a generalized damkohler number for classifying material ......a generalized damkohler number for...

16
Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013 www.hydrol-earth-syst-sci.net/17/1133/2013/ doi:10.5194/hess-17-1133-2013 © Author(s) 2013. CC Attribution 3.0 License. Hydrology and Earth System Sciences Open Access A generalized Damk ¨ ohler number for classifying material processing in hydrological systems C. E. Oldham 1 , D. E. Farrow 2 , and S. Peiffer 3 1 School of Environmental Systems Engineering, The University of Western Australia, 35 Stirling Hwy, 6009 Crawley, Western Australia 2 Mathematics and Statistics, Murdoch University, 6150 Murdoch, Western Australia 3 Department of Hydrology, University of Bayreuth, Bayreuth, Germany Correspondence to: C. E. Oldham ([email protected]), D. E. Farrow ([email protected]), and S. Peiffer ([email protected]) Received: 15 August 2012 – Published in Hydrol. Earth Syst. Sci. Discuss.: 18 September 2012 Revised: 1 February 2013 – Accepted: 17 February 2013 – Published: 15 March 2013 Abstract. Assessing the potential for transfer of pollutants and nutrients across catchments is of primary importance under changing land use and climate. Over the past decade the connectivity/disconnectivity dynamic of a catchment has been related to its potential to export material; however, we continue to use multiple definitions of connectivity, and most have focused strongly on physical (hydrological or hy- draulic) connectivity. In contrast, this paper constantly fo- cuses on the dynamic balance between transport and mate- rial transformation, and defines material connectivity as the effective transfer of material between elements of the hydro- logical cycle. The concept of exposure timescales is devel- oped and used to define three distinct regimes: (i) which is hydrologically connected and transport is dominated by ad- vection; (ii) which is hydrologically connected and transport is dominated by diffusion; and (iii) which is materially iso- lated. The ratio of exposure timescales to material process- ing timescales is presented as the non-dimensional number, N E , where N E is reaction-specific and allows estimation of relevant spatial scales over which the reactions of interest take place. Case studies within each regime provide exam- ples of how N E can be used to characterise systems accord- ing to their sensitivity to shifts in hydrology and gain insight into the biogeochemical processes that are signficant under the specified conditions. Finally, we explore the implications of the N E framework for improved water management, and for our understanding of biodiversity, resilience and chemical competitiveness under specified conditions. 1 Introduction Assessing the potential for export of nutrients and pollutants from catchments is of primary importance under changing land use and climatic conditions (Pringle, 2003a), yet our ability to predict export relies on a detailed understanding of both the transport and biogeochemical processing taking place (Hornberger et al., 2001). One of the challenges for the prediction of material transfer through catchments is the high degree of heterogeneity (of source distributions, trans- port media and biogeochemical processes), and its depen- dence on scale. Dooge (1986) suggested that scaling issues were fundamental to the development of hydrological the- ory, specifically the development of scaling relations across catchment scales and up-scaling from small-scale processes (Battiato and Tartakovsky, 2011). Since that time there has been substantial research on small-scale process-based dis- tributed model descriptions, however the application of these models to large catchments has at times been problematic and the preferred modes of investigation continue to vary across scales (Table 1). In response to this, McDonnell et al. (2007) challenged hydrologists to move “beyond heterogeneity and process complexity” to better facilitate the characterisation of catch- ments; however, approaches using either an ecological or hydrological lens tend to remain embedded in discipline- specific process complexity. In a study that investigated in- teractions between vegetation and hydrology, Thompson et al. (2011b) highlighted the co-evolution of vegetation and Published by Copernicus Publications on behalf of the European Geosciences Union.

Upload: others

Post on 09-Dec-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013www.hydrol-earth-syst-sci.net/17/1133/2013/doi:10.5194/hess-17-1133-2013© Author(s) 2013. CC Attribution 3.0 License.

EGU Journal Logos (RGB)

Advances in Geosciences

Open A

ccess

Natural Hazards and Earth System

Sciences

Open A

ccess

Annales Geophysicae

Open A

ccess

Nonlinear Processes in Geophysics

Open A

ccess

Atmospheric Chemistry

and Physics

Open A

ccess

Atmospheric Chemistry

and Physics

Open A

ccess

Discussions

Atmospheric Measurement

Techniques

Open A

ccess

Atmospheric Measurement

Techniques

Open A

ccess

Discussions

Biogeosciences

Open A

ccess

Open A

ccess

BiogeosciencesDiscussions

Climate of the Past

Open A

ccess

Open A

ccess

Climate of the Past

Discussions

Earth System Dynamics

Open A

ccess

Open A

ccess

Earth System Dynamics

Discussions

GeoscientificInstrumentation

Methods andData Systems

Open A

ccess

GeoscientificInstrumentation

Methods andData Systems

Open A

ccess

Discussions

GeoscientificModel Development

Open A

ccess

Open A

ccess

GeoscientificModel Development

Discussions

Hydrology and Earth System

SciencesO

pen Access

Hydrology and Earth System

Sciences

Open A

ccess

Discussions

Ocean Science

Open A

ccess

Open A

ccess

Ocean ScienceDiscussions

Solid Earth

Open A

ccess

Open A

ccess

Solid EarthDiscussions

The Cryosphere

Open A

ccess

Open A

ccess

The CryosphereDiscussions

Natural Hazards and Earth System

Sciences

Open A

ccess

Discussions

A generalized Damkohler number for classifying materialprocessing in hydrological systems

C. E. Oldham1, D. E. Farrow2, and S. Peiffer3

1School of Environmental Systems Engineering, The University of Western Australia, 35 Stirling Hwy,6009 Crawley, Western Australia2Mathematics and Statistics, Murdoch University, 6150 Murdoch, Western Australia3Department of Hydrology, University of Bayreuth, Bayreuth, Germany

Correspondence to:C. E. Oldham ([email protected]), D. E. Farrow ([email protected]),and S. Peiffer ([email protected])

Received: 15 August 2012 – Published in Hydrol. Earth Syst. Sci. Discuss.: 18 September 2012Revised: 1 February 2013 – Accepted: 17 February 2013 – Published: 15 March 2013

Abstract. Assessing the potential for transfer of pollutantsand nutrients across catchments is of primary importanceunder changing land use and climate. Over the past decadethe connectivity/disconnectivity dynamic of a catchment hasbeen related to its potential to export material; however,we continue to use multiple definitions of connectivity, andmost have focused strongly on physical (hydrological or hy-draulic) connectivity. In contrast, this paper constantly fo-cuses on the dynamic balance between transport and mate-rial transformation, and defines material connectivity as theeffective transfer of material between elements of the hydro-logical cycle. The concept of exposure timescales is devel-oped and used to define three distinct regimes: (i) which ishydrologically connected and transport is dominated by ad-vection; (ii) which is hydrologically connected and transportis dominated by diffusion; and (iii) which is materially iso-lated. The ratio of exposure timescales to material process-ing timescales is presented as the non-dimensional number,NE, whereNE is reaction-specific and allows estimation ofrelevant spatial scales over which the reactions of interesttake place. Case studies within each regime provide exam-ples of howNE can be used to characterise systems accord-ing to their sensitivity to shifts in hydrology and gain insightinto the biogeochemical processes that are signficant underthe specified conditions. Finally, we explore the implicationsof theNE framework for improved water management, andfor our understanding of biodiversity, resilience and chemicalcompetitiveness under specified conditions.

1 Introduction

Assessing the potential for export of nutrients and pollutantsfrom catchments is of primary importance under changingland use and climatic conditions (Pringle, 2003a), yet ourability to predict export relies on a detailed understandingof both the transport and biogeochemical processing takingplace (Hornberger et al., 2001). One of the challenges forthe prediction of material transfer through catchments is thehigh degree of heterogeneity (of source distributions, trans-port media and biogeochemical processes), and its depen-dence on scale. Dooge (1986) suggested that scaling issueswere fundamental to the development of hydrological the-ory, specifically the development of scaling relations acrosscatchment scales and up-scaling from small-scale processes(Battiato and Tartakovsky, 2011). Since that time there hasbeen substantial research on small-scale process-based dis-tributed model descriptions, however the application of thesemodels to large catchments has at times been problematicand the preferred modes of investigation continue to varyacross scales (Table 1).

In response to this, McDonnell et al. (2007) challengedhydrologists to move “beyond heterogeneity and processcomplexity” to better facilitate the characterisation of catch-ments; however, approaches using either an ecological orhydrological lens tend to remain embedded in discipline-specific process complexity. In a study that investigated in-teractions between vegetation and hydrology, Thompson etal. (2011b) highlighted the co-evolution of vegetation and

Published by Copernicus Publications on behalf of the European Geosciences Union.

Page 2: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

1134 C. E. Oldham et al.: A generalized Damkohler number for classifying material processing

Table 1.Different preferred modes of investigation used for different scales of elements in hydrological systems.

Soil pore scale Superficial Vegetation scale Riparian scale Catchment scaleaquifer scale

Vertical length scale O (mm–cm) O (m–10 m) O (10 cm–1 m) O (10 cm–1 m) O (1–10 m)Horizontal length scale O (mm–cm) O (10 m–100 km) O (10 cm–1 m) O (10–100 m) O (10–1000 km)Timescale O (min) O (days–decades) O (min–seasons) O (h–yr) O (days–decades)Typical modes of investigation Laboratory/numerical Numerical/field Laboratory/field Field/numerical Numerical/field

hydrological spatial organization, yet continued to use a hy-drological lens, via a water balance approach, to characterisesystems across scales. An alternative approach is to adopt adual lens, in which the focus remains on the dynamic bal-ance between hydrological and ecological processes. A sin-gle lens approach assumes the ecosystem responds to hydro-logical forcing; a dual lens approach assumes the ecosystemresponds to shifts in the balance between hydrological andecological processes. The range of ecological processes ofrelevance to catchment studies is vast; hereon we focus onmaterial fate as expressed by biogeochemical reactions.

