Open Access
How to translate text using browser tools
1 February 2014 Peatland Initiation, Carbon Accumulation, and 2 ka Depth in the James Bay Lowland and Adjacent Regions
James R. Holmquist, Glen M. MacDonald, Angela Gallego-Sala
Author Affiliations +
Abstract

Peatlands surrounding Hudson and James Bays form the second largest peatland complex in the world and contain major stores of soil carbon (C). This study utilized a transect of eight ombrotrophic peat cores from remote regions of central and northern Ontario to quantify the magnitude and rate of C accumulation since peatland initiation and for the past 2000 calendar years before present (2 ka). These new data were supplemented by 17 millennially resolved chronologies from a literature review covering the Boreal Shield, Hudson Plains, and Taiga Shield bordering Hudson and James Bays. Peatlands initiated in central and northern Ontario by 7.8 ka following deglaciation and isostatic emergence of northern areas to above sea level. Total C accumulated since inception averaged 109.7 ± (std. dev.) 36.2 kg C m-2. Approximately 40% of total soil C has accumulated since 2 ka at an average apparent rate of 20.2 ± 6.9 g C m-2 yr-1. The 2 ka depths correlate significantly and positively with modern gridded climate estimates for mean annual precipitation, mean annual air temperature, growing degree-days > 0 °C, and photosynthetically active radiation integrated over days > 0 °C. There are significantly shallower depths in permafrost peatlands. Vertical peat accumulation was likely constrained by temperature, growing season length, and photosynthetically active radiation over the last 2 ka in the Hudson Bay Lowlands and surrounding regions.

Introduction

Northern peatlands have acted as a terrestrial carbon (C) sink since the end of the last glacial maximum (MacDonald et al., 2006) and are estimated to have stored over 30% of global soil C in only 2–3% of the earth's land surface (Gorham, 1991; Turetsky et al., 2002; Bridgham et al., 2006; Yu et al., 2009). Short growing seasons, low mean annual temperatures, poorly drained land, and complex ecohydrological processes create conditions where net ecosystem productivity can remain positive (Clymo, 1984; van Breemen, 1995; Blodau, 2002; Dise, 2009; Yu et al., 2009). This positive net ecosystem productivity may increase, decrease, or reverse due to projected anthropogenic climate change (Moore and Knowles, 1989; Davidson and Janssens, 2006; Tarnocai, 2006; Beilman et al., 2009; Loisel et al., 2012). Therefore, it is important to understand peatland C storage so that soil C is accounted for in international climate change agreements (Roulet, 2000; Waddington et al., 2009; Dunn and Freeman, 2011) and in improving vegetation, soil, and climate models (Wania et al., 2009a, 2009b), and in evaluating possible mitigation options for projected climate change (Carlson et al., 2010; Freeman et al., 2012).

At present, peatlands store 270 to 450 Pg of C (Gorham, 1991). Global reports state that northern latitude peatlands store an average of 2.3 m depth times 130 kg C m-2 of peat (Gorham, 1991). Long-term net rates of C accumulation in Canadian peatlands typically range from 10 to 35 g C m-2yr-1 (Ovenden, 1990; Gorham et al., 2003). These figures are comparable to other large peatland complexes in the West Siberian Lowlands (WSL), which vary from 5.4 to 35.9 g C m-2 yr-1 (Beilman et al., 2009), as well as various other North American sites. Other North American sites range from 16 to 80 g peat m-2 yr-1 (Gorham et al., 2003), approximately 8 to 40 g C m-2 yr-1 if we assume a ratio of 0.5 g C for every 1 g peat (Turunen et al., 2002). Due to C storage, peatlands are estimated to have had a net global cooling effect of -0.2 to -0.5 W m-2 since their initiation, after an initial net warming effect of approximately 0.1 W m-2 due to CO2 and CH4 emissions (Frolking and Roulet, 2007).

This paper focuses on the Boreal Shield and the Hudson Plains of central and northern Ontario, and it synthesizes data from the Boreal Shield, Hudson Plains, and Taiga Shield bordering the Hudson and James Bays. We present data on the initiation, development, and patterns of C storage in ombrotrophic peatlands during the course of the Holocene, and since 2000 cal. yr BP (2 ka). Although the Arctic and subarctic have experienced variation in climatic conditions over the past 2000 years, this time period represents the period of the Holocene during which natural radiative forcing, boundary conditions, and Arctic climate are most similar to the present (Kaufman et al., 2009). Data on peatland depth can be useful to make baseline estimates of peat production relative to decay and the potential impacts of climate change (Beilman et al., 2009).

Precipitation and surface moisture may be important drivers of peat accumulation in Sphagnum bogs. Sphagnum mosses do not have stomata and cannot regulate water loss during CO2 exchange, making them vulnerable to large shifts in surface moisture (Loisel et al., 2012). Effective moisture has been found to be a significant driver of Sphagnum growth in some continental sites (Loisel et al., 2012). However, in a review of North American peat accumulation, precipitation was found to be significantly, but inversely, correlated with peat depth and the rate of accumulation (Gorham et al., 2003). There is also some evidence that Sphagnum bogs can regulate their own water table depth because of their complex structure and ecohydrology; therefore they may not be sensitive to precipitation-related water stress within their ecological limits (Laiho, 2006; Dise, 2009). This internal feedback relationship has been included in many different peatland development models (Clymo, 1984; Belyea and Baird, 2006; Eppinga et al., 2009; Frolking et al., 2010).

Growing season length and thermal characteristics such as air temperature are possible important drivers of Sphagnum growth and C accumulation. Sphagnum growth rate increases with temperature (Gunnarsson, 2005; Breeuwer et al., 2008). Sphagnum productivity has also been globally linked to photosynthetically active radiation, and growing season length (Loisel et al., 2012). Studies have also observed significant, positive correlations worldwide between mean growing degreedays above 0 °C (GDD0) and rates of peat accumulation (Clymo et al., 1998). As with GDD0, photosynthetically active radiation integrated over the growing season (PAR0) was also shown to be a driver of the rate of post-1 ka apparent C accumulation in a global peatland database (Charman et al., 2013). In an analysis of a widespread network of Sphagnum cores in the WSL, the depths of 2 ka peat correlate significantly and positively with mean annual air temperature (MAAT; Beilman et al., 2009). In the WSL, 2 ka depths are also significantly shallower if permafrost is present, suggesting productivity, rather than low decomposition rates, may exercise greater control on peat accumulation at these sites (Beilman et al., 2009).

The peatlands surrounding Hudson and James Bays in Quebec, Ontario, and Manitoba in Canada (Figs. 1 and 2), are second to the WSL as the largest continuous peatland complex in the world (Riley, 2005, 2011). Despite the importance of this area to global C cycling, only a few studies have estimated average apparent rate of C accumulation over the broad regions of northern Ontario (O'Reilly, 2011; Bunbury et al., 2012), mainly due to inaccessibility by surface transport. In addition to this, no studies have synthesized current patterns, or analyzed connections between peat accumulation and climate over the entire area surrounding the Hudson and James Bays.

FIGURE 1.

