Targets and tools for the maintenance of forest biodiversity in actual

advertisement
ВІСНИК ЛЬВІВ. УН-ТУ
Серія географічна. 2004. Вип. 31. С. 43–55
VISNYK LVIV UNIV
Ser.Geogr. 2004. №31. Р.43–55
ТЕОРЕТИЧНІ ПИТАННЯ ЛАНДШАФТОЗНАВЧИХ ДОСЛІДЖЕНЬ
УДК 911:2
TOWARDS TARGETS AND TOOLS FOR THE MAINTENANCE OF FOREST
BIODIVERSITY IN ACTUAL LANDSCAPES
P. Angelstam 1, J. Törnblom 2
1
Department of conservation biology and Faculty of forest science, Swedish university of
agricultural sciences, SE-730 91 Riddarhyttan, Sweden.
2
Department of natural sciences, Örebro university, SE-701 82 Örebro, Sweden.
Succeeding in maintaining forest biodiversity can even be viewed as an acid test of
sustainability as a whole. The principle of sustainable forest management has stimulated a
proliferation of a number of criteria and indicators. However, to achieve ecological
sustainability, it is vital that the monitoring of a suite of relevant indicators are compared
with targets to assess both status and, if repeated, the trends in actual landscapes. We first
describe how traditional measurement tools for describing wood resources need to be
complemented by monitoring of the elements of biodiversity including species, habitats and
functions at multiple spatial scales. Second, we review examples of empirical non-linear
relationships between presence and fitness of species’ populations and different levels of
anthropogenic change in their respective habitats at different spatial scales, and how this
can be used to formulate science-based performance targets for indicators. Finally, using
the results from monitoring and with relevant targets, it is possible to make assessments of
the status of a certain criterion, such as biodiversity. In this section examples of practical
assessment tools such as gap analysis and habitat models are presented for strategic and
tactic planning of operational management for protection, management and re-creation of
different elements of biodiversity. Additionally, the need to assure communication with
iterated feed-back of the results of assessments to managers and policy-makers is discussed.
We finally stress the need for international co-operation, for example by establishing a
network of case studies in the form of “landscape laboratories” in gradients of forest
alteration and different governance systems together with managers of forests and
woodland representing different trajectories towards SFM.
Key words: Sustainable forest management, biodiversity, landscapes, spatial
scales.
Sustainable forest management (SFM) represents a vision for the use of forests
based on satisfying ecological, economic and social values [69]. The current starting point
for trajectories towards the SFM vision, however, varies considerably among countries and
regions with different socio-economic settings and ecosystems [12]. Sweden, Switzerland
and Russia provide three contrasting European examples where the focus has been and is
on quite different SFM criteria. In Sweden, where forests have been subject to intensive
management for sustained yield of wood for a long time, biodiversity interpreted as the
________________________
© Angelstam P., Törnblom J., 2004
44
P.Angelstam, J.Törnblom
maintenance of viable populations of naturally occurring species has been a major driver of
changes towards SFM during the past decades [3]. By contrast, Switzerland’s steep terrain
has promoted management for protective functions of the conifer-dominated mountain
forests, and recently with an additional focus on biodiversity [38, 57]. Finally, in remote
parts of Russia such as in the Komi Republic in northeasternmost Europe, large-scale
logging of forests started only recently, and large intact forest areas still remain [88]. Here a
major current challenge is to use not previously managed forest landscapes for the
development of human welfare in a sustainable manner [64], but also to maintain the
functionality of the last remaining large intact forest areas.
The international and national policy arenas, the forest sector, non-governmental
organisations and scientists are the major actors trying to develop and interpret international
and national policies on sustainable development in forests. In Europe the Ministerial
Conference for the Protection of Forests in Europe (MCPFE) has derived a reasonably
complete set of indicators defining different SFM criteria, including biodiversity and forest
health [69]. The MCPFE’s criteria regarding ecological dimensions of SFM ultimately can
be viewed as proxies of the natural capital [23]. Consequently, the maintenance of
biodiversity and resulting products and services is a prerequisite for satisfying the economic
and social benefits of forests and woodland including both terrestrial and aquatic systems
[87].
Defining forest biodiversity. The natural potential vegetation of a large part of the
terrestrial ecosystems in Europe is forest [51]. Forest biodiversity is made up by species,
habitat structures and processes found in ecosystems with trees, and can therefore be
maintained in both natural forests and in remnants of pre-industrial cultural woodland
landscapes.
First, policies related to biodiversity of European forests and woodland make
explicit reference to the concept of naturalness [e.g., 69]. In spite of the ambiguity of this
concept [27], it is obvious that forest biodiversity indicators should represent elements
found in naturally dynamic forests [60, 8]. Second, the maintenance of ecological values
found in pre-industrial cultural landscapes are highlighted [37]. Reference areas for both
visions are characterised by the presence of habitat elements such as dead wood, large old
trees, a diversity of tree species, old-growth stands and the ecological integrity of aquatic
systems [e.g., 36, 82]. While ‘laissez-faire’ management usually can enhance the
naturalness vision, the maintenance of cultural landscapes require a certain amount of social
and cultural capital [62]. In other words the maintenance of forest biodiversity encompasses
two sets of broad visions depending on the history of forests and woodland in the actual
landscape [e.g., 1]. The development of SFM should reflect both these visions.