In an attempt to identify parameters that resolved the“scale” issue, Vache and McDonnell (2006) flagged that wa-ter residence times were integrative and meaningful across allscales, and went on to highlight the challenges in character-ising flow paths, and therefore residence times, that were ofrelevance to material transport; the residence time is a criticalparameter within the dual lens approach. The residence timedistribution (RTD) has been used as fundamental character-istic of aquatic systems with free surfaces (Carleton, 2002;Werner and Kadlec, 2000) and more recently of catchments(Beven, 2001; McDonnell et al., 2010; McGuire and Mc-Donnell, 2006). Botter et al. (2011) highlighted the differ-ence between the probability density function (pdf) of wa-ter parcel travel/transit time within a catchment, and the pdfof catchment residence times. They also noted the temporalvariability of these pdfs as a function of meteorological andhydrological forcing (Botter et al., 2010; Hrachowitz et al.,2010). All of these authors noted the importance of the resi-dence and transit times, or their distributions, for the exportof pollutants and nutrients from catchments, and recently,Basu et al. (2011a,b) explored the relationships between pdfsand pollutant and nutrient export from catchments, whereasThompson et al. (2011a) used ratios of the coefficients ofvariation of chemical concentration and discharge to classifychemical export from catchments.

Estimates of characteristic residence times have been usedin dimensionless ratios (Carleton, 2002; Hornberger et al.,2001; Ocampo et al., 2006) to improve our understanding ofcatchment hydro-chemical responses and chemical cycling.The dimensionless Damkohler number,Da, has been usedextensively in the discipline of chemical engineering to clas-sify a system according to the balance between transport andreaction, and is given as

Da =τT

τR(1)

whereτT is the transport (or residence) timescale andτR isthe reaction timescale. Thus,Da can be considered an impor-tant non-dimensional number within the dual lens approach.

Michalak and Kitanidis (2000) usedDa to characterise theappearance of chemical sorption peaks in simulation exper-iments. Wehrer and Totsche (2003) usedDa to characterisewhen breakthrough of contaminants took place under non-equilibrium conditions, Ocampo et al. (2006) explored theremoval of nitrate by riparian zones and demonstrated that50 % nitrate removal occurred whenDa< 1 and increased tonearly 100 % atDa= 2–20. Kurtz and Peiffer (2011) demon-strated that surface complexation needs to be consideredto explain the increase of FeOOH turnover rates, by dis-solved S (II), with increasingDa numbers. Battiatoc andTartakovsky (2011) usedDa in combination with multiplescale expansions to upscale from micro (pore) scale reaction-diffusion-advection processes to derive macro (averaged)governing equations.

The residence time distributions, as explored by Botter etal. (2011), create a Damkohler number distribution (DND)for a given aquatic system and Carleton (2002) exploredthe relationship between RTD and the DND in a wetlandtreatment system. In an exploration of hydrological versusbiogeochemical controls, Basu et al. (2011b) used ratios oftransport to reaction timescales to characterise nutrient andpollutant export from agricultural catchments. While Basu etal. (2011b) did not specifically define their regimes in termsof Da, their approach is similar. Given that theDa and theDND have been successfully used to characterise the transferof material under a range of transport conditions and acrossa range of spatial scales from pores to wetlands and riparianzones, we suggest thatDa could be used to characterise hy-drological elements in a catchment. We note, however, thatby definition the use ofDa assumes hydrological (or hy-draulic) connectivity.

The hydrological connectivity/disconnectivity dynamic isa key characteristic of many catchments (Meerkerk et al.,2009; Michaelides and Chappell, 2009; Ocampo et al., 2006;Pringle, 2003a,b) and is considered critical for ecologi-cal function, adaptation and evolution (Klein et al., 2009;Opperman et al., 2010; Pringle, 2003b). Over the pastdecade there have been numerous attempts to characterise

Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013 www.hydrol-earth-syst-sci.net/17/1133/2013/

Page 3: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

C. E. Oldham et al.: A generalized Damkohler number for classifying material processing 1135

the connectivity/disonnectivity dynamic (Ali and Roy, 2010;Meerkerk et al., 2009; Michaelides and Chappell, 2009;Tetzlaff et al., 2007). Pringle (2003b) noted the variety ofdefinitions of connectivity, used across and within disci-plines. In recent years, a commonly used definition is theability of energy, matter and organisms to transfer withinand between elements (for example sub-catchments, ripar-ian zones, wetlands, streams) of the hydrologic cycle (Dettyand McGuire, 2010; Pringle, 2003b). We note the historicaltendency to define connectivity in terms of water movement,i.e. hydrological or hydraulic connectivity, however underthis definition quantifying the threshold of hydrological dis-connectivity remains problematic. Are hydrological elementsthat are partially or fully saturated (with water) consideredconnected? Does hydrological connectivity imply transportbetween hydrological elements? If so, is there a minimumtransit time between hydrological elements, below which theelements are considered disconnected? Hydrological connec-tivity, within these contexts and across different scales, isthus operationally defined.

Additional complexity arises when we are interested in thefate of reactive material being transferred across hydrolog-ically connected landscapes. To our knowledge there havebeen no attempts in the hydrological literature to charac-terise connectivity in terms of theeffectivenessof transfer-ring material. Such a characteristic might be termed “mate-rial connectivity” and must take into account both the hydro-logical/hydraulic transport and the biological and/or chem-ical fate or processing; such a definition is needed under adual lens approach. Multiple elements of the hydrological cy-cle may be fully connected hydrologically, yet if material isbeing removed, or settles under gravity, the material trans-fer between elements is not effective, so by our definitionabove, the material connectivity between those elements ispoor. The effectiveness of material connectivity is a criticalparameter for ecological processes and outcomes.

To illustrate our point we present two examples. Firstly, wetake the example of Fe(II)-rich anoxic groundwaters beingdischarged into a well-oxygenated stream which then flowsinto a lake. The groundwater, stream and lake are hydro-logically and hydraulically fully connected; however, Fe(II)undergoes oxidation on exposure to oxygenated waters, andlikely precipitates as Fe(III)-oxide, which may affect the ben-thic ecology. The balance between the rate of transport, therate of Fe(II) oxidation, the rate of precipitation and set-tling, and the rate of Fe(II) production from reductive disso-lution will determine the effectiveness of material connectiv-ity, with respect to Fe(II), between the groundwater and lake.If all Fe(II) is oxidized before reaching the lake, the systemsare materially disconnected with respect to Fe(II). Under thisscenario, the reaction timescale (oxidation, precipitation andsettling) is much shorter that the transport timescale,Da� 1,ensuring material disconnectivity even under hydrologicallyconnected conditions. Secondly, we take the example of aphytoplankton-rich wetland under flooding conditions, being

flushed out across a floodplain and into a lake. The wetland,the floodplain and the lake are hydraulically connected, how-ever in transit the phytoplankton may grow, aggregate, settleand decompose. The balance between the rate of transport,the rate of particle settling under gravity and the phytoplank-ton net productivity will determine the extent of material con-nectivity, with respect to the phytoplankton, between the wet-land and the lake. If all phytoplankton settle out (or decom-pose) before reaching the lake, the lake and wetland are ma-terially disconnected with respect to phytoplankton. Underthis scenario, the reaction timescale (settling and decompo-sition) is much shorter that the transport timescale,Da� 1,ensuring material disconnectivity even under hydraulicallyconnected conditions.

We note that in the two examples provided above, hy-drological or hydraulic connectivity is preserved,Da is de-fined and therefore material connectivity can be estimated.When systems are no longer hydrological/hydraulically con-nected,Da is no longer defined, yet intermittent hydrologi-cal connectivity/disconnectivity characterises many ecosys-tems. This paper explores the extension of theDa frameworkto conceptually characterise hydrological systems, includ-ing systems undergoing intermittent connectivity, and thusto provide insights into processes controlling the export ofmaterial from catchments. To populate this framework, weprovide examples where a standard Da approach may beused (i.e. under hydrological connectivity) but use the non-dimensional Peclet number to determine appropriate lengthand timescales. We also provide examples of disconnectedsystems where a new non-dimensional number must be de-fined. For clarity, we modify the broader connectivity defi-nition of Pringle (2003b) to a definition of material connec-tivity as the ability of material to transfer between elementsof the hydrologic cycle, while subject to biogeochemical orbiological processing.

2 Conceptual framework

2.1 Governing equations

We have defined material connectivity as the ability to trans-fer matter between hydrological elements. We therefore be-gin the development of our framework with the scalar trans-port equation, or advection-diffusion-reaction (ADR) equa-tion, that describes transport of chemicalB via a homoge-neous fluid (constant density and viscosity) which is given as

∂[B]

∂t= −ui

∂[B]

∂xi+Deff

∂2[B]

∂xi ∂xi−k′

[B]

I II III(2)

where [B] is the concentration of chemicalB (mmol m−3),ui is the mean velocity in thei-direction (m s−1), t is time (s),xi is distance in thei-direction (m),Deff is theeffectivediffu-sion/dispersion coefficient (m2 s−1) andk′ is an appropriate

www.hydrol-earth-syst-sci.net/17/1133/2013/ Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013

Page 4: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

1136 C. E. Oldham et al.: A generalized Damkohler number for classifying material processing

rate constant (s−1). Outside of diffusive boundary layers,Deff is likely to be significantly larger than molecular dif-fusivity and is a simplified but commonly used descriptionof stochastic processes such as turbulent diffusion, bioturba-tion or mechanical dispersion (Bolster et al., 2012; Cirpkaand Attinger, 2003). Term I describes advective transport ofchemicalB, Term II describes Fickian transport of chemi-calB, and Term III describes the rate at which chemicalB isconsumed.

