1,774
Views
79
CrossRef citations to date
0
Altmetric
Review

Mountain Treelines: A Roadmap for Research Orientation

, , , , , , & show all
Pages 167-177 | Accepted 01 Dec 2010, Published online: 16 Jan 2018

Abstract

For over 100 years, mountain treelines have been the subject of varied research endeavors and remain a strong area of investigation. The purpose of this paper is to examine aspects of the epistemology of mountain treeline research—that is, to investigate how knowledge on treelines has been acquired and the changes in knowledge acquisition over time, through a review of fundamental questions and approaches. The questions treeline researchers have raised and continue to raise have undoubtedly directed the current state of knowledge. A continuing, fundamental emphasis has centered on seeking the general cause of mountain treelines, thus seeking an answer to the question, “What causes treeline?” with a primary emphasis on searching for ecophysiological mechanisms of low-temperature limitation for tree growth and regeneration. However, treeline research today also includes a rich literature that seeks local, landscape-scale causes of treelines and reasons why treelines vary so widely in three-dimensional patterns from one location to the next, and this approach and some of its consequences are elaborated here. In recent years, both lines of research have been motivated greatly by global climate change. Given the current state of knowledge, we propose that future research directions focused on a spatial approach should specifically address cross-scale hypotheses using statistics and simulations designed for nested hierarchies; these analyses will benefit from geographic extension of treeline research.

Introduction

Mountain treeline ecotones represent the spatial transition from forested to treeless mountain landscapes and are, therefore, considered among the most prominent features of mountain environments. Mountain treeline ecotones also represent the upper physiological limits of tree species and a lower boundary (though not a physiological limit) for alpine herbaceous species. The importance of this boundary for nutrient fluxes and biodiversity, the variety and complexity of spatial patterns found at treeline, and the intriguing and still enigmatic life-form limit it represents, has given rise to numerous and diverse research endeavors from a broad disciplinary base for over 100 years. An important motivation for many recent studies of treeline is ongoing global climate change due to their hypothesized use as an indicator of climate change and to potential loss of biodiversity and ecosystem function as tundra is replaced by woody species. The usefulness of treeline dynamics as indicators of climate change (or of general ecological processes and relations) will depend on how they are approached and thus what we believe we know about this topic.

Our purpose here is not to review what is known, or thought to be known, about the geography, ecology, and plant physiology of mountain treeline ecotones; CitationHoltmeier (2009) has already covered this ground from a worldwide perspective. Instead, our aim is to examine the nature of knowledge, through a review of the paths treeline researchers have followed to reach the current state of knowledge about alpine treeline ecotones, and also how the goals for knowledge acquisition differ. For simplicity, we refer to these as studies of mountain “treeline,” and distinguish a conceptual line versus ecotone only where necessary.

APPROACHES

From the variety of research conducted at mountain treelines worldwide, two major research approaches emerge: searches for global-scale and landscape-scale causes, wherein the latter are within and modify the limits of the former (CitationTroll, 1973; CitationWardle, 1974, Citation1993; CitationHoltmeier, 2009). CitationKörner (2003) identified these as fundamental/global and modulative/regional, and we also see these approaches as inquiry at different spatial scales within a geographic hierarchy but with the modulation being more local than regional. Studies exploring the broad causes of treeline attempt to answer the fundamental question, “What causes treelines?” Such studies attempt to understand the factors that cause treelines to exist on a global level, and we refer to it here as the global approach. Furthermore, these studies focus primarily on limitations to tree growth from an ecophysiological perspective because ecophysiology is replicated globally, but ecophysiological questions are also examined locally. Studies examining the fine-scale (or local) causes of treelines attempt to answer the question, “How and why do treelines differ across locations?” This approach focuses on landscape patterns, especially the effects of topography and treeline history (such as climatic changes, human impact, and wildfire), because these factors differ within regions, and we refer to it as the landscape approach. It is not a coincidence that these two approaches have different foci in terms of treeline studies, with the former most concerned with the upper elevational limit of ‘trees,’ i.e., full upright stems (or at least 2 m; e.g., CitationWieser and Tausz, 2007), and less with tree seedlings, while the latter is most concerned with the three-dimensional gradient or pattern across the ecotone with seedlings at the fore (e.g., CitationJohnson et al., 2004).

Global treeline research is currently dominated by the search for mechanisms of low-temperature limitations on tree growth (e.g., CitationKörner, 1998a). Here the discussion centers on the role of carbon acquisition versus use (i.e. source versus sink limitation) at low temperatures (particularly root zone temperature) and/or limitations related to the utilization of carbon within trees. In spite of abundant sampling in recent years (CitationLi et al., 2002; CitationHoch and Körner, 2003; CitationPiper et al., 2006; CitationBansal and Germino, 2008; CitationShi et al., 2008), there is generally no evidence for a poor carbon status in treeline trees in the long term, although it may influence local pattern (CitationCairns and Malanson, 1997; CitationCairns, 2005). Therefore, the attention is now turned to mechanisms of direct growth limitations, with an emphasis on understanding how a mean growing season temperature can best correlate with treeline positions worldwide (CitationKörner and Paulsen, 2004), in spite of this mean being composed of widely differing temperature regimes (CitationHoch and Körner, 2009).

Investigations following the landscape approach have typically studied treelines from the perspective of spatial pattern and process (CitationWalsh et al., 1997). Landscape change, linkages among biotic and physical elements of the environment, and human/environment interactions are often studied (CitationHoltmeier, 1974; CitationBroll et al., 2007; CitationHoltmeier, 2009). Landscape studies may also look at specific ecophysiological causes of treeline patterns, e.g., studies that address establishment limitations rather than growth in adult trees (e.g., CitationGermino et al., 2002; CitationPiper et al., 2006; CitationJohnson and Smith, 2007; CitationBader et al., 2007a; CitationBansal and Germino, 2008; CitationSmith et al., 2009). Such studies often find that landscape position and facilitation by neighbors are very important.

Though a divergence in knowledge ‘camps’ currently exists, present-day research has evolved from a similar line of inquiry—one that originated and continues in the consideration of global-scale causes of treelines. A search for broad-scale causes of treelines evolved from the work of great observationists. After Leonardo da Vinci (1452–1519) had noticed the altitudinal belts and their specific organisms on Monte Rosa (Italian Alps), Conrad Gesner (1516–1565) from Zurich mentioned in his “descriptio montis fracti” (1555) that the altitudinal position of the vegetation belts is related to the decrease of temperature and length of growing season with altitude (CitationHoltmeier, 1965). It is likely that Alexander von Humboldt (Citationvon Humboldt and Bonpland, 1805) was among the first researchers to mention the existence of alpine treelines in a manner that connected observation with causality. His foundational work in physical geography and meteorology was the first to document isotherms and their relationship with elevation and plant geography of mountain slopes (CitationMarek, 1910; CitationTroll, 1962). Additional work included the early identification of heat (CitationDäniker, 1923) and tree physiology (e.g., CitationMichaelis, 1934; CitationSteiner, 1935; CitationSchmidt, 1936; CitationPisek and Cartellieri, 1939), and more recent summaries combine the two (CitationTranquillini, 1979; CitationWieser and Tausz, 2007). CitationHoltmeier (2009) provides a more complete treatment of more recent research.

Here, we devote most of our analysis to the fundamental questions and themes in the landscape approach. We conclude by outlining needs for future research on mountain treelines conducted with a spatial perspective. In this paper, our emphasis on the landscape perspective in no way discredits the excellent contributions to understanding treelines offered by the global approach, and reflects, rather, the collective training and expertise of the authors.

LANDSCAPE FOUNDATIONS

The landscape approach, our main focus, is first and foremost a part of landscape ecology. Much of the early impetus for work in the Alps was from the perspective of landscape level management problems (CitationHoltmeier, 2009), the root of European landscape ecology. Much of the pattern-process paradigm in landscape ecology derived from island biogeography, and so it blossomed in the 1970s and boomed in the 1980s (CitationTurner, 1989). This latter history is mirrored in the increasing interest in three-dimensional pattern in treeline studies (e.g., CitationHumphries et al., 2008; see ). However, landscape ecology has been affected by ideas formed in mountain environments, perhaps beginning in the mid-20th century with CitationTroll's (1971) emphasis on ‘geo-ecology.’

FIGURE 1 Hedges, illustrating positive spatial association, and individual seedlings, illustrating association with microrelief, extend into tundra above the contiguous forest in the treeline ecotone at Lee Ridge, Glacier National Park, Montana, U.S.A. Photo: G. P. Malanson.

FIGURE 1 Hedges, illustrating positive spatial association, and individual seedlings, illustrating association with microrelief, extend into tundra above the contiguous forest in the treeline ecotone at Lee Ridge, Glacier National Park, Montana, U.S.A. Photo: G. P. Malanson.

The landscape approach, second, includes hierarchy theory in ecology, which developed in the 1980s (e.g., CitationAllen and Starr, 1982) and has been applied to treeline research since the early 1990s. The core idea here is that processes and patterns develop at multiple spatial and temporal scales. Together they are comprised of interactions of processes at finer scales, while constrained by patterns at coarser scales (CitationO'Neill et al., 1989). For example, the establishment of a forest patch in the treeline ecotone would depend on fine-scale processes such as seedling establishment but is also constrained by coarse-scale patterns such as mountain topography (slope aspect, exposure to wind). Also, the general temperature control on treelines can be seen as ‘merely’ a coarser scale constraint on the focal three-dimensional pattern dynamics driven by population level processes (e.g., CitationBrown et al., 1994; CitationWalsh et al., 1997).