Our study sites were located in the James Bay Lowland (JBL), and were supplemented by a review of 17 additional age-depth models from surrounding regions in Canada. Pictured are also maps of peatland extent (Tarnocai et al., 2011), and the borders of discontinuous and continuous permafrost zones (Brown et al., 2001).

f01_19.jpg

This study examines the rates and magnitudes of Holocene and post-2 ka peat and C accumulation in central and northern Ontario based on new data. The study also investigates the relationship between C accumulation and climate based on this new information, and synthesizes existing studies. Radiocarbon (14C)-based estimates of peatland initiation; timing; estimates of average Holocene, post-2 ka, and pre-2 ka apparent rate of C accumulation; and Holocene, post-2 ka, and pre-2 ka C mass totals are presented for eight previously undescribed ombrotrophic peatlands in Ontario (Fig. 1; Table 1). These data are combined with an analysis of 17 other cores from ombrotrophic peat sites from three ecozones surrounding the Hudson and James Bays, which have l4C chronologies of approximately millennial resolution (Dredge and Mott, 2003; Glaser et al., 2004; Arlen-Pouliot and Bhiry, 2005; Kuhry, 2008; Beaulieu-Audy et al., 2009; Sannel and Kuhry, 2009; Loisel and Garneau, 2010; van Bellen et al., 2011; Bunbury et al., 2012). The analysis is used to determine if rates and magnitudes of accumulation at the new sites are representative of 2 ka peat depth, and to gauge the wider effects of climate on 2 ka peat depth.

FIGURE 2.

Landscapes that are representative of the surface cover in the JBL, including: (a) Picea mariana dominated stands, (b) uncovered Sphagnum bogs, (c) minerotrophic thin fens, and (d) open water. Photos are courtesy of David Beilmen, and are used with permission (Beilman, personal communication).

f02_19.jpg

There were two major objectives of this study. The first was to quantify the initiation time, soil C density, and apparent rate of C accumulation in bog peat in central and northern Ontario for both pre-2 ka and post-2 ka. The second was to determine the effect that climate has had on vertical peat accumulation over the larger area surrounding the Hudson and James Bays. We hypothesized that apparent C accumulation and 2 ka depth would correlate significantly and inversely with modern mean annual precipitation (MAP), and positively with modern MAAT, GDD0, and PAR0. We also hypothesized that permafrost occurrence has had a significant effect on peat depth, and 2 ka peatland depth with permafrost would be significantly shallower than non-permafrost locations.

TABLE 1

Site information for 8 new James Bay Lowland and adjacent regions sites collected in the summer of 2008. Sites are listed from south to north.

t01_19.gif

Materials and Methods

FIELD SITES

The Ontario James Bay Lowland (JBL), in which most of the study sites are located (Fig. 1), covers 221,000 km2; 36% of the wetlands present are bogs dominated primarily by Sphagnum (Riley, 2005). The Hudson Plains contain Picea mariana-dominated forests in the south (Fig. 2) and mixed P. mariana and Larix larcina forests in the north (Riley, 2005, 2011). These are interspersed with unforested bog and fen complexes dominated by Sphagnum mosses and Carex sp., respectively (Fig. 2). Northern permafrost bogs have drier surfaces and are often typified by a surface cover of Cladonia. The Hudson Plains formed after the last glacial maximum following the retreat of the Laurentide ice sheet commencing at approximately 10 ka. The Hudson Plains remained largely covered with the remaining ice sheet and glacial lakes Agassiz and Ojibway until the first land emergence around 9 ka (Dyke et al., 2003). Many peatlands in the southern lowlands were formed after the catastrophic drainage of the glacial lakes due to the breaking of the Laurentide ice dam around 8.4 ka (Lajeunesse and St-Onge, 2008). The rest of the land gradually became available for peatlands due to isostatic rebound. The Hudson Plains have had some of the fastest rates of isostatic uplift on earth, with some areas rising an average of 1.2 m century-1 (Webber et al., 1970).

Permafrost occurs in the Hudson Bay coast as well as northern regions of the Hudson Bay and James Bay Lowlands. It dominates the landscape in the north within ∼80 km of the coast, and occurs along the coast in the south within a slimming margin approximately 20 to 40 km wide (Fig. 1; Riley, 2005). Along the coast, continuous permafrost and subarctic vegetation can be found anywhere where soil temperatures remain below 0 °C annually, and discontinuously south of this isotherm due to the efficient insulating qualities of peat soils (Riley, 2005).

The peatlands of central and northern Ontario are accessible almost exclusively by helicopter, boat, or winter ice road, and remain both sparsely populated and understudied. Some peatland basal dates and paleoecological information are available from studies focused near the Albany River (Glaser et al., 2004), Churchill Rail Line (Dredge and Mott, 2003), and Victor Mine (O'Reilly, 2011; Bunbury et al., 2012). We collected eight bog cores from eight sites by helicopter during the summer of 2008 (Fig. 1; Table 1). The sites form a latitudinal transect between 50°27′N and 55°24′N (Fig. 1; Table 1). All cores were collected from raised, ombrotrophic peat landforms (Table 1). MAP and MAAT data from three climate stations within the approximate range of the sites are included in Table 2 (Fig. 1; Environment Canada, 2013). In 2008 these climate stations were within the peatlands' temperature-precipitation space described in Yu et al. (2009), ranging from -4.88 to 3.03 °C, and 607.8 to 683.5 mm (Table 2; Environment Canada, 2013).

TABLE 2

Climate data for 2008 from 3 stations in the JBL and adjacent regions.

t02_19.gif

Six cores—JBL1, JBL2, JBL3, JBL4, JBL7, JBL8— were from Sphagnum fuscum-dominated hummock surfaces and were permafrost free (Table 1). These hummocks existed in complexes with S. angustifolium hollows and intermediate S. magellanicum sections. Hummock-hollow patterning reflects Sphagnum species optimal growth and community ecology (Gunnarsson, 2005). Notable vascular species include tree species (Picea mariana, Larix larcina), shrub species (Chamaedaphne sp., Vaccinium sp., Salix sp., Betula sp., Eriophorum sp.), as well as Carex sp. and Equisetum sp. in hollows. JBL2 and JBL3 existed as bog islands in larger poorfen complexes. JBL4 was from the discontinuous permafrost zone and contained examples of boreal vegetation as well as subarctic Cladonia sp. on higher microsites. Two cores, JBL5 and JBL6, were from Cladonia-dominated surfaces underlain by Sphagnum-dominated permafrost peat (Table 1). JBL5 was a degraded permafrost plateau with large cracks filled with meltwater. JBL5's surface cover was predominantly Cladonia sp., Ledum sp., and unvegetated surfaces of decayed peat.

At each core site, the first 0.8 to 1 m was sampled with a box corer and the remainder with a Russian auger in non-permafrost peatlands and a motorized SIPRE drill in permafrost. Seven of the cores contained complete profiles from the surface to the mineral/ peat interface. JBL6 was recovered without the basal section and is therefore omitted from the pre-2 ka C accumulation and total C accumulation data sets. Cores were wrapped in plastic wrap and aluminum foil and placed in a freezer on arrival at the University of California, Los Angeles. All cores were subsampled, while frozen, in 1 cm increments using a bandsaw.

We supplemented our 8 cores with a review of 17 other cores from a broader geographical region encompassing the Boreal Shield, Hudson Plains region, and Taiga Shield surrounding the Hudson and James Bays. The entire data set ranges in latitude from 50°27′N to 60°50′N (Fig. 1; Tables 1 and 3). MAP for the entire network ranged from to 238.0 to 861.50 mm, and MAAT ranged from -8.72 to 0.35 °C from 1975 to 2005 (Matsuura and Willmott, 2009). Out of the combined data set, 19 of the sites are from ombrotrophic bogs, whereas 6 are from permafrost plateaus or polygonal peat formations (Tables 1 and 3). We avoided analyzing cores from fens or sites with chronologies that were less than approximately millennially resolved.