We apply Eq. (2) to an idealized one-dimensional controlvolume or hydrological element that has a physically con-strained horizontal length scale of1x and a through-flowvelocityux . The general solution to Eq. (2) is then

[B] = exp(−k′ t)

[α0 +

∞∑n=0

(αn cos

(nπ

1x(x − ux t)

)+βn sin

(nπ

1x(x − ux t)

))exp

(−

n2π2Deff

1x2t

)](3)

wheren is the summation index andαn andβn are Fouriercosine and sine coefficients, respectively, which are deter-mined by the initial and boundary conditions that apply tothe particular element being modelled.

2.2 Timescales

2.2.1 Exposure timescales

Most authors have defined connectivity as a hydrological orhydraulic continuum and the transport timescales are derivedfrom this definition. However, in many environments thereare times when elements of the hydrological cycle are hy-drologically, hydraulically or materially disconnected or iso-lated. We have defined material connectivity as the abilityto transfer matter between hydrological elements; however,it is important to consider that during periods of isolation,material processing can continue within an hydrological el-ement and material will accumulate or decline, creating anisolated material reactor in the environment. One of our ob-jectives is to develop a framework that encompasses both pe-riods of connectivity and isolation, making a clear distinctionbetween hydrological versus material connectivity.

To facilitate this objective, we now conceptually define anexposure timescale,τE, as the timescale over which materialhas theopportunityto be processed. This opportunity mayoccur during transport (i.e. when elements are hydrologicallyconnected) or during hydrological disconnection. The con-cept of an exposure timescale has been previously used insurface renewal theory (Dankwerts, 1951) and chemical en-gineering (Asarita, 1967). We stress here the distinction be-tween the commonly used transport timescale (defined onlyduring periods of hydrological connection) and the exposuretimescale (defined during both hydrological connection andisolation). The exposure timescale is used to describe thetransit time between elements, or the isolation timescale of an

Fig. 1. The relationship between characteristic length,x, and time,t , scales under varying Peclet number conditions. UnderPe� 1,advective processes dominate (Regime I); underPe� 1 diffusiveprocesses dominate (Regime II); whenPe∼ O(1), the characteristiclength and timescales are set where the advection and diffusive linesintersect.

individual element. Both timescales provide opportunity forexposure. Averaging of transit and isolation timescales mayprovide an “effective” exposure timescale, however biogeo-chemical processes are frequently different under connectedversus disconnected conditions. If we are aiming to extendthe Da approach, it is critical that the relevant biogeochemi-cal processes are identified for each condition and thus tem-poral averaging across conditions may not be appropriate.

The choice of appropriate control volume determines rel-evant time and length scales as well as boundary condi-tions and thus becomes critical for application of our gov-erning equations. An initial assumption is that the controlvolume and therefore the characteristic length scale,1xi , arephysically constrained, typically by geomorphology (for ex-ample by the catchment, stream or wetland). The sizes ofhydrologically isolated systems are by definition physicallyconstrained.

For hydrologically connected systems, the relevant expo-sure timescale,τE, is then determined by the length scaleand the transport process (Fig. 1). Note that exposure lengthscales can be highly variable between individual flow pathsand require an appropriate definition of the hydrologicalproblem. In many cases it may be convenient to use a meanlength scale for a specific system. When a control volume isnot physically constrained, it may be defined operationally,depending on the objectives; an example of this approach isgiven in Sect. 2.3.

The exposure timescale for hydrologically connected ele-ments under advective conditions is then given as

τAE =

1x

ux

. (4)

Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013 www.hydrol-earth-syst-sci.net/17/1133/2013/

Page 5: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

C. E. Oldham et al.: A generalized Damkohler number for classifying material processing 1137

Fig. 2. Regime I: (a) transport and fluxes of chemicalB acrosshydrologically or hydraulically connected elements are dominatedby advection. Examples of this regime are a river under high flowconditions(b), a wetland with groundwater throughflow under alarge hydraulic gradient(c) and porous media flow under a largehydraulic gradient(d).

Examples of aquatic systems that are dominated by advectivetransport are shown in Fig. 2.

The exposure timescale for hydrologically connected ele-ments under diffusive conditions, is given as

τDE =

1x2i

Deff. (5)

Examples of aquatic systems that are dominated by diffu-sive transport are shown in Fig. 3. We note that exposuretimescales may be defined differently by stochastic transportmodels; any carefully formulated definition may be utilizedin the framework.

When hydrological elements are materially disconnectedor isolated, the exposure timescale in the isolated elementis equivalent to theisolation timescale,τ I

E. If the element

Fig. 3. Regime II: (a) transport and fluxes of chemicalB acrosshydrologically or hydraulically connected elements are dominatedby diffusion. Examples of this regime are a river under very lowflow conditions, where for example transverse transport exceedslongitudinal transport(b), a wetland at equilibrium with the localgroundwater(c) and porous media flow under minimal hydraulicgradient(d).

undergoes periodic connection and disconnection (for ex-ample on an annual basis),τ I

E is related to the connectionor disturbance frequency,kD, a concept used extensively interrestrial ecology (Huston, 1994). Probability density func-tions can be used to characterise repeated stochastic events(e.g. flooding, fire, rainfall) and the periods between events,providing quantification ofτ I

E distributions.

www.hydrol-earth-syst-sci.net/17/1133/2013/ Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013

Page 6: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

1138 C. E. Oldham et al.: A generalized Damkohler number for classifying material processing

We stress again that isolation may be defined both hydro-logically and materially. As commented above, connectivity(and therefore isolation) has previously been defined basedon hydrology. When water flows between elements, they arehydrologically connected. When water no longer flows be-tween elements, they are hydrologically isolated. However,if material is consumed or settled prior to reaching the adja-cent element, material isolation of elements is possible evenduring hydrological connectivity. For material isolation, theefficiency of material transfer will be poor. Our definition of“poor material transfer” is arbitrary and will vary with ap-plication. In some contexts, a 50 % transfer of material be-tween elements may define them as materially isolated. Inother applications, 1 % material transfer between elementsmay define isolation. Examples of aquatic systems that arematerially isolated are shown in Fig. 4.

2.2.2 Material processing timescales

Term III in Eq. (2) contains a rate constant for material pro-cessing, such as chemical reaction, seed dispersal and recruit-ment, nutrient uptake, biogeochemical cycling, etc. We con-tinue below with a focus on biogeochemical cycling. Whileacknowledging that many biogeochemical processes havemore complex kinetics, we initially assume our chemical re-action of interest can be described as

A + Bk

−→ C + D. (6)

Equation (6) can be treated as a second order reaction wherek is the second order rate constant (m3 mmol−1 s−1). But ifwe assumeA to be in excess ofB, then Eq. (6) can be con-sidered “pseudo-first-order” with respect toB, where [B] isnot limiting and

d[B]

dt= −k′

[B], (7)

wherek′ is the pseudo-first-order rate constant (s−1) and isused in Eq. (2). The material (in this case chemical) process-ing timescale is then

τP = 1/k′. (8)

For reactions mediated by microbes and enzymes, we canuse a Monod kinetic model (Park and Jaffe, 1995; Schafer etal., 1998), which describes the rate of microbial growth as afunction of substrate concentration:

d[X]

dt= µmax

[B]

Ks + [B][X], (9)

whereµmax is the maximum specific growth rate (day−1),[X] is the concentration of cells (cells L−1) andKs is the sub-strate half-saturation constant (mmol L−1). The correspond-ing equation for the rate of substrate consumption is

d[B]

dt= −qmax

[B]

Ks + [B][X] (10)

Fig. 4. Regime III: (a) the elements of interest are hydrologicallyand hydraulically disconnected. Examples of this regime are anoxbow lake, which has disconnected from its river(b), and a wet-land perched above the groundwater table(c).

where qmax is the maximum specific consumption rate(mmol cell−1 day−1) and

qmax = µmaxY,

whereY is the cell yield (mmol cell−1). We define the char-acteristic timescale as being the time at which [B] = Ks andthus

τP =2

µmax. (11)

Note that processing timescales areintrinsic timescales de-rived from rate laws describing reactions taking place at thescale of the reactants (e.g. microorganisms, microbial con-sortia, biofilms, mineral surfaces) with full access to sub-strate or under conditions of complete mixing. Under con-ditions where short exposure timescales constrain access tosubstrate and/or reaction, theeffectiveprocessing timescalesmay be significantly longer than intrinsic timescales (andhence, under 1st order conditions, effective rate constantswill be less that intrinsic rate constants). Careful analysis isrequired to clarify whether intrinsic or effective processingtimescales are desirable or useful to a specific problem; theuse of effective processing timescales may remove the signa-ture of key drivers of a system.

2.3 Non-dimensional numbers

Non-dimensional numbers have long been used to charac-terise natural and built systems; some common examples are

Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013 www.hydrol-earth-syst-sci.net/17/1133/2013/

Page 7: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

C. E. Oldham et al.: A generalized Damkohler number for classifying material processing 1139

the Mach number, the Reynolds number and the hydraulicgradient. Battiato and Tartakovsky (2011) proposed the useof non-dimensional numbers in hydrological upscaling ap-proaches and to provide quantitative measures of the valid-ity of upscaling approximations. We now explore two non-dimensional variables that make up the key components ofour connectivity framework.

The non-dimensional Peclet number,Pe, provides the bal-ance between advection and diffusion under hydrologicalconnectivity as

Pe =τD

E

τAE

=ui 1xi

Deff, (12)

where1xi is the characteristic length scale over which trans-port processes are operating. ThePe is used to determinewhether Terms I or II in Eq. (2) are retained.

