S.E. Connor et al. Forgotten land-use impacts in NW Australia

1 Manuscript for Journal of Vegetation Science (special issue edited by Thomas Giesecke, Triin Reitalu

2 and Petr Kuneš).

3

4 Title:

5 Forgotten impacts of European land-use on riparian and savanna vegetation in North-Western

6 Australia

7

8 Author names and addresses:

9 Simon E. Connor, Larissa Schneider, Jessica Trezise, Susan Rule, Russell L. Barrett, Atun Zawadzki &

10 Simon G. Haberle

11

12 Connor, S.E. (corresponding author, [email protected]) 1,2

13 Schneider, L. ([email protected]) 3

14 Trezise, J. ([email protected]) 3

15 Rule, S. ([email protected]) 3

16 Barrett, R.L. ([email protected]) 4,5

17 Zawadzki, A. ([email protected]) 6

18 Haberle, S.G. ([email protected]) 3

19 1School of Geography, 221 Bouverie Street, University of Melbourne, 3010 VIC, Australia;

20 2CIMA-FCT, University of the Algarve, Campus de Gambelas, 8005-139 Faro, Portugal;

1

S.E. Connor et al. Forgotten land-use impacts in NW Australia

21 3Centre of Excellence in Australian Biodiversity & Heritage, and Department of Archaeology and

22 Natural History, Australian National University, Fellows Rd, Canberra, 2600 ACT, Australia;

23 4National Herbarium of New South Wales, Royal Botanic Gardens and Domain Trust, Mrs Macquaries

24 Road, Sydney NSW 2000, Australia;

25 5College of Medicine, Biology and Environment, Research School of Biology, Australian National

26 University, Canberra, 2601 ACT, Australia;

27 6Institute for Environmental Research, Australian Nuclear Science and Technology Organisation, PMB

28 1 Menai, 2234 NSW, Australia.

29

30 Printed journal page estimate:

31 8597 words including references (10.7 pages), figures 1.3 pages, total 12 pages.

32

33 Abstract

34 Questions

35 Fire and livestock grazing are regarded as current threats to biodiversity and landscape integrity in

36 Northern Australia, yet it remains unclear what biodiversity losses and habitat changes occurred in

37 the 19–20th centuries as livestock and novel fire regimes were introduced by Europeans. What

38 baseline is appropriate for assessing current and future environmental change?

39 Location

40 Australia's Kimberley region is internationally recognised for its unique biodiversity and cultural

41 heritage. The region is home to some of the world's most extensive and ancient rock art galleries,

42 created by Aboriginal peoples since their arrival on the continent 65,000 years ago. The Kimberley is

2

S.E. Connor et al. Forgotten land-use impacts in NW Australia

43 considered one of Australia's most intact landscapes and its assumed natural vegetation has been

44 mapped in detail.

45 Methods

46 Interpretations are based on a continuous sediment record obtained from a waterhole on the

47 Mitchell River floodplain. Sediments were analysed for geochemical and palynological proxies of

48 environmental change and dated using Lead-210 and Carbon-14 techniques.

49 Results

50 We show that the present-day vegetation in and around the waterhole is very different to its pre-

51 European counterpart. Pre-European riparian vegetation was dominated by ghaesembilla

52 and Banksia dentata, both of which declined rapidly at the beginning of the 20th century. Soon after,

53 savanna density around the site declined and grasses became more prevalent. These vegetation

54 shifts were accompanied by geochemical and biological evidence for increased grazing, local

55 burning, erosion and eutrophication.

56 Conclusions

57 We suggest that the Kimberley region's vegetation, while maintaining a 'natural' appearance, has

58 been altered dramatically during the last hundred years through grazing and fire regime changes.

59 Landscape management should consider whether the current (impacted) vegetation is a desirable or

60 realistic baseline target for biodiversity conservation.

61

62 Keywords geochemistry, palaeoecology, human impact, riverine vegetation, savanna, fire regimes,

63 grazing, Australian Tropics, Potential Natural Vegetation

64

65 Nomenclature Western Australian Herbarium (2017); Mucina & Daniel (2013).

3

S.E. Connor et al. Forgotten land-use impacts in NW Australia

66

67 Running head Forgotten land-use impacts in NW Australia

68

69 Introduction

70 Human impacts associated with European colonisation beginning in the 15th century

71 profoundly altered ecosystems around the globe. Impacts included the introduction of exotic

72 species, manipulation of fire regimes, deforestation, major land-use changes, displacement of

73 indigenous land management practices and alterations to biogeochemical cycles (Kirkpatrick 1999;

74 Ireland and Booth 2012; Johnson et al. 2017). As these impacts preceded a widespread

75 understanding of ecosystem ecology and research, in many parts of the world it is difficult to

76 understand how ecosystems functioned prior to European colonisation. This raises questions about

77 how to best manage altered ecologies in an era of rapid environmental change (Johnson et al. 2017).

78 A common baseline for large-scale ecosystem management is the concept of Potential

79 Natural Vegetation (PNV). PNV takes information from areas of assumed natural vegetation (late-

80 successional) and extrapolates to other (impacted) areas with similar environmental conditions

81 (Loidi & Fernández-González 2012). The PNV approach has been criticised because its predictions

82 sometimes conflict with the fossil record (Carrión & Fernández 2009; Rull 2015; Abraham et al.

83 2015). PNV also fails to account for ecosystem dynamics (Chiarucci et al. 2010) and is unable to

84 simulate ecological structure in cultural landscapes (Strona et al. 2016). Despite these constraints,

85 PNV remains an accessible baseline for conservation and management decisions.

86 Northern Australia is regarded as one of the largest areas of intact tropical savanna

87 worldwide (Woinarski et al. 2007; Bowman et al. 2010; Ziembicki et al. 2015). If this is true, PNV

88 maps for the region should constitute a valuable baseline (Beard et al. 2013; Mucina & Daniel 2013).

89 Yet the ecosystems of Northern Australia have been extensively modified since European

4

S.E. Connor et al. Forgotten land-use impacts in NW Australia

90 colonisation through livestock grazing, disruption of Indigenous land curation and the introduction of

91 invasive species, among other impacts (Hnatiuk & Kenneally 1981; Lonsdale 1994; Vigilante &

92 Bowman 2004; Douglas et al. 2015; Radford et al. 2015). Impacts are linked to extinctions, declining

93 native mammal populations, altered stream ecology, eutrophication, erosion and catchment

94 instability (Payne et al. 2004; Woinarski et al. 2007; Ziembicki et al. 2015).

95 Taking a different view, recent research links the recent decline of fire-sensitive species

96 across Northern Australia to interactions between fire and flammable grasses (Trauernicht et al.

97 2013; Bowman et al. 2014). Uncertainties surrounding the causes of recent species losses remain

98 obscure in the absence of reliable historical evidence for 19th and early-20th century environmental

99 change in this sparsely populated region. Such evidence could provide a critical test of ecosystem

100 integrity and reveal long-term drivers of species decline.

101 Northern Australia is home to high levels of biodiversity, much of which remains

102 undocumented (Barrett & Barrett 2011, 2015; Barrett 2013, 2015; Maslin et al. 2013; Moritz et al.

103 2013). Phylogenetic research reveals extraordinary levels of genetic diversity and endemism,

104 perhaps equal to the recognised biodiversity hotspots of Eastern Australia (Moritz et al. 2013). The

105 Kimberley region in NW Australia is internationally recognised for its extensive rock art galleries,

106 which are some of the most ancient examples of human artistic expression globally (Aubert 2012)

107 and are situated in an ancient cultural landscape created by Indigenous Australians (Rangan et al.

108 2015; Hiscock et al. 2016). There is a pressing need to understand and document the region’s

109 biological and cultural heritage as economic development pressures on the Northern Australian

110 environment intensify.

111 In this paper, we confront the assumed natural vegetation of the Kimberley region with

112 fossil evidence, aiming to assess the ecological integrity of the region’s tropical savanna and riparian

113 vegetation types. Our data pertain to the savanna woodlands and riparian thickets (Mucina & Daniel

114 2013) of the Mitchell Plateau, one of the last areas of the Australian continent to experience

5

S.E. Connor et al. Forgotten land-use impacts in NW Australia

115 European colonisation (McGonigal 1990). We compare pollen data with independent geochemical

116 data to understand the timing, direction and drivers of two centuries of environmental change.

117

118 Methods

119 Study area

120 The Kimberley Region is situated in NW Australia and constitutes a biogeographically and

121 geologically distinct entity within the Australian Monsoonal Tropics (Pepper & Keogh 2014). The

122 region is characterised by complex and ancient geology, which has combined with long periods of

123 subaerial exposure and weathering to create unique landforms and topography (Pillans 2007; Tyler

124 2016). The Kimberley region is broadly divided into three geological basins: the extensive Kimberley

125 Basin in the north, the Ord Basin to the south and east, and the Canning Basin to the south and west.

126 These basins are separated by two orogenic belts of metamorphosed and intrusive rocks. The

127 western portion of the Kimberley Basin is dominated by the King Leopold Sandstone (erosion-

128 resistant arenites) and Carson Volcanics (weathered basalts), while the Warton and Pentecost

129 Sandstones dominate the eastern portion. These rocks have been subjected to only minor faulting

130 and deformation since their deposition some 1840–1800 million years ago and thick, erosion-

131 resistant laterites developed over the basalts 70–50 million years ago (Tyler 2016). Geology has a

132 defining influence on the distribution of soil types and vegetation units, as well as biogeographical

133 and phylogenetic patterns (Hnatiuk & Kenneally 1981; Pepper and Keogh 2014; Mucina & Daniel

134 2013; Moritz et al. 2013; Barrett 2015; Barrett & Barrett 2015).