TABLE 3

Site information from 17 Hudson Bay Lowlands-JBL and adjacent regions ombrotrophic peat profiles with millennially resolved age-depth models, listed by author.

t03_19.gif

SOIL C ESTIMATES

We estimated percentage C for every 1 cm increment as the product of dry bulk density (BD), loss on ignition at 550 °C (LOI550), and a %C assumption of 0.5 g C 1 g OM-1 (Turunen et al., 2002). Samples for BD and LOI550 were taken from subsampled cores with a stainless steel tube with a 1 cm diameter. Samples were measured lengthwise with digital calipers to calculate volume. Dry BD was measured by dividing the dry weight by the sample volume after drying to a constant mass at 105 °C. LOI550 was calculated as the mass lost after one hour ignition in a muffle furnace at 550 °C (Sheng et al., 2004).

AGE DEPTH MODELING

Chronologies for the eight new cores were established using multiple 14C dates (Table 4). Dates were selected in order to have approximately millennially resolved chronologies before 1 ka, and bicentennially resolved chronologies since 1 ka. Plant macrofossils were separated from peat with distilled H2O. Fungal hyphae and plant rootlets were removed using forceps under a dissecting microscope. Macrofossils were identified down to the genus level when possible using a stereomicroscope and Canadian macrofossil identification guides (Lévesque et al., 1988). Bulk peat was used when no single macrofossils were identified. Samples were pre-treated using acid-base-acid treatments at 65 °C to remove carbonates, humic acids, and dissolved organic C (Olsson, 1986). Samples were vacuum-sealed in quartz tubes with CuO powder and Ag wire and combusted for 4 h at 900 °C. Samples were graphitized, and atomized in an accelerator mass spectrometer at the Keck Laboratory at the University of California, Irvine. The ratio of 14C to l2C was calculated relative to blank and standards from dendrochronologically dated wood.

Multiple 14C dates (Table 4; Appendix Table A1) were calibrated, and an age-depth model was created for each core using ‘Bacon’ (Fig. 3 and Appendix Fig. A1), a flexible Bayesian age-depth modeling software (Blaauw and Christen, 2011) that is coded using the computer language R (R Development Core Team, 2012). Dates were calibrated using INTCAL09 (Reimer et al., 2009), and NH1 for post-bomb dates (Hua and Barbetti, 2007). We assumed a date of -58 BP for the surface of each new core except for JBL5, which had surface peat that dated pre-modern (15 14C age; Table 4). Seventeen age-depth models from the previously published sites were recalculated using “Bacon” to standardize methods. The surface date was used as a calibration point if it was listed in the literature.

FIGURE 3.

‘Bacon’ age-depth models for 8 new JBL cores. The best estimates for calibrated age are shown in black. The upper and lower estimates are shown in gray.

f03_19.jpg

We preferred ‘Bacon’ to linear interpolation between dates because the latter can be too restrictive if cores are longer than 1 m and do not have high-resolution dating (Blaauw and Heegaard, 2012). ‘Bacon’ deals with outliers in a standardized way using a robust Student's t method (Christen and Pérez, 2011). Bayesian algorithms require priors. We used the program's default settings for the priors: shape (2), memory strength (4), and memory mean (0.7). For the accumulation rate prior we used a default value of 20 yr cm-1, and used 50 yr cm-1 if the default did not produce a parsimonious age-depth model. Only JBL7 required a different accumulation prior of 40 yr cm-1.

AVERAGE APPARENT RATE OF C ACCUMULATION

We calculated the mass of C accumulated from initiation to the surface, 2 ka depth to the surface, and initiation to 2 ka (Beilman et al., 2009). For average Holocene apparent rate of C accumulation, we summarized the mass of C and divided the mass by the amount of time since initiation. For post-2 ka apparent rate of C accumulation, we summarized the amount of C accumulated since the closest approximation to 2 ka and divided the mass by that closest approximation. For pre-2 ka apparent rate of C accumulation, we summarized the mass of C accumulated from initiation to 2 ka and divided it by the difference between initiation date and the closest 2 ka date.

STATISTICAL ANALYSIS

We statistically tested hypotheses regarding the relationships between climate and 2 ka depths, as well as permafrost occurrence on both total depths, and 2 ka depths. To test the hypotheses that MAP, MAAT, GDD0, and PAR0 correlate significantly with 2 ka depth, we used linear regressions, and climate data from 0.5 × 0.5° integrated gridded databases for MAP and MAAT (1975–2005; Matsuura and Willmott, 2009). Initial analysis and previous research (Beilman et al., 2009) indicated the possibility of an exponential relationship between MAAT and 2 ka depth, so we included an exponential regression in our analysis of these two variables. For measurements involving seasonality we used 0 °C as a base temperature because Sphagnum species are adapted to growth at low temperatures (Asada et al., 2003). We calculated a 0.5 × 0.5° gridded database for GDD0 from daily temperature values and PAR0 from latitude, orbital parameters, and the fraction of sunshine hours (Prentice et al., 1993; Hijmans et al., 2005), using the CLIMATE 2.2 database (Kaplan et al., 2003). We also tested the hypothesis that permafrost peatlands are significantly shallower than nonpermafrost peatlands, using a one-way ANOVA with permafrost occurrence as the independent variable, and total depth, or 2 ka depth, as the dependent variable. All statistics were calculated in R (R Development Core Team, 2012).

TABLE 4

14C-AMS samples, depth, 14C ages, and best fit “Bacon” model calibration estimates for 8 new southwest JBL cores.

t04a_19.gif

Continued.

t04b_19.gif

TABLE 5

Peatland basal depths, basal ages, and 2 ka depths for 17 review and 8 new JBL and adjacent regions cores. Data for loss on ignition at 550 °C (LOI550), mean bulk density (BD), and apparent rate of C accumulation; C mass data since inception, post-2 ka, and pre-2 ka are displayed as well.

t05_19.gif

Results

PEAT INITIATION ESTIMATES, DEPTHS, AND CHRONOLOGIES

A full report of 14C dates in the eight new cores is available in Figure 1 and Table 4. The timing of peat initiation in the eight new sites followed major deglaciation, glacial lake drainage, and land emergence around 8.34 ka (Dyke et al., 2003). Basal dates from the eight new cores vary between 7.8 ka and 4.4 ka, and have an average initiation time of 6.3 ± (std. dev.) 1.2 ka (Fig. 1; Table 5). Peat depth averaged 256 ± 98 cm and ranged from 141 to 422 cm (Table 5). Core lengths from previously published northeastern Canadian studies ranged from 70 to 483 cm, and had a mean of 266 ± 109 cm (Table 5). These mean depths roughly correspond to the average depths estimated for northern peatlands (2.3 m; Gorham, 1991).

FIGURE 4.

Post-2 ka C accumulated in the 8 southwest JBL sites arranged by latitude. (-) Represent the upper and lower estimates of Post-2 ka C masses based on upper and lower estimates of 2 ka by ‘Bacon’ models. (*) Represents permafrost peat.

f04_19.jpg

SOIL C MASS IN CENTRAL AND NORTHERN ONTARIO

BD in the new cores ranged from 0.0034 to 0.62 g cm-3 and averaged 0.093 ± 0.041 g cm-3 (Table 5). LOI550 ranged from 8.7 to 100.0 %OM and averaged 94.5 ± 6.3 %OM. The highest LOI_50 generally corresponded to the low-density acrotelm and active layer sections. The exception to this is JBL5, which has a densely packed rootlet/lichen active layer (Table 5). The lowest LOI550 values generally correspond to highest BD values at the mineral dominated basal sections of JBL3 and JBL5. Our mean organic matter density, 0.087 ± 0.03 is slightly lower than the mean value for fens and bogs in western Canada, 0.094 g cm-3 (Vitt et al., 2000a), and the range of measurements from the Mackenzie River Basin, Finland, and the WSL (0.092–0.094 g cm-3; Makila, 1994; Sheng et al., 2004; Beilman et al., 2009). There were generally lower BD and higher LOI550 in the top meter of peat, and higher BD and lower LOI550 in deeper peat due to decay and compaction over time in deeper peat.

