Vous êtes sur la page 1sur 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/250390611

Review of Water Chemistry Research in Natural and Disturbed Peatlands

Article in Canadian Water Resources Journal / Revue canadienne des ressources hydriques · December 2009
DOI: 10.4296/cwrj3404393

CITATIONS READS
73 2,062

1 author:

Richard A Bourbonniere
Environment & Climate Change Canada
75 PUBLICATIONS 4,088 CITATIONS

SEE PROFILE

All content following this page was uploaded by Richard A Bourbonniere on 30 May 2014.

The user has requested enhancement of the downloaded file.


Review of Water Chemistry Research in Natural
and Disturbed Peatlands

Richard A. Bourbonniere

Abstract: The chemical components in water, and changes thereof, can control the ecology of peatlands
and impact the accumulation of peat; while at the same time ecological characteristics, hydrology
and interactions with the atmosphere and geosphere control peatland water chemistry. This review
summarizes results from Canadian and northern European research published over the past two decades
for major and minor, inorganic and organic constituents dissolved in surface or pore waters associated
with bogs and fens. First those studies that describe natural peatlands are discussed, followed by those
that involve peatlands that are disturbed by various direct or indirect anthropogenic influences. Finally
several recommendations are made regarding future studies and how they can assist in modelling efforts
that can inform peatland management and further scientific inquiry.

Résumé : Les composantes chimiques de l’eau, et les changements qui en découlent, peuvent contrôler
l’écologie des tourbières et influer sur l’accumulation de la tourbe; parallèlement les caractéristiques
écologiques, l’hydrologie et les interactions avec l’atmosphère et la géosphère contrôlent l’hydrochimie de la
tourbière. Cette étude résume les résultats tirés de recherches menées au Canada et en Europe du Nord et
publiées au cours des deux dernières décennies à l’égard des composés inorganiques et organiques, majeurs
et mineurs, dissous dans les eaux interstitielles ou de surface associées aux tourbières et aux marais. Sont
d’abord traitées les études qui décrivent les tourbières naturelles, suivies de celles qui portent sur les tourbières
perturbées par diverses influences anthropiques directes ou indirectes. Enfin, plusieurs recommandations
sont formulées à l’égard de futures études et de la façon dont elles pourraient contribuer aux efforts de
modélisation pouvant éclairer la gestion des tourbières et les futures recherches scientifiques.

Introduction as indicated by poorly drained soils, hydrophytic


vegetation and various kinds of biological activity
A definition of wetlands that is widely accepted in which are adapted to a wet environment.” This working
Canada was given by the National Wetlands Working group further stated that wetlands can be “organic” or
Group (1988): “. . . land that is saturated with water “mineral”. This review is limited to organic wetlands
long enough to promote wetland or aquatic processes which are often referred to as “peatlands”. To further

Richard A. Bourbonniere1

1
Water Science and Technology Directorate, Aquatic Ecosystem Impacts Research Division, Environment
Canada, Burlington, ON L7R 4A6

Submitted September 2009; accepted September 2009. Written comments on this paper will be accepted until
June 2010.

Canadian Water Resources Journal Vol. 34(4): 393–414 (2009) © 2009 Canadian Water Resources Association
Revue canadienne des ressources hydriques
394 Canadian Water Resources Journal/Revue canadienne des ressources hydriques

refine the definition peatlands contain more than and classification of peatlands. A global and up-to-date
40 cm of organic soil or “peat” (National Wetlands view of all aspects of wetlands, including peatlands,
Working Group, 1997). Throughout this review the can be found in the encyclopaedic text by Mitsch and
term “water chemistry” most often is used to describe Gosselink (2000). Reference will be made to these
the chemical content of peatland waters as most studies reviews, but little repetition of them will be found in
are descriptive in nature. Some studies involve chemical the present paper; instead the reader is encouraged to
reactions, either directly or implied, and “chemistry” seek out the original works.
occurs in the more formal sense.
To understand their water chemistry one must
have some familiarity with the hydrological, botanical, Peatlands – Bogs and Fens
geological, and ecological aspects of peatlands,
their formation and their susceptibility to change. Often the term “peatlands” includes both bogs and
Fortunately several reviews which cover these aspects fens. Here we will not discuss minerogenous systems
in depth pre-date this one (see forthcoming section) such as marshes, swamps, and pond wetlands. Fens
so this review can focus primarily on the water are included because they are intermediate containing
chemistry of bogs and fens and will concentrate on the many of the characteristics of bogs, but are situated in
past two decades. Although much of this review will the landscape such that they experience some influence
be presented in a Canadian context, many examples of surface or groundwater flows (National Wetlands
from other boreal systems (e.g., northern Europe) will Working Group, 1997). Bogs and fens are common
be used. In many ways the water chemistry of bogs in the cool boreal zones where excess moisture is
and fens is a necessarily descriptive endeavour. To be abundant. Bogs are isolated from groundwater; they
more useful this review will attempt to link peatland get all of their water and nutrition from atmospheric
water chemistry changes to natural processes and sources so they are called ombrotrophic. Because of
anthropogenic impacts so that readers can be guided by their interaction with groundwater fens can contain
example to assist interpretation of their own results. higher concentrations of mineral-sourced components.
The scientific understanding of peatlands and True fens are thus minerotrophic (rich fens) and those
their development has grown over a long period with less groundwater contact are transitional and are
of study by botanists and ecologists as well as called “poor fens” (Mitsch and Gosselink, 2000).
geochemists and geologists and other related sciences.
It is beyond the scope of this review to delve in detail
into the botanical and geological aspects of peatland Importance of Hydrology to the Ecology of
development but fortunately several excellent and Peatlands
comprehensive reviews have already appeared. For a
botanical and ecological view the book by Moore and When classified according to hydrologic criteria
Bellamy (1974) is highly recommended, especially wetlands are either ombrogenous or minerogenous
since it synthesizes much of the early work and thus systems. Ombrogenous systems receive water solely
provides an historical context. In the same way the from precipitation, are isolated from surface or
inorganic geochemistry of peatlands and especially the groundwater flows and occur only where precipitation
water they contain was comprehensively reviewed by exceeds evaporation during the growing season
Shotyk (1988). This work also delved rigorously and (National Wetlands Working Group, 1997). A positive
critically into much of the historical understanding water balance and an appropriate annual distribution
and examined peatland chemistry. A more recent of precipitation, e.g., the absence of hot dry summers,
review by Wheeler and Proctor (2000) set out to clarify are hydrological conditions necessary for a peatland
some of the ecological understanding and terminology to develop. This means precipitation is greater than
in current use by workers with diverse interests in evaporation on an annual basis; evaporation ranging
peatlands. Their review is narrower in scope than the from 50–70% of precipitation (Mitsch and Gosselink,
two mentioned above but no less rigorous and useful. 2000). Bogs are the only true ombrogenous wetlands
It delves into our understanding of how ecological (National Wetlands Working Group, 1997).
gradients are primarily responsible for the character

© 2009 Canadian Water Resources Association


Bourbonniere 395

The water table of peatlands often varies seasonally Peat, Organic Soils and Water
according to climate variations and because of changes
in evapotranspiration. This seasonality generally means On a dry weight basis organic soils contain at least
lower water table in the summer and results in the 20% but often more than 35% organic matter and
vertical structure of peatlands. The upper portion is peat is nearly all organic content. When compared
termed the acrotelm and its depth is equal to the level of to mineral soils, organic soils and peat exhibit: (i)
the lowest annual water table. The living plants occupy lower bulk density and higher porosity, (ii) widely
the top few centimetres of the acrotelm, below which ranging but generally lower hydraulic conductivity,
are degraded plant remains. Below the acrotelm the (iii) nutrients and minerals that are bound to organic
peat column is called the catotelm, which is perpetually matter thus not bioavailable, and (iv) higher cation
saturated so it becomes anoxic. The acrotelm is thus the exchange capacity. These properties broadly determine
oxic portion of the peat. the biogeochemistry of peatlands and the character of
the water contained in them (Mitsch and Gosselink,
2000).
Successional Development of Peatlands Water is also an important agent in the overall
biogeochemical cycling in peatlands. Chemical
Peatlands can be seen as transitional between aquatic components in peatland waters are more mobile and
and forest ecosystems, and such a view constitutes the can be more reactive while in solution than when
autogenic process of succession whereby the transition they are associated with the solid phases. Water
happens due to environmental modification by the biota chemistry is also linked to the hydrology of peatlands.
in place. Allogenic succession has also been inferred For instance the diffusion rate of oxygen in aqueous
from paleoecological studies where evidence suggests media, like the water filled pore spaces in saturated
that factors external to the ecosystem undergoing peat, is 10-4 times that in air filled pore spaces. This
succession, such as the rise of the regional water table, leads to the consumption of dissolved oxygen, resulting
were driving development (Mitsch and Gosselink, in its depletion (hypoxia and anoxia) and with time
2000). a sequence of other redox transformation occur, e.g.,
A comprehensive discussion of the development nitrate and sulphate reduction (Mitsch and Gosselink,
of bogs and fens is beyond the scope of this review, 2000).
but descriptions of the development of two peatlands
in Elk Island National Park (near Edmondton, AB)
by Nicholson and Vitt (1994) provide by example two Water Chemistry of Natural Peatlands
common autogenic successional paths leading in one
case to a bog and in another to a rich fen. The succession Ecologists are often concerned first with the species
was discerned by macrofossil analysis of cores collected composition of the biota in peatlands. To a large extent
from the centre of each peatland. Both peatlands the biota that exist in any peatland is dependent upon
began as aquatic systems but the sediment phase at the the chemical composition of the aqueous phase, and
bottom of the core from the rich fen contains mineral conversely, the biota present can influence the water
grains and fossils that indicate a permanent water chemistry. This section will review some of the patterns
body including: (i) abundant seeds of aquatic plants, and controls that have been observed on bogs and
(ii) low abundance of sedge and bryophyte fossils, fens over the past two decades. The review includes
and (iii) presence of ostracod and gastropod shells. inorganic and organic constituents whether they are
In contrast the sediments at the bottom of the bog major or minor components.
core contain more bryophyte fragments and no fossil
shells suggesting a less permanent aquatic phase and
the transition from aquatic through rich fen to bog Major Constituents
occurred more quickly (Nicholson and Vitt, 1994).
Acidity and pH
Shotyk (1988) reviews the differences between
“acidity” or “total acidity” and “pH”; and provides

