• Review Paper
  • Open access
  • Published: 14 January 2015

Climate change impacts and adaptation in forest management: a review

  • Rodney J. Keenan 1  

Annals of Forest Science volume  72 ,  pages 145–167 ( 2015 ) Cite this article

52k Accesses

351 Citations

45 Altmetric

Metrics details

Key message

Adaptation of forest management to climate change requires an understanding of the effects of climate on forests, industries and communities; prediction of how these effects might change over time; and incorporation of this knowledge into management decisions. This requires multiple forms of knowledge and new approaches to forest management decisions. Partnerships that integrate researchers from multiple disciplines with forest managers and local actors can build a shared understanding of future challenges and facilitate improved decision making in the face of climate change.

Climate change presents significant potential risks to forests and challenges for forest managers. Adaptation to climate change involves monitoring and anticipating change and undertaking actions to avoid the negative consequences and to take advantage of potential benefits of those changes.

This paper aimed to review recent research on climate change impacts and management options for adaptation to climate change and to identify key themes for researchers and for forest managers.

The study is based on a review of literature on climate change impacts on forests and adaptation options for forest management identified in the Web of Science database, focusing on papers and reports published between 1945 and 2013.

One thousand one hundred seventy-two papers were identified in the search, with the vast majority of papers published from 1986 to 2013. Seventy-six percent of papers involved assessment of climate change impacts or the sensitivity or vulnerability of forests to climate change and 11 % (130) considered adaptation. Important themes from the analysis included (i) predicting species and ecosystem responses to future climate, (ii) adaptation actions in forest management, (iii) new approaches and tools for decision making under uncertainty and stronger partnerships between researchers and practitioners and (iv) policy arrangements for adaptation in forest management.

Conclusions

Research to support adaptation to climate change is still heavily focused on assessing impacts and vulnerability. However, more refined impact assessments are not necessarily leading to better management decisions. Multi-disciplinary research approaches are emerging that integrate traditional forest ecosystem sciences with social, economic and behavioural sciences to improve decision making. Implementing adaptation options is best achieved by building a shared understanding of future challenges among different institutions, agencies, forest owners and stakeholders. Research-policy-practice partnerships that recognise local management needs and indigenous knowledge and integrate these with climate and ecosystem science can facilitate improved decision making.

1 Introduction

Anthropogenic climate change presents potential risks to forests and future challenges for forest managers. Responding to climate change, through both mitigation and adaptation, may represent a paradigm shift for forest managers and researchers (Schoene and Bernier 2012 ). Climate change is resulting in increasing air temperature and changing precipitation regimes, including changes to snowfall and to the timing, amount and inter-annual variability of rainfall (IPCC 2013 ). Forests are widespread, long-lived ecosystems that are both intensively and extensively managed. They are potentially sensitive to these longer term climatic changes, as are the societies and economies that depend on them (Bernier and Schöne 2009 ). Climate change increases the potential consequences of many existing challenges associated with environmental, social or economic change.

Whilst forest ecosystems are resilient and many species and ecosystems have adapted historically to changing conditions, future changes are potentially of such magnitudes or will occur at rates that are beyond the natural adaptive capacity of forest species or ecosystems, leading to local extinctions and the loss of important functions and services, including reduced forest carbon stocks and sequestration capacity (Seppälä et al. 2009 ).

Recent global warming has already caused many changes in forests (Lucier et al. 2009 ). Aspects of climate change may be positive for some tree species in some locations. Tree growth is observed to be increasing in some locations under longer growing seasons, warmer temperatures and increased levels of CO 2 . However, many projected future changes in climate and their indirect effects are likely to have negative consequences for forests. Observed shifts in vegetation distribution (Kelly and Goulden 2008 ; Lenoir et al. 2010 ) or increased tree mortality due to drought and heat in forests worldwide (Allen et al. 2010 ) may not be due to human-induced climate change but demonstrate the potential impacts of rapid climate change. These impacts may be aggravated by other human-induced environmental changes such as increases in low elevation ozone concentrations, nitrogenous pollutant deposition, the introduction of exotic insect pests and pathogens, habitat fragmentation and increased disturbances such as fire (Bernier and Schöne 2009 ). Other effects of climate change may also be important for forests. Sea level rise is already impacting on tidal freshwater forests (Doyle et al. 2010 ) and tidal saltwater forests (mangroves) are expanding landward in sub-tropical coastal reaches taking over freshwater marsh and forest zones (Di Nitto et al. 2014 ).

With projected future change, species ranges will expand or contract, the geographic location of ecological zones will shift, forest ecosystem productivity will change and ecosystems could reorganise following disturbances into ecological systems with no current analogue (Campbell et al. 2009 ; Fischlin et al. 2009 ). Forests types differ in their sensitivity to climatic change. Bernier and Schöne ( 2009 ) considered boreal, mountain, Mediterranean, mangrove and tropical moist forests most vulnerable to climate change. However, there has been recent debate about the vulnerability of tropical moist forests (Corlett 2011 ; Huntingford et al. 2013 ; Feeley et al. 2012 ), and temperate forests in areas subject to drier climates may be more at risk (Choat et al. 2012 ).

Adapting to these changing and uncertain future conditions can be considered from a number of perspectives (McEvoy et al. 2013 ). Policy and management might be directed at avoiding or reducing the impact of climate-related events, reducing vulnerability to future climatic conditions, managing a broader suite of climate ‘risks’ or increasing resilience and capacity in forest ecological and production systems to recover from climate ‘shocks’.

Adapting forest management to climate change involves monitoring and anticipating change and undertaking actions to avoid the negative consequences or take advantage of potential benefits of those changes (Levina and Tirpak 2006 ). Adopting the principles and practices of sustainable forest management (SFM) can provide a sound basis for addressing the challenges of climate change. However, Innes et al. ( 2009 ) pointed out that our failure to implement the multi-faceted components of sustainable forest management in many forests around the world is likely to limit capacity to adapt to climate change. Forest managers will need to plan at multiple spatial and temporal scales and adopt more adaptive and collaborative management approaches to meet future challenges.

Whilst forest managers are accustomed to thinking in long time scales—considering the long-term implications of their decisions and factoring in uncertainty and unknowns into management—many are now responding to much shorter term social or economic imperatives. Local forestry practices are often based on an implicit assumption that local climate conditions will remain constant (Guariguata et al. 2008 ). Other social and economic changes will also continue to drive changes in forest management (Ince et al. 2011 ). For example, a growing global population, rapid economic development and increased wealth are driving demand for food and fibre crops and forest conversion to agriculture in many developing countries (Gibbs et al. 2010 ). Climate change mitigation objectives are increasing demands for wood-based bioenergy and the use of wood in construction and industrial systems. Increasing urbanisation is changing the nature of social demands on forests, and decreasing rural populations is limiting the availability of labour and capacity for intensive forest management interventions.

Ecosystem-based adaptation is being promoted as having the potential to incorporate sustainable management, conservation and restoration of ecosystems into adaptation to climate change (IUCN 2008 ). This can be achieved more effectively by integrating ecosystem management and adaptation into national development policies through education and outreach to raise societal awareness about the value of ecosystem services (Vignola et al. 2009 ).

Kimmins ( 2002 ) invoked the term ‘future shock’, first coined by Toffler ( 1970 ) to describe the situation where societal expectations from forests were changing faster than the institutional capacity for change in forest management organisations. The pace of climate change is likely to intensify this phenomenon. Empirically based management based on traditional ‘evidence-based’ approaches therefore will potentially not develop quickly enough for development of effective future management options. How can managers consider rapid change and incorporate the prospect of very different, but uncertain, future climatic conditions into their management decisions? What types of tools are needed to improve decision making capacity?

This study aimed to review the literature on studies to support forest management in a changing climate. It builds on the major review of Seppala ( 2009 ), in particular Chapter 6 of that report by Innes et al. ( 2009 ).

The study involved a systematic assessment of the literature based on the database Web of Science (Thomson-Reuters 2014 ), an online scientific citation indexing service that provides the capacity to search multiple databases, allowing in-depth exploration of the literature within an academic or scientific discipline.

The following search terms were used in the titles of publications:

(forest* or tree* or (terrestrial and ecosystem)) and climat* and (adapt* or impact* or effect* or respons*) and

(forest* or tree*) and climat* and vulnerabilit* or sensitivit*)

The search was restricted to publications between 1945 and 2013. References related solely to climate change mitigation were excluded, as were references where the word ‘climate’ simply referred to a study in a particular climatic zone. This left a database of 1172 publications for analyses (a spreadsheet of the papers revealed in the search can be obtained from the author). References were classified into various types of studies and different regions, again based on the titles. Not all papers identified in the search are referenced. The selection of themes for discussion and papers for citation was a subjective one, based on scanning abstracts and results from relevant individual papers. The focus was important themes from key papers and literature from the last 5 years. The review includes additional papers not revealed in the search relating to these themes including selected papers from the literature in the year 2014.

Of the published papers relating to climate impacts or adaptation selected for analysis, the vast majority of papers were published from 1986 onwards. The earliest paper dated from 1949 (Gentilli 1949 ) analysing the effects of trees on climate, water and soil. Most studies prior to 1986 (and even some published later) focused on the effects of trees on local or wider regional climate (Lal and Cummings 1979 ; Otterman et al. 1984 ; Bonan et al. 1992 ), the implications of climate variability (Hansenbristow et al. 1988 ; Ettl and Peterson 1995 ; Chen et al. 1999 ), studies of tree and forest responses across climatic gradients (Grubb and Whitmore 1966 ; Bongers et al. 1999 ; Davidar et al. 2007 ) or responses to historical climate (Macdonald et al. 1993 ; Huntley 1990 ; Graumlich 1993 ).

One thousand twenty-six papers specifically addressed future climate change (rather than historical climate or gradient analysis). Of these, 88 % studied impacts, effects, vulnerability or responses to climate change in tree species, forests, forest ecosystems or the forest sector (Fig.  1 ). The first study analysing the potential impacts of future climate change on terrestrial ecosystems was published in 1985 (Emanuel et al. 1985 ) with other highly cited papers soon after (Pastor and Post 1988 ; Cannell et al. 1989 ).

Publication numbers by publication year for publications relating to climate change and forests from a search of the Web of Science database to the end of 2013 (1025 in total, 896 publications studied climate change impacts, responses or vulnerability, 129 studied adaptation)

Twelve percent of papers (129) considered adaptation options, including 10 papers on adaptation in the forest sector. The first papers to focus on adaptation in the context of climate change were in 1996 with a number of papers published in that year (Kienast et al. 1996 ; Kobak et al. 1996 ; Dixon et al. 1996 ). Publications were then relatively few each year until the late 2000s with numbers increasing to 11 in 2009, 22 in 2010 and 27 in 2011. Publications on adaptation dropped to 14 papers in 2013. The ratio of adaptation-related papers has increased more recently, with 19 % of total publications on adaptation in the last 5 years. Most papers considering adaptation since the early 2000s have related to the integration of adaptation and forest management (e.g. Lindner 2000 ; Spittlehouse 2005 ; Kellomaki et al. 2008 ; Guariguata 2009 ; Bolte et al. 2009 ; Keskitalo 2011 ; Keenan 2012 ; Temperli et al. 2012 ).

Analyses of the implications of climate change for the forest sector have focused heavily on North America: Canada (Ohlson et al. 2005 ; Van Damme 2008 ; Rayner et al. 2013 ; Johnston et al. 2012 ) and the USA (Joyce et al. 1995 ; Blate et al. 2009 ; Kerhoulas et al. 2013 ); and Europe (Karjalainen et al. 2003 ; von Detten and Faber 2013 ). There has been a stronger consideration in recent years of social, institutional and policy issues (Ogden and Innes 2007b ; Kalame et al. 2011 ; Nkem et al. 2010 ; Spies et al. 2010 ; Somorin et al. 2012 ) and the assessment of adaptive capacity in forest management organisations and in society more generally (Keskitalo 2008 ; Lindner et al. 2010 ; Bele et al. 2013a ).

Regionally, there have been relatively few published journal articles on impacts or adaptation in forests in the Southern Hemisphere (Hughes et al. 1996 ; Williams 2000 ; Pinkard et al. 2010 ; Gonzalez et al. 2011 ; Mok et al. 2012 ; Breed et al. 2013 ), although there have been more studies in the grey literature for Australian forests (Battaglia et al. 2009 ; Cockfield et al. 2011 ; Medlyn et al. 2011 ; Stephens et al. 2012 ). There have been some valuable analyses for the tropics (Guariguata et al. 2008 , 2012 ; Somorin et al. 2012 ; Feeley et al. 2012 ).

Analysis of the publications identified the following key themes: (i) predicting species and ecosystem responses to future climate, (ii) adaptation actions in forest management, (iii) new approaches and tools for decision making under uncertainty and stronger partnerships between researchers and practitioners and (iv) policy arrangements for adaptation in forest management. These are discussed in more detail below.

3.1 Predicting species and ecosystem responses to future climate

Forest managers have long used climatic information in a range of ways in planning and decision making. Climate information has been used extensively to define and map vegetation types and ecological zones and for modelling habitat distributions of vertebrates and invertebrates (Daubenmire 1978 ; Pojar et al. 1987 ; Thackway and Cresswell 1992 ), for species and provenance selection (Booth et al. 1988 ; Booth 1990 ) and seed zone identification (Johnson et al. 2004 ), for forest fire weather risk assessment and fire behaviour modelling (Carvalho et al. 2008 ), for modelling forest productivity (Battaglia et al. 2004 ) and analysing the dynamics of a range of ecological processes (Anderson 1991 ; Breymeyer and Melillo 1991 ). Predicting species responses to future climate change presents a different set of challenges, involving consideration of predictions of future climate that are often outside the historical range of variability of many species. These challenges are discussed in the next section.

3.1.1 Species responses to climate

Aitken et al. ( 2008 ) argued that there were three possible fates for forest tree populations in rapidly changing climatic conditions: persistence through spatial migration to track their ecological niches, persistence through adaptation to new conditions in current locations or the extirpation of the species. Predicting the potential fate of populations in these conditions requires the integration of knowledge across biological scales from individual genes to ecosystems, across spatial scales (for example, seed and pollen dispersal distances or breadth of species ranges) and across temporal scales from the phenology of annual developmental cycle traits to glacial and interglacial cycles.

Whilst there has been widespread use of climatic information to predict future distributions in species distribution models (SDMs, Pearson and Dawson 2003 ; Attorre et al. 2008 ; Wang et al. 2012 ; Ruiz-Labourdette et al. 2013 ), understanding of the range of climatic and non-climatic factors that will determine the future range of a particular species remains limited. Many now feel that SDMs are of limited value in adaptation decision making or species conservation strategies. Some of these limitations are summarised in Table  1 .

For example, models indicate significant shifts in patterns of tree species distribution over the next century but usually without any intrinsic consideration of the biological capacity of populations to move, internal population dynamics, the extent and role of local adaptation or the effects of climate and land use (Aitken et al. 2008 ; Thuiller et al. 2008 ). In a recent study, Dobrowski et al. ( 2013 ) found that the predicted speed of movement of species to match the predicted rate of climate change appears to be well beyond the historical rates of migration. Whilst modelled outputs suggest that migration rates of 1000 m per year or higher will be necessary to track changing habitat conditions (Malcolm et al. 2002 ), actual migration rates in response to past change are generally considered to have been less than 100 m per year. This was reinforced by model predictions that incorporate species dispersal characteristics for five tree species in the eastern USA indicated very low probabilities of dispersal beyond 10–20 km from current species boundaries by 2100 (Iverson et al. 2004 ). Corlett and Westcott ( 2013 ) also argued that plant movements are not realistically represented in models used to predict future vegetation or carbon-cycle feedbacks and that fragmentation of natural systems is likely to slow migration rates.

However, these estimates do not account for the role of humans in influencing tree species distributions, which they have done for thousands of years (Clark 2007 ), and managed translocation may be an option for conserving many tree species, but there are significant unresolved technical and social questions about implementing translocation at a larger scale (Corlett and Westcott 2013 ).

Most early SDMs relied primarily on temperature envelopes to model future distribution, but factors such as precipitation and soil moisture are potentially more limiting and more important in determining distribution patterns (Dobrowski et al. 2013 ). Aitken et al. ( 2008 ) found that the degree to which variation in precipitation explains phenotypic variation among populations is greater in general for populations from continental than from maritime climates and greater for lower latitude than higher latitude populations. However, precipitation alone is often not a good predictor of variation and there is often a strong interaction with temperature (Andalo et al. 2005 ). Heat to moisture index or aridity is probably more important in determining future distribution or productivity than precipitation alone (Aitken et al. 2008 ; Harper et al. 2009 ; Wang et al. 2012 ). Soil properties (depth, texture and organic matter content) have a major influence on plant-available water, but few SDMs incorporate these.

Future precipitation is proving more difficult to model than temperature, due to the complex effects of topography, and there are more widely varying estimates between global circulation models (GCMs) of future change in precipitation (IPCC 2013 ). As such, there is more uncertainty around the extent to which moisture stress will change with warming and the extent to which natural selection pressures will change as a result. Even without changes in precipitation, increased temperatures will increase the length of growing season and potential evapotranspiration (PET) resulting in more water use over the year and greater risk plant water shortage and drought death.

Changes in the intervals of extreme events (extreme heat, cold, precipitation, humidity, wind) may also matter more than changes in the mean. Current forecasting approaches that produce future climate averages may make it difficult to detect non-linear ecosystem dynamics, or threshold effects, that could trigger abrupt ecosystem change (Campbell et al. 2009 ). Zimmermann et al. ( 2009 ) found that predictions of spatial patterns of tree species in Switzerland were improved by incorporating measures of extremes in addition to means in SDMs.

The risks of future climate will also depend on the management goal. If the aim is simply to conserve genetic diversity, risks of extinction or reduction in genetic diversity may be overstated by SDMs because much of the genetic variation within tree species is found within rather than among their populations, and the extinction of a relatively large proportion of a population is generally likely to result in relatively little overall loss of genetic diversity (Hamrick 2004 ). Local habitat heterogeneity (elevation, slope aspect, moisture, etc.) can preserve adaptive genetic variation that, when recombined and exposed to selection in newly colonised habitats, can provide for local adaptation. The longevity of individual trees can also retard population extinction and allow individuals and populations to survive until habitat recovery or because animal and wind pollination can provide levels of pollen flow that are sufficient to counteract the effects of genetic drift in fragmented populations. Consequently, widespread species with large populations, high fecundity and higher levels of phenotypic plasticity are likely to persist and adapt and have an overall greater tolerance to changing climates than predicted by SDMs (Alberto et al. 2013 ).

Tree species distributions have always been dynamic, responding to changing environmental conditions, and populations are likely to be sub-optimal for their current environments (Namkoong 2001 ; Wu and Ying 2004 ). These lag effects are important in predicting species responses to climate change. In a modelling study of Scots pine and silver birch, Kuparinen et al. ( 2010 ) predicted that after 100 years of climate change, the genotypic growth period length of both species will lag more than 50 % behind the climatically determined optimum. This lag is reduced by increased mortality of established trees, whereas earlier maturation and higher dispersal ability had comparatively minor effects. Thuiller et al. ( 2008 ) suggest that mechanisms for incorporating these ‘trailing edge’ effects into SDMs are a major area of research potential.

Trees are also capable of long-distance gene flow, which can have both adaptive evolution benefits and disadvantages. Kremer et al. ( 2012 ) found that there may be greater positive effects of gene flow for adaptation but that the balance of positive to negative consequences of gene flow differs for leading edge, core and rear sections of forest distributions.

Epigenetics—heritable changes that are not caused by changes in genetic sequences but by differences in the way DNA methylation controls the degree of gene expression—is another complicating factor in determining evolutionary response to climate change (Brautigam et al. 2013 ). For example, a recent study in Norway spruce ( Picea abies ) showed that the temperature during embryo development can dramatically affect cold hardiness and bud phenology in the offspring. In some cases, the offspring’s phenotype varied by the equivalent of 6° of latitude from what was expected given the geographic origin of the parents. It remains uncertain whether these traits are persistent, both within an individual’s lifetime and in its offspring and subsequent generations (Aitken et al. 2008 ). It is suggested that analysis of the epigenetic processes in an ecological context, or ‘ecological epigenetics’, is set to transform our understanding of the way in which organisms function in the landscape. Increased understanding of these processes can inform efforts to manage and breed tree species to help them cope with environmental stresses (Brautigam et al. 2013 ). Others argue that whilst investigating this evolutionary capacity to adapt is important, understanding responses of species to their changing biotic community is imperative (Anderson et al. 2012 ) and ‘landscape genomics’ may offer a better approach for informing management of tree populations under climate change (Sork et al. 2013 ).

These recent results indicate the importance of accounting for evolutionary processes in forecasts of the future dynamics and productivity of forests. Species experiencing high mortality rates or populations that are subject to regular disturbances such as storms or fires might actually be the quickest to adapt to a warming climate.

Species life history characteristics are also not usually well represented in most climate-based distribution models. Important factors include age to sexual maturity, fecundity, seed dispersal, competition or chilling or dormancy requirements (Nitschke and Innes 2008b ).

Competitive relationships within and between species are likely to be altered by climate change. Most models also assume open site growth conditions, rather than those within a forest, where the growth environment will be quite different. However, increased disturbance associated with climate change may create stand reinitiation conditions more often than has occurred in the past, altering competitive interactions.

Process-based models of species range shifts and ecosystem change may capture more of the life history variables and competition effects that will be important in determining responses to climate change (Kimmins 2008 ; Nitschke and Innes 2008a , b ). These can provide the basis for a more robust assessment framework that integrates biological characteristics (e.g. shade tolerance and seedling establishment) and disturbance characteristics (e.g. insect pests, drought and fire topkill). Matthews et al. ( 2011 ) integrated these factors into a decision support system that communicates uncertainty inherent in GCM outputs, emissions scenarios and species responses. This demonstrated a greater diversity among species to adapt to climate change and provides a more practical assessment of future species projections.

In summary, whilst SDMs and other climate-based modelling approaches can provide a guide to potential species responses, the extent to which future climate conditions will result in major range shifts or extinction of species is unclear and the value of this approach in adaptation and decision making is limited. The evidence from genetic studies seems to suggest that many species are reasonably robust to potential future climate change. Those with a wide geographic range, large populations and high fecundity may suffer local population extinction but are likely to persist and adapt whilst suffering adaptational lag for a few generations. For example, Booth ( 2013 ) considered that many eucalyptus species, some of which are widely planted around the world, had a high adaptive capacity even though their natural ranges are quite small.

However, large contractions or shifts in distribution could have significant consequences for different forest values and species with small populations, fragmented ranges, low fecundity or suffering declines due to introduced insects or diseases may have a higher sensitivity and are at greater risk in a changing climate (Aitken et al. 2008 ).

3.1.2 Ecosystem responses to climate

Projecting the fate of forest ecosystems under a changing climate is more challenging than for species. It has been well understood for some time that species will respond individualistically to climate change, rather than moving in concert, and that this is likely to result in ‘novel’ ecosystems, or groups of species, that are not represented in current classifications (Davis 1986 ). Forecasts need to consider the importance of these new species interactions and the confounding effects of future human activities.

Climate change affects a wide range of ecosystem functions and processes (Table  2 ). These include direct effects of temperature and precipitation on physiological and reproductive processes such as photosynthesis, water use, flowering, fruiting and regeneration, growth and mortality and litter decomposition. Changes in these processes will have effects on species attributes such as wood density or foliar nutrient status. Indirect effects will be exhibited through changing fire and other climate-driven disturbances. These will ultimately have impacts on stand composition, habitat structure, timber supply capacity, soil erosion and water yield.

Most early studies of forest ecosystem responses to climate change were built around ecosystem process models at various scales (Graham et al. 1990 ; Running and Nemani 1991 ; Rastetter et al. 1991 ). A number of recent studies have investigated the effects of past and current climate change on forest processes, often with surprising effects (Groffman et al. 2012 ).

Observed forest growth has increased recently in a number of regions, for example over the last 100 years in Europe (Pretzsch et al. 2014 ; Kint et al. 2012 ), and for more recent observations in Amazon forests (Phillips et al. 2008 ). In a major review, Boisvenue and Running ( 2006 ) found that at finer spatial scales, a trend is difficult to decipher, but globally, based on both satellite and ground-based data, climatic changes seemed to have a generally positive impact on forest productivity when water was not limiting. However, there can be a strong difference between species, complicating ecosystem level assessments (Michelot et al. 2012 ), and there are areas with little observed change (Schwartz et al. 2013 ). Generally, there are significant challenges in detecting the response of forests to climate change. For example, in the tropics, the lack of historical context, long-term growth records and access to data are real barriers (Clark 2007 ) and temperate regions also have challenges, even with well-designed, long-term experiments (Leites et al. 2012 ).