The Damkohler number (Eq. 1) has been used extensivelyin chemical engineering and more recently in environmen-tal and hydrological contexts. However, the definition ofDaimplicitly assumes hydrological or hydraulic connectivity.Without hydrological connectivity, there is no transport ofchemicals between hydrological elements, therefore we areunable to defineτT. For example,Da could not be used tocharacterise hydrologically disconnected systems that are de-fined by the isolation timescaleτ I

E.To characterise material fate during both hydrological con-

nection and disconnection, we propose the use of the moregeneral form

NE =τE

τP(13)

whereτE is the generalized exposure timescale,τP is a gen-eralized processing timescale, andNE gives an indication ofthe extent of material processing possible within the expo-sure timescale. For hydrologically connected elements, whenNE � 1, material is transferred conservatively between ele-ments; whenNE � 1, material will have plenty of opportu-nity to be processed within the exposure times. For a hydro-logically isolated element, whenNE � 1, material is unlikelyto be processed while that element is isolated; whenNE � 1,material will be processed during isolation.

For systems that are intermittently connected, we suggestseparate estimation ofNE for the connected and disconnectedconditions. This forces us to define exposure and processingtimescales for both conditions, allows assessment of the eco-logical significance of each condition (is the isolated statemore significant? or is it the connected state that is importantfor the system? or both?) and facilitates decisions on appro-priate spatial or temporal averaging of exposure timescales.Even if the isolation timescale� the transport timescale,their relevance for material connectivity depends onNE. IfNE � 1 during isolation, andNE � 1 during transport, thenthe isolation period will be ecologically relevant even if it isshorter than the subsequent transport timescale. A priori av-eraging to obtain an “effective” exposure timescale may in

fact remove critical information about fundamental controlson the ecosystem.

There are times when the length scale of a control volumeis not physically constrained, instead the physical dimensionsof the control volume may be arbitrarily sized to ensure, forexample, that 90 % of material of interest within the controlvolume is processed. If we take an example of a pseudo-first-order reaction occurring across a hydrologically connectedsystem, and using Eq. (13), then Eq. (7) can be integrated toobtain

[B]∗ = [B]0 exp(−k′ τE

)= [B]0 exp(−NE)

(14)

where[B]∗ is [B] at t = τE and[B]0 is [B] at t = 0.When chemicalB is 90 % removed,

[B]∗

[B]0= 0.1 = exp(−NE) , (15)

thenNE ∼ 2.3. So for a system whereNE � 2.3, we expectmore than 90 % removal of chemical. For a system whereNE � 2.3, we expect less than 90 % removal of chemical.This framework also has application for the definition of ma-terially isolated systems and highlights that material isolationis a function ofNE, rather than an absolute concept.

Examples of the use ofNE for specific contexts are pre-sented below and as case studies in the discussion. The def-inition of boundary and initial conditions are required asthe framework is applied, and their magnitude and charac-ter (e.g. constant or varying with time) will depend on ap-plication. This is analogous to the application of the non-dimensional Reynolds number,Re, which describes the bal-ance between inertial and viscous forces and is used to char-acterise environmental flows.Reis defined by the context orapplication, requiring careful definition of the system, its rel-evant length scales, and thus velocities. These velocities maybe estimated via order of magnitude estimates, via experi-mental data or via complex numerical modeling. The scale-dependentRe provides a framework for conceptualizing aproblem and provides a degree of generalization across sys-tems. We envisage a similar utility ofNE.

2.4 Special cases

2.4.1 Regime I: advection dominated

For systems that are hydraulically or hydrologically con-nected andPe� 1, i.e. the system is advection dominated(Fig. 2), Term II in Eq. (2) drops out, leaving

∂[B]

∂t= ui

∂[B]

∂xi−k′

[B].

I II(16)

Substituting Eq. (4) andk′ into Eq. (13) gives

NAE =

τE

τP=

k′ 1x

ux

. (17)

www.hydrol-earth-syst-sci.net/17/1133/2013/ Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013

Page 8: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

1140 C. E. Oldham et al.: A generalized Damkohler number for classifying material processing

WhenNE � 1, τE � τP, which indicates that the control vol-ume exposure time is sufficiently long for chemicalsA andB

to react. WhenNE � 1, τT � τP, which indicates that chem-icalsA andB do not have sufficient time to react before theyare advected out of the control volume; in this case chemi-calsA andB may be treated as if they are conservative.

In the largePe limit, the general solution Eq. (3) can bere-written as

[B] = exp(−k′ t)f (x − ux t) (18)

where the functionf (x) represents the initial (t = 0) distri-bution of [B] in the element. As time increases, the initialprofile of [B] is advected at speedux between the elementswith a reduction in magnitude due to the reaction term.

This solution and its behaviour are shown in Fig. 5a and b.Figure 5a shows the case where there is no significant con-sumption ofB (NE � 1) before it has been advected betweenthe elements. The magnitude of the initial profile has beenpreserved. Figure 5b shows a reduction in the magnitude ofthe profile due to significant consumption ofB (NE � 1).Note, however, thatB has not spread during transport be-tween the elements since this is the advection-dominatedregime (i.e. diffusion is minimal).

2.4.2 Regime II: diffusion dominated

For systems that are hydraulically or hydrologically con-nected andPe� 1 (i.e. system is diffusion dominated,Fig. 3), Term I in Eq. (2) drops out; so we have

∂[B]

∂t= Deff

∂2[B]

∂xi ∂xi−k′

[B],

II III(19)

and given Eq. (5), then

NDE =

k′ 1x2

Deff. (20)

WhenPe� 1, the general solution Eq. (3) can be re-writtenas

[B] = exp(−k′ t)

[α0 +

∞∑n=0

(αn cos

(nπ

1xi

x

)

+βn sin

(nπ

1xi

x

))exp

(−

n2π2Deff

1x2i

t

)]. (21)

In the absence of significant reaction (NE � 1), B will even-tually diffuse through the entire system without significantloss ofB. This is shown in Fig. 5c. Alternatively, ifNE � 1,then much ofB will be used up before it has a chance todiffuse through the system, which is shown in Fig. 5d.

The idealized solutions discussed above for Regimes Iand II have assumed some initial distribution of [B] and pe-riodic boundary conditions. More general boundary condi-tions (for example fixed flux or fixed concentration boundary

Fig. 5. Example profiles of [B] from the model problem for small(solid line) and large (dashed line) times illustrating the evolution of[B] under different conditions:(a) Pe� 1 and NE� 1, (b) Pe� 1and NE� 1, (c) Pe� 1 and NE� 1, (d) Pe� 1 and NE� 1. Notethat in this figure,B and x have been normalized for illustrativepurposes.

conditions) do not affect the general framework and regimesoutlined above. The fate of inflows at the boundaries of theelement is determined by the element’s regime.

2.4.3 Regime I and II boundary

We have shown how to usePeandNE to characterise hydro-logical systems, however porous media are frequently char-acterised by their dispersivity,

αi =Deff

ui

. (22)

If we normalize the dispersivity by the characteristic lengthscale (i.e. the length of the control volume), then

αi

1xi

=Deff

ui 1xi

=1

Pe. (23)

Note that this relationship is dependent on scale and the em-pirically derived relationship

αi = 0.01751x1.46i (24)

can be applied to a flow path length less than 3500 m (Toddand Mays, 2005).

Reported horizontal field dispersivities considered of highreliability, typically range from 10−2 to 102 m, over lengthscales ranging from 100 to 103 m (Gelhar et al., 1992), andusing Eq. (23) we obtainPe∼ O(1)− O(102), indicating thatboth advection and diffusion are significant in the devel-opment of [B] at the upper end of the range. Under this

Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013 www.hydrol-earth-syst-sci.net/17/1133/2013/

Page 9: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

C. E. Oldham et al.: A generalized Damkohler number for classifying material processing 1141

condition, advection and diffusion act on similar timescalesand either Eq. (17) or Eq. (20) can be used.

It has been demonstrated that in porous media, whenPe∼ O(1), the progress of a reaction can be limited by in-complete mixing of flow paths containing the different reac-tants (e.g. O2 and Fe(II)) (Cirpka and Attinger, 2003). Incom-plete mixing results in a reduced exposure timescale (whichis a function of the mixing frequency) and therefore the effec-tiveness of a reaction (or process) is diminished even thoughthe processing timescale may be quite short or even instanta-neous (Gramling et al., 2002). This highlights the need to ac-curately determine exposure timescales and the caution withwhich we should utilize spatially or temporally averaged ex-posure timescales.

2.4.4 Regime III – hydrological disconnection

When hydrological elements become hydrologically isolatedwith no water fluxes in or out, the chemical reaction shown inEq. (6) may quickly become limited, by either the depletionof reactants or the build up of products. The consequencesof hydrological/hydraulic isolation are no resupply of reac-tants and no flushing of products. Under these conditions, thechange in concentration ofB cannot be described by Eq. (7),however to illustrate our conceptual framework we will as-sume that until chemicalB is 10 % consumed, pseudo-firstorder kinetics can be utilized, i.e. the reaction is not limitedby [A]. In this case we revert to using

∂[B]

∂t= −k′

[B]. (25)

The general solution Eq. (3) can then be re-written simply as

[B]t =[B]0 exp(−k′ t) (26)

where[B]0 is the spatially averaged concentration ofB att = 0, and then

N IE = τI k

′. (27)

3 Discussion

3.1 Application of NE approach across scales

In the previous sections we outlined howNE and Pe canbe used to characterise hydrological systems according tothe fate of chemicalB. We will now examine this conceptacross different temporal and spatial scales. We will focuson the oxidation of Fe(II), which plays a prominent rolein linking metabolic activities across anaerobic and aero-bic components of aquatic systems (Raiswell and Canfield,2012). Fe(II) oxidation is an ecologically significant reactionboth under neutral and acidic conditions, however the rateis strongly dependent on pH (Stumm and Morgan, 1996).