The cover and types of forests and woodland are dynamic, including both
degradation and restoration related to socio-economic changes [58]. Consequently,
monitoring and assessment of SFM should not only encompass the area covered by forest
and woodland at present, but rather a geographically contiguous units representing actual
landscapes or ideally natural units such as watersheds. Such areas, hereafter called
landscapes, have a wide range of implementing actors representing different institutions
ranging from the forest and wood industry, small private landowners and commons to
different public interests. However, even if the forest cover is constant, the relative
proportion of different governance systems varies considerably among landscapes, regions
and countries [e.g., 13]. This should be expected to have strong effects on the ways of and
extent to which different policies can be implemented using different management and
planning tools [30, 11].
TOWARDS TARGETS AND TOOLS FOR THE MAINTENANCE ...
45
Monitoring indicators of biodiversity. Monitoring the elements of forest
biodiversity can be made at multiple spatial scales. At the international and national policy
levels, indicators aim at communicating the status and trends of biodiversity to policymakers and the general public. However, to allow effective operations, indicators should
also be developed and applied at the level of forest management units. Such practical
indicators need to be adapted to the local conditions and resources available to different end
users ranging from corporate companies to the owners of small non-industrial private
forests.
Because it is impossible to measure all aspects of biodiversity, there is a need for
cost-efficient monitoring tools. At the policy level, international reporting is based on
individual countries providing data like those collected in national forest inventories. In
general, however, such programmes provide neither sufficient data on compositional (e.g.
occurrence of specialised species), structural (e.g. quality of habitat networks), nor
functional elements of biodiversity (e.g. ecosystem processes) to allow comprehensive
monitoring [14]. Conversely, at the management unit level, the development of
comprehensive biodiversity monitoring systems is still in its infancy. Based on the MCPFE
indicators, attempts to build a system for biodiversity monitoring based on the composition,
structure and function of natural forest ecosystems have been proposed [e.g., 44].
Evaluations of this system in European land use history gradients showed promising results
at the scale of both stands and landscapes [65, 8, 10]. Additionally, there is a need for
continuous evaluation both of the scientific validity of indicators and of the degree to which
the results from indicator systems can be interpreted and communicated to stakeholders at
all relevant levels [85]. These factors probably explain why real-life applications of such
monitoring systems are still rare.
Performance targets. Habitat loss is a major factor affecting directly or indirectly
the global decline of biodiversity [33]. With a biodiversity conservation perspective, the
evaluation of hypotheses claiming species-specific “extinction thresholds” defined as the
minimum amount of habitat required for the persistence of species in the landscape is an
urgent task [54, 55, 28]. Appearing empirical evidence show that human-driven landscape
changes have resulted in the trespassing of such critical levels of habitat loss. This applies
to structural elements such as dead wood [20, 21], large habitat patches [56] and reduced
amount of certain tree species [35]. In Europe an obvious consequence of this is that
countries with a lower intensity and shorter history of forest use still host populations of
species specialising on natural forest structures, while other countries do not [52, 65, 15].
Even though research on thresholds remains in its infancy, ecologically-based
targets inspired from such thresholds could be used to postulate management and
conservation strategies. We stress, however, the need for explicitly recognising uncertainty
and, rather than proposing target numbers, there should be a focus on probabilistic targets
defined using a variety of indicators, and on the associated “zones of risk” [e.g. 55, 61]. A
general procedure for identifying thresholds to be used in the determination of conservation
targets in forests was proposed by Angelstam et al. [6]: 1. Stratify the forests into broad
cover types as a function of their natural disturbance regimes; 2. Describe the historical
spread of different anthropogenic impacts in the forest region of concern that moved the
system away from the reference conditions of naturalness or pre-industrial cultural
landscapes; 3. Identify appropriate response variables (e.g. focal species, functional groups
or ecosystem processes) that are affected by habitat loss and fragmentation; 4. For each
forest type identified in step 1, combine steps 2 and 3 to look for the presence of non-linear
responses and to identify zones of risk and uncertainty. 5. Identify the “currencies” (i.e.
46
P.Angelstam, J.Törnblom
species, habitats, and processes) which are both relevant and possible to communicate to
stakeholders. 6. Combine information from different indicators selected.
Assessment of status and trends. Management of sustainable wood production as
well as management for protection, management and restoration of the elements of
biodiversity require planning at multiple scales. The approach used in most planning
systems for large-scale forestry is hierarchical within a forest management unit (FMU) [24].
The planning problem is usually divided into three sub-processes: strategic, tactical and
operational. Strategic planning means to decide on long-term goals covering an entire
rotation and tactical planning to select among different alternatives based on the strategic
goals, but on a shorter time horizon. Operational planning involves determining the actual
operations. The same logic can be used to build a toolbox of analytic tools for the
assessment of the structural elements of biodiversity being the focus in conservation,
management, and restoration [11].
