Download Effect of increased fire activity on global warming in the

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Effects of global warming on humans wikipedia , lookup

2009 United Nations Climate Change Conference wikipedia , lookup

Climate change and agriculture wikipedia , lookup

Economics of global warming wikipedia , lookup

Surveys of scientists' views on climate change wikipedia , lookup

Climate engineering wikipedia , lookup

Public opinion on global warming wikipedia , lookup

Attribution of recent climate change wikipedia , lookup

Climate governance wikipedia , lookup

Climate change, industry and society wikipedia , lookup

Global warming wikipedia , lookup

Effects of global warming on human health wikipedia , lookup

Carbon pricing in Australia wikipedia , lookup

Climate change and poverty wikipedia , lookup

Effects of global warming on Australia wikipedia , lookup

Climate change in Saskatchewan wikipedia , lookup

Climate-friendly gardening wikipedia , lookup

Mitigation of global warming in Australia wikipedia , lookup

Low-carbon economy wikipedia , lookup

Climate change in the United States wikipedia , lookup

Politics of global warming wikipedia , lookup

Years of Living Dangerously wikipedia , lookup

Solar radiation management wikipedia , lookup

Citizens' Climate Lobby wikipedia , lookup

Carbon Pollution Reduction Scheme wikipedia , lookup

IPCC Fourth Assessment Report wikipedia , lookup

Climate change feedback wikipedia , lookup

Business action on climate change wikipedia , lookup

Biosequestration wikipedia , lookup

Transcript
206
REVIEW
Effect of increased fire activity on global warming in the
boreal forest
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
France Oris, Hugo Asselin, Adam A. Ali, Walter Finsinger, and Yves Bergeron
Abstract: Forest fires are an important disturbance in the boreal forest. They are influenced by climate, weather, topography,
vegetation, surface deposits, and human activities. In return, forest fires affect the climate through emission of gases and
aerosols, and changes in surface albedo, soil processes, and vegetation dynamics. The net effect of these factors is not yet well
established but seems to have caused a negative feedback on climate during the 20th century. However, an increase in boreal
forest fires is predicted by the end of the 21st century, possibly changing the effect of fires on climate change to a positive feedback that
would exacerbate global warming. This review presents (1) an overview of fire regimes and vegetation succession in boreal forests;
(2) the effects on climate of combustion emissions and post-fire changes in ecosystem functioning; (3) the effects of fire regime
variations on climate, especially on carbon stock and surface albedo; (4) an integrative approach of fire effects on climate
dynamics; and (5) the implications of increased fire activity on global warming by calculating the radiative forcing of several
factors by 2100 in the boreal region, before discussing the results and exposing the limits of the data at hand. Generally, losses
of carbon from forest fires in the boreal region will increase in the future and their effect on the carbon stock (0.37 W/m2/decade)
will be greater than the effect of fire on surface albedo (−0.09 W/m2/decade). The net effect of aerosol emissions from boreal fires
will likely cause a positive feedback on global warming. This review emphasizes the importance of feedbacks between fires and
climate in the boreal forest. It presents limitations and uncertainties to be addressed in future studies, particularly with regards
to the effect of CO2 fertilization on forest productivity, which could offset or mitigate the effect of fire.
Key words: fire, radiative forcing, carbon cycle, surface albedo, carbonaceous aerosols.
Résumé : Les feux de forêt constituent une importante perturbation en forêt boréale. Le climat, la température, la topographie, la
végétation, les dépôts de surface et les activités humaines représentent autant de facteurs pouvant les influencer. À l'inverse, les feux
de forêt affectent le climat par l'émission de gaz et d'aérosols et modifient l'albédo de surface, les processus édaphiques et la
dynamique de la végétation. On ne connaît pas encore très bien l'effet net de ces facteurs, mais ils semblent avoir exercé une
rétroaction négative sur le climat au cours du XXe siècle. Cependant, on prédit une augmentation des feux de forêt vers la fin du XXIe
siècle susceptible de modifier leurs effets sur le climat vers une rétroaction positive capable d'exacerber réchauffement planétaire.
Cette synthèse présente (1) une revue d'ensemble des régimes des feux et de la succession de la végétation en forêts boréales, (2) les
effets sur le climat des émissions de combustion et des changements du fonctionnement des écosystèmes après feu, (3) les effets des
variations du régime de feu sur le climat, particulièrement sur les changements de stock de carbone et d'albédo de surface, (4) une
approche intégrée des effets des feux sur la dynamique du climat et (5) l'implication de l'augmentation de l'activité des feux sur le
réchauffement planétaire en calculant le forçage radiatif de plusieurs facteurs vers 2100 en région boréale, avant de discuter les
résultats et d'exposer les limites des données. Généralement, les pertes en carbone occasionnées par les feux de forêt en région boréale
augmenteront dans le futur et leur effet sur les stocks de carbone (0,37 W/m2/décennie) sera plus grand que l'effet du feu sur l'albédo
de surface (−0,09 W/m2/décennie). L'effet net des émissions d'aérosols venant des feux en forêt boréale causera vraisemblablement une
rétroaction positive sur le réchauffement planétaire. Cette synthèse met l'accent sur l'importance des rétroactions entre les feux et le
climat en forêt boréale. Elle présente les limites et les incertitudes à aborder dans les prochaines études, surtout en relation avec les
effets de la fertilisation par le CO2 sur la productivité forestière, laquelle pourrait contrebalancer ou atténuer les effets du feu. [Traduit
par la Rédaction]
Mots-clés : feu, forçage radiatif, cycle du carbone, albédo de surface, aérosols carbonés.
Introduction
Boreal ecosystems store large amounts of carbon in plant biomass, as well as in soils and peatlands (Apps et al. 1993; Pan et al.
2011). They play an important role in climate regulation by storing
anthropogenic carbon emissions (Bonan 2008). Forest fires, however, could counterbalance this effect by releasing stored carbon
into the atmosphere (as CO2) through combustion of plant biomass and soil organic matter (Seiler and Crutzen 1980; Carcaillet
et al. 2002; van der Werf et al. 2010; Hayes et al. 2011). When
combustion is incomplete, emissions can include other gases such
as carbon monoxide (CO), methane (CH4), and aerosols, with various impacts on climate (Carslaw et al. 2010; Simpson et al. 2011).
Carbon emissions by boreal forest fires represent 9% of global fire
emissions, behind the contribution of savanna, grassland, and
tropical forest fires (van der Werf et al. 2010). However, while
tropical forests are poor in soil organic matter, 40%–90% of carbon
Received 12 September 2013. Accepted 9 December 2013.
F. Oris, H. Asselin, and Y. Bergeron. Institut de recherche sur les forêts, Université du Québec en Abitibi-Témiscamingue, 445 boul. de l'Université,
Rouyn-Noranda, QC J9X 5E4, Canada.
A.A. Ali and W. Finsinger. Centre for Bio-Archaeology and Ecology (UMR5059 CNRS/UM2/EPHE), Institut de Botanique, 163 rue Broussonet, F-34090
Montpellier, France.
Corresponding author: France Oris (e-mail: [email protected]).
Environ. Rev. 22: 206–219 (2014) dx.doi.org/10.1139/er-2013-0062
Published at www.nrcresearchpress.com/er on 13 December 2013.
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
Oris et al.
emissions from boreal forest fires result from the combustion of
forest soil organic layers and peat (Kasischke et al. 1995; Amiro
et al. 2001b; van der Werf et al. 2010). Carbon accumulation in soils
and peatlands occurs over hundreds to thousands of years (Wirth
et al. 2002; Thornley and Cannell 2004; Kashian et al. 2006;
Lecomte et al. 2006) but can be released almost instantaneously
during a forest fire. By altering vegetation cover, forest fires also
change surface albedo (Liu et al. 2005; Amiro et al. 2006), thus
causing negative or positive radiative forcing (cooling or heating)
by decreasing or increasing the amount of solar radiation absorbed in the climate system (Randerson et al. 2006).
Increases in area burned are predicted in response to climate
change in Canada (Flannigan et al. 2005) and Alaska (Bachelet
et al. 2005; Balshi et al. 2009). In addition, areas with high fire risk
will increase by 100% in Russia (Stocks et al. 1998). These predictions raise the question “Will increased forest fire activity in the
boreal forest exacerbate global warming?”. This question is relevant, as the boreal forest carbon sink has decreased since 1997 and
is believed to store 70% less carbon than in the previous decades,
mainly because of climate change and more frequent forest fires
(Hayes et al. 2011). Moreover, compound disturbances may be
more common in response to global warming. Such disturbances
(e.g., an insect outbreak closely followed by a fire) have caused
regeneration failures and induced a change from spruce–moss
forests to lichen woodlands (Payette et al. 2000; Jasinski and
Payette 2005).
Radiative forcing (RF) is often defined as a perturbation to
the radiative energy budget relative to the pre-industrial state
(Ramaswamy et al. 2001). In the case of fire disturbance, RF is used
to quantify the per-unit area impact of forest fire relative to prefire state. Here we present a literature review to quantify the RF
associated with boreal forest fires through changes in carbon
stock, surface albedo, and aerosol emissions. Previous reviews
have focused on one or the other of these three variables (Kasischke
et al. 1995, 2000a, 2000b; Flannigan et al. 2009; Euskirchen et al.
2010); but to our knowledge, this is the first review to simultaneously consider the radiative forcing due to fire-induced changes in
carbon stock, surface albedo, and aerosol emissions in the boreal
forest. The review is organized as follows: (1) We provide an overview of fire regimes and vegetation succession in boreal forests; (2)
we expose the effects on climate of combustion emissions and
post-fire changes in ecosystem functioning; (3) we describe the
effects of fire regime variations on climate, especially on carbon
stock and surface albedo; (4) we examine integrated fire effects on
climate dynamics; and (5) we estimate the implications of increased fire activity on global warming in the boreal region, before discussing the results and exposing the limits of the data at
hand.
Fire regimes and vegetation succession in boreal
forests
Databases on forest fires are available for Alaska and Canada
since at least 1950 (Murphy et al. 2000; Stocks et al. 2002). In
Russia, fire maps are only produced for regions where fires are
suppressed, accounting for two thirds of Russian forests (Conard
and Ivanova 1997). Nevertheless, accurate satellite data are available in Russia since 1996 (Sukhinin et al. 2004) and have been used
in studies on the boreal forest (Balshi et al. 2007; Hayes et al. 2011).
Fires in North American boreal coniferous forests are mostly
crown fires that remove most of the trees and are thus called
stand-replacing fires (Johnson 1992). Unburned, residual stands
can account for 1%–25% of area burned (Delong and Tanner 1996;
Madoui et al. 2010; Dragotescu and Kneeshaw 2012). Large fires
(>200 ha) represent only 3% of total fires but contribute 97% of the
area burned in Canada (Stocks et al. 2002). In Russia, about 80% of
the burned areas are from surface fires, except during severe fire
seasons where crown fires may represent more than 50% of all
207
fires (Conard et al. 2002; Soja et al. 2004). In the North American
closed-crown boreal forest dominated by black spruce, the fire
return interval (FRI) can vary from 100 to 500 years, whereas lichen woodland experiences lower FRIs (70–110 years) (Payette
1992). Most of the variation in area burned in the Canadian boreal
forest is explained by the blockage of dry air masses in the troposphere which promotes dry conditions and thus large forest fires
(Skinner et al. 1999). At the regional scale, dry soils and coarser
surface deposits have been associated with shorter fire cycles
(Harden et al. 2003; Mansuy et al. 2010). In Russia, the FRI was
estimated at 90–130 years in larch (Larix sibirica), fir (Abies sibirica),
and spruce (Picea obavata) forests, and 25–50 years in pine (Pinus
sylvestris) forests (Conard and Ivanova 1997). Just after snowmelt,
the soil is still moist, promoting low-severity fires; whereas the
soil progressively dries until summer when severe fires occur
(Turetsky et al. 2011). Peat fires can also occur (van Bellen 2011) and
are mostly limited to the end of the fire season when peat is drier.
Peat fires in western Canada account for 10%–15% of all fires
(Turetsky et al. 2004).
In boreal forests, post-fire tree regeneration is mainly controlled by species composition prior to disturbance (seed availability), competition (shade tolerance), surface deposit, as well as fire
severity and frequency (Johnstone and Kasischke 2005; Johnstone
and Chapin 2006a; Greene et al. 2007). An increase in fire frequency creates a younger landscape and promotes the installation
of early successional, shade-intolerant broadleaved species (e.g.,
Populus tremuloides and Betula papyrifera) and fire-adapted pioneer
species (e.g., Picea mariana and Pinus contorta var. latifolia) (Johnstone
and Chapin 2006a). In the spruce–moss forests of eastern Canada,
increased fire frequency over the past 50 years caused a shift from
