Open Access
How to translate text using browser tools
1 February 2014 Effects of Climate Change on Peatlands in the Far North of Ontario, Canada: A Synthesis
Jim McLaughlin, Kara Webster
Author Affiliations +
Abstract

The Hudson Bay Lowlands (HBL) is the largest peatland complex in North America. More than 75% of the HBL occurs in Ontario, where the provincial government mandates that ecosystem carbon storage and sequestration be considered in land-use planning. Accomplishing this task requires identifying carbon indicators and assessing their responses to changing ecosystem processes, such as succession, permafrost thaw, and evapotranspiration (ET). Therefore, we synthesized information on peat carbon indicators and ecosystem process from the literature. Findings indicate that the long-term carbon accumulation, carbon dioxide (CO2) sequestration, peat depth, and peatland age were similar (p > 0.10) between dry and wet peatland features. Furthermore, CO2 sequestration displayed the highest variability and ponds were net CO2 emitters. Recent carbon accumulation, CH4 emission, and ET were highest (p < 0.01) in wet features, with CH4 emission displaying wide variation. Increased active layer thickness (105 ± 92 cm per 100 years) in permafrost was the most variable ecosystem process analyzed in this study, while variation in permafrost loss (53 ± 23% per 100 years) was similar to that of carbon accumulation and ET rates. Processes creating wet and pond conditions may increase landscape-scale CO2 and CH4 emissions to the atmosphere, weakening peatland carbon sinks. Dry conditions may reduce CH4 emissions but potentially increase peatland susceptibility to fire. Knowledge of these changes should be useful for climate change vulnerability and adaptation assessments for large landscapes. However, better understanding of variability in CO2 sequestration, CH4 emission, and permafrost dynamics is required to design such assessments for small landscapes.

Introduction

Northern peatlands account for approximately 25% of the global soil carbon (C) pool, with most of the peatlands occurring in Russia and Canada (Yu, 2012). Peatlands are categorized as ombrotrophic bogs (receiving water primarily from precipitation) and minerotrophic fens (receiving water from precipitation and groundwater inflow). Bogs and fens cycle C differently (Bauer et al., 2009), and their C sink strength may change as peat warms and becomes drier or wetter (Webster et al., 2013). However, vulnerability and adaptation of peatland C budgets in a changing climate are largely unknown (Frolking et al., 2011).

The Far North of Ontario (located between approximately 50 and 57°N, 79 and 94°W) is a highly variable landscape that consists of about 45 million hectares distributed relatively evenly between the Canadian Shield and Hudson Bay Lowlands (HBL). Here, peatlands are the dominant land class, accounting for nearly 50% of the area (Fig. 1). These peatlands store an estimated 36 Gt (1 Gt = 1 Pg = 1015 g) of C as peat (Far North Science Advisory Panel, 2010), which is approximately 25% of the C stored in all of Canada's peatlands. The HBL, which contains nearly 75% of the C stored in the Far North, is the largest peatland complex and has the southernmost extent of non-alpine permafrost (subsurface earth materials that remain below 0 °C for two or more consecutive years [Brown, 1967]; see Table 1 for permafrost features and their definitions) in North America (Martini, 2006).

Peatlands in the HBL have been important in Holocene climate cooling because since deglaciation they have sequestered large amounts of atmospheric carbon dioxide (CO2) (van Bellen et al., 2011; Bunbury et al., 2012). However, when disturbed, these peatlands may release much of this C back to the atmosphere as CO2 or methane (CH4), which may speed global warming (Webster and McLaughlin, 2011). For example, although warmer and drier peat may increase tree and shrub biomass (Weltzin et al., 2003), peat decomposition and CO2 loss may dominate over its gain through net primary production (NPP). In contrast, warmer and wetter peat enhances sedge and moss biomass, characteristic of fens that support rapid C accumulation in, and CH4 emission from, peat (Webster et al., 2013). Therefore, although CO2 is sequestered and accumulated as peat, CH4 losses may offset C gain and enhance atmospheric global warming potential (GWP; sum CO2 and CH4 fluxes expressed as CO2 equivalents].

Ontario House Bill 191, the Far North Act ( http://www.ontla.on.ca/bills/bills-files/39_Parliament/Session2/b191ra.pdf), which achieved Royal Assent in October 2010, describes four land-use planning objectives. One of those objectives is “the maintenance of biological diversity, ecological processes and ecological functions, including the storage and sequestration of carbon in the Far North.” Furthermore, Far North land-use planning teams may request the development of several policy statements, one of which concerns “ecological systems processes and functions, including considerations for cumulative effects and for climate change adaptation and mitigation.”

Achieving the act's planning objective and developing an appropriate policy statement for land-use planning requires an understanding of how climate change affects peatland C storage and sequestration at the landscape scale. Spatial configurations of permafrost and peatland types, succession and plant communities, along with subsequent evapotranspiration (ET) are important controls on landscape C storage and sequestration. For example, plant communities alter peatland C cycling through plant photosynthesis and the quantity and quality (i.e., decomposability) of litter added to peat (Laiho, 2006). Permafrost amount and thaw rates can also alter landscape-scale C budgets by creating drier or wetter conditions (Malmer et al., 2005). The latter inhibits organic matter decomposition relative to plant NPP, but enhances CH4 emissions, whereas the opposite occur when peat dries (Johansson et al., 2006b). Furthermore, ET is positively correlated with CO2, sequestration under wet conditions (Sonnentag et al., 2010). However, dry peat created by prolonged elevated ET relative to water input is also more susceptible to fires that burn deeper into peat (Turetsky et al., 2004) than occur in current HBL landscapes (Balshi et al., 2007), further contributing net peatland C losses to the atmosphere.

FIGURE 1.

Ontario land cover map. Solid white line approximates the Far North of Ontario boundary (adapted from OMNR, 2000, with permission).

f01_84.jpg

TABLE 1

Features common in the continuous (CPZ) and discontinuous (DPZ) permafrost zones in the Far North of Ontario.

t01_84.gif

In this paper, we synthesize the processes controlling peatland C accumulation and greenhouse gas (GHG) exchange in the HBL. The approach was to compile, analyze, and discuss the data on peatland C fluxes, succession, permafrost, and ET from the literature. We conclude by presenting a climate change vulnerability and adaptation assessment framework for HBL peatlands, discussing challenges to its development and its application in land-use planning.

Peatland Carbon Indicators

Successful communication among researchers, policymakers, and land managers requires common terminology. For this paper, peat C content is defined as mass of C contained in a unit area (e.g., kg C m-2) of dry peat. Carbon content is calculated as the product of C concentration, peat bulk density, and sample increment thickness. Peat C storage is defined as the product of C content, area of peatland in a landscape, and peat depth—the latter not well documented in subarctic peatlands. Carbon dioxide sequestration is defined as short-term (annual to decadal) balance of atmospheric CO2 uptake by plants and its release through plant respiration and organic matter decomposition. Carbon dioxide sequestration is commonly measured using gas flux chambers and eddy covariance flux towers. Carbon accumulation is defined as net C gain in peat at century to millennial scales. Peat C accumulation is measured by (1) dating of surface peat with specified tracers, such as 210Pb and 137Cs, that provide decadal to century (e.g., 50 to 200 years) time scales of peat accumulation (i.e., recent rate of C accumulation [RERCA]; Oldfield et al., 1995) and (2) radiocarbon (14C) dating of basal peat and peat profile age/depth relationships to estimate century to millennial C accumulation (i.e., long-term rates of C accumulation [LORCA]; Kuhry and Vitt, 1996). These rates are considered apparent because without intensively dating individual peat cores and identifying wetness using paleoecological indicators, timelines for and causes of LORCA and RERCA changes during peatland development cannot be inferred (Yu, 2012). However, differentiating sequestration from accumulation is critical in climate change vulnerability assessments because CO2 sequestration in plant biomass and surface litter does not equate to long-term C accumulation as peat.

TABLE 2

Number of peatland studies reporting data on carbon accumulation and sequestration and methane release that were used in this synthesis by peatland type/feature.

t02_84.gif

FIGURE 2.

Box plots of long-term rates of carbon accumulation (LORCA) and recent rates of carbon accumulation (RERCA), where the box height indicates the interquartile range, the solid white line within the box represents the median value, and the dashed lines from the top and bottom extend to the highest and lowest values, respectively. Different letters for LORCA and RERCA, respectively, indicate statistical differences at p < 0.05.

f02_84.jpg

Data sets for peatland C indicators analyzed in this study were compiled from 79 published papers; the number of studies and data points available for LORCA (g C m-2 yr-1), RERCA (g C m-2 yr-1), CO2 sequestration (g C m-2 d-1), and CH4 emission (mg C m-2 d-1) are shown in Table 2. A total of 680 values were retrieved from the publications, with approximately 70% corresponding to GHG exchange and the remainder to C accumulation rates. Mean values were calculated for each study reporting multiple rates by peatland type; this resulted in 129 data points to conduct the statistical analyses described below. Because of small sample sizes and infrequent differences among bog and fen hummocks, permafrost palsas, peat plateaus, and polygonal peatlands, they were classified as “dry features.” Similarly, collapse scars, lawns, carpets, hollows, and pools in bogs and fens were classified as “wet features.” Although only four ponds were represented in the final data set for CO2 sequestration, they were considered a separate category because all were net CO2 and CH4 sources to the atmosphere. A meta-analysis to assess the effects of experimental manipulations on C fluxes was beyond the scope of this paper, as the primary interest was an initial evaluation of potential C indicator variability. Meta-analysis is, however, recommended as the next step in peatland C indicator selection.

Differences in LORCA and RERCA between dry and wet features were assessed with the Mann-Whitney test, while a Kruskal-Wallis (Conover, 1980) test was used to assess differences in CO2 sequestration and CH4 emission from wet, dry, and pond features. Coefficient of variation (CV) and sample sizes were used to calculate the power (probability of failing to reject a false null hypothesis; Type II error) of the statistical tests. In addition, sample sizes required to detect a 20% difference in mean values of each indicator at the 0.05 level of significance (α) and a power of 0.70 were calculated following the methods of McLaughlin and Phillips (2006). Quartile analysis was used to define the bounds of natural variation for the indicators using the inter-quartile range between the first through third quartiles (25th through 75th percentiles), to which data compiled for HBL peatlands were compared.

