laughing Kookaburra were the Traditional resources of the South Alligator only conspicuous birds in heavily-burnt stands. floodplain: utilisation and management. Pre-fire abundance was reached 2 years post-fire, Australian Nature Conservation Agency. and marginally exceeded at 3 years post-fire, wetlands; tussock grasslands; Aboriginal knowledge. though some honeyeaters (e.g. Crescent and New Bioregion: Top End Coastal Holland Honeyeaters) and Beautiful Firetail were Some traditional burning was aimed at increasing still well below pre-fire levels. Most localised the abundance of foodplants for people and species (e.g. Southern Emu-wren, Beautiful important birds (Magpie Goose, Brolga). Early Firetail) had returned to the sites that they had dry season burning had to be undertaken carefully occupied pre-fire. out of consideration for ground-nesting birds, such as Partridge Pigeon.

Loyn, R. H., Hewish, M. J., & Considine, M. (1992b). Short-term effects of fuel reduction Lucas, K., & Lucas, D. (1993). Aboriginal fire burning on bird populations in Wombat State management of the Woolwonga wetlands in Forest. In K. Tolhurst & D. Flinn (Eds.), Kakadu National Park. Australian Nature Ecological impact of fuel reduction burning in Conservation Agency. dry sclerophyll forest: first progress report (pp. wetlands; tussock grasslands; Aboriginal knowledge. 5.1-5.11). Melbourne: Department of Bioregion: Top End Coastal Conservation and Environment (Research The grasslands should be burnt annually at Report no. 349). appropriate times of year. One aim of burning is eucalypt open forest; control burning; community to maintain or increase abundance of wild rice Bioregion: Victorian Midlands Oryza because it is an important food for Magpie Over 3 years, bird populations were assessed Goose. Prefer not to burn too early in the dry before and after fuel-reduction burns at 3 areas season because some birds (notably Partridge (each with control, spring burn and autumn burn Pigeon, White-throated Grasswren, , sites) in eucalypt open forest. Scarlet and Flame finches and whistle-ducks) are nesting then on the Robins were substantially more common in ground or in grass. Small fires may be relatively autumn-burnt areas. Birds which fed to some safe then, and may prevent very large destructive extent on bare ground (Laughing Kookaburra, fires later. Grey Shrike-thrush, Superb Fairy-wren) generally increased post-fire; those that fed in shrubby

145 Australia’s Biodiveristy - Responses to Fire

Lyndon, E. (1977). Aftermath of fire at Matthew, J. (1994). The status, distribution Waratah Bay - and Lyrebirds. Victorian and habitat of the Slender-billed Thornbill Naturalist, 94, 18-19. Acanthiza iredalei in South Australia. South Superb Lyrebird; heath; eucalypt open forest Australian Ornithologist, 32, 1-19. Bioregion: South east Coastal Plain Slender-billed Thornbill; heath; mallee. Anecdotal report of lyrebirds occurring in Bioregion: Murray Darling Depression regrowth 2-3 years post-fire. The subspecies of Slender-billed Thornbill occurring in mallee heath (Acanthiza iredalei hedleyi) may be considered vulnerable to fires, but MacGillivray, W. D. K. (1910). Along the fire is not a threatening process for subspecies Great Barrier Reef. Emu, 10, 216-233. occurring in samphire. seabirds Bioregion: Cape York Peninsula Fire reported from Fife Island, killing breeding Mawson, P. R., & Long, J. L. (1994). Size and seabirds. age parameters of nest trees used by four species of parrot and one species of cockatoo in south-west Australia. Emu, 94, 149-155. Main, A. R. (1981). Fire tolerance of heathland hollows; eucalypt woodland; eucalypt open forest; Regent animals. In R. L. Specht (Ed.), Ecosystems of Parrot; Red-capped Parrot; Western Rosella; Port the World. 9B. Heathlands and related Lincoln Ringneck; Western Long-billed Corella shrublands. Analytical studies. Amsterdam: Bioregions: Avon Wheatbelt; Jarrah Forest Elsevier. Hot fires may create, enlarge or destroy hollows. review; heath Dimensions of hollows (and size and age of trees) Birds tend to flee in advance of smoke and flame used by Regent Parrot, Red-capped Parrot, front, in contrast to mammals, reptiles and Western Rosella, Port Lincoln Ringneck and which burrow or re-enter burnt areas Western Long-billed Corella are given. Trees through gaps in fire front. Ravens, crows, magpies used are very old (minimum 73 years, maximum and currawongs forage in recently burnt areas. >1300 years). Clearing is main cause of reduction Swallows and kestrels may feed in front of the fire. in hollow availability, but frequent burning of roadside verges prevents seedling establishment and hence tree recruitment. In some remnant Marchant, S. (1985). Breeding of the Eastern patches, complete absence of fire may also prevent Yellow Robin. In A. Keast, H. F. Recher, H. germination. Ford, & D. Saunders (Eds.), Birds of eucalypt forests and woodlands: ecology, conservation, management (pp. 231-240). Chipping Norton: McAllan, I. A. W. (1996). The Flock Surrey Beatty. Bronzewing Phaps histrionica in New South Eastern Yellow Robin; eucalypt open forest; eucalypt Wales, with comments on its biology. woodland; wildfire. Australian Bird Watcher, 16, 175-204. Bioregion: South East Corner Flock Bronzewing; review; tussock grasslands An 8 year study of banded population of Eastern Bioregions: Mitchell Grass Downs; Channel Yellow Robin at two plots, both of which were Country; Darling Riverine Plains; Mulga Lands; burnt (to various degrees) in the sixth year. Fire Broken Hill Complex had little immediate effect on Robins, with all Although exclusion of fire by pastoralists in birds apparently surviving. In the year after fire, Mitchell grasslands may have reduced seed nests were generally higher and clutches smaller, availability, this can “be only a minor problem” but breeding success was similar to pre-fire. At for Flock Bronzewings. two years post-fire, fewer birds nested and some territories were deserted. This was assumed to be because a dense grass layer had developed which hindered foraging. Fire effects may have been more serious if the fire was during the breeding season.

146 Fire and Australian birds

McCaw, L., Maher, T., & Gillen, K. (1992). Golden-headed Cisticola, Red-backed Fairy-wren Wildfires in the Fitzgerald River National Park, and Southern Emu-wren were most abundant in Western Australia, December 1989. Department mid to late ages. Pheasant Coucal and Eastern of Conservation and Land Management. Grass Owl showed no relationships with age. Past wildfire; mortality practice of burning every c3 years has advantaged Bioregion: Esperance Plains opportunistic species, but disadvantaged most of The remains of a large number of small birds were the resident species. Longer intervals between found in a woodland burnt by relatively low burns (e.g. 7-8 years) and mosaic burns are intensity wildfire. Following a patchy wildfire, recommended to maintain resident species. unburnt vegetation was important as refuge for less mobile species. Large flocks of granivorous birds were observed well within an McFarland, D. C. (1989). The Ground Parrot extensively-burnt area within days following Pezoporus wallicus (Kerr) in Queensland: wildfire. habitat, biology and conservation. Department of Conservation, Parks & Wildlife, Queensland. McCulloch, E. M. (1966). Swifts and Ground Parrot; threatened species; heath; isolate bushfires. Emu, 65, 290. Bioregion: South Eastern Queensland White-throated Needletail; Fork-tailed Swift; wildfire Ground Parrots were most abundant in heaths Both White-throated Needletail and Fork-tailed from 5-8 years post-fire, and are threatened by Swifts flocks fed in insects in smoke over fire intervals of less than this. This threat is bushfires. exacerbated by fragmentation.

McFarland, D. C. (1988). The composition, McFarland, D. C. (1991). The biology of the microhabitat use and response to fire of the Ground Parrot, Pezoporus wallicus, in avifauna of subtropical heathlands in Coloola Queensland. III. Distribution and abundance. National Park, Queensland. Emu, 88, 249-257. Wildlife Research, 18, 199-213. heath; community; succession; management Ground Parrot; heath; threatened species; sedgeland; Bioregion: South Eastern Queensland isolate Bird species composition and richness was Bioregion: South Eastern Queensland determined at sites 0,2.5,5.5,6.5 and 10.5 years Describes distribution and habitat preferences for post-fire. Older heaths had fewer “inconsistent” Ground Parrot in subtropical heaths. Although species (irregular visitors, migrants, etc.). Post-fire Ground Parrots were recorded at a site 2 months recolonisation depends upon extent and intensity post-fire, they were not recorded regularly until of fire and availability of unburnt patches. Species >9 months post-fire. Densities peaked at 5-8 years richness was higher (39 spp.) in the first year after post-fire, when the number of food plants, and fire than in the oldest site (14 species), although standing crop of food, is maximum, and when many of the species in the young site were these are most stable seasonally. Decline of migrants or visitors from nearby habitats. These numbers of Ground Parrots in older heath reflects latter were attracted by post-fire flowering or reduced seed availability. High intensity fires may seeding or exposed and dead invertebrates and be more deleterious than fuel-reduction burns. vertebrates immediately after fire (e.g. for raptors, Ground Parrots recolonise rapidly after either Laughing Kookaburra, Torresian Crow, summer or winter fires. Increasing heathland Straw-necked Ibis). Responses to fire age were isolation increases the possibility of complete related to food and shelter. Most structural burns with catastrophic fire and reduces the attributes either peaked or reached a plateau by chance of recolonisation. 3-6 years post-fire. Brown Quail used seeding grasses in open regeneration. King Quail, Brush Bronzewing and Ground Parrot preferred thicker vegetation of middle-aged heathlands (which also held greatest abundance of their seed resources). Swallows, martins and Richard’s Pipit preferred open young sites. Tawny Grassbird,

147 Australia’s Biodiveristy - Responses to Fire

McFarland, D. C. (1992). Fire and the McFarland, D. C. (1994). Notes on the Brush management of ground parrot habitat. In B. Bronzewing Phaps elegans and Southern R. Roberts (Ed.), Fire Research in rural Emu-wren Stipiturus malachurus in Cooloola Queensland (pp. 483-495). Toowomba: National Park. Sunbird, 24, 14-17. University of Southern Queensland. heath; Brush Bronzewing; Southern Emu-wren; Ground Parrot; heath; management; threatened species isolate Bioregion: South Eastern Queensland Bioregion: South Eastern Queensland Densities of the fire-sensitive and fire-dependent Heaths of 0-10.5 years post-fire were searched. Ground Parrots in subtropical heaths peak at 5-8 No Brush Bronzewings were recorded from years post-fire, with no birds recorded at >15 years heaths younger than 2 years or older than 10.5 post-fire. A management plan for Ground Parrots years. Southern Emu-wrens occurred in heaths of in heathlands of southern Queensland is described: age 1.5 to 9 years, with highest density in sites of it involves moderate to low intensity burning at 6-8 years. Fragmentation and generally high past 8-10 years during winter. Impacts of this regime fire frequency may have eliminated Emu-wrens on other heathland biota are considered. Of 12 from what now appears suitable heaths. species of breeding resident birds, 7 reach peak densities at or >6 years post-fire. McIlroy, J. C. (1978). The effects of forestry practices on wildlife in Australia: a review. McFarland, D. C. (1993). Fire and bird Australian Forestry, 41, 78-94. conservation. In C. P. Catterall, P. V. Driscoll, review; eucalypt open forest; management; forestry K. Hulsman, D. Muir, & A. Taplin (Eds.), Most Australian fauna is fire-adapted. High Birds and their habitats: status and conservation intensity fires and frequent low intensity fires in Queensland (pp. 41-44). St Lucia: homogenise forests, leading to elimination of Queensland Ornithological Society Inc. species dependent on shrubs, litter and logs. Of 48 review; Ground Parrot; heath; threatened species; bird species in a native forest, 12 occurred community; management following clear-felling. All of these remained after Bioregion: South Eastern Queensland the felled area was burnt. In southeastern Queensland, Ground Parrots nest only in dry heaths that are at least 3-4 years post-fire. Fire is the main factor affecting density. McKean, J. L., & Martin, K. C. (1985). Ground Parrots can recolonise burnt heaths Distribution and status of the Carpentarian within 12 months, densities peak at 5-8 years Grasswren Amytornis dorotheae. Conservation post-fire, and no birds have been recorded in Commission of the Northern Territory. heaths >15 years post-fire (but few such areas Carpentarian Grass-wren; hummock grasslands; occur or have been searched). This pattern management; threatened species correlates with seed availability and cover. Other Bioregion: Gulf Fall and Uplands declining species in Queensland affected by fire Hot extensive wildfires are the most serious threat regime include Golden-shouldered Parrot, to the Carpentarian Grass-wren. Eastern Bristlebird, Rufous Scrub-bird, Carpentarian and Striated Grass-wrens. In heathland, raptors, egrets and ibis move in during McKean, J. L., & Martin, K. C. (1989). and just after fires. Within 1 year of fire, there is Distribution and status of the Carpentarian usually massive flowering of Xanthorrhoea, which Grass-wren Amytornis dorotheae. Northern attracts many honeyeaters. Granivores colonise Territory Naturalist, 11, 12-19. somewhat later. The total number of species is Carpentarian Grass-wren; hummock grassland; highest within 1 year post-fire, but then declines. management; threatened species The number of breeding species peaks between Bioregion: Gulf Fall and Uplands 3-8 years post-fire, followed by decline, especially Extensive wildfires are the most severe threat to for granivorous species. Carpentarian Grass-wren, and recent fires have Controlling fire to benefit one bird species led to its decline and contraction may have deleterious impact on other species. In heathlands of southeastern Queensland, mosaic winter burning at intervals of 8-10 years is recommended.

148 Fire and Australian birds

McLaughlin, J. (1992). The floristic and Menkhorst, P. W., & Bennett, A. F. (1990). structural features of Black-eared Miner Vertebrate fauna of mallee vegetation in Manorina melanotis habitat. RAOU Report No. southern Victoria. In J. C. Noble, P. J. Joss, & 84. Royal Australasian Ornithologists Union. G. K. Jones (Eds.), The mallee lands: a conser- Black-eared Miner; threatened species; old-growth. vation perspective (pp. 39-53). Melbourne: Bioregion: Murray-Darling Depression. CSIRO. Mallee of at least 55-60 years post-fire is required, mallee with younger mallee possessing few of the Bioregion: Murray Darling Depression structural features (decorticating bark) preferred Mallee vegetation is highly flammable and for by this species. some birds certain seral stages in the post-fire succession offer a higher quality habitat. Species favouring climax vegetation include Black-eared McLaughlin, J. (1994). Searches for the Miner and Malleefowl. Fire management since Black-eared Miner Manorina melanotis in the European settlement has reduced habitat Victorian Murray Mallee. RAOU Report No. 93. heterogeneity and especially the extent of Royal Australasian Ornithologists Union. old-growth mallee. This is especially so in isolated Black-eared Miner; mallee; threatened species patches which fire can reduce to a single age class. Bioregion: Murray-Darling Depression The endangered Black-eared Miner is restricted to long-unburnt mallee remote from edges. Meredith, C. W. (1982). The research input to fire management in natural areas with particular reference to wildlife research in the McNamara, E. (1946). Field notes on the semi-arid lands of Victoria and New South Eastern Bristlebird. Emu, 45, 260-265. Wales. In A. Heislers, P. Lynch, & B. Walters Eastern Bristlebird; heath; threatened species (Eds.), Fire ecology in semi-arid lands (10pp). Bioregion: South East Corner Deniliquin: CSIRO. The causes of the rarity of Eastern Bristlebirds are mallee; community; management obscure. Bushfires may be the chief reason, as Bioregion: Murray Darling Depression heaths are vulnerable to fire, and birds would have Comprehensive studies are required for input into trouble surviving in burnt areas. But much of the management, else casual observations and myth heath where they occur now has been burnt in the substitute for knowledge. Notes that fire regime past. Unburnt patches may be the key to their includes season, intensity, pattern and area survival. affected, as well as frequency; that fire frequency should be a statistical distribution rather than an McNee, S. (1986). Surveys of the Western inflexible repetition of a mid-point; that Whipbird and Western Bristlebird in Western “optimum” fire regimes vary between species and Australia. RAOU Report No. 18. Royal even between short- and long-term within species; Australasian Ornithologists Union. and that unplanned random burning is not the Western Whipbird; Western Bristlebird; heath; mallee; same as mosaic burning. Mallee woodland threatened species; management. endemic birds are most common in areas >20-25 Bioregions: Esperance Plains; Warren; Jarrah years post-fire. Specialist species with restricted Forest distributions tend to be found in older vegetation The range and abundance of the Western while widespread generalists are typical of early Whipbird and Western Bristlebird have been regeneration. However some specialists may substantially reduced by clearing and too frequent favour early regrowth. While Malleefowl require fires. Management of these species will require long period between fires, it may be that very old long intervals without fire (at least 30 years and vegetation has a declining food supply. Shy preferably 50 years for Western Whipbird). Heath-wren, Red-lored Whistler, Western Western Whipbird is known to survive fire, but Whipbird and Striated Grass-wren may be then not persist in burnt areas. associated with early stages of regeneration. Research needs include study of fire relationships of other mallee endemic birds, and the effect of fire on hollow formation and seed and nectar production.

149 Australia’s Biodiveristy - Responses to Fire

Meredith, C. W. (1983). Fire and birds. The Meredith, C. W. (1984c). Recent records of result of two studies and their relevance to the Ground Parrot Pezoporus wallicus in fuel reduction burning. In E. H. M. Ealey western Victoria. Geelong Naturalist, 21, 3-4. (Eds.), Fighting fire with fire (pp. 193-202). Ground Parrot; heath; threatened species; control Melbourne: Graduate School of burning Environmental Science Monash University. Bioregion: South east Coastal Plain mallee; community; Ground Parrot; heath; sedgelands; The disappearance of Ground Parrot from some control burning; threatened species sites was related to old age of the heaths. The low Considers two examples: bird communities of numbers in another site were a result of extensive mallee and Ground Parrot in coastal heathlands. and frequent fuel reduction burns. In both habitats, there is little functional difference between wildfire and fuel reduction burns. Burning frequency in mallee of <20-30 Meredith, C. W. (1988). Fire in the Victorian years would result in 25% or more reduction in environment - a discussion paper. Conservation bird species richness. Habitat endemic species Council of Victoria. would be those most likely to be lost. These review; management include those requiring abundant litter (e.g. Short-term effects of fire on bird communities Malleefowl), hollows (Chestnut-rumped vary with fire intensity (and the extent of Thornbill, parrots) or dense large spinifex vegetation change). However if fire regimes lead (Striated Grass-wren, Mallee Emu-wren). Bird to long-term change in vegetation characteristics, numbers may be greatest at about 15 years there will be long-term change in bird post-fire. Species common in early regrowth communities. If fire frequency is too great, birds include mainly widespread opportunists (Kestrel, dependent on older stages will be lost. These are Australian Magpie, White-fronted Honeyeater, often specialist species with restricted some thornbills). Management should concentrate distributions. Notes that no long-term studies of on localised burning at borders. Ground Parrot the relationships between birds and fire have been can survive in sedgelands for indefinite period, but conducted in any Australian habitat; and that in heathlands it requires fire every 20-25 years. previous studies (e.g. Kimber 1974, Christensen & Currently, fuel-reduction burning is too frequent, Kimber 1975) have restricted analysis to common and may eliminate this species from graminoid species in the communities they studied. heathlands. Meredith, C. W., & Isles, A. C. (1980). A study Meredith, C. W. (1984a). The Ground Parrot. of the Ground Parrot (Pezoporus wallicus) in RAOU Conservation Statement, 1. Victoria. report No. 304. Environmental Ground Parrot; heath; sedgeland; threatened species Studies Division of the Ministry for Inappropriate fire regimes, compounded by Conservation, Victoria. fragmentation of populations, are the main threats Ground Parrot; heath; threatened species. to the Ground Parrot Bioregions: South East Corner; South east Coastal Plain; Naracoorte Coastal Plain Ground Parrot abundance peaks at 5-10 years Meredith, C. W. (1984b). Management of the post-fire; heaths then eventually become Ground Parrot Pezoporus wallicus in Victoria. unsuitable. Fisheries & Wildlife Division, Victoria. Ground Parrot; heath; sedgelands; threatened species; management Detailed recommendations for fire management at sites of varying floristics in Victoria, in order to conserve Ground Parrot populations.

150 Fire and Australian birds

Meredith, C. W., & Jaremovic, R. (1990). Milledge, D. R., & Palmer, C. L. (1990). The Current status and management of the Ground Sooty Owl in Mountain Ash forests in the Parrot in Victoria. Arthur Rylah Institute for Victorian Central Highlands. Department of Environmental Research. Conservation and Environment. Ground Parrot; heath; wildfire; control burning; Sooty Owl; hollows; eucalypt open forest; old-growth; threatened species; management management; threatened species Bioregions: South East Corner; South east Bioregion: South Eastern Highlands Coastal Plain Sooty Owls are virtually confined to forests that Hot fire in a coastal heathland reduced the are at least 150 years old, because of requirements Ground Parrot population at this site by 75%. for high density of hollows (for their own nesting Elsewhere, widespread and frequent and that of their prey). fuel-reduction burns in heathland led to substantial and lasting reductions in Ground Parrot densities. A model is developed to predict Milledge, D. R., Palmer, C. L., & Nelson, J. L. population changes post fire, and hence to design (1991). “Barometers of change”: the management burning regimes. distribution of large owls and gliders in Mountain Ash forests of the Victorian Central Highlands and their potential as management Meredith, C. W., Gilmore, A. M., & Isles, A. C. indicators. In D. Lunney (Ed.), Conservation of (1984). The Ground Parrot (Pezoporus wallicus Australia’s Forest Fauna (pp. 53-65). Sydney: Kerr) in south-eastern Australia: a Royal Zoological Society of NSW. fire-adapted species? Australian Journal of eucalypt open forest; old-growth; hollows; threatened Ecology, 9, 367-380. species Ground Parrot; heath; sedgeland Bioregion: South Eastern Highlands Bioregions: South east Corner; South east Coastal Survey of owls at four age classes (50 to >250 Plain; Naracoorte Coastal Plain. years). Sooty Owl was associated with large areas Examined Ground Parrot abundance at a series of of old-growth (>250 years) forest, though did sites of known age since fire. Fire age is the major occur in 50-80 year forests (where fire-killed stags determinant of abundance in heathlands, mainly remained). through its influence on food resources. The production of seeds eaten by Ground Parrot remains relatively constant over time in Milligan, A. W. (1904). Notes on a trip to the sedgelands but varies in heathland with time since Wongan Hills, Western Australia. Emu, 4, burning. Long unburnt (>20 years) and very 2-11. frequently burnt (<6-8 years) heathlands are Western Whipbird; threatened species; heath. unsuitable for Ground Parrot. Fires are Bioregion: Avon Wheatbelt uncommon in sedgelands, and Ground Parrots Repeated fires probably led to the disappearance show no relationship with sedgleland age. Ground of Western Whipbirds from the Wongan Hills Parrots may disperse to unusual habitats after fire, area. and immatures may move long distances. Ground Parrots are not fire-adapted, but rather they are Mollison, B. C., & Green, R. H. (1962). adapted to a diet dominated by the seeds of Mist-netting Tree-Martins on charcoal graminoid sedges. In habitats where the patches. Emu, 61, 277-280. availability of these seeds is unrelated to fire (e.g. Tree Martin; charcoal sedgelands) Ground Parrot abundance is also Tree Martins observed to collect and maybe unrelated to fire, but in heathlands, the Ground consume charcoal and ash. Parrot is a fire-requiring species because of its adaptation to a particular diet.

151 Australia’s Biodiveristy - Responses to Fire

Morton, S. R., & Brennan, K. G. (1991). Nichols, O. G., & Muir, B. (1989). Vertebrates Birds. In C. D. Haynes, M. G. Ridpath, & of the jarrah forest. In B. Dell, J. J. Havel, & M. A. J. Williams (Eds.), Monsoonal Australia: N. Malajczuk (Eds.), The Jarrah Forest landscape, ecology and man in the northern (pp. 133-153). Dordrecht: Kluwer. lowlands. (pp. 133-149). Rotterdam: Balkema. eucalypt open forest; review review; tropical eucalypt open forest; management Bioregion: Jarrah Forest Little doubt that changed fire regimes have affected Understorey birds are affected for 1-2 years by birds. If current burning regimes have higher cool burns in Jarrah forests; canopy species show frequency of late dry season hot fires (as is likely) little response. The Jarrah forest fauna is adapted then birds associated with shrubby understorey may to a great variety of fire regimes, many irregular have declined considerably. Research into the and unpredictable. management of fire for birds needs high priority.

Noske, R. (1988). The status and biology of the Nelson, J. L., & Morris, B. J. (1994). Nesting White-throated Grass-wren. Australian requirements of the Yellow-tailed National Parks & Wildlife Service. Black-cockatoo, Calyptorhynchus funereus, in White-throated Grasswren; hummock grassland Eucalyptus regnans forest, and implications for Bioregion: Pine Creek Arnhem forest management. Wildlife Research, 21, Fire may decrease some populations of 267-278. White-throated Grass-wrens, but the status of the Yellow-tailed Black-cockatoo; hollows; management species is probably secure. Fires affect populations Bioregion: South eastern Highlands for two reasons: Grass-wrens rely on living The mean estimated age of trees used for nesting vegetation both directly and indirectly for plant by Yellow-tailed Black-cockatoos was 221 years and insect food; and they may be killed by fire (with youngest at 162 years). Live trees are better because of their ground-dwelling habits and poor nesting sites than dead stags as they are better able flying ability. Fire frequency may be more to withstand fire. important than intensity.

Newsome, A. E., McIlroy, J., & Catling, P. Noske, R. A. (1992a). Do Grasswrens have the (1975). The effects of extensive wildfire on numbers? Reply to Woinarski (1992). populations of twenty ground vertebrates in Northern Territory Naturalist, 13, 5-8. south-east Australia. Proceedings of the White-throated Grasswren; hummock grassland Ecological Society of Australia, 9, 107-123. Bioregion: Pine Creek Arnhem wildfire; heath; eucalypt open forest Most sites where White-throated Grasswrens Bioregion: South East Corner were recorded had been burnt in the previous few Tracks of large vertebrates were counted before and years. “Indeed I was impressed by the ability of after an extensive hot wildfire. No changes were Grasswrens to survive in areas of recent severe detected for Superb Lyrebird (the only bird fires.” Long-term absence of fire may be considered), though there were few records to disadvantageous. compare. Fire may reset fauna succession in dry sclerophyll forests, but “it is too early to think of utilising fire as a tool for management of fauna in Noske, R. A. (1992b). The status and ecology National Parks”. Wildfire is not necessarily the of the White-throated Grasswren Amytornis bogey widely imagined. woodwardi. Emu, 92, 39-51. White-throated Grasswren; hummock grassland Bioregion: Pine Creek Arnhem Fire is important in the ecology of White-throated Grasswrens, but information on its effects is inconclusive. Regular burning may benefit this species through maintenance of hummock grasslands at the expense of more fire-sensitive plants. The species was recorded across the range of fire ages studied (6 months

152 Fire and Australian birds

post-fire to “many years” post-fire). Fire may have Pescott, T. (1983). Beach-washed birds after immediate impacts of reduction in food resources the Ash Wednesday fire. Geelong Naturalist, and nest sites. Absence from some sites may be 20, 17-19. due to high frequency of fires. Recent changes in eucalypt open forest; wildfire; mortality fire regime may suggest that its secure status Bioregion: South east Coastal Plain cannot be assumed. Large numbers of bush birds were washed up on the beach following hot fire. The birds generally did not show any signs of having been burnt. A Paton, J. B., & Paton, D. C. (1977). Seabird total of 2183 individuals of 66 species were Islands: No. 52. Wright Island, South recorded. The most numerous dead birds were Australia. Corella, 1, 68-69. Crimson Rosella (984 individuals), New Holland seabirds Honeyeater, Red Wattlebird, Currawongs, Bioregion: Eyre and Yorke Blocks Sulphur-crested Cockatoo, White-eared A fire lit to control exotic Boxthorn spread Honeyeater, Ravens, Australian Magpies and through all vegetation. The resultant vegetation White-naped Honeyeater. loss caused erosion and damage to penguin burrows, and discouraged Silver Gulls from breeding, but probably had little impact on Fairy Pescott, T. (1985). Black Honeyeaters and Terns. charcoal. Geelong Naturalist, 22, 37-39. charcoal; Black Honeyeater Black Honeyeaters observed to repeatedly collect Pattemore, V. (1980). Effects of the pulpwood and consume charcoal. industry on wildlife in Tasmania. National Parks and Wildlife Service, Tasmania. forestry; management; eucalypt open forest Pescott, T. W. (1976). Seabird Islands: No. 27. Mainly concerned with effects of forestry Lady Julia Percy Island, Victoria. Australian practices, but considers interaction with fire Bird Bander, 14, 29-31. regimes. seabirds; tussock grassland Bioregion: South east Coastal Plain Occasional fires have had considerable impact on Pedler, L. (1991). Rare bird survey: implications vegetation, at times leading to rapid degeneration for fire management for Uluru National Park. of petrel and penguin burrows. Australian National Parks and Wildlife Service. hummock grassland; mallee; Striated Grasswren; Porter, J. W., & Henderson, R. (1983). Birds threatened species; management and burning histories of open forest at Bioregion: Great Sandy Desert Gundiah, southeastern Queensland. Sunbird, Examines distribution of rare bird species around 13, 61-69. Uluru. Striated Grasswren requires mature eucalypt open forest; historic change; management spinifex. Recommends a fire management strategy Bioregion: South Eastern Queensland which produces a fine-grained diversity of fire Bird communities were sampled in forests which ages in hummock grassland and mallee, and the had been subjected to three fire regimes over retention of long-unburnt patches. 29 years: burnt annually, burnt periodically at 2-5 year intervals, and unburnt. The annually burnt forest had developed a grassy understorey Pedler, L., & Burbidge, A. H. (1995). The and the unburnt forest had a shrubby understorey. range and status of the Nullabor Quail-thrush. Four species (Pheasant Coucal, , South Australian Ornithologist, 32, 45-52. Brown Treecreeper and Australian Magpie) were Nullabor Quail-thrush; chenopod shrubland; most abundant in the annually burnt forest. Two management; threatened species species (Forest Kingfisher, White-throated Bioregion: Nullabor Honeyeater) were most abundant in the forest Fire is a threat to Nullabor Quail-thrush as its burnt at longer intervals. Six species (Eastern preferred bluebush is killed in severe fires, and Yellow Robin, Golden Whistler, Variegated subsequent regeneration may be affected by Fairy-wren, White-throated Treecreeper, Little herbivore grazing.

153 Australia’s Biodiveristy - Responses to Fire

Wattlebird and Yellow-faced Honeyeater) were many repeatedly. Extensive areas of long-unburnt most abundant in the unburnt forest. 15 of 37 mallee are now rare in NSW. Fire threatens common species showed no apparent preference. mallee remnants in a different way to extensive The total abundance and richness of birds did not stands, with the isolation of remnants offering differ appreciably between treatments. The some protection, but when fires occur in them they current periodic burning of these forests is are likely to homogenise the remnant completely. probably restricting encroachment of rainforests and their associated birds. Priddel, D., & Wheeler, R. (1990). Survival of Malleefowl Leipoa ocellata chicks in the Press, A. J. (1987). Fire management in absence of ground-dwelling predators. Emu, Kakadu National Park: the ecological basis for 90, 81-87. the active use of fire. Search, 18, 244-248. Malleefowl; mallee; threatened species management Bioregion: Murray Darling Depression Bioregions: Pine Creek Arnhem; Top End Coastal Habitat quality may be the most critical element In the absence of management, up to 80% of the in the recruitment and survival of Malleefowl. area can be burnt annually. Fires attract falcons, This may be related to rainfall. “Alternatively, the kites and Tree Martin, which hawk in front of the age of the mallee community (i.e. time since last fire. Immediately after fire, Grey Butcherbird, fire) may be the critical determinant of habitat Magpie-lark, Straw-necked Ibis, Torresian Crow, suitability”. Red-tailed Black-cockatoo, owls and nightjars feed in the burnt area. Pyke, G. H., Saillard, R., & Smith, J. (1995). Abundance of Eastern Bristlebirds in relation Priddel, D. (1989). Conservation of rare fauna: to habitat and fire history. Emu, 95, 106-110. the Regent Parrot and the Malleefowl. In J. C. Eastern Bristlebird; eucalypt open forest; heath; Noble & R. A. Bradstock (Eds.), management Mediterranean landscapes in Australia: mallee Bioregion: South East Corner ecosystems and their management (pp. 243-249). Sampled Eastern Bristlebirds in vegetation at sites Melbourne: CSIRO. (forest, woodland, “mallee-heath” and heath) from Malleefowl; mallee; threatened species; management 0-14 years post-fire (all fires were control burns). Bioregion: Murray Darling Depression Bristlebird numbers increase with increasing time Active management of mallee habitat is required since fire up to about 9 years and then remain for the maintenance of its biota. Important food relatively constant for at least 4 years. Bristlebirds sources for the Mallee-fowl are available only after occurred in recently burnt areas, but these were the plants reach a certain age. Homogeneous mostly small and close to unburnt areas, Fire habitat (such as following extensive fire) may be management is essential for the conservation of detrimental to it. Bristlebirds. Numbers can be maintained if the interval between fires is at least 10 years, and if unburnt patches are retained during control Priddel, D. (1990). Conservation of the burning. Malleefowl in New South Wales: an experimental management study. In J. C. Noble, P. J. Joss, & G. K. Jones (Eds.), Pyne, S. J. (1991). Burning bush: a fire history The mallee lands: a conservation perspective of Australia. New York: Henry Holt and (pp. 71-74). Melbourne: CSIRO. Company. Malleefowl; mallee; threatened species; isolate; review management Provides a broad review of the history and Bioregion: Murray Darling Depression character of fire in Australia, with a few review One of four given threats to the long-term examples of responses of birds to fire. survival of Malleefowl is fire. There is a correlation between breeding density and age since fire. Almost all extensive stands of mallee in NSW have been burnt within the last 30 years,

154 Fire and Australian birds

Quin, B. R. (1990) Conservation and status of Ratkowsky, A. V. (1979). The bird species of the Turquoise Parrot (Neophema pulchella, Mt. Nelson in relation to microhabitat and Platycercidae) in Chiltern State Park and recent bushfires. Tasmanian Naturalist (57), adjacent areas. M.Sc., Department of Zoology, 12-18. LaTrobe University. eucalypt open forest; wildfire; community; control Turquoise Parrot; threatened species; management. burning Bioregion: South Eastern Highlands Bioregion: Freycinet Fires (wildfire and control burns) destroy the No. of bird species declined from 11.6 before a stumps often used by nesting Turquoise Parrots. November controlled burn in dry eucalypt forest to an average of 6.9 in the 3 months after fire. For 19 weeks after a hot October fire, bird Quin, B. R., & Baker-Gabb, D. (1993). richness in burnt areas was far less (5.7) than in Conservation and management of the Turquoise adjacent unburnt areas (16.2 species). Species Parrot Neophema pulchella in north-east absent from, or rarer in, the burnt areas included Victoria. Department of Conservation and Swift Parrot, Green Rosella, Fantailed Cuckoo, Environment. Shining Bronze-cuckoo, Olive Whistler, Golden Turquoise Parrot; eucalypt woodland; threatened Whistler, Satin Flycatcher, Superb Fairy-wren, species; management; control burning; hollows Brown Thornbill, Yellow Wattlebird, Bioregions: South Eastern Highlands; NSW Yellow-throated Honeyeater, Strong-billed Southwestern Slopes Honeyeater, Black-headed Honeyeater, Crescent Dead trees and stumps, favoured for nesting by Honeyeater, New Holland Honeyeater, Spotted Turquoise Parrots, are susceptible to destruction by Pardalote, Silvereye, Grey Butcherbird and Forest fire. These should be protected from control Raven. Species more common in the burnt area burning (or if burning must occur, the vicinity of all were Common Bronzewing and Dusky nest sites should be first slashed). Management may Wood-swallow. Common species showing little be able to impose a fire regime which increases the difference included Black-faced Cuckoo-shrike, dominance of native perennial grasses. Blackbird, Grey Shrike-thrush and European Goldfinch. More prolonged impact of this fire, compared to the milder fire considered by Ratkowsky, A. V. (1978). The effect of a spring Ratkowsky (1978), were because this severe fire fire on the number of bird species. Tasmanian had consumed much more of the vegetation, Naturalist (53), 11-12. including tree canopies. This “indicates clearly eucalypt open forest; wildfire; community that intensity of fire is a very important factor in Bioregion: Freycinet determining the rate at which avifauna will Birds were counted in burnt and unburnt dry repopulate a burnt area”. eucalypt forest (either side of a road) from 5 days to 13 weeks after a moderate intensity October wildfire (foliage up to about 8m was destroyed). In Ratkowsky, A. V. (1985). The effect of a severe the first 11 weeks post-fire, there was an average fire on the number of bird species in a wet of 2.7 species. Then this suddenly increased to 7.0 sclerophyll environment. Tasmanian species, equal to that in the unburnt area. During Naturalist, 8-9. the first week post-fire, Fantailed Cuckoo, eucalypt open forest; community; wildfire Black-headed Honeyeater and Black-faced Bioregion: Freycinet Cuckoo-shrike remained in the burnt area, but Following hot fire in January, more bird species they then disappeared from it. After 11 weeks, were observed in unburnt area (17.5 spp.) than in Black-headed Honeyeater, Brown Thornbill, burnt area (6.5 spp) in October-December. Superb Fairy-wren and, to a lesser extent, Dusky Species absent from the burnt area included Wood-swallow, Black-faced Cuckoo-shrike, and Shining Bronze-cuckoo, Olive Whistler, Golden Striated Pardalote returned to the burnt area. At Whistler, White-browed Scrubwren and this period there was considerable regrowth at the Strong-billed Honeyeater. Species less common in base and trunks of eucalypts and recovery of other burnt area included Peregrine Falcon, plants. There was no change observed for Forest Yellow-tailed Black-cockatoo, Black-faced Raven, Yellow-throated Honeyeater, Grey Cuckoo-shrike, Black-headed Honeyeater, Shrike-thrush, , Blackbird or Silvereye, Swift Parrot, Green Rosella, Blackbird, Dusky Robin. 155 Australia’s Biodiveristy - Responses to Fire

Yellow-throated Honeyeater, Crescent logging may be compounded if they are associated Honeyeater and Spotted Pardalote. The Striated with fire. Relatively inflexible imposed fire Pardalote was the only species common in the regimes may be particularly detrimental to birds. burnt area (though not more common than in the unburnt area). Results generally similar to those reported following severe fire in drier eucalypt Recher, H. F., & Christensen, P. E. (1981). open forest (Ratkowsky 1979). Fire and the evolution of the Australian biota. In A. Keast (Ed.), Ecological biogeography in Australia (pp. 135-162). The Hague: Junk. Recher, H. F. (1981). Death of an Australian review myth: fire and its effects on wildlife. In Birds easily survive mild fires: in a controlled fire P. J. Stanbury (Ed.), Bushfires: their effect on in shrub woodland, birds avoided the flames and Australian life and landscape. (pp. 39-48). moved into an adjacent unburnt heath, then Sydney: Macleay Museum, University of returned to their original sites immediately after Sydney. the fire had passed. However, there may be review; wildfire appreciable mortality in extensive wildfire. “The effects of fire on animals … have been Because of their greater mobility and higher neglected by Australian scientists … There are reproductive potential, birds respond to the few long term studies.” In mild fires, birds easily effects of fire more rapidly than do mammals. avoid flames. Some species are attracted to the fire Changes are most pronounced for birds of the front. In a mild burn, birds were observed in the ground and understorey vegetation. After mild still smoking forest within moments of the flame fires, bird numbers decline initially but then (1-3 passing. In extensive wildfires, immediate effect is years post-fire) quickly return to or exceed more severe. But birds respond rapidly because of pre-fire levels. Reproductive output of birds may their high mobility and high reproductive be increased in the year following fire (probably potential. Within a year of burning, birds are due to greater insect availability). numerous in forest. Many will have survived in small unburnt patches but others probably move in from distant areas unaffected by fire. Birds may Recher, H. F., & Lim, L. (1990). A review of respond to flush of insects on regrowth current ideas of the extinction, conservation vegetation. Forest canopy birds are generally less and management of Australia’s terrestrial affected than ground-dwelling birds or birds of vertebrate fauna. Proceedings of the Ecological the shrub layer. Nectarivorous birds may be Society of Australia, 16, 287-301. affected if fire frequency leads to decline in review; management nectariferous plants. A diversity of fires (rather Some birds (such as grass-wrens and scrub-birds) than regular prescriptive burns) is needed to which were restricted to specialised habitats, have maintain the greatest range of bird species. been affected by habitat change through alteration of fire regimes or grazing. Of 13 bird species which declined in King’s Park over the previous Recher, H. F. (1991). The conservation and 60 years, 10 were ground-foragers. This change management of eucalypt forest birds: resource was due to alteration of ground vegetation and requirements for nesting and foraging. In litter caused by spread of weeds and changed fire D. Lunney (Ed.), Conservation of Australia’s regimes. Frequent fuel-reduction burns in forests forest fauna. (pp. 25-34). Sydney: Royal are a concern, as they change understorey and Zoological Society of NSW. litter characteristics and lead to reduction in eucalypt open forest; management; forestry available energy. These fire-induced changes will Many bird species have very idiosyncratic have long-term effects on many insectivorous requirements for nesting or foraging (e.g. spiders birds. webs, dead limbs). Fire (and logging) simplifies forest environments to the detriment of many bird species. Loss of shrubby understorey may result in greater nest predation or removal of species which forage in particular shrub species. Effects of

156 Fire and Australian birds

Recher, H. F., & Serventy, D. L. (1991). Long Recher, H. F., Allen, D., & Gowing, G. (1985). term changes in the relative abundance of The impact of wildfire on birds in an birds in Kings Park, Perth, Western Australia. intensively logged forest. In A. Keast, Conservation Biology, 5, 90-102. H. F. Recher, H. Ford, & D. Saunders (Eds.), historic change; eucalypt open forest; management Birds of eucalypt forests and woodlands: ecology, Bioregion: Swan Coastal Plain conservation, management (pp. 283-290). Over nearly 60 years of intermittent survey, Chipping Norton: Surrey Beatty. change in ground cover (development of dense eucalypt open forest; community; wildfire; understorey, loss of eucalypt canopy and management; forestry proliferation of weeds), associated with fire Bioregion: South East Corner regime, has led to marked change in the bird One year after an intense and extensive wildfire, assemblage of Kings Park. Nine of 16 species bird populations (and invertebrate numbers) were which declined (Western Spinebill, Pallid Cuckoo, censused in 12 burnt and unburnt sites, including Grey Butcherbird, Laughing Kookaburra, Golden some which had been logged, across a range of Bronze-cuckoo, Scarlet Robin, Indian Turtledove, forest types. Invertebrates were generally more Western Yellow Robin, Senegal Dove, abundant in foliage in burnt forests. Burnt and White-tailed Black-cockatoo, Golden Whistler, logged forests had fewer bird species and Sacred Kingfisher, Tree Martin, Purple-crowned individuals than unburnt and unlogged forests. Lorikeet, Whistling Kite, Willie Wagtail) were The combined effect of these two disturbances ground-foraging birds. Prescription burning from was greater than either alone. Brown Thornbill, the 1930s to 1960s has exacerbated problems Striated Thornbill, White-throated Treecreeper, (spread of weeds and debilitation of remaining Golden Whistler and Grey Fantail were less eucalypts). abundant in burnt forests. Nectarivores (Musk Lorikeet, Yellow-faced Honeyeater, Yellow-tufted Honeyeater, White-naped Honeyeater), Recher, H. F., Lunney, D., & Posamentier, H. granivores (Diamond Dove, Common (1975). A grand natural experiment - the Bronzewing, Forest Bronzewing, Brown Quail, Nadgee wildfire. Australian Natural History, Painted Quail, Red-browed Finch, Beautiful 18, 150-163. Firetail) and Spotted Pardalote were generally wildfire; heath; eucalypt open forest; mortality more abundant in burnt forest. Bioregion: South East Corner “Considering the frequency with which fires Around 70,000 ha (including almost all of Nadgee occur in eucalypt forests and woodlands and their reserve) of heath and open forest were burnt in a dramatic impact on the landscape, there is wildfire, the first major fire in the area for 40 years. remarkably little information about the effects of The beach was littered with dead small birds. One fire on fauna or the long-term consequences of year after fire, all pre-fire bird species were present burning on forest ecosystems." Generally fire except for Crescent Honeyeater, although some effects are most substantial on species occurring in were less common than previously. Multiple lower substrates, and the timing of their responses nesting was evident for some insectivorous birds. is related to vegetation recovery. Proposes a Management should ensure that fires do not burn model that nutrients released after fire promote the whole reserve, and that a diversity of fires growth of nutrient-rich foliage and hence an occurs. increase in invertebrates (and thus insectivorous birds). However, following this intense fire, the relatively sparse foliage in burnt areas may have led to increased predation and to lower density of invertebrates per land area (cf. per area of foliage), and hence birds were relatively less common.