dense forests to open woodlands (Girard et al. 2008). Fire severity
affects vegetation dynamics by altering the thickness of the soil
organic layer (Johnstone and Chapin 2006b). High seedling establishment has been reported on organic soils less than 2.5 cm thick
(Johnstone and Chapin 2006b). However, a negative correlation
between fire severity and post-fire recruitment was found in boreal larch forests of northeastern China (Cai et al. 2013) and a black
spruce forest of Alaska (Johnstone and Kasischke 2005). After
high-severity fires, pioneer broadleaved species can rapidly recolonize the sites (Greene and Johnson 1999; Johnstone and
Kasischke 2005). Response to fire severity also depends on habitat
type, with neutral and negative responses in lichen woodlands
and moist conifer forests, respectively (Johnstone and Chapin
2006b).
Emissions from wildfires
Combustion proceeds through a sequence of stages (ignition;
flaming, glowing, and pyrolysis; glowing and pyrolysis (smoldering); glowing; and extinction) (Lobert and Warnatz 1993). Combustion through flaming mostly produces CO2, NO, and N2O, while
smoldering mostly produces CO and CH4 (Lobert and Warnatz
1993).
Carbon emissions
Estimates of total carbon released from fire (Ct) can be obtained
from a widely used equation requiring quantification of area
burned (A), fuel loading (biomass per unit area), and proportion of
biomass consumed (Seiler and Crutzen 1980; French et al. 2000):
(1)
Ct ⫽ A[(Cg × ␤g) ⫹ (Ba × fc × ␤a)]
where Ba is aboveground biomass; fc is the fraction of carbon in
aboveground biomass (percent); Cg is density of the carbon in
ground-layer components; and ␤a and ␤g are, respectively, the
aboveground and ground-layer carbon fraction consumed during
a fire.
Published by NRC Research Press
208
Environ. Rev. Vol. 22, 2014
Table 1. Estimated annual carbon emissions by fires in different boreal regions.
Region
Period
Emission
(Tg C/year)
Burned area
(106 ha/year)
Northern
Hemisphere
Boreal area
1977−1990
101
9.4
1998
1995−2003
1980−1994
1997−2004
1997−2006
1997−2009
1997−2006
1997−2004
1997−2010
Based on FRI
1998
1998−2002
1998−2010
1996−2002
1996−2002
1959−1999
1990−2008
1950−1999
1996−2002
290−383
106−209
53
44
51
54
255
133
128
194
135−190
116−520
92
208.8−215.7
32.2−32.9
27
23
4.5
13.7−13.9
17.9
3−23.6
2.6
2.3
2.57
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
BONA
BOAS
Russia
Siberia
Eurasia
Canada
Alaska
7.15
9.7
12
13.3
6.88−11.17
5.35
5.383
3.9
2
0.66
0.22
0.289
Note
Reference
Auclair and Carter 1993
Fire severity range, including peat fires
Fire severity range
Weighting by fire severity
CASA model
TEM model
CASA model
TEM model
CASA model
CASA model
Adjusted from Dixon and Krankina (1993)
Fire severity range
Fire severity range
ILIS
TEM; with or without CO2 fertilization
TEM; with or without CO2 fertilization
Different fuel types
Inventory in managed forest
GIS-based model
TEM; with or without CO2 fertilization
Kasischke and Bruhwiler 2003
Kasischke et al. 2005
French et al. 2000
van der Werf et al. 2006
Hayes et al. 2011
van der Werf et al. 2010
Hayes et al. 2011
van der Werf et al. 2006
van der Werf et al. 2010
Conard and Ivanova 1997
Conard et al. 2002
Soja et al. 2004
Shvidenko et al. 2011
Balshi et al. 2007
Balshi et al. 2007
Amiro et al. 2001b
Stinson et al. 2011
French et al. 2002
Balshi et al. 2007
Note: BONA, boreal North America; BOAS, boreal Asia; FRI, fire return interval; CASA, Carnegie–Ames–Stanford Approach; TEM, terrestrial ecosystem model;
ILIS, integrated land information system.
Variations in estimates of total carbon released among different
boreal regions (Table 1) are probably due to (1) the availability of
reliable data on area burned, especially in Russia where data are
scarce; (2) spatial variability of fire severity; and (3) inclusion of
fire severity in the models (% combustion of plant biomass and soil
organic matter). Most studies, however, reported similar values,
except Balshi et al. (2007) who reported greater carbon emissions
in Alaska. This can be explained by their use of the carbon consumption of soil organic matter estimates from French et al.
(2000) in their study, which reported higher consumption for the
Alaska sites.
Table 2. Emission factors (g/kg) for different fire types in Canada.
Gaseous emissions
To precisely characterize fire emissions, the emission factor (EF)
is frequently used. It is reported as the amount (in grams) of gas
emitted by kilogram dry matter burned (Andreae and Merlet
2001). Forest fires in boreal regions represent 18% of greenhouse
gas emissions in North America and Northern Asia (Amiro et al.
2001a), with 2.5 times as much emissions from Asia than from
North America (van der Werf et al. 2010). About 99% of the carbon
released from fires is composed of CO2, CO, and CH4 (Cofer et al.
1998; Simpson et al. 2011). The CO2 EF from boreal forest fires
varies between 1210 and 1616 g/kg in Canada (Cofer et al. 1998;
Simpson et al. 2011), considerably more important than the EFs of
CO and CH4 (Table 2). CO, CH4, and NOx (NO and NO2) can influence atmosphere chemistry, including a contribution to the formation of tropospheric ozone (O3), an important greenhouse gas
(Crutzen and Carmichael 1993; Sudo and Akimoto 2007). Emissions from forest fires contributed 13.8% and 12.4% of global fire
emissions of CO and CH4 in 1998, respectively (Kasischke and
Bruhwiler 2003). Variations in gas emissions by forest fires explained 85% of the variation in atmospheric CO concentration in
the northern hemisphere (excluding tropical forests) between
1995 and 2003 (Kasischke et al. 2005).
effects on climate depend on many parameters such as optical
depth, particle size, and single scattering albedo (SSA) (Hansen
et al. 1980; Reid et al. 2005; Myhre et al. 2013). Particles emitted by
forest fires contain more than 50% carbon, including both organic
carbon (OC) and elemental carbon (EC) (Lobert and Warnatz 1993).
Elemental carbon (also known as black carbon) is the main lightabsorbing constituent in aerosols and has the second largest radiative power at the global scale, with 0.2– 1.1 W/m2 (Jacobson 2001;
Bond et al. 2013). Its elimination could reduce global average
temperature by 0.5–1.0 °C (Ramanathan and Carmichael 2008).
However, many factors can contribute to uncertainties in the estimates of the direct forcing by aerosols, including a lack in EC
emission estimates (Bond et al. 2013). Indeed, EC emissions varied
from 0.016 Tg EC/year in 1960–1969 to 0.041 Tg EC/year in 1990–
1997 (Lavoué et al. 2000) and to 0.12 Tg EC/year in 1997–2000 for
Canadian fires (Park et al. 2003). Organic carbon scatters solar
radiation and cools the climate, except for some OC components
which are sunlight absorbers (Forster et al. 2007; Myhre et al. 2013).
Emission factor values of OC range between 8.6 and 9.7 g/kg,
greater than for EC with 0.56 ± 0.19 g/kg. Organic carbon is the
largest component of extratropical forest biomass burning aerosols (Andreae and Merlet 2001).
Aerosols also have indirect or semi-direct effects on climate.
Large clouds of smoke emitted by fires reduce solar radiation
that reaches the surface (Randerson et al. 2006; Ramanathan
and Carmichael 2008). In addition, deposition of light-absorbing
particles on snow and ice reduces the albedo and may accelerate
warming. Flanner et al. (2007) estimated the radiative forcing of
EC from boreal fires and pollution deposited on snow in 1998
Aerosol emissions
During a forest fire, aerosols (also called particulate emissions)
play an important role in climate regulation at the regional scale
due to their short half-life (hours to weeks). Direct radiative forcing by fire-emitted aerosols was equivalent to 17 ± 30 W/m2 during
the first year after a boreal fire (Randerson et al. 2006). Aerosol
CO2
CO
CH4
NOx
Mixed
combustiona
Flaming
combustionb
Smoldering
combustionb
1616±180
113±72
4.7±2.9
0.97±0.12
1530±80
100±10
2.4±0.6
n/a
1210±100
184±15
7.8±2
n/a
Note: n/a, not available.
aSmoke and flame (Simpson et al. 2011).
bCofer et al. 1998.
Published by NRC Research Press
Oris et al.
209
Region
Respiration
increase?
Alaska
Alaska
Alaska
Alaska
Alaska
Finland
Siberia
Yes
Yes
Yes
Yes
Yes
No
Yes
Note
Reference
Severe fires in black spruce forest; net loss of 20 t C/ha 15–20 years after a fire
Rh measured only in the first year, release of 202 g C/m2/year in black spruce forest
Release of 1.8−11.0 Mg C/ha, 7–15 years after fire in black spruce forest
Release of 100 g C/m2/year in a young black spruce population
Release of 14.7 t C/ha in black spruce stands over a decade after fire
Seven-fold decrease during 3 years in a Scots pine forest
Increase in Rh after fire in a larch forest correlated with increase in temperature
and organic carbon content
Kasischke et al. 2000b
Randerson et al. 2006
O'Neill et al. 2003
Bond-Lamberty et al. 2004
Kim and Tanaka 2003
Fritze et al. 1994
Sawamoto et al. 2000
and 2001. Isolating the contribution of boreal fires (12% in 2001
and 20% in 1998), the radiative forcing ranged from −0.0055 W/m2
(in 2001) to −0.01001 W/m2 (in 1998). This forcing is low, but its
effectiveness is greater than that of direct absorption by EC because of snow melting. Aerosols can also act as cloud condensation nuclei (CCN), form small cloud drops, and thus cool the
climate by increasing cloud albedo (Chuang et al. 2002; Lohmann
and Feichter 2005). This indirect effect is called the cloud albedo
effect. A cooling of 0.34 W/m2 at the global scale has been attributed to EC emissions from fires and pollution causing the formation of CCN (Spracklen et al. 2011). Moreover, smaller drops need
longer to reach sizes at which they can fall as rain, thus increasing
cloud cover and resulting in additional cooling. This second indirect effect is termed cloud lifetime effect (Albrecht 1989). Elemental carbon can also decrease the relative humidity of the cloud
layer by a semi-indirect effect, leading to evaporation of cloud
drops, decreased cloud cover and albedo and thus climate warming (Hansen et al. 1997; Ramanathan and Carmichael 2008). However, this effect depends on several factors, including cloud type
and the altitude of the EC relative to the cloud (Koch and Del
Genio 2010).
The single scattering albedo (SSA) is defined as the ratio of
radiation scattering to the sum of scattering and absorption coefficients. If SSA is less than 0.85, aerosols are responsible for a
positive (heating) radiative forcing. If it is greater than 0.95, the
forcing is negative (cooling). Between 0.85 and 0.95, the effect can
be positive or negative depending on cloud cover and surface
albedo (Ramanathan et al. 2001). For boreal and temperate forest
fires, a synthesis of in situ measurements of optical properties of
fresh smoke particles placed this ratio between 0.6 and 0.97 (Reid
et al. 2005). It varied according to fuel type, fire intensity, flame
height, and combustion phase (Radke et al. 1991; Reid et al. 2005).
Post-fire ecosystem functioning
Carbon fluxes
Carbon dioxide fluxes from vegetation regeneration (photosynthesis and autotrophic respiration (Ra)) and soil (heterotrophic
respiration (Rh)) are key elements in the post-fire restoration of
the boreal forest carbon pool (Chapin et al. 2006). Tree mortality
causes a decrease in autotrophic respiration (Ra) (Kasischke et al.
2000b; Sawamoto et al. 2000; Zhuang et al. 2002; Kim and Tanaka
2003; Bergner et al. 2004). At the microbial scale, however, the
pattern is unclear (Amiro et al. 2003). Higher soil temperatures
following fire increase microbial soil respiration (Rh), while the
removal of micro-organisms in severe fires could cause a decrease
in respiration. However, most studies find an increase in Rh after
fire (Table 3; Fig. 1), which is included in process-based climate
models (Auclair and Carter 1993; Dixon and Krankina 1993). These
indirect biogenic carbon losses could be greater than the direct
losses from combustion during fire (Auclair and Carter 1993;
Dixon and Krankina 1993; Kasischke et al. 1995). Greenhouse gases
(CO2 and CH4) emitted in the first year after a fire were responsible
for a 6 and a 8.3 ± 3 W/m2 positive radiative forcing in Manitoba
and Alaska, respectively (Randerson et al. 2006). Hence, boreal
Fig. 1. Conceptual net ecosystem production (NEP), net primary
production (NPP), and heterotrophic respiration (Rh) dynamics after
a fire in the boreal forest.
Carbon fluxes (g C/m2 /year )
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
Table 3. Heterotrophic respiration after fire in boreal forests.
Sink
NPP
NEP
Rh
Source
Time since last fire
Years
forest ecosystems remain carbon sources for several years after
fire because of increased soil respiration, decomposition of dead
wood, and absence of vegetative cover (Fig. 1). Usually, net primary
productivity (NPP = photosynthesis – Ra) first increases and culminates to high values at an early age, before later declining (Gower
et al. 1996; Chen et al. 2002; Goulden et al. 2011; Alexander et al.
2012). However, old-growth forests are not necessarily a carbon
source and have great sequestration potential (Luyssaert et al.
2008; Wei et al. 2013). Net ecosystem productivity (NEP = NPP – Rh)
returns to its pre-fire level after 10–40 years (Amiro 2001; Chen
et al. 2002; Amiro et al. 2003; Bond-Lamberty et al. 2004; Goulden
et al. 2011). However, even after NEP has reached its pre-fire level,
recovery of the initial carbon stock can take up to 235 years
(Kashian et al. 2006). Long-term carbon accumulation varies
between forest types and soil drainage. A modeling exercise by
Harden et al. (2000) showed that drier forests generally had
greater carbon losses to fire than wetter forests.
In boreal regions characterized by permafrost, increased soil
temperature following fire may lead to permafrost degradation
(Brown et al. 1983; Burn 1998; Burn et al. 2009). After the 1968
Inuvik fire in the Northwest Territories, a 50 cm increase of the