135 On the Mitchell Plateau, savanna woodlands on basalt tend to have a ground layer of

136 Allopteropsis semialata, Chrysopogon fallax, Heteropogon contortus, Sehima nervosum, Sorghum

137 plumosum and many other grasses, with a tree layer of eucalypts ( bigalerita, E. tectifera,

138 Corymbia greeniana, C. bella, C. disjuncta), ironwood trees (Erythrophloemum aff. chlorostachys) and

6

S.E. Connor et al. Forgotten land-use impacts in NW Australia

139 terminalia (Terminalia canescens, T. fitzgeraldii) depending on topography (Mucina & Daniel 2013).

140 Savanna woodlands on sandstone-derived soils have dominant Sorghum and Triodia grasses and a

141 dominant tree layer of eucalypts (Eucalyptus tetradonta, E. miniata, E. apodophylla, E. houseana,

142 Corymbia latifolia, C. nesophila, C. torta) and palms (Livistona eastonii) (Hnatiuk & Kenneally 1981;

143 Mucina & Daniel 2013). Sandstone outcrops have an altogether different flora, having a high

144 proportion of species considered to have Gondwanan heritage, as well as many endemics (Dunlop &

145 Webb 1991). Key species of these scrubs/heaths include gonocarpa, A. kelleri, A. arida, A.

146 tumida, Bossiaea arenitensis, brownii, C. exstipulata, C. achaeta, Corymbia polycarpa,

147 Gardenia pyriformis and Grevillea agrifolia (Mucina & Daniel 2013). Rainforest remnants (vine

148 thickets) occur in fire-protected locations where moisture is available all year (Hnatiuk & Kenneally

149 1981; Ondei et al. 2017). Albizia lebbeck, Atalaya variifolia, Cochlospermum fraseri, Sersalisia sericea

150 and Wrightia pubescens are important canopy species in these rainforests (Beard 1976). Wetlands

151 and riparian zones are dominated by Pandanus and Melaleuca species (P. aquaticus, P. spiralis, M.

152 leucadendron, M. nervosa, M. viridiflora). Nauclea orientalis, Timonius timon and the deciduous

153 shrub Antidesma ghaesembilla are also common in these areas.

154 The Kimberley region has a dry tropical monsoonal climate, with rainfall seasonality driven

155 by the Indo-Australian Summer Monsoon. Doongan Station, near our study site, receives an average

156 of 1205 mm of annual precipitation, 95% of which falls from November to April (Bureau of

157 Meteorology, http://www.bom.gov.au/climate/data/ accessed 5 June 2017). Rainfall events usually

158 occur in the form of prolonged tropical storms and/or cyclones during the humid ‘wet’ season,

159 flooding the region’s watercourses (Mucina & Daniel 2013). A steep rainfall gradient exists across the

160 Kimberley region – on average, over 1400 mm of precipitation falls annually on the seaward slopes

161 of the Mitchell Plateau, while the southern parts of the Kimberley adjoining the Great Sandy Desert

162 receive less than 400 mm. Maximum daily temperatures average over 30°C in most parts of the

163 Kimberley Region throughout the year (Bureau of Meteorology,

164 http://www.bom.gov.au/climate/data/ accessed 5 June 2017).

7

S.E. Connor et al. Forgotten land-use impacts in NW Australia

165

166 Sample collection and dating

167 A sediment core was collected from a small (approx. 100 m2) elliptical waterhole on the

168 floodplain of the Mitchell River (15°10’30.5”S, 125°53’29.3”E; Fig. 1) using a mud–water interface

169 corer operated from a floating platform. The site (MP11A) is 150 m from the main river channel and

170 forms part of a complex of lakes and swamps where the floodplain widens as it follows the

171 geological contact between basalt and sandstone. The waterhole is fringed by Melaleuca

172 leucadendron to the south and surrounded by savanna woodland on all sides. Nymphaea violacea

173 and Eleocharis grow in the water, which was approximately 50 cm deep at the time of core

174 collection.

175 The core was dated using a combination of 210Pb and 14C techniques (Wright et al. 2017).

176 Eight 210Pb samples were selected from the upper 35 cm of the core. Macroscopic remains

177 were removed before the samples were freeze dried, ground and weighed. Subsamples of >1 g were

178 dated at the Australian Nuclear Science and Technology Organisation (ANSTO) using alpha

179 spectrometry. Radiocarbon (14C) dating was performed on three samples >5 g using Accelerator

180 Mass Spectrometry at DirectAMS laboratories (Seattle, USA). Macroscopic plant material was

181 removed prior to further pretreatment. An age–depth model was constructed using cubic spline

182 regression in Clam 2.2 (Blaauw 2010).

183 Surface sediment samples are routinely collected in palaeoecological research to calibrate

184 pollen–vegetation relationships (Morris et al. 2017). In the Kimberley Region, surface samples have

185 been analysed to aid the interpretation of pollen records (Proske et al. 2014; Field et al. 2017), but

186 multi-site calibration has not been attempted. We collected surface sediment samples from ten

187 small lakes and wetlands across the region (Fig. 1). Multiple representative stands of the

188 surrounding vegetation were surveyed using a Haglöf HEC-R electronic clinometer and a GRS

189 Densitometer to estimate canopy height, tree volume and density. Composition was measured on

8

S.E. Connor et al. Forgotten land-use impacts in NW Australia

190 the DAFOR scale (Kent 2012). Satellite imagery was used to estimate the proportion of each

191 vegetation type within a 100 m radius of the coring site. Stand-based estimates were weighted using

192 these proportions to obtain vegetation estimates for comparison with pollen assemblages.

193

194 Pollen and charcoal analysis

195 Seasonally variable environments often have less suitable conditions for pollen preservation

196 than more stable environments (Head & Fullager 1992), requiring relatively large sediment samples.

197 Samples of 5 ml of sediment were treated according to standard palynological techniques (Jones &

198 Rowe 1999), including removal of clays through settling and decantation, sieving to remove large

199 particles (>120 μm), treatment with 10% KOH at 90°C for 20 minutes to remove humic material,

200 heavy liquid separation (s.g. 2.0) to remove minerogenic material, acetolysis (a 9:1 mixture of

201 C4H6O3to H2SO4), carbonate removal with 10% HCl, HF (50%) to remove fine silica, dehydration in

202 100% ethanol, and mounting in glycerol on glass slides. Lycopodium spore tablets were added early

203 in pretreatment to enable calculation of pollen concentrations and accumulation rates. Pollen,

204 spores and selected fungal and algal remains were identified at 400 x magnification, using the

205 extensive Kimberley collection in the Australasian Pollen and Spore Atlas (Rowe 2006) and published

206 guides (Macphail & Stevenson 2004; van Geel & Aptroot 2006). Identifications were agreed upon by

207 three analysts (JT, SR and SH).

208 Charcoal particles are an indicator of fire occurrence, with macroscopic particles considered

209 indicative of local fires and microscopic particles indicative of regional burning (Whitlock & Larsen

210 2001). Microscopic charcoal was quantified on pollen slides by counting opaque black particles 10–

211 120 μm. Macroscopic charcoal was sieved from a known volume of sediment using a 120 μm mesh

212 and particles identified using a binocular microscope (Stevenson & Haberle 2005). Results were

213 converted to accumulation rates for comparison with pollen data.

9

S.E. Connor et al. Forgotten land-use impacts in NW Australia

214 The pollen sequence was divided into relatively homogenous timespans using zonation

215 (CONISS), implemented in Tilia (version 2.0.41, Eric Grimm, http://www.TiliaIT.com). Pollen–

216 vegetation relationships were assessed by comparing the abundance of each pollen type to the

217 abundance of pollen-producing within a 100-m radius (the most commonly-used sampling

218 radius in comparable studies: Bunting et al. 2013).

219

220 Geochemical analysis

221 To examine the relationship between past vegetation and changes in the floodplain

222 environment, we analysed elemental concentrations of 34 geochemical elements. We used the

223 alkaline elements sodium (Na), potassium (K), magnesium (Mg), calcium (Ca) and barium (Ba) to

224 track erosion. These elements are relatively water-soluble and their mobility during erosion

225 processes results in elevated concentrations in water-body sediments (Davies et al. 2015). The

226 lithogenic element titanium (Ti) was used as a control for the mobile elements as it is geochemically

227 stable and hosted by resistant minerals (Davies et al. 2015). The element phosphorus (P) was used to

228 track livestock grazing. Cattle directly affect the local P cycle because the phosphorous they

229 consume is returned to the environment as dung (Parham et al. 2002). This increases P

230 concentrations in the soil and ultimately contributes to P enrichment in water bodies and sediments.

231 Sediment samples (33 from the core and 10 surface samples) of 0.2 g each were digested

232 using 1 ml of HNO3 (Aristar, BDH, Australia) and 2 ml HCl (Merck Suprapur, Germany) and analysed

233 for metals using an inductively coupled plasma mass spectrometer (ICP-MS) (Perkin-Elmer; model

234 DRC-e) with an AS-90 auto-sampler (Telford et al. 2008). Certified reference materials (NRC-BCSS1,

235 NRC-Mess1, NIST-1646, NRC-PCAS1) and blanks were used as to check the quality and traceability of

236 metals. Principal Components Analysis (PCA) was used to link geochemical variables to pollen

237 assemblages.

10

S.E. Connor et al. Forgotten land-use impacts in NW Australia

238

239

240 Results

241 Figure 2 shows the adopted age–depth model for the MP11A sediment record with dated

242 levels and associated errors. Two 14C dates were excluded from the model (30.5 and 40.5 cm depth).