More C was older than 2 ka than was younger in the eight new cores. This is not surprising as the pre-2 ka portion of the cores typically represents >5 ka to 2.4 ka of deposition. Pre-2 ka stocks range from 28.3 to 108.7 kg C m-2 and averaged 67.5 ± 30.0 kg C m-2. Post-2 ka C stocks range from 16.3 to 62.5 kg C m-2 and averaged 40.5 ± 14.3 kg C m-2 (Fig. 4 and Table 5). In central and northern Ontario, about 40% of the total C present in the cores is younger than 2 ka. Notable exceptions to this trend were JBL4 and JBL8 where pre-2 ka apparent rate of C accumulation was drastically lower than post-2 ka (Table 5).

Apparent rate of C accumulation was generally higher post-2 ka than total in the eight new cores. Average apparent rate of C accumulation since initiation ranged from 12.8 to 26.7 g C m-2 yr-1, with a mean of 17.4 ± 5.0 g C m-2 yr-1 (Table 5). These values correspond to estimates of 13 to 30 g C m-2 yr-1 for North American peat bogs (Gorham, 1991; Turunen et al., 2002; Kuhry and Turunen, 2006), and fall in the range of observed global peatland C accumulation (8.4 to 38.0 g C m-2 yr-1; Yu et al. 2009). Post-2 ka apparent rate of C accumulation ranged from 8.5 to 30.8 g C m-2 yr-1 and had an average of 20.2 ± 6.9 g C m-2 yr-1. The lowest post-2 ka estimate in JBL5 (8.5 g C m-2 yr-1) roughly corresponds with projected rates for subarctic Canada (9 g C m-2 yr-1; Gorham, 1991), and the lowest observed arctic estimates from a global synthesis (8.4 g C m-2 yr-1; Yu et al. 2009). Pre-2 ka apparent rate of C accumulation was generally lower than post-2 ka apparent rate of C accumulation and ranged from 8.2 to 24.8 g C m-2 yr-1 and averaged 15.0 ± 5.6 g C m-2 yr-1 (Table 5).

In the eight new sites, both apparent rate of C accumulation and 2 ka depth correlate significantly and positively with MAP with r2 values of 0.56 and 0.58, respectively (p <0.05; Table 6). 2 ka depths also correlate significantly and positively with MAAT (p < 0.05; r2 = 0.52; Table 6).

CLIMATE AND 2 ka DEPTH

Based on the combined data set of our 8 new cores and the 17 previously published cores, peat depth at 2 ka correlated significantly and positively with MAP, MAAT, GDD0, and PAR0. Depths at 2 ka ranged from 8 to 173 cm with a mean of 88.5 ± 41.1 cm (Table 5). Overall peat depth ranged from 70 to 483 cm, meaning that the average proportion of peat vertical accumulation since 2 ka was 36.9 ± 19.1%, roughly similar to the 40% of C present that was younger than 2 ka in our eight new sites. Depth at 2 ka correlates significantly and positively with MAP (p < 0.05; r2 = 0.20; Fig. 5; Table 6), MAAT (p < 0.001; r2 = 0.46; Fig. 5; Table 6), GDD0 (p < 0.0001; r2 = 0.60; Fig. 5; Table 6), and PAR0 (p < 0.0001; r2 = 0.62; Fig. 5; Table 6). An exponential regression between MAAT and 2 ka depth increased the r2 value to 0.49 (p < 0.0001; Fig. 5; Table 6). A linear regression defining PAR0 as a predictor of 2 ka depth describes the most variance out of all of the variables we tested (Fig. 5; Table 6).

PERMAFROST OCCURRENCE AND PEAT DEPTH

Permafrost occurrence was found to have a significant effect on 2 ka depths, and on total depths in the 25 new and review cores. A one-way ANOVA determined that there was a statistically significant difference between the depths of permafrost peatlands, and non-permafrost peatlands (p < 0.05). A one-way ANOVA showed that permafrost peatlands have a significantly shallower 2 ka depth than non-permafrost peatlands (p < 0.0001; Fig. 6). Permafrost total depths averaged 182.8 ± 31.6 cm, whereas non-permafrost total depth averages 290.6 ± 112.0 cm. Permafrost 2 ka depths averaged 41.3 ± 23.6 cm for permafrost peatlands, while non-permafrost bogs averaged 103.4 ± 33.4 cm (Fig. 6).

Discussion

PEATLAND INITIATION

Peatland initiation followed major deglaciation in central and northern Ontario. The oldest site (JBL3) occurred in an area that deglaciation models indicate was at, or near, major glacial lake drainage (Dyke et al., 2003). Basal dates show a lag between land availability and peat initiation in North America (Hasley et al., 2000; Gorham et al., 2007), and this is likely the case in eastern Canada. Previous studies describe an average 4000-year lag between deglaciation and widespread peat development due to the time it takes for land to become amenable to peat formation, and due to the random chance of dispersal and colonization of peat-forming species (Hasley et al., 2000). The drainage of post-glacial lakes around 8.34 ka is a likely lag factor (Hasley et al., 2000). Another factor may be a delayed Holocene Thermal Maximum in the North Atlantic caused by cooling from the remainder of the decaying Laurentide ice sheet as late as 6 ka (Kaplan and Wolfe, 2006).

FIGURE 5.

Four scatterplots correlating 2 ka depth as a function of environmental variables. Figures are arranged from left to right and top to bottom in order of the linear model's r2 value. Solid lines represent the best fit for linear regressions. Dashed lines represent the best fit for an exponential regression. Triangles represent new JBL site data. Circles represent review data.

f05_19.jpg

TABLE 6

Correlation statistics for James Bay Lowland (JBL) peatland 2 ka apparent rate of C accumulation, 2 ka depths, and Hudson Bay Low-lands (HBL)–JBL synthesis with mean annual precipitation (MAP), mean annual air temperature (MAAT), growing degree days above 0 °C (GDD0), and photosynthetically active radiation integrated over the growing season (PAR0).

t06_19.gif

FIGURE 6.

Bar and whisker plot of 2 ka depths in permafrost and non-permafrost peat. Central lines represent medians. Box edges represent the 25% upper and lower quantiles. Whisker lines represent least and greatest values.

f06_19.jpg

There was a high rate of peat initiation between 6 and 7.8 ka in central and northern Ontario (Fig. 1). However, the eight new initiation estimates do not show a clear stratification with newer initiation occurring closest to the James Bay coast. The youngest core (JBL8) occurs at the southernmost end of the transect, whereas the oldest (JBL3) occurs in the middle (Fig. 1). Stratification of initiation was discussed by a study closer to the southwest coast of the JBL near the Albany River, where rates of isostatic uplift are the highest (Fig. 1; Glaser et al., 2004). This difference is likely due to the fact that catastrophic drainage of glacial lakes Agassiz and Ojibway were responsible for land emergence in some of the eight new sites, rather than gradual isostatic uplift that drives initiation in the Albany River region (Glaser et al., 2004).

C ACCUMULATION IN CENTRAL AND NORTHERN ONTARIO

C accumulation measurements in the eight new sites showed that 40% of C was younger than 2 ka in Sphagnum bogs. This is roughly comparable to estimates for WSL, where 41% of C is younger than 2 ka (Sheng et al., 2004; Beilman et al., 2009). Only two of our sites, JBL4 and JBL8, have more post-2 ka C than pre-2 ka C, which may have been caused by low relative pre-2 ka accumulation, relatively high belowground C mineralization, or a combination of both. Apparent rate of C accumulation is the result of autogenic drivers, allogenic drivers, and complex internal ecohydrological feedbacks, and has the potential to vary substantially among peatlands (Belyea and Baird, 2006).