Long-term C accumulation rate is summarized in Table 3; it was the most promising indicator (power = 0.78) analyzed in this study. No difference (p = 0.281) occurred between dry and wet features; therefore, data were combined into one peatland class. Natural variation of LORCA ranged from 18.0 to 24.5 g C m-2 yr-1 (Fig. 2), and its mean value in HBL peatlands (16.4 ± 4.9 g C m-2 yr-1 ; n = 11) was below the 25th percentile (18.0 g C m-2 yr-1) of the overall data set. Although the power of the Mann-Whitney test was 0.40 for RERCA, it was higher (p = 0.009) in wet than dry features (Fig. 2). Furthermore, mean RERCA in HBL peatlands (77 ± 42 g C m-2 yr-1; n = 9) was similar to the 50th percentile (78 g C m-2 yr-1) of the overall data set. The above results suggest RERCA may be a viable decade-to-century indicator of C accumulation. However, it would be beneficial to conduct additional RERCA measurements in boreal and subarctic peatlands, including the HBL, to increase statistical power.

TABLE 3

Recent and long-term carbon accumulation rates reported for northern peatlands.

t03_84.gif

Greenhouse gas fluxes are summarized in Table 4; CO2 sequestration (power = 0.17) and CH4 emission (power = 0.24) were characterized by high variation (Fig. 3, parts a and b) and low statistical power. Furthermore, CO2 sequestration did not differ (p = 0.766) between dry and wet features, and they were weak CO2 sinks (Fig. 3, part a). Ponds may be a CO2 source; however, statistical inferences are limited by the sample size (n = 4) and caution is recommended when interpreting the pond CO2 exchange data. For example, ponds covering mineral sediments are large CO2 sources to the atmosphere, whereas those over organic sediments alternate between being CO2 sinks and sources and tend to be smaller sources when assessed on an annual basis (Macrae et al., 2004). Differences in mineral and organic sediments should, therefore, be considered in scaling and modeling CO2 sequestration in peatland ponds.

Mean CO2 sequestration in dry features of the HBL (-0.62 ± 1.04 g CO2-C m-2 d-1) was similar to the 50th percentile (-0.64 g CO2-C m-2 d-1) of the overall data set (Fig. 3, part a). However, 7 of the 11 data points from the HBL analyzed here came from the 1990 Northern Wetland Studies in the southern HBL (Glooschenko et al., 1994) and may not be applicable to the entire HBL or its current landscape structure. Nearly all the peatlands (97%) analyzed in this study were net emitters of CH4 to the atmosphere, with wet features and ponds releasing approximately 25 times more CH4than dry features (p < 0.0001; Fig. 3, part b). Variation in CH4 emission was least in ponds (CV = 0.55), followed by wet (CV = 0.77) and dry (CV = 0.91) features. In the HBL, mean CH4 emission from wet (41 ± 41 mg CH4-C m-2 d-1, n = 10) and dry (5.4 ± 4.4 mg CH4-C m2 d-1, n = 8) features, as well as ponds (87 ± 41 mg CH4-C m-2 d-1 n = 8), fell within the bounds of natural variation established in this study (Fig. 3, part b).

FIGURE 3.

Box plots of greenhouse gas fluxes from boreal and subarctic peatlands to the atmosphere calculated from the literature: (a) CO2, where the box height indicates the interquartile range, the solid white line within the box represents the median value, and the dashed lines from the top and bottom extend to the highest and lowest values, respectively, and (b) CH4, where the box height indicates the interquartile range, the solid white line within the box represents the median value, and the dashed lines from the top and bottom extend to the highest and lowest values, respectively. Different letters for CO2 and CH4, respectively, indicate statistical differences at p < 0.05.

f03_84.jpg

Ecological Processes and Carbon Associations

PEATLAND SUCCESSION

Peatland development in the HBL was primarily controlled by recession of the Tyrell Sea and subsequent isostatic rebounding of the land (Glaser et al., 2004). These peatlands emerged more than 7000 years ago, although 14C basal peat dates produced variable ages (Table 5). The results reported in Table 5 were summarized from 54 peat depth and 42 14C basal peat dates, respectively, collected in the Hudson Bay region. These data produced mean peat depth and age of 236 ± 92 cm and 4826 ± 1356 14C years, respectively, and neither depth (p = 0.256) nor age (p = 0.523) differed among peatland types; statistical power was near 0.70 for both depth and age. Natural variation for the combined data was 168 to 266 cm for depth and 4003 to 5810 for 14C age. Variation in both depth (CV = 0.39) and age (CV = 0.28) was comparable to variation of LORCA.

In the HBL, peat and C accumulation rates were generally high, between 6500 and 4200 cal yr BP (calibrated years before present), then decreased as rich and intermediate fens transformed to poor fens, with Larix laricina (larch) and ericaceous plants present (Arlen-Pouliot and Bhiry, 2005; Bunbury et al., 2012). Since 2500 cal yr BP, poor fens generally transformed to bogs, and in the northern HBL bogs may have contained permafrost. In general, bogs may have accumulated the least amount of peat during development (Arlen-Pouliot and Bhiry, 2005). Further south, low peat and C accumulation rates were also reported during bog and poor fen successional stages (Bunbury et al., 2012). Here, however, the fastest accumulation rates occurred between 1000 and 600 cal yr BP and have since declined. At the southernmost portion of the HBL (Eastmain region of Québec), high C accumulation rates occurred in bogs during dry conditions because of high bulk density and C concentration in buried wood (Loisel and Garneau, 2010). Also, rapid accumulation rates during wet conditions were dependent on the presence of hummock-mosses because of their slow decomposition rates (van Bellen et al., 2011).

Peat temperature and wetness are important controls on peatland succession, and plant communities can evolve rapidly in response to modified peatland and permafrost conditions. For example, sampling in a permafrost peatland complex in northern Sweden over 30 to 50 years showed more trees, ericaceous shrubs, and hummock-mosses in drier areas after permafrost thaw (Malmer et al., 2005). However, sedges expanded at the landscape scale due to spreading of wet areas at the cost of hummocks as permafrost thawed, contributing to increased CO2 sequestration and ch4 emissions (Johansson et al., 2006b). Resampling of another Swedish peatland complex indicated that tree cover increased from less than 5% of plots initially sampled to approximately 30% of plots sampled 42 years later (Gunnarsson et al., 2002). Percentage of plots containing mosses also increased, whereas those containing sedges decreased. In a large-scale water table and temperature manipulation experiment in northern Minnesota (Weltzin et al., 2003), ericaceous shrub cover increased 50% four years after peat in a bog mesocosm was warmed from 1.6 to 4.1 °C; again, sedge cover decreased. Furthermore, above- (30%) and belowground (100%) biomass increased when water tables were low (Weltzin et al., 2003). Higher water table levels in a fen mesocosm resulted in increased sedge cover, whereas lower water table levels produced shrub results similar to those in the bog.

Remote sensing studies also indicated increased tree and shrub vegetation under warmer and drier conditions, or sedges, hollow-mosses, and tall shrubs under wetter conditions within the past 50 years (Cornelissen et al., 2001; Tape et al., 2006). Therefore, establishing plant function types (PFTs) should facilitate scaling and modeling C budget responses to landscape-level drying or wetting. Combining remote sensing of peatland PFTs and peat accumulation rates and GHG exchange measured at research sites and monitoring networks allows the use of calibrated and validated models to estimate C balances (St-Hilaire et al., 2010). The PFTs used depend on scale. For large landscapes, vascular (e.g., treed bogs and fens) and non-vascular (e.g., open bogs and fens) plant groups have been used (Dorrepaal, 2007). Results presented in this paper suggest dry, wet, and pond classifications may also be useful for scaling and modeling current and future C accumulation and GHG changes of HBL landscapes. Peatlands may also be more narrowly defined using vegetation classes, such as (1) mosses, sedges, shrubs, and trees or (2) genera- and species-level PFTs for smaller areas (Dorrepaal, 2007). For example, sedges, tall shrubs, and Sphagnum mosses have been used successfully to scale CH4 emissions at site (Johansson et al., 2006b) and regional (Dorrepaal, 2007) scales.

TABLE 4

Carbon dioxide (CO2) sequestration and methane (CH4) emissions for northern peatlands.

t04_84.gif

TABLE 5

Basal peat ages in peatlands in permafrost zones of Canada, Peat depth Age

t05_84.gif

PERMAFROST PATTERNS

Permafrost distribution based on mean annual air temperature (MAAT) is generally well mapped across Canada (Fig. 4, part a). However, its association with mean annual ground temperature (MAGT) is complex, and only sparse data exist that describe MAAT and MAGT associations (Zhang et al., 2008). Furthermore, aggrading and degrading permafrost features (Table 1) are present in various stages of maturation within a climatic zone (Payette et al., 2004). These patterns result from local variations in vegetation, topography, snow cover, and soil conditions (i.e., peat versus mineral soil) that largely control MAGT and subsequent permafrost presence (Johansson et al., 2006a). Variations in environmental conditions are important when MAAT is close to 0 °C, as in the HBL (Fig. 4, part b). This temperature range may be associated with the fastest peat C accumulation rates in northern peatlands; rates decrease northward because of low NPP and southward because of faster peat decomposition (Beilman et al., 2009).

In the northern hemisphere, during the 21st century permafrost is expected to disappear from 16 to 90% of the land base, and active layer (surface peat layer experiencing seasonal thaw that is the location of rapid biogeochemical cycling and water fluxes) thickness (ALT) may increase by 30 to 300 cm (Table 6). Mean (± standard deviation) percent permafrost loss calculated from data in Table 6 was 53% ± 25% per 100 years, with a CV of 0.47 and a statistical power of 0.65; increasing the sample size from the 37 compiled in this study to 43 would increase the power to 0.70. Furthermore, natural variation in permafrost loss ranged from 33 to 73% per 100 years. Mean increase in ALT was 105 ± 92 cm per 100 years, with a CV of 0.88, a power of 0.24, and natural variation between 50 and 100 cm; sample size would need to be 143 to attain a power of 0.70. However, models project that the HBL has already surpassed its temperature threshold for maintaining permafrost (Koven et al., 2011), or will do so by 2020 (Zhang et al., 2008). Permafrost thawing is, therefore, an important concern because MAAT in the HBL is projected to rise between 2.5 and 8.0 °C, with the most rapid warming occurring in the most permafrostdense area in the region (Gagnon and Gough, 2005; McKenney et al., 2010). Yet, variable ALT may impede documenting current and future implications to landscape C balances.

Spatial configurations of permafrost features depend on a variety of internal (e.g., peat accumulation and acidification) and external (e.g., climate and fire behavior) factors (Johansson et al., 2006a). In the northernmost regions, palsa/peat plateaus and thermokarst features dominate landscape patterns, along with nonpermafrost bogs and fens (Payette et al., 2004; Kuhry, 2008). In contrast, mostly non-permafrost bog and fen complexes occur in southern regions of the HBL, with forested bogs that are susceptible to burning (Turetsky et al., 2004), covering much of the area (Klinger and Short, 1996). Spatial distribution of permafrost extent implies that during the 21st century southern and northern portions of the HBL may respond differently to climate change. For example, the boreal zone is projected to warm and dry, increasing the likelihood of faster peat decomposition and more frequent and severe burning of forested palsas, peat plateaus, and bogs. Subsequently, elevated CO2 losses may weaken the C sink strength of southern HBL peatlands (e.g., Turetsky et al., 2004). In contrast, permafrost thawing in the northern portion of the HBL may maintain current peat moisture levels, or increase the amount of saturated soils on the landscape (Zhang et al., 2008), with warmer temperatures enhancing peatland CO2 sequestration and CH4 emission (Johansson et al., 2006b). However, thermokarst pond expansion may partially offset peat CO2 sequestration and further enhance CH4 emission to the atmosphere (Macrae et al., 2004; Laurion et al., 2010).