At the strategic level gap analysis is a tool for assessing the extent to which
environmental policies succeed in maintaining biodiversity by protection, management and
restoration of habitats [80]. Originally developed in the USA, gap analyses have been used
in terrestrial systems to increase society’s awareness about conservation needs and to guide
the practical implementation of such policies. The rationale for focusing on habitat (i.e.
structural elements of biodiversity) is that it serves as a proxy for the maintenance of viable
populations of species, vital ecosystem processes and resilience to external disturbance
[e.g., 36]. Originally gap analyses focused on representation i.e., that the different types of
conservation areas should reflect the natural composition of different ecosystems [80].
Angelstam and Andersson [4] and Lõhmus et al. [49] developed the idea for Sweden and
Estonia, respectively, further by combining measurements of the habitat area with
information about thresholds for the amount and quality of habitats needed to maintain
viable populations within an ecoregion (Table 1).
Table 1
The ABC of gap analysis for strategic conservation planning
Explanation
Reference/Benchmark (e.g., conditions such as found in naturally
dynamic forests of pre-industrial cultural landscapes)
The present situation
Science-based threshold
Representation (e.g., analysis of representation for different types of
forest and wooded grassland)
Long-term goal
Area gap (e.g., identification of area gaps for certain types of
vegetation)
Code
A
B
C
A-B
A*C
B-(A*C)
Given a policy which can be interpreted scientifically, like maintaining viable
populations of naturally occurring species, reference conditions (A) such as found in
naturally dynamic forests of pre-industrial cultural landscapes, can be quantified. By
comparing the present situation (B) with A for different types of forest and wooded
grassland, analyses of representation can be made. Finally, with knowledge about the
quantitative requirements at the population level, expressed as a proportion of A, long-term
targets can be formulated and compared with B, allowing the identification of area gaps
for a certain type of vegetation. Next, to assure functional connectivity of the total area,
TOWARDS TARGETS AND TOOLS FOR THE MAINTENANCE ...
47
spatially explicit analyses need to be done for the tactical decisions regarding protection,
management and restoration.
When gap analysis has been performed within a particular ecoregion, the forest
types for which area gaps have been identified also need to be evaluated as to the extent to
which they actually provide functional habitat for the specialised focal species. One
approach to evaluate the functionality of existing networks of patches of different forest
types is habitat suitability modelling [e.g., 81, 7]. This means combining spatially explicit
land cover data with quantitative knowledge about the requirements of specialised species
and producing spatially explicit maps describing the probability that a species is found in a
landscape. With adequate quantitative data defining habitat variables and parameter values
for a suite of particular focal species carefully selected to represent all forest types of
concern, a series of predictive models can be built to assess the functionality of different
habitat networks. This requires quantitative information on the habitat requirements of the
species at different spatial scales. In general, a habitat model for a given species should
build on the following variables: land cover type(s) constituting habitat, habitat patch size,
landscape-scale proportion of suitable habitat, and habitat duration [14]. Using, for
example, neighbourhood analysis techniques in Geographic Information Systems, the
functionality of the network of each representative habitat (one or several land cover types)
can be evaluated. Because a landscape usually contains a range of types of forest
vegetation, a suite of species need to be modelled [74, 73]
The procedure suggested above provides a general basis for the assessment and
subsequent planning of habitat networks. The development of practical tools using focal
species is, however, subject to uncertainty depending on the knowledge about the different
parameters included in the models. Another factor influencing the development of practical
tools is the thematic and spatial resolution of the land cover data available to the planner
[89]. For example, depicting the habitat of species dependent on dead wood (e.g. many
species of woodpeckers, beetles, and wood-decay fungi) require spatially explicit data on
the occurrence of this resource across the landscape. Such data is not currently available
from forest management maps or classified satellite images, and therefore additional
ancillary data needs to be collected in the field. Until such data become available, surrogate
measures such as vicinity to roads as a proxy for the amount of dead wood could be used
[21].
Ideally, focal species should be chosen among the most demanding species for a
range of landscape attributes [43]. Since the most demanding species vary among habitats
and scales, the suite of focal species should include representatives from a number of
different taxa with different ecologies or functional groups [e.g., 2, 58]. Finally, each model
should be validated in order to test how reliably one can predict occurrences of the focal
species in real-world landscapes [81].
The need for spatially explicit forest management. Landscapes are not constant
[17]. The variation among different European regions in the trajectories of the development
towards the SFM vision is a reflection of this. Because most of Europe’s landscapes have
an origin as forests or wooded grasslands, forests and forestry must be seen in a landscape
perspective [29]. Current driving forces affecting European landscapes include the
macroeconomic development affecting human population migration from the periphery to
centre, the active expansion of the transport infrastructure, and the energy sector. Climatic
change is another, but less predictable factor. As shown in the following three examples the
effects of different elements of forest biodiversity through changes in the land cover and the
spatial configuration are complex.
48
P. Angelstam, J. Törnblom
The implementation of the EU Habitats Directive by establishing a network of
conservation areas with a favourable conservation status is one example. The appearing
knowledge about thresholds for the amount of habitat viable populations of species need,
i.e. reflecting the resources they require, has clear implications for biodiversity
management. Even if still under development, it is fair to state that the maintenance of the
species listed in the EU Birds and Habitats Directive requires functional networks of
suitable habitats. Establishment of functional habitat networks may both suffer and benefit
from the current land cover changes in Europe. Abandonment of agricultural land in the
periphery of economic development lead to increased cover and connectivity for forest
species [5]. The effects on the future functionality of habitat networks – or “green
infrastructure” cannot be understood and planned without spatially explicit analyses.
