You are on page 1of 7

Landscape Ecol (2009) 24:149155 DOI 10.

1007/s10980-008-9315-7

PERSPECTIVE

Landscapes in Time and Space


Lindsey Gillson

Received: 23 July 2008 / Accepted: 11 December 2008 / Published online: 8 January 2009 Springer Science+Business Media B.V. 2009

Abstract Landscape ecology has a temporal dimension, and the role of past processes in shaping landscapes is increasingly recognised. To date, the interface between landscape ecology and palaeoecology has proved most productive in understanding the impacts of climate change and in discovering the extent of past human impacts on ecosystems. Further areas of synergy are emerging. This Perspective gives selected examples of ve main areas of synergy between palaeoecology and landscape ecology: dynamic landscape mosaics; resilience and thresholds; biocomplexity; adaptive cycles; and in the landscape ecology of invasive spread. Keywords Landscape mosaics Biocomplexity Resilience Thresholds Adaptive cycles Palaeo-invasions

Introduction Landscape ecology has a temporal dimension, and the role of past processes in shaping landscapes is increasingly recognised (see the special issue of
L. Gillson (&) Plant Conservation Unit, Botany Department, University of Cape Town, Private Bag X3, Rondebosch 7701, South Africa e-mail: Lindsey.Gillson@uct.ac.za

Landscape Ecologywhy history matters in landscape ecology (Rhemtulla and Mladenoff 2007)). Like the Roman god of gateways and doorways, Landscape ecology, must be a Janus faced discipline (Fig. 1), which looks both to the past and present (sensu Pickett et al. 1992). Knowledge of the past elucidates how landscape patterns and processes evolved, and can help to predict and exibly inform future ecosystem management in a changing world. To date, the interface between landscape ecology and palaeoecology has proved most productive in understanding the impacts of climate change and in discovering the extent of past human impacts on ecosystems (Foster 2002; Gillson and Willis 2004); by looking to the past we have uncovered links between pattern and process in landscapes and thereby enriched and improved our understanding of changes in climate and land-use (Willis and Birks 2006). Landscapes are dynamic and heterogeneous through both space and time and further synergies between landscape ecology and palaeoecology are now emerging in the ecological and palaeoecological literature. Advances in theories of non-equilibrium ecology or ecological ux lend themselves to both the spatially explicit, scale-specic discipline of landscape ecology, as well as a re-reading of the Quaternary-time perspectives afforded by long-term palaeoecological data. Landscape processes are nested in a spatio-temporal hierarchy, from large-scale, slow processes like geological change, to smaller-scale, fast

123

150

Landscape Ecol (2009) 24:149155

Fig. 1 Like the Roman god Janus, landscape ecologists need to look to both the past and the future to spatial processes as a continuum through time

processes like plant-plant competition and succession. Comparisons between theoretical models and the palaeoecological understanding of different scales of temporal processes in the landscape may illuminate the complexities of the landscape and in turn feed back into ecological theory. Interdisciplinary studies of landscapes, utilising archaeological, palaeoecological and historical data provide an enriched understanding of landscape development and future trajectories, and the role of humans in shaping the landscapes that we see today. In this review, I will consider and give selected examples of ve emerging areas of synergy between palaeoecology and landscape ecology. First, in the area of dynamic landscape mosaics; second in the area of resilience and thresholds; how landscape processes can change dramatically when an ecological or environmental threshold is reached. Third, in biocomplexity; the link between landscape patterns and social systems. Fourth, adaptive cyclesthe four phases of system re-organisation. Finally, in the landscape ecology of invasive spread (With 2002). Scale, heterogeneity and the landscape mosaic To begin to understand the complexity of landscape response to disturbance patterns, a hierarchical perspective is needed, which can accommodate processes across a range of spatial and temporal scales (Allen

