12.07.2015 Views

Bromate Formation and Control During Ozonation of Low Bromide ...

Bromate Formation and Control During Ozonation of Low Bromide ...

Bromate Formation and Control During Ozonation of Low Bromide ...

SHOW MORE
SHOW LESS
  • No tags were found...

Create successful ePaper yourself

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

AWWA Microbial/DisinfectionResearch By-Products ResearchFoundation council<strong>Bromate</strong> <strong>Formation</strong><strong>and</strong> <strong>Control</strong> <strong>During</strong><strong>Ozonation</strong> <strong>of</strong> <strong>Low</strong><strong>Bromide</strong> WatersSubject Area:Water Treatment


<strong>Bromate</strong> <strong>Formation</strong><strong>and</strong> <strong>Control</strong> <strong>During</strong><strong>Ozonation</strong> <strong>of</strong> <strong>Low</strong><strong>Bromide</strong> Waters


The mission <strong>of</strong> the Awwa Research Foundation is to advance the science <strong>of</strong> water to improvethe quality <strong>of</strong> life. Funded primarily through annual subscription payments from over 1,000 utilities, consulting firms, <strong>and</strong> manufacturers in North America <strong>and</strong> abroad, AwwaRF sponsorsresearch on all aspects <strong>of</strong> drinking water, including supply <strong>and</strong> resources, treatment, monitoring<strong>and</strong> analysis, distribution, management, <strong>and</strong> health effects.From its headquarters in Denver, Colorado, the AwwaRF staff directs <strong>and</strong> supports the efforts<strong>of</strong> over 500 volunteers, who are the heart <strong>of</strong> the research program. These volunteers, serving onvarious boards <strong>and</strong> committees, use their expertise to select <strong>and</strong> monitor research studies to benefit the entire drinking water community.Research findings are disseminated through a number <strong>of</strong> technology transfer activities, including research reports, conferences, videotape summaries, <strong>and</strong> periodicals.


<strong>Bromate</strong> <strong>Formation</strong><strong>and</strong> <strong>Control</strong> <strong>During</strong><strong>Ozonation</strong> <strong>of</strong> <strong>Low</strong><strong>Bromide</strong> Waters___Prepared by:Thomas Gillogly, Issam NajmMontgomery Watson, Pasadena, California 91101Roger Minear, Benito Marinas, Mark Urban, Jae Hong Kim, Shinya EchigoUniversity <strong>of</strong> Illinois at Urbana-Champaign, Urbana, Illinois 61801Gary Amy, Christopher Douville, Brian DawUniversity <strong>of</strong> Colorado at Boulder, Boulder, Colorado 80309Robert Andrews, Ron H<strong>of</strong>mannUniversity <strong>of</strong> Toronto, Toronto, Ontario, M5S 1A4, Canada<strong>and</strong>Jean-Philippe CroueUniversite de Poitiers, 86022 Poitiers Cedex, FranceSponsored by:Microbial/Disinfection By-Products Research CouncilJointly funded by:Awwa Research Foundation6666 West Quincy AvenueDenver, CO 80235-3098U.S. Environmental Protection AgencyWashington, DC 20460<strong>and</strong>California Urban Water Agencies455 Capitol Mall, Suite 705Sacramento, CA 95814Published by theAwwa Research Foundation <strong>and</strong>American Water Works Association


DisclaimerThis study was jointly funded for the Microbial/Disinfection By-Products Research Council (M/DBP) by the Awwa ResearchFoundation (AwwaRF) the U.S. Environmental Protection Agency (USEPA), <strong>and</strong> the California Urban Water Agencies (CUWA)under Cooperative Agreement No. CX819540. AwwaRF, M/DBP, USEPA <strong>and</strong> CUWA assume no responsibility for the content<strong>of</strong> the research study reported in this publication or for the opinions or statements <strong>of</strong> fact expressed in the report. The mention<strong>of</strong> trade names for commercial products does not represent or imply the approval or endorsement <strong>of</strong> AwwaRF, M/DBP, USEPA,or CUWA. This report is presented solely for informational purposes.Library <strong>of</strong> Congress Cataloging-in-Publication Data has been applied for.Copyright 2001byAwwa Research Foundation<strong>and</strong>American Water Works AssociationPrinted in the U.S.A.ISBN 1-58321-155-1Printed on recycled paper


CONTENTSLIST OF TABLES......................................................................................................................ixLISTOFnOURES...........................................................................................................-........^!PREFACE ...........................................................................................................................xvFOREWORD ..........................................................................................................................xviiACKNOWLEDGMENTS.........................................................................................................xixEXECUTIVE SUMMARY....................................................................................................... xxiCHAPTER 1. INTRODUCTION................................................................................................ 11.1 Regulatory Background....................................................................................................... 11.2 Ozone Doses for Cryptosporidium Inactivation.................................................................. 11.2.1 Semi-Batch Cryptosporidiumparvum Inactivation..................................................... 11.2.2 Multi-Utility Cryptosporidium Inactivation Study...................................................... 21.3 Ozone Reaction Chemistry Background............................................................................. 31.4 Information Gaps................................................................................................................. 51.5 Project Objectives ............................................................................................................... 61.6 Approach............................................................................................................................. 6CHAPTER 2. MATERIALS AND METHODS.......................................................,.................?2.1 Source Waters ..................................................................................................................... 72.2 Reactors ..............................................................................................................................72.2.1 True-Batch Reactors.................................................................................................... 82.2.2 Semi-Batch Reactor................................................................................................... 112.2.3 Laboratory-Scale Continuous-Flow Ozone Contactor.............................................. 112.2.4 Full-Scale <strong>and</strong> Pilot-Scale Los Angeles Aqueduct Filtration Plant OzoneContactors.................................................................................................................. 132.2.5 Pilot-Scale Britannia Water Purification Facility Ozone Contactor.......................... 152.3 Calculation <strong>of</strong> Ozone Contact........................................................................................... 162.4 Analytical Methods........................................................................................................... 162.4.1 <strong>Bromide</strong> <strong>and</strong> <strong>Bromate</strong>................................................................................................ 172.4.2 Organic Bromine .......................................................................................................172.4.3 Total <strong>and</strong> Dissolved Organic Carbon......................................................................... 182.4.4 Ultraviolet Absorbance.............................................................................................. 182.4.5 NOMFractionation.................................................................................................... 182.4.6 Size Exclusion Chromatography ............................................................................... 192.4.7 Alkalinity...................................................................................................................192.4.8 Ammonia................................................................................................................... 202.4.9 pH ............................................................................................................................202.4.10 Ozone......................................................................................................................... 202.4.11 Cryptosporidium parvum.. .........................................................................................202.4.12 Quality Assurance <strong>and</strong> Quality <strong>Control</strong>..................................................................... 212.5 Model Development.......................................................................................................... 25


2.5.1 Cryptosporidiumparvum Inactivation Kinetics ........................................................ 252.5.2 Ozone Decomposition <strong>and</strong> <strong>Bromate</strong> <strong>Formation</strong> Mechanisms................................... 262.5.3 Ozone Contactor Modeling........................................................................................ 29CHAPTER 3. BROMATE SURVEY....................................................................................... 313.1 Participation...................................................................................................................... 313.2 Organization......................................................................................................................323.3 Details <strong>of</strong> Analyses............................................................................................................ 343.4 Results ............................................................................................................................353.4.1 Individual Survey Rounds ......................................................................................... 353.4.2 Composite Survey Database...................................................................................... 423.5 Discussion .........................................................................................................................453.5.1 United States <strong>Bromate</strong> <strong>Formation</strong> Survey................................................................. 453.5.2 Comparison to European <strong>Bromate</strong> <strong>Formation</strong> Survey............................................... 45CHAPTER 4. TRUE-BATCH INFLUENCE OF NOM AND TEMPERATURE ONBROMATE FORMATION................................................................................ 494.1 Influence <strong>of</strong> NOM on <strong>Bromate</strong> <strong>Formation</strong>........................................................................ 494.1.1 Source Waters............................................................................................................ 504.1.2 True-Batch <strong>Ozonation</strong> Experiments.......................................................................... 504.1.3 Results.......................................................................................................................^!4.1.4 Statistical Analysis..................................................................................................... 644.2 Temperature Effects on <strong>Bromate</strong> <strong>Formation</strong> <strong>and</strong> Disinfection......................................... 654.2.1 Temperature <strong>Control</strong>led True Batch <strong>Ozonation</strong>........................................................ 654.2.2 Results........................................................................................................................ 664.2.3 Discussion.................................................................................................................. 674.3 Summary ...........................................................................................................................68CHAPTER 5. EFFECT OF WATER QUALITY AND MINIMIZATION APPROACHESON BROMATE FORMATION......................................................................... 705.1 <strong>Bromate</strong> <strong>Formation</strong> ...........................................................................................................705.1.1 Effect <strong>of</strong> Cryptosporidium Inactivation..................................................................... 705.1.2 Temperature Effects................................................................................................... 735.2 <strong>Bromate</strong> Minimization...................................................................................................... 755.2.1 pH Depression ...........................................................................................................755.2.2 Ammonia Addition.................................................................................................... 775.2.3 Cumulative Effects <strong>of</strong> Ammonia Addition <strong>and</strong> pH Depression................................ 805.2.4 Hydroxyl Radical Scavenger Addition...................................................................... 825.3 Summary........................................................................................................................... 82CHAPTER 6. HYDRODYNAMIC IMPACTS ON BROMATE FORMATION.................... 856.1 True-Batch Staged <strong>Ozonation</strong>........................................................................................... 866.2 Pilot-Scale Staged <strong>Ozonation</strong>............................................................................................ 896.3 Co-Current Versus Counter-Current <strong>Ozonation</strong> ............................................................... 926.4 Summary........................................................................................................................... 95CHAPTER 7. A COMPARISON OF BROMATE FORMATION WITHIN DIFFERENTSCALES OF OZONE CONTACTORS............................................................. 977.1 Los Angeles Aqueduct Filtration Plant............................................................................. 97VI


7.1.1 Testing Plan...............................................................................................................977.1.2 True-Batch Ozone Reactor........................................................................................ 987.1.3 Laboratory-Scale Continuous-Flow Ozone Contactor............................................ 1007.1.4 Full-Scale <strong>and</strong> Pilot-Scale Ozone Contactors.......................................................... 1017.1.5 Comparison <strong>of</strong> Full-, Pilot-, <strong>and</strong> Bench-Scales....................................................... 1057.2 Neuilly sur Marne Water Treatment Plant...................................................................... 1067.3 Britannia Water Purification Facility.............................................................................. 1097.4 Conclusions..................................................................................................................... IllCHAPTER 8. OZONE CONTACTOR MODELING............................................................. 1138.1 Batch <strong>and</strong> Semi-Batch Reactor Modeling....................................................................... 1138.2 Bench-Scale Flow-through Ozone Contactor Modeling................................................. 1148.3 Pilot-Scale Ozone Contactor Modeling........................................................................... 1188.4 Full-scale ozone contactor modeling............................................................................... 1218.5 Summary......................................................................................................................... 122CHAPTER 9. SUMMARY AND CONCLUSIONS............................................................... 124APPENDIX .......................................................................................................................... 127REFERENCES......................................................................................................................... 147ABBREVIATIONS.................................................................................................................. 153vn


LIST OF TABLESTable 1.1: Proposed CTs Required for the Inactivation <strong>of</strong> Cryptosporidium parvum oocystsWith Ozone Applied to a Semi-Batch Reactor................................................................... 2Table 1.2: Proposed CTs Required for the Inactivation <strong>of</strong> Cryptosporidium parvum OocystsWith Ozone Applied to a Continuous-Flow Reactor.......................................................... 3Table 2.1: Source Water Quality Parameters................................................................................. 8Table 2.2: Source Water Quality Parameters After Dilution with Stock Ozone Solution........... 10Table 2.3: Experimental Matrix #1.............................................................................................. 11Table 2.4: Experimental Matrix #2 .............................................................................................. 12Table 2.5: First Round-Robin Results.......................................................................................... 21Table 2.6: Second Round-Robin Results..................................................................................... 22Table 2.7: Third Round-Robin <strong>Bromate</strong> Results ......................................................................... 23Table 2.8: Third Round-Robin <strong>Bromide</strong> Results......................................................................... 24Table 2.9: Analytical Method Reporting Levels <strong>and</strong> Coefficients <strong>of</strong> Variation.......................... 24Table 2.10: Experimental Quality <strong>Control</strong> for Laboratory-Scale Continuous-Flow OzoneContactor........................................................................................................................... 25Table 2.11: Ozone Decomposition <strong>and</strong> <strong>Bromate</strong> <strong>Formation</strong> Mechanism.................................... 26Table 3.1: Qualitative Summary <strong>of</strong> Survey Participants.............................................................. 33Table 3.2: Survey Utility Participant <strong>Ozonation</strong> Information...................................................... 34Table 3.3: Water Quality <strong>and</strong> Treatment Condition Summary <strong>of</strong> Sampling Round One............ 36Table 3.4: Water Quality <strong>and</strong> Treatment Condition Summary <strong>of</strong> Sampling Round Two ........... 36Table 3.5: Water Quality <strong>and</strong> Treatment Condition Summary <strong>of</strong> Sampling Round Three ......... 37Table 3.6: Pearson Correlation Matrix for Sampling Round One................................................ 38Table 3.7: Pearson Correlation Matrix for Sampling Round Two...............................................39Table 3.8: Pearson Correlation Matrix for Sampling Round Three............................................. 40Table 3.9: Top Five Highest Correlating Parameters to <strong>Bromate</strong> for Single-LinearRegression Analysis <strong>of</strong> Round One ..................................................................................41Table 3.10: Top Five Highest Correlating Parameters to <strong>Bromate</strong> for Single-LinearRegression Analysis <strong>of</strong> Round Two..................................................................................41Table 3.11: Top Five Highest Correlating Parameters to <strong>Bromate</strong> for Single-LinearRegression Analysis <strong>of</strong> Round Three................................................................................ 41Table 3.12: Cumulative Water Quality <strong>and</strong> Treatment Condition Summary .............................. 42Table 3.13: Cumulative Pearson Correlation Matrix...................................................................44Table 3.14: Top Five Highest Correlating Parameters to <strong>Bromate</strong> for Single-LinearRegression Analysis <strong>of</strong> the Cumulative Data.................................................................... 44Table 3.15: Water Quality <strong>and</strong> Treatment Condition Summary <strong>of</strong> European Union <strong>Bromate</strong>Survey................................................................................................................................46Table 3.16: Pearson Correlation Matrix for European Union <strong>Bromate</strong> Survey........................... 47Table 4.1: NOM Characterization <strong>of</strong> Raw Source Waters........................................................... 52Table 4.2: Summary <strong>of</strong> <strong>Bromate</strong> <strong>Formation</strong> Potentials <strong>and</strong> Ozone Exposure Results ................ 56Table 4.3: True-Batch Post-<strong>Ozonation</strong> Water Quality Parameters.............................................. 59Table 4.4: Estimated <strong>Bromate</strong> <strong>Formation</strong> for Cryptosporidium Inactivation (20 C, pH 7)........ 62Table 4.5: Pearson Correlation Matrix for <strong>Bromate</strong> <strong>Formation</strong> Potential Experiments .............. 63Table 4.6: Results for True-Batch <strong>Ozonation</strong> at 10 C................................................................. 66IX


LIST OF FIGURESFigure 1.1: <strong>Bromate</strong> <strong>Formation</strong> Pathways.....................:................................................................4Figure 2.1: True-Batch <strong>Ozonation</strong> Equipment Setup..................................................................... 9Figure 2.2: How-Through Reactor Schematic............................................................................. 13Figure 2.3: Schematic <strong>of</strong> the Full-Scale Ozone Contactor at the Los Angeles AqueductFiltration Plant................................................................................................................... 14Figure 2.4: Schematic <strong>of</strong> Pilot-Scale Ozone Contactor at the Los Angeles AqueductFiltration Plant................................................................................................................... 14Figure 2.5: Britannia Water Purification Facility Pilot-Scale Ozone Contactor.......................... 15Figure 3.1: Participation Map for Survey Participants................................................................. 32Figure 3.2: Cumulative Distribution <strong>of</strong> <strong>Bromate</strong> <strong>Formation</strong> in 78 Samples................................ 43Figure 4.1: Source Water Ultraviolet Adsorbance Spectra.......................................................... 53Figure 4.2: Source Water DOC <strong>and</strong> SUVA Values..................................................................... 53Figure 4.3: Distribution <strong>of</strong> NOM Fractions ................................................................................. 54Figure 4.4: Size Exclusion Chromatography Chromatograms for CCD Water........................... 55Figure 4.5: <strong>Low</strong> DOC Source Water Ozone Decay Curves (DOC 3 mg/L)............................. 57Figure 4.7: ASUVA Spectra <strong>of</strong> Ozonated Project Waters............................................................ 58Figure 4.8: Linear Relationship Between <strong>Bromate</strong> <strong>Formation</strong> <strong>and</strong> Ozone Exposure (LAW)..... 60Figure 4.9: <strong>Bromate</strong> <strong>Formation</strong> Potential <strong>and</strong> Ozone Exposure (20 C, pH 7)............................. 61Figure 4.10: Ozone Exposure versus <strong>Bromate</strong> <strong>Formation</strong> Potential for 2-LogCryptosporidium Inactivation at 10 <strong>and</strong> 20 C .................................................................. 68Figure 5.1: Impact <strong>of</strong> Ozone Exposure, Expressed as Log-Cryptosporidium Inactivation, on<strong>Bromate</strong> <strong>Formation</strong> in the Laboratory-Scale Continuous-Flow Ozone Contactor (pH7. 15 C)............................................................................................................................. 71Figure 5.2: Impact <strong>of</strong> Ozone Exposure, Expressed as Log-Cryptosporidium Inactivation, on<strong>Bromate</strong> <strong>Formation</strong> in the Laboratory-Scale Continuous-Flow Ozone Contactor (pH8. 15 C)............................................................................................................................. 72Figure 5.3: Impact <strong>of</strong> Ozone Exposure, Expressed as Log-Cryptosporidium Inactivation, onthe Percent Conversion <strong>of</strong> <strong>Bromide</strong> to <strong>Bromate</strong> (pH 7,15 C).......................................... 72Figure 5.4: Effect <strong>of</strong> Temperature on <strong>Bromate</strong> <strong>Formation</strong> (2-Log CryptosporidiumInactivation, pH 7)............................................................................................................. 73Figure 5.5: Effect <strong>of</strong> Temperature on <strong>Bromate</strong> <strong>Formation</strong> (1-Log CryptosporidiumInactivation, pH 8)............................................................................................................. 74Figure 5.6: Effect <strong>of</strong> pH on <strong>Bromate</strong> <strong>Formation</strong> (2-Log Cryptosporidium Inactivation, 15 C).. 76Figure 5.7: Effect <strong>of</strong> pH on <strong>Bromate</strong> <strong>Formation</strong> (1-Log Cryptosporidium Inactivation, 15 C).. 76Figure 5.8: Effect <strong>of</strong> pH on TOBr <strong>Formation</strong> (2-Log Cryptosporidium Inactivation, 15 C)...... 77Figure 5.9: Effect <strong>of</strong> Ammonia Addition on <strong>Bromate</strong> <strong>Formation</strong> (2-Log CryptosporidiumInactivation, 15 C, pH 7).................................................................................................. 78Figure 5.10: Effect <strong>of</strong> Ammonia Addition on <strong>Bromate</strong> <strong>Formation</strong> (1-Log CryptosporidiumInactivation, 15 C, pH 8).................................................................................................. 78Figure 5.11: Effect <strong>of</strong> Ammonia on <strong>Bromate</strong> <strong>Formation</strong> (Ottawa Pilot Plant, 0.5 C, pH 8.5,0.1 mg/L Bf)..................................................................................................................... 80XI


Figure 5.12: Effect <strong>of</strong> pH Depression <strong>and</strong> Ammonia Addition on <strong>Bromate</strong> <strong>Formation</strong> inCCD (1-Log Cryptosporidiwn Inactivation, 15 C).......................................................... 81Figure 5.13: Effect <strong>of</strong> pH Depression <strong>and</strong> Ammonia Addition on <strong>Bromate</strong> <strong>Formation</strong> inSPW (1-Log Cryptosporidium Inactivation, 15 C).......................................................... 81Figure 5.14: Effect <strong>of</strong> a Radical Scavenger (Tertiary Butanol) Addition on <strong>Bromate</strong><strong>Formation</strong> in the Laboratory-Scale Continuous-Flow Ozone Contactor (15 C).............. 82Figure 6.1: Impact <strong>of</strong> Staged <strong>Ozonation</strong> on Bench-Scale Batch <strong>Bromate</strong> <strong>Formation</strong>................. 89Figure 6.2: Ozone Pr<strong>of</strong>ile Through Britannia Pilot Ozone Contactor <strong>During</strong> Staged<strong>Ozonation</strong> (pH 8.5, 0.5 C)................................................................................................ 90Figure 6.3: Cumulative CT Through Britannia Pilot Ozone Contactor <strong>During</strong> Staged<strong>Ozonation</strong> (pH 8.5,0.5 C)................................................................................................ 91Figure 6.4: <strong>Bromate</strong> <strong>Formation</strong> Through Britannia Pilot Ozone Contactor <strong>During</strong> Staged<strong>Ozonation</strong> (pH 8.5, 0.5 C)................................................................................................ 91Figure 6.5: <strong>Bromate</strong> <strong>Formation</strong> as a Function <strong>of</strong> CT(pH 8.5, 0.5 C)..:...................................... 92Figure 6.6: Ozone Pr<strong>of</strong>ile Versus Contactor Depth for Counter- <strong>and</strong> Co-Current Operation ..... 93Figure 6.7: <strong>Bromate</strong> Pr<strong>of</strong>ile Across Contactor Depth For Counter- <strong>and</strong> Co-Current Operation. 93Figure 6.8: Calculated Crfor Co-<strong>and</strong> Counter-Current <strong>Ozonation</strong>............................................94Figure 6.9: <strong>Bromate</strong> Concentration as a Function <strong>of</strong> CT for Co- <strong>and</strong> Counter-Current<strong>Ozonation</strong>..........................................................................................................................95Figure 7.1: <strong>Bromate</strong> <strong>Formation</strong> in a True-Batch Reactor (LAW, pH 7, 20 C)........................... 99Figure 7.2: True-Batch Reactor <strong>Bromide</strong> Reaction <strong>and</strong> <strong>Bromate</strong> <strong>Formation</strong> (LAW, pH 7,20 C)................................................................................................................................. 99Figure 7.3: <strong>Bromate</strong> <strong>Formation</strong> in Laboratory-Scale Continuous-Flow Reactor....................... 100Figure 7.4: <strong>Bromide</strong> Reaction <strong>and</strong> <strong>Bromate</strong> <strong>Formation</strong> in a Laboratory-Scale Continuous-How Contactor................................................................................................................ 101Figure 7.5: Effect <strong>of</strong> CT Calculation <strong>of</strong> <strong>Bromate</strong> <strong>Formation</strong> Trends......................................... 103Figure 7.6: Los Angeles Aqueduct Filtration Plant <strong>Bromate</strong> Production.................................. 103Figure 7.7: Los Angeles Aqueduct Filtration Pilot Plant <strong>Bromate</strong> Production.......................... 104Figure 7.8: CTALL-ISO Development Along Full-Scale Contactor............................................... 104Figure 7.9: CTALL-t50 Development Along Pilot-Scale Contactor.............................................. 105Figure 7.10: Results <strong>of</strong> Los Angeles Aqueduct Filtration Plant Multiple-Scale ComparisonStudy................................................................................................................................ 106Figure 7.11: Neuilly sur Marne Comparison <strong>of</strong> Ozone Contactors (May 1998)....................... 108Figure 7.12: Neuilly sur Marne Comparison <strong>of</strong> Ozone Contactors (October 1998).................. 108Figure 7.13: Britannia Water Purification Facility Ozone Contactor Comparison (October1999, pH 8)...................................................................................................................... 110Figure 7.14: Temperature Effect on Laboratory-Scale Reactors (August 1999, pH 8)............. IllFigure 8.1: Effect <strong>of</strong> pH on Experimental <strong>and</strong> Fitted Kinetics <strong>of</strong> the Inactivation <strong>of</strong>Cryptosporidiumparvum Oocysts (20 C, Semi-Batch Reactor).................................... 113Figure 8.2: Batch Ozone Decay Kinetics for Selected Natural Waters (pH 7, 20 C)................ 114Figure 8.3: Fitted Batch Reactor <strong>Bromate</strong> <strong>Formation</strong> Potential................................................. 115Figure 8.4: Tracer Test Results for the Bench-Scale Flow-Through Reactor <strong>and</strong>Corresponding ADR Modeling....................................................................................... 116Figure 8.5: Cryptosporidium parvum Inactivation Results <strong>and</strong> Corresponding ModelPredictions for the Laboratory-Scale Continuous-Flow Ozone Contactor..................... 117xn


Figure 8.6: <strong>Bromate</strong> <strong>Formation</strong> <strong>and</strong> Prediction for the Bench-Scale Flow-Through OzoneContactor (LAW & OTT Spiked to 90 ug/L Br') ............................................................117Figure 8.7: Britannia Pilot Plant Counter-Current Tracer Test <strong>and</strong> Corresponding ADRModeling (Qwater = 5L/min,Qgas= 1-42 L/min,0.5 C)....................................................119Figure 8.8: Britannia Pilot Plant Measured <strong>and</strong> Predicted <strong>Bromate</strong> <strong>Formation</strong>, Ozone Pr<strong>of</strong>ile<strong>and</strong> Cryptosporidium Inactivation (pH 8.2,100 jig/L Br", 2.6 mg/L DOC, 0.5 C,Qwater = 5L/min, Qgas - 1.42 L/min).................................................................................120Figure 8.9: Los Angeles Aqueduct Filtration Plant Tracer Tests <strong>and</strong> Corresponding ADRModeling (Qwa,er = 85 MOD, Qgas = 17.8 scfh)................................................................121Figure 8.10: Los Angeles Aqueduct Filtration Plant Measured <strong>and</strong> Predicted Ozone Pr<strong>of</strong>ile,<strong>Bromate</strong> <strong>Formation</strong> <strong>and</strong> Cryptosporidium Inactivation (pH 8.2, 33 |ag/L Br", 1.9 mg/LDOC, 9 C Q^er = 85 MOD, Qgas = 17.8 scfh)................................................................122xni


PREFACEThe Microbial/Disinfection By-Products Research Council was established in 1995 as avehicle for the selection <strong>and</strong> funding <strong>of</strong> research to provide scientific information in the areas <strong>of</strong>health effects, exposure assessment, risk assessment, <strong>and</strong> prevention <strong>and</strong> control <strong>of</strong>contamination by microbes <strong>and</strong> disinfection by-products in drinking water. The council iscomposed <strong>of</strong> representatives designated by the U.S. Environmental Protection Agency (USEPA),the Awwa Research Foundation (AwwaRF) Board <strong>of</strong> Trustees, the Association <strong>of</strong> State DrinkingWater Administrators, the National Resources Defense Council, the National EnvironmentalHealth Association, or their designees. Sources <strong>of</strong> funding for this research include the USEPA<strong>and</strong> AwwaRF, along with other interested parties. The council disburses these funds for researchdeemed to be <strong>of</strong> the highest urgency <strong>and</strong> importance in resolving critical research issues indrinking water.xv


FOREWORDThe Awwa Research Foundation is a nonpr<strong>of</strong>it corporation that is dedicated to theimplementation <strong>of</strong> a research effort to help utilities respond to regulatory requirements <strong>and</strong>traditional high-priority concerns <strong>of</strong> the industry. The research agenda is developed through aprocess <strong>of</strong> consultation with subscribers <strong>and</strong> drinking water pr<strong>of</strong>essionals. Under the umbrella <strong>of</strong>a Strategic Research Plan, the Research Advisory Council prioritizes the suggested projectsbased upon current <strong>and</strong> future needs, applicability, <strong>and</strong> past work; the recommendations areforwarded to the Board <strong>of</strong> Trustees for final selection. The foundation also sponsors researchprojects through the unsolicited proposal process; the Collaborative Research, ResearchApplications, <strong>and</strong> Tailored Collaboration programs; <strong>and</strong> various joint research efforts withorganizations such as the US Environmental Protection Agency, the US Bureau <strong>of</strong> Reclamation,<strong>and</strong> the Association <strong>of</strong> California Water Agencies.This publication is a result <strong>of</strong> one <strong>of</strong> these sponsored studies, <strong>and</strong> it is hoped that its findings willbe applied in communities throughout the world. The following report serves not only as ameans <strong>of</strong> communicating the results <strong>of</strong> the water industry's centralized research program but alsoas a tool to enlist the further support <strong>of</strong> the nonmember utilities <strong>and</strong> individuals.Projects are managed closely from their inception to the final report by the foundation's staff <strong>and</strong>large cadre <strong>of</strong> volunteers who willingly contribute their time <strong>and</strong> expertise. The foundationserves a planning <strong>and</strong> management function <strong>and</strong> awards contracts to other institutions such aswater utilities, universities, <strong>and</strong> engineering firms. The funding for this research effort comesprimarily from the Subscription Program, through which water utilities subscribe to the researchprogram <strong>and</strong> make an annual payment proportionate to the volume <strong>of</strong> water they deliver <strong>and</strong>consultants subscribe based on their annual billings. The program <strong>of</strong>fers a cost-effective <strong>and</strong> fairmethod for funding research in the public interest.A broad spectrum <strong>of</strong> water supply issues is addressed by the foundation's research agenda:resources, treatment <strong>and</strong> operations, distribution <strong>and</strong> storage, water quality <strong>and</strong> analysis,toxicology, economics, <strong>and</strong> management. The ultimate purpose <strong>of</strong> the coordinated effort is toassist water suppliers to provide the highest possible quality <strong>of</strong> water economically <strong>and</strong> reliably.The true benefits are realized when the results are implemented at the utility level. Thefoundation's trustees are pleased to <strong>of</strong>fer this publication as a contribution toward that end.Strategies for minimizing bromate formation under ozone dosages capable <strong>of</strong> inactivatingCryptosporidium remain a challenge. Since the median level <strong>of</strong> bromide in U.S. drinking watersis approximately 100 ng/L, it is important to know if ozonating natural waters with bromidelevels less than 100 ng/L would form bromate in excess <strong>of</strong> the 10 jag/L MCL. The datagenerated form this research should help water utilities underst<strong>and</strong> some <strong>of</strong> the factorscontrolling bromate formation <strong>and</strong> identify selected methods to minimize its formation.Edmund G. Archuleta, P.E. James F. Manwaring, P.E.Chair, Board <strong>of</strong> Trustees Executive DirectorAwwa Research Foundation Awwa Research Foundationxvn


ACKNOWLEDGMENTSThe authors would like to recognize the numerous agencies <strong>and</strong> people that were involved in thisproject. The project would not have been possible without the cooperation <strong>and</strong> support <strong>of</strong> thefollowing 13 water agencies that participated in the study:• Alameda County Water District• City <strong>of</strong> Amarillo• City <strong>of</strong> Ann Arbor• City <strong>of</strong> Dallas• City <strong>of</strong> Fan-field• City <strong>of</strong> Houston• Compagnie Generale des Eaux• Contra Costa Water District• Los Angeles Department <strong>of</strong> Water <strong>and</strong> Power• Metropolitan Water District <strong>of</strong> Southern California• Municipality <strong>of</strong> Ottawa• New Jersey American Water Company• Palm Beach County Public Utility DepartmentThe authors would also like to express their gratitude to all <strong>of</strong> the 24 water agencies thatanonymously participated in the survey <strong>of</strong> ozonation facilities.The authors also gratefully acknowledge the support <strong>and</strong> assistance <strong>of</strong> Kenan Ozekin, AwwaRFProject Officer, <strong>and</strong> the AwwaRF Project Advisory Committee (PAC): Michael Elovitz(Environmental Protection Agency, Cincinnati, OH); Rengao Song (Louisville Water Company,Louisville, KY); <strong>and</strong> William Soucie (Central Lake County Joint Action Water Agency, LakeBluff, IL). The contributions <strong>of</strong> Elaine Archibald, CUWA Project Officer, <strong>and</strong> the CUWA PAC:Stuart Krasner (Metropolitan Water District <strong>of</strong> Southern California, La Verne, CA), Susan Teefy(Alameda County Water District, Freemont, CA) <strong>and</strong> Larry McCollum (Contra Costa WaterDistrict, Concord, CA), are greatly appreciated.xix


EXECUTIVE SUMMARYIn 1992, the Environmental Protection Agency (EPA) initiated a negotiated rulemaking processthat eventually led to the Stage 1 Disinfectants <strong>and</strong> Disinfection By-Products (D/DBP) Rule(Federal Register, December 16, 1999). Included in the Stage 1 D/DBP rule was a maximumcontaminant level (MCL) for bromate (BrO3~) <strong>of</strong> 10 ug/L <strong>and</strong> an MCL Goal (MCLG) <strong>of</strong> zero.Based on available information, EPA determined that the theoretical 10~4 cancer risk estimate forbromate was 5 ug/L. The MCL <strong>of</strong> 10 ug/L, therefore, represents a risk level higher than EPA'sapproach <strong>of</strong> regulating carcinogens within the 10"4 to 10"6 risk range.While the 10 ug/L MCL is anticipated to impact a limited number <strong>of</strong> water agencies currentlyusing ozone as the primary disinfectant to inactivate Giardia <strong>and</strong> viruses, a greater number <strong>of</strong>water agencies will be impacted by this MCL when compliance with the Long-Term 2 EnhancedSurface Water Treatment Rule (LT2ESWTR) is required. The agreement in principle for thisrule recommends additional treatment requirements for up to 2.5-log reduction inCryptosporidium, beyond that already required by the Interim Enhanced Surface WaterTreatment Rule. Ozone was one <strong>of</strong> the treatment tools identified that could be used to inactivateCryptosporidium. While the Cryptosporidium inactivation tables have not yet been establishedby the EPA, preliminary information suggests CT values 5 to 30 times higher than that requiredfor Giardia inactivation will be required depending on the temperature <strong>of</strong> the water (Renneckeret al 1999; Oppenheimer et al., 2000).To address some <strong>of</strong> the concerns related to the use <strong>of</strong> ozone for disinfection, several researchershave examined the influences <strong>of</strong> water quality <strong>and</strong> water treatment variables on bromateformation with a focus on process modifications to minimize bromate formation duringozonation (Amy et al., 1997; Glaze et al., 1993; Krasner et al., 1993). However, most <strong>of</strong> theresearch on underst<strong>and</strong>ing bromate formation <strong>and</strong> control has utilized source waters containingmedium to high levels <strong>of</strong> bromide (>100 ug/L) ozonated at low to medium Os doses (


• Comparing bench-, pilot- <strong>and</strong> full-scale bromate formation trends at facilities orparticipating utilities having two or more scales available.BROMATE SURVEYA survey <strong>of</strong> operating ozone facilities in North America <strong>and</strong> Europe was initially conducted tounderst<strong>and</strong> bromate occurrence under current levels <strong>of</strong> ozone application. The average bromateconcentration at 24 full-scale ozonation plants, based on 78 samples from three separatesampling campaigns for the existing levels <strong>of</strong> ozonation, was 3.9 ug/L. The average percentconversion <strong>of</strong> bromide to bromate was 6.7 percent. The 10th, 50th (median), <strong>and</strong> 90th percentilebromate values was determined to be 0.2, 1.2, <strong>and</strong> 13 ug/L, respectively. The survey indicatedthat there were some bromate issues with current ozonation practices at the full-scale (whichcurrently target Giardia inactivation). Specifically, several utilities were forming bromate atlevels in excess <strong>of</strong> the Stage 1 D/DBP Rule MCL <strong>of</strong> 10 ug/L; 11 percent <strong>of</strong> the samples analyzedwould be considered out <strong>of</strong> compliance. Future compliance may be difficult for a greaternumber <strong>of</strong> utilities when operating at disinfection levels capable <strong>of</strong> Cryptosporidiuminactivation.BROMATE FORMATION AND MINIMIZATIONThe background natural organic matter <strong>of</strong> 14 natural waters was characterized by measurements<strong>of</strong> the DOC, UV254 absorbance <strong>and</strong> UV2oo-soo absorbance. The NOM was also fractionated intohydrophobic, hydrophilic <strong>and</strong> transphilic fraction using macroporous, nonionic resins. Sizeexclusion chromatography was also used to determine the apparent molecular weight <strong>of</strong> thenatural organic matter. These various parameters were then statistically correlated to bromateformation under a set <strong>of</strong> st<strong>and</strong>ard ozonation conditions (true-batch ozone reactor; temperature =20°C; ozone dose = 1 mg ozone per mg DOC). For the 14 waters studied, the DOCconcentrations ranged from 1.4 to 7.3 mg/L <strong>and</strong> had a median bromide concentration <strong>of</strong> 42 ug/L.The SUVA values ranged from 1.4 to 3.8 L/mg-m. The average hydrophobic fraction accountedfor 50 percent <strong>of</strong> the NOM, with the transphilic <strong>and</strong> hydrophilic fractions averaging 21 <strong>and</strong> 29percent <strong>of</strong> the NOM, respectively. While all <strong>of</strong> the NOM fractions exerted an ozone dem<strong>and</strong>, thehydrophobic fraction was found to exert the highest dem<strong>and</strong> per unit DOC. Based on thest<strong>and</strong>ard bromate formation testing the hydrophobic fraction was negatively correlated withbromate formation, while the transphilic <strong>and</strong> hydrophilic fractions were positively correlated.<strong>Bromate</strong> formation in these waters was evaluated over a range (0.5-log to 3.0-logs) <strong>of</strong> predictedCryptosporidium inactivation levels. At 1-log Cryptosporidium inactivation, approximately half<strong>of</strong> the waters were expected to produce bromate in excess <strong>of</strong> 10 ug/L. It was noted that whilethese waters had a range <strong>of</strong> water qualities <strong>and</strong> bromide concentrations, the higher theconcentration <strong>of</strong> bromide the higher the percent <strong>of</strong> the bromide that was converted to bromate.To reduce the concentration <strong>of</strong> bromate formed, pH depression, ammonia addition <strong>and</strong> hydroxylradical scavenger addition were evaluated. By depressing the pH from 8 to 7 to 6, a generalreduction in bromate formation <strong>of</strong> 30 to 50 percent per unit decrease in pH was observed. Thisreduction in bromate formation resulted in an increase in the concentration <strong>of</strong> total organicxxn


omide formed. The addition <strong>of</strong> 0.5 mg/L ammonia-nitrogen to these waters resulted in areduction in bromate formation for all <strong>of</strong> the waters that would normally form over 10 ug/L <strong>of</strong>bromate. Increasing the ammonia concentration to 1.0 mg/L ammonia-nitrogen resulted in someadditional reduction in bromate formation. It was estimated that ammonia addition might be aneffective bromate formation control strategy for those waters within 10 to 20 |ig/L <strong>of</strong> the MCL.It was also revealed that the cumulative effects pH depression together with ammonia additioncould be used for waters in which bromate formation was particularly problematic.A preliminary investigation <strong>of</strong> the addition <strong>of</strong> a hydroxyl radical scavenger revealed the addition<strong>of</strong> 1 mM <strong>of</strong> t-butanol prevented bromate formation. While the addition <strong>of</strong> t-butanol to municipaldrinking water supply might not currently be an acceptable practice, it does illustrate theeffectiveness <strong>of</strong> this approach. This illustrates the need for additional research to identify anacceptable hydroxyl radical scavenger <strong>and</strong> determine the minimum amount necessary toeffectively inhibit bromate formation. It was just as important to note that since the addition <strong>of</strong> ahydroxyl radical scavenger completely inhibited bromate formation, it appeared that bromateformation in natural waters does not proceed through the "direct pathway" as discussed by Songet al. (1997), but must proceed through pathways in which hydroxyl radicals are needed.Consequently, the radical scavenging by natural organic matter likely plays a significant role inthe differences observed in bromate formation between natural waters.The temperature was observed to reduce bromate formation in true-batch, semi-batch <strong>and</strong>continuous-flow laboratory reactors. However, its impacts on the amount <strong>of</strong> bromate formed fora given level <strong>of</strong> Cryptosporidium disinfection varied between these reactors. In the true-batchexperiments, decreasing the temperature from 20°C to 10°C resulted in about a 40 percentreduction in the bromate formed for 2-logs <strong>of</strong> Cryptosporidium inactivation. Using thelaboratory-scale continuous flow reactor, decreasing the temperature from 25°C to 5°C, resultedin a 15 to 72 percent increase in the amount <strong>of</strong> bromate formed for 2-logs <strong>of</strong> Cryptosporidiuminactivation. An approximate 50 percent increase in the amount <strong>of</strong> bromate formed for 2-logs <strong>of</strong>Cryptosporidium inactivation was also observed in a semi-batch reactor as the temperature wasdecreased from 22°C to 7°C. It is unclear why different results were observed for these differentreactors.HYDRODYNAMIC IMPACTSIn addition to chemical approaches to bromate minimization, hydrodynamic strategies were alsoinvestigated. However, through bench-scale <strong>and</strong> pilot-scale experiments, no compelling reasonwas identified to employ staged/tapered ozonation for bromate minimization. It was alsodiscovered that bromate formation could not be reduced by operating the ozone contactors ineither a co-current or counter-current mode. Nevertheless, the operation <strong>of</strong> the ozone contactorin a co-current mode might have provided subtle opportunities for optimizing disinfection.These tests provided additional indications that bromate formation in natural waters ispredominantly controlled by chemical conditions <strong>and</strong> not hydraulic parameters.xxin


COMPARISON BETWEEN REACTOR SCALESMany <strong>of</strong> the bromate formation trends were developed through bench-scale tests. Concerns overthe applicability <strong>of</strong> these results to larger scale (pilot-scale <strong>and</strong> full-scale) reactors wereaddressed through a series <strong>of</strong> comparability tests. It was shown that bench-scale reactors couldprovide reasonable simulations <strong>of</strong> pilot-scale <strong>and</strong> full-scale bromate formation results providedan accurate estimate <strong>of</strong> the amount <strong>of</strong> ozone contact at the larger scale could be generated. Assuch, the approximation <strong>of</strong> ozone contact by CT, as defined by the SWTR, resulted in a lack <strong>of</strong>comparability. By calculating ozone contact with the hydraulic retention time instead <strong>of</strong> tio, <strong>and</strong>by giving credit to the first cell <strong>of</strong> the pilot-scale <strong>and</strong> full-scale reactors correlations to the benchscaleresults could be developed. Likewise, pilot-scale simulations could be used to provideapproximations <strong>of</strong> full-scale results. Such simulations could be useful in assessing trade<strong>of</strong>fsbetween bromate formation <strong>and</strong> disinfection (CT) under various scenarios. However, theseresults also illustrated that while the simulations could be used for trending purposes, the specificbromate concentrations formed for given amounts <strong>of</strong> ozone contact do not always match betweendifferent contactors. The results from these tests also lent additional credence to the theory thatchemical conditions, as opposed to hydraulic conditions, control the formation <strong>of</strong> bromate.INTEGRATED MODELThe results <strong>of</strong> these reactor scale comparability tests were also used to validate the performance<strong>of</strong> a series <strong>of</strong> integrated models. These models could be used to predict ozone decomposition,bromate formation <strong>and</strong> Cryptosporidium inactivation, <strong>and</strong> were presented for four different types<strong>of</strong> reactors: true-batch ozone reactor, laboratory-scale continuous-flow ozone contactor withinternal recirculation, pilot-scale ozone contactor <strong>and</strong> full-scale ozone contactor. Pr<strong>of</strong>iles <strong>of</strong>ozone concentration, bromate formation <strong>and</strong> Cryptosporidium inactivation through the BritanniaWater Purification Facility pilot-scale ozone contactor <strong>and</strong> the Los Angeles Aqueduct FiltrationPlant full-scale ozone contactor revealed similar trends to the experimental data. It wasrecognized that while limitations existed, the model simulations could be used to underst<strong>and</strong> theperformance <strong>of</strong> the ozone contactors at various configurations <strong>and</strong> operating conditions.xxiv