© 2009 Canadian Water Resources Association


396 Canadian Water Resources Journal/Revue canadienne des ressources hydriques

some methodological cautions. Through examples in Major Cations


the historical and recent literature he shows correct In addition to the division of bogs and fens by pH
and incorrect observations regarding acidity and pH and bicarbonate content the higher concentration of
and changes thereof. The pH of water is a measure of Ca+2 ions found in fen pore waters is often used as a
hydrogen ions (H+) dissolved in it, formally the activity distinguishing parameter because, like bicarbonate,
of H+ and is expressed as: pH = -log [H+], where [H+] Ca+2 content is related to contact with mineral soils
is a concentration in moles per L. The concentration of (Gorham et al., 1985; Vitt et al., 1995; Wheeler and
H+ is often confused with the total amount of available Proctor, 2000). Three other base cations are important
H+ in the peat and/or water, including those that components in peatland waters: Mg+2, Na+ and K+.These
are still bound to ligands (i.e., total acidity) which is four elements constitute the bulk of the cations found
determined by titration with base (e.g., NaOH) and in bog and fen waters in the northern hemisphere.
can be precisely described as the base neutralizing Shallow (acrotelm) pore water concentrations of these
capacity (Kramer, 1982). four major cations, published over the past two decades,
Bimodality of the pH found in peatlands has been were separated into four peatland types according to
noted by several authors (Gorham et al., 1985 – citing the description of the original authors, or using pH
earlier work of Gorham and co-workers; Shotyk, ranges guided by the results of Vitt et al. (1995): Bog
1988; Sjörs and Gunnarsson, 2002; Urban et al., 1989; (pH ≤4.3); Poor Fen (pH 4.3-5.5); Moderately Rich
Wheeler and Proctor, 2000). One mode with pH values Fen (pH 5.5-6.2); and Rich Fen (pH 6.2-7.0). Table 1
ranging from 3.5-4.5 is characteristic of bogs, and the lists the ranges of base cation concentrations found in
fen mode ranges from 5.5-7.5 (Urban et al., 1989). The peatlands in Canada and the northern United States
explanation of bimodality is related to the differing and for northern and central Europe. There is a lot of
buffer systems operating in each mode. Bog water pH overlap between the geographic regions and also among
is controlled by the carboxylic acid (COOH) content the wide concentration ranges determined for fens.
of dissolved organic matter (DOM) originating in the Differences among peatland types for all base cations
peat, and to a variable extent by ion exchange with are less pronounced for the minimum values than
living plants, e.g., bryophytes (Sjörs and Gunnarsson, for the maximum values. It is instructive to plot the
2002). Bicarbonate (HCO3-) and carbonate (CO3-2) Northern Hemisphere maxima for these base cations
originating in the groundwater buffers the pH of fens according to peatland type (Figure 1). Bogs and poor
(Shotyk, 1988; Sjörs and Gunnarsson, 2002; Urban et fens are similar in terms of the maximum base cation
al., 1989; Wheeler and Proctor, 2000). Vitt et al. (1995) concentrations found in them, and the progressively
show how pH and bicarbonate alkalinity are the most increasing maximum concentrations of base cations in
critical chemical components to use when studying the moderately rich and rich fens are considerable and
peatlands along the bog-fen gradient, correlating well exponential in nature.
with vegetation-based classes. Pore waters from rich fens showed base cation
Many authors commented that the pH of concentrations three to five times higher than
peatland waters is the most stable chemical parameter poor fens (Bendell-Young and Pick, 1997) and in
measured, i.e., it shows less temporal variability in another study Bendell-Young (2003) notes that the
seasonal studies (e.g., Hájková and Hájek, 2004; Vitt Ca:Mg ratio of peatland pore waters from poor
et al., 1995). However recent work by Tahvanainen and moderately rich fens approached that of rainwater
and Tuomaala (2003) showed that water sampling rather than bedrock and the reverse was true for rich
location, sampling method, sample treatment and fens. Based on major ion chemistry Malmer et al.
holding time, and even time of day can have a large (1992) concluded that the difference between poor
effect on the pH determined. Their work confirmed and rich fens is greater than between bog and poor
some of the cautions posed previously (see Shotyk, fen (see also Figures 1 and 2). Similarly Bendell-
1988) and they recommend sampling by pre-inserted Young (2003) used principal components analysis
wells with minimal vacuum, storage with no headspace of Ca+2, Mg+2 and alkalinity to show that surface
and measuring pH within two hours. water in poor fens and moderately rich fens could be
distinguished from that in rich fens.

© 2009 Canadian Water Resources Association


Bourbonniere 397

450

400
Calcium
350 Magnesium
Sodium
300 Potassium

250

200

150

100

50

0
Bog Poor Fen Mod. Rich Rich Fen
Fen

Figure 1. Maximum concentrations in mg/L of base cations determined in surface water or acrotelmic pore
water from Northern Hemisphere bogs and fens in the past two decades. Peatland classifications by original
authors (see Table 1) or guided by Vitt et al. (1995).

Table 1. Concentrations of major (base) cations reported for natural waters from bogs and fens in the Northern
Hemisphere. Values are ranges, expressed in mg/L, determined from those reported by the cited sources
which cover the past two decades of peatland research.

Cation Bogs Poor Fens Mod. Rich Fens Rich Fens


a
Canada and Northern United States:
Ca+2 0.06-6.8 0.10-16 0.80-179 1.4-428
Mg+2 0.04-2.8 0.16-4.6 0.10-29 0.27-47
Na+ 0.03-16.5 0.03-4.9 0.05-76 0.80-36
K+ 0.02-1.4 0.02-3.4 0.05-11 0.40-23
b
Northern and Central Europe:
Ca+2 0.10-5.0 3.1-10.9 12.9-24 29-408
Mg+2 0.09-2.6 0.41-0.95 1.0-17 1.1-125
Na+ 0.40-21 0.05-2.6 2.3-23 2.9-132
K+ 0.05-2.1 0.30-3.3 1.0-21 0.19-185

Sources:
a
Belland and Vitt (1995); Bendell-Young (2003); Bendell-Young and Pick (1997); Cooper and Andrus (1994); Foster et al. (1988);
Glaser et al. (2004); Godwin et al. (2002); Gorham et al. (1985); Malmer et al. (1992); Reeve et al. (1996); Shotyk et al. (1992);
Szumigalski and Bayley (1997); Vitt and Bayley (1984); Vitt et al. (1995); Weider (1985)
b
Bragazza and Gerdol (1999); Bragazza and Gerdol (2002); Bragazza et al. (2005); Gabka and Lamentowicz (2008); Gorham et al. (1985);
Hájek et al. (2002); Navratilova et al. (2006); Proctor (2003); Tahvanainen (2004); Verhoven and Arts (1992); Wassen and Joosten (1996)

© 2009 Canadian Water Resources Association


398 Canadian Water Resources Journal/Revue canadienne des ressources hydriques

900

800 Bicarbonate
Chloride
700 Sulphate
600

500

400

300

200

100

0
Bog Poor Fen Mod. Rich Rich Fen
Fen

Figure 2. Maximum concentrations in mg/L of major anions determined in surface water or acrotelmic pore
water from Northern Hemisphere bogs and fens in the past two decades. Peatland classifications by original
authors (see Table 2) or guided by Vitt et al. (1995).

