MT Iacopo F 03-02-2015

114
Università degli Studi di Milano – Bicocca SCUOLA DI SCIENZE Dipartimento di Scienze dell’Ambiente e del Territorio e di Scienze della Terra Corso di Laurea Magistrale in Scienze e Tecnologie per l’Ambiente e il Territorio The relationships between water table and redox potential in peatlands ______________________________________________________________________________ _________ Anno Accademico 2013 /2014 Relatore: Prof. Roberto Comolli Tesi di Laurea di: Iacopo Federico Ferrario Correlatore: Prof. Ruurd Van Diggelen Matricola: 770166

description

tesi

Transcript of MT Iacopo F 03-02-2015

  • Universit degli Studi di Milano Bicocca

    SCUOLA DI SCIENZE

    Dipartimento di Scienze dellAmbiente e del Territorio e di Scienze della Terra

    Corso di Laurea Magistrale in Scienze e Tecnologie per lAmbiente e il Territorio

    The relationships between water

    table and redox potential in

    peatlands

    ______________________________________________________________________________

    _________

    Anno Accademico 2013 /2014

    Relatore:

    Prof. Roberto Comolli

    Tesi di Laurea di:

    Iacopo Federico Ferrario

    Correlatore:

    Prof. Ruurd Van Diggelen

    Matricola:

    770166

  • 2

    TABLE OF CONTENTS

    1 INTRODUCTION ................................................................................................................... 4

    1.1 Peatlands ............................................................................................................................... 4

    1.2 Ecohydrology ....................................................................................................................... 6

    1.2.1 Height above water table .............................................................................................. 9

    1.2.2 Micro-topography ....................................................................................................... 10

    1.3 Carbon cycle ....................................................................................................................... 10

    1.3.1 Decomposition ........................................................................................................... 11

    1.3.2 Spatial process of decomposition ............................................................................... 12

    1.4 Microbial community ......................................................................................................... 15

    1.5 Vegetation season: drying and rewetting ........................................................................... 15

    1.6 Redox potential .................................................................................................................. 17

    1.7 Redox potential is a complex indicator .............................................................................. 22

    1.7.1 Assembling the picture ............................................................................................... 33

    1.8 Aim and research questions ................................................................................................ 35

    1.8.1 Hypotheses ................................................................................................................. 35

    2 MATERIALS AND METHODS .......................................................................................... 36

    2.1 Field work .......................................................................................................................... 36

    2.1.1 Site description ........................................................................................................... 36

    2.1.2 Experimental design ................................................................................................... 37

    2.1.3 Redox measurements .................................................................................................. 41

    2.1.4 Water table measurements ......................................................................................... 42

    2.1.5 Pore water sampling ................................................................................................... 43

    2.1.6 Decomposition ........................................................................................................... 44

    2.1.7 Peatsoil sampling ........................................................................................................ 45

    2.2 Laboratory work ................................................................................................................. 45

    2.2.1 Water chemistry analysis ........................................................................................... 45

    2.2.2 Peat analysis ............................................................................................................... 46

    2.2.3 Data calculation .......................................................................................................... 53

    3 RESULTS AND DISCUSSION ........................................................................................... 54

    3.1 Introduction ........................................................................................................................ 54

  • 3

    3.2 Redox potential and water table ......................................................................................... 56

    3.3 pH, Dissolved Inorganic Carbon and Dissolved Organic Carbon ..................................... 68

    3.4 Cellulose decomposition and TBI ...................................................................................... 78

    3.5 Peat quality and degradability ............................................................................................ 83

    4 CONCLUSIONS ................................................................................................................... 93

    5 Acknowledges ....................................................................................................................... 98

    6 Appendices ............................................................................................................................ 99

    7 References ........................................................................................................................... 102

  • 4

    1 INTRODUCTION

    1.1 Peatlands

    Peatland ecosystems are terrestrial environments where, over the long term, net primary

    production exceeds organic matter decomposition, leading to the accumulation of a deposit of

    poorly decomposed organic matter, named peat, thicker than 30 cm (Wieder and Vitt, 2006).

    Peat formation is the result of complex interaction among anoxic conditions, low

    decomposability of the plant material and hydrology. Peatlands are distinguished by the presence

    of water close to the surface, they often have unique soil conditions that differ from the adjacent

    uplands, they support vegetation adapted to wet conditions and they are characterised by an

    absence of flooding-intolerant vegetation. Climate and geomorphology control at the larger scale

    the degree to which peatland can exist, but hydrology underpins its ultimate development,

    affecting the physicochemical environment and influencing the biota (Gosselink and Mitsch,

    2000). The persistence of peatlands depends from the constant supply of water, while source of

    water influences its structure and function. Origin of water has been long employed for

    classifying peatlands. When water comes from surrounding or underlying mineral soil, peatlands

    are termed minerogenous (or geogenous). If peat has grown thick enough to progressively

    become isolated from the mineral soil, precipitation becomes the only source of water and

    peatlands are termed ombrogenus. Minerogenous water carries cations, anions and nutrients and

    the resulting chemical composition has great effects on influencing flora, vegetation and

    ecosystem functions. Minerotrophy is the term describing this condition and forms the ecological

    basis for the peatland type named fen. On the contrary, ombrogenous water has very low

    dissolved minerals and provides the ombrotrophic condition for the development of the peatland

    type named bog.

    Two main environmental gradients are responsible for the differentiation of peatland

    habitat types. One gradient follows the variation in moisture and aeration as a function of water

    table position, which changes in time and space, and of the pore structure of the peat. The

    distance between the soil surface and the water table is the height above water table (HWT) and

    represents roughly the depth of the aerated zone, which has high biological and ecological

    relevance. The other is a complex gradient that combines the variation of pH, base richness and

    nutrient availability. The gradient of nutrient availability and increasing productivity is described

    using three terms borrowed from limnology, which are respectively oligotrophic, mesotrophic

    and eutrophic. This gradient changes independently from the mineral gradient composed of pH,

  • 5

    base cations and associated anions (Wieder and Vitt, 2006), indeed nitrogen, potassium and

    phosphorus have their own chemistry and variation and all assumptions of strong correlation

    with pH and base richness are not obvious (Rydin and Jeglum, 2013). The factors governing the

    differentiation into the main habitats are the same as those causing vegetation differentiation, so

    that a useful system to describe habitat variation in peatlands relies on vegetation pattern.

    Therefore, the floristic composition mirrors change in pH and base richness, creating the bog-

    poor fen-rich series. Bogs are always oligotrophic while rich fens can be either eutrophic or

    oligotrophic. As the peatland becomes more isolated from the groundwater, the pH usually

    decreases because Sphagnum mosses have the ability to acidify water, which cannot be

    contrasted by the low buffering capacity of rainwater. The pH of bogs is comprised between 3.5

    - 4.2, while it increases along the poor-fen rich-fen gradient from 4 to 8 (Rydin and Jeglum,

    2013). The complete absence of bicarbonate alkalinity below pH 5.5 is a fundamental dividing

    point in the habitat limits of many peatland species. Below 5.5 pH fens are dominated by

    oligotrophic species of Sphagnum, while above 5.5 pH Sphagnum abundance decreases and true

    mosses mostly dominate (Wieder and Vitt, 2006). Mire ecologists observed that the measure of

    the water table from the surface (HWT) was a very strong predictor of vegetational gradients in

    peatlands. The water table gradient allows the micro-topography of peatlands to be classified in

    the following elements, which were first described in Sjrs (1948; cf. in Rydin and Jeglum,

    2013):

    Hummocks are mounds of 20-50 cm raised above the lowest surface level. The thick

    aerobic peat supports dwarf shrubs and vascular plants.

    Lawns are intermediate microforms, 5-20 cm above the water table, where graminoids

    are most common. Mosses reach the highest species diversity in lawns. Firm surfaces that

    you can walk on.

    Carpets are from 5 cm below to 5 cm above the water table. They are characterised by a

    soft, green layer of mosses and a sparse cover of cyperaceous plants. Softer surface into

    which your feet sinks.

    Mud-bottoms are most of the time submerged; they lack vascular plants and are

    dominated by mosses or they can also expose the bare peat, usually covered by algae.

    Pools are water-filled depressions.

    Within the peatland, a vegetational gradient develops largely in parallel to HWT, which is deeper

    at the uplands and shallower towards the open mire, passing through the mire margin and the

    mire expanse. The mire margin has usually thinner peat and vascular plants can reach the

  • 6

    mineral layer, so that small, creeping trees and shrubs can grow. pH is lower in the centre and

    higher in margin and calcium ion shows the same pattern. In bogs and especially in raised bogs,

    where the peat meets the upland soil, usually a narrow fen called lagg develops. The lagg is a

    feature of the margin that surrounds the bog and receives water both from the bog and from the

    surrounding mineral soils.

    The peculiar shape of a peatland, its surface morphology and patterning is the result of

    interactions between substrate, climate, hydrology and vegetation. Raised bogs are ombrotrophic

    mires raised above the level of surrounding uplands, which are usually at least 500 m in diameter

    and with a convex cupola that can be several metres higher than the edges and surrounding

    uplands (Rydin and Jeglum, 2013). Water slowly flows from the centre to the edge. Where raised

    bogs develop over round-shaped lakes, they grow circular and hummocks and hollows form

    concentric patterns, perpendicular to the lateral through-flow.