thickness of the active layer (surface layer of permafrost that
thaws in the summer) occurred over an 8-year period (Brown et al.
1983; Mackay 1995). This effect likely depends on the residual
depth of the organic layer. Residual depths >7–12 cm should protect permafrost from melting (Yoshikawa et al. 2003). Permafrost
melting can restore the soil respiration process, with major implications for the carbon cycle. A recent modeling study estimated
that future fires will lead to a 1.1 m increase of the thickness of the
active layer by 2100 in Alaska (Genet et al. 2013). With warming
Published by NRC Research Press
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
210
added to changes in fire regime, they projected a 9.8 kg C/m2
decrease of the soil carbon stock by 2100 (Genet et al. 2013).
Solid elemental carbon, more commonly known as charcoal, is
formed during the combustion of plant material. It represents a
potential carbon sink because it is resistant to decomposition (de
Lafontaine and Asselin 2011). Fire temperature, burn time, and
type and size of fuels influence charcoal production (Czimczik
et al. 2005; Fréjaville et al. 2013). Between 0.7% and 2.0% of the
initial organic matter can be converted to charcoal during boreal
forest fires (Hart and Luckai 2013). An estimated 8%–10% of soil
carbon is estimated to be under the form of charcoal (Hart and
Luckai 2013), up to 20%–24% in some Siberian soils (Schulze et al.
1999).
Albedo
Albedo is the fraction of solar radiation reflected back into
space. It plays an important role in climate regulation: the higher
the albedo, the cooler the climate (Eugster et al. 2000). Boreal
forests have a dark canopy (with a low albedo) that obscures the
snow-covered ground in winter. The albedo is higher for deciduous forests compared with coniferous forests (Betts and Ball 1997;
Lukeš et al. 2013). By removing vegetation cover, fire acts on albedo in two different ways. Albedo increases in winter and decreases in summer, respectively cooling and warming the climate.
As the difference in albedo before and after fire is greater in
winter than in summer, the net effect is that albedo increases in
the early years after a fire (5–55 years; Amiro et al. 2006;
Randerson et al. 2006; O’Halloran et al. 2012), and it decreases
later on due to the afforestation process (Liu et al. 2005). However,
if post-fire succession is dominated by broad-leaved species, an
increase in albedo will be observed in contrast to the pre-fire
coniferous stand (Liu and Randerson 2008).
Evapotranspiration
Evapotranspitation is the heat flux from the Earth's surface to
the atmosphere that is associated with evaporation or transpiration of water at the surface and subsequent condensation of water
vapor in the troposphere. This latent heat flux has a subsequent
impact on climate by affecting cloud cover and precipitation, and
thus the radiative energy balance. Higher evapotranspiration
causes a cooling of the regional climate (Eugster et al. 2000; Liu
et al. 2005). In Alaska, the loss of moss that helds soil moisture and
an increase in soil temperature have explained a 33% decrease in
annual evapotranspiration 3 years after fire (Liu et al. 2005). Thereafter, if post-fire succession is dominated by broad-leaved species,
an increase in evapotranspiration will be observed (Liu and
Randerson 2008).
Environ. Rev. Vol. 22, 2014
However, vegetation recovery may take longer when very severe
fires occur on thin organic soils (Asselin et al. 2006). Even if the
annual rate of storage was greater after a high intensity fire, a
theoretical study estimated that the carbon stock returns to its
pre-fire value more rapidly if fire severity is low (Thornley and
Cannell 2004). North American boreal forests are characterized by
infrequent severe fires and FRI long enough to allow the forest to
return to a mature state before the next fire (Payette 1992). In
Russia, fires are less severe but more frequent (Conard and
Ivanova 1997; Wooster and Zhang 2004). For a comparable area,
Russian fires may burn less intensely and produce fewer emissions than those in North America (Wooster and Zhang 2004).
Satellite images (MODIS) from Alaska and central western Canada showed that albedo was higher in severely burned areas due
to a higher proportion of deciduous regeneration (Beck et al.
2011; Jin et al. 2012). However, in areas where jack pine colonizes
burned stands, the post-fire albedo is lower than that in the prefire stand (Goetz et al. 2007). Increased fire severity could offset
the negative albedo feedback by releasing more carbon (Turetsky
et al. 2011). The shift from spruce–moss forest to lichen woodlands
due to compound disturbances in the eastern Canadian boreal
forest (Jasinski and Payette 2005; Girard et al. 2008) causes an
increase in albedo that offsets the carbon loss due to reduced tree
density and thus causes a net radiative forcing of −0.094 mW/m2
(Bernier et al. 2011).
Impact of fire regime on climate at broader scales
As the number and size of fires vary from year to year, carbon
losses (or gains) also change each year (van der Werf et al. 2006).
For example, emissions from forest fires in Siberia in 1998 accounted for 14%–20% of global emissions by fire (Conard et al.
2002), whereas in “normal” years they represent only 6.4% of total
emissions (van der Werf et al. 2010). Other studies suggest that the
strength of the carbon sink in the North American boreal forest
has declined in recent decades (Kurz and Apps 1999; Balshi et al.
2007; Hayes et al. 2011) due to an increase in forest fires.
A decrease in the time interval between fires across the boreal
forest could lead to a decrease in carbon stock and even a shift
from the boreal forest being a carbon sink to it being a source. The
fire regime threshold beyond which the boreal forest becomes a
carbon source depends on forest productivity and is thus difficult
to estimate. Significant increase in NEP would be required for a
100% increase in area burned predicted in Canada (Flannigan et al.
2005; Kurz et al. 2008; Metsaranta et al. 2010).
Impact of fire regime variations on climate
Integrative approach of fire effects on climate
dynamics
Impact of fire regime on climate at the stand scale
A decrease in the time interval between fires results in a
decrease of the carbon stock (Kasischke et al. 1995; Wirth et al.
2002; O'Neill et al. 2003; Thornley and Cannell 2004). In addition,
younger landscapes have been associated with lower carbon
stocks as biomass is lower in young populations than in older ones
(Harmon et al. 1990; Pregitzer and Euskirchen 2004; Wang et al.
2011). However, as the fire return interval (FRI) decreases, less
carbon is released per fire event because the ecosystem does not
have the time to accumulate carbon (Wirth et al. 2002; Zhuang
et al. 2006). Thus, if fire frequency increases, forests might continue to store carbon but will inevitably be a weaker carbon sink.
This hypothesis was studied by Brown and Johnstone (2011) who
measured the consumption of organic soil under different FRIs
(long: 94 years; short: 15 years) in Yukon. They found that the
carbon stock of the forest under short FRI represented only 60% of
the post-fire carbon found under long FRI.
Vegetation recovery is usually denser and quicker after severe
fires (Lecomte et al. 2006; Simard et al. 2007; Shenoy et al. 2011).
Only a few studies have estimated the combined boreal forest
fire effects on climate (Table 4). Most studies focused on the stand
scale, and data are scarcer at the boreal forest biome scale
(Table 4). The negative radiative forcing of post-fire changes in
albedo (−4.2 W/m2; Randerson et al. 2006) was estimated to be
greater than the positive radiative forcing of changes in NEP
(+1.6 W/m2; Randerson et al. 2006) in Alaska over 80 years. In
Manitoba, the average net radiative forcing over 100 years was
estimated at −0.9 ± 2.4 W/m2 (O’Halloran et al. 2012). However,
integrated post-fire changes in albedo and NEP have caused a
short-term positive radiative forcing (Randerson et al. 2006;
O’Halloran et al. 2012), but the trend rapidly (<5 years) reversed in
Alaska in contrast to Manitoba ca. after 50 years (Table 4). In
addition to post-fire changes in albedo and NEP, Randerson et al.
(2006) estimated that the effect of aerosols released from fires and
the cumulative fire effect caused a positive radiative forcing during more than 10 years. Hence, combined boreal fire effects seem
responsible for climate warming shortly after fire, followed by
climate cooling 80–100 years after fire (Table 4).
Published by NRC Research Press
Region
Studied effects
Carbon stock changes − boreal forest scale
Canada
Net ecosystem carbon fluxes
Period
Results
Feedback
Temporal scale
Reference
1920−1989
Increase in carbon release and
decrease in carbon stock since
1980
Decrease in carbon stock with an
increase in fire frequency and
release of an average of about
0.1 Gt C/year
Loss of 10%−30% in NPP after fire
Positive
9 years
Kurz and Apps
1999
Positive
500 years
Thornley and
Cannell 2004
Positive
6500 years
Harden et al. 2000
Saskatchewan
Fire frequency on NPP and
carbon storage
Fire return interval of
100 years
Boreal forest of North
America
Boreal forest
Fire-free interval
Last 6500 years
NEP
1997−2006
73% decrease insink strength
compared to 20th century
Positive
9 years
Hayes et al. 2011
Radiative budget − stand scale
Alaska
Post-fire change in albedo
1950−2004
Negative
50 years
following fire
Lyons et al. 2008
Alaska, Saskatchewan
and Manitoba
150-year forest
chronosequence
Surface radiative forcing of
−6.2 W/m2 compared with
mature coniferous forest
Decrease of 12−24 W/m2 of net
surface radiation in summer
Negative
1−3 years
following fire
Amiro et al. 2006
Negative
15 years
following fire
Liu et al. 2005
Increase in albedo due to shift
Negative
from black spruce–moss forest
to spruce–lichen woodland
responsible for a radiative
forcing of −0.094 mW/m2
Greater deciduous biomass in
Negative
more severely burned area;
winter and summer albedo were
higher in high-severity burns
Increases in spring albedo above
Negative
pre-fire levels in all burn severity
classes, and higher for high
severities
After successive
fires
Bernier et al. 2011
10−50 years
following
fires
Beck et al. 2011
5−7 years
following fire
Jin et al. 2012
80-year fire
cycle
Randerson et al.
2006
100-year fire
cycle
O’Halloran et al.
2012
Alaska
Radiative budget, post-fire
change in albedo and
evapotranspiration
Radiative budget
1-year forest chronosequence
in three sites 80, 15, and
3 years following a fire
2000−2008
Boreal forest of
eastern North
America
Radiative budget/post-fire
change in albedo
Alaska
Post-fire change in albedo
along gradients of burn
severity
1950−1991, 60-year forest
chronosequence
Central western
Canada
Post-fire change in albedo
along gradients of burn
severity
2000−2009
Combined effects − stand scale
Boreal forest -Alaska
GHG, aerosols, albedo, and
black carbon deposition
on snow
Boreal forest CO2 and post-fire change in
Manitoba
albedo
Decrease of 17 W/m2 in annual net
surface radiation
Calibrated on a 1999 fire in
Alaska
During the first year: 34 ± 31 W/m2
Over an 80-year fire cycle: −2.3 W/m2
150-year forest
chronosequence
First 50 years: 0.59 W/m2
Last 50 years: −2.4 W/m2
Over 100 years: −0.9 W/m2
Positive the first year/
negative over 80
years
Positive the first 50
years/negative over
100 years
Note: A positive feedback refers to a warming, whereas a negative feedback refers to a cooling. NPP, net primary production; NEP, net ecosystem productivity.
211
Published by NRC Research Press
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
Oris et al.
Table 4. Estimations of fire effects on climate at different scales in boreal forests.
212
Environ. Rev. Vol. 22, 2014
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
Table 5. Summary of future fire activity in North America and Russia. A1B and A2 are IPCC scenarios with the most intense climate forcing,
whereas B1 and B2 are scenarios with intermediate climate forcing (IPCC 2007).
Region
Period or scenario
Result
Reference
Canada
1959−97 vs. 2080−99 (3×CO2)
Flannigan et al. 2005
Canada
2×CO2; 3×CO2
Alaska
Western North America
1922−96 vs. 2025−99
1991−2000 vs. 2091−2100
(range of two IPCC scenarios:
A2 and B2)
2001−2007 vs. 2091−2100 (range of
scenarios A2, A1B, and B1)
1980−1989 vs. 2×CO2
74%–118% increase in area burned
(initial value of 1779 Mha)
34% increase in area burned and 3% increase
in soil consumption (kg/m2) by fire;
93% increase in area burned and 6% increase
in soil consumption (kg/m2) by fire
14%–34% increase in area burned
3.5- to 5-fold increase in area burned
Western Canada
Russia
Russia
Central Russia
Quebec
Eastern Canada
Quebec
1981−2000 vs. 2100
2001−2007 vs. 2091−2100 (range of
scenarios A2, A1B, and B1)
2×CO2; 3×CO2
1959−99 vs. 2046−65 and 2081−2100
(range of three IPCC scenarios:
A2, A1B, and B1)
1971−2000 vs. 2071−2000
(IPCC A2 without or with
unlimited tree dispersion)
54%–69% of all fires as crown fires
(initial value of 57%)
2- to 3-fold increase in high fire risk area in
June and July
2.5-fold increase in high fire risk area
6%–22% of all fires as crown fires
(initial value of 6%)
0.2389% increase in annual burn rate;
0.4441% increase in annual burn rate
(initial value of 0.1725%/year)
0.29−0.40 increase in annual burn rate;
0.35−0.45 increase in annual burn rate
(initial value of 0.22%)
10%−25% increase in fire occurrence;
under unlimited dispersal scenarios,
increases in fire occurrence are predicted
to be less important in the boreal forest,
and trends could even be reversed
In both studies, the fire was severe and induced a more productive 40-year deciduous succession phase with an albedo
higher than the pre-fire black spruce forest (Randerson et al. 2006;
O’Halloran et al. 2012). In addition, calculations of radiative forcing