The effect of land cover on aquatic systems is another example [83]. Interestingly
enough, the EC Water Framework Directive has recently reinforced a drainage basin
perspective on water issues and aquatic biodiversity. In order to maintain and restore
surface and groundwater to “Good Ecological Status” dead wood is a key structure in
stream order 1-4 [50]. Degerman et al. [25] studied the relationship between brown trout
(Salmo trutta) and dead wood in Swedish streams and found a positive relationship
between the abundance and size of trout and dead wood. The gap between the present
amount of dead wood and the amount found in reference landscape is, however, about 1-2
orders of magnitude [46]. With limited resources to leave harvestable wood in the forest to
restore the quality of aquatic systems, the spatial effects of retention on the functionality
will be important.
Finally, because different forest vegetation types host different species, the
maintenance of functional networks for species with different specialisations should be seen
as separate and not necessarily overlapping green infrastructures. The coniferous and
deciduous component in a landscape can serve as an example. In the watershed of the lake
Hjälmaren in south-central Sweden, the coniferous forest is managed and forms a stable
patch dynamics for at least species not requiring old-growth elements. The deciduous forest
originates from abandoned wooded grasslands, and around the lake from the lowering of
the lake level in the late 19th century [76, 77]. In contrast to the coniferous forest, the
deciduous forest is the result of a series events driven by socio-economic change [53]. To
maintain species of the deciduous forest in the long term, the deciduous component needs
to be restored in what is now coniferous forest. However, dense populations of moose and
deer severely hamper the recruitment of at least the most important tree species for
specialised species (aspen, sallow and rowan) [16]. Another barrier is the poor integration
between the management of trees in forest and on the agricultural landscapes.
Towards integrated and transdisciplinary approaches. Science develops
indicators because they are required for the policy implementation process. The MCPFE
criteria and indicators focus on the state of a system. However, such results need to be put
into the context of continuous evolution of policies. Indicators should thus be seen as
describing the success of a policy implementation feedback loop that begins with a Pressure
leading to a State and resulting in a Response [70]. The PSR model, and subsequent
elaboration of it, has been successful in helping structure the use of indicators [48]. The
socio-economic context and the associated governance systems drive the state. Based on
monitoring of the state of productive functions and biodiversity, and performance targets
for the different indicators, assessments can be made. If the outcome of such assessments
indicate the need for active response resulting in gradual modification of the state of the
landscapes in the desired direction, management must be planned and implemented.
ЩОДО ЦІЛЕЙ ТА ЗАСОБІВ ПІДТРИМКИ ЛІСОВОГО БІОРІЗНОМАНІТТЯ ...
49
There is a growing insight that there are complex interactions between the parts of
different ecosystems and institutions, which require transdisciplinary landscape-scale
approaches [66, 79]. In Europe the EC Water Framework Directive stresses this. In spite of
the presence of relevant tools from the natural and social sciences [11, 19], effective use of
them in a transdisciplinary fashion to facilitate the implementation of sustainable
development policies is rare in the real world [18, 26]. Apparently, working across disciplines in landscape analyses is a major challenge. In a comparison of two case studies
Jakobsen et al. [34] revealed a set of similar individual-based, group-based and organisation
culture-based barriers. However, even if they proposed a number of recommendations to
scientists across disciplines, the limited number of case studies precludes thorough analyses
of the effects of ecological, institutional and cultural contexts on both barriers and
facilitators to bridge them.
The “Landscape Lab” approach. Even with a wide-spread insight that spatial
forestry is necessary, the variation in ownership patterns and governance systems may
provide both barriers and bridges to the application of spatially explicit assessment and
planning. A major challenge is to achieve integration among actors. Researchers and
managers accomplish most of their work in isolation and then present their results to
decision-makers. There are hence a number of barriers, in particular when attempting to
apply a landscape approach to the conservation of biodiversity [e.g., 31]. Using landscapes
as laboratories is one approach. To describe this Kohler [39, p. 212] used the concept
‘practices of place’ whereby it is “…the arrangement of spatial elements that provides
critical evidence of relations between creatures and their environment…”. Places are thus to
the field ecologist what experimental set-ups are to laboratories. Co-ordinated case studies
based on the idea of ‘Practices of place’ can thus be designed stratified in replicated land
use history gradients. This can be made both in time and space.
For example, the historical occurrence of species dependent on dead wood can be
compared with the decline of dead wood over time [e.g., 47], and the presence today can be
made in landscapes located forest history gradients [e.g., 8, 10, 82]. This is consistent with
the combination of case studies and quantitative data termed triangulation used in social
sciences [19]. This approach may actually make it possible to “look into the future” to see
what new pressures on forest ecosystems which can be expected. The gradient of
commercial thinning is one example. While this has a very long tradition in Central Europe,
this forest history phase reached northern Sweden in the 1960s and is now entering Russia.