MAAT, GDD0, AND PAR0 AS DRIVERS OF POST-2 KA VERTICAL PEAT ACCUMULATION

In our analysis, differing scales showed different correlations between post—2 ka apparent rate of C accumulation, 2 ka depth, and four climate variables. In the eight new sites, MAP correlated significantly with both post—2 ka apparent rate of C accumulation and 2 ka depth. This correlation was also significant and positive in our review data set, but was low relative to MAAT, GDD0, and PAR0. MAP may exercise some local control over peat accumulation in the Boreal Shield and Hudson Plains of Ontario, but temperature, seasonality, and photosynthetically active radiation are likely the major drivers of vertical peat accumulation in all peatlands bordering the Hudson and James Bays. This supports the conclusions of a recent global synthesis of peatland C accumulation rates. Water table depths need to be persistently high for peat to form; however, if they are present, they do not explain any of the variance of C accumulation rates (Charman et al., 2013). It is also possible that small sample size in the new sites (n = 8) may lower the data set's descriptive power.

In the overall synthesis, there was a significant and positive correlation between 2 ka and MAP (p < 0.05; Table 6); however, we did not observe the inverse relationship with precipitation similar to that reported by a broad survey of North American peatlands (Gorham et al., 2003). This may be due to the topographic controls in southern sites (Gorham et al., 2003) that are not present in the more topographically homogeneous Hudson Plains, which we heavily sampled. This was also not a robust relationship because the correlation was heavily influenced by EL1 site data (MAP = 208 mm, 2 ka depth = 8 cm). If that point is deleted, the r2 value lowers to 0.087, and the statistical significance is eliminated.

Our study showed that MAAT had a significant effect on vertical peat accumulation since 2 ka in northeastern Canada (Fig. 5; Table 6). This is analogous to the relationship found between MAAT and 2 ka depth in WSL; however, there are key differences between the two areas. In WSL, data fits an exponential regression that explains more variation (r2 = 0.82; Beilman et al., 2009) than does our data's exponential regression (r2 = 0.49; Table 6). In WSL there was less variability in 2 ka depth for permafrost sites, as well as deeper 2 ka depths in southern non-permafrost bogs (Beilman et al., 2009).

Our hypotheses, that GDD0, and PAR0 are major drivers of peat accumulation, were supported by significant and positive correlations between these variables and 2 ka depths. PAR0 had the highest correlation (p < 0.0001, r2 = 0.62; Table 6). In a previous study that analyzed global Sphagnum productivity, a linear regression between PAR0 and Sphagnum productivity explained a significant but comparatively small amount of variance (r2 = 0.23; Loisel et al., 2012). PAR0 integrates photosynthetically active radiation over the growing season length and can potentially influence two different aspects of peat formation. Photosynthetically active radiation and growing season length may both drive plant growth, and thus litter input (Clymo et al., 1998). Growing season length also affects the length of time each year that the ground is unfrozen and biomass can pass from the acrotelm to the catotelm (Clymo et al., 1998; Gorham et al., 2003). Our results also support a global study of peatland apparent rate of C accumulation over the last 1000 years, in which GDD0 and PAR0 were found to correlate significantly with post-1 ka apparent rate of C accumulation (Charman et al., 2013). However, the significance of radiation as a driver of Sphagnum growth has been debated because of the high degree of shading in boreal peat bogs, and because of photoinhibition in Sphagnum photosynthetic tissues (Harley et al., 1989; Murray et al., 1993).

The relationship of growing season length and photosynthetically active radiation on vertical peat accumulation is important because northern latitudes will be disproportionately affected by global warming. Models of future anthropogenic climate forcing predict disproportionate seasonality changes in northern latitudes, and increases in the length of the growing season (Pachauri and Reisinger, 2007). Between 1982 and 1999 the effective start of the growing season advanced 5.4 days in the northern hemisphere (Jeong et al., 2011). Remote sensing images indicate that the plant growth in the Canadian Low Arctic has increased 0.46–0.67% yr1 from 1982 to 2006 (Jia et al., 2009). However, there will likely be complications modeling PAR0 as a predictive bioclimatic variable because of the uncertainty in model prediction of future cloud cover (Charman et al., 2013).

THE EFFECT OF PERMAFROST ON PEAT DEPTH

Significantly shallower basal depths and 2 ka depths in permafrost peatlands compared to peatlands that were permafrost-free supported our hypothesis that permafrost occurrence inhibits peat accumulation and that permafrost occurrence significantly decreased post-2 ka vertical accumulation. This was supported by significantly shallower 2 ka depths in permafrost peat, compared to non-permafrost peat (p < 0.0001). Previous studies indicated that permafrost likely formed in western Canada during the Little Ice Age (Vitt et al., 2000b) during the late Holocene. However, the timing of permafrost formation in eastern Canada is not as well understood. Overall shallower permafrost peat may be explained either by permafrost establishment earlier than 2 ka, or by general lower productivity due to constantly lower MAAT, GDD0, and PAR0 without the direct influence of permafrost.

There was consistent vertical growth in peatlands post-2 ka. Only two cores, JBL5 and EL1, contained near-surface macrofossils that dated pre-modern. JBL5 had dates of 204 and 207 yr BP at 8 and 12 cm, respectively (Table 4), and EL1 had a 13 cm depth dating near 3.0 ka (Appendix Table A1). Besides these two sites, there is little evidence for the recent shutdown by permafrost of 2 ka C accumulation described in the WSL (Beilman et al., 2009). This study showed reduced but ongoing vertical accumulation, similar to the described activity in a survey of subarctic palsas (Olefeldt et al., 2012).

Controls by permafrost of late Holocene peat growth show that productivity, rather than decay, currently drives the significant climatically related differences in peat growth observed here. However, the resulting changes in hydrology and CH4 emissions from melting permafrost will likely result in a greater uncertainty in peatland C cycle feedbacks (Pachauri and Reisinger, 2007). Past studies of northern and southern regions in the Hudson Plains show that CH4 emissions were lower than previously estimated (Roulet et al., 1994). If non-permafrost peatlands store C faster than permafrost peatlands, then peatlands could act as a negative feedback to future warming (Turetsky et al., 2007; Beilman et al., 2009) or at least as transient reactions to warming. The effect of permafrost melt on hydrology is also controversial. Some studies project that permafrost melts will increase drainage (Jorgenson and Osterkamp, 2005; Riordan et al., 2006), while others claim that it will increase peatland extent (Payette et al., 2004).

Conclusion

At eight new study sites from the Boreal Shield and Hudson Plains of Ontario, peatland initiation was concentrated between 6 ka and 7.8 ka and lagged land emergence following deglaciation, major glacial lake drainage, and isostatic rebound. Following initiation, peatlands have acted as a sink of atmospheric CO2. Younger initiations are also present (as late as 3.4 ka) in our data set. Of the eight peatlands that we studied, approximately 40% of their total soil C, 16.3 to 62.5 kg C m-2, is younger than 2 ka. Average rates of apparent C accumulation post-2 ka are 20.2 ± 6.9 g C m-2 yr-1. Across the Boreal Shield, Hudson Plains, and Taiga Shield surrounding the Hudson and James Bays, vertical peat growth since 2 ka correlates significantly and positively with modern mean annual precipitation. Depth at 2 ka also correlates significantly and positively with atmospheric thermal properties, and most closely with photosynthetically active radiation integrated over the growing season. There is also evidence for significantly reduced, although continuing, post-2 ka peat accumulation in permafrost bogs. These two pieces of evidence suggest the possibility that initial increased productivity of northern peatlands due to projected climate change in the 21st century and the arctic amplification of warming could produce increased C storage and a potential negative feedback to potential climate warming at least in a transient fashion until the climatic envelope for peatland maintenance is exceeded in northern regions.