At coarse landscape scales, permafrost thaw has been modeled for the Hudson Bay region using both equilibrium and disequilibrium assumptions between MAAT and MAGT, although small sample sizes preclude statistical analyses. Gagnon and Gough (2005) assumed MAAT and MAGT equilibrium would result in 35 to 67% losses in permafrost by 2100 in the Hudson Bay region. In contrast, Zhang et al. (2008) assumed non-equilibrium between MAAT and MAGT and reported only minor losses (about 16%) in permafrost occurrence in the HBL during the 21st century, primarily in the southern fringe of the discontinuous permafrost zone (DPZ). Gagnon and Gough's (2005) results were generally within the range of natural variability defined in this study, although some climate models produced results greater than that for the 75th percentile. In contrast, results from Zhang et al. (2008) were below the 25th percentile. That study simulated permafrost loss using the processbased Northern Ecosystem Soil Temperature (NEST; Zhang et al., 2003) model, which projects less permafrost loss by accounting for plant structure, snow depth, and soil organic layer thickness effects on ground thermal dynamics.

Aerial photographic and remote sensing studies provide evidence that permafrost thawing is changing HBL landscapes. For example, at Boniface River, Québec, permafrost decreased from approximately 80% of the land area in 1957 to 20% in 2003 (Payette et al., 2004). This decrease was accompanied by increases in (1) fen area from less than 5% in 1957 to approximately 50% in 2003 and (2) thermokarst pond area from 20 to 40% during the same time frame. Similar changes have occurred at the Stordalen mire complex in northern Sweden, also weakening landscape C and GWP sink strengths (Malmer et al., 2005; Johansson et al., 2006b). Those results were consistent with this current synthesis, where wet features and ponds emitted approximately 25 times more CH4 than dry features. Combined with potentially more CO2 release from the landscape, pond (and wet feature) expansion will likely play a vital role in C balances in northern HBL landscapes.

Carbon and GWP sink strengths over decadal to century time frames following thermokarst formation are relatively unknown. Some evidence supports less CH4 emission from Sphagnum-dominated collapse scars as they age (Prater et al., 2007), but landscape ch4 emissions may remain similar as ponds transform to sedgedominated fens because of the latter's rapid CH4 emission (Webster et al., 2013). Moreover, Frolking et al. (2011) suggested that by 2100, 30 to 100 Mt CO2-C yr-1 may be sequestered and up to 35 Mt CH4-C yr-1 released from permafrost regions (beyond current values), which was consistent with coupled carbon-climate-change biospheric terrestrial model projections (Koven et al., 2011; Schaefer et al., 2011).

PEATLAND EVAPOTRANSPIRATION

Evapotranspiration is reported as actual (ETa) and potential (ETp), and northern peatlands frequently evaporate water at rates lower than their potential (Rouse, 2000); the rate of ET is strongly influenced by the surface energy balance (i.e., latent heat flux; Qe). Evapotranspiration affects peatland C exchange at landscape scales primarily through its contribution to water balances. For example, when water was abundant, ET and gross primary production in bogs were positively correlated (Sonnentag et al., 2010), but under water-stressed conditions, photosynthetic depressions were linked to elevated ET in a variety of peatland plants (Euskirchen et al., 2012). Furthermore, elevated ET relative to water inputs may contribute to long-term peatland drying, leading to higher frequency and severity of fires, potentially contributing further to weakening peatland C sink strengths (Turetsky et al., 2004). Because of its relationship with C exchange, accurate estimates of ET are required to construct water balances and assess their potential responses to climate change. However, the amount of water withdrawn through ET varies between 20 and 80% of the water balance, depending on terrain unit (e.g., site type), permafrost conditions, and temperature and precipitation regimes (Rouse, 1998).

Since large areas of the northern HBL occur as palsa-bogfen-pool/pond complexes and small lakes (Martini, 2006), ET estimates for dominant features are needed before a scaled value can be derived for a specific peatland complex (e.g., Petrone et al., 2000). Although ET rates vary (Table 7), tundra ponds, small lakes, and wet sites including fens, marshes, and tundra wetlands displayed the highest ET rates. Dry features, including spruce forests, non-forested bogs, lichen/heath, heath/shrub, tussock tundra, and peat plateaus/polygonal tundra had the lowest ET rates. Analysis of data compiled from 28 studies (Tables 7 and 8) indicated ET was higher (p = 0.002) in wet and pond than in dry features (power = 0.61). Evapotranspiration rates in wet features in the HBL (2.8 to 3.5 mm d-1) were within the range of natural variability (2.5 to 3.5 mm d-1) calculated from the overall data set. Wet features (0.71 ± 0.15) also partitioned more energy (p = 0.003) to ET than dry features (0.49 ± 0.08), as defined by the Qe-to-net radiation (Q*) ratio (Table 8). The statistical power was 0.58, but the smallp value suggests Qe/Q* should be further pursued as an indicator of wet and dry feature energy partitioning and water budgets in developing methods to assess climate change vulnerability and adaptation.

Looking at estimates of how expected changes in climate might influence annual ET patterns in a sedge fen at Churchill, Manitoba, Rouse (1998) reported that a two times CO2 climate warming scenario with a MAAT increase of 4 °C and no precipitation change increased summer water deficit, triggered mainly by greater ET during May. The water deficit could be counterbalanced by a 23% increase in summer rainfall (Rouse, 1998). Similar results were recorded in the James Bay Lowland (Lafleur et al., 1993). Roulet et al. (1992) assumed a 3.5 °C increase in temperature and 1 mm day-1 increase in precipitation for several northern fens and reported an increase in the soil moisture deficit in excess of 100%. Other studies across Canadian peatlands have since shown that thresholds of about a 25% increase in water supplied by precipitation or permafrost thawing would be required to maintain soil moisture levels under warminginduced enhancement of ET (Quinton et al., 2011).

FIGURE 4.

(a) Permafrost distribution across Canada illustrating permafrost zones, which are defined by the proportion of area underlain by permafrost. (b) Mean annual near-surface temperature distribution across Canada. Image (a) adapted from Department of Natural Resources Canada, < http://geogratis.gc.ca/api/en/nrcan-rncan/ess-sst/092b663d-198b-5c8d-9665-fa3f5970al4f.html>. Image (b) from < http://www.ccin.ca/home/ccw/permafrost/future>.

f04_84.jpg

TABLE 6

Predicted change in permafrost area and active layer depth during the 21st century in Canada.

t06_84.gif

Effects of climate change on ET are more difficult to predict for dry than for wet features. This has been attributed to the phenology of vascular plants, predominantly leaf area index (LAI), and its control on energy partitioning in sparsely vegetated bogs (LAI < 1.0) (Admiral et al., 2006). Leaf area index is minimal in spring, when peat is generally saturated (Lafleur and Humphreys, 2008). This corresponds to the time of highest solar energy reaching bog surfaces and high Qe/Q* values due to Sphagnum mosses, which account for 50 to 70% of the total annual ET from bogs (Kellner, 2001). Increasing LAI decreases the amount of radiation reaching the surface (Admiral et al., 2006). At this time, ET switches from moss- to vascular plant—dominated, and subsequently dries the peat (Kellner, 2001).

Understanding moss and vascular plant interactions with projected earlier snowmelt is critical to comprehend future ET trends of dry features because of the earlier onset of ET by mosses. However, few studies have documented the competition between mosses and vascular plants following snowmelt. Therefore, it is uncertain if earlier snowmelt will only shift moss ET to earlier in the year or if soil warming during spring will enhance nutrient mineralization and subsequent vascular plant biomass and LAI, prolonging soil moisture deficits. Most model results project that during the 21st century vascular plant biomass and LAI will increase. Such increases have been positively correlated with aboveground biomass and soil C accretion (Frolking et al., 2011). However, LAI increases are also positively correlated with fire frequency and area burned because of dry surface peat (Turetsky et al., 2004). As such, we view (1) moss and vascular plant competition following earlier snowmelt and (2) extended growing season, energy partitioning, and peatland drying on C accumulation rates and GHG exchange rates as key research areas, where data are vital to understand and improve future bog C budget projections.

Permafrost is a strong heat sink and generally has a larger ground heat flux (Qg) than non-permafrost peatlands, and changes in temperature and moisture may alter permafrost thaw rates (Quinton et al., 2011). However, few data exist to confidently describe and model the energy balance in permafrost features. Boike et al. (2008) reported that during a wet year in polygonal tundra of the continuous permafrost zone (CPZ), energy partitioning remained typical of wet features (Table 8). However, when soil was dry, a relatively even partitioning of energy occurred (Table 8). In addition, during dry years, energy partitioning was not affected by differences in air temperature. However, during wet years Qe increased with increasing temperatures (Boike et al., 2008). Similar patterns were shown for wet and dry tussock tundra underlain by permafrost (Harazono et al., 1998) and mesic and heath/shrub tundra (Lafleur and Humphreys, 2008).

Climate Change Vulnerability and Adaptation Assessment Framework to Assess Peatland Carbon Indicators

The large contribution of the HBL to provincial and national C budgets implies that understanding peatland C accumulation GWP gains and losses is vital to sustaining C storage and sequestration in the ever-changing HBL landscape. Although scant, evidence is convincing that PFTs, permafrost, and ET are changing in the HBL. However, rates and direction of peatland changes are variable and depend on whether peat becomes drier or wetter as it warms (Fig. 5). Some changes, such as thermokarst formation (Arlen-Pouliot and Bhiry, 2005) and transformations within fen types as they convert to bogs (McLaughlin and Webster, 2010), occur rapidly (i.e., decades); those two processes have opposite effects on peatland C balances (Fig. 5). However, successional trajectories in the HBL are poorly documented (although recent summaries of data collected from the 1970s and early 1980s are available [Riley, 2011]). Also, the remoteness of the region hinders large-scale sampling endeavors due to high costs.

