01.12.2012 Views

Applied numerical modeling of saturated / unsaturated flow and ...

Applied numerical modeling of saturated / unsaturated flow and ...

Applied numerical modeling of saturated / unsaturated flow and ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

<strong>Applied</strong> <strong>numerical</strong> <strong>modeling</strong> <strong>of</strong> <strong>saturated</strong> /<br />

un<strong>saturated</strong> <strong>flow</strong> <strong>and</strong> reactive contaminant transport<br />

-<br />

evaluation <strong>of</strong> site investigation strategies<br />

<strong>and</strong> assessment <strong>of</strong> environmental impact<br />

Dissertation<br />

zur Erlangung des Grades eines Doktors der Naturwissenschaften<br />

der Geowissenschaftlichen Fakultät<br />

der Eberhard-Karls-Universität Tübingen<br />

vorgelegt von<br />

Christ<strong>of</strong> Beyer<br />

aus Braunschweig<br />

2007<br />

1


2<br />

Tag der mündlichen Prüfung: 23.02.2007<br />

Dekan: Pr<strong>of</strong>. Dr. Peter Grathwohl<br />

1. Berichterstatter: Pr<strong>of</strong>. Dr.-Ing. Olaf Kolditz<br />

2. Berichterstatter: Priv. Doz. Dr. rer. nat. Sebastian Bauer<br />

3. Berichterstatter: Pr<strong>of</strong>. James F. Barker


Abstract<br />

<strong>Applied</strong> <strong>numerical</strong> <strong>modeling</strong> <strong>of</strong> <strong>saturated</strong> /<br />

un<strong>saturated</strong> <strong>flow</strong> <strong>and</strong> reactive contaminant transport -<br />

evaluation <strong>of</strong> site investigation strategies<br />

<strong>and</strong> assessment <strong>of</strong> environmental impact<br />

CHRISTOF BEYER<br />

In this thesis <strong>numerical</strong> models <strong>of</strong> variably <strong>saturated</strong> <strong>flow</strong> <strong>and</strong> reactive transport<br />

processes in porous media are employed as assessment tools in two different fields <strong>of</strong><br />

application.<br />

The first thematic complex studied is the computer based evaluation <strong>of</strong> investigation<br />

strategies for contaminated soils <strong>and</strong> aquifers. For this purpose the “Virtual Aquifer”<br />

(VA) concept is introduced <strong>and</strong> demonstrated by an assessment <strong>of</strong> the so called center<br />

line method for contaminant plume investigation. Errors <strong>and</strong> uncertainties in the<br />

estimation <strong>of</strong> contaminant degradation rates from center line data, which is <strong>of</strong>ten<br />

collected at field sites, are analysed. This is done by application at synthetic<br />

contaminated aquifers models, which are generated in the computer by <strong>numerical</strong><br />

simulation <strong>of</strong> contaminant spreading. Monte Carlo simulations <strong>and</strong> sensitivity studies<br />

are performed to quantify the influences <strong>of</strong> sampling error magnitude, aquifer<br />

heterogeneity <strong>and</strong> model parameterisation on the estimated rate constants. In a second<br />

application example, the VA concept is used for the development <strong>and</strong> testing <strong>of</strong> a new<br />

approach for biodegradation parameter estimation. The performance <strong>of</strong> this method is<br />

studied in heterogeneous synthetic aquifers. Also, the propagation <strong>of</strong> errors <strong>and</strong><br />

uncertainties from estimated rate parameters to a prognosis <strong>of</strong> the contaminant plume<br />

lengths is studied to obtain an indicator for the significance <strong>of</strong> the estimated degradation<br />

potential. The performance <strong>of</strong> the new parameter estimation method is assessed by<br />

comparison to frequently used approximations by first order kinetics.<br />

The second thematic complex addressed in this thesis is the evaluation <strong>of</strong> environmental<br />

impact <strong>of</strong> contaminant emissions from road constructions by type scenario modelling.<br />

Two general concepts for modelling <strong>of</strong> <strong>flow</strong> <strong>and</strong> transport, i.e. the Eulerian <strong>and</strong> the<br />

Lagrangian points <strong>of</strong> view, are combined in this study to assess the extent <strong>of</strong> leaching<br />

from contaminated demolition waste within road structures to the groundwater surface.<br />

Transport simulations are performed for a number <strong>of</strong> typical subsoil units <strong>of</strong> Germany<br />

to analyse the sensitivity <strong>of</strong> contaminant transport behaviour on subsoil properties. The<br />

relevant contaminant transport <strong>and</strong> attenuation processes in road constructions <strong>and</strong> the<br />

different subsoils are identified. The study allows to draw important conclusions on how<br />

these mechanisms could be used or enhanced to reduce contaminant leaching to groundwater.<br />

3


Angew<strong>and</strong>te numerische Modellierung der gesättigten / ungesättigten<br />

Strömung und des reaktiven Schadst<strong>of</strong>ftransports-<br />

Evaluierung von Strategien der St<strong>and</strong>orterkundung<br />

und Umweltwirkungsprognose<br />

Kurzfassung<br />

CHRISTOF BEYER<br />

Diese Arbeit stellt die Verwendung numerischer Modelle gesättigt / ungesättigter<br />

Strömungs- und reaktiver St<strong>of</strong>ftransportprozesse in porösen Medien als<br />

Bewertungsinstrument in zwei verschiedenen Anwendungsfeldern vor.<br />

Der erste hier betrachtete Themenkomplex beh<strong>and</strong>elt die computerbasierte Evaluierung<br />

von Erkundungsstrategien für kontaminierte Böden und Grundwasserleiter. Hierzu wird<br />

das Konzept der „Virtuellen Aquifere“ (VA) eingeführt und anh<strong>and</strong> einer Bewertung<br />

der sogenannten Center-Line Methodik zur Schadst<strong>of</strong>ffahnenerkundung demonstriert.<br />

Die Center-Line Methode wird in der Praxis häufig zur Erhebung von Daten<br />

angewendet, auf deren Grundlage Schadst<strong>of</strong>f-Abbauratenkonstanten abgeschätzt<br />

werden können. Die Bewertung dieser Vorgehensweise wird durch eine Analyse der<br />

dabei auftretenden Fehler und Unsicherheiten vorgenommen, indem die Methode bei<br />

synthetischen kontaminierten Aquifermodellen angewendet wird, die am Computer<br />

durch numerische Simulation der Schadst<strong>of</strong>fausbreitung generierten wurden. Durch<br />

Monte-Carlo Simulationen und Sensitivitätsanalysen werden die Einflüsse von<br />

Messfehlern, Aquifer-Heterogenität und Modellparametrisierung quantifiziert. In einem<br />

zweiten Anwendungsbeispiel wird die VA-Methodik zur Ableitung und Prüfung eines<br />

neuen Ansatzes zur Schätzung von Parametern mikrobieller Abbaukinetiken<br />

angewendet. Die Eignung des Verfahrens wird in synthetischen heterogenen Aquiferen<br />

geprüft. Darüber hinaus wird die Fehler- und Unsicherheitspropagation von den<br />

geschätzten Abbauparametern zu mit diesen prognostizierte Schadst<strong>of</strong>ffahnenlängen<br />

untersucht, um Indikatoren für die Aussagekraft der Parameter zu erhalten. Das<br />

Verfahren wird durch einen Vergleich mit häufig angewendeten Näherungen durch<br />

Kinetiken erster Ordnung bewertet.<br />

Der zweite im Rahmen dieser Arbeit beh<strong>and</strong>elte Themenkomplex ist die Bewertung der<br />

Umweltauswirkungen von Schadst<strong>of</strong>fausträgen aus Straßenbauten durch Typszenarienmodellierung.<br />

Hier werden zwei allgemeine Konzepte zur Modellierung von Strömung<br />

und Transport, der Eulerische und der Larangesche Ansatz, mitein<strong>and</strong>er kombiniert.<br />

Das Ausmaß des Schadst<strong>of</strong>feintrags aus belastetem im Straßenbau eingesetztem<br />

Recycling-Bauschutt ins Grundwasser wird durch Einsatz der mit ein<strong>and</strong>er gekoppelten<br />

numerischen Modelle GeoSys/Rock<strong>flow</strong> und SMART bewertet. Die Transportsimulationen<br />

werden für eine Reihe typischer Unterbodeneinheiten Deutschl<strong>and</strong>s<br />

durchgeführt, um die Sensitivität des Schadst<strong>of</strong>ftransportverhaltens auf die<br />

Unterbodeneigenschaften zu untersuchen. Die relevanten Schadst<strong>of</strong>ftransport- und<br />

Attenuierungsprozesse im Straßenaufbau und den Unterböden werden identifiziert. Die<br />

Studie ermöglicht wichtige Schlüsse im Hinblick eine Nutzung und Förderung dieser<br />

Mechanismen zu einer weiteren Reduktion der Schadst<strong>of</strong>feinträge ins Grundwasser.<br />

5


Vorwort<br />

Die zurückliegenden drei Jahre am ZAG in Tübingen waren für mich eine sehr spannende Zeit<br />

und nicht nur im Hinblick auf meine Promotion sehr lehrreich. Aus diesem Grund möchte ich<br />

hier die Gelegenheit nutzen, ein paar persönliche Worte zu verlieren und einigen Personen zu<br />

danken, ohne deren Beteiligung diese Arbeit so sicher nicht hätte entstehen können.<br />

Die im Rahmen dieser Arbeit vorgestellten Ergebnisse wurden innerhalb der beiden Projekte<br />

“Virtueller Aquifer” (Förderkennzeichen 033 05 12/033 05 13) und “Übertragung der Ergebnisse<br />

des BMBF - Förderschwerpunktes „Sickerwasserprognose“ auf repräsentative Fallbeispiele”<br />

(Förderkennzeichen 02WP0517) erarbeitet. Beide Projekte wurden durch das Bundesministerium<br />

für Bildung und Forschung (BMBF) finanziell gefördert, w<strong>of</strong>ür diesem hiermit<br />

gedankt sei.<br />

Besonders bedanken möchte ich mich bei den Betreuern dieser Arbeit. Herrn Pr<strong>of</strong>. Dr.-Ing. Olaf<br />

Kolditz danke ich sehr für die Ermöglichung meiner Promotion, das in mich gesetztes Vertrauen<br />

und die stete Unterstützung. Seine immer optimistische Sicht der Dinge hat sich im Laufe der<br />

Zeit auch ein Wenig auf mich übertragen. Herr Dr. Sebastian Bauer ist mit Sicherheit die<br />

Person, die die Richtung meiner Arbeit am stärksten mitgeprägt hat. Sein aufrichtiges Interesse<br />

(auch abseits der fachlichen Dinge), viele ausführliche Diskussionen und immer konstruktive<br />

Kritik haben maßgeblich zu ihrem Gelingen beigetragen. Eine bessere Betreuung für seine<br />

Doktorarbeit kann man sich eigentlich kaum wünschen.<br />

Herrn Pr<strong>of</strong>. Dr. James F. Barker danke ich für die Erstellung des dritten Gutachtens dieser<br />

Arbeit.<br />

Herzlich danken möchte ich Herrn Pr<strong>of</strong>. Dr. Peter Grathwohl und Herrn Pr<strong>of</strong>. Dr. Rudolf Liedl,<br />

die mir die Bearbeitung des Sickerwasserprognose-Projekts anvertrauten und mir so zu einem in<br />

jeder Hinsicht spannenden letzten Dreivierteljahr verhalfen.<br />

Ein herzlicher Dank geht an meine Kollegen Dr. Cui Chen, Dipl.-Geol. Jan Gronewold, Dr.<br />

Wilfried Konrad, Dr. Thomas Kalbacher, Dr. Chan Hee Park, Dr. Wenqing Wang, Dr. Chris<br />

McDermott, Dr. Martin Beinhorn, Robert Walsh, M.Sc., Dr. Dirk Schäfer von der Uni Kiel, Dr.<br />

Hermann Rügner und Dr. Peter Dietrich vom UFZ in Leipzig, sowie an zahlreiche weitere<br />

Kollegen für angenehm reibungslose Zusammenarbeit, viele hilfreiche Diskussionen und<br />

gemeinsame Pausen vom Zahlen hin und her schubsen.<br />

Mein größter Dank gilt meinen Eltern, da Ihr mir durch Eure Unterstützung diesen Weg erst<br />

ermöglicht habt, und Frauke für Deine große Geduld, Dein Verständnis, Deine Liebe und dafür,<br />

dass Du es trotz aller Umstände und Entfernungen gemeinsam mit mir bis hierhin geschafft<br />

hast.<br />

7


Table <strong>of</strong> Contents<br />

List <strong>of</strong> abbreviations <strong>and</strong> mathematical symbols<br />

1. Introduction 1<br />

2. Mathematical models 2<br />

2.1. Saturated / un<strong>saturated</strong> <strong>flow</strong> 2<br />

2.2. Transport processes 3<br />

2.3. Reactive processes 4<br />

2.4. Numerics <strong>and</strong> s<strong>of</strong>tware methods 6<br />

3. Modeling applications 9<br />

3.1. Evaluation <strong>of</strong> investigation strategies for contaminated aquifers using<br />

the Virtual Aquifer concept 9<br />

3.2. Development <strong>and</strong> testing <strong>of</strong> a new approach to estimating biodegradation<br />

parameters from field data 16<br />

3.3. Prognosis <strong>of</strong> long term contaminant leaching from recycling materials in<br />

road constructions 18<br />

4. Conclusions <strong>and</strong> outlook 21<br />

References<br />

Enclosed Publications<br />

23<br />

9


List <strong>of</strong> abbreviations <strong>and</strong> mathematical symbols<br />

a empirical sorption constant [-]<br />

b empirical sorption constant [-]<br />

C concentration [M L -3 ]<br />

C0 equilibrium concentration [M L -3 ]<br />

Cl liquid phase concentration [M L -3 ]<br />

Cs solid phase concentration [M M -1 ]<br />

Corg soil organic carbon content [%]<br />

Cw(�) water capacity function<br />

D tensor <strong>of</strong> hydrodynamic dispersion<br />

[L² T -1 ]<br />

Da aqueous molecular diffusion<br />

coefficient [L² T -1 ]<br />

Dae effective diffusion coefficient [L² T -1 ]<br />

Dap apparent diffusion coefficient [L² T -1 ]<br />

De tensor <strong>of</strong> effective hydrodynamic<br />

dispersion [L² T -1 ]<br />

Dm tensor <strong>of</strong> mechanical dispersion<br />

[L² T -1 ]<br />

DW demolition waste<br />

FE finite element<br />

FEM finite element method<br />

g travel time probability density<br />

function<br />

h hydraulic head [L]<br />

h´ erroneous head measurement [L]<br />

I identity tensor<br />

IC inhibition concentration [M L -3 ]<br />

K tensor <strong>of</strong> hydraulic conductivity<br />

[L T -1 ]<br />

kmax maximum degradation rate<br />

[M L -3 T -1 ]<br />

l pore connectivity parameter [-]<br />

m empirical Van Genuchten<br />

parameter [-]<br />

MC half-saturation concentration [M L -3 ]<br />

MM Michaelis-Menten<br />

n empirical Van Genuchten<br />

parameter [-]<br />

OOP object-oriented programming<br />

PDE partial differential equation<br />

pdf probability density function<br />

Q source or sink term [M L -3 T -1 ]<br />

r radial distance [L]<br />

REV representative elementary volume<br />

S specific storativity [L -1 ]<br />

Sr relative saturation [-]<br />

t time [T]<br />

V magnitude <strong>of</strong> velocity vector<br />

[L T -1 ]<br />

vmax maximum growth rate [T -1 ]<br />

v average linear velocity [L T -1 ]<br />

v vector <strong>of</strong> average linear velocity<br />

[L T -1 ]<br />

VA Virtual Aquifer<br />

X microbial population [M L -3 ]<br />

Y yield coefficient [-]<br />

z elevation [L]<br />

� empirical Van Genuchten<br />

parameter [L -1 ]<br />

�L<br />

�T<br />

longitudinal dispersivity [L]<br />

transverse dispersivity [L]<br />

� reaction function<br />

� tortuosity related coefficient [-]<br />

� r<strong>and</strong>om number<br />

� intraparticle porosity [-]<br />

��h maximum measurement error [L]<br />

� empirical sorption constant [-]<br />

�e effective porosity [-]<br />

� volumetric water content [-]<br />

�r residual water content [-]<br />

<strong>saturated</strong> water content [-]<br />

�s<br />

� empirical sorption constant<br />

[M 1-� L 3� M -1 ]<br />

� first order degradation rate<br />

constant [T -1 ]<br />

� microbial decay term<br />

� density [M L -3 ]<br />

2<br />

� Y ln(K) variance<br />

� travel time [T]<br />

tortuosity factor [-]<br />

�f<br />

� matric head [L]<br />

11


<strong>Applied</strong> <strong>numerical</strong> <strong>modeling</strong> <strong>of</strong> <strong>saturated</strong> / un<strong>saturated</strong><br />

<strong>flow</strong> <strong>and</strong> reactive contaminant transport –<br />

evaluation <strong>of</strong> site investigation strategies <strong>and</strong> assessment<br />

<strong>of</strong> environmental impact<br />

1. Introduction<br />

In the field <strong>of</strong> subsurface hydrology, mathematical<br />

models are used to simulate fluid<br />

<strong>flow</strong> <strong>and</strong> solute transport by translating<br />

physical <strong>and</strong> biogeochemical processes into<br />

mathematical equations, which can be<br />

solved by either analytical or <strong>numerical</strong><br />

methods. Underst<strong>and</strong>ing <strong>of</strong> individual processes<br />

in domains <strong>of</strong> simple geometry is<br />

already a challenging task for itself. In large<br />

scale applications, however, we are faced<br />

with heterogeneous environments <strong>and</strong> interactions<br />

<strong>of</strong> many different types <strong>of</strong> spatially<br />

<strong>and</strong> temporally variable processes, which<br />

leave the <strong>numerical</strong> treatment <strong>of</strong> such complex<br />

coupled problems <strong>of</strong>ten as the only<br />

way to reach meaningful conclusions<br />

(Zheng <strong>and</strong> Bennett, 1995). As a benefit <strong>of</strong><br />

the rapid development <strong>of</strong> computational capabilities<br />

(e.g. high performance parallel<br />

computing, specialized s<strong>of</strong>tware implementation<br />

methods, data pre- <strong>and</strong> post-processing<br />

tools, graphical display routines) the<br />

<strong>numerical</strong> simulation <strong>of</strong> complex coupled<br />

problems in subsurface hydrology is continuously<br />

advanced. In general, <strong>numerical</strong><br />

models <strong>of</strong> <strong>flow</strong> <strong>and</strong> transport in geosystems<br />

are used as tools for<br />

� qualitative <strong>and</strong> quantitative analysis <strong>of</strong><br />

single or coupled processes<br />

� identification <strong>of</strong> relevant parameters<br />

� parameter estimation / inverse <strong>modeling</strong><br />

� sensitivity <strong>and</strong> uncertainty analysis<br />

� prediction <strong>of</strong> system response to changes<br />

in initial or boundary conditions<br />

These capabilities as well as the spectrum<br />

<strong>of</strong> application would not have been<br />

achieved without the ever growing dem<strong>and</strong><br />

for groundwater resources <strong>and</strong> the concern<br />

about its quality. Groundwater is one <strong>of</strong> the<br />

main drinking water supplies <strong>and</strong> increasingly<br />

used for agricultural field irrigation<br />

(Morris et al., 2003). At the same time<br />

pollution from industrial activities, waste<br />

disposal, agricultural use <strong>of</strong> fertilizers or<br />

pesticides <strong>and</strong> urban waste waters (Scheidleder<br />

et al., 1999) poses a serious threat to<br />

our groundwater resources.<br />

In this thesis <strong>numerical</strong> <strong>modeling</strong> is used in<br />

three applications within the context <strong>of</strong> contaminant<br />

hydrology. The term applied <strong>numerical</strong><br />

<strong>modeling</strong> here emphasizes the field<br />

scale application <strong>of</strong> computational methods<br />

to obtain solutions to systems <strong>of</strong> partial differential<br />

equations describing <strong>flow</strong> <strong>and</strong><br />

reactive transport process interactions in<br />

porous media. Application 1 introduces the<br />

“Virtual Aquifer” (VA) concept, in which<br />

<strong>numerical</strong> <strong>modeling</strong> is used as a tool for the<br />

evaluation <strong>of</strong> investigation <strong>and</strong> remediation<br />

strategies for contaminated soils <strong>and</strong> aquifers.<br />

The concept is demonstrated by an<br />

assessment <strong>of</strong> the so called center line method<br />

for site investigation. In application 2<br />

the VA concept is applied for the development<br />

<strong>and</strong> testing <strong>of</strong> a new approach for biodegradation<br />

parameter estimation. Both<br />

applications make use <strong>of</strong> the GeoSys /<br />

Rock<strong>flow</strong> code (Kolditz et al., 2006) for the<br />

<strong>numerical</strong> simulations. In application 3 the<br />

Eulerian <strong>and</strong> Lagrangian concepts for contaminant<br />

transport <strong>modeling</strong> are combined.<br />

GeoSys / Rock<strong>flow</strong> is coupled with the<br />

SMART model (Finkel et al., 1998) <strong>and</strong><br />

used for type scenario <strong>modeling</strong> to assess<br />

the environmental impact <strong>of</strong> recycling<br />

materials in road constructions.<br />

This synthesis is organized as follows:<br />

Chapter 2 presents the mathematical process<br />

models <strong>and</strong> <strong>numerical</strong> schemes used<br />

1


for the application studies outlined. The<br />

chapter is based on fundamental publications<br />

in the field <strong>of</strong> computational hydrology.<br />

However, it is not intended to<br />

serve as a comprehensive review <strong>of</strong> processes<br />

<strong>and</strong> model concepts, as this would be<br />

beyond the scope <strong>of</strong> this synthesis. Chapter<br />

3 <strong>and</strong> its subsections present the results<br />

<strong>and</strong> conclusions <strong>of</strong> the three application<br />

examples. Chapter 4 closes this synthesis<br />

with general conclusions <strong>and</strong> an outlook.<br />

2. Mathematical models<br />

The three-dimensional structure <strong>of</strong> natural<br />

porous media is manifested in its composition<br />

<strong>of</strong> three constituting phases, i.e. the<br />

solid (mineral or biophase), water <strong>and</strong> gaseous<br />

phases. At scales larger than the pore<br />

scale, a description <strong>of</strong> porous medium geometry<br />

becomes very complex <strong>and</strong> thus infeasible<br />

for <strong>modeling</strong> applications. To<br />

underst<strong>and</strong> <strong>and</strong> formulate the dynamics <strong>of</strong><br />

fluids in the subsurface the so called representative<br />

elementary volume (REV) concept<br />

is introduced (Bear, 1972): In the<br />

transition from the microscale to a larger<br />

macroscale material, parameters are averaged<br />

over a volume which is sufficiently<br />

large to describe the porous medium at that<br />

larger scale (see Fig. 1).<br />

Fig. 1: Representative elementary volume<br />

concept (Bear <strong>and</strong> Bachmat, 1990).<br />

This may also require a reformulation <strong>of</strong> the<br />

mathematical process descriptions. The<br />

derivation <strong>of</strong> representative or effective<br />

new material parameters <strong>and</strong> the corresponding<br />

governing equations is termed<br />

2<br />

upscaling. Within the REV the detailed<br />

structure <strong>of</strong> the medium is lost <strong>and</strong> becomes<br />

a continuous field. Parameters like porosity,<br />

permeability or dispersivity are considered<br />

constant over the averaging volume. In the<br />

following sections material parameters <strong>and</strong><br />

governing equations are based on this<br />

continuum approach.<br />

2.1. Saturated / un<strong>saturated</strong> <strong>flow</strong><br />

The dynamics <strong>of</strong> the water in fully <strong>saturated</strong><br />

three dimensional porous media can be described<br />

by the combination <strong>of</strong> the mass<br />

balance for the water phase (eq. (1)) <strong>and</strong><br />

Darcy’s law (eq. (2)) as a constitutive equation<br />

(Bear, 1972)<br />

�h<br />

S � �� �q<br />

� Q<br />

(1)<br />

�t<br />

q � �K�h<br />

(2)<br />

where S [m -1 ] is specific storativity, h [m] is<br />

the hydraulic head, given as the sum <strong>of</strong> elevation<br />

z [m] <strong>and</strong> the pressure head � [m], t<br />

[s] is time, q [m s -1 ] is the Darcy flux vector,<br />

K [m s -1 ] is the tensor <strong>of</strong> hydraulic conductivity<br />

<strong>and</strong> Q [s -1 ] is a source or sink<br />

term. The governing equation for groundwater<br />

<strong>flow</strong> under transient conditions is<br />

thus given by (Bear, 1972)<br />

�h<br />

S � � � �K �h��Q<br />

(3)<br />

�t<br />

which at steady state converts to<br />

� �h���Q<br />

� � K (4).<br />

This model <strong>of</strong> steady state <strong>flow</strong> in <strong>saturated</strong><br />

porous media is employed in applications 1<br />

<strong>and</strong> 2 (sections 3.1, 3.2) <strong>of</strong> this synthesis.<br />

For un<strong>saturated</strong> conditions, which are prevalent<br />

in application 3 (section 3.3), a more<br />

general form <strong>of</strong> eq. (2), the Buckingham-<br />

Darcy-law, can be used (Jury et al., 1991)<br />

q � �K(<br />

� ) �h<br />

(5)<br />

where K is a function <strong>of</strong> the pressure (or<br />

matric) head �, which itself depends on the<br />

volumetric water content � [-] <strong>of</strong> the porous<br />

medium. As for <strong>saturated</strong> conditions, the


Buckingham-Darcy law is combined with<br />

mass balance principles to yield the governing<br />

equation <strong>of</strong> <strong>flow</strong> for un<strong>saturated</strong><br />

conditions, i.e. the Richards equation. This<br />

equation exists in three main forms with �,<br />

� or both quantities as dependent variables<br />

(Jury et al., 1991). In GeoSys / Rock<strong>flow</strong><br />

the �-based form (Freeze <strong>and</strong> Cherry,<br />

1979) is implemented (Du et al., 2005)<br />

C w<br />

��<br />

�t<br />

��� � � � �K ( � ) ��<br />

�<br />

�K<br />

�<br />

�z<br />

��� � Q<br />

(6)<br />

where Cw(�) is the water capacity function<br />

defined by d� /d� <strong>and</strong> with z positive in a<br />

downward direction. For the functional<br />

description <strong>of</strong> un<strong>saturated</strong> hydraulic properties<br />

different mathematical formulations<br />

have been proposed in literature. A frequently<br />

used constitutive relation is the<br />

Van-Genuchten-Mualem model (Van Genuchten,<br />

1980) which is based on the<br />

statistical pore space model <strong>of</strong> Mualem<br />

(1976) <strong>and</strong> is given by<br />

� � � �2 �m<br />

m<br />

1�<br />

1�<br />

Sr<br />

l<br />

K(<br />

S ) � KS<br />

(7)<br />

Sr<br />

r<br />

� ��<br />

r � �<br />

� ��<br />

s<br />

r<br />

r<br />

1<br />

� � � �m n<br />

1�<br />

��<br />

(8)<br />

m � 1� 1/<br />

n<br />

(9)<br />

where Sr [-] is defined as the relative<br />

saturation, l [-] is a pore connectivity parameter,<br />

�, �r <strong>and</strong> �s are the actual, residual<br />

<strong>and</strong> <strong>saturated</strong> volumetric water contents, �<br />

[m -1 ], n [-], <strong>and</strong> m [-] are empirical parameters.<br />

Other constitutive relationships comprise<br />

approaches such as the Brooks-Corey<br />

model (Brooks <strong>and</strong> Corey, 1966), the Haverkamp<br />

model (Haverkamp et al., 1977), potential<br />

functions as introduced by Huyakorn <strong>and</strong><br />

Pinder (1983) or the multimodal model <strong>of</strong><br />

Durner (1994). Recently, also free form parameterizations<br />

were suggested (Bitterlich et al.,<br />

2004). An overview on the prevalent approaches<br />

is given e.g. by Durner <strong>and</strong> Flühler (2005).<br />

2.2. Transport processes<br />

The most fundamental transport process <strong>of</strong><br />

dissolved substances in <strong>saturated</strong> porous<br />

media is advection. Advection is passive<br />

with the <strong>flow</strong>ing water. Purely advective<br />

transport <strong>of</strong> a solute plume is free <strong>of</strong> interference<br />

or mixing with the surrounding<br />

ambient water <strong>and</strong> is described with the advection<br />

equation (Zheng <strong>and</strong> Bennett, 1995)<br />

�C<br />

�t<br />

� �v<br />

�C<br />

� Q<br />

(10)<br />

where C [kg m -3 ] is the concentration <strong>of</strong> a<br />

dissolved species, v [m s -1 ] is the vector <strong>of</strong><br />

average linear velocity which is given by<br />

division <strong>of</strong> q with the effective porosity �e<br />

[-], <strong>and</strong> Q [kg m -3 s -1 ] is a source or sink<br />

term for species C.<br />

In natural aquifers or soils, purely advective<br />

transport is practically not established as<br />

dissolved molecules migrate from high to<br />

low concentration regions by Brownian<br />

motion. This concentration gradient driven<br />

mass transport is termed molecular diffusion<br />

<strong>and</strong> occurs even when the fluid itself is<br />

stagnant. For transient systems the diffusion<br />

process in water can be described using<br />

Fick’s 2 nd law, (Fetter, 1993)<br />

2<br />

�C<br />

� C<br />

� �Da<br />

2<br />

�t<br />

�x<br />

(11)<br />

where Da is the molecular diffusion coefficient<br />

in water [m 2 s -1 ]. In porous media the<br />

diffusion process is hindered by the presence<br />

<strong>of</strong> the solid phase matrix <strong>and</strong> the tortous<br />

nature <strong>of</strong> the pores. Thus an effective<br />

diffusion coefficient Dae is derived as<br />

(Grathwohl, 1998)<br />

D<br />

� �<br />

e<br />

ae a<br />

f<br />

D � (12)<br />

�<br />

where � [-] is the constrictivity <strong>and</strong> �f [-] the<br />

tortuosity <strong>of</strong> the porous medium. Under<br />

un<strong>saturated</strong> conditions Dae can also be<br />

related to � (e.g. Olsen <strong>and</strong> Kemper, 1968).<br />

As water moves through a porous medium,<br />

single streamline velocities can be greater<br />

or less than v. This effect is due to different<br />

3


path lengths <strong>of</strong> water molecules that bypass<br />

mineral grains <strong>of</strong> different size <strong>and</strong> shape,<br />

different pore diameters as well as inner<br />

pore friction which results in velocity contrasts<br />

along pore cross sections. The consequential<br />

divergence <strong>of</strong> transport velocities<br />

for dissolved solutes causes mixing with the<br />

ambient water along the <strong>flow</strong> path <strong>and</strong> thus<br />

results in solute spreading longitudinally<br />

<strong>and</strong> transversally to the main <strong>flow</strong> direction.<br />

This process is termed mechanical dispersion.<br />

For its mathematical description usually<br />

an analogy to the diffusion process is<br />

assumed. According to Bear (1972) the tensor<br />

<strong>of</strong> mechanical dispersion Dm is given by<br />

D m<br />

4<br />

� � I � �<br />

��L�� V<br />

TV<br />

T<br />

vv<br />

(13)<br />

where �L <strong>and</strong> �T [m] are longitudinal <strong>and</strong><br />

transverse dispersivities, V is the magnitude<br />

<strong>of</strong> the velocity vector, I is the identity<br />

tensor <strong>and</strong> vv is the dyadic <strong>of</strong> the velocity<br />

vector. The tensor <strong>of</strong> hydrodynamic dispersion<br />

D [m 2 s -1 ] combines the dispersion <strong>and</strong><br />

diffusion processes <strong>and</strong> is calculated by<br />

D m ae<br />

� D � D I<br />

(14).<br />

The mathematical formulation <strong>of</strong> advectivedispersive<br />

transport in fully <strong>saturated</strong> porous<br />

media assuming constant porosity is<br />

given by the sum <strong>of</strong> the advective <strong>and</strong><br />

dispersive fluxes, i.e. the advection-dispersion-equation<br />

(Zheng <strong>and</strong> Bennett, 1995)<br />

�C<br />

� �� � �v C�����D�C��Q<br />

(15).<br />

�t<br />

For un<strong>saturated</strong> conditions the total solute<br />

flux in the water phase is described by<br />

��C<br />

� �� � �q�C������De�C��Q(16) �t<br />

where the effective hydrodynamic disper�<br />

sion tensor De is used, as besides Dae also<br />

�L <strong>and</strong> �T depend on � (Bear, 1979).<br />

2.3. Reactive processes<br />

The source or sink terms Q in eq. (15) <strong>and</strong><br />

(16) represent a large variety <strong>of</strong> processes<br />

other than advection or hydrodynamic dispersion,<br />

which may cause temporal changes<br />

in the solute concentration C. Hence, Q<br />

may represent transfer <strong>of</strong> species between<br />

solid, water, gaseous or biophase (e.g. volatilization,<br />

non aqueous phase liquid dissolution,<br />

sorption), or equilibrium <strong>and</strong> kinetic<br />

reactions <strong>of</strong> (geo-)chemical or biochemical<br />

nature. According to Rubin (1983) (Fig. 2)<br />

reactive processes can be classified by<br />

� reaction velocity <strong>and</strong> reversibility (equilibrium<br />

or non-equilibrium; level A)<br />

� involvement <strong>of</strong> only a single or several<br />

phases (homogeneous / heterogeneous;<br />

level B)<br />

� reaction type: surface (e.g. sorption) or<br />

“classical” chemical reaction (level C)<br />

For the sake <strong>of</strong> brevity, here only process<br />

concepts relevant for the model applications<br />

<strong>of</strong> chapter 3 are explained in more detail.<br />

Fig. 2: Classification <strong>of</strong> reactions in porous<br />

media (Rubin, 1983).<br />

Kinetic sorption<br />

Transfer <strong>of</strong> dissolved species from the<br />

mobile phase to the solid matrix by<br />

physico-chemical processes is termed<br />

sorption. Sorbed species are immobilized<br />

<strong>and</strong> not transported with the water flux.<br />

Sorption is a reversible process, i.e. sorbed<br />

species can be remobilised by desorption.<br />

Temporary immobilisation by sorption results<br />

in lowered solute concentrations <strong>and</strong><br />

retarded transport velocities. The manifold<br />

processes contributing to sorption phenomena<br />

include physical as well as chemical<br />

mechanisms (e.g. ion exchange <strong>and</strong> surface<br />

complexation through Coulomb or van der<br />

Waals forces, hydrogen-, hydrophobic- or


covalent bonding). To mathematically describe<br />

the sorption / desorption behaviour <strong>of</strong><br />

a species so called sorption isotherms can<br />

be used, in which the sorbed amount <strong>of</strong> the<br />

species is a function <strong>of</strong> its dissolved<br />

concentration. A general nonlinear sorption<br />

model can be formulated as<br />

C<br />

s<br />

� �a�Cl� � � � abC �<br />

� 1<br />

(17)<br />

l<br />

where Cs [kg kg -1 ] is the sorbed solid phase<br />

<strong>and</strong> Cl [kg m -3 ] the liquid phase concentration,<br />

�, a,� b <strong>and</strong> � are empirical constants.<br />

For b = 0, a = 1 [-] <strong>and</strong> � = 1 [-] eq.<br />

(17) is the linear Henry isotherm, in this<br />

case � [m³ kg -1 ] is a simple equilibrium<br />

constant. For � = 1 <strong>and</strong> b = 1 eq. (17) is the<br />

Langmuir isotherm with � [kg kg -1 ] being<br />

the maximum amount <strong>of</strong> a species which<br />

can be sorbed to the solid phase <strong>and</strong> a<br />

[m³ kg -1 ] an adsorption constant. For b = 0,<br />

a = 1 [-] <strong>and</strong> � � 1 [-] (usually � < 1) the<br />

Freundlich isotherm is obtained, where �<br />

[kg 1-� m 3� kg -1 ] is the Freundlich coefficient<br />

<strong>and</strong> � [-] is the Freundlich exponent.<br />

In general, sorption processes may be treated<br />

as equilibrium reactions because sorption<br />

is fast compared to transport. However,<br />

there are exceptions to this rule because for<br />

some solute species as well as soils or<br />

aquifers, equilibration is a slow process.<br />

Possible mechanisms include slow diffusion<br />

into intraparticle pores accompanied by<br />

equilibrium sorption to surfaces within the<br />

pores or slow diffusion in organic matter<br />

(Ball <strong>and</strong> Roberts, 1991; Grathwohl, 1998;<br />

Rügner et al., 1999). Hence, from a<br />

macroscopic point <strong>of</strong> view sorption equilibrium<br />

may not be reached within the available<br />

contact time between mobile <strong>and</strong> solid<br />

phases. Assuming spherical particles, intraparticle<br />

diffusion kinetics can be described<br />

by Fick´s 2 nd law in radial coordinates (e.g.<br />

Grathwohl, 1998)<br />

2<br />

� C ��<br />

C 2 � C �<br />

� Dap�<br />

� 2<br />

� t<br />

�<br />

� � r r � r �<br />

(18)<br />

where r [m] is the radial distance from the<br />

particle center <strong>and</strong> Dap [m 2 s -1 ] is the appa-<br />

rent diffusion coefficient, which is calculated<br />

from Da by<br />

D<br />

D �<br />

a<br />

ap � (19)<br />

( � ��<br />

�)<br />

� f<br />

where � [-] is the intraparticle porosity, �<br />

the linear equilibrium sorption coefficient<br />

<strong>and</strong> � [kg m -3 ] the particle density. Other<br />

approaches to describe slow sorption / desorption<br />

kinetics comprise first- or secondorder,<br />

two-stage models (e.g. Brusseau <strong>and</strong><br />

Rao, 1989; Ma <strong>and</strong> Selim, 1994; Streck et<br />

al., 1995). A comparison <strong>of</strong> first-order <strong>and</strong><br />

diffusion limited approaches was recently<br />

published by Altfelder <strong>and</strong> Streck (2006).<br />

Kinetic degradation<br />

Degradation, whether biotic or abiotic, is<br />

the only process that reduces the overall<br />

mass <strong>of</strong> contaminants in natural porous<br />

media without transfer to other phases.<br />

Biological degradation mechanisms are numerous,<br />

complex <strong>and</strong> by far not completely<br />

understood nor even identified. The vast<br />

amount <strong>of</strong> different types <strong>of</strong> microorganisms<br />

in the subsurface provides many<br />

metabolic pathways for contaminant degradation<br />

under aerobic <strong>and</strong> anaerobic conditions.<br />

Through successive oxidation or<br />

reduction reactions contaminants can be<br />

transformed to innocuous compounds like<br />

methane, chloride, water or carbon dioxide<br />

(Wiedemeier et al. 1999). However, intermediate<br />

products can even be <strong>of</strong> significantly<br />

higher toxicity <strong>and</strong> persistence than<br />

their parent compounds (e.g., dechlorination<br />

<strong>of</strong> dichloromethane to vinyl chloride;<br />

Wiedemeier et al. 1999).<br />

Kinetic growth <strong>and</strong> decay <strong>of</strong> a microbial<br />

population X [kg m -3 ] can be described by a<br />

generalized Monod-type equation as given<br />

e.g. in Schäfer et al. (1998)<br />

�X<br />

� v<br />

�t<br />

�<br />

max<br />

X<br />

�<br />

I<br />

�<br />

I<br />

Ci<br />

M � C<br />

i Ci<br />

i<br />

C j<br />

� C<br />

j C j j<br />

��<br />

�X� (20)<br />

5


where vmax [s -1 ] is a maximum growth rate,<br />

Ci [kg m -3 ] is the concentration <strong>of</strong> the i th<br />

substrate, MCi [kg m -3 ] is the corresponding<br />

half velocity concentration [kg m -3 ], Cj<br />

[kg m -3 ] is the concentration <strong>of</strong> the j th substance<br />

inhibiting microbial growth, ICj is the<br />

corresponding inhibition concentration<br />

[kg m -3 ] <strong>and</strong> �(X) is a microbial decay term,<br />

which is <strong>of</strong>ten modeled as being <strong>of</strong> first<br />

order. Consumption <strong>of</strong> substrates or production<br />

<strong>of</strong> k metabolites Ck [kg m -3 ] is both<br />

coupled to microbial growth via (Schäfer et<br />

al., 1998)<br />

�Ck �1<br />

��X<br />

�<br />

�<br />

�t<br />

Y �<br />

� �t<br />

�<br />

�<br />

6<br />

(21)<br />

k growth<br />

where Yk [-] is the yield coefficient for<br />

substrate or metabolite Ck, <strong>and</strong> [·]growth<br />

refers to the growth term only in eq. (20).<br />

From the generalized Monod-type equation<br />

different more simple kinetic formulations<br />

can be derived. For a temporally constant<br />

microbial population, i.e. growth <strong>and</strong> decay<br />

terms are constant <strong>and</strong> <strong>of</strong> equal magnitude,<br />

no inhibition <strong>and</strong> dependence on only a<br />

single substrate, eq. (20) <strong>and</strong> (21) can be<br />

combined to yield the Michaelis-Menten<br />

(MM) kinetics model (Simkins <strong>and</strong><br />

Alex<strong>and</strong>er, 1984)<br />

dC<br />

dt<br />

C<br />

� �kmax<br />

(22)<br />

C � M C<br />

where k max is the maximum degradation rate<br />

[kg m -3 s -1 ] <strong>and</strong> M C is the MM half-saturation<br />

concentration [kg m -3 ]. This approximation<br />

may be applicable e.g. when aquifer<br />

sediments have been exposed to contaminants<br />

for several years (Bekins et al., 1998)<br />

<strong>and</strong> is used in application 2 (section 3.2).<br />

Often, contaminant degradation is also<br />

described by simple first order kinetics, e.g.<br />

to simulate abiotic degradation reactions<br />

like hydrolysis <strong>and</strong> dehydrohalogenation <strong>of</strong><br />

halogenated compounds (Wiedemeier et al.,<br />

1999). First order kinetics can be derived<br />

from eq. (22) for C > MC eq. (22)<br />

approaches zero order kinetics. The first<br />

order model is used in application 1 <strong>and</strong> 3<br />

(sections 3.1 <strong>and</strong> 3.3). Extensive reviews on<br />

kinetic models <strong>of</strong> biodegradation can be<br />

found e.g. in Baveye <strong>and</strong> Valocchi (1989),<br />

Rittmann <strong>and</strong> VanBriesen (1996) or Islam<br />

et al. (2001).<br />

2.4. Numerics <strong>and</strong> s<strong>of</strong>tware<br />

methods<br />

Numerical solution <strong>of</strong> the governing<br />

equations<br />

The governing equations for <strong>flow</strong> <strong>and</strong> reactive<br />

transport presented in sections 2.1 - 2.3<br />

belong to the group <strong>of</strong> partial differential<br />

equations (PDE), containing derivatives <strong>of</strong><br />

first order in time <strong>and</strong> <strong>of</strong> first as well as<br />

second order in space. The classification <strong>of</strong><br />

PDE can be based on mathematical aspects<br />

(highest order derivatives in the dependent<br />

variables) or on a physical point <strong>of</strong> view<br />

(problem type <strong>of</strong> physical process) (Kolditz,<br />

2002). Parabolic PDE are used for timedependent<br />

problems with dissipation process,<br />

such as diffusion (eq. (11)) or transient<br />

groundwater <strong>flow</strong> (eq. (3)), which<br />

convert to elliptic PDE for steady state<br />

conditions (eq. (4)). A third class <strong>of</strong> PDE<br />

are hyperbolic equations like the linear advection<br />

equation (eq. (10)), which are used<br />

to describe time-dependent problems without<br />

dissipation process. The transport equations<br />

(15) <strong>and</strong> (16) are <strong>of</strong> mixed type with a<br />

parabolic dispersion-diffusion term <strong>and</strong> a<br />

hyperbolic advection term. Their behaviour<br />

for a particular problem depends on the<br />

relative magnitudes <strong>of</strong> these flux components.<br />

In general the transport equations are<br />

<strong>of</strong> parabolic character which changes to<br />

hyperbolic for large ratios <strong>of</strong> v/�L as in this<br />

case the advective flux term is dominant<br />

(Kolditz, 2002).<br />

In general, PDE describing physical<br />

problems are well-posed when appropriate<br />

initial <strong>and</strong> boundary conditions are specified<br />

for the domain where a solution is<br />

required. While analytical solutions can be


found for a number <strong>of</strong> problems with<br />

simple geometries <strong>and</strong> boundary conditions<br />

(e.g. Bear, 1979; Van Genuchten <strong>and</strong> Alves,<br />

1982; Kinzelbach, 1983), for complex nonlinear<br />

problems exact solutions may not<br />

exist <strong>and</strong> thus approximate <strong>numerical</strong><br />

solutions must be obtained. Given the<br />

governing equations with appropriate initial<br />

<strong>and</strong> boundary conditions for a specified<br />

problem, the general strategy for a <strong>numerical</strong><br />

solution is to first convert the PDE into<br />

a system <strong>of</strong> discrete algebraic equations <strong>and</strong><br />

then to find the exact solution <strong>of</strong> the latter,<br />

which is the approximate solution <strong>of</strong> the<br />

PDE. An overview <strong>of</strong> approximation<br />

methods for the solution <strong>of</strong> PDE is given in<br />

Fig. 3.<br />

Fig. 3: Overview <strong>of</strong> approximation methods<br />

for PDE (Kolditz, 2002).<br />

Among the many possible approaches for<br />

the <strong>numerical</strong> simulation <strong>of</strong> fluid dynamics<br />

<strong>and</strong> transport, finite difference, finite element<br />

<strong>and</strong> finite volume methods are the<br />

most frequently used. Their theory <strong>and</strong> implementation<br />

is content <strong>of</strong> numerous text<br />

books (e.g. Pinder <strong>and</strong> Gray, 1977; Baker,<br />

1983; Zienkiewicz <strong>and</strong> Taylor, 1993;<br />

Helmig, 1997; Kolditz, 2002). Here, a short<br />

description <strong>of</strong> the finite element method<br />

(FEM) only is given, as it is the <strong>numerical</strong><br />

scheme which is implemented in the<br />

GeoSys / Rock<strong>flow</strong> code (Kolditz et al.,<br />

2006) which is used in all applications<br />

presented in chapter 3.<br />

The concept <strong>of</strong> the FEM is based on the<br />

spatial discretization (“meshing”) <strong>of</strong> continuous<br />

structures into discrete substructures,<br />

i.e. the finite elements (FE). In comparison<br />

to finite differences, the big advantage <strong>of</strong><br />

the FEM is its ability to also h<strong>and</strong>le com-<br />

plex geometries by unstructured or arbitrarily<br />

shaped grids (Anderson <strong>and</strong><br />

Woessner, 1992). The governing PDE are<br />

discretized by deriving integral formulations,<br />

e.g. by the method <strong>of</strong> weighted<br />

residuals. By introducing weighting functions<br />

the approximate solution is forced to<br />

satisfy the condition that the weighted<br />

average residual <strong>of</strong> the unknown true solution<br />

at the nodes over the computational<br />

domain is equal to zero. For any location<br />

within the domain the solution is obtained<br />

by linear combination <strong>of</strong> local interpolation<br />

functions <strong>and</strong> the solution at the nodes.<br />

With the st<strong>and</strong>ard Galerkin method<br />

(Huyakorn <strong>and</strong> Pinder, 1983) weighting <strong>and</strong><br />

interpolation functions are selected<br />

identically. The FEM is globally mass<br />

conservative, locally, however, mass<br />

conservation problems may occur. Therefore<br />

mixed FEM approaches draw increasing<br />

attention (e.g. Starke, 2000; Knabner<br />

<strong>and</strong> Schneid, 2002; Korsawe et al., 2006).<br />

The governing PDE <strong>of</strong> <strong>flow</strong> <strong>and</strong> reactive<br />

transport in porous media outlined in the<br />

previous sections are formulated from the<br />

point <strong>of</strong> view <strong>of</strong> an fixed observer with the<br />

fluid <strong>and</strong> solute moving on a fixed spatial<br />

grid. This approach is termed the Eulerian<br />

concept <strong>and</strong> is able to h<strong>and</strong>le dispersiondominated<br />

problems accurately <strong>and</strong> efficiently.<br />

For advection-dominated problems,<br />

however, the Eulerian concept is susceptible<br />

to <strong>numerical</strong> dispersion <strong>and</strong> artificial<br />

oscillations (Zheng <strong>and</strong> Bennett, 1995), For<br />

the price <strong>of</strong> increased computational effort<br />

this problem can be limited by sufficiently<br />

fine spatial <strong>and</strong> temporal discretization.<br />

By contrast, the Lagrangian concept is well<br />

suited for advection-dominated problems,<br />

but problematic when advection <strong>and</strong> dispersion<br />

must be solved together (Thorenz,<br />

2001). In the Lagrangian concept concentrations<br />

are not associated with fixed spatial<br />

points but rather with moving particles, that<br />

are transported with the prevailing <strong>flow</strong><br />

velocity. The <strong>numerical</strong> model SMART<br />

(Finkel et al., 1998) is based on the<br />

Langrangian concept <strong>of</strong> Cvetkovic <strong>and</strong><br />

Dagan (1996). In SMART the model<br />

7


domain is discretized along advective <strong>flow</strong><br />

paths <strong>of</strong> the Eulerian <strong>flow</strong> field by the travel<br />

time � [s] <strong>of</strong> inert particles between an<br />

injection plane <strong>and</strong> a control plane, both<br />

oriented normal to the mean <strong>flow</strong> direction.<br />

Each particle trajectory is regarded as a<br />

separate one-dimensional stream tube <strong>of</strong> the<br />

<strong>flow</strong> field with an infinitesimal cross<br />

section. The probability density function<br />

(pdf) g(�, x) <strong>of</strong> all particles travel times<br />

completely reflects all hydraulic heterogeneities<br />

<strong>of</strong> the model domain. Influences <strong>of</strong><br />

reactions (e.g. biodegradation, intraparticle<br />

diffusion, sorption, etc.) are quantified by<br />

means <strong>of</strong> the reaction function �(�, t),<br />

which is evaluated by the BESSY model<br />

(Jäger <strong>and</strong> Liedl, 2000) implemented in<br />

SMART. With given g <strong>and</strong> � the normalized<br />

breakthrough curve <strong>of</strong> a reactive<br />

solute at a control plane is calculated by<br />

(Finkel et al. 1998)<br />

8<br />

�<br />

� �<br />

0<br />

�x, t�<br />

g��,<br />

x����t�d�<br />

C ,<br />

(24).<br />

To overcome the limitations <strong>of</strong> both the<br />

Eulerian <strong>and</strong> the Lagrangian concepts,<br />

mixed Eulerian-Lagrangian methods can be<br />

used, which take advantage <strong>of</strong> the particular<br />

appropriateness <strong>of</strong> both concepts for <strong>modeling</strong><br />

advective <strong>and</strong> dispersive transport (e.g.<br />

Neumann, 1981; Thorenz, 2001; Park et al.,<br />

2006). In application 3 (section 3.3) a<br />

combination <strong>of</strong> the s<strong>of</strong>tware codes GeoSys /<br />

Rock<strong>flow</strong> <strong>and</strong> SMART is used for a combined<br />

application <strong>of</strong> the Eulerian <strong>and</strong> the<br />

Lagrangian concepts. GeoSys / Rock<strong>flow</strong> is<br />

used to model the Eulerian <strong>flow</strong> field in a<br />

heterogeneous two dimensional domain <strong>and</strong><br />

to derive the representative pdf g(�, x). The<br />

SMART model then utilizes the pdf for the<br />

simulation <strong>of</strong> reactive transport in the<br />

model domain.<br />

Object- <strong>and</strong> process-oriented methods<br />

The GeoSys / Rock<strong>flow</strong> code, which is used<br />

for most <strong>of</strong> the <strong>numerical</strong> simulations<br />

described in chapter 3, is written in the C++<br />

language <strong>and</strong> thus allows the implementation<br />

by object oriented programming (OOP)<br />

methods. The OOP concept is especially<br />

helpful for the development <strong>of</strong> complex<br />

s<strong>of</strong>tware in programmer teams, as encapsulation<br />

<strong>and</strong> class-structures render the code<br />

more stable <strong>and</strong> errors are easier to detect.<br />

In GeoSys / Rock<strong>flow</strong> the OOP concept is<br />

met by so called process orientation<br />

(Kolditz <strong>and</strong> Bauer, 2004), which allows<br />

the coupling <strong>of</strong> two-phase <strong>flow</strong>, heat transport,<br />

mass transport, chemical reactions <strong>and</strong><br />

deformation in an efficient way (Wang et<br />

al., 2006). The basic idea <strong>of</strong> process orientation<br />

is that between each physical process<br />

(e.g. single species transport) <strong>and</strong> its<br />

<strong>numerical</strong> approximation by an algebraic<br />

equation system (AES) exists a direct correspondence<br />

(Fig. 4). The AES originates<br />

from the temporal <strong>and</strong> spatial discretization<br />

<strong>of</strong> the PDE on the computational grid. For<br />

its solution the following steps are performed:<br />

� AES assemblage <strong>and</strong> incorporation <strong>of</strong><br />

initial conditions<br />

� determination <strong>of</strong> element matrices<br />

� incorporation <strong>of</strong> boundary conditions<br />

<strong>and</strong> source terms<br />

� solving the AES by appropriate solvers<br />

This procedure can be generalized for any<br />

physical process regardless <strong>of</strong> its specific<br />

type in an object oriented way by introducing<br />

the process object (Kolditz <strong>and</strong> Bauer,<br />

2004) (Fig. 4).<br />

Transport <strong>of</strong> non-reactive<br />

species in water<br />

„PROCESS“<br />

Multifield problem, if many<br />

mobile species are<br />

transported<br />

Solution <strong>of</strong> a PDE<br />

system<br />

PROCESS-OBJECT<br />

System <strong>of</strong> PDE<br />

solved by<br />

Multi-Process-Method<br />

Fig. 4: Process analogy <strong>and</strong> process object,<br />

shown for an instance <strong>of</strong> a transport<br />

process (Kolditz <strong>and</strong> Bauer, 2004).


The process object has access to all<br />

required data structures <strong>and</strong> functions, <strong>and</strong><br />

thus is self configuring, executing <strong>and</strong><br />

destructing.<br />

Contaminant transport problems usually<br />

involve a number <strong>of</strong> different processes as<br />

outlined in sections 2.1 – 2.3. The resulting<br />

multi-field problems can be approached by<br />

a multi-process algorithm, where one instance<br />

<strong>of</strong> the process object is created for<br />

each process considered. Solution <strong>of</strong> any<br />

number <strong>of</strong> <strong>flow</strong> or transport equations is<br />

fully automatic <strong>and</strong> encapsulated, guaranteeing<br />

high efficiency <strong>and</strong> flexibility.<br />

In GeoSys / Rock<strong>flow</strong> <strong>saturated</strong> <strong>and</strong> un<strong>saturated</strong><br />

<strong>flow</strong> as well as conservative transport<br />

are solved using st<strong>and</strong>ard Galerkin FE. A<br />

non-iterative operator splitting technique<br />

for the coupling <strong>of</strong> conservative transport<br />

<strong>and</strong> (bio-)chemical reaction processes is<br />

used (Xie et al., 2006; Bauer et al., 2006b).<br />

First, the <strong>flow</strong> field is solved followed by<br />

conservative transport for all species. In the<br />

third step the calculation <strong>of</strong> kinetic<br />

biochemical reactions is performed. Finally<br />

chemical equilibrium reactions are calculated.<br />

This approach allows an easy h<strong>and</strong>ling<br />

<strong>of</strong> any number <strong>of</strong> species <strong>and</strong> reaction<br />

processes as well as employing optimised<br />

mathematical methods for the solution <strong>of</strong><br />

the corresponding equation systems. The<br />

non-iterative approach, however, is limited<br />

to small time steps in order to avoid<br />

<strong>numerical</strong> instabilities. It is also known not<br />

to converge necessarily to the exact solution<br />

(Carrayrou et al., 2004). These limitations<br />

can be overcome using a computationally<br />

more dem<strong>and</strong>ing iterative operator splitting<br />

approach (e.g. Kinzelbach et al., 1991).<br />

3. Modeling applications<br />

In this chapter three application examples<br />

<strong>of</strong> <strong>numerical</strong> models for <strong>flow</strong> <strong>and</strong> reactive<br />

transport as established in chapter 2 are<br />

presented. Section 3.1 introduces the Virtual<br />

Aquifer (VA) concept, in which <strong>numerical</strong><br />

<strong>modeling</strong> is used for the evaluation <strong>of</strong><br />

investigation strategies for contaminated<br />

sites. In section 3.2 the VA concept is<br />

applied to derive <strong>and</strong> test a novel method<br />

for the estimation <strong>of</strong> biodegradation kinetic<br />

parameters from measured field data. Section<br />

3.3 uses <strong>numerical</strong> <strong>modeling</strong> as a tool<br />

to predict the environmental impact <strong>of</strong> demolition<br />

waste used in road constructions<br />

3.1. Evaluation <strong>of</strong> investigation<br />

strategies for contaminated<br />

aquifers using the Virtual<br />

Aquifer concept<br />

Due to the limited accessibility <strong>of</strong> the<br />

subsurface, measurements <strong>of</strong> piezometric<br />

heads <strong>and</strong> pollutant concentrations at contaminated<br />

sites are sparse <strong>and</strong> may not be<br />

representative <strong>of</strong> the heterogeneous hydrogeologic<br />

conditions. Any site investigation<br />

is thus subject to uncertainty, reflecting the<br />

limited knowledge on aquifer properties<br />

<strong>and</strong> the extent <strong>of</strong> the contamination. Three<br />

main sources <strong>of</strong> uncertainty can be identified<br />

for site investigation, which are illustrated<br />

in Fig. 5. Conceptual model errors<br />

result from an incorrect identification <strong>of</strong> the<br />

governing processes at a site. Heterogeneity<br />

<strong>of</strong> the site causes an incomplete or wrong<br />

description <strong>of</strong> the relevant parameter distributions.<br />

For variables measured at observation<br />

wells like heads or concentrations<br />

measurement errors are inevitable. Due to<br />

this uncertainty, field investigation methods<br />

for plume screening <strong>and</strong> measuring <strong>of</strong><br />

hydraulic conductivity or degradation rates<br />

can hardly be tested or verified in the field.<br />

The VA approach is particularly aimed to<br />

overcome this problem. Its basic idea is the<br />

computer based evaluation <strong>of</strong> the performance<br />

<strong>and</strong> reliability <strong>of</strong> field investigation<br />

methods by application in heterogeneous<br />

synthetic (i.e. virtual) aquifers. In this it<br />

resembles the concept <strong>of</strong> “virtual realities”<br />

(Schäfer et al., 2002) which are used e.g. in<br />

car industry (“virtual crash test”), education<br />

(flight simulators) or medicine (interactive<br />

operation planning).<br />

9


conceptional<br />

model error:<br />

wrong process<br />

description<br />

10<br />

measurement<br />

error<br />

resulting<br />

investigation error<br />

conceptual site model<br />

investigation<br />

virtual aquifer<br />

Fig. 5: Virtual aquifer concept <strong>and</strong> possible sources <strong>of</strong> investigation error.<br />

An application <strong>of</strong> the VA concept requires<br />

the definition <strong>of</strong> a synthetic site model <strong>and</strong><br />

its translation into a <strong>numerical</strong> model for<br />

the simulation <strong>of</strong> the identified relevant processes.<br />

Synthetic site models are generated<br />

based on statistical properties <strong>of</strong> natural<br />

aquifers. A defined contaminant source is<br />

then introduced <strong>and</strong> the evolution <strong>of</strong> aquifer<br />

contamination is simulated by <strong>numerical</strong><br />

<strong>modeling</strong>, thus generating a realistic<br />

contaminant distribution in the synthetic<br />

aquifer. In comparison to the "real world",<br />

the unique advantage <strong>of</strong> the synthetic aquifer<br />

is that the spatial distribution <strong>of</strong> all<br />

physical <strong>and</strong> geochemical properties <strong>and</strong><br />

parameters as well as contaminant concentrations<br />

are exactly known. Once the<br />

synthetic contaminated aquifer is generated,<br />

it can be studied by st<strong>and</strong>ard monitoring<br />

<strong>and</strong> investigation techniques, e.g. by emplacement<br />

<strong>of</strong> observation wells. Although<br />

the parameter distribution <strong>of</strong> the synthetic<br />

aquifer is known a priori, only the data<br />

“measured” at wells (i.e. hydraulic heads or<br />

concentrations) are used <strong>and</strong> interpreted.<br />

This is done because in a real site investigation<br />

also only a limited amount <strong>of</strong> measured<br />

data would be available. Finally, the results<br />

are compared to the “true” parameter distribution<br />

known from the synthetic aquifer,<br />

allowing an evaluation <strong>of</strong> the accuracy <strong>of</strong><br />

the investigation method used. Using the<br />

VA concept, sources <strong>of</strong> uncertainty or error<br />

can be considered individually <strong>and</strong> the<br />

heterogeneity:<br />

incomplete or<br />

wrong description<br />

<strong>of</strong> parameter<br />

distribution<br />

sensitivity <strong>of</strong> investigation results on these<br />

can be studied. Stochastic approaches like<br />

the Monte-Carlo method are applied to<br />

study the propagation <strong>of</strong> parameter variability<br />

<strong>and</strong> uncertainty into the investigation<br />

results. The VA has been first introduced<br />

by Schäfer et al. (2002) <strong>and</strong> was applied by<br />

Schäfer et al. (2004, 2006b), Bauer <strong>and</strong><br />

Kolditz (2006), Bauer et al. (2005 [EP 1];<br />

2006a [EP 2], 2007 [EP 4]) <strong>and</strong> Beyer et al.<br />

(2006 [EP 3], 2007a [EP 5]). An overview<br />

<strong>of</strong> VA applications is given in Bauer et al.<br />

(2006b) <strong>and</strong> Schäfer et al. (2006a).<br />

Fig. 6: Virtual investigation <strong>of</strong> a heterogeneous<br />

contaminant plume by the center<br />

line method (Bauer et al., 2005 [EP 1]).<br />

The VA concept is used here to study errors<br />

<strong>and</strong> uncertainties in degradation rate constants<br />

estimated from data typically collected<br />

by site investigation with the so called


center line method (Fig. 6). This method is<br />

frequently used in field studies when natural<br />

attenuation is considered as a remediation<br />

alternative <strong>and</strong> is based on observation<br />

wells that are placed along the presumed<br />

center line <strong>of</strong> the contaminant plume.<br />

Influence <strong>of</strong> measurement errors<br />

The first aspect studied here is the influence<br />

<strong>of</strong> measurement errors in hydraulic heads<br />

on degradation rate constant estimates<br />

(Bauer et al., 2007 [EP 4]). The VA concept<br />

used in this study is based on a two<br />

dimensional conceptual model <strong>of</strong> the<br />

groundwater body using a homogeneous<br />

distribution <strong>of</strong> hydraulic conductivity K.<br />

The development <strong>of</strong> the contaminant plume<br />

originating from a rectangular source zone<br />

is simulated until steady state conditions are<br />

established. The <strong>numerical</strong> simulations are<br />

performed using the GeoSys / Rock<strong>flow</strong><br />

code, which was introduced in the previous<br />

chapter. The contaminant is subject to a<br />

first order kinetics (eq. (23)) degradation<br />

reaction using a uniform degradation rate<br />

constant �. Additionally, a conservative<br />

tracer is emitted from the source zone. The<br />

contaminant plume thus generated is then<br />

investigated by the center line method.<br />

From the hydraulic heads measured at three<br />

initial observation wells (one being located<br />

directly in the center <strong>of</strong> the source zone)<br />

first the direction <strong>of</strong> groundwater <strong>flow</strong> is<br />

estimated by construction <strong>of</strong> a hydrogeological<br />

triangle. Head measurements are<br />

obtained by reading the model output at the<br />

respective well locations <strong>and</strong> adding a<br />

r<strong>and</strong>om measurement error by<br />

h � � h � ���<br />

(25)<br />

h<br />

where h´ <strong>and</strong> h are the erroneous <strong>and</strong> exact<br />