emissions were based on the release of about 1.5–1.7 kg C/m2 during
fire. This is close to Canada's 1959–1999 average of 1.3 kg C/m2 (Amiro
et al. 2001a), but far from the average release estimates of 2.95 ±
0.12 and 6.15 ± 0.41 kg C/m2 between 1983 and 2005 in early-season
and late-season Alaskan fires, respectively (Turetsky et al. 2011).
Although albedo is higher after more severe fires (Beck et al. 2011;
Jin et al. 2012), the question is to know if there is a fire severity
threshold beyond which there would be no change in albedo but
still an increase in carbon emissions.
Future feedbacks between fire and climate
Climate change effects on future boreal forest fire
dynamics
Global circulation models (GCMs) predict increased temperatures in boreal ecosystems. In Canada, the greatest increases in
temperature are predicted for the north and central regions during the period 2041–2070 (Price et al. 2001). Winter minima will be
the most impacted with increases of 6 °C. A 5% increase in mean
annual precipitation is predicted across the country, with considerable regional variation and greater increases in summer (IPCC
2007). In Quebec, winter temperatures are expected to increase
between 5 and 9 °C and summer temperatures between 2 and
3.5 °C by the end of the 21st century (Logan et al. 2011). In Alaska,
a 4.5–10 °C warming is expected for 2100, along with a 21% increase in precipitation (Christensen et al. 2007). In Russia, temperature and precipitations are projected to increase by 4.3 °C and
15%, respectively (Christensen et al. 2007).
Many studies have evaluated the effect of climate change on
fires, especially in North America (Bachelet et al. 2005; Flannigan
et al. 2005; Bergeron et al. 2006, 2010; Balshi et al. 2008; Amiro
Amiro et al. 2009
Bachelet et al. 2005
Balshi et al. 2009
de Groot et al. 2013b
Stocks et al. 1998
Malevsky-Malevich et al. 2008
de Groot et al. 2013b
Bergeron et al. 2006
Bergeron et al. 2010
Terrier et al. 2013
et al. 2009; de Groot et al. 2013b), and to a lesser extent in Russia
(Stocks et al. 1998; Malevsky-Malevich et al. 2008; de Groot et al.
2013b) (Table 5). These studies suggest that the increase in precipitation will not be sufficient to counteract the increase in temperature, thus resulting in increased fire incidence in the boreal
forest. This increase is often estimated in terms of area burned,
with a range of increase across the boreal forest from 14% to 450%
(Table 5).
Effect of increased fire frequency on climate
In Quebec, carbon release estimates 3000 years before present were 80% higher than current emission levels (Bremond et al.
2010). It is thus likely that past boreal wildfires made higher contributions to the global carbon stock than current ones. With an
increased fire activity due to climate change, boreal wildfires
could again make major contributions to the global carbon stock.
Studies in Canada (Kasischke et al. 1995; Zhuang et al. 2006;
Amiro et al. 2009; Balshi et al. 2009), Alaska (Kasischke et al. 1995;
Bachelet et al. 2005; Zhuang et al. 2006; Balshi et al. 2009), and
Russia (Kasischke et al. 1995; Zhuang et al. 2006) have so far mostly
focused on future carbon emissions compared with other effects
of fire on climate (Table 6). An increase of 24%–450% in carbon
emissions was predicted depending on regions and studies.
A compilation of data available from the literature allowed us to
calculate the regional radiative forcing of albedo and carbon stock
changes due to fire for the present and future (end of 21st century)
in North America and in the whole boreal region (Fig. 2). Currently, the radiative forcing due to fires is an average of 0.04 ± 0.02
(mean ± standard deviation) W/m2/decade in the boreal region. At
the end of the 21st century, this effect will be almost ten times
greater, with an average of 0.37 ± 0.03 W/m2/decade. Other factors
influence carbon stocks, such as climate change and the CO2 fertilization effect on photosynthesis. Currently, adding the effects
of climate and fire, carbon stock changes have a cooling effect of
Published by NRC Research Press
Oris et al.
213
Table 6. Summary of the impacts of future increase in boreal forest fires on the carbon stock, albedo, and aerosols.
Region
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
Carbon stock changes
North America
Studied effect
Result
Reference
Theoretical productivity and respiration
with different growth rates
Boreal forest could be a carbon sink in the
future if decomposition rate does not
change
313 Tg of CO2 equivalent released per year
in the horizon 2100 compared to the
current release of 162 Tg of CO2
equivalent per year
473 Tg C released per year in 2100
compared to the current release of 276
Tg C/year
Loss of 27.1– 51.9 Pg carbon if there is a
50% increase in area burned
Loss of 17−19 Tg C/year in 2025−2099
Kurz et al. 2008
Canada
Carbon release by fires
Northern high latitude
ecosystems
Carbon emissions considering effects of
permafrost, CO2 fertilization, and fire
Boreal forest
Carbon dynamics depending on area
burned
Carbon dynamics with climate, fire, and
ecosystem interactions
Carbon stock changes depending on
two scenarios of area burned
Alaska
Managed forest in
Canada
North America
Radiative budget
Alaska
Carbon emissions and carbon stock
changes
Canadian forest will be a carbon sink with
a probability of 70% and 35% if there is
no change in area burned and if there is
an increase by a factor of 2–4 by 2100,
respectively
2.5- to 4.4-fold increase in carbon
emissions by 2091−2100; loss of 18−25 Tg
C/year in 2025−2099 in Alaska
North American boreal
forest
Changes in albedo under future fire
regimes (100%−200% increase in area
burned)
Radiative budget changes due to forest
fires
Changes in forest age distributions
increase summer albedo responsible for
a cooling of −0.9 W/m2/decade
A doubling of area burned would cool the
surface by 0.23 ± 0.09 °C during winter
and spring months (0.43 ± 0.12 °C for a
quadrupling of area burned)
Aerosols
Canada
Black carbon emissions
Emissions of 45 and 65 Gg/year for 2×CO2
and 3×CO2 scenarios respectively,
compared to the current value of 34 Gg/
year
Minor role in the reduction of continental
snow cover area; except in Quebec and
eastern Siberia when all the aerosol
forcings and atmospheric feedbacks
were included by 2050
Radiative forcing of −0.10 to 0.4 W/m2 by
2100 because of changes in future
aerosol abundance
90% increase of aerosol emissions by 2050
for a 54% increase in area burned
North of 30°N
Changes in snow cover due to black
carbon emissions
Global scale
Aerosol emissions
Western United States
Aerosol emissions
0.055 ± 0.096 W/m2/decade in the boreal region and 0.065 ±
0.062 W/m2/decade in North America. In the future, the effect of
climate change on carbon stock might be positive (Balshi et al.
2009) or negative (Kasischke et al. 1995; Zhuang et al. 2006). The
difference between these studies is the inclusion (or not) of carbon
release from permafrost thawing. The CO2 fertilization effect on
photosynthesis always has a positive effect (carbon storage). As a
result, only scenarios including the CO2 fertilization effect
have the potential to offset carbon losses from fires (and permafrost thawing), with an average cooling of 0.27 ± 0.14 W/m2/
decade in the future. The average of the other scenarios
(without CO2 fertilization) corresponds to a warming of 0.21 ±
0.04 W/m2/decade.
A higher proportion of deciduous forests in the landscape after
fire will change the surface albedo and cause climate cooling
(Table 6). In Alaska, this radiative forcing was estimated to
−0.09 W/m2/decade in 2003–2100 (Euskirchen et al. 2009). In North
America, changes in albedo and evapotranspiration due to a dou-
Amiro et al. 2009
Zhuang et al. 2006
Kasischke et al. 1995
Bachelet et al. 2005
Metsaranta et al. 2010
Balshi et al. 2009
Euskirchen et al.
2009
Rogers et al. 2013
Amiro et al. 2009
Ménégoz et al. 2013
Carslaw et al. 2010
Spracklen et al. 2009
bling of the area burned would cool the surface by 0.23 ± 0.09 °C
during winter and spring months (Rogers et al. 2013). The effect of
forest fires on carbon stock will be higher than that of albedo
(Fig. 2). Their sum will be a warming between 0.19 and 0.41 W/m2/
decade. Adding the climate and CO2 fertilization effects on carbon
stock change scenarios, future radiative forcing will range between −0.72 and 0.21 W/m2/decade (Fig. 2).
Uncertainty around the future increase in area burned leads to
important variations in estimates of radiative forcing by aerosols
(Fig. 3). On average, the direct effect of fire-produced aerosols will
exert a positive radiative forcing of 0.18 W/m2 by 2100 in the
boreal region. The effect will be smaller in Canada and Russia
compared with that in Alaska (Fig. 3).
Discussion
The fire-induced CO2 radiative forcing is estimated between
0.34 and 0.50 W/m2/decade at the end of the 21st century (Fig. 2).
Published by NRC Research Press
214
Environ. Rev. Vol. 22, 2014
Fig. 2. Summary of radiative forcing (W/m2/decade) due to changes in carbon uptake and (or) emission and land surface albedo in boreal
forests. Solid symbols represent positive radiative forcing and open symbols represent negative radiative forcing. Carbon releases were
converted in W/m2 as in Zhuang et al. (2006). Where appropriate, mean and standard errors were calculated. HAE, high anthropogenic CO2
emissions scenario; MAE, medium anthropogenic CO2 emissions scenario; *, with CO2 fertilization effect.
2
References
Radiative forcing (W/m /decade)
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
-1
-0.5
0
0.5
Balshi et al. (2007) Boreal region*
Hayes et al. (2011) Boreal region*
Metsanrata et al. (2010) Canada
Balshi et al. (2009) Scenario A2*
Balshi et al. (2009) Scenario B2*
Balshi et al. (2009) Scenario A2
Balshi et al. (2009) Scenario B2
Balshi et al. (2007) Boreal region*
Hayes et al. (2011) Boreal region*
Balshi et al. (2007) Boreal region
Zhuang et al. (2006) Boreal region
Zhuang et al. (2006) Scenario HAE*
Zhuang et al. (2006) Scenario MAE*
Zhuang et al. (2006) Scenario HAE
Zhuang et al. (2006) Scenario MAE
Kasischke et al. (1995) Boreal region
Hayes et al. (2011) North America*
Balshi et al. (2007) North America*
Balshi et al. (2007) North America
Balshi et al. (2009) Scenario B2*
Balshi et al. (2009) Scenario A2*
Balshi et al. (2009) Scenario A2
Balshi et al. (2009) Scenario B2
Euskirschen et al. (2009) Albedo
Present
Present
Future
Future
Canada
Boreal region
average
0.05
0.10
0.15
0.20
Fire effect
Combined effect of fire
and climate, with (*)
or without CO2 fertilization
(boreal region)
Future
Russia
0.00
2
Future
Alaska
-0.05
1.5
Present
Fig. 3. Summary of predicted radiative forcing (W/m2) due to
aerosols emitted by fires between today and 2100 in Alaska, Canada,
Russia, and the boreal region as a whole. Values do not represent
the radiative effect of the total burden but refer to the change in the
abundance of aerosols due to fire between today and 2100. To
calculate the radiative forcing, the hypothesis that changes in
aerosol emissions depend on changes in area burned was used. The
direct effect of aerosols was estimated using the method of Carslaw
et al. (2010). Future changes in aerosol concentration were
calculated using the ratio of the future change in emissions (90%) to
the future change in surface organic carbon concentration due to
fires (30% according to Spracklen et al. 2009). Changes in future
aerosol concentration were then multiplied by the current direct
radiative forcing of aerosols estimated by Forster et al. (2007) (from
–0.05 to 0.20 W/m2).
-0.10
1
0.25
0.30
0.35
0.40
Radiative forcing by aerosols (W/m 2 )
The impact on climate of carbon emissions from fires is primarily
measured by changes in carbon stock, which equate to the difference between carbon emissions and carbon storage by forests.
As CO2 is needed for photosynthesis, a CO2 fertilization effect
could increase the stock of terrestrial carbon captured by plants
(Amthor 1995; Long et al. 2004). Accounting for such a CO2 fertilization effect on photosynthesis, carbon storage could offset
warming by carbon emissions due to fire (Fig. 2; Zhuang et al.
2006). This negative feedback is uncertain, however, because the
response of vegetation to increased atmospheric CO2 concentration varies according to other factors (water and nutrient availability and species-specific responses) and a saturation point can
be reached (Cao and Woodward 1998; Canadell et al. 2007). Some
studies also suggested that CO2 fertilization might not be enough
to offset carbon losses due to disturbances (Kurz et al. 2008;
Combined effect of fire
and climate, with (*)
or without CO2 fertilization
(North America)
Albedo (Alaska)
Metsaranta et al. 2010). Moreover, permafrost thawing in response
to climate change will release into the atmosphere carbon that
has been trapped in soils for millennia (Schuur et al. 2009). When
permafrost thawing is included in the models, CO2 fertilization
does not offset carbon losses under low and medium anthropogenic CO2 emission scenarios for the 21st century (Zhuang et al.
2006). Balshi et al. (2009) did not include this effect in their model
and, therefore, estimated positive carbon stock changes.
Fire severity is an important parameter to consider when calculating carbon emissions (Turetsky et al. 2011). Studies (Kasischke
et al. 1995; Bachelet et al. 2005; Zhuang et al. 2006; Kurz et al. 2008;
Balshi et al. 2009; Metsaranta et al. 2010) on the carbon stock
changes do not take into account changes in fire severity within
the fire season, nor the average increase in fire severity due to
global warming, estimated at 6% by 2100 in Canada (Amiro et al.
2009). The radiative forcing due to fire-induced emissions of GHGs
is, therefore, underestimated. However, increased fire severity
could also induce an increase in albedo (Beck et al. 2011; Jin et al.
2012) and, therefore, climate cooling.