243 These samples yielded ages that were the same or older than the lowermost 14C age in the core.

244 Redeposition of older carbon by floods is a common problem when dating recent floodplain deposits

245 (Ely et al. 1992). The chronology established by 210Pb has low errors and indicates relatively constant

246 sedimentation rates. The apparent increase in sedimentation toward the top of the core is likely to

247 reflect the relatively uncompacted nature of the more recent sediments.

248 Figure 3 shows key changes in the vegetation of the MP11A site and its surrounds through the

249 last 200 years (see Supporting Information for complete pollen record). Four main phases were

250 defined by zonation, described below in relation to changes in charcoal and geochemistry:

251 1. Phase 1 (1823–1907 AD) is characterised by relatively high proportions of Corymbia, other

252 , Callitris, Terminalia, Banksia and Antidesma ghaesembilla pollen. Phase 1 pollen

253 assemblages are very different to those of surface samples collected in modern waterholes

254 (see PCA, Supporting Information). Microscopic charcoal accumulation rates were relatively

255 high in Phase 1 compared with later phases, whilst macroscopic charcoal accumulation rates

256 were low (average 4% total charcoal). Geochemical indicators show a dip in concentrations

257 of weathering resistant Ti later in this phase and an increase in silicon (Si).

258 2. Phase 2 (1907–1948 AD) exhibits a steep decline in Antidesma ghaesembilla and Banksia

259 pollen. Pollen of the canopy dominants fluctuated, with Corymbia and Callitris attaining

260 their highest proportions in the entire sequence before rapidly decreasing toward the end

261 of the phase. Botryococcus decreased substantially at the same time and Sordaria fungal

11

S.E. Connor et al. Forgotten land-use impacts in NW Australia

262 spores increased in abundance. Charcoal declined substantially toward the end of this

263 phase, reaching its lowest concentrations in the entire record. In contrast, the mobile

264 elements (Na, K and Ba) increased while Ti concentrations declined. The end of this phase is

265 marked by a sharp increase in P, reaching levels over 10 times background concentrations.

266 3. Phase 3 (1948–1993 AD) begins with major increases in Poaceae and Pandanus pollen.

267 These changes were followed by a major increase in both macroscopic and microscopic

268 charcoal, reaching the highest concentrations in the record. Tree pollen decreased to an

269 average of 34% (compared to 50% in the previous zone). Phase 3 pollen assemblages are

270 similar to those of surface samples from modern waterholes (see PCA, Supporting

271 Information). Geochemical data show a temporary decline in concentrations of Na, K and Ba

272 and an increase in Ti, Mg and Ca.

273 4. Phase 4 (1993–2016 AD) is differentiated from the previous phase by a further decline in

274 tree pollen (average: 30%) and the highest levels of Poaceae, Acacia and Dodonaea pollen

275 for the entire record. Macroscopic charcoal is abundant compared to microscopic charcoal

276 (average 30% total charcoal). Sordaria remained abundant through this phase. Geochemical

277 indicators show relatively high concentrations for all elements except P, which decreased.

278

279 Figure 4 shows the relationships between major pollen taxa and vegetation parameters for

280 surface samples (data in Supporting Information). Corymbia pollen correlates strongly with area-

281 weighted plant abundance (r2: 0.87), savanna basal area (r2: 0.74) and savanna canopy cover (r2:

282 0.48). Pandanus pollen was less strongly related to abundance (r2: 0.37) and had no relationship to

283 savanna basal area (i.e. tree density). Poaceae produced a short abundance gradient (being

284 dominant in vegetation surveys), but displayed an inverse relationship with savanna density (r2:

285 0.42) and canopy cover (r2: 0.27). Corymbia:Poaceae ratios correlated positively with both Corymbia

286 abundance (r2: 0.64) and savanna density (r2: 0.67).

12

S.E. Connor et al. Forgotten land-use impacts in NW Australia

287

288 Discussion

289 Our multi-proxy results provide a glimpse of historical changes that occurred in the savannas

290 and inland floodplains of the Kimberley Region. A longer record of palaeovegetation change comes

291 from Black Springs, 75 km SW of site MP11A (15°38’S; 126°23’23”E), showing Pandanus expansion

292 during Holocene wet phases and aridification cycles linked to weakening (Field et al.

293 2017). This record, however, has low temporal resolution (centennial-scale) and uncertain dating for

294 recent centuries. The MP11A record provides decadal resolution, revealing a sequence of

295 environmental changes in the upper Mitchell River area since European colonisation. The 210Pb

296 chronology shows no signs of flood disturbance (Arnaud et al. 2002). Nevertheless, the assigned ages

297 should be viewed as approximate.

298 The sequence of events is summarised in Figure 3. An initial decline of the riparian shrub

299 Antidesma ghaesembilla began around 1900 (Phase 1), followed by Banksia decline a decade later.

300 Corymbia, indicative of savanna density (Fig. 4), increased around 1912 and Callitris followed in the

301 1920s (Phase 2). At the same time, regional fires decreased (indicated by lower levels of microscopic

302 charcoal) and grazing intensified (indicated a rapid increase in dung fungal spores, such as Sordaria).

303 The 1930s brought greater grazing intensity, increasing erosion (indicated by rising Ba and Na

304 concentrations) and a major decline in savanna trees (Corymbia and Callitris) and regional burning.

305 Eutrophication followed in the 1940s, with a pronounced peak in P lasting until the 1970s (Phase 3).

306 During this interval, grass abundance increased and Pandanus replaced A. ghaesembilla as the

307 dominant riparian element. Fire occurrence rose substantially in the late 1970s, particularly the local

308 fires indicated by macroscopic charcoal and Neurospora fungal spores (Whitlock & Larsen 2001; van

309 Geel & Aptroot 2006). This increase in local burning was accompanied by evidence of greater grazing

310 intensity (dung fungi), peaking around 1990. The last two decades have seen further expansion of

13

S.E. Connor et al. Forgotten land-use impacts in NW Australia

311 grasses and a decline in savanna density, as well as the disappearance of Pandanus from the

312 waterhole margins (Phase 4).

313 Observed changes in savanna fire regime, savanna vegetation structure and riparian zone

314 vegetation are discussed in the following sections in relation to major drivers of environmental

315 change.

316

317 Changes in savanna fire regime

318 On a global scale, savannas are thought to exist as an alternative stable state to forests,

319 maintained by frequent fire (Bond et al. 2005). Changing fire regimes are implicated in major

320 vegetation shifts across the Northern Australian savanna zone (O’Neill et al. 1993; Russell-Smith et

321 al.2003; Bowman 2003), including the simplification of vegetation structure (Craig 1997),

322 compositional change (Trauernicht et al. 2013; Bowman et al. 2014), shifting forest–savanna

323 boundaries (Hnatiuk & Kenneally 1981; Banfai & Bowman 2005) and negative impacts on native

324 mammals (Radford et al. 2015; Ziembicki et al. 2015). The effects of fire regime change in Australia

325 are not completely understood, given a lack of information on historical fire regimes and fire ecology

326 (Silcock et al. 2016). Fire regime change is often linked to the dispossession of Aboriginal peoples of

327 their land and consequent loss of traditional burning practices, followed by the different burning

328 practices of European colonists (Russell-Smith et al. 2003; Ritchie 2009). Alternatively, fire regime

329 change may stem from the invasion of fire-promoting grass species, creating a grass–fire positive

330 feedback cycle (Bowman et al. 2014).

331 Our data demonstrate profound changes in fire regime around the MP11A site over 200

332 years. Phase 1 samples are dominated by microscopic charcoal. This suggests that regional savanna

333 fires were common, while local fires in the riparian zone were apparently rare. This agrees with

334 ethnographic evidence (Russell-Smith et al. 1997; Vigilante 2001). The major reduction in regional

14

S.E. Connor et al. Forgotten land-use impacts in NW Australia

335 burning in the 1930s and 1940s (Phase 2) coincides with the decline of continuous Aboriginal

336 occupation of the Mitchell Plateau, ending in the 1950s (Wilson 1981). Data from the 1960s onwards

337 (Phases 3–4) show that local fires became more prevalent, impacting both savanna and riparian

338 zones (Fig. 3). Historically, Aboriginal fire management is associated with light/early-season fires

339 compared to unmanaged areas, where severe dry-season fires are more frequent (O’Neill et al.

340 1993).

341 The nature of the MP11A site limits the conclusions that may be drawn from charcoal

342 analysis. Fluvial transport is a major source of primary charcoal (direct products of fire) and

343 secondary charcoal (redeposited by erosion : Patterson et al. 1987; Whitlock & Larsen 2001).

344 Overlapping radiocarbon dates at different sediment depths could indicate contamination by older

345 (secondary) charcoal (Fig. 2). However, the macroscopic charcoal record is validated by a similar

346 trend in Neurospora, a biological proxy for local fire (van Geel & Aptroot 2006). Charcoal abundance

347 is not clearly related to erosion proxies (Fig. 3) and there is no indication of a major shift in sediment

348 source that explains the dramatic increase in macroscopic charcoal since the 1960s. Hence, the

349 MP11A charcoal record is cautiously interpreted as a reflection of fire occurrence in the riparian

350 zone (macroscopic charcoal) and surrounding savanna (microscopic charcoal). The changes are

351 consistent with a transition from Aboriginal to European fire management (Haberle et al. 2006), but

352 remain to be corroborated at other sites.

353

354 Changes in savanna structure

355 A critical question in understanding savanna vegetation dynamics is how global-scale factors

356 interact with the local environment (Case & Staver 2017). Previous research has debated the role of

357 fire versus local edaphic factors in determining vegetation structure in Northern Australia (e.g.