Acknowledgments

The authors acknowledge the U.S. National Science Foundation for funding this research (NSF-0843685; NSF-0628598). We also acknowledge Dave Beilman for contributions to fieldwork; Matt Zebrowski for mapping assistance; and Siduo Zhang, Luis Rodriguez, Jennifer Kim, Karly Wagner, and Nicolai Kondov for contributions to laboratory processing. We also thank Nigel T. Roulet and one anonymous reviewer for their constructive criticisms and suggestions for improving this manuscript. We finally thank the people of the communities of Thunder Bay, Pickle Lake, and Big Trout Lake for their assistance and hospitality.

References Cited

1.

Y. Arlen-Pouliot , and N. Bhiry , 2005: Palaeoecology of a palsa and a filled thermokarst pond in a permafrost peatland, subarctic Québec, Canada. The Holocene , 15(3): 408–419. Google Scholar

2.

T. Asada , B. G. Warner , and A. Banner , 2003: Growth of mosses in relation to climate factors in a hypermaritime coastal peatland in British Columbia, Canada. The Bryologist , 106(4): 516–527. Google Scholar

3.

V. Beaulieu-Audy , M. Garneau , P. J. Richard , and H. Asnong , 2009: Holocene palaeoecological reconstruction of three boreal peatlands in the La Grande Riviere region, Quebec, Canada. The Holocene , 19(3): 459–476. Google Scholar

4.

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(1): GB1012, doi:  http://dx.doi.org/10.1029/2007GB003112Google Scholar

5.

L. R. Belyea , and A. J. Baird , 2006: Beyond “the limits to peat bog growth”: cross-scale feedback in peatland development. Ecological Monographs , 76(3): 299–322. Google Scholar

6.

M. Blaauw , and J. A. Christen , 2011: Flexible paleoclimate age-depth models using an autoregressive gamma process. Bayesian Analysis , 6(3): 457–474. Google Scholar

7.

M. Blaauw , and E. Heegaard , 2012: Estimation of age-depth relationships. In H. J. B. Birks , A. F. Lotter , S. Juggins , and J. P. Smol (eds.), Tracking Environmental Change Using Lake Sediments. Vol. 5, Data Handling and Numerical Techniques. Dordrecht, Springer, 379–413. Google Scholar

8.

C. Blodau , 2002: Carbon cycling in peatlands: a review of processes and controls. Environmental Reviews , 10(2): 111–134. Google Scholar

9.

A. Breeuwer , M. M. Heijmans , B. J. Robroek , and F. Berendse , 2008: The effect of temperature on growth and competition between Sphagnum species. Oecologia , 156(1): 155–167. Google Scholar

10.

S. D. Bridgham , J. P. Megonigal , J. K. Keller , N. B. Bliss , and C. Trettin , 2006: The carbon balance of North American wetlands. Wetlands , 26(4): 889–916. Google Scholar

11.

J. Brown, O. J. Ferrians Jr., J. A. Heginbottom, and E. S. Melnikov, 1998 revised February 2001]: Circum-Arctic map of permafrost and ground-ice conditions. Boulder, Colorado: National Snow and Ice Data Center/World Data Center for Glaciology. Digital Media (accessed 13 August 2012). Google Scholar

12.

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

13.

M. Carlson , J. Chen , S. Elgie , C. Henschel , Á. Montenegro , N. Roulet , N. Scott , C. Tarnocai , and J. Wells , 2010: Maintaining the role of Canada's forests and peatlands in climate regulation. The Forestry Chronicle , 86(4): 434–443. Google Scholar

14.

D. J. Charman , D. W. Beilman , M. Blaauw , R. K. Booth , S. Brewer , F. M. Chambers , J. A. Christen , A. Gallego-Sala , S. P. Harrison , P. D. M. Hughes , S. T. Jackson , A. Korhola , D. Mauquoy , F. J. G. Mitchell , I. C. Prentice , M. van der Linden , F. De Vleeschouwer , Z. C. Yu , J. Alm , I. E. Bauer , Y. M. C. Corish , M. Garneau , V. Hohl , Y. Huang , E. Karofeld , G. Le Roux , J. Loisel , R. Moschen , J. E. Nichols , T. M. Nieminen , G. M. MacDonald , N. R. Phadtare , N. Rausch , Ü. Sillasoo, G. T. Swindles , E.-S. Tuittila , L. Ukonmaanaho , M. Väliranta , S. van Bellen , B. van Geel , D. H. Vitt , and Y Zhao , 2013: Climate-related changes in peatland carbon accumulation during the last millennium. Biogeosciences , 10: 929–944. Google Scholar

15.

J. A. Christen , and S. Pérez , 2011: A new robust statistical model for radiocarbon data. Radiocarbon , 51(3): 1047–1059. Google Scholar

16.

R. S. Clymo , 1984: The limits to peat bog growth. Philosophical Transactions of the Royal Society of London. B, Biological Sciences , 303(1117): 605–654. Google Scholar

17.

R. S. Clymo , J. Turunen , and K. Tolonen , 1998: Carbon accumulation in peatland. Oikos , 81: 368–388. Google Scholar

18.

E. A. Davidson , and I. A. Janssens , 2006: Temperature sensitivity of soil carbon decomposition and feedbacks to climate change. Nature , 440(7081): 165–173. Google Scholar

19.

N. B. Dise , 2009: Peatland response to global change. Science , 326(5954): 810. Google Scholar

20.

L. A. Dredge , and R. J. Mott , 2003: Holocene pollen records and peatland development, northeastern Manitoba. Géographie physique et Quaternaire , 57(1): 7–19. Google Scholar

21.

C. Dunn , and C. Freeman , 2011: Peatlands: our greatest source of carbon credits? Carbon Management , 2(3): 289–301. Google Scholar

22.

A. S. Dyke , A. Moore , and L. Robertson , 2003: GEOSCAN Database. Natural Resources Canada, < http://geoscan.ess.nrcan.gc.ca> (accessed 5 February 2010). Google Scholar

23.

Environment Canada, 2013: National Climate Data and Information Archive. Environment Canada, < http://www.climate.weatheroffice.gc.ca> (accessed 7 January 2013). Google Scholar

24.

M. B. Eppinga , P. C. de Ruiter , M. J. Wassen , and M. Rietkerk , 2009: Nutrients and hydrology indicate the driving mechanisms of peatland surface patterning. The American Naturalist , 173(6): 803–818. Google Scholar

25.

C. Freeman , N. Fenner , and A. H. Shirsat , 2012: Peatland geoengineering: an alternative approach to terrestrial carbon sequestration. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences , 370(1974): 4404–4421. Google Scholar

26.

S. Frolking , and N. T. Roulet , 2007: Holocene radiative forcing impact of northern peatland carbon accumulation and methane emissions. Global Change Biology , 13(5): 1079–1088. Google Scholar

27.

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

28.

P. H. Glaser , B. 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(6): 1036–1053. Google Scholar

29.

E. Gorham , 1991: Northern peatlands: role in the carbon cycle and probable responses to climatic warming. Ecological applications , 1(2): 182–195. Google Scholar

30.

E. Gorham , J. A. Janssens , and P. H. Glaser , 2003: Rates of peat accumulation during the postglacial period in 32 sites from Alaska to Newfoundland, with special emphasis on northern Minnesota. Canadian Journal of Botany , 81(5): 429–438. Google Scholar

31.