Third, complexity science has, explicitly or implicitly, become part of many landscape-scale treeline studies. Complexity science is a body of concepts that examines how higher order pattern or structure in systems is produced by few, simple, but nonlinear interactions at a lower level, and thus includes hierarchy (e.g., CitationSole and Bascompte, 2006). The higher order patterns could not be predicted from the properties of the lower order units, but instead the ‘emergence’ of structure must be determined by observing the evolution of the system. Thus, a standard reductionist approach is overturned. Such systems can be said to be self-organized in that the higher order structure can be reduced to fewer, but still endogenous, nonlinear dimensions or drivers. Complexity science was developed and distinctly promoted by physicists (CitationGell-Mann, 1994; CitationBak, 1996), but quickly spread to other disciplines. CitationMalanson (1999) discussed its applicability to treeline studies, and the work by CitationRietkerk et al. (2002) in spatial ecology influenced later advances by CitationZeng and Malanson (2006) and CitationBader et al. (2008). Advances in all three domains developed simultaneously and were clearly being used together by landscape ecologists. The work of O'Neill (O'Neill et al., 1982, Citation1986, Citation1989, Citation1992, inter alia) alone illustrates this synergy, and work in landscape ecology (e.g., CitationLi, 2002) led to the initiation of a new journal, Ecological Complexity, in 2004.

One can see why linkages among these three areas (landscape ecology, hierarchy theory, and complexity) have been identified and are relevant to treeline research. They link spatial heterogeneity and feedbacks (CitationTroll, 1971), scale dependence in a geographic hierarchy (CitationPattee, ed., 1973), and self-organization (CitationHaken, 1975). The relations between processes and patterns, in this case between advances or retreats of treelines on mountain slopes and the spatial structure of patches and edges (), can be seen to have characteristics in common with other spatial phenomena at higher orders (e.g., fractal patterns or power-law distributions). Ecologists have found such higher order phenomena may explain important characteristics such as metabolic capacity (CitationWest et al., 1999) for which evolutionary processes may be revealed. For biogeography (e.g., CitationDeng et al., 2006), including ecotones (CitationMilne et al., 1996), the existence of these higher order patterns may indicate constraints on the variability and explanations possible (i.e., self-organization; CitationZeng and Malanson, 2006). We might hope that these linkages would provide predictive power for the response of treelines to climate change, but prediction may be limited. While self-organization, and thus fractal patterns, hold under strong exogenous forcing from either climate change or underlying geomorphic patterns (CitationZeng, 2010), the constraint that patterns will remain fractal does not provide the kind of prediction that ecologists or landscape managers desire.

Nevertheless, from these three perspectives we have learned to look at treelines at multiple scales, seeing organisms within patches within landscapes in three dimensions, and how the coarser levels constrain the finer scale dynamics that create and reproduce them (e.g., CitationMalanson et al., 2007). We further see the importance of dynamics as a conceptual framework, with both straightforward and higher order patterns of change as objects of study and both endogenous feedbacks and exogenous drivers as hypothetical processes.

These epistemological foundations were enabled, if not driven, by methodological developments. Common to landscape ecology in general, the landscape approach in treeline studies was advanced by remote sensing—aerial photography at first (e.g., CitationTroll, 1939), followed by satellite-based imagery (e.g., CitationBaker and Weisberg, 1995; CitationAllen and Walsh, 1996; CitationWalsh et al., 2003) and geographic information systems (GIS) analysis (e.g., CitationBrown, 1994; CitationHörsch, 2003). Quantification of pattern was an essential contributor to the concepts, and these technologies greatly increased the available data on spatial pattern and simplified its analysis. Computer power itself enabled the development of hierarchy theory and complexity science in general (Pagels, 1988). For example, cellular automata, or cellular models with stochasticity added to neighborhood effects, were core tools in the development of the complexity approach at ecotones in general (e.g., CitationLoehle et al., 1996; CitationLi, 2002) and treelines in particular (CitationAlftine and Malanson, 2004; CitationMalanson and Zeng, 2004; CitationZeng and Malanson, 2006; CitationWiegand et al., 2006; CitationZeng et al., 2007; CitationBader et al., 2008).

Understanding Landscape-Scale Relations

In landscape-scale treeline studies, an understanding of the treeline phenomenon and its local causes is obtained by focusing on variations of treeline spatial and temporal patterns (e.g., CitationBlüthgen, 1942; CitationGriggs, 1946; CitationHoltmeier, 1965; CitationSmith et al., 2003; CitationBroll et al., 2007; CitationMalanson et al., 2009; CitationButler et al., 2009a; see CitationHoltmeier, 2009, for further references). Like broad-scale approaches to treeline, treeline dynamics under global climate change scenarios have been a major theme of current research from a landscape perspective. However, for landscape-scale treeline scientists (who acknowledge that heat deficiency is the ultimate limit on tree growth at high elevation), the problem is that treeline ecotones present complex patterns in three dimensions that are beyond the explanatory power of temperature alone. The ribbons, hedges, and other patchy patterns (see CitationHoltmeier, 1982, and ), while constrained by temperature limits, require explanation. The landscape approach to treeline is motivated by interest in these three-dimensional patterns, and processes are seen as nested at multiple scales even within the landscape-scale domain (e.g., Elliott and CitationKipfmueller, 2010).

LOCAL DRIVERS

The primary fine-scale modulators of treeline patterns may be those that affect heat deficiency, but fine-scale modulators have effects on water, nutrients, and disturbance. Temperature deficiency is not only a broad-scale cause of treelines, but can also be a landscape-scale modulator of treeline pattern. At a landscape scale it is affected by local modulators such as topography, and it can also affect other landscape-scale modulators, such as distribution, depth, and duration of the winter snowpack. Factors that directly reduce temperatures at local scale are wind, cold air drainage, snow cover in summer, and shade. Temperature affects snow cover directly, in terms of determining the mix of rain vs. snow and the rate of melting and, indirectly, by determining snow density and thus snow removal and redeposition by wind.

Landscape-scale modulators not directly related to heat are quite diverse, but the one most important in ecophysiology is water (e.g., CitationBrodersen et al., 2006). Treeline areas vary in the soil moisture available for plant growth at landscape scales, and much depends on the removal and redeposition of snow by wind. Some areas are scoured of snow and so experience drought conditions much beyond what the regional precipitation pattern would indicate, whereas other areas have multiples of the regional precipitation due to snow collection and eventual melt (e.g., CitationWalsh et al., 1994; CitationHiemstra et al., 2002, Citation2006; CitationGeddes et al., 2005; CitationHoltmeier, 2005). Other effects of excessive snow cover, which may impede or even prevent successful seedling establishment, include shortened growing season, physical damage through creeping and settling snow, and snow fungi, which may affect seedlings and saplings of evergreen conifers (e.g., CitationCunningham et al., 2006).

Geomorphology is also a landscape-scale modulator of treeline patterns and dynamics, exerting its influence through more direct causes like snow, soil, and disturbance (e.g., CitationKullman, 1997; CitationWalsh et al., 2003; CitationHoltmeier and Broll, 2005; CitationButler et al., 2007; CitationZeng et al., 2007; CitationHumphries et al., 2008; CitationButler et al., 2009a, Citation2009b; CitationBekker and Malanson, 2009; CitationMunier et al., 2010). Local topography may provide shelter from the wind, as on the lee sides of small ridges, solifluction terraces, rocky outcrops and other convex sites (e.g., CitationResler et al., 2005, in the Rocky Mountains; CitationHoltmeier et al., 2003; CitationAnschlag et al., 2008, in Finland; see ). Depressions may be covered too long with snow, however, and waterlogging can be an additional adverse factor in such poorly drained places. On wind-exposed convex topography with little or no winter snowpack, wind actions (physiologically, mechanically) may cause serious damage to seedlings and saplings (CitationCairns and Malanson, 1998). Temperatures—their pattern as well as limits—also affect local geomorphic processes such as solifluction and frost heaving that may affect seedling establishment (CitationButler et al., 2004, Citation2009b).

Further local causes of treeline not directly related to heat are due to biotic factors such as pathogens, insects, mycorrhizae, birds, and mammals. Pathogen outbreaks can cause regional mortality of trees that could ultimately influence spatial pattern at alpine treeline. Cronartium ribicola, the introduced pathogen that causes blister rust in five-needled white pines, has caused widespread mortality in whitebark pine (Pinus albicaulis), a foundation and keystone species of northern Rocky Mountain subalpine and treeline ecosystems (CitationResler and Tomback, 2008), which has potentially serious and cascading consequences (CitationTomback and Resler, 2007). Mass outbreaks of the autumnal moth (Epirrita autumnata) have repeatedly destroyed vast areas of mountain birch (Betula pubescens ssp. czerepanovii) forests up to the treeline in Finland, which has been followed by severe soil erosion on sandy substrates that now impedes the re-establishment of this species (CitationHoltmeier, 2002; CitationHoltmeier et al., 2003; CitationBroll et al., 2007). Mycorrhizae, which increase nutrient availability, may be important to successful seedling establishment, tree growth, and afforestation at and above the treeline (CitationHasselquist et al., 2005; CitationGermino et al., 2006). Birds affect trees in the treeline ecotone by consuming seeds and buds, and dispersing and caching seeds (e.g., the nutcrackers [Nucifraga caryocatatces and subspecies in Eurasia, Nucifraga columbiana in North America]; CitationHoltmeier, 1966, Citation2002; CitationTomback, 1977; CitationMattes, 1978, Citation1982, Citation1985). The impacts of burrowing animals can be positive because they expose the mineral soil and thus create open patches that may facilitate the establishment of seedlings (e.g., CitationButler et al., 2009b; CitationButler and Butler, 2009); conversely they can destroy seedlings and saplings by girdling and by pushing seedlings out of the ground (Holtmeier, 1987, Citation2002). At the other end of the size scale, grizzly bears tear swaths of tundra with similar mixed effects (CitationButler 1992, Citation1995). Also, the activities of wild-living ungulates are detrimental to treelines (CitationHoltmeier, 2002, Citation2009; CitationCairns and Moen, 2004; CitationCairns et al., 2007). These phenomena are spatially variable and their relations with temperature are not well known.