    1.2 Ecohydrology of raised bogs

    Raised bogs are ombrotrophic mires that depend only on precipitation for their supply of water,

    so that they can be found only in climatic regions where the input of water exceed the loss.

    Raised bogs are characterised by vertical oscillation of the surface during drying and rewetting

    (the so called mire breath), small temporal fluctuation of the water table, reduction of

    evapotranspiration occurring at shallow water level and large storage coefficients (Schaaf, 1999).

    In saturated condition, the content of water in undisturbed bog peatsoil can range between 88-97

    % (Ivanov, 1981).

    Peatlands such as raised bogs are not made of a uniform body, but they are characterised

    by a double layer, and mire researchers introduced the terms acrotelm and catotelm to define this

    particular feature (figure 1.1).

    Figure 1.1 Bogs are formed by a double layer. The acrotelm, thin and biologically active and the catotelm, thick and inactive.

  • 7

    The acrotelm is the superficial layer where water table fluctuation occurs and where

    biogeochemical cycles and biological processes are more active, while under a hydrological

    point of view, the high hydraulic conductivity and low degree of humification identifies the

    acrotelm as an aquifer (Rydin and Jeglum, 2013; Schaaf, 1999). The acrotelm thickness can vary

    from 7-8 cm in herb rich fen to 60-70 cm in moss-rich raised mire (Ivanov, 1981). The catotelm

    is the main body of a bog and can be about several metres depth. It is the less active layer where

    biological processes are slow, peat is generally well decomposed and hydraulic conductivity is

    so low that, hydrologically, it is described as an aquitard. The classic Ingrams theory describes

    raised bogs as raised water mounds, with the catotelm as a body that loses water by lateral

    outflow and that is compensated for by infiltration by the overlying acrotelm (figure 1.2). The

    mound shape is a function of the hydraulic conductivity of the catotelm (assumed low and

    spatially constant) and the infiltration rate from the acrotelm (assumed constant).

    Figure 1.2 Precipitation is the only source of water in raised bogs. Rainwater flows outwards from the centre via runoff, lateral through flow and groundwater flow.

    The theory describes raised bogs as fed only by precipitation and isolated from the mineral

    surrounding groundwater. However, recent studies have demonstrated that the isolation from

    groundwater may not be always a rule in bogs. Relatively high hydraulic conductivity, ranging

    from 10-2 to 10-1 m d-1, was measured even in the catotelm (Chason and Siegel, 1986) so that, if

    there is no net hydrological distinction between acrotelm and catotelm, there might be significant

    downward and upward exchanges of water (Sirin et al., 1996 cited in Rydin and Jeglum, 2013)

    (figure 1.3). Internal hydrological mechanisms, the change in precipitation, evapotranspiration

    and the degree by which water table fluctuation affects the hydraulic head in the catotelm govern

    the process and can cause the reversal of hydraulic gradient (Fraser et al., 2001; Devito et al.,

    1997). Consequently, during droughts, the water table drops and mineral groundwater can move

    upward reaching 1-2 m depth from the surface thus affecting pore water chemistry (Gosselink

    and Mitsch, 2000).

  • 8

    Figure 1.3 For raised bog with particular climatic and peat conditions upwelling of mineral water can occur.

    Hydraulic conductivity (K) is the property that governs the flow of water and transport of

    solutes in a porous medium. In a bog, hydraulic conductivity varies with peat type, degree of

    humification and bulk density. Peat type is characterised by different structures influencing

    hydraulic conductivity, for instance K is lowest for Sphagnum peat and higher for lignoid peat

    (Rydin and Jeglum, 2013). Generally, K correlates negatively to degree of humification. The

    degree of humification can be measured with the von Posts method, which assigns an increasing

    number to the change of some diagnostic features that correlate with degree of decomposition.

    Humification can vary either vertically on a scale of decimetres as horizontally on a scale of few

    metres (Schaaf, 1999). Humification, however, does not increase necessarily with depth because

    it also depends on different factors: vegetation (Silamikele et al., 2007), peat forming conditions,

    different decaying rate in the catotelm, physical conditions, micro-topography and climatic

    conditions (Schaaf, 1999). Hydraulic conductivity correlates negatively to bulk density. Bulk

    density tends to increase with depth since compaction caused by overlaying peat layers reduces

    the volume of pores. All the factors described above vary spatially so that also K is spatially

    dependent, despite the Ingrams theory assumed constant hydraulic conductivity throughout the

    catotelm. Indeed, more recent works showed that in the catotelm K tends to have a vertical

    decreasing trend with depth while it tend to decrease from the centre to the bog margin (Beer et

    al., 2008; Schaaf, 1999), invalidating the assumption of the Ingrams theory. The way K varies in

    a bog controls the vertical and horizontal flow of water. In bogs, water has a preferential

    horizontal flow because vertical hydraulic conductivity can be as three orders of magnitude

    lower than horizontal K (Beer et al., 2008). In the acrotelm, compaction and decomposition

    cause the hydraulic conductivity to decrease strongly with depth from about 105 m d-1 at the

    surface to 110 m d-1 at some decimetres below it (Schaaf, 2004; 1999). Fraser et al. (2001)

    measured a great decrease of K that reached 10-3 m d-1 at 45 cm below the surface. Yet, hydraulic

  • 9

    conductivity can be dynamic. Drying can decrease K compressing the pores while precipitation

    can cause peat to swell, increasing its hydraulic conductivity. Given the vertical trend of K and

    the peat properties change during drying-rewetting cycles, the rate of lateral subsurface flow in

    the acrotelm depends on water table position, so that flow decreases as water table decreases.

    This mechanism provides an important feedback that augments the water retention power in bogs

    during droughts. In the acrotelm, hydraulic conductivity is usually 4-5 times greater horizontally

    than vertically (Beer et al., 2008; Schaaf, 1999). Runoff can occurs if the rate of precipitation

    fills the storage capacity of hollows and exceeds the rate of infiltration. In ecosystems like raised

    bogs, characterised by a small slope and a well developed hummock-hollow topography, runoff

    occurs as sheet flow and channel flow in micro channels. The micro-topography adds complexity

    to the flow pattern because depressions can be connected during wetting and can be disconnected

    during drying and the flow pattern can reverse due to the different filling and emptying pattern of

    depressions while some depression can also not participate to surface flow (van der Ploeg et al.,

    2011). The transition between subsurface flow and runoff is controlled by complex threshold

    mechanisms, and Frei et al. (2010) found that surface flow is mainly a function of depression

    storage and that runoff occurred only during intense rainfall. Subsurface flow is then the

    dominant flow during dry and wet periods while surface flow occurs only during intense rain

    events because micro-topography acts as inhibitors of surface flow (Frei et al. 2010).

    1.2.1 Height above water table

    Height above water table (HWT, i.e. the distance from peat surface to the water table) is

    an important indicator used to predict a number of important eco-hydrological variables in

    peatland hydrology, ecology and biogeochemistry including run-off, saturation, redox potential,

    biodiversity, soil structure, methane emission, peat quality and organic matter decomposition

    (Waddington et al., 2014). Water table governs the water availability to bog plants and

    delimitates the zones of aerobic and anaerobic respiration in peat, so that it is considered a main

    factor controlling vegetation distribution (Rydin and Jeglum, 2013), microbial diversity and

    niche differentiation (Andersen et al., 2013). Water table fluctuation governs the oxygen

    penetration in peat and affects redox potential (Niedermeier and Robinson, 2007; Mansfeldt,

    2003; Seybold et al., 2002), thus determining the spatial distribution of aerobic conditions. That

    condition enhances decomposition not only because aerobic respiration is more efficient than

    anaerobic respiration, but also because it allows the degradation of anti-microbial Sphagnum

    phenols in shallow peat (Abbott et al., 2013; Fenner and Freeman, 2011). Water table level is

  • 10

    also the most important control on the relative importance of methane production in peat and of

    the amount of CH4 fluxes out of peatlands (Bridgham et al., 2013). Yet, from a hydrological

    point of view, water table level can provoke flow reversal in peatlands (Fraser et al., 2001) and it

    is an indicator of water discharge and solute transport out of the peatland into the watershed.

    1.2.2 Micro-topography

    The surface of raised bogs is characterised by distinct hummocks and hollows micro-topographic

    elements. Micro-topography features can alter hydrological, physicochemical and biological

    attributes (Courtwright et al., 2011; Baldwin et al., 2007; Haraguchi, 1992), adding spatial and

    temporal heterogeneity to the turnover of redox-sensitive solutes in peatlands. Therefore, they

    are not just depending on peat quality, peat properties and temperature. The most important

    difference between hummocks and hollows is water-table depth (Bragazza et al., 1998).