Even if Europe is becoming more and more integrated in a political sense, the
states of the forests and woodland ecosystems are still highly variable in different regions
and countries and range from large intact natural areas to remnants of cultural woodlands.
Additionally there is considerable variation as determined by the type of ownership and
resulting governance system. These two dimensions should form the basis for a design for
communication, research and development towards SFM in case studies representing
gradients in social-ecological systems (Table 2). At the pan-European level this matrix
would then cover the gradients in regional macroeconomic development, rural-urban
transitions and with different sets of problems related to biodiversity and ecosystem
integrity. To promote this idea we encourage the development of case studies not only as a
research tool, but also as a tool for demonstration of bridges to deal with implementation
obstacles. One approach is the Canadian model forest network, which together forms a
partnership between individuals and organisations sharing the common goal of sustainable
forest management (see www.modelforest.net). Such a network of forest management units
consisting of actual landscapes with their characteristic ecosystems, actors and economic
activities can be used as the sites for spreading good examples.
50
P. Angelstam, J. Törnblom
Forest and woodland system
Table 2
Idealised matrix for selecting “Landscape Labs”
(Based on [86, 67, 40, 84, 29, 32, 68, 88, 41, 63, 72, 75, 22, 45, 53, 71, 83, 6, 42, 57, 78])
Non-Industrial
Private
Wooded Agricultural subgrasslan sidies and land
d
abandonment
Plantati Energy forests
ons on
formerly
cleared
land
Altered Southern
tree
Fennoscandia
species
composi
tion
SemiBaltic States
natural
manage
d
Benchm Pre-industrial
ark/
cultural
referenc landscapes
e
Governance system
Commercial Company Public
management
Ecological meat
Urban green
production
space and
forestry
Eucalypt forests in
Exotic conifer
Portugal
plantations in
Scotland
Protected
Recreation
forest
Habitat
restoration
in Western
Europe
Intensive forest
management for fiber
and wood
Transformation
of coniferous to
deciduous forest
Removal of
undesired
tree species
Managed forests in
Fennoscandia
State-owned
forests in
Fennoscandia
Protection
forests in
mountains
Woodland Key
Habitats
Remote parts of
Russia
Russian
Strict State
Reserves
Ideally, adaptive management teams [18] should be formed. This means that the
level of the actual case study participatory experts and stakeholders including researchers,
land managers and policy-makers share decisions and responsibilities toward the success or
failure of the strategy they jointly adopted. To put this bottom-up reflexive iterative
procedure into action and to design management applications, we suggest the development
of an international network of adaptive management teams. This network should be charged
with testing different approaches to the management of forests that will ensure that
biodiversity is restored in areas where it has been lost and maintained where forestry
intensification has yet to occur.
Acknowledgements. This paper is based on the plenary lecture presented at the
symposium in Florence, the work with the follow-up volume of Ecological Bulletins [9] to
the EC-funded BEAR project reported as Ecological Bulletins 50 [44], the development of
Aquatic Gap analyses with the landscape ecology group at Örebro university, and
discussions during a workshop held within the COST Action E 25 “European Network for
long-term Forest Ecosystem and Landscape Research (ENFORS)” with Peter Biber, Hubert
Hasenauer Norbert Kräuchi, Paul Tabbush and Uwe Schneider.
ЩОДО ЦІЛЕЙ ТА ЗАСОБІВ ПІДТРИМКИ ЛІСОВОГО БІОРІЗНОМАНІТТЯ ...
51
________________________
1. Agnoletti M. Introduction: the development of forest history research // Methods
and approaches in forest history. – Wallingford, 2000.
2. Angelstam P. Maintaining and restoring biodiversity by developing natural
disturbance regimes in European boreal forest // Journal of Vegetation Science.
1998. No 9(4).
3. Angelstam P. Forest biodiversity management - the Swedish model // Towards
Forest Sustainability. Canberra, Washington, 2003.
4. Angelstam P., Andersson L. Estimates of the needs for nature reserves in Sweden
// Scandinavian Journal of Forestry Supplement. 2001. No 3.
5. Angelstam P., Boresjö-Bronge L., Mikusinski G. et al. Assessing village
authenticity with satellite images – a method to identify intact cultural
landscapes in Europe // Ambio. 2003c. No 33(8).
6. Angelstam P., Boutin S., Schmiegelow F. et al. Targets for boreal forest
biodiversity conservation – a rationale for macroecological research and adaptive
management // Ecological Bulletins. 2004e. No 51.
7. Angelstam P., Bütler R., Lazdinis M. et al. Habitat thresholds for focal species at
multiple scales and forest biodiversity conservation – dead wood as an example
// Annales Zoologici Fennici. 2003b. No 40.
8. Angelstam P., Dönz-Breuss M. A system for indicators of forest biodiversity at
the stand scale (INFOBIOS) - differences in naturalness within European forest
history gradients. // Ecological Bulletins. 2004. No 51.
9. Angelstam P., Dönz-Breuss M., Roberge J.-M. (eds.) Targets and tools for the
maintenance of forest biodiversity // Ecological Bulletins. 2004f. 51
10. Angelstam P., Edman T., Dönz-Breuss M., Wallis deVries M. Land management
data and terrestrial vertebrates as indicators of forest biodiversity at the
landscape scale // Ecological Bulletins. 2004c. 51.