and Starr 1982; Gillson 2004; Kotliar and Wiens 1990; ONeill et al. 1986; Pickett et al. 1987, 1989; Urban et al. 1987; Willis and Whittaker 2002; Wu and Loucks 1995). In most landscapes, large-scale patterns of geological, topographical and catastrophic disturbance are overlaid by smaller-scale variations in microclimate and disturbance patterns. This structure was proposed for savanna landscapes, in which small scale patterns of disturbance by re and herbivores were nested within landscape and regional scale processes of climate and topography (Coughenour and Ellis 1993). Palaeoecological evidence was consistent with a hierarchical organisation, in which stability of vegetation at landscape scales could persist because woodland-grassland transitions at smaller spatial scales were out of phase (Gillson 2004; Watt 1947). Another good example of how palaeoecology has elucidated the hierarchical structure of processes is in Garua and Numundo, Papua New Guinea, where the effects of catastrophic volcanic events were overlain by differences in vegetation response at smaller spatial scales (Boyd et al. 2005). The effects of major eruptions at c. 5,900, 3,600, 1,700 and 1,400 cal yr. B.P. varied at smaller spatial scales, because of differences in seed banks, and the inuence of prehistoric people on the landscape mosaic (Boyd et al. 2005). An interdisciplinary synthesis of archaeology and phytolith analysis compared vegetation responses at local, sub-regional and regional scales. The authors deduced that topography and other local factors caused differential tephra deposition and this pattern was overlain by human partitioning of the landscape. On Garua island, when tephra accumulation continued after the initial volcanic eruption, landscapes were abandoned by people and impacts on vegetation were severe, whereas after a low impact eruption forest persisted and human occupation continued. On the mainland, however, recovery of the coastal lowland varied spatially, and showed a mosaic pattern that reected the modication of the landscapes by humans prior to eruptions (Boyd et al. 2005). Gardens and forest clearings, dominated by grasses and other pioneer species, recovered more slowly and remained abandoned by people for longer (Boyd et al. 2005). Here, we see a not only the effect of scale of processes but also a feedback between human activity and landscape pattern, process and

123

Landscape Ecol (2009) 24:149155

151

scale, a feature that will be explored further in the section on biocomplexity. The hierarchical approach to landscape ecology is further exemplied by research in the southern Appalachian mountains, where processes of ecological change were studied at micro-, meso-, macro-, and mega-scales using an array of techniques, including vegetation and geomorphic mapping, archaeology, ethnobotany, historical records of logging and other disturbances, and palaeoecological analyses (fossil pollen, plant macro-fossils and charcoal particles) (Delcourt and Delcourt 1988, 1998). In this area, the ruling hypothesis was that modern species composition and distribution resulted from two main factors: tectonic stability over 260 million years, which had provided habitat continuity; and topographic variability (slope aspect, altitude) combined with modern (post-European settlement) disturbance regimes, which created a diversity of habitats. However, the multi-proxy study revealed that myriad factors, including Holocene climate change, progressive disturbance by Native Americans and changes in geomorphic disturbance regimes had all contributed to the development of species richness. Alternate landscape states resulted from the passage of climatic and geomorphic thresholds, and these changes shaped todays landscape mosaic (Delcourt and Delcourt 1988). Furthermore, by studying the effects of Holocene climate change on the position of ecotones between alpine, boreal and temperate ecosystems, Delcourt and Delcourt (1988) identied the presence of critical thresholds of temperature and disturbance. This enabled the authors to predict the possible effects future climate change scenarios on ecotones. They suggest that future rapid climate warming could result in geomorphic instability and increased habitat heterogeneity, alterations which will favour disturbance adapted species. Feedbacks, thresholds and resilience Concepts of resiliencethe ability of ecosystems to absorb disturbanceand thresholdsthe points at which ecosystems rapidly change from one quasistable state to anotherhave emerged as important theoretical constructs that provide a framework for interpreting heterogeneity over space and time (Cumming et al. 2005; Groffman et al. 2006;