    Hydrologically, in continental bogs, above the water table, hummocks have a dominance of

    macropores associated with vascular plant roots, so that they have higher hydraulic conductivity

    than hollows and the water flow is predominantly vertical. The difference in peat physical

    properties and hydraulic conductivity accounted by microforms become less important with

    depth to water table. Branham and Strack (2014) showed that hydraulic conductivity was not

    influenced any more by microforms at a depth of 20 cm below the water table. Frei et al. (2012)

    showed that micro-topography could induce water flow patterns that eventually determined

    biogeochemical hot spots, defined as sites characterised by higher biogeochemical activity than

    the surrounding areas. For example, when water table increases, hummocks can be the only place

    where oxidation can occur. Indeed, a higher concentration of nitrate is usually found in

    hummocks (Frei et al. 2012; Wolf et al., 2011; Bragazza et al., 2005). Through the infiltration

    zone inside the hummocks, oxidised nutrients can be transported by precipitation to deeper and

    more redox-reducing levels where they can be reduced. In this occasion and at the water table

    level, hummocks can become hot spots for reducing reactions (Frei et al. 2012).

    1.3 Carbon cycle

    The study of biogeochemical cycles is of primary importance to unravel the ecological role of

    peatlands at different spatial and temporal scales. Regarding the carbon cycle (figure 1.4),

    Northern peatlands store half (approximately 1672 Gt) of the global soil carbon pool though

    cover an estimated 3.6x106 km2 (Tarnocai et al., 2009), which is equal to about 3 % of the total

  • 11

    lands. Peatlands can be source, sink or transformers of carbon compounds, and many efforts

    were directed to measure its fluxes and to understand their functioning.

    Figure 1.4 Carbon cycle in bogs. Circle boxes represent gas phase; broken lines represent microbial-mediated processes.

    1.3.1 Decomposition

    Measurement of decomposition rate (k) is important for the biogeochemistry of nutrients and

    carbon fluxes (Prescott, 2010). The availability of nutrients is due in large part to the decay

    dynamics of the organic matter. The process also supports diversity in the microbial population

    by supplying a rich set of intermediate degradation products (Berg and McClaugherty, 2014).

    Bogs receive very low amount of nutrients from wet and dry deposition that made up their entire

    external sources. For that reason, decomposition and recycling of nutrients become an essential

    internal source of nutrients in bogs (Bragazza et al., 2013).

    Zhang et al. (2008) reviewed the factors that controlled decomposition rate. These were:

    (i) climatic factors (mean annual temperature, MAT; mean annual precipitation, MAP; annual

    actual evapotranspiration, AET); (ii) litter quality (nitrogen content; carbon:nitrogen ratio, C:N;

    lignin content, LIGN; lignin:N ratio, LIGN:N); (iii) vegetation and litter types; (iv) geographical

    variables (latitude, LAT and altitude, ALT). The authors found that C:N ratio and the total

    nutrient content of the litter (both litter quality factors) were the two most important factors

    influencing decaying rate on a global scale. A threshold-based mechanism dominates the relation

    between decomposition rate and its factors. There is not a single factor that is, in every

  • 12

    conditions, more important than the other ones, so that decomposition can slow down despite all

    the factors but one is adequate (Prescott, 2010).

    In bogs, the litter quality of Sphagnum overwhelms the importance of other factors.

    Sphagnum species forming hummocks have relatively higher content of recalcitrant structural

    carbohydrates (lignin-like polymeric phenolics) than hollows species. Thus, despite hummocks

    are well-aerated and aerobic decomposition rate is higher than anaerobic decomposition, the

    decomposition of Sphagnum litter is lower in hummocks (Bragazza et al., 2013). After litter

    quality, moisture (i.e. HWT) is the most important factor for decomposition that can show a

    threshold effect. In fact, a lower water table caused by droughts can desiccate the litter

    hampering decomposition. Laiho (2006), in a review on decomposition constraint in northern

    peatlands, included pH as important factor because it can limit the activity of phenol oxidase.

    Finally, several studies have shown that temperature increased decomposition only under non-

    limiting moisture condition (Bragazza et al., 2013; Mkiranta et al., 2009).

    Hydrologic dynamics, interactions between surface ombrogenous water and deeper

    groundwater, influence the distribution and transformation of nutrients in wetlands (Frei et al.

    2010) and the decomposition rate in the catotelm (Beer and Blodau, 2007). If in the catotelm

    diffusion dominates over advection there will be a lack of solute transport and an accumulation

    of product of microbial metabolism, such as CH4, dissolved inorganic carbon (DIC) and

    recalcitrant dissolved organic carbon (DOC) (Beer et al., 2008). Beer and Blodau (2007) showed

    that these phenomena can lead to a thermodynamic constraint for microorganism metabolism

    that slows or even cuts off the decay of peat in anoxic layers. It follows that the groundwater

    residence time becomes a fundamental factor in assessing the spatial decomposition rate in bogs

    (Morris et al., 2011). Advective flux like upwelling of mineral water or downwelling of

    rainwater can break the constraint. The influence of rainwater and upwelling depends on the

    profile of hydraulic conductivity. It is reported that in bogs at most the first 100 cm can be

    affected by rainwater (Dobrovolskaya, 2013; Morris et al., 2011).

    1.3.2 Spatial process of decomposition

    Main inputs of carbon in peatland are through photosynthetic production of mosses and then

    vascular plants. In peatlands, plant biomass can be divided into above ground, rhizome and

    coarse roots, and fine roots (diameter < 0.5 mm). Carbon input below ground takes place through

    root exudations and decaying vascular plant materials. Most of the organic matter decomposition

    occurs in the acrotelm, where rate of decomposition has been estimated to be one hundred times

  • 13

    higher than in the catotelm. There is an important biogeochemical connection between the upper

    layers and the deep peat, so that about 10 % of litter mass produced in the acrotelm reaches the

    catotelm (Frolking et al. 2001 cited in Beer et al. 2008), albeit the occurrence of this

    translocation is actually strongly dependent on hydraulic conductivity (Beer et al. 2008).

    Products of decompositions are CO2 and CH4 gases and DOC. The relative importance of

    aerobic and anaerobic respiration depends on the HWT, which controls the oxygen supply into

    the peatsoil. The principal product of aerobic respiration is CO2, while anaerobic respiration

    produces DOC, CH4 and CO2 (Fenner and Freeman, 2011). Oxic conditions favour CO2

    production and inhibit methanogenesis, while anaerobic respiration favours lower CO2

    production but greater CH4 production rate (Estop-Aragons and Blodau, 2012). Respiration rate

    is higher at water table level and then decreases (Shoemaker et al., 2012) so that most of the

    organic matter is decomposed in the acrotelm (Beer and Blodau, 2007). In anoxic layers,

    respiratory pathway depends on Eh conditions, peat quality and relative concentration of electron

    acceptors. Microorganisms compete for electron acceptors in presence of nitrate, ferric ion,

    sulphate and finally CO2. Since bogs are fed only by precipitations, the common electron

    acceptors and nutrients have very low concentrations compared to other minerogenic wetlands

    and they even decrease with depth (Beer et al., 2008; Proctor, 2003; Steinmann and Shotyk,

    1996; Lundin and Bergquist, 1990). Keller and Bridgham (2007) measured that Fe(III) and

    nitrate reduction accounted for less than 1% of anaerobic carbon mineralisation, while sulphate

    reduction was responsible for 6 26 %. These results are consistent with the frequent

    observation that reactive inorganic sulphur (RIS) pool can be very dynamic in bogs (Wieder and

    Lang, 1988). Many studies have stressed that respiration driven by common electron acceptors

    accounts for only a minor fraction of the CO2 produced in bogs. The sources for the unexplained

    CO2 can be the regeneration of electron acceptors during water table drawdown and/or other

    respiratory pathways like bacterial respiration with humic substances (Knorr et al., 2009; Deppe

    et al., 2009; Heitmann et al., 2007), with organic sulphur species (Kertesz, 2000), and

    fermentative processes in absence of electron acceptors. Fermentation can be very important in

    bogs and can lead to high production of DOC (Vile et al., 2003). Fermenting microorganisms are

    important in degrading complex polymers yielding simpler products used in methanogenesis.

    Fermentation, together with DOC reduction, is addressed as responsible of the high fraction of

    unexplained carbon mineralization in anoxic layers (Keller and Bridgham, 2007). There is an

    increasing need to assess the use of organic electron acceptors (DOC) by microorganisms in

    anoxic environment. Comparing sites with different hydrology (bog and fen) Shoemaker et al.

    (2012) found a high production of CO2 in anaerobic layer at the ombrotrophic site and found a

  • 14

    local correlation with concentration of DOC, suggesting that DOC may have driven anaerobic

    respiration. A further complication to the role of DOC is that the chemical composition rather

    than its amount distinguishes its role as electron donor or electron acceptor (Keller et al., 2009).

    The vertical profile of DIC generally increases with depth (Deppe et al., 2009). Several

    authors observed that DIC increased at all depths during summer months, along with

    temperature. Despite the temperature drop during autumn, DIC is reported to be still increasing

    and Shoemaker et al. (2012) reported that the profile of DIC decreased only after frequent rain

    events occurred, since rain dilutes bog water and favours CO2 degassing. The DIC profile

    depends also by emission to atmosphere, which can occur by diffusive flux or by non-diffusive

    flux via roots of vascular plant or via ebullition (Glaser et al., 2004; Tokida et al., 2007).