TABLE 7

Evapotranspiration (ET) rates for common site types in high boreal and subarctic drainage basins north of 50° latitude.

t07_84.gif

TABLE 8

Energy balance partitioning, Bowen ratio [latent heat flux (Qe)/net radiation (Q*); sensible (Qh)/Q*; ground heat flux (Qg)/Q*; lake heat storage Qw/Q*], Bowen ratio energy balance (β), and Priestly-Taylor alpha (α) for different site types in peatlands and permafrost features north of 50° latitude.

t08_84.gif

Remote sensing using various platforms and sensors and modeling techniques have been employed to assess PFTs, permafrost thawing, ET, and C indicators at the landscape scale (Dorrepaal, 2007). Furthermore, the ecosystem processes presented here have consistently been correlated with various components of C budgets and successfully scaled across landscapes (e.g., Johansson et al., 2006b; Beilman et al., 2009). However, few researchers have attempted to develop correlations between climate change and peatland C indicators for the HBL. Significant efforts are required to sample peat and calibrate satellite and aerial photographs for PFTs, permafrost coverage, and ET to calibrate permafrost and peatland C models.

In terms of C indicators assessed in this study, low variation and high statistical power in LORCA comparisons point to it being a promising C indicator. Coupled with relatively small variability in peat depth and 14C basal peat age, long-term peatland processes may be effective metrics in climate change assessments. However, LORCA in the HBL peatlands was on the low end of the natural variation calculated across boreal and subarctic peatlands. Furthermore, LORCA values were from a limited number of locations and peatland types in the HBL; thus, additional measurements in the HBL are warranted to further evaluate LORCA variability. Although statistical power was relatively low for RERCA and CH4 emissions, rates were much higher in wet (and CH4 for ponds) than dry features, suggesting their potential usefulness in evaluating landscape-scale peatland C balances in response to decade-tocentury permafrost and PFT changes. However, much of the data from the HBL analyzed in this synthesis pre-dates the mid-1990s and may not be applicable across the region or to its current landscape structure. In addition, high variation and low statistical power in CH4 emission (and CO2 sequestration) hinders construction of viable GHG balances in HBL peatlands. Assigning to landscape GHG changes may be possible but requires further analyses.

Current and future peatland C accumulation and GHG emissions in northern peatlands may also be underestimated due to (1) poorly defined climate change rates and subsequent peatland adaptation and (2) whether new climate states will be within the range of Holocene variation (Frolking et al., 2011). As indicated by Frolking et al. (2011), the only approach available to address climate change and C budgets is rudimentary process-based modeling, the results of which are inconsistent with one another and field data. For example, warming may increase peatland C losses by 40 to 85% in coming centuries (Ise et al., 2008), or peatland C stocks may stabilize in response to drying (Frolking et al., 2010). However, none of the available permafrost (e.g., NEST) and C (e.g., McGill Wetland Model [St-Hilaire et al., 2010], Wetland-DNDC [Zhang et al., 2002]) models allow plant communities or spatial patterns in permafrost and peatland features to vary during simulations, significantly limiting their applicability to functional landscapes.

Information about the quality (decomposability) of peat exposed to aerated or flooded conditions with warming temperatures is also lacking. Deeply buried peat is primarily “recalcitrant” material due to prior decomposition in aerobic surface peat layers (Laiho, 2006). However, the effects of warmer and drier (or wetter) conditions on decomposition rates of “recalcitrant” peat are conflicting, with some studies reporting faster decomposition (Dorrepaal, 2007) and others showing no rate changes (Davidson and Janssens, 2006) under warmer temperatures. Furthermore, peat enclosed in permafrost tends to be less decomposable than that of enclosed mineral soil organic matter (Turetsky et al., 2007). However, data on short- and long-term decomposition of C released through ice melting in permafrost regions are scarce. Finally, dissolved organic C (DOC) may be an important regulator of microbial decomposition in peatlands (Laiho, 2006). The quality of DOC to support decomposition is known to differ among peatland types and between dry and wet conditions (Webster and McLaughlin, 2010). How those changes will be manifested in future organic matter decomposition in peatlands, however, is virtually unknown.

Other uncertainties and challenges also impede the use of climate and C indicators for assessing climate change vulnerability and adaptation. For climate change, lack of weather and permafrost monitoring networks in the HBL significantly limits understanding of past and current temperature, precipitation, and permafrost regimes, contributing to highly variable projections in future conditions. For example, mean monthly precipitation projections differed by more than 200% (Gagnon and Gough, 2005; McKenney et al., 2010). Precipitation amounts are important to landscape permafrost coverage within the same temperature zone (Rouse, 2000; Johansson et al., 2006a). Interactions between air temperature and precipitation with peat temperature and moisture dynamics are also some of the most poorly calibrated coefficients used to project permafrost and peatland C dynamics (Zhang et al., 2008; St-Hilaire et al., 2010).

Few valid estimates of permafrost amount and its thaw rate exist for the HBL. Payette et al. (2004) reported that between 1957 and 2003 percent annual permafrost loss increased from 2.5 to 5.3%, which is much higher than the natural range of variation (e.g., 0.33% yr-1 to 0.73% yr-1 assuming a 100-year time frame over the 21st century) calculated from coarse-scale climate and process-based models. Current models, such as NEST, may constrain variability by accounting for plant structure, snow depth, and soil organic layer thickness effects on energy and water transfer in soil-vegetation-atmosphere systems, and subsequent ground thermal dynamics. However, all models fail to address changes in landscape pattern, which is a critical modeling gap. For example, assuming a 50% increase in fen and 20% increase in pond areas in the northern HBL (Payette et al., 2004) during the 21st century, landscape-level CO2 and CH4 emissions to the atmosphere may increase significantly and would not be detectable in coarsescale projections. Such changes have been shown to significantly weaken C and GWP sinks (Malmer et al., 2005; Johansson et al., 2006b).

FIGURE 5.

Conceptual model of permafrost and non-permafrost peatland succession (arrows) in response to changing peat conditions (gray boxes) in the Hudson Bay Lowlands of Ontario. C = carbon, CO2 = carbon dioxide, CH4 = methane, GWP = global warming potential.

f05_84.jpg

In conclusion, this synthesis indicates that changes in C accumulation rates and CH4 emissions in response to altered ecosystem processes, such as succession, permafrost loss, and ET may be useful to assess peatland climate change vulnerability and adaptation for large HBL landscapes. Furthermore, C accumulation and CH4 emission may be evaluated through remote sensing identification of changes in wet, dry, and pond landscape features. However, an improved understanding of the interactions between GHG fluxes and permafrost dynamics (loss and ALT) is needed before such assessments are possible for small landscapes. Therefore, enhanced research and monitoring studies are needed to estimate current and project future peatland C storage and sequestration with reasonable certainty to support land-use planning in the HBL.

Acknowledgments

Support was provided by the Ontario Ministry of Natural Resources’ Applied Research and Development Branch and Far North Branch under the auspices of projects CC-008 and CC-021 and FNIKM 028, respectively, to McLaughlin and Natural Resources Canada-Canadian Forest Service A-base to Webster. The authors also thank Lisa Buse, Sandra Wawryszyn, Trudy Vaittinen, Maara Packalen, Ngaire Eskelin, Benoit Hamel, John Ralston, Adam Kinnunen, Mark Crofts, Sheri-Ann Kuiper, and Ravi Kanipayor for their assistance with various aspects of the project.

References Cited

1.

A. C. Adkinson , K. H. Syed , and L. B. Flanagan , 2011: Contrasting responses of growing season ecosystem CO2 exchange to variation in temperature and water table depth in two peatlands in northern Alberta, Canada. Journal of Geophysical Research , 116: G01004, doi:  http://dx.doi.org/10.1029/2010JG001515Google Scholar

2.

S. W. Admiral , P. M. Lafleur , and N. T. Roulet , 2006: Controls on latent heat flux and energy partitioning at a peat bog in eastern Canada. Agricultural and Forest Meteorology, 140: 308–321. Google Scholar

3.

A. A. Ali , B. Ghaleb , M. Garneau , H. Asnong , and J. Loisel , 2008: Recent peat accumulation rates in minerotrophic peatlands of the James Bay region, eastern Canada, inferred by Pb-210 and Cs-137 radiometric techniques. Applied Radiation and Isotopes, 66: 1350–1358. Google Scholar

4.

J. Alm , K. Schulman , L. Walden , H. Nykanen , P. J. Martikainen , and J. Silvola , 1999: Carbon balance of a boreal bog during a year with an exceptionally dry summer. Ecology, 80: 161–174. Google Scholar

5.

Y. Arlen-Pouliot , and N. Bhiry , 2005: Paleoecology of a paisa and a filled thermokarst pond in a permafrost peatland, subarctic Québec, Canada. The Holocene, 15: 408–419. Google Scholar

6.

A. Arneth , J. Kurbatova , O. Kolle , O. Shibistova , J. Lloyd , N. Vygodskaya , and E. D. Schulze , 2002: Comparative ecosystem atmosphere exchange of energy and mass in a European Russian and a central Siberian bog. II. Interseasonal and interannual variability of CO2 fluxes. Tellus Series B, 54: 514–530. Google Scholar

7.

M. Aurela , T. Riutta , T. Laurila , J.-P. Tuovinen, T. Vesala , E.-S. Tuittila , J. Rinne , S. Haapanala , and T. Laine , 2007: CO2 exchange of a sedge fen in southern Finland—The impact of a drought period. Tellus Series B, 59: 826–837. Google Scholar

8.

K. Bäckstrand , P. M. Crill , M. Jackowicz-Korczynski , M. Mastepanov , T. R. Christensen , and D. Bastviken , 2010: Annual carbon gas budget for a subarctic peatland, northern Sweden. Biogeosciences, 7: 95–108. Google Scholar

9.

M. S. Balshi , A. D. McGuire , Q. Zhuang , J. Melillo , D. W. Kicklighter , E. Kasischke , C. Wirth , M. Flannigan , J. Harden , J. S. Clein , T. J. Burnside , J. McAllister , W. A. Kurz , M. Apps , and A. Shvidenko , 2007: The role of historical fire disturbance in the carbon dynamics of the pan-boreal region: a process-based analysis. Journal of Geophysical Research, 112: G02029, doi:  http://dx.doi.org/10.1029/2006JG000380Google Scholar

10.

I. E. Bauer , J. S. Bhatti , C. Swanston , R. K. Wieder , and C. M. Preston , 2009: Organic matter accumulation and community change at the peatland-upland interface: inferences from 14C and 210Pb dated profiles. Ecosystems, 12: 636–653. Google Scholar

11.

D. W. Beilman , G. M. MacDonald , L. C. Smith , and P. J. Reimer , 2009: Carbon accumulation in peatlands of West Siberia over the last 2000 years. Global Biogeochemical Cycles, 23: GB1012, doi:  http://dx.doi.org/10.1029/2007GB003112Google Scholar

12.