Additionally, competition with alpine herbaceous species can limit or facilitate seedling establishment. Seedlings are facilitated by a moderate amount of herb cover and limited by too little or too much (CitationGermino et al., 2002). These relations can affect pattern development (CitationMalanson and Butler, 1994). Moreover, allelopathic effects of the associated vegetation (e.g., some lichen species or dwarf shrubs), may impair germination, mycorrhiza development and seedling development (CitationHoltmeier, 2009, further references therein).

SPATIAL DYNAMICS

Change through time is integral to explaining three-dimensional treeline patterns in a spatial hierarchy (CitationArmand, 1992; CitationWiegand et al.. 2006). The feedbacks of growing tree populations on their neighborhood become increasingly important as the size of individuals and patches increases (e.g., CitationHoltmeier, 1999, Citation2005, Citation2009), but fade as these patches merge into contiguous forest because of reduced edge (CitationZeng and Malanson, 2006). These feedbacks imply a strong effect of existing patterns on dynamics and indicate self-organization, and more generally, an important role for landscape history.

Patterns and Feedbacks

Treelines frequently exhibit very discrete patterns, consisting, for instance, of abrupt forest edges or distinct dwarf tree or krummholz patches (e.g., CitationWalsh et al., 1992; CitationHumphries et al., 2008). The variety of three-dimensional patterns observed are unlikely to emerge on environmental gradients of temperature due to lapse rates unless positive feedback processes amplify initial environmental differences (i.e., a positive feedback switch sensu CitationWilson and Agnew, 1992) or growth responses are nonlinear (CitationBatllori et al., 2009). The most notable driver in the pattern-process feedback is wind (e.g., CitationMarr, 1977; CitationAkhalkatsi et al., 2006; CitationHoltmeier and Broll, 2010); shade and protection from sky exposure is probably next (e.g., CitationÖrlander, 1993; CitationGermino and Smith, 2000; CitationGermino et al., 2002; CitationSmith et al., 2003; CitationSlot et al., 2005); followed by nutrient enrichment (CitationHoltmeier and Broll, 1992; CitationSeastedt and Adams, 2001; CitationShiels and Sanford, 2001; CitationLiptzin and Seastedt, 2009) and albedo. The negative feedback (shading and cooler soils) identified by CitationKörner (1998a, Citation1998b) acts in the opposite direction. However, tree seedlings at treeline are often found preferentially beneath tree canopies (e.g., CitationGriggs, 1946; CitationBall et al., 1991; CitationGermino and Smith, 1999; CitationResler and Fonstad, 2009), indicating that the positive feedback through shelter and protection from radiative stress is more important for treeline spatial dynamics than the negative feedback of lower soil temperatures.

Several treeline studies have investigated the role of landscape position, including effects of neighboring plants, on demographic processes (e.g., CitationDaly and Shankman, 1985; CitationBekker, 2005; CitationResler et al., 2005; CitationMaher and Germino, 2006; CitationBatllori et al., 2009, Citation2010; CitationHughes et al., 2009). However, few of these studies have included indices or direct measurements of ecophysiological parameters like photosynthetic capacity (CitationGermino and Smith, 1999; CitationMaher et al., 2005) or pigmentation (CitationAkhalkatsi et al., 2006), thus indicating the mechanisms by which spatial associations have probably arisen (e.g., protection from cold-induced photoinhibition). Combinations of such detailed mechanistic knowledge could strongly improve predictions of response to climate change. Another interesting issue is the effect of landscape position on the growth and general performance of adult trees, bringing local detail to the general ecophysiological principles that may (or may not) be applicable at treelines globally (CitationLi and Yang, 2004; CitationWilmking et al., 2004).

Given the difficulty of capturing feedbacks in space and time with field data, several authors have used simulations to try to understand the effects of feedbacks between spatial patterns and dynamics (e.g., CitationMalanson, 1997; CitationMalanson et al., 2001; CitationAlftine and Malanson, 2004; CitationMalanson and Zeng, 2004; CitationWiegand et al., 2006; CitationBader et al., 2008; CitationElliott, 2009; CitationDiaz-Varela et al., 2010). These computer simulations predict that alpine treelines exhibit unusual dynamics. CitationZeng and Malanson (2006) found that a model that included both positive and negative feedback could generate many observed patterns (notably fractals, cf. CitationAllen and Walsh, 1996) in a single long-term realization driven only by the endogenous feedback. CitationZeng et al. (2007) further found that such self-organization maintained higher order pattern relations even when exogenous geomorphic patterns might be expected to alter it (at least within realistic ranges). CitationBader et al. (2008) found that positive feedbacks could lead to abrupt transitions that decouple rates of advance from climate change (). These modeling efforts are best taken as hypothesis generators, rather than tests, and indicate the likely important feedbacks.

FIGURE 2 Emergence of abrupt treeline transitions in time and space. (a) Tree growth is determined by the microclimate, which is determined by the external macroclimate [e] and modifications caused by the vegetation itself [i]. Depending on the strength of i relative to e, the vegetation will be more (large e) or less (large i) coupled to the external climate and hence more or less sensitive to climatic changes. (b) Hysteresis fold demonstrating the alternative stable states of forest and alpine vegetation existing under the same external environmental conditions (gray area; at treelines the stress factor can be e.g., freezing temperatures, wind, or high solar radiation). Abrupt (‘catastrophic’) transitions from forest to alpine vegetation or vice versa can occur at threshold conditions or due to disturbances. (c) Changes in microclimatic conditions as a result of vegetation cover. Some stress factors can be aggravated just outside tree stands (gray graph) due to redirection of wind and snow. Adapted from CitationBader (2007).

FIGURE 2 Emergence of abrupt treeline transitions in time and space. (a) Tree growth is determined by the microclimate, which is determined by the external macroclimate [e] and modifications caused by the vegetation itself [i]. Depending on the strength of i relative to e, the vegetation will be more (large e) or less (large i) coupled to the external climate and hence more or less sensitive to climatic changes. (b) Hysteresis fold demonstrating the alternative stable states of forest and alpine vegetation existing under the same external environmental conditions (gray area; at treelines the stress factor can be e.g., freezing temperatures, wind, or high solar radiation). Abrupt (‘catastrophic’) transitions from forest to alpine vegetation or vice versa can occur at threshold conditions or due to disturbances. (c) Changes in microclimatic conditions as a result of vegetation cover. Some stress factors can be aggravated just outside tree stands (gray graph) due to redirection of wind and snow. Adapted from CitationBader (2007).

Historical Legacy

While self-organization may be maintained in principle, the evolution of pattern is also contingent on exogenous forces. The position and structures of present treelines often are the result of historical legacy rather than of the present climate (e.g., CitationHoltmeier, 1974, Citation2009; CitationHoltmeier and Broll, 2007). Extreme natural events such as severe storms, drought, extremely snow-rich or poor winters, natural and human-induced forest fires, mass outbreaks of leaf-eating insects, debris flows, snow avalanches, rock avalanches, and volcanic eruptions have long-lasting effects on current treeline ecotones (e.g. CitationButler and Walsh, 1994; CitationDaniels and Veblen, 2003; CitationStueve et al., 2009; CitationColombaroli et al., 2010). For example, the destruction of the soil organic layer by severe fires can result in an almost complete loss of nutrient supply, reduced water-holding capacity of the soils, and consequent increased surface runoff and soil erosion (e.g., CitationHoltmeier, 2009; CitationHoltmeier and Broll, 2005; CitationHoltmeier et al., 2003; CitationBroll et al., 2007). Thus, the legacy of specific patterns in specific situations becomes a dominant local control. At a broader temporal scale, many regions of the Rocky Mountains possess ‘relict treelines’ formed by long-lived pines (Pinus aristata, P. albicaulis, P. flexilis), subalpine fir (Abies lasiocarpa), and Engelmann spruce (Picea engelmannii), which became established at higher elevations under a warmer-than-present climate many centuries or even millennia ago (e.g., CitationIves, 1973; CitationIves and Hansen-Bristow, 1983; CitationHoltmeier, 1985, Citation1999, Citation2009). During the subsequent less favorable climatic conditions, subalpine fir and Engelmann spruce were able to reproduce by layering (i.e., formation of adventitious roots) in krummholz form. Some of these trees now produce viable seeds and facilitate seedling establishment by providing shelter from strong winds.

Apart from natural landscape processes and feedbacks, human land use has in many regions exerted a strong influence on treeline patterns. In the European Alps and many other Eurasian high mountains, which were already settled in prehistoric time, treeline has been lowered through pastoral use, mining, and burning the high-elevation forest. The present upper limit of the forest has become an ecological boundary that is as distinct as was the original climatic forest limit, at least in the Alps, Pyrenees, and Andes (CitationHoltmeier, 1965, Citation1974; CitationCamarero and Gutierrez, 2002; CitationDi Pasquale et al., 2008). A treeline depression by 150 to 300 m below the uppermost postglacial level of the climatic treeline can be taken for an average value (CitationHoltmeier, 1974, Citation1986; CitationBurga, 1988; CitationTinner et al., 1996; CitationCarcaillet et al., 1998; CitationBurga and Perret, 2001; CitationKaltenrieder et al., 2005). In tropical mountains the history of human settlement and its impact on treeline habitats is less clear. Humans are thought to have spread through South America before the beginning of the Holocene (e.g., CitationJackson et al., 2007), and the earliest evidence of fires at current treeline altitudes stem from this time, although clear signals of regional agriculture only appear in the second half of the Holocene (CitationDi Pasquale et al., 2008). Paleoecological records of past treeline altitudes are heterogeneous, but it is tempting to assume that humans have used tropical alpine habitats from very early times and at least locally have slowed or prevented a rise in treeline altitude from late Pleistocene levels (CitationHorn, 1993; CitationDi Pasquale et al., 2008). In some regions, however, recent destruction of tropical mountain forests by human land use is evident and has depressed treeline altitude considerably or in some cases has combined with deforestation from below to remove the forest belt altogether (e.g., CitationMiehe and Miehe, 2000).

