Open Access
How to translate text using browser tools
1 November 2008 Mountain Birch Seedlings in the Treeline Ecotone, Subarctic Finland: Variation in Above- and Below-Ground Growth Depending on Microtopography
Kerstin Anschlag, Gabriele Broll, Friedrich-Karl Holtmeier
Author Affiliations +
Abstract

The treeline ecotone in northern Finnish Lapland is characterized by a mosaic of sites with highly varying environmental conditions. Density, age structure, growth, and root systems of mountain birch seedlings (Betula pubescens ssp. czerepanovii [Orlova] Hämet-Ahti) were studied in different microsite classes (deflation, lichen heath, dwarf shrub heath, hummock, willow shrub, sedge mire). On wind-exposed convex topography characterized by shortage of moisture and nutrients, seedling establishment is impeded, as is indicated by low seedling densities and lack of very young seedlings as well as by high rooting depths and root/shoot ratios of the few individuals. In sedge mires, birch seedlings occasionally occur in great numbers but die off at an early stage. Extremely shallow root systems point to anoxia as the main reason. Severe shoot damage is common to almost all saplings. This is likely due to grazing by reindeer (Rangifer t. tarandus). As a result, height growth is suppressed even on sites with otherwise relatively suitable conditions (e.g. willow shrub sites). Future effects of climate change which could locally improve conditions for germination and establishment of birch young growth might be overridden by the effects of high reindeer density.

Introduction

In northern Finnish Lapland, treeline is formed by mountain birch (Betula pubescens ssp. czerepanovii [Orlova] Hämet-Ahti). The treeline is generally determined by thermal deficiency (e.g. Holtmeier, 1994; Wielgolaski, 1997; Körner, 1998). However, edaphic, topographic, anthropogenic, and biological factors have modifying effects (e.g. Oksanen et al., 1995; Sveinbjörnsson, 2000). Within the treeline ecotone, extending from the closed forest to the uppermost and usually crippled specimens of the tree species, site conditions vary strongly within a few meters distance (Holtmeier, 2003). The resulting microsite pattern (e.g. microclimate, soil conditions, duration of the winter snowpack) is a key factor for the regeneration of birch in the harsh treeline environment (Holtmeier, 1971; Kullman, 1984; Holtmeier et al., 2003; Holtmeier and Broll, 2005). Successful germination and establishment of tree seedlings rather than increasing growth of mature trees appear to be the most critical processes in treeline advance to higher altitude and more northern positions (Smith et al., 2003; Holtmeier and Broll, 2005, 2007; Karlsson et al., 2005). Thus, when assessing treeline response to climate change (cf. e.g. Seppä, 1996; Autio and Heikkinen, 2002; Kullman, 2002) tree regeneration and development of young growth in the treeline ecotone must particularly be considered.

Studies on mountain birch regeneration at Fennoscandian treelines are numerous (e.g. Sonesson and Hoogesteger, 1983; Holm, 1993; Kullman, 2002, 2003). Besides studies regarding stand structure and dynamics, investigations on root systems can substantially contribute to understanding the influence of the site pattern on birch performance (cf. e.g. Laitakari, 1935). However, although rooting zone conditions are particularly important for tree growth in the treeline ecotone, only a few studies concerning this matter exist for northern Fennoscandia (Sveinbjörnsson, 1987; cf. Sveinbjörnsson et al., 2002). Results from manipulative experiments with Betula pubescens ssp. czerepanovii seedlings (e.g. Karlsson and Nordell, 1996; Weih and Karlsson, 1997, 1999, 2002; Weih, 1998; Karlsson et al., 2000) are transferable to natural regeneration to a limited extent only. Thus, the objective of this study is to examine (1) how birch seedling occurrence is influenced by microtopography, (2) how birch root and shoot systems are influenced by environmental conditions depending on microtopography, and (3) how successful natural seed-based birch regeneration is in the study area.

Methods

Study Area

The field studies (2002/2003) were carried out in northern Utsjoki in Finnish Lapland (Fig. 1). The climate of the area is subcontinental (Oksanen and Virtanen, 1995). Mean annual air temperature at the nearest meteorological station (Kevo, 107 m a.s.l.) is −1.7 °C, mean monthly air temperature is −14.7 °C in January and 9.7 °C in June, and mean annual precipitation is 416 mm (data for 1962–2006). The acid bedrock of the study area (granulite and granodiorite) is covered by sandy till (2–3 m) (Geologinen Tutkimuslaitos, 1981).

Figure 1

Location of the study sites. 1  =  Rodjanoaivi, 2  =  Koahppeloaivi, 3  =  Staloskaidi.

i1523-0430-40-4-609-f01.gif

The study area is situated in the Paistunturin cooperative herding district and is intensively grazed year-round by semi-domesticated reindeer (Rangifer t. tarandus), with additional winter feeding. Reindeer population has rapidly increased since the 1970s and peaked in the 1980s and early 1990s (Helle and Kojola, 2006). Recent summer stocking densities in the study area are 3.5–4.5 reindeer km−2 (Solberg et al., 2005).

All plots are located in the upper treeline ecotone near the uppermost mountain birch trees (≥2 m) between 260 and 380 m a.s.l. on the slopes of Rodjanoaivi (509 m a.s.l.), Koahppeloaivi (419 m a.s.l.), and Staloskaidi (340 m a.s.l.) (Fig. 1).