Lindenmayer and Luck 2005). Ecologists are increasingly aware that the resilience of ecosystems is nite; if environmental change or disturbances reach a critical threshold, small changes in environmental conditions or disturbance patterns can cause a disproportionately dramatic response. Ecosystems or species may respond continuously to disturbance, but equally they may buffer the change and remain essentially the same, due to ecological resilience. Furthermore, ecosystem re-organisation is most likely when extreme events, or Large Infrequent Disturbances (LIDs), like droughts, oods, hurricanes and wildres exceed critical thresholds and cause major changes in ecological behaviour (Gillson 2006; Landres et al. 1999; Turner et al. 1998; Turner and Dale 1998). When a threshold is crossed, systems may return to their original state (engineering resilience), or move to a new quasi-stable state or phase (Delcourt and Delcourt 1987, 1998; Frelich and Reich 1998; Holling 1973; Romme et al. 1998; Theurillat and Guisan 2001; Thomas 2001). Palaeoecological studies can help in evaluating the buffering capacity, or resilience, of ecosystems and in determining the position of critical ecological thresholds. In arctic regions, shallow freshwater ponds are hotspots of biodiversity, and are also important resources for migratory birds and water-dependent mammals. Paleolimnological data indicate that these ponds have been permanent features of the landscape for millennia, but unprecedented change began in the 19th century, and increased evaporation/precipitation ratios in recent decades have led to complete desiccation, a new stable state (Smol and Douglas 2007). Furthermore, desiccation of the surrounding vegetation has increased the probability of re, potentially turning these habitats from carbon sinks to carbon sources. Identifying the causes and consequences of ecological thresholds is important to policy makers, subsistence users, and ecosystem managers, because rapid transitions provide less time and opportunity for adaptation (Chapin et al. 2004). Ecosystem managers need to know the position of critical ecological thresholds, so that they can maintain desired states or facilitate benecial changes. In the Kruger National Park (KNP), South Africa, ecologists have developed process-orientated management goals, based on ecosystem properties, known as Thresholds of Potential Concern. These thresholds are points along a

123

152

Landscape Ecol (2009) 24:149155

continuum of ecological or environmental change, at which managers either intervene to guide ecosystem change, or at which they decide whether to alter their management objectives (Biggs and Rogers 2003). Concern for the impact of elephants on trees and shrubs led to the development of a TPC for elephant management; if woody vegetation cover drops below 80% of its highest ever value, managers will control elephant covers by culling or translocation, or will adjust the TPC. This TPC then raises the question of how dense woody cover was in the past, and how far present-day tree cover deviates from the highest ever value. Using fossil pollen data and landscape modelling software, Gillson and Dufn (2007) were able to show that at two study sites, woody vegetation cover remained above 20% of its highest value over 5,000 and 1,400 year periods. These data showed that recent increases in elephant abundance did not appear to have exceeded the resilience of the woody vegetation, since no dramatic decreases in woody vegetation had occurred. The research also highlighted alternative ways of thinking about elephant TPCs; for example, the value of highest ever depends on the length of the temporal record, and a more useful TPC might be developed based on changes in mean tree abundance over time, and/or changes in the variability of tree cover; ecosystems that begin to vary rapidlyindicated by an increasing amplitude or co-efcient of variation are more likely to cross a critical threshold and undergo rapid re-organisation (Brock and Carpenter 2006; Willis et al. 2007). Furthermore, the variability and abundance of tree cover were different at the two study sites, highlighting the importance of TPCs that are site specic, and sensitive to local conditions (Gillson and Dufn 2007). Biocomplexity; the link between ecological and social systems As illustrated in the above example from Garua and Numondo (Boyd et al. 2005), people make use of and modify the landscape mosaic in order to preserve or accentuate favoured habitat types. This feedback between ecological, geomorphological and social systems is increasingly recognised in the literature on linked socialecological systems and biocomplexity (Berkes and Folke 1998; Callicott et al. 2007;

Gunderson and Holling 2001), and can also be applied in interpreting palaeoecological and archaeological data in terms of feedbacks between anthropogenic activities and other landscape processes. In Siberia, productive herbivore-maintained mosaic dominated by steppe grassland was resilient through many glacial-interglacial cycles, but switched to a less productive mosaic dominated by moss-shrub tundra at the start of the Holocene, when mega-fauna disappeared from the landscape, possibly because of northward migration of huntergatherers (Chapin et al. 2004). A study of re-adapted species in the southern Appalachian highlands, North Carolina, USA, by Delcourt and Delcourt (1997) showed that the preColombian selective re management, in turn formed by the physiological characteristics have shaped the present-day heterogeneous landscapes. In this study, archaeological records were compared with pollen and charcoal data to address the question of how Native American use of re may have contributed to the persistence of re adapted species like Pinus rigida and Pinus pungens in a landscape where lightning induced res are rare (Delcourt and Delcourt 1997). The authors concluded that in pre-Columbian times, Native Americans selectively utilised different elements of the landscape, cultivating maize and other crops in alluvial valley bottoms and used re on upper slopes and ridges, where they hunted, and gathered acorns and chestnuts. Prescribed burning on selected portions of the topographic-edaphic gradient restored oak orchards and promoted regeneration of re adapted endemic pines, while north-facing slopes and more mesic areas of the escarpment probably did not burn, providing refuges for mesic hardwoods, that require disturbance by windthrow to regenerate, but are intolerant of re. This interaction between landscape characteristics and human management, elucidated through long-term data, helps us to understand the development and maintenance of a heterogeneous landscape, in which both readapted and re intolerant species could persist (Delcourt and Delcourt 1997). The ndings also have relevance to present landscape management, indicating how a carefully managed re regime, that is responsive to the landscape, and its history and topography, can promote heterogeneity and thereby biodiversity.