R. Bello , and I. D. Smith , 1990: The effect of weather variability on the energy balance of a lake in the Hudson Bay Lowlands, Canada. Arctic and Alpine Research, 22: 98–107. Google Scholar

13.

N. Bhiry , and E. C. Robert , 2006: Reconstruction of changes in vegetation and trophic conditions of a palsa in a permafrost peatland, subarctic Québec, Canada. Écoscience, 13: 56–65. Google Scholar

14.

N. Bhiry , S. Payette , and E. C. Robert , 2007: Peatland development at the arctic tree line (Québec, Canada) influenced by flooding and permafrost. Quaternary Research, 67: 426–437. Google Scholar

15.

J. Boike , C. Wille , and A. Abnizova , 2008: Climatology and summer energy and water balance of polygonal tundra in the Lena River Delta, Siberia. Journal of Geophysical Research, 113: G03025, doi:  http://dx.doi.org/10.1029/JG000540Google Scholar

16.

W. Borren , W. Bleuten , and E. D. Lapshina , 2004: Holocene peat and carbon accumulation rates in the southern taiga of western Siberia. Quaternary Research, 61: 42–51. Google Scholar

17.

L. D. Boudreau , and W. R. Rouse , 1995: The role of individual terrain units in the water balance of wetland tundra. Climatic Research, 5: 31–47. Google Scholar

18.

R. J. E. Brown , 1967: Permafrost in Canada. Ottawa, Ontario: Natural Resources Council of Canada, Geological Survey of Canada Map 1246a and Division of Building Research Map NRC-9769. Google Scholar

19.

J. L. Bubier , 1995: The relationship of vegetation to methane emission and hydrochemical gradients in northern peatlands. Journal of Ecology, 83: 403–420. Google Scholar

20.

J. L. Bubier , A. Costello , T. R. Moore , N. T. Roulet , and K. Cavage , 1993: Microtopography and methane flux in boreal peatlands, northern Ontario, Canada. Canadian Journal of Botany, 71: 1056–1063. Google Scholar

21.

J. L. Bubier , T. R. Moore , L. Bellisario , R. T. Comer , and P. M. Crill , 1995: Ecological controls on methane emissions from a northern peatland complex the zone of discontinuous permafrost, Manitoba, Canada. Global Biogeochemical Cycles, 9: 455–470. Google Scholar

22.

J. L. Bubier , S. Frolking , P. M. Crill , and E. Linder , 1999: Net ecosystem exchange and its uncertainty in a diverse boreal peatland. Journal of Geophysical Research, 104: 27683–27692. Google Scholar

23.

J. Bunbury , S. A. Finkelstein , and J. Bollman , 2012: Holocene hydroclimate and effects on carbon accumulation inferred for a peat bog in the Attawapiskat River watershed, Hudson Bay Lowlands, Canada. Quaternary Research, 78: 275–284. Google Scholar

24.

K. L. Burton , A. R. Rouse , and L. D. Boudreau , 1996: Factors affecting the summer carbon dioxide budget of subarctic wetland tundra. Climate Research, 6: 203–213. Google Scholar

25.

P. Camill , J. A. Lynch , J. S. Clark , J. B. Adams , and B. Jordan , 2001: Changes in biomass, aboveground net primary production, and peat accumulation following permafrost thaw in the boreal peatlands of Manitoba, Canada. Ecosystems , 4: 461–478. Google Scholar

26.

L. Chasmer , A. Kenward , W. Quinton , R. Petrone , 2012: CO2 exchanges within zones of rapid conversion from permafrost plateau to bog and fen land cover types. Arctic, Antarctic, and Alpine Research, 44: 399–411. Google Scholar

27.

T. R. Christensen , S. Josasson , T. V. Callahan , and M. Havström , 1995: Spatial variation in high-latitude methane flux along a transect across Siberian and European tundra environments.Journal of Geophysical Research, 100: 21035–21045. Google Scholar

28.

T. R. Christensen , M. Jackowicz-Korczynski , M. Aurela , P. Crill , M. Heliasz , M. Mastepanov , and T. Friborg , 2012: Monitoring the multi-year carbon balance of a subarctic palsa mire with micrometeorological techniques. Ambio , 41(Suppl. 3): 207–217. Google Scholar

29.

W. J. Conover , 1980. Practical Nonparametric Statistics. New York: John Wiley & Sons. Google Scholar

30.

J. H. C. Cornelissen , T. V. Callaghan , J. M. Alatalo , A. Michelsen , E. Graglia , A. E. Hartley , D. S. Hik , S. E. Hobbie , M. C. Press , H. Robinson , G. H. R. Henry , G. R. Shaver , G. K. Phoenix , D. G. Jones , S. Jonasson , F. S. Chapin III, U. Molau , C. Neill , J. A. Lee , J. M. Melillo , B. Sveinbjörnsson , and R. Aerts , 2001: Global change and arctic ecosystems: Is lichen decline a fonction of increases in vascular plant biomass? Journal of Ecology , 89: 984–994. Google Scholar

31.

C. Corradi , O. Kolle , K. Walter , S. A. Zimov , and E. D. Schultze , 2005: Carbon dioxide and methane exchange of a north-east Siberian tussock tundra. Global Change Biology, 11: 1910–1925. Google Scholar

32.

E. A. Davidson , and I. A. Janssens , 2006: Temperature sensitivity of soil carbon decomposition and feedbacks to climate change. Nature, 440: 165–173. Google Scholar

33.

G. den Hartog , H. H. Neumann , K. M. King , and A. C. Chipanshi , 1994: Energy budget measurements using eddy correlation and Bowen ratio techniques at the Kinosheo Lake tower site during the Northern Wetlands Study. Journal of Geophysical Research, 99: 1539–1549. Google Scholar

34.

A. R. Desyatkin , F. Takakai , P. P. Fedorov , M. C. Nikolaeva , R. V. Desyatkin , and R. Hatano , 2009: CH4 emission from different stages of thermokarst formation in central Yakutia, East Siberia. Soil Science and Plant Nutrition, 55: 558–570. Google Scholar

35.

R. E. Dickinson , K. W. Oleson , G. Bonan , F. Hoffman , P. Thornton , M. Vertenstein , Z.-L. Yang , and X. Zeng , 2006: The Community Land Model and its climate statistics as a component of the Community Climate System Model. Journal of Climate, 19: 2302–2324. Google Scholar

36.

E. Dorrepaal , 2007: Are plant growth-form-based classifications useful in predicting northern ecosystem carbon cycling feedbacks to climate change? Journal of Ecology , 95: 1167–1180. Google Scholar

37.

A. K. Eaton , W. R. Rouse , P. M. Lafleur , P. Marsh , and P. D. Blanken , 2001: Surface energy balance of the western and central Canadian subarctic: variations in the energy balance among five major terrain types. Journal of Climate, 14: 3692–3703. Google Scholar

38.

W. Eugster , W. R. Rouse , R. A. Pielke , J. P. McFadden , D. D. Baldocchi , T. G. F. Kittel , F. S. Chapin , G. E. Liston , P. L. Vidale , E. Vaganov , and S. Chambers , 2000: Land-atmosphere energy exchange in Arctic tundra and boreal forest: available data and feedbacks to climate. Global Change Biology, 6(Suppl. 1): 84–115. Google Scholar

39.

E. S. Euskirchen , A. D. McGuire , D. W. Kicklighter , Q. Zhuang , J. S. Clein , R. J. Dargaville , D. G. Dye , J. S. Kimball , K. C. McDonald , J. M. Meillo , V. E. Romanovsky , and N. V. Smith , 2006: Importance of recent shifts in soil thermal dynamics on growing season length, productivity, and carbon sequestration in terrestrial high latitude ecosystems. Global Change Biology, 12: 731–750. Google Scholar

40.

E. S. Euskirchen , M. S. Bret-Hart , G. J. Scott , C. Edgar , and G. R. Shaver , 2012: Seasonal patterns of carbon dioxide and water fluxes in three representative tundra ecosystems in northern Alaska. Ecosphere , 3: article 4, doi:  http://dx.doi.org/10.1890/ES11-00202.1Google Scholar

41.

Far North Science Advisory Panel, 2010: Science for a Changing Far North. The Report of the Far North Science Advisory' Panel. A report submitted to the Ontario Ministry of Natural Resources, < http://www.ontario.ca\farnorth>. Google Scholar

42.

H. Flessa , A. Rodionoc , G. Guggenberger , H. Fuchs , P. Magdon , O. Shibistova , G. Zrazhevskaya , N. Mikheyeva , O. A. Kasansky , and C. Blodau , 2008: Landscape controls of CH4 fluxes in a catchment of the forest tundra ecotone in northern Siberia. Global Change Biology, 14: 2040–2056. Google Scholar

43.

T. Friborg , H. Søegaard , T. R. Christensen , C. R. Lloyd , and N. S. Panikov , 2003: Siberian wetlands: where a sink is a source. Geophysical Research Letters , 30: 2129, doi:  http://dx.doi.org/10.1029/2003GL017797Google Scholar

44.

S. Frolking , N. T. Roulet , E. Tuittila , J. L. Bubier , and A. Quillet , 2010: A new model of Holocene peatland net primary production, decomposition, water balance and peat accumulation. Earth Systems Dynamics, 1: 115–167. Google Scholar

45.

S. Frolking , J. Talbot , N. Roulet , M. Jones , C. Treat , J. B. Kauffman , and E. Tuittila , 2011: Peatlands in the Earth's 21st century coupled climate-carbon system. Environmental Reviews, 19: 371–396. Google Scholar

46.

A. S. Gagnon , and W. A. Gough , 2005: Climate change scenarios for the Hudson Bay region: an intermodel comparison. Climatic Change, 69: 269–297. Google Scholar

47.

P. H. Glaser , B. S. C. Hansen , D. I. Siegel , A. S. Reeve , and P. J. Morin , 2004: Rates, pathways and drivers for peatland development in the Hudson Bay Lowlands, northern Ontario, Canada. Journal of Ecology, 92: 1036–1053. Google Scholar

48.

A. J. Glenn , L. B. Flanagan , K. H. Syed , and P. J. Carlson , 2006: Comparison of net ecosystem CO2 exchange in two peatlands in western Canada with contrasting dominant vegetation, Sphagnum and Carex. Agricultural and Forest Meteorology, 140: 115–135. Google Scholar

49.

W. A. Glooschenko , N. T. Roulet , L. A. Barrie , H. I. Schiff , and H. G. McAdie , 1994: The Northern Wetlands Study (NOWES): an overview. Journal of Geophysical Research, 99: 1423–1428. Google Scholar

50.

T. J. Griffis , W. R. Rouse and J. M. Waddington , 2000: Interannual variability of net ecosystem CO2 exchange at a subarctic fen. Global Biogeochemical Cycles, 14: 1109–1121. Google Scholar

51.