Microsite Classification

Microtopography varies locally (basal till dissected by brooks; esker), and it strongly influences microclimate (wind exposure, snow distribution, soil moisture, soil temperature). As a result, very dry and wind-eroded ridges with hardly any vegetation alternate with more sheltered lichen heath and dwarf shrub heath areas. In valleys, where high water tables cause peat formation and waterlogging, willow coppice and sedge mires can be found, and hummock fields have been formed by cryoturbation (for a comprehensive description, cf. e.g. Haapasaari, 1988; Broll et al., 2007). Microsites were classified (Fig. 2) using local topography and vegetation as indicators (Holtmeier et al., 2003). Microsite classes are (1) deflation, (2) lichen heath, (3) dwarf shrub heath, (4) hummock, (5) willow shrub, and (6) sedge mire.

Figure 2

Topography-dependent microsite pattern (schematic).

i1523-0430-40-4-609-f02.gif

Across the three studied mountains, six investigation areas with highly variable microtopography were delimited (size of the areas: 1–4 ha; distance between areas: several hundred meters to several kilometers). Within an area, several individual plots per microsite class were examined. Individual and aggregated plot size for each microsite class varied between areas, depending on existence and spatial extent of the respective class. Emphasis was placed on heath sites which cover vast fell areas. Thus, deflation sites, lichen heath, and dwarf shrub heath were studied in all six areas. Hummock, willow shrub, and sedge mire sites were studied in three areas. Total square meters of studied plots varies between microsite classes because birch seedling availability is of different orders of magnitude (approximate sums: deflation 4500 m2; lichen heath 1200 m2; dwarf shrub heath 1200 m2; hummock 270 m2; willow shrub 120 m2; and sedge mire 120 m2).

Total vegetation cover and height (Table 1) were recorded in 20 representative plots per microsite class. Plot size was 4 m2 (lichen heath, dwarf shrub heath, hummock) or 9 m2 (deflation, willow shrub, sedge mire). Vegetation cover is very sparse at deflation sites and increases across heath sites to hummocks, willow shrub, and sedge mires. Vegetation height is about 20–40 cm on average, with lower vegetation at deflation sites and higher vegetation at willow shrub sites.

Table 1

Mean cover and maximum height of the vegetation (±SE in parentheses) in the different microsite classes (n  =  20 per microsite class).

i1523-0430-40-4-609-t01.gif

Soil profiles were described for all microsite classes following the German manual for soil mapping (AG Boden, 1994). Soil organic carbon and total nitrogen were measured by an elemental analyzer (EuroVector EA 3000). Soil conditions vary considerably between microsite classes (for representative soil profiles cf. Figs. 3a and 3b). Generally, topsoils at deflation sites were eroded and former Podzol subsoils became exposed. Consequently, organic carbon is frequently and total nitrogen is always below the detection limit at deflation sites (cf. Fig. 3a). At lichen heath and dwarf shrub heath sites, shallow Podzols occur, with soils under dwarf shrub heath being more developed. At these heath sites, detectable nitrogen is limited to the organic topsoils. Hummock sites are characterized by Podzols influenced by ground water, a relic peat cover, horizons partly mixed by cryoturbation, and normally wide C/N ratios (cf. Fig. 3b). Histosols of willow shrub sites and sedge mires show comparatively narrow C/N ratios. Usually, the ground water table is higher in sedge mires than in willow shrub sites.

Figure 3a

Organic carbon and total nitrogen in representative soil profiles (b.d.l.: below detection limit, n.d.: not determined) A: deflation; B: lichen heath; C: dwarf shrub heath. No nitrogen could be detected in the deflation site profile. Nomenclature of soil horizons after the German soil classification (AG Boden, 1994).

i1523-0430-40-4-609-f301.gif

Figure 3b

Organic carbon and total nitrogen in representative soil profiles (b.d.l.: below detection limit, n.d.: not determined). A: hummock; B: willow shrub; C: sedge mire. Nomenclature of soil horizons after the German soil classification (AG Boden, 1994).

i1523-0430-40-4-609-f302.gif

Soil moisture of the topmost 6 cm was measured by TDR on five (deflation, lichen heath, dwarf shrub heath) or three (hummock, willow shrub, sedge mire) representative plots once a week (10 random measurements per date and plot; 24 June–28 August 2003). Soil moisture (Table 2) differs significantly with microsite class [H (5)  =  1742, p < 0.001]. It is lowest at excessively drained deflation sites and increases across lichen and dwarf shrub heath and hummocks to poorly drained willow shrub sites and sedge mires.

Table 2

Soil moisture (topmost 6 cm) at the different microsite classes (total number of measurements in the respective class: 1n  =  500, 2n  =  300) (b.d.l.  =  below detection limit).

i1523-0430-40-4-609-t02.gif

Soil temperatures were recorded (Hotdog DT1, Elpro; September 2002–August 2003) in 2.5 cm depth in one representative deflation site, lichen heath site, and dwarf shrub heath site, respectively. Additionally, temperature was recorded on a brookside close to a willow shrub site and a sedge mire site; however, transferability to those site classes is restricted (ground water table, vegetation structure). Duration of snow cover was estimated from soil temperature data. Soil temperatures (Table 3) are most extreme at the studied deflation site (annual and daily amplitudes). There is little snow cover, and freeze-thaw cycles are frequent. Conditions are less extreme underneath lichen heath. Dwarf shrub heath is characterized by long-lasting snow and relatively few freeze-thaw cycles, but the soil is generally cold during the growing season. The brookside soil is generally rather warm, in spite of a long lasting winter snow cover. Freeze-thaw cycles underneath the snow cover are frequent.