The direct effect of aerosols is higher than the variation in
carbon stock by fire emissions of GHGs and changes in albedo.
However, the effect of aerosols is temporary while the others are
long-lasting. Aerosols interact with the atmosphere in complex
ways, and thus important variations exist in the estimation of
radiative forcing, including negative and positive values (between
−0.075 and 0.33 W/m2 in 2100; Fig. 3). The effect of elemental
carbon deposition on snow and ice remains the only effect of
aerosols producing a positive radiative forcing, which would not
be offset in the future. This effect is, however, believed to be small
compared with the direct and indirect effects, but have greater
efficiency by accelerating snowmelt (Flanner et al. 2007). According to a recent simulation study, a future increase in biomass
burning will play a minor role in the reduction of continental
snow cover through snow darkening if atmospheric feedbacks are
not included (Ménégoz et al. 2013). However, in a scenario considering all the aerosol forcings and atmospheric feedbacks, a significant decrease of mean number of days with snow at the surface
was simulated in Quebec and eastern Siberia (Ménégoz et al. 2013).
If boreal forests do not support the predicted increase in fire frequency and (or) area burned, they might shift to a new, non-forest
stable state (Girard et al. 2008). Hence, the predicted increase in
forest fires might not be sustained. The Alaskan boreal forest was
Published by NRC Research Press
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
Oris et al.
likely not resilient to a climate change that occurred 5500 years ago
and has grown from a canopy dominated by white spruce to a dominance by black spruce (Hu et al. 2006). In eastern Canada, the boreal
forests gradually changed during the Holocene (e.g., 8000 years) and
have generally been resilient to changes in fire regime (Carcaillet
et al. 2010). However, a gradual opening of the northern Quebec
boreal landscape has resulted from higher fire frequencies during
the last 3000 years (Asselin and Payette 2005). In eastern Canada, the
increase in burn rate predicted by the end of the 21st century remains within the natural range of variability recorded in the past
(Bergeron et al. 2010). The effect of compound disturbances on forest
resilience, which may be more common in response to global warming, is also an important issue (Paine et al. 1998; Payette and Delwaide
2003). Compound disturbances (e.g., an insect outbreak closely followed by a fire) have caused regeneration failure which caused a shift
from a closed forest to an open woodland ecosystem (Payette et al.
2000; Jasinski and Payette 2005). Such an ecosystem change would
likely have an impact on climate through changes in albedo.
Uncertainties and limitations
The interaction between fire and vegetation is a major limitation in assessing the influence of boreal forest on climate. With
increasing temperature, species typical of temperate environments could migrate northward (Burton and Cumming 1995;
Terrier et al. 2013). These temperate species would act negatively
on fire spread due to their high moisture content and low flammability (Hély et al. 2000). A high proportion of broad-leaved species could have offset the effect of a warmer and drier climate on
fire activity during the mid-Holocene in eastern Canadian boreal
forests (Girardin et al. 2013). Models used to estimate the area
burned or future carbon emissions do not take into account
changes in vegetation (e.g., conifer versus broad-leaved). Yet, fire
occurrence and the resulting carbon emissions vary depending on
fuel type. To better understand the effect of increased fire on
climate, it is necessary to develop better prediction models of area
burned, taking into account vegetation dynamics.
A second challenge is to determine what amount of carbon will be
released as a result of soil decomposition and permafrost thawing
(Goulden et al. 1998; Gruber et al. 2004; Allison and Treseder 2011;
Koven et al. 2011). An increase in the decomposition rate of soil organic matter, which is directly proportional to ground temperature
(Lloyd and Taylor 1994), is predicted (Davidson and Janssens 2006;
Euskirchen et al. 2006; Allison and Treseder 2011), implying a potential decrease in net ecosystem productivity. The amplitude of this
positive feedback is uncertain (Dufresne et al. 2002), but could be
higher than the negative feedback due to CO2 fertilization (Cox et al.
2000). The 1%–26% loss of permafrost expected during the 21st century (Koven et al. 2012) would also release significant amounts of
carbon (Davidson and Janssens 2006; Allison and Treseder 2011;
Koven et al. 2011). O'Donnell et al. (2011) estimated that NPP would
have to increase by 7%–14% in the next two to three centuries to offset
carbon losses due to the interaction between fire and permafrost.
The third challenge is to take into account spatial variability in fire
regimes. Variations between and within countries in terms of temperature, precipitation, area burned, fire severity, carbon emissions
(with different severity and type of fuel), and carbon storage (forest
productivity in mixed, deciduous, coniferous forests and dense and
open forests) need further investigations. For example, Russia differs
from North America in various ways (Wooster and Zhang 2004; de
Groot et al. 2013a). Future research efforts should focus on the Russian boreal forest, which received considerably less attention so far.
Temporal variability is also important to take into account, as fire
regimes can differ markedly from year to year or over longer time
periods (Larsen 1997; Duncan et al. 2003; Bergeron et al. 2004;
Balzter et al. 2005; Kasischke and Turetsky 2006).
Finally, to determine the effect of aerosols on climate is a big
challenge. Difficulty in separating the origins of aerosols (anthro-
215
pogenic versus natural) complicates model calibration. Aerosol
production depends on numerous variables (e.g., temperature,
wind speed, precipitation, changes in antioxidants, trace gases,
and radiation) and is difficult to predict regionally. In addition,
the effect of aerosols depends on the relationship between their
abundance and concentration at the surface, which is not linear
(Spracklen et al. 2009). Coupled climate–chemistry models should
thus be developed. Data on long-term aerosol emissions by fire are
needed to calibrate the models.
Conclusion
It is not yet possible to simultaneously take into account all
fire effects on climate, but some conclusions can nevertheless be
made. In general, research findings point towards a future increase
in carbon emissions from boreal forest fires, with a likely greater
effect on climate than that of post-fire albedo change. However, possible limits to the CO2 fertilization effect on photosynthesis make
uncertain the future of the carbon stock in boreal forests. The carbon
stock also varies according to other disturbances, such as insect infestations and diffuse mortality (Kurz and Apps 1999; De Grandpré
et al. 2008). For example, the Canadian boreal forest is currently a
source of carbon because of the combined effects of fire and mountain pine beetle outbreaks (Natural Resources Canada 2012). Only the
direct aerosol effect was calculated in this review with a large range
of radiative forcings on climate. The indirect aerosol effect (CCN)
could be equivalent or greater, but with negative values (Forster et al.
2007). Elemental carbon deposited on snow and ice would thus be
the only uncompensated positive radiative forcings due to aerosols.
The partial results presented here suggest that boreal forest fires are
more likely to warm the climate than to cool it in the future, especially in North America.
Acknowledgements
Funding was provided by the Natural Sciences and Engineering
Research Council of Canada (NSERC Strategic Project Grant #41344411) and the Fonds québécois de recherche – Nature et Technologies
(FQRNT).
References
Albrecht, B.A. 1989. Aerosols, cloud microphysics, and fractional cloudiness.
Science, 245(4923): 1227–1230. doi:10.1126/science.245.4923.1227. PMID:
17747885.
Alexander, H., Mack, M., Goetz, S., Loranty, M., Beck, P.A., Earl, K., Zimov, S.,
Davydov, S., and Thompson, C. 2012. Carbon accumulation patterns during
post-fire succession in cajander larch (Larix cajanderi) forests of Siberia. Ecosystems, 15(7): 1065–1082. doi:10.1007/s10021-012-9567-6.
Allison, S.D., and Treseder, K.K. 2011. Climate change feedbacks to microbial
decomposition in boreal soils. Fungal Ecol. 4(6): 362–374. doi:10.1016/j.
funeco.2011.01.003.
Amiro, B.D. 2001. Paired-tower measurements of carbon and energy fluxes following disturbance in the boreal forest. Glob. Change Biol. 7(3): 253–268.
doi:10.1046/j.1365-2486.2001.00398.x.
Amiro, B.D., Todd, J.B., Wotton, B.M., Logan, K.A., Flannigan, M.D., Stocks, B.J.,
Mason, J.A., Martell, D.L., and Hirsch, K.G. 2001a. Direct carbon emissions
from Canadian forest fires, 1959–1999. Can. J. For. Res. 31(3): 512–525. doi:10.
1139/x00-197.
Amiro, B.D., Stocks, B.J., Alexander, M.E., Flannigan, M.D., and Wotton, B.M.
2001b. Fire, climate change, carbon and fuel management in the Canadian
boreal forest. Int. J. Wildland Fire, 10(3–4): 405–413. doi:10.1071/WF01038.
Amiro, B.D., MacPherson, J.I., Desjardins, R.L., Chen, J.M., and Liu, J. 2003. Postfire carbon dioxide fluxes in the western Canadian boreal forest: Evidence
from towers, aircraft and remote sensing. Agric. For. Meteorol. 115(1–2): 91–
107. doi:10.1016/S0168-1923(02)00170-3.
Amiro, B.D., Orchansky, A.L., Barr, A.G., Black, T.A., Chambers, S.D.,
Chapin, F.S., III, Goulden, M.L., Litvak, M., Liu, H.P., McCaughey, J.H.,
McMillan, A., and Randerson, J.T. 2006. The effect of post-fire stand age on the
boreal forest energy balance. Agric. For. Meteorol. 140(1–4): 41–50. doi:10.1016/
j.agrformet.2006.02.014.
Amiro, B.D., Cantin, A., Flannigan, M.D., and de Groot, W.J. 2009. Future emissions from Canadian boreal forest fires. Can. J. For. Res. 39(2): 383–395. doi:
10.1139/X08-154.
Amthor, J.S. 1995. Terrestrial higher-plant response to increasing atmospheric
[CO2] in relation to the global carbon cycle. Glob. Change Biol. 1(4): 243–274.
doi:10.1111/j.1365-2486.1995.tb00025.x.
Published by NRC Research Press
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
216
Andreae, M.O., and Merlet, P. 2001. Emission of trace gases and aerosols from
biomass burning. Global Biogeochem. Cycles, 15(4): 955–966.
Apps, M.J., Kurz, W.A., Luxmoore, R.J., Nilsson, L.O., Sedjo, R.A., Schmidt, R.,
Simpson, L.G., and Vinson, T.S. 1993. Boreal forests and tundra. Water, Air,
Soil Pollut. 70(1): 39–53. doi:10.1007/BF01104987.
Asselin, H., and Payette, S. 2005. Late Holocene opening of the forest tundra
landscape in northern Québec, Canada. Glob. Ecol. Biogeogr. 14(4): 307–313.
doi:10.1111/j.1466-822X.2005.00157.x.
Asselin, H., Belleau, A., and Bergeron, Y. 2006. Factors responsible for the cooccurrence of forested and unforested rock outcrops in the boreal forest.
Landsc. Ecol. 21(2): 271–280. doi:10.1007/s10980-005-1393-1.
Auclair, A.N.D., and Carter, T.B. 1993. Forest wildfires as a recent source of CO2 at
northern latitudes. Can. J. For. Res. 23(8): 1528–1536. doi:10.1139/x93-193.
Bachelet, D., Lenihan, J., Neilson, R., Drapek, R., and Kittel, T. 2005. Simulating
the response of natural ecosystems and their fire regimes to climatic variability in Alaska. Can. J. For. Res. 35(9): 2244–2257. doi:10.1139/x05-086.
Balshi, M.S., McGuire, A.D., Zhuang, Q., Melillo, J., Kicklighter, D.W.,
Kasischke, E., Wirth, C., Flannigan, M., Harden, J., Clein, J.S., Burnside, T.J.,
McAllister, J., Kurz, W.A., Apps, M., and Shvidenko, A. 2007. The role of
historical fire disturbance in the carbon dynamics of the pan-boreal region:
A process-based analysis. J. Geophys. Res. Biogeosci. 112(2). doi:10.1029/
2006JG000380.
Balshi, M.S., McGuire, A.D., Duffy, P., Flannigan, M., Walsh, J., and Melillo, J.
2008. Assessing the response of area burned to changing climate in western
boreal North America using a Multivariate Adaptive Regression Splines
(MARS) approach. Glob. Change Biol. 15(3): 578–600. doi:10.1111/j.1365-2486.
2008.01679.x.
Balshi, M.S., McGuire, A.D., Duffy, P., Flannigan, M., Kicklighter, D.W., and
Melillo, J. 2009. Vulnerability of carbon storage in North American boreal
forests to wildfires during the 21st century. Glob. Change Biol. 15(6): 1491–
1510. doi:10.1111/j.1365-2486.2009.01877.x.
Balzter, H., Gerard, F.F., George, C.T., Rowland, C.S., Jupp, T.E., McCallum, I.,
Shvidenko, A., Nilsson, S., Sukhinin, A., Onuchin, A., and Schmullius, C.
2005. Impact of the Arctic Oscillation pattern on interannual forest fire
variability in Central Siberia. Geophys. Res. Lett. 32(14): L14709.
Beck, P.S.A., Goetz, S.J., Mack, M.C., Alexander, H.D., Jin, Y., Randerson, J.T., and
Loranty, M.M. 2011. The impacts and implications of an intensifying fire
regime on Alaskan boreal forest composition and albedo. Glob. Change Biol.
17(9): 2853–2866. doi:10.1111/j.1365-2486.2011.02412.x.
Bergeron, Y., Flannigan, M., Gauthier, S., Leduc, A., and Lefort, P. 2004. Past,
current and future fire frequency in the Canadian boreal forest: implications
for sustainable forest management. AMBIO, 33(6): 356–360. doi:10.1579/00447447-33.6.356. PMID:15387074.
Bergeron, Y., Cyr, D., Drever, C.R., Flannigan, M., Gauthier, S., Kneeshaw, D.,
Lauzon, È., Leduc, A., Le Goff, H., Lesieur, D., and Logan, K. 2006. Past, current, and future fire frequencies in Quebec's commercial forests: Implications for the cumulative effects of harvesting and fire on age-class structure
and natural disturbance-based management. Can. J. For. Res. 36(11): 2737–
2744. doi:10.1139/x06-177.
Bergeron, Y., Cyr, D., Girardin, M.P., and Carcaillet, C. 2010. Will climate change
drive 21st century burn rates in Canadian boreal forest outside of its natural