157 Australia’s Biodiveristy - Responses to Fire

Recher, H. F., Davis, W. E., & Holmes, R. T. Reid, J., & Fleming, M. (1992). The (1987a). Ecology of Brown and Striated conservation status of birds in arid Australia. Thornbills in forests of south-eastern New Rangelands Journal, 14, 65-91. South Wales, with comments on forest review; management; historic change; hummock management. Emu, 87, 1-13. grasslands; mallee Brown Thornbill; Striated Thornbill; eucalypt open Suggests that fire regime and extent of forest; wildfire; forestry; management long-unburnt patches have changed in arid Bioregions: South east Corner; South eastern environments generally since European Highlands settlement, though evidence is equivocal. There is Brown and Striated Thornbills were more more convincing evidence of changed fire regime abundant in unburnt open forest than in forest in hummock grassland deserts, to the possible burnt 1 year previously, regardless of logging detriment of species such as Striated Grass-wren. history. However, pastoralism is regarded as a more serious threat for most birds and most environments (e.g. riparian habitats, tussock Recher, H. F., Shields, J., Kavanagh, R., & grasslands and chenopod shrublands). Webb, G. (1987b). Retaining remnant mature forest for nature conservation at Eden, New South Wales: a review of theory and practice. Reid, J. R. W., Kerle, J. A., & Morton, S. R. In D. A. Saunders, G. W. Arnold, A. A. (1993a). Birds. In J. R. W. Reid, J. A. Kerle, & Burbidge, & A. J. M. Hopkins (Eds.), Nature S. R. Morton (Eds.), Uluru fauna: the conservation: the role of remnants of native distribution and abundance of vertebrate fauna vegetation (pp. 177-194). Chipping Norton: of Uluru (Ayers Rock - Mount Olga) National Surrey Beatty. Park, N.T. (pp. 36-57). Canberra: Australian eucalypt open forest; community; isolate; management; National Parks and Wildlife Service. forestry mallee; hummock grassland; Acacia woodland; Bioregion: South Eastern Corner community; wildfire Six creek reserves (retained strips post-logging) Bioregion: Great Sandy Desert and two sites in unlogged areas were censused for Recently burnt areas are favoured by nomadic birds over 8 years. The widest reserve was burnt birds (e.g. Budgerigar, White-winged Triller, in the fourth year and four others were burnt in Black Honeyeater, Crimson Chat, Zebra Finch, the seventh year. The fire had little effect on bird Masked Wood-swallow) and the rarer Banded numbers in the widest strip, but total bird Whiteface. A few species (e.g. Striated population and number of forest birds declined Grass-wren, Rufous-crowned Emu-wren) are precipitously in the year post-fire for the four largely restricted to mature hummock grasslands. smaller burnt strips. The decline was only partly By 10 years post-fire, mulga woodlands were due to fire, but may have been compounded by recolonised by most sedentary mulga birds, drought. Open country birds increased in the though much older woodlands support more strips post-fire. mistletoe and hence greater numbers of White-fronted Honeyeaters. Recommends a patch-burn strategy as advocated by Aboriginal Recher, H. F., Hutchings, P. A., & Rosen, S. residents. Extensive fires probably have a more (1993). The biota of the Hawkesbury-Nepean homogenising effect. Rainfall after fires may catchment: reconstruction and restoration. substantially influence the rate of regrowth and Australian Zoologist, 29, 3-41. bird responses. historic change Bioregion: Sydney Basin Describes environmental change since European colonisation. Many bird species have declined, associated with a complex mix of factors including change in fire regimes. Increased fire frequencies and more frequent wildfires of human origin are probably the major causes of decline of Rock Warbler and Ground Parrot.

158 Fire and Australian birds

Reid, J. R. W., Kerle, J. A., & Morton, S. R. Seed-eating Bronzewings did not appear until 2 (1993b). Ecological patterns and processes of years post-fire. Another group showed an initial importance. In J. R. W. Reid, J. A. Kerle, & S. increase, generally followed by decrease (Flame R. Morton (Eds.), Uluru fauna: the distribution Robin, Scarlet Robin, Australian Magpie, ravens). and abundance of vertebrate fauna of Uluru Breeding was recorded within the first year (Ayers Rock - Mount Olga) National Park, N.T. post-fire for Flame Robin, Buff-rumped (pp. 133-148). Canberra: Australian National Thornbill, Grey Currawong and Australian Parks and Wildlife Service. Magpie. By two years post-fire all species present hummock grasslands; mallee; Acacia woodland; pre-fire had been recorded, though abundance management was generally still slightly less. Some species Bioregion: Great Sandy Desert changed foraging behaviours in burnt areas. “Mulga is killed by fire, and given the importance Rufous Bristlebirds began returning 2 years of this species to the associated rich bird post-fire, possibly from refuges in town gardens community, it would seem that the destruction of (where they had been unreported before the fire). extensive areas of mulga by fire would decimate Fire probably poses the greatest threat to this bird populations, at least until the mulga had species. “If fire is to be used as a management regenerated to a certain age.” However, a mulga tool, frequent pockets of unburnt country of stand 11-14 years post-fire supported birds typical sufficient size are vital … The time of burning of tall old mulga, as well as many nomadic species ought also to be considered so that its impact on characteristic of early regrowth. This may be breeding populations is minimal.” because the site was close to old mulga. Management should aim to avoid very extensive fires, and fires in regrowth mulga not old enough Reilly, P. (1991b). The effect of wildfire on to have produced seed (c15 years). In spinifex, bush bird populations in six Victorian coastal Striated Grasswren is confined to mature habitats. Corella, 15, 134-142. vegetation, while Banded Whiteface (and a range heath; thicket; eucalypt open forest; community; of nomadic species) are associated with recently wildfire; management; succession burnt areas. Management require a fine-scale Bioregions: South eastern Highlands, South east patch-burn strategy, but with special attempt to Coastal Plain maintain old vegetation. Six sites across a range of habitats were monitored at 6-monthly intervals from 14 months after a severe bushfire to 56 months post-fire. No Reilly, P. (1991a). The effect of wildfire on unburnt controls were available in the area. For bird populations in a Victorian coastal habitat. some species, return showed considerable Emu, 91, 100-106. variation between sites. Rufous Whistlers initially eucalypt open forest; wildfire; community; did not appear to be greatly affected by the fire, management; isolate; succession but became extremely rare or absent 32-56 Bioregion: South east Coastal Plain months post-fire. Southern Emu-wrens Bird censuses were conducted in coastal recolonised heath 38 months and 56 months vegetation from several months before a hot and post-fire, possibly from a nucleus in a small patch extensive wildfire (with very few unburnt patches) that was unburnt. Rufous Bristlebirds returned at to nearly six years post-fire. There were marked 14 months post-fire at one site (near an unburnt differences between species in responses. A group patch), but had not returned to another site (of of species were not or little affected (including former occurrence) 7.5 years post-fire. Crimson Rosella, Rufous Whistler, Grey White-throated Treecreepers recolonised sites at Shrike-thrush, Grey Fantail, Striated Pardalote 15, 27 and 39 months post-fire. Tawny-crowned and Pied Currawong). Another group showed an Honeyeaters were absent until 56 months initial decline with recovery after one or more post-fire. For some species, colonisation was not years (including Golden Whistler, Superb from unburnt refuges to burnt margins to interior Fairy-wren, Striated Thornbill, Crescent of burnt areas, but rather they first appeared deep Honeyeater, New Holland Honeyeater, Spotted in the burnt forests. Fire effects are greatest for Pardalote for recovery after one year, and Eastern ground-dwelling birds, especially those with Yellow Robin, White-throated Treecreeper and restricted distributions, narrow habitat ranges and Eastern Spinebill for recovery after 2-3 years). limited dispersal ability (e.g. Rufous Bristlebird,

159 Australia’s Biodiveristy - Responses to Fire

Southern Emu-wren). Too frequent fires alter Kookaburra, Pied Currawong, Australian Magpie, vegetation and change bird assemblages. If fire is Australian Raven, Magpie-lark, Eastern Rosella, to be used as a management tool, pockets of Crimson Rosella. These fed on the seeds shed unburnt country are vital. Burning during the after fire (the rosellas) or invertebrate and breeding season should be avoided. vertebrate prey killed or exposed by the fire. White-eared Honeyeaters increased in the burnt heath from occasional visitors to nesting residents Ridpath, M. G. (1972). The effects of fire on about 1 year post-fire. fauna. In R. J. Hooper & M. Rowell (Eds.), Proceedings of the 1971 Tropical and Arid Fire Symposium. (pp. 64-66). Darwin: NT Robertson, B. I. (1981). Seabird Islands: No. Government Printer. 102. , Furneaux Group, historical change; review; Orange-footed Scrub-fowl; Tasmania. Corella, 5, 49-52. Wedge-tailed Eagle; Ground Parrot seabirds; tussock grasslands; heath Fire regimes change habitat configurations and Bioregion: Furneaux relative extent. For example, abandoned mounds The island was burnt by quail-shooters in 1968, of Orange-footed Scrubfowl indicate that affecting breeding seabirds. rainforests have declined over the last 8,000 years because of Aboriginal burning and/or climate change. In arid environments, burning may Robinson, A. C., Casperson, K. D., & Copley, reduce the number of rabbits, leading to local P. B. (1990). Breeding records of the decline in Wedge-tailed Eagle population. Malleefowl (Leipoa ocellata) and Ground Parrots need heaths burnt every 4-5 years. Scarlet-chested Parrot (Neophema splendida) within the Yellabinna Wilderness area, South Australia. South Australian Ornithologist, 31, Ridpath, M. G. (1974). The ecological 8-12. consequences of fire for animal communities. Scarlet-chested Parrot; mallee; hummock grassland. In R. E. Fox (Ed.), Report on the use of fire in Bioregion: national parks and reserves. (pp. 48-53). Scarlet-chested Parrot may show some association Darwin: Department of the Northern with young regrowth (3-5 years post-fire). Territory. review Robinson, D. (1992). Habitat use and foraging behaviour of the Scarlet Robin and the Flame Roberts, P. E. (1970). Some effects of a Robin at a site of breeding-season sympatry. bushfire on heathland birdlife. Proceedings of Wildlife Research, 19, 377-395. the Royal Zoological Society of New South Wales, Flame Robin; Scarlet Robin 89, 40-43. Bioregion: South eastern Highlands heath; wildfire Wildfire or fuel reduction burning may change Bioregion: Sydney Basin the relative competitive ability of these two similar Up to 2.5 years after fire, eight previously resident species. bird species disappeared or declined: Eastern Whipbird, White-cheeked Honeyeater (occasional visits to flowering Xanthorrhoea since Robinson, D., & Woinarski, J. C. Z. (1992). A fire), Little Wattlebird, Eastern Spinebill review of records of the Northern Shrike-tit (occasionally feeding in regrowth heath), Falcunculus frontatus whitei in northwestern Variegated Fairy-wren, Chestnut-rumped Australia. South Australian Ornithologist, 31, Heath-wren (one pair returned to regrowth heath 111-117. 1 year after fire), Brown Thornbill (declined in Crested Shrike-tit; tropical eucalypt open forest; abundance in heath), Red-browed Finch threatened species (completely disappeared from the locality). Crested Shrike-tits favour old forests in temperate Several birds from adjacent woodlands Australia. Frequent burning of tropical eucalypt temporarily moved into the heath immediately forests may be one reason for the rarity of this (and up to 3 months) after fire: Laughing subspecies, through loss of large trees. More data

160 Fire and Australian birds

are needed on the long-terms effects of fire on Rowley, I., & Brooker, M. (1987). The and hence bark-foraging birds. response of a small insectivorous bird to fire in heathlands. In D. A. Saunders, G. W. Arnold, A. A. Burbidge, & A. J. M. Hopkins Rounsevell, D. E., & Woinarski, J. C. Z. (1983). (Eds.), Nature Conservation: the role of Status and conservation of the Forty-spotted remnants of native vegetation (pp. 211-218). Pardalote, Pardalotus quadragintus (Aves: Chipping Norton: Surrey Beatty. Pardalotidae). Australian Wildlife Research, 10, Splendid Fairy-wren; heath; wildfire; isolate; 343-349. management Forty-spotted Pardalote; threatened species; eucalypt Bioregion: Jarrah Forest open forest; isolate Six fires (one of which was intense) burnt parts of Bioregion: Freycinet the study area over a 12 year study of Wildfire may threaten all populations of colour-banded Splendid Fairy-wrens. In most Forty-spotted Pardalote. One small isolated cases, unburnt patches were left. But in both population may have disappeared after fire. territories that were completely burnt out, most of However, population numbers at another site the previous occupiers disappeared (and were were similar 2 years post-fire to pre-fire levels presumed to have perished). The incidence of (possibly because of proximity of unburnt areas parasitism in the years following fire was low. with relatively large populations). Fuel-reduction With the exception of adult survival, all measures burns are probably not detrimental. of density, productivity and survival increased during fire-free periods. The major change occurred in the fourth year post-fire and was Rowley, I. (1987). Conservation of the sustained thereafter. High adult survival allows Purple-crowned Fairy-wren Malurus coronatus population maintenance despite several years of in northern Australia. World Wildlife Fund low productivity post-fire. However, frequent fires (Australia). eliminated some populations at the site, with Purple-crowned Fairy-wren; riparian vegetation; subsequent restoration of territories arising from threatened species; pandanus; tussock grassland; tropical recruitment from beyond the site. eucalypt savanna woodland This study also reports short-term (9 Bioregions: Victoria Bonaparte; Ord-Victoria months) response to an intense wildfire which Plains; ; North Kimberley; consumed 95% of the vegetation. The population Gulf Fall and Uplands; Gulf Coastal survived the very hot (>600oC at 10cm above Fire may destroy or degrade some of the riparian ground) fire well, showed remarkable site tenacity, strips on which this species depends, though the and showed similar or even reduced mortality in main threat is degradation of this habitat by cattle. the months post-fire than in previous years. The better short-term survival after a hot fire than Rowley, I. (1993). The Purple-crowned after cooler burns may have been due to the Fairy-wren Malurus coronatus. I. History, timing of the hot fire, which was several months distribution and present status. Emu, 93, after breeding season. 220-234. Frequent burns had a deleterious effect on Purple-crowned Fairy-wren; pandanus; tussock this population. If the site had been a remnant, grassland; threatened species; tropical eucalypt savanna the Fairy-wrens may well have been eliminated by woodland this fire regime. Optimum density may be Bioregions: Victoria Bonaparte; Ord-Victoria maintained with at least 12 year intervals between Plains; Central Kimberley; North Kimberley; fires, which should not be during the breeding Gulf Fall and Uplands; Gulf Coastal season. The Purple-crowned Fairy-wren is found in riparian vegetation (Pandanus and cane-grasses) in northern Australia. Heavy grazing by cattle and burning by pastoralists removes both shelter and foraging substrate for this species, forcing them to abandon this habitat at critical times of the year, leading to local declines and extinction.

161 Australia’s Biodiveristy - Responses to Fire

Rowley, I., & Russell, E. (1993). The a number of minor wildfires and one which burnt Purple-crowned Fairy-wren Malurus 95% of the study area. Annual survival decreased coronatus. II. Breeding biology, social in the two years after fire. Females laid more organisation, demography and management. clutches when predation or brood-parasitism was Emu, 93, 235-251. high, particularly in the 2 years post-fire, but Purple-crowned Fairy-wren; tussock grassland; productivity was low for several years post-fire. pandanus; management; threatened species Territories were not vacated initially after fire, but Bioregions: Victoria Bonaparte; Ord-Victoria by 3-5 years post-fire some territories were vacant Plains; Central Kimberley; North Kimberley; and group sizes decreased. Population decline Gulf Fall and Uplands; Gulf Coastal occurred 3-5 years post-fire due to decreased Purple-crowned Fairy-wrens require riparian reproductive output and replacement of older habitat which is easily damaged by stock or fire. breeding birds by less productive novices. Fire This habitat should be protected from fire at all directly affected natality and juvenile survival and times. indirectly affected population density, age structure, sex ratio and group composition.

Rowley, I., & Russell, E. (1995). The breeding biology of the White-winged Fairy-wren Russell-Smith, J. (1985). A record of change: Malurus leucopterus leuconotus in a Western studies of Holocene vegetation history in the Australian coastal heathland. Emu, 95, South Alligator Region, Northern Territory. 175-184. Proceedings of the Ecological Society of White-winged Fairy-wren; heath; wildfire Australia., 13, 191-202. Bioregion: Swan Coastal Plain Orange-footed Scrubfowl; historic change; rainforest All White-winged Fairy-wrens from a 32ha study Bioregion: Top End Coastal site (77 adults banded in the 4 years before fire) Historic changes in fire regime have led to could not be relocated shortly after a wildfire, and contraction of monsoon rainforests, as evidenced only one wren was seen there 3 months after the by old scrubfowl mounds occurring in eucalypt fire. open forests.

Rowley, I., Russell, E., Brown, R., & Brown, Russell-Smith, J. (1995). Fire management. In M. (1988). The ecology and breeding biology T. Press, D. Lea, A. Webb, & A. Graham of the Red-winged Fairy-wren Malurus (Eds.), Kakadu: natural and cultural heritage elegans. Emu, 88, 161-176. and management (pp. 217-237). Darwin: Red-winged Fairy-wren; control burning; eucalypt Australian Nature Conservation Agency and open forest North Australia Research Unit. Bioregion: Warren management; review; tropical eucalypt open forest; In the long term, fuel-reduction burns must have tussock grassland; Aboriginal knowledge; Aboriginal severe effects on populations of Red-winged burning Fairy-wren, because nearly half of the nest sites in Bioregions: Pine Creek - Arnhem; Top End this study were in “dead brush”, the fuel that such Coastal burns seek to reduce. This species occurs in a very Reviews burning regimes used traditionally by stable environment and has very low reproductive Aboriginal people and currently by park output: it may be a very poor recoloniser. managers; and some studies of the effects of fire on birds in this region.

Russell, E. M., & Rowley, I. (1993). Demography of the cooperatively breeding Saunders, D. A. (1979). The availability of tree Splendid Fairy-wren, Malurus splendens hollows for use as nest sites by White-tailed (Maluridae). Australian Journal of Zoology, 41, Black Cockatoos. Australian Wildlife Research, 475-505. 6, 202-216. Splendid Fairy-wren; wildfire hollows; White-tailed Black-cockatoo Bioregion: Swan Coastal Plain Studied a population for 18 years, which included

162 Fire and Australian birds

Saunders, D. A. (1985). Human impact: the Schodde, R., Mason, I. J., & Wood, J. T. response of forest and woodland bird (1993). Geographical differentiation in the communities. Whither the future? A Glossy Black-Cockatoo Calyptorhynchus synthesis. In A. Keast, H. F. Recher, H. Ford, lathami (Temminck) and its history. Emu, 93, & D. Saunders (Eds.), Birds of eucalypt forests 156-166. and woodlands: ecology, conservation, Glossy Black-Cockatoo; threatened species; management (pp. 355-357). Chipping Norton: management; hollows Surrey Beatty. Management for the Glossy Black-Cockatoo review; eucalypt open forest; management should include developing a fire regime that Much more research is needed on the role of fire prevents severe wildfires (and hence elimination in bird conservation and management. “Minor of food sources), and maintaining availability of changes in the fire regime may be as damaging as hollows. the widespread clearing for agriculture.”

Scotts, D. (1994). Sustaining sensitive wildlife Saunders, D. A., Smith, G. T., & Rowley, I. within temperate forest landscapes: regional (1982). The availability and dimensions of tree systems of retained habitat as a planning hollows that provide nest sites for cockatoos framework. In T. W. Norton & S. R. Dovers (Psittaciformes) in Western Australia. (Eds.), Ecology and sustainability of southern Australian Wildlife Research, 9, 541-556. temperate ecosystems (pp. 85-106). Canberra: hollows CSIRO. The smallest trees suitable for nesting by any of 5 old-growth; forestry; management; eucalypt open forest species of cockatoos considered was at least 130 Mainly concerned with management of forestry, years old. Such trees are becoming scarcer due to but provides a review of old-growth characteristics vegetation clearing and prevention of and the association of birds with these. regeneration (due to stock trampling or grazing).

Scotts, D. J. (1991). Old-growth forests: their Schodde, R. (1982). The Fairy-wrens. A ecological characteristics and value to monograph of the Maluridae. Melbourne: forest-dependent vertebrate fauna of Lansdowne. south-east Australia. In D. Lunney (Ed.), Mallee Emu-wren; Carpentarian Grass-wren; Conservation of Australia’s forest fauna (pp. threatened species; management; hummock grasslands; 147-159). Sydney: Royal Zoological Society of review. NSW. The Mallee Emu-wren and Carpentarian eucalypt open forest; old-growth; forestry; hollows Grass-wren are probably threatened by fire. Both Bioregions: South Eastern Highlands; South East may recover very slowly after fire has consumed Corner their highly flammable hummock grassland Old-growth forests are characterised by abundant habitat. This may be especially threatening where litter, logs and hollows, and distinctive floristic populations are highly fragmented. Southern composition. 17 bird species are listed as finding Emu-wren and Striated Grass-wren may also be optimum habitat in old-growth forests. These adversely affected by wildfire, though this does include hollow-nesting species, honeyeaters, not yet threaten the species existence. Both may trunk-gleaning insectivores, frugivores and some take many years to recolonise. The Southern insectivores of the canopy foliage. Emu-wren is reported to well before the fire front.

163 Australia’s Biodiveristy - Responses to Fire

Silveira, C. E. (1993). Recovery plan for Skira, I. J., & Brothers, N. P. (1988a). Seabird threatened mallee birds - addressing fire regimes. Islands: No. 183. , Royal Australasian Ornithologists Union. Furneaux Group, Tasmania. Corella, 12, mallee; heath; threatened species; management; review; 80-81. Malleefowl; Pink Cockatoo; Regent Parrot; Mallee seabirds; tussock grassland Emu-wren; Striated Grass-wren; Black-eared Miner; Bioregion: Furneaux Slender-billed Thornbill; Western Whipbird; Red-lored The island is regularly fired, and burning and Whistler. grazing by sheep have caused much deterioration Bioregions: Murray-Darling Depression; Mallee of vegetation and soil, affecting breeding Provides a review of the observed association of shearwaters. threatened mallee birds (Malleefowl, Pink Cockatoo, Regent Parrot, Mallee Emu-wren, Striated Grass-wren, Black-eared Miner, Skira, I. J., & Brothers, N. P. (1988b). Seabird Slender-billed Thornbill, Western Whipbird and Islands: No. 184. Great Dog Island, Furneaux Red-lored Whistler) with fire, and research and Group, Tasmania. Corella, 12, 82-84. management requirements. For the Malleefowl, a seabirds; tussock grassland high frequency of fire is a confirmed threat in Bioregion: Furneaux mallee of eastern Australia, but has not been Frequent burning has caused fewer areas to be established in western mallee. For the Pink used by breeding shearwaters, however burning is Cockatoo, primary threats are the loss of suitable now less frequent due to management control by hollows through natural attrition and the state conservation agency. prevention of formation of new hollows due to increased frequency of fire. The Mallee Smith, G. T. (1977). The effect of Emu-wren and Striated Grass-wren are absent in environmental change on six rare birds. Emu, mallee with hummock grassland understorey for 77, 173-179. 5-6 years post-fire, but persist then until the review; historic change; Aboriginal burning; Noisy vegetation is long-unburnt. Recolonisation may Scrub-bird; Rufous Scrub-bird; Western Bristlebird, be a problem for these species with poor Eastern Bristlebird, Rufous Bristlebird; Western dispersive ability following large-scale fires which Whipbird; threatened species; management eliminate populations. Black-eared Miners Examines status, habitat requirements and history apparently prefer long-unburnt mallee for of Noisy Scrub-bird, Rufous Scrub-bird, Western breeding, though may feed in younger regrowth. Bristlebird, Eastern Bristlebird, Rufous Bristlebird In southwestern Australia, Western Whipbirds and Western Whipbird. The Scrub-birds and prefer mallee-heath older than 14 years, and may Bristlebirds are Tertiary relicts, more common in prefer much older vegetation. closed forests of the Tertiary, and now largely restricted to very narrow specialised habitats. Silveira, C. E. (1995). The Black-eared Miner. Aboriginal entry to Australia had a major impact Australian Bird Watcher, 16, 96-109. on the fauna during the Pleistocene and Recent by Black-eared Miner; threatened species; mallee “prolonged, constant and widespread use of fire”. Bioregion: Murray-Darling Depression One consequence was a possible increase in the In contrast to previous claims (McLaughlin 1992) area of heath (to the advantage of several that the Black-eared Miner is restricted to mallee heath-favouring birds, including Western at least 55 years post-fire, this paper reports Bristlebird and Eastern Bristlebird). However, foraging by Black-eared Miners in mallee of 3-11 burning regime of heaths became much more years post-fire (though these birds apparently frequent with European colonisation and led to nested in adjacent long-unburnt mallee). the decline of bristlebirds and scrub-birds. Western Bristlebirds require heath of at least 8-10 years post-fire before habitat is suitable. For Noisy Scrub-birds fire intervals of <5 years or >30 years render heath/thicket unsuitable.

164 Fire and Australian birds

Smith, G. T. (1979a). The Noisy Scrub-bird. Smith, G. T. (1985a). Fire effects on In M. J. Tyler (Ed.), The status of endangered populations of the Noisy Scrub-bird Australasian wildlife (pp. 117-121). Adelaide: (Atrichornis clamosus), Western Bristle-bird Royal Zoological Society of South Australia. (Dasyornis longirostris) and Western Noisy Scrub-bird; heath; thicket; isolate; threatened Whip-bird (Psophodes nigrogularis). In J. R. species; eucalypt open forest; management; historic Ford (Ed.), Fire ecology and management in change; Aboriginal burning Western Australian ecosystems (pp. 95-102). Bioregion: Jarrah Forest Perth: Western Australian Institute of Aboriginal use of fire probably had little impact Technology. on Noisy Scrub-bird habitat (as such fires were Noisy Scrub-bird; Western Bristlebird; Western probably small and lit after the breeding season). Whipbird; heath; thicket; eucalypt open forest; After European colonisation, heaths were burnt management; historic change; threatened species; every 2-3 years to provide grass for cattle, while Aboriginal burning attempts made to exclude fire in forests probably Bioregions: Esperance Plains; Warren; Swan resulted in higher frequency of occasional hot Coastal Plain; Jarrah Forest wildfires. These changes would have led to the There have been few studies of the effects of fire decline and local extinction of Noisy Scrub-birds, on birds, and most are opportunistic and and these losses would have further fragmented short-term. Western Whipbird, Western populations and made recolonisations after fires Bristlebird and Noisy Scrub-birds are all poor less likely. Recent increases are due to absence of dispersers. Change in fire regime from that used fire. However, lack of fire may allow successional by Aborigines to that used by settlers resulted in changes that render gullies unsuitable for severe contraction of range. These three species Scrub-birds (this has already happened at two survive at Two Peoples Bay, largely through small sites). topographic protection from fire. Recent fire control in this area has led to the expansion of populations of all three species. Vegetation has to Smith, G. T. (1979b). The status of Australian be 4-10 years post-fire before being suitable for parrots. In M. J. Tyler (Ed.), The status of Noisy Scrub-birds (being faster in wet gullies endangered Australasian wildlife (pp. 101-108). where dominant eucalypts have not been Adelaide: Royal Zoological Society of South destroyed). Reproductive output is low, hence Australia. recruitment after fire may be very gradual. Wet Ground Parrot; review; heath; threatened species; heath may be suitable for Western Bristlebirds 3 management years post-fire and dry heath by 6-10 years Too frequent burning renders heath unsuitable for post-fire, though grazing pressure from Grey Ground Parrot, and heath over-protected from Kangaroos (especially if burnt areas are limited) fire also becomes unsuitable as it becomes too tall may extend these periods. For Western and dense. Management needs to mosaic burn at Whipbirds, minimum periods before vegetation 8-10 year intervals. was suitable were 4-6 and 7 years post-fire. In heath, there is a bird succession from Richard’s Pipit to Striated Field-wren to Western Bristlebird to Western Whipbird. The long-unburnt part of this succession is fairly speculative as there are few very old heaths. Some >45 year post-fire heaths retain Bristlebirds, but at densities less than those in 20 year old heaths. Less is known of later succession in the low eucalypt forests used by Noisy Scrub-birds. A fire interval of at least 20 years (and possibly of the order of 50 years) should be maintained, and accompanied by considered monitoring.

165 Australia’s Biodiveristy - Responses to Fire

Smith, G. T. (1985b). The Noisy Scrub-bird Smith, G. T. (1987a). The changing Atrichornis clamosus. Does it’s past suggest a environment for birds in the south-west of future? In A. Keast, H. F. Recher, H. Ford, & Western Australia; some management D. Saunders (Eds.), Birds of eucalypt forests and implications. In D. A. Saunders, G. W. Arnold, woodlands: ecology, conservation, management A. A. Burbidge, & A. J. M. Hopkins (Eds.), (pp. 301-308). Chipping Norton: Surrey Nature Conservation: The role of remnants of Beatty. native vegetation (pp. 269-277). Chipping Noisy Scrub-bird; threatened species; eucalypt open Norton: Surrey Beatty. forest; wetlands; heath; management; Aboriginal historic change; eucalypt open forest; heath; thicket; burning; historic change Noisy Scrub-bird; Western Whipbird; Western Bioregions: Jarrah Forest; Warren Bristlebird; management; threatened species; Compares Aboriginal fire regimes (small, patchy) Aboriginal burning with settler fire regimes (frequent in wet heaths Bioregions: Esperance Plains; Jarrah Forest; and swamp margins; hot summer fires and more Warren frequent occurrence of bushfires in open forest) in Change from Aboriginal burning regimes to those the restricted habitat of Noisy Scrub-bird. associated with Europeans led to epidemic of Topographic protection has resulted in low intense fires in the early years of settlement. In frequency of fire at one site and hence the survival coastal areas, early pastoralists fired the heaths every of one relict population. Minimum age of 2-3 years. Clearing and fragmentation have left the vegetation to be suitable is known, but there is remnants more vulnerable to catastrophic fire. also indication that long-term fire exclusion may Changes in burning regime are thought to be the result in reduction of habitat suitability. main cause of decline for Noisy Scrub-bird, Western Bristlebird, Western Whipbird and possibly Ground Parrot; and probably for the Smith, G. T. (1985c). Population and habitat regional extinction of Rufous Bristlebird. Discusses selection of the Noisy Scrub-bird, Atrichornis management of Two Peoples Bay for the threatened clamosus, 1962-83. Australian Wildlife Noisy Scrub-bird, Western Bristlebird and Western Research, 12, 479-485. Whipbird. Recent large wildfires in nearby areas Noisy Scrub-bird; heath; thicket; eucalypt open forest; severely reduced or wiped out populations of threatened species Western Bristlebird. Notes the importance of Bioregion: Jarrah Forest natural firebreaks (rocky hills, dissescted gullies) in Exclusion of fire has led to substantial vegetation providing protection from fire for these remnant change and marked increases in populations of populations. Recolonisation period post-fire Noisy Scrub-bird. Six years after a hot fire, depends on a number of factors (edaphic, proximity regeneration was suitable in some, if not all, areas of source populations, refuge areas). Fine scale for Noisy Scrub-bird. But the first male to habitat manipulation of old areas may maintain recolonise was not recorded until 9 years post-fire, their suitability for Noisy Scrub-birds. Older heaths suggesting problems of dispersal from unburnt (45 years post-fire) may have decreased productivity areas. Nothing is known about how long and hence lowered suitability for Western vegetation remains suitable, but it is likely to be Whipbirds and Western Bristlebirds. Fire intervals greater than 40 years. of <10 years will lead to local decline and possible extinction for these 3 species. Fire intervals of >50 years may also be deleterious. Fire management should maintain firebreaks and prevent one fire burning the whole reserve. Long intervals between fires will disadvantage some other early successional birds (e.g. Richard’s Pipit, Striated Fieldwren). Long-term fire exclusion may also change the relative extent of heaths and thickets to the disadvantage of one or other of the threatened species. The longevity of many Australian birds means that many which survive fire may live long enough to recolonise. Describes three requirements

166 Fire and Australian birds

for management: detailed information on habitat Smith, G. T., & Forrester, R. I. (1981). The requirements, data on type and rate of vegetation status of the Noisy Scrub-bird Atrichornis change post-fire, and information on the clamosus. Biological Conservation, 19, 239-254. relationship of birds with fire regime. Noisy Scrub-bird; heath; eucalypt open forest; thicket; threatened species; management Bioregion: Jarrah Forest Smith, G. T. (1987b). Observations on the Fire suppression at remaining colony sites has biology of the Western Bristlebird Dasyornis helped to increase the numbers of Noisy longirostris. Emu, 87, 111-118. Scrub-birds. Western Bristlebird; heath; thicket; threatened species; historic change Bioregions: Esperance Plains; Jarrah Forest Smith, G. T., & Robinson, F. N. (1976). The Western Bristlebirds have disappeared from much Noisy Scrub-bird: an interim report. Emu, 76, of their range over the last 100 years, due to past 37-42. practice of burning the coastal heaths and thickets Noisy Scrub-bird; eucalypt open forest; threatened every two or three years to provide cattle feed. species; thicket; heath; historic change Clearing and draining swamps have also Bioregion: Jarrah Forest contrubuted to the decline. Home ranges were The dramatic disappearance of the Noisy established 9 years post-fire (adjacent to gullies) Scrub-bird soon after European colonisation and 11-14 years post-fire (drier areas) at sites certainly resulted from frequent devastating fires relatively distant from existing populations, and and the clearing and burning of eucalypt forests. 4-6 years post-fire at sites adjacent to or within However total fire exclusion may eventually 2km of existing populations. The preferred short render habitat unsuitable. closed heaths are suitable 3 years after fire (in wet areas) or 6-10 years after fire (drier areas). Oldest suitable age is not known, but birds are present in Smith, L. H. (1994). A critical analysis of the 45 year heath (although less abundant than in 20 factors responsible for the decline of the year old heath). While old heaths are probably Superb Lyrebird Menura novaehollandiae in structurally suitable, their productivity may have Sherbrooke Forest, Victoria. Australian Bird declined. Thickets and more swampy vegetation Watcher, 15, 238-249. may be less vulnerable to fire and provide Superb Lyrebird; eucalypt open forest; control burning; important refuges after fire. management Bioregion: South eastern Highlands Population of Superb Lyrebirds has declined, Smith, G. T. (1991). Ecology of the Western partly through fox predation and through Whipbird Psophodes nigrogularis in Western understorey change (reduction in open leaf litter Australia. Emu, 91, 145-157. and increase in grass cover), during a period when Western Whipbird; mallee; heath; thicket former frequent burning was replaced by fire Bioregions: Esperance Plains; Jarrah Forest exclusion. A firebreak maintained by burning and Western Whipbirds in southwestern Australia are slashing was abandoned by Lyrebirds when it most abundant in dense coastal thickets. became overgrown following cessation of burning. Population increases at some sites are due to absence of fire (up to at least 50 years). In general fire intervals of less than 10 years will lead, or has Smith, P. (1989). Changes in a forest bird led, to extinction of local populations. Notes that community during a period of fire and fire effects will differ according to vegetation type drought near Bega, New South Wales. (for example between Victorian mallee and Australian Journal of Ecology, 14, 41-54. southwestern Australian heath). eucalypt open forest; community; rainforest; wildfire; succession Bioregion: South East Corner A forest bird community was monitored for 3 years, during which it was partly burnt by wildfire and affected by severe drought. The fire affected mainly the undergrowth, especially on ridges:

167 Australia’s Biodiveristy - Responses to Fire

gullies were mainly left unburnt. Changes in the Turquoise Parrot, Southern Scrub-robin, Shy bird community were limited. Differences in Heath-wren, Chestnut Quail-thrush, species were restricted to rare visitors. The Black-winged Currawong, Gilbert’s Whistler, distribution of insectivorous birds contracted (and Jacky Winter and Golden Whistler). the species declined) in the year following fire; nectarivorous and frugivorous species expanded and became more abundant (due to prolific Smith, P. J., Smith, J. E., & Pressey, R. L. flowering and fruiting in the second year of the (1995). Birds of particular conservation concern study, possibly because of the fire). 10 species in the Western Division of New South Wales: (Eastern Yellow Robin, Crested Shrike-tit, distribution, habitats and threats. NSW Golden Whistler, Rufous Fantail, Superb National Parks and Wildlife Service. Fairy-wren, White-browed Scrubwren, Brown threatened species; review; mallee; eucalypt woodland; Gerygone, Brown Thornbill, Lewin’s Honeyeater, Acacia woodland; hummock grassland; tussock Silvereye) declined in burnt areas in the 2 years grassland; management following fire: all, other than Crested Shrike-tit, Bioregions: Murray-Darling Depression; Broken were birds which feed, nest or shelter in dense Hill Complex; Mulga Lands; Channel Country; shrubby understorey. For seven resident species, Simpson-Strzelecki Dunefields. colour-banding allowed assessment of changes in Reviews the threatening processes for threatened territory. For 6 of these species (especially Brown bird taxa in the Western Division of NSW. Fire is Thornbill), the number of territories declined one of the major threats, particularly for mallee post-fire, with loss of colour-banded birds from birds. the site and/or contraction to unburnt patches. In the year post-fire many species bred at atypical Smith, S., & Baker-Gabb, D. (1993). Rufous times, possibly related to flush of resources. Bristlebird Dasyornis broadbenti. Department Recovery of populations had not reached pre-fire of Conservation and Natural Resources. levels two years after fire, which is slower than Rufous Bristlebird; threatened species; thicket; control that reported in other studies. This may have been burning; wildfire. because of drought, and suggests that a range of Bioregions: Naracoorte Coastal Plain; South east variables may affect recolonisation. The resilience Coastal Plain of the community was related to the role of Reviews information on the species (in Victoria). unburnt areas, particularly wet gullies. Wildfire or inappropriate (control) burning regimes are a threat throughout its range. Smith, P. J., Pressey, R. L., & Smith, J. E. Management requires determination and (1994). Birds of particular conservation implementation of a preferred fire regime, concern in the Western Division of New however this may conflict with protection of South Wales. Biological Conservation, 69, property (or current legislation). 315-338. threatened species; review; mallee; eucalypt woodland; Sonter, C. (1984). The Yellow-rumped Acacia woodland; management Pardalote - a declining bird in Sunraysia? Bioregions: Murray Darling Depression; Broken Australian Bird Watcher, 10, 234-235. Hill Complex; Mulga Lands; Channel Country; Yellow-rumped Pardalote; wildfire; mallee; eucalypt Simpson-Strzelecki Dunefields woodland Reviews change in abundance and threats for bird Bioregion: Murray-Darling Depression species in the Western Division of New South Combination of bushfires, clearing and drought Wales. The most frequently-mentioned threats may have led to the decline of this species. are overgrazing, introduced predators, clearing and altered fire regimes (in that order). Fire is a threat particularly to mallee birds, especially the frequent occurrence of extensive wildfires. Altered fire regime is listed as a threat for 15 taxa (Black-eared Miner, Malleefowl, Red-lored Whistler, Regent Parrot, Scarlet-chested Parrot, Striated Grass-wren, Glossy Black-cockatoo,

168 Fire and Australian birds

Specht, R. L. (1981). Responses to fires in Starks, J. (1987). The status and distribution of heathlands and related shrublands. In A. M. the Black-eared Miner (Manorina melanotis) in Gill, R. H. Groves, & I. R. Noble (Eds.), Fire Victoria. Technical report no.49. Arthur Rylah and the Australian biota (pp. 395-415). Insitute of Environmental Research. Canberra: Australian Academy of Science. Black-eared Miner; mallee; threatened species heath; review Bioregion: Murray Darling Depression Yellow-tailed Black-cockatoos visited a heathland The endangered Black-eared Miner is associated soon after fire to feed from grubs attracted to the with long-unburnt mallee, and the limited extent fire-induced flowering of Xanthorrhoea. of such old vegetation is a factor in its threatened Honeyeaters recolonised when woody plants status. replaced the more herbaceous short-lived plants. Black cockatoos returned to mature heaths to tear apart Banksia flowers. Stephens, S. (1992). Endangered species and communities and threatening processes in the Murray Mallee. Australian National Parks and Stanton, J. P. (1992). J.P. Thomson oration. Wildlife Service. The neglected lands: recent changes in the mallee; review; management; threatened species ecosystems of Cape York Peninsula and the Bioregion: Murray Darling Depression challenge of their management. Journal of the Altered fire regime has contributed to the decline Queensland Geographical Society, 7, 1-18. of 8 (sic) bird species (Beautiful Firetail, Cassowary; management; threatened species; tropical Black-eared Miner, Chestnut Quail-thrush, eucalypt open forest; rainforest; hollows. Chestnut-rumped Heath-wren, Crested Bioregion: Cape York Peninsula Shrike-tit, King Quail, Mallee Emu-wren, The maintenance of the preferred habitat of the Mallee-fowl, Night Parrot, Orange-bellied Cassowary (ecotone between rainforest and wet Parrot, Red-lored Whistler, Red-tailed eucalypt open forest) requires a regime including Black-cockatoo, Rufous Bristlebird and Rufous frequent fire. Preservation of trees bearing Field-wren) in the Murray Mallee (the hollows suitable for nesting birds may require fourth-ranking threat after overgrazing, control burning at times when fuel moisture is introduced predators and clearing). Includes high. discussion of management and research options and priorities.

Stanton, P. (1995). A tropical Queensland perspective. In D. B. Rose (Ed.), Country in Stocker, G. C. (1971). The age of charcoal flames. Proceedings of the 1994 symposium on from old jungle fowl nests and vegetation biodiversity and fire in North Australia (pp. change on Melville Island. Search, 2, 28-30. 71-76). Canberra: Department of the Orange-footed Scrubfowl; charcoal; historic change; Environment, Sport and Territories, and rainforest; Aboriginal burning North Australia Research Unit. Bioregion: Top End Coastal tropical eucalypt open forest; rainforest; tussock Recent changes in fire regimes have led to retreat grassland; management; threatened species; of rainforests, marked by the occurrence of Golden-shouldered Parrot; Palm Cockatoo; Cassowary. abandoned mounds of scrubfowl in areas which Bioregion: Cape York Peninsula are now eucalypt open forests. Present fire regime is leading to expansion of rainforest at the expense of wet eucalypt forests, to the detriment of Palm Cockatoo and Cassowary. Stokes, T. (1975). The effect of a bushfire on It is also leading to invasion of grassy flats by the banding of Flame Robins in the Melaleuca, threatening the Golden-shouldered Brindabella Ranges. Australian Bird Bander, Parrot. 13, 75-76. Flame Robin; wildfire; eucalypt open forest Bioregion: South Eastern Highlands In 28 months following hot wildfire in montane eucalypt forest, Flame Robins were far more abundant in burnt than unburnt areas.

169 Australia’s Biodiveristy - Responses to Fire

Suckling, G. C., & MacFarlane, M. A. (1983). Taylor, R. J. (1991). Fauna conservation in The effects of fire on fauna - a review. In E. production forests in Tasmania. Hobart: Ealey (Ed.), Fighting fire with fire. (pp. Forestry Commission Tasmania. 107-128). Melbourne: Monash University. review; forestry; eucalypt open forest; control burning; review; eucalypt open forest; wildfire; control burning; management management Reviews impacts of control burning and wildfire Few studies of effects of fire on fauna other than on birds (and other fauna) in eucalypt forests. In mammals. Summarises existing data for birds, in general, effects of mild control fires are terms of effect of fire intensity, frequency and short-term and minor, though long-term changes season. Few birds are killed by low intensity fires, in understorey as a result of frequent control fires but many may subsequently die through predation may lead to more substantial changes in the bird or starvation. In contrast, immediate mortality assemblage. Regeneration of wet sclerophyll may be high with intense fires. Fire intensity eucalypt forests may require hot burns. Species affects vegetation structure, which has more such as and hollow-nesting species important long-term implications for birds. More require, or are more abundant in, old-growth refuge areas may be left by low intensity fires. forests. Some birds (e.g. Flame Robin) prefer severely burnt areas to unburnt forests. Return of bird communities is probably quicker after Tidemann, S. C. (1990). Relationships low-intensity fires. In open forests, understorey between finches and pastoral practices in may require more frequent fires than overstorey, northern Australia. In J. Pinowski & J. D. and the number of birds in the understorey may Summers-Smith (Eds.), Granivorous birds and decline in old forests. Ground-feeding birds (e.g. agriculture (pp. 305-315). Warsaw: Superb Lyrebird, Superb Fairy-wren, Bassian PWN-Polish Scientific Publishers. Thrush, Eastern Yellow Robin) may require tropical eucalypt savanna woodland; tussock grassland; periodic fires. Other species (e.g. Noisy management Scrub-bird) cannot survive frequent fire. There Bioregions: Victoria Basin; Ord-Victoria Plains; are no data on the time of year at which birds are Sturt Plateau most susceptible to fire, though effects may be Across 21 trapping sites, there was no relationship greatest during the breeding season. between number of finches and a general measure of fire extent on pastoral properties (although this measure was not related closely to the fire history Swanson, N. M. (1976). Seabird Islands: No. of the actual site); in contrast, there was a 32. Mutton Bird Island, New South Wales. significant negative relationship with cattle Australian Bird Bander, 14, 88-91. density. seabirds; tussock grassland Bioregion: NSW North Coast Fires have led to erosion and spread of exotic Tidemann, S. C. (1992). Conservation of the grasses. Nesting birds took about 4 years to fully Gouldian Finch, N.T. World Wide Fund for re-use a site that had been severely burnt. Nature. Gouldian Finch; threatened species; tropical eucalypt savanna woodland; management; control burning Tarr, H. E. (1965). The Mallee-Fowl in Bioregions: Daly Basin; Victoria-Bonaparte; Wyperfeld National Park. Australian Bird Ord-Victoria Plains Watcher, 2, 140-144. Management of the Gouldian Finch should Malleefowl; threatened species. require controlled burning of the breeding sites Bioregion: Murray-Darling Depression early in the dry season, to prevent late dry season Mallee may be unsuitable for Malleefowl for up to destructive fires. 10-20 years post-fire.