Table 3

Soil temperatures (2.5 cm depth) at representative plots (brookside data are transferable to willow shrub and sedge mire to a limited extent).

i1523-0430-40-4-609-t03.gif

Birch Sampling

In the present paper, the term “seedlings” refers to birches up to 50 cm tall grown from seeds. Taller young growth, sprouts from old root stocks, and hybrids (Betula pubescens ssp. czerepanovii × Betula nana) were excluded from the study. Hybridization becomes apparent from leaf morphology (e.g. Jonsell, 2000). Shoot length was mapped in the field for all seedlings in the study plots (total n  =  736). Seedling density was calculated for the combined plots of each microsite class.

A subsample of seedlings was randomly chosen (5 seedlings per area for deflation, hummock, willow shrub, and sedge mire; 12 seedlings per area for lichen heath and dwarf shrub heath; plus a total of 6 seedlings from preliminary tests). All visible shoot lesions per seedling were counted and later related to the seedling's age (to account for differences of seedling age distribution between microsite classes). Plants were excavated (total n  =  225), and depth of the rooting systems was determined. For calculation of root/shoot ratios, leafless seedlings were dried (75 °C, 48 h) and separated into above- and below-ground compartments (in case of adventitious root formation at the level of the uppermost root). Age of excavated plants was determined from stem discs under the stereomicroscope after staining with safranine solution and chalk (Iseli and Schweingruber, 1989). Age values are approximate due to narrow and obscure rings, frequent fungal damage, and possibly false and missing rings. Thus, correlation with climatic data is not possible.

Statistical Analysis

Data were compared between microsite classes with one-way ANOVA and Student-Newman-Keuls test or with Kruskal-Wallis test and Nemenyi test for unequal sample sizes when parametric assumptions were not met (Lozán and Kausch, 1998). The significance level is 0.05. Since the root/shoot ratio does not differ significantly with plant age, the complete data set is used for analysis. As seedling age influences maximum rooting depth, only data from age class 5–12 years are presented here.

Results

Mean densities of birch young growth (Fig. 4) are lowest at deflation sites where seedlings grow mainly close to the surrounding heath. On lichen heath sites, young growth often occurs at disturbed patches with exposed mineral soil, which are common because of steep slopes. Quite low densities are recorded in dwarf shrub heath where seedlings occasionally grow clustered in places with comparatively sparse vegetation. In hummock fields, birches usually grow at the sides of the hummocks, while they do not occur on dry hummock tops or in the wet troughs between the hummocks. Willow shrub sites are characterized by fairly high numbers of birch seedlings and also constant occurrence over the various field plots. Very high densities can be found in sedge mires.

Figure 4

Mean seedling density in the different microsite classes across all study plots. DE  =  deflation, LI  =  lichen heath, DS  =  dwarf shrub heath, HU  =  hummock, WI  =  willow shrub, SE  =  sedge mire.

i1523-0430-40-4-609-f04.gif

Median age of birch seedlings at deflation sites (17 yrs.), lichen heath (15 yrs.), and dwarf shrub heath (15 yrs.) is significantly greater than at willow shrub sites (7 yrs.) and in sedge mires (9 yrs.), with intermediate values at the hummock sites (13 yrs.) [H (5)  =  40.05; p < 0.001]. Seedling establishment after the onset of the 1990s is especially low at the deflation sites; by contrast, none of the examined specimens from the willow shrub and sedge mire sites became established earlier than approximately 1990 (Fig. 5). At the willow shrub sites, however, birches taller than 0.5 m and presumably accordingly older than the examined seedlings could occasionally be found.

Figure 5

Years of birch establishment in dependence of microsite class.

i1523-0430-40-4-609-f05.gif

At deflation sites, root/shoot ratio of birch seedlings (Fig. 6A) is significantly higher than at all other sites (median: 2.90 g g−1). Relative investment into root systems is also high at lichen heath sites (median: 1.66 g g−1). There are no significant differences between the root/shoot ratios of the remaining site classes (medians from 0.88 to 1.28 g g−1), but birch seedlings at willow shrub sites show the tendency towards a very low relative root mass.

Figure 6

Root/shoot ratio (R/S) (A); maximum rooting depth in age class 5–12 years (B); shoot length (C); and number of shoot lesions in relation to plant age (D) in dependence of microsite class. Circles  =  outliers, asterisks  =  extreme cases. DE  =  deflation, LI  =  lichen heath, DS  =  dwarf shrub heath, HU  =  hummock, WI  =  willow shrub, SE  =  sedge mire.

i1523-0430-40-4-609-f06.gif

Maximum rooting depth of birch seedlings (5–12 yrs.; Fig. 6B) is on average greatest at the deflation sites, but significantly differs only from the dwarf shrub sites and sedge mires. No shallow root systems occur at the deflation sites and the willow shrub sites. By contrast, root systems in sedge mires are uniformly shallow (with frequent formation of adventitious roots). At lichen heath and especially at dwarf shrub heath sites, both shallow root systems (restricted to the topsoil rich in organic carbon) and deep reaching root systems occur. Older seedlings (data not shown) reach rooting depths of up to 50 cm at deflation sites, 43 cm in lichen heath, and 35 cm in dwarf shrub heath. Seedlings at deflation sites and also at lichen heath sites often show intense root ramification underneath boulders or along channels of old, decaying roots.