variability: collating global climate model experiments with sedimentary
charcoal data. Int. J. Wildland Fire, 19(8): 1127–1139. doi:10.1071/WF09092.
Bergner, B., Johnstone, J.F., and Treseder, K.K. 2004. Experimental warming and
burn severity alter soil CO2 flux and soil functional groups in a recently
burned boreal forest. Glob. Change Biol. 10(12): 1996–2004. doi:10.1111/j.13652486.2004.00868.x.
Bernier, P.Y., Desjardins, R.L., Karimi-Zindashty, Y., Worth, D., Beaudoin, A.,
Luo, Y., and Wang, S. 2011. Boreal lichen woodlands: A possible negative
feedback to climate change in eastern North America. Agric. For. Meteorol.
151(4): 521–528. doi:10.1016/j.agrformet.2010.12.013.
Betts, A.K., and Ball, J.H. 1997. Albedo over the boreal forest. J. Geophys. Res.:
Atmos. 102(D24): 28901–28909. doi:10.1029/96JD03876.
Bonan, G.B. 2008. Forests and climate change: forcings, feedbacks, and the climate benefits of forests. Science, 320(5882): 1444–1449. doi:10.1126/science.
1155121. PMID:18556546.
Bond, T.C., Doherty, S.J., Fahey, D.W., Forster, P.M., Berntsen, T., DeAngelo, B.J.,
Flanner, M.G., Ghan, S., Kärcher, B., Koch, D., Kinne, S., Kondo, Y.,
Quinn, P.K., Sarofim, M.C., Schultz, M.G., Schulz, M., Venkataraman, C.,
Zhang, H., Zhang, S., Bellouin, N., Guttikunda, S.K., Hopke, P.K.,
Jacobson, M.Z., Kaiser, J.W., Klimont, Z., Lohmann, U., Schwarz, J.P.,
Shindell, D., Storelvmo, T., Warren, S.G., and Zender, C.S. 2013. Bounding the
role of black carbon in the climate system: A scientific assessment. J. Geophys. Res.: Atmos. 118(11): 5380–5552. doi:10.1002/jgrd.50171.
Bond-Lamberty, B., Wang, C., and Gower, S.T. 2004. Net primary production and
net ecosystem production of a boreal black spruce wildfire chronosequence.
Glob. Change Biol. 10(4): 473–487. doi:10.1111/j.1529-8817.2003.0742.x.
Bremond, L., Carcaillet, C., Favier, C., Ali, A.A., Paitre, C., Bégin, Y., Bergeron, Y.,
and Richard, P.J.H. 2010. Effects of vegetation zones and climatic changes on
fire-induced atmospheric carbon emissions: A model based on paleodata. Int.
J. Wildland Fire, 19(8): 1015–1025. doi:10.1071/WF09096.
Brown, C.D., and Johnstone, J.F. 2011. How does increased fire frequency affect
Environ. Rev. Vol. 22, 2014
carbon loss from fire? A case study in the northern boreal forest. Int. J.
Wildland Fire 20(7): 829–837. doi:10.1071/WF10113.
Brown, R.J.E., Wein, R.W., and Maclean, D.A. 1983. Effects of fire on the permafrost ground thermal regime. John Wiley & Sons, Chichester, UK.
Burn, C.R. 1998. The response (1958–1997) of permafrost and near-surface
ground temperatures to forest fire, Takhini River valley, southern Yukon
Territory. Can. J. Earth Sci. 35(2): 184–199. doi:10.1139/e97-105.
Burn, C.R., Mackay, J.R., and Kokelj, S.V. 2009. The thermal regime of permafrost
and its susceptibility to degradation in upland terrain near Inuvik, N.W.T.
Permafrost and Periglacial Processes, 20(2): 221–227. doi:10.1002/ppp.649.
Burton, P.J., and Cumming, S.G. 1995. Potential effects of climatic change on
some western Canadian forests, based on phenological enhancements to a
patch model of forest succession. Water, Air, Soil Pollut. 82(1–2): 401–414.
doi:10.1007/BF01182850.
Cai, W., Yang, J., Liu, Z., Hu, Y., and Weisberg, P.J. 2013. Post-fire tree recruitment
of a boreal larch forest in Northeast China. For. Ecol. Manage. 307: 20–29.
doi:10.1016/j.foreco.2013.06.056.
Canadell, J.G., Pataki, D., Gifford, R., Houghton, R.A., Lou, Y., Raupach, M.R.,
Smith, P., and Steffen, W. 2007. Saturation of the terrestrial carbon sink. In
Terrestrial Ecosystems in a Changing World. Edited by J.G. Canadell, D. Pataki,
and L. Pitelka. The IGBP Series. Springer-Verlag, Berlin Heidelberg. pp. 59–78.
Cao, M., and Woodward, F.I. 1998. Dynamic responses of terrestrial ecosystem
carbon cycling to global climate change. Nature, 393(6682): 249–252. doi:10.
1038/30460.
Carcaillet, C., Almquist, H., Asnong, H., Bradshaw, R.H.W., Carrión, J.S.,
Gaillard, M.J., Gajewski, K., Haas, J.N., Haberle, S.G., Hadorn, P., Müller, S.D.,
Richard, P.J.H., Richoz, I., Rösch, M., Sánchez Goi, M.F., Von Stedingk, H.,
Stevenson, A.C., Talon, B., Tardy, C., Tinner, W., Tryterud, E., Wick, L., and
Willis, K.J. 2002. Holocene biomass burning and global dynamics of the carbon cycle. Chemosphere, 49(8): 845–863. doi:10.1016/S0045-6535(02)00385-5.
PMID:12430662.
Carcaillet, C., Richard, P.J.H., Bergeron, Y., Fréchette, B., and Ali, A.A. 2010.
Resilience of the boreal forest in response to Holocene fire-frequency
changes assessed by pollen diversity and population dynamics. Int. J. Wildland Fire, 19(8): 1026–1039. doi:10.1071/WF09097.
Carslaw, K.S., Boucher, O., Spracklen, D.V., Mann, G.W., Rae, J.G.L.,
Woodward, S., and Kulmala, M. 2010. Atmospheric aerosols in the earth
system: A review of interactions and feedbacks. Atmos. Chem. Phys. 10:
1701–1737. doi:10.5194/acp-10-1701-2010.
Chapin, F.S., III, Woodwell, G.M., Randerson, J.T., Rastetter, E.B., Lovett, G.M.,
Baldocchi, D.D., Clark, D.A., Harmon, M.E., Schimel, D.S., Valentini, R.,
Wirth, C., Aber, J.D., Cole, J.J., Goulden, M.L., Harden, J.W., Heimann, M.,
Howarth, R.W., Matson, P.A., McGuire, A.D., Melillo, J.M., Mooney, H.A.,
Neff, J.C., Houghton, R.A., Pace, M.L., Ryan, M.G., Running, S.W., Sala, O.E.,
Schlesinger, W.H., and Schulze, E.D. 2006. Reconciling carbon-cycle concepts, terminology, and methods. Ecosystems, 9(7): 1041–1050. doi:10.1007/
s10021-005-0105-7.
Chen, W., Chen, J.M., Price, D.T., and Cihlar, J. 2002. Effects of stand age on net
primary productivity of boreal black spruce forests in Ontario, Canada. Can.
J. For. Res. 32(5): 833–842. doi:10.1139/x01-165.
Christensen, J.H., Hewitson, B., Busuioc, A., Chen, A., Gao, X., Held, I., Jones, R.,
Kolli, R.K., Kwon, W.-T., Laprise, R., Magaña Rueda, V., Mearns, L., Menéndez,
C.G., Räisänen, J., Rinke, A., Sarr, A., and Whetton, P. 2007. Regional Climate
Projections. In Climate Change 2007: The Physical Science Basis. Contribution of Working Group I to the 4th Assessment Report of the Intergovernmental Panel on Climate Change. Edited by S. Solomon, D. Qin, M. Manning,
Z. Chen, M. Marquis, K.B. Averyt, M. Tignor, and H.L. Miller. Cambridge
University Press, Cambridge, United Kingdom and New York, NY, USA.
Chuang, C.C., Penner, J.E., Prospero, J.M., Grant, K.E., Rau, G.H., and
Kawamoto, K. 2002. Cloud susceptibility and the first aerosol indirect forcing: Sensitivity to black carbon and aerosol concentrations. J. Geophys. Res.:
Atmos. 107(D21): AAC 10-11–AAC 10-23. doi:10.1029/2000JD000215.
Cofer, W.R., III, Winstead, E.L., Stocks, B.J., Goldammer, J.G., and Cahoon, D.R.
1998. Crown fire emissions of CO2, CO, H2, CH4, and TNMHC from a dense
jack pine boreal forest fire. Geophys. Res. Lett. 25(21): 3919–3922. doi:10.1029/
1998GL900042.
Conard, S.G., and Ivanova, G.A. 1997. Wildfire in Russian boreal forests –
Potential impacts of fire regime characteristics on emissions and global carbon balance estimates. Environ. Pollut. 98(3): 305–313. doi:10.1016/S02697491(97)00140-1.
Conard, S.G., Sukhinin, A.I., Stocks, B.J., Cahoon, D.R., Davidenko, E.P., and
Ivanova, G.A. 2002. Determining effects of area burned and fire severity on
carbon cycling and emissions in Siberia. Clim. Change 55(1–2): 197–211. doi:
10.1023/A:1020207710195.
Cox, P.M., Betts, R.A., Jones, C.D., Spall, S.A., and Totterdell, I.J. 2000. Acceleration of global warming due to carbon-cycle feedbacks in a coupled climate
model. Nature, 408(6809): 184–187. doi:10.1038/35041539. PMID:11089968.
Crutzen, P.J., and Carmichael, G.J. 1993. Modeling the influence of fires on atmospheric chemistry. In Fire in the environment: the ecological, atmospheric, and climatic importance of vegetation fires. Edited by P.J. Crutzen,
and J.G. Goldammer. Wiley, New York. pp. 89–105.
Czimczik, C.I., Schmidt, M.W.I., and Schulze, E.D. 2005. Effects of increasing
fire frequency on black carbon and organic matter in Podzols of Siberian
Published by NRC Research Press
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
Oris et al.
Scots pine forests. Eur. J. Soil Sci. 56(3): 417–428. doi:10.1111/j.1365-2389.
2004.00665.x.
Davidson, E.A., and Janssens, I.A. 2006. Temperature sensitivity of soil carbon
decomposition and feedbacks to climate change. Nature, 440(7081): 165–173.
doi:10.1038/nature04514. PMID:16525463.
De Grandpré, L., Gauthier, S., Allain, C., Cyr, D., Périgon, S., Pham, A.T., Boucher, D.,
Morissette, J., Reyes, G., Aakala, T., and Kuuluvaien, T. 2008. Vers un aménagement écosystémique de la forêt boréale de la Côte Nord. In Aménagement écosystémique en forêt boréale. Edited by S. Gauthier, M.-A. Vaillancourt, A. Leduc, L.
De Grandpré, D. Kneeshaw, H. Morin, P. Drapeau, and Y. Bergeron. Les Presses de
l'Université du Québec, Québec, Canada. pp. 242–268.
de Groot, W.J., Cantin, A.S., Flannigan, M.D., Soja, A.J., Gowman, L.M., and
Newbery, A. 2013a. A comparison of Canadian and Russian boreal forest fire
regimes. For. Ecol. Manage. 294: 23–34. doi:10.1016/j.foreco.2012.07.033.
de Groot, W.J., Flannigan, M.D., and Cantin, A.S. 2013b. Climate change impacts
on future boreal fire regimes. For. Ecol. Manage. 294: 35–44. doi:10.1016/j.
foreco.2012.09.027.
de Lafontaine, G., and Asselin, H. 2011. Soil charcoal stability over the Holocene
across boreal northeastern North America. Quat. Res. 76(2): 196–200. doi:10.
1016/j.yqres.2011.06.006.
Delong, C.S., and Tanner, D. 1996. Managing the pattern of forest harvest:
lessons from wildfire. Biodivers. Conserv. 5(10): 1191–1205. doi:10.1007/
BF00051571.
Dixon, R.K., and Krankina, O.N. 1993. Forest fires in Russia: carbon dioxide emissions to the atmosphere. Can. J. For. Res. 23(4): 700–705. doi:10.1139/x93-091.
Dragotescu, I., and Kneeshaw, D.D. 2012. A comparison of residual forest following fires and harvesting in boreal forests in Quebec, Canada. Silva Fenn.
46(3): 365–376.
Dufresne, J.L., Fairhead, L., Le Treut, H., Berthelot, M., Bopp, L., Ciais, P.,
Friedlingstein, P., and Monfray, P. 2002. On the magnitude of positive feedback between future climate change and the carbon cycle. Geophys. Res. Lett.
29(10): 43-41–43-44. doi:10.1029/2001GL013777.
Duncan, B.N., Martin, R.V., Staudt, A.C., Yevich, R., and Logan, J.A. 2003. Interannual and seasonal variability of biomass burning emissions constrained by
satellite observations. J. Geophys. Res.: Atmos. 108(D2): 4100. doi:10.1029/
2002JD002378.
Eugster, W., Rouse, W.R., Pielke, R.A., Sr., McFadden, J.P., Baldocchi, D.D.,
Kittel, T.G.F., Chapin, F.S., III, Liston, G.E., Vidale, P.L., Vaganov, E., and
Chambers, S. 2000. Land-atmosphere energy exchange in Arctic tundra and
boreal forest: Available data and feedbacks to climate. Glob. Change Biol.
6(Suppl. 1): 84–115. doi:10.1046/j.1365-2486.2000.06015.x.
Euskirchen, E.S., McGuire, A.D., Kicklighter, D.W., Zhuang, Q., Clein, J.S.,
Dargaville, R.J., Dye, D.G., Kimball, J.S., McDonald, K.C., Melillo, J.M.,
Romanovsky, V.E., and Smith, N.V. 2006. Importance of recent shifts in soil
thermal dynamics on growing season length, productivity, and carbon sequestration in terrestrial high-latitude ecosystems. Glob. Change Biol. 12(4):
731–750. doi:10.1111/j.1365-2486.2006.01113.x.
Euskirchen, E.S., McGuire, A.D., Rupp, T.S., Chapin, F.S., III, and Walsh, J.E. 2009.
Projected changes in atmospheric heating due to changes in fire disturbance
and the snow season in the western Arctic, 2003-2100. J. Geophys. Res. Biogeosci. 114(4). doi:10.1029/2009JG001095.
Euskirchen, E.S., McGuire, A.D., Chapin, F.S., III, and Rupp, T.S. 2010. The changing effects of Alaskas boreal forests on the climate system. Can. J. For. Res.
40(7): 1336–1346. doi:10.1139/X09-209.
Flanner, M.G., Zender, C.S., Randerson, J.T., and Rasch, P.J. 2007. Present-day
climate forcing and response from black carbon in snow. J. Geophys. Res.:
Atmos. 112(11). doi:10.1029/2006JD008003.
Flannigan, M.D., Logan, K.A., Amiro, B.D., Skinner, W.R., and Stocks, B.J. 2005.
Future area burned in Canada. Clim. Change, 72(1–2): 1–16. doi:10.1007/
s10584-005-5935-y.
Flannigan, M.D., Krawchuk, M.A., de Groot, W.J., Wotton, B.M., and
Gowman, L.M. 2009. Implications of changing climate for global wildland
fire. Int. J. Wildland Fire, 18(5): 483–507. doi:10.1071/WF08187.
Forster, P., Ramaswamy, V., Artaxo, P., Berntsen, T., Betts, R., Fahey, D.W.,
Haywood, J., Lean, J., Lowe, D.C., Myhre, G., Nganga, J., Prinn, R., Raga, G.,
Schulz, M., and Van Dorland, R. 2007. Changes in atmospheric constituents
and in radiative forcing. In Climate Change 2007: The Physical Science Basis.
Contribution of Working Group I to the 4th Assessment Report of the Intergovernmental Panel on Climate Change. Edited by S. Solomon, D. Qin, M. Manning,
Z. Chen, M. Marquis, K.B. Averyt, M. Tignor, and H.L. Miller. Cambridge University Press, Cambridge, United Kingdom and New York, NY, USA.
Fréjaville, T., Carcaillet, C., and Curt, T. 2013. Calibration of charcoal production
from trees biomass for soil charcoal analyses in subalpine ecosystems. Quat.
Int. 289: 16–23. doi:10.1016/j.quaint.2012.02.043.
French, N.H.F., Kasischke, E.S., Stocks, B.J., Mudd, J.P., Martell, D.L., and Lee, B.S.
2000. Carbon release from fires in the north american boreal forest. In
Fire, Climate Change, and Carbon Cycling in the Boreal Forest. Edited by
E.S. Kasischke, and B.J. Stocks. Springer, New York. pp. 377–388.
French, N.H.F., Kasischke, E.S., and Williams, D.G. 2002. Variability in the emission of carbon-based trace gases from wildfire in the Alaskan boreal forest.
J. Geophys. Res.: Atmos. 107(D1): FFR 7-1–FFR 7-11. doi:10.1029/2001JD000480.
Fritze, H., Smolander, A., Levula, T., Kitunen, V., and Mälkönen, E. 1994. Woodash fertilization and fire treatments in a Scots pine forest stand: Effects on
217
the organic layer, microbial biomass, and microbial activity. Biol. Fertil.
Soils, 17(1): 57–63. doi:10.1007/BF00418673.
Genet, H., McGuire, A.D., Barrett, K., Breen, A., Euskirchen, E.S., Johnstone, J.F.,