123

Landscape Ecol (2009) 24:149155

153

Adaptive Change According to the theory of adaptive change, periods of building and conservation of system processes are punctuated by collapse, reorganisation, innovation and rapid growth, known as adaptive cycles (Holling et al. 2001). Adaptive cycles often take decades, centuries or millennia, and it is only by considering the longterm pattern and process of change that they can be identied. A recent paper by Dearing (2008) mapped the millennial-scale patterns of land-use, erosion, and monsoonal intensity (reected in speleothem, pollen, magnetic susceptibility and sand content) in Yunnan, SW China, onto the adaptive cycle of conservation, collapse, reorganisation and rapid growth (Holling et al. 2001). Surface erosion showed two distinct phases, the rst was resilient to monsoon intensity and corresponded to landscapes undisturbed by people between 2,960 and 1,430 cal yr. B.P. From 800 cal yr. B.P., however, erosional intensity increases with monsoonal intensity, indicating a loss of resilience in a more open landscape, inuenced by anthropogenic activity. Critically, the loss of resilience occurred not when intensive agriculture was initiated, but during periods of social upheaval, when agricultural landscape was abandoned, allowing rapid erosion of steep slopes unbuffered by well-maintained terrace systems. Furthermore, this loss of ecosystem resilience appears to be hysteretic (irreversible), even with reforestation, because of erosion gullies in the degraded landscape. The study thereby elegantly demonstrated how changes in social capital and the corresponding changes in land-use interacted with environmental variables to drive an ecosystem across a threshold of reorganisation, to a new phase, itself maintained by emergent properties (Dearing 2008). Palaeoinvasions Alien species are one of the main causes of extinction. They cause biotic homogenisation, as well as huge economic losses (Olden et al. 2004; Perrings et al. 2005). Palaeo-data can reveal past variability in species distribution and abundance, information which helps to distinguish apparent invasions from cyclical changes, phase shifts and recovery from past disturbance (Landres et al. 1999). Palaeodata has revealed that numerous incidences of apparent

invasion are in fact a recovery from past disturbance, or part of a cyclical pattern of change, rather than an unprecedented and linear increase in the abundance of one species at the expense of others. It is only with a long-term perspective that areas where forests are still recovering from past anthropogenic disturbance can be distinguished from unprecedented tree invasions that could threaten ancient grassland ecosystems. In the Chaco Canyon, New Mexico, for example, increasing abundance of pinyon pine (Pinus edulis) in some areas was found to reect a recovery from past overexploitation that occurred 800 to 1,000 years ago (Allen et al. 2003; Swetnam et al. 1999). A further synergy between palaeoecology and invasion ecology is in understanding the landscape ecology of invasive spread (Gillson et al. 2008; With 2002). The relevance of palaeo-invasions to presentday invasions lies with the process of invasion itself; while the means and frequency of species introductions are unprecedented, the process of spread of species through a landscape remains comparable to the past (Rejmanek 1999; Richardson et al. 2000) and this raises possibilities for using the palaeo-literature to look for patterns and processes predicted by contemporary theory on invasion ecology (Gillson et al. 2008). The palaeo-record reveals the continuous behaviour of invasion over time (Mitchell et al. 2006), enabling the process of invasion to be studied temporally, and the shifting interplay of environmental and ecological interactions over time (With 2002). The palaeo-literature provides numerous examples of patchily distributed populations that were a precursor to invasive spread. Examples include beech (Fagus sylvatica) in Europe and eastern hemlock (Tsuga Canadensis) in the U.S.A. In the case of beech, expansion could only occur when ground res created suitable seed beds (Bjorkman and Bradshaw 1996; Bradshaw and Lindbladh 2005), while the spread of hemlock from small isolated populations could only occur when a critical climatic threshold had been crossed (Parshall 2002). In both cases, patchily distributed populations were a precursor to invasive spread, providing an interesting test case for percolation theory that predicts how species spread through landscapes based on the interaction between their dispersal abilities, and disturbances that create spaces into which species can spread (With 2002, 2004).