Shoot length of the birch seedlings is on average 6 cm (Fig. 6C). Only at the willow shrub sites, birch seedlings show a higher average shoot length. Some young growth larger than 0.5 m growing at the willow shrub sites was not included in this study. However, none of those birches grew higher than the canopy of the surrounding willows that obviously had been browsed.

The overall number of visible shoot lesions relative to plant age (Fig. 6D) is highest at the deflation sites, though not differing very much from numbers in lichen heath and dwarf shrub heath. Damage is low in sedge mires and also in hummock fields. At deflation sites, birches often show partly exposed roots systems with up to 30 cm of the main root laying bare. Depending on grain size at the individual deflation patches, 15% (sandy-skeletal substrate) to more than 50% (sandy substrate) of birch seedlings show this particular trait.

Discussion

Germination of birch is extremely poor at deflation sites under current conditions, as may be concluded from seedling density and age structure. One reason might be the lack of birch seeds. As these are dispersed by the wind, they may accumulate in more sheltered places (Sveinbjörnsson et al., 2002; Seppälä, 2004). Also, dryness of deflation sites enforced by the highly permeable substrate as well as frost damage and frost heave because of sparse snow cover and frequent freeze-thaw cycles are especially detrimental to very young seedlings with weakly developed root systems (Holtmeier et al., 2003; Forbes et al., 2005; Holtmeier and Broll, 2005; Karlsson et al., 2005).

Soil profiles clearly show that the deflation sites were formerly covered by Podzols (Holtmeier et al., 2003). The frequent occurrence of birch young growth near the edges of deflation patches as well as partly exposed roots suggest that many of the seedlings may have become established before the deflation areas reached their current extent. Hardly any birch seedlings date from after the beginning of the 1990s, which is also the time when reindeer population peaked in the study area (Helle and Kojola, 2006). The rapid increase since the mid 1970s was due to the use of off-road vehicles for herding, additional winter feeding, veterinarian care, decrease of the number of predators, and restricted slaughtering because of the Chernobyl nuclear disaster (Kashulina et al., 1997; Colpaert et al., 2003). It is possible that the deflation areas reached their present extent at the time of the population peak, since high reindeer numbers are often considered to be responsible for soil erosion on exposed topography (e.g. Kashulina et al., 1997; Broll, 2000; Löffler, 2000; Uhlig and Zink, 2006). On the other hand, a dense plant cover also hampers germination (e.g. Forbes et al., 2005). This can be seen from the occasional clustering of seedlings in small patches with sparser vegetation in heath sites. In this respect, reindeer grazing creating gaps in the field layer can be beneficial for germination, because competition is reduced, soil temperature increases, and N mineralization is enhanced (Broll, 2000; Cairns and Moen, 2004; Karlsson et al., 2005).

Occasionally extremely high birch germination in sedge mires is not followed by successful long-term establishment, although damage to shoot systems is negligible. The reason for this could be high water tables and thus a lack of oxygen in the soil, as indicated by shallow root systems of seedlings and frequently high soil moisture recordings (cf. Fitter, 1991; Jackson and Ricard, 2003).

The examined root systems show a strong response to microtopographic gradients in water and nutrient availability. Because of the excessively drained sandy substrate and the lack of organic topsoil at deflation sites, water and nutrient supply are very poor. This also holds true for plant available phosphorus, potassium, and cation exchange capacity (Holtmeier et al., 2003). Drought stress and nutrient shortage are commonly reported to cause deep reaching root systems and high root/shoot ratios (Cairns et al., 1997; Davies and Bacon, 2003; Ovaska et al., 2005), just as they are displayed by birch seedlings at the deflation sites and, to a lesser extent, at the lichen heath sites. Also, intense root branching that frequently occurs along former root channels and underneath boulders can be explained by enhanced nutrient supply and water percolation in such places (cf. Broll, 1994). The fact that the root systems of some seedlings at the lichen heath and dwarf shrub heath sites grow exclusively in organic topsoils clearly reflects the importance of humus content for water and nutrient availability at these sites.

Reindeer grazing may enhance or slow down nutrient cycling in tundra, depending on timing and intensity of the grazing (Olofsson et al., 2004). For soils comparable to the heath site soils in this study (poor and dry sites), a reduction of carbon and nitrogen content has been shown (Broll et al., 1999) which further diminishes nutrient and water supply to seedlings. This is also how herbivory can indirectly influence root/shoot ratios in the long term. By contrast, direct grazing impacts on the root/shoot ratio soon get counterbalanced by tissue growth in order to readjust to resource availability (e.g. water, nutrients, light) (Farrar and Jones, 2003).

Seedlings at sites with good resource availability usually show small investments into the root systems (Karlsson et al., 2000; Hutchings and John, 2003). For example, lowest root/shoot ratios occur at the willow shrub sites, which generally are comparatively productive sites (Hobbie and Chapin, 1998). This can be attributed to lateral nutrient input with (melt-)water and to relatively high soil temperatures under a thick snow cover and thus ongoing N mineralization during winter (Karlsson et al., 2005).

In contrast to the root system characteristics, average shoot size does not correspond to any gradient in microclimate or resource availability. The only microsite class displaying relatively larger birch seedlings (willow shrub) mainly differs from all the other microsite classes regarding vegetation structure (comparatively dense and high). A possible explanation for the enhanced shoot size might be more effective shelter from herbivory (cf. Cairns and Moen, 2004). A connection between intensive reindeer grazing and trampling on the one hand and low birch seedling densities and growth rates on the other hand has been documented by several studies (e.g. Eilertsen et al., 2002; den Herder and Niemelä, 2003; Lempa et al., 2005). A comparable shelter effect of willow coppice on birch seedlings at the treeline is reported from Norwegian sheep pastures (Löffler et al., 2004). Another reason for larger shoot systems at the willow shrub sites might be stronger competition for light, which usually triggers higher shoot growth (Farrar and Jones, 2003).