    Substrate quality decreases with depth and the main substrate for decomposition in

    deeper layer becomes DOC. DOC concentration is higher in surface layer and decreases with

    depth (Deppe et al., 2009; Beer et al., 2008; Fraser et al., 2001). Pore water concentration of

    DOC usually increases in summer due to evapotranspiration and enhanced decomposition

    (Waddington and Roulet, 1997).

    Supply of high quality substrate and water table depth (Rydin and Jeglum, 2013) are the

    main factors controlling the production of methane in peatlands. In general, acetoclastic

    methanogenesis is favoured in the upper peat layers where abundant labile organic carbon from

    decaying vegetation are found, and H2 and CO2-dependent methanogenesis predominates in the

    more recalcitrant deeper peat layers (Beer and Blodau, 2007). The zone of most active CH4

    production was measured about 10 cm below the average water table (Sundh et al., 1994), where

    there is input of fresh organic matter. This zone is closely related to the source of products from

    anaerobic respiration, in fact Shoemaker et al. (2012) found a peak of methane production just

    below the respiration rate peak near the water table surface (often between 0-5 cm below the

    water table). Regarding the role of micro-topography, the hummocks are reported to have lower

    methanogenesis than hollows because there is less quality input of substrate (Bubier et al., 1993).

    It is also arguable that, being oxygenated, the hummocks provide more energetically favourable

    microbial pathways.

    Peak of oxidation of methane is found at the water table level and above it and it is higher

    in hummocks than in hollows (Frenzel and Karofeld, 2000). Occasionally methanotrophy can be

    found in saturated layer. In this case it is driven by oxygen released by roots and by anaerobic

    oxidation of methane.

  • 15

    1.4 Microbial community

    Mires are active systems characterized by organic matter turnover (Biester et al. 2012).

    Microorganisms play a key role in wetland soils. They are involved in fundamental

    biogeochemical cycles and transformation of elements. The rate and extent of their role in these

    processes depend on the variability and change of environmental variables on which their

    metabolism and growth depend (Andersen et al., 2012). The activity of microorganism depends

    on temperature, pH, hydrological regime, oxygen, peat quality, nutrients availability and redox

    potential. These variables are spatially dependent and, in peatlands, different habitats can show

    changes in the distribution of microbial groups and in their total biomass. Distribution of

    microorganism is also depth dependent. Microorganisms face increasing energetic constraints

    with depth, caused by a combination of factors such as the availability of oxygen and

    thermodynamic disadvantage that are characteristic of a stratified peat with redox and peat

    quality change (Robroek et al., 2013). Fungi, as general rule, are dominant in oxic and acid

    habitats, so they are abundant in the aerated surface layer. Microhabitats as hollows and

    hummocks have different vegetation, hydroperiod and redox dynamic. In Robroek et al. (2013)

    an analysis of microbial community between microhabitats has revealed a change in the

    fungal/bacterial ratio toward higher values in hummocks than hollows. Generally fungi decreases

    in biomass with depth (Andersen et al., 2012), but recently Jassey et al. (2011) have found larger

    number in lower peat layer, a finding described also in Dobrovilskaya (2014). Bacteria instead

    dominate in anoxic and neutral environment. The bacterial abundance generally decreases with

    depth, but some studies have revealed increasing biomass with depth (Golovchenko et al., 2007),

    or peaks at some depth (Grodnitskaya and Trusova (2009) cited in Rydin and Jeglum, 2013).

    In a peatland, microorganisms respond to alteration of hydrological regime following the

    time and space changes and the intensity of the event. Persistent drought could trigger

    vegetational shifts and indirectly affect microorganisms. During drought methanogenic bacteria

    can survive in anoxic pores still present in aerobic layers (Kotiaho et al., 2013) and aerobic

    bacteria can increase in number following the lowering of water table (Karsisto, 1979).

    Methanotrophic bacteria have been found throughout the profile but in greater concentration at

    certain depth so that under particular condition the population can become active.

    Methanotrophic bacteria, which can tolerate extended periods of anoxia, can resume methane

    oxidation within few hours of re-exposure to oxygen.

    1.5 Vegetation season: drying and rewetting

  • 16

    Drought is a period of scarcity of water that is statistically derived from climatic data. Regarding

    the effect on biogeochemical processes in peatlands, droughts can be defined according to their

    degree of intensity. Severe droughts can affect microorganisms either directly or indirectly, on a

    long-term range, after driving a shift in vegetation. Short droughts, although having minor effect

    on vegetation, can also affect directly the microbial community. Droughts influence microbes by

    lowering the moisture content of peat (i.e. water table) and enabling the penetration of oxygen in

    deeper peat layers so that aerobic respiration can establish. If drought is severe, shortage of water

    can lead to microbial mortality (Mettrop et al., 2014). Regarding climate change, the effects of

    droughts on the carbon cycle is under high concern since an increase of their severity or

    frequency could change the role of peatlands from being sinks of carbon to being sources.

    Drought intensity and following rewetting are the most important factors that can

    enhance decomposition of peat soil. Water table drawdown allows oxygen to reach previous

    anaerobic layers of peat. Usually, although the pattern is very peat type-dependent, CO2

    increases during drying until an optimum moisture level is reached and afterward it decreases

    (Estop-Aragons and Blondau, 2012), because moisture becomes limiting. The higher

    decomposition rate occurring during droughts, allows the release of nutrients in the peat, a

    process called eutrophication. During the rewetting phase, these nutrients can be transported

    downward and can serve to enhance microbial decomposition in low nutrient layers.

    Fenner and Freeman (2011) have proposed a theory, called the enzymatic latch theory,

    which aims to explain the mechanism underlying the bogs response to droughts. As mentioned

    above, the persistence of bogs is due mainly to waterlogged anoxic condition and recalcitrance of

    Sphagnum. Sphagnum spp. have high amount of phenolic compounds, with inhibiting effects on

    microbes, that can be degraded only in presence of oxygen by phenol oxidases enzymes. When

    water table is high, oxygen cannot penetrate in peat and that can prevent the peat deposit to be

    released as CO2. When drought introduces oxygen, phenol oxidase can remove phenolic

    inhibitors, enabling hydrolases to resume normal mineralization of organic matter and increase

    decomposition.

    Consequently, droughts influence microbial activity and lead to a hydrochemical shift in

    surface water. As mentioned above, eutrophication releases nutrients, while another important

    effect of aeration is the re-oxidation and regeneration of electron acceptors (Deppe et al. 2009).

    At the same time, the oxidation of reduced redox couples by aerobic respiration releases

    hydrogen ions bringing acidification (Brouns et al., 2014; Mettrop et al., 2014; Clark et al.,

    2009), which in turn further limits microbial activity. In Fenner and Freeman (2011), sulphate

    and nitrate were released after 23 days of drought while potassium and P after 48 days. Sulphate

  • 17

    was released within weeks in Brouns et al. (2014) and Proctor (1994), followed by acidification.

    The time scale of response of oxidising and reducing processes depends on process, depth and

    peat heterogeneity (Knorr et al., 2009). According to the enzymatic latch theory, if a drought is

    not intense enough, it is possible that phenolics will not be degraded and then eutrophication and

    increasing of CO2 will not occur (Mettrop et al., 2014). Fenner and Freeman (2011) have

    observed a strong increase of CO2 after a severe drought. Despite the increasing production of

    CO2, in the zone of aeration, DIC can be very low because of degassing (Deppe et al., 2009). The

    production of DOC seems related to drought intensity. During aeration, DOC becomes the

    substrate of aerobic respiration (Hughes et al. 1997) and it decreases (Mettrop et al., 2014;

    Fenner and Freeman, 2011). However, in a mesocosm experiment, Mettrop et al. (2014)

    observed that after strong desiccation DOC greatly increased, which the authors proposed that

    this is a consequences of microbial mortality or a change in microbial community composition.

    Rewetting in bogs occurs by means of precipitations. The infiltration of rainwater to

    lower layer is important after droughts (Deppe et al. 2009) because causes the pH to increase

    (Fiedler et al. 2007). Fenner and Freeman (2011) measured increasing flux of CO2 after

    rewetting, which the authors explained as an effect of eutrophication and removal of pH

    constraints. Deppe et al (2009) measured a rapid increase of DIC and CH4 after flooding,

    whereas drying and following rewetting had not strong effects on DOC. Clark et al. (2009)

    measured that the drying-rewetting cycles influenced DOC down to 55 cm depth. Proctor (2003)

    observed a sharp peak of SO4 after rewetting (Proctor, 2003).