(i.e. simulated) heads, respectively, � is an<br />

evenly distributed r<strong>and</strong>om number from the<br />

interval [-1, 1] <strong>and</strong> ��h is the maximum<br />

measurement error. Along the estimated<br />

(<strong>and</strong> potentially erroneous) <strong>flow</strong> direction<br />

three new observation wells are installed,<br />

one at every 10 m. These wells were then<br />

used to measure local (erroneous) heads,<br />

contaminant concentrations <strong>and</strong> hydraulic<br />

conductivities along the presumed plume<br />

center line. From the hydraulic head difference,<br />

true porosity <strong>and</strong> well positions<br />

groundwater <strong>flow</strong> velocities are calculated.<br />

Together with the concentration data this<br />

allows the determination <strong>of</strong> � using any <strong>of</strong><br />

the analytical models presented in Tab. 1.<br />

As hydraulic conductivity K is distributed<br />

homogeneously <strong>and</strong> concentrations are<br />

assumed to be measured precisely, the only<br />

source <strong>of</strong> error here is the measured head.<br />

Tab. 1: Analytical models for the estimation <strong>of</strong> the first order degradation rate constant �.<br />

method formula description reference<br />

1<br />

2<br />

3<br />

4<br />

v � �<br />

a C ( x)<br />

� � � �<br />

�<br />

�<br />

�<br />

1 ln<br />

�x<br />

� C0<br />

�<br />

v � � �<br />

a � C ( x)<br />

C0<br />

� � �<br />

�<br />

2 ln<br />

�x<br />

� C � �<br />

� 0 C ( x)<br />

�<br />

with<br />

�C( x)<br />

C �<br />

2<br />

v �<br />

�<br />

a ��<br />

ln<br />

0 �<br />

�<br />

� �<br />

3 �<br />

�<br />

�1<br />

� 2�<br />

L<br />

� 1<br />

4�<br />

�<br />

L ��<br />

�x<br />

� �<br />

�C( x)<br />

( C � ) �<br />

2<br />

v �<br />

�<br />

a ��<br />

ln<br />

0 �<br />

�<br />

� �<br />

4 �<br />

�<br />

�1<br />

� 2�<br />

L<br />

� 1<br />

4�<br />

�<br />

L ��<br />

�x<br />

� �<br />

� �<br />

�<br />

W S<br />

� � erf<br />

�<br />

� �<br />

� 4 � T � x �<br />

analyt. solution <strong>of</strong> 1D advection<br />

equation with first order<br />

degradation<br />

same as method 1; concentrations<br />

normalized by conservative tracer<br />

analyt. solution <strong>of</strong> 1D advectiondispersion<br />

equation with first<br />

order degradation<br />

analyt. solution <strong>of</strong> 2D advectiondispersion<br />

equation with first<br />

order degradation <strong>and</strong> accounting<br />

for the source area width .<br />

Newell et al.<br />

(2002)<br />

Wiedemeyer<br />

et al. (1996)<br />

Buscheck<br />

<strong>and</strong> Alcantar<br />

(1995)<br />

Zhang <strong>and</strong><br />

Heathcote<br />

(2003)<br />

11


The conceptual model used is a rigorous<br />

simplification <strong>of</strong> the processes observed in<br />

natural aquifer systems, as K varies in space<br />

<strong>and</strong> contaminant degradation usually follows<br />

more complicated laws, depends on<br />

microorganism growth <strong>and</strong> may be limited<br />

or inhibited by other substances. The<br />

simplifications assumed here are however<br />

necessary to study the sole effects <strong>of</strong> measurement<br />

error on rate constant estimation<br />

under otherwise ideal conditions for the<br />

application <strong>of</strong> the center line method <strong>and</strong><br />

the analytical models <strong>of</strong> Tab. 1.<br />

In this synthesis only results for method 1<br />

<strong>of</strong> Tab. 1 are presented. The degradation<br />

rate estimated from evaluation <strong>of</strong> measured<br />

heads <strong>and</strong> concentrations is divided by the<br />

true rate constant used in the <strong>numerical</strong><br />

model yielding normalized overestimation<br />

factors. To assess the range <strong>of</strong> uncertainty<br />

resulting from the r<strong>and</strong>om measurement<br />

error a Monte Carlo analysis is conducted<br />

for five increasing values <strong>of</strong> ��h, each with<br />

a sample size <strong>of</strong> 100. Fig. 7 presents the<br />

rate constants thus estimated versus ��h.<br />

Without measurement error the correct rate<br />

constant is obtained. However, already for<br />

very small ��h < 1 cm, a significant<br />

overestimation can be observed. This has<br />

two main reasons: Firstly, an erroneous<br />

head yields an incorrect transport velocity<br />

<strong>and</strong> thus results in rate constant<br />

overestimation for too high velocities <strong>and</strong> in<br />

underestimation for too low velocities.<br />

Secondly, erroneous heads result in an incorrect<br />

derivation <strong>of</strong> the <strong>flow</strong> direction<br />

using the hydrogeologic triangle. Thus the<br />

observation wells installed downgradient <strong>of</strong><br />

the source may be placed <strong>of</strong>f the true center<br />

line position (compare Fig. 6). Hence, concentrations<br />

measured in <strong>of</strong>f center line wells<br />

are too low, indicating an overly high rate<br />

<strong>of</strong> degradation. Overestimation <strong>of</strong> the<br />

degradation rate increases with the maximum<br />

head error <strong>and</strong> reaches factors <strong>of</strong> more<br />

than 20 in the worst cases.<br />

In Bauer et al. (2007 [EP 4]) also the<br />

influence <strong>of</strong> concentration measurement<br />

errors was studied. It was found that also<br />

12<br />

this type <strong>of</strong> error results in overestimation<br />

<strong>of</strong> the rate constant on average. Erroneous<br />

heads, however, were found to have a larger<br />

impact on the rate constant estimates. For<br />

real field applications therefore much care<br />

has to be taken when measuring hydraulic<br />

heads for the derivation <strong>of</strong> the plume center<br />

line position.<br />

Fig. 7: Influence <strong>of</strong> head measurement<br />

error on rate constant estimates (Bauer et<br />

al., 2007 [EP 4]). The reference rate<br />

constant is indicated by the horizontal line,<br />

small symbols represent single estimates,<br />

big symbols are ensemble averages with<br />

st<strong>and</strong>ard deviation as error bars.<br />

Influence <strong>of</strong> aquifer heterogeneity<br />

The second aspect studied using the VA<br />

concept is the influence <strong>of</strong> spatially heterogeneous<br />

hydraulic conductivity distributions<br />

on the accuracy <strong>of</strong> rate constant<br />

estimation methods 1 – 4 <strong>of</strong> Tab. 1. For this<br />

end the conceptual model used so far is<br />

modified by assuming all head, concentration<br />

<strong>and</strong> K measurements to be free <strong>of</strong><br />

measurement error <strong>and</strong> regarding K as a<br />

spatial r<strong>and</strong>om variable. This is done to<br />

study the sole influence <strong>of</strong> heterogeneous<br />

conductivity. Multiple realizations <strong>of</strong> heterogeneous<br />

K fields for four degrees <strong>of</strong><br />

aquifer heterogeneity characterized by the<br />

ln-conductivity variance 2<br />

� Y were generated,<br />

using a Monte Carlo approach to study<br />

the range <strong>of</strong> investigation result uncertainty.<br />

� at least 100 different reali-<br />

For each 2<br />

Y


zations were generated. For all realizations<br />

the spreading <strong>of</strong> a conservative <strong>and</strong> a<br />

reactive contaminant plume subject to first<br />

order degradation was simulated. The resulting<br />

heterogeneous plumes were investigated<br />

as explained above. Fig. 8 presents<br />

normalized estimated rate constants for<br />

2<br />

methods 1 – 4 (Tab. 1) versus � Y (Bauer et<br />

al., 2005 [EP 1]). Clearly, most rate<br />

constants are larger than 1, i.e. the rate<br />

constant is generally overestimated. Single<br />

realizations show overestimation by several<br />

orders <strong>of</strong> magnitude. It is obvious that an<br />

increase in K heterogeneity causes higher<br />

overestimation. Also the spread <strong>of</strong> the 100<br />

realizations <strong>and</strong> the resulting ensemble<br />

st<strong>and</strong>ard deviations increase significantly,<br />

causing higher uncertainty in the rate constant<br />

estimate. The main reasons for the<br />

observed overestimation are identified as<br />

deviation <strong>of</strong> observation wells from the true<br />

plume center line position, an incorrect<br />

approximation <strong>of</strong> the transport velocity <strong>and</strong><br />

no or inadequate consideration <strong>of</strong> concen-<br />

normalized deg. rate constant [-]<br />

normalized deg. rate constant [-]<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

(a) method 1 (b) method 2<br />

(c) method 3 (d) method 4<br />

0 1 2 3 4 5<br />

tration reduction by longitudinal <strong>and</strong><br />

transverse dispersion. Comparing the performance<br />

<strong>of</strong> methods 1 – 4 yields that<br />

method 2 is the most accurate <strong>and</strong> reliable<br />

among the four approaches. The superiority<br />

<strong>of</strong> method 2 follows from the correction <strong>of</strong><br />

contaminant concentrations by normalization<br />

to the concentrations <strong>of</strong> a conservative<br />

tracer, which is spread from the same<br />

source. The tracer correction successfully<br />

accounts for the effects <strong>of</strong> dispersion <strong>and</strong><br />

measuring <strong>of</strong>f the center line. Method 3 explicitly<br />

accounts for longitudinal, method 4<br />

for both, longitudinal <strong>and</strong> transverse dispersion.<br />

However, for each realization investigated,<br />

method 3 yields a higher rate constant<br />

estimate than methods 1 or 2. Longitudinal<br />

dispersion <strong>of</strong> a degrading contaminant<br />

results in a stronger spreading <strong>of</strong> the solute<br />

downstream <strong>and</strong> thus in higher concentrations<br />

along the center line <strong>of</strong> a steady state<br />

plume. To model an observed concentration<br />

reduction with a one-dimensional model<br />

like method 3 which accounts for �L only<br />

0 1 2 3 4 5<br />

�2 �2 Y Y<br />

Fig. 8: Estimated degradation rate constants versus aquifer heterogeneity for methods 1 (a),<br />

2 (b), 3 (c) <strong>and</strong> 4 (d) <strong>of</strong> Tab. 1 (Bauer et al., 2005 [EP 1]). Small symbols represent single<br />

realization results, large symbols ensemble averages <strong>of</strong> 100 realizations.<br />

13


equires therefore a higher degradation rate<br />

constant. Method 4 yields significantly<br />

lower rate constant estimates than<br />

method 1, when aquifer heterogeneity is<br />

low, but almost the same results as method<br />

3 for high heterogeneity. The rate constant<br />

underestimation for low heterogeneity is<br />

due to an “over correction” for transverse<br />

dispersion by the error function term � in<br />

the rate equation. Hence, both methods fail<br />

to yield closer rate constant estimates than<br />

the simpler approach <strong>of</strong> method 1, which<br />

completely neglects the dispersion process.<br />

Influence <strong>of</strong> dispersivity parameterization<br />

Since the results <strong>of</strong> methods 3 <strong>and</strong> 4 depend<br />

on longitudinal <strong>and</strong> transverse dispersivities,<br />

an adequate parameterization is crucial<br />

for their success. In this study, parameterization<br />

<strong>of</strong> �L as well as �T is based on the<br />

scale <strong>of</strong> the contamination problem, as<br />

common with many field applications.<br />

Clearly from the results in Fig. 8 derivation<br />

<strong>of</strong> �L <strong>and</strong> �T from the assumed length <strong>of</strong> the<br />

contaminant plume is not appropriate. As<br />

known from stochastic hydrogeology (e.g.<br />

Dagan, 1989) �L as well as �T strongly<br />

depend on travel time <strong>and</strong> distance as well<br />

as on the correlation structure <strong>of</strong> hydraulic<br />

conductivity <strong>and</strong> <strong>flow</strong> velocity. Therefore<br />

the influence <strong>of</strong> dispersivity parameterization<br />

on the performance <strong>of</strong> methods 3 <strong>and</strong> 4<br />

was analyzed (Bauer et al., 2006a [EP 2]).<br />

From the <strong>of</strong>ten very limited amount <strong>of</strong> data<br />

on the degree <strong>of</strong> aquifer heterogeneity <strong>and</strong><br />

the spatial correlation structure <strong>of</strong> hydraulic<br />

conductivity available for real field sites,<br />

the inference <strong>of</strong> dispersivities by methods<br />

<strong>of</strong> stochastic theory is rarely possible or at<br />

least afflicted by high uncertainty.<br />

Therefore a sensitivity analysis is performed<br />

to cover a wide range <strong>of</strong> values for<br />

�L <strong>and</strong> �T. Varying both parameters the<br />

heterogeneous plume realizations were reevaluated<br />

using the different parameterizations<br />

<strong>of</strong> method 4. Fig. 9 shows results <strong>of</strong><br />

method 4 in terms <strong>of</strong> ensemble medians <strong>of</strong><br />

the 100 estimated rate constants for each <strong>of</strong><br />

the four different degrees <strong>of</strong> heterogeneity<br />

14<br />

<strong>and</strong> all combinations <strong>of</strong> �L <strong>and</strong> �T considered.<br />

As method 4 converges with<br />

method 3 for �T = 0, results for �T = 0 are<br />

also representative for method 3.<br />

For all degrees <strong>of</strong> heterogeneity, decreasing<br />

median degradation rates are found with<br />

increasing �T. For low <strong>and</strong> medium heterogeneity<br />

(Fig. 9 (a) <strong>and</strong> (b)) combinations <strong>of</strong><br />

�L <strong>and</strong> �T can be found, which allow for an<br />

optimal estimation <strong>of</strong> the degradation rate<br />

constant, i.e. the results are <strong>of</strong> comparable<br />

accuracy as those <strong>of</strong> methods 1 or 2. For the<br />

highest degree <strong>of</strong> heterogeneity (Fig. 9 (d)),<br />

no such parameter combination is found<br />

within the range <strong>of</strong> dispersivities considered<br />

here. However, with an inadequate parameterization<br />

also for low heterogeneities<br />

severe over- as well as underestimation is<br />

possible. Moreover, with the exception <strong>of</strong><br />

the low heterogeneity case, only unphysical<br />

combinations <strong>of</strong> low �L <strong>and</strong> high �T yield<br />

close estimates <strong>of</strong> the rate constant. These<br />

parameters do not represent actual dispersivities,<br />

but must be considered as mere<br />

lumped fitting parameters, as they are used<br />

to correct for <strong>of</strong>f center line measurements<br />

<strong>and</strong> not the dispersion process only. As the<br />

magnitude <strong>of</strong> bias caused by <strong>of</strong>f center line<br />

measurements is not known <strong>and</strong> the K<br />

correlation structure <strong>and</strong> degree <strong>of</strong> heterogeneity<br />

are usually not properly characterized<br />

at many field sites, choosing<br />

dispersivities for an application <strong>of</strong> method 4<br />

would be highly uncertain <strong>and</strong> arbitrary.<br />

From the results presented here, it can be<br />

concluded that for contaminated sites,<br />

where the assumption <strong>of</strong> first order degradation<br />

kinetics can be justified, method 2<br />

should be preferably applied, if the degradation<br />

rate constant is to be derived with<br />

the center line approach. In situations,<br />

where a correction <strong>of</strong> concentrations by a<br />

conservative tracer is not possible,<br />

alternatively method 1 should be used, as<br />

its application implies the least amount <strong>of</strong><br />

parameterization uncertainty.


Fig. 9: Influence <strong>of</strong> dispersivity parameterization on estimated degradation rate constants for<br />

2<br />

method 4 (<strong>and</strong> method 3 with �T = 0) <strong>of</strong> Tab. 1 <strong>and</strong> four degrees <strong>of</strong> aquifer heterogeneity � Y<br />

(Bauer et al., 2006a [EP 2]). Symbols represent ensemble medians <strong>of</strong> 100 realizations.<br />

Extensive monitoring networks<br />

For the previous studies, linearly arranged<br />

center line observation wells have been<br />

used for the VA investigation (compare Fig.<br />

6). Contaminant plumes in natural heterogeneous<br />

aquifers may however show a<br />

substantial amount <strong>of</strong> me<strong>and</strong>ering (Wilson<br />

et al., 2004) resulting in non-linear center<br />

line orientations. As documented, an<br />

inappropriate assumption <strong>of</strong> a linear center<br />

line may result in misplacement <strong>of</strong> the<br />

observation wells <strong>and</strong> in significant overestimation,<br />

if degradation rate constants are to<br />

be derived. Moreover, only measurements<br />

in observation wells located on the plume<br />

axis are evaluated <strong>and</strong> additional information<br />

possibly at h<strong>and</strong> (i.e. well data downgradient<br />

from the source but not on the<br />

center line) is not explicitly accounted for<br />

in the rate estimation. Therefore the performance<br />

<strong>of</strong> methods 1, 3 <strong>and</strong> 4 <strong>of</strong> Tab. 1 was<br />

Fig. 10: Site investigation for degradation<br />

rate constant evaluation by (A) non-linearly<br />

positioned center line wells <strong>and</strong> (B) using<br />

all observation wells <strong>of</strong> the monitoring<br />

network (Beyer et al., 2007a [EP 5]). Small<br />

dark squares show observation well<br />

locations, larger squares show presumed<br />

center line well locations, contour lines are<br />

concentrations calculated analytically with<br />

the approach <strong>of</strong> Stenback et al. (2004).<br />

15


studied for heterogeneous sites with extensive<br />

monitoring networks allowing the estimation<br />

<strong>of</strong> the degradation rate constant<br />

based on freely positioned observation<br />

wells from which a center line is constructed<br />

(strategy A, Fig. 10 (a)). Results are<br />

compared to a two-dimensional inverse<br />

<strong>modeling</strong> approach <strong>of</strong> Stenback et al.<br />

(2004), which accounts for information<br />

from all observation wells <strong>of</strong> a monitoring<br />

network (strategy B, Fig. 10 (b)). Both rate<br />

constant estimation strategies are applied to<br />

a set <strong>of</strong> synthetic contaminated sites with<br />

independently designed extensive monitoring<br />

networks (Beyer et al., 2007a [EP 5]).<br />

a)<br />

b)<br />

16<br />

norm. degradation rate constant [-]<br />

norm. degradation rate constant [-]<br />

100<br />

10<br />

1<br />

0.1<br />

100<br />

10<br />

1<br />

0.1<br />

SW = 4 m<br />

SW = 16 m<br />

meth. 1<br />

(A)<br />

meth. 3<br />

(A)<br />

meth. 4<br />

(A)<br />

investigation strategy<br />

Stenback<br />

(B)<br />

Fig. 11: Rate constant estimates obtained<br />

for investigation strategy A <strong>and</strong> methods 1,<br />

3 <strong>and</strong> 4 as well as strategy (B) for a source<br />

width WS <strong>of</strong> 4 (a) <strong>and</strong> 16 m (b) (Beyer et al.,<br />

2007a [EP 5]). Small symbols represent<br />

single realization results, large symbols<br />

ensemble averages.<br />

Two different types <strong>of</strong> plumes are considered<br />

in this comparison: narrow contaminant<br />

plumes with a small source area width<br />

WS = 4 m (Fig. 11 (a)) <strong>and</strong> wider plumes<br />

with WS = 16 m (Fig. 11 (b)). For small<br />

source widths overestimation by strategy B<br />

is comparable to or at best slightly less than<br />

for approach A. For large source widths <strong>and</strong><br />

wider plumes, however, strategy B yields<br />

closer estimates <strong>of</strong> the degradation rate<br />

constant than strategy A on average. The<br />

results <strong>of</strong> this study suggest that incorporation<br />

<strong>of</strong> <strong>of</strong>f center line information for the<br />

estimation <strong>of</strong> the degradation rate constant<br />

can improve results <strong>of</strong> the plume investigation<br />

significantly.<br />

3.2. Development <strong>and</strong> testing <strong>of</strong> a<br />

new approach to estimating<br />

biodegradation parameters<br />

from field data<br />

In the previous section, the VA method was<br />

applied to evaluate the performance <strong>of</strong><br />

different analytical models for the derivation<br />

<strong>of</strong> first order degradation rate constants<br />

from center line investigation data in heterogeneous<br />

aquifers. From the literature,<br />

however, it is well known that the use <strong>of</strong><br />

first order kinetics may be problematic in<br />

some situations, as it is an inaccurate representation<br />

<strong>of</strong> the processes occurring in<br />

contaminated aquifers. Usage <strong>of</strong> a first<br />

order model outside its range <strong>of</strong> validity<br />

may result either in significant under- or<br />

overestimation <strong>of</strong> the attenuation potential<br />

at a site (Bekins et al., 1998). In an<br />

extension to this study, therefore a new<br />

approach for the estimation <strong>of</strong> degradation<br />

parameters kmax <strong>and</strong> MC for Michaelis-<br />

Menten (MM) kinetics (eq. 22) from the<br />

same plume investigation strategy was<br />

developed <strong>and</strong> tested in Beyer et al. (2006<br />

[EP 3]) using the VA method.<br />

In its integral form eq. (22) can be<br />

rearranged to


v<br />

�<br />

�C�Cx) �<br />

� C0<br />

�<br />

ln��<br />

C�x���<br />

M � � 1<br />

�<br />

�<br />

( k C � C(<br />

x)<br />

k<br />

x C<br />

(25).<br />

a 0<br />

max 0<br />

max<br />

With the same type <strong>of</strong> center line<br />

investigation data used for the estimation <strong>of</strong><br />

the first order degradation rate constant (i.e.<br />

local concentrations, heads, hydraulic conductivities),<br />

eq. (25) can be utilized to<br />

estimate the MM parameters kmax <strong>and</strong> MC<br />

by linear regression.<br />

The Monte Carlo scenario definition <strong>and</strong><br />

site investigation procedure for this study<br />

are similar to those explained in section 3.1.<br />

Numerical simulations <strong>of</strong> plume development<br />

in homogeneous <strong>and</strong> heterogeneous<br />

aquifers were performed with the GeoSys /<br />

Rock<strong>flow</strong> code. Here, however, the contaminant<br />

plumes investigated were generated<br />

using MM instead <strong>of</strong> first order degradation<br />

kinetics. The parameters k max <strong>and</strong> M C<br />

estimated with eq. (25) for the different<br />

plume realizations were normalized to the<br />

true values used in the <strong>numerical</strong><br />

simulations <strong>and</strong> are shown in a 3D-scatterplot<br />

versus aquifer heterogeneity (Fig. 12).<br />

Fig. 12: Normalized Michaelis-Menten parameters<br />

(given as overestimation factors)<br />

versus aquifer heterogeneity.<br />

In general an overestimation <strong>of</strong> both k max<br />

<strong>and</strong> M C is observed, which increases with<br />

heterogeneity. An overestimation <strong>of</strong> k max increases<br />

the velocity <strong>of</strong> contaminant degradation<br />

as long as concentrations are much<br />

higher than MC. The simultaneous overestimation<br />

<strong>of</strong> MC counterbalances this effect<br />

because the concentration threshold is<br />

raised at which the kinetic begins to show a<br />

dependence on concentration <strong>and</strong> transits<br />

from zero to first order <strong>and</strong> hence decreases<br />

the rate <strong>of</strong> degradation.<br />

To obtain an indicator for the significance<br />

<strong>of</strong> the estimated degradation potential, the<br />

MM parameters determined were used in an<br />

analytical transport model to estimate the<br />

contaminant plume lengths. These then<br />

were compared to the respective true plume<br />

lengths from the <strong>numerical</strong> simulations<br />

(Fig. 13 (a)). As a consequence <strong>of</strong> overestimating<br />

the degradation parameters, calculated<br />

plume lengths for high heterogeneities<br />

are estimated to about 75 % <strong>of</strong> the true<br />

length on average <strong>and</strong> thus are not conservative.<br />

For low heterogeneities, however,<br />

the suggested regression approach on average<br />

yields good estimates <strong>of</strong> the plume<br />

length <strong>and</strong> the degradation potential.<br />

In addition to the effect <strong>of</strong> aquifer heterogeneity<br />

on estimated MM parameters <strong>and</strong> the<br />

resultant plume length estimates, also the<br />

effect <strong>of</strong> a wrong process identification<br />

(compare Fig. 5) is studied in Beyer et al.<br />

(2006 [EP 3]). Although it is well known<br />

that contaminant degradation in natural<br />

aquifers is governed by complex processes<br />

<strong>and</strong> kinetic laws, simple first order models<br />

are routinely used at many field sites. This<br />

study therefore highlights some <strong>of</strong> the<br />

problems that result from an insufficient<br />

wrong process identification. For this end<br />

investigation <strong>of</strong> the plumes following MM<br />

degradation kinetics is repeated, assuming<br />

the appropriateness <strong>of</strong> a first order rate law<br />

to approximate the contaminant degradation<br />

behaviour. Hence the methods <strong>of</strong> Tab. 1<br />

were used to derive first order rate constants<br />

for the multiple plume realizations.<br />

As for the estimated MM parameters the<br />

estimated first order rate constants were<br />

evaluated by analytical transport models to<br />

yield estimates <strong>of</strong> the contaminant plume<br />

length. Results for method 1 (Tab. 1) are<br />

presented in Fig. 13 (b).<br />

17


Fig. 13: Plume length overestimation factors<br />

versus aquifer heterogeneity (Beyer et<br />

al., 2006 [EP 3]). Plume lengths were estimated<br />

for plumes following Michaelis-Menten<br />

degradation kinetics estimated using the<br />

Michaelis-Menten model (a) <strong>and</strong> assuming<br />

validity <strong>of</strong> a first order rate law (b).<br />

In comparison with Fig. 13 (a) an additional<br />

error is introduced which stems from the<br />

first order approximation. Uncertainty as<br />

well as bias increase significantly, as can be<br />

seen by the wider spread <strong>of</strong> single realization<br />

results around the ensemble medians.<br />

Estimated plume lengths here are found to<br />

be less than 40 % <strong>of</strong> the true length on<br />

average even for mildly heterogeneous<br />

aquifers. Plume lengths calculated using the<br />

MM parameters in general are significantly<br />

closer to the correct length compared to<br />

those obtained by a first order approximation.<br />

This approach is therefore recom-<br />

18<br />

mended, if field data collected along the<br />

center line <strong>of</strong> a plume give evidence <strong>of</strong> MM<br />

type degradation kinetics.<br />

3.3. Prognosis <strong>of</strong> long term contaminant<br />

leaching from recycling<br />

materials in road<br />

constructions<br />

The third application <strong>of</strong> the <strong>numerical</strong><br />

models <strong>and</strong> methods presented in section 2<br />

is focussed on the prognosis <strong>of</strong> contaminant<br />

leaching <strong>and</strong> transport by seepage water<br />

from pollutant loaded recycling materials,<br />

which are used in earthworks or road<br />

constructions. According to the German<br />

federal soil protection decree (BBodSchV,<br />

1999) such a prognosis is required for<br />

contaminated sites as well as for constructions<br />

or depositions <strong>of</strong> contaminated materials<br />

in order to assess the extent <strong>and</strong> environmental<br />

impact <strong>of</strong> potential contaminant<br />

leaching through the vadose zone to the<br />

groundwater. In such a prognosis, the<br />

relevant attenuation processes need to be<br />

considered <strong>and</strong> quantified, as significant<br />

contaminant attenuation could result in less<br />

restrictive utilization criteria without<br />

compromising the protection <strong>of</strong> groundwater<br />

resources. For this end, the application<br />

<strong>of</strong> process based <strong>numerical</strong> transport<br />

models is favorable, as complex geometries<br />

<strong>of</strong> the model scenarios <strong>and</strong> possible process<br />

interactions limit the applicability <strong>of</strong> analytical<br />

models or expertise founded “verbalargumentative”<br />

assessments.<br />

In this study, process based type scenario<br />

<strong>modeling</strong> is used as a tool to assess<br />

contaminant leaching from recycled demolition<br />

waste (DW) material. The type<br />

scenarios are based on three different case<br />

studies for the utilization <strong>of</strong> the DW, i.e.<br />

recycling as base <strong>and</strong> subbase layers <strong>of</strong> a<br />

parking lot, a noise protection dam <strong>and</strong> a<br />

road dam (Fig. 14) (Beyer et al., 2007b<br />

[EP 6]). Instead <strong>of</strong> regarding the full spectrum<br />

<strong>of</strong> contaminants typically embodied in<br />

DW three model substances are considered<br />

in the type scenarios: a conservative tracer


1.3m<br />

0.3m<br />

as a representative for highly soluble salts,<br />

naphthalene for moderately sorbing <strong>and</strong><br />

phenanthrene for strongly sorbing organic<br />

compounds. Contaminant leaching from the<br />

DW to the groundwater surface is studied<br />

with six different characteristic subsoil<br />

units <strong>of</strong> Germany (BGR, 2006) to be able to<br />

compare the influence <strong>of</strong> hydraulic <strong>and</strong><br />

basic physico-chemical soil properties on<br />

contaminant attenuation. Fig. 14 presents<br />

the conceptual model for the road dam.<br />

Here coarse grained DW is used as unbound<br />

base / subbase layers below the<br />

asphalt surface <strong>of</strong> the road. According to<br />

German road construction regulations, the<br />

base / subbase layers are covered by low<br />

<strong>and</strong> high permeable soil layers along the<br />

embankment (Fig. 14).<br />

RCB<br />

Körnung 0/32<br />

b horizon<br />

c horizon<br />

symmetry axis<br />

10m 1.5m<br />

asphalt layer<br />

impermeable low permeable soil<br />

3%<br />

4%<br />

12%<br />

1:1. 5<br />

2.3m<br />

10 cm high<br />

permeable soil<br />

2m<br />

groundwater surface<br />

(point <strong>of</strong> compliance)<br />

Fig. 14: Road construction with demolition<br />

waste recycled in base / subbase layers<br />

(Beyer et al., 2007b [EP 6]).<br />

The simulation strategy for this study<br />

combines the Eulerian <strong>and</strong> Lagrangian<br />

frameworks <strong>of</strong> transport <strong>modeling</strong>. Un<strong>saturated</strong><br />

<strong>flow</strong>, i.e. the hydraulics <strong>of</strong> the constructions,<br />

is modeled with GeoSys /<br />

Rock<strong>flow</strong> using st<strong>and</strong>ard FEM. Fig. 15<br />

shows that the two-dimensional scenarios<br />

exhibit complex <strong>flow</strong> patterns under un<strong>saturated</strong><br />

conditions. The velocity vectors at<br />

the element nodes <strong>of</strong> the FEM mesh for the<br />

road dam presented in Fig. 15 indicate<br />

unhindered infiltration <strong>of</strong> water from the<br />

low permeable soil on top <strong>of</strong> the embankment<br />

into the DW material. Along the<br />

sloped material boundary with high permeable<br />

soil on top <strong>of</strong> the coarse DW, however,<br />

strong capillary barrier effects are observed.<br />

These cause a concentration <strong>of</strong> the water<br />

flux on top <strong>of</strong> the DW <strong>and</strong> generation <strong>of</strong><br />

lateral run<strong>of</strong>f, almost completely bypassing<br />

the DW.<br />

Fig. 15: Element nodes <strong>and</strong> velocity vectors<br />

<strong>of</strong> un<strong>saturated</strong> water <strong>flow</strong> in a section <strong>of</strong> the<br />

model domain (Beyer et al., 2007b [EP 6]).<br />

Reactive transport <strong>of</strong> the model compounds<br />

is simulated using the stream tube concept<br />

<strong>of</strong> the SMART code, to take full advantage<br />

<strong>of</strong> the reactive process models implemented<br />

in SMART (degradation, intraparticle<br />

diffusion kinetics, sorption, etc.). Coupling<br />

between GeoSys / Rock<strong>flow</strong> <strong>and</strong> SMART is<br />

achieved through the travel time pdf. These<br />

are generated by simulation <strong>of</strong> conservative<br />

tracer breakthrough curves by GeoSys /<br />

Rock<strong>flow</strong> using the FEM. The pdf are used<br />

as input for the reactive transport simulations<br />

with SMART. The model output <strong>of</strong><br />

SMART for the three model substances at<br />

the groundwater surface represents concentrations<br />

integrated along the cross-section<br />

<strong>of</strong> the contaminant transport path only.<br />

These breakthrough curves for the tracer in<br />

the road dam are displayed in Fig. 16 as<br />

grey curves. The three black curves<br />

represent the same tracer breakthrough<br />

concentrations but are integrated along the<br />

groundwater surface <strong>of</strong> the overall model<br />

domain. Hence, they account for dilution by<br />

uncontaminated seepage water which<br />

bypasses the actual transport path due to the<br />

capillary barrier.<br />

Breakthrough curves in relative concentrations<br />

C/C0 [-] are given here for three <strong>of</strong><br />

the six soils regarded in this study, i.e. for a<br />

cambisol, a podzol <strong>and</strong> a chernozem. Comparing<br />

the concentration breakthrough for<br />

the transport path (grey curves), it is found<br />

that the earliest breakthrough time for the<br />

19


concentration maximum is for the cambisol,<br />

followed by the podzol <strong>and</strong> the chernozem.<br />

In general, however, breakthrough times are<br />

very similar for the three soils. Also the<br />

maximum concentrations observed are <strong>of</strong><br />

comparable magnitude. As the tracer is conservative,<br />

concentration reductions <strong>of</strong> about<br />

70 % can be attributed to the dispersion<br />

process. Integration <strong>of</strong> breakthrough concentrations<br />

along the overall lower model<br />

boundary (black curves) yields a further<br />

reduction <strong>of</strong> concentrations, as about 30 %<br />

<strong>of</strong> the infiltration bypasses the contaminant<br />

transport path.<br />

Fig. 16: Tracer breakthrough curves at the<br />

groundwater surface below the road dam<br />

for three different subsoil types (Beyer et<br />

al., 2007b [EP 6]).<br />

For naphthalene <strong>and</strong> phenanthrene, contaminant<br />

transport is regarded with <strong>and</strong><br />

without degradation. For the cases with<br />

degradation, a simple first order kinetics is<br />

used. Sorption <strong>of</strong> both compounds is<br />

modeled by a linear isotherm, where the<br />

equilibrium sorption coefficient is derived<br />

from the soil organic carbon content <strong>and</strong> the<br />

distribution coefficient between organic<br />

carbon <strong>and</strong> the aqueous phase (Beyer et al.,<br />

2007b [EP 6]). Sorption kinetics are quantified<br />

using the intraparticle diffusion model<br />

(eq. 18). Concentration breakthroughs are<br />

presented in Fig. 17. For the moderately<br />

sorbing naphthalene without decay the influence<br />

<strong>of</strong> soil organic carbon Corg is clearly<br />

20<br />

visible. The cambisol, which is almost free<br />

<strong>of</strong> Corg (0.01 %), shows the earliest breakthrough<br />

<strong>of</strong> the three soils. For the podzol<br />

with a little higher Corg (0.21 %), the maximum<br />

concentration breakthrough is slightly<br />

retarded. The latest breakthrough time is<br />

observed for the chernozem, which is the<br />

soil with the highest Corg (0.88 %) regarded<br />

here. For the strongly sorbing phenanthrene<br />

this behaviour is even more characteristic.<br />

For the chernozem, the maximum concentration<br />

breakthrough is not yet observed<br />

within 275 a <strong>of</strong> simulated contaminant leaching.<br />

In contrast to the tracer, for which<br />

source concentrations are depleted within a<br />

few years, retardation within the DW causes<br />

naphthalene <strong>and</strong> phenanthrene leachate<br />

concentrations to stay on an almost<br />

unreduced level throughout the simulation<br />

period <strong>of</strong> 275 a (Beyer et al., 2007b [EP 6]).<br />

As high contaminant concentrations are<br />

constantly delivered from the source<br />

material, dispersive concentration reductions<br />

remain ineffective. Hence for the<br />

transport path maximum concentration<br />

breakthrough between 70 <strong>and</strong> 90 % <strong>of</strong> C0 is<br />

observed. As for the tracer, these are<br />

reduced by about 30 % when integrated<br />

along the overall lower model boundary.<br />

In comparison to the other type scenarios<br />

(parking lot, noise protection dam) studied<br />

in Beyer et al. (2007b [EP 6]), the contaminant<br />

residence times in the road dam are<br />

rather short, as high amounts <strong>of</strong> run<strong>of</strong>f<br />

water from the road asphalt which infiltrate<br />

along the embankment result in comparably<br />

high <strong>flow</strong> velocities. The short residence<br />

times reduce the effectiveness <strong>of</strong> degradation.<br />

Hence, naphthalene <strong>and</strong> phenanthrene<br />

concentrations are reduced by factors<br />

between 2.5 <strong>and</strong> 5, respectively, while for<br />

the other two type scenarios concentration<br />

reductions by factors between 10 <strong>and</strong> 150<br />

were observed. From this type scenario<br />

<strong>modeling</strong> study the relevant transport <strong>and</strong><br />

attenuation processes for contaminant<br />

leachate from DW used in road constructions<br />

could be identified <strong>and</strong> quantified for<br />

the assumed model structure.


C/C 0 [-]<br />

C/C 0 [-]<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

1<br />

0.1<br />

0.01<br />

0.001<br />

cambisol (dil.)<br />

podzol (dil.)<br />

chernozem (dil.)<br />

cambisol<br />

podzol<br />

chernozem<br />

0 50 100 150 200 250<br />

time [a]<br />

a) naphthalene b) phenanthrene<br />

c) naphthalene<br />

with degradation<br />

d) phenanthrene<br />

with degradation<br />

0 50 100 150 200 250<br />

time [a]<br />

Fig. 17: Concentration breakthrough curves at the groundwater surface below the road dam<br />

for three different subsoil types showing naphthalene without (a) <strong>and</strong> with degradation (c),<br />

<strong>and</strong> phenanthrene without (b) <strong>and</strong> with degradation (d) (Beyer et al., 2007b [EP 6]).<br />

The results allow important conclusions on<br />

mechanisms (e.g. capillary barrier effects)<br />

<strong>and</strong> complementary design criteria for road<br />

constructions, which can be used to reduce<br />

contaminant leaching to groundwater.<br />

4. Conclusions <strong>and</strong> outlook<br />

In this thesis, the utilization <strong>of</strong> process<br />

based <strong>numerical</strong> models <strong>of</strong> <strong>saturated</strong> / un<strong>saturated</strong><br />

<strong>flow</strong> <strong>and</strong> reactive contaminant transport<br />

is demonstrated for two different fields<br />

<strong>of</strong> application.<br />

The Virtual Aquifer (VA) concept is<br />

introduced, which uses <strong>numerical</strong> <strong>modeling</strong><br />

as a tool for the computer based evaluation<br />

<strong>of</strong> investigation <strong>and</strong> remediation strategies<br />

for contaminated soils <strong>and</strong> aquifers. In the<br />

application examples presented, the VA<br />

concept proves its usefulness for assessing<br />

the uncertainty involved in the investigation<br />

<strong>of</strong> heterogeneous sites <strong>and</strong> the parameterization<br />

<strong>of</strong> degradation process models.<br />

Moreover, the VA concept is successfully<br />

applied to test a newly developed approach<br />

for the inference <strong>of</strong> biodegradation parameters<br />

from data typically collected during site<br />

investigation. Both applications exemplify<br />

the importance but also the pitfalls <strong>of</strong> a<br />

careful <strong>and</strong> accurate collection <strong>of</strong> investigation<br />

data, if these are to be used for<br />

prognosis <strong>of</strong> the contaminant behaviour at a<br />

site. The main advantage <strong>of</strong> the VA is, that<br />

individual factors, such as hydraulic heterogeneity<br />

or conceptual model errors can be<br />

studied in detail at low costs <strong>and</strong> without<br />

high effort, either individually or in combination<br />

<strong>and</strong> under otherwise ideal <strong>and</strong><br />

controlled conditions. As the sum <strong>of</strong> these<br />

21


possibilities can not be provided neither by<br />

large scale field experiments nor in the<br />

laboratory, the VA concept can be<br />

considered a valuable contribution complementing<br />

state <strong>of</strong> the art experimental<br />

methods. Future applications <strong>of</strong> the VA<br />

concept will incorporate more realistic degradation<br />

kinetics <strong>and</strong> structures <strong>of</strong> aquifer<br />

heterogeneity, in order to study the<br />

influence <strong>of</strong> <strong>flow</strong> <strong>and</strong> transport channelling<br />

on the effectiveness <strong>of</strong> natural contaminant<br />

attenuation processes.<br />

In the second field <strong>of</strong> interest regarded here,<br />

<strong>numerical</strong> models are applied for an<br />

assessment <strong>of</strong> environmental impact <strong>of</strong><br />

recycling materials used in road constructions.<br />

In this study, the reactive streamtube<br />

model SMART is used for the first time in<br />

combination with the finite element model<br />

GeoSys / Rock<strong>flow</strong> to simulate the extent<br />

<strong>of</strong> contaminant leaching from contaminated<br />

demolition waste within road structures to<br />

the groundwater surface. The coupling<br />

between GeoSys / Rock<strong>flow</strong> <strong>and</strong> SMART<br />

combines the Eulerian <strong>and</strong> Langrangian<br />

frameworks <strong>of</strong> <strong>flow</strong> <strong>and</strong> transport. In so<br />

doing, the complex hydraulic behaviour <strong>of</strong><br />

the two-dimensional model geometries are<br />

successfully modeled by travel time<br />

probability density functions, which allows<br />

to take full advantage <strong>of</strong> the process models<br />

implemented in the SMART code. For the<br />

assessment <strong>of</strong> contaminant leaching consequences<br />

on groundwater recharge quality,<br />

type scenario <strong>modeling</strong> is used. The<br />

relevant transport <strong>and</strong> attenuation processes<br />

in road constructions are identified. The<br />

study allows important conclusions on how<br />

these mechanisms could be used or<br />

enhanced to further reduce contaminant<br />

leaching to groundwater. As an example,<br />

hydraulic processes like capillary barrier<br />

formation from layered composition <strong>of</strong><br />

granular base / subbase materials in a road<br />

dam could be exploited to reduce water<br />

fluxes through contaminant loaded recycling<br />

materials. This study demonstrates<br />

that process based <strong>numerical</strong> models can<br />

provide valuable tools for optimizing<br />

22<br />

design <strong>and</strong> construction with regard to both<br />

functionality <strong>and</strong> environmental impacts.<br />

In as much as the application <strong>of</strong> <strong>numerical</strong><br />

models provides preeminent <strong>and</strong> cost<br />

effective means <strong>of</strong> assessment <strong>and</strong><br />

prognosis, their value <strong>and</strong> credibility relies<br />

on a thorough collection <strong>and</strong> preparation <strong>of</strong><br />

the required parameters <strong>and</strong> input data.<br />

Only if utmost care is taken in the stage <strong>of</strong><br />

setting up the model, reliable results can be<br />

expected. As the data <strong>and</strong> parameter acquisition<br />

is <strong>of</strong> fundamental importance to a<br />

successful <strong>modeling</strong> <strong>of</strong> hydrogeosystems -<br />

be it by direct measurements in laboratory<br />

<strong>and</strong> field experiments, research in literature<br />

<strong>and</strong> databases or utilization <strong>of</strong> expert knowledge<br />

- both processes are closely related<br />

<strong>and</strong> therefore should, if possible, go h<strong>and</strong> in<br />

h<strong>and</strong>.


References<br />

Altfelder, S., Streck, T. (2006): Capability <strong>and</strong> limitations <strong>of</strong> first-order <strong>and</strong> diffusion approaches to<br />

describe long-term sorption <strong>of</strong> chlortoluron in soil. J. Contam. Hydrol., 86, 279-298,<br />

10.1016/j.jconhyd.2006.04.002.<br />

Anderson, M., Woesner, W. (1992): <strong>Applied</strong> groundwater <strong>modeling</strong>. Simulation <strong>of</strong> <strong>flow</strong> <strong>and</strong><br />

advective transport. Academic Press, New York, 381 p.<br />

Baker, A.J. (1983): Finite Element Computational Fluid Mechanics McGraw-Hill Publ., /<br />

Hemisphere, New York, 1983, 544 p.<br />

Ball, W.P., Roberts, P.V. (1991): Long-term sorption <strong>of</strong> halogenated organic chemicals by aquifer<br />

material: 2. Intraparticle diffusion. Environ. Sci. Technol. 25, 1237–1249,<br />

10.1021/es00019a003.<br />

Bauer, S., Kolditz, O. (2006): Assessing contaminant mass <strong>flow</strong> rates by the integral groundwater<br />

investigation method by using the virtual aquifer approach. In: Bierkens, M.F.P., Gehrels, J.C.,<br />

Kovar, K. (eds.), Calibration <strong>and</strong> Reliability in Groundwater Modelling: From Uncertainty to<br />

Decision Making. Proceedings <strong>of</strong> ModelCARE’2005, The Hague, The Netherl<strong>and</strong>s, June 2005.<br />

IAHS Publ. 304, 123–129.<br />

Bauer, S., Beyer, C., Kolditz, O. (2005): Assessing measurements <strong>of</strong> first order degradation rates by<br />

using the virtual aquifer approach. In: Thomson, N.R. (ed.), GQ2004, Bringing Groundwater<br />

Quality Research to the Watershed Scale. Proceedings <strong>of</strong> a Symposium held at Waterloo,<br />

Canada, July 2004. IAHS Publ. 297, 274–281.<br />

Bauer, S., Beyer, C., Kolditz, O. (2006a): Assessing measurement uncertainty <strong>of</strong> first-order<br />

degradation rates in heterogeneous aquifers. Water Resour. Res., 42, W01420,<br />

10.1029/2004WR003878.<br />

Bauer, S., Beyer, C., Kolditz, O. (2007): Einfluss von Heterogenität und Messfehler auf die<br />

Bestimmung von Abbauraten erster Ordnung - eine Virtueller Aquifer Szenarioanalyse.<br />

(Influence <strong>of</strong> heterogeneity <strong>and</strong> measurement error on the determination <strong>of</strong> first order<br />

degradation rates by us1ing the virtual aquifer approach.), Grundwasser, 12, 3–14,<br />

10.1007/s00767-007-0019-8<br />

Bauer, S., Beyer, C., Chen, C., Gronewold, J., Kolditz, O., (2006b): Virtueller Aquifer (VA) -<br />

Computergestützte Bewertung von Erkundungs-, Sanierungs- und Monitoringstrategien im<br />

Hinblick auf das "Natural Attenuation" (NA) und "Enhanced Natural Attenuation" (ENA) -<br />

Potenzial kontaminierter Böden und Grundwässer. Statusseminar des KORA-TV 7, 8.6.2006,<br />

Dresden. Gemeinsame Mitteilungen des DGFZ e.V. und seiner Partner, 3/2006, Dresden, 93-<br />

113.<br />

Baveye, P., Valocchi, A. (1989): An evaluation <strong>of</strong> mathematical models <strong>of</strong> the transport <strong>of</strong><br />

biologically reacting solutes in <strong>saturated</strong> soils <strong>and</strong> aquifers. Water Resour. Res., 25, 1413–1421,<br />

10.1029/89WR00367.<br />

Bear, J. (1972): Dynamics <strong>of</strong> Fluids in Porous Media. American Elsevier Publishing Co., New York,<br />

761 p.<br />

Bear, J. (1979) Hydraulics <strong>of</strong> Groundwater. McGraw-Hill Book Co., London 567 p.<br />

Bear, J., Bachmat, Y. (1990): Introduction to Modeling <strong>of</strong> Transport Phenomena in Porous Media.<br />

Kluwer Academic Publ., 554 p.<br />

BBodSchV (1999): Bundes-Bodenschutz- und Altlastenverordnung vom 16. Juli 1999. Bundesgesetzblatt<br />

Jahrgang 1999, Teil I Nr. 36, 1554-1682.<br />

Bekins, B.A., Warren, E., Godsy, E.M. (1998): A comparison <strong>of</strong> zero-order, first-order, <strong>and</strong> Monod<br />

biotransformation models. Ground Water, 36, 261-268.<br />

Beyer, C., Bauer, S., Kolditz, O. (2006): Uncertainty Assessment <strong>of</strong> Contaminant Plume Length<br />

Estimates in Heterogeneous Aquifers. J. Contam. Hydrol., 87, 73-95,<br />

10.1016/j.jconhyd.2006.04.006.<br />

Beyer, C., Chen, C., Gronewold, J., Kolditz, O., Bauer, S. (2007a): Determination <strong>of</strong> first order<br />

degradation rate constants from monitoring networks. (accepted by Ground Water.).<br />

Beyer, C., Konrad, W., Park, C.H., Bauer, S., Rügner, H., Liedl, R., Grathwohl, P. (2007b):<br />

Modellbasierte Sickerwasserprognose für die Verwertung von Recycling-Baust<strong>of</strong>f in<br />

23


technischen Bauwerken. (Model based prognosis <strong>of</strong> contaminant leaching for reuse <strong>of</strong><br />

demolition waste in construction projects.) (accepted by Grundwasser <strong>and</strong> published online<br />

via SpringerLink), 10.1007/s00767-007-0025-x.<br />

BGR (Bundesanstalt für Geowissenschaften und Rohst<strong>of</strong>fe) (2006): Nutzungsdifferenzierte Bodenübersichtskarte<br />

der Bundesrepublik Deutschl<strong>and</strong> 1:1.000.000 (BÜK 1000 N2.3). Auszugskarten<br />

Acker, Grünl<strong>and</strong>, Wald; Digit. Archiv FISBo BGR; Hannover und Berlin.<br />

Bitterlich, S., Durner, W., Iden, S.C., Knabner, P. (2004): Inverse Estimation <strong>of</strong> the Un<strong>saturated</strong> Soil<br />

Hydraulic Properties from Column Out<strong>flow</strong> Experiments Using Free-Form Parameterizations,<br />

Vadose Zone J., 3, 971-981.<br />

Brooks, R.H., Corey, A.T. (1966): Properties <strong>of</strong> porous media affecting fluid <strong>flow</strong>. Proc. Amer. Soc.<br />

Civ. Eng., 92 [IR2], 61–87.<br />

Brusseau, M.L., Rao, P.S.C. (1989): Sorption nonideality during organic contaminant transport in<br />

porous media. Crit. Rev. Environ. Control, 19, 33–99.<br />

Buscheck, T.E., Alcantar, C.M. (1995): Regression techniques <strong>and</strong> analytical solutions to demonstrate<br />

intrinsic bioremediation. In Hinchee, R.E., Wilson, T.J., Downey, D. (Eds.), Intrinsic<br />

Bioremediation. Battelle Press, Columbus, OH, 109-116.<br />

Carrayrou, J., Mosé, R., Behra, P. (2004): Operator-splitting procedures for reactive transport <strong>and</strong><br />

comparison <strong>of</strong> mass balance errors. J. Contam. Hydrol., 68: 239-268, 10.1016/S0169-<br />

7722(03)00141-4.<br />

Celia, M., Russel, T., Herrera, I. Ewing, R. (1990): An Eulerian-Lagrangian localized adjoint method<br />

for the advection-diffusion equation. Adv. Water Resour., 13, 187-206, 10.1016/0309-<br />

1708(90)90041-2<br />

Cvetkovic, V., Dagan, G. (1996): Reactive transport <strong>and</strong> immiscible <strong>flow</strong> in geological media. 2<br />

Applications.- Proc. Royal Soc. London A, 452, 303-328.<br />

Dagan, G. (1989): Flow <strong>and</strong> Transport in Porous Formations. Springer, Heidelberg. 465 p.<br />

Du, Y., Wang, W. Kolditz, O. (2005): Benchmarking <strong>of</strong> Richards Model. GeoSys – Preprint [2005-<br />

12], GeoSystemsResearch, Center for <strong>Applied</strong> Geosciences, University <strong>of</strong> Tübingen.<br />

Durner, W. (1994): Hydraulic conductivity estimation for soils with heterogeneous pore structure,<br />

Water Resour. Res., 30, 211-223, 10.1029/93WR02676.<br />

Durner, W., Flühler, H. (2005): Chapter 74: Soil Hydraulic Properties, in: Anderson M.G. <strong>and</strong> J. J.<br />

McDonnell (eds.), Encyclopedia <strong>of</strong> Hydrological Sciences, John Wiley & Sons, Ltd., Chapter<br />

74, 1103-1120,<br />

Fetter, C.W. (1993): Contaminant Hydrogeology, Prentice Hall, New Jersey, 500 p.<br />

Finkel, M., Liedl, R., Teutsch, G. (1998): Modelling surfactant-enhanced remediation <strong>of</strong> polycyclic<br />

aromatic hydrocarbons. J. Environ. Modelling & S<strong>of</strong>tware, 14, 203-211. 10.1016/S1364-<br />

8152(98)00071-1.<br />

Freeze, R.A., Cherry, J.A. (1979): Groundwater. Prentice Hall, Englewood Cliffs, New Jersey, 604 p.<br />

Grathwohl, P. (1998): Diffusion in Natural Porous Media: Contaminant Transport,<br />

Sorption/Desorption <strong>and</strong> Dissolution Kinetics. Kluwer Academic Publishers, 224 p.<br />

Haverkamp, R., Vaclin, M., Touma, J., Wierenga, P.J., Vachaud, G. (1977): A comparison <strong>of</strong><br />

<strong>numerical</strong> simulation models for one-dimensional infiltration. Soil Sci. Soc. Am. J., 41, 285-<br />

294.<br />

Helmig, R. (1997): Multiphase Flow <strong>and</strong> Transport Processes in the Subsurface: A Contribution to<br />

the Modeling <strong>of</strong> Hydrosystems. Springer, Berlin, Heidelberg, 255 p.<br />

Huyakorn, P.S., Pinder, G.F. (1983): Computational Methods in Subsurface Flow. Academic Press,<br />

New York, 473 p.<br />

Islam, J., Naresh, S., O’Sullivan, M., (2001): Modeling biogeochemical processes in leachatecontaminated<br />

soils: A review. Transp. Porous Media, 43, 407-440, 10.1023/A:1010737825232.<br />

Jäger, R., Liedl, R. (2000): Prognose der Sorptionskinetik organischer Schadst<strong>of</strong>fe in heterogenem<br />

Aquifermaterial. Grundwasser, 2, 57-66.<br />

Jury, W.A., Gardner, W.R., Gardner, W.H. (1991): Soil Physics. John Wiley & Sons, New York,<br />

328 p.<br />

Kinzelbach, W. (1983): Analytische Lösungen der Schadst<strong>of</strong>ftransportgleichung und ihre Anwendung<br />

auf Schadensfälle mit flüchtigen Chlorkohlenwasserst<strong>of</strong>fen. In: Methoden zur rechnerischen<br />

24


Erfassung und hydraulischen Sanierung von Grundwasserkontaminationen, Mitteilungen des<br />

Instituts für Wasserbau, Universität Stuttgart, Heft 54, 115-199.<br />

Kinzelbach, W., Schäfer, W., Herzer, J. (1991): Numerical <strong>modeling</strong> <strong>of</strong> natural <strong>and</strong> enhanced<br />

denitrification processes in aquifers. Water Resour. Res., 27, 1123-1136, 10.1029/91WR00474.<br />

Knabner, P., Schneid, E. (2002): Adaptive Hybrid Mixed Finite Element Discretization <strong>of</strong><br />

Instationary Variably Saturated Flow in Porous Media. In: Breuer, M.,. et al., (eds.), High<br />

Performance Scientific <strong>and</strong> Engineering Computing, Springer Verlag, Berlin, 37-44.<br />

Korsawe, J., Starke, G., Wang, W., Kolditz, O. (2006): Finite Element Analysis <strong>of</strong> Poro-Elastic<br />

Consolidation in Porous Media: St<strong>and</strong>ard <strong>and</strong> Mixed Approaches. Comput. Methods Appl.<br />

Mech. Engrg., 195, 1096-1115, 10.1016/j.cma.2005.04.011<br />

Kolditz, O. (2002): Computational methods in environmental fluid dynamics. Springer, Heidelberg,<br />

400 p.<br />

Kolditz, O., Bauer, S. (2004): A process-oriented approach to computing multi-field problems in<br />

porous media. J. Hydroinf., 6, 225-244.<br />

Kolditz, O., Xie, M., Kalbacher, T., Bauer, S., Wang, W., McDermott, C., Chen, C., Beyer, C.,<br />

Gronewold, J., Kemmler, D., Walsh, R., Du, Y., Park, C.H., Hess, M., Amanidis, P. (2006):<br />

GeoSys/Rock<strong>flow</strong> version 4.4.03 - Theory <strong>and</strong> users manual, Center for <strong>Applied</strong> Geoscience,<br />

University <strong>of</strong> Tübingen.<br />

Ma, L., Selim, H.M. (1994): Predicting atrazine adsorption–desorption in soils: a modified secondorder<br />

kinetic model. Water Resour. Res., 30, 447–456. 10.1029/93WR02478<br />

Morris, B., Lawrence, A., Chilton, P., Adams, B., Calow, R., Klinck, B. (2003): Groundwater <strong>and</strong> its<br />

susceptibility to degradation: a global assessment <strong>of</strong> the problem <strong>and</strong> options for management.<br />

Nairobi: United Nations Environmental Programme, 140 p.<br />

Mualem, Y. (1976): A New model for predicting the hydraulic conductivity <strong>of</strong> un<strong>saturated</strong> porous<br />

media, Water Resour. Res., 12, 513 - 522.<br />

Neuman, S. (1981): An Eulerian-Lagrangian <strong>numerical</strong> scheme for the dispersion convection<br />

equation using conjugate space time grids, J. Comput. Phys. 41, 270-285.<br />

Newell, C.J., Rifai, H.S., Wilson, J.T., Connor, J.A., Aziz, J.A., Suarez, M.P. (2002): Calculation <strong>and</strong><br />

use <strong>of</strong> first-order rate constants for monitored natural attenuation studies. U.S. EPA Ground<br />

Water Issue, U.S. EPA/540/S-02/500.<br />

Olsen, S.R., Kemper, W.D. (1968): Movement <strong>of</strong> nutrients to plant roots. Adv. Agronomy, 30, 91-<br />

151.<br />

Park, C.H., Beyer, C., Bauer, S., Kolditz, O. (2006): An effcient method <strong>of</strong> r<strong>and</strong>om walk particle<br />

tracking: accuracy <strong>and</strong> resolution. (submitted to Adv. Water Resour.)<br />

Pinder, G.F., Gray, W.G. (1977): Finite element simulation in surface <strong>and</strong> subsurface hydrology.<br />

Academic Press, New York, 295 p.<br />

Rittmann, B., VanBriesen, J.M. (1996): Microbiological processes in reactive <strong>modeling</strong>. In: Lichtner,<br />

P., Steefel, C., Oelkers, E. (eds.), Reactive Transport in Porous Media. Reviews in Mineralogy,<br />

34, Mineralogical Society <strong>of</strong> America, Washington, DC, 311-334.<br />

Rügner, H., Kleineidam, S., Grathwohl, P. (1999): Long term sorption kinetics <strong>of</strong> phenanthrene in<br />

aquifer materials. Environ. Sci. Technol., 33, 1645–1651, 10.1021/es980664x<br />

Rubin, Y. (1983): Transport <strong>of</strong> reacting solutes in porous media: Relationship between mathematical<br />

nature <strong>of</strong> problem formulation <strong>and</strong> chemical nature <strong>of</strong> reactions. Water Res. Resour., 19, 1231-<br />

1252.<br />

Schäfer, D., Schäfer, W., Kinzelbach, W. (1998): Simulation <strong>of</strong> Reactive Processes Related to<br />

Biodegradation in Aquifers. 1. Structure <strong>of</strong> the 3D Reactive Transport Model, J. Contam.<br />

Hydrol., 31, 167-186, 10.1016/S0169-7722(97)00060-0<br />

Schäfer, D., Schlenz, B. Dahmke, A. (2004): Evaluation <strong>of</strong> exploration <strong>and</strong> monitoring methods for<br />

verification <strong>of</strong> natural attenuation using the virtual aquifer approach. Biodegradation, 15, 453-<br />

465, 10.1023/B:BIOD.0000044600.81216.00.<br />

Schäfer, D., Schlenz, B., Dahmke A. (2006a): Virtuelle Aquifere – ein Werkzeug zur Simulation von<br />

Natural Attenuation und zur Bewertung von Monitoringstrategien. Statusseminar des KORA-<br />

TV 7, 8.6.2006, Dresden. Gemeinsame Mitteilungen des DGFZ e.V. und seiner Partner, 3/2006,<br />

Dresden, 115-139.<br />

25


Schäfer, D., Dahmke, A., Kolditz, O., Teutsch, G. (2002): "Virtual Aquifers": A concept for<br />

evaluation <strong>of</strong> exploration, remediation <strong>and</strong> monitoring strategies. In: Kovar, K., Hrkal, Z. (eds.),<br />

Calibration <strong>and</strong> Reliability in Groundwater Modelling: A Few Steps Closer to Reality.<br />

Proceedings <strong>of</strong> the ModelCARE 2002 Conference, Prague, Czech Republic, June 2002. IAHS<br />

Publication 277, 52-59.<br />

Schäfer, D., Hornbruch, G. Schlenz, B., Dahmke A. (2006b): Schadst<strong>of</strong>fausbreitung unter Annahme<br />

verschiedener kinetischer Ansätze zur Modellierung mikrobiellen Abbaus. Grundwasser (in<br />

print).<br />

Scheidleder, A., Grath, J., Winkler, G., Stärk., U, Koreimann., C, Gmeiner, C., Nixon, S., Casillas, J.,<br />

Gravesen, P., Leonard, J., Elvira, M. (1999): Groundwater quality <strong>and</strong> quantity in Europe.<br />

European Environment Agency, Copenhagen, 123 p.<br />

Simkins, S., Alex<strong>and</strong>er, M. (1984): Models for mineralization kinetics with variables <strong>of</strong> substrate<br />

concentration <strong>and</strong> population density. Appl. Envir. Microbiol., 47, 1299-1306.<br />

Starke, G. (2000): Least-squares mixed finite element solution <strong>of</strong> variably <strong>saturated</strong> subsurface <strong>flow</strong><br />

problems. SIAM J. Sci. Comput. 21, 1869-1885, 10.1137/S1064827598339384<br />

Stenback, G.A., Ong, S.K., Rogers, S.W., Kjartanson, B.H. (2004): Impact <strong>of</strong> transverse <strong>and</strong><br />

longitudinal dispersion on first-order degradation rate constant estimation. J. Contam. Hydrol.,<br />

73, 3-14, 10.1016/j.jconhyd.2003.11.004.<br />

Streck, T., Poletika, N.N., Jury, W.A. Farmer, W.J. (1995): Description <strong>of</strong> simazine transport with<br />

rate-limited, two-stage linear <strong>and</strong> nonlinear sorption. Water Resour. Res., 31, 811-822,<br />

10.1029/94WR02822<br />

Thorenz, C. (1999): Model Adaptive Simulation <strong>of</strong> Multiphase <strong>and</strong> Density Driven Flow in Fractured<br />

<strong>and</strong> Porous Media. Dissertation, Institut für Strömungsmechanik, Universität Hannover, Bericht<br />

Nr. 62/2001.<br />

Van Genuchten, M.T. (1980): A closed form equation for predicting the hydraulic conductivity <strong>of</strong><br />

un<strong>saturated</strong> soils. Soil Sci. Soc. Am. J., 44, 892-898.<br />

Van Genuchten, M.T., Alves, W.J. (1982): Analytical solutions <strong>of</strong> the one-dimensional convective<br />

dispersive solute transport equation. USDA ARS Technical Bulletin Number 1661. U.S.<br />

Salinity Laboratory, 4500 Glenwood Drive, Riverside, CA 92501.<br />

Wang, W., Datcheva, M., Schanz, T., Kolditz, O. (2006): A sub-stepping approach for elastoplasticity<br />

with rotational hardening. Computational Mechanics. (in print). 10.1007/s00466-005-<br />

0710-5<br />

Wiedemeier, T.H., Swanson, M.A., Wilson, J.T., Kampbell, D.H., Miller, R.N., Hansen, J.E. (1996):<br />

Approximation <strong>of</strong> biodegradation rate constants for monoaromatic hydrocarbons (BTEX) in<br />

ground water. Ground Water Monit. Remed., 16 (3),186-194.<br />

Wiedemeier, T.H., Rifai, H.S., Newell, C.J., Wilson, J.T. (1999): Natural Attenuation <strong>of</strong> Fuels <strong>and</strong><br />

Chlorinated Solvents in the Subsurface. John Wiley <strong>and</strong> Sons, New York, 632 p.<br />

Wilson, R.D., Thornton, S.F., Mackay, D.M. (2004): Challenges in monitoring the natural attenuation<br />

<strong>of</strong> spatially variable plumes, Biodegradation, 15, 459-469,<br />

10.1023/B:BIOD.0000044591.45542.a9<br />

Xie, M., Bauer, S., Kolditz, O., Nowak, T., Shao, H. (2006): Numerical simulation <strong>of</strong> reactive<br />

processes in an experiment with partially <strong>saturated</strong> bentonite, J. Contam. Hydrol., 83, 122-147.<br />

10.1016/j.jconhyd.2005.11.003<br />

Zhang, Y.-K., Heathcote, R.C. (2003): An improved method for estimation <strong>of</strong> biodegradation rate<br />

with field data. Ground Water Monit. Remed., 23 (3), 112-116.<br />

Zheng, C., Bennett, G.D., (1995): <strong>Applied</strong> Contaminant Transport Modeling: Theory <strong>and</strong> Practice.<br />

Van Nostr<strong>and</strong> Reinhold, New York, 440 p.<br />

Zienkiewicz, O.C., Taylor, R. (1991): The Finite Element Method, Vol. 1, 2 (4th edition). McGraw-<br />

Hill, London, 459 p.<br />

26


Enclosed publications<br />

[EP 1] Bauer, S., Beyer, C., Kolditz, O., (2005): Assessing measurements <strong>of</strong> first order<br />

degradation rates by using the Virtual Aquifer approach. In: Thomson, N.R. (Ed.), GQ2004,<br />

Bringing Groundwater Quality Research to the Watershed Scale. Proceedings <strong>of</strong> a<br />

Symposium held at Waterloo, Canada, July 2004. IAHS Publication 297, IAHS Press,<br />

Wallingford, 274-281.<br />

[EP 2] Bauer, S., Beyer, C., Kolditz, O. (2006a): Assessing measurement uncertainty <strong>of</strong> first-order<br />

degradation rates in heterogeneous aquifers. Water Resour. Res., 42, W01420,<br />

10.1029/2004WR003878.<br />

[EP 3] Beyer, C., Bauer, S., Kolditz, O. (2006): Uncertainty assessment <strong>of</strong> contaminant plume<br />

length estimates in heterogeneous aquifers. J. Contam. Hydrol., 87, 73-95,<br />

[EP 4]<br />

10.1016/j.jconhyd.2006.04.006.<br />

Bauer, S., Beyer, C., Kolditz, O. (2007): Einfluss von Heterogenität und Messfehler auf die<br />

Bestimmung von Abbauraten erster Ordnung - eine Virtueller Aquifer Szenarioanalyse.<br />

(Influence <strong>of</strong> heterogeneity <strong>and</strong> measurement error on the determination <strong>of</strong> first order<br />

degradation rates by us1ing the virtual aquifer approach.), Grundwasser, 12, 3–14,<br />

10.1007/s00767-007-0019-8.<br />

[EP 5] Beyer, C., Chen, C., Gronewold, J., Kolditz, O., Bauer, S. (2007a): Determination <strong>of</strong> first<br />

order degradation rate constants from monitoring networks. (accepted by Ground Water.).<br />

[EP 6] Beyer, C., Konrad, W., Park, C.H., Bauer, S., Rügner, H., Liedl, R., Grathwohl, P. (2007b):<br />

Modellbasierte Sickerwasserprognose für die Verwertung von Recycling-Baust<strong>of</strong>f in<br />

technischen Bauwerken. (Model based prognosis <strong>of</strong> contaminant leaching for reuse <strong>of</strong><br />

demolition waste in construction projects.) (accepted by Grundwasser <strong>and</strong> published online<br />

via SpringerLink), 10.1007/s00767-007-0025-x.


Enclosed Publication 1<br />

Bauer, S., Beyer, C., Kolditz, O. (2005): Assessing measurements <strong>of</strong> first order degradation<br />

rates by using the Virtual Aquifer approach. In: Thomson, N.R. (Ed.), GQ2004, Bringing<br />

Groundwater Quality Research to the Watershed Scale. Waterloo, Canada, July 2004. IAHS<br />

Publication 297, IAHS Press, Wallingford, 274-281.<br />

The enclosed article was reproduced <strong>and</strong> is made available with the permission <strong>of</strong> IAHS<br />

Press.<br />

It can be obtained from IAHS Press at http://www.cig.ensmp.fr/~iahs/redbooks/297.htm.


274<br />

INTRODUCTION<br />

Bringing Groundwater Quality Research to the Watershed Scale (Proceedings <strong>of</strong> GQ2004, the 4th International<br />

Groundwater Quality Conference, held at Waterloo, Canada, July 2004). IAHS Publ. 297, 2005.<br />

Assessing measurements <strong>of</strong> first-order degradation<br />

rates through the virtual aquifer approach<br />

SEBASTIAN BAUER, CHRISTOF BEYER & OLAF KOLDITZ<br />

Center for <strong>Applied</strong> Geoscience, University <strong>of</strong> Tübingen, Sigwartstrasse 10, D 72076 Tübingen,<br />

Germany<br />

sebastian.bauer@uni-tuebingen.de<br />

Abstract The principal idea behind the “virtual aquifer” is to simulate <strong>and</strong><br />

evaluate investigation strategies for contaminated sites by modelling typical<br />

contamination scenarios. In this paper, first-order degradation rates using<br />

various methods were the focus <strong>of</strong> study. A virtual reality <strong>of</strong> a contaminated<br />

aquifer was generated by simulating the spreading <strong>of</strong> a plume, originating<br />

from a defined source zone, subject to first-order degradation. This plume was<br />

investigated through monitoring wells placed along the plume centre-line.<br />

Using information such as head measurements, concentration <strong>and</strong> hydraulic<br />

conductivity, first-order degradation rates were calculated <strong>and</strong> compared to the<br />

true predefined value. This comparison was conducted for varying degrees <strong>of</strong><br />

heterogeneity, represented by ln(KF), r<strong>and</strong>omly distributed conductivity fields.<br />

It was found that when heterogeneity was increased, “measured” degradation<br />

rates overestimated the true degradation rate by several orders <strong>of</strong> magnitude.<br />

The range <strong>of</strong> degradation rates obtained roughly corresponds to the range<br />

stated in literature values.<br />

Key words first-order degradation; modelling; natural attenuation; virtual reality<br />

At a real contaminated site, the true hydrogeological properties (e.g. conductivity,<br />

porosity, recharge rates, source position, degradation rates, etc.) are generally unknown<br />

(spatially). The basic idea behind the “virtual aquifer” is to produce a “virtual” contaminated<br />

site, where the spatial distribution <strong>of</strong> parameters is exactly known. By using a<br />

process-based <strong>flow</strong> <strong>and</strong> transport model, the fate <strong>of</strong> contaminants in the subsurface can<br />

be simulated, including plume development. The second step for creating a virtually<br />

contaminated site is to examine the virtual aquifer properties using st<strong>and</strong>ard investigative<br />

procedures (e.g. interpolating hydraulic head <strong>and</strong> contaminant concentrations<br />

measured at monitoring wells). The result <strong>of</strong> this virtual investigation can be compared<br />

to the true hydrogeological property distribution <strong>of</strong> the virtual aquifer because, contrary<br />

to an actual contaminated site, the exact distribution is known. The investigation<br />

techniques used can thus be tested <strong>and</strong> evaluated with respect to certain influences, i.e.<br />

sensitivity to aquifer heterogeneity or variation <strong>of</strong> other parameters. When the virtual<br />

plume is examined, only the data obtained by these investigation techniques are used,<br />

i.e. only the data that would also be measured at a real field site. Data such as hydraulic<br />

head <strong>and</strong> contaminant concentrations are “measured” in the virtual aquifer by “reading”<br />

the model output. The “virtual aquifer” approach <strong>of</strong>fers the ability to test <strong>and</strong> evaluate<br />

site investigation techniques, which cannot be performed in the real world. In this<br />

paper, four methods for determining first-order degradation rate constants, all based on<br />

the plume centre-line method, are examined by the “Virtual Aquifer” approach.