The slight decrease of shoot lesions in relation to plant age from deflation sites across lichen heath to dwarf shrub heath might be at least partly attributed to the microclimatic gradient (e.g. less abrasion by sand and ice particles, Holtmeier, 2003, 2005; Holtmeier et al., 2004). However, microclimatic differences cannot explain the large discrepancy in the amount of damage between those sites that are all similarly sheltered by a thick snow cover in winter (dwarf shrub heath, hummock fields, willow shrub sites, and sedge mires). Accessibility to herbivores is likely to be at least one reason (cf. Cairns and Moen, 2004). During the summer, reindeer usually avoid areas in wind-sheltered depressions because of molestation by insects (Helle, 2001). On the other hand, willow shrub foliage is an attractive food source for reindeer in low-productivity dwarf shrub heath (den Herder et al., 2004). In willow shrub sites, birch seedlings are therefore more exposed to browsing than in sedge mires or hummock fields, while they are still more protected from browsing than in dwarf shrub heath, for example, because of the denser vegetation structure. Also, other herbivores such as snow hare (Lepus timidus) and willow grouse (Lagopus lagopus) might affect birch seedlings.

Climate warming is generally supposed to be very pronounced in Northern Europe (IPCC, 2007). In principle, this is expected to cause a treeline rise (e.g. Callaghan et al., 2002; Holtmeier, 2003; Holtmeier and Broll, 2005). However, the influence of local site conditions will have a modifying effect on establishment and survival of birch seedlings as it is also evident in the present study. For Northern Europe, an increase of both precipitation and evapotranspiration is expected in the future, and snow depth and duration of the snow season might be reduced due to higher temperatures (IPCC, 2007). The resulting changes in soil moisture and snow distribution will determine the suitability of microsites for treeline advance. For example, in case of drier conditions, birch treeline advance is most likely in concave topography providing sufficient soil moisture and winter snowpack (cf. Kullman, 2004, 2005; Holtmeier and Broll, 2005). However, in case of higher soil moisture, conditions for birch regeneration might deteriorate at these sites due to waterlogging. At the same time, moister conditions (e.g. because of more frequent rainfall events) might improve regeneration on permeable substrates covered by only thin organic topsoils, i.e. at sites suboptimally supplied with water for seedling establishment at present (cf. Holtmeier and Broll, 2005). Nevertheless, increased precipitation cannot compensate for the loss of nutrients and water storage capacity caused by removal of the topsoil on convex topography. Thus, birch will invade such sites with great delay.

Another important factor regarding treeline development is a possible future increase in mass outbreaks of the autumnal moth Epirrita autumnata. In warmer winters, more eggs will survive (Virtanen et al., 1998; Neuvonen et al., 1999). Additionally, the winter moth Operophtera brumata has recently defoliated birch forests in northern Utsjoki for the first time. Mass outbreaks destroying forest stands in junction with high reindeer grazing pressure preventing birch regeneration have caused a treeline decline before (Kallio and Lehtonen, 1975; Lehtonen and Heikkinen, 1995). This is witnessed in the actual study area by numerous old remains of birch root stocks above the present treeline (Holtmeier and Broll, 2006). The same could easily happen again in the future. So, even though climatic change might improve the conditions for regeneration at some of the microsites in the treeline ecotone, successful seed-based regeneration of birch in this study area is likely to depend on future reindeer pasture management.

Conclusions

Locally varying environmental conditions strongly affect below-ground characteristics of birch young growth. Root systems clearly reflect differences in soil moisture and nutrient supply along a microtopographical gradient. Shoot system characteristics, however, though partly showing an influence of microclimate, seem to be strongly affected by herbivory. Consequently, a positive effect of rising temperatures on the future treeline position in the study area might be overridden by the impact of intensive reindeer grazing. This has to be taken into account if the treeline is considered as an indicator for climate change.

Acknowledgments

For help in the field and constructive discussions, we thank Hans-Jörg Brauckmann, Birgit Drees, and Sabine Wald. Many thanks go to the staff of the Kevo Subarctic Research Station, especially its former leader Seppo Neuvonen and Saini Heino. We also thank Lillian Harris, George Malanson, and two anonymous reviewers for helpful comments on the manuscript. We gratefully acknowledge financial support by the EU (LAPBIAT) and by the Deutsche Forschungsgemeinschaft (DFG).

References Cited

1.

AG Boden 1994. Bodenkundliche Kartieranleitung Hannover Schweizerbart'sche Verlagsbuchhandlung. pp. Google Scholar

2.

J. Autio and O. Heikkinen . 2002. The climate of northern Finland. Fennia 180:61–66. Google Scholar

3.

G. Broll 1994. Influence of the soil mosaic on biodiversity at heath sites in the European subarctic. In International Society of Soil Science , editor. 15th World Congress of Soil Science, Acapulco, Mexico, July 1994. Transactions Vol. 4a: Commission III: Symposia 220–231. Google Scholar

4.