358 Bowman 1988; Craig 1997; Douglas et al. 2015; Ondei et al. 2017). These studies often used space-

359 for-time substitution or repeat aerial photography to analyse vegetation changes. Space-for-time

15

S.E. Connor et al. Forgotten land-use impacts in NW Australia

360 substitution is limited by confounding variables such as differences in soil type, climate, species

361 pools and land use at different sampling locations. Aerial photography provides long-term data on a

362 broad geographic scale, but has limited temporal scope (usually late 20th century) and provides little

363 detail on changes in understorey vegetation beneath the canopy where the impacts of savanna fire

364 are presumably greatest. Palaeoecological data from small basins, such as the MP11A record, reveal

365 vegetation changes at a small spatial scale over long timescales and are ideally placed to inform

366 debates on vegetation structure (Mariani et al. 2017; Morris et al. 2017).

367 Our results point to a long-term decline in savanna tree density. The decline is registered in

368 decreasing Corymbia pollen abundances and Corymbia:Poaceae ratios, proxies strongly associated

369 with savanna basal area (Fig. 4). Tree density decline was particularly acute in the 1930s. This

370 evidence, albeit from a single site, contrasts with observations of 20th-century ‘woody thickening’

371 across Northern Australia (e.g. Bowman et al. 2010; Tng et al. 2011, cf. Murphy et al. 2014).

372 Observed changes in vegetation density are dependent on the baseline used. The earliest aerial

373 photography for most of Northern Australia dates to the 1940s and 50s. Savanna density change at

374 MP11A since the 1940s has been negligible (Fig. 3). Compared to a baseline of the 19th or early 20th

375 century, however, loss has been substantial.

376 Around MP11A, grass cover increased as savanna density declined (Fig. 3). Historically,

377 livestock have caused major changes in grassland species composition in the Kimberley Region

378 (Petherham et al. 1986). Cattle, sheep, pigs, horses and donkeys were introduced by European

379 graziers from the 1880s (McGonigal 1990). Livestock numbers increased rapidly from less than

380 100,000 cattle in the 1890s to more than 600,000 in 1914 (Bolton 1953), and were linked to

381 environmental degradation from the 1930s (Payne et al. 2004). Cattle grazing can trigger a transition

382 from perennial to annual grasses in Australian savannas (Ash et al. 1997; Fensham & Skull 1999). The

383 same type of transition is linked to increasing fire frequency (Bowman et al. 2014). The histories of

16

S.E. Connor et al. Forgotten land-use impacts in NW Australia

384 grazing and fire at MP11A are so long and intertwined that a single cause of recent savanna

385 structural change is unlikely to emerge.

386 Perennial Triodia grasses, if left unburnt for long periods, can gain significant size and

387 density, providing habitat to a range of vertebrates and invertebrates. Old tussocks, often exceeding

388 two metres in height and five metres in breadth, were frequent along the Fitzroy Crossing Road until

389 the mid-1980s. Recent changes in fire regime have greatly reduced the extent of unburnt Triodia

390 tussocks, a change that may be linked to the decline of small mammals in the region. Pollen data

391 cannot be used to interrogate the history of specific grass species, because different species produce

392 pollen that is morphologically indistinguishable. Analysis of phytoliths or ancient DNA may help

393 address this shortfall.

394 Callitris intratropica has been the target of considerable research into fire impacts in

395 Northern Australian savannas (e.g. Bowman & Panton 1993; Bowman et al. 2014). Callitris pollen is

396 morphologically distinct and the MP11A pollen record provides insight into the recent history of this

397 fire-sensitive conifer. Callitris apparently experienced a major decline in the 1930s and has since

398 recovered slightly. C. intratropica trees were not sufficiently abundant in our vegetation surveys to

399 permit a quantitative test of the grass–fire cycle hypothesis (Bowman et al. 2014). Theoretically, if

400 the expansion of grasses is linked to Callitris decline, the ratio of Callitris:Poaceae pollen should

401 decrease over time. Instead there is an increase since the mid-20th century at MP11A (Fig. 3). The

402 purported link between fire and grass cover finds little support in our data, with charcoal and

403 Poaceae pollen following different trajectories. Our data cannot answer questions about grazing–

404 fire–grass interactions on a regional scale, but point to a profound legacy of 20th-century grazing and

405 fire impacts in the Mitchell Plateau savannas.

406

407 Changes in the riparian zone

17

S.E. Connor et al. Forgotten land-use impacts in NW Australia

408 Data from the MP11A waterhole provide strong evidence for the 20th-century deterioration

409 of a riparian shrub layer comprising Banksia and Antidesma. Banksia pollen represents B. dentata,

410 the only Banksia species extant in the Kimberley region. This non-serotinous species occurs in

411 seasonally inundated areas and swamp margins on sandstone bedrock, where it grows alongside

412 Eucalyptus apodophylla (Hnatiuk & Kenneally 1981). Extensive aerial surveys across the Kimberley

413 confirm this almost exclusive habitat restriction (R. Barrett, pers. obs.). Our data suggest that

414 Banksia dentata also occupied riparian zones in the past (i.e. Melaleuca leucadendron alliance of

415 Hnatiuk & Kenneally, 1981; Riparian Thickets group of Mucina & Daniel 2013).

416 Antidesma ghaesembilla is a fire-sensitive riparian shrub that bears edible fruits and is

417 known to expand rapidly in areas of unburnt woodland (Bowman et al. 1988; Russell-Smith et al.

418 2003). It occurs frequently along the Mitchell River floodplain today. The MP11A pollen record

419 suggests A. ghaesembilla was dominant around this site until the early 20th century. Its decline

420 follows a peak in macroscopic charcoal, implicating local fires as the cause. According to

421 ethnographic sources, Aboriginal people avoided burning riparian vegetation (Russell-Smith et al.

422 1997; Vigilante 2001). Post-fire recovery may have been impeded by grazing, as Antidesma species

423 are more palatable to cattle than browse- and burning-resistant Pandanus.

424 Riparian and wetland zones are where cattle and feral animal impacts in the Kimberley are

425 most acute (Legge et al. 2011). Grazing and trampling have had significant detrimental impacts on

426 riparian ‘bamboo’ (Phragmites karka), once common in the Kimberley, but now listed as a species of

427 conservation concern in . Enhanced erosion is attested geochemically by increased

428 concentrations of mobile versus immobile elements (Bierman et al. 1998). At MP11A, mobile Na, K

429 and Ba increased as relatively immobile Ti decreased during the 20th century. Erosion is further

430 indicated by Pseudoschizaea palynomorphs (Lopez-Marino et al. 2010), which peak in Phases 3–4.

431 Grazing may cause P enrichment from faeces and losses of Ca and Mg from urine patches (Early et al.

432 1993; Parham et al. 2002). In the MP11A record, the early–mid 20th century is characterised by an

18

S.E. Connor et al. Forgotten land-use impacts in NW Australia

433 increase in P (Fig. 3) and a decrease in Ca and Mg (Supporting Information). With the additional

434 support of proxies for livestock dung (Fig. 3), these geochemical changes may be linked to grazing

435 impacts.

436 Some of the changes witnessed by the MP11A record could be explained by other

437 environmental shifts occurring in the Mitchell River catchment . Larsen et al. (2016) demonstrated

438 how the geomorphic process of knick-point migration caused riparian forest loss in Australia’s

439 . Sand slugs released by upstream erosion cause long-lasting impacts on

440 downstream ecosystems in many Australian rivers (Prosser et al. 2001). These explanations seem

441 inadequate in the case of MP11A – aerial photographs of the study area in 1949 and 2017 show no

442 obvious change in riparian forest and floodplain geomorphology (see Supporting Information).

443 There are no major flooding events corresponding to the time of peak impact at MP11A

444 (Gillieson et al. 1991; Wohl et al. 1994). Meteorological records indicate some higher-than-average

445 rainfall years in the 1940s and 1950s, but much higher values in recent decades (Bureau of

446 Meteorology, http://www.bom.gov.au/climate/data/ accessed 5 June 2017), part of a wetting trend

447 across Northern Australia (Bowman et al. 2010). The decades-long environmental transformation

448 evident at MP11A cannot be explained by a single disturbance event. Agreement between

449 geochemical and biological proxies for grazing and fire, alongside clear evidence for the loss of

450 savanna tree density and the shrub stratum within the riparian zone, suggest geomorphic processes

451 are an unlikely cause of historic vegetation changes in the study area.

452

453 Implications for conservation and vegetation mapping

454 Our evidence casts doubt on the notion that Northern Australia’s landscapes and

455 ecosystems are intact and unmodified (Woinarski et al. 2007; Bowman et al. 2010; Ziembicki et al.

456 2015). Just as the intact canopy of ‘Australian’ eucalypts disguises a savanna flora with strong SE

19

S.E. Connor et al. Forgotten land-use impacts in NW Australia

457 Asian affinities (Haynes et al. 1991), the same canopy may disguise the pervasive ecological

458 transformations that have occurred over the last 100 years. The forgotten impacts of early 20th-

459 century fire and grazing have profoundly altered savanna and riparian ecosystems on the Mitchell

460 Plateau and perhaps more widely.

461 Our data call for caution in adopting mid-20th-century ecological baselines for assessing

462 ecological change, given that the impacts of European colonisation often manifested themselves

463 much earlier. Potential Natural Vegetation (PNV) maps have similar limitations. Because PNV maps

464 are populated from data collected in the current (impacted) environment, there are likely to be

465 knowledge shortfalls in reconstructing pre-European vegetation (e.g. Beard et al. 2013). Given the

466 duration and magnitude of historical impact, it could be asked whether restoration to pre-European

467 conditions is desirable or realistic.

468 There is a clear need for replication beyond the single palaeoecological record presented

469 here to better understand spatial and temporal variation in the Kimberley region’s ecosystems. It is

470 acknowledged that obtaining palaeoecological data for vegetation types existing under more arid

471 conditions may be difficult (Head & Fullager 1992). Hence PNV provides a useful ‘null model’

472 (Somodi et al. 2012) that can only be improved as better data and new sources of information come

473 to hand.