Projections of net primary productivity (NPP) under climate change indicate that there is likely to be a high level of regional variation (Zhao et al. 2013 ). Using a process model and climate scenario projections, Peters et al. ( 2013 ) predicted that average regional productivity in forests in the Great Lakes region of North America could increase from 67 to 142 %, runoff could potentially increase from 2 to 22 % and net N mineralization from 10 to 12 %. Increased productivity was almost entirely driven by potential CO 2 fertilization effects, rather than by increased temperature or changing precipitation. Productivity in these forests could shift from temperature limited to water limited by the end of the century. Reyer et al. ( 2014 ) also found strong regional differences in future NPP in European forests, with potential growth increases in the north but reduced growth in southern Europe, where forests are likely to be more water limited in the future. Again, assumptions about the impact of increasing CO 2 were a significant factor in this study.

In a different type of study using analysis of over 2400 long-term measurement plots, Bowman et al. ( 2014 ) found that there was a peaked response to temperature in temperate and sub-tropical eucalypt forests, with maximum growth occurring at a mean annual temperature of 11 °C and maximum temperature of the warmest month of 25–27 °C. Lower temperatures directly constrain growth, whilst high temperatures primarily reduced growth by reducing water availability but they also appeared to exert a direct negative effect. Overall, the productivity of Australia’s temperate eucalypt forests could decline substantially as the climate warms, given that 87 % of these forests currently experience a mean annual temperature above the ‘optimal’ temperature.

Incorporating the effects of rising CO 2 in models of future tree growth continues to be a major challenge. The sensitivity of projected productivity to assumptions regarding increased CO 2 was high in modelling studies of climate change impacts in commercial timber plantations in the Southern Hemisphere (Kirschbaum et al. 2012 ; Battaglia et al. 2009 ), and a recent analysis indicated a general convergence of different model predictions for future tree species distribution in Europe, with most of the difference between models due to the way in which this effect is incorporated (Cheaib et al. 2012 ). Increased CO 2 has been shown to increase the water-use efficiency of trees, but this is unlikely to entirely offset the effects of increased water stress on tree growth in drying climates (Leuzinger et al. 2011 ; Booth 2013 ). In general, despite studies extending over decades and improved understanding of biochemical processes (Franks et al. 2013 ), the impacts of increased CO 2 on tree and stand growth are still unresolved (Kallarackal and Roby 2012 ).

Integrating process model outputs with spatially explicit landscape models can improve understanding and projection of responses and landscape planning and this could provide for simulations of changes in ecological processes (e.g. tree growth, succession, disturbance cycles, dispersal) with other human-induced changes to landscapes (Campbell et al. 2009 ).

Investigation of current species responses to changing climate conditions may also guide improved prediction of patterns of future change in ecosystem distribution. For example, Allen et al. ( 2010 ) suggest that spatially explicit documentation of environmental conditions in areas of forest die-off is necessary to link mortality to causal climate drivers, including precipitation, temperature and vapour pressure deficit. Better prediction of climate responses will also require improved knowledge of belowground processes and soil moisture conditions. Assessments of future productivity will depend on accurate measurements of rates (net ecosystem exchange and NPP), changes in ecosystem level storage (net ecosystem production) and quantification of disturbances effects to determine net biome production (Boisvenue and Running 2006 ).

Hydrological conditions, runoff and stream flow are of critical importance for humans and aquatic organisms, and many studies have focused on the implications of climate change for these ecosystem processes. However, most of these have been undertaken at small catchment scale (Mahat and Anderson 2013 ; Neukum and Azzam 2012 ; Zhou et al. 2011 ) with few basin-scale assessments (van Dijk and Keenan 2007 ). However, the effects of climate and forest cover change on hydrology are complicated. Loss of tree cover may increase stream flow but can also increase evaporation and water loss (Guardiola-Claramonte et al. 2011 ). The extent of increasing wildfire will also be a major factor determining hydrological responses to climate change (Versini et al. 2013 ; Feikema et al. 2013 ).

Changing forest composition will also affect the habitat of vertebrate and invertebrate species. The implications of climate change for biodiversity conservation have been subject to extensive analysis (Garcia et al. 2014 ; Vihervaara et al. 2013 ; Schaich and Milad 2013 ; Clark et al. 2011 ; Heller and Zavaleta 2009 ; Miles et al. 2004 ). An integrated analytical approach, considering both impacts on species and habitat is important. For example, in a study of climate change impacts on bird habitat in the north-eastern USA, the combination of changes in tree distribution and habitat for birds resulted in significant impacts for 60 % of the species. However, the strong association of birds with certain vegetation tempers their response to climate change because localised areas of suitable habitat may persist even after the redistribution of tree species (Matthews et al. 2011 ).

Understanding thresholds in changing climate conditions that are likely to result in a switch to a different ecosystem state, and the mechanisms that underlie ecosystem responses, will be critical for forest managers (Campbell et al. 2009 ). Identifying these thresholds of change is challenging. Detailed process-based ecosystem research that identifies and studies critical species interactions and feedback loops, coupled with scenario modelling of future conditions, could provide valuable insights (Kimmins et al. 1999 , 2008 ; Walker and Meyers 2004 ). Also, rather than pushing systems across thresholds into alternative states, climate change may create a stepwise progression to unknown transitional states that track changing climate conditions, requiring a more graduated approach in management decisions (Lin and Petersen 2013 ).

Ultimately, management decisions may not be driven by whether we can determine future thresholds of change, but by observing the stressors that determine physiological limits of species distributions. These thresholds will depend on species physiology and local site conditions, with recent research demonstrating already observed ecosystem responses to climate change, including die-back of some species (Allen et al. 2010 ; Rigling et al. 2013 ).

3.1.3 Fire, pests, invasive species and disturbance risks

Many of the impacts of a changing future climate are likely to be felt through changing disturbance regimes, in particular fire. Forest fire weather risk and fire behaviour prediction have been two areas where there has been strong historical interaction between climate science and forest management and where we may see major tipping points driving change in ecosystem composition (Adams 2013 ). Fire weather is fundamentally under the control of large-scale climate conditions with antecedent moisture anomalies and large-scale atmospheric circulation patterns, further exacerbated by configuration of local winds, driving fire weather (Brotak and Reifsnyder 1977 ; Westerling et al. 2002 , 2006 ). It is therefore important to improve understanding of both short- and long-term atmospheric conditions in determining meteorological fire risk (Amraoui et al. 2013 ).

Increased fuel loads and changes to forest structure due to long periods of fire exclusion and suppression are increasing fire intensity and limiting capacity to control fires under severe conditions (Williams 2004 , 2013 ). Increasing urbanisation is increasing the interface between urban populations and forests in high fire risk regions, resulting in greater impacts of wildfire on human populations, infrastructure and assets (Williams 2004 ). Deforestation and burning of debris and other types of human activities are also introducing fire in areas where it was historically relatively rare (Tacconi et al. 2007 ).

In a recent study, Chuvieco et al. ( 2014 ) assessed ecosystem vulnerability to fire using an index based on ecological richness and fragility, provision of ecosystem services and value of houses in the wildland–urban interface. The most vulnerable areas were found to be the rainforests of the Amazon Basin, Central Africa and Southeast Asia; the temperate forest of Europe, South America and north-east America; and the ecological corridors of Central America and Southeast Asia.

In general, fire management policies in many parts of the world will need to cope with longer and more severe fire seasons, increasing fire frequency, and larger areas exposed to fire risk. This will especially be the case in the Mediterranean region of Europe (Kolström et al. 2011 ) and other fire-prone parts of the world such as South Eastern Australia (Hennessy et al. 2005 ). This will require improved approaches to fire weather modelling and behaviour prediction that integrate a more sophisticated understanding of the climate system with local knowledge of topography, vegetation and wind patterns. It will also require the development of fire management capacity where it had previously not been necessary. Increased fire weather severity could push current suppression capacity beyond a tipping point, resulting in a substantial increase in large fires (de Groot et al. 2013 ; Liu et al. 2010 ) and increased investment in resources and management efforts for disaster prevention and recovery.

Biotic factors may be more important than direct climate effects on tree populations in a changing climate. For example, insects and diseases have much shorter generation length and are able to adapt to new climatic conditions more rapidly than trees. However, if insects move more rapidly to a new environment whilst tree species lag, some parts of the tree population may be impacted less in the future (Regniere 2009 ).

The interaction of pests, diseases and fire will also be important. For example, this interaction will potentially determine the vulnerability of western white pine ( Pinus monticola ) ecosystems in Montana in the USA. Loehman et al. ( 2011 ) found that warmer temperatures will favour western white pine over existing climax and shade tolerant species, mainly because warmer conditions will lead to increased frequency and extent of wildfires that facilitates regeneration of this species.

3.2 Adaptation actions in forest management

The large majority of published studies relating to forests and climate change have been on vulnerability and impacts. These have increased understanding of the various relationships between forest ecosystems and climate and improved capacity to predict and assess ecosystem responses. However, managers need greater guidance in anticipating and responding to potential impacts of climate change and methods to determine the efficiency and efficacy of different management responses because they are generally not responding sufficiently to potential climate risks.

3.2.1 Adaptation actions at different management levels

A number of recent reviews have described adaptation actions and their potential application in different forest ecosystems being managed for different types of goods or services (Bernier and Schöne 2009 ; Innes et al. 2009 ; Lindner et al. 2010 ; Kolström et al. 2011 ), and adaptation guides and manuals have been developed (Peterson et al. 2011 ; Stephens et al. 2012 ) for different types of forest and jurisdictions. Adaptation actions can be primarily aimed at reducing vulnerability to increasing threats or shocks from natural disasters or extreme events, or increasing resilience and capacity to respond to progressive change or climate extremes. Adaptation actions can be reactive to changing conditions or planned interventions that anticipate future change. They may involve incremental changes to existing management systems or longer term transformational changes (Stafford Smith et al. 2011 ). Adaptation actions can also be applied at the stand level or at ownership, estate or national scales (Keskitalo 2011 ).

Recent research at the stand level in forests in the SE USA showed that forest thinning, often recommended in systems that are likely to experience increased temperature and decreased precipitation as a result of climate change, will need to be more aggressive than traditionally practised to stimulate growth of large residual trees, improve drought resistance and provide greater resilience to future climate-related stress (Kerhoulas et al. 2013 ).

An analysis of three multi-aged stand-level options in Nova Scotia, Canada, Steenberg et al. ( 2011 ) found that leaving sexually immature trees to build stand complexity had the most benefit for timber supply but was least effective in promoting resistance to climate change at the prescribed harvest intensity. Varying the species composition of harvested trees proved the most effective treatment for maximising forest age and old-growth area and for promoting stands composed of climatically suited target species. The combination of all three treatments resulted in an adequate representation of target species and old forest without overly diminishing the timber supply and was considered most effective in minimising the trade-offs between management values and objectives.

An estate level analysis of Austrian Federal Forests indicated that management to promote mixed stands of species that are likely to be well adapted to emerging environmental conditions, silvicultural techniques fostering complexity and increased management intensity might successfully reduce vulnerability, with the timing of adaptation measures important to sustain supply of forest goods and services (Seidl et al. 2011 ).

Whilst researchers are analysing different management options, the extent to which they are being implemented in practice is generally limited. For example, in four regions in Germany, strategies for adapting forest management to climate change are in the early stages of development or simply supplement existing strategies relating to general risk reduction or to introduce more ‘nature-orientated’ forest management (Milad et al. 2013 ). Guariguata et al. ( 2012 ) found that forest managers across the tropics perceived that natural and planted forests are at risk from climate change but were ambivalent about the value of investing in adaptation measures, with climate-related threats to forests ranked below others such as clearing for commercial agriculture and unplanned logging.

Community-based management approaches are often argued to be the most successful approach for adaptation. An analysis of 38 community forestry organisations in British Columbia found that 45 % were researching adaptation and 32 % were integrating adaptation techniques into their work (Furness and Nelson 2012 ). Whilst these community forest managers appreciated support and advice from government for adaptation, balancing this advice with autonomy for communities to make their own decisions was considered challenging.

In a study of communities impacted by drought in the forest zone of Cameroon, Bele et al. ( 2013b ) identified adaptive strategies such as community-created firebreaks to protect their forests and farms from forest fires, the culture of maize and other vegetables in dried swamps, diversifying income activities or changing food regimes. However, these coping strategies were considered to be incommensurate with the rate and magnitude of change being experienced and therefore no longer seen as useful. Some adaptive actions, whilst effective, were resource inefficient and potentially translate pressure from one sector to another or generated other secondary effects that made them undesirable.

3.2.2 Integrating adaptation and mitigation

In considering responses to climate change, forest managers will generally be looking for solutions that address both mitigation objectives and adaptation. To maintain or increase forest carbon stocks over the long term, the two are obviously inextricably linked (Innes et al. 2009 ). Whilst there are potentially strong synergies, Locatelli et al. ( 2011 ) identified potential trade-offs between actions to address mitigation and the provision of local ecosystem services and those for adaptation. They argued that mitigation projects can facilitate or hinder the adaptation of local people to climate change, whereas adaptation projects can affect ecosystems and their potential to sequester carbon.

Broadly, there has been little integration to date of mitigation and adaptation objectives in climate policy. For example, there is little connection between policies supporting the reducing emissions from deforestation and forest degradation plus (REDD+) initiatives and adaptation. Integrating adaptation into REDD+ can advance climate change mitigation goals and objectives for sustainable forest management (Long 2013 ). Kant and Wu ( 2012 ) considered that adaptation actions in tropical forests (protection against fire and disease, ensuring adequate regeneration and protecting against coastal impacts and desertification) will improve future forest resilience and have significant climate change mitigation value.

3.2.3 Sector-level adaptation

Analyses of climate change impacts and vulnerability at the sector level have been undertaken for some time (Lindner et al. 2002 ; Johnston and Williamson 2007 ; Joyce 2007 ). However, it has recently been argued (Wellstead et al. 2014 ) that these assessments, which focus on macro system-level variables and relationships, fail to account for the multi-level or polycentric nature of governance and the possibility that policy processes may result in the non-performance of critical tasks required for adaptation.

Joyce et al. ( 2009 ) considered that a toolbox of management options for the US National Forests would include the following: practices focused on reducing future climate change effects by building resistance and resilience into current ecosystems and on managing for change by enabling plants, animals and ecosystems to adapt to climate change. Sample et al. ( 2014 ) demonstrated the utility of this approach in a coniferous forest management unit in northwestern USA. It provided an effective means for guiding management decisions and an empirical basis for setting budgetary and management priorities. In general, more widespread implementation of already known practices that reduce the impact of existing stressors represents an important ‘no regrets’ strategy.

Johnston and Hesseln ( 2012 ) found that barriers to implementing adaptation across forest sector managers in Canada included inflexible tenure arrangements and regulatory environments which do not support innovation. Echoing calls for wider implementation of SFM as a key adaptation strategy (Innes et al. 2009 ), they argued that forest certification systems, participating in the Canadian model forest programme, and adopting criteria and indicators of SFM can support sectoral level adaptation.

Decentralised management approaches are considered to be a more appropriate governance arrangement for forest management, but Rayner et al. ( 2013 ) argued that a decentralised forest policy sector in Canada has resulted in limitations where policy, such as adaptation, requires a coherent national response. Climate change adaptation has led to an expansion of departmental mandates that is not being addressed by better coordination of the available policy capacity. Relevant federal agencies are not well represented in information networks, and forest policy workers report lower levels of internal and external networking than workers in related policy subsectors.

Economic diversification can be a valuable strategy to improve resilience to climate-related shocks. This can take a range of forms: developing new industries or different types of forest-based industries based on different goods or services. For the timber sector, the value of diversification as a risk management strategy for communities is open to question. Ince et al. ( 2011 ) pointed out that the forest sector operates in an international market and is susceptible to changes in the structure of this market. In the US forest sector, globalization has accelerated structural change, favouring larger and more capital-intensive enterprises and altering historical patterns of resource use. They suggest that future markets for timber will be driven by developments in these larger scale enterprises and may not lead to expansion of opportunities for smaller scale forest enterprises because development of niche markets or customised products is likely to be pursued aggressively by larger globally oriented enterprises to develop branding, product identity and product value. How to best diversify for adaptation therefore remains an open question.

Consequently, whilst policies that support economic diversification will be important, this may involve diversification well beyond traditional sectors. For example, in areas where there is a high probability that forests will be lost in favour of other ecosystems, such as grasslands, managers should recognise early on that their efforts and resources may best be focused outside forests (Innes et al. 2009 ). These adjustments will involve taking into account the perceptions of climate risk by various stakeholders, including individuals, communities, governments, private institutions and organisations (Adger et al. 2007 ). Vulnerability assessments and adaptation measures also need to be developed in a framework that takes into account the vulnerabilities and actions in other sectors that are linked to the forest sector, such as food, energy, health and water (Sonwa et al. 2012 ).

3.3 New approaches to decision making

Climate change presents new challenges for forest managers. Change is likely to happen faster than traditional, empirical approaches can provide evidence to support changes in management. Uncertainties in a range of aspects of future climate may also not be reduced through investment in research. Given that management for activities such as timber production can no longer be based solely on empirically derived growth and yield trajectories and management plans must incorporate uncertainty and the increased probability of extreme events, what types of tools are available to support these approaches? This section presents key points from the literature on decision making under uncertainty, adaptive management and resilience as a guide to future decision making in forest management.

3.3.1 Decision making under uncertainty

The future conditions for forest managers are subject to a high degree of uncertainty, and the future prospects for reducing these large uncertainties are limited. There is uncertainty regarding the trajectory of future increases in atmospheric greenhouse gases, what kind of effects these might have on the climate system and the effects of climatic changes on ecological and social systems and their capacity to adapt (see Fig.  2 ) (Wilby and Dessai 2010 ).

The cascade of uncertainty (Wilby and Dessai 2010 )

Consequently, many forest managers consider that the future situation is too uncertain to support long-term and potentially costly decisions that may be difficult to reverse. Dessai and Hulme ( 2004 ) argued that uncertainty per se should not be a reason for inaction. However, the critical issue for managers is deciding the types of actions to take and the timing and conditions under which they should be taken (Ogden and Innes 2007a ). A more reactive ‘wait and see’ approach (or ‘purposeful procrastination’) might be justified if uncertainty or costs are high relative to the expected impacts and risks, or if it is cheaper to implement interventions by waiting until after a significant disturbance (e.g. replanting an area with more fire- or drought-resistant tree species after a wildfire or drought-induced insect outbreak).

Effective adaptation requires setting clear objectives. Managers and policy makers need to decide whether they are trying to facilitate ecosystem adaptation through changing species composition or forest structure or trying to engineer resistance to change through proactive management strategies (Joyce et al. 2008 ). Establishing objectives often depends on the integration of the preferences of different stakeholders (Prato 2008 ), but changing social preferences presents another source of potential uncertainty.

Risk assessment and management provide a foundation for decision making in considering climate change in natural resource management. This approach provides both a qualitative and quantitative framework for evaluating climate change effects and adaptation options. Incorporating risk management approaches into forest management plans can provide a basis for managers to continue to provide forest conditions that meet a range of important values (Day and Perez 2013 ).

However, risk approaches generally requiring assigning probabilities to future events. In a comprehensive review, Yousefpour et al. ( 2011 ) identified a growing body of research literature on decision making under uncertainty, much of which has focused on price uncertainty and variation in timber production but is extending to multiple forest management objectives and other types of risk. They argue that we are actually in a stochastic transition from one known stable (but variable) climate state to a new but largely unknown and likely more rapidly changing set of future conditions.

Decision makers themselves may also not be the rational actors assumed by these models, with their decisions taken according to quite different assumptions, preferences and beliefs (Ananda and Herath 2009 ; Couture and Reynaud 2008 ). Therefore, the communication approach will be important in determining whether the information is acted on. In a recent study, Yousefpour et al. ( 2014 ) considered that the speed with which decision makers will form firm beliefs about future climate depends on the divergence among climate trajectories, the speed of change and short-term climate variability. Using a Bayesian modelling approach, they found that if a large change in climate occurs, the value of investing in knowledge and taking an adaptive approach would be positive and higher than a non-adaptive approach. In communicating about uncertainty, it may be better to focus discussion on the varying time in the future when things will happen, rather than on whether they will happen at all (Lindner et al. 2014 ).

Increased investment in climate science and projections or species distribution modelling may not necessarily decrease uncertainty in climate projections or impacts. Climate models are best viewed as heuristic tools rather than as accurate forecasts of the future (Innes et al. 2009 ). Trajectories of change in many other drivers of forest management (social, political or economic) are also highly uncertain (Keskitalo 2008 ) and the effects of these on the projected performance of management can be the same order of magnitude, requiring an integrated social-ecological perspective to adaptation (Seidl and Lexer 2013 ).

In a more ‘decision-centred’ approach, plausible scenarios of the potential range of future conditions are required. These can be derived from climate models but do not need to be accurate and precise ‘predictions’ of future climate states (Wilby and Dessai 2010 ). To support this type of approach, research needs to focus on improved understanding of tree and ecosystem responses and identifying those aspects of climate to which different forest types are most sensitive.

Devising strategies that are able to meet management objectives under a range of future scenarios is likely to be the most robust approach, recognising that these strategies are unlikely to be optimal under all future conditions. In some cases, the effect of different scenarios on forest growth may not be that great and differences in the present value of different management options are relatively small. For example, Eriksson et al. ( 2011 ) found that there was limited benefit in attempting to optimise management plans in accordance with future temperature scenarios.

Integration of climate change science and adaptation in forest management planning is considered important for building resilience in forest social and ecological systems (Keskitalo 2011 ; D’Amato et al. 2011 ; Chmura et al. 2011 ; Parks and Bernier 2010 ; Lindner et al. 2014 ). Forest restoration is becoming a more prominent aspect of forest management in many parts of the world and restoration approaches will also need to integrate understanding of future climate change to be successful (Stanturf et al. 2014 ).

3.3.2 Adaptive management, resilience and decisions

Adaptive management provides a mechanism to move forward when faced with future uncertainty (Innes et al. 2009 ). It can be viewed as a systematic process for continually improving management policies and practices by monitoring and then learning from the outcomes of operational programmes as a basis for incorporating adaptation actions into forest management. Whilst many management initiatives purport to implement these principles, they often lack essential characteristics of the approach (Innes et al. 2009 ).

However, effective adaptation to changing climate cannot simply involve adaptive management as it is currently understood. The pace of climate change is not likely to allow for the use of management as a tool to learn about the system by implementing methodologies to test hypotheses concerning known uncertainties (Holling 1978 ). Future climatic conditions may result in system states and dynamics that have never previously existed (Stainforth et al. 2007 ), so observation of past experience may be a poor guide for future action. Management will need to be more ‘forward-looking’, considering the range of possible future conditions and planning actions that consider that full range.

How does this translate into the practical guidance forest managers are seeking on how to adapt their current practices and, if necessary, their goals (Blate et al. 2009 )? Managers will need to consider trade-offs between different objectives under different conditions. For example, Seidl et al. ( 2011 ) showed that, to keep climate vulnerability in an Austrian forest low, Norway spruce will have to be replaced almost entirely by better adapted species. However, indicator weights that favoured timber production over C storage or biodiversity exerted a strong influence on the results. Wider social implications of imposing such drastic changes in forest landscapes will also deserve stronger consideration in decision making.

Ecosystem management will need to be reframed to accommodate the risks of a changing climate. Adaptive strategies, even without specific information on the future climate conditions of a target ecosystem, would enhance social and ecological resilience to address the uncertainties due to changing climate (Mori et al. 2013 ). These are likely to be more subject to change over the short to medium term, in response to more rapidly changing conditions.

Analysis of ecosystem resilience can provide a framework for these assessments. Resilience can be defined as ‘the capacity of ecosystems to absorb disturbance and reorganise so as to retain essentially the same function, structure and feedbacks – to have the same identity’ (Walker and Salt 2012 ). It is a function of the capacity of an ecosystem to resist change, the extent and pace of change and the ability of an ecosystem to reorganise following disturbance. The concept of resilience holds promise for informing future forest management, but Rist and Moen ( 2013 ) argue that its contributions are, so far, largely conceptual and offer more in terms of being a problem-framing approach than analytical or practical tools. There may also be trade-offs involved with focusing on resilience through retention of current species composition or using a more adaptation-oriented management approach after disturbances (Buma and Wessman 2013 ). Complexity theory and concepts can provide an appropriate framework for managing resilience (Messier et al. 2013 ).