170 Fire and Australian birds

Tidemann, S. C. (1993a). Management of a Tingay, A., & Tingay, S. R. (1982a). Seabird threatened species: the Gouldian Finch Islands: No. 113. Middle Island, Archipelago example. In C. P. Catterall, P. V. Driscoll, K. of the Recherche, Western Australia. Corella, Hulsman, D. Muir, & A. Taplin (Eds.), Birds 6, 49-50. and their habitats: status and conservation in seabirds; tussock grassland; heath; eucalypt open forest Queensland (pp. 123-131). St Lucia: Bioregion: Esperance Plains Queensland Ornithological Socity Inc. A fire burnt uncontrolled on the Island for nearly Gouldian Finch; tropical eucalypt savanna woodland; 4 weeks in January-February 1977, destroying tropical eucalypt open forest; threatened species; most vegetation. “The effect of the holocaust on management; control burning breeding seabirds is not known.” Sorghum seeds are important in the diet of the threatened Gouldian Finch, but their abundance is reduced by early dry season burning, and early Tingay, A., & Tingay, S. R. (1982b). Seabird wet season burning can eliminate annual Islands: No. 118. Hood Island, Archipelago of Sorghum. the Recherche, Western Australia. Corella, 6, 59-60. seabirds; heath Tidemann, S. C. (1993b). Where are Gouldian Bioregion: Esperance Plains Finches after the breeding season? Victorian “Fire is an everpresent threat” to seabirds Naturalist, 110, 238-243. breeding on Hood Island. Gouldian Finch; tropical eucalypt savanna woodland; threatened species; control burning; wildfire Bioregion: Daly Basin Tingay, A., & Tingay, S. R. (1982c). Seabird Within weeks of a hot fire, Gouldian Finch Islands: No. 120. Sandy Hook Island, activity was monitored in a cool burnt (April) and Archipelago of the Recherche, Western a hot burnt (October) area. Early in the morning, Australia. Corella, 6, 63-64. birds tended to feed in the hot burnt area (where seabirds; tussock grassland; heath fallen seeds were less hidden by regrowth), but Bioregion: Esperance Plains would fly to the cool area (which had more foliage The main threat to the breeding seabirds is the in canopies) when disturbed. Later in the risk of fire caused by human visitors. morning, birds moved to feed in the cool burnt area. Temperatures on the ground were hotter in Tingay, A., & Tingay, S. R. (1984). Bird the hot burnt area (to 56°C). Cool fires occur communities in the karri forest of Western during the nesting season, but the hollow-nesting Australia. Australian Conservation Gouldian Finches are unaffected. Foundation. eucalypt open forest; community; forestry; control Tidemann, S. C., McArtney, J., & Smith, I. burning; management (1993). Queensland Gouldian Finches Bioregion: Warren Erythrura gouldiae and air-sac mite Bird distributions were monitored in a series of Sternostoma tracheacolum. Sunbird, 23, 36-40. plots of varying age in Karri forests, especially Gouldian Finch; tropical eucalypt savanna woodland; relating to forestry. Cool burns in mature forest threatened species; management; control burning initially cause a short-term decrease in total Bioregion: Mount Isa Inlier abundance of birds, followed by an increase of Pastoralists burning during the early Wet season small insectivores occurring in low vegetation. may have reduced the availability of Sorghum, These then decline as the shrubs regenerate. reducing the survival of Gouldian Finch. Species richness and abundance peaked at 6 years post-fire. However some species were more common in long-unburnt forests. The bird communities in mature forests showed more seasonal stability than those in 50-year old regeneration.

171 Australia’s Biodiveristy - Responses to Fire

Tolhurst, K. (1996). Effects of fuel reduction Bee-eater, Speckled Warbler, White-browed burning on fauna in a dry sclerophyll forest. In Babbler, White-plumed Honeyeater, Little DEST (Ed.), Fire and biodiversity: the effects Friarbird, Yellow-faced Honeyeater, and effectiveness of fire management. Chestnut-breasted Mannikin, White-browed Proceedings of the conference held 8-9 October Wood-swallow, Superb Fairy-wren and 1994, Footscray, Melbourne (pp. 113-121 Mistletoebird. Species more common in the burnt (&129)). Canberra: Department of the areas included Emu, White-winged Triller and Environment, Sports and Territories. Dusky Wood-swallow. At 8 years 8 months control burning; eucalypt open forest post-fire, the abundance and richness of birds was Describes experiment with 15ha treatments and greater in burnt area (7.3 birds per observer-hour) replicates, with range of controlled fire regimes. than unburnt (3.8 birds per observer-hour). Species Only results following 1 year of treatment are more common in burnt area included: Galah, presented. “Bird abundance and species Cockatiel, Turquoise Parrot, Rainbow Bee-eater, composition remained remarkably stable on burnt Black-faced Cuckoo-shrike, White-winged Triller, and unburnt areas. The main changes observed Superb Fairy-wren, Western Gerygone, Rufous after burning were influxes of some species Songlark, Grey Fantail, Brown Treecreeper, Varied inhabiting understorey, and influx of some species Sittella, Grey-fronted Honeyeater, White-naped (e.g. scarlet robin, flame robin) which fed on Honeyeater, Noisy Friarbird and Dusky eucalypt nectar. The latter is of some interest, as Wood-swallow. Golden Whistler and Speckled such influxes have not been reported in previous Warbler were more abundant in the unburnt forest. studies of prescribed burning”. Notes the caveat These differences were considered to be due to that the study area was small, and no part of any release of nutrients after fire, which allowed burnt site was more than 300m from unburnt vigorous regrowth of wattles, other shrubs and forest. Also notes that the pre-fire understorey was eucalypts and which probably supported greater relatively open - hence the responses may be more abundance of insects. pronounced in forests with denser understorey.

Turner, R. J. (1992). Effect of wildfire on birds Trickett, T. (1983). Geelong garden birds on at Weddin Mountain, New South Wales. the days of dust-storm and bushfire. Geelong Corella, 16, 65-74. Naturalist, 20, 28-29. eucalypt woodland; Callitris woodland; wildfire; wildfire community; succession Bioregion: South east Coastal Plain Bioregion: NSW South western Slopes Lorikeets and some other birds were behaving Birds were surveyed 8 months and 8.7 years after strangely in a garden on the day of extreme winds, wildfire, including comparison with adjacent smoke and nearby fire. unburnt areas. At 8 months post-fire there were fewer species and individuals in burnt (50 spp, 225 individuals) than unburnt (59 spp, 400 individuals) Turner, R. J. (1987). Effect of fire on birds - areas. At 8.7 years post-fire, there were more Weddin Mountain. In Disappearing islands. species and individuals in the burnt area (87 spp, Proceedings of a seminar on conservation and 660 individuals) than unburnt (63 spp, 340 co-operation in the Central West (pp. 66-86). individuals) area. Ground-foraging, Bathurst: National Parks and Wildlife Service, foliage-gleaning, aerial, nectarivorous, NSW. granivorous, trunk-gleaning and frugivorous birds Callitris woodland; eucalypt woodland; wildfire; increased significantly from 8 months to 8.7 years community; succession post-fire. Recruitment may have been limited Bioregion: NSW South western Slopes because the site was largely surrounded by cleared Burnt and unburnt vegetation was surveyed 8 farmlands. The intense fire at this site led to months, and 8 years 8 months, after an intense increase in the dominance of eucalypts at the wildfire. At 8 months post-fire, the abundance and expense of Callitris. By 8 years post-fire, Rufous richness of bird species was significantly less in Songlark, parrots, Black-faced Cuckoo-shrike, burnt than in unburnt vegetation. Species which Superb Fairy-wren, Grey Fantail, Varied Sittella were more abundant in unburnt vegetation and were more common in the burnt included: Crested Pigeon, cuckoos, Rainbow area than unburnt sites.

172 Fire and Australian birds

Venn, D. R., & Fisher, J. (1993). Red-tailed Wardell-Johnson, G., & Christensen, P. Black-Cockatoo Calyptorhynchus banksii (1992). A review of the effects of disturbance graptogyne. Department of Conservation and on wildlife of the karri forest. In Research on Natural Resources. the impact of forest management in south-west Red-tailed Black-Cockatoo; eucalypt woodland; Western Australia (pp. 33-57). Perth: wildfire; management; threatened species; hollows. Department of Conservation and Land Bioregions: Naracoorte Coastal Plain; Management (Occasional Paper 2/92). Murray-Darling Depression. eucalypt open forest; management; review; control Reviews information on this species. Wildfire can burning have a devestating effect on this species, leading to Bioregion: Warren death of River Red Gum and Yellow Gum, and Historic changes in the bird community in hence removing food sources and nesting trees. south-western Australia were mainly from heaths Impacts of fuel reduction burning are not known, as a result of frequent burning. Reviews studies on but hot fires may reduce the availability of fruits the effects upon birds of fuel-reduction burning in (of Brown Stringybark) for up to 3 years. These Karri forests. Generally, small decrease in factors should be further researched. Firewood abundance of understorey birds for a few months collection is likely to be detrimental to this species. after burning, followed by increase to above pre-burn levels within 1-2 years post-fire. Many individuals remain in the same locality after fire. Verbeek, N. A. M., Braithwaite, R. W., & Boasson, R. (1993). The importance of Pandanus spiralis to birds. Emu, 93, 53-58. Wardell-Johnson, G., & Nichols, O. (1991). pandanus; community Forest wildlife and habitat management in Bioregion: Top End Coastal southwestern Australia: knowledge, research Disturbance due to increasingly intense fires may and direction. In D. Lunney (Ed.), have reduced pandanus thickets and their Conservation of Australia’s forest fauna (pp. associated birds, though little evidence is available. 161-192). Mosman: Royal Zoological Society of NSW. eucalypt open forest; review; control burning; wildfire; Walker, T. A., & Hulsman, K. (1993). Seabird management Islands: No. 221. Wilson Island, Great Barrier Bioregions: Jarrah Forest; Swan Coastal Plain; Reef, Queensland. Corella, 17, 155-157. Warren seabirds; pandanus; tussock grassland Reviews studies dealing with fire, bird Bioregion: South Eastern Queensland communities and their relationship in forests of Previous dominance by Pandanus has been much southwestern Australia. Describes an ongoing reduced by burning. This reduction has led to study of birds in 67 plots, of which nearly half decline of Wedge-tailed Shearwaters. The island have been burnt. was set on fire in the 1960s and 1970s, which would have caused mortality to breeding seabirds. Bridled Terns were probably most affected. Wardell-Johnson, G., McCaw, W. L., & Masey, K. G. (1989). Critical data requirements for the effective management of fire on nature Wall, C. A. (1989) Post-fire dynamics of the conservation lands in south Western Australia. vegetation, habitat and population of the Ground In N. Burrows, L. McCaw, & G. Friend (Eds.), Parrot at Barren Grounds Nature Reserve, New Fire management on nature conservation lands South Wales. B.Sc. (Hons.), University of New Perth: Department of Conservation and Land South Wales. Management. Ground Parrot; heath; threatened species eucalypt open forest; heath; review; management; Bioregion: South East Corner threatened species; control burning Bioregions: Jarrah Forest; Warren; Swan Coastal Plain Fire management plans need to be associated with monitoring of threatened and key species.

173 Australia’s Biodiveristy - Responses to Fire

Watkins, D. (1985). Report of the R.A.O.U. planned burning within the breeding range should Ground Parrot survey in Western Australia. occur between May and July, and that more Royal Australasian Ornithologists Union. research be directed at the effects of fire. Ground Parrot; heath; threatened species Bioregions: Esperance Plains; Jarrah Forest; Warren Wegener, A. (1984a). Animals killed by Ground Parrots were recorded only in diverse bushfires in coastal Victoria, February 1983. shrub-heathlands unburnt for at least 16 years. Australian Bird Watcher, 10, 248-250. mortality; wildfire Bioregions: South east Coastal Plain; South Watkins, D., & Burbidge, A. H. (1992). Eastern Highlands Conservation of the Ground Parrot in Dead birds found on the beach following the Ash Western Australia. In L. Joseph (Ed.), Issues in Wednesday bushfires were counted. 55 species the Conservation of Parrots in Australasia and were recorded, the most common being Crimson Oceania: Callenges to Conservation Biology. Rosella, Red Wattlebird and New Holland Proceedings of the RAOU/WWF Scientific Day Honeyeater. Many were badly burnt, but others and Workshop, Sydney, 22-23 September 1990. showed no signs of burns. The northerly winds (pp. 46-49). Melbourne: RAOU. blew the fire towards the coastline and many birds Ground Parrot; threatened species; management; had tried to escape by flying out to sea. heath; wildfire Bioregion: Esperance Plains In WA, the Ground Parrot has been recorded from Wegener, A. (1984b). Survey of animal species heaths of 15-35 years after fire, though individuals killed by the Ash Wednesday fires along the may sometimes forage in the edge of adjacent areas Great Ocean Road. Geelong Naturalist, 21, that have been burnt more recently. These 13-16. preferred ages are appreciably greater than those mortality; eucalypt open forest reported for mainland eastern Australia. Wildfire is Bioregion: South east Coastal Plain the major threat to this subspecies. Recent Count of 55 bird species washed up along the extensive wildfires have probably greatly reduced beach immediately after hot wildfire. The most populations, including in national parks. common dead birds were Crimson Rosella, New Holland Honeyeater and Red Wattlebird. Many were badly burnt, but some appeared undamaged Webster, R., & Ahern, L. (1992). Management and were presumably asphyxiated. for conservation of the Superb Parrot (Polytelis swainsonii) in New South Wales and Victoria. Department of Conservation and Natural Whelan, R. J. (1995). The ecology of fire. Resources. Cambridge: Cambridge University Press. Superb Parrot; threatened species; eucalypt woodland; review control burning; management Reviews many aspects of fire ecology. For birds, Bioregion: Riverina recommends more long-term studies with banded For the Superb parrot, the impacts of control birds and research into possible ecological burning are unclear, but may include reduction in disadvantages of control burning. food availability during the breeding or post-breeding period (and hence reduced breeding White, G. (1979a). Seabird islands: No. 75. success), change in vegetation structure or Louisa Island, Tasmania. Corella, 3, 61-62. composition (possibly including invasion by exotic seabirds; mortality; tussock grassland plants), and damage or destruction of nest sites. Bioregion: West and South West However, control burning may provide a tool for “A very severe fire was lit on the island in recent increasing food resources or other habitat years and burnt for several weeks extending over characteristics. Recommends that nest sites be most parts and reportedly killing countless given priority protection during fire suppression, thousands of breeding shearwaters.” that control burns not occur around nest sites (unless slashing is done first), that extensive control burns not occur in major feeding sites, that

174 Fire and Australian birds

White, G. (1979b). Seabird Islands: No. 76. Williams, J. E., & Gill, A. M. (1995). The , Tasmania. Corella, 3, 63-65. impact of fire regimes on native forests in eastern seabirds; eucalypt open forest; heath; tussock grassland; New South Wales. NSW National Parks and mortality Wildlife Service. Bioregion: West and South West review; eucalypt open forest; management; control Fishermen frequently set fire to the vegetation burning; wildfire; forestry; historic change; Aboriginal causing heavy casualties among the penguin burning colonies. Three fires in 1975 and 1976 reduced Bioregions: South East Corner; South Eastern the population of breeding penguins by almost Highlands; Sydney Basin; NSW North Coast; one thousand, with even greater toll of chicks. New England Tableland Broad review of ecological impacts of fire in eastern NSW. For birds, responses “depend to White, G. (1979c). Seabird Islands: No. 77. some extent on fire intensity … Unburnt patches Flat Witch Island, Tasmania. Corella, 3, 66-67. provide important resources … Species occupying seabirds; mortality; heath; tussock grassland the lower vegetation strata are … likely to be most Bioregion: West and South West affected by the changes in vegetation structure Fires are occasionally lit during the seabird caused by fire … Following fire of low to nesting season and undoubtedly kill many moderate intensity, a rapid recovery in both seabirds. numbers and species composition is generally observed, whereas the recovery process after more intense fires is usually slower.” White, G. (1979d). Seabird Islands: No. 79. Flat Island, Tasmania. Corella, 3, 70-72. seabirds; tussock grassland; mortality Wilson, B. A. (1996). Fire effects on vertebrate Bioregion: West and South West fauna and implications for fire management Fires are uncommonly lit during the seabird and conservation. In DEST (Ed.), Fire and breeding season, leading to some losses of biodiversity: the effects and effectiveness of fire breeding birds. management. Proceedings of the conference held 8-9 October 1994, Footscray, Melbourne (pp. 131-147). Canberra: Department of the Whitley, G. P. (1944). Fire and Petrels: the Environment, Sports and Territories. mystery of Mondrain Island. Emu, 44, 6-7. review seabirds; tussock grassland; heath; mortality Reviews studies on responses of birds (and other Bioregion: Esperance Plains vertebrates) to fire. Recommends increased Many breeding shearwaters were killed on synthesis and modelling (across a range of scales); Mondrain Island in an extensive fire. longer-term studies; effects of fire interacting with other factors (e.g. predation); and improved Whitley, G. P. (1971). Field notes on birds by communication between researchers and Thomas Carter. Western Australian Naturalist, managers. Also, notes far more studies for plants 12, 41-44. than for animals. Western Whipbird; heath; threatened species; historic change. Wilson, R. I. (1981). The woodchip industry Bioregion: Warren and Tasmanian birds. Tasmanian Bird Report, Frequent firing of heaths probably led to the local 11, 11-14. extinctions of Western Whipbirds. forestry; eucalypt open forest; Spotted Quail-thrush; management; control burning Whittell, H. M. (1936). The Bristle-birds of Fire regimes may exacerbate or ameliorate Western Australia. Emu, 35, 197-201. forestry effects. For example, the Spotted Western Bristlebird; Rufous Bristlebird; heath; Quail-thrush may better tolerate logging if the wildfire; threatened species forest is allowed to regenerate without use of fire. Bioregion: Warren Reports from Whitlock that Western Bristlebird disappeared from an area after severe fire.

175 Australia’s Biodiveristy - Responses to Fire

Wilson, S. J. (1995). Survival of Brown and Woinarski, J. C. Z. (1989b). The vertebrate Striated Thornbills in the , fauna of broombush Melaleuca uncinata Australian Capital Territory. Corella, 19, vegetation in north-western Victoria, with 138-146. reference to effects of broombush harvesting. Brown Thornbill; Striated Thornbill; eucalypt open Australian Wildlife Research, 16, 217-238. forest; wildfire. mallee; heath; community; succession; management Bioregion: South Eastern Highlands Bioregion: Murray Darling Depression Wildfire burnt about half of a study site in the Bird assemblages were sampled in mallee middle of a 20-year study. Banding and survival broombush stands 4,8,26,40 and 60-80 years data are presented on a year-by-year basis, though post-fire, 3 years after a severe frost, and 1-2,3,4-5 the effects of fire are not analysed. and 6-10 years after vegetation harvesting. Invertebrates were least abundant in old vegetation. There was substantial change in bird Woinarski, J. C. Z. (1987). Notes on the status species composition from young to old vegetation. and ecology of the Red-lored Whistler Species characteristic of areas burnt recently (<10 Pachycephala rufogularis. Emu, 87, 224-231. years) included Chestnut Quail-thrush and Shy Red-lored Whistler; mallee; heath; threatened species Heath-wren. Species most common in Bioregion: Murray Darling Depression intermediate ages (10-30 years post-fire) included Red-lored Whistlers prefer mallee vegetation of Southern Scrub-robin and Red-lored Whistler. 5-30 years post-fire, probably because they Species most common in older vegetation require dense low cover under relatively open included Malleefowl and Gilbert’s Whistler. mallee canopies. Consequently, long intervals of There were some differences in bird species fire exclusion may lead to the loss of this species. composition between sites regrowing for comparable periods after fire, frost and cutting (with burnt sites having fewest species and Woinarski, J. C. Z. (1989a). Broombush individuals, possibly because more trees were harvesting in southeastern Australia. In J. C. retained with frost and cutting disturbances). Noble & R. A. Bradstock (Eds.), Mediterranean landscapes in Australia: mallee ecosystems and their management (pp. 362-378). Woinarski, J. C. Z. (1990). Effects of fire on Melbourne: CSIRO. the bird communities of tropical woodlands mallee; heath; management; community; succession and open forests in northern Australia. Bioregion: Murray Darling Depression Australian Journal of Ecology, 15, 1-22. Although few bird species were restricted to tropical eucalypt open forest; tropical eucalypt savanna particular age classes of vegetation, for 18 of 26 woodland; community; succession commonly-recorded species, there was significant Bioregions: Top End Coast; Daly Basin. variation in density between sites with different Two studies are reported: a short-term response to post-disturbance age. Shy Heath-wren, fires at 3 times during the dry season, and Tawny-crowned Honeyeater and Chestnut longer-term responses to four fire regimes Quail-thrush favoured young (<10 years) maintained over 14 years. Of 11 common species, regrowth; Yellow Thornbill, White-browed 6 showed significant differences between Babbler and Malleefowl favoured old (>30 years) long-term fire treatments (fire exclusion, annual vegetation. early dry season burn, annual late dry season burn, biennial early burn): White-throated Honeyeater, Weebill and Lemon-bellied Flycatcher were all significantly more common in unburnt plots and then in plots burnt biennially, and Pied Butcherbird, Blue-winged Kookaburra and Magpie-lark were absent in unburnt plots. In general, granivores were most common in the early burnt plots, carnivores were least common in unburnt plots, and foliage-gleaners, branch-gleaners and nectarivores were most common in unburnt plots. These differences were

176 Fire and Australian birds

related to variation in extent and composition of Woinarski, J. C. Z. (1993a). Australian tropical understorey. With short-term responses, early dry savannas, their avifauna, conservation status season burns were relatively cool and localised, and threats. In C. P. Catterall, P. V. Driscoll, and a wide range of bird species (Gouldian Finch, K. Hulsman, D. Muir, & A. Taplin (Eds.), Long-tailed Finch, Masked Finch, Galah, Birds and their habitats: status and conservation Cockatiel, Magpie-lark, Grey-crowned Babbler, in Queensland (pp. 45-63). St Lucia: Pied Butcherbird, Black-faced Wood-swallow and Queensland Ornithological Socity Inc. Black-tailed Treecreeper) were more common in tropical eucalypt savanna woodland; topical eucalypt burnt areas from 3 days to 4 months after fire. open forest; tussock grassland; hummock grassland; This attraction was less pronounced beyond 4 review; management months post-fire, possibly because of greater Fire regimes in monsoonal savannas have changed extent then of burnt areas outside the study site. markedly since European colonisation. Limited Several other species (Tree Martin, research in savanna woodlands suggests that early White-breasted Wood-swallow, Black Kite, dry season fires attract many birds, those later in Torresian Crow and Brown Falcon) moved into the dry season are more detrimental. Fire the study site when fires were burning. Only one exclusion may benefit species which feed or nest in species (Red-backed Fairy-wren) showed shrubby understorey. No research has been short-term decline in burnt areas. There was little undertaken on fire and birds in floodplains. Fire is short-term response to the late fire, either because usually excluded from Mitchell grasslands, which it destroyed more of the seed resources, or may be disadvantageous for granivorous birds. because of the greater regional extent of burnt Recommends more research on fire and birds, areas. Many species in this region probably track particularly in Mitchell and floodplain grasslands. fires to take advantage of the consequent rich food resources or greater access to food (through clearing of the dense grass layer). In comparison Woinarski, J. C. Z. (1993b). A cut-and-paste to temperate forests, there appears to be a greater community: birds of monsoon rainforests in positive response of birds to cool burns, and a less Kakadu National Park, Northern Territory. obvious long-term successional response. The Emu, 93, 100-120. latter may be because the prevalence of fire rainforest; community prevents any substantial vegetation succession in Bioregion: Pine Creek Arnhem tropical eucalypt open forests. Relative to larger patches, small rainforest patches tend to be more disturbed by fire (and other factors), and have fewer obligate rainforest bird Woinarski, J. C. Z. (1992). The conservation species. The abundance of several open forest or status of the White-throated Grasswren rainforest margin species (e.g. Pied Butcherbird, Amytornis woodwardi, an example of problems Bar-shouldered Dove) within monsoon rainforests in status designation. Northern Territory was positively correlated with recent extent of fire Naturalist, 13, 1-5. in rainforest patches. White-throated Grasswren; hummock grassland Bioregion: Pine Creek Arnhem White-throated Grasswren densities were less in Woinarski, J. C. Z., & Fisher, A. (1995a). recently burnt (<1yr) sites, and some populations Wildlife of lancewood (Acacia shirleyi) thickets may have been eliminated by repeated hot fires. and woodlands in northern Australia: 1. The fire regime of the Arnhem Land massif is variation in vertebrate species composition undergoing substantial change, probably to the across the environmental range occupied by detriment of this species, and its existence may be lancewood vegetation in the Northern threatened by the current regime. Territory. Wildlife Research, 22, 379-411. Acacia woodlands; community; wildfire Bioregions: Sturt Plateau; Gulf Fall and Uplands; Ord-Victoria Plains Many lancewood patches showed impact of fire, and wildfire is the major threat to lancewood. The abundance of 11 bird species (including Peaceful Dove, Singing Honeyeater, Double-barred Finch

177 Australia’s Biodiveristy - Responses to Fire

and Pied Butcherbird) was significantly negatively Woinarski, J. C. Z., Eckert, H. J., & associated with fire impact, whereas only 2 species Menkhorst, P. W. (1988a). A review of the showed positive correlations. distribution, habitat and conservation status of the Western Whipbird Psophodes nigrogularis leucogaster in the Murray mallee. South Woinarski, J. C. Z., & Fisher, A. (1995b). Australian Ornithologist, 30, 146-153. Wildlife of lancewood (Acacia shirleyi) thickets Western Whipbird; mallee; heath; threatened species; and woodlands in northern Australia: 2. management comparisons with other environments of the Bioregion: Murray Darling Depression region (Acacia woodlands, Eucalyptus savanna Previous studies have differed in characterising woodlands and monsoon rainforests). Wildlife the preferred age of vegetation used by the Research, 22, 413-443. Western Whipbird. This study also shows no Acacia woodland; rainforest; tropical eucalypt savanna consistent pattern, with records from 2 years woodland; historic changes post-fire to at least 40 years post-fire. However, The distributions of bird species offers some intermediate (10-25 years) ages seem most support to a formerly extensive continuum from suitable, but this may change with different monsoon ranforests to Acacia thickets. Climate floristic associations, or with localised change and/or changes in fire regime may have idiosyncracies (such as insect outbreaks). Other fragmented this vegetation, with consequent terrestrial bird species showing association with expansion of Eucalyptus savanna woodlands. seral stages in mallee include Shy Heath-wren and Chestnut Quail-thrush (<10 years post-fire), Southern Scrub-robin (10-30 years), Malleefowl Woinarski, J. C. Z., & Tidemann, S. C. (1991). and White-browed Babbler (>30 years). Habitat The bird fauna of a deciduous woodland in the fragmentation may exacerbate fire effects, wet-dry tropics of northern Australia. Wildlife reducing chances of refuge and recolonisation. Research, 18, 479-500. Small-area mosaic burning to maintain a diversity tropical eucalypt savanna woodland; community of fire ages is recommended. Bioregion: Daly Basin Fire, rain, flowering and seeding produce a dynamic patchwork of environments, to which Woinarski, J. C. Z., Tidemann, S. C., & Kerin, bird species respond by mobility and shifting S. (1988b). Birds in a tropical mosaic: the composition at any site. distribution of bird species in relation to vegetation patterns. Australian Wildlife Research, 15, 171-196. Woinarski, J. C. Z., & Tidemann, S. C. (1992). tropical eucalypt open forest; community Survivorship and some population parameters Bioregion: Top End Coastal for the endangered Gouldian Finch Erythrura Tropical eucalypt open forests with a dense gouldiae and two other finch species at two shrubby understorey have far richer bird sites in tropical northern Australia. Emu, 92, assemblages than those without. This structural 33-38. complexity is probably related to fire regime, Gouldian Finch; tropical eucalypt savanna woodland; though may also be influenced by moisture threatened species availability. Bioregions: Daly Basin; Victoria-Bonaparte Extensive nomadism in Gouldian Finch may be a response to environmental heterogeneity caused Woinarski, J. C. Z., Whitehead, P. J., Bowman, by fire, rainstorms, floristic and substrate D. M. J. S., & Russell-Smith, J. (1992). patterning. Conservation of mobile species in a variable environment: the problem of reserve design in the Northern Territory, Australia. Global Ecology and Biogeography Letters, 2, 1-10. Gouldian Finch; threatened species; management; tropical eucalypt savanna woodland; tussock grassland The endangered Gouldian Finch (and other more common species) may exploit landscape patchiness

178 Fire and Australian birds

by following episodic fires or occasional Dry Wouters, M. (1996). Developing fire season rainstorms. Changes in distribution related management planning and monitoring. In to fire constrain reservation options for this DEST (Ed.), Fire and biodiversity: the effects species and conservation planning generally. and effectiveness of fire management. Proceedings of the conference held 8-9 October 1994, Footscray, Melbourne (pp. 235-239). Wooller, R. D., & Brooker, K. S. (1980). The Canberra: Department of the Environment, effects of controlled burning on some birds of Sports and Territories. the understorey in Karri forest. Emu, 80, control burning; heath; Ground Parrot; management; 165-166. threatened species eucalypt open forest; control burning; community Bioregion: South East Corner Bioregion: Warren Outlines a fire management plan for far East Birds were mist-netted and banded in Karri forest Gippsland, which sets out the fire management in the year before and after a fuel-reduction burn. required to maintain the ecology of coastal The species and number of birds caught were very heathland, in particular Ground Parrot habitat, similar before and after fire. Some species showed and how this can be accommodated within high quantitative changes. Fairy-wrens were less protection priority zones (where ecological values common, possibly because of changed food may need to be compromised). Also notes that resources or less cover. Black-faced Cuckoo-shrike some mallee communities should have a fire and Rufous Treecreeper increased, possibly exclusion management for conservation of because of more open understorey. Six individuals Malleefowl, but such straightforward cases are (of 5 spp) were retrapped at the same location rare - in most cases, management generally has to after fire. be conservative, adopt a range of fire management regimes and monitor consequences. Wooller, R. D., & Calver, M. C. (1988). Changes in an assemblage of small birds in the Young, J. (1991). Dealing with fire in parks understorey of dry sclerophyll forest in and protected areas. In Tropics under fire: fire south-western Australia after fire. Australian management on Cape York Peninsula (pp. Wildlife Research, 15, 331-338. 43-45). Cairns: Cairns and Far North eucalypt open forest; control burning Environment Centre. Bioregion: Jarrah Forest hollows; tropical eucalypt open forest; rainforest; Mist-netting and banding of understorey birds management was conducted 2 months before a cool burn, then Bioregion: Cape York Peninsula 2,7,10,22 and 34 months post-fire in a Jarrah Hot fires (including control burns) can destroy the forest. Although the fire was mild, limited in hollows important for nesting birds (especially extent and left much surrounding vegetation owls and parrots), and have been reported unburnt, the total number of birds caught after destroying nests and young. the fire was about half that of before fire. 22% of marked birds were recaptured after fire. After the fire, birds ate relatively more ants but fewer beetles, than before fire. Prey size was smaller after the fire than before. Dietary diversity of insectivorous birds declined post-fire. Abundance of all prey types (including ants) decreased after fire. One common species (Splendid Fairy-wren) pre-fire disappeared for 10 months post-fire. One species absent before fire (Scarlet Robin) became common within 2 months of the fire. Otherwise most changes were of relative abundance rather than change in species presence.

179 Australia’s Biodiveristy - Responses to Fire

LOCATION (AND NUMBER) OF PRIMARY STUDIES OF EFFECTS OF FIRE ON BIRDS.

Numbers are based on a 1° x 1° grid. Studies extending over more than 5 cells were not counted. In some cases where study areas were not clearly defined, there may be some minor imprecision in my attempt to locate them.

180 Forest Research and Development Division State Forests of NSW

LONG-TERM EFFECTS OF REPEATED PRESCRIBED BURNING ON FOREST INVERTEBRATES: Management Implications for the Conservation of Biodiversity

Consultancy Report to the Department of the Environment and Heritage

Alan York

Final Report April 1996 ACKNOWLEDGEMENTS

Full credit must be given to Adrian van Loon Urquart who assisted in the laboratory work and who, in the 1950s, conceived and established a with reference material, and to Traecey Brassil field trial which was sufficiently well designed and and Alison Towerton for their assistance with the robust to not only be addressing questions still data analyses and presentation. Taxonomic relevant 40 years later, but also sufficiently flexible support was provided by specialists Mike Gray, to now facilitate questions probably not even Dan Bickel, John Lawrence, Robert Taylor and conceived at that time. Similarly, much is owed to Gerry Cassis who also provided enthusiastic the Forestry Commission of New South Wales guidance and advice. Thanks also to Andy Beattie (State Forests) and its staff for the long-term and Ian Oliver from the Key Centre for maintenance of the trial, in particular Bill Buckler Biodiversity & Bioresources at Macquarie who methodically and reliably undertook routine University for their ongoing support and counsel, measurements for over 20 years, and to Hugh particularly in regard to the development and use Dowden, Bob Bridges and others who maintained of methodologies for Rapid Biodiversity and verified the enormous database. Assessment. For my part in this project, which began in I also wish to sincerely thank Neal Hardy for 1991, I wish to thank Bill Buckler, Bill Chapman, his patience and perseverance, and the Patrick Murphy and Traecey Brassil for their Department of the Environment and Heritage for assistance with fieldwork, and Rita Holland, Ruth their financial support. Burton, Graeme Price and Darren Waterson for the tremendous effort and dedication involved in sorting in excess of 55,000 specimens. My appreciation also to Debbie Kent and Chris-Ann

182 EXECUTIVE SUMMARY

BACKGROUND THIS REPORT

Infrequent, periodic forest fires (bushfires) are an Little is known about the effects of repeated integral part of the modern physical environment of hazard-reduction burning over long time scales. Australian sclerophyll forests. Low-intensity fires The research reported here was conducted as part are extensively used in managed sclerophyll forests of the Bulls Ground Frequent Burning Study to stimulate regeneration, manipulate wildlife (Experiment F8/2.9), located on the mid-north habitat and in particular, to reduce fuel levels with coast of New South Wales and established in the intention of minimising the extent and severity 1969. The terrestrial invertebrate component of of wildfires. In Australia, the use of deliberate fire to the project commenced in 1991 following 20 years prevent high-intensity wildfires has become of repeated low-intensity fire, and was undertaken probably the most extensive use of fire in land with the following aims: management. The inherent variability in natural fire • to identify the effects of long-term repeated regimes generally results in a mosaic of habitats burning on terrestrial invertebrate with vegetation at different stages of floristic and biodiversity, structural post-fire succession, each potentially • to identify species and/or species groups supporting particular animal communities. Changes most affected by this management practice to the components of the fire regime (fire intensity, and to devise strategies to ensure their frequency and season of occurrence), as a continued conservation, consequence of forest management practices, have • to identify species which, due to the nature the potential to alter the composition and structure of their response, may be useful “indicators” of natural communities. The research reported here of environmental disturbance and deals with the impact of frequent low-intensity fire degradation. (“hazard-” or “fuel-reduction burning”) on the The primary outcome was therefore to assess abundance, richness, composition and structure of whether frequent hazard reduction burning is a terrestrial invertebrate communities. sustainable long-term management practice with regard to the conservation of our forest WHY INVERTEBRATES? biodiversity. Invertebrates (insects, spiders, mites, worms, snails, FREQUENT FIRE & centipedes etc.) are the most diverse and abundant HABITAT STRUCTURE animals in most natural systems, but their importance in sustaining those systems is If frequent fire reduces the diversity of post-fire commonly not appreciated. This multitude of environments, then it has the potential to impact organisms constitutes the bulk of the biodiversity upon animal communities dependent upon this within forests and plays an essential role in primary habitat mosaic. This research indicated that production, nutrient cycling and uptake, population frequent burning resulted in a simplification of & community level interactions and energy storage large-scale spatial patterning in the litter (fine-fuel) & transfer. Through their contribution to environment. The components (, twigs, bark ecosystem function, these organisms also enable etc) that give the leaf litter its physical structure forest ecosystems to provide benefits to humanity. changed with regard to their relative abundance These benefits include amenity values in the form and spatial distribution. There were marked of aesthetics, recreation and education; heritage changes to the amount (cover) of vegetation in the values as forests contribute to long-term security understorey and its spatial patterning. While the for catchment protection, air and water quality and quantity of vegetation closest to the ground nature conservation; and economic values including (ground herbs & small shrubs) was not affected by timber production, grazing and ecotourism. The frequent burning, there was a decrease in the maintenance of biodiversity is a fundamental spatial heterogeneity (patchiness) of these layers. principle underlying the ecologically sustainable Conversely, the cover of tall and very tall shrubs management (ESM) of these environments.

183 Australia’s Biodiveristy - Responses to Fire

was substantially reduced and showed an increase with broad tolerances, or adaptations, to drier and in spatial heterogeneity. Top-soil moisture levels more open environments. were, on average, 18% lower following 20 years of These shifts in community composition were frequent burning, whereas the amount of light substantial and suggested that the extensive and reaching ground level had increased (on average) frequent application of fuel-reduction burning by 125% and become more spatially homogeneous could result in a reduction in terrestrial (less patchy). A number of habitat components (eg. invertebrate biodiversity at a regional scale, with top-soil hardness, the distribution of large sticks & this decrease potentially as high as 50%. Current logs) however showed no significant response to fuel management strategies which limit the extent frequent burning. of frequent burning will ameliorate these impacts, however there remains a need to establish secure TERRESTRIAL INVERTEBRATE refuges for species with specialist requirements and COMMUNITIES limited dispersal abilities, and provide links (ie corridors) between habitat patches to facilitate This study revealed a rich terrestrial invertebrate recolonisation. The effectiveness of similar fauna with representatives from the Chelicerata strategies developed to conserve vascular plants (spiders, ticks & mites, pseudoscorpions, and vertebrates remains untested however for the harvestmen), Crustacea (landhoppers, slaters), groups which actually constitute the bulk of our Chilopoda (centipedes), Diplopoda (millipedes), forest biodiversity. Realistically, the conservation and a diverse array of Insect Orders and Families. of biodiversity cannot be achieved without Numerically, the most abundant groups overall consideration of the important role that were the springtails (33%), ticks & mites (24%) invertebrates play, both through their involvement and ants (23%), with these three groups in ecological processes, and as a significant representing 80% of all individuals caught. For 10 component of the overall richness of biotic broad taxonomic groups there were sufficient data communities. to statistically test the effects of frequent burning. The results indicated a variety of responses with Community Structure & statistically significant decreases in abundance for Ecosystem Function ↓ ↓ ticks & mites ( 31%), insect larvae ( 35%), The biological structure of a community involves ↓ ↓ ( 58%) and beetles ( 31%). Many of these groups species composition (diversity and relative are associated with leaf litter and it is likely that abundance) and the relationships between species - their numbers have been influenced by the their ecological role. It was demonstrated here that episodic removal of this resource. Three groups considerable additional detail concerning, and showed substantial increases in abundance insight into, the nature of invertebrate community ↑ following frequent burning; bugs ( 77%), ants changes could be provided by the inclusion of fairly ↑ ↑ ( 250%) and spiders ( 33%), probably as a general information concerning habitat and dietary response to both changes in habitat suitability and preferences. It was apparent that frequent burning increased ease of capture in a simplified leads to a change in the structure of the environment. invertebrate community. Within species Biodiversity assemblages there were shifts based on feeding strategy and habitat preference. While the impact Using ants, beetles, flies, spiders & bugs as of these changes on ecosystem function was beyond representative groups and potential indicators of the scope of this study, substantial measured environmental degradation, this research changes in the structure of invertebrate demonstrated that although overall species assemblages and the loss of species associated with α richness at specific sites ( -diversity) did not the decomposer cycle implies frequent burning change with frequent burning, all groups showed may be impacting upon nutrient cycling and substantial changes in the composition of species transfer within these forests. If this is the case, it assemblages. There was a loss of taxa dependent would have serious implications with regard to the upon a substantial litter layer and stable moist maintenance of ecological sustainability. conditions, and these species were frequently habitat or dietary specialists and often uncommon or “rare”. The overall diversity of frequently burnt areas was maintained by the addition of species

184 Bushfire and forest invertebrates

Biodiversity Indicators Note: Indices used to gauge the success of ecologically Following the preparation of this report, there sustainable management practices need to be have been some taxonomic revisions and interpretable, significant and cost efficient. They associated morphospecies corrections of the also need to account for variability in space and data. These have been independently published, time, and be appropriate for the scale of however they were of a minor nature and do not management. The research reported here identified alter the outcomes of the analyses or the the limited usefulness of data obtained using coarse- conculusion drawn in this report. scale taxonomic classification (eg. Family or Order), with the cost-effectiveness of abundance data alone shown to be low. This research also identified substantial limitations with regard to the use of a single index, species richness, as a measure of change and/or environmental impact. Species richness (α-diversity) is frequently used to describe and compare communities, however in this case it was found to provide a deceptive summary of community characteristics and severely restrict the level of interpretation that could be derived for impact assessment purposes. The application of Rapid Biodiversity Assessment (RBA) methodology here demonstrated that the study of the composition and structure of communities is likely to prove more rewarding in this regard. The identification of individuals to distinct “morphospecies” facilitated the incorporation of broad-level ecological information into the assessment, and interpretation, of environmental impact. This in turn enabled the development of management recommendations consistent with the conservation of biological diversity.