Major Anions the upper pH limit for rich fens under consideration
here is pH=7.0. Wetzel (1983), citing previous workers,
Fewer data are reported for the major anions HCO3-, Cl- shows graphically and in tabular form the distribution
and SO4-2 than for the major cations. Reported pore of “Total CO2” (H2CO3 + HCO3- + CO3-2) in pure
water (acrotelm) concentrations were separated by water. At pH=7 HCO3- makes up 79.2% and 20.8%
peatland type and geographic location as was done for is H2CO3; CO3-2 is nil, so for this review we need not
the base cations and the resulting ranges are shown be concerned with it. It has often been suggested that
in Table 2. Here again there is much overlap and the H2CO3 is a significant contributor to the acidity of
overall differences between bog and poor fen are far peatland waters as suggested by the pure water example
less than among those two types and the richer fens. above. However work by Urban et al. (1989) proved
Confirmation of the ability of HCO3- to distinguish the often repeated statement that organic acids are
mineral soil contact in much the same way as Ca+2 responsible for the acidity of bogs and fens (see below).
(Vitt et al., 1995), is shown in Figure 2. Also shown in The maximum concentrations of Ca+2 and HCO3- are
Figure 2 where the maxima are plotted, it is clear that inversely related to the concentration ranges of H+ used
lower Cl- contents are typical of bog, poor fen, and rich for defining the peatland type (Figure 3), underscoring
fen pore waters with moderately rich fens somewhat the utility of these three water constituents as
higher. Only rich fen pore waters are high in SO4-2 distinguishing properties related to contact with
content. mineral soil. The majority of peatland studies involve
It should be noted that the concentration of some measurement of pH and the four base cations in
HCO3- is most often estimated by alkalinity titration. Table 1, and to lesser extent the three major anions in
Such titrations are commonly used in aquatic chemistry Table 2. Many of these studies also involved surveys
and several texts cover this material well and go way of plant species present and statistical analyses were
beyond the scope required for this review (Morel, 1983; often used to merge chemical and floristic components
Stumm and Morgan, 1981; Wetzel, 1983). Recall that resulting in ecological classification of peatlands (e.g.,

© 2009 Canadian Water Resources Association


Bourbonniere 399

Table 2. Concentrations of major anions reported for natural waters from bogs and fens in the Northern
Hemisphere. Values are ranges, expressed in mg/L, determined from those reported by the cited sources
which cover the past two decades of peatland research.

Anion Bogs Poor Fens Mod. Rich Fens Rich Fens


a
Canada and Northern United States:
HCO3- 0.001-12 0.03-32 0.55-688 6.3-805
Cl- 0.07-29 0.10-7.0 0.10-119 2.1-27
SO4 -2
0.03-11 0.03-1.9 0.03-29 1.4-620
b
Northern and Central Europe:
HCO3- 0.001-1.2 30-43 19.7 -135 nd - nd
Cl-
0.90-35 1.3-11 11.5 -62 0.03-15.6
SO4-2 0.03-27 6.3-37 9.5 -40 5.7-78

nd = no data available
Sources:
a
Bendell-Young (2003); Bendell-Young and Pick (1997); Glaser et al. (2004); Godwin et al. (2002); Gorham et al. (1985); Malmer
et al. (1992); Reeve et al. (1996); Szumigalski and Bayley (1997); Vitt et al. (1995); Vitt and Bayley (1984); Wieder (1985)
b
Bragazza and Gerdol, (1999); Bragazza and Gerdol (2002); Bragazza et al. (2005); Gabka and Lamentowicz (2008); Gorham et
al. (1985); Hájek et al. (2002); Navratilova et al. (2006); Proctor (2003); Tahvanainen (2004); Verhoven and Arts (1992); Wassen
and Joosten (1996)

Figure 3. Inverse relationships between H+ concentration ranges (moles/L) characteristic of bogs and fens (Vitt
et al., 1995) and the maximum concentrations of Ca+2 and HCO3- from Tables 1 and 2 plotted as: 1/mg L-1.

© 2009 Canadian Water Resources Association


400 Canadian Water Resources Journal/Revue canadienne des ressources hydriques

Table 3. Concentrations of nitrate, ammonium, and phosphate reported for natural waters from bogs and fens
in the Northern Hemisphere. Values are ranges, expressed in mg/L, determined from those reported by the
cited sources which cover the past two decades of peatland research.

Nutrient Bogs Poor Fens Mod. Rich Fens Rich Fens


a
Canada and Northern United States:
NO3- 0.02-1.7 0.01-10 0.01-24 0.01-28
NH4+ 0.01-2.3 0.01-8.0 0.01-2.9 0.01-0.07
PO4 -3
0.01-0.05 0.08-2.0 0.01-11 0.01-12
b
Northern and Central Europe:
NO3- 0.02-2.4 0.02-2.4 0.01 -9.3 0.01-23
NH4 +
0.01-0.13 0.01-1.67 0.01-13 0.01-0.09
PO4-3 0.01-0.1 0.01-0.37 0.01 -5.4 0.01-18

Sources:
a
Glaser et al. (2004); Godwin et al. (2002); Kingsbury and Moore (1989); Szumigalski and Bayley (1997); Vitt and Bayley (1984);
Vitt et al. (1995); Wieder (1985)
b
Bragazza and Gerdol (1999); Bragazza and Gerdol (2002); Bragazza et al. (2005); Gabka and Lamentowicz (2008); Hájek et al.
(2002); Navratilova et al. (2006); Proctor (2003); Tahvanainen (2004); Verhoven and Arts (1992); Wassen and Joosten (1996)

30
Nitrate
25 Ammonium
Phosphate
20

15

10

0
Bog Poor Fen Mod. Rich Rich Fen
Fen

Figure 4. Maximum concentrations in mg/L of N and P species determined in surface water or acrotelmic pore
water from Northern Hemisphere bogs and fens in the past two decades. Peatland classifications by original
authors (see Table 3) or guided by Vitt et al. (1995).

© 2009 Canadian Water Resources Association


Bourbonniere 401

Belland and Vitt, 1995; Glaser et al., 2004; Hájková Metals


and Hájek, 2003). In this way defining the bog-fen Bendell-Young (2003) showed that pore water
gradient using pH, reduced conductivity, HCO3- and concentrations of Mn and Fe come from groundwater
Ca+2 became well accepted (Vitt et al., 1995). Sjörs and exchange at the bottom of the peat profile in a poor
Gunnarsson (2002) stressed that the bog-fen gradient fen. Similarly Bragazza et al. (2005) suggested that
in water chemistry should be considered continuous pore water concentrations of metals in bogs mirror
with considerable overlap rather than one with discrete the composition of the groundwater particularly at
boundaries. the margins. Al and Fe were found to be enriched in
Proctor (2003) used five years of water chemistry bogs relative to fens by Vitt et al. (1995) and this was
data from an ombrotrophic system and was able to attributed to chelation by DOM. Helmer et al. (1990)
learn much about the sources and seasonal fluctuations showed that interaction with DOM was important to
of the major constituents—cations and anions with model the speciation of Al in bog waters.
factor analysis. As well ratios of cations and anions were Wetlands were shown to yield methylmercury
useful in detecting seasonal changes, e.g., evaporation (MeHg) from 26-79 times higher than from upland
and biological activity in the long-term data that are portions of boreal catchments (St. Louis et al., 2004).
often lost in the variability of short-term data. Cation Sulphate was shown to stimulate production of MeHg
and anion results from this study suggested that ion by sulphate reducing bacteria which can have important
exchange between the peat and the pore water can be consequences wherever SO4-2 concentrations are
responsible for short-term changes in water chemistry. increased by atmospheric or other sources (Branfireun
The equivalent sum of the base cations, other et al., 1999). Studying whether microtopography affects
cations determined and H+ is often compared to the the chemical constituents of bog and fen pore water,
equivalent sum of the major anions and other anions Branfireun (2004) found MeHg at the water table to
determined in an attempt to balance the ionic charge. be 3.5 times higher in shallow hollows compared to
In peatlands this always results in an excess of cations deep hollows, and the sequence of these concentrations
or an “anion deficit” which is “balanced” by organic among four topographic locations was shallow hollows
anions in the DOM (see below and Proctor, 2003; > hummock > lawn > deep hollow. At 25 cm below
Urban et al., 1989). the water table there were no significant differences
in MeHg concentration. More detailed profiles of
pore water chemistry (pH, SO4-2, DOC, and MeHg)
Minor Constituents along the microtopographic gradient suggested that
hydrology and biogeochemistry were also important
Nutrients controlling factors in the complex relationship.
Table 3 lists the reported ranges for acrotelm pore water Because of its ombrotrophy, bog vegetation is used
NO3-, NH4+ and PO4-3 according to peatland type. to monitor changes in atmospheric deposition of metals
Even the maximum values are considered low for all (Martinez-Cortizas et al., 1999; Shotyk et al., 1998).
peatland types. The maximum NO3- and PO4-3 values Bragazza (2006) investigated whether bog surface
show a progressive increase along the bog-fen gradient water could be used as a monitoring medium instead
and those for NH4+ appear to be highest for the poor of vegetation and found a significant relationship
fen (Figure 4). These data illustrate the long held notion between bog water and rainwater concentrations of
that peatland pore waters are nutrient poor and thus may several trace metals (Cd, Cr, Cu, Zn, and Pb). At his
be nutrient sinks (Mitsch and Gosselink, 2000). Vitt et study site in the Italian Alps he noted the same patterns
al. (1995) showed that nutrient elements (N and P) in bog water and vegetation concentrations but mosses
correlate somewhat with peatland types and exhibited exhibited higher concentrations.
strong seasonal variations. Gabka and Lamentowicz
(2008) studied peatlands along a bog to moderately rich
fen gradient and also found significant relationships Dissolved Organic Matter in Peatlands
between nutrients to peatland type. Proctor (2003)
showed that NO3- tends to be highest during the colder Dissolved organic matter (DOM) is produced in peat
months when biological activity is low. and soils by the microbial degradation of plant material

