24.11.2012 Views

RBF_Cover (for eps) - National Water Research Institute

RBF_Cover (for eps) - National Water Research Institute

RBF_Cover (for eps) - National Water Research Institute

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

2<strong>RBF</strong><br />

The <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong> Presents<br />

The Second International<br />

Riverbank Filtration Conference<br />

Riverbank Filtration:<br />

The Future Is NOW!<br />

PROGRAM & ABSTRACTS<br />

Edited by:<br />

GINA MELIN, <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong><br />

September 16-19, 2003<br />

Hilton Cincinnati Netherlands Plaza ✦ Cincinnati, Ohio USA


2<strong>RBF</strong><br />

The <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong> Presents<br />

The Second International<br />

Riverbank Filtration Conference<br />

Riverbank Filtration :<br />

The Future Is NOW!<br />

PROGRAM & ABSTRACTS<br />

Edited by:<br />

GINA MELIN, <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong><br />

September 16-19, 2003<br />

Hilton Cincinnati Netherlands Plaza ✦ Cincinnati, Ohio USA


Published by the<br />

NATIONAL WATER RESEARCH INSTITUTE<br />

NWRI-2003-10<br />

10500 Ellis Avenue ✦ P.O. Box 20865<br />

Fountain Valley, Cali<strong>for</strong>nia 92728-0865<br />

(714) 378-3278 ✦ Fax: (714) 378-3375<br />

www.NWRI-USA.org


Conference Planning Committee<br />

Chair: CHITTARANJAN RAY, Ph.D, P.E., University of Hawaii at Mañoa<br />

Conference Coordinator: GINA MELIN, <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong><br />

EDWARD J. BOUWER, Ph.D.,<br />

The Johns Hopkins University<br />

PAUL ECKERT, Ph.D.,<br />

Stadtwerke Düsseldorf<br />

WILLIAM D. GOLLNITZ,<br />

Greater Cincinnati <strong>Water</strong> Works<br />

THOMAS GRISCHEK, Ph.D.,<br />

University of Applied Sciences<br />

Dresden<br />

DAVID L. HAAS, P.E.,<br />

Jordan, Jones & Goulding<br />

THOMAS HEBERER, Ph.D.,<br />

Technical University of Berlin<br />

STEPHEN HUBBS, P.E.,<br />

Louisville <strong>Water</strong> Company<br />

Conference Sponsors<br />

HENRY C. HUNT, CPG,<br />

Collector Wells International, Inc.<br />

RUDOLF IRMSCHER, Ph.D.,<br />

Stadtwerke Düsseldorf AG<br />

RONALD B. LINSKY,<br />

<strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong><br />

JÜRGEN SCHUBERT,<br />

Stadtwerke Düsseldorf AG<br />

RODNEY A. SHEETS,<br />

United States Geological Survey<br />

THOMAS SPETH, Ph.D., P.E.,<br />

United States<br />

Environmental Protection Agency<br />

✦ Environmental & <strong>Water</strong> Resources <strong>Institute</strong> of the American Society of Civil Engineers<br />

✦ Greater Cincinnati <strong>Water</strong> Works<br />

✦ International Association of <strong>Water</strong>works in the Rhine Catchment Area<br />

✦ Jordan, Jones & Goulding<br />

✦ Louisville <strong>Water</strong> Company<br />

✦ Stadtwerke Düsseldorf AG<br />

✦ United States Environmental Protection Agency<br />

✦ United States Geological Survey<br />

in collaboration with<br />

✦ International Association of Hydrogeologists Commission<br />

on Management of Aquifer Recharge<br />

i


Foreword<br />

In 1999, the <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong> (NWRI) invited American and European<br />

experts to the first International Riverbank Filtration Conference, held in Louisville,<br />

Kentucky, to discuss and promote riverbank filtration, a low-cost, relatively unknown water<br />

treatment technology in the United States. The result of this conference was Riverbank<br />

Filtration: Improving Source-<strong>Water</strong> Quality, a 365-page book written by over 30 experts and<br />

jointly published by NWRI and Kluwer Academic Publishers in 2002.<br />

The editors of Riverbank Filtration — namely, Chittaranjan Ray, Gina Melin, and Ronald<br />

Linsky — saw a need to expand on the topics raised in the book, specifically because the<br />

United States Environmental Protection Agency is developing the proposed Long Term 2<br />

Enhanced Surface <strong>Water</strong> Treatment Rule, which applies to public water-supply systems that<br />

use either surface water or groundwater under the direct influence of surface water as their<br />

raw-water source. As a result, NWRI organized the Second International Riverbank<br />

Filtration Conference, which featured over 40 presenters from around the world and was held<br />

in Cincinnati, Ohio, in September 2003.<br />

This conference came about through the dedication and assistance of numerous individuals<br />

and organizations worldwide. NWRI gratefully acknowledges the ef<strong>for</strong>ts of all those involved<br />

with planning, organizing, and sponsoring the conference, especially the 15-member<br />

Conference Planning Committee. NWRI also extends special thanks to the conference<br />

speakers, moderators, and panelists, whose expertise provided invaluable insight into the<br />

status and needs of riverbank-filtration technology. Finally, NWRI would specifically like to<br />

thank the Greater Cincinnati <strong>Water</strong> Works and Jordan, Jones & Goulding <strong>for</strong> providing<br />

countless hours and manpower towards organizing this conference.<br />

The extended abstracts provided at the conference were the contributions of conference<br />

presenters. Abstracts were edited only when obvious errors were detected or when printing<br />

requirements necessitated action. The opinions expressed within the abstracts are those of<br />

individual authors and do not necessarily reflect those of the sponsors.<br />

NWRI would like to extend sincere thanks to Gina Melin, Editor and Conference<br />

Coordinator, and Tim Hogan, Graphics Designer, <strong>for</strong> their ef<strong>for</strong>ts in bringing the abstracts to<br />

press and ensuring that the quality of each and every abstract reached its fullest potential.<br />

Ronald B. Linsky<br />

Executive Director<br />

<strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong><br />

iii


Conference Schedule & Abstracts Contents<br />

Wednesday, September 17, 2003<br />

All sessions will take place in the Continental Room<br />

7:00 am Registration Mezzanine Level<br />

7:00 am Speaker Breakfast Rosewood Room<br />

8:00 am Welcome Continental Room<br />

Ronald B. Linsky, <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong>, Cali<strong>for</strong>nia<br />

8:15 am Keynote Address<br />

Riverbank Filtration: The American Experience . . . . . . . . . . . . . 1<br />

Edward J. Bouwer, Ph.D., The Johns Hopkins University, Maryland<br />

8:45 am Session 1: Costs<br />

Moderated by William D. Gollnitz, Greater Cincinnati <strong>Water</strong> Works, Ohio<br />

The Costs and Benefits of Riverbank-Filtration Systems . . . . . . . . . . . . . . 3<br />

Stephen A. Hubbs, P.E., Louisville <strong>Water</strong> Company, Kentucky<br />

9:15 am Session 2: Operations<br />

Moderated by Chittaranjan Ray, Ph.D., P.E.,<br />

University of Hawaii at Mañoa, Hawaii<br />

Bridging <strong>Research</strong> and Practical Design Applications . . . . . . . . . . . . . . . . . 7<br />

David L. Haas, P.E., Jordan, Jones & Goulding, Inc., Georgia<br />

Construction and Maintenance of Wells <strong>for</strong> Riverbank Filtration . . . . . . . 17<br />

Henry C. Hunt, CPG, Collector Wells International, Inc., Ohio<br />

Aquifer Storage and Recharge Pretreatment:<br />

Synergies of Bank Filtration, Ozonation, and Ultraviolet Disinfection . . . 23<br />

Robert S. Cushing, Ph.D., Carollo Engineers, Florida<br />

Evolution from a Conventional Well Field<br />

to a Riverbank-Filtration System . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29<br />

John D. North, Cedar Rapids <strong>Water</strong> Department, Iowa<br />

11:15 am Session 3A: Hydraulic Aspects<br />

Moderated by Rudolf Irmscher, Ph.D., Stadtwerke Düsseldorf, Germany<br />

Groundwater Flow and <strong>Water</strong>-Quality – A Flowpath Study<br />

in the Seminole Well Field, Cedar Rapids, Iowa . . . . . . . . . . . . . . . . . . . . . 35<br />

Douglas J. Schnoebelen, Ph.D., United States Geological Survey, Iowa<br />

The Use of Aquifer Testing and Groundwater Modeling<br />

to Evaluate Changes in Aquifer/River Hydraulics at<br />

the Louisville <strong>Water</strong> Company . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39<br />

David C. Schafer, David Schafer & Associates, Minnesota<br />

12:15 pm Lunch Rosewood Room<br />

v


vi<br />

1:30 pm Session 3B: Hydraulic Aspects<br />

Moderated by Rudolf Irmscher, Ph.D., Stadtwerke Düsseldorf, Germany<br />

Plugging in Riverbank-Filtration Systems:<br />

Evaluating Yield-Limiting Factors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43<br />

Stephen A. Hubbs, P.E., Louisville <strong>Water</strong> Company, Kentucky<br />

Application of Different Tracers to Evaluate the Flow Regime<br />

at Riverbank-Filtration Sites in Berlin, Germany . . . . . . . . . . . . . . . . . . . . 49<br />

Dr. Gudrun Massman, Free University of Berlin, Germany<br />

2:30 pm Session 4: Siting<br />

Moderated by Henry C. Hunt, CPG,<br />

Collector Wells International, Inc., Ohio<br />

Siting and Testing Procedures <strong>for</strong> Riverbank-Filtration Systems . . . . . . . . 57<br />

Samuel M. Stowe, P.G., CPG, International <strong>Water</strong> Consultants, Inc., Ohio<br />

<strong>Water</strong> Quality Management <strong>for</strong> Existing Riverbank Filtration Sites<br />

along the Elbe River in Germany. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63<br />

Prof. Dr.-Ing. Thomas Grischek, University of Applied Sciences Dresden, Germany<br />

3:30 pm Session 5: Dynamics<br />

Moderated by Prof. Dr.-Ing. Thomas Grischek,<br />

University of Applied Sciences Dresden, Germany<br />

Using Models to Predict Filtrate Quality at Riverbank-Filtration Sites –<br />

What Is the Adequate Level of Modeling?. . . . . . . . . . . . . . . . . . . . . . . . . . 69<br />

Chittaranjan Ray, Ph.D., P.E., University of Hawaii at Mañoa, Hawaii<br />

The 100-Year Flood of the Elbe River in 2002<br />

and Its Effects on Riverbank-Filtration Sites . . . . . . . . . . . . . . . . . . . . . . . 81<br />

Dipl.-Ing. Matthias Krueger, Fernwasserversorgung Elbaue-Ostharz GmbH, Germany<br />

Temporal Changes of Natural Attenuation Processes<br />

During Bank Filtration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87<br />

Paul Eckert, Ph.D., Stadtwerke Düsseldorf AG, Germany<br />

An Update of the City of Guelph’s Response to Regulation 459/00:<br />

Effective Natural In Situ Filtration of Several Groundwater<br />

Under the Direct Influence of Surface-<strong>Water</strong> Supplies . . . . . . . . . . . . . . . 91<br />

Dennis E. Mutti, P.E., Associated Engineering Limited, Canada<br />

On Bank Filtration and Reactive Transport Modeling . . . . . . . . . . . . . . . . 93<br />

Dr.-Ing. Ekkehard Holzbecher, Humboldt University, Germany<br />

6:00 pm Reception Pavilion Foyer<br />

6:30 pm Dinner Pavilion Ballroom<br />

Dinner Speaker:<br />

Hydraulic Sensitivities and Reduction Potential Correlated<br />

with the Distance Between the Riverbank and Production Well . . . . . . . . 99<br />

Bernhard Wett, Ph.D., University of Innsbruck, Austria


Thursday, September 18, 2003<br />

All sessions will take place in the Continental Room<br />

7:00 am Speaker Breakfast Rosewood Room<br />

8:00 am Keynote Address<br />

Riverbank Filtration: The European Experience . . . . . . . . . . . . . .<br />

Prof. Dr.-Ing. Martin Jekel, Technical University of Berlin, Germany<br />

105<br />

8:30 am Session 6: Microorganisms<br />

Moderated by Edward J. Bouwer, Ph.D., The Johns Hopkins University, Maryland<br />

Using Microscopic Particulate Analysis <strong>for</strong> Riverbank Filtration . . . . . . . 111<br />

Jennifer L. Clancy, Ph.D., Clancy Environmental Consultants, Inc., Vermont<br />

Transport and Removal of Cryptosporidium Oocysts<br />

in Subsurface Porous Media. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 115<br />

Menachem Elimelech, Ph.D., Yale University, Connecticut<br />

Laboratory and Field Strategies <strong>for</strong> Assessing Pathogen Removal<br />

by Riverbank Filtration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 117<br />

Monica B. Emelko, Ph.D., University of <strong>Water</strong>loo, Canada<br />

Fate of Disinfection Byproduct Precursors and Microorganisms<br />

During Riverbank Filtration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 123<br />

W. Joshua Weiss, The Johns Hopkins University, Maryland<br />

Assessment of the Microbial Removal Capabilities<br />

of Riverbank Filtration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129<br />

Vasiliki Partinoudi, University of New Hampshire, New Hampshire<br />

11:00 am Session 7: Organics Removal<br />

Moderated by Richard J. Miltner, P.E.,<br />

United States Environmental Protection Agency, Ohio<br />

Riverbank Filtration: A Very Efficient Treatment Process<br />

<strong>for</strong> the Removal of Organic Contaminants?. . . . . . . . . . . . . . . . . . . . . . . . . 137<br />

Dr.-Ing. Heinz-Jürgen Brauch, DVGW-Technologiezentrum Wasser, Germany<br />

Organics Removal by Riverbank Filtration<br />

at the Greater Cincinnati <strong>Water</strong> Works Site . . . . . . . . . . . . . . . . . . . . . . . . 143<br />

Jeffrey Vogt, Greater Cincinnati <strong>Water</strong> Works, Ohio<br />

12:00 pm Lunch Pavilion Ballroom<br />

Lunch Speaker:<br />

Potential Uses of Riverbank Filtration<br />

<strong>for</strong> Regulatory Compliance. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .<br />

Stig Regli, United States Environmental Protection Agency, Washington, D.C.<br />

149<br />

vii


viii<br />

1:15 pm Session 8: Emerging Contaminants Removal<br />

Moderated by Monica B. Emelko, Ph.D., University of <strong>Water</strong>loo, Canada<br />

Transport and Attenuation of Pharmaceutical Residues<br />

During Bank Filtration . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151<br />

Andy Mechlinski, Technical University of Berlin, Germany<br />

Attenuation of Pharmaceuticals During Riverbank Filtration . . . . . . . . . . 155<br />

Traugott Scheytt, Ph.D., Technical University of Berlin, Germany<br />

The Fate of Bulk Organics and Emerging Contaminants<br />

During Soil-Aquifer Treatment . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 159<br />

Dr. Jörg E. Drewes, Colorado School of Mines, Colorado<br />

Ethylenediaminetetraacetic Acid Occurrence and Removal<br />

Through Bank Filtration in the Platte River, Nebraska . . . . . . . . . . . . . . . 163<br />

Jason R. Vogel, Ph.D., United States Geological Survey, Nebraska<br />

3:15 pm Session 9: Public Policy and Regulatory<br />

Moderated by Richard J. Miltner, P.E.,<br />

United States Environmental Protection Agency, Ohio<br />

Riverbank Filtration as a Regional Supply Option<br />

<strong>for</strong> the United States . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 167<br />

Leo Gentile, P.G., CPG, Jordan, Jones & Goulding, Inc., Georgia<br />

Application of the Long Term 2 Enhanced Surface <strong>Water</strong><br />

Treatment Rule Microbial Toolbox at Existing <strong>Water</strong> Plants . . . . . . . . . . . 173<br />

Richard A. Brown, Environmental Engineering and Technology, Inc., Virginia<br />

Draft Protocol <strong>for</strong> the Demonstration of Effective Riverbank Filtration . . . . 175<br />

William D. Gollnitz, Greater Cincinnati <strong>Water</strong> Works, Ohio<br />

Source <strong>Water</strong> Protection and Riverbank Filtration<br />

in the Dyje River Basin . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 181<br />

Prof.-Dr. Petr Hlavinek, Brno University of Technology, Czech Republic<br />

5:15 pm Session 10: <strong>Research</strong> Needs Panel Discussion<br />

Moderated by Ronald B. Linsky, <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong>, Cali<strong>for</strong>nia<br />

Panelists:<br />

Kellogg J. Schwab, Ph.D.,<br />

Johns Hopkins Bloomberg School of Public Health, Maryland<br />

Peter Fox, Ph.D., Arizona State University, Arizona<br />

Monica B. Emelko, Ph.D., University of <strong>Water</strong>loo, Canada<br />

Prof. Dr.-Ing. Thomas Grischek,<br />

University of Applied Sciences Dresden, Germany<br />

Prof. Vladimir Rojanschi, Ecological University Bucharest, Romania<br />

6:30 pm Reception Rosewood Room


Friday, September 19, 2003<br />

7:00 am Speaker Breakfast Rosewood Room<br />

8:00 am Session 11: Case Studies “Lessons Learned”<br />

Moderated by Stephen Hubbs, P.E., Louisville <strong>Water</strong> Company, Kentucky<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Flowpath Study Field Design: Methodology and Evaluation. . . . . . . . . . . . 187<br />

Bruce Whitteberry, P.G., Greater Cincinnati <strong>Water</strong> Works, Ohio<br />

The Hungarian Experience with Riverbank Filtration . . . . . . . . . . . . . . . . 193<br />

Ferenc Laszlo, Ph.D.,<br />

<strong>Institute</strong> <strong>for</strong> <strong>Water</strong> Pollution Control, <strong>Water</strong> Resources <strong>Research</strong> Centre, Hungary<br />

Nitrate Pollution of a <strong>Water</strong> Resource –<br />

15 18<br />

N and 0 Study of Infiltrated Surface <strong>Water</strong> . . . . . . . . . . . . . . . . . . . . . 197<br />

Frantisek Buzek, Ph.D., Czech Geological Survey, Czech Republic<br />

Microbial Growth in Artificially Recharged Groundwater:<br />

Experiences from a 4-Year Project . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 203<br />

Ilkka T. Miettinen, Ph.D., <strong>National</strong> Public Health <strong>Institute</strong>, Finland<br />

Evaluation of the Existing Per<strong>for</strong>mance of Infiltration Galleries<br />

in the Alluvial Deposits of the Parapeti River . . . . . . . . . . . . . . . . . . . . . . 207<br />

Dip.-Eng. Alvaro Camacho Garnica,<br />

Bolivian Association of Sanitary Engineers, Bolivia<br />

Sensitivity and Implication of Microscopic Particulate Analysis –<br />

A Collector Well Owner’s Perspective. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 215<br />

Barry C. Beyeler, City of Boardman, Oregon<br />

Combined Use of Surface <strong>Water</strong> and Groundwater<br />

<strong>for</strong> Drinking-<strong>Water</strong> Production in the Barcelona Metropolitan Area. . . . . 217<br />

Jordi Martín-Alonso, Barcelona’s <strong>Water</strong> Company, Spain<br />

11:30 am Conference Wrap-Up<br />

Edward J. Bouwer, Ph.D., Johns Hopkins University, Maryland<br />

Martin Jekel, Ph.D., Technical University of Berlin, Germany<br />

ix


Acronyms<br />

ADA ß-alaninediacetic acid<br />

AOC Assimilable organic carbon<br />

AOX Adsorbable organic halogen<br />

ASR Aquifer Storage and Recovery<br />

DBP Disinfection byproduct<br />

DOC Dissolved organic carbon<br />

DTPA Diethylenetrinitrilopenataacetic acid<br />

EDTA Ethylenediaminetetraacetic acid<br />

GAC Granular activated carbon<br />

GC Gas chromatography<br />

GWUDI Groundwater under the direct influence of surface water<br />

HAA Haloacetic acid<br />

HPI Hydrophilic carbon<br />

HPO-A Hydrophobic acids<br />

LT2ESWTR Long Term 2 Enhanced Surface <strong>Water</strong> Treatment Rule<br />

MAP Microbially available phosphorous<br />

MS Mass spectrometry<br />

NASRI Natural and Artificial Systems <strong>for</strong> Recharge and Infiltration<br />

NTA Nitrilotriacetic acid<br />

NOM Natural organic matter<br />

NPEC Nonylphenolpolyethoxycarboxylate<br />

O&M Operation and maintenance<br />

PDTA 1.3-propylenedinitrilotetraacetic acid<br />

PhAC Pharmaceutically active compound<br />

PHREEQC pH redox equilibrium equation<br />

<strong>RBF</strong> Riverbank filtration<br />

THM Trihalomethane<br />

TOC Total organic carbon<br />

USEPA United States Environmental Protection Agency<br />

USGS United States Geological Survey<br />

UV Ultraviolet<br />

xi


xii<br />

Units of Measure<br />

cfu Colony-<strong>for</strong>ming unit.<br />

ft Foot<br />

m Meter<br />

m 3 /d Cubic meters per day<br />

m 3 /s Cubic meters per second<br />

MGD Million gallons per day<br />

mg/L Milligrams per liter<br />

ntu Nephelometric turbidity unit<br />

pCi/L Picocurie per liter<br />

pfu Plaque-<strong>for</strong>ming unit.<br />

µg/L Micrograms per liter


Conference<br />

Abstracts


Keynote Presentation<br />

Riverbank Filtration: The American Experience<br />

Edward J. Bouwer, Ph.D.<br />

The Johns Hopkins University<br />

Baltimore, Maryland<br />

Riverbank filtration (<strong>RBF</strong>) is a process during which surface water is subjected to subsurface flow<br />

prior to extraction from vertical or horizontal wells. Most <strong>RBF</strong> systems are located along<br />

riverbanks. Alternative systems can involve lakes and infiltration ponds. The objective of this<br />

keynote address is to provide an overview of the water-quality improvements possible with <strong>RBF</strong>,<br />

the motivation <strong>for</strong> <strong>RBF</strong> and its promise in the United States, and some of the remaining<br />

challenges <strong>for</strong> the reliable implementation of this technology.<br />

During infiltration and soil passage, surface water is subjected to a combination of physical,<br />

chemical, and biological processes, such as filtration, dilution, sorption, and biodegradation, that<br />

can significantly improve raw-water quality. Transport through alluvial aquifers is associated with<br />

a number of water-quality benefits, including the removal of microbes, pesticides, total organic<br />

carbon (TOC), dissolved organic carbon (DOC), nitrate, and other contaminants. In comparison<br />

to most groundwater sources, alluvial aquifers that are hydraulically connected to rivers are<br />

typically easier to exploit (shallow) and more highly productive <strong>for</strong> drinking-water supplies. As<br />

reflected by several recently published reviews (Journal of Hydrology, 2002; Ray et al., 2002a;<br />

Tufenkji et al., 2002; and Ray et al., 2002b), <strong>RBF</strong> is receiving increased attention, especially in the<br />

United States. One motivation <strong>for</strong> the increased applications of <strong>RBF</strong> is the need <strong>for</strong> drinking-water<br />

utilities to meet increasingly stringent drinking-water regulations, especially with regard to:<br />

• The provision of multiple barriers <strong>for</strong> protection against microbial pathogens.<br />

• Tighter regulations <strong>for</strong> disinfection byproducts (DBPs), such as trihalomethanes (THMs)<br />

and haloacetic acids (HAAs).<br />

A second motivation is the ability of <strong>RBF</strong> to provide continuous treatment and to buffer against<br />

accidental spills and terrorist events.<br />

Since <strong>RBF</strong> is a natural treatment system and has a long history of use in Europe, it is sometimes<br />

viewed as a simple process. This simplicity is appealing; however, the geochemical, biological, and<br />

hydrologic factors that control the removal of dissolved and particulate contaminants during <strong>RBF</strong><br />

are complex. We have learned from experience with groundwater remediation that the effectiveness<br />

of subsurface processes is inherently site-specific. Many of the papers presented at this conference<br />

will address the factors influencing water-quality improvements possible with <strong>RBF</strong> systems.<br />

One question of interest to users can be stated as, “Is <strong>RBF</strong> with post-disinfection an acceptable, or<br />

even preferable, alternative to conventional drinking-water treatment?” In my laboratory, the<br />

reduction of DBP precursors upon <strong>RBF</strong> was compared with that obtained using a bench-scale<br />

Correspondence should be addressed to:<br />

Edward J. Bouwer, Ph.D.<br />

Professor, Department of Geography and Environmental Engineering<br />

The Johns Hopkins University<br />

3400 N. Charles Street • Baltimore, Maryland 21218 USA<br />

Phone: (410) 516-7437 • Fax: (410) 516-8996 • Email: Bouwer@jhu.edu<br />

1


2<br />

conventional treatment train on corresponding river waters. The river waters were subjected to a<br />

treatment train consisting of coagulation, flocculation, sedimentation, filtration, and ozonation.<br />

This research showed that <strong>RBF</strong> per<strong>for</strong>ms as well as or better than a bench-scale conventional<br />

treatment train (based on coagulation chemistry <strong>for</strong> optimum turbidity removal) with respect to<br />

the removal of natural organic matter (NOM), particularly precursor material <strong>for</strong> THM4 and<br />

HAA6 concentrations. Consequently, a potential major benefit of <strong>RBF</strong> is as a pretreatment step<br />

<strong>for</strong> controlling DBPs. A shift from chlorinated to brominated DBPs occurred during <strong>RBF</strong>. Since<br />

brominated DBPs tend to have greater toxicity than chlorinated DBPs, the shift from chlorinated<br />

to brominated DBP species caused reductions in the calculated risk <strong>for</strong> bank-filtered waters in this<br />

study to be lower than corresponding reductions in THM concentrations. Nonetheless, the data<br />

demonstrate the ability of <strong>RBF</strong> to reduce theoretical risk due to THMs <strong>for</strong>med upon chlorination<br />

in all cases and with substantially better per<strong>for</strong>mance than the bench-scale conventional<br />

treatment train.<br />

A remaining challenge <strong>for</strong> the reliable implementation of <strong>RBF</strong> as a technology is the dynamic<br />

nature of these complex hydrologic systems. Transient flows in rivers influence flow and removal<br />

processes during ground passage, which can affect the quality of bank-filtered waters. A more<br />

detailed understanding of the fundamental processes that govern contaminant transport in <strong>RBF</strong><br />

systems will lead to the reliable design, implementation, and operation of <strong>RBF</strong> systems.<br />

REFERENCES<br />

Journal of Hydrology (2002). Special Issue on Bank Filtration, 266(3-4): 139-284.<br />

Ray, C., T. Grischek, J. Schubert, J.Z. Wang, and T.F. Speth (2002a). “A perspective of riverbank filtration.”<br />

Jour. AWWA, 94(4): 149-160.<br />

Ray, C., G. Melin, and R.B. Linsky, editors (2002b). Riverbank Filtration: Improving Source <strong>Water</strong> Quality,<br />

Kluwer Academic Publishers, Dordrecht.<br />

Tufenkji, N., J.N. Ryan, and M. Elimelech (2002). “The promise of bank filtration.” Environmental Science<br />

and Technology, 36(21): 423A-428A.<br />

ED BOUWER has taught environmental engineering courses at The Johns Hopkins<br />

University since 1985. His research interests include factors that influence the biotrans<strong>for</strong>mation<br />

of organic contaminants, bioremediation <strong>for</strong> the control of organic contaminants<br />

at waste sites, biofilm kinetics, the interaction between biotic and abiotic processes,<br />

groundwater contamination, biological processes design in wastewater, industrial and<br />

drinking-water treatment, and the transport and fate of microorganisms in porous media.<br />

At present, he is an editor <strong>for</strong> or is on the editorial board of several journals, including the<br />

Journal of Contaminant Hydrology, Biodegradation, and Environmental Engineering Sciences. Bouwer received a<br />

B.S. in Civil Engineering from Arizona State University, and both an M.S. and Ph.D. in Environmental<br />

Engineering and Science from Stan<strong>for</strong>d University.


Session 1: Costs<br />

The Costs and Benefits of Riverbank-Filtration Systems<br />

Stephen A. Hubbs, P.E.<br />

Louisville <strong>Water</strong> Company<br />

Louisville, Kentucky<br />

Henry C. Hunt, CPG<br />

Collector Wells International, Inc.<br />

Columbus, Ohio<br />

Jürgen Schubert<br />

Stadtwerke Düsseldorf<br />

Düsseldorf, Germany<br />

The Benefits of Riverbank Filtration<br />

The history of <strong>RBF</strong> in modern times is connected to the experience of disease outbreaks in Europe<br />

in the 1890s, with specific reference to the cholera epidemic in Hamburg, Germany, in 1892. The<br />

preference <strong>for</strong> groundwater, <strong>RBF</strong>, and artificial recharge to surface water stems from this<br />

experience, noting much more wholesome water from these sources than from surface water.<br />

A much earlier reference to <strong>RBF</strong> (albeit far less scientific) is found in the Bible:<br />

The fish that were in the Nile died, and the Nile became foul, so that the Egyptians<br />

could not drink water from the Nile … So all the Egyptians dug around the Nile <strong>for</strong><br />

water to drink, <strong>for</strong> they could not drink of the water of the Nile. (Exodus 7, 21-24)<br />

Thus, it appears that the benefits of <strong>RBF</strong> are anything but new!<br />

The benefits of <strong>RBF</strong> have been cataloged in recent publications to include:<br />

• Particle removal.<br />

• Pathogen removal.<br />

• Organic and inorganic chemical removal.<br />

• Peak smoothing in spills.<br />

• Reduction in DBP <strong>for</strong>mation.<br />

• Production of a more biologically stable water.<br />

When faced with the decision to choose an advanced treatment technology <strong>for</strong> its two treatment<br />

plants, the Louisville <strong>Water</strong> Company in Louisville, Kentucky, sought to compare these benefits<br />

against the benefits of in-plant treatment techniques from both a treatment efficiency perspective<br />

and cost-based perspective; however, a comprehensive, quantitative measure was difficult to develop.<br />

Correspondence should be addressed to:<br />

Stephen A. Hubbs, P.E.<br />

Vice President, New Technology<br />

Louisville <strong>Water</strong> Company<br />

550 South Third Street • Louisville, Kentucky 40202 USA<br />

Phone: (502) 569-3675 • Fax: (502) 569-0813 • Email: SHubbs@lwcky.com<br />

3


4<br />

The Louisville <strong>Water</strong> Company also wanted to involve the public in the overall decision process,<br />

so a simple communication tool was desired that was capable of ranking various treatment<br />

technologies with regards to the risks they were designed to reduce. Treatment effectiveness was<br />

simply identified by a series of “+” signs, with one “+” being effective and two “++” being highly<br />

effective. A negative assignment (“–”) indicated that the treatment process had an overall<br />

negative impact on the risk of a selected component, and a “0” indicated no impact.<br />

The risks evaluated in the analysis included: pathogens, such as Giardia and Cryptosporidium;<br />

DBPs; tastes and odor (2-methylisoborneol and Geosmin); synthetic organic chemicals like<br />

atrazine and Aloclor; and river-borne spills of industrial chemicals. Operational benefits included:<br />

reduced regrowth in the distribution system, reduced main breaks from avoidance of extremely<br />

cold water, avoidance of zebra mussels, and reliability associated with simple operation. These risks<br />

and benefits were used to compare the <strong>RBF</strong> process to other treatment technologies, including<br />

conventional treatment, ozone, ultraviolet light (UV), granular activated carbon (GAC), biological<br />

active carbon, combinations of all of these, and membranes.<br />

This matrix was recently modified to include risks associated with the presence of radon and<br />

arsenic in groundwater. Radon has been detected in <strong>RBF</strong> water in Louisville at levels near the<br />

maximum contaminant level of 300 picocuries per liter (pCi/L). The radon level leaving the<br />

treatment plant is below the detection limit of 50 pCi/L, with the reduction the result of<br />

out-gassing in open treatment basins and blending with surface water. Arsenic has not been<br />

detected in <strong>RBF</strong> water at Louisville. Radon and arsenic risk factors do not exist in Ohio River<br />

source water. The treatments <strong>for</strong> radon (aeration) and arsenic (flocculation with ferric) effectively<br />

reduce these risk factors at an increase in treatment costs.<br />

A matrix comparing these risks and operational benefits is presented in Table 1. This analysis<br />

indicates which treatment combinations provide specific water-quality benefits. From this analysis,<br />

it was obvious that <strong>RBF</strong> was capable of matching the benefits of in-plant treatment options.<br />

Table 1. Comparison of Risks and Operational Benefits<br />

River + Ozone River + GAC River + GAC<br />

Benefits <strong>RBF</strong> + UV and UV + UV + Membranes<br />

Particle Removal + + + +<br />

Microbial Removal + + + +<br />

DBP Reduction + + + +<br />

Taste and Odor + + + +<br />

Spill Dampening ++ 0 ++ +<br />

Iron/Manganese – + + 0<br />

SOC Removal + + + +<br />

AOC/BDOC Control + 0 + +<br />

Nitrification Control + + + +<br />

Temperature ++ 0 0 0<br />

Operability + 0 0 +<br />

Residuals 0 0 0 +<br />

Multiple Barriers ++ ++ ++ ++<br />

Reduced Chemicals + 0 0 +<br />

Radon – 0 0 0<br />

Arsenic 0 0 0 0<br />

TOTAL 13 9 12 13<br />

AOC = Assimilable organic carbon. BDOC = Biodegradable organic carbon. SOC = Synthetic organic chemical.


Cost-Evaluation of Riverbank Filtration<br />

The capital cost of a <strong>RBF</strong> system depends on many factors, including aquifer characteristics, type<br />

of well-screen installation (vertical or horizontal), aesthetic considerations in facility design, and<br />

distance to the population served. The operational costs vary as a function of water quality and<br />

required treatment, lift required in pumping, and pump and well-screen maintenance costs. The<br />

ability of a stream to support a given well-field capacity can be calculated as the yield per unit<br />

length of riverbank. This capacity is influenced by the composition of the riverbed, riverbed<br />

scouring characteristics, stream-water quality, and width of the river.<br />

Large river systems situated in glacial sands and gravels (such as the Ohio, Mississippi, and Rhine<br />

rivers) can sustain yields up to 8 million gallons per day (MGD) per 1,000 feet (ft) of riverbank.<br />

Typical installations in these aquifers include 1.5-MGD vertical wells spaced on 200-ft centers, or<br />

15-MGD horizontal collector wells spaced at 2,000-ft centers.<br />

The 20-MGD horizontal collector well system constructed in Louisville in 1999 cost $5 million.<br />

The system included:<br />

• Seven laterals that are 200 to 240 ft in length.<br />

• A 21-ft diameter, 100-ft deep caisson.<br />

• A pump house and controls with one constant speed and one variable speed pump<br />

(both 10 MGD).<br />

• Two thousand feet of 42-inch discharge piping to the plant.<br />

The pump house was designed to include architectural features complimentary to surrounding<br />

residential neighborhoods. This system can peak at over 20 MGD in warm weather and is operated<br />

year-round at 17 MGD.<br />

When this system was being considered, alternative treatment techniques <strong>for</strong> surface-water treatment<br />

were also evaluated. Critical treatment functions included efficiency in removing Giardia and<br />

Cryptosporidium, ability to remove 2-methylisoborneol and Geosmin, and DBP reduction. Suitable<br />

alternatives were determined from the matrix and evaluated <strong>for</strong> cost. This process was repeated<br />

twice, by two separate consultants: once in the initial planning stages of the 20-MGD project<br />

(1995), and again just be<strong>for</strong>e a contract was let to design a 45-MGD expansion of the system<br />

(2002).<br />

The surface-water treatment process selected in the final cost analysis as providing comparable<br />

benefits to the overall benefits of <strong>RBF</strong> included: conventional treatment, ozone and biological<br />

treatment in GAC-capped filters, and UV/chlorine/chloramine disinfection. Treatment costs,<br />

however, were also estimated <strong>for</strong> separate elements of this treatment scheme (as was the cost of<br />

membranes) <strong>for</strong> the 180-MGD treatment plant at Crescent Hill.<br />

The results of this cost comparison are included in Table 2. The capital costs were amortized over<br />

20 years in this analysis, and the operation and maintenance (O&M) costs include the treatment<br />

costs to reduce hardness from 220 to 160 milligrams per liter (mg/L). Based on this analysis, the<br />

decision was made to proceed with the development of <strong>RBF</strong> <strong>for</strong> the entire capacity of the 60-MGD<br />

B.E. Payne <strong>Water</strong> Treatment Plant, and to continue considering <strong>RBF</strong> <strong>for</strong> the larger 180-MGD<br />

capacity Crescent Hill Treatment Plant.<br />

5


6<br />

Treatment Capital Cost Annual O&M Cost Present Worth Cost<br />

Alternative ($ million) ($ million) ($ million)<br />

<strong>RBF</strong> Conventional 103 1.82 112<br />

<strong>RBF</strong> + UV 116 2.34 130<br />

River + Ozone UV 70 6.36 152<br />

River + GAC + UV 114 5.46 160<br />

River + Membranes<br />

+ GAC + UV<br />

Table 2. Results of the Cost Comparison<br />

204 4.96 251<br />

Capital Costs of Alternative Well-Field Design: Hard-Rock Tunnel Collector<br />

The difficulty of obtaining riverfront property in developed areas prompted the Louisville <strong>Water</strong><br />

Company to consider alternative designs to the traditional well-field configuration. The design<br />

selected <strong>for</strong> constructing the 45-MGD addition to the existing 20-MGD <strong>RBF</strong> capacity includes<br />

the construction of a large-diameter (10- to 12-ft) horizontal tunnel in bedrock, below the waterbearing<br />

sand and gravel aquifer. Vertical wells will be constructed at 200-ft centers, and will<br />

penetrate the bedrock and discharge by gravity into the tunnel. A single pump station located at<br />

the treatment plant will extract up to 45 MGD from the tunnel, connecting the 30 vertical wells<br />

in the system. This construction technique minimizes the impact on landowners, with visible<br />

construction activities limited to drilling and developing the individual wells. This design also<br />

allows total flow to be distributed to each well, evenly distributing riverbed plugging stresses across<br />

6,000 ft of riverbank.<br />

The cost of this system was compared to the cost of developing a conventional well field with<br />

submersible pumps and a collecting header. The cost-estimate <strong>for</strong> the bedrock tunnel system was highly<br />

impacted by the assigned cost of the hard-rock tunnel. The final design included a cement-lined tunnel<br />

to protect against intermittent layers of shale encountered in the massive limestone <strong>for</strong>mation.<br />

The current estimate <strong>for</strong> this hard-rock tunnel-vertical well extraction system is $33 million <strong>for</strong><br />

45-MGD capacity. The system involves approximately 6,000 ft of riverbank and hard-rock tunnel.<br />

The comparable conventional vertical well field has been estimated at $23 million. Current plans<br />

are to design and construct the hard-rock vertical well system, noting advantages in expandability,<br />

future connection to an additional 180-MGD system <strong>for</strong> the larger treatment plant, and general<br />

constructability in the developed riverfront area.<br />

STEVE HUBBS is a Professional Engineer with 28 years of experience at the Louisville<br />

<strong>Water</strong> Company in Louisville, Kentucky. In the early 1980s, he began researching riverbank<br />

filtration as an alternate source of water <strong>for</strong> the Louisville <strong>Water</strong> Company, specifically<br />

looking at the reduction in disinfection byproduct precursors, river-borne organics, and<br />

mutagenicity in the riverbank-filtration process. His work continued in the 1990s with a<br />

focus on pathogen reduction, and his research is now being conducted on the hydraulic<br />

connection between the riverbed and aquifer, with a focus on riverbed plugging dynamics<br />

and their influence on sustainable yields from high-capacity riverbank-filtration systems. Hubbs received an<br />

M.S. in Environmental Engineering from the University of Louisville, and is currently enrolled in the Ph.D.<br />

program in Civil and Environmental Engineering at the University of Louisville, focusing on the hydraulics<br />

of riverbank-filtration systems.


Session 2: Operations<br />

Bridging <strong>Research</strong> and Practical Design Applications<br />

David L. Haas, P.E.<br />

Jordan, Jones & Goulding, Inc.<br />

Norcross, Georgia<br />

Michael J. Robison, P.E.<br />

Jordan, Jones & Goulding, Inc.<br />

Norcross, Georgia<br />

David R. Wilkes, P.E.<br />

Jordan, Jones & Goulding, Inc.<br />

Norcross, Georgia<br />

Background<br />

The Louisville <strong>Water</strong> Company operates two water filtration plants: the B.E. Payne <strong>Water</strong> Treatment<br />

Plant (Payne Plant), which has a treatment capacity of 60 MGD (227,000 cubic meters per day [m 3 /d])<br />

and is located in eastern Jefferson County, Kentucky; and the Crescent Hill <strong>Water</strong> Treatment<br />

Plant, which is located closer to downtown Louisville and has a treatment capacity of 180 MGD<br />

(682,000 m 3 /d). Both are conventional surface-water treatment plants drawing their supply from<br />

the Ohio River. The Payne Plant has two 60-inch (1.5-meter [m]) raw-water intake pipes located<br />

on plant property. The intake <strong>for</strong> the Crescent Hill <strong>Water</strong> Treatment Plant is located near the<br />

intersection of Zorn Avenue and River Road, approximately 7 miles (4.4 kilometers) downstream<br />

from the Payne Plant intake.<br />

In 1999, the Louisville <strong>Water</strong> Company started operating a horizontal collector well at its<br />

Payne Plant. This horizontal collector well consists of a 16-ft (4.9-m) diameter caisson.<br />

The caisson was constructed down to the top of bedrock and is approximately 100-ft (30.5-m)<br />

deep. There are seven 12-inch (30.5-centimeter) collector laterals radiating out horizontally from<br />

the caisson. A pump station was constructed on top of the caisson above the flood plain to pump<br />

water to the Payne Plant <strong>for</strong> further treatment. The design capacity of this well is 15 MGD<br />

(57,000 m 3 /d).<br />

Jordan, Jones & Goulding, Inc. was retained by the the Louisville <strong>Water</strong> Company to implement<br />

<strong>RBF</strong> <strong>for</strong> the remainder of the Payne Plant. This second phase of <strong>RBF</strong> will provide an additional<br />

45 MGD (171,000 m 3 /d) of capacity. Ultimately, the Louisville <strong>Water</strong> Company desires to extend<br />

<strong>RBF</strong> technology to the Crescent Hill <strong>Water</strong> Treatment Plant.<br />

This paper provides a summary of issues that were faced in taking the concept of providing<br />

additional <strong>RBF</strong> capacity <strong>for</strong> the Payne Plant through design. There were three primary areas that<br />

Correspondence should be addressed to:<br />

David L. Haas, P.E.<br />

Senior Project Manager<br />

Jordan, Jones & Goulding, Inc.<br />

6801 Governors Lake Parkway • Norcross, Georgia 30071 USA<br />

Phone: (678) 333-0242 • Fax: (678) 333-0828 • Email: DHAAS@JJG.com<br />

7


8<br />

design engineers needed to address:<br />

• Type of <strong>RBF</strong> collection system.<br />

• Hydrogeologic yield.<br />

• <strong>Water</strong> quality and treatment.<br />

Type of <strong>RBF</strong> Collection System<br />

First, it was necessary to select the most appropriate type of <strong>RBF</strong> technology to satisfy the technical<br />

needs of the project economically, as well as other goals such as community acceptance of the<br />

project. The construction of additional horizontal collector wells, similar to the existing well, was<br />

considered unacceptable because 15 to 20 of these collector wells would be required to satisfy the<br />

total treatment capacity <strong>for</strong> both plants. The construction of this many wells, each with an aboveground<br />

pump station, would be costly, as well as aesthetically unpleasing.<br />

Three options were considered that would use tunnels in conjunction with <strong>RBF</strong> to satisfy the<br />

needs of the project:<br />

• Option 1 consisted of driving a tunnel through the alluvial deposits that overlay bedrock<br />

and installing collector laterals from within the tunnel (soft-ground tunnel option). This<br />

design concept is depicted in Figure 1.<br />

• Option 2 included installing a series of horizontal collector wells (Figure 2) that would<br />

be capped at grade and connected together using a deep tunnel through the bedrock.<br />

• Option 3 consisted of installing vertical collectors connected together using a tunnel<br />

through the bedrock (Figure 3).<br />

With any of these options, the number of above-ground pump stations would be minimized,<br />

because water from horizontal or vertical collectors would be conveyed through the tunnel to a<br />

single pump station located on plant property.<br />

In addition to cost, factors that were considered in selecting the type of <strong>RBF</strong> collection system<br />

included construction and maintenance.<br />

Construction Considerations<br />

The two main construction issues that were evaluated <strong>for</strong> Option 1 were the feasibility of<br />

constructing the tunnel in the alluvial deposits along the riverbank and the feasibility of installing<br />

laterals safely from within the tunnel. With regard to Options 2 and 3, the main construction issue<br />

evaluated was rock quality.<br />

To evaluate these factors, a detailed subsurface investigation was conducted that included the<br />

following elements:<br />

• Large-diameter bucket auger borings to bedrock.<br />

• Grain-size analysis of composite samples obtained from the bucket auger borings.<br />

• Core borings of bedrock materials.<br />

• Examination of rock outcrops near the site.<br />

Large-diameter bucket auger borings were used to collect representative samples of alluvium.<br />

Because alluvium contains significant amounts of coarse gravel, cobbles, and possibly boulders, the


Ohio River<br />

Southern Caisson<br />

(capped at grade)<br />

19-ft Diameter<br />

Soft Soft-Ground<br />

Ground Tunnel unnel<br />

Tunnel<br />

Caisson<br />

Capped<br />

at Grade<br />

Ground Surface<br />

Pump<br />

Station<br />

Lagoon<br />

Existing Collector Well<br />

Existing 48-Inch<br />

Raw <strong>Water</strong> Transmission Line<br />

New 54-Inch<br />

Raw <strong>Water</strong> Transmission Line<br />

Lagoon<br />

Lagoon<br />

Clay<br />

Lagoon<br />

Sand and Gravel<br />

14-ft Diameter Tunnel with 25 Laterals<br />

Bedrock<br />

Plan View 1A<br />

Existing (Two)<br />

60-Inch Diameter<br />

<strong>Water</strong> Intake Pipes<br />

Northern Caisson<br />

(Pump Station)<br />

32-ft Diameter Shaft<br />

200-ft Laterals<br />

Existing B.E. Payne<br />

<strong>Water</strong> Treatment Plant<br />

15+0 10+0 5+00 0+00<br />

Figure 1. Soft ground tunnel concept (Option 1).<br />

Section View 1B<br />

large-diameter bucket auger was considered the most appropriate sample collection method.<br />

Grain-size distribution curves were developed based on sieve analyses of the bucket auger samples.<br />

The results of the rock-core boring program indicated the presence of limestone and shale. A<br />

particularly good layer of limestone occurred from 130- to 157-ft (40- to 48-m) below ground<br />

surface. Limestone beds in this unit tend to be on the order of 2- to 3-ft (0.6- to 0.9-m) thick,<br />

interbedded with 3- to 4-inch (7- to 10-centimeter) thick layers of shale. All of the limestone in<br />

this layer has a rock quality designation in the range of 90 to 100 percent, which according to<br />

Deere and Deere (1989) is classified as “excellent.”<br />

9


10<br />

Ohio River<br />

Pump Station<br />

Collector Well<br />

19-ft Diameter<br />

Drop Shaft<br />

4-ft Diameter<br />

Based on the results of the subsurface investigation, the construction of either type of tunnel<br />

system (soft ground or bedrock) was determined to be feasible.<br />

Maintenance Considerations<br />

Collector<br />

Well<br />

Collector<br />

Hard Rock<br />

Hard Roc Rock Tunnel unnel<br />

Lagoon<br />

200-ft<br />

Diameter Laterals<br />

11 per Collector Well<br />

7-ft Diameter<br />

Tunnel<br />

Existing<br />

Collector Well<br />

Lagoon<br />

Existing 48-Inch<br />

Raw <strong>Water</strong><br />

Transmission Line<br />

Lagoon<br />

Lagoon<br />

Pump Station<br />

32-ft Diameter Shaft<br />

Existing (2) 60-Inch Diameter<br />

<strong>Water</strong> Intake Pipes<br />

New 54-Inch Raw <strong>Water</strong><br />

Transmission Line<br />

200-ft<br />

Diameter Laterals<br />

11 per Collector Well Sand and<br />

Gravel<br />

24-ft Diameter<br />

Shaft<br />

in Bedrock<br />

Collector Well<br />

19-ft Diameter<br />

8+00 4+00 0+00<br />

Figure 2. Hard-rock tunnel concept using horizontal collector wells (Option 2).<br />

Existing B.E. Payne<br />

<strong>Water</strong> Treatment Plant<br />

Ground<br />

Surface<br />

Bedrock<br />

Plan View 2A<br />

One of the unique challenges of Option 1 would be the long-term maintenance of collector<br />

laterals. Over time, it is anticipated that the laterals would need to be cleaned to restore their<br />

yield, as would be the case with any well screen. Several options were considered that would allow<br />

individual laterals to be taken out of service <strong>for</strong> cleaning while the rest of the tunnel system<br />

remained in operation. The selected option consisted of a “wet/dry” tunnel system, with water<br />

being conveyed through carrier pipes in the “wet” side and O&M being per<strong>for</strong>med from within<br />

Clay<br />

Drop Shaft<br />

4-ft Diameter<br />

Section View 2B


TBM Exit<br />

Construction<br />

Vertical Collectors<br />

Installedon 200-ft Centers<br />

<strong>RBF</strong> Pump Station<br />

Ohio River<br />

Existing B.E. Payne<br />

<strong>Water</strong> Treatment Plant<br />

Capped Well Capped Well<br />

Clay<br />

Sand &<br />

Gravel<br />

Aquifer<br />

Bedrock<br />

Well Casing<br />

<strong>Water</strong> Table Level<br />

Wellscreen<br />

Tunnel<br />

Figure 3. Hard-rock tunnel concept using vertical collectors (Option 3).<br />

TBM Construction<br />

Existing Collector<br />

Well/Pump Station<br />

Approx. 2,000 ft<br />

Tunnel Extension<br />

(no wells)<br />

Existing Low Lift<br />

Pump Station<br />

Plan View 3A<br />

the “dry” side. This construction option would allow laterals to be maintained individually, while<br />

all other laterals remain in service. Figure 4 illustrates the “wet/dry” tunnel concept.<br />

Maintenance requirements <strong>for</strong> Option 2 would involve taking an entire caisson out of service <strong>for</strong><br />

lateral rehabilitation. This concept would reduce the effective capacity of the overall system by<br />

the capacity of an entire caisson, or approximately one-third of a three-caisson system.<br />

Option 3 would consist of individual wells that could be individually taken out of service and<br />

rehabilitated. A packer could be used to isolate the vertical collector from the tunnel. Based on<br />

maintenance considerations, Option 3 was preferred.<br />

3B<br />

To Pump<br />

Station<br />

Section View<br />

11


12<br />

Figure 4. Wet/dry tunnel concept.<br />

Hydrogeologic Yield<br />

Computer modeling was used to predict the safe yield of the aquifer and to optimize the placement<br />

of collectors <strong>for</strong> all of the options. For Option 1, collector laterals would be positioned 60 ft<br />

(18.3 m) on center on the riverside of the tunnel (Schafer, 2000). Using a 1,500-ft (457-m) long<br />

tunnel with laterals located on 60-ft (18.3-m) centers resulted in an estimated yield ranging from<br />

44 to 61 MGD (167,000 to 231,000 m 3 /d), depending on water temperature. For Option 2, the<br />

optimum number of laterals per collector well was determined to be 10 to 11 (Schafer, 2000) and<br />

the estimated yield ranged from 37 to 51 MGD (140,000 to 193,000 m 3 /d), depending on water<br />

temperature, <strong>for</strong> a system consisting of three new collector wells. For Option 3, approximately<br />

30 vertical collectors would be installed approximately 200-ft (67-m) apart, resulting in an<br />

estimated yield ranging from 38 to 44 MGD (144,000 to 167,000 m 3 /d) (Shafer, 2002).<br />

Data from the existing collector well and pumping tests were used as the basis <strong>for</strong> selecting<br />

leakance and transmissivity values that were used in calculating the predicted yields. Figure 5<br />

shows measured leakance data from the existing collector. As shown, leakance declined over the<br />

initial 2-year period of operation from over 2.0 to 0.148 inverse days due to clogging in the<br />

riverbed. Within the past year, measured leakance values appear to have stabilized and increased<br />

slightly. To be conservative in predicting yields <strong>for</strong> the new <strong>RBF</strong> addition, the lowest observed<br />

leakance value was used.<br />

<strong>Water</strong> Quality and Treatment<br />

Tunnel Cross-Section<br />

Gate Valves with<br />

Removable Elbow<br />

Precast<br />

Concrete<br />

Tunnel Segments<br />

12-Inch Lateral<br />

Concrete Fill<br />

Two 48-Inch Carrier Pipes<br />

<strong>Water</strong> quality obtained from the collector wells in many locations has been documented as having<br />

many advantages over that obtained directly from surface-water sources (Sontheimer, 1980;<br />

Hubbs, 1981; Wang et al., 1995; <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong>, 1999; Kuehn and Mueller,<br />

2000). In the Louisville <strong>Water</strong> Company’s case, the existing <strong>RBF</strong> collector well has shown many


Leakance (inverse days)<br />

10<br />

1<br />

0.1<br />

10/1/1999 3/31/2000 9/30/2000 3/31/2001 9/30/2001 3/31/2002 9/30/2002<br />

favorable improvements in water quality compared to the Ohio River source (Wang et al., 2002),<br />

including:<br />

• Lower turbidity.<br />

• Lower TOC.<br />

• Increased minimum temperature.<br />

• Reduced microbial contaminants.<br />

Date<br />

Figure 5. Measured leakance values corrected to 68-degrees Fahrenheit (October 1999).<br />

Table 1 summarizes typical water quality from the collector well compared to local groundwater<br />

and Ohio River sources.<br />

Table 1. Characteristics of River <strong>Water</strong> and Groundwater in the Project Area<br />

Infiltrated Bedrock<br />

Parameters River <strong>Water</strong> Groundwater Groundwater<br />

pH 1 7.7 to 7.9 7.4 to 7.5 7.2 to 7.3<br />

Total Hardness (mg/L) 1 90 to 205 280 to 290 530 to 582<br />

Total Alkalinity (mg/L) 1 50 to 110 235 to 250 260 to 280<br />

TOC (mg/L) 1 2.1 to 4.9 0.3 to 0.6 0.4 to 0.7<br />

Turbidity (ntu) 1 2 to 1,500 5.0


14<br />

The turbidity of well water is typically less than 0.08 nephelometric turbidity units (ntu), which<br />

could allow the Louisville <strong>Water</strong> Company to bypass the coagulation treatment process and use<br />

direct filtration <strong>for</strong> treating water. Reduced turbidity will result in lower coagulant dosages and less<br />

sludge production. The reduction of organic material in the water will provide lower levels of<br />

DBPs in the finished water.<br />

With minimum water temperatures of the Ohio River source near freezing, water main breaks are an<br />

issue during the winter months. By having a more moderate water temperature and using more <strong>RBF</strong><br />

water, this will ultimately result in fewer main breaks. The Louisville <strong>Water</strong> Company also noted the<br />

lack of Geosmin and 2-methylisoborneol in well water during a taste and odor episode that occurred<br />

in 1999, resulting in the ability to eliminate the need to add powdered activated carbon.<br />

Well water, however, has other water-quality issues that need to be addressed in the design of the<br />

system. Data from the Louisville <strong>Water</strong> Company shown in Table 1 indicates the following<br />

treatment issues:<br />

• Increased hardness.<br />

• Presence of radon.<br />

• Low dissolved oxygen.<br />

Additionally, water quality of the aquifer downstream from the Payne Plant has been documented<br />

as having elevated levels of iron (Schafer, 2000). This will be a consideration as the Louisville<br />

<strong>Water</strong> Company proceeds with implementing <strong>RBF</strong> <strong>for</strong> the Crescent Hill <strong>Water</strong> Treatment Plant,<br />

but is not a factor <strong>for</strong> the current design project at the Payne Plant. Because the Payne Plant is<br />

already a softening plant, the increased hardness of <strong>RBF</strong> water will be treated using existing basins.<br />

Radon levels of 180 pCi/L have been noted in the existing collector well. While this is below the<br />

anticipated maximum contaminant level of the future Radon Rule (300 pCi/L), the Louisville<br />

<strong>Water</strong> Company desires to set a goal of 0 pCi/L, because the surface-water source currently<br />

supplying water to its customers does not contain radon. Treatment <strong>for</strong> radon will involve the<br />

installation of an in-line aerator.<br />

Dissolved oxygen from well water is less than 0.1 mg/L. Currently, as water flows through<br />

the basins of the existing treatment plant, oxygen levels are naturally increased up to about<br />

5 mg/L. The same in-line aerator used <strong>for</strong> radon removal can be used to impart higher levels of<br />

oxygen into the water, if needed.<br />

Summary<br />

The results of the subsurface evaluation show that both tunnel options are feasible <strong>for</strong> <strong>RBF</strong>. With<br />

either Option 1 or 3, the collector laterals would be evenly spaced along the entire length of the<br />

collection system, providing <strong>for</strong> more even hydraulic head distribution across the riverbank and,<br />

thus, better overall yield from the installation. With Option 2, the hydraulic head would be<br />

focused at the collector wells, resulting in less than maximum yield from the aquifer system.<br />

Maintenance flexibility would be greater <strong>for</strong> either Option 1 or 3, which would allow each<br />

lateral/vertical collector to be maintained independently of the operation of the remaining<br />

laterals/vertical collectors. Cost was a determining factor in selecting Option 3. In optimizing the<br />

design of the selected <strong>RBF</strong> system, the design engineers considered both water-quantity and waterquality<br />

issues. <strong>Water</strong>-quantity predictions were made using conservative assumptions based on<br />

data from the existing collector well. <strong>Water</strong>-quality issues were identified and appropriate<br />

treatment technologies are being incorporated into the design.


REFERENCES<br />

Deere, D.U., and D.W. Deere (1989). “Rock Quality Designation (RQD) After Twenty Years.” Contract<br />

Report No. 6L-89-1, U.S. Army Corps of Engineers.<br />

Hubbs, S.A. (1981). “Organic reduction — Riverbank Infiltration at Louisville.” Proceedings, American<br />

<strong>Water</strong> Works Association Annual Conference, St. Louis, Missouri.<br />

Kuehn, W., and U. Mueller (2000). “Riverbank Filtration: An Overview.” Journal AWWA, 92(12): 60.<br />

<strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong> (1999). Abstracts, International Riverbank Filtration Conference, November<br />

4-5, 1999, Louisville, Kentucky.<br />

Schafer, D. (2001). Evaluation of Collector Well Production Capacity B.E. Payne <strong>Water</strong> Treatment Plant, David<br />

Schafer & Associates, Inc., Project Report.<br />

Schafer, D. (2000). Hydraulics Analysis of Groundwater Extraction at the B.E. Payne <strong>Water</strong> Treatment Plant,<br />

David Schafer & Associates, Inc., Project Report.<br />

Sontheimer, H. (1980). “Experience with Riverbank Filtration Along the Rhine River.” Journal AWWA,<br />

72(7): 386.<br />

Wang, J., J. Smith, and L. Dooley (1995). “Evaluation of Riverbank Infiltration as a Process <strong>for</strong> Removing<br />

Particles and DBP Precursors.” Proceedings, American <strong>Water</strong> Works Association <strong>Water</strong> Quality Technology<br />

Conference, New Orleans, Louisiana.<br />

Wang, J.Z., S.A. Hubbs, and R. Song (2002). Evaluation of Riverbank Filtration as a Drinking <strong>Water</strong> Treatment<br />

Process, American <strong>Water</strong> Works Association <strong>Research</strong> Foundation Report Number 90922.<br />

Wang, J.Z. (2003). Personal communication.<br />

DAVID HAAS is a Senior Project Manager with Jordan, Jones & Goulding, Inc., an<br />

Atlanta-based consulting firm that offers engineering, management, and planning<br />

services. Haas has over 18 years of experience in municipal water supply, treatment, and<br />

distribution system projects. Currently, he is the Project Manager <strong>for</strong> the Louisville <strong>Water</strong><br />

Company’s 45-million gallons per day riverbank-filtration project at the B.E. Payne <strong>Water</strong><br />

Treatment Plant in Louisville, Kentucky. He is also a contributing author of Riverbank<br />

Filtration: Improving Source-<strong>Water</strong> Quality, jointly published by Kluwer Academic<br />

Publishers and the <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong> in 2002. Haas received both a B.S. and M.S. in<br />

Environmental Engineering from the University of Louisville. He is a Professional Engineer in the States of<br />

Georgia, Kentucky, and Tennessee.<br />

15


Session 2: Operations<br />

Construction and Maintenance of Wells<br />

<strong>for</strong> Riverbank Filtration<br />

Henry C. Hunt, CPG<br />

Collector Wells International, Inc.<br />

Columbus, Ohio<br />

By description, <strong>RBF</strong> implies that we are designing something that will allow water to be infiltrated<br />

from a surface-water source through riverbank (and river-bottom) deposits. This allows physical<br />

(e.g., suspended) particles to be filtered out of source water in an attempt to “pretreat” raw water<br />

be<strong>for</strong>e it reaches a treatment plant or enters a distribution system with some sort of primary<br />

treatment, such as chlorination.<br />

While the term “<strong>RBF</strong>” is a relatively new term in the United States, well and gallery systems have<br />

been used in international settings to develop water supplies using induced infiltration dating back<br />

to the 1800s. Many well and infiltration systems have relied on <strong>RBF</strong> to provide recharge into<br />

alluvial aquifers to replace groundwater pumped at sites all across the United States, but only<br />

recently have these sites been coined as “<strong>RBF</strong>” sites. <strong>Water</strong>-supply facilities at such sites typically<br />

include vertical wells, infiltration galleries, and radial collector wells.<br />

What makes <strong>RBF</strong> work is that the water level in the aquifer is drawn down by pumping from a<br />

well or gallery system located adjacent to a surface-water source, and water from the surface-water<br />

body is then induced to infiltrate into the aquifer as hydraulic gradients are developed from<br />

pumping. The schematic in Figure 1 shows this general relationship <strong>for</strong> horizontal collector wells<br />

and conventional vertical wells.<br />

Plan<br />

Elevation<br />

Screened<br />

Pipe<br />

River River<br />

Horizontal Vertical<br />

Figure 1. Well systems develop capacity through induced infiltration (Ray et al., 2002a).<br />

Flow<br />

Correspondence should be addressed to:<br />

Henry C. Hunt, CPG<br />

Senior Project Manager/Hydrogeologist<br />

Collector Wells International, Inc.<br />

6360 Huntley Road • Columbus, Ohio 43229 USA<br />

Phone: (614) 888-6263 • Fax: (614) 888-9208 • Email: hchunt@collectorwellsint.com<br />

Flow<br />

17


18<br />

Riverbank Filtration Suitability<br />

A <strong>RBF</strong> system is designed to infiltrate water from an adjacent surface-water source, using streambed<br />

and riverbank deposits to naturally filter out suspended materials from source water. The first (and<br />

obvious) requirement is that the facilities be placed in close proximity to a source of recharge, such<br />

as a river. During the feasibility and siting stages of a project, a number of criteria must be<br />

considered, including:<br />

• Availability of water from a surface-water source that can recharge the aquifer.<br />

• An efficient hydraulic interconnection between the river and aquifer.<br />

• Suitable water quality in the surface-water source.<br />

• Sustainable flow in the river to match anticipated withdrawal rates.<br />

A detailed hydrogeologic investigation is required to verify that aquifer conditions are favorable<br />

<strong>for</strong> considering a <strong>RBF</strong> facility to meet project water demands. The investigation must evaluate the<br />

geology of the aquifer and the interconnection between the river and groundwater in the aquifer<br />

to determine the feasibility of inducing infiltration from the river, evaluate possible well designs,<br />

and determine likely well yields. This type of investigation typically includes exploratory test<br />

drilling, aquifer (pumping) testing, and data analysis to develop the needed in<strong>for</strong>mation <strong>for</strong> each<br />

prospective project site. Based on the results of this investigation, well designs are developed and<br />

alternatives are compared <strong>for</strong> feasibility, yield, and cost. These alternative designs are generally<br />

discussed below.<br />

Vertical Wells<br />

Vertical wells represent the most conventional method <strong>for</strong> developing a groundwater supply in the<br />

country. These wells consist of a vertical borehole that is usually completed with a screen and riser<br />

casing to allow water to enter from the <strong>for</strong>mation and be pumped from the borehole via a pumping<br />

system installed within the riser casing. A diagram of a typical vertical well constructed in an<br />

unconsolidated (e.g., sand and gravel) aquifer is shown in Figure 2. Vertical wells can be used<br />

effectively when small to moderate yields are needed, or when a system is growing slowly over a<br />

longer period of time, such that adding a well to the system every year or so meets growing water<br />

demands. To meet larger demands, a series of vertical wells must be installed, spreading along<br />

riverfront areas or grouped into well-field clusters.<br />

Vertical wells can be constructed using a variety of drilling methods, including mud rotary, reverse<br />

circulation, cable tool, bucket-auger, dual-rotary, and air rotary. The method selected <strong>for</strong> each site<br />

will take into account a combination of well criteria, including well depths, diameters, water-table<br />

elevations, screen requirements, and potential problems caused by geologic conditions<br />

encountered. These wells are constructed by drilling a vertical borehole, and then installing the<br />

desired well riser casing and screen in the borehole. In some cases, an artificial gravel-pack filter<br />

is also added to help prevent the intrusion of sand and silt from the <strong>for</strong>mation during pumping.<br />

Collector Wells<br />

A collector well, also called a horizontal collector or radial well, differs from a vertical well in that<br />

the well screen is installed horizontally into the aquifer <strong>for</strong>mation from a central rein<strong>for</strong>ced<br />

concrete caisson that serves as the wet well pumping station. These wells are constructed by<br />

sinking sections of rein<strong>for</strong>ced concrete (called lifts) into the aquifer adjacent to the river until the<br />

lower portion of the caisson reaches the design elevation <strong>for</strong> installing the well screen. Individual


Land<br />

Surface<br />

lengths of well screens are then projected out into the aquifer in a variety of patterns near the base<br />

of the alluvial aquifer or at another pre-selected horizon within the aquifer. Where <strong>RBF</strong> is desired,<br />

the pattern of lateral well screens is predominantly beneath the river. This allows the “pumping<br />

center” <strong>for</strong> the collector well to be shifted closer to the river, usually in an attempt to increase the<br />

percentage of river water that is infiltrated into the aquifer. This design capability usually permits<br />

the highest percentage of river water to be achieved in a riverbank setting. The ability to install<br />

the well screens horizontally in the aquifer beneath the river permits longer lengths of well screen<br />

to be installed, per site, typically maximizing the yield possible from each well site. It is common<br />

<strong>for</strong> a collector well to produce a yield equivalent to multiple vertical wells from the same well site.<br />

A general schematic of a typical collector well is shown in Figure 3.<br />

Well Selection<br />

<strong>Water</strong><br />

Pump to<br />

<strong>Water</strong><br />

Treatment<br />

Plant<br />

Well<br />

Casing<br />

Pump House<br />

Electric Motor<br />

Approximately<br />

1-m High<br />

Rotating<br />

Shaft<br />

Cone of Depression<br />

Pump Bowl<br />

<strong>for</strong> a Turbine Pump<br />

Well Screen<br />

Gravel Pack<br />

<strong>Water</strong><br />

Figure 2. Typical components of a vertical well (Ray et al., 2002b).<br />

Original <strong>Water</strong> Surface<br />

The hydrogeologic investigation determines the hydraulic characteristics of the aquifer <strong>for</strong>mation<br />

necessary to determine the potential yield possible from either a collector well or a series of<br />

vertical wells. This data allows a comparison to be made of well designs to meet project water<br />

demands, usually considering a single collector well versus a series of vertical wells, with<br />

connecting pipeline, electrical service, access roads, etc., to produce the equivalent yield (it is<br />

common <strong>for</strong> a collector well to produce a yield equal to 5 to 10 vertical wells). As the capital costs<br />

to install the “complete system” are compared, it is often very competitive to consider a collector<br />

well to meet moderate to very large capacities, and more competitive to consider a vertical well<br />

when lower capacities are required (<strong>for</strong> example, when incremental increases in capacity are<br />

projected over a number of years).<br />

River<br />

19


20<br />

Laterals<br />

Alternate Systems<br />

In addition to vertical wells and radial collector wells, infiltration galleries can be used to develop<br />

filtered surface-water supplies through <strong>RBF</strong>. These can include trenching parallel or beneath the<br />

river and installing screened gallery pipes that can deliver filtered surface water to sumps and wet<br />

wells. It is also possible to use horizontal directionally drilled wells to induce infiltration from an<br />

adjacent river. These systems are often installed under low head conditions, such that lower per<br />

foot yields are obtained, requiring longer gallery lengths to meet higher capacity needs. It is often<br />

difficult to per<strong>for</strong>m effective well-screen maintenance on these systems.<br />

Well Maintenance<br />

Pump Shaft<br />

Pump<br />

Pump House<br />

Central<br />

Collection<br />

Caisson<br />

Well Screens<br />

(in Laterals)<br />

Figure 3. Typical components of a radial collector well (Ray et al., 2002b).<br />

Land Surface<br />

It is normal <strong>for</strong> well screens to become plugged with chemical (mineral) precipitates and biological<br />

growths (e.g., iron bacteria) over time in alluvial aquifers. The rate of plugging can be exacerbated<br />

if the screen design creates excessive entrance velocities through the slot openings in the well<br />

screens, so it is important that proper screen design considers the water quality and hydraulic<br />

characteristics of the aquifer to maintain entrance and approach velocities within acceptable<br />

ranges to minimize this rate of plugging and extend the intervals between well cleanings. Since<br />

the lineal footage of well screen in a collector well is longer than <strong>for</strong> a vertical well, entrance<br />

velocities are minimized so that the intervals between required maintenance can be extended.<br />

This typically results in lower O&M costs over the life of the collector well.<br />

Well maintenance can be accomplished using a variety of methods including mechanical,<br />

chemical, or a combination of methods. The most effective method to restore well-screen<br />

openings and well efficiency will vary from well field to well field, and is selected based upon<br />

details of the well construction (e.g., screen design), groundwater quality, past results, the nature<br />

of plugging (mineralogic versus biological), and other factors. The use of an ongoing monitoring<br />

and record-keeping program should allow you to track well per<strong>for</strong>mance, identify operating trends,<br />

and predict optimal times <strong>for</strong> per<strong>for</strong>ming maintenance.