185

Bushfire and forest invertebrates

1. INTRODUCTION

The concept of Ecologically Sustainable there is growing concern that repeated low- Development (ESD) was defined by the United intensity burning, as a management prescription, Nations in 1987 as “… development that meets the may have a negative influence on plant and animal needs of the present without compromising the ability of communities. Frequent firing may remove future generations to meet their own needs” vegetation species that rely on seed production for (“Bruntland Report” - WCED 1987). This their persistence (Gill 1981; Bradstock and concept has been developed and refined regularly Myerscough 1981; Benson 1985; Fox and Fox since that time, most recently at the “Earth 1986), often leading to dominance by herbaceous Summit”, the United Nations Conference on fire-tolerant species (Cary and Morrison 1995). Environment and Development held in Rio de Fire frequency becomes a significant factor for Janeiro in 1992. ESD forms part of the World plant species requiring a long period of time Conservation Strategy (IUCN 1980) and is the (relative to the interval between fires) to reach basis for the National Conservation Strategy for reproductive maturity (Zedler et al. 1983; Australia (Commonwealth of Australia 1983). Nieuwenhuis 1987). Changes in habitat structure The conservation of biological diversity is a as a consequence of frequent burning are likely to foundation of ESD and is one of the three core disadvantage many native mammal and bird objectives of the Australian National Strategy for species (Catling 1991; Whelan 1995). Ecologically Sustainable Development. Biological While sclerophyll forests, woodlands and diversity refers to the variety of all life forms - the heaths are dominated by plant species with different plants, animals and micro-organisms, the adaptive responses to fire that enable them to genes they contain and the ecosystems of which survive exposure to periodic burning (see for they form part. Australia has ratified the example Gill 1981; Noble and Slatyer 1981), the Convention on Biological Diversity arising from impact of such fires on terrestrial invertebrates is the Earth Summit, and is now developing strategies poorly understood. The consumption of some or to assess and protect its biodiversity. The all of the leaf litter by flame, short-lived but conservation of biological diversity is a major substantial rises in soil temperature during fire, objective of the National Forest Policy Statement and post-fire changes in the surface radiation (NFPS 1992), to be achieved through the budget, mean that soil and litter fauna are protection of ecosystems (reserve strategies) and substantially affected by fire in the short-term complementary off-reserve management (Bornemissza 1969; Springett 1979; Moulton (Ecologically Sustainable Management — ESM). 1982; Coy 1996). Recovery from a single fire may In New South Wales, State Forests has put forward take up to 3-5 years (Metz and Farrier 1973; ESM as a major objective in its 1992–5 Corporate Seastedt 1984; Neumann and Tolhurst 1991), Plan (Forestry Commission of NSW 1992). This however the timing and intensity of burning is concept has been widely adopted by other land important, as is the mobility and recolonising management agencies throughout Australia and ability of particular species (Morris 1975). Given forms part of the National Strategy for the the patchy nature of low-intensity fuel-reduction Conservation of Australia’s Biological Diversity. burns, and the protection afforded by small Low-intensity fires are extensively used in habitat refuges and within the soil, it has been managed sclerophyll forests to stimulate suggested that periodic fires used for fuel regeneration, manipulate wildlife habitat and in management purposes have few long-term effects particular, to reduce fuel levels with the intention on most soil and litter invertebrates (Majer 1980; of minimising the extent and severity of wildfires. Campbell and Tanton 1981; Abbott et al. 1984). In Australia, the use of deliberate fire to prevent There is little information on the effects of high-intensity wildfires has become probably the fire frequency on forest invertebrates, but Abbott most extensive use of fire in land management et al. (1984) suggest that periodic low intensity (Whelan 1995). While infrequent, periodic fires fires have few permanent effects on most of the (bushfires) are an integral part of the modern invertebrate taxa present in the litter and soil of environment of Australian sclerophyll forests, the Jarrah forest. Long-term studies of spiders

187 Australia’s Biodiveristy - Responses to Fire

(Huhta 1971; Merrett 1976) and ants (York 1994, thereby making a substantial contribution to our 1996) suggest that, in the years following fire, National biodiversity (New 1984; CONCOM there is a replacement series of groups of species 1989). Realistically, the conservation of related to their particular habitat requirements biodiversity cannot be achieved without being met as the habitat changes in structure over consideration of the important role that time. A number of species persist throughout this invertebrates play, both through their involvement period, but show changes in relative dominance in ecological processes, and as a significant within the community. York (1996) suggested that, component of the overall richness of biotic for ants, the use of regular widespread fires for communities. fuel reduction was likely to result in a truncation The research reported here was therefore of these successional patterns and an associated undertaken with the following aims: loss of regional biodiversity. • to identify the effects of long-term repeated Periodic low-intensity fire (hazard-reduction burning on terrestrial invertebrate burning) is a conspicuous management strategy in biodiversity, virtually all of Australia’s dry forest communities. • to identify species and/or species groups While it is primarily used to reduce fuel levels, most affected by this management practice little is known about the effects of its repeated use and to devise strategies to ensure their on natural ecosystems over long time-scales. On continued conservation, the east coast of NSW, extensive wildfires in • to identify species which, due to the nature January 1994 have led to calls for increased use of of their response, may be useful “indicators” hazard-reduction burning, however the impacts of of environmental disturbance and the resulting increase in fire frequency are poorly degradation. understood in the very forest environments this The primary outcome is therefore to assess management strategy seeks to protect. The whether frequent hazard reduction burning is a paucity of information available on the effects of sustainable long-term management practice with increased fire frequency on forest invertebrates is regard to the conservation of our forest of considerable concern. Invertebrates constitute biodiversity. 95% of known species of fauna in Australia,

188 Bushfire and forest invertebrates

2. METHODOLOGY

2.1 STUDY AREA whenever fuel build-up permitted, generally every 3 years (1970, 1973, 1977, 1980, 1983, 1986, As part of the F8 series of “fire effects” studies, 1989, 1992). This burning regime is ongoing. State Forests has an ongoing experimental project A program was instituted to monitor aspects which is particularly suitable for addressing of the response of this forest to repeated low- questions relating to repeated disturbance and intensity fire. A number of parameters were biodiversity conservation. The F8/2.9 Frequent regularly measured on each research plot: tree Burning Study is located in even-aged coastal growth, major and minor understorey vegetation, blackbutt Eucalyptus pilularis regeneration in litterfall, and fine and heavy fuel. These Compartment 70, Bulls Ground State Forest, measurements were made systematically between Kendall Management Area on the mid-north 1970 and 1987, when the project was reviewed; coast of New South Wales (31°33'S, 152°38'E, and then less frequently until 1992. 240m ASL.). The stand was logged and silviculturally treated in 1958–9, with seed trees 2.2 EXPERIMENTAL DESIGN retained singly and in groups, and unmerchantable trees culled in line with Timber From an inspection (by the author) of the area in Stand Improvement (T.S.I.) techniques. The area 1990 it was apparent that twenty years of repeated has experienced no further management burning had resulted in substantial changes in treatment (except experimental fuel-reduction macro- and micro-habitat parameters. It was burning) since that time. hypothesised that these changes would have had a In 1969 twenty-one 0.225 acre (0.1 ha.) significant effect on terrestrial invertebrate temporary plots were established in openings communities. In 1991, two years after the last fire, created by the logging treatment which carried a a project was initiated to assess the impact of good stocking of young blackbutt regrowth long-term fuel reduction burning on terrestrial (11 years old). These areas were found to support invertebrates, and to investigate the possibility of a mean number of 339 stems per hectare (Van using this faunal group as monitoring agents in Loon 1970), consisting mainly of blackbutt (48%) the assessment of ecologically sustainable and bloodwood E. gummifera (31%). The management. The overall approach was to view remainder (21%) consisted most commonly of this single sample period as a “snapshot” of the turpentine Syncarpia glomolifera, red mahogany effects of 20 years of prescribed burning by E. resinifera, white mahogany E. acmeniodes and comparing burnt and unburnt replicates. While grey gum E. punctata. Following an assessment of this does not enable a description of changes over stand parameters, a number of these plots were time, it does provide a unique opportunity to selected on the basis of their similarity for a long- assess the long-term impact of this management term fire study (F8/2.9), which was formally practice. initiated in March 1970. Twelve of the fourteen plots were selected as Fourteen 0.1 ha. permanent research plots suitable, six within each treatment (unburnt & were established, 7 randomly allocated as burning burnt). Plots 7A and 7B were excluded as they treatments (burnt), the remaining 7 as control contained rocky outcrops and were subjectively (unburnt) plots from which fire was excluded assessed to be different to other plots. (7×2 randomised block design). These study plots Randomised assignment of treatments to were located within similarly treated forest blocks experimental units ensured “true” replication of of approximately 1 ha. and separated by cleared treatment effects (see Hurlbert 1984). In order to buffer areas to protect them from wildfire (see increase the sensitivity of the experiment by Figure 2.1). For the remainder of this report the increasing the “precision” with which properties term “plot” refers to the 1 ha. treated forest areas, of each experimental unit (plot) and hence each while “research plot” refers to the 0.1ha study treatment were estimated, it was necessary to take plots defined in 1970 (see Figure 2.2). Fuel multiple samples from each plot. Four 20m reduction burning was implemented in Autumn transects were therefore established within each

189 Australia’s Biodiveristy - Responses to Fire

Figure 2.1 F8/2.Frequent Burning Study, Bulls Ground State Forest. Location of study plots. (Plots 7A and 7B not shown).

190 Bushfire and forest invertebrates

Figure 2.2 Schematic layout of study “plot” plot (a “nested” design), each on a randomly- Table 2.1 Slope and aspect of study plots oriented compass bearing starting from each corner Unburnt Burnt × of the established “research plot”. A 20m 10m sub- Plot Sub-plot Slope° Aspect° Slope° Aspect° plot was then centred on this transect (see Figure 1 1 1 290 1 220 2.2) in order to assess the small-scale variability of 2 1 300 1 220 measured parameters. 3 0 325 2 245 The general physical characteristics of each 4 1 330 3 210 sub-plot were summarised by measurements of 2 1 1 250 2 320 ground slope and aspect. The average slope of the 2 3 240 5 300 site in degrees below the horizontal was 3 3 260 2 340 determined with a hand-held clinometer, while the 4 0 360 3 310 aspect was determined by use of a compass. The 3 1 2 290 2 310 sub-plots had low slope angles (0–9) and 2 4 330 2 315 predominantly north-west to south-west aspects 3 1 290 1 290 (see Table 2.1). On average, burnt sites had slightly 4 0 280 4 310 steeper slopes, primarily plots 4, 5 & 6. The 4 1 1 210 2 250 differences in slope however were slight and reflect 2 3 240 9 285 the ridge-top nature of the study area. The range of 3 3 195 5 265 aspects was similar for both treatments. 4 1 200 5 270 5 1 1 225 2 275 2 2 280 4 270 3 0 270 2 255 4 2 275 6 270 6 1 0 360 7 225 2 1 230 7 200 3 1 210 2 200 4 1 220 3 220 Range 0–4 195–360 1–9 200–340 Mean±s.e. 1.4±0.2 269±10 3.4±0.4 26±69

191 Australia’s Biodiveristy - Responses to Fire

2.3 MEASUREMENT OF Table 2.2 Structual vegetation components ENVIRONMENTAL PARAMETERS Height class Structural component 0–20 cm Ground herbs The environmental framework within which 20–50 cm Small shrubs terrestrial invertebrate communities function 50–100 cm Mid-sized shrubs primarily involves elements of the vegetation 100–150 cm Tall shrubs understorey, the top-soil and litter components. A 150–200 cm Very tall shrubs number of parameters were quantitatively assessed to evaluate their possible influence on species richness and community structure. 2.3.2 The Litter Environment The distribution of data for most variables suggested that the sample mean was the best Ground-dwelling invertebrates have been shown estimate of average conditions at each sub-plot. to be sensitive to levels of forest litter, particularly Because individual samples were randomly drawn during post-fire recovery (Bornemissza 1969; from within replicates, a measure of variability Springett 1976; Seastedt 1984). Five randomly 2 about the mean also provided information about placed samples (0.1 m ) of litter (including sticks the spatial variability (“patchiness”) of the variables up to 2.5cm diameter) were collected from each concerned. The coefficient of variation (CV = sub-plot, sieved with 1mm soil sieves to dislodge standard deviation/mean × 100%) was selected as soil material, and then dried in an oven at 105° for the most appropriate measure here due to the 72 hours. Material was then sorted into 5 large fluctuations in mean values and the observed components and weighed: twigs 0–5mm & twigs dependence of the standard deviation on the mean. 6–25mm diameter, bark, leaves, and very fine fuel In order to satisfy the assumptions underlying (miscellaneous decomposing matter). This particular statistical procedures, variables were approach was consistent with that used to estimate appropriately transformed as required. the “fine fuel” fraction over the previous 20 years. 2.3.1 Understorey Vegetation Structure 2.3.3 Sticks and Logs The physical structure of the vegetation The incidence and diameter of all sticks & logs understorey in this forest environment consists (“heavy fuel”, >2.5cm) was recorded along 2 primarily of a shrub stratum and a herb stratum orthogonal 20m transects centred on each sub- whose heights and spatial distribution are a function plot. Following an examination of the frequency of fire history. A structural classification of the distribution of values, data were grouped for vegetation was chosen because it allowed a relatively subsequent analyses into the following five quick and consistent assessment of the sites (48 sub- diameter categories: 2.5–9.9, 10–24.9, 25–49.9, plots in total) to be made in an environment which 50–74.9, 75+cm. is floristically diverse (Doug Binns pers. comm.). 2.3.4 Insolation Vegetation structure is of direct significance in ecological studies of soil and soil-surface Levels of insolation have been shown to be critical invertebrates because the amount and distribution factors determining the abundance and of vegetation determines both the physical distribution of certain terrestrial invertebrates. framework within which activity takes place, and the The amount of light reaching the forest floor was food availability and hence carrying capacity of the used as an index of insolation levels, and measured environment (Greenslade and Thompson 1981). using a Lunasix 3 Gossen exposure meter, fitted Vegetation structure was quantitatively with incident light cone. Twenty measurements assessed using the “cover-board” technique (see were taken systematically within each sub-plot MacArthur & MacArthur 1961; Fox 1979). and expressed as a percentage of available light as Percentage cover was measured at 20 points measured outside the forest at that time of day. systematically located along each transect for five 2.3.5 The Soil Environment structural components of the understorey (see Table 2.2). The mean of the 20 measurements was The underlying geology of the site consists used as an estimate of percentage cover for each primarily of conglomerate, sandstones and shales. vegetation layer at each sub-plot, and the These have weathered to form shallow soils coefficient of variation (CV) as an estimate of (yellow earths & brown podzolics) which are spatial variability. relatively low in nutrients. Two aspects of the soil

192 Bushfire and forest invertebrates

physical environment were assessed: top-soil beetles (Coleoptera), spiders (Araneae), bugs moisture and top-soil hardness. (Hemiptera) and flies (Diptera) were subsequently 2.3.5.1 Top-soil Moisture sorted to “morphospecies” using the protocols Five samples (10cm diameter 3cm deep) per sub- described in Oliver & Beattie (1993), with final plot were collected and kept in sealed containers. taxonomic verifications being performed by Mike Samples for each sub-plot were pooled in the Gray (spiders), Gerry Cassis (bugs) and Dan laboratory, weighed and dried in an oven at 105° Bickel (flies) of the Australian Museum, John for 72 hours. An estimate of “field moisture Lawrence (beetles) and Robert Taylor (ants) of the content” for each sub-plot was calculated in the CSIRO. following manner (see Lambert 1982): Oliver and Beattie (1996a) have shown that morphospecies can provide a robust estimate of % Moisture Content = species richness across a variety of habitats. This study provides a substantial test of the hypothesis air-dried weight - oven dried weight × 100 air-dried weight that the lack of knowledge concerning so many Australian invertebrates, the so called “taxonomic impediment” (Taylor 1983), no longer prevents 2.3.5.2 Top-soil Hardness the inclusion of invertebrates in biodiversity An index of top-soil hardness (0–5cm) was assessment and studies of management impacts. obtained using a Geonor inspection vane, which measures soil shear strength. Twenty 2.5 ANALYTICAL PROCEDURES measurements were taken within each sub-plot, the mean value representing the average shear All initial analytical procedures were performed strength and the coefficient of variation (CV) an using the SPSS statistical package on a 486PC at indication of spatial variability. SFNSW’s Research Division. Data distributions were examined using exploratory data analysis 2.4 TERRESTRIAL INVERTEBRATE techniques (EXAMINE) and transformed (as COMMUNITIES required) for subsequent analyses (MANOVA and REGRESSION). Canonical Correspondence Epigaeic (surface active) invertebrate communities Analyses (CCA - Ter Braak 1986) were performed were assessed by a single summer pitfall trapping using programs written in Splus on a Sun program in February 1991. Nine points were Workstation. established and marked along the 20m transect within each sub-plot. At each point a 6.5cm 2.5.1 Treatment, Exposure diameter 9cm deep plastic cup was sunk flush with and Position Effects the ground surface and half-filled with a non- Plots had been allocated to one of two treatments: attractive preservative solution. Pitfall traps were burnt (1B-6B) or unburnt (1A-6A). Aspect values left open for a period of 7 days (5–12th February were coded from 1-6 to reflect the relative 1991), reducing the effect of temporal changes in “exposure” of sub-plots to solar radiation; with abundance and activity on estimates of species 300–330° = 1 (highest), 270–300° = 2, richness and community composition (York 1989). 330–360° = 3, 240–270° = 4, 210–240° = 5, and Weather during this period was typical for that 180–210° = 6 (lowest). Sites intermediate between time of year; temperatures ranged from 17–35°C categories were allocated an average (mean) code. and 27mm of rain fell between the 7th and 8th. To evaluate any large-scale spatial trend in Samples were returned to the laboratory and habitat variables (and species’ responses) a new examined with a binocular microscope where variable (position) was generated to reflected the material was sorted to the taxonomic level of north-south location of plots along Sandy Order. A number of groups were chosen for more Hollow Road (see Figure 2.1). The value of aspect detailed investigation based on the criteria of and position for each plot ranged from 1-6 (see sufficient numbers for statistical analysis, the Table 2.3). ability to recognise and define species, and the likely appropriateness of the sampling methodology. Ants (: Formicidae),

193 Australia’s Biodiveristy - Responses to Fire

Table 2.3 Values of exposure and position for component delineates the largest pattern of each sub-plot. Aspect was coded from 1-6 to relationships in the data (defines the greatest reflect the relative exposure of sub-plots to solar amount of variation in the data); the second radiation (see text). Position was coded from 1 delineates the next largest pattern and so on. (north) to 6 (south) to reflect location along Sandy 2.5.3 Terrestrial Invertebrate Hollow Road (see Figure 2.1) so as to detect Communities possible spatial patterns. Samples were sorted to Order using Exposure (sub-plots) morphological characteristics and general Treatment Plot Position 1 2 3 4 taxonomic keys. Relative abundance of individuals 1A 1 2 1.5 1 2 within these groups at plots differing in treatment 2A 2 4 4.5 4 3 and position were examined using Analysis of Unburnt 3A 3 2 2 2 2 Variance (ANOVA) procedures. 4A 5 5.5 4.5 6 6 5A 4 5 2 3 2 2.5.3.1 Biodiversity Selected taxa (see 2.4) were described in terms of 6A 6 3 5 5.5 5 the relative abundance of individuals within 1B 2 5 5 4 5.5 constituent groups (families, sub-families, genera 2B 1 1 1.5 3 1 etc. as appropriate), and their species richness (as Burnt 3B 3 1 1 2 1 defined by morphospecies). Patterns in these 4B 4 4 2 4 3 community descriptors at plots differing in 5B 5 2 3 4 3 treatment and position were examined graphically, 6B 6 5 6 6 5 and using Analysis of Variance (ANOVA) procedures. Patterns in environmental parameters at plots differing in treatment and position were 2.5.3.2 Community Composition examined graphically, and using Analysis of Patterns of species’ responses to treatments are Variance (ANOVA) procedures. For frequency illustrated in tables of relative abundance. This data, the degree of association between variables enabled broad “assemblages” of species, with was examined using contingency tables similar responses to disturbance, to be identified. (crosstabulation), with significant associations 2.5.3.3 Environmental Determinants of tested using the χ2 statistic. Community Composition 2.5.2 Inter-relationships Between The relative importance or ability of the measured Environmental Variables habitat variables to explain the composition of invertebrate assemblages was assessed using The environmental (habitat) variables were Canonical Correspondence Analysis (CCA, Ter subsequently analysed using an ordination Braak 1986, 1991). This method arranges species procedure (Principal Components Analysis - along environmental gradients by constructing PCA) in order to untangle linear relationships linear combinations of environmental factors between variables, and reflect inherent structural which result in maximal separation of species’ patterns. In this analysis, each pattern appears as a distributions in ordination species-space. These component delineating a distinct cluster of analyses were performed to determine whether interrelated data. Components are rotated any differences in turnover or spatial similarity of orthogonally (VARIMAX procedure) to clarify the assemblages among taxa might be explained by the definition of these clusters by maximising or different taxa responding to environmental minimising correlations between variables and gradients. Ter Braak (1986) fully describes the components. The projection (the loadings) of underlying assumptions and strengths of this each variable on the component axes defines the method. The main assumption is that individual clusters of variables. Kaiser’s criterion (only the species response models are all similar and of components with eigenvalues greater than one) unimodal, Gaussian form. Although it is doubtful was used to determine the number of components whether this assumption is reasonable for all to be extracted. Eigenvalues measure the amount species, CCA has been shown to be robust to of variation accounted for by a pattern, while the moderate violations of assumptions (Palmer 1993) loadings measure the degree to which variables and offers the potentially most powerful method are involved in the component pattern. The first

194 Bushfire and forest invertebrates

available in revealing patterns of community Broad groups of this kind (functional groups) were composition in relation to environmental factors. identified in this analysis by reference to the It also has the advantage that the results are literature (eg. see Andersen 1990) and following unaffected by correlations among environmental discussions with relevant taxonomists. The variables. numbers of morphospecies within these groups Results of the CCA ordination were was graphically presented and examined in order displayed as “bi-plots” which show the to detect those which may be sensitive to configuration of the variables, the scatter of sub- microhabitat features associated with structural plots, and the relationship between the two. This characteristics of the environment. gives an overview of how community composition Additionally, the relative abundance of varies with the environment (Ter Braak 1986). species recorded in one treatment only was The interpretation of the results of the bi-plots displayed in tabular form and discussed in relation was simplified by using a sub-set of the to their likely ecological roles. environmental variables in the analyses. This sub- 2.5.3.5 Biodiversity Indicators set was composed of representative variables from Decisions regarding conservation evaluation often each of the eight independent patterns identified are based upon the diversity (species richness) of by the Principle Components Analysis (see 2.5.2), the area under concern (see Margules et al. 1988). with the additional inclusion of two largely Similarly, species richness is a common independent variables: aspect and insolation. “performance indicator” used for monitoring 2.5.3.4 Community Structure disturbance impacts (see Kremen 1992). In order Analyses of species’ “assemblages” often fails to to simplify these processes, it is often adequately account for rare species, which are hypothesised that one taxonomic or functional frequently represented by too few records to allow group may reflect the response of other taxa, and any meaningful patterns to be determined (see hence function as “indicator” or “umbrella” York 1994). One common means to overcoming species. this problem, at least to some extent, is to group To test whether the richness of particular species according to some ecological invertebrate taxa could be useful in predicting characteristic, so that the collective behaviour of overall invertebrate biodiversity, the relationship the group can be assessed. At the species level between species richness of selected taxa was there is insufficient ecological information for investigated using correlation analyses (Pearson’s most groups to do this with confidence, however Product-Moment and Spearman’s Rank). broad grouping may be identified at higher taxonomic levels, such as sub-family or family.

195 Australia’s Biodiveristy - Responses to Fire

Forest area one day after a low-intensity fuel-reduction burn. The small areas of leaf litter and unburnt understorey vegetation remaining indicate the patchy nature of such burns. These represent potential refuges for terrestrial invertebrates; refuges that are reduced in number and extent by frequent fire. /Alan York

Dry eucalypt forest that has remained unburnt for over 25 years. These forests are characterised by deep leaf litter and low light levels. /Alan York

196 Bushfire and forest invertebrates

Dry eucalypt forest that has been frequently burnt for the past 25 years. These forests are very open and characterised by low leaf litter levels and high light levels. /Alan York

A spider from the Family Salticidae. These “jumping spiders” hunt for their food on understorey vegetation, trees and logs. The two species from this Family were only found on unburnt plots, suggesting that they prefer habitats with more structurally complex vegetation in A spider from the Family Zodariidae. These spiders which to hunt. /Alan York typically live under stones, logs and in leaf litter. There were four times as many species from this Family on frequently burnt plots, suggesting that they prefer these more open habitats in which to hunt. /Alan York

197 Australia’s Biodiveristy - Responses to Fire

An ant from the Probolomyrmex. It is a rare “cryptic” species that was only found in leaf litter samples collected from one site. Little is known about its habitat preferences, although it is not thought to be disadvantaged by frequent burning at this stage. /Alan York

A spider from the Family Lycosidae. These “wolf spiders” are ground hunters. All three species from this Family were only found on frequently burnt plots, suggesting that they prefer these more open habitats in which to hunt. /Alan York

This is a species of ant known as Rhytidoponera An ant from the genus Orectognathus. These ants are metallica which is an “opportunist” commonly found in specialist predators who use their long mandibles to disturbed habitats. In this study it was 500 time more catch soft-bodied insects such as Springtails abundant on frequently burnt sites, potentially indicating (Collembola). Because of their specialist habitat that frequent burning is having a negative impact on the requirements, they were not caught in pitfall traps but environment. /Alan York only in leaf litter samples collected from near large logs on unburnt sites. They were not found on frequently burnt sites. In these forests they could be considered an uncommon species with high conservation status. /Alan York

198 Bushfire and forest invertebrates

3. RESULTS

The results for the various components of the with patterns quite variable and independent of project are reported separately, and then discussed treatment and but not spatial location (position) in terms of their relevance to the existing within the study area (see Figure 3.1B). experimental fire regime. 3.1.1.3 Mid-sized Shrub Layer The cover of Mid-sized Shrubs (50–100cm) on 3.1 ENVIRONMENTAL PARAMETERS sub-plots showed considerable variation overall The following sections summarise the results of (means ranged from 2–47%) with average an investigation into the effects of frequent (mean±s.e.) values for unburnt and burnt plots burning on environmental variables as they relate 14.3±0.8 and 14.5±0.8 respectively. Cover of Mid- to components of terrestrial invertebrate habitat. sized Shrubs was similarly variable (“patchy”) on Where differences are described as “significant”, burnt and unburnt plots. this refers to statistical significance at a probability Mean values of Mid-sized Shrub cover were level of 0.05. Where results were not considered not significantly different between burnt and to be statistically significant, the general nature of unburnt plots, however there were significant any observed patterns is described. “Box and spatial trends, with patterns quite variable and Whisker” plots are utilised to graphically independent of treatment and but not spatial represent variation in environmental variables. location (position) within the study area (see Figure The box represents the interquartile range 3.1C). (25th–75th percentile) with the median shown. 3.1.1.4 Tall Shrub Layer The whiskers indicate the range of values which The cover of Tall Shrubs (100–150cm) on sub- lie within 1.5 box lengths of the upper and lower plots showed moderate variation overall (means quartile (75th and 25th percentile respectively). ranged from 2–12%) with average (mean±s.e.) values for unburnt and burnt plots 5.5±0.4 and 3.1.1 Understorey Vegetation Structure 1.9±0.2 respectively. Cover of Tall Shrubs was less 3.1.1.1 Ground Herb Layer variable (“patchy”) on unburnt plots. The cover of Ground Herbs (0–20cm) on sub- Mean values of Tall Shrub cover were plots showed considerable variation overall significantly different between burnt and (means ranged from 29–98%) with average unburnt plots, with substantially lower cover on (mean±s.e.) values for unburnt and burnt plots frequently burnt plots. There were however 72.5±1.5 and 80.1±1.3 respectively. Cover of significant spatial trends, with patterns quite Ground Herbs was less variable (“patchy”) on variable and independent of treatment and but not burnt plots. spatial location (position) within the study area (see Mean values of Ground Herb cover were not Figure 3.1D). significantly different between burnt and 3.1.1.5 Very Tall Shrub Layer unburnt plots, however there were significant The cover of Very Tall Shrubs (150–200cm) on spatial trends, with patterns quite variable and sub-plots showed moderate variation overall independent of treatment and spatial location (means ranged from 0–10%) with average (position) within the study area (see Figure 3.1A). (mean±s.e.) values for unburnt and burnt plots 3.1.1.2 Small Shrub Layer 2.8±0.3 and 0.2±0.6 respectively. Cover of Very The cover of Small Shrubs (20–50cm) on sub- Tall Shrubs was less variable (“patchy”) on plots showed considerable variation overall unburnt plots. (means ranged from 12–91%) with average Mean values of Very Tall Shrub cover were (mean±s.e.) values for unburnt and burnt plots significantly different between burnt and unburnt 49.5±1.7 and 57.7±1.6 respectively. Cover of Small plots, with substantially lower cover on frequently Shrubs was less variable (“patchy”) on burnt plots. burnt plots. There were however significant spatial Mean values of Small Shrub cover were not trends, with patterns quite variable and significantly different between burnt and unburnt independent of treatment and spatial location plots, however there were significant spatial trends, (position) within the study area (see Figure 3.1E).

199 Australia’s Biodiveristy - Responses to Fire

Figure 3.1 Understorey Vegetation Structure

200 Bushfire and forest invertebrates

Figure 3.2 Effects of repeated burning upon Litter components.

201 Australia’s Biodiveristy - Responses to Fire

3.1.1.6 Summary 3.1.2.2 Bark The amount (cover) of vegetation in the first Mean biomass of Bark on sub-plots varied from metre above the ground (Ground Herbs, Small 0.50–2.52 t.ha.-1, with the mean (±s.e.) values for Shrubs & Mid-sized Shrubs) was not significantly unburnt and burnt plots 1.35 (±0.06) and 0.86 different between burnt and unburnt plots. (±0.06) t.ha.-1 respectively. Bark biomass was quite Significant interactions between the three factors variable by nature and considerably more variable (treatment, position & sub-plot) however reflects the (“patchy”) on burnt plots. spatially variable nature of these understorey Mean values of Bark biomass were vegetation components, irrespective of fire history significantly different between burnt and (treatment) and large- (position) & small-scale (sub- unburnt plots, with unburnt plots having (on plot) location, within the study area. Cover of average) higher Bark biomass. There were no Ground Herbs and Small Shrubs was however less obvious broad spatial trends with this variable (see spatially variable (“patchy”) on burnt plots, while Figure 3.2B), however there was substantial the cover of Mid-sized Shrubs was similarly within-plot variation, irrespective of treatment, with spatially variable on burnt and unburnt plots. primarily a greater level of spatial variability In contrast, the amount (cover) of vegetation within burnt plots compared to unburnt plots. in the second metre above the ground (Tall & 3.1.2.3 Twigs 0-6mm Very-tall Shrubs) was significantly different Values of mean Twig biomass (0–6mm diam.) on between burnt and unburnt plots. There was sub-plots varied from 0.99–5.22 t.ha.-1, with the substantially lower cover on frequently burnt mean (±s.e.) values for unburnt and burnt plots plots, and it was more spatially variable. There 3.65 (±0.01) and 1.84 (±0.08) t.ha.-1 respectively. were however significant interactions between the Twig biomass (0–6mm) was quite variable by three factors (treatment, position & sub-plot) again nature and slightly more variable (“patchy“) on reflecting the spatially variable nature of these burnt plots (see Figure 3.2C). understorey vegetation components, irrespective Mean values of Twig biomass were of fire history (treatment) and large- (position) & significantly different between burnt and small-scale (sub-plot) location, within the study unburnt plots, with unburnt plots having (on area. average) higher Twig biomass. There was however 3.1.2 Litter Biomass substantial within-plot variation, irrespective of treatment, reflecting the substantial spatial For purposes of analysis, Dry Litter Biomass was “patchiness” of this variable. considered in 5 categories compatible with other studies of “fine fuel”. These categories were: 3.1.2.4 Twigs 6–25mm Leaves, Bark, Twigs 0–6mm & 6–25mm diameter, Values of mean Twig biomass (6–25mm diam.) on -1 and Miscellaneous (very fine) Material. Green sub-plots varied from 0.35–6.19 t.ha. , with the (live) and cured (dead) vegetation data collected as mean (±s.e.) values for unburnt and burnt plots -1 part of the fuel studies were not used in this 2.72 (±0.21) and 1.53 (±0.16) t.ha. respectively. analysis. Twig biomass (0-6mm) was quite variable by nature and slightly more variable (“patchy”) on burnt 3.1.2.1 Leaves plots (see Figure 3.2D). Mean biomass of Leaves on sub-plots varied from Mean values of Twig biomass were -1 2.02–9.84 t.ha. , with the mean (±s.e.) values for significantly different between burnt and unburnt and burnt plots 7.54 (±0.19) and 4.23 unburnt plots, with unburnt plots having (on -1 (±0.14) t.ha. respectively. Biomass of Leaves was average) higher Twig biomass. There was however similarly variable (“patchy”) on burnt and unburnt substantial within-plot variation, particularly plots. within burnt plots. Mean values of biomass of Leaves were significantly different between burnt and 3.1.2.5 Very Fine (Miscellaneous) Material unburnt plots, with unburnt plots having (on Mean biomass of Miscellaneous Material on sub- -1 average) higher leaf biomass. There was an plots varied from 0.33–3.94 t.ha. , with the mean obvious spatial (N-S) trend with regard to this (±s.e.) values for unburnt and burnt plots 2.28 -1 variable, which was more apparent for unburnt (±0.11) and 0.85 (±0.04) t.ha. respectively. plots (see Figure 3.2A). Miscellaneous Material biomass was quite variable by nature but similarly variable (“patchy”) on burnt and unburnt plots.

202 Bushfire and forest invertebrates

Mean values of Miscellaneous Material were 3.1.3 Sticks & Logs significantly different between burnt and Sticks & Logs in excess of 2.5cm diameter (“heavy unburnt plots, with unburnt plots having (on fuel”) ranged overall from 2.5cm to 170cm average) higher amounts of Miscellaneous diameter, with a number of large logs still present Material. There were obvious broad spatial trends on the forest floor following the 1959 post- with this variable, particularly for burnt plots (see logging silvicultural treatment (T.S.I.). Figure 3.2E), however there was substantial There were 474 Sticks & Logs recorded within-plot variation, irrespective of treatment. overall on unburnt plots, ranging from 3.1.2.6 Summary 2.5–170cm diameter (mean±s.e.= 15.2±0.9cm) Twig (0–6mm & 6–25mm) and Bark biomass were with a similar pattern apparent on burnt plots significantly higher on unburnt plots, however (423 overall, ranging from 2.5–150cm diameter, there was substantial within-plot variation (spatial mean±s.e.= 13.6±0.9cm). Numbers of Sticks & “patchiness”) of these variables, irrespective of Logs per 40m transect on each sub-plot varied treatment. overall from 3–47, with average numbers slightly The biomass of Leaves was also significantly higher on unburnt sub-plots. higher on unburnt plots, with a spatial trend For purposes of analysis, size (diameter) of apparent, irrespective of treatment. This trend Sticks & Logs were divided into 5 size classes based was more apparent however on unburnt plots. on a visual inspection of the frequency distribution. Values for the very fine litter component Mean (±s.e.) numbers of Sticks & Logs in these 5 (Miscellaneous Material) were significantly higher size categories are shown in Table 3.1. on unburnt plots, with a spatial patterning Analysis of frequency distributions within apparent on burnt but not unburnt plots. nominated size categories revealed that there was Twig biomass (0–6mm & 6–25mm) was no obvious overall spatial pattern with regard to slightly more variable (“patchy”), and Bark the distribution of Sticks & Logs on either burnt biomass considerably more variable on burnt and unburnt plots, although considerable plots. In contrast, levels of within-plot variation variation was apparent, particularly for sticks and were similar for both the biomass of Leaves and of small logs between 10 and 50cm diameter. Field Miscellaneous Material on burnt and unburnt inspections revealed a change in character plots. however for larger logs on burnt plots with considerable charring and desiccation of the outer surface.

Table 3.1 Numbers of sticks & logs per sub-plot in 5 size classes (mean±s.e.) Diameter Unburnt (control) plots mean Frequently burnt plots mean Class 123456±se123456±se 2.5–10cm 9.8 6.3 10.3 12.5 14.0 12.5 10.9 9.0 8.3 9.0 16.3 8.9 11.0 10.3 ± 1.1 ± 2.1 ± 0.6 ± 3.0 ± 2.9 ± 1.1 ± 1.1 ± 3.1 ± 2.0 ± 1.5 ± 3.7 ± 1.7 ± 3.2 ± 1.1 10–25cm 6.5 4.0 7.3 5.3 5.8 6.3 5.9 5.0 3.0 5.3 10.5 3.8 4.3 5.3 ± 2.9 ± 2.0 ± 1.5 ± 1.7 ± 1.5 ± 3.2 ± 0.8 ± 2.4 ± 1.8 ± 1.1 ± 3.2 ± 0.9 ± 1.1 ± 0.9 25–50cm 2.8 1.0 1.0 2.3 1.5 2.5 1.8 1.5 0.8 1.8 1.8 2.0 0.5 1.4 ± 1.5 ± 0.7 ± 0.7 ± 1.1 ± 0.6 ± 1.6 ± 0.4 ± 0.5 ± 0.3 ± 1.4 ± 0.8 ± 1.1 ± 0.3 ± 0.3 50–75cm 1.0 1.0 0.8 0.3 0.8 0.3 0.7 0.8 0.3 0.5 0.3 0.3 0.5 0.4 ± 0.4 ± 0.4 ± 0.3 ± 0.3 ± 0.5 ± 0.3 ± 0.1 ± 0.5 ± 0.3 ± 0.3 ± 0.3 ± 0.3 ± 0.3 ± 0.1 75+ cm 0.5 0.5 0.0 1.0 0.8 0.3 0.5 0.5 0.5 0.3 0.0 0.0 0.5 0.3 ± 0.3 ± 0.5 ± 0.0 ± 0.7 ± 0.5 ± 0.3 ± 0.2 ± 0.5 ± 0.3 ± 0.3 ± 0.0 ± 0.0 ± 0.3 ± 0.1 Total 20.5 12.5 19.3 21.3 23.3 21.8 19.8 16.8 12.8 16.8 28.8 14.0 16.8 17.6 ± 5.2 ± 2.1 ± 2.8 ± 4.1 ± 3.7 ± 8.9 ± 1.9 ± 5.9 ± 3.0 ± 2.6 ± 5.1 ± 3.4 ± 4.7 ± 1.9

203 Australia’s Biodiveristy - Responses to Fire

3.1.4 Insolation 3.1.5.2 Top-soil Hardness Overall measures of the mean Top-soil Hardness Overall measures of the mean Insolation Index (shear) index varied from 1.9–10.0, with mean (Percentage of Light at Ground Level) varied from (±s.e.) values for unburnt and burnt plots 5.3 1.1 - 35.5%, with the mean (±s.e.) values for (±0.1) and 6.3 (0.2) respectively. Mean values of unburnt and burnt plots 7.3 (±0.7) and 16.4 (±1.1) Top-soil Hardness were not significantly different respectively. Insolation was extremely “patchy” by between burnt and unburnt plots. nature with the CV ranging from 31.5 - 68.6% There was a significant position effect (mean = 45.8%) on unburnt plots and 20.4 - reflecting a general north-south spatial pattern in 51.6% (mean = 30.6%) on burnt plots. Insolation soil hardness (see Figure 3.4B). There was was therefore less spatially variable (“patchy”) on however a slight tendency for a treatment/position unburnt plots. interaction caused by substantially higher values Mean values of Insolation were significantly on two burnt plots, 2 and 4 (positions 1 and 4). different between burnt and unburnt plots, with burnt plots having (on average) higher percentage light levels at ground level. There was however a general trend for increasing Insolation from north to south on burnt plots, with a different pattern evident on unburnt plots. This position effect was significant and is illustrated in Figure 3.3. 3.1.5 The Soil Environment

3.1.5.1 Top-soil Moisture Overall measures of mean Top-soil Moisture content on sub-plots varied from 4.5–17.7%, with mean (±s.e.) values for unburnt and burnt plots of 10.3% (±0.6) and 8.4% (±0.4) respectively. Mean values of Top-soil Moisture were significantly different between burnt and unburnt plots, with unburnt plots having (on average) higher percentage moisture levels. There was a significant large-scale spatial (position) effect, irrespective of treatment, with a slight increase apparent from position 1 to 3, and a gradual decline thereafter (see Figure 3.4A). Figure 3.4 Top-soil Moisture and Top-soil Hardness

3.2 INTER-RELATIONSHIPS BETWEEN ENVIRONMENTAL VARIABLES

The Principle Components Analysis (PCA) identified 8 significant themes in the environmental data. The 31 habitat variables were thus simplified to 8 new variables (Components) representing independent (uncorrelated) “patterns” in the environment. These Figure 3.3 Percentage Light at Ground Level on burnt Components are displayed in a matrix which and unburnt plots. shows the loadings of each variable for each Component (Table 3.2). These 8 Components described 75.1% of the overall variance in the environmental data set, thus substantially simplifying and clarifying structural patterns.

204 Bushfire and forest invertebrates .645 –.341 –.591 .529 .377 .721 .853 .847 .851 –.818 –.319 .373 .549 .638.861 .817 –.437 23.0 21.2 7.0 6.1 5.4 4.8 4.0 3.6 Logs 50+cm CV Topsoil hardness CV Topsoil CV Leaves tall shrubsCV Very –.376 Table 3.2 Significant Principle Components (eigenvalues > 1) showing variable loadings 0.3 Table VariableGround herbsCV Ground herbsSmall shrubsCVSmall shrubsMid-sized shrubsCV Mid-sized shrubsAspect Component 1 .946 –.949 Component 2 –.938 .940 .753 –.779 Component 3 .596 Component 4 Component 5 Component 6Component 7 Component 8 .367 SlopeEigenvalue explained % Variance Cumulative Variance 23.0 7.127 44.2 6.570 –.447 51.2 2.174 57.3 1.903 1.665 62.7 1.482 67.5 1.244 71.5 1.120 75.1% Twigs 0–6mmTwigs 6–10mmTwigs LeavesBarkCV BarkFine litter moistureTopsoil PLAG .311CV PLAG shrubsTall shrubsCV Tall tall shrubsVery –.438Sticks 0.5–10cm .614 Logs 10–25cm .901 Logs 25–50cm hardnessTopsoil –.349 –.550 .832 .567 CV Fine litter .686 .777CV twigs 0–6mm –.316 6–10mm CV Twigs .357 .614 –.684 .354 .352 .722 .378 .696 –.734 .339

205 Australia’s Biodiveristy - Responses to Fire

The first Component concerns the absence Table 3.3 Interpretation of environmental and spatial patchiness of Ground Herbs, Small Principal Components (based on identification Shrubs and Mid-sized Shrubs, and explains 23% of variables with high loadings - see Table 3.2). of the variance in the environmental data set (see Component Interpretation Table 3.3). This Component describes a pattern in the data whereby the first three structural 1 Absence and spatial patchiness of vegetation layers (0–1m) are positively correlated Ground Herbs, Small Shrubs and (ie. vary together). As the amount of vegetation in Mid-sized Shrubs. these layers increases, it becomes more spatially 2 Abundance of litter (Twigs, Leaves, homogenous. Conversely, low amounts of these Bark & Miscellaneous Material [very structural vegetation Components imply spatial fine litter]); high Top-soil Moisture; patchiness of the vegetation. There is a weak low and spatially variable amounts of tendency for there to be greater amounts of Insolation at ground level. vegetation in the first metre above the ground on 3 Presence and spatial homogeneity of sub-plots with more exposed (north-westerly) Tall Shrubs, presence of Very Tall aspects. Shrubs. The second Component concerns the 4 Presence of Sticks & Logs abundance of litter (Twigs, Bark, Leaves & (0.5–10cm, 10–25cm & 25–50cm Miscellaneous Material [very fine litter]), high diameter). Top-soil Moisture, and low and spatially patchy 5 Low and spatially variable Top-soil amounts of Insolation at ground level. This Hardness. describes a pattern that with increasing litter 6 Spatial patchiness of Leaves and levels (all Components), top-soil moisture levels Miscellaneous Material [very fine increase and the amount of insolation reaching litter]. ground level decreases and becomes more 7 Spatial patchiness of Twigs (0-6, spatially variable. This Component explains an 6–10mm) and Very Tall Shrubs. additional 21.2% of the variance in the 8 Presence of large logs (50+cm environmental data set. diameter) and low slope angles. The third Component concerns the presence and spatial homogeneity of Tall Shrubs, and presence of Very Tall Shrubs, and explains an As the first 3 Components together explain additional 7% of the variance in the environmental over half (51.2%) of the variance in the data set. This Component describes a pattern environmental data set (Table 3.3), their whereby when vegetation occurs in the upper Component scores were plotted to see how the understorey layers (1–2m) it tends to occur in both individual sub-plots were responding (Figure 3.5). elements of these layers, however it is spatially The complexity of 3-dimensions is displayed in patchy. two plots for clarity. The “elevation” view Subsequent Components do not explain illustrates the projection of Components 1 & 2, substantial additional variance in the representing 44.2% of the variance in environmental data set, however they do illustrate environmental variables. The “plan” view that many of the patterns in the habitat data are illustrates the projection of Components 1 & 3, independent (ie. uncorrelated). For example, the representing 23% and 7% of the variance number of Sticks & Logs in certain size categories respectively. (Component 4) are inter-correlated and unrelated A number of features are apparent. Firstly, in to the amount of Litter biomass at sub-plots the elevation view, there is a distinct separation of (Component 2). Top-soil Hardness (Component burnt and unburnt sub-plots along Component 5) is site-specific and independent of other 2. This reflects the distinct treatment effect environmental variables. The presence of large detected for Litter biomass (see 3.1.2). This logs on sites (Component 8) is also site-specific, separation is not apparent along Components 1 although there is a tendency for a greater number and 3, where there are no significant treatment on sub-plots with lower slope angles. effects. That is, the amount of vegetation in the first metre above ground is independent of treatment (see 3.1.1.6). Secondly, sub-plots within each plot are only loosely grouped together,

206 Bushfire and forest invertebrates

reflecting the lack of spatial patterning (position) in 3.3 TERRESTRIAL INVERTEBRATE understorey vegetation cover (see 3.1.1) and Litter COMMUNITIES Biomass (see 3.1.2). The relationship of these environmental In excess of 55,000 individuals from 24 broad patterns to the abundance and distribution of taxonomic groups were collected during pitfall terrestrial invertebrates is explored and reported trapping. These indicated a rich fauna with in the following sections of this report. representatives from the Chelicerata (spiders, ticks & mites, pseudoscorpions, harvestmen), Crustacea (landhoppers, slaters), Chilopoda (centipedes), Diplopoda (millipedes), and a large number of Insect Orders & Families (see Table 3.4). 3.3.1 Ordinal Diversity Of the 24 Orders recorded overall, ordinal diversity on sub-plots varied from 11–17, with mean (±s.e.) values for unburnt and burnt plots 14.0 (±0.3) and 12.7 (±0.2) respectively. A 2-way ANOVA indicated a significant effect (F1,5 = 20.65 P = 0.006) of treatment, but not position (F5,36 = 0.61 P = 0.693) on the number of Orders represented on sub-plots. Unburnt sub-plots were significantly more diverse at the Ordinal level than burnt plots, although the actual mean difference was slight (1 Order on average). Interaction between the factors treatment and position was not significant (F5,36 = 0.61 P = 0.693). This result suggests that frequent burning has slightly reduced the terrestrial invertebrate diversity at the Ordinal level (see Figure 3.6).

Figure 3.5 Habitat conditions on sub-plots as shown by scores on Components 1, 2 & 3.