© 2009 Canadian Water Resources Association


402 Canadian Water Resources Journal/Revue canadienne des ressources hydriques

and subsequent dissolution in the pore waters (Moore Thurman, 1985; Thurman and Malcolm, 1981 among
and Dalva, 2001) and related to temperature, roughly others). Recent attempts to unravel the nature of DOM
doubling with each 10ºC increase (Moore et al., have moved away from fractionation and returned to
2008). Increased primary production can also result in viewing DOM as a whole, utilizing modern sensitive
increased DOM concentrations (Freeman et al., 2004). techniques that help to understand the biogeochemical
The concentration of DOM is generally expressed function of DOM or its component parts.
as the dissolved organic carbon (DOC) content of
the water. Since DOM is approximately 55% C, the Organic Acidity and DOM
approximate conversion is: The acidity of DOM is one property that interests
ecologists and chemists alike because of its relationship
DOM = 1.8 × DOC. to the pH of peatlands and their floristic composition,
and to the interaction of DOM with metals.The “organic
The range of DOC concentration typical for bog acidity” of DOM and its fractions is determined most
associated waters is illustrated by the results obtained often by base titration to pH=7.0 (e.g., Kramer, 1982;
from Mer Bleue in Ontario: 30-60 mgC/L (Fraser et McKnight et al., 1985; Oliver et al., 1983; Urban et al.
al., 2001). The natural leaching processes that occur 1989). Organic acidity is often expressed as charge or
in peatlands increase the DOC concentration, and carboxylic density (µeq COOH/ mg DOC) and values
the concentration of DOC exported from peatlands ranging from 7 – 24 were reported by Urban et al.
is generally greater than that from precipitation and (1989). Evidence that FA and HA fractions of DOM
upland sources (Vasander and Kettunen, 2006). are polyfunctional, made up of a complex mixture of
Chemically DOM is a complex mixture of organic monomers and polymers, is the monotonous shape of
compounds varying widely in molecular weight from base titration curves. This suggests that the titratable
small molecules like organic acids (some volatile) and H+ in DOM exist in a multitude of environments;
sugars to intermediate polymers like hemicellulose that acids with dissociation constants characteristic of
and large polycondensates usually called humic matter. strong and weak acids co-exist in the mixture. Urban
Some early workers isolated and identified particular et al. (1989) illustrated this using multiple samples of
organic constituents from peat (see Shotyk (1988) bog water and two prominent modelling techniques.
for some fascinating examples), but in the past two They determined acid dissociation constants expressed
decades aquatic and soil chemists and ecologists have as pKa ranging from 3.5 to 7.4.
viewed DOM in many of the same ways they did soil The organic acidity, or total carboxylic acid
organic matter or DOM in lakes, stressing its complex concentration determined by base titration of natural
(humic) nature. The study of humic matter originated peatland waters is often equal to the anion deficit
in soil chemistry and has a 200 year history. Humic calculated by subtracting the total anion concentration
matter is isolated from soil, sediment and peat by from that of the total cations determined by a variety
various methods and the resulting fractions - fulvic of methods. This correspondence between two
acid (FA), humic acid (HA) and humin (HU) – have independent measures, is taken as evidence that the
been studied extensively for their physical and chemical [H+] in natural peatland waters is provided mostly
properties, reactivity in the environment, relationship by protons released from DOM. A tight relationship
to soil fertility as well as theories of their formation, between anion deficit and DOC concentration
All of this and more has been extensively reviewed was shown by Gorham et al. (1985) for their North
– see: Kononova (1966), Poveledo and Golterman American bog transect and by Marin et al. (1990) in a
(1973), Schnitzer and Khan (1972) and Stevenson Wisconsin poor fen. The same authors also found good
1982) among many others. relationships between [H+] and DOC concentration.
Aquatic chemists began about 40 years ago to A further indication that DOM was responsible for
consider the DOM in natural waters as “aquatic humic low pH in bog and poor fen waters was provided by
matter” and developed isolation techniques so that degrading it photochemically and noting the resulting
many of the soil scientists’ methods for studying FA increase in pH (Gorham et al., 1985; Marin et al.,
and HA could be applied to those fractions of DOM 1990).
(Aiken et al., 1979, 1985, 1992; Leenheer, 1981;

© 2009 Canadian Water Resources Association


Bourbonniere 403

Characterization of Peatland DOM suggesting that over winter degradation and early
The character of DOM in peatlands was investigated summer production affect DOM character and that
utilizing a comprehensive resin based fractionation more HA character is dominant in the late summer
technique initially used by Leenheer (1981) which (Moore, 1987). Zaccone et al. (2009) characterized bog
separated the DOM into several hydrophobic and pore waters using E4/E6 and, though highly variable,
hydrophilic acid, base and neutral sub-fractions. found higher ratios in the upper portions of a one
Interstitial waters from bogs in Nova Scotia contained meter bog core (FA like) than at the bottom portions
59-64% of their DOC in the acidic sub-fractions (HA like).
of DOM: HA, hydrophobic acid (HPOA) and Specific UV absorbance (SUVA) has also found
hydrophilic acid (HPIA) (Bourbonniere, 1989) and utility as a measure of the aromatic character of DOM
a continental bog contained 82% (Moore and Matos, in soils and has sometimes been used in peatland
1999). Moore et al. (2003) found 85-93% of the DOC waters. Frequently measured at 254 or 280 nm,
in these acidic fractions in shallow bog water from SUVA is the ratio of absorbance per m of path length
the Experimental Lakes Area in Ontario. The sums normalized to DOC in mg/L [units: L mgC-1 m-1];
of these operationally defined “acid” fractions are high higher SUVA indicates higher aromatic character or
enough to account for the organic acidity and anion aromaticity. Chin et al. (1994) showed that SUVA
deficit of peatland waters discussed above. These was related to total aromaticity and weight average
same workers also found that deeper bog waters were molecular weight determined by high pressure size
usually higher in HA content, reflecting the diagenetic exclusion chromatography. Sachse et al. (2001) found
modification that occurs as peat accumulates and is higher aromaticity in fen pore waters than in adjacent
buried (Bourbonniere, 1989; Moore et al., 2003). The ponds.
higher summer HA concentrations in bog drainage Fluorescence spectra have been widely used to
water and streams that contain bogs in their watersheds characterize the DOM in many natural water samples
suggests that as water table drops seasonally, older HA (cf. McKnight et al., 2001) including DOM from soils
enriched water is a greater proportion of the drainage and peatlands. Techniques vary in terms of the types
than in other seasons. and amount of pre-treatment (e.g., isolation on resins,
DOM can also be characterized by size exclusion see above), but Kalbitz et al. (2000) showed that UV
chromatography into fractions defined by molecular absorbance at 285 nm and a humification index (HIX)
weight. Sachse et al. (2001) studied the interstitial water obtained by synchronous fluorescence spectroscopy
in a poor fen and reported depth-averaged results: 72% were typically lower by about 15% when applied to
of the DOC in their highest molecular weight humic DOM without any sample preparation compared to
substances fraction, 3% in the intermediate molecular resin isolated FA. Humification indices proposed by
weight polysaccharides fraction and 4% in the low Kalbitz et al. (2001) are ratios of emission intensity
molecular weight acids fraction. in synchronous fluorescence spectra (e.g., 390/355
Spectroscopy in the ultraviolet and visible spectral nm, 400/360 nm, 470/360 nm, 470/400 nm) and are
ranges has a long history in the characterization DOM similar to the fluorescence index proposed for aquatic
from peatlands. Measures that came from the study of FA by McKnight et al. (2001). The ratio of emission
soil organic matter like the E4/E6 ratio (absorbance at intensities 450/500 nm, at 370 nm excitation showed
465 nm / absorbance at 665 nm) can indicate the fulvic a strong negative linear relationship to aromaticity
vs. humic character of DOM: HA 3-5 and FA >8 which derived from 13C NMR spectroscopy for aquatic FA
in turn indicate molecular weight and aromaticity (see isolated from a broad range of water samples with
Chen et al., 1977). Moore (1987) (using absorbances microbial DOM and terrestrial DOM as end members.
at 400 and 600 nm) found that peatland waters with The fluorescence index is suggested for use between
higher E4/E6 (13.2) also contained a greater proportion 1.9 (microbial) and 1.2 (humic), but should be used in
of low molecular weight components, and those waters conjunction with SUVA (McKnight et al., 2001).
with higher molecular weight constituents gave lower Fraser et al. (2001) used the fluorescence index
E4/E6-ratio (5.5-7.3). These subarctic peatland waters of McKnight et al. (2001) on pore water from Mer
also exhibited marked seasonal differences in E4/E6- Bleue and found increasing values from 1.1 to 1.6 with
higher earlier in the growing season than at the end increasing depth to 4.5 m in the peatland, indicating