    1.6 Redox potential

    Redox reactions are chemical reactions that involve the transfer of electrons between two

    species, molecules, atoms or ions, so that their oxidation state changes. Oxidation is the loss of

    electrons or an increase of oxidation state and reduction is the gain of electrons or a decrease in

    oxidation state. Reduction and oxidation must occur simultaneously, since, in a redox reaction,

    the reducing agent transfers the electrons to an oxidised agent and free electrons cannot exist in

    solution. Conceptually, redox reactions are described as two half-reactions, one releasing

    electrons and the other gaining electrons, combined to form a whole reaction. The two related

    species that exchange electrons and change their oxidation state in a redox reaction are called

    redox pair or redox couple. In analogy with pH, the hypothetical electron activity of a solution is

    given by p = - log{e-}. The quantity p measures the relative tendency of a solution (with one or

    more redox couples) to accept or transfer electrons and it is a measure of the Gibbs free energy

    of the redox reaction. Also, the redox potential (Eh) is a measure of the electron availability in a

  • 18

    solution, though it is made with an electrochemical cell, where the potential is expressed in

    relation to the standard hydrogen electrode. In making an electrochemical measurement of the

    redox intensity, an electro motive force (Volts) is measured. Eh is related to P by

    hERT

    Fp

    3.2 Eq. 1.1

    Where Eh is the redox potential [V] of the solution (in relation to reference electrode); F is the

    faraday constant (= 96940 Cmol-1 e-); R is the gas constant (= 8.314 J mol-1K-1); T is

    temperature in K.

    The thermodynamic definition of the redox potential of a solution is given by the

    Nernsts equation, which describes the relationship between Eh and the activities of oxidised and

    reduced species

    d

    Ox

    nF

    RThEEh

    Relog3.2 Eq. 1.2

    Where Eh is the redox potential [V] of the solution (in relation to reference electrode); Eh is the

    redox potential [V] under standard conditions (all activities = 1, pH2= 1atm, [H+] = 1M) and is

    related to the free energy change for the cell reaction (G); F = 1 is the faraday constant (=

    96940 Cmol-1); n is the number of exchanged electrons; R is the gas constant (= 8.314 J mol-

    1K-1); T is the temperature in K; 2.3RT/F = 0.059 V (at 25 C); {Ox}/{Red}is the activity of

    oxidised/reduced couples.

    From equation 1.2, it follows that Eh depends on the ratio of oxidized and reduced forms

    (i.e. the relative activities) and not on their absolute quantities. For that reason, Eh is considered

    an intensity factor for the reduction/oxidation (i.e. the greater the ratio the more oxidised the

    system). Contrarily, redox capacity, a function of the amount of oxidised and reducing

    compounds, measures the buffer effect of the solution. The intensity factor determines the

    relative ease of the reduction/oxidation whereas the capacity factor denotes the extent to which

    the shift will take place. In aqueous natural environments, the water solvent exerts a levelling

    influence on the system and restricts the range of accessible Eh between the intensity of water

    reduction (E= - 0.83 V) and the intensity of water oxidation (E= + 1.23 V).

    Redox reactions are common in nature and they are primary driver of biogeochemical

    processes. The metabolism of living organisms relies on redox reactions for energy and for

    provision of building blocks, such as for example respiration. Microbes are an important control

  • 19

    of redox reactions in soils. They catalyse the kinetic of redox reactions, transferring electrons

    from reduced inorganic or organic matter to inorganic or organic terminal electron acceptors

    (TEAs), to produce energy for their metabolism. The kind of redox couple, electron donor and

    electron acceptor, determines the amount of energy (i.e. Gibbs energy) gained by the

    microorganisms out of the process. In a closed aquatic system with organic matter as energy

    source and microbes, it is possible to calculate the sequence of oxidation reaction coupled to

    TEAs expected on the base of their thermodynamic possibility. Oxygen is the most abundant and

    the strongest oxidised agent that is readily available in the atmosphere so that, in aerobic

    condition, microbial respiration uses oxygen as final electron acceptor to oxidise organic matter

    and produce energy. As water table rises, water saturates the soil and slows down the rate of

    oxygen diffusion. Consequently, microorganisms consume oxygen more rapidly than it is

    supplied, and, after oxygen is completely depleted, the oxidation of organic matter is coupled to

    other TEAs. The temporal sequence predicted on the basis of decreasing Gibbs free energy is the

    following: NO3-, Fe(III), Mn(IV), SO4

    2- and finally CO2. To put it differently, nitrate has a better

    affinity for electrons than sulphate so that the microbes require a higher electron pressure (lower

    Eh) to transfer electrons from organic matter to sulphate.

    Each of these reactions occurs within defined values of redox potential, or thresholds,

    which can be used to predict the occurrence of the relative redox reactions in soils, though there

    are currently arguments about where to draw these boundaries and the matter needs to be verified

    more extensively. According to Gosselink and Mitsch (2007), local factors like pH and

    temperature, influence the values of these thresholds. Nevertheless, if caution is taken in

    interpreting the results they can still be useful in characterising redox condition in soils (Reddy

    and DeLaune, 2008). In addition, it has to be borne in mind that such boundaries are rather

    smooth stripes than straight lines, therefore they should not be taken as exact borders between

    soil redox reactions. According to Gosselink and Mitsch (2007), the microbial oxidation of

    organic substrate uses oxygen as terminal electron acceptor at a redox potential of between +400

    and +600 mV (Reddy and DeLaune (2008) suggested that below 300 mV oxygen is completely

    absent). Below +400 mV nitrate starts to being used as electron acceptor; at about +225 mV

    manganese is reduced; between +100 and -100 mV iron is transformed from ferric to ferrous

    form; from -100 to -200 mV sulphate is reduced to sulphide; eventually, below -200 mV carbon

    dioxide is reduced to methane. There are different thresholds and classifications of redox zones

    in literature. Figure 1.5 shows a comparison among other classifications.

  • 20

    Figure 1.5. Redox zones defined by different authors, based on Fiedler (2007).

    As mentioned before, an electrochemical cell measures the difference in potential

    (electron motive force, emf) between an inert indicator electrode in contact with a redox couple

    in solution and a reference electrode. The indicator electrode is usually an inert metal like

    platinum and the reference electrode can be H+/H2 or Ag/AgCl 1mol/l KCl. The reaction of the

    cell where the reference electrode is hydrogen is

    Pt, H2 (pH2=1)| H+ (a=1) | Mn+ | M

    The overall cell reaction is

    Mn+ + nH2 M(s) + 2nH+

    If the electrode potential is positive, the above reaction is the spontaneous reaction in the

    direction left to right. If the electrode potential is negative, the spontaneous reaction is in

    the opposite direction. The voltage of the cell, when the activities of all ions in the cell are

  • 21

    unity, when gases are at 1 atm pressure and solids are in their most stable form at 25C, is

    calculated as

    Ecell = Eh(ox-red) Eref Eq. 1.3

    Rearranging eq. 1.3 gives

    Eh(ox-red) = Ecell + Eref Eq. 1.4

    That is the redox potential for the solution measured.

    The measurement of redox potential with platinum electrodes has wide applications. The

    assessment of anaerobic condition in soils and sediments is important for many disciplines such

    as ecology, ecotoxicology, soil science and agronomy. Redox potential influences the

    availability of redox sensitive nutrients, their removal and/or translocation in soil profiles and the

    efflux of solutes in percolating water (Chadwick and Chorover, 2001). For example, redox

    condition influences the nitrogen cycle in soils, determining the nitrogen emission as N2O, N2 or

    NH3, or the plant uptake as NH+

    4 + or NO3-, with obvious consequence for agriculture and water

    quality. Regarding the carbon cycle and global warming, methane, a strong green house gas, is

    produced in wetlands depending on redox condition. Most agricultural systems rely on aerated

    soil so that roots respiration can occur. Reducing conditions can produce toxic compounds for

    plants (reduced forms of Fe and Mn, cyanogenic compounds, ethanol, lactic acid, acetaldehyde

    and aliphatic acids such as formic, acetic, butyric acids, and H2S) impairing their growth and the

    crops yield. Again, the fate of toxic compounds in the environment depends on redox potential

    since the mobility of heavy metals, and that of non-metal like arsenic as well, is redox-sensitive.

    Moreover, the degradation of organic pollutants and pesticides in groundwater are affected by

    oxidising/reducing condition.

    The U.S. National Academy of Sciences Definition states that the minimum essential

    characteristics of a wetland are the recurrent, sustained inundation or saturation at or near the

    surface and the presence of physical, chemical, and biological features reflective of recurrent,

    sustained inundation or saturation. (Gosselink and Mitsch, 2007). In peatlands, where water

    saturation and anaerobic conditions are essential characteristics, it is possible to use an electrode

    to measure oxygen penetration, in order to characterise the redox conditions and to relate them to

    the dominant redox processes (Pezeshki, 2001). When Eh time series is measured, and the

  • 22

    relative frequency of redox potential values taken at certain depths is calculated, is possible to

    determinate to which redox zones the peat layer corresponds. For instance, Fiedler and Sommer

    (2004) related the frequency of Eh to diagnostic features in hydric soil, finding that the

    thresholds for Mn reduction was Eh < 450 mV, for Fe(III) reduction was Eh < 170 mV and for

    CH4 oxidation was Eh > 75 mV. Measured Eh must be seen as an integrated operational

    parameter, which is influenced, as stated, by the activity of living microorganisms but also by

    external factors such as variation of water table, precipitation, source of water and chemistry,

    pH, temperature and organic matter (Fielder et al., 2007). So far, a number of problems have

    hindered the broader application of Eh direct measurement. First, the employment of platinum

    electrode probe in in-situ condition requires rugged electrode device to tackle harsh

    environmental conditions like storms, prolonged submersion, extreme temperatures and animal

    disturbance. Second, irreversibility of coating reactions at the electrode, slow reaction kinetics

    and mixed potentials can complicate the interpretation of redox measurements (Peiffer et al.,

    1992; Stumm and Morgan, 1981). Last but not least, redox potential shows high spatial and

    temporal variability. To conclude, on one hand unpredictable incidents or technical problems

    could prevent the collection of reliable data and, on the other hand, the high number of

    interplaying variables might prevent simple interpretation of redox potential (Husson, 2013). To

    date these issues have slowed the measurement of redox potential in wetlands, leading to a

    scarcity of studies (De Mars and Wassen, 1999). Frequently, when a study on redox potential

    was performed, it was designed without permanent continuous measurements or without vertical

    profiles measurements (Shoemaker et al., 2013; Fiedler et al., 2007). The common use of single

    time measurement with intervals in the order of days or weeks cannot account for the variability

    expressed by wetland systems. Indeed, in wetland soil redox potential may have periodical or

    occasional fluctuations within time of hours and at different depths (Vorenhout et al., 2004).