Underexplored Areas and Directions for Future Research

As noted, a motivation for much ongoing treeline research is anticipated global climate change. The key question resulting from this motivation is, “Will a warmer world result in globally comparable responses of treelines?” Given the hypothesized general causal control, we would expect a comparable upward movement of alpine treeline ecotones worldwide—comparable in the sense that the rise in elevation of controlling isotherms, though variable, would produce rises in treelines. However, given the landscape-scale controls, an increase in temperature will change the broad constraint, but the response in any one area will vary through the interaction of the fine- and broad-scale controls. Such a varied response is indeed observed when comparing treelines worldwide (CitationHarsch et al., 2009). A good example is the observed dieback of treeline stands due to drought conditions accompanying temperature increases in the Sierra Nevada (CitationLloyd and Graumlich, 1997; cf. CitationJohnson et al., 2004; CitationBrodersen et al., 2006; CitationJohnson and Smith, 2007; CitationMillar et al., 2007). Differential responses, such as the limited advance of treeline in hedges on some areas with only densification but no advance in others in Glacier National Park, U.S.A., indicates that local-scale controls are more important than global temperature control here, at least in the short term (CitationButler et al., 1994; CitationKlasner and Fagre, 2002; CitationAlftine et al., 2003; CitationBekker, 2005). So although topography will change the spatial expression of the rise in elevation of any isotherm, the nonlinear relations created by positive feedbacks, in the context of existing spatial patterns and their legacies, will further complicate the dynamics (e.g., CitationBader and Ruijten, 2008; and CitationBader et al., 2008; CitationKharuk et al., 2010).

One productive area for new research would be in differentiating the responses to current climate change from those of past human impact; both climate and land use are major aspects of global change (CitationVitousek, 1994). Where treelines have been lowered by human activities such as grazing and burning, their response to release from these impacts may have similarities with responses to climatic warming. Differentiating the two responses could be informative in terms of understanding the relative importance of processes and how they relate to ecological theory as well as providing a sound basis for monitoring and mitigating climate change impacts. To pursue this line of research will require analyses that compare geographic areas with different impacts. A starting point exists in current research, primarily in those locations with long records of human occupation, such as the European Alps (e.g., CitationDidier, 2001; CitationHeiri et al., 2006; CitationDullinger et al., 2003) and in the Andes (e.g., CitationYoung, 1993; CitationSarmiento, 2000; CitationSarmiento and Frolich, 2002; CitationYoung and León, 2007, and references therein), but treelines in Asia and Africa also are part of unique cultural landscapes. Because treeline forms are historically contingent, extensive sampling will be needed to gain enough statistical power to make sense of these interactions. While some such expansions could be within continents (e.g., CitationBader et al., 2007b [plus an island]; CitationWeiss, 2009), expansion across continents to understudied areas, such as the southern hemisphere and remote tropical areas, especially in Africa and Asia, is potentially most fruitful (cf. CitationOhsawa, 1990; CitationSchmidt-Vogt, 1990; CitationMiehe and Miehe, 1994; CitationRundel et al., 1994; CitationSchickhoff, 1995, Citation2005; CitationWinkler, 1997; CitationWardle et al., 2001; CitationDiaz et al., 2003; CitationHofstede et al., 2003; CitationBaker and Moseley, 2007). Moreover, investigation of the treelines on oceanic islands, where the altitudinal treeline position is usually several hundred meters lower than the continental high-elevation treeline at the same latitude, needs to be intensified (e.g., Azores, Canary Islands, Hawaii, etc.; cf. CitationHenning, 1974; CitationLeuschner and Schulte, 1991; CitationLeuschner, 1996; CitationBader et al., 2007b). Not least, the upper treeline in New Guinea would be a valuable site for field research because it is the largest tropical island with a treeline located above 3000 m and human-induced fires play an important role for treeline physiognomy and dynamics (CitationPaijmans and Löffler, 1972).

Given the three major conceptual domains that help define and inform the landscape-scale approach to treeline research (landscape ecology, hierarchy theory, and complexity science), addressing this geographical and historical variation will require a research program that reaches across scales. Though a multitude of studies exist that address either, there is a scarcity of research aimed at bridging the gap between general and local patterns and causes of treeline (but see CitationHarsch et al., 2009; CitationHarsch and Bader, 2011). To build such a bridge, methodologies can be shared between the two approaches and specifically multiscale analyses can be adopted.

In more general terms, we propose that more formal hierarchical statistical methods (e.g., using Bayesian and/or multilevel statistics) be applied to link the two approaches that we have discerned (e.g., CitationBeever et al., 2006; CitationClark and Gelfand, 2006; CitationQian and Shen, 2007). Multilevel regression could link approaches because of their nested, geographically hierarchical relationship. Multilevel analyses will be most informative where the levels of the hierarchy cross functionally important differences in underlying and constraining variables. Some work on treeline illustrates this approach but more variables are needed (e.g., CitationHarsch et al., 2009; but contrast CitationGellrich et al., 2007).

Research likely to emerge includes reformulations of existing treeline models that incorporate higher spatial and temporal resolution data sets. Increasing the level of detail and geographic specificity within the present understanding of alpine treeline ecology is warranted because, generally speaking, global-scale controls on treelines are well understood, (that is, temperature is an important control on treelines worldwide, e.g., CitationKörner and Paulsen, 2004). In contrast, landscape-scale treeline analyses vary greatly in their foci and level of detail, which often makes them challenging to compare directly and to synthesize across large geographic areas. Particularly desirable treeline analyses include those that (1) use theoretically and methodologically consistent analytical approaches to better define geographic variability in treeline pattern-process relationships; (2) explicitly assess the role that local climatic conditions (e.g., the timing of events like first snowfalls, spring thaw dates, and late and early frosts) play in tree establishment and survival; and (3) combine sites with different histories of land use. Such analyses are needed to provide a basis for assessing localized ecotonal responses to geographically variable climate changes expected to occur in coming decades.

Acknowledgments

This paper was initiated at a workshop organized by the Mountain GeoDynamics Research Group that was supported by cooperative agreement 04CRAG0030 with the U.S. Geological Survey.