Summary<br />

Well systems can be constructed in alluvial aquifer systems at many locations to induce infiltration<br />

from adjacent surface-water sources to provide a pre-filtered raw-water supply. The design of these<br />

systems should be selected after evaluating site-specific aquifer conditions, water-quality<br />

objectives, and project water demands to ensure that the most effective system is selected. There<br />

are many well systems operating in the United States and overseas, demonstrating that <strong>RBF</strong> can<br />

be used effectively to meet water-quality and capacity objectives.<br />

REFERENCES<br />

Ray, C., T. Grischek, J. Schubert, J. Wang, and T. Speth (2002a). “A perspective of riverbank filtration.”<br />

Journal AWWA, 94(4): 149-160.<br />

Ray, C., G. Melin, and R. Linsky, editors (2002b). Riverbank Filtration: Improving Source-<strong>Water</strong> Quality,<br />

Kluwer Academic Publishers, Dordrecht.<br />

HENRY HUNT is a Senior Project Manager and Hydrogeologist with Collector Wells<br />

International, Inc., which specializes in the design, construction, and rehabilitation of<br />

water-supply systems <strong>for</strong> public drinking water, industrial process, or power plant cooling<br />

water. Hunt has over 25 years experience involving riverbank-filtration systems, ranging<br />

from the inspection and rehabilitation of existing well and infiltration systems to the<br />

design and construction of new water wells and riverbank-filtration facilities. He has<br />

special expertise concerning the inspection, evaluation, siting, and testing of radial<br />

collector wells along many alluvial valleys across the United States and overseas. In addition, he has authored<br />

a number of papers regarding water-supply development, infiltration galleries, radial collector wells, and<br />

seawater collector wells, including several chapters in the 2002 publication Riverbank Filtration: Improving<br />

Source-<strong>Water</strong> Quality and the newly revised American <strong>Water</strong> Works Association Manual 21 on Groundwater.<br />

Hunt received a B.A. in Geology from Lafayette College in Pennsylvania, with a minor in Civil Engineering.<br />

21


Session 2: Operations<br />

Aquifer Storage and Recovery Pretreatment:<br />

Synergies of Bank Filtration, Ozonation,<br />

and Ultraviolet Disinfection<br />

Robert Stanley Cushing, Ph.D., P.E.<br />

Carollo Engineers<br />

Sarasota, Florida<br />

R. David G. Pyne, P.E.<br />

ASR Systems<br />

Gainesville, Florida<br />

David R. Wilkes, P.E.<br />

Jordan, Jones & Goulding<br />

Norcross, Georgia<br />

Introduction<br />

Aquifer Storage and Recovery (ASR) is a technique in which water is stored in subsurface aquifers<br />

during high-supply, low-demand periods <strong>for</strong> use during low-supply, high-demand periods. As<br />

indicated in Figure 1, water is pumped into a confined aquifer, displacing the native water, and<br />

<strong>for</strong>ms a zone in the aquifer comprised primarily of injected water. <strong>Water</strong> is subsequently extracted<br />

from the aquifer when needed to supplement finished water-production capacity.<br />

Native<br />

Ground<br />

<strong>Water</strong><br />

Buffer<br />

Zone<br />

Stored<br />

<strong>Water</strong><br />

ASR Well<br />

Confining Layer Confining Layer<br />

Figure 1. Schematic representation of ASR.<br />

Stored<br />

<strong>Water</strong><br />

Target Storage Volume<br />

Confining Layer<br />

Correspondence should be addressed to:<br />

Robert S. Cushing, Ph.D., P.E.<br />

Partner<br />

Carollo Engineers, P.C.<br />

401 North Cattlemen Road • Suite 306 • Sarasota, Florida 34232 USA<br />

Phone: (941) 371-9832 • Fax: (941) 371-9873 • Email: RCushing@carollo.com<br />

Buffer<br />

Zone<br />

Native<br />

Ground<br />

<strong>Water</strong><br />

23


24<br />

Traditionally, ASR has been used to store potable water to meet seasonal limitations in source-water<br />

supply or to optimize the use of water-treatment capacity. For example, the Peace River Regional <strong>Water</strong><br />

Supply Authority in Florida draws and treats water from the Peace River at a rate limited by minimum<br />

flows and levels. By storing treated water through ASR during high river flow periods, the Authority<br />

has been able to more than double the usable yield from this source, while avoiding adverse impacts to<br />

sensitive downstream ecosystems. Mt. Pleasant <strong>Water</strong> Works in South Carolina supplements their<br />

water supply using reverse-osmosis treatment of brackish groundwater. ASR allows the reverse-osmosis<br />

plant to operate at a relatively constant rate, minimizing the production cost from this facility.<br />

Emerging ASR applications include storing water types other than finished potable water, including<br />

partially treated surface water and reclaimed wastewater, as well as different uses <strong>for</strong> ASR production<br />

water (such as water-treatment plant source-water, irrigation, and environmental applications).<br />

Biological and physiochemical processes during ASR storage can remove certain contaminants,<br />

including THMs and HAAs, leading to ASR applications that provide treatment, as well as storage.<br />

<strong>Water</strong>-quality goals <strong>for</strong> water to be injected into an ASR well vary with application. Generally, the<br />

least restrictive goal is to produce a water with chemical and particle characteristics that prevent<br />

excessive loss of permeability in the injection well aquifer. Some applications require injection water<br />

to meet primary (or public health-related drinking water) standards to protect aquifer quality or to<br />

meet ASR production requirements. Treatment to secondary or aesthetic water-quality standards may<br />

also be required. The most restrictive water-quality goals require treatment <strong>for</strong> trace, unregulated<br />

contaminants (e.g., low levels of endocrine disrupting compounds). Treatment system complexity,<br />

O&M requirements, and cost are all strongly related to source-water quality and finished water-quality<br />

goals, with ASR pretreatment costs ranging well over an order of magnitude, depending on the<br />

particular application.<br />

One general treatment approach <strong>for</strong> meeting ASR pretreatment goals consists of bank filtration<br />

followed by ozonation and/or UV disinfection (Figure 2). This process combination offers the<br />

improved reliability of a multiple-barrier approach while maintaining cost and operational<br />

efficiencies due to the synergistic nature of the individual unit processes. The following discussion<br />

provides a summary of treatability study results, engineering analysis, and cost-estimates <strong>for</strong> this<br />

process configuration.<br />

Lake Okeechobee<br />

Figure 2. Integrated bank filtration/ozonation/UV disinfection process train.<br />

Discussion<br />

Bank Filtration<br />

Ozone Contactor/<br />

Reaction Basin<br />

UV Disinfection<br />

Traditional treatment approaches, such as membrane filtration or conventional treatment,<br />

produce process residuals that generally have substantial cost and operational implications. In<br />

addition, these traditional approaches require substantial chemicals <strong>for</strong> process O&M, an aspect


that is important when considering both reliability and security. A process train comprised of bank<br />

filtration followed by ozonation and/or UV disinfection produces no residuals and has minimal<br />

chemical requirements. This process combination offers a number of benefits, including cost,<br />

flexibility, and multiple barriers to pathogens and other contaminants.<br />

For this process combination, the barriers to pathogens operate by distinctly different mechanisms,<br />

enhancing the robustness of the multiple-barrier approach and providing the first example of<br />

process synergy. Bank filtration provides physical removal and physical/chemical/biological<br />

inactivation of pathogens, ozonation provides chemical disinfection, and UV provides<br />

disinfection through a fundamentally different mechanism.<br />

While bank filtration removes color and taste- and odor-causing compounds, ozonation provides<br />

a means of further improving these water-quality characteristics. Ozonation also increases<br />

UV transmittance, decreasing the capital and operational cost of the UV disinfection process.<br />

Case Study Description and Objectives<br />

As part of the Comprehensive Everglades Restoration Program, the U.S. Army Corps of Engineers<br />

— in conjunction with the South Florida <strong>Water</strong> Management District — have initiated pilot<br />

studies to evaluate potential treatment options <strong>for</strong> an ASR project that includes hundreds of ASR<br />

wells. Treatment systems will be designed to treat surface water prior to injection into groundwater<br />

storage zones, with ultimate capacity approaching 1.5-billion gallons of water per day.<br />

The purpose of the pilot-testing project was to demonstrate surface-water treatment technologies<br />

to meet multiple water-quality goals prior to injection into ASR wells. Pilot-test data and<br />

engineering analysis were used to develop optimized design criteria, as well as capital and O&M<br />

costs <strong>for</strong> full-scale treatment systems. The specific primary objectives of the study were to:<br />

• Determine raw-water quality during the pilot study.<br />

• Collect and analyze unit process and treatment train per<strong>for</strong>mance data.<br />

• Establish process feasibility and design criteria <strong>for</strong> full-scale water treatment systems.<br />

• Determine preliminary capital and operating cost estimates <strong>for</strong> each of the unit processes<br />

considered.<br />

Methodology<br />

Two parallel treatment trains were operated during most of the study:<br />

• Treatment Train 1 — Simulated bank filtration/ozonation.<br />

• Treatment Train 2 — Simulated bank filtration/UV disinfection.<br />

During later stages of the study, two additional trains were tested. Simulated bank filtration was<br />

bypassed and a direct surface-water intake (with wire mesh filtration followed by ozonation and<br />

UV disinfection) was evaluated. Figure 3 illustrates the treatment trains, with dashed lines labeled<br />

“Alternative Treatment Train” showing the supplemental testing.<br />

<strong>Water</strong> was taken from the St. Lucie Canal (eastern shore of Lake Okeechobee) through an intake<br />

header that consisted of slotted polyvinyl chloride pipe. The water was pumped from the intake<br />

pipe through a centrifugal pump and into the simulated bank-filtration system, consisting of a<br />

wedge-wire filter (Parker TruClean filtration system) followed by a gravity sand filter<br />

(10-ft × 10-ft area, with 36-inches of 0.45 to 0.55-millimeter diameter sand). Note that original<br />

25


26<br />

plans <strong>for</strong> a pilot-scale bank-filtration intake were changed by the U.S. Army Corps of Engineers<br />

and South Florida <strong>Water</strong> Management District. Simulated bank-filtration results were analyzed<br />

directly and extrapolated to consider the impacts of full-scale bank-filtration intake.<br />

Results<br />

Decant<br />

<strong>Water</strong> Source Submersible<br />

Pump<br />

Figure 3. Pilot process flow schematic.<br />

Residual<br />

Wedge Wire<br />

Filter<br />

Settling Tank<br />

Residual<br />

Sand Filter<br />

Solids Land<br />

Applied On-Site<br />

Alternate Treatment Train<br />

Holding<br />

Tank<br />

Alternate Treatment Train<br />

Raw-water quality during the study (August 2002 to October 2002) was generally consistent with<br />

historical trends. TOC levels were high, averaging 27 mg/L, with a maximum concentration of<br />

49 mg/L. Correspondingly, true color was also elevated, averaging 96 color units, with a maximum<br />

of 236 color units. Turbidity was variable, ranging from 5 to 57 ntu, with an average of 15 ntu.<br />

The simulated bank-filtration unit process provided particle removal averaging 91 percent and<br />

turbidity removal averaging 78 percent. The removal of dissolved constituents (e.g., TOC and<br />

color) was minimal. As expected, due to the short residence time in the simulated bank-filtration<br />

system (1 to 3 hours), particle removal through the simulated system was less than expected<br />

through a full-scale bank-filtration system, and microbial activity and the associated removal of<br />

organic material was not significant.<br />

Ozone was capable of removing color to meet and exceed the secondary maximum contaminant<br />

level of 15 color units and to reduce UV absorbance. Due to the high ozone doses required to<br />

maintain a residual and rapid decay rates, achieving disinfection credit through ozonation was not<br />

feasible. In addition, with bromide concentrations approaching 300 micrograms per liter (µg/L),<br />

achieving significant disinfection and maintaining bromate concentrations below the maximum<br />

contaminant level of 10 micrograms per milliliter would be infeasible; however, ozone doses used<br />

to reduce color and improve UV transmittance (from a minimum of 13-percent UV transmittance<br />

in raw water to greater than 65-percent UV transmittance in ozonated water).<br />

Due to the synergistic relationship between ozone and UV disinfection, full-scale design criteria<br />

<strong>for</strong> the two processes were analyzed concurrently to minimize water production costs. As ozone<br />

dose goes up, the cost to purchase and operate the ozone system rises; however, due to the increase<br />

in UV transmittance with higher doses, UV-disinfection system costs decrease. An average ozone<br />

dose of approximately 6 mg/L and maximum dose of 11 mg/L produced a UV transmittance of<br />

65 to 72 percent and a minimum production cost (amortized capital, plus O&M). The UV system<br />

UV<br />

O 3<br />

Effluent<br />

to Canal


was designed to deliver a dose of 140 millijoules per square centimeter to meet all disinfection<br />

goals <strong>for</strong> Giardia, Cryptosporidium, and viruses at a design UV transmittance of 60 percent.<br />

The cost of a 5-MGD bank filtration/ozonation/UV-disinfection treatment facility is approximately<br />

$4,759,000 <strong>for</strong> design and construction and $147,000 per year <strong>for</strong> O&M (or $0.20 per 1,000 gallons<br />

of water produced). This translates into a normalized capital cost of $0.95 per gallons per day of<br />

capacity and a production cost of $0.70.<br />

BOB CUSHING has 14 years of experience in applied environmental science and<br />

engineering. Throughout his career, he has coupled fundamental concepts with sound<br />

engineering practices to provide creative, innovative, and enduring solutions to challenges<br />

faced by water and wastewater utilities. Cushing has been responsible <strong>for</strong> numerous<br />

successful treatment facility planning and design projects, as well as studies and programs<br />

<strong>for</strong> improving distribution system water quality. Cushing has practiced nationally,<br />

providing service to a broad cross-section of the industry, from some of the largest and most<br />

visible utilities (e.g., New York City and Washington, D.C.) to very small applications with important and<br />

unique issues (e.g., Oray <strong>National</strong> Fish Hatchery, Utah). He has also been responsible <strong>for</strong> introducing and<br />

applying advanced technologies, most notably ultraviolet disinfection <strong>for</strong> potable water treatment. An<br />

internationally recognized expert in ultraviolet disinfection, he is responsible <strong>for</strong> seminal applied research in<br />

this area, and is project manager and a primary author <strong>for</strong> the United States Environmental Protection<br />

Agency’s Ultraviolet Disinfection Guidance Manual. Cushing received a B.S. in Petroleum Engineering and an<br />

M.S. and Ph.D. in Civil Engineering from the University of Texas at Austin.<br />

27


Session 2: Operations<br />

Evolution from a Conventional Well Field<br />

to a Riverbank-Filtration System<br />

John D. North<br />

Cedar Rapids <strong>Water</strong> Department<br />

Cedar Rapids, Iowa<br />

Douglas J. Schnoebelen, Ph.D.<br />

United States Geological Survey<br />

Iowa City, Iowa<br />

Shelli Grapp<br />

Cedar Rapids <strong>Water</strong> Department<br />

Cedar Rapids, Iowa<br />

Roy E. Hesemann<br />

Cedar Rapids <strong>Water</strong> Department<br />

Cedar Rapids, Iowa<br />

The City of Cedar Rapids is located in east-central Iowa and has a population of 121,000. The<br />

Cedar River alluvial aquifer is the sole source of drinking water <strong>for</strong> the City. Within the last few<br />

years, the Cedar River — which drains a major agricultural watershed — has experienced extended<br />

periods of elevated levels of nitrate and herbicides (notably, atrazine and its degradation<br />

byproducts) that were above the maximum contaminant level <strong>for</strong> drinking water. Fortunately,<br />

during these extended periods of compromised water quality in the river, <strong>RBF</strong> enabled the Cedar<br />

Rapids <strong>Water</strong> Department to obtain and treat adequate quantities of water that met all drinkingwater<br />

standards. The Cedar Rapids <strong>Water</strong> Department and the United States Geological Survey<br />

(USGS) have been collaborating in an ongoing study of the river and its hydraulic interconnection<br />

with the City’s wells. A primary focus of this study is to identify operational and development<br />

strategies that will optimize <strong>RBF</strong> and the water-quality benefits it af<strong>for</strong>ds.<br />

The Cedar River originates in southern Minnesota and generally flows in a southeasterly direction<br />

through east-central Iowa until it discharges to the Iowa River in Louisa County. The Cedar River<br />

drains a major agricultural watershed, which also includes several urban areas. The entire<br />

watershed encompasses 7,800 square miles, in which approximately 6,500 square miles are<br />

upstream from Cedar Rapids’ well fields. Land use in the watershed is predominantly agricultural<br />

(approximately 90 percent), with major crops being corn and soybeans (Kalkhoff et al., 2000;<br />

Schnoebelen and Schulmeyer, 1998; Schulmeyer, 1995). Major urban areas in the watershed<br />

include:<br />

• Albert Lea and Austin in Minnesota.<br />

• Mason City, Clear Lake, Charles City, Forest City, Waverly, Cedar Falls/<strong>Water</strong>loo, and<br />

Cedar Rapids/Marion in Iowa.<br />

Correspondence should be addressed to:<br />

John D. North<br />

<strong>Water</strong> Utility Director, Cedar Rapids <strong>Water</strong> Department<br />

1111 Shaver Road NE<br />

Cedar Rapids, Iowa 52402 USA<br />

Phone: (319) 286-5912 • Fax: (319) 286-5911 • Email: J.North@Cedar-Rapids.org<br />

29


30<br />

The Cedar River and its water quality have been the subject of extensive studies and monitoring<br />

by the Cedar Rapids <strong>Water</strong> Department, USGS, Iowa Geological Survey Bureau, and other local<br />

agencies. These studies have documented several major water-quality issues that affect the river’s<br />

suitability <strong>for</strong> the three designation uses established by the Iowa Department of Natural Resources:<br />

• Primary contact recreation (e.g., swimming).<br />

• Wildlife, fish, and aquatic life.<br />

• Potable water source.<br />

<strong>Water</strong>-quality challenges come from both point and non-point sources and include excessive soil<br />

erosion/sedimentation, elevated levels of nitrate and phosphorous, and fecal coli<strong>for</strong>m bacteria<br />

counts above the acceptable levels <strong>for</strong> recreational swimming. The 57-mile segment of the Cedar<br />

River upstream from the City’s well field to LaPorte City has been placed on the State’s list of<br />

impaired streams due to elevated levels of nitrate and fecal coli<strong>for</strong>m bacteria.<br />

The most daunting challenge <strong>for</strong> the Cedar Rapids <strong>Water</strong> Department is the increasing trend of<br />

elevated levels of nitrate, especially in the spring and early summer. During these periods, monthly<br />

nitrate levels are routinely in the range of 10.0 to 14.7 mg/L as nitrogen (N). The drinking-water<br />

standard <strong>for</strong> nitrate is 10.0 mg/L (as N), and a single exceedance constitutes a violation.<br />

Fortunately, as illustrated in Figure 1, a 2- to 3-mg/L reduction in nitrate levels is generally<br />

accomplished as the water moves from the river to the wells. This natural reduction, combined<br />

with the monitoring and management of individual wells, has enabled the Cedar Rapids <strong>Water</strong><br />

Department to avoid any violations of the drinking-water standard <strong>for</strong> nitrate.<br />

16.0<br />

14.0<br />

12.0<br />

10.0<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

0.0<br />

Jan-02<br />

Feb-02<br />

Mar-02<br />

Apr-02<br />

May-02<br />

Jun-02<br />

Monthly Nitrate Results (mg/L as N)<br />

Highest Reported Values<br />

Jul-02<br />

Aug-02<br />

Sep-02<br />

Oct-02<br />

Nov-02<br />

Dec-02<br />

Jan-03<br />

Feb-03<br />

Mar-03<br />

Apr-03<br />

May-03<br />

Jun-03<br />

Jul-03<br />

Cedar River<br />

NW <strong>Water</strong> Plant<br />

MCL Standard<br />

Figure 1. Highest recorded nitrate levels: Cedar River as compared to the well water supplies to Cedar<br />

Rapids’ NW <strong>Water</strong> Plant (Source: Cedar Rapids <strong>Water</strong> Department).<br />

Prior to 1963, the City of Cedar Rapids obtained its water supplies directly from the Cedar River.<br />

River water underwent treatment processes similar to those used today — lime softening,<br />

disinfection by chloramination, filtration, and the addition of fluoride and phosphate <strong>for</strong> corrosion<br />

control. This provided <strong>for</strong> safe drinking water, but failed to avoid the intermittent aesthetic<br />

problems (taste and odor) primarily associated with elevated algae levels in the river. This<br />

prompted the City to construct and convert its water supplies to a series of shallow wells along the<br />

Cedar River.


Today, Cedar Rapids obtains all of its water supplies from a series of shallow wells constructed in<br />

the sand and gravel alluvium along the Cedar River. The City now has four well fields consisting<br />

of 45 vertical wells and four horizontal collector wells. Total current production capacity of the<br />

wells is generally assumed to be approximately 65 MGD, but output will vary depending on river<br />

level and recent operating conditions. Withdrawal rates now average about 36 MGD, with a<br />

maximum demand of 50 MGD. Approximately 80 percent of the water is used <strong>for</strong> grain processing<br />

and other wet industries.<br />

Vertical wells are drilled to the top of the bedrock with depths ranging from 42 to 72 ft and are<br />

located about 30 to 900 ft from the river. The two original horizontal collector wells placed in<br />

service in 1995 consist of a 13-ft diameter center column extending to a depth of approximately<br />

60 ft (about 2 ft above bedrock) and six lateral screens extending about 200 ft from the central<br />

column. The two collector wells placed in service this year are of similar construction, except <strong>for</strong><br />

the use of a 16-ft diameter center column. All of the well laterals were constructed so as not to<br />

extend under the river channel except under high water conditions.<br />

The USGS and other agencies have conducted extensive studies of the Cedar River and the<br />

alluvium that serves as the drinking-water supply <strong>for</strong> Cedar Rapids. These studies have determined<br />

that the Cedar River is generally a “gaining stream” — that is, water will typically move through<br />

alluvium to the river; however, pumping of the wells will reverse the normal flow pattern and<br />

induce the infiltration of water from the Cedar River into the alluvium. The USGS found that<br />

induced filtration from the Cedar River supplies approximately 74 percent of the recharge to the<br />

City’s wells. The balance of the recharge is from the underlying aquifer (21 percent) and the<br />

infiltration of precipitation (5 percent). The USGS also determined that the travel times <strong>for</strong> the<br />

movement of water from the Cedar River to the nearest vertical well ranged from 7 to 17 days<br />

(Schulmeyer, 1995; Schulmeyer and Schnoebelen, 1998).<br />

The movement of water through sand and gravel <strong>for</strong>mations due to pumping is commonly called<br />

induced <strong>RBF</strong>. <strong>RBF</strong> has been widely used in Europe as a pretreatment process <strong>for</strong> almost 100 years,<br />

and there has been recent interest and research concerning its use in North America. The natural<br />

filtration and biological activity that occur during the movement of water through the alluvium<br />

af<strong>for</strong>ds several water-quality benefits. These include a reduction in turbidity and microbiological<br />

improvements with the removal/reduction of viruses, bacteria, and protozoan organisms<br />

(e.g., Giardia and Cryptosporidium), as well as a reduction in the levels of nitrates, herbicides, and<br />

other potential contaminants. USGS studies of Cedar Rapids’ wells have confirmed that, “The<br />

filtering efficiency of the aquifer is equivalent to a 3-log reduction rate or 99.99-percent reduction<br />

in particulates” (Schulmeyer, 1995).<br />

When the Cedar Rapids <strong>Water</strong> Department and USGS initiated their cooperative study in 1992, it<br />

was restricted to a 231-square mile area along the Cedar River that encompasses the City’s four well<br />

fields. Initially, the primary focus was to develop an understanding of the hydraulic interconnection<br />

of the river and wells and the factors that affect the quantity and quality of the water drawn from the<br />

wells. For example, studies regarding new well development would simply attempt to identify sites<br />

that would yield significant water with relatively low levels of iron and manganese. There was very<br />

little consideration given to “<strong>RBF</strong>” and enhancing the water-quality benefits it af<strong>for</strong>ds.<br />

Beginning about 1995, the scope of the study was expanded to include additional physical and<br />

chemical parameters, as well as the addition of biological monitoring (e.g., Microscopic Particulate<br />

Analysis). This was prompted by two primary considerations. The ongoing cooperative study<br />

confirmed that well water was consistently of higher quality than the river; however, the study also<br />

confirmed that the movement of recharge water through alluvium could also readily transport<br />

31


32<br />

contaminants from the river to the wells. Additionally, Iowa regulatory officials issued a preliminary<br />

determination that the newly constructed collector wells were groundwater under the direct<br />

influence of surface water (GWUDI) sources.<br />

To date, major study activities and work products of the USGS/Cedar Rapids <strong>Water</strong> Department<br />

cooperative study include:<br />

• Identification and mapping of contaminant sources in the immediate vicinity of the<br />

City’s wells.<br />

• Development of a regional groundwater model covering 231 square miles, which<br />

simulates groundwater flow under steady-state conditions.<br />

• Construction of a detailed groundwater flow model, which simulates groundwater flow<br />

and well recharge under transient conditions (a computer model is now complete and<br />

being tested).<br />

• Completion of extensive water-quality monitoring (physical and chemical parameters) of<br />

the river and wells.<br />

This research, along with the Cedar Rapids <strong>Water</strong> Department’s Source <strong>Water</strong> Assessment study,<br />

has documented the importance of the river and the potential contaminant threats it poses to our<br />

water supplies. They have also confirmed the protection and other benefits af<strong>for</strong>ded by <strong>RBF</strong>.<br />

Consequently, the cooperative study was recently expanded to include the entire Cedar River<br />

<strong>Water</strong>shed. The Cedar Rapids <strong>Water</strong> Department, USGS, and Iowa Geological Survey Bureau<br />

have partnered to conduct three separate comprehensive synoptic studies of water quality of<br />

samples at 64 locations throughout the watershed. A dye tracer/time-of-travel study has been<br />

completed on the lower main stem of the Cedar River, with plans <strong>for</strong> more dye tracing on other<br />

sections at low flow conditions, as well as possible Lagrangian sampling. In Lagrangian sampling,<br />

the same mass of water is sampled as it moves downstream. The objective of this study will be to<br />

validate time-of-travel models and to determine the fate of nitrogen compounds as they are<br />

transported down the river (D.J. Schnoebelen, 2002).<br />

In summary, the Cedar Rapids <strong>Water</strong> Department has attempted to develop a better understanding<br />

of the river and its wells, which would allow the implementation of management and<br />

protection programs <strong>for</strong> its drinking-water supplies. Although we did not fully understand its role<br />

until just recently, <strong>RBF</strong> is a valuable pretreatment process and is an integral component of our<br />

multi-barrier strategy <strong>for</strong> protecting water quality. It has enabled the Cedar Rapids <strong>Water</strong><br />

Department to avoid any violations of the drinking-water standard <strong>for</strong> nitrate that would<br />

necessitate the construction and operation of costly nitrate removal facilities.<br />

Without <strong>RBF</strong>, the Cedar Rapids <strong>Water</strong> Department would not have been able to supply quality<br />

water that meets all drinking-water standards while maintaining the competitive rates required by<br />

its major customers, the local industries. Although very beneficial, it must be acknowledged that<br />

<strong>RBF</strong> does not completely remove contaminants, as evidenced by the presence of nitrates and<br />

herbicides in Cedar Rapids’ well supplies. <strong>RBF</strong>, along with watershed protection programs, must<br />

be a key element of Cedar Rapids’ multi-barrier approach to the protection of its water supplies<br />

and public health.


REFERENCES AND ACKNOWLEDGEMENTS<br />

The following references and publications were used in the research and preparation of this report<br />

(though not all are cited in the text):<br />

<strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong><br />

• Ray, C., G. Melin, and R.B. Linsky, editors (2002). Riverbank Filtration: Improving Source-<strong>Water</strong><br />

Quality, Kluwer Academic Publishers, Dordrecht.<br />

United States Geological Survey<br />

• Becher, K.D., S.J. Kalkhoff, D.J. Schnoebelen, K.K. Barnes, and V.E. Miller (2001). <strong>Water</strong>-Quality<br />

Assessment of the Eastern Iowa Basins – Nitrogen, Phosphorous, Suspended Sediment and Organic<br />

Carbon in Surface <strong>Water</strong>, 1996-98, U.S. Geological Survey <strong>Water</strong>-Resources Investigations Report<br />

01-4175.<br />

• Kalkhoff, S.J., K.K. Barnes, K.D. Becher, M.E. Savoca, D.J. Schnoebelen, E.M. Sadorf, S.D. Porter,<br />

and D.J. Sullivan (2000). <strong>Water</strong> Quality in the Eastern Iowa Basins, Iowa and Minnesota, 1996-98,<br />

U.S. Geological Survey Circular 1210.<br />

• Schnoebelen, D.J. (2002). Written correspondence.<br />

• Schnoebelen, D.J., D.E. Christiansen, and K.D. Becher (2002). “City of Cedar Rapids: Source<br />

<strong>Water</strong> Assessment of Public Drinking-<strong>Water</strong> Supplied from Ground-<strong>Water</strong>.”<br />

• Schulmeyer, P.M. (1995). Effect of the Cedar River on the Quality of the Ground-<strong>Water</strong> Supply <strong>for</strong> City<br />

Rapids, Iowa, U.S. Geological Survey <strong>Water</strong>-Resources Investigations Report 94-4211.<br />

• Schulmeyer, P.M., and D.J. Schnoebelen (1998). Hydrogeology and <strong>Water</strong> Quality in the Cedar Rapids<br />

Area, Iowa, 1992-1996, U.S. Geological Survey <strong>Water</strong>-Resources Investigations Report 97-4261.<br />

Iowa Department of Natural Resources<br />

• “Progress Report One — Cedar River Assessment Survey,” June 2001.<br />

• “Source <strong>Water</strong> Assessment and Protection Program and Implementation Strategy <strong>for</strong> the State of<br />

Iowa,” March 2000.<br />

Iowa Department of Natural Resources – Geological Survey Bureau<br />

• Seigely, L.S., and M.P. Skopec. Written reports of three comprehensive synoptic monitoring programs<br />

conducted at 62 sites on the Cedar River on August 26, 2000; June 2, 2001; and June 7, 2003.<br />

Cedar Rapids <strong>Water</strong> Department<br />

• Grapp, S.J., J.D. North, and R.E. Hesemann (2002). “City of Cedar Rapids Source <strong>Water</strong> Assessment.”<br />

JOHN NORTH has 30 years of experience in water-quality monitoring and in the<br />

operation and maintenance of water systems. For the past 11 years, he has worked <strong>for</strong> the<br />

City of Cedar Rapids in Iowa, which draws its water supplies from a series of shallow<br />

alluvial wells located along the Cedar River and drains a major agricultural watershed. At<br />

present, he is the <strong>Water</strong> Utility Director <strong>for</strong> the City of Cedar Rapids <strong>Water</strong> Department.<br />

Under his direction, the Cedar Rapids <strong>Water</strong> Department has conducted several<br />

comprehensive water-quality research projects and other studies designed to ensure the<br />

continued availability of an ample supply of safe drinking water <strong>for</strong> Cedar Rapids. Since 1993, the utility has<br />

partnered with the United States Geological Survey in a comprehensive ongoing study of the City’s well<br />

fields and their hydraulic interconnection with the Cedar River. A major focus of the study has been to<br />

optimize the benefits of <strong>RBF</strong>. North received a B.S. in Chemistry from Loras College. He has dualcertification<br />

as a Grade IV water and wastewater plant operator.<br />

33


Session 3: Hydraulic Aspects<br />

Groundwater Flow and <strong>Water</strong> Quality – A Flowpath<br />

Study in the Seminole Well Field, Cedar Rapids, Iowa<br />

Douglas J. Schnoebelen, Ph.D.<br />

United States Geological Survey<br />

Iowa City, Iowa<br />

Michael J. Turco<br />

United States Geological Survey<br />

Lincoln, Nebraska<br />

John D. North<br />

Cedar Rapids <strong>Water</strong> Department<br />

Cedar Rapids, Iowa<br />

In Iowa, alluvial aquifers near major rivers are a source of water <strong>for</strong> many communities. The City<br />

of Cedar Rapids withdraws water from wells completed in the Cedar River alluvium, a shallow<br />

alluvial aquifer adjacent to the Cedar River. The City of Cedar Rapids is located within Linn County<br />

in east-central Iowa, and water <strong>for</strong> the City is supplied by four well fields (East, Northwest,<br />

Seminole, and West well fields) along the Cedar River. The City has a population of about 121,000,<br />

and several large industries are major water users. Currently, per capita water usage in the City is<br />

nearly three times the national average. The City is committed to providing both a high quality<br />

and quantity of water to its customers. The USGS and Cedar Rapids <strong>Water</strong> Department have<br />

been working together in an ongoing research program to better understand water quality and flow<br />

in the Cedar River and alluvial well fields. Work has been done on both a basin and well-field<br />

approach and has involved dye tracing/time-of-travel studies on the Cedar River, water-quality<br />

sampling, geochemical modeling, and groundwater-flow modeling.<br />

The effect of land use in the Cedar River Basin on both surface-water and groundwater quality is<br />

an important issue. The Cedar River Basin upstream from Cedar Rapids is approximately<br />

6,500 square miles. Upstream land use in the Cedar River Basin is over 90-percent agriculture.<br />

Corn and soybeans are the major crops. Livestock raised in the area include beef and dairy cattle,<br />

as well as hogs. Runoff from agriculture is of concern, particularly during the spring and early<br />

summer when many chemicals are applied to cropland. Triazine and acetanilide herbicides are<br />

commonly applied in the Cedar River Basin, and these herbicides are water soluble and can be<br />

transported to streams and infiltrate to groundwater. In addition, several studies in eastern Iowa<br />

have identified nutrients as a major contaminant that has impaired water quality (Goolsby and<br />

Battaglin, 1993; Hallberg et al., 1996; Schnoebelen et al., 1999; Kalkhoff et al., 2000). In general,<br />

the majority of nitrogen inputs in the Cedar River Basin are from chemical fertilizers and animal<br />

manure (Becher et al., 2000). High nitrate levels (greater than 10.0 mg/L) in the Cedar River are<br />

of particular concern to municipal water operators. The lower Cedar River is listed on the Iowa<br />

total maximum daily load list <strong>for</strong> nitrate upstream of Cedar Rapids, Iowa. The Cedar River is the<br />

Correspondence should be addressed to:<br />

Douglas J. Schnoebelen, Ph.D.<br />

<strong>Research</strong> Hydrologist/<strong>Water</strong>-Quality Specialist<br />

United States Geological Survey<br />

Federal Building, Room 269 • 400 South Clinton • Iowa City, Iowa 52240 USA<br />

Phone: (319) 358-3617 • Fax: (319) 358-3606 • Email: djschnoe@usgs.gov<br />

35


36<br />

source of most nitrate detected in the Cedar River alluvial aquifer because of induced infiltration<br />

from the river due to pumping (Schulmeyer and Schnoebelen, 1998; Boyd, 2000).<br />

An unconsolidated surficial layer of glacial till, loess, and Cedar River alluvium (alluvial aquifer)<br />

overlies carbonate bedrock of Devonian and Silurian age (bedrock aquifer) in the study area. The<br />

alluvial aquifer typically consists of a sequence of coarse sand and gravel at the base, grading<br />

upwards to finer sand, silt, and clay near the surface. The sand and gravel contain carbonate, shale,<br />

and ferro-magnesium rich rock fragments. The thickness of the alluvial aquifer ranges from about<br />

2 to 30 m. The alluvial aquifer is recharged by the infiltration of water from the Cedar River,<br />

precipitation, and seepage from the underlying bedrock and adjacent hydrogeologic units. In areas<br />

under the influence of municipal pumping, groundwater flow is from the Cedar River toward the<br />

well fields; in areas outside the influence of municipal pumping, groundwater flow is toward the<br />

Cedar River. Results from a regional groundwater flow model indicated that approximately<br />

74 percent of the water pumped from the alluvial well fields is recharged from the Cedar River,<br />

approximately 21 percent is recharged from adjacent underlying hydrogeologic units, and approximately<br />

5 percent of the water is from infiltrating precipitation (Schulmeyer and Schnoebelen,<br />

1998). Currently, a more detailed groundwater model in the study area indicates that, in some<br />

places, up to 90 percent of the water pumped from the alluvial well fields is recharged from the<br />

Cedar River.<br />

The water quality in the alluvial aquifer within the well field has been characterized with samples<br />

collected from both monitoring and municipal wells at various times since 1992 (Boyd, 2000;<br />

Schulmeyer and Schnoebelen, 1998; Schnoebelen and Schulmeyer, 1996). Calcium, magnesium,<br />

and bicarbonate are the dominant ions. In addition, nitrate, sulfate, silica, iron, and manganese are<br />

present in significant concentrations in certain wells or at certain times of the year. Previous work in<br />

the Seminole well field indicated some detections of herbicides and their degradates (breakdown<br />

products) in shallow monitoring wells (3.8- to 6-m deep) completed in the alluvium as water moved<br />

from the river into the alluvial aquifer (Boyd, 2000). Atrazine was the most commonly detected<br />

herbicide in this study. Acetochlor, cyanazine, and metolachlor were also detected, but at smaller<br />

concentrations than atrazine. Acetanilide degradates were detected at greater frequencies and at<br />

greater concentrations than their corresponding parent compounds. Fewer numbers of detections of<br />

herbicide compounds were found in wells completed deeper in the alluvium.<br />

The infiltration of water with large nitrate concentrations into the alluvial aquifer from the Cedar<br />

River affects groundwater quality. Recent research was conducted along a flowpath to study <strong>RBF</strong><br />

through a natural wetland area. Groundwater modeling helped locate the flowpath study. The<br />

study examined the role of a natural wetland in reducing nitrate concentrations as water moves<br />

from the Cedar River. A real challenge <strong>for</strong> the Cedar Rapids <strong>Water</strong> Department is the increasing<br />

trend of nitrate concentrations in the Cedar River. Nitrate concentrations in the Cedar River<br />

during the spring are often more than 10 mg/L and can reach 20 mg/L. A 2- to 3-mg/L reduction<br />

in nitrate often occurs as water moves from the river to the well, but in some wells, this may not<br />

reduce nitrate concentrations below the 10.0-mg/L maximum contaminant level. Sampling in<br />

wells along a flowpath occurred quarterly over a period of about 4 years. A comparison of water<br />

chemistry was made from water analyses from:<br />

• The river.<br />

• A monitoring well upgradient of the wetland area and river.<br />

• Wells in the wetland area.<br />

• Wells between the wetland area and river.


In addition, a comparison of water-chemistry data from a municipal well located near the wetland<br />

area and one located nearest the river were compared in terms of water chemistry from previous<br />

sampling work (Schulmeyer and Schnoebelen, 1998). Results show that nitrate concentrations<br />

were 4 to 6 times lower in samples from monitoring wells completed in the wetland area than in<br />

the Cedar River or groundwater in the upland area; however, iron and manganese concentrations<br />

in samples from the monitoring wells in the wetland areas were an order of magnitude higher<br />

when compared to the river or upland well. <strong>Water</strong> samples from the wells and the Cedar River<br />

generally displayed similar trends (high in the spring and low in the fall), while iron and manganese<br />

concentrations were more variable.<br />

As water moves from the river towards the monitoring wells, microorganisms obtain energy <strong>for</strong><br />

metabolic processes by catalyzing the oxidation of organic matter with a progressive series of<br />

reducing reactions (Stumm and Morgan, 1981). Nitrate can be reduced to elemental nitrogen (N 2)<br />

by denitrification (Equation 1) or to ammonium (NH 4+) by reduction (Equation 2). Since<br />

ammonium was only detected in small quantities (less than 0.80 mg/L), denitrification most likely<br />

is the predominant process.<br />

4NO 3 – + 5CH2O + 4H + = 2N 2(g) + 5CO 2 + 7H 2O (Equation 1)<br />

NO 3 – + 2CH2O + 2H + = NH 4 + + H2O + 2CO 2<br />

(Equation 2)<br />

Reduction then proceeds from nitrate (NO 3 – ) to Mn +4 , Fe +3 , SO4 –2 , CO2, and N 2. The reduced<br />

<strong>for</strong>ms of iron (Fe II) and manganese (Mn II) are more soluble in water and are more mobile than<br />

oxidized <strong>for</strong>ms (Hem, 1985) and, under anoxic conditions, are stable. As nitrate in groundwater<br />

is depleted, iron and manganese reduction begins. The reduction of Fe +3 to Fe +2 and Mn +4 to<br />

Mn +2 from aquifer grain coatings can cause large concentrations of these ions in groundwater.<br />

Ferrihydrite and manganite (MnOOH) occurring as oxyhydroxide coatings on clay and silt<br />

particles are the most likely oxidized <strong>for</strong>ms of iron (Fe +3 ) and manganese (Mn +3 and Mn +4 ) in the<br />

alluvial aquifer. Oxidized <strong>for</strong>ms of iron and manganese might occur in the aquifer as crystalline<br />

minerals, such as hematite (Fe 2O 3) and hausmannite (Mn 3O 4). Iron and manganese may<br />

co-precipitate with carbonate minerals to cause well fouling.<br />

<strong>Research</strong> in the Seminole well field indicates that the location of a well in or near natural wetland<br />

areas may benefit from the natural reduction of nitrate concentrations, with the disadvantage of<br />

increased iron and manganese concentrations. Future expansions of the well fields may take<br />

advantage of natural wetland areas to help reduce nitrate concentrations. In Iowa, most wetlands<br />

have been drained, but alluvial wetlands associated with bottomland <strong>for</strong>ested and oxbow lake<br />

areas may persist as they are subject to periodic flooding and are often not suitable <strong>for</strong> sustained<br />

agriculture.<br />

REFERENCES<br />

Becher, K.D., D.J. Schnoebelen, and K.K.B. Akers (2000). “Nutrients discharged to the Mississippi River<br />

from Eastern Iowa <strong>Water</strong>sheds, 1996-97.” Journal AWWA, 36(1): 161-173.<br />

Boyd, R.A. (2000). “Herbicides and herbicide degradates in shallow groundwater and the Cedar River near<br />

a municipal well field, Cedar Rapids, Iowa.” The Science of the Total Environment, 241-253.<br />

Goolsby, D.A., and W.A. Battaglin (1993). “Occurrence, distribution, and transport of agricultural chemicals<br />

in surface water of the Midwestern United States.” Selected papers on agricultural chemicals in water resources of<br />

the midecontinental United States, D.A. Goolsby, L.L. Boyer, and G.E. Mallard, compilers, U.S. Geological<br />

Survey Open-File Report 93-418, p. 1-25.<br />

37


38<br />

Hallberg, G.R., D.G. Riley, J.R. Kantamneni, P.J. Weyer, and R.D. Kelley (1996). Assessment of Iowa safe<br />

drinking water act monitoring data, 1988-1995, Iowa City, University of Iowa Hygienic Laboratory <strong>Research</strong><br />

Report 97-1, 132 p.<br />

Hem, J.D. (1985). Study and interpretation of the chemical characteristics of natural water, third edition, U.S.<br />

Geological Survey <strong>Water</strong>-Supply Paper 2254, p. 264.<br />

Kalkhoff, S.J., K.K. Barnes, K.D. Becher, M.E. Savoca, D.J. Schnoebelen, E.M. Sadorf, S.D. Porter, and D.J.<br />

Sullivan (2000). <strong>Water</strong> Quality in the Eastern Iowa Basins, Iowa and Minnesota, 1996-98, U.S. Geological<br />

Survey Circular 1210, 37 p.<br />

Schnoebelen, D.J., K.D. Becher, M.W. Bobier, and T. Wilton (1999). Selected nutrients and pesticides in streams<br />

of the Eastern Iowa Basins, 1970-95, U.S. Geological Survey <strong>Water</strong>-Resources Investigations Report 99-4028,<br />

105 p.<br />

Schnoebelen, D.J., and P.M. Schulmeyer (1996). Selected hydrogeologic data in the Cedar Rapids area, Benton<br />

and Linn Counties, Iowa, October 1992 through March 1996, U.S. Geological Survey Open-File Report<br />

96-471, 172 p.<br />

Schulmeyer, P.M., and D.J. Schnoebelen (1998). Hydrogeology and water-quality in the Cedar Rapids Area,<br />

Iowa, 1992-96, U.S. Geological Survey <strong>Water</strong> Resources Investigations Report 97-4261, 77 p.<br />

Stumm W., and J.J. Morgan (1981). Aquatic Chemistry — An introduction emphasizing chemical equilbria in<br />

natural waters, second edition, Wiley-Interscience Publisers, New York, 780 p.<br />

DOUG SCHNOEBELEN is a <strong>Research</strong> Hydrologist with the United States Geological<br />

Survey and has worked on a variety of groundwater and surface-water projects over the last<br />

13 years. He has served as the groundwater specialist on the White River <strong>National</strong> <strong>Water</strong><br />

Quality Assessment project in Indiana and as the surface-water specialist <strong>for</strong> the Eastern<br />

Iowa Basins <strong>National</strong> <strong>Water</strong> Quality Assessment project in Iowa. In addition, he has been<br />

the Iowa District <strong>Water</strong>-Quality Specialist since 1994, and is an Adjunct Professor in the<br />

Geoscience Department at the University of Iowa. His research has focused on the use of<br />

isotopes and bore hole geophysics in groundwater and, more recently, on the fate and transport of agricultural<br />

chemicals in surface water and groundwater across eastern Iowa. In particular, he has focused on the fate and<br />

transport of several pesticide degradate compounds. He has been involved with ongoing research in riverbank<br />

filtration and geochemical modeling in the Cedar Rapids well field since 1999. Schnoebelen received a B.S.<br />

in Geology from the University of Iowa, an M.S. in Geology from the University of Tennessee, and a Ph.D.<br />

in Geology, with a minor in Environmental Science, from Indiana University.


Session 3: Hydraulic Aspects<br />

The Use of Aquifer Testing and Groundwater Modeling<br />

to Evaluate Changes in Aquifer/River Hydraulics at the<br />

Louisville <strong>Water</strong> Company<br />

David C. Schafer<br />

David Schafer & Associates<br />

Stillwater, Minnesota<br />

In 1999, the Louisville <strong>Water</strong> Company in Louisville, Kentucky, constructed a 20-MGD radial<br />

collector well as part of a pilot study to evaluate using <strong>RBF</strong> to reduce treatment costs and to ensure<br />

high-quality water production. The water company currently withdraws up to a total of 240 MGD<br />

of surface water from the Ohio River at the Zorn Pumping Station and B.E. Payne <strong>Water</strong> Treatment<br />

Plant (both in Louisville, Kentucky), and is exploring ways to further improve water-supply<br />

quality.<br />

A primary objective of the pilot project was to identify processes that would meet or exceed<br />

expected water-quality regulations. Considerations included:<br />

• Removing synthetic organics.<br />

• Reducing turbidity.<br />

• Eliminating taste and odor during summer months.<br />

• Attenuating water temperature extremes in the distribution system.<br />

Although several options were evaluated (such as GAC and membrane filtration), <strong>RBF</strong> was the<br />

only one that addressed all of the Louisville <strong>Water</strong> Company’s concerns.<br />

In addition to evaluating water quality, a secondary objective was to assess aquifer/river hydraulics<br />

and monitor the hydraulic per<strong>for</strong>mance over time to detect whether or not changes in capacity<br />

occurred. Although it was not expected that the well structure would lose hydraulic efficiency,<br />

there was some concern that riverbed materials adjacent to the well could clog or compact in<br />

response to groundwater withdrawal.<br />

The collector well was constructed about 120 ft from the shore of the Ohio River at the B.E. Payne<br />

<strong>Water</strong> Treatment Plant site. It incorporated a 16-ft inside diameter concrete caisson extending to<br />

a depth of about 105-ft below land surface. The caisson penetrated the glacial sand and gravel<br />

aquifer at the site, bottoming out on the underlying shale and limestone bedrock. Seven 12-inch<br />

diameter stainless steel horizontal well-screen laterals were installed near the base of the sand and<br />

gravel aquifer — four laterals 240 ft in length and three 200 ft in length. The four longer laterals<br />

were oriented toward the Ohio River, while the other three shorter laterals extended 1) upriver,<br />

2) downriver, and 3) away from the river.<br />

Correspondence should be addressed to:<br />

David C. Schafer<br />

President<br />

David Schafer & Associates<br />

9955 North 101st Street • Stillwater, Minnesota 55082 USA<br />

Phone: (651) 762-8281 • Fax: (651) 762-8335 • Email: dschafer@prodigy.net<br />

39


40<br />

Two 10-MGD capacity pumps were installed to produce water from the collector well. A variablefrequency<br />

driver was used to power one of the two pump motors to allow the pumping plant to<br />

readily adapt to changing flow and head requirements.<br />

In addition to the collector well, about a dozen on-land piezometers were installed around the<br />

collector well at distances up to 1,200-ft away. Also, a nested piezometer was installed about<br />

80-ft offshore, adjacent to the well site. The piezometers were used to monitor aquifer water levels<br />

during the testing and operation of the collector well to provide the data needed to quantify<br />

aquifer characteristics.<br />

Following well construction, a 70-day constant-rate pumping test was conducted at a pumping rate<br />

of 19.4 MGD. <strong>Water</strong> levels were monitored in both the collector well and observation wells. The<br />

maximum drawdown observed in the collector well during the test was about 27 ft. The pumping<br />

test data were used to quantify hydraulic properties — transmissivity and river leakance — and to<br />

identify baseline per<strong>for</strong>mance characteristics.<br />

The data were analyzed by constructing a groundwater flow model of the site and adjusting model<br />

input parameters until the model faithfully replicated water levels observed during the pumping<br />

test. The USGS groundwater flow code, MODFLOW, was selected <strong>for</strong> the modeling study. The<br />

aquifer and Ohio River were represented in a model grid covering an area 4,000 ft × 14,000 ft. The<br />

model domain was subdivided into 119 rows, 53 columns, and 8 model layers. The use of a large<br />

number of model layers permitted the realistic simulation of vertical resistance to groundwater<br />

flow — both water exiting the river and entering collector well laterals. Results of the pumping<br />

test and modeling ef<strong>for</strong>t revealed an aquifer transmissivity of 204,000 gallons per day per foot and<br />

a river leakance of 2.35 inverse days at the prevailing groundwater and surface-water temperatures.<br />

The collector well was used as the pumping well in five subsequent pumping tests conducted over<br />

a 3-year period from 1999 to 2002. Data from each pumping test were analyzed using groundwater<br />

modeling to update the computed values of transmissivity and leakance. The prevailing groundwater<br />

and surface-water temperatures were different <strong>for</strong> each test that was conducted. Because water<br />

temperature affects flow characteristics, it was necessary to correct the results of each test <strong>for</strong><br />

temperature effects so that per<strong>for</strong>mance could be compared from one test to another.<br />

In the model calculations, a simplifying assumption was made that leakance was constant<br />

everywhere <strong>for</strong> any given model run. This simplification allowed a relative comparison of one test<br />

with another. It is known that leakance declines in a non-uni<strong>for</strong>m manner, with greater reduction<br />

near the pumped well and less reduction away from the well; however, the exact nature of<br />

leakance distribution is not known, and trying to estimate it with the available data would have<br />

been speculative. There<strong>for</strong>e, the assumption of a single leakance value was made. The resulting<br />

leakance value can be thought of as an “effective leakance” <strong>for</strong> comparison purposes.<br />

Throughout the period of testing, transmissivity remained constant while effective leakance<br />

declined steadily. Over time, the rate of leakance decline diminished and finally stabilized based on<br />

the last pumping test conducted in July 2002, after nearly 3 years of well operation. Measured<br />

leakance values, corrected <strong>for</strong> water temperature, are listed in Table 1.<br />

River scour associated with a bank full-flood event in Spring 2002 may have contributed to the<br />

apparent stabilization (actually, a slight increase in leakance). The average measured leakance in<br />

July 2002, corrected to the same water temperature as the original pumping test, was about<br />

0.18 inverse days, more than an order of magnitude lower than the original leakance value.


Table 1. Measured Leakance Values<br />

Date of Test Effective Leakance<br />

Data Acquisition (inverse days)<br />

October 1999 2.35<br />

March 2000 0.72<br />

October 2000 0.25<br />

April 2001 0.20<br />

September 2001 0.15<br />

July 2002 0.18<br />

The corresponding well capacity had declined by about a third in response to reduced river<br />

leakance.<br />

The reduction in river leakance is being evaluated on an ongoing basis, but is believed to be<br />

caused by clogging/compaction of riverbed sediments in response to collector well operation.<br />

Regular pumping tests and leakance evaluation will be continued in the future to further monitor<br />

the effects of riverbed clogging and periodic river scour associated with flood events on the Ohio<br />

River. The Louisville <strong>Water</strong> Company continues to do research in the area of riverbed clogging<br />

and leakance reduction.<br />

Hydrologist DAVID SCHAFER has over 30 years of experience in the groundwater industry.<br />

He has designed hundreds of high-capacity water supply wells, has analyzed hundreds of<br />

pumping tests, and is consulted regularly on well development and rehabilitation procedures.<br />

In addition, he has done extensive groundwater modeling using numerical models, analytic<br />

element models, and proprietary analytical models that he has developed. At present, he is<br />

President of his own company, David Schafer & Associates, which he founded in 1999.<br />

Schafer has published numerous articles on dewatering, well hydraulics, well development,<br />

and well rehabilitation, and contributed a portion of the well hydraulics section of Groundwater and Wells,<br />

Second Edition. Schafer received both a B.S. in Mathematics and an M.S. in Computer Science from the<br />

University of Minnesota.<br />

41


Session 3: Hydraulic Aspects<br />

Plugging in Riverbank-Filtration Systems:<br />

Evaluating Yield-Limiting Factors<br />

Stephen A. Hubbs, P.E.<br />

Louisville <strong>Water</strong> Company<br />

Louisville, Kentucky<br />

The <strong>RBF</strong> process involves the passage of river water containing dissolved and suspended solids<br />

through the sands and gravels of a connected aquifer. In this process, the removal of suspended solids<br />

involves dynamics that are similar to slow sand filtration, with both biological filtration and physical<br />

filtration acting to remove particles. If the entrained particles are not removed, the hydraulic<br />

conductivity of the riverbed will decrease due to particles plugging the pores of the aquifer.<br />

Work done at the Louisville <strong>Water</strong> Company in Louisville, Kentucky, and elsewhere have indicated<br />

that the bulk of particle removal takes place in the first few feet of the aquifer, which is consistent<br />

with behavior in a slow sand filter. The dynamics of how a riverbed restores its filtration capacity<br />

are not fully understood. Anecdotal in<strong>for</strong>mation from those operating large-capacity well fields<br />

indicates that these well fields require 3 to 5 years to “settle in” to their sustainable yield, which<br />

is usually within 50 to 75 percent of the initial capacity of the well field. Recent research at the<br />

Louisville <strong>Water</strong> Company has focused on developing a better understanding of the dynamics of<br />

riverbed plugging and predicting sustainable capacity from high-capacity <strong>RBF</strong> systems.<br />

Predicting Sustainable Yield<br />

The 20-MGD horizontal collector well at Louisville was constructed in 1999, and data on various<br />

parameters have been collected since that time. These data have been analyzed by traditional<br />

modeling techniques (Schafer, 2003) and, more recently, by regression techniques, with the<br />

specific interest of gaining more in<strong>for</strong>mation about the process of riverbed plugging.<br />

Data indicate a cyclic pattern in specific yield, assumed to be a function of water viscosity as<br />

impacted by water temperature. Data also indicate a decreasing trend with time <strong>for</strong> specific yield.<br />

The possible causes of this decrease include:<br />

• Plugging of the riverbed-aquifer interface.<br />

• Decreased conductance of the bulk of the aquifer.<br />

• Decreased conductance in the area of the well screen.<br />

Various parameters from the entire 4-year database at the Louisville <strong>Water</strong> Company were<br />

subjected to regression analysis, with specific yield as the dependent variable. Variables included:<br />

• River stage.<br />

• <strong>Water</strong> elevation in the collector caisson.<br />

Correspondence should be addressed to:<br />

Stephen A. Hubbs, P.E.<br />

Vice President, New Technology<br />

Louisville <strong>Water</strong> Company<br />

550 South Third Street • Louisville, Kentucky 40202 USA<br />

Phone: (502) 569-3675 • Fax: (502) 569-0813 • Email: SHubbs@lwcky.com<br />

43


44<br />

• River and caisson water temperature.<br />

• Time (linear), the square root of time, and natural log of time.<br />

The influence of the driving head from the river was factored into the specific yield value by<br />

subtracting river stage from water level in the caisson. This eliminated the need to carry a term<br />

<strong>for</strong> river level in the analysis. Thus, the term <strong>for</strong> specific yield was defined from the data as:<br />

(Well output in MGD)/(River stage – Caisson water level).<br />

The square root of time was selected as a parameter based on the hypothesis that:<br />

• Plugging of the aquifer was the result of the deposition of suspended sediment at the<br />

aquifer-riverbed interface.<br />

• The rate of this plugging process would decrease as the recharge area extended farther<br />

into the river.<br />

• A balance between plugging and riverbed scouring would eventually be established.<br />

Subsequent analyses indicated that a better regression fit was realized when the natural log of time<br />

was used as a parameter, as opposed to linear time or the square root of time.<br />

When the entire data set was analyzed, it was apparent that typical trends were skewed during<br />

times of flooding. The database was modified to exclude flooding periods and then re-analyzed.<br />

Results of this analysis are presented in Figure 1. The regression model provides a good fit with<br />

actual data, but varies the most at the initiation of pumping and during the period of Weeks 160<br />

to 180, following a significant flooding event.<br />

Specific Yield (normalized to mean of 0.491 MGD/ft)<br />

0.500<br />

0.400<br />

0.300<br />

0.200<br />

0.100<br />

0.000<br />

-0.100<br />

-0.200<br />

-0.300<br />

Ln = Natural log.<br />

Figure 1. Regression with Natural Log of Time (Cleaned Data)<br />

SpYield (MGD/ft) = (-0.013) + .0045 (Well Temp. F) + .0018 (River Temp F) - .095 (Ln (Time week))<br />

0 20 40 60 80 100<br />

Period (week)<br />

120 140 160 180 200<br />

Actual Data Predicted Data<br />

R = 0.96<br />

F = 583<br />

Specific Yield immediate following<br />

flood event is greater than predicted<br />

Period of flooding<br />

It is suggested that the parameter of natural log of time in this analysis is a good surrogate <strong>for</strong> the<br />

measure of plugging in the system. This analysis was extended to include an extrapolation of an<br />

additional 5 years of data, with the insertion of “jumps” representing the impact of periodic


iverbed scouring. This analysis shows the system stabilizing after 8 years of operation, with a<br />

specific yield cycling between 0.25 and 0.4 MGD/ft drawdown. This type of extrapolation is highly<br />

speculative, but fits anecdotal in<strong>for</strong>mation on sustainable yields from high-capacity <strong>RBF</strong> systems.<br />

Riverbed Shear Stress as an Indicator of Riverbed Scouring<br />

Riverbed scouring plays an essential role in determining the sustainable yield in <strong>RBF</strong> systems. The<br />

shear stresses exerted on the riverbed during periods of high flow provide <strong>for</strong> the transport of<br />

riverbed materials and the re-suspension and transport of particles trapped during the <strong>RBF</strong> process.<br />

The extent of riverbed scouring can be estimated as a function of riverbed shear stress exerted<br />

during high-flow events.<br />

Streams typically exert higher riverbed shear stresses near headwaters, with decreasing stresses<br />

exerted near the terminus of a large water body (such as an ocean). This implies that riverbed<br />

scouring will decrease near the terminus of a stream, consistent with the <strong>for</strong>mation of river deltas.<br />

Because of this tendency to deposit fine materials near the mouth of streams, these locations are<br />

typically not well suited <strong>for</strong> high-capacity <strong>RBF</strong> systems.<br />

Two techniques <strong>for</strong> estimating riverbed shear stresses were used in this analysis:<br />

• Measured velocity profiles across the river cross-section.<br />

• Stream-surface slope measurements.<br />

The shear stresses exerted on a streambed can also be inferred by looking at the riverbed material in<br />

transport during high-flow events, as indicated by the particle-size distribution of riverbed sediments.<br />

A compilation of particle-size data from the Ohio River in the Louisville area is presented in<br />

Figure 2. Data to the left represent suspended solids, while data to the center and to the right<br />

indicate bed material. The river at the point where these data were collected is on a bend, with<br />

Percent Passing<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

Figure 2. Particle Size Distribution Analysis – Riverbed and Suspended Solids<br />

(USGS Data, 1979–1982, Ohio River at Louisville)<br />

Flow= 239,000 cfs<br />

TSS=408-466 mg/l<br />

Suspended Solids Indiana side<br />

Riverbed<br />

Flow = 301,000 to 434,000 cfs<br />

TSS = 482 to 698 mg/l<br />

Main channel and Kentucky side Riverbed<br />

0<br />

0.001 0.01 0.1 1 10 100<br />

Particle Size<br />

SS -11/30/1979<br />

SS -6/10/1981<br />

SS-1/27/1982<br />

bed-1375' from KY bank<br />

bed-1650' from KY bank<br />

bed-1850' from KY bank<br />

bed-2000' from KY bank<br />

bed-2200' from KY bank<br />

bed-2500' from KY bank<br />

bed-2700' from KY bank<br />

bed-2900' from KY bank<br />

bed-3900' from KY bank<br />

bed-4120' from KY bank<br />

SS-6/2/82 mid-stream<br />

SS-6/2/82 Ind side<br />

TSS = Total suspended solids. cfs = Cubic feet per second. SS = Suspend solids. KY = Kentucky. Ind = Indiana.<br />

45


46<br />

the Kentucky side of the river on the outside of the bend. The prevalence of courser material in<br />

the riverbed on the outside of the bend reflects higher riverbed shear stress exerted at the outside<br />

face of the bend.<br />

These data indicate that the shear stresses exerted against the riverbed are capable of transporting<br />

riverbed media in the size range of 1 to 10 millimeters. According to available literature, this<br />

would require a streambed shear stress in the range of 5 Newtons per square meter.<br />

Stream velocity profiles and stream surface slopes of the Ohio River at high flow conditions<br />

(383,000 cubic feet per second) were also secured from USGS. Velocity profile data were reported<br />

down to about 4 ft, the approximate limit of the acoustic doppler current profiler technology used<br />

<strong>for</strong> this data collection ef<strong>for</strong>t.<br />

The velocity profiles used in this study <strong>for</strong> calculating boundary shear stress would be expected to<br />

reflect a maximum shear stress <strong>for</strong> any point in the river, as the velocity profile selected was the<br />

one with the highest velocity in the cross-section. The slope calculation technique, on the other<br />

hand, would be expected to reflect an average boundary shear stress across both the cross-section<br />

and length of river selected <strong>for</strong> calculation. Comparable data sets <strong>for</strong> these two methods of<br />

calculation revealed that the velocity profile (using the manipulation yielding the greatest shear<br />

velocity) provided a lower and less consistent estimate of boundary shear stress than did the<br />

method based on measured stream slope. These data are summarized in Table 1 from a 2000 USGS<br />

database:<br />

Table 1. Comparison of Shear Stress Calculation Methods at Three Locations<br />

in the Ohio River Near Louisville<br />

Location Profile-Calculated Shear Stress Slope-Calculated Shear Stress<br />

(N/m 2 ) (N/m 2 )<br />

12 Mile Island 1.63 3.21<br />

BEPWTP 1.68 5.33<br />

Zorn Avenue 5.79 5.62<br />

BEPWTP = B.E. Payne <strong>Water</strong> Treatment Plant. N/m 2 = Newtons per square meter.<br />

The inconsistency and lower results from the stream profile calculations may be the results of the<br />

higher degree of variability obtained from the acoustic doppler measurement technique <strong>for</strong><br />

velocity and, particularly, the inability <strong>for</strong> this technique to measure velocity at the lower portions<br />

of the stream. It is also noted that video images of the riverbed indicated surface irregularities (or<br />

“pocks” on the river bottom). These depressions would be expected to change the velocity profile<br />

as compared to a theoretical surface where friction is exerted by media roughness only.<br />

The slope-calculated shear stress values, on the other hand, were consistent and logical <strong>for</strong> all of<br />

the measured values. The higher stream flows were associated with higher shear stresses, and shear<br />

stresses tended to be greater where the river was deeper and narrower.<br />

The data from the Ohio River were compared to data from the Rhine River. The maximum<br />

boundary shear stress observed on the Ohio River during the high-flow event was<br />

7.5 Newtons per square meter. This compares to the average shear stress of the Rhine River of<br />

10 Newtons per square meter, as reported by Schubert (2002).


REFERENCES<br />

Schafer, D.C. (2003). “The Use of Aquifer Testing and Groundwater Modeling to Evaluate Changes in<br />

Aquifer/River Hydraulics at the Louisville <strong>Water</strong> Company.” Program and Abstracts, Second International<br />

Riverbank Filtration Conference, G. Melin, ed., <strong>National</strong> <strong>Water</strong> <strong>Research</strong> <strong>Institute</strong>, Fountain Valley.<br />

Schubert, J. (2002). “Hydraulic aspects of riverbank filtration: Field studies.” Journal of Hydrology, 266 (3-4):<br />

154-161.<br />

STEVE HUBBS is a Professional Engineer with 28 years of experience at the Louisville<br />

<strong>Water</strong> Company in Louisville, Kentucky. In the early 1980s, he began researching riverbank<br />

filtration as an alternate source of water <strong>for</strong> the Louisville <strong>Water</strong> Company, specifically<br />

looking at the reduction in disinfection byproduct precursors, river-borne organics, and<br />

mutagenicity in the riverbank-filtration process. His work continued in the 1990s with a<br />

focus on pathogen reduction, and his research is now being conducted on the hydraulic<br />

connection between the riverbed and the aquifer, with a focus on riverbed plugging<br />

dynamics and their influence on sustainable yields from high-capacity riverbank-filtration systems. Hubbs<br />

received an M.S. in Environmental Engineering from the University of Louisville, and is currently enrolled<br />

in the Ph.D. program in Civil and Environmental Engineering at the University of Louisville, focusing on<br />

the hydraulics of riverbank-filtration systems.<br />

47


Session 3: Hydraulic Aspects<br />

Application of Different Tracers to Evaluate<br />

the Flow Regime at Riverbank-Filtration Sites<br />

in Berlin, Germany<br />

Dr. Gudrun Massmann<br />

Free University of Berlin<br />

Berlin, Germany<br />

Dipl. Geol. Andrea Knappe<br />

Alfred-Wegener-<strong>Institute</strong> Potsdam<br />

Pottsdam, Germany<br />

Doreen Richter<br />

Free University of Berlin<br />

Berlin, Germany<br />

Dr. Jürgen Sültenfuß<br />

University of Bremen<br />

Bremen, Germany<br />

Prof. Asaf Pekdeger<br />

Free University of Berlin<br />

Berlin, Germany<br />

Introduction<br />

The inhabitants of metropolitan Berlin (Germany) rely on drinking water derived from groundwater<br />

within the City’s boundary. Production well galleries are generally located adjacent to the surfacewater<br />

system and artificial infiltration ponds, and around 70 percent of abstracted groundwater is<br />

estimated to originate from bank filtration and artificial groundwater recharge (Pekdeger and<br />

Sommer-von Jarmerstedt, 1998). Berlin’s water production is a semi-closed water cycle with an<br />

indirect potable reuse system: local wastewater treatment plants discharge treated effluent into the<br />

surface-water system (Figure 1). Induced by well abstraction, surface water infiltrates into the<br />

ground and is used <strong>for</strong> drinking-water production. The raw water requires only minimal treatment<br />

to ensure good drinking-water quality. After use, water is redirected into wastewater treatment<br />

plants and released into the surface-water system again after treatment. Hence, the quality of the<br />

raw water is influenced by a number of factors, including:<br />

• Temporal and spatial variations of surface-water quality, largely depending on location in<br />

relation to wastewater treatment plants.<br />

• Presence, permeability, and thickness of the colmation layer.<br />

• Lithology, permeability, and geochemistry of the aquifer sediment.<br />

• Nature of the wells (location, length, and depth of filter screens).<br />

Correspondence should be addressed to:<br />

Dr. Gudrun Massmann<br />

Scientific Colleague<br />

Free University of Berlin<br />

Hydrogeology Group • Malteserstr. 74-100 • 12249 Berlin Germany<br />

Phone: +49-30-83870472 • Fax: +49-30-83870742 • Email: massmann@zedat.fu-berlin.de<br />

49


50<br />

TS GWA Tegel<br />

TS Wannsee<br />

TS Tegeler See<br />

WW Spandau<br />

Spandau WW Jungfernheide<br />

Charlottenburg<br />

KW Ruhleben<br />

Unterschleuse<br />

WW Kladow<br />

Havel<br />

WW Stolpe KW Schönerlinde<br />

Tegeler See<br />

Spree<br />

WW Tiefwerder<br />

WW und PEA<br />

Beelitzhof<br />

Kleinmachnow<br />

KW Stahnsdorf<br />

WW Tegel<br />

Tegeler Flieβ<br />

PEA-Tegel<br />

input summer<br />

KW Ruhleben<br />

Telfowkanal<br />

Nordgraben<br />

Legend<br />

flow direction<br />

waterworks<br />

sewage treatment plant<br />

surface water treatment plant<br />

transect / field site<br />

weir<br />

lock<br />

Figure 1. Overview of the surface-water system, flow directions, waterworks, wastewater and surface-water<br />

treatment plants, and study sites in the western part of Berlin.<br />

• Hydraulic regime resulting from natural gradients and pumping per<strong>for</strong>mances.<br />

• <strong>Water</strong>-quality changes that occur during bank filtration.<br />

A major advantage of the application of bank filtration is the capability of the subsurface to<br />

remove contaminants during underground passage, either by physical filtration and/or (bio)chemical<br />

processes, such as adsorption, reduction, or degradation. In addition, the application of bank<br />

filtration spares natural groundwater resources, which inhibits the rise of deeper, more saline<br />

groundwater into local freshwater aquifers; however, since a proportion of the surface water in<br />

Berlin originates from treated effluent released by wastewater treatment plants, it is necessary to<br />

constantly monitor the fate of potential contaminants during bank filtration.<br />

Processes accompanying bank filtration and artificial recharge are currently studied in Berlin<br />

within a multidisciplinary cooperative project at the Berlin Centre of Competence called Natural<br />

and Artificial Systems <strong>for</strong> Recharge and Infiltration (NASRI) (KWB, 2002). The project focuses<br />

on the behavior and removal of, <strong>for</strong> example, pathogens, microcystins, and organic pollutants, as<br />

well as pharmaceutically active compounds, during underground passage. To interpret the<br />

behavior of these non-conservative water constituents, the local hydrogeological and hydraulic<br />

situation has to be well understood. Three field sites are the focus of current investigations, all in<br />

the western part of the City (see Figure 1). Two sites are located next to natural lakes (Tegeler<br />

See/Tegel Lake and Wannsee/Wannsee Lake) and one next to an artificial recharge pond (GWA<br />

Tegel). Only examples of the Wannsee Lake site will be discussed.<br />

Several tracers have been successfully applied in Berlin <strong>for</strong> various purposes. Some of these are<br />

useful to study water movement and derive mean residence times (e.g., delta deuterium [δD],<br />

delta oxygen 18 [δ 18 O], temperature [T], chloride [ C1 – ]), while others may be particularly useful<br />

in identifying the proportion of treated wastewater in surface water (e.g., boron [B], Cl – , δD, δ 18 O)<br />

or the proportion of either bank filtrate (e.g., ethylenediaminetetraacetic acid [EDTA], galdolinium<br />

[Gd], strontium [Sr]) or deeper saline groundwater (e.g., B, Cl-, sodium [Na+]) in raw water.<br />

Panke<br />

Landwehrkanal<br />

(KW Marienfelde)


Methodology<br />

Transects generally reach from the lake or artificial recharge pond to a production well parallel to<br />

the flow direction. These transects contain a number of observation wells, usually one below the<br />

lake, some between the lake and production well, and some beyond the well. A brief overview on<br />

the analysis discussed here is given in Table 1.<br />

The T-He age dating method uses the ratio of the concentration of radioactive tritium ( 3 H or T)<br />

derived from atmospheric nuclear bomb testing and its decay product, Helium ( 3 He), in<br />

groundwater to determine a groundwater age (i.e., the time passed since water had its last contact<br />

with the atmosphere).<br />

In case the abstracted water is a mixture of surface water and background groundwater only, the<br />

percentage of bank filtrate in the well (X) can be calculated as:<br />

X = [C w–C GW)/(C SW–C GW)]•100 [%]<br />

With C as the concentration of a suitable tracer in groundwater (C GW), well water (C w), or surface<br />

water (CSW).<br />

The simple mixing <strong>for</strong>mula can be used under the premise that the differences between groundwater<br />

and surface water are large and relatively stable, and well water is a mixture of two water components<br />

only.<br />

Results<br />

Surface-<strong>Water</strong> System<br />

Table 1. Overview of Analytical Methods<br />

Parameter Analytical Method Laboratory Details Provided In<br />

4 He, 20 Ne, Mass Spectrometry <strong>Institute</strong> of Environmental Physics, Sültenfuß et al.<br />

and 22 Ne Pfeiffer QMG 112 University of Bremen (in preparation)<br />

(Age Dating)<br />

3 He and 4 He Mass Spectrometry <strong>Institute</strong> of Environmental Physics, Sültenfuß et al.<br />

(Age Dating) MAP215-50 University of Bremen (in preparation)<br />

Cl – Photometry Free University of Berlin<br />

(Autoanalyser,<br />

Technicon)<br />

B ICP-OCE Free University of Berlin<br />

δ 2 H, δ 18 O Delta S MS Alfred-Wegener-<strong>Institute</strong>, Meyer et al. (2000)<br />

Potsdam<br />

EDTA DIN 38413-PO3 Berlin <strong>Water</strong> Works<br />

4 He = Helium 4. 20 Ne = Neon 20. 22 Ne = Neon 22. 3 He = Helium 3. 4 He = Helium 4. ICP-OCE = Ionic Coupled Plasma.<br />

The strain on surface water becomes larger as it flows through the City and is substituted by<br />

considerable amounts of treated wastewater. The wastewater influence provokes high<br />

concentrations of, <strong>for</strong> example, Cl – , B, anthropogenic Gd, high conductivities and temperatures<br />

(in winter), and more negative isotope signatures. The highest influence and largest fractions of<br />

treated wastewater can be seen in the Teltow Canal and Nordgraben (a ditch). The Havel River<br />

51


52<br />

contains the lowest concentration of wastewater indicators. For example, Figure 2 shows the<br />

B concentration and the proportions of treated wastewater in surface water in September 2002. The<br />

influence of the Nordgraben extends to transect Tegel Lake, while the field site Wannsee Lake is<br />

under the influence of the Teltow Canal. The percentage of treated wastewater is higher during<br />

the summer months, when the Ruhleben Treatment Plant discharges into the Teltow Canal.<br />

Figure 2. (a) Percentage of treated wastewater in surface water (calculated with discharge only; evaporation,<br />

storage changes, and precipitation neglected). (b) Boron concentration indicating the influence of<br />

the Teltow Canal in the south and Nordgraben in the northeast <strong>for</strong> September 2002.<br />

By combining the results of the discharge measurements with the concentrations of wastewater<br />

indicators, proportions of treated wastewater can be calculated with a mixing <strong>for</strong>mula <strong>for</strong> each<br />

sampling point. In Figure 3, results are shown <strong>for</strong> Tegel Lake in front of the transect. In October<br />

2001, a pumping device started operation, which pumps Havel River water into the PEA-Tegel, a<br />

phophate elimination plant (see Figure 1), when the natural discharge of the Nordgraben is very<br />

low in summer. This led to a strong decrease of treated wastewater in Lake Tegel near the transect<br />

(from 33 to 46 percent [Fritz, 2002] to an average of 11.7 +/– 1.9 percent). For surface water near<br />

transect Wannsee Lake, values of 20.1 percent in the summer (when Ruhleben discharges into the<br />

Teltow Canal) and 8.5 percent in the winter were calculated.<br />

Bank-Filtration System (Wannsee Lake)<br />

At Wannsee Lake, two transects exist, running perpendicular to the shore of Wannsee Lake<br />

between drinking-water Production Wells 3 (TS Wannsee 2) and 4 (TS Wannsee 1). The wells<br />

of the local gallery (Beelitzhof) have several filter screens in different porous aquifers separated by<br />

aquitards. A schematic overview is given in Figure 4. Also shown are the T and 4 He concentrations<br />

originating from hydrogen bomb testing in the 1960s, as well as uranium and thorium decay<br />

within the aquifer. In the two deeper aquifers, groundwater is considerably older than 50 years,<br />

since T is already decayed and the 4 He values are high. In contrast, the shallow wells reflect the<br />

present atmospheric concentrations of T, while 4 He could not be detected. The resulting age is less<br />

than the detection limit of 3 months. The production well is a mixture of all aquifers.