    Moreover, most of the available studies have focused on minerogenous systems with an even

    greater lack of publications on redox measurements in ombrogenous bogs.

    In the next section, the factors that control redox potential in peatlands are examined, in

    order to clarify, where possible, the nature of the relation, the relative importance and the

    potential feedbacks.

    1.7 Redox potential is a complex indicator

    Redox potential in peatland depends on several factors, namely: peat property and quality,

    temperature, pH, microbial community, type of vegetation, hydrology and water chemistry

  • 23

    (figure 2.2). Since redox potential is an indicator of several processes, the aim here is to

    understand how each factor can influence it. The method is based on a review of previous

    researches in order to trace emerging trends, which will be discussed in the light of the theory

    discussed above and the peculiar characteristic of bog peatlands. In this section, each factor will

    be analysed separately for the sake of clarity, though it must be stressed that, in nature, they are

    not isolated but they closely interact.

    Eh

    Hydrology

    Water table

    Precipitation

    Groundwater

    Microbial community

    Temperature

    pH

    Peat properties and

    peat quality

    Vegetation

    Water chemistry

    Figure 1.6 Controls of redox potential.

    Redox potential has a highly dynamic nature and is spatially dependent. The discussion

    starts with an evaluation of spatial and temporal variability in the measurement of redox

    potential.

    Spatial and temporal variability of Eh

    In typical wetland soils, Eh values vary from 300 to 700 mV (Reddy and DeLaune, 2008).

    According to the work of Fiedler et al. (2007), the range of temporal variation of redox potential

    in wetlands can be identified in: short-term, diurnal, single event, seasonal and annual variations.

    Periodic diurnal fluctuations can be explained in term of Nernsts equation and the vant Hoffs

    law. In this case, the fluctuations are driven by temperature with a temperature maximum

    followed by Eh minimum. Plant living roots have daily metabolic cycles. In saturated anaerobic

    soils, if aerenchymatous roots are enough dense, redox potential may fluctuate following the

  • 24

    release of oxygen that has its peak during the highest photosynthetic rate (i.e. light intensity).

    Periodical redox variation can also be driven by periodic physical phenomenon. For example,

    Catallo et al. (1999) studied the effect of a 12-hours tide fluctuation in a salt marsh and found

    that it corresponded to redox potential fluctuation of about 40-300 mV. In Fiedler et al. (2007),

    drying and rewetting are considered single event changes that may bring redox potential to vary

    up to 900 mV. Seasonal changes can be related to gradients in soil temperature that affects

    microbial activity. Short-term changes are the most difficult to characterise. They can be caused

    by soil chemistry dynamics that may result from the production of carbon dioxide, the release of

    hydroxyls in ferric iron reduction, the chemistry of precipitation and from inputs of redox

    sensitive species. Measurement of redox potential often showed short-term peaks or daily

    variation, though so far, there have not been studies that addressed and explained the underlying

    causes.

    Redox potential can vary with respect to centimetres or even millimetres in soils. This

    variability is induced by partial water saturation of soil structure, absorption on soil minerals and

    organic matters, plant roots and microbial activity (Yang et al., 2006). Each of these factors

    varies spatially and the soil heterogeneity may cause the electrodes positioned at the same site

    and depth to show simultaneously a wide range of Eh values (Urquhart et al, 1972). However,

    De Mars and Wassen (1999) investigated the role of heterogeneity in different peatlands and,

    placing from 4 to 12 electrode replicates within an area of 10 m2 and at 15 cm below surface,

    showed that the variability of Eh among the replicates was low, even at low water level. Yang et

    al. (2006) measured that spatial redox variability is higher vertically than horizontally, which

    seems reasonable given that peat properties that are reported to vary more vertically than

    horizontally (Schaaf, 1999). In conclusion, the degree and scale of heterogeneity will affect the

    variability of redox potential, though in permanently saturated soils or in areas with minimal

    hydrological fluctuations the variability could be minimal (Reddy and DeLaune, 2008).

    In wetlands, redox potential generally follows a depth gradient. The vertical profile is

    created by local hydrology and addition of electron acceptors and donors and typically decreases

    with depth due to oxygen intrusion from the surface (Reddy and DeLaune, 2008). Nonetheless,

    other studies showed quite different redox profile patterns. For instance, Vorenhout et al. (2011)

    observed that redox potential at deeper soil layers occasionally increased and exceeded the redox

    potential at upper layers and that some sensors located at greater depth showed steadily higher

    redox potential than overlying layers. Urquhart et al. (1972) observed in a bog a similar pattern,

    where some of the redox potential profiles (0 - 30 cm) showed deeper layers having higher Eh

    than overlying layers. It follows that redox profile in peatlands can be dynamic and that a simple

  • 25

    decreasing gradient is not the general pattern. Hydrology and water circulation are important

    factors that can influence the vertical profile and can disrupt the vertical redox stratification that

    might be predicted by oxygen intrusion alone (Thompson et al., 2009). However, as mentioned

    before, there have not yet been direct field investigations of these eventual relations in wetlands.

    Finally, a bogs surface is characterised by micro-topography that adds complexity and affects

    the variability of redox potential (Haraguchi, 1992).

    Water table

    Water table level controls the diffusion of oxygen in peatlands and several studies have found a

    good correlation between oxygen concentration and redox potential at low (6-15%) O2 contents

    (Callebaut et al., 1982). Almost all the literature examined in the present work involves research

    carried out in wetlands other than bogs, highlighting a general lack of studies on bogs. Generally,

    redox potential decreases when water table increases so that they are negatively correlated

    (Niedermeier and Robinson, 2007; De Mars and Wassen, 1999) and the change of redox

    potential trails shortly the drop of water table (Seybold et al., 2002). In a mineral Calcaric

    Gleysol soil, Mansfeldt et al. (2003) showed that the principal variable explaining temporal and

    spatial variation of redox potential through a vertical profile was water table (r = - 0.97). De

    Mars and Wassen (1999) studied different peatlands and showed that the temporal variation

    explained by water table level was 2/3 of the total variance. The authors suggested that soil

    conditions like physical properties (capillarity), pH and nutritional status (quality of organic

    matter) might have explained the residual variance. Rezanezhad et al. (2014) set up two columns

    filled with homogenised riparian soil: in one column they imposed a fluctuating water table

    regime whereas in the other the water table was kept stable. Redox potential was measured at

    10 cm and - 30 cm every 60 seconds for 75 days and they clearly showed that the imposed

    regime controlled the spatial and temporal distribution of the soil redox potential. The authors

    observed also short-term spikes during high water table extending for hundreds of mV and

    probably caused by gas transport and heterogeneity of water composition. Interestingly, these

    spikes did not show up in the column with stable water table, so that soil water circulation could

    have been responsible of their generation.

    While the degree to which water table level influences redox potential is depth

    dependent, the major part of studies on redox potential in wetland soils did not carry out

    measurements with depth resolution (Shoemaker et al., 2013; Fiedler et al., 2007). The oxygen

    concentration in soil depends on diffusion rate and consumption rate by microbial activity. The

    process of diffusion as described by Ficks law is driven by the concentration of oxygen at

  • 26

    interfaces, diffusion path length and diffusion coefficient. Water table influences indirectly redox

    potential lowering the rate of diffusion of O2 by a factor of 10000 in respect to the gaseous

    phase (Gosselink and Mitsch, 2007). Soil properties, like porosity, are also important to develop

    the degree of soil aeration; in fact, a strong capillarity can support saturation above the water

    table level (Knorr et al., 2009; Thompson et al., 2009; Thompson et al., 2007). Previous studies

    have measured that oxygen penetration length in peat can be very short; for example Armstrong

    and Boatman (1967) found that oxygen, even in extreme cases, would not diffuse further than 6

    centimetres. Benstead and Lloyd (1995) measured oxygen in different hollows in a Sphagnum

    bog, when water table was above and below 2 cm from the surface, and showed that, at 2 cm

    depth, almost all hollows lacked O2; only in one site, the authors detected oxygen at depth of 5

    centimetres. These studies suggest that water table efficiently prevents oxygen to penetrate in

    peat, though local peat properties and microbial activity can still be important to determine the

    length of the vertical extinction curve of oxygen. It is worth adding that during summer the water

    table drop might not entail oxygen penetration to deeper layers if oxygen in the upper layer has

    been consumed because of high microbiological activity (Barber et al., 2004).