Figure 3.6 Spatial Variation in Ordinal Diveristy

3.3.2 Invertebrate Abundance Numerically, the most abundant groups overall were the springtails (33.1%), ticks & mites (23.9%), ants (23.1%), bugs (4.2%), beetles (4.0%), bees & wasps (2.8%), insect larvae (2.7%), flies (2.6%) and spiders (2.2%), with these nine groups making up 98.6% of the total number of

207 Australia’s Biodiveristy - Responses to Fire iance results testing the ction 3.3.2). mean (s.e.) [n] mean (s.e.) [n] (millipedes) - - [2] - - - [3] [5] (centipedes) - - [6] - - - [7] [13] Diplopoda Araneae (spiders)Acarina (ticks & mites)Pseudoscorpionida (pseudoscorpions)Opilionida (harvestmen)CRUSTACEA Amphipoda (landhoppers)Isopoda (slaters) -UNIRAMIA Chilopoda 227.8 28.7Hexapoda Collembola (springtails) - -Diplura (diplurans) (20.3) -Blattodea (cockroaches) (1.5)Isoptera () [5,468]Dermaptera (earwigs) [9]Orthoptera (grasshoppers, crickets) 6.4 -Embioptera (embiids) [688]Psocoptera (booklice) 352.2 -Hemiptera (bugs)Thysanoptera (thrips) 2.8 (1.0) [0] 0.7Neuroptera (lacewings) (21.6)Coleoptera (beetles) P,T - 1Diptera (flies) - 0.7 [153] (caddisflies)Tricoptera TP - (0.5) [8,452]Lepidoptera (moths, ) (0.2)Hymenoptera (bees, wasps) - 328.9Hymenoptera (ants) - - (0.2) [67] - 61.9 2.1 [17] 21.5 - (28.7) - {T} 37.9 TP - [16] - [1]{} indicates not tested but trend apparent from data. - 29.0 - (7.1) (0.5) (1.5) - [7,895] 17.8 [7] (3.2) - 2.3 {T} 417.8 [1,487] [0] 10.3 {T} [13,363] - [50] (2.0) [0] [517] 416.2 (1.4) - [909] {T} (0.5) - (28.7) [0] (0.9) [4] [696] [12] (20.1) [1,205] 1.2 [427] 2.0 - [10,028] [2] T - [54] [248] [9,988] 2.4 - [13] (0.2) - [1] (0.3) P,T, - [1] [18,440] - [401] (0.4) - - [55] TP T,P, 35.0 T,TP [28] [49] - 1.2 [1] - 55.5 - - 119.5 (3.2) [58] - 42.3 - 34.9 [95] (0.4) (3.6) [66] - - (12.3) [7] - [841] (2.9) - [74] (2.7) [1,332] [0] - [29] [2,867] - [2,328] [1] [1,016] [8] [3] [839] [2,241] - [12,895] [7] [1] [79] [1,443] [9] [1,535] [1] [3] [0] [1] [21] [2] Table 3.4 Abundance of terrestrial invertebrates on frequently burnt and unburnt (control) plots, showing 3-way Analysis Var Table of treatment (burnt/unburnt) and position (N-S). effects TaxaCHELICERATA Burnt Plots * Anova#Larvae (insect)Totals* For Burnt and Unburnt replicates: values represent mean (± standard error) for n = 24, and [total caught] P = Position, TP indicates significant interaction term (refer Se # Anova results show significant effects where T = Treatment, Unburnt Plots * 24.2 Total (3.2) [580] [29,078] T 37.5 (3.1) [899] [1,479] [26,696] [55,774]

208 Bushfire and forest invertebrates

organisms caught. The first three groups were no significant position or interaction effects (springtails, ticks & mites, and ants) represented although values for burnt sub-plots at position 1 80% of individuals caught. were, on average, much lower (see Figure 3.7). The numbers collected from several groups were insufficient to comment about possible effects of frequent burning. These were the pseudoscorpions, harvestmen, centipedes, millipedes, diplurans, termites, embiids, booklice, lacewings, caddisflies, moths and butterflies. For these groups the trapping method used may not have been the most appropriate and has potentially contributed to the low capture rate. While the low numbers collected for several other taxa precluded statistical analysis, a general level of knowledge concerning their habitat requirements permits a few preliminary observations to be drawn. Frequent burning Figure 3.7 Abundance of Collembola on burnt and appears to have led to a reduction in the numbers unburnt plots of amphipods, cockroaches and earwigs, and an increase in the numbers of grasshoppers & crickets, and thrips. 3.3.2.3 Ticks & Mites For ten broad taxonomic groups there were Ticks and mites (Acarina) were numerically the sufficient data to permit statistical testing. A two- second most abundant organisms found during way Analysis of Variance procedure (ANOVA) was this trapping program, accounting for about 24% used to investigate treatment (burnt vs unburnt) of individuals. A 2-way Analysis of Variance and position (large spatial scale) effects, and (ANOVA) procedure (with ln(x+1) possible interactions between the factors. transformation) indicated that the effect of Significant treatment and position interaction treatment was significant (F1,5 = 21.03 P = 0.006), effects were detected for a number of taxa. with approximately 31% less individuals caught 3.3.2.1 Isopods (on average) on burnt plots. Variation in relative abundance between Overall only 401 slaters (Isopoda) were caught in plots within each treatment was high, with a pitfall traps. Although approximately 38% less significant position effect (F individuals were caught (on average) on burnt 1,36 = 1.11 P<0.001), plots, the effect of treatment was not significant reflecting a spatial pattern in abundance, and (F = 2.84 P = 0.153). Variation in relative greater numbers of individuals being caught on 1,5 unburnt abundance between plots within each treatment was plots at position 5 & 6, with no obvious high, with a significant treatment/position interaction treatment effect observed at other positions (see Figure 3.8). (F5,36 = 3.78 P = 0.007). This was due to greater numbers of individuals being caught on unburnt plots at position 1, 4, 5 & 6, and on the burnt plot at position 2. Only 4 species were detected (Fiona Lewis, Macquarie University pers. comm.) precluding further reliable analysis of these data. 3.3.2.2 Springtails Springtails (Collembola) were numerically the most abundant organisms found during this trapping program, accounting for about 33% of individuals. Approximately 15% less individuals were caught (on average) on burnt plots, however a 2-way Analysis of Variance (ANOVA) indicated a non-significant effect of treatment (F1,5 = 2.79 P = 0.156). Variation in relative abundance between Figure 3.8 Abundance of Acarina on burnt and unburnt plots within each treatment was low, and there plots

209 Australia’s Biodiveristy - Responses to Fire

At this stage taxonomic difficulties preclude 3.3.2.6 Hemiptera detailed examination at the species level, however Overall 2,328 individual bugs (Hemiptera) were the matter is currently being discussed with the trapped, which represented 2.2% of the total Australian Museum. number of organisms caught. Hemipteran 3.3.2.4 Hymenoptera (excluding ants) abundance on sub-plots varied from 9–138 As a component of the Hymenoptera, bees and individuals, with the distribution of values wasps are not effectively sampled by pitfall positively skewed (mean±s.e.= 48.5±4.3, median = trapping. However, over 1,500 individuals were 37.5). The abundance of Hemiptera overall was caught, with no significant effects of treatment or more variable on burnt (23–138) compared to position. Values on unburnt sub-plots at position 2 unburnt (9-88) sub-plots. were substantially higher than at burnt sub-plots, Of the total number of Hemiptera caught however they were within the range experienced 841 (36%) were trapped on unburnt sub-plots by other plots from that treatment (see Figure 3.9). and 1,487 (64%) on burnt sub-plots. Mean (±s.e.) & median values for unburnt and burnt plots were 35.0 (±3.2) & 30.5 and 61.9 (±7.1) & 46.5 respectively. A 2-way Analysis of Variance (ANOVA) procedure (with ln(x+1) transformation) indicated that the effect of treatment was significant (F1,5 = 10.93 P=0.021) with approximately 77% more individuals caught (on average) on burnt plots (Figure 3.11). Variation in relative abundance between plots within each treatment was high, although the effect of position was not significant (F5,36 = 2.21 P = 0.075). Abundance was higher on all burnt plots except position 6, although interaction between Figure 3.9 Abundance of Hymenoptera (non-ants) on treatment and position was not significant (F = burnt and unburnt plots 5,36 2.91 P = 0.241). 3.3.2.5 Insect Larvae For insect larvae, a 2-way Analysis of Variance (ANOVA) procedure indicated that the effect of treatment was significant (F1,5 = 14.67 P = 0.012), with approximately 35% less individuals caught (on average) on burnt plots. Variation in relative abundance between plots within each treatment was low with no significant position or interaction effects (see Figure 3.10).

Figure 3.11 Abundance of Hemiptera on burnt and unburnt plots

3.3.2.7 Diptera Although pitfall trapping may not appear to be the preferred method of sampling flies (Diptera), 1,443 individuals were caught, representing 2.6% of the total number of organisms. Dipteran abundance on sub-plots varied from 8–64 individuals, with the distribution of values slightly Figure 3.10 Abundance of Insect Larvae on burnt and positively skewed (mean = 30.1±2.4, median = unburnt plots

210 Bushfire and forest invertebrates

28.0). The abundance of Diptera overall was more 21.5 (±1.5) & 22.0 and 28.7±1.5 & 28.5 variable on unburnt (9–64) compared with burnt respectively. There were (on average) 33% more (8–34) sub-plots. spiders on frequently burnt sub-plots. Of the total number of Diptera caught 1,016 A 2-way ANOVA indicated a non-significant (70%) were trapped on unburnt sub-plots and effect of treatment (F1,5 = 1.48 P = 0.278) and 427 (30%) on burnt sub-plots. Mean (±s.e.) & position (F5,36 = 1.45 P = 0.229), with a significant median values for unburnt and burnt plots were interaction detected between these two factors 42.3 (±2.9) & 43.0 and 17.8 (±1.4) & 16.0 (F5,36 = 3.24 P = 0.016). respectively. Mean values of spider abundance were A 2-way Analysis of Variance (ANOVA) therefore not significantly different between procedure indicated that the effect of treatment burnt and unburnt plots, although burnt plots was significant (F1,5 = 24.30 P = 0.004), with have (on average) higher numbers of individuals approximately 58% less individuals caught (on (see Figure 3.13). The significant interaction average) on burnt plots. between treatment and position reflects the reversal Variation in relative abundance between in the pattern of the treatment effect at position 3 plots within each treatment was high, with a and 4, where burnt plots had slightly lower spider marginally non-significant position effect (F5,36 = abundance than unburnt, and at position 6 where 2.22 P = 0.073) and a significant treatment/position the mean value on burnt plots (11.3±1.7) is interaction effect (F5,36 = 3.49 P = 0.011). This substantially higher than on unburnt plots was due to greater numbers of individuals being (4.8±0.5). caught on unburnt plots at position 2, 3, 4, 5 & 6, and on the burnt plot at position 1 (see Figure 3.12).

Figure 3.13 Abundance of Spiders on burnt and unburnt plots

Figure 3.12 Abundance of Diptera on burnt and unburnt plots 3.3.2.9 Beetles Overall 2,145 individual beetles were trapped, 3.3.2.8 Spiders which represented 4.0% of the total number of Overall 1,205 individual spiders (Araneae) were organisms caught. Beetle abundance on sub-plots trapped, which represented 2.2% of the total varied from 13–88 individuals, with the with the number of organisms caught. Spider abundance on distribution of values slightly positively skewed sub-plots varied from 8–42 individuals, with the (mean = 44.7, median = 43.5). The abundance of distribution of values slightly positively skewed beetles overall was slightly less variable on burnt (mean ±s.e= 25.1±1.1, median = 23.0). The (13–67) compared with unburnt (29–88) sub-plots. abundance of spiders overall was slightly more Of the total number of beetles caught, 1,269 variable on unburnt (8-40) compared with burnt (59%) were trapped on unburnt sub-plots and (16–42) sub-plots. 876 (41%) on burnt sub-plots. Mean (±s.e.) & Of the total number of spiders caught 515 median values for unburnt and burnt plots were (43%) were trapped on unburnt sub-plots and 52.9 (±3.2) & 47.5 and 36.5 (±2.9) & 34.0 688 (57%) on burnt sub-plots. Mean (±s.e.) & respectively. There were (on average) 31% less median values for unburnt and burnt plots were beetles on frequently burnt sub-plots.

211 Australia’s Biodiveristy - Responses to Fire

A 2-way ANOVA (with logex position (F5,36 = 12.01 P<0.001), with a significant transformation) indicated a significant effect of interaction detected between these two factors treatment (F1,5 = 8.75 P = 0.032) and position (F5,36 = 2.82 P = 0.030). (F5,36 = 6.65 P<0.001), with a significant Mean values of ant abundance were interaction detected between these two factors therefore significantly different between burnt (F5,36 = 3.44 P = 0.012). and unburnt plots, with burnt plots having (on Mean values of beetle abundance were average) much higher numbers of individuals. The therefore significantly different between burnt position effect and treatment/position interaction and unburnt plots, with burnt plots having (on reflects the strong spatial trend in ant abundance average) lower numbers of individuals. The with higher values at position 1, particularly for position effect reflects the strong spatial trend in burnt plots (see Figure 3.15). beetle abundance, particularly for burnt plots (see Figure 3.14). The significant interaction effect reflects the reversal in the pattern of the treatment effect at position 4, where burnt plots had slightly higher beetle abundance than unburnt plots.

Figure 3.15 Abundance of Ants on burnt and unburnt plots

3.3.2.11 Summary This forest environment has a abundant and Figure 3.14 Abundance of Beetles on burnt and unburnt plots diverse terrestrial invertebrate fauna with in excess of 55,000 individuals from 24 broad taxonomic groups collected during a one week sampling 3.3.2.10 Ants period using pitfall traps. Numerically, the most Ants (Hymenoptera:Formicidae) represented the abundant groups overall were the springtails third most abundant group trapped (12,895 (33.1%), ticks & mites (23.9%), ants (23.1%), bugs individuals), accounting for 23% of the total (4.2%), beetles (4.0%), bees & wasps (2.8%), insect catch. Ant abundance on sub-plots varied from larvae (2.7%), flies (2.6%) and spiders (2.2%). 24–778 individuals, with the with the distribution Due to their low numbers, it was not of values slightly positively skewed (mean±s.e.= possible to comment on the effects of frequent 268.6±26.7, median = 219.0). The abundance of burning for: pseudoscorpions, harvestmen, ants overall was considerably more variable on centipedes, millipedes, diplurans, termites, burnt (155–778) compared with unburnt embiids, booklice, lacewings, caddisflies, moths (24–300) sub-plots. and butterflies. For these groups the trapping Of the total number of ants caught 2,867 method used may not have been the most (22%) were trapped on unburnt sub-plots and appropriate and has potentially contributed to the 10,028 (78%) on burnt sub-plots. Mean (±s.e.) & low capture rate. While the low numbers median values for unburnt and burnt plots were collected for several other taxa precluded 119.5 (±12.3) & 102.5 and 417.8 (±28.7) & 401.5 statistical analysis, frequent burning appears to respectively. There were (on average) 250% more have led to a reduction in the numbers of ants on frequently burnt sub-plots. amphipods, cockroaches and earwigs, and an A 2-way ANOVA indicated a significant increase in the numbers of grasshoppers & effect of treatment (F1,5 = 77.82 P < 0.001) and crickets, and thrips.

212 Bushfire and forest invertebrates

Overall, the number of invertebrate Orders The groups studied at morphospecies level had been significantly reduced on sub-plots were the: experiencing frequent burning, although the Hemiptera (bugs): magnitude of this decrease was small (average ≈ 1 as most are terrestrial and phytophagous Order). (plant-feeding), they are a group which have For ten broad taxonomic groups there were a close association with plant communities. sufficient data to permit statistical testing using Diptera (flies): the Analysis of Variance (ANOVA) procedure to although highly mobile as adults, the flies investigate the effects of frequent burning have particular requirements with regard to (treatment) and patterns due to large-scale spatial larval food sources; usually moist, decaying effects (position). These results are summarised in plant and animal material. Many species are Table 3.5 and indicate a variety of responses to parasitic on the larvae of other insect orders. frequent burning. Seven groups (isopods, Araneae (spiders): springtails, ticks & mites, bees & wasps, insect spiders are a major group of predators in larvae, flies and beetles) showed substantial forest ecosystems exploiting a variety of decreases in abundance following frequent habitats. They live in burrows or crevices in burning. These decreases ranged from 15 to 58%, the ground, amongst leaf litter or in but were only statistically significant for ticks & vegetation. mites, insect larvae, flies and beetles. High spatial Coleoptera (beetles): variability in abundance for isopods, springtails, beetles utilise a diverse range of habitats & and bees & wasps possibly contributed to the lack micro- habitats, with a variety of feeding of statistical significance. strategies (adults include herbivores, Three groups showed substantial increases predators & scavengers, while larval forms in abundance following frequent burning. These feed either internally or externally on plants were statistically significant for bugs (77%) and ants and fungal products). (250%), but not for spiders (33%). Both spiders Formicidae (ants): and ants showed considerable spatial variability in ants are one of the most numerous and their numbers. widespread groups in Australian ecosystems. They have a diverse diet, and utilise a variety 3.3.3 Invertebrate Species Richness of feeding strategies from predators and Five invertebrate groups were identified to scavengers, to plant eaters and morphospecies (see Section 2.4) in order to feeders, with frequent and varied further investigate the impact of repeated burning interactions with other invertebrate groups. on species richness, and the related aspects of These groups are discussed separately with community composition and structure. These related issues considered in Sections 3.3.3.6 groups utilise a diversity of micro-habitats and & 7. niches and are representative of the range of terrestrial invertebrates found in these forest environments. Table 3.5 Changes in mean abundance following frequent burning for selected terrestrial invertebrate taxa. Change with Statistically Large-scale Taxa Common name frequent burning significant ? (P<0.05) spatial patterns ? Isopoda slaters, pill-bugs ⇓ 38% no yes Collembola springtails ⇓ 15% no no Acarina ticks & mites ⇓ 31% yes yes Hymenoptera (excl. ants) bees & wasps ⇓ 17% no yes Insecta insect larvae ⇓ 35% yes no Hemiptera bugs ⇑ 77% yes no Diptera flies ⇓ 58% yes no* Araneae spiders ⇑ 33% no yes* Coleoptera beetles ⇓ 31% yes yes* Formicidae ants ⇑ 250% yes yes* * indicates a statistically significant interaction between treatment and position effects.

213 Australia’s Biodiveristy - Responses to Fire

3.3.3.1 Hemiptera A 2-way ANOVA indicated a significant Overall, 44 morphospecies of Hemiptera were effect of treatment (F1,5 = 14.84 P = 0.012) and collected, 25 on unburnt plots and 26 on frequently position (F5,36 = 24.67 P = 0.031), with no burnt plots. Hemipteran richness on sub-plots significant interaction detected between these two varied from 1–8, with a normal distribution of factors (F5,36 = 1.82 P = 0.134). values (mean = 3.4 (±0.2), median = 3.0). Mean values of fly richness were therefore Mean (±s.e.) & median values for unburnt significantly different between burnt and and burnt plots were 3.2 (±0.3) & 3.0 and 3.7 unburnt plots, with burnt plots having (on (±0.3) & 3.0 respectively. There were (on average) average) lower numbers of morphospecies (see 16% more Hemipteran species on frequently Figure 3.17). The weak position effect identified burnt sub-plots. The number of morphospecies for fly abundance (see 3.3.2.7) was more evident was similarly variable on unburnt (1–8) and for fly species richness, but only for burnt plots. burnt (1–7) sub-plots. Values of species richness were quite consistent A 2-way ANOVA (with logex between unburnt plots. transformation) indicated a non-significant effect of treatment (F1,5 = 1.99 P = 0.217) and position (F5,36 = 1.49 P = 0.218), with a non-significant interaction detected between these two factors (F5,36 = 0.60 P = 0.699). Mean values of Hemipteran richness were therefore not significantly different between burnt and unburnt plots, with both treatments having (on average) similar numbers of morphospecies (see Figure 3.16).

Figure 3.17 Richness of Flies on burnt & unburnt plots

3.3.3.3 Spiders Overall, 63 morphospecies of spiders were collected, 32 on unburnt plots and 47 on frequently burnt plots. Spider richness on sub- plots varied from 1–10, with a normal distribution of values (mean = 4.9 (±0.3), median = 5.0). Mean (±s.e.) & median values for unburnt and Figure 3.16 Richness of Hemiptera on burnt & unburnt burnt plots were 4.4 (±0.4) & 4.0 and 5.6 (±0.4) & plots 6.0 respectively. There were (on average) 27% more spider species on frequently burnt sub-plots. The 3.3.3.2 Diptera number of morphospecies was similarly variable on Overall, 77 morphospecies of flies were collected, 66 unburnt (1–8) and burnt (2–10) sub-plots. on unburnt plots and 46 on frequently burnt plots. A 2-way ANOVA indicated a non-significant Fly richness on sub-plots varied from 0 –16, with a effect of treatment (F1,5 = 2.43 P = 0.180) and normal distribution of values (mean = 7.9 (±0.6), position (F5,36 = 0.92 P = 0.482), with a weak median = 8.0). interaction detected between these two factors Mean (±s.e.) & median values for unburnt (F5,36 = 2.27 P = 0.069). and burnt plots were 10.1 (±0.6) & 10.0 and 5.7 Mean values of spider richness were (±0.8) & 6.0 respectively. There were (on average) therefore not significantly different between 44% fewer fly species on frequently burnt sub- burnt and unburnt plots, although burnt plots plots. The number of morphospecies was similarly have (on average) slightly higher numbers of variable on unburnt (6–16) and burnt (0–12) sub- morphospecies (see Figure 3.18). The weak plots. position/treatment interaction effect identified for

214 Bushfire and forest invertebrates

spider abundance (see 3.3.2.8) was also evident for Mean values of beetle richness were spider species richness, largely due to higher therefore significantly different between burnt richness on burnt sub-plots at position 6. and unburnt plots, with burnt plots having (on average) lower numbers of morphospecies (see Figure 3.19). The strong position effect identified for beetle abundance (see 3.3.2.9) was not evident for beetle species richness, although there is a slight N-S decline evident on unburnt plots. The significant interaction reflects the reversal in the pattern of the treatment effect at position 3, where burnt sub-plots had slightly higher species richness (see Figure 3.19). 3.3.3.5 Ants Overall, 88 morphospecies of ants were collected, 70 on unburnt plots and 68 on frequently burnt Figure 3.18 Richness of Spiders on burnt and unburnt plots. Ant richness on sub-plots varied from plots 11–27, with a normal distribution of values (mean = 19.7 (±0.6), median = 19.5). Mean (±s.e.) & median values for unburnt 3.3.3.4 Beetles and burnt plots were 17.4 (±0.6) & 18.0 and 22.0 Overall, 139 morphospecies of beetles were (±0.6) & 22.0 respectively. There were (on collected, 86 on unburnt plots and 92 on average) 26% more ant species on frequently frequently burnt plots. Beetle richness on sub- burnt sub-plots. The number of morphospecies plots varied from 5–20, with a normal distribution was similarly variable on unburnt (11–24) and of values (mean = 11.3 (±0.5), median = 11.0). burnt (17–27) sub-plots. Mean (±s.e.) & median values for unburnt A 2-way ANOVA indicated a significant and burnt plots were 13.1 (±0.6) & 13.0 and 9.5 effect of treatment (F1,5 = 11.22 P = 0.020) but not (±0.5) & 9.0 respectively. There were (on average) position (F5,36 = 0.52 P = 0.757), with a significant 27% fewer beetle species on frequently burnt interaction detected between these two factors sub-plots. The number of morphospecies was (F5,36 = 2.88 P = 0.028). similarly variable on unburnt (8–20) and burnt Mean values of ant richness were therefore (5–15) sub-plots. significantly different between burnt and A 2-way ANOVA indicated a significant unburnt plots, with burnt plots having (on effect of treatment (F1,5 = 8.81 P = 0.031) but not average) higher numbers of morphospecies (see position (F5,36 = 2.24 P = 0.071), with a significant Figure 3.20). The strong position effect identified interaction detected between these two factors for ant abundance (see 3.3.2.10) was not evident (F5,36 = 2.86 P = 0.028). for ant species richness. The significant interaction reflects a lack of treatment effect at positions 1 and 4. 3.3.3.6 Scale Effects It is also important to recognise the effect that sampling intensity (scale) has on the detection of treatment effects for different taxa. With the bugs, results were consistent across a range of scales of measurement. The magnitude and direction of differences between unburnt and burnt species richness results for sub-plot (3.2 vs 3.7) and plot (7.5 vs 7.7) means and treatment totals (25 vs 26) were similar at these three scales (see Table 3.6). Similar patterns were apparent for flies, with a Figure 3.19 Richness of Beetles on burnt and unburnt similar magnitude of difference detected at the plots scale of sub-plot (10.1 vs 5.7), plot (25.8 vs 16.5) and treatment (66 vs 46).

215 Australia’s Biodiveristy - Responses to Fire

For ants, the magnitude of the difference in species richness detected between unburnt and burnt areas at the scale of sub-plot (17.4 vs 22.0) and plot (33.3 vs 38.7) were similar, although the magnitude of the difference was reduced at the scale of plot (compared with other taxa). The direction of the difference was however substantially reversed at the scale of treatment (70 vs 68). This would suggest a different situation to that with the spiders and beetles, with the species’ assemblages on burnt plots more similar (less diverse) than those on unburnt plots. Diversity Figure 3.20 Richness of Ants on burnt and unburnt plots on sub-plots within both unburnt and burnt plots would appear to be less similar than with For spiders however, while the magnitude other taxa, suggesting the differences lie at less and direction of species richness at the scale of than the scale of plot ( 1 hectare). These patterns sub-plot (4.4 vs 5.6) and plot (12.5 vs 15.8) were will be further explored in Section 3.3.4.5. similar, considerably more species were found 3.3.3.7 Summary overall on burnt compared to unburnt plots (47 Overall, 411 morphospecies were identified from vs 32). This would suggest that the species’ the five groups studied in detail. The beetles were assemblages on burnt plots are less similar (more the most species rich (139 species), followed by diverse) than those on unburnt plots, resulting in the ants (88), flies (77), spiders (63), and bugs (44). higher (between-habitat) diversity on frequently The results of analyses (ANOVA) burnt plots compared to unburnt plots. Diversity investigating the effects of frequent burning on sub-plots within both unburnt and burnt (treatment) and patterns due to large-scale spatial plots would appear to be similar, suggesting the effects (position) are summarised in Table 3.7, and differences lie at the scale of plot ( 1 hectare). indicate a variety of responses to frequent These patterns will be further explored in Section burning. Two groups, flies and beetles, 3.3.4.3. experienced a significant reduction in species For beetles, the magnitude of the difference richness on sub-plots following frequent burning detected between unburnt and burnt at the scale (44% and 27% reduction respectively). Two of sub-plot (13.1 vs 9.5) and plot (29.3 vs 27.0) were groups, the bugs and the spiders, showed an similar, however the direction was reversed at the increase in species richness on sub-plots (16% and scale of treatment (86 vs 92). This would suggest a 27% respectively), although these results were not similar situation as to that with the spiders, where statistically significant. The ants experienced a the species’ assemblages on burnt plots are less significant increase in sub-plot richness (26%) similar (more diverse) than those on unburnt plots. following repeated burning. Diversity on sub-plots within both unburnt and It was apparent that estimates of species burnt plots would appear to be similar, suggesting richness were influenced by the spatial scale of the differences lie at the scale of plot ( 1 hectare). measurement, with associated implications for the These patterns will be further explored in Section interpretation of observed treatment effects for the 3.3.4.4.

Table 3.6 Comparison of estimates of species richness at different scales of measurement. Taxa sub-plot (mean±s.e.) plot (mean±s.e.) treatment total unburnt burnt unburnt burnt unburnt burnt Bugs 3.2±0.3 3.7±0.3 7.5±0.5 7.7±0.9 25 26 Flies 10.1±0.6 5.7±0.8 25.8±0.6 16.5±3.6 66 46 Spiders 4.4±0.4 5.6±0.4 12.5±1.5 15.8±1.7 32 47 Beetles 13.1±0.6 9.5±0.5 29.3±0.8 27.0±3.1 86 92 Ants 17.4±0.6 22.0±0.6 33.3±1.3 38.7±1.1 70 68

216 Bushfire and forest invertebrates

different taxa. For bugs and flies results were These spatial patterns in estimates of species consistent across a range of scales of richness are a consequence of the composition of measurement, with the magnitude and direction invertebrate assemblages (communities) at the of differences between unburnt and burnt results varying scales of investigation. The nature of for sub-plot, plot and treatment similar. these patterns, and their interaction with For spiders, while the magnitude and environmental variables, will be further explored direction of species richness at the scale of sub- in Section 3.3.4. plot and plot were similar, considerably more 3.3.4 Community Composition species were found overall on burnt compared to unburnt plots. This suggested that species’ 3.3.4.1 Hemiptera assemblages on burnt plots were more diverse The 44 Hemipteran morphospecies were than those on unburnt plots, resulting in higher representative of 14 families, the infra-order (between-habitat) diversity. Diversity on sub-plots Dipsocoromorpha (not readily discernible to within both unburnt and burnt plots would family) and an unidentifiable Homopteran (see appear to be similar, suggesting the differences lie Table 3.8). The most diverse groups overall were at the scale of plot ( 1 hectare). the family Cicadellidae and the infra-order For beetles, the magnitude of the difference Dipsocoromorpha, containing 27% and 25% of detected between unburnt and burnt at the scale overall Hemipteran morphospecies respectively. of sub-plot and plot were similar, however the All three sub-orders were represented: the direction was reversed at the scale of treatment. Sternorrhyncha (5 morphospecies), the This would suggest a similar situation as to that Auchenorrhyncha (13 morphospecies) and the with the spiders, where the species’ assemblages Heteroptera (26 morphospecies). The on burnt plots are more diverse than those on Sternorrhyncha are mainly sedentary, often living unburnt plots. Diversity on sub-plots within both under waxy secretions or inside galls induced in the unburnt and burnt plots would appear to be host plants. Four morphospecies from this sub- similar, suggesting the differences lie at the scale order were found on unburnt plots and three on of plot (1 hectare). burnt plots (see Table 3.8). The Auchenorrhyncha For ants, the magnitude of the difference in (leaf- and plant-hoppers) are all plant-sap feeders as species richness detected between unburnt and adults and generally spend the bulk of their time on burnt areas at the scale of sub-plot and plot were plant foliage. Seven morphospecies from this sub- similar, although the magnitude of the difference order were found on unburnt plots and nine on was reduced at the scale of plot (compared with burnt plots. Most of the “true bugs”, the other taxa). The direction of the difference was Heteroptera, are plant-sap feeders, although some however substantially reversed at the scale of groups, such as the Reduviidae, are predatory. treatment, suggesting a different situation to that Thirteen morphospecies from this sub-order were with the spiders and beetles, with the species’ found on unburnt plots and fourteen on burnt assemblages on burnt plots less diverse. Diversity plots. Overall, a similar number of morphospecies on sub-plots within both unburnt and burnt plots were found on unburnt and burnt plots (24 and 26 would appear to be less similar than with other respectively) with similar proportions from the taxa, suggesting the differences lie at less than the three sub-orders. scale of plot ( 1 hectare). Although the two treatments are similarly diverse, both at the sub-ordinal and

Table 3.7 Changes in mean species richness following frequent burning for selected terrestrial invertebrate taxa. Taxa Common name Change with Statistically Large-scale frequent burning significant ? (P<0.05) spatial patterns ? Hemiptera bugs ⇑ 16% no no Diptera flies ⇓ 44% yes yes Araneae spiders ⇑ 27% no no* Coleoptera beetles ⇓ 27% yes no* Formicidae ants ⇑ 26% yes no* * indicates a statistically significant interaction between treatment and position effects.

217 Australia’s Biodiveristy - Responses to Fire

Table 3.8 Breakdown of Hemipteran morphospecies by sub-order and family. # Morphospecies Sub-order Family Common name unburnt burnt total Sternorrhyncha Coccidae scale insects 3 3 4 Homoptera* - 1 0 1 Auchenorrhyncha Cicadellidae leaf-hoppers 6 8 12 Fulgoroidea plant-hoppers 1 1 1 Heteroptera Dipsocoromorpha* - 8 4 11 Enicocephalidae - 1 0 1 Nabidae - 1 0 1 Tingidae lace bugs 1 0 1 Pentatomidae shield bugs 1 0 1 Lygaeidae seed bugs 1 1 2 Anthocoridae flower bugs 0 1 1 Miridae - 0 1 1 Thaumastocoridae - 0 1 1 Plataspidae - 0 1 1 Reduviidae assassin bugs 0 5 5 Totals 24 26 44 * Not readily discernible to Family level morphospecies level (see Section 3.3.3.1), only 7 sub-plot of a treatment, it is difficult to identify morphospecies (16%) were common to both clear patterns from these tabulated data (at the treatments (Group A - Table 3.9). When the species level). At this stage it is apparent however morphospecies are arranged to reflect their that unburnt plots have more species from the distribution across plots and treatments, it is clear infra-order Dipsocoromorpha (7 vs 3), while that there is a group of species (18) found only on burnt plots have greater numbers of species from unburnt sub-plots (41% - Group B) and a the family Reduviidae (5 vs 0). differenct group of species (19) found only on Although this table of relative abundance burnt sub-plots (43% - Group C). Frequent enables broad “assemblages” of species with burning would therefore appear to have resulted similar responses to disturbance to be identified, in the loss of up to 18 species of Hemipteran, these data are more clearly displayed in the form however the changed environment supports upto of a “bi-plot” derived from the CCA ordination 19 new species not found in unburnt areas. (see 2.5.3.3). This graphical display (Figure 3.21) Morphospecies found on both treatments shows the configuration of the environmental are representative of the families Cicadellidae variables, the scatter of sub-plots, and the (3 species), Coccidae (2 species), Fulgoroidea relationship between the two, giving an overview (1 species) and the infra-order Dipsocoromorpha of how community composition varies with the (1 species). Morphospecies found only on environment (Ter Braak 1986). unburnt sub-plots are representative of the A number of features are apparent from this families Cicadellidae (4 species), Coccidae bi-plot. Firstly, the minimal overlap of unburnt and (1 species), Nabidae (1 species), Enicocephalidae burnt sub-plots in ordination space reflects the (1 species), Lygaeidae (1 species), Pentatomidae largely dissimilar species assemblages of these two (1 species), Tingidae (1 species), the infra-order treatments. Secondly, the tighter clustering of burnt Dipsocoromorpha (7 species) and an unnamed sub-plots indicates a lower within-treatment Homopteran species. Morphospecies found only diversity compared to unburnt sub-plots (ie. a on burnt sub-plots are representative of the lower β-diversity). Richness values on burnt sub- families Cicadellidae (5 species), Coccidae plots were on average 16% higher than on unburnt (1 species), Miridae (1 species), Reduviidae sub-plots (see Section 3.3.3.1), however the high (5 species), Lygaeidae (1 species), Plataspidae similarity of sub-plots for this treatment mean that (1 species), Thaumastocoridae (1 species), the overall richness of unburnt and burnt plots was Anthocoridae (1 species) and the infra-order similar (25 vs 26 species). Unburnt sub-plots have Dipsocoromorpha (3 species). lower richness (α-diversity) but are less similar, As many of these families are represented resulting in higher “turnover” between sub-plots only by single individuals or found only on one (higher β-diversity). 218 Bushfire and forest invertebrates Table 3.9 Presence of Hemipteran morphospecies on burnt and unburnt plots Table

219 Australia’s Biodiveristy - Responses to Fire

The third feature concerns the contribution Brachycera (59 morphospecies). Nematocera adults of environmental variables to the differences in are generally slender with long legs, and have species composition for the two treatments. In the aquatic larvae (eg. mosquitoes, midges & sandflies) bi-plot (Figure 3.21) the length of the arrow or are gall makers. Twenty-five morphospecies from signifies the relative contribution of that variable this sub-order were found on unburnt plots and to species composition, and the direction signifies eighteen on burnt plots (see Table 3.10). its contribution to the differences between Brachycera adults are heavier set with relatively treatments. Unburnt sub-plots are characterised short legs (eg. House flies and March flies) and have by high levels of litter and high cover of tall and mainly terrestrial larvae (often found in damp soil very tall shrubs. Due to the correlation between and rotting vegetation). Forty-one morphospecies variables (see Section 3.2) these plots are also from this sub-order (Brachycera) were found on characterised by high top-soil moisture levels and unburnt plots and twenty-eight on burnt plots. low and spatially variable amounts of insolation at Overall, 30% less morphospecies were found on ground level. These environmental variables make burnt plots compared with unburnt plots (46 and the greatest contribution to the differences in 66 respectively) with similar trends for both sub- species composition between unburnt and burnt orders. plots. Burnt sub-plots are characterised by high The unburnt treatment was considerably levels of insolation at ground level, and to a lesser more diverse than the burnt treatment at the extent, steeper slopes, greater top-soil hardness morphospecies level (see Section 3.3.3.2) however 35 and greater cover of the herb & shrub component morphospecies (45%) were common to both of the understorey vegetation. Other treatments (Group A - Table 3.11). When the environmental variables make only a minor morphospecies are arranged to reflect their contribution to the observed differences in distribution across plots and treatments, it is clear Hemipteran species composition between burnt that there is a group of species (31) found only on and unburnt treatments. unburnt sub-plots (40% - Group B) and a different group of species (11) found only on burnt sub-plots (14% - Group C). Frequent burning would therefore appear to have resulted in the loss of up to 31 species of Diptera, however the changed environment supports up to 11 new species not found in unburnt areas. Morphospecies found on both treatments are representative of the families Phoridae (12 species), Cecidomyiidae and Sciaridae (5 species each), Chloropidae (3 species), Tachydromiinae, Ceratopogonidae, Chironomidae, Muscidae (2 species each), Dolichopodidae and Scatopsidae (1 species each). Morphospecies found only on Figure 3.21 Bi-plot from CCA ordination of Hemipteran unburnt sub-plots are representative of the families presence/absence data. (Ellipses represent 1 standard Tachydromiinae and Phoridae (4 species each), deviation unit around treatment centroids) Ceratopogonidae and Sciaridae (3 species each), Chloropidae, Drosophilidae, Muscidae, 3.3.4.2 Diptera and Tachinidae (2 species each), Syrphidaea, Therevidae, Micropezidae, Tipulidae, The 77 Dipteran (fly) morphospecies were Cecodomyiidae, Mycetophilidae and Chironomidae representative of 2 sub-orders and 20 families (see (1 species each). Morphospecies found only on Table 3.10). The most diverse groups overall were burnt sub-plots are representative of the families the families Phoridae and Sciaridae, containing Ceratopogonidae (2 species), Tachinidae, 22% and 10% of morphospecies respectively. The Piophilidae, Calliphoridae, Cecidomyiidae, families Cecodomyiidae, Ceratopogonidae and Chloropidae, Muscidae, Tachydromiinae and Empididae each contained 9% of overall Sphaeroceridae (1 species each). morphospecies. As many of these families are represented Both Australian sub-orders were represented: only by single individuals or found only on one the Nematocera (28 morphospecies) and the

220 Bushfire and forest invertebrates

Table 3.11 Presence of Dipteran morphospecies on burnt and unburnt plots.