References Cited

  • Akhalkatsi, M. , O. Abdaladze , G. Nakhutsrishvili , and W. K. Smith . 2006. Facilitation of seedling microsites by Rhododendron caucasicum extends the Betula litwinowii alpine treeline, Caucasus Mountains, Republic of Georgia. Arctic, Antarctic, and Alpine Research 38:481–488.
  • Alftine, K. J. and G. P. Malanson . 2004. Directional positive feedback and pattern at an alpine tree line. Journal of Vegetation Science 15:3–12.
  • Alftine, K. J. , G. P. Malanson , and D. B. Fagre . 2003. Feedback-driven response to multi-decadal climatic variability at an alpine forest-tundra ecotone. Physical Geography 24:520–535.
  • Allen, T. F. H. and T. B. Starr . 1982. Hierarchy. Perspectives for Ecological Complexity. Chicago University of Chicago Press.
  • Allen, T. R. and S. J. Walsh . 1996. Spatial and compositional pattern of alpine treeline, Glacier National Park, Montana. Photogrammetric Engineering and Remote Sensing 62:1261–1268.
  • Anschlag, K. , G. Broll , and F-K. Holtmeier . 2008. Mountain birch seedlings in the treeline ecotone, subarctic Finland: variation in above- and below-ground growth depending on microtopography. Arctic, Antarctic, and Alpine Research 40:609–616.
  • Armand, A. D. 1992. Sharp and gradual mountain timberlines as a result of species interaction. In Hansen, A. J. and F. di Castri . (eds.). Landscape Boundaries: Consequences for Biotic Diversity and Ecological Flows. New York Springer. 360–378.
  • Bader, M. Y. 2007. Tropical Alpine Treelines: How Ecological Processes Control Vegetation Patterning and Dynamics Ph.D. thesis. Wageningen Universiteit.
  • Bader, M. Y. and J. J. A. Ruijten . 2008. A topography based model of forest cover at the alpine tree line in the tropical Andes. Journal of Biogeography 35:711–723.
  • Bader, M. Y. , I. Van Geloof , and M. Rietkerk . 2007a. High solar radiation hinders tree regeneration above the alpine treeline in northern Ecuador. Plant Ecology 191:33–45.
  • Bader, M. Y. , M. Rietkerk , and A. K. Bregt . 2007b. Vegetation structure and temperature regimes of tropical alpine treelines. Arctic, Antarctic, and Alpine Research 39:353–364.
  • Bader, M. Y. , M. Rietkerk , and A. K. Bregt . 2008. A simple spatial model exploring positive feedbacks at tropical alpine treelines. Arctic, Antarctic, and Alpine Research 40:269–278.
  • Bak, P. 1996. How Nature Works. New York Springer.
  • Baker, B. B. and R. K. Moseley . 2007. Advancing treeline and retreating glaciers: implications for conservation in Yunnan, P.R. China. Arctic, Antarctic, and Alpine Research 39:200–209.
  • Baker, W. L. and P. J. Weisberg . 1995. Landscape analysis of the forest-tundra ecotone in Rocky Mountain National Park, Colorado. Professional Geographer 47:361–375.
  • Ball, M. C. , V. S. Hodges , and G. P. Laughlin . 1991. Cold-induced photoinhibition limits regeneration of snow gum at tree-line. Functional Ecology 5:663–668.
  • Bansal, S. and M. J. Germino . 2008. Carbon balance of conifer seedlings at timberline: relative changes in uptake, storage, and utilization. Oecologia 158:217–227.
  • Batllori, E. , J. J. Camarero , J. M. Ninot , and E. Gutierrez . 2009. Seedling recruitment, survival and facilitation in alpine Pinus uncinata tree line ecotones. Implications and potential responses to climate warming. Global Ecology and Biogeography 18:460–472.
  • Batllori, E. , J. J. Camarero , and E. Gutierrez . 2010. Current regeneration patterns at the tree line in the Pyrenees indicate similar recruitment processes irrespective of the past disturbance regime. Journal of Biogeography 37:1938–1950.
  • Beever, E. A. , M. Huso , and D. A. Pyke . 2006. Multiscale responses of soil stability and invasive plants to removal of non-native grazers from an arid conservation reserve. Diversity and Distributions 12:258–268.
  • Bekker, M. F. 2005. Positive feedback between tree establishment and patterns of subalpine forest advancement, Glacier National Park, Montana, U.S.A. Arctic, Antarctic, and Alpine Research 37:97–107.
  • Bekker, M. F. and G. P. Malanson . 2009. Modeling feedback effects on linear patterns of subalpine forest advancement. In Butler, D. R. , G. P. Malanson , S. J. Walsh , and D. B. Fagre . (eds.). The Changing Alpine Treeline: the Example of Glacier National Park, Montana, USA. Amsterdam Elsevier. 167–190.
  • Blüthgen, J. 1942. Die polare Baumgrenze Veröffentlichungen des Deutschen Wissenschaftlichen Instituts zu Kopenhagen, Reihe 1, Arktis 10.
  • Brodersen, C. R. , M. J. Germino , and W. K. Smith . 2006. Photosynthesis during an episodic drought in Abies lasiocarpa and Picea engelmannii across an alpine treeline. Arctic, Antarctic, and Alpine Research 38:34–41.
  • Broll, G. , F-K. Holtmeier , K. Anschlag , H-J. Brauckmann , S. Wald , and B. Drees . 2007. Landscape mosaic in the treeline ecotone on Mt. Rodjanoaivi, subarctic Finland. Fennia 1985:89–105.
  • Brown, D. G. 1994. Predicting vegetation at treeline using topography and biophysical disturbance variables. Journal of Vegetation Science 5:641–656.
  • Brown, D. G. , D. M. Cairns , G. P. Malanson , S. J. Walsh , and D. R. Butler . 1994. Remote sensing and GIS techniques for spatial and biophysical analyses of alpine treeline through process and empirical models. In Michener, W. K. , S. Stafford , and J. Brunt . (eds.). Environmental Information Management and Analysis: Ecosystem to Global Scales. Philadelphia Taylor and Francis. 453–481.
  • Burga, C. A. 1988. Swiss vegetation history during the last 1800 years. New Phytologist 110:581–602.
  • Burga, C. A. and R. Perret . 2001. Monitoring of eastern and southern Swiss alpine timberline ecotones. In Burga, C. and A. Kratochwil . (eds.). Biomonitoring: General and Applied Aspects on Regional and Global Scale. Dordrecht Kluwer. 179–194.
  • Butler, D. R. 1992. The grizzly bear as an erosional agent in mountainous terrain. Zeitschrift für Geomorphologie 36:179–189.
  • Butler, D. R. 1995. Zoogeomorphology: Animals as Geomorphic Agents. Cambridge Cambridge University Press.
  • Butler, D. R. and W. D. Butler . 2009. The geomorphic effects of gophers on soil characteristics and sediment compaction: a case study from alpine treeline, Sangre de Cristo Mountains, Colorado, USA. The Open Geology Journal 3:82–89.
  • Butler, D. R. and S. J. Walsh . 1994. Site characteristics of debris flows and their relationship to alpine treeline. Physical Geography 15:181–199.
  • Butler, D. R. , G. P. Malanson , and D. M. Cairns . 1994. Stability of alpine treeline in northern Montana, USA. Phytocoenologia 22:485–500.
  • Butler, D. R. , G. P. Malanson , and L. M. Resler . 2004. Turf-banked terrace tread and risers, turf exfoliation, and possible relationships with advancing treeline. Catena 58:259–274.
  • Butler, D. R. , G. P. Malanson , and D. B. Fagre . 2007. Influences of geomorphology and geology on alpine treeline in the American West—More important than climatic influences? Physical Geography 28:434–450.
  • Butler, D. R. , G. P. Malanson , S. J. Walsh , and D. B. Fagre . 2009a. The Changing Alpine Treeline, the Example of Glacier National Park, MT, USA. Amsterdam Elsevier.
  • Butler, D. R. , G. P. Malanson , L. M. Resler , S. J. Walsh , F. D. Wilkerson , G. L. Schmid , and C. F. Sawyer . 2009b. Geomorphic patterns and processes at alpine treeline. In Butler, D. R. , G. P. Malanson , S. J. Walsh , and D. B. Fagre . (eds.). The Changing Alpine Treeline: the Example of Glacier National Park, Montana, USA. Amsterdam Elsevier. 63–84.
  • Cairns, D. M. 2005. Simulating carbon balance at treeline for krummholz and dwarf tree growth forms. Ecological Modelling 187:314–328.
  • Cairns, D. M. and G. P. Malanson . 1997. Examination of the carbon balance hypothesis of alpine treeline location, Glacier National Park, Montana. Physical Geography 18:125–145.
  • Cairns, D. M. and G. P. Malanson . 1998. Environmental variables influencing carbon balance at the alpine treeline ecotone: a modeling approach. Journal of Vegetation Science 9:679–692.
  • Cairns, D. M. and J. Moen . 2004. Herbivory influences tree lines. Journal of Ecology 92:1019–1024.
  • Cairns, D. M. , C. Lafon , J. Moen , and A. Young . 2007. Influences of animal activity on treeline position and pattern: implications for treeline responses to climate change. Physical Geography 28:419–433.
  • Camarero, J. J. and E. Gutierrez . 2002. Plant species distribution across two contrasting treeline ecotones in the Spanish Pyrenees. Plant Ecology 162:247–257.
  • Carcaillet, A. L. , B. Talon , and M. Barbero . 1998. Pinus cembra et incendies pendant l'Holocene, 300 m au-dessus de la limite actuelle des arbres dans le massif de la Vanoise (Alpes due nourd-ouest). Ecologie 29:227–282.
  • Clark, J. S. and A. E. Gelfand . 2006. A future for models and data in environmental science. Trends in Ecology & Evolution 21:375–380.
  • Colombaroli, D. , P. D. Henne , P. Kaltenrieder , E. Gobet , and W. Tinner . 2010. Species responses to fire, climate and human impact at tree line in the Alps as evidenced by palaeo-environmental records and a dynamic simulation model. Journal of Ecology 98:1346–1357.
  • Cunningham, C. , N. E. Zimmermann , V. Stoeckli , and H. Bugmann . 2006. Growth response of Norway spruce saplings in two forest gaps in the Swiss Alps to artificial browsing, infection with black snow mold, and competition with ground vegetation. Canadian Journal of Forest Research 36:2782–2793.