CHAPTER 1. INTRODUCTION1.1 REGULATORY BACKGROUNDIn 1992 the Environmental Protection Agency (EPA) initiated a negotiated rulemaking processthat eventually led to the proposed Stage 1 Disinfectants <strong>and</strong> Disinfection By-Products (D/DBP)Rule (Federal Register, July 29 1994). Included in the proposed Stage 1 D/DBP rule was aproposed maximum contaminant level (MCL) for bromate (BrCV) <strong>of</strong> 10 ug/L <strong>and</strong> an MCL Goal(MCLG) <strong>of</strong> zero. Based on available information, EPA determined that the theoretical 10cancer risk estimate for bromate was 5 ug/L. The proposed MCL <strong>of</strong> 10 ug/L, therefore,represented a risk level higher than EPA's approach <strong>of</strong> regulating carcinogens within the 10"4 to10"6 risk range. In part, this decision was driven by available analytical methods that EPAconsidered, which had a practical quantification limit (PQL) <strong>of</strong> 10 |ig/L. In the preamble to theproposed Stage 1 D/DBP Rule, EPA requested public comment on whether the PQL for bromatecould be lowered <strong>and</strong> whether a lower MCL could be established, hi 1997, EPA formed a newAdvisory Committee <strong>and</strong> the consensus from that committee was that there was no newinformation currently available to justify lowering the MCL. Consequently, the final bromateMCL was promulgated at 10 ng/L (Federal Register, December 16,1999).While the 10 |ng/L is anticipated to impact a limited number <strong>of</strong> utilities currently using ozone asthe primary disinfectant to inactivate Giardia <strong>and</strong> viruses, a greater number <strong>of</strong> utilities will beimpacted by this MCL when compliance with the Long-Term 2 Enhanced Surface WaterTreatment Rule (LT2ESWTR) is required. In September 2000, an EPA Federal AdvisoryCommittee signed an agreement in principle for recommendations <strong>of</strong> the LT2ESWTR. In thisagreement, additional treatment requirements for reductions in Cryptosporidium are identified.Beyond the Cryptosporidium reduction requirements identified in the Interim Enhanced SurfaceWater Treatment Rule, up to an additional 2.5-logs <strong>of</strong> Cryptosporidium reduction may berequired based on source water monitoring. Ozone was one <strong>of</strong> the treatment tools identified thatcould be used to inactivate Cryptosporidium. While the Cryptosporidium inactivation tableshave not yet been established by the EPA, preliminary information suggests CT values 5 to 30times higher than that required for Giardia inactivation will be required (Rennecker et al. 1999;Oppenheimer et al., 2000).1.2 OZONE DOSES FOR CRYPTOSPORIDIUM INACTIVATION1.2.1 Semi-Batch Cryptosporidium parvum InactivationThe kinetics <strong>of</strong> Cryptosporidium parvum inactivation with ozone have been investigated byseveral research groups (Peeters et al., 1989; Perrine et al, 1990; Korich et al., 1990; Finch etal., 1993; Liyanage et al, 1997; Rennecker et al, 1999) in batch or semi-batch reactors. Untilrecently, there has been lack <strong>of</strong> consistency among the various results reported presumably due


to disparities in the methods used to assess oocyst viability. In-vitro excystation methods havetypically produced slower inactivation rates than those based on animal infectivity assays (Finchet al., 1993; Owens et al., 1994). However, a modified in-vitro excystation approach based onsporozoites enumeration has been shown to be consistent with animal infectivity data presentedin the literature for the same C. parvum strain (Rennecker et al., 1999). The CTs required for0.5-log to 3-logs inactivation <strong>of</strong> Iowa strain C. parvum oocysts in distilled-deionized water forthe temperature range from 0.5°C to 30°C obtained by modified in-vitro excystation (Renneckeret al,. 1999) are presented in Table 1.1.Table 1.1: Proposed CTs Required for the Inactivation <strong>of</strong> Cryptosporidium parvum oocystsWith Ozone Applied to a Semi-Batch ReactorTemperature(°C)-Log(AWV0)0.51.01.52.02.53.00.51932445668805111825313845Source: Adapted from Rennecker et al., 199910 15 20CT [mg-min/L]5.9 3.2 1.89.5 5.2 2.913 7.3 4.117 9.3 5.221 11 6.324 13 7.51.2.2 Multi-Utility Cryptosporidium Inactivation StudyOppenheimer et al. (2000) have extended the Cryptosporidium inactivation findings <strong>of</strong> otherresearchers obtained for ozone in laboratory-grade water <strong>and</strong> static- or semi-batch reactors, tonatural water conditions. For this study, sixteen different North American <strong>and</strong> European naturalwaters were seeded with viable Cryptosporidium parvum (C. parvum) oocysts. Inactivation wasthen measured by animal infectivity following the application <strong>of</strong> different disinfectants ordisinfectant combinations. The results presented here focus on the ozone disinfection portion <strong>of</strong>the study. All <strong>of</strong> the C. parvum inactivation experiments by ozone were conducted using atemperature controlled continuous flow contacting system that was designed to simulate thehydraulic conditions <strong>of</strong> full-scale ozone contacting systems.Analysis <strong>of</strong> data from these experiments showed that modeling the results by the linear ChickWatson model, the non-linear Chick Watson model or the Horn model, yielded similar modelfits. As a result the simple, one-parameter, linear Chick Watson model was used to compare theresults obtained for each <strong>of</strong> the water sources. Using the linear Chick Watson inactivationconstants derived for several <strong>of</strong> the project waters at more than one experimental temperature, itwas calculated that the CT requirements more than quadrupled (4.6 multiplier) for every 10°Cdecrease in temperature. This impact <strong>of</strong> temperature on ozone inactivation <strong>of</strong> C. parvum may beseverely underestimated if the temperature corrections used for the ozone inactivation <strong>of</strong> Giardialamblia are used (i.e. a doubling <strong>of</strong> the CT requirements for every 10°C decrease in temperature).The results obtained were used to generate a set <strong>of</strong> CT values for Cryptosporidium parvuminactivation between 0.5-log <strong>and</strong> 3.0-logs at temperatures ranging from 1 to 30°C (see Table1.2).251.01.72.33.03.64.3300.61.01.41.72.12.5


Table 1.2: Proposed CTs Required for the Inactivation <strong>of</strong> Cryptosporidium parvum OocystsWith Ozone Applied to a Continuous-Flow ReactorTemperature(°C)-Log(AWVo)0.51.01.52.02.53.0. 111213242536355.71117232934Source: Adapted from Oppenheimer et al., 200010 15 20CT [mg-min/L]2.7 1.2 0.65.3 2.5 1.28.0 3.7 1.711 5.0 2.313 6.2 2.916 7.5 3.5Obvious differences exist between the data presented in the literature, <strong>and</strong> numerous hypotheseswhy these differences exist. As the specific levels <strong>of</strong> ozone CT required to achieve a specificlevel <strong>of</strong> Cryptosporidium inactivation was not the focus <strong>of</strong> this study, the differences between theproposed data presented above <strong>and</strong> in the literature will not be debated here. For the purposes <strong>of</strong>this study, a range <strong>of</strong> CT values will be used to illustrate the impacts <strong>of</strong> the increased levels <strong>of</strong>ozone exposure. Additionally, the modeling efforts presented in this study utilized the highervalues proposed by Rennecker et al. (1999) as a conservative estimate <strong>of</strong> the levels <strong>of</strong> CTrequired for Cryptosporidium inactivation.1.3 OZONE REACTION CHEMISTRY BACKGROUNDThe move to non-chlorine disinfection processes has not eliminated the production <strong>of</strong>halogenated by-products. Much <strong>of</strong> the recent attention has focused on the presence <strong>of</strong> bromateion resulting from alternative disinfection processes, including but not limited exclusively toozonation (Siddiqui et al., 1995). Considerable effort has been directed towards investigatingbromate formation during the ozonation <strong>of</strong> bromide (Br~) containing waters. A number <strong>of</strong>models have been proposed for the corresponding reaction mechanisms (Yates <strong>and</strong> Stenstrom,1993; von Gunten <strong>and</strong>Hoigne, 1993, 1995; Westerh<strong>of</strong>f, 1995; Song, 1996; Song et al, 1997).<strong>Bromate</strong>, formed during the ozonation <strong>of</strong> source waters containing bromide ion (bromide), hasbeen reported to be affected by both water quality <strong>and</strong> treatment conditions. The water qualityconditions that have been reported as having either a positive or negative effect on bromateformation have included: natural organic matter (NOM); bromide; pH; carbonate alkalinity;temperature; <strong>and</strong> ammonia-nitrogen (NHs-N). Influential treatment conditions include:transferred ozone (Os) dose; hydraulic residence time (HRT); dissolved ozone residual (DOs);CT; acid addition; <strong>and</strong> ammonia addition.<strong>Bromate</strong> formation has been well defined in laboratory water (Milli-Q water: deionized, carbonfiltered,0.45 um filtered water) by recent efforts. These studies have shown that bromateformation resulted from a complex combination <strong>of</strong> molecular ozone <strong>and</strong> free radical mechanismsinitiated by the hydroxyl radical (-OH) formed through ozone decomposition. Experimentalresults were explained quite well with a deterministic modeling approach (Song, 1996). Asshown in Figure 1.1, Song et al. (1997) identified three distinct pathways with prominent250.30.50.81.11.41.6300.10.30.40.50.60.8


contributions to bromate formation in laboratory water. These pathways included a truemolecular ozone route <strong>and</strong> two combination pathways, one initiated by molecular ozone <strong>and</strong> theother initiated by the hydroxyl radical.OBrVHOBr < °3— Br" ——^1—* OBrVHOBrOH I OHOBr . Br BrO2"Disproportionation Q3BKV OBr Br°BrOs" Br02-DisproportionationBrO3'Direct/Indirect Indirect/Direct DirectPathway Pathway PathwaySource: Song et al. 1997. Reprinted from Journal AWWA, Vol. 89, No. 6 (June 1997), bypermission. Copyright ©1997, American Water Works AssociationFigure 1.1: <strong>Bromate</strong> <strong>Formation</strong> PathwaysThe deterministic approach has been complicated in natural waters by the presence <strong>of</strong> organicsubstances. Organic compounds become involved in the ozone <strong>and</strong> free radical reactions, as wellas the bromine reaction pathway through the formation <strong>of</strong> organic bromine compounds.Consequently, in addition to bromate formation concerns, the background organic substancesmay also provide a sink for bromide. Based on results with radical scavengers in the presence <strong>of</strong>NOM, the true molecular ozone pathway loses significance or is eliminated (Song, 1996).Several researchers have recently examined the influences <strong>of</strong> water quality <strong>and</strong> water treatmentvariables on bromate formation with a focus on process modifications to minimize bromateformation during ozonation (Amy et al., 1997; Glaze et al., 1993; Krasner et al., 1993). Thiswork has led to the observation that in many cases, there is a trade-<strong>of</strong>f between bromatereduction <strong>and</strong> increases in total organic bromide (TOBr) formation. Furthermore, thecomposition <strong>of</strong> the TOBr formed has been poorly characterized which has hampered efforts toassess the toxicity <strong>of</strong> this group <strong>of</strong> brominated DBFs (Bull et al., 1995). It has beendemonstrated by Song (1996), that TOBr may form rapidly from hypobromous acid (HOBr). Inaddition, the production <strong>of</strong> HOBr from ozonation is also rapid <strong>and</strong> the resultant HOBr can persistlong after the ozone has dissipated. This has been indirectly demonstrated through the cessation<strong>of</strong> bromate formation with the depletion <strong>of</strong> the residual ozone concentration, while TOBrcontinues to form thereafter (Westerh<strong>of</strong>f, 1995; Song, 1996).


1.4 INFORMATION GAPSUntil now, an ozone dose <strong>of</strong> 1 mg Os for every mg dissolved organic carbon (DOC) hasgenerally been accepted by the water industry as an upper limit for meeting Giardia <strong>and</strong> virusinactivation requirements under the Surface Water Treatment Rule. At these ozone doses, waterswith low concentrations <strong>of</strong> bromide (e.g. < 50 ug/L) generally would meet a 10 ug/L bromateMCL. However, most <strong>of</strong> the research on underst<strong>and</strong>ing bromate formation <strong>and</strong> control hasutilized source waters containing medium to high levels <strong>of</strong> bromide (>100 |o.g/L) ozonated at lowto medium Oa doses (


1.5 PROJECT OBJECTIVESThe primary objective <strong>of</strong> this research was to evaluate the formation <strong>of</strong> bromate <strong>and</strong> the efficacy<strong>of</strong> control strategies for low to moderate (


CHAPTER 2. MATERIALS AND METHODSTo accomplish the objectives <strong>of</strong> the research, many analytical methods <strong>and</strong> experimentalprocedures were employed. Analytical methods were used to assess organic <strong>and</strong> inorganic waterquality parameters. Experimental procedures were used to operationally characterize NOM, <strong>and</strong>evaluate bromate formation under true-batch ozonation, staged ozonation, <strong>and</strong> temperaturecontrolledozonation conditions. <strong>During</strong> the course <strong>of</strong> research, the methods <strong>and</strong> procedureswere subjected to quality control to identify variability/error within the analytical <strong>and</strong>experimental results.2.1 SOURCE WATERSFourteen different North American <strong>and</strong> European natural water sources were utilized for thebench-scale, pilot-scale <strong>and</strong> full-scale tests performed as a part <strong>of</strong> this study. These surfacewater, groundwater, <strong>and</strong> surface/groundwater mixed sources included the following:• Lake Houston (HOU), Houston, Texas• Colorado River (CRW), La Verne, California• Los Angeles Aqueduct (LAW), Los Angeles, California• Marne River (CGE), Paris, France• Delaware River (NJA), New Jersey• Lake Meredith <strong>and</strong> Ogallala Aquifer mix (AMA), Amarillo, Texas• Trinity River-Ehn Fork (DAL), Dallas, Texas• State Project (SPW), La Verne, California• Sacramento River (SAC), California• Sacramento/San Joaquin Delta (CCD), Concord, California• Huron River (ANN), Ann Arbor, Michigan• Ottawa River (OTT), Ottawa, Canada• South Bay Aqueduct (ACD), Fremont, California• Floridian Aquifer (WPB), West Palm Beach, FloridaSource water quality parameters are summarized in Table 2.1. All waters were untreated, exceptfor the DAL <strong>and</strong> OTT samples, which were taken immediately prior to the point <strong>of</strong> intermediateozonation within municipal drinking water treatment facilities.2.2 REACTORSThroughout this study, several different types <strong>and</strong> scales <strong>of</strong> reactors were utilized. A description<strong>of</strong> each type <strong>of</strong> reactor <strong>and</strong> its operation is given below.


Table 2.1: Source Water Quality ParametersID<strong>Bromide</strong>AmbientpHAlkalinity(mg/L as CaCO3)Ammonia-N(mg/L)DOC(mg/L)UVA254(cm" 1 )SUVA(L/mg-m)AMAWPBSPWCRWANNACDCCDLAWCGEHOUNJADALOTTSACMeanMedianStd DevRange200170145726555393323212121178.26436628.2 - 2007.97.07.48.18.47.58.17.97.57.36.98.16.17.07.57.50.66.1 - 8.41832307213033151821102102838754567104796728 - 2300.020.770.030.060.140.070.060.070.010.20.230.530.060.090.170.070.220.01 - 0.772.810.42.82.72.53.31.81.91.59.43.04.42.61.73.62.82.81.5 - 10.40.0400.3900.0670.0440.0480.0710.0620.0440.0370.3300.0800.0890.0560.0440.1000.0590.1120.037 -0.3901.43.82.41.61.92.23.42.32.53.52.72.02.12.62.52.40.71.4-3.82.2.1 True-Batch Reactors2.2.1.1 True-Batch <strong>Ozonation</strong>The true-batch ozonation procedure consisted <strong>of</strong> the transfer <strong>of</strong> an aliquot <strong>of</strong> ozone stock solutioninto a completely-mixed batch reactor, resulting in a 100 percent transfer efficiency <strong>of</strong> aqueousozone to the sample. The addition <strong>of</strong> ozone stock, however, incurs a dilution <strong>of</strong> the originalsample <strong>and</strong> alters the concentrations <strong>of</strong> the solutes in the sample. This batch reactor simulatesthe conditions <strong>of</strong> a plug-flow reactor (PFR). By modeling kinetic data derived from the truebatchreactor over time, concentration versus distance pr<strong>of</strong>iles for plug-flow reactors could bedeveloped.The true-batch reactor consisted <strong>of</strong> a 500-mL glass graduated cylinder fitted with a glasssampling port at the bottom (Figure 2.1). Experiments were carried out by initially buffering(1.0 mM PC'43-) <strong>and</strong> adjusting the pH <strong>of</strong> a natural water sample to 7.0 ±0.1. Calculated volumes<strong>of</strong> the prepared natural water <strong>and</strong> ozone stock solution were sequentially added to the reactor.Ozone stock was transferred to the reactor via a glass pipette <strong>and</strong> nalgene tubing. The totalvolume added to the reactor was 500 mL, with an ozone to DOC dose <strong>of</strong> Img O3/mg DOC. Anadjustable Teflon disc was then placed on the top <strong>of</strong> the water surface to prevent gas transfer at


the air-water interface (i.e., no headspace). Ozone decomposition kinetics were monitored byanalyzing aliquots <strong>of</strong> the solutions for the residual dissolved ozone concentration atpredetermined times. A magnetic stir bar <strong>and</strong> plate were used to provide continuous mixing <strong>of</strong>the ozonated sample.Teflon DiskOzone StockRaw Water Sample(adj usted/bufferedto pH = 7.0)Stir BarFigure 2.1: True-Batch <strong>Ozonation</strong> Equipment SetupOnce the measured ozone concentrations were observed to decay below 0.05 mg/L, the Teflondisc was removed from the reactor <strong>and</strong> the sample was allowed to rapidly mix for a period <strong>of</strong> 5to 10 minutes to dissipate the remaining residual ozone concentration. Aliquots <strong>of</strong> the ozonatedsample were then collected in 40-mL amber vials, labeled <strong>and</strong> stored at 4°C until they wereanalyzed for bromate concentration.While the majority <strong>of</strong> these tests were performed at 20°C, selected true-batch experiments wereperformed at 10 ± 1°C to assess the effects <strong>of</strong> low temperature on bromate formation. For theseexperiments, the temperature was controlled by recirculating chilled water through a water jacketsurrounding the glass reactor.The addition <strong>of</strong> the ozone stock solution to samples <strong>of</strong> natural water resulted in dilutions <strong>of</strong> theoriginal water quality parameters summarized in Table 2.1. The indicated dilution percentages<strong>and</strong> "diluted" water quality characteristics are summarized in Table 2.2 for bromide, DOC,ammonia, <strong>and</strong> alkalinity. The parameter pH was not shown since all true-batch ozonation


experiments were performed at a constant buffered pH (7.0 ±0.1). Subsequent data analyseswere based on these diluted water quality characteristics.Table 2.2: Source Water Quality Parameters After Dilution with Stock Ozone SolutionIDAMAWPBSPWCRWANNACDCCDLAWCGEHOUNJADALOTTSACMeanMedianStd DevRange<strong>Bromide</strong> Alkalinity Ammonia(ug/L) (mg/L as CaCO3) (mg/L)182119131666048373022151818154758425115-182 167161651203013277101200203365416391715720 - 200 0.020.540.030.060.130.060.060.060.010.140.200.460.050.080.140.0630.160.01 - 0.54* Based on assumption <strong>of</strong> Beer's Law applicabilityDOC(mg/L)2.557.282.522.482.312.881.701.751.436.772.643.782.341.603.02.51.81.4-7.3UVA254(cm' 1 )*0.0360.2730.0600.0400.0440.0620.0590.0410.0350.2340.0700.0760.0500.0410.0800.0600.0720.036 - 0.273Dilution(%)8.730.09.77.77.713.06.18.24.828.013.013.08.96.0129.37.54.8 - 30Ozone stock solutions were prepared in Milli-Q water. To do this, a stream <strong>of</strong> ozone gasgenerated (OREC: Model O3V5-O, Phoenix, AZ) from medical grade oxygen, was continuouslypassed through ice-cooled Milli-Q water for a minimum <strong>of</strong> 120 minutes. The resultant aqueousozone concentrations ranged from 20 to 35 mg/L.2.2.1.2 Staged-Batch <strong>Ozonation</strong>Staged true-batch ozonation consisted <strong>of</strong> applying aliquots <strong>of</strong> ozone stock solution in two stagesto assess the difference in bromate formed as compared to the single-stage true-batch ozonationexperiments. In this procedure, an aliquot <strong>of</strong> ozone stock solution was added to the sampleduring the first stage at a 0.5:1 ozone to DOC mass ratio. As in the procedure described above,ozone decay was tracked using ozone residual measurements at timed intervals. Once the ozoneresidual had decayed to below detection limits, the sample was allowed to continue to react for aperiod <strong>of</strong> 5 to 10 minutes before the samples were collected in 40-mL amber vials.10


For the second stage, the new DOC concentration was calculated based on the diluted DOC <strong>of</strong>the sample after the addition <strong>of</strong> ozone stock in stage one. An aliquot <strong>of</strong> ozone stock solutionswas once again added at a dose <strong>of</strong> 0.5:1 ozone to DOC (mg/mg). As with the first stage, ozoneresiduals were recorded over time. Once the system had been allowed to react for 5 to 10minutes after the ozone residual had decayed to below detection limits, samples were collected inamber vials, labeled <strong>and</strong> stored at 4°C until the bromate analyses could be performed.2.2.2 Semi-Batch ReactorThe inactivation kinetics <strong>of</strong> C. parvum oocysts <strong>and</strong> selected bromate formation results withOttawa River water were determined by performing semi-batch experiments. Similar proceduresto those described by Hunt <strong>and</strong> Marinas (1997) were followed. A reactor containing 200 mL <strong>of</strong>0.01M phosphate buffer solution at pH 6.5, 7.5, or 8.5 was immersed in a 20°C water bath. Aconcentration <strong>of</strong> dissolved ozone inside the reactor was kept constant by continuous bubbling <strong>of</strong>ozone gas. The ozone gas, generated (PCI Ozone Corp: Model GL-1, NJ) from 99.9% pureoxygen, was passed through an 8-L reservoir bottle to dampen any fluctuations in ozoneconcentration during the experiment. When the dissolved ozone concentration inside the reactorreached a steady value <strong>of</strong> about 0.5 mg/L, approximately 7xl05 oocysts were injected. Eachinactivation experiment was stopped after a predetermined contact time to target a specific CTvalue by adding a 2-mL aliquot <strong>of</strong> a 1.67 percent (w/v) sodium thiosulfate solution to the reactor.This procedure was repeated at various CTs to cover an inactivation range between 0 <strong>and</strong> 99.5percent.2.2.3 Laboratory-Scale Continuous-Flow Ozone ContactorA number <strong>of</strong> experiments were performed on each source water in a laboratory-scale continuousflowozone contactor to develop an orthogonal matrix to evaluate bromate formation under avariety <strong>of</strong> pH, temperature <strong>and</strong> ammonia conditions. This same contactor was used in a limitednumber <strong>of</strong> C. parvum inactivation experiments. A subset <strong>of</strong> radical scavenger additionexperiments was completed using two concentrations <strong>of</strong> tertiary butanol. Two slightly differentorthogonal matrices were employed. The first matrix, as shown as Table 2.3, was used for LakeHouston water (HOU), Los Angeles Aqueduct water (LAW), Delaware River water (NJA),Trinity River water (DAL), Sacramento/San Joaquin Delta water (CCD), Huron River water(ANN), Ottawa River water (OTT), <strong>and</strong> Floridian Aquifer water (WPB). The correspondingmatrix for Colorado River water (CRW), State Project water (SPW), Sacramento River water(SAC), <strong>and</strong> South Bay Aquifer water (ACD) is listed in Table 2.4.Table 2.3: Experimental Matrix #1Parameter_________________Value (* denotes baseline value)______pH 6, 7*, 8Temperature 5,15*,25°CAmmonia Ambient*, 0.5,1.0 mg/L-NLog Inactivation_____________1,2*, 3 log__________________________11


Table 2.4: Experimental Matrix #2Parameter________________Value (* denotes baseline value)pH 6, 7, 8*Temperature 5,15*, 25°CAmmonia Ambient*, 0.5, l.Omg/L-NLog Inactivation_____________0.5, 1.0*, 1.5 log_________2.2.3.1 Sample PreparationFor all <strong>of</strong> these tests, each natural water sample was initially adjusted to the desired temperature,<strong>and</strong> then pH. For the two ammonia addition experiments with each water, enough ammoniumchloride (NILjCl) was added to achieve total ammonia concentrations <strong>of</strong> 0.5 <strong>and</strong> 1.0 mg/L as N.The ammonia concentration was measured immediately before ammonia addition by the methoddescribed in Section 2.4.8. For the radical scavenger addition experiments, tertiary butanol wasadded to achieve concentrations <strong>of</strong> 1 or 3 mM.2.2.3.2 ApparatusThe continuous-flow ozone contactor consisted <strong>of</strong> several components: bubble contactor, simpledosing column, ozone generator, sample feed pump, sample recirculation pump, water bath,back-flow catch flask, <strong>and</strong> KI trap. A schematic <strong>of</strong> the system is shown in Figure 2.2. Thecontactor was 1.2m long <strong>and</strong> had an inner diameter <strong>of</strong> 0.02 m. The contactor column <strong>and</strong>internal recirculation line had volumes <strong>of</strong> 440 mL <strong>and</strong> 85 mL, respectively. A water jacketsurrounding the contactor was used for temperature control. A glass-frit gas-diffuser placed atthe bottom <strong>of</strong> the contactor was secured in place by a plastic fitting with a Teflon o-ring.Another glass elbow fits into the top <strong>of</strong> the column. This elbow simply accepts the <strong>of</strong>f-gas fromthe column. The simple dosing column was a 1.2-m long glass tube with an inner diameter <strong>of</strong>0.02 m. A glass diffuser <strong>and</strong> <strong>of</strong>f-gas elbow identical to those for the contactor column fit into thebottom <strong>and</strong> top <strong>of</strong> the simple column. The simple dosing column was used only for calibration<strong>of</strong> the system.In this reactor bulk, undiluted water samples were continuously passed through <strong>and</strong> ozonated bya co-current stream <strong>of</strong> bubbles created by a frit at the base <strong>of</strong> the contactor. The ozonated waterwithin the contact column was recirculated at a rate approximately ten times the sample feedrate. A flow meter was used to regulate the amount <strong>of</strong> ozone gas that entered the contactor fromthe ozone generator. Water was pumped into the bottom <strong>of</strong> the reactor <strong>and</strong> exited at the top.The system was allowed to stabilize in terms <strong>of</strong> experimental temperature <strong>and</strong> liquid <strong>and</strong> gaseousflow rates. It was determined that true system stability had been achieved once the ozoneconcentration at each <strong>of</strong> three sample ports along the length <strong>of</strong> the column was the same <strong>and</strong>constant (stable) with time. Once the ozone residual had stabilized, an aliquot <strong>of</strong> the ozonatedwater sample exiting the reactor was quenched with sodium thiosulfate (two times the theoreticalstoichiometric amount) <strong>and</strong> stored in the dark at 4°C until the bromate analyses could beperformed.12


——— Liquid Flow•—••••• Gas FlowO Two way valveO Three way valveI Gas sampling portQ.,Ozone Exhaust KITraplie ^•m........... .1^Water BathtSampleFeed Pump npSampleReservoir"5 -: rS-*-* •*g \t— l ts *il———— !". §feKO ^; ,Q^ ft-.-I' |iri:;l1 "' " •t- •~!- .'T^*$"'$?•':"*iI^''"-


with the full-scale contactor, ozone gas was introduced into the first cell <strong>and</strong> traveled upwards,counter-current to the flow <strong>of</strong> the water. The height <strong>of</strong> the columns was much greater than thecolumn's diameter, creating near plug-flow hydrodynamics.CelllCell 2 Cell 3Cell 4Rapid MixersContactor"Influent.* .v'~y-v'.'«..;-. .•'?*!• :•rf .""' •"' I*."- •' .'-;!.' '•'"•••'-•'«> £ .' 7 ;.•._'.-• •':''.'};.-TapC>TapF4 iL b»PaF >GrhTap!i»&&&&?Tap A4TapDOzone IntroductionContactorEffluentFigure 2.3: Schematic <strong>of</strong> the Full-Scale Ozone Contactor at the Los Angeles AqueductFiltration PlantContactorInfluent«»« >i>^ 1-ContactorEffluent/ ^\/\ /te e0 »°.rn< >Ozone Introductioni»« >4 >•Sample TapLocationFigure 2.4: Schematic <strong>of</strong> Pilot-Scale Ozone Contactor at the Los Angeles AqueductFiltration Plant14


For both the full- <strong>and</strong> pilot-scale contactors, the tio/tso ratio was provided as information aboutthe hydrodynamics <strong>of</strong> the contactors. Because an ideal plug-flow reactor (PFR) has a tio/tso ratio<strong>of</strong> 1.0, it was determined that the pilot-scale system was hydraulically similar a PFR (tio/tso <strong>of</strong>0.87 based on tracer study information). The full-scale contactor, however, hydraulicallybehaved more like a continuous-stirred tank reactor (CSTR) with a tio/tso <strong>of</strong> 0.49.2.2.5 Pilot-Scale Britannia Water Purification Facility Ozone ContactorPilot-scale experiments were also conducted at the Britannia Water Purification Facility (Ottawa,Canada). The ozone contactor consisted <strong>of</strong> three glass columns operated in series (3 m high <strong>and</strong>0.15 m diameter). Ozone could then be selectively applied to either, or both, <strong>of</strong> the first twocolumns, as shown in Figure 2.5. The water used was plant clarified (alum <strong>and</strong> activated silicacoagulation, fiocculation, <strong>and</strong> sedimentation) to represent mid-train ozonation. Experimentswere conducted at both the ambient pH (6.1), <strong>and</strong> at pH 8.5. Elevating the pH to 8.5 wasachieved by adding a caustic (NaOH) solution via a peristaltic pump to the feed linesimmediately upstream <strong>of</strong> the contactor. The alkalinity <strong>of</strong> the pH 8.5 experiments was alsoincreased from the ambient 10 mg/L to 75 mg/L as CaCOs using NaHCOs. CT values rangingfrom 3 to 45 mg-min/L were obtained by varying the water flow rate (8 <strong>and</strong> 5 L/min) <strong>and</strong> ozonegas concentration (0.45 to 2.09 % by weight). Ozone was produced from pure oxygen using aModel GL-1 Generator (PCI Ozone <strong>and</strong> <strong>Control</strong> System Inc., West Coldwell, NJ).Water OutFigure 2.5: Britannia Water Purification Facility Pilot-Scale Ozone Contactor15


Prior to sample collection, the system was allowed to stabilize for a minimum <strong>of</strong> four HRTs.Samples for bromate <strong>and</strong> bromide were collected from ports located at the outlet <strong>of</strong> each column.These samples were quenched with using a 10:1 stoichiometric addition <strong>of</strong> ethylenediaminesolution. Ozone residuals were measured by the indigo method, as described in Section 2.4.10.Samples for determining the dissolved ozone concentration were taken from approximately thebottom, middle, <strong>and</strong> top <strong>of</strong> each column, using a 10-mL bottle-top dispenser attached to Teflontubing submerged in the columns.2.3 CALCULATION OF OZONE CONTACTThe terms ozone exposure (OE) <strong>and</strong> CTwere both used to represent levels <strong>of</strong> ozonation, <strong>and</strong> thedifferences between the two must be clarified. The term OE was used to describe the results <strong>of</strong>the true-batch experiments <strong>and</strong> represented a complete time-based integration <strong>of</strong> the ozone decaycurve. OE was also used to describe the laboratory-scale continuous-flow ozone contactor. Dueto the high rate <strong>of</strong> recirculation, the ozone residual measured at multiple points along the height<strong>of</strong> the contactor were almost identical. As a result, OE was calculated by multiplying the steadystatereactor effluent ozone residual by the theoretical HRT.The term CT was used to describe the results derived from contactors where there was not acomplete spatial pr<strong>of</strong>ile <strong>of</strong> ozone through the contactor, but rather only discrete measurements <strong>of</strong>ozone residuals corresponding to sampling locations. CT values for pilot- <strong>and</strong> full-scalecontactors were calculated according to four different methods: (i) conventional CT (CTswiR-tio)using tio contact time times <strong>and</strong> effluent ozone residual for each cell <strong>of</strong> the ozone contactorexcluding the first cell; (ii) all-cells CT (CTALL-tio) using an "all cells" approach where CT creditfor the first cell (effluent ozone residual multiplied by the first cell tio) was added to the CTswiRtio;(iii) CTswTR-t50 using tso (HRT) instead <strong>of</strong> tio for the corresponding contact time for each cellexcluding the first, multiplied by the corresponding cell's effluent ozone residual; <strong>and</strong> (iv)CTALL-ISO, calculated by summing the tso multiplied by the effluent ozone residual for each cell inthe contactor. The Surface Water Treatment Rule (SWTR) guidelines do not allow for CT creditfrom the first cell (USEPA, 1989; USEPA, 1990), however, CTSWTR-tio is a conservativeindicator <strong>of</strong> ozonation because there is some disinfection occurring within the first cell <strong>of</strong> anozone contactor <strong>and</strong> hence the additional use <strong>of</strong> CTALL-UO- Meanwhile, the use <strong>of</strong> tio versus tso isalso a conservative approach <strong>and</strong> hence the comparative use <strong>of</strong> CTswra-tso <strong>and</strong> CTALL-tso-2.4 ANALYTICAL METHODSAnalytical methods <strong>and</strong> procedures follow the accepted st<strong>and</strong>ards as set forth in St<strong>and</strong>ardMethods <strong>and</strong> Procedures for Analysis <strong>of</strong> Water <strong>and</strong> Wastewater (APHA, AWWA, <strong>and</strong> WEF,1998). All samples to be analyzed for various water quality parameters were stored in amberglass containers at 4°C. All st<strong>and</strong>ards <strong>and</strong> solutions were prepared using reagent-gradechemicals <strong>and</strong> Milli-Q water. Milli-Q water was the product <strong>of</strong> tap water passed through deionizingcartridges (Millipore, Bedford, MA) followed by a carbon adsorber <strong>and</strong> a 0.2-ummembrane filter. This treated water consistently tested at 0.2 mg/L for dissolved organic carbon.16


Adjustments to pH for analytical purposes were made using phosphoric acid or 2.5% sodiumhydroxide.2.4.1 <strong>Bromide</strong> <strong>and</strong> <strong>Bromate</strong><strong>Bromide</strong> <strong>and</strong> bromate samples were prepared for analysis by filtering the samples with 0.2-umsyringe filters. Dionex silver (OnGuard-Ag) <strong>and</strong> hydrogen-exchange (OnGuard-H) filters wereused to reduce chloride interference with bromate detection; however, the filters also removebromide. Thus, separate aliquots <strong>of</strong> the samples were analyzed to determine bromide <strong>and</strong>bromate concentrations. The detection limits for bromate <strong>and</strong> bromide were 2.0 ng/L <strong>and</strong> 10Hg/L, respectively. If the peak for bromate was not detected or unresolved from other peaks, apost-column derivatization method was performed (Echigo et al., 2000) to lower the detectionlimit to 0.3The ion chromatographic conditions for the conductivity method were as follows: analyticalcolumn, lonPac® AS9-HC (4 x 250 mm, Dionex, Sunnyvale, CA); guard column, lonPac® AG9-HC (4 x 50 mm, Dionex); eluent, 9 mM Na2COs; eluent flow rate, 1.0 mL/min; sample loop size,500 ]iL; detector, PED-2 (Dionex); suppressor, ASRS-Ultra (Dionex); suppression mode,external water mode. Samples <strong>of</strong> high chloride or carbonate concentrations, were pretreated withOn Guard-H <strong>and</strong> On Guard-Ag pretreatment cartridges <strong>and</strong> sparged with nitrogen gas for 5 min(at 3 psi).For the post-column derivatization analysis, the same ion chromatographic conditions as theconductivity method were used except for the detection method. The following conditions wereused for the derivatization <strong>and</strong> detection: post-column derivatization reagent, 1.0 M NaBrsolution containing 10 mg/L <strong>of</strong> NOa"; reagent flow rate, 1.0 mL/min; acidification, 1.5 M HaSO^reagent delivery, a gradient pump (GMP-2, Dionex) with 50/50 delivery; reaction coil size, 375uL; reaction temperature 68°C; UV detector path length, 6 mm; detection wavelength, 267 nm.2.4.2 Organic BromineBased on St<strong>and</strong>ard Method 5320 (APHA, AWWA, <strong>and</strong> WEF, 1998), the analysis <strong>of</strong> total organicbromine (TOBr) was measured using a Dohrman Model DX-20A TOX analyzer with a granularactivated carbon (GAC) adsorption module. All the glassware was washed in an acid bath (1:1nitric acid) to minimize background interference, <strong>and</strong> then dried in an oven at 100°C. Tominimize background contamination, the GAC, packing material, <strong>and</strong> packing tools were storedin a desiccator under vacuum, <strong>and</strong> columns were packed in a room <strong>of</strong> relatively low organicbackground level.Sodium thiosulfate was used as the quenching agent for excess bromine. Eliminating theheadspace in the sample vials minimized the loss <strong>of</strong> volatile DBFs. The sample was thenacidified to a pH less than 2 to convert many <strong>of</strong> the acidic constituents <strong>of</strong> TOBr to theirprotinated (non-ionic) form. After acidification, 140 mL <strong>of</strong> sample was passed through twoGAC columns in series. The carbon column was then washed with 5 mL <strong>of</strong> 5,000 mg/L sodiumnitrate solution to remove inorganic halides trapped in the columns. The GAC was subsequently17


combusted at 800PC for 12 minutes. The HBr from the pyrolysis <strong>of</strong> brominated organiccompounds was trapped in a 70% aqueous acetic acid solution, <strong>and</strong> quantified bymicrocoulometric titration, yielding a minimum TOBr reporting level <strong>of</strong> 10 ug/L.2.4.3 Total <strong>and</strong> Dissolved Organic CarbonDissolved organic carbon (DOC) is operationally defined as the portion <strong>of</strong> the total organiccarbon (TOC) that passes through a 0.45 |im filter. Both TOC <strong>and</strong> DOC were analyzed as perSt<strong>and</strong>ard Method 5310C (APHA, AWWA, <strong>and</strong> WEF, 1998), using a Sievers (Boulder, CO) 800portable total organic carbon analyzer. The Sievers analyzer uses the ultraviolet-persulfateoxidation method, yielding measurements <strong>of</strong> total organic carbon, inorganic carbon <strong>and</strong> totalcarbon (sum <strong>of</strong> preceding measurements). TOC <strong>and</strong> DOC were reported in mg <strong>of</strong> carbon perliter (mg-C/L), each with a minimum reporting levels <strong>of</strong> 0.1 mg-C/L.2.4.4 Ultraviolet AbsorbanceUltraviolet absorbance (UVA) at 254 nm was used as a surrogate measurement <strong>of</strong> NOMaromatic character, as it responds to aromatic <strong>and</strong> carbon-carbon double bonds. A Shimadzu(Columbia, MD) UV160U UV-visible recording spectrophotometer was used to analyze UVA254using a 1-cm path length, as described in St<strong>and</strong>ard Method 5910 (APHA, AWWA, <strong>and</strong> WEF,1998). As part <strong>of</strong> the characterization <strong>of</strong> NOM, UVA254 (measured in cm" 1) was measured for0.45 um filtered water samples. In addition, the ultraviolet spectrum from 200 to 400 nm(UV2oo-40o)was recorded to provide an untreated water "signature" to compare with ozonatedwater spectra.Specific UVA (SUVA) is a correlative parameter <strong>of</strong> the UVA normalized by the DOCconcentration <strong>of</strong> the sample, <strong>and</strong> is reported as L/(mg-m). Both UVA <strong>and</strong> SUVA were used toobserve changes in NOM character due to ozonation.In addition, differential UVA254 (AUVA254) <strong>and</strong> differential UVA spectra (AUVA2oo-40o)wereused to characterize the change in NOM as a result <strong>of</strong> ozonation. These different values weredetermined by mathematically subtracting one value (UVA254) or a spectrum (UVA2oo-40o) fromanother, resulting in a positive values when the after-ozone UVA254 value or spectrum (UVA2oo-400) was subtracted from the before-ozone values. A change in UVA spectra implies reactivity <strong>of</strong>the NOM with ozone.2.4.5 NOM FractionationNOM was fractionated using macroporous, nonionic resins (Amberlite XAD-4 <strong>and</strong> XAD-8)according to the method described by Aiken et al. (1988). Filtered (0.45 urn), acidified (pH = 2)samples are passed sequentially through XAD-8 <strong>and</strong> XAD-4 resin columns, at flow rates <strong>of</strong> 1.3to 2.0 mL/min <strong>and</strong> 1.0 to 1.5 mL/min, respectively. These conditions provided an adsorptioncoefficient, k, <strong>of</strong> 50. Samples from the effluent <strong>of</strong> each column were collected after the samplewas processed. Pre-XAD, XAD-8 effluent <strong>and</strong> XAD-4 effluent samples were analyzed forNOM, operationally defined by the DOC results <strong>of</strong> the sample analyses.18


2.4.12 Quality Assurance <strong>and</strong> Quality <strong>Control</strong>Quality assurance (QA) <strong>and</strong> quality control (QC) for the project consisted <strong>of</strong> a series <strong>of</strong>comparative analyses for bromate <strong>and</strong> bromide through a round robin procedure among all <strong>of</strong> theproject research laboratories <strong>and</strong> the internal evaluations <strong>of</strong> replication error for analyses <strong>and</strong>experimental reproducibility.2.4.12.1 Interlaboratory ComparisonsFormal comparisons <strong>of</strong> bromate <strong>and</strong> bromide analysis were administered through a series <strong>of</strong>round robin tests, which served as a quality assurance <strong>and</strong> quality control check between the fouruniversities. After preliminary method comparisons, three comprehensive round robin tests wereconducted during the data collection period <strong>of</strong> this project.For the first two round robin tests, samples were prepared by the University <strong>of</strong> Illinois (UI) <strong>and</strong>sent blindly to the University <strong>of</strong> Colorado (UC), University <strong>of</strong> Toronto (UT) <strong>and</strong> University <strong>of</strong>Poitiers (UP), for bromate <strong>and</strong> bromide analysis. Natural <strong>and</strong> Milli-Q water samples were spikedwith different levels <strong>of</strong> bromide, bromate <strong>and</strong> selected interfering ions. Samples werecategorized in terms <strong>of</strong> initial water matrix, presence <strong>of</strong> interfering ions (chloride - Cl", <strong>and</strong>sulfate - SC


Sample1234567891012345678910Milli-QMilli-Q + SaltsfMilli-QMilli-Q + Salts1Milli-QMilli-Q + Salts1Lake Houston Blank*Lake Houston *Lake Houston*Lake Houston*Milli-QMilli-Q + SaltsfMilli-QMilli-Q + Salts1Milli-QMilli-Q + SaltsfLake Houston Blank*Lake Houston *Lake Houston*Lake Houston*Table 2.6: Second Round-Robin ResultsSpike<strong>Bromate</strong>1.31.36.46.43838-+1.3+6.4+38<strong>Bromide</strong>5515155050-+5+15+501(ug/L)1.21.15.45.6333328252957(Hg/L)5.410549.737356.3191565* Preozonated2'f 50 mg/L Cr + 1 00 mg/L SO42* Insufficient sample volume for analysisUniversity2 3


particular interferences. If interferences were detected, st<strong>and</strong>ard additions were used to minimizethe error <strong>of</strong> analysis.For the samples with low bromate concentrations, less than the minimum reporting level resultsfrom several <strong>of</strong> the labs were considered generally accurate for electrical conductivity detection.For these same samples, the post-column reaction system worked reasonably well.For the third comprehensive round robin test, blind samples were prepared by MontgomeryWatson <strong>and</strong> sent to UI, UC, UT <strong>and</strong> UP for bromate <strong>and</strong> bromide analysis. In addition to theseblind samples, two sets <strong>of</strong> st<strong>and</strong>ard bromate solutions were included by Montgomery Watson.The first series <strong>of</strong> st<strong>and</strong>ard bromate solutions were prepared from known masses <strong>of</strong> sodiumbromate salt. The second set <strong>of</strong> st<strong>and</strong>ard solutions were prepared from a concentrated bromatesolution used in the European Community (EC) interlaboratory bromate comparability study.The universities were asked to calculate the bromate concentrations <strong>of</strong> the blind samples using:1) their own st<strong>and</strong>ard solutions, 2) the Montgomery Watson st<strong>and</strong>ard solutions, <strong>and</strong> 3) the ECst<strong>and</strong>ard solutions. No additional bromide st<strong>and</strong>ard solutions were provided by MontgomeryWatson. The blind samples were Milli-Q water spiked with different levels <strong>of</strong> bromide, bromate<strong>and</strong>/or interfering ions (chloride <strong>and</strong> sulfate).Interlaboratory agreement was generally good for bromate analyses, with <strong>and</strong> without interferingions, even at these low concentrations <strong>of</strong> bromate (Table 2.7 <strong>and</strong> Table 2.8). There were somepreviously acknowledged temporary difficulties with the bromide analysis at the University <strong>of</strong>Illinois during the period that the round robin results were being sent out. It is believed that thisproblem was quickly rectified. <strong>During</strong> this time <strong>of</strong> difficulty, the samples for bromide analysiswere sent from University <strong>of</strong> Illinois to University <strong>of</strong> Colorado. The other three universitiesshowed good interlaboratory agreement for these bromide analyses.Sample Spike(US/L)A 1.4BC 8.3D 5.3E 2.9Table 2.7: Third Round-Robin <strong>Bromate</strong> Results11 MWT EC1.9 2.0 2.0nd nd nd9.4 10.1 9.76.4 6.9 6.63.3 3.6 3.4Univi;rsity22 MWT EC1.0 1.1 1.1nd nd nd7.9 8.3 8.95.4 5.6 6.02.5 2.6 2.8A - Deionized water + 54 mg/L NaCl + 51 mg/L NaSO4B - Deionized water + 54 mg/L NaCl + 51 mg/L NaSO4C - Deionized water + 54 mg/L NaCl + 51 mg/L NaSO4D - Deionized water + 54 mg/L NaCl + 51 mg/L NaSO4E - Deionized waterMontgomery Watson33 MWT EC1.1 1.4 1.3nd nd nd6.8 7.4 7.54.2 4.8 4.22.9 3.2 2.744 MWT EC