221 Australia’s Biodiveristy - Responses to Fire

Table 3.10 Breakdown of Dipteran morphospecies by sub-order and family. # Morphospecies Sub-order Family Common name unburnt burnt total Nematocera Sciaridae - 8 5 8 Cecidomyiidae gall midges 6 6 7 Ceratopogonidae sand flies 5 4 7 Chironomidae midges 3 2 3 Scatopsidae - 1 1 1 Tipulidae crane flies 1 0 1 Mycetophilidae fungus gnats 1 0 1 Brachycera Phoridae - 16 13 17 Empididae* - 6 3 7 Chloropidae - 5 4 6 Muscidae bush flies 4 3 5 Sphaeroceridae - 2 1 3 Tachinidae - 2 1 3 Dolichopodidae - 1 1 1 Drosophilidae vinegar flies 2 0 2 Therevidae - 1 0 1 Syrphidae hover flies 1 0 1 Micropezidae stilt-legged flies 1 0 1 Piophilidae - 0 1 1 Calliphoridae blowflies 0 1 1 Totals 66 46 77 * all morphospecies from sub-family Tachydromiinae sub-plot of a treatment, it is difficult to identify clear patterns from these tabulated data (at the species level). At this stage it is apparent however that unburnt plots have more species from the families Sciaridae (8 vs 5), Phoridae (16 vs 13) and Empididae (6 vs 3). Although this table of relative abundance enables broad “assemblages” of species with similar responses to disturbance to be identified, these data are more clearly displayed in the form of a “bi-plot” derived from the CCA ordination (see 2.5.3.3). This graphical display (Figure 3.22) shows the configuration of the environmental variables, the scatter of sub-plots, and the Figure 3.22 Bi-plot from CCA ordination of Dipteran relationship between the two, giving an overview presence/absence data. (Ellipses represent 1 standard of how community composition varies with the deviation unit around treatment centroids) environment (Ter Braak 1986). A number of features are apparent from this were on average 44% lower than on unburnt sub- bi-plot. Firstly, the substantial overlap of unburnt plots (see Section 3.3.3.2), however the slightly and burnt sub-plots in ordination space reflects higher similarity of unburnt sub-plots means that, the relatively large number of morphospecies (35) overall, burnt plots had 30% less species than shared by the two treatments. Secondly, the loose unburnt plots (46 vs 66). Unburnt sub-plots have clustering of both unburnt and burnt sub-plots higher richness (α-diversity) and are more similar, indicates similar within-treatment diversity. Burnt resulting in lower “turnover” between sub-plots sub-plots are slightly more diverse (spread-out) (lower β-diversity). indicating a higher turnover (β-diversity) for this The third feature concerns the contribution treatment. Richness values on burnt sub-plots of environmental variables to the differences in

222 Bushfire and forest invertebrates

species composition for the two treatments. In the Gnaphosidae and Corinnidae, containing 14, 13 bi-plot (Figure 3.22) the length of the arrow and 13% of overall morphospecies respectively signifies the relative contribution of that variable (see Table 3.12). Most (62%) of the families were to species composition, and the direction signifies represented by only 1 or 2 species. its contribution to the differences between Overall, burnt plots had a greater number of treatments. Unburnt sub-plots are characterised morphospecies than unburnt plots (48 and 31 by high levels of litter and high cover of tall and respectively), with 16 morphospecies (25%) very tall shrubs. Due to the correlation between common to both treatments (Group A - Table variables (see Section 3.2) these plots are also 3.13). When the morphospecies are arranged to characterised by high top-soil moisture levels and reflect their distribution across plots and low and spatially variable amounts of insolation at treatments, it is clear that there is a group of ground level. Burnt sub-plots are characterised by species (15) found only on unburnt sub-plots high levels of insolation at ground level, and to a (24% - Group B) and a different group of species lesser extent, steeper slopes and greater exposure (32) found only on burnt sub-plots (51% - Group (more north-westerly aspects). The amount of C). Frequent burning would therefore appear to litter and the level of insolation at ground level have resulted in the loss of up to 15 species of make the greatest contribution to the differences spider, however the changed environment supports in species composition between unburnt and up to 32 new species not found in unburnt areas. burnt plots. Other environmental variables make Morphospecies found on both treatments only a minor contribution to the observed were representative of the families Corinnidae differences in Dipteran species composition (3 species), Textricellidae and Zodariidae (2 species between burnt and unburnt treatments. each), Hahniidae, Linyphiidae, Gnaphosidae, 3.3.4.3 Spiders Ctenidae, Micropholcommatidae, Oonopidae, Thomisidae, Toxopidae and Theridiidae (1 species The 63 spider morphospecies were representative each). Morphospecies found only on unburnt sub- of 21 families, with the most diverse groups plots are representative of the families Theridiidae, overall being the families Zodariidae,

Table 3.12 Breakdown of spider morphospecies by family. # Morphospecies Family Common name Ecological Information unburnt burnt total Zodariidae - terrestrial; under stones, logs & litter 2 9 9 Gnaphosidae - terrestrial; under stones, logs & litter 3 6 8 Corinnidae - generalised hunters & ant specialists 4 7 8 Salticidae jumping spiders terrestrial hunters on foliage, trees & logs. 3 3 6 Linyphiidae tent spiders build webs in foliage & near ground 2 4 5 Theridiidae - web builders; near ground level 3 2 4 Textricellidae - moist habitats; litter dwellers 2 3 3 Clubionidae ant-mimicking spiders nocturnal hunters; live in rolled-up leaves & in litter 1 1 2 Tekellidae - moist habitats; litter dwellers 1 1 2 Oonopidae - cryptic; occur under stones, logs & litter. 2 1 2 Ctenidae - terrestrial; vagrant hunters on ground & in litter 1 1 1 Hahniidae - build sheet webs in litter & foliage 1 1 1 Micropholcommatidae - moist habitats; litter dwellers 1 1 1 Thomisidae crab/flower spiders occur on foliage & bark 1 1 1 Toxopidae - moist habitats 1 1 1 Lycosidae wolf spiders terrestrial; ground hunters 0 3 3 Prododomidae - dry habitats 0 1 1 Heteropodidae - bark, foliage & litter dwellers 0 1 1 Dictynoidea - sheet web builders on ground, litter & bark 0 1 1 Malkaridae - moist habitats; litter dwellers 2 0 2 Stiphidiidae - sheet web builders 1 0 1 TOTAL 31 48 63

223 Australia’s Biodiveristy - Responses to Fire Table 3.13 Presence of Spider morphospecies on burnt and unburnt plots. Table

224 Bushfire and forest invertebrates

Salticidae, Gnaphosidae and Malkaridae (2 species A number of features are apparent from this each), Oonopidae, Tekellidae, Stiphidiidae, bi-plot. Firstly, the minimal overlap of unburnt and Linyphiidae, Clubionidae and Corinnidae burnt sub-plots in ordination space reflects the (1 species each). Morphospecies found only on largely dissimilar species assemblages of these two burnt sub-plots are representative of the families treatments. Secondly, the tighter clustering of Zodariidae (7 species), Gnaphosidae (5 species), unburnt sub-plots indicates a lower within- Corinnidae (4 species), Salticidae and Lycosidae treatment diversity compared to burnt sub-plots. (3 species each), Linyphiidae (2 species) and Burnt sub-plots are slightly more diverse (spread- Tekellidae, Textricellidae, Theridiidae, out) indicating a higher turnover (β-diversity) for Clubionidae, Dictynoidea, Heteropodidae and this treatment. Richness values on burnt sub-plots Prododomidae (1 species each). were on average were 27% higher than on unburnt As many of these families are represented sub-plots (see Section 3.3.3.3), however the greater only by single individuals or found only on one dissimilarity of burnt sub-plots means that, overall, sub-plot of a treatment, it is difficult to identify burnt plots had 55% more species than unburnt clear patterns from these tabulated data (at the plots. Burnt sub-plots have higher richness (α- species level). At this stage it is apparent however diversity) and are less similar, resulting in higher that unburnt plots have more species from the “turnover” between sub-plots (higher β-diversity). family Malkaridae (2 vs 0), while burnt plots have The third feature concerns the contribution greater numbers of species from the families of environmental variables to the differences in Zodariidae (9 vs 2), Gnaphosidae (6 vs 3), species composition for the two treatments. In the Corinnidae (7 vs 4), Linyphiidae (4 vs 2) and bi-plot (Figure 3.23) the length of the arrow Lycosidae (3 vs 0). signifies the relative contribution of that variable to Although this table of relative abundance species composition, and the direction signifies its enables broad “assemblages” of species with contribution to the differences between treatments. similar responses to disturbance to be identified, Unburnt sub-plots are characterised by high levels these data are more clearly displayed in the form of litter and high cover of tall and very tall shrubs. of a “bi-plot” derived from the CCA ordination Due to the correlation between variables (see (see 2.5.3.3). This graphical display (Figure 3.23) Section 3.2) these plots are also characterised by shows the configuration of the environmental high top-soil moisture levels and low and spatially variables, the scatter of sub-plots, and the variable amounts of insolation at ground level. relationship between the two, giving an overview Burnt sub-plots are characterised by high levels of of how community composition varies with the insolation at ground level, and to a lesser extent, environment (Ter Braak 1986). steeper slopes and greater cover of the herb & shrub component of the understorey vegetation. The amounts of insolation and litter make the greatest contribution to the differences in species composition between unburnt and burnt plots. Other environmental variables make only a minor contribution to the observed differences in spider species composition. 3.3.4.4 Beetles The 139 beetle morphospecies were representative of nine super-families and 25 families (see Table 3.14). The most diverse groups were the families Staphylinidae and Curculionidae, containing 22% and 17% of overall morphospecies respectively. Nine super-families were represented: the Figure 3.23 Bi-plot from CCA ordination of spider Staphylinoidea (56 morphospecies), the presence/absence data. (Ellipses represent 1 standard Curculionoidea (23 morphospecies), Caraboidea, deviation unit around treatment centroids) Scarabaeoidea, Cucujoidea (12 morphospecies each), Chrysomeloidea (10 morphospecies), Tenebrionoidea (9 morphospecies), Bostrichoidea

225 Australia’s Biodiveristy - Responses to Fire

Table 3.14 Breakdown of Beetle morphospecies by super-family and family. # Morphospecies Super-family Family Common name unburnt burnt total Staphylinoidea Staphylinidae Rove beetles 21 20 31 Pselaphidae - 6 7 11 Scydmaenidae - 5 7 8 Ptiliidae - 3 1 3 Leiodidae - 3 1 3 Scarabaeoidea Scarabaeidae - 6 7 9 Trogidae - 1 0 1 Hybosoridae - 1 1 1 Passalidae - 0 1 1 Elateroidea Elateridae Click Beetles 1 1 2 Cucujoidea Nitidulidae - 2 1 3 Corylophidae - 1 2 3 Endomychidae - 1 1 2 Lathridiidae - 2 1 2 Silvanidae - 0 1 1 Phalacridae - 0 1 1 Chrysomeloidea Chrysomelidae Leaf Beetles 4 8 9 Cerambycidae Longicorn Beetles 0 1 1 Curculionoidea Curculionidae Weevils 9 18 23 Caraboidea Carabidae Ground Beetles 11 8 12 Bostrichoidea Anobiidae - 2 2 3 Tenebrionoidea Tenebrionidae Darkling Beetles 3 2 5 Aderidae - 2 0 2 Oedemeridae - 1 0 1 Anthicidae - 1 0 1 Totals 86 92 139

(3 morphospecies), and the Elateroidea on unburnt plots and nine on burnt plots. Many (2 morphospecies). Tenebrionoidea are scavengers inhabiting the litter Most Staphylinoidea are general predators, layer. Seven morphospecies were found on with some feeding on decomposing fruits. Thirty- unburnt plots and two on burnt plots. The eight morphospecies from this super-family were Bostrichoidea were represented by one family, the found on unburnt plots and thirty-six on burnt Anobiidae, which feed on a variety of plant & plots (see Table 3.14). The Curculionoidea were animal products. Three morphospecies from this represented by one family, Curculionidae (the family were each found, two on unburnt and two Weevils), with nine morphospecies found on on burnt plots. The Elateroidea were represented unburnt plots and eighteen on burnt plots. The by a single family, the Elateridae (Click Beetles), Caraboidea were represented by one family, which are mainly predatory. Two morphospecies Caraboidea (the Ground Beetles), which are mainly from this family were found, one on unburnt and predatory on plant-inhabiting insects. Eleven one on burnt plots (see Table 3.14). morphospecies from this family were found on Overall, slightly more morphospecies were unburnt plots and eight on burnt plots. The found on burnt plots compared with unburnt Scarabaeoidea commonly feed on eucalypt leaves as plots (92 and 86 respectively), although on average, adults. Eight morphospecies from this super-family burnt sub-plots had 27% fewer species (see Section were found on unburnt plots and nine on burnt 3.3.3.4). Thirty-nine morphospecies (28%) were plots. Many Cucujoidea are known to feed on fungi common to both treatments (Group A - Table growing on leaf surfaces. Six morphospecies from 3.15). When the morphospecies are arranged to this super-family were found on unburnt plots and reflect their distribution across plots and seven on burnt plots. The Chrysomeloidea feed on treatments, it is clear that there is a group of species leaves and other vegetative parts of plants, both as (47) found only on unburnt sub-plots (34% - larvae and adults. Four morphospecies were found Group B) and a different group of species (53)

226 Bushfire and forest invertebrates

found only on burnt sub-plots (38% - Group C). sub-plots suggests a lower within-treatment Frequent burning would therefore appear to have diversity compared to unburnt sub-plots (ie. a resulted in the loss of up to 47 species of Coleoptera, lower β-diversity). Richness values on burnt sub- however the changed environment supports up to 53 plots were, on average, 27% lower than on unburnt new species not found in unburnt areas. sub-plots (see Section 3.3.3.4), with the apparently Morphospecies found on both treatments are high similarity of sub-plots for this treatment representative of the families Staphylinidae meaning that the overall richness of burnt plots (10 species), Carabidae (7 species), Scarabaeidae, should be much lower than for unburnt plots. In Scydmaenidae, Curculionidae (4 species each), fact, overall, burnt plots had slightly more species Chrysomelidae (3 species), Pselaphidae (2 species), than unburnt plots (92 vs 86). This clustering is Ptiliidae, Hybosoridae, Lathridiidae, Anobiidae and therefore an artefact of the two-dimensional Leiodidae (1 species each). Morphospecies found representation of the data, with an inspection of only on unburnt sub-plots are representative of the Table 3.15 revealing a greater number of species families Staphylinidae (11 species), Curculionidae unique to burnt plots (53 vs 47). While unburnt (5 species), Carabidae, Pselaphidae (4 species each), sub-plots have higher richness (α-diversity), burnt Tenebrionidae (3 species), Aderidae, Scarabaeidae, plots are less similar, resulting in higher “turnover” Ptiliidae and Nitidulidae (2 species each), between sub-plots (higher β-diversity). Scydmaenidae, Chrysomelidae, Lathridiidae, The third feature concerns the contribution Anobiidae, Leiodidae, Trogidae, Elateridae, of environmental variables to the differences in Corylophidae, Endomychidae, Oedemeridae and species composition for the two treatments. In the Anthicidae (1 species each). Morphospecies found bi-plot (Figure 3.24) the length of the arrow only on burnt sub-plots are representative of the signifies the relative contribution of that variable to families Curculionidae (14 species), Staphylinidae species composition, and the direction signifies its (10 species) Chrysomelidae, Pselaphidae (5 species contribution to the differences between treatments. each), Scarabaeidae, Scydmaenidae (3 species each), Unburnt sub-plots are characterised by high levels Corylophidae, Tenebrionidae, (2 species each), of litter and high cover of tall and very tall shrubs. Passalidae, Elateridae, Nitidulidae, Endomychidae, Due to the correlation between variables (see Silvanidae, Phalacridae, Cerambycidae, Carabidae Section 3.2) these plots are also characterised by and Anobiidae (1 species each). high top-soil moisture levels and low and spatially As many of these families are represented variable amounts of insolation at ground level. only by single individuals or found only on one Burnt sub-plots are characterised by high levels of sub-plot of a treatment, it is difficult to identify insolation at ground level, and to a lesser extent, clear patterns from these tabulated data (at the steeper slopes. These environmental variables make species level). At this stage it is apparent however the greatest contribution to the differences in that unburnt plots have more species from the species composition between unburnt and burnt family Carabidae (11 vs 8), while burnt plots have plots, with other environmental variables make greater numbers of species from the families only a minor contribution to the observed Curculionidae (18 vs 9) and Chrysomelidae (8 vs 4). differences in beetle species composition. Although this table of relative abundance enables broad “assemblages” of species with similar responses to disturbance to be identified, these data are more clearly displayed in the form of a “bi-plot” derived from the CCA ordination (see 2.5.3.3). This graphical display (Figure 3.24) shows the configuration of the environmental variables, the scatter of sub-plots, and the relationship between the two, giving an overview of how community composition varies with the environment (Ter Braak 1986). A number of features are apparent from this bi-plot. Firstly, the small overlap of unburnt and burnt sub-plots in ordination space reflects the Figure 3.24 Bi-plot from CCA ordination of Beetle largely dissimilar species assemblages of these two presence/absence data. (Ellipses represent 1 standard treatments. Secondly, the tighter clustering of burnt deviation unit around treatment centroids)

227 Australia’s Biodiveristy - Responses to Fire Table 3.15. Presence of Beetle morphospecies on burnt and unburnt plots. Table

228 Bushfire and forest invertebrates

to be scanned

229 Australia’s Biodiveristy - Responses to Fire

3.3.4.5 Ants plots (70 and 68 respectively) with the sub-families The 88 ant morphospecies were representative of generally equally represented at the morphospecies 5 sub-families and 34 genera (see Table 3.16). The level: Myrmeciinae 3 & 4, 18 & 20, most diverse groups overall were the genera Dolichoderinae 7 & 9, and 26 & 22 Pheidole, and Camponotus, each respectively. The exception was the Ponerinae, with containing 8% of overall morphospecies. 22 morphospecies on unburnt plots and 11 on Five sub-families were represented: Ponerinae burnt plots (see Table 3.16). (11 genera, 19 morphospecies), Myrmeciinae (1 A substantial proportion (50) of genus, 5 morphospecies), Myrmicinae (11 genera, morphospecies (57%) were common to both 24 morphospecies), Dolichoderinae (4 genera, 10 treatments (Group A - Table 3.17). When the morphospecies) and Formicinae (7 genera, 30 morphospecies are arranged to reflect their morphospecies). Overall, a similar number of distribution across plots and treatments, it is clear morphospecies were found on unburnt and burnt that there is a group of species (20) found only on

Table 3.16 Breakdown of Ant morphospecies by sub-family and genus. # Morphospecies Sub-family Genus Ecological role* unburnt burnt total

Ponerinae Rhytidoponera opportunist 4 4 4 Bothroponera solitary forager 1 1 1 Trachymesopus cryptic in soil/litter 1 1 1 Cerapachys climate specialist 2 1 2 Hypoponera cryptic in soil/litter 4 1 4 Heteroponera cryptic in soil/litter 2 1 2 Amblyopone cryptic in soil/litter 1 0 1 Discothyrea cryptic in soil/litter 1 0 1 Ponera cryptic in soil/litter 1 0 1 Leptogenys specialist predator 0 1 1 Sphinctomyrmex cryptic in soil/litter 0 1 1 Myrmeciinae Myrmecia solitary forager 3 4 5 Myrmicinae Pheidole generalist 5 6 7 Solenopsis cryptic in soil/litter 3 2 3 Crematogaster generalist 2 2 2 Strumigenys cryptic in soil/litter 1 1 1 Mayriella opportunist 1 1 1 Meranoplus climate specialist 1 1 1 Tetramorium opportunist 2 1 2 Epopostruma specialist predator 0 1 1 Podomyrma climate specialist 1 2 2 Monomorium generalist 1 2 2 Colobostruma solitary forager 0 2 2 Dolichoderinae Iridomyrmex dominant 5 7 7 Tapinoma cryptic in soil/litter 1 1 1 Technomyrmex opportunist 1 1 1 Leptomyrmex dominant 0 1 1 Formicinae Camponotus sub-dominant 6 6 7 Paratrechina opportunist 4 3 4 Polyrhachis sub-dominant 5 3 6 Melophorus climate specialist 5 3 5 Prolasius climate specialist 3 2 3 Stigmacros cryptic in soil/litter 2 3 3 Notoncus climate specialist 1 2 2 Totals 70 68 88

* ecological functional groups as defined by Andersen (1990)

230 Bushfire and forest invertebrates

unburnt sub-plots (23% - Group B) and a different Morphospecies found on both treatments are group of species (18) found only on burnt sub-plots representative of the genera: Camponotus, (20% - Group C). Frequent burning would Iridomyrmex (5 species each), Rhytidoponera, Pheidole therefore appear to have resulted in the loss of up to (4 species each), Paratrechina, Melophorus (3 species 20 species of ants, however the changed each), Myrmecia, Solenopsis, Crematogaster, Prolasius, environment supports up to 18 new species not Polyrhachis, Stigmacros (2 species each), Bothroponera, found in unburnt areas. Trachymesopus, Cerapachys, Hypoponera, Heteroponera, Table 3.17. Presence of Ant morphospecies on burnt and unburnt plots.

231 Australia’s Biodiveristy - Responses to Fire

Strumigenys, Mayriella, Meranoplus, Tetramorium, Podomyrma, Monomorium, Tapinoma, Technomyrmex and Notoncus (1 species each). Morphospecies found only on unburnt sub-plots are representative of the genera: Polyrhachis, Hypoponera (3 species each), Melophorus (2 species), Cerapachys, Heteroponera, Amblyopone, Discothyrea, Ponera, Myrmecia, Pheidole, Solenopsis, Tetramorium, Camponotus, Paratrechina and Prolasius (1 species each). Morphospecies found only on burnt sub-plots are representative of the genera: Iridomyrmex, Myrmecia, Pheidole, Colobostruma (2 species each), Rhytidoponera, Leptogenys, Sphinctomyrmex, Epopostruma, Figure 3.25 Bi-plot from CCA ordination of Ant relative Podomyrma, Monomorium, Leptomyrmex, Camponotus, abundance data. (Ellipses represent 1 standard Polyrhachis, Stigmacros and Notoncus (1 species each). deviation unit around treatment centroids) As many of these families are represented only by single individuals or found only on one indicates a lower within-treatment diversity sub-plot of a treatment, it is difficult to identify compared to unburnt sub-plots (ie. a lower β- clear patterns from these tabulated data (at the diversity). Richness values on burnt sub-plots were individual species level). At this stage it is apparent on average 26% higher than on unburnt sub-plots however that unburnt plots have more species (see Section 3.3.3.5), however the high similarity of from the genera Cerapachys (7 vs 1) and Hypoponera sub-plots for this treatment means that the overall (4 vs 1), while burnt plots have greater numbers richness of unburnt and burnt plots was similar (70 of species from the genus Colobostruma (2 vs 0). vs 68 species). Unburnt sub-plots have lower Although this table of relative abundance richness (α-diversity) but are less similar, resulting enables broad “assemblages” of species with in higher “turnover” between sub-plots (higher β- similar responses to disturbance to be identified, diversity). these data are more clearly displayed in the form The third feature concerns the contribution of a “bi-plot” derived from the CCA ordination of environmental variables to the differences in (see 2.5.3.3). This graphical display (Figure 3.25) species composition for the two treatments. In the shows the configuration of the environmental bi-plot (Figure 3.25) the length of the arrow variables, the scatter of sub-plots, and the signifies the relative contribution of that variable to relationship between the two, giving an overview species composition, and the direction signifies its of how community composition varies with the contribution to the differences between treatments. environment (Ter Braak 1986). Unburnt sub-plots are characterised by high A number of features are apparent from this levels of litter and high cover of tall and very tall bi-plot. Firstly, the lack of any overlap of unburnt shrubs. Due to the correlation between variables and burnt sub-plots in ordination space reflects the (see Section 3.2) these plots are also characterised largely dissimilar species assemblages of these two by high top-soil moisture levels and low and treatments. Although 57% of morphospecies were spatially variable amounts of insolation at ground found on both treatments, their relative abundance level. Burnt sub-plots are characterised by high on each treatment differs substantially, leading to levels of insolation at ground level, and to a lesser fundamentally different species assemblages. extent, steeper slopes, greater cover of the herb & Secondly, the tighter clustering of burnt sub-plots shrub component of the understorey vegetation,

Table 3.18 Distribution of invertebrate morphospecies richness by treatment Invertebrate both unburnt only burnt only total Taxa # % # % # % # Hemiptera (bugs) 7 16 18 41 19 43 44 Diptera (flies) 35 45 31 40 11 15 77 Araneae (spiders) 16 25 15 24 32 51 63 Coleoptera (beetles) 39 28 47 34 53 38 139 Formicidae (ants) 50 57 20 23 18 20 88 Total 147 - 131 - 133 - 411

232 Bushfire and forest invertebrates

greater top-soil hardness and greater spatial biodiversity of frequently burnt areas was patchiness of twigs (0–10mm) and very tall shrubs. maintained by the addition of species not The level of insolation at ground level, the recorded on unburnt plots. The changed amount of litter and the cover of tall shrubs make the environment was supporting an additional 133 greatest contribution to the differences in species morphospecies (19 bugs, 11 flies, 32 spiders, 53 composition between unburnt and burnt plots. beetles and 18 ants). Other environmental variables make only a minor Many of the morphospecies apparently lost contribution to the observed differences in ant species. from frequently burnt sites were however only 3.3.4.6. Summary detected on a single sub-plot or represented by a single individual on unburnt plots. These could be The five groups studies in detail proved to be genuinely rare or uncommon species which were extremely diverse. Beetles had the richness fauna missed purely by chance when sampling burnt overall with 139 beetle morphospecies plots. For this reason it is difficult to identify clear representative of nine super-families and 25 patterns (at the species level) from the relative families. The ants were the second richness group abundance data alone. Some general trends were with 88 morphospecies representative of 5 sub- apparent however when morphospecies data were families and 34 genera. They were followed by the grouped into a higher taxonomic level. For bugs, flies with 77 morphospecies representative of 2 unburnt plots had more species from the infra- sub-orders and 20 families, the spiders with 63 order Dipsocoromorpha (7 vs 3), while burnt plots morphospecies representative of 21 families, and have greater numbers of species from the family the bugs with 44 morphospecies representative of Reduviidae (5 vs 0). For flies, unburnt plots had 16 family (or similar) groups. more species from the families Sciaridae (8 vs 5), Overall, the same number of morphospecies Phoridae (16 vs 13) and Empididae (6 vs 3). For (279) were collected from unburnt and burnt spiders, unburnt plots had more species from the plots. Richness on unburnt sub-plots was, on family Malkaridae (2 vs 0), while burnt plots had average, 48.2 morphospecies, which was similar to greater numbers of species from the families the average value on burnt sub-plots (46.5 Zodariidae (9 vs 2), Gnaphosidae (6 vs 3), morphospecies). This initially suggests that Corinnidae (7 vs 4), Linyphiidae (4 vs 2) and frequent burning had not reduced biodiversity in Lycosidae (3 vs 0). For beetles, unburnt plots had this forest environment. An analysis of the more species from the family Carabidae (11 vs 8), richness of individual groups (see Section 3.3.3) while burnt plots had greater numbers of species has shown this not to be the case, with groups from the families Curculionidae (18 vs 9) and responding differently to frequent burning. This Chrysomelidae (8 vs 4). For ants, unburnt plots Section (3.3.4) examined the nature of that had more species from the genera Cerapachys (7 vs response by looking at the composition of faunal 1) and Hypoponera (4 vs 1), while burnt plots had assemblages (communities). greater numbers of species from the genus An examination of the distribution of Colobostruma (2 vs 0). The implication of these morphospecies across sub-plots for each changes for community organisation and treatment, detected a consistent pattern. ecosystem function are considered in Section 3.4. Morphospecies fell into one of three groups: found Although an examination of relative abundance on both treatments (Group A), found only on patterns enables broad “assemblages” of species with unburnt plots (Group B), or found only on burnt similar responses to disturbance to be identified, plots (Group C). The relative proportions of these data are more clearly displayed and interpreted morphospecies in each category however varied in the form of a “bi-plot” derived from the CCA substantially between taxonomic groups. For ordination. These graphical displays show the Hemiptera (bugs) the proportions were 16, 41 and configuration of the environmental variables, the 43% for both, unburnt and burnt respectively; for scatter of sub-plots, and the relationship between the Diptera (flies) 45, 40 and 15%; for spiders 25, 24 two, giving an overview of how community and 51%; for beetles 28, 34 and 38%; and for ants composition varies with the environment. An 57, 23 and 20% (see Table 3.18). examination of the bi-plots revealed a number of These results also suggest that frequent consistent features. Firstly, the degree of overlap of burning had led to the loss of up to 131 species unburnt and burnt sub-plots in ordination space (18 bugs, 31 flies, 15 spiders, 47 beetles and 20 reflected the similarity (or dissimilarity) of the ants), which represents 47% of the morphospecies species’ assemblages of the two treatments. For bugs, known from the unburnt areas. Overall 233 Australia’s Biodiveristy - Responses to Fire

spiders, beetles and ants there was little or no level and, to a lesser extent, steeper slopes. These overlap, indicating low similarity of the two environmental variables make the greatest assemblages. For flies however there was a contribution to the differences in species composition substantial overlap, reflecting the relatively large between unburnt and burnt plots. For bugs, burnt number of morphospecies shared by the two sub-plots were also characterised by greater top-soil treatments. hardness and greater cover of the herb & shrub Secondly, the degree of clustering of the sub- component of the understorey vegetation, with herb plots from each treatment indicates the relative and shrub cover also important for spiders and ants. similarity of species’ assemblages on sub-plots and For flies, burnt sites were also characterised by plots within each treatment. The tighter clustering greater exposure (more north-westerly aspects). of burnt sub-plots for bugs, beetles and ants Other environmental variables made only a minor indicated a lower within-treatment diversity contribution to the observed differences in species compared to unburnt sub-plots (ie. a lower β- composition between burnt and unburnt treatments diversity). The converse applied for spiders, with for all taxa. Communities were therefore influenced the tighter clustering of unburnt sub-plots by a combination of site-dependent (slope and aspect) indicating a lower β-diversity compared to burnt and treatment-dependent (litter, insolation, herb & sub-plots. The situation for flies indicated similar shrub cover, top-soil moisture & hardness) within-treatment diversity for both treatments, with environmental variables. loose clustering of both unburnt and burnt sub- plots. This interaction between point richness 3.3.5 Community Structure (α-diversity) and spatial “turnover” of species While it is possible to describe and assess (β-diversity) has substantial implications for the communities using indices such as species interpretation of the apparent effect of repeated richness, or to compare the relative abundance of burning and were previously identified as “scale species using similarity indices, multi-variate effects” in Section 3.3.3.6. Similar patterns were approaches and/or through graphical apparent for bugs and ants where richness values representation, these contribute little to an (α-diversity) on burnt sub-plots were on average understanding of the processes underlying their higher than on unburnt sub-plots, however the differences. In order to simplify and interpret the high similarity of assemblages on burnt sub-plots complexity of ecological systems, one approach (low β-diversity) meant that the overall richness of has been to group species into “guilds” or both treatments were similar. Unburnt sub-plots “functional groups”. These groups recognise the had lower richness (α-diversity) but are less similar, ecological rather than the taxonomic affinity of resulting in higher “turnover” between sub-plots species. In this Section morphospecies were (higher β-diversity), increasing overall species allocated to guilds following reference to the richness for that treatment. For flies and beetles relevant literature and discussions with taxonomic burnt plots had lower richness (α-diversity) but experts. To maintain comparability between broad were less similar, resulting in higher β-diversity. taxonomic groups, the number of species from Spiders exhibited a different pattern with both each group was averaged across the 24 sub-plots higher α- and β-diversity for the burnt treatment, for each treatment. This method does not take resulting in a large number of species (32) unique to into account the relative abundance of individuals frequently burnt sites. of species and therefore is not unfairly biased by a The third feature of the bi-plots concerns the few very abundant species. contribution of environmental variables to the 3.3.5.1 Hemiptera differences in species composition for the two Morphospecies were classified into one of 5 groups treatments. In the bi-plots (Figures 3.21-5) the length based on the known habits and requirements at the of the arrows signify the relative contribution of that family level (see Table 3.19). These groups were variable to species composition, and the direction primarily based on feeding strategy, and secondarily signifies their contribution to the differences between habitat preferences. The groups were; primarily treatments. For all taxa unburnt sub-plots were phytophagous (Cicadellidae, Coccidae, Fulgoridae characterised by high levels of litter, high cover of tall and the unidentified Homopteran), moist habitat and very tall shrubs, high top-soil moisture levels and specialists (Dipsocoromorpha), primarily predacious low and spatially variable amounts of insolation at (Nabidae, Reduvidae) and others. This last category ground level. Similarly, burnt sub-plots were included morphospecies from the families characterised by high levels of insolation at ground Enicocephalidae, Anthocoridae, Miridae, Tingidae, 234 Bushfire and forest invertebrates

Table 3.19 Comparison of Hemipteran community structure on burnt and unburnt plots. Data represent mean percentage of morphospecies in each category. Ecological role Taxa Unburnt Burnt Primarily phytophagous Cicadellidae 48.3 75.7 Coccidae, Fulgoridae, & Homoptera 15.4 3.3 Moist habitat specialists Dipsocoromorpha 22.9 3.9 Primarily predacious Nabidae 11.7 0.0 Reduvidae 0.0 10.1 Others (see text) 1.7 7.0 Total 100 100

Thaumastocoridae, Lygaeidae, Plataspididae and family Reduviidae. The greatest change concerns Pentatomidae which were only found on one sub- habitat specialists from the sub-order plot and therefore offering limited information to Dipsocoromorpha where there has been, on this analysis. average, an 83% reduction in the number of On average, 64% of morphospecies on morphospecies. unburnt plots were primarily phytophagous (plant 3.3.5.2 Diptera eating) and 12% predacious on other invertebrates. Morphospecies were classified into one of 7 groups Of the remaining 24%, approximately 23% are based on the known habits and requirements at the known to be moist habitat specialists from the sub- family level (see Table 3.20). These groups were order Dipsocoromorpha utilising a variety of primarily based on feeding strategy, and feeding strategies. On average, 79% of secondarily habitat preferences. The groups were; morphospecies on burnt plots were primarily primarily phytophagous (Cecidomyiidae), phytophagous and 10% predacious on other primarily predacious (Tachydromiinae), fungal invertebrates. Of the remaining 11%, approximately feeders (Sciaridae, Drosophilidae, Mycetophilidae 4% are known to be moist habitat specialists from & Scatopsidae), generalists and scavengers the sub-order Dipsocoromorpha. The remaining (Phoridae & Chloropidae), moist habitat specialists 7% of morphospecies on burnt plots were from the (Ceratopogonidae & Chironomidae), litter families Anthocoridae, Miridae, Thaumastocoridae, dwellers (Sphaeicieiidae & Tipilidae) and wide- Lygaeidae, Plataspidae and Pentatomidae (1 species ranging “tourists” (Piophilidae, Micropezidae, each). Dolichopodidae, Calliphoridae, Syrphidae, A graphical comparison of these data (Figure Thereuidae, Muscidae & Tachinidae). 3.26) indicates that frequent burning has resulted On average, 29% of morphospecies on in a fundamental shift in Hemipteran community unburnt plots were primarily generalists and structure. With regard to feeding strategy, there scavengers, 23% feeders on fungal products, 17% has been, on average, a 15% increase in the moist habitat specialists, and 11% phytophagous number of phytophagous species. While the (plant eating). Of the remaining 20%, number of predacious species has remained approximately 13% are known to be wide-ranging largely unchanged, there has been a total shift “tourists”, 6% predators, and 2% litter dwellers. from species from the family Nabidae to the On average, 26% of morphospecies on unburnt plots were primarily generalists and scavengers, 13% feeders on fungal products, 18% moist habitat specialists, and 26% phytophagous (plant eating). Of the remaining morphospecies, approximately 7% are known to be wide-ranging “tourists”, 6% predators, and 1% litter dwellers. A graphical comparison of these data (Figure 3.27) indicates that frequent burning has resulted in a fundamental shift in Dipteran community structure. With regard to feeding strategy, there has been, on average, a 44% decrease in the number of morphospecies feeding on fungal Figure 3.26 Comparison of Hemipteran community products. This change was most marked in the structure on burnt and unburnt plots families Scaridae and Scatopsidae. The proportion

235 Australia’s Biodiveristy - Responses to Fire

Table 3.20 Comparison of Fly community structure on burnt and unburnt plots. Data represent mean percentage of morphospecies in each category. Ecological role Taxa Unburnt Burnt Fungal feeders Sciaridae, Drosophilidae, Mycetophilidae & Scatopsidae 23.2 12.9 Litter dwellers Sphaeicieiidae & Tipilidae 2.0 0.8 Tourists Piophilidae, Micropezidae, Dolichopodidae, Calliphoridae, Syrphidae, Thereuidae, Muscidae & Tachinidae 12.6 7.2 Moist habitat specialists Ceratopogonidae & Chironomidae 17.0 17.8 Generalists/scavengers Phoridae & Chloropidae 29.1 25.9 Predators Tachydromiinae 5.8 5.8 Phytophagous Cecidomyiidae 10.3 25.8 Total 100 100

With regard to micro-habitat preferences, the number of morphospecies specifically utilising the litter layer has, on average, decreased by 60%. This was primarily due to the absence of the family Tipulidae on frequently burnt sub-plots. The number of moist habitat specialists remained similar, although the family Ceratopogonidae was more commonly represented on burnt sub-plots and the family Chironomidae on unburnt sub-plots. 3.3.5.3 Spiders Morphospecies were classified into one of 5 groups Figure 3.27 Comparison of Fly community structure on based on the known habits and requirements at the burnt and unburnt plots family level (see Table 3.21). As all spiders are fundamentally predacious, these groups were of predacious species on sub-plots has remained primarily based on habitat preferences. The groups similar, however the number of phytophagous were; moist habitat specialists (Theridiidae, species has, on average, increased by 140%. These Toxopidae, Oonopidae, Malkaridae & Tekellidae), additional species were from the family those with a known preference for dry habitats Cecidomyiidae. The number of generalists and (Gnaphosidae & Corinnidae), litter dwellers scavengers has decreased slightly (11%). The (Hahniidae & Textricellidae), open & disturbed number of morphospecies regarded as wide- habitat specialists (Linyphiidae & Zodariidae) and ranging “tourists” was, on average, reduced by others. This last category included generalised 43% on frequently burnt plots. hunters from the families Ctenidae, Dictynoidea, Heteropodidae, Micropholcommatidae,

Table 3.21 Comparison of Spider community structure on burnt and unburnt plots. Data represent mean percentage of morphospecies in each category. Ecological role Family Unburnt Burnt Theridiidae 22.2 1.8 Moist habitat Toxopidae 11.4 1.2 specialists Oonopidae 8.1 0.4 Malkaridae 4.0 0.0 Tekellidae 4.3 2.2 Dry habitat Gnaphosidae 5.8 8.1 preference Corinnidae 14.6 18.4 Litter Hahniidae 4.5 10.4 dwellers Textricellidae 5.5 10.3 Open/disturbed Linyphiidae 3.1 15.7 habitat specialists Zodariidae 2.1 20.1 Others (see text) 14.3 11.3 Total 100 100

236 Bushfire and forest invertebrates

Stiphidiidae, Clubionidae, Lycosidae, Prododomidae, Salticidae and Thomisidae. Many of these were only found on one sub-plot, therefore offering limited information to this analysis. On average, 50% of morphospecies on unburnt plots were primarily moist habitat specialists, 20% have a known preference for dry habitats, 10% are litter dwellers, and 5% are known open/disturbed habitat specialists. The remaining 14% are generalist hunters with more flexible habitat requirements. On average, 6% of morphospecies on unburnt plots were primarily moist habitat specialists, 27% have a known preference for dry habitats, 21% are litter dwellers, Figure 3.28 Comparison of Spider community structure and 36% are known open/disturbed habitat on burnt and unburnt plots specialists. The remaining 11% are generalist primarily based on feeding strategy and were; hunters with more flexible habitat requirements. primarily predacious (Staphylinidae, Scydmaenidae, A graphical comparison of these data (Figure Pselaphidae and Carabidae), fungal feeders 3.28) indicates that frequent burning has resulted in (Leiodidae), generalists (Scarabaeidae), a fundamental shift in spider community structure. phytophagous (Chrysomelidae & Curculionidae) With regard to habitat preference, there has been, and others. This last category included on average, an 88% decrease in the number of morphospecies from the families Ptiliidae, moist habitat specialists, and an 35% increase in the Silvanidae, Endomychidae, Corylophidae, number of species with a known preference for dry Phalacridae & Lathridiidae (fungal feeders), habitats. The change in moist habitat specialists Trogidae, Hybosoridae, Elateridae, Nitidulidae, was primarily due to a 90-95% decrease in species Tenebrionidae and Anthicidae (generalists), and from the families Theridiidae, Toxopidae and Cerambycidae, Oedemeridae, Aderidae & Oonopidae. The number of litter dwelling species Anobiidae (phytophagous). They were not included has, on average, increased by 110% with an directly into the above groupings because most equivalent increase from the families Hahniidae morphospecies were only found on one sub-plot and Textricellidae. The number of species known and therefore offer limited information to this to prefer open and disturbed habitats has, on analysis. average, increased by over 600% due primarily to 7 On average, 70% of morphospecies on species from the family Zodariidae which were only unburnt plots were primarily predacious, 5% found on burnt sub-plots. fungal feeders, 7% generalists and 5% 3.3.5.4 Beetles phytophagous (plant eating). On average, 57% of Morphospecies were classified into one of 5 groups morphospecies on unburnt plots were primarily based on the known habits and requirements at the predacious, 3% fungal feeders, 12% generalists family level (see Table 3.22). These groups were and 16% phytophagous.

Table 3.22 Comparison of Beetle community structure on burnt and unburnt plots. Data represent mean percentage of morphospecies in each category. Ecological role Family Unburnt Burnt Staphylinidae 33.1 15.3 General Scydmaenidae 5.6 6.3 predators Pselaphidae 4.0 5.6 Carabidae 27.0 29.7 Fungal feeders Leionidae 4.8 3.3 Generalists Scarabaeidae 6.9 12.4 Phytophagous Chrysomelidae 1.3 6.6 Curculionidae 3.3 9.5 Others (see text) 14.0 11.3 Total 100 100

237 Australia’s Biodiveristy - Responses to Fire

A graphical comparison of these data (Figure Table 3.23 Comparison of Ant community 3.29) indicates that frequent burning has resulted in structure on burnt and unburnt plots. Data a fundamental shift in beetle community structure. represent mean percentage of morphospecies in With regard to feeding strategy, there has been, on each category. average, an 18% decrease in the number of predator Ecological role Taxa Unburnt Burnt species, due primarily to a 54% reduction in the Dominants Iridomyrmex 15.6 12.0 number of morphospecies from the family Leptomyrmex 0.0 0.7 Staphylinidae. The average number of Sub-dominants Camponotus 8.7 10.7 morphospecies feeding on fungal products (as Polyrhachis 1.3 1.8 represented by the family Leiodidae) has declined Cerapachys 1.2 0.8 by 31%, with some other fungal feeders (families Melophorus 3.3 3.3 Ptiliidae & Endomychidae) decreasing by over Climate Meranoplus 2.1 3.1 200%. The generalists, as represented by the family Specialists Notoncus 2.1 4.5 Podomyrma 0.2 0.4 Scarabaeidae, have increased on average by 79%. Prolasius 8.2 6.2 Phytophagous species have increased dramatically Amblyopone 0.5 0.0 (up by 250%), particularly morphospecies from the Discothyrea 0.2 0.0 family Chrysomelidae. For this family, 56% of its Heteroponera 1.6 1.1 overall morphospecies were only found on Hypoponera 1.7 0.2 frequently burnt sub-plots. Cryptic Ponera 0.2 0.0 Species Solenopsis 4.6 4.4 Sphinctomyrmex 0.0 0.4 Stigmacros 3.6 4.1 Tapinoma 2.5 0.4 Trachymesopus 0.8 2.3 Mayriella 0.9 0.8 Paratrechina 9.9 8.4 Opportunists Rhytidoponera 3.7 10.5 Technomyrmex 0.5 0.2 Tetramorium 3.8 4.6 Crematogaster 4.9 3.9 Generalists Monomorium 0.8 3.8 Pheidole 11.5 7.4 Bothroponera 2.8 0.7 Colobostruma 0.0 0.6 Solitary/ Epopostruma 0.0 0.2 Figure 3.29 Comparison of Beetle community structure Specialists Leptogenys 0.0 0.8 on burnt and unburnt plots Myrmecia 1.4 2.5 Total 100 100 3.3.5.5 Ants Morphospecies were classified into one of 7 On average, 16% of morphospecies on “functional groups” groups based on the known unburnt plots were dominants, 10% sub- habits and requirements at the generic level (see dominants, 18% climate specialists, 16% cryptic Andersen 1990). These groups were; dominants species, 19% opportunists, 17% generalists, and 4% (Iridomyrmex & Leptomyrmex), sub-dominants solitary/ specialist species. On average, 16% of (Camponotus & Polyrhachis), climate specialists morphospecies on unburnt plots were dominants, (Cerapachys, Melophorus, Meranoplus, Notoncus, 10% sub-dominants, 18% climate specialists, 16% Podomyrma & Prolasius), cryptic species cryptic species, 19% opportunists, 17% generalists, (Amblyopone, Discothyrea, Heteroponera, Hypoponera, and 4% solitary/specialist species. Ponera, Solenopsis, Sphinctomyrmex, Stigmacros, A graphical comparison of these data (Figure Tapinoma and Trachymesopus), opportunists 3.30) indicates that frequent burning has resulted in (Mayriella, Paratrechina, Rhytidoponera, a slight shift in ant community structure. The Technomyrmex, Tetramorium, Crematogaster, number of morphospecies within the “dominant” Monomorium & Pheidole) and solitary/specialist functional group has, on average, decreased by 23%, species (Bothroponera, Colobostruma, Epopostruma, primarily through a reduction in the occurrence of Leptogenys & Myrmecia). The average number of Iridomyrmex species. The number of “sub- morphospecies from these genera on each dominants” has increased slightly (25%), largely treatment are shown in Table 3.23. because of an increase in the occurrence of 238 Bushfire and forest invertebrates

3.3.5.6 Summary This Section examined community structure by grouping morphospecies into ecological groups based upon their feeding strategies and habitat preferences. By classifying large numbers of species into smaller, more manageable groups, it is possible to substantially reduce the apparent complexity of ecological systems and provide a basis for evaluating environmental change (Andersen 1990). For all groups studied, frequent burning Figure 3.30 Comparison of Ant community structure on resulted in a change in the structure of the burnt and unburnt plots community. With regard to feeding strategy, there was on average, a 15, 140 and 250% increase Camponotus species. With regard to “climate (respectively) in the number of phytophagous specialists”, they have remained largely stable (plant feeding) species of bugs, flies and beetles. (increased by 4%) as a group, although individual With regard to the proportion of predator species, genera did vary in their response (see Table 3.23). it remained unchanged for flies and bugs, however There was a decrease (19%) in the average number for bugs there was a total shift from the family of “cryptic” morphospecies, particularly with regard Nabidae to the family Reduviidae. With beetles to the genera Hypoponera (90%) and Tapinoma there was, on average, an 18% decrease in the (86%). A number of morphospecies from this group number of predator species, due primarily to a 54% however increased in their average occurrence; reduction in the number of morphospecies from Stigmacros (13%) and Trachymesopus (196%), the family Staphylinidae. For groups feeding although the actual number of species involved is primarily on fungal products there was, on average, small (overall 3 and 1 respectively). a 44% (flies) and 31% (beetles) decrease in the There was a substantial increase overall in number of morphospecies. For flies this was most the average number of “opportunist” marked in the families Scaridae and Scatopsidae, morphospecies (30%), primarily due to one genus, and for beetles, in the families Leiodidae, Ptiliidae Rhytidoponera, which increased by 180%. In fact & Endomychidae. For flies, the number of this was attributable to a single species: generalists and scavengers has decreased slightly Rhytidoponera metallica, which was extremely (11%), while for beetles the proportion of numerous on burnt sub-plots and contributed to generalists has increased, on average, by 79%, the high ant abundance detected for this treatment primarily through an increase in morphospecies (see Section 3.3.2.10). The number of “generalist” from the family Scarabaeidae. With ants, the morphospecies, on average, decreased by 11% proportion of generalist morphospecies decreased, following frequent burning. Results were not on average, by 11% following frequent burning. consistent within the group with the genera Results were not consistent within the group with Pheidole and Crematogaster decreasing by 35 and the genera Pheidole and Crematogaster decreasing by 19% respectively, and the genus Monomorium 35 and 19% respectively, and the genus increasing by 388%. The number of larger Monomorium increasing by 388%. For flies, the “solitary/specialist” morphospecies increased number of morphospecies regarded as wide- slightly (13%), although the results were quite ranging “tourists” was, on average, reduced by 43% variable within the group. Three new genera were on frequently burnt plots. found on burnt plots (Colobostruma, Epopostruma & When groups were compared with regard to Leptogenys) while the numbers of species of habitat preferences, it was apparent that there Myrmecia increased, on average, by 87%. The were substantial changes in community structure occurrence of the solitary forager Bothroponera sp. for some taxa. For bugs, numbers of moist habitat A decreased substantially (76%). specialists from the sub-order Dipsocoromorpha were, on average, reduced by 83%. Amongst spiders, moist habitat specialists were reduced by 88%, primarily due to a 90-95% decrease in numbers of morphospecies from the families

239 Australia’s Biodiveristy - Responses to Fire

Theridiidae, Toxopidae and Oonopidae. With the morphospecies, due primarily to a single species: flies, the number of moist habitat specialists Rhytidoponera metallica, which was extremely remained similar, although the family numerous on burnt sub-plots. Ceratopogonidae was more commonly It has been observed that the structure of ant represented on burnt sub-plots and the family communities, in particular, may be influenced by Chironomidae on unburnt sub-plots. The the relative abundance of particular “dominant” proportion of temperature-dependent (climate) and “sub-dominant’ groups (Fox & Fox 1982, specialists among the ants remained largely stable Andersen 1990). Following frequent fire, the (increased by 4%), although individual genera did number of morphospecies within the “dominant” vary in their response. For spiders, there was, on functional group had, on average, decreased by average, and an 35% increase in the number of 23%, primarily through a reduction in the species with a known preference for dry habitats. occurrence of Iridomyrmex species. The number of For groups primarily inhabiting the litter layer “sub-dominants” had increased slightly (25%), there was a variety of responses to frequent burning. largely because of an increase in the occurrence of For flies, the number of morphospecies specifically Camponotus species. The number of larger utilising the litter layer had, on average, decreased “solitary/specialist” morphospecies, which interact by 60%. This was primarily due to the absence of only slightly with other groups, had increased the family Tipulidae on frequently burnt sub-plots. slightly (13%), although the results were quite With ants there was a decrease (19%) in the average variable within the group. Three new genera were number of “cryptic” morphospecies inhabiting the found on burnt plots (Colobostruma, Epopostruma litter and soil, particularly with regard to the genera & Leptogenys) while the numbers of species of Hypoponera (90%) and Tapinoma (86%). For spiders Myrmecia increased, on average, by 87%. The however, the number of litter dwelling species occurrence of a solitary forager, Bothroponera sp.A, increased, on average, by 110% with an equivalent had decreased substantially (76%). increase from the families Hahniidae and Textricellidae. 3.3.6 Biodiversity Indicators Many groups of terrestrial invertebrates are It is often postulated that one group of invertebrates adapted to exploit disturbed habitats. For spider may act as an “indicator” or “umbrella” group for morphospecies known to prefer open and others, thereby allowing inferences to be made on disturbed habitats, the number of morphospecies the impact of disturbance regimes. The relationship increased, on average, by over 600% following between species richness of selected taxa was frequent burning, This was due primarily to the investigated here using Pearson’s Product-Moment occurrence of seven species from the family correlation analyses. Table 3.24 gives the correlation Zodariidae only on burnt sub-plots. For the ants coefficients (and probability values) for the there was a substantial increase overall (30%) in relationship between richness values for pairs of taxa the average proportion of “opportunist” at the sub-plot scale. The values in the bottom left