© 2009 Canadian Water Resources Association


404 Canadian Water Resources Journal/Revue canadienne des ressources hydriques

a decrease in aromaticity of the DOM with depth. This can be as profound as complete degradation
Using the HIX (390/355 nm) Zaccone et al. (2009) by drainage, utilization or even urbanization. More
also showed decreasing aromatic character of bog moderate and much slower changes in water chemistry
pore water DOM with depth. HIX (470/400 nm) can occur because of direct and indirect anthropogenic
and SUVA (280 nm) in bog waters decreased after influences that gradually modify the relationships that
removal of HA, suggesting a more humified character exist among peatland vegetation, groundwater and the
of HA than the bulk DOM (Glatzel et al., 2003). atmosphere.
Taken together these results suggest that some of the
HA produced by humification of peat at depth may Disturbances by Atmospheric Deposition
precipitate from the pore water resulting in decreased Atmospheric components are potentially important
aromatic character of DOM at depth. contaminants of peatlands, particularly ombrotrophic
ones (Gorham et al., 1985; Shotyk, 1988). Airborne
Role of DOM in the Peatland Carbon Budget SO4-2 and NO3- from anthropogenic and natural
As recent interest in the global carbon budget focussed sources might decrease the pH of peatlands and if
a lot of peatland research on the cycling of carbon they are not assimilated or otherwise utilized in
within them, an appreciation of the role that DOM biogeochemical reactions, they may replace a portion
plays has emerged (Moore et al., 1998; Vasander and of the natural organic acidity. Over recent decades
Kettunen, 2006). Hydrology is usually the controlling increased atmospheric deposition of SO4-2 and NO3-
factor determining DOM export from peatlands in northern Sweden, though only slight compared to
(Fraser et al., 2001; Freeman et al., 2001; Hinton et other locations, impacted the pH and Ca+2 content
al., 1997; Koprivjnak and Moore, 1992) with greater of bog waters. At all sites studied Ca+2 increased over
export generally associated with higher discharge. 50 years and pH decreased for the fen sites while
Several studies have quantified the export of DOM remaining largely unchanged at the bog sites (Sjörs
(as DOC) from northern bogs and fens during the and Gunnarsson, 2002). The water chemistry of bogs,
growing season which ranged from 0.8–29.4 gC/m2, fens and associated pools in central Ontario were
the highest values from bogs on the Precambrian analyzed for 27 chemical variables and the results
Shield and the lowest from subarctic fens (Moore, of factor analysis yielded three main factors that
1987; Koprivnjak and Moore, 1992; Moore et al., together explained 81% of the variance (Blancher and
2003). There is some uncertainty in estimating annual McNicol, 1987). The first factor (39% of variance)
DOC exports, particularly if the estimation is based on was attributed to DOM influence and metals like
three to four month growing season data (Koprivnjak Al, Fe, other trace metals, colour, and anion deficit
and Moore, 1992), but where annual budgets were were positively weighted on this factor. Factor 2 (28%
determined they ranged from 1.5–40 gC/m2 (Schiff et of the variance) was heavily weighted on the base
al., 1998; Urban et al., 1989; Fraser et al., 2001). The cations, pH and alkalinity and was interpreted to
magnitude of annual DOC export from peatlands represent mineral influence or buffering capacity. The
is about 2-8% of the annual CO2 exchange by soil third factor (14% of the variance) are heavily loaded
respiration and photosynthesis and about twice the on SO4-2, Mn and Ni and negatively loaded on pH;
emissions of CH4 (Moore et al., 1998). Waddington all indicating local sources of atmospheric deposition.
and Roulet (1997) showed that movement of dissolved Proctor and Maltby (1998) noted a strong correlation
components (DOC, CH4 and CO2) within a peatland between surface water pH of bogs and rainwater
may have implications for carbon cycling. pH and a slightly less strong correlation with total
deposited [H+] when dry deposition was included.
Similarly non-marine sulphate in wet and total
Water Chemistry of Disturbed Peatlands deposition, a surrogate for air pollution, was correlated
with surface water pH. Such close correlation would
Potential Changes in Water Chemistry be expected in ombrotrophic systems and the authors
suggest that they are likely better buffered against
Much of the recent interest in peatland water chemistry moderate air pollution than more neutral systems like
is related to questions of anthropogenic disturbance. fens (Proctor and Maltby, 1998).

© 2009 Canadian Water Resources Association


Bourbonniere 405

Drainage from acid impacted bogs was studied by higher concentration of all components determined;
Hruska et al. (1996) in the Czech Republic and though chemistry more like a moderately rich fen than a bog
organic acids were similar in character (carboxylic (Wind-Mulder and Vitt, 2000).
density, pKa) to sites in North America and northern DOC export was studied for three years at a
Europe, the calculated anion deficit (OA-) comprised drained and cutover bog in Quebec in conjunction
only 21-32% of the total anion charge because of high with a parallel study of a restored portion of the same
SO4-2 contents from acid deposition. The pH of these bog (Waddington et al., 2008). The drainage occurred
drainage streams is controlled by mineral acids because 28 years prior to the start of the DOC study and
it was highly correlated with SO4-2 and not with OA-. harvesting activity ended 19 years prior; the site was
Laboratory experiments using peat and soils confirmed abandoned for those years. Bog restoration involved
that the increased discharge of DOM recorded at a removing invasive vegetation, ditch blockage, milling,
reservoir was related to the increase in pH which seeding with Sphagnum, fertilization, and protecting
occurred because of reductions in acid deposition to with straw (Rochefort et al., 2003); none of this
the previously highly acidified and disturbed watershed was done to the cutover portion. The study focussed
(Scheithauer et al., 2005). on the comparison for three years and showed that
DOC export from the restored portion was 2-2.5
times less than from the cutover portion reflecting
Changes Brought about by Peatland Development and the decreased discharge. DOC concentrations in
Restoration discharge and surface waters from the restored portion
Peatlands may undergo development for forestry, were consistently higher than those from the cutover
agriculture, mining, urbanization, waste disposal or portion reflecting increased leaching of peat, vegetation
harvesting of peat. All of these uses require drainage and the straw used to protect the Sphagnum diaspores
which constitutes a profound disturbance of the natural during the initial restoration procedure. At a block
functioning of the peatland (i.e., acrotelm destruction). cut operation DOC concentrations increased during
Once a decision is made to drain a peatland for one harvesting and decreased after the initial harvesting.
of these uses, interest turns to the quality, chemical Blocking the drainage ditches to stimulate restoration
composition, and fate of the drained water as it may resulted in the highest DOC concentrations observed
impact upon downstream uses and water bodies. As (Glatzel et al., 2003).
well, in some cases there is interest in recording the Strack et al. (2008) report on the short and long-
retention or loss of certain components (e.g., nutrients) term changes in DOC concentrations in surface and
from the drained peatland. pore water, and export after artificial lowering of the
Lundin and Bergquist (1990) reported on a long- water table in a poor fen complex. DOC concentration
term controlled bog drainage study in a single watershed in pools was increased immediately after drawdown
containing two bogs. Changes in chemical composition and export was also high initially. In subsequent seasons
were monitored in discharge from an upstream control after drawdown, DOC concentrations in surface and
basin and a drained basin as well as small catchments pore waters remained higher than an undrained control
within the drained basin. Most dissolved components fen for up to 11 years. Seasonal DOC export from the
(e.g., base cations, SO4-2, DOC) showed net losses in experimental fen decreased after the first season to
both basins but greater losses occurred in the impacted values that were not significantly different from those
basin. A deviation from this was the retention of N in of a fen drained nine years previously.
both sub-basins which occurred because of a decrease
of export of organic N. Impacts of Adjacent Land-Use
Drainage of a small bog in Ohio appears to have Peatlands adjacent to agricultural areas such as the
increased groundwater inflow and the chemistry of mid continental prairies of North America are
the bog water became more like a moderately rich influenced by soil derived dust with consequential
fen (Miletti et al., 2005). Similarly an Alberta bog increases in Ca+2 concentrations. On their North
that had been drained and harvested was compared American transect Gorham et al. (1985) found a five
to a nearby natural bog. The groundwater chemistry times decrease in bog water Ca+2 concentrations from
and that of the water in the drainage ditches showed Minnesota east to Newfoundland.

© 2009 Canadian Water Resources Association


406 Canadian Water Resources Journal/Revue canadienne des ressources hydriques

Clay et al. (2009) investigated managed burning expressed in studies of carbon cycling in peatlands that
of the uplands in blanket bogs to determine whether are primarily interested in exchange of CO2 and CH4
it had a detrimental effect on the quantity and quality with the atmosphere as a warming climate is expected
of DOM exported. Although an initial increase in to lower water tables in peatlands (Roulet, 2000).
DOM quantity occurred in the weeks after burning, it Some studies discuss the potential changes in water
was short-lived and there was no significant difference chemistry, particularly DOM export that may occur as
between DOM quantity in the year after the burn water tables drop (Moore et al., 1998 among others).
compared to the year before. Twenty years post burn, Simulated summer drought (16-20 weeks in
there was a decrease in DOM colour compared to an duration, followed by re-wetting for 32-36 weeks)
unburned control. decreased the usual summer peaks of DOC and Fe
Blanket bogs in Ireland that were utilized for concentrations, increased the SO4-2 concentrations
silviculture were studied for the effects on water in autumn and decreased pH (Hughes et al., 1997).
quality after clear-felling and reforestation (Cummins Drawdown of a poor fen to a water table position
and Farrell, 2003a; b). Generally higher pH and equal to the anticipated amount that would result from
concentrations of base cations, NH4+, NO3-, DOC warming and higher evapotranspiration, caused an
and alkalinity were found in receiving streams and initial increase in DOC concentrations and discharge.
drains from clear cut areas than from control areas. Subsequently DOC concentrations decreased, though
Cl-, SO4-2, and conductivity showed little difference; they remained higher than an undrained control after
aluminum was complicated by seasonality between three years. DOC discharge also decreased and the level
organic and inorganic forms (Cummins and Farrell, reached in the third year equalled that of a site that was
2003b). Phosphorus showed large increases (Cummins eleven years from drainage (Strack et al., 2008). Using
and Farrell, 2003a) which was expected because of the a combination of DOC discharge measurements and
fertilizers applied to the forest. modelled CO2 and CH4 fluxes Clair et al. (2002)
Highway development near or through peatlands suggest that DOC export from a poor fen might
can affect not only their hydrology but also their water double under a two times CO2 warming scenario.
chemistry. A common problem in northern areas is Peat monoliths in mesocosms and incubation
the effect of road salt. Pugh et al. (1996) showed that studies were used by Blodau et al. (2004) for a
Na+ and Cl- concentrations as well as K+, Ca+2 and comprehensive investigation of the effect of water table
Mg+2 in surface and pore waters of a poor fen decrease on the processing of carbon in bogs. They determined
exponentially from a major Maine highway that had rates of CO2 and CH4 exchange and DOC production
been salted in winter for 30 years. They found evidence under aerobic (lowered water table) and anaerobic
that cation exchange processes have been impacted so (saturated) conditions. Under saturated conditions
that cation ratios also change regularly with distance DOC production dominated the carbon turnover, and
from the highway, suggesting a long-term impact on its production rate did not change when the water
peat chemistry. The bulk of applied road salt is however table was lowered; but soil respiration and microbial
flushed through the system during a few weeks of biomass carbon rates increased. All production rates
snowmelt every spring. In western Massachusetts decreased with depth under saturated conditions
concentrations of Na+ and Cl- from road salt also indicating that fresh organic matter is an important
decreased with distance from a highway (Richburg et perquisite for release of carbon as DOC, or dissolved
al., 2001). In northwestern Indiana a bog was impacted CO2 and CH4.
by runoff from an uncovered salt storage facility such DOC concentrations in rivers and streams have
that pH, alkalinity and base cation levels were all increased (Worrall et al., 2004). Freeman et al. (2004)
elevated (Wilcox, 1986). inferred that CO2 enrichment results in increased
DOC concentration found in outflows from a drought
Impacts of Climate Change experiment by increasing net primary productivity and
The potential impacts of climate change on peatlands thus the leaching of fresh biomass. This explanation
(and the potential impacts of peatlands on climate) follows after showing how three alternate explanations:
are objects of current research and international (i) warming, (ii) increased river discharge, and (iii)
negotiations (IPCC, 2007). Much of the interest is shifting rainfall patterns are unable to account for the