2.4.12.2 Individual Research LaboratoriesQuality control was undertaken during the project to assess analytical <strong>and</strong> experimental error.Coefficients <strong>of</strong> variation (CV) were calculated using the means <strong>and</strong> st<strong>and</strong>ard deviations <strong>of</strong>analytical <strong>and</strong> experimental results. Table 2.9 represents the calculated CVs. Analytical errorwas assessed by measuring 10 percent <strong>of</strong> the analyses in triplicate.Sample ID*ABCDETable 2.8: Third Round-Robin <strong>Bromide</strong> ResultsSpikeGig/L)102129-11926409.6ndUniversity2 31123292.11818A - Deionized water + 54 mg/L NaCl + 51 mg/L NaSO4B - Deionized water + 54 mg/L NaCl + 51 mg/L NaSO4C - Deionized water + 54 mg/L NaCl + 51 mg/L NaSO4D - Deionized water + 54 mg/L NaCl + 51 mg/L NaSO4E - Deionized water1224322.0214122127


experiments, a relatively high degree <strong>of</strong> reproducibility was observed for both bromate <strong>and</strong> totalorganic bromine given the inherent difficulties in attaining the exact same OE value <strong>and</strong> otherexperimental conditions.Table 2.10: Experimental Quality <strong>Control</strong> for Laboratory-Scale Continuous-Flow OzoneContactorReplicate12345*6*789101112OE(mg-min/L)112.3102.519.419.657.364.156.866.062.861.863.862.819.620.018.119.160.456.760.364.964.568.720.820.4<strong>Bromate</strong>(Hg/L)21.322.35.75.413.39.14.82.41.21.70.71.138.440.224.123.642.448.313.315.83.03.2116.292.4Total OrganicBromine(Hg/L)789116ND169351334401348109NDND7898NDNDNDND* Note: Replicates 5 <strong>and</strong> 6 were performed on DAL whichhas a high ambient TOX due to chlorinated compounds.2.5 MODEL DEVELOPMENT2.5.1 Cryptosporidium parvum Inactivation KineticsThe Chick-Watson law has been widely used for the mathematical representation <strong>of</strong> thedisinfection kinetics <strong>of</strong> various microorganisms. However, the inactivation kinetics <strong>of</strong> C.parvum oocysts have been shown by some researchers to deviate from the Chick-Watson law.Rennecker et al. (1999) presented C. parvum inactivation curves characterized by an initial lag25


phase during which little or no inactivation occurred followed by a pseudo first-order rate lawinactivation. Several empirical expressions have been suggested for inactivation kinetics thathave an initial time lag (Collins <strong>and</strong> Selleck, 1972; Haas 1980; Severin et al., 1984; Rennecker etal., 1999). The delayed Chick-Watson model (Equation 2.1) proposed by Rennecker et al.(1999) was used in this study for modeling purposes.JLNnNnifCT>CTIag = --(2.1)Where: N/No - survival ratio; N\INo - lag phase factor; fa = pseudo first-orderinactivation rate constant [I/M^T" 1 ]; CT/ag = CT required to overcome theinitial time lag [MTL~3]. The parameters that determine inactivation kinetics,such as Ni/No <strong>and</strong> fa, can be obtained from fitting experimental data obtainedfrom a semi-batch reactor.2.5.2 Ozone Decomposition <strong>and</strong> <strong>Bromate</strong> <strong>Formation</strong> MechanismsOzone in aqueous phase decomposes through a chain <strong>of</strong> reactions including initiation byhydroxide ion attack followed by propagation steps involving superoxide radical, ozonideradical, <strong>and</strong> hydroxyl radical as intermediate species. The ozone decomposition mechanismdeveloped by Staehelin <strong>and</strong> Hoigne (1982), Biihler et al. (1984), <strong>and</strong> Staehelin et al. (1984) wasadopted (Reactions 1-11, Table 2.11) with minor modification. In this study, the species HCV,whose existence was suggested by Staehelin et al. (1984) as an intermediate for the reaction <strong>of</strong>Os <strong>and</strong> -OH ultimately producing HCV <strong>and</strong> 62, was neglected due to the lack <strong>of</strong> evidenceshowing that this hypothetical species exists.Table 2.11: Ozone Decomposition <strong>and</strong> <strong>Bromate</strong> <strong>Formation</strong> MechanismNo. Reaction Rate Constant* ReferencesOzone Decomposition1 O3 + OH' -> O2'- + HO2 -k, = 70Staehelin <strong>and</strong> Hoigne (1982)2 HO2 - O2-> H+pKa = 4.8Staehelin <strong>and</strong> Hoigne (1982)3 O3 + O2-- -> O3"- + O2ka = 1.6x 109 Buhlere/a/.(1984)4 HO3 - H202 + O2k9 = 5 x 109 Staehelin et al. (1984)10 O2'- + HO3 - -» OH' + 2O2k, 0 =l x 10 10 Staehelin et al. (1984)11 HO3 - + HO3 - -» H2O2 + 2O2 k,, = 5x io9 Staehelin et al (1984)(continued)26


Table 2.11: Ozone Decomposition <strong>and</strong> <strong>Bromate</strong> <strong>Formation</strong> Mechanism (cont'd)No. Reaction Rate Constant* ReferencesMolecular Ozone Pathway20 Br" + O3 -» O2 + OBr"k20 =16021 OBr" + O3 -» 2O2 + Brk21 = 33022 OBr" + O3 -> BrO2" + O2k22 =10023 BrO 2" + O 3 -» BrO 3" + O2k23 =lx!0 5Hydroxyl Radical Pathway24a Br + -OH-> BrOH"-24b BrOH"- -» Br" + -OH25a BrOH"- -» Br- + OH"25b Br- + OH" -> BrOH'-26 BrOH"-+ H+ -> Br- + H2O27 BrOH"- + Br" -» Br2"- + OH"28a Br- + Br' -» Br2"-28b Br2'- -> Br- + Br"29 Br2"- + Br2"- -> Br3" + Br"30a Br3" -> Br2 + Br"30b Br2 + fir" -> Br3"31a Br2 + H2O-» HOBr + H+ + Br"3 Ib HOBr + H+ + Br" -» Br2 + H2O32 HOBr BrO" + H+33 O3 + Br- -> O2 + BrO-34 Br- + OBf -> BrO- + Br"35 -OH + OBr" -» BrO- + OH"36 -OH + HOBr -> BrO- + H2O37 O2"- + HOBr -» O2 + Br- + OH"38 2BrO- +H2O->. OBr" +BrO2" + 2iT39 BrO- + BrO2" -» OEf + BrO2 -40 Br2"- + BrO2 " -> OBr" + BrO- + Br"41 -OH + BrO2- -> BrO2 - + OH'42 -OH + BrO2 - -» BrO3 " + H*43a BrO2 - + BrO2 - -> Br2O443b Br2O4 -> BrO2 - + BrO2 -44 Br2O4 +OH" -^ BrO3" + BrO2" + H+Carbonate50 H2CO3* H+ + HC03"51 HCXV


Table 2.11: Ozone Decomposition <strong>and</strong> <strong>Bromate</strong> <strong>Formation</strong> Mechanism (cont'd)No. ReactionPhosphate Buffer60 H3PO4


omate reduction potential <strong>of</strong> adding ammonia or J-butanol was incorporated into the modelwith the Reactions 80-82 <strong>and</strong> Reaction 90, respectively. Finally, the influence <strong>of</strong> natural organicmatter on ozone decomposition <strong>and</strong> bromate formation was empirically formulated throughReactions 100-102, which are modified from those proposed in previous studies (Staehelin <strong>and</strong>Hoigne, 1985; Bezbarua <strong>and</strong> Reckhow, 1996; Westerh<strong>of</strong>f et al, 1997). The rate constants for allthe reactions in Table 2.11 were used as reported in the literature. However, the empiricalparameters that reflect the characteristics <strong>and</strong> the quantity <strong>of</strong> NOM in natural waters, such as a,P, <strong>and</strong> y, were determined from fitting the experimental ozone decomposition <strong>and</strong> bromateformation data collected from the true-batch reactor experiments.2.5.3 Ozone Contactor Modeling2.5.3.1 Batch Ozone ContactorA total <strong>of</strong> N chemical species undergoing chemical reactions in an ideal batch reactor can berepresented mathematically with a system <strong>of</strong> N coupled first-order nonlinear ordinary differentialequations with the general form:=/,.(C 1 ,C 2 ,--.,C JV ) i = l,2,---,N (2.2)Where: C, [ML"3] = concentration <strong>of</strong> chemical species; [M] = mass; [L] = length; t [T] =time;fj = overall rate function corresponding to the relevant chemical reactions.Though commercially available s<strong>of</strong>tware, such as Acuchem, have been widely used for solvingrate expressions for the sets <strong>of</strong> chemical reactions, their application is confined only to the batchreactor configuration. Therefore, s<strong>of</strong>tware that could also be used to represent chemicalreactions in flow-through reactor systems was developed as a part <strong>of</strong> this study. The abovesystem <strong>of</strong> ordinary differential equations was solved by a semi-implicit extrapolation method toaddress the stiffness problem that was experienced when applying explicit methods, such asRunge-Kutta. The solver algorithm was designed to provide an adaptive step size control overthe period <strong>of</strong> integration in order to achieve desired accuracy with minimal computational effort.Major integration routines were taken from Press et al. (1992), modified, <strong>and</strong> encoded for C++programming language. The program code is included in the appendix.2.5.3.2 Continues -Flow ContactorThe computer code for the batch reactor was modified to obtain the steady-state solution for anozone bubble-diffuser reactor with internal recirculation. In this case, both the reactor (mainbubble column) <strong>and</strong> the recirculation line were modeled as ideal plug-flow reactors (PFRs).Specifically, calculation along the bubble column <strong>and</strong> recirculation line was iterated untilsufficient convergence criteria were met at the inlet <strong>of</strong> the reactor. Additional mass balances forgaseous phase ozone <strong>and</strong> the interface mass transfer <strong>of</strong> gaseous phase ozone to liquid phase werealso incorporated into the model. The overall volumetric mass transfer coefficient for thespecific operating conditions were estimated from the empirical correlations described byMarinas et al. (1993). Average bubble size in the bubble-diffuser column, one deciding29


parameter in mass transfer evaluation, was measured photographically. The program code isincluded in the appendix.2.5.3.3 Pilot- <strong>and</strong> Full-Scale Ozone ContactorsEach chamber <strong>of</strong> a multi-chamber bubble diffuser contactor can be modeled as an axialdispersion reactor (ADR), where a single parameter (i.e. dispersion number) determines anoverall characteristic <strong>of</strong> the non-ideal reactor hydrodynamics. In general, the steady-state massbalance for each chemical species can be represented by the following second-order non-linearordinary differential equation as follows:dz dzC2 ,--,CN ) i = l,2,--,N (2.3)Where: d [dimensionless] = dispersion number; z [dimensionless] = normalizeddownward distance from the water surface; 9 [T] = hydraulic residence time inthe chamber. In the case <strong>of</strong> a bubble-diffuser column, interface mass transferterms are included in the reaction ft for the gas <strong>and</strong> liquid phases as described inMarinas et al. (1993). The sign <strong>of</strong> the second term was determined by theoperation mode (i.e. counter-current or co-current). In this model, it wasassumed that d = 0 in gas phase (<strong>and</strong> thus the equation is first-order) <strong>and</strong> the gasvolume fraction throughout the bubble-diffuser column was negligible. Theabove set <strong>of</strong> second order non-linear ordinary differential equation was subject to2N-1 boundary conditions <strong>of</strong> the form:i = l,2,-, AT (2.4)dC:outlet= 0 i =1,2,-, AT (2.5)Each second order ordinary differential equation except that for gas phase wasreplaced by a set <strong>of</strong> two first-order ordinary differential equations by introducingthe auxiliary variable d*.dz d(2.7)The resulting set <strong>of</strong> 2N-1, first-order differential equations were substituted with finite-differenceequations on a mesh <strong>of</strong> points that covered the range <strong>of</strong> integration. An iterative calculation wasperformed until the deviations from the finite-difference approximation <strong>of</strong> each equation at eachmesh point were sufficiently small. The calculation algorithm, called the relaxation method, wasadopted mostly from Press et al. (1992) <strong>and</strong> modified <strong>and</strong> programmed with C++ programminglanguage. The code is included in the appendix.30


CHAPTERS. BROMATE SURVEYA North American <strong>and</strong> European survey <strong>of</strong> ozonation utilities was conducted to underst<strong>and</strong>bromate occurrence (formation) at full-scale water treatment plants under current levels <strong>of</strong> ozoneapplication. While bromate occurrence information was the primary intent <strong>of</strong> the survey,additional information was extracted from a large water quality <strong>and</strong> treatment conditiondatabase. Specifically, relationships between key water quality parameters <strong>and</strong> bromateformation were investigated.3.1 PARTICIPATIONA list <strong>of</strong> c<strong>and</strong>idate participants was established by using two existing water treatment plantdatabases that identify ozonation facilities: the American Water Works Association (AWWA)WaterStats database <strong>and</strong> the International Ozone Association (IOA) database. A master list <strong>of</strong>over sixty potential survey c<strong>and</strong>idates was generated. A survey participation letter <strong>and</strong>questionnaire was sent to each c<strong>and</strong>idate, explaining the survey's intent <strong>and</strong> outlining the details<strong>of</strong> participation. Two rounds <strong>of</strong> solicitation by mail resulted in the participation <strong>of</strong> twenty-fourutilities, yielding a response rate <strong>of</strong> 40 percent. At the request <strong>of</strong> the majority <strong>of</strong> the utilities,participation was kept anonymous.The goal <strong>of</strong> the survey solicitation was to assemble a representative cross-section <strong>of</strong> ozonationutilities as participants. A geographic distribution <strong>of</strong> survey participants is presented in Figure3.1. There was a reasonable distribution among the participating utilities. Of the twenty-fourutilities that responded, twenty-one were located in the United States <strong>and</strong> three were located inFrance. While participation from California appears to be geographically biased, the largernumber <strong>of</strong> participants is somewhat justified due to the number <strong>of</strong> utilities that utilize ozonewithin the state.Survey participation was also summarized in terms <strong>of</strong> type <strong>of</strong> ozonation (i.e. point <strong>of</strong> ozoneapplication) <strong>and</strong> source water type. A qualitative summary <strong>of</strong> utility participation can be foundin Table 3.1. Pre-ozonation (ozonation before coagulation) was the dominant type <strong>of</strong> ozonationrepresented, <strong>and</strong> the majority <strong>of</strong> participants were ozonating surface waters.In order to provide some background in terms <strong>of</strong> the physical configuration <strong>of</strong> each utility'sozone process, the main details <strong>of</strong> the individual ozone contactors were summarized. Thisinformation is displayed in Table 3.2 in terms <strong>of</strong> location <strong>of</strong> ozone application, contactor type,<strong>and</strong> number <strong>of</strong> cells. All <strong>of</strong> the participating utilities utilized various configurations <strong>of</strong> finebubble diffusers to transfer the ozone into the bulk water; no use <strong>of</strong> other contactor types (i.e.turbine, packed column, direct injection or deep U-tube) was reported. Some contactors had upto six cells or individual chambers, while others consisted <strong>of</strong> a single column or a single largebasin, where ozone gas was applied. While many utilities have the capability for multi-stageozonation (i.e. multiple locations <strong>of</strong> ozone addition within one contactor), almost all <strong>of</strong> thecontactors are operated in a single-stage fashion.31


= utility location24 utilities total (21 U.S. <strong>and</strong> 3 French)Figure 3.1: Participation Map for Survey Participants3.2 ORGANIZATIONAfter responses to the survey questionnaire were received, the sampling effort was launched.The survey was organized into three rounds. The three separate sampling campaigns (June 1998,November 1998, <strong>and</strong> June 1999) were administered to obtain sample pairs <strong>of</strong> before <strong>and</strong> afterozone application <strong>and</strong> evaluate seasonal variability. The before-ozonation samples werecollected from the influent <strong>of</strong> the ozone contactor immediately prior to ozonation, <strong>and</strong> the afterozonationsamples were collected from the effluent <strong>of</strong> the ozone contactor. Thus, the true effects32


<strong>of</strong> the ozone contactor could be investigated, without the cumulative effect <strong>of</strong> multiple treatmentprocesses.Table 3.1: Qualitative Summary <strong>of</strong> Survey ParticipantsType <strong>of</strong> ozonation:Pre-ozonationIntermediate-ozonationDual-ozonation (2 points <strong>of</strong> application)Post-ozonationGeographic distribution:North-WestSouth-WestNorth-CentralSouth-CentralNorth-EastSouth-EastFranceSource water type:Surface water, river or aqueductSurface water, lake or reservoirGroundwater* 24 participants total in 3 sampling campaignsNumber <strong>of</strong> participants*11571The first <strong>and</strong> third rounds <strong>of</strong> samples were collected <strong>and</strong> analyzed in the late spring/earlysummer, which is generally representative <strong>of</strong> primarily late run<strong>of</strong>f conditions for surface watersupplies. The second round <strong>of</strong> sampling <strong>and</strong> analysis took place in late fall/early winter, whensurface-water hydrologic systems generally experience baseflow conditions. Because the thirdround <strong>of</strong> the survey was administered during the same time <strong>of</strong> year as the first round, these twodata sets were used to evaluate potential year-to-year temporal variation in source water quality.The second round <strong>of</strong> sampling could be compared to round one <strong>and</strong>/or three to examine seasonalvariability.While there were twenty-four survey participants, the number <strong>of</strong> samples received during eachround was less than this value. Round one included twenty-two utilities, round two had twentyoneparticipants, <strong>and</strong> round three received a response from twenty-two utilities. Two sample setswere requested from the dual-ozonation facilities. Both influent <strong>and</strong> effluent ozone sampleswere obtained for each <strong>of</strong> the two points <strong>of</strong> ozone application in an attempt to further underst<strong>and</strong>the stepwise bromate formation in this facility. This fluctuating number was due tocircumstantial inclusion or withdrawal <strong>of</strong> utilities from one round to another. Nineteen <strong>of</strong> theutilities participated in all three rounds <strong>of</strong> the survey.5442423138333


Table 3.2: Survey Utility Participant <strong>Ozonation</strong> InformationID A B C D E F G H I J K L M N 0 Type <strong>of</strong> <strong>Ozonation</strong>Pre Pre/Int Pre Pre Pre Pre/Int Pre Post Pre Pre Pre/Int Pre Pre/Int Int Int Contactor Type Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Diffuse bubble Flow ConfigurationOver/under Over/under Over/under Over/under Over/under Over/under Over/under Over/under Over/under Counter-current Cross-flow Over/under Over/under Over/under Over/under Cells' 2 4/4 5 4 6 2/6 4 5 3 1 1/1 5 4/4 3 5 Comments2 contactors in parallelCan apply Os in first 3 cells2 contactors in parallel4 contactors in parallelSingle-stage ozonation2 contactors in parallelCan apply Os in any <strong>of</strong> 4 cellsOzone is primary disinfectantMulti-stage applicationSingle columnSingle large undivided contact basin4 contactors in parallelSingle-stage ozonationSingle-stage ozonationCan apply Os in first 3 cellsP Int Diffuse bubble Over/under 4 Can apply Os in first 2 cellsQ Pre Diffuse bubble Over/under 2 Can apply Os in first 2 cellsR Pre Diffuse bubble Over/under 6 Ozone applied in first cellS T U Pre/Int Pre/Post Pre Diffuse bubble Diffuse bubble Diffuse bubble Over/under Over/under Over/under 4/4 4/4 6 Multi-stage application2nd Os application after filtersOzone is primary disinfectantV Int Diffuse bubble Over/under 4 Can apply Os in first 2 cellsw Pre/Int Diffuse bubble Over/under 1/4 Can apply Os in first 2 cells <strong>of</strong> int cont.X Int Diffuse bubble Over/under 3 Can apply Os in first 2 cellsnumber <strong>of</strong> cells in contactor; two numbers reflect pre- <strong>and</strong> int-(or post-)ozonation3.3 DETAILS OF ANALYSESThe information derived from the survey was organized into two groups: water qualityparameters <strong>and</strong> treatment conditions. Measured water quality parameters consisted <strong>of</strong> pH,temperature, DOC, UVA254, alkalinity, ammonia-nitrogen, bromide, <strong>and</strong> bromate. Calculatedwater quality parameters included SUVA, AUVA254, <strong>and</strong> percent bromide conversion to bromate.The percent bromide conversion to bromate was calculated as follows:% <strong>Bromide</strong> Conversion to <strong>Bromate</strong> = (BrO3 ~ as Br~) / Br"before * 100% (3.1)After receiving the samples, analyses <strong>of</strong> DOC, alkalinity <strong>and</strong> ammonia were performed on theinfluent sample only; UVA254 <strong>and</strong> bromide were conducted on the both the influent <strong>and</strong> effluentsamples; <strong>and</strong> bromate was only analyzed for in the effluent sample. Since the effluent sampleswere not quenched at the time <strong>of</strong> sampling, it is important to note that the data reflect thebromate formation after complete dissipation <strong>of</strong> ozone residual. These bromate values would be34


generally be representative unless the utility practiced ozone residual quenching (utilities werenot queried about quenching). Values for pH <strong>and</strong> temperature were measured by the respectiveutilities at the time <strong>of</strong> sampling.Treatment conditions monitored <strong>and</strong> provided by the utility included ozone dose (either astransferred dose or applied dose with estimated percent transfer efficiency) <strong>and</strong> hydraulicresidence time (HRT). When available, dissolved ozone concentrations at the exit <strong>of</strong> each cell <strong>of</strong>the ozone contactor <strong>and</strong> tio contact time estimates from each cell were provided. Also, estimates<strong>of</strong> CT were requested. If CT was not provided, but adequate ozone residual <strong>and</strong> tio data wereavailable, CTswia was calculated by SWTR guidelines (Oa residual x tio for each cell, excludingthe first cell). Although some informative data were obtained, the feedback from utility to utilityvaried in terms <strong>of</strong> type <strong>of</strong> information <strong>and</strong> degree <strong>of</strong> detail. For example, some utilities simplyprovided an overall HRT <strong>of</strong> the contactor along with transferred ozone dose, while othersprovided a complete breakdown <strong>of</strong> tio contact times <strong>and</strong> dissolved ozone residuals for each cellwithin the ozone contactor.3.4 RESULTSResults from the survey are presented for each round <strong>of</strong> sampling. Also, an overall summarywas created based on all three sampling campaigns. A collection <strong>of</strong> descriptive statistical tools,as well as simple-linear regression, was used to analyze each data set. The number <strong>of</strong> samplesreported for each round was greater than the number <strong>of</strong> participants because <strong>of</strong> the dualozonationutilities. For the dual-ozonation utilities, the influent to each ozone contactor wascharacterized in terms <strong>of</strong> water quality <strong>and</strong> treatment conditions. For statistical purposes, anynon-detect measurement was assigned a value <strong>of</strong> one-half times the minimum detection limit.3.4.1 Individual Survey Rounds3.4.1.1 Descriptive Statistical AnalysisInitially, a variety <strong>of</strong> descriptive statistics was used to summarize the data from each <strong>of</strong> thesampling rounds. The number <strong>of</strong> samples, the sample mean, st<strong>and</strong>ard deviation, <strong>and</strong> range invalues (minimum to maximum), as well as the 10th, 50th (median), <strong>and</strong> 90th percentile values <strong>of</strong>all the survey parameters were summarized. The results <strong>of</strong> the descriptive statistics are presentedin Table 3.3, Table 3.4, <strong>and</strong> Table 3.5, for sample rounds one, two <strong>and</strong> three, respectively. Theparameter <strong>of</strong> CT was not requested during round one <strong>of</strong> sampling <strong>and</strong>, therefore, was not a part<strong>of</strong> the statistical analysis for that round.The average bromate was highest during round three (4.7 ug/L), followed by round one (4.1ug/L) <strong>and</strong> lowest during round two (2.7 ug/L). Average percent bromide conversion to bromatefor rounds one, two <strong>and</strong> three were 7.3, 5.3, <strong>and</strong> 6.7 percent, respectively.35


Table 3.3: Water Quality <strong>and</strong> Treatment Condition Summary <strong>of</strong> Sampling Round OneParameterTemperature (°C)Br (ug/L)BrO3 " (ug/L)10th % 50th % 90th %Br effluent (ug/L)Br' conversion to BrO3"(%) PHAlkalinity (mg/L as CaCO3) NH3-N (mg/L)O3 dose (mg/L)Mean 21 66 4.1 62 7.3 7.7 111 0.17 1.9 Std. Dev.4.9525.653140.7690.191.0HRT (min)18 18CTswiR (mg/L-min)*- -DOC (mg /L)UVA254(cm- 1 )SUVA (L/mg-min)UVAeffluent (cm" 1)AUVA254 (cm' 1 )3.1 0.069 2.1 0.032 0.037 1.50.0500.70.0160.041* data not collected during the first roundRange8.8 - 272.5 - 1800.2- 182.5-1800.2 - 726.5 - 10.68.7 - 2700.01 - 0.700.5 - 5.02.0 - 82-1.3-8.60.012 - 0.2600.9 - 3.40.011-0.079-0.010-0.18114140.27.90.27.1290.020.83.0-1.50.0231.20.0170.00121541.6472.27.71100.082.015-2.70.0542.00.0280.0272717014170168.32000.522.730-4.50.1203.00.0520.098Table 3.4: Water Quality <strong>and</strong> Treatment Condition Summary <strong>of</strong> Sampling Round TwoParameterTemperature (°C)Br- (ug/L)Br03- (ug/L)10th % 50th % 90th %Br effluent (ug/L)Br" conversion to BrO3"(%)pHAlkalinity (mg/L as CaCO3)NH3-N (mg/L)O3 dose (mg/L)HRT (min)CTswiR (mg/L-min)*DOC (mg/L)UVA254(cm- 1 )SUVA (L/mg-min)UVAeffluent (cm' 1 )AUVA254 (cm' 1)Mean17492.7465.37.7950.141.5231.23.00.0682.10.0350.033Std. Dev.4.4407.039120.8610.230.9141.61.90.0580.630.0340.027Range7.0 - 243.7-1500.1-366.4-1500.08 - 536.3 - 9.96.8-2800.01 - 1.10.3-3.14.2 - 540.0 - 4.71.0-8.90.017 - 0.2470.9 - 3.26.011-0.1540.002-0.1169.0110.28.00.36.6390.010.3100.01.50.0241.40.0130.01018360.6331.67.6840.071.3180.32.50.0542.30.0240.022221204.1120198.31600.332.7463.54.90.0902.90.0670.065Slightly higher average transferred ozone doses were observed during rounds one <strong>and</strong> three (1.9mg/L <strong>and</strong> 2.0 mg/L, respectively), when compared to round two (1.5 mg/L). The vast majority<strong>of</strong> participants reported transferred ozone dose as opposed to applied dose with percent transferefficiency. HRT is a characterization <strong>of</strong> the contactor volume for a given flow, <strong>and</strong> served as abasis from which ti 0 contact data were derived. The limited amount <strong>of</strong> tio data provided for36


individual cells <strong>of</strong> contactors was determined from tracer tests <strong>and</strong> was dependent on the specifichydrodynamics <strong>of</strong> the individual contactors. The average HRT for each round ranged from 18 to23 minutes. The reported CTswiR averaged 1.23 mg-min/L <strong>and</strong> 0.68 mg-min/L for rounds two<strong>and</strong> three, respectively. These averages were calculated from all <strong>of</strong> the available data, though notall <strong>of</strong> the utilities responded to all the questions.Table 3.5: Water Quality <strong>and</strong> Treatment Condition Summary <strong>of</strong> Sampling Round ThreeParameterTemperature (°C)Bf (ug/L)Br03" (ug/L)10th % 50th % 90th %Br'effluent (Hg/L)Br" conversion to BrO3~(%)PHAlkalinity (mg/L as CaCO3)NH3-N (mg/L)O3 dose (mg/L)HRT (min)CTswxR (mg/L-min)*DOC (mg /L)UVA254(cm~ 1)SUVA(L/mg-min)UVAgffluent (cm" 1)AUVA254 (cm' 1 )Mean21454.7456.77.5930.222.0210.73.00.0712.30.0410.031Std. Dev.4.7381037110.7610.321.59.90.91.80.0540.60.0280.028Range9.0-27 .2.0 - 1300.2 - 403.2-1100.12-486.2 - 9.46.5 - 2800.01 - 1.20.1-7.96.8 - 490.0-3.10.8 - 8.40.013 -0.2441.1 -3.20.011 -0.1160.002-0.128135.90.28.10.26.5290.030.5120.01.30.0181.50.0130.00522291.1262.57.7780.071.9190.22.60.0652.30.0360.0252610022110278.31700.913.0352.25.00.1183.10.0860.052The organic characterization <strong>of</strong> the survey samples was limited to measurements <strong>of</strong> DOC,UVA254, SUVA, <strong>and</strong> AUVA254. Based on these four indicators, the average organic character <strong>of</strong>the ozone contactor influent samples did not appear to vary significantly between the samplingrounds. Mean values were very similar for DOC (from 3.01 to 3.10 mg/L), UVA (from 0.068 to0.071 cm" 1), <strong>and</strong> SUVA (from 2.08 to 2.25 L/mg-m). Even the average AUVA254, whichembodies the oxidative effects <strong>of</strong> ozonation, did not appear to change significantly (from 0.031to 0.037 cm" 1).3.4.1.2 Single-Linear Regression AnalysisSpecific relationships between water quality parameters <strong>and</strong> treatment conditions were exploredusing simple-linear regression. The direct relationship <strong>of</strong> each parameter compared to everyother parameter was evaluated through the development <strong>of</strong> a Pearson correlation matrix. Theresultant symmetric correlation matrix lists correlation coefficients (r values) for each parameteras a result <strong>of</strong> individual linear regression on a variable per variable basis. The Pearsoncorrelation matrices for rounds one, two, <strong>and</strong> three are shown in Table 3.6, Table 3.7, <strong>and</strong> Table3.8, respectively. The importance <strong>of</strong> the correlation coefficients was not only in the magnitude<strong>of</strong> the value but also the sign, which relays a direct (positive value) or inverse (negative value)relationship between parameters.37


Table 3.6: Pearson Correlation Matrix for Sampling Round OneooParameter 0.73 0.87 -0.350.70 0.96 -0.370.39 0.74 -0.32DOC UVA254 SUVA NH3-N Alk Br" Temp pH O3 dose HRT CTSWTR Br'efnuent Br-conver UVA254 efnuent AUVA254 BrO3"DOC 1.00 0.940.25 -0.09 0.69 -0.10 n/a 0.01 -0.21 UVA2S4 1.000.18 -0.22 0.71 -0.12 n/a -0.15 -0.23 SUVA-0.05 -0.32 0.46 -0.23 n/a -0.21 -0.20 NH3-NAlkalinityBrTemperaturePHO3 doseHRTCTs\VTRBr effluentD_-01 converUVA2 54_efflUentAUVA2S40.48 0.58 0.29 0.250.72 0.47 0.22 0.091.00 0.190.10 0.151.00 0.26 0.611.00 0.391.000.540.140.351.000.24-0.050.310.341.00BrO/Samples 26 26 26 26 26 26 26 26 25 25 26 26 26 26 26n/a = not applicable0.470.320.170.330.051.00-0.27-0.02-0.28-0.64-0.22-0.101.00n/an/an/an/an/an/an/an/a0.490.190.930.280.37-0.07-0.33n/a1.00-0.26-0.13-0.33-0.40-0.36-0.180.73n/a-0.321.000.370.240.240.00-0.100.46-0.05n/a0.06-0.281.000.430.180.020.21-0.240.70-0.13n/a-0.20-0.170.471.00-0.130.410.180.030.050.110.10n/a0.130.30-0.38-0.301.00


Table 3.7: Pearson Correlation Matrix for Sampling Round TwoOJParameter 0.92 0.87 -0.040.95 0.92 -0.130.43 0.61 -0.250.88 0.61 -0.050.24 0.25 0.090.52 0.40 0.200.51 0.44 -0.020.29 0.14 0.090.52 0.61 0.230.01 0.05 -0.07-0.41 -0.54 0.560.44 0.30 0.13-0.12 -0.15 0.501.00 0.75 -0.091.00 -0.17DOC UVA254 SUVA NH3-N Alk Br Temp pH 6^HRT CTSWTR Kr emaent Br-conver UVA254emuent AUVA254 BrO3doseDOC 1.00 0.96 0.31 0.82 0.28 0.57 0.52 0.27 0.61 -0.04 -0.43 0.48 -0.16 UVA254 1.00 0.55 0.81 0.26 0.50 0.51 0.24 0.60 0.03 -0.48 0.41 -0.14 SUVANH3-NAlkalinity1.00 0.20 1.00 0.06 0.14 0.02 0.53 1.00 0.05 1.00 0.16 0.49 0.11 0.38 0.07 0.39 0.04 0.44 0.41 0.49 0.02 0.70 0.03 0.02 -0.22 -0.13 -0.63 -0.26 -0.16 -0.39 0.13 0.53 -0.16 0.94 -0.01 -0.18 -0.24 -0.23 TemperaturepHO3 doseHRTCTswTR1.00 0.11 1.00 0.25 0.23 1.00 0.05 -0.33 0.04 1.00 -0.37 -0.29 -0.38 0.48 1.00 0.28 0.54 0.67 -0.18 -0.37 -0.44 -0.29 0.03 0.25 0.41 Br effluent1.00 -0.23 **'" conver1.00 UVA254.efnuentAUVA254_________________________________________________________________________________1.00Samples_____25 25 25 25 25 25 24 25 24 21 17 25 25_____25______25 25


Table 3.8: Pearson Correlation Matrix for Sampling Round ThreeParameter UVA2S4 efnuent AUVA254 BrO3'0.94 0.87 -0.150.95 0.95 -0.170.49 0.55 -0.080.73 0.46 -0.22DOC UVA254 SUVA NH3-N Alk Br' Temp pH O3 dose HRT Br'effluent DOC 1.00 0.96 0.31 0.65 0.34 0.36 0.27 0.45 0.25 0.15 -0.31 0.35 -0.25 UVA2S4 1.00 0.54 0.62 0.34 0.28 0.20 0.39 0.22 0.15 -0.28 0.26 -0.28 SUVA1.00 0.26 0.09 0.06 -0.08 0.19 0.12 0.04 -0.07 0.05 -0.19 NH3-N1.00 0.03 0.20 0.33 0.48 0.13 -0.03 -0.34 0.34 -0.29 Alkalinity1.00 0.43 0.15 0.23 -0.18 0.07 0.01 0.26 0.19 BrTemperaturepHO 3 doseHRTCTswTRBr effluent"1" converAUVA254Br031.00 0.24 1.00 0.29 0.20 1.00 0.12 -0.04 -0.11 1.00 0.07 -0.40 -0.48 0.63 1.00 0.35 -0.33 -0.16 -0.18 0.32 1.00 0.93 0.24 0.43 0.15 -0.02 0.30 1.00 -0.06 0.19 0.03 -0.05 0.02 0.30 -0.17 1.00 0.31 0.30 0.27 0.45 0.18 0.10 -0.35 0.30 -0.37 1.00 0.34 0.23 0.11 0.30 0.23 0.19 -0.19 0.19 -0.17 0.81 1.00 0.190.320.250.130.140.070.500.190.82-0.25-0.071.00Samples 27 27 27 27 27 27 27 27 26 20 23 27 27 27 27 27


A ranking scheme for the correlation coefficients was developed to qualitatively identify the topfive most influential parameters on bromate formation. An r value with a magnitude <strong>of</strong> 0.65 orgreater received a triple index ranking (+++ or —), an r value between 0.40 <strong>and</strong> 0.65 received adouble index rank (++ or --), <strong>and</strong> values less than 0.40 received a single index rank (+ or -). Thefive most influential parameters, along with their corresponding r values, are listed in Table 3.9,Table 3.10, <strong>and</strong> Table 3.11, for sampling rounds one, two <strong>and</strong> three, respectively. Since percentbromide conversion to bromate represents an alternative way to express bromate formation, itwas not included in the ranking. For round one, alkalinity showed the highest correlation tobromate (r = 0.41), while CTSWTR had the highest correlation with bromate for rounds two <strong>and</strong>three with r values <strong>of</strong> 0.56 <strong>and</strong> 0.50, respectively. Even though these correlations were thelargest for each <strong>of</strong> the three rounds, none <strong>of</strong> these relationships were considered strong.Table 3.9: Top Five Highest Correlating Parameters to <strong>Bromate</strong> for Single-LinearRegression Analysis <strong>of</strong> Round OneParameter Rank Correlation Index r valueAlkalinity I ++ 0.41UVAeffluent 2 - -0.38UVA 3 - -0.37DOC 4 - -0.35SUVA_____________5__________-_______-0.32Table 3.10: Top Five Highest Correlating Parameters to <strong>Bromate</strong> for Single-LinearRegression Analysis <strong>of</strong> Round TwoParameterCTsWTRSUVAOzone Dose<strong>Bromide</strong>AUVA254Rank12345Correlation Index r value++ 0.56-0.25+ 0.23+ 0.20-0.17Table 3.11: Top Five Highest Correlating Parameters to <strong>Bromate</strong> for Single-LinearRegression Analysis <strong>of</strong> Round ThreeParameter Rank Correlation Index r valueCTswiR 1 ++ 0.50<strong>Bromide</strong> 2 + 0.32Temperature 3 + 0.25UVA254-effluem 4 - . -0.25Alkalinity___________5__________+_______0.19In general, the results from the simple-linear regression analysis revealed that there were nooverwhelming trends between any one variable <strong>and</strong> bromate formation for the waters studied.However, the promotive or inhibitive trends from each parameter with respect to bromate41


formation were consistent. For example, bromide (+), SUVA (-), <strong>and</strong> CTswra (++) were shownin multiple rounds to be similarly correlated to bromate formation. Also, trends <strong>of</strong> temperature(+) <strong>and</strong> ozone dose (+) were shown as expected. DOC was relatively influential, as identifiedduring the first sampling round correlation ranking, <strong>and</strong> displayed a slight inhibitive effect onbromate formation.3.4.2 Composite Survey Database3.4.2.1 Descriptive Statistical AnalysisAs with each <strong>of</strong> the individual rounds, the cumulative survey database was summarized in theform <strong>of</strong> descriptive statistics. Table 3.12 displays the summary <strong>of</strong> these statistical indicators forthe overall pool <strong>of</strong> 78 samples. The average water temperature <strong>and</strong> pH were 20°C <strong>and</strong> 7.7,respectively. The average ammonia (0.18 mg/L as N) <strong>and</strong> alkalinity (100 mg/L as CaCO3) levelswere considered to be moderate <strong>and</strong> comparable to national averages for these parameters.Table 3.12: Cumulative Water Quality <strong>and</strong> Treatment Condition SummaryMean Std. Dev. Range10th % 50th % 90th %20 5.0 7.0 - 27 12 20 2653 44 2.0-180 6.0 42 1203.87 7.7 0.1-40.0 0.2 1.2 12.751 44 2.5 - 180 8.0 39 1206.5 13 0.1-72 0.2 2.0 187.7 0.7 6.2- 11 6.6 7.7 8.3100 63.3 6.5-280 30 98 1800.18 0.25 0.01 - 1.22 0.02 0.08 0.511.8 1.2 0.1-7.9 0.5 1.5 3.00.06 0.10 0.00 - 0.50 0.00 0.02 0.2021 15 2.0 - 82 5.0 18 350.9 1.3 0.0 - 4.7 0.0 0.3 3.03.1 1.73 0.8 - 8.9 1.5 2.6 4.80.069 0.053 0.012 - 0.260 0.023 0.061 0.1202.2 0.63 0.9 - 3.4 1.4 2.1 2.90.036 0.027 0.011-0.154 0.013 0.029 0.0660.034 0.032 -0.010-0.181 0.005 0.025 0.078ParameterTemperature (°C)Br" (ug/L)Br03'(ng/L)Br effluent (Ug/L)Br conversion to BrO3~(%)pHAlkalinity (mg/L as CaCO3)NH3-N (mg/L)O3 dose (mg/L)O3 residual (mg/L)HRT (min)CT SWTR (mg/L-min)*DOC (mg/L)UVA254(cm- 1)SUVA (L/mg-min)T *~> T\/ v -^effluent A r-ri ( v^/iil r*Tf\ ~ \ ^AUVA254 (cm' 1 )data not collected during the first sampling round<strong>Bromate</strong> levels varied significantly, from below the detection limit <strong>of</strong> 0.3 ug/L to an upper limit<strong>of</strong> 40 ug/L, with an average bromate concentration <strong>of</strong> 3.9 ug/L. <strong>Bromide</strong> conversion rangedfrom 0.1 to 72 percent, with an average conversion <strong>of</strong> 6.5 percent. In order to gain a betterunderst<strong>and</strong>ing <strong>of</strong> bromate occurrence, a cumulative frequency distribution graph was developedfor each <strong>of</strong> the three rounds as well as the cumulative survey database. The graph, which isshown in Figure 3.2, plots bromate concentration (log-scale) on the abscissa <strong>and</strong> cumulativedistribution percentile on the ordinate. Percentile values <strong>of</strong> bromate occurrence can bedetermined from the plot as the percentage <strong>of</strong> values less than or equal to a given percent. The42


50l percentile value for the cumulative database is 1.2 ug/L, while the individual roundsproduced 50th percentile values <strong>of</strong> 1.6, 0.6,.<strong>and</strong> 1.1 ug/L. The average bromide concentrationwas determined to be 53 ug/L.10080 -60I 400>CM2000.1 1 10<strong>Bromate</strong> ConcentrationSecond Sampling Round° Third Sampling Round• First SampKg RoundAll Three Rounds100Figure 3.2: Cumulative Distribution <strong>of</strong> <strong>Bromate</strong> <strong>Formation</strong> in 78 SamplesThe waters in this study (at the point <strong>of</strong> ozonation) were not high in humic content, asrepresented by the average SUVA <strong>of</strong> 2.2 L/mg-min. DOC values ranged from 0.8 mg/L to 8.9mg/L, with an average value <strong>of</strong> 3.1 mg/L.3.4.2.2 Single-Linear Regression AnalysisLinear regression was performed on the cumulative database, <strong>and</strong> the results were againpresented in the form <strong>of</strong> a Pearson correlation matrix. The correlation matrix for all theparameters <strong>of</strong> the sample pool is listed in Table 3.13. Similar to the individual rounds, the topfive correlating parameters were identified using the +/- ranking scheme (as before, bromideconversion was excluded). The summary <strong>of</strong> the correlation index ranking is listed in Table 3.14;the overall most influential parameter was CTSWTR-CTswTR was shown to have the highest correlation with bromate as represented by an r value <strong>of</strong>0.48. None <strong>of</strong> the other parameters show a statistically significant correlation with bromate.These results indicate that bromate formation may be simultaneously dependent on numerousparameters.43


Table 3.13: Cumulative Pearson Correlation MatrixParameter DOC UVA2S4 SUVA NH3-N Alk fir' Temp pH O3 dose HRT CTSWTR Br-emuent Hi-"JJI converTTVA*J T ^254 effluent AUVA2S4 BrO 3"DOC 1.00 0.95 0.35 0.67 0.30 0.37 0.33 0.23 0.43 -0.02 -0.36 0.26 -0.20 0.87 0.83 -0.15UVA254 1.00 0.60 0.63 0.27 0.26 0.28 0.14 0.44 -0.01 -0.39 0.14 -0.21 0.87 0.91 -0.19SUVA 1.00 0.23 0.00 -0.02 0.00 -0.04 0.29 -0.09 -0.40 -0.06 -0.14 0.43 0.63 -0.18NH3-N1.00 0.08 0.37 0.42 0.36 0.30 -0.11 -0.29 0.40 -0.23 0.71 0.45 -0.14Alkalinity1.00 0.32 0.14 0.08 0.03 -0.07 -0.08 0.13 -0.06 0.23 0.25 0.21Br"1.00 0.33 0.35 0.25 -0.20 -0.04 0.94 -0.22 0.31. 0.18 0.22Temperature1.00 0.19 0.20 -0.39 -0.38 0.28 -0.19 0.27 0.23 0.14pH1.00 0.02 -0.30 -0.20 0.44 -0.22 0.24 0.04 0.09O3 dose1.00 0.12 -0.27 0.18 -0.07 0.31 0.45 0.16HRT1.00 0.44 -0.26 0.48 0.02 -0.04 0.03CTgWTR1.00 -0.08 0.38 -0.39 Br effluent**1 converAUVA2S4BrO3"1.00 -0.24 1.00 0.24 -0.22 1.00 -0.37 0.03 -0.16 0.60 1.00 0.480.140.54-0.19-0.161.00Samples 78 78 78 78 78 78 77 78 75 66 40 78 78 78 78 78Table 3.14: Top Five Highest Correlating Parameters to <strong>Bromate</strong> for Single-Linear Regression Analysis <strong>of</strong> the CumulativeDataParameterCTswTRfir'AlkalinityUVA254UVA254-effluentRank Correlation Index r value1++ 0.482+ 0.223+ 0.214-0.195-0.19


3.5 DISCUSSIONWith the exception <strong>of</strong> the three French waters, the composite results from this survey should begenerally indicative <strong>of</strong> levels <strong>of</strong> bromate formation <strong>and</strong> the conditions under which bromate isformed, in the United States. This study was complimentary to a European study conductedearlier, entitled "A Survey <strong>of</strong> <strong>Bromate</strong> Ion in European Drinking Water" (Legube, 1996). Theresults <strong>of</strong> the European study were compared against those <strong>of</strong> this study.3.5.1 United States <strong>Bromate</strong> <strong>Formation</strong> SurveyOn average, bromate formation in 88 percent <strong>of</strong> the survey waters with current treatment (basedon Giardia inactivation, at ambient pH) was less than the Stage 1 Disinfectant / Disinfection By-Product 10 ng/L bromate maximum contaminant level. Conversely, 12 percent <strong>of</strong> the samplesanalyzed were shown to be in excess <strong>of</strong> the MCL indicating some form <strong>of</strong> bromate controlstrategy may need to be implemented. It was also observed that under current treatmentconditions, typically less than 10 percent <strong>of</strong> the bromide is converted to bromate (mass bromidebasis).The average CTswra value reported (a subset <strong>of</strong> 40 samples) was 0.91 mg/L-min. This can bethought <strong>of</strong> as an average CT value under current ozonation levels targeted for Giardiainactivation. Because Cryptosporidium is becoming the controlling organism dictating ozonedisinfection levels, higher CT values will be needed in order to achieve an adequate log-kill. Acurrent estimate <strong>of</strong> CTfor 1-log inactivation <strong>of</strong> Cryptosporidium at 15°C using ozone is 2.5 mgmin/L(Oppenheimer, et al., 2000). This fact, coupled with the observation that CT was the mostinfluential parameter on bromate formation, could result in an increased average bromate levelproduced during disinfection at water treatment plants that use ozone as the new levels <strong>of</strong>disinfection are implemented.3.5.2 Comparison to European <strong>Bromate</strong> <strong>Formation</strong> SurveyThe European Union (EU) <strong>Bromate</strong> Survey was comprised <strong>of</strong> thirty-eight European waterworks<strong>and</strong> spanned two sampling rounds, the first during the summer <strong>and</strong> fall <strong>of</strong> 1993 <strong>and</strong> the secondduring the spring <strong>of</strong> 1994 (Legube, 1996). The EU survey included the same water quality <strong>and</strong>treatment condition parameters as the US survey, except for the UV analyses (UVA, SUVA,UVA after, <strong>and</strong> AUVA254) <strong>and</strong> CT. Thus, the EU <strong>and</strong> US survey were compared based on the 12remaining parameters monitored. A composite analysis <strong>of</strong> the EU data is presented in Table3.15.Similar values <strong>of</strong> bromate formation were measured in both surveys. The average concentrationswere 3.6 ug/L <strong>and</strong> 3.9 ug/L, for the EU <strong>and</strong> this study, respectively. The average temperature,pH <strong>and</strong> DOC values were generally consistent. The primary differences in water quality wereobserved with the alkalinity (139 mg/L for EU vs. 100 mg/L for US), ammonia-N (0.08 mg/L vs.0.18 mg/L) <strong>and</strong> bromide ion concentrations (84 ug/L EU vs. 53 ug/L US).45