11. Angelstam P., Mikusinski G., Rönnbäck B.-I. et al. Two-dimensional gap
analysis: a tool for efficient conservation planning and biodiversity policy
implementation // Ambio. 2003a. No 33.
12. Angelstam P., Persson R., Schlaepfer R. The sustainable forest management
vision – barriers and bridges for the maintenance of biodiversity // Ecological
Bulletins. 2004a. No 51
13. Angelstam P., Pettersson B. Principles of present Swedish forest biodiversity
management // Ecol. Bull. 1997. No 46.
14. Angelstam P., Roberge J.-M., Dönz-Breuss M. et al. Monitoring of forest
biodiversity at multiple spatial scales – from the policy level to actual
management units // Ecological Bulletins. 2004b. No 51.
15. Angelstam P., Roberge J.-M., Lõhmus A. et al. Habitat modelling as a tool for
landscape-scale conservation – a review of parameters for focal forest birds //
Ecological Bulletins 2004d. No 51.
16. Angelstam P., Wikberg P. E, Danilov P. et al. Effects of moose density on timber
quality and biodiversity restoration in Sweden, Finland and Russian Karelia //
Alces. 2000. No 36.
17. Bengtsson J., Angelstam P., Elmqvist T. et al. Reserves, resilience and dynamic
landscapes // Ambio. 2003. No 32 (6).
18. Boutin S. et al. The active adaptive management experimental team: a
collaborative approach to sustainable forest management // Advances in forest
management: from knowledge to practise. Edmonton, 2002.
52
P. Angelstam, J. Törnblom
19. Bryman A. Social research methods. – Oxford, 2001.
20. Bütler R., Angelstam P., Schlaepfer R. Quantitative snag targets for the threetoed woodpecker, Picoides tridactylus // Ecological Bulletins. 2004. No 51.
21. Bütler R., Angelstam P., Ekelund P., Schlaepfer R. Dead wood threshold values
for the Three-toed woodpecker in boreal and sub-alpine forest // Biological
Conservation. 2004. In press.
22. Carey P. D., Short C., Morris C. et al. The multi-disciplinary evaluation of a
national agri-environment scheme // Journal of Environmental Management.
2003. No 69.
23. Costanza R., Daly H. E., Bartholomew J. A. Goals, agenda, and policy recommendations for ecological economics // Ecological Economic. – New York,
1991.
24. Davis L. S., Johnson K. N., Bettinger P. S., Howard T. E. Forest management. To
sustain ecological, economic, and social values. – Boston, 2001.
25. Degerman E., Sers B., Törnblom J., Angelstam P. Large woody debris and
brown trout in small forest streams – towards targets for assessment and
management of riparian landscapes // Ecological Bulletins. 2004. No 51.
26. Duinker P. N., Trevisan L. M. Adaptive management: progress and prospects for
Canadian forests // Towards sustainable management of the boreal forest. –
Ottawa, 2003.
27. Egan D., Howell E.A. The historical ecology handbook. – Covelo, 2001.
28. Fahrig, L. Effect of habitat fragmentation on the extinction threshold: a synthesis
// Ecological Applications. 2002. No 12.
29. Farrell E. P., Führer E., Ryan D. et al. European forest ecosystems: building the
future on the legacy of the past // Forest Ecology and Management. 2000. No 132.
30. Fries C. et al. A review of conceptual landscape planning models for
multiobjective forestry in Sweden // Can. J. For. Res. 1998. No 28.
31. Gutzwiller K. J. (ed.) Applying landscape ecology in biological conservation.
Springer, 2002.
32. Hansson L. Key habitats in Swedish managed forests // Scandinavian Journal of
Forest Research Supplement. 2001. No 3.
33. Heywood V.H. (ed.) Global biodiversity assessment. – Cambridge. 1995.
34. Jakobsen C. H., Hels T. McLaughlin W. J. Barriers and facilitators to integration
among scientists in transdisciplinary landscape analyses: a cross-country
comparison // Forest Policy and Economics. 2004. No 6.
35. Jansson G., Angelstam P. Threshold levels of habitat composition for the
presence of the long-tailed tit (Aegithalos caudatus) in a boreal landscape //
Landscape Ecology, 1999. No 14.
36. Karr J. R. Health, integrity and biological assessment: the importance of
measuring whole things // Ecological integrity. – Washington, 2000.
37. Kirby K. J., Watkins C. The ecological history of European forests. Wallingford,
1998.
38. Kräuchi N., Brang P., Schönenberger W. Forests of mountainous regions: gaps in
knowledge and research needs // Forest Ecology and Management. 2000. No 132.
39. Kohler R. E. Landscapes and labscapes. Exploring the lab-field border in
biology. Chicago, 2002.
40. Korhonen K.-M., Laamanen R., Savonmäki S. Environmental guidelines to
practical forest management. Helsinki, 1998.
ЩОДО ЦІЛЕЙ ТА ЗАСОБІВ ПІДТРИМКИ ЛІСОВОГО БІОРІЗНОМАНІТТЯ ...