123

154

Landscape Ecol (2009) 24:149155 change? Ambio 33:361365. doi:10.1639/0044-7447 (2004)033[0361:GCATBF]2.0.CO;2 Coughenour MB, Ellis JE (1993) Landscape and climatic control of woody vegetation in a dry tropical ecosystem: Turkana District, Kenya. J Biogeogr 20:383398. doi: 10.2307/2845587 Cumming GS, Barnes G, Perz S, Schmink M, Sieving KE, Southworth J, Binford M, Holt RD, Stickler C, Van Holt T (2005) An exploratory framework for the emprirical measurement of resilience. Ecosystems (N Y, Print) 8:975987. doi:10.1007/s10021-005-0129-z Dearing JA (2008) Landscape change and resilience theory: a palaeoenvironmental assessment from Yunnan, SW China. Holocene 18:117127. doi:10.1177/095968360708 5601 Delcourt HR, Delcourt PA (1988) Quaternary landscape ecology: relevant scale in space and time. Landscape Ecol 2:2344. doi:10.1007/BF00138906 Delcourt HR, Delcourt PA (1997) Pre-Columbian Native American use of re on southern Appalachian landscapes. Conserv Biol 11:10101014. doi:10.1046/j.1523-1739. 1997.96338.x Delcourt PA, Delcourt HR (1987) Late-Quaternary dynamics of temperate forests: applications of paleoecology to issues of global environmental change. Quat Sci Rev 6:129146 Delcourt PA, Delcourt HR (1998) Paleoecological insights on conservation of biodiversity: a focus on species, ecosystems, and landscapes. Ecol Appl 8:921934 Foster DR (2002) Conservation issues and approaches for dynamic cultural landscapes. J Biogeogr 29:15331535. doi:10.1046/j.1365-2699.2002.t01-1-00788.x Frelich LE, Reich PB (1998) Disturbance severity and threshold reponses in the Boreal Forest. Conserv Ecol 2:7 Gillson L (2004) Evidence of hierarchical patch dynamics in an East African Savanna? Landscape Ecol 19:883894. doi: 10.1007/s10980-004-0248-5 Gillson L, Willis KJ (2004) As Earths testimonies tell: wilderness conservation in a changing world. Ecol Lett 7:990998. doi:10.1111/j.1461-0248.2004.00658.x Gillson L (2006) A large infrequent disturbance in an East African savanna. Afr J Ecol 44:458467. doi: 10.1111/j.1365-2028.2006.00662.x Gillson L, Dufn KI (2007) Thresholds of potential concern as benchmarks in the management of African savannahs. Philos Trans R Soc Lond B Biol Sci 362:309319. doi: 10.1098/rstb.2006.1988 Gillson L, Ekblom A, Willis KJ, Froyd C (2008) Holocene Palaeo-invasions: the link between pattern, process and scale in invasion ecology? Landscape Ecol 23(7):757769 Groffman PM, Baron JS, Blett T, Gold AJ, Goodman I, Gunderson LH, Levinson BM, Palmer MA, Paerl HW, Peterson GD, Poff NL, Rejeski DW, Reynolds JF, Turner MG, Weathers KC, Wiens J (2006) Ecological thresholds: the key to successful environmental management or an important concept with no practical application? Ecosystems (N Y, Print) 9(1):113. doi:10.1007/s10021-0030142-z Gunderson LH, Holling CS (eds) (2001) Panarchy: understanding transformations in human and natural systems. Island Press, Washington

Conclusions Landscape ecology is a discipline integrally concerned with space and time; understanding landscape patterns involves consideration of scale and process, heterogeneity, patch mosaics and resilience. Like Janus, the divine keeper of time in Roman mythology, paleoecologists and landscape ecologists stand in the threshold between past and future, and need both perspectives if we are to adapt to todays changing climate, land-use pressures and conservation goals.