G. Broll 2000. Influence of overgrazing by reindeer on soil organic matter and soil microclimate of well-drained soils in the Finnish subarctic. In R. Lal , editor. Global climate change and cold regions ecosystems. Advances in soil science Boca Raton CRC/Lewis Publishers. 163–172. Google Scholar

5.

G. Broll, R. Ohtonen, and S. Stark . 1999. Effects of reindeer grazing on soil properties in Finnish Lapland. In M. Turunen, J. Hukkinen, O. W. Heal, N. R. Saelthun, and J. I. Holten , editors. A terrestrial transect for Scandinavia/northern Europe: proceedings of the international SCANTRAN Conference, 19.–23.3.1998, Rovaniemi, Finland. Ecosystems Research Report 31 Brussels European Commission. 56. Google Scholar

6.

G. Broll, 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 185:89–105. Google Scholar

7.

D. M. Cairns and J. Moen . 2004. Herbivory influences tree lines. Journal of Ecology 92:1019–1024. Google Scholar

8.

M. A. Cairns, S. Brown, E. H. Helmer, and G. A. Baumgardner . 1997. Root biomass allocation in the world's upland forests. Oecologia 111:1–11. Google Scholar

9.

T. V. Callaghan, B. R. Werkman, and R. M. M. Crawford . 2002. The tundra-taiga interface and its dynamics: concepts and applications. Ambio Special Report 12:6–14. Google Scholar

10.

A. Colpaert, J. Kumpula, and M. Nieminen . 2003. Reindeer pasture biomass assessment using satellite remote sensing. Arctic 56:147–158. Google Scholar

11.

W. J. Davies and M. A. Bacon . 2003. Adaptation of roots to drought. In H. de Kroon and J. W. Visser , editors. Root ecology Berlin Springer. 173–192. Google Scholar

12.

M. den Herder and P. Niemelä . 2003. Effects of reindeer on the re-establishment of Betula pubescens subsp. czerepanovii and Salix phylicifolia in a subarctic meadow. Rangifer 23:3–12. Google Scholar

13.

M. den Herder, R. Virtanen, and H. Roininen . 2004. Effects of reindeer browsing on tundra willow and its associated insect herbivores. Journal of Applied Ecology 41:870–879. Google Scholar

14.

S. M. Eilertsen, U. Schjelderup, and S. D. Mathiesen . 2002. Early season grazing effects on birch, grass, herbs and plant litter in coastal meadows used by reindeer: a short-term case study. Rangifer 22:123–131. Google Scholar

15.

J. F. Farrar and D. L. Jones . 2003. The control of carbon acquisition by and growth of roots. In H. de Kroon and J. W. Visser , editors. Root ecology Berlin Springer. 91–124. Google Scholar

16.

A. H. Fitter 1991. The ecological significance of root system architecture: an economic approach. In D. Atkinson , editor. Plant root growth: an ecological perspective London Blackwell. 229–243. Google Scholar

17.

B. C. Forbes, A. Tolvanen, F. E. Wielgolaski, and K. Laine . 2005. Rates and processes of natural regeneration in disturbed habitats. In F. E. Wielgolaski , editor. Plant ecology, herbivory, and human impact in Nordic mountain birch forests Berlin Springer. 193–202. Google Scholar

18.

Geologinen Tutkimuslaitos 1981. General geological map of Finland. Quaternary deposits Helsinki. Geologinen Tutkimuslaitos (Geological Survey of Finland), 1:400.000, No. 39 + 49 Utsjoki. Google Scholar

19.

M. Haapasaari 1988. The oligotrophic heath vegetation of northern Fennoscandia and its zonation. Acta Botanica Fennica 135:1–219. Google Scholar

20.

T. Helle 2001. Mountain birch forests and reindeer husbandry. In F. E. Wielgolaski , editor. Nordic mountain birch ecosystems New York UNESCO and The Parthenon Publishing Group. 279–291. Google Scholar

21.

T. Helle and I. Kojola . 2006. Population trends of semi-domesticated reindeer in Fennoscandia—Evaluation of explanations. In B. C. Forbes, M. Bölter, L. Müller-Wille, J. Hukkinen, F. Müller, N. Gunslay, and Y. Konstantinov , editors. Reindeer management in northernmost Europe: linking practical and scientific knowledge in social-ecological systems Berlin Springer. 319–339. Google Scholar

22.

S. E. Hobbie and F. S. Chapin III . 1998. An experimental test of limits to tree establishment in arctic tundra. Journal of Ecology 86:449–461. Google Scholar

23.

S. O. Holm 1993. Regeneration of Betula pendula and B. pubescens coll. above and below the natural altitudinal distribution limit of B. pendula in south-east Norway. Oecologia Montana 2:1–6. Google Scholar

24.

F-K. Holtmeier 1971. Waldgrenzstudien im nördlichen Finnisch-Lappland und angrenzenden Nordnorwegen. Reports from the Kevo Subarctic Research Station 8:53–62. Google Scholar

25.

F-K. Holtmeier 1994. Ecological aspects of climatically-caused timberline fluctuations. Review and outlook. In M. Beniston , editor. Mountain environments in changing climates London Routledge. 220–233. Google Scholar

26.

F-K. Holtmeier 2003. Mountain timberlines. Ecology, patchiness, and dynamics Dordrecht Kluwer. pp. Google Scholar

27.

F-K. Holtmeier 2005. Change in the timberline ecotone in northern Finnish Lapland during the last thirty years. Reports from the Kevo Subarctic Research Station 23:97–113. Google Scholar

28.