Kasischke, E.S., Melvin, A.M., Bennett, A., Mack, M.C., Rupp, T.S.,
Schuur, A.E.G., Turetsky, M.R., and Yuan, F. 2013. Modeling the effects of fire
severity and climate warming on active layer thickness and soil carbon storage of black spruce forests across the landscape in interior Alaska. Environ.
Res. Lett. 8(4): 045016. doi:10.1088/1748-9326/8/4/045016.
Girard, F., Payette, S., and Gagnon, R. 2008. Rapid expansion of lichen woodlands
within the closed-crown boreal forest zone over the last 50 years caused by
stand disturbances in eastern Canada. J. Biogeogr. 35(3): 529–537. doi:10.1111/
j.1365-2699.2007.01816.x.
Girardin, M.P., Ali, A.A., Carcaillet, C., Blarquez, O., Hely, C., Terrier, A.,
Genries, A., and Bergeron, Y. 2013. Vegetation limits the impact of a warm
climate on boreal wildfires. New Phytol. 199(4): 1001–1011. doi:10.1111/nph.
12322. PMID:23691916.
Goetz, S.J., MacK, M.C., Gurney, K.R., Randerson, J.T., and Houghton, R.A. 2007.
Ecosystem responses to recent climate change and fire disturbance at northern high latitudes: Observations and model results contrasting northern
Eurasia and North America. Environ. Res. Lett. 2(4): 045031. doi:10.1088/17489326/2/4/045031.
Goulden, M.L., Wofsy, S.C., Harden, J.W., Trumbore, S.E., Crill, P.M., Gower, S.T.,
Fries, T., Daube, B.C., Fan, S.-M., Sutton, D.J., Bazzaz, A., and Munger, J.W.
1998. Sensitivity of boreal forest carbon balance to soil thaw. Science,
279(5348): 214–217. doi:10.1126/science.279.5348.214. PMID:9422691.
Goulden, M.L., McMillan, A.M.S., Winston, G.C., Rocha, A.V., Manies, K.L.,
Harden, J.W., and Bond-Lamberty, B.P. 2011. Patterns of NPP, GPP, respiration, and NEP during boreal forest succession. Glob. Change Biol. 17(2): 855–
871. doi:10.1111/j.1365-2486.2010.02274.x.
Gower, S.T., McMurtrie, R.E., and Murty, D. 1996. Aboveground net primary
production decline with stand age: potential causes. Trends Ecol. Evol. 11(9):
378–382. doi:10.1016/0169-5347(96)10042-2. PMID:21237883.
Greene, D.F., and Johnson, E.A. 1999. Modelling recruitment of Populus tremuloides, Pinus banksiana, and Picea mariana following fire in the mixedwood
boreal forest. Can. J. For. Res. 29(4): 462–473. doi:10.1139/x98-211.
Greene, D.F., Macdonald, S.E., Haeussler, S., Domenicano, S., Noël, J., Jayen, K.,
Charron, I., Gauthier, S., Hunt, S., Gielau, E.T., Bergeron, Y., and Swift, L.
2007. The reduction of organic-layer depth by wildfire in the North American
boreal forest and its effect on tree recruitment by seed. Can. J. For. Res. 37(6):
1012–1023. doi:10.1139/X06-245.
Gruber, N., Friedlingstein, P., Field, C.B., Valentini, R., Heimann, M., Richey, J.E.,
Romero-Lankao, P., Schulze, E.-D., and Chen, C.-T.A. 2004. The vulnerability
of the carbon cycle in the 21st Century: An assessment of carbon-climatehuman interactions. In The global carbon cycle: integrating humans, climate, and the natural world. Edited by C.B. Field, and M.R. Raupach. Island
Press, Washington, D.C. pp. 45–76.
Hansen, J.E., Lacis, A.A., Lee, P., and Wang, W.-C. 1980. Climatic effects of atmospheric aerosols. Ann. N.Y. Acad. Sci. 338(1): 575–587. doi:10.1111/j.1749-6632.
1980.tb17151.x.
Hansen, J., Sato, M., and Ruedy, R. 1997. Radiative forcing and climate response.
J. Geophys. Res.: Atmos. 102(D6): 6831–6864. doi:10.1029/96JD03436.
Harden, J.W., Trumbore, S.E., Stocks, B.J., Hirsch, A., Gower, S.T., O’Neill, K.P.,
and Kasischke, E.S. 2000. The role of fire in the boreal carbon budget. Glob.
Change Biol. 6(Suppl. 1): 174–184. doi:10.1046/j.1365-2486.2000.06019.x.
Harden, J.W., Meier, R., Silapaswan, C., Swanson, D.K., and McGuire, A.D. 2003. Soil
drainage and its potential for influencing wildfires in Alaska. In Studies by the
US Geological Survey in Alaska, 2001, US Geological Survey Professional Paper
1678. Edited by J. Galloway. US Geological Survey, Menlo Park, CA. pp. 139–144.
Harmon, M.E., Ferrell, W.K., and Franklin, J.F. 1990. Effects on carbon storage of
conversion of old-growth forests to young forests. Science, 247(4943): 699–
702. doi:10.1126/science.247.4943.699. PMID:17771887.
Hart, S., and Luckai, N. 2013. Charcoal function and management in boreal
ecosystems. J. Appl. Ecol. 50(5): 1197–1206. doi:10.1111/1365-2664.12136.
Hayes, D.J., McGuire, A.D., Kicklighter, D.W., Gurney, K.R., Burnside, T.J., and
Melillo, J.M. 2011. Is the northern high-latitude land-based CO2 sink weakening? Global Biogeochem. Cycles, 25(3). doi:10.1029/2010GB003813.
Hély, C., Bergeron, Y., and Flannigan, M.D. 2000. Effects of stand composition on
fire hazard in mixed-wood Canadian boreal forest. J. Veg. Sci. 11(6): 813–824.
doi:10.2307/3236551.
Hu, F., Brubaker, L., Gavin, D., Higuera, P., Lynch, J., Rupp, T., and Tinner, W.
2006. How climate and vegetation influence the fire regime of the alaskan
boreal biome: The holocene perspective. Mitig. Adapt. Strategies Glob.
Change, 11(4): 829–846. doi:10.1007/s11027-005-9015-4.
IPCC. 2007. Climate Change 2007: Synthesis Report, Contribution of Working
Groups I, II and III to the 4th Assessment Report of the Intergovernmental
Panel on Climate Change. Edited by Core Writing Team, R.K. Pachauri, and
A. Reisinger. IPCC, Geneve, Switzerland.
Jacobson, M.Z. 2001. Strong radiative heating due to the mixing state of black
carbon in atmospheric aerosols. Nature, 409(6821): 695–697. doi:10.1038/
35055518. PMID:11217854.
Jasinski, J.P., and Payette, S. 2005. The creation of alternative stable states in the
southern boreal forest, Québec, Canada. Ecol. Monogr. 75(4): 561–583. doi:10.
1890/04-1621.
Published by NRC Research Press
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
218
Jin, Y.F., Randerson, J.T., Goetz, S.J., Beck, P.S.A., Loranty, M.M., and
Goulden, M.L. 2012. The influence of burn severity on postfire vegetation
recovery and albedo change during early succession in North American boreal forests. J. Geophys. Res.: Biogeosci. 117. doi:10.1029/2011jg001886.
Johnson, E.A. 1992. Fire and vegetation dynamics: studies from the North American boreal forest. Cambridge University Press.
Johnstone, J.F., and Chapin, F.S., III. 2006a. Fire interval effects on successional
trajectory in boreal forests of northwest Canada. Ecosystems, 9(2): 268–277.
doi:10.1007/s10021-005-0061-2.
Johnstone, J.F., and Chapin, F.S., III. 2006b. Effects of soil burn severity on postfire tree recruitment in boreal forest. Ecosystems, 9(1): 14–31. doi:10.1007/
s10021-004-0042-x.
Johnstone, J.F., and Kasischke, E.S. 2005. Stand-level effects of soil burn severity
on postfire regeneration in a recently burned black spruce forest. Can. J. For.
Res. 35(9): 2151–2163. doi:10.1139/x05-087.
Kashian, D.M., Romme, W.H., Tinker, D.B., Turner, M.G., and Ryan, M.G. 2006.
Carbon storage on landscapes with stand-replacing fires. BioScience, 56(7):
598–606. doi:10.1641/0006-3568(2006)56[598:CSOLWS]2.0.CO;2.
Kasischke, E.S., and Bruhwiler, L.P. 2003. Emissions of carbon dioxide, carbon
monoxide, and methane from boreal forest fires in 1998. J. Geophys. Res.:
Atmos. 108(1): FFR 2-1–FFR 2-14. doi:10.1029/2001JD000461.
Kasischke, E.S., Christensen, N.L., Jr., and Stocks, B.J. 1995. Fire, global warming,
and the carbon balance of boreal forests. Ecol. Appl. 5(2): 437–451. doi:10.
2307/1942034.
Kasischke, E.S., and Turetsky, M.R. 2006. Recent changes in the fire regime
across the North American boreal region – Spatial and temporal patterns of
burning across Canada and Alaska. Geophys. Res. Lett. 33(9). doi:10.1029/
2006GL025677.
Kasischke, E.S., Stocks, B.J., O’Neill, K., French, N.H.F., and Bourgeau-Chavez, L.L.
2000a. Direct effects of fire on the boreal forest carbon budget. In Biomass
Burning and Its Inter-Relationships with the Climate System. Edited by J.L.
Innes, M. Beniston, and M.M. Verstraete. Springer, Dordrecht. pp. 51–68.
Kasischke, E.S., O’Neill, K., Bourgeau-Chavez, L., and French, N. 2000b. Indirect
and long-term effects of fire on the boreal forest carbon budget biomass
burning and its inter-relationships with the climate system. In Biomass Burning and Its Inter-Relationships with the Climate System. Edited by J.L. Innes,
M. Beniston, and M.M. Verstraete. Springer Netherlands. pp. 263–280.
Kasischke, E.S., Hyer, E.J., Novelli, P.C., Bruhwiler, L.P., French, N.H.F.,
Sukhinin, A.I., Hewson, J.H., and Stocks, B.J. 2005. Influences of boreal fire
emissions on Northern Hemisphere atmospheric carbon and carbon monoxide. Global Biogeochem. Cycles, 19(1): 1–16. doi:10.1029/2004GB002300.
Kim, Y., and Tanaka, N. 2003. Effect of forest fire on the fluxes of CO2, CH4 and
N2O in boreal forest soils, interior Alaska. J. Geophys. Res.: Atmos. 108(D1).
doi:10.1029/2001jd000663.
Koch, D., and Del Genio, A.D. 2010. Black carbon semi-direct effects on cloud
cover: review and synthesis. Atmos. Chem. Phys. 10(16): 7685–7696. doi:10.
5194/acp-10-7685-2010.
Koven, C.D., Ringeval, B., Friedlingstein, P., Ciais, P., Cadule, P.,
Khvorostyanov, D., Krinner, G., and Tarnocai, C. 2011. Permafrost carbonclimate feedbacks accelerate global warming. Proc. Natl. Acad. Sci. U.S.A.
108(36): 14769–14774. doi:10.1073/pnas.1103910108. PMID:21852573.
Koven, C.D., Riley, W.J., and Stern, A. 2012. Analysis of permafrost thermal
dynamics and response to climate change in the CMIP5 Earth system models.
J. Clim. 26(6): 1877–1900. doi:10.1175/JCLI-D-12-00228.1.
Kurz, W.A., and Apps, M.J. 1999. A 70-year retrospective analysis of carbon fluxes
in the Canadian forest sector. Ecol. Appl. 9(2): 526–547. doi:10.1890/10510761(1999)009[0526:AYRAOC]2.0.CO;2.
Kurz, W.A., Stinson, G., and Rampley, G. 2008. Could increased boreal forest
ecosystem productivity offset carbon losses from increased disturbances?
Philos. Trans. R. Soc. Acad. B, 363(1501): 2261–2269. doi:10.1098/rstb.2007.
2198. PMID:18029308.
Larsen, C.P.S. 1997. Spatial and temporal variations in boreal forest fire frequency in northern Alberta. J. Biogeogr. 24(5): 663–673. doi:10.1111/j.1365-2699.
1997.tb00076.x.
Lavoué, D., Liousse, C., Cachier, H., Stocks, B.J., and Goldammer, J.G. 2000.
Modeling of carbonaceous particles emitted by boreal and temperate wildfires at northern latitudes. J. Geophys. Res.: Atmos. 105(D22): 26871–26890.
doi:10.1029/2000JD900180.
Lecomte, N., Simard, M., Fenton, N., and Bergeron, Y. 2006. Fire severity and
long-term ecosystem biomass dynamics in coniferous boreal forests of eastern Canada. Ecosystems, 9(8): 1215–1230. doi:10.1007/s10021-004-0168-x.
Liu, H., and Randerson, J.T. 2008. Interannual variability of surface energy
exchange depends on stand age in a boreal forest fire chronosequence.
J. Geophys. Res. 113(G1): G01006. doi:10.1029/2007JG000483.
Liu, H., Randerson, J.T., Lindfors, J., and Chapin, F.S., III. 2005. Changes in the surface
energy budget after fire in boreal ecosystems of interior Alaska: an annual
perspective. J. Geophys. Res.: Atmos. 110(13). doi:10.1029/2004JD005158.
Lloyd, J., and Taylor, J.A. 1994. On the temperature dependence of soil respiration. Funct. Ecol. 8(3): 315–323. doi:10.2307/2389824.
Lobert, J.M., and Warnatz, J. 1993. Emissions from the combustion process in
vegetation. In Fire in the environment: the ecological, atmospheric, and climatic
importance of vegetation fires Edited by P.J. Crutzen, and J.G. Goldammer. Wiley,
New York. pp. 15–37.
Environ. Rev. Vol. 22, 2014
Logan, T., Charron, I., Chaumont, D., and Houle, D. 2011. Atlas de scénarios
climatiques pour la forêt québécoise. Ouranos and MRNF.
Lohmann, U., and Feichter, J. 2005. Global indirect aerosol effects: a review.
Atmos. Chem. Phys. 5(3): 715–737. doi:10.5194/acp-5-715-2005.
Long, S.P., Ainsworth, E.A., Rogers, A., and Ort, D.R. 2004. Rising atmospheric
carbon dioxide: plants face the future. Annu. Rev. Plant Biol. 55(1): 591–628.
doi:10.1146/annurev.arplant.55.031903.141610. PMID:15377233.
Lukeš, P., Stenberg, P., and Rautiainen, M. 2013. Relationship between forest
density and albedo in the boreal zone. Ecol. Model. 261–262: 74–79. doi:10.
1016/j.ecolmodel.2013.04.009.
Luyssaert, S., Schulze, E.D., Borner, A., Knohl, A., Hessenmoller, D., Law, B.E.,
Ciais, P., and Grace, J. 2008. Old-growth forests as global carbon sinks. Nature,
455(7210): 213–215. doi:10.1038/nature07276. PMID:18784722.
Lyons, E.A., Jin, Y., and Randerson, J.T. 2008. Changes in surface albedo after fire
in boreal forest ecosystems of interior Alaska assessed using MODIS satellite
observations. J. Geophys. Res. 113(G2): G02012. doi:10.1029/2007JG000606.
Mackay, J.R. 1995. Active layer changes (1968 to 1993) following the forest-tundra
fire near Inuvik, N.W.T., Canada. Arct. Alp. Res. 27(4): 323–336. doi:10.2307/
1552025.
Madoui, A., Leduc, A., Gauthier, S., and Bergeron, Y. 2010. Spatial pattern analyses of post-fire residual stands in the black spruce boreal forest of western
Quebec. Int. J. Wildland Fire, 19(8): 1110–1126. doi:10.1071/WF10049.
Malevsky-Malevich, S.P., Molkentin, E.K., Nadyozhina, E.D., and Shklyarevich, O.B.
2008. An assessment of potential change in wildfire activity in the Russian
boreal forest zone induced by climate warming during the twenty-first century. Clim. Change 86(3–4): 463–474. doi:10.1007/s10584-007-9295-7.
Mansuy, N., Gauthier, S., Robitaille, A., and Bergeron, Y. 2010. The effects of
surficial deposit–drainage combinations on spatial variations of fire cycles in
the boreal forest of eastern Canada. Int. J. Wildland Fire, 19(8): 1083–1098.
doi:10.1071/WF09144.
Ménégoz, M., Krinner, G., Balkanski, Y., Cozic, A., Boucher, O., and Ciais, P. 2013.
Boreal and temperate snow cover variations induced by black carbon emissions in the middle of the 21st century. Cryosphere, 7(2): 537–554. doi:10.5194/
tc-7-537-2013.
Metsaranta, J.M., Kurz, W.A., Neilson, E.T., and Stinson, G. 2010. Implications of future
disturbance regimes on the carbon balance of Canada's managed forest
(2010–2100). Tellus B, 62(5): 719–728. doi:10.1111/j.1600-0889.2010.00487.x.
Murphy, P.r., Mudd, J.P., Stocks, B.J., Kasischke, E.S., Barry, D., Alexander, M.E.,