Figure 3. Percentage of treated wastewater calculated with B and Cl <strong>for</strong> surface water in front of transect<br />

Tegel Lake.<br />

W E<br />

3339<br />

3338<br />

3337<br />

BEE201OP<br />

BEE201UP<br />

3336<br />

3334<br />

3335<br />

Brunnen 4<br />

BEE200UP<br />

BEE200OP<br />

3332<br />

m below m above<br />

ground sea-level<br />

Figure 4. Tritium and terrigenic 4 He in different observation wells and Production Well 4 of TS Wannsee 1.<br />

Mean residence times are estimated with the help of (<strong>for</strong> example) δD and δ 18 O breakthrough<br />

curves. The surface water shows a clear seasonal signal, with more negative signatures in winter<br />

(Figures 5a and 6). The exemplary breakthrough curves of Observation Wells 3,337 and 3,335<br />

(see Figure 4) reflect the surface-water signal. Here, groundwater is made of bank filtrate only. The<br />

curves illustrate that residence times are rather low at this particular site. Travel times are around<br />

1 month on the distance from the lake to Observation Well 3,337, which is approximately<br />

two-thirds of the way to the production well. Well 4 does not show a seasonal signal and has a more<br />

negative signature.<br />

53


54<br />

δ 18 O [ 0⁄00 versus Standard Mean Ocean <strong>Water</strong>]<br />

EDTA [µg/L]<br />

–6<br />

–7<br />

–8<br />

–9<br />

7<br />

6<br />

5<br />

4<br />

3<br />


similar in the lake and Wells 3 and 5. Mixing calculations assuming a residence time of 1 to 2 months<br />

result in an average percentage of bank filtrate of 61 to 97 percent in Well 3, 10 to 47 percent in<br />

Well 4, and 88 to 96 percent in Well 5. The large variations are mainly due to the detection limit<br />

of 2-µg/L EDTA, which results in uncertainties in the actual concentrations of Well 4 and<br />

background groundwater. An evaluation of other tracers suggest that the proportion of bank<br />

filtrate in Well 4 is more likely to be in the order of magnitude of 10 percent.<br />

The differences between the water bodies become clear in the plot of δD versus δ 18 O (Figure 6),<br />

where Well 4 plots next to the deeper aquifer samples. In contrast, Well 3 and Well 5 plot within<br />

the surface-water samples (just like the shallow observation wells), but show less seasonal variation.<br />

δD [ 0⁄00 versus Standard Mean Ocean <strong>Water</strong>]<br />

–46<br />

–48<br />

–50<br />

–52<br />

–54<br />

–56<br />

–58<br />

–60<br />

–62<br />

–64<br />

–66<br />

–68<br />

–70<br />

Figure 6. δD versus δ 18 O at the Wannsee Lake field sites. Analysis of deeper aquifer samples was done from<br />

September 2000 to November 2001, with remaining data from May 2002 to March 2003.<br />

Conclusions<br />

Mar 03<br />

Jan 03<br />

Feb 03<br />

May 02<br />

Dec 02<br />

–9 –8 –7 –6 –5<br />

δD [ 0⁄00 versus Standard Mean Ocean <strong>Water</strong>]<br />

Jun 02<br />

Nov 02<br />

Sep 02<br />

Aug 02<br />

Jul 02<br />

The combination of different tracers enables the interpretation of the flow regime at sites where<br />

bank-filtration processes are currently studied in Berlin. With the help of T/He analysis, the ages<br />

of different water bodies can be estimated. The analysis of tracers showing distinct seasonal<br />

variations is used to estimate travel times, while water constituents — which are either mainly<br />

present in bank filtrate or background water — are used <strong>for</strong> mixing calculations. The results of the<br />

tracer studies can be used to interpret the fate and behavior of potential contaminants (e.g., drug<br />

residues, organic pollutants, etc.).<br />

55


56<br />

Acknowledgements<br />

We would like to thank the German <strong>Research</strong> Foundation, Veolia <strong>Water</strong>, and the Berlin <strong>Water</strong><br />

Company <strong>for</strong> financing the NASRI Project.<br />

REFERENCES<br />

Fritz, B. (2002). Untersuchungen zur Uferfiltration unter verschiedenen wasserwirtschaftlichen, hydrogeologischen<br />

und hydraulischen Bedingungen, Ph.D. Thesis, University of Berlin, Berlin, 203 pp.<br />

KWB (2002). NASRI Natural and Artificial Systems <strong>for</strong> Recharge and Infiltration, First Progress Report, reporting<br />

period May 2002 – December 2002.<br />

Pekdeger, A., and C. Sommer-von Jarmerstedt (1998). Einfluß der Oberflächenwassergüte auf die Trinkwasserversorgung<br />

Berlins, Forschungspolitische Dialoge in Berlin, Geowissenschaft und Geotechnik, Berlin, pp. 33-41.<br />

Meyer, T., L. Schönicke, U. Wand, H.W. Hubberten, and H. Friedrichsen (2000). Isotope studies of hydrogen<br />

and oxygen in ground ice – Experiences with the equilibration technique, Isotopes Environ. Health Stud., p 133-149.<br />

Sültenfuß, J., G. Massmann, and A. Pekdeger (in preparation). Datierung mit der He3-T-Methode am Beispiel<br />

der Uferinfiltration im Oderbruch.<br />

GUDRUN MASSMANN studied Geology at both the Universities of Bremen and<br />

Edinburgh, specializing in Hydrogeology. After receiving her Diploma, she worked as an<br />

occupational trainee at the Center <strong>for</strong> Groundwater Studies at the Commonwealth<br />

Scientific & Industrial <strong>Research</strong> Organization (CSIRO) Land and <strong>Water</strong> in Adelaide,<br />

Australia, gaining experience on Artificial Storage and Recovery. For the past 4 years, she<br />

has worked as a scientific colleague <strong>for</strong> the Hydrogeology Group of the Free University of<br />

Berlin. Her research interests include hydraulic modeling, hydrochemistry, tracer analysis,<br />

and isotopes. Massmann finished her Ph.D. in summer 2002 at the Free University of Berlin on a project<br />

dealing with evaluating and modeling hydraulic and hydrochemical processes accompanying bank filtration<br />

in a polder region in Germany. She then continued to word on bank filtration, but changed the field site to<br />

Berlin, where bank filtration plays an important role in drinking-water production.


Session 4: Siting<br />

Siting and Testing Procedures<br />

<strong>for</strong> Riverbank-Filtration Systems<br />

Samuel M. Stowe, P.G., CPG<br />

International <strong>Water</strong> Consultants, Inc.<br />

Columbus, Ohio<br />

Phased, multi-tasked investigations can be structured to collect pertinent data to evaluate yield,<br />

quality, and (ultimately) the design of <strong>RBF</strong> systems. The initial phase should:<br />

• Define project objectives (yield, quality, water use).<br />

• Identify project concerns (treatment and regulatory issues).<br />

• Collect available data <strong>for</strong> preliminary hydrogeological screening/feasibility.<br />

If the initial phase indicates a reasonable probability that potential favorable conditions exist<br />

(given project objectives), then subsequent phases can be designed to collect site-specific data <strong>for</strong><br />

comprehensive analysis. Subsequent phases should include:<br />

• Test drilling.<br />

• Hydraulic interval testing.<br />

• <strong>Water</strong>-quality screening.<br />

• Streambed characterization.<br />

• <strong>Water</strong>-level monitoring.<br />

• Detailed aquifer and water-quality testing.<br />

The key parameters to any <strong>RBF</strong> evaluation are aquifer transmissivity and streambed permeability.<br />

A good understanding of these parameters, along with the hydrogeological setting, will result in<br />

the proper evaluation of the expected yield and quality of a <strong>RBF</strong> system and will allow <strong>for</strong> the<br />

thorough evaluation of potential means of development. Understanding the ability of the aquifer<br />

to provide sufficient <strong>RBF</strong> to recharge water pumped from a <strong>RBF</strong> system is key to ensuring that<br />

long-term capacities can be sustained and that target water quality can be maintained through a<br />

balance of infiltrated surface water and groundwater.<br />

Two case studies are presented to provide examples of recent <strong>RBF</strong> investigations. One is <strong>for</strong> a<br />

low-yield system (1.0 MGD), with very difficult access conditions in New Mexico, where rafts,<br />

helicopters, and portable drilling equipment were used. The other is a high-yield system<br />

(50 MGD) along a large river in the Great Plains, where more standard access and drilling methods<br />

(rotary, rotosonic) were used.<br />

Correspondence should be addressed to:<br />

Samuel M. Stowe, P.G., CPG<br />

President<br />

International <strong>Water</strong> Consultants, Inc.<br />

6360 Huntley Road • Columbus, Ohio 43229 USA<br />

Phone: (614) 888-6263 • Fax: (614) 888-9208 • Email: smstowe@collectorwellsint.com<br />

57


58<br />

Phase 1 Investigations<br />

Once the project objectives have been adequately defined, existing data that pertain to the<br />

hydrogeology and hydrology of the region of interest should be collected, assimilated, and assessed.<br />

This research should focus on the feasibility of developing the required capacity from<br />

unconsolidated deposits. In<strong>for</strong>mation collected <strong>for</strong> evaluation would include:<br />

• Well logs.<br />

• Bridge boring logs.<br />

• Steam flow and quality data.<br />

• Records of production wells <strong>for</strong> other water utilities in the region.<br />

• Reports on file with state or local agencies.<br />

• Other in<strong>for</strong>mation that pertains to the hydrogeological setting of the area.<br />

Additionally, sites that are identified as potentially favorable should be visited and inspected.<br />

Phase 1 findings should include the following:<br />

• Listings of the available data reviewed and other resources used.<br />

• Discussions of the data and results of any preliminary analyses that were possible, with the<br />

confidence level of data obtained.<br />

• Identification of areas that appear to have a favorable potential <strong>for</strong> testing.<br />

• Preliminary rankings of prospective sites, with a discussion of relative advantages and<br />

disadvantages <strong>for</strong> siting.<br />

• Evaluations regarding water quality that may be expected from a <strong>RBF</strong> system located in<br />

this area, considering depth, proximity to the river, other regional water-quality problems<br />

experienced, and local features that may impact quality, etc.<br />

• Assessment of permits that may be required <strong>for</strong> installation and withdrawal using a<br />

<strong>RBF</strong> system.<br />

• Descriptions and budgets <strong>for</strong> work procedures recommended to be completed in<br />

subsequent phases.<br />

Phase 2 Investigations<br />

Phase 2 tasks would include drilling, sampling, and testing exploratory borings/observation wells<br />

at sites identified in Phase 1. Given logistics and site layouts, surface geophysical surveys (seismic,<br />

electrical resistivity, electromagnetic) could precede actual drilling. Geophysical surveys can be<br />

helpful in preliminary screening activities. These sites would be selected based upon access (land<br />

ownership, permission, etc.) and the potential <strong>for</strong> favorable saturated aquifer thicknesses near the<br />

river. The primary objective of this Phase is to locate the most favorable site(s) <strong>for</strong> detailed aquifer<br />

testing (Phase 3) and the collection of data <strong>for</strong> hydraulic and quality evaluations <strong>for</strong> site ranking.<br />

The selection of appropriate drilling and sampling methods are imperative during this phase.<br />

Drilling methods should consider:<br />

• Representative soil samples.<br />

• Expected drilling conditions (materials, depths).<br />

• Rate of advancement (timely drilling).<br />

• Cost-effectiveness.


• Potential <strong>for</strong> hydraulic testing and water-quality sampling.<br />

• Site access conditions.<br />

• Monitoring well installation.<br />

Potential drilling/sampling methods include:<br />

• Hollow stem auger with split-spoon sampling.<br />

• Cable tool with bailed samples.<br />

• Direct rotary with wash and split-spoon sampling.<br />

• Rotosonic with dual-tube continuous sampling.<br />

• Reverse rotary with wash samples.<br />

Generally, it has been found that rotosonic drilling provides the best overall results <strong>for</strong> representative<br />

soil sampling, speed, water sampling, and hydraulic testing. Drawbacks have been costs and<br />

access. In the rotosonic drilling method, a drill casing is advanced into the ground using<br />

rotary/vibrasonic techniques. This method does not require the use of drilling mud, so there is no<br />

mud to dispose of, and ground-surface disturbance is minimal. The rotosonic drilling method<br />

produces nearly continuous samples of the materials penetrated by the sample tube, and the<br />

method produces representative samples from unconsolidated, granular deposits.<br />

Sieve analyses should be per<strong>for</strong>med on selected lithologic samples collected from test borings to<br />

determine optimum well-screen design and to help evaluate hydraulic conductivity. Hydraulic<br />

interval testing can also be conducted on test borings to determine the hydraulic conductivity of<br />

selected intervals and to evaluate groundwater quality.<br />

Following the collection of all field data, the data are compiled and analyzed to determine the<br />

most favorable locations <strong>for</strong> detailed aquifer testing (Phase 3). This determination is based upon<br />

saturated thicknesses, hydraulic conductivities, and logistics. From Phase 2 results, it can be<br />

determined with a reasonable degree of certainty that one or more of the sites tested is capable of<br />

yielding the desired quantity of water.<br />

Phase 3 Investigations<br />

At the completion of Phase 2, if results are favorable, a site (or sites) <strong>for</strong> detailed aquifer testing is<br />

selected. Phase 3 studies generally include the installation of additional observation wells <strong>for</strong><br />

sampling subsurface materials, monitoring water levels, and conducting a detailed aquifer test,<br />

along with a test pumping well. Additionally, the evaluation of streambed conditions (width,<br />

depth, permeability) is generally conducted during this phase.<br />

Observation wells are positioned in a pattern around the test pumping well at appropriate locations<br />

and distances selected to facilitate data analysis, with emphasis on <strong>RBF</strong>. Following the installation<br />

of observation wells, a temporary test pumping well is installed. It is important that the test<br />

pumping well be constructed to produce enough water to adequately stress the aquifer. Well points<br />

can be installed in the river (conditions permitting) to further evaluate streambed permeability.<br />

The monitoring of water levels in the wells and river should be conducted prior to any pumping<br />

to evaluate antecedent trends and the correlation of stream level and groundwater levels, with<br />

levels converted to elevation datum <strong>for</strong> analysis. Following the collection of background water<br />

levels, a long-term (2- to 10-days) constant rate test is conducted. A general guideline is to run<br />

the test 24 hours after steady-state conditions are reached. At the conclusion of the test, the<br />

59


60<br />

pumping is discontinued and the recovery of water levels is monitored until at least 90-percent<br />

recovery is obtained in the observation wells. During the test, samples of the water pumped (and<br />

river) can be obtained <strong>for</strong> laboratory analyses. Additionally, the pump discharge should be<br />

monitored every 6 to 12 hours and river quality every 24 hours <strong>for</strong> field screening of temperature,<br />

pH, iron, hardness, turbidity, and specific conductance.<br />

All data from Phase 3 are analyzed to determine aquifer properties, streambed permeability, aquifer<br />

character, yield, anticipated quality, and conceptual <strong>RBF</strong> system design. Yield estimates should<br />

consider the impacts of river stage/flow and water temperature variations. Based upon the results,<br />

the preferred location of a <strong>RBF</strong> system can be recommended based upon an evaluation of benefits<br />

and drawbacks, including refinements in design. Phase 3 testing may also identify additional<br />

testing that may be necessary to finalize design.<br />

Case Studies<br />

Two case studies are briefly discussed to illustrate the phased approach to siting <strong>RBF</strong> systems and<br />

critical evaluation items. Case Study Number 1 was completed in north-central New Mexico along<br />

the Rio Grande (Figure 1), where difficult access conditions presented challenges <strong>for</strong> completion<br />

using conventional approaches. Project objectives were to locate a site that could yield at least<br />

1.0 MGD of infiltrated surface water from the Rio Grande <strong>for</strong> potable use (surface-water rights to<br />

1,200 acre-feet per year). Phase 1 studies (literature review, site reconnaissance) identified up to<br />

six potentially favorable sites within the canyon. Given difficult site access, the initial Phase 2 study<br />

involved surface geophysics (electrical resistivity) at the six locations, along with a control site.<br />

Access to the sites was obtained using an all terrain vehicle and rafts, with base camps set up <strong>for</strong><br />

overnight stays. Based upon the survey, two sites were selected <strong>for</strong> test drilling.<br />

Figure 1. Rio Grande Valley, New Mexico.<br />

Drilling was conducted using a portable modular core rig with wash samples and 2-inch<br />

outside-diameter split-spoon samples. At Site A, nine borings were drilled, six 1.5-inch diameter<br />

observation wells installed, and hydraulic testing (high-efficiency suction pumps) completed. As<br />

results were favorable, Phase 3 aquifer testing was completed with additional observation wells and<br />

a 66-hour constant rate test (0.144 MGD).<br />

Following the completion of testing at Site A, the equipment and base camp were mobilized to<br />

Site B using rafts and a helicopter. Phase 2 testing included the drilling of four borings, interval


testing, and a short constant rate test. As results here were not as favorable as at Site A, detailed<br />

aquifer testing was not completed.<br />

Case Study Number 2 was completed along the Missouri River in North Dakota (Figure 2), where<br />

more conventional investigative methods could be used. Project objectives were to locate a site<br />

that could yield up to 50 MGD of potable water from a <strong>RBF</strong> system. The study was being<br />

completed to evaluate the possibility of replacing an existing direct surface-water intake. Phase 1<br />

investigations (desktop study, site visit) determined that several sites existed within the area,<br />

which were promising. Phase 2 investigations were targeted at two sites, given the primary<br />

logistical concern of distance from existing treatment. Phase 2 included the drilling of three<br />

borings and conducting one hydraulic test using rotosonic methods.<br />

Figure 2. Missouri River Valley, North Dakota.<br />

Given favorable Phase 2 results, Phase 3 investigations were completed at one site with the<br />

installation of four additional observation wells and a test pumping well. Because of budget<br />

concerns, Phase 3 drilling was completed using mud rotary methods and a local drilling contractor.<br />

Testing was completed with the running of a 72-hour constant rate test (1.9 MGD). The analysis<br />

of data indicated a transmissive aquifer (40,000 square feet per day) in good communication with<br />

the river. Preliminary results indicate that the site will yield 30 MGD from a single horizontal<br />

collector well.<br />

SAM STOWE is President of International <strong>Water</strong> Consultants, Inc., a subsidiary of<br />

Collector Wells International, Inc. A hydrogeologist, Stowe has nearly 30 years of diverse<br />

experience in the groundwater industry. He has been in charge of projects involving<br />

aquifer-test analyses, riverbank filtration and recharge evaluation, groundwater quality,<br />

well design, groundwater management, numerical modeling, contamination investigation,<br />

and remedial action. He has also been involved in groundwater-supply projects <strong>for</strong> yields<br />

of nearly 100 million gallons per day and in contamination evaluations ranging from<br />

industrial organic pollution from landfills to the environmental effects of strip and deep-coal mining. In<br />

addition, Stowe is highly experienced in evaluations of induced infiltration potential from streams and<br />

oceans. He has completed hundreds of hydrogeological evaluations <strong>for</strong> horizontal collector wells in regard to<br />

yield, quality, and design, having been responsible <strong>for</strong> siting and designing many horizontal collector wells<br />

with individual yields ranging from 2 to 40 million gallons per day and is a recognized expert in their<br />

application. Stowe received a B.A. in Geology from Miami University and an M.S. in Geology from Ohio<br />

State University.<br />

61


Session 4: Siting<br />

<strong>Water</strong>-Quality Management <strong>for</strong> Existing Riverbank-<br />

Filtration Sites along the Elbe River in Germany<br />

Prof. Dr.-Ing. Thomas Grischek<br />

University of Applied Sciences Dresden<br />

Dresden, Germany<br />

Introduction<br />

In <strong>for</strong>mer times, the management of <strong>RBF</strong> sites in Germany focused on water quantity — how much<br />

water could be abstracted and what groundwater levels would result. Control measures were based<br />

on measurements of river and groundwater levels and pumping rates and on black-box models to<br />

estimate the portion of riverbank filtrate abstracted (e.g., Luckner and Nestler, 1982). The<br />

development of computer programs <strong>for</strong> groundwater flow and transport simulations resulted in<br />

increasing management applications <strong>for</strong> <strong>RBF</strong> sites (e.g., Koster et al., 1994; Heinzmann, 1998;<br />

Eckert et al., 2000).<br />

The number of published papers dealing with the complex management of water quality at <strong>RBF</strong><br />

sites is very low. Besides work done by Sontheimer (1991) and Schubert (1999) at sites along the<br />

Rhine River, a control and management concept <strong>for</strong> the Hengsen site on the Ruhr River<br />

(Schöttler and Sommer, 1992) must be highlighted. At the end of the 1980s, water-quality<br />

management also became a subject <strong>for</strong> <strong>RBF</strong> sites on the Elbe River due to problems with river and<br />

raw-water quality. Initial work done by Müller, Schwan, and others between 1985 and 1989 was<br />

continued by Nestler et al. (1998).<br />

<strong>Water</strong>-quality management must be based on detailed knowledge of groundwater flow conditions<br />

and monitoring measures to obtain sufficient data on water quality; however, this is not the case<br />

<strong>for</strong> every site. Of course, <strong>for</strong> small waterworks, such investigations and investments may be higher<br />

than the cost savings resulting from water-quality management. But, the optimization of water<br />

abstraction can have long-term effects on raw-water quality and treatment cost savings. <strong>Water</strong>quality<br />

management systems are well established at sites where problems with contamination have<br />

occurred (e.g., high concentrations of nitrate, organic halogens). In such cases, experiences may<br />

not be published to give the impression of a “no problem waterworks” or because a clear<br />

description is not easy due to the large amount of data needed to understand the complex system<br />

and site-specific boundary conditions.<br />

General <strong>Water</strong>-Quality Management Measures<br />

The first step in water-quality management is to clarify which advantage of <strong>RBF</strong> is the most<br />

important and to identify the main aims or problems. At one site, the nitrate concentration in the<br />

raw water should be decreased; at another site, the concentration of DOC, dissolved iron, or an<br />

organic contaminant should decrease.<br />

Correspondence should be addressed to:<br />

Prof. Dr.-Ing. Thomas Grischek<br />

Professor<br />

<strong>Institute</strong> <strong>for</strong> Geotechnics & <strong>Water</strong> Sciences<br />

University of Applied Sciences Dresden • Friedrich-List-Platz 1 • 01069 Dresden, Germany<br />

Phone: +49 351 4623350 • Fax: +49 351 4623567 • Email: grischek@htw-dresden.de<br />

63


64<br />

<strong>Water</strong>-quality management could include:<br />

• Optimization of the operation of existing abstraction wells.<br />

• Building additional abstraction wells.<br />

• Technical measures in the riverbed.<br />

• (Political) activities to improve river-water quality.<br />

• Contracts with farmers to change land use in a catchment zone.<br />

This paper focuses on the operation of existing wells affected by mixing processes in an aquifer,<br />

flow times and flow path lengths, and the catchment area <strong>for</strong> landside groundwater. Special<br />

measures to face problems resulting from contaminant shock loads in a river, as well as droughts<br />

or floods, are not covered because these are the subjects of other contributions in this volume.<br />

Strategies based on the operational control of existing abstraction wells could include:<br />

• Well operation to achieve a maximum volume of utilized aquifer (long flow paths and<br />

retention times).<br />

• Preferential operation of selected wells with best raw-water quality.<br />

• Operation of a limited number of wells with high abstraction rates to increase the<br />

proportion of bank filtrate in raw water.<br />

• Continuous operation of selected wells to meet the mean water demand and additional<br />

operation of other wells in peak periods.<br />

• Periodic operation of wells to increase the effects of dispersion and mixing in the aquifer.<br />

An additional technical measure could be a change of the well filter length and depth in an<br />

aquifer.<br />

Riverbank Filtration along the Elbe River<br />

At present, there are eight water facilities along the Elbe River that use <strong>RBF</strong> to supply more than<br />

1.5-million people with drinking water and industry with process water. Three of these sites have<br />

been chosen to carry out further investigations on water-flow and water-quality changes during<br />

<strong>RBF</strong>, especially due to the changing water quality of the Elbe River in the 1990s. Between 1992<br />

and 2002, different research programs focused mainly on the behavior of DOC, EDTA, sulfur<br />

organic compounds, and carbonic acids. Besides this specialized research, daily problems among<br />

water companies were investigated.<br />

All <strong>RBF</strong> sites along the Elbe River are operated under anoxic conditions. Reducing conditions<br />

result in the dissolution of iron and manganese along the flowpath between the river and<br />

abstraction wells. At some sites, denitrification occurring during <strong>RBF</strong> was found to compensate <strong>for</strong><br />

high nitrate concentrations in groundwater from other sources.<br />

Due to the closure of large industries and a new water pricing system after the reunification of<br />

Germany, water demand decreased significantly between 1989 and 1998. Lower abstraction rates<br />

<strong>for</strong> bank filtrate and an improvement in river-water quality allow <strong>for</strong> efficient groundwater<br />

management and the optimization of pumping schemes to obtain good raw-water quality.


Torgau Case Study<br />

The highly productive <strong>RBF</strong> scheme near Torgau in Saxony, eastern Germany, is situated within the<br />

Elbe River Basin. The basin is filled with Pleistocene deposits to a depth of 10 to 55 m that comprise<br />

interfingered glaciofluviatile sediments ranging from fine sand and silt to medium sand and gravel<br />

(Grischek et al., 1998). The deposits are overlain by Holocene river gravels (5- to 8-m thick) and<br />

meadow loam (2-m thick). The hydraulic conductivity values range from 0.6-2 × 10 –3 m per second.<br />

The Elbe River is in direct hydraulic contact with the aquifer. Riverbed clogging is low. The <strong>RBF</strong><br />

system consists of 42 vertical wells arranged in nine well groups (Figure 1) and has a capacity of<br />

150,000 m 3 /d. The distance between abstraction wells and the riverbank is about 300 m. The flow<br />

time of bank filtrate is between 80 and 300 days. The wells have a 20-m long filter placed in the<br />

lower layer of the aquifer at 30- to 50-m below ground surface. The site is well equipped with two<br />

sampling profiles along flowpaths between the river and abstraction wells, with monitoring wells<br />

in the landside catchment zone.<br />

5712<br />

5710<br />

5708<br />

5706<br />

2.3*<br />

2.0<br />

3.2<br />

2.6*<br />

3.2<br />

4.4<br />

4.1*<br />

1.7<br />

3.1<br />

I<br />

1.0<br />

1.0<br />

3.1<br />

II<br />

1.2<br />

1.9<br />

2.6<br />

2.7<br />

III<br />

1.3<br />

1.1<br />

3.0 1.0<br />

1.4 4.1<br />

0.8<br />

3.3 1.0<br />

0.9<br />

IV<br />

1.7<br />

3.3<br />

2.9<br />

2.8<br />

Mixed water<br />

DOC < 2.3 mg/L<br />

1.1<br />

0.9<br />

2.8<br />

0.8<br />

Lakes<br />

0.8<br />

1.5<br />

1.8<br />

3.3<br />

2.7<br />

4.6*<br />

2.8<br />

3.0<br />

5.1<br />

Mixed water 4.2<br />

DOC 2.3...3.0 mg/L<br />

Elbe River<br />

V VI<br />

VII<br />

1.6<br />

2.0<br />

4.3<br />

2.6<br />

5.2 5.0<br />

Mixed water 4.8<br />

DOC > 3 mg/L<br />

Legend<br />

Abstraction well<br />

3.7<br />

3.2<br />

3.0 3.3<br />

1.9<br />

1.7<br />

4571 4573 4575 4577<br />

Figure 1. Mean DOC concentration in milligrams per liter in groundwater in the catchment zone of the<br />

Torgau <strong>Water</strong>works from 1995 to 1997.<br />

The main aims of water-quality management include the maximum attenuation of organic<br />

compounds during aquifer passage and low concentrations of DOC, dissolved iron, and nitrate in<br />

raw water. Results from long-term monitoring programs and special field experiments provided an<br />

excellent database to draw conclusions on the effect of water-quality management measures. A<br />

detailed description of redox conditions and removal rates of organics is given in Grischek et al.<br />

(1998, 2000). Table 1 summarizes the effects of different management measures <strong>for</strong> the <strong>RBF</strong> site<br />

at Torgau.<br />

2.6<br />

3.2<br />

4.4<br />

4.4*<br />

Observation well with three<br />

sampling depths; mean DOC<br />

concentration in mg/L, 1995-97<br />

Affected by local infiltration<br />

of surface water<br />

VIII<br />

1 km<br />

IX<br />

1.5<br />

old branch<br />

65


66<br />

Table 1. Effect of Control Measures on Raw-<strong>Water</strong> Quality at the <strong>RBF</strong> Site in Torgau<br />

(Grischek, 2002)<br />

Control Measure Bank- Effect on Raw-<strong>Water</strong> Quality<br />

Filtrate DOC Trace NO3 – Fe 2+ NH4 + SO4 2–<br />

Portion<br />

Organics Mn 2+<br />

Continuous operation of wells<br />

with increased abstraction rate<br />

Continuous operation of wells<br />

with decreased abstraction rate<br />

Periodic operation of selected wells<br />

Preferential operation of selected<br />

wells with favorable catchment area<br />

Operation of every second or third<br />

well within a group of wells<br />

Technical measure<br />

Change of the well filter depth<br />

➚<br />

➚<br />

➚<br />

➚<br />

— — —<br />

➚<br />

➚ ➚➚➚<br />

➚ ➚ ➚<br />

➚<br />

➚<br />

➚➚➚<br />

➚<br />

— — — — —<br />

— Effect is negligible or only temporary. Concentration increase. Concentration decrease.<br />

—<br />

Readily degradable and highly adsorbable organic compounds are attenuated in the biologically<br />

active riverbed. Long flow paths and long retention times of the bank filtrate in the aquifer allow<br />

further attenuation of poorly degradable and some polar organics.<br />

A field test was done to study the effect of an increase in water abstraction resulting in a decrease<br />

in retention time of bank filtrate in the aquifer. Over a period of 1.5 years, water abstraction from<br />

five selected wells was increased by about 40 percent. Excluding some slight changes in water<br />

quality near the bank line, no significant effect due to the decrease in retention time was observed.<br />

DOC removal and denitrification along the whole flowpath were not affected (Grischek, 2002).<br />

A continuous operation of selected wells has the advantage of lower concentrations of dissolved<br />

iron in raw water. High concentrations of dissolved iron were not measured in the bank filtrate,<br />

but were measured in the landside groundwater. Switching off the abstraction wells results in<br />

groundwater flow towards the river and the transport of dissolved iron into the aquifer zone<br />

between the wells and river. When the pumps are switched on again, higher iron concentrations<br />

are observed in raw water. Furthermore, periodic well operation leads to higher well clogging.<br />

A groundwater flow and transport model has been used to simulate the change of the well filter<br />

depth and its effect on groundwater flow, especially groundwater flow from the opposite side of the<br />

river beneath the riverbed to the wells. Due to the long distance between the wells and bank line,<br />

the effect was found to be negligible. A change in the filter depth should only be considered if the<br />

aquifer consists of layers with very different hydraulic conductivities and if the well is located at a<br />

short distance from the riverbank (the distance is less than aquifer thickness).<br />

Surprisingly, the most important factor was the selection of abstraction wells according to their<br />

catchment zone <strong>for</strong> landside groundwater. Based on the dense net of observation wells, it was possible<br />

to identify zones with different concentrations of DOC and dissolved iron in landside groundwater.<br />

Figure 1 shows the range of DOC concentrations in groundwater in the catchment zones behind<br />

the wells.<br />

—<br />

—<br />

—<br />

—<br />

➚<br />

—<br />

➚ ➚➚<br />

➚ ➚<br />

➚<br />

➚<br />

➚ ➚➚➚


The mean concentration in the “mixed water” (see Figure 1) was calculated from the concentrations<br />

determined at different depths in the aquifer. Depth-dependent groundwater sampling was a<br />

key factor in understanding groundwater flow and quality changes in the catchment zone. The<br />

operation of Well Groups II to IV results in lower DOC concentration in the raw water compared<br />

to the operation of Well Groups VIII and IX. Fortunately, low DOC concentrations are associated<br />

with low dissolved iron concentrations. Thus, the advantage of appropriately selecting abstraction<br />

wells covers both parameters. Because the water-quality change <strong>for</strong> bank filtrate was very similar<br />

in all wells, an improvement of raw-water quality can be achieved mainly by the selection of wells<br />

abstracting the proportion of landside groundwater with the best quality. At the Torgau<br />

<strong>Water</strong>works, the preferential operation of these wells has already resulted in cost savings,<br />

especially <strong>for</strong> the removal of dissolved iron during the water-treatment process that requires iron<br />

sludge disposal.<br />

Summary<br />

The main aim at all <strong>RBF</strong> sites along the Elbe River is the attenuation of organic compounds and<br />

low concentrations of DOC, dissolved iron, and manganese in raw water. Long flow paths and<br />

retention times promote the attenuation of organics, but were found to have only a relatively small<br />

effect on iron and manganese concentrations. In general, the continuous pumping of selected<br />

wells should be preferred over periodic operation. At all sites, mixing ratios of bank filtrate and<br />

groundwater were found to be of main importance <strong>for</strong> the concentration of nitrate, sulfate,<br />

dissolved iron, and manganese in raw water. Due to different groundwater qualities within the<br />

whole catchment zone, a selection of wells having a catchment zone with good groundwater<br />

quality offered a significant improvement in raw-water quality. Thus, monitoring systems <strong>for</strong> <strong>RBF</strong><br />

sites should not only focus on bank filtrate, but should include observation wells landside of the<br />

abstraction wells. Due to site-specific boundary conditions, a detailed investigation of groundwater<br />

flow conditions and proportions of bank filtrate in raw water is very important <strong>for</strong> drafting<br />

effective water-quality management measures.<br />

REFERENCES<br />

Eckert, P., C. Blömer, J. Gotthardt, S. Kamphausen, D. Liebich, and J. Schubert (2000). “Correlation<br />

between the well field catchment area and transient flow conditions.” Proceedings, International Riverbank<br />

Filtration Conference, W. Jülich and J. Schubert (eds.), IAWR Rheinthemen 4, 103-113.<br />

Grischek, T., KM. Hiscock, T. Metschies, P.F. Dennis, and W. Nestler (1998). “Factors affecting denitrification<br />

during infiltration of river water into a sand and gravel aquifer in Saxony, Germany.” Wat. Res.,<br />

32(2): 450-460.<br />

Grischek, T., E. Worch, and W. Nestler (2000). “Is bank filtration under anoxic conditions feasible?”<br />

Proceedings, International Riverbank Filtration Conference, W. Jülich and J. Schubert (eds.), IAWR<br />

Rheinthemen 4, 57-65.<br />

Grischek, T. (2002). Zur Bewirtschaftung von Uferfiltratfassungen an der Elbe (Management of riverbank filtration<br />

sites along the River Elbe), Ph.D. thesis, Dresden University of Technology, <strong>Institute</strong> <strong>for</strong> Groundwater<br />

Management (in German).<br />

Heinzmann, B. (1998). “Beispiele für integriertes Management von Wasserressourcen in der Region Berlin-<br />

Brandenburg. (Examples of integrative management of water resources in the Berlin-Brandenburg region).”<br />

Wasserwirtschaft in urbanen Räumen. Schriftenreihe Wasser<strong>for</strong>schung, 3: 171-197 (in German).<br />

Koster, N., U. Willme, and K. Döhmen (1994). “Wasserwirtschaftliche Betriebsanalyse am Beispiel einer<br />

Wassergewinnungsanlage im Ruhrtal (Analysis of water management at a waterworks in the River Ruhr<br />

valley).” Wasser Abwasser Praxis, 5: 19-23 (in German).<br />

67


68<br />

Luckner, L., and W. Nestler (1982). “Zur Methodik des Aufbaus und der Nutzung von Kontroll- und<br />

Steuerungsprogrammen von Grundwasserfassungen (On the methodology of structuring and use of control<br />

programs <strong>for</strong> groundwater abstraction wells).” Wasserwirtschaft-Wassertechnik, 32(7): 219-223 (in German).<br />

Nestler, W., W. Walther, F. Jacobs, R. Trettin, and K. Freyer (1998). Wassergewinnung in Talgrundwasserleitern<br />

der Elbe (<strong>Water</strong> production in alluvial aquifers along the River Elbe), UFZ-<strong>Research</strong> Report 7 (in German).<br />

Schöttler, U., and H. Sommer (1992). “Optimierung einer Wassergewinnungsanlage mit Hilfe der Modellrechnung<br />

und ihre Auswirkungen auf die Grundwassergüte (Optimisation of a water abstraction system with<br />

the help of model calculations and the effects on groundwater quality).” Steuerung in der Wasserwirtschaft,<br />

DFG-Forschungsbericht, VCH-Verlagsges., Weinheim, 178-195 (in German)<br />

Schubert, J. (1999). “Riverbank filtration — Field studies, modeling, monitoring.” Proceedings, International<br />

Riverbank Filtration Conference, 4-6 Nov 1999, Louisville, Kentucky, 39-42.<br />

Sontheimer, H. (1991). Trinkwasser aus dem Rhein? Bericht über ein Verbund<strong>for</strong>schungsvorhaben zur Sicherheit<br />

der Trinkwassergewinnung aus Rheinuferfiltrat (Drinking water from the River Rhine? Report on a collaborative<br />

research project on safety of drinking water production from bank filtrate from the River Rhine), Academia Verlag,<br />

Sankt Augustin, Germany (in German).<br />

THOMAS GRISCHEK has 15 years of research experience in the field of riverbank<br />

filtration. He published eight papers on riverbank filtration in refereed journals as principal<br />

author and has contributed to more than 10 papers as co-author. His main research<br />

interests are the interaction of groundwater and surface water, especially riverbank<br />

filtration and artificial groundwater recharge, groundwater management, and diffusing<br />

groundwater pollution. Currently, he is Professor of <strong>Water</strong> Science at the University of<br />

Applied Sciences Dresden. Prior to this position, Grischek spent 8 years as a <strong>Research</strong><br />

Associate <strong>for</strong> the <strong>Institute</strong> <strong>for</strong> Groundwater Management at the Dresden University of Technology, as well<br />

as at the University of Applied Sciences Dresden. In addition, he also taught at the <strong>Institute</strong> <strong>for</strong> <strong>Water</strong><br />

Chemistry at the Dresden University of Technology. In 2003, he worked as a Referee at the Saxon State<br />

Agency <strong>for</strong> Environment and Geology. Grischek studied <strong>Water</strong> Management in the Department of <strong>Water</strong><br />

Sciences at Dresden University of Technology, where he graduated as a diploma engineer. He also received<br />

a Ph.D. in Environmental Engineering from the Dresden University of Technology in 2002, where he<br />

researched the “Management of Riverbank-Filtration Sites along the Elbe River.”


Session 5: Dynamics<br />

Using Models to Predict Filtrate Quality<br />

at Riverbank-Filtration Sites –<br />

What Is the Adequate Level of Modeling?<br />

Chittaranjan Ray, Ph.D., P.E.<br />

University of Hawaii at Mañoa<br />

Honolulu, Hawaii<br />

Henning Prommer, Ph.D.<br />

Delft University of Technology<br />

Delft, The Netherlands, and<br />

Center <strong>for</strong> Groundwater Studies<br />

Perth, Australia<br />

Prof. Dr.-Ing. Thomas Grischek<br />

University of Applied Sciences<br />

Dresden, Germany<br />

<strong>RBF</strong> is an efficient and low-cost treatment technology <strong>for</strong> drinking-water production that has<br />

been in use <strong>for</strong> more than a century in Europe and nearly half a century in the United States.<br />

There has been a recent surge in the design and construction of <strong>RBF</strong> facilities in the United States<br />

due to the potential to receive 0.5- to1.0-log removal credit <strong>for</strong> protozoa Cryptosporidium filtration<br />

under the upcoming Enhanced Surface <strong>Water</strong> Treatment Rule (USEPA, 2002). While most small<br />

water utilities use <strong>RBF</strong> as the sole treatment system (with the exception of disinfection), most<br />

medium to large utilities use <strong>RBF</strong> as a pretreatment system, which helps process per<strong>for</strong>mance in<br />

the final treatment system.<br />

The design and operation of <strong>RBF</strong> systems are facilitated by the use of flow and transport models<br />

<strong>for</strong> estimating yield and water quality of the filtrate; however, the level of the modeling ef<strong>for</strong>t can<br />

vary depending upon the utility, data availability, and problems to be solved. While water-quality<br />

modeling is the primary focus, flow simulation is the first step prior to undertaking transport<br />

simulations. In this paper, we present various scenarios involved in flow and transport modeling<br />

at <strong>RBF</strong> sites and discuss what level of modeling is adequate <strong>for</strong> what problems.<br />

Flow Simulations<br />

Most utilities undertake some sort of flow simulation to estimate yield from <strong>RBF</strong> wells. These<br />

simulations are broadly divided into (a) analytical and (b) numerical models. Analytical models<br />

(Ferris et al., 1962; Hantush, 1959) assume that the river fully penetrates the aquifer and there is<br />

no additional resistance to flow at the river-aquifer interface. Hantush (1959) also presented a<br />

simple procedure to handle partial penetration of the river by moving the recharge boundary some<br />

distance away from the actual location with respect to the pumping well. Other analytical models,<br />

Correspondence should be addressed to:<br />

Chittaranjan Ray, Ph.D., PE<br />

Associate Professor<br />

Department of Civil & Environmental Engineering<br />

University of Hawaii at Mañoa • 2540 Dole Street, 383 Holmes Hall • Honolulu, Hawaii 96822 USA<br />

Phone: (808) 956-9652 • Fax: (808) 956-5014 • Email: cray@hawaii.edu<br />

69


70<br />

as described by Dillon and Ligett (1983) and Wilson (1993), include transient effects and<br />

clogging. Their application is mostly limited to two-dimensional problems. Three-dimensional<br />

flow and hydraulic and material heterogeneity cannot be easily handled by these methods. Conrad<br />

and Beljin (1996) provide a result comparison from the application of analytical and numerical<br />

models <strong>for</strong> different sites.<br />

Nowadays, numerical models are most commonly used to simulate stream-aquifer interaction.<br />

MODFLOW (Harbaugh et al., 2000) and MODPATH (Pollock, 1994) are some of the commonly<br />

used models <strong>for</strong> estimating hydraulic heads and path lines of neutrally buoyant particles.<br />

MODFLOW can handle material heterogeneity and various boundary conditions. The River and<br />

Stream/Aquifer Interaction packages are typically used in MODFLOW to examine the interaction<br />

between surface water and groundwater. In the River package, the hydraulic heads in the aquifer<br />

and the stage of water in the river are prescribed. Certain block-centered grids serve as the river.<br />

If the hydraulic head in the aquifer is lower than river stage, then the flow is from the river to the<br />

aquifer and vice versa. The hydraulic conductivity of riverbed material controls the flow into or<br />

out of the river. In contrast to the River package, where the hydraulic heads are specified, the flow<br />

is routed through the channel in the Stream/Aquifer Interaction package using the Manning<br />

equation, channel geometry, roughness, and other parameters.<br />

MODPATH is used to conduct advective tracking of neutrally buoyant particles in the flow field using<br />

<strong>for</strong>ward or reverse particle tracking techniques. For example, reverse particle tracking from the screen<br />

zone of a well can determine the areas contributing to flow to a well. Similarly, <strong>for</strong>ward tracking of a<br />

set of particles from a riverbed would indicate if any of the particles will be captured by pumping wells<br />

located on riverbanks. Path lines can be delineated <strong>for</strong> steady and transient flow conditions.<br />

In MODPATH simulations, the hydraulic gradient and well location can play important roles. For<br />

example, in a stream-aquifer simulation, the path lines <strong>for</strong> a single well can be seen in Figure 1.<br />

This figure shows the impact of the placement of one or more wells on the flow field. In the left<br />

side of this figure, a portion of the flow to the pumping well originates from the river and the rest<br />

comes from upgradient locations; however, if a second well is installed to the northwest of this well<br />

(figure to the right), a hydraulic divide is created. As apparent, most of the flow to this new well<br />

comes from the aquifer.<br />

In most river-aquifer simulation models, the nearest riverbank or the mid-point of the river is often<br />

considered as the model boundary. This assumption may hold as long as the river is a significant<br />

Figure 1. Flow fields <strong>for</strong> one and two pumping wells located on the bank of the Illinois River.


source of induced infiltration to the well; however, if the stream bottom is constituted with low<br />

permeability material and the pumping rate is high, a portion of the pumped water could come<br />

from the other side of the river. Figure 2 shows a case study <strong>for</strong> the <strong>RBF</strong> wells of Meissen-<br />

Siebeneichen, Germany, in which the path lines go past the mid-point of the river and extend to<br />

the other side. Cross-sections A-A and B-B show that only particles starting in the upper layer of<br />

the aquifer actually reach the river. All other particles are turned away near the river, but travel<br />

towards the production borehole. Thus, high nitrate concentrations in the lower layer of the<br />

aquifer between the river and production borehole can be explained by groundwater flow from the<br />

opposite side of the river. Further, the impact of riverbed clogging on the proportion of pumped<br />

river water was investigated using the model. Under site-specific geological conditions, observed<br />

clogging of the river bed (1 × 10 –5 m per second of the 0.1-m thick clogging layer) causes a<br />

decrease in <strong>RBF</strong> by only about 5 percent, as compared to the “no clogging” case. In general,<br />

geological anisotropy was found to have a stronger effect on the proportion of groundwater flow<br />

beneath the riverbed than riverbed clogging. Even under conditions where riverbed clogging does<br />

not occur, groundwater can flow from the opposite side of the river beneath the riverbed towards<br />

the production boreholes (Grischek et al., 2002).<br />

Model Grid<br />

98.0<br />

98.1<br />

98.2<br />

A A<br />

98.3<br />

The Meissen-Siebeneichen example demonstrates that if groundwater on the other side of the river<br />

contained dissolved contaminants, the contaminants could possibly appear in the filtrate. Thus, while<br />

designing <strong>RBF</strong> systems, land use on the other bank of the river should be considered and potential<br />

pollution sources must be identified. The calibration of the flow model heavily depends on the<br />

accuracy of the hydraulic conductivity of riverbed material. The heterogeneity of riverbed material<br />

can produce uneven leakage through the riverbed. Further, hydraulic parameters of the riverbed<br />

material can change depending on the flow status of the river due to scouring and sedimentation<br />

process. To date, there is no reliable and verifiable method available to estimate riverbed hydraulic<br />

conductivity of large rivers in situ and to estimate its variability with the flow regime of the river. An<br />

overview on methods and recent developments is given by Macheleidt et al. (2002).<br />

Single-Species Transport Modeling<br />

To estimate the amount of river water entering a well, mass balance methods are generally<br />

employed. In such methods, the concentrations of a conservative chemical in the river, aquifer,<br />

and well are used to estimate the mass fractions coming from the river and aquifer:<br />

C well = xC river +(1 – x)C aquifer<br />

98.0<br />

98.1<br />

98.2<br />

98.4<br />

Flow Paths (Plan View)<br />

98.3<br />

Adjacent<br />

Granitic<br />

Rocks<br />

98.4<br />

(Equation 1)<br />

98.5<br />

98.5<br />

Elbe River<br />

Legend<br />

Production Well 500 m<br />

–98– Piezometric Contour<br />

in Meters Above Mean Sea Level<br />

98.6<br />

A Flow Path A<br />

(Side View)<br />

Well<br />

Figure 2. Location map of a <strong>RBF</strong> site of the Meissen-Siebeneichen <strong>Water</strong>works in Germany (after Grischek et al., 2002).<br />

71


72<br />

In Equation 1, the fraction of river water (x) can be calculated if the three concentrations are known.<br />

Similarly, if x is known from prior investigations, the concentration of a conservative chemical in<br />

the filtrate may be estimated; however, the equation is only valid <strong>for</strong> steady-state conditions. In other<br />

words, concentrations in the river and aquifer do not change with time. Also, it is assumed that<br />

infiltrating water from the river at a given moment in time is of the same quality as that reaching the<br />

well and that the quality of the bank filtrate and groundwater does not change over the whole<br />

thickness of the aquifer. Such assumptions hamper the use of Equation 1 <strong>for</strong> spill events, even <strong>for</strong><br />

conservative chemicals. Transient effects such as lag time (or travel time) issues are not considered.<br />

Further, Equation 1 cannot be used <strong>for</strong> non-conservative contaminants that are subject to<br />

degradation, sorption, or other reactions. For such contaminants, knowing the concentrations in the<br />

river and aquifer, as well as the mass fraction of the water derived from the river, one would not be<br />

able to predict the concentration at the well. For transient problems, the application of numerical<br />

models is typically required. In those, the advection-dispersion equation is solved with or without<br />

chemical reactions, depending upon the type of contaminants. The advection-dispersion equation<br />

can be solved analytically if the flow field is steady and reaction terms are linear. In cases where the<br />

use of three-dimensional numerical models is beyond the ability of water utilities, simple<br />

one-dimensional analytical models can be used to estimate chemical concentrations in pumped<br />

water, especially from shock events or chemical spills. Mälzer et al. (2003) used the analytical<br />

solution to the one-dimensional advection-dispersion equation to estimate the concentration of a<br />

chemical reaching a well. In their approach, they approximated the flow field between the river and<br />

aquifer using five flow tubes. For each tube, they solved the advection-dispersion equation with<br />

(sorption and first order degradation) to calculate the concentration at Observation Well M1<br />

(Figure 3); however, the assumption of steady flow through these flow tubes may not be valid <strong>for</strong> all<br />

cases. Further, estimating the concentrations <strong>for</strong> a contaminant at the pumping well may not be easy<br />

without knowing the amount of groundwater contribution to the well.<br />

Meters Above Sea Level<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

Vertical Section at Wittlaer<br />

Rhine River<br />

Aquifer<br />

For transient three-dimensional simulations, MT3D/MT3DMS (Zheng and Wang, 1999) is<br />

probably the most common transport simulator that is used in conjunction with MODFLOW.<br />

MT3D uses the same grid of MODFLOW and solves advection-dispersion equation, along with<br />

sorption and degradation reactions, to estimate the concentration of a contaminant at a given<br />

location in the model domain. For estimating the concentration at the pumping well, a flow model<br />

must be built and calibrated to provide a good description of the velocity field. As mentioned<br />

earlier, the hydraulic conductivity of bed material at the river-aquifer interface can significantly<br />

23 m<br />

27 m<br />

Figure 3. Vertical section of the aquifer at Wittlaer between the Rhine River<br />

and Observation Well M1 and a simplified model assumption of five<br />

parallel tubes (after Mälzer et al., 2003).<br />

M1<br />

High<br />

Low


affect the amount of contaminant entering from the river to the aquifer and, eventually, to the<br />

pumping well(s). For flow fields affected by pumping wells, the assumption of equilibrium sorption<br />

(in contrast to kinetically controlled sorption) near the screen zone may not be valid due to fast<br />

travel rates of water particles. Ray et al. (2002) simulated the impact of riverbed/bank material<br />

properties on filtrate quality during a flood pass-through event at a <strong>RBF</strong> facility on the banks of the<br />

Illinois River. They used equilibrium sorption and first order degradation from literature-reported<br />

data. Also, in certain simulations, they considered the contaminant to be conservative. This<br />

assumption enveloped the two extremes. For atrazine, with a highly conductive bank, the<br />

concentrations in groundwater and the pumped well are slightly attenuated when the effects of<br />

sorption and degradation are considered (Figure 4A); however, without sorption and degradation,<br />

the concentrations can be close to that found in river water (Figure 4B).<br />

Atrazine Concentration, µg/L<br />

Atrazine Concentration, µg/L<br />

4<br />

3<br />

2<br />

1<br />

0<br />

4<br />

3<br />

2<br />

1<br />

0<br />

(A) with sorption and decay<br />

River water<br />

SP-1<br />

Caisson<br />

SP-3 is below detection limit<br />

0 20 40 60 80 100<br />

River water<br />

Days Since Start of Simulation<br />

(B) without sorption and decay<br />

SP-3<br />

SP-1<br />

Caisson<br />

0 20 40 60 80 100<br />

Days Since Start of Simulation<br />

Figure 4. Atrazine transport from the river to the aquifer and the pumping well<br />

with and without sorption and decay reactions <strong>for</strong> a highly conductive bed<br />

and bank (after Ray et al., 2002).<br />

73


74<br />

When the hydraulic conductivity of the bed and bank are reduced to moderate values (see sets<br />

A-A and B-B in Ray et al., 2002), there is a significant attenuation of atrazine at the collector well<br />

caisson (Figure 5). It is clear that the transient effects call <strong>for</strong> the use of complex three-dimensional<br />

models to account <strong>for</strong> the spatial and temporal variability of streamlines. Based on hydraulic head<br />

measurements and water-quality observations, the calibration of flow and transport models is<br />

essential <strong>for</strong> an accurate description and prediction of the fate of individual or multiple chemicals.<br />

The current version of the standard MT3DMS code (Zheng and Wang, 1999) accounts <strong>for</strong> the<br />

transport of multiple species and a small range of basic reactions; however, <strong>for</strong> cases where<br />

chemical species are assumed to interact in a more complex manner, MT3DMS-based packages<br />

such as RT3D (Clement, 1997) or PHT3D (Prommer et al., 2003a), which account <strong>for</strong> a greater<br />

variety of biogeochemical processes, are available.<br />

Atrazine Concentration, µg/L<br />

4<br />

3<br />

2<br />

1<br />

0<br />

River water<br />

Caisson (set B-B)<br />

Caisson (set A-A)<br />

Multi-Species and Multi-Component Transport Modeling<br />

No reaction – filtrate<br />

With reaction – filtrate<br />

River water<br />

0 20 40 60 80 100<br />

Days Since Start of Simulation<br />

Figure 5. Atrazine transport from the river to the aquifer and the pumping well <strong>for</strong> a case<br />

with low permeability riverbed/bank (after Ray et al., 2002).<br />

In those scenarios where the reaction progress of a dissolved chemical strongly depends on<br />

the concentration of one or more other dissolved species, reactive multi-species models can<br />

typically provide a better process description. This is particularly important if models are used in a<br />

predictive mode. Furthermore, if water-quality changes are additionally affected by water-sediment<br />

interactions, such as mineral dissolution/precipitation and/or ion-exchange reactions, a reactive<br />

multi-component transport model might need to be applied to explain specific field observations.<br />

Multi-component models typically use an extensive reaction database. Until recently, those<br />

databases were (in most cases) confined to the definition of thermodynamic equilibrium reactions,<br />

which made those models only applicable to systems where (all) reactions proceed relatively fast in<br />

relation to groundwater flow velocity (local equilibrium assumption); however, in most of the<br />

recently published models, a wide range of different kinetic reactions and processes can be defined.<br />

Models with those capabilities can be used to study complex process interactions that lead to non-


intuitive system behavior. In the case of <strong>RBF</strong>, such examples might be:<br />

• The reductive dissolution of iron oxides by DOC and the resulting release of sorbed<br />

heavy metals.<br />

• The assessment of pesticide mobility during flood events.<br />

Denitrification, degradation of pesticides, and dissolution of minerals are closely linked to DOC,<br />

a key component of river water whose concentration can vary depending on the season and flow<br />

events. Because of the reaeration process, river water is oxygenated compared to groundwater.<br />

During the induced infiltration process, oxygen-rich river water enters the aquifer, as does DOC.<br />

On its travel path, DOC from river water is oxidized and either mineralizes completely or is<br />

trans<strong>for</strong>med to intermediates through bacterial catalysis, together with the organic carbon that is<br />

perhaps naturally abundant in the aquifer as sediment-bound organic matter. Oxygen in<br />

the invading water is used as an electron acceptor in the process. Normally, there is sufficient<br />

carbon available <strong>for</strong> microbial use; however, oxygen can become in short supply along the flow<br />

path. Once microbes consume the oxygen, an anoxic zone develops where the nitrate of the<br />

infiltrating river water and groundwater is used as a substitute electron acceptor. This leads to the<br />

reduction of nitrate along the flow path. Once nitrate is also depleted, thermodynamically less<br />

favorable oxidized iron and manganese minerals and/or sulfate might act as alternative electron<br />

acceptors. Note that the simulation of the oxidation of one or multiple organic substrates using a<br />

sequence of electron acceptors is routinely applied in the field of bioremediation modeling, mainly<br />

where the transport and natural attenuation of oxidizable organic contaminants is simulated<br />

(Barry et al., 2002).<br />

The simultaneous simulation of <strong>RBF</strong>-typical denitrification and mineral dissolution reactions can<br />

be handled by a number of existing codes (<strong>for</strong> examples, see Table 1). One such code is EASY-<br />

LEACHER (Stuyfzand and Lüers, 2000), which was used to simulate reactions along a transect of<br />

the Torgau <strong>RBF</strong> site on the Elbe River in Germany. EASY-LEACHER is a two-dimensional<br />

reactive transport code in EXCEL spreadsheet, combining chemical principles with empirical<br />

rules in an expert system. The code was found useful to attain a first estimate of water quality in<br />

the production well <strong>for</strong> a <strong>RBF</strong> site where the operation of wells is started and <strong>for</strong> an easy<br />

calculation of different boundary conditions. In contrast to EASY LEACHER, which uses a<br />

collection of (one-dimensional) flow tubes to account <strong>for</strong> the transport of chemicals, the<br />

FEREACT model (see Tebes-Stevens et al., 1998; Tebes-Stevens and Valocchi, 2000) is a<br />

Table 1. Examples of Reactive Multi-Component Transport Models<br />

Model Reference<br />

CRUNCH Steefel (2001)<br />

EASY-LEACHER Stuyfzand and Lüers (2000)<br />

FEREACT Tebes-Stevens et al. (1998)<br />

PHT3D Prommer et al. (2003a)<br />

PHAST Parkhurst et al. (1995)<br />

MIN3P Mayer (1999)<br />

TBC Schäfer et al. (1998)<br />

HBGC123D Salvage and Yeh (1998)<br />

75


76<br />

two-dimensional, finite element-based transport model that can simulate biodegradation and<br />

geochemical reactions along, <strong>for</strong> instance, a vertical transect of the river-aquifer interface. Although<br />

the model cannot handle the true three-dimensional (and perhaps transient) flow dynamics<br />

experienced at a <strong>RBF</strong> site, it accounts <strong>for</strong> all typical geochemical and microbial reactive processes.<br />

An example <strong>for</strong> a fully three-dimensional model is the MODFLOW/MT3DMS-based code<br />

PHT3D (Prommer, 2002, Prommer et al., 2003a). It combines the previously mentioned<br />

MT3DMS (Zheng and Wang, 1999) with PHREEQC-2 (Parkhurst and Appelo, 1999), whereby<br />

the <strong>for</strong>mer solves the advection-dispersion equation and the latter accounts <strong>for</strong> all geochemical<br />

reactions. Prommer et al. (2003b) recently applied the model to simulate the transport and<br />

reactive processes that affect the fate of atrazine near a <strong>RBF</strong> scheme during a flood event.<br />

For that modeling study, the hydrogeological setting and hydrological characteristics described by<br />

Ray et al. (2002) were used to demonstrate the potential influence of a microbial lag effect and<br />

the redox-dependency of atrazine degradation on abstraction water quality during the flood event.<br />

They considered kinetically controlled mineralization of DOC, dissolution of sediment-bound<br />

organic matter, growth and decay of atrazine degraders, and microbially mediated atrazine<br />

degradation. During the flood event, DOC from the river water modifies the redox patterns in the<br />

aquifer along the flow path to the well screens. The simulations demonstrated that with the flood,<br />

the concentration of atrazine reaches a peak in a similar pattern to that found in floodwater and<br />

the concentration in groundwater remains high; however, once atrazine degraders are active, the<br />

concentration drops significantly (Figure 6, Case 1). On the other hand, if the DOC of the river<br />

water is somewhat increased, this leads to the <strong>for</strong>mation of an anoxic zone, which promotes<br />

denitrification, but inhibits atrazine degradation (Figure 6, Case 2).<br />

Head (m)<br />

142<br />

140<br />

138<br />

136<br />

134<br />

<strong>Water</strong> level<br />

Concluding Remarks<br />

River<br />

Obs. Pt.<br />

132<br />

0 50<br />

Time (days)<br />

100<br />

C (mol/l)<br />

C (mol/l)<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

x 104<br />

5<br />

4<br />

3<br />

2<br />

1<br />

x 107<br />

1<br />

Oxygen<br />

reactive case 1<br />

reactive case 2<br />

0<br />

0 50 100<br />

Atrazine<br />

conservative<br />

reactive case 1<br />

reactive case 2<br />

0<br />

0 50<br />

Time (days)<br />

100<br />

0<br />

0 50<br />

Time (days)<br />

100<br />

It is apparent that the level of modeling can range from a very simple mass balance to very<br />

complex studies that involve transient flow and biogeochemical reactions. The objective of the<br />

modeling study, as well as the type, quantity, and quality of data, will dictate what level of<br />

sophistication is needed. For example, if yield determination is the primary purpose, MODFLOW<br />

simulations will mostly be adequate. Preliminary investigations to delineate the sources of water<br />

1.5<br />

1<br />

0.5<br />

x Nitrate<br />

104<br />

8<br />

6<br />

4<br />

2<br />

0<br />

0 50 100<br />

x 108<br />

2<br />

reactive case 1<br />

reactive case 2<br />

Atrazine degraders<br />

reactive case 1<br />

reactive case 2<br />

Figure 6. Schematic representation of the 3-D model, hydraulic heads, and selected simulated concentrations<br />

of oxygen, nitrate, atrazine, and atrazine degraders.


entering the well can be carried out with MODPATH or comparable particle tracking tools;<br />

however, the accuracy of the simulation strongly depends on how well the true values of hydraulic<br />

conductivity of riverbed/bank materials are represented in the models. Especially <strong>for</strong> transient<br />

simulations, accurate estimates of riverbed and bank hydraulic conductivity are critical. For<br />

transport simulations, simple mixing models (as used by many utilities) may not provide good<br />

estimates of contaminant concentrations at <strong>RBF</strong> wells due to lag times involved, transient flow,<br />

and heterogeneity issues. Models using MT3D <strong>for</strong> simulating the transport of single or multiple<br />

contaminants may be adequate <strong>for</strong> specific problems (<strong>for</strong> example, worst-case estimates); however,<br />

the assumption of equilibrium reactions near well screens may not be valid. On the other hand,<br />

<strong>for</strong> kinetic reactions, the availability of appropriate rate parameters might not always be<br />

warranted. Finally, biogeochemical models such as PHT3D are most comprehensive and are ideal<br />

to study both steady and transient effects; however, the input data sets <strong>for</strong> such models can be<br />

extensive, and field data sets that underpin the simulations may not always be readily available.<br />

For such complex models, input parameters are rarely found in literature and are very much sitespecific.<br />

On the other hand, the determination of such parameters at the laboratory-scale requires<br />

both significant time and ef<strong>for</strong>t and, additionally, the usefulness of the results <strong>for</strong> field-scale<br />

simulations might still not be warranted due to scale-issues. Furthermore, the application of such<br />

models typically requires users with advanced modeling skills and experience in the areas of<br />

geochemistry and (groundwater) hydrology. After all, the complexity found in some of the models<br />

is only a reflection of the complexity and diversity found in nature and natural materials. To find<br />

and make use of adequate simplifications should be part of the development of the (site-specific)<br />

conceptual model. Each complex numerical model can also be used as a “simpler” model by fixing<br />

those parameters that were excluded in the corresponding simple model. Also, it must be noted<br />

that the parameters used in simpler models are by far not less site-specific than those of more<br />

complex models.<br />

A typical example of the “how much complexity is adequate” question is DOC, since it consists<br />

of thousands of single organic compounds with different biodegradation behavior. The question is<br />

how many different fractions (concerning biodegradability and sorption behavior) should be<br />

identified and used in the model to be adequate <strong>for</strong> the selected level of model discretization,<br />

reactions included, and kinetic parameters. Based on the application of different models <strong>for</strong><br />

chosen field-sites and qualified sensitivity analyses (benchmarking), a guideline should be<br />

developed to propose the right type of model <strong>for</strong> a specific task. The development of reduced<br />

models, such as the spreadsheet program EASY-LEACHER (Stuyfzand and Lüers, 2000), and the<br />

improvement of multi-species reaction models should be carried out simultaneously.<br />

Furthermore, <strong>for</strong> detailed modeling, the hydraulic component must be able to handle scour,<br />

deposition, high water levels, flooding, etc. For detailed transport and reaction modeling, varied<br />

properties of the riverbed material and aquifer material (e.g., organic carbon content, enzymatic<br />

activity) should be considered.<br />

Finally, we must mention that besides the widely known models mentioned in Table 1, there are<br />

many other models that have been applied successfully <strong>for</strong> the simulation of processes at <strong>RBF</strong> sites<br />

<strong>for</strong> the specific evaluation purposed.<br />

77


78<br />

REFERENCES<br />

Barry, D.A., H. Prommer, C.T. Miller, P. Engesgaard, and C. Zheng (2002). “Modeling the fate of oxidisable<br />

organic contaminants in groundwater.” Adv. <strong>Water</strong> Resour., 25: 945-983.<br />

Clement, T.P. (1997). RT3D - A Modular Computer Code <strong>for</strong> Simulating Reactive Multi-species Transport in 3-<br />

Dimensional Groundwater Aquifers, Battelle Pacific Northwest <strong>National</strong> Laboratory <strong>Research</strong> Report, PNNL-<br />

SA-28967.<br />

Conrad, L.P., and M.S. Beljin (1996). “Evaluation of an induced infiltration model as applied to glacial<br />

aquifer systems.” Wat. Resour. Bull., 32(6): 1,209-1,220.<br />

Dillon, P.J., and J.A. Liggett (1983). “An ephemeral stream-aquifer interaction model.” Wat. Resour. Res.,<br />

19(3): 621-626.<br />

Ferris, J.G., D.B. Knowles, R.H. Brown, and R.W. Stallmann (1962). Theory of Aquifer Tests, U.S. Geological<br />

Survey <strong>Water</strong> Supply Paper 1536-E, Government Printing Office, Washington, D.C.<br />

Grischek, T., D. Schoenheinz, and W. Nestler (2002). “Unexpected groundwater flow beneath the River Elbe at the<br />

bank filtration site Meissen, Germany.” <strong>Water</strong> resources and environment research, ICWRER 2002, Vol. 1: 116-120.<br />

Hantush, M.S. (1959). “Analysis of data from pumping wells near a river.” Jour. Geophys. Res., 64(11): 1,921-1,931.<br />

Harbaugh, A.W., E.R. Banta, M.C. Hill, and M.G. McDonald (2000). MODFLOW-2000, the U.S. Geological<br />

Survey modular ground-water model — User guide to modularization concepts and the Ground-<strong>Water</strong> Flow Process,<br />

U.S. Geological Survey Open-File Report 00-92, 121 p.<br />

Macheleidt, W., W. Nestler, and T. Grischek (2002). “Determination of hydraulic boundary conditions <strong>for</strong><br />

the interaction between surface water and groundwater.” Sustainable groundwater development, Geological<br />

Society, London, Special Publ. 193: 235-243.<br />

Mälzer, H., J. Schubert, R. Gimbel, and C. Ray (2003). “Effectiveness of riverbank filtration sites to mitigate<br />

shock loads.” Riverbank Filtration: Improving Source <strong>Water</strong> Quality, Kluwer Academic Publishers, Dordrecht,<br />

The Netherlands.<br />

Mayer, K.U. (1999). A numerical model <strong>for</strong> multicomponent reactive transport in variably saturated porous media,<br />

Ph.D. thesis, University of <strong>Water</strong>loo, <strong>Water</strong>loo, Ontario, Canada.<br />

Parkhurst, D.L., P. Engesgaard, and K.L. Kipp (1995). “Coupling the geochemical model PHREEQC with a<br />

3D multi-component solute transport model.” Fifth Annual V.M. Goldschmidt Conference, Penn State<br />

University, University Park Pennsylvania, USA, May 1995.<br />

Parkhurst, D.L., and C.A.J. Appelo (1999). User’s guide to PHREEQC - A computer program <strong>for</strong> speciation,<br />

reaction-path, 1D-transport, and inverse geochemical calculations: Technical Report 99-4259, U.S. Geological<br />

Survey <strong>Water</strong>-Resources Investigations Report.<br />

Pollock, D.W. (1994). User’s Guide <strong>for</strong> MODPATH/MODPATH-PLOT, Version 3: A particle tracking postprocessing<br />

package <strong>for</strong> MODFLOW, the U.S. Geological Survey finite-difference ground-water flow model,<br />

U.S. Geological Survey Open-File Report 94-464.<br />

Prommer, H. (2002). PHT3D: A Reactive Multicomponent Transport Model <strong>for</strong> Saturated Media, User’s Manual<br />

Version 1.0, Contaminated Land Assessment and Remediation <strong>Research</strong> Centre, The University of Edinburgh, UK,<br />

http://www.pht3d.org.<br />

Prommer, H., D.A. Barry, and C. Zheng (2003a). “PHT3D - A MODFLOW/MT3DMS based reactive multicomponent<br />

transport model.” Ground <strong>Water</strong>, 42(2): 247-257.<br />

Prommer, H., J. Greskowiak, P.J. Stuyfzand, and C. Ray (2003b). “Geochemical transport modeling of water<br />

quality changes during managed artificial recharge.” MODFLOW and More 2003: Understanding through<br />

Modeling, Proceedings of the International Ground <strong>Water</strong> Modeling Conference, Golden, Colorado USA,<br />

16-19 September 2003.