Table 3.15: Water Quality <strong>and</strong> Treatment Condition Summary <strong>of</strong> European Union<strong>Bromate</strong> SurveyParameterStd Dev Range10th % 50th % 90th %Temperature (°C)9.0-•26.8Bf(ng/L)10-•400BrO3' (ug/L)0.05 -19.62.5--430Br'after (f^g/L)Br conversion to BrO3'(%)pHAlkalinity (mg/L)NH3-N (mgVL)O3 dose (mg/L)HRT (min)DOC (mg/L)Mean17.4843.6874.77.31390.081.6112.73.6871.33.7482.65.580.5676.20.130.987.521.630.15.76.50.0010.13.00.25-30-8.6-277-0.60-4.2-40-8.512.5201.0130.56.4250.000.74.01.417.0602.0592.67.51510.051.59.02.222.01558.8187127.92490.203.0205.0Number6769706869686766605569As a final comparison, a Pearson correlation matrix was calculated for the EU survey to probethe most influential parameters in relation to their effect on bromate (Table 3.16). This analysisindicated that the most influential parameters were determined to be ozone dose, HRT <strong>and</strong>temperature, followed by pH <strong>and</strong> ammonia-nitrogen. It generally appeared that for the waterssampled in both surveys, bromate formation appeared to be influenced more by ozone treatmentconditions than water quality.46


Table 3.16: Pearson Correlation Matrix for European Union <strong>Bromate</strong> SurveyParameter -0.16 0.06-0.24 -0.180.02 0.03-0.43 -0.100.36 0.400.17 0.240.25 0.470.33 0.45-0.38 -0.061.00 0.84DOC NH3 Alkalinity Br' Temperature pH O3 dose HRT Br"after Br" conver BrO3"DOC 1.00 0.30 -0.13 0.37 0.00 -0.16 0.43 0.08 0.43 NH3-N 1.00 -0.27 0.19 -0.10 -0.46 -0.04 -0.13 0.19 AlkalinityBrTemperaturePHO3 doseHRTBr"after**r conver1.00 0.00 1.00 -0.09 -0.24 1.00 0.43 -0.01 0.16 1.00 -0.04 0.11 0.09 0.07 1.00 0.08 -0.07 0.35 0.14 0.25 1.00 0.08 0.95 -0.24 0.00 0.22 -0.01 1.00 BrO3 _______________________________________________________________1.00Samples_____69 66 67 69_____67_____68 60 55 68_____69_____70


CHAPTER 4. TRUE-BATCH INFLUENCE OF NOM AND TEMPERATURE ONBROMATE FORMATIONThis chapter emphasizes the use <strong>of</strong> true-batch simulations to probe important issues related todeveloping a comprehensive underst<strong>and</strong>ing <strong>of</strong> bromate formation for different levels <strong>of</strong>disinfection, as represented by CT^or ozone exposure (OE). These include the influence <strong>of</strong> bothNOM's concentration <strong>and</strong> character on bromate formation during disinfection <strong>and</strong> temperatureeffects on both bromate formation <strong>and</strong> disinfection. Each <strong>of</strong> these issues was probed through theuse <strong>of</strong> completely-mixed true-batch reactors to estimate bromate formation potential (FP) atdifferent levels <strong>of</strong> ozone exposure.4.1 INFLUENCE OF NOM ON BROMATE FORMATIONUnderst<strong>and</strong>ing the influences <strong>of</strong> NOM is critical to elucidating the mechanisms that contribute tobromate formation <strong>and</strong> affect disinfection (CT or OE). Distinct formation pathways have beenidentified including the molecular ozone pathway (direct oxidation) <strong>and</strong> the hydroxyl radicalpathway (indirect oxidation). Initial research attributed the majority <strong>of</strong> bromate formation innatural waters to the direct pathway, with a minimal influence from the indirect pathway (vonGunten <strong>and</strong> Hoigne, 1994). Recent research has indicated evidence to the contrary, suggestingthat the hydroxyl radical pathway can contribute up to 100 percent <strong>of</strong> the bromate formation insource natural waters (Ozekin et al., 1998; von Gunten & Oliveras, 1998).The effort to develop bromate minimization strategies requires an underst<strong>and</strong>ing <strong>of</strong> the influence<strong>of</strong> specific pathways for bromate formation. Natural organic matter can affect the chemistry <strong>of</strong>ozone processes in several ways. Direct reactions between ozone <strong>and</strong> NOM can effectivelyreduce the extent <strong>of</strong> ozone <strong>and</strong> bromide reactions. Specific NOM sites (e.g., humic <strong>and</strong> fulvicacids) can react with hydroxyl radicals to produce chain carriers <strong>of</strong> ozone decomposition, thusincreasing the hydroxyl radical formation rate. Other NOM sites may react with hydroxylradicals to produce compounds that do not react with ozone, <strong>and</strong> thus inhibit the ozone decaycycle without further production <strong>of</strong> oxidizing agents. Overall, variations in NOM can influencethe ratio <strong>of</strong> hydroxyl radical exposure <strong>and</strong> ozone exposure with natural waters, resulting invarying bromate formation rates even in those waters with similar water qualities (Elovitz et al.,1999b).The dependence <strong>of</strong> bromate formation on various inorganic <strong>and</strong> organic parameters has beenextensively investigated. However, organic parameters have mainly consisted <strong>of</strong> DOCconcentrations, specific NOM fractions, or molecular size <strong>of</strong> NOM. The role <strong>of</strong> functional NOMfractions has not been investigated, except for limited research on hydrophobic organic acids(Westerh<strong>of</strong>f, 1995). One objective <strong>of</strong> this research effort was to correlate NOM characteristicswith bromate formation, specifically looking at the role <strong>of</strong> functional fractions <strong>and</strong> averagemolecular size. A correlative statistical analysis was used to determine the influence <strong>of</strong>inorganic <strong>and</strong> organic parameters using a cumulative database <strong>of</strong> the fourteen project sourcewaters.49


4.1.1 Source WatersFourteen source waters were evaluated representing a geographic distribution <strong>of</strong> ozone plantsthroughout North America <strong>and</strong> Europe. Samples were collected from utilities in California (6);Texas (3); Florida (1); New Jersey (1); Michigan (1); Ottawa; Canada (1); <strong>and</strong> Paris, France (1).One utility used a groundwater source, one used a combined surface/ground water mix, <strong>and</strong> theremaining twelve utilities used surface water sources.4.1.2 True-Batch <strong>Ozonation</strong> ExperimentsTrue-batch ozonation experiments were used to assess ozone exposure <strong>and</strong> the resultant bromateformation potential. As indicated in Chapter 2, experiments were conducted under constantconditions for pH (7.0), temperature (20°C) <strong>and</strong> ozone dose (1 mg-Oa/L per 1 mg-DOC/L).After dosing the samples with ozone stock solution, kinetic data were collected as ozone residualmeasurements with time. Ozone residual measurements were taken every minute for the firstfive minutes <strong>and</strong> then were gradually spaced over longer intervals. Integration <strong>of</strong> the generatedozone decay curve was used to estimate OE. Once the ozone residual had decayed to below theminimum detection limit, samples were collected from the reactor <strong>and</strong> analyzed for bromate.The true-batch reactor was mathematically equivalent to plug-flow conditions, relating ozonedecay over time. Moreover, kinetics for ozone decay can be represented by a first-order rate law:where, [Os] = ozone concentration, mg/Lk = experimentally derived first-order rate constant for ozone decayThe relationship between bromate <strong>and</strong> OE has been shown to be approximately linear under nonbromidelimiting conditions. To compare bromate formation between source waters at a givenOE, the linearity assumption was used to develop a relationship between the experimentallyderived OE <strong>and</strong> bromate concentrations. A linear function passing through the origin wascalculated for each set <strong>of</strong> experimental results. The resultant linear function is represented as:[BrO3"]=K*OE;where, K = slope coefficient, pg/mg-minThe different source waters were then compared against one another using the K value as anindicator <strong>of</strong> expected bromate formation at a given OE.The linear representation <strong>of</strong> the bromate formation versus OE or CT relationship was meant to bea simple model for trending purposes. For a given reactor, the slope (K = BKV/OE (or CT)) isreflective <strong>of</strong> the set <strong>of</strong> water quality <strong>and</strong> ozone dose conditions that affect both bromateformation <strong>and</strong> disinfection (CT or OE). For different reactors, the slope (K) is also reflective <strong>of</strong>the unique hydrodynamic conditions. This simple model was not intended to be used forpredictive purposes but rather for trending comparisons among different water quality conditions50


for a given reactor, <strong>and</strong> among different reactors for a given set <strong>of</strong> water quality conditions.Departures from linearity are most strongly influenced by bromide-limiting conditions.4.1.3 Results4.1.3.1 NOM CharacterizationBefore investigating trends in the formation <strong>of</strong> bromate at various levels <strong>of</strong> disinfection, initialefforts were focused on characterizing the natural organic matter in the various project watersamples to determine if these characteristics might affect bromate formation. Thecharacterization <strong>of</strong> NOM properties consisted <strong>of</strong> measures <strong>of</strong> DOC concentrations, spectroscopicproperties, fractional DOC distribution, <strong>and</strong> molecular size distribution. Results <strong>of</strong> the analysesare presented in Table 4.1; refer to Section 2.4 for extended descriptions <strong>of</strong> the methods <strong>and</strong>analyses used to characterize the parameters. These parameters <strong>and</strong> properties collectivelydescribe the size, structure, <strong>and</strong> functionality <strong>of</strong> NOM; these are important because NOM exertsan ozone dem<strong>and</strong> <strong>and</strong> serves as both a scavenger <strong>and</strong> promoter <strong>of</strong> hydroxyl radicals. All sampleswere filtered through 0.45 urn filters <strong>and</strong> initially analyzed for DOC <strong>and</strong> UVA.Representative UVAioo-soo spectra for each water are presented in Figure 4.1. UVAaoo-soo can beconsidered to reflect the specific character <strong>of</strong> water in terms <strong>of</strong> aromatic carbon content. TheCGE spectrum is characterized by a very high absorbance below 240 nm; this phenomenon maybe due to interference by inorganics (i.e., NOs") present in the matrix. The nitrate level in theCGE sample was determined to be 19.6 mg/L, corroborating the noticeable difference inabsorbance compared to the other waters.Specific ultraviolet absorbance was calculated by normalizing the UVA254 with DOC. SUVAhas been shown to correlate well with both aromatic carbon content <strong>and</strong> aliphatic carbon content(inversely) <strong>of</strong> NOM isolates (Westerh<strong>of</strong>f et al., 1999). SUVA is a parameter that has been usedto relate relative characteristics <strong>of</strong> humic content; a high SUVA represents NOM that largelyconsists <strong>of</strong> humic material, <strong>and</strong> a lower SUVA indicates a less-humic nature <strong>of</strong> the NOM.Conventional treatment practices (i.e., coagulation) <strong>and</strong> ozone treatment are known to reduce thehumic nature <strong>of</strong> a water, effectively shifting the molecular size distribution <strong>and</strong> humic nature tolower molecular sizes <strong>and</strong> a non-humic nature. The source waters, arranged from high SUVAdescending to low SUVA, are presented in Figure 4.2. The range <strong>of</strong> SUVA values observed forthe water sources was from 3.8 L/mg-m (WPB) to 1.4 L/mg-m (AMA). Generally, waters withhigher SUVA are more reactive with ozone.Through the use <strong>of</strong> macroporous, nonionic resins, the NOM was fractionated into hydrophobic,transphilic, <strong>and</strong> hydrophilic fractions. Examples <strong>of</strong> the compounds in the transphilic (TPI)fraction includes carboxylic acids, amino acids, carbohydrates, <strong>and</strong> volatile hydrocarbons. Thehydrophilic (HPI) fraction is characterized by sugar acids, volatile fatty acids, hydroxy acids, <strong>and</strong>complex polyelectrolytic acids with many carboxylic <strong>and</strong> hydroxyl functional groups. Thehydrophobic (HPO) fraction is mostly made up <strong>of</strong> humic <strong>and</strong> fulvic substances that exhibit anaromatic nature <strong>and</strong> contain electron rich double bonds <strong>and</strong> conjugate bonds that arepreferentially oxidized by ozone (Westerh<strong>of</strong>f et al, 1999; Thurman, 1985).51


K)Table 4.1: NOM Characterization <strong>of</strong> Raw Source WatersWater SUVA Natural Organic Matter FractionHydrophobic Transphilic Hydrophilic Hydrophobic TransphilicWPBHOUCCDNJASACCGESPWLAWACDOTTDALANNCRWAMAAverageMedianStd DevRange(L/mg-m) (% NOM)3.83.53.42.72.62.52.42.32.22.12.01.91.61.42.52.40.71.4-3.8666254515347455449524245453550507.935-66(% NOM)181720211719232021202224242721212.917-27(% NOM)162126282934322631283630313729305.616-37(mg-C/L)6.95.81.01.50.90.71.31.00.81.4.9.1.2.0.9.2.90.7 - 6.9(mg-C/L)1.91.60.40.60.30.30.60.40.40.51.00.60.70.80.7 •0.60.50.3-1.9Hydrophilic WeightAveraged(mg-C/L) (Daltons)1.72.00.50.80.50.50.90.50.50.71.60.80.81.00.90.80.50.5-2.0Molecular Weight PolydispersivityNumberAveraged(Daltons)1350 10201.31250 7201.71250 9601.31020 5801.81210 9151.3810 6401.31090 8501.31300 9401.41590 11001.4885 7301.2985 7101.4680 5201.3980 6301.6990 7901.31099 7931.41055 7601.3241 175 0.2680-1590 520-1100 1.2-1.8


u11.00.8°- 60.40.21.CGE2.WPB3.HOU4.NJA5.SPW6.CCD7.AMA8. CRW9.ACD10.DAL11. ANN12.0TT13. SAC14. LAW0.0200- 220 240 260Wavelength (nm)280 300Figure 4.1: Source Water Ultraviolet Adsorbance Spectra129- 10E• SUVA(L/mg-m)D DOC (mg/L)uo/——NIWPB HOU CCD NJA SAC CGE SPW LAW ACD OTT DAL ANN CRW AMAFigure 4.2: Source Water DOC <strong>and</strong> SUVA Values53


NOM fractions were calculated on both a relative distribution basis (mass <strong>of</strong> fraction per mass <strong>of</strong>bulk DOC) <strong>and</strong> a mass (mg-C/L) basis to account for the difference in DOC concentrations. Forthose waters with higher DOC concentrations, the mass <strong>of</strong> the fractions were much larger thanfor waters having a low DOC. For example, HOU <strong>and</strong> SAC waters have equal percentages <strong>of</strong>the TPI fraction (17 percent), but the mass varies between, 1.6 mg-C/L for HOU water <strong>and</strong> 0.3mg-C/L for SAC water.The percent <strong>of</strong> NOM fractions for all <strong>of</strong> the waters are presented in Figure 4.3. The HPOfraction predominates (average = 50 percent), followed by the HPI (29 percent) <strong>and</strong> TPI (21percent) fractions. The distribution <strong>of</strong> fraction percentages was relatively consistent except forthree waters; WPB <strong>and</strong> HOU were observed to reflect up to 66 percent <strong>of</strong> the HPO fraction <strong>and</strong>AMA analyses indicated a balance between the HPO <strong>and</strong> HPI fractions. Both <strong>of</strong> these watershad DOC concentrations significantly greater than the other waters. The fractions generallyfollow a trend relating to SUVA. The waters with high SUVA show a predominant percentage<strong>of</strong> HPO matter; waters lower in the SUVA range exhibit an increasing percentage <strong>of</strong> TPI matter<strong>and</strong>, to a somewhat lesser extent, HPI matter.100%80%Io03Io>g60%40%20%0%WPB HOU CCD NJA SAC CGE SPW LAW ACD OTT DAL ANN CRW AMASource WaterFigure 4.3: Distribution <strong>of</strong> NOM FractionsApparent molecular weight (AMW) results ranged from 680 to 1590 daltons based on weightaveraged (Mw) values, with a median <strong>of</strong> 1055 daltons. The range for the number averagedAMW (Mn) was observed to be 520 to 1100 daltons with a median <strong>of</strong> 760 daltons. Arepresentative size exclusion chromatogram for CCD water is presented in Figure 4.4. As withthe fraction percentages, the molecular weight distribution is observed to follow similar trends,hi general, as SUVA decreases, the AMW, calculated as Mw, decreases. The AMW calculatedas Mn did not portray a similar trend <strong>and</strong> was variable for all waters. Polydispersivity was usedas a measure <strong>of</strong> NOM heterogeneity. The observed polydispersivities ranged from 1.2 (OTT) to54


1.8 (NJA). A value <strong>of</strong> 1.0 indicates ideal homogeneity <strong>of</strong> the AMWs; therefore, values closer to1.0 indicate more homogeneous <strong>and</strong> uniform NOM.8000i [ i i inMwAve =1250 daltonsStDev = 80 daltons6000 h CV = 6.1%I ICCDRunlCCD Run 2CCD Run 3oaoC3i_40002000Ave = 960 daltons•StDev = 35 daltonsCV = 3.5%PolydispersivityAve =1.3StDev = 0.07CV = 5.3%010 100 1000Molecular Weight (daltons)Figure 4.4: Size Exclusion Chromatography Chromatograms for CCD Water1000C4.1.3.2 True-Batch <strong>Ozonation</strong> ResultsPrepared aliquots <strong>of</strong> raw water samples were ozonated per the true-batch procedure explained inSection 2.2.1.1. The data derived from these experiments were used to estimate ozone decay,bromate formation potential (FP), <strong>and</strong> OE. The slope coefficient, K, for the bromate FP versusOE relationship was calculated based on the experimental data. Table 4.2 displays the overallresults for the true-batch ozonation <strong>of</strong> the fourteen source waters.Ozone Decay KineticsOzone decay rates were estimated by fitting a first-order rate equation to the measured ozonedecay data. The decay curves developed for the fourteen waters <strong>and</strong> Milli-Q water are presentedin Figure 4.5 <strong>and</strong> Figure 4.6. The curves for low DOC waters (3 mg/L), are presented separately. For all waters, a rapid initial decay rate was observed,characterized by a linear (constant) decay. Linearity for this region is assumed since the ozonedecay was very rapid <strong>and</strong> accurate measurements could not be reproducibly made until oneminute had passed. This initial phase could probably be modeled as a steep exponential drop inozone residual. For consistency, however, the first stage <strong>of</strong> ozone decay was characterized byozone decomposition as expressed by:55


= [O3]o - [O3]twhere, Ainjtiai represents the amount <strong>of</strong> ozone loss during this phase <strong>of</strong> the reaction<strong>and</strong> reflects a "pseudo" zero-order reaction (Westerh<strong>of</strong>f, 1995). The initialperiod persisted up to three minutes, followed by an apparent pseudo-firstorder rate <strong>of</strong> ozone decay.Table 4.2: Summary <strong>of</strong> <strong>Bromate</strong> <strong>Formation</strong> Potentials <strong>and</strong> Ozone Exposure ResultsWaterBr"OE Br03 " K* Dilution£TAOs initial-')(Hg/L) (mg-min/L) (Hg/L) (Hg/mg-min) (%) (mg/L) (xlO"3 secAMA WPB SPW CRW ANN ACD CCD LAW CGE HOU NJA DAL OTT SAC Mean Median 182119 13166 6048 37 3022 15 18 18 15 47 5842 18 3.3 12 27 1123 5.0 12 7.9 4.5 3.5 6.9 4.5 6.0 10 7.4 35 4.2 23 18.1 11 17 8.5 7.6 1.2 1.9 1.5 2.4 1.4 7.7 107.7 1.90 1.27 1.92 0.660.970.73 1.70 0.650.150.42 0.43 0.35 0.31 1.28 10.7 9 30 10 8 8 13 6 8 5 28 12 14 10 6 12 9.5 1.3 6.6 1.0 1.7 1.0 1.5 0.5 0.9 0.5 6.7 1.6 2.2 1.2 0.6 1.9 1.3 1.3428.42.491.022.911.025.231.311.9322.711.54.968.483.2373.1StdDev 51 7.7 10.0 0.6 7.7 2.0 8.5Range 15-182 3.3 - 27 1.0-35 0.15-1.92 5-30 0.6-6.7 1.02-28.4* Based on diluted sample bench-scale experimentst First order decay constants estimated using ozone decay data <strong>and</strong> exponential curve fitNatural organic matter has been shown to exert an influence on ozone decay (Westerh<strong>of</strong>f et al.,1999). Organic matter <strong>of</strong> an aromatic nature has been shown to be highly reactive with ozone.The waters with HPO fractions greater than 50 percent (WPB, HOU, CCD, NJA, SAC, OTT,LAW) were generally observed to have higher ozone decay rate constants ranging from 2.49x10"sec" to 28.4x10"3 sec" 1 . The range <strong>of</strong> rate constants for those waters with less than 50 percentHPO (CGE, SPW, ACD, DAL, ANN, CRW, <strong>and</strong> AMA) was less variable, ranging between1.02x10" sec" 1 <strong>and</strong> 4.96x10"3 sec" 1 . These decay constants are in agreement with published datafor natural waters (Elovitz et al, 1999b).56


Residual Ozone Concentration (mg/L) Residual Ozone Concentration (mg/L)99IS3ereIInCfQ erOonC/5Orsre1 ooorereOsoBreOrens9onVcroIO NOorerersne9oIoo


After ozonation, the sample was measured for AUVA254. These results are presented in Table4.3. The change in the aromatic character <strong>of</strong> the NOM was represented by the decrease in theUVA254 after ozonation. Representative spectra, normalized to diluted DOC, are presented inFigure 4.7 for comparison. The difference <strong>of</strong> UVA at a particular wavelength is indicative <strong>of</strong> thereactivity <strong>of</strong> specific NOM sites with ozone; decreases in UVA254 is indicative <strong>of</strong> the reactivearomatic portions <strong>of</strong> the NOM molecules (Westerh<strong>of</strong>f et al., 1999). The observed AUVA254ranged from 0.018 cm" 1 for AMA to 0.139 cm" 1 for WPB. The relative change in the WPB <strong>and</strong>HOU waters was far greater than the other AUVA254 values observed for the remaining waters(0.018 cm" 1 to 0.046 cm" 1). Both the WPB <strong>and</strong> HOU waters were characterized by predominantHPO fractions. The CGE spectrum was influenced by a high nitrate concentration <strong>and</strong> thereforeshould not be considered as a true representation <strong>of</strong> NOM reactivity with ozone.I1. LAW2. HOU3. WPB4. CGE5. NJA6. CRW7. ACD8. SPW9. ANN10. OTT11. SAC12. CCD13. AMA14.DALP0200 220 240 260Wavelength (nm)280 300<strong>Bromate</strong> <strong>Formation</strong> PotentialsFigure 4.7: ASUVA Spectra <strong>of</strong> Ozonated Project WatersOnce the ozone residual concentration in each <strong>of</strong> the bromate formation potential tests (dosed 1mg Os per mg DOC) had decayed to below detection limits, sample aliquots were collected forbromide <strong>and</strong> bromate analyses. The results are presented in Table 4.3. In order to attempt toclose the bromine mass balance, total organic bromine (TOBr) was modeled using a relationshipdeveloped by Amy et al. (1997). The modeling results predicted only 0.0 to 2.5 ug/L <strong>of</strong> thebromide might be incorporation into TOBr. Alone, these predicted levels <strong>of</strong> bromideincorporation into TOBr would not be able to explain the amount <strong>of</strong> unaccounted bromide.58


Dilution(%)302861265108131014889121085-30WaterWPBHOUCCDNJASACCGESPWLAWACDOTTDALANNCRWAMAMeanMedianStd Dev.RangeAUVA254(cm' 1)0.1390.1460.0260.0350.0230.0210.0400.0420.0460.0310.0400.0320.0230.0180.0470.0330.0410.018-0.146InitialBr'(Hg/L)1702139215023145335517216572200Table 4.3: True-Batch Post-<strong>Ozonation</strong> Water Quality ParametersResidualBr'«L)926.0261543156413406.39.758301103828336.0-11BrO3(Hg/L)4.21.98.51.57.7'1.2237.6131.42.4111835107.7101.2-35UnaccountedBr'*(Hg/L)75.413.87.75.12.27.366.615.37.09.89.80.0*30.768.1ModeledTOBr**(ug/L as Br')3.40.10.20.10.00.01.90.10.40.10.10.70.62.90.80.21.10.0 - 2.5Br' to BrO3'Conversion(%)2.27.9145.1103.411161768.312171210114.92.2-17Br"Incorporation(%)236029198315157175946355403635203-60* "0.0" value indicates that the calculated Br" recovery is greater than the estimated Initial <strong>Bromide</strong> concentration (adjusted for dilution);Unaccounted Br"=[Initial Br'*(l-dilution)]-(residual Br')-(BrO :r as Br")]** TOBr estimated using Amy et al. (1997) TOBr Model where: [TOBr]=1.004(Br-/ 765(DOC)-055 '(pH)-3 -916(NH3-N)c 13(O3) U35(TIC)-°'".;where, TIC is total inorganic carbon (mg/L)


The average bromide conversion to bromate was 10 percent, with a range from 2.0 (WPB) to 17percent (CRW). The total reduction in bromide concentration, presumably as result <strong>of</strong>incorporation into bromate, hypobromous acid, hypobromite <strong>and</strong> TOBr, ranged from 3 (ANN) to60 percent (HOU), with an average <strong>of</strong> 36 percent. In other research, HOBr had been observed toaccount for 5 to 50 percent <strong>of</strong> initial Br" molar concentration (Westerh<strong>of</strong>f, 1998).For these ozonated waters, the calculated ozone exposure values ranged from 3.3 mg-min/L forWPB to 27 mg-min/L for CRW. <strong>Bromate</strong> concentrations measured after the true-batchozonation were referred to as the formation potential (FP) concentrations. As discussed earlier,the bromate formation potential was plotted against the ozone exposure for the given water <strong>and</strong>fit with a linear function passing through the origin. Past research has shown an approximatelylinear relationship between bromate formation <strong>and</strong> ozone contact if bromide was not limiting.Additional tests were performed with LAW at elevated ozone doses to determine if this linearrelationship would be observed in the true-batch reactor. These bench-scale tests run with LAWappeared to support this general trending approach, as shown Figure 4.8.0 5 10 1520Ozone Exposure (mg-min/L)Figure 4.8: Linear Relationship Between <strong>Bromate</strong> <strong>Formation</strong> <strong>and</strong> Ozone Exposure (LAW)<strong>Bromate</strong> formation potential <strong>and</strong> ozone exposure for each water are plotted in Figure 4.9. Themaximum observed bromate FP was 35 ug/L (AMA) with an OE <strong>of</strong> 18 mg-min/L, <strong>and</strong> theminimum was 1.2 ug/L (CGE) with an OE <strong>of</strong> 7.9 mg-min/L. The mean FP <strong>and</strong> OE values were10 ug/L <strong>and</strong> 10 mg-min/L, respectively, though the corresponding median values were 7.7 ug/L<strong>and</strong> 7.4 mg-min/L. Only the data from the LAW 1:1 ozone dose ratio were included in theanalysis. The slope coefficients, K, for each water were calculated using the linear function, orquite simply [BrO3 ~ FP]/OE (Table 4.2). Assuming linearity, the slope coefficients can be usedto estimate bromate concentration as a function <strong>of</strong> OE. These coefficients ranged from 0.1560


ug/mg-min for CGE to 1.92 ug/mg-min for SPW. High K values indicate a water's propensityto form more bromate with an incremental increase in OE when compared to lower K values.oOHesoos2ca25201510SPW ———AMACCD ----- SACWPB x ANN---- ACD ———CRW*~~LAW -^—NJAa— HOU —o— DAL+--OTT —•—CGEl-Log Cryptosporidium Inactivation(1.2 - 2.9 mg-irdn/L @ 20°Q10 ^g/L <strong>Bromate</strong> MCL0 3 6 9 12Ozone Exposure (mg-min/L)Figure 4.9: <strong>Bromate</strong> <strong>Formation</strong> Potential <strong>and</strong> Ozone Exposure (20°C, pH 7)Figure 4.9 also shows a representative inactivation range for 1-log inactivation <strong>of</strong>Cryptosporidium, 1.2 to 2.9 mg-min/L, based on Oppenheimer et al. (2000) <strong>and</strong> Rennecker et al.(1999) at 20°C. Given the 10 ug/L MCL <strong>and</strong> using the upper end <strong>of</strong> this representativeCryptosporidium inactivation range (2.9 mg-min/L), bromate formation should not be a problemfor any water with a slope coefficient K consistently less than 3.4 ng/mg-min at 20°C. Thelargest K value was only 1.9, 44 percent lower than the predicted maximum. Consequently,bromate minimization strategies may not need to be investigated for these waters at 20°C if only1-log Cryptosporidium inactivation was deemed necessary. However, this might not be the casefor higher levels <strong>of</strong> Cryptosporidium inactivation or ozone exposure.Compliance AssessmentThe data derived from the true-batch experiments were used to assess the impacts <strong>of</strong> increasedCT requirements for the inactivation <strong>of</strong> Cryptosporidium. Ozone has been shown to be aneffective disinfectant for inactivation <strong>of</strong> target pathogens, particularly Cryptosporidium (Finch<strong>and</strong> Li, 1999). With the expected promulgation <strong>of</strong> up to 2.5-log <strong>of</strong> additional Cryptosporidiumremoval, ozone CT will need to be increased to achieve these levels <strong>of</strong> compliance. Currentresearch efforts are attempting to discern the necessary CT to provide ozone inactivation.Oppenheimer et al. (2000) <strong>and</strong> Rennecker et al (1999) used the Chick-Watson model to developsuggested CT values for inactivation across a range <strong>of</strong> temperatures from 0.5°C to 30°C <strong>and</strong> log-61


kills from 0.5 to 6.0. Using the K values determined for these project waters, bromate formationwas predicted using the suggested Oppenheimer et al. (2000) <strong>and</strong> Rennecker et al. (1999) CTvalues for 1-log, 2-log <strong>and</strong> 3-log inactivation at 20°C. The results from this analysis arepresented in Table 4.4. The analogous experimental parameter to CT is the OE calculated as afunction <strong>of</strong> observed ozone decay data. The CT value incorporates ozone decay into estimation<strong>of</strong> inactivation for pilot- <strong>and</strong> full-scale applications; OE incorporates ozone decay into estimation<strong>of</strong> true-batch inactivation <strong>and</strong> bromate formation at the bench-scale.Table 4.4: Estimated <strong>Bromate</strong> <strong>Formation</strong> for Cryptosporidium Inactivation (20°C, pH 7)Source WaterSPWAMACCDSAC*WPBANNACDCRWLAWHOUDALOTTNJACGEspiked to 47 ng/L Br"K1.921.901.701.281.270.970.730.660.650.420.350.310.300.15~l-Log Inactivation(1. 2-2.9 mg-min/L)2.3-5.62.3 - 5.52.0-4.91.5-3.71.5-3.71.2-2.80.9-2.10.8-1.90.8-1.90.5 - 1.20.4-1.00.4 - 0.90.4 - 0.90.2 - 0.4<strong>Bromate</strong> <strong>Formation</strong> Potential~2-Log Inactivation(2.3-5.2 mg-min/L)4.4 - 104.4 - 9.93.9-8.82.9 - 6.72.9 - 6.62.2 - 5.01.7-3.81.5-3.41.5-3.41.0-2.20.8- 1.80.7-1.60.7-1.60.3 - 0.8~3-Log Inactivation(3.5-7.5 mg-min/L)6.7 - 146.7 - 146.0 - 134.5 - 9.64.4 - 9.53.4 - 7.32.6 - 5.52.3 - 5.02.3 - 4.91.5-3.21.2-2.61.1-2.31.1-2.30.5-1.1The results shown in Table 4.4, indicate that depending on which data the EPA decides to utilizefor the development <strong>of</strong> the ozone CT values necessary for various levels <strong>of</strong> Cryptosporidiuminactivation, bromate formation may or may not be an issue for many water sources. Based onusing the suggested Oppenheimer et al. (2000) CT values, bromate formation at 20°C would notbe a concern for any <strong>of</strong> these project waters at 20°C, even in excess <strong>of</strong> credit for 3-logs <strong>of</strong>inactivation. The more conservative Rennecker et al. (1999) suggested CT values indicated thatbromate formation at 20°C might exceed 10 ug/L for some waters (SPW, AMA, CCD) at orabove 2-logs <strong>of</strong> Cryptosporidium inactivation.Should a more stringent 5 ng/L MCL be implemented <strong>and</strong> the more conservative Rennecker etal. (1999) suggested CT values be used, SPW, AMA <strong>and</strong> CCD would be impacted at 1-log <strong>of</strong>Cryptosporidium inactivation by ozone. At 2-logs, the spiked SAC, WPB, <strong>and</strong> ANN would jointhe list <strong>of</strong> impacted project waters. At 3-logs <strong>of</strong> inactivation, nine <strong>of</strong> the project waters wouldneed to have some form <strong>of</strong> bromate minimization strategy in place at 20°C.62


BrConv OE BrO3-FP-0.11 0.31 0.27-0.51 -0.41 -0.390.02 0.31 0.73-0.44 -0.30 -0.25-0.48 -0.39 -0.32-0.44 -0.68 -0.55-0.35 -0.54 -0.620.43 0.63 0.770.29 0.46 0.47-0.49 -0.40 -0.33-0.46 -0.29 -0.16-0.42 -0.27 -0.160.29 0.10 0.040.26 0.06 0.170.01 0.05 -0.20-0.39 -0.35 -0.331.00 0.74 0.571.00 0.711.00o\U)Parameter Alk NH3 Br DOC UVA SUVAAlk 1.00 0.21 0.54 0.15 0.17 0.01NH3 1.00 0.19 0.74 0.72 0.46Br 1.00 0.28 0.24 -0.11DOC 1.00 0.98 0.61UVA 1.00 0.73SUVA 1.00HPO%TPI%HPI%HPO-DOCTPI-DOCHPI-DOCAMW-MwAMW-MnPolyAUVABr ConvOEBrO3- FPTable 4.5: Pearson Correlation Matrix for <strong>Bromate</strong> <strong>Formation</strong> Potential ExperimentsHPO%-0.030.44-0.210.640.750.881.00TPI%0.07-0.270.53-0.38-0.50-0.78-0.841.00HPI% HPO-DOC0.02 0.15-0.48 0.730.01 0.25-0.69 0.98-0.79 0.99-0.83 0.69-0.96 0.730.66 -0.461.00 -0.781.00TPI-DOC0.140.770.380.970.930.500.50-0.21-0.600.951.00HPI-DOC-0.020.690.250.880.800.340.30-0.12-0.360.830.931.00AMW-Mw0.220.190.070.350.390.460.51-0.42-0.460.320.180.071.00AMW-Mn Poly0.38 -0.310.14 0.080.23 -0.280.17 0.270.23 0.240.36 0.140.38 0.17-0.32 -0.08-0.33 -0.180.16 0.240.04 0.25-0.11 0.330.88 0.121.00 -0.361.00AUVA0.080.630.160.970.98.0.710.76-0.52-0.790.970.890.790.430.230.281.00


The results show that at 20°C, the overall impact <strong>of</strong> increased CT on low-bromide waters (


4.2 TEMPERATURE EFFECTS ON BROMATE FORMATION AND DISINFECTIONWhile the influence <strong>of</strong> temperature on disinfection can be represented by temperature-dependentCT (or OE) values, the effects <strong>of</strong> temperature on bromate formation have not been extensivelystudied; few studies report results for variations (e.g., seasonal) in temperature during ozonation.Recent efforts to characterize the impacts <strong>of</strong> variable temperatures have focused on thedecomposition <strong>of</strong> ozone <strong>and</strong> the resultant ozone exposure associated with stabilized ozoneresiduals at lower temperatures. Elovitz et al. (1999b) studied the effects <strong>of</strong> reaction temperatureon the ozonation <strong>of</strong> natural waters <strong>and</strong> observed that ozone decay rates decreased more than anorder <strong>of</strong> magnitude, from l.lxlO"2 sec" 1 to 6.0X10"4 sec" 1 , as temperatures were lowered from35°C to 5°C. To assess the impact <strong>of</strong> temperature on the indirect pathway, these authors used theRet concept (=[OH»]/[O3J) to conclude that temperature does not have an impact on hydroxylradical exposure. This was attributed to the relatively low activation energy for reactions withhydroxyl radicals, which are typically in the 5 to 10 kJ/mol range.<strong>Bromate</strong> formation under variable temperature has been suggested to be a function <strong>of</strong> theincreased reaction rates with bromide species (Siddiqui <strong>and</strong> Amy, 1993; Song, 1996). Siddiqui<strong>and</strong> Amy (1993) observed an increase in bromate formation <strong>of</strong> 20 percent, on an ozone dosebasis, as temperature was increased from 20°C to 30°C in semi-batch reactor experiments. Otherstudies utilizing various batch reaction systems indicated similar increases in bromateconcentrations over similar temperature ranges, up to a 50 percent increase between experimentsperformed at 5°C <strong>and</strong> 20°C (Siddiqui et a., 1995). It should be noted, however, that the bromideconcentrations for these experiments were 1,000 jj.g/L or greater. The effects <strong>of</strong> temperature onlow bromide waters have not been studied, but it would be expected that these waters wouldexhibit similar trends in bromate reduction with decreased temperature.Of paramount practical interest is the effect <strong>of</strong> temperature variation on the inactivation <strong>of</strong>pathogens. Cryptosporidiwn has been shown to be more resistant to inactivation with decreasingtemperature (Oppenheimer et al., 2000; Finch <strong>and</strong> Li, 1999; Rennecker et al., 1999). SuggestedCT values reflect this resistance as they increase three to ten-fold over a temperature range from20°C to 5°C; the implications on bromate formation are obvious. With increased CTrequirements, the disinfectant dose or contact time must be increased, which in turn increases theprobability <strong>of</strong> forming bromate concentrations in excess <strong>of</strong> observed baseline conditions. Testswere conducted in this study to investigate the impact <strong>of</strong> temperature on bromate formation <strong>and</strong>corresponding disinfection (CTor OE).4.2.1 Temperature <strong>Control</strong>led True Batch <strong>Ozonation</strong>Five waters were chosen to be used in temperature-controlled true-batch ozonation experiments:AMA, SPW, CCD, ANN <strong>and</strong> OTT. Water quality parameters for these waters were presentedpreviously in Chapter 2. The five samples were adjusted to a pH <strong>of</strong> 7.0 <strong>and</strong> buffered with 5 mL<strong>of</strong> 1 mM phosphate buffer. The reactor was constructed similar to the true-batch reactor, withthe exception <strong>of</strong> a glass water jacket. Initial tests were run on the reactor <strong>and</strong> cooling system toassess the control <strong>of</strong> temperature over time. Cooling water was recirculated through a coolerequipped with a recirculating pump. With this system, the resulting reactor sample could be65


maintained at 10 ±1°C for up to four hours. Temperature variations within the reactor wereobserved to vary less than 0.3°C along its height. <strong>Ozonation</strong> was performed in the same manneras the preceding true-batch experiments discussed in Section 4.1, <strong>and</strong> ozone kinetic data wererecorded over time.4.2.2 ResultsResults <strong>of</strong> the temperature-controlled true-batch ozonation experiments were used to assessbromate formation potential <strong>and</strong> estimate bromate concentrations based on increased CTrequirements for Cryptosporidium inactivation at various temperatures. Results from the 10°Cexperiments for the five water samples are presented in Table 4.6.ParameterInitial <strong>Bromide</strong>* (ug/L)Br(V(ug/L)Residual Br" (ug/L)Unaccounted Br"* (fig/L)Table 4.6: Results for True-Batch <strong>Ozonation</strong> at 10°CBr' to BrO3~ Conversion (%)Total Br" Incorporation (%)AUVA254 (cm'1)SUVA (L/mg-cm)OE (mg-min/L)AO3 initial (mg/L)Ozone Decay Rate (1(T3 sec'1)Dilution (%)AMA20020137347250.0190.7571.570.388SPW14511110165170.0331481.210.488Water SourceCCD394.3304.77200.0321.5140.951.155ANN655.242136280.0211.1311.180.7210OTT1715.88.94620.0221.1251.29* Initial <strong>Bromide</strong> concentration adjusted for dilution in calculations (i.e.unaccounted Br~ = (200*( 100-8)7100) - 137 - 20*(79.9/127.9) = 34 ng/L)4.2.2.1 <strong>Bromate</strong> <strong>Formation</strong> Potentials <strong>and</strong> Ozone ExposuresThe ozone kinetic data obtained from the 10°C experiments were expected to reflect decreasedozone decay rates. Significant decreases in the ozone decay rate were detected at this lowertemperature. For example, the decay rate was observed to decrease from 2.91xlO'3 sec' 1 at 20°Cto 7.2x10 sec' at 10°C for the ANN sample. Consequently, the OE values for a given ozonedose at 10°C were observed to be three to six-times higher when compared to true-batch resultsat 20°C. The highest observed OE was 57 mg-min/L for the AMA sample; the lowest was 14mg-min/L for CCD water.1.181066


<strong>Bromate</strong> formation potentials showed marked decreases as much as 54 percent for SPW (23|^g/L to 12 fJ.g/L) to 29 percent for OTT (1.4 ug/L to 1.0 ug/L). An average bromate reduction <strong>of</strong>46 percent was observed.The bromate <strong>and</strong> OE relationship was derived as presented in the previous Section 4.1. Thecalculation <strong>of</strong> the slope coefficient, K, was determined by fitting a linear function passingthrough the origin to the data point. A comparison <strong>of</strong> K values for the 10°C <strong>and</strong> 20°Cexperiments is presented in Table 4.7. The maximum K values for both temperatures wasobserved for AM A, while the minimum values were observed for OTT.Table 4.7: Effect <strong>of</strong> Temperature on True-Batch <strong>Formation</strong> <strong>of</strong> <strong>Bromate</strong>Temp 10°C20°CWaterSourceAMASPWCCDANNOTTBrO3"(Ug/L)20114.35.21.0True-Batch ExperimentOE(mgmin/L)57 0.3548 0.2214 0.3031 0.1725 0.04Slope, K*(mg/mg-min)Predicted<strong>Bromate</strong>for 2-Loghiact3.9-6.02.4 - 3.83.3. -5.11.8-2.80.4 - 0.7BrO3 "(Ug/L)True-Batch ExperimentOE(mgmin/L)18 1.9012 1.925.0 1.7011 1.004.5 0.3135238.511.31.4Slope, K*(mg/mg-min)Predicted<strong>Bromate</strong>for 2-LogInact4.4 - 9.94.4-103.9-8.82.3 - 5.20.7-1.64.2.2.2 <strong>Bromate</strong> <strong>Formation</strong> Potential <strong>and</strong> Cryptosporidium InactivationThe respective linear relationships representing bromate formation as a function <strong>of</strong> OE arepresented in Figure 4.10. The suggested ranges for 2-log Cryptosporidium inactivation at 10°C<strong>and</strong> 20°C by Oppenheimer et al. (2000) <strong>and</strong> Rennecker et al. (1999), have been included asreferences to assess the impact <strong>of</strong> increased CT requirements at lower temperatures. At 20°C,this figure shows bromate formation in SPW, AMA, <strong>and</strong> CCD close to or at the 10 ug/L MCLfor the higher CT requirements suggested by Rennecker et al. (1999). Even though the requiredCT increased dramatically to achieve the same level <strong>of</strong> inactivation once the temperature wasdecreased to 10°C, the level <strong>of</strong> bromate formation decreased just as dramatically. Using the highend <strong>of</strong> the suggested inactivation range presented in Figure 4.10, bromate formation in AMA haddecreased by 41 percent to 5.9 ug/L for the same level <strong>of</strong> inactivation. The other four watersalso underwent this dramatic reductions in bromate formation from 42 percent for CCD to 62percent for SPW.4.2.3 DiscussionIncreasing CT requirements will undoubtedly increase the potential for bromate formation.However, the impacts <strong>of</strong> temperature on bromate formation were characterized by a lower Kvalue, indicating less bromate formation per unit OE. The observations made in this analysisindicated that low bromide waters may exhibit trends that result in less bromate formation upon a67


decrease in temperature. As seasonal variations are taken into account, less bromate would beexpected to form during the winter than in the summer, based on this analysis. Ultimately, thebalance for inactivation must take into account that the strategy employed during winterconditions may result in exceeding the MCL if used during summer conditions. Thus, seasonalstrategies would be a prudent measure for facilities to maintain compliance with both pathogeninactivation <strong>and</strong> bromate formation requirements./•^K15SPW (20°C¥ AMA (20°C)y=1.92>/ y = 1.90xI.233aoPL,oa'•a12ANN (20°QCCD(20°C) 0= 1.70xAMA (10°C)y = 0.35xI2cc0 10 15 20OE (mg-min/L)25 30Figure 4.10: Ozone Exposure versus <strong>Bromate</strong> <strong>Formation</strong> Potential for 2-LogCryptosporidium Inactivation at 10 <strong>and</strong> 20°C4.3 SUMMARYThe influence <strong>of</strong> natural organic matter on the formation <strong>of</strong> bromate was illustrated through apair-wise linear regression analysis. The analysis related bromate formation potential to variousNOM characterization methods. The bromate formation potential was determined using a truebatchozone reactor operated under a st<strong>and</strong>ard set <strong>of</strong> testing conditions: 20°C; pH 7; 1 mg ozoneper mg DOC. The characterization <strong>of</strong> NOM was developed through measurements <strong>of</strong> DOC, UVabsorbance <strong>and</strong> SUVA. The NOM was also operationally fractionated into hydrophobic,transphilic <strong>and</strong> hydrophilic fractions using macroporous, nonionic resins. The apparentmolecular weight distribution <strong>of</strong> these organics was obtained through size exclusionchromatography.68