474 Failure to recognise the Kimberley Region’s historically-altered ecological state could place

475 the remaining biodiversity and ecosystems at greater risk than could be expected under truly ‘intact’

476 ecological conditions.

477

478 Acknowledgements

479 We acknowledge the traditional owners of the land on which our research was conducted. Sincere

480 thanks to Cecelia Myers, Susan Bradley and Doongan Station staff for facilitating and supporting our

20

S.E. Connor et al. Forgotten land-use impacts in NW Australia

481 research. Thanks to Janelle Stevenson, Peter Kershaw and an anonymous reviewer for constructive

482 reviews. Funding was provided by the Kimberley Foundation of Australia and the ARC

483 (DP140103591) project ‘Environmental Transformations linked to Early Human Occupation in

484 Northern Australia’. Research is also supported by the ARC Centre of Excellence for Australian

485 Biodiversity and Heritage (CE170100015).

486

487 Author contributions

488 SEC led the writing and conducted fieldwork

489 LS led the geochemical and statistical analyses and contributed to the manuscript

490 JT analysed the palynological data and led the literature review

491 SR analysed additional palynological data and oversaw pollen

492 RLB contributed to the manuscript, particularly its introduction and discussion sections

493 AZ provided Pb-210 dating and expertise

494 SGH coordinated the research, conducted fieldwork and contributed to the manuscript

495

496 References

497 Abraham, V., Kuneš, P., Petr, L., Svitavská-Svobodová, H., Jamrichová, E., Švarcová, M.G. & Pokorný,

498 P. 2016. A pollen-based quantitative reconstruction of the Holocene vegetation updates a

499 perspective on the natural vegetation in the Czech Republic and Slovakia. Preslia 88: 409-

500 434.

21

S.E. Connor et al. Forgotten land-use impacts in NW Australia

501 Arnaud, F., Lignier, V., Revel, M., Desmet, M., Beck, C., Pourchet, M., Charlet, F., Trentesaux, A. &

502 Tribovillard, N. 2002. Flood and earthquake disturbance of 210Pb geochronology (Lake

503 Anterne, NW Alps). Terra Nova 14: 225-232.

504 Ash, A.J., McIvor, J.G., Mott, J.J. & Andrew, M.H. 1997. Building grass castles: integrating ecology and

505 management of Australia's tropical tallgrass rangelands. The Rangeland Journal 19: 123-144.

506 Aubert, M. 2012. A review of rock art dating in the Kimberley, Western Australia. Journal of

507 Archaeological Science 39: 573–577.

508 Banfai, D.S. & Bowman, D.M.J.S. 2005. Dynamics of a savanna-forest mosaic in the Australian

509 monsoon tropics inferred from stand structures and historical aerial photography. Australian

510 Journal of 53: 185-194.

511 Barrett, R.L. & Barrett, M.D. 2011. Two new species of Triodia (Poaceae: Triodieae) from the

512 Kimberley region of Western Australia. Telopea 13: 57–67.

513 Barrett, R.L. & Barrett, M.D. 2015. Twenty-seven new species of vascular plants from Western

514 Australia. Nuytsia 26: 21–87.

515 Barrett, R.L. 2013. Solanum zoeae (Solanaceae), a new species of bush tomato from the North

516 Kimberley, Western Australia. Nuytsia 23: 5–21.

517 Barrett, R.L. 2015. Fifty new species of vascular plants from Western Australia—celebrating fifty

518 years of the Western Australian Botanic Garden at Kings Park. Nuytsia 26: 3–20.

519 Beard, J.S. 1976. The monsoon forests of the Admiralty Gulf, Western Australia. Vegetatio 31: 177-

520 192.

521 Beard, J.S., Beeston, G.R., Harvey, J.M., Hopkins, A.J.M. & Shepherd, D.P. 2013. The vegetation of

522 Western Australia at the 1:3,000,000 scale. Explanatory memoir. Second edition.

523 Conservation Science Western Australia 9: 1–152.

22

S.E. Connor et al. Forgotten land-use impacts in NW Australia

524 Bierman, P.R., Albretch, A., Bothner, M.H., Brown, E.T., Bullen, T.D., Gray, L. & Turpin, L. 1998.

525 Erosion, weathering and sedimentation. In: Kendall, C. & McDonnell, J.J. (eds.), Isotope

526 Tracers in Catchment Hydrology, pp. 647–678. Elsevier, Amsterdam.

527 Blaauw, M., 2010. Methods and code for 'classical' age-modelling of radiocarbon

528 sequences. Quaternary Geochronology 5: 512-518.

529 Bolton, G.C. 1953. A survey of the Kimberley pastoral industry from 1885 to the present. University of

530 Western Australia, Perth.

531 Bond, W.J., Woodward, F.I. & Midgley, G.F. 2005. The global distribution of ecosystems in a world

532 without fire. New Phytologist 165: 525-538.

533 Bowman, D.M.J.S. 1988. Stability amid turmoil?: towards an ecology of north Australian eucalypt

534 forests. Proceedings of the Ecological Society of Australia 15: 149-158.

535 Bowman, D.M.J.S. 2003. Australian landscape burning: a continental and evolutionary perspective.

536 In: Abbott, I. & Burrows, N. (eds.) Fire in ecosystems of south-west Western Australia:

537 impacts and management, pp. 107–118. Backhuys Publishers, Leiden, The Netherlands.

538 Bowman, D.M.J.S., Brown, G.K., Braby, M.F., Brown, J.R., Cook, L.G., Crisp, M.D., Ford, F., Haberle,

539 S.G., Hughes, J., Isagi, Y., Joseph, L., McBride, J., Nelson, G. & Ladiges, P.Y. 2010.

540 Biogeography of the Australian monsoon tropics. Journal of Biogeography 37: 201–216.

541 Bowman, D.M.J.S., MacDermott, H.J., Nichols, S.C. & Murphy, B.P. 2014. A grass–fire cycle eliminates

542 an obligate-seeding tree in a tropical savanna. Ecology and Evolution 4: 4185-4194

543 Bowman, D. M. J. S., and W. J. Panton. 1993. Decline of Callitris intratropica Baker, R.T. and Smith,

544 H.G. in the Northern Territory: implications for pre-European and post-European

545 colonization fire regimes. Journal of Biogeography 20: 373-381.

23

S.E. Connor et al. Forgotten land-use impacts in NW Australia

546 Bowman, D.M.J.S., Wilson, B.A. & Hooper, R.J. 1988. Response of Eucalyptus forest and woodland to

547 four fire regimes at Munmarlary Northern Territory Australia. Journal of Ecology 76: 215-

548 232.

549 Bunting, M.J., Farrell, M., Brostrom, A., Hjelle, K.L., Mazier, F., Middleton, R., Nielsen, A.B., Rushton,

550 E., Shaw, H. & Twiddle, C.L. 2013. Palynological perspectives on vegetation survey: a critical

551 step for model-based reconstruction of Quaternary land cover. Quaternary Science Reviews

552 82: 41-55.

553 Carrión, J.S. & Fernández, S. 2009. The survival of the ‘natural potential vegetation’ concept (or the

554 power of tradition). Journal of Biogeography 36: 2202-2203.

555 Case, M.F. & Staver, A.C. 2017. Fire prevents woody encroachment only at higher-than-historical

556 frequencies in a South African savanna. Journal of Applied Ecology 54: 955-962.

557 Chiarucci, A., Araújo, M.B., Decocq, G., Beierkuhnlein, C. & Fernández-Palacios, J.M. 2010. The

558 concept of potential natural vegetation: an epitaph? Journal of Vegetation Science 21: 1172–

559 1178.

560 Craig, A.B. 1997. A review of information on the effects of fire in relation to the management of

561 rangelands in the Kimberley high-rainfall zone. Tropical Grasslands 31: 161–187.

562 Davies, S., Lamb, H. & Roberts, S.J. 2015. Micro-XRF core scanning in palaeolimnology: recent

563 developments. In: Croudace I., Rothwell R. (eds) Micro-XRF Studies of Sediment Cores.

564 Developments in Paleoenvironmental Research, vol 17. Springer, Dordrecht. pp. 189-226.

565 Douglas, M.M., Setterfield, S.A., McGuinness, K. & Lake, P.S. 2015. The impact of fire on riparian

566 vegetation in Australia’s tropical savanna. Freshwater Science 34: 1351–1365.

567 Dunlop, C.R. & Webb, L.R. 1991. Flora and vegetation. In: Haynes, C.D., Ridpath, M.G. & Williams,

568 M.A.J. (eds) Monsoonal Australia: Landscape, Ecology and Man in the Northern Lowlands,

569 pp.41-60. A.A. Balkema, Rotterdam.

24

S.E. Connor et al. Forgotten land-use impacts in NW Australia

570 Early, M.S.B., Cameron, K.C. & Fraser, P.M. 1998. The fate of potassium, calcium, and magnesium in

571 simulated urine patches on irrigated dairy pasture soil. New Zealand Journal of Agricultural

572 Research 41: 117-124.

573 Ely, L.L., Webb, R.H. & Enzel, Y. 1992. Accuracy of post-bomb 137Cs and 14C in dating fluvial deposits.

574 Quaternary Research 38: 196-204.

575 Fensham, R.J & Skull, S.D. 1999. Before cattle: a comparative floristic study of Eucalyptus savanna

576 grazed by macropods and cattle in North , Australia. Biotropica 31: 37-47.

577 Field, E., McGowan, H.A., Moss, P.T. & Marx, S.K. 2017. A late Quaternary record of monsoon

578 variability in the northwest Kimberley, Australia. Quaternary International 449: 119-135.