  • Daly, C. and D. Shankman . 1985. Seedling establishment by conifers above tree limit on Niwot Ridge, Front Range, Colorado, U.S.A. Arctic and Alpine Research 17:389–400.
  • Daniels, L. D. and T. T. Veblen . 2003. Regional and local effects of disturbance and climate on altitudinal treelines in northern Patagonia. Journal of Vegetation Science 14:733–742.
  • Däniker, A. 1923. Biologische Studien über Wald- und Baumgrenze, insbesondere über die klimatischen Ursachen und deren Zusammenhänge. Vierteljahresschrift der Naturforschenden Gesellschaft Zürich 63.
  • Deng, J-M. , G-X. Wang , E. C. Morris , X-P. Wei , D-X. Li , B-M. Chen , C-M. Zhao , J. Liu , and Y. Wang . 2006. Plant mass-density relationship along a moisture gradient in north-west China. Journal of Ecology 94:953–958.
  • Di Pasquale, G. , M. Marziano , S. Impagliazzo , C. Lubritto , A. De Natale , and M. Y. Bader . 2008. The Holocene treeline in the northern Andes (Ecuador): first evidence from soil charcoal. Palaeogeography Palaeoclimatology Palaeoecology 259:17–34.
  • Diaz, H. F. , M. Grosjean , and L. Graumlich . 2003. Climate variability and change in high-elevation regions: past, present and future. Climatic Change 59:1–4.
  • Diaz-Varela, R. A. , R. Colombo , M. Meroni , M. S. Calvo-Iglesias , A. Buffoni , and A. Tagliagerri . 2010. Spatio-temporal analysis of alpine ecotones: a spatial explicit model targeting altitudinal vegetation shifts. Ecological Modelling 221:621–633.
  • Didier, L. 2001. Invasion patterns of European larch and Swiss stone pine in subalpine pastures in the French Alps. Forest Ecology and Management 145:67–77.
  • Dullinger, S. , T. Dirnbock , and G. Grabherr . 2003. Patterns of shrub invasion into high mountain grasslands of the Northern Calcareous Alps, Austria. Arctic, Antarctic, and Alpine Research 35:434–441.
  • Elliott, G. P. 2009. Multi-scale Influences of Climate on Upper Treeline Dynamics along an Altitudinal Gradient in the Rocky Mountains, U.S.A Ph.D. thesis. University of Minnesota.
  • Elliott, G. P. and K. E. Kipfmueller . 2010. Multi-scale influences of slope aspect and spatial pattern on ecotonal dynamics at upper treeline in the southern Rocky Mountains, USA. Arctic, Antarctic, and Alpine Research 42:45–56.
  • Geddes, C. A. , D. G. Brown , and D. B. Fagre . 2005. Topography and vegetation as predictors of snow water equivalent across the alpine treeline ecotone at Lee Ridge, Glacier National Park, Montana, U.S.A. Arctic, Antarctic, and Alpine Research 37:197–205.
  • Gell-Mann, M. 1994. The Quark and the Jaguar. New York W. H. Freeman.
  • Gellrich, M. , P. Baur , B. Kock , and N. E. Zimmermann . 2007. Agricultural land abandonment and natural forest re-growth in the Swiss mountains: a spatially explicit economic analysis. Agriculture, Ecosystems and Environment 118:93–108.
  • Germino, M. J. and W. K. Smith . 1999. Sky exposure, crown architecture, and low-temperature photoinhibition in conifer seedlings at alpine treeline. Plant, Cell and Environment 22:407–415.
  • Germino, M. J. and W. K. Smith . 2000. Differences in microsite, plant form, and low-temperature photoinhibition of photosynthesis in alpine plants. Arctic, Antarctic, and Alpine Research 32:388–396.
  • Germino, M. J. , W. K. Smith , and A. C. Resor . 2002. Conifer seedling distribution and survival in an alpine-treeline ecotone. Plant Ecology 162:157–168.
  • Germino, M. J. , N. J. Hasselquist , T. McGonigle , W. K. Smith , and P. P. Sheridan . 2006. Landscape- and age-based factors affecting fungal colonization of conifer seedling roots at the alpine tree line. Canadian Journal of Forest Research 36:901–909.
  • Griggs, R. F. 1946. The timberlines of northern America and their interpretation. Ecology 27:275–289.
  • Haken, H. 1975. Cooperative phenomena in systems far from thermal equilibrium and in nonphysical systems. Reviews of Modern Physics 47:67–121.
  • Harsch, M. A. , P. E. Hulme , M. S. McGlone , and R. P. Duncan . 2009. Are treelines advancing? A global meta-analysis of treeline response to climate warming. Ecology Letters 12:1040–1049.
  • Harsch, M. A. and M. Y. Bader . 2011. Treeline form: a potential key to understanding treeline dynamics? Global Ecology and Biogeography 20. in press.
  • Hasselquist, N. , M. J. Germino , T. McGonigle , and W. K. Smith . 2005. Variability of Cenococcum colonization and its ecophysiological significance for young conifers at alpine-treeline. New Phytologist 165:867–873.
  • Heiri, C. , H. Bugmann , W. Tinner , O. Heiri , and H. Lischke . 2006. A model-based reconstruction of Holocene treeline dynamics in the Central Swiss Alps. Journal of Ecology 94:206–216.
  • Henning, I. 1974. Geookologie der Hawaii-Inseln. Erdwissenschaftliche Forschung IX. Wiesbaden Steiner Verlag.
  • Hiemstra, C. A. , G. E. Liston , and W. A. Reiners . 2002. Snow redistribution by wind and interactions with vegetation at upper treeline in the Medicine Bow Mountains, Wyoming, USA. Arctic, Antarctic, and Alpine Research 34:262–273.
  • Hiemstra, C. A. , G. E. Liston , and W. A. Reiners . 2006. Observing, modelling, and validating snow redistribution by wind in a Wyoming upper treeline landscape. Ecological Modelling 197:35–51.
  • Hoch, G. and C. Körner . 2003. The carbon charging of pines at the climatic treeline: a global comparison. Oecologia 135:10–21.
  • Hoch, G. and C. Körner . 2009. Growth and carbon relations of tree line forming conifers at constant vs. variable low temperatures. Journal of Ecology 97:57–66.
  • Hofstede, R. , P. Segarra , and P. Mena Vásconez . (eds.). 2003. Los páramos del Mundo. Proyecto Atlas Mundial de los Páramos. Quito Global Peatland Initiative/NC-IUCN/EcoCiencia.
  • Holtmeier, F-K. 1965. Die Waldgrenze im Oberengadin in ihrer physiognomischen und ökologischen Differenzierung Dissertation Mathematisch-Naturwissenschaftliche Fakultät. Rheinische Friedrich-Wilhelms-Universität Bonn.
  • Holtmeier, F-K. 1966. Die ökologische Funktion des Tannenhähers im Zirben-Lärchenwald und an der Waldgrenze im Oberengadin. Journal für Ornithologie 4:337–345.
  • Holtmeier, F-K. 1974. Geoökologische Beobachtungen und Studien an der subarktischen und alpine Waldgrenze in vergleichender Sicht (nördliches Fennoskandien/Zentralalpen. Erdwissenschaftliche Forschung 8. Wiesbaden Steiner Verlag.
  • Holtmeier, F-K. 1982. ‘Ribbon-forest’ und ‘Hecken’. Streifenartige Verbreitungsmuster des Baumwuchses an der oberen Waldgrenze in den Rocky Mountains. Erdkunde 36:142–153.
  • Holtmeier, F-K. 1985. Die klimatische Waldgrenze—Linie oder Übergangssaum (Ökoton)?—ein Diskussionsbeitrag unter besonderer Berücksichtigung der Waldgrenzen in den mittleren und hohen Breiten der Nordhalbkugel. Erdkunde 39:271–285.
  • Holtmeier, F-K. 1986. Die obere Waldgrenze unter dem Einfluß von Klima und Mensch. Abhandlungen aus dem Landesmuseum für Naturkunde 48:395–412.
  • Holtmeier, F-K. 1999. Ablegerbildung im Hochlagenwald und an der oberen Waldgrenze in der Front Range, Colorado. Mitteilungen der Deutschen Dendrologischen Gesellschaft 84:39–61.
  • Holtmeier, F-K. 2002. Tiere in der Landschaft, Einfluß und ökologische Bedeutung. Stuttgart Ulmer Verlag.
  • Holtmeier, F-K. 2005. Relocation of snow and its effects in the treeline ecotone—With special regard to the Rocky Mountains, the Alps and northern Europe. Die Erde 136:343–373.
  • Holtmeier, F-K. 2009. Mountain Timberlines—Ecology, Patchiness, and Dynamics. Advances in Global Change Research, 36. Dordrecht Springer Science + Business Media B. V.
  • Holtmeier, F-K. and G. Broll . 1992. The influence of tree islands and microtopography on pedoecological conditions in the forest alpine tundra ecotone on Niwot Ridge, Colorado Front Range, USA. Arctic and Alpine Research 24:216–228.
  • Holtmeier, F-K. and G. Broll . 2005. Sensitivity and response of northern hemisphere altitudinal and polar treelines to environmental changes at landscape and local scales. Global Ecology and Biogeography 14:395–410.
  • Holtmeier, F-K. and G. Broll . 2007. Treeline advance driving processes and adverse factors. Landscape Online 1:1–33. doi:10.3097/LO.200701.
  • Holtmeier, F-K. and G. Broll . 2010. Wind as an ecological agent at treelines in North America, in North America, the Alps, and the European Subarctic. Physical Geography 31:203–233.
  • Holtmeier, F-K. , G. Broll , A. Mütherthies , and K. Anschlag . 2003. Regeneration of trees in the treeline ecotone: northern Finnish Lapland. Fennia 181:103–128.
  • Horn, S. P. 1993. Postglacial vegetation and fire history in the Chirripo Paramo of Costa Rica. Quaternary Research 40:107–116.
  • Hörsch, B. 2003. Modelling the spatial distribution of montane and subalpine forests in the central alps using digital elevation models. Ecological Modelling 168:267–282.
  • Hughes, N. M. , D. M. Johnson , M. Akhalkatsi , and O. Abdaladze . 2009. Characterizing Betula litwinowii seeding microsites at the alpine-treeline ecotone, central Greater Caucasus Mountains, Georgia. Arctic, Antarctic, and Alpine Research 41:112–118.
  • Humphries, H. C. , P. S. Bourgeron , and L. R. Mujica-Crapanzano . 2008. Tree spatial patterns and enviornmental relationships in the forest-alpine tundra ecotone at Niwot Ridge, Colorado, USA. Ecological Research 23:589–605.
  • Ives, J. D. 1973. Studies in high altitude geoecology of the Colorado Front Range. A review of the research program of the Institute of Arctic and Alpine Research. Arctic and Alpine Research 5:65–66.