U. Gunnarsson , N. Malmer , and H. Rydin , 2002: Dynamics or constancy in Sphagnum dominated mire ecosystems? A 40-year study. Ecography, 25: 685–704. Google Scholar

52.

J. D. Hamilton , C. A. Kelly , J. W. M. Rudd , R. H. Hesslein , and N. T. Roulet , 1994: Flux to the atmosphere of CH4 and CO2, from wetland ponds on the Hudson Bay Lowlands (HBL). Journal of Geophysical Research, 99: 1495–1510. Google Scholar

53.

Y. Harazono , M. Yoshimoto , M. Mano , G. L. Vourlitis , and W. C. Oechel , 1998: Characteristics of energy and water budgets over wet sedge and tussock tundra ecosystems at North Slope in Alaska. Hydrological Processes, 12: 2163–2183. Google Scholar

54.

J. E. P. Heikkinen , M. Maijanen , M. Aurela , K. J. Hargreaves , and P. J. Martikainen , 2002: Carbon dioxide and methane dynamics in a sub-Arctic peatland in northern Finland. Polar Research, 21: 49–62. Google Scholar

55.

E. R. Humphreys , P. M. Lafleur , L. B. Flanagan , N. Hedstrom , K. H. Syed , A. J. Glenn , and R. Granger , 2006: Summer carbon dioxide and water vapor fluxes across a range of northern peatlands. Journal of Geophysical Research-Biogeosciences, 111: G04011, doi:  http://dx.doi.org/10.1029/2005JG000111Google Scholar

56.

T. Ise , A. L. Dunn , S. C. Wofsy , and P. R. Moorcroft , 2008: High sensitivity of peat decomposition to climate change through watertable feedback. Nature Geoscience, 1: 763–766. Google Scholar

57.

M. Jackowicz-Korczyski , T. R. Christensen , K. Bäckstrand , P. Crill , T. Friborg , M. Mastepanov , and L. Ström , 2010: Annual cycle of methane emission from a subarctic peatland. Journal of Geophysical Research , 115: G03009, doi:  http://dx.doi.org/10.1029/2008JG000913.  Google Scholar

58.

T. Johansson , T. R. Christensen , H. J. Åkerman , and T. V. Callaghan , 2006a: What determines the current presence or absence of permafrost in the Torneträsk region, a sub-arctic landscape in northern Sweden. Ambio, 35: 190–197. Google Scholar

59.

T. Johansson , N. Malmer , P. M. Crill , T. Friborg , J. H. Åkerman , M. Mastepanov , and T. R. Christensen, 2006b: Decadal vegetation changes in a northern peatland, greenhouse gas fluxes and net radiative forcing. Global Change Biology, 12: 2352–2369. Google Scholar

60.

D. W. Joiner , P. M. Lafleur , J. H. McCaughey , and P. A. Bartlett , 1999: Interannual variability in carbon dioxide exchanges at a boreal wetland in the BOREAS northern study area. Journal of Geophysical Research, 104: 663–672. Google Scholar

61.

M. C. Jones , G. Grosse , B. M. Jones , and K. W. Anthony , 2012: Peat accumulation in drained thermokarst lake basins in continuous, ice-rich permafrost, northern Seward Peninsula, Alaska. Journal of Geophysical Research, 117: G00M07, doi:  http://dx.doi.org/10.1029/2011JG001766Google Scholar

62.

E. Kellner , 2001: Surface energy fluxes and control of evapotranspiration from a Swedish Sphagnum mire. Agricultural and Forest Meteorology, 110: Google Scholar

63.

L. M. Kettles , M. Garneau, and H. Jette , 2000: Macrofossil, pollen, and geochemical records of peatlands in the Kinosheo Lake and Detour Lake areas, northern Ontario. Ottawa, Canada: Geological Survey of Canada, Bulletin 545, 24 pp. Google Scholar

64.

E. S. Klein , R. K. Booth , Z. Yu , B. G. Mark , and N. D. Stansell , 2013: Hydrology-mediated differential response of carbon accumulation to late Holocene climate change at two peatlands in southcentral Alaska. Quaternary Science Reviews, 64: 61–75. Google Scholar

65.

L. E. Klinger , and S. K. Short , 1996: Succession in the Hudson Bay Lowland, northern Ontario, Canada. Arctic and Alpine Research, 28: 172–193. Google Scholar

66.

L. F. Klinger , P. R. Zimmerman , J. P. Greenberg , L. E. Heidt , and A. B. Guenther , 1994: Carbon trace gas fluxes along a successional gradient in the Hudson Bay Lowland. Journal of Geophysical Research, 99: 1469–1494. Google Scholar

67.

U. Kokfeit , N. Reuss , E. Strug , M. Sonesson , M. Rundgren , G. Skog , P. Rosén , and D. Hammarlund , 2010: Wetland development, permafrost history and nutrient cycling inferred from late Holocene peat and lake sediment records in subarctic Sweden. Journal of Paleolimnology , 44: 327–342. Google Scholar

68.

C. D. Koven , B. Ringeval , P. Friedlingstein , P. Ciais , P. Cadule , D. Khvorostyanov , G. Krinner , and C. Tarnocai , 2011: Permafrost carbon-climate feedbacks accelerate global warming. Proceedings of the National Academy of Sciences, 108: 14769–14774. Google Scholar

69.

H. Kubiw , M. Hickman , and D. H. Vitt , 1989: The development history of peatlands at Muskiki and Marguerite Lakes, Alberta. Canadian Journal of Botany, 67: 3534–3544. Google Scholar

70.

P. Kuhry , 1994: The role of fire in the development of Sphagnumdominated peatlands in western boreal Canada. Journal of Ecology, 82: 899–910. Google Scholar

71.

P. Kuhry , 2008: Palsa and peat plateau development in the Hudson Bay Lowlands, Canada: timing, pathways and causes. Boreas, 37: 316–327. Google Scholar

72.

P. Kuhry , and D. H. Vitt , 1996: Fossil carbon/nitrogen ratios as a measure of peat decomposition. Ecology, 77: 271–275. Google Scholar

73.

P. Kuhry , A. Halsey , S. E. Bayley , and D. H. Vitt , 1992: Peatland development in relation to Holocene climatic change in Manitoba and Saskatchewan (Canada). Canadian Journal of Earth Sciences, 29: 1070–1090. Google Scholar

74.

J. Kurbatova , A. Arneth , N. N. Vygodskaya , O. Kolle , I. M. Milyukova , N. M. Tchebakova , E.-D. Schulze , and J. Lloyd , 2002: Comparative ecosystem-atmosphere exchange of energy and mass in a European Russian and central Siberian bog I. Interseasonal and interannual variability of energy and latent heat fluxes during the snowfree period. Tellus Series B, 54: 497–513. Google Scholar

75.

P. M. Lafleur , 1992: Energy balance and evapotranspiration from a subarctic forest. Agricultural and Forest Meteorology, 58: 163–175. Google Scholar

76.

P. M. Lafleur , and E. R. Humphreys , 2008: Spring warming and carbon dioxide exchange over Low Arctic tundra in central Canada. Global Change Biology, 14: 1–17. Google Scholar

77.

P. M. Lafleur , A. V. Renzetti , and B. Bello , 1993: Seasonal changes in the radiation balance of subarctic forest and tundra. Arctic and Alpine Research, 25: 33–36. Google Scholar

78.

P. M. Lafleur , N. T. Roulet , J. L. Bubier , S. Frolking , and T. R. Moore , 2003: Interannual variability in the peatland-atmosphere carbon dioxide exchange at an ombrotrophic bog. Global Biogeochemical Cycles, 17: 1036, doi:  http://dx.doi.org/10.1029/2002GB001983Google Scholar

79.

R. Laiho , 2006: Decomposition in peatlands: reconciling seemingly contrasting results on the impacts of lowered water levels. Soil Biology and Biochemistry, 38: 2011–2024. Google Scholar

80.

A. Lamarre , M. Garneau , H. Asnong , 2012: Holocene paleohydrological reconstruction and carbon accumulation of a permafrost peatland using testate amoeba and macrofossil analyses, Kuujjuarapik, subarctic Québec, Canada. Review of Paleobotany and Palynology, 186: 131–141. Google Scholar

81.

I. Laurion , W. F. Vincent , S. MacIntyre , L. Retamal , C. Dupont , P. Francuc , and R. Pienitz , 2010: Variability in greenhouse gas emissions from permafrost ponds. Limnology and Oceanography, 55: 115–133. Google Scholar

82.

D. M. Lawrence , and A. G. Slater , 2005: A projection of severe near-surface permafrost degradation during the 21st century. Geophysical Research Letters, 32: L24401, doi:  http://dx.doi.org/10.1029/2005GL025080Google Scholar

83.

D. M. Lawrence , and A. G. Slater , 2010: The contribution of snow conditions trends to future ground climate. Climate Dynamics, 34: 969–981. Google Scholar

84.

D. M. Lawrence , A. G. Slater , R. A. Thomas , M. M. Holland , and C. Deser , 2008: Accelerated arctic land warming and permafrost degradation during rapid sea ice loss. Geophysical Research Letters 35: L11506, doi:  http://dx.doi.org/10.1029/2008GL033985Google Scholar

85.

M. Leppälä , A. M. Laine , M.-J. Seväkivi , and E.-S. Tuittila , 2011: Differences in CO2 dynamics between successional mire plant communities during wet and dry summers. Journal of Vegetation Science , 22: 357–366. Google Scholar

86.

L. K. Libilik , T. R. Moore , J. L. Bubier , and S. D. Robinson , 1997: Methane emissions from wetlands in the zone of discontinuous permafrost: Fort Simpson, Northwest Territories, Canada. Global Biogeochemical Cycles, 11: 485–494. Google Scholar

87.

J. Loisel , and M. Garneau , 2010: Late Holocene paleoecohydrology and carbon accumulation estimates from two boreal peat bogs in eastern Canada: potential and limits of multi-proxy archives. Palaeogeography, Palaeoclimatology, Palaeoecology, 291: 493–533. Google Scholar

88.

L. Maanavilja , T. Riutta , M. Aurela , M. Pulkkinen , T. Laurila , and E.-S. Tuittila , 2011 : Spatial variation in CO2 exchange at a northern aapa mire. Biogeochemistry, 104: 325–345. Google Scholar

89.

M. L. Macrae , R. L. Bello , and L. A. Molot , 2004: Long-term carbon storage and hydrological control of CO2 exchange in tundra ponds in the Hudson Bay Lowland. Hydrological Processes, 18: 2051–2069. Google Scholar

90.

M. Mäkilä , M. Saarnisto , and T. Kankainen , 2001: Aapa mires as a carbon sink and source during the Holocene. Journal of Ecology , 89: 589–599. Google Scholar

91.