579 Gillieson, D., Smith, D.I., Greenaway, M. & Ellaway, M. 1991. Flood history of the limestone ranges in

580 the Kimberley region, Western Australia. Applied Geography 11: 105-123.

581 Haberle, S.G., Tibby, J., Dimitriadis, S. & Heijnis, H. 2006. The impact of European occupation on

582 terrestrial and aquatic ecosystem dynamics in an Australian tropical rain forest. Journal of

583 Ecology 94: 987-1004.

584 Haynes, C.D., Ridpath, M.G. & Williams, M.A.J. (eds) Monsoonal Australia: Landscape, Ecology and

585 Man in the Northern Lowlands. A.A. Balkema, Rotterdam

586 Head, L. & Fullager, R. 1992. Palaeoecology and archaeology in the East Kimberley. Quaternary

587 Australasia 10: 27-31.

588 Henderson, M.K. & Keith, D.A. 2002. Correlation of burning and grazing indicators with composition

589 of woody understorey flora of dells in a temperate eucalypt forest. Austral Ecology 27: 121-

590 131.

591 Hiscock, P., O’Connor, S., Balme, J. & Maloney, T. 2016. World’s earliest ground-edge axe production

592 coincides with human colonisation of Australia. Australian Archaeology 82: 2-11.

25

S.E. Connor et al. Forgotten land-use impacts in NW Australia

593 Hnatiuk, R.J. & Kenneally, K.F. 1981. A survey of the vegetation and flora of Mitchell Plateau,

594 Kimberley, Western Australia. In: Biological survey of Mitchell Plateau and Admiralty Gulf,

595 Kimberley, Western Australia, pp. 13–94. Western Australian Museum, Perth.

596 Ireland, A.W. & Booth, R.K. 2012. Upland deforestation triggered an ecosystem state-shift in a kettle

597 peatland. Journal of Ecology 100, 586–596.

598 Johnson, C.N., Balmford, A., Brook, B.W., Buettel, J.C., Galeti, M., Guangchun, L. & Wilmshurst, J.M.

599 2017. Biodiversity losses and conservation responses in the Anthropocene. Science 356: 270-

600 275.

601 Jones, T.P. & Rowe, N.P. (eds) 1999. Fossil plants and spores: modern techniques. Geological Society,

602 London.

603 Kent, M. 2012. Vegetation Description and Analysis: a practical approach. 2nd ed. Wiley-Blackwell,

604 Chichester.

605 Kirkpatrick J. 1999. A Continent Transformed: Human Impact on the Natural Vegetation of Australia.

606 2nd ed. Oxford University Press, South Melbourne.

607 Larsen, A., May, J.-H., Moss, P. & Hacker, J. 2016. Could alluvial knickpoint retreat rather than fire

608 drive the loss of alluvial wet monsoon forest, tropical northern Australia? Earth Surface

609 Processes and Landforms 41: 1583–1594.

610 Legge, S., Murphy, S., Kingswood, R., Maher, B. & Swan, D. 2011. EcoFire: restoring the biodiversity

611 values of the Kimberley region by managing fire. Ecological Management and Restoration

612 12: 84–92.

613 Loidi, J. & Fernández-González, F. 2012. Potential natural vegetation: reburying or reboring? Journal

614 of Vegetation Science 23: 596-604.

26

S.E. Connor et al. Forgotten land-use impacts in NW Australia

615 Lonsdale, W.M. 1994. Inviting trouble: introduced pasture species in northern Australia. Australian

616 Journal of Ecology 19: 345-354.

617 López-Marino, L., Martínez Cortizas, A. & López Sáez, J.A. 2010. arly agriculture and

618 palaeoenvironmental history in the North of the Iberian Peninsula: a multi-proxy analysis of

619 the Monte Areo mire (Asturias, Spain). Journal of Archaeological Science 37: 1978-1988.

620 Macphail, M. & Stevenson, J. 2004. Fungal Spores in Archaeological Context. PalaeoWorks Technical

621 Report 3. Department of Archaeology and Natural History, Australian National University,

622 Canberra.

623 Mariani, M., Connor, S.E., Fletcher, M.-S., Theuerkauf, M., Kuneš, P., Jacobsen, G., Saunders, K.M. &

624 Zawadzki, A. 2017. How old is the Tasmanian cultural landscape? A test of landscape

625 openness using quantitative land-cover reconstructions. Journal of Biogeography doi:

626 10.1111/jbi.13040

627 Maslin, B.R., Barrett, M.D. & Barrett, R.L. 2013. A baker’s dozen of new wattles highlights significant

628 Acacia (: ) diversity and endemism in the north-west Kimberley

629 region of Western Australia. Nuytsia 23: 543–587.

630 McGonigal. D. 1990. The Australian Geographic Book of the Kimberley. Australian Geographic, Terrey

631 Hills NSW.

632 Moritz, C., Ens, E.J., Potter, S. & Catullo, R.A. 2013. The Australian monsoonal tropics: An opportunity

633 to protect unique biodiversity and secure benefits for Aboriginal communities. Pacific

634 Conservation Biology 19: 343-355.

635 Morris, J.L., Higuera, P.E., Haberle, S.G. and Whitlock, C. 2017. Modern pollen from small hollows

636 reflects Athrotaxis cupressoides density across a wildfire gradient in subalpine forests of the

637 Central Plateau, Tasmania, Australia. The Holocene doi: 10.1177/0959683617702228

27

S.E. Connor et al. Forgotten land-use impacts in NW Australia

638 Mucina, L. & Daniel, G. (eds) 2013. Vegetation mapping in Northern Kimberley, Western Australia.

639 Curtin University, Bentley.

640 Murphy, B.P., Lehmann, C.E.R., Russell-Smith, J. & Lawes, M.J. 2014. Fire regimes and woody

641 biomass dynamics in Australian savannas. Journal of Biogeography 41: 133–144.

642 Ondei, S., Prior, L.D., Williamson, G.J., Vigilante, T. & Bowman, D.M.J.S. 2017. Water, land, fire, and

643 forest: multi-scale determinants of rainforests in the Australian monsoon tropics. Ecology

644 and Evolution 7: 1592–1604.

645 O’Neill, A.L., Head, L.M. & Marthick, J.K. 1993. Integrating remote sensing and spatial analysis

646 techniques to compare Aboriginal and pastoral fire patterns in the East Kimberley, Australia.

647 Applied Geography 13: 67-85.

648 Parham, J., Deng, S., Raun, W. & Johnson, G. 2002. Long-term cattle manure application in soil I.

649 Effect on soil phosphorus levels, microbial biomass C, and dehydrogenase and phosphatase

650 activities. Biology and Fertility of Soils 35: 328-337,

651 Patterson, W.A., Edwards, K.J. & Maguire, D.J. 1987. Microscopic charcoal as a fossil indicator of fire.

652 Quaternary Science Reviews 6: 3-23.

653 Payne, A.L., Watson, I.W. & Novelly, P.E. 2004. Spectacular Recovery in the Ord River Catchment.

654 Department of Agriculture and Food, Western Australia, Perth. Report 17/2004

655 Pepper, M. & Keogh, J.S. 2014. Biogeography of the Kimberley, Western Australia: a review of

656 landscape evolution and biotic response in an ancient refugium. Journal of Biogeography 41:

657 1443–1455.

658 Petheram, R.J., Kok, B. & Bartlett-Torr, E. 1986. Plants of the Kimberley region of Western Australia.

659 Revised ed. Western Australian Department of Agriculture, South Perth.

28

S.E. Connor et al. Forgotten land-use impacts in NW Australia

660 Pillans, B. 2007. Pre-Quaternary landscape inheritance in Australia. Journal of Quaternary Science 22:

661 439-447.

662 Proske, U., Heslop, D. & Haberle, S. 2014. A Holocene record of coastal landscape dynamics in the

663 eastern Kimberley region, Australia. Journal of Quaternary Science 29: 163–174.

664 Prosser, I.P., Rutherfurd, I.D., Olley, J.M., Young, W.J., Wallbrink, P.J. & Moran, C.J. 2001. Large-scale

665 patterns of erosion and sediment transport in river networks, with examples from Australia.

666 Marine and Freshwater Research 52: 81-99.

667 Radford, I.J., Gibson, L.A., Corey, B., Carnes, K. & Fairman, R. 2015. Influence of fire mosaics, habitat

668 characteristics and cattle disturbance on mammals in fire-prone savanna landscapes of the

669 northern Kimberley. PLoS ONE 10(6): e0130721.

670 Rangan, H., Bell, K.L., Baum, D.A., Fowler, R., McConvell, P., Saunders, T., Spronck, S., Kull, C.A. &

671 Murphy, D.J. 2015. New genetic and linguistic analyses show ancient human influence on

672 baobab evolution and distribution in Australia. PLoS ONE 10(4): e0127582

673 Ritchie, D. 2009. Things fall apart: the end of an era of systematic Indigenous fire management. In:

674 Russell-Smith, J., Whitehead, P.J. & Cooke, P.M. (eds.) Culture, ecology and economy of fire

675 management in north Australian savannas: rekindling the Wurrk tradition, pp. 23–40. CSIRO

676 Publishing, Collingwood.

677 Rowe, C. 2006. The Australasian Pollen and Spore Atlas User Document. PalaeoWorks Technical

678 Report 8. Department of Archaeology and Natural History, Australian National University,

679 Canberra.

680 Rull, V. 2015. Long-term vegetation stability and the concept of potential natural vegetation in the

681 Neotropics. Journal of Vegetation Science 26: 603-607.

682 Russell-Smith, J., Lucas, D., Gapindi, M., Gunbunuka, B., Kapirigi, N., Namingum, G., Lucas, K.,

683 Giuliani, P. & Chaloupka, G. 1997. Aboriginal resource utilization and fire management

29

S.E. Connor et al. Forgotten land-use impacts in NW Australia

684 practice in Western Arnhem Land, Monsoonal Northern Australia: notes for prehistory,

685 lessons for the future. Human Ecology 25: 159-195.