E. Gorham , C. Lehman , A. Dyke , J. Janssens , and L. Dyke , 2007: Temporal and spatial aspects of peatland initiation following deglaciation in North America. Quaternary Science Reviews , 26(3): 300–311. Google Scholar

32.

U. Gunnarsson , 2005: Global patterns of Sphagnum productivity. Journal of Bryology , 27(3): 269–279. Google Scholar

33.

L. A. Halsey , D. H. Vitt , and L. D. Gignac , 2000: Sphagnum-dominated peatlands in North America since the last glacial maximum: their occurrence and extent. The Bryologist , 103(2): 334–352. Google Scholar

34.

P. C. Harley , J. D. Tenhunen , K. J. Murray , and J. Beyers , 1989: Irradiance and temperature effects on photosynthesis of tussock tundra Sphagnum mosses from the foothills of the Philip Smith Mountains, Alaska. Oecologia , 79(2): 251–259. Google Scholar

35.

R. J. Hijmans , S. E. Cameron , J. L. Parra , P. G. Jones , and A. Jarvis , 2005: Very high resolution interpolated climate surfaces for global land areas. International Journal of Climatology , 25(15): 1965–1978. Google Scholar

36.

Q. Hua , and M. Barbetti , 2007: Review of tropospheric bomb 14C data for carbon cycle modeling and age calibration purposes. Radiocarbon , 46(3): 1273–1298. Google Scholar

37.

S. Jeong , C. Ho , H. Gim , and M. E. Brown , 2011: Phenology shifts at start vs. end of growing season in temperate vegetation over the northern hemisphere for the period 1982–2008. Global Change Biology , 17(7): 2385–2399. Google Scholar

38.

G. J. Jia , H. E. Epstein , and D. A. Walker , 2009: Vegetation greening in the Canadian Arctic related to decadal warming. Journal of Environmental Monitoring , 11(12): 2231–2238. Google Scholar

39.

M. T. Jorgenson , and T. E. Osterkamp , 2005: Response of boreal ecosystems to varying modes of permafrost degradation. Canadian Journal of Forest Research , 35(9): 2100–2111. Google Scholar

40.

J. O. Kaplan , N. H. Bigelow , P. J. Bartlein , T. R. Christensen , W. Cramer , S. P. Harrison , N. V. Matveyeva , A. D. McGuire , D. F. Murray , I. C. Prentice , V. Y. Razzhivin , B. Smith , D. A. Walker , P. M. Anderson , A. A. Andreev , L. B. Brubaker , M. E. Edwards , A. V. Lozhkin , and J. Ritchie , 2003: Climate change and Arctic ecosystems II: modeling, palaeodata-model comparisons, and future projections. Journal of Geophysical Research-Atmospheres , 108: 8171, doi:  http://dx.doi.org/10.1029/2002JD002559Google Scholar

41.

M. R. Kaplan , and A. P. Wolfe , 2006: Spatial and temporal variability of Holocene temperature in the North Atlantic region. Quaternary Research , 65(2): 223–231. Google Scholar

42.

D. S. Kaufman , D. P. Schneider , N. P. McKay , C. M. Ammann , R. S. 20Bradley , K. R. Briffa , G. H. Miller , B. L. Otto-Bliesner , J. T. Overpeck , B. M. Vinther , Arctic Lakes 2k Project Members: M. Abbott , Y. Axford , B. Bird , H. J. B. Birks , A. E. Bjune , J. Briner , T. Cook , M. Chipman , P. Francus , K. Gajewski , Á. Geirsdóttir , F. S. Hu , B. Kutchko , S. Lamoureux , M. Loso , G. MacDonald , M. Peros , D. Porinchu , C. Schiff , H. Seppä , and E. Thomas , 2009: Recent warming reverses long-term Arctic cooling. Science , 325(5945): 1236–1239. Google Scholar

43.

P. Kuhry , 2008: Palsa and peat plateau development in the Hudson Bay Lowlands, Canada: timing, pathways and causes. Boreas , 37(2): 316–327. Google Scholar

44.

P. Kuhry , and J. Turunen , 2006: The postglacial development of boreal and subarctic peatlands. Boreal and Peatland Ecosystems , 188: 21–46. Google Scholar

45.

R. Laiho , 2006: Decomposition in peatlands: reconciling seemingly contrasting results on the impacts of lowered water levels. Soil Biology and Biochemistry , 38(8): 2011–2024. Google Scholar

46.

P. Lajeunesse , and G. St-Onge , 2008: The subglacial origin of the Lake Agassiz-Ojibway final outburst flood. Nature Geoscience , 1(3): 184–188. Google Scholar

47.

P. E. M. Lévesque , H. Dinel , and A. Larouche , 1988: Guide to the Identification of Plant Macrofossils in Canadian Peatlands. Ottawa: Agriculture Canada, Land Resource Research Centre, Research Branch. Google Scholar

48.

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(3): 493– 533. Google Scholar

49.

J. Loisel , A. V. Gallego-Sala , and Z. Yu , 2012: Global-scale pattern of peatland Sphagnum growth driven by photosynthetically active radiation and growing season length. Biogeosciences , 9(84): 2169–2196. Google Scholar

50.

G. M. MacDonald , D. W. Beilman , K. V. Kremenetski , Y. Sheng , L. C. Smith , and A. A. Velichko , 2006: Rapid early development of circumarctic peatlands and atmospheric CH4 and CO2 variations. Science , 314(5797): 285–288. Google Scholar

51.

M. Makila , 1994: Calculation of the energy content of mires on the basis of peat properties. Geological Survey of Finland, Report of Investigation , 121: 1–73. Google Scholar

52.

K. Matsuura , and C. J. Willmott , 2009: Terrestrial Air Temperature and Precipitation: Monthly Climatologies, version 4.01 (1975–2005). < http://climate.geog.udel.edu/~climate/html_pages/Global2_Clim/README.global2_clim.html>, (accessed 4 September 2012). Google Scholar

53.

T. R. Moore , and R. Knowles , 1989: The influence of water table levels on methane and carbon dioxide emissions from peatland soils. Canadian Journal of Soil Science , 69(1): 33–38. Google Scholar

54.

K. J. Murray , J. D. Tenhunen , and R. S. Nowak , 1993: Photoinhibition as a control on photosynthesis and production of Sphagnum mosses. Oecologia , 96(2): 200–207. Google Scholar

55.

D. Olefeldt , N. T. Roulet , O. Bergeron , P. Crill , K. Bäckstrand , and T. R. Christensen , 2012: Net carbon accumulation of a high-latitude permafrost palsa mire similar to permafrost-free peatlands. Geophysical Research Letters , 39(3): L03501, doi:  http://dx.doi.org/10.1029/2011GL050355Google Scholar

56.

I. U. Olsson , 1986: A study of errors in 14C dates of peat and sediment. Radiocarbon , 28(2A): 429–435. Google Scholar

57.

B. O'Reilly, 2011: Paleoecological and Carbon Accumulation Dynamics of a Fen Peatland in the Hudson Bay Lowlands Northern Ontario, from the Mid-Holocene to Present. Master's thesis, Department of Geography, University of Toronto, Toronto, Canada, < https://tspace.library.utoronto.ca/bitstream/1807/31374/l/O'Reilly_Benjamin_C_20111l_MSc_ Thesis.pdf>. Google Scholar

58.

L. Ovenden , 1990: Peat accumulation in northern wetlands. Quaternary Research , 33(3): 377–386. Google Scholar

59.