N. Malmer , T. Johansson , M. Olsrud , and T. R. Christensen , 2005: Vegetation, climatic changes and net carbon sequestration in a northScandinavian subarctic mire over 30 years. Global Change Biology, 11: 1895–1909. Google Scholar

92.

S. Marchenko , V. Romanovsky , and G. Tipenko , 2008: Numerical modeling of spatial permafrost dynamics in Alaska. In M. Philips , S. M. Springman , and L. U. Arenson (eds.), Proceedings 8th International Conference on Permafrost. Boca Raton, Florida: Taylor and Francis, 1125–1130. Google Scholar

93.

I. P. Martini , 2006: The cold climate peatlands of the Hudson Bay Lowland, Canada: brief overview of recent work. In I. P. Martini , A. M. Cortizas , and W. Chesworth (eds.), Peatlands: Evolution and Records of Environmental and Climate Changes. Amsterdam: Elsevier, 53–84. Google Scholar

94.

N. A. McEnroe , N. T. Roulet , T. R. Moore , and M. Garneau , 2009: Do pool surface area and depth control CO2 and CH4 fluxes from an ombrotrophic raised bog, James Bay, Canada? Journal of Geophysical Research , 114: G01001, doi:  http://dx.doi.org/10.1029/2007JG000639Google Scholar

95.

D. W. McKenney , J. H. Pedlar , K. Lawrence , P. A. Gray , S. J. Colombo , and W. J. Crins , 2010: Current and projected future climatic conditions for ecoregions and selected Natural Heritage Areas in Ontario. Sault Ste. Marie, Ontario, Canada: Ontario Ministry of Natural Resources, Climate Change Research Report CCRR-16,42 pp. Google Scholar

96.

J. W. McLaughlin , and S. A. Phillips , 2006: Soil carbon, nitrogen, and base cation cycling 17 years after whole-tree harvesting in a low-elevation red spruce (Picea rubens)—balsam fir (Abies balsamea) forested watershed in central Maine. Forest Ecology and Management, 222: 234–253. Google Scholar

97.

J. W. McLaughlin , and K. L. Webster , 2010: Alkalinity and acidity cycling and fluxes in an intermediate fen peatland in northern Ontario. Biogeochemistry, 99: 143–155. Google Scholar

98.

I. H. Meyers-Smith , A. D. McGuire , J. W. Harden , and R. C. Chapin III, 2007: Influence of disturbance on carbon exchange in a permafrost collapse and adjacent burned forest. Journal of Geophysical Research, 112: G04017, doi:  http://dx.doi.org/10.1029/2007JG000423Google Scholar

99.

T. Moore , N. T. Roulet , and R. Knowles , 1990: Spatial and temporal variation of methane flux from subarctic fens. Global Biogeochemical Cycles , 4: 29–46. Google Scholar

100.

K. E. Moore , D. R. Fitzjarrald , S. C. Wofsy , B. C. Daube , J. W. Munger , and P. S. Bakwin , 1994a: A season of heat, water vapour, total hydrocarbon and ozone fluxes at a subarctic fen. Journal of Geophysical Research , 99: 1937–1952. Google Scholar

101.

T. R. Moore , A. Heyes and N. T. Roulet , 1994b: Methane emissions from wetlands, southern Hudson Bay Lowland. Journal of Geophysical Research , 99: 1455–1467. Google Scholar

102.

H. H. Neumann , G. den Hartog , K. M. King , and A. C. Chipanshi , 1994: Carbon dioxide fluxes over a raised open bog at the Kinosheo Lake tower site during the Northern Wetlands Study (NOWES). Journal of Geophysical Research , 99: 1529–1538. Google Scholar

103.

B. J. Nicholson , and D. H. Vitt , 1990: The paleoecology of a peatland complex in western Canada. Canadian Journal of Botany, 68: 121–138. Google Scholar

104.

H. Nykanen , J. E. P. Heikkinen , L. Pirinen , K. Tiilikainen , and P. J. Martikainen , 2003: Annual CO2 exchange and CH4 fluxes on a subarctic palsa mire during climatically different years. Global Biogeochemical Cycles, 17: 1018, doi:  http://dx.doi.org/10.1029/2002GB001861Google Scholar

105.

F. Oldfield , N. Richardson , and P. G. Abbleby , 1995: Radiometric dating (210Pb, 137Cs, 241Am) of recent ombrotrophic peat accumulation and evidence for changes in mass balance. The Holocene, 5: 141–148. Google Scholar

106.

OMNR [Ontario Ministry of Natural Resources], 2000:28 class Ontario provincial land cover raster (computer file). Peterborough, Canada: Ontario Ministry of Natural Resources. Google Scholar

107.

S. Payette , A. Delwaide , M. Caccianiga , and M. Beauchemin , 2004: Accelerated thawing of subarctic peatland permafrost over the last 50 years. Geophysical Research Letters, 31: L18208, doi:  http://dx.doi.org/10.1029/2004GL020358Google Scholar

108.

L. Pelletier , T. R. Moore , N. T. Roulet , M. Garneau , and V. Beaulieu-Audy , 2007: Methane fluxes from three peatlands in the La Grande Rivière watershed, James Bay Lowland, Canada. Journal of Geophysical Research, 112: G01018, doi:  http://dx.doi.org/10.1029/2006JG000216Google Scholar

109.

I. F. Pendea , and G. L. Chmura , 2012: A high-resolution record of carbon accumulation rates during boreal peatland initiation. Biogeosciences, 9: 2711–2717. Google Scholar

110.

R. M. Petrone , W. R. Rouse , and P. Marsh , 2000: Comparative surface energy budgets in western Canada and central subarctic regions of Canada. International Journal of Climatology, 20: 1131–1148. Google Scholar

111.

J. L. Prater , J. P. Chanton , and G. J. Whiting , 2007 : Variation in methane production pathways associated with permafrost decomposition in collapse scar bogs of Alberta, Canada. Global Biogeochemical Cycles, 21: GB4004, doi:  http://dx.doi.org/10.1029/2006GB002866Google Scholar

112.

J. S. Price , 1991: Evaporation from a blanket bog in a foggy coastal environment. Boundary-Layer Meteorology, 57: 391–406. Google Scholar

113.

J. S. Price , and K. M. Woo , 1988: Studies of subarctic coastal marsh: I. Hydrology. Journal of Hydrology, 103: 275–292. Google Scholar

114.

J. S. Price , D. A. Maloney , and F. G. Downey , 1991: Peatlands of the Lake Melville Coastal Plain, Labrador. In T. D. Prowse , and C. S. L. Ommanney (eds.), Northern Hydrology: Selected Perspectives Proceedings of the Northern Hydrology Symposium No. 6. Ottawa, Canada: Environment Canada, Canadian Government Publication Centre , Supply and Services Canada, 293–302. Google Scholar

115.

W. L. Quinton , M. Hayashi , and L. E. Chasmer , 2011: Permafrostthaw-induced land-cover change in the Canadian subarctic: implications for water resources. Hydrological Processes, 25: 152–158. Google Scholar

116.

J. W. Raich , E. B. Rasetter , J. M. Mellilo , D. W. Kicklighter , P. A. Struedler , B. J. Peterson , A. L. Grace , B. Moore III, and C. J. Vorosmarty, 1991: Potential net primary production in South America: an application of a global model. Ecological Applications, 1: 399–429. Google Scholar

117.

H. Rask , J. Schoenau , and D. Anderson , 2002: Factors influencing methane flux from a boreal forest wetland in Saskatchewan, Canada. Soil Biology and Biochemistry, 34: 434–443. Google Scholar

118.

J. L. Riley , 2011: Wetlands of the Ontario Hudson Bay Lowland: a regional overview. Toronto, Canada: Nature Conservancy of Canada, 156 pp., < http://www.natureconservancy.ca>. Google Scholar

119.

S. D. Robinson , 2006: Carbon accumulation in peatlands, southwestern Northwest Territories, Canada. Canadian Journal of Soil Science, 86: 305–319. Google Scholar

120.

S. D. Robinson , and T. R. Moore , 1999: Carbon and peat accumulation over the past 1200 years in a landscape with discontinuous permafrost, northwestern Canada. Global Biogeochemical Cycles, 13: 591–601. Google Scholar

121.

S. D. Robinson , and T. R. Moore , 2000: The influence of permafrost and fire upon carbon accumulation in high boreal peatlands, Northwest Territories, Canada. Arctic, Antarctic, and Alpine Research, 32: 155–166. Google Scholar

122.

C. L. Roehm , and N. T. Roulet , 2003: Seasonal contribution of CO2 fluxes in the annual C budget of a northern bog. Global Biogeochemical Cycles, 17: 129, doi:  http://dx.doi.org/10.1029/2002GB001889Google Scholar

123.

N. T. Roulet , and M.-k. Woo , 1986: Wetland and lake evaporation in the low Arctic. Arctic and Alpine Research, 18: 195–200. Google Scholar

124.

N. Roulet , T. Moore , J. Bubier , and P. Lafleur , 1992: Northern fens: Methane flux and climate change. Tellus Series B, 44: 100–105. Google Scholar

125.

N. T. Roulet , A. Jano , C. A. Kelly , L. F. Klinger , T. R. Moore , R. Protz , J. A. Ritter , and W. R. Rouse , 1994: Role of the Hudson Bay Lowland as a source of atmospheric methane. Journal of Geophysical Research, 99: 1439–1454. Google Scholar

126.

W. R. Rouse , 1998: A water balance model for a subarctic sedge fen and its application to climate change. Climatic Change, 38: 107–234. Google Scholar

127.

W. R. Rouse , 2000: The energy and water balance of high-latitude wetlands: controls and extrapolation. Global Change Biology, 6: 59–68. Google Scholar

128.

W. R. Rouse , S. G. Hardill , and P. Lafleur , 1987: The energy balance in the coastal environment of James Bay and Hudson Bay during the growing season. Journal of Climatology , 1: 165–179. Google Scholar

129.

W. R. Rouse , A. K. Eaton , R. M. Petrone , L. D. Boudreau , P. Marsh , and T. J. Griffis, 2003: Seasonality in the surface energy balance of tundra in the lower Mackenzie River Basin. Journal of Hydrometeorology, 4: 673–679. Google Scholar

130.

J. Sagerfors , A. Lindroth , L. Klemedtsson , P. Weslien , and M. Nilsson , 2008: Annual CO2 exchange between a nutrient poor, minerotrophic, boreal mire and the atmosphere. Journal of Geophysical Research, 113: G01001, doi:  http://dx.doi.org/10.1029/2006JG000306Google Scholar

131.

F. St-Hilaire , J. Wu , N. T. Roulet , S. Frolking , P. M. Lafleur , E. R. Humphreys , and V. Arora , 2010: McGill wetland model: evaluation of a peatland carbon simulator developed for global assessments. Biogeosciences, 7: 3517–3530. Google Scholar

132.