686 Russell-Smith, J., Yates, C., Edwards, A., Allan, G.E., Cook, G.D., Cooke, P., Craig, R., Heath, B. &

687 Smith, R. 2003. Contemporary fire regimes of northern Australia, 1997–2001: change since

688 Aboriginal occupancy, challenges for sustainable management. International Journal of

689 Wildland Fire 12: 283–297.

690 Skroblin, A. & Legge, S. 2012. Influence of fine-scale habitat requirements and riparian degradation

691 on the distribution of the purple-crowned fairy-wren (Malurus coronatus coronatus) in

692 northern Australia. Austral Ecology 37: 874-884.

693 Somodi, I., Molnar, Z. & Ewald, J. 2012. Towards a more transparent use of the potential natural

694 vegetation concept – an answer to Chiarucci et al. Journal of Vegetation Science 23: 590–

695 595.

696 Silcock, J.L., Witt, G.B. & Fensham, R.J. 2016. A 150-year fire history of mulga (Acacia aneura F.

697 Muell. ex Benth.) dominated vegetation in semiarid Queensland, Australia. The Rangeland

698 Journal 38: 391–415.

699 Stevenson, J. and Haberle, S.G. (2005) Macro Charcoal Analysis: A modified technique used by the

700 Department of Archaeology and Natural History. PalaeoWorks Technical Report 5.

701 Department of Archaeology and Natural History, Australian National University, Canberra.

702 Strona, G., Mauri, A., Veech, J.A., Seufert, G., San Miguel-Ayanz, J. & Fattorini, S. 2016. Far from

703 naturalness: how much does spatial ecological structure of European tree assemblages

704 depart from Potential Natural Vegetation? PLoS ONE 11(12): e0165178.

705 Telford, K., Maher, W., Krikowa, F., & Foster, S. 2008. Measurement of total antimony and antimony

706 species in mine contaminated soils by ICPMS and HPLC-ICPMS. Journal of Environmental

707 Monitoring 10: 136-140.

30

S.E. Connor et al. Forgotten land-use impacts in NW Australia

708 Tng, D.Y.P., Murphy, B.P., Weber, E., Sanders, G., Williamson, G.J., Kemp, J. & Bowman, D.J.M.S.

709 2011. Humid tropical rain forest has expanded into eucalypt forest and savanna over the last

710 50 years. Ecology and Evolution 2: 34–45.

711 Trauernicht, C., Murphy, B.P., Tangalin, N. & Bowman, D.M.J.S. 2013. Cultural legacies, fire ecology,

712 and environmental change in the Stone Country of Arnhem Land and Kakadu National Park,

713 Australia. Ecology and Evolution 3: 286–297.

714 Tyler, I. 2016. Geology and Landforms of the Kimberley. Bush Books, Department of Parks and

715 Wildlife, Kensington WA.

716 van Geel, B. & Aptroot, A. 2006. Fossil ascomycetes in Quaternary deposits. Nova Hedwigia 82: 313-

717 329.

718 Vigilante, T. 2001. Analysis of explorers’ records of Aboriginal landscape burning in the Kimberley

719 Region of Western Australia. Australian Geographical Studies 39: 135-155.

720 Vigilante, T. & Bowman, D.M.J.S. 2004. Effects of fire history on the structure and floristic

721 composition of woody vegetation around Kalumburu, north Kimberley, Australia: a

722 landscape-scale natural experiment. Australian Journal of Botany 52: 381–404.

723 Warren, S.D., Thurow, T.L., Blackburn, W.H. & Garza, N.E. 1986. The influence of livestock trampling

724 under intensive rotation grazing on soil hydrologic characteristics. Journal of Range

725 Management 39: 491–495.

726 Western Australian Herbarium 2017. FloraBase – The Western Australian Flora. Department of Parks

727 and Wildlife. http://florabase.dpaw.wa.gov.au/ [accessed 31 May 2017]

728 Whitlock, C. & Larsen, C. 2001. Charcoal as a fire proxy. In: Smol, J.P., Birks, H.J.B. & Last, W.M. (eds)

729 Tracking Environmental Change using Lake Sediments, Volume 3: Terrestrial, Algal and

730 Siliceous Indicators, pp. 75-97. Kluwer, Dordrecht.

31

S.E. Connor et al. Forgotten land-use impacts in NW Australia

731 Wilson, B.R. 1981. General introduction. In: Biological survey of Mitchell Plateau and Admiralty Gulf,

732 Kimberley, Western Australia, pp. 1-11. Western Australian Museum, Perth

733 Wohl, E.E., Fuertsch, S.J. & Baker, V.R. 1994. Sedimentary records of late Holocene floods along the

734 Fitzroy and Margaret Rivers, Western Australia. Australian Journal of Earth Sciences 41: 273-

735 280.

736 Woinarski, J.C.Z., Mackey, B., Nix, H.A. & Traill, B.J. 2007. The nature of northern Australia - natural

737 values, ecological processes and future prospects. ANU Press, Canberra.

738 Wright, A.J., Edwards, R.J., van de Plasschea, O., Blaauw, M., Parnell, A.C., van der Borge, K., de

739 Jonge, A.F.M., Roec, H.M., Selbyf, K. & Blackg, K. 2017. Reconstructing the accumulation

740 history of a saltmarsh sediment core: which age-depth model is best? Quaternary

741 Geochronology 39: 35–67.

742 Yibarbuk, D., Whitehead, P.J., Russell-Smith, J., Jackson, D., Godjuwa, C., Fisher, A., Cooke, P.,

743 Choquenot, D. & Bowman, D.M.J.S. 2001. Fire ecology and Aboriginal land management in

744 central Arnhem Land, northern Australia: A tradition of ecosystem management. Journal of

745 Biogeography 28: 325-343.

746 Ziembicki, M.R., Woinarski, J.C.Z., Webb, J.K., Vanderduys, E., Tuft, K., Smith, J., Ritchie, E.G.,

747 Reardon, T.B., Radford, I.J. (…) & Burbidge, A.A. 2015. Stemming the tide: progress towards

748 resolving the causes of decline and implementing management responses for the

749 disappearing mammal fauna of northern Australia. Therya 6: 169-225.

750

32

S.E. Connor et al. Forgotten land-use impacts in NW Australia

751

752 Fig. 1 Location of the main study site (MP11A -

753 triangle) and surface sampling sites (circles) in

754 relation to regional vegetation structure, NW

755 Kimberley Region, Western Australia.

756 Vegetation units after Mucina and Daniel (2013)

757

758

759

760

761

762

763 Fig. 2 Age-depth model for the MP11A

764 sediment core, showing calibrated ages and

765 associated errors from 210Pb and 14C isotopic

766 dating, with smooth spline interpolation

33

S.E. Connor et al. Forgotten land-use impacts in NW Australia

767

768 Fig. 3 Summary of major changes in environmental proxies at site MP11A, Mitchell Plateau, Western

769 Australia (see Appendix S2 for complete palynological and geochemical proxy diagrams)

34

S.E. Connor et al. Forgotten land-use impacts in NW Australia

770

771 Fig. 4 Scatterplots of plant abundance and savanna density parameters versus pollen abundances in

772 surface samples

773

774 Supporting Information

775 Appendix S1. Map of sampling locations, NW Kimberley Region

35

S.E. Connor et al. Forgotten land-use impacts in NW Australia

776 777

778

36

S.E. Connor et al. Forgotten land-use impacts in NW Australia

779 Appendix S2. Complete pollen and geochemical diagrams, MP11A record

e a la e e e d il a a c e b e e ia a h rn c c n ll s e m a a n y a f e n b a h u ) s a e a d p q te s a d e l a d F e r o S e e a s e o e e / a u g / ila l e a r h a u a S c t e e e M 2 y te e m s s u a r r a e g p e a / a la a a s s c / e e Z a e a u p u ic iz la a o a a a e e i c a m r c i ra e m u e e s s e a a a a / n n e a s ( c p h l a r p r h a a li a a a ta n o ig a e ia n o a r c c s c ia a ra a h in s e e ia i e ia c . s c s c ie r o s o c i in a e e s a e t a c m t c a b to h e e ti a u a r r e t l i c c r m m r a p m lo te o s o a h o p a s b is r n a c iu le c fe n ia a s s o e r g p c p e o n th la G G c n e ra g a a a r y la h s o g e c e o rm p ri p in s n s b s m r a i n a n ia il a a n m e lo n c o n o a s r iz a l ll ra ra m e ra r r a ic e h u p s z o l ro g d o s a o s o li e u b y lit u m o i o c v v m h y r id a o a h ri z c i A d an e a ge e a s o e e h s t c c ri ty i o y n u r o d c a r u Depth cal. BP r r r l s r d b a e l u c e t t d b p r i n 4 n r g r t m o l p p p r e i m h n t o e o d r r l u s re h e o a a e o u im c r a ri u re p n e o a u a h re e P 3 a o la m a s m o r a y y y e d ra tr y te la p o o p s p o o e e e s T S H C MyrtaceaeC C T D R T A BanksiaG M T B B S A P C F E B R A M C C P L F S L A A CaryophyllaceaeC D H C C T P In T U MyriophyllumN P TrileteP typeO M6 B S P S P S C G N A 0 -70 5 -60 10 -50 -40 15 -30 20 -20 -10 25 0 10 30 20 30 40 35 50 60 70 40 80 90 100 45 110 120 780 130 20 40 60 80 100 20 40 20 20 20 20 40 20 20 20 40 20 20 20 20 20 40 60 80 10002000300040005000 781 782

s s b e u Depth BP cal. r re h a a g T S Herbs Li Be Na Mg Al Si P K C Ti V Cr Fe Mn Co Ni Cu Zn Ge As Se Rb Sr Zr Cd Sn Sb B Au H Pb Bi Th U 0 -70 5 -60 10 -50 -40 15 -30 20 -20 -10 25 0 10 30 20 30 40 35 50 60 40 70 80 90 100 45 110 120 783 130 20 40 60 80 100 20 100 200 300 1000200030004000 200 400 600 800 500 1000 1500 1000 2000 3000 200 400 600 500 100015002000 200 400 600 50 100 150 200 100 200 300 500100015002000 100 200 300 400 20 20 40 60 80 20 40 60 20 40 60 0 20 20 20 40 0 100 200 300 400 0 0 20 0 784 785 786