References
Allen CD, Betancourt JL, Swetnam TW (2003) Landscape changes in the Southwestern United States: techniques, long-term data sets, and trends page. Land Use History of North America.http://biology.usgs.gov/luhna/chap9.html Allen TFH, Starr TB (1982) Hierarchy perspectives for ecological complexity. University of Chicago Press, Chicago Berkes F, Folke C (eds) (1998) Linking social and ecological systems: management practices and social mechanisms for building resilience. Cambridge University Press, New York Biggs HC, Rogers KH (2003) An adaptive system to link science, monitoring, and management in practice. In: duToit JT, Rogers KH, Biggs HC (eds) The Kruger experience: ecology and management of savanna heterogeneity. Island Press, Washington, pp 5980 Bjorkman L, Bradshaw R (1996) The immigration of Fagus sylvatica L., Picea abies (L.) Karst. into a natural forest stand in southern Sweden during the last 2000 years. J Biogeogr 23:235244. doi:10.1046/j.1365-2699.1996. 00972.x Boyd WE, Lentfer CJ, Parr J (2005) Interactions between human activity, volcanic eruptions and vegetation during the holocene at Garua and Numundo, West New Britain, PNG. Quat Res 64:384398. doi:10.1016/j.yqres.2005.08. 017 Bradshaw RHW, Lindbladh M (2005) Regional spread and stand-scale establishment of Fagus sylvatica and Picea abies in scandinavia : paleoperspective in ecology. Ecology 86:16791686. doi:10.1890/03-0785 Brock WA, Carpenter SR (2006) Variance as a leading indicator of regime shift in ecosystem services. Ecology and Society 11 online Callicott JB, Rozzi R, Delgado L, Monticino M, Acevedo M, Harcombe P (2007) Biocomplexity and conservation of biodiversity hotspots: three case studies from the Americas Philosophical Transactions of the Royal Society of London Series. B-Biological Sciences 362:321333 Chapin FS, Callaghan TV, Bergeron Y, Fukuda M, Johnstone JF, Juday G, Zimov SA (2004) Global change and the boreal forest: thresholds, shifting states or gradual

123

Landscape Ecol (2009) 24:149155 Holling CS (1973) Resilience and stability of ecological systems. Annu Rev Ecol Syst 4:123. doi:10.1146/annurev. es.04.110173.000245 Holling CS, Gunderson LH, Ludwig D (2001) Chapter 1: In search of a theory of adaptive change. In: Gunderson LH, Holling CS (eds) Panarchy: understanding transformations in human and natural systems. Island press, Washington Kotliar N, Wiens JA (1990) Multiple scales of patchiness and patch structure: a hierarchical framework for the study of heterogeneity. Oikos 59:253260. doi:10.2307/3545542 Landres PB, Morgan P, Swanson FJ (1999) Overview of the use of natural variability concepts in managing ecological systems. Ecol Appl 9:11791188 Lindenmayer DB, Luck GW (2005) Synthesis: thresholds in conservation and management. Biol Conserv 124:351 354. doi:10.1016/j.biocon.2005.01.041 Mitchell CE, Agrawal AA, Bever JD, Gilbert GS, Hufbauer RA, Klironomos JN, Maron JL, Morris WF, Parker IM, Power AG, Seabloom EW, Torchin ME, Vazquez DP (2006) Biotic interactions and plant invasions. Ecol Lett 9(6):726740 ONeill RV, De Angelis D, Waide J, Allen T (1986) A hierarchical concept of ecosystems. Princeton University Press, Princeton Olden JD, Poff NL, Douglas MR, Douglas ME, Fausch KD (2004) Ecological and evolutionary consequences of biotic homogenization. Trends Ecol Evol 19:1824. doi: 10.1016/j.tree.2003.09.010 Parshall T (2002) Late-Holocene stand-scale invasion by hemlock (Tsuga canadensis) at its western range limit. Ecology 83:13861398 Perrings C, Dehnen-Schmutz K, Touza J, Williamson M (2005) How to manage biological invasions under globalization. Trends Ecol Evol 20:212215. doi:10.1016/j.tree.2005.02. 011 Pickett S, Collins S, Armesto J (1987) A hierarchical consideration of causes and mechanisms of succession. Vegetatio 69:109114. doi:10.1007/BF00038691 Pickett S, Kolasa J, Armesto J, Collins S (1989) The ecological concept of disturbance and its expression at various hierarchical levels. Oikos 54:129136. doi:10.2307/3565258 Pickett ST, Parker VT, Fiedler PL (1992) The new paradigm in ecology: implications for conservation biology above the species level. In: Fiedler PL, Jain SK (eds) conservation biology. Chapman and Hall, New York, pp 6588 Rejmanek M (1999) Holocene invasions: nally the resolution ecologists were waiting for!. Trends Ecol Evol 14:810. doi:10.1016/S0169-5347(98)01517-1 Rhemtulla JM, Mladenoff DJ (2007) Why history matters in landscape ecology. Landscape Ecol 22:13. doi:10.1007/ s10980-007-9163-x Richardson DM, Pysek P, Rejmanek M, Barbour MG, Panetta FD, West CJ (2000) Naturalization and invasion of alien