Assessing measurements <strong>of</strong> first-order degradation rates through the virtual aquifer approach 275<br />

METHODS<br />

Virtual aquifers were produced by generating r<strong>and</strong>om fields <strong>of</strong> hydraulic conductivity<br />

for the model area (dimensions: 184 × 64 m, Fig. 1). Flow direction was from left to<br />

right, with a mean hydraulic gradient <strong>of</strong> 0.003. For this application, a mean hydraulic<br />

conductivity <strong>of</strong> 7.2 × 10 -5 m s -1 was assumed, with ln(KF) variances <strong>of</strong> 0.38, 1.71, 2.7<br />

<strong>and</strong> 4.5; the variances were chosen to simulate different degrees <strong>of</strong> heterogeneity. An<br />

exponential variogram model, with an integral scale <strong>of</strong> 2.33 m, was used to describe<br />

spatial correlation. A virtual contaminant source zone <strong>of</strong> widths 4, 8 <strong>and</strong> 16 m was<br />

introduced into the aquifer, emitting a contaminant which was subject to a first-order<br />

degradation constant, λ, <strong>of</strong> 1 year -1 . A conservative tracer was also released. By using<br />

first order degradation kinetics for the reactive contaminant, the plume evolving<br />

corresponds to the methods used to estimate the first order degradation rate. This is<br />

certainly not true in reality, where the degradation <strong>of</strong> a contaminant follows changing<br />

<strong>and</strong> more complex kinetics. The plume was simulated using a process-based <strong>numerical</strong><br />

<strong>flow</strong> <strong>and</strong> transport model, assuming longitudinal <strong>and</strong> transverse dispersivities <strong>of</strong> 0.25<br />

<strong>and</strong> 0.05, respectively. Thus, the virtual contaminated aquifer was generated. In the<br />

second step, the plume’s properties were examined using the centre-line approach. For<br />

this investigation, not the full data <strong>of</strong> the model is used but only the values which are<br />

obtained by the centreline approach. There were three initial observation wells; one<br />

was directly in the source zone, while the other two were outside <strong>of</strong> the source (Fig. 1).<br />

At these three wells, the hydraulic heads were measured by reading the model output.<br />

A hydrogeological triangle was constructed <strong>and</strong> the direction <strong>of</strong> groundwater <strong>flow</strong> was<br />

thus determined. Along the estimated direction <strong>of</strong> groundwater <strong>flow</strong>, new observation<br />

wells were installed at every 10 m. These wells were then used to measure (arrows in<br />

Fig. 1) hydraulic head, contaminant concentrations, tracer concentrations <strong>and</strong> local<br />

hydraulic conductivity. From the hydraulic head difference, the true porosity <strong>and</strong> the<br />

well positions the respective groundwater <strong>flow</strong> velocities are calculated. Together with<br />

the concentration data, this allows the determination <strong>of</strong> the degradation rate constant<br />

Fig. 1 Method used to obtain plume centreline concentrations. The upper graph<br />

depicts the investigation procedure <strong>and</strong> the lower graph is the virtual site.<br />

virtual reality investigation


276<br />

S. Bauer et al.<br />

Table 1 Methods used for calculating first-order rate constants λ. va is the transport velocity, ∆x is the<br />

observation well distance, C(x) is the downstream concentration <strong>and</strong> C0 the source concentration, while<br />

αL <strong>and</strong> αT are the longitudinal <strong>and</strong> transverse dispersivities, WS is the source width <strong>and</strong> erf is the error<br />

function.<br />

Method Formula for degradation rate Description<br />

1<br />

va � C(<br />

x)<br />

�<br />

λ1<br />

= − ln� �<br />

∆x<br />

� �<br />

� C0<br />

�<br />

Analytical solution to 0-D transport equation<br />

(batch reactor)<br />

2<br />

v � ∗ �<br />

a � C(<br />

x)<br />

C0<br />

λ = −<br />

�<br />

2 ln<br />

∆x<br />

� C ∗ �<br />

� 0 C(<br />

x)<br />

�<br />

Concentration normalized to a non-degrading cocontaminant,<br />

thus accounting for dilution <strong>and</strong><br />

dispersion<br />

3<br />

�<br />

2<br />

v<br />

( ) �<br />

a ��<br />

ln C(<br />

x)<br />

C0<br />

�<br />

λ = � − α<br />

� −1�<br />

3 1 2 L<br />

4α<br />

�<br />

�<br />

L ��<br />

∆x<br />

� �<br />

Analytical solution to the 1-D transport equation.<br />

Accounts for longitudinal dispersion.<br />

4<br />

�<br />

2<br />

v<br />

( ) �<br />

a ��<br />

ln C(<br />

x)<br />

( C0β)<br />

�<br />

λ = � − α<br />

� −1�<br />

3 1 2 L<br />

4α<br />

�<br />

�<br />

L ��<br />

∆x<br />

� �<br />

Analytical solution to the 2-D transport equation.<br />

Accounts for longitudinal as well as transverse<br />

dispersion <strong>and</strong> a finite source width.<br />

� �<br />

with: � WS<br />

β = erf �<br />

� �<br />

� 4 αT<br />

∆x<br />

�<br />

first-order degradation; modelling; natural attenuation; virtual reality. The setup is thus<br />

designed to resemble ideal conditions for the application <strong>of</strong> the four methods for<br />

estimating the degradation rate constant. The only uncertainty <strong>and</strong> variability is<br />

introduced by the aquifer heterogeneity. The data used for the centreline method is<br />

thus only the data which can be measured at the three initial <strong>and</strong> the three downstream<br />

wells. Thus neither the mean hydraulic conductivity nor the variance or correlation<br />

length is known.<br />

The four centre-line methods used are provided in Table 1. Method 1 (Wiedemeier<br />

et al., 1996) is the batch solution to a first-order degradation (i.e. no transport is<br />

included). Method 2 was proposed by Wilson et al. (1994) (see also Wiedemeier et al.,<br />

1996) <strong>and</strong> is similar to Method 1, except amended concentrations are used. The<br />

measured concentrations for the reactive contaminant are corrected by using the ratio<br />

<strong>of</strong> the conservative co-contaminant at the observation well. This method corrects for<br />

dispersion <strong>and</strong> measurements taken outside <strong>of</strong> the plume centre-line at the three new<br />

observation wells. Method 3 was proposed by Buscheck & Alcantar (1995) <strong>and</strong> is<br />

based on the solution <strong>of</strong> a one-dimensional (1-D) transport equation with a first-order<br />

decay constant. This method accounts explicitly for longitudinal dispersion <strong>of</strong> the<br />

plume. To account for transverse dispersion (Method 4), Stenback et al. (2004) suggest<br />

using the analytical solution for a 2-D transport equation with first-order decay. In<br />

order to carry out Methods 3 <strong>and</strong> 4, it is required that the longitudinal <strong>and</strong> the transverse<br />

dispersivities be known. The following dispersivities were used according to<br />

Wiedemeier et al. (1999): 0.1 <strong>of</strong> the plume length for the longitudinal dispersivity, <strong>and</strong><br />

0.33 <strong>of</strong> longitudinal dispersivity for the transverse dispersivity. For each <strong>of</strong> the four<br />

approaches, first-order degradation rates were calculated <strong>and</strong> compared to the value<br />

used in generating the plume. For each degree <strong>of</strong> heterogeneity, 100 realizations were<br />

evaluated to obtain a statistical measure <strong>of</strong> the error introduced by the heterogeneity <strong>of</strong>


Assessing measurements <strong>of</strong> first-order degradation rates through the virtual aquifer approach 277<br />

the hydraulic conductivity. For each realization, the procedures described above were<br />

followed <strong>and</strong> a degradation rate was calculated for each method, at each downstream<br />

well <strong>and</strong> for every source width.<br />

RESULTS AND DISCUSSION<br />

Figure 2 illustrates the results for the calculated first-order rate constants. Calculated rate<br />

constants were reported as normalized rate constants (i.e. the calculated rate constant<br />

was divided by the true rate constant used in the <strong>numerical</strong> simulation). The normalized<br />

rate constant can thus be interpreted as an overestimated factor or an underestimated<br />

factor. Inspection <strong>of</strong> Fig. 2 yields that most calculated rates are higher than one (i.e. the<br />

degradation rate is overestimated). This conclusion is quite concerning for single<br />

realizations, where overestimations can be <strong>of</strong> several orders <strong>of</strong> magnitude. On the lefth<strong>and</strong><br />

side <strong>of</strong> Fig. 2(a), the variation <strong>of</strong> the calculated normalized rate with the source<br />

zone width is illustrated. It is clear that for Method 1, the calculated rates improve when<br />

the source zone width is increased; this is because Method 1 does not account for<br />

dilution, dispersion or measurements outside <strong>of</strong> the plume. These factors become less<br />

relevant with increasing source width since the basic assumptions inherent in Method 1<br />

are better fulfilled, <strong>and</strong> the overestimation factor drops accordingly. On the right-h<strong>and</strong><br />

side <strong>of</strong> Fig. 2(a), the dependence <strong>of</strong> the calculated rate on the degree <strong>of</strong> heterogeneity<br />

(given as variance) is shown. It is obvious that an increase <strong>of</strong> σ²ln(KF) leads to an<br />

overestimation <strong>of</strong> the calculated degradation rate. Furthermore, the st<strong>and</strong>ard deviation <strong>of</strong><br />

the mean calculated degradation rate increases, leading to greater uncertainty in the<br />

calculated rate. For the smallest degree <strong>of</strong> heterogeneity the mean overestimation is a<br />

factor a bit smaller than 2, which increases to values between 3 <strong>and</strong> 5 for medium to<br />

high heterogeneity, <strong>and</strong> 10 for very high heterogeneity.<br />

Figure 2(b) shows the results for Method 2. Degradation rates for this method were<br />

also overestimated. However, when comparing this method to Method 1, the overestimation<br />

factor <strong>and</strong> st<strong>and</strong>ard deviation are generally smaller (i.e. both the error <strong>and</strong><br />

the uncertainty are lower compared to Method 1). When examining the left-h<strong>and</strong> side<br />

<strong>of</strong> Fig. 2(b), the calculated rates show no dependence on source width. This effect is<br />

inherent to the method, since Method 2 accounts for dispersion, dilution <strong>and</strong><br />

measurements taken outside <strong>of</strong> the plume. Method 3 depicts results similar to Method<br />

1 regarding the rate dependence on source width <strong>and</strong> on the degree <strong>of</strong> heterogeneity<br />

(Fig. 2(c)). However, the normalized degradation rates for Method 3 are higher than<br />

for Method 1 (<strong>and</strong> also higher than Method 2); this is due to the dispersivity term. In<br />

comparison to Method 1, a portion <strong>of</strong> the concentration reduction from the source<br />

observation well to the downstream observation well is attributed to dispersion <strong>and</strong><br />

corrected for, <strong>and</strong> thus a higher degradation rate is estimated. Method 4 (Fig. 2(d))<br />

displays behaviour similar to Method 3, except that the rate values are slightly lower.<br />

Lower rate values are attributed to the additional term in the rate equation, which<br />

accounts for transverse dispersion. It should also be noted that at the lowest degree <strong>of</strong><br />

heterogeneity, the normalized degradation rates were actually underestimated; this is<br />

due to an “over correction” <strong>of</strong> the effects for transverse dispersion. To effectively<br />

illustrate the over <strong>and</strong> underestimation <strong>of</strong> the degradation rates for all four methods,<br />

the degradation rates where calculated (for a homogeneous hydraulic conductivity) <strong>and</strong>


278<br />

(a) Method 1<br />

(b) Method 2<br />

(c) Method 3<br />

(d) Method 4<br />

S. Bauer et al.<br />

Fig. 2 “Measured” first-order degradation rate constants normalized to the true<br />

degradation rate constant vs source width (left) <strong>and</strong> degree <strong>of</strong> heterogeneity (right) for<br />

(a) Method 1, (b) Method 2, (c) Method 3 <strong>and</strong> (d) Method 4. All figures show results<br />

for all observations (small symbols) as well as their mean value (large symbols) <strong>and</strong><br />

the corresponding st<strong>and</strong>ard deviations (error bars).


Assessing measurements <strong>of</strong> first-order degradation rates through the virtual aquifer approach 279<br />

plotted as small horizontal bars for σ²ln(KF) <strong>of</strong> 0 on the right-h<strong>and</strong> side graphs <strong>of</strong><br />

Fig. 2. It is obvious that for Method 1 <strong>and</strong> Method 2, the normalized degradation rate<br />

is exactly 1 for the homogeneous case (i.e. these methods yield the correct result).<br />

Method 3 shows a slight overestimation, while Method 4 yields very small, normalized<br />

degradation rates. The smallest rate, from Method 4, was obtained for the smallest<br />

source width, as then the correction is largest (compare Method 4 in Table 1). For large<br />

source widths, the argument <strong>of</strong> the error function <strong>of</strong> method 4 approaches 1.<br />

As mean <strong>and</strong> st<strong>and</strong>ard deviations are true for the ensemble mean, but not for single<br />

observations, an alternative method was chosen for comparing the four methods. The<br />

four methods were contrasted by plotting the probability <strong>of</strong> success against the error<br />

factor; the results are illustrated in Fig. 3. An error factor <strong>of</strong> 10 corresponds to an<br />

interval <strong>of</strong> 0.1 to 10 for normalized degradation rates, i.e. the interval that is obtained<br />

by multiplying 1 with the error factor <strong>and</strong> dividing 1 by the error factor (“within one<br />

order <strong>of</strong> magnitude”). An error factor <strong>of</strong> 5 thus corresponds to an interval <strong>of</strong> 0.2 to 5.<br />

These plots illustrate the probability that the measured degradation rate is within the<br />

interval <strong>of</strong> the corresponding error factor. Figure 3(a) represents the lowest degree <strong>of</strong><br />

heterogeneity <strong>and</strong> shows that the probability <strong>of</strong> calculating the degradation rate with an<br />

error factor <strong>of</strong> less than 2 (i.e. “within a factor <strong>of</strong> 2”) is about 0.7 for Method 1, 0.9 for<br />

Method 2, 0.55 for Method 3 <strong>and</strong> 0.3 for Method 4. When increasing the error factor<br />

(a) (b)<br />

(c) (d)<br />

Fig. 3 Probability plots for all methods for σ²ln(KF) <strong>of</strong>: (a) 0.38, (b) 1.71, (c) 2.7 <strong>and</strong><br />

(d) 4.5, resembling the four degrees <strong>of</strong> heterogeneity, shown for a source width <strong>of</strong><br />

4 m. The probability <strong>of</strong> method success is plotted against the error factor.


280<br />

S. Bauer et al.<br />

to 5, Methods 1, 2 <strong>and</strong> 3 show a success probability <strong>of</strong> about 1, while Method 4 yields<br />

a probability <strong>of</strong> 0.7. Figure 3 also illustrates that all methods exhibit a decreasing<br />

probability <strong>of</strong> success when the degree <strong>of</strong> heterogeneity is increased. An error factor<br />

<strong>of</strong> 5 yields a success probability <strong>of</strong> about 1 for the lowest degree <strong>of</strong> heterogeneity<br />

(0.38) for Method 1; this probability decreases to 0.7, 0.5 <strong>and</strong> 0.35 for a σ²ln(KF) <strong>of</strong><br />

1.71, 2.7 <strong>and</strong> 4.5, respectively. For Method 2 the corresponding success probabilities<br />

are 1.0, 0.9, 0.8 <strong>and</strong> 0.6; Method 3 yields values <strong>of</strong> 1.0, 0.55, 0.35 <strong>and</strong> 0.25; <strong>and</strong><br />

Method 4 0.7, 0.7, 0.6 <strong>and</strong> 0.4.<br />

To achieve the degradation rate within a factor <strong>of</strong> 10 <strong>of</strong> the correct degradation<br />

rate (for high hydraulic heterogeneity, Fig. 3(c)), the success probabilities were found<br />

to be 0.8, 0.95, 0.8 <strong>and</strong> 0.6 for Methods 1 through 4, respectively. For all degrees <strong>of</strong><br />

heterogeneity, Method 2 yielded the highest probability for achieving the correct<br />

degradation rate. For medium to very high heterogeneity, Method 4 was second best,<br />

<strong>and</strong> was the worst for aquifers that were only slightly heterogeneous (Fig. 3(a)). Method<br />

1, although the simplest method, yielded similar probabilities as Method 4, except<br />

Method 1 works well for aquifers <strong>of</strong> low heterogeneity. Method 3 yielded the lowest<br />

success probabilities, with the exception <strong>of</strong> an aquifer <strong>of</strong> low heterogeneity.<br />

CONCLUSIONS<br />

From the study presented, it can be concluded that the four methods for examining<br />

decay coefficients behave differently when heterogeneity is increased. All methods<br />

show a decrease in success probability with increasing heterogeneity. Figure 2<br />

illustrated that the decrease in success probability is attributed to an overestimation <strong>of</strong><br />

the degradation rate constant. The overestimation was largest for Method 3, yielding<br />

the lowest success probability. Method 2 was the least affected by an increase in<br />

heterogeneity <strong>and</strong> was also the method that depicted the lowest overestimation <strong>of</strong><br />

degradation rates. Methods 3 <strong>and</strong> 4, the most realistic since they are based on the 1-D<br />

<strong>and</strong> 2-D transport equations, illustrated low success probabilities <strong>and</strong> high overestimation<br />

<strong>of</strong> degradation rates, Method 3 being the worst. Methods 3 <strong>and</strong> 4 were<br />

prone to errors due to the introduction <strong>of</strong> longitudinal <strong>and</strong> transverse dispersivities,<br />

which were required to calculate the degradation rate constant. Method 1, the simplest<br />

method, yielded results comparable to Method 4.<br />

It can be concluded from this study, that Method 2 (using a non-reactive cocontaminant)<br />

is the preferred method for field investigations, where first-order<br />

degradation rates are to be estimated. If this co-contaminant is not available, then<br />

Method 1 is preferred as then no uncertainty regarding the dispersivities is introduced.<br />

Although widely used <strong>and</strong> published in the literature, the method <strong>of</strong> Buscheck &<br />

Alcantar (1995), Method 3, yielded the worst results in this study.<br />

Acknowledgements This work is funded by the German Ministry <strong>of</strong> Education <strong>and</strong><br />

Research as part <strong>of</strong> the KORA priority programme, sub-project 7.2.


Assessing measurements <strong>of</strong> first-order degradation rates through the virtual aquifer approach 281<br />

REFERENCES<br />

Buscheck, T. E. & Alcantar, C. M. (1995) Regression techniques <strong>and</strong> analytical solutions to demonstrate intrinsic<br />

bioremediation, In: Intrinsic Bioremediation (ed. by R. E. Hinchee, T. J. Wilson, & D. Downey), 109–116. Batelle<br />

Press, Columbus, Ohio, USA.<br />

Stenback, G. A., Ong, S. K., Rogers, S. W. & Kjartonson, B. H. (2004) Impact <strong>of</strong> transverse <strong>and</strong> longitudinal dispersion on<br />

first-order degradation rate constant estimation. J. Contam. Hydrol. 73, 3–14.<br />

Wiedemeier, T. H., Swanson, M. A., Wilson, J. T., Kampbell, D. H., Miller, R. N. & Hansen, J. E. (1996) Approximation<br />

<strong>of</strong> biodegradation rate constants for monoaromatic hydrocarbons (BTEX) in ground water. Ground Water<br />

Monitoring Remed. 16(3), 186–194.<br />

Wiedemeier, T. H., Rifai, H. S., Wilson, J. T. & Newell, C. (1999) Natural Attenuation <strong>of</strong> Fuels <strong>and</strong> Chlorinated Solvents<br />

in the Subsurface. Wiley, New York, USA.<br />

Wilson, J. T., Pfeffer, F. M., Weaver, J. W., Kampbell, D. H., Wiedemeier, T. H., Hansen, J. E., & Miller, R. N. (1994)<br />

Intrinsic bioremediation <strong>of</strong> JP-4 jet fuel. In: Symposium on Intrinsic Bioremediation <strong>of</strong> Ground Water (Denver,<br />

Colorado, USA), 60–72. US-EPA/540 R-94/515, Washington DC, USA.


Enclosed Publication 2<br />

Bauer, S., Beyer, C., Kolditz, O. (2006a): Assessing measurement uncertainty <strong>of</strong> first-order<br />

degradation rates in heterogeneous aquifers, Water Resour. Res., 42, W01420,<br />

doi:10.1029/2004WR003878. Copyright 2006 American Geophysical Union.<br />

Reproduced by permission <strong>of</strong> American Geophysical Union.<br />

The enclosed article can be obtained online from AGU at<br />

http://www.agu.org/pubs/crossref/2006.../2004WR003878.shtml.


WATER RESOURCES RESEARCH, VOL. 42, W01420, doi:10.1029/2004WR003878, 2006<br />

Assessing measurement uncertainty <strong>of</strong> first-order<br />

degradation rates in heterogeneous aquifers<br />

Sebastian Bauer, Christ<strong>of</strong> Beyer, <strong>and</strong> Olaf Kolditz<br />

Center for <strong>Applied</strong> Geoscience, University <strong>of</strong> Tübingen, Tübingen, Germany<br />

Received 7 December 2004; revised 7 October 2005; accepted 18 October 2005; published 31 January 2006.<br />

[1] The principal idea <strong>of</strong> this paper is to simulate <strong>and</strong> evaluate the determination <strong>of</strong><br />

first-order degradation rate constants at heterogeneous contaminated sites under realistic<br />

conditions. First, a set <strong>of</strong> heterogeneous <strong>and</strong> contaminated synthetic aquifers is generated;<br />

second, the spreading <strong>of</strong> a solute plume subject to first-order degradation is simulated.<br />

Third, this plume is investigated using ‘‘monitoring wells’’ placed along the presumed<br />

plume center line. Using only piezometric heads, concentrations <strong>and</strong> hydraulic<br />

conductivities obtained at these monitoring wells, first-order degradation rate constants are<br />

calculated by methods typically used in field applications. The estimated rate constants<br />

are compared to the ‘‘real’’ value known from the simulations. This comparison is<br />

conducted for different degrees <strong>of</strong> heterogeneity, represented by lognormally distributed<br />

r<strong>and</strong>om conductivity fields. The results indicate that, with increasing degree <strong>of</strong><br />

heterogeneity, ‘‘measured’’ degradation rate constants become uncertain with a high<br />

variability around the true constant. Measured rate constants tend to overestimate the true<br />

constant by up to one order <strong>of</strong> magnitude. A sensitivity analysis <strong>of</strong> the influences <strong>of</strong> source<br />

width, transport velocity, <strong>and</strong> dispersivity shows that (1) with increasing source width,<br />

measured rate constants decrease their relative error <strong>and</strong> increase their accuracy; (2) the<br />

choice <strong>of</strong> dispersivity can produce both over- <strong>and</strong> under-estimation <strong>of</strong> the true rate<br />

constant; <strong>and</strong> (3) that large-scale measurements <strong>of</strong> hydraulic conductivity yield better<br />

estimates <strong>of</strong> <strong>flow</strong> velocities as compared to local scale measurements. These results<br />

explain in part the high variability <strong>of</strong> field measured degradation rate constants reported in<br />

the literature.<br />

Citation: Bauer, S., C. Beyer, <strong>and</strong> O. Kolditz (2006), Assessing measurement uncertainty <strong>of</strong> first-order degradation rates in<br />

heterogeneous aquifers, Water Resour. Res., 42, W01420, doi:10.1029/2004WR003878.<br />

1. Introduction<br />

[2] This work studies the uncertainty involved in estimating<br />

first order degradation rate constants by the plume<br />

center line method for the assessment <strong>of</strong> natural attenuation<br />

at contaminated groundwater sites. Natural attenuation, also<br />

known as intrinsic bioremediation, refers to the observed<br />

reduction in contaminant concentration via natural processes<br />

as contaminants migrate from the source into environmental<br />

media [U.S. Environmental Protection Agency (EPA), 1999;<br />

Wiedemeier et al., 1999]. The processes contributing to<br />

natural attenuation include dilution, dispersion, sorption,<br />

volatilization <strong>and</strong> biodegradation, where biodegradation is<br />

the only process that decreases the total contaminant mass.<br />

The relative efficiencies <strong>of</strong> the attenuation processes active<br />

at a contaminated site must be carefully assessed before<br />

natural attenuation can be adopted as a cleanup remedy or<br />

risk reduction strategy. Thus degradation rates <strong>of</strong> the contaminants<br />

under consideration may play an important role in<br />

decision making <strong>and</strong> site management, when natural attenuation<br />

is considered as a remedial alternative or a remedial<br />

step in contaminated site management. Degradation rate<br />

constants can be used to estimate (1) the total overall natural<br />

Copyright 2006 by the American Geophysical Union.<br />

0043-1397/06/2004WR003878$09.00<br />

W01420<br />

attenuation potential <strong>of</strong> an aquifer, (2) contaminant plume<br />

lengths <strong>and</strong> (3) downstream concentrations. They can also be<br />

used for identifying potential receptors <strong>and</strong> exposure levels<br />

in case <strong>of</strong> a risk analysis.<br />

[3] Several approaches for estimating biodegradation<br />

rates in ground water in the field are commonly used,<br />

including mass balances, in situ microcosm studies <strong>and</strong><br />

the use <strong>of</strong> concentration-distance relations obtained along<br />

the plume center line [Chapelle et al., 1996; Wiedemeier et<br />

al., 1999]. The latter include a batch-reaction solution<br />

[Wiedemeier et al., 1996], normalization to a recalcitrant<br />

co-contaminant [Wiedemeier et al., 1996, 1999] <strong>and</strong> the<br />

method <strong>of</strong> Buscheck <strong>and</strong> Alcantar [1995]. The method <strong>of</strong><br />

Buscheck <strong>and</strong> Alcantar [1995] utilizes contaminant concentrations<br />

measured along the plume center line, which are<br />

evaluated by an analytical solution to the one-dimensional<br />

transport equation with first-order degradation. The firstorder<br />

degradation rate is calculated from the concentrations<br />

<strong>and</strong> an assumed longitudinal dispersivity. An<br />

additional requirement is, that the plume has reached<br />

steady state. This approach has been used by a number<br />

<strong>of</strong> authors, e.g., Chapelle et al. [1996], Wiedemeier et al.<br />

[1996], Zamfirescu <strong>and</strong> Grathwohl [2001], Suarez <strong>and</strong> Rifai<br />

[2002] or Bockelmann et al. [2003]. Recently, two- <strong>and</strong><br />

three-dimensional approaches were suggested [Zhang <strong>and</strong><br />

Heathcote, 2003; Stenback et al., 2004] as extensions to the<br />

1<strong>of</strong>14


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES W01420<br />

method by Buscheck <strong>and</strong> Alcantar [1995], which are based<br />

on analytical solutions for transport in two <strong>and</strong> three dimensions<br />

<strong>and</strong> account for finite source widths as well as<br />

transverse dispersion. By a method comparison with the<br />

original data Zhang <strong>and</strong> Heathcote [2003] showed that the<br />

method <strong>of</strong> Buscheck <strong>and</strong> Alcantar [1995] overestimates<br />

the degradation rate by 21% <strong>and</strong> 65% in case <strong>of</strong> a two<strong>and</strong><br />

three-dimensional plume, respectively. McNab <strong>and</strong><br />

Dooher [1998] reported that the method by Buscheck <strong>and</strong><br />

Alcantar [1995] is easily subject to misinterpretation, as<br />

transverse dispersivities <strong>and</strong> temporal effects can produce<br />

center line concentration pr<strong>of</strong>iles which resemble a degrading<br />

contaminant, even in the absence <strong>of</strong> degradation.<br />

[4] The spatial variability <strong>of</strong> aquifer properties has a<br />

significant influence on the distribution <strong>of</strong> contaminants<br />

<strong>and</strong> plume development. As a consequence, the methods for<br />

the estimation <strong>of</strong> degradation rates presented above are<br />

prone to effects <strong>of</strong> hydraulic heterogeneity, as they rely on<br />

concentration samples along the (presumed) plume center<br />

line as well as on estimations <strong>of</strong> site specific dispersivity. As<br />

Wilson et al. [2004] point out, the center line <strong>of</strong> a plume can<br />

easily be missed by monitoring wells installed based on<br />

assumed, but incorrect, groundwater <strong>flow</strong> directions. Moreover,<br />

contaminant plumes may w<strong>and</strong>er in all three dimensions<br />

due to macroscale heterogeneities [Wilson et al.,<br />

2004]. However, so far no study has been reported in<br />

literature which investigates these effects. Aim <strong>of</strong> this work<br />

is therefore to assess the influence <strong>of</strong> spatially heterogeneous<br />

hydraulic conductivities on the determination <strong>of</strong> firstorder<br />

degradation rates using sets <strong>of</strong> synthetic aquifer<br />

models.<br />

[5] Owing to the limited accessibility <strong>of</strong> the subsurface,<br />

measurements <strong>of</strong> piezometric heads <strong>and</strong> contaminant concentrations<br />

at contaminated sites are sparse <strong>and</strong> may not be<br />

representative <strong>of</strong> the heterogeneous hydrogeologic conditions.<br />

Therefore site investigation is subject to uncertainty,<br />

reflecting the limited knowledge on the aquifer properties<br />

<strong>and</strong> the extent <strong>of</strong> the contamination. Owing to this uncertainty,<br />

field investigation methods for plume screening or<br />

measuring hydraulic conductivity or degradation rates can<br />

neither be tested nor verified in the field. The only way <strong>of</strong><br />

assessing the performance <strong>and</strong> reliability <strong>of</strong> field investigation<br />

methods is by studying them in synthetic aquifers<br />

within a Monte Carlo framework. By applying the investigation<br />

method under consideration in the synthetic contaminated<br />

<strong>and</strong> heterogeneous aquifer, the method results can be<br />

compared to the true values. These are known from the<br />

synthetic aquifer, unlike in reality, where the true values are<br />

unknown.<br />

[6] This approach uses synthetic aquifer models, which<br />

are generated as the first step based on statistical properties<br />

<strong>of</strong> real aquifers <strong>and</strong> have a defined source <strong>of</strong> contamination.<br />

A reactive transport model is then used to simulate the<br />

spreading <strong>of</strong> the plume, resulting in realistic concentration<br />

distributions in the synthetic aquifer. In comparison to the<br />

‘‘real world,’’ the unique advantage <strong>of</strong> the synthetic aquifer<br />

is that the spatial distribution <strong>of</strong> all physical <strong>and</strong> geochemical<br />

properties <strong>and</strong> parameters as well as the contaminant<br />

concentrations are exactly known. In the second step, the<br />

synthetic aquifer is investigated by st<strong>and</strong>ard monitoring <strong>and</strong><br />

investigation techniques. In this step, only the data obtained<br />

by the investigation methods, i.e., heads <strong>and</strong> concentrations<br />

2<strong>of</strong>14<br />

at the observation wells, is used, because in case <strong>of</strong> a real<br />

site investigation the true parameter distribution is unknown.<br />

In the third step, the results from the investigation<br />

are compared to the true values, which allows to test <strong>and</strong><br />

evaluate the investigation method used. Using synthetic<br />

aquifers <strong>of</strong>fers furthermore the possibility to single out the<br />

influence <strong>of</strong> different parameters, such that sources <strong>of</strong><br />

uncertainty <strong>and</strong> error for the investigation method can be<br />

studied individually. Owing to this possibility <strong>of</strong> extensive<br />

<strong>and</strong> detailed scenario analysis <strong>and</strong> visualization, this approach<br />

is well suited to explore the uncertainty involved in<br />

hydrogeologic investigation <strong>and</strong> management. It has been<br />

applied under the term ‘‘virtual aquifer’’ by Schäfer et al.<br />

[2002, 2004], Bauer et al. [2005] <strong>and</strong> Bauer <strong>and</strong> Kolditz<br />

[2006].<br />

[7] This paper uses synthetic heterogeneous <strong>and</strong> contaminated<br />

aquifers in a Monte Carlo approach to assess for the<br />

first time the influence <strong>of</strong> spatially heterogeneous hydraulic<br />

conductivities on the determination <strong>of</strong> first-order degradation<br />

rates. To this end, plumes formed by contaminants<br />

degrading according to a first-order degradation rate in<br />

aquifers <strong>of</strong> different degrees <strong>of</strong> heterogeneity are investigated<br />

by the center line approach. By comparison <strong>of</strong> the<br />

estimated degradation rate constant with the true degradation<br />

rate constant the methods are tested <strong>and</strong> evaluated. This<br />

is performed by individually studying the influence <strong>of</strong><br />

aquifer heterogeneity, source width, <strong>flow</strong> velocity <strong>and</strong><br />

dispersivity on the estimated rate constant.<br />

2. Methods<br />

2.1. Model Domain<br />

[8] The model domain used for the <strong>numerical</strong> investigation<br />

is a two-dimensional aquifer with 184 m length <strong>and</strong><br />

64 m width (Figure 1). Flow is from left to right, with a<br />

mean hydraulic gradient <strong>of</strong> 0.003, which is induced by<br />

fixed head boundary conditions on the left <strong>and</strong> the right<br />

h<strong>and</strong> side <strong>of</strong> the model domain. No <strong>flow</strong> boundary<br />

conditions are assigned to all other sides <strong>of</strong> the model<br />

domain. The model domain is discretized with a grid<br />

density <strong>of</strong> 0.5 m in both directions. A contaminant source<br />

is emplaced 11.5 m downstream <strong>of</strong> the in<strong>flow</strong> boundary in<br />

the center <strong>of</strong> the aquifer, emitting a contaminant subject to<br />

first-order degradation with a degradation rate constant l<br />

<strong>of</strong> 1 a 1 (one per year). The contaminant source is<br />

represented by a fixed concentration boundary condition<br />

at the source position. Neither sorption, i.e., retardation,<br />

nor volatilization or dilution by recharge are accounted for.<br />

Additionally, a conservative compound is emitted from the<br />

source. The model setup is thus designed to provide ideal<br />

conditions for the application <strong>of</strong> the four center line<br />

methods to be studied. This is certainly not the case in<br />

nature, where the reaction kinetics will follow more<br />

complicated laws <strong>and</strong> may be spatially dependent, or<br />

influences from sorption <strong>and</strong> dilution have to be accounted<br />

for. However, these assumptions are used here to be able<br />

to study the st<strong>and</strong>ard methods closely <strong>and</strong> evaluate individually<br />

the influence <strong>of</strong> heterogeneity <strong>of</strong> the hydraulic<br />

conductivity. Further studies will use model setups which<br />

incorporate, e.g., different degradation kinetics.<br />

[9] A plume is generated using a process based <strong>numerical</strong><br />

<strong>flow</strong> <strong>and</strong> reactive transport model. The simulation code


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES<br />

Figure 1. Model area <strong>of</strong> the synthetic aquifer <strong>and</strong><br />

boundary conditions applied.<br />

GeoSys/RockFlow [Kolditz, 2002; Kolditz et al., 2004] is<br />

used here, which solves the <strong>flow</strong> <strong>and</strong> transport equations by<br />

st<strong>and</strong>ard Galerkin finite element methods [e.g., Huyakorn<br />

<strong>and</strong> Pinder, 1983] <strong>and</strong> using implicit Euler time stepping.<br />

The governing equations are given as [e.g., Bear, 1972]:<br />

<strong>and</strong><br />

S @h<br />

@t<br />

¼rðKrhÞþq ð1Þ<br />

@C<br />

@t ¼ varC þrðDrCÞ lC ð2Þ<br />

where S is the storage coefficient, h is the piezometric head,<br />

K is the tensor <strong>of</strong> hydraulic conductivity, q are sources <strong>and</strong><br />

sinks <strong>of</strong> water, C is concentration, v a is the transport<br />

velocity, D is the dispersion tensor, l is the first order<br />

degradation rate constant <strong>and</strong> t is time. The model<br />

parameters used in this study are given in Table 1. Details<br />

on <strong>numerical</strong> <strong>and</strong> s<strong>of</strong>tware issues can be found in the work<br />

<strong>of</strong> Kolditz [2002] <strong>and</strong> Kolditz <strong>and</strong> Bauer [2004]. The<br />

simulation code has been used for ground water <strong>flow</strong> <strong>and</strong><br />

transport simulations by Kolditz et al. [1998], Diersch <strong>and</strong><br />

Kolditz [1998, 2002], Thorenz et al. [2002] <strong>and</strong> Beinhorn et<br />

al. [2005].<br />

[10] To study the effects <strong>of</strong> spatially variable hydraulic<br />

conductivity, K is regarded as a r<strong>and</strong>om variable following a<br />

lognormal distribution with an expected value <strong>of</strong> E[Y =<br />

ln(K)] = 9.54. This corresponds to an effective hydraulic<br />

conductivity K ef <strong>of</strong> 7.2 10 5 ms 1 using the geometric<br />

mean [Rubin, 2003]. Using a porosity n <strong>of</strong> 0.33, the mean<br />

transport velocity is given by 6.5 10 7 ms 1 . The spatial<br />

correlation structure is characterized by an isotropic exponential<br />

covariance function C Y = s Y 2 exp( Dh/lY), with an<br />

integral scale <strong>of</strong> l Y = 2.67 m <strong>and</strong> the variance s Y 2 . Four<br />

different cases <strong>of</strong> increasing heterogeneity with ln(K) variances<br />

s Y 2 <strong>of</strong> 0.38, 1.71, 2.70 <strong>and</strong> 4.50 are considered,<br />

representing mildly to highly heterogeneous conductivity<br />

fields. The value <strong>of</strong> s Y 2 = 0.38 as well as the integral scale lY<br />

is taken from the Borden field site [Sudicky, 1986]. The<br />

value <strong>of</strong> 1.71 stems from an alluvial valley aquifer in<br />

southern Germany [Herfort, 2000]. The values <strong>of</strong> 2.70<br />

<strong>and</strong> 4.50 were reported for the Columbus Air Force Base<br />

site [Rehfeldt et al., 1992]. The geostatistical s<strong>of</strong>tware tool<br />

gstat2.4 [Pebesma <strong>and</strong> Wesseling, 1998] is used to generate<br />

100 realizations <strong>of</strong> the r<strong>and</strong>om field for each value <strong>of</strong> s Y 2 by<br />

unconditional sequential Gaussian simulation. The r<strong>and</strong>om<br />

3<strong>of</strong>14<br />

K values are generated over a two-dimensional grid <strong>of</strong><br />

density 0.5 m, exactly matching the <strong>numerical</strong> grid. Thus,<br />

following a rule <strong>of</strong> thumb <strong>of</strong> Ababou et al. [1989], a<br />

sufficient resolution <strong>of</strong> 5.33 > 1 + sY 2 grid nodes per integral<br />

scale is ensured.<br />

[11] To generate steady state plumes, as required by the<br />

methods under consideration, a stationary <strong>flow</strong> field is<br />

assumed. The time development <strong>of</strong> the plume is calculated,<br />

until the plume has reached steady state. A local longitudinal<br />

dispersivity aL = 0.25 m <strong>and</strong> a local transversal<br />

dispersivity <strong>of</strong> aT = 0.05 m are used for the <strong>numerical</strong><br />

simulations (compare Table 1).<br />

2.2. Center Line Method<br />

[12] Four methods for the determination <strong>of</strong> first-order<br />

degradation rate constants are investigated here, which are<br />

all based on the plume center line method. Method 1 is<br />

based on the one-dimensional transport equation, considering<br />

advection <strong>and</strong> first-order degradation only. The steady<br />

state solution for the concentration pr<strong>of</strong>ile can be rearranged<br />

to yield the first-order degradation rate constant for method<br />

1, i.e., l 1 [T 1 ]as:<br />

l1 ¼ va<br />

Dx<br />

Cx<br />

ln ðÞ<br />

C0<br />

where va [L T 1 ] is the transport velocity, Dx [L] is the<br />

distance between the observation wells, <strong>and</strong> C0 <strong>and</strong> C(x)<br />

[M L 3 ] are the upstream <strong>and</strong> downstream contaminant<br />

concentrations at the observation wells. In this formulation,<br />

all concentration changes resulting from processes other<br />

than degradation, i.e., diffusion, dispersion <strong>and</strong> dilution,<br />

are attributed to degradation. Therefore the rate constant<br />

l1 determined with method 1 can be considered rather<br />

an overall (or bulk) attenuation rate than a degradation<br />

rate constant [Newell et al., 2002]. Also, if the<br />

downstream observation well is not placed on the plume<br />

center line, the measured concentration is smaller than on<br />

the plume center line <strong>and</strong> the degradation rate constant is<br />

overestimated.<br />

[13] Method 2 was proposed by Wiedemeier et al. [1996]<br />

<strong>and</strong> is based on the same transport equation as method 1.<br />

However, to overcome the above mentioned drawbacks,<br />

amended concentrations are used: The measured concentrations<br />

<strong>of</strong> the reactive contaminant are corrected by the ratio<br />

<strong>of</strong> upgradient concentration C* 0 to downgradient concentration<br />

C(x)* [M L 3 ] <strong>of</strong> a nondegrading co-contaminant at the<br />

same observation wells. Thus the method corrects for<br />

dispersion <strong>of</strong> the plume or for the effects <strong>of</strong> unintended<br />

measurements <strong>of</strong>f the plume center line. The degradation<br />

Table 1. Model Parameters Used in the Simulations<br />

Parameter Value<br />

Kef 7.2 10 5 ms 1<br />

lY sy<br />

2.67 m<br />

2<br />

n<br />

0, 0.38, 1.71, 2.7, 4.5<br />

0.33<br />

l 1a 1<br />

S, q 0<br />

aL 0.25 m<br />

0.05 m<br />

a T<br />

W01420<br />

ð3Þ


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES W01420<br />

rate constant for method 2 is then calculated as [Wiedemeier<br />

et al., 1996]:<br />

l2 ¼ va<br />

Dx<br />

Cx<br />

ln ðÞ<br />

C0<br />

C* 0<br />

C* ðÞ x<br />

A prerequisite for the application <strong>of</strong> method 2 is that<br />

physicochemical properties <strong>of</strong> degradable <strong>and</strong> recalcitrant<br />

compounds like Henry’s Law constants <strong>and</strong> sorption<br />

coefficients must be comparable. A group <strong>of</strong> substances<br />

that has been proven to be well suited for the normalization<br />

<strong>of</strong> downgradient concentrations in BTEX plumes under<br />

anaerobic conditions are several trimethylbenzene (TMB)<br />

isomers [EPA, 1998; Wiedemeier et al., 1996, 1999]. When<br />

TMB is subject to biodegradation the estimated rate<br />

constant will be less than the actual value. For chlorinated<br />

solvent plumes, inorganic compounds like chloride may be<br />

appropriate substances for the normalization. Reductive<br />

dechlorination results in the production <strong>of</strong> chloride along<br />

the <strong>flow</strong> path, which by means <strong>of</strong> a mass balance can be<br />

used to derive the correction factor [EPA, 1998].<br />

[14] The third method investigated here was proposed by<br />

Buscheck <strong>and</strong> Alcantar [1995]. It is based on the steady<br />

state solution to the one-dimensional transport equation,<br />

accounting for advection, dispersion <strong>and</strong> first-order degradation.<br />

In comparison to method 1, method 3 accounts<br />

additionally for effects <strong>of</strong> longitudinal dispersion <strong>and</strong> thus<br />

requires an estimate <strong>of</strong> the longitudinal dispersivity aL [m].<br />

The degradation rate constant for method 3 is given by<br />

Buscheck <strong>and</strong> Alcantar [1995]:<br />

l3 ¼<br />

va<br />

ln<br />

1 2aL<br />

4aL<br />

Cx ðÞ<br />

00<br />

B<br />

C0<br />

@@<br />

Dx<br />

1<br />

A<br />

2<br />

1<br />

ð4Þ<br />

C<br />

1A<br />

ð5Þ<br />

Method 4 used here is the modified method <strong>of</strong> Buscheck <strong>and</strong><br />

Alcantar [1995], as proposed by Zhang <strong>and</strong> Heathcote<br />

[2003]. Since in this study a two-dimensional synthetic<br />

aquifer is used to assess the different approaches, for method<br />

4 the analytical solution to the two-dimensional transport<br />

equation including first order decay [Domenico, 1987] is<br />

adopted. Method 4 accounts for a finite source width as well<br />

as longitudinal <strong>and</strong> transverse dispersion. Therefore longitudinal<br />

<strong>and</strong> transverse dispersivities aL [m] <strong>and</strong> aT [m] as<br />

well as the source width WS [m] perpendicular to the average<br />

<strong>flow</strong> direction have to be known or estimated. Given these<br />

prerequisites, the degradation rate constant is given as<br />

[Zhang <strong>and</strong> Heathcote, 2003]:<br />

l4 ¼<br />

va<br />

1<br />

4aL<br />

ln<br />

2aL<br />

Cx ðÞ<br />

00<br />

B<br />

@@<br />

C0b<br />

Dx<br />

b ¼ erf<br />

WS<br />

1<br />

A<br />

2<br />

1<br />

C<br />

1A<br />

4 ffiffiffiffiffiffiffiffiffiffiffi p ð6Þ<br />

aT Dx<br />

2.3. Investigation Scenario<br />

[15] The site investigation mimicked in this study is<br />

depicted in Figure 2, where a steady state plume has<br />

evolved from a contaminant source. This plume is investigated<br />

by the center line approach. Figure 2a shows the<br />

initial situation, where three observation wells are present in<br />

4<strong>of</strong>14<br />

Figure 2. Representation <strong>of</strong> the investigation method to<br />

obtain plume center line concentrations: (a) initial<br />

situation, (b) estimation <strong>of</strong> <strong>flow</strong> direction by application<br />

<strong>of</strong> a hydrogeologic triangle, (c) measured concentrations on<br />

the inferred center line, <strong>and</strong> (d) comparison to true<br />

concentrations <strong>and</strong> heads.<br />

the aquifer. One <strong>of</strong> these wells (solid circle) is in the source<br />

<strong>and</strong> concentrations are high, while the other two show no<br />

concentration. This situation is the starting point for the<br />

investigation scenario for all realizations <strong>and</strong> represents the<br />

initial knowledge on the site. In the first investigation step,<br />

the <strong>flow</strong> direction is determined (Figure 2b). Hydraulic<br />

heads are measured in the three wells (by reading the model<br />

output at the well positions), a hydrogeologic triangle is<br />

constructed <strong>and</strong> the hydraulic gradient is calculated. In the<br />

second investigation step, three new observation wells are<br />

installed every 10 m along the estimated direction <strong>of</strong> <strong>flow</strong><br />

(Figure 2c). These wells are then used to obtain hydraulic<br />

heads <strong>and</strong> concentrations <strong>of</strong> the contaminant <strong>and</strong> the nonreactive<br />

co-contaminant as well as local hydraulic conductivities<br />

at the observation wells by using the model input for<br />

hydraulic conductivity at the corresponding location. From<br />

the head difference, the true porosity <strong>and</strong> the well positions<br />

the respective groundwater <strong>flow</strong> velocities are calculated.<br />

An effective conductivity Kef between each upstream <strong>and</strong><br />

downstream well is estimated by calculating the geometric<br />

mean <strong>of</strong> the local conductivities. Together with the concentration<br />

data, this information allows for the determination <strong>of</strong>


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES<br />

the first-order degradation rate constants by the four methods<br />

presented above. The investigation setup is designed to<br />

resemble ideal conditions for the application <strong>of</strong> the four<br />

methods for estimating the degradation rate constant. All<br />

measurements are assumed to be exact, which means that<br />

there is no measurement error involved. The only uncertainty<br />

<strong>and</strong> variability is introduced by the heterogeneity <strong>of</strong><br />

hydraulic conductivity. For methods 3 <strong>and</strong> 4, additionally<br />

aL <strong>and</strong> aT have to be known. These are estimated following<br />

Wiedemeier et al. [1999] as 0.1 <strong>of</strong> the plume length for aL,<br />

with aT being about 0.33 <strong>of</strong> the longitudinal dispersivity. As<br />

plume length the maximum distance covered by the observation<br />

wells, i.e., 30 m, is used. As the plumes are generally<br />

longer, this assumption yields rather low dispersivities.<br />

Thus aL <strong>and</strong> aT are estimated to be 3.0 <strong>and</strong> 1.0 m,<br />

respectively. These estimates <strong>of</strong> dispersivities are not optimal,<br />

as they are not based on the heterogeneity <strong>of</strong> the<br />

hydraulic conductivity. However, dispersivities based on<br />

results from stochastic hydrogeology are difficult to obtain,<br />

as for most field sites structure <strong>and</strong> degree <strong>of</strong> heterogeneity<br />

are not well known. Also, the four samples taken in this<br />

investigation scenario do not allow for an estimation <strong>of</strong> the<br />

correlation length or the ln(K) variances. Both the correct<br />

source width <strong>and</strong> the correct porosity are used. In the last<br />

step, by each <strong>of</strong> the four approaches the corresponding<br />

first-order degradation rate constants l 1 through l 4 are<br />

calculated. These values can be compared to the value used<br />

to generate the plume (l =1a 1 ). For each realization, the<br />

investigation procedure described above is followed <strong>and</strong> a<br />

degradation rate is calculated for each method, each downstream<br />

well <strong>and</strong> for each source width. For each <strong>of</strong> the four<br />

classes <strong>of</strong> heterogeneity used in this study (ln(K) variances<br />

s Y 2 <strong>of</strong> 0.38, 1.71, 2.70 <strong>and</strong> 4.50) a minimum <strong>of</strong> 100 realizations<br />

is evaluated. Thus statistical measures <strong>of</strong> the errors<br />

<strong>and</strong> uncertainties introduced by the heterogeneity <strong>of</strong> the<br />

hydraulic conductivity are obtained. Additionally, also the<br />

impact <strong>of</strong> the width <strong>of</strong> the source zone is studied. Here it<br />

is expected, that for increasing source width the onedimensional<br />

methods yield better results, as then the<br />

investigated situation corresponds better to the assumptions<br />

<strong>of</strong> the method. Source widths W S <strong>of</strong> 4 m, 8 m <strong>and</strong> 16 m are<br />

used, corresponding to 1.5, 3 <strong>and</strong> 6 integral scales l Y. Then<br />

methods for estimating the <strong>flow</strong> velocity are elucidated for<br />

the different degrees <strong>of</strong> heterogeneity. This is because the<br />

goodness <strong>of</strong> the calculated value for lambda is directly<br />

related to estimated transport velocity accuracy. Finally the<br />

influence <strong>of</strong> estimated longitudinal <strong>and</strong> transversal dispersivities<br />

on results by methods 3 <strong>and</strong> 4 is studied in a<br />

sensitivity analysis.<br />

2.4. Numerical Tests<br />

[16] Convergence <strong>of</strong> the Monte Carlo simulation with<br />

regard to the sample size N <strong>of</strong> estimated degradation rate<br />

constants was tested by a procedure following Goovaerts<br />

[1999]. The test is only conducted for the highest degree <strong>of</strong><br />

heterogeneity used in this study (sY 2 = 4.5) <strong>and</strong> the smallest<br />

source width <strong>of</strong> 4 m, as this is the case <strong>of</strong> highest variability.<br />

A total <strong>of</strong> 1000 realizations <strong>of</strong> the r<strong>and</strong>om conductivity field<br />

was generated. For each realization, plume development<br />

was simulated <strong>and</strong> the degradation rate constant l1 was<br />

calculated using method 1. The resulting set <strong>of</strong> 1000<br />

degradation rates is assumed to be sufficiently large to<br />

Figure 3. Influence <strong>of</strong> Monte Carlo sample size N on the<br />

average subset mean (l1), the st<strong>and</strong>ard deviation <strong>of</strong> subset<br />

means (sl1 ), <strong>and</strong> the average st<strong>and</strong>ard deviation <strong>of</strong> the<br />

subset population (sl1 ) for the highest degree <strong>of</strong> heterogeneity<br />

(sY 2 ).<br />

represent the global population. The global population <strong>of</strong><br />

l1 was r<strong>and</strong>omly sampled with a sample size <strong>of</strong> N =2,<br />

yielding a subset <strong>of</strong> two l1. For this subset the mean<br />

degradation rate as well as the st<strong>and</strong>ard deviation were<br />

calculated. R<strong>and</strong>om sampling was repeated 999 times, resulting<br />

in 1000 subsets <strong>of</strong> N = 2. From these subsets, the average<br />

subset mean l1, the st<strong>and</strong>ard deviation <strong>of</strong> subset means sl1 <strong>and</strong> the average st<strong>and</strong>ard deviation <strong>of</strong> the subset population<br />

sl1 are calculated. R<strong>and</strong>om sampling was repeated with<br />

increasing subset sizes N =3,4,..., 1000, resulting in 999<br />

triplets <strong>of</strong> the statistics, one for each subset size. Dependence<br />

<strong>of</strong> the three statistics on N is shown in Figure 3.<br />

[17] The middle curve in Figure 3 displays the average<br />

subset mean l1, which shows almost no dependence on N<br />

<strong>and</strong> yields values very close to the global mean <strong>of</strong> 8.9. The<br />

st<strong>and</strong>ard deviation <strong>of</strong> the subset means (s , lower curve)<br />

l1<br />

shows a strong decrease from 12.3 (N = 2) to 2.2 (N = 50)<br />

<strong>and</strong> 1.5 (N = 100), with a significantly reduced decrease for<br />

larger subset sizes. In relation to the global mean <strong>of</strong> 8.9, the<br />

variation among the subsets is therefore small for N 100.<br />

The upper curve in Figure 3 shows the average st<strong>and</strong>ard<br />

deviation <strong>of</strong> the r<strong>and</strong>om sample subsets sl1 , which strongly<br />

increases with subset size for small N, but with a much<br />

smaller increase for N > 50. For N = 100 a value <strong>of</strong> 15.5 is<br />

found, which is 95% <strong>of</strong> the st<strong>and</strong>ard deviation <strong>of</strong> the global<br />

population, as obtained for N = 1000. The observed reduction<br />

in increase <strong>of</strong> sl1 with N indicates the redundancy <strong>of</strong><br />

additional realizations with regard to the subset variability.<br />

As the rate <strong>of</strong> decrease <strong>of</strong> s as well as the rate <strong>of</strong> increase<br />

l1<br />

<strong>of</strong> sl1 becomes small for more than 100 realizations, we feel<br />

confident that a sample size <strong>of</strong> N = 100 is sufficient to yield<br />

stable ensemble averaged rate coefficients. Since the analysis<br />

was conducted for the largest degree <strong>of</strong> heterogeneity,<br />

5<strong>of</strong>14<br />

W01420


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES W01420<br />

Figure 4. Estimated first-order degradation rate constants Li (normalized to the true rate constant l)<br />

versus degree <strong>of</strong> heterogeneity sY 2 for (a) method 1, (b) method 2, (c) method 3, (d) <strong>and</strong> method 4. All<br />

figures show results for all single realizations (small symbols) as well as their ensemble means (large<br />

crosses) with their corresponding st<strong>and</strong>ard deviations (error bars) <strong>and</strong> ensemble medians (large diamonds).<br />

The reference rate constant used in the <strong>numerical</strong> simulations is indicated by the horizontal line.<br />

for lower values <strong>of</strong> s Y 2 convergence can be expected already<br />

at lower sample sizes.<br />

[18] Another check was performed concerning the mean<br />

flux over the <strong>flow</strong> domain. Although deviations between<br />

single realizations are quite distinct <strong>and</strong> increase with s Y 2 ,<br />

the ensemble averages for each s Y 2 match the theoretical<br />

value with less than a 1% error. Furthermore, the correct<br />

operation <strong>of</strong> the investigation methods was verified by<br />

applying the investigation procedure for a source <strong>of</strong> infinite<br />

width, i.e., a width equal to the model area, <strong>and</strong> by<br />

assuming the aquifer is homogeneous. Then all methods<br />

reduce to the one-dimensional method <strong>and</strong> yield the correct<br />

degradation rate constant. This result was obtained <strong>and</strong> thus<br />

the correct operation verified.<br />

3. Results <strong>and</strong> Discussion<br />

3.1. Influence <strong>of</strong> Heterogeneous Conductivity<br />

[19] To examine the influence <strong>of</strong> heterogeneity on the<br />

estimation <strong>of</strong> the rate constants, a contaminant source <strong>of</strong><br />

6<strong>of</strong>14<br />

width W S = 4 m perpendicular to the average <strong>flow</strong> direction<br />

is emplaced in the synthetic aquifer <strong>and</strong> <strong>flow</strong> as well as<br />

reactive transport are simulated. With the investigation<br />

scenario described in section 2.3 applied to a single realization,<br />

each <strong>of</strong> the four methods yields differing rate<br />

constants, l i, for each <strong>of</strong> the three center line observation<br />

wells (10 m, 20 m, <strong>and</strong> 30 m distance from the source). For<br />

the assessment <strong>of</strong> the four methods, these rate constants<br />

l i,10, l i,20 <strong>and</strong> l i,30 are averaged to yield one single<br />

estimated l i for each method. This procedure is repeated<br />

for all realizations. Figure 4 presents results <strong>of</strong> the calculated<br />

rate constants l i versus the degree <strong>of</strong> heterogeneity<br />

(given as ln conductivity variance s Y 2 ). Calculated rates li<br />

are reported as normalized rates L i, i.e., the calculated rate<br />

l i is divided by the true rate l used in the <strong>numerical</strong><br />

simulation. The normalized rate constants L i thus can be<br />

interpreted as over- or under-estimation factors. The homogeneous<br />

case (s Y 2 = 0) is included for reference.<br />

[20] Figure 4a presents the results for method 1, i.e., based<br />

on the one-dimensional advection-degradation solution


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES<br />

(equation (3)). It can be clearly seen, that an increase <strong>of</strong><br />

sY 2 leads to an increase in spread <strong>of</strong> the single realizations.<br />

Quite a number <strong>of</strong> realizations exhibit normalized rate<br />

constants L1 <strong>of</strong> more than 10 up to about 100, i.e., the<br />

true rate constant is severely overestimated by a factor <strong>of</strong><br />

10 to 100. Also, L1 increases with increasing sY 2 as well<br />

as the st<strong>and</strong>ard deviations <strong>of</strong> the ensemble means. This<br />

points to an increase in uncertainty <strong>and</strong> reflects the spread<br />

<strong>of</strong> the single realizations. Mean overestimation increases<br />

from a factor <strong>of</strong> about 1.6 for sY 2 = 0.38 to about 10.2 in<br />

the case <strong>of</strong> sY 2 = 4.50. In the homogeneous case (sY 2 =0),<br />

method 1 yields the correct result. Although on average l<br />

is overestimated, even for high values <strong>of</strong> sY 2 in single<br />

realizations l may actually be underestimated by L1. For<br />

the highest degree <strong>of</strong> heterogeneity, the L1 <strong>of</strong> the single<br />

realizations span about three orders <strong>of</strong> magnitude. Comparison<br />

<strong>of</strong> ensemble means with the corresponding<br />

medians shows that in all cases the medians are significantly<br />

lower. The populations <strong>of</strong> estimated l1 are positively<br />

skewed as some exceedingly large values <strong>of</strong> l1<br />

shift the means to high values. However, the general trend<br />

<strong>of</strong> increasing overestimation with heterogeneity is also<br />

distinct for the medians.<br />

[21] Figure 4b shows the corresponding results obtained<br />

with method 2, i.e., using the one dimensional advectiondegradation<br />

equation with normalization to a recalcitrant cocontaminant<br />

(equation (4)) [Wiedemeier et al., 1996]. As for<br />

method 1, the spread <strong>of</strong> the single realizations increases<br />

with increasing s Y 2 . Compared to method 1, however, the<br />

spread is smaller <strong>and</strong> more equally distributed about L 2 =1.<br />

Therefore the average overestimation factors as well as the<br />

st<strong>and</strong>ard deviations are much smaller than for method 1 <strong>and</strong><br />

both the error <strong>and</strong> the uncertainty are lower. Average rate<br />

constants L 2 are 1.1 for s Y 2 = 0.38, increasing to 3.3 for sY 2 =<br />

4.50. For homogeneous conditions, also method 2 yields the<br />

correct result, i.e., L 2 = 1 for s Y 2 = 0. Ensemble medians are<br />

just slightly above the true l, i.e., deviating less than a<br />

factor <strong>of</strong> 2.<br />

[22] Degradation rate constants calculated with method<br />

3, i.e., the one-dimensional method introduced by<br />

Buscheck <strong>and</strong> Alcantar [1995] (equation (5)), are displayed<br />

in Figure 4c. They exhibit a similar general<br />

behavior as found with method 1, i.e., increasing spread<br />

<strong>and</strong> increasing overestimation <strong>of</strong> the true rate constant for<br />

higher s Y 2 . However, mean L3 are significantly higher<br />

than the corresponding L 1, with ensemble means <strong>of</strong> 2.0<br />

for s Y 2 = 0.38 increasing to 29.4 for sY 2 = 4.50. In the<br />

homogeneous case, the true rate constant is slightly<br />

overestimated, i.e., L 3 =1.25fors Y 2 = 0. Spread in the<br />

single realizations is higher compared to method 1, now<br />

spanning nearly four orders <strong>of</strong> magnitude for the largest<br />

variance value <strong>of</strong> s Y 2 .<br />

[23] Results for method 4, i.e., the two-dimensional<br />

solution (equation (6)) suggested by Zhang <strong>and</strong> Heathcote<br />

[2003], are depicted in Figure 4d. Compared to the other<br />

methods, method 4 displays the largest spread <strong>of</strong> calculated<br />

L 4 around the mean values. For the highest degree <strong>of</strong><br />

heterogeneity, the spread <strong>of</strong> the single realizations covers<br />

nearly five orders <strong>of</strong> magnitude. Ensemble means <strong>of</strong> L 4<br />

increase from 0.6 for s Y 2 = 0.38 to about 23.0 for sY 2 = 4.5,<br />

i.e., for low s Y 2 the normalized rate constant L4 is actually<br />

underestimated in most realizations, while for larger varian-<br />

7<strong>of</strong>14<br />

W01420<br />

ces the ensemble averages approach the results <strong>of</strong> method 3.<br />

In the homogeneous case (sY 2 = 0), the estimated rate<br />

constant l4 is about two orders <strong>of</strong> magnitude lower than<br />

the true rate constant l. Ensemble medians for method 4<br />

show a similar behavior as the ensemble means, with their<br />

values closer to the true rate constant than for methods 1<br />

<strong>and</strong> 3 for large heterogeneities (sY 2 = 1.71). This reflects the<br />

fact, that the spread <strong>of</strong> the single realizations is distributed<br />

symmetrically around L4 = 1, however, the spread <strong>of</strong> the<br />

populations <strong>and</strong> thus the uncertainty is significantly larger<br />

than for methods 1 <strong>and</strong> 3.<br />

[24] Method 1 is based on the one-dimensional solution<br />

to first-order biodegradation <strong>and</strong> advection. Therefore it is<br />

expected that rate constants estimated with method 1<br />

overestimate the true rate constant due to two effects.<br />

Firstly, method 1 does not account for measuring <strong>of</strong>f the<br />

center line. So if an observation well is placed <strong>of</strong>f the plume<br />

center line, concentrations sampled there will be smaller<br />

than on the center line <strong>and</strong> therefore the degradation rate<br />

constant will be estimated too high. Secondly, as method 1<br />

is based on a one-dimensional solution <strong>of</strong> the transport<br />

equation, it does not account for transverse dispersion,<br />

which lowers concentrations on the plume center line. This<br />

second effect also causes an overestimation <strong>of</strong> the rate<br />

constant. Both effects together cause the overestimation<br />

<strong>of</strong> the rate constant as shown in Figure 4a. Method 2<br />

tries to overcome these two problems by normalization<br />

to a conservative tracer. Both above effects also determine<br />

the concentration <strong>of</strong> the nonreactive component, <strong>and</strong> are<br />

thus corrected for by the normalization. As is shown in<br />

Figure 4b, results <strong>of</strong> method 2 are considerably better than<br />

<strong>of</strong> method 1, both considering spread <strong>and</strong> ensemble averages.<br />

However, as can be seen from Figure 4b, method 2<br />

does not correct for all effects, as overestimation is observed<br />

with increasing s Y 2 . Effects <strong>of</strong> dispersion <strong>and</strong> measuring <strong>of</strong>f<br />

the center line are accounted for by method 2, so the<br />

deviation seen for method 2 has to have a hydraulic cause.<br />

This deviation is introduced by the determination <strong>of</strong> the<br />

average <strong>flow</strong> velocity between the observation wells, which<br />

is calculated using an averaged value <strong>of</strong> the hydraulic<br />

conductivity at the two observation wells. This averaged<br />

value may not be representative <strong>of</strong> the <strong>flow</strong> path between the<br />

two wells <strong>and</strong> bias may be introduced into the calculation <strong>of</strong><br />

degradation rate constants. This effects is studied closely<br />

below in section 3.3.<br />

[25] Method 3 is based on the one-dimensional transport<br />

equation including advection, degradation <strong>and</strong> longitudinal<br />

dispersion. Results from method 3, as shown in Figure 4c,<br />

display a higher spread <strong>and</strong> higher ensemble averages<br />

compared to method 1. Because method 3 includes longitudinal<br />

dispersion, it should be closer to reality <strong>and</strong> advantageous<br />

over method 1. The differences in estimated rate<br />

constants between method 1 <strong>and</strong> 3 are therefore due to the<br />

longitudinal dispersivity a L in method 3. With the onedimensional<br />

transport model used, pronounced longitudinal<br />

dispersion <strong>of</strong> a degrading contaminant results in a stronger<br />

spreading <strong>of</strong> the solute downstream <strong>and</strong> thus in higher<br />

concentrations along the plume center line compared to an<br />

advection only case. Therefore a larger rate constant is<br />

calculated to accomplish a given concentration decrease<br />

between the upgradient <strong>and</strong> the downgradient observation<br />

well. l 3 grows linearly with a L <strong>and</strong> is always larger than l 1,


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES W01420<br />

as can be seen by exp<strong>and</strong>ing the squared brackets in<br />

equation (3) <strong>and</strong> using C(x) C0:<br />

l3 ¼ va aL<br />

va<br />

lnð Cx ðÞ=C0<br />

Dx<br />

lnð Cx ðÞ=C0<br />

Dx<br />

2<br />

¼ l1<br />

!<br />

lnðCx ðÞ=C0Þ<br />

Dx<br />

[26] If a L = 0, method 3 reduces to the advection only<br />

case, i.e., method 1. Method 3 still does not account for<br />

transverse dispersion, which is the process causing smaller<br />

concentrations on the plume center line. Method 4 is based<br />

on a two-dimensional solution to the transport equation<br />

including advection, longitudinal <strong>and</strong> transverse dispersion<br />