Table 3.24 Correlations between species richness values for five taxa. Values on the bottom-left represent standard Pearson’s Product-Moment correlation coefficients (n=48), those on the top-right are partial coefficients (n=43) controlling for the effects of the other taxa in each comparison. Pairs of data represent correlation coefficient (top) and probability (bottom) values. UNBURNT ANTS BEETLES BUGS SPIDERS FLIES ANTS -0.053 0.131 0.292 -0.343 0.727 0.391 0.052 0.021 BEETLES -0.238 0.071 -0.160 0.352 0.103 0.643 0.293 0.018 BUGS 0.174 0.005 0.055 -0.058 0.2360.974 0.722 0.704 SPIDERS 0.312 -0.189 0.093 0.122 0.031 0.199 0.529 0.424 FLIES -0.394 0.395 -0.098 -0.062 0.0060.0060.5060.676

240 Bushfire and forest invertebrates

of the table represent standard correlation coefficients, while those in the top-right are “partial” coefficients, controlling for the possible effect of other variables (taxa). This tests, for example, whether a spurious relationship between two taxa exists because of a separate relationship they may have individually with another taxa. A examination of the correlation coefficients reveals three significant relationships. Firstly, species richness values for spiders and ants are positively correlated (r = 0.312, P = 0.031, n = 48). This implies that as the richness of ants increases so does the richness of spiders. One could be used to predict the other, however the predictive power Figure 3.31 Relationship between species richness is low with only 9.6% of the variance in one taxa values for ants and spiders contributed by the variance of the other. In addition, this coefficient decreases slightly when controlling for the effect of other taxa (r = 0.292, ants, and between flies and beetles. Given this P = 0.052, n = 43). situation, and the previously identified differences Secondly, species richness values for flies and between burnt and unburnt areas, both in regard ants are negatively correlated (r = -0.394, P = to habitat conditions (see Section 3.1) and their 0.006, n = 48). This implies that as the richness of faunal assemblages (see Section 3.3), it was ants increases the richness of flies decreases (and appropriate to test correlations separately for the vice versa). One could be used to predict the two treatments. other, however the predictive power is low with Table 3.25 gives the correlation coefficients only 15.5% of the variance in one taxa (and probability values) for the relationship between contributed by the variance of the other. In richness values for pairs of taxa. The values in the addition, this coefficient decreases slightly when bottom-left of the table represent correlation controlling for the effect of other taxa (r = -0.343, coefficients for burnt sites, while those in the top- P = 0.021, n = 43). right are coefficients for unburnt sites. In other Thirdly, species richness values for flies and words, this approach “controls for” the possible beetles are positively correlated (r= 0.395, P = 0.006, effect of treatment. n = 48). This implies that as the richness of flies When the data are analysed in this way it is increases so does the richness of beetles. One could obvious that there are no significant relationships be used to predict the other, however the predictive between the species richness values for the five power is low with only 15.6% of the variance in one taxa examined. Therefore, none of these groups of taxa contributed by the variance of the other. In terrestrial invertebrates would be a reliable addition, this coefficient decreases slightly when “indicator” or “umbrella” group for any other, and controlling for the effect of other taxa (r = 0.352, P therefore are inappropriate for predicting overall = 0.018, n = 43). None of the other combinations of biodiversity at this scale. taxa have statistically significant correlations (see Given the effects of sampling “scale” on Table 3.24). species richness identified in Section 3.3.3.6, and To further examine the nature of the the influence of α- and β-diversity patterns relationship between richness values for ants and (Section 3.3.4.6), the richness data were re- spiders, the data were examined graphically analysed using the plot-based values (sum of 4 (Figure 3.31). It is obvious that the apparent sub-plots) for each treatment. Table 3.26 gives the positive correlation between these two taxa stems correlation coefficients (and probability values) for primarily from the disparity between richness the relationship between richness values for pairs values on burnt and unburnt sub-plots. Both taxa of taxa at the plot scale (n=6). The values in the have significantly higher values on burnt sub-plots bottom-left of the table represent correlation producing a relationship which is in fact an artefact coefficients for burnt sites, while those in the top- of this difference in richness. Similar patterns were right are coefficients for unburnt sites. There are exhibited for the relationship between flies and two statistically significant coefficients, ants &

241 Australia’s Biodiveristy - Responses to Fire

Table 3.25 Correlations between species richness values for five taxa (controlling for treatment). Values on the bottom-left represent standard Pearson’s Product-Moment Correlation Coefficients (n=24) for burnt sub-plots, those on the top-right are for unburnt sub-plots. Pairs of data represent correlation coefficient (top) and probability (bottom) values. UNBURNT ANTS BEETLES BUGS SPIDERS FLIES ANTS 0.042 0.086 0.064 -0.077 0.847 0.689 0.765 0.721 BEETLES 0.267 0.123 -0.213 0.386 0.208 0.568 0.318 0.063 BUGS 0.105 0.098 0.275 -0.035 0.624 0.648 0.193 0.871 SPIDERS 0.260 0.195 -0.181 0.175 0.220 0.361 0.397 0.414 FLIES 0.076 -0.104 0.024 0.101 0.724 0.628 0.912 0.640

Table 3.26 Correlations between species richness values for five taxa (controlling for treatment). Values on the bottom-left represent standard Pearson’s Product-Moment Correlation Coefficients (n=6) for burnt sub- plots, those on the top-right are for unburnt sub-plots. Pairs of data represent correlation coefficient (top) and probability (bottom) values. UNBURNT ANTS BEETLES BUGS SPIDERS FLIES ANTS -0.719 -0.968 0.603 0.353 0.107 0.001 0.205 0.492 BEETLES 0.751 0.664 -0.818 0.161 0.085 0.150 0.0460.760 BUGS 0.046 -0.185 -0.591 -0.388 0.930 0.7260.2160.447 SPIDERS 0.155 -0.189 -0.644 0.273 0.770 0.719 0.167 0.600 FLIES -0.120 -0.112 -0.021 0.186 0.821 0.832 0.968 0.724

bugs and beetles & spiders, but only on unburnt used to predict the other, and the predictive plots. None of the other combinations of taxa power is high with 66.9% of the variance in one have statistically significant correlations. taxa contributed by the variance of the other. On unburnt plots only, species richness These results suggest that, at least for these 2 values for ants and bugs are negatively correlated pairs of taxa, and at the 1ha sampling scale, it may (r= -0.968, P = 0.001, n = 6). This implies that as be possible and reliable to use one group as an the richness of ants increases, the richness of bugs indicator of the biodiversity of the other. decreases (and vice versa). One could be used to predict the other, and the predictive power is high with 93.7% of the variance in one taxa contributed by the variance of the other. Similarly, on unburnt plots only the species richness values for beetles and spiders are negatively correlated (r= -0.818, P = 0.046, n = 6). This implies that as the richness of beetles increases, the richness of spiders decreases (and vice versa). One could be

242 4. DISCUSSION

There is good theoretical and growing empirical 4.1 HABITAT STRUCTURE evidence to support the role of biodiversity in the maintenance of ecological processes within Low intensity fires used for fuel control generally forests. The multitude of organisms that result in incomplete combustion of surface litter constitute biodiversity play an essential role in and understorey vegetation (Tolhurst et al. 1992; primary production, nutrient cycling and uptake, Williams and Gill 1995). A mosaic of habitat population and community level interactions and patches results at a small scale, with these patches energy storage and transfer (see Majer 1992b; influencing the spatial distribution of surviving Woodward 1993; Beattie 1995). Through their terrestrial invertebrates and their ability to contribution to ecosystem function, these recolonise burnt areas. If repeated low-intensity organisms also enable forest ecosystems to fires reduce this spatial heterogeneity (see Fox and provide benefits to humanity. These include Fox 1986; Nieuwenhuis 1987) then this practice amenity values in the form of aesthetics, recreation may have long-term consequences for the survival and education; heritage values as forests contribute of invertebrate populations. In this study, 31 to long-term security for catchment protection, environmental variables were measured in order air & water quality and nature conservation; and to firstly, assess the long-term effect of frequent economic values including timber production, burning on the habitat, and secondly, to grazing and ecotourism (see Hobbs 1992; York investigate the nature of relationships between 1993; New 1995). habitat components and invertebrate biodiversity. The maintenance of biodiversity is a Four general trends with regard to fundamental principle underlying ecologically environmental variables were detected: large-scale sustainable management (NSESD 1992). State spatial patterns, site-dependent patterns, Forests of New South Wales, through its Corporate treatment-dependent patterns, and general Plan (1992), has stated that it will manage its patterns independent of site and treatment. At the forests on an ecologically sustainable basis using largest scale (the study area), there was a gradual best forest practices. Biodiversity conservation, N-S decrease in the mean biomass of leaves as a and hence ecological sustainability, cannot be component of the leaf litter. This pattern was only achieved without consideration of the important evident however on unburnt plots and role that invertebrates play, both through their disappeared with frequent burning, resulting in a involvement in ecological processes, and by their simplification of the large-scale spatial patterning substantial contribution to the overall richness of of this habitat component. Top-soil moisture biological communities. Invertebrates are the content tended to rise to a peak towards to centre most diverse and abundant animals in most of the N-S road transect and then decline towards natural ecosystems, but their importance in the southern end, with frequent burning having sustaining those systems is commonly not no effect on this spatial trend. appreciated (New 1995). Periodic low-intensity Frequent low-intensity fire had resulted in a fire (hazard-reduction burning) is a conspicuous number of treatment-related changes in measured management strategy in virtually all of Australia’s habitat parameters. Within the “fine-fuel” dry forest communities. While it is primarily used component, there was (on average) a 44% to reduce fuel levels, little is known about the reduction in leaf biomass, and a reduction in effects of its repeated use on natural ecosystems large-scale spatial trends (see above). There was over long time-scales. The primary objective of (on average) a 63% reduction in the very fine this study was therefore to assess whether the litter component, with the appearance of spatial frequent use of this forest management practice patterning not evident on unburnt plots. With the was compatible with the conservation of a major twig component, there had been (on average) a component of biodiversity: the terrestrial 50% reduction in the biomass of twigs 0-6mm, invertebrates. and a 44% reduction in the biomass of twigs 6- 25mm. Bark biomass had decreased (on average)

243 Australia’s Biodiveristy - Responses to Fire

by 36%, and increased in spatial heterogeneity. sized shrubs, the spatial homogeneity of certain These reductions in litter biomass largely reflect a litter components (biomass of leaves and very fine response to the most recent fuel-reduction burn material) and top-soil hardness (but see site- (2 years previously) and would be expected to dependent effects mentioned above). There was change with time as fuel continues to accumulate also no obvious change with regard to the (see Birk and Bridges 1989). Over the 20 year distribution of sticks & logs (>2.5cm) within period of frequent hazard-reduction burning, nominated size categories, however some changes litter biomass (fine fuel) had fluctuated between to the external nature of logs (charring) was 15 and 23 tonnes ha-1 on unburnt plots, and apparent. between 4 and 20 tonnes ha-1 on burnt plots There were a number of “general patterns” (York 1996). Burning removed between 46 and with regard to the response of environmental 73% of litter (by weight), but by 3 years post-fire variables to frequent fire. As the amount of litter levels had usually reached and often vegetation in the first metre above the ground exceeded that achieved before the previous fire increased, it became more spatially homogeneous (range = 72–136%). At the time of this study mean (less “patchy”). The amount of leaf litter (twigs, litter biomass was 9.3 tonnes ha-1 on burnt plots, bark, leaves and very fine material) at study sites representing about 50% of levels on unburnt was independent of the amount of understorey plots. vegetation, but was correlated (positively) with Other treatment-related effects involved the top-soil moisture levels and (negatively) with the amount (cover) of vegetation in the understorey amount and spatial variability of light reaching the and its spatial distribution. While the quantity of ground. This suggests that frequent fire has an vegetation in the layers closest to the ground independent influence on vegetation understorey (ground herbs and small shrubs) was not affected and leaf-litter habitats, and that (in this forest) the by frequent burning (see below), there was an leaf-litter environment exerts the primary control decrease in the spatial heterogeneity on top-soil moisture and surface insolation levels. (“patchiness”) of these layers. Conversely, there Christensen (1985) found that the removal of the was a substantial reduction in the cover of tall and litter layer by fire caused increased surface very tall shrubs (on average, 65% and 93% heating, leading to greater evaporation and lower respectively). Both these layers showed an increase moisture in the upper soil. This situation would in spatial heterogeneity with frequent burning. be expected to change with the time-related Top-soil moisture levels were, on average, 18% accumulation of litter after fire. lower following 20 years of frequent burning, While the changes to the amount of leaf whereas the amount of light reaching ground level litter, understorey vegetation and top-soil had increased (on average) by 125%, and moisture may only reflect a time-since-last-fire increased in its spatial heterogeneity. phenomenon, changes to the spatial variability Spatial patterning of a number of following frequent burning may reflect a decrease environmental variables was largely site- in habitat heterogeneity which, in turn, could dependent, with no overall large-scale or impact upon terrestrial invertebrate communities. treatment-related features. Top-soil hardness was There were measured reductions in the large- site-specific and independent of other variables. scale spatial patterning of the leaf litter, and There was a weak tendency for there to be greater changes to its physical structure with an increase amounts of vegetation in the first metre above in the patchiness of bark and very fine litter ground on sub-plots with more exposed (north- components. The increased patchiness of ground westerly) aspects, and there was a slight tendency insolation levels would appear primarily to be a for there to be a greater number of large logs on response to these changes. There was a decrease sub-plots with lower slope angles. in the spatial heterogeneity of ground herbs & A number of habitat components showed no small shrubs but an increase for tall & very tall significant response to frequent burning and did shrubs. This potentially reflects a change in the not appear to exhibit patterns that could be composition of the understorey vegetation attributed to large-scale spatial influences. These following frequent burning (see Fox and Fox were: the amount of vegetation in the first metre 1986; Nieuwenhuis 1987) and a removal of taller above the ground (ground herbs, small and mid- (older) shrubs due to the short interval between sized shrubs), the spatial homogeneity of mid- fires. The relationship of these environmental

244 Bushfire and forest invertebrates

patterns to the abundance and distribution of open habitats (CSIRO 1991) such as those typical terrestrial invertebrates is explored and discussed of frequently burnt areas, although Dolva (1993) in the following sections. found that wood crickets (Gryllidae) were more abundant in unburnt areas, probably in response 4.2 TERRESTRIAL INVERTEBRATE to the structure of the litter layer. Thrips are COMMUNITIES generally associated with vegetation and may be responding to the marked changes in the structure This study revealed a rich terrestrial invertebrate and cover of the understorey. fauna with representatives from the Chelicerata For the ten broad taxonomic groups where (spiders, ticks & mites, pseudoscorpions, there were sufficient data to permit statistical harvestmen), Crustacea (landhoppers, slaters), testing, the results indicated a variety of responses Chilopoda (centipedes), Diplopoda (millipedes), to frequent burning. Seven groups (isopods, and a large number of Insect Orders & Families. springtails, ticks & mites, bees & wasps, insect 4.2.1 Invertebrate Abundance larvae, flies and beetles) showed substantial decreases in abundance following frequent Numerically, the most abundant groups overall burning. These decreases ranged from 15 to 58% were the springtails (33.1%), ticks & mites (see Section 3.3.2), but were only statistically (23.9%), ants (23.1%), bugs (4.2%), beetles significant for ticks & mites (31%), insect larvae (4.0%), bees & wasps (2.8%), insect larvae (2.7%), (35%), flies (58%) and beetles (31%). Many of flies (2.6%) and spiders (2.2%), with these nine these groups are associated with leaf litter and it is groups making up 98.6% of the total number of likely that their numbers have been influenced by organisms caught. The first three groups the episodic removal of this resource, and the fact (springtails, ticks & mites, and ants) represented that litter levels on frequently burnt plots were, on 80% of individuals caught. average, 50% of that on unburnt plots. Newman Due to their low numbers, it was not and Tolhurst (1991) considered that reductions in possible to comment on the effects of frequent abundance of Collembola and Diptera following a burning for: pseudoscorpions, harvestmen, single fire event were in response to reduced litter centipedes, millipedes, diplurans, termites, (fuel) levels. Terrestrial mites are exceedingly embiids, booklice, lacewings, caddisflies, moths common in soil and leaf litter and are generally and butterflies. For these groups the trapping predatory, feeding on small invertebrates (Harvey method used may not have been the most and Yen 1989). Collembola communities have appropriate and has potentially contributed to the been shown to be initially particularly sensitive to low capture rate. While the low numbers fire (Campbell 1973), and an increase in fire collected for several other taxa precluded frequency may reduce population sizes and alter statistical analysis, frequent burning appears to community structure (Metz and Dindal 1975; have led to a reduction in the numbers of Dindal and Metz 1977). It was interesting to note amphipods, cockroaches and earwigs, and an that the large-scale patterns in abundance increase in the numbers of grasshoppers & exhibited by ticks & mites and the beetles crickets, and thrips. Terrestrial amphipods live in paralleled patterns in top-soil moisture content and feed on decaying litter of the forest floor, (see Figures 3.4, 3.8 and 3.14). The observed requiring a relatively moist environment because significant reduction in abundance of these groups they are susceptible to desiccation (Friend and (both down by 31%) may be in response to the Richardson 1986). Densities of some species have 18% decrease (on average) in top-soil moisture been correlated with litter thickness (eg., Duncan levels apparent on sites experiencing frequent fire. 1969) suggesting that the reduction in leaf litter High spatial variability in abundance for isopods, associated with hazard-reduction burning is likely springtails, and bees & wasps possibly contributed to result in lower population sizes for this group. to the lack of statistical significance in this study. Similarly, cockroaches feed mainly on the detritus Three groups showed substantial increases associated with leaf litter while the majority of in abundance following frequent burning. These earwigs feed on live or decaying plant matter were statistically significant for bugs (77%) and (Zborowski and Storey 1995). Both groups would ants (250%), but not for spiders (33%). This may be impacted upon by a reduction in the amount of be due to increased ease of movement (increased this resource. Surface active groups such as “trapability” - see Majer 1980, Andersen 1988) for grasshoppers and crickets tend to prefer more

245 Australia’s Biodiveristy - Responses to Fire

surface active groups such as spiders and ants, as some lacewings feeding on honeydew or pollen, well as changes in habitat suitability. The relative while others prey on aphids and scale-insects abundance of many groups was influenced not (Zborowski and Storey 1995). This group may be only by fire history, but also by site-specific more abundant in the more structurally complex habitat conditions (see below), with both spiders vegetation of unburnt sites, however a sampling and ants showing considerable spatial variability in regime concentrating on understorey vegetation their numbers. This suggested that large sample would be required to more accurately assess this. sizes would be required to detect management In order to test the utility of RBA effects when using coarse-scale taxonomic methodology, and further investigate the impact classification (eg. Family or Order). While the of repeated burning on species richness and the uncertain taxonomy of many groups precludes related aspects of community composition and finer scale resolution, recent developments in structure, five taxa were investigated in detail by Rapid Biodiversity Assessment (RBA) may permit analysis to morphospecies level. These groups subsets of the fauna to be investigated more fully utilise a diversity of micro-habitats and niches and (Oliver and Beattie 1993; Beattie and Oliver are representative of the range of terrestrial 1994). This would allow the use of smaller sample invertebrates found in these forest environments. sizes and produce a more cost-effective outcome These were: firstly, the Hemiptera (bugs), a (see below). mostly terrestrial and phytophagous (plant- feeding) group which have a close association with 4.2.2 Invertebrate Species Richness plant communities. Secondly, the Diptera (flies), Diversity (richness) at the Ordinal level varied which although highly mobile as adults, have from 11–17 broad taxa on individual study plots, particular requirements with regard to larval food with frequent burning significantly reducing sources; usually moist, decaying plant and animal diversity at this scale. While on average this material. Many species are parasitic on the larvae decrease was slight (≈ 1 Order per sub-plot), 4 of other insect orders with specialist habitat taxa were missing overall from frequently burnt requirements for oviposition. Thirdly, the plots. These were the Opilionida (harvestmen), Araneae (spiders), a major group of predators in Embioptera (embiids), Psocoptera (booklice) and forest ecosystems exploiting a variety of habitats. Neuroptera (lacewings). The Opilionida They live in burrows or crevices in the ground, (harvestmen) are small to medium Arachnids amongst leaf litter or in vegetation, and are a (<10mm body length) which are usually found in group with many habitat specialists. Fourthly, the moist leaf litter, or under rocks, logs and bark. Coleoptera (beetles), which utilise a diverse range Most feed on smaller invertebrates but some also of habitats & micro-habitats, with a variety of consume plant material (Harvey and Yen 1989). feeding strategies (adults include herbivores, Given their habitat requirements, it is expected predators & scavengers, while larval forms feed that they would be disadvantaged by the drier either internally or externally on plants and fungal conditions found on frequently burnt areas. products). Beetles are a rich and diverse group Embiids (web-spinners) are small to medium which are active in the litter layer. Lastly, the (4–15mm body length) insects usually living under Formicidae (ants), which are one of the most rocks, bark or leaf litter. They feed on leaves, numerous and widespread groups in Australian bark, mosses and lichens (Zborowski and Storey ecosystems. They have a diverse diet, and utilise a 1995). The Psocoptera (booklice, psocids) are variety of feeding strategies from predators and minute to small (<1–10mm body length) insects scavengers, to plant eaters and fungus feeders, which live on vegetation, or under bark or stones. with frequent and varied interactions with other They feed on minute organic items such as plant invertebrate groups. Ants nest in the soil and litter spores, algae, lichen and fungi. For both groups, and therefore are response to disturbance of these these food resources are likely to be more habitats, and they are functionally important abundant in the moister litter and soil conditions within the forest ecosystem. prevalent in infrequently burnt forest. The Overall, 411 morphospecies were identified Neuroptera (lacewings) are small to large from the five groups studied in detail. The beetles (wingspan 5–150mm) insects with generally were the most species rich (139 morphospecies), active, long-legged predacious larvae (antlions). followed by the ants (88), flies (77), spiders (63), Adults may be predacious or omnivorous, with and bugs (44). The results of analyses (ANOVA)

246 Bushfire and forest invertebrates

investigating the effects of frequent burning and observation is the lack of concordance in spatial patterns due to large-scale spatial effects indicated patterns of richness between taxonomic groups. a variety of responses to frequent burning. Two This suggests that groups are responding groups, flies and beetles, experienced a significant differently to environmental factors and to the reduction in species richness on sub-plots agent of disturbance (frequent fire). This has following frequent burning (44% and 27% implications not only for invertebrate sampling reduction respectively). A further two groups, the strategies, but also for the use of a single, or bugs and the spiders, showed an increase in limited group of taxa as a surrogate in biodiversity species richness on sub-plots (16% and 27% assessment. This supports the findings of Oliver respectively), although these results were not (1995) who found that different taxa responded to statistically significant. The ants experienced a the disturbance of forest logging in distinct ways, significant increase in sub-plot richness (26%) and that it was not appropriate to use any one following repeated burning. Few studies have taxon as a surrogate for the richness of any others dealt with the impact of fire on invertebrate in conservation evaluation, environmental species richness. Leonard (1972) found that the monitoring or impact assessment. species richness of leaf litter fauna may drop by It was apparent that estimates of species 50% immediately after fire. Recovery after fire richness were also influenced by the spatial scale may be rapid (Leonard 1972), or take several years of measurement, with associated implications for (Moulton 1982), depending upon the season of the interpretation of observed treatment effects burn and the meteorological conditions following for the different taxa. For bugs and flies results the fire and the influence these factors have on were consistent across a range of scales of litter accumulation, and the mobility and measurement, with the magnitude and direction recolonising ability of particular species (Morris of differences between unburnt and burnt results 1975). Long-term studies of spiders (Huhta 1971; for sub-plot, plot and treatment similar. For Merrett 1976) and ants (Brian et al. 1976; York spiders, while the magnitude and direction of 1994) have shown a replacement of species in the species richness at the scale of sub-plot and plot years after fire which is related to their particular were similar, considerably more species were habitat requirements being met as the vegetation found overall on burnt compared to unburnt structure changes over time. Species richness may plots. This suggested that species assemblages on stay largely unchanged (Merrett 1976) or decline burnt plots were more diverse than those on (York 1994). unburnt plots, resulting in higher β- (between- At the scale at which richness (α-diversity) habitat) diversity. Diversity on sub-plots within was estimated in this study there were large-scale both unburnt and burnt plots would appear to be spatial patterns exhibited by some groups. For similar, suggesting the differences lie at, or above, bugs, spiders, beetles and ants these were not the scale of plot ( 1 hectare). For beetles, the statistically significant overall, but nevertheless magnitude of the difference detected between have implications for future projects attempting to unburnt and burnt at the scale of sub-plot and measure disturbance impacts. For ants, observed plot were similar, however the direction was spatial variation in richness appeared to reflect reversed at the scale of treatment. This would random variation expected within a sampling suggest a similar situation as to that with the program such as this. For other groups observed spiders, where the species assemblages on burnt variation was distinctly non-random and appeared plots are more diverse than those on unburnt to reflect underlying environmental patterns. For plots. Diversity on sub-plots within both unburnt bugs similar trends in richness were apparent at and burnt plots would appear to be similar, paired sites along the road transect suggesting that suggesting the differences lie at, or above, the a smaller number of replicates would have been scale of plot ( 1 hectare). For ants, the magnitude sufficient to detect impact (or lack of impact). For of the difference in species richness detected flies, spiders and beetles this was not the case. between unburnt and burnt areas at the scale of Exhibited spatial patterns were often quite sub-plot and plot were similar, although the different between the two treatments for a single magnitude of the difference was reduced at the taxa (eg. Flies - Section 3.3.3.2), reinforcing the scale of plot (compared with other taxa). The need for sufficient experimental replication in direction of the difference was however order to detect real differences. A most important substantially reversed at the scale of treatment,

247 Australia’s Biodiveristy - Responses to Fire

suggesting a different situation to that with the proportions of morphospecies in each category spiders and beetles, with the species assemblages however varied substantially between taxonomic on burnt plots less diverse. Diversity on sub-plots groups. For Hemiptera (bugs) the proportions within both unburnt and burnt plots would appear were 16, 41 and 43% for both, unburnt and burnt to be less similar than with other taxa, suggesting respectively; for Diptera (flies) 45, 40 and 15%; the differences lie at less than the scale of plot ( 1 for spiders 25, 24 and 51%; for beetles 28, 34 and hectare). 38%; and for ants 57, 23 and 20%. The overall These spatial patterns in estimates of species biodiversity of frequently burnt areas was richness are a consequence of the composition of maintained by the addition of species not invertebrate assemblages (communities), and their recorded on unburnt plots. The changed response to habitat conditions, at the varying environment was supporting an additional 133 scales of investigation. The nature of these morphospecies (19 bug, 11 fly, 32 spider, 53 beetle patterns, and their interaction with environmental and 18 ant species). It is notable that a large variables, will be further explored in the following proportion of species (16-47%) are apparently section. indifferent to disturbance history and habitat structure within the limits sampled in this survey, 4.2.3 Community Composition although changes to relative abundance need to be The five groups studies in detail proved to be taken into consideration. extremely diverse. Beetles had the richness fauna Although an examination of relative overall with 139 beetle morphospecies abundance patterns enables broad “assemblages” representative of nine super-families and 25 of species with similar responses to disturbance to families. The ants were the second richness group be identified, these different patterns were more with 88 morphospecies representative of 5 sub- clearly apparent from a comparison of bi-plots families and 34 genera. They were followed by the derived from the CCA ordination procedure. For flies with 77 morphospecies representative of 2 bugs, spiders, beetles and ants there was little or sub-orders and 20 families, the spiders with 63 no overlap of unburnt and burnt sub-plots in morphospecies representative of 21 families, and ordination space, indicating low similarity of the the bugs with 44 morphospecies representative of species assemblages of the two treatments. For 16 family (or similar) groups. flies however there was a substantial overlap, Overall, the same number of morphospecies reflecting the relatively large number of (279) were collected from unburnt and burnt morphospecies shared by the two treatments. plots, with average (mean) richness values similar Morphospecies found in both burnt and unburnt on both treatments (48.2 and 46.5 morphospecies areas can be regarded as habitat “generalists”, respectively). This initially suggests that frequent largely resilient to frequent disturbance and burning had not reduced biodiversity in this forest therefore of lesser concern with regard to environment. An analysis of the richness (α- and biodiversity conservation. Of greater importance β-diversity) of individual taxonomic groups has are those species absent from frequently burnt shown this not to be the case, with groups sites (potential habitat “specialists”). The results responding differently to frequent burning. The from this study suggest that frequent burning had nature and, potentially, the mechanisms behind led to the loss of up to 131 species (18 bugs, 31 this difference can only be elucidated by an flies, 15 spiders, 47 beetles and 20 ants), which examination of the species composition of faunal represents 47% of the morphospecies known from assemblages (communities) and would not be the unburnt areas. Many of the morphospecies apparent from an examination of data at a higher apparently lost from frequently burnt sites were taxonomic level (for example: Neumann and however only detected on a single sub-plot or Tolhurst 1991; Neumann 1992; Coy 1996). represented by a single individual on unburnt An inspection of the distribution of plots. These could be genuinely rare or morphospecies across sub-plots for each treatment uncommon species which were missed purely by detected a consistent pattern, irrespective of chance when sampling burnt plots. For this reason faunal group. Morphospecies fell into one of three it is difficult to identify clear patterns (at the groups: found on both treatments (Group A), species level) from the relative abundance data found only on unburnt plots (Group B), or found alone. Some general trends were apparent only on burnt plots (Group C). The relative however when morphospecies data were arranged

248 Bushfire and forest invertebrates

by genus (ants) or family (bugs, flies, spiders and following frequent burning, and lower top-soil beetles) utilising general information available on moisture levels, would appear to offer an their biology and ecology at these taxonomic explanation to the reduction in the number of levels (“guild” / “functional group” approach). species from the families Sciaridae, Phoridae and For bugs, unburnt plots had more species Empididae. Results of the CCA ordination from the infra-order Dipsocoromorpha (7 vs 3), support these conclusions with unburnt sub-plots while burnt plots have greater numbers of species characterised by high levels of litter, high cover of from the family Reduviidae (5 vs 0). The tall and very tall shrubs, high top-soil moisture Dipsocoromorpha include species (described & levels and low and spatially variable amounts of undescribed) primarily known from leaf litter and insolation at ground level. Burnt sub-plots were other moist environments (CSIRO 1991). The characterised by high levels of insolation at observed substantially lower (50%) amounts of ground level, greater exposure (more north- leaf litter and changes to its spatial distribution westerly aspects), and to a lesser extent, steeper following frequent burning would appear to offer slopes. These drier, more exposed conditions an explanation to the reduction in would be less favourable for species from these fly Dipsocoromorpha species. Results of the CCA families. This supports the findings of Delettre ordination support these conclusions with (1994) who, in a study of a heathland chironomid unburnt sub-plots characterised by high levels of (midge) community, found that species litter, high cover of tall and very tall shrubs, high composition was best explained by fire-related top-soil moisture levels and low and spatially changes to the vegetation structure and soil variable amounts of insolation at ground level. moisture levels. Burnt sub-plots were characterised by high levels For spiders, unburnt plots had more species of insolation at ground level, greater top-soil from the family Malkaridae (2 vs 0), while burnt hardness and greater cover of the herb & shrub plots had greater numbers of species from the component of the understorey vegetation. The families Zodariidae (9 vs 2), Gnaphosidae (6 vs 3), Reduviidae (Assassin Bugs) are predacious on Corinnidae (7 vs 4), Linyphiidae (4 vs 2) and other invertebrates and found generally on Lycosidae (3 vs 0). The Malkaridae are moist vegetation and on the ground. There is no clear habitat specialists, commonly dwelling in the leaf explanation as to their absence from unburnt litter (M.Gray pers. com.). The Lycosidae (wolf forest and they may be useful disturbance spiders) and Zorariidae are small to large, ground “indicators” (see 4.2.5). living, hunting spiders. Species from the For flies, unburnt plots had more species Zorariidae are vagrant hunters, frequenting areas from the families Sciaridae (8 vs 5), Phoridae (16 with an open vegetation structure and low litter vs 13) and Empididae (6 vs 3). The Sciaridae levels. The Gnaphosidae and Corinnidae are both (black fungus gnats) are often associated with generalised hunters, tolerant of drier conditions. decaying material, with their larvae are often The Linyphiidae (tent spiders) are good found in rotting vegetable matter or highly colonisers of disturbed habitats and can “balloon- organic soils (CSIRO 1991). The Phoridae in” from shrubs some 50m distant (M.Gray pers. (humpbacked flies) are active scavengers on com.). The observed substantially lower amounts foliage and litter, with the larvae generally of leaf litter and changes to its spatial distribution scavengers in carrion and other decomposing following frequent burning, and lower top-soil matter. Adults generally oviposit in carrion and moisture levels, would appear to offer an organic material on the ground (D.Bickel pers. explanation to the reduction in the number of com.) All the morphospecies of Empididae found species from the family Malkaridae, and the in this study were from the sub-family increase in species from the families Zodariidae, Tachydromiinae, which are mainly terrestrial and Gnaphosidae, Corinnidae, Linyphiidae and rarely fly. As adults they are generally predacious Lycosidae. Results of the CCA ordination support on smaller arthropods and frequent moist places, these conclusions with unburnt sub-plots commonly amongst vegetation. The larvae are characterised by high levels of litter, high cover of probably predacious, living in the soil and within tall and very tall shrubs, high top-soil moisture decaying vegetation and the leaf litter (CSIRO levels and low and spatially variable amounts of 1991). The observed substantially lower amounts insolation at ground level. Similarly, burnt sub- of leaf litter and changes to its spatial distribution plots were characterised by high levels of

249 Australia’s Biodiveristy - Responses to Fire

insolation at ground level and greater cover of the following frequent burning, and lower top-soil herb & shrub component of the understorey moisture levels, would appear to offer an vegetation. As a number of Corinnidae species are explanation to the reduction in the number of ant specialists, the substantial increase in ant species from the genera Cerapachys and abundance on frequently burnt sites may have also Hypoponera. Similarly, the structurally simplified contributed to the increased richness of this environment found on frequently burnt areas family through increased prey availability. would provide suitable habitats for large, solitary For beetles, unburnt plots had more species foragers such as Colobostruma spp. Results of the from the family Carabidae (11 vs 8), while burnt CCA ordination support these conclusions with plots had greater numbers of species from the unburnt sub-plots characterised by high levels of families Curculionidae (18 vs 9) and litter, high cover of tall and very tall shrubs, high Chrysomelidae (8 vs 4). The Carabidae (Ground top-soil moisture levels and low and spatially Beetles) are mainly predatory, both as adults and variable amounts of insolation at ground level. larvae, on plant-inhabiting insects. The Similarly, burnt sub-plots were characterised by Curculionidae (Weevils), as adults, feed on the high levels of insolation at ground level and stems, roots, seeds and fruits of plants, with larvae greater cover of the herb & shrub component of usually feeding on wood and other plant parts. the understorey vegetation. The Chrysomelidae (Leaf Beetles) feed on leaves Across the five groups studied in detail; bugs, and other (living) vegetative parts of plants, both flies, spiders beetles and ants, there was a as larvae and adults. The richness of the beetle consistent pattern with regard to the changes that fauna here would appear to reduce the ability to occur to community composition following generalise at the level of family. Results of the frequent burning. Although species richness (α- CCA ordination suggest a reasonable separation diversity) decreased by 44% and 27% for flies and of assemblages on burnt and unburnt plots, with beetles, and increased by 16%, 27% and 26% for unburnt sub-plots characterised by high levels of bugs, spiders and ants (respectively), all groups litter, high cover of tall and very tall shrubs, high experienced a loss of species with frequent top-soil moisture levels and low and spatially burning. The results from this study suggest that variable amounts of insolation at ground level. frequent burning had led to the loss of up to 131 Similarly, burnt sub-plots were characterised by species (18 bugs, 31 flies, 15 spiders, 47 beetles high levels of insolation at ground level. It is and 20 ants), which represents 47% of the reasonable to suggest that ground-dwelling morphospecies known from the unburnt areas. species from the Carabidae are influenced by The losses were disproportionate across the substantially lower amounts of leaf litter and groups with percentage reductions ranging from changes to its spatial distribution following 41% (bugs), to 40% (flies), 34% (beetles), 24% frequent burning. Changes to the structure of the (spiders) and 23% (ants). The species lost would vegetation community with frequent burning appear to be from those groups dependent upon a appear to have provided additional habitats for substantial litter layer and stable moist conditions. plant-dependent species from the families The overall biodiversity of frequently burnt areas Curculionidae and Chrysomelidae. was maintained by the addition of species not For ants, unburnt plots had more species recorded on unburnt plots. The changed from the genera Cerapachys (7 vs 1) and Hypoponera environment was supporting an additional 133 (4 vs 1), while burnt plots had greater numbers of morphospecies (19 bugs, 11 flies, 32 spiders, 53 species from the genus Colobostruma (2 vs 0). beetles and 18 ants). These species would appear Species of Cerapachys are considered by Andersen to have broad tolerances, or adaptations, to drier (1990) to be “climate specialists” and are known to and more open environments. Overall, the be specialist predators, often on other ants composition of terrestrial invertebrate (Holldobler and Wilson 1990). Species from the communities was therefore influenced by a genus Hypoponera are cryptic, nesting and foraging combination of site-dependent (slope and aspect) within the soil and leaf litter. Members of the and treatment-dependent (litter, insolation, herb genus Colobostruma are large solitary & shrub cover, top-soil moisture & hardness) foragers/specialist predators (Andersen 1990). environmental variables. The observed substantially lower amounts of leaf This research has also shown that patterns of litter and changes to its spatial distribution relative abundance and species richness are not

250 Bushfire and forest invertebrates

concordant between broad taxa. Analysis of increased for spiders, but appeared unchanged for community composition, as illustrated by the flies. For bugs, beetles and ants it would appear CCA ordination, also illustrates that patterns of that the degree of heterogeneity of the vegetation invertebrate community organisation show varied near the ground plays a role in maintaining responses to environmental disturbance. In the bi- biodiversity, probably by providing additional plots derived from the CCA ordination, the habitats and an associated increase in food and degree of clustering of the sub-plots from each other resources. Structural heterogeneity may also treatment indicates the relative similarity of imply a greater diversity of plant species, or of species assemblages on sub-plots and plots within growth stages of existing species. Increased levels each treatment. The tighter clustering of burnt of shading and changes in surface and nest sub-plots for bugs, beetles and ants indicated a temperature with increased vegetation cover has lower within-treatment diversity compared to been shown to reduce ant species richness unburnt sub-plots (ie. a lower β-diversity). The (Goldstein 1975, Greenslade and Mott 1979), and converse applied for spiders, with the tighter alter community composition as the clustering of unburnt sub-plots indicating a lower environmental conditions become sub-optimal for β-diversity compared to burnt sub-plots. The certain species (Welch 1978, Elmes and Wardlaw situation for flies indicated similar within- 1982, York 1994). Moisture, light and temperature treatment diversity for both treatments, with loose have been identified as important factors in the clustering of both unburnt and burnt sub-plots. determination of the composition of invertebrate This interaction between point richness (α- assemblages (Huhta et al. 1967, Punttila et al. diversity) and spatial “turnover” of species (β- 1991, McIver et al. 1992). Frequent burning led to diversity) has substantial implications for the greater spatial heterogeneity in the bark and twig interpretation of the apparent effect of repeated components of the litter layer, with a likely impact burning (see “scale effects” in Section 3.3.3.6). on its structural complexity. The composition of Similar patterns were apparent for bugs and ants spider communities has previously been shown to where richness values (α-diversity) on burnt sub- be influenced by the structure of the litter layer plots were on average higher than on unburnt following fire (Huhta 1971). The implication of sub-plots, however the high similarity of these changes for community organisation and assemblages on burnt sub-plots (low β-diversity) ecosystem function are considered in the meant that the overall richness of both treatments following section. were similar. Unburnt sub-plots had lower richness (α-diversity) but are less similar, resulting 4.2.4 Community Structure in higher “turnover” between sub-plots (higher β- The biological structure of a community involves diversity), increasing overall species richness for species composition and abundance, temporal that treatment. For flies and beetles burnt plots changes in communities, and the relationships had lower richness (α-diversity) but were less between species in communities. This in turn similar, resulting in higher β-diversity. Spiders exerts strong influences on the functioning of the exhibited a different pattern with both higher α- community, in other words, how the community and β-diversity for the burnt treatment, resulting works as a processor of energy and nutrients in a large number of species (32) unique to (Krebs 1985). The continued functioning of frequently burnt sites. communities and their ecological processes is a An explanation for the different levels of primary goal of ecologically sustainable species turnover (β-diversity) for the different management (ESDWG-Forest Use 1991). groups is likely to be found in patterns of While it is possible to describe and assess environmental heterogeneity and its effect on communities using indices such as species richness, micro-habitat diversity. Frequently burnt areas or to compare the relative abundance of species were shown to have more spatially homogeneous using similarity indices, multi-variate approaches ground herb, small shrub, tall and very tall shrub and/or through graphical representation, these layers, more spatially heterogeneous levels of provide little information concerning the processes insolation at ground level, and more spatially underlying these differences or any indication as to heterogeneous bark and twig components of the the relative sensitivity of species to landscape litter layer. The β-diversity of bug, beetle and ant change (Samways 1994). In order to simplify and communities was reduced with frequent burning, interpret the complexity of ecological systems, one

251 Australia’s Biodiveristy - Responses to Fire

approach has been to group species into “guilds” dung. Among the beetles, the Leiodidae are or “functional groups”. These groups recognise abundant in decaying organic matter and also the ecological rather than the taxonomic affinity of occur in carrion and fungal fruiting bodies. Many species. To test the applicability of this approach, are general scavengers, but certain groups are morphospecies (see Oliver and Beattie 1993) were associated with particular fungi (Lawrence and allocated to guilds based upon their feeding Britton 1994). The Ptiliidae are minute beetles strategies and habitat preferences, following which are relatively abundant in decaying organic reference to the relevant literature and discussions matter, including leaf litter, where their major with taxonomic experts. By classifying large food source appears to be fungal spores and numbers of species into smaller, more manageable hyphae. The Endomychidae feed on a variety of groups, it is possible to substantially reduce the fungi, with many occurring in leaf litter in moist apparent complexity of ecological systems and habitats. provide a basis for evaluating environmental For flies, the number of generalists and change (Andersen 1990). scavengers decreased slightly (11%), while for For all groups studied, frequent burning beetles the proportion of generalists increased, on resulted in a change in the structure of the average, by 79%, primarily through an increase in community. With regard to feeding strategy, there morphospecies from the family Scarabaeidae. was on average, a 15, 140 and 250% increase Scarab beetles always live in concealed habitats, (respectively) in the number of phytophagous feeding on roots, dung or decaying vegetable (plant feeding) species of bugs, flies and beetles. matter (McQuillan 1985). The reason for their With regard to the proportion of predator species, dramatic increase here is not apparent, although it remained unchanged for flies and bugs, however with their generally large size, their mobility may for bugs there was a total shift from the family be enhanced in the more open environment of Nabidae to the family Reduviidae. The Nabidae frequently burnt areas. With ants, the proportion are a family of predacious bugs whose eggs are of generalist morphospecies decreased, on oviposited into grass stems (CSIRO 1991). This average, by 11% following frequent burning. suggests that a change in this component of the Results were not consistent within the group with vegetation may be influencing the suitability of the genera Pheidole and Crematogaster decreasing the habitat for these species, which are then by 35 and 19% respectively, and the genus replaced by another group of predators from a Monomorium increasing by 388%. In these dry different family. With beetles there was, on forests, species from the genus Monomorium are average, an 18% decrease in the number of major seed predators (Andersen 1985, Andersen predator species, due primarily to a 54% and Ashton 1985), and may be responding to reduction in the number of morphospecies from changes in the vegetation composition that the family Staphylinidae. Most species from this accompany a frequent fire regime (Zedler et al. family are small, often cryptic and live as 1983, Nieuwenhuis 1987, Cary and Morrison predators hidden in soil and leaf litter. Others are 1995). For flies, the number of morphospecies associated with dung, carrion or fungi (Zborowski regarded as wide-ranging “tourists” was, on and Storey 1995). It is anticipated that the average, reduced by 43% on frequently burnt reduction in litter levels and decrease in soil plots. moisture associated with frequent burning would When groups were compared with regard to disadvantage this group. habitat preferences, it was apparent that there For groups feeding primarily on fungal were substantial changes in community structure products there was, on average, a 44% (flies) and for some taxa. For bugs, numbers of moist habitat 31% (beetles) decrease in the number of specialists from the sub-order Dipsocoromorpha morphospecies. For flies this was most marked in were, on average, reduced by 83%. Amongst the families Scaridae and Scatopsidae, and for spiders, moist habitat specialists were reduced by beetles, in the families Leiodidae, Ptiliidae & 88%, primarily due to a 90-95% decrease in Endomychidae. Species from the family Scaridae numbers of morphospecies from the families oviposit in fungi in the soil and also feed on fungal Theridiidae, Toxopidae and Oonopidae. The products, while the Scatopsidae are generally Theridiidae build their webs in leaf litter, the found in moist forest environments with their Toxopidae are a moist-adapted group, and the larvae occurring in rotting vegetable matter and Oonopidae mostly inhabit the litter in moist