R. K. Pachauri , and A. Reisinger , 2007: Climate Change 2007: Synthesis Report. Contribution of Working Groups I, II and III to the Fourth Assessment Report of the Intergovernmental Panel on Climate Change. Geneva: IPCC. Google Scholar

60.

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(18): doi:  http://dx.doi.org/10.1029/2004GL020358Google Scholar

61.

I. C. Prentice , M. T. Sykes , and W. Cramer , 1993: A simulation model for the transient effects of climate change on forest landscapes. Ecological Modeling , 65(1): 51–70. Google Scholar

62.

R Development Core Team, 2012: R: A Language and Environment for Statistical Computing. Vienna, Austria: R Foundation for Statistical Computing. ISBN 3-900051-07-0, < http://www.R-project.org/>. Google Scholar

63.

P. J. Reimer , M. G. L. Baillie , E. Bard , A. Bayliss , B. J. Warren , P. G. Blackwell , C. B. Ramsey , C. E. Buck , G. S. Burr , R. L. Edwards , M. Friedrich , P. M. Grootes , T. P. Guilderson , I. Hajdas , T. J. Heaton , A. G. Hogg , K. A. Hughen , K. F. Kaiser , B. Kromer , F. G. McCormac , S. W. Manning , R. W. Reimer , D. A. Richards , J. R. Southon , S. Talamo , C. S. M. Turney , J. van der Plicht , and C. E. Weyhenmeyer , 2009: IntCal09 and Marine09 radiocarbon age calibration curves, 0–50,000 cal years BP. Radiocarbon , 51: 1111–1150. Google Scholar

64.

J. L. Riley , 2005: Flora of the Hudson Bay Lowlands and Its Postglacial Origins. Ottawa, Canada: National Research Council of Canada. ISBN 0-660-18941-0. Google Scholar

65.

J. L. Riley , 2011: Wetlands of the Ontario Hudson Bay Lowland: a Regional Overview. Toronto, Canada: Nature Conservancy of Canada. Google Scholar

66.

B. Riordan , D. Verbyla , and A. D. McGuire , 2006: Shrinking ponds in subarctic Alaska based on 1950-2002 remotely sensed images. Journal of Geophysical Research-Biogeosciences, 111(G4): doi:  http://dx.doi.org/l0.1029/2005JG000150Google Scholar

67.

N. T. Roulet , 2000: Peatlands, carbon storage, greenhouse gases, and the Kyoto protocol: prospects and significance for Canada. Wetlands , 20(4): 605–615. Google Scholar

68.

N. T. Roulet , A. Jano , 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(D1): 1439–1454. Google Scholar

69.

A. B. K Sannel , and P. Kuhry , 2009: Holocene peat growth and decay dynamics in sub-arctic peat plateaus, west-central Canada. Boreas , 38(1): 13–24. Google Scholar

70.

Y. Sheng , L. C. Smith , G. M. MacDonald , K. V. Kremenetski , K. E. Frey , A. A. Velichko , M. Lee , D. W. Beilman , and P. Dubinin , 2004: A high-resolution GIS-based inventory of the West Siberian peat carbon pool. Global Biogeochemical Cycles, 18(3): doi:  http://dx.doi.org/10.1029/2003GB002190Google Scholar

71.

C. Tarnocai , 2006: The effect of climate change on carbon in Canadian peatlands. Global and planetary Change , 53(4): 222–232. Google Scholar

72.

C. Tarnocai , I. M. Kettles , B. Lacelle , 2011: Geological Survey of Canada, Open File 6561. Ottawa: Natural Resources Canada, < http://geoscan.nrcan.gc.ca/starweb/geoscan/servlet.starweb?path=geoscan/fulle.web&search1=R=288786#tphp> (accessed 13 August 2012). Google Scholar

73.

M. Turetsky , K. Wieder , L. Halsey , and D. Vitt , 2002: Current disturbance and the diminishing peatland carbon sink. Geophysical Research Letters , 29(11): 21–1 to 21–4. Google Scholar

74.

M. R. Turetsky , R. K. Wieder , D. H. Vitt , R. J. 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(9): 1922–1934. Google Scholar

75.

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(1): 69–80. Google Scholar

76.

S. van Bellen , M. Garneau , and R. K. Booth , 2011 : Holocene carbon accumulation rates from three ombrotrophic peatlands in boreal Quebec, Canada: impact of climate-driven ecohydrological change. The Holocene , 21(8): 1217–1231. Google Scholar

77.

N. van Breemen , 1995: How Sphagnum bogs down other plants. Trends in Ecology and Evolution , 10(7): 270–275. Google Scholar

78.

D. H. Vitt , L. A. Halsey , I. E. Bauer , and C. Campbell , 2000a: Spatial and temporal trends in carbon storage of peatlands of continental western Canada through the Holocene. Canadian Journal of Earth Sciences , 37(5): 683–693. Google Scholar

79.

D. H. Vitt , L. A. Halsey , and S. C. Zoltai , 2000b: The changing landscape of Canada's western boreal forest: the current dynamics of permafrost. Canadian Journal of Forest Research , 30(2): 283–287. Google Scholar

80.

J. M. Waddington , J. Plach , J. P. Cagampan , M. Lucchese , and M. Strack , 2009: Reducing the carbon footprint of Canadian peat extraction and restoration. AMBIO: a Journal of the Human Environment , 38(4): 194–200. Google Scholar

81.

R. Wania , I. Ross , and I. C. Prentice , 2009a: Integrating peatlands and permafrost into a dynamic global vegetation model: 1. Evaluation and sensitivity of physical land surface processes. Global Biogeochemical Cycles , 23(3): GB3014, doi:  http://dx.doi.org/10.1029/2008GB003412Google Scholar

82.

R. Wania , I. Ross , and I. C. Prentice , 2009b: Integrating peatlands and permafrost into a dynamic global vegetation model: 2. Evaluation and sensitivity of vegetation and carbon cycle processes. Global Biogeochemical Cycles, 23(3): GB3014, doi:  http://dx.doi.org/10.1029/2008GB003413Google Scholar

83.

P. J. Webber , J. W. Richardson , and J. T. Andrews , 1970: Post-glacial uplift and substrate age at Cape Henrietta Maria, southeastern Hudson Bay, Canada. Canadian Journal of Earth Sciences , 7(2): 317–325. Google Scholar

84.

Z. Yu , D. W. Beilman , and M. C. Jones , 2009: Sensitivity of northern peatland carbon dynamics to Holocene climate change. In A. J. Baird , L. R. Belyea , X. Comas , A. S. Reeve , and L. D. Slater (eds.), Carbon Cycling in Northern Peatlands . Washington, D.C.: American Geophysical Union, Geophysical Monograph Series, 184: 55–69. Google Scholar

Appendices

APPENDIX

TABLE A1

14C samples, depth, 14C ages, and best-fit “Bacon” model calibration estimates for 17 previously published HBL-JBL cores.

tA01a_19.gif

Continued.

tA01b_19.gif

Continued.

tA01c_19.gif

FIGURE 1A.

“Bacon” age-depth models for 17 previously published peat cores. The best estimates for calibrated age (ybp) are shown in black. The upper and lower estimates are shown in gray.

fA01a_19.jpg

Continued.

fA01b_19.jpg
© 2014 Regents of the University of Colorado
James R. Holmquist, Glen M. MacDonald, and Angela Gallego-Sala "Peatland Initiation, Carbon Accumulation, and 2 ka Depth in the James Bay Lowland and Adjacent Regions," Arctic, Antarctic, and Alpine Research 46(1), 19-39, (1 February 2014). https://doi.org/10.1657/1938-4246-46.1.19
Accepted: 1 April 2013; Published: 1 February 2014
Back to Top