Ray, C., T.W. Soong, Y.Q. Lian, and G.S. Roadcap (2002). “Effect of flood-induced chemical load on filtrate<br />

quality at bank filtration sites.” J. Hydrol., 266: 235-258.<br />

Salvage, K.M., and G.T. Yeh (1998). “Development and application of a numerical model of kinetic and<br />

equilibrium microbiological and geochemical reactions (BIOKEMOD).” J. Hydrol., 209(1-4): 27-52.<br />

Schäfer, D., W. Schäfer, and W. Kinzelbach (1998). “Simulation of processes related to biodegradation of<br />

aquifers 1. structure of the 3D transport model.” J. Contam. Hydrol., 31(1-2): 167-186.<br />

Steefel, C.I. (2001). GIMRT, Version 1.2: Software <strong>for</strong> Modeling Multicomponent, Multidimensional Reactive<br />

Transport. Users Guide, Technical Report UCRL-MA-143182, Lawrence Livermore <strong>National</strong> Laboratory,<br />

Livermore, Cali<strong>for</strong>nia, 2001.<br />

Stuyfzand, P.J., and F. Lüers (2000). “Modelling the quality changes upon artificial recharge and bank<br />

infiltration.” Principles and user`s guide of EASY-LEACHER 4.5, KIWA-report SWI 99.199.<br />

Tebes-Stevens, C.L., and A.J. Valocchi (2000). “Calculation of reaction parameter sensitivity coefficients in<br />

multicomponent subsurface transport models.” Adv. <strong>Water</strong> Resour., 23: 591-611.<br />

Tebes-Stevens, C.L., A.J. Valocchi, J.M. VanBriesen, and B.E. Rittmann (1998). “Multicomponent transport<br />

with coupled geochemical and microbiological reactions: Model description and example simulations.”<br />

J. Hydrol., 209: 8-26.<br />

United States Environmental Protection Agency (2002). Long-Term 1 Enhanced Surface <strong>Water</strong> Treatment<br />

Rule, Final Rule. Federal Register, 67:9:1812 (January 14, 2002).<br />

van Breukelen, B.M., C.A.J. Appelo, and T.N. Oltshoorn (1998). “Hydrogeochemical transport modeling of<br />

24 years of Rhine water infiltration in the dunes of the Amsterdam water supply.” J. Hydrol., 209: 281-296.<br />

Wilson, J.L. (1993). “Induced infiltration in aquifers with ambient flow.” Wat. Resour. Res., 29(10): 3,503-3,512.<br />

Zheng, C., and P.P. Wang (1999). MT3DMS: A modular three-dimensional multispecies model <strong>for</strong> simulation of<br />

advection, dispersion and chemical reactions of contaminants in groundwater systems, Documentation and User’s<br />

Guide, Contract Report SERDP-99-1, U.S. Army Engineer <strong>Research</strong> and Development Center, Vicksburg, MS.<br />

CHITTARANJAN RAY is an Associate Professor in the Department of Civil &<br />

Environmental Engineering and an Associate <strong>Research</strong>er with the <strong>Water</strong> Resources<br />

<strong>Research</strong> Center at the University of Hawaii at Mañoa. His current research interests<br />

include riverbank filtration, pesticides in drinking-water wells, and the flow and transport<br />

of pathogens and chemicals in saturated/unsaturated media. Over the past 2 years, he has<br />

edited two books on riverbank filtration and written a monograph on pesticides in<br />

domestic wells. Among his honors, he is a recipient of the Fulbright faculty scholarship <strong>for</strong><br />

conducting riverbank filtration-related research in Nepal, India, and Bangladesh. Prior to joining the<br />

University, Ray worked as a staff engineer with the groundwater consulting firm of Geraghty & Miller, Inc.<br />

(now called Arcadis Geraghty & Miller) and in the Groundwater Section of Illinois State <strong>Water</strong> Survey,<br />

conducting various groundwater quantity and quality investigations, including bank filtration. Ray received<br />

a B.S. in Agricultural Engineering from Orissa University of Agriculture and Technology in India, an M.S.<br />

in Agricultural Engineering from the University of Manitoba in Canada, an M.S. in Civil Engineering from<br />

Texas Tech University, and Ph.D. Civil and Environmental Engineering from the University of Illinois at<br />

Urbana-Champaign.<br />

79


Session 5: Dynamics<br />

The 100-Year Flood of the Elbe River in 2002<br />

and Its Effects on Riverbank-Filtration Sites<br />

Dipl.-Ing. Matthias Krueger<br />

Fernwasserversorgung Elbaue-Ostharz GmbH<br />

Torgau, Germany<br />

Dipl.-Ing. Ingbert Nitzsche<br />

Fernwasserversorgung Elbaue-Ostharz GmbH<br />

Torgau, Germany<br />

Introduction<br />

To historians, Torgau, Germany, is mainly known as the <strong>for</strong>mer residence of the Saxon Electors<br />

and the cultural capital of sixteenth-century Saxony (Figure 1). A short time later, Martin Luther’s<br />

work gave the town on the Elbe the nickname, “Nursemaid of the Re<strong>for</strong>mation.” Finally, many<br />

people know Torgau as the historic site where the defeat of Nazi Germany was symbolized by the<br />

meeting of Soviet and American troops on the Elbe River on April 25, 1945.<br />

Figure 1. Air view of Torgau, Germany, 2 days be<strong>for</strong>e the flood peak.<br />

Torgau is the headquarters of the Fernwasserversorgung Elbaue-Ostharz GmbH, one of the largest<br />

drinking-water supply companies in Germany. The company was founded about 50 years ago and<br />

manages one waterworks at a reservoir in the Harz Mountains, five bank-filtration waterworks in the<br />

Elbe River Basin near Torgau, and a 700-kilometer long distribution network. In 2002, the<br />

drinking-water production was 76-cubic meters. About 3.5-million people were supplied with<br />

drinking water of high quality.<br />

Correspondence should be addressed to:<br />

Dipl.-Ing. Matthias Krueger<br />

Head of Laboratory<br />

Fernwasserversorgung Elbaue-Ostharz GmbH<br />

Naundorferstraße 46 • 04860 Torgau, Germany<br />

Phone: +49 3421 757511 • Fax: +49 3421 757522 • Email: matthias.krueger@fwv-torgau.de<br />

81


82<br />

Riverbank Filtration Near Torgau<br />

In this system, raw-water abstraction via bank filtration in the Elbe River Basin is a key element.<br />

The production boreholes are located in an alluvial sand and gravel aquifer with a thickness of<br />

40 to 60 m, covered by a 2- to 5-m thick layer of meadow loam. The meadow loam provides an<br />

important protection against pollution and infiltration during flooding. Due to erosive conditions,<br />

there is only low clogging of the riverbed. The good hydraulic connection between the Elbe River<br />

and the adjacent aquifer ensures stable water abstraction during low flow periods.<br />

Figure 2 shows the biggest waterworks, Torgau-Ost, which has a maximum capacity of<br />

120,000 m 3 /d. Raw water is abstracted from 42 production boreholes on the western bank of the<br />

Elbe River. At a distance of about 300 m between the production boreholes and the riverbank,<br />

the flow time of the bank filtrate is between 60 and 200-plus days.<br />

Figure 2. Torgau-Ost <strong>Water</strong>works and the Elbe River.<br />

The produced water has a good quality due to the natural purification ability of the aquifer. Two<br />

monitoring cross-sections positioned along assumed flowpaths to the production boreholes have<br />

been installed to control borehole operation and to guarantee flexibility in response to environmental<br />

impacts, both seasonal and long-term. These monitoring cross-sections include up to five<br />

observation stations between the production boreholes and the river. Each observation station<br />

consists of three to five monitoring points at varying depths.<br />

Raw water is purified by aeration and deacidification, pre-purification by 20 tube sedimentation<br />

basins, and fine purification by 16 open sand filters. Iron, manganese, and carbon dioxide are the<br />

main constituents of raw water that require treatment. These constituents are eliminated by<br />

adding hydrated lime and potassium permanganate. Prior to pumping into the public-water supply<br />

network, the water is further treated with small amounts of chlorine and chlorine dioxide to<br />

prevent bacteriological deterioration during the long transportation process to the consumer.<br />

The 100-Year Flood of the Elbe River in 2002<br />

In Torgau, the mean discharge of the Elbe River is about 330 cubic meters per second (m 3 /s). In<br />

August 2002, a so-called “Flood of the Century” occurred. The event was caused by a “Vb” (extreme)<br />

weather situation. Precipitation reached 300 millimeters per day, with 25 to 30 millimeters per hour<br />

in the Ore Mountains. This intensity had never been measured be<strong>for</strong>e in this region. In Dresden,<br />

the Elbe River exceeded the 1845 flood level of 8.77 m, hitting a record of 9.40 m on August 17,<br />

2002. The towns of Meissen, Torgau, Wittenberg, and Dessau were also partially flooded.


In Torgau, the water level of the Elbe River increased within 10 days by about 8 m up to 9.47 m<br />

on August 19, a level that has never been observed be<strong>for</strong>e. The peak discharge of the Elbe River<br />

in Torgau was estimated to have been 4,295 m 3 /s. The City of Torgau had to be evacuated<br />

completely. Technical staff, soldiers, and helpers built anti-flood barricades with thousands of<br />

sandbags.<br />

To ensure a safe drinking-water production, the water company <strong>for</strong>med an emergency task <strong>for</strong>ce<br />

on August 13 to organize necessary measures.<br />

The Effects of the Flood<br />

Most bank-filtration waterworks near Torgau are located at elevated levels at a sufficient distance<br />

to the river and are normally not affected by flooding. Nearly all production boreholes were flooded,<br />

but still functioned due to their special construction. A problem arose with the power-supply system<br />

<strong>for</strong> the boreholes. Trans<strong>for</strong>mer stations are located behind the dikes. But, due to the high water<br />

level and risk of dikes breaking, the stations had to be extra protected; there<strong>for</strong>e, the stations were<br />

housed-in using sheet piling and equipped with pumps <strong>for</strong> dewatering and additional emergency<br />

power generators (Figure 3).<br />

Figure 3. Protected trans<strong>for</strong>mer station behind a dike.<br />

As a result of these measures, the power supply was maintained at most locations. Only two small<br />

waterworks in the north of Torgau had to be abandoned due to water influx at trans<strong>for</strong>mer stations.<br />

Besides the aspect of water quantity, there is water quality. As a result of <strong>for</strong>mer bank-filtration<br />

research programs at the Torgau site, there is an excellent knowledge of flow processes and<br />

water-quality changes along the flowpath. Based on this knowledge, no significant change in<br />

raw-water quality was expected, mainly because the flow time of the bank filtrate ranges between<br />

60 and 300-plus days. For control purposes, the continuous monitoring of raw-water quality was<br />

intensified during the flood, and a special groundwater-monitoring program <strong>for</strong> the cross-sections<br />

was set up after the flood.<br />

All results of measurements at the outflow of the waterworks proved that drinking-water quality<br />

was not at risk at any time. The quality of the production water met the German standards <strong>for</strong><br />

83


84<br />

drinking-water quality. Drinking-water disinfection was done by 0.2 mg/L chlorine dioxide (ClO 2)<br />

and 0.6 mg/L chlorine (Cl 2). There was no problem with bacterial contamination.<br />

But what about raw-water quality? The flood caused a strong increase in turbidity, organic carbon<br />

concentration, and number of microorganisms in river water. The increase in DOC from 5 mg/L<br />

to more than 10 mg/L in river water did not affect the DOC concentration in the raw water<br />

(Figure 4). The increase in DOC in river water was mainly caused by an increase of the<br />

biodegradable fraction of DOC, which is consumed along the flowpath of the bank filtrate. Thus,<br />

biodegradation and mixing in the aquifer prevented an increase in DOC concentration in the raw<br />

water and an increase in DBPs.<br />

40<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

Increases in concentrations of organic trace compounds such as polynuclear aromatic hydrocarbon,<br />

chlor-organics, and pesticides were not observed in Elbe River water during the flood. Despite<br />

many reported inputs of contaminants, the huge discharge caused an effective dilution and<br />

minimized the risk. After the flood, concentrations were found to be around the mean annual<br />

values. Despite these facts, measurements of the sum of adsorbable organic halogenated<br />

compounds (AOX) were used to check <strong>for</strong> the potential contamination of river water and raw<br />

water. The AOX concentration in the raw water remained at a low level of less than 30 µg/L,<br />

giving no indication of contamination (see Figure 4). Measured DOC and AOX concentrations<br />

in bank-filtrate samples were found to be within long-term concentrations ranges.<br />

Increased heavy metal concentrations in the river water also did not affect raw-water quality.<br />

Conclusions<br />

5<br />

0<br />

Elbe River<br />

4/29/02<br />

Elbe River<br />

8/20/02<br />

Borehole 22<br />

9/4/02<br />

Raw water<br />

10/7/02<br />

AOX (µg/l)<br />

DOC (mg/l)<br />

Figure 4. DOC and AOX concentrations in Elbe River water and raw water<br />

The 100-year flood of the River Elbe in August 2002 became the greatest challenge <strong>for</strong><br />

Fernwasserversorgung Elbaue-Ostharz GmbH in its history. Maintaining the power supply <strong>for</strong><br />

pumps in the production boreholes was the most important factor, whereas the quality of<br />

abstracted bank filtrate was never at risk due to the design of bank-filtration sites. Based on the<br />

measures taken and management of efficient distribution network, the water company ensured a


stable water supply <strong>for</strong> millions of people during the flood without restrictions to quantity and<br />

quality. The available technology <strong>for</strong> drinking-water treatment was successful in producing<br />

high-quality drinking water.<br />

Since 1990, MATTHIAS KRUEGER — a process engineer and chemist — has worked <strong>for</strong><br />

the Fernwasserversorgung Elbaue-Ostharz GmbH, a water company that provides drinking<br />

water <strong>for</strong> over 3.5-million people in Germany. Early on, he was responsible <strong>for</strong> water<br />

treatment technology, then became Head of Laboratory in 2001. Under his leadership, the<br />

Laboratory has recently been involved in large research projects on physical and chemical<br />

processes during bank filtration. Prior to joining Fernwasserversorgung Elbaue-Ostharz<br />

GmbH, Krueger was an Engineer at Galvanotechnic in Leipzig, Germany, <strong>for</strong> 7 years. His<br />

areas of interest include flow path, residence times, and redox conditions in the behavior of pollutants during<br />

riverbank filtration. Krueger received a diploma (Dipl.-Ing.) in Chemistry from the Technical University<br />

Ilmenau and completed post-graduated studies in analytics and spectroscopy at Leipzig University.<br />

85


Session 5: Dynamics<br />

Temporal Changes of Natural Attenuation Processes<br />

During Bank Filtration<br />

Paul Eckert, Ph.D.<br />

Stadtwerke Düsseldorf AG<br />

Düsseldorf, Germany<br />

Rudolf Irmscher, Ph.D.<br />

Stadtwerke Düsseldorf AG<br />

Düsseldorf, Germany<br />

<strong>RBF</strong> is a well-proven natural purification step in water supply. Sustainable water management<br />

should be based specifically on natural purification methods, as declared by the International<br />

Association of <strong>Water</strong>works in the Rhine Catchment Area, in their 2003 memorandum. To<br />

achieve this aim, a profound knowledge of the purification capacity of bank filtration is essential.<br />

At the Düsseldorf <strong>Water</strong>works in Germany, the influence of long-term — as well as periodic —<br />

changes of both hydraulics and river-water quality on natural attenuation processes were<br />

investigated.<br />

The improvement of Rhine River water quality over the last 30 years enabled the Düsseldorf<br />

<strong>Water</strong>works to reduce their technical treatment expenses. Temperature variations throughout the<br />

year and flood events significantly influenced the purification capacity of bank filtration. This<br />

rein<strong>for</strong>ces the need <strong>for</strong> flexible technical treatment methods capable of adapting to changing rawwater<br />

quality.<br />

Even though the complete replacement of subsequent technical treatment st<strong>eps</strong> might be seen as<br />

an unreachable vision, the substantial knowledge we have acquired on the purification capacity<br />

of bank filtration enables the design of tailor-made treatment methods.<br />

Site Description and Treatment Concept<br />

The City of Düsseldorf is situated in northwestern Germany, in the lower Rhine Valley (Figure 1).<br />

The Düsseldorf <strong>Water</strong>works supply 600,000 inhabitants with treated bank filtrate. A multiprotective<br />

barrier concept ensures the constant production of high-quality drinking water (Figure 2).<br />

Natural attenuation processes during bank filtration <strong>for</strong>m the first and most efficient protective<br />

barrier. The subsequent protective barrier is raw-water treatment, including ozonation, biological<br />

active filtration, and active carbon adsorption.<br />

The Rhine River has a length of 1,320 kilometers and a catchment area of 185,000 square kilometers;<br />

it is the third biggest river and the largest source of drinking water in Europe. The mean discharge<br />

of the Rhine at Düsseldorf is 2,200 m 3 /s, while the <strong>Water</strong>works use less than 2 m 3 /s. During flood<br />

events, discharge increases up to 9,900 m 3 /s.<br />

Correspondence should be addressed to:<br />

Paul Eckert, Ph.D.<br />

Head of the <strong>Water</strong> Management Department<br />

Stadtwerke Düsseldorf AG<br />

Abt. Wasserwirtschaft • Höherweg 100 • 40233 Düsseldorf, Germany<br />

Phone: +0211/8218359 • Fax: +0211/821778359 • Email: peckert@swd-ag.de<br />

87


88<br />

Protection Zones<br />

II<br />

IIIa<br />

IIIb<br />

Local Frontier<br />

0 1 2 3 4 5km<br />

Rhine<br />

Krefeld<br />

The production wells, situated at a distance of 50 to 350 m of the riverbank, discharge water from<br />

a sandy and gravel aquifer. Depending on the hydraulic situation, the residence time of bank<br />

filtrate in the aquifer varies between weeks to several months.<br />

From a water-resources perspective, bank filtration is characterized by an improvement in water<br />

quality (Kühn and Müller, 2000). The most important effects of bank filtration include:<br />

• Removal of particles and turbidity.<br />

Duisburg<br />

Neuss<br />

Figure 1. Geographic location of the Düsseldorf <strong>Water</strong>works.<br />

Corg<br />

Inactivation<br />

of virusis and<br />

pathogena<br />

CO 2<br />

Staad <strong>Water</strong>works<br />

Düsseldorf<br />

Lörick <strong>Water</strong>works<br />

Flehe <strong>Water</strong>works<br />

Auf dem Grind Well Field<br />

Ozonation<br />

NBG GmbH<br />

Figure 2. The multi-protection barrier concept <strong>for</strong> drinking-water production.<br />

Mettmann<br />

Mettmann<br />

Rhine<br />

N<br />

Biological active filtration<br />

Activated carbon adsorption<br />

upper layer<br />

lower layer


• Equalization of fluctuating concentrations in river water.<br />

• Removal of biodegradable compounds.<br />

• Removal of bacteria, viruses, and parasites.<br />

The efficiency of these natural attenuation processes is influenced by river-water quality and the<br />

hydraulic situation. Continuous and periodic changes of water quality must be considered together<br />

with fluctuating river-water levels.<br />

Temporal Changes Affecting Bank Filtration<br />

During the last 30 years, Rhine River water quality has improved significantly. Many actions to<br />

reduce nutrients and pollutants were necessary. These measures comply with the best available<br />

technology in production, as well as wastewater treatment along the Rhine. The return of salmon<br />

in the year 2000 is a visible sign of the success of this program.<br />

These quality improvements were accompanied by a decrease of ammonia and DOC in river<br />

water. Oxygen concentrations reached saturation. The higher oxidation capacity of bank filtrate<br />

is linked to more efficient natural attenuation processes within the aquifer. This enabled the<br />

<strong>Water</strong>works to reduce treatment expenses.<br />

Natural attenuation processes are affected periodically by flood events. The water level increases<br />

up to 5 m, which leads to a higher influx of river water into the aquifer. Thus, oxygen and organic<br />

carbon flux also increase by a factor of 3 to 4. The aerobic microbes adapted to the lower flux are<br />

not able to degrade the entire organic carbon during flood events. Another decrease of natural<br />

attenuation processes becomes evident in the removal of bacteria. While mostly raw water already<br />

fulfills the European Drinking <strong>Water</strong> Standard, higher colony counts are observed in the<br />

production wells following flood events (Irmscher and Teermann, 2002; Schubert, 2002). These<br />

findings explain the need of subsequent purification processes: the so-called “second protective<br />

barrier” (see Figure 2).<br />

Another effect on natural attenuation processes during bank filtration is caused by changing water<br />

temperatures throughout the year. As the temperature of the river rises quickly in the spring and<br />

summer, the temperature of the water in the aquifer also rises, but not as quickly due to a certain<br />

time delay. Microbial activity related to temperature leads, there<strong>for</strong>e, to a higher degradation rate<br />

of organic carbon. The organic carbon is removed more efficiently; however, it must be considered<br />

that, under aerobic conditions, the bank filtrate becomes more aggressive versus calcite. A flexible<br />

treatment step must be adaptable to these changing conditions.<br />

Conclusions<br />

To reduce technical treatment expenses, river water of high quality, together with a profound<br />

knowledge of the natural attenuation processes within the aquifer, is essential. At the Rhine River,<br />

successful ef<strong>for</strong>ts have improved water quality significantly. More efficient natural attenuation<br />

processes accompanied this improvement during bank filtration, which enabled the <strong>Water</strong>works<br />

to reduce expenses in water treatment.<br />

Temporal changes of river-water quality and hydraulics still influence natural attenuation processes<br />

during bank filtration. They have to be well-understood to design adequate treatment st<strong>eps</strong> and<br />

to define specific target values on river-water quality. The multi-protective barrier concept,<br />

including both natural and technical purification, is still necessary to ensure drinking water of<br />

continuously high standards.<br />

89


90<br />

REFERENCES<br />

Irmscher, R., and I. Teermann (2002). “Riverbank filtration <strong>for</strong> drinking water supply – A proven method,<br />

perfect to face today’s challenges.” <strong>Water</strong> Science and Technology, 2(5-6): 1-8.<br />

Kühn, W., and U. Müller (2000). “Riverbank filtration.” Journal AWWA, 92(12): 60-69.<br />

Schubert, J. (2002). “<strong>Water</strong>-Quality improvement with riverbank filtration at Düsseldorf <strong>Water</strong>works in<br />

Germany.” Riverbank Filtration: Improving Source-<strong>Water</strong> Quality, C. Ray, G. Melin, and R.B. Linsky (eds.),<br />

Kluwer Academic Publishers, Dordrecht.<br />

PAUL ECKERT has more than 10 years of professional experience in water-management<br />

projects. He was involved in the project <strong>for</strong> bank filtration (risk assessment of shock loads)<br />

funded by the Federal Ministry <strong>for</strong> <strong>Research</strong> and Technology (BMFT). Currently, as Head<br />

of the <strong>Water</strong> Management Department, he is responsible <strong>for</strong> protection zone management<br />

at Düsseldorf <strong>Water</strong>works, where special investigations are per<strong>for</strong>med in the field of bank<br />

filtration and groundwater remediation. A hydrogeochemist, he specializes in<br />

hydraulic/hydrochemical modeling, the design of field studies, and the assessment of<br />

groundwater-quality problems. Extended experience includes the application of a geographic in<strong>for</strong>mation<br />

system database <strong>for</strong> groundwater. Eckert received both an M.S. (Diplom) and Ph.D. from the University of<br />

Bochum in Germany, where he researched the efficacy of enhanced natural attenuation processes in a BTEXcontaminated<br />

aquifer with nitrate.


Session 5: Dynamics<br />

An Update of the City of Guelph’s Response<br />

to Regulation 459/00: Effective Natural In Situ<br />

Filtration of Several Groundwater Under the Direct<br />

Influence of Surface-<strong>Water</strong> Supplies<br />

Dennis E. Mutti, P.E.<br />

Associated Engineering Limited<br />

Toronto, Ontario, Canada<br />

Peter Busatto<br />

City of Guelph <strong>Water</strong>works Division<br />

Guelph, Ontario, Canada<br />

Douglas H. Stendahl<br />

City of Guelph <strong>Water</strong>works Division<br />

Guelph, Ontario, Canada<br />

Caroline Korn, P.E.<br />

Associated Engineering Limited<br />

Toronto, Ontario, Canada<br />

Elia Edwards, P.E.<br />

Associated Engineering Limited<br />

Toronto, Ontario, Canada<br />

Associated Environmental Limited developed a protocol <strong>for</strong> determining GWUDI status to assist<br />

the City of Guelph in Ontario, Canada, in assessing treatment requirements <strong>for</strong> their municipal<br />

water supply. This protocol is based on protocols established by the United States Environmental<br />

Protection Agency (USEPA) and further defined in American <strong>Water</strong> Works Association <strong>Research</strong><br />

Foundation Project #605. The protocol entails three stages of assessment. Stage 1 is a review of<br />

existing in<strong>for</strong>mation to characterize the well as a true groundwater or GWUDI. This data includes<br />

well construction and maintenance records, sanitary condition of the well, hydrogeological data<br />

including time-of-travel estimates (well-head delineation), geological characteristics, and waterquality<br />

data. Stage 2 is initiated if there is insufficient data to make a characterization of the nature<br />

of the well in Stage 1 and consists of a period of data collection that will allow the characterization<br />

to be completed. Wells that are characterized as GWUDI are then subject to a Stage 3 assessment,<br />

which consists of a data collection period followed by an assessment of the level of natural or in situ<br />

filtration that is occurring. If sufficient evidence of in situ filtration is shown in the Stage 3<br />

assessment, then a case can be made <strong>for</strong> waiving the chemically assisted filtration component of the<br />

minimum treatment requirement <strong>for</strong> GWUDI in the Province of Ontario. The minimum treatment<br />

requirements in the Province of Ontario <strong>for</strong> GWUDI sources with effective in situ filtration is 3-log<br />

inactivation <strong>for</strong> Giardia cysts and 4-log inactivation <strong>for</strong> viruses from disinfection alone.<br />

Correspondence should be addressed to:<br />

Dennis Mutti, P.E.<br />

Senior <strong>Water</strong> Treatment and Supply Engineer<br />

Associated Engineering Limited<br />

525-21 Four Seasons Place • Toronto, Ontario M9B 6J8 Canada<br />

Phone: (416) 622-9502 • Fax: (416) 622-6249 • Email: muttid@.ae.ca<br />

91


92<br />

The City of Guelph’s supply system consists of a network of 21 groundwater wells and a spring<br />

collector system, augmented by artificial recharge. The Stage 1 assessment was completed <strong>for</strong> all<br />

of the City’s wells as a part of the provincially mandated Engineers’ Report. Two bedrock wells<br />

(Burke Well and Downey Well) and an overburden well system (Carter Wells 1 and 2) were<br />

identified as requiring a Stage 2 assessment due to insufficient existing data. The collector system<br />

(Arkell Spring and Glen Collector) was identified as GWUDI by definition and required a Stage 3<br />

assessment. The City of Guelph decided to expedite the process and collect data required <strong>for</strong> a<br />

Stage 3 assessment <strong>for</strong> the three wells, as well as <strong>for</strong> the surface-water recharge and collector system<br />

in case the wells were determined to be GWUDI.<br />

The results of the GWUDI assessment were presented at a stakeholders meeting with key<br />

provincial regulators. Consensus was reached as follows:<br />

• Burke Well and Downey Well are representative of true groundwater supplies and, as such,<br />

must provide disinfection as mandated by the Drinking <strong>Water</strong> Protection Regulation<br />

(2.0-log virus inactivation).<br />

• The treatment requirement to provide chemically assisted filtration <strong>for</strong> the Carter Wells,<br />

Arkell Recharge System, and Glen Collector System should be waived as effective<br />

natural in situ filtration is provided through the subsurface. As such, the treatment<br />

requirements <strong>for</strong> Giardia cysts (3-log inactivation) and viruses (4-log inactivation) may<br />

be provided by disinfection only.<br />

The City of Guelph is currently negotiating treatment requirements, as well as the final wording<br />

in its Consolidated Certificate of Approval, with the Ontario Ministry of the Environment.<br />

DENNIS MUTTI has 13 years of progressive experience as a Project Manager and Process<br />

Engineer, the last 5 of these years with Associated Engineering, which provides environmental<br />

engineering consulting services to the Ontario, Canada, market. His focus has been on<br />

process evaluation and development, and the design, implementation, and optimization of<br />

water supply and treatment systems. Mutti has worked on all types of water-supply and<br />

treatment projects, including water-supply master plans, water treatment plant designs and<br />

upgrades, groundwater supplies, automation and Supervisory Control and Data Acquisition<br />

(SCADA) system implementation, and start-up and commissioning. He received both a B.S. in Chemical<br />

Engineering and an M.S. in Civil Engineering from the University of <strong>Water</strong>loo.


Session 5: Dynamics<br />

On Bank Filtration and Reactive Transport Modeling<br />

Dr.-Ing. Ekkehard Holzbecher<br />

Humboldt University<br />

Berlin, Germany<br />

Prof.-Dr. Gunnar Nützmann<br />

Humboldt University<br />

Berlin, Germany<br />

Modeling can enhance the understanding of the influence of hydraulic, transport, and biogeochemical<br />

processes <strong>for</strong> the development of bank filtration as an applied technology. Because a<br />

coupled three-dimensional approach, including all biogeochemical trans<strong>for</strong>mations, speciation,<br />

and kinetics, is not feasible nowadays in applied projects, a simpler procedure is proposed. In<br />

the first step, flow is calculated using analytical solutions or numerical models in two- or<br />

three-dimension. In the second step, reactive transport is simulated in one-dimension along<br />

flowpaths. How these two st<strong>eps</strong> are per<strong>for</strong>med in a project depends on the chosen software <strong>for</strong> flow<br />

and reactive transport. A simple approach is demonstrated that uses analytical solutions <strong>for</strong> flow and<br />

the pH redox equilibrium equation (PHREEQC) <strong>for</strong> reactive transport. This approach is appropriate<br />

<strong>for</strong> understanding processes in porous media in the direct vicinity of a surface-water body.<br />

Flow Modeling<br />

The simulation of reactive transport at a field site is always based on a flow model. Flow models<br />

can account <strong>for</strong>:<br />

• Inhomogeneities.<br />

• Anisotropies.<br />

• Site-specific design of the well system.<br />

• Irregular (usually) setting of the boundaries.<br />

Rather complex two-dimensional or three-dimensional flow patterns emerge.<br />

Although the capability of software has increased in conjunction with the power of hardware,<br />

there are still limiting factors <strong>for</strong> solving general three-dimensional reactive transport models.<br />

Transport models require a small spatial resolution (e.g., a high number of nodes or blocks) in<br />

which dependent variables are computed. Biogeochemistry requires a small temporal resolution to<br />

capture the changes of some variables in response to changes in the system.<br />

Numerical errors and their propagation when solving huge linear and/or nonlinear systems are<br />

hard to predict. If some well-known constraints (grid Péclet-number criterion, Courant-number<br />

and Neumann-number criteria) are not fulfilled, numerical errors increase dramatically. The<br />

Correspondence should be addressed to:<br />

Dr.-Ing. Ekkehard Holzbecher<br />

Senior Scientist<br />

Humboldt University Berlin<br />

<strong>Institute</strong> of Freshwater Ecology and Inland Fisheries • Müggelseedamm 310 • 12587 Berlin, Germany<br />

Phone: +0049-30-64181 667 • Fax: +0049-30-64181 663 • E-Mail: holzbecher@igb-berlin.de<br />

93


94<br />

resulting numerical dispersion or oscillations disturb results in that the results cannot be used <strong>for</strong><br />

comparisons with observations.<br />

To couple the result of a flow model with a general geochemical approach, as given by the<br />

PHREEQC code, the flow system must be reduced to one-dimension. Flowpaths can be taken from a<br />

two- or three-dimensional model and be used <strong>for</strong> a one-dimensional reactive transport simulation.<br />

The procedure is applicable <strong>for</strong> bank-filtration systems, as concentration gradients in transverse<br />

directions can be assumed to be much smaller than in longitudinal directions.<br />

In the first step towards such a simulation, flowpaths are taken from analytical solutions <strong>for</strong><br />

idealized well gallery systems. MATLAB (2002) software is applied <strong>for</strong> the analytical solution.<br />

Systems with different numbers of wells, different distances from the surface-water body, and<br />

different pumping rates can be investigated. It is also possible to introduce a groundwater base<br />

flow. Moreover, different types of boundary conditions at the bank can be considered.<br />

Figures 1 and 2 show the results of a simulation <strong>for</strong> two wells. The calculations with analytical<br />

solutions, as well as the graphical representation, were made using MATLAB.<br />

Figure 1. Hydraulic potential φ (m 3 /s) <strong>for</strong> two wells with different pumping rates and distances from a<br />

surface-water body near a gaining stream.<br />

Some basic results can be obtained <strong>for</strong> a single well near a gaining stream. There is a critical<br />

pumping rate, Q crit , below which only groundwater is pumped. By increasing the pumping rate<br />

above Q crit , the share of surface water (filtrate) increases. An equal share between groundwater<br />

and surface water is given <strong>for</strong>:<br />

Q = 6.1878 •Q crit<br />

(Equation 1)


Figure 2. Streamlines <strong>for</strong> two wells with different pumping rates and distances from a surface-water body near<br />

a gaining stream.<br />

We intend to further explore the use of analytical solutions. A numerical algorithm, which<br />

calculates flowpaths and travel times from the surface-water body to the wells, is currently being<br />

tested. If combined with geochemical computations, the approach can provide a highly important<br />

tool <strong>for</strong> the design of well galleries.<br />

Travel times along flowpaths are obtained by numerical integration.<br />

Reactive Transport Modeling<br />

With the outlined strategy, the well-established PHREEQC code (new version: PHREEQC2) can<br />

be applied <strong>for</strong> reactive transport modeling to reduce the number of spatial dimensions. The<br />

PHREEQE code (which is the origin of PHREEQC) was mainly a geochemical speciation code<br />

and originally not intended to be combined with a transport model. Velocity is not an input to be<br />

specified. Instead, lengths (∆x), time step (∆t), cells (number of blocks), and shifts (number of time<br />

st<strong>eps</strong>) must be given as parameters (Parkhurst, 1995). Nevertheless, the simulated velocity v<br />

results from the input parameters:<br />

ν = ∆x / ∆t<br />

(Equation 2)<br />

The so-called “mixing cell approach” <strong>for</strong> advection, which uses the operator-splitting technique, is<br />

combined with the finite difference method <strong>for</strong> diffusion and proves to be competitive with advanced<br />

numerical techniques. Errors from the discretization advection term are successfully suppressed.<br />

PHREEQC can also be used <strong>for</strong> variable velocity along flowpaths. The time step in Equation 2 is<br />

fixed, but grid spacing varies locally, taking velocity changes into account. Small velocities result<br />

in small ∆x and high velocities in large ∆x.<br />

95


96<br />

As an alternative to PHREEQC <strong>for</strong> certain tasks, a new model based on MATLAB is currently in<br />

development. The model is being tested in a number of benchmark studies. Figure 3 shows the<br />

results of the MATLAB model <strong>for</strong> a test of redox sequences, as compared with PHREEQC output.<br />

Concentration [mol/L]<br />

5.00E-03<br />

4.50E-03<br />

4.00E-03<br />

3.50E-03<br />

3.00E-03<br />

2.50E-03<br />

2.00E-03<br />

1.50E-03<br />

1.00E-03<br />

5.00E-04<br />

CH20 MATLAB<br />

N(0) PHREEQE<br />

S-2 MATLAB<br />

02 PHREEQC<br />

CH20 PHREEQC<br />

N(-3) MATLAB<br />

S(-2) PHREEQC<br />

pH MATLAB<br />

N(5) MATLAB<br />

N(-3) PHREEQC<br />

HCO3 MATLAB<br />

pH PHREEQC<br />

N(3) PHREEQC<br />

S(6) MATLAB<br />

HCO3 PHREEQC<br />

pe MATLAB<br />

Figure 3. Reactive transport (redox) test case: Redox zoning after a 1-year simulation timeframe.<br />

N(0) MATLAB<br />

S(6) PHREEQC<br />

O2 MATLAB<br />

pe PHREEQC<br />

0.00E-00<br />

–6.00<br />

0.00 10.00 20.00 30.00 40.00 50.00<br />

x [m]<br />

60.00 70.00 80.00 90.00 100.00<br />

In the MATLAB model, valence electron balances are obtained using an adequate conceptualization<br />

of operational valence electron balancing. Redox equilibrium modeling, including kinetics<br />

<strong>for</strong> organic carbon biodegradation and operational valence electron balancing to the equilibrium<br />

module, yields automatically to a correct choice of electron acceptor.<br />

For non-redox equilibrium systems, which incorporate proton and component mass balances and<br />

mass action equations, the operational valence electron balance and the electron (e – ) as <strong>for</strong>mal<br />

species entering mass action equations — derived from redox half reactions — are considered.<br />

Since free electrons are generally not observed in aqueous solutions (Thorstenson, 1984),<br />

operation valence electron balancing has to be carried out by adapting the share of different<br />

valence states of heterovalent components (in a sense, valence electron “bookkeeping”) (Appelo<br />

and Postma, 1996).<br />

In inorganic redox systems, the sum of mobile operational valence electrons is an additional<br />

transport species. In the presence of biodegradation redox reactions (which supply additional<br />

carbonate species to the system), the actual operation electron balance has to be corrected by the<br />

carbonate species arising from biodegradation, as pointed out by Brun and Engesgaard (2002).<br />

A test case including hydrodynamic transport to the redox system was set up using MATLAB. At<br />

the inlet of the model column, a solution that is rich in mobile organic matter infiltrates into an<br />

aquifer void of organic matter, but contains oxygen, nitrate, and sulphate as electron acceptors.<br />

Unlike Redox Test Case 1, ammonium (N[–3]) is included in the equilibrium reaction framework.<br />

During infiltration, a steady-state redox zone is simulated similar to the redox zone observed at<br />

16.00<br />

14.00<br />

12.00<br />

10.00<br />

8.00<br />

6.00<br />

4.00<br />

2.00<br />

0.00<br />

–2.00<br />

–4.00<br />

pH. pE [–]


well galleries near the Unterhavel (Berlin). The distance from the inlet at which the redoxcline<br />

(a large jump in the redox state) is expected depends on:<br />

• Hydrodynamic parameters in relation to the first-order organic matter decay constant.<br />

• Initial redox state (operational valence electron balance).<br />

• Initial electron-acceptor concentrations.<br />

• Availability of reactive organic matter.<br />

These are the main sensitivity parameters that determine the extent of redox zoning.<br />

Figure 3 shows the simulated steady-state species and pH/oxidation-reduction (redox) potential (pE)<br />

distribution along the model transect. The simulation reached steady-state concentrations after a<br />

150-day simulation time. The pE profile shows the typical variation caused by aerobic respiration<br />

(pE buffering at +15), denitrification (pE buffering at +12 to +14), and sulphate reduction<br />

(pE buffering at –3). The profiles of oxygen (O 2) concentrations, nitrate (N[5]), and sulphate<br />

(S[6]) concentrations also are figured so that the changes of pE are explained straight<strong>for</strong>wardly.<br />

Because ammonium (N[–3]) is incorporated into the speciation, elemental nitrogen (N[0]) will<br />

not be the end step of the nitrogen species, but will be reduced to ammonium in conjunction with<br />

the sulphate-reduction step.<br />

Outlook<br />

The outlined coupling strategy <strong>for</strong> flow and reactive transport is intended to be applied on real<br />

sites. Flow simulation will be done using MODFLOW (Harbaugh et al., 2000) or FEFLOW (2002)<br />

<strong>for</strong> site-specific models and by analytical solutions <strong>for</strong> conceptual models concerning well-gallery<br />

design. For reactive transport, there is the choice between PHREEQC and MATLAB.<br />

REFERENCES<br />

Appelo, C.A., and D. Postma (1996). Geochemistry, groundwater and pollution, Balkema, Rotterdam.<br />

Brun, A., and P. Engesgaard (2002). “Modeling of transport and biogeochemical processes in pollution<br />

plumes: literature review and model development.” Journ. of Hydrology, 256: 211-227.<br />

FEFLOW (2002). Finite element, subsurface flow & transport simulation system, WASY, Berlin, Germany.<br />

Harbaugh, A.W., E.R. Banta, M.C. Hill, and M.G. McDonald (2000). MODFLOW-2000, the U.S. Geological<br />

Survey modular ground-water model — User guide to modularization concepts and the Ground-<strong>Water</strong> Flow Process,<br />

U.S. Geological Survey Open-File Report 00-92, 121p.<br />

MATLAB (2002). MATLAB the language of technical computing, The MathWorks, Inc., Natick, MA.<br />

Parkhurst, D.L. (1995). PHREEQC: A computer program <strong>for</strong> speciation, reaction-path, advective transport, and<br />

inverse geochemical calculations, U.S. Geological Survey, <strong>Water</strong> Res. Invest. Report 95-4227, Lakewood, 143p.<br />

Thorstenson, D.C. (1984). The concept of electron activity and its relation to redox potentials in aqueous<br />

geochemical systems, Open-File Report 84-072, U.S. Geological Survey Denver, Colorado.<br />

97


98<br />

Mathematician EKKEHARD HOLZBECHER has worked the field of groundwater<br />

science and applications since 1984. Most of his projects are related to numerical<br />

modeling. Past projects include investigating the safety of nuclear waste repositories in<br />

Germany, saltwater intrusion in the Nile-Delta in Egypt, geothermal flow in Japan, and<br />

abandoned industrial sites in Germany. Currently, he is investigating various problems<br />

regarding eco-hydrology. Since 2002, Holzbecher has been a Senior Scientist with<br />

Humboldt University in Berlin, where he teaches groundwater and transport modeling<br />

courses, and participates in the Natural and Artificial Systems <strong>for</strong> Recharge and Infiltration project. He was<br />

also involved in the international Groundwater Hydrology Modeling Strategies <strong>for</strong> Per<strong>for</strong>mance Assessment of<br />

Nuclear Waste Disposal (HYDROCOIN) project and is a member of the United Nations Educational,<br />

Scientific and Cultural Organization (UNESCO) working group <strong>for</strong> the Development and Calibration of<br />

Coupled Hydrological/Atmospheric Models. Holzbecher received a Ph.D. from the Civil Engineering<br />

Department at the Technical University Berlin and the German degree of Habilitation (qualification <strong>for</strong> a<br />

tenure professorship) at the Faculty of Geosciences at the Free University Berlin, where he is now a<br />

Privatdozent.


Dinner Presentation<br />

Hydraulic Sensitivities and Reduction Potential<br />

Correlated with the Distance Between<br />

the Riverbank and Production Well<br />

Bernhard Wett, Ph.D.<br />

University of Innsbruck<br />

Innsbruck, Austria<br />

Objectives<br />

Beginning with a detailed case study of a <strong>RBF</strong> system at the alpine river, Enns, in Austria, some<br />

aspects of the question, “In what ways do specified hydraulic parameters affect water quality (e.g.,<br />

what is the appropriate well distance from the riverbank)?” are elaborated. In addition to<br />

monitoring data, numerical sensitivity analyses will be applied to display interactions between<br />

well distances, filtrate production, and reduction processes.<br />

Methodology and Site Description<br />

The investigated <strong>RBF</strong> well is situated about 50 m from the bank of the oligotrophic alpine river,<br />

Enns, at the beginning of the 5-kilometer long reservoir of the HPP Garsten power plant (Figure 1).<br />

At this particular location, the river stage (as determined by a dam) is 302.0 m above sea level and<br />

showed only minor elevations when discharge varied from about 70 to 540 m 3 /s during a 1-year<br />

measurement period. Since the riverbed is cut into the dense flysch zone, river water infiltrates<br />

almost exclusively through the bank and not the bottom. Between the river and well, aquifer<br />

thickness is about 5 m, and the total thickness of the gravel layer is 15 m (Ingerle et al., 1999;<br />

Wett et al., 2002).<br />

Organic loading is very low (the DOC concentration varies between 1 and 2 mg/L), and river<br />

water is saturated with oxygen (greater than 10 mg/L). Groundwater quality reflects the trends of<br />

land use and the intensity of agriculture along the river. The further downstream of the river, the<br />

higher the nitrate and pesticides concentrations are in groundwater. In the region of the filtration<br />

site, groundwater shows nitrate concentrations of about 50 mg/L. The enrichment of groundwater<br />

by river filtrate (less than 5 mg/L nitrate) offers a solution to obtain nitrate concentrations in<br />

agreement with drinking-water standards.<br />

A high-grade steel box with windows <strong>for</strong> visual observation and video recording of riverbed clogging<br />

was installed in the riverbank. One set of five probes was installed at the upstream side of the box in<br />

natural sediment (multi-level Probe F at depths between 0.1 and 0.9 m beneath the riverbed surface),<br />

and two sets of five probes (D and E) were installed downstream in specified filter sand (0 to 4 mm)<br />

to measure hydraulic heads and obtain water samples. <strong>Water</strong> levels in the probes correspond with the<br />

levels of graduated glass pipes within the steel box. An analysis of hydraulic head data and relative<br />

variations of head differences allows <strong>for</strong> filter velocity calculations. Higher hydraulic conductivity in<br />

Correspondence should be addressed to:<br />

Bernhard Wett, Ph.D.<br />

Scientific <strong>Research</strong>er and Lecturer<br />

<strong>Institute</strong> of Environmental Engineering<br />

University of Innsbruck • Technikerstr. 13 • A-6020 Innsbruck, Austria<br />

Phone: +43 512 507 6926 • Email: Bernhard.Wett@uibk.ac.at<br />

99


100<br />

Figure 1. Schematic overview and cross-section of the <strong>RBF</strong> site on the Enns River, a tributary of<br />

the Danube River.<br />

the filter sand causes a significantly higher filter velocity than in natural sediment (the mean filter<br />

velocity is 1.99 × 10 –5 m per second and 0.83 × 10 –5 m per second, respectively).<br />

Results<br />

In general, microbial metabolism and degradation processes require both electron donators and<br />

acceptors. Depending on the site and its inherent boundary conditions, an oligotrophic river<br />

would have carbon as a limiting factor. The lack of electron donators can be compensated <strong>for</strong> by<br />

long hydraulic retention in organic-rich sediment layers at high water temperatures. A long<br />

retention time (represented by low infiltration rates in Figure 2) in the most biologically active<br />

zone near the water-sediment interface drives the balance between the availability of electron<br />

donators and acceptors towards higher oxygen consumption. The arrow indicating the influence<br />

of hydraulic parameters refers to the distance d from the river to the production well, hydraulic<br />

Figure 2. Limitations of microbial metabolism in the riverbank and their balancing factors.


conductivity k (especially of the riverbed), aquifer thickness H, and production rate Q BF.<br />

Additionally, an increase in temperature causes a decrease in water viscosity and, there<strong>for</strong>e, creates<br />

higher hydraulic conductivity (this is a 50-percent increase in the presented case study).<br />

It is a known fact that particulate organic matter deposited in the riverbed represents an important<br />

carbon pool (Bretschko and Moser, 1993). The metabolized mass of particulate organic carbon can<br />

be estimated from the respiration of the sediment community to identify the dominant carbon<br />

source. The difference between total hyporeic respiration and DOC respiration, as depicted in<br />

Figure 3, describes a substantial gap in the mass balance between carbon fluxes in and out of the<br />

riverbed. In the case of increased DOC loads of the filtrate, the microbial community can adapt<br />

as long as electron acceptors are available (e.g., basin maturation <strong>for</strong> soil-aquifer treatment of<br />

percolated secondary effluent occurs within a few months) (Quanrud et al., 2003). Reactive<br />

transport models can describe the relatively quick development from electron donator limitation<br />

towards electron acceptor limitation (Lensing et al., 1994).<br />

DOC immobilization / microbial C respiration<br />

(µ mol C per L)<br />

DOC immobilization C respiration<br />

Station C Station D<br />

Station F Station E<br />

month (1997/1998)<br />

(non-DOC)-C respiration<br />

Figure 3. Seasonal variation of DOC immobilization (∆DOC), total hyporheic carbon respiration (HCR),<br />

and the difference between ∆DOC and HCR in the riparian zone (0.9 to 1.0 m from the sedimentwater<br />

interface, with Stations C and F in natural sediment and Stations D and E in specified filter<br />

sand) (Brugger et al., 2001).<br />

Figure 4 demonstrates influences on microbial reduction processes both by temperature and flow<br />

velocity: the oxygen and nitrate of the electron acceptors show the highest peaks during the winter<br />

season (the oxygen concentration approaches saturation at water temperatures of 2-degrees Celsius<br />

in February) and drop to low points during the summer season (8.5-mg/L oxygen [O 2] in Station E<br />

and 4.5-mg/L O 2 in Station F in July, when water temperature reaches 15-degrees Celsius). This<br />

period of increased oxygen depletion corresponds with concentration peaks of iron and manganese.<br />

The mean oxygen depletion during the summer season in the riverbed achieves 4.0-mg/L O 2 at<br />

Bore F and 2.8-mg/L O 2 at Bore D. These respiration rates include a portion <strong>for</strong> DOC immobilization<br />

of 25 and 19 percent, respectively (see Figure 3). Hence, both DOC immobilization and<br />

(non-DOC)-C respiration<br />

(µ mol C per L)<br />

101


102<br />

Figure 4. Courses of oxygen, nitrate, iron, and manganese concentrations at Station E in riverbed zones with<br />

low hydraulic conductivity and at Station F with higher conductivity during the 1-year monitoring<br />

period (0.9 to 1.0 m from the sediment-water interface).<br />

total hyporheic carbon respiration are significantly higher in Infiltration Area F, where the filter<br />

velocity is 58 percent lower than at Bore D. These results suggest that monitored degradation<br />

processes are limited by available retention time and that the hydrolyses of particulate organic<br />

matter is the time-limiting process step.<br />

The spatial distribution of oxygen, as presented in Figure 5, clearly shows the influence of<br />

retention time on reduction processes. During the migration time in the first meter of the flowpath<br />

in natural sediment, the oxygen concentration drops to a minimum level and the subsequent<br />

recovery is attributed to mixing with oxygen-enriched water.<br />

Figure 5. Oxygen concentration profiles along the flowpath from the river to the well,<br />

starting from four different infiltration areas (Stations D and E in natural<br />

sediment, July 27, 1998).<br />

In the following section, a calibrated model of the presented site serves as the initial situation (plot<br />

and cross-section in Figure 6) of a systematic parameter investigation. The linear sensitivity<br />

analysis uses a function, δ v,p , to quantify the sensitivity of the model response to a unit change in


parameter value:<br />

Figure 6. Initial conditions of numerical sensitivity analyses (MODFLOW software).<br />

dv/ v<br />

δv,p = where p = parameter and v = output variable (model response).<br />

dp/ p<br />

The larger the value of the function, the more significant the specific parameter is <strong>for</strong> model<br />

behavior. The applicability of this <strong>for</strong>m is limited to linear cause-and-effect relationships or small<br />

parameter variations. The results in Table 1 outline the major influence of riverbed clogging<br />

k A/k BF on filtrate production Q BF and well distance d on the total migration time HRT and<br />

maximum infiltration rate v max.<br />

Table 1. Comparison of Parameter Sensitivities to Hydraulic Properties of the Considered <strong>RBF</strong> Site<br />

d v,p k A / k BF k A Q Prod. d H<br />

HRT [d] 0.24 0.33 –1.01 1.89 1.20<br />

Q BF [%] –0.25 –0.15 0.03 –0.18 0.17<br />

v max [m/h] –0.44 –0.44 0.89 –0.76 –0.89<br />

HRT = Hydraulic retention time. Q BF = Filtrate portion. v max = Maximum infiltration rate.<br />

103


104<br />

Conclusions<br />

Biologically mediated degradation processes mainly occur at the first meter of the flowpath from<br />

the river to the well. Hydraulic retention in this layer and infiltration rates, respectively, are<br />

crucial parameters of additional particulate carbon respiration. Infiltration velocity, in turn,<br />

depends on the site-specific hydraulic setting, especially on production rate, well distance, and<br />

aquifer thickness. The appropriate selection of these parameters, there<strong>for</strong>e, should aim towards<br />

achieving preferable hyporheic conditions.<br />

Acknowledgements<br />

Basic results presented in this paper have been achieved by close interdisciplinary cooperation in the<br />

course of a research project coordinated (Ch. Hasenleithner) and funded by the Ennskraft Company.<br />

REFERENCES<br />

Brettschko, G., and H. Moser (1993). “Transport and retention of matter in riparian ecotones.” Hydrobiologia,<br />

251: 95-102.<br />

Brugger, A., B. Wett, I. Kolar, B. Reitner, and G.J. Herndl (2001). “Immobilization and bacterial utilization<br />

of dissolved organic carbon entering the riparian zone of the alpine Enns River, Austria.” Aquatic Microbial<br />

Ecology, 24(2): 129-142.<br />

Ingerle, K., A.P. Blaschke, A. Brugger, C. Hasenleithner, G.J. Herndl, H. Jarosch, I. Kolar, H.J. Lensing, S.<br />

Pöschl, N. Quéric, B. Reitner, F. Schöller, R. Sommer, and B. Wett (1999). Forschungsprojekt Uferfiltrat<br />

(<strong>Research</strong> project bank filtration), <strong>Research</strong> Initiative Verbund, Vienna, 60, 43-78.<br />

Lensing, H.J., M. Vogt, and B. Herrling (1994). “Modeling of biologically meditated redox processes in the<br />

subsurface.” J. of Hydrology, 159: 125-143.<br />

Quanrud, D.M., J. Hafer, M.M. Karpiscak, J. Zhang, K.E. Lansey, and R.G. Arnold (2003). “Fate of organics<br />

during soil-aquifer treatment: sustainability of removals in the field.” Wat. Res., 37: 3,401-3,411.<br />

Wett, B., H. Jarosch, and K. Ingerle (2002). “Flood induced infiltration affecting a bank filtrate well at the<br />

River Enns, Austria.” J. of Hydrology, 266(3-4): 222-234.<br />

BERNHARD WETT has been a Scientific <strong>Research</strong>er at the Department of Environmental<br />

Engineering at the University of Innsbruck in Austria <strong>for</strong> 9 years. His recent scientific<br />

work has focused on the separate biological treatment of rejection water — a major<br />

interaction between the wastewater and the sludge lane of a treatment plant. With regard<br />

to riverbank filtration and its relation to flooding events, he has conducted an<br />

interdisciplinary project together with various companies and universities over the past<br />

several years. The central question of these investigations was the vulnerability of<br />

riverbank-filtration waterworks, in regard to both hydraulic and water-quality aspects. Wett participates in<br />

working groups of national water associations (Wastewater Technology Association, Austrian <strong>Water</strong>-and<br />

Waste Management Association) and in the EU-COST-624 Action. He lectures in various fields of<br />

environmental engineering, including numerical methods of groundwater flow modeling. Wett received an<br />

M.S. in Civil Engineering and a Ph.D. in Engineering Science-Environmental Engineering from the<br />

University of Innsbruck.