Through these analyses, the fourteen project waters evaluated ranged in DOC from 1.4 to 7.3mg/L. On average 50 percent <strong>of</strong> the DOC was classified as hydrophobic. The remainder was onaverage 29 percent hydrophobic <strong>and</strong> 21 percent hydrophilic. The SUVA values calculated forthese waters ranged from 1.4 to 3.8 L/mg-m.The effects <strong>of</strong> NOM on bromate formation were confounded by NOM effects on the ozonedem<strong>and</strong>, rate <strong>of</strong> ozone decay <strong>and</strong> concomitant disinfection (CT or OE). NOM properties such asSUVA appeared to have a negative correlation with both bromate formation <strong>and</strong> ozone exposure.Consequently, the obvious conclusion was that NOM removal before ozone application would bebeneficial, thereby, indicating advantages <strong>of</strong> intermediate-ozonation over pre-ozonation. Theresults presented in Section 4.1.4, however, indicated that different NOM fractions might haveeither inhibitory or promotional effects on bromate formation. As optimized conventionaltreatment preferentially removes hydrophobic (inhibitory effect) <strong>and</strong> some transphilic(promotional effect) compounds, the observed benefits <strong>of</strong> intermediate-ozonation bromateformation may be a result <strong>of</strong> a change in the ozone dem<strong>and</strong> <strong>and</strong> rate <strong>of</strong> ozone decay.The results from the true-batch reactor experiments, performed at 20°C <strong>and</strong> 10°C, indicatedabout an 80 percent reduction in bromate formation potential at the lower temperature. Due tothe sensitivity <strong>of</strong> Cryptosporidium inactivation to temperature, about three times as much ozonecontact would be required at 10°C as would be required at 20°C. Even with the increasedamount <strong>of</strong> ozone contact, approximately 40 percent less bromate was formed at 10°C for thesame level <strong>of</strong> Cryptosporidium inactivation. This result was surprising as it is commonlybelieved that more bromate would be formed at a lower temperature for a given level <strong>of</strong>Cryptosporidium inactivation.69


CHAPTER 5. EFFECT OF WATER QUALITY AND MINIMIZATION APPROACHESON BROMATE FORMATIONThis chapter focuses on bromate formation <strong>and</strong> minimization strategies investigated using benchscale<strong>and</strong> pilot-scale flow-through ozone contactors. Exp<strong>and</strong>ing on the results obtained usingbench-scale batch reactors, these hydraulically different flow-through ozone contactors wereused to evaluate the effects <strong>of</strong> increasing levels <strong>of</strong> Cryptosporidium inactivation, temperature,ammonia addition, pH depression, <strong>and</strong> hydroxyl radical scavenger addition on bromateformation. The trade<strong>of</strong>fs between bromate minimization <strong>and</strong> the formation <strong>of</strong> brominatedorganic compounds were also explored.5.1 BROMATE FORMATION5.1.1 Effect <strong>of</strong> Cryptosporidium InactivationOzone exposure values for a range <strong>of</strong> Cryptosporidium inactivation were developed specific tothe unique hydraulic characteristics <strong>of</strong> the laboratory-scale continuous-flow reactor, <strong>and</strong> arediscussed further in Chapter 8. Actual Cryptosporidium parvum experiments were performed toverify the relationship between ozone exposure <strong>and</strong> Cryptosporidium inactivation for thecontinuous-flow reactor. The results <strong>of</strong> these inactivation experiments are presented later inSection 8.2.For each <strong>of</strong>.the project waters, additional testing was performed to evaluate the effect <strong>of</strong>increasing levels <strong>of</strong> Cryptosporidium inactivation on bromate formation. The targeted ozoneexposure values used for the waters presented in Figure 5.1 corresponded to 1-, 2- <strong>and</strong> 3-logCryptosporidium inactivation at 15°C in the laboratory-scale continuous-flow reactor used forthe project. These data demonstrate that at pH 7 <strong>and</strong> 15°C, the majority <strong>of</strong> the low bromidewaters formed less than 7 ug/L bromate at 1-log Cryptosporidium inactivation. The two watersin which problematic levels <strong>of</strong> bromate formation was observed at this level <strong>of</strong> disinfection wereCCD <strong>and</strong> WPB, at 28 <strong>and</strong> 12 ug/L <strong>of</strong> bromate, respectively. It was anticipated that WPB mightform higher concentrations <strong>of</strong> bromate as it contained 170 ug/L bromide. Based on basic waterquality parameters <strong>and</strong> NOM fractionation, it was unclear why CCD formed significantly higherconcentrations <strong>of</strong> bromate as compared to other waters with similar water qualities.As the ozone exposure in these waters was increased to achieve 2- <strong>and</strong> then 3-logs <strong>of</strong>Cryptosporidium inactivation, an increase in bromate formation was observed each time.Increasing the inactivation <strong>of</strong> Cryptosporidium from 1- to 3- logs resulted in a 2.3 to 16-foldincrease in bromate formation. The lowest percent increase, from 0.7 to 1.6 ug/L, was observedin DAL water, <strong>and</strong> the largest percent increase, from 0.6 to 9.8 ug/L, was observed with HOUwater.For SPW, CRW <strong>and</strong> SAC waters presented in Figure 5.2, a range <strong>of</strong> Cryptosporidiuminactivation between 0.5-log <strong>and</strong> 1.5-log was evaluated. The measured bromate formation for 1-log Cryptosporidium inactivation in each <strong>of</strong> these three waters at the higher pH <strong>of</strong> 8, was higher70


than the 10 ug/L Stage 1 D/DBP Rule MCL. The concentrations <strong>of</strong> bromate formed rangedbetween 24 ug/L for SAC water (initially spiked to 47 jig/L bromide) <strong>and</strong> 38 ug/L for SPWwater. While reducing the ozone exposure reduced bromate formation to in each <strong>of</strong> the waters,SPW still formed 12 ng/L bromate at 0.5-log Cryptosporidium inactivation. Underst<strong>and</strong>ably, theinitial 145 fig/L bromide concentration in SPW contributed to this elevated level <strong>of</strong> bromate.CRW water <strong>and</strong> the bromide spiked SAC water (47 ug/L bromide) formed less than the 10 |ig/LMCL, at 3.2 <strong>and</strong> 9.5 ug/L bromate, respectively.Log Cryptosporidium InactivationFigure 5.1: Impact <strong>of</strong> Ozone Exposure, Expressed as Log-Cryptosporidium Inactivation,on <strong>Bromate</strong> <strong>Formation</strong> in the Laboratory-Scale Continuous-Flow Ozone Contactor (pH 7,15°C)Figure 5.3 shows percent conversion <strong>of</strong> bromide to bromate in the laboratory-scale continuousflowozone contactor for selected waters. By normalizing bromate formation by the initialbromide concentrations, differences in the propensity <strong>of</strong> the different waters to form bromate canbe easily identified. Most notably at pH 7 <strong>and</strong> 15°C, only between 2 <strong>and</strong> 5 percent <strong>of</strong> thebromide in the DAL water was converted to bromate at 1- <strong>and</strong> 3-log Cryptosporidiuminactivation, respectively. It is likely that the 0.53 mg/L ambient ammonia-nitrogenconcentration was sufficient to minimize bromate formation under these conditions. Thepotential benefits <strong>of</strong> ammonia addition are discussed later in Section 5.2.2. At the other extreme,approximately 100 percent <strong>of</strong> the bromide was converted to bromate in the CCD water sample at3-logs Cryptosporidium inactivation. It is possible that even at pH 7, initial concentrations <strong>of</strong>ammonia-nitrogen (0.06 mg/L) <strong>and</strong> DOC (1.8 mg/L) were not high enough to provide sinks forbromide. Consequently, all <strong>of</strong> the bromide was available for conversion to bromate, <strong>and</strong> did soonce a sufficient amount <strong>of</strong> ozone contact had been provided. For the rest <strong>of</strong> the waters, thepercent bromide conversion to bromate ranged between 2 <strong>and</strong> 21 percent <strong>and</strong> between 20 <strong>and</strong> 60percent at 1-log <strong>and</strong> 3-logs <strong>of</strong> Cryptosporidium inactivation, respectively.71


0.5 1.0Log Cryptosporidium Inactivation1.5Figure 5.2: Impact <strong>of</strong> Ozone Exposure, Expressed as Log-Cryptosporidium Inactivation,on <strong>Bromate</strong> <strong>Formation</strong> in the Laboratory-Scale Continuous-Flow Ozone Contactor (pH 8,15°C)-•-CCD-0-WPB-A- LAW-&-NJA-•-ANN-•-OTT-0-DALLog Cryptosporidium InactivationFigure 5.3: Impact <strong>of</strong> Ozone Exposure, Expressed as Log-Cryptosporidium Inactivation,on the Percent Conversion <strong>of</strong> <strong>Bromide</strong> to <strong>Bromate</strong> (pH 7,15°C)72


5.1.2 Temperature EffectsAs previously stated, both the rate <strong>of</strong> ozone decay <strong>and</strong> the sensitivity <strong>of</strong> Cryptosporidium toozone are dependent on the temperature <strong>of</strong> the system. The true-batch ozone decay experimentsperformed at 20°C <strong>and</strong> 10°C had indicated significant decreases in the ozone decay rate asdecreases in the temperature were made. These experiments also predicted lower levels <strong>of</strong>bromate formation at the lower temperature for the same predicted level <strong>of</strong> Cryptosporidiuminactivation. However, the opposite trend in bromate formation was observed using thelaboratory-scale continuous-flow reactor. For a given level <strong>of</strong> Cryptosporidium inactivation,decreasing temperatures generally resulted in increasing amounts <strong>of</strong> bromate formation, asshown in Figure 5.4 <strong>and</strong> Figure 5.5.-•-WPB-0-CCD-if- LAW-•-NJA-&- ANN-CHHOU10 15 2025 30Temperature (°C)Figure 5.4: Effect <strong>of</strong> Temperature on <strong>Bromate</strong> <strong>Formation</strong> (2-Log CryptosporidiumInactivation, pH 7)Since the inactivation kinetics <strong>of</strong> Cryptosporidium parvum are temperature dependent, differentamounts <strong>of</strong> ozone exposure were required to achieve the same level <strong>of</strong> inactivation at differenttemperatures. The specific levels <strong>of</strong> ozone exposure for each temperature are discussed inSection 8.2. At pH 7 <strong>and</strong> a predicted 2-logs <strong>of</strong> Cryptosporidium at each temperature evaluated,increasing the temperature from 5°C to 25°C resulted in a 28 to 85 percent reduction in bromateformation for most <strong>of</strong> the waters, as shown in Figure 5.4. The corresponding HOU analysesindicated the opposite trend with a 113 percent increase in bromate formation for the sameincrease in reaction temperature. Compared to the other waters, HOU had a high DOC (9.4mg/L), a correspondingly rapid ozone decay curve (Figure 4.6) <strong>and</strong> the lowest alkalinity (28mg/L). However, only NJA was impacted by the temperature with respect to compliance withthe 10 ng/L MCL. For the rest <strong>of</strong> the waters, those that formed bromate in excess <strong>of</strong> the MCL73


emained above the MCL at all temperatures evaluated, <strong>and</strong> for those waters that did not have aproblem with bromate formation remained that way, for this level <strong>of</strong> inactivation.15012090<strong>of</strong>a3OSOpq603000 10 15 20Temperature (°Q25 30Figure 5.5: Effect <strong>of</strong> Temperature on <strong>Bromate</strong> <strong>Formation</strong> (1-Log CryptosporidiumInactivation, pH 8)At pH 8 <strong>and</strong> sufficient ozone exposure to inactivate a predicted 1-log <strong>of</strong> Cryptosporidium at eachtemperature evaluated, similar trends were observed. The percent reduction in bromateformation between 5°C <strong>and</strong> 25°C ranged between 17 <strong>and</strong> 83 percent for the three waters shownin Figure 5.5. Even at 25°C, all <strong>of</strong> these waters were in excess <strong>of</strong> the 10 ug/L MCL.The opposing trends observed between the true-batch reactor <strong>and</strong> the laboratory-scalecontinuous-flow ozone contactor illustrated the complexity <strong>of</strong> comparing effects on bromateformation between hydraulically different systems. The trends in temperature effects on bromateformation appear to have been significantly influenced by the hydraulic characteristics <strong>of</strong> thesystem controlling the ozone exposure requirements. Increased ozone exposure will undoubtedlyincrease the potential for bromate formation, but as seasonal variations are taken into account itbecomes unclear from this analysis if more or less bromate would be expected during the winteras compared to the summer. The results from the full-scale ozone facility survey presented inChapter 3 were also unclear. While a slightly lower average bromate level was observed duringthe winter, differences in water quality <strong>and</strong> levels <strong>of</strong> inactivation were recorded.As temperature control cannot economically be used as a bromate formation control strategy <strong>and</strong>seasonal differences in water quality generally exist for those waters that experience significanttemperature variation, utilities might be best suited to develop site-specific seasonal bromateformation information using representative water qualities. The data presented here illustrate,74


however, that this testing should be performed in a system hydraulically similar to the utility'santicipated or existing ozonation system.5.2 BROMATE MINIMIZATION5.2.1 pH DepressionIt has been reported that the depression <strong>of</strong> pH to below 7 prior to ozonation will result in thereduction in bromate formation as compared to bromate formation in water above pH 7 (e.g.,Song, 1996). This reduction has been attributed to two factors: at pH less than 7, oxidizedbromide will primarily be found as hypobromous acid (HOBr), limiting the amount <strong>of</strong>hypobromite (OBr~) available for reaction with ozone (HOBr 4* OBr" + H+; pKa = 8.7); at thesedepressed pHs, the generation <strong>of</strong> hydroxyl radicals is reduced, resulting in more stable ozoneresiduals, limiting the amount <strong>of</strong> bromate formed through the hydroxyl radical related bromateformation pathways. The increased ozone stability at the depressed pH has the added benefit <strong>of</strong>requiring a lower initial ozone dose to achieve the same level <strong>of</strong> disinfection as compared to theozone dose required at ambient pH. At lower pH, the ratio <strong>of</strong> hydroxyl radical, the main oxidantfor bromate formation in natural water to ozone tends to be lower than at higher pH (Eloviz etal., 1999b).Each <strong>of</strong> the waters was again evaluated using the laboratory-scale continuous-flow ozonecontactor. As anticipated, reductions in bromate formation were observed as the pH wasdepressed from 8 to 7 to 6, as shown in Figure 5.6 <strong>and</strong> Figure 5.7. The reduction <strong>of</strong> bromate perunit decrease <strong>of</strong> pH, generally ranged between 30 to 50 percent. The decrease in bromateformation generally followed a linear trend for the range <strong>of</strong> pH values evaluated. It is importantto note that while significant reductions in bromate formation were realized through pHdepression, the concentrations <strong>of</strong> bromate formed at pH 6 for 2-logs <strong>of</strong> Cryptosporidiuminactivation in several <strong>of</strong> the project waters still exceeded the 10 ug/L MCL. In CCD, WPB <strong>and</strong>LAW waters 17, 16 <strong>and</strong> 11 ug/L bromate was formed. The other waters presented in Figure 5.6,formed less than 5 ug/L under these conditions. At pH 6 <strong>and</strong> 1-log Cryptosporidiuminactivation, less than 6 ug/L bromate formed in SPW, CRW <strong>and</strong> SAC waters. This indicatedthat pH reduction appears to be a successful approach to mitigate bromate formation for manywaters. However, in a few waters, it may be necessary to implement additional or alternativebromate formation reduction strategies to ensure the MCL is not exceeded at the higher levels <strong>of</strong>CT required for Cryptosporidium inactivation.While lowering the pH helps to mitigate the problematic formation <strong>of</strong> bromate, this bromateminimization strategy appears to enhance the formation <strong>of</strong> various brominated organiccompounds. The concentration <strong>of</strong> TOBr in the ozonated samples was measured to determine thesignificance <strong>of</strong> pH depression on TOBr formation. The concentration <strong>of</strong> TOBr formed duringthe experiments tended to be quite low. The typical range <strong>of</strong> TOBr formation was from belowthe 4 ug/L as Br" detection limit to 20 ug/L as Br", as shown in Figure 5.8. WPB water hadhigher TOBr values in the range <strong>of</strong> 25 to 35 ug/L as Br", which was likely due to the high DOC(10.4 mg/L) <strong>and</strong> bromide (170 ug/L) concentrations. For utilities looking at pH depression tocontrol bromate, TOBr may become problematic if it becomes a regulated class <strong>of</strong> compounds.75


I•o12090-WPB -0-CCD-LAW -0-ANN-NJA -&-HOU-OTTu.aCS60O03678PHFigure 5.6: Effect <strong>of</strong> pH on <strong>Bromate</strong> <strong>Formation</strong> (2-Log Cryptosporidium Inactivation,15°C)-•-ACD-D-SPWSAC-A-CRW678pHFigure 5.7: Effect <strong>of</strong> pH on <strong>Bromate</strong> Formadon (1-Log Cryptosporidium Inactivation,15°C)76


DC-•-CCD3•o


elevated further to 1.0 mg/L as N, though it was not as significant as the net change observedwith the spike to 0.5 mg/L ammonia-nitrogen.WPB-A- LAW-A-NJA-•-ANN-D-HOU-0-DAL00.5Ammonia Concentration (mg/L-N)Figure 5.9: Effect <strong>of</strong> Ammonia Addition on <strong>Bromate</strong> <strong>Formation</strong> (2-Log CryptosporidiumInactivation, 15°C, pH 7)&3•oI ota2cc0 0.5Ammonia Concentration (mg/L-N)Figure 5.10: Effect <strong>of</strong> Ammonia Addition on <strong>Bromate</strong> <strong>Formation</strong> (1-Log CryptosporidiumInactivation, 15°C, pH 8)78


At 2-logs <strong>of</strong> Cryptosporidium inactivation <strong>and</strong> pH 7, ammonia addition to 0.5 mg/L as N lead toa reduction in bromate formation <strong>of</strong> approximately 50 percent in LAW, NJA <strong>and</strong> ANN waters.This level <strong>of</strong> reduction was sufficient for these waters to reduce bromate formation to below the10 ng/L MCL. For CCD water, the 30 percent reduction in bromate formation that correspondedto the 0.5 mg/L ammonia-nitrogen, was not sufficient to reduce the bromate concentration tobelow 10 ug/L. Ammonia addition did not appear to significantly impact those waters withbromate formation already below 10 ug/L.At 1-log Cryptosporidium inactivation <strong>and</strong> pH 8, the elevated 0.5 mg/L ammonia-nitrogenconcentration resulted in an 88 <strong>and</strong> 69 percent reduction in bromate formation for CRW <strong>and</strong>bromide spiked SAC (47 ug/L Br") waters, respectively. This was sufficient to reduce the finalbromate concentration to below 10 ug/L. In SPW a 58 percent reduction in bromate formationwas observed at the 0.5 mg/L ammonia-nitrogen concentration. While increasing the ammonianitrogenconcentration to 1.0 mg/L resulted in a total reduction in bromate formation <strong>of</strong> 70percent, about 12 ug/L bromate was formed.As the addition <strong>of</strong> ammonia can provide a temporary bromide sink to reduce bromate formation,additional analyses were performed to determine if this effect would extend to the formation <strong>of</strong>brominated organic compounds. Unfortunately, only four samples yielded TOBr concentrationsat or above the minimum reporting level. Consequently, it was not possible to determine ifammonia addition had an effect on TOBr formation.5.2.2.1 Pilot-Scale Plant InvestigationAmmonia was used at the Britannia pilot plant (Ottawa, Canada) to reduce bromate formationduring ozonation. The pilot-scale ozone contactor consisted <strong>of</strong> three glass columns in series, asdescribed in Section 2.2.5. The water used for these experiments was adjusted to promotebromate formation by raising the pH from 6.1 to 8.5, <strong>and</strong> spiking bromide to 0.1 mg/L, asdescribed in Section 2.2.5. It was necessary to take this step since no bromate formation wasobserved under ambient conditions (pH 6.1, 0.03 mg/L Br", 0.5°C, 30 mg-min/L CT).Three separate trials were conducted to evaluate ammonia addition as a bromate minimizationstrategy at pilot-scale. The initial trial served as the control without ammonia addition (ambientammonia < 0.05 mg/L). Ammonia (0.05 <strong>and</strong> 0.1 mg/L as N) was added to the second <strong>and</strong> thirdtrials. The data shown in Figure 5.11 suggest that adding as little as 0.05 mg/L <strong>of</strong> ammoniaresulted in about a 50 percent reduction in bromate concentrations in the water leaving the ozonecontactor. Doubling the amount <strong>of</strong> ammonia applied resulted in a smaller marginalimprovement, reducing bromate by an additional 20 percent. It is likely that the addition <strong>of</strong> morethan 0.1 mg/L <strong>of</strong> ammonia would continue the reduction in bromate formation, but only by asmall amount.5.2.2.2 ConclusionThe important conclusion from these ammonia addition experiments was that the application <strong>of</strong> arelatively small amount <strong>of</strong> ammonia could result in significant reductions in bromate formation.The magnitude <strong>of</strong> this benefit, however, was not consistent for all <strong>of</strong> the waters tested. Since79


many water treatment plants already use ammonia for chloramination, ammonia may be a readilyavailable method for reducing bromate formation without sacrificing CT targets. The presence<strong>of</strong> ammonia can interfere with the bromate formation pathway by temporarily scavenginghypobromous acid <strong>and</strong> forming bromamine species. These species are subsequently oxidized byozone, freeing the bromide so that it could again be oxidized to hypobromite ion/hypobromousacid. This mechanism can limit the availability <strong>of</strong> hypobromite ion/hypobromous acid until all<strong>of</strong> the ammonia is consumed. As an additional benefit, the limited availability <strong>of</strong> hypobromiteion/hypobromous acid due to ammonia addition, also appears to reduce TOBr formation.5040.1 30•^«I 20Ammonia Concentration-*- O.05 mg/L as N-a- 0.05 mg/L as N-•—0.1 mg/L as N10Column 1 Column 2 ColumnsFigure 5.11: Effect <strong>of</strong> Ammonia on <strong>Bromate</strong> <strong>Formation</strong> (Ottawa Pilot Plant, 0.5°C, pH 8.5,0.1 mg/L BO5.2.3 Cumulative Effects <strong>of</strong> Ammonia Addition <strong>and</strong> pH DepressionAs the stability <strong>of</strong> the bromamines is relative to the particular species formed, which primarilydepends on the pH, ammonia concentration, <strong>and</strong> bromide concentration, additional tests wereperformed to evaluate the effects <strong>of</strong> both pH adjustment <strong>and</strong> ammonia addition. A full factorialset <strong>of</strong> experiments for pHs 6, 7, <strong>and</strong> 8, <strong>and</strong> ammonia concentrations <strong>of</strong> ambient, 0.5, <strong>and</strong> 1.0mg/L as N were conducted on CCD <strong>and</strong> SPW waters. The results <strong>of</strong> these experiments arepresented in Figure 5.12 <strong>and</strong> Figure 5.13. Significant reductions in bromate formation wereobserved in both waters for the strategies employing both pH depression <strong>and</strong> ammonia addition.A cumulative effect was observed as pH depression <strong>and</strong> ammonia addition reduced bromateformation more than either alone. For example, in SPW water at pH 8 without ammoniaaddition, 38 ng/L bromate was formed. When the pH was depressed to 6 or 1.0 mg/L ammonia80


was added separately, 5 <strong>and</strong> 12 ng/L bromate was formed, respectively. If the pH was depressedto 6 <strong>and</strong> 1.0 mg/L ammonia was added less than 1 jig/L bromate was formed.1ota.8832MpHOmg/L-N0.5 mg'L-N6 l.Omg^-N AmmoniaAdditionFigure 5.12: Effect <strong>of</strong> pH Depression <strong>and</strong> Ammonia Addition on <strong>Bromate</strong> <strong>Formation</strong> inCCD (1-Log Cryptosporidium Inactivation, 15°C)aoto asOfacs£10Omg/L-N1. Omg/L-NAmmoniaAdditionFigure 5.13: Effect <strong>of</strong> pH Depression <strong>and</strong> Ammonia Addition on <strong>Bromate</strong> <strong>Formation</strong> inSPW (1-Log Cryptosporidium Inactivation, 15°C)81


5.2.4 Hydroxyl Radical Scavenger AdditionThe addition <strong>of</strong> acid or ammonia has focused on bromate formation pathways requiringhypobromite. Even with these bromate minimizing strategies in place, bromate formation cancontinue to proceed through hydroxyl radical initiated pathways. It is likely that the addition <strong>of</strong> ahydroxyl radical scavenger, such as tertiary butanol, could limit bromate formation reactions thatrequire hydroxyl radicals. A subset <strong>of</strong> waters was tested with tertiary butanol additions <strong>of</strong> 1 <strong>and</strong>3 mM, as shown in Figure 5.14. <strong>Bromate</strong> formation was, for practical purposes, completelyinhibited in the presence <strong>of</strong> 1 mM tertiary butanol. Obviously, higher concentrations <strong>of</strong> tertiarybutanol did not provide any additional benefit. These initial experiments indicate there ispotential for the minimization <strong>of</strong> bromate formation through the addition <strong>of</strong> a hydroxyl radicalscavenger. Additional experiments should be performed to determine the minimum amount <strong>of</strong>scavenger addition needed to effectively inhibit bromate formation <strong>and</strong> identify hydroxyl radicalscavengers suited for use in drinking water treatment plants. As the mechanism <strong>of</strong>Cryptosporidium inactivation is still being investigated, it is not clear if the scavenging <strong>of</strong>hydroxyl radicals would adversely impact its inactivation kinetics.3 40<strong>of</strong>e£cs30-•-WPB (pH7)-D-SPW (pH8)-A-CRW(pH8)-A-NJA (pH7)-O-DAL (pH7)o201000.1 2 3t-BuOH (mM)Figure 5.14: Effect <strong>of</strong> a Radical Scavenger (Tertiary Butanol) Addition on <strong>Bromate</strong><strong>Formation</strong> in the Laboratory-Scale Continuous-Flow Ozone Contactor (15°C)5.3 SUMMARYAs expected the higher levels <strong>of</strong> CT required for Cryptosporidium inactivation resulted in higherlevels <strong>of</strong> bromate formation. Based on the results presented in this chapter, however, bromate82


formation should not be a problem for many utilities utilizing low bromide (


formation observed in these laboratory-scale continuous-flow ozone contactors indicated thatammonia addition by itself might only be sufficient for those utilities within 10 to 20 jig/L <strong>of</strong> the10 jug/L MCL. However, it was also shown that together, the cumulative bromate reductioneffects <strong>of</strong> ammonia addition <strong>and</strong> pH depression could be used to control bromate formation inproblematic waters.It is important to note that these bromate minimization strategies also appeared to affect theformation <strong>of</strong> brominated organic compounds. While pH depression clearly enhanced TOBrformation, ammonia addition appeared to reduce TOBr formation. Should TOBr become aregulated group <strong>of</strong> compounds in the future, utilities ozonating under low pH conditions by mayhave to consider alternative bromate minimization strategies.84


CHAPTER 6. HYDRODYNAMIC IMPACTS ON BROMATE FORMATIONTo determine if hydrodynamic conditions (staged/tapered versus single-stage; counter-currentversus co-current) would significantly impact bromate formation, a series <strong>of</strong> true-batch <strong>and</strong> pilotscaleexperiments were performed. Staged ozonation was evaluated with both <strong>of</strong> the scales <strong>of</strong>reactors, while the flow configuration was only evaluated at pilot-scale. It was hypothesized thatif a good correlation could be developed between ozone contact (CT or OE) <strong>and</strong> bromateformation, regardless <strong>of</strong> the hydrodynamic configuration, then the tests would indicate thechemical conditions were more important than the hydrodynamic conditions for bromateformation. If significant differences were observed between the different hydrodynamicconditions then it would be expected that the hydrodynamics <strong>of</strong> the ozone contactor could beused as a bromate formation minimization strategy.Krasner et al. (1993) studied the effect <strong>of</strong> staging ozone application within a contactor, i.e. ozonedosed in more than one cell, to minimize ozone residual <strong>and</strong>, thus, potentially reduce bromateformation; no appreciable difference in bromate formation was observed with staged ozonation.Targeting specific ozone residual concentrations, an ozone contactor was dosed in three separateexperiments: (1) single stage ozonation with a dose <strong>of</strong> 1.4 mg/L, (2) staged ozonation with 1.0mg/L added in the first stage <strong>and</strong> 0.4 mg/L added in the second stage, <strong>and</strong>, (3) staged ozonationwith 1.4 mg/L added in the first stage <strong>and</strong> 0.5 mg/L added in the second stage to meet CTrequirements for direct filtration. Analyses revealed bromate formations <strong>of</strong> 5 ug/L,


6.1 TRUE-BATCH STAGED OZONATIONStaged ozone tests were performed in a true-batch reactor on selected waters: NJA, DAL, SPW,ANN, <strong>and</strong> OTT. Samples were prepared for the single-stage true-batch experiments by initiallybuffering with 1 mM phosphate solution <strong>and</strong> adjusting to a pH <strong>of</strong> 7.0. A total ozone dose <strong>of</strong> 1:1mg 63 to mg DOC was applied to the water in two stages, with half <strong>of</strong> the total ozone applied toeach stage. The ozone residuals were allowed to decay below the minimum detection limit aftereach stage <strong>of</strong> ozone addition, at which point, aliquots <strong>of</strong> the ozonated samples were collected forbromate analysis. Kinetic data were used to calculate the ozone exposure for each stage.Additional discussion <strong>of</strong> the staged ozonation method is outlined in Section 2.2.1.2.The concentration <strong>of</strong> bromate formed in each <strong>of</strong> the staged ozonation experiments is presented inTable 6.1. The reported bromate concentrations were based on the cumulative bromateformation during each stage <strong>of</strong> ozonation. For example, the stage two results included thebromate formed in the first stage plus the additional bromate formed during the second stage.The residual bromide concentrations were measured after each stage. Cumulative bromateformation (stage 1 + stage 2) was observed to range from 0.7 ug/L for OTT water to 28 ug/L forSPW water. As little as 0.1 ug/L, or 14 percent <strong>of</strong> the final bromate concentration, was formedduring the second stage (OTT). The maximum second stage formation occurred in SPW water,where the bromate formed was calculated to be 15 ug/L, accounting for 53 percent <strong>of</strong> the totalamount <strong>of</strong> bromate formation.The effect <strong>of</strong> staged ozonation on bromide concentrations must be taken into account whenconsidering the second stage bromate formation. As the first stage was completed, the bromideconcentration was effectively reduced because <strong>of</strong> oxidation to bromate <strong>and</strong> reactions with NOMto form TOBr. Thus, when the second ozone dose was applied, there was less bromide availablefor reaction. It is possible that the dilution <strong>of</strong> the water by the addition <strong>of</strong> the ozone stocksolution may have also affected the subsequent formation <strong>of</strong> bromate.Ozone exposure values for each stage were calculated separately <strong>and</strong> summed to provide acumulative OE value for the second stage. The maximum observed cumulative OE was 10 mgmin/Lfor SPW; the minimum was 4.9 mg-min/L for NJA. The second stage OE values weregreater than the first stage values for all the waters. The reduction <strong>of</strong> the ozone dem<strong>and</strong> duringthe first stage (i.e. oxidation <strong>of</strong> highly reactive hydrophobic compounds) was expected toincrease OE for the second stage. The second stage values <strong>of</strong> OE accounted for more than 55percent <strong>of</strong> the total OE for all <strong>of</strong> the waters. For each <strong>of</strong> the waters, the conversion <strong>of</strong> bromide tobromate (calculated as a function <strong>of</strong> initial bromide concentration adjusted for dilution) wasnearly equivalent for each stage. For example, ANN showed a conversion <strong>of</strong> bromide tobromate 4.1 percent <strong>and</strong> 8.0 percent, respectively.86


ooTable 6.1: <strong>Bromate</strong> <strong>Formation</strong> Comparison Between Bench-Scale True-Batch Staged <strong>and</strong> Single-Stage <strong>Ozonation</strong> (20°C)WaterNJADALSPWANNOTTStage Cumulative Initial* Residual CumulativeOE <strong>Bromide</strong> <strong>Bromide</strong> <strong>Bromate</strong>(mg-min/L) (ug/L) (ug/L) QigfL)12Single12Single12Single12Single12Single1.04.93.54.19.56.92.810123.89.0111.95.24.521 14131521 17149.7145 1331106465 58535817 12116.31.02.01.50.61.02.41328234.17.6110.60.71.4Unaccounted<strong>Bromide</strong>(Hg/L)4.63.75.11.52.89.80.00.066.61.81.50.03.43.79.8Modeled TOBrt(jig/L as Br")0.040.030.100.030.020.101.21.01.90.250.220.700.020.020.10Total Br to BrO3"Conversion*(%)* undiluted** "0.0" value indicates that the calculated bromide recovery is greater than the estimated unozonated water bromide concentration* TOBr estimated using Song (1996) TOBr Model where: [TOBr]=1.004(Br)' 765(DOCy0551(pHy3916(NH3-N)° ^(Os) 1135(TIC)'0153* estimate <strong>of</strong> all bromide species other than bromide; adjusted for dilution by ozone stock solution addition3.26.95.12.03.68.36.014114.18.0122.32.96.0Total <strong>Bromide</strong> DilutionIncorporation*(%) (%)272719102046114517113242759714121017147121049851110


Table 6.2 lists the bromate formation <strong>and</strong> staged ozone exposure relationships developed foreach <strong>of</strong> the water samples. The relationship for each individual stage was first calculated <strong>and</strong>identified as Kstage with KI representing stage one values <strong>and</strong> Ka representing stage two values.The KI values were calculated by assuming a linear relationship between the origin (zerobromate formation at zero OE) <strong>and</strong> the first stage bromate concentration <strong>and</strong> OE measurement.The Ka value was calculated as the slope <strong>of</strong> the linear relationship between the first <strong>and</strong> thesecond stage bromate concentrations <strong>and</strong> the first <strong>and</strong> second stage measured values <strong>of</strong> OE. TheKa value was consistently less than the KI value, indicating that less bromate formed per unit OEduring the second stage. It should be noted that the available bromide concentrations <strong>of</strong> thesecond stage were lower due to bromate <strong>and</strong> TOBr formation <strong>and</strong> the addition <strong>of</strong> ozone stocksolution during the first stage. Consequently, the slope coefficients for the second stage werereflective <strong>of</strong> the decreased bromide availability.Table 6.2: <strong>Bromate</strong> <strong>Formation</strong> Potential For Bench-Scale Batch Staged <strong>Ozonation</strong>Water Stage <strong>Bromate</strong> Cumulative OE Kstage* Overall K** K^*(mg-min/L)NJADALSPWANNOTT12121212121.02.00.61.013284.17.60.60.71.04.94.19.52.810.13.89.01.95.21.000.260.150.074.642.051.080.670.320.030.43—0.11—2.91—0.88—0.16—0.43—0.35—1.92—0.97—0.22~* Slope value, K (ug/mg-min), is based on diluted sample bench-scale experiments. KI value calculatedas slope <strong>of</strong> origin <strong>and</strong> Stage 1 OE <strong>and</strong> BrO 3". K2 value calculated as slope between the Stage 1 <strong>and</strong> Stage2 data. KSmgie value calculated as slope between the origin <strong>and</strong> the non-staged single ozone dose OE <strong>and</strong>BrO3" value.** Overall K value is the slope <strong>of</strong> the best fit line between the origin <strong>and</strong> the Stage 2 OE <strong>and</strong> BrO3" value.Figure 6.1 graphically presents the cumulative bromate formation results <strong>of</strong> the staged ozonationexperiments compared to the single dose true-batch experiments initially presented in Section4.1.3.2. The comparison indicated inconclusive results with respect to the reduction <strong>of</strong> bromateformation by staged ozonation. An increase in the level <strong>of</strong> bromate formed was observed forSPW water <strong>and</strong> no effect was observed in NJA water. For the rest <strong>of</strong> the waters, the low levels<strong>of</strong> bromate formed during either staged or single-stage application <strong>of</strong> ozone limited any furthercomparison.88


•SPW Staged• ANN Staged• NJA Staged•DAL Staged• OTT StagedSPW Single StageANN Single Stage-0- - NJA Single Stage-Q- - DAL Single StageX OTT Single Stage0 12OE (mg-min/L)Figure 6.1: Impact <strong>of</strong> Staged <strong>Ozonation</strong> on Bench-Scale Batch <strong>Bromate</strong> <strong>Formation</strong>6.2 PILOT-SCALE STAGED OZONATIONPilot-scale experiments conducted at the Britannia Water Purification Plant (Ottawa, Canada)were used to explore contactor configuration options for minimizing bromate. As a result <strong>of</strong> thepilot work being performed during the winter, all <strong>of</strong> the pilot-scale tests were performed at awater temperature <strong>of</strong> 0.5°C. Preliminary experiments with the ambient water (pH 6.1, 17 (ig/Lbromide) revealed that at this temperature <strong>and</strong> pH, no measurable bromate formed even at veryhigh CT levels. To obtain measurable bromate formation, the pH <strong>of</strong> the water was raised to 8.5by NaOH addition, <strong>and</strong> the bromide concentration was raised to 100 |ig/L using KBr. While it isacknowledged that the intent <strong>of</strong> this research was to investigate bromate formation in lowbromide waters, it was decided that the objective <strong>of</strong> determining the impacts <strong>of</strong> contactorconfigurations took precedence for these tests. Furthermore, the need to artificially raise the pH<strong>and</strong> bromide levels to obtain measurable bromate formation at high CT values in very cold waterwas an important consideration in itself. It suggests that bromate formation at low pH (6.1),bromide (17 [ig/L) <strong>and</strong> temperature (0.5°C), would not be a problem for Ottawa River watereven when ozonating to CT values as high as 50 mg-min/L.To evaluate staged ozonation, ozone was applied to the first two columns by splitting a fixed gasflow from the ozone generator. By varying the split ratio <strong>of</strong> the ozone gas, different residualozone concentration pr<strong>of</strong>iles were developed through the ozone contactor, as shown in Figure6.2. Both contactors were operated with ozone applied in counter-current mode. The data showthat the residual ozone concentration exiting the first column was approximately proportional tothe fraction <strong>of</strong> the total ozone applied to the column. In the two subsequent columns, however, aconvergence <strong>of</strong> the residual ozone measured for each <strong>of</strong> the four ozone split ratios was observed.89


In particular, the ozone residuals at the effluent <strong>of</strong> the third column were similar for the fourscenarios, all in the range <strong>of</strong> 2.0 to 2.4 mg/L. This trend might be expected because the coldwater (0.5°C) <strong>and</strong> relatively low DOC (2.6 mg/L) ensured a slow rate <strong>of</strong> ozone decay. Underthese conditions, applying ozone in two stages in rapid succession instead <strong>of</strong> one stage wouldappear to result in approximately the same total downstream ozone concentration, as wasobserved in the third column effluent.4r 3«5S 'tt>6 24>c o§ 1ao>O 3 Split Ratio (%/%)75/25 Col 1&250/50 Col 1&225/75 Col 1&2W 0Column 1 Column 2 ColumnSFigure 6.2: Ozone Pr<strong>of</strong>ile Through Britannia Pilot Ozone Contactor <strong>During</strong> Staged<strong>Ozonation</strong> (pH 8.5,0.5°C)With little ozone decay evident, adding more ozone in the first contactor resulted in higher totalCT values due to greater contact time, as seen in Figure 6.3. Cryptosporidium inactivation usinga given amount <strong>of</strong> ozone would therefore be maximized in this system by applying 100 percentto the first column. (It should be noted that due to the cold water <strong>and</strong> the height <strong>of</strong> the column,ozone mass transfer was not limited in this system when applying all <strong>of</strong> the ozone to onecolumn.) As shown in Figure 6.4, while CJwas greatest when adding all <strong>of</strong> the ozone to the firstcolumn, bromate formation was also maximized when using this approach. In fact, bymodifying Figure 6.4 to plot bromate against the cumulative CT through each column <strong>of</strong> theozone contactor, a consistent correlation between bromate <strong>and</strong> CT was observed regardless <strong>of</strong> thesplit ratios used, except perhaps for the "100% ozone in column 1" scenario. However, even forthe "100%" scenario, the effluent <strong>of</strong> the ozone contactor yielded a CT<strong>and</strong> bromate concentrationconsistent with the other split ratios.Based on these pilot-scale results presented here <strong>and</strong> the bench-scale results presented in Section6.1, there is no strong evidence suggesting that splitting the ozone dose between multiple cells inan ozone contactor would enable bromate formation to be reduced while maintaining the sameCT. The data in Figure 6.5 indicate that with very little variation, there exists a unique bromateconcentration for a given CT value, regardless <strong>of</strong> the split ratio used.90


25.3EwoSsU2015100Column 1 Column 2 ColumnSO3 Split Ratio (%/%)-•-100% in Coll-o- 75/25 Col 1 & 2-•-50/50 Col 1&2-^25/75 Col 1&2Figure 6.3: Cumulative CT Through Britannia Pilot Ozone Contactor <strong>During</strong> Staged<strong>Ozonation</strong> (pH 8.5,0.5°C)Q 3 Split Ratio (%/%)-•-100% in Coll-°-75/25 Col 1&2-•-50/50 Col 1&2-*- 25/75 Col 1&2Column 1 Column 2 ColumnSFigure 6.4: <strong>Bromate</strong> <strong>Formation</strong> Through Britannia Pilot Ozone Contactor <strong>During</strong> Staged<strong>Ozonation</strong> (pH 8.5, 0.5°C)91


O 3 Spft Ratio (%/%)-*-100% in Coll-o-75/25 Col 1&2-•- 50/50 Col 1 & 2-^-25/75 Col 1&20 5 10 15CT (mg-min/L)20 25Figure 6.5: <strong>Bromate</strong> <strong>Formation</strong> as a Function <strong>of</strong> CT (pH 8.5,0.5°C)6.3 CO-CURRENT VERSUS COUNTER-CURRENT OZONATIONAdditional pilot-scale studies were conducted at the Britannia Water Purification Plant toexamine whether co-current or counter-current ozonation would yield lower bromateconcentrations for similar CT values. For these experiments ozone was only applied to thesecond column <strong>of</strong> the three-column pilot-scale ozone contactor.The same dose <strong>of</strong> ozone was applied for two runs, with the contactor operating in co-currentmode in one run, <strong>and</strong> counter-current in the other. Figure 6.6 shows the ozone concentrationpr<strong>of</strong>ile under each <strong>of</strong> these conditions. As would be expected, the ozone concentration was moreuniform in co-current application as compared to counter-current (Hull, 1995). It is important tonote that the ozone concentration leaving the column was fairly similar in both cases: in countercurrentmode, the ozonated water left the contactor at the bottom (height = 0 m) with an ozoneconcentration <strong>of</strong> 1.6 mg/L, while in co-current operation, the water left at the top <strong>of</strong> the contactorwith an ozone concentration <strong>of</strong> 1.9 mg/L.As shown in Figure 6.7, the formation <strong>of</strong> bromate along the depth <strong>of</strong> the contactor was also verysimilar for both modes <strong>of</strong> operation. At the entrance <strong>of</strong> the contactor (at the bottom for cocurrentmode, <strong>and</strong> the top for counter-current), the bromate concentration was about 4 to 6 ug/L.However, at the effluent <strong>of</strong> the column the bromate was approximately 14 to 15 ug/L forcounter-current <strong>and</strong> co-current modes, respectively. Since both the bromate <strong>and</strong> residual ozoneconcentrations entering the following column were similar for both modes <strong>of</strong> operation, it wasnot surprising that the final bromate concentrations at the effluent <strong>of</strong> the third column wereapproximately the same (22 <strong>and</strong> 24 ug/L for co-current <strong>and</strong> counter-current ozonation,respectively). Overall, there was no evidence that operating the contactor in a co-current orcounter-current mode would have any significant impact on bromate formation for the sameapplied ozone dose.92


Counter-CurrentCo-Current0123Ozone (mg/L)Figure 6.6: Ozone Pr<strong>of</strong>ile Versus Contactor Depth for Counter- <strong>and</strong> Co-CurrentOperations ,s 3oii2. 2• Counter-Current• Co-Current00 4 8 12<strong>Bromate</strong> <strong>Formation</strong> (ug/L)16Figure 6.7: <strong>Bromate</strong> Pr<strong>of</strong>ile Across Contactor Depth For Counter- <strong>and</strong> Co-CurrentOperationFigure 6.8 shows that the mode <strong>of</strong> ozonation did, however, affect the CT achieved within thecontactor. When operating in co-current mode, more than twice the CT was obtained than whenoperating in a counter-current fashion. This was a result <strong>of</strong> the different ozone concentrationpr<strong>of</strong>iles for the two operational configurations, as well as the shorter contact time for counter-93


current flow. Based on tracer tests a tio <strong>of</strong> approximately 3 minutes was calculated for countercurrentflow, <strong>and</strong> 4.5 minutes was calculated for co-current flow. In this particular contactor, itwould therefore be expected that a greater level <strong>of</strong> disinfection would be obtained during cocurrentoperation for the same applied ozone dose. Similar results have been reported in otherstudies (Hull, 1995).10-^averageCeffluentis average fromfour depths in contactoraEtwoHU642-•average0Co-CurrentCounter-CurrentFigure 6.8: Calculated CTfor Co- <strong>and</strong> Counter-Current <strong>Ozonation</strong>The CT values for co-current <strong>and</strong> counter-current ozonation were shown in Figure 6.8 usingdifferent methods <strong>of</strong> defining the ozone concentration. CT was always calculated based on thetio <strong>of</strong> the reactor. The Surface Water Treatment Rule (SWTR) allows C to be considered as (i)the average ozone concentration from multiple readings taken within the contact chamber(Caverage), (ii) the effluent ozone concentration for the chamber (Ceffl Uent) for co-current flow only,or (iii) the effluent concentration divided by 2 (CefnUent/2) for a counter-current chamber. Asshown in Figure 6.8, a difference was observed when comparing the calculated CT valuesdepending on the calculation method used, especially for the counter-current mode, hi countercurrentmode, use <strong>of</strong> Caverage led to a calculated CT that was 65 percent higher than when usingthe more easily measured Cout/2- While such differences would be site-specific, this result servesto illustrate that consideration should be given to such details when evaluating the performance<strong>of</strong> an ozone contactor, since the calculated CT may influence ozone dose requirements. Also,when designing a contactor it may be useful <strong>and</strong> cost-effective to include internal samplinglocations to permit the determination <strong>of</strong> Caverage, rather than relying exclusively on Ceffluent orCeffluent/2-While co-current operation provided a higher CT compared to counter-current operation, whenthe additional CT from the third column was added, the total CT for the system was essentiallythe same for both modes <strong>of</strong> operation (22 vs. 23 mg-min/L), as shown in Figure 6.9. This was aresult <strong>of</strong> the total CT value being largely controlled by the CTin the third column. Since the CT94


for the third column was about the same for co-current <strong>and</strong> counter-current flows, this tended tomask the differences in CT from the second column.5040Counter-CurrentCo-Current•2 30Oto0)*^OSOPQ2010Contactor ColumnEffluentsReaction ColumnEffluents010 1520 25CT (mg-min/L)Figure 6.9: <strong>Bromate</strong> Concentration as a Function <strong>of</strong> CT for Co- <strong>and</strong> Counter-Current<strong>Ozonation</strong>hi summary, for the overall pilot-scale ozone contactor there was no evidence that operating ineither co-current or counter-current mode was preferable in terms <strong>of</strong> reducing bromate formationwithout sacrificing CT. In this study, approximately the same bromate concentration wasobserved (22 ug/L vs. 24 ug/L) for the same total CT value (22 mg-min/L vs. 23 mg-min/L),when operating in co-current or counter-current mode, respectively.6.4 SUMMARYBased on these bench-scale <strong>and</strong> pilot-scale tests, staged (tapered) ozonation does not appear toreduce the levels <strong>of</strong> bromate formed compare to single stage ozonation, for a fixed level <strong>of</strong>disinfection. It was initially hypothesized that pilot-scale experiments might provide evidencethat contactor configuration, either staged ozonation or co-/counter-current application, wouldsuggest a means to minimize bromate formation while maintaining the same CT. However, therewas no evidence to support this hypothesis. Throughout these experiments, a fixed relationshipbetween bromate formation <strong>and</strong> ozone contact (CT or OE) existed, regardless <strong>of</strong> whether theozone was applied in one or two stages, or in co- or counter-current flow. The consistentrelationship between bromate formation <strong>and</strong> ozone contact indicates that the chemical conditionsare more important to bromate formation than hydrodynamic conditions.95