155 plants: concepts and denitions. Divers Distrib 6:93107. doi:10.1046/j.1472-4642.2000.00083.x Romme WH, Everham EH, Frelich FL, Mortiz MA, Sparks RE (1998) Are large, infrequent disturbances qualitatively different from small, frequent disturbances? Ecosystems (N Y, Print) 1:524534. doi:10.1007/s100219900048 Smol JP, Douglas MSV (2007) Crossing the nal ecological threshold in high Arctic ponds. Proc Natl Acad Sci USA 104:1239512397. doi:10.1073/pnas.0702777104 Swetnam TW, Allen CD, Betancourt JL (1999) Applied historical ecology: using the past to manage for the future. Ecol Appl 9:11891206. doi:10.1890/1051-0761(1999) 009[1189:AHEUTP]2.0.CO;2 Theurillat J-P, Guisan A (2001) Potential impact of climate change on vegetation in the European Alps: a review. Clim Change 50:77109. doi:10.1023/A:1010632015572 Thomas MF (2001) Landscape sensitivity in time and space-an introduction. Catena 42:8398. doi:10.1016/S0341-8162 (00)00133-8 Turner MG, Baker WL, Peterson CJ, Peet RK (1998) Factors inuencing succession: lessons from large, infrequent natural disturbances. Ecosystems (N Y, Print) 1:511523. doi:10.1007/s100219900047 Turner MG, Dale VH (1998) Comparing large, infrequent disturbances: what have we learned? Ecosystems (N Y, Print) 1:493496. doi:10.1007/s100219900045 Urban D, ONeill RV, Shugart HH Jr (1987) Landscape Ecology. A hierarchical perspective can help scientists understand spatial patterns. Bioscience 37:119127. doi: 10.2307/1310366 Watt AS (1947) Pattern and process in the plant community. J Ecol 35:122. doi:10.2307/2256497 Willis KJ, Whittaker RJ (2002) Species diversity: scale matters. Science 295:12451246. doi:10.1126/science.1067 335 Willis KJ, Birks HJB (2006) What is natural? The need for a long-term perspective in biodiversity conservation science. Science 314:12611265. doi:10.1126/science.1122 667 Willis KJ, Kleczkowski A, New M, Whittaker RJ (2007) Testing the impact of climate variability on European plant diversity: 320, 000 years of water-energy dynamics and its longterm inuence on plant taxonomic richness. Ecol Lett 10:673679. doi:10.1111/j.1461-0248.2007.01056.x With KA (2002) The landscape ecology of invasive spread. Conserv Biol 16:11921203. doi:10.1046/j.1523-1739. 2002.01064.x With KA (2004) Assessing the risk of invasive spread in fragmented landscapes. Risk Anal 24:803815. doi: 10.1111/j.0272-4332.2004.00480.x Wu J, Loucks O (1995) From balance of nature to hierarchical patch dynamics: a paradigm shift in ecology. Q Rev Biol 70:439466. doi:10.1086/419172

123

You might also like