    As mentioned before, many studies have measured decreasing redox potential with

    increasing depth. Fiedler et al. (2004) found decreasing redox potential in a wetland mineral soil,

    while Thomas et al. (2009) found a decreasing gradient in Florida Everglades wetlands. The

    relative importance of sources of water shapes the vertical redox profile: for instance, lateral

    water flowing from uplands may be responsible of changes of Eh (Wheeler and Shaw, 1995

    cited in Thompson et al., 2009) and upwelling of groundwater (when occurs) was suggested to

    be responsible of local increasing of Eh in deeper layers. Considering redox profile again, when

    two or more redox probes were employed, the measurements have shown that spatial and

    temporal variability of Eh decreased with depth (Thomas et al., 2009; Mansfeldt et al., 2003;

    Fiedler et al., 2004). However, Mansfeldt et al. (2003) found that the highest variation of redox

    potential was at 60 cm below the surface, where the soil most frequently changed between

    saturated and unsaturated condition. Therefore, standard deviation is associated to the extent of

    water table fluctuation, because it affects the relative proportion of air and water in the pores. In

    conclusion, the effects of water table on redox potential should be limited to the zone of water

    table fluctuation, or at most, should extend very shortly above or below the water level.

    Contrarily, below that zone, other factors should play a major role.

    Precipitation

  • 27

    Rainfall is a source of dissolved oxygen and oxidised nutrients, like nitrate and sulphate, which

    may temporally increase redox potential in surface peat layers. Rainfall also raises water table

    and if rainfall rate exceeds storage capacity, runoff water is produced. The effect of rainfall on

    oxidation-reduction cycles may be important in peatland and can be greater during dry periods,

    because rain can reach deeper anoxic layers (Deppe et al. 2009). Niedermeier and Robinson

    (2006) measured redox potential in a fen during two summer rain events, when water table was

    below 20 cm. The authors observed a sharp increase of Eh at 10 cm depth during the first rain

    (45 mm d-1) and a broad peak in Eh at 10 cm during the second rain (70 mm d-1). The peaks were

    about 400-500 mV. There were no effects at 30 cm depth, below the water table. Precipitation

    also raised water table lowering the rate of oxygen diffusion and counteracting the former

    oxidising effect of rainwater (Niedermeier and Robinson, 2006). The influence of rain is not

    always outstanding, in fact Mitchell et al. (2005) has observed just little Eh increment during

    rainfall events in the surface of a peatland, and no influence at layers deeper than 15 centimetres.

    Contrarily, in the first 50 cm of peat, Haavisto (1974) has measured an averaged decrease of 47

    mV, although the procedure used may have produced unreliable values. Fiedler (1999) observed

    that at depths > 30 cm, precipitation influenced the potentials only indirectly by raising the water

    table.

    Water chemistry

    Peatland bog soils are electron acceptors limited, but have plenty of electron donors (i.e. organic

    matter). Addition of easy degradable organic matter can enhance the electron pressure and can

    reduce redox potential while the input of oxidised species can increase it, as explained by the

    Nernsts equation (Reddy and DeLaune, 2008). The redox potential measured at the electrode is

    characterised by the dominant redox couple, and it depends on standard rate constant for the

    redox couple, concentration of oxidised and reduced species, number of electrons transferred per

    molecule and electrode surface (Peiffer et al, 1992). Some redox couples can be present in very

    low concentration in oligotrophic bog water. For example, Fe(III)/Fe(II) has a high standard rate

    constant, so that is often dominant in mineral soils, but total dissolved Fe is very low in bogs

    (Wieder and Vitt, 2006) and the major part is bound to and stabilised by DOC (Steinmann and

    Shotyk, 1996), so that its reaction may be insignificant (Keller and Bridgham, 2007). The same

    can be said for manganese. Reduced inorganic sulphur (RIS) is much less abundant than

    organically bound S in bogs, but it is reported to have a very dynamic turnover (Wieder and

    Lang, 1988). The concentration of RIS is higher than iron and it can be more important in a bog,

    explaining up to 30 % of anaerobic respiration (Keller and Bridgham, 2007). DOC is not only an

  • 28

    electron donor but it can also act as electron acceptor for anaerobic respiration. There is an

    increasing need of assessing the use of organic electron acceptors by microorganisms in anoxic

    environment. Many studies hypothesised their involvement to account for the great fraction of

    unexplained carbon mineralization in low nutrient bogs (Knorr et al., 2009; Deppe et al., 2009).

    However, the great importance of DOC has been evaluated only in laboratory and, so far, it has

    not been studied directly in the field (Blodau et al., 2009). To complicate the issue, the DOC

    chemical composition rather than its amount distinguishes its reducing/oxidising role in

    peatlands (Keller et al., 2009).

    Microbial community

    Microorganisms control oxidation-reduction reactions in soils, modifying their environment by

    consuming TEAs and lowering redox potential, which, at the same time, determines the

    functional microbial type emergent in the bacterial community (Husson, 2013). Despite the low

    nutrient content and pH and the recalcitrance of Sphagnum (Dobrovolskaya, 2014), microbial

    community in bogs is frequently reported to be as active as other richer peatlands fens

    (Mettrop et al., 2014; Fisk et al., 2003). In bogs, microbial activity is greater in surface layers (0-

    15 cm below surface) than subsurface ones (15-30 cm below surface) (Fisk et al. 2003). In

    general, the physical-chemical and environmental factors that influence microbial activity will

    cause the redox potential to change. Studies that took into account the role of microorganisms in

    driving Eh changes in bogs are lacking, albeit an important role for them, as observed in other

    ecosystems, might be hypothesized. In a tidal marsh, Catallo et al. (1999) showed that microbial

    activity alone produced quick Eh variation to the extent of more than 370 mV in less than 48

    hours. Oligotrophy of bog ecosystems represents a constraint to microbial activity and input of

    nutrients from deposition or other sources can alter this constraint. Bragazza et al. (2012) have

    shown that N deposition could alter microbial communities and favour bacterial growth.

    Temperature can influence microbial community composition, growth rate, enzyme synthesis

    and response. As a rule of thumb, the decomposition rate of organic matter doubles every 10C

    increment, following the Arrhenius equation. However, other environmental constraints affect

    the rate of decomposition. These physical and chemical constraints are themselves affected by

    temperature and climatic factor like flooding, droughts and freezing, which have the effect to

    making the relation complex (Davidson and Janssens, 2006) and giving different Q10 for

    different ecosystems (Peng et al., 2009). Nevertheless, regardless of the extent of their effect,

    warmer conditions increase microbial activity both in oxic and anoxic environments (Estop-

    aragons and Blondau, 2012). Acidity affects the microbial community. Acidic pH means a

  • 29

    greater amount of energy spent by bacteria pumping hydrogen ions out of their cells to survive.

    Enwall et al. (2007), showed a negative correlation between soil pH and the microbial metabolic

    quotient, which indicate a decreased efficiency to convert organic carbon in microbial biomass in

    acidic soils. Notwithstanding these constraints, N-mineralisation and P-mineralisation can be

    higher in bogs than fens (Mettrop et al. 2014; Verhoeven et al. 1990 cited in Wieder and Vitt,

    2006). It is possible that bacterial groups in bogs are not constrained by pH because they evolved

    for that specific acid condition (Dobrovolskaya, 2014). Moreover, it has to be kept in mind that

    fungi are less affected by low pH than bacteria.

    Vegetation

    Plants have the potential to change redox stratification in soil through at least three distinct

    processes: (1) release of O2 through roots into the rhizosphere, (2) primary production,

    increasing the quantity of labile organic carbon and releasing roots exudates and (3) direct

    nutrients uptake into roots, rhizomes, stems and leaves. The plant species can account for

    differences in electron acceptor renewal in anaerobic soils. It is reported that in bogs this process

    is usually less important than in fens because of lack of sedges (Deppe et al., 2009). During the

    day, photosynthetic activity and active transport of oxygen create an oxic environment in the

    rhizosphere. Nikolaustz et al. (2008) observed that, in an artificial Juncus effusus wetland,

    reducing condition during night (-320 mV) and oxic condition during day (+300 mV) were a

    function of light intensity and dissolved oxygen in the rhizosphere. Instead, microbial

    consumption of electron acceptors could explain the reducing condition measured at night. The

    authors suggested that this pattern might be applicable to all aerenchymatous plants in wetlands.

    Respiration of roots produces a dynamic trend of carbon dioxide in soil. Benstead and Lloyd

    (1996) incubated a solid peat core extracted from a Sphagnum-Eriophorum bog and observed a

    diurnal fluctuation of CO2 at the depth of 15 cm and 20 cm. They found a minimum at 18:00 and

    a maximum at 7:00. The authors placed the same peat in dark and fluctuations disappeared,

    suggesting a key role of vegetation in the process. Shoemaker et al. (2012) build a mesocosmos

    with Myriophyllum acquaticum and Leersia oryzoides and found that an increase of temperature

    occurred almost simultaneously with a rise of redox potential leading to daily fluctuation of

    about 100 mV in the top 10 cm. The authors did not find the same pattern in a non-vegetated

    mesocosm, which suggested that the observed fluctuations were produced by vegetation roots

    and not by temperature.