<strong>and</strong> first-order degradation. Results from method 4<br />

(Figure 4d) show an underestimation <strong>of</strong> the true rate<br />

constant for homogeneous or slightly heterogeneous conditions<br />

(sY 2<br />

1.71), while for high degrees <strong>of</strong> heterogeneity<br />

(sY 2<br />

2.7), the ensemble averages <strong>of</strong> the estimated rate<br />

constants approach the respective values obtained with<br />

method 3. The underestimation for low heterogeneities is<br />

a consequence <strong>of</strong> the correction for transverse dispersion,<br />

represented by the error function b in equation (4). The<br />

effect <strong>of</strong> lower concentrations along the plume center line<br />

due to transverse dispersion is strong for small source<br />

widths, large transversal dispersivities <strong>and</strong> large well spacings.<br />

b is always less than 1 <strong>and</strong> asymptotically approaches<br />

unity for arguments <strong>of</strong> the error function larger than 2, i.e.,<br />

l4 converges toward l3 for small aT or large WS. For<br />

arguments


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES<br />

Figure 5. Estimated first-order degradation rate constants versus degree <strong>of</strong> heterogeneity sY 2 for<br />

(a) method 1, (b) method 2, (c) method 3, <strong>and</strong> (d) method 4 <strong>and</strong> source widths WS <strong>of</strong> 4, 8, <strong>and</strong> 16 m,<br />

respectively. Solid lines show ensemble means normalized to the true degradation rate constant (left<br />

axis), dashed lines show the corresponding st<strong>and</strong>ard deviations (right axis).<br />

[29] In the following, the influence <strong>of</strong> different approaches<br />

for estimating K (<strong>and</strong> thus va) on the determination <strong>of</strong> li<br />

is analyzed. As the plume samples only part <strong>of</strong> the full<br />

domain, the hydraulic conductivity needed is not the effective<br />

conductivity for the complete domain, but an equivalent<br />

hydraulic conductivity valid for the <strong>flow</strong> path along the<br />

center line wells. This equivalent hydraulic conductivity<br />

may differ from the global effective value. During the field<br />

investigation by the center line method, only the local<br />

hydraulic conductivities measured at the observation wells<br />

placed along the inferred plume center line or only the<br />

global effective hydraulic conductivity are known. Neither<br />

the underlying statistical parameters nor the complete conductivity<br />

field are known. The first approach represents the<br />

situation, where local hydraulic conductivities have been<br />

obtained in the observation wells, e.g., by slug tests or sieve<br />

analysis, <strong>and</strong> are averaged to obtain an estimate <strong>of</strong> the<br />

equivalent hydraulic conductivity between the two observation<br />

wells used for rate determination. In addition to the<br />

geometric mean Kg, as used so far, also the arithmetic mean<br />

Ka <strong>and</strong> the harmonic mean Kh are used, as they present<br />

9<strong>of</strong>14<br />

W01420<br />

upper <strong>and</strong> lower limits for Kef [e.g., Renard <strong>and</strong> de Marsily,<br />

1997]. As hydraulic conductivity data is available only at<br />

four wells, a full characterization <strong>of</strong> the conductivity field is<br />

not possible. Often at a real site, values <strong>of</strong> hydraulic<br />

conductivity are not available at the locations <strong>of</strong> the<br />

observation wells at the plume center line, but only from<br />

a single well not on the plume center line <strong>and</strong> thus not used<br />

for rate determination. This case is simulated by using a<br />

value KS <strong>of</strong> hydraulic conductivity from a well at point<br />

[55.10 m, 55.20 m] (see Figure 1). In the last approach, the<br />

true global effective conductivity KG <strong>of</strong> the synthetic<br />

aquifer is known, i.e., from a large-scale pumping test,<br />

<strong>and</strong> KG is used as an estimator for the local equivalent<br />

hydraulic conductivity at the observation wells.<br />

[30] Figure 6 shows the ensemble averages (Figure 6a)<br />

<strong>and</strong> medians (Figure 6b) <strong>of</strong> the degradation rate constants<br />

estimated using the above five approaches. Only results for<br />

method 2 <strong>and</strong> WS = 16 m are presented here, because<br />

method 2 is only affected by the hydraulic error due to a<br />

wrong estimation <strong>of</strong> <strong>flow</strong> velocity. Inspection <strong>of</strong> Figure 6a<br />

shows that all approaches lead to an overestimation <strong>of</strong> the


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES W01420<br />

Figure 6. Ensemble (a) means <strong>and</strong> (b) medians <strong>of</strong><br />

calculated L 2 for the different approaches <strong>of</strong> average <strong>flow</strong><br />

velocity estimation <strong>and</strong> (c) the corresponding coefficients <strong>of</strong><br />

variation.<br />

ensemble averaged L2, which increases with sY 2 . This<br />

increase is most pronounced for Ka <strong>and</strong> KS. For sY 2 = 4.50<br />

the average L2,K a is about 4.3 times larger than L2,K g ,<br />

whereas for single realizations the difference may reach a<br />

10 <strong>of</strong> 14<br />

factor <strong>of</strong> 70 (not shown here). If only the single KS value is<br />

used, overestimation is even larger for all degrees <strong>of</strong><br />

heterogeneity. In comparison to the geometric average,<br />

using the harmonic average Kh or the true global effective<br />

conductivity KG reduces the overestimation <strong>of</strong> L2 considerably.<br />

Both approaches result in nearly identical ensemble<br />

averages. However, inspecting ensemble medians as shown<br />

in Figure 6b, harmonic averaging <strong>of</strong> K measurements results<br />

in an underestimation <strong>of</strong> L2 in more than 50% <strong>of</strong> all<br />

realizations. Taking this into account, usage <strong>of</strong> the global<br />

KG seems to be the best approach for the estimation <strong>of</strong> the<br />

local average <strong>flow</strong> velocity between two observation wells<br />

along the plume center line. This finding is supported by<br />

Figure 6c, which shows coefficients <strong>of</strong> variation (CV) <strong>of</strong>L2<br />

as a measure <strong>of</strong> uncertainty. Using KG yields the smallest<br />

spread. Comparison <strong>of</strong> all five approaches yields: CVK G <<br />

CVK g CVK h < CVK a < CVK S for all sY 2 . These findings<br />

indicate that for the conditions given in this work the best<br />

estimate <strong>of</strong> local <strong>flow</strong> velocities <strong>and</strong> <strong>of</strong> the degradation rate<br />

constant is given by the global geometric average <strong>of</strong> the<br />

hydraulic conductivity. This result is expected for large well<br />

distances, as then ergodic conditions have been reached.<br />

Our results show that also for nonergodic conditions due to<br />

small well distances <strong>of</strong> just a few integral scales the global<br />

geometric average yields better estimates <strong>of</strong> local <strong>flow</strong><br />

velocity compared to using locally measured hydraulic<br />

conductivities. The spread observed in Figure 6 is thus<br />

due to the nonergodic conditions <strong>of</strong> single realizations, i.e.,<br />

the fact that the plume has sampled only a part <strong>of</strong> the<br />

domain. Thus for a field case, a single large-scale pumping<br />

test would be preferable to small-scale local information<br />

obtained directly at the sampling wells.<br />

3.4. Influence <strong>of</strong> Dispersivity Parameterization<br />

[31] As demonstrated in the previous sections, the parameterization<br />

<strong>of</strong> macrodispersivities a L <strong>and</strong> a T solely<br />

based on the scale <strong>of</strong> the contaminant plume according to<br />

Wiedemeier et al. [1999] using 0.1 <strong>of</strong> the assumed plume<br />

length for a L <strong>and</strong> a T = a L/3 introduces a significant<br />

additional error when rate constants are estimated with<br />

methods 3 or 4. For method 3, this error is always toward<br />

higher rate constants, while for method 4 errors in both<br />

directions may occur. It is known that aL <strong>and</strong> aT do not only<br />

depend on plume scale, because the variability <strong>of</strong> <strong>flow</strong><br />

velocity resulting from the heterogeneity <strong>of</strong> hydraulic conductivity<br />

is also important for the spreading during solute<br />

transport. If additional information on the spatial distribution<br />

<strong>and</strong> variability <strong>of</strong> hydraulic conductivity is available,<br />

i.e., from an geologically analogous aquifer, where these<br />

parameters have been determined, this information can be<br />

used to obtain estimates <strong>of</strong> the dispersivity values based on<br />

aquifer heterogeneity. So far in the manuscript, only estimates<br />

based on plume length have been used.<br />

[32] In a two-dimensional isotropic domain with zero<br />

local dispersivity <strong>and</strong> for ergodic conditions, the asymptotic<br />

large time longitudinal dispersivity is given by aL = sY 2 lY,<br />

whereas the asymptotic limit <strong>of</strong> aT is 0 [Dagan, 1989;<br />

Rubin, 2003]. For the two-dimensional model setup investigated<br />

in this manuscript <strong>and</strong> the values <strong>of</strong> sY 2 (0.38, 1,71,<br />

2.7, 4.5) <strong>and</strong> lY (2.67) m, the corresponding values for the<br />

asymptotic limit <strong>of</strong> aL are 1.02, 4.57, 7.21 <strong>and</strong> 12.02 m,<br />

respectively. Close to the source, i.e., in the preasymptotic


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES<br />

Figure 7. Ensemble medians <strong>of</strong> estimated first-order rate constants L 4 for (a) s Y 2 = 0.38, (b) 1.71,<br />

(c) 2.7, <strong>and</strong> (d) 4.5 for different values <strong>of</strong> a L versus a T. Values for method 3 are obtained using<br />

L 3(a L)=L 4(a L, a T = 0). Medians for L 1 <strong>and</strong> L 2 are depicted in each diagram by dashed <strong>and</strong> dash-dotted<br />

horizontal lines for reference.<br />

region, aL will show lower values. aT grows with travel<br />

distance until aT reaches peak values at about 2.5 lY <strong>of</strong> 0.1,<br />

0.44, 0.69 <strong>and</strong> 1.15 m, respectively. The asymptotic limit<br />

<strong>of</strong> aT is 0. For nonzero but small local dispersivities (aL,T<br />

l 1<br />

Y 1), aL is unaffected or slightly reduced, while aT<br />

increases <strong>and</strong> a non zero long time limit will establish<br />

[Zhang <strong>and</strong> Neuman, 1996]. As the observation wells used<br />

for the investigation <strong>of</strong> the plumes are located within 3.8 to<br />

11.2 lY from the contaminant source, it is expected, that the<br />

asymptotic limits <strong>of</strong> aL <strong>and</strong> aT are not yet reached. Thus<br />

for methods 3 <strong>and</strong> 4, aL values below those listed above<br />

should be used. Similarly, aT values in the range between<br />

the peak values <strong>and</strong> the asymptotic limit should be used<br />

for method 4. As this rough estimation already shows,<br />

applying the stochastically derived values in a real field<br />

case may introduce uncertainty, as the local conditions are<br />

generally not known exactly.<br />

[33] To study the influence <strong>of</strong> dispersivity parameterization<br />

on rate constants estimated with methods 3 <strong>and</strong> 4, the<br />

sensitivity <strong>of</strong> L3 <strong>and</strong> L4 on different values <strong>of</strong> aL <strong>and</strong> aT is<br />

investigated. Because in a real field application these values<br />

11 <strong>of</strong> 14<br />

W01420<br />

will be always uncertain, this investigation is performed as a<br />

sensitivity analysis, which allows to cover a wider range <strong>of</strong><br />

values for aL <strong>and</strong> aT. For aL values <strong>of</strong> 12, 9, 6, 3, 0.8 <strong>and</strong><br />

0.25 m are used, while for aT values <strong>of</strong> 2, 1.15, 1, 0.75, 0.5,<br />

0.3, 0.07 <strong>and</strong> 0 m are used. Thus the range <strong>of</strong> reasonable<br />

values <strong>of</strong> aL <strong>and</strong> aT up to the maximum values for each sY 2<br />

is well represented. Figure 7 presents for each sY 2 the<br />

medians <strong>of</strong> estimated rate constants L3, using L3(aL) =<br />

L4(aL, aT = 0), <strong>and</strong> L4(aL, aT) in dependence on aT for the<br />

different values <strong>of</strong> aL. For reference, also medians for L1<br />

<strong>and</strong> L2 (compare Figure 4) are included. Because method 2<br />

corrects for dispersion as well as for measuring <strong>of</strong>f the<br />

plume center line, but not for errors in the estimated<br />

transport velocity, <strong>and</strong> the same value <strong>of</strong> the transport<br />

velocity is used for all four methods, L2 can be seen as<br />

the optimal lower limit for methods 3 <strong>and</strong> 4. Method 1<br />

involves no correction for longitudinal <strong>and</strong> transverse dispersion<br />

or measuring <strong>of</strong>f the center line. Thus a good<br />

parameterization <strong>of</strong> aL <strong>and</strong> aT for method 4 should result<br />

in rate constant estimates L4 < L1. However, L4 < L2 would<br />

indicate over correction for transverse dispersion.


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES W01420<br />

[34] Results <strong>of</strong> the sensitivity study are presented in<br />

Figure 7. For all degrees <strong>of</strong> heterogeneity, decreasing<br />

median degradation rates are found with increasing aT.<br />

Differences between the different aL are most distinct for<br />

low values <strong>of</strong> aT. Considering method 3, it is clearly shown<br />

that L3 = L1 only for aL = 0, <strong>and</strong> L3 > L1 for aL >0.<br />

This demonstrates again that method 3 always yields<br />

higher estimated degradation rate constants as compared<br />

to method 1 (compare equation (7)). Considering that<br />

generally l is overestimated in our study, accounting only<br />

for aL by using method 3 aggravates this problem.<br />

[35] For method 4 it is found that for low <strong>and</strong> medium<br />

heterogeneity values for aL <strong>and</strong> aT exist, which allow for<br />

an optimal estimation <strong>of</strong> the degradation rate constant, i.e.,<br />

L4 = 1, <strong>and</strong> thus L4 L2. For sY 2 = 2.70, L4 > 1.0<br />

always, but L4 < L2 can be achieved for large values <strong>of</strong><br />

aT <strong>and</strong> small values <strong>of</strong> aL. However, for the highest<br />

degree <strong>of</strong> heterogeneity, L4 > L2 > 1 always. For small<br />

values <strong>of</strong> aT, the degradation rate is overestimated for all<br />

degrees <strong>of</strong> heterogeneity, while for low <strong>and</strong> medium<br />

heterogeneity L4 < 1 is possible for large values <strong>of</strong> aT.<br />

[36] For sY 2 = 0.38 <strong>and</strong> aT = 0.07 m, all medians are larger<br />

than L 1 (when a L > 0.8), at a T = 0.3 all medians are 0.5 m (Figure 7c). However, even<br />

for an unrealistically large a T <strong>of</strong> 2 m, L 4 is still significantly<br />

larger than L 2. A similar behavior is found for s Y 2 =4.5<br />

(Figure 7d), where only using a L = 9 m <strong>and</strong> a T >1.15m<br />

will result in L 4 being closer to the true rate constant than<br />

the corresponding L 1.<br />

[37] These results show, that a wide range <strong>of</strong> dispersivities<br />

can <strong>and</strong> must be used to obtain better estimates using<br />

method 4 than using method 1 or method 2. It is also found<br />

that the theoretical values (as given above) lead only for the<br />

case <strong>of</strong> low heterogeneity to considerably better estimates<br />

than using method 1. Therefore a large fraction <strong>of</strong> the<br />

observed overestimation must result from <strong>of</strong>f center line<br />

measurements, as it cannot be corrected for by reasonable<br />

values <strong>of</strong> a T. Especially for the cases <strong>of</strong> high heterogeneity,<br />

the effects <strong>of</strong> dispersion seem to be minor in comparison to<br />

the effect <strong>of</strong> missing the center line.<br />

[38] As shown above, method 4 could yield estimated<br />

rate coefficients that are as close or even closer to the true<br />

rate constant than rate coefficients estimated with method 2,<br />

regardless <strong>of</strong> the degree <strong>of</strong> heterogeneity. However, this<br />

requires unreasonably low values for a L <strong>and</strong> very high<br />

values for a T, as these parameters would have to correct for<br />

the <strong>of</strong>f center line measurement errors. In this case, a L<br />

<strong>and</strong> a T would no longer represent the actual dispersivities,<br />

but are lumped fitting parameters. As the magnitude <strong>of</strong><br />

bias introduced by missing the center line as well as the<br />

exact values <strong>of</strong> s Y 2 or lY are usually not known at a real<br />

field site, choosing dispersivities is highly uncertain <strong>and</strong><br />

may cause over- as well as under-estimation <strong>of</strong> the<br />

degradation rate constant. Thus estimation <strong>of</strong> dispersivity<br />

12 <strong>of</strong> 14<br />

introduces an additional error into the estimation <strong>of</strong> degradation<br />

rate constants using methods 3 or 4. Only for aquifers<br />

<strong>of</strong> low heterogeneity method 4 yields better estimates than<br />

method 1. Better estimates than using method 2 are only<br />

possible by assuming unphysical dispersivity values.<br />

4. Summary <strong>and</strong> Conclusions<br />

[39] In this paper the performance <strong>of</strong> four different<br />

methods for the estimation <strong>of</strong> degradation rate constants<br />

in an aquifer with a heterogeneous distribution <strong>of</strong> the<br />

hydraulic conductivity is studied. All four methods are<br />

based on the center line approach. The results demonstrate<br />

that a heterogeneous distribution <strong>of</strong> the hydraulic conductivity<br />

may lead to severe overestimation <strong>of</strong> the ensemble<br />

averaged degradation rate constant. Furthermore, the single<br />

realizations show a large spread <strong>and</strong> a large st<strong>and</strong>ard<br />

deviation, indicating that results obtained from any one<br />

estimation are highly uncertain. Mean overestimation as<br />

well as spread increase with degree <strong>of</strong> aquifer heterogeneity.<br />

By method comparison, the main reasons are identified as<br />

‘‘measuring <strong>of</strong>f the plume center line’’ <strong>and</strong> effects <strong>of</strong><br />

transverse dispersion. Best method performance is observed<br />

for method 2, which is based on the one-dimensional<br />

transport equation including advection <strong>and</strong> first order degradation.<br />

By normalizing the measured concentrations <strong>of</strong> the<br />

degrading contaminant to a nonreactive co-contaminant<br />

emitted by the same source, the above mentioned effects<br />

are corrected for. Method 2 thus shows the lowest spread<br />

<strong>and</strong> the lowest overestimation <strong>of</strong> estimated degradation rate<br />

constants <strong>of</strong> all four methods. However, the presence <strong>of</strong> the<br />

recalcitrant co-contaminant needed for the normalization<br />

approach may not always be given at a site. Second best<br />

performance is observed for method 1, which yields consistently<br />

higher spread <strong>and</strong> degradation rate constant overestimation<br />

as compared to method 2. Methods 3 <strong>and</strong> 4,<br />

although more realistic in the sense that they base on the<br />

one-dimensional <strong>and</strong> two-dimensional transport equation,<br />

respectively, show higher spread <strong>and</strong> larger overestimation.<br />

Both methods are prone to errors introduced by estimating<br />

longitudinal <strong>and</strong> transverse dispersivities. The choice <strong>of</strong><br />

these values introduces additional uncertainty without yielding<br />

substantially better results than methods 1 or 2. The<br />

ensemble averaged degradation rate constant is highest for<br />

method 3, due to the longitudinal dispersivity term in<br />

equation (3), while performance <strong>of</strong> method 4 crucially<br />

depends on the choice <strong>of</strong> an appropriate transverse dispersivity<br />

value a T.Ifa T is chosen too small with respect to the<br />

real macrodispersivity at the field site under consideration,<br />

the degradation rate constant may be overestimated. If a T is<br />

chosen too large, the degradation rate may be underestimated.<br />

For all methods, a high spread <strong>of</strong> the results from the<br />

single realizations is found, causing a high uncertainty <strong>of</strong><br />

the estimated degradation rate constant for all methods. For<br />

a single realization, the estimated degradation rate may<br />

deviate by one or even two orders <strong>of</strong> magnitude from the<br />

correct value. This deviation is caused only by the heterogeneity<br />

<strong>of</strong> the hydraulic conductivity. The spread observed<br />

here may contribute to the spread observed in degradation<br />

rate constants observed in the field. Wiedemeier et al. [1999]<br />

<strong>and</strong> Aronson <strong>and</strong> Howard [1997] report measured degradation<br />

rate constants for PCE ranging over two orders <strong>of</strong>


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES<br />

magnitude from 0.07 a 1 to 1.2 a 1 <strong>and</strong> for TCE ranging<br />

from 0.05 a 1 to 4.75 a 1 . For benzene, toluene <strong>and</strong> xylene<br />

rate constants <strong>of</strong> 0.07–3.0 a 1 , 0.36–21.0 a 1 <strong>and</strong> 0.32–<br />

76.0 a 1 , respectively, are reported [Wiedemeier et al.,<br />

1999]. Method performance increases with increasing<br />

source width for all methods. For sources very wide with<br />

respect to the integral scale <strong>of</strong> the hydraulic conductivity<br />

field, all methods yield reasonable results. In reality, however,<br />

when sources are heterogeneous or formed by a<br />

complex combination <strong>of</strong> a number <strong>of</strong> zones, the total source<br />

width may be difficult to estimate. If degradation rates are<br />

used for assessing the NA potential at a contaminated site,<br />

overestimation <strong>of</strong> the degradation rates is a critical point.<br />

Overestimation <strong>of</strong> the degradation rate constant leads to an<br />

overestimation <strong>of</strong> the overall natural attenuation potential. If<br />

plume lengths are calculated with too high degradation rate<br />

constants, then estimated plume lengths are too short.<br />

Remediation times as well as downgradient concentrations<br />

may be underestimated. The results presented show that<br />

determination <strong>of</strong> degradation rate constants suffers from two<br />

main sources <strong>of</strong> error, i.e., sampling <strong>of</strong>f the plume center<br />

line <strong>and</strong> an incorrect estimate <strong>of</strong> the average transport<br />

velocity. The first can be overcome by using method 2,<br />

the second can be resolved by conducting tracer tests or<br />

additional measurements <strong>of</strong> the hydraulic conductivity. A<br />

tracer test would furthermore prove, that the observation<br />

wells under consideration are sampling the same <strong>flow</strong> path.<br />

Further work on this subject will include the effects <strong>of</strong><br />

measurement error on the estimated degradation rates, both<br />

in measuring hydraulic head <strong>and</strong> contaminant concentration.<br />

Also effects <strong>of</strong> different formulations <strong>of</strong> the kinetic reactions<br />

used to simulate the plume will be investigated.<br />

[40] Acknowledgments. This work is funded by the German Ministry<br />

<strong>of</strong> Education <strong>and</strong> Research as part <strong>of</strong> the KORA priority program,<br />

subproject 7.1 Virtual Aquifer. We would like to acknowledge the thoughtful<br />

reviews <strong>of</strong> three anonymous reviewers. Their comments have greatly<br />

improved this manuscript.<br />

References<br />

Ababou, R., D. McLaughlin, A. L. Gutjahr, <strong>and</strong> A. F. B. Tompson (1989),<br />

Numerical simulation <strong>of</strong> three-dimensional <strong>saturated</strong> <strong>flow</strong> in r<strong>and</strong>omly<br />

heterogeneous porous media, Transp. Porous Media, 4, 549–565.<br />

Aronson, D., <strong>and</strong> P. H. Howard (1997), Anaerobic biodegradation <strong>of</strong> organic<br />

chemicals in groundwater: A summary <strong>of</strong> field <strong>and</strong> laboratory<br />

studies, SRC TR-97-0223F, Sci. Cent. Rep., Syracuse Res. Corp., New<br />

York.<br />

Bauer, S., <strong>and</strong> O. Kolditz (2006), Assessing contaminant mass <strong>flow</strong> rates<br />

obtained by the integral groundwater investigation method by using the<br />

virtual aquifer approach, IAHS Publ., in press.<br />

Bauer, S., C. Beyer, <strong>and</strong> O. Kolditz (2005), Assessing measurements <strong>of</strong> first<br />

order degradation rates by using the Virtual Aquifer approach, IAHS<br />

Publ., 297, 274–281.<br />

Bear, J. (1972), Dynamics <strong>of</strong> fluids in Porous Media, Elsevier, New York.<br />

Beinhorn, M., O. Kolditz, <strong>and</strong> P. Dietrich (2005), 3-D <strong>numerical</strong> evaluation<br />

<strong>of</strong> density effects on tracer tests, J. Contam. Hydrol., 81, 89–105,<br />

doi:10.1016/j.jconhyd.2005.08.001.<br />

Bockelmann, A., D. Zamfirescu, T. Ptak, P. Grathwohl, <strong>and</strong> G. Teutsch<br />

(2003), Quantification <strong>of</strong> mass fluxes <strong>and</strong> natural attenuation rates<br />

at an industrial site with a limited monitoring network: A case study,<br />

J. Contam. Hydrol., 60, 97–121, doi:10.1016/S0169-7722(02)00060-8.<br />

Buscheck, T. E., <strong>and</strong> C. M. Alcantar (1995), Regression techniques <strong>and</strong><br />

analytical solutions to demonstrate intrinsic bioremediation, in Intrinsic<br />

Bioremediation, edited by R. E. Hinchee, T. J. Wilson, <strong>and</strong> D. Downey,<br />

pp. 109–116, Battelle Press, Columbus, Ohio.<br />

Chapelle, F. H., P. M. Bradley, D. R. Lovley, <strong>and</strong> D. A. Vroblesky (1996),<br />

Measuring rates <strong>of</strong> biodegradation in a contaminated aquifer using field<br />

<strong>and</strong> laboratory methods, Ground Water, 34(4), 691–698.<br />

13 <strong>of</strong> 14<br />

W01420<br />

Dagan, G. (1989), Flow <strong>and</strong> Transport in Porous Formations, Springer,<br />

New York.<br />

Diersch, H.-J., <strong>and</strong> O. Kolditz (1998), Coupled groundwater <strong>flow</strong> <strong>and</strong><br />

transport: 2. Thermohaline <strong>and</strong> 3-D convection processes, Adv. Water<br />

Resour., 21(5), 401–425.<br />

Diersch, H.-J., <strong>and</strong> O. Kolditz (2002), Variable-density <strong>flow</strong> <strong>and</strong> transport<br />

in porous media: Approaches <strong>and</strong> challenges, Adv. Water Resour., 25(8–<br />

12), 899–944.<br />

Domenico, P. A. (1987), An analytical model for multidimensional transport<br />

<strong>of</strong> a decaying contaminant species, J. Hydrol., 91, 49–59.<br />

Dykaar, B. B., <strong>and</strong> P. K. Kitanidis (1992), Determination <strong>of</strong> the effective<br />

hydraulic conductivity for heterogeneous porous media using a <strong>numerical</strong><br />

spectral approach: 2. Results, Water Resour. Res., 28(4), 1167–<br />

1178.<br />

Goovaerts, P. (1999), Impact <strong>of</strong> the simulation algorithm, magnitude <strong>of</strong><br />

ergodic fluctuations <strong>and</strong> number <strong>of</strong> realizations on the spaces <strong>of</strong> uncertainty<br />

<strong>of</strong> <strong>flow</strong> properties, Stochastic Environ. Res. Risk Assess., 13(3),<br />

161–182, doi:10.1007/s004770050037.<br />

Herfort, M. (2000), Reactive transport <strong>of</strong> organic compounds within a<br />

heterogeneous porous aquifer, Tübinger Geowiss. Arbeiten, Univ. <strong>of</strong><br />

Tübingen, Tübingen.<br />

Huyakorn, P. S., <strong>and</strong> G. F. Pinder (1983), Computational Methods in Subsurface<br />

Flow, Academic, San Diego, Calif.<br />

Kolditz, O. (2002), Computational Methods in Environmental Fluid Dynamics,<br />

Springer, New York.<br />

Kolditz, O., <strong>and</strong> S. Bauer (2004), A process-oriented approach to computing<br />

multi-field problems in porous media, J. Hydroinf., 6(3), 225–<br />

244.<br />

Kolditz, O., R. Ratke, H.-J. Diersch, <strong>and</strong> W. Zielke (1998), Coupled<br />

groundwater <strong>flow</strong> <strong>and</strong> transport: 1. Verification <strong>of</strong> variable density <strong>flow</strong><br />

<strong>and</strong> transport models, Adv. Water Resour., 21(1), 27–46.<br />

Kolditz, O., et al. (2004), GeoSys—Theory <strong>and</strong> users Manual, Release 4.1.<br />

GeoSystems Res., Cent. for Appl. Geosci., Univ. <strong>of</strong> Tübingen, Tübingen.<br />

(Available at http://www.uni-tuebingen.de/zag/geohydrology.)<br />

McNab, W. W., <strong>and</strong> B. P. Dooher (1998), A critique <strong>of</strong> a steady-state<br />

analytical method for estimating contaminant degradation rates, Ground<br />

Water, 36(6), 983–987.<br />

Newell, C. J., H. S. Rifai, J. T. Wilson, J. A. Connor, J. A. Aziz, <strong>and</strong> M. P.<br />

Suarez (2002), Calculation <strong>and</strong> use <strong>of</strong> first-order rate constants for monitored<br />

natural attenuation studies, U.S. EPA Ground Water Issue, EPA/<br />

540/S-02/500, U.S. Environ. Protect. Ag., Washington, D. C.<br />

Pebesma, E. J., <strong>and</strong> C. G. Wesseling (1998), Gstat: A program for geostatistical<br />

<strong>modeling</strong>, prediction <strong>and</strong> simulation, Comput. Geosci., 24(1),<br />

17–31, doi:10.1016/S0098-3004(97)00082-4.<br />

Rehfeldt, K. R., J. M. Boggs, <strong>and</strong> L. W. Gelhar (1992), Field study <strong>of</strong><br />

dispersion in a heterogeneous aquifer: 3. Geostatistical analysis <strong>of</strong> hydraulic<br />

conductivity, Water Resour. Res., 28(12), 3309–3324.<br />

Renard, P., <strong>and</strong> G. de Marsily (1997), Calculating equivalent permeability:<br />

A review, Adv. Water Resour., 20(5–6), 253–278.<br />

Rubin, Y. (2003), <strong>Applied</strong> Stochastic Hydrogeology, Oxford Univ. Press,<br />

New York.<br />

Schäfer, D., A. Dahmke, O. Kolditz, <strong>and</strong> G. Teutsch (2002), ‘‘Virtual<br />

Aquifers’’: A concept for evaluation <strong>of</strong> exploration, remediation <strong>and</strong><br />

monitoring strategies, IAHS Publ., 277, 52–59.<br />

Schäfer, D., B. Schlenz, <strong>and</strong> A. Dahmke (2004), Evaluation <strong>of</strong> exploration<br />

<strong>and</strong> monitoring methods for verification <strong>of</strong> natural attenuation using the<br />

virtual aquifer approach, Biodegradation J., 15(6), 453–465,<br />

doi:10.1023/b:biod.0000044600.81216.00.<br />

Stenback, G. A., S. K. Ong, S. W. Rogers, <strong>and</strong> B. H. Kjartonson (2004),<br />

Impact <strong>of</strong> transverse <strong>and</strong> longitudinal dispersion on first-order degradation<br />

rate constant estimation, J. Contam. Hydrol., 73, 3 – 14, doi:10.1016/<br />

j.jconhyd.2003.11.004.<br />

Suarez, M. P., <strong>and</strong> H. S. Rifai (2002), Evaluation <strong>of</strong> BTEX remediation by<br />

natural attenuation at a coastal facility, Ground Water Monit. Remed.,<br />

22(1), 62–77.<br />

Sudicky, E. A. (1986), A natural gradient experiment on solute transport<br />

in a s<strong>and</strong> aquifer: Spatial variability <strong>of</strong> hydraulic conductivity <strong>and</strong> its<br />

role in the dispersion process, Water Resour. Res., 22(13), 2069 –<br />

2082.<br />

Thorenz, C., G. Kosakowski, O. Kolditz, <strong>and</strong> B. Berkowitz (2002), An<br />

experimental <strong>and</strong> <strong>numerical</strong> investigation <strong>of</strong> saltwater movement in<br />

coupled <strong>saturated</strong>-partially <strong>saturated</strong> systems, Water Resour. Res.,<br />

38(6), 1069, doi:10.1029/2001WR000364.<br />

U.S. Environmental Protection Agency (EPA) (1998), Technical protocol<br />

for evaluating natural attenuation <strong>of</strong> chlorinated solvents in groundwater,<br />

Rep. EPA/600/R/128, Washington, D. C.


W01420 BAUER ET AL.: ASSESSING FIRST-ORDER RATES W01420<br />

U.S. Environmental Protection Agency (EPA) (1999), Use <strong>of</strong> monitoring<br />

natural attenuation at Superfund, RCRA Corrective Action, <strong>and</strong> Underground<br />

Storage Tank Sites, Office <strong>of</strong> Solid Waste <strong>and</strong> Emergency Response<br />

Directive 9200.4-17, Washington, D. C.<br />

Wiedemeier, T. H., M. A. Swanson, J. T. Wilson, D. H. Kampbell, R. N.<br />

Miller, <strong>and</strong> J. E. Hansen (1996), Approximation <strong>of</strong> biodegradation rate<br />

constants for monoaromatic hydrocarbons (BTEX) in ground water,<br />

Ground Water Monit. Remed., 16(3), 186–194.<br />

Wiedemeier, T. H., H. S. Rifai, T. J. Wilson, <strong>and</strong> C. Newell (1999), Natural<br />

Attenuation <strong>of</strong> Fuels <strong>and</strong> Chlorinated Solvents in the Subsurface, John<br />

Wiley, Hoboken, N. J.<br />

Wilson, R. D., S. F. Thornton, <strong>and</strong> D. M. Mackay (2004), Challenges<br />

in monitoring the natural attenuation <strong>of</strong> spatially variable plumes, Biodegradation<br />

J., 15(6), 459–469, doi:10.1023/b:biod.0000044591. 45542.a9.<br />

Zamfirescu, D., <strong>and</strong> P. Grathwohl (2001), Occurrence <strong>and</strong> attenuation <strong>of</strong><br />

specific organic compounds in the groundwater plume at a former<br />

gasworks site, J. Contam. Hydrol., 53, 407–427, doi:10.1016/S0169-<br />

7722(01)00176-0.<br />

14 <strong>of</strong> 14<br />

Zhang, D., <strong>and</strong> S. P. Neuman (1996), Effect <strong>of</strong> local dispersion on solute<br />

transport in r<strong>and</strong>omly heterogeneous media, Water Resour. Res., 32,<br />

2715–2723.<br />

Zhang, Y.-K. (2003), Non-ergodic solute transport in physically <strong>and</strong><br />

chemically heterogeneous porous media, Water Resour. Res., 39(7),<br />

1197, doi:10.1029/2003WR002116.<br />

Zhang, Y.-K., <strong>and</strong> R. C. Heathcote (2003), An improved method for estimation<br />

<strong>of</strong> biodegradation rate with field data, Ground Water Monit.<br />

Remed., 23(3), 112–116.<br />

S. Bauer, C. Beyer, <strong>and</strong> O. Kolditz, Center for <strong>Applied</strong> Geoscience,<br />

University <strong>of</strong> Tübingen, Sigwartstr. 10, D-72076 Tübingen, Germany.<br />

(sebastian.bauer@uni-tuebingen.de; christ<strong>of</strong>.beyer@uni-tuebingen.de;<br />

kolditz@uni-tuebingen.de)


Enclosed Publication 3<br />

Beyer, C., Bauer, S., Kolditz, O. (2006): Uncertainty Assessment <strong>of</strong> Contaminant Plume<br />

Length Estimates in Heterogeneous Aquifers. J. Contam. Hydrol., 87, 73-95, doi:<br />

10.1016/j.jconhyd.2006.04.006.<br />

Reprinted from Journal <strong>of</strong> Contaminant Hydrology, 87, Beyer, Christ<strong>of</strong>, Bauer, Sebastian <strong>and</strong><br />

Olaf Kolditz, Uncertainty Assessment <strong>of</strong> Contaminant Plume Length Estimates in<br />

Heterogeneous Aquifers, 73-95, Copyright 2006, with permission from Elsevier.<br />

The enclosed article can be obtained online via ScienceDirect at<br />

http://www.sciencedirect.com/science/journal/01697722.


Uncertainty assessment <strong>of</strong> contaminant plume length<br />

estimates in heterogeneous aquifers<br />

Abstract<br />

Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

Christ<strong>of</strong> Beyer ⁎ , Sebastian Bauer, Olaf Kolditz<br />

www.elsevier.com/locate/jconhyd<br />

Center for <strong>Applied</strong> Geoscience, University <strong>of</strong> Tübingen, Sigwartstraße 10, D 72076 Tübingen, Germany<br />

Received 4 November 2005; received in revised form 14 April 2006; accepted 25 April 2006<br />

Available online 16 June 2006<br />

The Virtual Aquifer approach is used in this study to assess the uncertainty involved in the estimation <strong>of</strong><br />

contaminant plume lengths in heterogeneous aquifers. Contaminant plumes in heterogeneous twodimensional<br />

conductivity fields <strong>and</strong> subject to first order <strong>and</strong> Michaelis–Menten (MM) degradation kinetics<br />

are investigated by the center line method. First order degradation rates <strong>and</strong> plume lengths are estimated<br />

from point information obtained along the plume center line. Results from a Monte-Carlo investigation<br />

show that the estimated rate constant is highly uncertain <strong>and</strong> biased towards overly high values. Uncertainty<br />

<strong>and</strong> bias amplify with increasing heterogeneity up to maximum values <strong>of</strong> one order <strong>of</strong> magnitude.<br />

Calculated plume lengths reflect this uncertainty <strong>and</strong> bias. On average, plume lengths are estimated to about<br />

50% <strong>of</strong> the true plume length. When plumes subject to MM degradation kinetics are investigated by using a<br />

first order rate law, an additional error is introduced <strong>and</strong> uncertainty as well as bias increase, causing plume<br />

length estimates to be less than 40% <strong>of</strong> the true length. For plumes with MM degradation kinetics,<br />

therefore, a regression approach is used which allows the determination <strong>of</strong> the MM parameters from center<br />

line data. Rate parameters are overestimated by a factor <strong>of</strong> two on average, while plume length estimates are<br />

about 80% <strong>of</strong> the true length. Plume lengths calculated using the MM parameters are thus closer to the<br />

correct length, as compared to the first order approximation. This approach is therefore recommended if<br />

field data collected along the center line <strong>of</strong> a plume give evidence <strong>of</strong> MM kinetics.<br />

© 2006 Elsevier B.V. All rights reserved.<br />

Keywords: Natural Attenuation; Heterogeneity; Plume length; Center line method; Uncertainty analysis; Monte-Carlo;<br />

Numerical modelling<br />

⁎ Corresponding author. Tel.: +49 7071 29 73176; fax: +49 7071 5059.<br />

E-mail address: christ<strong>of</strong>.beyer@uni-tuebingen.de (C. Beyer).<br />

0169-7722/$ - see front matter © 2006 Elsevier B.V. All rights reserved.<br />

doi:10.1016/j.jconhyd.2006.04.006


74 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

1. Introduction<br />

One major requirement for the implementation <strong>of</strong> natural attenuation (NA) as a remedial <strong>and</strong> risk<br />

reduction strategy for contaminated aquifers is an assessment <strong>of</strong> the dimensions <strong>of</strong> contaminant<br />

plumes <strong>and</strong> to predict their fate. Down gradient contaminant concentrations, i.e. plume lengths, must<br />

be calculated or estimated to identify potential receptors <strong>and</strong> predict exposure levels. For this purpose<br />

analytical <strong>and</strong> <strong>numerical</strong> solute transport models (e.g. Bioscreen (Newell et al., 1996), Biochlor<br />

(Aziz et al., 2000), Bioplume III (Rifai et al., 1998)) are routinely used. The rate <strong>of</strong> contaminant<br />

removal through biodegradation is a key parameter, as concentrations <strong>and</strong> the modeled plume<br />

lengths are highly sensitive to the degradation rate (McNab, 2001; Suarez <strong>and</strong> Rifai, 2004). Although<br />

very detailed mathematical descriptions <strong>of</strong> contaminant degradation in the subsurface are available<br />

(Baveye <strong>and</strong> Valocchi, 1989; Rittmann <strong>and</strong> VanBriesen, 1996; Wiedemeier et al., 1999; Islam et al.,<br />

2001), for applications in the field, usually simplified approaches are used because the identification<br />

<strong>of</strong> a large number <strong>of</strong> parameters <strong>and</strong> processes from field data is <strong>of</strong>ten impossible. Due to its<br />

mathematical simplicity, its ease <strong>of</strong> implementation into transport models <strong>and</strong> the necessity <strong>of</strong><br />

determining only a single parameter, the most frequently used degradation model is first order<br />

kinetics (Wiedemeier et al., 1999). Field methods for the determination <strong>of</strong> biodegradation rates in<br />

ground water include mass balance calculations, in-situ microcosm studies <strong>and</strong> the center line<br />

method (Chapelle et al., 1996; Wiedemeier et al., 1999). The latter is frequently used for plume<br />

monitoring <strong>and</strong> degradation rate evaluation (e.g. Chapelle et al., 1996; Wiedemeier et al., 1996;<br />

Zamfirescu <strong>and</strong> Grathwohl, 2001; Suarez <strong>and</strong> Rifai, 2002; Wilson <strong>and</strong> Kolhatkar, 2002; Bockelmann<br />

et al., 2003), <strong>and</strong> is based on contaminant concentrations measured in observation wells<br />

installed along the presumed center line <strong>of</strong> a plume. This approach, however, is only applicable for<br />

plumes that have reached a (quasi-) steady state, i.e. the plume is neither shrinking nor exp<strong>and</strong>ing <strong>and</strong><br />

the measured concentrations do not change with time. The concentration-distance relations thus<br />

obtained for a steady state plume can be used to estimate the first order rate constant λ. This<br />

parameter can then be used with an appropriate transport model to estimate the contaminant<br />

distribution in the aquifer. However, as the spatial variability <strong>of</strong> aquifer properties has a substantial<br />

influence on the distribution <strong>of</strong> contaminants <strong>and</strong> plume development, also the results <strong>of</strong> such an<br />

assessment are affected. Wilson et al. (2004) point out that the approach is prone to errors because the<br />

center line <strong>of</strong> a plume may be missed by monitoring wells if the inferred ground water <strong>flow</strong> direction<br />

is incorrect or the contaminant plume me<strong>and</strong>ers in all three dimensions due to macro-scale<br />

heterogeneities. McNab <strong>and</strong> Dooher (1998) demonstrated that, even in a homogeneous aquifer,<br />

transverse dispersion can produce center line concentration pr<strong>of</strong>iles <strong>of</strong> recalcitrant compounds that<br />

exhibit characteristics consistent with first order degradation; this can easily lead to misinterpretation<br />

<strong>of</strong> the monitoring data. The result <strong>of</strong> these complicating factors is that the degradation potential may<br />

be severely overestimated, causing underestimation <strong>of</strong> plume length or contaminant mass <strong>and</strong> an<br />

over optimistic prognosis <strong>of</strong> down gradient concentrations <strong>and</strong> exposure levels. Moreover, it is well<br />

known that the use <strong>of</strong> first order kinetics may be problematic in some situations, as it is a poor<br />

representation <strong>of</strong> the processes occurring in contaminated aquifers. Usage <strong>of</strong> a first order model<br />

outside its range <strong>of</strong> validity may result either in significant under- or overestimation <strong>of</strong> the<br />

attenuation potential at a site (Bekins et al., 1998). In a <strong>numerical</strong> experiment, Schäfer et al. (2004a)<br />

demonstrated that for specific points in time, first order kinetics may be able to approximately<br />

reproduce mass <strong>and</strong> dimensions <strong>of</strong> contaminant plumes that follow from a far more complex<br />

degradation model. For a long term prognosis, however, the first order approximation proved<br />

inappropriate, resulting in an underestimation <strong>of</strong> plume length <strong>and</strong> contaminant mass. Recently,<br />

Bauer et al. (2006) performed a sensitivity study on the influences <strong>of</strong> aquifer heterogeneity on first


order degradation rate constants estimated from using the center line method. This study demonstrated<br />

that aquifer heterogeneity introduces significant uncertainty in the estimated rate constants<br />

<strong>and</strong> may cause a severe overestimation <strong>of</strong> the degradation potential.<br />

Since the determination <strong>of</strong> degradation rates is usually only an intermediate step for the<br />

characterization <strong>of</strong> contaminated sites, the present paper takes the approach <strong>of</strong> Bauer et al. (2006) one<br />

step further. Here, the uncertainty involved in the estimation <strong>of</strong> contaminant plume lengths in<br />

heterogeneous aquifers is evaluated using the Virtual Aquifer concept. Three different scenarios are<br />

studied in detail. In case A, synthetic contaminant plumes following first order degradation kinetics<br />

are investigated. First order rate constants are estimated by methods typically used in field<br />

applications. The rate constants are then used to calculate the corresponding contaminant plume<br />

lengths with analytical transport models. As the first order degradation model results in theoretically<br />

infinite plumes, a relative concentration contour line is defined as the plume length here. Results are<br />

analysed with regard to errors <strong>and</strong> uncertainty in the rate constants <strong>and</strong> their propagation to the plume<br />

length estimates. In case B, the additional error is studied that arises when a first order approximation<br />

is used although degradation kinetics deviate from a first order rate law. Here the attenuation<br />

potential for plumes following Michaelis–Menten (MM) degradation kinetics is assessed using the<br />

same first order methods employed in case A. In case C, a regression approach is used to estimate the<br />

MM parameters <strong>and</strong> plume lengths for the plumes with MM degradation kinetics. Results are<br />

compared to cases A <strong>and</strong> B to allow conclusions about potentials <strong>and</strong> limitations <strong>of</strong> this approach.<br />

2. Virtual Aquifer concept<br />

C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

Due to the limited accessibility <strong>of</strong> the subsurface, measurements <strong>of</strong> piezometric heads <strong>and</strong><br />

pollutant concentrations at contaminated sites are sparse <strong>and</strong> may not be representative <strong>of</strong> the<br />

heterogeneous hydrogeologic conditions. Any site investigation is thus subject to uncertainty,<br />

reflecting the limited knowledge <strong>of</strong> the aquifer properties <strong>and</strong> the extent <strong>of</strong> the contamination. Due<br />

to this uncertainty, field investigation methods for plume screening <strong>and</strong> measuring <strong>of</strong> hydraulic<br />

conductivity or degradation rates can neither be tested nor verified in the field. One appropriate<br />

method <strong>of</strong> assessing the performance <strong>and</strong> reliability <strong>of</strong> field investigation methods is by studying<br />

them in heterogeneous synthetic (virtual) aquifers. With this approach the results <strong>of</strong> a particular<br />

method can be compared to the “true” values, as these values are known from the synthetic aquifer.<br />

The “Virtual Aquifer” concept is a combination <strong>of</strong> different methodologies, tools <strong>and</strong> techniques,<br />

particularly aimed at this type <strong>of</strong> problem. Its two key components are (1) a flexible <strong>and</strong> efficient<br />

modelling system, allowing the <strong>numerical</strong> simulation <strong>of</strong> reactive multi-component transport in the<br />

subsurface, <strong>and</strong> (2) an extensive database, containing statistical information, physical <strong>and</strong> (bio-)<br />

geochemical data from a large number <strong>of</strong> well investigated sites <strong>and</strong> aquifers. Moreover, the concept<br />

comprises a collection <strong>of</strong> analytical <strong>and</strong> <strong>numerical</strong> methods, that are commonly used for the<br />

investigation <strong>of</strong> contaminated sites <strong>and</strong> aquifers or the interpretation <strong>of</strong> measured data. The synthesis<br />

<strong>of</strong> both, the database <strong>and</strong> the simulation system allows a proper definition <strong>and</strong> computer based<br />

evaluation <strong>of</strong> scenarios <strong>and</strong> case studies, focussed on investigation strategies, redevelopment <strong>and</strong><br />

monitoring at contaminated sites.<br />

As a first step for such an analysis, synthetic aquifer models are generated based on the statistical<br />

properties <strong>of</strong> real aquifers. Thus, to a certain degree, these aquifer models represent realistic analogues<br />

<strong>of</strong> existing sites. A defined source <strong>of</strong> contamination is then introduced <strong>and</strong> a reactive transport model<br />

is used to simulate the evolution <strong>of</strong> the plume, resulting in realistic concentration distributions in the<br />

synthetic aquifer. In comparison to the “real world”, the unique advantage <strong>of</strong> the synthetic aquifer is<br />

that the spatial distribution <strong>of</strong> all physical <strong>and</strong> geochemical properties <strong>and</strong> parameters as well as the<br />

75


76 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

contaminant concentrations are exactly known. In the second step, the synthetic aquifer is investigated<br />

by st<strong>and</strong>ard monitoring <strong>and</strong> investigation techniques. Although the parameter distribution <strong>of</strong> the<br />

synthetic aquifer is known a priori, only the data “measured” at the observation wells (i.e. hydraulic<br />

heads <strong>and</strong> concentrations) are used <strong>and</strong> interpreted. This is done because in a real site investigation<br />

also only a limited amount <strong>of</strong> measured data would be available, with the amount <strong>of</strong> information<br />

depending on investigation intensity <strong>and</strong> project finances. In the third step, the investigation results are<br />

compared to the “true” values, allowing an evaluation <strong>of</strong> the accuracy <strong>of</strong> the investigation method. In<br />

addition, the use <strong>of</strong> synthetic aquifers <strong>of</strong>fers the possibility to assess the influence <strong>of</strong> different<br />

parameters, such that sources <strong>of</strong> uncertainty <strong>and</strong> error for the investigation method can be considered<br />

individually <strong>and</strong> the sensitivity <strong>of</strong> investigation results on these can be studied. Stochastic simulation<br />

techniques like the Monte-Carlo method are applied to study the propagation <strong>of</strong> parameter variability<br />

<strong>and</strong> uncertainty into the investigation results. Due to the ability to perform extensive <strong>and</strong> detailed<br />

scenario analysis <strong>and</strong> visualization, this approach is well suited to the exploration <strong>of</strong> the uncertainty<br />

involved in hydrogeologic investigation <strong>and</strong> management. The methodology has been applied under<br />

the term “Virtual Aquifer” by Schäfer et al. (2002, 2004b), Bauer <strong>and</strong> Kolditz (2005) <strong>and</strong> Bauer et al.<br />

(2005, 2006).<br />

3. Scenario definition<br />

In this study, the Virtual Aquifer concept is used in a Monte-Carlo framework to assess the<br />

influence <strong>of</strong> spatially heterogeneous hydraulic conductivities on the estimation <strong>of</strong> degradation<br />

rates <strong>and</strong> contaminant plume lengths. Multiple plume realizations <strong>of</strong> contaminants degrading<br />

according to a first order degradation kinetics or Michaelis–Menten kinetics in aquifers with<br />

different degrees <strong>of</strong> heterogeneity are investigated using the center line approach. By comparison<br />

<strong>of</strong> the estimated degradation rates <strong>and</strong> plume lengths with the respective virtual reality data the<br />

investigation methods are tested <strong>and</strong> evaluated. Three different cases are studied in detail:<br />

In case A, four different st<strong>and</strong>ard methods for the determination <strong>of</strong> the first order rate constant λ<br />

are applied to concentration vs. distance data obtained from investigation <strong>of</strong> synthetic contaminant<br />

plumes following first order degradation kinetics. Accordingly, four different rate constants are<br />

estimated for each plume realization. For each λ the length <strong>of</strong> the contaminant plume is estimated<br />

using an analytical transport model. The four methods <strong>and</strong> the corresponding analytical transport<br />

models are introduced in Sections 4.1 <strong>and</strong> 4.2. The main objectives <strong>of</strong> case A are to test the<br />

applicability <strong>and</strong> performance <strong>of</strong> the four different methods <strong>of</strong> determining the first order degradation<br />

rate constant in heterogeneous aquifers <strong>and</strong> to analyse the propagation <strong>of</strong> errors <strong>and</strong><br />

uncertainty from the rate constant to the plume length estimate.<br />

In case B, the same four methods are evaluated with regard to their ability to approximate the<br />

degradation potential <strong>and</strong> estimate the plume length, when the true degradation kinetics deviate<br />

from first order. Here plumes following MM degradation kinetics are investigated in an analogous<br />

manner to case A. The additional error that arises from the first order approximation is studied. The<br />

motivation behind this scenario is that although it is well known that contaminant degradation in<br />

natural aquifers may follow far more complicated processes <strong>and</strong> kinetic laws than a simple first<br />

order model, the latter is routinely used at many field sites. Therefore, this scenario highlights some<br />

<strong>of</strong> the problems that result from this discrepance.<br />

In case C, a regression approach is studied which allows the estimation <strong>of</strong> MM kinetic parameters<br />

from plume center line data. This method is developed in Section 4.1 <strong>and</strong> tested under the influence<br />

<strong>of</strong> aquifer heterogeneity in case C. Here the plumes following MM degradation kinetics are<br />

investigated. As for cases A <strong>and</strong> B the propagation <strong>of</strong> errors <strong>and</strong> uncertainty from the estimated MM


Table 1<br />

Overview <strong>of</strong> scenarios studied<br />

Case Contaminant plume<br />

following<br />

parameters to the plume lengths is analysed. Results are set in relation to cases A <strong>and</strong> B to discuss<br />

advantages <strong>and</strong> disadvantages, potentials <strong>and</strong> limitations <strong>of</strong> the MM parameter estimation approach.<br />

An overview <strong>of</strong> the three different cases is given in Table 1.<br />

4. Methods<br />

C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

Degradation rate<br />

determined with<br />

4.1. Estimation <strong>of</strong> degradation rate constants<br />

Plume length<br />

determined with<br />

A First order kinetics First order kinetics (Eqs. (1)–(4)) First order kinetics (Eqs. (7)–(9)) 6.1<br />

B Michaelis–Menten<br />

kinetics<br />

First order kinetics (Eqs. (1)–(4)) First order kinetics (Eqs. (7)–(9)) 6.2<br />

C Michaelis–Menten<br />

kinetics<br />

Michaelis–Menten kinetics (Eq. (5)) Michaelis–Menten kinetics (Eq. (10)) 6.3<br />

Results in<br />

section<br />

Four different st<strong>and</strong>ard methods for the determination <strong>of</strong> the first order degradation rate<br />

constant λ (Table 2, Eqs. (1)–(4)) are compared in this study. Each <strong>of</strong> the four methods requires<br />

concentration-distance relations obtained by measuring contaminant concentrations in several<br />

observation wells along the center line <strong>of</strong> a steady state plume. The degradation rate constant λ is<br />

estimated by fitting a linear function to the logarithms <strong>of</strong> concentration vs. distance from the source<br />

by linear regression. Methods 1–4 are introduced only briefly here, more details are given in Bauer<br />

et al. (2006). Furthermore, a regression approach that allows the estimation <strong>of</strong> MM kinetics<br />

parameters from center line data is developed <strong>and</strong> tested as method 5 (Table 2, equation (5)).<br />

Method 1 (Table 2, equation (1)) is based on the one dimensional transport equation, considering<br />

advection <strong>and</strong> first order degradation only. Rate constants determined with method 1 can<br />

be considered rather an overall or bulk attenuation rate than a degradation rate constant (Newell et<br />

al., 2002) as all concentration changes that result from processes other than degradation, such as<br />

diffusion, dispersion, volatilization <strong>and</strong> dilution, are attributed to the degradation process. Method<br />

2(Table 2, equation (2); Buscheck <strong>and</strong> Alcantar, 1995) is based on the steady state solution to the<br />

one dimensional transport equation considering advection, longitudinal dispersion <strong>and</strong> first order<br />

degradation. Method 2 thus requires an estimate <strong>of</strong> longitudinal dispersivity αL [m]. Method 3<br />

(Table 2, equation (3)) was proposed by Zhang <strong>and</strong> Heathcote (2003) <strong>and</strong> represents a twodimensional<br />

modification <strong>of</strong> method 2. A correction term derived from the analytical solution <strong>of</strong><br />

the two-dimensional transport equation including first order decay (Domenico, 1987) is used to<br />

account for lateral spreading <strong>and</strong> the width <strong>of</strong> the source zone. A similar approach, which also<br />

allows the inclusion <strong>of</strong> measurements <strong>of</strong>f the plume center line was presented by Stenback et al.<br />

(2004). Method 4 (Table 2, equation (4); Wilson et al., 1994; Wiedemeier et al., 1996) is based on<br />

the same transport equation as method 1. However, measured concentrations <strong>of</strong> the reactive<br />

contaminant are scaled by up <strong>and</strong> down gradient concentration C0 ⁎ <strong>and</strong> C(x)⁎ [M L − 3 ] <strong>of</strong> a nondegrading<br />

conservative solute spreading from the same source. Since dispersion <strong>and</strong> measuring <strong>of</strong>f<br />

the center line also affect the measured concentrations <strong>of</strong> non-reactive solutes, this procedure<br />

allows a correction for both effects. Method 5 (Table 2, equation (5)) is valid for contaminant<br />

plumes following Michaelis–Menten (MM) degradation kinetics (Beyer et al., 2005). When the<br />

77


78 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

Table 2<br />

Estimation <strong>of</strong> first order degradation rate constants (methods 1–4) <strong>and</strong> Michaelis–Menten degradation kinetics parameters (method 5) from center line data<br />

Formula Description Parameter<br />

estimated<br />

Method<br />

(equation)<br />

λ<br />

1D transport equation with advection <strong>and</strong><br />

first order degradation; advective velocity,<br />

va, source concentration C0, down gradient<br />

concentration, C(x), first order degradation<br />

rate constant, λ, distance, x<br />

(1) k1 ¼ − va CðxÞ<br />

ln<br />

Dx C0<br />

!<br />

2<br />

λ<br />

1D transport equation with advection,<br />

longitudinal dispersion <strong>and</strong> first order<br />

degradation; longitudinal dispersivity, αL<br />

−1<br />

lnðCðxÞ=C0Þ<br />

Dx<br />

1−2aL<br />

(2) k2 ¼ va<br />

4aL<br />

λ<br />

!<br />

2D transport equation with advection,<br />

longitudinal <strong>and</strong> transverse dispersion,<br />

source width <strong>and</strong> first order degradation;<br />

transverse dispersivity, αT, source area<br />

width, WS 2<br />

−1<br />

lnðCðxÞ=ðC0bÞÞ Dx<br />

1−2aL<br />

k3 ¼ va<br />

4aL<br />

(3)<br />

4 ffiffiffiffiffiffiffiffiffiffi p<br />

aTDx<br />

C * 0<br />

CðxÞ *<br />

!<br />

WS<br />

with b ¼ erf<br />

λ<br />

Same as method 1; contaminant<br />

concentrations normalized with<br />

regard to conservative solute<br />

concentrations, C0⁎, C(x)⁎<br />

CðxÞ<br />

ln<br />

C0<br />

(4) k4 ¼ − va<br />

Dx<br />

k max, M C<br />

1D transport equation with advection <strong>and</strong><br />

Michaelis–Menten degradation kinetics;<br />

maximum degradation rate, kmax,<br />

half saturation concentration, MC<br />

1<br />

þ<br />

kmax<br />

lnðC0=CðxÞÞ C0−CðxÞ<br />

MC<br />

¼<br />

kmax<br />

Dx<br />

ð Þ<br />

(5)<br />

va C0−CðxÞ


Table 3<br />

Calculation <strong>of</strong> contaminant plume lengths for first order <strong>and</strong> Michaelis–Menten degradation kinetics<br />

Method Formula Equation<br />

1 L1 ¼ Dx ¼ − va<br />

k lnðCðxÞ=C0Þ (7)<br />

2<br />

lnðCðxÞ=C0Þ<br />

L2 ¼ Dx ¼ 2aL<br />

1−ð1 þ 4kaL=vaÞ 0:5<br />

( " sffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi#<br />

)<br />

(8)<br />

3, 4 L3;4 ¼ Dxi; for CðxÞ<br />

5 LMM ¼ Dx ¼ va<br />

degradation <strong>of</strong> a contaminant is not limited by electron acceptor availability <strong>and</strong> the microbial<br />

density is assumed to be constant with time, Eq. (6) applies (Simkins <strong>and</strong> Alex<strong>and</strong>er, 1984):<br />

dC<br />

dt<br />

C<br />

¼ −kmax<br />

C þ MC<br />

C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

where kmax is the maximum degradation rate [M L − 3 T − 1 ] <strong>and</strong> MC is the MM half-saturation<br />

concentration [ML − 3 ]. This approximation may be applicable when aquifer sediments have been<br />

exposed to contaminants for several years (Bekins et al., 1998). The integral form <strong>of</strong> Eq. (6) can be<br />

rearranged to yield equation (5) <strong>of</strong> Table 2. According to Robinson (1985), equation (5) is the most<br />

reliable <strong>of</strong> several different formulations <strong>of</strong> the integrated MM model for estimation <strong>of</strong> kmax <strong>and</strong><br />

MC. Both parameters are estimated by a linear least squares fit <strong>of</strong> Δx/[va(C0−C(x))] vs. ln(C0/C<br />

(x))/(C0−C(x)). Thus kmax is obtained as the reciprocal <strong>of</strong> the intercept <strong>of</strong> the linear function <strong>and</strong><br />

MC as its slope multiplied by kmax. Application <strong>of</strong> method 5 assumes advective transport only (see<br />

also Parlange et al., 1984). As Robinson (1985) points out, application <strong>of</strong> a linearized integrated<br />

MM model for least squares estimation <strong>of</strong> its parameters may be problematic, because measured<br />

concentrations C(x) appear in the dependent as well as in the independent variable. A preliminary<br />

examination <strong>of</strong> several nonlinear least squares approaches for fitting the MM parameters to the<br />

concentration vs. distance data showed that the parameters determined by method 5 were on<br />

average more accurate than those obtained by other methods.<br />

4.2. Estimation <strong>of</strong> contaminant plume lengths<br />

C0<br />

kmax<br />

−exp Dxi<br />

2aL<br />

1−<br />

1 þ 4kaL<br />

va<br />

Given a first order degradation rate constant λ (by estimation with one <strong>of</strong> the four center line<br />

methods presented above), equations (1)–(3) (Table 2) can be rearranged to calculate the length <strong>of</strong><br />

the steady state contaminant plume (see Table 3). The plume length here is defined as the largest<br />

distance between the source <strong>and</strong> the concentration isoline for concentration CPL [ML − 3 ]. Equation<br />

(7) gives this distance for the purely advective case <strong>of</strong> method 1, equations (8) <strong>and</strong> (9) correspond<br />

with methods 2 <strong>and</strong> 3, respectively. As the rate constant estimated with method 4 is corrected for<br />

dispersion, this process has to be accounted for when the plume length is calculated. Therefore<br />

equation (9) is also used for calculation <strong>of</strong> plume lengths based on λ4. A steady state plume length<br />

can also be calculated using the MM parameters estimated using method 5. Rearrangement <strong>of</strong><br />

equation (5) <strong>of</strong> Table 2 yields equation (10) <strong>and</strong> gives the distance L MM at which concentrations fall<br />

erf<br />

WS<br />

4 ffiffiffiffiffiffiffiffiffiffiffi p ¼ 0 (9)<br />

aTDxi<br />

MCln C0<br />

CðxÞ þ C0−CðxÞ (10)<br />

79<br />

ð6Þ


80 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

below CPL. To define the plume length for this study, the 1% relative concentration contour line <strong>of</strong><br />

the contaminants is used, i.e. CPL=C(x)/C0=0.01.<br />

4.3. Numerical Monte-Carlo simulations<br />

To study the influence <strong>of</strong> heterogeneous hydraulic conductivity on the investigation results<br />

two-dimensional virtual aquifers are used. The model domain has dimensions <strong>of</strong> 184 m length <strong>and</strong><br />

64 m width (Fig. 1). A mean hydraulic gradient I <strong>of</strong> 0.053 is induced by fixed head boundary<br />

conditions on the left <strong>and</strong> the right h<strong>and</strong> side <strong>of</strong> the model domain. No <strong>flow</strong> boundary conditions<br />

are assigned to all other sides. Flow conditions are at steady state.<br />

The model domain is discretized with a grid density <strong>of</strong> 0.5 m in both directions. A contaminant<br />

source <strong>of</strong> 3 m×8 m, represented by a fixed concentration boundary condition, is centered at [11.5 m;<br />

32.0 m] down stream <strong>of</strong> the in<strong>flow</strong> boundary. The source emits two reactive contaminants <strong>and</strong> a<br />

conservative compound, each with a unit concentration <strong>of</strong> 1. The first reactive contaminant is<br />

degraded by first order kinetics with a rate constant λ=5.87·10 −7 s −1 . Degradation <strong>of</strong> the second<br />

reactive contaminant follows MM kinetics. MM parameters are taken from Bekins et al. (1998)<br />

(kmax=3.9·10 −9 g L −1 s −1 <strong>and</strong> MC=1.33·10 −3 g L). Using the source concentration <strong>of</strong> 2.68·10 −2 g<br />

L −1 given in Bekins et al. (1998) these parameters were scaled to a dimensionless source<br />

concentration <strong>of</strong> 1.0, as used here, yielding relative values (in normalized units) <strong>of</strong><br />

kmax=1.45·10 −7 s −1 <strong>and</strong> MC=4.97·10 −2 . Thus the first order <strong>and</strong> MM plume lengths for both<br />

compounds are equal in a two-dimensional homogeneous aquifer for CPL=C(x)/C0=0.01. Neither<br />

growth <strong>of</strong> microorganisms nor limitation or inhibition <strong>of</strong> degradation by other substances is<br />

considered here. All compounds are not retarded <strong>and</strong> show no volatilization. The conceptual model<br />

used in this study is a rigorous simplification <strong>of</strong> the processes observed in natural aquifer systems,<br />

where degradation follows more complicated laws <strong>and</strong> is spatially dependent. The model setup is thus<br />

designed to provide ideal conditions for the application <strong>of</strong> the center line methods to be studied. This<br />

is certainly not the case in nature, where the reaction kinetics will follow more complicated laws, may<br />

be spatially dependent, be steered by the availability <strong>of</strong> electron donors <strong>and</strong> acceptors, or additional<br />

influences from transient effects <strong>and</strong> dilution have to be accounted for. However, these simplifying<br />

assumptions are used here to be able to study the st<strong>and</strong>ard methods closely <strong>and</strong> evaluate individually<br />

the influence <strong>of</strong> heterogeneity <strong>of</strong> the hydraulic conductivity <strong>and</strong> the influence <strong>of</strong> degradation kinetics<br />

on the performance <strong>of</strong> the methods under otherwise ideal conditions. Case B (Table 1) is the case were<br />

we study the combination <strong>of</strong> errors stemming from hydraulics <strong>and</strong> from reaction kinetics.<br />

The hydraulic conductivity K <strong>of</strong> the virtual aquifers is regarded as a spatial r<strong>and</strong>om variable,<br />

following a lognormal distribution with an expected value <strong>of</strong> E[Y=ln(K)]=−9.54, which corresponds<br />

to an effective conductivity Kef <strong>of</strong> 7.19·10 −5 ms −1 using the geometric mean. An isotropic<br />

exponential covariance function with an integral scale lY<strong>of</strong> 2.67 m is used for the spatial correlation<br />

Fig. 1. Virtual Aquifer model domain <strong>and</strong> boundary conditions.


structure. Four different cases with increasing heterogeneity, i.e. ln(K) variances σY 2 <strong>of</strong> 0.38, 1.71, 2.7<br />

<strong>and</strong> 4.5, respectively, are considered in this study, representing mildly to highly heterogeneous<br />

conductivity fields. The parameters lY <strong>and</strong> σY 2 =0.38 are taken from the Borden field site (Sudicky,<br />