K. Saito , M. Kimoto , T. M. Zhang , K. Takata , and S. Emori , 2007: Evaluating a high-resolution climate model: simulated hydrothermal regimes in frozen ground regions and their change under the global warming scenario. Journal of Geophysical Research, 112: F02S11, doi:  http://dx.doi.org/10.1029/2006JF000577Google Scholar

133.

A. B. K Sannel ,. and P. Kuhry , 2008: Long-term stability of permafrost in subarctic peat plateaus, west-central Canada. The Holocene, 18: 589–601. Google Scholar

134.

T. S. Sazonova , E. V. Romanovsky , J. E. Walsh , and D. O. Sergueev , 2004: Permafrost dynamics in the 20th and 21st centuries along the East Siberian transect. Journal of Geophysical Research, 109: D01108, doi:  http://dx.doi.org/10.1029/2003JD003680Google Scholar

135.

K. Schaefer , G. J. Collatz , P. Trans , A. S. Denning , I. Baker, J. Berry , L. Prihodko , N. Suits , and A. Philpott , 2008: Combined Simple Biosphere/Carnegie-Ames Stanford Approach terrestrial carbon cycle model. Journal of Geophysical Research-Biogeosciences, 113: G03034, doi:  http://dx.doi.org/10.1029/2007JG000603Google Scholar

136.

K. Schaefer , T. Zhang , L. Bruhwiler , and A. P. Barrett , 2011: Amount and timing of permafrost carbon release in response to climate warming. Tellus Series B, 63: 165–180. Google Scholar

137.

C. P. Schreader , W. R. Rouse , T. J. Griffis , L. D. Boudreau , and P. D. Blanken , 1998: Carbon dioxide fluxes in a northern fen during a hot, dry summer. Global Biogeochemical Cycles , 12: 729–740. Google Scholar

138.

O. Sonnentag , G. V. D. Kamp , A. G. Barr , and M. Chen , 2010: On the relationship between water table depth and water vapor and carbon dioxide fluxes in a minerotrophic fen. Global Change Biology, 16: 1762–1776. Google Scholar

139.

S. Spence , W. R. Rouse , D. Worth , and C. Oswald , 2003: Energy budget processes of a small northern lake. Journal of Hydrometeorology, 4: 694–701. Google Scholar

140.

R. B. Stewart , and W. R. Rouse , 1976: A simple method for determining the evaporation from shallow lakes and ponds. Water Resources Research, 12: 623–628. Google Scholar

141.

M. Strack , J. M. Waddington , L. Rochefort , and E.-S. Tuittila , 2006: Response of vegetation and net ecosystem carbon dioxide exchange at different peatland microforms following water table drawdown. Journal of Geophysical Research, 111: G02006, doi:  http://dx.doi.org/10.1029/2005JG000145Google Scholar

142.

S. L. Strilesky , and E. R. Humphreys , 2012. A comparison of the net ecosystem exchange of carbon dioxide and evapotranspiration for treed and open portions of a temperate peatland. Agricultural and Forest Meteorology, 153:45–53. Google Scholar

143.

A. E. Suyker , S. B. Verman , and T. J. Arkebauer , 1997: Season-long measurement of carbon dioxide exchange in a boreal fen. Journal of Geophysical Research, 102: 29021–29028. Google Scholar

144.

K. H. Syed , L. B. Flanagan , P. J. Carlson , A. J. Glenn , and K. E. Van Gaalen , 2006: Environmental control of net ecosystem CO2 exchange in a treed, moderately rich fen in northern Alberta. Agricultural and Forest Meteorology, 140: 97–114. Google Scholar

145.

K. Takata , S. Emori , and T. Watanabe , 2003: Development of the minimal advanced treatments of surface interaction and runoff. Global Planetary Change, 38: 209–222. Google Scholar

146.

K. Tape , M. Strum , and C. Racine , 2006: The evidence for shrub expansion in northern Alaska and the pan-arctic. Global Change Biology, 12: 686–702. Google Scholar

147.

K. Tolonen , and J. Turunen , 1996: Accumulation rates of carbon in mires in Finland and implications for climate change. The Holocene, 6: 171–178. Google Scholar

148.

M. R. Turetsky , R. K. Wieder , C. J. Williams , and D. H. Vitt , 2000: Organic matter accumulation, peat chemistry, and permafrost thawing in peatlands of boreal Alberta. Écoscience, 7: 379–392. Google Scholar

149.

M. R. Turetsky , R. K. Wieder , D. H. Vitt , R. Evans , and K. D. Scott , 2002: Boreal peatland C fluxes under varying permafrost regimes. Soil Biology and Biochemistry, 34: 907–912. Google Scholar

150.

M. R. Turetsky , B. D. Amiro , E. Bosch , and J. S. Bhatti , 2004: Historical burn area in western Canadian peatlands and its relationship to fire weather indices. Global Biogeochemical Cycles, 18: 1–9. Google Scholar

151.

M. R. Turetsky , R. K. Wieder , D. H. Vitt , R. Evans , and K. D. Scott , 2007: The disappearance of relict permafrost in boreal North America: effects on peatland carbon storage and fluxes. Global Change Biology, 13: 1922–1934. Google Scholar

152.

J. Turunen , T. Tahvanainen , and K. Tolonen , 2001: Carbon accumulation in West Siberian mires, Russia. Global Biogeochemical Cycles, 15: 285–296. Google Scholar

153.

J. Turunen , E. Tomppo , K. Tolonen , and A. Reinikainen , 2002: Estimating carbon accumulation rates of undrained mires in Finland—Application to boreal and subarctic regions. The Holocene, 12: 69–80. Google Scholar

154.

J. Turunen , N. T. Roulet , and T. R. Moore , 2004: Nitrogen deposition and increased carbon accumulation in ombrotrophic peatlands in eastern Canada. Global Biogeochemical Cycles, 18: GB3002, doi:  http://dx.doi.org/10.1029/2003GB002154Google Scholar

155.

S. van Bellen , M. Garneau , and R. K. Booth , 2011: Holocene carbon accumulation rates from three ombrotrophic peatlands in boreal Québec, Canada: impact of climate-driven ecohyrological change. The Holocene, 21: 1217–1231. Google Scholar

156.

S. R. Vardy , B. G. Warner , J. Turunen , and R. Aravena , 2000: Carbon accumulation in permafrost peatlands in the Northwest Territories and Nunavut, Canada. The Holocene 10: 273–280. Google Scholar

157.

G. L. Vourlitis , and W. C. Oechel , 1997: Climate change in northern latitudes: alterations in ecosystem structure and function and effects on carbon sequestration. In W. C. Oechel , T. Callaghan , T. Gilmanov , J. I. Holten , B. Maxwell , U. Molau , B. Sveinbjörnsson , (eds.), Global Change and Arctic Terrestrial Ecosystems. Berlin: Springer, 381–401. Google Scholar

158.

K. L. Webster , and J. W. McLaughlin , 2010: Importance of water table in controlling dissolved carbon along a fen nutrient gradient. Soil Science Society of America Journal, 74: 2254–2266. Google Scholar

159.

K. L. Webster , and J. W. McLaughlin , 2011: Carbon cycling. In K. M. Abraham , L. M. McKinnon , Z. Jumean , S. M. Tully , L. R. Walton , and H. M. Stewart (eds.), Hudson Plains Ecozone+ Status and Trends Assessment. Canadian Biodiversity: Ecosystem Status and Trends 2010. Ottawa, Canada: Canadian Council of Resource Ministers, Technical Ecozone+ Report, 311–321. Google Scholar

160.

K. L. Webster , J. W. McLaughlin , Y. Kim , M. Packalen , and C. S. Li , 2013: Modelling carbon dynamics and response to environmental change along a boreal fen nutrient gradient. Ecological Modelling, 248: 148–164. Google Scholar

161.

J. F. Weltzin , S. D. Bridgham , J. Pastor , J. Chen , and C. Harth , 2003: Potential effects of warming and drying on peatland community composition. Global Change Biology, 9: 141–151. Google Scholar

162.

K. P. Wickland , R. G. Striegl , J. C. Neff , and T. Sachs , 2006: Effects of permafrost melting on CO2 and CH4 exchange of a poorly drained black spruce lowland. Journal of Geophysical Research, 111: G02011, doi:  http://dx.doi.org/10.1029/2005JG000099.  Google Scholar

163.

M.-K. Woo , and P. D. DiCenzo , 1989: Hydrology of small tributary streams in a subarctic wetland. Canadian Journal of Earth Sciences, 26:1557–1566. Google Scholar

164.

J. Wu , L. Kurtzbah , D. Jager , C. Wille , and M. Wilmking , 2010: Evapotranspiration dynamics in a boreal peatland and its impact on the water and energy balance. Journal of Geophysical Research, 115: G04038, doi:  http://dx.doi.org/10.1029/2009JG001075Google Scholar

165.

Z. Yu , 2006: Holocene carbon accumulation of fen peatlands in boreal western Canada: a complex ecosystem response to climate variation and disturbance. Ecosystems, 9: 1278–1288. Google Scholar

166.

Z. Yu , I. D. Campbell , C. Campbell , D. H. Vitt , G. C. Bond , and M. J. Apps , 2003: Carbon sequestration in western Canadian peat highly sensitive to Holocene wet-dry climate cycles at millennial timescales. The Holocene, 13: 801–808. Google Scholar

167.

Z. C. Yu , 2012: Northern peatland carbons stocks and dynamics: a review. Biogeosciences, 9: 4071–4085. Google Scholar

168.

Y. Zhang , C. Li , C. C. Trettin , H. Li , G. Sun , 2002: An integrated model of soil, hydrology, and vegetation for carbon dynamics in wetland ecosystems. Global Biogeochemical Cycles, 16: 1061, doi:  http://dx.doi.org/10.1029/2001GB001838Google Scholar

169.

Y. Zhang , W. Chen , and J. Cihlar , 2003: A process-based model for quantifying the impact of climate change on permafrost thermal regimes. Journal of Geophysical Research, 108: 4695, doi:  http://dx.doi.org/10.1029/2002JD003354Google Scholar

170.

Y. Zhang , W. Chen , and D. W. Riseborough , 2008: Transient projections of permafrost distribution in Canada during the 21st century under scenarios of climate change. Global Planetary Change, 60: 443–456. Google Scholar
Jim McLaughlin and Kara Webster "Effects of Climate Change on Peatlands in the Far North of Ontario, Canada: A Synthesis," Arctic, Antarctic, and Alpine Research 46(1), 84-102, (1 February 2014). https://doi.org/10.1657/1938-4246-46.1.84
Accepted: 1 June 2013; Published: 1 February 2014
Back to Top