CHAPTER 7. A COMPARISON OF BROMATE FORMATION WITHIN DIFFERENTSCALES OF OZONE CONTACTORS<strong>Bromate</strong> formation was evaluated with several different scales <strong>of</strong> ozone contactors, to determineif the bench-scale reactors could be used to assess trade<strong>of</strong>fs between bromate formation <strong>and</strong>disinfection for the larger pilot-scale <strong>and</strong> full-scale systems. These tests were also used todetermine if the precise concentrations <strong>of</strong> bromate formed in the pilot-scale <strong>and</strong> full-scalesystems could be predicted at bench-scale. These tests utilized several different types <strong>of</strong> ozonereactors, including: completely-mixed true-batch bench-scale ozone reactor, semi-batch benchscaleozone reactor, bench-scale continuous-flow ozone contactor, multi-column counter-currentpilot-scale ozone contactors, <strong>and</strong> under-over counter-current multi-chamber full-scale ozonecontactors. The full-scale <strong>and</strong> pilot-scale experiments took place at the Los Angeles AqueductFiltration Plant (Sylmar, California), the Neuilly sur Mame Water Treatment Plant (Paris,France) <strong>and</strong> the Britannia Water Purification Plant (Ottawa, Canada).7.1 LOS ANGELES AQUEDUCT FILTRATION PLANTA multiple-scale comparison study <strong>of</strong> ozone contactors was conducted in conjunction with theLos Angeles Department <strong>of</strong> Water <strong>and</strong> Power (Los Angeles Aqueduct Filtration Plant) in earlyFebruary <strong>of</strong> 1999. The purpose <strong>of</strong> the study was to probe bromate formation at different levels<strong>of</strong> ozonation at each <strong>of</strong> three contactor scales, <strong>and</strong> to determine the comparability. Two differentbench-scale contactors were compared with pilot-scale <strong>and</strong> full-scale ozone contactorperformance. The bromate formation relationships developed at full-scale <strong>and</strong> pilot-scale werebased on terms <strong>of</strong> disinfection credit, or CT. For the bench-scale testing, the level <strong>of</strong> ozonationwas reported in both terms <strong>of</strong> CT <strong>and</strong> ozone exposure. The different contactors were thencompared based on the amount <strong>of</strong> bromate formed for a given amount <strong>of</strong> ozone contact (CT orOE).The full-scale <strong>and</strong> pilot-scale tests were performed on the same day using water from the LosAngeles Aqueduct. <strong>During</strong> the testing period, a large batch <strong>of</strong> the untreated water was collectedfor use in all <strong>of</strong> the laboratory-scale tests. This approach was used to minimize water qualitydifferences between the ozonation tests performed at each scale. Additional testing on this batch<strong>of</strong> water was presented throughout this report as data collected on LAW water. A summary <strong>of</strong>the water quality is presented in Chapter 2.7.1.1 Testing PlanOn-site ozone testing <strong>and</strong> sampling was performed at LAAFP by plant staff. Four experimentswere run each at full-scale <strong>and</strong> pilot-scale. At each scale, the experimental testing conditionswere varied by adjusting the gaseous ozone flow to either 100, 75, 50, or 25 percent <strong>of</strong> the totalozone output. The inlet <strong>and</strong> outlet <strong>of</strong> each cell was monitored for dissolved ozone residual usingHACK AccuVac vials to instantaneously quench dissolved ozone <strong>and</strong> a laboratoryspectrophotometer to measure ozone residual. Contact times were estimated from tracer studyinformation coupled with the liquid flow rates <strong>and</strong> the volumes <strong>of</strong> the contactor cells. Tracer97


studies were used to determine tio (the time for 10 percent <strong>of</strong> a tracer mass to exit the contactor)<strong>and</strong> tso (assumed equal to HRT). Utilizing the ozone concentration, HRT <strong>and</strong> tio contact time,CT values were calculated for the overall contactors <strong>and</strong> for each cell.Samples for residual ozone concentration <strong>and</strong> bromate concentration were taken at six locationsin the full-scale contactor <strong>and</strong> eleven locations in the pilot-scale ozone contactor. The sample taplocations are shown on the full-scale <strong>and</strong> pilot-scale schematics, presented in Chapter 2.As a residual ozone concentration existed at many <strong>of</strong> the sampling points, all <strong>of</strong> the bromatesamples were collected in 40-mL vials containinglOO uL <strong>of</strong> a 5 percent (by volume) solution <strong>of</strong>diethylamine (C^nN - an "organic ammonia") to immediately quench any ozone residual. Byquenching the ozone residual, bromate formation through the ozone contactors could beevaluated. Without quenching the residual ozone concentration, bromate formation would beexpected to continue, resulting in bromate measurements not representative <strong>of</strong> the bromatepr<strong>of</strong>ile through the ozone contactor at the time <strong>of</strong> sampling. These vials were then shipped to theauthors for analysis.The batch <strong>of</strong> water collected for the laboratory-scale experiments was tested using thecompletely-mixed true-batch ozone reactor <strong>and</strong> the laboratory-scale continuous-flow ozonecontactor. The details <strong>of</strong> the experimental methodologies are presented in Chapter 2. In order tocover a range <strong>of</strong> ozone exposures, three experiments were performed with the true-batch reactor<strong>and</strong> five experiments were performed with the continuous-flow contactor. <strong>Bromate</strong>concentrations were measured following each separate experiment. The CT <strong>and</strong> OE versusbromate relationships generated for each <strong>of</strong> these reactors were compared with the bromateformation relationships developed from the full-scale <strong>and</strong> pilot-scale testing. It should also benoted that the true-batch experiments were performed at a st<strong>and</strong>ard 20°C while the continuousflowexperimentation for this scale comparison was conducted at 9°C, the water temperature <strong>of</strong>both the full-scale <strong>and</strong> pilot-scale tests.7.1.2 True-Batch Ozone ReactorIndividual relationships for the four contactors/reactors were developed. These individualrelationships were initially explored, then overall conclusions were drawn based on thesimultaneous comparison <strong>of</strong> the three scales.A series <strong>of</strong> true-batch ozonation experiments were performed to determine bromate formation forselected ozone doses: 2:1, 1.5:1 <strong>and</strong> 1:1 mg ozone to mg DOC. For each experiment, the ozoneexposure was calculated to quantify the level <strong>of</strong> ozonation <strong>and</strong> the corresponding concentration<strong>of</strong> bromate formed was measured. The relationship between bromate formation <strong>and</strong> ozoneexposure is shown in Figure 7.1. The 1:1, 1.5:1 <strong>and</strong> 2:1 mg ozone to mg DOC doses produced7.5, 16 <strong>and</strong> 19 ug/L bromate, respectively.For all <strong>of</strong> the ozone doses evaluated, unreacted bromide was present after ozonation. At the 2:1mg ozone to mg DOC dose, 5.6 ng/L bromide remained after the ozone residual had decayed tobelow detection. The presence <strong>of</strong> unreacted bromide indicates that additional bromate couldhave been formed if additional ozone contact was provided. However, Figure 7.2 suggests that98


there might be the beginning <strong>of</strong> a slowing or plateau in bromate formation as the ozone dose wasincreased to the 1.5:1 mg ozone to mg DOC dose.25| 20y=0.63xR2 = 0.90a•2 15«£ 102 50900 5 10 15 20 25 30 35 40Ozone Exposure (mg-min/L)Figure 7.1: <strong>Bromate</strong> <strong>Formation</strong> in a True-Batch Reactor (LAW, pH 7,20°C)A <strong>Bromate</strong>A <strong>Bromide</strong>0 10 20 3040Ozone Exposure (mg-min/L)Figure 7.2: True-Batch Reactor <strong>Bromide</strong> Reaction <strong>and</strong> <strong>Bromate</strong> <strong>Formation</strong> (LAW, pH 7,20°C)99


7.1.3 Laboratory-Scale Continuous-Flow Ozone ContactorFive experiments were performed with the laboratory-scale continuous-flow ozone contactor.Ozone exposure was calculated as the product <strong>of</strong> the hydraulic residence time (HRT or tso) <strong>and</strong>the steady-state effluent ozone residual. A summary <strong>of</strong> the experimental details are presented inTable 7.1. It can be seen that a lower sample flow rate produced a higher ozone exposure. Thiswas attributed to the corresponding increase in HRT due to the increased contact times within thereactor. The resulting bromate concentrations for each <strong>of</strong> the experiments are graphicallypresented in Figure 7.3.Table 7.1: Bench-Scale Flow-Through Reactor Scaling ExperimentsExp. Qwater Qrecirc Oa Resid HRT OE <strong>Bromide</strong> <strong>Bromate</strong> Accounted Br"No. (mL/min) (mL/min) (mg/L) (min) (mg-min/L) (ug/L)-----033 01001 72 300 1.97 6.3 12197.4722 3 4 5 6 71 58 48 32 18 300 300 300 200 200 2.87 2.78 3.31 3.43 3.96 6.4 7.8 9.4 14 25 182231489916 15151611168.6161816796276836420a•B<strong>of</strong>a1I1510y=0.56xPossible <strong>Bromide</strong> Limiting Conditions0020 40 60Ozone Exposure (mg-min/L)80 100Figure 7.3: <strong>Bromate</strong> <strong>Formation</strong> in Laboratory-Scale Continuous-Flow Reactor100


The bromate concentration measured for the experiment with an ozone exposure <strong>of</strong> 99 mg-min/Lwas no greater than the bromate levels formed at an OE <strong>of</strong> 18 mg-min/L. This result indicated aplateau at about 16 ug/L bromate, <strong>and</strong> the subsequent exclusion <strong>of</strong> the higher OE data pointsfrom the OE versus bromate relationship. It is interesting to note that this plateau was observedin approximately the same range as for the true-batch ozonation experiments. As with the truebatchreactor experiments, it is unclear why a plateau was observed with a significant amount <strong>of</strong>unreacted bromide remaining in the continuous-flow system, as shown in Figure 7.4. Up to thisplateau, however, the bromide concentration decreased with increasing levels <strong>of</strong> ozonation. Thebromide <strong>and</strong> bromate data, indicated 70 to 80 percent <strong>of</strong> the bromide was accounted for betweenthe residual bromide concentration <strong>and</strong> the bromide incorporated into bromate. It is expectedthat the portion <strong>of</strong> bromide that was not accounted for in either the unreacted bromide or bromateformed, had been incorporated into brominated organic compounds or existed as free bromine(HOBr/OBr).A <strong>Bromide</strong>A <strong>Bromate</strong>0 *0 20 40 60Ozone Exposure (mg-min/L)80 100Figure 7.4: <strong>Bromide</strong> Reaction <strong>and</strong> <strong>Bromate</strong> <strong>Formation</strong> in a Laboratory-Scale Continuous-Flow Contactor7.1.4 Full-Scale <strong>and</strong> Pilot-Scale Ozone ContactorsA summary <strong>of</strong> the results from each <strong>of</strong> the full-scale <strong>and</strong> pilot-scale test is presented in Table 7.2.As before, larger applied ozone doses ultimately resulted in greater CT values <strong>and</strong> higherbromate concentrations. Based on these data, the amount <strong>of</strong> ozone contact was calculated fourdifferent ways to compare bromate formation for the full-scale <strong>and</strong> pilot-scale contactors. TheEPA accredited CT (CTswiR-tio) was calculated as the product <strong>of</strong> the ti 0 contact time <strong>and</strong> the101


ozone residual for each cell <strong>of</strong> the ozone contactor, excluding the first cell. Surface WaterTreatment Rule (SWTR) guidelines do not allow for CT credit from the first cell (USEPA, 1989;USEPA, 1990). However, this is likely a conservative indicator for ozonation because it isbelieved that some disinfection occurs within the first cell <strong>of</strong> the ozone contactors. Therefore,CT was also calculated with an "all cells" approach in which credit from the first cell was addedto CTswjR-tio- This parameter was labeled CTALL-UO- Additional CT representations, CTswrR-tso<strong>and</strong> (^TALL-ISO, were calculated using tso (HRT) contact times instead <strong>of</strong> tio.Table 7.2: Los Angeles Aqueduct Filtration Plant, Full-Scale <strong>and</strong> Pilot-Scale CT (9°C)___________Ozone__________ HRT tio <strong>Bromate</strong> CTSwTR-tio CTALL-tsoOutput Applied Transferred Residual(mg/L) (mg/L) (mg/L) (min) (min) (ug/L) (mg-min/L) (mg-min/L)Full-Scale Plant Summary100%75%50%25%100%75%50%25%2.952.231.550.732.552.111.360.652.722.051.400.622.552.111.360.651.161.030.700.268.88 4.378.71 4.348.92 4.428.75 4.35Pilot-Scale Summary1.15 8.38 7.270.80 8.38 7.270.53 8.38 7.270.18 8.38 7.277.03.71.20.53.51.90.80.32.552.211.210.403.703.441.340.418.377.354.291.636.785.752.870.92For the ozone conditions evaluated at full-scale, the 100, 75, 50, <strong>and</strong> 25 percent experimentsproduced bromate levels <strong>of</strong> 7.0, 3.7, 1.2, <strong>and</strong> 0.5 p.g/L, respectively. The four experimentsconducted on the pilot-scale contactor resulted in final bromate concentrations <strong>of</strong> 3.5, 1.9, 0.8,<strong>and</strong> 0.3 ug/L for the 100, 75, 50, <strong>and</strong> 25 percent tests, respectively. The CT versus bromaterelationship for each method <strong>of</strong> CT calculation has been presented in Figure 7.5. As expected,the correlation between CT <strong>and</strong> bromate formation was lower for each CTALL than for thecorresponding CTswTR due the additional CT "credit" obtained in the first cell for the samebromate concentration.Along with the overall ozone contact versus bromate trends, the data from the Los AngelesAqueduct Filtration Plant full-scale <strong>and</strong> pilot-scale components <strong>of</strong> the study yielded insight intothe bromate formation trends within the contactors. <strong>Bromate</strong> concentrations were plottedspatially along the contactor (expressed in terms <strong>of</strong> cumulative HRT through the contactor) toprovide a pattern <strong>of</strong> bromate formation while the water was ozonated <strong>and</strong> traveled from cell tocell. The full-scale <strong>and</strong> pilot-scale bromate formation trends are presented in Figure 7.6 <strong>and</strong>Figure 7.7, respectively. A large portion <strong>of</strong> the bromate appeared to form within the later stages<strong>of</strong> the contactors for the 75 <strong>and</strong> 100 percent tests, for both scales.The corresponding trends in CTALL-t50 development were also generated for both the full-scale<strong>and</strong> pilot-scale tests, as seen in Figure 7.8 <strong>and</strong> Figure 7.9, respectively. It is interesting to notethat an almost linear relationship was detected between HRT <strong>and</strong> CT ALL-ISO, revealing the slowrate <strong>of</strong> ozone decay.102


y=1.29x y=1.03x•'--£ xy=0.76x y = 0.66x y = 0.49x y = 0.42xR =0.88 R =0.88 R =0.89 R =0.89• FullSWTRtlOAFullALLtlO* Full SWTR150X Full ALL t50n Pilot SWTRtlOo Pilot SWTRtSOA Pilot ALL tlOX Pilot ALL t5002468CT (mg-min/L)10Figure 7.5: Effect <strong>of</strong> CT Calculation <strong>of</strong> <strong>Bromate</strong> <strong>Formation</strong> Trends3a•-§esoto"«o•D0 4O D D0 2 4 6 8 10Relative Contactor Location as Cumulative HRT (min)Figure 7.6: Los Angeles Aqueduct Filtration Plant <strong>Bromate</strong> Production103


4c1| oto2CO0 OD-0-CD-O-A-DO-2468Relative Contactor Location as Cumulative HRT (min)10Figure 7.7: Los Angeles Aqueduct Filtration Pilot Plant <strong>Bromate</strong> Production0 2 4 6 8Relative Contactor Location as Cumulative HRT (min)10Figure 7.8: CTALL-tso Development Along Full-Scale Contactor104


.3BJD4 -U002468Relative Contactor Location as Cumulative HRT (min)10Figure 7.9: CTALL-tso Development Along Pilot-Scale Contactor7.1.5 Comparison <strong>of</strong> Full-, Pilot-, <strong>and</strong> Bench-ScalesFigure 7.10 shows the combined results <strong>of</strong> the multiple-scale comparison study. Sixrelationships are shown: CTswiR-no <strong>and</strong> CTALL-tso representations for both full- <strong>and</strong> pilot-scaleozone contactors, <strong>and</strong> the two bench-scale reactor relationships. CTswra-tio <strong>and</strong> CTALL-t50 werechosen because these calculated parameters represent the extremes in CT calculations. CTswrRtio,as per the SWTR guidelines, is a conservative estimate <strong>of</strong> the ozone contact so that factor <strong>of</strong>safety is incorporated to insure that public health is protected against microbial contamination.CTALL-tso might be considered more representative <strong>of</strong> the chemical reactions, as the dissolvedozone was in contact with the water constituents in all <strong>of</strong> the cells <strong>and</strong> for longer than the tiocontact time. Conceptually, CTALL-tso approximates the ozone exposure calculation used for thelaboratory reactors.As expected the full-scale system CTswxR-tio generated the highest relationship between bromateformation <strong>and</strong> calculated ozone contact. By the same calculation method, the pilot-scale systemperformed better. It was believed that this difference could be attributed to the differenthydraulics <strong>of</strong> the two systems. The full-scale contactor performed similar to a completely mixedsystem with a characteristic tio/tso <strong>of</strong> 0.50, determined by tracer testing. The hydraulics <strong>of</strong> thepilot-scale ozone contactor were closer to that <strong>of</strong> a plug-flow system, with a tio/tso <strong>of</strong> 0.87. Byutilizing the best estimate <strong>of</strong> ozone contact (CTALL-tso), the bromate formation relationshipbetween the pilot-scale <strong>and</strong> full-scale contactors converged. This convergence between105


hydraulically disparate contactors indicates that a chemically controlled relationship betweenbromate formation <strong>and</strong> ozone contact exists.2016a£ ceE<strong>of</strong>eesoCQ12840l-Log Cryptosporidium Inactivation(2.5 -10 ipg-min/L @ 9°C)""•••' Yd [ig/LBfomate M'CL" • FullSWTRtlO0 10 15Ozone Exposure or CT (mg-miii/L)APilotSWTRtlOo Full ALL t50A Pilot ALL t50X Bench-Scale (TB)+ Bench-Scale(CF)20Figure 7.10: Results <strong>of</strong> Los Angeles Aqueduct Filtration Plant Multiple-Scale ComparisonStudyThese relationships were also compared to the laboratory reactor relationships. <strong>Bromate</strong>formation versus ozone exposure in the true-batch ozone reactor <strong>and</strong> laboratory-scalecontinuous-flow ozone contactor, respectively, were essentially the same. This is believed tohave been a coincidental occurrence resulting from the scatter in the data (see Figure 7.2 <strong>and</strong>Figure 7.3), as the true-batch <strong>and</strong> continuous-flow experiments were performed at differenttemperatures, 20°C <strong>and</strong> 9°C, respectively.Therefore, these data indicate that the bench-scale systems could be used to approximate theresults <strong>of</strong> the larger pilot-scale <strong>and</strong> full-scale systems, provided the methods used to calculateozone contact approximate the total exposure <strong>of</strong> ozone to the water (i.e. OE <strong>and</strong> CTALL-ISO)- Thedata also illustrate that a pilot-scale system could be used to accurately predict bromateformation in a hydraulically disparate full-scale ozone contactor.7.2 NEUILLY SUR MARNE WATER TREATMENT PLANTThe same true-batch ozone reactor <strong>and</strong> laboratory-scale continuous-flow ozone contactor wereused to generate small-scale results for a comparison to full-scale <strong>and</strong> pilot-scale ozonecontactors at Compagnie Generale des Eaux's (CGE) Neuilly sur Mame Water Treatment Plant(WTP). While the testing at the Los Angeles Aqueduct Filtration Plant was carried out byvarying the ozonation conditions <strong>of</strong> one contactor, the full-scale tests performed at the Neuilly106


sur Marne WTP involved parallel contactors simultaneously operated at different experimentalconditions. The full-scale tests involved monitoring <strong>of</strong> 3 or 4, <strong>of</strong> 5 parallel counter-currentcontactors. Ozone experiments were conducted at each scale using Neuilly sur Mame WTPclarified <strong>and</strong> s<strong>and</strong>-filtered Mame River water, in May <strong>and</strong> October <strong>of</strong> 1998. The water qualityconditions measured during the two sets <strong>of</strong> ozone experiments are presented in Table 7.3.Table 7.3: Water Quality <strong>of</strong> Clarified <strong>and</strong> S<strong>and</strong>-Filtered Marne River WaterParameter _____Testing PeriodTemperature (°C)<strong>Bromide</strong> (ng/L)pHAlkalinity (mg/L as CaCO3) DOC (mg/L)SUVA (L/mg-m)May 1998 19427.62301.12.3October 199815307.82201.52.0A linear relationship was observed between bromate formation <strong>and</strong> ozone exposure for the truebatchozone reactor <strong>and</strong> the laboratory-scale continuous-flow ozone contactor in the clarified <strong>and</strong>s<strong>and</strong>-filtered Marne River water. Figure 7.11 <strong>and</strong> Figure 7.12 also show that this linear trendalso appeared to extend to the pilot-scale ozone contactor at the Neuilly sur Marne WTP testingas well. The limited amount <strong>of</strong> full-scale data was not sufficient to determine if the linearrelationship would be observed at this scale as well. The full-scale results can be viewed asreplicate experiments between ozone contactors operated under similar hydraulic <strong>and</strong> waterquality conditions. The ozone contact for the true-batch reactor <strong>and</strong> laboratory-scale continuousflowozone contactor was quantified by calculating the ozone exposure. For the pilot-scale <strong>and</strong>full-scale results, however, the quantity <strong>of</strong> ozone contact was expressed asFor the same ozone exposure, bromate formation was consistently greater for the laboratoryscalecontinuous-flow ozone contactor than for true-batch ozone reactor. <strong>During</strong> both testingperiods, the full-scale results were generally bounded by the results obtained from the twolaboratory reactors. The same bromate formation relationship was calculated from the pilotscale<strong>and</strong> continuous-flow ozone contactors during the May 1998 testing period (1.2 ug/mgmin).However, a slightly lower production <strong>of</strong> bromate per unit <strong>of</strong> ozone contact was observedin the continuous-flow ozone contactor as compared to the pilot contactor (0.93 versus 1.1ug/mg-min) during the October 1998 testing period. Overall, the slightly higher temperature (19as compared 15°C) <strong>and</strong> bromide concentration (42 as compared to 30 ug/L) in the May testingperiod resulted in slightly higher bromate formation relationships, if any difference wasobserved. As a consistent relationship between the different ozone contactors utilizing Neuillysur Marne Water Treatment Plant clarified <strong>and</strong> s<strong>and</strong>-filtered Mame River water, it was notpossible to determine why the differences existed. However, either the pilot or continuous-flowozone contactors could be used to provide a conservative (high) estimate <strong>of</strong> bromate formation atfull-scale.107


Full-Scab• Pibt-ScaleA Continuous-Flow* True-Batch0 10 20 30 40or Ozone Exposure (mg-min/L)50Figure 7.11: Neuilly sur Marne Comparison <strong>of</strong> Ozone Contactors (May 1998)FuD-Scale• Pilot-ScaleA Continuous-Flow* True-Batch0 10 20 30 40or Ozone Exposure (mg-min/L)50Figure 7.12: Neuilly sur Marne Comparison <strong>of</strong> Ozone Contactors (October 1998)108


7.3 BRITANNIA WATER PURIFICATION FACILITYAdditional tests were performed during two testing periods using clarified Ottawa River water tocompare the bromate formation in different scales <strong>of</strong> ozone contactors. For these tests, resultsobtained from the pilot-scale ozone contactor at the Britannia Water Purification Facility werecompared to the data collected in a semi-batch ozone reactor <strong>and</strong> the laboratory-scalecontinuous-flow ozone contactor. The characteristics <strong>of</strong> the water quality during the two testingperiods are shown in Table 7.4.Table 7.4: Water Quality <strong>of</strong> Clarified Ottawa River WaterParameter ______________ August 1999 _______ October 1999Ambient Temperature (°C) 22 11<strong>Bromide</strong> (ug/L) 1 5 (spiked to 3 5) 15 (spiked to 3 5)UV254(cm' 1 ) 0.109 0.089TOC(mg/l) 3.7 2.8pH 6.18 (adjusted to 8) 6.15 (adjusted to 8)Alkalinity (mg/L as CaCO3) 9.2 (adjusted to 55) 10 (adjusted to 60)Turbidity (NTU)_____________ 0.08 ____________ 0.07 ______The experiments to directly compare bench-scale <strong>and</strong> pilot-scale ozonation results wereconducted in October 1999, when the Ottawa River temperature was 11°C. Additional benchscaletests were conducted in both cold water (7°C) <strong>and</strong> warm water (22°C) to examinetemperature effects.Preliminary experiments at 1 1°C showed no bromate formation at the ambient bromide level (15ug/L). Consequently, an additional 20 ug/L <strong>of</strong> bromide was added to bring the concentration upto 35 ug/L. At ambient pH (6.2), however, no detectable amount <strong>of</strong> bromate was formed, evenat a CTALL-tso <strong>of</strong> 50 mg-min/L. As a result, the pH was raised to 8 by adding NaOH <strong>and</strong> thealkalinity was adjusted with 55 mg/L <strong>of</strong> NaHCOs. The increase in alkalinity was sufficient tomaintain an approximate equilibrium with atmosphericThe bromate formation results for these adjusted water quality conditions are shown in Figure7.13. In each case, a simple linear correlation between bromate formation <strong>and</strong> ozone contact wasobserved. At the 95 percent confidence interval, no statistical difference existed between thepilot-scale <strong>and</strong> continuous-flow ozone contactors. While similar increases in bromateconcentration for a given unit change in ozone contact was observed in the semi-batch results,the absolute bromate concentrations were approximately 3 ug/L higher than the measurementsobtained with the other contactors.An additional series <strong>of</strong> bench-scale experiments were conducted in August 1999 to determinehow the performance <strong>of</strong> the reactors would compare at temperatures above <strong>and</strong> below 1 1°C. Forthese tests, the pH, alkalinity <strong>and</strong> bromide adjusted water was tested at ambient temperature(22°C) in the continuous-flow <strong>and</strong> semi-batch reactors. The water was then cooled to 7°C <strong>and</strong>109


the tests were repeated. In Figure 7.14, the linear relationship between bromate formation <strong>and</strong>ozone contact was maintained for both reactors at both temperatures. Approximately half asmuch bromate was formed at 7°C when compared to 22°C for the same CTALL-ISO- It isinteresting that the colder water resulted in only about a 50 percent reduction in bromateformation for a given CTALL-t50 value.2015aE 10<strong>of</strong>a£pao-•-Semi-Batch-A- Continuous-Flow-•-Pifot-Scale0 10 20 30 40CTALL-tso or Ozone Exposure (mg-min/L)50Figure 7.13: Britannia Water Purification Facility Ozone Contactor Comparison (October1999,pH8)Proposed CT requirements for a target level <strong>of</strong> Cryptosporidium inactivation are approximately 6times higher at 7°C than at 22°C (Rennecker et al., 1999; Oppenheimer et al, 2000). Thisimplies that a plant may have to increase the amount <strong>of</strong> ozone contact by 600 percent to maintainthe same level <strong>of</strong> Cryptosporidium control for these temperatures, while bromate formation mayonly be reduced by 50 percent for the same temperature difference. These particular resultssuggest the potential for greater bromate formation in colder water than in warm water existswhen ozonating to a specific level <strong>of</strong> inactivation. While these findings are consistent with thosedescribed in Section 5.1.2, they are contradictory to the batch tests described in Section 4.2. It isunclear why different results were observed for these different reactors.Excellent agreement between the results from the semi-batch reactor <strong>and</strong> the continuous-flowcontactor were observed in Figure 7.14, in contrast to the lack <strong>of</strong> agreement observed in Figure7.13. The experiments in Figure 7.13 <strong>and</strong> Figure 7.14 differed not only in temperature, but alsothe time <strong>of</strong> year that the water was collected (summer versus fall). As excellent agreement wasobtained at both 7°C <strong>and</strong> 22°C, the lack <strong>of</strong> agreement at 11°C was not expected to be attributableto temperature, but may depend on the specific water matrix. A reasonable correlation wasobserved between the laboratory-scale continuous-flow ozone contactor <strong>and</strong> the pilot-scale ozonecontactor.110


3020• Continuous-Flow (22°C)A Continuous-Flow (7°C)o Semi-Batch (22°C)A Semi-Batch (7°C)<strong>of</strong>e* 10esI0010 20 30Ozone Exposure (mg-min/L)40 50Figure 7.14: Temperature Effect on Laboratory-Scale Reactors (August 1999, pH 8)7.4 CONCLUSIONSIt was shown that bench-scale reactors could provide reasonable simulations <strong>of</strong> pilot-scale <strong>and</strong>full-scale bromate formation trends provided an accurate estimate <strong>of</strong> the amount <strong>of</strong> ozone contactcould be generated. Likewise, pilot-scale simulations can provide useful approximations <strong>of</strong> fullscaleresults. Such simulations would be useful in assessing trade<strong>of</strong>fs between bromateformation <strong>and</strong> disinfection (CT) under various scenarios. However, these results also illustratedthat while the simulations could be used for trending purposes, the specific bromateconcentrations formed for given amounts <strong>of</strong> ozone contact do not always match betweendifferent contactors. The results from these tests also imply that chemical conditions, as opposedto hydraulic conditions, control the formation <strong>of</strong> bromate.Comparisons between full-scale ozonation at the Los Angeles Aqueduct Filtration Plant, pilotscaletests performed at the same facility, <strong>and</strong> two types <strong>of</strong> laboratory reactors indicated that themethod <strong>of</strong> calculating the ozone contact is critical to developing accurate predictions betweendifferent contactors. If ozone contact is calculated as CTper the SWTR guidelines (CTswiR-tio),the total amount <strong>of</strong> ozone contact is underestimated <strong>and</strong> hydrodynamic differences can result inlarge discrepancies between the concentrations <strong>of</strong> bromate formed in different contactors. As themethod <strong>of</strong> calculating ozone contact is improved, the relationship between bromate formation<strong>and</strong> ozone contact in hydrodynamically varied contactors tends to converge. At the Los AngelesAqueduct Filtration Plant, identical results were obtained between full-scale <strong>and</strong> pilot-scale onceozone contact was calculated at CTALL-tio; where CT credit for the first cell is included <strong>and</strong> the111


hydraulic residence time (HRT or tso) is used to characterize the contact time instead <strong>of</strong> tio.Results from the laboratory-scale continuous-flow ozone contactor also provided a reasonableestimate <strong>of</strong> the bromate formation at both <strong>of</strong> the larger scales.Using CTALL-tio as an estimate <strong>of</strong> the ozone contact for full-scale <strong>and</strong> pilot-scale system at theNeuilly sur Mame Water Treatment Plant <strong>and</strong> the Britannia Water Purification Facility,reasonable predictions <strong>of</strong> bromate formation could be made using the laboratory-scalecontinuous-flow ozone contactor. For both <strong>of</strong> these water agencies, the use <strong>of</strong> either a true-batchor semi-batch reactor could not consistently simulate bromate formation in the other reactors.Additional experiments performed at 7°C <strong>and</strong> 22°C in clarified Ottawa River water, indicatedthat excellent correlations between reactors could be obtained at different temperatures. Thesedata also indicated that the potential for greater bromate formation in colder water than warmwater exists when ozonating to a specific level <strong>of</strong> inactivation.112


CHAPTER 8. OZONE CONTACTOR MODELING8.1 BATCH AND SEMI-BATCH REACTOR MODELINGThe initial step in modeling Cryptosporidiwn parvum oocyst inactivation for the various reactorsutilized in this study was to determine C. parvum oocyst inactivation kinetics using ozone.Results obtained for multiple C. parvum oocysts inactivation experiments with the semi-batchreactor are presented in Figure 8.1. The inactivation levels presented here were normalized toviability <strong>of</strong> the control sample to which no ozone was exposed. The results were found to beindependent <strong>of</strong> the solution pH within the pH range <strong>of</strong> 6.5-8.5, indicating that molecular ozonewas responsible for inactivation. The lag phase factor Ni/N0 <strong>and</strong> pseudo first-order inactivationrate constant k^, corresponding to the Delayed Chick-Watson model, were determined to be 3.6<strong>and</strong> 1.77 (L/mg-min) at 20°C, respectively. These parameters were used for the modelingexperimental results described throughout this chapter.-10OX)o-10 1 2 3Ozone Exposure (mg-min/L)Figure 8.1: Effect <strong>of</strong> pH on Experimental <strong>and</strong> Fitted Kinetics <strong>of</strong> the Inactivation <strong>of</strong>Cryptosporidium parvum Oocysts (20°C, Semi-Batch Reactor)Batch ozonation tests were performed with all the natural waters investigated in this study. Theozone decomposition kinetics in selected samples are shown in Figure 8.2. Also shown in thefigure are the curves obtained from fitting the data with the kinetic model corresponding to thereactions represented in Table 2.11. The empirical parameters that determine the characteristics<strong>of</strong> NOM in specific water, such as a, (3, y> <strong>and</strong> fa>» were calibrated so that the overalldecomposition kinetics matched the experimental results. These parameters were also included113


in Figure 8.2. The NOM fraction responsible for Reaction 100 in Table 2.11 was obtained fromozone decomposition curve <strong>and</strong> represented as a fraction <strong>of</strong> DOC (Fo). The correspondingpredicted bromate formation potential for each natural water tested was compared to theexperimental data <strong>and</strong> presented in Figure 8.3. While the model prediction <strong>of</strong> ozonedecomposition was accurate, <strong>and</strong> that <strong>of</strong> bromate formation was generally acceptable, there werelarge discrepancies for a few natural waters. This limitation in bromate formation predictioncould be the result <strong>of</strong> inaccuracies in bromate measurement, imperfections in the empiricalapproach to model NOM, or a combination <strong>of</strong> these two factors.coUa-10 20 30Time [min]40 5 10 15Time [min]20 40Time [min]Figure 8.2: Batch Ozone Decay Kinetics for Selected Natural Waters (pH 7,20°C)8.2 BENCH-SCALE FLOW-THROUGH OZONE CONTACTOR MODELINGTo model the laboratory-scale continuous-flow ozone contactor described in Section 2.2.3, it wasnecessary to underst<strong>and</strong> the hydrodynamic characteristics <strong>of</strong> this reactor. A tracer test wasperformed with Rhodamine WT as a tracer compound for this purpose. The concentration <strong>of</strong>Rhodamine WT was determined with an Aminco Bowman Series 2 Spectr<strong>of</strong>luorometer usingexcitation <strong>and</strong> emission wavelengths <strong>of</strong> 555nm <strong>and</strong> 580nm, respectively. The tests wereperformed on the bubble column <strong>and</strong> recirculation line separately. Tracer solution was injectedas a pulse input into the inlet <strong>of</strong> the bubble column or the recirculation line after reaching steadystate conditions. Two-milliliter effluent samples were collected at maximum time intervals <strong>of</strong> 5114


seconds for an overall period <strong>of</strong> at least five times the theoretical hydraulic residence time <strong>of</strong>each component.I2520aI"• Experimental!D Model FitI 50SPW* CRW ACD* ANN* CCD SAC LAW* WPB DAL HOU NJA CGE OTT* Ammonia level was adjusted to fit the bromate formation potentialFigure 8.3: Fitted Batch Reactor <strong>Bromate</strong> <strong>Formation</strong> PotentialTracer test results with normalized time <strong>and</strong> concentration scales are presented in Figure 8.4.Actual times were divided by the corresponding hydraulic residence time, <strong>and</strong> actualconcentrations were divided by the ratio <strong>of</strong> total tracer mass added to reactor component volume.Tracer curves were analyzed to obtain the mean residence time <strong>and</strong> st<strong>and</strong>ard deviation. Thesevalues were then used to determine axial dispersion numbers assuming closed vessel boundaryconditions. The model tracer curve was generated from an axial-dispersion reactor (ADR) modelprogram developed in C++. hi this program, a finite difference approach was used to integratethe second-order partial differential equations. Both tracer curves were accurately described bythe axial dispersion model by a dispersion number <strong>of</strong> 0.01, corresponding to Peclet number <strong>of</strong>100, as shown in Figure 8.4 The tracer curves could be also modeled as a CSTR cascade <strong>of</strong> 50reactors. This analysis revealed that each component <strong>of</strong> the bench-scale flow-through reactorbehaves similar to an ideal plug-flow reactor (PFR). Therefore, the bench-scale flow-throughreactor was modeled as a PFR-side PFR, in which PFR represents the main bubble column <strong>and</strong>side-PFR represents the recirculation line.The results obtained for experiments performed with the bench-scale flow-through reactor usingthree selected natural waters are presented in Figure 8.5 <strong>and</strong> Figure 8.6. The inactivation <strong>of</strong> C.parvum oocysts <strong>and</strong> formation <strong>of</strong> bromate was investigated simultaneously in these tests. Thebromide level was adjusted to 90 ^ig/L for LAW <strong>and</strong> OTT waters. The temperature was kept at20°C <strong>and</strong> pH was adjusted to 7.5 for all <strong>of</strong> the experiments. Figure 8.5 shows the inactivationlevel <strong>of</strong> C. parvum oocyst at different CT. The level <strong>of</strong> inactivation was independent <strong>of</strong> the type<strong>of</strong> natural water tested as long as the same CT was attained, once again confirming that C.parvum oocyst inactivation is only affected by exposure to dissolved molecular ozone. The115


integrated model prediction for each water showed reasonable agreement with experimental dataover the CT range investigated. These results validated the use <strong>of</strong> the PFR-side-PFR model torepresent the bench-scale flow-through reactor. The corresponding bromate formationpredictions are also shown in Figure 8.6. Contrary to C. parvum inactivation, bromate formationwas greatly affected by the nature <strong>of</strong> NOM, even though the initial bromide levels for LAW <strong>and</strong>OTT were the same. This large discrepancy in bromate formation was indicative <strong>of</strong> differencesin ozone consumption <strong>and</strong> hydroxyl radical inhibition by NOM. This trend was reflected in thebromate formation model predictions shown in Figure 8.6. The model showed good agreementwith the experimental data.1.0Contactor Column ExperimentADR Model (d=0.01)0.5satuaoU•ees0.01.0Recirculation ColumnExperiment-ADR Model (d=0.01)0.50.00Normalized TimeFigure 8.4: Tracer Test Results for the Bench-Scale Flow-Through Reactor <strong>and</strong>Corresponding ADR Modeling116


A SPWLAWOTT- - - - SPW Model PredictionLAW Model PredictionOTT Model Prediction10 20CT (mg-min/L)30Figure 8.5: Cryptosporidium parvum Inactivation Results <strong>and</strong> Corresponding ModelPredictions for the Laboratory-Scale Continuous-Flow Ozone Contactoraooto8521601208040A SPWa LAW» OTT— - - SPW Model Prediction——— LAW Model Prediction—------ OTT Model Prediction0 10 20CT (mg-min/L)30Figure 8.6: <strong>Bromate</strong> <strong>Formation</strong> <strong>and</strong> Prediction for the Bench-Scale Flow-Through OzoneContactor (LAW & OTT Spiked to 90 ug/L fir")The model-predicted CT requirements for inactivation <strong>of</strong> Iowa strain C. parvum oocysts atdifferent temperatures <strong>of</strong> the laboratory-scale continuous-flow ozone contactor are summarizedin Table 8.1. The ozone concentration was assumed to be constant throughout the reactor <strong>and</strong>117


hydraulic residence time was used to calculate CT. It is important to note that the CT required toobtain the same level <strong>of</strong> inactivation in the laboratory-scale continuous-flow ozone contactor isdramatically higher than that required in a batch reactor (Table 1.1). This results from the uniquehydrodynamic behavior <strong>of</strong> the laboratory-scale continuous-flow ozone contactor that deviatesfrom ideal plug-flow. The CT table for a CSTR, the other extreme ideal case, is presented inTable 8.2. The CT requirements for this case are much greater than that for a batch orlaboratory-scale continuous-flow ozone contactor.Table 8.1: Required CJfor C.parvum Oocysts Inactivation in Laboratory-ScaleContinuous-Flow Reactor (Iowa strain Oocysts with Inactivation Kinetics Shown in Table1.1, Recirculation Ratio <strong>of</strong> 10)Temperature(°C)-Log(N/N0)0.511.522.530.5 5 10 15 20 25 30Cr[mg-min/L]41.4 126 23.2 71.0 12.5 38.2 6.87 21.0 3.85 11.8 2.20 6.74 1.283.92246 388 538 690 138 218 302 387 74.5 117 162 208 40.964.5 89.3 114 . 23.0 36.2 50.1 64.3 13.1 20.7 28.7 36.7 7.6512.016.721.4Table 8.2: Required CJfor C.parvum Oocysts Inactivation in CSTR (Iowa strain Oocystswith Inactivation Kinetics Shown in Table 1.1)Temperature(°C)-Log(N/N0)0.51.01.52.02.53.00.5 5 10 15 20 25 30CT frtig-min/Ll56.1 200 656 2.09x1 03 31.5 112 368 1 .ISxlO3 17.0 60.5 198.2 633.6 9.32 33.2 108 348 5.23 18.7 61.1 195 2.99 10.7 34.9 112 1.746.2120.465.16.65xl032.llxlO43 .74x1 03 1 .ISxlO4 2.01x10 3 6.36x1 03 1 .llxlO3 3 .50x1 03 620 1 .96x1 03 355 1.12xl03 2076548.3 PILOT-SCALE OZONE CONTACTOR MODELINGIntegrated modeling <strong>of</strong> C. parvum inactivation <strong>and</strong> bromate formation was further tested on thepilot-scale ozone contactor at the Britannia Water Purification Facility. The experimental datawere obtained with the pilot-scale ozone contactor operated in both the counter-current <strong>and</strong> cocurrentmode. In each case ozone was only applied to the second column <strong>and</strong> allowed todissipate through the third column. The hydrodynamics <strong>of</strong> this reactor were first characterizedwith a tracer test using fluoride as a tracer compound. The experimental data for the case <strong>of</strong>counter-current operation mode are presented in Figure 8.7 along with the corresponding axialdispersionreactor model fit. The dispersion number within the second column during ozoneapplication was determined to be 0.6, while third column chamber had a much lower dispersion118


number <strong>of</strong> 0.04. The significantly higher dispersion number in the second column wasattributable to the induced mixing by the rising gas bubbles. <strong>During</strong> co-current operation <strong>and</strong> atthe same water <strong>and</strong> gas flow rates, the tracer test was used to calculate dispersion numbers <strong>of</strong> 0.3<strong>and</strong> 0.06 for the second <strong>and</strong> third column, respectively.• Bubble diffuserD Reacting chamberADR model0.5 1.5 2 2.5Normalized Time3.5Figure 8.7: Britannia Pilot Plant Counter-Current Tracer Test <strong>and</strong> Corresponding ADRModeling (Qwater = 5L/min, Qgas = 1.42 L/min, 0.5°C)The ADR model predictions for dissolved ozone concentration pr<strong>of</strong>ile, bromate formation <strong>and</strong> C.parvum inactivation in the bubble column <strong>and</strong> the reacting chamber were compared to theBritannia Pilot Plant experimental results, as shown in Figure 8.8. Although the model deviatedfrom the experimental data, shape <strong>of</strong> the predicted ozone pr<strong>of</strong>iles was similar to those <strong>of</strong> the data.It is believed that the primary major reason for the deviation was that the model was based onreaction rate constants only available at 20°C while the experiments were performed at 0.5°C.As the ozone residuals were underpredicted by the model at this low temperature, it was notsurprising that the corresponding bromate formation was underpredicted. It was believed that thepredicted level <strong>of</strong> concomitant disinfection was underpredicted as well.119