Management decisions will ultimately depend on the costs and benefits of different options, but there are few examples of decision making frameworks that compare the costs of future impacts with the costs of different actions and the efficacy of those actions in reducing impacts. Ogden and Innes ( 2009 ) used a structured decision making process to identify and assess 24 adaptation options that managers considered important to achieve their regional goals and objectives of sustainable forest management in light of climate change. In the analysis of options for biodiversity conservation, Wintle et al. ( 2011 ) found that the amount of funding available for adaptation was a critical factor in deciding options aimed at minimising species extinctions in the mega-diverse fynbos biome of South Africa. When the available budget is small, fire management was the best strategy. If the budget is increased to an intermediate level, the marginal returns from more fire management were limited and the best strategy was added habitat protection. Above another budget threshold, increased investment should go into more fire management. By integrating ecological predictions in an economic decision framework, they found that making the choice of how much to invest is as important as determining what actions to take.

3.3.3 Adaptation as a social learning process

Whilst adaptation has been defined as ‘adjustment in natural or human systems in response to actual or expected climatic stimuli or their effects’ (Levina and Tirpak 2006 ), adaptation is essentially about meeting future human needs (Spittlehouse and Stewart 2003 ; Hahn and Knoke 2010 ). Consequently, it is inherently a social process. Forest landscapes are social-ecological systems that involve both nature and society (Innes et al. 2009 ), and resolving trade-offs between different management objectives to meet the different needs in society is an important element of sustainable forest management. As Kolström et al. ( 2011 ) pointed out, some proposed adaptation measures may change the balance between current objectives and stakeholder interests, and it will be important to consider the relative balance of different measures at the stand, management unit and landscape scales.

Those investigating adaptive management also recognise that it goes beyond the focus on scientific methods, statistical designs or analytical rigour favoured by its early proponents and that there is now an expectation of much greater stakeholder involvement, with the concept being renamed by some as adaptive, collaborative management (Innes et al. 2009 ). SFM and adaptation are as much about those who inhabit, work in or utilise forests as it is about managing the forest ecosystems themselves (White et al. 2010 ; Pramova et al. 2012 ; Fischer et al. 2013 ).

The choice of adaptation options will thus likely be relatively complex, even in cases where information and policy have been developed, and communication measures for forest management have been well formulated. Making such choices may require considerable knowledge, competence and commitment for implementation at the local level (Keskitalo 2011 ). Effective adaptation will require much greater cooperation between stakeholders, more flexibility for management actions and commitment of time to develop the social license for action in the absence of conclusive evidence or understanding. This will require venues for sharing perspectives on the nature of the problem (Fig.  3 ).

Adaptation as a social learning process. There is a need to provide situations to share different viewpoints on the nature of the problem as a basis for developing shared solutions (image source: John Rowley, http://ch301.cm.utexas.edu/learn/ )

3.3.4 Local and indigenous knowledge

The promotion of community-based forest management may increase local adaptive capacity by putting decisions in the hands of those people who first feel the effects of climate change (Gyampoh et al. 2009 ). In this context, local knowledge systems based on long-term observation and experience are likely to be of increasing importance in decision making. Adaptation strategies can benefit from combining scientific and indigenous knowledge, especially in developing countries (Gyampoh et al. 2009 ), with the translation of local forest knowledge into the language of formal forest science being considered an important step towards adaptation (Roberts et al. 2009 ). However, conservation and natural resource managers in government agencies have often discounted traditional local management systems (Scott 2005 ), although Spathelf et al. ( 2014 ) provided a useful approach for capturing local expert knowledge. Linking this type of knowledge with broader scientific understanding of ecosystem functioning and the global climate system will be a major challenge, requiring consideration of both technical and cultural issues (Caverley 2013 ), including intellectual property concerns of indigenous people (Lynch et al. 2010 ).

3.4 Policy arrangements for adaptation

Increasingly, many are arguing that effectively responding to climate change will require polycentric and multi-level governance arrangements (Peel et al. 2012 ). However, Nilsson et al. ( 2012 ) found that institutionalising of knowledge and knowledge exchange regarding climate change adaptation in Sweden was weak and that improved mechanisms are required for feedback from the local to the national level. Recent studies have described stronger relationships between scientific research and forest management to assess trade-offs and synergies, for participatory decision making and for shared learning (Blate et al. 2009 ; Littell et al. 2012 ; Klenk et al. 2011 ).

Many papers emphasised the need for greater flexibility in the policies, cultures and structures of forest management organisations (Brown 2009 ; von Detten and Faber 2013 ; Rayner et al. 2013 ). Because no single community or agency can prepare on their own for future impacts, inter-sectoral policy coordination will be required to ensure that policy developments in related policy sectors are not contradictory or counterproductive. Greater integration of information, knowledge and experience and collaborative projects involving scientists, practitioners and policy makers from multiple policy communities could increase focus on resilience, identify regions of large-scale vulnerability and provide a more rigorous framework for the analysis of vulnerability and adaptation actions (Thomalla et al. 2006 ).

There is also likely to be a greater need for cross-border implementation of different forest management options, requiring greater coordination between nation states and sub-national governments (Keenan 2012 ). Policy is the product of both ‘top-down’ and ‘bottom-up’ processes and these might sometimes be in conflict. Simply having ‘good policy’ in place is unlikely to be sufficient, as a great deal of what takes place at ‘street level’ is not determined by formal aims of central policy (Urwin and Jordan 2008 ). Having the right policies can send a strong political signal that adaptation needs to be considered seriously but flexibility in policy systems will be required to facilitate adaptive planning.

4 Discussion and conclusions

This broad survey of the literature indicated that, whilst there has been considerable development in research and thinking about adaptation in forest management over the last 10 years, research is still strongly focused on assessment of future impacts, responses and vulnerability of species and ecosystems (and in some cases communities and forest industries) to climate change. There has been some movement from a static view of climate based on long-term averages to a more detailed understanding of the drivers of different climate systems and how these affect the factors of greatest influence on different forest ecosystems processes, such as variability and extremes in temperature or precipitation or fire disturbance. For example, Guan et al. ( 2012 ) demonstrated that quasi-periodic climate variation on an inter-annual (ENSO) to inter-decadal (PDO) time scale can significantly influence tree growth and should be taken into account when assessing the impact of climate changes on forest productivity.

Adaptation is, in essence, about making good decisions for the future, taking into account the implications of climate change. It involves recognising and understanding potential future climate impacts and planning and managing for their consequences, whilst also considering the broader social, economic or other environmental changes that may impact on us, individually or collectively. To effectively provide a role in mitigation, delivering associated ecosystem services and benefits in poverty reduction (Eliasch 2008 ) forest management will have to adapt to a changing and highly variable climate. In achieving this, the roles and responsibilities of different levels of government, the private sector and different parts of the community are still being defined.

The broader literature emphasises that adaptation is a continuous process, involving a process of ‘adapting well’ to continuously changing conditions (Tompkins et al. 2010 ). This requires organisational learning based on past experience, new knowledge and a comprehensive analysis of future options. This can take place through ‘learning by doing’ or through a process of search and planned modification of routines (Berkhout et al. 2006 ). However, interpreting climate signals is not easy for organisations, the evidence of change is ambiguous and the stimuli are not often experienced directly within the organisation. For example, many forest managers in Australia currently feel little need to change practices to adapt to climate change, given both weak policy signals and limited perceived immediate evidence of increasing climate impacts (Cockfield et al. 2011 ). To explain and predict adaptation to climate change, the combination of personal experience and beliefs must be considered (Blennow et al. 2012 ). ‘Climate smart’ forest management frameworks can provide an improved basis for managing forested landscapes and maintaining ecosystem health and vitality based on an understanding of landscape vulnerability to future climatic change (Fig. 4 ) (Nitschke and Innes 2008a ).

Components of climate smart forest management (after Nitschke and Innes 2008a , b )

Many are now asking, do we really need more research to start adapting forest management to climate change? Whilst adaptation is often considered ‘knowledge deficit’ problem—where scientists provide more information and forest managers will automatically make better decisions—the reality is that the way in which this information is presented and how it is interpreted and received, will play major roles in determining potential responses. Successful adaptation will require dissemination of knowledge of potential climate impacts and suitable adaptation measures to decision makers at both practice and policy levels (Kolström et al. 2011 ) but it needs to go well beyond that.

Adaptation is, above all, a social learning process. It requires an understanding of sense of place, a capacity for individuals and society to consider potential future changes and what they mean for their circumstances. Leaders in forest management organisations will need to support a greater diversity of inputs into decision making, avoid creating rigid organisational hierarchies that deter innovation, and be inclusive, open and questioning (Konkin and Hopkins 2009 ). They will need to create more opportunities for interaction between researchers, managers and the community and space for reflection on the implications and the outcomes of management actions and unplanned events. Researchers will need to develop new modes of communication, providing knowledge in forms that are appropriate to the management decision and suitable for digestion by a range of different audiences.

From this analysis, key gaps in knowledge for adaptation may not be improved climate scenarios or better understanding of the biophysical responses of individual tree species or forest ecosystems to future climate. Knowledge gaps lie more in understanding the social and community attitudes and values that drive forest management and the decision making processes of forest managers, in order to work out how ‘climate intelligence’ can be built in to these processes.

The impacts of changing climate will vary locally. Consequently, managers must be given the flexibility to respond in ways that meet their particular needs and capacity to choose management options that are applicable to the local situation (Innes et al. 2009 ). This may not be consistent with rigid indicator-driven management assessment processes like forest certification. Whilst policy to support climate change mitigation is primarily a task for national governments and international agreements and processes, responsibility for supporting adaptation will fall more to sub-national and local governments, communities and the private sector. More active management will be required if specific values are to be maintained, particularly for forests in conservation reserves. This will require additional investment, but there has been little analysis to support the business case for investment in adaptation or to determine who should pay, particularly in developing countries.

We need to strengthen the relationship between climate science, forest research, forest managers and the community. Key challenges will include the setting of objectives for desired future conditions and accepting that we may not be able to maintain everything that forests have traditionally provided. It is important to discuss and agree on common goals in order to cope with, or benefit from, the challenges of future climates. Actively managing our forest ecosystems effectively and intelligently, using the best available knowledge and foresight capacity, can make those goals a reality.

Adams MA (2013) Mega-fires, tipping points and ecosystem services: managing forests and woodlands in an uncertain future. For Ecol Manag 294:250–261. doi: 10.1016/j.foreco.2012.11.039

Google Scholar  

Adger WN, Agrawala S, Mirza MMQ, Conde C, O’Brien K, Pulhin J, Pulwarty R, Smit B, Takahashi K (2007) Assessment of adaptation practices, options, constraints and capacity. In: Parry ML, Canziani OF, Palutikof JP, van der Linden PJ, Hanson CE (eds) Climate change 2007: impacts, adaptation and vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the Intergovernmental Panel of Climate Change (IPCC). Cambridge University Press, Cambridge, pp 717–743

Aitken SN, Yeaman S, Holliday JA, Wang T, Curtis-McLane S (2008) Adaptation, migration or extirpation: climate change outcomes for tree populations. Evol Appl 1:95–111. doi: 10.1111/j.1752-4571.2007.00013.x

PubMed Central   PubMed   Google Scholar  

Alberto FJ, Aitken SN, Alia R, Gonzalez-Martinez SC, Hanninen H, Kremer A, Lefevre F, Lenormand T, Yeaman S, Whetten R, Savolainen O (2013) Potential for evolutionary responses to climate change evidence from tree populations. Glob Chang Biol 19:1645–1661. doi: 10.1111/gcb.12181

Allen CD, Macalady AK, Chenchouni H, Bachelet D, McDowell N, Vennetier M, Kitzberger T, Rigling A, Breshears DD, Hogg EH, Gonzalez P, Fensham R, Zhang Z, Castro J, Demidova N, Lim J-H, Allard G, Running SW, Semerci A, Cobb N (2010) A global overview of drought and heat-induced tree mortality reveals emerging climate change risks for forests. For Ecol Manag 259:660–684. doi: 10.1016/j.foreco.2009.09.001

Amraoui M, Liberato MLR, Calado TJ, DaCamara CC, Coelho LP, Trigo RM, Gouveia CM (2013) Fire activity over Mediterranean Europe based on information from Meteosat-8. For Ecol Manag 294:62–75. doi: 10.1016/j.foreco.2012.08.032

Ananda J, Herath G (2009) A critical review of multi-criteria decision making methods with special reference to forest management and planning. Ecol Econ 68:2535–2548. doi: 10.1016/j.ecolecon.2009.05.010

Andalo C, Beaulieu J, Bousquet J (2005) The impact of climate change on growth of local white spruce populations in Québec, Canada. For Ecol Manag 205:169–182. doi: 10.1016/j.foreco.2004.10.045

Anderson JM (1991) The effects of climate change on decomposition processes in grassland and coniferous forests. Ecol Appl 1:326–347. doi: 10.2307/1941761

Anderson JT, Panetta AM, Mitchell-Olds T (2012) Evolutionary and ecological responses to anthropogenic climate change. Plant Physiol 160:1728–1740. doi: 10.1104/pp. 112.206219

PubMed Central   CAS   PubMed   Google Scholar  

Attorre F, Francesconi F, Scarnati L, De Sanctis M, Alfo M, Bruno F (2008) Predicting the effect of climate change on tree species abundance and distribution at a regional scale. For Biogeosci For 1:132–139. doi: 10.3832/ifor0467-0010132

Battaglia M, Bruce J, Brack C, Baker T (2009) Climate change and Australia’s plantation estate: analysis of vulnerability and preliminary investigation of adaptation options.

Battaglia M, Sands P, White D, Mummery D (2004) CABALA: a linked carbon, water and nitrogen model of forest growth for silvicultural decision support. For Ecol Manag 193:251–282

Bele MY, Sonwa DJ, Tiani AM (2013a) Supporting local adaptive capacity to climate change in the Congo basin forest of Cameroon a participatory action research approach. Int J Clim Chang Strateg Manag 5:181–197. doi: 10.1108/17568691311327587

Bele MY, Tiani AM, Somorin OA, Sonwa DJ (2013b) Exploring vulnerability and adaptation to climate change of communities in the forest zone of Cameroon. Clim Chang 119:875–889. doi: 10.1007/s10584-013-0738-z

Berkhout F, Hertin J, Gann D (2006) Learning to adapt: organisational adaptation to climate change impacts. Clim Chang 78:135–156. doi: 10.1007/s10584-006-9089-3

Bernier P, Schöne D (2009) Adapting forests and their management to climate change: an overview. Unasylva 60:5–11

Blate GM, Joyce LA, Littell JS, McNulty SG, Millar CI, Moser SC, Neilson RP, O’Halloran K, Peterson DL (2009) Adapting to climate change in United States national forests. Unasylva 60:57–62

Blennow K, Persson J, Tome M, Hanewinkel M (2012) Climate change: believing and seeing implies adapting. PLoS ONE 7:e50182. doi: 10.1371/journal.pone.0050182

Boisvenue C, Running SW (2006) Impacts of climate change on natural forest productivity—evidence since the middle of the 20th century. Glob Chang Biol 12:862–882. doi: 10.1111/j.1365-2486.2006.01134.x

Bolte A, Eisenhauer DR, Ehrhart HP, Gross J, Hanewinkel M, Kolling C, Profft I, Rohde M, Rohe P, Amereller K (2009) Climate change and forest management—accordances and differences between the German states regarding assessments for needs and strategies towards forest adaptation. Landbauforschung Volkenrode 59:269–278

Bonan GB, Pollard D, Thompson SL (1992) Effects of boreal forest vegetation on global climate. Nature 359:716–718. doi: 10.1038/359716a0

Bongers F, Poorter L, Van Rompaey R, Parren MPE (1999) Distribution of twelve moist forest canopy tree species in Liberia and Cote d’Ivoire: response curves to a climatic gradient. J Veg Sci 10:371–382. doi: 10.2307/3237066

Booth TH (1990) Mapping regions climatically suitable for particular tree species at the global scale. For Ecol Manag 36:47–60

Booth TH (2013) Eucalypt plantations and climate change. For Ecol Manag 301:28–34. doi: 10.1016/j.foreco.2012.04.004

Booth TH, Nix HA, Hutchinson MF, Jovanovic T (1988) Niche analysis and tree species introduction. For Ecol Manag 23:47–59

Bowman DMJS, Williamson GJ, Keenan RJ, Prior LD (2014) A warmer world will reduce tree growth in evergreen broadleaf forests: evidence from Australian temperate and subtropical eucalypt forests. Glob Ecol Biogeogr 23:925–934. doi: 10.1111/geb.12171

Brautigam K, Vining KJ, Lafon-Placette C, Fossdal CG, Mirouze M, Marcos JG, Fluch S, Fraga MF, Guevara MA, Abarca D, Johnsen O, Maury S, Strauss SH, Campbell MM, Rohde A, Diaz-Sala C, Cervera MT (2013) Epigenetic regulation of adaptive responses of forest tree species to the environment. Ecol Evol 3:399–415. doi: 10.1002/ece3.461

Breed MF, Stead MG, Ottewell KM, Gardner MG, Lowe AJ (2013) Which provenance and where? Seed sourcing strategies for revegetation in a changing environment. Conserv Genet 14:1–10. doi: 10.1007/s10592-012-0425-z

Breymeyer A, Melillo JM (1991) Global climate change—the effects of climate change on production and decomposition in coniferous forests and grasslands. Ecol Appl 1:111–111. doi: 10.2307/1941804

Brooker RW, Travis JMJ, Clark EJ, Dytham C (2007) Modelling species’ range shifts in a changing climate: the impacts of biotic interactions, dispersal distance and the rate of climate change. J Theor Biol 245:59–65. doi: 10.1016/j.jtbi.2006.09.033

PubMed   Google Scholar  

Brotak EA, Reifsnyder WE (1977) Predicting major wildland fire occurrence. Fire Manag Notes 38:5–8

Brown HCP (2009) Climate change and Ontario forests: prospects for building institutional adaptive capacity. Mitig Adapt Strateg Glob Chang 14:513–536. doi: 10.1007/s11027-009-9183-8

Buma B, Wessman CA (2013) Forest resilience, climate change, and opportunities for adaptation: a specific case of a general problem. For Ecol Manag 306:216–225. doi: 10.1016/j.foreco.2013.06.044

Campbell EM, Saunders SC, Coates KD, Meidinger DV, MacKinnon A, O’Neil GA, MacKillop DJ, DeLong SC, Morgan. DG (2009) Ecological resilience and complexity: a theoretical framework for understanding and managing British Columbia’s forest ecosystems in a changing climate. B.C. Min. For. Range, For. Sci. Prog., Victoria, B.C.

Cannell MGR, Grace J, Booth A (1989) Possible impacts of climatic warming on trees and forests in the United Kingdom—a review. Forestry 62:337–364. doi: 10.1093/forestry/62.4.337

Carvalho A, Flannigan MD, Logan K, Miranda AI, Borrego C (2008) Fire activity in Portugal and its relationship to weather and the Canadian Fire Weather Index System. Int J Wildl Fire 17:328–338. doi: 10.1071/WF07014

Castagneri D, Motta R (2010) A research gap in the interactive effects of climate and competition on trees growth. In: Karam WP (ed) Tree growth: influences, layers and types. Nova Science, Hauppauge, pp 93–102

Caverley N (2013) The guardians of Mother Earth: a qualitative study of aboriginal knowledge keepers and their views on climate change adaptation in the South Selkirks region. Nativ Stud Rev 21:125–150

Cheaib A, Badeau V, Boe J, Chuine I, Delire C, Dufrêne E, François C, Gritti ES, Legay M, Pagé C, Thuiller W, Viovy N, Leadley P (2012) Climate change impacts on tree ranges: model intercomparison facilitates understanding and quantification of uncertainty. Ecol Lett 15:533–544. doi: 10.1111/j.1461-0248.2012.01764.x

Chen WJ, Black TA, Yang PC, Barr AG, Neumann HH, Nesic Z, Blanken PD, Novak MD, Eley J, Ketler RJ, Cuenca A (1999) Effects of climatic variability on the annual carbon sequestration by a boreal aspen forest. Glob Chang Biol 5:41–53. doi: 10.1046/j.1365-2486.1998.00201.x

CAS   Google Scholar  

Chmura DJ, Anderson PD, Howe GT, Harrington CA, Halofsky JE, Peterson DL, Shaw DC, St. Brad Clair J (2011) Forest responses to climate change in the northwestern United States: ecophysiological foundations for adaptive management. For Ecol Manag 261:1121–1142. doi: 10.1016/j.foreco.2010.12.040

Choat B, Jansen S, Brodribb TJ, Cochard H, Delzon S, Bhaskar R, Bucci SJ, Feild TS, Gleason SM, Hacke UG, Jacobsen AL, Lens F, Maherali H, Martinez-Vilalta J, Mayr S, Mencuccini M, Mitchell PJ, Nardini A, Pittermann J, Pratt RB, Sperry JS, Westoby M, Wright IJ, Zanne AE (2012) Global convergence in the vulnerability of forests to drought. Nature 491:752–755

CAS   PubMed   Google Scholar  

Chuvieco E, Martinez S, Roman MV, Hantson S, Pettinari ML (2014) Integration of ecological and socio-economic factors to assess global vulnerability to wildfire. Glob Ecol Biogeogr 23:245–258. doi: 10.1111/geb.12095

Clark DA (2007) Detecting tropical forests’ responses to global climatic and atmospheric change: current challenges and a way forward. Biotropica 39:4–19. doi: 10.1111/j.1744-7429.2006.00227.x

Clark JS, Bell DM, Hersh MH, Nichols L (2011) Climate change vulnerability of forest biodiversity: climate and competition tracking of demographic rates. Glob Chang Biol 17:1834–1849. doi: 10.1111/j.1365-2486.2010.02380.x

Cockfield G, Maraseni T, Buys L, Sommerfeld J, Wilson C, Athukorala W (2011) Socioeconomic implications of climate change with regard to forests and forest management. Contribution of Work Package 3 to the Forest Vulnerability Assessment. National Climate Change Adaptation Research Facility, Gold Coast

Corlett RT (2011) Impacts of warming on tropical lowland rainforests. Trends Ecol Evol 26:606–613

Corlett RT, Westcott DA (2013) Will plant movements keep up with climate change? Trends Ecol Evol 28:482–488. doi: 10.1016/j.tree.2013.04.003

Couture S, Reynaud A (2008) Multi-stand forest management under a climatic risk: do time and risk preferences matter? Environ Model Assess 13:181–193

D’Amato AW, Bradford JB, Fraver S, Palik BJ (2011) Forest management for mitigation and adaptation to climate change: insights from long-term silviculture experiments. For Ecol Manag 262:803–816. doi: 10.1016/j.foreco.2011.05.014

Daubenmire RF (1978) Plant geography. Academic, New York

Davidar P, Rajagopal B, Mohandass D, Puyravaud JP, Condit R, Wright SJ, Leigh EG (2007) The effect of climatic gradients, topographic variation and species traits on the beta diversity of rain forest trees. Glob Ecol Biogeogr 16:510–518. doi: 10.1111/j.1466-8238.2007.00307.x

Davis MB (1986) Climatic instability, time lags and community disequilibrium. In: Diamond J, Case TJ (eds) Community ecology. Harper and Row, New York, pp 269–284

Day JK, Perez DM (2013) Reducing uncertainty and risk through forest management planning in British Columbia. For Ecol Manag 300:117–124. doi: 10.1016/j.foreco.2012.11.035

de Groot WJ, Flannigan MD, Cantin AS (2013) Climate change impacts on future boreal fire regimes. For Ecol Manag 294:35–44. doi: 10.1016/j.foreco.2012.09.027

Dessai S, Hulme M (2004) Does climate adaptation policy need probabilities? Clim Pol 4:107–128. doi: 10.1080/14693062.2004.9685515

Di Nitto D, Neukermans G, Koedam N, Defever H, Pattyn F, Kairo JG, Dahdouh-Guebas F (2014) Mangroves facing climate change: landward migration potential in response to projected scenarios of sea level rise. Biogeosciences 11:857–871. doi: 10.5194/bg-11-857-2014

Dixon RK, Krankina ON, Kobak KI (1996) Global climate change adaptation: examples from Russian boreal forests. Adapting to climate change: an international perspective.