© 2009 Canadian Water Resources Association


Bourbonniere 407

increase in stream DOC concentration (Freeman et al., papers reviewed here are: Clay et al. (2009), Freeman et
2004; Worrall et al., 2004). al. (2004), Hughes et al. (1997), Lundin and Bergquist
(1990), Strack et al. (2008), and Waddington et al.
(2008).
Concluding Remarks and Recommendations Although models are beneficiaries of long and
short-term field studies and manipulations, those
Over the past two decades research on the water studies often cannot provide all the data required for
chemistry of peatlands has strengthened our model building. Field data are integrative in nature so
understanding of the links among botany, hydrology, they do not represent small scale properties or processes
and water chemistry. In addition studies began to look well. When models require information on smaller
at other environmental drivers that can alter water scale processes and on the environmental factors that
chemistry (e.g., climate, atmospheric deposition) and control them, mesocosm and microcosm studies are
opened up a whole new box of questions. These new useful. In this review several papers reported on such
lines of research also brought into the field techniques studies (Blodau et al., 2004; Moore and Dalva, 2001;
and methods that focussed on understanding peatland Moore et al., 2008).
functioning. Thus the view that peatlands function Many of the modern spectroscopic techniques
as separate ecosystems is gradually being replaced (e.g., SUVA, HIX) are useful to characterize and
by the more holistic idea that peatlands function as quantify the aromatic content of peatland DOM.
components of larger ecosystems. What we are lacking are techniques to reliably
Peatland research has a long tradition as a science determine the character and quantity of hydrophilic
where direct field observations are the mainstay of components that are not aromatic in character. An
studies because practice evolved from methods based in indirect indication of the quantity of “non-aromatic”
ecology. Continuation of the field tradition is necessary components can be inferred by a low SUVA value, but
to improve our understanding of the place that peatlands this is of limited utility. Fractionation methods address
have in the overall landscape, learning more about the hydrophilic content and attempt to distinguish
how peatlands interact with the atmosphere and other acids, bases and neutral components (e.g., Aiken et al.,
landscape components upstream, downstream, and at 1992; Bourbonniere, 1989; Leenheer, 1981; Moore et
their margins. In depth understanding of ecosystem al., 2003), but these techniques require much sample
functioning and appreciation for the ecosystem handling and preparation possibly losing volatile
services provided by peatlands is necessary to inform hydrophilic components in the process. All techniques
policy when development threatens to encroach. Field require filtration, so loss of volatiles is common (see
based observations provide the framework from which Shotyk, 1988). Several authors suggested that “easily
to build models that can assist management decisions degradable” organic matter is important in peatland
and inform future scientific endeavours. biogeochemistry (Sachse et al., 2001) including
Long-term studies have the greatest utility in model promoting the production of CO2 (Blodau et al., 2004;
building because they are more likely to encompass Glatzel et al., 2003). It is highly likely that these are
the range of conditions experienced by the peatland. hydrophilic components of DOM.
In this review several papers reported on data that had Finally when the opportunity arises comprehensive
been monitored for a period of years, some examples studies should be undertaken regarding peatland water
are: Lundin and Bergquist (1990), Proctor (2003), chemistry that links to larger ecosystem scale studies (e.g.,
Strack et al. (2008), and Waddington et al. (2008). Glaser et al., 2004). We should mobilize to take advantage
Manipulation studies are another important avenue of of existing infrastructure and conduct water chemistry
research that can inform modelling efforts. These are research at a common site where other environmental
often tied to field studies and have a lot of potential to variables are under study (e.g., FLUXNET Canada sites
predict impacts on real systems but are often difficult to studied by Moore et al., 2008).
design and implement. Sometimes the design involves
compromises that can yield unexpected or ambiguous
results. Examples of manipulation studies reported in

© 2009 Canadian Water Resources Association


408 Canadian Water Resources Journal/Revue canadienne des ressources hydriques

Acknowledgements Blodau, C., N. Basiliko, and T. R. Moore. 2004.


Carbon turnover in peatland mesocosms exposed
I would like to thank Kasey Livingston for her assistance to different water table levels. Biogeochemistry 67:
tracking down hard to find papers. Three anonymous 331-351.
reviewers helped to improve the quality of this paper
and their efforts and insight are appreciated. Bourbonniere, R. A. 1989. Distribution patterns of
dissolved organic matter fractions in natural
waters from eastern Canada. Organic Geochemistry
References 14: 97-107.

Aiken, G. R., D. M. McKnight, R. L. Wershaw, and P. Bragazza, L. 2006. Heavy metals in bog waters: An
MacCarthy (Eds.). 1985. Humic substances in soils, alternative way to assess atmospheric precipitation
sediment, and water. John Wiley & Sons, New quality? Global and Planetary Change 53: 290-298.
York, 692pp.
Bragazza, L. and R. Gerdol. 1999. Hydrology,
Aiken, G. R., D. M. McKnight, K. A. Thorn, and E. M. groundwater chemistry and peat chemistry in
Thurman. 1992. Isolation of hydrophilic organic relation to habitat conditions in a mire on the
acids from water using nonionic macroporous south-eastern Alps of Italy. Plant Ecology 144:
resins. Organic Geochemistry 18: 567-573. 243-256.

Aiken, G. R., E. M. Thurman, R. L. Malcolm, and H. F. Bragazza, L. and R. Gerdol. 2002. Are nutrient
Walton. 1979. Comparison of XAD macroporous availability and acidity-alkalinity gradients related
resins for the concentration of fulvic acid from in Sphagnum-dominated peatlands? Journal of
aqueous solutions. Analytical Chemistry 51: 1799- Vegetation Science 13: 473-482.
1803.
Bragazza, L., H. Rydin, and R. Gerdol. 2005. Multiple
Belland, R. J. and D. H. Vitt. 1995. Bryophyte gradients in mire vegetation: A comparison of a
vegetation patterns along environmental gradients Swedish and an Italian bog. Plant Ecology 177:
in continental bogs. Ecoscience 2: 395-407. 223-236.

Bendell-Young, L. 2003. Peatland interstitial water Branfireun, B. A. 2004. Does microtopography influence
chemistry in relation to that of surface pools along subsurface pore-water chemistry? Implications for
a peatland mineral gradient. Water, Air, and Soil the study of methylmercury in peatlands. Wetlands
Pollution 143: 363-375. 24: 207–211.

Bendell-Young, L. and F. R. Pick. 1997. Base cation Branfireun, B. A., N. T. Roulet, C. A. Kelly, and J. W.
composition of pore water, peat and pool water M. Rudd. 1999. In situ sulphate stimulation of
of fifteen Ontario peatlands: Implications for mercury methylation in a boreal peatland: Toward
peatland acidification. Water, Air, and Soil Pollution a link between acid rain and methylmercury
96: 155-173. contamination in remote environments. Global
Biogeochemical Cycles 13: 743-750.
Blancher, P. J. and D. K. McNicol. 1987. Peatland
water chemistry in central Ontario in relation to Chen, Y., N. Senesi, and M. Schnitzer. 1977.
acid deposition. Water, Air, and Soil Pollution 35: Information provided on humic substances by E4/
217-232. E6 ratios. Soil Science Society of America Journal 41:
352-358.

© 2009 Canadian Water Resources Association


Bourbonniere 409

Chin, Y.-P., G. Aiken, and E. O’Loughlin. 1994. Freeman, C., N. Fenner, N. J. Ostle, H. Kang, D. J.
Molecular weight, polydispersity, and spectroscopic Dowrick, B. Reynolds, M. A. Lock, D. Sleep, S.
properties of aquatic humic substances. Hughes, and J. Hudson. 2004. Export of dissolved
Environmental Science and Technology 28: 1853- organic carbon from peatlands under elevated
1858. carbon dioxide levels. Nature 430: 195–198.