The oxidation process typically implies precipitation of fer-ric oxides and through that the generation of acidity. Un-der well-buffered conditions, such as anaerobic ground orpore waters, the pH remains relatively constant. Under lowor no-alkalinity conditions however, oxidation of Fe(II) sig-nificantly contributes to acidification of surface waters. Ex-amples of such systems are acidic mining lakes (Peine et al.,2000) or acid sulphate soils (Burton et al., 2006) where acid-ity has been generated through the oxidation of pyrite, butwhere acidic iron cycling, i.e. the combined reduction of fer-ric and oxidation of ferrous iron, maintains acidic conditions(Peine et al., 2000). Here we will compare anaerobic aquaticsystems containing Fe(II) and how they respond to oxygenexposure under different transport regimes. We have cho-sen two pH conditions: neutral (pH 7) conditions as foundin many anaerobic environments, and slightly acidic (pH 4)conditions as found in many anaerobic systems receivingwaters that have been exposed to pyrite oxidation (Blodau,2006).

The oxidation of Fe(II) can be modelled using the rate lawgiven by Stumm and Lee (1961).

d[Fe(II)]

dt= −kabioPO2

[OH−

]2[Fe(II)] (28)

where PO2 is the partial pressure of oxygen (0.21 atm)and kabio is the abiotic rate constant (1.5× 1013 L2 mol−2

atm−1 min−1 at 25◦C).This rate law predicts extremely low oxidation rates under

acidic conditions. Typically, Fe(II) oxidation will be accel-erated by bacteria. Pesic et al. (1989) reported a decrease inmicrobial oxidation rate with increasing pH between pH 2.2and pH 3. Extrapolation of their rate law (Kirby et al., 1999)demonstrates that at pH 4 the abiotic rate is still higher thenthe microbial rate. Hence, in our examination we will useEq. (28) for pH 4 and pH 7. The rate constantk in Eq. (28)can be converted into a pseudo 1st-order rate constant forambient temperature and pressure conditions, and for a pre-defined pH, as

k′= kabioPO2

[OH−

]2(29)

wherek′ yields the reaction times,τR = τP (Table 2), whichwill be used for the analyses below.

3.1.1 Case I: porewater transport in deep lakesediments

The sediment of deep eutrophic lakes (pH neutral con-ditions) or mining lakes (acidic conditions) may be hy-drologically connected to the surrounding groundwater,but if the hydraulic conductivity of the sediment materialis extremely low, groundwater inflow velocities could be< 10−9 m s−1. We assume a molecular diffusion coefficient,Deff of 10−9 m2 s−1.

Under this scenario, the characteristic length scale is notphysically constrained. We can, however, estimate a depth at

www.hydrol-earth-syst-sci.net/17/1133/2013/ Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013

Page 10: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

1142 C. E. Oldham et al.: A generalized Damkohler number for classifying material processing

Table 2.Pseudo-first order rate constants,k′, and processing (reac-tion) timescales,τP, for the oxidation of iron(II) under different pHconditions.

pH k′ (min−1) τP (s)

7 3× 10−2 2× 103

4 3× 10−8 2× 109

which the O2 concentration has been decreased to a specifiedlevel. In order to fulfil the validity of pseudo-first order kinet-ics, we assume a reduction by only 10 %, i.e.1PO2 = 0.021atm which corresponds to1[O2] ≈ 0.03 mmol L−1 at ambi-ent temperature, and stoichiometrically that corresponds toan Fe(II) consumption of1[FeII] = 0.03 mmol L−1. Hence,the thickness of a layer will be estimated, in which the oxy-gen concentration does not drop below 0.27 mmol L−1.

Fe(II) concentrations in pH neutral lake sediments rarelyexceed 0.1 mmol L−1 (see e.g. Baccini, 1985; Gelhar et al.,1992), while in acidic environments Fe(II) concentrationsof up to 10 mmol L−1 may be reached (Burton et al., 2006;Peine et al., 2000). Therefore, a1[FeII] of 0.03 mmol L−1 isa 30 % removal at pH 7 and a 0.3 % removal at pH 4. UsingEq. (15),NE ≈ 0.4 and 0.003 for pH 7 and 4, respectively.

In the next step we can estimate1x10% by combiningEqs. (5) and (20) as

1x10% =

√ND

E · τP · Deff, (30)

which yields 8× 10−4 m at pH 7, and 8× 10−2 m at pH 4.The corresponding exposure timescales,τD

E , within thesecontact zones are 7× 102 s at pH 7, and 6× 106 s at pH 4.Using these length scales, a groundwater inflow velocityof 10−9 m s−1 and the molecular diffusion coefficient,Pe,ranges from 8.2× 10−4 at pH 7, to 7.6× 10−2 at pH 4 (Ta-ble 3), which indicates diffusive conditions and supports ap-plication of Eq. (5).

The above calculations have two specific implications.Firstly, the thickness of the layer that remains close to O2 sat-uration varies significantly between the two pH regimes. Thezone of Fe(II) production from dissimilatory iron reduction(a biogeochemical process sensitive to O2) will be shiftedto deeper sediments under acidic conditions. Secondly andmore importantly, the timescale of exposure to oxygen satu-ration is 4 orders of magnitudes longer under acidic condi-tions and lasts for several weeks. Note that the calculatedτEis a lower estimate and may be even longer at lower Fe(II)concentrations. This long timescale provides ample opportu-nities for interference from other biogeochemical processesthat involve oxygen or consume Fe(II).

Hence,NE can be regarded as a process-specific parame-ter that allows us to characterise hydrological systems basedon chemical fate. WhenNE ∼ O(1), as was calculated un-der pH 7, the balance between exposure and processing is

Table 3. Estimates of relevant parameters and non-dimensionalnumbers at pH 4 and 7.

pH Fe(II) in NE 1x10% (m) τE (s) Peporewater

(mmol L−1)

7 0.1 0.4 8× 10−4 7× 102 8.2× 10−4

4 10 0.003 8× 10−2 6× 106 7.6× 10−2

critical and the system will be sensitive to shifts in hydro-logical regime. WhenNE � O(1), as was calculated underpH 4, the system will not be sensitive to shifts in hydrologi-cal regime.

3.1.2 Case II: Surface water – groundwater exchange

The advective recharge of oxygenated surface waters intoFe(II)-rich groundwater is very common in many ripariansystems under losing conditions. While the majority of thesesystems are pH neutral, wetlands affected by acid sulphatesoils (Johnston et al., 2011) and mining lakes (Fleckensteinet al., 2009) may lose acidic water to groundwater.

Again under this scenario, the characteristic length scaleis not physically constrained, and we can again estimate thelength scales over which the O2 concentration has been de-creased by 10 %. The Fe(II) boundary conditions are selectedas in Case I.1x10% can now be calculated using Eqs. (4)and (17) as

1x10% = ux τPNAE (31)

whereux is set to 10−5 m s−1 for this scenario.The distance from the riparian zone over which 10 % of

O2 is consumed is 6.8× 10−3 m at pH 7, and 5.7× 101 m atpH 4. In other words, after 57 m of flow across a riparian zoneat pH 4 the system would still be close to oxygen saturation,allowing O2 consuming reactions.

To calculatePenumbers for Case II, we can use Eq. (23).In porous-media flow,Deff becomes a scale-dependent dis-persion coefficient, which can be estimated from the dis-persivity, α, of the aquifer into which the lake waterpenetrates (see Eqs. 22 and 24). We getαi = 1× 10−5 mand Deff = 1× 10−10 m2 s−1 for pH 7, andαi = 6.4 m andDeff = 6× 10−5 m2 s−1 for pH 4. Thus at pH 7, i.e. withinthe short distance of 6 mm, no dispersion occurs. The corre-spondingPe ranges from 5.7× 102 at pH 7 (indicating ad-vective transport through the O2 containing layer) to 8.9 atpH 4 (indicating advective transport with contributions fromdispersive mixing).

Equations (30) and (31) can be directly applied for wa-ter management purposes. For example, we propose that thewidth of a riparian zone, in which no agricultural activ-ity is allowed, can be calculated using the framework pre-sented here to ensure minimal nitrate export from ground-water to the river. With knowledge of the dominant transport

Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013 www.hydrol-earth-syst-sci.net/17/1133/2013/

Page 11: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

C. E. Oldham et al.: A generalized Damkohler number for classifying material processing 1143

conditions (and how they vary with time), we could size theriparian width to the length scale,1x, that is required to re-move, for instance, 90 % of the nitrate that has infiltrated togroundwater from agricultural land. To some extent the “50-day line” typically applied to constrain drinking water zonesimplicitly uses this concept already.

3.1.3 Case III: temporarily disconnected systems

Floodplain wetlands and playas frequently become temporar-ily disconnected from surface and sub-surface water flowpathways. In Western Australia, shallow groundwater-fedwetlands affected by acid sulphate soils fill with Fe(II)-richgroundwater, which is then exposed to O2 (Nath et al., 2013).Seasonal lowering of the groundwater table disconnects thewetlands from the source of Fe(II), and its subsequent oxida-tion via ongoing exposure to O2 is controlled by the isolationtimescale, i.e.τ I

E. Depending on the buffering capacity of thewetland mineralogy, both acidic and neutral conditions canbe observed.

Under this scenario, the characteristic length andtimescales are physically constrained. If we assume a typi-calτ I

E ∼ 100 days (∼ 107 s), then using Eq. (14), we estimatethe extent of Fe(II) oxidation as 100 % at pH 7 and 1 % atpH 4. The differences between Fe(II) oxidation at the differ-ent pHs can also be demonstrated byNE. Using Eq. (13),NEis 5× 103 at pH 7 (indicating the disconnection timescaleprovides sufficient exposure opportunity for the reaction toproceed) and 5× 10−3 at pH 4 (indicating that the disconnec-tion timescale does not provide enough time for the reactionto proceed). Of course these calculations assume the wetlandwaters are fully mixing at all times. Any density stratificationwithin the wetland will reduceτ I

E, i.e. the timescales overwhich Fe(II)-rich bottom waters are exposed to O2-rich sur-face waters. Density stratification may also alter the controlvolume utilized and therefore the spatial averaging of chem-ical concentrations.