Dobrowski SZ, Abatzoglou J, Swanson AK, Greenberg JA, Mynsberge AR, Holden ZA, Schwartz MK (2013) The climate velocity of the contiguous United States during the 20th century. Glob Chang Biol 19:241–251. doi: 10.1111/gcb.12026

Doyle TW, Krauss KW, Conner WH, From AS (2010) Predicting the retreat and migration of tidal forests along the northern Gulf of Mexico under sea-level rise. For Ecol Manag 259:770–777. doi: 10.1016/j.foreco.2009.10.023

Eliasch J (2008) Climate change: financing global forests. The Eliasch Review. HMSO

Emanuel WR, Shugart HH, Stevenson MP (1985) Climatic-change and the broad-scale distribution of terrestrial ecosystem complexes. Clim Chang 7:29–43. doi: 10.1007/bf00139439

Eriksson LO, Backéus S, Garcia F (2011) Implications of growth uncertainties associated with climate change for stand management. Eur J For Res 131:1199–1209. doi: 10.1007/s10342-011-0591-4

Ettl GJ, Peterson DL (1995) Extreme climate and variation in tree growth—individualistic response in sub-alpine fir ( Abies-lasiocarpa ). Glob Chang Biol 1:231–241. doi: 10.1111/j.1365-2486.1995.tb00024.x

Feeley KJ, Rehm EM, Machovina B (2012) Perspective: the responses of tropical forest species to global climate change: acclimate, adapt, migrate, or go extinct? Frontiers of Biogeography 4 (2)

Feikema PM, Sherwin CB, Lane PNJ (2013) Influence of climate, fire severity and forest mortality on predictions of long term streamflow: potential effect of the 2009 wildfire on Melbourne’s water supply catchments. J Hydrol 488:1–16. doi: 10.1016/j.jhydrol.2013.02.001

Fischer AP, Paveglio T, Carroll M, Murphy D, Brenkert-Smith H (2013) Assessing social vulnerability to climate change in human communities near public forests and grasslands: a framework for resource managers and planners. J For 111:357–365. doi: 10.5849/jof. 12-091

Fischlin A, Ayres M, Karnosky D, Kellomäki S, Louman B, Ong C, Plattner G-K, Santoso H, Thompson I, Booth TH, Marcar N, Scholes B, Swanston C, Zamolodchikov D (2009) Future environmental impacts and vulnerabilities. In: Seppälä R, Buck A, Katila P (eds) Adaptation of forests and people to climate change: a global assessment report, vol 22. IUFRO World Series, Helsinki, pp 53–100

Franks PJ, Adams MA, Amthor JS, Barbour MM, Berry JA, Ellsworth DS, Farquhar GD, Ghannoum O, Lloyd J, McDowell N, Norby RJ, Tissue DT, von Caemmerer S (2013) Sensitivity of plants to changing atmospheric CO2 concentration: from the geological past to the next century. New Phytol 197:1077–1094. doi: 10.1111/nph.12104

Furness E, Nelson H (2012) Community forest organizations and adaptation to climate change in British Columbia. For Chron 88:519–524

Garcia RA, Cabeza M, Rahbek C, Araújo MB (2014) Multiple dimensions of climate change and their implications for biodiversity. Science 344 (6183). doi:10.1126/science.1247579

Gentilli J (1949) Forest influences: the effects of woody vegetation on climatic water, and soil, with applications to the conservation of water and the control of floods and erosion. Geogr Rev 39:164–164. doi: 10.2307/211169

Gibbs HK, Ruesch AS, Achard F, Clayton MK, Holmgren P, Ramankutty N, Foley JA (2010) Tropical forests were the primary sources of new agricultural land in the 1980s and 1990s. Proc Natl Acad Sci 107:16732–16737. doi: 10.1073/pnas.0910275107

Gilman RT, Fabina NS, Abbott KC, Rafferty NE (2012) Evolution of plant–pollinator mutualisms in response to climate change. Evol Appl 5:2–16. doi: 10.1111/j.1752-4571.2011.00202.x

Gonzalez ME, Lara A, Urrutia R, Bosnich J (2011) Climatic change and its potential impact on forest fire occurrence in south-central Chile (33 degrees-42 degrees S). Bosque 32:215–219. doi: 10.4067/s0717-92002011000300002

Graham RL, Turner MG, Dale VH (1990) How increasing CO2 and climate change affect forests—at many spatial and temporal scales, there will be forest responses that will be affected by human activities. Bioscience 40:575–587. doi: 10.2307/1311298

Graumlich LJ (1993) Response of tree growth to climatic variation in the mixed conifer and deciduous forests of the upper Great-Lakes region. Can J For Res 23:133–143. doi: 10.1139/x93-020

Groffman PM, Rustad LE, Templer PH, Campbell JL, Christenson LM, Lany NK, Socci AM, Vadeboncoeur MA, Schaberg PG, Wilson GF, Driscoll CT, Fahey TJ, Fisk MC, Goodale CL, Green MB, Hamburg SP, Johnson CE, Mitchell MJ, Morse JL, Pardo LH, Rodenhouse NL (2012) Long-term integrated studies show complex and surprising effects of climate change in the northern hardwood forest. Bioscience 62:1056–1066. doi: 10.1525/bio.2012.62.12.7

Grubb PJ, Whitmore TC (1966) A comparison of montane and lowland rain forest in Ecuador. 2. Climate and its effects on distribution and physiognomy of forests. J Ecol 54:303. doi: 10.2307/2257951

Guan BT, Wright WE, Chung C-H, Chang S-T (2012) ENSO and PDO strongly influence Taiwan spruce height growth. For Ecol Manag 267:50–57. doi: 10.1016/j.foreco.2011.11.028

Guardiola-Claramonte M, Troch PA, Breshears DD, Huxman TE, Switanek MB, Durcik M, Cobb NS (2011) Decreased streamflow in semi-arid basins following drought-induced tree die-off: a counter-intuitive and indirect climate impact on hydrology. J Hydrol 406:225–233. doi: 10.1016/j.jhydrol.2011.06.017

Guariguata MR (2009) Tropical forest management and climate change adaptation. Rev Estud Soc 32:98–112

Guariguata MR, Cornelius JP, Locatelli B, Forner C, Sánchez-Azofeifa GA (2008) Mitigation needs adaptation: tropical forestry and climate change. Mitig Adapt Strateg Glob Chang 13:793–808. doi: 10.1007/s11027-007-9141-2

Guariguata MR, Locatelli B, Haupt F (2012) Adapting tropical production forests to global climate change: risk perceptions and actions. Int For Rev 14:27–38

Gyampoh BA, Amisah S, Idinoba M, Nkem J (2009) Using traditional knowledge to cope with climate change in rural Ghana. Unasylva 60:70–74

Hahn WA, Knoke T (2010) Sustainable development and sustainable forestry: analogies, differences, and the role of flexibility. Eur J For Res 129:787–801. doi: 10.1007/s10342-010-0385-0

Hamrick JL (2004) Response of forest trees to global environmental changes. For Ecol Manag 197:323–335. doi: 10.1016/j.foreco.2004.05.023

Hansenbristow KJ, Ives JD, Wilson JP (1988) Climatic variability and tree response within the forest alpine tundra ecotone. Ann Assoc Am Geogr 78:505–519. doi: 10.1111/j.1467-8306.1988.tb00221.x

Harper RJ, Smettem KRJ, Carter JO, McGrath JF (2009) Drought deaths in Eucalyptus globulus (Labill.) plantations in relation to soils, geomorphology and climate. Plant Soil 324:199–207. doi: 10.1007/s11104-009-9944-x

Heller NE, Zavaleta ES (2009) Biodiversity management in the face of climate change: a review of 22 years of recommendations. Biol Conserv 142:14–32. doi: 10.1016/j.biocon.2008.10.006

Hennessy K, Lucas C, Nicholls N, Bathols J, Suppiah R, Ricketts J (2005) Climate change impacts on fire-weather in south-east Australia. CSIRO Division of Marine and Atmospheric Research, Aspendale

Holling CS (1978) Adaptive environmental assessment and management. Wiley, Chichester

Hughes L, Cawsey EM, Westoby M (1996) Climatic range sizes of Eucalyptus species in relation to future climate change. Glob Ecol Biogeogr Lett 5:23–29. doi: 10.2307/2997467

Huntingford C, Zelazowski P, Galbraith D, Mercado LM, Sitch S, Fisher R, Lomas M, Walker AP, Jones CD, Booth BBB, Malhi Y, Hemming D, Kay G, Good P, Lewis SL, Phillips OL, Atkin OK, Lloyd J, Gloor E, Zaragoza-Castells J, Meir P, Betts R, Harris PP, Nobre C, Marengo J, Cox PM (2013) Simulated resilience of tropical rainforests to CO2-induced climate change. Nat Geosci 6:268–273

Huntley B (1990) European postglacial forests—compositional changes in response to climatic-change. J Veg Sci 1:507–518. doi: 10.2307/3235785

Ince PJ, Kramp AD, Skog KE, Yoo DI, Sample VA (2011) Modeling future U.S. forest sector market and trade impacts of expansion in wood energy consumption. J For Econ 17:142–156. doi: 10.1016/j.jfe.2011.02.007

Innes J, Joyce LA, Kellomäki S, Louman B, Ogden A, Parrotta J, Thompson I, Ayres M, Ong C, Santoso H, Sohngen B, Wreford A (2009) Management for adaptation. In: Seppälä R, Buck A, Katila P (eds) Adaptation of forests and people to climate change: a global assessment report, vol World Series Volume 22. IUFRO Helsinki, pp 135–186

IPCC (2013) Climate Change 2013: The physical science basis. Contribution of Working Group I to the Fifth Assessment Report of the Intergovernmental Panel on Climate Change [Stocker, T.F., D. Qin, G.-K. Plattner, M. Tignor, S.K. Allen, J. Boschung, A. Nauels, Y. Xia, V. Bex and P.M. Midgley (eds.)]. Cambridge, United Kingdom and New York, NY, USA

IUCN (2008) Ecosystem-based adaptation: an approach for building resilience and reducing risk for local communities and ecosystems. Submission to the Chair of the AWG-LCA with respect to the Shared Vision and Enhanced Action on Adaptation. International Union for the Conservation of Nature,

Iverson LR, Schwartz MW, Prasad AM (2004) How fast and far might tree species migrate in the eastern United States due to climate change? Glob Ecol Biogeogr 13:209–219

Johnson G, Sorensen FC, St Clair JB, Cronn RC (2004) Pacific Northwest forest tree seed zones: a template for native plants? Nativ Plants J 5:131–140

Johnston M, Hesseln H (2012) Climate change adaptive capacity of the Canadian forest sector. For Policy Econ 24:29–34. doi: 10.1016/j.forpol.2012.06.001

Johnston M, Lindner M, Parrotta J, Giessen L (2012) Adaptation and mitigation options for forests and forest management in a changing climate. For Policy Econ 24:1–2. doi: 10.1016/j.forpol.2012.09.007

Johnston M, Williamson T (2007) A framework for assessing climate change vulnerability of the Canadian forest sector. For Chron 83:358–361

Joyce LA (2007) The impacts of climate change on forestry. In: Adams DM, Haynes RW (eds) Resource and market projections for forest policy development: twenty-five years of experience with the US RPA Timber Assessment, vol 14. Managing Forest Ecosystems. pp 449–488

Joyce LA, Blate GM, Littell JS, McNulty SG, Millar CI, Moser SC, Neilson RP, O’Halloran K, Peterson DL (2008) National forests. Preliminary review of adaptation options for climate-sensitive ecosystems and resources. A report by the U.S. Climate Change Science Program and the Subcommittee on Global Change Research. US Environmental Protection Agency, Washington, DC

Joyce LA, Blate GM, McNulty SG, Millar CI, Moser S, Neilson RP, Peterson DL (2009) Managing for multiple resources under climate change: national forests. Environ Manag 44:1022–1032. doi: 10.1007/s00267-009-9324-6

Joyce LA, Mills JR, Heath LS, McGuire AD, Haynes RW, Birdsey RA (1995) Forest sector impacts from changes in forest productivity under climate change. J Biogeogr 22:703–713. doi: 10.2307/2845973

Kalame FB, Luukkanen O, Kanninen M (2011) Making the National Adaptation Programme of Action (NAPA) more responsive to the livelihood needs of tree planting farmers, drawing on previous experience in dryland Sudan. Forests 2:948–960. doi: 10.3390/f2040948

Kallarackal J, Roby TJ (2012) Responses of trees to elevated carbon dioxide and climate change. Biodivers Conserv 21:1327–1342. doi: 10.1007/s10531-012-0254-x

Kant P, Wu S (2012) Should adaptation to climate change be given priority over mitigation in tropical forests? Carbon Manag 3:303–311. doi: 10.4155/cmt.12.29

Karjalainen T, Pussinen A, Liski J, Nabuurs GJ, Eggers T, Lapvetelainen T, Kaipainen T (2003) Scenario analysis of the impacts of forest management and climate change on the European forest sector carbon budget. For Policy Econ 5:141–155. doi: 10.1016/s1389-9341(03)00021-2

Keenan RJ (2012) Adaptation of forests and forest management to climate change: an editorial. Forests 3:75–82. doi: 10.3390/f3010075

Kellomaki S, Peltola H, Nuutinen T, Korhonen KT, Strandman H (2008) Sensitivity of managed boreal forests in Finland to climate change, with implications for adaptive management. Phil Trans R Soc B 363:2341–2351. doi: 10.1098/rstb.2007.2204

Kelly AE, Goulden ML (2008) Rapid shifts in plant distribution with recent climate change. Proc Natl Acad Sci 105:11823–11826. doi: 10.1073/pnas.0802891105

Kerhoulas LP, Kolb TE, Hurteau MD, Koch GW (2013) Managing climate change adaptation in forests: a case study from the US Southwest. J Appl Ecol 50:1311–1320. doi: 10.1111/1365-2664.12139

Keskitalo EC (2008) Vulnerability and adaptive capacity in forestry in northern Europe: a Swedish case study. Clim Chang 87:219–234. doi: 10.1007/s10584-007-9337-1

Keskitalo ECH (2011) How can forest management adapt to climate change? Possibilities in different forestry systems. Forests 2:415–430. doi: 10.3390/f2010415

Kienast F, Brzeziecki B, Wildi O (1996) Long-term adaptation potential of Central European mountain forests to climate change: a GIS-assisted sensitivity assessment. For Ecol Manag 80:133–153. doi: 10.1016/0378-1127(95)03633-4

Kimmins JP (2002) Future shock in forestry—where have we come from; where are we going; is there a “right way” to manage forests? Lessons from Thoreau, Leopold, Toffler, Botkin and Nature. For Chron 78:263–271

Kimmins JP (2008) From science to stewardship: harnessing forest ecology in the service of society. For Ecol Manag 256:1625–1635. doi: 10.1016/j.foreco.2008.02.057

Kimmins JP, Blanco JA, Seely B, Welham C, Scoullar K (2008) Complexity in modelling forest ecosystems: how much is enough? For Ecol Manag 256:1646–1658. doi: 10.1016/j.foreco.2008.03.011

Kimmins JP, Mailly D, Seely B (1999) Modelling forest ecosystem net primary production: the hybrid simulation approach used in forecast. Ecol Model 122:195–224

Kint V, Aertsen W, Campioli M, Vansteenkiste D, Delcloo A, Muys B (2012) Radial growth change of temperate tree species in response to altered regional climate and air quality in the period 1901–2008. Clim Chang 115:343–363. doi: 10.1007/s10584-012-0465-x

Kirschbaum MUF, Watt MS, Tait A, Ausseil A-GE (2012) Future wood productivity of Pinus radiata in New Zealand under expected climatic changes. Glob Chang Biol 18:1342–1356. doi: 10.1111/j.1365-2486.2011.02625.x

Klenk NL, Adams BW, Bull GQ, Innes JL, Cohen SJ, Larson BC (2011) Climate change adaptation and sustainable forest management: a proposed reflexive research agenda. For Chron 87:351–357

Kobak KI, Turchinovich IY, Kondrasheva NY, Schulze ED, Schulze W, Koch H, Vygodskaya NN (1996) Vulnerability and adaptation of the larch forest in eastern Siberia to climate change. Water Air Soil Pollut 92:119–127

Kolström M, Lindner M, Vilén T, Maroschek M, Seidl R, Lexer MJ, Netherer S, Kremer A, Delzon S, Barbati A, Marchetti M, Corona P (2011) Reviewing the science and Implementation of climate change adaptation measures in European forestry. Forests 2:961–982. doi: 10.3390/f2040961

Konkin D, Hopkins K (2009) Learning to deal with climate change and catastrophic forest disturbances. Unasylva 60:17–23

Kremer A, Ronce O, Robledo-Arnuncio JJ, Guillaume F, Bohrer G, Nathan R, Bridle JR, Gomulkiewicz R, Klein EK, Ritland K, Kuparinen A, Gerber S, Schueler S (2012) Long-distance gene flow and adaptation of forest trees to rapid climate change. Ecol Lett 15:378–392. doi: 10.1111/j.1461-0248.2012.01746.x

PubMed Central   Google Scholar  

Kuparinen A, Savolainen O, Schurr FM (2010) Increased mortality can promote evolutionary adaptation of forest trees to climate change. For Ecol Manag 259:1003–1008. doi: 10.1016/j.foreco.2009.12.006

Lal R, Cummings DJ (1979) Clearing a tropical forest. 1. Effects on soil and micro-climate. Field Crop Res 2:91–107. doi: 10.1016/0378-4290(79)90012-1

Leites LP, Rehfeldt GE, Robinson AP, Crookston NL, Jaquish B (2012) Possibilities and limitations of using historic provenance tests to infer forest species growth responses to climate change. Nat Resour Model 25:409–433. doi: 10.1111/j.1939-7445.2012.00129.x

Lenoir J, Gegout JC, Dupouey JL, Bert D, Svenning JC (2010) Forest plant community changes during 1989–2007 in response to climate warming in the Jura Mountains (France and Switzerland). J Veg Sci 21:949–964. doi: 10.1111/j.1654-1103.2010.01201.x

Leuzinger S, Luo Y, Beier C, Dieleman W, Vicca S, Körner C (2011) Do global change experiments overestimate impacts on terrestrial ecosystems? Trends Ecol Evol 26:236–241

Levina E, Tirpak D (2006) Adaptation to climate change: key terms. OECD/IEA, Paris

Lin BB, Petersen B (2013) Resilience, regime shifts, and guided transition under climate change: examining the practical difficulties of managing continually changing systems. Ecology and Society 18 (1). doi:10.5751/es-05128-180128

Lindner M (2000) Developing adaptive forest management strategies to cope with climate change. Tree Physiol 20:299–307

Lindner M, Fitzgerald JB, Zimmermann NE, Reyer C, Delzon S, van der Maaten E, Schelhaas MJ, Lasch P, Eggers J, van der Maaten-Theunissen M, Suckow F, Psomas A, Poulter B, Hanewinkel M (2014) Climate change and European forests: what do we know, what are the uncertainties, and what are the implications for forest management? J Environ Manag 146:69–83. doi: 10.1016/j.jenvman.2014.07.030

Lindner M, Maroschek M, Netherer S, Kremer A, Barbati A, Garcia-Gonzalo J, Seidl R, Delzon S, Corona P, Kolström M, Lexer MJ, Marchetti M (2010) Climate change impacts, adaptive capacity, and vulnerability of European forest ecosystems. For Ecol Manag 259:698–709. doi: 10.1016/j.foreco.2009.09.023

Lindner M, Sohngen B, Joyce LA, Price DT, Bernier PY, Karjalainen T (2002) Integrated forestry assessments for climate change impacts. For Ecol Manag 162:117–136. doi: 10.1016/S0378-1127(02)00054-3

Littell JS, Peterson DL, Millar CI, O’Halloran KA (2012) U.S. national forests adapt to climate change through science-management partnerships. Clim Chang 110:269–296. doi: 10.1007/s10584-011-0066-0

Liu Y, Stanturf J, Goodrick S (2010) Trends in global wildfire potential in a changing climate. For Ecol Manag 259:685–697. doi: 10.1016/j.foreco.2009.09.002

Locatelli B, Evans V, Wardell A, Andrade A, Vignola R (2011) Forests and climate change in Latin America: linking adaptation and mitigation. Forests 2:431–450. doi: 10.3390/f2010431

Loehman RA, Clark JA, Keane RE (2011) Modeling effects of climate change and fire management on Western White Pine (Pinus monticola) in the Northern Rocky Mountains, USA. Forests 2:832–860. doi: 10.3390/f2040832

Long A (2013) REDD plus, adaptation, and sustainable forest management: toward effective polycentric global forest governance. Trop Conserv Sci 6:384–408

Lucier A, Ayres M, Karnosky D, Thompson I, Loehle C, Percy K, Sohngen B (2009) Forest responses and vulnerabilities to recent climate change. In: Seppälä R, Buck A, Katila P (eds) Adaptation of forests and people to climate change: a global assessment report, vol World Series Volume 22. IUFRO Helsinki, pp 29–52

Lynch AJJ, Fell DG, McIntyre-Tamwoy S (2010) Incorporating Indigenous values with ‘Western’ conservation values in sustainable biodiversity management. Aust J Environ Manag 17:244–255

Macdonald GM, Edwards TWD, Moser KA, Pienitz R, Smol JP (1993) Rapid response of treeline vegetation and lakes to past climate warming. Nature 361:243–246. doi: 10.1038/361243a0

Mahat V, Anderson A (2013) Impacts of climate and catastrophic forest changes on streamflow and water balance in a mountainous headwater stream in Southern Alberta. Hydrol Earth Syst Sci 17:4941–4956. doi: 10.5194/hess-17-4941-2013

Malcolm JR, Markham A, Neilson RP, Garaci M (2002) Estimated migration rates under scenarios of global climate change. J Biogeogr 29:835–849. doi: 10.1046/j.1365-2699.2002.00702.x

Matthews SN, Iverson LR, Prasad AM, Peters MP (2011) Changes in potential habitat of 147 North American breeding bird species in response to redistribution of trees and climate following predicted climate change. Ecography 34:933–945. doi: 10.1111/j.1600-0587.2011.06803.x

McEvoy D, Fünfgeld H, Bosomworth K (2013) Resilience and climate change adaptation: the importance of framing. Plan Pract Res 28:280–293. doi: 10.1080/02697459.2013.787710

Medlyn B, Zeppel M, Brouwers N, Howard K, O’Gara E, Hardy G, Lyons T, Li L, Evans B (2011) Biophysical impacts of climate change on Australia’s forests. Contribution of Work Package 2 to the Forest Vulnerability Assessment. National Climate Change Adaptation Research Facility, Gold Coast

Messier C, Puettmann K, Coates DJ (2013) Managing forests as complex adaptive systems: building resilience to the challenge of global change. Earthscan, London

Michelot A, Breda N, Damesin C, Dufrene E (2012) Differing growth responses to climatic variations and soil water deficits of Fagus sylvatica, Quercus petraea and Pinus sylvestris in a temperate forest. For Ecol Manag 265:161–171. doi: 10.1016/j.foreco.2011.10.024

Milad M, Schaich H, Konold W (2013) How is adaptation to climate change reflected in current practice of forest management and conservation? A case study from Germany. Biodivers Conserv 22:1181–1202. doi: 10.1007/s10531-012-0337-8

Miles L, Grainger A, Phillips O (2004) The impact of global climate change on tropical forest biodiversity in Amazonia. Glob Ecol Biogeogr 13:553–565. doi: 10.1111/j.1466-822X.2004.00105.x

Mok H-F, Arndt SK, Nitschke CR (2012) Modelling the potential impact of climate variability and change on species regeneration potential in the temperate forests of South-Eastern Australia. Glob Chang Biol 18:1053–1072. doi: 10.1111/j.1365-2486.2011.02591.x

Mori AS, Spies TA, Sudmeier-Rieux K, Andrade A (2013) Reframing ecosystem management in the era of climate change: issues and knowledge from forests. Biol Conserv 165:115–127. doi: 10.1016/j.biocon.2013.05.020

Namkoong G (2001) Forest genetics: pattern and complexity. Can J For Res 31:623–632. doi: 10.1139/cjfr-31-4-623