53
41. Kumm K.-I. Sustainability of organic meat production under Swedish conditions
// Agriculture, Ecosystems and Environment. 2002. No 88.
42. Kurlavicius P., Kuuba R., Lukins M. et al. Identifying biologically valuable
forests in the Baltic States from forest databases // Ecological Bulletins. 2004.
No 51.
43. Lambeck R. J. Focal species: a multi-species umbrella for nature conservation //
Conservation Biology. 1997. No 11.
44. Larsson T.-B. et al. Biodiversity evaluation tools for European forests // Ecol.
Bull. 2001. No 50.
45. Larsson K., Simonsson G. Den halländska skogen – människa och mångfald //
Lännstyrelsen i Halland, Meddelande. 2003 No 7.
46. Liljaniemi P. et al. Habitat characteristics and macroinvertebrate assemblages in
boreal forest streams: relations to catchment silvicultural activities //
Hydrobiologia. 2002. No 474.
47. Linder P. and Östlund L. Structural changes in three mid-boreal Swedish forest
landscapes, 1885-1996. // Biological Conservation. 1998. No 85.
48. Linser S. Critical analysis of the basics for the assessment of sustainable
development indicators // Freiburger forstliche Forschung. 2001. Band 17.
49. Lõhmus A., Kohv K., Palo A., Viilma K. Loss of old-growth, and the minimum
need for strictly protected forests in Estonia // Ecological Bulletins. 2004. No 51.
50. Marcus W. A. et al. Mapping the spatial and temporal distributions of woody
debris in streams of the Greater Yellowstone Ecosystem, USA //
Geomorphology. 2002. No 44.
51. Mayer H. Wälder Europas. – Stuttgart, 1984.
52. Mikusinski G., Angelstam P. Economic geography, forest distribution, and
woodpecker diversity in Central Europe // Conservation Biology. 1998. No 12.
53. Mikusinski G., Angelstam P., Sporrong U. Distribution of deciduous stands in
villages located in coniferous forest landscapes in Sweden // Ambio. 2003. No 33(8)
54. Mönkkönen M., Reunanen, P. On critical thresholds in landscape connectivity: a
management perspective // Oikos. 1999. No 84.
55. Muradian R. Ecological thresholds: a survey // Ecological Economics. 2001.
no 38.
56. Mykrä S., Kurki S., Nikula A. The spacing of mature forest habitat in relation to
species-specific scales in manages boreal forests in NE Finland // Annales
Zoologici Fennici. 2000. No 37.
57. Neet C., Bolliger M. Biodiversity management in Swiss mountain forests // Ecol.
Bull. 2004. No 51.
58. Nilsson S.G., Hedin J., Niklasson M. Biodiversity and its assessment in boreal
and nemoral forest // Scandinavian Journal of Forest Research Supplement.
2001. No 3.
59. Nilsson S., Sallnäs O., Duinker P. Future forest resources of western and eastern
Europe. – Carnforth, 1992.
60. Peterken G. Natural Woodland: Ecology and Conservation in Northern
Temperate Regions. – Cambridge, 1996.
61. Phillis Y. A., Andriantiatsaholiniana L. A. Sustainability: an ill-defined concept
and its assessment using fuzzy logic // Ecological Economics, 2001. No 37.
62. Pierce Colfer C. J., Byron Y. (eds) People managing forests. – Bogor, 2001.
63. Pietsch S. A., Hasenauer H. Using mechanistic modeling within forest
ecosystem restoration // Forest Ecology and Management. 2002. No 159.
54
P. Angelstam, J. Törnblom
64. Pollard D. High conservation value forests in practise. – Glan, 2003.
65. Puumalainen J. Structural, compositional and functional aspects of forest
biodiversity in Europe. // Geneva timber and forest discussion papers. – New
York, 2001.
66. Rabeni C. F., Sowa S. P. A landscape approach to maintaining the biota of
streams // Integrating landscape ecology into natural resource management. –
Cambridge, 2002.
67. Rackham O., Moody J. The making of the Cretan landscape. – Manchester, 1996.
68. Raivio S., Normark E., Pettersson B., Salpakivi-Salomaa P. Science and
management of boreal forest biodiversity // Scand. J. For. Res. Suppl. 2001. No 3.
69. Rametsteiner E., Mayer P. Sustainable forest management and pan-European
forest policy // Ecol. Bull. 2004. No 51.
70. Rapport D. J., Friend A. M. Towards a comprehensive framework for
environmental statistics: a stress-response approach. – Ottawa, 1979.
71. Reed D. D., Jones E. A., Tomé M. and Araújo M.C. Models of potential height
and diameter for Eucalyptus globulus in Portugal // Forest Ecology and
Management. 2003. No 172.
72. Richardson J., Björheden R., Hakkila P. et al. (eds.) Bioenergy from sustainable
forestry. Guiding principles and practice. – Dordrecht, 2002.
73. Roberge J.-M., Angelstam P. Usefulness of the umbrella species concept as a
conservation tool // Conservation Biology. 2004. No 18(1).