3.2 Competiveness of reactions

We have shown thatNE numbers are reaction specific; thecorresponding spatial scales are also reaction dependent andnot always controlled by geomorphology. In a first approach,we can therefore argue that lowNE for a certain reaction im-plies ample opportunity for alternative reaction pathways toproceed between materially connected elements or within amaterially isolated element. Inversely, highNE implies pre-dominance of the biogeochemical process of interest.NE al-lows us to compare the effectiveness of different materialprocessing reactions across a certain temporal (τE) or spatial(1x) scale.

More generally, lowerNE implies lower competiveness ofa specific reaction. In the examples discussed above, the ox-idation of Fe(II) is much more competitive at pH 7 than atpH 4. In this case competiveness is clearly controlled by the

large differences ink′. However,NE is of course also affectedby the exposure timescale. We therefore propose that for agiven system,NE can be used as a general parameter to com-pare competiveness between different reactions of interest.

Competiveness of redox reactions is typically discussed interms of thermodynamic arguments, for example the gain infree energy,1Gf , obtained from oxidation of organic matter(Zehnder and Stumm, 1988). This concept works very well indiffusion controlled systems such as the sediment porewatersof the ocean or deep lakes. Attempts to adapt this concept togroundwater systems have been problematic. Rather, zonesof intensive redox activity are observed to be spatially dis-tributed or located at plume fringes (see e.g. Prommer et al.,2006). Considering that the rate of a reaction is proportionalto 1Gf , thermodynamics are also related to the reactiontimescale. Considering further that all dissolved substanceshave appropriately the same diffusion coefficients and thattheir residence times under diffusion controlled conditionsare approximately identical, then it becomes clear that1GfandNE are interrelated predictors for competiveness underdiffusion-controlled (or well-mixed) conditions.

Under advection or dispersion controlled conditions, how-ever, the timescale of exposure along a flow path variesstrongly depending on external forcing. Of course the flowpaths and therefore exposure timescales may be distributedin space, similar to residence times, and this spatial vari-ability will give rise to a patchwork ofNE values, parallel-ing the concept of “residence time distributions” with eachcontrolled by specific biogeochemical regimes. Under theseconditions, chemical competiveness becomes dependent onthe ratio of processing/reaction to exposure timescales;NEtherefore can be used as a general parameter to compare com-petiveness under all regimes.

3.3 Intermittent connectivity and ecologicalconsequences

Hydrological systems transition between the regimes out-lined above, according to external forcing. Of particular rel-evance may be shifts between Regime I and II or intermit-tent transition from Regime III, i.e. the disconnected state,to Regime I or II. Estimates ofNE for the different regimesprovide a framework for predicting the biogeochemical andecological significance of the shifts (Table 4).

As an example, riparian zones are frequently characterisedby such regime shifts. Material connectivity between receiv-ing streams and adjacent riparian soils or wetlands is oftenstrongly related to storm events leading to rapid activationof groundwater and its discharge into streams. This is an ex-ample of systems that routinely shift between periods char-acterised by long exposure times and diffusion-controlledtransport (Regime I), and periods triggered by rainfall eventsand characterised by strong advection and short exposuretimes (Regime II). This shift in regime results in substan-tial transfer of solutes, e.g. DOC, Fe, NO−

3 or other species.

www.hydrol-earth-syst-sci.net/17/1133/2013/ Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013

Page 12: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

1144 C. E. Oldham et al.: A generalized Damkohler number for classifying material processing

Table 4.A summary of system characteristics under the different regimes and as a function ofNE. Note that once elements becomeconnected after a period of isolation, it is possible for them to rapidly move across regimes. The balance between exposure timescales (bothconnected and isolated) and processing timescales determines the ecological signficance of the period of isolation and the regime shift.

39

Table 4: A summary of system characteristics under the different regimes and as a 798

function of NE. Note that once elements become connected after a period of isolation, 799

it is possible for them to rapidly move across regimes. The balance between 800

exposure timescales (both connected and isolated) and processing timescales 801

determines the ecological signficance of the period of isolation and the Regime shift. 802

803

804 805

NE>> 1TE>> TP

NE<< 1TE<< TP

NE ~O(1)TE ~ TP

Reg

ime

I or R

egim

e II

Reg

ime

III

Not sufficient exposuretime for reaction.Reactants can beconsidered conservative.Materialconnectivity definedhydrologically. Sensitive to change inhydrological regime.

Highly dynamic.Balance very senstiive toconditions.Need to characterise bothhydrology andbiogeochemistry.Sensitive to change inhydrological regime.

Reactants must beconsiderednon-conservative.Materialconnectivity definedbiogeochemically.Ecological hotspotexpected.

Highly dynamic.Balance very senstiive toconditions.Need to characterise bothhydrology andbiogeochemistry.Sensitive to change inhydrological regime.

Not sufficient exposuretime for reaction.Sensitive to change inhydrological regime. Ecological hotspot notexpected.

ISOLATION IMPORTANTECOLOGICALLY

ISOLATION NOT IMPORTANTECOLOGICALLY

Reactants must beconsiderednon-conservative.Materialconnectivity controlledbiogeochemically.

possibleEcological hotspot

Hyd

rolo

gica

l l c

onne

ctiv

ityH

ydro

logi

cal o

r mat

eria

ldi

scon

nect

ivity

E!ective m

aterial tr

ansfer, i

solatio

n importa

nt eco

logically

Lim

ited

mat

eria

l tra

nfer

pos

sivle

with

in e

xpos

ure

times

cale

Conn

ectiv

ity le

ss im

port

ant e

colo

gica

llyThe magnitude of such pulses typically depends stronglyon the antecedent conditions (e.g. Inamdar et al., 2004).Knorr (2013) observed that Fe and DOC were exported tostreams during rainfall events only if the preceding periodallowed a pool of Fe and DOC to be created (Regime I). In-versely, after a rainfall event, a rapid transfer of Fe to thestream would require the prevention of oxidation and reten-tion at the site (Regime II). If we apply theNE framework tothis example, we can see that if the Regime I (pre-event)NEfor the formation of Fe(II) is high, a large pool of dissolvedFe can build up. In contrast, effective transfer of Fe into thereceiving stream during a rainfall event requires the Fe(II)oxidation timescale to be longer than the (typically short)exposure timescale, i.e.NE < 1. As in our case studies, theoxidation timescale depends strongly on pH. The oxidationkinetics may also be dependent on DOC-Fe(II) complexationwhich stabilizes Fe(III) as colloids and thus promotes exportof Fe associated with DOC (Llang et al., 1993). During peri-ods of shifting exposure times, changes inNE usefully high-light the different controls on material connectivity across a

hydrological system and its impact on material transfer outof the system.

An interesting example for our discussion of material con-nectivity is given by Basu et al. (2011b) in an explorationof hydrological versus biogeochemical controls on the catch-ment scale, which can be reanalysed usingNE numbers. Basuet al. (2011b) termed export of substances from a catchment“chemostatic”, (it showed little variation of concentrationwith flow) if the rate of mass input from storage areas wascomparable to the rates of degradation. Under these con-ditions the storage pool does not become depleted and theconcentration of the substance becomes buffered. Using ourframework at the catchment scale, chemostatically behavingsubstances (nitrogen, weathering products) haveNE ∼ 1. Incontrast, episodic (event driven) responses are predicted tooccur if degradation rates of substances (e.g. pesticides) arehigh during periods of hydrological disconnection in the soilsurface and prevent accumulation of these substances (Basuet al., 2011b). Using our framework, episodic events trig-ger material connectivity that transfers the substances into

Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013 www.hydrol-earth-syst-sci.net/17/1133/2013/

Page 13: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

C. E. Oldham et al.: A generalized Damkohler number for classifying material processing 1145

surface waters, and the transfer rate fluctuates with the sizeof the accumulated pool. For such episodic events,NE � 1during time of disconnection in the soil surface, andNE � 1at the catchment scale, allowing detection of the substanceduring subsequent connection.

Gradients ofNE in either space or time may also be in-dicative of sinks or sources. WhenNE for a specific pro-cess and a specific hydrological element is large relative toit neighbouring elements, a concentration gradient may beestablished and a “hotspot” of material processing formed.WhenNE for an element is large relative to the preceding orsubsequent conditions, a “hot moment” may form. We sug-gest that it is the shift inNE that is critical for hot spot orhot moment formation, rather than a regime shift. Some ofthe possible ecological consequences of shifts inNE are ex-plored in Table 4.

So far we have constrained ourselves to considering spe-cific biogeochemical reactions occurring within elements ofthe hydrological cycle. Finally, we broaden our focus to con-sider landscape-scale ecological processes and their charac-teristic timescales. The extent to which an ecosystem is re-silient to environmental perturbations is likely controlled bythe relative rates at which a biological community can re-spond to environmental change or disturbance (i.e.k′) andthe exposure timescale of the relevant element. If we setthe biogeochemical rate,k′, to be analogous to populationgrowth rate and the exposure timescale,τE, to be analogousto exogenous timescales (Hastings, 2010) or the reciprocalof the disturbance frequency, then our two timescales havedirect analogues to concepts in terrestrial ecology. For ex-ample, Huston (1994) suggested that the ratio of the popula-tion growth rate to the disturbance frequency, i.e.NE in ourframework, impacts on biodiversity and ecological resilience(Fig. 6).