F-K. Holtmeier and G. Broll . 2005. Sensitivity and response of northern hemisphere altitudinal and polar treelines to environmental change at landscape and local scales. Global Ecology and Biogeography 14:395–410. Google Scholar

29.

F-K. Holtmeier and G. Broll . 2006. Radiocarbon-dated peat and wood remains from the Finnish subarctic: evidence of treeline and landscape history. The Holocene 16:743–751. Google Scholar

30.

F-K. Holtmeier and G. Broll . 2007. Treeline advance—Driving processes and adverse factors. Landscape Online 1:1–33. Google Scholar

31.

F-K. Holtmeier, G. Broll, A. Müterthies, and K. Anschlag . 2003. Regeneration of trees in the treeline ecotone: northern Finnish Lapland. Fennia 181:103–128. Google Scholar

32.

F-K. Holtmeier, G. Broll, and K. Anschlag . 2004. Winderosion und ihre Folgen im Waldgrenzbereich und in der alpinen Stufe einiger nordfinnischer Fjelle. Geoöko 25:203–224. Google Scholar

33.

M. J. Hutchings and E. A. John . 2003. Distribution of roots in soil, and root foraging activity. In H. de Kroon and J. W. Visser , editors. Root ecology Berlin Springer. 33–60. Google Scholar

34.

IPCC 2007. Climate Change 2007: the physical basis. Contribution of Working Group I to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change Cambridge Cambridge University Press. pp. Google Scholar

35.

M. Iseli and F. H. Schweingruber . 1989. Sichtbarmachen von Jahrringen für dendrochronologische Untersuchungen. Dendrochronologia 7:145–157. Google Scholar

36.

M. B. Jackson and B. Ricard . 2003. Physiology, biochemistry and molecular biology of plant root systems subjected to flooding of the soil. In H. de Kroon and J. W. Visser , editors. Root ecology Berlin Springer. 193–213. Google Scholar

37.

B. Jonsell Flora Nordica Vol. 1 2000. Stockholm The Bergius Foundation. pp. Google Scholar

38.

P. Kallio and J. Lehtonen . 1975. On the ecocatastrophy of birch forests caused by Oporinia autumnata (Bkh.) and the problem of reforestation. In F. E. Wielgolaski , editor. Fennoscandian tundra ecosystems. Part 2: animals and system analysis Berlin Springer. 174–180. Google Scholar

39.

P. S. Karlsson and K. O. Nordell . 1996. Effects of soil temperature on the nitrogen economy and growth of mountain birch seedlings near its presumed low temperature distribution limit. Ecoscience 3:183–189. Google Scholar

40.

P. S. Karlsson, L. F. Schleicher, and M. Weih . Seedling growth characteristics in three birches originating from different environments. Ecoscience 2000. 7:80–85. Google Scholar

41.

P. S. Karlsson, M. Weih, and C. Borg . 2005. Mountain birch growth in relation to climate and herbivores. F. E. Wielgolaski , editor. In. Plant ecology, herbivory, and human impact in Nordic mountain birch forests Berlin Springer. 71–86. Google Scholar

42.

G. Kashulina, C. Reimann, T. E. Finne, J. H. Halleraker, M. Äyräs, and V. A. Chekushin . 1997. The state of the ecosystems in the central Barents region: scale, factors and mechanism of disturbance. The Science of the Total Environment 205:203–225. Google Scholar

43.

C. Körner 1998. Worldwide positions of alpine treelines and their causes. In M. Beniston and J. L. Innes , editors. The impacts of climate variability on forests Berlin Springer. 221–229. Google Scholar

44.

L. Kullman 1984. Transplantation experiments with saplings of Betula pubescens ssp. tortuosa near the tree-limit in central Sweden. Holarctic Ecology 7:289–293. Google Scholar

45.

L. Kullman 2002. Rapid recent range-margin rise of tree and shrub species in the Swedish Scandes. Journal of Ecology 90:68–77. Google Scholar

46.

L. Kullman 2003. Recent reversal of neoglacial climate cooling trend in the Swedish Scandes as evidenced by mountain birch tree-limit rise. Global and Planetary Change 36:77–88. Google Scholar

47.

L. Kullman 2004. Tree-limit landscape evolution at the southern fringe of the Swedish Scandes (Dalarna province)—Holocene and 20th century. Fennia 182:73–94. Google Scholar

48.

L. Kullman 2005. Wind-conditioned 20th century decline of birch treeline vegetation in the Swedish Scandes. Arctic 58:286–294. Google Scholar

49.

E. Laitakari 1935. The root system of birch (Betula verrucosa and odorata). Acta Forestalia Fennica 41:1–216. Google Scholar

50.

J. Lehtonen and R. K. Heikkinen . 1995. On the recovery of mountain birch after Epirrita damage in Finnish Lapland, with a particular emphasis on reindeer grazing. Ecoscience 2:349–356. Google Scholar

51.

K. Lempa, S. Neuvonen, and H. Tømmervik . 2005. Effects of reindeer grazing on pastures in a mountain birch ecosystem. In F. E. Wielgolaski , editor. Plant ecology, herbivory, and human impact in Nordic mountain birch forests Berlin Springer. 157–164. Google Scholar

52.

J. Löffler 2000. High mountain ecosystems and landscape degradation in northern Norway. Mountain Research and Development 20:356–363. Google Scholar

53.