1986); the value <strong>of</strong> 1.71 was found at the Testfeld Süd in southern Germany (Herfort, 2000); the<br />

values <strong>of</strong> 2.7 <strong>and</strong> 4.5 were reported for the Columbus Air Force Base site (Rehfeldt et al., 1992). A<br />

constant porosity n <strong>of</strong> 0.33 is used, resulting in a mean <strong>flow</strong> velocity va <strong>of</strong> 1.16·10 −5 ms −1 . For each<br />

<strong>of</strong> the four degrees <strong>of</strong> heterogeneity, ensembles <strong>of</strong> 100 realizations <strong>of</strong> the r<strong>and</strong>om field are generated<br />

by unconditional Gaussian simulation. The Monte-Carlo strategy is chosen in order to obtain<br />

statistical measures <strong>of</strong> the errors <strong>and</strong> uncertainties introduced by the heterogeneity <strong>of</strong> K. Plumes <strong>of</strong><br />

the three compounds are generated in each virtual aquifer using a process based <strong>numerical</strong> <strong>flow</strong> <strong>and</strong><br />

reactive transport model. The GeoSys/RockFlow simulation code (Kolditz et al., 2004)isusedhere,<br />

which solves the <strong>flow</strong> <strong>and</strong> transport equations by finite element methods. The governing equations<br />

are given as (e.g. Bear, 1972; Kolditz, 2002):<br />

jðKjhÞ ¼0 ð11Þ<br />

∂C<br />

∂t ¼ −vajC þ jðDjCÞ−C ð12Þ<br />

with h as the piezometric height, K the tensor <strong>of</strong> hydraulic conductivity, C concentration, D the<br />

dispersion tensor, t time <strong>and</strong> Γ a sink term representing first order or MM degradation kinetics. For<br />

local dispersivities αL <strong>and</strong> αT values <strong>of</strong> 0.25 m <strong>and</strong> 0.05 m were used. Details <strong>of</strong> <strong>numerical</strong> <strong>and</strong><br />

s<strong>of</strong>tware issues can be found in Kolditz (2002) <strong>and</strong> Kolditz <strong>and</strong> Bauer (2004). All model parameters<br />

are summarized in Table 4.<br />

5. Investigation <strong>of</strong> the synthetic plumes<br />

The steady state contaminant plumes in each virtual aquifer are investigated by the center line<br />

method (see Fig. 2). Initially three observation wells are present in the aquifer (Fig. 2 (a)), one being<br />

located directly in the source in the center <strong>of</strong> the aquifer (full circle) at [13.0 m; 32.0 m], showing<br />

high concentrations. This setup is the starting point for the investigation <strong>of</strong> all realizations. The initial<br />

knowledge on the site comprises only the hydraulic heads at the three wells. The full concentration,<br />

Table 4<br />

Model parameters used in the <strong>numerical</strong> simulations<br />

Parameter Value<br />

Kef Effective conductivity 7.19·10 − 5 − 1<br />

ms<br />

σY 2<br />

ln(K)-variance 0, 0.38, 1.71, 2.7, 4.5<br />

lY Integral scale 2.67 m<br />

n Porosity 0.33<br />

I Hydraulic gradient 0.053<br />

αL Longitudinal dispersivity 0.25 m<br />

α T Transverse dispersivity 0.05 m<br />

λ First order degradation rate constant 5.87·10 − 7 − 1<br />

s<br />

kmax<br />

Maximum degradation velocity a<br />

1.45·10 − 7 − 1<br />

s<br />

MC<br />

C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

Half saturation concentration a<br />

a In normalized units, see explanation in the text.<br />

0.0497<br />

81


82 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

Fig. 2. Investigation <strong>of</strong> a virtual plume by the center line approach: (a) initial situation, (b) estimation <strong>of</strong> <strong>flow</strong> direction by<br />

application <strong>of</strong> a hydrogeologic triangle, (c) observation wells on the inferred center line, (d) comparison to virtual reality<br />

(concentration <strong>and</strong> heads).<br />

head <strong>and</strong> conductivity distributions <strong>of</strong> the virtual sites are assumed unknown. In the first investigation<br />

step, the <strong>flow</strong> direction is determined (Fig. 2 (b)). A hydrogeologic triangle is constructed<br />

<strong>and</strong> the hydraulic gradient is calculated using the heads measured at the three wells. After this, five<br />

new observation wells are installed along the estimated <strong>flow</strong> direction with distances <strong>of</strong> 15, 50, 75,<br />

100 <strong>and</strong> 125 m from the source (Fig. 2 (c)). At these <strong>and</strong> at the well at the source hydraulic heads <strong>and</strong><br />

concentrations <strong>of</strong> the three compounds are measured. Additionally, local hydraulic conductivities are<br />

determined (e.g. by a slug test). The geometric mean <strong>of</strong> these six K values is used as an estimator for<br />

the effective conductivity Kef along the <strong>flow</strong> path. From the head difference, the true porosity <strong>and</strong> Kef<br />

an average va is calculated. For methods 3 <strong>and</strong> 4, estimates <strong>of</strong> dispersivities αL <strong>and</strong> αT are required.<br />

Following Wiedemeier et al. (1999) αL is taken as 0.1 <strong>of</strong> the plume length <strong>and</strong> αT is assumed to be 0.1<br />

<strong>of</strong> αL. As the true plume length is unknown at this stage <strong>of</strong> the site investigation, the maximum<br />

distance covered by the observation wells, i.e. 125 m, is used. Consequently, α L <strong>and</strong> α Tare estimated<br />

to be 12.5 <strong>and</strong> 1.25 m, respectively. Of course, such rather rough estimates <strong>of</strong> α L <strong>and</strong> α T are not<br />

optimal, as they are not based on the heterogeneity structure <strong>of</strong> the aquifer. In practice, however,<br />

dispersivities based on results from stochastic hydrogeology are difficult to obtain, as for most field<br />

sites structure <strong>and</strong> degree <strong>of</strong> heterogeneity are not well characterized. Also with this scenario, the


samples taken at all eight wells do not allow for an estimation <strong>of</strong> lY<strong>and</strong> σY 2 , which would be required<br />

to derive αL <strong>and</strong> αT. A detailed study on the effects <strong>of</strong> dispersivity parameterization on the<br />

performance <strong>of</strong> methods 2 <strong>and</strong> 3 is presented in Bauer et al. (2006).<br />

The information obtained by the site investigation allows the application <strong>of</strong> the five methods for<br />

estimation <strong>of</strong> the degradation kinetic parameters <strong>and</strong> the subsequent calculation <strong>of</strong> contaminant<br />

plume lengths as presented in Sections 4.1 <strong>and</strong> 4.2. The investigation setup is designed to resemble<br />

ideal conditions for this purpose. All measurements are assumed to be exact, i.e. without measurement<br />

error, <strong>and</strong> are obtained by reading the model output at the respective well positions. The<br />

only uncertainty <strong>and</strong> variability is introduced by the heterogeneity <strong>of</strong> the hydraulic conductivity.<br />

6. Results <strong>and</strong> discussion<br />

C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

As outlined in Section 3, three different cases A, B <strong>and</strong> C are investigated here (see Table 1). The<br />

next three Sections 6.1, 6.2 <strong>and</strong> 6.3 present detailed results <strong>and</strong> discussions for each case studied. A<br />

detailed comparison <strong>and</strong> discussion <strong>of</strong> the performance <strong>of</strong> the different approaches is given in<br />

Section 6.4.<br />

6.1. Case A: estimation <strong>of</strong> first order degradation rate constants <strong>and</strong> plume lengths for plumes<br />

following a first order degradation kinetics<br />

6.1.1. Estimation <strong>of</strong> rate constants<br />

In case A methods 1–4 are tested <strong>and</strong> compared based on their ability to estimate the first order<br />

degradation rate constant λ <strong>and</strong> the contaminant plume length in heterogeneous aquifers.<br />

Therefore, plumes following first order degradation kinetics are investigated here. The four<br />

estimated rate constants λi are divided by the true rate constant to yield corresponding normalized<br />

Λi, which can be interpreted as an over- or underestimation factor. Fig. 3 presents ensemble means<br />

with corresponding st<strong>and</strong>ard deviations as error bars, medians, coefficients <strong>of</strong> variation <strong>and</strong> the<br />

single realization results for methods 1–4 against the aquifer heterogeneity σY 2 .<br />

Results for method 1 (Fig. 3 (a)) show that Λ1=1 in the homogeneous case (σY 2 =0), thus in this case<br />

the true rate is obtained. However, already for the lowest degree <strong>of</strong> heterogeneity (σY 2 =0.38), about<br />

77% <strong>of</strong> the rate constants estimated fall above the reference line which indicates the true rate constant,<br />

overestimating λ up to factors <strong>of</strong> 4.79 in the worst case. In the remaining realizations λ is<br />

underestimated. The spread <strong>of</strong> the ensemble <strong>of</strong> 100 realizations covers about one order <strong>of</strong> magnitude.<br />

When σY 2 is further increased, spread <strong>and</strong> st<strong>and</strong>ard deviation also increase. For σY 2 =4.5 the Λ1 cover<br />

almost two orders <strong>of</strong> magnitude with single realizations showing overestimation factors N10. Mean Λ1<br />

is 3.3 with st<strong>and</strong>ard deviation <strong>and</strong> coefficient <strong>of</strong> variation <strong>of</strong> 3.65 <strong>and</strong> 1.11, respectively. Increasing the<br />

heterogeneity <strong>of</strong> the aquifer thus results in a higher uncertainty <strong>of</strong> λ <strong>and</strong> increases the probability <strong>of</strong> a<br />

significant overestimation. For method 2 normalized rate constants Λ2 are shown in Fig. 3 (b). As seen<br />

for method 1 a systematic overestimation <strong>of</strong> λ can be observed, increasing with σ Y 2 . However, the Λ2<br />

are significantly larger than the corresponding Λ 1. In the homogeneous case method 2 yields Λ 2=1.7,<br />

increasing to 9.38 for σY 2 =4.50. In extreme cases, overestimation <strong>of</strong> λ is larger than a factor <strong>of</strong> 50. Also<br />

spread <strong>and</strong> st<strong>and</strong>ard deviations are larger than for Λ1. Estimated rate constants Λ3 obtained by method<br />

3(Fig. 3 (c)) are only slightly lower than those <strong>of</strong> method 2, ranging from 1.17 in the homogeneous<br />

case to 7.39 for σY 2 =4.5. Method 4 (Fig. 3(d)) yields the true rate for homogeneous conditions, i.e.<br />

Λ4=1. As for the other methods, spread <strong>and</strong> uncertainty increase with σY 2 . Compared to methods 1–3<br />

the spread is smaller <strong>and</strong> balanced around the true rate constant. For σY 2 ≤2.7 the ensemble averages<br />

<strong>and</strong> medians match the true rate constant well. For σ Y 2 =4.50 the average Λ4 reaches 1.73.<br />

83


84 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

The overestimation observed for the different Λi results from a combination <strong>of</strong> several effects.<br />

Method 1, which is the simplest approach used in this study, is based on the one dimensional transport<br />

equation only accounting for advection <strong>and</strong> degradation. Concentration reductions on the plume<br />

center line caused by transverse dispersion therefore are incorrectly attributed to the degradation<br />

process <strong>and</strong> Λ1 is overestimated. Moreover, when the inferred center line deviates from the true center<br />

line, concentrations measured are too low, which further increases rate overestimation. A third source<br />

<strong>of</strong> error is the estimate <strong>of</strong> v a (see Section 5), which may not be representative for the <strong>flow</strong> path.<br />

Overestimation is caused if v a is estimated too high, while a too low value <strong>of</strong> v a causes underestimation<br />

<strong>of</strong> the rate constant. All three error types are also relevant for method 2. The additional bias <strong>of</strong> method<br />

2 towards too large rate constants in comparison with method 1 is a consequence <strong>of</strong> accounting for αL<br />

in equation (2) <strong>of</strong> Table 2. Longitudinal dispersion <strong>of</strong> a degrading contaminant results in a stronger<br />

spreading <strong>of</strong> the solute down stream <strong>and</strong> consequently in higher concentrations along the center line <strong>of</strong><br />

a steady state plume. Equation (2) can be rearranged to show that λ2=λ1+αLva(ln(C(x)/C0)/Δx) 2 , i.e.<br />

λ2 grows linearly with αL <strong>and</strong> is always larger than λ1 for αLN0.Method3isaffectedbyerrorsinva<br />

<strong>and</strong> <strong>of</strong>f center line measurements. By accounting for transverse dispersion, rate constant estimates are<br />

improved compared to method 2, because βb1.0 in equation (4) <strong>and</strong> thus Λ 3bΛ 2 always. Because β<br />

approaches unity for arguments N2, Λ3 converges with Λ2 for small αT, short transport distances Δx<br />

Fig. 3. Estimated first order degradation rate constants Λi (normalized to the true rate constant λ, indicated by the<br />

horizontal line) vs. heterogeneity <strong>of</strong> the aquifer σ Y 2 for methods 1 (a), 2 (b), 3 (c) <strong>and</strong> 4 (d) for case A.


C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

<strong>and</strong> large source widths WS. A surprising result is that method 1, despite its simplicity, yields closer<br />

estimates <strong>of</strong> the true rate constant than the more comprehensive description by method 3. Since<br />

method 3 depends on longitudinal <strong>and</strong> transverse dispersivities, an adequate parameterization is<br />

crucial for its success. From stochastic hydrogeology it is known that αL as well as αT strongly<br />

depends on travel time <strong>and</strong> distance as well as on the correlation structure <strong>of</strong> hydraulic conductivity<br />

<strong>and</strong> <strong>flow</strong> velocity (e.g. Dagan, 1989). Consequently, a uniform parameterization solely based on the<br />

scale <strong>of</strong> the contaminant problem as used in this study (<strong>and</strong> with many field applications) is not<br />

adequate. A detailed sensitivity study on the influence <strong>of</strong> dispersivity parameterization on the<br />

performance <strong>of</strong> methods 2 <strong>and</strong> 3 is presented in Bauer et al. (2006). It is found that for method 2 no<br />

value <strong>of</strong> αL <strong>and</strong> for method 3 only very high <strong>and</strong> thus unphysical values <strong>of</strong> αT yield the correct<br />

degradation rate constant. The required values, however, cannot be deduced from aquifer heterogeneity<br />

σY 2 alone, as the other errors also influence the estimated degradation rate constant.<br />

Method 4 circumvents this problem <strong>and</strong> corrects for transverse dispersion as well as for measuring<br />

<strong>of</strong>f the center line by normalizing concentrations to a conservative tracer. The bias towards too large<br />

degradation rate constants observed for the other methods is significantly reduced yielding the closest<br />

estimates <strong>of</strong> λ <strong>of</strong> the four methods. The remaining deviation from the true rate constant is due to the<br />

hydraulic error introduced by the approximation <strong>of</strong> va. For low heterogeneities there is no evidence for<br />

a systematic bias towards either too high or too low rate constants. A prerequisite which may limit the<br />

applicability <strong>of</strong> method 4 is the presence <strong>of</strong> a suited normalization compound. A discussion <strong>of</strong> potential<br />

normalization compounds is provided in U.S. EPA (1998) <strong>and</strong> Wiedemeier et al. (1996, 1999).<br />

6.1.2. Estimation <strong>of</strong> plume lengths<br />

During site characterization, estimating degradation rate constants rarely is a goal per se. Here, the<br />

kinetics <strong>of</strong> contaminant degradation are quantified to be used for prediction <strong>of</strong> the steady state length<br />

<strong>of</strong> the plumes. In site assessment, such information could be used to identify potential receptors <strong>and</strong><br />

exposure levels. Rate constants λ1, λ2 <strong>and</strong> λ3 are evaluated using the respective corresponding<br />

equations (7), (8) <strong>and</strong> (9) <strong>of</strong> Table 3 yielding plume length estimates L1, L2 <strong>and</strong> L3. Forλ4 also<br />

equation (9) is used yielding L4. To be able to compare the results <strong>of</strong> all realizations, the Li are<br />

normalized by the respective true length L read from the model output. Resulting over- respectively<br />

underestimation factors against aquifer heterogeneity σY 2 are presented in Fig. 4 (single realization<br />

results, ensemble means with st<strong>and</strong>ard deviations as error bars, medians, coefficients <strong>of</strong> variation).<br />

Plume lengths L 1 <strong>and</strong> L 2, calculated from λ 1 <strong>and</strong> λ 2, show exactly identical results, although<br />

the λ2 show a stronger overestimation than the corresponding λ1 for all realizations. The reason<br />

for the equivalence <strong>of</strong> L1 <strong>and</strong> L2 is that the bias introduced by estimating λ2 with a one<br />

dimensional model accounting for longitudinal dispersion only is reversed by using the same<br />

transport equation to calculate the plume length. While for homogeneous conditions the true<br />

plume length is obtained, L is underestimated in most realizations for all degrees <strong>of</strong> heterogeneity.<br />

Mean L1 <strong>and</strong> L2 decrease to 0.59 for σY 2 =4.5. As for λ, spread <strong>and</strong> uncertainty <strong>of</strong> L increase with<br />

σ Y 2 . The overestimation <strong>of</strong> the degradation rate constant is thus reflected in an underestimation <strong>of</strong><br />

the plume length.<br />

Plume lengths L3 calculated with the two-dimensional transport equation on average are lower<br />

than the corresponding L1 <strong>and</strong> L2. This is a consequence <strong>of</strong> the β term in equation (5) (Table 2),<br />

which is used to correct down gradient concentrations for transverse dispersion. When estimating<br />

the rate constant with method 3, each down gradient concentration C(x) is scaled by a different<br />

value β, as the correction factor is dependent on the distance from the source. This scaling is not<br />

fully reversed when the plume length is calculated using equation (9) (Table 3), as then only one<br />

single Δx is used.<br />

85


86 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

As shown before, degradation rate constants λ4 in general are significantly better estimated<br />

than those obtained by the other three approaches. This is also reflected in plume lengths L4 (Fig.<br />

4 (c)). For all σY 2 the ensemble means almost perfectly match the true L. Of all four approaches the<br />

L4 consistently show the lowest coefficients <strong>of</strong> variation, indicating the lowest spread <strong>and</strong> thus the<br />

lowest uncertainty.<br />

The plume length results presented for case A demonstrate that the determination <strong>of</strong><br />

degradation rate constants using the center line method as well as the subsequent calculation <strong>of</strong><br />

contaminant plume lengths both are subject to uncertainty induced by the heterogeneity <strong>of</strong> the<br />

medium. However, the uncertainty observed for the rate constants is only partially transferred to<br />

the calculated plume lengths. For all approaches the observed spread <strong>of</strong> the ensembles <strong>of</strong> estimated<br />

plume lengths is smaller than for the corresponding rate constants. This is most obvious for the L4,<br />

which also show the best agreement with the true plume lengths. Moreover, the L4 are unbiased,<br />

showing equal amounts <strong>of</strong> under- as well as overestimation. However, both types <strong>of</strong> error are<br />

undesirable. Underestimation <strong>of</strong> the contaminant plume length, i.e. a “non-conservative” result,<br />

may pose a threat to down gradient receptors. On the other h<strong>and</strong>, a “conservative” result, i.e. an<br />

overestimation <strong>of</strong> plume dimensions, might result in wrong decisions regarding the necessity <strong>and</strong><br />

dimensioning <strong>of</strong> engineered remediation measures with unnecessary financial expenses. In<br />

Fig. 4. Plume lengths Li calculated with degradation rate constants λ1 through λ4. The Li are normalized by the true plume<br />

length L (indicated by the horizontal line) for case A.


C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

contrast to L4, plume lengths L1, L2 <strong>and</strong> L3 show a clear tendency <strong>of</strong> underestimation, demonstrating<br />

the non-conservativeness <strong>of</strong> too large rate constants. Comparing methods 1, 2 <strong>and</strong> 3,<br />

differences in calculated plume lengths are not as pronounced as for the rate constants. Thus it can<br />

be concluded, that the same underlying equation should be used for rate constant <strong>and</strong> plume length<br />

estimation. However, if e.g. a rate constant estimated with method 2 is used in a two- or threedimensional<br />

(<strong>numerical</strong>) transport model, results are susceptible to the bias in method 2, resulting<br />

in a significantly stronger underestimation <strong>of</strong> the plume length.<br />

6.2. Case B: estimation <strong>of</strong> first order degradation rate constants <strong>and</strong> plume lengths for plumes<br />

following Michaelis–Menten degradation kinetics<br />

In case B the additional error is studied, that arises when the four methods for the estimation <strong>of</strong><br />

first order degradation rate constants (Table 2, equations (1)–(4)) <strong>and</strong> the respective equations for<br />

plume lengths (Table 3, equations (7)–(9)) are used for plumes with a degradation kinetics<br />

deviating from first order but following MM degradation kinetics instead (see Table 1).<br />

Since a direct comparison <strong>of</strong> estimated rate constants λ 1 through λ 4 with the MM parameters k max<br />

<strong>and</strong> MC used in the <strong>numerical</strong> simulations is not possible, the evaluation here is based on calculated<br />

Fig. 5. Normalized plume lengths Li calculated with degradation rate constants λ1 through λ4 for the contaminant plumes<br />

following Michaelis–Menten degradation kinetics.<br />

87


88 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

plume lengths only. The estimated contaminant plume lengths Li are displayed in Fig. 5 (single<br />

realization results, ensemble means with st<strong>and</strong>ard deviations as error bars, medians, coefficients <strong>of</strong><br />

variation). As for case A, L1 <strong>and</strong> L2 (Fig. 5 (a)) yield very similar results. However, in comparison to<br />

case A, underestimation is clearly increased. This is most obvious for homogeneous conditions, where<br />

L1 <strong>and</strong> L2 are only about 50% <strong>of</strong> the true length <strong>and</strong> thus no longer yield the correct result.<br />

Underestimation <strong>of</strong> L increases with heterogeneity, yielding mean values <strong>of</strong> L1 <strong>and</strong> L2 <strong>of</strong> 0.45 for<br />

σ Y 2 =4.5. Spread is about 0.75 orders <strong>of</strong> magnitude for low heterogeneity <strong>and</strong> one order <strong>of</strong> magnitude<br />

for high heterogeneities, which is similar to case A. Plume lengths L 3 (Fig. 5 (b)) show the same<br />

general behaviour as the L1 <strong>and</strong> L2, with mean values being slightly lower. In contrast to this, L is<br />

overestimated for most realizations by L4 (Fig. 5 (c)). While for homogeneous conditions a too short<br />

L4 <strong>of</strong> 0.65 is obtained, L4 increases to values larger than one for higher degrees <strong>of</strong> heterogeneity.<br />

Spread <strong>of</strong> single realizations is significantly increased in comparison to case A (compare Figs. 4 (c)<br />

<strong>and</strong> 5(c)), as a spread <strong>of</strong> about one order <strong>of</strong> magnitude can be observed for all degrees <strong>of</strong> heterogeneity.<br />

The additional error introduced by using methods 1 through 4 for plumes following Michaelis–<br />

Menten kinetics degradation is most pronounced for low heterogeneities. Compared to case A,<br />

plume lengths are underestimated to a larger extent, as can be seen by the lower mean values <strong>and</strong><br />

Fig. 6. Normalized Michaelis–Menten kinetics parameters kmax vs. MC estimated for σY 2 =0.38 (a), 1.71 (b), 2.7 (c) <strong>and</strong> 4.5<br />

(d), respectively.


C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

medians. Only for L4 plume length overestimation is observed <strong>and</strong> uncertainty is increased in<br />

comparison to case A.<br />

6.3. Case C: estimation <strong>of</strong> Michaelis–Menten kinetics parameters <strong>and</strong> plume lengths for plumes<br />

following Michaelis–Menten degradation kinetics<br />

6.3.1. Estimation <strong>of</strong> rate parameters<br />

As demonstrated in case B, using a first order kinetics approximation for plumes following MM<br />

degradation kinetics introduces an additional error, yielding less conservative estimates <strong>of</strong> plume<br />

lengths (L1−L3) as well as higher uncertainty (L4). Therefore, method 5 for the estimation <strong>of</strong> the<br />

MM parameters is tested by application to the contaminant plumes following MM degradation<br />

kinetics (see Table 1). Using the same concentration vs. distance data as in Section 6.2, MM<br />

parameters kmax <strong>and</strong> MC are estimated by a linear fit as explained in Section 4.1. These parameters<br />

then are used to calculate the length <strong>of</strong> the contaminant plumes LMM by equation (10) (Table 3).<br />

Fig. 6 presents the results <strong>of</strong> the parameter estimation in single diagrams for each σ Y 2 (results<br />

for single realizations, ensemble means with corresponding st<strong>and</strong>ard deviations as error bars,<br />

medians). Since for each realization the maximum degradation rate kmax <strong>and</strong> half-saturation<br />

concentration MC are estimated simultaneously, the parameters are shown in scatter plots. For the<br />

homogeneous case (not shown), kmax is slightly overestimated with a value <strong>of</strong> 1.06, while for MC<br />

an underestimation is observed with MC=0.91. This error results from neglecting the dispersion<br />

process. Fig. 6 shows that normalized kmax <strong>and</strong> MC are increasingly overestimated with increasing<br />

σY 2 . Also uncertainty increases with σY 2 . This is a similar behaviour as found for the first order rate<br />

constants in case A. Approximate orientation <strong>of</strong> the data points along a diagonal axis with positive<br />

slope gives evidence <strong>of</strong> a weak positive correlation between k max <strong>and</strong> M C. An overestimation <strong>of</strong><br />

kmax increases the degradation rate as long as concentrations are higher than MC. However,<br />

overestimation <strong>of</strong> MC raises the concentration threshold at which kinetic degradation transits from<br />

zeroth to slower first order, which counter-balances the effects <strong>of</strong> the kmax overestimation.<br />

Fig. 7. Normalized plume lengths LMM calculated with Eq. (10) based on estimated Michaelis–Menten kinetics parameters<br />

k max <strong>and</strong> M C.<br />

89


90 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

6.3.2. Estimation <strong>of</strong> plume lengths<br />

For the determination <strong>of</strong> contaminant plume lengths, the estimated parameters kmax <strong>and</strong> MC are<br />

evaluated with equation (10) (Table 3). Normalized plume lengths LMM are presented in Fig. 7<br />

(single realization results, ensemble means with corresponding st<strong>and</strong>ard deviations as error bars,<br />

medians, coefficients <strong>of</strong> variation). For the homogeneous case the estimated <strong>and</strong> true plume<br />

lengths agree well with LMM=0.97. Also for σY 2 =0.38 the mean yields the correct result with<br />

L MM=1.0 <strong>and</strong> spread is about 0.5 orders <strong>of</strong> magnitude. For higher heterogeneities, however,<br />

plume lengths again tend to be underestimated with L MM decreasing to 0.80 for σ Y 2 =4.5. Spread<br />

for all degrees <strong>of</strong> heterogeneity is smaller than one order <strong>of</strong> magnitude. When the spread <strong>of</strong><br />

estimated kmax <strong>and</strong> MC values is compared to the spread <strong>of</strong> calculated MM plume lengths, it is<br />

found – as for case A – that the uncertainty in MM parameters is not fully propagated to the<br />

plume lengths. Comparing MM plume lengths with those determined for the plumes subject to<br />

first order degradation in case A (see Fig. 4), the LMM outperform L1, L2 or L3 with regard to<br />

accuracy as well as uncertainty. Only the L4 are closer to the true lengths on average, while the<br />

spread <strong>of</strong> single realizations is comparable. It is thus found in this study that the proposed method<br />

<strong>of</strong> estimating MM kinetics parameters is adequate <strong>and</strong> performs better than the methods for<br />

estimating first order rate constants.<br />

6.4. Comparison <strong>and</strong> discussion <strong>of</strong> cases B <strong>and</strong> C<br />

In this section the plume lengths LMM calculated in case C using the MM model (Table 3,<br />

equation (10)) are compared to the L1 −L4 <strong>of</strong> case B, where plume lengths were calculated using<br />

the first order kinetics approximations <strong>of</strong> equations (7) through (9) <strong>of</strong> Table 3. Table 5 compares<br />

the accuracy <strong>of</strong> calculated plume lengths using MM parameters estimated by method 5 with the<br />

four different first order approaches. Given are the percentages <strong>of</strong> realizations for which LMM<br />

constitutes a closer estimate <strong>of</strong> L than L1, L2, L3, orL4. In comparison to L1, L2 <strong>and</strong> L3 it is found<br />

that LMM is the closer estimate for 93 up to 99% <strong>of</strong> all realizations, depending on the degree <strong>of</strong><br />

heterogeneity <strong>and</strong> the first order method used. In comparison with L4 LMM constitutes the closer<br />

estimate still for 53 up to 72% <strong>of</strong> all realizations.<br />

As a more quantitative criterion, plume length error factors EF are calculated for each<br />

realization <strong>and</strong> all five estimation methods:<br />

EF ¼ L *<br />

L<br />

a<br />

with<br />

a ¼ −1jL * bL<br />

a ¼þ1jL * zL<br />

where L ⁎ <strong>and</strong> L are the estimated <strong>and</strong> true plume lengths, respectively, <strong>and</strong> the exponent a is used to<br />

obtain a st<strong>and</strong>ardized measure for over- as well as for underestimation. Hence, EF gives the degree <strong>of</strong><br />

accuracy <strong>of</strong> the plume length estimate L ⁎. For each ensemble <strong>of</strong> Li <strong>and</strong> LMM cumulative empirical<br />

Table 5<br />

Percentage <strong>of</strong> realizations for which LMM is a closer estimate <strong>of</strong> L than L1−L4<br />

LMM vs. L1, L2 % LMM vs. L3 % LMM vs. L4 %<br />

0.38 97 99 72<br />

1.71 99 99 57<br />

2.7 98 99 59<br />

4.5 94 93 53<br />

σ Y 2<br />

ð13Þ


C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

Fig. 8. Cumulative empirical distribution functions <strong>of</strong> plume length error factors for estimated plume lengths L1 through L4<br />

<strong>and</strong> LMM <strong>and</strong> aquifer heterogeneities σY 2 <strong>of</strong> 0.38 (a), 1.71 (b), 2.7 (c) <strong>and</strong> 4.5 (d). The diagrams yield the probability <strong>of</strong><br />

obtaining an estimate <strong>of</strong> the plume length L with an accuracy (EF) as given on the abscissa.<br />

distribution functions (edf) <strong>of</strong> the EF were calculated. These are presented in Fig. 8 (a) through (d) for<br />

each degree <strong>of</strong> heterogeneity. The edf give the probability <strong>of</strong> obtaining an estimate <strong>of</strong> the plume<br />

length with an EF less than or equal to the associated quantile on the abscissa. In Fig. 8 (a) with<br />

σY 2 =0.38, for example, the probability <strong>of</strong> estimating L by LMM, given an accuracy <strong>of</strong> factor 2, i.e.<br />

allowing a maximum underestimation by 50% or overestimation by 100%, is about 0.98. In contrast<br />

to this, the probability <strong>of</strong> obtaining L1 (=L2)orL3 as accurate as a factor <strong>of</strong> 2 is only 0.68 <strong>and</strong> 0.55,<br />

respectively, while for L4 the probability is approximately 0.89. When aquifer heterogeneity<br />

increases, the probabilities are reduced, as can be seen by the shift towards higher EF <strong>and</strong> the<br />

flattening <strong>of</strong> the slopes (note the logarithmic scale <strong>of</strong> the abscissa). Moreover, differences between<br />

LMM <strong>and</strong> L4 diminish. Thus for σY 2 =4.5 (Fig. 8 (d)) edf for both approaches almost coincide <strong>and</strong><br />

show the same degree <strong>of</strong> accuracy. However, LMM still yields significantly higher probabilities for a<br />

given EF than L1, L2 or L3.<br />

The results obtained in this comparison clearly support that for plumes following MM degradation<br />

kinetics usage <strong>of</strong> the MM parameter estimation approach allows a distinct improvement <strong>of</strong><br />

plume length estimates over those obtained using a first order approximation, especially when<br />

aquifer heterogeneity is not too high. For the majority <strong>of</strong> realizations investigated in this study, L MM<br />

91


92 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

constitutes a more accurate estimate <strong>of</strong> the true plume length than any <strong>of</strong> the first order methods used.<br />

Only when aquifer heterogeneity is very high, L4 is able to yield plume length estimates <strong>of</strong> almost<br />

comparable accuracy. This is mainly, because increased transverse dispersion produces concentration<br />

pr<strong>of</strong>iles that allow a log-linear fit <strong>of</strong> a first order rate constant.<br />

7. Summary <strong>and</strong> conclusions<br />

In this study the Virtual Aquifer concept is used to assess the uncertainty involved in estimating<br />

down stream contaminant concentrations <strong>and</strong> plume lengths in heterogeneous aquifers. For such<br />

an analysis a key element is the quantification <strong>of</strong> the degradation rate at the site under study.<br />

Therefore, the main focus <strong>of</strong> this work is on the influence <strong>of</strong> this parameter on the estimated plume<br />

lengths. Three different scenarios are analysed.<br />

In case A, four different st<strong>and</strong>ard field methods based on the center line investigation strategy<br />

are tested <strong>and</strong> compared with regard to their capability <strong>of</strong> estimating first order degradation rate<br />

constants <strong>and</strong> the contaminant plume length in heterogeneous aquifers. The four methods are<br />

applied to plumes following first order degradation kinetics. It is found that both, the estimated<br />

degradation rate constants <strong>and</strong> the calculated plume lengths, are subject to high uncertainty. On<br />

average, estimated rate constants exceed the true degradation rate constant, causing calculated<br />

plume lengths that are too short. Both bias <strong>and</strong> uncertainty <strong>of</strong> estimated degradation rate constants<br />

increase with the degree <strong>of</strong> heterogeneity to about a factor in the order <strong>of</strong> a magnitude, respectively.<br />

However, the uncertainty observed in the rate constants does not fully propagate to the plume length<br />

estimates. On average plume lengths are underestimated by about 50% <strong>of</strong> the true plume length, <strong>and</strong><br />

up to a factor <strong>of</strong> ten in the worst cases. Of the four different methods, the approach <strong>of</strong> Wiedemeier et<br />

al. (1996) using concentrations normalized to a conservative tracer yields the best results for the rate<br />

constants <strong>and</strong> for the plume lengths. Consequently, this method should preferably be used.<br />

However, the presence <strong>of</strong> a suitable recalcitrant compound may not always be given. In this case,<br />

the most simple <strong>of</strong> the four approaches, which is based on the advection equation <strong>and</strong> neglects<br />

dispersive processes should be used to determine the degradation potential, as this method yields<br />

closer estimates <strong>of</strong> the first order rate constant than the approaches using the one- (Buscheck <strong>and</strong><br />

Alcantar, 1995) <strong>and</strong> two-dimensional advection dispersion equations (Zhang <strong>and</strong> Heathcote, 2003).<br />

The non-conservative plume length estimates might cause threats to down stream receptors, as the<br />

risk <strong>of</strong> contamination could be underrated. The high uncertainty when estimating plume dimensions<br />

could result in incorrect decisions regarding the necessity <strong>and</strong> dimensioning <strong>of</strong> engineered remediation<br />

measures or when considering the applicability <strong>of</strong> natural attenuation. All four methods are only<br />

applicable to steady state plumes. In reality, however, contaminant plumes <strong>of</strong>ten show temporal<br />

variations in extent <strong>and</strong> orientation as the plume exp<strong>and</strong>s, the source slowly depletes, or the <strong>flow</strong><br />

regime changes over time. For exp<strong>and</strong>ing plumes, application <strong>of</strong> methods 1–4 wouldresultinan<br />

overestimation <strong>of</strong> the rate constant, <strong>and</strong> thus in underestimation <strong>of</strong> the plume length at steady state. This<br />

is because the down stream transient contaminant concentrations are lower than those at steady state,<br />

causing a higher concentration decrease along the center line, which would be falsely attributed to the<br />

degradation process. For the same reason rate constants would also be overestimated for shrinking<br />

plumes. An examination <strong>of</strong> the current state <strong>of</strong> the plume is necessary, when one <strong>of</strong> the four methods is<br />

to be applied at a site. An overview <strong>of</strong> techniques for this purpose is given by Newell et al. (2002).<br />

Case B analyses the additional error that results from application <strong>of</strong> the four methods, although<br />

the true degradation kinetics deviate from first order. Here plumes following Michaelis–Menten<br />

degradation kinetics are investigated. Results show that for the three approaches without<br />

correction <strong>of</strong> concentrations to a conservative tracer, an additional underestimation occurs which


is largest for low heterogeneities. For the approach using the tracer correction, the estimated<br />

plume lengths still match the true lengths on average. However, the uncertainty <strong>of</strong> the plume<br />

lengths is significantly increased as compared to case A.<br />

In case C, a regression approach is introduced to estimate the parameters <strong>of</strong> the MM kinetics <strong>and</strong> to<br />

calculate the plume lengths for the same plumes as investigated in case B. Since longitudinal <strong>and</strong><br />

transverse dispersion are neglected in this method, the MM parameters are overestimated on average.<br />

Overestimation increases from less than a factor <strong>of</strong> two for low degrees <strong>of</strong> heterogeneity to a factor <strong>of</strong><br />

roughly four for high heterogeneity. Consequently an underestimation <strong>of</strong> the corresponding plume<br />

lengths is observed for most realizations. For low heterogeneity, the plume length is estimated<br />

precisely with low uncertainty, for high heterogeneity the average estimated plume length is about<br />

80% <strong>of</strong> the true plume length, with estimates as low as 40–35% in the worst cases. However, in<br />

comparison with the first order approximation investigated in case B the error resulting from this<br />

approach is significantly reduced, with the uncertainty <strong>of</strong> the plume length estimates being reduced by<br />

a factor <strong>of</strong> three. Therefore, if field data collected along the center line <strong>of</strong> a plume gives evidence <strong>of</strong><br />

MM kinetics (linear behaviour <strong>of</strong> concentrations vs. distance in a linear plot, concave down behaviour<br />

in a semi-logarithmic plot (Bekins et al., 1998)), this approach is recommended.<br />

Generally, the Virtual Aquifer concept has proven useful in assessing the performance <strong>of</strong><br />

methods for investigating first order rate constants <strong>and</strong> plume lengths from plume center line<br />

measurements. The main advantage is, that individual factors, as here the hydraulic heterogeneity<br />

<strong>and</strong> the assumption <strong>of</strong> a wrong plume kinetics, can be studied either individually in detail or in<br />

combination <strong>and</strong> under otherwise ideal conditions. However, the restrictions from the simplified<br />

setup <strong>of</strong> the model scenario have to be kept in mind when drawing conclusions for field<br />

applications. In reality, contaminant degradation follows more complicated rate laws, depending<br />

e.g. on electron acceptor <strong>and</strong> donor availability, may include transient effects, or dilution <strong>and</strong><br />

phase changes to the un<strong>saturated</strong> zone. Further studies will thus incorporate more realistic<br />

degradation kinetics as well as influences from e.g. measurement errors.<br />

Acknowledgements<br />

This work is funded by the German Ministry <strong>of</strong> Education <strong>and</strong> Research (BMBF) under grant<br />

033 05 12/033 05 13 as part <strong>of</strong> the KORA priority program, sub-project 7.2. We wish to thank<br />

Robert Walsh for his helpful comments on the manuscript. We also wish to thank our project<br />

partners at the Christian-Albrechts-University Kiel Andreas Dahmke <strong>and</strong> Dirk Schäfer for their<br />

support in our research. We acknowledge Uwe Wittmann, Iris Bernhardt <strong>and</strong> Ludwig Luckner for<br />

coordination <strong>of</strong> the project work. Furthermore we would like to acknowledge the thoughtful<br />

reviews <strong>of</strong> the anonymous reviewers <strong>and</strong> the Editor. Their comments have greatly improved the<br />

manuscript.<br />

References<br />

C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

Aziz, C.E., Newell, C.J., Gonzales, A.R., Haas, P., Clement, T.P., Sun, Y., 2000. BIOCHLOR Natural Attenuation<br />

Decision Support System User's Manual. U.S. EPA, Office <strong>of</strong> Research <strong>and</strong> Development, EPA/600/R-00/008,<br />

Washington D.C.<br />

Bauer, S., Kolditz, O., 2005. Assessing contaminant mass <strong>flow</strong> rates by the integral groundwater investigation method by<br />

using the virtual aquifer approach. In: ModelCARE 2005, Fifth International Conference on Calibration <strong>and</strong> Reliability<br />

in Groundwater Modelling. From Uncertainty to Decision Making. Pre-published Proceedings, The Hague<br />

(Scheveningen), The Netherl<strong>and</strong>s, June 2005, pp.301–307.<br />

93


94 C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

Bauer, S., Beyer, C., Kolditz, O., 2005. Assessing measurements <strong>of</strong> first order degradation rates by using the virtual aquifer<br />

approach. In: Thomson, N.R. (Ed.), GQ2004, Bringing Groundwater Quality Research to the Watershed Scale.<br />

Proceedings <strong>of</strong> a Symposium held at Waterloo, Canada, July 2004. IAHS Publication, vol. 297. IAHS Press,<br />

Wallingford, pp. 274–281.<br />

Bauer, S., Beyer, C., Kolditz, O., 2006. Assessing measurement uncertainty <strong>of</strong> first-order degradation rates in<br />

heterogeneous aquifers. Water Resour. Res. 42, W01420. doi:10.1029/2004WR003878.<br />

Baveye, P., Valocchi, A., 1989. An evaluation <strong>of</strong> mathematical models <strong>of</strong> the transport <strong>of</strong> biologically reacting solutes in<br />

<strong>saturated</strong> soils <strong>and</strong> aquifers. Water Resour. Res. 25 (6), 1413–1421.<br />

Bear, J., 1972. Dynamics <strong>of</strong> Fluids in Porous Media. Elsevier, Amsterdam. 784 pp.<br />

Bekins, B.A., Warren, E., Godsy, E.M., 1998. A comparison <strong>of</strong> zero-order, first-order, <strong>and</strong> Monod biotransformation<br />

models. Ground Water 36 (2), 261–268.<br />

Beyer, C., Bauer, S., Kolditz, O., 2005. Uncertainty assessment <strong>of</strong> degradation rate measurements in heterogeneous media<br />

using the virtual aquifer approach. In: Kolditz, O., Bauer, S., Gronewold, J. (Eds.), Proceedings <strong>of</strong> the 5th Workshop<br />

“Porous Media”, Blaubeuren, Germany, December 2004. ZAG Publisher, Tübingen.<br />

Bockelmann, A., Zamfirescu, D., Ptak, T., Grathwohl, P., Teutsch, G., 2003. Quantification <strong>of</strong> mass fluxes <strong>and</strong> natural<br />

attenuation rates at an industrial site with a limited monitoring network: a case study. J. Contam. Hydrol. 60 (1–2), 97–121.<br />

Buscheck, T.E., Alcantar, C.M., 1995. Regression techniques <strong>and</strong> analytical solutions to demonstrate intrinsic<br />

bioremediation. In: Hinchee, R.E., Wilson, T.J., Downey, D. (Eds.), Intrinsic Bioremediation. Battelle Press,<br />

Columbus, OH, pp. 109–116.<br />

Chapelle, F.H., Bradley, P.M., Lovley, D.R., Vroblesky, D.A., 1996. Measuring rates <strong>of</strong> biodegradation in a contaminated<br />

aquifer using field <strong>and</strong> laboratory methods. Ground Water 34 (4), 691–698.<br />

Dagan, G., 1989. Flow <strong>and</strong> Transport in Porous Formations. Springer, Heidelberg. 465 pp.<br />

Domenico, P.A., 1987. An analytical model for multidimensional transport <strong>of</strong> a decaying contaminant species. J. Hydrol.<br />

91 (1–2), 49–59.<br />

Islam, J., Naresh, S., O'Sullivan, M., 2001. Modeling biogeochemical processes in leachate-contaminated soils: a review.<br />

Transp. Porous Media 43 (3), 407–440.<br />

Herfort, M., 2000. Reactive Transport <strong>of</strong> Organic Compounds Within a Heterogeneous Porous Aquifer, PhD Thesis,<br />

University <strong>of</strong> Tübingen, Tübinger Geowissenschaftliche Arbeiten (TGA), 54, Tübingen. 76 pp.<br />

Kolditz, O., 2002. Computational Methods in Environmental Fluid Dynamics. Springer, Heidelberg. 400 pp.<br />

Kolditz, O., Bauer, S., 2004. A process-oriented approach to computing multifield problems in porous media. J. Hydroinform.<br />

6(3),225–244.<br />

Kolditz, O., de Jonge, J., Beinhorn, M., Xie, M., Kalbacher, T., Wang, W., Bauer, S., McDermott, C., Chen, C., Beyer, C.,<br />

Gronewold, J., Kemmler, D., Manabe, T., Legeida, D., Adamidis, P., 2004. GeoSys- Theory <strong>and</strong> users manual. release<br />

4.1. GeoSystems Research, Center for <strong>Applied</strong> Geoscience, University <strong>of</strong> Tübingen, http://www.uni-tuebingen.de/zag/<br />

geohydrology, Tübingen.<br />

McNab Jr., W.W., 2001. A Monte Carlo simulation method for assessing biotransformation effects on groundwater fuel<br />

hydrocarbon plume lengths. Comput. Geosci. 27 (1), 31–42.<br />

McNab Jr., W.W., Dooher, B.P., 1998. A critique <strong>of</strong> a steady-state analytical method for estimating contaminant<br />

degradation rates. Ground Water 36 (6), 983–987.<br />

Newell, C.J., McLeod, R.K., Gonzales, J.R., 1996. BIOSCREEN Natural Attenuation Decision Support System User's<br />

Manual. Version 1.3. U.S. EPA, Office <strong>of</strong> Research <strong>and</strong> Development, EPA/600/R-96/087, Washington D.C.<br />

Newell, C.J., Rifai, H.S., Wilson, J.T., Connor, J.A., Aziz, J.A., Suarez, M.P., 2002. Calculation <strong>and</strong> use <strong>of</strong> first-order rate<br />

constants for monitored natural attenuation studies. U.S. EPA Ground Water Issue, U.S. EPA, Office <strong>of</strong> Research <strong>and</strong><br />

Development, EPA/540/S-02/500, Washington D.C.<br />

Parlange, J.-Y., Starr, J.L., Barry, D.A., Braddock, R.D., 1984. Some approximate solutions <strong>of</strong> the transport equation with<br />

irreversible reactions. Soil Sci. 137, 434–442.<br />

Rehfeldt, K.R., Boggs, J.M., Gelhar, L.W., 1992. Field study <strong>of</strong> dispersion in a heterogeneous aquifer, 3, geostatistical<br />

analysis <strong>of</strong> hydraulic conductivity. Water Resour. Res. 28 (12), 3309–3324.<br />

Rifai, H.S., Newell, C.J., Gonzales, J.R., Dendrou, S., Kennedy, L., Wilson, J.T., 1998. BIOPLUME III Natural<br />

Attenuation Decision Support System Version 1.0 User's Manual. U.S. EPA, Office <strong>of</strong> Research <strong>and</strong> Development,<br />

EPA/600/R-98/010, Washington D.C.<br />

Rittmann, B., VanBriesen, J.M., 1996. Microbiological processes in reactive <strong>modeling</strong>. In: Lichtner, P., Steefel, C.,<br />

Oelkers, E. (Eds.), Reactive Transport in Porous Media. Reviews in Mineralogy, vol. 34. Mineralogical Society <strong>of</strong><br />

America, Washington, DC, pp. 311–334.<br />

Robinson, J.A., 1985. Determining microbial kinetic parameters using nonlinear regression analysis. Adv. Microb. Ecol. 8,<br />

61–114.


C. Beyer et al. / Journal <strong>of</strong> Contaminant Hydrology 87 (2006) 73–95<br />

Schäfer, D., Dahmke, A., Kolditz, O., Teutsch, G., 2002. “Virtual Aquifers”: a concept for evaluation <strong>of</strong> exploration,<br />

remediation <strong>and</strong> monitoring strategies. In: Kovar, K., Hrkal, Z. (Eds.), Calibration <strong>and</strong> Reliability in Groundwater<br />

Modelling: A Few Steps Closer to Reality. Proceedings <strong>of</strong> the ModelCARE 2002 Conference, Prague, Czech<br />

Republic, June 2002. IAHS Publication, vol. 277. IAHS Press, Oxfordshire, pp. 52–59.<br />

Schäfer, D., Manconi, A., Dahmke, A., 2004a. Identification <strong>and</strong> consequences <strong>of</strong> different kinetic approaches for simulation<br />

<strong>of</strong> microbial degradation. In: Aagaard, P., Bedbur, E., Bidoglio, G., C<strong>and</strong>ela, L., Nuetzmann, G., Trevisan, M., Vanclooster,<br />

M., Viotti, P. (Eds.), Proceedings <strong>of</strong> the International Workshop: Saturated <strong>and</strong> Un<strong>saturated</strong> Zone, Integration <strong>of</strong> Process<br />

Knowledge into Effective Models. University <strong>of</strong> Rome La Sapienza“, Rome, Italy, May 2004, pp. 291–296.<br />

Schäfer, D., Schlenz, B., Dahmke, A., 2004b. Evaluation <strong>of</strong> exploration <strong>and</strong> monitoring methods for verification <strong>of</strong> natural<br />

attenuation using the virtual aquifer approach. Biodegradation 15 (6), 453–465.<br />

Simkins, S., Alex<strong>and</strong>er, M., 1984. Models for mineralization kinetics with variables <strong>of</strong> substrate concentration <strong>and</strong><br />

population density. Appl. Environ. Microbiol. 47 (6), 1299–1306.<br />

Stenback, G.A., Ong, S.K., Rogers, S.W., Kjartanson, B.H., 2004. Impact <strong>of</strong> transverse <strong>and</strong> longitudinal dispersion on<br />

first-order degradation rate constant estimation. J. Contam. Hydrol. 73, 3–14.<br />

Suarez, M.P., Rifai, H.S., 2002. Evaluation <strong>of</strong> BTEX remediation by natural attenuation at a coastal facility. Ground Water<br />

Monit. Remediat. 22 (1), 62–77.<br />

Suarez, M.P., Rifai, H.S., 2004. Modeling natural attenuation <strong>of</strong> total BTEX <strong>and</strong> benzene plumes with different kinetics.<br />

Ground Water Monit. Remediat. 24 (3), 53–68.<br />

Sudicky, E.A., 1986. A natural gradient experiment on solute transport in a s<strong>and</strong> aquifer: spatial variability <strong>of</strong> hydraulic<br />

conductivity <strong>and</strong> its role in the dispersion process. Water Resour. Res. 22 (13), 2069–2082.<br />

U.S. EPA, 1998. Technical protocol for evaluating natural attenuation <strong>of</strong> chlorinated solvents in groundwater, U.S. EPA,<br />

Office <strong>of</strong> Research <strong>and</strong> Development, EPA/600/R/128, Washington, D.C.<br />

Wiedemeier, T.H., Swanson, M.A., Wilson, J.T., Kampbell, D.H., Miller, R.N., Hansen, J.E., 1996. Approximation <strong>of</strong><br />

biodegradation rate constants for monoaromatic hydrocarbons (BTEX) in ground water. Ground Water Monit.<br />

Remediat. 16 (3), 186–194.<br />

Wiedemeier, T.H., Rifai, H.S., Wilson, T.J., Newell, C., 1999. Natural Attenuation <strong>of</strong> Fuels <strong>and</strong> Chlorinated Solvents in the<br />

Subsurface. Wiley, New York. 632 pp.<br />

Wilson, J.T., Kolhatkar, R., 2002. Role <strong>of</strong> natural attenuation in life cycle <strong>of</strong> MTBE plumes. J. Environ. Eng. 128 (9),<br />

876–882.<br />

Wilson, J.T., Pfeffer, F.M., Weaver, J.W., Kampbell, D.H., Wiedemeier, T.H., Hansen, J.E., Miller, R.N., 1994. Intrinsic<br />

bioremediation <strong>of</strong> JP-4 jet fuel. Symposium on Intrinsic Bioremediation <strong>of</strong> Ground Water, Denver, Colorado, EPA/<br />

540/R-94/515, pp. 60–72.<br />

Wilson, R.D., Thornton, S.F., Mackay, D.M., 2004. Challenges in monitoring the natural attenuation <strong>of</strong> spatially variable<br />

plumes. Biodegradation 15 (6), 459–469.<br />

Zamfirescu, D., Grathwohl, P., 2001. Occurrence <strong>and</strong> attenuation <strong>of</strong> specific organic compounds in the groundwater plume<br />

at a former gasworks site. J. Contam. Hydrol. 53 (1–2), 407–427.<br />

Zhang, Y.-K., Heathcote, R.C., 2003. An improved method for estimation <strong>of</strong> biodegradation rate with field data. Ground<br />

Water Monit. Remediat. 23 (3), 112–116.<br />

95


Enclosed Publication 4<br />

Bauer, S., Beyer, C., Kolditz, O. (2007): Einfluss von Heterogenität und Messfehler auf die<br />

Bestimmung von Abbauraten erster Ordnung - eine Virtueller Aquifer Szenarioanalyse.<br />

(Influence <strong>of</strong> heterogeneity <strong>and</strong> measurement error on the determination <strong>of</strong> first order<br />

degradation rates by us1ing the virtual aquifer approach.), Grundwasser, 12, 3–14, doi:<br />

10.1007/s00767-007-0019-8<br />

The enclosed article is made available with the permission <strong>of</strong> Springer <strong>and</strong> was published in<br />

the journal Grundwasser, 12, 3-14 (2007). Copyright © 2007 Springer.<br />

The article can be obtained online via SpringerLink at<br />

http://www.springerlink.com/openurl.asp?genre=journal&eissn=1432-1165.


Einfluss von Heterogenität und Messfehler auf die<br />

Bestimmung von Abbauraten erster Ordnung – eine<br />

Virtueller Aquifer Szenarioanalyse ∗<br />

Influence <strong>of</strong> heterogeneity <strong>and</strong> measurement error on the determination<br />

<strong>of</strong> first order degradation rates - a virtual aquifer scenario analysis<br />

Sebastian Bauer, Christ<strong>of</strong> Beyer, Olaf Kolditz<br />

Zentrum für Angew<strong>and</strong>te Geowissenschaften, Universität Tübingen, Sigwartstraße 10,<br />

D 72076 Tübingen. Tel: 07071-2973171. Fax: 07071 5059<br />

e-mail: sebastian.bauer@uni-tuebingen.de, christ<strong>of</strong>.beyer@uni-tuebingen.de, kolditz@uni-tuebingen.de<br />

Header: Heterogenität und Messfehler bei Abbauratenbestimmung<br />

Kurzfassung<br />

Die grundlegende Idee des Virtuellen Aquifers ist, durch numerische Modellierung von typischen<br />

Schadensfällen Erkundungsstrategien zu simulieren und zu bewerten. In diesem Beitrag wird die<br />

Bestimmung von Abbauraten erster Ordnung untersucht. Eine Schadensquelle wird dabei in einen<br />

virtuellen Aquifer eingebracht und die stationäre Schadst<strong>of</strong>ffahne unter der Annahme einer<br />

Abbaukinetik erster Ordnung simuliert. Diese Fahne wird dann durch Beobachtungspegel entlang der<br />

Zentrallinie der Fahne untersucht. Anh<strong>and</strong> von vier typischen Methoden werden Abbauraten erster<br />

Ordnung berechnet und mit dem vorgegebenen Wert verglichen. Dieser Vergleich wird für<br />

unterschiedlich stark ausgeprägte hydraulische Heterogenitäten durchgeführt. Dabei zeigt sich, dass<br />

mit zunehmender Heterogenität die ermittelten Abbauraten die tatsächliche Abbaurate um Größenordnungen<br />

überschätzen können und sie somit sehr unsicher sind. Bei der Untersuchung der<br />

Messfehler wurden Abweichungen bei der Bestimmung der Piezometerhöhe und der Konzentration<br />

angenommen. Hierbei ergibt sich, dass Messfehler ebenfalls zu einer hohen Unsicherheit der<br />

Ratenkonstante führen können, wobei Messfehler der Piezometerhöhe einen stärkeren Einfluss<br />

haben.<br />

Abstract<br />

The principal idea <strong>of</strong> the Virtual Aquifer is to simulate <strong>and</strong> evaluate monitoring strategies <strong>and</strong><br />

remediation options for contaminated sites by modelling <strong>of</strong> typical contamination scenarios. Here the<br />

determination <strong>of</strong> first order degradation rates is studied. A virtual reality is generated by simulating the<br />

spreading <strong>of</strong> a plume, originating from a defined source <strong>and</strong> subject to first order degradation. This<br />

plume is investigated using monitoring wells placed along the plume center line. From the information<br />

thus obtained first order degradation rates are calculated by methods typically used <strong>and</strong> are then<br />

compared to the predefined value. This comparison is conducted for different degrees <strong>of</strong><br />

heterogeneity. It is found that with increasing heterogeneity degradation rates overestimate the real<br />

degradation rate by up to orders <strong>of</strong> magnitude <strong>and</strong> show a high uncertainty. Then measurement errors<br />

are introduced for piezometric head <strong>and</strong> concentration measurements. It is found that deviations <strong>of</strong> the<br />

estimated first order rate constant from the true one <strong>of</strong> up to orders <strong>of</strong> magnitude can occur, with<br />

errors <strong>of</strong> the piezometric head measurement causing the dominant uncertainty.<br />

Keywords: natural attenuation, heterogeneity, first-order degradation rate, virtual aquifer, scenario<br />

analysis<br />

∗ Bauer, S., Beyer, C., Kolditz, O. (2007): Einfluss von Heterogenität und Messfehler auf die Bestimmung von Abbauraten erster<br />

Ordnung - eine Virtueller Aquifer Szenarioanalyse. (Influence <strong>of</strong> heterogeneity <strong>and</strong> measurement error on the determination <strong>of</strong><br />

first order degradation rates by us1ing the virtual aquifer approach.), Grundwasser, 12, 3–14, doi: 10.1007/s00767-007-0019-8.<br />

Der Artikel wurde in der Zeitschrift Grundwasser publiziert und mit Erlaubnis von Springer reproduziert. Copyright © 2007 Springer.<br />

Der Artikel ist online abrufbar via SpringerLink: http://www.springerlink.com/openurl.asp?genre=journal&eissn=1432-1165.<br />

1


Einführung<br />

In dieser Arbeit wird eine „Virtueller Aquifer“ Szenarioanalyse verwendet, um die Unsicherheit bei der<br />

Bestimmung von Abbauratenkonstanten erster Ordnung anh<strong>and</strong> von Messstellen auf der Zentrallinie<br />

der Schadst<strong>of</strong>ffahne („Center line method“) im Kontext von Natural Attenuation zu untersuchen und zu<br />

quantifizieren. Natural Attenuation (NA), auch als Bioremediation bekannt, bezieht sich auf die Abnahme<br />

von Schadst<strong>of</strong>fkonzentrationen durch natürliche Abbauprozesse mit zunehmendem Abst<strong>and</strong> von<br />

der Quelle (US-EPA, 1999; WIEDEMEIER et al., 1999). Dabei werden Dispersion, Verdünnung,<br />

Sorption, Ausgasung und Bioabbau betrachtet, wobei der Bioabbau der einzige Prozess ist, der zu<br />

einer Verringerung der Schadst<strong>of</strong>fmasse führt. Die an einem St<strong>and</strong>ort ablaufenden Prozesse müssen<br />

sorgfältig charakterisiert werden, um Aussagen über NA treffen zu können. Hierbei können<br />

insbesondere die Abbauraten der betrachteten Schadst<strong>of</strong>fe für mögliche Sanierungen und das<br />

St<strong>and</strong>ortmanagement eine Rolle spielen. Die Abbauraten werden verwendet, um das gesamte NA-<br />

Potential am St<strong>and</strong>ort zu charakterisieren, um die Länge von Schadst<strong>of</strong>ffahnen in der Zukunft zu<br />

prognostizieren und um unterstromige Konzentrationen zu berechnen, die für eine Auswirkungsprognose<br />

benötigt werden (WIEDEMEIER et al., 1999).<br />

Zur Bestimmung von Abbauraten im Feld stehen derzeit mehrere Methoden zur Verfügung, wie z.B.<br />

Massenbilanzen, in-situ Mikrokosmosstudien oder die Verwendung von Konzentrations-Abst<strong>and</strong>s-<br />

Beziehungen, die auf der Fahnenzentrallinie einer stationären Schadst<strong>of</strong>ffahne ermittelt wurden<br />

(CHAPELLE, 1996). Für letztere sind in der Literatur vier unterschiedliche Methoden beschrieben. Die<br />

erste beruht auf der eindimensionalen Transportgleichung mit Advektion und Abbau erster Ordnung<br />

(WIEDEMEIER et al., 1996). Die zweite Methode ist eine Erweiterung der ersten, indem die<br />

Konzentrationen des betrachteten Schadst<strong>of</strong>fes auf die Konzentrationen eines nichtreaktiven Mitkontamin<strong>and</strong>en<br />

normiert werden (WIEDEMEIER et al., 1996, 1999; WILSON et. al., 1994). Die dritte<br />

Methode, von BUSCHECK & ALCANTAR (1995) vorgeschlagen, beruht auf der eindimensionalen<br />

Transportgleichung mit Advektion, Dispersion und Abbau erster Ordnung und wurde bereits an einigen<br />

St<strong>and</strong>orten eingesetzt (CHAPELLE et al., 1996; WIEDEMEIER et al., 1996; ZAMFIRESCU & GRATHWOHL,<br />

2001; SUAREZ & RIFAI, 2002; BOCKELMANN et al., 2003). In letzter Zeit wurden als Erweiterung zur<br />

Methode von BUSCHECK & ALCANTAR (1995) zwei- und dreidimensionale Methoden entwickelt (ZHANG<br />

& HEATHCOTE, 2003; STENBACK et al., 2004). ZHANG & HEATHCOTE (2003) konnten zeigen, dass die<br />

eindimensionale Methode durch Vernachlässigung der Querdispersion die Abbauraten um 21% im<br />

Vergleich zum zweidimensionalen und um 65% im Vergleich zum dreidimensionalen Fall überschätzt.<br />

MCNAB Jr. & DOOHER (1998) beschrieben, wie transversale Dispersion und instationäre Strömungszustände<br />

Konzentrationsverteilungen erzeugen können, die durch die Methode von BUSCHECK &<br />

ALCANTAR (1995) dann fälschlich als Bioabbau klassifiziert werden können, obwohl am St<strong>and</strong>ort kein<br />

Abbau stattfindet.<br />

Aufgrund der eingeschränkten Zugänglichkeit des Untergrundes sind Beobachtungen an<br />

kontaminierten St<strong>and</strong>orten nur an einzelnen räumlichen Punkten möglich. Die aus einzelnen<br />

Messpunkten abgeleiteten Ergebnisse und Aussagen sind daher immer mit Unsicherheit behaftet, die<br />

das eingeschränkte und punktuelle Wissen über den St<strong>and</strong>ort widerspiegelt. Eine „Virtuelle Aquifer“<br />

Szenarioanalyse kann verwendet werden, um diese Unsicherheit näher zu betrachten und zu quantifizieren<br />

(SCHÄFER et al., 2002; SCHÄFER et al., 2004; BAUER et al., 2006; BEYER et al., 2005a; BAUER &<br />

KOLDITZ, 2005). Eine Untersuchung des Einflusses der Parametrisierung der Abbaukinetik auf die<br />

prognostizierte Fahnenlänge wird in SCHÄFER et al. (2005) durchgeführt. Dabei werden numerische<br />

synthetische Modelle typischer Aquifere generiert. Mithilfe eines reaktiven Transportmodells können<br />

dann typische Schadensszenarien, wie beispielsweise die Ausbreitung einer Schadst<strong>of</strong>ffahne ausgehend<br />

von einer Schadst<strong>of</strong>fquelle, simuliert werden. Der große Vorteil der Virtuellen Aquifere ist, dass<br />

die so erhaltene realistische räumliche Verteilung aller Parameter, wie z.B. Piezometerhöhe oder<br />

Konzentration, exakt bekannt ist. Die virtuelle Schadst<strong>of</strong>fahne wird dann in einem zweiten Schritt<br />

durch typische Erkundungsstrategien (hier: Zentrallinienmethode) untersucht. Eine Messung im<br />

virtuellen Aquifer bedeutet, die entsprechenden Werte aus der Modellausgabedatei zu lesen. Bei<br />

dieser Erkundung der virtuellen Schadst<strong>of</strong>ffahne werden nur die Messwerte (hier: Piezometerhöhe,<br />

Konzentration) berücksichtigt, die auch an einem echten St<strong>and</strong>ort erhalten werden können.<br />

Ausgehend von diesen (virtuellen) Messwerten werden dann weitere Parameter ermittelt (hier:<br />

hydraulische Durchlässigkeiten, Abbauraten). Bei diesem zweiten Schritt ist also die richtige<br />

Parameterverteilung nicht bekannt. Indem das Ergebnis der virtuellen Erkundung im dritten Schritt mit<br />

der wahren Parameterverteilung (hier: Abbauratenkoeffizient) verglichen wird, können die<br />

verwendeten Untersuchungsmethoden getestet und bewertet werden.<br />

Die räumliche Heterogenität von Aquiferparametern hat einen erheblichen Einfluss auf die<br />

Fahnenentwicklung und die resultierende Konzentrationsverteilung eines Schadst<strong>of</strong>fes im Aquifer. Die<br />

2


oben erwähnten Methoden zur Bestimmung von Abbauratenkonstanten erster Ordnung unterliegen<br />

daher dem Einfluss der Heterogenität, da sie auf Messdaten entlang der vermuteten Fahnenzentrallinie<br />

und auf Abschätzungen der Dispersivität beruhen. Wird die Grundwasserfließrichtung<br />

falsch abgeschätzt, kann die Fahnenzentrallinie leicht verfehlt werden (WILSON et al., 2004). Bisher<br />

existiert in der Literatur keine Studie, die den Einfluss der Heterogenität auf die Ermittlung von<br />

Abbauratenkonstanten erster Ordnung untersucht. Bei der Probenahme im realen Fall können darüber<br />

hinaus noch Messfehler auftreten, die ebenfalls die Ratenkonstanten beeinflussen. Diese können<br />

sowohl bei der Pegeleinmessung, bei der Bestimmung der hydraulischen Leitfähigkeit, bei Wasserst<strong>and</strong>smessungen<br />

und insbesondere bei Konzentrationsmessungen auftreten. Auch hierzu existieren<br />

noch keine Untersuchungen in der Literatur. Ziel dieser Arbeit ist daher a) die Genauigkeit von<br />

Abbauraten, wie sie aus Feldmessungen gewonnen werden, zu ermitteln, und b) die verschiedenen<br />

Methoden zu Bestimmung von Abbauraten zu bewerten. Dazu wird der Einfluss von hydraulischer<br />

Heterogenität und Messfehlern anh<strong>and</strong> einer „Virtueller Aquifer“ Szenarioanalyse untersucht. Hierzu<br />

werden in einem synthetischen Aquifer mit unterschiedlich stark ausgeprägter Heterogenität virtuelle<br />

stationäre Fahnen simuliert, die einer Abbaukinetik erster Ordnung unterliegen. Die anh<strong>and</strong> der oben<br />

genannten vier Methoden bestimmten Abbauratenkonstanten erster Ordnung werden mit dem wahren<br />

Wert verglichen und so Aussagen über die Zuverlässigkeit und Unsicherheit der Methoden und der<br />

resultierenden Ratenkonstanten abgeleitet. Bei der Untersuchung des Einflusses des Messfehlers auf<br />

die Ratenkonstante wird anschließend von einem homogenen Aquifer ausgegangen. Der Einfluss von<br />

Messfehlern bei der Bestimmung der Piezometerhöhe und bei der Konzentrationsmessung wird<br />

separat untersucht und quantifiziert.<br />

Methodik<br />

Das verwendete Modellgebiet ist zweidimensional und misst 184 m auf 64 m (vergl. Abb. 1). Das<br />