Keynote Presentation<br />

Riverbank Filtration: The European Experience<br />

Prof. Dr.-Ing. Martin Jekel<br />

Technical University of Berlin<br />

Berlin, Germany<br />

Prof. Dr.-Ing. Thomas Grischek<br />

University of Applied Sciences<br />

Dresden, Germany<br />

Introduction<br />

The first historical citation of a small bank filtration system <strong>for</strong> a village dates back to 1798 in a<br />

German text entitled, About <strong>Water</strong>. It describes a construction procedure <strong>for</strong> artificial bank<br />

filtration into a dug well, about 2 m from a turbid creek. The intermediate space is filled with sand<br />

to remove pollution during gravity flow into the well. The sand is replaced twice a year to<br />

guarantee sufficient yield. Thus, we interpret this treatment as a combination of <strong>RBF</strong> and slow<br />

sand filtration, two techniques still widely used and known as effective physical and biological<br />

filters with many similarities.<br />

In Europe, more than 130 years of experience exist in the O&M of large bank-filtration schemes.<br />

The oldest operating <strong>RBF</strong> waterworks were founded in the 1870s in Germany and The Netherlands<br />

during the introduction of centralized water supply in urban areas. The most important sites are<br />

situated along the Danube, Rhine, and Elbe rivers, and the lakes in the City of Berlin area. A large<br />

variety of schemes has been designed and operated according to site-specific conditions. Many<br />

research projects have been conducted within the last 20 years to study attenuation processes<br />

during bank filtration.<br />

Relevance of Riverbank Filtration in Europe<br />

Groundwater derived from infiltrating river water provides 45 percent of drinking-water supplies in<br />

Hungary, 16 percent in Germany, 5 percent in The Netherlands, and 50 percent in the Slovak<br />

Republic. In Germany, the City of Berlin depends on bank filtration <strong>for</strong> about 60 percent of its<br />

drinking water. The City of Düsseldorf, situated on the River Rhine in Germany, is entirely supplied<br />

with drinking water derived from bank filtration. Budapest, the capital of Hungary, also depends<br />

100 percent on bank filtration. <strong>Water</strong>works in many other cities (e.g., Cologne, Dresden, Zurich,<br />

Lindau, and Maribor) rely on bank filtrate as an important resource. In Finland, 217 waterworks use<br />

bank-filtration techniques as part of their water-treatment process (Kivimäki, 1995).<br />

Besides known bank-filtration sites, there are many groundwater works where bank filtration is<br />

not planned and is unintentional, especially within periods of high surface-water levels during<br />

floods.<br />

Correspondence should be addressed to:<br />

Prof. Dr.-Ing. Martin Jekel<br />

Professor, Department of <strong>Water</strong> Quality Control<br />

Technical University of Berlin<br />

Sekr. KF 4 • Strasse des 17. Juni 135 • 10623 Berlin, Germany<br />

Phone: +(30) 314 - 25058 • Fax: +(30) 314 - 23313 / 23850 • Email: jekel@itu201.ut.TU-Berlin.DE<br />

105


106<br />

Geohydraulic Characteristics<br />

Grischek et al. (2002) provide an overview on the geohydraulic characteristics of different<br />

bank-filtration sites in Europe. Typical aquifers used <strong>for</strong> bank filtration consist of alluvial sand and<br />

gravel deposits and have a hydraulic conductivity higher than 1 × 10 –4 m per second. The<br />

thickness of the exploited alluvial aquifers covers a wide range. Along the Danube River in the<br />

Slovak Republic and Elbe River in Germany, aquifer thickness exceeds 50 m. Along the Lot River<br />

in France and Neckar River in Germany, boreholes are operated in shallow aquifers of about 5-m<br />

thickness. The distance between production wells and the bank of the river or lake is, in general,<br />

more than 50 m. Excluding some sites in The Netherlands and along the Danube River in the<br />

Slovak Republic, travel times are mostly between 20 and 300 days.<br />

In some countries, a travel time minimum of 50 days is proposed because an accepted rule of<br />

thumb assumes that about 50 days are sufficient to obtain pathogen-free water. But, recent findings<br />

underline that other factors, such as surface area (flowpath length), riverbed properties, and redox<br />

conditions, play an important role in removing pathogens.<br />

At most sites, vertical wells are in operation. At sites with a high aquifer thickness (<strong>for</strong> example,<br />

Kalinkovo in the Slovak Republic), vertical wells have screen lengths of 40 m, allowing high<br />

abstraction rates. Older systems often include siphon pipes connected to vertical well galleries<br />

with low abstraction rates (<strong>for</strong> example, Karany in the Czech Republic and Dresden-Tolkewitz in<br />

Germany). Horizontal wells are used at sites with high abstraction rates, such as in Düsseldorf,<br />

Germany, and Budapest, Hungary.<br />

Riverbeds at <strong>RBF</strong> sites are normally cut into an underlying sand and gravel aquifer, resulting in<br />

direct hydraulic contact of the river and aquifer. At many <strong>RBF</strong> sites, erosive conditions in the river<br />

limit the <strong>for</strong>mation of a siltation layer. Detailed research to understand the specific mechanisms<br />

controlling clogging has been undertaken on the Rhine River (Schubert, 2001), Enns River<br />

(Ingerle et al., 1999), and Elbe River (Heeger, 1987). At most rivers, infiltration occurs in the<br />

areas between the bank adjacent to the wells and the middle of the river, sometimes over the<br />

whole width of the river. At the Rhine River, partial clogging of the riverbed limits infiltration<br />

near the bank adjacent to the wells. The riverbed infiltration zone is fairly clean, but other areas<br />

are coated with a layer of about 1-millimeter thickness. Sand ripples also develop. Under rapidly<br />

changing river levels, the riverbed is cleaned (Schubert, 2000). The other case is observed at<br />

impounded river lakes in Berlin, where infiltration occurs mainly via the bank because thick<br />

organic mud layers at the bottom of lakes have low hydraulic conductivity.<br />

Surface-<strong>Water</strong> Quality<br />

In general, organic carbon concentrations (TOC) in river waters used <strong>for</strong> <strong>RBF</strong> are between 1 and 6 mg/L.<br />

The lakes of Berlin are higher in TOC, up to 10 mg/L, due to natural fulvic acids, effluent organic<br />

matter, and algal growth. In Finland, where NOM concentrations contribute to TOC values of<br />

4 to 15 mg/L, problems are associated with natural humic substances and DBP <strong>for</strong>mation.<br />

Temperature variations in most rivers range from zero to 25-degrees Celsius, and the pH is between<br />

6 and 8.<br />

Since the late 1950s, the water quality of large rivers in Europe began to deteriorate. High<br />

wastewater input threatened the use of bank filtrate. Furthermore, spectacular spills (<strong>for</strong> example,<br />

the Sandoz accident on November 1, 1986 [Sontheimer, 1991]), underlined the need <strong>for</strong><br />

sanitation measures and pollution control. The activities of waterworks and water-industry<br />

associations, authorities and industries, transborder programs, and the closure of industries all


esulted in a significant improvement of river-water quality (e.g., the Rhine River since 1980 to<br />

1985 and the Elbe River since 1990). Historic water-quality records show a strong decrease in the<br />

concentrations of phosphate, polynuclear aromatic hydrocarbon compounds, pesticides such as<br />

atrazine and mecoprop, and biodegradable organic carbon.<br />

Raw-<strong>Water</strong> Quality<br />

A summary evaluation of elimination processes at bank-filtration sites in The Netherlands showed<br />

an effective, sustainable, and redox-independent elimination of polycyclic aromatic hydrocarbons,<br />

polychlorinated biphenyls, chloroorganic pesticides, bromo<strong>for</strong>m, dichloroaniline, nitrobenzene,<br />

nitrotoluene, and chlornitrobenzene (Stuyfzand, 1998). Anoxic conditions cause an effective<br />

elimination of pesticides, such as atrazine, diurone, and simazine, which are less (or not) degraded<br />

under aerobic conditions.<br />

During bank filtration at the Rhine River, the concentrations of DOC, AOX, and sulphur<br />

compounds in Rhine water are reduced by more than 50 percent during a mean retention time of<br />

the infiltrate in the aquifer of 6 to 8 weeks (Brauch and Kuehn, 1988). Concentrations of<br />

malodorous compounds such as geosmin and monoterpenes decrease through degradation<br />

processes during bank filtration by 80 to 99 percent (Juettner, 1995). During bank filtration at<br />

Schlachtensee, Wannsee, and Tegeler See lakes in Berlin, Germany, algae-born terpenes causing<br />

malodor of lake waters were not found in bank filtrate (Chorus et al., 1992). Complex studies on<br />

biogeochemical reactions are reported by Bourg and Bertin (1993), Sontheimer (1991), and<br />

Nestler et al. (1998).<br />

The long-term monitoring of water quality has demonstrated the effective removal of a range of<br />

contaminants, including ammonium and polar compounds. Results from monitoring programs at<br />

sites on the Rhine, Elbe, and Danube rivers and at Tegel Lake have been recently summarized by<br />

Kuehn and Mueller (2000), Brauch et al. (2000), Grischek (2002), Mucha et al. (2002), and Fritz<br />

(2002), respectively.<br />

Drinking-<strong>Water</strong> Treatment<br />

In the 1960s and 1970s, one of the major problems was bad odor of drinking water derived from<br />

bank filtrate. The main ways to solve this problem were a reducing the organic load of the river<br />

water and using GAC filtration as a post-treatment step (Sontheimer, 1980).<br />

At some sites, advanced technologies such as ozone treatment, biological filtration, or GAC<br />

adsorption were established to treat the pumped infiltrate (e.g., along the Rhine River). At other<br />

sites, simple treatment technologies, such as pH-adjustment, aeration with sand filtration <strong>for</strong> iron<br />

and manganese removal, or disinfection by chlorine, are sufficient <strong>for</strong> meeting today’s drinkingwater<br />

standards (e.g., along the Elbe River). The necessary treatment st<strong>eps</strong> are still in discussion,<br />

especially due to recent findings of persistent organic trace compounds in bank filtrate.<br />

In Germany, different tests have been developed to classify compounds into compound classes that<br />

are not removed during aquifer passage and not removed by common drinking-water treatment<br />

(Sontheimer, 1991; Mueller et al., 1993). This concept has been used to identify compounds that<br />

are problematic in terms of drinking-water quality and to initiate source control measures <strong>for</strong><br />

reducing the input of these mobile substances into receiving rivers.<br />

107


108<br />

Problems<br />

At several sites (e.g., Budapest, Hungary and Dresden, Germany, or at the lower Rhine River), the<br />

quality of backside groundwater beneath the river floodplain, especially regarding the nitrate<br />

concentration, is becoming a greater problem than bank-filtrate quality.<br />

Problems with persistent organic compounds (e.g., pharmaceuticals) are reported <strong>for</strong> bank<br />

filtration in Berlin, Germany, where the percentage of surface waters comprised of municipal<br />

sewage effluents is relatively high (Ziegler et al., 2000, 2002). Recent work has shown that<br />

pharmaceuticals, with sources including feed additives, animal drugs, and human medical care,<br />

can be persistent in the aquatic environment. Trace-level analysis shows the presence of drug<br />

residues such as clofibric acid (a metabolite of a blood lipid regulator in medical care) in municipal<br />

sewage effluents and in receiving surface waters. As polar compounds, clofibric acid and several<br />

other pharmaceuticals, such as diclofenac, propyphenazone, carbamazepine, and primidone, are<br />

highly mobile (Heberer, 2002). Findings of pharmaceutical residues in drinking water are not<br />

desirable from a hygienic point of view; nevertheless, from today’s knowledge, these low<br />

concentrations do not have any toxicological impacts on human health. A limited removal of the<br />

polar drug residues is possible using activated carbon filtration. In summary, polar-peristent<br />

synthetic organic contaminants such as pharmaceuticals and industrial compounds require greater<br />

attention to assess the possibility, if any, of attenuation during bank filtration.<br />

This topic is one of the major issues in a large-scale research program of the new Berlin Centre of<br />

Competence on <strong>Water</strong>, called NASRI, which commenced in May 2002. <strong>Research</strong> activities include<br />

several field-site investigations with bank filtration and recharge, as well as semi-technical and<br />

laboratory studies on hydrogeology, modeling, algal toxins, viruses and bacteria, natural and<br />

anthropogenic organic compounds, and the clogging processes. Results will be presented in this<br />

survey, as well as in other presentations out of the NASRI-group at this conference.<br />

Conclusions<br />

For the future, bank filtration is viewed as a promising approach to increase limited groundwater<br />

resources and to provide a sustainable pretreatment step with multiple-barrier functions regarding<br />

chemical and microbial parameters.<br />

A new period in the exchange of worldwide experience started in November 1999 with the<br />

Inernational Riverbank Filtration Conference in Louisville, Kentucky (United States), followed by a<br />

workshop on the Attenuation of Groundwater Pollution by Bank Filtration in June 2000 in Dresden,<br />

Germany, an International Riverbank Filtration Conference in November 2000 in Düsseldorf, Germany,<br />

and a NATO Advanced <strong>Research</strong> Workshop on Riverbank Filtration on Understanding Contaminant<br />

Biogeochemistry and Pathogen Removal, which was held in September 2001 in Tihany, Hungary.<br />

REFERENCES<br />

Bourg, A.C.M., and C. Bertin (1993). “Biogeochemical processes during the infiltration of river water into<br />

an alluvial aquifer.” Environ. Sci. Technol., 27(4): 661-666.<br />

Brauch, H.-J., and W. Kuehn (1988). “Organische Spurenstoffe im Rhein und bei der Trinkwasseraufbereitung.”<br />

gwf Wasser Abwasser, 129(3): 37-44 (in German).<br />

Brauch, H.-J., U. Mueller, and W. Kuehn (2000). “Experiences with riverbank filtration in Germany.”<br />

Proceedings, International Riverbank Filtration Conference, IAWR, 4: 33-39.<br />

Chorus, I., G. Klein, J. Fastner, and W. Rotard (1992). “Off-flavors in surface waters - How efficient is bank<br />

filtration <strong>for</strong> their abatement in drinking water?” Wat. Sci. Technol., 25(2): 251-258.


Fritz, B. (2002). Uferfiltration unter verschiedenen wasserwirtschaftlichen, hydrogeologischen und hydraulischen<br />

Bedingungen. (Investigations of river bank filtration influenced by different hydrogeological, hydraulic and water<br />

management systems), PhD thesis, Free University of Berlin (in German).<br />

Grischek, T. (2002). Zur Bewirtschaftung von Uferfiltratfassungen an der Elbe. (Management of bank filtration<br />

sites along the River Elbe), PhD thesis, Dresden University of Technology (in German).<br />

Grischek, T., D. Schoenheinz, E. Worch, and K. Hiscock (2002). “Bank filtration in Europe - An overview<br />

of aquifer conditions and hydraulic controls.” Management of aquifer recharge <strong>for</strong> sustainability, P. Dillon (ed.),<br />

Swets & Zeitlinger, Balkema, Lisse, 485-488.<br />

Heberer, T. (2002). “Occurrence, fate, and removal of pharmaceutical residues in the aquatic environment:<br />

a review of recent date.” Toxicology Letters, 131: 5-17.<br />

Heeger, D. (1987). Investigations on clogging of river beds, Unpublished PhD-thesis (in German).<br />

Ingerle, K., A.P. Blaschke, A. Brugger, C. Hasenleithner, G.J. Herndl, H. Jarosch, I. Kolar, H.J. Lensing, S.<br />

Pöschl, N. Queric, B. Reitner, F. Schoeller, R. Sommer, and B. Wett (1999). Forschungsprojekt Uferfiltrat<br />

(<strong>Research</strong> Project Bank Filtration), <strong>Research</strong> Initiative Verbund, Vienna 60 (in German).<br />

Kivimäki, A.L. (1995). “Production of artificially recharged groundwater using bank filtration” Publication of<br />

the <strong>Water</strong> and Environment Administration 573, Helsinki, Finland (in Finnish).<br />

Kuehn, W., and U. Mueller (2000). “Riverbank filtration: An overview.” Journal AWWA, 92, (12): 60-69.<br />

Mucha, I., D. Rodak, Z. Hlavaty, and L. Bansky (2002). “Groundwater quality processes after bank<br />

infiltration from the Danube at Cunkovo.” Riverbank Filtration: Understanding Contaminant Biogeochemistry<br />

and Pathogen Removal, C. Ray (ed.), Kluwer Academic Publishers, The Netherlands, 177-219.<br />

Mueller, U., B. Wricke, and H. Sontheimer (1993). “Wasserwerks- und trinkwasserrelevante Substanzen in<br />

der Elbe.” Vom Wasser 81, 371-386 (in German).<br />

Nestler, W., W. Walther, F. Jacobs, R. Trettin, and K. Freyer (1998). “<strong>Water</strong> production in alluvial aquifers<br />

along the River Elbe.” UFZ-<strong>Research</strong> Report 7, 203 p. (in German).<br />

Schubert, J. (2001). “How does it work? Field studies on riverbank filtration.” Proceedings, International<br />

Riverbank Filtration Conference, 2-4 Nov. 2000, Düsseldorf, Germany, IAWR-Rheinthemen 4, 41-55.<br />

Schubert, J. (2000). “Entfernung von Schwebstoffen und Mikroorganismen sowie Verminderung der<br />

Mutagenität bei der Uferfiltration (Removal of suspended matter and microorganisms and reduction of<br />

mutagenity during bank filtration).” gwf Wasser Abwasser, 14(1/4): 218-225 (in German).<br />

Sontheimer, H. (1980). “Experience with riverbank filtration along the Rhine River.” Journal AWWA,<br />

72: 386-390.<br />

Sontheimer, H. (1991). Drinking water from the River Rhine? Academia Verlag, Sankt Augustin (in German).<br />

Stuyfzand, P.J. (1998). “Fate of pollutants during artificial recharge and bank filtration in the Netherlands.”<br />

Artificial Recharge of Groundwater, Balkema, Rotterdam, 119-125.<br />

Ziegler, D., C. Hartig, S. Wischnack, and M. Jekel (2000). “Behaviour of dissolved organic compounds and<br />

pharmaceuticals during lake bank filtration in Berlin.” Proceedings, International Riverbank Filtration<br />

Conference, IAWR, 4: 151-160.<br />

Ziegler, D., C. Hartig, S. Wischnack, and M. Jekel (2002). “Organic substances in partly closed water cycles.”<br />

Management of aquifer recharge <strong>for</strong> sustainability, P. Dillon (ed.), Swets & Zeitlinger, Balkema, Lisse, 161-167.<br />

109


110<br />

<strong>Water</strong> chemist and treatment engineer MARTIN JEKEL has about 28 years of experience<br />

in the field of water and wastewater treatment, as well as in water-quality analysis. Since<br />

1988, he has been a full Professor <strong>for</strong> water-quality control at the <strong>Institute</strong> <strong>for</strong><br />

Environmental Engineering of the Technical University of Berlin. His research interests<br />

include the development of the “Muelheim Process” <strong>for</strong> drinking-water treatment, basic<br />

and applied studies on the preoxidation mechanisms in coagulation of water treatment,<br />

advanced treatment processes <strong>for</strong> indirect potable water reuse, a new adsorption technique<br />

with Granular Ferric Hydroxide <strong>for</strong> solving the worldwide problem of arsenic in groundwater, and the<br />

analysis and fate of new organic trace organics, such as iodinated X-ray contrast agents. During the last<br />

5 years, he has conducted several studies on the processes of organics removal in lake bank filtration in the<br />

Berlin area, and he is Scientific Coordinator of Natural and Artificial Systems <strong>for</strong> Recharge and Infiltration,<br />

one of the largest bank-filtration and recharge research programs worldwide. Jekel received a diploma<br />

(M.Sc.) in Chemistry, a Ph.D. in Chemical Engineering, and the German degree of Habilitation<br />

(qualification <strong>for</strong> a tenure professorship) at the University of Karlsruhe, Germany.


Session 6: Microorganisms<br />

Using Microscopic Particulate Analysis<br />

<strong>for</strong> Riverbank Filtration<br />

Jennifer L. Clancy, Ph.D.<br />

Clancy Environmental Consultants, Inc.<br />

St. Albans, Vermont<br />

William D. Gollnitz<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

Microscopic Particulate Analysis has been used since the mid-1980s as a tool to examine the<br />

nature of biological particulates in water. This methodology is used <strong>for</strong> two purposes:<br />

• To examine groundwater <strong>for</strong> the presence of biological surface-water indicators. These<br />

species may indicate that the groundwater source is GWUDI and, hence, subject to the<br />

requirements of the Surface <strong>Water</strong> Treatment Rule.<br />

• To determine filtration efficiency by comparing the types of particulates in raw and<br />

finished waters. The log removal of particulates is calculated to determine the efficiency<br />

of the filtration process. Both full-scale and pilot processes can be evaluated using<br />

Microscopic Particulate Analysis.<br />

Microscopic Particulate Analysis samples are collected by filtering a relatively large volume of<br />

water through a cartridge filter. The sample is generally several hundred gallons, but volume can<br />

be adjusted depending on the type of sample. The filter is shipped overnight on ice to the laboratory.<br />

The filter is cut apart and the fibers washed or backwashed directly (Envirochek capsule) to<br />

remove particulate matter. The particulates are concentrated into a pellet, a portion of which may<br />

be subjected to buoyant density gradient centrifugation to separate biological particulates from<br />

heavier debris. This material is examined in two ways:<br />

• Direct microscopic observation using brightfield and phase or differential interference<br />

contrast to identify, enumerate, and classify biological particulates.<br />

• A portion of the sample may be stained using fluorescent monoclonal antibodies specific<br />

to Giardia cysts and Cryptosporidium oocysts. The sample is examined using epifluorescence<br />

microscopy to detect the presence of these parasites.<br />

The data are reported as the numbers and types of biological particulates. Categories include, but<br />

are not limited to, algae, diatoms, rotifers, crustaceans, insects, protozoa, plant debris, Giardia cysts,<br />

and Cryptosporidium oocysts. The log reduction of each bio-indicator category is determined by<br />

comparing the numbers of particulates in raw and finished water samples. A filtration per<strong>for</strong>mance<br />

rating is determined, and the significance of any additional data or observations is explained.<br />

Other in<strong>for</strong>mation that can be obtained includes the presence of alum/polymer floc, carbon fines,<br />

Correspondence should be addressed to:<br />

Jennifer L. Clancy, Ph.D.<br />

President<br />

Clancy Environmental Consultants, Inc.<br />

P.O. Box 314 • St. Albans, Vermont 05478 USA<br />

Phone: (802) 527-2460 • Fax: (802) 524-3909 • Email: jclancy@clancyenv.com<br />

111


112<br />

and other material in the finished water. The overall microscopic quality of the water can be readily<br />

assessed using Microscopic Particulate Analysis. If a surface-water sample has been analyzed<br />

simultaneously, a comparison of particulates in surface-water and groundwater samples can be made.<br />

The rationale in examining groundwater using Microscopic Particulate Analysis is that if<br />

surface-water indicators are observed, then the source may be at risk <strong>for</strong> contamination by Giardia<br />

cysts and Cryptosporidium oocysts. Groundwaters that are at risk of surface-water influence are<br />

considered to be surface waters under the Surface <strong>Water</strong> Treatment Rule and are required to<br />

comply with filtration and disinfection requirements. An obvious extension of this technique is to<br />

examine water samples in systems using <strong>RBF</strong>. Similar to conducting a filtration plant efficiency<br />

study, a river or source sample can be compared to samples from the groundwater collection<br />

device, and a determination of the level of in situ natural filtration occurring through the aquifer<br />

can be made. This idea of natural filtration to improve water quality is not new. The concept of<br />

using the aquifer as a natural filtration device was termed “<strong>RBF</strong>” over 100 years ago, although the<br />

process has been used <strong>for</strong> centuries. <strong>RBF</strong> is a standard surface-water treatment practice in Europe.<br />

In 1997, the authors first proposed considering natural filtration when per<strong>for</strong>ming GWUDI<br />

evaluations of collection devices located in porous media aquifers. In a long-term study of<br />

groundwater collection devices in the North Platte Alluvial Aquifer in Casper Wyoming, the<br />

authors found that natural filtration through the aquifer produced finished water of superior<br />

quality to that produced by the water filtration plant, both using the North Platte River as a<br />

source. Surface-water indicators (algae and diatoms) were found in some of the groundwater<br />

collection devices, and always at significantly lower concentrations than in surface water. Even<br />

when algae and diatoms were present, cysts and oocysts were never observed in any of the<br />

groundwater samples. The reduction of algae and diatoms through the aquifer ranged from 4.2 to<br />

greater than 5 log, while the removal range in the plant, which consistently met all compliance<br />

requirements, ranged from 0.3 to 1.5 log.<br />

The key to providing adequate <strong>RBF</strong> is the aquifer type. Natural filtration through porous media<br />

aquifers, particularly those with a matrix in the sand- and gravel-size range, is very effective <strong>for</strong><br />

particulate removal. Natural filtration is similar to engineered filtration in that we expect to see a<br />

reduction of particulates through the process; the end result is not necessarily the observation of<br />

no surface-water indicators, but adequate removal so as to minimize the risks to public health. In<br />

the Surface <strong>Water</strong> Treatment Rule, this is defined as 2-log removal of Giardia cysts through<br />

filtration, and in the Long Term 2 Enhanced Surface <strong>Water</strong> Treatment Rule (LT2ESWTR), 3-log<br />

Cryptosporidium removal through some combination of processes, including <strong>RBF</strong> removal credit.<br />

This paper examines the filtration efficiencies of six surface water treatment plants and six systems<br />

using in situ natural filtration or <strong>RBF</strong>. Each system was monitored regularly using Microscopic<br />

Particulate Analysis <strong>for</strong> a minimum of 1 to 11 years. Samples were collected in pairs (raw and<br />

treated surface water or river water and groundwater collection device) and analyzed by a single<br />

group of analysts, minimizing some of the inherent variability noted between laboratories and<br />

analysts. Sample sites are located in the United States and Canada, and the selected sites represent<br />

a broad geographic region (Cali<strong>for</strong>nia to Connecticut). Over 1,400 samples were analyzed in this<br />

comparison.<br />

As noted in previous studies, surface-water samples showed the largest diversity and concentrations<br />

of microorganisms. Algae and diatoms occurred in the highest concentrations, ranging from l0 2 to<br />

greater than 10 9 per 100 gallons. The reduction of indicators through engineered filtration ranged<br />

from no reduction to greater than 6 log 10. Log 10 reduction of biological indicators varied within<br />

treatment plants, and poorer filtration per<strong>for</strong>mance noted using Microscopic Particulate Analysis


was independent of finished water turbidity. All plants in the study were considered well-operated<br />

and met the requirements of the Surface <strong>Water</strong> Treatment Rule. In some cases, the levels of algae<br />

and diatoms in finished water were actually higher than that observed in raw water. This may be<br />

due to sampling anomalies, growth of algae in the filters, poor filtration per<strong>for</strong>mance, or vagaries<br />

associated with the Microscopic Particulate Analysis method itself.<br />

For <strong>RBF</strong> groundwater collection devices, log 10 reduction of algae and diatoms ranged from about<br />

4 log 10 to greater than 7 log 10. Reductions across the aquifer through natural filtration were far<br />

more consistent, and the minimum level of reduction was always greater than that required <strong>for</strong><br />

filtration per<strong>for</strong>mance under the Surface <strong>Water</strong> Treatment Rule (2-log 10 Giardia and 3-log 10<br />

Cryptosporidium). No Giardia cysts or Cryptosporidium oocysts were noted in any of the <strong>RBF</strong><br />

groundwater collection device samples.<br />

While all samples met turbidity requirements, the biological water quality measured using<br />

Microscopic Particulate Analysis was far superior in the <strong>RBF</strong> groundwater collection device<br />

samples than that noted in the treatment plant effluents. This same observation has been noted<br />

when comparing surface-water treatment plant effluents and groundwater samples in previous<br />

studies. <strong>RBF</strong> is a more effective process than engineered filtration <strong>for</strong> removing biological particles<br />

from surface-water sources.<br />

JENNIFER CLANCY is a microbiologist with 25 years of experience in environmental<br />

microbiology, focusing on water. She is currently President of Clancy Environmental<br />

Consultants, Inc., which provides microbiological testing and consulting services to the<br />

water and wastewater industries. Clancy has spent years studying the issue of groundwater<br />

under the direct influence of surface water in systems in the United States and Canada.<br />

Prior to <strong>for</strong>ming Clancy Environmental Consultants, Inc., in 1994, she was the Director<br />

of <strong>Water</strong> Quality at the Erie County <strong>Water</strong> Authority in Buffalo, New York, and was<br />

responsible <strong>for</strong> administration of the <strong>Water</strong> Quality Department. Clancy received a B.S. in Microbiology<br />

from Cornell University, an M.S. in Microbiology and Biochemistry from the University of Vermont, a Ph.D.<br />

in Microbiology and Immunology from McGill University, and an M.S. in Environmental Law from Vermont<br />

Law School.<br />

113


114


Session 6: Microorganisms<br />

Transport and Removal of Cryptosporidium Oocysts<br />

in Subsurface Porous Media<br />

Menachem Elimelech, Ph.D.<br />

Yale University<br />

New Haven, Connecticut<br />

Garrett Miller<br />

Yale University<br />

New Haven, Connecticut<br />

Zachary Kuznar<br />

Yale University<br />

New Haven, Connecticut<br />

Cryptosporidium parvum contamination is considered one of the most important water-supply<br />

problems today. Disinfectants commonly used by water treatment plants are ineffective at reducing<br />

Cryptosporidium parvum risk, and deep-bed filtration represents one of the main barriers to oocyst<br />

contamination. <strong>RBF</strong>, which is gaining popularity in the United States, presents an alternative<br />

method of pretreating water to remove oocysts. Very little research has been done so far to<br />

understand the mechanisms controlling the transport and filtration behavior of Cryptosporidium in<br />

saturated porous media under chemical and physical conditions relevant to <strong>RBF</strong>. The objective of<br />

this study was to elucidate the role of electrostatic double layer interactions in the attachment and<br />

transport of Cryptosporidium in flow through saturated porous media. Well-controlled column<br />

experiments were carried out using ultra-clean quartz sand and Cryptosporidium parvum oocysts.<br />

Complimentary experiments using a stagnation point flow system have been conducted under<br />

identical chemical and hydrodynamic conditions. Initial bacterial cell deposition rates are compared<br />

with column breakthrough curves, and the results are used to highlight the controlling mechanisms<br />

of Cryptosporidium parvum adhesion and transport.<br />

MENACHEM ELIMELECH is the Llewellyn West Jones Professor of Environmental<br />

Engineering and Director of the Environmental Engineering Program at Yale University.<br />

His research interests center on problems involving physicochemical and colloidal<br />

processes in engineered and natural systems, and he has worked on the dynamics of colloid<br />

transport and deposition in porous media, transport and fate of microbial particles (viruses,<br />

bacteria, and Cryptosporidium) in the subsurface environment, and contaminant removal<br />

by membrane processes. Elimelech is the author of over 90 journal publications and is the<br />

principal author of Particle Deposition and Aggregation (1995). Among his recent honors, he was the recipient<br />

of the Association of Environmental Engineering and Science Professors (AEESP) Outstanding Paper Award<br />

in 2002 and the AEESP Doctoral Dissertation Award, with his graduate student Eric M.V. Hoek, in 2002.<br />

He also has served on the Editorial Advisory Boards of Environmental Science & Technology, Environmental<br />

Engineering Science, Desalination, and Journal of Colloid and Interface Science. Elimelech received both a B.S. in<br />

Soil and <strong>Water</strong> Sciences and an M.S. in Environmental Science and Technology from the Hebrew<br />

University in Jerusalem and a Ph.D. in Environmental Engineering from The Johns Hopkins University.<br />

Correspondence should be addressed to:<br />

Menachem Elimelech, Ph.D.<br />

Llewellyn West Jones Professor of Environmental Engineering<br />

Department of Chemical Engineering • Environmental Engineering Program<br />

Yale University • P.O. Box 208286 • New Haven, Connecticut 06520-8286 USA<br />

Phone: (203) 432-2789 • Fax: (203) 432-2881 • Email: menachem.elimelech@yale.edu<br />

115


116


Session 6: Microorganisms<br />

Laboratory and Field Strategies <strong>for</strong> Assessing<br />

Pathogen Removal by Riverbank Filtration<br />

Monica B. Emelko, Ph.D.<br />

University of <strong>Water</strong>loo<br />

<strong>Water</strong>loo, Ontario, Canada<br />

Mark T. Watling<br />

University of <strong>Water</strong>loo<br />

<strong>Water</strong>loo, Ontario, Canada<br />

Martin M. Côté<br />

University of <strong>Water</strong>loo<br />

<strong>Water</strong>loo, Ontario, Canada<br />

Introduction<br />

<strong>RBF</strong> systems can significantly reduce the concentrations of many surface-water pollutants<br />

(Shubert, 2000; Wang et al., 2000; Kuehn and Mueller, 2000); however, predicting and<br />

quantifying those reductions is difficult. An understanding of subsurface pathogen transport,<br />

particularly Cryptosporidium and viruses, is critical <strong>for</strong> utilities faced with potential GWUDI of<br />

surface-water wells. In the United States, the LT2ESWTR prescribes Cryptosporidium removal<br />

credits (0, 0.5, or 1.0 log) based on aquifer grain-size distribution and the distance between the<br />

well and riverbed (USEPA, 2001). Regulations in Ontario, Canada, require utilities to<br />

demonstrate in situ filtration per<strong>for</strong>mance using particle counts (less than 100 particles greater<br />

than or equal to 10 micrometers per milliliter) and qualitative assessments of potential temporal<br />

changes in well effluent particle counts and raw-water quality (Ontario Ministry of the<br />

Environment, 2001). While these approaches represent an attempt at demonstrating the efficacy<br />

of <strong>RBF</strong> as a treatment technology <strong>for</strong> reducing pathogen concentrations in drinking-water sources,<br />

it is generally acknowledged that they do not provide adequate assessments of pathogen removal<br />

by <strong>RBF</strong>.<br />

Assessments of pathogen removal by <strong>RBF</strong> are difficult. Although column studies may approximate<br />

some <strong>RBF</strong> per<strong>for</strong>mance, they are of limited value because they often fail to adequately represent<br />

temporal changes in regional groundwater conditions or microbiological activity, geologic<br />

heterogeneity, and physical scale (filtration length) of the true system. While full-scale<br />

investigations are possible and have been conducted, they are costly and plagued by analytical<br />

limitations such as low indigenous pathogen concentrations (that preclude representative<br />

sampling) and, in the case of Cryptosporidium, unreliable analytical methods with low and highly<br />

variable recoveries (Nieminski et al., 1995; Clancy et al., 1999). Full-scale investigations are<br />

further complicated by the use of more readily analyzed, but unproven, surrogates (which is<br />

Correspondence should be addressed to:<br />

Monica B. Emelko, Ph.D.<br />

Assistant Professor<br />

Department of Civil Engineering<br />

University of <strong>Water</strong>loo • <strong>Water</strong>loo, Ontario N2L 3G1 Canada<br />

Phone: (519) 888-4567 ext. 2208 • Fax: (519) 888-6197 • Email: mbemelko@uwaterloo.ca<br />

117


118<br />

necessitated because of the a<strong>for</strong>ementioned analytical limitations). As a result, the outcomes of<br />

full-scale per<strong>for</strong>mance or surrogate validation studies are difficult to interpret and compare,<br />

hindering the development of widely accepted protocols <strong>for</strong> assessing pathogen removal by <strong>RBF</strong>.<br />

Objectives<br />

In this paper, we highlight some recent laboratory and analytical advances that are useful <strong>for</strong><br />

improving assessments of pathogen removal by <strong>RBF</strong>. The specific objectives of this work include:<br />

• Investigating the impact of column orientation (and the associated impacts of<br />

gravitational <strong>for</strong>ce) on pathogen removal in laboratory studies investigating <strong>RBF</strong><br />

per<strong>for</strong>mance.<br />

• Quantitatively demonstrating the impact of the actual number of discrete particles<br />

(e.g., pathogens or surrogates) counted from a sample on the uncertainty of inferences<br />

drawn from experimental and monitoring data.<br />

• Proposing pathogen or surrogate count targets so that reasonably certain conclusions can<br />

be drawn from experimental or monitoring data.<br />

Methodology<br />

Column Studies<br />

It has been suggested that column orientation (horizontal versus vertical versus angled) may be an<br />

important parameter in mimicking subsurface transport and filtration due to the impacts of<br />

sedimentation (Harvey, 1997). To investigate the impact of the gravitational <strong>for</strong>ce on pathogen<br />

removal during column studies, four 10-cm filter columns with 6-cm internal diameters were<br />

packed with sieved aquifer material and were operated at conditions typical of those that may be<br />

encountered during <strong>RBF</strong> (flow of ~1 milliliter per minute). Two of the columns were operated in<br />

a horizontal mode and two of the columns were operated in an upflow vertical mode. Preliminary<br />

investigations were conducted at water-quality conditions that do not favor pathogen removal<br />

(i.e., low ionic strength).<br />

After the columns were flushed with degassed water <strong>for</strong> several hours, the influent reservoir —<br />

which fed the four columns simultaneously — was spiked with both Cryptosporidium oocysts and<br />

oocyst-sized polystyrene microspheres. The influent oocyst and microsphere concentrations were<br />

determined from influent reservoir samples. Effluent samples were collected at intervals of<br />

0.25-pore volumes until at least 6-pore volumes of influent had passed through the columns.<br />

Microsphere removals from samples collected during the passage of 0.3- to 5.2-pore volumes<br />

through the columns are discussed herein.<br />

Polycarbonate membranes (25-millimeter, 0.40-micrometer nominal porosity) were used to filter<br />

the samples. Samples were filtered directly on a manifold with a vacuum of ~5-inches of mercury.<br />

Microsphere enumeration was per<strong>for</strong>med at 100× and 400× magnification (Nikon Labophot 2A,<br />

Nikon Canada Inc., Toronto, Ontario, Canada). The filtered sample volumes were selected to<br />

yield between 10 and 1,000 microspheres per membrane.<br />

Data Reliability and Interpretation<br />

The occurrence of waterborne pathogens such as Cryptosporidium is difficult to measure precisely<br />

because they are present in varied concentrations, often at concentrations so low that detection<br />

is difficult (Lisle and Rose, 1995). Not surprisingly, current methods <strong>for</strong> quantifying


Cryptosporidium are unreliable, laborious, and expensive. Analytical recoveries are often low and<br />

highly variable (Nieminski et al., 1995; Clancy et al., 1999).<br />

Nahrstedt and Gimbel (1996) proposed a model <strong>for</strong> the process of estimating the concentration<br />

of Cryptosporidium oocysts in a body of water; it addresses uncertainty associated with sampling and<br />

enumeration. The authors used:<br />

• A Poisson distribution to model true sample counts (representative sampling).<br />

• A binomial distribution to model the recovered fraction of oocysts (random analytical error).<br />

• A Beta distribution to describe non-constant analytical recovery.<br />

The authors provided only limited in<strong>for</strong>mation on how to apply their model to obtain<br />

concentration estimates and to quantify the uncertainty in these estimates.<br />

To obtain both concentration and removal estimates and to quantify the associated uncertainties,<br />

Emelko (2001) presented a Bayesian analysis of the Nahrstedt and Gimbel (1996) model and<br />

utilized the Gibbs sampler to provide easy access to a wide variety of estimated quantities<br />

(e.g., point estimates, probability [confidence] intervals, probability density functions, etc.). In<br />

brief, this approach results in a joint posterior probability density function <strong>for</strong> the true<br />

concentration of pathogens in the water body (c), the probability of pathogen recovery from the<br />

water sample (p), and the number of pathogens in the water sample (N) given the observed<br />

pathogen count (X). The uncertainty of recovery is accounted <strong>for</strong> by the Beta distribution’s shape<br />

parameters (a and b). Replicate sampling is incorporated, if applicable (n replicate samples). The<br />

posterior probability density function, after simplification and suppressing constant multipliers, is:<br />

Df [(c,p,Ni|Xi),i = 1,…, n] ∝ c –1+ Σ N i×e –c+Σ V i×p a–1+Σ X i×(1–p) b–1+Σ(N i –X i ) ×Π<br />

Since the parameter of interest is the true concentration c, the remaining (nuisance) parameters<br />

may be integrated out (Box and Tiao, 1973). This operation finds the mathematical expectation<br />

of c over all values of the nuisance parameters. The posterior distribution given by Equation 1 is<br />

very difficult, as it stands, to integrate or to use directly to make inferences about c. The Gibbs<br />

sampler, a Monte Carlo method that produces an indefinitely long Markov chain of vectors of<br />

parameter values (Geman and Geman, 1984), can handle these problems with relative ease.<br />

Applicable to all discrete particles, this technique was used in the present investigation; details<br />

regarding the development of this approach were provided in Emelko (2001).<br />

The impact of the actual number of pathogens or surrogates counted from a sample on uncertainty<br />

(confidence interval <strong>for</strong> pathogen concentration) is demonstrated herein using a concentration of<br />

one pathogen per 100 liters. Counts between one and 1,000 pathogens (and the associated sample<br />

volumes that would yield a concentration of one pathogen per 100 liters) were utilized. The recovery<br />

data and associated Beta parameters (a = 28.12, b = 8.43) used in this assessment correspond to oocyst<br />

recoveries ranging from 69 to 85 percent and average 77 percent, as reported by Emelko (2001).<br />

Results<br />

Column Studies<br />

n<br />

i=1<br />

n<br />

i=1<br />

Preliminary investigations of Cryptosporidium oocyst-sized polystyrene microsphere removal by<br />

horizontal and upflow vertical columns are summarized in a box and whisker plot (Figure 1). This<br />

figure illustrates that a consistent level of microsphere removal was achieved during the period<br />

n<br />

i=1<br />

n<br />

i=1<br />

n<br />

i=1<br />

Vi N i<br />

(N i –N i)<br />

(1)<br />

119


120<br />

Microsphere Removal (log 10 )<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

Vertical #1 Vertical #2 Horizontal #1 Horizontal #2<br />

Figure 1. Cryptosporidium-sized microsphere removal by 10-centimeter columns operated in upflow vertical<br />

and horizontal orientations.<br />

that 0.3- to 5.2-pore volumes of water passed through the columns. In addition, the data suggest<br />

slightly higher removal of oocyst-sized microspheres by the horizontal columns. Moreover,<br />

excellent reproducibility between duplicate columns was achieved; this is also evident in the<br />

breakthrough curves (not shown). It should be noted that some microsphere loss occurred in the<br />

sample line prior to reaching the columns. The overall microsphere removals are, there<strong>for</strong>e,<br />

somewhat high; however, given that the columns received the same influent water (losses<br />

occurred be<strong>for</strong>e the feed lines were split), the relative relationship between the columns should be<br />

unaffected. Continued experiments will be useful in determining whether these differences in<br />

microsphere removal as a function of column orientation are statistically significant.<br />

Data Reliability and Interpretation<br />

n=8 n=8 n=8 n=8<br />

Column<br />

The impact of the actual number of pathogens or surrogates counted from a sample on uncertainty<br />

(confidence interval <strong>for</strong> pathogen concentration) was demonstrated using a concentration of<br />

one pathogen per 100 liters (Figure 2). Figure 2 illustrates that the uncertainty associated with<br />

pathogen concentration data can be considerably reduced by increasing the number of pathogens<br />

that are counted in a sample; naturally, this can be achieved by increasing the processed sample<br />

volume. Similar to the present analysis, Emelko and Reilly (2002) demonstrated the same impact<br />

on uncertainty when examining confidence intervals <strong>for</strong> pathogen removal (as opposed to<br />

concentration) data. To avoid confidence intervals that extend beyond an order of magnitude,<br />

counts of approximately 10 or more pathogens (preferably, approximately 50 pathogens) are<br />

required from an individual sample. Only limited improvements in uncertainty are observed at<br />

counts greater than 50 pathogens per sample.<br />

It is important to note that the outcomes in Figure 2 are associated with the recovery profile of<br />

one analytical method. Somewhat different outcomes with respect to the range of the confidence<br />

intervals would be expected with different analytical methods (and their associated recovery<br />

profiles); however, the same general trend of increased counts resulting in decreased uncertainty


Pathogen Concentration (pathogens/100 Liters)<br />

10.<br />

1.<br />

0.1<br />

0.01<br />

0.001<br />

Figure 2. Ninety-five percent confidence intervals obtained <strong>for</strong> pathogen concentrations using the<br />

Cryptosporidium recovery profile of Emelko (2001).<br />

would be expected. This type of statistical analysis <strong>for</strong> already reported data can be somewhat<br />

difficult as adequate recovery data (actual counts) are often lacking from the reported literature,<br />

underscoring the need <strong>for</strong> guidance on how to adequately report pathogen recovery in<strong>for</strong>mation.<br />

The analysis in Figure 2 does not address the issues of replicate sampling or pooled data. Emelko<br />

and Reilly (2002) briefly discussed this issue and demonstrated that the largest improvements<br />

associated with decreasing data uncertainty were associated with increased total counts, regardless<br />

of the number of samples from which they were derived. The impact of pathogen recovery profiles<br />

on this relationship remains to be fully demonstrated.<br />

Conclusions<br />

1 2 5 10 50 100 500 1000<br />

Pathogen Count in Sample<br />

Several preliminary conclusions have resulted from this work. They include the following:<br />

• Pathogen removal appears to be slightly higher in horizontal laboratory columns as<br />

compared to upflow vertical columns. As more data are collected, the statistical significance<br />

of this outcome must be determined.<br />

• Column studies that are used to investigate <strong>RBF</strong> should be conducted using a column<br />

orientation that is as representative of the site as possible; in most cases, this orientation<br />

should be at or near horizontal.<br />

• To minimize the uncertainty associated with pathogen data so that it ranges over less<br />

than an order of magnitude, total counts of between 10 and 50 pathogens are required.<br />

This can be achieved via increased sample volume processing or increased replication.<br />

121


122<br />

REFERENCES<br />

Box, G.E., and G.C. Tiao (1973). Bayesian Inference in Statistical Analysis, Addison Wesley Publishing<br />

Company, Reading, MA.<br />

Clancy, J.L., Z. Bukhari, R.M. McCuin, Z. Matheson, and C.R. Fricker (1999). “USEPA Method 1622.”<br />

Journal AWWA, 91(9): 60-68.<br />

Emelko, M.B. (2001). Removal of Cryptosporidium parvum by granular media filtration, Ph.D. Dissertation,<br />

University of <strong>Water</strong>loo, <strong>Water</strong>loo, Ontario, Canada.<br />

Emelko, M.B., and P.M. Reilly (2002). “Reporting and Regulating Cryptosporidium Concentrations and<br />

Removals.” Proceedings, AWWA WQTC, AWWA, Denver, CO.<br />

Geman, S., and D. Geman (1984). “Stochastic Relaxation, Gibbs Distributions, and the Bayesian Restoration<br />

of Images.” IEEE Trans. Pattn. Anal. Bach. Intell., 6: 721.<br />

Harvey, R.W. (1997). “In Situ Laboratory Methods to Study Subsurface Microbial Transport.” Manual of<br />

Environmental Microbiology, C.J. Hurst, G.R. Knudsen, M.J. McInerney, L.D. Stetzenbach, and M.V. Walter<br />

(eds.), ASM Press, Washington, D.C., p. 586-589.<br />

Kuehn, W., and U. Mueller (2000). “Riverbank filtration: An overview.” Journal AWWA, 82(12): 60-69.<br />

Lisle, J.T., and J.B. Rose (1995). “Cryptosporidium Contamination of <strong>Water</strong> in the U.S.A. and U.K.: A Mini<br />

Review.” Jour. <strong>Water</strong> SRT–Aqua, 44(3): 103.<br />

Nahrstedt, A., and R. Gimbel (1996). “A Statistical Method <strong>for</strong> Determining the Reliability of the Analytical<br />

Results in the Detection of Cryptosporidium and Giardia in <strong>Water</strong>.” Jour. <strong>Water</strong> SRT – Aqua, 45(3): 101.<br />

Nieminski, E.C., F.W. Schaefer, and J.E. Ongerth (1995). “Comparison of two Methods <strong>for</strong> Detection of<br />

Giardia Cysts and Cryptosporidium Oocysts in <strong>Water</strong>.” Appl. and Envir. Microbiol., 61(5): 1714.<br />

Ontario Ministry of the Environment (MOE) (2001). Terms of Reference, Hydrogeological Study to Examine<br />

Groundwater Sources Potentially Under Direct Influence of Surface <strong>Water</strong>, Ontario Ministry of the Environment.<br />

Shubert, J. (2000). “How does it work? Field Studies on Riverbank Filtration.” Proceedings, International<br />

Riverbank Filtration Conference, Düsseldorf, 2-4 November 2000, Internationale Arbeitsgemeinschaft der<br />

Wasserwerke im Rheineinsugsgebiet (IAWR), Düsseldorf.<br />

USEPA (2001). <strong>National</strong> Primary Drinking <strong>Water</strong> Regulations: Long Term 2 Enhanced Surface <strong>Water</strong> Treatment<br />

Rule, 40 CFR Parts 9, 141 and 142, United States Environmental Protection Agency.<br />

Wang, J.Z., R. Song, and S.A. Hubbs (2000). “Particle removal through riverbank filtration process.”<br />

Proceedings, International Riverbank Filtration Conference, Düsseldorf, 2-4 November 2000, Internationale<br />

Arbeitsgemeinschaft der Wasserwerke im Rheineinsugsgebiet (IAWR), Düsseldorf.<br />

MONICA EMELKO has research expertise in microbial pathogen and surrogate transport<br />

and removal by porous media, developing analytical methods <strong>for</strong> enumerating<br />

microorganisms during porous media investigations, and developing statistical tools <strong>for</strong><br />

describing analytical uncertainty. She was awarded the American <strong>Water</strong> Works<br />

Association Academic Achievement Award <strong>for</strong> her doctoral dissertation, and has over<br />

25 publications on pathogen transport and filtration in porous media systems. Emelko has<br />

been an advisor in the development of the Long-Term 2 Enhanced Surface <strong>Water</strong><br />

Treatment Rule and serves on several American <strong>Water</strong> Works Association <strong>Research</strong> Foundation advisory<br />

committees <strong>for</strong> projects focused on characterizing filter effluents and filtration per<strong>for</strong>mance; she is also the<br />

Vice-Chair of the Particulate Contaminants <strong>Research</strong> Committee <strong>for</strong> the American <strong>Water</strong> Works Association.<br />

At present, she is an Assistant Professor in the Department of Civil Engineering at the University of<br />

<strong>Water</strong>loo in Canada. Emelko received B.S degrees in Chemical Engineering and Environmental Engineering<br />

from the Massachusetts <strong>Institute</strong> of Technology, an M.S. in Civil Engineering from the University of<br />

Cali<strong>for</strong>nia, Los Angeles, and a Ph.D. in Civil Engineering from the University of <strong>Water</strong>loo.


Session 6: Microorganisms<br />

Fate of Disinfection Byproduct Precursors<br />

and Microorganisms During Riverbank Filtration<br />

W. Joshua Weiss<br />

The Johns Hopkins University<br />

Baltimore, Maryland<br />

Edward J. Bouwer, Ph.D.<br />

The Johns Hopkins University<br />

Baltimore, Maryland<br />

William P. Ball<br />

The Johns Hopkins University<br />

Baltimore, Maryland<br />

Charles R. O’Melia, Ph.D.<br />

The Johns Hopkins University<br />

Baltimore, Maryland<br />

Kellogg J. Schwab, Ph.D.<br />

Bloomberg School of Public Health, The Johns Hopkins University<br />

Baltimore, Maryland<br />

Binh T. Le<br />

Bloomberg School of Public Health, The Johns Hopkins University<br />

Baltimore, Maryland<br />

Ramon Aboytes, D.V.M., Ph.D.<br />

American <strong>Water</strong>, Belleville Laboratory<br />

Belleville, Illinois<br />

Experience with <strong>RBF</strong> in Europe and more recently in the United States has demonstrated<br />

significant improvements in raw-water quality, including the removal of NOM, biodegradable<br />

compounds, pesticides, microbes, and other water-quality contaminants and compensation <strong>for</strong><br />

shock loads of chemical contaminants (Kuehn and Mueller, 2000; Ray et al., 2002a and 2002b;<br />

Tufenkji et al., 2002; Weiss et al., 2003a and 2003b). Because of these potential improvements,<br />

regulators and utilities in the United States have recently looked more strongly at <strong>RBF</strong> as a means<br />

<strong>for</strong> providing high-quality sources <strong>for</strong> drinking water; however, little data are available to compare<br />

the per<strong>for</strong>mance of <strong>RBF</strong> with that of conventional drinking-water treatment processes more<br />

commonly used in the United States (e.g., coagulation, flocculation, sedimentation) from identical<br />

river-water sources, especially with regard to the removal of organic DBP precursor material. In<br />

addition, little is known about the extent to which <strong>RBF</strong> may serve to reliably remove Giardia,<br />

Cryptosporidium, and other pathogens (e.g., bacteria, viruses) from river water. In particular, data<br />

are needed on the transport of microbial pathogens through riverbank systems relative to that of<br />

more easily measured indicator parameters, such as particles and coli<strong>for</strong>m bacteria.<br />

Correspondence should be addressed to:<br />

W. Joshua Weiss<br />

Ph.D. Student/<strong>Research</strong> Assistant<br />

Department of Geography and Environmental Engineering<br />

The Johns Hopkins University • 3400 N. Charles Street • Baltimore, Maryland 21218 USA<br />

Phone: (410) 516-6220 • Fax: (410) 516-8996 • Email: jweiss@jhu.edu<br />

123


124<br />

In the above context, research was conducted to document water-quality benefits during <strong>RBF</strong> at<br />

three major river sources in the Midwestern United States (the Ohio River at Jeffersonville,<br />

Indiana; Wabash River at Terre Haute, Indiana; and Missouri River at Parkville, Missouri),<br />

specifically with regard to reducing DBP precursor organic matter and microbial pathogens.<br />

Specific objectives were to:<br />

1. Evaluate the merits of <strong>RBF</strong> <strong>for</strong> removing organic DBP precursor material.<br />

2. Evaluate whether <strong>RBF</strong> can improve finished drinking-water quality by removing and/or<br />

altering NOM in a manner that is not otherwise accomplished through conventional<br />

processes of drinking-water treatment (e.g., coagulation, flocculation, sedimentation).<br />

3. Evaluate changes in the character of NOM upon ground passage from the river to wells.<br />

4. Evaluate the merits of <strong>RBF</strong> <strong>for</strong> removing pathogenic microorganisms.<br />

5. Compare the transport of microbial pathogens with that of some potential surrogate or<br />

indicator parameters (e.g., particles, turbidity, coli<strong>for</strong>ms, aerobic and anaerobic spores,<br />

diatoms, bacteriophage).<br />

To address Objectives 1, 2, and 3, samples of river source waters and bank-filtered well waters from<br />

the three study sites were analyzed <strong>for</strong> a range of water-quality parameters, including:<br />

• TOC.<br />

• DOC.<br />

• UV-absorbance at 254-nm (UV254). • Biodegradable dissolved organic carbon.<br />

• Biological AOC.<br />

• Inorganic species.<br />

• DBP <strong>for</strong>mation potential.<br />

In the second year of the project, river waters were subjected to a bench-scale conventional<br />

treatment train consisting of coagulation, flocculation, sedimentation, glass-fiber filtration, and<br />

ozonation. The treated river waters were compared with the bank-filtered waters in terms of TOC,<br />

DOC, UV 254, and DBP <strong>for</strong>mation potential. In the third and fourth years of the project, NOM<br />

from the river and well waters was characterized using the XAD-8 resin adsorption fractionation<br />

method (Leenheer, 1981; Thurman and Malcolm, 1981). XAD-8 adsorbing (hydrophobic) and<br />

non-adsorbing (hydrophilic) fractions of the river and well waters were compared with respect to<br />

DOC, UV 254, and DBP <strong>for</strong>mation potential to determine whether <strong>RBF</strong> alters the character of the<br />

source-water NOM upon ground passage and, if so, which fractions are preferentially removed.<br />

The ongoing research to address Objectives 4 and 5 consists of:<br />

• Field studies at the three study sites to document actual changes in microorganism<br />

concentrations upon subsurface travel between the rivers and wells.<br />

• Parallel laboratory column studies with riverbank aquifer media to provide insights into<br />

process mechanisms so that reliable treatment credits <strong>for</strong> pathogen removal can be<br />

established and the suitability of using indicator parameters <strong>for</strong> pathogens that are<br />

difficult to measure in these systems can be determined.<br />

The results of the DBP-precursor study demonstrate the effectiveness of <strong>RBF</strong> at removing organic<br />

precursors to potentially carcinogenic DBPs. When compared to a bench-scale conventional


treatment train optimized <strong>for</strong> turbidity removal, <strong>RBF</strong> per<strong>for</strong>med as well as treatment at one of the<br />

sites and substantially better than treatment at the other two sites in terms of removing organic<br />

carbon and DBP-precursor material. Removals of TOC and DOC upon <strong>RBF</strong> at the three sites<br />

generally ranged from 30 to 70 percent, compared to 20- to 50-percent removals upon bench-scale<br />

treatment of river waters. Reductions in precursor material <strong>for</strong> a variety of DBPs (THMs, HAAs,<br />

haloacetonitriles, haloketones, chloral hydrate, and chloropicrin) upon <strong>RBF</strong> ranged from<br />

50 to 100 percent using both the <strong>for</strong>mation potential (FP) and uni<strong>for</strong>m <strong>for</strong>mation conditions<br />

(UFC) tests (Standard Methods, 1998; Summers et al., 1996), while reductions upon bench-scale<br />

treatment were generally in the range of 40 to 80 percent. The substantially higher reductions of<br />

the DBP precursors relative to those of TOC and DOC indicate a preferential reduction upon<br />

ground passage in the NOM that reacts with chlorine to <strong>for</strong>m DBPs.<br />

Upon both bench-scale conventional treatment and <strong>RBF</strong>, a shift was observed in DBP <strong>for</strong>mation<br />

from the chlorinated to the more brominated species due to the removal of DOC relative to<br />

bromide upon treatment or <strong>RBF</strong>. Brominated THMs have greater toxicity than chloro<strong>for</strong>m, so the<br />

shift from chlorinated to brominated DBP species means that the reduction in risk is not directly<br />

proportional to the reduction in DBP <strong>for</strong>mation. Risk calculations demonstrated the ability of<br />

<strong>RBF</strong> in all cases to reduce the theoretical excess cancer risk due to THMs <strong>for</strong>med upon<br />

chlorination, and with substantially better per<strong>for</strong>mance than the bench-scale treatment train.<br />

The results of the NOM characterization study indicate that <strong>RBF</strong> appears to be equally capable of<br />

removing material of different character. The different removal mechanisms in the subsurface<br />

(e.g., sorption, biodegradation, filtration) combine to provide similar removal of the operationally<br />

defined hydrophilic and hydrophobic fractions of organic material upon ground passage. Thus, the<br />

observed reductions in DBP <strong>for</strong>mation upon <strong>RBF</strong> were largely the result of a decrease in the NOM<br />

concentration rather than a consistent change in NOM character.<br />

Field monitoring of a number of microorganisms on a tri-monthly basis between 1999 and 2000<br />

(Table 1) indicated that <strong>RBF</strong> may also serve as a significant barrier <strong>for</strong> removing microbial<br />

contaminants, including human pathogens. The monitoring data demonstrated:<br />

• Greater than 3-log removal of Clostridium spores.<br />

• Greater than 2.5-log removal of E. coli C bacteriophage (somatic phage host).<br />

• Greater than1.9-log removal of E. coli Famp bacteriophage (male-specific host).<br />

Log removals were calculated using “average” concentrations determined <strong>for</strong> each organism as the<br />

sum of the counts over all sampling rounds divided by the sum of the volumes sampled over all<br />

sampling rounds (Parkhurst and Stern, 1998); non-detects were treated as being zero (in the event<br />

of no detects over all sampling rounds, the average concentration is reported as being less than<br />

one count divided by the sum of the volumes sampled over all sampling rounds). Assuming that<br />

these indicator organisms can be used as surrogates <strong>for</strong> Giardia cysts and human enteric viruses,<br />

<strong>RBF</strong> at the three study sites met or surpassed the per<strong>for</strong>mance requirements in the United States<br />

<strong>for</strong> conventional coagulation, sedimentation, and filtration (i.e., 2.5-log removal <strong>for</strong> Giardia cysts<br />

and 2.0-log removal of viruses). More recent monthly field-monitoring results (Table 2) indicate<br />

greater than 0.8-log reduction of Bacillus, greater than 3-log reduction of Clostridium, greater than<br />

1.8-log reduction of bacteriophage MS-2, and greater than 3.7-log reduction of bacteriophage<br />

φX174 concentrations in bank-filtered waters relative to river waters. Cryptosporidium and Giardia<br />

were occasionally detected in river waters and were below the detection limits during all monthly<br />

sampling events in well waters.<br />

125


126<br />

Table 1. Results from 1999 to 2000 Tri-Monthly Field Monitoring<br />

(Seven Total Sampling Rounds) a,b<br />

Clostridium E. coli C<br />

Bacteriophage<br />

c E. coli F-ampc Average counts/<br />

100 mL<br />

Indiana American <strong>Water</strong> at Jeffersonville, Indiana<br />

pfu /100 mL pfu /100 mL<br />

Ohio River 122 49 12<br />

Well 9 3.2] 2.8] 0.09 [2.1]<br />

Well 2 3.2] 2.8] 2.2]<br />

Indiana American <strong>Water</strong> at Terre Haute, Indiana<br />

Wabash River 183 147 13<br />

Collector Well 0.07 [3.4] 3.3] 2.3]<br />

Well 3 3.4] 3.3] 2.3]<br />

Missouri American <strong>Water</strong> at Parkville, Missouri<br />

Missouri River 143 31 6<br />

Well 4 3.3] 2.6] 1.9]<br />

Well 5 3.3] 2.6] 1.9]<br />

a<br />

Concentrations calculated as Σ (counts <strong>for</strong> all sampling rounds)/Σ (volume sampled <strong>for</strong> all sampling rounds); see text.<br />

b<br />

Log removals shown in brackets.<br />

c<br />

E. coli C and E. coli F-amp are the host organisms.<br />

pfu = Plaque-<strong>for</strong>ming unit.<br />

Because of low and variable concentrations of pathogens such as Cryptosporidium in river waters,<br />

it is difficult to evaluate log removals <strong>for</strong> such parameters. Laboratory-scale column studies with<br />

riverbank media are currently underway to compare the behavior of pathogens (viruses Poliovirus<br />

and Feline Calicivirus; bacteria E. coli; and protozoans Giardia and Cryptosporidium) with that of<br />

a number of potential surrogate parameters. The potential surrogate or indicator parameters being<br />

studied include:<br />

• Bacteriophage MS-2 and PRD-1 (<strong>for</strong> human viruses).<br />

• Particles (latex and native river particles).<br />

• Turbidity.<br />

• Aerobic and anaerobic spores.<br />

Column studies are being conducted under a variety of physical/chemical conditions, with pH,<br />

ionic strength, multi-valent cation concentration (Ca 2+ ), NOM concentration, and flow rate as the<br />

variables, with laboratory-prepared water and river water collected from full-scale <strong>RBF</strong> systems.<br />

Column experiments indicate that bacteriophage can travel through riverbank sediment columns<br />

to a greater extent than Poliovirus, suggesting that bacteriophage may be useful as conservative<br />

indicators of human viruses. Continuing studies will further explore this relationship under a<br />

variety of physical/chemical conditions. Similar studies will explore the relationship between the<br />

transport of particle counts, turbidity, and aerobic and anaerobic spores with that of E.coli,<br />

Cryptosporidium, and Giardia.


Acknowledgements<br />

We gratefully acknowledge the support of the USEPA Office of <strong>Research</strong> and Development<br />

(Project CR-826337; Thomas F. Speth, Project Officer), USEPA Science to Achieve Results<br />

(STAR) program (Grant #R82901101-0; Angela Page, Project Officer), and the utility<br />

subsidiaries of American <strong>Water</strong>.<br />

REFERENCES<br />

Table 2. Results from January to December 2002 Monthly Field Monitoring<br />

(12 Total Sampling Rounds) a,b<br />

Bacillus Clostridium Total E. coli Bacterio- Bacterio- Crypto- Giardia<br />

(cfu/L) (cfu/L) Coli<strong>for</strong>ms (MPN/L) phage phage sporidium (cysts/L)<br />

(MPN/L) MS-2 PhiX174 (oocysts/L)<br />

(pfu/L) (pfu/L)<br />

Indiana-American <strong>Water</strong> Company at Jeffersonville, Indiana<br />

Ohio River 8.9 × 10 4 3.8 × 10 2 9.2 × 10 4 2.6 × 10 3 6 × 10 1 1.7 × 10 3 2.5 × 10 –2 3 × 10 –2<br />

Well 9 2.5 × 103 0.5]<br />

Well 2 1.3 × 104


128<br />

Ray, C., G. Melin, and R.B. Linsky, editors (2002b). Riverbank Filtration: Improving Source-<strong>Water</strong> Quality,<br />

Kluwer Academic Publishers, Dordrecht.<br />

Standard Methods <strong>for</strong> the Examination of <strong>Water</strong> and Wastewater, Twentieth Edition (1998). APHA, AWWA,<br />

& WEF, Washington.<br />

Summers, R.S., S.M. Hooper, H.M. Shukairy, G. Solarik, and D.M. Owen (1996). “Assessing DBP Yield:<br />

Uni<strong>for</strong>m Formation Conditions.” Journal AWWA, 88(6): 80.<br />

Thurman, E.M., and R.L. Malcolm (1981). “Preparative Isolation of Aquatic Humic Substances.” Envir. Sci.<br />

& Technol., 15(4): 463.<br />

Tufenkji, N., J.N. Ryan, and M. Elimelech (2002). “The Promise of Bank Filtration.” Envir. Sci. & Technol.,<br />

36(21): 423A.<br />

Weiss, W.J., E.J. Bouwer, W.P. Ball, C.R. O’Melia, M.W. LeChevallier, H. Arora, and T.F. Speth (2003a).<br />

“Riverbank Filtration: Fate of Disinfection By-product Precursors and Selected Microorganisms.” Journal<br />

AWWA, in publication (expected October, 2003).<br />

Weiss, W.J., E.J. Bouwer, W.P. Ball, C.R. O’Melia, H. Arora, and T.F. Speth (2003b). “Riverbank Filtration:<br />

Comparison with Bench-Scale Conventional Treatment <strong>for</strong> NOM and DBP Precursor Reductions.” Journal<br />

AWWA, in publication (expected December, 2003).<br />

JOSH WEISS is a <strong>Research</strong> Assistant in the Department of Geography and Environmental<br />

Engineering at The Johns Hopkins University, where he is currently investigating waterquality<br />

improvements during riverbank filtration at three Midwestern drinking-water<br />

utilities. He expects to receive a Ph.D. in 2004 after the completion of his thesis,<br />

“Reduction in Disinfection Byproduct Precursors and Microorganisms During Riverbank<br />

Filtration.” Weiss is the co-author of seven publications — including a chapter in Riverbank<br />

Filtration: Improving Source-<strong>Water</strong> Quality (2002) — and several conference proceedings,<br />

and was the 2000 recipient of the Chesapeake Section American <strong>Water</strong> Works Association Student Paper<br />

Award. Prior to attending Johns Hopkins, he researched the geochemical impact of proposed developments<br />

on a Florida barrier island <strong>for</strong> the Georgia <strong>Institute</strong> of Technology and investigated the bioremediation of sites<br />

contaminated by nitroaromatic compounds <strong>for</strong> Argonne <strong>National</strong> Laboratory. Weiss received a B.S. in Civil<br />

Engineering from the Georgia <strong>Institute</strong> of Technology and a M.S. in Environmental Engineering from The<br />

Johns Hopkins University.


Session 6: Microorganisms<br />

Assessment of the Microbial Removal Capabilities<br />

of Riverbank Filtration<br />

Vasiliki Partinoudi<br />

New England <strong>Water</strong> Treatment Technology Assistance Center<br />

Department of Civil Engineering, University of New Hampshire<br />

Durham, New Hampshire<br />

M. Robin Collins, Ph.D., P.E.<br />

New England <strong>Water</strong> Treatment Technology Assistance Center<br />

Department of Civil Engineering, University of New Hampshire<br />

Durham, New Hampshire<br />

Aaron B. Margolin, Ph.D.<br />

New England <strong>Water</strong> Treatment Technology Assistance Center<br />

Department of Civil Engineering, University of New Hampshire<br />

Durham, New Hampshire<br />

Larry K. Brannaka, Ph.D., P.E.<br />

New England <strong>Water</strong> Treatment Technology Assistance Center<br />

Department of Civil Engineering, University of New Hampshire<br />

Durham, New Hampshire<br />

Introduction<br />

Riverbank filtrate includes both groundwater and river water that has percolated through the<br />

riverbank or riverbed to an extraction well. One of the primary objectives of this study was to<br />

assess the microbial removal capabilities of <strong>RBF</strong> independent of any groundwater dilution (i.e., a<br />

worse-case scenario). This study monitored total coli<strong>for</strong>ms, E. coli, and aerobic spore-<strong>for</strong>ming<br />

bacteria along with other water-quality parameters over a 12-month period in the following<br />

locations:<br />

• Pembroke, New Hampshire.<br />

• Mil<strong>for</strong>d, New Hampshire.<br />

• Jackson, New Hampshire.<br />

• Louisville, Kentucky.<br />

• Cedar Rapids, Iowa.<br />

This study also monitored the removal of male-specific coliphage achieved by <strong>RBF</strong> in Louisville,<br />

Kentucky, and Cedar Rapids, Iowa. Samples were collected at both of these sites <strong>for</strong> the detection<br />

of enteric viruses.<br />

Correspondence should be addressed to:<br />

Vasiliki Partinoudi<br />

Graduate Student<br />

Environmental Technology Building • Department of Civil Engineering • Environmental <strong>Research</strong> Group<br />

University of New Hampshire • 25 Colovos Road • Durham, New Hampshire 03824 USA<br />

Phone: (603) 862-1197 • Fax: (603) 862-3957 • Email: Vp4@cisunix.unh.edu<br />

129


130<br />

Objectives<br />

The overall study objectives were to:<br />

• Quantify the contributions of river water and groundwater to <strong>RBF</strong>-extraction water.<br />

• Assess <strong>RBF</strong> as a viable treatment and pretreatment option.<br />

• Compare <strong>RBF</strong> to slow sand filtration in regards to particulate, organic precursors, and<br />

microbiological removal capabilities expressed in terms of log-removal credits.<br />

The focus of this paper will be on microbial removals achieved by <strong>RBF</strong>.<br />

Characterization of Sampling Sites<br />

Five <strong>RBF</strong> sites (three in New Hampshire, one in Kentucky, and one in Iowa) were chosen that<br />

possessed different characteristics, in terms of the degree of hydraulic connection between the<br />

river and <strong>RBF</strong> extraction well, pumping rates, well construction criteria, geology, water source and<br />

quality, and distance of the <strong>RBF</strong> extraction well from the river. The sites were chosen to enable<br />

the researchers to assess the abilities of <strong>RBF</strong> technology in different settings. The study sites in<br />

Louisville, Kentucky, and Cedar Rapids, Iowa, were chosen because these particular sites have<br />

been characterized in detail in previous studies.<br />

Pembroke, New Hampshire: The well site lies within a stratified-drift aquifer located within the<br />

Soucook River Valley. The eastern aquifer boundary is delineated by contact with glacial till. The<br />

aquifer consists of layers of fine to course sand of varying thickness grading with fine gravel to<br />

boulders in some locations. In the area of the <strong>RBF</strong> extraction well, aquifer thickness ranges from<br />

9.7 to 18.9 m, with a saturated thickness of 7.6 to 17.4 m. The distance between the <strong>RBF</strong><br />

extraction well and river is approximately 55 m. The <strong>RBF</strong> extraction well has a diameter of<br />

30.5 centimeters and is 17-m deep. The subsurface material at that depth was dense silt and very<br />

fine sand.<br />

Mil<strong>for</strong>d, New Hampshire: The Mil<strong>for</strong>d State Fish Hatchery is located along the Souhegan River<br />

near its confluence with Purgatory Brook. The Souhegan River Valley contains rich deposits of<br />

glacial sand, gravel, and silt. The well is a 61-centimeter diameter gravel-pack well, constructed<br />

to a depth of 19.8 m. The screen is 50.8 centimeters in diameter and the length is 4.6 m. The<br />

aquifer at the well site consists mainly of sand and gravel. The distance between the river and the<br />

well is 23 m. The well is pumped continuously and delivers 1.6 MGD.<br />

Jackson, New Hampshire: The Jackson <strong>Water</strong> Precinct infiltration gallery was constructed in the<br />

early 1980s and is located on the banks of the Ellis River. Five infiltration galleries are connected<br />

to the river to provide adequate flow to the <strong>RBF</strong> extraction well. The five infiltration gallery<br />

intakes are each 6.1-m long, 1.2-m deep, and 1.2-m wide, and are located underneath the riverbed.<br />

They were placed 1.2-m apart, thus covering a total area of 10.8 m. A 12.2-m long polyvinyl<br />

chloride pipe connects each gallery to an intake pipe leading to the 20.3-centimeter diameter,<br />

7.3-m deep <strong>RBF</strong> extraction well equipped with a submersible pump. A new infiltration gallery is<br />

now under construction.<br />

Louisville, Kentucky: The construction of the collector well was completed in March 1999. The<br />

Louisville <strong>Water</strong> Company has extensively characterized this site. The <strong>RBF</strong> extraction well is on<br />

the bank of the Ohio River, about 30.5 m from the river. The depth of the well is 12.2-m below<br />

the river bottom and the pumping rate is 75.7-million liters per day. The well has seven horizontal<br />

laterals, which are each 73.1-m long and constructed of 30.5-centimeter stainless steel, with well


screens along the entire lateral length. Of the seven laterals, L4 is the lateral underneath the river<br />

separated by at least 40 ft of aquifer material; L1 is the furthest (perpendicular) to the river; and<br />

L2 is parallel to the river.<br />

Cedar Rapids, Iowa: The Cedar River is a meandering steam that has cut into the bedrock<br />

surface, exposing steep valley walls in the study area (Schulmeyer, 1995), where a total of<br />

53 vertical and horizontal wells have been installed. Seminole Valley Park Well Number 22 is<br />

located on the banks of the Cedar River, and is 19.5 m from the crest of the riverbank. The <strong>RBF</strong><br />

extraction well was drilled in 1991 to a total depth of 17.4-m below grade. The borehole diameter<br />

was 10.7 centimeters to depth, and a 7.6-centimeter screen and casing were installed.<br />

Microbial Analyses<br />

Total Coli<strong>for</strong>ms and E. coli: The IDEXX Quanti-Tray/2000 method was used <strong>for</strong> this analysis.<br />

Aerobic Spore-Forming Bacteria: Procedures were followed as outlined by Ballester and Margolin (2000).<br />

Virus Indicators: Male-specific and somatic coliphage viruses were monitored intensively <strong>for</strong><br />

2 weeks in December 2002 at Louisville, Kentucky, and Cedar Rapids, Iowa, using a single agar<br />

overlay (Method 1602, 1999) and a two-step enrichment method. Antibiotic-resistant strains<br />

E. coli F-amp (resistant to Streptomycin and Ampicillin) and E .coli CN-13 (resistant to Nalidixic<br />

Acid) were used as hosts <strong>for</strong> F-specific coliphage and somatic coliphage, respectively. Analyses<br />

followed the method as outlined by Ballester and Margolin (2000).<br />

Viruses: The intensive coliphage monitoring was followed by the collection of samples to detect<br />

human enteric viruses (Adenovirus type 40 and 41, Astrovirus, Poliovirus, Coxsackie virus,<br />

Rotavirus, and Echovirus) using positive microwound filters. The virus samples were analyzed<br />

using the Integrated Cell Culture-Reverse Transcription-Nested Polymerase Chain Reaction<br />

(ICC-RT-nPCR) method, due to its high specificity and sensitivity (Chapron et al., 2000).<br />

Problems with Assessing <strong>RBF</strong> Removal Capabilities<br />

It is essential to understand the extent and magnitude of river-aquifer interaction to address waterquality<br />

and supply issues, as well as to ensure the health of ecosystems (Winter et al., 1998). There<br />

are many questions to be answered as to how <strong>RBF</strong> works, how to establish travel time, and how to<br />

assess the mixing ratio of river water to groundwater in an <strong>RBF</strong> extraction well. These questions<br />

need to be answered to determine the removal efficiency of the <strong>RBF</strong> process independently of<br />

groundwater dilution.<br />

What Is the Exact Travel Time from the River to the Well?<br />

Table 1 provides a summary of travel time determinations <strong>for</strong> river water to reach the <strong>RBF</strong><br />

extraction well and the selected method used <strong>for</strong> each site in this study.<br />

How Much Removal Is Due to Filtration and How Much Is Due to Dilution with Groundwater?<br />

A variety of water-quality parameters (Table 2) were analyzed to assess the contribution of river<br />

and groundwater to <strong>RBF</strong>-extracted water. Knowledge of the mixing ratio would be necessary to<br />

determine removals achieved solely by subsurface filtration. The degree of mixing depends on<br />

factors such as hydraulic head gradient in the aquifer, aquifer properties, and hydrostatic pressure<br />

in the river, as well as the degree of connectivity between the river and parts of the aquifer.<br />

131


132<br />

The contribution of river water and groundwater to each of the <strong>RBF</strong> extraction wells was<br />

evaluated based on a different parameter at each site due to different types of in<strong>for</strong>mation available<br />

at each site. The average source contributions to the <strong>RBF</strong>-extracted water during this study period<br />

is summarized in Table 3.<br />

Results and Discussion<br />

Table 1. Travel Times from the River to the <strong>RBF</strong> Extraction Well<br />

Sampling Site Travel Time Evaluation of Travel Time<br />

(from river<br />

to <strong>RBF</strong> well)<br />

Pembroke, 5 days Darcy’s Law in terms of seepage velocity<br />

New Hampshire<br />

Mil<strong>for</strong>d, 1 day Darcy’s Law in terms of seepage velocity<br />

New Hampshire<br />

Jackson, 1 day Infiltration gallery. Estimated from in<strong>for</strong>mation based<br />

New Hampshire on the construction details of the gallery<br />

Louisville, 1 day Estimated from in<strong>for</strong>mation provided by the<br />

Kentucky Louisville <strong>Water</strong> Company (initial pumping test data)<br />

Cedar Rapids, 5 days Estimated from in<strong>for</strong>mation provided by the<br />

Iowa City of Cedar Rapids (Schulmeyer, 1995)<br />

Table 2. Sampling Parameters Used to Assess the Mixing Ratio in the <strong>RBF</strong> Extraction Well<br />

Sulfate Color Hardness<br />

(true)<br />

Chloride UV 254 absorbance Alkalinity<br />

pH Particle counts Redox potential<br />

Temperature Selected radioisotopes (222Ra) Conductivity<br />

<strong>RBF</strong> reduces DOC by as much as 82 percent and reduces temperature spikes by more than<br />

46 percent in the summer and 96 to 97 percent in the winter months. Turbidity was reduced to<br />

well below 1 ntu (80- to 86-percent removal). The reductions were computed from field values<br />

presented in Table 4.<br />

Typical river-water concentrations ranged between:<br />

• Below detection limit to 24,192 colony-<strong>for</strong>ming units (cfu)/100 milliliters <strong>for</strong> total coli<strong>for</strong>ms.<br />

• Below detection limit to 1,031 cfu/100 milliliters <strong>for</strong> E. coli.<br />

• Eighty-four to 145,000 cfu/100 milliliters <strong>for</strong> aerobic spore-<strong>for</strong>ming bacteria (see Table 4).<br />

All three of these microbial concentrations were below detection limit (less than 1 cfu/100 milliliters)<br />

in <strong>RBF</strong>-extraction well water, even after eliminating the “dilution” effects with groundwater.