(a) Bubble Column Simulation1.00.90.80.7I)E 0.6•u


8.4 FULL-SCALE OZONE CONTACTOR MODELINGFull-scale ozone contactor performance was also simulated with the integrated bromateformation <strong>and</strong> C. parvum inactivation model. The tracer test data obtained from the Los AngelesAqueduct Filtration Plant was analyzed for the evaluation <strong>of</strong> hydrodynamic characteristics.Tracer tests were performed using Rhodamine WT introduced as a step input. The axialdispersionmodel was applied to estimate a dispersion number <strong>of</strong> 0.054 for the entire reactor, asshown in Figure 8.9. This single dispersion number was used for the modeling the differentsections <strong>of</strong> the contactor (ozone application cell, ozone reaction cells, rapid mixers) even thoughdifferent hydraulic conditions were expected, because separate tracer data was not available.1.0o oaoa>awa•aVeso0.5• Positive Step Inputo Negative Step InputADR Model0.00 1 2Normalized TimeFigure 8.9: Los Angeles Aqueduct Filtration Plant Tracer Tests <strong>and</strong> Corresponding ADRModeling (Qwater = 85 MGD, Q gas = 17.8 scfh)Figure 8.10 shows the dissolved ozone <strong>and</strong> bromate concentrations observed at various samplingtaps <strong>of</strong> the contactor. The graph is presented in terms <strong>of</strong> normalized cumulative volume <strong>of</strong> theozone contactor including rapid mixers. The model was initialized using the kinetic informationobtained from the batch ozonation test for LAW water. Simulations were performed for each <strong>of</strong>the experimental testing conditions: 100, 75, 50, <strong>and</strong> 25 percent <strong>of</strong> the ozone generation capacity.The model simulations had the same trends as those observed for experimental data. It isbelieved that the deviations from the data are much smaller than those observed with the pilotscaleozone contactor at the Britannia Water Purification Facility because the temperature in thiscase was 9°C, closer to the 20°C derived reaction rate constants. Based on C. parvum oocystinactivation modeling, this plant would be expected to achieve about 3-log inactivation for 75percent capacity operation at 20°C.121


IU°2.01.0 -(a)Liquid PhaseOzoneConcentrationPr<strong>of</strong>ile0.5 -(c)C. parvumInactivation00.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0Normalized Cumulative Volume• Experimental datafor 100% capacityO Experimental datafor 75% capacityT Experimental datafor 50% capacityV Experimental datafor 25% capacity——— Model simulationfor 100% capacity— • • • Model simulationfor 75% capacity— — Model simulationfor 50% capacity— Model simulationfor 25% capacityFigure 8.10: Los Angeles Aqueduct Filtration Plant Measured <strong>and</strong> Predicted OzonePr<strong>of</strong>ile, <strong>Bromate</strong> <strong>Formation</strong> <strong>and</strong> Cryptosporidium Inactivation (pH 8.2,33 ng/L Br", 1.9mg/L DOC, 9°C Qwater = 85 MGD, Qgas = 17.8 scfh)8.5 SUMMARYIntegrated models for ozone decomposition, bromate formation, <strong>and</strong> C. parvum inactivation weredeveloped for four different types <strong>of</strong> reactors <strong>and</strong> contactors: true-batch ozone reactor,laboratory-scale continuous-flow ozone contactor with external recirculation, pilot-scale ozonecontactor <strong>and</strong> full- scale ozone contactor. While employing the same reaction mechanisms,122


different programs were developed to solve to specific mathematical problems corresponding tothe different hydrodynamic schemes. Once the empirical kinetic evaluation <strong>of</strong> NOM wasperformed in batch tests, this information was utilized to simulate the ozone contactor at eachscale. The integrated modeling provided a good strategy to simulate the ozone contactorperformance during scale-up. This simulation provided not only the final effluent watercharacteristics, but also generated the chemical species concentration pr<strong>of</strong>iles <strong>and</strong> C. parvuminactivation levels along the ozone contactors. However, there were modeling limitationsidentified that should be addressed in future research.• Simulations for ozone decomposition <strong>and</strong> bromate formation were possible only at 20°C, thetemperature at which the rate constants for each reaction were evaluated.• Empirical approach to model the interactions <strong>of</strong> NOM with ozone <strong>and</strong> secondary oxidantswas not adequate for some natural waters.• Accurate measurement <strong>of</strong> primary species such as ozone, bromine, <strong>and</strong> bromate, as well assecondary species, such as hydroxyl radical, were needed for more robust modeling atmechanistic level.Despite these limitations, this study suggests that the model simulation is a powerful tool tounderst<strong>and</strong> the performance <strong>of</strong> ozone contactors at various configurations <strong>and</strong> operatingconditions. The model was especially useful in assessing the actual C. parvum inactivation levelin full-scale ozone contactors for which testing with actual microorganisms is currentlyimpossible. Finally, it is expected that the model should provide a promising approach tooptimize <strong>and</strong> design ozone contactors.123


CHAPTER 9. SUMMARY AND CONCLUSIONSFacilities attempting to comply with regulations governing pathogen inactivation <strong>and</strong>disinfection by-product minimization are faced with the challenge <strong>of</strong> balancing their systems <strong>and</strong>optimizing their treatment processes. Impending regulations are expected to require up to 2.5-logs <strong>of</strong> additional Cryptosporidium treatment, potentially by disinfection. Ozone has beenshown to be an effective disinfectant for Cryptosporidium. The obvious implication to the use <strong>of</strong>ozone is the formation <strong>of</strong> bromate in excess <strong>of</strong> the current maximum contaminant level (MCL),with the potential for the MCL to be further decreased. Full-scale treatment plants will berequired to optimize processes to meet both requirements in reducing the acute <strong>and</strong> chronic risksassociated with Cryptosporidium <strong>and</strong> bromate.A survey <strong>of</strong> operating ozone facilities in North America <strong>and</strong> Europe was initially conducted tounderst<strong>and</strong> bromate occurrence under current levels <strong>of</strong> ozone application. The average bromateconcentration at 24 full-scale ozonation plants, based on 78 samples from three separatesampling campaigns for the existing levels <strong>of</strong> ozonation, was 3.9 ug/L. The average percentconversion <strong>of</strong> bromide to bromate was 6.7 percent. The 10th, 50th (median), <strong>and</strong> 90th percentilebromate values was determined to be 0.2, 1.2, <strong>and</strong> 13 ug/L, respectively. The survey indicatedthat there were some bromate issues with current ozonation practices at the full-scale (whichcurrently target Giardia inactivation). Specifically, several utilities were forming bromate atlevels in excess <strong>of</strong> the Stage 1 D/DBP Rule MCL <strong>of</strong> 10 ug/L; 11 percent <strong>of</strong> the samples analyzedwould be considered out <strong>of</strong> compliance. Future compliance may be difficult for a greaternumber <strong>of</strong> utilities when operating at disinfection levels capable <strong>of</strong> Cryptosporidiuminactivation.<strong>Bromate</strong> formation can be affected by many factors ranging from naturally occurring compounds(i.e., NOM), water quality <strong>and</strong> temperature. Underst<strong>and</strong>ing the influence <strong>of</strong> these parameters isparamount to developing minimization strategies. The ubiquitous presence <strong>of</strong> NOM in sourcewaters has implications to ozone reaction kinetics that influence bromate formation. Specificfractions (i.e., hydrophobic) are known to exert a substantial ozone dem<strong>and</strong> while producingproducts that continue the ozone decomposition cycle. The heterogeneity <strong>of</strong> NOM providesreactive sites for both the inhibition <strong>and</strong> the promotion <strong>of</strong> ozone decay, which in turn influencesthe oxidation potential <strong>of</strong> bromide, by ozone, into intermediate products that lead to theformation <strong>of</strong> bromate. Underst<strong>and</strong>ing the role <strong>of</strong> NOM <strong>and</strong> its operationally <strong>and</strong> functionallydefinedfractions contributes to the elucidation <strong>of</strong> specific control strategies that will minimizebromate formation.The background natural organic matter <strong>of</strong> 14 natural waters was characterized by measurements<strong>of</strong> the DOC, UV254 absorbance <strong>and</strong> UV2oo-30o absorbance. The NOM was also fractionated intohydrophobic, hydrophilic <strong>and</strong> transphilic fraction using macroporous, nonionic resins. Sizeexclusion chromatography was also used to determine the apparent molecular weight <strong>of</strong> thenatural organic matter. These various parameters were then statistically correlated to bromateformation under a set <strong>of</strong> st<strong>and</strong>ard ozonation conditions (true-batch ozone reactor; temperature =20°C; ozone dose = 1 mg ozone per mg DOC). For the 14 waters studied, the DOCconcentrations ranged from 1.4 to 7.3 mg/L <strong>and</strong> had a median bromide concentration <strong>of</strong> 42 ug/L.124


The SUVA values ranged from 1.4 to 3.8 L/mg-m. The average hydrophobic fraction accountedfor 50 percent <strong>of</strong> the NOM, with the transphilic <strong>and</strong> hydrophilic fractions averaging 21 <strong>and</strong> 29percent <strong>of</strong> the NOM, respectively. While all <strong>of</strong> the NOM fractions exerted an ozone dem<strong>and</strong>, thehydrophobic fraction was found to exert the highest dem<strong>and</strong> per unit DOC. Based on thest<strong>and</strong>ard bromate formation testing the hydrophobic fraction was negatively correlated withbromate formation, while the transphilic <strong>and</strong> hydrophilic fractions were positively correlated.<strong>Bromate</strong> formation in these waters was evaluated over a range (0.5-log to 3.0-logs) <strong>of</strong> predictedCryptosporidium inactivation levels. At 1-log Cryptosporidium inactivation, approximately half<strong>of</strong> the waters were expected to produce bromate in excess <strong>of</strong> 10 ug/L. It was noted that whilethese waters had a range <strong>of</strong> water qualities <strong>and</strong> bromide concentrations, the higher theconcentration <strong>of</strong> bromide the higher the percent <strong>of</strong> the bromide that was converted to bromate.To reduce the concentration <strong>of</strong> bromate formed, pH depression, ammonia addition <strong>and</strong> hydroxylradical scavenger addition were evaluated. By depressing the pH from 8 to 7 to 6, a generalreduction in bromate formation <strong>of</strong> 30 to 50 percent per unit decrease in pH was observed. Thisreduction in bromate formation resulted in an increase in the concentration <strong>of</strong> total organicbromide formed. The addition <strong>of</strong> 0.5 mg/L ammonia-nitrogen to these waters resulted in areduction in bromate formation for all <strong>of</strong> the waters that would normally form over 10 ug/L <strong>of</strong>bromate. Increasing the ammonia concentration to 1.0 mg/L ammonia-nitrogen resulted in someadditional reduction in bromate formation. It was estimated that ammonia addition might be aneffective bromate formation control strategy for those waters within 10 to 20 ug/L <strong>of</strong> the MCL.It was also revealed that the cumulative effects pH depression together with ammonia additioncould be used for waters in which bromate formation was particularly problematic.A preliminary investigation <strong>of</strong> the addition <strong>of</strong> a hydroxyl radical scavenger revealed the addition<strong>of</strong> 1 mM <strong>of</strong> t-butanol prevented bromate formation. While the addition <strong>of</strong> t-butanol to municipaldrinking water supply might not currently be an acceptable practice, it does illustrate theeffectiveness <strong>of</strong> this approach. This illustrates the need for additional research to identify anacceptable hydroxyl radical scavenger <strong>and</strong> determine the minimum amount necessary toeffectively inhibit bromate formation. It was just as important to note that since the addition <strong>of</strong> ahydroxyl radical scavenger completely inhibited bromate formation, it appeared that bromateformation in natural waters does not proceed through the "direct pathway" as discussed by Songet al. (1997), but must proceed through pathways in which hydroxyl radicals are needed.Consequently, the radical scavenging by natural organic matter likely plays a significant role inthe differences observed in bromate formation between natural waters.In addition to these chemical approaches to bromate minimization, hydrodynamic strategies werealso investigated. However, through bench-scale <strong>and</strong> pilot-scale experiments, no compellingreason was identified to employ staged/tapered ozonation for bromate minimization. It was alsodiscovered that bromate formation could not be reduced by operating the ozone contactors ineither a co-current or counter-current mode. Nevertheless, the operation <strong>of</strong> the ozone contactorin a co-current mode might have provided subtle opportunities for optimizing disinfection. As aresult, these tests provided indications that bromate formation was predominantly controlled bychemical conditions <strong>and</strong> not hydraulic parameters.125


The temperature was observed to reduce bromate formation in true-batch, semi-batch <strong>and</strong>continuous-flow laboratory reactors. However, its impacts on the amount <strong>of</strong> bromate formed fora given level <strong>of</strong> Cryptosporidium disinfection varied between these reactors. In the true-batchexperiments, decreasing the temperature from 20°C to 10°C resulted in about a 40 percentreduction in the bromate formed for 2-logs <strong>of</strong> Cryptosporidium inactivation. Using thelaboratory-scale continuous flow reactor, decreasing the temperature from 25°C to 5°C, resultedin a 15 to 72 percent increase in the amount <strong>of</strong> bromate formed for 2-logs <strong>of</strong> Cryptosporidiuminactivation. An approximate 50 percent increase in the amount <strong>of</strong> bromate formed for 2-logs <strong>of</strong>Cryptosporidium inactivation was also observed in a semi-batch reactor as the temperature wasdecreased from 22°C to 7°C. It is unclear why different results were observed for these differentreactors.Many <strong>of</strong> the above trends were developed through bench-scale tests. Concerns over theapplicability <strong>of</strong> these results to larger scale (pilot-scale <strong>and</strong> full-scale) reactors were addressedthrough a series <strong>of</strong> comparability tests. It was shown that bench-scale reactors could providereasonable simulations <strong>of</strong> pilot-scale <strong>and</strong> full-scale bromate formation results provided anaccurate estimate <strong>of</strong> the amount <strong>of</strong> ozone contact at the larger scale could be generated. As such,the approximation <strong>of</strong> ozone contact by CT, as defined by the SWTR, resulted in a lack <strong>of</strong>comparability. By calculating ozone contact with the hydraulic retention time instead <strong>of</strong> tio, <strong>and</strong>by giving credit to the first cell <strong>of</strong> the pilot-scale <strong>and</strong> full-scale reactors correlations to the benchscaleresults could be developed. Likewise, pilot-scale simulations could be used to provideapproximations <strong>of</strong> full-scale results. Such simulations could be useful in assessing trade<strong>of</strong>fsbetween bromate formation <strong>and</strong> disinfection (CT) under various scenarios. However, theseresults also illustrated that while the simulations could be used for trending purposes, the specificbromate concentrations formed for given amounts <strong>of</strong> ozone contact do not always match betweendifferent contactors. The results from these tests also lent additional credence to the theory thatchemical conditions, as opposed to hydraulic conditions, control the formation <strong>of</strong> bromate.The results <strong>of</strong> these reactor scale comparability tests were also used to validate the performance<strong>of</strong> a series <strong>of</strong> integrated models that could be used to predict ozone decomposition, bromateformation <strong>and</strong> Cryptosporidium inactivation in four different types <strong>of</strong> reactors. While limitationsexisted, it was demonstrated that the model simulations could be used to underst<strong>and</strong> theperformance <strong>of</strong> the ozone contactors at various configurations <strong>and</strong> operating conditions.126


APPENDIX1 . Solver routines for bromate formation <strong>and</strong> Cryptosporidium parvum inactivation in batch <strong>and</strong>PFR-side PFR reactor. These routines are utilized by a main function that should beprogrammed for the specific purposes. Main functions for different modeling were notincluded./*Solve the set <strong>of</strong> n Linear Equations AX=B/*******************#*******#****#*********/void ludcmp(double a[][N+l], int *indx, double *d){int i,imaxj,k;double big,dum,sum,temp;double w[N+l];*d=1.0;big=0.0;for(j=la big) big=temp;if (big = 0.0) break;w[i]=1.0/big;sum=a[i][j];for (k=l;k


for(i=l;i


if (k = kmax || k = kopt+1) {red=SAFE2/err[km];break;}else if (k = kopt && alf[kopt-l][kopt] < err[km]) {red=1.0/err[km];break;}else if (kopt = kmax && alf[km][kmax-l]


2. Routines that provide kinetic information for bromate formation. These routine are utilizedonly for the case <strong>of</strong> batch <strong>and</strong> PFR-PFR simulations _____________________double k24a = l.OelO;/*********##********#***#*#*#***********#/ double k24b = 3.3e7;/*Variable Definitiondouble k25a = 4.0e6;/****** #***************#**#**************/ double k25b = 1.3elO;double k26 = 4.4e 10;[03]double k27 = 2.0e8;// y[2] = CT_H03double k28a = l.OelO;// y[3] = VACANTdouble k28b=1.0e5;// y[4] = CT_HO2double k29=2.0e9;// y[5] = VACANTdouble k30a = 8.3e8;// y[6] = [OH]Radicaldouble k30b = l.OelO;CT_H2O2double k31 a = 4.6e2;// y[8] = CT_C03double k31b = 7.9elO;// y[9] = CT_C03rdouble k33 = 1.5e8;//y[10] =CT_PO4double k34 = 4.0e9;= [Br-]double k35=4.5e9;= [Br]Radicaldouble k36 = 2.0e9;// y[13] = [BrOH-]Radicaldouble k37 = 3.5e9;//y[14] = [Br2-]Radicaldouble k38 = 5.0e9;//y[15] = [Br3-]double k39 = 3.4e8;//y[16] =CT_HOBrdouble k40 = 8.0e7;= [Br2]double k41 = 1.9e9;= [BrO]Radicaldouble k42=2.0e9;= [BrO2-]double k43 a = 1.4e9;y[20] = [BrO2]Radicaldouble k43b = 7.0e7;= [Br2O4]double k44 = 7.0e8;// y[22] = [BrO3-]double k53 = 8.5e6;// y[23] = [t-BuOH]double k54 = 3.9e8;//y[24] =Ct_NH3double k55=4.3e7;// y[25] = [NH2Br]double k56 = l.le8;// y[26] = NOM1double k63 = 2.0e4;//y[27] =NOM2double k64 = 1.5e5;//y[28] =NOM3double k71=2.8e6;//y[29] =TOBrdouble k72=2.7e7;double k73 = 7.5e9;/****************************************/ double k74 = 7.6e8;/*Rate Constants <strong>of</strong> Reactionsdouble k75= 3.5e9;/*************************#****#*********/ double k81 = 8.0e7;double k82=4.0el;double kl = 70;double k90 = 5.9e8;double k3 = 1.6e9;double klOO;double k5 = Lle5;double klOl =4.0e8;double k6 = 2.6e8;double kl 02 = 1.0e3;double k7 = 5.0e9;double Alpha;// Fraction to Hydroxyl Radicaldouble k8 = l.OelO;double Beta;// Fraction to Superoxide Radicaldouble k9 = 5.0e9;double Gamma;// Fraction to <strong>Bromide</strong>double klO = l.OelO;double PRT, HYD;double kll = 5.0e9;double aO_HO3, al_HO3;double k20 = 1.6e2;double aO_HO2, al_HO2;double k21 = 3.3e2;double aO_Peroxide, al_Peroxide;double k22 = 1.0e2;double aO_CO3r, al_CO3r;double k23 = 1.0e5;double aO_HOBr, al_HOBr;130


double aO_CO3, al_CO3, a2_CO3;double aO_PO4, al_PO4, a2_PO4, a3_PO4;double aO_NH3, al_NH3;/****************************************//* Acid-Base Equilibrium Calculation/****************************************/void pHequilibrium(double pH) {// Hydrogen Ion ConcentrationPRT=pow(10,-pH);HYD =pow(10,-14.17)/PRT;//H03/03-PairaO_HO3 = PRT / ( pow (10, -8.2) + PRT );al_HO3 = l.F-aO_HO3;//H02/02-PairaO_HO2 = PRT / ( pow (10, -4.8) + PRT );al_HO2 = l.F-aO_HO2;// H2O2/HO2- PairaO_Peroxide = PRT /( pow (10, -1 1 .6) + PRT );al_Peroxide = l.F - aO_Peroxide;// HCO3.radical/CO3-.radical PairaO_CO3r = PRT / ( pow (10, -9.6) + PRT );al_CO3r=l.F-aO_CO3r;//HOBr/OBr-PairaO_HOBr = PRT / ( pow (10, -8.8) + PRT );al_HOBr = l.F - aO_HOBr;// Carbonate SystemaO_CO3 = PRT * PRT / ( PRT * PRT + PRT * pow(10, -6.35) + pow (10, -6.35) * pow (10, -10.35) );al_CO3 = aO_CO3 / PRT * pow(10, -6.35);a2_CO3 = l.F - aO_CO3 - al_CO3;// Phosphate SystemaO_PO4 = PRT * PRT * PRT /( PRT * PRT * PRT +PRT * PRT * pow(10, -2.3) + PRT * pow(10,-2.3) *pow(10,-7.2) + pow(10,-2.3) * pow(10,-7.2) *pow(10,-12.3));al_PO4 = aO_PO4 / PRT * pow(10, -2.3);a2_PO4 = al_PO4 / PRT * pow(10, -7.2);a3_PO4 = l.F - aO_PO4 - al_PO4 - a2_PO4;//NH4+/NH3PairaO_NH3 = PRT / ( pow (10, -9.3) + PRT );al_NH3 = l.F-aO_NH3;/****************************************//* Analytical Evaluation <strong>of</strong> Jacobiany**# *************************************/void jacobn(double x, double y[], double dfdx[],double dfdy[][N+l])int i;for(i=l;i


dfdy[2][9]=0.0;dfdy[2][10]=0.0;dfdy[2][ll] = 0.0;dfdy[2][12]=0.0;dfdy[2][13] = 0.0;dfdy[2][14] = 0.0;dfdy[2][15]=0.0;dfdy[2][16] = 0.0;dfdy[2][17] = 0.0;dfdy[2][18]=0.0;dfdy[2][19]=0.0;dfdy[2][20] = 0.0;dfdy[2][21] = 0.0;dfdy[2][22] = 0.0;dfdy[2][23]=0.0;dfdy[2][24] = 0.0;dfdy[2][25]=0.0;dfdy[2][26] = Alpha * klOO * y[l];dfdy[2][27] = 0.0;dfdy[2][28]=0.0;dfdy[2][29] = 0.0;dfdy[3][l]=0.0;dfdy[3][2]=0.0;dfdy[3][3] = 0.0;dfdy[3][4]=0.0;dfdy[3][5]=0.0;dfdy[3][6] = 0.0;dfdy[3][7] = 0.0;dfdy[3][8]=0.0;dfdy[3][9] = 0.0;dfdy[3][10]=0.0;dfdy[3][ll]=0.0;dfdy[3][12] = 0.0;dfdy[3][13] = 0.0;dfdy[3][14]=0.0;dfdy[3][15]=0.0;dfdy[3][16] = 0.0;dfdy[3][17]=0.0;dfdy[3][18]=0.0;dfdy[3][19]=0.0;dfdy[3][20] = 0.0;dfdy[3][21] = 0.0;dfdy[3][22]=0.0;dfdy[3][23] = 0.0;dfdy[3][24] = 0.0;dfdy[3][25] = 0.0;dfdy[3][26] = 0.0;dfdy[3][27] = 0.0;dfdy[3][28] = 0.0;dfdy[3][29] = 0.0;dfdy[4][l] = 2*kl * HYD - k3 * al_HO2 * y[4] + k6* y[6] + k71 * al_Peroxide * y[7];dfdy[4][2] = - klO * al_HO2 * y[4] * aO_HO3;dfdy[4][3]=0.0;dfdy[4][4] = - k3 * y[l] * al_HO2 - k8 * y[6] *al_HO2 - klO * al_HO2 * aO_HO3 * y[2] - k37 *al_H02*aO_HOBr*y[16];dfdy[4][5]=0.0;dfdy[4][6] = k6 * y[l] - k8 * al_HO2 * y[4] + k72 *aO_Peroxide * y[7] + k73 * al_Peroxide * y[7] +Beta * klOl * y[27];dfdy[4][7] = k71 * y[l] * al_Peroxide + k72 *aO_Peroxide * y[6] + k73 * al_Peroxide * y[6];dfdy[4][8]=0.0;dfdy[4][9] = 0.0;dfdy[4][10] = 0.0;dfdy[4][ll]=0.0;dfdy[4][12]=0.0;dfdy[4][13]=0.0;dfdy[4][14]=0.0;dfdy[4][15]=0.0;dfdy[4][16] = - k37 * al_HO2 * y[4] * aO_HOBr;dfdy[4][17]=0.0;dfdy[4][18]=0.0;dfdy[4][19]=0.0;dfdy[4][20] = Q.O;dfdy[4][21] = 0.0;dfdy[4][22] = 0.0;dfdy[4][23] = 0.0;dfdy[4][24] = 0.0;dfdy[4][25]=0.0;dfdy[4][26]=0.0;dfdy[4][27] = Beta * klOl * y[6];dfdy[4][28] = 0.0;dfdy[4][29] = 0.0;dfdy[5][l]=0.0;dfdy[5][2]=0.0;dfdy[5][3]=0.0;dfdy[5][4] = 0.0;dfdy[5][5]=0.0;dfdy[5][6] = 0.0;dfdy[5][7]=0.0;dfdy[5][8]=0.0;dfdy[5][9]=0.0;dfdy[5][10] = 0.0;dfdy[5][ll] = 0.0;dfdy[5][12]=0.0;dfdy[5][13]=0.0;dfdy[5][14]=0.0;dfdy[5][15] = 0.0;dfdy[5][16]=0.0;dfdy[5][17]=0.0;dfdy[5][18] = 0.0;dfdy[5][19]=0.0;dfdy[5][20] = 0.0;dfdy[5][21]=0.0;dfdy[5][22]=0.0;132


dfdy[5][23]=0.0;dfdy[5][24] = 0.0;dfdy[5][25] = 0.0;dfdy[5][26] = 0.0;dfdy[5][27] = 0.0;dfdy[5][28] = 0.0;dfdy[5][29] = 0.0;dfdy[6][l] = - k6 * y[6] + k71 * alJPeroxide * y[7];dfdy[6][2] = k5 * aO_HO3 - k9 * y[6] * aO_HO3;dfdy[6][3] = 0.0;dfdy[6][4] = - k8 * y[6] * al_HO2;dfdy[6][5] = 0.0;dfdy[6][6] = - k6 * y[l] - 4*k7 * y[6] - k8 * al_HO2* y[4] - k9 * aO_HO3 * y[2] - k24a * y[l 1] - k35 *al_HOBr * y[16] - k36 * aO_HOBr * y[16] - k41 *y[19] - k42 * y[20] - k53 * al _CO3 * y[8] - k54 *a2_CO3 * y[8] - k63 * al_PO4 * y[10] - k64 *a2_PO4 * y[10] - k72 * aO_Peroxide * y[7] - k73 *al_Peroxide * y[7] - k90 * y[23] - klOl * y[27];dfdy[6][7] = k71 * y[l] * al_Peroxide - k72 *aO_Peroxide * y[6] - k73 * al J>eroxide * y[6];dfdy[6][8] = - k53 * al_CO3 * y[6] - k54 * a2_CO3* y[6];dfdy[6][9]=0.0;dfdy[6][10] = - k63 * y[6] * al_PO4 - k64 * y[6] *a2_PO4;dfdy[6][ll]=-k24a*y[6];dfdy[6][12] = 0.0;dfdy[6][13]=k24b;dfdy[6][14] = 0.0;dfdy[6][15]=0.0;dfdy[6][16] = - k35 * y[6] * al_HOBr - k36 * y[6] *aO_HOBr;dfdy[6][17] = 0.0;dfdy[6][18]=0.0;dfdy[6][19]=-k41 * y[6];dfdy[6][20] = - k42 * y[6];dfdy[6][21]=0.0;dfdy[6][22]=0.0;dfdy[6][23]=-k90*y[6];dfdy[6][24] = 0.0;dfdy[6][25] = 0.0;dfdy[6][26] = 0.0;dfdy[6][27]=-k!01*y[6];dfdy[6][28]=0.0;dfdy[6][29] = 0.0;dfdy[7][l] = - k71 * al_Peroxide * y[7];dfdy[7][2] = k9 * y[6] * aO_HO3 + 2* kl 1 * aO_HO3* aO_H03 * y[2];dfdy[7][3] = 0.0;dfdy[7][4] = 0.0;dfdy[7][5] = 0.0;dfdy[7][6] = 2 * k7 * y[6] + k9 * aO_HO3 * y[2] -k72 * aO_Peroxide * y[7] - k73 * al_Peroxide * y[7];dfdy[7][7] = - k71 * y[l] * alJPeroxide - k72 *aO_Peroxide * y[6] - k73 * al_Peroxide * y[6] - k74* al_Peroxide * aO_HOBr * y[16];dfdy[7][8] = 0.0;dfdy[7][9] = 0.0;dfdy[7][10]=0.0;dfdy[7][ll]=0.0;dfdy[7][12.]=0.0;dfdy[7][13]=0.0;dfdy[7][14] = 0.0;dfdy[7][15]=0.0;dfdy[7][16] = - k74 * al_Peroxide * y[7] *aOJHOBr;dfdy[7][17]=0.0;dfdy[7][18]=0.0;dfdy[7][19] = 0.0;dfdy[7][20] = 0.0;dfdy[7][21]=0.0;dfdy[7][22]=0.0;dfdy[7][23] = 0.0;dfdy[7][24]=0.0;dfdy[7][25] = 0.0;dfdy[7][26] = 0.0;dfdy[7][27] = 0.0;dfdy[7][28] = 0.0;dfdy[7][29] = 0.0;dfdy[8][l]=0.0;dfdy[8][2] = 0.0;dfdy[8][3] = 0.0;dfdy[8][4] = 0.0;dfdy[8][5] = 0.0;dfdy[8][6] = - k53 * al_CO3 * y[8] - k54 * a2_CO3* y[8];dfdy[8][7] = 0.0;dfdy[8][8] = - k53 * al_CO3 * y[6] - k54 * a2_CO3*y[6];dfdy[8][9] = k55 * al_CO3r * al_HOBr * y[16] +k56*al_CO3r*y[19];dfdy[8][10]=0.0;dfdy[8][ll] = 0.0;dfdy[8][12] =0.0;dfdy[8][13] = 0.0;dfdy[8][14] = 0.0;dfdy[8][15]=0.0;dfdy[8][16] =k55 * al_CO3r * y[9] * al_HOBr;dfdy[8][17]=0.0; •dfdy[8][18] = 0.0;dfdy[8][19] = k56 * al_CO3r * y[9];dfdy[8][20]=0.0;dfdy[8][21] = 0.0;dfdy[8][22] = 0.0;dfdy[8][23]=0.0;dfdy[8][24] = 0.0;dfdy[8][25] = 0.0;dfdy[8][26] = 0.0;133


dfdy[8][27]=0.0;dfdy[8][28] = 0.0;dfdy[8][29] = 0.0;dfdy[9][l]=0.0;dfdy[9][2]=0.0;dfdy[9][3] = 0.0;dfdy[9][4]=0.0;dfdy[9][5]=0.0;dfdy[9][6] = k53 * al_CO3 * y[8] + k54 * a2_CO3 *y[8];dfdy[9][7]=0.0;dfdy[9][8] = k53 * al_CO3 * y[6] + k54 * a2_CO3 *y[6];dfdy[9][9] = - k55 * al_CO3r * al_HOBr * y[16] -k56 * al_C03r * y[19];dfdy[9][10] = 0.0;dfdy[9][ll] = 0.0;dfdy[9][12] = 0.0;dfdy[9][13] = 0.0;dfdy[9][14] = 0.0;dfdy[9][15] = 0.0;dfdy[9][16] = - k55 * al_CO3r * y[9] * al_HOBr;dfdy[9][17] = 0.0;dfdy[9][18] = 0.0;dfdy[9][19] = - k56 * al_CO3r * y[9];dfdy[9][20] = 0.0;dfdy[9][21]=0.0;dfdy[9][22] = 0.0;dfdy[9][23] = 0.0;dfdy[9][24] = 0.0;dfdy[9][25] = 0.0;dfdy[9][26] = 0.0;dfdy[9][27] = 0.0;dfdy[9][28] = 0.0;dfdy[9][29]=0.0;dfdy[10][l] = 0.0;dfdy[10][2] = 0.0;dfdy[10][3]=0.0;dfdy[10][4]=0.0;dfdy[10][5]=0.0;dfdy[10][6] = - k63 * al_PO4 * y[10] - k64 *a2_PO4 * y[10];dfdy[10][7]=0.0;dfdy[10][8] = 0.0;dfdy[10][9] = 0.0;dfdy[10][10] = - k63 * y[6] * al_PO4 - k64 * y[6] *a2_PO4;dfdy[10][ll] = 0.0;dfdy[10][12]=0.0;dfdy[10][13] = 0.0;dfdy[10][14] = 0.0;dfdy[10][15]=0.0;dfdy[10]f!6]=0.0;dfdy[10][17] = 0.0;dfdy[10][18] = 0.0;dfdy[10][19] = 0.0;dfdy[10][20] = 0.0;dfdy[10][21] = 0.0;dfdy[10][22] = 0.0;dfdy[10][23] = 0.0;dfdy[10][24] = 0.0;dfdy[10][25] = 0.0;dfdy[10][26] = 0.0;dfdy[10][27] = 0.0;dfdy[10][28] = 0.0;dfdy[10][29] = 0.0;dfdy[ll][l] = - k20 * y[ll] + k21 * al_HOBr * y[16]+ k82 * y[25];dfdy[ll][2] = 0.0;dfdy[ll][3] = 0.0;dfdy[ll][4]=0.0;dfdy[ll][5]=0.0;dfdyfl I]f6] = -k24a*y[ll];dfdy[l 1][7] = k74 * al_Peroxide * aO_HOBr * y[16];dfdy[ll][8]=0.0;dfdyfl 1][9] = 0.0;dfdy[ll][10] = 0.0;dfdy[l 1][1 1] = - k20 * y[l] - k24a * y[6] - k27 *y[13] - k28a * y[12] - k30b * y[17] - k31b *aO_HOBr * y[16] * PRT;dfdyfl l]f 12] = -k28a*y[ll]+k34*al_HOBr*dfdyfl 1][13] = k24b - k27 * y[ll];dfdyfl 1][14] = k28b + 2 * k29 * y[14] + k40 * y[19];dfdyfl I][15] = k30a;dfdy[ll][16] = k21 * y[l] * al_HOBr-k31b *aO_HOBr * PRT * y[l 1] + k34 * y[12] * al_HOBr +k74 * al_Peroxide * y[7] * aO_HOBr + Gamma *k!02 * y[28];dfdy[l 1][17] = - k30b * y[l 1] + k31a;dfdyfl 1][1 8] = 0.0;dfdyfl I][19] = k40*y[14];dfdyfl 1][20] = 0.0;dfdyfl 1][21] = 0.0;dfdy[ll][22] = 0.0;dfdyfl 1][23] = 0.0;dfdyfl 1][24] = 0.0;dfdyfl I][25] = k82*y[l];dfdyfll][26] = 0.0;dfdyfl 1][27] = 0.0;dfdyfl 1][28] = Gamma * k!02 * y[16];dfdy[ll][29] = 0.0;dfdy[12][l]=-k33*y[12];dfdy[12][2]=0.0;dfdy[12][3] = 0.0;dfdy[12][4] = k37 * al_HO2 * aO_HOBr * y[16];dfdy[12][5]=0.0;dfdyfl 2] [6] = 0.0;134


dfdy[12][7]=0.0;dfdy[12][8]=0.0;dfdy[12][9] = 0.0;dfdy[12][10]=0.0;dfdy[12][ll] = -k28a*y[12];dfdy[12][12] = - k25b * HYD - k28a * y[l 1] - k33 *y[l]-k34*al_HOBr*y[16];dfdy[12][13] = k25a + k26 * PRT;dfdy[12][14]=k28b;dfdy[12][15]=0.0;dfdy[12][16] = - k34 * y[12] * al_HOBr + k37 *al_HO2 * y[4] * aOJHOBr;dfdy[12][17] = 0.0;dfdy[12][18] = 0.0;dfdy[12][19] = 0.0;dfdy[12][20] = 0.0;dfdy[12][21] = 0.0;dfdy[12][22] = 0.0;dfdy[12][23] = 0.0;dfdy[12][24] = 0.0;dfdy[12][25] = 0.0;dfdy[12][26] = 0.0;dfdy[12][27] = 0.0;dfdy[12][28] = 0.0;dfdy[12][29] = 0.0;dfdy[13][l] = 0.0;dfdy[13][2] = 0.0;dfdy[13][3] = 0.0;dfdy[13][4]=0.0;dfdy[13][5] = 0.0;dfdy[13][6]=k24a*y[ll];dfdy[13][7] = 0.0;dfdy[13][8] = 0.0;dfdy[13][9] = 0.0;dfdy[13][10] = 0.0;dfdy[13][l 1] = k24a * y[6] - k27 * y[13];dfdy[13][12] = k25b * HYD;dfdy[13][13] = - k24b - k25a - k26 * PRT - k27 *dfdy[13][14] = 0.0;dfdy[13][15] = 0.0;dfdy[13][16] = 0.0;dfdy[13][17] = 0.0;dfdy[13][18]=0.0;dfdy[13][19] = 0.0;dfdy[13][20] = 0.0;dfdy[13][21] = 0.0;dfdy[13][22] = 0.0;dfdy[13][23] = 0.0;dfdy[13][24]=0.0;dfdy[13][25] = 0.0;dfdy[13][26] = 0.0;dfdy[13][27]=0.0;dfdy[13][28] = 0.0;dfdy[13][29]=0.0;dfdy[14][l]=0.0;dfdy[14][2] = 0.0;dfdy[14][3]=0.0;dfdy[14][4]=0.0;dfdy[14][5] = 0.0;dfdy[14][6] = 0.0;dfdy[14][7]=0.0;dfdy[14][8] = 0.0;dfdy[14][9]=0.0;dfdy[14][10] = 0.0;dfdy[14][l 1] = k27 * y[13] + k28a * y[12];dfdy[14][12] = k28a*y[ll];dfdy[14][13] = k27*y[ll];dfdy[14][14] = - k28b - 4*k29 * y[14] - k40 *dfdy[14][15] = 0.0;dfdy[14][16] = 0,0;dfdy[14][17] = 0.0;dfdy[14][18] = 0.0;dfdy[14][19] = -k40*y[14];dfdy[14][20] = 0.0;dfdy[14][21] = 0.0;dfdy[14][22] = 0.0;dfdy[14][23]=0.0;dfdy[14][24] = 0.0;dfdy[14][25] = 0.0;dfdy[14][26] = 0.0;dfdy[14][27] = 0.0;dfdy[14][28] = 0.0;dfdy[14][29] = 0.0;dfdy[15][l] =dfdytl5][2] =dfdy[15][3] =dfdy[15][4] =dfdy[15][5] =dfdy[15][6] =dfdy[15][7] =dfdy[15][8] =0.0;0.0;0.0;0.0;0.0;0.0;0.0;0.0;dfdy[15][9] = 0.0;dfdy[15][10]dfdy[15][ll]dfdy[15][12]dfdy[15][13]dfdy[15][14]dfdy[15][15]dfdy[15][16]dfdy[15][17]dfdy[15][18]dfdy[15][19]dfdy[15][20]dfdy[15][21]dfdy[15][22]dfdy[15][23]dfdy[15][24]dfdy[15][25]= 0.0;= k30b*y[17];= 0.0;= 0.0;= 2*k29*y[14];= - k30a;= 0.0;= k30b*y[ll];= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;135


dfdy[15][26] = 0.0;dfdy[15][27] = 0.0;dfdy[15][28] = 0.0;dfdy[15][29] = 0.0;dfdy[16][l] =k20 * y[ll] -k21 * al_HOBr * y[16] -k22*al_HOBr*y[16];dfdy[16][2] = 0.0;dfdy[16][3] = 0.0;dfdy[16][4] = - k37 * al_HO2 * aO_HOBr * y[16];dfdy[16][5]=0.0;dfdy[16][6] = - k35 * al_HOBr * y[16] - k36 *aO_HOBr*y[16];dfdy[16][7] = - k74 * al_Peroxide * aO_HOBr *dfdy[16][8] = 0.0;dfdy[16][9] = - k55 * al_CO3r * al_HOBr * y[16];dfdy[16][10] = 0.0;dfdy[16][ll] =k20 * y[l] - k31b * aO_HOBr * y[16]*PRT;dfdy[16][12] = - k34 * al_HOBr * y[16];dfdy[16][13] = 0.0;dfdy[16][14]=k40*y[19];dfdy[16][15] = 0.0;dfdy[16][16] = - k21 * y[l] * al_HOBr - k22 * y[l]* al_HOBr - k3 Ib * aO_HOBr * PRT * y[l 1] - k34 *y[12] * al_HOBr - k35 * y[6] * al_HOBr - k36 *y[6] * aO_HOBr - k37 * al_HO2 * y[4] * aO_HOBr -k55 * al_CO3r * y[9] * al_HOBr - k74 *al_Peroxide * y[7] * aO_HOBr - k81 * aO_HOBr *al_NH3 * y[24] - k!02 * y[28];dfdy[16][17] = k31a;dfdy[16][18] = 2 * k38 * y[18] + k39 * y[19];dfdy[16][19] = k39 * y[18] + k40 * y[14];dfdy[16][20] = 0.0;dfdy[16][21] = 0.0;dfdy[16][22]=0.0;dfdy[16][23] = 0.0;dfdy[16][24] = - k81 * aO_HOBr * y[16] * al_NH3;dfdy[16][25] = 0.0;dfdy[16][26] = 0.0;dfdy[16][27]=0.0;dfdy[16][28] = - k!02 *dfdy[16][29] = 0.0;dfdy[17][l] = 0.0;dfdy[17][2] = 0.0;dfdy[17][3] = 0.0;dfdy[17][4] = 0.0;dfdy[17][5] = 0.0;dfdy[17][6]=0.0;dfdy[17][7] = 0.0;dfdy[17][8] = 0.0;dfdy[17][9]=0.0;dfdy[17][10] = 0.0;dfdy[17][l 1] = - k30b * y[17] + k31b * aO_HOBr *y[16]*PRT;dfdy[17][12] = 0.0;dfdy[17][13]= 0.0;dfdy[17][14] = 0.0;dfdy[17][15] = k30a;dfdy[17][16] = k31b * aO_HOBr * PRT * y[l 1];dfdy[17][17] = - k30b * y[l 1] - k31a;dfdy[17][18] = 0.0;dfdy[17][19] = 0.0;dfdy[17][20] = 0.0;dfdy[17][21] = 0.0;dfdy[17][22] = 0.0;dfdy[17][23] = 0.0;dfdy[17][24] = 0.0;dfdy[17][25] = 0.0;dfdy[17][26] = 0.0;dfdy[17][27] = 0.0;dfdy[17][28] = 0.0;dfdy[17][29] = 0.0;dfdy[18][l]=k33*y[12];dfdy[18][2]=0.0;dfdy[18][3]=0.0;dfdy[18][4]=0.0;dfdy[18][5]=0.0;dfdy[18][6] = k35 * al_HOBr * y[16] + k36 *aO_HOBr*y[16];dfdy[18][7] = 0.0;dfdy[18][8]=0.0;dfdy[18][9] = k55 * al_CO3r * al_HOBr * y[16];dfdy[18][10] = 0.0;dfdy[18][ll] = 0.0;dfdy[18][12] = k33 * y[l] + k34 * al_HOBr * y[16];dfdy[18][13] = 0.0;dfdy[18][14] = k40*y[19];dfdy[18][15] = 0.0;dfdy[18][16] = k34 * y[12] * al_HOBr + k35 * y[6]* al_HOBr + k36 * y[6] * aO_HOBr + k55 *al_C03r * y[9] * al_HOBr;dfdy[18][17] = 0.0;dfdy[18][18] = - 4 * k38 * y[18] - k39 * y[19];dfdy[18][19] = - k39 * y[18] + k40 * y[14];dfdy[18][20] = 0.0;dfdy[18][21] = 0.0;dfdy[18][22] = 0.0;dfdy[18][23] = 0.0;dfdy[18][24] = 0.0;dfdy[18][25] = 0.0;dfdy[18][26] = 0.0;dfdy[18][27] = 0.0;dfdy[18][28] = 0.0;dfdy[18][29] = 0.0;dfdy[19][l] = k22 * al_HOBr * y[16] - k23 * y[19];136