  • Ives, J. D. and K. J. Hansen-Bristow . 1983. Stability and instability of natural and modified upper timberline landscape in the Colorado Rocky Mountains, USA. Mountain Research and Development 3:1149–155.
  • Jackson, D. , C. Mendez , R. Seguel , A. Maldonado , and G. Vargas . 2007. Initial occupation of the Pacific coast of Chile during late Pleistocene times. Current Anthropology 48:725–731.
  • Johnson, D. M. and W. K. Smith . 2007. Limitations to photosynthetic carbon gain in timberline Abies lasiocarpa seedlings during prolonged drought. Canadian Journal of Forest Research 37:568–579.
  • Johnson, D. M. , M. J. Germino , and W. K. Smith . 2004. Abiotic factors limiting photosynthesis in Abies lasiocarpa and Picea engelmannii seedlings below and above the alpine timberline. Tree Physiology 24:377–386.
  • Kaltenrieder, P. , W. Tinner , and B. Ammann . 2005. Zur Langzeitokologie des Larchen-Arvengurtels inden sudlichen Walliser Alpen. Botanica Helvetica 115:137–154.
  • Kharuk, V. I. , K. J. Ransom , S. T. Im , and A. S. Vdovin . 2010. Spatial distribution and temporal dynamics of high-elevation forest stands of southern Siberia. Global Ecology and Biogeography 19:822–830.
  • Kihlman, O. 1890. Pflanzenbiologische Studien aus Russisch Lappland. Acta Societas Fauna et Flora Fennica VI (3):
  • Kipfmueller, K. F. and M. W. Salzer . 2010. Linear trend and climate response of five-needle pines in the western United States related to treeline proximity. Canadian Journal of Forest Resources 40:134–142.
  • Klasner, F. L. and D. B. Fagre . 2002. A half century of change in alpine treeline patterns at Glacier National Park, Montana, USA. Arctic, Antarctic, and Alpine Research 34:49–56.
  • Körner, C. 1998a. A re-assessment of high elevation treeline positions and their explanation. Oecologia 115:445–459.
  • Körner, C. 1998b. Worldwide positions of alpine treelines and their causes. In Beniston, M. and J. L. Innes . (eds.). The Impacts of Climatic Variability on Forests. Heidelberg Springer. 221–229.
  • Körner, C. 2003. Alpine Plant Life. Berlin Springer.
  • Körner, C. and J. Paulsen . 2004. A world-wide study of high altitude treeline temperatures. Journal of Biogeography 31:713–732.
  • Kullman, L. 1997. Tree-limit stress and disturbance. A 25-year survey of geoecological change in the Scandes Mountains of Sweden. Geografiska Annaler A 79:139–165.
  • Leuschner, C. 1996. Timberline and alpine vegetation on the tropical and warm-temperate oceanic islands of the world: elevation, structure and floristics. Vegetatio 123:193–206.
  • Leuschner, C. and M. Schulte . 1991. Microclimatological investigations in the tropical alpine scrub of Maui, Hawaii: evidence for a drought-induced alpine timberline. Pacific Science 45:152–168.
  • Li, B. L. 2002. A theoretical framework of ecological phase transitions for characterizing tree-grass dynamics. Acta Biotheoretica 50:141–154.
  • Li, M. H. and J. Yang . 2004. Effects of microsite on growth of Pinus cembra in the subalpine zone of the Austrian Alps. Annals of Forest Science 61:319–325.
  • Li, M. H. , G. Hoch , and C. Körner . 2002. Source/sink removal affects mobile carbohydrates in Pinus cembra at the Swiss treeline. Trees–Structure and Function 16:331–337.
  • Liptzin, D. and T. R. Seastedt . 2009. Patterns of snow, deposition, and soil nutrients at multiple spatial scales at a Rocky Mountain tree line ecotone. Journal of Geophysical Research–Biogeosciences 114:article G04002.
  • Lloyd, A. H. and L. J. Graumlich . 1997. Holocene dynamics of treeline forests in the Sierra Nevada. Ecology 78:1199–1210.
  • Loehle, C. , B. L. Li , and R. C. Sundell . 1996. Forest spread and phase transitions at forest-prairie ecotones in Kansas, USA. Landscape Ecology 11:225–235.
  • Maher, E. L. and M. J. Germino . 2006. Microsite differentiation among conifer species during seedling establishment at alpine treeline. Ecoscience 13:334–341.
  • Maher, E. L. , M. J. Germino , and N. J. Hasselquist . 2005. Interactive effects of tree and herb cover on survivorship, physiology, and microclimate of conifer seedlings at the alpine tree-line ecotone. Canadian Journal of Forest Research 35:567–574.
  • Malanson, G. P. 1997. Effects of feedbacks and seed rain on ecotone patterns. Landscape Ecology 12:27–38.
  • Malanson, G. P. 1999. Considering complexity. Annals, Association of American Geographers 89:746–753.
  • Malanson, G. P. and D. R. Butler . 1994. Tree–tundra competitive hierarchies, soil fertility gradients, and the elevation of treeline in Glacier National Park, Montana. Physical Geography 15:166–180.
  • Malanson, G. P. and Y. Zeng . 2004. Uncovering spatial feedbacks at alpine treeline using spatial metrics in evolutionary simulations. In Atkinson, P. M. , G. Foody , S. Darby , and F. Wu . (eds.). Geodynamics. Boca Raton CRC Press. 137–150.
  • Malanson, G. P. , N. Xiao , and K. J. Alftine . 2001. A simulation test of the resource-averaging hypothesis of ecotone formation. Journal of Vegetation Science 12:743–748.
  • Malanson, G. P. , D. R. Butler , D. B. Fagre , S. J. Walsh , D. F. Tomback , L. D. Daniels , L. M. Resler , W. K. Smith , D. J. Weiss , D. L. Peterson , A. G. Bunn , C. A. Hiemstra , D. Liptzin , P. S. Bourgeron , Z. Shen , and C. I. Millar . 2007. Alpine treeline of western North America: linking organism-to-landscape dynamics. Physical Geography 28:378–396.
  • Malanson, G. P. , D. G. Brown , D. R. Butler , D. M. Cairns , D. B. Fagre , and S. J. Walsh . 2009. Ecotone dynamics: invasibility of alpine tundra by tree species from the subalpine forest. In Butler, D. R. , G. P. Malanson , S. J. Walsh , and D. B. Fagre . (eds.). The Changing Alpine Treeline: the Example of Glacier National Park, Montana, USA. Amsterdam Elsevier. 35–61.
  • Marek, R. 1910. Waldgrenzstudien in den österreichischen Alpen. Petermanns Geographische Mitteilungen. Ergänzungsheft 168.
  • Marr, J. W. 1977. The development and movement of tree islands near the upper limit of tree growth in the southern Rocky Mountains. Ecology 58:1159–1164.
  • Mattes, H. 1978. Der Tannenhäher im Engadin. Studien zu seiner Ökologie und Funktion im Arvenwald. Münstersche Geographische Arbeiten 2.
  • Mattes, H. 1982. Die Lebensgemeinschaft von Tannenhäher und Arve. Eidgenössische Anstalt für das forstliche Versuchswesen, Berichte 241.
  • Mattes, H. 1985. The role of animals in cembran pine forest regeneration. In Turner, H. and W. Tranquillini . (eds.). Establishment and Tending of Subalpine Forests: Research and Management Proceedings of the 3rd IUFRO workshop P 1.07-00 1984. Eidgenössische Anstalt für das forstliche Versuchswesen, Berichte 270:195–205.
  • Michaelis, P. 1934. Ökologische Studien an der alpinen Baumgrenze V. Osmotischer Wert und Wassergehalt während des Winters in verschiedenen Höhenlagen. Jahrbuch für wissenschaftliche Botanik 80:337–362.
  • Miehe, G. and S. Miehe . 1994. Zur oberen Waldgrenze in tropischen Gebirge. Phytocoenologia 24:53–110.
  • Miehe, G. and S. Miehe . 2000. Comparative high mountain research on the treeline ecotone under human impact. Erdkunde 54:34–50.
  • Millar, C. I. , R. D. Westfall , and D. L. Delany . 2007. Response of high-elevation limber pine ( Pinus flexilis ) to multiyear droughts and 20th-century warming, Sierra Nevada, California, USA. Canadian Journal of Forest Research 37:2508–2520.
  • Milne, B. T. , A. R. Johnson , T. H. Keitt , C. A. Hatfield , J. David , and P. T. Hraber . 1996. Detection of critical densities associated with pinon-juniper woodland ecotones. Ecology 77:805–821.
  • Munier, A. , L. Hermanutz , J. D. Jacobs , and K. Lewis . 2010. The interacting effects of temperature, ground disturbance, and herbivory on seedling establishment: implications for treeline advance with climate warming. Plant Ecology 210:19–30.
  • Ohsawa, M. 1990. An interpretation of latitudinal patterns of forest limits in south and east Asian mountains. Journal of Ecology 78:326–339.
  • O'Neill, R. V. , D. L. DeAngelis , J. B. Waide , and G. E. Allen . 1986. A Hierarchical Concept of Ecosystems. Princeton Princeton University Press.
  • O'Neill, R. V. , A. R. Johnson , and A. W. King . 1989. A hierarchical framework for the analysis of scale. Landscape Ecology 3:193–205.
  • O'Neill, R. V. , R. H. Gardner , and M. G. Turner . 1992. A hierarchical neutral model for landscape analysis. Landscape Ecology 7:55–61.
  • Örlander, G. 1993. Shading reduces both visible and invisible frost damage to Norway spruce seedlings in the field. Forestry 66:27–36.
  • Pagels, H. R. 1989. The Dreams of Reason: the Computer and the Rise of the Sciences of Complexity. New York Bantam Books.
  • Paijmans, K. and E. Löffler . 1972. High altitude forests and grasslands on Mt. Edward, New Guinea. Journal of Tropical Geography 34:58–64.
  • Pattee, H. H. (ed.). 1973. Hierarchy Theory: the Challenge of Complex Systems. New York Braziller.
  • Piper, F. I. , L. A. Cavieres , M. Reyes-Díaz , and L. J. Corcuera . 2006. Carbon sink limitation and frost tolerance control performance of the tree Kageneckia aungustifolia D. Don (Rosaceae) at the treeline in central Chile. Plant Ecology 185:29–39.
  • Pisek, A. and E. Cartellieri . 1939. Zur Kenntnis des Wasserhaushaltes der Pflanzen, IV. Bäume und Sträucher. Jahrbuch für wissenschaftliche Botanik 88:22–68.
  • Qian, S. S. and Z. Shen . 2007. Ecological applications of multilevel analysis of variance. Ecology 88:2489–2495.