Table 3. Average Distribution of River <strong>Water</strong> and Groundwater to <strong>RBF</strong>-Extracted <strong>Water</strong><br />

Percent of Percent of Parameter<br />

River <strong>Water</strong> Groundwater Upon Which<br />

in <strong>RBF</strong> Wells in <strong>RBF</strong> Wells the Ratio Is Based<br />

Pembroke, 40.7 ± 3.7 59.3 ± 3.7 Conductivity<br />

New Hampshire<br />

Mil<strong>for</strong>d, 40.8 ± 6.4 59.2 ± 6.4 Sulfate<br />

New Hampshire<br />

Jackson, 100 0 Infiltration gallery receiving only<br />

New Hampshire river water (due to an impermeable<br />

barrier that prevents groundwater<br />

from entering the well)<br />

Louisville, 78.1 ± 4.4 21.9 ± 4.4 Hardness<br />

Kentucky<br />

Cedar Rapids, 70 30 Based on groundwater flow modeling<br />

Iowa<br />

Parameter/<br />

Site<br />

Pembroke,<br />

New Hampshire<br />

Table 4. Typical Ranges of Selected Analytes of Interest<br />

Mil<strong>for</strong>d,<br />

New Hampshire<br />

Jackson,<br />

New Hampshire<br />

Louisville,<br />

Kentucky<br />

Cedar Rapids, Iowa<br />

Date 8/01 to 11/03 11/01 to 11/02 5/02 to 11/02 9/01 to 12/02 9/02 to 1/03<br />

River <strong>RBF</strong> GW River <strong>RBF</strong> GW River <strong>RBF</strong> River <strong>RBF</strong> GW River <strong>RBF</strong> GW<br />

pH 6.3–7.3 6.1–6.6 5.6–6.2 6.3–7.6 5.8–6.6 5.9–6.7 6.3–7.3 6.0–7.5 NA NA NA 8.2–8.5 7.6–7.7 7.3–7.4<br />

DOC (mg/L) 1.7–6.6 0.3–1.2 0.3–0.7 1.6–6.4 0.5–1.3 0.3-0.8 2.4–2.7 1.7–1.8 3.0–5.0 1.7–2.0 1.3–1.8 2.8–4.3 1.5–2.5 0.1–0.3<br />

Temperature (°C) 0.3–25.7 8.5–13.8 8.3–11.8 0.4–26.1 7.3–16.2 6.7-15.3 10–21.1 9.5–11 3.6–29.7 13.3–26.415.5–16.6 0.4–20.6 8–22.9 9.4–13.1<br />

Conductivity 42–150 193–269 294–395 119–214 107–152 113-194 23–83 52–69 NA NA NA 454–666 473–624 468–658<br />

(µS/cm)<br />

Total 141– BDL BDL 423– BDL BDL 52–1,356 16–160 10– BDL BDL 75– BDL BDL<br />

Coli<strong>for</strong>ms 1,399 2,431 24,192 2,572<br />

(cfu/100 mL)<br />

E. coli 4–108 BDL BDL 6–119 BDL BDL 2.0–20 2.0–10 2–1,031 BDL BDL 23–56 BDL BDL<br />

(cfu/100 mL)<br />

Aerobic 84– BDL BDL 124-1,093 BDL BDL BDL BDL 3,500– BDL BDL 39–217 BDL BDL<br />

Spore-Forming 1,997 145,000<br />

Bacteria<br />

(cfu/100 mL)<br />

NA = Not available. µS/cm = MicroSemens per centimeter. GW = Groundwater. BDL = Below detection limit.<br />

Typical concentrations of aerobic spore-<strong>for</strong>ming bacteria, total coli<strong>for</strong>ms, E. coli, and male-specific<br />

coliphage in river water and <strong>RBF</strong>-extracted water can be seen in Table 5. The total removal is<br />

calculated by subtracting the concentration of an analyte in the river from that in the <strong>RBF</strong><br />

extraction well. Total coli<strong>for</strong>ms and E. coli were reduced on average by at least 2.6- and 0.8-log removal<br />

credits, respectively. These values were considered artificially low due to low numbers observed in<br />

river water. <strong>RBF</strong> achieved a 4.4-log credit reduction of total coli<strong>for</strong>ms on May 20, 2002, in<br />

133


134<br />

Louisville, Kentucky. Aerobic spore-<strong>for</strong>ming bacteria were removed by at least 1.9-log credit in<br />

Pembroke, New Hampshire, and up to 5.2-log credits in Louisville, Kentucky. These values are in<br />

line with those reported <strong>for</strong> Louisville, Kentucky, by Wang et al. (2002).<br />

The male-specific coliphages ranged between 3,453 and 4,622 plaque-<strong>for</strong>ming units (pfu)/100 mL<br />

in river water. The concentration of the male-specific coliphages was reduced at least 80 percent<br />

by riverbank passage at all the study sites, as is indicated in the reduction summary presented in<br />

Table 5. Although there was a high concentration of male-specific coliphages in the river water<br />

and <strong>RBF</strong>-extracted water, the virus samples collected in Cedar Rapids in December 2003 and<br />

Louisville in March 2003 were negative (ICC-nPCR method) <strong>for</strong> the viruses of interest.<br />

Conclusions<br />

At each study site, the water quality of the source water, aquifer material, distance from the river<br />

to the <strong>RBF</strong> extraction well, travel time, and mixing ratio between river and groundwater were<br />

evaluated and found to differ from site to site. Travel times and mixing ratios were evaluated by<br />

different methods based on the in<strong>for</strong>mation available at each site. <strong>RBF</strong> was found to be a sitespecific<br />

process and should be treated as such. Based on differences between sites in source-water<br />

quality, aquifer material, and other parameters, it will be difficult to develop guidelines applicable<br />

to all <strong>RBF</strong> sites.<br />

In summary, the sites evaluated in this study indicated the conservative effectiveness of <strong>RBF</strong> in<br />

removing bacteria and virus indicators (any groundwater dilution with <strong>RBF</strong> extract should<br />

contribute to even lower microbial concentrations). Aerobic spore-<strong>for</strong>ming bacteria, total<br />

coli<strong>for</strong>ms, E. coli, and male-specific bacteriophage were removed by at least an average of 1.9, 1.0,<br />

0.3 and 0.2 logs, respectively. Based on this study, <strong>RBF</strong> shows potential to be a viable pretreatment<br />

and treatment process and warrants additional study.<br />

Acknowledgements<br />

• USEPA <strong>for</strong> funding this project through the New England <strong>Water</strong> Treatment Technology<br />

Assistance Center at the University of New Hampshire.<br />

• Jackson, New Hampshire <strong>Water</strong>works.<br />

• Nicola A. Ballester and Justin H. Fontaine of the University of New Hampshire.<br />

• Mil<strong>for</strong>d, New Hampshire Fish Hatchery personnel.<br />

• Cedar Rapids <strong>Water</strong> Department, Iowa.<br />

• Pembroke, New Hampshire <strong>Water</strong>works.<br />

• Louisville <strong>Water</strong> Company, Kentucky.<br />

• Melisa A. Smith of the University of New Hampshire.


Location Number<br />

of<br />

Sampling<br />

Events [1]<br />

Table 5. Average Reductions Achieved by <strong>RBF</strong><br />

River [2] <strong>RBF</strong> [2] Background<br />

Well [2]<br />

Total<br />

Log<br />

Removal<br />

Average<br />

Percent<br />

Removal<br />

Achieved by<br />

Subsurface<br />

Filtration<br />

Average<br />

Percent<br />

Removal<br />

Achieved by<br />

Groundwater<br />

Dilution<br />

Aerobic Spore Forming Bacteria (cfu/100 mL)<br />

Pembroke, 19 502 BDL BDL >1.9 100% 0%<br />

New Hampshire ± 11<br />

Mil<strong>for</strong>d, 13 673 BDL BDL >2.1 100% 0%<br />

New Hampshire ± 55<br />

Jackson,<br />

New Hampshire<br />

3 32 ± 1 BDL NA BDL 100% 0%<br />

Louisville, 11 33, 404 BDL BDL >3.5 100% 0%<br />

Kentucky ± 5,672<br />

Cedar Rapids,<br />

Iowa<br />

5 101 ± 8 BDL BDL >2.6 100% 0%<br />

Total Coli<strong>for</strong>ms (cfu /100 mL)<br />

Pembroke, 19 521 BDL BDL >2.1 100% 0%<br />

New Hampshire ± 63<br />

Mil<strong>for</strong>d, 13 1,091 BDL BDL >2.6 100% 0%<br />

New Hampshire ± 114<br />

Jackson, 3 504 85 NA >0.5 100% 0%<br />

New Hampshire ± 42 ± 12<br />

Louisville, 11 3,921 BDL BDL >1 100% 0%<br />

Kentucky ± 182<br />

Cedar Rapids, 5 1,391 BDL BDL >1.4 100% 0%<br />

Iowa<br />

E. coli (cfu /100 mL)<br />

± 25<br />

Pembroke,<br />

New Hampshire<br />

19 27 ± 4 BDL BDL >0.6 100% 0%<br />

Mil<strong>for</strong>d, New<br />

Hampshire<br />

13 55 ± 10 BDL BDL >0.8 100% 0%<br />

Jackson,<br />

New Hampshire<br />

3 30 ± 1 6.5 ± 1 NA >0.4 100% 0%<br />

Louisville,<br />

Kentucky<br />

11 5 ± 2 BDL BDL >0.3 100% 0%<br />

Cedar Rapids,<br />

Iowa<br />

5 16 ± 1 BDL BDL >0.7 100% 0%<br />

Virus Indicators (Male-Specific Coliphage) (pfu/100 mL)<br />

Louisville, 4 4,342 3,703 3,402 >0.2 80.2% 19.80%<br />

Kentucky ± 24 ± 22 ± 18<br />

Cedar Rapids, 5 3,438 753 BDL >0.7 100% 0%<br />

Iowa ± 21 ± 9<br />

Where: [1] = One sampling event includes the collection of a river water, groundwater, and <strong>RBF</strong>-extracted water sample.<br />

The <strong>RBF</strong>-extracted water sample was sampled with travel time taken into consideration.<br />

Where: [2] = This value shows the average concentration of the microorganism of interest throughout the study ± analytical error.<br />

BDL = Below detection limit. NA = Not available.<br />

135


136<br />

REFERENCES<br />

Ballester, N.A., and A.B. Margolin (2000). Detection of Bacillus Spores in Environmental Samples, University<br />

of New Hampshire <strong>Water</strong>borne Disease Laboratory SOP Manual.<br />

Chapron, C.D., N.A. Ballester, J.H. Fontaine, C.N. Frades, and A.B. Margolin (2000). “Detection of<br />

Astroviruses, Enteroviruses, and Adenovirus types 40 and 41 in Surface <strong>Water</strong>s Collected and Evaluated by<br />

the In<strong>for</strong>mation Collection Rule and an Integrated Cell Culture-Nested PCR Procedure.” Applied and<br />

Environmental Microbiology, 66(6): 2,520-2,525.<br />

Method 1602 (1999). Male-specific (F+) and Somatic Coliphages in <strong>Water</strong> by Single Agar Layer Procedure.<br />

United States Environmental Protection Agency-821-R-00-00X. Draft: July 1999.<br />

Schulmeyer P.M. (1995). Effect of the Cedar River on the quality of the groundwater supply <strong>for</strong> Cedar Rapids,<br />

Iowa, U.S. Geological Survey <strong>Water</strong> Resources Investigations Report 94-4211.<br />

Wang J.Z., S.A. Hubbs, and R. Song (2002). Evaluation of riverbank filtration as a drinking water treatment<br />

process, American <strong>Water</strong> Works Association <strong>Research</strong> Foundation and American <strong>Water</strong> Works Association.<br />

Winter T.C., J.W. Harvey, O.L. Franke, and W.M. Alley (1998). Groundwater and surface water – A single<br />

resource, U.S. Geological Survey Circ. 1139, 79 pp.<br />

VASO PARTINOUDI is a graduate student at the University of New Hampshire who is<br />

working on her Master’s thesis, “Riverbank Filtration as a Viable Treatment and<br />

Pretreatment Process.” She has worked as a <strong>Research</strong> Assistant at the New England <strong>Water</strong><br />

Treatment Technology Assistance Center (WTTAC) at the University of New Hampshire<br />

and has participated in WTTAC projects, per<strong>for</strong>ming various tasks such as installing and<br />

maintaining water monitoring equipment and pilot slow sand filters, and sampling and<br />

per<strong>for</strong>ming various water-quality tests in the laboratory and in the field. She has presented<br />

her thesis-related work at the New England <strong>Water</strong> Works Association conference in 2002 (paper titled,<br />

“Riverbank Filtration as a Viable Treatment and Pretreatment Method”) and in 2003 at the American<br />

Geophysical Union conference in Nice, France (poster presentation titled, “Assessment of the microbial<br />

removal capabilities of Riverbank Filtration”). Partinoudi has received a B.S. in Civil Engineering from the<br />

University of Brighton, United Kingdom, and an M.S. in European Construction Engineering from the<br />

University of Coventry, United Kingdom.


Session 7: Organics Removal<br />

Riverbank Filtration: A Very Efficient Treatment<br />

Process <strong>for</strong> the Removal of Organic Contaminants?<br />

Dr.-Ing. Heinz-Jürgen Brauch<br />

DVGW-Technologiezentrum Wasser<br />

Karlsruhe, Germany<br />

Dr. rer. nat. Frank Sacher<br />

DVGW-Technologiezentrum Wasser<br />

Karlsruhe, Germany<br />

Prof. Dr. Wolfgang Kühn<br />

DVGW-Technologiezentrum Wasser<br />

Karlsruhe, Germany<br />

Introduction<br />

In Germany, <strong>RBF</strong> has been used <strong>for</strong> more than 100 years as a natural treatment process <strong>for</strong> the<br />

production of drinking water (Sontheimer, 1980 and 1991; Kühn and Müller, 2000; Sacher and<br />

Brauch, 2002). Nowadays, the major raw-water resource <strong>for</strong> drinking-water supplies in Germany<br />

is groundwater (about 64 percent), whereas bank-filtrated (or infiltrated) water is about 16 percent<br />

(Sacher and Brauch, 2002; Brauch et al., 2001). Compared to this, the direct abstraction of river<br />

water is of minor importance (less than 1 percent). In many cases (and mostly along larger rivers),<br />

a clear distinction between bank-filtrated water and groundwater is difficult, and the raw water<br />

used <strong>for</strong> drinking-water production is bank-filtrated water blended with groundwater.<br />

Bank-filtrated water as a source of raw water provides high-quality river water, and its subsoil passage<br />

guarantees an efficient and lasting removal of suspended matter, microorganisms, and organic<br />

micropollutants. Hence, the occurrence and fate of organic contaminants in river water and bankfiltrated<br />

water are of great concern <strong>for</strong> water suppliers worldwide using surface water or artificial<br />

groundwater as a drinking-water resource. Due to the huge number of possible contaminants in river<br />

water, a necessary restriction has to be made on organic substances that may be relevant to drinkingwater<br />

production (Sacher and Brauch, 2002 and 1999; Sacher et al., 2001a). These target<br />

compounds are characterized by criteria like microbial biodegradability, adsorbability onto activated<br />

carbon and onto soil material, behavior versus oxidation agents, bioaccumulation, and groundwater<br />

mobility, as well as specific data about production and consumption quantities. Toxicological data, if<br />

available, are also important, but are not the main criterion.<br />

Methodology<br />

Within the last few years, the behavior of selected hydrophilic organic micropollutants during<br />

<strong>RBF</strong> was studied in waterworks on the lower Rhine River, as well as by laboratory-scale<br />

Correspondence should be addressed to:<br />

Dr.-Ing. Heinz-Jürgen Brauch<br />

Head of the Analytical Department<br />

DVGW-Technologiezentrum Wasser (TZW)<br />

Karlsruher Strasse 84 • D-76139 Karlsruhe, Germany<br />

Phone: +49(0)721/9678-150 • Fax: +49(0)721/9678-104 • Email: brauch@tzw.de<br />

137


138<br />

experiments. Most of the hydrophilic compounds under investigation are industrial chemicals<br />

with high production volumes and a broad range of applications; there<strong>for</strong>e, they are likely to enter<br />

river water if they are not totally removed in industrial or municipal wastewater treatment plants.<br />

In this paper, data on their fate during <strong>RBF</strong> will be presented, whereby the results of both lab-scale<br />

experiments and long-time measurements of river water and bank-filtrated water will be given.<br />

For the simulation of microbial degradation during <strong>RBF</strong>, a so-called “test filter” is used. This<br />

closed-loop apparatus was developed and used by Sontheimer and Völker (1987) <strong>for</strong> the<br />

characterization of fractions of surrogate parameters from wastewater effluents. In the last few<br />

years, the method was adjusted and optimized to study the biodegradation of single compounds at<br />

concentration levels relevant <strong>for</strong> the environment (Karrenbrock et al., 1999; Knepper at al.,<br />

1999). Details of the experimental set-up are given elsewhere (Karrenbrock et al., 1999).<br />

Several waterworks along the lower Rhine River were selected to measure the behavior of organic<br />

micropollutants during <strong>RBF</strong>. All use bank-filtrated water from the Rhine River as raw water <strong>for</strong><br />

drinking-water production (Schubert, 2000; Denecke, 1997; Brauch et al., 2000), whereby their<br />

wells are situated in a distance of 30 to 50 m from the Rhine River. In all cases, the raw water is a<br />

mixture between bank-filtrated water and groundwater, whereby the groundwater fraction ranges<br />

between 5 and 40 percent (i.e., the raw water is predominantly bank-filtrated water from the<br />

Rhine River). Regular measurements were per<strong>for</strong>med over a time period of several years to attain<br />

reliable and significant data on the removal of organic compounds during underground passage<br />

(Brauch et al., 2000).<br />

Results<br />

Complexing Agents<br />

Aminopolycarbonic acids like nitrilotriacetic acid (NTA), EDTA, or diethylenetrinitrilopentaacetic<br />

acid (DTPA) are used as chelating agents in detergents and industrial cleaners, as well as in the<br />

photo, textile, and pulp- and paper-making industries. Due to their widespread use, NTA and<br />

EDTA are permanently found in the Rhine River in concentration levels of more than 1 µg/L<br />

(ARW and AWBR annual reports; Sacher et al., 1998). Besides these compounds, other<br />

complexing agents like ß-alaninediacetic acid (ADA) or 1.3-propylenedinitrilotetraacetic acid<br />

(PDTA) are used <strong>for</strong> special applications or as substitutes <strong>for</strong> EDTA. To test the biodegradation of<br />

these synthetic compounds, test-filter experiments were per<strong>for</strong>med in which water from the Rhine<br />

River at Karlsruhe was spiked with NTA, EDTA, DTPA, PDTA, and ADA at a concentration<br />

level of 10-µg/L each. These experiments showed that the concentration of NTA decreased quite<br />

rapidly, indicating a fast microbial degradation of this complexing agent. The concentration of<br />

ADA decreased much slower and, after 35 days, about 30 percent of the initial concentration was<br />

still present. The concentrations of EDTA, PDTA, and DTPA seemed to be more or less constant,<br />

indicating that these complexing agents are persistent under the conditions of the test-filter<br />

experiment.<br />

Looking at the concentrations of NTA, EDTA, and DTPA in the Rhine River and in the raw<br />

water of the waterworks, which consists of at least 90-percent bank-filtrated water from the Rhine<br />

River, it is clear that in correspondence to the results of the test-filter experiments, NTA is nearly<br />

totally removed during <strong>RBF</strong> and is only sporadically found in raw water. On the other hand,<br />

EDTA proved to be recalcitrant and was present in the raw water under investigation. DTPA was<br />

found only once in raw water, but concentrations in the Rhine River are quite near to the limit<br />

of determination (which was 2 µg/L until 1996 and is currently 1 µg/L), and mixing with some


uncontaminated groundwater could lead to an apparent removal of DTPA; there<strong>for</strong>e, based on<br />

environmental data, no definite statement can be given on the behavior of DTPA during <strong>RBF</strong>.<br />

ADA is only sporadically found in the Rhine River, and PDTA could not be detected in the Rhine<br />

River (but is often found in the Neckar River, <strong>for</strong> instance).<br />

Aromatic Sulfonates<br />

Aromatic sulfonates are the corresponding bases to sulfonic acids and, due to their permanent<br />

negative charge, are highly soluble in water. Benzenesulfonates are mainly used as intermediates<br />

during the manufacture of azo dyestuffs, optical brighteners, ion-exchange resins, plasticizers, and<br />

pharmaceuticals. Amino- and hydroxynaphthalene-sulfonates and anthraquinonesulfonates are<br />

important building blocks <strong>for</strong> azo dyestuffs. A major source of sulfonated stilbenes is the production<br />

of fluorescent whitening agents <strong>for</strong> laundry products and paper. Naphthalenesulfonates and their<br />

condensates with <strong>for</strong>maldehyde are large-scale products that have widespread applications,<br />

including paper chemicals, superplasticisers <strong>for</strong> concrete, textile auxiliaries, and synthetic leather<br />

tanning agents (Redin et al., 1999).<br />

Test-filter experiments with 2-naphthalene-sulfonate and 1,5-naphthalenedisulfonate showed<br />

that the behavior of the 2-naphthalene-sulfonates is different. Whereas 2-naphthalene-sulfonate<br />

degraded very fast and, after 2 days, could not be detected in water, 1,5-naphthalenedisulfonate<br />

proved to be persistent and no change in concentration could be observed even after 30 days. The<br />

behavior of 1,5-naphthalenedisulfonate is characteristic <strong>for</strong> many two- or threefold sulfonated<br />

naphthalene compounds. Naphthalene-sulfonates with two sulfo groups in alpha position seem to<br />

be especially persistent (Lange et al., 1995).<br />

As could be expected from the results of the test-filter experiments, 1,5-naphthalenedisulfonate,<br />

which is almost always present in the Rhine River, is not removed during <strong>RBF</strong> and was always<br />

found in the raw water of the waterworks under investigation. The same behavior was found <strong>for</strong><br />

2-amino-1,5- and 2-amino-4,8-naphthalene-disulfonate, <strong>for</strong> 1,3,5- and 1,3,6-naphthalenetrisulfonate,<br />

<strong>for</strong> cis-4,4’-dinitro-2,2’-stilbene-disulfonate, and <strong>for</strong> 8,8’-methylenebis-2-naphthalenesulfonate<br />

(Lange et al., 1995).<br />

Pharmaceutical Compounds<br />

Drugs were produced, prescribed, and used in quantities up to some hundred tons per year (in<br />

Germany). Due to an incomplete elimination in wastewater treatment plants, residues of<br />

pharmaceutical products have recently been found in surface waters and groundwaters (Ternes,<br />

1998; Heberer, 2002; Sacher et al., 2001b). In the Rhine River, compounds like diclofenac (an<br />

antirheumatic and analgesic), carbamazepine (an antiepileptic that is also used as antidepressant),<br />

and clofibric acid and bezafibrate (two lipid-regulating agents) are most often found with<br />

concentrations in the 10- to 100-nanograms per liter range. To study their behavior during <strong>RBF</strong>,<br />

test-filter experiments were per<strong>for</strong>med, yielding that only bezafibrate is biodegradable under the<br />

conditions of the test-filter experiment. The concentration of carbamazepine decreases as a<br />

function of time, but even after 30 days, it is present in the test-filter system. The concentrations<br />

of diclofenac and clofibric acid are more or less constant, indicating that the respective<br />

compounds are not easily biodegradable. Monitoring data on the behavior of diclofenac and<br />

carbamazepine during <strong>RBF</strong> under environmental conditions show that that diclofenac was never<br />

detected in the raw water under investigation, although it was nearly always found in the Rhine<br />

River. Qualitatively, the same results were found <strong>for</strong> bezafibrate and clofibric acid (even if clofibric<br />

acid was only found a few times in the Rhine River). For bezafibrate, this finding is in accordance<br />

with the results of the test-filter experiments. For diclofenac and clofibric acid, there is a<br />

139


140<br />

contradiction to the results of the test-filter experiment, indicating that these two compounds are<br />

not biodegradable during underground passage. A satisfying explanation <strong>for</strong> the more effective<br />

elimination of diclofenac and clofibric acid under environmental conditions cannot be given.<br />

Furthermore, it can be seen from the monitoring data that carbamazepine is always present in the<br />

raw waters of all waterworks taking bank-filtrated water from the Rhine River, confirming the<br />

poor biodegradability of this compound found in the test-filter experiment.<br />

Conclusions<br />

Many organic micropollutants present in the Rhine River (and other river waters) are eliminated<br />

during <strong>RBF</strong>. Test-filter experiments prove that this elimination is mainly due to a microbial<br />

degradation of the compounds. Until now, less is known about metabolites and their behavior in<br />

the environment and during treatment processes in waterworks. Only few compounds are not<br />

totally removed during <strong>RBF</strong> and enter the raw waters used <strong>for</strong> drinking-water production. For the<br />

waterworks under investigation at the Rhine River, the compounds found in raw water could be<br />

totally removed by subsequent treatment st<strong>eps</strong> like ozonation or GAC filtration, in most cases.<br />

Hence, the use of bank-filtrated water instead of river water is advisable if industrial or municipal<br />

wastewaters may affect river-water quality. The use of bank-filtrated water, however, cannot<br />

replace further treatment st<strong>eps</strong>.<br />

REFERENCES<br />

Annual reports of Arbeitsgemeinschaft Rhein-Wasserwerke e.V. (ARW) and Arbeitsgemeinschaft Wasserwerke<br />

Bodensee-Rhein (AWBR).<br />

Brauch, H.-J., F. Sacher, E. Denecke, and T. Tacke (2000). “Wirksamkeit der Uferfiltration für die Entfernung<br />

von polaren organischen Spurenstoffen.” gwf-Wasser/Abwasser, 141: 226-234.<br />

Brauch, H.-J., U. Müller, and W. Kühn (2001). “Experiences with riverbank filtration in Germany.”<br />

Proceedings, International Riverbank Filtration Conference, Rheinthemen 4, 33-39.<br />

Denecke, E. (1997). “Auswertung langzeitlicher Messreihen zur aeroben Abbauleistung der Uferpassage einer<br />

Wassergewinnungsanlage am Niederrhein. Z.” Wasser-Abwasser-Forschung, 25: 311-318.<br />

Heberer, T. (2002). “Occurrence, fate, and removal of pharmaceutical residues in the aquatic environment:<br />

A review of recent research data.” Toxicology Letters, 131: 5-17.<br />

Karrenbrock, F., T.P. Knepper, F. Sacher, and K. Lindner (1999). “Entwicklung eines standardisierten<br />

Testfilters zur Bestimmung der mikrobiellen Abbaubarkeit von Einzelsubstanzen.” Vom Wasser, 92: 361-371.<br />

Knepper, T.P., F. Sacher, F.T. Lange, H.-J. Brauch, F. Karrenbrock, O. Rörden, and K. Lindner (1999).<br />

“Detection of polar organic substances relevant <strong>for</strong> drinking water.” Waste Management, 19: 77-99.<br />

Kühn, W. and U. Müller (2000). “Riverbank filtration — An overview.” Journal AWWA, 92: 60-69.<br />

Lange, F.T., M. Wenz, and H.-J. Brauch (1995). “The behaviour of aromatic sulfonates in drinking water<br />

production from River Rhine water and bank filtrate.” Analytical Methods and Instrumentation, 2: 277-284.<br />

Redin, C., F.T. Lange, H.-J. Brauch, and S.H. Eberle (1999). “Synthesis of sulfonated naphthalene<strong>for</strong>maldehyde<br />

condensates and their trace-analytical determination in wastewater and river water.” Acta<br />

Hydrochim. Hydrobiol., 27: 136-142.<br />

Sacher, F., E. Lochow, and H.-J. Brauch (1998). “Synthetische organische Komplexbildner - Analytik und<br />

Vorkommen in Oberflächenwässern.” Vom Wasser, 90: 31-41.<br />

Sacher, F., and H.-J. Brauch (1999). “Bewertung organischer Einzelstoffe im Hinblick auf ihr Verhalten bei<br />

der Wasseraufbereitung.” Veröffentlichungen aus dem Technologiezentrum Wasser, 7: 111-127.


Sacher, F., H.-J. Brauch, and W. Kühn (2001a). “Fate studies of organic micropollutants in riverbank<br />

filtration.” Proceedings, International Riverbank Filtration Conference, Rheinthemen, 4: 139-148<br />

Sacher, F., F.T. Lange, H.-J. Brauch, and I. Blankenhorn (2001b). “Pharmaceuticals in groundwaters —<br />

Analytical methods and results of a monitoring program in Baden-Württemberg, Germany.” J. Chromatogr.,<br />

A 938: 199-210.<br />

Sacher, F., and H.-J. Brauch (2002). “Experiences on the fate of organic micropollutants during riverbank<br />

filtration.” Understanding contaminant biogeochemistry and pathogen removal, C. Ray (ed.), Kluwer Academic<br />

Publishers, The Netherlands, p. 135-151.<br />

Schubert, J. (2000). “Entfernung von Schwebstoffen und Mikroorganismen sowie Verminderung der<br />

Mutagenität bei der Uferfiltration.” gwf-Wasser/Abwasser, 141: 218-225.<br />

Sontheimer, H. (1980). “Experiences with riverbank filtration along the Rhine River.” Journal AWWA, 72: 3,386-3,392.<br />

Sontheimer, H., and E. Völker (1987). Charakterisierung von Abwassereinleitungen aus der Sicht der Trinkwasserversorgung.<br />

Veröffentlichungen des Bereichs und Lehrstuhls für Wasserchemie am Engler-Bunte-Institut der<br />

Universität Karlsruhe 31.<br />

Sontheimer, H. (1991). Trinkwasser aus dem Rhein? Bericht über ein Verbund<strong>for</strong>schungsvorhaben zur Sicherheit<br />

der Trinkwassergewinnung aus Rheinuferfiltrat, Academia Verlag, Sankt Augustin.<br />

Ternes, T.A. (1998). “Occurrence of drugs in German sewage treatment plants and rivers.” Wat. Res.,<br />

32: 3,245-3,260.<br />

HEINZ-JÜRGEN BRAUCH has over 20 years of experience in resolving water-quality<br />

problems, with special emphasis in drinking-water production. He has conducted many<br />

research projects on the development and optimization of analytical determination<br />

methods <strong>for</strong> organic micropollutants, including emerging contaminants and fate studies on<br />

ground surface and drinking water, as well as the design and implementation of monitoring<br />

strategies <strong>for</strong> controlling hazardous substances. In addition, he has completed several<br />

projects in cooperation with water utilities to improve and optimize the treatment<br />

techniques <strong>for</strong> removing organic substances. Brauch has published numerous articles and papers on modern<br />

analytical techniques <strong>for</strong> detecting organic micropollutants, fate and behavior of organics in drinking-water<br />

treatment, occurrence and distribution of organic substances between water, and sediment and soil, as well<br />

as on water-quality problems in surface water and groundwater. Brauch has been the Head of the Analytical<br />

Department of DVGW-Technologiezentrum Wasser (Technology Center <strong>for</strong> <strong>Water</strong>) in Karlsruhe, Germany<br />

since 1990, and is the Chairman or Member of many national and international working groups in this field.<br />

He received a Ph. D. in Chemical Engineering from the University of Karlsruhe.<br />

141


142


Session 7: Organics Removal<br />

Organics Removal by Riverbank Filtration<br />

at the Greater Cincinnati <strong>Water</strong> Works Site<br />

Jeffrey Vogt<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

William Fromme<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

Bruce Whitteberry, P.G.<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

William D. Gollnitz<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

Objective<br />

With the upcoming promulgation of the Stage 2 Disinfectants and Disinfection By-Product Rule,<br />

water utilities are required to reduce DBPs in their finished water (USEPA, 2003). The<br />

concentration of NOM in source waters is directly related to the concentration of DBPs <strong>for</strong>med<br />

in finished waters (Owen et al., 1993). The challenge <strong>for</strong> water utilities is to reduce precursors and<br />

NOM prior to disinfection. The natural process of <strong>RBF</strong> may be an effective method to remove<br />

NOM. As part of a research study, TOC, UV 254, and THM <strong>for</strong>mation potential were used to<br />

evaluate the organic removal capabilities of <strong>RBF</strong>. This project was a cooperative study between the<br />

USGS, Miami University in Ox<strong>for</strong>d, Ohio, and the Greater Cincinnati <strong>Water</strong> Works. This was a very<br />

large project with many objectives. The objective of this paper is to evaluate the removal of NOM<br />

by <strong>RBF</strong>.<br />

Background<br />

The Greater Cincinnati <strong>Water</strong> Works owns and operates a 40-MGD drinking-water treatment<br />

plant located in southwestern Ohio. The source water <strong>for</strong> the plant comes from 10 production<br />

wells located in an alluvial aquifer system adjacent to the Great Miami River. The aquifer is a<br />

mixture of sand and gravel and has a hydraulic conductivity of 200 to 300 ft (62 to 91 m) per day.<br />

Five monitoring wells were drilled adjacent to the river and close to a production well (CMB1).<br />

Each monitoring well was drilled at varying depths in relation to the production well. Well 1A is<br />

the shallowest vertical well and is the nearest to the river. Well 1D is the deepest well and closest<br />

to CMB1. Wells 1B and 1C were placed at intermediate depths between Wells 1A and 1D.<br />

Correspondence should be addressed to:<br />

Jeffrey Vogt<br />

Chemist<br />

Greater Cincinnati <strong>Water</strong> Works<br />

5651 Kellogg Ave • Cincinnati, Ohio 45228 USA<br />

Phone: (513) 624-5624 • Fax: (513) 624-5670 • Email: Jeff.Vogt@gcww.cincinnati-oh.gov<br />

143


144<br />

Well 1C is located at the top of the production well screen, and Well 1D is located at the bottom<br />

of the production well screen. To identify discrete sections of the aquifer, the monitoring wells<br />

were developed with 2-ft (0.6 m) screened interval. An inclined well (Well 1I) was drilled from<br />

the top of the riverbank at a 20- to 30-degree angle from the horizontal plane. This well allowed<br />

the study team to collect samples approximately 5- to 10-ft (1.5- to 3.1-m) below the streambed.<br />

Samples were collected and analyzed on a weekly basis and, later, at monthly intervals from<br />

September 1999 until May 2001.<br />

Materials and Methods<br />

TOC and UV 254 were used to evaluate NOM removal. TOC analyses were analyzed with a<br />

Tekmar/Dohrman Phoenix 8000 using the UV-Persulfate Method. UV absorbance at 254-mn<br />

wavelength was per<strong>for</strong>med with a Hach DR4000. UV absorbance is primarily related to the humic<br />

fraction of NOM. UV at a wavelength of 254 nm is absorbed by double bonds and/or aromatic<br />

structures mostly produced from the breakdown of plant and animal matter (Owen et al., 1993).<br />

THM <strong>for</strong>mation potential were used to evaluate the removal of DBP precursors. The concept<br />

behind the analysis is to maximize the THM <strong>for</strong>mation reaction. Conditions (chlorine dose, pH,<br />

hold time, and high temperature) are set so that the reaction is “pushed” to <strong>for</strong>m THMs. A dose<br />

of 15-mg/L free chlorine was applied to the samples using a sodium hypochlorite solution. The<br />

sample pH was adjusted to 9.5 with a borate buffer and then incubated at 35-degrees Celsius <strong>for</strong><br />

7 days. After incubation, residual chlorine was measured. Samples were poured off and analyzed<br />

<strong>for</strong> THMs with a Varian 3400 Gas Chromatograph using USEPA Method 502.2.<br />

Results and Discussion<br />

All three of the parameters demonstrated the ability of <strong>RBF</strong> to remove organic material. Table 1<br />

represents water-quality data obtained from the river, inclined well, monitoring wells, and production<br />

well. The removal values represent removals from the Great Miami River through the wells. The<br />

average river results show relatively high concentrations of NOM as compared to drinking-water<br />

standards. Organic material was removed through the streambed (Well 1I), and is continually<br />

reduced as it migrates downward into the aquifer. Peaks in organic concentrations in the Great<br />

Miami River correlated with peak concentrations in Well 1I, yet at a reduced concentration (Figure 1).<br />

TOC is reduced 37 percent from the Great Miami River to Well 1I. An additional 20 percent of<br />

TOC was removed to Well 1C. This produces a total reduction in TOC of 59 percent. UV and<br />

THM <strong>for</strong>mation potential demonstrated even better reductions of, respectively, 63 and 72 percent.<br />

Table 1. Average Values and Percent Removals from the Great Miami River<br />

Over the Entire Study Period<br />

Great Well 1I Well 1A Well 1B Well 1C Production Well 1D<br />

Miami Well<br />

River CMB1<br />

DEPTH (ft) 5 to 10 31 46 60 60 88<br />

Avg. Avg. % Avg. % Avg. % Avg. % Avg. % Avg. %<br />

TOC 5.48 3.36 39 2.50 54 2.32 58 2.27 59 1.17 79 0.63 89<br />

UV 254 0.147 0.092 37 0.063 57 0.057 61 0.055 63 0.023 84 0.007 95<br />

THMFP 924 404 56 277 70 267 71 260 72 131 86 55.3 94<br />

THMFP = THM <strong>for</strong>mation potential.


TOC (mg/L)<br />

9<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

Site 1 Total Organic Carbon<br />

GMR FP1i FP1A FP1B FP1C CMB1 FP1D<br />

0<br />

9/1/99 12/2/99 3/3/00 6/3/00 9/3/00<br />

Date<br />

12/4/00 3/6/01 6/6/01 9/6/01<br />

Figure 1. TOC <strong>for</strong> the Great Miami River and well samples <strong>for</strong> the entire study period.<br />

Figure 2 presents UV 254 data <strong>for</strong> all study locations. The bulk of UV reduction is observed between<br />

the Great Miami River and Well 1I (37 percent). An additional 20-percent reduction is seen<br />

between Wells 1I and 1A. The reduction between Wells 1A and 1B is about 4 percent. Between<br />

Wells 1B and 1C, there is about 2-percent reduction. The downward trend of UV reduction<br />

continues through all of the locations, but the zone surrounding Wells 1A, 1B, and 1C<br />

demonstrates minor change. The similarities within that zone can be observed in the maximum,<br />

average, and minimum values <strong>for</strong> those wells. Additional reductions are seen at Well 1D. Results<br />

<strong>for</strong> Well 1D showed very low organic concentrations in all measured parameters when compared<br />

to others wells. It is likely that this water is a combination of regional groundwater and water with<br />

a much longer flow path to the production well. This downward trend or pattern <strong>for</strong> reduction can<br />

be observed in data plots of all three parameters.<br />

CMB1 has a 30-ft (9-m) intake screen. It draws in water from a much larger capture zone than the<br />

monitoring wells due to a much longer screen and significantly higher pumping rates. The water<br />

entering the production well is a mixture of water passing through Wells 1D and 1C, as well as<br />

other parts of the aquifer. The results <strong>for</strong> CMB1 fall between the results <strong>for</strong> both Wells 1C and 1D<br />

in three parameters.<br />

CMB1 showed the least amount of variation in the data, as compared to Wells 1A, 1B, and 1C in<br />

Figure 2. UV results of the samples collected at CMB1 showed a maximum value of 0.029 cm –1 ,<br />

an average value of 0.023 cm –1 , and a minimum value of 0.016 cm –1 . Figure 2 shows the UV results<br />

<strong>for</strong> CMB1 plotting very close together. The calculated standard deviation <strong>for</strong> CMB1 UV data is<br />

0.003 cm –1 . Comparing that value to the standard deviation of Well 1C (0.009 cm –1 ), it is<br />

observed that CMB1 shows little variation in its results. This trend of consistent results between<br />

the data points can be seen in the THM <strong>for</strong>mation potential and TOC results <strong>for</strong> CMB1. This is<br />

likely due to the large capture zone normalizing the concentration.<br />

145


146<br />

Absorbance (cm –1 )<br />

0.30<br />

0.25<br />

0.20<br />

0.15<br />

0.10<br />

0.05<br />

0.00<br />

UV 254 Absorbance<br />

Maximum Average Minimum<br />

GMR FP1i FP1A FP1B FP1C CMB1 FP1D<br />

Location<br />

Figure 2. Maximum, average, and minimum UV values at sample locations.<br />

Looking again at Figure 1, all of the wells except <strong>for</strong> the inclined well demonstrated very<br />

consistent results. The average results <strong>for</strong> Wells 1A, 1B, and 1C are similar <strong>for</strong> all of the parameters<br />

(see Figure 2). These data demonstrate that <strong>RBF</strong> is very effective in reducing any organic shock<br />

load from the river. High concentration peaks in the river and inclined well are unnoticed in the<br />

remaining wells. Others (Kuehn et al., 2000) have also demonstrated this process when using <strong>RBF</strong>.<br />

THM <strong>for</strong>mation potential can be used as a surrogate <strong>for</strong> THM concentrations in finished water<br />

(Owen et al., 1993). The ultimate goal of the Stage 2 Disinfectants and Disinfection By-Product<br />

Rule is to reduce THMs in finished waters produced by water treatment utilities across the<br />

country. UV and TOC measure bulk concentrations of organic matter, but THM <strong>for</strong>mation<br />

potential is a measurement of actual THM concentrations. The organic matter, which directly<br />

reacts with the chlorine to <strong>for</strong>m THMs, was observed to be the most reduced through the aquifer.<br />

Fifty-six percent of that matter was reduced at Well 1I, with an additional reduction of 21 percent<br />

through the 1A, 1B, and 1C zone. At CMB1, the matter was reduced to a total reduction of<br />

86 percent. These results suggest that <strong>RBF</strong> does an excellent job in reducing organic matter that<br />

<strong>for</strong>ms THMs.<br />

The data from this project also supports the concept that the natural process of <strong>RBF</strong> produces<br />

consistent water quality through time. Through this study period, there was no apparent breakthrough<br />

of organic material. This would indicate that the processes, which remove NOM in the<br />

aquifer, do not reach a saturation point, as would be expected in an engineered system, such as<br />

GAC.


Conclusions<br />

<strong>RBF</strong> is very effective in reducing organic matter in source waters. The reduction of NOM was<br />

demonstrated using TOC and UV 254 absorbance. THM precursor removal was demonstrated<br />

using THM <strong>for</strong>mation potential.<br />

• TOC reductions ranged from 39 to 79 percent between the Great Miami River and<br />

CMB1.<br />

• UV254 absorbance reductions ranged from 37 to 84 percent between the Great Miami<br />

River and CMB1.<br />

• THM <strong>for</strong>mation potential demonstrated the greatest range of reductions between the<br />

Great Miami River and CMB1, with values of 56 to 86 percent.<br />

• <strong>RBF</strong> is very effective in reducing organic shock loading. The high concentrations seen in<br />

the river were not seen in the 1A, 1B, or 1C monitoring wells. During the study period,<br />

there was never a breakthrough of organic matter. The results <strong>for</strong> the three parameters<br />

remained consistent when compared to varying concentrations in the Great Miami River.<br />

• Production well (CMB1) data remained very consistent in all of the measured parameters.<br />

REFERENCES<br />

Owen, D.M., G.L. Amy, and Z.K. Chowdhury (1993). Characterization of Natural Organic Matter and Its<br />

Relationship to Treatability, American <strong>Water</strong> Works <strong>Research</strong> Foundation, Denver, Colorado.<br />

USEPA (2003). Stage 2 Disinfectants/Disinfection By Product Rule, United States Environmental Protection<br />

Agency, Washington, D.C.<br />

Kuehn, W., and U. Mueller (2000). “Riverbank filtration: An Overview.” Journal AWWA, 92(12): 60-69.<br />

JEFFREY VOGT is a Chemist with the Greater Cincinnati <strong>Water</strong> Works, where he has<br />

worked since 1992. For the last 5 years, he has been involved in riverbank-filtration<br />

issues/research and was largely involved in the 3-year Flowpath Study. His responsibilities<br />

during the study involved planning, sampling, analyses, data evaluation, and interpretation<br />

of the results. In 2002, he presented and published at the American <strong>Water</strong><br />

Resources Association’s Groundwater/Surface <strong>Water</strong> Interactions Conference in<br />

Keystone, Colorado. Some of his current responsibilities include managing continuing<br />

riverbank-filtration issues, parasite-monitoring program <strong>for</strong> the Long Term 2 Enhanced Surface <strong>Water</strong><br />

Treatment Rule at the Greater Cincinnati <strong>Water</strong> Works’ Richard Miller Treatment Plant, and treatment<br />

studies <strong>for</strong> the Greater Cincinnati <strong>Water</strong> Works’ three treatment plants. Vogt holds a Class 3 <strong>Water</strong><br />

Treatment license in the State of Ohio and received a Science degree from the University of Cincinnati.<br />

147


148


Lunch Presentation<br />

Potential Uses of Riverbank Filtration<br />

<strong>for</strong> Regulatory Compliance<br />

Stig Regli<br />

United States Environmental Protection Agency<br />

Washington, D.C.<br />

<strong>RBF</strong> may have its most general application <strong>for</strong> systems seeking compliance with the LT2ESWTR.<br />

On August 11, 2003, the USEPA proposed the LT2ESWTR and included provisions by which<br />

<strong>RBF</strong> could be used as one of the compliance options <strong>for</strong> providing Cryptosporidium removal credits<br />

(USEPA, 2003). While the USEPA has previously recognized (through guidance implementation<br />

decisions) that <strong>RBF</strong> is a technology that can achieve pathogen removal, the LT2ESWTR is the<br />

first United States drinking-water regulation that specifically recognizes <strong>RBF</strong> as a compliance<br />

technology option.<br />

Under the proposed LT2ESWTR, filtered systems must monitor source water <strong>for</strong> Cryptosporidium<br />

to determine what source-water bin concentration category it belongs in and whether additional<br />

treatment is required. As part of this determination, the USEPA provides a “toolbox” of technologies<br />

by which systems can assess their total removal/inactivation credits <strong>for</strong> Cryptosporidium.<br />

The proposed LT2ESWTR recognizes <strong>RBF</strong> as a “toolbox” pretreatment technique that can<br />

provide a system 0.5- or 1.0-log additional pretreatment credit, if it meets specified design criteria<br />

and monitoring criteria.<br />

For <strong>RBF</strong> to be eligible <strong>for</strong> credit as a pretreatment technique, the following proposed criteria must<br />

be met:<br />

• Wells must be drilled in an unconsolidated, predominantly sandy aquifer, as determined<br />

by grain-size analysis of recovered core material — the recovered core must contain<br />

greater than 10-percent fine-grained material (grains less than 1.0-millimeter diameter)<br />

in at least 90 percent of its length.<br />

• Wells must be located at least 25 ft (in any direction) from the surface-water source to be<br />

eligible <strong>for</strong> 0.5-log credit; wells located at least 50 ft from surface water are eligible <strong>for</strong><br />

1.0-log credit.<br />

• The wellhead must be continuously monitored <strong>for</strong> turbidity to ensure that no system<br />

failure is occurring. If the monthly average of daily maximum turbidity values exceeds<br />

1 ntu, the system must report this finding to the State. The system must also conduct an<br />

assessment to determine the cause of high turbidity levels in the well and consult with<br />

the State to determine whether the previously allowed credit is still appropriate.<br />

Systems using <strong>RBF</strong> as pretreatment to a filtration plant at the time that the system is required to<br />

monitor <strong>for</strong> Cryptosporidium must sample the well effluent <strong>for</strong> the purpose of determining the bin<br />

Correspondence should be addressed to:<br />

Stig Regli<br />

Environmental Engineer<br />

United States Environmental Protection Agency<br />

OGWDW (4607M) • 1200 Pennsylvania Avenue NW • Washington, D.C. 20460 USA<br />

Phone: (202) 564-5270 • Fax: (202) 564-3767 • Email: regli.stig@epa.gov<br />

149


150<br />

classification. Where bin classification is based on monitoring the well effluent, systems are not<br />

eligible to receive additional credit <strong>for</strong> <strong>RBF</strong>. The rationale <strong>for</strong> the above proposed criteria and<br />

opportunity <strong>for</strong> public comment are described in detail in the Federal Register (68FR47692) and<br />

are available on the web at http://www.regulations.gov/fredpdfs/03-18295.pdf.<br />

<strong>RBF</strong> also provides the opportunity <strong>for</strong> directly reducing organic DBP precursor levels or indirectly<br />

facilitating the application of advanced precursor removal technologies, such as nanofiltration.<br />

The USEPA is proposing the Stage 2 Disinfection By-Product Regulation to mitigate concerns <strong>for</strong><br />

the potential risk of developmental and reproductive effects from DBPs. Since the compliance<br />

dates of the LT2ESWTR will coincide with those of the Stage 2 Disinfection By-Product<br />

Regulation, there may be opportunities <strong>for</strong> utilities to use <strong>RBF</strong> <strong>for</strong> helping to achieve compliance<br />

with both regulations.<br />

The simultaneous reduction of other regulated contaminants by irreversible adsorption, biodegradation,<br />

dilution with groundwater, or attenuation mechanisms is also possible with <strong>RBF</strong>. This is<br />

clearly a site-specific issue whose success cannot be assumed without adequate testing/monitoring.<br />

REFERENCE<br />

USEPA (2003). <strong>National</strong> Primary Drinking <strong>Water</strong> Regulations: Long Term 2 Enhanced Surface <strong>Water</strong> Treatment<br />

Rule; Propose Rule, Federal Register, 68(154): 47,691-47,696.<br />

STIG REGLI is an Environmental Engineer <strong>for</strong> the Office of Ground <strong>Water</strong> and Drinking<br />

<strong>Water</strong> of the United States Environmental Protection Agency. He has been with the<br />

United States Environmental Protection Agency since 1979 and is involved with<br />

developing national drinking-water regulations <strong>for</strong> public water systems. His major focus<br />

has been as a Regulation Manager (1985 to 1996) and, more recently, as a Senior Science<br />

Advisor pertinent to the control of pathogens and disinfection byproducts. He has also<br />

been on extended leave to work on drinking-water related projects in Somalia and<br />

Thailand. Prior to working at the United States Environmental Protection Agency, he taught environmental<br />

engineering courses as a Peace Corps volunteer at Kabul University in Kabul, Afghanistan. Regli received<br />

both a B.S. in Mechanical Engineering and an M.S. in Civil Engineering from Duke University.


Session 8: Emerging Contaminants Removal<br />

Transport and Attenuation of<br />

Pharmaceutical Residues During Bank Filtration<br />

Andy Mechlinski<br />

<strong>Institute</strong> of Food Chemistry<br />

Technical University of Berlin<br />

Berlin, Germany<br />

Thomas Heberer, Ph.D.<br />

<strong>Institute</strong> of Food Chemistry<br />

Technical University of Berlin<br />

Berlin, Germany<br />

Bank filtration and artificial groundwater recharge are important, effective, and cheap techniques<br />

to treat surface water and remove microbes and inorganic and (some) organic contaminants;<br />

however, the purification capacity of these techniques varies and is limited in the removal of some,<br />

but not all, potential impurities. Thus, bank-filtration research began investigating pharmaceutically<br />

active compounds (PhACs) when a number of groundwater samples from bank filtration sites<br />

positively detected these compounds. To date, the mechanisms <strong>for</strong> removing impurities and<br />

chemical reactions of the water components have not sufficiently been understood. These subjects<br />

are currently addressed in a new research project called NASRI. In this interdisciplinary project,<br />

the fate and transport of some new emerging contaminants during bank filtration are investigated<br />

at Tegel and Wannsee Lakes and at a groundwater replenishment infiltration pond in Berlin,<br />

Germany. The locations of the field-sites are shown in Figure 1.<br />

The field sites are equipped with different types of monitoring wells that screen at various depths<br />

and are drilled between the infiltration bank and water-supply wells, as well as behind the watersupply<br />

wells. An example of the transects is shown in Figure 2.<br />

These transects are sampled monthly, which allows the fate and transport of PhACs during<br />

groundwater recharge to be monitored. The samples are analyzed by solid phase extraction,<br />

chemical derivatization, and gas chromatography-mass spectrometry (GC/MS) applying selected<br />

ion monitoring. Two novel analytical methods detect PhACs even in complex environmental<br />

samples (Reddersen, 2003). Additionally, some other PhACs (such as antibiotics and estrogenic<br />

steroids) are analyzed by high-pressure liquid chromatography mass spectrometry (HPLC-MS/MS).<br />

In Berlin surface waters, PhACs are found up to the microgram-per-liter level as highly persistent<br />

residues. Six PhACs (Figure 3), including the analgesic drugs diclofenac and propyphenazone, the<br />

antiepileptic drugs carbamazepine and primidone, and the drug metabolites clofibric acid and<br />

1-acetyl-1-methyl-2-dimethyl-oxamoyl-2-phenylhydrazide (AMDOPH), were found to leach from<br />

contaminated watercourses into groundwater aquifers. They also occur at low concentrations in<br />

receiving water-supply wells. Bank filtration was, however, also found to decrease concentrations<br />

(e.g., of diclofenac) or even remove some PhACs (bezafibrate, indomethacine, antibiotics, and<br />

estrogens). Other PhACs (such as carbamazepine and, especially, primidone) were identified as<br />

being excellent tracers of sewage contamination in surface water and groundwater.<br />

Correspondence should be addressed to:<br />

Andy Mechlinski<br />

Graduate Student<br />

<strong>Institute</strong> of Food Chemistry<br />

Technical University of Berlin • Sekr. TIB 4/3-1 • Gustav-Meyer-Allee 25 • 13355 Berlin, Germany<br />

Phone: +49 (30) 314-72267 • Fax: +49 (30) 314-72823 • Email: andymechlinski@web.de<br />

151


152<br />

Acknowledgements<br />

The authors would like to thank Veolia <strong>Water</strong> and the Berlin <strong>Water</strong> Company <strong>for</strong> financing the<br />

NASRI project.<br />

REFERENCE<br />

TS Tegeler See<br />

GWA Tegel<br />

3339<br />

3338<br />

Sickerbecken<br />

TS Lieper Bucht<br />

Unterhavel<br />

TEG365 TEG248<br />

TEG364<br />

TEG247<br />

3337 3335<br />

4<br />

3336 3334<br />

TS Wannsee<br />

Galerie Wannsee<br />

3333<br />

3332<br />

3311<br />

3310<br />

Br.20<br />

Tegeler See<br />

3301<br />

3302<br />

3309<br />

3308<br />

3307<br />

3306 12<br />

3303 13 3305<br />

14<br />

Bernauerst.<br />

3304<br />

Galene West<br />

WW Kladow<br />

Testfeld<br />

Marienfelde<br />

WW Spandau<br />

Spandau<br />

Havel<br />

WW Stolpe<br />

Kleinmachnow<br />

KW Stahnsdorg<br />

Tegeler See<br />

PEA-Tegel<br />

WW Tegel<br />

WW Jungfernheide<br />

Spree<br />

Langsamsandfilter<br />

Uferfiltration<br />

8 Peilrohre<br />

Tegeler Fließ<br />

Charlottenburg<br />

KW Ruhleben<br />

Unterschleuse<br />

WW Tiefwerder<br />

WW und PEA<br />

Beelitzhof<br />

Einleitung<br />

KW Ruhleben<br />

Teitkowkanal<br />

Messstellen<br />

KW Schönerlinde<br />

(KW Marienfelde)<br />

KW Waßmannsdorf<br />

KW Münchehofe<br />

WW Friedrichshagen<br />

Reddersen, K., and T. Heberer (2003). “Multi-methods <strong>for</strong> the trace-level determination of pharmaceutical<br />

residues in sewage, surface and ground water samples applying GC-MS.” J. Sep. Sci. (in press).<br />

Nordgraben<br />

Panke<br />

Speicherteich<br />

Mühlendamm<br />

Oberschleuse<br />

Landwehrkanal<br />

Neukölln<br />

Spree<br />

Panke<br />

Tettowkanal<br />

MHG<br />

KW Falkenberg<br />

WW Wuhlheide<br />

Wuhle<br />

WW Johannisthal<br />

TS Müggelsee E.<br />

Dahme<br />

Erpe<br />

Wasserwerk waterworks<br />

Klärwerk sewage treatment plant<br />

Phospat-Eliminationsanlage<br />

surface water treatment plant<br />

TS Transsekte Testfield<br />

Wehr weir<br />

Schleuse lock<br />

Nottekanal<br />

Neue Mühle<br />

Fredersdorfer<br />

Fließ<br />

Wernsdorf<br />

© Free University of Berlin<br />

Federal Environmental Agency (UBA)<br />

TS Müggelsee C.<br />

Flakenfließ und<br />

Löcknitz<br />

Figure 1. Locations of the Tegel transect, Wannsee transect, and groundwater recharge area in Berlin, Germany.<br />

SW<br />

? ?<br />

existing observation wells<br />

Pegel Tegler See<br />

3310 3309<br />

3311<br />

?<br />

3301 3308<br />

TEG371OP<br />

TEG371UP<br />

3302 3307<br />

approximately 120 m<br />

TEG372<br />

3303<br />

3306<br />

new observation wells (08/2002)<br />

glacial till<br />

Br 13<br />

NE<br />

3304 3305<br />

Figure 2. Profile of the Tegel transect with 14 observation wells and <strong>Water</strong> Works Well Br. 13.<br />

Woltersdorf<br />

Große Tränke<br />

depth below<br />

ground [m]<br />

0 m<br />

5 m<br />

10 m<br />

15 m<br />

20 m<br />

25 m<br />

30 m<br />

35 m<br />

40 m<br />

45 m<br />

veritas<br />

iustitia<br />

libertas<br />

Freie Universität Berlin


H 3C O O N<br />

N<br />

H3C N CH3 Cl<br />

AMDOPH Bezafibrate<br />

O<br />

H 5C 6<br />

H 5C 2<br />

Cl<br />

CH 3<br />

H<br />

N<br />

O<br />

CH 3<br />

Clofibric Acid<br />

NH<br />

CH 3<br />

C C COOH<br />

CH 3<br />

N<br />

O<br />

C<br />

N<br />

HOOC<br />

Primidone Propyphenazone Indometacine<br />

Figure 3. Structures of some detected PhACs.<br />

NH-CH 2CH 2<br />

O<br />

ANDY MECHLINSKI is a Ph.D. student at the <strong>Institute</strong> of Food Chemistry of the<br />

Technical University of Berlin in Germany. His research interests include the analysis of<br />

pharmaceutically active compounds by gas chromatography-mass spectrometry and the<br />

investigation of the environmental fate and transport of pharmaceuticals during<br />

groundwater recharge at bank-filtration sites in Berlin. In 2001, he took the first part of<br />

the final examination covering food chemistry at the Technical University of Berlin. After<br />

this, he prepared a diploma thesis concerning the transport and attenuation of pharmaceuticals<br />

during bank filtration at two field sites in Berlin. From 2002 until 2005, he will be involved in the<br />

interdisciplinary Natural and Artificial Systems <strong>for</strong> Recharge and Infiltration project. Mechlinski recieved<br />

his undergraduate degree at the <strong>Institute</strong> <strong>for</strong> Food Chemistry of the Technical University of Berlin, where he<br />

investigated the behavior of several pharmaceuticals and other contaminants during riverbank filtration at<br />

two field sites in Berlin.<br />

H 3C<br />

CH 3<br />

C<br />

CH 3<br />

O<br />

CH 3<br />

CH 2<br />

H<br />

N<br />

Cl<br />

Diclofenac<br />

N<br />

O<br />

Carbamazepine<br />

Cl<br />

C<br />

2H 2<br />

CH 3<br />

N<br />

C CH 2<br />

O<br />

COOH<br />

Cl<br />

153


154


Session 8: Emerging Contaminants Removal<br />

Attenuation of Pharmaceuticals<br />

During Riverbank Filtration<br />

Traugott Scheytt, Ph.D.<br />

<strong>Institute</strong> of Applied Geosciences<br />

Technical University Berlin<br />

Berlin, Germany<br />

Petra Mersmann<br />

<strong>Institute</strong> of Applied Geosciences<br />

Technical University Berlin<br />

Berlin, Germany<br />

Marcus Leidig<br />

<strong>Institute</strong> of Applied Geosciences<br />

Technical University Berlin<br />

Berlin, Germany<br />

Thomas Heberer, Ph.D.<br />

<strong>Institute</strong> of Food Chemistry<br />

Technical University Berlin<br />

Berlin, Germany<br />

Objective<br />

Occurrences of PhACs from human medical care are reported in groundwater not only from the<br />

Berlin (Germany) area (Heberer et al., 1998; Scheytt et al., 1998), but also from several places<br />

worldwide (Heberer, 2002). These findings initiated more specific investigations concerning the fate<br />

and transport of those pharmaceuticals under water-saturated and unsaturated conditions (Scheytt<br />

et al., in preparation). Field investigations at the bank infiltration site at Lake Tegel (Berlin,<br />

Germany) show the presence of several classes of pharmaceuticals, such as antirheumatics<br />

(e.g., diclofenac), analgesics (e.g., propyphenazone), and blood lipid regulators (clofibric acid) in<br />

both surface water and groundwater (Verstraeten et al., 2003). At Lake Tegel, clofibric acid was found<br />

at concentrations up to 290 nanograms per liter and propyphenazone up to 250 nanograms per liter,<br />

whereas concentrations of diclofenac were around detection limit.<br />

These results will be compared to preliminary results from an ongoing study at the Santa Ana<br />

River in Orange County, Cali<strong>for</strong>nia. Using its <strong>for</strong>ebay facilities, the Orange County <strong>Water</strong> District<br />

Correspondence should be addressed to:<br />

Traugott Scheytt, Ph.D. (after December 2003)<br />

Department of Civil and Environmental Engineering<br />

609 Davis Hall • University of Cali<strong>for</strong>nia • Berkeley, Cali<strong>for</strong>nia 94720 USA<br />

Phone: (510) 642-0151 • Fax: (510) 642-7483 • Email: scheytt@ce.berkeley.edu<br />

Traugott Scheytt, Ph.D. (until December 2003)<br />

Associate Professor <strong>for</strong> Hydrogeology<br />

<strong>Institute</strong> of Applied Geosciences<br />

Technical University Berlin • Ackerstr. 71-76 • 13355 Berlin, Germany<br />

Phone: +49 30 314-72417 • Fax: +49 30 314-25674 • Email: traugott.scheytt@tu-berlin.de<br />

155


156<br />

recharges surface water from the Santa Ana River into the Orange County Groundwater Basin.<br />

The fact that a high amount of the Santa Ana River’s flow is of wastewater origin has prompted<br />

concern that PhACs that survive secondary and tertiary treatment and groundwater transport may<br />

end up in the aquifer.<br />

To understand the transport and attenuation processes of those substances, laboratory column<br />

experiments were conducted (Mersmann et al., 2002). Based on these results, it is possible to<br />

identify and predict the transport and fate of PhACs during bank infiltration and to identify the<br />

pharmaceuticals, which may enter the aquifer via bank infiltration.<br />

Materials and Methods<br />

Pleistocene sediments from a well drilling campaign carried out by the Berlin <strong>Water</strong> Works were<br />

used to prepare the soil column <strong>for</strong> the column leaching experiment. The sediment was sampled<br />

at a depth of approximately 60-m below ground level and consisted of medium-grained sand. The<br />

sediment was manually packed into a stainless steel column measuring 35 × 14 centimeters (inner<br />

diameter). A gauze net and 0.5-diameter globes were placed at both the top and bottom of the<br />

column to prevent soil particles from leaching. The column was pre-wetted and, afterwards,<br />

equilibrated with groundwater originating from the same location and depth like the sediment<br />

itself. Groundwater was led through the column with a flow rate of about 0.3 m per day and a<br />

bottom-to-top flow direction to ensure saturated conditions in the column.<br />

Equilibration of the column with pure groundwater took about 5 days (4.83 pore volumes) be<strong>for</strong>e<br />

pharmaceutical compounds were applied. The groundwater used <strong>for</strong> the column experiment<br />

represents a typical groundwater from the Berlin area not contaminated by any PhAC residues.<br />

Lithium chloride (LiCl) (used as tracer) and the pharmaceuticals were applied to the same<br />

groundwater, which was then passed through the column <strong>for</strong> approximately 10 days. Beside the<br />

tracer and pharmaceutical compounds, all other parameters (e.g., flow rate) were kept the same<br />

during all three phases of the study. The experiments took place at a room temperature of<br />

approximately 20-degrees Celsius and all parts of the column experiment, including the tank, were<br />

protected against exposure to light. The eluted liquid was collected in fractions of approximately<br />

25 milliliters and analyzed <strong>for</strong> contents of anions, cations, pharmaceutical chemicals, and the<br />

lithium chloride used as a tracer.<br />

The concentration of the lithium chloride tracer was 10 mg/L, and the pharmaceuticals had a<br />

concentration of 10 µg/L in the spiked water. Physico-chemical parameters (redox potential, pH,<br />

temperature, oxygen saturation, specific conductance) were measured every 10 minutes using<br />

respective electrodes coupled to a data logger. Lithium chloride was chosen as a tracer because the<br />

background concentration of lithium was definitely below detection limit in all sediment and<br />

groundwater samples used <strong>for</strong> the experiments. Lithium also shows a transport behavior comparable<br />

to a nonreactive tracer and can be analyzed in a rapid and cost-effective manner.<br />

For the analysis of pharmaceutical compounds, water samples were adjusted to a pH of 2 and then<br />

extracted by solid-phase extraction using a non-endcapped reversed phase adsorbent<br />

(RP-C18 Bakerbond Polar Plus). Then the analytes and surrogate standard were derivatized,<br />

making them amendable to gas chromatographic separation (Heberer et al., 1998).<br />

Two microliters of the sample extracts (100 microliters <strong>for</strong> each sample) were analyzed by capillary<br />

GC-MS with selected ion monitoring. Depending on the sample volume (100 to 1,000 milliliters),<br />

the limits of determination were between 1 and 10 nanograms per liter, and the limits of quantitation<br />

were between 5 and 25 nanograms per liter. The analytical recoveries range between 80 and 120 percent.<br />

For further analytical details, refer to Heberer et al. (1998).