Grundwasser strömt von links nach rechts, der mittlere hydraulische Gradient beträgt 0.003 und wird<br />

durch Festpotentialr<strong>and</strong>bedingungen am linken und rechten Modellr<strong>and</strong> erzeugt. Alle <strong>and</strong>eren Ränder<br />

sind undurchlässig. Das Modellgebiet ist mit einer Gitterweite von 0.5 m regelmäßig diskretisiert. 11.5<br />

m vom oberstromigen R<strong>and</strong> ist eine Schadst<strong>of</strong>fquelle, durch eine Festkonzentrationsr<strong>and</strong>bedingung<br />

dargestellt, eingebracht, die einen Schadst<strong>of</strong>f emittiert, der einer Abbaukinetik erster Ordnung folgend<br />

mit einer Ratenkonstanten λ abgebaut wird. Zusätzlich wird ein nichtreaktiver St<strong>of</strong>f aus der Quelle<br />

freigesetzt. Es werden Quellbreiten von 4, 8 und 16 m betrachtet. Weder Sorption noch Ausgasung<br />

werden berücksichtigt. Der Schadst<strong>of</strong>f verhält sich also genau so, wie es von den Methoden zur<br />

Bestimmung der Abbaurate angenommen wird – eine Annahme, die in der Realität nur annähernd<br />

erfüllt ist, da die Abbauraten komplexeren Gesetzmäßigkeiten folgen. Diese Abstraktion ist hier<br />

notwendig, um die vier Methoden möglichst genau untersuchen zu können. Als longitudinale und<br />

transversale Dispersionslängen werden im Modell 0.25 m und 0.05 m angenommen. Diese stellen die<br />

lokale Dispersion dar, in die effektive Dispersion geht noch die Heterogenität der hydraulischen<br />

Durchlässigkeit ein, die hier im Modell explizit dargestellt wird. Mit Hilfe des Programms GeoSys<br />

(KOLDITZ et al., 2005) wird dann eine stationäre Schadst<strong>of</strong>ffahne erzeugt. Um den Einfluss räumlicher<br />

Heterogenität untersuchen zu können, wird die hydraulische Durchlässigkeit K als eine ln –<br />

normalverteilte Zufallsvariable aufgefasst mit einem Erwartungswert von ln(K) = -9.54, was einer<br />

mittleren hydraulischen Durchlässigkeit von 7.2 10 -5 m s -1 entspricht. Mit der verwendeten Porosität<br />

von 0.33 ergibt sich so eine mittlere Transportgeschwindigkeit von 6.5·10 -7 m s -1 . Die räumliche<br />

Struktur wird durch ein isotropes exponentielles Covarianzmodell Cln(K) = σ 2 exp (-∆h/l) abgebildet,<br />

wobei eine integrale Länge von l = 2.67 m verwendet wird, was einer Korrelationslänge von 8.0 m<br />

entspricht (RUBIN, 2003). Vier Klassen von hydraulischer Heterogenität werden betrachtet, die durch<br />

ln(K)-Varianzen σ 2 von 0.38, 1.71, 2.70 und 4.50 gegeben sind und das Spektrum von wenig bis stark<br />

heterogenen Bedingungen abdecken. Der Wert σ 2 = 0.38, das Covarianzmodell, die mittlere<br />

hydraulische Durchlässigkeit und Porosität als auch die integrale Länge l entsprechen dem St<strong>and</strong>ort<br />

Borden (SUDICKY, 1986). Der Wert 1.71 wurde in einem alluvialen Aquifer in Süddeutschl<strong>and</strong> bestimmt<br />

(HERFORT, 2000), während die Werte 2.70 und 4.50 von Untersuchungen der Columbus Air Force<br />

Base stammen (REHFELDT et al., 1992). Die geostatistische S<strong>of</strong>tware gstats2.4 (PEBESMA &<br />

WESSELING, 1998) wurde zur Generierung der Zufallsverteilungen der hydraulischen Durchlässigkeit<br />

durch unkonditionierte Gauss’sche Simulation verwendet.<br />

Die so erzeugten Fahnen werden nun anh<strong>and</strong> der Zentrallinien-Methode untersucht. Dabei sind<br />

anfänglich drei Beobachtungsbrunnen vorgegeben, von denen einer direkt in der Schadst<strong>of</strong>fquelle<br />

liegt, während die <strong>and</strong>eren keine Kontamination zeigen (Abb.1 a). An diesen drei Brunnen werden nun<br />

die Piezometerhöhen „gemessen“, indem die Modellausgabe gelesen wird. Durch Konstruktion eines<br />

hydrogeologischen Dreiecks wird die lokale Fließrichtung an der Schadensquelle bestimmt (Abb.1b).<br />

3


Entlang der so bestimmten Grundwasserfließrichtung werden nun drei neue Messstellen im Abst<strong>and</strong><br />

von je 10 m installiert. An diesen Messstellen werden dann Piezometerhöhen, Konzentrationen des<br />

reaktiven und des nichtreaktiven Schadst<strong>of</strong>fes und die lokale hydraulische Durchlässigkeit „gemessen“<br />

(Abb. 1c). Aus der Differenz der Piezometerhöhen, den hydraulischen Durchlässigkeiten und der<br />

wahren, d.h. der auch im Modell verwendeten, Porosität wird die Grundwasserfließgeschwindigkeit<br />

bestimmt. Zusammen mit den gemessenen Schadst<strong>of</strong>fkonzentrationen können dann Abbauraten<br />

erster Ordnung anh<strong>and</strong> von vier Methoden berechnet werden. Der Virtuelle Aquifer und das Messszenario<br />

sind so ausgelegt, dass optimale Bedingungen für eine Bestimmung der Abbauraten<br />

vorliegen. Die einzige Unsicherheit wird durch die heterogene Verteilung der hydraulischen<br />

Durchlässigkeit erzeugt.<br />

4<br />

Anfangssituation<br />

Erkundungsschritt 1<br />

Erkundungsschritt 2<br />

Virtuelle Realität<br />

Abb. 1 Verwendete Methodik, um Konzentrationen von der Zentrallinie der Fahne zu erhalten: a)<br />

Anfangszust<strong>and</strong>, b) Abschätzung der Fließrichtung anh<strong>and</strong> des hydrogeologischen Dreiecks, c)<br />

„gemessene“ Konzentrationen auf der ermittelten Zentrallinie, d) Vergleich mit der Virtuellen Realität<br />

(Piezometerhöhen und Fahne).<br />

Für den zweiten Teil dieses Beitrags, in dem der Einfluss von Messfehlern untersucht wird, wird der<br />

Aquifer als homogen angenommen. Die Unsicherheit wird nun also nicht durch das heterogene<br />

Fließfeld erzeugt, sondern durch fehlerhaftes „Messen“ von Piezometerhöhe und Schadst<strong>of</strong>fkonzentration.<br />

Im Falle der Piezometerhöhe lautet das Fehlermodell:<br />

[1] h′ = h + z∆hmax<br />

wobei h’ die fehlerbehaftete, hh die wahre Piezometerhöhe und ∆hmax der maximale Messfehler für die<br />

Piezometerhöhe ist. z ist eine gleichverteilte Zufallszahl aus dem Intervall [-1, 1], mit der ∆hmax


multipliziert wird, um einen fehlerbehafteten Messwert für die Piezometerhöhe aus dem Intervall [h-<br />

∆hmax, h+∆hmax ] zu erhalten. Die fehlerbehafteten Messwerte streuen also symmetrisch um den<br />

wahren Wert. ∆hmax wird zwischen 0 und 5 cm variiert. Bei der Ermittlung der Ratenkonstante wird nun<br />

bei der Probenahme für die Piezometerhöhe ein Messfehler gemäß des vorgestellten Fehlermodells<br />

berücksichtigt. Die gesamte Auswertung wird einhundert mal durchgeführt, um statistisch<br />

repräsentative Aussagen über den Einfluss des Messfehlers auf die Ratenkonstante zu erlangen.<br />

Da für die Messung von Konzentrationen ein höherer Messfehlerbereich zu erwarten ist, wurde das<br />

Fehlermodell angepasst:<br />

a<br />

[2] c ′ = c z∆c<br />

)<br />

5<br />

( max<br />

wobei c’ die fehlerbehaftete, ch die wahre Konzentration und ∆cmax der maximale Messfehlerfaktor für<br />

die Konzentration ist. Der Exponent a ist –1 oder 1 und wird zufällig bestimmt, z ist eine gleichverteilte<br />

Zufallszahl aus dem Intervall [0, 1], mit der ∆cmax multipliziert wird. Man erhält so einen fehlerbehaftet<br />

Messwert für die Konzentration aus dem Intervall [c/∆cmax, c∆cmax ]. ∆cmax wird zwischen 1 und 100<br />

variiert. Dieses Verfahren ist ähnlich zu dem für die Piezometerhöhe angewendeten Verfahren und ein<br />

Messfehlerfaktor von 2 entspricht dem umgangsprachlichen „auf einen Faktor 2 genau“. Für die<br />

Bestimmung des Einflusses von Konzentrationsmessfehlern wurden ein homogenes Strömungsfeld<br />

und keine Messfehler für die Piezometerhöhe angenommen. Für den konservativen und den reaktiven<br />

St<strong>of</strong>f wird dasselbe Messfehlermodell angenommen.<br />

Die vier Methoden, anh<strong>and</strong> derer die Abbauraten bestimmt werden, sind in Tabelle 1 aufgeführt.<br />

Methode 1 stellt die Lösung zur eindimensionalen Advektionsgleichung mit Abbau erster Ordnung dar.<br />

Methode 2 wurde von WIEDEMEIER et al. (1996) vorgeschlagen, und baut auf Methode 1 auf.<br />

Allerdings werden die Konzentrationen des reaktiven St<strong>of</strong>fes auf die Konzentrationen des nichtreaktiven<br />

St<strong>of</strong>fes bezogen. Dadurch berücksichtigt Methode 2 Dispersion, Verdünnung und „Aus der<br />

Fahne Messen“ an Beobachtungspegeln, die nicht genau auf der Zentrallinie der Fahne liegen. Die<br />

dritte Methode wurde von BUSCHECK & ALCANTAR (1995) vorgestellt und basiert auf der eindimensionalen<br />

Transportgleichung mit Advektion, Dispersion und Abbau erster Ordnung. Diese Methode<br />

beinhaltet somit explizit die longitudinale Dispersion. ZHANG & HEATHCOTE (2003) beschrieben die<br />

vierte Methode, die auf der analytischen Lösung zur zweidimensionalen Transportgleichung beruht<br />

und zusätzlich zur longitudinalen auch die transversale Dispersion berücksichtigt.<br />

Tab. 1 Methoden zur Bestimmung von Abbauratenkonstanten erster Ordnung λ. va ist die<br />

Transportgeschwindigkeit, ∆x ist der Abst<strong>and</strong> der verwendeten Beobachtungspegel, C(x) ist die unterstromige<br />

und C0 die Quellkonzentration. αL und αT sind die longitudinale und transversale Dispersivität, WS ist die<br />

Quellbreite und erf ist die Fehlerfunktion.<br />

Methode Formel für Abbauratenkonstante Beschreibung<br />

1 ⎟ v ⎛ ⎞<br />

a C(<br />

x)<br />

λ = −<br />

⎜<br />

1 ln<br />

∆x<br />

⎝ C0<br />

⎠<br />

2<br />

3<br />

4<br />

v ⎛ ∗ ⎞<br />

a ⎜ C(<br />

x)<br />

C0<br />

λ = −<br />

⎟<br />

2 ln<br />

∆x<br />

⎜ C ∗ ⎟<br />

⎝ 0 C(<br />

x)<br />

⎠<br />

( C(<br />

x)<br />

C )<br />

2<br />

v ⎛<br />

⎞<br />

a ⎜⎛<br />

ln<br />

0 ⎞<br />

λ<br />

− ⎟<br />

3 =<br />

⎜<br />

⎜1−<br />

2α<br />

L<br />

⎟ 1<br />

4α<br />

⎟<br />

L ⎝⎝<br />

∆x<br />

⎠ ⎠<br />

( C(<br />

x)<br />

( C β ) )<br />

2<br />

v ⎛<br />

⎞<br />

a ⎜⎛<br />

ln<br />

0 ⎞<br />

λ<br />

− ⎟<br />

3 =<br />

⎜<br />

⎜1−<br />

2α<br />

L<br />

⎟ 1<br />

4α<br />

⎟<br />

L ⎝⎝<br />

∆x<br />

⎠ ⎠<br />

mit:<br />

⎛ ⎞<br />

⎜<br />

WS<br />

β = erf ⎟<br />

⎜ ⎟<br />

⎝ 4 αT<br />

∆x<br />

⎠<br />

Analytische Lösung der 1D<br />

Advektionsgleichung mit Abbau erster<br />

Ordnung<br />

Wie Methode 1, aber Konzentration normiert<br />

auf einen nichtreaktiven Mitkontamin<strong>and</strong>,<br />

berücksichtigt Verdünnung und Dispersion<br />

Analytische Lösung der 1D Transportgleichung<br />

mit longitudinaler Dispersion<br />

Analytische Lösung der 2D Transport-<br />

gleichung. Berücksichtigt sind longitudinale<br />

und transversale Dispersion und die<br />

Quellbreite


Für die letztgenannten Methoden müssen die longitudinale und die transversale Dispersivität bekannt<br />

sein. Gemäß einem Ansatz in WIEDEMEIER et al. (1999) wurde die longitudinale Dispersivität als 10%<br />

der Fahnenlänge angenommen, die transversale Dispersivität beträgt 33% der longitudinalen<br />

Dispersivität. Als Fahnenlänge wurde der Abst<strong>and</strong> von der Quelle zur entferntesten Messstelle<br />

angenommen. Dieser Abst<strong>and</strong> beträgt 30 m, die longitudinale Dispersivität somit 3 m und die<br />

Transversale Dispersivität 1 m. Die Dispersivitäten sind damit recht gering geschätzt, eine detaillierte<br />

Untersuchung des Einflusses der angenommenen Dispersivitäten ist in BAUER et al. (2005) zu finden.<br />

Mithilfe der vier vorgestellten Ansätze werden Abbauraten erster Ordnung ermittelt und mit dem<br />

Modelleingabewert verglichen. Für jede Klasse der hydraulischen Heterogenität werden je 100<br />

Realisierungen betrachtet, um ein statistisches Maß für die Unsicherheit zu erhalten, die durch die<br />

hydraulische Heterogenität erzeugt wird. Für jede Realisierung wurde die oben beschriebene<br />

Auswertung für jede Methode und jede Quellbreite durchgeführt. Dabei wurden jeweils die Abbauraten<br />

erster Ordnung ausgehend von der Quelle für die unterstromigen Brunnen im Abst<strong>and</strong> von 10 m, 20 m<br />

und 30 m bestimmt und anschließend arithmetisch gemittelt. Für Methode 4 wird für die Ermittlung der<br />

Abbaurate die wahre Quellbreite angenommen.<br />

Ergebnisse und Diskussion<br />

Einfluss der Heterogenität<br />

Abb. 2 zeigt die Ergebnisse der Berechnung der Abbauratenkonstante erster Ordnung. Die Raten<br />

werden normalisiert dargestellt, d.h. die berechnete Ratenkonstante wird durch die wahre Ratenkonstante<br />

geteilt. Die normalisierte Rate kann so als Überschätzungs- bzw. Unterschätzungsfaktor<br />

interpretiert werden. Abb. 2 zeigt, dass die meisten berechneten Ratenkonstanten größer als 1.0 sind,<br />

d.h. dass die Ratenkonstante überschätzt wird. Dieser Effekt ist in einzelnen Realisierungen sehr<br />

stark, wo Überschätzungen der Ratenkonstante von einigen Größenordnungen auftreten können.<br />

Abb. 2a zeigt links die Variation der berechneten normalisierten Ratenkonstante mit der Quellbreite<br />

der emittierenden Schadst<strong>of</strong>fquelle. Es ist deutlich, dass mit zunehmender Quellbreite die berechnete<br />

Ratenkonstante sich der wahren Ratenkonstante annähert, d.h. sich dem Wert 1 annähert, was am<br />

deutlichsten für große Heterogenitäten sichtbar ist. Dies ist in Methode 1 begründet, die weder<br />

Verdünnung noch Dispersion oder „Aus der Fahne messen“ berücksichtigt. Diese Effekte werden mit<br />

zunehmender Quellbreite weniger signifikant, da dann die Annahmen zur Anwendung von Methode 1<br />

besser erfüllt sind. Auf der rechten Seite der Abb. 2 ist die Abhängigkeit der berechneten Ratenkonstante<br />

von der verwendeten Heterogenitätsklasse, durch die zugehörige Varianz σ²ln(K) bezeichnet,<br />

dargestellt. Es ist zu erkennen, dass eine Erhöhung von σ²ln(K) im Mittel zu einer Überschätzung der<br />

Ratenkonstante führt. Zusätzlich zu diesem Trend nimmt auch die St<strong>and</strong>ardabweichung der<br />

berechneten Ratenkonstanten zu, was die Zunahme der Spanne der ermittelten Ratenkonstanten<br />

wiederspiegelt. Diese Zunahme der Spanne kann als Zunahme der Unsicherheit der berechneten<br />

Ratenkonstante verst<strong>and</strong>en werden. Die Überschätzung beträgt im Mittel ca. 2 für geringe Heterogenität,<br />

steigt jedoch auf Werte von 4 bis 10 für mittlere und hohe Heterogenität und erreicht einen<br />

Wert von ca. 100 im Falle der sehr hohen hydraulischen Heterogenität.<br />

Abb. 2b stellt die Ergebnisse für Methode 2 dar. Auch anh<strong>and</strong> von Methode 2 werden die<br />

Ratenkonstanten überschätzt, jedoch sind, verglichen mit Methode 1, die Überschätzungen als auch<br />

die St<strong>and</strong>ardabweichung generell geringer. Methode 2 liefert also bessere und mit weniger<br />

Unsicherheit behaftete Ergebnisse. Wie man an der linken Grafik in Abb. 2b erkennen kann, ergibt<br />

sich für Methode 2 keine Abhängigkeit von der Quellbreite. Dies ist in der Methode begründet, da sie<br />

Effekte von Dispersion, Verdünnung und „Aus der Fahne messen“ durch die Normierung explizit<br />

berücksichtigt. Methode 3 zeigt ein ähnliches Verhalten wie Methode 1, sowohl in Abhängigkeit von<br />

der Quellbreite als auch in Abhängigkeit von der Heterogenität (Abb. 2c). Jedoch sind die<br />

normalisierten Abbauratenkonstanten höher als für Methode 1, was an der Berücksichtigung der<br />

Dispersion in Methode 3 liegt.<br />

Im eindimensionalen Aquifer mit einer Festkonzentration als R<strong>and</strong>bedingung führt die<br />

Berücksichtigung der longitudinalen Dispersion zu höheren Konzentrationen entlang der stationären<br />

Fahne, verglichen mit dem Fall ohne Dispersion. Dies wird durch den zusätzlichen dispersiven<br />

Massenaustrag aus der Schadst<strong>of</strong>fquelle verursacht. Um eine gemessene Konzentrationsabnahme<br />

zwischen Quelle und unterstromigen Brunnen mit Methode 3 erklären zu können ist also eine höhere<br />

Abbauratenkonstante notwendig als mit Methode 1. Methode 4 (Abb. 2d) schließlich zeigt ein sehr<br />

ähnliches Bild wie Methode 3. Aufgrund der in Methode 4 berücksichtigten Querdispersivität sind die<br />

bestimmten Abbauratenkonstanten etwas geringer als bei Methode 3. Für geringe Heterogenität wird<br />

die Ratenkonstante durch Methode 4 sogar unterschätzt. Dies liegt an einer zu hohen Korrektur durch<br />

6


den Querdispersionsterm, der die Effekte der Querdispersion für geringe Heterogenität überschätzt.<br />

Für den Fall geringer oder keiner Heterogenität sind die hier gewählten und in Methode 4<br />

verwendeten Dispersivitäten zu groß.<br />

a)<br />

b)<br />

c)<br />

d)<br />

normalisierte Ratenkonstante [-]<br />

normalisierte Ratenkonstante [-]<br />

normalisierte Ratenkonstante [-]<br />

normalisierte Ratenkonstante [-]<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

0 4 8 12 16 20<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

0 4 8 12 16 20<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

0 4 8 12 16 20<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

0 4 8 12 16 20<br />

Quellbreite [m]<br />

7<br />

0 1 2 3 4 5<br />

0 1 2 3 4 5<br />

0 1 2 3 4 5<br />

0 1 2 3 4 5<br />

σ 2 ln(K)<br />

Abb. 2 “Gemessene” Abbauratenkonstanten erster Ordnung, normalisiert auf die wahre<br />

Abbauratenkonstante, aufgetragen gegen Quellbreite (links) und Heterogenität (rechts). a) Methode 1,<br />

b) Methode 2, c) Methode 3 und d) Methode 4. Alle Abbildungen zeigen die Einzelergebnisse aller<br />

Realisierungen (kleine Symbole), ihren Mittelwert (große Symbole) und die zugehörige<br />

St<strong>and</strong>ardabweichung (Fehlerbalken). Unterschiedliche Grautöne geben die Quellbreite,<br />

unterschiedliche Symbole die verwendeten σ²ln(K) an.


Um diesen Effekt zu verdeutlichen sind in Abb. 2 für alle Methoden noch die für den homogenen Fall<br />

bestimmten Ratenkonstanten eingetragen. Die Ergebnisse sind für σ²ln(K) = 0 in den Abbildungen<br />

rechter H<strong>and</strong> als kleine horizontale Balken dargestellt. Für Methode 1 und Methode 2 beträgt die<br />

normalisierte Ratenkonstante im homogenen Fall genau 1.0, d.h. diese Methoden liefern exakt die<br />

wahre Ratenkonstante. Mit Methode 3 ergibt sich eine geringe Überschätzung, mit Methode 4 eine<br />

deutliche Unterschätzung der wahren Ratenkonstante. Die geringste Ratenkonstante wird für die<br />

geringste Quellbreite ermittelt, da hier die Korrektur durch β am größten ist (vergl. Tab 1).<br />

Da die bisher gezeigten Mittelwerte und St<strong>and</strong>ardabweichungen repräsentativ für das Ensemblemittel<br />

sind, nicht jedoch für die einzelnen Realisierungen, wurde ein weiteres Maß zum Vergleich der<br />

anh<strong>and</strong> der vier Methoden bestimmten Ratenkonstanten entwickelt. Abb. 3 zeigt die Wahrscheinlichkeit,<br />

mit der eine Methode zum Erfolg führen kann, worunter hier verst<strong>and</strong>en wird, dass die Abbaurate<br />

anh<strong>and</strong> der Methode mit einer gewünschten Genauigkeit ermittelt wird. Die gewünschte Genauigkeit<br />

wird als sogenannten Fehlerfaktor angegeben. Ein Fehlerfaktor von 10 entspricht dem Intervall 0.1 bis<br />

10 der normierten Ratenkonstanten, wobei dieses Intervall durch Division und Multiplikation von 1.0<br />

mit dem Fehlerfaktor (10) ermittelt wird. Dies entspricht somit der umgangssprachlichen Formulierung<br />

„... innerhalb einer Größenordnung ...“. Ein Fehlerfaktor von 5 entspricht somit dem Intervall 0.2 bis 5<br />

der normierten Ratenkonstanten. Abb. 3 gibt daher die Wahrscheinlichkeit dafür an, dass eine<br />

„gemessene“ Ratenkonstante innerhalb des durch den Fehlerfaktor aufgespannten Intervalls liegt.<br />

Abb. 3a zeigt für die geringste Heterogenität, dass die Wahrscheinlichkeit, die Abbauratenkonstante<br />

mit einem Fehlerfaktor kleiner als 2.0 zu bestimmen („ ... bis auf einen Faktor zwei ...“), für Methode 1<br />

ca. 70 %, für Methode 2 ca. 90%, für Methode 3 ca. 55% und für Methode 4 ca. 30% beträgt. Wird der<br />

Fehlerfaktor auf 5 erhöht, dann erhält man mit Methoden 1 bis 3 eine Erfolgswahrscheinlichkeit von<br />

100%, nur mit Methode 4 beträgt die Erfolgswahrscheinlichkeit ca. 70%. Je geringer die Werte einer<br />

Methode in Abb. 3 sind, desto geringer ist die Wahrscheinlichkeit, die Abbauratenkonstante mit der<br />

gewünschten Genauigkeit bestimmen zu können. Wie Abb. 3 zeigt, sinkt für alle Methoden die Erfolgswahrscheinlichkeit<br />

mit zunehmender Heterogenität. Beträgt die Erfolgswahrscheinlichkeit von Methode<br />

1 für einen Fehlerfaktor von 5 noch ca. 100% für die geringste Heterogenität, so sinkt diese<br />

Wahrscheinlichkeit auf 70%, 50% und schließlich 35% für die höheren Heterogenitätsklassen. Für<br />

Methoden 2 bis 4 sind die Werte analog Abbildung 3 zu entnehmen.<br />

Abb. 3 erlaubt daher einen direkten Vergleich der in dieser Arbeit untersuchten vier Methoden. Wird<br />

die Ratenkonstante in einem stark heterogenen Aquifer beispielsweise auf einen Faktor zehn genau<br />

benötigt, betragen die Erfolgswahrscheinlichkeiten ca. 80%, 95%, 80% und 60% für Methoden 1 bis 4<br />

(Abb. 3c). Für alle Heterogenitätsklassen liefert Methode 2 die höchste Wahrscheinlichkeit, das<br />

richtige Ergebnis zu bekommen. Für mittlere bis hohe Heterogenitäten folgt Methode 4 an zweiter<br />

Stelle, während diese Methode für geringe Heterogenität die schlechtesten Erfolgswahrscheinlichkeiten<br />

aufweist. Obwohl Methode 1 die einfachste Methode ist, liefert sie sehr ähnliche Ergebnisse wie<br />

Methode 4. Für geringe Heterogenitäten ergeben sich mit Methode 1 sogar die besseren<br />

Abschätzungen der Ratenkonstante. Methode 3 zeigt – bis auf geringe Heterogenität – immer die<br />

geringste Erfolgswahrscheinlichkeit.<br />

Für größere Quellbreiten ergeben sich qualitativ dieselben Ergebnisse. Methode 2 ist unabhängig von<br />

der Quellbreite, daher ist auch für Quellbreiten von 8 m oder 16 m die Erfolgswahrscheinlichkeit von<br />

Methode 2 dieselbe wie im Falle von 4 m. Für die <strong>and</strong>eren Methoden ergeben sich mit zunehmender<br />

Quellbreite höhere Erfolgswahrscheinlichkeiten. So steigt z.B. die Erfolgswahrscheinlichkeit von<br />

Methode 3 für σ²ln(K) = 2.7 von 35% für eine Quellbreite von 4 m (Abb. 3c) auf 40% für eine<br />

Quellbreite von 8 m und ca. 50% für eine Quellbreite von 16 m. Somit steigt die Erfolgswahrscheinlichkeit<br />

für Methoden 1, 3 und 4 mit der Quellbreite an, erreichen aber dennoch nicht die<br />

Erfolgswahrscheinlichkeit von Methode 2.<br />

8


1 a) b) 1<br />

c)<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

Methode 1<br />

Methode 2<br />

Methode 3<br />

Methode 4<br />

0<br />

1 10 100 1000<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

Methode 1<br />

Methode 2<br />

Methode 3<br />

Methode 4<br />

1 10 100 1000<br />

Fehlerfaktor<br />

9<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

d)<br />

Methode 1<br />

Methode 2<br />

Methode 3<br />

Methode 4<br />

0<br />

1 10 100 1000<br />

1<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0<br />

Methode 1<br />

Methode 2<br />

Methode 3<br />

Methode 4<br />

1 10 100 1000<br />

Fehlerfaktor<br />

Abb. 3 Erfolgswahrscheinlichkeit für alle vier Methoden gegen den Fehlerfaktor für σ²ln(K) a) 0.38, b)<br />

1.71, c) 2.7 <strong>and</strong> d) 4.5. Die Quellbreite beträgt 4 m.<br />

Einfluss des Messfehlers<br />

Bei der Untersuchung des Einflusses des Messfehlers wird von einem homogenen Aquifer ausgegangen.<br />

Die Unsicherheit wird nun nicht durch das heterogene Fließfeld erzeugt, sondern durch<br />

falsches „Messen“ von Piezometerhöhe und Schadst<strong>of</strong>fkonzentration.<br />

Da in der vorigen Untersuchungen gezeigt wurde, dass Methode 4 mit der gewählten transversalen<br />

Dispersivität im homogenen Fall zu kleine Abbauratenkonstanten liefert, wurde die transversale<br />

Dispersivität, die für die Auswertung mit Methode 4 (Tabelle 1) verwendet wird, auf 0.15 m verringert.<br />

Damit ergibt sich im homogenen Fall die richtige Abbauratenkonstante.<br />

In Abb. 4 ist die auf die wahre Ratenkonstante normierte fehlerbehaftete Ratenkonstante gegen den<br />

maximal möglichen Fehler bei der Bestimmung der Piezometerhöhe, ∆hmax, gesondert für die vier<br />

Auswertemethoden aufgetragen. In Abb. 4a erkennt man ein deutliches Ansteigen der normierten<br />

Ratenkonstanten mit zunehmendem ∆hmax für die Auswertung mit Methode 1. Bereits für einen<br />

maximalen Messfehler von 1 cm ergibt sich eine mittlere Überschätzung von ca. 2, der für ein ∆hmax<br />

von 5 cm auf ca. 5 steigt. Außerdem ist zu erkennen, dass für einzelne Messungen Überschätzungen<br />

der Ratenkonstante von mehr als 10 möglich sind. Generell liegt eine Überschätzung vor, da eine<br />

fehlerhaft bestimmte Piezometerhöhe zu einer falschen Abschätzung der Fließrichtung führt. Die neu<br />

platzierten unterstromigen Brunnen liegen dann nicht mehr auf der Zentrallinie, d.h. die gemessenen<br />

Konzentrationen sind kleiner als auf der Zentrallinie und dies führt zu einer Überschätzung der<br />

Ratenkonstanten. Es wird jedoch nicht nur die Fließrichtung falsch ermittelt, sondern auch die<br />

Fließgeschwindigkeit entlang der (angenommenen) Zentrallinie, da diese aus den gemessenen<br />

Piezometerhöhen, Brunnenabständen und hydraulischen Durchlässigkeiten berechnet wird. Der<br />

hieraus resultierende Fehler kann sowohl zu große als auch zu kleine Ratenkonstanten produzieren.<br />

Für Methode 2 (Abb. 4b) ergibt sich ein <strong>and</strong>eres Bild. Die mittlere Ratenkonstante ist auch für große<br />

∆hmax sehr nahe bei 1, d.h. es liegt kein Trend zu Über- oder Unterschätzung vor. Die Unsicherheit der<br />

bestimmten Abbaurate ist kleiner als bei Methode 1, wie an den Fehlerbalken zu erkennen ist. Da<br />

Methode 2 das „aus der Fahne messen“ korrigiert, lässt sich dieser Fehler für Methode 2 nicht<br />

beobachten (Abb. 4b). Es kommt sowohl zu einer Über- als auch Unterschätzung der Ratenkonstante,<br />

also einem eher symmetrischen Fehler ohne generelle Tendenz, der auf die falsche Abschätzung der<br />

Fließgeschwindigkeit zurückzuführen ist. Mit zunehmendem ∆hmax nimmt auch für Methode 2 die


Unsicherheit zu. Für Methode 3 ( Abb. 4c) und Methode 4 (Abb. 4d) ergibt sich – wie schon für die<br />

hydraulische Heterogenität – ein sehr ähnliches Bild wie für Methode 1, jedoch mit höheren<br />

Überschätzungen und Unsicherheiten als für Methode 1. Für Methode 3 und 4 beträgt die mittlere<br />

Überschätzung der Ratenkonstanten für ∆hmax = 1 cm bereits ca. 5, jeweils verbunden mit einer hohen<br />

Unsicherheit.<br />

normalisierte Abbauratenkonstante [-]<br />

normalisierte Abbauratenkonstante [-]<br />

100<br />

10<br />

1<br />

0.1<br />

100<br />

10<br />

1<br />

0.1<br />

0 1 2 3 4 5 6<br />

10<br />

100<br />

a) b)<br />

0 1 2 3 4 5 6<br />

maximaler Messfehler Piezometerhöhe [cm]<br />

10<br />

1<br />

0.1<br />

100<br />

c) d)<br />

10<br />

1<br />

0.1<br />

0 1 2 3 4 5 6<br />

0 1 2 3 4 5 6<br />

maximaler Messfehler Piezometerhöhe [cm]<br />

Abb. 4 Normalisierte Abbauratenkonstante, aufgetragen gegen den maximalen Messfehler bei der<br />

Bestimmung der Piezometerhöhe für a) Methode 1, b) Methode 2, c) Methode 3 und d) Methode 4.<br />

Dargestellt sind die Einzelergebnisse (kleine Symbole), die Mittelwerte (durch Linie verbundene<br />

Punkte) und die zugehörige St<strong>and</strong>ardabweichung (Fehlerbalken).<br />

Für Konzentrationsmessungen wurde ein höherer Messfehlerfaktor von 100 und eine exakte Messung<br />

der Piezometerhöhe angenommen (siehe Kapitel Methodik). Ein Messfehlerfaktor von 100 ist<br />

sicherlich sehr hoch gegriffen und wird hier nur zu Demonstrationszwecken eingesetzt. Das Ergebnis<br />

für Methode 1 zeigt Abb. 5a, in der die normierte Ratenkonstanten gegen den Fehlerfaktor<br />

aufgetragen sind. Man erkennt, dass für geringe Messfehlerfaktoren die Abbaukonstante sehr gut<br />

bestimmt werden kann, für einen Messfehlerfaktor von 2 erhält man im Mittel noch das richtige<br />

Ergebnis, für einen Messfehlerfaktor von 5 erhält man eine mittlere Überschätzung der Ratenkonstante<br />

von ca. 3. Die maximale mittlere Überschätzung der Ratenkonstante von ca. 5 wird bei einem<br />

Messfehlerfaktor von ca. 10 erreicht und steigt auch für größere Fehlerfaktoren nicht weiter an.<br />

Allerdings nimmt die Unsicherheit, dargestellt als St<strong>and</strong>ardabweichung, mit dem Messfehlerfaktor<br />

stark zu und man kann sowohl zu große als auch zu kleine Ratenkonstanten erhalten. Für Methode 2<br />

(Abb. 5b) erhält man generell eine geringere Überschätzung der Ratenkonstante von maximal ca. drei,<br />

jedoch eine ähnlich hohe Unsicherheit wie mit Methode 1. Generell werden bei Methode 2 eher<br />

kleinere Ratenkonstanten erzeugt als bei Methode 1. Methoden 3 und 4 (Abb 5.c und Abb. 5d) zeigen<br />

einen deutlichen Anstieg der ermittelten Überschätzung mit dem Messfehlerfaktor, maximale Werte<br />

liegen hier bei etwa 10. Ebenso ist die Unsicherheit dieser Werte sehr hoch.


normalisierte Abbauratenkonstante [-]<br />

normalisierte Abbauratenkonstante [-]<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

a)<br />

1 10 100<br />

c)<br />

1 10 100<br />

maximaler Messfehlerfaktor Konzentration [-]<br />

11<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

1000<br />

100<br />

10<br />

1<br />

0.1<br />

0.01<br />

b)<br />

1 10 100<br />

d)<br />

1 10 100<br />

maximaler Messfehlerfaktor Konzentration [-]<br />

Abb. 5 Normalisierte Abbauratenkonstante, aufgetragen gegen den maximalen Messfehlerfaktor bei<br />

der Bestimmung der Konzentration für a) Methode 1, b) Methode 2, c) Methode 3 und d) Methode 4.<br />

Dargestellt sind die Einzelergebnisse (kleine Symbole), die Mittelwerte (durch Linie verbundene<br />

Punkte) und die zugehörige St<strong>and</strong>ardabweichung (Fehlerbalken).<br />

Schlussfolgerungen<br />

Die vorgestellten Ergebnisse zeigen, dass sich die anh<strong>and</strong> der vier untersuchten Methoden ermittelten<br />

Abbauratenkonstanten erster Ordnung sowohl für zunehmende hydraulische Heterogenität als auch<br />

für zunehmende Messfehler bezüglich der Piezometerhöhe und der Konzentration unterscheiden. Alle<br />

Methoden zeigen geringere Erfolgswahrscheinlichkeiten mit zunehmender Heterogenität (Abb.3)<br />

sowie eine erhöhte Unsicherheit (Abb. 2). Aus Abb. 2 ist ersichtlich, dass diese Abnahme der<br />

Erfolgswahrscheinlichkeit durch eine generelle Überschätzung der Ratenkonstante verursacht wird.<br />

Diese Überschätzung ist am größten für Methode 3. Methode 2 wird dagegen am wenigsten von der<br />

hydraulischen Heterogenität beeinflusst und zeigt die geringsten Überschätzungen (Abb. 2) und<br />

Unsicherheiten. Obwohl Methoden 3 und 4 realitätsnäher sind, da sie auf der eindimensionalen bzw.<br />

zweidimensionalen Transportgleichung beruhen, zeigen sie eine geringere Erfolgswahrscheinlichkeit<br />

und eine größere Überschätzung der Ratenkonstanten als Methoden 1 und 2. Beide Methoden sind<br />

anfällig für Fehler, die durch die Abschätzung der longitudinalen und transversalen Dispersivitäten<br />

erzeugt werden können. Methode 1 als die einfachste Methode, da sie weder die Abschätzung der<br />

Dispersivität noch einen nichtreaktiven Mitkontamin<strong>and</strong>en benötigt, zeigt bessere oder ähnlich gute<br />

Ergebnisse wie Methode 4. Sowohl mittlere Überschätzung als auch Unsicherheit steigen mit<br />

zunehmender Größe des Messfehlers. Die Untersuchung des Messfehlers der Piezometerhöhe zeigt,<br />

dass bereits geringe Messfehler von 1 cm zu deutlicher Überschätzung der Ratenkonstanten und<br />

einer erhöhten Unsicherheit führen können. Der Einfluss des Messfehlers der Konzentration auf die<br />

Ratenkonstante ist generell geringer. Auch bei der Untersuchung der Messfehler erhält man, wie<br />

schon im Falle von hydraulischer Heterogenität, die besten Ergebnisse mit Methode 2. Methoden 3<br />

und 4 zeigen sehr ähnliche Ergebnisse und weisen größere Überschätzungen auf als Methode 2 oder<br />

Methode 1.<br />

Insgesamt ergibt sich, dass die Verwendung von Methode 2, die auf der Normierung der<br />

Schadst<strong>of</strong>fkonzentration mit einem nichtreaktiven Mitkontamin<strong>and</strong>en beruht, zu den besten<br />

Ergebnissen bei der Bestimmung von Abbauratenkonstanten erster Ordnung führt. Daher sollte diese<br />

Methode, wenn möglich, angewendet werden. Ist kein nichtreaktiver Mitkontamin<strong>and</strong> vorh<strong>and</strong>en, sollte<br />

Methode 1 verwendet werden, da sie die Abbauratenkonstante besser als Methoden 3 und 4 vorher-


sagt sowie die zusätzliche Problematik des Abschätzens der Dispersivitäten umgeht. In BAUER et al.<br />

(2005) wird der Einfluss der geschätzten Dispersivitäten auf die Abbauratenkonstante sowohl für die<br />

longitudinale als auch transversale Dispersivität anh<strong>and</strong> einer Sensitivitätsstudie eingehend untersucht.<br />

Es zeigt sich dabei, dass mit Methode 3 nie, mit Methode 4 nur unter Annahme sehr hoher<br />

transversaler Dispersivitäten, die nicht mit der Heterogenität (σ²ln(K)) des Aquifers begründet werden<br />

können, die korrekten Abbauraten ermittelt werden. In diesem Aufsatz werden weiterhin der Einfluss<br />

der Quellbreite sowie verschiedene Methoden zur Berechnung der Fliessgeschwindigkeit untersucht.<br />

Es zeigt sich, dass die Anordnung der Messpegel, abgeleitet aus dem hydraulischen Dreieck, großen<br />

Einfluss auf die geschätzte Abbaurate hat, wie er sich auch in Abbildung 1 widerspiegelt. Ebenso ist<br />

die Mittelung der an den einzelnen Messpegeln ermittelten hydraulischen Durchlässigkeiten von<br />

Bedeutung. Hier ergibt sich, dass eine geometrische Mittelung zu bevorzugen ist (BAUER et al., 2006).<br />

Die gefundene generelle Überschätzung der Ratenkonstante ist im Hinblick auf eine Prognose der<br />

Fahnenlänge problematisch, da eine zu hohe Ratenkonstante den Abbau überschätzt und daher zu<br />

geringe prognostizierte Fahnenlängen verursacht. Dieser Fehler ist daher nicht konservativ und kann<br />

zu einer zu optimistischen Einschätzung der St<strong>and</strong>ortverhältnisse führen. Im Folgenden ist die<br />

Fahnenlänge definiert als die Strecke zwischen Schadensquelle und dem Ort auf der<br />

Fahnenzentralachse, an dem das Verhältnis C/C0 einen vorgegebenen Wert erreicht hat<br />

(beispielsweise 10 -3 ), wobei C die Konzentration in der Fahne und C0 die Konzentration in der Quelle<br />

ist. Für eindimensionale Lösungen ohne Dispersion ergibt sich, dass Abbaurate und Fahnenlänge<br />

umgekehrt proportional zuein<strong>and</strong>er sind und die Fahnenlänge linear mit dem Kehrwert der Abbaurate<br />

wächst (Vergleiche Methoden 1 und 2 in Tabelle 1). Eine Überschätzung der Abbaurate um den<br />

Faktor 5 verursacht also eine Unterschätzung der Fahnenlänge ebenfalls um den Faktor 5. Im<br />

eindimensionalen Fall mit Dispersion sowie im zweidimensionalen Fall ist die Fahnenlänge annähernd<br />

umgekehrt proportional zum Kehrwert der Quadratwurzel der Abbaurate, die genauen Werte hängen<br />

von den gegebenen Dispersivitäten, der Quellbreite und dem Verhältnis von Abst<strong>and</strong>sgeschwindigkeit<br />

zu Abbaurate ab (vergleiche Methoden 3 und 4 in Tabelle 1). Eine Überschätzung der Fahnenlänge<br />

um den Faktor 5 bzw. 10 kann beispielsweise eine Unterschätzung der Fahnenlänge um den Faktor 3<br />

bzw. 5 verursachen. Der Einfluss der Dispersion verringert die Sensitivität der Fahnenlänge auf die<br />

Abbaurate. Eine genauere Untersuchung des Einflusses der ermittelten Abbaurate auf die<br />

prognostizierte Fahnenlänge wurde von BEYER et al., (2005b) durchgeführt. Dabei zeigte sich eine<br />

durchschnittliche Unterschätzung der Fahnenlängen um 50 %. Für einzelne Realisierungen ergaben<br />

sich zum Teil jedoch erheblich größere Fehler, sodass in ungünstigen Fällen die geschätzte<br />

Fahnenlänge lediglich 10% des tatsächlichen Wertes betrug. Dabei zeigt sich, dass die mit Methoden<br />

1, 3 und 4 berechneten Abbauraten prinzipiell zu ähnlichen Abschätzungen der Fahnenlänge führen,<br />

wenn diese mit einem mit der Abbaurate korrespondierenden Transportmodell berechnet wird. Die<br />

durch Vernachlässigung bzw. falsche Parametrisierung der Längs- und Querdispersion verursachten<br />

Fehler in der Abbaurate werden bei der Fahnenlängenprognose teilweise wieder aufgehoben. Als<br />

problematisch stellte es sich jedoch heraus, Abbauraten, die mit der eindimensionalen Lösung<br />

berechnet wurden, in einem zwei- oder dreidimensionalen Transportmodell zur Prognose der<br />

Fahnenlänge zu verwenden (BEYER et al., 2005b).<br />

Danksagung: Der Beitrag entst<strong>and</strong> im Rahmen des Projekts TV 7.1 „Modellierung und<br />

Prognose“ des BMBF- Förderschwerpunktes “Kontrollierter natürlicher Rückhalt und Abbau<br />

von Schadst<strong>of</strong>fen bei der Sanierung kontaminierter Grundwässer und Böden (KORA)“ an den<br />

Universitäten Kiel und Tübingen. Den Projektträgern und dem BMBF sei für die hierbei<br />

gewährte Unterstützung gedankt.<br />

Literaturangaben<br />

Bauer, S., Beyer, C. , Kolditz, O. (2006): Assessing measurement uncertainty <strong>of</strong> first order degradation<br />

rates in heterogeneous aquifers. - Water Resour. Res., im Druck.<br />

Bauer, S., Beyer, C., Kolditz, O. (2005): Assessing measurements <strong>of</strong> first order degradation rates by<br />

using the Virtual Aquifer approach. - IAHS Publication, 297: 274-281.<br />

Bauer, S., Kolditz, O. (2005): Uncertainty assessment <strong>of</strong> integral pumping tests in heterogeneous<br />

aquifers. - Angenommener Beitrag für eine IAHS Publikation.<br />

Beyer, C., Bauer S., Kolditz, O. (2005a): Uncertainty assessment <strong>of</strong> degradation rate measurements in<br />

heterogeneous media using the virtual aquifer approach. - In: Kolditz, O., Bauer, S., <strong>and</strong><br />

Gronewold, J. (Eds.): Proceedings <strong>of</strong> the 5th Workshop “Porous Media”, Blaubeuren, Germany,<br />

December 2004,. ZAG Publisher, Tübingen.<br />

12


Beyer, C., Bauer S., Kolditz, O. (2005b): Using the Virtual Aquifer approach for uncertainty<br />

assessment <strong>of</strong> contaminant plume length estimates. GeoSys Preprint, GeoSystemForschung,<br />

Tübingen. http://www.uni-tuebingen.de/zag/geohydrology/index.html<br />

Bockelmann, A., Zamfirescu, D., Ptak, T., Grathwohl, P., Teutsch, G. (2003): Quantification <strong>of</strong> mass<br />

fluxes <strong>and</strong> natural attenuation rates at an industrial site with a limited monitoring network: A case<br />

study. - J. Contam. Hydrol., 60: 97-121.<br />

Buscheck, T.E., Alcantar, C.M. (1995): Regression techniques <strong>and</strong> analytical solutions to demonstrate<br />

intrinsic bioremediation. - In: Proceedings <strong>of</strong> the 1995 Battelle International Conference on In-Situ<br />

<strong>and</strong> On Site Bioreclamation, Batelle, CA, April 1995. 109-116.<br />

Chapelle, F.H., Bradley, P.M., Lovley, D.R., Vroblesky, D.A. (1996): Measuring rates <strong>of</strong> biodegradation<br />

in a contaminated aquifer using field <strong>and</strong> laboratory methods. -Ground Water, 34(4): 691-698.<br />

Herfort, M. (2000): Reactive Transport <strong>of</strong> Organic Compounds Within a Heterogeneous Porous<br />

Aquifer. - Tübinger Geowissenschaftliche Arbeiten, 54, Universität Tübingen, Tübingen.<br />

Kolditz, O., de Jonge, J., Beinhorn, M., Xie, M., Kalbacher, M., Wang, W., Bauer, S., McDermott, C.,<br />

Chen, C., Beyer, C., Gronewold, J., Kemmler, D., Manabe, T., Legeida D., Adamidis, P. (2005):<br />

GeoSys - Theory <strong>and</strong> users manual, release 4.2. GeoHydrology / HydroInformatics, Center for<br />

<strong>Applied</strong> Geoscience, Universität Tübingen, Tübingen.<br />

McNab Jr, W. W., Dooher, B.P. (1998): A critique <strong>of</strong> a steady-state analytical method for estimating<br />

contaminant degradation rates. - Ground Water 36(6): 983–987.<br />

Pebesma, E. J., Wesseling, C. G. (1998): Gstat: A program for geostatistical <strong>modeling</strong>, prediction <strong>and</strong><br />

simulation. - Computers & Geosciences, 24(1): 17-31, doi:10.1016/S0098- 3004(97)00082-4.<br />

Rehfeldt, K.R., Boggs, J.M., Gelhar, L.W. (1992): Field study <strong>of</strong> dispersion in a heterogeneous aquifer,<br />

3, geostatistical analysis <strong>of</strong> hydraulic conductivity. - Water Resour. Res., 28(12): 3309–3324.<br />

Rubin,Y. (2003): <strong>Applied</strong> Stochastic Hydrogeology. - 416 S.; Oxford University Press, New York, NY.<br />

Schäfer, D., Dahmke, A., Kolditz, O., Teutsch, G. (2002): "Virtual Aquifers": A concept for evaluation <strong>of</strong><br />

exploration, remediation <strong>and</strong> monitoring strategies. - In: Calibration <strong>and</strong> Reliability in Groundwater<br />

Modelling: A Few Steps Closer to Reality (Proceedings <strong>of</strong> the ModelCARE 2002 Conference held<br />

in Prague, Czech Republic, June 2002), edited by K. Kovar & Z. Hrkal, IAHS Publication, 277: 52-<br />

59.<br />

Schäfer, D., Schlenz B., Dahmke, A. (2004): Evaluation <strong>of</strong> exploration <strong>and</strong> monitoring methods for<br />

verification <strong>of</strong> natural attenuation using the virtual aquifer approach. - Biodegradation Journal<br />

15(6): 453-465.<br />

Schäfer, D., Hornbruch, G., Schlenz B., Dahmke, A. (2005): Untersuchung kinetischer Ansätze zur<br />

Modellierung mikrobieller Abbauprozesse mit Hilfe des Virtuelle Aquifere Ansatzes. - Eingereicht<br />

bei Grundwasser.<br />

Suarez, M. P., Rifai, H.S. (2002): Evaluation <strong>of</strong> BTEX remediation by natural attenuation at a coastal<br />

facility. - Ground Water Monit. Remed. 22(1): 62-77.<br />

Stenback, G.A., Ong, S.K., Rogers, S.W., Kjartanson, B.H. (2004): Impact <strong>of</strong> transverse <strong>and</strong> longitudinal<br />

dispersion on first-order degradation rate constant estimation. - J. Cont. Hydrol. 73: 3-14.<br />

Sudicky, E. A. (1986): A natural gradient experiment on solute transport in a s<strong>and</strong> aquifer: Spatial<br />

variability <strong>of</strong> hydraulic conductivity <strong>and</strong> its role in the dispersion process. - Water Resour. Res.<br />

22(13): 2069-2082.<br />

U.S. Environmental Protection Agency (1999): Use <strong>of</strong> monitoring natural attenuation at Superfund,<br />

RCRA Corrective Action, <strong>and</strong> Underground Storage Tank Sites, - Office <strong>of</strong> Solid Waste <strong>and</strong><br />

Emergency Response Directive 9200.4-17, Washington, D.C.<br />

Wiedemeier, T.H., Swanson, M.A., Wilson, J.T., Kampbell, D.H., Miller, R.N., Hansen,J.E. (1996):<br />

Approximation <strong>of</strong> biodegradation rate constants for monoaromatic hydrocarbons (BTEX) in ground<br />

water. - Ground Water Monit. Remed., 16(3): 186-194.<br />

Wiedemeier, T.H., Rifai, H.S., Wilson, J.T., Newell, C. (1999): Natural Attenuation <strong>of</strong> Fuels <strong>and</strong><br />

Chlorinated Solvents in the Subsurface. - 617 S.; Wiley, New York, NY.<br />

Wilson, J. T., Pfeffer, F. M., Weaver, J. W., Kampbell, D. H., Wiedemeier, T. H., Hansen, J. E., Miller,<br />

R. N. (1994): Intrinsic Bioremediation <strong>of</strong> JP-4 Jet Fuel. - In: Symposium on Intrinsic<br />

Bioremediation <strong>of</strong> Ground Water, Denver, Colorado, US-EPA: 60-72.<br />

Wilson, R. D., Thornton, S.F. Mackay, D.M. (2004): Challenges in monitoring the natural attenuation <strong>of</strong><br />

spatially variable plumes. - Biodegradation Journal 15(6): 459-469.<br />

Zamfirescu, D., Grathwohl, P.(2001): Occurrence <strong>and</strong> attenuation <strong>of</strong> specific organic compounds in the<br />

groundwater plume at a former gasworks site. - J. Contam. Hydrol. 53: 407-427.<br />

Zhang, Y.-K. , Heathcote, R.C. (2003): An improved method for estimation <strong>of</strong> biodegradation rate with<br />

field data. - Ground Water Monit. Remed., 23(3): 112-116.<br />

13


Enclosed Publication 5<br />

Beyer, C., Chen, C., Gronewold, J., Kolditz, O., Bauer, S. (2007a): Determination <strong>of</strong> first<br />

order degradation rate constants from monitoring networks. (accepted by Ground Water.).<br />

The enclosed article was accepted for publication in the journal Ground Water. Copyright ©<br />

2007 Blackwell Publishing.<br />

The definite version will be available at www.blackwell-synergy.com.


Determination <strong>of</strong> first order degradation rate constants from<br />

monitoring networks ∗<br />

Christ<strong>of</strong> Beyer, Cui Chen, Jan Gronewold, Olaf Kolditz, Sebastian Bauer<br />

Center for <strong>Applied</strong> Geoscience, Eberhard-Karls-University <strong>of</strong> Tübingen,<br />

Sigwartstraße 10, D 72076 Tübingen, Germany<br />

Tel.: +49 7071 29 73176; Fax: +49 7071 5059. E-mail: christ<strong>of</strong>.beyer@uni-tuebingen.de<br />

Abstract<br />

In this paper different strategies for estimating first order degradation rate constants from measured field data are<br />

compared by application to multiple synthetic contaminant plumes. The plumes were generated by <strong>numerical</strong> simulation<br />

<strong>of</strong> contaminant transport <strong>and</strong> degradation in virtual heterogeneous aquifers. These sites then were individually <strong>and</strong><br />

independently investigated on the computer by installation <strong>of</strong> extensive networks <strong>of</strong> observation wells. From the data<br />

measured at the wells, i.e. contaminant concentrations, hydraulic conductivities <strong>and</strong> heads, first order degradation rates<br />

were estimated by three one-dimensional center line methods, which use only measurements located on the plume axis,<br />

<strong>and</strong> a two-dimensional method, which uses all concentration measurements available downgradient from the<br />

contaminant source. Results for both strategies show that the true rate constant used for the <strong>numerical</strong> simulation <strong>of</strong> the<br />

plumes in general tends to be overestimated. Overestimation is stronger for narrow plumes from small source zones with<br />

an average overestimation factor <strong>of</strong> about 5 <strong>and</strong> single values ranging from 0.5 to 20, decreasing for wider plumes, with<br />

an average overestimation factor <strong>of</strong> about 2 <strong>and</strong> similar spread. Reasons for this overestimation are identified in the<br />

velocity calculation, the dispersivity parameterization <strong>and</strong> <strong>of</strong>f-center-line measurements. For narrow plumes the onedimensional<br />

<strong>and</strong> the two-dimensional strategies show approximately the same amount <strong>of</strong> overestimation. For wide<br />

plumes, however, incorporation <strong>of</strong> all measurements in the two-dimensional approach reduces the estimation error. No<br />

significant relation between the number <strong>of</strong> observation wells in the monitoring network <strong>and</strong> the quality <strong>of</strong> the estimated<br />

rate constant is found for the two-dimensional approach.<br />

Introduction<br />

Although detailed mathematical descriptions <strong>of</strong> contaminant degradation in the subsurface are available (e.g. Rittmann<br />

<strong>and</strong> VanBriesen 1996; Wiedemeier et al. 1999; Islam et al. 2001), in field studies <strong>of</strong>ten simplified approaches are used for<br />

contaminant transport <strong>modeling</strong>. This is mainly because the identification <strong>of</strong> a large number <strong>of</strong> parameters <strong>and</strong> processes<br />

from field data <strong>of</strong>ten is impossible. Due to its mathematical simplicity, its easy implementation into transport models <strong>and</strong><br />

the necessity <strong>of</strong> determining only a single parameter, the biodegradation model most frequently used is first order kinetics<br />

(Wiedemeier et al. 1999). Field methods for the determination <strong>of</strong> biodegradation rates in ground water comprise a variety<br />

<strong>of</strong> approaches, including mass balance calculations, in-situ microcosm studies <strong>and</strong> the so called center line method<br />

(Chapelle et al. 1996). With the center line method, only concentration measurements in observation wells located on the<br />

plume axis are evaluated <strong>and</strong> additional information possibly at h<strong>and</strong> (e.g. well data downgradient from the source but<br />

not on the center line) is not explicitly accounted for in the rate estimation. Thus, the center line method is an essentially<br />

one-dimensional approach to rate constant estimation, as <strong>flow</strong>, transport <strong>and</strong> degradation are only evaluated for a single<br />

streamline. Assuming that contaminant biodegradation can be approximated by a first order kinetics, the degradation rate<br />

constant can be calculated with analytical transport models to yield the sought for first order rate constant. An overview<br />

<strong>of</strong> common methods for estimating degradation rate constants from field data within the context <strong>of</strong> monitored natural<br />

attenuation is given by Newel et al. (2002). The authors discuss approaches based on (point-) concentration over time<br />

data <strong>and</strong> different concentration vs. distance relations, where dispersion is completely neglected as well as accounted for.<br />

Important to note is that the different approaches yield different types <strong>of</strong> degradation rates that should be used for<br />

destined purposes only. Rate constants estimated from concentration vs. time data for single monitoring wells for<br />

example represent point decay rates that are typically used to assess source decay or the time required to reach defined<br />

remediation goals at a particular location (Newell et al. 2002). Complete neglect <strong>of</strong> dispersion in the rate estimation<br />

procedure for concentration vs. distance data along the center line <strong>of</strong> the plume yields bulk attenuation rates, which<br />

quantify the reduction <strong>of</strong> contaminant concentrations with distance from the source due to the combined effects <strong>of</strong><br />

dispersion, dilution (e.g. by recharge) <strong>and</strong> degradation processes. The frequently used method <strong>of</strong> Buscheck <strong>and</strong> Alcantar<br />

∗<br />

Beyer, C., Chen, C., Gronewold, J., Kolditz, O., Bauer, S. (2007): Determination <strong>of</strong> first order degradation rate constants from monitoring networks.<br />

(accepted by Ground Water.).<br />

The manuscript was accepted for publication in the Journal Ground Water. Copyright © 2007 Blackwell Publishing. The definite version will be<br />

available at www.blackwell-synergy.com.<br />

1


(1995) yields a “hybrid” rate constant between a bulk attenuation <strong>and</strong> a pure biodegradation rate constant, as it accounts<br />

for longitudinal dispersion whereas the effects <strong>of</strong> transverse dispersion are still reflected in the rate constant. A pure<br />

biodegradation rate constant can be estimated analytically by normalizing contaminant concentrations to concentrations<br />

<strong>of</strong> a recalcitrant tracer, if such a compound is emitted from the same source zone as the contaminant <strong>of</strong> concern<br />

(Wiedemeier et al. 1996). If this is not the case, an approach <strong>of</strong> Zhang <strong>and</strong> Heathcote (2003) may be used, in which the<br />

Buscheck <strong>and</strong> Alcantar (1995) method is extended to account for dispersion in two or three dimensions. Pure<br />

biodegradation rate constants exclusive <strong>of</strong> dispersion or other attenuation processes (<strong>and</strong> only those) may be used in<br />

contaminant transport models for prognoses <strong>of</strong> plume trends. It is important to note that these center line based rate<br />

constant estimation approaches are only applicable to contaminant plumes that have reached steady state conditions, as<br />

for still exp<strong>and</strong>ing plumes the rate constant would be overestimated.<br />

Although it is known, that biodegradation rate estimates obtained from an investigation <strong>of</strong> the plume center line are<br />

subject to substantial uncertainty, this strategy is frequently used in practice. Even in homogeneous aquifers vertical <strong>and</strong><br />

horizontal transverse dispersion can produce center line concentration pr<strong>of</strong>iles <strong>of</strong> recalcitrant compounds that could be<br />

mistaken as following from first order degradation (McNab Jr. <strong>and</strong> Dooher 1998). Moreover, the plume axis may easily<br />

be missed by monitoring wells when the inferred ground water <strong>flow</strong> direction is incorrect, changes over time due to<br />

transient <strong>flow</strong> behavior or when the contaminant plume shows large scale me<strong>and</strong>ering due to aquifer heterogeneity<br />

(Newell et al. 2002; Wilson et al. 2004). Therefore, biodegradation rate constants obtained from such field data should be<br />

taken as rough estimates only (Chapelle et al. 2003). As first order rates calculated from center line data include all<br />

effects <strong>and</strong> processes that lower local contaminant concentrations, the precarious result <strong>of</strong> the different sources <strong>of</strong><br />

uncertainty is that the degradation potential may be severely overestimated (Rittmann 2004), causing underestimation <strong>of</strong><br />

plume length <strong>and</strong> contaminant mass as well as a too optimistic prognosis <strong>of</strong> down gradient concentrations <strong>and</strong> exposure<br />

levels. These aspects were recently studied in-depth by Bauer et al. (2006a) <strong>and</strong> Beyer et al. (2006) by means <strong>of</strong><br />

<strong>numerical</strong> experiments in two-dimensional synthetic heterogeneous contaminated aquifers. The <strong>numerical</strong> experiments<br />

were based on coarse monitoring networks with 6 to 8 wells, which were all placed along an inferred plume center line.<br />

In reality, however, monitoring networks typically are designed to suit multiple <strong>and</strong> sometimes conflicting requirements<br />

<strong>and</strong> objectives, which may also change with time, depending on the stage <strong>of</strong> the site investigation. Objectives in this<br />

context are the detection <strong>of</strong> ground water contamination (e.g. Storck et al. 1997), site characterization <strong>and</strong> spatial<br />

delineation <strong>of</strong> the contamination (e.g. McGrath <strong>and</strong> Pinder 2003) or the long term monitoring <strong>of</strong> the plume behavior (e.g.<br />

Wu et al. 2006). These aims require spatially more extensive monitoring networks with larger numbers <strong>of</strong> wells<br />

compared to the relatively simple center line well configurations necessary to estimate a degradation rate constant.<br />

Recently, Stenback et al. (2004) demonstrated that additional <strong>of</strong>f center line measurements can be incorporated in the<br />

estimation <strong>of</strong> the degradation rate, when a two-dimensional analytical transport model is fitted to contaminant<br />

concentrations <strong>of</strong> all monitoring wells <strong>of</strong> an extensive monitoring network downgradient from the source. In a field<br />

application example Stenback et al. (2004) showed, that accounting for the additional information on contaminant<br />

concentrations <strong>and</strong> distribution significantly reduced rate constant estimates obtained from the conventional center line<br />

approach to about 50 %, pointing out the well known problem <strong>of</strong> rate constant overestimation with the 1D center line<br />

method.<br />

This paper therefore studies the performance <strong>of</strong> several rate constant estimation approaches based on center line<br />

investigation data for sites with extensive monitoring networks <strong>and</strong> compares the results to the two-dimensional approach<br />

<strong>of</strong> Stenback et al. (2004). Adopting the terminology <strong>of</strong> Newell et al. (2002) the types <strong>of</strong> degradation rate constants<br />

regarded here comprise bulk attenuation, biodegradation <strong>and</strong> “hybrid” rate constants. Point decay rates are not addressed<br />

in this paper. Both strategies for rate constant estimation, i.e. the investigation <strong>of</strong> the plume center line <strong>and</strong> the approach<br />

<strong>of</strong> Stenback et al. (2004), are applied to a set <strong>of</strong> synthetic sites with extensive monitoring networks, which were<br />

independently designed <strong>and</strong> installed by individual test persons engaged in hydrogeological research <strong>and</strong> consulting. The<br />

networks were installed for a general characterization <strong>and</strong> quantification <strong>of</strong> the contaminant plume (Chen et al. 2005;<br />

Bauer et al. 2006b), <strong>and</strong> are used here as a basis for degradation rate estimation. Estimated rate constants for both<br />

strategies are compared with regard to magnitude <strong>of</strong> errors <strong>and</strong> variability to draw conclusions on their limitations in<br />

view <strong>of</strong> the monitoring network used. Thereby the studies <strong>of</strong> Bauer et al. (2006a) <strong>and</strong> Beyer et al. (2006) are considerably<br />

extended, as the monitoring networks used in this paper are representative <strong>of</strong> real field situations by using all installed<br />

observation wells, <strong>and</strong> are not restricted to an one-dimensional plume center line. Moreover, this analysis allows an<br />

evaluation <strong>of</strong> the “human factor” on estimated rate constants resulting from individual notions <strong>of</strong> “sufficient accuracy” in<br />

plume investigation.<br />

Background <strong>and</strong> Scope<br />

This study is based on a set <strong>of</strong> synthetic contaminated two-dimensional aquifers, generated by multiple stochastic<br />

simulations <strong>of</strong> heterogeneous hydraulic conductivity fields <strong>and</strong> subsequent <strong>numerical</strong> simulation <strong>of</strong> contaminant<br />

spreading from a defined source zone in the synthetic aquifers. The evolved virtual plumes were independently<br />

investigated by a number <strong>of</strong> German scientists engaged in hydrogeological <strong>and</strong> environmental research using an<br />

2


interactive plume investigation <strong>and</strong> mapping s<strong>of</strong>tware (Chen et al. 2005) with any number <strong>of</strong> wells considered necessary<br />

for a characterization <strong>of</strong> the synthetic sites <strong>and</strong> the delineation <strong>of</strong> the contaminant plume. The plume investigations were<br />

performed within a related project (Bauer et al. 2006b) <strong>and</strong> are described in detail below. The results <strong>of</strong> these virtual<br />

plume investigations yield individually <strong>and</strong> realistically investigated plumes. The “measured” concentrations, hydraulic<br />

heads <strong>and</strong> conductivities at the observation well networks <strong>of</strong> the different plumes investigated are used as a basis for the<br />

determination <strong>of</strong> the degradation rate constants. Two different strategies are compared here:<br />

• Strategy A - One-dimensional center line approach: From all observation wells installed the concentration<br />

distribution is analyzed <strong>and</strong> wells along the approximate center line <strong>of</strong> the contaminant plume are identified (see<br />

Figure 1(a)). Concentrations measured in these wells are evaluated using three different methods for the<br />

inference <strong>of</strong> a degradation rate constant (Newell et al. 2002; Buscheck <strong>and</strong> Alcantar 1995; Zhang <strong>and</strong> Heathcote<br />

2003). Thus, here only a subset <strong>of</strong> concentrations measured in the original monitoring network is used for rate<br />

constant estimation.<br />

• Strategy B - Two-dimensional evaluation: An approximate solution for the two-dimensional steady state<br />

concentration distribution is fitted to concentrations measured in all observation wells <strong>of</strong> the monitoring network<br />

downgradient <strong>of</strong> the source zone using a residual least squares criterion (Stenback et al. 2004) (Figure 1(b)).<br />

Fitting parameters are the first order rate constant <strong>and</strong> the source concentration.<br />

Figure 1: a) Strategy A <strong>and</strong> b) strategy B for the inference <strong>of</strong> the degradation rate constant using an existing monitoring<br />

network. Shown is the true virtual plume (unknown to the investigator), the monitoring network (small squares) <strong>and</strong> the<br />

identified center line wells (larger squares) (strategy A) as well as an example <strong>of</strong> a fitted 2D analytical plume (contour<br />

lines) (strategy B).<br />

Generation <strong>of</strong> synthetic sites<br />

The data basis <strong>of</strong> this study consists <strong>of</strong> 20 different realizations <strong>of</strong> heterogeneous steady state contaminant plumes. The<br />

plumes were generated by <strong>numerical</strong> simulation <strong>of</strong> ground water <strong>flow</strong> <strong>and</strong> contaminant transport in 20 different twodimensional<br />

horizontal aquifer realizations with heterogeneous hydraulic conductivity distributions. Into ten <strong>of</strong> these<br />

conductivity fields a source with a width <strong>of</strong> 4 m perpendicular to the mean <strong>flow</strong> direction was introduced, for the other 10<br />

aquifer realizations, a source width <strong>of</strong> 16 m was used. For the <strong>numerical</strong> simulations the GeoSys/Rock<strong>flow</strong> code (Kolditz<br />

et al. 2006; Kolditz <strong>and</strong> Bauer 2004) was used, which solves the <strong>flow</strong> <strong>and</strong> transport equations by finite element methods.<br />

The governing equations for steady state <strong>flow</strong> conditions are given as (e.g. Bear 1972):<br />

∇( K ∇h)<br />

= 0<br />

(1)<br />

∂C<br />

= −va∇C<br />

+ ∇(<br />

D∇C)<br />

− λC<br />

(2)<br />

∂t<br />

with h [L] the hydraulic head, K [L T -1 ] the tensor <strong>of</strong> hydraulic conductivity, C [M L -3 ] concentration, D [L 2 T -1 ] the<br />

dispersion tensor which is calculated acording to Bear (1972), va [L T -1 ] the transport velocity, t [T] time <strong>and</strong> λ [T -1 ]<br />

representing the first order degradation rate constant. Equation (2) was solved over a two-dimensional <strong>numerical</strong> grid <strong>of</strong><br />

184 m length by 64 m width with node spacing <strong>of</strong> 0.5 m in both directions <strong>and</strong> setting the left h<strong>and</strong> side <strong>of</strong> equation (2) to<br />

0. This guarantees that the simulated plumes are at steady state. For local dispersivities αL <strong>and</strong> αT values <strong>of</strong> 0.25 m <strong>and</strong><br />

0.05 m are used. A mean hydraulic gradient I <strong>of</strong> 0.003 is induced by fixed head boundary conditions on the left <strong>and</strong> the<br />

right h<strong>and</strong> side <strong>of</strong> the model domain (Figure 2). Flow conditions are at steady state. Hydraulic conductivity K <strong>of</strong> the<br />

virtual aquifers is regarded as a spatial r<strong>and</strong>om variable (Figure 2), following a lognormal distribution with an expected<br />

value <strong>of</strong> E[Y = ln(K)] = -9.54, which corresponds to an effective conductivity Kef <strong>of</strong> 7.19·10 -5 m s -1 using the geometric<br />

mean. As the aquifer models are two-dimensional <strong>and</strong> horizontal, an isotropic exponential covariance function with an<br />

3


2<br />

integral scale lY <strong>of</strong> 2.67 m <strong>and</strong> ln-conductivity variance σ Y = 2.7 is used for the spatial correlation structure. Kef <strong>and</strong> lY<br />

2<br />

are taken from the Borden field site (Sudicky 1986), whereas the degree <strong>of</strong> heterogeneity σ Y was reported for the<br />

Columbus Air Force Base site (Rehfeldt et al. 1992). Porosity n is set to 0.33, resulting in va = 6.54 10 -7 m s -1 . For local<br />

dispersivities αL <strong>and</strong> αT values <strong>of</strong> 0.25 m <strong>and</strong> 0.05 m are used. The contaminant source introduced in the aquifers is <strong>of</strong><br />

rectangular shape centered at [11.5 m; 32.0 m] downstream <strong>of</strong> the in<strong>flow</strong> boundary <strong>and</strong> has an area <strong>of</strong> either 3 * 4 m or 3<br />

* 16 m, corresponding to a source width WS <strong>of</strong> 1.5 or 6 correlation lengths transverse to the mean <strong>flow</strong> direction. It emits<br />

a contaminant with the contaminant concentration fixed in the source area at C / C0 = 1.0. The dissolved contaminant is<br />

subject to a first order kinetics degradation reaction with a rate constant λ <strong>of</strong> 1.59·10 -8 s -1 (0.5 a -1 = 0.5 yr -1 ). This value is<br />

well within the range <strong>of</strong> reported / recommended first order degradation rate constants for chlorinated solvents as well as<br />

petroleum hydrocarbons under anaerobic conditions listed in Aronson <strong>and</strong> Howard (1997). The model parameters used<br />

are summarized in Table 1.<br />

contaminant source: C = 1.0<br />

0<br />

fixed head<br />

h = 0.552 m<br />

4<br />

no <strong>flow</strong><br />

no <strong>flow</strong><br />

fixed head<br />

h = 0.0 m<br />

-1<br />

K [m s ]<br />

Figure 2: A single realization <strong>of</strong> the spatially correlated r<strong>and</strong>om hydraulic conductivity field <strong>and</strong> model boundary<br />

conditions.<br />

The conceptual model used here is a rigorous simplification <strong>of</strong> contaminant transport <strong>and</strong> degradation observed in natural<br />

aquifer systems, where biodegradation is a function <strong>of</strong> electron acceptor <strong>and</strong> donor availability <strong>and</strong> the aquifer structure<br />

<strong>and</strong> thus reaction kinetics follow more complicated laws, show spatial dependence <strong>and</strong> may include transient effects,<br />

dilution or phase changes to the un<strong>saturated</strong> zone. Furthermore, the contaminant is not retarded <strong>and</strong> shows no<br />

volatilization. However, as the methods assume a uniform rate constant to be valid, only using such simplified<br />

representations <strong>of</strong> reality as virtual test sites allows for a detailed analysis <strong>of</strong> the influence <strong>of</strong> heterogeneous hydraulic<br />

conductivity on the different methods under otherwise ideal conditions.<br />

1.0*10 -2<br />

1.0*10 -3<br />

1.0*10 -4<br />

1.0*10 -5<br />

1.0*10 -6<br />

1.0*10 -7<br />

1.0*10 -8<br />

1.0*10 -9<br />

Table 1: Model parameters used in the <strong>numerical</strong> simulations.<br />

parameter description value<br />

Kef effective conductivity 7.19·10 -5 m s -1<br />

2<br />

σ Y<br />

ln(K)-variance 2.7<br />

lY integral scale 2.67 m<br />

n porosity 0.33<br />

αL longitudinal dispersivity 0.25 m<br />

αT transverse dispersivity 0.05 m<br />

I hydraulic gradient 0.003<br />

λ first order degradation rate constant 1.59·10 -8 s -1<br />

Investigation <strong>of</strong> synthetic sites<br />

The plumes were independently investigated by 85 different test persons, engaged in hydrogeological <strong>and</strong> environmental<br />

research, consulting or administration. These investigators were confronted with the scenario <strong>of</strong> contaminant migration<br />

downstream from a source zone in a virtual aquifer. A two-dimensional top view <strong>of</strong> the site, the mean ground water <strong>flow</strong><br />

direction <strong>and</strong> the approximate location <strong>of</strong> the contaminant source were given as the only prior information for the site<br />

investigation. Neither the plume nor the hydraulic heads calculated are known at this stage. The task <strong>of</strong> the investigators<br />

was an as exact as possible characterization <strong>of</strong> the contaminated aquifer by the following procedure:<br />

• Step 1) Emplacing observation wells into the virtual aquifer: Using an interactive graphical user interface, the<br />

investigator positions observation wells on the virtual site. At the wells local contaminant concentrations <strong>and</strong><br />

hydraulic heads are measured.<br />

• Step 2) Regionalization <strong>of</strong> local measurements: Using different interpolation schemes (e.g. Kriging or Inverse<br />

Distance Weighting) the investigator interpolates contaminant concentrations <strong>and</strong> hydraulic heads measured at<br />

the observation wells to the virtual aquifer.