Of particular interest is the conceptual understandingthat community biodiversity is maximal when there is abalance between population growth rate (or biogeochem-ical reaction rate) and disturbance frequency, i.e. whenNE ∼ O(1) (Huston, 1994, Fig. 6). The biodiversity in turnaffects ecosystem resilience (Gunderson, 2000; Scheffer andCarpenter, 2003); however, the stability space within whichecosystem resilience operates is determined by external fac-tors, and has been related to characteristic spatial or tempo-ral scales of external forcing (Scheffer and Carpenter, 2003).We have shown above thatNE is process-specific so thequestion arises whether we can describe the resilience of asystem in terms of itsNE distribution. This requires furtherinvestigation.

Fig. 6.The relationship betweenNE and biodiversity. This is likelyto hold true for microbial diversity as well as landscape-scale bio-diversity. Modified from Huston (1994).

4 Conclusions

The concepts discussed above provide a general and holis-tic framework to analyse and characterise system behaviourin terms of the effectiveness of material processing, definedby the non-dimensional numberNE, as the ratio of the expo-sure timescale to the material processing timescale. The useof this framework has multiple benefits: (a) it focuses our at-tention on clearly defining our system, its spatial scales andthe material processing (or reaction) of interest; (b) it high-lights that “transport limitation” is a relative concept and de-pends on the material processing of interest; (c) it allows usto incorporate isolated systems within a hydrological systemsunderstanding; (d) it provides insights into the likely ecolog-ical significance of shifts in hydrological forcing; (e) it canbe used for clarification of whether the balance in control-ling mechanisms is constant with scale-up; (f) it can be usedfor water management as it allows us to appropriately sizewater remediation options; and (g) it can be used in the de-sign of monitoring programs to ensure that monitoring scalescapture the key dynamics of a system.

TheNE framework is simple in definition but challengingto apply. Identification of appropriate exposure and process-ing timescales for the system of interest is not trivial. Appli-cation of theNE framework requires a clear definition of theproblem and a sound rationale for the choice of timescalesand the methods used for their determination. These meth-ods can range from numerical simulations to geochemical orisotopic analysis. TheNE framework certainly helps to con-strain a research problem, and we expect that, with on-goinguse and development of this framework by the scientific

www.hydrol-earth-syst-sci.net/17/1133/2013/ Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013

Page 14: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

1146 C. E. Oldham et al.: A generalized Damkohler number for classifying material processing

Appendix A

Notation.

Variable Units Description

αi m Dispersivity ini-directionαn dimensionless Fourier cosine coefficientβn dimensionless Fourier sine coefficient[B] mmol L−1 Concentration of chemicalB[B]max mmol L−1 Maximum expected concentration of chemicalB

[B]0 mmol L−1 Initial concentration of chemicalBDa dimensionless Damkohler numberDeff m2 day−1 Effective diffusion coefficientk′ s−1 First order rate constantkabio L2 mol−2 atm−1 min−1 Abiotic rate constant for iron oxidationn dimensionless Summation indexNE dimensionless Ratio of exposure to processing timescalesNA

E dimensionless Ratio of exposure to processing timescales (advective regime)ND

E dimensionless Ratio of exposure to processing timescales (diffusive regime)N I

E dimensionless Ratio of exposure to processing timescales (isolated regime)Pe dimensionless Peclet numberPO2 atm Partial pressure of oxygenqmax mmol L−1 day−1

t unit of time TimeτT unit of time Transport timescaleτR unit of time Reaction timescaleτAE unit of time Characteristic exposure timescale (asdvective regime)

τDE unit of time Characteristic exposure timescale (diffusive regime)

τ IE unit of time Characteristic exposure timescale (isolated regime)

τP unit of time Characteristic processing timescaleui m day−1 Velocity in i-directionux m day−1 Velocity in x-directionxi m Distance ini-direction1x m Characteristic length scale inx-direction1x10% m Characteristic length scale required for 10 % of reactant to be removed

community, protocols for the delineation of timescales willdevelop.

This work has opened up an array of questions but per-haps the most intriguing question is how we most effectivelyanalyse intermittently disconnected systems. Is it preferrableto take a stochastic approach, using an ensemble of exposuretimescales? The parameter analysed could vary dependingon assessed ecological significance, but could involve (a) theisolation timescales, (b) the connection timescales, and/or(c) the ratio of the isolation to connection timescales.

We expect that in many systems there will also be an en-semble of reaction timescales, so a pdf would also be re-quired. We also expect the exposure timescales to have sig-nificantly greater variance than the reaction timescales. Theratio of these two pdfs would create anNE pdf, which wouldin turn characterise material connectivity between systems.A full exploration of the stochastic nature ofNE is neededand requires analysis of multiple exemplar data sets.

Acknowledgements.The authors would like to express thanks toa large number of people who over the years have contributed tolively discussion on the use of Da and other non-dimensionlessnumbers for characterising catchments, including Carlos Ocampo,Murugesu Sivapalan, Jeff McDonnell, Paul Lavery, Ursula Salmonand Wolfgang Kurtz. This collaboration was supported by DFGgrant Pe 438/20-1.

Edited by: S. Thompson

References

Ali, G. A. and Roy, A. G.: Shopping for hydrologicallyrepresentative connectivity metrics in a humid temper-ate forested catchment, Water Resour. Res., 46, 1–24,doi:10.1029/2010WR009442, 2010.

Asarita, G.: Mass transfer and chemical reactions, Elsevier, Ams-terdam, 1967.

Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013 www.hydrol-earth-syst-sci.net/17/1133/2013/

Page 15: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

C. E. Oldham et al.: A generalized Damkohler number for classifying material processing 1147

Baccini, P.: Phosphate interactions at the sediment-water interface,in: Chemical Processes in Lakes, edited by: Stumm, W., WileyOnline Library, 1985.

Basu, N. B., Rao, P. S. C., Thompson, S. E., Loukinova, N. V., Don-ner, S. D., Ye, S., and Sivapalan, M.: Spatiotemporal averagingof in-stream solute removal dynamics, Water Resour. Res., 47,W00J06,doi:10.1029/2010WR010196, 2011a.

Basu, N. B., Thompson, S. E., and Rao, P. S. C.: Hydrologicand biogeochemical functioning of intensively managed catch-ments: A synthesis of top-down analyses, Water Resour. Res.,47, W00J15,doi:10.1029/2011WR010800, 2011b.

Battiato, I. and Tartakovsky, D. M.: Applicability regimesfor macroscopic models of reactive transport inporous media, J. Contam. Hydrol., 120-121, 18–26,doi:10.1016/j.jconhyd.2010.05.005, 2011.

Beven, K. J.: Rainfall-runoff modelling: The primer, Wiley, NewYork, 2001.

Blodau, C.: A review of acidity generation and consumption inacidic coal mine lakes and their watersheds, Sci. Total Environ.,369, 307–332, 2006.

Bolster, D., de Anna, P., Benson, D. A., and Tartakovsky, A. M.:Incomplete mixing and reactions with fractional dispersion, Adv.Water Resour., 37, 86–93,doi:10.1016/j.advwatres.2011.11.005,2012.

Botter, G., Bertuzzo, E., and Rinaldo, A.: Transport in the hydro-logic response: Travel time distributions, soil moisture dynam-ics, and the old water paradox, Water Resour. Res., 46, 1–18,doi:10.1029/2009WR008371, 2010.

Botter, G., Bertuzzo, E., and Rinaldo, A.: Catchment residence andtravel time distributions: The master equation, Geophys. Res.Lett., 38, 1–6,doi:10.1029/2011GL047666, 2011.

Burton, E., Bush, R., and Sullivan, L.: Sedimentary iron geochem-istry in acidic waterways associated with coastal lowland acidsulfate soils, Geochim. Cosmochim. Acta, 79, 5455–5468, 2006.

Carleton, J. N.: Damkohler number distributions and constituent re-moval in treatment wetlands, Ecol. Eng., 19, 233–248, 2002.

Cirpka, O. A. and Attinger, S.: Effective dispersion in heteroge-neous media under random transient flow conditions, Water Re-sour. Res., 39, 1–15,doi:10.1029/2002WR001931, 2003.

Dankwerts, P. V.: Absorption by simultaneous diffusion and chemi-cal reaction into particles of various shapes and into falling drops,Trans. Faraday Soc., 47, 1014–1023, 1951.

Detty, J. M. and McGuire, K. J.: Topographic controls on shallowgroundwater dynamics: implications of hydrologic connectivitybetween hillslopes and riparian zones in a till mantled catchment,Hydrol. Process., 24, 2222–2236,doi:10.1002/hyp.7656, 2010.

Dooge, J. C. I.: Looking for hydrologic laws, Water Resour. Res.,22, 46–58, 1986.

Fleckenstein, J., Neumann, C., Volze, N., and Beer, J.: Raumzeit-muster des Grundwasser-Seeaustausches in einem Tagebaurest-see, Grundwasser, 14, 207–217, 2009.

Gelhar, L. W. L., Welty, C., and Rehfeldt, K. R. K.: A critical reviewof data on field-scale dispersion in aquifers, Water Resour. Res.,28, 1955–1974, 1992.

Gramling, C. M., Harvey, C. F., and Meigs, L. C.: Reactive transportin porous media: a comparison of model prediction with labora-tory visualization, Environ. Sci. Technol., 36, 2508–2514, 2002.

Gunderson, L. H.: Ecological resilience – in theory and application,Annu. Rev. Ecol. Syst., 31, 425–439, 2000.

Hastings, A.: Timescales, dynamics and ecological understanding,Ecology, 91, 3471–3480, 2010.

Hornberger, G. M., Scanlon, T. M., and Raffensperger, J. P.: Mod-elling transport of dissolved silica in a forested headwater catch-ment: the effect of hydrological and chemical time scales on hys-teresis in the concentration-discharge relationship, Hydrol. Pro-cess., 15, 2029–2038,doi:10.1002/hyp.254, 2001.