252 Bushfire and forest invertebrates

environments. With the flies, the number of moist there was a substantial increase overall (30%) in habitat specialists remained similar, although the the average proportion of “opportunist” family Ceratopogonidae was more commonly morphospecies, due primarily to a single species: represented on burnt sub-plots and the family Rhytidoponera metallica, which was extremely Chironomidae on unburnt sub-plots. The numerous on burnt sub-plots. R. metallica is a well proportion of temperature-dependent (climate) known coloniser of disturbed habitats (Yeatman specialists among the ants remained largely stable and Greenslade 1980). (increased by 4%), although individual genera did It has been observed that the structure of ant vary in their response. For spiders, there was, on communities, in particular, may be influenced by average, and an 35% increase in the number of the relative abundance of particular “dominant” species with a known preference for dry habitats. and “sub-dominant’ groups (Fox & Fox 1982, For groups primarily inhabiting the litter Andersen 1990). Following frequent fire, the layer there was a variety of responses to frequent number of morphospecies within the “dominant” burning. For flies, the number of morphospecies functional group had, on average, decreased by specifically utilising the litter layer had, on 23%, primarily through a reduction in the average, decreased by 60%. This was primarily occurrence of Iridomyrmex species. The number of due to the absence of the family Tipulidae on “sub-dominants” had increased slightly (25%), frequently burnt sub-plots. Tipulidae (crane flies) largely because of an increase in the occurrence of use moist soil for breeding, with their larvae Camponotus species. The number of larger common in decaying vegetation, and therefore “solitary/specialist” morphospecies, which interact would be disadvantaged by frequent burning. only slightly with other groups, had increased With ants there was a 19% decrease in the slightly (13%), although the results were quite average number of “cryptic” morphospecies variable within the group. Three new genera were inhabiting the litter and soil, particularly with found on burnt plots (Colobostruma, Epopostruma & regard to the genera Hypoponera (90%) and Leptogenys) while the numbers of species of Tapinoma (86%). Species from the genus Myrmecia increased, on average, by 87%. Hypoponera are specialist predators, feeding largely Myrmecia have been shown to be more common on Collembola which were approximately 15% in recently burnt habitats in these forests (York less abundant on frequently burnt sites. Tapinoma 1994, 1996). The occurrence of a solitary forager, species are cryptic omnivores in the litter, with Bothroponera sp.A, had decreased substantially some arboreal nesters. Habitat availability for this (76%). Bothroponera are predacious, often on genus would be reduced by frequent burning. For termites (Holldobler and Wilson 1990), a group spiders however, the number of litter dwelling of organisms not found on frequently burnt sites species increased, on average, by 110% with an in this study. While a number of mounds equivalent increase from the families Hahniidae were observed on these areas, they did not appear and Textricellidae. Species from the family to be active. Hahniidae are small spiders that construct small sheet webs in litter and foliage. It would appear 4.2.5 Biodiversity Indicators that they are not disadvantaged by lower litter Programs in land appraisal and applied resource levels (M.Gray pers. com.). Spiders in the family management increasingly utilise "environmental Textricellidae are also very small and live deep in indicators" to facilitate and simplify assessment the litter layer near the litter/soil interface. As and decision-making procedures. Indicators may only 46-73% of litter is removed in each fire event take the form of an index which concisely in these forests (York 1996), this group may not be summarises some property of the system, such as disadvantaged by frequent fire. abundance or species richness (diversity); or Many groups of terrestrial invertebrates are describes the community via its species adapted to exploit disturbed habitats. For spider composition, the relative abundance of individuals morphospecies known to prefer open and within constituent species (evenness etc), or its disturbed habitats, the number of morphospecies organisation or “structure” (eg. number of guilds increased, on average, by over 600% following or functional groups). An alternative (or frequent burning, This was due primarily to the complimentary) approach may be to use occurrence of seven species from the family “indicator taxa”: an organism (or group of Zodariidae only on burnt sub-plots. For the ants organisms) that reveals important aspects of the

253 Australia’s Biodiveristy - Responses to Fire

structure and function for some part of the assessment of the impact of fire (see Campbell and ecosystem without exhaustive study of that part Tanton 1981; Majer 1984; Friend 1996). Temporal (Cornaby 1977). variability following single fires has been shown to A wide range of terrestrial invertebrates have be substantial, with taxa responding to seasonal and been used as “indicators” in Australia. Examples meteorological cues (see Neumann 1992; Coy include spiders (Mawson 1986), springtails 1996). Where sampling is required for comparative (Greenslade 1984, 1985; Greenslade and purposes only, for example burnt/unburnt, Greenslade 1987), termites (Nichols and Bunn logged/unlogged, rehabilitated/not rehabilitated 1980; Greenslade 1985), beetles (Greenslade contrasts, then the influence of temporal variability 1985; Yen 1987) and ants (Weir 1978; Majer can be reduced (controlled for) by simultaneous 1980b, 1984, 1985; Whelan et al. 1980; Yeatman sampling in the various treatment categories under and Greenslade 1980; Majer et al. 1982, 1984; examination (see Yeatman and Greenslade 1980; Andersen and McKaige 1987). Arthropods have a Majer et al. 1984; Burbidge et al. 1992). York significant role in the forest community, affecting (1994) showed that estimates of ant species richness both primary production by their grazing from a single summer sample of chronosequence activities, and the turnover of nutrients in their sites provided an accurate representation of long- role as decomposers (Lowman 1982). Forest floor term changes over time at a single site. Oliver and arthropods regulate microfloral decomposer Beattie (1996b) demonstrated that the richness of populations by their feeding on bacteria and ants recorded from a single summer pitfall sample fungal colonies, their transport of spores, and by was significantly correlated with richness values the contribution of their faeces and bodies for assessed by other sampling methods and seasons of decomposition (Van der Drift 1958, Engelmann sampling. This research therefore utilised a single 1961, MacFadyen 1962). Their requirements of a summer sample to assess the implications of spatial source of cover and food has led arthropods to variation in α- and β-diversity for investigations develop a sensitivity and responsiveness to system into frequent disturbance. structure which has made them useful indicators In this study, estimates of abundance were of system status and condition (Mattson 1977). shown to be influenced, not only by fire This research used two complimentary (treatment), but by large-scale spatial effects approaches to investigate the applicability of (position), with frequent interactions between invertebrate indicators for environmental impact these two factors. This meant that the particular assessment. This was undertaken firstly by looking fire effect (or lack thereof) was not always at spatial characteristics of commonly used consistent across the spatial range of sites biodiversity indices (abundance & richness) and (replicates). These interactions were significant (ie. by examining the concordance of these indices not as a consequence of random variation) for between taxa. It is often postulated that one group spiders, isopods (slaters), flies, beetles and ants. At of invertebrates may act as an indicator or times spatial patterns were evident, potentially in “umbrella” group for others, thereby allowing response to underlying environmental trends. This inferences to be made on the impact of was the case for ticks & mites and beetles. disturbance regimes. This assertion however relies Similarly, estimates of richness (α-diversity) were on as yet untested assumptions which flow from shown to be influenced by large-scale spatial particular sampling strategies, in particular, spatial effects (position) for flies and spiders, and patterns in species richness (α-diversity) and interactions between treatment and position effects species turnover (β-diversity). The second (spiders, beetles and ants). These patterns, approach involved a study of the composition and combined with the often considerable spatial structure of these communities in order to variability exhibited within a single treatment, identify species, groups of species, or community suggests that large sample sizes would be required “descriptors” which were useful as a means of to detect disturbance effects when using coarse- assessing potential effects on ecosystem function. scale taxonomic classification (eg Family or This was undertaken to determine the Order). Neumann (1992) recognised that while applicability of these methods for assessing such a broad taxonomic approach gives a ecological sustainability. cumulative estimate of the responses of the species High levels of spatial and temporal variability within each taxon, the behaviour of individual in invertebrate populations severely complicate an species remain unknown, thereby limiting the level

254 Bushfire and forest invertebrates

of interpretation that could be applied to the sustainable management. When differences in results. From an analysis of species-level data, this composition between burnt and unburnt areas research has shown those limitations to be were compared, it was apparent that frequent substantial when dealing with biodiversity issues. burning had led to a marked decline in species The concordance between the species with particular habitat preferences. The groups richness (α-diversity) of selected taxa was which were indicative of this decline were the investigated here using correlation analyses. infra-order Dipsocoromorpha and the families While a number of statistically significant Sciaridae, Phoridae and Empidae (flies), the correlations were detected, the predictive power families Malkaridae (spiders), Carabidae (beetles), of these relationships was weak. In addition it was and for ants, the genera Cerapachys and found that apparently significant correlations Hypoponera. Conversely, a number of groups were detected when using the whole dataset (spiders & “disturbance indicators” through marked increases ants - positive, flies & ants - negative, flies & in their richness on frequently burnt areas. These beetles - positive) did not hold when the two were the families Reduvidae (bugs), Zorariidae, treatments (burnt and unburnt) were analysed Gnaphosidae, Corinnidae, Linyphiidae and separately. It was evident that the apparent Lycosidae (spiders), Curculionidae and correlations between these taxa were an artefact of Chrysomelidae (beetles), and for ants, the genus the differences between average richness values on Colobostruma. The habitat parameters most burnt and unburnt sub-plots. Therefore, none of consistently associated with these changes were these groups of terrestrial invertebrates would be the amount and spatial distribution of leaf litter, a reliable “indicator” or “umbrella” group for any top-soil moisture levels, and the amount and other, and therefore are inappropriate for spatial heterogeneity of insolation levels at the predicting overall biodiversity at this scale. This ground surface. These environmental factors were supports the findings of Oliver and Beattie shown to vary primarily as a response to frequent (1996b) who demonstrated non-concordance burning, that is, under management control. between richness estimates for ants, beetles and Accompanying these habitat-related changes spiders. in community composition were a number of Given the identified influence of sampling shifts in ecological structure and function. For “scale” and spatial patterns of α- & β-diversity on some taxa the guild structure was generally estimates of species richness (see 4.2.2), maintained, but with changes to the suite of concordance between biodiversity estimates for species comprising each guild. This was most selected taxa were examined at a larger scale (1 marked with the ants, where the functional group hectare). While two statistically significant structure remained largely unaltered, but relationships were identified (ants & bugs - community composition changed substantially. negative, beetles & spiders - negative), these were This suggests that ant communities are resilient to only apparent on unburnt plots. These results fire-related disturbances (although individual suggest that, at least for these 2 pairs of taxa, and species may be fire sensitive). A similar pattern at the 1ha sampling scale, it may be possible and was evident in the bugs, with a shift within the reliable to use one group as an indicator of the phytophagous (plant-eating) guild from the biodiversity of the other. The fact that disturbance Coccidae, Fulgoridae and Homoptera to the changes the nature of these relationships casts Cicadellidae; and within the predacious guild doubt however on the value of this in studies of from Nabidae to Reduvidae. With the flies impact assessment. however there was a marked reduction (44%) in It was therefore considered that a second fungal feeders (Sciaridae, Drosophilidae, approach, a study of the composition and Mycetophilidae and Scatopsidae) and a substantial structure of these communities, may prove more increase (140%) in phytophagous species useful. If species, or groups of species, could be (primarily Cecidomyiidae), resulting in a found that typified the response of whole taxa, considerable change in community structure. The then these may serve as useful “bio-indicators”. beetles were similarly affected with a 31% An assessment of the effects of disturbance on reduction in the richness of the fungal-feeding ecosystem function would then be feasible, guild (primarily Leionidae), and a 250% increase providing a valuable tool for monitoring progress in richness of phytophagous species (primarily towards, and compliance with, ecologically Chrysomelidae and Curculionidae). The loss of

255 Australia’s Biodiveristy - Responses to Fire

species associated with the decomposer cycle given that they cannot be unambiguously defined implies the frequent burning may be impacting (Caughley and Gunn 1996). It is recognised that a upon nutrient cycling and transfer within these modern conservation strategy cannot however be forests. based solely on areas managed purely for While grouping species into guilds or conservation purposes. Forests (and other functional groups does achieve its intended aim of environments) outside the reserve system will simplifying complex systems, it is at the expense of continue to play an important and complementary considerable important detail. Due to the richness role in meeting conservation objectives with of the communities involved, and the associated respect to biodiversity, even though these forests taxonomic difficulties, this research adopted a may be available for the production of timber and “morphospecies” approach where the biology and other commercial uses in an ecologically ecology of individual species was not examined. sustainable way. York (1996) has shown that, for ants, a knowledge This research has indicated that, through the of the ecology of the constituent species at the development of appropriate management level of genus enables considerable insight into the strategies, we have the knowledge to ensure that mechanisms of change following disturbance. biodiversity is adequately conserved. Strategies Given the aims of this study, the morphospecies required to conserve invertebrate biodiversity are (RBA) approach has however proved to be fundamentally consistent with those used to successful. The limitations of this approach are protect other groups, although the emphasis is acknowledged, and it is considered that additional likely to be placed on the protection of habitats insight could be gained by a more detailed rather than individual species or assemblages (see ecological investigation of individual taxa. This is Samways 1994; New 1995). With regard to proposed in a series of future studies. hazard-reduction burning, the extensive and The limitations of grouping species into frequent use of this management practice has the taxonomic or ecological units are particularly potential to substantially reduce regional evident in the area of conservation biology where biodiversity. The development of strategies which a “species-by-species” management approach may set aside (unburnt) refuges, maintain a diversity of be required, with the ecological needs of each habitats at various stages in the post-fire species addressed separately in Plans of succession, permit variability in other components Management. The incredible diversity of of the fire regime (season of burn and intensity), terrestrial invertebrates may preclude this strategy and allow connectivity between different habitats in Australian forests, with reserve strategies being (corridors), requires urgent attention. Current developed based primarily on vascular plants and draft strategies in forest zoning (eg. SFNSW vertebrates. Whether a conservation strategy 1995) are currently addressing some of these developed using this approach is appropriate for issues, however it is of some concern that other the conservation of invertebrates is questionable, land management agencies still advocate frequent given the lack of concordance between the “broad-acre” burning as a panacea for hazard richness and composition of these very disparate reduction and protection from wildfire without groups of organisms (see Oliver 1995). Similarly, adequate consideration of important biodiversity the use of “ecosystems” or “communities” as issues. ecological units of conservation is problematic

256 Bushfire and forest invertebrates

5. CONCLUSIONS

Infrequent, periodic forest fires (bushfires) are an substantial changes in the composition of species integral part of the modern physical environment assemblages following frequent disturbance of Australian sclerophyll forests. The inherent however, with a loss of taxa dependent upon a variability in natural fire regimes generally results substantial litter layer and stable moist conditions. in a mosaic of habitats with vegetation at different The overall diversity of frequently burnt areas was stages of floristic and structural post-fire maintained by the addition of species with broad succession, each potentially supporting particular tolerances, or adaptations, to drier and more open animal communities. Changes to the components environments. Shifts in community composition of the fire regime (fire intensity, frequency and were best explained by the changes in the amount season of occurrence), as a consequence of forest of leaf litter and insolation at the ground surface, management practices, have the potential to alter habitat elements shown to be dramatically the composition and structure of natural modified by frequent burning. This suggests that communities. The research reported here deals the extensive application of this management with the impact of frequent low-intensity fire practice could result in a reduction in terrestrial (“hazard-reduction burning”) on the abundance, invertebrate biodiversity at a regional scale (γ- richness, composition and structure of terrestrial diversity), with this decrease potentially as high as invertebrate communities. This group, which 50%. constitutes a major component of the overall It was demonstrated that considerable biodiversity in these forests, plays a substantial additional detail concerning, and insight into, the role in the maintenance of ecosystem processes. nature of these changes could be provided by the The ability of the Australian forest industry to inclusion of fairly general information concerning achieve Ecologically Sustainable Management the habitat and dietary preferences of the groups (ESM) depends therefore on a better under investigation. It was apparent that frequent understanding of the impact of commonly used burning leads to a change in the structure of the management strategies on this important invertebrate community. Within species component of the ecosystem. assemblages there were shifts in feeding strategy, While a number of habitat components were with substantial increases in the proportion of responding to large-scale environmental patterns, phytophagous species for bugs, flies and beetles, frequent burning was shown to be impacting upon and a reduction in fly and beetles groups reliant the amount, structure and spatial distribution of primarily on fungal products. With regard to surface leaf litter, the structure and spatial predator guilds, there was a substantial decrease in heterogeneity of components of the vegetation proportional representation for beetles, primarily understorey, moisture levels in the top-soil and the in relation to the family Staphylinidae. For bugs amount and patchiness of insolation reaching the however, the proportion of predators remained ground. While some aspects of these changes are largely unchanged, however there was a total shift likely to reflect post-fire successional trends, the from the family Nabidae to the family Reduviidae. truncation of successional patterns by frequent fire Within one major group of predators, the spiders, and subsequent reduction in environmental there was a substantial reduction in the number of heterogeneity can be expected to impact upon moist-habitat and leaf litter specialists, and a terrestrial invertebrate communities. dramatic increase in the number of species known Using ants, beetles, flies, spiders & bugs as to prefer dry and open environments, particularly representative groups and potential indicators of from the family Zorariidae. Similar patterns were environmental degradation, this research exhibited by the ants, with changes in functional demonstrated that although overall species group representation in response to habitat richness (α-diversity) may not change with alteration. While the impact of these changes on frequent disturbance, species turnover (β- ecosystem function was beyond the scope of this diversity) does. There is a lack of concordance study, substantial changes in the structure of however between groups in the magnitude and invertebrate assemblages and the loss of species direction of these responses. All groups showed associated with the decomposer cycle implies

257 Australia’s Biodiveristy - Responses to Fire

frequent burning may be impacting upon nutrient The application of Rapid Biodiversity Assessment cycling and transfer within these forests. If this is (RBA) methodology demonstrated that the study the case, it would have serious implications with of the composition and structure of communities regard to the maintenance of ecological is likely to prove more rewarding in this regard. sustainability. The identification of individuals to distinct In New South Wales, State Forests has, as a “morphospecies”, while requiring additional stated objective of its Corporate Plan (1992), that it laboratory time and taxonomic expertise, will achieve ecologically sustainable management facilitated the incorporation of broad-level (ESM) by refining concepts and developing ecological information into the assessment and measurable indicators of ecologically sustainable interpretation of environmental impact. use. To be useful therefore, ESM indicators need to Information currently available at the level of be interpretable, significant, cost efficient, and need Family or Genus, but unavailable at the level of to account for variability in space and time, and be species, was sufficient to enable a meaningful appropriate for the scale of management (Turner interpretation of data in relation to impacts on 1993, York 1993). The research reported here community structure and ecological function. supports previous findings concerning the high This in turn enabled the development of spatial variability of invertebrate population management recommendations consistent with numbers, and confirmed the limited use of data the conservation of biological diversity. obtained using coarse-scale taxonomic classification In studies such as this, the accurate (eg. Family or Order). The cost-effectiveness of assessment of the impact of long-term frequent using abundance data alone was shown to be low, disturbance is potentially confounded by short- with high spatial variability and spatial patterning term responses to the most recent perturbation. It requiring large sample sizes to detect management is likely that the rate of post-fire successional effects. This research also identified substantial change will be greater on the frequently disturbed limitations with regard to the use of a single index, plots compared to the unburnt (control) plots, and species richness, as an measure of change and/or the results obtained (and hence assessment of environmental impact. Species richness (α- impact) will be, to some extent, dependent upon diversity) is frequently used to describe and the current successional stage (see Huhta 1971; compare communities, however in this case it was Merrett 1976; York 1994, 1996). In this study the found to provide a deceptive summary of disturbed ares were sampled 2 years after fire and community characteristics. The lack of at a time when there was sufficient fuel to support concordance of richness and abundance patterns another low-intensity fire. With regard to habitat within and between taxa, and the identified modification therefore, this study represents relationships between estimates of richness (α- (potentially) a “worse-case-scenario” typifying one diversity) and turnover (β-diversity) and the spatial end of the spectrum of responses to frequent fire. scale of measurement, meant that the use of these Nevertheless it typifies many areas of dry indices (alone) for impact assessment will sclerophyll forest which are regularly burnt for substantially restrict the level of interpretation that the purposes of hazard reduction. Of more can be derived from the data. Variable spatial concern is the fact that the experimental design patterns in these indices between disturbed and used here substantially “down-plays” the effect of undisturbed sites, and between taxa, also limits recolonisation ability, with potential refuge their applicability in impact assessment. (unburnt) areas no more than 20m from It is often postulated that one group of frequently burnt sites. In a managed forest invertebrates may act as an “indicator” or environment where fuel-reduction burning is “umbrella” group for others, thereby allowing often spatially extensive, the “habitat inferences to be made on the impact of fragmentation” effect is likely to be more disturbance regimes. In this project, it was shown pronounced. Given that many invertebrate species the lack of correlation between taxa with regard to within a community are of low abundance (see richness indices (α- and β-diversity) restricted York 1994) or are habitat or dietary specialists their utility in this regard, primarily as a (York 1996), the risk of local extinction is high. consequence of the non-concordant spatial While local extinction of invertebrates is likely to patterning of these community descriptors and be a regular occurrence in natural systems associated implications for sampling effectiveness. (Samways 1994), systems prone to

258 Bushfire and forest invertebrates

anthropomorphic disturbance require the strategies, from fire exclusion to frequent burning, establishment of adequate measures for the in-situ can be applied to an area in response to protection of successional stages and their management and conservation needs (see Ridley constituent invertebrate fauna (eg. refuges & 1993). Given that the various successional states reserves) and the establishment of links (ie can provide optimal habitats for certain species corridors) to facilitate recolonisation. This need to (York 1994), and that disturbance is an intrinsic provide undisturbed and secure refuges for species and necessary feature in most natural systems with specialist requirements and limited dispersal (Pickett and White 1985), broad-scale fire abilities is the same dilemma facing those exclusion is not a practical management and concerned with the conservation of flowering conservation option. With the increasing plants and vertebrates (see Caughley and Gunn awareness however of the importance of 1996), the difference however is that the groups maintaining environmental and biological involved may be responding to different diversity at a range of spatial scales, the environmental cues and strategies developed for development and implementation of appropriate one taxon may not adequately conserve others (see fire “regimes” which take into consideration the Oliver 1995). scale (ie. frequency and extent) of disturbance, is The development of a “Forest Zoning within the grasp of forest management agencies. System” in New South Wales (SFNSW 1995) should be seen as a necessary response to this situation, whereby a variety of management

259

Bushfire and forest invertebrates

REFERENCES

Abbott, I., Van Heurck, P. and Wong, L. (1984) Bradstock, R.A. and Myerscough, P.J. (1981) Fire Responses to long-term fire exclusion : effects on seed release and the emergence physical, chemical and faunal features of and establishment of seedlings in Banksia litter and soil in a Western Australian forest. ericifolia L.f. Aust. J. Bot. 29:521-532. Aust. For. 47:237-242. Brian, M.V., Mountford, M.D., Abbot, A. and Andersen, A.N. (1985) Seed predation by insects in Vincent, S. (1976) The changes in ant sclerophyllous vegetation at Wilson’s Promontory, species distribution during ten years post- Victoria, with particular reference to ants. fire regeneration of a heath. J. Anim. Ecol. Unpubl. PhD Thesis, School of Botany, 45:115-133. . 241 pp. Burbidge, A.H., Leicester, K., McDavitt, S. and Andersen, A.N. (1988) Immediate and longer- Majer, J.D. (1992) Ants as indicators of term effects of fire on seed predation by ants disturbance at Yanchep National Park, in sclerophyllous vegetation in south-eastern Western Australia. J.Royal Soc.West.Aust. Australia. Aust. J. Ecol. 13:285-293. 75:89-95. Andersen, A.N. (1990) The use of ant Campbell, A.J. (1973) The Effects of Prescribed communities to evaluate change in Burning on Surface Active Invertebrate Fauna Australian terrestrial ecosystems: a review in Pine and Eucalypt Forest Within the and a recipe. Proc. Ecol. Soc. Aust. 16, 347- Australian Capital Territory. BSc. (Hons) 357. Thesis. ANU. Canberra, ACT. Australia. Andersen, A.N. and Ashton, D.H. (1985) Rates of Campbell, A.J. and Tanton, M.T. (1981) Effect of seed removal by ants at heath and woodland fire on the invertebrate fauna of soil and sites in southeastern Australia. Aust.J.Ecol. litter of a eucalypt forest. In: Fire and the 10:381-390. Australian Biota. (Eds Gill, A.M., Groves, Andersen, A.N. and McKaige, M.E. (1987) Ant R.H. and Noble, I.R.). Australian Academy communities at Rotamah Island Victoria, of Science, Canberra. with particular reference to disturbance and Cary, G.J. and Morrison, D.A. (1995) Effects of Rhytidoponera tasmaniensis. Proc. Roy. Soc. Vict. fire frequency on plant species composition 99: 141-146. of sandstone communities in the Sydney Beattie, A.J. (1995) (ed.) Biodiversity, Australia’s region: Combinations of inter-fire intervals. Living Wealth. Research Unit for Aust.J.Ecol. 20:418-426. Biodiversity and Bioresources, School of Catling, PC. (1991) Ecological effects of Biological Sciences, Macquarie University, prescribed burning practices on the Australia. 131pp. mammals of southeastern Australia. In: Beattie, A.J. and Oliver, I. (1994) Taxonomic Conservation of Australia’s Forest Fauna. (ed minimilisation. Trends in Ecology and Lunney, D.), pp. 353-63. Royal Zoological Evolution 9:488-490. Society of NSW, Mosman. Benson, D.H. (1985) Maturation periods for fire Caughley, G. and Gunn, A. (1996) Conservation sensitive shrub species in Hawkesbury Biology in Theory and Practice. Blackwell sandstone vegetation. Cunninghamia 1:339- Science, Massachusetts USA. 459pp. 349. Christensen, N. (1985) Shrubland fire regimes Birk, E.M. and Bridges, R.G. (1989) Recurrent and their evolutionary consequences. Pp86- fires and fuel accumulation in even-aged 99 in S.T.A. Pickett and P.S.White (Eds) The blackbutt (Eucalyptus pilularis) forests. For. Ecology of Natural Disturbance and Patch Ecol. and Manage. 29:59-79. Dynamics. Academic Press, Orlando, Florida. Bornemissza, G.F. (1969) The re-invasion of burnt woodland areas by insects and mites. Proc. Ecol. Soc. Aust. 4:138.

261 Australia’s Biodiveristy - Responses to Fire

Commonwealth of Australia (1983) National Forestry Commission of NSW (1992) 1992-1995 Conservation Strategy for Australia. Living Corporate Plan. For.Com.NSW, Sydney. Resource Conservation for Sustainable 26pp. Development. Department of Home Affairs Fox, B.J. (1979) An objective method of and Environment. Aust.Govt.Pub.Service, measuring the vegetation structure of animal Canberra. habitats. Aust.Wildl.Res. 6:297-303. CONCOM (1989) Council of Nature Fox, M.D. and Fox, B.J. (1982) Evidence for Conservation Ministers - Australian interspecific competition influencing ant statement on invertebrates. Myrmecia 25, species diversity in regenerating heathland. 123-125. pp.99-110 in Buckley, R.C. (1982) (ed.) Ant- Cornaby, B.W. (1977) Saprophagous organisms Plant Interactions in Australia. Junk and problems in applied resource Publishers. The Hague. partitioning. pp. 96-100 in Mattson, W.J. Fox, M.D. and Fox, B.J. (1986) The effect of fire (ed) (1977) The Role of Arthropods in Forest frequency on the structure and floristic Ecosystems. Springer-Verlag, New York. composition of a woodland understorey. Coy, R. (1996) The impact of fire on soil Aust. J. Ecol. 11:77-85. invertebrates in E.regnans forest at Friend, G. (1996) Fire ecology of invertebrates - Powelltown, Victoria. In: Fire and implications for nature conservation, fire Biodiversity. The Effects and Effectiveness of Fire management and future research. In Fire and Management. pp. 183-198. Biodiversity Biodiversity. The Effects and Effectiveness of Fire Series, Paper No. 8. Biodiversity Unit, Management. Pp. 155-162. Biodiversity Department of the Environment Sport and Series, Paper No. 8. Biodiversity Unit, Territories. Canberra, Australia. Department of the Environment Sport and CSIRO (1991) The Insects of Australia. Second Territories. Canberra, Australia. Edition. Melbourne University Press. Friend, J.A. and Richardson, A.M.M. (1986) Dindal, D.L. and Metz, L.J. (1977) Community Biology of terrestrial amphipods. Ann. Rev. structure of Collembola affected by fire Entomol. 31: 25-48. frequency. pp.88-95 in The Role of Arthropods Gill, A.M. (1981) Adaptive responses of Australian in Forest Ecosystems. Mattson, W.J. (ed) vascular plant species to fires. In: Fire and the Springer-Verlag, New York. Australian Biota (eds. A.M. Gill, R.H. Groves Dolva, G. (1993) The effect of fire on the ecology and and I.R. Noble), pp.243-272. Aust. Acad. life-history of the wood cricket Nambungia Sci., Canberra. balyarta (Nemobiinae:Gryllidae). Unpublished Goldstein, E.L. (1975) Island biogeography of Masters Thesis, University of Western ants. Evolution 29:750-762. Australia. Greenslade, P. and Greenslade, P.J.M. (1987) Duncan, K.W. (1969) The ecology of two species Ecological strategies in collembola: a new of terrestrial Amphipoda (Crustaceae: approach to the use of terrestrial Family Tallitridae) living in waste grassland. invertebrates in environmental assessment. Pedobiologia 9:323-341. In: B.R. Striganova (Ed.), Soil Fauna and Soil Elmes, G.W. and Wardlaw, J.C. (1982) A fertility, pp.245-252. Nauka, Moscow. population study of the ants Myrmica Greenslade, P.J.M. (1984 ) Invertebrates and sabuletti and Myrmica scabrinodis, living at two environmental assessment. Environment and sites in the south of England. II. Effect of Planning 3:13-15. above-nest vegetation. J.Anim.Ecol. 51:665- Greenslade, P.J.M. (1985) Pterygote insects and 680. the soil: their diversity, their effects on soils Engelmann, M.D. (1961) The role of soil and the problem of species identification. arthropods in the energetics of an old field Quaest. Ent. 21:571-585. community. Ecol. Monogr. 31:221-238. ESDWG-Forest Use (1991) Ecologically Sustainable Development Working Groups. Final Report - Forest Use. Australian Government Publishing Service, Canberra. 227pp.

262 Bushfire and forest invertebrates

Greenslade, P.J.M. and Mott, J.J. (1979) Ants of Kremen, C. (1992) Assessing the indicator native and sown pastures in the Katherine properties of species assemblages for natural area, Northern Territory, Australia. areas monitoring. Ecological Applications (Hymenoptera:Formicidae). Pp.153-156 in 2:203-217. Crosby, T.K. and Pottinger, R.P. (eds) Lambert, M.J. (1982) Methods for Chemical proceedings of The 2nd Australasian Analysis. Forestry Commission of NSW. Conference on Grassland Invertebrate Ecology. Sydney, Australia. 3rd Edition. Palmerston North, New Zealand. 22-26th Lawrence, J.F. and Britton, E.B. (1994) Australian May, 1978. Wellington Govt. Printer. Beetles. Melbourne University Press. 192pp. Greenslade, P.J.M. and Thompson, C.H. (1981) Leonard, B.V. (1972) The effect of fire upon selected Ant distribution, vegetation and soil small mammals and leaf litter fauna in relationships in the Coolola-Noosa River sclerophyll forest in southern Australia. Unpubl. area, Queensland. In: Vegetation Classification MSc. Thesis. Monash University, Victoria, in Australia (eds A.N. Gillson & D.J. Australia. Anderson), pp. 192-207. CSIRO and Lowman, M.D. (1982) Seasonal variation in insect Australian National University Press, abundance among three Australian rain Canberra. forests, with particular reference to Harvey, M.S. and Yen, A.L. (1989) Worms to phytophagous types. Aust. J. Ecol. 7:353-361. Wasps. An Illustrated Guide to Australia’s MacArthur, R.H. and MacArthur, J.W. (1961) On Terrestrial Invertebrates. Oxford University bird species diversity. Ecology 42:594-598. Press. 203pp. MacFadyen, A. (1962) Soil sampling. Hobbs, R.J. (1992) Function of biodiversity in Adv. Ecol. Res. 1:1-34. Mediterranean ecosystems in Australia: MacQuillan, P.B. (1985) The identification of definitions and background. In: Biodiversity of root-feeding cockchafer larvae (Coleoptera: Mediterranean Ecosystems in Australia (ed. R.J. Scarabaeidae) found in pastures in Tasmania. Hobbs), pp. 1-25. Surrey Beatty & Sons. Aust.J.Zool. 33: 509-46. Holldobler, B. and Wilson, E.O. (1990) The Ants. Majer, J.D. (1980) Report on a study of Springer-Verlag, Berlin. 732pp. invertebrates in relation to the Kojonup Huhta, V. (1971) Succession in the spider Nature Reserve fire management plan. Bull. communities of the forest floor after clear- Dept. Biol. West. Aust. Inst. Tech. 2:1-22. cutting and prescribed burning. Ann. Zool. Majer, J.D. (1984) Ant return in rehabilitation Fennici. 8:483-542. mines - an indication of ecosystem resilience. Huhta, V.E., Karpinnen, E., Nurminen, M. and In : B.Dell (ed.), Medicos IV:Proceedings 4th Valpus, A. (1967) Effects of silvicultural International Conference on Mediterranean practices upon arthropod annelid and Ecosystems, pp.105-106, Botany Dept., nematode populations in coniferous forest University of Western Australia, Nedlands. soil. Anneles Zoologie Fennici 4:87-135. Majer, J.D. (1985) Recolonisation by ants of Hurlbert, S.H. (1984) Pseudoreplication and the rehabilitated mineral sand mines on north design of ecological field experiments. Ecol. Stradbroke Island, Queensland, with Monogr. 54:187-211. particular reference to seed removal. Aust. J. IUCN (1980) International Union for the Ecol. 10:31-48. Conservation of Nature, United Nations Majer, J.D. (1992) Ant recolonisation of Environment Program (UNEP) and World rehabilitated bauxite mines of Pocos de Wildlife Fund (WWF) - World Caldas, Brazil. J.Trop.Ecol. 8:97-108. Conservation Strategy - Living Resource Majer, J.D., Satori, M., Stone, R. and Perriman, Conservation for Sustainable Development. W.S. (1982) Recolonisation by ants and IUCN, Gland, Switzerland. other invertebrates in rehabilitated mineral Krebs, C.J. (1985) Ecology. The Experimental sand mines near Eneabba, Western Australia. Analysis of Distribution and Abundance. Third Reclaim. Reveg. Res. 1:63-81. Edition. Harper & Row, New York. 800pp. Majer, J.D., Day, J.E., Kabay, E.D. and Perriman, W.S. (1984) Recolonisation by ants in bauxite mines rehabilitated by a number of different methods. J. Appl. Ecol. 21:355-375.

263 Australia’s Biodiveristy - Responses to Fire

Margules, C.R., Nicholls, A.O. and Pressey, R.L. NFPS (1992) National Forest Policy Statement. A (1988) Selecting networks of reserves to New Focus for Australia’s Forests. maximise biological diversity. Biological Commonwealth of Australia. Canberra. Conservation 43:63-76. Nichols, O.G. and Bunn, S. (1980) Termite Mattson, W.J. (ed) (1977) The Role of Arthropods in utilisation of rehabilitated bauxite mined Forest Ecosystems. Springer-Verlag, New York. areas. Alcoa of Australia Limited. Environ. Mawson, P.R. (1986) A comparative study of Res. Bull.No.9. arachnid communities in rehabilitated Nieuwenhuis, A. (1987) The effect of fire bauxite mines. W.A.I.T. School of Biology frequency on the sclerophyll vegetation of Bulletin No.14. the West Head, New South Wales. McIver, J.D., Parsons, G.L. and Moldenke, A.R. Aust.J.Ecol. 12:373-385. (1992) Litter spider succession after clear- Noble, I.R. and Slatyer, R.O. (1981) Concepts and cutting in a western coniferous forest. models of succession in vascular plant Canad.J.For.Res. 22:984-992. communities subject to recurrent fire. In Fire Merrett, P. (1976) Changes in the ground-living and the Australian Biota (Gill, A.M., Groves, spider fauna after heathland fires in Dorset. R.H. and Noble, I.R. eds.), pp.311-315. Bull. Br. Arachnol. Soc. 3:214-221. Aust.Acad.Sci., Canberra. Metz, L.J. and Dindal, D.L. (1975) Collembola NSESD (1992) National Strategy for Ecologically populations and prescribed burning. Environ. Sustainable Development. Commonwealth Entomol. 4:583-587. of Australia. 128pp. Metz, L.J. and Farrier, M.H. (1973) Prescribed Oliver, I. (1995) Rapid Biodiversity Assessment and burning and populations of soil mesofauna. its Application to Invertebrate Conservation in Environ. Entomol. 2:433-440. Production Forests. Unpubl. PhD Thesis. Morris, M.G. (1975) Preliminary observations on School of Biological Sciences, Macquarie the effects of burning on the Hemiptera University, Australia. (Homoptera and Auchenorhyncha) of Oliver, I. and Beattie, A. J. (1993) A possible limestone grassland. Biol. Conserv. 7:312-319. method for the rapid assessment of Moulton, T.P. (1982) The effect of prescribed burning biodiversity. Cons.Biol. 7:562-568. and simulated burning on soil and litter Oliver, I. and Beattie, A. J. (1996a) Invertebrate arthropods in open forest at Cordeaux, N.S.W. morphospecies as surrogates for species: a Australia. Unpubl. PhD. Thesis. Macquarie case study. Cons. Biol. 10:99–109. University. Oliver, I. and Beattie, A. J. (1996b) Designing a Neumann, F.G. (1992) Effects of low intensity cost-effective invertebrate survey: a test of prescribed fire on invertebrates in mixed methods for rapid biodiversity assessment. eucalypt forest - Wombat State Forest. In Ecological Applications 6:594-607. Tolhurst, K. And Flinn, D. (eds.) Ecological Palmer, M.W. (1993) Putting things in even better Impacts of Fuel Reduction Burning in Dry order: The advantages of canonical Sclerophyll Forest: First Progress Report. correspondence analysis. Ecology 74:2215- Conservation and Environment, Victoria, 2230. Australia. Pickett, S. T. A. and White, P. S. (1985) (Eds) The Neumann, F.G. and Tolhurst, K. (1991) Effects of Ecology of Natural Disturbance and Patch fuel reduction burning on epigeal arthropods Dynamics. Academic Press, Orlando, Florida. and earthworms in dry sclerophyll eucalypt Punttila, P., Haila, Y., Payunen, T and Tukia H. forest of west-central Victoria. Aust. J. Ecol. (1991) Colonisation of clearcut forests by 16, 315-330. ants in the southern Finnish taiga: a New, T.R. (1984) Insect Conservation: An Australian quantitative survey. Oikos 61, 250-262. Perspective. Dr. W.Junk Publishers, Ridley, D. (1993) Fire fuel management in Eden Netherlands. State Forest District. Pp. 91-101 in Ross, J. New, T.R. (1995) Introduction to Invertebrate (Ed) The Burning Question: Fire Management Conservation Biology. Oxford University in NSW. Proceedings of a Conference held Press, New York. 194pp. at the Department of Continuing Education, University of New England, Armidale.

264 Bushfire and forest invertebrates

Samways, M. J. (1994) Insect Conservation Biology. WCED (1987) World Commission on Chapman & Hall, London. 358 pp. Environment and Development. Our Seastedt, T. R. (1984) Below-ground Common Future. Oxford University Press, macroarthropods of annually burned and Oxford, UK. 400pp. unburned tallgrass prairie. Am. Midl. Nat. Weir, E.G. (1978) Development of a Fire Policy. 111:405-408. Parks and Wildlife, Vol.2. National Parks Springett, J. A. (1976) The effect of prescribed and Wildlife Service of N.S.W. burning on the soil fauna and litter Welch, R.C. (1978) Changes in the distribution of decomposition in Western Australian forests. the nests of Formica rufa L. Aust.J.Ecol. 1:77-82. (Hymenoptera:Formicidae) at Brian Woods Springett, J. A. (1979) The effects of a single hot National Nature Reserve, Kent, during the summer fire on soil fauna and on litter decade following coppicing. Insectes Sociaux decomposition in jarrah (Eucalyptus 25:173-186. marginata) forest in Western Australia. Aust. Whelan, R.J., Langedyk, W. and Pashby, A.S. J. Ecol. 4:279-91. (1980) The effects of wildfire on arthropod SFNSW (1995) Forest Management Zoning in State populations in a Jarrah - Banksia woodland. Forests (Draft). State Forests of New South W.A. Naturalist 14:214-220. Wales. Sydney, Australia. Whelan, R.J. (1995) The Ecology of Fire. Taylor, R. W. (1983) Systematics and ecology: Cambridge University Press. 346pp. reflections from the interface. Pp50-79 in Williams, J.E. and Gill, A.M. (1995) The Impact of E.Highley and R.W. Taylor (eds) Australian Fire Regimes on Native Forests in Eastern New Systematic Entomology: A Bicentenary South Wales. Environmental Heritage Perspective. CSIRO, Melbourne. Monograph Series No. 2. NSW National Ter Braak, C. J. F. (1986) Canonical Parks and Wildlife Service. 68pp. Correspondence Analysis: a new eigenvector Woodward, F.I. (1993) How many species are technique for multivariate direct gradient required for a functioning ecosystem? In analysis. Ecology 67:1167-1179. Ecosystem Function of Biodiversity. Eds. E.D. Ter Braak, C. J. F. (1991) CANOCO version 3.12. Schultz and H.A. Mooney. Springer-Verlag, Agricultural Mathematics Group, DLO Box Heidelberg. 100, 6700 AC, Wageningen, The Yeatman, E.M. and Greenslade, P.J.M. (1980) Netherlands. Ants as indicators of habitat in three Tolhurst, K. G., Kellas, J. D. and Wilson, A. A. conservation parks in South Australia. (1992) Low intensity fire behaviour and fuel S.Aust.Nat. 55:20-26, 30. dynamics in dry sclerophyll forest, Wombat Yen, A.L. (1987) A preliminary assessment of the State Forest. In Tolhurst, K. And Flinn, D. correlation between plant, vertebrate and (eds.) Ecological Impacts of Fuel Reduction Coleoptera communities in the Victorian Burning in Dry Sclerophyll Forest: First mallee. In : Majer, J.D. (ed.), The Role of Progress Report. Conservation and Invertebrates in Conservation and Biological Environment, Victoria, Australia. Survey pp.73-88. W.A. Dept. of Turner, J. (1993) Research and Ecologically Conservation and Land Management Sustainable Forest Management. A Discussion Report. Paper. Research Division, State Forests of York, A. (1989) The Response of Ant Communities to New South Wales. 26pp. Fire-induced Habitat Disturbance. Van der Drift, J. (1958) The role of the soil fauna Unpublished PhD. Thesis, School of in the decomposition of forest litter. pp.357- Biological Sciences, University of New 360 in Proc. 15th Intern. Congr. Zool. South Wales. Van Loon, A.P. (1970) Further Investigations in York, A. (1993) Towards ecologically sustainable the Effects of Prescribed Burning on Young forest use. In Rapid Biodiversity Assessment (ed Even-aged Blackbutt. Taree Forestry Office Beattie, A.J.), pp. 51-55. Proceedings of the Research Report. Biodiversity Workshop Macquarie University - 1993. Research Unit for Biodiversity and Bioresources, School of Biological Sciences, Macquarie University.

265 Australia’s Biodiveristy - Responses to Fire

York, A. (1994) The long-term effects of fire on forest ant communities: management implications for the conservation of biodiversity. Memoirs of the Queensland Museum 36, 231-239. York, A. (1996) Long-term effects of fuel reduction burning on invertebrates in a dry sclerophyll forest. In Fire and Biodiversity. The Effects and Effectiveness of Fire Management. Pp. 163-181. Biodiversity Series, Paper No. 8. Biodiversity Unit, Department of the Environment Sport and Territories. Canberra, Australia. Zborowski, P. and Storey, R. (1995) A Field Guide to Insects in Australia. Reed Books, Sydney. 207pp. Zedler, P.H., Gautier, C.R. and McMaster, G.S. (1983) Vegetation change in response to extreme events: the effect of a short interval between fires in California and coastal scrub. Ecology 64:809-818.

266