Steps 1 <strong>and</strong> 2 could be repeated as <strong>of</strong>ten as desired by the investigator, until the interpolated contaminant plume was<br />

deemed to be investigated accurately enough to properly characterize the contaminant distribution. Thus the interactive<br />

site investigation is an iterative procedure <strong>and</strong> is evaluated by a comparison <strong>of</strong> the “true” plume with the investigation<br />

result. More details on this procedure can be found in Chen et al. (2005).<br />

Using this methodology, a total number <strong>of</strong> 85 individual investigation results were obtained. 47 <strong>of</strong> these investigations<br />

were conducted for plumes originating from a source <strong>of</strong> width 4 m, while the plumes with a source width <strong>of</strong> 16 m were<br />

investigated 38 times. Accordingly, the majority <strong>of</strong> the 20 different plumes were investigated three or four times by<br />

different investigators.<br />

The configuration <strong>and</strong> the number <strong>of</strong> monitoring wells varies substantially between the different realizations <strong>and</strong> even for<br />

the same realization but different investigators (12 – 93 wells), as the decision about how many wells were needed for an<br />

accurate characterization <strong>of</strong> the site was left to the individual investigators. For two contaminant plumes (one for each<br />

source width) the investigation was repeated 13 times. Each investigation was performed by a different investigator. In<br />

addition to the general comparison <strong>of</strong> strategies A <strong>and</strong> B for the inference <strong>of</strong> the degradation rate constant, this subset <strong>of</strong><br />

the whole data set allows for an analysis <strong>of</strong> estimated rate constant variability for a single site, resulting from different<br />

notions <strong>of</strong> “sufficient accuracy” <strong>of</strong> the virtual plume investigation.<br />

Rate constant estimation<br />

Using the monitoring networks installed by the investigators <strong>of</strong> the virtual plumes, first order rate constants were<br />

estimated following strategies A <strong>and</strong> B. Methods A1, A2 <strong>and</strong> A3 are used in strategy A <strong>and</strong> are described in detail in<br />

Bauer et al. (2006a). Therefore, only a brief description is given here:<br />

Method A1 (equation (3)) is based on the one-dimensional transport equation, considering advection <strong>and</strong> first order<br />

degradation only. The steady state solution for the concentration pr<strong>of</strong>ile is rearranged to yield λA1, i.e. the first order<br />

degradation rate constant for method A1:<br />

va ⎛ C(<br />

x)<br />

⎞<br />

λ A1<br />

= − ln⎜ ⎟<br />

∆x<br />

⎜ ⎟<br />

(3)<br />

⎝ C0<br />

⎠<br />

va [L T -1 ] is the transport velocity, ∆x [L] the distance between the observation wells <strong>and</strong> C0 <strong>and</strong> C(x) [M L -3 ] are<br />

upstream <strong>and</strong> downstream contaminant concentrations at the observation wells, respectively. λA1 can be considered rather<br />

an overall or bulk attenuation rate than a degradation rate constant (Newell et al. 2002), as all concentration changes due<br />

to diffusion, dispersion, volatilization or dilution are attributed to the degradation process.<br />

The method introduced by Buscheck <strong>and</strong> Alcantar (1995) is the second approach applied in this study (equation (4)). It is<br />

based on the steady state solution <strong>of</strong> the one-dimensional transport equation considering advection, longitudinal<br />

dispersion <strong>and</strong> first order degradation. Method A2 requires an estimate <strong>of</strong> longitudinal dispersivity αL [L] <strong>and</strong> yields a<br />

“hybrid” rate constant between a bulk attenuation <strong>and</strong> a pure biodegradation rate constant, as the effects <strong>of</strong> transverse<br />

dispersion are still reflected in the rate constant estimate λA2.<br />

⎛<br />

2<br />

v<br />

( ) ⎞<br />

a ⎜⎛<br />

ln C(<br />

x)<br />

C0<br />

⎞<br />

λ = ⎜1<br />

− 2<br />

⎟ −1⎟<br />

A2<br />

α L<br />

4α<br />

⎜<br />

⎟<br />

L ⎝⎝<br />

∆x<br />

⎠ ⎠<br />

(4)<br />

Zhang <strong>and</strong> Heathcote (2003) proposed modifications to the method <strong>of</strong> Buscheck <strong>and</strong> Alcantar (1995) to improve the<br />

estimation <strong>of</strong> λ with regard to transverse dispersion. Correction terms derived from analytical solutions to the two- <strong>and</strong><br />

three-dimensional advection dispersion equations including first order decay are used to account for lateral spreading <strong>and</strong><br />

the width <strong>of</strong> the source zone WS [L] in two <strong>and</strong> three dimensions, respectively. Therefore information about WS, αL <strong>and</strong><br />

αT are required for this approach. Method A3 (equation (5)) is the two-dimensional form <strong>of</strong> the method by Zhang <strong>and</strong><br />

Heathcote (2003) <strong>and</strong> yields the biodegradation rate constant estimate λA3:<br />

2<br />

v ⎛ ( ) ⎞<br />

a ⎜⎛<br />

ln C(<br />

x)<br />

( C0β<br />

) ⎞<br />

⎛ ⎞<br />

λ = − ⎟<br />

A3<br />

⎜<br />

⎜1−<br />

2α<br />

L<br />

⎟ 1 with<br />

⎜ WS<br />

β = erf ⎟ (5)<br />

4α<br />

⎟<br />

L ⎝⎝<br />

∆x<br />

⎠<br />

⎜ ⎟<br />

⎠<br />

⎝ 4 αT<br />

∆x<br />

⎠<br />

For evaluation strategy B, method B (equation (6)) is used, which corresponds to the approach <strong>of</strong> Stenback et al. (2004).<br />

An approximate solution for the steady state concentration distribution derived from the two-dimensional advectiondispersion<br />

equation with first order degradation (Domenico <strong>and</strong> Schwartz 1990) is fitted to measured concentrations:<br />

C ⎪⎧<br />

⎛<br />

⎞⎪⎫<br />

⎪<br />

⎧ ⎛ ⎞ ⎛ ⎞⎪<br />

⎫<br />

0 ⎛ x ⎞<br />

⎨<br />

⎜ + ⎟ ⎜ −<br />

⎨ ⎜<br />

4λ<br />

⎟⎬<br />

−<br />

⎟<br />

⎜<br />

⎟<br />

Bα<br />

L 2y<br />

WS<br />

2y<br />

WS<br />

C(<br />

x,<br />

y)<br />

= exp 1−<br />

1+<br />

erf<br />

erf ⎬ (6)<br />

2 ⎪⎩ ⎝ 2α<br />

⎠<br />

⎜<br />

⎟<br />

⎪⎭ ⎪⎩<br />

⎜ ⎟ ⎜ ⎟<br />

L ⎝<br />

va<br />

⎠ ⎝ 4 αT<br />

x ⎠ ⎝ 4 αT<br />

x ⎠⎪⎭<br />

5


The approximate solution employed has the same basis as method 3. Here, however, a real two-dimensional approach is<br />

used, as equation (6) is fitted to the concentrations <strong>of</strong> all observation wells <strong>and</strong> not only to those measured along the<br />

center line.<br />

In strategy A the first step is an analysis <strong>of</strong> the concentration distribution based on all measurements made at the<br />

observation wells <strong>of</strong> the monitoring network. Starting at the well with the highest measured concentration, those<br />

downgradient wells are identified, that best represent the center line <strong>of</strong> the plume. These then are used to estimate the rate<br />

constant. Only locally measured quantities are used here, i.e. local hydraulic conductivities, hydraulic heads <strong>and</strong><br />

contaminant concentrations are “measured” at these wells by reading the model data at the respective nodes <strong>of</strong> the<br />

<strong>numerical</strong> grid. The transport velocity va between each pair <strong>of</strong> center line wells is approximated by:<br />

∆h<br />

va = K ef<br />

(7)<br />

n∆x<br />

with Kef the effective conductivity <strong>of</strong> local hydraulic conductivities at up <strong>and</strong> down gradient wells, n the porosity, ∆h the<br />

head difference <strong>and</strong> ∆x the distance between the wells. According to Rubin (2003) in stationary isotropic twodimensional<br />

domains with gaussian probability density functions <strong>of</strong> Y = ln(K) the effective conductivity Kef can be<br />

calculated as the geometric mean (cf. Bauer et al. 2006a). The porosity is assumed to be known correctly. With va, λ can<br />

then be calculated for each pair <strong>of</strong> center line wells using methods A1 – A3. For a set <strong>of</strong> k center line wells thus k-1 rate<br />

constants are calculated for one method. These are averaged to yield an estimate <strong>of</strong> the mean degradation rate constant λ.<br />

As an alternative to using only the locally measured conductivities, rate constant estimation is also performed using a<br />

global estimate <strong>of</strong> Kef, which is obtained from the geometric mean value <strong>of</strong> all hydraulic conductivities measured at all<br />

observation wells.<br />

In strategy B, method B (equation (6)) is fitted to measured concentrations <strong>of</strong> all observation wells <strong>of</strong> the monitoring<br />

network. Both the biodegradation rate constant λ <strong>and</strong> the source concentration C0 are varied simultaneously to achieve<br />

correspondence <strong>of</strong> measured <strong>and</strong> calculated concentrations, as a preliminary analysis (in agreement with results <strong>of</strong><br />

Stenback et al. (2004)) showed that this procedure on average yields closer estimates <strong>of</strong> the true rate constant than fitting<br />

only λ with a single fixed estimate <strong>of</strong> the source concentration. A least squares criterion for the concentration residuals is<br />

used in the fitting procedure. As in strategy A the <strong>flow</strong> velocity va is approximated using equation (7). Kef is calculated as<br />

the geometric mean <strong>of</strong> hydraulic conductivities measured at all wells <strong>of</strong> the network. The average hydraulic gradient over<br />

the entire site is approximated by fitting a linear trend surface to all head measurements by ordinary least squares<br />

regression.<br />

For methods A2, A3 <strong>and</strong> B estimates <strong>of</strong> longitudinal <strong>and</strong> transverse dispersivities are required. Practical guidance on<br />

estimating these parameters at the field scale is given e.g. by Wiedemeier et al. (1999), where one suggestion is to use 0.1<br />

times the plume length for αL <strong>and</strong> αT as 0.1 αL. Here, however, an alternative strategy is employed: Macrodispersivities<br />

2<br />

αL <strong>and</strong> αT are derived from correlation scale, aquifer heterogeneity σ Y <strong>and</strong> travel distance (Dagan 1984; Hsu 2003;<br />

Rubin et al. 2003). Thus, αL is taken as 7 m, which roughly corresponds to the large time asymptotic limit for the given<br />

conductivity distribution, while αT is taken as the approximate peak value <strong>of</strong> transverse macrodispersivity, calculated as<br />

0.7 m (which thus also corresponds to the frequently used relationship as αT ≈ 0.1 αL). These values are well within the<br />

ranges <strong>of</strong> dispersivities commonly used for the field scale <strong>modeling</strong> <strong>of</strong> contaminant transport. The true value <strong>of</strong> the<br />

source width WS is assumed to be known from the site investigation. These approximations <strong>and</strong> assumptions were made<br />

to ensure that the error introduced in estimated rate constants due to the parameterization <strong>of</strong> methods A2, A3 <strong>and</strong> B is as<br />

small as possible.<br />

Results <strong>and</strong> Discussion<br />

Strategy A - One-dimensional center line approach<br />

The rate constants λA1 - λA3 estimated with equations (3) – (5) are divided by the “true” value used in the <strong>numerical</strong><br />

simulations to yield normalized rate constants ΛA1 - ΛA3, which can directly be interpreted as overestimation or<br />

underestimation factors. Results in terms <strong>of</strong> mean values, medians, st<strong>and</strong>ard deviations <strong>and</strong> coefficients <strong>of</strong> variation (cv)<br />

as well as the number <strong>of</strong> realizations (N) used for strategy A are presented in Table 2 <strong>and</strong> Figure 3. Comparing methods<br />

A1 - A3 for the small source width WS = 4 m yields that all approaches on average result in a distinct overestimation <strong>of</strong> λ.<br />

For method A1 λ is overestimated on average by a factor <strong>of</strong> 6.88, while for A2 an mean ΛA2 = 8.24 is observed. Hence,<br />

method A1 performs better than A2. Method A3 yields a slightly lower mean <strong>of</strong> ΛA3 = 6.82, while the spread <strong>of</strong> results<br />

for A3 is noticeably larger, as can also be seen by the higher cv. Three main error sources for the observed<br />

overestimation exist:<br />

• neglect <strong>of</strong> dispersion by method A1, which thus is attributed to the degradation process (λA1 represents a bulk<br />

attenuation rate constant)<br />

6


• deviation <strong>of</strong> sampling well locations from the true center line position, which causes the sampling <strong>of</strong> too low<br />

concentrations<br />

• unrepresentative estimates <strong>of</strong> the local va along the <strong>flow</strong> path, as the estimated λ increases <strong>and</strong> decreases with va<br />

(cf. equations (3) – (5)).<br />

These effects are discussed in detail in Bauer et al. (2006a). For method A2 the additional bias towards too large rate<br />

constants is a consequence <strong>of</strong> accounting only for αL in equation (4), as with a one-dimensional transport model<br />

longitudinal dispersion <strong>of</strong> a degrading contaminant results in a stronger spreading <strong>of</strong> the solute downstream <strong>and</strong><br />

consequently in higher concentrations along the center line <strong>of</strong> a steady state plume compared to an advection only case.<br />

Therefore a larger rate constant is needed to fit a given concentration decrease <strong>and</strong> the “hybrid” rate constant estimate ΛA2<br />

is always larger than the bulk attenuation rate constant ΛA1 (cf. Bauer et al. 2006a), which appears contradictory.<br />

Table 2: Normalized degradation rate constants estimated<br />

with strategy A.<br />

WS = 4 m<br />

method A1 method A2 method A3<br />

mean 6.88 8.24 6.82<br />

median 4.43 5.50 2.65<br />

stdv. 6.36 7.41 10.09<br />

cv 0.92 0.90 1.48<br />

N 47 47 47<br />

WS = 16 m<br />

method A1 method A2 method A3<br />

mean 4.46 5.33 4.99<br />

median 2.66 3.23 2.45<br />

stdv. 4.81 6.29 9.59<br />

cv 1.08 1.18 1.92<br />

N 38 38 36<br />

For the source width WS = 16 m all methods improve (Table 2, Figure 3). Still, method A1 shows the closest estimates <strong>of</strong><br />

the rate constant <strong>and</strong> the lowest variability <strong>of</strong> single realization results in comparison to methods A2 <strong>and</strong> A3. With<br />

method A3 for two out <strong>of</strong> 38 plumes no reasonable rate constant estimate is obtained. This effect is due to an overcorrection<br />

for transverse dispersion in the β term <strong>of</strong> equation (5), which can cause the corrected C(x) to be very close to<br />

or even larger than the respective upgradient concentration C0, yielding very small or even negative λA3.<br />

norm. degradation rate constant [-]<br />

100<br />

10<br />

1<br />

W S = 4 m W S = 16 m<br />

single realization result<br />

ensemble mean<br />

0.1<br />

A1 1 A2 2 A3 3<br />

A1 1 A2 2 A3 3<br />

method<br />

method<br />

Figure 3: Normalized degradation rate constants estimated with strategy A <strong>and</strong> methods A1, A2 <strong>and</strong> A3 for source widths<br />

<strong>of</strong> 4 m (left diagram) <strong>and</strong> 16 m (right diagram). Small symbols represent results <strong>of</strong> individual realizations, large symbols<br />

the mean <strong>of</strong> all realizations. Kef used for rate estimation is calculated from measurements at center line wells only.<br />

7


As Bauer et al. (2006a) demonstrated, the local <strong>flow</strong> velocity estimate is a crucial parameter for the estimation <strong>of</strong> the<br />

degradation rate constant. In a sensitivity study the authors found that inclusion <strong>of</strong> field scale information on hydraulic<br />

conductivity, which is more representative for the entire site, can improve the accuracy <strong>of</strong> estimated rate constants over<br />

usage <strong>of</strong> local information only. Therefore the center line data obtained in strategy A are re-evaluated using a field scale<br />

estimate <strong>of</strong> the effective conductivity at the investigated sites: Kef is calculated as the geometric mean <strong>of</strong> local<br />

conductivity values measured at all observation wells <strong>of</strong> the site <strong>and</strong> not only using those measured on the center line.<br />

Results for this approximation are presented in Table 3 <strong>and</strong> Figure 4. It is found that consideration <strong>of</strong> all conductivity<br />

measurements available improves the rate constants estimated for all three methods <strong>and</strong> both source widths. On average<br />

all estimates are closer to the true rate than those estimated using only local conductivity information. The average<br />

improvement is between a factor <strong>of</strong> 1.5 for WS = 4 m <strong>and</strong> a factor <strong>of</strong> two for WS = 16 m. Also the st<strong>and</strong>ard deviations<br />

from the mean values <strong>and</strong> the cv are lower. This finding is true for the settings investigated here, i.e. a correlation length<br />

smaller than the average distance between observation wells <strong>and</strong> a stationary K distribution. If the correlation length is on<br />

the order or longer than the average observation well distance or K is not stationary, this finding is probably not<br />

transferable.<br />

norm. degradation rate constant [-]<br />

100<br />

10<br />

1<br />

0.1<br />

Table 3: Normalized degradation rate constants estimated<br />

with strategy A, re-evaluated using a field scale estimate <strong>of</strong><br />

Kef.<br />

WS = 4 m<br />

method A1 method A2 method A3<br />

mean 4.75 5.73 4.14<br />

median 4.04 4.67 2.66<br />

stdv. 3.75 4.33 4.06<br />

cv 0.79 0.76 0.98<br />

N 47 47 47<br />

WS = 16 m<br />

method A1 method A2 method A3<br />

mean 2.16 2.53 2.42<br />

median 1.14 1.50 0.90<br />

stdv. 2.08 2.45 3.98<br />

cv 0.96 0.97 1.64<br />

N 38 38 36<br />

W S= 4 m W S = 16 m<br />

single realization result<br />

ensemble mean<br />

A1 1 A2 2 A3 3<br />

A1 1 A2 2 A3 3<br />

method<br />

method<br />

Figure 4: Normalized degradation rate constants estimated with strategy A <strong>and</strong> methods A1, A2 <strong>and</strong> A3 for source widths<br />

<strong>of</strong> 4 m (left diagram) <strong>and</strong> 16 m (right diagram), respectively. Kef used for rate estimation is calculated as a field scale<br />

estimate from all monitoring wells available for each investigated plume.<br />

8


An interesting observation is that the differences between the three methods <strong>of</strong> strategy A are not as distinct as observed<br />

in Bauer et al. (2006a). One explanation for this finding is that due to the larger number <strong>of</strong> observation wells used for the<br />

site investigation in this study, the plume center line positions are better identified on average <strong>and</strong> concentration samples<br />

are taken closer to the true plume axis. Methods A1, A2 <strong>and</strong> A3 show a different sensitivity on deviations <strong>of</strong><br />

measurement locations from the plume center line. This is because in method A1, the rate constant estimate is linearly<br />

related to ln( C( x)<br />

/ C0<br />

) / ∆x<br />

, while in method A2 this term appears in linear as well as squared form after rearrangement<br />

<strong>of</strong> equation (4). As a consequence, increasing deviations <strong>of</strong> observation well locations from the plume center line <strong>and</strong><br />

thus lower measured contaminant concentrations will result in increasingly stronger overestimation <strong>of</strong> λ by A2 relative to<br />

A1. For method A3, the correction factor β has to be taken into account additionally (cf. equation (5)). For significant<br />

deviations from the center line, however, the same effect as for A2 can be shown.<br />

In Figure 5 the degradation rate constants ΛA1 estimated with method A1 were plotted against the number <strong>of</strong> observation<br />

wells with relative concentrations C/C0 > 0.001 in the respective monitoring networks. A clear relationship between the<br />

number <strong>of</strong> wells <strong>and</strong> the accuracy <strong>of</strong> ΛA1 is neither observed for WS = 4 m nor for WS = 16 m. It seems, however, that the<br />

spread <strong>of</strong> estimated rate constants decreases slightly with increasing numbers <strong>of</strong> wells, but a larger number <strong>of</strong> samples,<br />

especially for numbers <strong>of</strong> monitoring wells > 30 would be required to derive more meaningful results. For the other<br />

methods A2 <strong>and</strong> A3, the similar observations are made (not shown here).<br />

norm. deg. rate constant Λ A1 [-]<br />

50<br />

10<br />

1<br />

0.1<br />

Y = -0.074 * X + 6.531; R<br />

0 20 40 60 80<br />

no. <strong>of</strong> wells with C/C0 > 0.001<br />

2 = 0.063<br />

Y = -0.043 * X + 3.174; R2 linear fits<br />

= 0.072<br />

Figure 5: Degradation rate constants ΛA1 for source widths <strong>of</strong> 4 m (grey diamonds) <strong>and</strong> 16 m (black crosses) versus the<br />

number <strong>of</strong> wells showing relative concentrations C/C0 > 0.001.<br />

9<br />

W S = 4 m<br />

W S = 16 m<br />

Strategy B - Two-dimensional evaluation<br />

Applying method B it is not always possible to obtain a rate constant λB > 0 when minimizing the sum <strong>of</strong> squared<br />

residuals <strong>of</strong> concentration. The concentration distribution calculated with the analytical model indicates absence <strong>of</strong><br />

contaminant degradation for the closest fit to the measured data for six out <strong>of</strong> 47 plumes when WS = 4 m. For the<br />

remaining 41 plumes, usage <strong>of</strong> method B on average yields ΛB = 4.51 (Table 4, Figure 6), which is considerably closer to<br />

the true rate constant than for methods A1 - A3 with using only local measurements <strong>of</strong> hydraulic conductivity along the<br />

center line (cf. Table 2). The st<strong>and</strong>ard deviation for method B, however, is slightly larger than for A1 <strong>and</strong> A2. With WS =<br />

16 m for five out <strong>of</strong> 38 plumes no ΛB > 0 is found. Here, the improvement over methods A1 – A3 is even more distinct<br />

with the mean ΛB = 1.92. Also the spread <strong>of</strong> single realizations is reduced for the larger source width. However, no<br />

definite improvement <strong>of</strong> ΛB over those obtained by strategy A using the field scale estimate <strong>of</strong> Kef (cf. Table 3) is<br />

observed. For WS = 4 m the mean ΛB is slightly lower than the mean ΛA1, but slightly larger than ΛA3. Also the variability<br />

<strong>of</strong> estimated ΛB is larger than for methods A1 – A3. For WS = 16 m the mean ΛB is slightly lower than for all methods <strong>of</strong><br />

strategy A, while the observed spread is only lower in comparison to A3.<br />

In the estimation <strong>of</strong> λ with method B the source concentration is included as a fitting parameter. On average the true<br />

source concentration is overestimated by a factor <strong>of</strong> three for WS = 4m, while for WS = 16m deviations are lower than 5<br />

%. As for strategy A a distinct relationship between the number <strong>of</strong> observation wells in the monitoring network <strong>and</strong> the<br />

quality <strong>of</strong> the degradation rate estimates is not observed.


Table 4: Normalized degradation rate<br />

constants estimated with method B.<br />

WS = 4 m WS = 16 m<br />

mean 4.51 1.92<br />

median 1.96 1.14<br />

stdv. 8.34 2.97<br />

cv 1.85 1.55<br />

N 41 33<br />

norm. degradation rate constant [-]<br />

100<br />

10<br />

1<br />

0.1<br />

WS = 4 m WS = 16 m<br />

Figure 6: Normalized degradation rate constants estimated with method B for source widths <strong>of</strong> 4 m <strong>and</strong> 16 m.<br />

10<br />

single realization result<br />

ensemble mean<br />

Variability <strong>of</strong> estimated rate constants for a single site<br />

For each source width, one single plume was repeatedly investigated 13 times by different investigators. For this subset<br />

<strong>of</strong> plumes the variability <strong>of</strong> estimated rate constants for a single site, resulting from different notions on “sufficient<br />

accuracy” <strong>of</strong> the plume investigation, is studied. As the incorporation <strong>of</strong> conductivity measurements from all observation<br />

wells <strong>of</strong> the monitoring network significantly improves rate constant estimation with methods A1, A2 <strong>and</strong> A3, this<br />

approach is used in this comparison. Results for strategies A <strong>and</strong> B are summarized in Table 5 <strong>and</strong> Figure 7.<br />

For source width WS = 4 m a slightly stronger overestimation <strong>of</strong> λ can be observed for strategy A (methods A1 - A3) in<br />

comparison to the results for the complete data set. Here, the mean ΛA1 resulting from 13 investigations <strong>of</strong> the single site<br />

is 5.26, while ΛA1 for the complete data set is 4.75. For methods A2 <strong>and</strong> A3 the same tendency <strong>of</strong> increased<br />

overestimation is found. The variability among the sets <strong>of</strong> 13 estimated ΛA1, ΛA2 <strong>and</strong> ΛA3, however, is much smaller. The<br />

cv here are only 0.3, 0.34 <strong>and</strong> 0.68, respectively. With strategy B, i.e. method B, a slightly smaller mean ΛB = 4.26 is<br />

found than for the complete data set <strong>and</strong> the cv <strong>of</strong> 0.78 here also is reduced. For a source width <strong>of</strong> 16 m, all four methods<br />

show a much lower mean value <strong>of</strong> normalized rate constants than for the complete data set. On average, method A1 <strong>and</strong><br />

A3 yield the correct result for the single investigated site, while method A2 <strong>and</strong> B show a slight overestimation.<br />

Variability <strong>of</strong> the estimated rate constants in terms <strong>of</strong> cv is lower than for the complete data set <strong>of</strong> WS = 16 m. For<br />

strategy B no clear dependence between the precision <strong>of</strong> the rate constant estimate <strong>and</strong> the absolute number <strong>of</strong> wells in<br />

the monitoring network nor the number <strong>of</strong> respective wells showing considerable contaminant concentrations (e.g.<br />

C/C0 > 0.001) could be observed. The absence <strong>of</strong> such an obvious relationship may be due to the low sample size <strong>of</strong> only<br />

13 investigations <strong>of</strong> both plume realizations. As described before, it seems, however, that the spread <strong>of</strong> estimated rate<br />

constants decreases slightly with increasing numbers <strong>of</strong> wells (not shown here).<br />

The stronger overestimation observed for WS = 4 m as well as the more precise estimation <strong>of</strong> λ for WS = 16 m do not<br />

allow for any general conclusions regarding the absolute magnitude <strong>of</strong> errors, as this might probably be an effect <strong>of</strong> the<br />

two individual realizations investigated <strong>and</strong> their respective plume geometry. The variability observed among the 13<br />

results for each <strong>of</strong> the two single plumes however shows, that for a single realization <strong>of</strong> a contaminant plume,<br />

substantially different investigation results, in this case degradation rate constants, may be obtained, when investigated<br />

by individual persons. For both, the narrow plume with a comparatively small source zone <strong>of</strong> 4 m width as well as for the<br />

wider plume with WS = 16 m, the variability <strong>of</strong> investigation results here has a magnitude <strong>of</strong> 25 % up to 69 % <strong>of</strong> the<br />

variability observed among different realizations <strong>of</strong> the plume, depending on the method used for rate constant<br />

estimation. This clearly shows, that a substantial part <strong>of</strong> rate constant estimation uncertainty is due to the individual<br />

configuration <strong>of</strong> the monitoring network.


norm. degradation rate constant [-]<br />

100<br />

10<br />

1<br />

0.1<br />

W S = 4 m W S = 16 m<br />

single realization result<br />

ensemble mean<br />

A1 1 A2 2 A3 3 4B<br />

A1 1 A2 2 A3 3 4B<br />

method<br />

method<br />

Figure 7: Comparison <strong>of</strong> normalized degradation rate constants estimated with methods A1 – A3 <strong>and</strong> B for two single<br />

plume realizations with different source widths (Ws = 4 m: left diagram; Ws = 16 m: right diagram). Both plume<br />

realizations were repeatedly <strong>and</strong> independently investigated by 13 individual persons each.<br />

Summary <strong>and</strong> Conclusions<br />

Table 5: Normalized degradation rate constants estimated with methods<br />

A1 – A3 <strong>and</strong> B for two individual plume realizations with different source<br />

widths (Ws = 4 m <strong>and</strong> 16 m). Both plume realizations were independently<br />

investigated by 13 individual persons each.<br />

WS = 4 m<br />

method A1 method A2 method A3 method B<br />

mean 5.26 6.25 4.56 4.26<br />

median 5.02 5.88 2.94 3.47<br />

stdv. 1.59 2.11 3.10 3.33<br />

cv 0.30 0.34 0.68 0.78<br />

N 13 13 13 13<br />

WS = 16 m<br />

method A1 method A2 method A3 method B<br />

mean 1.01 1.28 1.00 1.21<br />

median 0.97 1.15 0.87 1.20<br />

stdv. 0.31 0.51 0.48 0.51<br />

cv 0.31 0.40 0.48 0.42<br />

N 13 13 13 13<br />

In this study, the frequently used center line approach for estimation <strong>of</strong> degradation rate constants is compared to a<br />

methodology suggested by Stenback et al. (2004), which uses all concentration measurements downstream <strong>of</strong> the<br />

contaminant source zone to estimate the degradation rate. In a <strong>numerical</strong> experiment, both strategies are applied to a set<br />

<strong>of</strong> 85 synthetic contaminant plumes, subject to first order degradation, <strong>and</strong> investigated using extensive monitoring<br />

networks. Rate constants are estimated using concentrations, hydraulic heads <strong>and</strong> conductivities locally measured at the<br />

observation wells <strong>of</strong> the monitoring network.<br />

Results show that rate constants tend to be overestimated in general. Using the center line approach (strategy A) <strong>and</strong> a<br />

simplified observation well set up causes a general overestimation due to three factors:<br />

• biased estimates <strong>of</strong> the mean <strong>flow</strong> velocity,<br />

• inadequate parameterizations <strong>of</strong> longitudinal <strong>and</strong> transverse dispersivities <strong>and</strong><br />

• <strong>of</strong>f center line measurements.<br />

11


Comparing results for source widths <strong>of</strong> 4 <strong>and</strong> 16 m yields, that rate constant estimation improves significantly for wider<br />

plumes, as it is less likely, that observation wells are placed <strong>of</strong>f the center line. Estimated rate constants also become<br />

more accurate when more than only local information on hydraulic conductivity is used to approximate the effective<br />

conductivity along the <strong>flow</strong> path: Taking Kef as the geometric mean <strong>of</strong> all measured K values from the monitoring<br />

network reduces rate constant overestimation, although wells are not necessarily located on the <strong>flow</strong> path or even within<br />

the contaminant plume.<br />

For the two-dimensional approach (strategy B), sources <strong>of</strong> error are inadequate parameterizations <strong>of</strong> longitudinal <strong>and</strong><br />

transverse dispersivities <strong>and</strong> an erroneous approximation <strong>of</strong> the mean <strong>flow</strong> velocity. As all observation wells<br />

downgradient <strong>of</strong> the source zone are incorporated in the estimation procedure, <strong>of</strong>f center line measurements are not<br />

relevant for this strategy. However, when the contaminant plume shows significant me<strong>and</strong>ering, the applicability <strong>of</strong><br />

strategy B is hampered, as the analytical model which is fitted to the concentration measurements presumes a linear<br />

plume axis. This problem is observed for the small source width <strong>of</strong> 4 m, where overestimation by strategy B is<br />

comparable to the one-dimensional approaches using the global estimate <strong>of</strong> Kef. For larger source widths like the 16 m<br />

used in this study, however, strategy B yields closer estimates <strong>of</strong> the degradation rate constant than strategy A on<br />

average.<br />

One drawback <strong>of</strong> the two-dimensional approach is that source concentration <strong>and</strong> degradation rate constant have to be<br />

fitted simultaneously to obtain accurate estimates <strong>of</strong> the degradation rate constant. This proves problematic especially for<br />

the small source width <strong>of</strong> 4 m where for some realizations the source concentration is overestimated by up to 300 %. For<br />

the source width <strong>of</strong> 16 m, however, the source concentration is fitted precisely within 5% <strong>of</strong> the “true” source<br />

concentration. These results suggest that incorporation <strong>of</strong> <strong>of</strong>f center line information in the estimation <strong>of</strong> the degradation<br />

rate can improve results <strong>of</strong> the plume investigation considerably. Especially for wide plumes, the two-dimensional<br />

approach <strong>of</strong> Stenback et al. (2004) proves superior to the one-dimensional center-line investigation strategy, resulting in<br />

more precise estimates <strong>of</strong> the degradation rate constant.<br />

Studying the variability <strong>of</strong> estimated degradation rates due to different monitoring network configurations, i.e. the<br />

personal factor due to different “site investigators”, it is found that for both source widths the variability <strong>of</strong> investigation<br />

results <strong>of</strong> one realization ranges from 25 % up to 69 % <strong>of</strong> the variability observed among different realizations. This<br />

demonstrates that the configuration <strong>of</strong> the monitoring network following from the investigators individual site<br />

investigation approach represents a substantial part <strong>of</strong> the uncertainty <strong>of</strong> the estimated rate constant.<br />

Overestimation <strong>of</strong> the degradation rate at a contaminated site is a critical point if monitored natural attenuation is<br />

considered as an alternative to conventional engineered remediation measures because the overall NA potential is<br />

assessed too positive. As demonstrated by Beyer et al. (2006) plume lengths calculated with biased degradation rates will<br />

result in an underestimation <strong>of</strong> the contaminated regions <strong>of</strong> the aquifer. Such an application, however, might not be <strong>of</strong><br />

primary concern, when monitoring networks <strong>of</strong> sufficient density have already been established over a site. Nonetheless,<br />

a too high rate constant could falsely lead to the conclusion, that a plume is at steady state, when the present day length<br />

observed fits the calculated steady state plume length. In the assessment <strong>of</strong> contaminated sites, indication for a steady<br />

state plume is an important result, e.g. for the acceptance <strong>of</strong> NA as a remediation scheme. Also for well investigated sites,<br />

a reliable determination <strong>of</strong> contaminant degradation rates is important, when the rates are to be used as <strong>modeling</strong><br />

parameters, for comparison with other sites <strong>and</strong> for comparing <strong>and</strong> optimizing alternatives <strong>of</strong> remediation measures.<br />

Acknowledgements: This work is funded by the German Ministry <strong>of</strong> Education <strong>and</strong> Research (BMBF) under grant<br />

033 05 12 / 033 05 13 as part <strong>of</strong> the KORA priority program, sub-project 7.2. We wish to thank our project partners at the<br />

Christian-Albrechts-University Kiel Andreas Dahmke <strong>and</strong> Dirk Schäfer for their support in our research. We<br />

acknowledge the support <strong>of</strong> Uwe Wittmann, Iris Bernhardt <strong>and</strong> Ludwig Luckner in coordination <strong>of</strong> the project work. Last<br />

but not least we are grateful to Bernie Kueper <strong>and</strong> two anonymous reviewers for their thoughtful comments <strong>and</strong><br />

suggestions which considerably improved this paper.<br />

References:<br />

Aronson, D., <strong>and</strong> P.H. Howard. 1997. Anaerobic biodegradation <strong>of</strong> organic chemicals in groundwater: A summary <strong>of</strong><br />

field <strong>and</strong> laboratory studies, SRC TR-97-0223F, Science Center Report, Syracuse Research Corporation, New York.<br />

Bauer, S., C. Beyer, <strong>and</strong> O. Kolditz. 2006a. Assessing measurement uncertainty <strong>of</strong> first-order degradation rates in<br />

heterogeneous aquifers. Water Resources Research 42, no. 1: W01420. 10.1029/2004WR003878.<br />

Bauer, S., C. Beyer, C. Chen, J. Gronewold, <strong>and</strong> O. Kolditz. 2006b. Virtueller Aquifer (VA) - Computergestützte<br />

Bewertung von Erkundungs-, Sanierungs- und Monitoringstrategien im Hinblick auf das "Natural Attenuation" (NA)<br />

und "Enhanced Natural Attenuation" (ENA) -Potenzial kontaminierter Böden und Grundwässer. In Statusseminar<br />

des KORA-TV 7, 8.6.2006, Dresden. Gemeinsame Mitteilungen des DGFZ e.V. und seiner Partner, vol. 3/2006, 93-<br />

113. Dresden.<br />

Bear, J. 1972. Dynamics <strong>of</strong> Fluids in porous Media. Amsterdam: Elsevier.<br />

Beyer, C., S. Bauer, <strong>and</strong> O. Kolditz. 2006. Uncertainty Assessment <strong>of</strong> Contaminant Plume Length Estimates in<br />

12


Heterogeneous Aquifers. Journal <strong>of</strong> Contaminant Hydrology 87, no. 1-2: 73-95. 10.1016/j.jconhyd.2006.04.006.<br />

Buscheck, T.E., <strong>and</strong> C.M. Alcantar. 1995. Regression techniques <strong>and</strong> analytical solutions to demonstrate intrinsic<br />

bioremediation. In Intrinsic Bioremediation, ed. R.E. Hinchee, T.J. Wilson, <strong>and</strong> D. Downey, 109-116. Columbus<br />

OH: Battelle Press.<br />

Chapelle, F.H., P.M. Bradley, D.R. Lovley, <strong>and</strong> D.A. Vroblesky. 1996. Measuring rates <strong>of</strong> biodegradation in a<br />

contaminated aquifer using field <strong>and</strong> laboratory methods. Ground Water 34, no. 4: 691-698.<br />

Chapelle, F.H., Widdowson, M.A., Brauner, J.S., Mendez, E. <strong>and</strong> C.C. Casey. 2003. Methodology for estimating times <strong>of</strong><br />

remediation associated with monitored natural attenuation. USGS Water-Resources Investigations Report 03-4057.<br />

Chen, C., C. Beyer, S. Bauer, <strong>and</strong> O. Kolditz. 2005. Interactive visual framework to demonstrate the uncertainty <strong>of</strong><br />

contaminant plume investigation. In 5. Workshop Porous Media, 02.-03.12.2004 Blaubeuren, Workshop<br />

Proceedings, Center for <strong>Applied</strong> Geosciences, University <strong>of</strong> Tübingen. http://www.virtual-aquifer.unituebingen.de/pdf/2005_29_chen.pdf<br />

Dagan, G. 1984. Solute transport in heterogeneous porous formations. Journal <strong>of</strong> Fluid Mechanics 145: 151–177.<br />

Domenico, P.A., <strong>and</strong> F.W. Schwartz. 1990. Physical <strong>and</strong> Chemical Hydrogeology, 2nd ed. New York: John Wiley <strong>and</strong><br />

Sons.<br />

Hsu, K.-C. 2003. The influence <strong>of</strong> the log-conductivity autocovariance structure on macrodispersion coefficients. Journal<br />

<strong>of</strong> Contaminant Hydrology 65, no. 1-2: 65-77. 10.1016/S0169-7722(02)00231-0.<br />

Islam J., S. Naresh, <strong>and</strong> M. O’Sullivan. 2001. Modeling biogeochemical processes in leachate-contaminated soils: A<br />

review. Transport in Porous Media 43, no. 3: 407 – 440. 10.1023/A:1010737825232.<br />

Kolditz, O., <strong>and</strong> S. Bauer, 2004. A process-oriented approach to computing multifield problems in porous media. Journal<br />

<strong>of</strong> Hydroinformatics 6, no. 3: 225-244.<br />

Kolditz, O., M. Beinhorn, M. Xie, T. Kalbacher, S. Bauer, W. Wang, C. McDermott, C. Chen, C. Beyer, J. Gronewold,<br />

D. Kemmler, R. Walsh, Y. Du, C.H. Park, M. Hess, <strong>and</strong> C. Büurger <strong>and</strong> J.O. Delfs. 2006. GeoSys/Rock<strong>flow</strong><br />

version 4.4.03 - Theory <strong>and</strong> users manual. Center for <strong>Applied</strong> Geoscience, University <strong>of</strong> Tübingen.<br />

McGrath, W.A., <strong>and</strong> G.F. Pinder. 2003. Search strategy for groundwater contaminant plume delineation. Water<br />

Resources Research 39, no. 10: 1298. 10.1029/2002WR001636.<br />

McNab Jr., W.W., <strong>and</strong> B.P. Dooher. 1998. A critique <strong>of</strong> a steady-state analytical method for estimating contaminant<br />

degradation rates. Ground Water 36, no. 6: 983–987.<br />

Newell, C.J., H.S. Rifai, J.T. Wilson, J.A. Connor, J.A. Aziz, <strong>and</strong> M.P. Suarez. 2002. Calculation <strong>and</strong> use <strong>of</strong> first-order<br />

rate constants for monitored natural attenuation studies. U.S. EPA Ground Water Issue, U.S. EPA/540/S-02/500.<br />

Rehfeldt, K.R., J.M. Boggs, <strong>and</strong> L.W. Gelhar. 1992. Field study <strong>of</strong> dispersion in a heterogeneous aquifer, 3, geostatistical<br />

analysis <strong>of</strong> hydraulic conductivity. Water Resources Research 28, no. 12: 3309–3324. 10.1029/92WR01758.<br />

Rittmann, B.E. 2004. Definition, objectives, <strong>and</strong> evaluation <strong>of</strong> natural attenuation. Biodegradation 15, no. 6: 349-357.<br />

10.1023/B:BIOD.0000044587.05189.99.<br />

Rittmann, B.E., <strong>and</strong> J.M. VanBriesen. 1996. Microbiological processes in reactive <strong>modeling</strong>. In Reactive Transport in<br />

Porous Media. Reviews in Mineralogy, vol. 34, ed. P. Lichtner, C. Steefel, <strong>and</strong> E. Oelkers, 311-334. Mineralogical<br />

Society <strong>of</strong> America: Washington DC.<br />

Rubin, Y., A. Bellin, <strong>and</strong> A.E. Lawrence. 2003. On the use <strong>of</strong> block-effective macrodispersion for <strong>numerical</strong> simulations<br />

<strong>of</strong> transport in heterogeneous formations. Water Resources Research 39, no. 9: 1242, 10.1029/2002WR001727.<br />

Stenback, G.A., S.K. Ong, S.W. Rogers, <strong>and</strong> B.H. Kjartanson. 2004. Impact <strong>of</strong> transverse <strong>and</strong> longitudinal dispersion on<br />

first-order degradation rate constant estimation. Journal <strong>of</strong> Contaminant Hydrology 73, no. 1-4: 3-14.<br />

10.1016/j.jconhyd.2003.11.004<br />

Storck, P., J.W. Eheart, <strong>and</strong> A.J. Valocchi. 1997. A method for the optimal location <strong>of</strong> monitoring wells for detection <strong>of</strong><br />

groundwater contamination in three-dimensional aquifers. Water Resources Research 33, no. 9: 2081–2088.<br />

10.1029/97WR01704.<br />

Sudicky, E.A.. 1986. A natural gradient experiment on solute transport in a s<strong>and</strong> aquifer: Spatial variability <strong>of</strong> hydraulic<br />

conductivity <strong>and</strong> its role in the dispersion process. Water Resources Research 22, no. 13: 2069-2082.<br />

Wiedemeier, T.H., M.A. Swanson, J.T. Wilson, D.H. Kampbell, R.N. Miller, <strong>and</strong> J.E. Hansen. 1996. Approximation <strong>of</strong><br />

biodegradation rate constants for monoaromatic hydrocarbons (BTEX) in ground water. Ground Water Monitoring<br />

& Remediation 16, no. 3: 186-194.<br />

Wiedemeier, T.H., H.S. Rifai, T.J. Wilson, <strong>and</strong> C. Newell. 1999. Natural Attenuation <strong>of</strong> Fuels <strong>and</strong> Chlorinated Solvents<br />

in the Subsurface. New York: Wiley.<br />

Wilson, R.D., S.F. Thornton, <strong>and</strong> D.M. Mackay. 2004. Challenges in monitoring the natural attenuation <strong>of</strong> spatially<br />

variable plumes. Biodegradation 15, no. 6: 459-469. 10.1023/B:BIOD.0000044591.45542.a9.<br />

Wu, J., C. Zheng, C.C. Chien, <strong>and</strong> L. Zheng. 2006. A comparative study <strong>of</strong> Monte Carlo simple genetic algorithm <strong>and</strong><br />

noisy genetic algorithm for cost-effective sampling network design under uncertainty. Advances in Water Resources<br />

29, no. 6: 899–911. 10.1016/j.advwatres.2005.08.005<br />

Zhang, Y.-K., <strong>and</strong> R.C. Heathcote. 2003. An improved method for estimation <strong>of</strong> biodegradation rate with field data.<br />

Ground Water Monitoring & Remediation 23, no. 3: 112-116.<br />

13


Enclosed Publication 6<br />

Beyer, C., Konrad, W., Park, C.H., Bauer, S., Rügner, H., Liedl, R., Grathwohl, P. (2007b):<br />

Modellbasierte Sickerwasserprognose für die Verwertung von Recycling-Baust<strong>of</strong>f in<br />

technischen Bauwerken. (Model based prognosis <strong>of</strong> contaminant leaching for reuse <strong>of</strong><br />

demolition waste in construction projects.) (accepted by Grundwasser <strong>and</strong> published online<br />

via SpringerLink), doi:10.1007/s00767-007-0025-x<br />

The enclosed article is made available with the permission <strong>of</strong> Springer <strong>and</strong> was published in<br />

the journal Grundwasser, online. Copyright © 2007 Springer.<br />

The article can be obtained online via SpringerLink at<br />

http://www.springerlink.com/openurl.asp?genre=journal&eissn=1432-1165.


Modellbasierte Sickerwasserprognose für die Verwertung von<br />

Recycling-Baust<strong>of</strong>f in technischen Bauwerken *<br />

Model based prognosis <strong>of</strong> contaminant leaching for reuse <strong>of</strong> demolition waste<br />

in construction projects<br />

Christ<strong>of</strong> Beyer 1 , Wilfried Konrad 1 , Hermann Rügner 2 , Sebastian Bauer 1 , Park Chan Hee 1 , Rudolf<br />

Liedl 3 , Peter Grathwohl 1<br />

1 Eberhard-Karls-Universität Tübingen, Zentrum für Angew<strong>and</strong>te Geowissenschaften (ZAG), Sigwartstraße 10, 72076<br />

Tübingen; Telefon: 07071-29 73176, Telefax: 07071-5059, E-mail: christ<strong>of</strong>.beyer@uni-tuebingen.de<br />

2<br />

Umweltforschungszentrum Leipzig-Halle (UFZ), Permoserstraße 15,<br />

04318 Leipzig<br />

3<br />

TU Dresden, Institut für Grundwasserwirtschaft, 01062 Dresden<br />

Header: Sickerwasserprognose für Recycling-Baust<strong>of</strong>f<br />

Kurzfassung:<br />

In dieser Studie wird die in der BBodSchV rechtlich etablierte Sickerwasserprognose in Bezug auf die<br />

Beurteilung der von Recycling-Baust<strong>of</strong>f-Verwertungen im Straßenbau ausgehenden Schadst<strong>of</strong>feinträge ins<br />

Grundwasser weiterentwickelt. Anh<strong>and</strong> numerischer reaktiver St<strong>of</strong>ftransportsimulationen für drei<br />

praxisrelevante Verwertungsszenarien (Parkplatz, Lärmschutzwall, Straßendamm) sowie eine Auswahl<br />

regionaltypischer Unterbodeneinheiten Deutschl<strong>and</strong>s werden zeitliche Konzentrationsverläufe verschiedener<br />

St<strong>of</strong>fklassen an der Grundwasseroberfläche berechnet. Der Durchbruchszeitpunkt konservativer Tracer wird<br />

allein von den hydraulischen Eigenschaften der Unterböden gesteuert, für organische Schadst<strong>of</strong>fe sind vor<br />

allem deren KOC-Werte und die Corg-Gehalte der Unterböden ausschlaggebend. Signifikante dispersive<br />

Konzentrationsverminderungen ergeben sich nur bei deutlicher Abnahme der Quellstärke vor dem<br />

Durchbruch der Konzentrationspeaks. Bei lang anhaltend hohen Quellkonzentrationen relativ zur<br />

Transportzeit bleiben die Konzentrationsdurchbrüche unvermindert. Biologischer Schadst<strong>of</strong>fabbau führt zu<br />

deutlich reduzierten Durchbruchskonzentrationen. Für die Szenarien Lärmschutzwall und Straßendamm<br />

werden Kapillarsperreneffekte beobachtet, die zu einem teilweisen Umfließen der Schadst<strong>of</strong>fquelle führen.<br />

Bei Berücksichtigung des am Recyclingmaterial vorbeiströmenden Sickerwassers durch Konzentrationsmittelung<br />

über die gesamte Bauwerksbreite ergeben sich Konzentrationsminderungen um 30-40%.<br />

Abstract:<br />

In this study contaminant leaching from recycling materials in road constructions to groundwater is assessed<br />

by the “Sickerwasserprognose”. Numerical transport simulations for three scenarios (parking lot, noise<br />

protection dam, road) <strong>and</strong> a number <strong>of</strong> characteristic subsoils <strong>of</strong> Germany are performed to estimate the<br />

breakthrough <strong>of</strong> different contaminant classes at the groundwater table. Conservative tracer breakthrough<br />

times (BTT) primarily depend on subsoil hydraulic properties, for organic pollutants Koc <strong>and</strong> subsoil OC are<br />

the controlling parameters. Significant concentration reductions from dispersion only occur when source<br />

concentrations decrease prior to contaminant breakthrough. If source concentrations remain high for long<br />

periods relative to peak BTT, concentration breakthrough is undamped. Accounting for biodegradation<br />

reduces breakthrough concentrations significantly. For the scenarios "noise protection dam" <strong>and</strong> "road" capillary<br />

barrier effects cause the seepage water to partially bypass the recycling material. Accounting for this<br />

bypass <strong>flow</strong> <strong>and</strong> averaging spatially across the constructions reduces concentrations by about 30-40%.<br />

Keywords: ground water risk assessment; reuse; demolition waste; type-scenarios; road construction;<br />

modelling<br />

* Beyer, C., Konrad, W., Park, C.H., Bauer, S., Rügner, H., Liedl, R., Grathwohl, P. (2007b): Modellbasierte Sickerwasserprognose für<br />

die Verwertung von Recycling-Baust<strong>of</strong>f in technischen Bauwerken. (Model based prognosis <strong>of</strong> contaminant leaching for reuse <strong>of</strong> demolition<br />

waste in construction projects.) (accepted by Grundwasser), doi:10.1007/s00767-007-0025-x<br />

Der Artikel wurde von der Zeitschrift Grundwasser zur Veröffentlichung angenommen, online publiziert und mit Erlaubnis von Springer<br />

reproduziert. Copyright © 2007 Springer. Der Artikel ist online abrufbar via SpringerLink:<br />

http://www.springerlink.com/openurl.asp?genre=journal&eissn=1432-1165<br />

1


Einleitung und Zielsetzung<br />

In Deutschl<strong>and</strong> fallen jährlich etwa 250 Mio. t mineralischer Abfälle an, die in erheblichem Umfang im Erd-,<br />

Straßen- und Verkehrsflächenbau verwertet werden. Recycling-Baust<strong>of</strong>fe und Rückstände aus der<br />

industriellen Produktion (z. B. Hoch<strong>of</strong>enschlacken) oder aus der Abfallbeh<strong>and</strong>lung (z. B. Hausmüllverbrennungsaschen)<br />

finden unter <strong>and</strong>erem in Trag- und Frostschutzschichten beim Straßenbau, zum Bau von<br />

Lärmschutzwällen oder als Verfüllmaterial zunehmende Verwendung (Krass et al. 2004a; 2004b). Die<br />

Verwertung von Recycling-Baust<strong>of</strong>fen (RCB) in technischen Bauwerken wird in Deutschl<strong>and</strong> in<br />

Übereinstimmung mit der diesbezüglichen Politik der Europäischen Union gegenüber einer Deponierung<br />

prinzipiell vorgezogen (KrW-/AbfG 1994). Aufgrund der häufig vorliegenden Belastung von RCB durch<br />

organische und anorganische Schadst<strong>of</strong>fe (z.B. PAK, Salze, Schwermetalle) muss bei der Verwertung in<br />

technischen Bauwerken jedoch berücksichtigt werden, dass Inhaltsst<strong>of</strong>fe durch das Sickerwasser<br />

ausgewaschen werden und das Grund- und Oberflächenwasser belasten können. Aus diesem Grund ist<br />

eine Bewertung von Verwertungsmaßnahmen hinsichtlich ihrer Umweltauswirkungen notwendig. Das<br />

Schadst<strong>of</strong>faustragsverhalten technischer Bauwerke ist deshalb ein sowohl auf nationaler als auch auf<br />

internationaler Ebene aktuelles und intensiv bearbeitetes Forschungsgebiet. So führten Hjelmar et al. (2007)<br />

großskalige Feldstudien an einem Straßenabschnitt in Dänemark durch und beobachteten ein zeitliches<br />

Abklingen der Schadst<strong>of</strong>fquellstärke, welches bei der Umweltwirkungsprognose berücksichtigt werden sollte.<br />

Für einige Schadst<strong>of</strong>fe ergäben sich so weniger konservativen Prognosen, die weniger restriktive<br />

Grenzwerte erlauben würden, ohne den Schutz des Grundwassers zu gefährden (Hjelmar et al. 2007). Die<br />

Autoren wiesen zudem das Auftreten von Kapillarsperren in Straßendämmen nach, die zu einem Umfließen<br />

des Verwertungsmaterials und somit zu reduzierten Wasserflüssen durch die Schadst<strong>of</strong>fquelle führten.<br />

Kapillarsperreneffekte wurden auch von Hansson et al. (2006) bei numerischen Simulationen der<br />

Wasserströmung in Straßendämmen beobachtet. Susset (2007) führte Freil<strong>and</strong>lysimeteruntersuchungen mit<br />

RCB über Löss- und geringsorptiven S<strong>and</strong>bodenmonolithen durch und stellte einen Rückhalt von PAK über<br />

bisher 4 Jahre nach, wobei die Wasserdurchsatzraten übertragen auf Feldbedingungen mehreren<br />

Jahrzehnten entsprechen. Sulfatkonzentrationen an der Unterkante des RCB gingen im<br />

Beobachtungszeitraum kaum zurück und brachen in voller Höhe durch. Für Chlorid hingegen konnte eine<br />

dispersive Verdünnung der Maximalkonzentration beobachtet werden, da die „Lebensdauer“ der<br />

Chloridquelle deutlich kleiner als die Transportzeit durch die Lysimeter war. In umfangreichen Säulenexperimenten<br />

mit RCB über Böden unter feldnahen Bedingungen zeigten Stieber et al. (2006), dass<br />

organische Kontaminanten wie PAK in der ungesättigten Zone effektiv durch mikrobiellen Abbau eliminierbar<br />

sind, s<strong>of</strong>ern die dafür physiologisch notwendigen Milieubedingungen nicht durch das Sickerwasser nachteilig<br />

verändert werden.<br />

Die Vielzahl der zu berücksichtigenden hydraulischen, geochemischen und mikrobiologischen Prozesse<br />

sowie die Komplexität ihrer Interaktionen macht deutlich, dass die Anwendung von prozessbasierten<br />

Modellen bei der Bewertung von umwelt<strong>of</strong>fen verwerteten Recyclingmaterialien große Vorteile bietet. Vor<br />

diesem Hintergrund untersucht diese Arbeit in Anlehnung an die Methodik der Sickerwasserprognose nach<br />

BBodSchV (1999) anh<strong>and</strong> numerischer reaktiver St<strong>of</strong>ftransportsimulationen, welche Schadst<strong>of</strong>feinträge von<br />

im Straßenbau eingesetztem RCB über die ungesättigte Zone ins Grundwasser unter Berücksichtigung<br />

möglicher zeitabhängiger Festlegungs- und Abbauprozesse (Sorption, Intrapartikeldiffusion, Bioabbau)<br />

erfolgen können. Dazu wird hier erstmals eine Kopplung des Stromröhrenmodells SMART (Finkel 1998,<br />

Finkel et al. 1998) mit dem Finite-Elemente-Modell GeoSys/Rock<strong>flow</strong> (Kolditz & Bauer 2004; Kolditz et al.<br />

2006) eingesetzt. Die Kombination verschiedener repräsentativer Verwertungsszenarien (Parkplatz,<br />

Lärmschutzwall, Straßendamm), RCB-typischer Schadst<strong>of</strong>fklassen und regionaltypischer Unterböden führt<br />

zu einem umfangreichen Satz von Typ-Szenarien-Simulationen, deren Ergebnisse es erlauben,<br />

• den aus den Verwertungsszenarien zu erwartenden charakteristischen zeitlichen Verlauf der<br />

Schadst<strong>of</strong>feinträge in Boden und Grundwasser darzustellen,<br />

• die während des Transports für Schadst<strong>of</strong>frückhalt und Konzentrationsminderung relevanten<br />

Prozesse zu identifizieren,<br />

• Einflüsse der Unterbodeneigenschaften hinsichtlich des Filter- und Puffervermögens herauszustellen<br />

und<br />

• das Potential der modellbasierten Sickerwasserprognose praxisnah zu demonstrieren.<br />

Modellkonzept<br />

Die Berechnung der Schadst<strong>of</strong>feinträge aus dem RCB über die ungesättigte Zone ins Grundwasser wird in<br />

dieser Studie erstmalig mit Hilfe einer Kopplung zwischen dem Finite-Elemente-Modell GeoSys/Rock<strong>flow</strong><br />

2


und dem stochastischen Stromröhrenmodell SMART durchgeführt. Die Simulationsaufgabe wird hier in die<br />

zwei Teilprobleme Hydraulik/Strömung und reaktiver Transport zerlegt. Diese Vorgehensweise wurde<br />

gewählt, um die Auswirkung der räumlich variablen Sickerwasserströmung in den Verwertungsszenarien auf<br />

den Schadst<strong>of</strong>ftransport berücksichtigen sowie physikalisch realistische Schadst<strong>of</strong>ffreisetzungsprozesse in<br />

der Quelle und die durch Intrapartikeldiffusion zeitabhängige Sorption beim Transport quantifizieren zu<br />

können.<br />

Die Kopplung zwischen der Strömungssimulation mit GeoSys/Rock<strong>flow</strong> und der reaktiven<br />

Transportsimulation mit SMART erfolgt über Verteilungsfunktionen (pdf, "probability density function") der<br />

Verweilzeiten des Sickerwassers, welche die Einflüsse der hydraulischen Heterogenität auf das<br />

Strömungsfeld beschreiben (vgl. Utermann et al. 1990; Bold 2004). Die pdf sind Modellergebnis der<br />

Strömungssimulation und werden von SMART als Modelleingabe benötigt, um in der reaktiven<br />

Transportsimulation Massendurchbruchskurven an einer unterstromigen Kontrollebene (hier die<br />

Grundwasseroberfläche) zu berechnen. Da das SMART-Konzept zwar Transportsimulationen in heterogen<br />

zusammengesetzten Materialien erlaubt (Lithokomponenten-Ansatz nach Kleineidam et al. 1999a), entlang<br />

der Stromröhre jedoch eine homogene Materialverteilung vorausgesetzt wird, muss für jede Material- bzw.<br />

Modellschicht der aus mehreren Baumaterialien und Unterbodenhorizonten bestehenden Typ-Szenarien<br />

eine separate SMART-Simulation gerechnet werden. Das Modell ist somit als eine Reihe von hinterein<strong>and</strong>er<br />

geschalteten 1D-Stromröhren konzipiert (siehe Abb. 1).<br />

GeoSys/Rock<strong>flow</strong> SMART<br />

35cm<br />

Charakterisierung<br />

der Strömung<br />

konst. Infiltration<br />

Tragschicht<br />

RCB<br />

Körnung<br />

0/32<br />

Unterboden<br />

Untergrund<br />

Grundwasseroberfläche<br />

(GWO)<br />

Tracereingabe<br />

BTC Tragschicht<br />

Tracereingabe<br />

BTC Unterboden<br />

Tracereingabe<br />

BTC Untergrund<br />

pdf Tragschicht<br />

pdf Unterboden<br />

pdf Untergrund<br />

3<br />

reaktiver<br />

Transport<br />

Stromröhre<br />

Tragschicht<br />

Stromröhre<br />

Untergrund<br />

Modellergebnis<br />

BTC an GWO<br />

Abb. 1: Veranschaulichung des Modellkonzepts anh<strong>and</strong> des Parkplatzszenarios: Simulation der<br />

Tracerdurchbruchskurven zur Ableitung von Laufzeitverteilungen (pdf) der einzelnen Modellschichten mit<br />

GeoSys/Rock<strong>flow</strong>; reaktive Transportsimulationen mit SMART (BTC = Durchbruchskurve); Kopplung beider<br />

Modelle durch die pdf.<br />

Die modellierte Massendurchbruchskurve an der unterstromigen Kontrollebene jeder Stromröhre<br />

repräsentiert jeweils die oberstromige zeitabhängige Konzentrationsr<strong>and</strong>bedingung der nachfolgenden<br />

Stromröhre bzw. im Fall der untersten Modellschicht den Schadst<strong>of</strong>feintrag ins Grundwasser.<br />

Dementsprechend muss für jede Modellschicht eine eigene pdf abgeleitet werden. Aus Gründen der<br />

Modellvereinfachung werden die einzelnen pdf als unabhängig vonein<strong>and</strong>er betrachtet (vgl. V<strong>and</strong>erborght et<br />

al. 2007). Zur Ableitung der pdf wird in GeoSys/Rock<strong>flow</strong> für jede einzelne Materialschicht an der<br />

Schichtobergrenze ein konservativer Tracer mit konstanter Konzentration C0 in das Strömungsfeld<br />

eingegeben und dessen Durchbruch C(t) an der Schichtuntergrenze zeitlich registriert (siehe Abb. 1). Die pdf<br />

der Modellschicht ergibt sich als Ableitung der Durchbruchskurve dC/dt aufgetragen über die Zeit t.<br />

Die Kopplung beider Simulationsmodelle wurde als sogenannte lose Kopplung realisiert, d.h. die Quellcodes<br />

wurden nicht kombiniert sondern kommunizieren über Batch-Aufrufe und automatisch generierte Eingabe-<br />

/Ausgabedateien während bzw. nach der Programmausführung. Diese Art der Einbindung hat im Hinblick auf


die Anzahl Simulationsläufe den Vorteil, dass pro Kombination aus Verwertungsmaßnahme und<br />

Unterbodentyp nur eine Simulation des Strömungsfeldes zur Bestimmung aller pdf notwendig ist und diese<br />

jeweils für die darauf folgenden Transportsimulationen der einzelnen Schadst<strong>of</strong>fklassen mit SMART<br />

verwendet werden können.<br />

Numerische Simulation der Strömung und des reaktiven Transports<br />

Die Simulation der ungesättigten Strömung mit GeoSys/Rock<strong>flow</strong> zur Ableitung der pdf erfolgt auf Grundlage<br />

der Richards-Gleichung, zu deren Lösung das Van-Genuchten-Mualem-Modell (van Genuchten 1980;<br />

Mualem 1976) verwendet wird (Du et al. 2005). Auf eine umfassende Ausführung der allgemein bekannten<br />

Modellgleichung wird hier und im Folgenden verzichtet, eine ausführliche Erläuterung der verwendeten<br />

mathematischen Modelle ist in Grathwohl et al. (2006) zu finden. Die Berechnung der Tracerbewegung im<br />

ungesättigten Strömungsfeld zur Ableitung der pdf erfolgt auf Grundlage der Konvektions-<br />

Dispersionsgleichung. Die Beschreibung des reaktiven Transports mit SMART erfolgt dagegen wie oben<br />

beschrieben nicht entlang von Raumkoordinaten sondern in Abhängigkeit der Verweilzeit entlang einer<br />

Stromröhre (Finkel, 1998). Die Rückhalteprozesse beim Transport werden durch das in SMART integrierte<br />

numerische Modell BESSY (Jäger & Liedl 2000) quantifiziert. Die Berücksichtigung der den<br />

Sorptionsprozess zeitlich limitierenden Intrapartikel-Porendiffusion ist insbesondere für Böden mit hohem<br />

Kies- oder Skelettanteil notwendig (Grathwohl 1998; Rügner et al. 1997; 1999). Die Sorptionskinetik wird<br />

unter Annahme von sphärischen Partikeln durch das 2. Fick´sche Gesetz in Radialkoordinaten beschrieben:<br />

2<br />

∂ w ⎡∂<br />

Cw<br />

2 ∂ C<br />

= Da⎢<br />

+ 2<br />

∂ t ⎣ ∂ r r ∂ r<br />

C w<br />

r [m] bezeichnet den radialen Abst<strong>and</strong> vom Kornmittelpunkt, CW [kg m -3 ] die Konzentration in der wässrigen<br />

Phase der Intrapartikelporen und Da [cm 2 s -1 ] den scheinbaren Diffusionskoeffizienten, der gegenüber der<br />

Diffusion in Wasser (Daq) vermindert ist:<br />

D<br />

D<br />

ε<br />

4<br />

⎤<br />

⎥<br />

⎦<br />

(1)<br />

aq<br />

a = (2)<br />

( ε + Kd<br />

ρk<br />

) τ f<br />

ε [-] bezeichnet die Intrapartikelporosität, τf [-] den Tortuositätsfaktor, Kd [m -3 kg -1 ] den Gleichgewichts-<br />

Verteilungskoeffizienten und ρk [kg m -3 ] die Partikeldichte gemäß ρk = ( 1−<br />

ε )ρ mit ρ [kg m -3 ] der<br />

Mineraldichte.<br />

Die Verteilung der Schadst<strong>of</strong>fe zwischen dem in den Intrapartikelporen gelösten und dem sorbierten Anteil<br />

wird durch eine lineare Sorptionsisotherme beschrieben, für den Schadst<strong>of</strong>fabbau im mobilen Porenwasser<br />

wird eine Kinetik erster Ordnung angenommen. Somit ergibt sich unter stationären Fließbedingungen ein<br />

insgesamt lineares mathematisches Modell für die im Sickerwasser auftretenden Konzentrationen C und<br />

eine Proportionalität zwischen C und den anfänglich aus der Schadst<strong>of</strong>fquelle eluierenden Schadst<strong>of</strong>fkonzentrationen<br />

C0, bzw. den anfanglich sorbierten St<strong>of</strong>fmengen.<br />

Parametrisierung der Typ-Szenarien<br />

Im Rahmen dieser Studie werden drei repräsentative Verwertungsszenarien betrachtet: Parkplatz (PP),<br />

Lärmschutzwall (LSW) und Straßendamm (SD). Das in diesen zum Einsatz kommende Verwertungsmaterial<br />

besteht aus granularem RCB. Als Schadst<strong>of</strong>fe werden vier Modellsubstanzen verwendet, die in ihren<br />

Eigenschaften für RCB typische Schadst<strong>of</strong>fe repräsentieren: Naphthalin (NAP) und Phenanthren (PHE) als<br />

schwach bzw. mäßig sorbierende organische Schadst<strong>of</strong>fe, ein stark sorbierender für das durchschnittliche<br />

Verhalten der 15 EPA-PAK repräsentativer Summenparameter (Σ15 EPA-PAK) sowie ein konservativer<br />

Tracer als Beispiel leicht löslicher Salze wie Chlorid. Für die Transportstrecken unterhalb der Quellterme<br />

werden sechs regionaltypische Unterbodeneinheiten Deutschl<strong>and</strong>s berücksichtigt, um Charakteristika und<br />

Unterschiede hinsichtlich ihrer Filter- und Pufferkapazitäten herauszustellen. Die folgenden Abschnitte<br />

erläutern die Parametrisierung der Typ-Szenarien (für detailliertere Ausführungen siehe Grathwohl et al.<br />

2006).