Hrachowitz, M., Soulsby, C., Tetzlaff, D., Malcolm, I. A., andSchoups, G.: Gamma distribution models for transit time esti-mation in catchments: Physical interpretation of parameters andimplications for time-variant transit time assessment, Water Re-sour. Res., 46,doi:10.1029/2010WR009148, 2010.

Huston, M. A.: Biological diversity: The coexistence of species onchanging landscapes, Cambridge University Press, Cambridge,1994.

Inamdar, S. P., Christopher, S. F., and Mitchell, M. J.: Export mech-anisms for dissolved organic carbon and nitrate during summerstorm events in a glaciated forested catchment in New York,USA, Hydrol. Process., 18, 2651–2661,doi:10.1002/hyp.5572,2004.

Johnston, S., Keene, A., Bush, R., Sullivan, L., and Wong, V.:Tidally driven water column hydro-geochemistry in a remediat-ing acidic wetland, J. Hydrol., 409, 128–139, 2011.

Kirby, C., Thomas, H., Southam, G., and Donald, R.: Relative con-tributions of abiotic and biological factors in Fe(II) oxidation inmine drainage, Appl. Geochem., 14, 511–530, 1999.

Klein, C., Wilson, K., Watts, M., Stein, J., Berry, S., Carwardine,J., Smith, M. S., Mackey, B., and Possingham, H.: Incorporat-ing ecological and evolutionary processes into continental-scaleconservation planning, Ecol. Appl., 19, 206–217, 2009.

Knorr, K.-H.: DOC-dynamics in a small headwater catchment asdriven by redox fluctuations and hydrological flow paths – areDOC exports mediated by iron reduction/oxidation cycles?, Bio-geosciences, 10, 891–904,doi:10.5194/bg-10-891-2013, 2013.

Kurtz, W., and Peiffer, S.: The role of transport in aquatic redoxchemistry, in: Aquatic Redox Chemistry, edited by: Tratnyek,P.G., Grundl, T. J., and Haderlein, S. B., American Chemical So-ciety, Washington DC, USA, 559–580, 2011.

Llang, L., McNabb, J. A., Paulk, J. M., Guy, B., and McCarthy, J. F.:Kinetics of Fe(I1) oxygenation at low partial pressure of oxygenin the presence of natural organic matter, Environ. Sci. Technol.,27, 1864–1870, 1993.

McDonnell, J., McGuire, K., Aggarwal, P., Beven, K., Biondi, D.,Destouni, G., Dunn, S., James, A., Kirchner, J., Kraft, P., Lyon,S., Maloszewski, P., Newman, B., Pfister, L., Rinaldo, A., Rodhe,A., Sayama, T., Seibert, J., Solomon, K., Soulsby, C., Stew-art, M., Tetzlaff, D., Tobin, C., Troch, P., Weiler, M., West-ern, A., Worman, A., and Wrede, S.: How old is streamwa-ter? Open questions in catchment transit time conceptualiza-tion, modelling and analysis, Hydrol. Process., 24, 1745–1754,doi:10.1002/hyp.7796, 2010.

McDonnell, J., Sivapalan, M., Vache, K., Dunn, S., Grant, G., Hag-gerty, R., Hinz, C., Hooper, R., Kirchner, J., Roderick, M., Selker,J., and Weiler, M.: Moving beyond heterogeneity and processcomplexity: A new vision for watershed hydrology, Water Re-sour. Res., 43, 1–6,doi:10.1029/2006WR005467, 2007.

McGuire, K. and McDonnell, J.: A review and evaluation ofcatchment transit time modeling, J. Hydrol., 330, 543–563,doi:10.1016/j.jhydrol.2006.04.020, 2006.

www.hydrol-earth-syst-sci.net/17/1133/2013/ Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013

Page 16: A generalized Damkohler number for classifying material ......A generalized Damkohler number for classifying material¨ processing in hydrological systems C. E. Oldham1, D. E. Farrow2,

1148 C. E. Oldham et al.: A generalized Damkohler number for classifying material processing

Meerkerk, A. L., Van Wesemael, B., and Bellin, N.: Applicationof connectivity theory to model the impact of terrace failure onrunoff in semi-arid catchments, Hydrol. Process., 23, 2792–2803,doi:10.1002/hyp.7376, 2009.

Michaelides, K. and Chappell, A.: Connectivity as a concept forcharacterising hydrological behaviour Definitions of Connec-tivity, Hydrol. Process., 522, 517–522,doi:10.1002/hyp.7214,2009.

Michalak, A. and Kitanidis, P.: Macroscopic behavior and random-walk particle tracking of kinetically sorbing solutes, Water Re-sour. Res., 36, 2133–2146,doi:10.1029/2000WR900109, 2000.

Nath, B., Lillicrap, A., Ellis, L., Boland, D., and Old-ham, C.: Hydrological and chemical connectivity dy-namics in a groundwater-dependent ecosystem impactedby acid sulfate soils, Water Resour. Res., 49, 441–457,doi:10.1029/2012WR012760, 2013.

Ocampo, C. J., Oldham, C. E., and Sivapalan, M.: Nitrate at-tenuation in agricultural catchments: Shifting balances be-tween transport and reaction, Water Resour. Res., 42, 1–16,doi:10.1029/2004wr003773, 2006.

Opperman, J. J., Luster, R., McKenney, B. A., Roberts, M., andMeadows, A. W.: Ecologically Functional Floodplains: Con-nectivity, Flow Regime, and Scale, J. American Water Resour.Assoc., 46, 211–226,doi:10.1111/j.1752-1688.2010.00426.x,2010.

Park, S. S. and Jaffe, P. R.: Development of a sediment redox poten-tial model for the assessment of post-depositional metal mobility,Ecol. Model., 91, 169–181, 1995.

Peine, A., Tritschler, A., Kusel, K., and Peiffer, S.: Electron flow inan iron-rich acidic sediment – evidence for an acidity-driven ironcycle, Limnol. Oceanogr., 45, 1077–1087, 2000.

Pesic, B., Oliver, D., and Wichlacz, P. An electrochemical methodof measuring the oxidation rate of ferrous to ferric iron withoxygen in the presence of Thiobacillus ferrooxidans, Biotechnol.Bioeng., 33, 428–439, 1989.

Pringle, C.: The need for a more predictive understandingof hydrologic connectivity, Aquat. Conserv., 13, 467–471,doi:10.1002/aqc.603, 2003a.

Pringle, C.: What is hydrologic connectivity and why is itecologically important?, Hydrol. Process., 17, 2685–2689,doi:10.1002/hyp.5145, 2003b.

Prommer, H., Tuxen, N., and Bjerg, P.: Fringe-controlled naturalattenuation of phenoxy acids in a landfill plume: integration offield-scale processes by reactive transport modeling, Environ.Sci. Technol., 40, 4732–4738, 2006.

Raiswell, R. and Canfield , D.: The iron biogeochemical cycle pastand present, Geochem. Perspect., 1, 1–232, 2012.

Schafer, D., Schafer, W., and Kinzelbach, W.: Simulation of reactiveprocesses related to biodegradation in aquifers 1. Structure of thethree-dimensional reactive transport model, J. Contam. Hydrol.,31, 167–186, 1998.

Scheffer, M. and Carpenter, S. R.: Catastrophic regime shifts inecosystems: linking theory to observation, Trends Ecol. Evol.,18, 648–656,doi:10.1016/j.tree.2003.09.002, 2003.

Stumm, W. and Lee, G.: Oxygenation of Ferrous Iron, Indust. Eng.Chem. Res., 53, 143–146, 1961.

Stumm, W. and Morgan, J. J.: Aquatic chemistry: chemical equi-libria and rates in natural waters, John Wiley & Sons Inc., NewYork, USA, 1996.

Tetzlaff, D., Soulsby, C., Bacon, P. J., and Youngson, A. F.: Con-nectivity between landscapes and riverscapes – a unifying themein integrating hydrology and ecology in catchment science?, Hy-drol. Process., 1389, 1385–1389, 2007.

Thompson, S. E., Harman, C. J., Troch, P. A., Brooks, P. D.,and Sivapalan, M.: Spatial scale dependence of ecohydrologi-cally mediated water balance partitioning: A synthesis frame-work for catchment ecohydrology, Water Resour. Res., 47, 1–20,doi:10.1029/2010WR009998, 2011a.

Thompson, S. E., Basu, N. B., Lascurain, J., Aubeneau, A., andRao, P. S. C.: Relative dominance of hydrologic versus biogeo-chemical factors on solute export across impact gradients, WaterResour. Res., 47, W00J05,doi:10.1029/2010WR009605, 2011b.

Todd, D. and Mays, L.: Groundwater Hydrology, John Wi-ley & Sons, Chichester, 2005.

Vache, K. B. and McDonnell, J. J.: A process-based rejectionistframework for evaluating catchment runoff model structure, Wa-ter Resour. Res., 42, 1–15,doi:10.1029/2005WR004247, 2006.

Wehrer, M. and Totsche, K. U.: Detection of non-equilibrium con-taminant release in soil columns: Delineation of experimentalconditions by numerical simulations, J. Plant Nutr. Soil Sci., 166,475–483,doi:10.1002/jpln.200321095, 2003.

Werner, T. M. and Kadlec, R. H.: Wetland residence time distribu-tion modeling, Science, 15, 77–90, 2000.

Zehnder, A. and Stumm, W.: Geochemistry and biogeochemistry ofanaerobic habitats, in: Biology of Anaerobic Organisms, editedby: Zehnder, A., Wiley, New York, 1–38, 1988.

Hydrol. Earth Syst. Sci., 17, 1133–1148, 2013 www.hydrol-earth-syst-sci.net/17/1133/2013/