Results and Conclusion<br />

The Santa Ana River study will provide in<strong>for</strong>mation about the concentration of some high-volume<br />

pharmaceuticals from human medical care in the Santa Ana River, especially acetaminophen,<br />

carbamazepine, diclofenac, gemfibrozil, ibuprofen, metoprolol, naproxen, primidone, and propranolol.<br />

Laboratory experiments show that clofibric acid exhibits no degradation and almost no<br />

retardation (Rf = 1.1). After spiking groundwater with lithium chloride and diclofenac, the<br />

movement of diclofenac within the column is much slower than the movement of the lithium<br />

tracer (Figure 1), leading to a retardation factor of Rf = 2.6 <strong>for</strong> diclofenac. Propyphenazone<br />

(Rf = 2.0) is also retarded, whereas significant degradation was not observed <strong>for</strong> both pharmaceuticals,<br />

diclofenac and propyphenazone, under prevailing conditions in the soil column. Ibuprofen<br />

is degraded under aerobic conditions, whereas only little degradation was observed under<br />

anaerobic conditions. The retardation factor <strong>for</strong> ibuprofen was extrapolated to be 4.0.<br />

Carbamazepine shows no degradation in the soil column experiments, but significant retardation<br />

(Rf = 2.8) under prevailing conditions.<br />

C/C 0<br />

1.2<br />

1.0<br />

0.8<br />

0.6<br />

0.4<br />

0.2<br />

0.0<br />

0.0 5.0 10.0 15.0<br />

Pore Volume<br />

20.0 25.0 30.0<br />

Lithium Diclofenac<br />

Figure 1. Concentration of lithium and diclofenac at the outflow of the soil<br />

column; results from a single-substance laboratory soil column experiment.<br />

It was concluded that at least clofibric acid, propyphenazone, and carbamazepine are recalcitrant<br />

under groundwater conditions and will be transported within the aquifer at the Berlin bank-filtration<br />

site. Additionally, due to anaerobic conditions in the deeper part of the aquifer, diclofenac and<br />

ibuprofen may also occur in groundwater and are transported if these compounds are not completely<br />

degraded in surface water or in the aerobic part of the aquifer.<br />

Compared to the bank-infiltration site at Lake Tegel in Berlin, the situation is quite different at<br />

the Santa Ana River, especially in respect to the climate and hydrogeological setting. Among the<br />

already mentioned pharmaceuticals of interest at the Santa Ana River, acetaminophen is<br />

prescribed in high amounts, but it is expected that this compound will be attenuated during bank<br />

infiltration due to its high biodegradability. Additionally, due to mostly aerobic conditions in the<br />

aquifer at the Santa Ana River, diclofenac and ibuprofen will be degraded under these aerobic<br />

conditions. Among the pharmaceuticals that might be expected in groundwater are carbamazepine,<br />

primidone, gemfibrozil, metoprolol, naproxen, and propranolol.<br />

157


158<br />

Acknowledgments<br />

Deutsche Forschungsgemeinschaft has funded parts of this work, and the <strong>National</strong> <strong>Water</strong> <strong>Research</strong><br />

<strong>Institute</strong> is funding the research on attenuation of pharmaceuticals during recharge at the Santa<br />

Ana River. The authors appreciate the cooperation of the Berliner Wasserbetriebe in gaining<br />

access to waterworks wells and the help of the Orange County <strong>Water</strong> District, which has greatly<br />

contributed to the study.<br />

REFERENCES<br />

Heberer, T. (2002). “Occurrence, fate, and removal of pharmaceutical residues in the aquatic environment:<br />

A review of recent research data.” Toxicology Letters, 131: 5-17.<br />

Heberer, T., K. Schmidt-Bäumler, and H.-J. Stan (1998). “Occurrence and distribution of organic contaminants<br />

in the aquatic system in Berlin. Part I: Drug residues and other polar contaminants in Berlin surface<br />

and groundwater.” Acta Hydrochimica et Hydrobiologica, 26: 272-278.<br />

Mersmann, P., T. Scheytt, and T. Heberer (2002). “Column experiments on the transport behavior of<br />

pharmaceutically active compounds in the saturated zone.” Acta Hydrochimica et Hydrobiologica,<br />

30(5-6): 1-10.<br />

Scheytt, T., T. Mersmann, M. Leidig, A. Pekdeger, and T. Heberer (in preparation). “Transport of<br />

pharmaceutically active compounds (PhACs) clofibric acid, diclofenac, and propyphenazone under water<br />

saturated conditions.” Ground <strong>Water</strong>, Westerville, OH.<br />

Scheytt, T., S. Grams, and H. Fell (1998). “Occurrence and behavior of drugs in groundwater.” Gambling with<br />

groundwater – physical, chemical, and biological aspects of aquifer-stream relations, J.V. Brahana, Y. Eckstein,<br />

L.K. Ongley, R. Schneider, and J.E. Moore, eds., IAH/AIH Proc., St. Paul., MN.<br />

Verstraeten, I.M., T. Heberer, and T. Scheytt (2003). “Occurrence, characteristics, transport, and fate of pesticides,<br />

pharmaceutically active compounds, and industrial products and personal care products at bank-filtration<br />

sites.” Riverbank Filtration: Improving Source-<strong>Water</strong> Quality, C. Ray, G. Melin, and R.B. Linsky, eds.,<br />

Kluwer Academic Publishers, Dordrecht.<br />

TRAUGOTT SCHEYTT has been an Associate Professor <strong>for</strong> Hydrogeology in the<br />

Department of Civil Engineering and Applied Geosciences at the Technical University<br />

Berlin since 1996. His teaching duties include courses on groundwater chemistry,<br />

groundwater flow, and transport of organic contaminants. Current research areas include<br />

the transport and attenuation of pharmaceutically active compounds, natural attenuation<br />

of contaminants emanating from landfills, and prediction of transport and degradation of<br />

organic contaminants in groundwater. Presently, he is on sabbatical at University of<br />

Cali<strong>for</strong>nia, Berkeley, and will stay in Cali<strong>for</strong>nia until January 2004 to conduct research on the transport and<br />

attenuation of pharmaceutically active compounds during riverbank filtration at the Santa Ana River in<br />

Orange County, Cali<strong>for</strong>nia. Scheytt received an M.S. and Ph.D. in Geology from Christian-Albrechts-<br />

University Kiel in Germany.


Session 8: Emerging Contaminants Removal<br />

The Fate of Bulk Organics and Emerging Contaminants<br />

During Soil-Aquifer Treatment<br />

Dr. Jörg E. Drewes<br />

Colorado School of Mines<br />

Golden, Colorado<br />

Dipl.-Ing. Tanja Rauch<br />

Colorado School of Mines<br />

Golden, Colorado<br />

Introduction<br />

Wastewater reuse employing soil-aquifer treatment is becoming an increasingly important strategy<br />

<strong>for</strong> many utilities in the United States and abroad to augment local drinking-water sources where<br />

supplies are limited. One major water-quality issue associated with soil-aquifer treatment leading<br />

to indirect potable or nonpotable reuse of wastewater effluents is the fate and transport of organic<br />

constituents. Effluent-derived bulk organic matter can impair the quality of recovered water by<br />

interfering with post-treatment processes, such as coagulation, adsorption, or membrane<br />

treatment. Bulk organic matter is also a known precursor <strong>for</strong> DBPs. The presence of bioavailable<br />

organic carbon after soil-aquifer treatment could also increase the microbial regrowth potential in<br />

distribution systems. In addition, BOM might be comprised of organic micropollutants, which<br />

survive during soil-aquifer treatment and which are associated with potential adverse human<br />

health effects. The objective of this study was to investigate the fate and transport of bulk and<br />

trace organics present in reclaimed water during soil-aquifer treatment leading to indirect potable<br />

reuse. The American <strong>Water</strong> Works Association <strong>Research</strong> Foundation and USEPA funded this study.<br />

Methodology<br />

The fate of organics during soil-aquifer treatment was investigated using controlled biodegradation<br />

studies in adapted soil-columns and full-scale infiltration facilities in Arizona and Cali<strong>for</strong>nia. The<br />

study design followed a watershed-guided approach considering hydraulically corresponding<br />

samples of drinking-water sources, soil-aquifer treatment-applied wastewater effluents, and<br />

subsequent post-soil-aquifer treatment samples representing a series of different travel times in the<br />

subsurface. Extensive characterization of organic carbon in the different samples was per<strong>for</strong>med<br />

using state-of-the-art analytical techniques (such as size-exclusion chromatography with online<br />

DOC and UV absorbance detection, carbon-13 nuclear magnetic resonance spectroscopy, Fourier<br />

trans<strong>for</strong>m infrared spectroscopy, and elemental analysis) and biomass/bioactivity measurements<br />

(dehydrogenase activity; total viable biomass through phospholipid extraction; substrate induced<br />

respiration) to distinguish between primary removal mechanisms (biodegradation versus adsorption).<br />

The mechanisms contributing to bulk organic matter removal during initial recharge were<br />

Correspondence should be addressed to:<br />

Dr. Jörg E. Drewes<br />

Assistant Professor<br />

Environmental Science & Engineering Division<br />

Colorado School of Mines • Golden, Colorado 80401-1887 USA<br />

Phone: (303) 273-3401 • Fax: (303) 273-3413 • Email: jdrewes@mines.edu<br />

159


160<br />

identified by isolating three bulk fractions from treated wastewater effluents:<br />

• Hydrophilic carbon (HPI).<br />

• Hydrophobic acids (HPO-A).<br />

• Colloidal organic matter.<br />

HPI and HPO-A were isolated from a domestic secondary effluent by XAD-8 fractionation.<br />

Colloidal organic matter was operationally defined as organic matter in the size range of<br />

approximately 6,000 Dalton to 1 micrometer and was isolated using dialysis diffusion after a preconcentration<br />

using large volume rotary evaporation. Trace organics selected <strong>for</strong> this study were<br />

PhACs and personal care products, as well as selected endocrine disrupting compounds<br />

(17β-estradiol, estriol, and testosterone).<br />

Pharmaceutical compounds were selected <strong>for</strong> this study based on their production, per-capita<br />

consumption, and occurrence in domestic effluents and surface waters in the United States and<br />

Middle Europe. Caffeine has been a well-known PhAC of wastewater origin <strong>for</strong> more than<br />

20 years. In the category of analgesics/anti-inflammatory drugs, diclofenac, ibuprofen, ketoprofen,<br />

naproxen, fenoprofen, propyphenazone, meclofenamic acid, and tolfenamic acid were identified.<br />

Carbamazepine and primidone were identified as antiepileptic drugs. Pentoxifylline represented a<br />

common blood viscosity-affecting agent. Gemfibrozil, clofibric acid, and fenofibrate were selected<br />

to represent blood lipid regulators. Organic iodine was used as a surrogate parameter <strong>for</strong> the<br />

detection of X-ray contrast agents (triiodinated benzene derivates such as iopromide and<br />

diatrizoate) and their halogenated metabolites.<br />

Results<br />

Data derived from several soil-aquifer treatment field sites consistently indicate a substantial<br />

removal of DOC during travel through the subsurface. At one site employing tertiary treatment<br />

prior to recharge, average DOC concentrations of 5.6 mg/L in the tertiary effluent decreased from<br />

5.6 mg/L to approximately 1.45 mg/L in groundwater monitoring wells after traveling 6 to 12 months<br />

in the subsurface. At the same time, the specific UV absorbance increased, indicating a preferred<br />

removal of aliphatic compound groups within bulk organic matter. Further DOC removal<br />

occurred during long-term travel in the subsurface of 1 to 2 years from 1.45 mg/L to approximately<br />

1.1 mg/L. At another site employing only secondary treatment without nitrification prior to<br />

recharge, the DOC concentration was efficiently reduced to approximately 2 mg/L after<br />

percolating through the vadose zone. The specific UV absorbance increased only slightly from<br />

1.7 liters per milligram and meter (L/mg m) in the secondary treated effluent to 2.0 L/mg m in<br />

groundwater samples. These findings pointed to a fast and substantial removal of biodegradable<br />

organic carbon during the initial phase of groundwater recharge. The results of spectroscopic<br />

investigations generally demonstrated that naturally derived and effluent-derived organic matter<br />

after SAT overlap extensively in molecular weight distribution, amount and distribution of<br />

hydrophobic and hydrophilic carbon fractions, and chemical characteristics based on elemental<br />

analysis and structural analysis; however, the residual portion of DOC contained both<br />

effluent-derived organic matter and NOM.<br />

Results of this study indicate that during vadose zone treatment, HPI is removed by a combination<br />

of physical and biological mechanisms. HPO-A is least efficiently removed of all fractions during<br />

soil infiltration. Both adsorption and biotrans<strong>for</strong>mation seem to contribute to this removal. Low<br />

biomass activity responses in soil-columns fed with HPO-A indicated that these substances are<br />

refractory in nature. The fate of effluent-derived colloidal organic matter during groundwater


echarge is not well understood. Our studies give reason to believe that colloidal organic matter<br />

can be efficiently removed during initial soil recharge, and that this removal is based on the<br />

filtration mechanism and biodegradation. Colloidal organic matter stimulated high biomass<br />

activities, but was not fully mineralized. Remobilization and the breakthrough of colloidal organic<br />

matter was observed in column experiments fed with colloidal organic matter under higher<br />

infiltration rates. The removal of colloidal organic matter might, there<strong>for</strong>e, depend upon hydraulic<br />

regimes in the aquifer during recharge operation. The majority of biological removal in the HPI<br />

and HPO-A fractions occurred in the first 30-centimeters of soil infiltration.<br />

Findings of this study demonstrated that the dominating removal mechanism <strong>for</strong> steroids during<br />

soil-aquifer treatment is adsorption, although biodegradation is also taking place and is important<br />

<strong>for</strong> a sustainable operation avoiding compound accumulation in the system. The study showed<br />

that steroid removal was not dependent upon the type of organic background matrix present (HPI,<br />

HPO-A, colloidal organic matter) or redox and flox conditions (aerobic versus anoxic; saturated<br />

versus unsaturated). 17β-estradiol, estriol, and testosterone were efficiently removed from initial<br />

concentrations as high as 500 nanograms per liter after only 5.2 hours of contact with subsurface<br />

media in the presence of soil microbial activity.<br />

In addition, the study revealed that the stimulant caffeine, analgesic/anti-inflammatory drugs, and<br />

blood lipid regulators were efficiently removed to concentrations near or below the detection limit<br />

of the analytical method after retention times of less than 6 months during groundwater recharge.<br />

The antiepileptics carbamazepine and primidone were not removed during groundwater recharge<br />

under either anoxic saturated or aerobic unsaturated flow conditions during travel times of up to<br />

8 years. Organic iodine showed a partial removal only under anoxic saturated conditions (as<br />

compared to aerobic conditions) and persisted in recharged groundwater.<br />

Conclusions<br />

Findings of this study demonstrated that soil-aquifer treatment represents an efficient treatment<br />

step to reduce bulk organic carbon and to remove trace pollutants of concern present in reclaimed<br />

water.<br />

JÖRG DREWES is an Assistant Professor of Environmental Science and Engineering at<br />

the Colorado School of Mines. He has been actively involved in research in the area of<br />

water reclamation, water reuse, and groundwater recharge <strong>for</strong> more than 11 years. His<br />

research interests include water and wastewater treatment engineering; potable and nonpotable<br />

water reuse (soil-aquifer treatment and microfiltration/reverse osmosis); state-ofthe-art<br />

characterization of natural and effluent organic matter; contaminant transfer<br />

among environmental media; and the fate of endocrine disrupting compounds and<br />

pharmaceuticals in natural and engineered systems. Drewes has published more than 50 journal papers, book<br />

contributions, and conference proceedings. Among his honors, he was awarded the Willy-Hager Award in<br />

1997, Quentin Mees <strong>Research</strong> Award in 1999, and Dr. Nevis Cook Excellence in Teaching Award in 2003.<br />

Drewes received both an M.S. and Ph.D. in Environmental Engineering from the Technical University of<br />

Berlin, Germany.<br />

161


162


Session 8: Emerging Contaminants Removal<br />

Ethylenediaminetetraacetic Acid Occurrence and<br />

Removal Through Bank Filtration in the Platte River,<br />

Nebraska<br />

Jason R. Vogel, Ph.D.<br />

United States Geological Survey<br />

Lincoln, Nebraska<br />

Larry B. Barber, Ph.D.<br />

United States Geological Survey<br />

Boulder, Colorado<br />

Tyler B. Coplen, Ph.D.<br />

United States Geological Survey<br />

Reston, Virginia<br />

Ingrid M. Verstraeten, Ph.D.<br />

United States Geological Survey<br />

Baltimore, Maryland<br />

Thomas F. Speth, Ph.D., P.E.<br />

United States Environmental Protection Agency<br />

Cincinnati, Ohio<br />

Jerry Obrist, P.E.<br />

City of Lincoln <strong>Water</strong> System<br />

Lincoln, Nebraska<br />

The USGS, USEPA, and City of Lincoln <strong>Water</strong> System (Nebraska) have conducted a study to<br />

determine the occurrence and removal of EDTA, NTA, and nonylphenol monoethoxycarboxylate<br />

to nonylphenol pentaethoxycarboxylate (NP1EC-NP5EC) in the hydrologic system at the City of<br />

Lincoln well field. The objective of the study is to evaluate the occurrence and removal of EDTA,<br />

NTA, and total nonylphenolpolyethoxycarboxylate (NPEC) by bank filtration at the City of<br />

Lincoln well field. This presentation will discuss removal during two sampling periods — May and<br />

August 2002 — based upon surface-water fractions in the collector well calculated using stable<br />

isotope ratios of hydrogen and oxygen.<br />

The rationale <strong>for</strong> selecting compounds evaluated in this study was based on the hierarchical<br />

analytical approach and includes a range of compounds covering a spectrum of uses and effects.<br />

For example, EDTA is a low-toxicity, high-production volume chemical used in multiple<br />

domestic, commercial, and industrial applications to <strong>for</strong>m stable, water-soluble complexes with<br />

trace metals. Because of its uses and chemical characteristics, EDTA occurs at relatively high<br />

concentrations and can persist in the aquatic environment (Barber et al., 1996; Barber et al., 2000;<br />

Leenheer et al., 2001; Barber et al., 2003).<br />

Correspondence should be addressed to:<br />

Jason R. Vogel, Ph.D.<br />

Hydrologist<br />

United States Geological Survey<br />

Room 406, Federal Building • 100 Centennial Mall North • Lincoln, Nebraska 68508 USA<br />

Phone: (402) 437-5129 • Fax: (402) 437-5139 • Email: jrvogel@usgs.gov<br />

163


164<br />

Site Description<br />

There are 40 active production wells at the City of Lincoln well field. Two of these wells are<br />

horizontal collector wells screened in alluvial sand and gravel approximately 26-m below the<br />

Platte River; they provide approximately 50 percent of the municipal water during most times of<br />

the year. The remaining 38 wells are vertical production wells that are developed in alluvial<br />

sediments mainly consisting of sand and gravel. At this location, the quality of the river water has<br />

a large effect on the quality of the bank-filtered water obtained from the collector wells. Similarly,<br />

the water quality of the groundwater in the vertical production wells directly corresponds to the<br />

distances of the wells from the river. Because of the direct link between the collector wells and<br />

river, the collector well is usually turned off during the month of May and first part of June to avoid<br />

the flush of herbicides associated with spring planting in this agricultural area. After collection,<br />

well water is treated by ozonation, filtration, and chlorination be<strong>for</strong>e distribution.<br />

Sampling<br />

Representative surface-water samples were collected quarterly from the Platte River at the wellfield<br />

site using equal width-increment, flow-weighted sampling. Groundwater samples were also<br />

collected quarterly from one of the collector wells and from two of the vertical groundwater wells.<br />

One groundwater well was located relatively close to the river (within 100 m), with another<br />

located away from the river (1,000 m). All samples were filtered through 0.7-m glass fiber filters<br />

and collected in pre-cleaned amber glass bottles. Samples <strong>for</strong> EDTA, NTA, and NPEC analyses<br />

were preserved with 2-percent by volume <strong>for</strong>malin. This presentation will discuss the results of<br />

samples from May and August 2002.<br />

Analysis<br />

EDTA, NTA, and NP1EC-NP5EC were measured using a modification (Barber et al., 2000) of<br />

the method of Schaffner and Giger (1984). Samples (100 milliliters) were evaporated to dryness,<br />

acidified with 5-milliliter 50-percent (volume per volume) <strong>for</strong>mic acid/distilled water, and evaporated<br />

to dryness. Acetyl chloride/propanol (10-percent volume per volume) was added, the sample<br />

heated at 90-degrees Celsius <strong>for</strong> 1 hour to <strong>for</strong>m the propyl-esters, and the propyl-esters were<br />

extracted into chloro<strong>for</strong>m. The chloro<strong>for</strong>m extracts were evaporated to dryness and re-dissolved<br />

in toluene <strong>for</strong> analysis by GC/MS, as described below.<br />

The propyl-ester extracts were analyzed by electron-impact GC/MS in both the full-scan and<br />

select ion monitoring modes. The general GC conditions were:<br />

• Hewlett Packard (HP) 6890 GC.<br />

• Column-HP Ultra II (5-percent phenylmethyl silicone), 25 m × 0.2 millimeters,<br />

33-micrometer film thickness.<br />

• Carrier gas, ultra-high purity helium, with a linear-flow velocity of 27 centimeters per second.<br />

• Injection port temperature, 300-degrees Celsius.<br />

• Initial oven temperature, 50-degrees Celsius.<br />

• Split vent open, 0.75 minutes.<br />

• Ramp rate, 6-degrees Celsius per minute to 300-degrees Celsius.<br />

• Hold time, 15 minutes at 300-degrees Celsius.


The MS conditions were as follows:<br />

• HP 5973 Mass Selective Detector.<br />

• Tune with perflurotributylamine.<br />

• Ionization energy, 70 electron volt (eV).<br />

• Source pressure, 1 × 10 –5 torr.<br />

• Source temperature, 250-degrees Celsius.<br />

• Interface temperature, 280-degrees Celsius.<br />

• Full scan, 40 to 550 atomic mass units at one scan per second.<br />

Concentrations were calculated based on select ion monitoring data using diagnostic ions <strong>for</strong> each<br />

compound. Each compound was identified based on the matching of retention times (±0.02 minutes)<br />

and ion ratios (±20 percent) determined from the analysis of authentic standards. An eight-point<br />

calibration curve (typically ranging from 0.01 to 50 nanograms per microliter) and internal<br />

standard procedures were used <strong>for</strong> calculating concentrations. Surrogate standards were added to<br />

the samples prior to extraction and derivatization to evaluate compound recovery and whole<br />

method per<strong>for</strong>mance.<br />

Results<br />

In general, based upon results from May and August 2002, measured EDTA concentrations <strong>for</strong><br />

surface-water samples were larger than <strong>for</strong> groundwater samples. In addition, EDTA concentrations<br />

decreased and total NPEC concentrations increased in the groundwater as the distance of<br />

the well from the river increased. NTA was only detected in one surface-water sample at very low<br />

levels and not at all in groundwater samples.<br />

Using surface-water fractions in the collector well determined from deuterium and oxygen-18 ratios,<br />

the transport of EDTA was nearly conservative during these two sampling periods. Total NPEC<br />

concentrations were lower than predicted. Further analysis will be <strong>for</strong>thcoming in the presentation.<br />

REFERENCES<br />

Barber, L.B., G.K. Brown, and S.D. Zaugg (2000). “Potential endocrine disrupting organic chemicals in<br />

treated municipal wastewater and river water.” Analysis of Environmental Endocrine Disruptors, L.H. Keith,<br />

T.L. Jones-Lepp, and L.L. Needham, eds., American Chemical Society Symposium Series 747, American<br />

Chemical Society, Washington, DC, p. 97-123.<br />

Barber, L.B., E.T. Furlong, S.H. Keefe, G.K. Brown, and J.D. Cahill (2003). “Natural and Contaminant<br />

Organic Compounds in Boulder Creek, Colorado under High-Flow and Low-Flow Conditions, 2000.”<br />

Comprehensive water quality of the Boulder Creek <strong>Water</strong>shed, Colorado, under high-flow and low-flow conditions,<br />

S.F. Murphy, L.B. Barber, and P.L. Verplanck, eds., U.S. Geological Survey <strong>Water</strong> Resources Investigations<br />

Report 03-4045.<br />

Barber, L.B., J.A. Leenheer, W.E. Pereira, T.I. Noyes, G.A. Brown, C.F. Tabor, and J.H. Writer (1996).<br />

“Organic contamination of the Mississippi River from municipal and industrial wastewater.” U.S. Geological<br />

Survey, Circular 1133, p. 114-135.<br />

Leenheer, J.A., C.E. Rostad, L.B. Barber, R.A. Schroeder, R. Anders, and M.L. Davisson (2001). “Nature and<br />

chlorine reactivity of organic constituents from reclaimed water in groundwater, Los Angeles County,<br />

Cali<strong>for</strong>nia.” Environmental Science and Technology, 35: 3,869-3,876.<br />

Schaffner, C., and W. Giger (1984). “Determination of nitrilotriacetic acid in water by high-resolution gas<br />

chromatography.” Journal of Chromatography, 312: 413-421.<br />

165


166<br />

Hydrologist JASON VOGEL has been with the United States Geological Survey in<br />

Lincoln, Nebraska, <strong>for</strong> a little over a year. During that time, he has been the Project Chief<br />

<strong>for</strong> a bank-filtration study in cooperation with the United States Environmental Protection<br />

Agency and the City of Lincoln <strong>Water</strong> System, and is also Lead Scientist of the agricultural<br />

chemical transport team in the Nebraska District. Be<strong>for</strong>e joining the United States<br />

Geological Survey, Vogel was a research engineer in the Biosystems Engineering Department<br />

at Oklahoma State University <strong>for</strong> 5 years. He has published articles on a wide variety of<br />

topics, including geostatistics and stochastic design, vadose zone transport, and microbial transport in<br />

groundwater, and co-authored the chapter on Geostatistics in Statistical Methods in Hydrology, Second Edition,<br />

with C.T. Haan. Vogel received a B.S. in Biological Systems Engineering from the University of Nebraska,<br />

an M.S. in Agricultural Engineering from Texas A&M University, and Ph.D. in Biosystems Engineering from<br />

Oklahoma State University.


Session 9: Public Policy and Regulatory<br />

Riverbank Filtration as a Regional Supply Option<br />

<strong>for</strong> the United States<br />

Leo Gentile, P.G., CPG<br />

Jordan Jones & Goulding<br />

Norcross, Georgia<br />

David Haas, P.E.<br />

Jordan Jones & Goulding<br />

Norcross, Georgia<br />

D. Joseph Hagerty, Ph.D., P.E.<br />

University of Louisville, Kentucky<br />

Louisville, Kentucky<br />

Peggy H. Duffy, P.E.<br />

Hagerty Engineering<br />

Jeffersonville, Indiana<br />

Introduction<br />

<strong>RBF</strong> has been used on the lower Rhine River <strong>for</strong> over 130 years and, today, comprises 15 to 20 percent<br />

of the water supply in Germany as a whole (Schubert, 2000). In the United States, the majority<br />

of larger water systems (serving 10,000 or more) use surface water (Wang et al., 2002). Smaller<br />

systems supplying 10,000 or less often rely on groundwater <strong>for</strong> supply because of reduced treatment<br />

requirements. <strong>RBF</strong> is employed in communities in the United States such as Cedar Rapids, Iowa;<br />

Lincoln, Nebraska; Louisville, Kentucky; and Sonoma County, Cali<strong>for</strong>nia. Some communities<br />

such as St. Helens, Oregon, and Sioux Falls, South Dakota, essentially use <strong>RBF</strong> by drawing water<br />

from collector wells constructed adjacent to rivers, although this is usually recognized as GWUDI<br />

(Wang et al., 2002). Still, <strong>RBF</strong> is an underutilized supply whose benefits are likely available to<br />

more communities to help provide better and more uni<strong>for</strong>m raw-water quality. <strong>RBF</strong> potential is<br />

greatest <strong>for</strong> communities in the Midwestern United States that are located where glacial valleyfill<br />

aquifers occur that are interconnected to adjacent surface-water bodies.<br />

Conditions<br />

The need <strong>for</strong> water is the basic condition <strong>for</strong> <strong>RBF</strong> or any other supply source. The magnitude of<br />

the demand and physical conditions dictates what supply source is most feasible. If a community<br />

is located along a major river or stream, then surface-water intake is a relatively simple solution;<br />

however, when water treatment issues and costs are considered, then groundwater — which may<br />

require no more treatment than disinfection — is the next logical choice. <strong>RBF</strong> is an alternative<br />

Correspondence should be addressed to:<br />

Leo Gentile, P.G., CPG<br />

Senior Project Hydrogeologist<br />

Jordan, Jones & Goulding, Inc.<br />

6801 Governors Lake Parkway • Norcross, Georgia 30071 USA<br />

Phone: (678) 333-0148 • Fax: (770) 455-7391 • Email: LGENTILE@JJG.com<br />

167


168<br />

supply option when these conditions are not available that provides typically larger quantities<br />

than wells, and more uni<strong>for</strong>m and higher quality raw water compared to surface-water intake.<br />

Aquifer Thickness and Extent: To supply the millions of gallons a day needed <strong>for</strong> a typical<br />

municipal water supply, the aquifer needs to be sufficiently thick and extensive to sustain such a<br />

yield without depleting the aquifer. If the aquifer is thin (less than about 100-ft thick), the cone<br />

of depression and amount of drawdown from wells penetrating the aquifer will be narrow and<br />

steep, and the well will likely not be efficient in producing the required yields. Excessive<br />

drawdown indicates an inefficient well and/or an over-drafted aquifer. Constructing collector wells<br />

with multiple horizontal lateral collector galleries will help increase well efficiency and create a<br />

broader cone of depression with less drawdown. Still, if the aquifer is relatively thin and limited<br />

in aerial extent, the sustained (or safe yield) may not be sufficient to meet demand. With <strong>RBF</strong>,<br />

conventionally constructed wells, collector wells, or specially designed tunnels overcome the<br />

physical shortcomings of the aquifer. The leakance from an adjacent river or lake recharges the<br />

aquifer, augmenting the safe yield (Figure 1). Thus, a relatively thin and aerially limited aquifer<br />

can yield the water quantities needed <strong>for</strong> municipal supplies. <strong>RBF</strong> also yields water of more<br />

consistent and better overall quality.<br />

River or<br />

Recharge Pond<br />

<strong>Water</strong> Table<br />

A B<br />

NOT TO SCALE<br />

River or<br />

Recharge Pond<br />

NOT TO SCALE<br />

Pumping Well<br />

<strong>Water</strong> Table<br />

Modified from Gallaher and Prize, 1966<br />

EXPLANATION<br />

Figure 1. Groundwater flow recharge to a river (A) and an <strong>RBF</strong> well (B) in a sand and gravel aquifer (from<br />

Lloyd and Lyke, 1995).<br />

Aquifer Characteristics: Glacial and alluvial sand and gravel deposits are some of the most<br />

productive aquifers in the world (Todd, 1980). The hydraulic conductivities and amount of water<br />

in storage are often high. The recharge potential from adjacent streams, rivers, or lakes is also<br />

considered highly favorable as natural or induced vertical gradients, coupled with good hydraulic<br />

conductivity, allows <strong>for</strong> leakance from the surface-water body to the aquifer. Pleistocene-age sand<br />

and gravel deposits have been used <strong>for</strong> well fields and <strong>RBF</strong> sites. By comparison, high hydraulic<br />

conductivity and yields can also occur in bedrock aquifers with fracture or karst porosity; however,<br />

storage capacity may be limited. Furthermore, groundwater flow through fractures or karst porosity<br />

is essentially conduit flow. Recharge from adjacent rivers or lakes would be minimally filtered,<br />

reducing the benefit from <strong>RBF</strong>.<br />

Other Considerations: There must be effective interconnection between the aquifer and surface-water<br />

body. Occasionally, an aquifer is confined or semi-confined by a layer of finer sediments that<br />

reduces the potential <strong>for</strong> leakance from the surface to the aquifer (Figure 2). <strong>RBF</strong> sites are more<br />

efficient where there is good interconnection (over 90 percent of withdrawn water can be bank<br />

filtrate [Eckert et al., 2000]). Clogging of the interconnection through <strong>RBF</strong> use, or the natural<br />

accumulation of fines, also reduces the efficacy of <strong>RBF</strong>. The removal of streambed fines by water<br />

flow from naturally high-gradient streams and rivers or during floods will help restore infiltration.


FEET<br />

1,000<br />

900<br />

800<br />

700<br />

600<br />

VERTICAL SCALE GREATLY EXAGGERATED<br />

DATUM IS SEA LEVEL<br />

Modified from Gann and others, 1973<br />

0 1 2 MILES<br />

Figure 2. Glacial valley-fill aquifer overlain by fine grained sediments in the Great Miami River Valley, Ohio<br />

(from Miller and Appel, 1997).<br />

German researchers experimented with excavating a “window” by dredging accumulated<br />

sediment. While the measure increased <strong>RBF</strong> efficiency, it was considered temporary, lasting only<br />

weeks be<strong>for</strong>e silting over (Schubert, 2000). Rivers whose flow has been altered <strong>for</strong> flood control<br />

and navigation improvement, such as the Ohio River, are particularly prone to clogging. Steep<br />

gradients and floodwaters that may have scoured the riverbed in the past have been eliminated.<br />

Wells or specially designed tunnels used to extract groundwater and bank filtrate must also be<br />

properly located to maximize system efficiency. Rivers and streams that occur with the thumbprint<br />

of the most recent continental glacial period in the United States (e.g., Wisconsin ice age) have<br />

a similar terraced profile. The unaltered modern river channel carries normal stage flows within<br />

the primary terrace. This terrace corresponds to a 100-year flood plain. Where significantly altered<br />

<strong>for</strong> flood control or navigation, the river may now have inundated the primary terrace extending<br />

from bank to bank up to the secondary terrace. The secondary terrace extends outward to the<br />

500-year flood plain. This channel was cut into the bedrock by glacially fed rivers. The rivers were<br />

large braided streams laden with sand and gravel. As the glacial outwash diminished, the scour<br />

channels filled with sand and gravel to become aquifers. The tertiary terrace is the remnant of<br />

peak stage flooding and its resultant scouring into the bedrock. Relatively thin and finer grained<br />

sediments were deposited as the flood stage fell (see Figure 2). These sediments may have <strong>for</strong>med<br />

local, relatively less-productive aquifers that are not connected to the modern river channel. The<br />

secondary terrace is the most productive place to locate <strong>RBF</strong> wells (e.g., out of main navigation,<br />

interconnected to the river, relatively thicker aquifer, greater potential <strong>for</strong> effective filtration).<br />

Potential <strong>RBF</strong> Regions<br />

Grand River<br />

B<br />

0 1 2 KILOMETERS<br />

The greatest potential <strong>for</strong> <strong>RBF</strong> — and, in fact, where <strong>RBF</strong> has been implemented — in the United<br />

States is in the Midwestern states. Figure 3 shows the location of mid- to large-sized communities<br />

in the United States that are located along rivers and on alluvial aquifers. Due to site conditions,<br />

<strong>RBF</strong> is a viable water supply option <strong>for</strong> many, but not all, communities.<br />

Midwest stream-valley aquifers in Midwestern states west of the Mississippi River are important<br />

sources of water <strong>for</strong> communities and industries (Figure 4). These aquifers are unconsolidated sand<br />

and gravel deposits that are thicker, greater in extent, and more productive in the major river<br />

valleys of the region. The stream-valley aquifers are unconfined and in direct connection to adjacent<br />

streams and rivers. Average yields from wells range from 100 to 1,000 gallons per minute, with some<br />

wells yielding over 3,000 gallons per minute (Olcott, 1992). <strong>RBF</strong> opportunities may be available<br />

Buried Valley<br />

B<br />

EXPLANATION<br />

Clay, silt, and fine-grained sand<br />

Coarse-grained sand and gravel<br />

Sand, gravel, and boulders<br />

Till<br />

Bedrock<br />

169


170<br />

EXPLANATION<br />

Dune Sand<br />

NOT TO SCALE<br />

Recharge<br />

Discharge<br />

South<br />

N<br />

Figure 3. Potential <strong>RBF</strong> areas in the United States (from USGS, 1965).<br />

Evaporation<br />

Arkansas<br />

River<br />

Evapotranspiration<br />

<strong>Water</strong> Table<br />

Stream to<br />

Aquifer Leakage<br />

Aquifer<br />

Bedrock<br />

East<br />

using stream-valley aquifers that are located along major rivers such as the Arkansas, Des Moines,<br />

Grand (Michigan), Mississippi, Platte, Missouri, and Wisconsin rivers, plus numerous mediumsized<br />

streams that course the region (Miller and Appel, 1997).<br />

West<br />

Precipitation<br />

Canal Diversion System<br />

Irrigation<br />

Well<br />

Crop Consumptive Use<br />

Deep Percolation<br />

of Precipitation and<br />

Applied Irrigation <strong>Water</strong><br />

Pumpage<br />

Modified from Barker and Others, 1983<br />

North<br />

Figure 4. Midwestern stream valley site favorable <strong>for</strong> <strong>RBF</strong> (Arkansas River Valley) (from Miller and Appel,<br />

1997).<br />

Legend<br />

Metropolitan Areas with<br />

Bank Infiltration Potential<br />

Population 250,000–1,000,000<br />

Population 1,000,001–2,500,000<br />

Population greater than 2,500,000<br />

Unconsolidated Aquifers<br />

Alluvial aquifers suitable <strong>for</strong> bank infiltration<br />

Sand and gravel<br />

Consolidated Rock Aquifers<br />

Sandstone: includes some unconsolidated sand<br />

Carbonate rock: limestone and dolomite;<br />

and in Texas and Oklahoma, some gypsum<br />

Sandstone and carbonate rocks<br />

Volcanic rocks, chiefly basalt<br />

Crystalline rocks, igneous and metamorphic<br />

Withdrawals from wells


The surficial aquifers in Midwestern states that border the Ohio River and are east of the<br />

Mississippi River are similar to those described previously, but are primarily of glacial origin. These<br />

aquifers are very productive (1,000 gallons per minute) and supply almost 50 percent of the fresh<br />

groundwater produced in the region. The course-grained aquifers are divided into two categories:<br />

• Deposits at or near the land surface occurring in stream and river valleys.<br />

• Deposits buried by a layer of fine-grained material that occur in <strong>for</strong>mer river valleys cut<br />

into bedrock and filled with coarse-grained glacial outwash.<br />

The sands and gravels range in thickness from less than 100 to over 600 ft in some buried bedrock<br />

valleys. Large yields are possible from wells completed in the glacial outwash aquifers that are<br />

hydraulically connected to streams, rivers, and lakes, and the wells are sufficiently close to the surfacewater<br />

body (Lloyd and Lyke, 1995). Communities located near rivers such as the Illinois, Kaskaski,<br />

Wabash, White, Kankakee, Maumee, Great Miami, and Scioto rivers could benefit from <strong>RBF</strong>.<br />

Valley-fill glacial aquifers also occur in the Northeastern United States. While extensive, the<br />

aquifer thickness and productivity is less than those occurring in Midwestern states (Olcott,<br />

1995). In general, the valley-fill aquifers are less common along major streams and rivers, so the<br />

potential <strong>for</strong> <strong>RBF</strong> is considered lower than in the Midwestern United States. Alluvial aquifers<br />

occur along streams and rivers in the Northwestern states, but sparse population and relatively low<br />

water demands can be met through either surface-water intakes or groundwater wells. In the arid<br />

Southwest, few major rivers or streams are perennial, thereby limiting <strong>RBF</strong> potential. <strong>Water</strong><br />

supplies in Gulf Coast states are met through groundwater withdrawals from productive regional<br />

aquifers and surface water. In the Piedmont region, water needs are met through surface water, as<br />

rainfall is usually adequate to meet demands and aquifers are fractured crystalline rock, which are<br />

not considered viable <strong>RBF</strong> areas.<br />

REFERENCES<br />

Eckert, P., C. Blomer, J. Gothardt, S. Kamphausen, D. Liebich, and J. Schubert (2000). “Correlation between<br />

the Well Field Catchment and Transient Flow Conditions.” Proceedings, International Riverbank Filtration<br />

Conference, November 2 - 4, Dusseldorf, Germany.<br />

Lloyd, O.B., and W.L. Lyke (1995). “Ground <strong>Water</strong> Atlas of the United States, Segment 10 - Indiana,<br />

Illinois, Kentucky, Ohio.” U.S. Geological Survey Hydrologic Investigations Atlas 730 -K.<br />

Miller, J.A., and C.L. Appel (1997). “Ground <strong>Water</strong> Atlas of the United States, Segment 3- Kansas,<br />

Missouri, Nebraska.” U.S. Geological Survey Hydrologic Investigations Atlas 730 -D.<br />

Olcott, P.G. (1995). “Ground <strong>Water</strong> Atlas of the United States, Segment 12 - Connecticut, Maine,<br />

Massachusetts, New Hampshire, New York, Rhode Island, Vermont.” U.S. Geological Survey Hydrologic<br />

Investigations Atlas 730 -J.<br />

Olcott, P.G. (1992). “Ground <strong>Water</strong> Atlas of the United States, Segment 12 - Iowa, Michigan, Minnesota,<br />

Wisconsin.” U.S. Geological Survey Hydrologic Investigations Atlas 730 -M.<br />

Schubert, J. (2000). “How Does it Work: Field Studies on Riverbank Filtration.” Proceedings, International<br />

Riverbank Filtration Conference, November 2 - 4, Dusseldorf, Germany.<br />

Todd, D.K. (1980). Groundwater Hydrology, John Wiley & Sons, New York.<br />

United States Geological Survey (1965). Productive Aquifers in the Conterminous United States, United States<br />

Geological Survey.<br />

Wang, J.Z., S.A. Hubbs, and R. Song (2002). Evaluation of Riverbank Filtration as a Drinking <strong>Water</strong> Treatment<br />

Process, American <strong>Water</strong> Works Association <strong>Research</strong> Foundation and American <strong>Water</strong> Works Association.<br />

171


172<br />

LEO GENTILE is Senior Hydrogeologist and Group Manager with Jordan, Jones &<br />

Goulding in Atlanta, Georgia. He has a broad range of experience in applied geology and<br />

hydrogeology in the United States. His background includes over 20-years of experience<br />

in technical support and management on a broad range of projects, including waterresources<br />

development, remedial investigations, and feasibility studies at Comprehensive<br />

Environmental Resource Compensation and Liability Act sites; Resource Conservation<br />

and Recovery Act facility investigation and closure; investigation and remediation of<br />

agricultural chemical sites; and assessments and reclamation <strong>for</strong> mining and petroleum-related sites. Gentile<br />

received a B.S. in Geology/Mineralogy from Ohio State University, an M.S. in Petroleum Geology from<br />

Oklahoma State University, and an MBA in Enterprise Risk Management/Finance from Georgia State<br />

University. He is a Registered Professional Geologist in seven states and Puerto Rico, and is a Certified<br />

Professional Geologist by the American <strong>Institute</strong> of Professional Geologists.


Session 9: Public Policy and Regulatory<br />

Application of the Long Term 2 Enhanced Surface<br />

<strong>Water</strong> Treatment Rule Microbial Toolbox at Existing<br />

<strong>Water</strong> Plants<br />

Richard A. Brown<br />

Environmental Engineering and Technology, Inc.<br />

Newport News, Virginia<br />

All surface-water utilities in the United States will be required to comply with specific Cryptosporidium<br />

removal/inactivation targets in the LT2ESWTR, depending upon their bin classification as a<br />

result of raw-water Cryptosporidium occurrence levels. The “microbial toolbox” is intended to<br />

provide utilities with a range of treatment options <strong>for</strong> meeting LT2ESWTR compliance<br />

requirements.<br />

This presentation will include a general discussion of LT2ESWTR requirements, including two of<br />

its main components:<br />

• Cryptosporidium sampling.<br />

• Treatment credits from the microbial toolbox.<br />

In particular, this will include a discussion of how these requirements relate to systems with <strong>RBF</strong><br />

or other groundwater wells that are GWUDI. LT2ESWTR requirements <strong>for</strong> <strong>RBF</strong> are different <strong>for</strong><br />

systems in place at the time the Rule is published versus well systems installed later. New systems<br />

are eligible <strong>for</strong> treatment credits if they meet USEPA-defined requirements <strong>for</strong> media gradation,<br />

separation distance from the associated surface-water source, and turbidity monitoring. Existing<br />

<strong>RBF</strong> and GWUDI wells are not eligible <strong>for</strong> any direct credits, but will produce indirect credits if<br />

Cryptosporidium bin assignment samples are collected from well effluent. According to the<br />

LT2ESWTR, these Cryptosporidium samples must be collected from well effluent if well water is<br />

sent to a subsequent treatment process be<strong>for</strong>e being delivered to consumers. Public water-supply<br />

GWUDI or <strong>RBF</strong> wells that provide water directly to customers without subsequent physical<br />

treatment must collect Cryptosporidium bin assignment samples from the surface-water source.<br />

Additional benefits of <strong>RBF</strong> (removal or degradation of other contaminants and microorganisms,<br />

dampening of water quality and temperature spikes, etc.), as well as potential issues of concern<br />

(clogging of pores, scour associated with flooding/high stage events), will also be outlined.<br />

Other microbial toolbox alternatives will be outlined, though in less detail, and cost comparisons<br />

<strong>for</strong> different treatment strategies will also be discussed. Strategies <strong>for</strong> source-water sampling and<br />

the preparation of contingency plans will also be addressed.<br />

Correspondence should be addressed to:<br />

Richard A. Brown<br />

Engineer<br />

Environmental Engineering and Technology<br />

712 Gum Rock Court • Newport News, Virginia 23606 USA<br />

Phone: (757) 873-1534 • Fax: (757) 873-2392 • Email: rbrown@EETinc.com<br />

173


174<br />

RICHARD BROWN, an Environmental Engineer at Environmental Engineering and<br />

Technology, Inc. in Newport News, Virginia, has almost 20 years of experience working<br />

on surface-water and groundwater quality, regulatory compliance, and water-treatment<br />

projects. Recently, this has included extensive involvement with the American <strong>Water</strong><br />

Works Association, Association of Metropolitan <strong>Water</strong> Agencies, and various United<br />

States drinking-water utilities during negotiations with the United States Environmental<br />

Protection Agency regarding the Stage 2 Disinfection By-Products Rule and Long Term 2<br />

Enhanced Surface <strong>Water</strong> Treatment Plant. Prior, Brown worked <strong>for</strong> 11 years at the Los Angeles County<br />

Sanitation Districts, per<strong>for</strong>ming and directing numerous projects involving the investigation and<br />

interpretation of groundwater chemistry at six municipal solid waste landfills. The American <strong>Water</strong> Works<br />

Association <strong>Research</strong> Foundation study that is the subject of Brown’s presentation during the Second<br />

International Riverbank Filtration Conference will be published in the American <strong>Water</strong> Works Association’s<br />

e-Journal in September 2003. Brown received a B.S. in Civil Engineering from Purdue University and an<br />

M.S. in Environmental Engineering from the University of North Carolina at Chapel Hill.


Session 9: Public Policy and Regulatory<br />

Draft Protocol <strong>for</strong> the Demonstration of<br />

Effective Riverbank Filtration<br />

William D. Gollnitz<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

Verna J. Arnette, P.E.<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

Bruce L. Whitteberry<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

Introduction<br />

<strong>RBF</strong> is a natural in situ filtration process by which microbial pathogens in surface water are<br />

removed via the porous media of the streambed and aquifer under induced infiltration conditions<br />

created by a pumping well. With the promulgation of the LT2ESWTR, the USEPA will allow<br />

0.5- and 1.0-log treatment credit <strong>for</strong> Cryptosporidium removal using <strong>RBF</strong> (USEPA, 2003).<br />

Based upon a review of data from several <strong>RBF</strong> systems in the United States, it is not uncommon<br />

<strong>for</strong> <strong>RBF</strong> to achieve a removal per<strong>for</strong>mance better than the proposed 1.0-log reduction credit. The<br />

0.5- and 1.0-log credits are based on general conservative requirements obtainable by most <strong>RBF</strong><br />

systems; however, there are still several concerns with the effectiveness of <strong>RBF</strong> during worst-case<br />

hydrologic and water-quality conditions. For example, high flow events may significantly scour a<br />

streambed, thus reducing its filtration capabilities. To date, there has been a reluctance to provide<br />

additional credit because an acceptable procedure has not been developed that will allow <strong>for</strong> the<br />

demonstration of consistent removal during these periods of high risk.<br />

The Greater Cincinnati <strong>Water</strong> Works has developed the following draft protocol <strong>for</strong> demonstrating<br />

<strong>RBF</strong> pathogen removal. The draft protocol will:<br />

• Allow utilities a method arguing <strong>for</strong> the re-designation of a groundwater from GWUDI<br />

back to “groundwater.”<br />

• Allow <strong>for</strong> demonstration of 2.0-log Cryptosporidium removal in lieu of engineered filtration<br />

under the Interim Enhanced Surface <strong>Water</strong> Treatment Rule.<br />

• Allow <strong>for</strong> greater than 1.0-log treatment credit under LT2ESWTR.<br />

• Provide consistency <strong>for</strong> evaluating <strong>RBF</strong> sites.<br />

The draft protocol is based upon a review of several regulatory guidance publications dealing with<br />

GWUDI evaluations and “alternative filtration technology”(USEPA, 1991; USEPA, 1992;<br />

Correspondence should be addressed to:<br />

William D. Gollnitz<br />

Supervisor of Treatment, <strong>Water</strong> Quality & Treatment Division<br />

Greater Cincinnati <strong>Water</strong> Works<br />

5651 Kellogg Ave • Cincinnati, Ohio 45228 USA<br />

Phone: (513) 624-5657 • Fax: (513) 624-5670 • Email: William.Gollnitz@gcww.cincinnati-oh.gov<br />

175


176<br />

Wilson et al., 1996; USEPA, 2001). In addition, the Greater Cincinnati <strong>Water</strong> Works has<br />

reviewed data from an extensive flowpath study conducted in cooperation with the USGS at the<br />

Greater Cincinnati <strong>Water</strong> Works’ Charles M. Bolton Well Field (Sheets et al., 2002; Gollnitz et<br />

al., 2003), as well as published data from other <strong>RBF</strong> sites such as Louisville, Kentucky (Wang et<br />

al., 2002); Casper, Wyoming (Gollnitz et al., 1997); and others (Cote et al., 2002). The protocol<br />

is in response, in part, to requests <strong>for</strong> comments on additional <strong>RBF</strong> credit in the LT2ESWTR. It<br />

is hoped that the protocol will provide initial guidance to the USEPA, primacy agencies, and<br />

water utilities. The draft <strong>for</strong>mat of the protocol will facilitate comment and modification by<br />

regulatory and utility personnel.<br />

The <strong>RBF</strong> process is very complex and is still not fully understood; however, our knowledge to date<br />

does pinpoint several controlling factors that are primary with respect to the removal of microbial<br />

pathogens and surrogates. These factors are identified in Cote et al. (2002) and include:<br />

• Assessment of river and groundwater quality.<br />

• Flow velocity through the streambed and aquifer.<br />

• Dilution with regional groundwater.<br />

The <strong>RBF</strong> protocol has, there<strong>for</strong>e, been designed around:<br />

• Characterizing the aquifer and its capability as a <strong>RBF</strong> system.<br />

• Identifying conditions <strong>for</strong> testing during periods of high flow velocity.<br />

• Testing during periods of minimal dilution with regional groundwater.<br />

The latter two items involve identifying periods of maximum induced filtration during high stage<br />

events (as related to storm or reservoir release events) and maximum groundwater pumping. Theoretically,<br />

these two factors would correspond with periods of high velocity and minimum dilution.<br />

Pre-Demonstration Evaluation<br />

The purpose of the pre-demonstration evaluation is to characterize the aquifer and collection<br />

devices with respect to hydrology and water quality. This includes a determination of the worst-case<br />

conditions <strong>for</strong> minimal natural filtration of pathogenic protozoa. This evaluation will then identify<br />

the type of demonstration project needed <strong>for</strong> determining treatment credit. The pre-demonstration<br />

evaluation will organize and evaluate existing data and will identify what additional data is needed<br />

to complete this first phase. The pre-demonstration project will specifically look at the following<br />

important aspects of the natural filtration process:<br />

1. Surface-water and groundwater quality.<br />

• Filtration capability.<br />

• Contaminant and surrogate concentrations.<br />

• Continuous monitoring of turbidity.<br />

2. Potential infiltration rate variability.<br />

3. Time of travel between surface-water and groundwater collection devices.<br />

The first aspect of Item 1 is to characterize source-water quality with respect to its filtration<br />

capability. It is recommended that the utility per<strong>for</strong>m analyses to determine the ionic strength of<br />

both surface water and groundwater. The ionic strength may provide an indication of the filtration<br />

capability of the sediments. Cote et al. (2002) and others have suggested that a high ionic strength<br />

improves the removal of particles; however, its role is not yet fully understood. It should be noted


that multiple samples should be evaluated. Ionic strength may vary between low (primarily<br />

groundwater inflow) and high (primarily runoff) river flow conditions.<br />

The second aspect of Item 1 is to determine the level of contaminant (Giardia, Cryptosporidium)<br />

and surrogate (algae, diatoms, spores, particle counts) concentrations and their seasonal<br />

variability. If this data does not already exist, it is recommended that the utility per<strong>for</strong>m monthly<br />

Giardia/ Cryptosporidium monitoring using Method 1623 (the same as required under<br />

LT2ESWTR), monthly Microscopic Particulate Analysis using the USEPA Consensus Method<br />

(1992), and weekly endospore analysis (Rice, 1996). Another optional surrogate parameter is<br />

particle counts. Particle counts should be in the two size ranges <strong>for</strong> Giardia (7 to 10 micrometers)<br />

and Cryptosporidium (3 to 5 micrometers). Data collection <strong>for</strong> all parameters selected should be<br />

representative of one full hydrologic cycle (e.g., 1 year). The concentrations of pathogens,<br />

primarily Cryptosporidium, will determine the level of treatment needed. Because Giardia and<br />

Cryptosporidium concentrations are typically low in surface waters, combined with detection<br />

limitations during sample analysis, they cannot be used to determine log-reduction credit. Log<br />

reductions should be based upon levels of algae, diatoms (from Microscopic Particulate Analysis),<br />

endospores, and particle counts; however, prior to per<strong>for</strong>ming the demonstration project,<br />

measurable concentrations need to be established. Particle counts may be used; however, they<br />

should be considered secondary due to the fact that they detect sediment or natural microbial<br />

growth (e.g., iron bacteria) in groundwater from wells. Typically, log reductions using particle<br />

counts are lower as compared to algae and spores due to the measurement of other inert and<br />

microbial particles.<br />

The third aspect of Item 1 is the continuous monitoring of turbidity. Turbidity monitoring is a<br />

requirement <strong>for</strong> engineered filtration systems, and has been included as a requirement on each<br />

collection device <strong>for</strong> <strong>RBF</strong> credit under LT2ESWTR <strong>for</strong> the period that the device is in use.<br />

Groundwater from <strong>RBF</strong> sites should be consistently low (0.01 to 1.0 ntu). Significant excursions<br />

above the normal level, especially after high infiltration periods, should be investigated. As with<br />

particle counts, turbidity also measures inert geologic material and other non-pathogenic microbials.<br />

As part of the pre-demonstration phase, it is important to estimate the range of potential unit<br />

infiltration rates out in the active area of the streambed, articularly to estimate the highest value<br />

under conditions of high-river stage and high groundwater pumping during periods when river<br />

water is warm. Typically, infiltration rates at <strong>RBF</strong> sites are lower than slow sand filters, but should<br />

be investigated to see if they are higher. Infiltration rates can be calculated using Darcy’s Law, with<br />

input variables being river stage, groundwater elevations below the river stage, streambed permeability,<br />

streambed thickness, and surface-water temperature (affecting viscosity) (Gollnitz et al.,<br />

2002). River-stage data may be obtained from a local United States geological gaging station. If<br />

this data is not available, the maximum stage elevation may be estimated from the depth of the<br />

river channel. The groundwater elevation should be measured below the active infiltration area<br />

of the streambed using piezometers in the streambed or monitoring wells on each side of the river.<br />

Streambed permeability can be estimated using pump test analysis (preferred, but costly), seepage<br />

meters, temperature probes, and piezometers (each method has inherent problems and should be<br />

considered carefully be<strong>for</strong>e selection). Streambed thickness can be estimated by visually<br />

inspecting sediments from shallow excavations of the streambed during low flow. Temperature can<br />

be measured directly in surface water. Surface-water temperature typically reflects the average<br />

daily ambient air temperature. If this data is not available, it may be necessary to collect it during<br />

the pre-demonstration period. Another important aspect is to look at the frequency of the<br />

occurrence of high infiltration events. Frequency analysis requires long-term river stage data,<br />

which may be limited <strong>for</strong> most <strong>RBF</strong> sites.<br />

177


178<br />

<strong>Water</strong>-quality monitoring during the demonstration project should consider the travel time<br />

between the river and collection device. Travel time should be calculated <strong>for</strong> the shortest flowpath<br />

(e.g., fastest potential time of travel) during continuous pumping of the collection device. Time of<br />

travel can be estimated using various techniques ranging from tracer studies (conductivity,<br />

temperature, and chloride), analytical flow calculations, and groundwater flow modeling. Each<br />

method has inherited limitations that must be considered. It is recommended that the utility<br />

determine a sampling window based upon a realistic range of travel times using two or more methods.<br />

Another important consideration is the travel time of particles. Larger particles such as algae may<br />

have longer travel times due to retardation in the aquifer. Sampling windows should be long<br />

enough to take these into consideration. For example, Table 1 lists various travel-time estimates<br />

<strong>for</strong> Production Well 1 at the Charles M. Bolton Well Field, including limitations. The fastest<br />

theoretical time is 1 day; however, a more accurate estimate using conductivity indicates that the<br />

travel time of groundwater ranges from 10 to 14 days (Sheets et al., 2002). Interestingly, algae<br />

appear to arrive at the production well much later (approximately 38 days). In this case, a high<br />

infiltration-event monitoring sampling window may be from 1 to 60 days after the event peak. It<br />

is recommended that 10 to 15 samples be collected during this time period, with increased<br />

sampling frequency around the anticipated period <strong>for</strong> possible breakthrough of microbial surrogates.<br />

Table 1. Time of Travel Estimates between the Great Miami River and Production Well 1<br />

at the Charles M. Bolton Well Field<br />

TOT<br />

Methodology (Days) Comments<br />

Calculated Fixed Radius 1 Short TOT; considers only horizontal flowpath;<br />

no gradient<br />

Groundwater Flow Model 8 Limitations with model design and scale<br />

Conductivity 10 to 14 Real time measurement; range based upon multiple<br />

measurements; considered most accurate method<br />

Event Temperature Lag 26 Real-time measurement; problems due to<br />

thermodynamics<br />

Algae Event Peak 11 to 38 Limited data; considers flow characteristics of algae<br />

TOT = Time of travel. CMB = Charles M. Bolton Well Field.<br />

Demonstration Evaluation<br />

The demonstration evaluation is broken down into three options: 1-year monitoring, multi-event<br />

monitoring, or seasonal monitoring. One of these options is selected based upon the results of the<br />

pre-demonstration evaluation. Log reduction <strong>for</strong> all three options should be based upon the<br />

average of Microscopic Particulate Analysis and endospore data.<br />

The 1-year monitoring option is <strong>for</strong> systems with little or limited infiltration rate and water-quality data.<br />

It is recommended that utilities monitor river stage, groundwater elevations, and surface-water<br />

temperature on a daily basis. Daily calculations can be made with Darcy’s law using conservative<br />

values of permeability (e.g., highest measured value) and thickness (lowest measured value).<br />

These calculations can be easily made using a computer spreadsheet (Gollnitz et al., 2002). With<br />

respect to water quality, it is recommended that utilities collect weekly samples of turbidity and<br />

endospores from surface water, along with monthly Giardia, Cryptosporidium, and Microscopic<br />

Particulate Analysis. Each collection device being evaluated should be continuously monitored <strong>for</strong>


turbidity. Endospores, Giardia, Cryptosporidium, and Microscopic Particulate Analysis data should<br />

be sampled at each collection device on a period and frequency determined from the time-of-travel<br />

data. During the 1-year period, an attempt should be made to identify conditions <strong>for</strong> high<br />

infiltration and to monitor one or more events, if possible.<br />

The multi-event monitoring option is provided <strong>for</strong> systems with infiltration rates that are<br />

significantly high, as compared to engineered filtration systems, during storm events, upstream<br />

reservoir releases, and during periods of heavy groundwater pumping. <strong>Water</strong>-quality monitoring<br />

should include multiple turbidity, Giardia, Cryptosporidium, Microscopic Particulate Analysis, and<br />

endospore data collected from the river during and shortly after the river stage peak. As with all<br />

options, groundwater turbidity should be continuously monitored. Groundwater samples <strong>for</strong> the<br />

other parameters should be collected within a defined sampling window using the time-of-travel<br />

data. Endospores and particle counts should be collected more frequently due to their lower cost.<br />

A major advantage to this option is that monitoring resources can be concentrated during these<br />

periods, which in turn may reduce overall costs. Infiltration rate parameters should be monitored<br />

to determine the magnitude of the event.<br />

Option three is seasonal monitoring. Pre-demonstration evaluations may identify if a system is at<br />

risk during a 2- to 4-month period of the year. For example, a controlled river in the arid West<br />

may have a high stage only during the summer months, when water is released from reservoirs <strong>for</strong><br />

irrigation. <strong>Water</strong>-quality monitoring should be concentrated during this high-risk period, taking<br />

time of travel into consideration, as well as estimating the magnitude of infiltration.<br />

REFERENCES<br />

Cote, M.M., M.B. Emelko, and N.R.Thomson (2002). “Factors Influencing Prediction of Cryptosporidium<br />

Removal in Riverbank Filtration Systems: Focus on Filtration.” Proceedings, American <strong>Water</strong> Works Association<br />

WQTC Conference, Denver, CO.<br />

Gollnitz, W.D., J.L. Clancy, and S.C. Garner (1997). “Reduction of Microscopic Particulates by Aquifers.”<br />

Journal AWWA, 89(11).<br />

Gollnitz, W.D., J.L. Clancy, B.L. Whitteberry, and J.A. Vogt (2003). “Riverbank Filtration as a Microbial<br />

Treatment Process.” Journal AWWA (in press).<br />

Gollnitz, W.D., B.L. Whitteberry, and J.A. Vogt (2002). “Induced Infiltration Rate Variability and <strong>Water</strong><br />

Quality.” Proceedings, SW-GW Interactions Conference, American <strong>Water</strong> Resource Association, Keystone, CO.<br />

Rice, E.W., K.R. Fox, R.J. Miltner, D.A. Lytle, and C.H. Johnson (1996). “Evaluating Plant Per<strong>for</strong>mance<br />

with Endospores.” Journal AWWA, Denver, CO.<br />

Sheets, R.A., R.A. Darner, and B.L. Whitteberry (2002). “Lag Times of Bank Filtration at a Well Field,<br />

Cincinnati, Ohio, USA.” Journal of Hydrology, 233: 162-174.<br />

USEPA (1991). Guidance Manual <strong>for</strong> Compliance with the Filtration and Disinfection Requirements <strong>for</strong> Public<br />

<strong>Water</strong> Systems using Surface <strong>Water</strong> Sources, American <strong>Water</strong> Works Association, Denver, CO.<br />

USEPA (1992). Consensus Method <strong>for</strong> Determining Groundwaters under the Direct Influence of Surface <strong>Water</strong><br />

using Microscopic Particulate Analysis (MPA), Port Orchard, WA.<br />

USEPA (2001). Draft, Considerations <strong>for</strong> Evaluating the Effectiveness of Alternative Filtration at Central Wyoming<br />

Regional <strong>Water</strong> System (CWRWS), Region 8, Denver, CO (unpublished letter to CWRWS).<br />

USEPA (2003). <strong>National</strong> Primary Drinking <strong>Water</strong> Regulations: Long Term 2 Enhanced Surface <strong>Water</strong> Treatment<br />

Rule, 40 CFR 141 and 142, Washington, D.C.<br />

Wang, J.Z., S.A. Hubbs, and R. Song (2002). Evaluation of Riverbank Filtration as a Drinking <strong>Water</strong> Treatment<br />

Process, Tailored Collaboration with Louisville <strong>Water</strong> Co., American <strong>Water</strong> Works Association <strong>Research</strong><br />

Foundation, Denver, CO.<br />

Wilson, M.P., W.D. Gollnitz, S.N. Boutros, and W.T. Boria (1996). Determining Ground <strong>Water</strong> under the Direct<br />

Influence of Surface <strong>Water</strong>, American <strong>Water</strong> Works Association <strong>Research</strong> Foundation Project 605, Denver, CO.<br />

179


180<br />

BILL GOLLNITZ is a Supervisor of Treatment <strong>for</strong> the Greater Cincinnati <strong>Water</strong> Works,<br />

where he is responsible <strong>for</strong> water quality and treatment <strong>for</strong> both the Charles M. Bolton and<br />

Mason Ground <strong>Water</strong> Facilities. Gollnitz has been in the water supply industry <strong>for</strong> over<br />

30 years, having previously managed surface-water and groundwater utilities in New York<br />

and Rhode Island. He has also had extensive experience in water-supply protection and<br />

surface-water/groundwater interactions. In addition, he was a Project Manager and co-author<br />

on the American <strong>Water</strong> Works Association <strong>Research</strong> Foundation’s project “Determining<br />

Ground <strong>Water</strong> under the Direct Influence of Surface <strong>Water</strong>.” Gollnitz has completed groundwater under the<br />

direct influence of surface water evaluations in Connecticut, New York, Ohio, Rhode Island, and Wyoming,<br />

and he has been published in the Journal American <strong>Water</strong> Works Association and elsewhere. In 1992, he was<br />

honored with an American <strong>Water</strong> Works Association Best Paper Award. Gollnitz received a B.S. in Biology<br />

from Mount Union College in Alliance, Ohio, and a M.S. in Environmental Science-<strong>Water</strong> Resources from<br />

the State University of New York, College of Environmental Science and Forestry in Syracuse, New York.


Session 9: Public Policy and Registration<br />

Source <strong>Water</strong> Protection and Riverbank Filtration<br />

in the Dyje River Basin<br />

Prof.-Dr. Petr Hlavínek<br />

Brno University of Technology<br />

Brno, Czech Republic<br />

Prof.-Dr. Jaroslav Hlavác˘<br />

Vodárenská Akciová Spolec˘nost a.s.<br />

Brno, Czech Republic<br />

<strong>Water</strong> is not a commercial product like any other but, rather, a heritage that must be protected,<br />

defended, and treated as such. Achieving regional water-quality goals often involves substantial<br />

capital investments and changes in public attitudes concerning resource management. Economic<br />

impacts include:<br />

• The cost of facilities designed to reduce the discharge of contaminants into natural waters<br />

or to improve the quality of waste-receiving waters.<br />

• Limitations on economic activities and economic development in a particular region or<br />

river basin.<br />

Those responsible <strong>for</strong> the <strong>for</strong>mulation and adoption of water-quality plans and management<br />

policies must have a means of estimating and evaluating the temporal and spatial economics,<br />

environmental, and ecologic impacts of these plans and policies. This need has stimulated the<br />

development and application of a wide range of mathematical modeling techniques <strong>for</strong> predicting<br />

the impact of alternative pollution control plans.<br />

The Dyje River is one of the biggest boundary streams in the Czech Republic. It is located in<br />

southern Moravia, crosses the state boundary between Austria and the Czech Republic several<br />

times, and creates the state boundary (with a total length of approximately 50 kilometers). There<br />

are 86 urban areas (agglomerations) larger than 2,000 population equivalent, with a total of more<br />

than 1.2-million inhabitants in the area of interest. The area of the Dyje River Basin is<br />

approximately 12,000 square kilometers, and the monitored river network is approximately<br />

1,000-kilometers long (Figure 1). There is number of <strong>RBF</strong> plants along the Dyje River basin that<br />

are, in the long run, affected by river-water quality.<br />

The Dyje Project deals with the areas that do not yet meet European Union legislation and<br />

correspond with the Instrument <strong>for</strong> Structural Policies <strong>for</strong> Pre-Accession (ISPA) definition of<br />

agglomerations.<br />

Correspondence should be addressed to:<br />

Prof.-Dr. Petr Hlavínek<br />

Professor of Civil Engineering, Faculty of Civil Engineering<br />

Brno University of Technology<br />

Zizkova 17 • 602 00 Brno, Czech Republic<br />

Phone: 420-541147733 • Fax: 420-543245032 • Email: hlavinek.p@fce.vutbr.cz<br />

181


182<br />

Figure 1. Location of the Dyje River Basin in the Cezch Republic. A= Austria. SK = Slovakia.<br />

PL = Polland. D = Germany.<br />

The goal of the project is to ensure that:<br />

• The quality and volume of water collected from polluters and treated in wastewater<br />

treatment plants complies by the year 2005 and is within the area managed by Union<br />

water supply and sewer systems with the values set down by Directive No. 91/271/EEC.<br />

• All wastewater discharged to the sewer is collected.<br />

• Within any drained area serviced by a wastewater treatment plant, all wastewater is<br />

carried to that wastewater treatment plant and treated in compliance with the values set<br />

down by Directive No. 91/271/EEC.<br />

• <strong>National</strong> parks and landscape-protected areas within the region are protected in terms of<br />

surface-water and groundwater quality.<br />

A group of related projects covers the following priority measures, as classified below:<br />

Category A.1: Reconstruct and upgrade existing wastewater treatment plants in municipalities<br />

with populations over 2,000 (or equivalent).<br />

Category A.2: Refurbish existing large wastewater treatment plants (<strong>for</strong> populations over 10,000<br />

[or equivalent]), including equipment <strong>for</strong> the removal of nitrogen compounds and phosphorus.<br />

Category A.3: Reconstruct existing sewer systems connected to existing wastewater treatment<br />

plants to provide sufficient capacity and a high level of treatment.<br />

Category A.4: Complete sewer systems and connect them to existing wastewater treatment plants<br />

to provide sufficient capacity and a high level of treatment.<br />

Category A.5: Construct wastewater treatment plants in municipalities with populations over<br />

2,000 (or equivalent).<br />

Category A.6: Complete sewer systems and wastewater treatment plants in municipalities with<br />

populations over 2,000 (or equivalent).