Clair, T. A., P. Arp, T. R. Moore, M. Dalva, and F. Gabka, M. and M. Lamentowicz. 2008. Vegetation-
R. Meng. 2002. Gaseous carbon dioxide and environment relationships in peatlands dominated
methane, as well as dissolved organic carbon losses by Sphagnum fallax in western Poland. Folia
from a small temperate wetland under a changing Geobotanica 43: 413-429.
climate. Environmental Pollution 116: S143-S148.
Glaser, P. H., D. I. Siegel, A. S. Reeve, J. A. Janssens, and
Clay, G. D., F. Worrall, and E. D. G. Fraser. 2009. D. R. Jancky. 2004. Tectonic drivers for vegetation
Effects of managed burning upon dissolved patterning and landscape evolution in the Albany
organic carbon (DOC) in soil water and runoff River region of the Hudson Bay lowlands. Journal
water following a managed burn of a UK blanket of Ecology 92: 1054-1070.
bog. Journal of Hydrology 367: 41-51.
Glatzel, S., K. Kalbitz, M. Dalva, and T. Moore. 2003.
Cooper, D. J. and R. E Andrus. 1994. Patterns of Dissolved organic matter properties and their
vegetation and water chemistry in peatlands of the relationship to carbon dioxide efflux from restored
west-central Wind River Range, Wyoming, U.S.A. peat bogs. Geoderma 113: 397-411.
Canadian Journal of Botany 72: 1586-1597.
Godwin, K. S., J. P. Shallenberger, D. J. Leopold, and
Cummins, T. and E. P. Farrell. 2003a. Biogeochemical B. L. Bedford. 2002. Linking landscape properties
impacts of clearfelling and reforestation on blanket to local hydrogeologic gradients and plant species
peatland streams I. phosphorus. Forest Ecology and occurrence in minerotrophic fens of New York
Management 180: 545-555. State, USA: A hydrogeologic setting (HGS)
framework. Wetlands 22: 722-737.
Cummins, T. and E. P. Farrell. 2003b. Biogeochemical
impacts of clearfelling and reforestation on blanket Gorham, E., S. J. Eisenreich, J. Ford, and M. V.
peatland streams II. Major ions and dissolved Santlemann. 1985. The chemistry of bog waters.
organic carbon. Forest Ecology and Management Ch. 15 In Chemical Processes in Lakes, Stumm, W.
180: 557-570. (Ed.). Wiley-Interscience, New York, 339-363.

Foster, D. R., G. A. King, and M. V. Santleman. 1988. Hájek, M., P. Hekera, and P. Hájková. 2002. Spring
Patterned fens of western Labrador and adjacent fed vegetation and water chemistry in the western
Quebec: phytosociology, water chemistry, landform Carpathian flysch zone. Folia Geobot. 37: 205-
features, and dynamics of surface patterns. 244.
Canadian Journal of Botany 66: 2402-2418.
Hájková P. and M. Hájek. 2003. Species richness and
Fraser, C. J. D., N. T. Roulet, and T. R. Moore. above-ground biomass of poor and calcareous
2001. Hydrology and dissolved organic carbon spring fens in the flysch West Carpathians and
biogeochemistry in an ombrotrophic bog. their relationships to water and soil chemistry.
Hydrological Processes 15: 3151-3166. Preslia, Praha 75: 271–287.

Freeman, C., C. D. Evans, D. T. Monteith, B. Reynolds,


and N. Fenner. 2001. Export of organic carbon
from peat soils. Nature 412: 785.

© 2009 Canadian Water Resources Association


410 Canadian Water Resources Journal/Revue canadienne des ressources hydriques

Hájková P. and M. Hájek. 2004. Environmental Koprivnjak, J.-F. and T. R. Moore. 1992. Sources, sinks
factors and Carpathian spring fen vegetation: The and fluxes of dissolved organic carbon in subarctic
importance of scale and temporal variation. Ann. fen catchments. Arctic and Alpine Res. 24: 204-
Bot. Fennici 41: 249-262. 210.

Hinton, M. J., S. L. Schiff, and M. C. English. 1997. Kramer, J. 1982. Alkalinity and acidity. In Water
The significance of storms for the concentration Analysis Vol. 1, Inorganic Species Part 1. Minear, R.
and export of dissolved organic carbon from two A. and L. H. Keith (Eds.). Academic Press, New
Precambrian Shield catchments. Biogeochemistry York, 85-135.
36: 67–88.
Leenheer, J. A. 1981. Comprehensive approach
Helmer, E. H., N. R. Urban, and S. J. Eisenreich. 1990. to preparative isolation and fractionation of
Aluminum geochemistry in peatland waters. dissolved organic carbon from natural waters and
Biogeochemistry 9: 247-276. wastewaters. Environmental Science and Technology
15: 578–587.
Hruska, J., C. E. Johnson, and P. Kram. 1996. Role of
organic solutes in the chemistry of acid-impacted Lundin, L. and B. Bergquist. 1990. Effects on water
bog waters of the western Czech Republic. Water chemistry after drainage of a bog for forestry.
Resources Research 32: 2841-2851. Hydrobiologia 196: 167-181.

Hughes, S., B. Reynolds, J. A. Hudson, and C. Freeman. Malmer, N., D. G. Horton, and D. H. Vitt. 1992.
1997. Effects of summer drought on peat soil Element concentrations in mosses and surface
solution chemistry in an acid gully mire. Hydrology waters of western Canadian mires relative to
and Earth System Sciences 1: 661-669. precipitation chemistry and hydrology. Ecography
15: 114-128.
IPCC. 2007. Climate Change 2007: The Physical Science
Basis. Contributions of Working Group I to the Marin, L. E., T. M. Kratz, and C. J. Bowser. 1990. Spatial
Fourth Assessment Report of the Intergovernmental and temporal patterns in the hydrogeochemistry of
Panel on Climate Change. Solomon, S., D. Qin, a poor fen in northern Wisconsin. Biogeochemistry
M. Manning, Z. Chen, M. Marquis, K.B. Averyt, 11: 63-76.
M. Tignor, and H. L. Miller (Eds.). Cambridge
University Press, Cambridge, UK and New York, Martinez-Cortizas, A., X. Pontevedra-Pombal, E.
NY, USA, 996 pp. Garcia-Rodeja, J. C. Novoa Muños, and W. Shotyk.
1999. Mercury in Spanish peat bog: Archive of
Kalbitz, K., S. Geyer, and W. Geyer. 2000. A comparative climate change and atmospheric metal pollution.
characterization of dissolved organic matter by Science 284: 939–942.
means of original aqueous samples and isolated
humic substances. Chemosphere 40: 1305-1312. McKnight, D. M., E. W. Boyer, P. K. Westerhoff, P.
T. Doran, T. Kulbe, and D. T. Andersen. 2001.
Kingsbury, C. M. and T. R. Moore. 1989. Over-winter Spectrofluorometric characterization of dissolved
chemistry of subarctic fens, eastern Canada. Nordic organic matter for indication of precursor
Hydrology 20: 97-108. organic material and aromaticity. Limnology and
Oceanography 46: 38–48.
Kononova, M. 1966. Soil Organic Matter – 2nd Edition.
Pergamon Press Ltd., London, 505pp. McKnight, D., E. M. Thurman, and R. L. Wershaw.
1985. Biogeochemistry of aquatic humic substances
in Thoreau’s Bog, Concord, Massachusetts. Ecology
66: 1339–1352.

© 2009 Canadian Water Resources Association


Bourbonniere 411

Miletti, T. E., C. N. Carlyle, C. R. Picard, K. M. Mulac, National Wetlands Working Group. 1988. Wetlands
A. Landaw, and L. H. Fraser. 2005. Hydrology, of Canada. Ecological Land Classification Series,
water chemistry and vegetation characteristics of No. 24. Sustainable Development Branch,
a tamarack bog in Bath Township, Ohio: Towards Environment Canada, Ottawa, Ontario, and
restoration and enhancement. Ohio Journal of Polyscience Publications Inc., Montreal, Quebec,
Science 105: 21-30. 452pp.

Mitsch, W. J. and J. G. Gosselink. 2000. Wetlands - 3rd National Wetlands Working Group. 1997. The
Edition. John Wiley & Sons, New York, 920pp. Canadian Wetland Classification System – 2nd
Edition. Warner, B. G. and C. D. A. Rubec (Eds.).
Morel, F. M. M. 1983. Principles of Aquatic Chemistry. Wetlands Research Centre, Waterloo, Ontario,
John Wiley & Sons, New York, 446pp. 68pp.