J. Löffler, A. Lundberg, O. Rössler, A. Bräuning, G. Jung, R. Pape, and D. Wundram . 2004. The alpine treeline under changing land use and changing climate: approach and preliminary results from continental Norway. Norsk Geografisk Tidsskrift 58:183–193. Google Scholar

54.

J. L. Lozán and H. Kausch . 1998. Angewandte Statistik für Naturwissenschaftler Berlin Parey. pp. Google Scholar

55.

S. Neuvonen, P. Niemelä, and T. Virtanen . 1999. Climatic change and insect outbreaks in boreal forests: the role of winter temperatures. Ecological Bulletins 47:63–67. Google Scholar

56.

L. Oksanen and R. Virtanen . 1995. Topographic, altitudinal and regional patterns in continental and suboceanic heath vegetation of northern Fennoscandia. Acta Botanica Fennica 153:1–80. Google Scholar

57.

L. Oksanen, J. Moen, and T. Helle . 1995. Timberline patterns in northernmost Fennoscandia. Relative importance of climate and grazing. Acta Botanica Fennica 153:93–105. Google Scholar

58.

J. Olofsson, S. Stark, and L. Oksanen . 2004. Reindeer influence on ecosystem processes in the tundra. Oikos 105:386–396. Google Scholar

59.

J. Ovaska, J. Nilsen, F. E. Wielgolaski, H. Kauhanen, R. Partanen, S. Neuvonen, L. Kapari, O. Skre, and K. Laine . 2005. Phenology and performance of mountain birch provenances in transplant gardens: latitudinal, altitudinal and oceanity-continentality gradients. In F. E. Wielgolaski , editor. Plant ecology, herbivory, and human impact in Nordic mountain birch forests Berlin Springer. 99–115. Google Scholar

60.

H. Seppä 1996. Post-glacial dynamics of vegetation and tree-lines in the far north of Fennoscandia. Fennia 174:1–96. Google Scholar

61.

M. Seppälä 2004. Wind as a geomorphic agent. in. cold climates Cambridge Cambridge University Press. pp. Google Scholar

62.

W. K. Smith, M. J. Germino, T. E. Hancock, and D. M. Johnson . 2003. Another perspective on altitudinal limits of alpine timberlines. Tree Physiology 23:1101–1112. Google Scholar

63.

B. Solberg, H. Tømmervik, D. Thannheiser, and S. Neuvonen . 2005. Economic limits and possibilities for sustainable utilization of northern birch forests. In F. E. Wielgolaski , editor. Plant ecology, herbivory, and human impact in Nordic mountain birch forests Berlin Springer. 219–233. Google Scholar

64.

M. Sonesson and J. Hoogesteger . 1983. Recent tree-line dynamics (Betula pubescens Ehrh. ssp. tortuosa [Ledeb.] Nyman) in northern Sweden. Collection Nordicana 47:47–54. Google Scholar

65.

B. Sveinbjörnsson 1987. Biomass proportioning as related to plant size in juvenile mountain birch near Abisko, Swedish Lapland. Reports from the Kevo Subarctic Research Station 20:1–8. Google Scholar

66.

B. Sveinbjörnsson 2000. North American and European treelines: external forces and internal processes controlling position. Ambio 29:388–395. Google Scholar

67.

B. Sveinbjörnsson, A. Hofgaard, and A. Lloyd . 2002. Natural causes of the tundra-taiga boundary. Ambio Special Report 12:23–29. Google Scholar

68.

C. Uhlig and A. Zink . 2006. Changes in organic horizon soil properties due to reindeer herding and changing management. In B. C. Forbes, M. Bölter, L. Müller-Wille, J. Hukkinen, F. Müller, N. Gunslay, and Y. Konstantinov , editors. Reindeer management in northernmost Europe: linking practical and scientific knowledge in social-ecological systems Berlin Springer. 245–264. Google Scholar

69.

T. Virtanen, S. Neuvonen, and A. Nikula . 1998. Modelling topoclimatic patterns of egg mortality of Epirrita autumnata (Lepidoptera: Geometridae) with a geographical information system: predictions for current climate and warmer climate scenarios. Journal of Applied Ecology 35:311–322. Google Scholar

70.

M. Weih 1998. Seasonality of nutrient availability in soils of subarctic mountain birch woodlands, Swedish Lapland. Arctic and Alpine Research 30:19–25. Google Scholar

71.

M. Weih and P. S. Karlsson . 1997. Growth and nitrogen utilization in seedlings of mountain birch (Betula pubescens ssp. tortuosa) as related to plant nitrogen status and temperature: a two-year study. Ecoscience 4:365–373. Google Scholar

72.

M. Weih and P. S. Karlsson . 1999. The nitrogen economy of mountain birch seedlings: implications for winter survival. Journal of Ecology 87:211–219. Google Scholar

73.

M. Weih and P. S. Karlsson . 2002. Low winter soil temperature affects summertime nutrient uptake capacity and growth rate of mountain birch seedlings in the subarctic, Swedish Lapland. Arctic, Antarctic, and Alpine Research 34:434–439. Google Scholar

74.

F. E. Wielgolaski 1997. Fennoscandian tundra. In F. E. Wielgolaski , editor. Polar and alpine tundra Amsterdam Elsevier. 27–83. Google Scholar
Kerstin Anschlag, Gabriele Broll, and Friedrich-Karl Holtmeier "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(4), 609-616, (1 November 2008). https://doi.org/10.1657/1523-0430(07-087)[ANSCHLAG]2.0.CO;2
Accepted: 1 April 2008; Published: 1 November 2008
Back to Top