For the first phase of the Dyje Project, 10 agglomerations were chosen based on area investigations<br />

and comparisons with European Union standards. A technical solution was designed <strong>for</strong> those<br />

locations. Wastewater treatment and collection needs were identified and their standards<br />

compared with European Union standards. See Figure 2 and Table 1.<br />

1. Dyje – Moravská Dyje<br />

2. Jihlava – lower part<br />

3. Jihlava – upper part<br />

4. Oslava<br />

5. Bobrava<br />

6. Svratka – lower part<br />

7. Svratka – upper part<br />

8. Litava<br />

Figure 2. Sub-catchments 1 through 8 of the Dyje River.<br />

Table 1. Basic Data of the Dyje River Catchment Area<br />

Increase in Connected Inhabitants PE 203,606<br />

Increase in Removed Pollution PE 154,905<br />

Total Capacity of New and Upgraded WWTP PE 350,133<br />

Increase in Removed Pollution – SS ton/year 3,110<br />

Increase in Removed Pollution – BOD 5 ton/year 3,392<br />

Increase in Removed Pollution – COD ton/year 6,785<br />

Increase in Removed Pollution – N-NH4 ton/year 622<br />

Increase in Removed Pollution – Total ton/year 141<br />

Increase in Treated Wastewater m 3 /year 9,103,587<br />

Wastewater Treated on WWTPs m 3 /year 20,975,687<br />

Total Length of New and Upgraded WWTP kilometers 601<br />

WWTP = Wastewater treatment plant. SS= Suspended solid. BOD 5 = Biochemical oxygen demand.<br />

COD = Chemical oxygen demand. N-NH4 = Ammonia nitrogen. PE = Population equivalent.<br />

183


184<br />

As a result of improvements in pollution source-management at the Dyje River Basin, mathematical<br />

modeling techniques were used to assess stream-water quality. The study aimed to evaluate<br />

the effect of improvements on stream-water quality at main streams in the Dyje River basin, with<br />

special respect to the Dyje River transboundary profile (downstream of the City of Br˘eclav). The<br />

results were used as bases in the decision-making process. By doing this, financial sources can be<br />

distributed more efficiently in the basin and the application of remediation measures would better<br />

improve water quality while considering the whole spectrum of requirements connected to the<br />

process of approaching the European Union.<br />

There are a number of <strong>RBF</strong> plants along the Dyje River Basin. After decades of safe operation,<br />

some came under heavy pressure due to increasing river-water pollution. As an example, the <strong>RBF</strong><br />

plant at Vojkovice is discussed. See Figures 3 and 4.<br />

Concentration mg/L<br />

2.5<br />

2.0<br />

1.5<br />

1.0<br />

0.5<br />

0<br />

1979<br />

1981<br />

1983<br />

Average<br />

1985<br />

1987<br />

1989<br />

Date<br />

Maximum<br />

Figure 3. Maximum and average concentration of N-NH4 at the <strong>RBF</strong> plant at Vojkovice from 1979 to 2000.<br />

In 1967, 170,000 tons of biochemical oxygen demand (BOD 5) per year were discharged into rivers<br />

in the Czech Republic. From 1957 to 1970, 800 new wastewater treatment plants were built, and the<br />

load decreased to 142,000 tons of BOD 5 in 1975. Between 1970 and 1980, the construction of<br />

wastewater treatment plants was stifled; there<strong>for</strong>e, the load increased to 198,000 tons of BOD 5 in<br />

1980. In 1990, the load decreased to 148,000 tons of BOD 5. Due to the massive construction of<br />

wastewater treatment plants after 1990, the load decreased to 67,000 tons of BOD 5 in 1995 and<br />

to 65,000 tons of BOD 5 in 2000.<br />

Gradual improvements in quality are expected after the construction and completion of the<br />

project, “Protection of <strong>Water</strong> in the Dyje River Basin.” It is expected that a number of <strong>RBF</strong> plants<br />

will be immediately affected (Ivancice, Moravské Bránice), while others will be gradually affected<br />

(Lomnicka, Omice, Vojkovice).<br />

1991<br />

1993<br />

1995<br />

1997<br />

1999


Concentration mg.l 1<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

1979<br />

1981<br />

1983<br />

1985<br />

Average<br />

1987<br />

1989<br />

Date<br />

Maximum<br />

Figure 4. Maximum and average concentration of N-NO 3 in the <strong>RBF</strong> plant at Vojkovice from 1979 to 2000.<br />

The simulation models are a relatively crude approximation of interactions among various<br />

constituents that occur in water bodies. The best water-quality simulation model is the simplest<br />

one that will adequately predict water-quality impacts within a particular water body associated<br />

with a particular water-quality management policy. Yet, in spite of current limitations, simulation<br />

models are the only reasonable means available <strong>for</strong> predicting surface-water quality. The state-ofthe-art<br />

in water-quality modeling and the understanding of physical, chemical, and biological<br />

processes that affect water quality are improving rapidly. The model provides preliminary<br />

estimates; a more precise tool should be used after collecting all necessary data (i.e., especially<br />

more accurate pollution data, corresponding real discharge data at the time of sampling, more<br />

accurate flow hydrodynamics calculation, etc.). The activities mentioned are quite time-consuming<br />

and expensive, yet necessary when improvements to the receiving waters must be more precisely<br />

assessed and quantified.<br />

Regional projects <strong>for</strong> water protection are complex, complicated, multidisciplinary, and long-term.<br />

They are politically important, with complicated relations among authorities both at the local and<br />

international levels. Close cooperation of all involved parties and companies is absolutely<br />

necessary. The Dyje Project was established as a pilot project <strong>for</strong> a regional solution in the Czech<br />

Republic. The project, “<strong>Water</strong> Protection of the River Dyje Basin,” has been accepted <strong>for</strong><br />

financing from the ISPA fund.<br />

1991<br />

1993<br />

1995<br />

1997<br />

1999<br />

185


186<br />

PETR HLAVÍNEK has 20 years experience in wastewater treatment and water quality. He<br />

has 10 years of experience as a lead designer in the design company, HYDROPROJECT.<br />

At present, he is Professor of Civil Engineering at Brno University of Technology, Faculty<br />

of Civil Engineering. He also acts as Vice-President of the Czech Wastewater Treatment<br />

Experts Association, as well as Deputy Head of the <strong>Institute</strong> of Urban <strong>Water</strong> Management<br />

and as Member of the governing board of the Faculty of Civil Engineering at Brno<br />

University of Technology. Hlavínek has completed more than 200 research reports, expert<br />

opinions, and reports dealing with water quality and wastewater treatment. He has published more than 120<br />

articles dealing with wastewater treatment and water quality, and is author and co-author of six books,<br />

including Industrial Wastewater Treatment, Upgrading of Wastewater Treatment Plants, Hydraulics of WWTP,<br />

Wastewater Treatment-Examples of Calculation, Drainage of Urban Areas — Conceptual Approach and <strong>Water</strong><br />

Structures, and <strong>Water</strong> Structures. Hlavínek received an M.S. at Brno University of Technology, a Dipl. in<br />

Sanitary Engineering at IHE Delft Netherlands, and a Ph.D. at Brno University of Technology.


Session 11: Case Studies “Lessons Learned”<br />

Greater Cincinnati <strong>Water</strong> Works Flowpath Study<br />

Field Design: Methodology and Evaluation<br />

Bruce Whitteberry, P.G.<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

William D. Gollnitz<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

Jeffrey Vogt<br />

Greater Cincinnati <strong>Water</strong> Works<br />

Cincinnati, Ohio<br />

Objective<br />

The Greater Cincinnati <strong>Water</strong> Works, USGS, and Miami University of Ox<strong>for</strong>d, Ohio, entered<br />

into a joint research study to evaluate the effectiveness of <strong>RBF</strong> on an unconsolidated sand and<br />

gravel aquifer at the Greater Cincinnati <strong>Water</strong> Works’ Charles M. Bolton Well Field. One of the<br />

goals of this study was to better understand the aquifer’s effectiveness at reducing biological<br />

drinking-water contaminants. The purpose of this presentation is to describe and evaluate the<br />

study’s field design to determine its effectiveness <strong>for</strong> achieving study goals. This presentation will<br />

also discuss water-quality anomalies in the observed data and how they relate to field design.<br />

Methodology<br />

The Charles M. Bolton Well Field is located within the Great Miami Buried Valley Aquifer, which is<br />

composed of sand and gravel. The well field consists of 10 production wells located 50 to 600 ft<br />

from the Great Miami River. The well field is situated along the southern edge of the Great Miami<br />

Buried Valley Aquifer and is bounded by the bedrock valley wall to the south and the Great Miami<br />

River to the north. Due to this configuration, the well field is dependent upon the river as a major<br />

source of induced recharge or infiltration.<br />

Two study sites were chosen within the well field. The sites are adjacent to Production Well 1<br />

(referred to as Site 1) and to Production Well 8 (referred to as Site 8). At each site, four vertical<br />

monitoring wells with 2-ft long well screens were installed along a theoretical flowpath between<br />

the river and production well. Each well was installed progressively deeper, moving away from the<br />

river and toward the well. The two deepest wells were installed to correspond to the top and<br />

bottom of the production well screen. In addition, an inclined well was installed at each site. The<br />

inclined wells were installed at 20- to 30-degree angles from the horizontal and drilled beneath<br />

Correspondence should be addressed to:<br />

Bruce Whitteberry, P.G.<br />

Hydrogeologist<br />

Greater Cincinnati <strong>Water</strong> Works<br />

5651 Kellogg Ave • Cincinnati, Ohio 45228 USA<br />

Phone: (513) 624-5611 • Fax: (513) 624-5670 • Email: Bruce.Whitteberry@gcww.cincinnati-oh.gov<br />

187


188<br />

the river. This allowed water-sample collection from approximately 5- to 10-ft beneath the Great<br />

Miami River.<br />

Each of the vertical flowpath wells was constructed using 4-inch diameter polyvinyl chloride wells<br />

with 2 ft of 10-slot (0.010-inch) screen. The lengths of the screens were designed to provide the<br />

shortest interval necessary to accommodate a sample pump and downhole data.<br />

Once the wells were installed and developed (to clean the well of construction-induced silt), a<br />

data sonde was installed in each well. The data sonde was used to record temperature, conductivity,<br />

pH, dissolved oxygen, and water levels. Each well was fitted with a bladder pump to collect<br />

water samples. By using dedicated equipment <strong>for</strong> sampling, the possibility of introducing contamination<br />

into the well was minimized, sampling was more consistent and efficient, and eliminating<br />

equipment blank samples from the quality control program reduced analytical costs. The inclined<br />

wells were constructed similarly to vertical wells, but with a 6-inch diameter polyvinyl chloride<br />

well casing and 5 ft of 10-slot (0.010-inch) well screen. A submersible sampling pump and data<br />

sonde were installed in each inclined well.<br />

To monitor the river, USGS installed a stream gage station at Site 1. The gage continuously<br />

measured and recorded river stage. It was also fitted with an automated sampling pump and data<br />

sonde to collect readings of the same parameters measured in the wells.<br />

Prior to each sampling event, the monitoring wells were purged using low-flow (minimal<br />

drawdown) purging methodology. Each well was pumped at a rate of less than 1 liter per minute<br />

(1 to 3 liters per minute <strong>for</strong> the inclined wells), while minimizing drawdown in the wells (usually<br />

less than 0.02 ft). Temperature, pH, specific conductance, and dissolved oxygen were measured<br />

during purging, and the well was sampled when the parameters stabilized according to USGS<br />

<strong>National</strong> <strong>Water</strong>-Quality Assessment Program Sampling Protocol (Koterba et al., 1995). With<br />

low-flow techniques, sampling-induced turbidity problems can sometimes be minimized. Because<br />

the goal of this study was to sample discrete portions of the aquifer in close proximity, this<br />

technique was also desirable to minimize the mixing from portions of the aquifer above or below<br />

the screened area.<br />

Anomalous Data<br />

Because the purpose of this study was to evaluate the effectiveness of the aquifer in acting as a<br />

filtration system, significant anomalies in the data were of concern because of their potential to<br />

represent a preferred flowpath through which water is not effectively filtered. One anomaly<br />

observed was the concentrations of TOC in Well FP8B. Because this well is of intermediate depth,<br />

the TOC concentration was expected to be between the concentrations of shallower (FP8A) and<br />

deeper (FP8C) wells. This is the trend seen at Site 1; however, during the first 2 months of the<br />

study, FP8B had TOC concentrations several times higher than the river (Figure 1). Throughout<br />

the first 2 months of monitoring, TOC dropped steadily and stabilized, with concentrations<br />

between those of Wells FP8A and FP8C.<br />

Another anomaly identified in the study was high particle counts in Wells FP8B and FP8D. Well<br />

FP8B particle counts were lower than the river, but higher than FP8A (a shallower well closer to<br />

the river). Particle counts in Well FP8D, the deepest monitoring well, were higher or very near<br />

the levels of particles in the river (Figure 2).


Concentration (mg/L)<br />

35<br />

30<br />

25<br />

20<br />

15<br />

10<br />

5<br />

0<br />

Aug-99 Nov-99 Feb-00 May-00 Sep-00 Dec-00 Mar-01 Jul-01<br />

Figure 1. Total organic carbon at Site 8.<br />

Counts/100 ml<br />

10000000<br />

1000000<br />

100000<br />

10000<br />

1000<br />

100<br />

Discussion<br />

Production Well 8 FP8A FP8B FP8C River<br />

Date<br />

Production Well 8 FP8A FP8B FP8D River<br />

10<br />

Aug-99 Dec-99 Mar-00 Jun-00 Oct-00 Jan-01 Apr-01 Jul-01<br />

Date<br />

Figure 2. Particle counts (3 to 5 micrometers in size) at Site 8.<br />

In general, the field design was appropriate <strong>for</strong> this project. Several monitoring wells were intentionally<br />

placed close together to evaluate the aquifer in detail. Although appropriate <strong>for</strong> this study,<br />

the design was costly. Similar designs with several monitoring wells and intensive water-quality<br />

analyses may be prohibitively expensive <strong>for</strong> many utilities. Based on the results of this study,<br />

natural filtration at similar sites could be evaluated using fewer wells. With fewer wells, the wells<br />

can be designed with longer screens to accommodate higher pumping rates <strong>for</strong> analyses, such as<br />

189


190<br />

Microscopic Particulate Analysis. In addition, longer screen lengths may be less susceptible to<br />

local aquifer heterogeneities because water will be contributed from a larger portion of the aquifer.<br />

For other study designs, using data sondes that measure only temperature and specific conductance,<br />

as well as eliminating the inclined wells (depending on the goals of the study), would reduce costs.<br />

The construction of a simple stilling well fitted with a transducer, data sonde, and sampling pump<br />

would be adequate <strong>for</strong> short-term river monitoring.<br />

In general, the monitored parameters showed the most significant drop in concentration from the<br />

river through the riverbed to the inclined wells. Further reductions of concentrations continued<br />

(although not as dramatically) as the water traveled from below the streambed to the production<br />

well. Another significant drop in concentrations occurred from the “C” wells (nearest the<br />

production wells) to the production wells. This is believed to be caused by the dilution effect of<br />

regional groundwater as a result of the large capture zone of the production well.<br />

The high TOC concentrations observed in Well FP8B during the first 2 months of the study<br />

(described earlier) may be due to the growth of indigenous aerobic bacteria after oxygen was<br />

introduced during well installation and development. Once oxygen was consumed, the aerobic<br />

bacteria likely died off and TOC dropped to natural levels. This indicates that this particular well<br />

intercepted a zone of the aquifer with different biological properties than other parts of the aquifer.<br />

The presence of strong odors of hydrogen sulfide, indicative of sulfur-reducing bacteria, and the<br />

excessive corrosion of metal pump fittings are also indicative of biological activity in this well. Had<br />

these high concentrations been due to a preferential flow path, TOC concentrations would not<br />

have exceeded the levels of the river and would not have been expected to decline.<br />

If the high particle counts in Wells FP8B and FP8D were a result of poor filtration, other constituents<br />

also show poor reduction from the river to the well. Aside from the initial high concentrations in<br />

FP8B, TOC concentrations were between those of Wells FP8A and FP8C. Additional parameters<br />

such as UV 254, nitrate, and chloride (not shown) also reflected effective filtration through the<br />

aquifer. This demonstrates that the high particle counts in Well FP8B originated in the aquifer and<br />

were not a result of a preferential flowpath. In the case of FP8D, the particle levels are above the<br />

counts in the river, indicating a secondary source of particles. As in FP8B, other parameters in FP8D<br />

indicated the high particle counts were not a result of preferential flow.<br />

It is widely understood that even a relatively homogeneous aquifer contains localized heterogeneities<br />

in particle size, in situ biology, and water chemistry. These heterogeneities, however, do<br />

not necessarily compromise filtration through the aquifer. Some anomalies can be recognized and<br />

evaluated by incorporating sufficient monitoring points, lengthening well screens to intercept<br />

larger sections of the aquifer, and monitoring <strong>for</strong> sufficient parameters so that anomalies can be<br />

recognized and evaluated.<br />

Acknowledgement<br />

Funding <strong>for</strong> this project was provided by a grant from the Ohio <strong>Water</strong> Development Authority, a<br />

grant from USGS, and funds from the Greater Cincinnati <strong>Water</strong> Works. The authors thank each<br />

of these agencies <strong>for</strong> their contributions and support throughout this project.


REFERENCE<br />

Koterba, M.T, F.D. Wilde, and W.W. Lapham (1995). Ground-water data-collection protocols and procedures <strong>for</strong><br />

the <strong>National</strong> <strong>Water</strong>-Quality Assessment Program: Collection and documentation of water-quality and samples and<br />

related data, U.S. Geological Survey Open File Report 95-399.<br />

BRUCE WHITTEBERRY has been a Professional Hydrogeologist <strong>for</strong> Greater Cincinnati<br />

<strong>Water</strong> Works <strong>for</strong> the past 5 years. In addition to <strong>RBF</strong> research, Whitteberry oversees the<br />

wellhead protection programs <strong>for</strong> two well fields and addresses various groundwater<br />

quantity and quality concerns <strong>for</strong> the Greater Cincinnati <strong>Water</strong> Works. He is one of the<br />

Greater Cincinnati <strong>Water</strong> Works’ representatives <strong>for</strong> the Hamilton to New Baltimore<br />

Ground <strong>Water</strong> Consortium, and he also oversees the Consortium’s regional groundwater<br />

monitoring program, as well as serves in an advisory capacity on groundwater issues. Prior<br />

to joining the Greater Cincinnati <strong>Water</strong> Works, he spent several years in the groundwater industry, both in<br />

consulting and in a regulatory capacity. Whitteberry received a B.S. in Geology from Olivet Nazarene<br />

University and an M.S. in Hydrogeology from Wright State University.<br />

191


192


Session 11: Case Studies “Lessons Learned”<br />

The Hungarian Experience with Riverbank Filtration<br />

Ferenc Laszlo, Ph.D.<br />

<strong>Institute</strong> <strong>for</strong> <strong>Water</strong> Pollution Control<br />

<strong>Water</strong> Resources <strong>Research</strong> Centre<br />

Budapest, Hungary<br />

The water-supply systems of Hungary provide drinking water <strong>for</strong> 9.5-million inhabitants<br />

(95 percent of the population) in 2,100 settlements of the country. About 88 percent of the total<br />

volume supplied originates from groundwater sources. About 30 percent of this (approximately<br />

800,000 m 3 /d) is abstracted from <strong>RBF</strong> water resources.<br />

There is an additional 3-million m 3 /d free <strong>RBF</strong> capacity along the Hungarian section of the<br />

Danube River.<br />

The largest <strong>RBF</strong> system in Hungary provides drinking water <strong>for</strong> 2-million inhabitants in Budapest.<br />

The present capacity is about 900,000 m 3 /d. Approximately 850 <strong>RBF</strong> wells are located on two<br />

islands (Szentendre Island and Csepel Island) of the Danube River. Some parts of this <strong>RBF</strong> system<br />

have operated <strong>for</strong> more than 100 years (Laszlo and Homonnay, 1986). The hydraulic conductivity<br />

of the gravel terraces, the composition of the filtration layer, and the water quality of the river<br />

enables the production of high-quality drinking water.<br />

In recent years, considerable ef<strong>for</strong>ts were focused on investigating and developing methods of<br />

practical application <strong>for</strong> the reliable utilization and protection of bank-filtered water resources.<br />

The activities were mainly directed toward evaluating pollution impacts and exploring the<br />

processes and causes of water-quality changes in <strong>RBF</strong> systems.<br />

Studies in an <strong>RBF</strong> pilot area on Szentendre Island included:<br />

• An extended field investigation to provide adequate data <strong>for</strong> a comprehensive<br />

water-quality evaluation and <strong>for</strong> modeling activities.<br />

• Adsorption experiments in the laboratory.<br />

• Application and development of numerical models to assess and predict water-quality<br />

changes in the <strong>RBF</strong> system that are induced by various potential changes in the<br />

conditions of the river and filtration media.<br />

Numerical modeling was applied to:<br />

• Calibrate hydraulic and transport parameters.<br />

• Calculate the transport of various pollutants from the Danube River to the wells.<br />

• Assess quality changes in the water of the production wells.<br />

• Predict the consequences of accidental pollution in the Danube River.<br />

Correspondence should be addressed to:<br />

Ferenc Laszlo, Ph.D.<br />

Director, <strong>Institute</strong> <strong>for</strong> <strong>Water</strong> Pollution Control<br />

<strong>Water</strong> Resources <strong>Research</strong> Centre<br />

H-1095 Budapest • Kvassay ut 1., Hungary<br />

Phone: +36 1 215 9045 • Fax: +36 1 216 8140 • Email: laszloferenc@vituki.hu<br />

193


194<br />

During the field investigation program, hydrogeological, hydraulic, hydrological, chemical,<br />

physico-chemical, biological, and microbiological data were collected and evaluated. The results of<br />

the water-quality studies revealed characteristic quality changes in the filter media: the chemical<br />

oxygen demand concentration gradually decreased from the river towards abstraction sites,<br />

indicating the decomposition of organic pollutants. Hydro-biological studies indicated that the<br />

Danube River had considerable fecal coli<strong>for</strong>m pollution, but that bacteria removal during <strong>RBF</strong> was<br />

very efficient. Coliphages used as indicators of virus pollution were also measured. The results of<br />

these tests verified that coliphages were not able to pass through filter media.<br />

The retardation of specific organic micropollutants is very different depending on their biodegradability,<br />

sorption affinity, etc. The mobile, persistent organic micropollutants (e.g., atrazine,<br />

simazine, trichloroethene, and benzene) have low removal efficiency (less than 30 percent); on<br />

the other hand, some pesticides (e g., terbutrin, 2,4-D, and carbaryl) and petroleum hydrocarbons<br />

are removed substantially. The retardation or remobilization of some heavy metals is influenced<br />

by redox conditions.<br />

Adsorption experiments in the laboratory provided in<strong>for</strong>mation on the partition coefficients of<br />

various micropollutants between spiked Danube water and the material of filtration medium taken<br />

from <strong>RBF</strong> sites (Laszlo and Literathy, 2002). Two riverbed materials were used: one represented the<br />

sandy alluvium having low organic material; the other was taken from a sedimentation zone of the<br />

riverbed where river water infiltrates through sediment having high organic material content and fine<br />

particle size. The partition coefficient varied widely <strong>for</strong> the different organic micropollutants and<br />

heavy metals, depending mainly on the character of the pollutants. The type of bed material was also<br />

important: the partition coefficients were higher in silty riverbed material than in the sandy matrix<br />

<strong>for</strong> all pollutants.<br />

Numerical transport modeling — based on measured hydraulic, hydrogeological, and physicochemical<br />

parameters — concluded that relatively short-duration accidental pollution has limited<br />

risk <strong>for</strong> <strong>RBF</strong> schemes along the Danube River due to a wide range of travel times from the river to<br />

abstraction wells. Higher risk would be long-lasting, continuous pollution of the Danube.<br />

Another concern is the impacts of river training and gravel dredging on the quality of riverbankfiltered<br />

water (Laszlo et al., 1990). Training structures (groynes, parallel dykes) and dredging<br />

operations affect the hydraulic conditions in the river that are conducive to silting in areas with<br />

reduced flow velocities. Adverse hydrochemical changes occur in the silted filter layer, especially<br />

the dissolution of iron and manganese, and higher concentrations of ammonium are observable.<br />

Bed degradation owing to dredging causes changes in the inflow ratio to the abstraction wells,<br />

increasing the proportion of polluted groundwater from the background areas in the wells.<br />

Conclusions<br />

<strong>RBF</strong> has been an effective method <strong>for</strong> drinking-water supply in Hungary <strong>for</strong> a long time.<br />

The sustainability of <strong>RBF</strong> abstraction schemes depends on several factors, including the quality of<br />

river water, characteristics of the filtration media, retention time within the aquifer, and quality<br />

of the groundwater adjacent to the extraction site.


REFERENCES<br />

Laszlo , F., and Z. Homonnay (1986). “Study of effects determining the quality of bank-filtered well water.”<br />

Conjunctive <strong>Water</strong> Use, IAHS Publ. No. 156, 181-188.<br />

Laszlo, F., Z. Homonnay, and M. Zimonyi (1990). “Impacts of river training on the quality of bank-filtered<br />

waters.” Wat. Sci. Tech., 22 (5): 167-172.<br />

Laszlo, F., and P. Literathy (2002). “Laboratory and field studies of pollutant removal.” Riverbank filtration:<br />

Understanding contaminant biogeochemistry and pathogen removal, C. Ray, ed., Kluwer Academic Publishers,<br />

Dordrecht.<br />

FERENC LASZLO is Director of the <strong>Institute</strong> <strong>for</strong> <strong>Water</strong> Pollution Control of the <strong>Water</strong><br />

Resources <strong>Research</strong> Centre in Hungary. A chemical engineer, Laszlo has 30 years of<br />

experience in research related to water pollution and aquatic chemistry. His research<br />

interests include the protection of riverbank-filtration systems, investigation of<br />

water-quality changes in different water resources, study of micropollutants, simulation of<br />

the fate and transport of contaminants in the aquatic environment, development and<br />

operation of water-quality monitoring system, and accident emergency warning systems.<br />

In addition, he has provided short-term consulting and expert services to the World Health Organization and<br />

United Nations Development Program. Laszlo received an M.S. in Chemical Engineering and a University<br />

Doctorate Degree in Aquatic Chemistry from Veszprem University in Hungary.<br />

195


196


Session 11: Case Studies “Lessons Learned”<br />

Nitrate Pollution of a <strong>Water</strong> Resource – 15 N and 18 O<br />

Study of Infiltrated Surface <strong>Water</strong><br />

Frantisek Buzek, Ph.D.<br />

Czech Geological Survey<br />

Prague, Czech Republic<br />

Renata Kadlecova<br />

Czech Geological Survey<br />

Prague, Czech Republic<br />

Miroslav Knezek, Ph.D.<br />

Prague, Czech Republic<br />

This study was undertaken to determine the effects of agricultural and human activities at the<br />

village of Karany near Prague, the Czech Republic, on the quality of water resources. Three groups of<br />

wells use bank infiltration water from the Jizera River to produce about 1,000 liters per second of<br />

quality drinking water <strong>for</strong> Prague. Some wells have exhibited a steady increase in nitrate content<br />

in recent years (Figure 1), although river-water quality remains good. A question arises of the<br />

origin of nitrate contamination and its transportation to wells, and to what extent bank-filtered<br />

water contributes to produced water. Also of interest are other contributions and how long it takes<br />

<strong>for</strong> these contributions to reach the wells.<br />

Figure 1. Nitrate content (in milligrams per liter, 6-months average value) of contaminated capture wells.<br />

Correspondence should be addressed to:<br />

Frantisek Buzek, Ph.D.<br />

Head of Stable Isotope Lab<br />

Czech Geological Survey<br />

Geologicka 6 • 152 00 Prague 5, Czech Republic<br />

Phone: 420 251085345 • Fax: 420 251818748 • Email: buzek@cgu.cz<br />

197


198<br />

Site Description<br />

The study area is <strong>for</strong>med by Cretaceous sandstones and marlstones (the Jizera <strong>for</strong>mation) of Turonian<br />

age. The Jizera River represents a base <strong>for</strong> local drainage. A shallow aquifer is situated in Late<br />

Pleistocene and Holocene alluvial deposits. They consist of sands and sandy gravel 7- to 9.5-m thick.<br />

The Quaternary aquifer is 2.5- to 5-m thick. It is hydraulically connected with the Jizera River<br />

stream. The Quaternary sediments have a high degree of transmissivity and high permeability. The<br />

Cretaceous deposits have an average transmissivity and free groundwater surface.<br />

On the right bank of the Jizera River, where the Turonian aquifer has a sandy development, its<br />

transmissivity is only about 1 order inferior to that of the alluvial plain deposits. On the left bank<br />

of the Jizera River, the permeability is lower due to a higher proportion of marly component in the<br />

Turonian sediments. Contamined water Wells 227 to 299 (Skorkov capture wells) are located<br />

along the Jizera River at a distance of 200 to 300 m from the right bank (Wells 227 to 270) and<br />

left bank (Wells 271 to 299) (Figure 2). Sojovice capture wells are located on the left bank only;<br />

one part is recharged by artificial infiltration (Wells 120 to 187), another part is traditional bank<br />

infiltration at a distance 300 m from the river (Wells 188 to 226). The river terraces are used by<br />

local farmers <strong>for</strong> the cultivation of corn, potatoes, and maize. Local villages have poor sewer<br />

systems, and there are several landfills located in the area.<br />

Methods<br />

From 1999 to 2000, we took water samples at various time intervals from the river, precipitation<br />

in the recharge area, precipitation at the site, and wells <strong>for</strong> the oxygen isotope composition of<br />

water (δ 18 O - H 2O), and river and wells <strong>for</strong> the isotope composition of nitrogen in nitrate<br />

(δ 15 N-NO 3). 1<br />

δ 18 O data were used to:<br />

• Describe the river system.<br />

• Specify groundwater recharge.<br />

• Calculate the contribution of infiltrated river water to well production.<br />

• Model infiltration water flow and transportation time to the well.<br />

• Specify additional water contributing to well water.<br />

δ 15 N-NO 3 data were used to:<br />

• Specify the origin of nitrate and possible mixtures.<br />

• Trace seasonal trends in nitrate content.<br />

• Follow reactions in the groundwater system (denitrification).<br />

River System<br />

The mean residence time T equals 7.2 months, as calculated from variations of δ 18 O values of the<br />

river and precipitation in the recharge area. An average contribution of direct precipitation to<br />

runoff is about 13 percent, with a maximum of 36 percent during the highest flood event.<br />

1 Isotopic composition is measured in delta units, δ(in ‰) = Rx /R s -1) × 1000, where R denotes the ratio of the heavyto-light<br />

isotope (e.g., 18 O/ 16 O or 15 N/ 14 N), and R x and R s are the ratios in the sample and standard, respectfully. As<br />

international reference standard is used, Standard Mean Ocean <strong>Water</strong> (SMOW) stands <strong>for</strong> water and atmospheric N 2<br />

<strong>for</strong> nitrogen.


Figure 2. A sketch of the study area.<br />

Key: 1 = Forested area.<br />

2 = Jizera-Elbe watershed.<br />

3 = Capture well.<br />

4 = Groundwater flow.<br />

5 = Deep well.<br />

6 = Landfill.<br />

Nitrate content and δ 15 N-NO 3 of the Jizera River follow seasonal variations. δ 15 N-NO 3 varies from<br />

8 ‰ in the winter to about 1.5 ‰ (base flow only). The highest nitrate content (about 25 mg/L)<br />

was measured during snowmelt. Base flow nitrate content is very low (about 8 to 10 mg/L).<br />

Additional nitrate to base flow originates mostly from drained precipitation events (short time or<br />

fast component of runoff), flushing fertilizers bringing mostly organic nitrogen in the spring<br />

(manure applied in the winter period), and inorganic nitrogen during the growing season<br />

(fertilizers applied on leaves).<br />

199


200<br />

The nitrate content and δ 15 N-NO 3 of the Jizera River do not follow the same seasonal variation.<br />

δ 15 N-NO 3 varies from 8 ‰ in winter to about 1.5 ‰ (base flow only). The highest nitrate content<br />

(about 25 mg/L) was measured during snowmelt. Base flow nitrate content is very low (about 8 to<br />

10 mg/L). Figure 3 shows the nitrate content and isotopic composition of the Jizera River.<br />

Figure 3. Nitrate content and isotopic composition of the Jizera River.<br />

As the response of river runoff to precipitation is quite fast (from 3 to 5 days after a storm), a<br />

simplified two-component model of runoff generation is applicable. The discharged water consists<br />

of groundwater and short residence-time components. Short time components are <strong>for</strong>med by<br />

drained precipitation events flushing fertilizers — predominantly the organic nitrogen applied in<br />

winter — and by inorganic nitrogen applied during the growing season. Increasing δ 15 N-NO 3<br />

levels during the autumn corresponds to residual nitrate levels from the unsaturated soil zone<br />

below root level. A similar seasonal variability was observed during the next year as well.<br />

Infiltration System<br />

<strong>Water</strong> Balance<br />

The actual and modeled δ 18 O data were compared during a flood event and steady-state<br />

conditions. At flood conditions, the contribution of infiltrated river water increased up to<br />

100 percent, with a transit time of only about 5 days. At steady-state conditions, infiltrated river<br />

water <strong>for</strong>ms about 60 percent of infiltrated water, with a transit time of 28 days. The additional<br />

40 percent of infiltrated water cannot be explained by local groundwater only. It must originate<br />

from a shallow aquifer. Fast transit times of this recharging water were confirmed by a variability<br />

of δ 18 O in the wells, which is higher than in the river. Additional recharging water can be<br />

modeled by local precipitation input, with a delay in the order of weeks.<br />

Nitrate Contamination of Skorkov Wells<br />

From comparing the high nitrate content of contaminated wells and their seasonal δ 15 N-NO 3<br />

variability, it is obvious that:<br />

• Nitrate originates mostly from water other than infiltrated water.<br />

• Sources of nitrate are very similar to river sources (i.e., similar sequences of nitrogen<br />

inputs during the year.<br />

Wells have a higher concentration of nitrates than river water.


By analyzing the δ 15 N-NO 3 of single wells along the contaminated part of bank infiltration, we<br />

identified the possible sources of nitrate. Besides the seasonal application of fertilizers and manure,<br />

village sewer systems and landfills in the recharge area also contribute to water pollution in the<br />

wells (Figure 4).<br />

Figure 4. Nitrate content and isotopic composition of contaminated water (Skorkov single wells).<br />

The importance of local precipitation <strong>for</strong> recharging wells is obvious from the long-time variation<br />

of nitrate content of contaminated wells (see Figure 1). Minimal values on an otherwise<br />

continuously increasing plot correspond to years with very low precipitation input (25-percent less<br />

than average). With low precipitation input, wells are recharged preferentially by deeper<br />

groundwater with a lower nitrate content. Within a high precipitation period, wells are recharged<br />

by fast infiltrated local precipitation washing out nitrate in the unsaturated zone.<br />

Nitrate Contamination of Sojovice Wells<br />

Sojovice capture wells have two parts: one part is recharged by artificial infiltration (Wells 120 to<br />

187), another part is traditional bank filtration at a distance of 300 m from the river (Wells 188<br />

to 126). Contrary to the Skorkov system, the Sojovice wells are not affected directly by<br />

precipitation (Figure 5). Besides, both the bank infiltration and artificial infiltration parts are<br />

partially recharged by denitrified groundwater. From a time sequence of nitrate content and its<br />

δ 15 N-NO 3 on single wells, we could separate three components of the recharge:<br />

• River water.<br />

• Shallow groundwater with a nitrate source typical <strong>for</strong> infiltration in the left bank.<br />

• Deep groundwater with a reduction zone and denitrification.<br />

Flow paths of groundwater components are different — perpendicular to river flow (shallow<br />

component) and along the river (deep component).<br />

Conclusions<br />

Nitrate contamination of the bank infiltration systems originates in local sources of nitrate,<br />

including inappropriate uses of manure/fertilizers and leaking village sewers. A considerable<br />

201


202<br />

Figure 5. Nitrate isotopic composition of contaminated water (Sojovice single wells).<br />

increase of nitrate in the 1980s originated both because of the change in agriculture practices (due<br />

to an increased proportion of maize cultivation) and the change in hydrology dynamics of the<br />

system (due to the extensive use of water resources, new infiltration pathways were developed that<br />

recharge more shallower groundwater than be<strong>for</strong>e).<br />

FRANK BUZEK has worked <strong>for</strong> the Czech Geological Survey <strong>for</strong> about 25 years, and is<br />

Head of the Stable Isotope Group. His primary research interest includes the application<br />

of isotope geochemistry to environmental problems, hydrology, and organic geochemistry.<br />

As an all-round geochemist, he was responsible <strong>for</strong> stable isotope data on carbon and<br />

nitrogen cycling in <strong>for</strong>est soils (European Union international projects CANIF, NIPHYS,<br />

and FORCAST) and tracing atmospheric emissions and groundwater pollution<br />

(International Atomic Energy Agency project on isotope technique in groundwater<br />

pollution). Most of the recent case studies solved by Buzek and his team have dealt with nitrate in water<br />

resources. Buzek received an M.S. in Physical Chemistry from the <strong>Institute</strong> of Chemical Technology and a<br />

Ph.D. in Applied Surface Chemistry from the <strong>Institute</strong> of Chemical Processes Fundamentals at the Czech<br />

Academy of Sciences in Prague.


Session 11: Case Studies “Lessons Learned”<br />

Microbial Growth in Artificially Recharged Groundwater:<br />

Experiences from a 4-Year Project<br />

Ilkka T. Miettinen, Ph.D.<br />

<strong>National</strong> Public Health <strong>Institute</strong><br />

Kuopio, Finland<br />

Markku J. Lehtola, Ph.D.<br />

<strong>National</strong> Public Health <strong>Institute</strong><br />

Kuopio, Finland<br />

Prof. Terttu Vartiainen<br />

<strong>National</strong> Public Health <strong>Institute</strong><br />

Kuopio, Finland<br />

Prof. Pertti J. Martikainen<br />

Kuopio University<br />

Kuopio, Finland<br />

Introduction<br />

There are regions in Finland where large groundwater aquifers are not available and where the<br />

artificial groundwater recharge technique is an important alternative <strong>for</strong> drinking-water production.<br />

In Finland, 9 percent of drinking water is produced using the artificial recharge of groundwater or<br />

bank-filtration techniques. The production of artificially recharged groundwater has increased<br />

during the last decade. The high content of organic carbon (humus) present in surface water is a<br />

serious problem when this water is used to produce drinking water.<br />

Organic carbon can react with chlorine during disinfection, <strong>for</strong>ming halogenated byproducts. The<br />

high availability of organic carbon may also cause microbial problems in distribution networks.<br />

The fraction of organic carbon that microbes can use <strong>for</strong> growth (i.e., AOC) is considered to be<br />

the most important nutrient affecting microbial growth in drinking waters (van der Kooij, 1992).<br />

In boreal areas, phosphorus — in addition to organic carbon — has been shown to regulate<br />

microbial growth in drinking waters (Miettinen et al., 1996).<br />

Objectives<br />

A 4-year project examined how the artificial recharge of lake water affects the following:<br />

• Molecular weight distribution of organic matter.<br />

• AOC content.<br />

• Microbially available phosphorus (MAP) content.<br />

• Microbial growth potential in water.<br />

Correspondence should be addressed to:<br />

Ilkka Miettinen, Ph.D.<br />

<strong>Research</strong>er, Laboratory of Environmental Microbiology<br />

<strong>National</strong> Public Health <strong>Institute</strong><br />

Department of Environmental Health • P.O. Box 95 • FIN-70701 Kuopio Finland<br />

Phone: (358) 17-201371 • Fax: (358) 17-201155 • Email: Ilkka.Miettinen@ktl.fi<br />

203


204<br />

Methodology<br />

Experimental Sites<br />

Changes in water quality during artificial groundwater recharge were studied in five water works.<br />

All water works (<strong>Water</strong> Works A, B, C, D, and E) are located in esker areas, where the soil consists<br />

mainly of sand and gravel. Lake water is filtrated into the ground by basin infiltration in <strong>Water</strong><br />

Works A, C, D, and E. <strong>Water</strong> Works A, B, and E use sprinkling infiltration through <strong>for</strong>est soil,<br />

while <strong>Water</strong> Works B and C use the same lake as their raw-water source.<br />

<strong>Water</strong> Analyses<br />

The quality of raw-water samples and water samples from different monitoring wells at different<br />

filtration distances (raw water, the first monitoring point after infiltration, and the last monitoring<br />

point after filtration) were analyzed from 1998 to 2001. Filtration distance varied from 10 to 280 m in<br />

the first monitoring point and from 390 to 1,200 m in the last monitoring point (artificially recharged<br />

groundwater). The quality of organic carbon was analyzed in terms of its molecular weight distribution<br />

using the high-pressure size-exclusion chromatography (HPSEC) method (Vartiainen et al., 1987).<br />

AOC was measured with a modified bioassay (Miettinen et al., 1999) originally presented by van der<br />

Kooij et al. (1982), and MAP was analyzed by bioassay (Lehtola et al., 1999). Microbial growth<br />

potential (heterotrophic growth potential) was tested using a method by Miettinen et al. (1997).<br />

Results and Discussion<br />

Organic matter in lake waters consisted mainly of high and intermediate molecular-size fractions. These<br />

fractions decreased rapidly in the filtration process. In contrast, the lowest molecular weight fractions<br />

of organic matter decreased only slightly during filtration. The strongest reduction in the content of<br />

AOC occurred in the beginning of filtration. Later on, there was only a slight decrease in AOC<br />

content. On average, there was a 53-percent reduction in the concentration of AOC during the<br />

artificial recharge of groundwater (Table 1). Changes in MAP concentrations during filtration varied<br />

among water works. MAP content decreased strongly (89- to 99.9-percent removal) during filtration<br />

in <strong>Water</strong> Works A, B, and D. <strong>Water</strong> Works E was the only exception where MAP content increased<br />

during the filtration process, indicating that phosphorus was dissolving from soil into water (see Table 1).<br />

Table 1. Concentrations (Range and Mean Value in Parentheses) of AOC and MAP<br />

in Raw <strong>Water</strong> and Artificially Recharged Groundwater. Number of Observations: 4 to 11.<br />

<strong>Water</strong> AOC: µg Acetate eq. C/L MAP: µg MAP-P/L<br />

Works RW ARW RW ARW<br />

A 54 to 366 (134) 17 to 132 (46) 0.98 to 9.90 (3.84) 0.22 to 0.34 (0.27)<br />

B 15 to 120 (56) 15 to 73 (46) 1.79 to 5.31 (2.57) 0.15 to 0.37 (0.28)<br />

C 39 to 144 (78) 27 to 90 (49) Not Determined Not Determined<br />

D 27 to 288 (151)


Microbial growth was weak in all lake-water samples, despite the fact that these waters had high<br />

AOC and MAP contents; however, there was strong microbial growth in artificially recharged<br />

groundwater. Lower microbial growth in surface water could be associated with the grazing activity<br />

of protozoa (Hahn and Hofle, 2001). The microbial growth potential in groundwater was strongest<br />

in the first sampling point and weakened during filtration (Figure 1). The AOC content did not<br />

correlate with microbial growth potential; however, the maximum microbial counts correlated<br />

with the MAP content (r = 0.70, p = 0.000, n = 26) when raw-water samples and samples with<br />

MAP over 5 µg/L (no phosphorus limitation) were excluded from the data.<br />

Figure 1. Increase in microbial cell numbers (R2A plate counts) with time when water samples were<br />

incubated in the laboratory. The figure shows mean numbers in samples taken from raw water or<br />

artificially recharged groundwater of <strong>Water</strong> Works A, B, D, and E. Symbols: RW = Raw water.<br />

1.point = The first monitoring point after infiltration. 2. Point = The last monitoring point after<br />

filtration). Number of observations: 18 to 21.<br />

Conclusions<br />

• The quality of organic matter changed during artificial recharge: high and intermediate<br />

molecular weight organic compounds were removed from infiltrated lake water.<br />

• The decrease in AOC content was strongest in the beginning of infiltration. An increase<br />

in filtration distance had a minor effect on AOC removal.<br />

• MAP removal depended on soil characteristics. In most of the studied water works,<br />

removal was good.<br />

• Microbial growth was higher in artificially recharged groundwater than in raw water<br />

(surface water). The microbial growth potential decreased during filtration.<br />

• Because the removal efficiency <strong>for</strong> phosphorus was higher than that <strong>for</strong> carbon, phosphorus<br />

became the main nutrient regulating microbial growth in artificially recharged<br />

groundwater.<br />

205


206<br />

REFERENCES<br />

Hahn, M.W., and M.G. Höfle (2001). “Grazing of protozoa and its effect on populations of aquatic bacteria.”<br />

FEMS Microbiol. Ecol., 35: 113-121.<br />

Lehtola, M., I.T. Miettinen, T. Vartiainen, and P.J. Martikainen (1999). “A new sensitive bioassay <strong>for</strong><br />

determination of microbially available phosphorus in water.” Appl. Environ. Microbiol., 65(5): 2,032-2,034.<br />

Miettinen, I.T., T. Vartiainen, and P.J. Martikainen (1996). “Contamination of drinking water.” Nature,<br />

381: 654-655.<br />

Miettinen, I.T., T. Vartiainen, and P.J. Martikainen (1997). “Microbial growth and assimilable organic<br />

carbon in Finnish drinking waters.” Wat. Sci. Techn., 35(11/12): 301-306<br />

Miettinen, I.T., T. Vartiainen, and P.J. Martikainen (1999). “Determination of assimilable organic carbon<br />

(AOC) in humus-rich waters.” Wat. Res., 3(10): 277-2,282.<br />

Vartiainen, T., A. Liimatainen, and P. Kauranen (1987). “The use of size exclusion columns in determination<br />

of the quality and quantity of humus in raw waters and drinking waters.” Sci. Total Environ., 62: 75-84.<br />

van der Kooij, D., W.A.M. Hijnen, and A. Visser (1982). “Determinating the concentration of easily<br />

assimilable organic carbon.” Journal AWWA, 74: 540-545.<br />

van der Kooij, D. (1992). “Assimilable organic carbon as an Indicator of Bacterial Regrowth.” Journal AWWA,<br />

84(2): 57-66.<br />

ILKKA MIETTINEN has been a researcher <strong>for</strong> the <strong>National</strong> Public Health <strong>Institute</strong> in<br />

Finland since 1987. Currently, he serves as <strong>Research</strong>er of the Laboratory of Environmental<br />

Microbiology at the <strong>Institute</strong>, as well as <strong>Research</strong>er <strong>for</strong> Kuopio University. For the past<br />

3 years, his research has focused on the role of phosphorous in microbial growth potential<br />

and biofilm <strong>for</strong>mation in distribution networks. A recent project of his is called,<br />

“Surveillance and control of microbiological stability in drinking-water distribution<br />

networks.” Past research has focused on chemical and microbiological quality in humusrich<br />

lake water during bank filtration, as well as the effects of strong oxidants on the quantity and quality of<br />

organic matter and microbial regrowth in treated waters. Miettinen received a B.S. in Biochemistry, M.S. in<br />

Biotechnology, and Ph.D. in Environmental Sciences at the University of Kuopio, where he currently is<br />

Docent of Environmental Sciences.


Session 11: Case Studies “Lessons Learned”<br />

Evaluation of the Existing Per<strong>for</strong>mance of Infiltration<br />

Galleries in Alluvial Deposits of the Parapeti River<br />

Dip.-Eng. Alvaro Camacho<br />

Bolivian Association of Sanitary Engineers<br />

La Paz, Bolivia<br />

Introduction<br />

The Parapeti River is part of the drainage system of the Parapeti-Izozog watershed, which covers<br />

about 52,000 square kilometers and is the sole water resource <strong>for</strong> supplying drinking water to the<br />

City of Camiri, Bolivia. The strata of the basin consist of Quaternary deposits and the permeability<br />

is good; however, precipitation is low (only 700 to 800 millimeters) and, as this zone is located in<br />

the watershed between the Amazon and La Plata river systems, existing groundwater quantity is<br />

very low. The aquifer depth is between 100 to 200 m in the vicinity of Camiri (yield capacity is<br />

less than 1.0 liters per second). The City is located in the region of Santa Cruz, Bolivia, at an<br />

altitude of 810-m above sea level, in the southeast, and has a population of 25,000 inhabitants.<br />

The average temperature is 25-degrees Celsius, with a maximum of 38-degrees Celsius in the<br />

summer and a minimum of 11-degrees Celsius in the winter.<br />

The City’s water demand is supplied by five infiltration galleries buried below the riverbed<br />

(Figure 1), on the bedrock, at a depth of about 4 to 5 m. The galleries are interconnected through<br />

Gallery<br />

1<br />

Pumping<br />

Station<br />

Collector<br />

Well<br />

Parapeti River<br />

Gallery<br />

2 Gallery<br />

5<br />

Figure 1. Location of the study area.<br />

Correspondence should be addressed to:<br />

Dip.-Eng. Alvaro Camacho<br />

Consultant Engineer<br />

Bolivian Association of Sanitary Engineers<br />

Casilla 9348 • La Paz, Bolivia<br />

Phone and Fax: (591-2) 241- 6283 • Email: alcamachog@unete.com<br />

Study<br />

Area<br />

Camiri<br />

Location of the Infiltration Galleries<br />

Collector<br />

Well<br />

Collector<br />

Well<br />

Gallery<br />

3 Gallery<br />

4<br />

Collector<br />

Well<br />

Collector<br />

Well<br />

207


208<br />

a network and linked to a collector well sited on the riverbank. From there, the collected water is<br />

pumped to a disinfection tank prior to delivery to customers.<br />

The infiltration galleries are channels with a 1.0-m × 1.0-m section or drains of 12-inch covered<br />

by four layers (Table 1) of a filter medium (Figure 2)<br />

4 to 5 m<br />

Permeable<br />

Alluvial<br />

Deposits<br />

Figure 2. Infiltration gallery system.<br />

Table 1. Artificial Filterbed Layers of Infiltration Galleries<br />

Bed Layers Type/Size Height<br />

First Layer (Top) Coarse Sand 1.75 m<br />

Second Layer Gravel, 1 to 2 inches 1.0 m<br />

Third Layer Gravel, 2 to 3 inches 1.0 m<br />

Fourth Layer (Bottom) Small Stone, 3 to 6 inches 1.0 m<br />

Natural Sand River<br />

Section 1 m × 1 m<br />

<strong>Water</strong> Surface<br />

Gravel 1 to 2 inches<br />

Gravel 2 to 3 inches<br />

Gravel 3 to 6 inches<br />

Bed Rock<br />

River Bed<br />

In this area of the Parapeti River, where the galleries are in operation, the riverbed consists of<br />

permeable alluvial material, sand, and gravel of about 4- to 5-m deep, which creates favorable<br />

conditions <strong>for</strong> groundwater flow through the unconfined aquifer (Huisman, 1978). Because the<br />

movement of groundwater develops small velocities in granular material, the flow is laminar (Harr,<br />

1990) and this is the driving <strong>for</strong>ce <strong>for</strong> the sedimentation process in the sand and gravel layers of<br />

the river. The Parapeti is a mountain river whose water flow is subject to dramatic variations, both<br />

in quality and quantity, during the dry season (May to October) and rainy season (November to<br />

April). With a relatively steep slope (0.5 to 1 percent), intensive sediment transport rates occur<br />

in the rainy season (the maximum flow exceeds 1,000 m 3 /s, as compared to 5 m 3 /s in the dry<br />

season). Regarding water quality in the river, turbidity levels in the rainy season have been reported<br />

at about 15,000 ntu and suspended solid concentrations at 30,000 mg/L (Binnie & Partners<br />

Consultants, 1982).<br />

In spite of Parapeti’s water quality, infiltration galleries have worked continuously <strong>for</strong> more than<br />

15 years, producing water of good quality and using disinfection as the only treatment process,<br />

with a discharge of 48 to 50 liters per second. The only annual O&M consists of removing the<br />

finest particles of sand that have reached the bottom of the galleries.


Objectives<br />

The aim of the study was to evaluate the existing per<strong>for</strong>mance of infiltration galleries to improve<br />

on the water quality of surface waters. The study was oriented to answer the following question:<br />

“To what extent do infiltration galleries work to remove contaminants from surface-water supplies?”<br />

Materials and Methods<br />

A sampling program was implemented during the dry and rainy seasons. Two complete sets of water<br />

samples were collected and analyzed between 2002 and 2003 (Schubert, 2002). The first set of samples<br />

was taken in the dry season, from early October to November 2002, when suspended-solid<br />

concentrations in the river are low. The second set of samples was collected during the rainy<br />

season, from February and April, when the river carries higher suspended-solid concentrations. In<br />

both periods, water samples were taken over a period of 41 to 45 days. Grab samples were collected<br />

three times per day in the Parapeti River and in two of the five infiltration galleries (Numbers 1<br />

and 4). Raw-water samples collected in the river were taken from a depth of 0.30-m below the<br />

surface and 100-m upstream of the galleries. At the same time, water samples were taken from the<br />

effluents of the two infiltration galleries (Numbers 1 and 4) at the inlet point of the disinfection tank.<br />

Eight water-quality parameters were chosen: temperature, color, turbidity, suspended solids, total<br />

dissolved solids, pH, conductivity, and fecal coli<strong>for</strong>m counts. Grab samples <strong>for</strong> temperature and pH<br />

were collected in glass bottles and analyzed on site. Random samples <strong>for</strong> color, turbidity, suspended<br />

solids, total dissolved solids, and conductivity were collected in glass bottles <strong>for</strong> analysis in a<br />

laboratory. Microbiological samples were collected in sterile 50-mL Pyrex glass <strong>for</strong> analysis in a<br />

laboratory within 24 hours. The test was conducted using a portable OXFAM-DEL AGUA water<br />

test kit (membrane filtration technique). A duplicate was prepared from each sample. All the<br />

experimental tests were based on the “Standard Methods <strong>for</strong> the Examination of <strong>Water</strong> and<br />

Wastewater” (1975).<br />

In addition, exploration holes were dug into different layers of the coarse material to take samples from<br />

different filter media that cover the galleries. Each layer of the filter media, from Galleries 3, 4, and 5,<br />

was washed with distilled water to per<strong>for</strong>m a column-settling test. This experiment was conducted to<br />

determine the frequency distribution of settling velocities and particle diameters. Also, each sample<br />

was subject to physical analysis, such as mass density, porosity, and granular grain-size distribution.<br />

Results and Discussion<br />

Figures 3 to 5 illustrate seasonal variations in water quality in the Parapetí River and infiltration<br />

galleries. For the purposes of this paper, only two indicators of water quality were chosen: suspended<br />

solids and fecal coli<strong>for</strong>m.<br />

During the rainy season, raw water in the Parapeti River had suspended-solid levels ranging from<br />

45 to 4,850 mg/L, with a mean value of 618 mg/L. These figures differ substantially in comparison<br />

to the dry season, where the average value was 72 mg/L and the maximum was 410 mg/L. In the<br />

rainy season, because the river’s flow is high (from 5 to 1,000 m 3 /s, with a mean value of 30 m 3 /s),<br />

the elevated suspended-solid concentration is caused by erosion that occurs on the steep slopes of<br />

watershed highlands. As shown in Figures 3 to 8, the infiltration galleries that were evaluated<br />

during the same period had very low levels. The values were ranged from 0 to 7 mg/L, with an average<br />

of 2 mg/L far below the common standards. From this evidence, it has been demonstrated that the<br />

per<strong>for</strong>mance of the infiltration galleries reached a removal ratio of more than 90 percent. Turbidity<br />

of 5 ntu or less in 98 percent of the daily samples was reported in the effluents (see Figure 5).<br />

209


210<br />

6045<br />

5045<br />

4045<br />

3045<br />

2045<br />

1045<br />

45<br />

8<br />

7<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

River<br />

Gal. 1<br />

Gal. 4<br />

Suspended Solids (mg/L)<br />

Feb-03 Feb-03 Feb-03 Mar-03 Mar-03 Mar-03 Mar-03 Mar-03 Mar-03 Apr-03 Apr-03<br />

Figure 3. Raw water and infiltration gallery water quality.<br />

70150<br />

60150<br />

50150<br />

40150<br />

30150<br />

20150<br />

10150<br />

150<br />

200<br />

150<br />

100<br />

50<br />

0<br />

River<br />

Gal. 1<br />

Gal. 4<br />

Fecal Coli<strong>for</strong>ms (FC/100 mL)<br />

Feb-03 Feb-03 Feb-03 Mar-03 Mar-03 Mar-03 Mar-03 Mar-03 Apr-03<br />

Figure 4. Raw water and infiltration gallery water quality.<br />

% Cummulative Distribution<br />

120%<br />

100%<br />

80%<br />

60%<br />

40%<br />

20%<br />

Faecal Coli<strong>for</strong>ms in Galleries. Frequency Distribution.<br />

Rainy Season (2003)<br />

0%<br />

0 20 40 60 80 100 120 140 160 180<br />

NTU<br />

Figure 5. Quality of water produced in the infiltration galleries.<br />

Gallery 1<br />

Galeria 4


6000<br />

4000<br />

2000<br />

Suspended Solids (mg/L)<br />

0<br />

Minimum Mean Maximum<br />

Gallery 1 0 1 5<br />

Gallery 4 0 2 7<br />

River 45 618 4850<br />

Figure 6. A comparison of raw water and water produced by the galleries.<br />

80000<br />

60000<br />

40000<br />

20000<br />

FE Faecal RMS (FC/100 mL)<br />

0<br />

Minimum Mean Maximum<br />

Gallery 1 1 0 20<br />

Gallery 4 1 32 169<br />

River 160 8958 60685<br />

Figure 7. A comparison of raw water and water produced by the galleries.<br />

Microbiological water-quality parameters in the Parapeti River show poor quality with a wide range<br />

of 160 to 60,000 fecal coli<strong>for</strong>m counts/100 milliliters, with an average value of 9,000 fecal coli<strong>for</strong>m<br />

counts/100 milliliters. The infiltration gallery data illustrate that in one gallery (Gallery 1), the<br />

mean value was zero fecal coli<strong>for</strong>m counts/100 milliliters, with a maximum of 20 fecal coli<strong>for</strong>m<br />

counts/100 milliliters achieving a removal ratio of more than 90 percent. Three fecal coli<strong>for</strong>m<br />

counts /100 milliliters or less in 98 percent of the samples was found. In the other gallery (Gallery 4),<br />

the mean value was a concentration of 32 fecal coli<strong>for</strong>m counts/100 milliliters, with a maximum<br />

of 169 fecal coli<strong>for</strong>m/100 milliliters representing a removal ratio of more than 90 percent. In this<br />

gallery, 120 fecal coli<strong>for</strong>m counts/100 milliliters or less in 98 percent of the daily samples was found.<br />

These differences may be attributed to variations in gallery O&M.<br />

211


212<br />

Cummulative Distribution<br />

100%<br />

90%<br />

80%<br />

70%<br />

60%<br />

50%<br />

40%<br />

30%<br />

20%<br />

10%<br />

0%<br />

Conclusions<br />

Suspended Solids in Galleries. Frequency Distribution.<br />

Rainy Season (2003)<br />

0 1 2 3 4 5 6 7 8<br />

NTU<br />

Figure 8. Quality of water produced by the galleries.<br />

Gallery 4<br />

Gallery 1<br />

Infiltration galleries built in the permeable alluvial deposits of the rivers can be suitable systems<br />

<strong>for</strong> improving water quality from surface waters. Removal ratios of more than 90 percent <strong>for</strong><br />

turbidity, suspended solids, and fecal coli<strong>for</strong>m have been recorded (during this research project).<br />

Infiltration galleries produce high-quality and stable volumes of water independent of seasonal<br />

variations in the quality and quantity of raw-surface stream, minimizing environmental effects.<br />

It seems that there is a natural and self-cleaning filtration process at work here. The permeable<br />

deposits of the Parapeti River <strong>for</strong>m a highly efficient system <strong>for</strong> the pretreatment of water that<br />

contains high concentrations of solids.<br />

Moreover, the filtration qualities of the deposits are constantly maintained by the action of the<br />

river itself: during periods of flooding, the top layer of filtration material — the coarse sand — is<br />

stirred-up and cleaned by the sheer <strong>for</strong>ce of water flow. As flooding subsides, the clean, loose sand<br />

re-settles on the riverbed, thus preventing the natural process of clogging and hardening, which<br />

would otherwise impair its efficacy as a filter. It is, there<strong>for</strong>e, possible that the natural composition<br />

of the river deposits, assisted by the cycle of the river itself, are acting as an effective and<br />

self-cleaning filter of the solids contained in the river.<br />

With proper studies and depending on local conditions and the characteristics of raw-water<br />

quality, infiltration galleries are a suitable alternative <strong>for</strong> supplying drinking water to small and<br />

medium communities, with minimum capital costs and lower O&M costs. The system in Camiri<br />

has been running since the early 1980s, and no critical O&M problems have been observed<br />

(particularly clogging). It is essential that galleries must have easy access to facilitate the periodic<br />

cleaning of sediments from conduits (the minimum channel section should be 1.0-m × 1.0-m or<br />

more <strong>for</strong> manual cleaning). The infiltration gallery system in Camiri has shown that the amount<br />

of maintenance required on the galleries is very small indeed. For this reason, as well as the<br />

benefits of improved or maintained water quality, infiltration galleries have a significant advantage<br />

over other conventional systems.<br />

Further research is still needed to <strong>for</strong>m a complete understanding of the whole process and the<br />

mechanisms that are involved in removing contaminants at higher concentrations.


REFERENCES<br />

American Public Health Association (1975). Standard Methods <strong>for</strong> the Examination of <strong>Water</strong> and Wastewater,<br />

Fourteenth Edition, American Public Health Association, New York.<br />

Binnie & Partners Consultants (1982). Final Report on the design of the surface water treatment plant in Camiri,<br />

Binnie & Partners Consultants, Santa Cruz, Bolivia.<br />

Harr, M.E. (1990). Groundwater and Seepage, General Publishing Company, Toronto, Ontario.<br />

Huisman, L. (1978). Ground <strong>Water</strong> Recovery, International <strong>Institute</strong> <strong>for</strong> Hydraulic and Environmental<br />

Engineering, Technical University of Delft, The Netherlands.<br />

Schubert, J. (2002). “German Experience with Riverbank Filtration Systems Ray.” Riverbank Filtration<br />

Improving Source-<strong>Water</strong> Quality, C. Ray, G. Melin, and R.B. Linsky, editors, Kluwer Academic Publishers,<br />

Dordrecht.<br />

Since 2002, ALVARO CAMACHO has been a consultant engineer <strong>for</strong> the water and<br />

sanitation sectors, including the Bolivian Association of Sanitary Engineers. Prior, he was<br />

the Director General of <strong>Water</strong> Supply and Sanitation <strong>for</strong> the Ministry of Housing Services<br />

of the Republic of Bolivia, a Project Leader in the <strong>Water</strong> and Sanitation Program <strong>for</strong> rural<br />

areas in Bolivia <strong>for</strong> the World Bank, and Guest Lecturer of Sanitary Engineering at the<br />

Universidad Mayor de San Andrés in Bolivia. He is also the author of several reports on<br />

topics such as water-treatment plants designed within the multi-stage filtration concept,<br />

developing the water and sanitation sector in Bolivia, technical standards <strong>for</strong> the design of unconventional<br />

sewer systems in Bolivia, and guidelines <strong>for</strong> the design of water-treatment plants in small and medium<br />

communities. Camacho received a degree in Civil Engineering from the Universidad Mayor de San Andrés<br />

in Bolivia and a Dipl.-Engineer in Sanitary Engineering from the International <strong>Institute</strong> <strong>for</strong> Hydraulics and<br />

Environmental Engineering, in conjunction with the Technical University of Delft, The Netherlands. He is<br />

a M.S. researcher at the Universidad del Valle in Cali, Columbia, where he is conducting a field study on<br />

infiltration galleries <strong>for</strong> water supply.<br />

213


214


Session 11: Case Studies “Lessons Learned”<br />

Sensitivity and Implications of Microscopic Particulate<br />

Analysis – A Collector Well Owner’s Perspective<br />

Barry C. Beyeler<br />

City of Boardman<br />

Boardman, Oregon<br />

The City of Boardman, Oregon, owns two horizontal collector wells situated along the banks of<br />

the Columbia River, which provides the City’s water supply. The horizontal collector well system<br />

was chosen in 1975 <strong>for</strong> its unique ability to produce high volumes of high-quality water. The<br />

horizontal collector wells fill a role of vital importance in the City by addressing the balance of<br />

plentiful water supplies <strong>for</strong> industrial development and the need <strong>for</strong> high-quality drinking water<br />

<strong>for</strong> the citizens of the community.<br />

The City has been confident of the water quality produced by the collector and the filtration<br />

effectiveness it possesses; however, under provisions of the Surface <strong>Water</strong> Treatment Rule, the<br />

potential water-quality effects of the hydraulic connectivity to the Columbia River came under<br />

question. The City chose to per<strong>for</strong>m Microscopic Particulate Analysis on the collector and Columbia<br />

River in the attempt to indicate and quantify the filtration effectiveness of the horizontal collector<br />

well system.<br />

Per<strong>for</strong>mance of Microscopic Particulate Analysis in the field presented many challenges that<br />

needed to be addressed to ensure that a quality representative sample was taken. There are many<br />

ways this sample can be compromised, which could produce non-representative or inaccurate<br />

results. Understanding the sensitivity of the Microscopic Particulate Analysis process from start to<br />

finish, combined with an assessment of what the indicated results would trigger under the Safe<br />

Drinking <strong>Water</strong> Act concerning additional treatment and subsequent associated costs, required<br />

significant care and diligence.<br />

As a result of the sampling and analysis that the City per<strong>for</strong>med, the State of Oregon Department<br />

of Human Resources Drinking <strong>Water</strong> Program has determined that the City is operating a<br />

“groundwater” system. This is based on the quality of the water and assurances provided by<br />

accurate and representative data obtained through Microscopic Particulate Analysis. As the City<br />

brings its second collector online, the coordination of the Microscopic Particulate Analysis<br />

sampling ef<strong>for</strong>ts are being discussed with the Department of Human Resources Drinking <strong>Water</strong><br />

Program to ensure that this collector per<strong>for</strong>mance can be quantified appropriately. The Microscopic<br />

Particulate Analysis process will be accomplished over the next year to provide the in<strong>for</strong>mation<br />

to make appropriate decisions. The City will follow the same process of sampling the river and<br />

collector to indicate filtration effectiveness.<br />

In retrospect, there have been numerous other benefits to the process of Microscopic Particulate<br />

Analysis sampling. The increased knowledge of how the system works has been invaluable in<br />

Correspondence should be addressed to:<br />

Barry C. Beyeler<br />

Community Development Director<br />

City of Boardman, Oregon<br />

202 N Main Street • P.O. Box 229 • Boardman, Oregon 97818 USA<br />

Phone: (541) 481-9252 • Fax: (541) 481-3244 • Email: bbeyeler@cityofboardman.com<br />

215


216<br />

<strong>for</strong>ums concerning Endangered Species Act listings of salmon and steelhead populations in the<br />

Columbia and Snake River basins. As proposals <strong>for</strong> river operations are brought <strong>for</strong>ward, the City<br />

of Boardman can more accurately respond to what the potential impacts may be. The increased<br />

knowledge and communication between the City and State and Federal agencies has also been<br />

beneficial in allowing the City to get through permitting processes associated with this water right.<br />

Although there is no direct method to measure the impact Microscopic Particulate Analysis has<br />

had in this regard, there is no doubt that it has played a role in improving communication linkages<br />

and the level of understanding between the agencies and public involved.<br />

Since April 2003, BARRY BEYELER has been Community Development Director <strong>for</strong> the<br />

City of Boardman, Oregon, where he has been employed <strong>for</strong> over 22 years. Among his<br />

responsibilities, he coordinates and reviews all land-use activities within the City and its<br />

urban growth boundary. He also oversees all planning and land-use issues relating to<br />

utilities operations (water, wastewater, storm drainage, and traffic infrastructure), natural<br />

resource impacts and environmental policy, wellhead protection, and public education.<br />

Among his professional affiliations, Beyeler is a member of the Oregon <strong>Water</strong> Resources<br />

Department Ground <strong>Water</strong> Advisory Committee, where he is currently serving a third 4-year term and is<br />

<strong>for</strong>mer Committee Chair. Beyeler received the Northeastern Oregon Sub-Section American <strong>Water</strong> Works<br />

Association Activities Award in both 1992 and 1995.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!