Moore, P. D. and D. J. Bellamy. 1974. Peatlands. Navratilova, J., J. Navratil, and M. Hajek. 2006.
Springer-Verlag, New York, 221pp. Relationships between environmental factors and
vegetation in nutrient-enriched fens at fishpond
Moore, T. R. 1987. Patterns of dissolved organic matter margins. Folia Geobotanica 41: 353-376.
in subarctic peatlands. Earth Surface Processes and
Landforms 12: 387-397. Nicholson, B. J. and D. H. Vitt. 1994. Wetland
development at Elk Island National Park, Alberta,
Moore, T. R. and M. Dalva. 2001. Some controls on Canada. Journal of Paleolimnology 12: 19-34.
the production of DOC by plant tissues and soils.
Soil Science 166: 38-47. Oliver, B. G., E. M. Thurman, and R. L. Malcolm.1983.
The contribution of humic substances to the
Moore, T. R. and L. Matos. 1999. The influence of acidity of coloured natural waters. Geochimica et
source on the sorption of dissolved organic carbon Cosmochimica Acta 47: 2031-2035.
by soils. Canadian Journal of Soil Science 79: 321-
324. Poveledo, D. and H. L. Golterman (Eds.). 1973.
Humic substances: Their structure and function in
Moore, T. R., L. Matos, and N. T. Roulet. 2003. the biosphere. Centre for Agricultural Publishing
Dynamics and chemistry of dissolved organic and Documentation, Wageningen, Netherlands,
carbon in Precambrian Shield catchments and an 368pp.
impounded wetland. Canadian Journal of Fisheries
and Aquatic Sciences 60: 612-623. Proctor, M. C. F. 2003. Malham Tarn Moss: The
surface-water chemistry of an ombrotrophic bog.
Moore, T. R., D. Pare, and R. Boutin. 2008. Production Field Studies 10: 553-578.
of dissolved organic carbon in Canadian forest
soils. Ecosystems 11: 740-751. Proctor, M. C. F. and E. Maltby. 1998. Relations
between acid atmospheric deposition and the
Moore, T. R., N. T. Roulet, and J. M. Waddington. surface pH of some ombrotrophic bogs in Britain.
1998. Uncertainty in predicting the effect of Journal of Ecology 86: 329-340.
climatic change on the carbon cycling of Canadian
peatlands. Climatic Change 40: 229-245. Pugh, A. L., S. A. Norton, M. Schauffler, G. L. Jacobson,
J. S. Kahl, W. F. Brutsaert, and C. F. Mason. 1996.
Interactions between peat and salt-contaminated
runoff in Alton Bog, Maine, USA. Journal of
Hydrology 182: 83-104.

© 2009 Canadian Water Resources Association


412 Canadian Water Resources Journal/Revue canadienne des ressources hydriques

Reeve, A. S., D. I. Siegel, and P. H. Glaser. 1996. Shotyk, W., H. W. Nesbitt, and W. S. Fyfe. 1992.
Geochemical controls on peatland pore water from Natural and anthropogenic enrichments of trace
the Hudson Bay lowland: A multivariate statistical metals in peat profiles. International Journal of Coal
approach. Journal of Hydrology 181: 285-304. Geology 20: 49-84.

Richburg, J. A., W. A. Q. Patterson, and F. Lowenstein. Shotyk, W., D. Weiss, P. G. Appleby, A. K. Cheburkin,
2001. Effects of road salt and Phragmites R. Frei, M.Gloor, J. D. Kramers, S. Reese, and W.
australis invasion on the vegetation of a western O. van der Knaap. 1998. History of atmospheric
Massachusetts calcareous lake-basin fen. Wetlands lead deposition since 12,370 14C yr BP from a peat
21: 247-255. bog, Jura Mountains, Switzerland. Science 281:
1635–1640.
Rochefort, L., F. Quinty, S. Compeau, K. Johnson, and
T. Malterer. 2003. North American approach to Sjörs, H. and U. Gunnarsson. 2002. Calcium and pH
the restoration of Sphagnum dominated peatlands. in north and central Swedish mire waters. Journal
Wetland Ecology and Management 11: 3-20. of Ecology 90: 650-657.

Roulet, N. T. 2000. Peatlands, carbon storage, Stevenson, F. J. 1982. Humus Chemistry. John Wiley &
greenhouse gases, and the Kyoto Protocol: Sons, New York, 443pp.
Prospects and significance for Canada. Wetlands
20: 605-615. St. Louis, V. L., J. W. M. Rudd, C. A. Kelly, K. G. Beaty,
N. S. Bloom, and R. J. Flett. 1994. Importance of
Sachse, A., D. Babenzien, G. Ginzel, J. Gelbrecht, and wetlands as sources of methyl mercury to boreal
C. E. W. Steinberg. 2001. Characterization of forest ecosystems. Canadian Journal of Fisheries
dissolved organic carbon (DOC) in a dystrophic and Aquatic Sciences 51: 1065-1076.
lake and an adjacent fen. Biogeochemistry 54: 279-
296. Strack, M., J. M. Waddington, R. A. Bourbonniere,
E. L. Buckton, K. Shaw, P. Whittington, and
Scheithauer, J., K. Grunewald, and D. Pavlik. 2005. J. S. Price. 2008. Effect of water table drawdown
NOM quality and transfer from peat bogs due on peatland dissolved organic carbon export and
to environmental changes in the mountains dynamics. Hydrological Processes 22: 3373-3385.
–Erzgebirge- (Germany). Scientific Papers of the
University of Latvia 692: 127-137. Stumm, W. and J. J. Morgan 1981. Aquatic Chemistry
– 2nd Edition. John Wiley & Sons, New York,
Schiff, S., R. Aravena, E. Mewhinney, R. Elgood, 780pp.
B. Warner, P. Dillon, and S. Trumbore. 1998.
Precambrian Shield wetlands: Hydrologic control Szumigalski, A. R. and S. E. Bayley. 1997. Net
of the sources and export of dissolved organic aboveground primary production along a
matter. Climatic Change 40: 167-188. peatland gradient in central Alberta in relation to
environmental factors. Ecoscience 4: 385-393.
Schnitzer, M. and S. U. Khan. 1972. Humic Substances
in the Environment. Marcel Dekker Inc. New York, Tahvanainen, T. 2004. Water chemistry of mires in
327pp. relation to the poor-rich vegetation gradient and
contrasting geochemical zones of the north-
Shotyk, W. 1988. Review of the inorganic geochemistry eastern Fennoscandian Shield. Folia Geobotanica
of peats and peatland waters. Earth-Science Reviews 39: 353-369.
25: 95-176.

© 2009 Canadian Water Resources Association


Bourbonniere 413

Tahvanainen, T. and T. Tuomaala. 2003. The reliability Wassen, M. J. and H. J. Joosten. 1996. In search of a
of mire water pH measurements – A standard hydrological explanation for vegetation changes
sampling protocol and implications to ecological along a fen gradient in a Biebrza Upper Basin
theory. Wetlands 23: 701-708. (Poland). Vegetatio 124: 191-209.

Thurman, E. M. 1985. Organic Geochemistry of Natural Wetzel, R.G. 1983. Limnology – 2nd Edition. Saunders
Waters. Martinus Nijhoff/Dr. W. Junk, Dordrecht, Publishing, Fort Worth, TX, 825pp.
497pp.
Wheeler, B. D. and M. C. F. Proctor. 2000. Ecological
Thurman, E. M. and R. L. Malcolm. 1981. gradients, subdivisions and terminology of north-
Preparative isolation of aquatic humic substances. west European mires. Journal of Ecology 88: 187-
Environmental Science and Technology 15: 463- 203.
466.
Wieder, R. K. 1985. Peat and water chemistry at Big
Urban, N. R., S. E. Bayley, and S. J. Eisenreich. 1989. Run Bog, a peatland in the Appalachian mountains
Export of dissolved organic carbon and acidity of West Virginia, USA. Biogeochemistry 1: 277-
from peatlands. Water Resources Research 25: 1619- 302.
1628.
Wilcox, D. A. 1986. The effects of deicing salts on
Vasander, H. and A. Kettunen. 2006. Carbon in boreal water chemistry in Pinhook Bog, Indiana. Water
peatlands. Ch 9 in Boreal Peatland Ecosystems, Resources Bulletin 22: 57-65.
Wieder, R. K. and D. H. Vitt (Eds.). Springer-
Verlag, Berlin, 165-194. Wind-Mulder, H. and D. H. Vitt. 2000. Comparisons
of water and peat chemistries of a post-harvested
Verhoeven, J. T. A. and H. H. M. Arts. 1992. Carex and undisturbed peatland with relevance to
litter decomposition and nutrient release in mires restoration. Wetlands 20: 616-628.
with different water chemistry. Aquatic Botany 43:
365-377. Worrall, F., T. Burt, and J. Adamson. 2004. Can climate
change explain increases in DOC flux from upland
Vitt, D. H. and S. Bayley. 1984. The vegetation and peat catchments? Science of the Total Environment
water chemistry of four oligotrophic basin mires in 326: 95-112.
northwestern Ontario. Canadian Journal of Botany
62: 1485-1500. Zaccone, C., V. D’Orazio, W. Shotyk, and T. M. Miano.
2009. Chemical and spectroscopic investigation of
Vitt, D. H., S. E. Bayley, and T.-L. Jin. 1995. Seasonal porewater and aqueous extracts of corresponding
variation in water chemistry over a bog-rich fen peat samples throughout a bog core ( Jura
gradient in continental western Canada. Canadian Mountains, Switzerland). Journal of Soils and
Journal of Fisheries and Aquatic Sciences 52: 587- Sediments 9: 443-456.
606.

Waddington, J. M. and N.T. Roulet. 1997. Groundwater


flow and dissolved carbon movement in a boreal
peatland. Journal of Hydrology 191: 122-138.

Waddington, J. M., K. Toth, and R. Bourbonniere. 2008.


Dissolved organic carbon export from a cutover
and restored peatland. Hydrological Processes 22:
2215-2224.

© 2009 Canadian Water Resources Association


View publication stats

Vous aimerez peut-être aussi