Neukum C, Azzam R (2012) Impact of climate change on groundwater recharge in a small catchment in the Black Forest, Germany. Hydrogeol J 20:547–560. doi: 10.1007/s10040-011-0827-x

Nilsson AE, Gerger Swartling Å, Eckerberg K (2012) Knowledge for local climate change adaptation in Sweden: challenges of multilevel governance. Local Environ 17:751–767. doi: 10.1080/13549839.2012.678316

Nitschke CR, Innes JL (2008a) Integrating climate change into forest management in South-Central British Columbia: an assessment of landscape vulnerability and development of a climate-smart framework. For Ecol Manag 256:313–327. doi: 10.1016/j.foreco.2008.04.026

Nitschke CR, Innes JL (2008b) A tree and climate assessment tool for modelling ecosystem response to climate change. Ecol Model 210:263–277. doi: 10.1016/j.ecolmodel.2007.07.026

Nkem J, Kalame FB, Idinoba M, Somorin OA, Ndoye O, Awono A (2010) Shaping forest safety nets with markets: adaptation to climate change under changing roles of tropical forests in Congo Basin. Environ Sci Pol 13:498–508. doi: 10.1016/j.envsci.2010.06.004

Ogden AE, Innes J (2007a) Incorporating climate change adaptation considerations into forest management planning in the boreal forest. Int For Rev 9:713–733. doi: 10.1505/ifor.9.3.713

Ogden AE, Innes JL (2007b) Perspectives of forest practitioners on climate change adaptation in the Yukon and Northwest Territories of Canada. For Chron 83:557–569

Ogden AE, Innes JL (2009) Application of structured decision making to an assessment of climate change vulnerabilities and adaptation options for sustainable forest management. Ecology and Society 14 (1). doi:11

Ohlson DW, McKinnon GA, Hirsch KG (2005) A structured decision-making approach to climate change adaptation in the forest sector. For Chron 81:97–103

Otterman J, Chou MD, Arking A (1984) Effects of nontropical forest cover on climate. J Clim Appl Meteorol 23:762–767. doi: 10.1175/1520-0450(1984)023<0762:eonfco>2.0.co;2

Parks CG, Bernier P (2010) Adaptation of forests and forest management to changing climate with emphasis on forest health: a review of science, policies and practices. For Ecol Manag 259:657–659. doi: 10.1016/s0378-1127(09)00903-7

Pastor J, Post WM (1988) Response of northern forests to CO2-induced climate change. Nature 334:55–58. doi: 10.1038/334055a0

Pearson RG, Dawson TP (2003) Predicting the impacts of climate change on the distribution of species: are bioclimate envelope models useful? Glob Ecol Biogeogr 12:361–371. doi: 10.1046/j.1466-822X.2003.00042.x

Peel J, Godden L, Keenan RJ (2012) Climate change law in an era of multi-level governance. Transl Environ Law 1:245–280. doi: 10.1017/S2047102512000052

Peters EB, Wythers KR, Zhang SX, Bradford JB, Reich PB (2013) Potential climate change impacts on temperate forest ecosystem processes. Can J For Res 43:939–950. doi: 10.1139/cjfr-2013-0013

Peterson DL, Millar CI, Joyce LA, Furniss MJ, Halofsky JE, Neilson RP, Morelli TL (2011) Responding to climate change in national forests: a guidebook for developing adaptation options. General Technical Report. USDA Forest Service, Pacific Northwest Research Station

Phillips OL, Lewis SL, Baker TR, Chao K-J, Higuchi N (2008) The changing Amazon forest. Phil Trans R Soc B 363:1819–1827. doi: 10.1098/rstb.2007.0033

Pinkard EA, Battaglia M, Bruce J, Leriche A, Kriticos DJ (2010) Process-based modelling of the severity and impact of foliar pest attack on eucalypt plantation productivity under current and future climates. For Ecol Manag 259:839–847. doi: 10.1016/j.foreco.2009.06.027

Pojar J, Klinka K, Meidinger DV (1987) Biogeoclimatic ecosystem classification in British Columbia. For Ecol Manag 22:119–154

Pramova E, Locatelli B, Djoudi H, Somorin OA (2012) Forests and trees for social adaptation to climate variability and change. Wiley Interdiscip Rev Clim Chang 3:581–596. doi: 10.1002/wcc.195

Prato T (2008) Conceptual framework for assessment and management of ecosystem impacts of climate change. Ecol Complex 5:329–338

Pretzsch H, Biber P, Schütze G, Uhl E, Rötzer T (2014) Forest stand growth dynamics in Central Europe have accelerated since 1870. Nat Commun 5. doi:10.1038/ncomms5967

Rastetter EB, Ryan MG, Shaver GR, Melillo JM, Nadelhoffer KJ, Hobbie JE, Aber JD (1991) A general biogeochemical model describing the responses of the C and N cycles in terrestrial ecosystems to changes in CO2, climate, and N deposition. Tree Physiol 9:101–126

Rayner J, McNutt K, Wellstead A (2013) Dispersed capacity and weak coordination: the challenge of climate change adaptation in Canada’s forest policy sector. Rev Policy Res 30:66–90. doi: 10.1111/ropr.12003

Regniere J (2009) Predicting insect continental distributions from species physiology. Unasylva 60:37–42

Reyer C, Lasch-Born P, Suckow F, Gutsch M, Murawski A, Pilz T (2014) Projections of regional changes in forest net primary productivity for different tree species in Europe driven by climate change and carbon dioxide. Ann For Sci 71:211–225. doi: 10.1007/s13595-013-0306-8

Rigling A, Bigler C, Eilmann B, Feldmeyer-Christe E, Gimmi U, Ginzler C, Graf U, Mayer P, Vacchiano G, Weber P, Wohlgemuth T, Zweifel R, Dobbertin M (2013) Driving factors of a vegetation shift from Scots pine to pubescent oak in dry Alpine forests. Glob Chang Biol 19:229–240. doi: 10.1111/gcb.12038

Rist L, Moen J (2013) Sustainability in forest management and a new role for resilience thinking. For Ecol Manag 310:416–427. doi: 10.1016/j.foreco.2013.08.033

Roberts G, Parrotta J, Wreford A (2009) Current adaptation measures and policies. In: Seppälä R, Buck A, Katila P (eds) Adaptation of forests and people to climate change: a global assessment report, vol World Series Volume 22. IUFRO Helsinki, pp 123–134

Ruiz-Labourdette D, Schmitz MF, Pineda FD (2013) Changes in tree species composition in Mediterranean mountains under climate change: indicators for conservation planning. Ecol Indic 24:310–323. doi: 10.1016/j.ecolind.2012.06.021

Running SW, Nemani RR (1991) Regional hydrologic and carbon balance responses of forests resulting from potential climate change. Clim Chang 19:349–368. doi: 10.1007/bf00151173

Sample VA, Halofsky JE, Peterson DL (2014) US strategy for forest management adaptation to climate change: building a framework for decision making. Ann For Sci 71:125–130. doi: 10.1007/s13595-013-0288-6

Schaich H, Milad M (2013) Forest biodiversity in a changing climate: which logic for conservation strategies? Biodivers Conserv 22:1107–1114. doi: 10.1007/s10531-013-0491-7

Schoene DHF, Bernier PY (2012) Adapting forestry and forests to climate change: a challenge to change the paradigm. For Policy Econ 24:12–19. doi: 10.1016/j.forpol.2011.04.007

Schwartz MW, Dolanc CR, Gao H, Strauss SY, Schwartz AC, Williams JN, Tang Y (2013) Forest structure, stand composition, and climate-growth response in montane forests of Jiuzhaigou National Nature Reserve, China. Plos One 8 (8). doi:10.1371/journal.pone.0071559

Scott D (2005) Integrating climate change into Canada’s National Parks System. In: Lovejoy T, Hannah L (eds) Climate change and biodiversity. Yale University Press, New Haven, pp 343–345

Seidl R, Lexer MJ (2013) Forest management under climatic and social uncertainty: trade-offs between reducing climate change impacts and fostering adaptive capacity. J Environ Manag 114:461–469. doi: 10.1016/j.jenvman.2012.09.028

Seidl R, Rammer W, Lexer MJ (2011) Adaptation options to reduce climate change vulnerability of sustainable forest management in the Austrian Alps. Can J For Res 41:694–706. doi: 10.1139/x10-235

Seppala R (2009) A global assessment on adaptation of forests to climate change. Scand J For Res 24:469–472. doi: 10.1080/02827580903378626

Seppälä R, Buck A, Katila P (2009) Adaptation of forests and people to climate change: a global assessment report, vol World Series Volume 22. Helsinki, IUFRO

Six DL (2009) Climate change and mutualism. Nat Rev Microbiol 7:686–686

Somorin OA, Brown HCP, Visseren-Hamakers IJ, Sonwa DJ, Arts B, Nkem J (2012) The Congo Basin forests in a changing climate: Policy discourses on adaptation and mitigation (REDD+). Glob Environ Chang 22:288–298. doi: 10.1016/j.gloenvcha.2011.08.001

Sonwa DJ, Somorin OA, Jum C, Bele MY, Nkem JN (2012) Vulnerability, forest-related sectors and climate change adaptation: the case of Cameroon. For Policy Econ 23:1–9. doi: 10.1016/j.forpol.2012.06.009

Sork VL, Aitken SN, Dyer RJ, Eckert AJ, Legendre P, Neale DB (2013) Putting the landscape into the genomics of trees: approaches for understanding local adaptation and population responses to changing climate. Tree Genet Genome 9:901–911. doi: 10.1007/s11295-013-0596-x

Spathelf P, van der Maaten E, van der Maaten-Theunissen M, Campioli M, Dobrowolska D (2014) Climate change impacts in European forests: the expert views of local observers. Ann For Sci 71:131–137. doi: 10.1007/s13595-013-0280-1

Spies TA, Giesen TW, Swanson FJ, Franklin JF, Lach D, Johnson KN (2010) Climate change adaptation strategies for federal forests of the Pacific Northwest, USA: ecological, policy, and socio-economic perspectives. Landsc Ecol 25:1185–1199. doi: 10.1007/s10980-010-9483-0

Spittlehouse DL (2005) Integrating climate change adaptation into forest management. For Chron 81:691–695

Spittlehouse DL, Stewart RB (2003) Adaption to climate change in forest management. BCJ Ecosyst Manag 4:1–11

Stafford Smith M, Horrocks L, Harvey A, Hamilton C (2011) Rethinking adaptation for a 4 degrees C world. Philos Transact A Math Phys Eng Sci 369:196–216

Stainforth DA, Allen MR, Tredger ER, Smith LA (2007) Confidence, uncertainty and decision-support relevance in climate predictions. Philos Trans R Soc A Math Phys Eng Sci 365:2145–2161. doi: 10.1098/rsta.2007.2074

Stanturf JA, Palik BJ, Dumroese RK (2014) Contemporary forest restoration: a review emphasizing function. For Ecol Manag 331:292–323

Steenberg JWN, Duinker PN, Bush PG (2011) Exploring adaptation to climate change in the forests of central Nova Scotia, Canada. For Ecol Manag 262:2316–2327. doi: 10.1016/j.foreco.2011.08.027

Stephens M, Pinkard L, Keenan RJ (2012) Plantation forest industry climate change adaptation handbook. Australian Forest Products Association, Canberra

Tacconi L, Moore PF, Kaimowitz D (2007) Fires in tropical forests—what is really the problem? Lessons from Indonesia. Mitig Adapt Strateg Glob Chang 12:55–66. doi: 10.1007/s11027-006-9040-y

Temperli C, Bugmann H, Elkin C (2012) Adaptive management for competing forest goods and services under climate change. Ecol Appl 22:2065–2077

Thackway R, Cresswell ID (1992) Environmental regionalisations of Australia: a user-oriented approach. Environmental Resources Information Network, Canberra

Thomalla F, Downing T, Spanger-Siegfried E, Han GY, Rockstrom J (2006) Reducing hazard vulnerability: towards a common approach between disaster risk reduction and climate adaptation. Disasters 30:39–48. doi: 10.1111/j.1467-9523.2006.00305.x

Thomson-Reuters (2014) Web of Science. http://thomsonreuters.com/thomson-reuters-web-of-science/ . Accessed 21 Aug 2014

Thuiller W, Albert C, Araújo MB, Berry PM, Cabeza M, Guisan A, Hickler T, Midgley GF, Paterson J, Schurr FM, Sykes MT, Zimmermann NE (2008) Predicting global change impacts on plant species’ distributions: future challenges. Perspect Plant Ecol Evol Syst 9:137–152. doi: 10.1016/j.ppees.2007.09.004

Toffler A (1970) Future shock. Bantam, Toronto

Tompkins EL, Adger WN, Boyd E, Nicholson-Cole S, Weatherhead K, Arnell N (2010) Observed adaptation to climate change: UK evidence of transition to a well-adapting society. Glob Environ Chang Hum Policy Dimens 20:627–635. doi: 10.1016/j.gloenvcha.2010.05.001

Urwin K, Jordan A (2008) Does public policy support or undermine climate change adaptation? Exploring policy interplay across different scales of governance. Glob Environ Chang Hum Policy Dimens 18:180–191. doi: 10.1016/j.gloenvcha.2007.08.002

Van Damme L (2008) Can the forest sector adapt to climate change? For Chron 84:633–634

van Dijk AIJM, Keenan RJ (2007) Planted forests and water in perspective. For Ecol Manag 251:1–9. doi: 10.1016/j.foreco.2007.06.010

Versini PA, Velasco M, Cabello A, Sempere-Torres D (2013) Hydrological impact of forest fires and climate change in a Mediterranean basin. Nat Hazards 66:609–628. doi: 10.1007/s11069-012-0503-z

Vignola R, Locatelli B, Martinez C, Imbach P (2009) Ecosystem-based adaptation to climate change: what role for policy-makers, society and scientists? Mitig Adapt Strateg Glob Chang 14:691–696. doi: 10.1007/s11027-009-9193-6

Vihervaara P, D’Amato D, Forsius M, Angelstam P, Baessler C, Balvanera P, Boldgiv B, Bourgeron P, Dick J, Kanka R, Klotz S, Maass M, Melecis V, Petrik P, Shibata H, Tang JW, Thompson J, Zacharias S (2013) Using long-term ecosystem service and biodiversity data to study the impacts and adaptation options in response to climate change: insights from the global ILTER sites network. Curr Opin Environ Sustain 5:53–66. doi: 10.1016/j.cosust.2012.11.002

von Detten R, Faber F (2013) Organizational decision-making by German state-owned forest companies concerning climate change adaptation measures. For Policy Econ 35:57–65. doi: 10.1016/j.forpol.2013.06.009

Walker B, Meyers JA (2004) Thresholds in ecological and social-ecological systems: a developing database. Ecology and Society 9 (2)

Walker B, Salt D (2012) Resilience practice: engaging the sources of our sustainability. Island Press, Washington, DC

Wang T, Campbell EM, O’Neill GA, Aitken SN (2012) Projecting future distributions of ecosystem climate niches: uncertainties and management applications. For Ecol Manag 279:128–140. doi: 10.1016/j.foreco.2012.05.034

Wellstead A, Rayner J, Howlett M (2014) Beyond the black box: forest sector vulnerability assessments and adaptation to climate change in North America. Environ Sci Pol 35:109–116. doi: 10.1016/j.envsci.2013.04.002

Westerling AL, Gershunov A, Cayan DR, Barnett TP (2002) Long lead statistical forecasts of area burned in western US wildfires by ecosystem province. Int J Wildl Fire 11:257–266. doi: 10.1071/wf02009

Westerling AL, Hidalgo HG, Cayan DR, Swetnam TW (2006) Warming and earlier spring increase western U.S. forest wildfire activity. Science 313:940–943. doi: 10.1126/science.1128834

White A, Hatcher J, Khare A, Liddle M, Molnar A, Sunderlin WD (2010) Seeing people through the trees and the carbon: mitigating and adapting to climate change without undermining rights and livelihoods. Social dimensions of climate change: equity and vulnerability in a warming world: 277–301

Wilby RL, Dessai S (2010) Robust adaptation to climate change. Weather 65:180–185. doi: 10.1002/wea.543

Williams J (2013) Exploring the onset of high-impact mega-fires through a forest land management prism. For Ecol Manag 294:4–10. doi: 10.1016/j.foreco.2012.06.030

Williams JE (2000) The biodiversity crisis and adaptation to climate change: a case study from Australia’s forests. Environ Monit Assess 61:65–74. doi: 10.1023/a:1006361917359

Williams JT (2004) Managing fire-dependent ecosystems: we need a public lands policy debate. Fire Manag Today 64:6–11

Wintle BA, Bekessy SA, Keith DA, van Wilgen BW, Cabeza M, Schroder B, Carvalho SB, Falcucci A, Maiorano L, Regan TJ, Rondinini C, Boitani L, Possingham HP (2011) Ecological-economic optimization of biodiversity conservation under climate change. Nat Clim Chang 1:355–359. doi: 10.1038/nclimate1227

Wu HX, Ying CC (2004) Geographic pattern of local optimality in natural populations of lodgepole pine. For Ecol Manag 194:177–198. doi: 10.1016/j.foreco.2004.02.017

Yousefpour R, Jacobsen JB, Meilby H, Thorsen BJ (2014) Knowledge update in adaptive management of forest resources under climate change: a Bayesian simulation approach. Ann For Sci 71:301–312. doi: 10.1007/s13595-013-0320-x

Yousefpour R, Jacobsen JB, Thorsen BJ, Meilby H, Hanewinkel M, Oehler K (2011) A review of decision-making approaches to handle uncertainty and risk in adaptive forest management under climate change. Ann For Sci 69:1–15. doi: 10.1007/s13595-011-0153-4

Zhao D, Wu S, Yin Y (2013) Responses of terrestrial ecosystems’ net primary productivity to future regional climate change in China. PLoS ONE 8:e60849. doi: 10.1371/journal.pone.0060849

Zhou GY, Wei XH, Wu YP, Liu SG, Huang YH, Yan JH, Zhang DQ, Zhang QM, Liu JX, Meng Z, Wang CL, Chu GW, Liu SZ, Tang XL, Liu XD (2011) Quantifying the hydrological responses to climate change in an intact forested small watershed in Southern China. Glob Chang Biol 17:3736–3746. doi: 10.1111/j.1365-2486.2011.02499.x

Zimmermann NE, Yoccoz NG, Edwards TC, Meier ES, Thuiller W, Guisan A, Schmatz DR, Pearman PB (2009) Climatic extremes improve predictions of spatial patterns of tree species. Proc Natl Acad Sci 106:19723–19728. doi: 10.1073/pnas.0901643106

Download references

Acknowledgments

Thanks to Linda Joyce for her comments on an earlier draft of this paper, to a number of anonymous reviewers for their thoughtful suggestions and to many colleagues that I have discussed these ideas with over the past five years.

Author information

Authors and affiliations.

Department of Forest and Ecosystem Science, The University of Melbourne, 221 Bouverie St., Carlton, VIC, 3010, Australia

Rodney J. Keenan

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Rodney J. Keenan .

Additional information

Handling Editor: Erwin Dreyer

This research was partly undertaken during the period the author was Director of the Victorian Centre for Climate Change Adaptation Research and part-funded by the Victorian Government. It was completed with support from the University of Melbourne under a Special Studies Program grant.

Rights and permissions

Open Access This article is distributed under the terms of the Creative Commons Attribution License which permits any use, distribution, and reproduction in any medium, provided the original author(s) and the source are credited.

Reprints and permissions

About this article

Cite this article.

Keenan, R.J. Climate change impacts and adaptation in forest management: a review. Annals of Forest Science 72 , 145–167 (2015). https://doi.org/10.1007/s13595-014-0446-5

Download citation

Received : 01 September 2014

Accepted : 10 December 2014

Published : 14 January 2015

Issue Date : March 2015

DOI : https://doi.org/10.1007/s13595-014-0446-5

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Climate change
  • Vulnerability
  • Forest management

Annals of Forest Science

ISSN: 1297-966X

forest management research papers

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals

Forestry articles from across Nature Portfolio

Forestry is the study and practice of the cultivation and management of woodland. Forests provide goods and services such as timber, recreation and carbon sequestration.

forest management research papers

Wildlife boost in African forests certified for sustainable logging

Is there a conservation benefit if tropical forests that are affected by logging gain certification from the Forest Stewardship Council? An analysis of the biodiversity outcomes in such tropical forests provides answers.

  • Julia E. Fa

Latest Research and Reviews

forest management research papers

How will future climate change impact prescribed fire across the contiguous United States?

  • Julia Oliveto
  • Nicholas Skowronski

forest management research papers

Impact of species composition on fire-induced stand damage in Spanish forests

  • Marina Peris-Llopis
  • Blas Mola-Yudego
  • José Ramón González-Olabarria

forest management research papers

Maximizing carbon sequestration potential in Chinese forests through optimal management

The authors show China’s forests can sequester 172.3 million tons of carbon per year in biomass by 2100, with an additional 28.1 million tons from improved management practices, but neglecting wood harvest impacts will distort long-term future projections.

  • Shirong Liu
  • Evgenios Agathokleous

forest management research papers

FSC-certified forest management benefits large mammals compared to non-FSC

Camera-trap images of 55 mammal species in 14 logging concessions in western equatorial Africa reveal greater animal encounter rates in FSC-certified than in non-certified forests, especially for large mammals and species of high conservation priority.

  • Joeri A. Zwerts
  • E. H. M. Sterck
  • Marijke van Kuijk

forest management research papers

Current inequality and future potential of US urban tree cover for reducing heat-related health impacts

  • Robert I. McDonald
  • Tanushree Biswas
  • Joseph E. Fargione

forest management research papers

LiDAR-based reference aboveground biomass maps for tropical forests of South Asia and Central Africa

  • Suraj Reddy Rodda
  • Rakesh Fararoda
  • Pierre Ploton

Advertisement

News and Comment

forest management research papers

Indian forest act faces challenge in Supreme Court

Ecologists, bureaucrats and conservationists say India’s amended Forest Conservation Act will reduce biodiversity and harm livelihoods.

  • Gayathri Vaidyanathan

The value of forests

Forests are fascinating ecosystems that have accompanied our history and are part of our collective tales. Let’s protect them!

forest management research papers

Surge in extreme forest fires fuels global emissions

Climate change and human activities have led to more frequent and intense forest blazes over the past two decades.

  • Xiaoying You

forest management research papers

Sounds of recovery: AI helps monitor wildlife during forest restoration

System aids researchers measuring biodiversity levels in Ecuador, and how people can follow basic instructions while fast asleep.

  • Benjamin Thompson
  • Shamini Bundell

forest management research papers

Assessing the scale of rubber deforestation in southeast Asia

Understanding the extent of deforestation associated with agriculturally harvested crops has implications for conservation efforts. A method to assess satellite data offers an accurate way to estimate rubber deforestation.

  • Carlos Souza Jr

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

forest management research papers

forest management research papers

Academia.edu no longer supports Internet Explorer.

To browse Academia.edu and the wider internet faster and more securely, please take a few seconds to  upgrade your browser .

  •  We're Hiring!
  •  Help Center

Forest Management

  • Most Cited Papers
  • Most Downloaded Papers
  • Newest Papers
  • Save to Library
  • Last »
  • Sustainable forestry management Follow Following
  • Forest Ecology And Management Follow Following
  • Biological Conservation Follow Following
  • Intraorganismal Heterogeneity Follow Following
  • Forest Policy Follow Following
  • Forestry Follow Following
  • Operational Research Follow Following
  • Transaction Costs Follow Following
  • Restoration Ecology Follow Following
  • Agroforestry Follow Following

Enter the email address you signed up with and we'll email you a reset link.

  • Academia.edu Publishing
  •   We're Hiring!
  •   Help Center
  • Find new research papers in:
  • Health Sciences
  • Earth Sciences
  • Cognitive Science
  • Mathematics
  • Computer Science
  • Academia ©2024

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings

Preview improvements coming to the PMC website in October 2024. Learn More or Try it out now .

  • Advanced Search
  • Journal List
  • Springer Nature - PMC COVID-19 Collection

Logo of phenaturepg

A global systematic review of forest management institutions: towards a new research agenda

Jude ndzifon kimengsi.

1 Forest Institutions and International Development (FIID) Research Group, Chair of Tropical and International Forestry, Faculty of Environmental Sciences, Technische Universität Dresden, Dresden, Germany

2 Department of Geography, The University of Bamenda, Bamenda, Cameroon

Raphael Owusu

Shambhu charmakar, gordon manu.