74. Root K. V., Akcakya H. R., Ginzburg L. A multispecies approach to ecological
valuation and conservation // Conservation biology. 2002. No 17.
75. Roovers P., Hermy M., Gulinck H. Visitor profile, perceptions and expectations
in forests from a gradient of increasing urbanisation in central Belgium //
Landscape and Urban Planning. 2002. No 59.
76. Rydin H., Borgegård S.-O. Plant species richness on islands over a century of
primary succession: Lake Hjälmaren // Ecology. 1988. No 69(4).
77. Rydin H., Borgegård S.-O. Plant characteristics over a century of primary
succession on islands: Lake Hjälmaren. Ecology. 1991. No 72 (3).
78. Sandström U.G., Khakee A., Angelstam P. Urban planner’s knowledge of
biodiversity maintenance – an evaluation of six Swedish cities // Landscape and
Urban Planning. In press.
79. Schneider R.L., Mills E.L., Josephson D.C. Aquatic-terrestrial linkages and
implications for landscape management // Integrating landscape ecology into
natural resource management. – Cambridge, 2002.
80. Scott J.M., Tear T.H., Davis F.W. (eds.) Gap analysis: a landscape approach to
biodiversity planning. – Bethesda, 1996.
81. Scott J.M., Heglund P.J., Morrison M. et al. Predicting species occurrences:
issues of scale and accuracy. – Covelo, 2002.
82. Shorohova E., Tetioukhin S. Natural disturbances and the amount of large trees,
deciduous trees and coarse woody debris in the forests of Novgorod Region,
Russia // Ecological Bulletins. 2004. No 51.
83. Stålnacke P., Grimvall A., Libiseller C. et al. Trends in nutrient concentrations in
Latvian rivers and the response to the dramatic change in agriculture // Journal of
Hydrology. 2003. No 283.
84. Summers R. W., Mavors R. A. MacLennan A. M., Rebecca G. W. The structure of
ancient native pinewoods and other woodlands in the Highlands of Scotland //
Forest Ecology and Management. 1999. No 119.
ЩОДО ЦІЛЕЙ ТА ЗАСОБІВ ПІДТРИМКИ ЛІСОВОГО БІОРІЗНОМАНІТТЯ ...
55
85. Uliczka H., Angelstam P., Roberge, J.-M. Indicator species and biodiversity
monitoring systems for non-industrial private forest owners – is there a
communication problem? // Ecological Bulletins. 2004. No 51.
86. Wallis de Vries M. F. Large herbivores and the design of large-scale nature
reserves in western Europe // Conservation Biology. 1995. No 9.
87. Wiens J. Riverine landscapes: taking landscape ecology into the water //
Freshwater Biology. 2002. No 47.
88. Yaroshenko A.Yu., Potapov, P.V., Turubanova S.A. The intact forest landscapes
of Northern European Russia. – Moscow, 2001.
89. Young J. E., Sanchez-Azofeifa G. A. The role of geographical information
systems and optical remote sensing in monitoring boreal ecosystems //
Ecological Bulletins. 2004. No 51.
ЩОДО ЦІЛЕЙ ТА ЗАСОБІВ ПІДТРИМКИ
ЛІСОВОГО БІОРІЗНОМАНІТТЯ В АКТУАЛЬНИХ ЛАНДШАФТАХ
П. Ангельстам1, Й. Тьорнблм 2
Факультет лісівництва, Шведський університет сільськогосподарських наук
Ріддаргіттан, Швеція
2
Відділ природничих наук, Університет Оребро
Оребро, Швеція
1
Одним із компонентів екологічної стабільності є підтримка біорізноманіття
лісових ландшафтів шляхом поєднання заходів з охорони, менеджменту та
відновлення. Досягнення цієї цілі можна розглядати як головну умову сталого
розвитку в цілому. Впровадження принципу сталого лісового менеджменту
стимулювало появу багатьох критеріїв та індикаторів. Однак для досягнення
екологічної стабільності моніторинг відповідних показників повинен здійснюватися
у постійному порівнянні з цілями для того, щоб можна було оцінити теперішній
статус та тенденції розвитку ландшафтів. Показано, як традиційні засоби опису
лісових ресурсів повинні бути доповнені елементами моніторингу біорізноманіття на
різних просторових рівнях, включно з видами, умовами пробування та функціями.
Розглянуто приклади емпіричних нелінійних залежностей поміж
популяціями певних видів та рівнями антропогенних змін їхніх умов
місцепробування на різних просторових рівнях. З’ясовано, як це допомагає
сформулювати відповідні цілі. Отож, використовуючи результати моніторингу та
маючи відповідні цілі, можна оцінити статус певного критерію – такого, як
біорізноманіття. Наведено приклади оцінки за допомогою аналізу пробілів (gap
analysis) та моделі місцепробувань для стратегічного та тактичного планування
менеджменту для захисту, підтримки та відтворення різних елементів
біорізноманіття. Обговорено потребу постійного інформування менеджерів та
політиків щодо результатів оцінки.
Ключові слова: сталий лісовий менеджмент, біорізноманіття, ландшафти,
просторові рівні.
Стаття надійшла до редколегії 20.03.2004
Прийнята до друку 16.06.2004
Download