and French, N.H.F. 2000. Historical Fire Records in the North American Boreal Forest. In Fire, Climate Change, and Carbon Cycling in the Boreal Forest.
Edited by E.S. Kasischke and B.J. Stocks. Springer New York. pp. 274–288.
Myhre, G., Myhre, C.E.L., Samset, B.H., and Storelvmo, T. 2013. Aerosols and their
relation to global climate and climate sensitivity. Nature Education Knowledge, 4(5): 7.
Natural Resources Canada. 2012. The State of Canada's Forests (CFS). Annual
Report 2012. Ottawa. Natural Resources Canada.
O’Donnell, J.A., Harden, J.W., McGuire, A.D., Kanevskiy, M.Z., Jorgenson, M.T.,
and Xu, X. 2011. The effect of fire and permafrost interactions on soil carbon
accumulation in an upland black spruce ecosystem of interior Alaska: Implications for post-thaw carbon loss. Glob. Change Biol. 17(3): 1461–1474. doi:10.
1111/j.1365-2486.2010.02358.x.
O’Halloran, T.L., Law, B.E., Goulden, M.L., Wang, Z., Barr, J.G., Schaaf, C.,
Brown, M., Fuentes, J.D., Göckede, M., Black, A., and Engel, V. 2012. Radiative
forcing of natural forest disturbances. Glob. Change Biol. 18(2): 555–565.
doi:10.1111/j.1365-2486.2011.02577.x.
O’Neill, K.P., Kasischke, E.S., and Richter, D.D. 2003. Seasonal and decadal patterns of soil carbon uptake and emission along an age sequence of burned
black spruce stands in interior Alaska. J. Geophys. Res.: Atmos. 108(1): FFR
11-11–FFR 11-15. doi:10.1029/2001JD000443.
Paine, R.T., Tegner, M.J., and Johnson, E.A. 1998. Compounded perturbations
yield ecological surprises. Ecosystems, 1(6): 535–545. doi:10.1007/s100219900049.
Pan, Y., Birdsey, R.A., Fang, J., Houghton, R., Kauppi, P.E., Kurz, W.A.,
Phillips, O.L., Shvidenko, A., Lewis, S.L., Canadell, J.G., Ciais, P., Jackson, R.B.,
Pacala, S.W., McGuire, A.D., Piao, S., Rautiainen, A., Sitch, S., and Hayes, D.
2011. A large and persistent carbon sink in the World's forests. Science,
333(6045): 988–993. doi:10.1126/science.1201609. PMID:21764754.
Park, R.J., Jacob, D.J., Chin, M., and Martin, R.V. 2003. Sources of carbonaceous
aerosols over the United States and implications for natural visibility.
J. Geophys. Res.: Atmos. 108(D12): 4355. doi:10.1029/2002JD003190.
Payette, S. 1992. Fire as a controlling process in the North American boreal
forest. In A systems analysis of the global boreal forest. Edited by H.H. Shugart,
R. Leemans, and G.B. Bonan. Cambridge University Press, Cambridge, NY.
pp. 144–169.
Payette, S., and Delwaide, A. 2003. Shift of conifer boreal forest to lichen-heath
parkland caused by successive stand disturbances. Ecosystems, 6(6): 540–
550. doi:10.1007/PL00021507.
Payette, S., Bhiry, N., Delwaide, A., and Simard, M. 2000. Origin of the lichen
woodland at its southern range limit in eastern Canada: the catastrophic
impact of insect defoliators and fire on the spruce-moss forest. Can. J. For.
Res. 30(2): 288–305. doi:10.1139/x99-207.
Pregitzer, K.S., and Euskirchen, E.S. 2004. Carbon cycling and storage in world
forests: biome patterns related to forest age. Glob. Change Biol. 10(12): 2052–
2077. doi:10.1111/j.1365-2486.2004.00866.x.
Published by NRC Research Press
Environ. Rev. Downloaded from www.nrcresearchpress.com by Université du Québec à Montréal on 08/28/14
For personal use only.
Oris et al.
Price, D.T., McKenney, D.W., Caya, D., and Côté, H. 2001. Transient climate
change scenarios for high resolution assessment of impacts on Canada's
forest ecosystems. Natural Resources Canada, Great Lakes Forestry Centre.
Radke, L.F., Hegg, D.A., Hobbs, P.V., Nance, J.D., Lyons, J.H., Laursen, K.K., Weiss,
R.E., Riggan, P.J., and Ward, D.E. 1991. Particulate and trace gas emissions
from large biomass fires in North America. In Global Biomass Burning. Edited
by J. S. Levine. MIT Press, Cambridge, Massachusetts. pp. 209–216.
Ramanathan, V., and Carmichael, G. 2008. Global and regional climate changes
due to black carbon. Nat. Geosci. 1(4): 221–227. doi:10.1038/ngeo156.
Ramanathan, V., Crutzen, P.J., Kiehl, J.T., and Rosenfeld, D. 2001. Atmosphere:
aerosols, climate, and the hydrological cycle. Science, 294(5549): 2119–2124.
doi:10.1126/science.1064034. PMID:11739947.
Ramaswamy, V., Boucher, O., Haigh, J., Hauglustaine, D., Haywood, J., Myhre, G.,
Nakajima, T., Shi, G.Y., and Solomon, S. 2001. Radiative forcing of climate
change. In Climate Change 2001:The Scientific Basis. Edited by J.T. Houghton,
Y. Ding, D.J. Griggs, M. Noguer, P.J. van der Linden, X. Dai, K. Maskell, and C.A.
Johnson. Cambridge University Press, Cambridge, UK. pp. 349–416.
Randerson, J.T., Liu, H., Flanner, M.G., Chambers, S.D., Jin, Y., Hess, P.G.,
Pfister, G., Mack, M.C., Treseder, K.K., Welp, L.R., Chapin, F.S., III,
Harden, J.W., Goulden, M.L., Lyons, E., Neff, J.C., Schuur, E.A.G., and
Zender, C.S. 2006. The impact of boreal forest fire on climate warming.
Science, 314(5802): 1130–1132. doi:10.1126/science.1132075. PMID:17110574.
Reid, J.S., Eck, T.F., Christopher, S.A., Koppman, R., Dubovik, O., Eleuterio, D.P.,
Holben, B.N., Reid, E.A., and Zhang, J. 2005. A review of biomass burning
emissions part III: Intensive optical properties of biomass burning particles.
Atmos. Chem. Phys. 5(3): 827–849. doi:10.5194/acp-5-827-2005.
Rogers, B.M., Randerson, J.T., and Bonan, G.B. 2013. High-latitude cooling associated with landscape changes from North American boreal forest fires. Biogeosciences, 10(2): 699–718. doi:10.5194/bg-10-699-2013.
Sawamoto, T., Hatano, R., Yajima, T., Takahashi, K., and Isaev, A.P. 2000. Soil
respiration in Siberian taiga ecosystems with different histories of forest fire.
Soil Sci. Plant Nutr. 46(1): 31–42. doi:10.1080/00380768.2000.10408759.
Schulze, E.D., Lloyd, J., Kelliher, F.M., Wirth, C., Rebmann, C., Luhker, B., Mund, M.,
Knohl, A., Milyukova, I.M., Schulze, W., Ziegler, W., Varlagin, A.B., Sogachev, A.F.,
Valentini, R., Dore, S., Grigoriev, S., Kolle, O., Panfyorov, M.I., Tchebakova, N., and
Vygodskaya, N.N. 1999. Productivity of forests in the eurosiberian boreal region
and their potential to act as a carbon sink – a synthesis. Glob. Change Biol.
5(6): 703–722. doi:10.1046/j.1365-2486.1999.00266.x.
Schuur, E.A.G., Vogel, J.G., Crummer, K.G., Lee, H., Sickman, J.O., and
Osterkamp, T.E. 2009. The effect of permafrost thaw on old carbon release
and net carbon exchange from tundra. Nature, 459(7246): 556–559. doi:10.
1038/nature08031. PMID:19478781.
Seiler, W., and Crutzen, P.J. 1980. Estimates of gross and net fluxes of carbon
between the biosphere and the atmosphere from biomass burning. Clim.
Change, 2(3): 207–247. doi:10.1007/BF00137988.
Shenoy, A., Johnstone, J.F., Kasischke, E.S., and Kielland, K. 2011. Persistent
effects of fire severity on early successional forests in interior Alaska. For.
Ecol. Manage. 261(3): 381–390. doi:10.1016/j.foreco.2010.10.021.
Shvidenko, A., Schepaschenko, D., Sukhinin, A., Mccallum, I., and Maksyutov, S.
2011. Carbon emissions from forest fires in boreal Eurasia between 1998–2010. In
The 5th International Wildland Fire Conference, 9–13 May 2011, Sun City, South
Africa.
Simard, M., Lecomte, N., Bergeron, Y., Bernier, P.Y., and Paré, D. 2007. Forest
productivity decline caused by successional paludification of boreal soils.
Ecol. Appl. 17(6): 1619–1637. doi:10.1890/06-1795.1. PMID:17913128.
Simpson, I.J., Akagi, S.K., Barletta, B., Blake, N.J., Choi, Y., Diskin, G.S., Fried, A.,
Fuelberg, H.E., Meinardi, S., Rowland, F.S., Vay, S.A., Weinheimer, A.J.,
Wennberg, P.O., Wiebring, P., Wisthaler, A., Yang, M., Yokelson, R.J., and
Blake, D.R. 2011. Boreal forest fire emissions in fresh Canadian smoke plumes:
C1-C10 volatile organic compounds (VOCs), CO2, CO, NO2, NO, HCN and
CH3CN. Atmos. Chem. Phys. 11(13): 6445–6463. doi:10.5194/acp-11-6445-2011.
Skinner, W.R., Stocks, B.J., Martell, D.L., Bonsal, B., and Shabbar, A. 1999. The
association between circulation anomalies in the mid-troposphere and area
burned by wildland fire in Canada. Theor. Appl. Clim. 63(1–2): 89–105. doi:10.
1007/s007040050095.
Soja, A.J., Cofer, W.R., III, Shugart, H.H., Sukhinin, A.I., Stackhouse, P.W., Jr.,
McRae, D.J., and Conard, S.G. 2004. Estimating fire emissions and disparities
in boreal Siberia (1998–2002). J. Geophys. Res.: Atmos. 109(14). doi:10.1029/
2004JD004570.
Spracklen, D.V., Mickley, L.J., Logan, J.A., Hudman, R.C., Yevich, R.,
Flannigan, M.D., and Westerling, A.L. 2009. Impacts of climate change from
2000 to 2050 on wildfire activity and carbonaceous aerosol concentrations in
the western United States. J. Geophys. Res.: Atmos. 114(20). doi:10.1029/
2008JD010966.
Spracklen, D.V., Carslaw, K.S., Poschl, U., Rap, A., and Forster, P.M. 2011. Global
cloud condensation nuclei influenced by carbonaceous combustion aerosol.
Atmos. Chem. Phys. 11(17): 9067–9087. doi:10.5194/acp-11-9067-2011.
219
Stinson, G., Kurz, W.A., Smyth, C.E., Neilson, E.T., Dymond, C.C.,
Metsaranta, J.M., Boisvenue, C., Rampley, G.J., Li, Q., White, T.M., and
Blain, D. 2011. An inventory-based analysis of Canada's managed forest carbon dynamics, 1990 to 2008. Glob. Change Biol. 17(6): 2227–2244. doi:10.1111/
j.1365-2486.2010.02369.x.
Stocks, B.J., Fosberg, M.A., Lynham, T.J., Mearns, L., Wotton, B.M., Yang, Q.,
Jin, J.Z., Lawrence, K., Hartley, G.R., Mason, J.A., and McKenney, D.W. 1998.
Climate change and forest fire potential in Russian and Canadian boreal
forests. Clim. Change 38(1): 1–13. doi:10.1023/A:1005306001055.
Stocks, B.J., Mason, J.A., Todd, J.B., Bosch, E.M., Wotton, B.M., Amiro, B.D.,
Flannigan, M.D., Hirsch, K.G., Logan, K.A., Martell, D.L., and Skinner, W.R.
2002. Large forest fires in Canada, 1959–1997. J. Geophys. Res.: Atmos.
107(D1): 8149. doi:10.1029/2001JD000484.
Sudo, K., and Akimoto, H. 2007. Global source attribution of tropospheric ozone:
Long-range transport from various source regions. J. Geophys. Res.: Atmos.
112(D12): D12302. doi:10.1029/2006JD007992.
Sukhinin, A.I., French, N.H.F., Kasischke, E.S., Hewson, J.H., Soja, A.J.,
Csiszar, I.A., Hyer, E.J., Loboda, T., Conrad, S.G., Romasko, V.I.,
Pavlichenko, E.A., Miskiv, S.I., and Slinkina, O.A. 2004. AVHRR-based mapping of fires in Russia: New products for fire management and carbon cycle
studies. Remote Sens. Environ. 93(4): 546–564. doi:10.1016/j.rse.2004.08.011.
Terrier, A., Girardin, M.P., Périé, C., Legendre, P., and Bergeron, Y. 2013. Potential
changes in forest composition could reduce impacts of climate change on
boreal wildfires. Ecol. Appl. 23(1): 21–35. doi:10.1890/12-0425.1. PMID:
23495633.
Thornley, J.H.M., and Cannell, M.G.R. 2004. Long-term effects of fire frequency
on carbon storage and productivity of boreal forests: A modeling study. Tree
Physiol. 24(7): 765–773. doi:10.1093/treephys/24.7.765. PMID:15123448.
Turetsky, M.R., Amiro, B.D., Bosch, E., and Bhatti, J.S. 2004. Historical burn area
in western Canadian peatlands and its relationship to fire weather indices.
Global Biogeochem. Cycles, 18(4): 1–9. doi:10.1029/2004GB002222.
Turetsky, M.R., Kane, E.S., Harden, J.W., Ottmar, R.D., Manies, K.L., Hoy, E., and
Kasischke, E.S. 2011. Recent acceleration of biomass burning and carbon
losses in Alaskan forests and peatlands. Nat. Geosci. 4(1): 27–31. doi:10.1038/
ngeo1027.
van Bellen, S. 2011. Carbon accumulation in three ombrotrophic peatlands of the
Eastmain region, Quebec, Canada: influence of hydrology, vegetation and
fire in the holocene climatic context. Ph.D. yhesis, Université du Québec à
Montréal. p. 141.
van der Werf, G.R., Randerson, J.T., Giglio, L., Collatz, G.J., Kasibhatla, P.S., and
Arellano, A.F., Jr. 2006. Interannual variability in global biomass burning
emissions from 1997 to 2004. Atmos. Chem. Phys. 6(11): 3423–3441. doi:10.
5194/acp-6-3423-2006.
van der Werf, G.R., Randerson, J.T., Giglio, L., Collatz, G.J., Mu, M.,
Kasibhatla, P.S., Morton, D.C., Defries, R.S., Jin, Y., and Van Leeuwen, T.T.
2010. Global fire emissions and the contribution of deforestation, savanna,
forest, agricultural, and peat fires (1997–2009). Atmos. Chem. Phys. 10(23):
11707–11735. doi:10.5194/acp-10-11707-2010.
Wang, S., Zhou, L., Chen, J., Ju, W., Feng, X., and Wu, W. 2011. Relationships
between net primary productivity and stand age for several forest types and
their influence on China's carbon balance. J. Environ. Manage. 92(6): 1651–
1662. doi:10.1016/j.jenvman.2011.01.024. PMID:21339040.
Wei, Y., Li, M., Chen, H., Lewis, B.J., Yu, D., Zhou, L., Zhou, W., Fang, X., Zhao, W.,
and Dai, L. 2013. Variation in carbon storage and its distribution by stand age
and forest type in boreal and temperate forests in Northeastern China. PLoS
One, 8(8): 1–9. doi:10.1371/journal.pone.0072201. PMID:23977252.
Wirth, C., Schulze, E.D., Lühker, B., Grigoriev, S., Siry, M., Hardes, G., Ziegler, W.,
Backor, M., Bauer, G., and Vygodskaya, N.N. 2002. Fire and site type effects on
the long-term carbon and nitrogen balance in pristine Siberian Scots pine
forests. Plant Soil, 242(1): 41–63. doi:10.1023/A:1020813505203.
Wooster, M.J., and Zhang, Y.H. 2004. Boreal forest fires burn less intensely in
Russia than in North America. Geophys. Res. Lett. 31(20): L20505 20501–
20503. doi:10.1029/2004GL020805.
Yoshikawa, K., Bolton, W.R., Romanovsky, V.E., Fukuda, M., and Hinzman, L.D.
2003. Impacts of wildfire on the permafrost in the boreal forests of interior
Alaska. J. Geophys. Res.: Atmos. 108(1): FFR 4-1–FFR 4-14. doi:10.1029/
2001JD000438.
Zhuang, Q., McGuire, A.D., O’Neill, K.P., Harden, J.W., Romanovsky, V.E., and
Yarie, J. 2002. Modeling soil thermal and carbon dynamics of a fire chronosequence in interior Alaska. J. Geophys. Res.: Atmos. 108(1): FFR 3-1–FFR 3-26.
doi:10.1029/2001JD001244.
Zhuang, Q., Melillo, J.M., Sarofim, M.C., Kicklighter, D.W., McGuire, A.D.,
Felzer, B.S., Sokolov, A., Prinn, R.G., Steudler, P.A., and Hu, S. 2006. CO2 and
CH4 exchanges between land ecosystems and the atmosphere in northern
high latitudes over the 21st century. Geophys. Res. Lett. 33(L17403). doi:10.
1029/2006GL026972.
Published by NRC Research Press