3 Food and Agricultural Organization (FAO), Rome, Italy

Lukas Giessen

4 Chair of Tropical and International Forestry, Faculty of Environmental Sciences, Technische Universität Dresden, Dresden, Germany

Associated Data

Globally, forest landscapes are rapidly transforming, with the role of institutions as mediators in their use and management constantly appearing in the literature. However, global comparative reviews to enhance comprehension of how forest management institutions (FMIs) are conceptualized, and the varying determinants of compliance, are lacking. And so too, is there knowledge fragmentation on the methodological approaches which have and should be prioritized in the new research agenda on FMIs.

We review the regional variations in the conceptualization of FMIs, analyze the determinants of compliance with FMIs, and assess the methodological gaps applied in the study of FMIs.

A systematic review of 197 empirically conducted studies (491 cases) on FMIs was performed, including a directed content analysis.

First, FMIs literature is growing; multi-case and multi-country studies characterize Europe/North America, Africa and Latin America, over Asia. Second , the structure-process conceptualization of FMIs predominates in Asia and Africa. Third , global south regions report high cases of compliance with informal FMIs, while non-compliance was registered for Europe/North America in the formal domain. Finally, m ixed-methods approaches have been least employed in the studies so far; while the use of only qualitative methods increased over time, the adoption of only quantitative approaches witnessed a decrease.

Future research should empirically ground informality in the institutional set-up of Australia while also valorizing mixed-methods research globally. Crucially, future research should consider multidisciplinary and transdisciplinary approaches to explore the actor and power dimensions of forest management institutions.

Supplementary Information

The online version contains supplementary material available at 10.1007/s10980-022-01577-8.

Introduction

Globally, forests are at a crossroads—characterized by rapid transformation (Garcia et al. 2020 ). For instance, Global Forest Watch estimated that forest loss around the globe reached 29.7 million hectares as of 2016, indicating a 51% increase since 2015. For tropical forests, the loss was estimated at 12 million hectares (the size of Belgium) in 2018 (Weisse and Goldman 2017 ; Garcia et al. 2020 ). While such changes are linked to natural (e.g., climate change) and human-induced drivers such as land-use change (Rounsevell et al. 2006 ; Meyfroidt and Lambin 2011 ; Aguiar et al. 2016 ; Houghton and Nassikas 2018 ), they form part of a complex transformation system—mediated by socio-economic, political, and institutional forces (Malhi et al. 2014 ). This validates the role of institutions as a key enhancing or constraining factor in determining forest resources access, use and management (Cleaver 2017 ). Institutions are viewed as highly abstract and invisible conditions in the political environment. They constitute cognitive, normative, and regulatory structures which provide stability and meaning to social behaviour. Institutions are carried across multiple vehicles, including cultures, structures and routines, operating at multiple levels of jurisdiction (Scott 1995 ). The transformation of forest landscapes at various timescales is characterized by net tropical forest loss (Geist and Lambin 2002 ; Kissinger et al. 2012 ; Song et al. 2018 ). Furthermore, regional variations in the drivers (Curtis et al. 2018 ) exist: in Latin America, transformations are largely rooted in ranching and soybean expansion (Rudel et al. 2009 ; Verburg et al. 2014 ; Tyukavina et al. 2017 ), while subsistence agriculture drives the transformation process in Africa (Hosonuma et al. 2012 ; Tyukavina et al. 2018 ). Transformations in Asia are significantly linked to industrial processes and small-holder farming (Rudel et al. 2009 ; Turubanova et al. 2018 ).

While forests are declining (Weisse and Goldman 2017 ), their roles in the resolution of global socio-ecological challenges (e.g., climate change mitigation and poverty reduction) remain unrivalled (Oldekop et al. 2020 ; Nerfa et al. 2020 ). Scholars submit that governance mechanisms, especially the role of institutions, remain primordial in shaping forests access, use, and management. For this reason, institutions—the rules of the game—continually gain relevance (Agrawal and Gupta 2005 ; Dixon and Wood 2007 ; Kimengsi et al. 2021 ). Variations exist in the way institutions are conceptualized. For instance, following the structure process dichotomy (Fleetwood 2008a , b ), institutions relate to tissues of social relations linking groups and communities (structures) and a set of rules, conventions and values, among others (processes) (Fleetwood 2008a ; Bernardi et al. 2007 ). It is, however, difficult to provide a dividing line between the processes and structures; processes (rules) guide the formation of structures, while structures, on the other hand, oversee and enforce rules (Fleetwood 2008a ; Ntuli et al. 2021 ). However, structures differ from processes in terms of their functioning; structures could represent forest management organizations as an entity, and not the rules (processes) which they produce (Ntuli et al. 2021 ). Both structures and processes are subjected to a categorization as either formal (written and codified laws, largely state driven) and informal (unwritten or uncodified rules that transcend generations) (Osei-Tutu et al . 2014 ; Yeboah-Assiamah et al. 2017 ). Furthermore, and on the basis of source, institutions could be categorized following the endogenous—exogenous dichotomy; the former relates to community-specific complex and embedded rules, while the latter denotes institutions introduced by the state and international agencies (Yeboah-Assiamah et al. 2017 ; Kimengsi et al. 2022a , b , c ). By and large, these categories of institutions exist to provide order in the midst of ‘chaos’, with regards to the sustainable management of forest resources (Beunen and Patterson 2019 ).

While forest landscapes are transforming, institutions have also been subjected to several dimensions of change. Their evolution over time manifests through formation, reformation, disintegration, and modification in several contexts, including Africa (Haller et al. 2016 ; Friman 2020 ; Kimengsi et al. 2022a , b , c ), Asia (Haapal and White 2018 ; Steenbergen and Warren 2018 ), and Latin America (Faggin and Behagel 2018 ; Gebara 2019 ). This brings to fore the notion of ephemeral, intermittent and perennial institutions (Kimengsi et al. 2021 )—borrowed from the geographic classification of streams (Gomes et al. 2020 ). Ephemeral refers to short-term stream movements (institutional arrangements), intermittent is analogous to medium-term/seasonal streams (medium-term institutional arrangements), and perennial relates to streams that flow all through—analogous to more long-term, enduring institutions (Kimengsi et al. 2021 ). Therefore, the search for perennial (enduring) institutions is top on the scientific and policy agenda (Ostrom 1990 ; Kimengsi et al. 2021 ). This is important to support the attainment of objectives such as halting forest loss and improving forest cover and species diversity (Bare et al. 2015 ; Assa 2018 ), sustaining livelihoods and economic welfare (Buchenrieder and Balgah 2013 ; Foundjem-Tita et al. 2018 ), and engendering equity and fairness in the distribution of proceeds from forest systems (Faye et al. 2017 ). However, studies on institutions and institutional change are seemingly at an impasse; it seems difficult to proceed with the framing of forward-looking research questions linked to forest management institutions (FMIs). The impasse is rooted in the largely fragmented and unstructured institutional analysis around forest settings that harbor conflicts linked to emerging and persistent resource use inequalities (Gautam et al. 2004 ; Soliev et al. 2021 ). Additionally, the multiplicity of institutional variables and the lack of a consensus on which of the methods—qualitative or quantitative—is best suited for analyzing institutions and institutional change (Kimengsi et al. 2022a , b , c ) further validate the need to surmount this impasse. In this regard, a systematic review of the global knowledge base on FMIs is imminent. Furthermore, details on the methods to prioritize in future studies further validates the need for a review. Consequently, we seek answers to the following questions: (1) How have FMIs been conceptualized and analyzed globally? (2) How varied are the (non)compliance determinants and outcomes of FMIs? (3) How can we conceptually and methodologically advance research on FMIs? To provide answers to these interrogations, we undertake a review of FMIs. The study is inspired by an earlier review conducted in the context of sub-Saharan Africa (Kimengsi et al. 2022b ).

Materials and methods

Analytical framework.

In this review, we make use of the socio-ecological co-evolution framework (Pretzsch et al. 2014 ). The framework serves as a useful theoretical fundament to enhance understanding of the dynamics around forests and rural development. While allowing for the differentiation between humans and ecological subsystems, the framework also outlines the dynamic interactions between these two systems (Berkes et al. 1998 ; Pretzsch et al. 2014 ). The socio-ecological co-evolution framework is designed to enhance comprehension of the interactions between the social system (e.g., the community of forest users), the institutions that shape them, and the ecological system—forests. These interactions occur at the interface (management segment) of the framework. Besides providing a useful analytical lens to appreciate current levels of engagement in decision-making and the enforcement of institutional provisions, it also serves as a useful framework to understand how institutional change triggers the co-evolution of both ecological and social systems. The socio-ecological co-evolution framework is informed by the earlier works of Berkes et al. ( 1998 ), which bridged the hitherto divide between social research (centred around institutions), and ecological research, which emphasized cross-scale ecosystem dynamics. Worthy of note is the fact that other frameworks exist; for instance, the socio-ecological systems (SES) framework (Ostrom 2009 ) was proposed to explain complex systems involving resource systems (forests in this case), their resource units (e.g., timber), appropriators (e.g., timber exploiters), and governance systems (e.g., forest management rules) that continually interact to produce differential outcomes (Ostrom 2009 ). It explains that socio-ecological systems are constantly subjected to change. Some of these changes are rooted in institutions and institutional change processes (Rammel et al. 2007 ; Pretzsch et al. 2014 ). The socio-ecological co-evolution framework is employed for the following reasons: (1) with rapid transformations experienced in forest landscapes across the globe, scientific and policy circles need to extend their breadth of knowledge on how to further ‘marry’ social and ecological systems in forest management. (2) Institutional change is reflected through the decisions and actions of resource users at the interface of the framework. Therefore, understanding how these changes and their determinants precipitate (non)compliance is helpful in today’s dispensation, where forests are seen as crucial in stemming the upsurge of environmental crises. (3) The outcomes associated with the myriads of institutions need to be further appreciated to inform policy actors on the orientation of future FMIs. The socio-ecological co-evolution framework (Fig.  1 ) explains how changing societal demands and choices, influenced by the institutions in place, shape the type and magnitude of societal intervention in socio-ecological systems (e.g., forests).

An external file that holds a picture, illustration, etc.
Object name is 10980_2022_1577_Fig1_HTML.jpg

Analytical framework for the systematic review.

Source: Based on Cleaver ( 2017 ), Haller et al. ( 2016 ), North ( 1990 ), Ostrom ( 1990 , 2005), and Pretzsch et al. ( 2014 ). NTFPs Non-Timber Forest Product

The review, guided by the research questions, focuses on the management phase and the social segment of the socio-ecological co-evolution framework. The management phase represents an interface—a point where management decisions under different forest categories such as plantation forests, forest reserves, community forests and landscapes in want of restoration, are implemented. Institutions and institutional change processes drive such decisions. The management operations are construed as forest-linked activities which are informed by institutions regulating timber and NTFPs exploitation, ecotourism, medicinal plants’ extraction and forest conservation. Institutional arrangements in this socio-ecological system culminate in the derivation of different management approaches, such as co-management and community-based forestry with a focus on livelihoods and conservation. The social segment of the framework focuses on the conceptualization of forest management institutions (for instance, structures vs processes, formal vs informal, and endogenous vs exogenous). This segment also captured forest management institutional compliance with an emphasis on the variations and determinants. The segment on outcomes explored the ecological, economic, socio-cultural, and political outcomes of FMIs. The framework also has a segment which explores methodological approaches employed in the study of FMIs.

Methodology

Data collection.

The systematic review approach (Nightingale 2009a , b ; Mengist et al. 2019 ) was employed in this study. Systematic reviews follow an established and standardized protocol for the search, appraisal and inclusion (or exclusion) of literature for subsequent analysis (Boell and Cecez-Kecmanovic 2015 ). This is different from the general review of literature which consists of a non-structured and highly subjective method of literature search and analysis (Kraus et al. 2020 ). The procedure was employed as follows: First a list of search terms (Appendix) was developed and used in the article search process. We targeted the following databases: Scopus, Science Direct, Google Scholar and Web of Science. Search terms such as forest management, forest governance, institutions, rules, norms, norms, laws, policies, community-based organizations, NGOs, associations, compliance, determinants, and outcomes were repeatedly employed in the search. The terms were combined with the respective regions (Africa, Asia, Australia, Europe/North America, Latin America), over a 15-year period (2006–2021). It should be noted that Europe and North America were clustered due to the observed similarity in their societal fabric and culture. We considered this timespan good enough to mirror contemporary evidence on the question of forest management institutions (FMIs). The search led to the initial identification of 920 articles. Four hundred thirty articles were identified from Web of Science, 104 from Google Scholar, 348 from Scopus and 38 from the Science Direct database. The search on Google Scholar did not produce a lot of articles. This is because grey literature was not considered during the search. Our emphasis was to derive literature which were published in internationally recognized databases. We then proceeded to deduplicate the articles—the deduplication process led to a reduction to 680 articles. Furthermore, article screening was performed with emphasis on the abstracts. This informed the decision to include or exclude the paper. In the selection, we targeted journal articles that were published in English and were empirically grounded. In cases where the abstract could not provide these details, we proceeded to review the methods and conclusions to inform inclusion (or exclusion). We excluded all grey literature during the article selection. This reduced the number of manuscripts to 197 (see Supplementary Excel Sheet), from which we derived 491 case studies; the cases were derived by considering the number of study areas that were included for analysis. We use ArcMap 10.5 to generate the map of the globe and the regions and/or countries where most of the case studies in this review paper were concentrated.

Data analysis

The articles retained were further read, and following the analytical framework (Fig.  1 ), a directed content analysis was performed (Hsieh and Shannon 2005 ). The directed content analysis began with a relevant theoretical framework—in this case, the socio-ecological coevolution framework. This framework provided a clear focus for the research questions under review. The key variables which were outlined in the framework (Fig.  1 ) informed the clustering of the data generated from the selected articles. For the selected articles, we read the abstract, methods and conclusion sections to generate data. The dataset was compiled in an excel sheet and further read; key texts which contained variables of interest were highlighted. These variables were then clustered following the established questions and themes for further analysis (Mayring 2000 ). Therefore, the highlighted texts, which contained data corresponding to the four thematic sections, were extracted from each article and organized under the main themes: conceptualization of institutions, institutional compliance, outcomes of forest management institutions and methodological approaches . We approached the conceptualization of institutions following the structure-process dimension (Fleetwood 2008a ), the formal and informal dichotomy (North 1990 ), the endogenous vs exogenous institutional lens (Kimengsi et al. 2021 ) and the state vs community-based institutional dichotomy (Ntuli et al. 2021 ). Compliance denotes the extent to which forest users adhere to the institutional provisions in their communities. This translates to forest management outcomes which could be ecological, socio-economic and even political (Haller et al. 2016 ).

These were recorded in a Microsoft Excel sheet (Artmann and Sartison 2018 ). We considered this approach appropriate, considering that software extraction might ignore salient details owing to the complex nature of institutional variables. Besides narratives and content analysis, we used descriptive statistics to report the variations across the five regions. The descriptive analysis further aided in establishing institutional compliance and its determinants, the ecological, socio-cultural, economic, and political outcomes linked to forest management institutions, and the variations in methodological approaches employed.

Attributes of reviewed papers and case studies

The review indicated that most of the articles emanate from Africa and Latin America—home to two of the world’s major forest ecosystems. This was followed by Asia. Case-wise, the study captured a total of 491 cases drawn from 99 countries across the globe (Fig.  2 ). The highest number of cases emanate from Europe/North America and Africa. This suggests that multi-case and multi-country studies have been significantly prioritized in these regions compared to single case/country studies for Asia and Australia.

An external file that holds a picture, illustration, etc.
Object name is 10980_2022_1577_Fig2_HTML.jpg

Spatial distribution of case studies on forest management institutions (2006–2021)

In parts of Central Europe, studies to explore shifts towards new governance established that recent changes in institutional arrangements result from macro-political trends and the geopolitical strategy of some states (Sergent et al. 2018 ). In the United States and Canada, forest certification led to substantial changes in practices as enterprises embraced changes in forestry, environmental, social, and economic/system practices in the realm of forest certification (Moore et al. 2012 ). In the case of Africa, a comparative study of 38 countries reported that the activities of multinational corporations are associated with differential losses in forest cover—linked to weak governance (institutions) (Assa 2018 ). The review clearly shows that while political, geostrategic and religious forces defined the institutional change process in Europe/North America, economic interests through multinational companies shaped institutional change in Africa. The review established that some of the significant countries with regards to cases include Australia, Ecuador, and Germany (21 + cases). In addition, Bolivia, Brazil, Cameroon, Ghana, and Canada, India registered between 8 and 14 cases, while the remaining countries registered between 1 and 7 cases (Fig.  2 ).

Temporal evolution of papers and cases on forest management institutions

On the whole, the literature on institutions has grown over the last 15 years(Fig.  3 ). This could be linked to renewed interests to understand governance mishaps and to engage in getting institutions (including FMIs) right . In all, while the number of publications increased from 2006, the review shows that it witnessed a decline in 2009 and 2012. The growth in the literature is possibly explained by the interest to uncover institutional ‘relicts’ (in Africa) and rising environmental challenges in Latin America (bushfires and migration). Formal forest management institutions (structure and process) for forest products have received much attention in the 2000s literature. However, significant growth was observed in the literature of the 2010s, which included both formal (international) and informal (traditional and local) institutions and concepts such as ecosystem services, sustainable forest management, farm forestry, biodiversity, REDD + , and forest certification. This classification increasingly accommodated the use of endogenous and exogenous institutions, as well as state and community-based institutions. However, both classifications have been (mis)construed to represent formal and informal institutions.

An external file that holds a picture, illustration, etc.
Object name is 10980_2022_1577_Fig3_HTML.jpg

Temporal evolution of papers on forest institutions across the globe for the past 15 years

Conceptualization of forest management institutions

From a structural dimension, institutions have been most conceptualized as structures in Asia and Africa. These predominate the informal structures where Asia and Africa account for 35% and 28%, respectively, of the review’s literature reporting on informal institutional structures. Latin America closely follows them with 22%. Literature from Asia and Africa further dominates in the classification of formal institutional structures with 28% and 25%, respectively. This is closely followed by Europe/North America (23%). Asia and Africa are ‘pace-setters” in the implementation of new forest management paradigms such as community-based forest management (Kimengsi and Bhusal, 2022 ). The introduction of these models saw the multiplication of management structures to oversee them. This explains why the literature significantly captures the structural dimension of institutions. Process-wise, literature from Asia and Africa accounts for 48% and 30%, respectively, of the literature on informal institutions, while Australia surprisingly reports none. In the formal domain, Asia, Europe/North America, and Latin America account for over 60% of the literature reporting the formal conceptualization of institutions (Table ​ (Table1 1 ).

Total number of papers (studies) = 197

Literature from Africa and Asia showed similarities in the conceptualization of institutions (Table ​ (Table2); 2 ); informal structures, for instance, chieftaincy and women groups, define and enforce processes (rules) which are conceived, for example, as taboos, beliefs, traditions, and customary rules. Studies in Cameroon and Burkina Faso in Africa report on bricolage manifestations involving formal and informal institutions (Kimengsi and Balgah 2021 ; Friman 2020 ), and the spatial variations in traditional institutions (Kimengsi et al. 2021 , 2022b ). In Europe/North America, the literature shows that informal structural institutions are conceptualized sparingly to include community leadership and inter-community forestry associations. Formally, they are reported as forest owners’ associations, political parties, protected area management, timber industry associations, resident associations, and state forest management. Informally, processes are conceived as local rules, while formally, they represent forest management policy, regulations, legal framework, local community forest governance, forest management strategies and forest marketing strategies. In Latin America, forest user groups, indigenous organizations and community management committees are frequently used in informal characterization, while labour unions, national services of protected areas, REDD + working group and Community general assemblies are used formally.

Global conceptualization of forest management institutions

Local land management rules constitute the key informal process in Latin America, while conservation laws, community-based forest policy, forest codes, forest laws, forest tenure agreements, and decentralized environmental policy appear in the formal conception of institutions. On the whole, a more diverse conceptualization of institutions (structures and processes) appear in the literature from Africa and Asia, followed by Latin America. The diversity is rooted in the diverse ethnic arrangements which characterize these regions. Africa, is the most ethnically diverse region in the world (Fearon 2003 ). This diversity accounts for the diversity in the nomenclature employed for forest management institutions—leading to the diversity in their conceptualization. With more empirical cases emanating from these settings, it is plausible to suggest that more in-depth and varied analysis about forest management institutions has been explored in these settings. Furthermore, the plethora of governance challenges in the management of natural resources (forest in this case) which is associated with such settings further explains the multifarious typification of institutions. In another dimension, structures and processes are surprisingly only conceptualized formally in the context of Australia—suggesting a significant drift away from informality to the pursuit of more formal, state-sanctioned institutions.

Compliance with forest management institutions

From the review, Africa, Asia, and Latin America report the highest cases of compliance in the informal institutional set-up. These settings have had a history linked to traditional institutions which were made to interact with colonially shaped institutions during their history. However, some degree of closeness to cultural institutions could be reported for these regions. The existence of compliance in the literature for Africa, Asia and Latin America is enough pointer to the multiplicity of institutional structures and processes which require monitoring against (non)compliance. Additionally, the interaction between formal and informal institutions, including the fallouts of colonial influence, led to the multiplicity of institutions. This possibly explains why compliance predominates the literature in the three regions. In Africa, for instance, pre-colonial types of resource use included the royal hunting preserves of the amaZulu and amaSwati people, and the kgotla system of land management practiced by the Batswana people (Ghai 1992 ; Fabricius 2004 ). Further, the making of access and use rules for natural resources in Mali (Moorehead 1989 ) and Botswana (Ostrom 1990 ), all indicate how endogenous cultural institutions shaped forest use. Likewise, Khasi, Garo and Jaintia tribes in Meghalaya of India, and traditional customary organization “Lembaga Adat” in Indonesia have not only conserved forest resources but also ensured its capacity to deliver ecosystem goods and services in sustainable manner (Mehring et al. 2011 ; Tiwari et al. 2013 ). Europe/North America registered few studies on informal institutional compliance, while this was non-existent for Australia—apparently due to the non-reported case of informal arrangements. Regarding non-compliance, articles from Latin America reported the highest case of non-compliance. This could be explained by the progressive decline in the informal institutions due to globalization and market forces which seemed to have permeated communities around the Amazon (Blundo-Canto et al. 2020 ). In the formal domain, non-compliance was significantly registered for Europe/North America, Africa, and Latin America (Fig.  4 ).

An external file that holds a picture, illustration, etc.
Object name is 10980_2022_1577_Fig4_HTML.jpg

Global statistics of institutional compliance and non-compliance by regions. Left chart—papers ( N  = 197, n for compliance = 133, n for non-compliance = 64, Right chart—Cases ( N  = 491, n for compliance = 362, n for non-compliance = 129). Note Some papers reported compliance/compliance to formal and informal institutions simultaneously

It is important to note that some of the evoked reasons behind (non)compliance still require further investigation. For example, significant contextual variations in peoples’ attitudes and adherence to forest-sector institutions and governance in Asia, Africa, Latin America, and North America are directly linked to the disparities of key underlying and broader factors such as institutional models and policy frameworks for decentralization (Shackleton et al. 2002 ; Ribot 2003 ; Larson 2012 ; Mustalahti et al. 2020 ). Thus, a broader focus on institutional factors—rather than isolated reasons—is important to sufficiently explain (non-)compliance dynamics.

Forest management Institutional compliance determinants

From the analysis, political and economic factors were recurrent in the literature as key forces that influence institutional compliance. For instance, in Europe/North America, Latin America and Australia, political factors significantly influenced compliance (Table ​ (Table3). 3 ). Some of these key determinants include conflicts, policy enforcement, power relations and governance structure (Latin America) and actor network, policy development, and governance structure (Europe/North America). Politics and geostrategy contributed to defining natural resource (forest in this case) institutions in Europe and North America. However, in Latin America, the rise in challenges linked to migration and bush fires also stand as key determinants of forest management institutional compliance. Economic factors in Europe/North America and Africa determined compliance. In Africa, economic factors linked to private enterprises and market-based mechanisms significantly featured in the literature as determinants of institutional compliance. For instance, aspects linked to donor income/investment and material aid, community forest expenditure and benefits, poverty were common. In Europe/North America, the economic viability of forest land use, forest certification, amongst others, were reported in the articles. In Nigeria, economic incentives (incomes) from farming activities, NTFPs use and non-traditional employment shaped compliance (Ezebilo 2011 ). On the whole, ecological, socio-cultural and demographic factors did not significantly explain compliance with forest management institutions. The case of ecological determinants in surprising given the litany of ecological campaigns which have been introduced in Africa (e.g., Leventon et al. 2014 ; Senganimalunje et al. 2016 ), Asia (Gilani et al. 2017 ) and Latin America (Entenmann and Schmitt 2013 ; Kowler et al. 2020 ) for instance, to foster conservation. Furthermore, socio-cultural diversity as viewed in Africa warrants some diversity in the way people adhere to institutions—both formal and informal. In Tanzania, trust in institutions was a significant predictor of participation intensity of the households in forest management (Luswaga and Nuppenau 2020 ). However, in Europe, differences in the attitudes of actors with regard to pursuing sustainable development significantly shaped compliance with forest management institutions (Jankovska et al. 2010 ).