Verwertungsszenarien und -materialien<br />

Das PP-Szenario wird als vertikales 1D-Modell betrachtet und orientiert sich am Querschnitt für<br />

Verkehrsflächen der Bauklasse VI mit Pflasterdecke und ungebundener Tragschicht auf einer<br />

Frostschutzschicht (FSS) nach FGSV (2001). Für Tragschicht und FSS, die im Modell als einheitliche<br />

Schicht von 0.35 m betrachtet werden, wird RCB mit einer Körnung der Sieblinie 0/32 (FGSV 2004)<br />

eingesetzt. Das Pflaster selbst wird nicht berücksichtigt, da die geforderte Versickerungsleistung von 2.7*10 -5<br />

m s -1 (FGSV 1998) deutlich über den hier angenommenen Infiltrationsraten liegt (s.u.). Abb. 1 zeigt neben<br />

dem Modellkonzept auch den für die Modellierung vereinfachten PP-Aufbau.<br />

Die Szenarien LSW und SD werden als 2D-Vertikalschnitte betrachtet, bei denen die Simulationen aus<br />

Symmetriegründen nur für jeweils eine Hälfte des Querschnitts durchgeführt wurden. Das LSW-Modell (Abb.<br />

2) orientiert sich an einem von Mesters (1993) experimentell untersuchten LSW. Als Bodenabdeckung des<br />

Wallkerns aus RCB wird ein Lehmboden verwendet.<br />

0.5m<br />

3.5m<br />

1m 6m<br />

2m<br />

RCB<br />

Körnung 0/32<br />

Unterboden<br />

Untergrund<br />

Symmetrieachse<br />

1:1. 5<br />

Planum<br />

Abb. 2: Lärmschutzwall mit Wallkern aus Recycling-Baust<strong>of</strong>f (RCB).<br />

0.5m<br />

5<br />

Bodenabdeckung<br />

kulturfähiger<br />

Lehmboden<br />

Grundwasseroberfläche<br />

(GWO)<br />

Das SD-Modell (Abb. 3) orientiert sich am Regelquerschnitt RQ 26 für vierstreifige Autobahnen (FGSV<br />

1996). Das Bankett wird nach FGSV (2005) als schwach durchlässiger Boden SŪ* ausgeführt (entspr.<br />

Bodenart Su3), die Bodenabdeckung der Böschung als stark durchlässiger S<strong>and</strong> (entspr. Bodenart Ss). Die<br />

hydraulischen Eigenschaften der für Bankett und Böschung verwendeten Böden sind in Tab. 1 aufgeführt.<br />

Unterhalb der Asphaltdeckschicht schließt sich eine ungebundene Tragschicht an, für die wie für den<br />

darunter liegenden Dammkern RCB der Sieblinie 0/32 angenommen wird.<br />

1.3m<br />

0.3m<br />

RCB<br />

Körnung 0/32<br />

Unterboden<br />

Untergrund<br />

Symmetrieachse<br />

10m 1.5m<br />

Asphaltdeckschicht<br />

undurchlässig<br />

3%<br />

Planum 4%<br />

12%<br />

schwach durchlässiger<br />

Boden<br />

1:1. 5<br />

2.3m<br />

10 cm stark durchlässiger<br />

Boden<br />

2m<br />

Grundwasseroberfläche<br />

(GWO)<br />

Abb. 3: Straßendamm mit Frostschutzschicht/Tragschicht aus Recycling-Baust<strong>of</strong>f (RCB).<br />

Zu den ungesättigten hydraulischen Eigenschaften von Recyclingmaterialien im Straßenbau finden sich in<br />

der Literatur kaum experimentelle Angaben. Deshalb wurden die Van-Genuchten-Parameter für den RCB in<br />

den einzelnen Verwertungsszenarien mit einem Ansatz von Arya & Paris (1981) bzw. Mishra et al. (1989) auf


Grundlage des Korngrößenspektrums der 0/32-Siebline, der Lagerungsdichte ρb = ( V ⋅ ρ p ) und der<br />

Porosität η = ( 1−<br />

ρb<br />

/ ρ)<br />

abgeleitet. V [-] ist der Verdichtungsgrad und ρp die Proctordichte [g cm -3 ] (siehe<br />

Tab. 1). Dieser Ansatz wurde für natürliche Böden entwickelt, sollte nach Hansson et al. (2006) jedoch auch<br />

für relativ grobkörnige Materialien wie Tragschichtschotter im Straßenbau geeignet sein. Für die gesättigte<br />

Leitfähigkeit Ks von Tragschichten wird ein Mindestwert von 5.4*10 -5 m s -1 gefordert (FGSV 1998). Aufgrund<br />

der großen Spannbreite experimentell in Labor und in-situ bestimmter Durchlässigkeiten (z.B. Wörner et al.<br />

2001; Stoppka 2002; Kellermann 2003) wird dieser Wert hier in allen Verwertungsszenarien für den RCB<br />

angenommen. Abb. 4 zeigt die θ-ψ- sowie die θ-K-Beziehungen des RCB für PP, LSW und SD. Die<br />

Kurvenverläufe sind durchweg sehr ähnlich und durch sehr geringe Kapillarität gekennzeichnet, was zu einer<br />

schnellen Drainage der Tragschichten erforderlich ist.<br />

Tab. 1: Hydraulische Eigenschaften der Baumaterialien für Parkplatz, Lärmschutzwall<br />

und Straßendamm<br />

Parkplatz Lärmschutzwall<br />

RCB RCB Böschung ‡<br />

6<br />

Straßendamm<br />

RCB Bankett † Böschung †<br />

θs 0.27 0.31 0.43 0.25 0.36 0.37<br />

θr 0.00 0.00 0.08 0.00 0.00 0.04<br />

n 1.29 1.29 1.56 1.29 1.28 1.57<br />

α [m -1 ] 166 154 3.60 146 2.64 8.74<br />

l 0.50 0.50 0.50 0.50 0.50 0.50<br />

ρ [g cm -3 ] ¶ 2.66 2.66 2.65 2.66 2.65 2.65<br />

ρb [g cm -3 ] 1.94 1.84 1.51 2.00 1.69 1.67<br />

ρp [g cm -3 ] ¶ 1.94 1.94 - 1.94 - -<br />

V [-] § 1 0.95 - 1.03 - -<br />

Ks [m s -1 ] 5.40*10 -5 5.40*10 -5 1.00*10 -6 5.40*10 -5 1.00*10 -6 2.90*10 -6<br />

‡ : Carsell & Parish 1988; † : Hennings 2000;<br />

§ : FGSV 2002 (PP), 1997 (LSW), 2004 (SD); ¶ : Kellermann 2003<br />

α, n, l: empirische Van-Genuchten-Parameter<br />

θr, θs: residualer und gesättigter Wassergehalt<br />

ρ, ρb, ρp: Mineral-, Lagerungs- und Proctordichte<br />

V: Verdichtungsgrad; Ks: gesättigte hydraulische Leitfähigkeit<br />

ψ [m]<br />

10 2<br />

10 1<br />

10 0<br />

10 -1<br />

10 -2<br />

10 -3<br />

10 -4<br />

Parkplatz<br />

Lärmschutzwall<br />

Straßendamm<br />

0 0.1 0.2 0.3 0.4<br />

θ [-]<br />

Krel [-]<br />

10 0<br />

10 -1<br />

10 -2<br />

10 -3<br />

10 -4<br />

10 -5<br />

10 -6<br />

10 -7<br />

10 -8<br />

10 -9<br />

10 -10<br />

0 0.1 0.2 0.3 0.4<br />

θ [-]<br />

Abb. 4: Wassergehalts-SAugspannungs- (links) sowie Wassergehalts-Leitfähigkeits-Beziehungen (rechts) für<br />

Recycling-Baust<strong>of</strong>f in den drei Verwertungsszenarien (ψ = Matrixpotential, Krel = relative Leitfähigkeit, θ =<br />

Wassergehalt).


Die Kd-Werte von NAP, PHE und Σ15 EPA-PAK im RCB von 106, 496 bzw. 1333 l kg -1 sowie die<br />

Intrapartikelporosität ε = 0.015 wurden von Henzler (2004) experimentell bestimmt. Daq für PHE und NAP<br />

wurde mit 7.86*10 -10 bzw. 9.15*10 -10 m² s -1 nach Hayduk & Laudie (1974) abgeschätzt. Die heterogene<br />

Zusammensetzung des RCB aus verschiedenen Korngrößenfraktionen wurde durch den Lithokomponenten-<br />

Ansatz nach Kleineidam et al. (1999a) berücksichtigt. Wegen sehr langer Rechenzeiten wird der RCB durch<br />

zwei Korngrößenklassen modelliert (siehe auch Rügner et al. 2005), für die die kinetische Sorption jeweils<br />

separat berechnet wird. Auf die Fein- und Grobfraktion entfallen Anteile von 32.8 % bzw. 67.2 % mit<br />

„effektiven“ Kornradien von a = 0.25 mm bzw. a = 8 mm (siehe Henzler (2004)).<br />

Auswahl und Klassifizierung der Unterbodenpr<strong>of</strong>ile<br />

Die Charakterisierung der Unterbodenpr<strong>of</strong>ile erfolgte auf Basis der nutzungsdifferenzierten Bodenübersichtskarte<br />

1:1.000.000 (BÜK1000; BGR 2006). Aus deren 672 Referenzpr<strong>of</strong>ilen wurden sechs Pr<strong>of</strong>ile nach<br />

Flächenrepräsentanz und dem zu erwartenden charakteristischen Transportverhalten für die numerischen<br />

Simulationen des reaktiven St<strong>of</strong>ftransports ausgewählt (Braunerde und Podsol aus S<strong>and</strong>, Fahlerde aus<br />

Geschiebelehm, Schwarzerde und Parabraunerde aus Löss, Pelosol aus verwittertem Mergel- und<br />

Tonstein). Zur Vereinfachung der Simulationen wurde eine Reduktion der vertikalen Pr<strong>of</strong>ildifferenzierung<br />

durch Zusammenfassung mehrerer Horizonte vorgenommen, soweit eine einheitliche Betrachtung nach<br />

bodenkundlichen und geologischen Aspekten möglich schien. Bei allen Pr<strong>of</strong>ilen wurden somit Daten aus<br />

verschiedenen Horizonten nach deren Mächtigkeit gewichtet gemittelt (Tab. 2). Als zu berechnende<br />

Pr<strong>of</strong>iltiefe wurde für die Böden die in der BÜK1000 (BGR 2006) beschriebene Pr<strong>of</strong>iltiefe ohne den<br />

Oberboden angesetzt, da dieser bei der Bebauung abgetragen wird. Die resultierenden Pr<strong>of</strong>iltiefen (Tab. 2)<br />

sind im Sinne von Mindesttiefen zu verstehen, da die in der Praxis relevanten Grundwasserflurabstände<br />

häufig deutlich größer sein dürften.<br />

Zur Festlegung der Ks-Werte und Van-Genuchten-Parameter wurden Pedotransferfunktionen von Wösten et<br />

al. (1998) angewendet. Der Sättigungswassergehalt θs wurde für alle Bodenschichten der Porosität<br />

gleichgesetzt. Ks bezieht sich auf den Feinbodenanteil (< 2 mm) und wurden bei höheren Skelettgehalten um<br />

die Reduktion der Leitfähigkeit durch den Skelettanteil mit einem Verfahren von Brakensiek & Rawls (1994)<br />

korrigiert. Für die Simulation der Tracerversuche in GeoSys/Rock<strong>flow</strong> zur Ableitung der pdf wurden die<br />

Dispersivitäten mit αL = 0.1 m und αT = 0.01 m festgelegt.<br />

Die Kd–Werte wurden aus dem Gehalt an organischem Kohlenst<strong>of</strong>f foc [-] (= 0.01 Corg; vgl. Tab. 2) und dem<br />

K = K f abgeschätzt. Für<br />

auf den Corg-Gehalt normierten Verteilungskoeffizienten KOC [l kg -1 ] durch d OC oc<br />

Horizonte, die nach BÜK1000 (BGR 2006) als „humusfrei“ ausgewiesen sind, wurde als konservative<br />

Annahme für den Feinboden ein minimaler Corg von 0.01 % angenommen. Zur Abschätzung des KOC wurde<br />

die auf der Wasserlöslichkeit S [mol l -1 ] beruhende Korrelation von Seth et al. (1999) verwendet (für<br />

vergleichbare Ansätze siehe auch Allen-King et al. 2002):<br />

log KOC = − 0.<br />

88 log S + 0.<br />

07 (3)<br />

S beträgt für PHE 3.46*10 -5 , für NAP 8.74*10 -4 mol l -1 . Für Σ15 EPA-PAK wurde S über ein mittleres<br />

Molgewicht von 202 g mol −1 und einer effektiven Wasserlöslichkeit von 2.5 g l -1 (Grathwohl 2004) mit 1.14<br />

*10 -5 mol l -1 abgeschätzt. Für die verschiedenen Böden ergeben sich so Kd-Werte für NAP zwischen 0.06<br />

und 5.10 l kg -1 , für PHE zwischen 0.99 und 87.2 l kg -1 sowie für Σ15 EPA-PAK zwischen 2.45 und 217 l kg -1<br />

Für die Skelettanteile wurden KOC-Werte von 4.0 l kg -1 (NAP), 5.2 l kg -1 (PHE) und 5.7 l kg -1 (Σ15 EPA-PAK)<br />

angenommen (Rügner et al. 2005). Mit diesen vergleichsweise hohen Werten wird berücksichtigt, dass es<br />

sich beim Corg dieser Komponenten idR. um gealtertes Material mit höherer Sorptionskapazität h<strong>and</strong>elt<br />

(Kleineidam et al. 1999b). Der foc wurde mit 0.0005 angesetzt (Kleineidam et al. 1999b).<br />

Die Sorptionskapazität und -kinetik für den Feinboden wird vor allem durch das partikuläre organische<br />

Material bestimmt. Es wurden folgende effektive Parameter zugrunde gelegt: a = 11.7 µm, ε = 0.00175,<br />

ρ = 2.65 g cm -3 und die Tortuosität τ f = 1/<br />

ε (Grathwohl, 1992; Rügner et al. 1999). Für die Grobfraktion (> 2<br />

mm) wurden a = 1 cm und ε = 0.01 angenommen (Rügner et al. 1999; Kleineidam et al. 1999a).<br />

Die Transportsimulationen für NAP, PHE und Σ15 EPA-PAK wurden jeweils mit und ohne Berücksichtigung<br />

von Bioabbau durchgeführt, wobei nur die in Lösung vorliegenden Anteile als abbaubar betrachtet wurden.<br />

Zur Beurteilung der langfristigen Filterwirkung des Bodens bedarf es repräsentativer Langzeit-<br />

Ratenkonstanten (Henzler et al. 2006), deren quantitative Abschätzung in der ungesättigten Zone jedoch<br />

durch die Komplexität der Wechselwirkungen zwischen schadst<strong>of</strong>fspezifischen Eigenschaften, klimatischen<br />

und geochemischen R<strong>and</strong>bedingungen (Grathwohl et al. 2003, Höhener et al. 2006) bisher kaum möglich<br />

ist. Aus diesem Grund wurde ein relativ niedriger Wert von 1.15*10 -7 s -1 (Halbwertzeit = 70 d) angenommen.<br />

7


Dies ist als konservative Abschätzung zu betrachten, da viele Schadst<strong>of</strong>fe in der ungesättigten Zone unter<br />

Feldbedingungen deutlich schneller abgebaut werden können (Maier & Grathwohl 2005, Rügner et al. 2005).<br />

Tab. 2: Bodeneigenschaften der sechs betrachteten Unterböden aus der BÜK1000 (BGR 2006).<br />

Für vertikal aggregierte Horizonte wurden die Parameter über die Mächtigkeit der einzelnen<br />

Horizonte gewichtet gemittelt.<br />

Bodentyp Braunerde Podsol Fahlerde<br />

8<br />

Schwarzerde <br />

Parabraunerde<br />

Pelosol<br />

Fläche [km²] 1672 7237 6933 3455 14428 5614<br />

Einteilung Unterboden Unterboden Unterboden Untergrund Unterboden Unterboden Unterboden<br />

von - bis [m] 0.00-1.70 0.00-1.70 0.00-0.90 0.90-1.70 0.00-1.70 0.00-1.70 0.00-1.75<br />

Ton [%] 3.82 10.40 31.85 31.85 19.69 19.78 39.88<br />

Schluff [%] 14.54 20.48 20.48 20.48 69.40 70.28 37.00<br />

S<strong>and</strong> [%] 57.41 63.21 38.68 38.68 8.44 8.43 16.48<br />

Skelett [%] 24.24 5.91 9.00 9.00 2.47 1.50 6.64<br />

Corg [%] 0.01 ‡ 0.21 0.01 ‡ 0.01 ‡ 0.88 0.01 ‡ 0.03<br />

ρb [g cm -3 ] 1.71 1.61 1.62 1.75 1.42 1.63 1.60<br />

θs 0.29 0.36 0.35 0.34 0.45 0.38 0.39<br />

†<br />

θr<br />

0.01 0.01 0.01 0.01 0.01 0.01 0.01<br />

n † 1.34 1.26 1.17 1.08 1.15 1.13 1.16<br />

α † [m -1 ] 5.93 5.98 6.21 4.95 1.76 1.24 2.73<br />

l † 1.77 -0.30 -0.66 -3.57 -1.91 -0.32 -2.75<br />

Ks † [m s -1 ] 1.52*10 -6 3.37*10 -6 3.09*10 -6 8.46*10 -7 3.83*10 -6 1.38*10 -6 9.48*10 -7<br />

‡ : 0.00 nach BÜK1000; Annahme eines minimalen Corg-Gehaltes des Feinbodens von 0.01%<br />

† : abgeleitet nach Wösten et al. (1998); Ks korrigiert um Skelettanteil nach Brakensieck & Rawls (1994)<br />

θr, θs: residualer und gesättigter Wassergehalt; ρb: Lagerungsdichte<br />

α, n, l: empirische Van-Genuchten-Parameter; Ks: gesättigte hydraulische Leitfähigkeit<br />

Abschätzung der Sickerwasserrate<br />

Für eine langfristige Beurteilung der Schadst<strong>of</strong>fverlagerung ist es sinnvoll und zulässig von vereinfachten<br />

Fließverhältnissen auszugehen (Grathwohl & Susset 2001; Henzler et al. 2006). Aus diesem Grund wird eine<br />

stationäre ungesättigte Strömung angenommen. Die jährlichen mittleren Infiltrationsraten I [mm a -1 ] der<br />

Verwertungsszenarien wurden als Mittelwerte für Deutschl<strong>and</strong> abgeschätzt. Hierzu wurde das BAGLUVA-<br />

Verfahren (Glugla et al. 2003) angewendet. Als mittlerer korrigierter Niederschlag Njahr wurde ein Wert von<br />

859 mm a −1 angenommen (BMU 2000). I ergibt sich als Differenz von Njahr und der Jahressumme der<br />

Evapotranspiration ET. Diese wird als Funktion der maximalen ET und st<strong>and</strong>ortspezifischer Parameter<br />

bestimmt. Für den PP wurde I so mit 583 mm a -1 abgeschätzt, was im oberen Bereich für sickerfähige<br />

Pflasterdecken liegt (Richter 2003). Die Annahme wird jedoch durch Freil<strong>and</strong>lysimeter-Untersuchungen von<br />

Flöter (2006) gestützt, in denen der Anteil von I am Niederschlag über 4 a im Mittel rund 80 % betrug. Für<br />

den LSW wurde I mit 313 mm a -1 abgeschätzt. Für den SD ergibt sich die zu infiltrierende Wassermenge als<br />

Summe des Oberflächenabflusses der Asphaltdecke (Abflussbeiwert = 0.9; FGSV 2005) und des auf<br />

Bankett und Böschung fallenden Niederschlags abzüglich der ET. Für die räumliche Verteilung der Infiltration<br />

mussten vereinfachende Annahmen getr<strong>of</strong>fen werden, da der SD-Querschnitt eine neue Bauweise nach<br />

FGSV (2005) darstellt, bei der der Straßenabfluss zu größeren Teilen über die Böschung infiltrieren soll,<br />

bisher jedoch keine experimentellen Daten dazu vorliegen. Aus diesem Grund wurde angenommen, dass die<br />

Versickerung gleichmäßig über Bankett und Böschung erfolgt. Auf den Querschnitt bezogen ergibt sich I mit<br />

2318 mm a -1 . Für den Böschungsfuß wurden wie für den LSW I = 313 mm a -1 angenommen.


Ergebnisse und Diskussion<br />

Die im Folgenden vorgestellten zeitlichen Konzentrationsverläufe der Modellsubstanzen an der GWO stellen<br />

über den unteren Modellr<strong>and</strong> integrierte Durchbruchskurven dar. Dabei ist für das LSW- und SD-Szenario<br />

zwischen den mit SMART berechneten, auf den Transportpfad bezogenen (graue dicke Kurven, Abb. 5, 8, 9)<br />

und den über den gesamten Querschnitt gemittelten Konzentrationen (schwarze dünne Kurven, Abb. 5, 8, 9)<br />

zu unterscheiden (siehe auch Abb. 6).<br />

Die Tracer-Durchbruchskurven an der GWO für die drei Szenarien PP, LSW und SD sind in Abb. 5 (a)-(c)<br />

dargestellt. Für den PP erfolgt der Durchbruch des Tracers am frühesten bei Braunerde und Podsol, später<br />

bei Pelosol, Parabraunerde und Schwarzerde (Abb. 5 (a)), die eine geringere<br />

Porenwasserfließgeschwindigkeit als die S<strong>and</strong>böden aufweisen. Dispersion führt zu Verminderungen der<br />

Durchbruchskonzentrationen auf C/C0 zwischen 0.23 (Braunerde) und 0.16 (Schwarzerde). Die hier<br />

betrachteten Transportstrecken von bis zu 1.75 m ab der RCB-Unterkante dürften idR. kürzer als die in<br />

vielen Praxisfällen relevanten Grundwasserflurabstände sein, wodurch unter Umständen eine stärkere<br />

dispersive Konzentrationsreduktion möglich wäre.<br />

C/C0 [-]<br />

0.3<br />

0.2<br />

0.1<br />

a) Parkplatz<br />

Braunerde<br />

Podsol<br />

Schwarzerde<br />

Parabraunerde<br />

Fahlerde<br />

Pelosol<br />

0<br />

0 1 2 3<br />

Zeit [a]<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

0 5 10 15<br />

Zeit [a]<br />

9<br />

b) Lärmschutzwall<br />

Legende siehe<br />

Straßendamm<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

c) Straßendamm<br />

Braunerde (verd.)<br />

Podsol (verd.)<br />

Schwarzerde (verd.)<br />

Braunerde<br />

Podsol<br />

Schwarzerde<br />

0 1 2 3<br />

Zeit [a]<br />

Abb. 5: Durchbruchskurven des Tracers an der Grundwasseroberfläche für (a) Parkplatz-, (b)<br />

Lärmschutzwall- und (c) Straßendamm-Szenario. Gezeigt werden Durchbruchskurven für den<br />

Schadst<strong>of</strong>ftransportpfad ((b) und (c), graue dicke Kurven) bzw. auf das Gesamtbauwerk bezogene<br />

Durchbruchskurven unter Berücksichtigung der Verdünnung durch nicht belastetes Sickerwasser ((b) und<br />

(c), schwarze dünne Kurven).<br />

Die über den unteren Modellr<strong>and</strong> integrierten Tracer-Durchbruchskurven für das LSW-Szenario (Abb. 5 (b))<br />

zeigen im Vergleich zum PP ein stärkeres Tailing und über längere Zeiträume anhaltende hohe<br />

Konzentrationen C/C0. Aus Anschaulichkeitsgründen werden die Ergebnisse hier nur für Braunerde, Podsol<br />

und Schwarzerde vorgestellt. Das ausgeprägte Tailing resultiert aus der geringeren Infiltrationsrate sowie<br />

aus der Geometrie des LSW. Die Überdeckung des grobkörnigen RCB mit einem feinkörnigen Lehm führt an<br />

der Böschung zu einer Kapillarsperre, welche die Infiltration in den RCB vermindert und ein Umströmen des<br />

Wallkerns verursacht (siehe Abb. 6 (a)).<br />

Die Kapillarsperre bildet sich aus, da der RCB aufgrund geringer Kapillarität bereits bei geringen<br />

Saugspannungen einen Großteil des Porenwassers verliert (siehe Abb. 4) und die hydraulische Leitfähigkeit<br />

gegenüber dem Lehm deutlich stärker abnimmt. Das Sickerwasser kann so entlang der geneigten<br />

Materialgrenze in dem nun besser wasserleitenden Lehmboden abfließen. Mit zunehmender Distanz von der<br />

Wallkrone erhöhen sich die Wasserflüsse im Lehm, sodass in Abhängigkeit der Druckverhältnisse<br />

zunehmend Wasser infiltrieren kann (Ross 1990). Die räumliche Variabilität der Sickerwasserströmung führt<br />

zu Bereichen mit stark reduzierten Fließgeschwindigkeiten und so zu einem langsameren Auswaschen des<br />

Tracers aus dem RCB. Die Maximalkonzentrationen erreichen die GWO nach etwa 2 (Braunerde) bis 3.5 a<br />

(Schwarzerde) (Abb. 5 (b)). Im Vergleich zum PP kommt es zudem zu höheren Peakkonzentrationen C/C0<br />

für den Transportpfad (Braunerde: 0.43, Schwarzerde: 0.35), da zum Zeitpunkt des Peak-Durchbruchs an<br />

der GWO die Eluatkonzentrationen der Quelle noch deutlich höher als beim PP sind (siehe Abb. 7 (a)) und<br />

so die dispersive Konzentrationsminderung weniger effektiv ist. Der länger anhaltende hohen<br />

Eluatkonzentrationen resultieren aus den geringeren Sickerwassermengen, der Reduktion der Infiltration in<br />

den RCB durch Kapillarsperren und der größere Mächtigkeit des RCB gegenüber dem PP. Auf das gesamte<br />

Bauwerk bzw. auf die gesamte infiltrierende Wassermenge bezogen reduzieren sich die<br />

Maximalkonzentrationen durch kleinräumige Mittelung jeweils um ca. 40 %. Zu beachten ist in diesem<br />

Zusammenhang, dass die Ausprägung und Effektivität einer Kapillarsperre von einer Reihe von


R<strong>and</strong>bedingungen abhängig ist. Aus diesem Grund sind die berechneten Verdünnungsfaktoren nur für die<br />

hier betrachteten Geometrien, Infiltrationsraten und Materialkombinationen quantitativ gültig.<br />

(a) Lärmschutzwall (b) Straßendamm<br />

Modellergebnisse bezogen auf den<br />

Transportpfad (SMART-Output)<br />

über Gesamtbreite gemittelte Modellergebnisse über Gesamtbreite gemittelte Modellergebnisse<br />

10<br />

Modellergebnisse bezogen auf den<br />

Transportpfad (SMART-Output)<br />

Abb. 6: Geschwindigkeitsvektoren der ungesättigten Strömung an allen Elementknoten im Lärmschutzwall<br />

(a) und in einem Ausschnitt des Straßendammes (b). Die Konzentration des Flusses in den<br />

Bodendeckschichten ist Resultat des Kapillarsperreneffektes, wird durch die feinere Netzdiskretisierung<br />

jedoch überzeichnet.<br />

C/C0 [-]<br />

1<br />

0.1<br />

0.01<br />

a) Tracer<br />

Parkplatz<br />

Lärmschutzwall<br />

Straßendamm<br />

0.01 0.1 1 10<br />

Zeit [a]<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

b) Parkplatz<br />

Naphthalin<br />

Phenanthren<br />

Σ15 EPA PAK<br />

0.1 1 10 100<br />

Zeit [a]<br />

Abb. 7: Zeitliche Entwicklung der aus der Schadst<strong>of</strong>fquelle eluierenden Konzentrationen an der Recycling-<br />

Baust<strong>of</strong>f-Unterkante (RCB) für ausgewählte Schadst<strong>of</strong>f-Verwertungsszenario-Kombinationen: (a)<br />

konservativer Tracer für Parkplatz-, Lärmschutzwall- und Straßendammszenario; (b) Naphthalin,<br />

Phenanthren und Σ15 EPA-PAK für das Parkplatzszenario.<br />

Auch für das SD-Szenario ist ein Kapillarsperreneffekt zu beobachten (Abb. 6(b)). Während sich im<br />

Bankettbereich eine relativ gleichförmige Verteilung der Sickerwasserströmung zeigt, strömt in der<br />

Bodenabdeckung der Böschung ein Teil des Sickerwassers am RCB vorbei und infiltriert konzentriert am<br />

Böschungsfuß. Bei kleinräumiger Mittelung der Durchbruchskonzentrationen ergibt sich eine Verdünnung um<br />

ca. 31 %. Die Durchbruchskurven des Tracers (Abb. 5 (c)) zeigen ein steileres Ansteigen als beim LSW und<br />

fallen innerhalb von fünf Jahren auch deutlich schneller ab. Die maximalen Durchbruchskonzentrationen<br />

C/C0 liegen im kleinräumigen Mittel zwischen 0.20 (Braunerde) und 0.18 (Schwarzerde).<br />

Abb. 8 zeigt die Durchbruchskurven für NAP, PHE und Σ15 EPA-PAK jedoch ohne Abbau. Für das PP-<br />

Szenario zeigt sich in Abb. 8 (a)-(c) deutlich der Effekt der vom Corg des Unterbodens abhängigen Sorption.<br />

Während für die Corg-armen Böden NAP bereits nach kurzer Zeit durchbricht, wird es beim mäßig Corghaltigen<br />

Podsol leicht, bei der Corg-reichen Schwarzerde stärker retardiert (Abb. 8 (a)). Für PHE und<br />

Σ15 EPA-PAK (Abb. 8 (b) und (c)) zeigen auch die Corg-armen Böden eine deutliche Retardation. Der<br />

flachere Verlauf der Durchbruchskurven für Braunerde im oberen Konzentrationsbereich deutet zudem auf<br />

einen Einfluss der langsamen Sorptionskinetik des Skelettanteils hin. Für alle Böden bis auf die Schwarzerde<br />

wird innerhalb der Simulationszeit der Durchbruch der PHE-Peaks mit C/C0 zwischen 0.70 und 0.96 und eine<br />

500


anschließende Konzentrationsabnahme auf Grund der zeitlichen Abnahme der Quellstärke (siehe Abb. 7 (b))<br />

beobachtet. Für die Schwarzerde steigt C/C0 nach 500 a noch deutlich an.<br />

C/C0 [-]<br />

C/C0 [-]<br />

C/C0 [-]<br />

1<br />

0.1<br />

0.01<br />

1<br />

0.1<br />

0.01<br />

1<br />

0.1<br />

0.01<br />

a) Parkplatz<br />

Naphthalin<br />

d) Lärmschutzwall<br />

Naphthalin<br />

g) Straßendamm<br />

Naphthalin<br />

0.1 1 10 100<br />

Zeit [a]<br />

b) Parkplatz<br />

Phenanthren<br />

e) Lärmschutzwall<br />

Phenanthren<br />

h) Straßendamm<br />

Phenanthren<br />

0.1 1 10 100<br />

Zeit [a]<br />

11<br />

c) Parkplatz<br />

Σ15 EPA PAK<br />

Braunerde<br />

Podsol<br />

Schwarzerde<br />

Parabraunerde<br />

Fahlerde<br />

Pelosol<br />

f) Lärmschutzwall<br />

Σ15 EPA PAK<br />

Braunerde (verd.)<br />

Podsol (verd.)<br />

Schwarzerde (verd.)<br />

Braunerde<br />

Podsol<br />

Schwarzerde<br />

i) Straßendamm<br />

Σ15 EPA PAK<br />

Braunerde (verd.)<br />

Podsol (verd.)<br />

Schwarzerde (verd.)<br />

Braunerde<br />

Podsol<br />

Schwarzerde<br />

0.1 1 10 100<br />

Zeit [a]<br />

Abb. 8: Durchbruchskurven der sorbierbaren St<strong>of</strong>fe Naphthalin (schwach sorptiv, links) und Phenanthren<br />

(mäßig gut sorptiv, Mitte) und des Summenparameters Σ15 EPA-PAK (stark sorptiv, rechts) für die drei<br />

Szenarien Parkplatz ((a)-(c)), Lärmschutzwall ((d)-(f)) und Straßendamm ((g)-(i)). Gezeigt werden für<br />

Lärmschutzwall und Straßendamm jeweils Durchbruchskurven für den Schadst<strong>of</strong>ftransportpfad (graue dicke<br />

Kurven) bzw. auf das Gesamtbauwerk bezogene Durchbruchskurven unter Berücksichtigung der<br />

Verdünnung durch nicht belastetes Sickerwasser (schwarze dünne Kurven).<br />

Abb. 8 (d)-(f) zeigt die NAP-, PHE- und Σ15 EPA-PAK- Durchbruchskurven des LSW-Szenarios. Ein<br />

deutlicher Unterschied zum PP ist das längere Anhalten hoher Konzentrationen, welches aus den im<br />

Durchschnitt niedrigeren Strömungsgeschwindigkeiten resultiert. Für NAP werden unter Berücksichtigung<br />

der Verdünnung durch am RCB vorbeiströmendes Sickerwasser Maximaldurchbrüche von C/C0 = 0.56<br />

erreicht (Abb. 8 (d)). Für PHE und Σ15 EPA-PAK zeigen sich bei allen betrachteten Böden nach 500 a noch<br />

deutlich ansteigende Konzentrationen (Abb. 8 (e) und (f)).<br />

Im Vergleich dazu zeigen sich für das SD-Szenario aufgrund der stark erhöhten Infiltrationsraten steilere<br />

Durchbruchskurven und somit ein früheres Erreichen der Konzentrationspeaks (Abb. 8 (g)-(i)). NAP zeigt<br />

seine Maximaldurchbrüche bei der Braunerde bereits nach 19 a (C/C0 = 0.59), bei der Schwarzerde nach 46<br />

a (C/C0 = 0.60).<br />

Ohne Abbau erfolgt langfristig der Eintrag der gesamten Schadst<strong>of</strong>fmasse des RCB ins Grundwasser. Wird<br />

Abbau bei der Simulation mit berücksichtigt (Abb. 9 (a)– (i)), ist die Massenreduktion bei gegebener<br />

Abbauratenkonstante nur von der Verweilzeit der Substanzen im Bodenwasser abhängig. Retardation hat


keinen Einfluss auf die abgebaute Masse, da Mikroorganismen idR. nur zum Abbau in Lösung vorliegender<br />

Schadst<strong>of</strong>fe fähig sind. So zeigen Abb. 9 (a)-(c) für den PP, dass der Abbau bei der Parabraunerde trotz<br />

früherer Durchbruchszeiten effektiver ist, als z.B. bei der Braunerde oder dem Podsol. Dies liegt an der<br />

geringeren Abst<strong>and</strong>sgeschwindigkeit va = q / θ (= Transportgeschwindigkeit des Tracers) bzw. dem höheren<br />

Wassergehalt der feinkörnigen gegenüber den s<strong>and</strong>igen Böden bei vorgegebener Infiltrationsrate (=<br />

Darcyfluss q), woraus sich eine längere effektive Aufenthaltszeit der Substanzen im Bodenwasser des Pr<strong>of</strong>ils<br />

ergibt. Insgesamt ist bei der angenommenen Halbwertzeit von 70 d eine Reduktion der<br />

Durchbruchkonzentrationen für den PP um einen Faktor von bis zu 30 möglich.<br />

C/C0 [-]<br />

C/C0 [-]<br />

C/C0 [-]<br />

1<br />

0.1<br />

0.01<br />

0.001<br />

1<br />

0.1<br />

0.01<br />

0.001<br />

1<br />

0.1<br />

0.01<br />

0.001<br />

a) Parkplatz<br />

Naphthalin<br />

d) Lärmschutzwall<br />

Naphthalin<br />

Braunerde (verd.)<br />

Podsol (verd.)<br />

Schwarzerde (verd.)<br />

Braunerde<br />

Podsol<br />

Schwarzerde<br />

g) Straßendamm<br />

Naphthalin<br />

Braunerde (verd.)<br />

Podsol (verd.)<br />

Schwarzerde (verd.)<br />

Braunerde<br />

Podsol<br />

Schwarzerde<br />

0.1 1 10 100<br />

Zeit [a]<br />

b) Parkplatz<br />

Phenanthren<br />

Braunerde<br />

Podsol<br />

Schwarzerde<br />

Parabraunerde<br />

Fahlerde<br />

Pelosol<br />

e) Lärmschutzwall<br />

Phenanthren<br />

h) Straßendamm<br />

Phenanthren<br />

0.1 1 10 100<br />

Zeit [a]<br />

12<br />

c) Parkplatz<br />

Σ15 EPA PAK<br />

f) Lärmschutzwall<br />

Σ15 EPA PAK<br />

i) Straßendamm<br />

Σ15 EPA PAK<br />

0.1 1 10 100<br />

Zeit [a]<br />

Abb. 9: Durchbruchskurven der sorbierbaren St<strong>of</strong>fe Naphthalin (schwach sorptiv, links) und Phenanthren<br />

(mäßig gut sorptiv, Mitte) und des Summenparameters Σ15 EPA-PAK (stark sorptiv, rechts) unter<br />

Berücksichtigung des biologischen Abbaus für die drei Szenarien Parkplatz ((a)-(c)), Lärmschutzwall ((d)-(f))<br />

und Straßendamm ((g)-(i)). Gezeigt werden für Lärmschutzwall und Straßendamm jeweils<br />

Durchbruchskurven für den Schadst<strong>of</strong>ftransportpfad (graue dicke Kurven) bzw. auf das Gesamtbauwerk<br />

bezogene Durchbruchskurven unter Berücksichtigung der Verdünnung durch nicht belastetes Sickerwasser<br />

(schwarze dünne Kurven).<br />

Aufgrund der langsameren Strömungsgeschwindigkeiten ist beim LSW die durchschnittliche Verweilzeit der<br />

gelösten Schadst<strong>of</strong>fe im Boden höher als beim PP. Dies führt zu einer stärkeren Konzentrationsreduktion bei<br />

Abbau (Minderungsfaktoren zwischen 25 und 150; Abb. 9 (d)-(f)). Umgekehrt zeigen sich beim SD auf Grund<br />

höherer Strömungsgeschwindigkeiten deutlich geringere Aufenthaltszeiten, sodass sich trotz des Abbaus nur<br />

geringe Minderungsfaktoren zwischen 2.5 und 5 ergeben (Abb. 9 (g)–(i)).


Schlussfolgerungen<br />

Die in dieser Studie durchgeführten umfangreichen Typ-Szenarien-Simulationen ergänzen die im BMBF-<br />

Verbundprojekt „Sickerwasserprognose“ erarbeiteten wissenschaftlichen Grundlagen und methodischen<br />

Instrumentarien durch die Anwendung auf praxisrelevante Fallbeispiele. Aus den dabei gewonnenen<br />

Ergebnissen lassen sich folgende Schlussfolgerungen ableiten:<br />

• Dispersive Konzentrationsminderung ist nur bedingt wirksam. Eine nennenswerte Abschwächung der<br />

Durchbruchskonzentrationen ist nur für Fälle möglich, in denen das Abklingen der Quellkonzentrationen<br />

deutlich vor dem Durchbruch des Schadst<strong>of</strong>fpeaks an der GWO erfolgt. Bei länger anhaltenden<br />

Quellstärken (z.B. für Salze wie Sulfat) ist dagegen mit weitgehend unvermindertem<br />

Konzentrationsdurchbruch zu rechnen.<br />

• Für stärker sorbierende Schadst<strong>of</strong>fe (PHE, Σ15 EPA-PAK) zeigt sich schon bei geringem Corg der<br />

Unterböden eine Retardation der Peak-Durchbruchszeitpunkte um viele Jahrzehnte bis Jahrhunderte.<br />

Mit signifikanten St<strong>of</strong>feinträgen (in Abhängigkeit der Quellstärke) ins Grundwasser ist so bei „natürlichen“<br />

Grundwasserneubildungsverhältnissen erst nach sehr langen Zeiträumen zu rechnen.<br />

• Auch bei relativ niedrigen Ratenkonstanten ist eine deutliche Reduktion der Konzentrationen organischer<br />

Schadst<strong>of</strong>fe durch mikrobiellen Abbau möglich. Bezüglich der Effektivität des Abbaus ergeben sich<br />

jedoch erhebliche Unterschiede zwischen den drei betrachteten Verwertungsszenarien, die aus den<br />

unterschiedlichen Strömungs- bzw. Transportgeschwindigkeiten resultieren.<br />

• Die St<strong>of</strong>fverlagerung in den in zwei Raumdimensionen betrachteten Szenarien Lärmschutzwall und<br />

Straßendamm zeigt darüber hinaus einen ausgeprägten Einfluss der räumlich variablen<br />

Sickerwassermenge. Bereiche hoher Strömungsgeschwindigkeit führen zu früheren Ankunftszeiten der<br />

Schadst<strong>of</strong>fe im Grundwasser. Transportprognosen unter Vernachlässigung dieser Effekte können<br />

deshalb zu einer Unterschätzung der Schadst<strong>of</strong>fverlagerung führen und sind nicht konservativ.<br />

• Die Strömungsbilanzen für Lärmschutzwall und Straßendamm legen nahe, dass mit geeigneten<br />

Bauwerksgeometrien und Materialien eine Reduktionen der Infiltration in das Verwertungsmaterial und<br />

des Austrags von Schadst<strong>of</strong>fen ins Grundwasser durch das Ausnutzen von Kapillarsperreneffekten zu<br />

erzielen ist. Der hier vorgestellte Modellansatz schafft grundsätzlich die Möglichkeit, das Design von<br />

Verwertungsszenarien ohne großen Mehraufw<strong>and</strong> in der Umsetzung durch numerische Simulationen in<br />

Richtung einer möglichst geringen Umweltbelastung bei hoher Verwertungsquote optimieren zu können.<br />

Danksagung: Die Untersuchungen wurden im Rahmen des BMBF-Förderschwerpunkts<br />

„Sickerwasserprognose“ (Projektnummer 02WP0517) durchgeführt. Dem BMBF sei für die Förderung<br />

gedankt. Des Weiteren danken wir Herrn Dr. Utermann und Herrn Dr. Duijnisveld (BGR, Hannover), Herrn<br />

Dr. Susset und Herrn Dr. Leuchs (LANUV, Recklinghausen), Herrn Dr. Henzler (UFZ, Leipzig) und Frau Dr.<br />

Kocher (BaST, Bergisch Gladbach) für ausführliche und hilfreiche Diskussionen sowie Frau Dr. Kouznetsova<br />

und Herrn Duran für Hilfe bei der Durchführung der Simulationen.<br />

Literaturangaben<br />

Allen-King, R. M. Grathwohl, P., Ball, W. P.: New modelling paradigms for the sorption <strong>of</strong> hydrophobic organic<br />

chemicals to heterogeneous carbonaceous matter in soils, sediments <strong>and</strong> rocks.- Adv. Wat. Res. 25, 985-<br />

1016 (2002)<br />

Arya, L.M., Paris, J.F.: A physicoempirical model to predict soil moisture characteristics from particle-size<br />

distribution <strong>and</strong> bulk density data. Soil Sci. Soc. Am. J. 45, 1023-1030 (1981)<br />

BBodSchV: Bundes-Bodenschutz- und Altlastenverordnung vom 16. Juli 1999. Bundesgesetzblatt Jahrgang<br />

1999, Teil I Nr.36, 1554-1682 (1999)<br />

BGR (Bundesanstalt für Geowissenschaften und Rohst<strong>of</strong>fe): Nutzungsdifferenzierte Bodenübersichtskarte der<br />

Bundesrepublik Deutschl<strong>and</strong> 1:1.000.000 (BÜK 1000 N2.3). Auszugskarten Acker, Grünl<strong>and</strong>, Wald; Digit.<br />

Archiv FISBo BGR; Hannover und Berlin (2006)<br />

BMU (Bundesministerium für Umwelt, Naturschutz und Reaktorsicherheit): Hydrologischer Atlas von<br />

Deutschl<strong>and</strong>, 3. Lieferung 2003; Berlin (2000)<br />

Bold, S.: Processed-based prediction <strong>of</strong> long-term risk <strong>of</strong> groundwater pollution by organic non-volatile<br />

contaminants. Dissertation, Tübinger Geowissenschaftliche Arbeiten (TGA), C72, 76 S.; Tübingen (2004)<br />

Brakensiek, D.L., Rawls, W.J.: Soil containing rock fragments: Effects on infiltration.- Catena 23 (1-2), 99–110<br />

(1994)<br />

13


Carsell, R. F., Parish, R. S.: Developing joint probability distributions <strong>of</strong> soil water retention characteristics.<br />

Water Resour. Res. 24 (5), 755-769 (1988)<br />

Du, Y., Wang, W. Kolditz, O.: Richards Flow Modelling. 5. Workshop “Porous Media" Proceedings CD, 2.<br />

Ausgabe, Center for <strong>Applied</strong> Geosciences, University <strong>of</strong> Tübingen (2005)<br />

FGSV: RAS-Q 96, Richtlinien für die Anlage von Straßen – Teil: Querschnitte. 62 S.; Köln (1996)<br />

FGSV: ZTV E-StB 94, Zusätzliche Technische Vertragsbedingungen und Richtlinien für Erdarbeiten im<br />

Straßenbau. 108 S.; Köln (1997)<br />

FGSV: Merkblatt für wasserdurchlässige Befestigung von Verkehrsflächen. 36 S.; Köln (1998)<br />

FGSV: RStO 01, Richtlinien für die St<strong>and</strong>ardisierung des Oberbaues von Verkehrsflächen. 52 S.; Köln (2001)<br />

FGSV: ZTV T-StB 95, Zusätzliche Technische Vertragsbedingungen und Richtlinien für Tragschichten im<br />

Straßenbau. Ausgabe 1955 / Fassung 2002. 126 S.; Köln (2002)<br />

FGSV: ZTV SoB-StB 04, Zusätzliche Technische Vertragsbedingungen und Richtlinien für den Bau von<br />

Schichten ohne Bindemittel im Straßenbau. 48 S.; Köln (2004)<br />

FGSV: RAS-EW, Richtlinien für die Anlage von Straßen – Teil: Entwässerung. 82 S.; Köln (2005)<br />

Finkel, M.: Quantitative Beschreibung des Transports von polyzyklischen aromatischen Kohlenwasserst<strong>of</strong>fen<br />

(PAK) und Tensiden in porösen Medien. Tübinger Geowissenschaftliche Arbeiten (TGA), C47. 98 S.;<br />

Tübingen (1998)<br />

Finkel M., Liedl R., Teutsch G.: Modelling surfactant-enhanced remediation <strong>of</strong> polycyclic aromatic<br />

hydrocarbons. J. Environ. Modelling & S<strong>of</strong>tware 14, 203-211 (1998)<br />

Flöter, O.: Wasserhaushalt gepflasterter Straßen und Gehwege. Lysimeterversuche an drei Aufbauten unter<br />

praxisnahen Bedingungen unter Hamburger Klima. Dissertation, Hamburger Bodenkundliche Arbeiten,<br />

58, 329 S.; Hamburg (2006)<br />

Glugla, G., Jankiewicz, P., Rachimow, C., Lojek, K., Richter, K., Fürtig, G., Krahe, P.:<br />

Wasserhaushaltsverfahren zur Berechnung vieljähriger Mittelwerte der tatsächlichen Verdunstung und<br />

des Gesamtabflusses. BfG-Bericht, Nr. 1342, 106 S.; (2003)<br />

Grathwohl, P.: Diffusion controlled desorption <strong>of</strong> organic contaminants in various soils <strong>and</strong> rocks.- In Kharaka,<br />

K.Y., Maest, A.S. (Hrsg.): Water rock Interaction - Proceedings <strong>of</strong> the Utah Conference, 283-286 (1992)<br />

Grathwohl, P.: Diffusion in Natural Porous Media: Contaminant Transport, Sorption/Desorption <strong>and</strong><br />

Dissolution Kinetics. 224 S.; Kluwer Academic Publishers, Boston (1998)<br />

Grathwohl, P.: Gutachten zur Beschreibung von fachlichen Eckpunkten für die Festlegung von<br />

Zuordnungswerten der Einbauklasse 1.1 (Z 1.1) für organische Schadst<strong>of</strong>fe in mineralischen Abfällen.<br />

Umweltbundesamt (Hrsg.) Texte 37/2004. Umweltforschungsplan des Bundesministeriums für Umwelt,<br />

Naturschutz und Reaktorsicherheit, Forschungsbericht 36301047, UBA-FB 000721. 71 S.; Berlin (2004)<br />

Grathwohl, P., Susset, B.: Sickerwasserprognose - Grundlagen, Möglichkeiten, Grenzen, Modelle. In<br />

"Belastung von Böden und Gewässern" Gemeinschaftstagung ATV-DVWK, 28./29.05.01 Hannover<br />

(2001)<br />

Grathwohl, P., Liedl, R., Beyer, C. Konrad, W.: Übertragung der Ergebnisse des BMBF -<br />

Förderschwerpunktes „Sickerwasserprognose“ auf repräsentative Fallbeispiele. Abschlussbericht des<br />

Teilprojekts 1a des Forschungsvorhabens „Entwicklung und Validierung eines Modells zur Abschätzung<br />

der St<strong>of</strong>fkonzentration am Beurteilungsort“ im Rahmen des BMBF-Förderschwerpunkts<br />

„Sickerwasserprognose“ (Förderkennzeichen 02WP0517) (2006)<br />

Grathwohl, P., Halm, D., Bonilla, A., Broholm, M., Burganos, V., Christophersen, M., Comans, R., Gaganis, P.,<br />

Gorostiza, I., Höhener, P., Kjeldsen, P. <strong>and</strong> van der Sloot, H.: Guideline for Groundwater Risk<br />

Assessment at Contaminated Sites.- Final Report EU-Project EVK1-CT-1999-00029; (2003)<br />

Hansson, K., Lundin, L.C., Simunek, J.: Modeling water <strong>flow</strong> patterns in flexible pavements. Transportation<br />

Research Record 1936, 133-141 (2006)<br />

Hayduk, W., Laudie, H.: Prediction <strong>of</strong> diffusion coefficients for nonelectrolytes in dilute aqueous solutions.<br />

AIChE Journal 20 (3), 611-615 (1974)<br />

Hennings, V.: Methodendokumentation Bodenkunde. Auswertungsmethoden zur Beurteilung der<br />

Empfindlichkeit und Belastbarkeit von Böden. In: Geologisches Jahrbuch/ Sonderheft Reihe G, B<strong>and</strong> 1, 2.<br />

Auflage, 232 S.; (2000)<br />

Henzler R.: Quantifizierung und Modellierung der PAK-Elution aus verfestigten und unverfestigten<br />

Abfallmaterialien. Dissertation, Tübinger Geowissenschaftliche Arbeiten (TGA), C76, 118 S.; Tübingen<br />

(2004)<br />

Henzler, R., Rügner, H., Grathwohl, P.: Bewertung der Filter- und Pufferfunktion von Unterböden für<br />

organische Schadst<strong>of</strong>fe. Bodenschutz 1/2006, 8-14 (2006)<br />

Hjelmar, O., Holm J., Crillesen, K.: Utilisation <strong>of</strong> MSWI bottom ash as sub-base in road construction: First<br />

results from a large-scale test site. J. Hazard. Mater. 139 (3), 471-480 (2007)<br />

Höhener, P, Dakhel, N., Christophersen, M., Broholm, M, Kjeldsen, P.: Biodegradation <strong>of</strong> hydrocarbons<br />

vapors: Comparison <strong>of</strong> laboratory studies <strong>and</strong> field investigations in the vadose zone at the emplaced fuel<br />

source experiment, Airbase Værløse, Denmark. J. Contam. Hydrol. 88 (3-4), 337-358 (2006)<br />

14


Jäger, R., Liedl, R.: Prognose der Sorptionskinetik organischer Schadst<strong>of</strong>fe in heterogenem Aquifermaterial.<br />

Grundwasser 2, 57-66 (2000)<br />

Kellermann, C.: Zur Bewertung des Infiltrationsverhaltens von Tragschichten ohne Bindemittel. Dissertation,<br />

Schriftenreihe des Lehrstuhls für Verkehrswegebau der Ruhr-Universität Bochum, Heft 19, 165 S.;<br />

Bochum (2003)<br />

Kleineidam, S., Rügner, H., Grathwohl, P.: Impact <strong>of</strong> grain scale heterogeneity on slow sorption kinetics.<br />

Environ. Tox. Chem. 18 (8), 1673–1678 (1999a)<br />

Kleineidam, S., Rügner, H., Ligouis, B., Grathwohl, P.: Organic matter facies <strong>and</strong> equilibrium sorption <strong>of</strong><br />

phenanthrene. Environ. Sci. Tech. 33 (10), 1637–1644 (1999b)<br />

Kolditz, O., Bauer, S.: A process-oriented approach to computing multifield problems in porous media. J.<br />

Hydroinf. 6 (3), 225-244 (2004)<br />

Kolditz, O., Xie, M., Kalbacher, T., Bauer, S., Wang, W., McDermott, C., Chen, C., Beyer, C., Gronewold, J.,<br />

Kemmler, D., Legeida, D., Walsh, R., Du, Y., Park, C.H.: GeoSys/Rock<strong>flow</strong> version 4.3.21 - Theory <strong>and</strong><br />

users manual, Center for <strong>Applied</strong> Geoscience, University <strong>of</strong> Tübingen (2006)<br />

Krass, K., Brüggemann, M., Görener, E.: Anfall, Aufbereitung und Verwertung von Recycling-Baust<strong>of</strong>fen und<br />

industriellen Nebenprodukten im Wirtschaftsjahr 2001 – Teil 1; Recycling-Baust<strong>of</strong>fe. Straße + Autobahn<br />

4, 193-202 (2004a)<br />

Krass, K., Brüggemann, M., Görener, E.: Anfall, Aufbereitung und Verwertung von Recycling-Baust<strong>of</strong>fen und<br />

indstriellen Nebenprodukten im Wirtschaftsjahr 2001 – Teil 2; Industrielle Nebenprodukte. Straße +<br />

Autobahn 5, 275-281 (2004b)<br />

KrW/AbfG: Gesetz zur Förderung der Kreislaufwirtschaft und Sicherung der umweltverträglichen Beseitigung<br />

von Abfällen (Kreislaufwirtschafts- und Abfallgesetz – KrW/AbfG). BGBl I 1994, 2705 (1994)<br />

Maier, U., Grathwohl, P.: Natural Attenuation in the un<strong>saturated</strong> zone <strong>and</strong> shallow groundwater: coupled<br />

<strong>modeling</strong> <strong>of</strong> vapor phase diffusion, biogeochemical processes <strong>and</strong> transport across the capillary fringe. In:<br />

Nützmann, G., Viotti, P., Aagard, P.: Reactive transport in soil <strong>and</strong> groundwater., 141-155; Springer,<br />

Berlin (2005)<br />

Mesters, K.: Abschätzung der Mobilisierbarkeit von leichtlöslichen Salzen aus Müllverbrennungsasche am<br />

Beispiel eines Lärmschutzwalles.- Dissertation, Schriftenreihe des Lehrstuhls für Verkehrswegebau der<br />

Ruhr-Universität Bochum, Heft 5, 170 S.; Bochum (1993)<br />

Mishra, S., Parker, J. C. , Singhal, N.: Estimation <strong>of</strong> soil hydraulic properties <strong>and</strong> their uncertainty from particle<br />

size distribution data. J. Hydrol. 108, 1-18 (1989)<br />

Mualem, Y.: A new model for predicting the hydraulic conductivity <strong>of</strong> un<strong>saturated</strong> porous media. Water<br />

Resour. Res. 12, 513–522 (1976)<br />

Richter, T.: Regenwasserversickerung durch Pflasterdecken. Zement-Merkblatt Straßenbau S15<br />

http://www.vdz-online.de/fileadmin/gruppen/vdz/3LiteraturRecherche/<br />

Zementmerkblaetter/S15.pdf (2003)<br />

Ross, B.: The diversion capacity <strong>of</strong> capillary barriers. Water Resour. Res. 26 (10), 2625–2629 (1990)<br />

Rügner, H., Henzler, R., Grathwohl, P.: Beurteilung der Empfindlichkeit der Filter- und Pufferfunktion von<br />

Böden (i.B. Unterböden) nach Maßstäben des vorsorgenden Bodenschutzes für organische Schadst<strong>of</strong>fe -<br />

Abschlussbericht. Projekt: „LABO 2003 B 2.03" (Programm: "Wasser und Boden" 2002 der Bund/Länder-<br />

Arbeitsgemeinschaft Bodenschutz) (2005)<br />

Rügner, H., Kleineidam, S., Grathwohl, P.: Sorptions- und Transportverhalten organischer Schadst<strong>of</strong>fe in<br />

heterogenen Materialien am Beispiel des Phenanthrens. Grundwasser 3, 133-138 (1997)<br />

Rügner, H., Kleineidam, S., Grathwohl, P.: Long-term sorption kinetics <strong>of</strong> phenanthrene in aquifer materials.<br />

Environ. Sci. Tech. 33 (10), 1645–1651 (1999)<br />

Seth, R., Mackay, D., Muncke, J.: Estimating <strong>of</strong> organic carbon partition coefficient <strong>and</strong> its variability for<br />

hydrophobic chemicals. Environ. Sci. Technol. 33 (14), 2390-2394 (1999)<br />

Stieber, M., Kraßnitzer, S., dos Santos Coutinho, C., Tiehm, A.: Untersuchung der Bedeutung des<br />

mikrobiellen Abbaus für den Transport persistenter organischer Schadst<strong>of</strong>fe in der ungesättigten Zone.<br />

Abschlussbericht des Teilprojekts 3 des Forschungsvorhabens „Entwicklung und Validierung eines<br />

Modells zur Abschätzung der St<strong>of</strong>fkonzentration am Beurteilungsort“ im Rahmen des BMBF-<br />

Förderschwerpunkts „Sickerwasserprognose“ (Förderkennzeichen 02WP0515) (2006)<br />

Stoppka, B.: Prognose des Salzaustrags aus Straßenbaust<strong>of</strong>fen in der Sickerzone.- Dissertation,<br />

Schriftenreihe des Lehrstuhls für Verkehrswegebau der Ruhr-Universität Bochum, Heft 16, 232 S.;<br />

Bochum (2002)<br />

Susset, B.: Materialien zur Sickerwasserprognose.- In: Förstner, U., Grathwohl, P.: Ingenieurgeochemie,<br />

Technische Geochemie - Konzepte und Praxis. Textbeitrag und Folien im CD-Anhang der 2. neu<br />

bearbeiteten Auflage, 471 S.; Springer, Berlin (2007)<br />

Utermann, J., Kladivko, E.J., Jury, W.A.: Evaluation pesticide migration in tile drained soils with a transfer<br />

function model. J. Environ. Quality 19, 707-714 (1990)<br />

V<strong>and</strong>erborght, J., Vereecken, H.: One-Dimensional Modeling <strong>of</strong> Transport in Soils with Depth-Dependent<br />

Dispersion, Sorption <strong>and</strong> Decay. Vadose Zone J. 6, 140-148 (2007)<br />

15


van Genuchten, M.T.: A closed form equation for predicting the hydraulic conductivity <strong>of</strong> un<strong>saturated</strong> soils.<br />

Soil Sci. Soc. Am. J. 44, 892-898 (1980)<br />

Wörner, T., Löcherer, L., Westiner, E.: Untersuchung über das Verhalten von Recycling-Baust<strong>of</strong>fen in<br />

Tragschichten ohne Bindemittel unter längerer Verkehrsbeanspruchung. Beitrag zum Statusseminar<br />

Verbundvorhaben "Restst<strong>of</strong>fverwertung im Straßenbau", TV 4, FKZ 147 10 43 0, Bochum (2001)<br />

Wösten, J. H. M., Lilly, A., Nemes, A., Le Bas, C.: Using existing soil data to derive hydraulic parameters for<br />

simulation models <strong>and</strong> in l<strong>and</strong> use planning. – Final Report on the European Union Funded project. DLO-<br />

Staring Centre, Report 156, Wageningen /The Netherl<strong>and</strong>s (1998)<br />

16

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!