  • 30

    Peat structure

    Bulk density is measured as the dry weight of peat per unit volume (g cm-3). Bulk density

    increases with degree of humification and increases with depth. Porosity is the fraction of soil

    volume that is not filled by the solid phase. In Sphagnum peat near the surface porosity is very

    high, ranging between 88 and 97 % (Ivanov, 1981). Porosity decreases with increasing

    humification, bulk density and degree of compaction, so it decreases with depth. Porosity

    correlates positively to hydraulic conductivity (Schaaf, 1999). Porosity can decrease during

    water table drawdown (Schlotzhauer and Price, 1999). Peat of different origins can vary in

    porosity and in bulk density influencing the capillarity and the air-filled porosity of the peatsoil.

    Sphagnum peat may have low capillarity: indeed, Deppe et al. (2009) showed that air filled

    porosity sharply increases at 1 cm above the water table in a bog peatsoil. Fen peat with

    dominant sedges instead is denser than Sphagnum peat and has higher capillarity, which can

    bring to a disconnection between oxic/anoxic boundary and water table dynamic (Knorr et al.,

    2009).

    Peat chemistry

    Peat is composed of an enormous mixture of organic compounds, including carbohydrates

    (cellulose, hemicellulose), nitrogenous compounds (proteins, amino acids), phenolics (including

    lignin), lipids (waxes, resins, steroids, terpenes) and humic substances (Rydin and Jeglum,

    2013). The amount of organic matter, C quality and nutrient availability influence redox

    potential through microbial activity. Low input of C, lower nutrient availability and less labile

    organic matter can slow the redox pathway catalysed by microbes and the release of electrons,

    diminishing the electron pressure and resulting in higher Eh (De Mars and Wassen, 1999). The

    acid soluble components (e.g. cellulose) tend to be metabolised by microorganisms more readily

    than acid insoluble components (e.g. lignin), so that the relative fraction can be used as index of

    degradability. The C:N ratio is an important indices of degradability or recalcitrance, so that

    where it is high the decomposition is low (Biester et al., 2014). The relative C:N and C fraction

    in litter depends on vegetation type. In Sphagnum bogs, where vascular plants are presents there

    is a higher input of lignin (Biester et al. 2014). C:N ratio also varies with plant species, for

    example in Sphagnum spp. it is generally around 50-60 but it can reach value of 300 in S.fuscum

    litter (Rydin and Jeglum, 2013). Sphagnum tissue is recalcitrant not only because it has high C:N

    ratio, but also because it has high content of sphagnum acid in the cell walls (Rudolph and

    Sampland, 1985) and other phenolic compounds that act as anti-septic (Dobrovolskaya, 2014).

    Phenolic compounds are degraded by phenolic oxidase only in presence of oxygen. Therefore, in

  • 31

    anoxic conditions phenolic compounds can inhibit the hydrolases, the main enzyme involved in

    peat decomposition (Freeman et al., 2001). New fresh organic matter is added into the acrotelm

    from the layer of living Sphagnum spp. and from vascular plants. Different vegetation grows on

    hummocks and hollows and they have also different HWT, so that litter chemical precursors and

    importance of aerobic respiration differ between the two micro-structures, influencing the quality

    and quantity of organic matter that is added to the acrotelm. In the hollows, organic matter is

    added directly to the zone of water table fluctuation, while in the hummock it has to pass through

    the aerobic body where a great part of the labile carbon is decomposed to CO2. Therefore, by the

    time that litter has reached the water fluctuation zone, the amount of labile organic matter added

    would be lower below hummocks (Nilsson and quist, 2009). Moreover, the hummocks

    S.fuscum and vascular plants (e.g. Calluna vulgaris) should produce litter with higher lignin

    fraction and higher C:N ratio than hollows, increasing its recalcitrance. Since the effect of water

    table fluctuation on decomposition depends on the intrinsic liability of that zone to sustain

    decomposition in terms of nutrients, C quality and microbial community (Artz, 2009), also redox

    potential will be affected by these factors by showing different sensitivity to water table

    fluctuation. Ombrotrophic plants evolved mechanisms for dealing with a nutrient-deficient

    environment. Sphagnum spp. and Eriophorum spp. show strong resorption and recycling of

    nutrients during tissue senescence (Wang et al., 2014; Rydin and Jeglum, 2013; Bragazza et al.,

    2003), resulting in a nutrient-depleted litter. Microbial activity and decomposition rate will be

    controlled by the limiting nutrient, however it is also true that a greater amount of nutrients

    sustains a higher microbial activity and decomposition. Nutrients exert an indirect effect on Eh

    through microbial activity. Thomas et al. (2009) found that the phosphorous gradient in the

    Everglades influenced how sensitive the redox potential was to changing water table only in the

    surface layer. Thomas et al. (2009) found that the higher the P content was the lower was the Eh.

    So far, to the writers knowledge, there have not been studies that addressed in detail the

    difference in nutrients between micro-topographic structures. Regarding the pore-water

    chemistry, there is some interesting data in the work of Bragazza et al. (2005), who studied

    Ryggmossen bog, and found higher K+ and slightly higher orthophosphate, nitrate and ammonia

    in hummocks. However, the authors sampled water using wells rather than rhizons. Other studies

    found higher nitrate in hummocks (Frei et al., 2012; Wolf et al., 2011). Bragazza et al. (2005)

    also analysed peat samples and found that total P was two times higher in hummocks, total

    nitrogen was lower and total K was higher. Along the burial process of peat within Sphagnum

    bogs, the variation of litter fraction and C:N depends more on decomposition pattern (climate,

    hydrology) than plant species (Biester et al., 2014). In the acrotelm the C:N ratio usually

  • 32

    decreases with depth (Wang et al., 2014), while in more decomposed and deeper layer it

    increases (Silamikele et al., 2007). As decomposition proceeds the organic matter becomes

    enriched in lignin and N, with lignin suppressing decomposition rate and raised N-level

    suppressing lignin degradation (Berg and Meentemeyer, 2002). As mentioned before, the

    absence of signicant external P input and relatively elevated atmospheric N deposition in

    ombrotrophic systems requires efcient internal cycling of P so that the usual increase of C:P

    and N:P ratio with depth might be due to preference of P uptake (Wang et al., 2014). Another

    factor of recalcitrance varies with depth, Beer et al. (2008) observed in a bog that aromatic and

    phenolic functional groups of organic matter increased with depth, suggesting that degradability

    decreased accordingly. The humic fraction of peat is important in affecting redox potential and

    electron transfer to humic substances in anoxic systems is considered to competitively suppress

    reduction of other terminal electron acceptors (TEAs), including CO2 under methanogenic

    conditions (Klpfel et al., 2014). Gondar et al. (2005) measured that the extractable fraction of

    humic acid (HA) and fulvic acid (FA) was 10 % of the TOM in the fibric layer and 2 % in the

    sapric. Peat has high cations exchange capacity due to humified organic matters (Rydin and

    Jeglum, 2013). Most cations can be bound to organic matter diminishing the availability to

    microorganisms. For example, Fe(III) can be stabilised by complexation to DOC and can still be

    found in anoxic and reducing environment (Steinmann and Shotyk, 1996).

    Temperature

    Nernsts equation describes how temperature directly affects redox potential, but that the direct

    effect of temperature may not be relevant (Shoemaker et al., 2012). Temperature exerts a more

    important indirect effect triggering microbial activity (Fiedler et al., 2007). A more subtle

    indirect effect occurs through the influence on ionic activity, though in this case a shift from 3 to

    22 C produces only a shift of around 25-30 mV (de Mars and Wissen, 1999). It is possible that

    temperature effects are more important in explaining variance of Eh in anaerobic layer than in

    surface layer because the lack of oxygen in the former does not influence redox potential.

    Urquhart et al. (1972) studying four different bogs found significant negative relation between

    Eh and temperature at the depth of 30 cm. The authors argued that it was a consequence of

    microbial activity. The higher sensibility to temperature of decomposition rate of deep

    recalcitrant peat than surface peat (Hilasvuori et al., 2013) could corroborate the importance of

    temperature effect on redox potential in anaerobic peat. The heat wave varies in different soil

    ecosystem, and, for example, Barber et al (2004) did not find seasonal variation (i.e.

    temperature) of redox at depth of 40 cm in a wet grassland.

  • 33

    pH

    Bogs are acid environments and the main source of H+ stems from humic substances and

    Sphagnum phenolics. In surface layers, humic substances mainly supply hydrogen ions while at

    deeper layers carbonic acid becomes the main source of hydrogen ions (Steinmann and Shotyk,

    1996). Humic acids have a more acidic dissociation constant than carbonic acid and for this

    reason pH is reported to increase with depth in bogs (Deppe et al., 2009; Steinmann and Shotyk,

    1996; Lundin and Berquist, 1990). The Eh responds inversely to change in pH according to the