Forest management institutional compliance determinants to by regions

Forest management institutional outcomes

The review indicates that political outcomes were the most significant for Europe/North America, followed by Latin America and Australia (Table ​ (Table4). 4 ). Some key political outcomes included policy fragmentation, market formation failures, and reduced legitimacy of FSC certification (Europe/North America). This is understandable, considering that political and geostrategic forces were key determinants of institutional compliance. In Latin America,the setting up of provincial regulations which undermine enforcement of forest regime, rule breaking, challenges with the day-to-day operational institutions, inequitable benefit-sharing mechanism; the absence of law enforcement on sustainability of and access to non-wood forest products were common. Rule breaking is potentially triggered by increasing in-migration and the upsurge of bushfires. Bottazzi et al. ( 2014 ) showed how incentive-based systems of institutions facilitated the allocation and use of funds in REDD + programmes. In all these, deforestation persisted in the midst of lost and/or bypassed institutions (Carvalho et al. 2019 ). In Australia, divergent views characterized the seeking of solutions to enhance inter-departmental and inter-municipal coordination (Ordóñez et al. 2020 ). Positive ecological outcomes were significantly reported for Africa (forest or biodiversity protection/conservation, improved forest condition and surface water quality, sustainable forest or ecosystem management, planting of timber and fruit trees) and Latin America (fostering forest conservation, stabilization and/or decrease of deforestation, sustainable forest management). Furthermore, Europe/North America, Africa and Asia respectively reported positive economic outcomes linked to the generation of net monetary gains from parks, and from the wood harvesting and marketing (Europe/North America), higher incomes derived from certification, profits derived under community forestry, and the augmentation of household cash income (Africa). In Asia, studies report the positive outcomes linked to the forests’ substantial contribution to local livelihoods and income (Muhammed et al. 2008 ; Harada and Wiyono 2014 ; Barnes and van Laerhoven 2015 ).

Forest management institutional outcomes by regions

Some case studies have multiple institutional outcomes

Australia witnessed the most negative ecological outcomes. For instance, regional forest agreements were characterized by poor governance, leading to failures in biodiversity protection and ecosystem maintenance. This further precipitated the over-commitment of forest resources to wood production (Lindenmayer 2018 ). In Latin America, significant deforestation was observed for the Guarayos Indigenous Territory from 2000 to 2017—primarily driven by agricultural commodity production (He et al. 2019 ), while in Africa, Garekae et al. ( 2020 ) reported forest and wildlife decline in Botswana, linked to sectoral bias. Furthermore, the articles reported significant negative economic outcomes for Latin America, Asia, and Africa. Some of the reported outcomes include financial resource decline (Latin America), the timber-centric and market-oriented nature of community forests (Asia) (Bhusal et al. 2020 ), and the manifestations of elite capture in Africa. In Malawi, for instance, co-management programmes did not lead to positive outcomes, i.e., community organization, forest access, forest product availability and commercialization of forest products (Senganimalunje et al. 2016 ). On the whole socio-cultural outcomes were prevalent in Australia, Europe/North America, and Asia. Here, reported issues were linked to public acceptance of plantation policy, the improvement in communication of forest owners' associations and increased reliance on informal relationships. A case in point is linked to forest policy in Australia which led to several negative social impacts, including uncertainty, perceived injustice, and financial stress (Loxton et al. 2014 ). In the case of Asia, it was linked to inequalities among local actors, demographic changes and transformations in local social structures, gender inequality, successful collaboration between NGOs and community-based organizations , conflicts between communities and state forest enterprises (Adhikari and Lovett 2006 ; Barnes and van Laerhoven 2015 ).

Methodological approaches

The analysis reveals that globally mixed methods approaches have been least prioritized in the study of forest management institutions. For instance, between 2006 and 2021, we observed a growing trend in the application of qualitative methods; only 2 articles were reported in 2007, while this peaked to 99 in 2021. However, the use of quantitative and mixed methods approaches was significantly lower (Fig.  5 ). Considering the intricacies linked to the study of institutions, the prioritization of qualitative approaches is understandable. However, with growing interest in employing more robust data collection and analysis methods, it is only germane to report that studies have not prioritized mixed methods approaches so far (Malina et al. 2011 ; Karolina et al. 2021 ).

An external file that holds a picture, illustration, etc.
Object name is 10980_2022_1577_Fig5_HTML.jpg

Cumulative distribution of paper and adoption of methods

A slight increase in the application of mixed methods approaches is observed (Table ​ (Table5) 5 ) from 22% between 2006 and 2010 to 26% between 2016 and 2021, while there was a progressive decline in the sole application of quantitative methods from 37 to 23% within this time period. A slight decrease is also observed for qualitative methods, from 54 to 51% between 2011 and 2021.

Methods employed over time in the study of forest management institutions (% in parenthesis)

On the whole, while the use of only qualitative methods in the study of forest management institutions increased over time, the adoption of only quantitative approaches witnessed a decrease (Table ​ (Table5). 5 ). Studies in Africa have largely prioritized the sole application of quantitative methods, as 42% of the papers reported this approach (Table ​ (Table6), 6 ), while mixed methods (25%) were least prioritized. This could be linked to the growing ‘quantification revolution’ in research across the region. While quantitative analysis provides some pointers to institutional questions, they hide significant intricacies which could be revealed by solely qualitative or, better still, mixed methods analytical approaches. In Latin America, however, a significant proportion of the studies (40%) employed solely qualitative methods, followed by mixed methods (35%) and then quantitative methods (26%). Likewise, the highest proportion (58%) of studies draw from qualitative methods in Asia and Australia, whereas mixed-methods were prioritized as the second highest (37%) in Asia and the least (10%) in Australia (Table ​ (Table6 6 ).

Methods used by different continents (percentage in parenthesis)

On the whole, while studies in Africa employed more of only quantitative methods over qualitative ones, research on forest management institutions in Europe and North America prioritized only qualitative methods over only quantitative ones. In North America/Europe, 73% of the studies employed the qualitative approach, followed by quantitative (19%). The least employed approach is mixed-methods, as only 8% used this approach (Table ​ (Table6). 6 ). Overall, qualitative methods have been significantly employed globally except in Africa, while mixed methods were the least adopted in all regions except Asia and Latin America.

Perspectives on the conceptual and methodological advancement of research on FMIs

The literature so far presents a fragmented conceptualization of forest management institutions. For instance, institutions are broadly categorized as formal or informal on the one hand and as exogenous and endogenous on the other hand. This is based on the premise that not all endogenous institutions are informal institutions. A more detailed conceptualization which captures the formal and informal dimension, including the endogenous and exogenous categorization, is helpful to advance theoretical developments in the field of institutions in relation to forest management. Additionally, institutions seem to exhibit stream-like attributes; an approach which further conceptualizes them as ephemeral (very short-term arrangements made by forest actors to facilitate forest resource use conflict minimization), partially enduring ( arrangements that temporarily become a norm but fizzle out as new actors take over (Kimengsi et al. 2022b ); and enduring ( institutions are either codified (formal) and/or take the status of customs and values which transcend several generations (Ostrom 1990 ). In both cases, empirical studies geared towards establishing these proposed conceptual approaches are needed. Future research also needs to advance the “marriage” between actors and institutions.

From a methodological standpoint, further studies should prioritize methods based reviews (Palmatier et al. 2018 ) to enable researchers synthesize in detail, the design and instruments used so far, the approaches employed in data collection approaches and pros and cons linked to the methods employed. This will further inform the application of methodological approaches and instruments for future empirical studies on forest management institutions. Also, multi-country studies, employing mixed-methods approaches are needed to analyze institutions in forest use and management.

Review Limitations

This review provides an initial synthesis of the literature on forest management institutions from a global perspective. It is helpful in the identification of region-specific research needs in the ever-evolving field of institutions. A couple of limitations could be raised: Firstly, the conceptual analysis of institutions does not incorporate the exogenous versus endogenous dichotomy. With growing interest to further explore the typology and source of institutions, including whether management outcomes are a function of more endogenous or exogenous institutional arrangements (Kimengsi et al. 2022b ), future reviews and empirical studies should incorporate this dimension. Secondly, the regional clustering of institutions might shade details linked to how institutions are conceptualized and the outcomes they effectively produce. Although case studies are used, it is not possible in a single review to derive all these conceptual details, which might vary even within regions. Taking Africa, for instance, diversity in the region’s culture requires country-specific analysis of institutions. Latin America’s diversity precipitates ‘institutional shopping’ (Wartmann et al. 2016 Thirdly, institutions do not operate in isolation—therefore an actor-centred/power dimension is required to better appreciate institutional arrangements (Giessen et al. 2014 ; Ongolo et al. 2021 ). Therefore, a review of the actor and power dimensions of institutions is required to inform subsequent empirical studies. Fourthly, while the paper reports on compliance, the level of compliance is not reported, and the factors which militate for or against compliance. Fifthly, we selected articles which were exclusively published in English language and indexed in certain data bases. In doing so, we ignored papers which might have been published in French, Spanish, Amharic, Kiswahili, Nepali and other languages; such articles might have provided further compelling details on the region-specific dynamics of forest management institutions. We call for subsequent reviews to aim at valorizing such studies.

The current socio-ecological outcomes linked to the upsurge of pandemics (e.g. COVID-19) further justify the need to pay more attention to the management of forests and forest resources (Tollefson 2020 ; Saxena et al. 2021). These details, which vary over space and time, and may potentially assume a different dimension under the current COVID-19 scenario (Saxena et al. 2021), require extensive review and further empirical grounding. When pandemic prevention hinges on forest management to some extent, it is imperative to further explore the role of FMIs. Further reviews could emphasize the extent of compliance and the conditions under which (non)compliance prevails in the context of pandemics. Additionally, institutional change which is triggered by health crises (e.g., pandemics) still needs to be further established.

Finally, our review of the methods focused on providing a snapshot of the approaches, following the broad categorization of qualitative, quantitative and methods. This does not provide details on the specific qualitative methods employed (e.g., key informant interviews, participant observations, vignettes, focus group discussions). Future methods-based reviews should consider these.

To define conceptual and methodological pathways for future studies on forest management institutions (FMIs), this study undertakes a systematic review of the literature on FMIs using 197 papers (491 cases). From the study, the following conclusions are plausible: Firstly, while forest management institutions literature has witnessed a growth, this is most significant in Africa and Latin America. Secondly, the structure-process conceptualization of institutions (formal and informal) predominates in Asia and Africa. Process-wise, studies from Australia surprisingly did not report on a single process-linked institution. This merits further studies which pays attention to the identification of such institutions. The literature also reports on the drift away from informality to the pursuit of more formal, state-sanctioned institutional arrangements in Australia. Thirdly , global south regions—Africa, Asia, and Latin America—report the highest cases of compliance in the informal institutional set-up, while non-compliance was significantly registered for Europe/North America in the formal domain. Fourthly , politico-economic factors significantly influence institutional compliance in Europe/North America, while economic factors shape compliance in Africa. On the whole, ecological, socio-cultural, and demographic factors were reported to less significantly explain compliance with forest management institutions (FMIs). Fifthly , while forest management institutions in Europe/North America significantly contributed to determining politico-economic outcomes, those in Africa and Latin America contributed to positive ecological and negative economic outcomes. Finally, mixed methods approaches have been least prioritized in the study of forest management institutions; in Africa, the sole application of quantitative methods was prioritized. Future research needs to (1) extend the conceptualization of institutions, (2) increase multi-case and multi-country studies on FMIs especially for Asia and Australia, (3) empirically ground informality in the institutional set up of forest management in Australia, (4) establish in detail, the extent of (non)compliance, their spatio-temporal variations, and determinants, and (5) valorize the application of mixed-methods approaches in the study of FMIs across the globe.

Below is the link to the electronic supplementary material.

Open Access funding enabled and organized by Projekt DEAL. This research was funded by the Deutsche Forschungsgemeinschaft (DFG)—Projektnummer (437116427), Grant ID: F-010300-541-000-1170701.

Publisher's Note

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Contributor Information

Jude Ndzifon Kimengsi, Email: [email protected]_eduj , Email: [email protected] .

Raphael Owusu, Email: moc.oohay@usuwo60leahpar .

Shambhu Charmakar, Email: [email protected] .

Gordon Manu, Email: [email protected] .

Lukas Giessen, Email: [email protected] .