37

S.E. Connor et al. Forgotten land-use impacts in NW Australia

787 Appendix S3. Principal Components Analysis result

788

789

38

S.E. Connor et al. Forgotten land-use impacts in NW Australia

790 Appendix S4. Pollen and vegetation data from surface samples

Pollen Morgan Basalt MP34 MP10 MBS MP11A MP12A MP10A MP37 MP12 River Water Swamp Corymbia 17 18 50 77 5 29 10 8 0 22 Myrtaceae 41 53 102 111 6 32 31 26 0 50 Callitris 4 6 1 7 0 8 0 0 1 0 Casuarina 2 4 1 1 0 0 1 0 0 0 Terminalia 6 1 6 6 0 2 2 0 0 0 Dodonaea 0 0 4 1 0 2 2 4 0 9 Rubiaceae 0 0 2 0 0 0 1 0 0 0 Timonius 1 0 0 1 0 0 0 4 0 3 Acacia 0 0 1 0 0 0 0 1 1 3 Banksia 0 1 2 0 0 2 0 0 0 0 Grevillea 1 2 0 2 0 0 1 2 0 2 Breynia 1 0 2 0 0 0 0 1 0 0 Spermacoce 3 1 2 0 0 1 1 2 0 1 Antidesma ghaesembilla 0 0 1 1 0 2 1 0 0 1 Petalostigma 1 0 1 0 1 0 0 1 0 0 Celtis 1 0 0 0 0 0 0 0 0 0 0 1 0 0 0 0 0 0 0 0 Moraceae 0 1 0 0 0 0 0 0 0 0 Goodeniaceae 0 0 1 0 1 0 0 0 0 0 Fabaceae 0 0 3 0 0 1 7 2 0 1 Lamiaceae 0 0 0 0 2 0 0 0 0 3 Barringtonia 0 0 0 0 1 0 0 0 0 0 Malvaceae 1 1 1 1 0 0 1 0 0 0 Brachychiton 0 0 0 0 0 0 1 1 0 0 Euphorbiaceae 0 0 0 0 0 0 1 1 0 1 Solanaceae 0 0 0 0 0 0 3 0 0 0 Rapanea type 0 0 0 0 0 0 2 0 0 0 Gentianaceae/Canscora type 0 0 0 0 0 0 13 0 0 0 Pandanus 3 2 12 13 1 7 7 23 0 16 Jasminum 0 0 0 0 0 0 6 0 0 0 Loranthaceae 6 27 2 1 0 1 0 0 0 0 CP4 (reticulated) 0 0 0 0 0 1 0 0 0 0 Small Grass 32 53 19 21 47 16 22 27 2 17 Large Grass 39 49 28 21 8 33 4 15 14 34 Asteraceae 0 8 1 0 0 0 0 0 0 0

39

S.E. Connor et al. Forgotten land-use impacts in NW Australia

Asteraceae (liguliflorae) 0 0 1 0 0 0 0 0 0 0 Amaranthaceae 1 2 1 1 0 1 3 1 0 2 Haloragis 3 2 0 3 0 0 0 2 0 4 Cyperaceae 9 9 11 8 62 14 1 5 0 4 Cyperaceae 2 0 1 0 0 0 6 0 2 0 3 Indeterminate 0 7 0 2 3 2 3 5 0 2 Liliaceae 0 35 1 2 0 0 1 1 0 0 Persicaria/Zygophyllaceae 0 0 0 1 0 0 0 1 0 0 Crinum 1 0 1 0 0 0 0 0 0 0 Utricularia 0 0 1 0 0 0 0 0 0 0 Myriophyllum 2 2 1 7 2 3 1 3 1 5 Drosera 0 0 0 0 0 0 0 0 0 1 Nymphaceae 0 0 0 1 2 3 2 0 0 1 Nymphoides 0 0 2 0 0 0 1 0 0 3 Pteris sp. 0 0 0 0 1 0 0 2 0 1 Ophioglossum 0 0 1 0 0 0 0 2 0 1 Ceratopteris 0 0 0 0 0 0 0 0 0 1 Trilete type 6 0 0 0 0 0 1 0 0 6 0 Botyrococcus 5 9 37 32 4 73 0 3 0 3 Pseudoschizaea 1 0 2 1 15 0 2 1 0 6 Sporormiella 0 2 0 0 1 0 1 1 0 1 Podaspora 0 0 2 0 8 0 1 1 0 2 Sordaria 2 3 4 1 6 5 3 2 0 9 Cercophora 0 0 0 0 6 0 0 0 1 3 Delitschia 0 0 1 0 0 0 1 2 0 1 Neurospora 2 0 1 1 67 0 3 4 0 2 Gelasinospora 0 0 0 4 1 1 2 0 0 3 Vegetation parameters for surface samples Morgan Basalt MP34 MP10 MBS MP11A MP12A MP10A MP37 MP12 River Water Swamp Canopy cover % 18.75 24.75 40.3125 30 21.66667 51.25 28.87143 22 18.33333 28.87143 Basal Area 3.875 4.8 6.0625 7.166667 5.333333 6.1 4.918333 5.375 3.333333 4.918333 Height (average) 11.225 12.295 15.725 13.23333 13.53333 14.22 10.99133 12.6 11.9 10.99133 Tree volume estimate 20.475 29 43.5125 42.58333 25.7 42.22 24.51167 32.95 20.33333 24.51167 Corymbia cover (DAFOR) 3 3.35 4 5 3 5 3 2.666667 2.25 3 Eucalypt cover (DAFOR) 2.25 3.2 3.714286 1 2.5 1 1 0 1.25 1 Terminalia cover (DAFOR) 0.25 2.25 2.714286 0 0.5 0 0 0 1 0 Poaceae cover (DAFOR) 4.5 5 4.857143 5 4.5 5 5 2.666667 5 5 Pandanus cover (DAFOR) 0.25 0.25 1.285714 3 2 3 2 3 1 2 791

40

S.E. Connor et al. Forgotten land-use impacts in NW Australia

792 Appendix S5. Geochemical data from surface samples

mg/kg Morgan R Basalt MP34 MP10 MBS MP11A MP12A MP10A MP37 MP12 Water Swamp Li 9 20 15 17 18 17 15 3 20 12 Be 1 1 1 1 2 2 1 0 1 1 Na 416 823 220 238 972 181 136 22 992 118 Mg 1297 9168 3920 4018 2294 3762 4390 108 13920 3480 Al 28405 44515 53612 51856 27583 48798 43287 10242 78854 34060 Si 582 523 165 203 300 312 178 239 192 281 P 133 347 100 518 896 298 172 25 67 66 K 671 1725 809 655 1832 726 484 145 1172 360 Ca 600 24664 2604 580 3412 1749 978 36 10655 715 Ti 300 748 464 490 52 495 510 103 624 439 V 102 293 118 124 17 117 121 37 226 95 Cr 50 53 66 192 17 198 196 24 75 134 Fe 29804 77635 28896 56956 8180 22927 33336 1224 58074 23757 Mn 476 6048 268 134 293 464 286 3 791 107 Co 15 48 18 15 13 20 21 1 29 14 Ni 15 40 34 45 8 50 48 4 45 39 Cu 16 63 44 32 23 40 32 5 69 25 Zn 13 56 35 34 17 38 33 3 51 26 Ge 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 As 2 7 2 4 1 2 3 1 3 2 Se 0.001 2 0.001 1 0.001 1 0.001 1 1 0.001 Rb 7 20 15 9 13 7 6 2 30 4 Sr 10 96 20 12 34 21 14 3 70 9 Zr 17 43 29 27 5 28 26 3 48 19 Cd 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 Sn 1 1 1 1 1 1 1 0.001 1 0.001 Sb 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 Ba 58 530 91 100 90 173 87 13 154 61 Au 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 Hg 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 Pb 4 14 8 12 6 13 11 1 12 8 Bi 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 0.001 Th 5 4 6 7 6 5 5 2 9 4 U 1 1 1 3 5 2 2 1 2 1

41

S.E. Connor et al. Forgotten land-use impacts in NW Australia

793 Appendix S6. Aerial photographs of study area in 1949 and 2017

794

42

Minerva Access is the Institutional Repository of The University of Melbourne

Author/s: Connor, SE; Schneider, L; Trezise, J; Rule, S; Barrett, RL; Zawadzki, A; Haberle, SG

Title: Forgotten impacts of European land-use on riparian and savanna vegetation in northwest Australia

Date: 2018-05-01

Citation: Connor, S. E., Schneider, L., Trezise, J., Rule, S., Barrett, R. L., Zawadzki, A. & Haberle, S. G. (2018). Forgotten impacts of European land-use on riparian and savanna vegetation in northwest Australia. JOURNAL OF VEGETATION SCIENCE, 29 (3), pp.427-437. https://doi.org/10.1111/jvs.12591.

Persistent Link: http://hdl.handle.net/11343/213924

File Description: Accepted version