  • Rebertus, A. J. , T. Kitzberger , T. T. Veblen , and L. M. Roovers . 1997. Blowdown history and landscape patterns in the Andes of Tierra del Fuego, Argentina. Ecology 78:678–692.
  • Resler, L. M. and M. A. Fonstad . 2009. A Markov analysis of tree islands at alpine treeline. In Butler, D. R. , G. P. Malanson , S. J. Walsh , and D. B. Fagre . (eds.). 2009: The Changing Alpine Treeline: the Example of Glacier National Park, Montana, USA. Amsterdam Elsevier. 151–165.
  • Resler, L. M. and D. F. Tomback . 2008. Blister rust prevalence in krummholz whitebark pine: implications for treeline dynamics, northern Rocky Mountains, Montana, USA. Arctic, Antarctic, and Alpine Research 40:161–170.
  • Resler, L. M. , D. R. Butler , and G. P. Malanson . 2005. Topographic shelter and conifer establishment and mortality in an alpine environment, Glacier National Park, Montana. Physical Geography 26:112–125.
  • Rietkerk, M. , M. C. Boerlijst , F. van Langevelde , R. HilleRisLambers , J. van de Koppel , L. Kumar , H. H. T. Prins , and A. M. de Roos . 2002. Self-organization of vegetation in arid ecosystems. American Naturalist 160:524–530.
  • Rundel, P. W. , A. P. Smith , and F. C. Meinzer . (eds.). 1994. Tropical Alpine Environments: Plant Form and Function. Cambridge Cambridge University Press.
  • Sarmiento, F. O. 2000. Breaking mountain paradigms: ecological effects on human impacts in managed tropandean landscapes. Ambio 29:423–431.
  • Sarmiento, F. O. and L. M. Frolich . 2002. Andean cloud forest tree lines. Naturalness, agriculture and the human dimension. Mountain Research and Development 22:278–287.
  • Schickhoff, U. 1995. Himalayan forest-cover changes in historical perspective: a case study in the Kaghan valley, northern Pakistan. Mountain Research and Development 15:3–18.
  • Schickhoff, U. 2005. The upper timberline in the Himalayas, Hindu Kush and Karakorum: a review of geographical and ecological aspects. In Broll, G. and B. Keplin . (eds.). Mountain Ecosystems. Studies in Treeline Ecology. Berlin Springer. 275–354.
  • Schmidt, E. 1936. Baumgrenzstudien am Feldberg im Schwarzwald. Tharandter Forstliches Jahrbuch 87.
  • Schmidt-Vogt, D. 1990. Fire in high altitude forests of the Nepal Himalaya. In Goldhammer, J. G. and M. J. Jenkins . (eds.). Fire in Ecosystem Dynamics. The Hague SPB Academic. 191–199.
  • Seastedt, T. R. and G. A. Adams . 2001. Effects of mobile tree islands on alpine tundra soils. Ecology 82:8–17.
  • Shi, P. , C. Körner , and G. Hoch . 2008. A test of the growth-limitation theory for alpine tree line formation in evergreen and deciduous taxa of the eastern Himalayas. Functional Ecology 22:213–220.
  • Shiels, A. B. and R. L. Sanford . 2001. Soil nutrient differences between two krummholz-form tree species and adjacent alpine tundra. Geoderma 102:205–217.
  • Slot, M. , C. Wirth , J. Schumacher , G. M. Mohren , O. Shibistova , J. Lloyd , and J. Ensminger . 2005. Regeneration pattern in boreal Scots pine glades linked to cold-induced photoinhibition. Tree Physiology 25:1139–1150.
  • Smith, W. K. , M. J. Germino , T. E. Hancock , and D. M. Johnson . 2003. Another perspective on altitudinal limits of alpine timberlines. Tree Physiology 23:1101–1112.
  • Smith, W. K. , M. J. Germino , D. M. Johnson , and K. Reinhardt . 2009. The altitude of alpine treeline: a bellwether of climate change effects. Botanical Review 75:163–190.
  • Sole, R. V. and J. Bascompte . 2006. Self-Organization in Complex Ecosystems. Princeton Princeton University Press.
  • Steiner, M. 1935. Winterliches Bioklima und Wasserhaushalt an der alpinen Waldgrenze. Bioklimatische Beiblätter 2:57–65.
  • Stueve, K. M. , D. L. Cerney , R. M. Rochefort , and L. L. Kurth . 2009. Post-fire tree establishment patterns at the alpine treeline ecotone: Mount Rainier National Park, Washington, USA. Journal of Vegetation Science 20:107–120.
  • Tinner, W. , B. Ammann , and P. Germann . 1996. Treeline fluctuations recorded for 12,500 years by soil profiles, pollen, and plant macrofossils in the Central Swiss Alps. Arctic and Alpine Research 28:131–147.
  • Tomback, D. F. 1977. The Behavioral Ecology of Clark's Nutcracker (Nucifraga columbiana) in the Eastern Sierra Nevada Ph.D. thesis. University of California.
  • Tomback, D. F. and L. M. Resler . 2007. Invasive pathogens at alpine treeline: consequences for treeline dynamics. Physical Geography 28:397–418.
  • Tranquillini, W. 1979. Physiological Ecology of the Alpine Timberline: Tree Existance at High Altitudes with Special Reference to the European Alps. Berlin Springer-Verlag.
  • Troll, C. 1939. Luftbildplan und ökologische Bodenforschung. Berlin Zeitschrift der Gesellschaft für Erdkunde. 241–298.
  • Troll, C. 1962. Die dreidimensionale Landschaftsgliederung der Erde. Tübingen Hermann von Wissmann-Festschrift. 54–80.
  • Troll, C. 1971. Landscape ecology (geo-ecology) and biocoenology—A terminological study. Geoforum 8:43–46.
  • Troll, C. 1973. The upper timberline in different climatic zones. Arctic and Alpine Research 5:3–18.
  • Turner, M. G. 1989. Landscape ecology: the effect of pattern on process. Annual Review of Ecology and Systematics 20:171–197.
  • Vitousek, P. M. 1994. Beyond global warming: ecology and global change. Ecology 75:1861–1876.
  • von Humboldt, A. and A. Bonpland . 1805. Essai sur la Géographie des Plantes. Paris Chez Levrault, Schoell et Compagnie.
  • Walsh, S. J. , G. P. Malanson , and D. R. Butler . 1992. Alpine treeline in Glacier National Park, Montana. In Janelle, D. G. (ed.). Geographical Snapshots of North America. New York Guilford Press. 167–171.
  • Walsh, S. J. , D. R. Butler , T. R. Allen , and G. P. Malanson . 1994. Influence of snow patterns and snow avalanches on the alpine treeline ecotone. Journal of Vegetation Science 5:657–672.
  • Walsh, S. J. , D. R. Butler , and G. P. Malanson . 1997. An overview of scale, pattern, process relationships in geomorphology: a remote sensing and GIS perspective. Geomorphology 21:183–205.
  • Walsh, S. J. , D. R. Butler , G. P. Malanson , K. A. Crews-Meyer , J. P. Messina , and N. C. Xiao . 2003. Mapping, modeling, and visualization of the influences of geomorphic processes on the alpine treeline ecotone, Glacier National Park, MT, USA. Geomorphology 53:129–145.
  • Wardle, P. 1974. Alpine timberlines. In Ives, J. D. and R. G. Barry . (eds.). Arctic and Alpine Environments. London Methuen. 371–402.
  • Wardle, P. 1993. Causes of alpine timberline: a review of the hypotheses. In Alden, J. N. , S. Odum , and J. L. Mastrantonio . (eds.). Forest Development in Cold Climates. New York Springer. 89–103.
  • Wardle, P. , C. Ezcurra , C. Ramirez , and S. Wagstaff . 2001. Comparison of the flora and vegetation of the southern Andes and New Zealand. New Zealand Journal of Botany 39:69–108.
  • Weiss, D. J. 2009. Alpine Treeline Ecotones in the Western United States: a Multi-scale Comparative Analysis of Environmental Factors Influencing Pattern-Process Relations Ph.D. thesis. University of North Carolina, Chapel Hill.
  • West, G. B. , J. H. Brown , and B. J. Enquist . 1999. The fourth dimension of life: fractal geometry and allometric scaling of organisms. Science 284:1677–1679.
  • Wiegand, T. , J. J. Camarero , N. Rüger , and E. Gutiérrez . 2006. Abrupt population changes in treeline ecotones along smooth gradients. Journal of Ecology 94:880–892.
  • Wieser, G. and M. Tausz . 2007. Trees at their Upper Limit. Treelife Limitation at the Alpine Timberline. Dordrecht Springer.
  • Wilmking, M. , G. P. Juday , V. A. Barber , and H. S. J. Zald . 2004. Recent climate warming forces contrasting growth responses of white spruce at treeline in Alaska through temperature thresholds. Global Change Biology 10:1724–1736.
  • Wilson, J. B. and A. D. Q. Agnew . 1992. Positive-feedback switches in plant communities. Advances in Ecological Research 23:263–336.
  • Winkler, D. 1997. Waldvegetation in der Ostabdachung des Tibetanischen Hochlandes und die historiche und gegenwartige Entwaldung. Erdkunde 51:143–163.
  • Young, K. R. 1993. Tropical timberlines: changes in forest structure and regeneration between two Peruvian timberline margins. Arctic and Alpine Research 25:167–174.
  • Young, K. R. and B. León . 2007. Tree-line changes along the Andes: implications of spatial patterns and dynamics. Philosophical Transactions of the Royal Society B 362:263–272.
  • Zeng, Y. 2010. Modeling Complex Dynamics at Alpine Treeline Ecotones Ph.D. thesis. University of Iowa.
  • Zeng, Y. and G. P. Malanson . 2006. Endogenous fractal dynamics at alpine treeline ecotones. Geographical Analysis 38:271–287.
  • Zeng, Y. , G. P. Malanson , and D. R. Butler . 2007. Geomorphic limits to self organization in alpine forest-tundra ecotone vegetation. Geomorphology 91:378–392.

Reprints and Corporate Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

To request a reprint or corporate permissions for this article, please click on the relevant link below:

Academic Permissions

Please note: Selecting permissions does not provide access to the full text of the article, please see our help page How do I view content?

Obtain permissions instantly via Rightslink by clicking on the button below:

If you are unable to obtain permissions via Rightslink, please complete and submit this Permissions form. For more information, please visit our Permissions help page.