  • Agrawal A, Gupta K. Decentralization and participation: the governance of common pool resources in Nepal’s Terai. World Dev. 2005; 33 (7):1101–1114. doi: 10.1016/j.worlddev.2005.04.009. [ CrossRef ] [ Google Scholar ]
  • Aguiar APD, Vieira ICG, Assis TO, Dalla-Nora EL, Toledo PM, Santos-Junior RA, Batistella M, Coelho AS, Savaget EK, Aragaõ LEOC, et al. Land use change emission scenarios: anticipating a forest transition process in the Brazilian Amazon. Glob Change Biol. 2016; 22 :1821–1840. doi: 10.1111/gcb.13134. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Adhikari B, Lovett JC. Institutions and collective action: does heterogeneity matter in community-based resource management? J Dev Stud. 2006; 42 (3):426–445. doi: 10.1080/00220380600576201. [ CrossRef ] [ Google Scholar ]
  • Artmann M, Sartison K. The role of urban agriculture as a nature-based solution: a review for developing a systemic assessment framework. Sustainability. 2018; 10 (6):1937. doi: 10.3390/su10061937. [ CrossRef ] [ Google Scholar ]
  • Assa BSK. Foreign direct investment, bad governance and forest resources degradation: evidence in Sub-Saharan Africa. Economia Politica. 2018; 35 (1):107–125. doi: 10.1007/s40888-017-0086-y. [ CrossRef ] [ Google Scholar ]
  • Bare M, Kauffman C, Miller DC. Assessing the impact of international conservation aid on deforestation in sub-Saharan Africa. Environ Res Lett. 2015; 10 (12):125010. doi: 10.1088/1748-9326/10/12/125010. [ CrossRef ] [ Google Scholar ]
  • Barnes C, van Laerhoven F. Making it last? Analysing the role of NGO interventions in the development of institutions for durable collective action in Indian community forestry. Environ Sci Policy. 2015; 53 :192–205. doi: 10.1016/j.envsci.2014.06.008. [ CrossRef ] [ Google Scholar ]
  • Berkes F, Colding J, Folke C, editors. Linking social and ecological systems. Management practices and social mechanisms for building resilience. Cambridge: Cambridge University Press; 1998. [ Google Scholar ]
  • Bernardi F, Gonzalez JJ, Requena M. The sociology of social structure. In: Bryant B, Peck D, editors. 21st Century sociology: a reference handbook. Newbury: Sage; 2007. pp. 162–170. [ Google Scholar ]
  • Beunen R, Patterson JJ. Analysing institutional change in environmental governance: exploring the concept of ‘institutional work’ J Environ Plan Manage. 2019; 62 (1):12–29. doi: 10.1080/09640568.2016.1257423. [ CrossRef ] [ Google Scholar ]
  • Bhusal P, Karki P, Kimengsi JN. Timber distribution dynamics in scientifically managed community forests: learning from Nepal. Forests. 2020; 11 (10):1032. doi: 10.3390/f11101032. [ CrossRef ] [ Google Scholar ]
  • Blundo-Canto G, Cruz-Garcia GS, Talsma EF, Francesconi W, Labarta R, Sanchez-Choy J, et al. Changes in food access by mestizo communities associated with deforestation and agrobiodiversity loss in Ucayali, Peruvian Amazon. Food Secur. 2020; 12 (3):637–658. doi: 10.1007/s12571-020-01022-1. [ CrossRef ] [ Google Scholar ]
  • Boell SK, Cecez-Kecmanovic D. On being ‘systematic’ in literature reviews. In: Willcocks LP, Sauer C, Lacity MC, editors. Formulating Research Methods for Information Systems. London: Palgrave Macmillan; 2015. [ Google Scholar ]
  • Bottazzi P, Crespo D, Soria H, Dao H, Serrudo M, Benavides JP, et al. Carbon sequestration in community forests: trade-offs, multiple outcomes and institutional diversity in the Bolivian Amazon. Dev Chang. 2014; 45 (1):105–131. doi: 10.1111/dech.12076. [ CrossRef ] [ Google Scholar ]
  • Buchenrieder G, Balgah RA. Sustaining livelihoods around community forests: What is the potential contribution of wildlife domestication? J Modern Afr Stud. 2013; 51 (1):57–84. doi: 10.1017/S0022278X12000596. [ CrossRef ] [ Google Scholar ]
  • Carvalho WD, Mustin K, Hilário RR, Vasconcelos IM, Eilers V, Fearnside PM. Deforestation control in the Brazilian Amazon: a conservation struggle being lost as agreements and regulations are subverted and bypassed. Perspect Ecol Conserv. 2019; 17 (3):122–130. [ Google Scholar ]
  • Cleaver F. Development through bricolage: rethinking institutions for natural resource management. London: Routledge; 2017. [ Google Scholar ]
  • Curtis PG, Slay CM, Harris NL, Tyukavina A, Hansen MC. Classifying drivers of global forest loss. Science. 2018; 361 :1108–1111. doi: 10.1126/science.aau3445. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Dixon AB, Wood AP. Local institutions for wetland management in Ethiopia: sustainability and state intervention. In: van Koppen B, Giordano M, Butterworth J, editors. Community-based water law and water resource management reform in developing countries. Wallingford: CABI International; 2007. pp. 130–145. [ Google Scholar ]
  • Entenmann SK, Schmitt CB. Actors’ perceptions of forest biodiversity values and policy issues related to REDD+ implementation in Peru. Biodivers Conserv. 2013; 22 (5):1229–1254. doi: 10.1007/s10531-013-0477-5. [ CrossRef ] [ Google Scholar ]
  • Ezebilo EE. Local participation in forest and biodiversity conservation in a Nigerian rain forest. Int J Sust Dev World. 2011; 18 (1):42–47. doi: 10.1080/13504509.2011.544389. [ CrossRef ] [ Google Scholar ]
  • Fabricius C. Historical background to community-based natural resource management. In: Fabricius C, Koch E, editors. Rights, Resources and Rural Development Community-based Natural Resource Management in Southern Africa. Oxfordshire, New York: Earthscan; 2004. [ Google Scholar ]
  • Faggin J, Behagel J. Institutional bricolage of sustainable forest management implementation in rural settlements in Caatinga biome. Brazil. International Journal of the Commons. 2018; 12 (2):275–299. doi: 10.18352/ijc.872. [ CrossRef ] [ Google Scholar ]
  • Faye P, Haller T, Ribot R. Shaping rules and practice for more justice? Local conventions and local resistance in eastern Senegal. Hum Ecol. 2017; 8 (2017):1–11. [ Google Scholar ]
  • Fearon JD. Ethnic structure and cultural diversity by country. J Econ Growth. 2003; 8 (June):195–222. doi: 10.1023/A:1024419522867. [ CrossRef ] [ Google Scholar ]
  • Fleetwood S. Institutions and social structures. J Theory Soc Behav. 2008; 38 :30021–38308. doi: 10.1111/j.1468-5914.2008.00370.x. [ CrossRef ] [ Google Scholar ]
  • Fleetwood S. Structure, institution, agency, habit and reflexive deliberation. J Inst Econ. 2008; 4 (2):183–203. [ Google Scholar ]
  • Foundjem-Tita D, Duguma LA, Speelman S, Piabuo SM. Viability of community forests as social enterprises. Ecol Soc. 2018 doi: 10.5751/ES-10651-230450. [ CrossRef ] [ Google Scholar ]
  • Friman J. Gendered woodcutting practices and institutional bricolage processes: the case of woodcutting permits in Burkina Faso. For Policy Econ. 2020; 111 :102045. doi: 10.1016/j.forpol.2019.102045. [ CrossRef ] [ Google Scholar ]
  • Garcia CA, Savilaakso S, Verburg RW, Gutierrez V, Wilson SJ, Krug CB, Sassen M, Robinson BE, Moersberger H, Naimi B, Rhemtulla JM, Dessard H, Gond V, Vermeulen C, Trolliet F, Oszwald J, Quétier F, Pietsch SA, Bastin JF, Dray A, Araújo MB, Ghazoul J, Waeber PO. The global forest transition as a human affair. One Earth. 2020; 2 (5):417–428. doi: 10.1016/j.oneear.2020.05.002. [ CrossRef ] [ Google Scholar ]
  • Gautam AP, Shivakoti GP, Webb EL. A review of forest policies, institutions, and changes in the resource condition in Nepal. Int for Rev. 2004; 6 (2):136–148. [ Google Scholar ]
  • Gebara MF (2019) Understanding institutional bricolage: what drives behavior change towards sustainable land use in the Eastern Amazon? Int J Commons 13(1)
  • Geist HJ, Lambin EF. Proximate causes and underlying driving forces of tropical deforestation. Bioscience. 2002; 52 :143–150. doi: 10.1641/0006-3568(2002)052[0143:PCAUDF]2.0.CO;2. [ CrossRef ] [ Google Scholar ]
  • Ghai D (1992) Conservation, livelihood and democracy: social dynamics of environmental changes in Africa. Discussion Paper 33, United Nations Research Institute for Social Development. http://www.unrisd.org/unrisd/website/document.nsf/(httpPublications)/ . Accessed 10 Sept 2021
  • Giessen L, Krott M, Möllmann T. Increasing representation of states by utilitarian as compared to environmental bureaucracies in international forest and forest-environmental policy negotiations. Forest Policy Econ. 2014; 38 :97–104. doi: 10.1016/j.forpol.2013.08.008. [ CrossRef ] [ Google Scholar ]
  • Gilani HR, Yoshida T, Innes JL. A collaborative forest management user group's perceptions and expectations on REDD+ in Nepal. Forest Policy Econ. 2017; 80 :27–33. doi: 10.1016/j.forpol.2017.03.004. [ CrossRef ] [ Google Scholar ]
  • Gomes PIA, Wai OWH, Dehini GK. Vegetation dynamics of ephemeral and perennial streams in mountainous headwater catchments. J Mater Sci. 2020; 17 :1684–1695. [ Google Scholar ]
  • Haapal J, White P. Development through bricoleurs: portraying local personnel’s role in implementation of water resources development in rural Nepal. Water Alternat. 2018; 11 (3):979–998. [ Google Scholar ]
  • Harada K, Wiyono (2014) Certification of a community-based forest enterprise for improving institutional management and household income: a case from Southeast Sulawesi, Indonesia. Small-Scale For 13(1):47–64
  • Haller T, Acciaioli G, Rist S. Constitutionality: conditions for crafting local ownership of institution-building processes. Soc Nat Resour. 2016; 29 (1):68–87. doi: 10.1080/08941920.2015.1041661. [ CrossRef ] [ Google Scholar ]
  • He Y, Baldiviezo JP, Agrawal A, Candaguira V, Perfecto I. Guardians of the forests: how should an indigenous community in eastern bolivia defend their land and forests under increasing political and economic pressures? Case Stud Environ. 2019; 3 :1–14. doi: 10.1525/cse.2019.sc.946307. [ CrossRef ] [ Google Scholar ]
  • Hosonuma N, Herold M, De Sy V, De Fries RS, Brockhaus M, Verchot L, Angelsen A, Romijn E. An assessment of deforestation and forest degradation drivers in developing countries. Environ Res Lett. 2012; 7 :044009. doi: 10.1088/1748-9326/7/4/044009. [ CrossRef ] [ Google Scholar ]
  • Houghton RA, Nassikas AA. Negative emissions from stopping deforestation and forest degradation, globally. Glob Change Biol. 2018; 24 :350–359. doi: 10.1111/gcb.13876. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Hsieh HF, Shannon SE. Three approaches to qualitative content analysis. Qual Health Res. 2005; 15 (9):1277–1288. doi: 10.1177/1049732305276687. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Jankovska ILZE, Straupe INGA, Panagopoulos THOMAS. Professionals awareness in promotion of conservation and management of urban forests as green infrastructure of Riga, Latvia. WSEAS Trans Environ Dev. 2010; 6 (8):614–623. [ Google Scholar ]
  • Karolina V, Alif M, Sudharni S. The advantages and disadvantages of quantitative and qualitative approach for investigating washback in English language testing. EDUKATIF. 2021; 3 (5):2299–2310. [ Google Scholar ]
  • Kimengsi JN, Abam CE. Forje GW (2021) Spatio-temporal analysis of the ‘last vestiges’ of endogenous cultural institutions: implications for Cameroon’s protected areas. GeoJournal. 2021 doi: 10.1007/s10708-021-10517-z. [ CrossRef ] [ Google Scholar ]
  • Kimengsi JN, Balgah RA. Colonial hangover and institutional bricolage processes in forest use practices in Cameroon. Forest Policy Econ. 2021; 125 :102406. doi: 10.1016/j.forpol.2021.102406. [ CrossRef ] [ Google Scholar ]
  • Kimengsi JN, Mairomi HW. COVID-19 and natural resource use practices in Cameroon. In: Akumbu PW, Nzweundji JG, editors. Responding to disease outbreak in Cameroon: lessons from COVID-19. Köln: Rüdiger Köppe; 2021. [ Google Scholar ]
  • Kimengsi JN and Bhusal P (2022) Community forestry governance: lessons for Cameroon and Nepal. Soc Nat Resour 35(4):447–464
  • Kimengsi JN, Grabek J, Giessen L, Balgah RA, Buchenrieder G (2022a) Forest management institutions and actor-centered conflicts in sub-Saharan Africa: contemporary realities and future avenues for research and policy. Forest Policy Econ 144:102846
  • Kimengsi JN, Mukong AK, Giessen L, Pretzsch J. Institutional dynamics and forest use practices in the Santchou Landscape of Cameroon. Environ Sci Policy. 2022; 128 (2022):68–80. doi: 10.1016/j.envsci.2021.11.010. [ CrossRef ] [ Google Scholar ]
  • Kimengsi JN, Owusu R, Djenontin INS, Pretzsch J, Giessen L, Buchenrieder G, Pouliot M, Acosta AN. What do we (not) know on forest management institutions in sub-Saharan Africa? A regional comparative review. Land Use Policy. 2022; 114 :105931. doi: 10.1016/j.landusepol.2021.105931. [ CrossRef ] [ Google Scholar ]
  • Kissinger G, Herold M, De Sy V (2012) Drivers of deforestation and forest degradation: a synthesis report for REDD+ policymakers (Lexeme Consulting). https://www.cifor.org/knowledge/publication/5167/ . Accessed 18 Sept 2021
  • Kowler L, Kumar Pratihast A, Ojeda P, del Arco A, Larson AM, Braun C, Herold M. Aiming for sustainability and scalability: community engagement in forest payment schemes. Forests. 2020; 11 (4):444. doi: 10.3390/f11040444. [ CrossRef ] [ Google Scholar ]
  • Kraus S, Breier M, Dasí-Rodríguez S. The art of crafting a systematic literature review in entrepreneurship research. Int Entrep Manag J. 2020; 16 :1023–1042. doi: 10.1007/s11365-020-00635-4. [ CrossRef ] [ Google Scholar ]
  • Larson AM (2012) Democratic decentralization in the forestry sector: lessons learned from Africa, Asia and Latin America. In: The politics of decentralization. Routledge, London, pp 46–76
  • Leventon J, Kalaba FK, Dyer JC, Stringer LC, Dougill AJ. Delivering community benefits through REDD+: Lessons from joint forest management in Zambia. Forest Policy Econ. 2014; 44 :10–17. doi: 10.1016/j.forpol.2014.03.005. [ CrossRef ] [ Google Scholar ]
  • Lindenmayer DB. Flawed forest policy: flawed regional forest agreements. Aust J Enviro Manag. 2018; 25 (3):258–266. doi: 10.1080/14486563.2018.1466372. [ CrossRef ] [ Google Scholar ]
  • Loxton E, Schirmer J, Kanowski P. Social impacts of forest policy changes in Western Australia on members of the natural forest industry: implications for policy goals and decision-making processes. Forestry. 2014; 87 (3):363–376. doi: 10.1093/forestry/cpu011. [ CrossRef ] [ Google Scholar ]
  • Luswaga H, Nuppenau EA. Participatory forest management in West Usambara Tanzania: what is the community perception on success? Sustainability. 2020; 12 (3):921. doi: 10.3390/su12030921. [ CrossRef ] [ Google Scholar ]
  • Malhi Y, Gardner TA, Goldsmith GR, Silman MR, Zelazowski P. Tropical forests in the Anthropocene. Annu Rev Environ Resour. 2014; 39 :125–159. doi: 10.1146/annurev-environ-030713-155141. [ CrossRef ] [ Google Scholar ]
  • Malina MA, Nørreklit HSO, Selto FH. Lessons learned: advantages and disadvantages of mixed method research. Qual Res Account Manag. 2011; 8 :59–71. doi: 10.1108/11766091111124702. [ CrossRef ] [ Google Scholar ]
  • Mayring P (2000) Qualitative content analysis. Forum 1(2)
  • Mehring M, et al. Local institutions: regulation and valuation of forest use-evidence from Central Sulawesi, Indonesia. Land Use Policy. 2011; 28 (4):736–747. doi: 10.1016/j.landusepol.2011.01.001. [ CrossRef ] [ Google Scholar ]
  • Mengist W, Soromessa T, Legese G. Method for conducting systematic literature review and meta-analysis for environmental science research. MethodsX. 2019; 7 :100777. doi: 10.1016/j.mex.2019.100777(accessedon19.09.2021). [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Meyfroidt P, Lambin EF. Global forest transition: prospects for an end to deforestation. Annu Rev Environ Resour. 2011; 36 :343–371. doi: 10.1146/annurev-environ-090710-143732. [ CrossRef ] [ Google Scholar ]
  • Moore SE, Cubbage F, Eicheldinger C. Impacts of forest stewardship council (FSC) and sustainable forestry initiative (SFI) forest certification in North America. J Forest. 2012; 110 (2):79–88. doi: 10.5849/jof.10-050. [ CrossRef ] [ Google Scholar ]
  • Moorehead R. Changes taking place in common-property resource management in the Inland Niger Delta of Mali. In: Berkes F, editor. Common property resources. London: Belhaven; 1989. pp. 256–272. [ Google Scholar ]
  • Muhammed N, Koike M, Haque F. Forest policy and sustainable forest management in Bangladesh: an analysis from national and international perspectives. New for. 2008; 36 (2):201–216. doi: 10.1007/s11056-008-9093-8. [ CrossRef ] [ Google Scholar ]
  • Mustalahti I, Gutiérrez-Zamora V, Hyle M, Devkota BP, Tokola N. Responsibilization in natural resources governance: a romantic doxa? Forest Policy Econ. 2020; 111 :102033. doi: 10.1016/j.forpol.2019.102033. [ CrossRef ] [ Google Scholar ]
  • Nerfa L, Rhemtulla JM, Zerriffi H. Forest dependence is more than forest income: Development of a new index of forest product collection and livelihood resources. World Dev. 2020; 125 (2020):104689. doi: 10.1016/j.worlddev.2019.104689. [ CrossRef ] [ Google Scholar ]
  • Nightingale A. A guide to systematic literature reviews. Surg Infect (larchmt) 2009; 27 (9):381–384. [ Google Scholar ]
  • Nightingale A. A guide to systematic literature reviews. Surgery. 2009; 27 (9):381–384. [ Google Scholar ]
  • North D. Institutions, institutional change and economic performance. Cambridge: Cambridge University Press; 1990. [ Google Scholar ]
  • Ntuli H, Mukong AK, Kimengsi JN. Institutions and environmental resource extraction within local communities in Mozambique. Forest Policy Econ. 2021; 139 :102724. doi: 10.1016/j.forpol.2022.102724. [ CrossRef ] [ Google Scholar ]
  • Ordóñez C, Kendal D, Threlfall CG, Hochuli DF, Davern M, Fuller RA, et al. How urban forest managers evaluate management and governance challenges in their decision-making. Forests. 2020; 11 (9):963. doi: 10.3390/f11090963. [ CrossRef ] [ Google Scholar ]
  • Oldekop JA, Rasmussen LV, Agrawal A, et al. Forest-linked livelihoods in a globalized world. Nat Plants. 2020; 6 :1400–1407. doi: 10.1038/s41477-020-00814-9. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Ongolo S, Giessen L, Karsenty A, Tchamba M, Krott M. Forestland policies and politics in Africa: recent evidence and new challenges. For Policy Econ. 2021; 127 (2021):102438. doi: 10.1016/j.forpol.2021.102438. [ CrossRef ] [ Google Scholar ]
  • Osei-Tutu P, Pregernig M, Pokorny B. Legitimacy of informal institutions in contemporary local forest management: insights from Ghana. Biodivers Conserv. 2014; 23 (14):3587–3605. doi: 10.1007/s10531-014-0801-8. [ CrossRef ] [ Google Scholar ]
  • Ostrom E. Governing the commons: the evolution of institutions for collective action. New York: Cambridge University Press; 1990. [ Google Scholar ]
  • Ostrom E. A general framework for analyzing sustainability of social-ecological systems. Science. 2009; 2009 (325):419–422. doi: 10.1126/science.1172133. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Palmatier RW, Houston MB, Hulland J. Review articles: Purpose, process, and structure. J Acad Market Sci. 2018; 46 (1):1–5. doi: 10.1007/s11747-017-0563-4. [ CrossRef ] [ Google Scholar ]
  • Pretzsch J, Darr D, Uibrig H, Auch E, editors. Forests and rural development. Berlin Heidelberg: Springer-Verlage; 2014. [ Google Scholar ]
  • Rammel C, Stagl S, Wilfing H. Managing complex adaptive systems: a co-evolutionary perspective on natural resource management. Ecol Econ. 2007; 63 :9–21. doi: 10.1016/j.ecolecon.2006.12.014. [ CrossRef ] [ Google Scholar ]
  • Ribot JC. Democratic decentralization of natural resources: institutional choice and discretionary power transfers in Sub-Saharan Africa. Public Admin Develop. 2003; 23 (1):53–65. doi: 10.1002/pad.259. [ CrossRef ] [ Google Scholar ]
  • Rounsevell MDA, Reginster I, Araújo MB, Carter TR, Dendoncker N, Ewert F, House JI, Kankaapãã S, Leemans R, Metzger MJ, et al. A coherent set of future land-use change scenarios for Europe. Agric Ecosyst Environ. 2006; 114 :57–68. doi: 10.1016/j.agee.2005.11.027. [ CrossRef ] [ Google Scholar ]
  • Rudel TK, Defries R, Asner GP, Laurance WF. Changing drivers of deforestation and new opportunities for conservation. Conserv Biol. 2009; 23 :1396–1405. doi: 10.1111/j.1523-1739.2009.01332.x. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Scott WR. Organizations, rational, natural and open systems. 4. New Jersey: Prentice Hall; 1995. [ Google Scholar ]
  • Senganimalunje TC, Chirwa PW, Babalola FD, Graham MA. Does participatory forest management program lead to efficient forest resource use and improved rural livelihoods? Experiences from Mua-Livulezi Forest Reserve, Malawi. Agroforest Syst. 2016; 90 (4):691–710. doi: 10.1007/s10457-015-9826-6. [ CrossRef ] [ Google Scholar ]
  • Sergent A, Arts B, Edwards P. Governance arrangements in the European forest sector: Shifts towards ‘new governance’or maintenance of state authority? Land Use Policy. 2018; 79 :968–976. doi: 10.1016/j.landusepol.2016.08.036. [ CrossRef ] [ Google Scholar ]
  • Shackleton S, Campbell B, Wollenberg E, Edmunds D. Devolution and community-based natural resource management: creating space for local people to participate and benefit. Nat Resour Perspect. 2002; 76 (1):1–6. [ Google Scholar ]
  • Soliev I, Theesfeld I, Abert E, Schramm W. Benefit sharing and conflict transformation: Insights for and from REDD+ forest governance in sub-Saharan Africa. For Policy Econ. 2021; 133 :102623. doi: 10.1016/j.forpol.2021.102623. [ CrossRef ] [ Google Scholar ]
  • Song XP, Hansen MC, Stehman SV, Potapov PV, Tyukavina A, Vermote EF, Townshend JR. Global land change from 1982 to 2016. Nature. 2018; 560 :639–643. doi: 10.1038/s41586-018-0411-9. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Steenbergen DJ, Warren C. Implementing strategies to overcome social-ecological traps. Ecol Soc. 2018 doi: 10.5751/ES-10256-230310. [ CrossRef ] [ Google Scholar ]
  • Tiwari BK, et al. Institutional arrangement and typology of community forests of Meghalaya, Mizoram and Nagaland of North-East India. J for Res. 2013; 24 (1):179–186. doi: 10.1007/s11676-013-0337-x. [ CrossRef ] [ Google Scholar ]
  • Tollefson J. Why deforestation and extinctions make pandemics more likely. Nature. 2020; 584 (13):175–176. doi: 10.1038/d41586-020-02341-1. [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Turubanova S, Potapov PV, Tyukavina A, Hansen MC. Ongoing primary forest loss in Brazil, Democratic Republic of the Congo, and Indonesia. Environ Res Lett. 2018; 13 :074028. doi: 10.1088/1748-9326/aacd1c. [ CrossRef ] [ Google Scholar ]
  • Tyukavina A, Hansen MC, Potapov PV, Stehman SV, Smith-Rodriguez K, Okpa C, Aguilar R. Types and rates of forest disturbance in Brazilian Legal Amazon, 2000–2013. Sci Adv. 2017; 3 :e1601047. doi: 10.1126/sciadv.1601047. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Tyukavina A, Hansen MC, Potapov P, Parker D, Okpa C, Stehman SV, Kommareddy I, Turubanova S. Congo Basin forest loss dominated by increasing smallholder clearing. Sci Adv. 2018; 4 :t2993. doi: 10.1126/sciadv.aat2993. [ PMC free article ] [ PubMed ] [ CrossRef ] [ Google Scholar ]
  • Verburg R, Rodrigues Filho S, Lindoso D, Debortoli N, Litre G, Bursztyn M. The impact of commodity price and conservation policy scenarios on deforestation and agricultural land use in a frontier area within the Amazon. Land Use Policy. 2014; 37 :14–16. doi: 10.1016/j.landusepol.2012.10.003. [ CrossRef ] [ Google Scholar ]
  • Wartmann FM, Haller T, Backhaus N. “ Institutional shopping” for natural resource management in a protected area and indigenous territory in the Bolivian Amazon. Hum Organ. 2016; 75 (3):218–229. doi: 10.17730/1938-3525-75.3.218. [ CrossRef ] [ Google Scholar ]
  • Weisse M, Goldman ED (2017). Global tree cover loss rose 51 percent in 2016, World Resources Institute blog, October 23, 2017. http://www.wri.org/blog/2017/10/global-tree-cover-loss-rose-51-percent-2016 . Accessed 15 Sept 2021.
  • Yeboah-Assiamah E, Muller K, Domfeh KA. Institutional assessment in natural resource governance: a conceptual overview. Forest Policy Econ. 2017; 74 :1–12. doi: 10.1016/j.forpol.2016.10.006. [ CrossRef ] [ Google Scholar ]

Reimagining Design with Nature: ecological urbanism in Moscow

  • Reflective Essay
  • Published: 10 September 2019
  • Volume 1 , pages 233–247, ( 2019 )

Cite this article

forest management research papers

  • Brian Mark Evans   ORCID: orcid.org/0000-0003-1420-1682 1  

978 Accesses

2 Citations

Explore all metrics

The twenty-first century is the era when populations of cities will exceed rural communities for the first time in human history. The population growth of cities in many countries, including those in transition from planned to market economies, is putting considerable strain on ecological and natural resources. This paper examines four central issues: (a) the challenges and opportunities presented through working in jurisdictions where there are no official or established methods in place to guide regional, ecological and landscape planning and design; (b) the experience of the author’s practice—Gillespies LLP—in addressing these challenges using techniques and methods inspired by McHarg in Design with Nature in the Russian Federation in the first decade of the twenty-first century; (c) the augmentation of methods derived from Design with Nature in reference to innovations in technology since its publication and the contribution that the art of landscape painters can make to landscape analysis and interpretation; and (d) the application of this experience to the international competition and colloquium for the expansion of Moscow. The text concludes with a comment on how the application of this learning and methodological development to landscape and ecological planning and design was judged to be a central tenant of the winning design. Finally, a concluding section reflects on lessons learned and conclusions drawn.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price includes VAT (Russian Federation)

Instant access to the full article PDF.

Rent this article via DeepDyve

Institutional subscriptions

forest management research papers

Similar content being viewed by others

forest management research papers

The politics of designing with nature: reflections from New Orleans and Dhaka

forest management research papers

Acknowledgements

The landscape team from Gillespies Glasgow Studio (Steve Nelson, Graeme Pert, Joanne Walker, Rory Wilson and Chris Swan) led by the author and all our collaborators in the Capital Cities Planning Group.

Author information

Authors and affiliations.

Mackintosh School of Architecture, The Glasgow School of Art, 167 Renfrew Street, Glasgow, G3 6BY, UK

Brian Mark Evans

You can also search for this author in PubMed   Google Scholar

Corresponding author

Correspondence to Brian Mark Evans .

Rights and permissions

Reprints and permissions

About this article

Evans, B.M. Reimagining Design with Nature: ecological urbanism in Moscow. Socio Ecol Pract Res 1 , 233–247 (2019). https://doi.org/10.1007/s42532-019-00031-5

Download citation

Received : 17 March 2019

Accepted : 13 August 2019

Published : 10 September 2019

Issue Date : October 2019

DOI : https://doi.org/10.1007/s42532-019-00031-5

Share this article

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

  • Design With Nature
  • Find a journal
  • Publish with us
  • Track your research

IMAGES

  1. Forest Ecology and Management Reprinted from

    forest management research papers

  2. (PDF) Guidelines for the Development of a Criteria and Indicator Set

    forest management research papers

  3. (PDF) Exploring Options for Joint Forest Management in India

    forest management research papers

  4. (PDF) Handbook on Forest Certification

    forest management research papers

  5. (PDF) Opinion Paper: Forest management and biodiversity

    forest management research papers

  6. (PDF) Annals of Forest Science promotes multidisciplinary research

    forest management research papers

VIDEO

  1. [Unit 4] Forest management (Part 1)

  2. Forest Management

  3. Assessing forest management scenarios C

  4. Assessing forest management scenarios B

  5. Forest Management

  6. Forest Management Roundtable Discussion

COMMENTS

  1. 202396 PDFs

    Explore the latest full-text research PDFs, articles, conference papers, preprints and more on FOREST MANAGEMENT. Find methods information, sources, references or conduct a literature review on ...

  2. Forest Ecology and Management

    2. Novel ideas or approaches to important challenges in forest ecology and management; 3. Studies that address a population of interest beyond the scale of single research sites (see the editorial, Three key points in the design of forest experiments, Forest Ecology and Management 255 (2008) 2022-2023); 4. Review Articles on timely, important ...

  3. Climate change impacts and adaptation in forest management: a review

    Twelve percent of papers (129) considered adaptation options, including 10 papers on adaptation in the forest sector. The first papers to focus on adaptation in the context of climate change were in 1996 with a number of papers published in that year (Kienast et al. 1996; Kobak et al. 1996; Dixon et al. 1996).Publications were then relatively few each year until the late 2000s with numbers ...

  4. Full article: Identifying sustainable forest management research

    Introduction. Sustainable forest management is a traditional research field with an outstanding history. Since Von Carlowitz (Citation 1713) practically invented the term sustainability in the context of forest management, the general understanding of what sustainable forest management (SFM) is, has undertaken an evolutionary process.Between 1980 and 1990 research communities increasingly ...

  5. Forest landscape planning and management: A state-of-the-art review

    The review covers the world statistics of global forests, forest plantations, production, planning, decision support systems, landscape ecology, heuristics, meta-heuristics, and geospatial solutions. In the second part of this paper, we provide an extensive review of worldwide qualitative and quantitative indices.

  6. Journal of Forest Research

    Journal of Forest Research covers 4 sections with the following research areas; Socioecnomics, Planning and Management Section. - forest policy and social sciences, forest assessment, modelling and management, forest operations and forest engineering, landscape planning and design, forest education. Forest Environment Section.

  7. Home

    The Journal of Forestry Research is an international platform for original theoretical, experimental research and technical reviews on forestry.. Welcomes review articles, original papers, commentary, perspective and short communications etc. with high academic quality.; Publishes scientific articles related to forestry for a broad range of international scientists, forest managers and ...

  8. A global systematic review of forest management institutions ...

    Temporal evolution of papers and cases on forest management institutions. ... contemporary realities and future avenues for research and policy. Forest Policy Econ 144:102846. Kimengsi JN, Mukong AK, Giessen L, Pretzsch J (2022b) Institutional dynamics and forest use practices in the Santchou Landscape of Cameroon. Environ Sci Policy 128(2022 ...

  9. Advances in Forest Management Research in the Context of Carbon ...

    Using the bibliometrix R-package in R and CiteSpace for bibliometric analysis of research areas from general statistics and knowledge base perspectives, this study systematically reviewed the status, evolution, and research hotspots of forest management in the context of carbon neutrality based on 6112 papers published in this research area ...

  10. Forestry

    The authors show China's forests can sequester 172.3 million tons of carbon per year in biomass by 2100, with an additional 28.1 million tons from improved management practices, but neglecting ...

  11. Conservation, Management and Monitoring of Forest Resources ...

    The book is organized into five sections: (I) Forest Conservation Ecology (II) Forest Conservation and Society (III) Forest Management (IV) Forest Monitoring using GIS and Remote Sensing and (V) Human Wildlife Conflicts. It covers various research themes related to forestry, wildlife, habitat fragmentation, forest management and human-wildlife ...

  12. Post-Soviet forest fragmentation and loss in the Green Belt around

    While much research is focused on detection of forest loss through the use of remotely sensed data (Cohen et al., 1998, Lioubimtseva, 2003, Lopez Garcia and Caselles, 1991, Miller and Yool, 2002), relatively little research has been done on analyzing the spatial patterns of such loss by using landscape metrics to explicitly quantify its pattern ...

  13. Forest Management Research Papers

    However, progress may be facilitated with a systematic approach to forest management embracing the usual planning cycle: formulation of objectives, preparation of a strategy, planning, implementing, monitoring, and reappraisal. This requires a good understanding of each particular situation.

  14. The History and Impact of Forest Management

    Studying Forest Management History. The History of Forest Cover. From Forest Use to Forest Management. Forest Management by Non-Professionals. Forest Management by Professionals. The Impact of Forest Management on Forests. Conclusions: Leaving the Margins of History. References

  15. A global systematic review of forest management institutions: towards a

    Temporal evolution of papers and cases on forest management institutions. On the whole, the literature on institutions has grown over the last 15 years(Fig. 3). This could be linked to renewed interests to understand governance mishaps and to engage in getting institutions (including FMIs) right. In all, while the number of publications ...

  16. Multipurpose GIS Portal for Forest Management, Research, and Education

    The main objective of this research was to develop a web-based geographic information system (GIS) based on a detailed analysis of user preferences from the perspective of forest research, management and education. An anonymous questionnaire was used to elicit user preferences for a hardware platform and evaluations of web-mapping applications, geographic data, and GIS tools. Mobile GIS was ...

  17. Urban forests of Moscow: typological diversity, succession ...

    The forest cover of the study area experienced a strong anthropogenic impact (logging, land plowing) for several centuries. In the first half of the twentieth century and especially actively after the WW2 (1941-1945) there was a significant change of direction of the impact—active creation of forest plantations (mainly pine and spruce) on the former arable land, which greatly increased the ...

  18. Reimagining Design with Nature: ecological urbanism in Moscow

    The twenty-first century is the era when populations of cities will exceed rural communities for the first time in human history. The population growth of cities in many countries, including those in transition from planned to market economies, is putting considerable strain on ecological and natural resources. This paper examines four central issues: (a) the challenges and opportunities ...