dfdy[19][2]=0.0;dfdy[19][3] = 0.0;dfdy[19][4]=0.0;dfdy[19][5]=0.0;dfdy[19][6] = -k41*y[19];dfdy[19][7]=0.0;dfdy[19][8] = 0.0;dfdy[19][9] = - k56 * al_CO3r * y[19];dfdy[19][10] = 0.0;dfdy[19][ll] = 0.0;dfdy[19][12] = 0.0;dfdy[19][13] = 0.0;dfdy[19][14] = -k40*y[19];dfdy[19][15]=0.0;dfdy[19][16] = k22 * y[l] * alJJOBr;dfdy[19][17] = 0.0;dfdy[19][18] = 2 * k38 * y[18] - k39 * y[19];dfdy[19][19] = - k23 * y[l] - k39 * y[18] - k40 *y[14] - k41 * y[6] - k56 * al_CO3r * y[9];dfdy[19][20] = 0.0;dfdy[19][21J=k44*HYD;dfdy[19][22]=0.0;dfdy[19][23] = 0.0;dfdy[19][24] = 0.0;dfdy[19][25]=0.0;dfdy[19][26] = 0.0;dfdy[19][27] = 0.0;dfdy[19][28] = 0.0;dfdy[19][29]=0.0;dfdy[20][l] = 0.0;dfdy[20][2] = 0.0;dfdy[20][3] = 0.0;dfdy[20][4] = 0.0;dfdy[20][5] = 0.0;dfdy[20][6] = k41 * y[19] - k42 * y[20];dfdy[20][7] = 0.0;dfdy[20][8] = 0.0;dfdy[20][9] = k56 * al_CO3r * y[19];dfdy[20][10] = 0.0;dfdy[20][ll] = 0.0;dfdy[20][12] = 0.0;dfdy[20][13]=0.0;dfdy[20][14]=0.0;dfdy[20][15] = 0.0;dfdy[20][16] = 0.0;dfdy[20][17]=0.0;dfdy[20][18]=k39*y[19];dfdy[20][19] = k39 * y[18] + k41 * y[6] +k56 *al_C03r * y[9];dfdy[20][20] = - k42 * y[6] - 4*k43a * y[20];dfdy[20][21] = 2*k43b;dfdy[20][22] = 0.0;dfdy[20][23] = 0.0;dfdy[20][24] = 0.0;dfdy[20][25] = 0.0;dfdy[20][26]=0.0;dfdy[20][27] = 0.0;dfdy[20][28] = 0.0;dfdy[20][29]=0.0;dfdy[2 !][!]= 0.0;dfdy[21][2] = 0.0;dfdy[21][3] = 0.0;dfdy[21][4] = 0.0;dfdy[21][5]=0.0;dfdy[21][6]=0.0;dfdy[21][7]=0.0;dfdy[21][8]=0.0;dfdy[21][9] = 0.0;dfdy[21][10] = 0.0;dfdy[21][ll] = 0.0;dfdy[21][12] = 0.0;dfdy[21][13] = 0.0;dfdy[21][14] = 0.0;dfdy[21][15] = 0.0;dfdy[21][16] = 0.0;dfdy[21][17] = 0.0;dfdy[21][18] = 0.0;dfdy[21][19] = 0.0;dfdy[21][20] = 2 * k43a * y[20];dfdy[21][21] = - k43b - k44 * HYD;dfdy[21][22]=0.0;dfdy[21][23] = 0.0;dfdy[21][24] = 0.0;dfdy[21][25] = 0.0;dfdy[21][26] = 0.0;dfdy[21][27]=0.0;dfdy[21][28] = 0.0;dfdy[21][29] = 0.0;dfdy[22][l] = k23*y[19];dfdy[22][2] = 0.0;dfdy[22][3] = 0.0;dfdy[22][4] = 0.0;dfdy[22][5] = 0.0;dfdy[22][6] = k42* - y[20];dfdy[22][7] = 0.0;dfdy[22][8] = 0.0;dfdy[22][9] = 0.0;dfdy[22][10]dfdy[22][ll]dfdy[22][12]dfdy[22][13]dfdy[22][14]dfdy[22][15]dfdy[22][16]dfdy[22][17]dfdy[22][18]dfdy[22][19]dfdy[22][20]dfdy[22][21]= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= k23*y[l];= k42*y[6];= k44*HYD;137


dfdy[22][22] = 0.0;dfdy[22][23] = 0.0;dfdy[22][24] = 0.0;dfdy[22][25] = 0.0;dfdy[22][26] = 0.0;dfdy[22][27] = 0.0;dfdy[22][28] = 0.0;dfdy[22][29] = 0.0;dfdy[23][l] =dfdy[23][2] =dfdy[23][3] =dfdy[23][4] =dfdy[23][5] =dfdy[23][6] =dfdy[23][7] =dfdy[23][8] =dfdy[23][9] =dfdy[23][10]dfdy[23][ll]dfdy[23][12]dfdy[23][13]dfdy[23][14]dfdy[23][15]dfdy[23][16]dfdy[23][17]dfdy[23][18]dfdy[23][19]dfdy[23][20]dfdy[23][21]dfdy[23][22]dfdy[23][23]dfdy[23][24]dfdy[23][25]dfdy[23][26]dfdy[23][27]dfdy[23][28]dfdy[23][29]0.0;0.0;0.0;0.0;0.0;- k90 * y[23];0.0;0.0;0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= - k90 * y[6];= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;= 0.0;dfdy[24][l]=0.0;dfdy[24][2] = 0.0;dfdy[24][3] = 0.0;dfdy[24][4] = 0.0;dfdy[24][5] = 0.0;dfdy[24][6]=0.0;dfdy[24][7] = 0.0;dfdy[24][8]=0.0;dfdy[24][9] = 0.0;dfdy[24][10] = 0.0;dfdy[24][ll]=0.0;dfdy[24][12]=0.0;dfdy[24][13] = 0.0;dfdy[24][14] = 0.0;dfdy[24][15] = 0.0;dfdy[24][16] = - k81 * aO_HOBr * al_NH3 * y[24];dfdy[24][17] = 0.0;dfdy[24][18] = 0.0;dfdy[24][19] = 0.0;dfdy[24][20] = 0.0;dfdy[24][21] = 0.0;dfdy[24][22] = 0.0;dfdy[24][23] = 0.0;dfdy[24][24] = - k81 * aO_HOBr * y[16] * al_NH3;dfdy[24][25] = 0.0;dfdy[24][26] = 0.0;dfdy[24][27] = 0.0;dfdy[24][28] = 0.0;dfdy[24][29] = 0.0;dfdy[25][l]=-k82*y[25];dfdy[25][2]=0.0;dfdy[25][3]=0.0;dfdy[25][4]=0.0;dfdy[25][5]=0.0;dfdy[25][6]=0.0;dfdy[25][7]=0.0;dfdy[25][8]=0.0;dfdy[25][9]=0.0;dfdy[25][10] = 0.0;dfdy[25][ll] = 0.0;dfdy[25][12]=0.0;dfdy[25][13]=0.0;dfdy[25][14]=0.0;dfdy[25][15] = 0.0;dfdy[25][16] = k81 * aO_HOBr * al_NH3 * y[24];dfdy[25][17]=0.0;dfdy[25][18]=0.0;dfdy[25][19] = 0.0;dfdy[25][20] = 0.0;dfdy[25][21]=0.0;dfdy[25][22]=0.0;dfdy[25][23] = 0.0;dfdy[25][24] = k81 * aO_HOBr * y[16] * al_NH3;dfdy[25][25]=-k82*y[l];dfdy[25][26] = 0.0;dfdy[25][27] = 0.0;dfdy[25][28] = 0.0;dfdy[25][29] = 0.0;dfdy[26][l]=-klOO*y[26];dfdy[26][2] = 0.0;dfdy[26][3]=0.0;dfdy[26][4] = 0.0;dfdy[26][5] = 0.0;dfdy[26][6]=0.0;dfdy[26][7] = 0.0;dfdy[26][8]=0.0;dfdy[26][9]=0.0;dfdy[26][10] = 0.0;dfdy[26][ll]=0.0;dfdy[26][12]=0.0;dfdy[26][13]=0.0;138


dfdy[26][14]=0.0;dfdy[26][15] = 0.0;dfdy[26][16]=0.0;dfdy[26][17] = 0.0;dfdy[26][18]=0.0;dfdy[26][19] = 0.0;dfdy[26][20] = 0.0;dfdy[26][21]=0.0; 'dfdy[26][22] = 0.0;dfdy[26][23] = 0.0;dfdy[26][24] = 0.0;dfdy[26][25] = 0.0;dfdy[26][26] = - klOO *dfdy[26][27] = 0.0;dfdy[26][28] = 0.0;dfdy[26][29] = 0.0;dfdy[27][l] = 0.0;dfdy[27][2] = 0.0;dfdy[27][3] = 0.0;dfdy[27][4] = 0.0;dfdy[27][5] = 0.0;dfdy[27][6] = -klOl *y[27];dfdy[27][7] = 0.0;dfdy[27][8] = 0.0;dfdy[27][9] = 0.0;dfdy[27][10] = 0.0;dfdy[27][ll] = 0.0;dfdy[27][12] = 0.0;dfdy[27][13] = 0.0;dfdyf27][14] = 0.0;dfdy[27][15] = 0.0;dfdy[27][16] = 0.0;dfdy[27][17] = 0.0;dfdy[27][18] = 0.0;dfdy[27][19] = 0.0;dfdy[27][20] = 0.0;dfdy[27][21] = 0.0;dfdy[27][22] = 0.0;dfdy[27][23] = 0.0;dfdy[27][24] = 0.0;dfdy[27][25] = 0.0;dfdy[27][26] = 0.0;dfdy[27][27] = -k!01*y[6];dfdy[27][28] = 0.0;dfdy[27][29] = 0.0;dfdy[28][l]=0.0;dfdy[28][2] = 0.0;dfdy[28][3] = 0.0;dfdy[28][4]=0.0;dfdy[28][5] = 0,0;dfdy[28][6] = 0.0;dfdy[28][7] = 0.0;dfdy[28][8]=0.0;dfdy[28][9] = 0.0;dfdy[28][10]=0.0;dfdy[28][ll] = 0.0;dfdy[28][12] = 0.0;dfdy[28][13] = 0.0;dfdy[28][14] = 0.0;dfdy[28][15] = 0.0;dfdy[28][16] = -k!02*y[28];dfdy[28][17] = 0.0;dfdy[28][18] = 0.0;dfdy[28][19] = 0.0;dfdy[28][20]=0.0;dfdy[28][21] = 0.0;dfdy[28][22] = 0.0;dfdy[28][23] = 0.0;dfdy[28][24] = 0.0;dfdy[28][25]=0.0;dfdy[28][26] = 0.0;dfdy[28][27] = 0.0;dfdy[28][28] = -k!02*y[16];dfdy[28][29] = 0.0;dfdy[29][l]=0.0;dfdy[29][2]=0.0;dfdy[29][3]=0.0;dfdy[29][4]=0.0;dfdy[29][5] = 0.0;dfdy[29][6] = 0.0;dfdy[29][7] = 0.0;dfdy[29][8]=0.0;dfdy[29][9] = 0.0;dfdy[29][10] = 0.0;dfdy[29][ll] = 0.0;dfdy[29][12]=0.0;dfdy[29][13] = 0.0;dfdy[29][14] = 0.0;dfdy[29][15] = 0.0;dfdy[29][16] = (1-Gamma) * k!02 * y[28];dfdy[29][17]=0.0;dfdy[29][18] = 0.0;dfdy[29][19] = 0.0;dfdy[29][20] = 0.0;dfdy[29][21] = 0.0;dfdy[29][22]=0.0;dfdy[29][23]=0.0;dfdy[29][24] = 0.0;dfdy[29][25] =- 0.0;dfdy[29][26] = 0.0;dfdy[29][27] = 0.0;dfdy[29][28] = (1-Gamma) * k!02 * y[16];dfdy[29][29] = 0.0;/*********#*****#************************/// Calculation <strong>of</strong> Differential Steps. Kinetics Routine139


void derivs(double x, double y[], double dydx[j){//printf ("\n derivs");nrhs-H-;/*Rate Equations// Ozone decompositiondouble Rl = kl *y[l] *HYD;// kl[O3][OH-]double R3 = k3 * y[l] * al_HO2 * y[4];//k3[O3][O2-]double R5 = k5 * aO_HO3 *y[2];/7 k3[HO3]double R6 = k6 * y[l] * y[6];// k6[HO][O3]double R7 = k7 * y[6] * y[6];// k7[HO][HO]double R8 = k8 * y[6] * al_HO2 * y[4];//k8[HO][O2-]double R9 = k9 * y[6] * aO_HO3 * y[2];//k9[HO][H03]double RIO = klO * al_HO2 * y[4] * aO_HO3 *y[2];// klO[02-][H03]double Rl 1 = kl 1 * aO_HO3 * y[2] * aO_HO3 *kll[H03][H03]// <strong>Bromate</strong> <strong>Formation</strong>double R20 = k20 * y[l] * y[lI];// k20[O3][Br-]double R21 = k21 * y[l] * al_HOBr * y[16];//k21[03][OBr-]double R22 = k22 * y[l] * al_HOBr * y[16];//k22[O3][OBr-]double R23 = k23 * y[l] * y[19];// k23[O3][BrO2-]double R24a = k24a * y[6] * y[l I];// k24a[OH][Br-]double R24b = k24b * y[13];// k24b[BrOH-]double R25a = k25a * y[13];// k25a[BrOH-jdouble R25b = k25b * y[ 12] * HYD;//k25b[Br][OH-]double R26 = k26 * y[13] * PRT;// k26[BrOH-][H+]double R27 = k27 * y[13] * y[l 1]; // k27[BrOHdoubleR28a = k28a * y[12] * y[l I];// k28a[Br][Br-]double R28b = Jc28b * y[14];// k28b[Br2-]double R29 = k29 * y[14] * y[14];// k29[Br2-][Br2-]double R30a = k30a * y[15];// k30a[Br3-]double R30b = k30b * y[17] * y[lI];//k30b[Br2][Br-]double R31a = k31a * y[17];// k31a[Br2][H2O]double R31b = k31b * aO_HOBr * y[16] * PRT *y[ll];// k31b[HOBr][Br-][H+]double R33 =k33 * y[l] * y[12];// k33[O3][Br]double R34 = k34 * y[12] * al_HOBr * y[16];//k34[Br][OBr-]double R35 = k35 * y[6] * al_HOBr * y[16];//k35[OH][OBr-]double R36 = k36 * y[6] * aOJHOBr * y[16]; //k36[OH][HOBr]double R37 = k37 * al_HO2 * y[4] * aO_HOBr *y[16];// k37[O2-][HOBr]double R38 = k38 * y[18] * y[18];//k38[BrO][BrO][H20]double R39 = k39 * y[18] * y[19];//k39[BrO][BrO2-]double R40 = k40 * y[14] * y[19];// k40[Br2-][Br02-]double R41 = k41 * y[6] * y[19];// k41[OH][BrO2-]double R42 = k42 * y[6] * y[20];// k42[OH][BrO2]double R43a = k43a * y[20] * y[20];//k43a[Br02][Br02]double R43b = k43b * y[21];// k43b[Br2O4]double R44 = k44 * y[21] * HYD;//k44[Br2O4][OH-]// Carbonatedouble R53 = k53 * al_CO3 * y[8] * y[6];//k53[HCO3-][OH]double R54 = k54 * a2_CO3 * y[8] * y[6];//k54[C032-][OH]double R55 = k55 * al_CO3r * y[9] * al_HOBr *y[16];// k55[CO3-][OBr-]double R56 = k56 * al_CO3r * y[9] * y[19]; //k56[CO3-][BrO2-]// Phosphatedouble R63 = k63 * y[6] * al_PO4 * y[10];//k65[OH][H2PO4-]double R64 = k64 * y[6] * a2_PO4 * y[10];//k66[OH][HPO42-]// Hydrogen Peroxidedouble R71 = k71 * y[l] * al_Peroxide * y[7];//k71[O3][HO2-]double R72 = k72 * aO_Peroxide * y[7] * y[6];//k72[H2O2][OH]double R73 = k73 * al_Peroxide * y[7] * y[6];//k73[HO2-][OH]double R74 = k74 * al_Peroxide * y[7] * aO_HOBr *];// k74[HO2-][HOBr]// Ammoniadouble R81 = k81 * aO_HOBr * y[16] * al_NH3 *y[24];// k91[HOBr][NH3]double R82 = k82 * y[l] * y[25];// k92[NH2Br][O3]// Radical Scavengerdouble R90 = k90 * y[23] * y[6];// k95[t-BuOH][OH]140


Natural Organic Matterdouble R100 = klOO * y[26] * y[l];//klOO[NOMl][O3]double R101 = klOl * y[27] * y[6];//k!01[NOM2][OH]double R102 = k!02 * y[28] * y[16];//k!02[NOM3][HOBr]tot/*Differential Equationsdydx[l] = - Rl - R3 - R6 - R20 - R21 -R22 - R23 -R33-R71-3*R82-R100;dydx[2] = R3 - R5 - R9 - RIO - 2*R1 1 + Alpha *R100;dydx[3] = 0.0;dydx[4] = 2*R1 - R3 + R6 - R8 - RIO - R37 + R71 +R72 + R73 + Beta*R101;dydx[5] = 0.0;dydx[6] = R5 - R6 - 2*R7 - R8 - R9 - R24a + R24b -R35 - R36 - R41 - R42 - R53 - R54 - R63 - R64 +R71 - R72 - R73 - R90 - R101;dydx[7] = R7 + R9 + Rl 1 - R71 - R72 - R73 - R74;dydx[8] = - R53 - R54 + R55 + R56;dydx[9] = R53 + R54 - R55 - R56;dydx[10] = -R63-R64;dydx[l 1] = - R20 + R21 - R24a + R24b - R27 - R28a+ R28b + R29 + R30a - R30b + R31a - R3 Ib + R34+ R40 + R74 + R82 + Gamma * R102;dydx[12] = R25a - R25b + R26 - R28a + R28b - R33- R34 + R37;dydx[13] = R24a - R24b - R25a + R25b - R26 - R27;dydx[14] = R27 + R28a - R28b - 2*R29 - R40;dydx[15] = R29 - R30a + R30b;dydx[16] = R20 - R21 - R22 + R31a - R31b - R34 -R35 - R36 - R37 + R38 + R39 + R40 - R55 - R74 -R81-R102;dydx[17] = R30a - R30b - R31a + R31b;dydx[18] = R33 + R34 + R35 + R36 - 2*R38 - R39 +R40 + R55;dydx[19] = R22 - R23 + R38 - R39 - R40 - R41 +R44 - R56;dydx[20] = R39 + R41 - R42 - 2*R43a + 2*R43b+R56;dydx[21] = R43a - R43b - R44;dydx[22] = R23 + R42 + R44;dydx[23] = - R90;dydx[24] = -R81 ;dydx[25]=R81-R82;dydx[26] = - R100;dydx[27] = -R101;dydx[28] = -R102;dydx[29] = (1-Gamma) *.R102;3. Solver routines for bromate formation <strong>and</strong> Cryptosporidiwn parvum inactivation in AxialDispersion Model. These routines are utilized by main function that should be programmedfor the specific purposes. Main functions for different modeling were not included_____/****************************************/// pinvs(ic3,ic4,j5,j9Jcl,kl,c,s) Called from solvde/#****#**********#******************#****/void pinvs(int iel, int ie2, int jel, int jsf, int jcl, int k,double ***c, double **s){intjsljpivjpje2jc<strong>of</strong>f,j,irow,ipiv,id,ic<strong>of</strong>f,i,*indxr;double pivinv,piv,dum,big,*pscl;indxr=ivector(ie 1 ,ie2);pscl=vector(ie 1 ,ie2);je2=jel+ie2-iel;jsl=je2+l;for(i=iel;i


indxr[ipiv]=;jpiv;pivinv=l .0/s[ipiv][jpiv];for (j=Jel;j


inticl=l;int ic2=ne-nb;intic3=ic2+l;int ic4=ne;intjcl=l;intjcf=ic3;for(it=l;it


elseif(k>k2) {if (current=l) {for(i=l;i


double R24a = k24a * 0.5*(y[6][k]+y[6][k-l]) *0.5*(y[ll][k]+y[ll][k-l]);// k24a[OH][Br-]double R24b = k24b * 0.5*(y[13][k]+y[13][k-l]);//k24b[BrOH-]double R25a = k25a * 0.5*(y[13][k]+y[13][k-l]);//k25a[BrOH-Jdouble R25b = k25b * 0.5*(y[12][k]+y[12][k-l]) *HYD;// k25b[Br][OH-]double R26 = k26 * 0.5*(y[13][k]+y[13][k-l]) *PRT;// k26[BrOH-][H+]double R27 = k27 * 0.5*(y[13][k]+y[13][k-l]) *0.5*(y[ll][k]+y[ll][k-l]); // k27[BrOH-][Br-]double R28a = k28a * 0.5*(y[12][k]+y[12][k-l]) *0.5*(y[l l][k]+y[l l][k-l]);// k28a[Br][Br-]double R28b = k28b * 0.5*(y[14][k]+y[14][k-l]);//k28b[Br2-]double R29 = k29 * 0.5*(y[14][k]+y[14][k-l]) *0.5*(y[14][k]+y[14][k-l]);// k29[Br2-][Br2-]double R30a = k30a * 0.5*(y[15][k]+y[15][k-l]);//k30a[Br3-]double R30b = k30b * 0.5*(y[17][k]+y[17][k-l]) *0.5*(y[l l][k]+y[l l][k-l]);// k30b[Br2][Br-]double R3la = k31a * 0.5*(y[17][k]+y[17][k-l]);//k31a[Br2][H20]double R31b = k31b * aO_HOBr *0.5*(y[16][k]+y[16][k-l]) * PRT *0.5*(y[ll][k]+y[ll][k-l]);// k31b[HOBr][Br-][H+]double R33 =k33 * 0.5*(y[l][k]+y[l][k-l]) *0.5*(y[12][k]+y[12][k-l]);//k33[03][Br]double R34 = k34 * 0.5*(y[12][k]+y[12][k-l]) *al_HOBr * 0.5*(y[16][k]+y[16][k-l]);//k34[Br][OBr-]double R35 =k35 * 0.5*(y[6][k]+y[6][k-l]) *al_HOBr * 0.5*(y[16][k]+y[16][k-l]);//k35[OH][OBr-]double R36 = k36 * 0.5*(y[6][k]+y[6][k-l]) *aO_HOBr * 0.5*(y[16][k]+y[16][k-l]); //k36[OH][HOBr]double R37 = k37 * al_HO2 * 0.5*(y[4][k]+y[4][k-1]) * aO_HOBr * 0.5*(y[16][k]+y[16][k-l]);//k37[02-][HOBr]double R38 = k38 * 0.5*(y[18][k]+y[18][k-l]) *0.5*(y[18][k]+y[18][k-l]);// k38[BrO][BrO][H2O]double R39 =k39 * 0.5*(y[18][k]+y[18][k-l]) *0.5*(y[19][k]+y[19][k-l]);// k39[BrO][BrO2-]double R40 =k40 * 0.5*(y[14][k]+y[14][k-l]) *0-5*(y[19][k]+y[19][k-l]);// k40[Br2-][BrO2-]double R41 =k41 * 0.5*(y[6][k]+y[6][k-l]) *0.5*(y[19][k]+y[19][k-l]);// k41[OH][BrO2-]double R42 = k42 * 0.5*(y[6][k]+y[6][k-l]) *0.5*(y[20][k]+y[20][k-l]);// k42[OH][BrO2]double R43a = k43a * 0.5*(y[20][k]+y[20][k-l]) *0.5*(y[20][k]+y[20][k-l]);// k43a[BrO2][BrO2]double R43b = k43b * 0.5*(y[21][k]+y[21][k-l]);//k43b[Br2O4]double R44 = k44 * 0.5*(y[21][k]+y[21][k-l]) *HYD;// k44[Br2O4][OH-]// Carbonatedouble R53 =k53 * al_CO3 * 0.5*(y[8][k]+y[8][k-1]) * 0.5*(y[6][k]+y[6][k-l]);// k53[HCO3-][OH]double R54 = k54 * a2_CO3 * 0.5*(y[8][k]+y[8][k-1]) * 0.5*(y[6][k]+y[6][k-l]);// k54[CO32-][OH]double R55 = k55 * al_CO3r * 0.5*(y[9][k]+y[9][k-1]) * al_HOBr * 0.5*(y[16][k]+y[16][k-l]);//k55[C03-][OBr-]double R56 = k56 * al_CO3r * 0.5*(y[9][k]+y[9][k-1]) * 0.5*(y[19][k]+y[19][k-l]); // k56[CO3-][Br02-]// Phosphatedouble R63 = k63 * 0.5*(y[6][k]+y[6][k-l]) *al_PO4 * 0.5*(y[10][k]+y[10][k-l]);//k65[OH][H2PO4-]double R64 = k64 * 0.5*(y[6][k]+y[6][k-l]) *a2_PO4 * 0.5*(y[10][k]+y[10][k-l]);//k66[OH][HPO42-]// Hydrogen Peroxidedouble R71 = k71 * 0.5*(y[l][k]+y[l][k-l]) *al_Peroxide * 0.5*(y[7][k]+y[7][k-l]);//k71[03][H02-]double R72 = k72 * aO_Peroxide *0.5*(y[7][k]+y[7][k-l]) * 0.5*(y[6][k]+y[6][k-l]);//k72[H202][OH]double R73 = k73 * al_Peroxide *0.5*(y[7][k]+y[7][k-l]) * 0.5*(y[6][k]+y[6][k-l]);//k73[HO2-][OH]double R74 = k74 * al_Peroxide *0.5*(y[7][k]+y[7][k-l]) * aO_HOBr *]);// k74[HO2-][HOBr]// Ammoniadouble R81 = k81 * aO_HOBr *0.5*(y[16][k]+y[16][k-l]) * al_NH3 *0.5*(y[24][k]+y[24][k-l]);// k91[HOBr][NH3]double R82 = k82 * 0.5*(y[l][k]+y[l][k-l]) *0.5*(y[25][k]+y[25][k-l]);// k92[NH2Br][O3]// Radical Scavengerdouble R90 = k90 * 0.5*(y[23][k]+y[23][k-l]) *0.5*(y[6][k]+y[6][k-l]);// k95[t-BuOH][OH]// Natural Organic Matterdouble R100 = klOO * 0.5*(y[26][k]+y[26][k-l]) *0.5*(y[l][k]+y[l][k-l]);// klOO[NOMl][O3]double R101 =k!01 * 0.5*(y[27][k]+y[27][k-l]) *0.5*(y[6][k]+y[6][k-l]);// k!01[NOM2][OH]double R102 = k!02 * 0.5*(y[28][k]+y[28][k-l]) *]);// k!02[NOM3][HOBr]tot145


C. parvum Inactivationdouble RI1 =kNl * 0.5*(y[l][k]+y[l][k-l]) *0.5*(y[3][k]+y[3][k-l]);double RI2 = kN2 * 0.5*(y[l][k]+y[l][k-l]) *0.5*(y[5][k]+y[5][k-l]);// Mass Transferdouble Rka = ka * 0.5*(y[30][k]+y[30][k-l]) / m - ka- 0.5*(y[l][k]+y[l][k-l]);for(n=l;n


REFERENCESAiken, G., D. McKnight, K. Thom, <strong>and</strong> E. Thurman. (1992). "Isolation <strong>of</strong> hydrophilic organicacids from water using nonionic macroporous resins," Organic Geochemistry, 18:567-573.APHA, AWWA, WEF. (1998). St<strong>and</strong>ard Methods for the Examination <strong>of</strong> Water <strong>and</strong>Wastewater. 20th Edition. Washington, D.C.Amy, G., P.Westerh<strong>of</strong>f, R.Minear, <strong>and</strong> R.Song (1997). <strong>Formation</strong> <strong>and</strong> <strong>Control</strong> <strong>of</strong> BrominatedOzone By-Products. AwwaRF, Denver ,CO.Amy, G. <strong>and</strong> M. Siddiqui. (1999). Strategies to control bromate formation. AwwaRF, Denver,CO.Amy, G., M. Siddiqui, W. Zhai, <strong>and</strong> J. Debroux. (1993). National survey <strong>of</strong> bromide in drinkingwaters. Proceedings AWWA Annual Conference. Denver, CO.Bezbarua, B. K. <strong>and</strong> Reckhow, D. A. (1996). "Modeling Ozone Consumption by NaturalOrganic Matter." Proceedings Water Quality <strong>and</strong> Technology Conference, American WaterWorks Association, Boston, Nov 17-21.Buhler, R. E., Staehelin, J., <strong>and</strong> Hoigne, J. (1984). "Ozone Decomposition in Water Studied byPulse Radiolysis 1. HO2/O2~ <strong>and</strong> HOs/Os" as Intermediates." Journal <strong>of</strong> Physical Chemistry,88(12), 2560-2564.Bull, R. J., L. S., Birnbaum, K. P., Cantor, J. B. Rose, B. E. Butterworth, R., Pergram, <strong>and</strong> J.Tuomisto (1995). "Water Chlorination: Essential Process or Cancer Hazard?" Fundam. Appl.Toxicol. 28:155-166.Chin, Y., G. Aiken <strong>and</strong> E. O'Loughlin (1994). Molecular weight polydispersivity, <strong>and</strong>spectroscopic properties <strong>of</strong> aquatic humic substances. Environmental Science <strong>and</strong> Technology,28(11), pp. 1853.Chiou, C-F, Marinas B.J., <strong>and</strong> Adams, J.Q. (1995) Modified indigo method for gaseous <strong>and</strong>aqueous ozone analysis. Ozone Science <strong>and</strong> Engineering. 17:329-344.Collins, H. <strong>and</strong> Selleck, R. (1972). "Process kinetics <strong>of</strong> wastewater chlorination." SERL ReportNo 72-5, University <strong>of</strong> California, Berkeley.Croue, J.P., Koudkonou, B.K. <strong>and</strong> Legube, B. (1996). Parameters Affecting the <strong>Formation</strong> <strong>of</strong><strong>Bromate</strong> Ion <strong>During</strong> <strong>Ozonation</strong>. Ozone Science & Engineering., 18:1-18.Echigo, S., Yamada, H., Minear, R. A. (2000). Ultra-low bromate detection in drinking water, InS. E. Barrett (ed.) Natural Organic Mater <strong>and</strong> Disinfection By-Products, ACS Books. IN PRESS.Elovitz, M. <strong>and</strong> U. von Gunten. (1999a). "Hydroxyl radical/ozone ratios during ozonationprocesses. I. The Rct concept," Ozone Science <strong>and</strong> Engineering, 21:239-260.147


Elovitz, M., U. von Gunten, <strong>and</strong> H. Kaiser. (1999b). "Hydroxyl radical/ozone ratios duringozonation processes, n. The effect <strong>of</strong> temperature, pH, alkalinity, <strong>and</strong> DOM properties,"submitted to Ozone Science <strong>and</strong> Engineering, 1999.Finch, G. R., Black, E. K., Gytirek, L., <strong>and</strong> Belosevic, M. (1993). "Ozone inactivation <strong>of</strong> C.parvum in dem<strong>and</strong>-free phosphate buffer determined by in vitro excystation <strong>and</strong> animalinfectivity." Journal <strong>of</strong> Applied Environmental Microbiology, 59(12), 4203-4210.Finch, G. <strong>and</strong> H. Li. (1999). "Inactivation <strong>of</strong> Cryptosporidium at 1°C using ozone or chlorinedioxide," Ozone Science <strong>and</strong> Engineering, 21:477-486.Gelinet, K. (1999). Importance des Caracteristiques Physico-Chimiques des Eaux Naturelles surla <strong>Formation</strong> des Ions <strong>Bromate</strong> Lors de L'<strong>Ozonation</strong>. Ph.D. Dissertation. University <strong>of</strong> Poitiers,France.Glaze et al. 1993 Glaze, W.H., H.S. Weinberg <strong>and</strong> J.E. Cavanagh. (1993). "Evaluating the<strong>Formation</strong> <strong>of</strong> Brominated DBPs <strong>During</strong> <strong>Ozonation</strong>," Journal American Water Works Association85:1:96-103.Haag, W. R., <strong>and</strong> Hoigne, J. (1983). "<strong>Ozonation</strong> <strong>of</strong> <strong>Bromide</strong>-Containing Waters : Kinetics <strong>of</strong><strong>Formation</strong> <strong>of</strong> Hypobroumous Acid <strong>and</strong> <strong>Bromate</strong>." Environmental Science <strong>and</strong> Engineering,17(5), 261-267.Hagg, W. R., Hoigne J., <strong>and</strong> Bader, H. (1984). "Improved Ammonia Oxidation by Ozone In thePresence <strong>of</strong> <strong>Bromide</strong> Ion <strong>During</strong> Water Treatment." Water Research, 18(9), 1125-1128.Hass, C. N. (1980). "A mechanistic kinetic model for chlorine disinfection." EnvironmentalScience & Technology, 14(3) 339-340.Hull, C.S. (1995). Modeling <strong>of</strong> Ozone Contactors. Ph.D. Dissertation, University <strong>of</strong> NorthCarolina at Chapel Hill.H<strong>of</strong>mann, R. <strong>and</strong> Andrews, R.C. (2000). A <strong>Bromate</strong> <strong>Control</strong> Strategy Using Ammonia <strong>During</strong><strong>Ozonation</strong>. Proc. 9fft National Drinking Water Conference, Regina, Canada.Hunt N. K. <strong>and</strong> Marinas B. J. (1997). "Kinetics <strong>of</strong> Eschericia coli Inactivation with Ozone."Water Research, 21(6), 1355-1362.Korich, D. G., Mead, J. R., Madore, M. S., Sinclair, N. A., <strong>and</strong> Sterling, C. R. (1990). "Effect <strong>of</strong>ozone, chlorine dioxide, chlorine, <strong>and</strong> monochloramine on Cryptosporidium parvum in raw <strong>and</strong>finished water, Journal <strong>of</strong> American Water Works Association. 87(9), 54-68.Krasner, S., W. Glaze, H. Weinberg, P. Daniel, <strong>and</strong> I. Najm. (1993). "<strong>Formation</strong> <strong>and</strong> control <strong>of</strong>bromate during ozonation <strong>of</strong> waters containing bromide," Journal <strong>of</strong> American Water WorksAssociation, 85:1:73-81.Legube, Bernard (1996). "A Survey <strong>of</strong> <strong>Bromate</strong> Ion in European Drinking Water," OzoneScience <strong>and</strong> Engineering 18:325-348.148


Liyanage, L. R., Finch, G. R., <strong>and</strong> Belosevic, M. (1997). "Synergistic effect <strong>of</strong> sequentialexposure <strong>of</strong> Cryptosporidium oocysts to chemical disinfectants." 7997 International Symposiumon Waterborne Cryptosporidium Proceedings, March, 1997, Newport Beach, CA.Marinas B. J., Liang, S., <strong>and</strong> Aieta, M. E. (1993). "Modeling hydrodynamics <strong>and</strong> ozone residualdistribution in a pilot-scale ozone bubble-diffuser contactor." Journal <strong>of</strong> American Water WorksAssociation, 85(3), 90-99.Oppenheimer, J.A., E.M. Aieta, R.R. Trussell, J.G. Jacangelo, I.N. Najm (2000). Evaluation <strong>of</strong>Cryptosporidium Inactivation in Natural Waters. AwwaRF, Denver, CO.Owens, J. H., Miltner, R. J., Schaefer ffl, F. W., <strong>and</strong> Rice, E. W. (1994). "Pilot-scale inactivation<strong>of</strong> Cryptosporidium <strong>and</strong> Giardia." Proceedings AWWA Water Quality Technology Conference,November, 1994, San Francisco, CA.Ozekin, K. (1994). "<strong>Bromate</strong> <strong>Formation</strong>." Ph.D. Dissertation, University <strong>of</strong> Colorado, Boulder,CO.Ozekin, K., P. Westerh<strong>of</strong>f, <strong>and</strong> G. Amy. (1998). "Predicting bromate formation duringCryptosporidium inactivation," In Proc. AWWA Water Quality Technology Conference, SanDiego, CA.Ozekin, K., P. Westerh<strong>of</strong>f, G. Amy, <strong>and</strong> M. Siddiqui. (1998). "Molecular ozone <strong>and</strong> radicalpathways <strong>of</strong> bromate formation during ozonation," Journal <strong>of</strong> Environmental Engineering,124:5:456-462.Peelers, J. E., Mazas, E. A., Masschelein, W. J., Villacorta Martinez de Maturana, I., <strong>and</strong>Debacker, E. (1989). "Effect <strong>of</strong> disinfection <strong>of</strong> drinking water with ozone <strong>and</strong> chlorine dioxideon survival <strong>of</strong> Cryptosporidium parvum oocysts." Journal <strong>of</strong> Applied <strong>and</strong> EnvironmentalMicrobiology, 55(6) 1519-1522.Perrine, D., Georges, P., <strong>and</strong> Langlais, B. (1990). "Efficacite de 1'ozonation des eaux sur1'inactivation des oocysts de Cryptosporidium." Bull. Acad. Natle, Med., 174(6), 845-851.Press W. H., Teukolsky S. A., Vetterling W. T., <strong>and</strong> Flannery B. P. (1992) Numerical Recipes inC, Cambridge University Press, 2nd ed.Rennecker, J., B. Marinas, J. Owens, <strong>and</strong> E. Rice. (1999). "Inactivation <strong>of</strong> Cryptosporidiumparvum oocysts with ozone," Water Research, 33:11:2481-2488.Severin, B. F., Suidan, M. T., <strong>and</strong> Engelbrecht, R. S. (1984). "Series-event kinetic model forchemical disinfection." Journal <strong>of</strong> Environmental Engineering, 110(2), 430-439.Siddiqui, M, G. Amy, <strong>and</strong> R. Rice. (1995). "<strong>Bromate</strong> ion formation: a critical review," Journal<strong>of</strong> American Water Works Association, 87:10:58-70.Siddiqui, M. <strong>and</strong> G. Amy. (1993). "Factors affecting DBP formation during ozone-bromidereactions," Journal <strong>of</strong> American Water Works Association, 85:1:63-72.149


Singer, P. (1990). "Assessing ozonation research needs in water treatment," Journal <strong>of</strong>American Water Works Association, 85:10:78-88.Song, R., Amy, G.L., Westerh<strong>of</strong>f, P., <strong>and</strong> Minear, R. (1997). <strong>Bromate</strong> Minimization <strong>During</strong><strong>Ozonation</strong>. Journal <strong>of</strong> American Water Works Association, 89(6):69-78.Song, R., C. Donohoe, R. Minear, P. Westerh<strong>of</strong>f, K. Ozekin, <strong>and</strong> G. Amy. (1996). "Empriricalmodeling <strong>of</strong> bromate formation during ozonation <strong>of</strong> bromide-containing waters," WaterResearch, 30:5:1161-1168.Song R. (1996) Ozone-<strong>Bromide</strong>-NOM Interactions in Water Treatment, Ph.D. Dissertation,University <strong>of</strong> Illinois at Urbana-Champaign.Staehelin, J., Biihler, R. E., <strong>and</strong> Hoigne, J. (1984). "Ozone Decomposition in Water Studied byPulse Radiolysis. 2. OH <strong>and</strong> HO4 as Chain Intermediates." Journal <strong>of</strong> Physical Chemistry,88(24), 5999-6004.Staehelin, J., <strong>and</strong> Hoigne, J. (1982). "Decomposition <strong>of</strong> Ozone in Water : Rate <strong>of</strong> Initiation byHydroxide Ions <strong>and</strong> Hydrogen Peroxide." Environmental Science <strong>and</strong> Technology, 16(10), 676-681.Staehelin, J., <strong>and</strong> Hoigne, J. (1985). "Decomposition <strong>of</strong> Ozone in Water in the Presence <strong>of</strong>Organic Solutes Acting as Promoters <strong>and</strong> Inhibitors <strong>of</strong> Radical Chain Reactions." EnvironmentalScience <strong>and</strong> Engineering, 19(12), 1206-1213.Thurman, E.M. (1985). Organic geochemistry <strong>of</strong> natural waters, Martinus Nijh<strong>of</strong>f <strong>and</strong> Dr W.Junk Publishers, Boston, MA.U.S. Environmental Protection Agency (1989) "National Primary Drinking Water Regulations;Disinfection; Turbidity, Giardia lambia, Viruses, Legionella, <strong>and</strong> Heterotrophic Bacteria; FinalRule," Federal Register, 54: 27486-27541.U.S. Environmental Protection Agency (1990) The Guidance Manual for Compliance with theFiltration <strong>and</strong> Disinfection Requirements for Public Water Supplies, Using Surface WaterSources, Appendix O.von Gunten, U., <strong>and</strong> J. Hoigne. (1993). "<strong>Bromate</strong> <strong>Formation</strong> during <strong>Ozonation</strong> <strong>of</strong> <strong>Bromide</strong>-Containing Waters." International Water Supply Association International Conference,November 22-24, Paris, pp. 51-56.von Gunten, U., <strong>and</strong> J. Hoigne. (1994). "<strong>Bromate</strong> formation during ozonation <strong>of</strong> bromidecontainingwaters: interaction <strong>of</strong> ozone <strong>and</strong> hydroxyl radical reactions," Environmental Science& Technology, 28:7:1234-1242.von Gunten, U., <strong>and</strong> J. Hoigne. (1996). "<strong>Bromate</strong> <strong>Formation</strong> in advanced oxidation processes:role <strong>of</strong> hydrogen peroxide." 210th American Chemical Society National Meeting, Division <strong>of</strong>Environmental Chemistry, Chicago, IL, August 20-24, Vol. 35(2), pp. 661-663.150


von Gunten, U., <strong>and</strong> Oliveras, Y. (1997). "Kinetics <strong>of</strong> the Reaction Between Hydrogen Peroxide<strong>and</strong> Hypobromous Acid: Implication on Water Treatment <strong>and</strong> Natural Waters." Water Research,31(4), 900-906.von Gunten, U., <strong>and</strong> Oliveras, Y. (1998). "Advanced Oxidation <strong>of</strong> <strong>Bromide</strong>-Containing Waters:<strong>Bromate</strong> <strong>Formation</strong> Mechanisms." Environmental Science <strong>and</strong> Technology, 32(1), 63-70.Weinberg, H.S., Yamada, H. (1998). Post-Ion-Chromatography Derivatization for theDetermination <strong>of</strong> Oxyhalides at Sub-PPB Levels in Drinking Water. Analytical Chemistry. 70(1),1-6.Westerh<strong>of</strong>f, P. (1995). "Ozone oxidation <strong>of</strong> bromide <strong>and</strong> natural organic matter." PhDDissertation, University <strong>of</strong> Colorado, Boulder, CO.Westerh<strong>of</strong>f, P., R. Song, G. Amy <strong>and</strong> R. Minear. (1998). "NOM's role in bromine <strong>and</strong> bromateformation during ozonation," Journal <strong>of</strong> American Water Works Association, 89:11:82-94.Westerh<strong>of</strong>f, P., G. Aiken, G. Amy <strong>and</strong> J. Debroux. (1999). "Relationships between the structure<strong>of</strong> natural organic matter <strong>and</strong> its reactivity towards molecular ozone <strong>and</strong> hydroxyl radicals,"Water Research, 33:10:2265-2276.Westerh<strong>of</strong>f, P., Song, R., Amy, G., <strong>and</strong> Minear, R. (1997). "Application <strong>of</strong> OzoneDecomposition Models." Ozone Science <strong>and</strong> Engineering, 19(1), 55-74.Westerh<strong>of</strong>f, P., Song, R., Amy, G., <strong>and</strong> Minear, R. (1998). "Numerical Kinetic Models for<strong>Bromide</strong> Oxidation to Bromine <strong>and</strong> <strong>Bromate</strong>." Water Research, 32(5), 1687-1699.Woodmansee, D. B. (1987) "Studies <strong>of</strong> In Vitro Excystation <strong>of</strong> Cryptosporidium parvum fromCalves," Journal <strong>of</strong> Protozoology, 34, pp. 398-402.Yates, R. S., <strong>and</strong> Stenstrom, M. K. (1993). "<strong>Bromate</strong> Production in Ozone Contactors."Proceedings AWWA Annual Conference, San Antonio, TX, 475-499.151


ABBREVIATIONSACDADRAMAAMWANNAPHAAWWAAwwaRFBAG°CCCDCGEC,cmcm" 1CRWCSTRCTALL-tlOCT ALL-t50CTswTR-tlOCTsWTR-t50CUWACVdD/DBPDALDBFDO 3DOC^initialAUVA200-400AUVA254South Bay Aqueduct, Fremont, Californiaaxial dispersion reactorLake Meredith <strong>and</strong> Ogallala Aquifer mix, Amarillo, Texasapparent molecular weightHuron River, Ann Arbor, MichiganAmerican Public Health AssociationAmerican Water Works AssociationAwwa Research Foundationbiological activated carbondegrees CelsiusSacramento/San Joaquin Delta, Concord, CaliforniaMarne River, Paris, Franceconcentration <strong>of</strong> chemical speciescentimeterinverse centimeterColorado River, La Verne, Californiacontinuous-stirred tank reactoran "all cells" approach where CT credit for the first cell (effluent ozoneresidual multiplied by the first cell tio) was added to the CTswTR-tiocalculated by summing the tso multiplied by the effluent ozone residual foreach cell in the contactorCT required to overcome the initial time lagconventional CT using tio contact time times <strong>and</strong> effluent ozone residualfor each cell <strong>of</strong> the ozone contactor excluding the first cellusing tso (HRT) instead <strong>of</strong> tio for the corresponding contact time for eachcell excluding the first, multiplied by the corresponding cell's effluentozone residualCalifornia Urban Water Agenciescoefficient <strong>of</strong> variationdispersion numberDisinfectants <strong>and</strong> Disinfection By-ProductsTrinity River - Elm Fork, Dallas, Texasdisinfection by-productdissolved ozone residualdissolved organic carbonrepresents the amount <strong>of</strong> ozone loss during this phase <strong>of</strong> the reaction <strong>and</strong>reflects a "pseudo" zero-order reactiondifferential ultraviolet absorbance spectra between 200 <strong>and</strong> 400nanometersdifferential ultraviolet absorbance at 254 nanometers153


ECEUFDfiFPGACHBrHOBrHOUHPIHPLCHPOHRTIOAkKkJ/molkNEuropean CommunityEuropean Unionfraction <strong>of</strong> dissolved organic carbonoverall rate function corresponding to the relevant chemical reactionsformation potentialgranular activated carbonhydrogen bromidehypobromous acidLake Houston, Houston, Texashydrophilichigh performance liquid chromatographyhydrophobichydraulic residence timeInternational Ozone Associationexperimentally derived first-order rate constant for ozone decayslope coefficientkilojoule per molepseudo first-order inactivation rate constantL/mg-mL/minLAWLT2ESWTRmMM/DBPMCLMCLGmg NH3-N/Lmg-min/Lmg/Lmg/L/minmg-C/Lmg-CaCO3/Lmg-DOC/Lmg-Os/LminmLmL/minliter per milligram per meterliter per minuteLos Angeles Aqueduct, Los Angeles, CaliforniaLong-Term 2 Enhanced Surface Water Treatment RulemetermolarMicrobial/Disinfection By-Productmaximum contaminant levelmaximum contaminant level goalmilligram ammonia as nitrogen per litermilligram minute per litermilligram per litermilligram per liter per minutemilligram <strong>of</strong> carbon per litermilligram calcium carbonate per litermilligram dissolved organic carbon per litermilligram ozone per literminutemillilitermilliliter per minute154


mMMRLMWu,mug/LN/NONJAnmNOM•OHOEOTTPACPFRPQLpsiPSSQAQCSACSECsec" 1SPWSUVASWTRttiotsoTOBrTOCTPI6UCUIUPUSEPAUTmillimolarminimum reporting levelmolecular weightmicrometermicrogram per litersurvival ratiolag phase factorDelaware River, New Jerseynanometernatural organic matterhydroxyl radicalozone exposureOttawa River, Ottawa, CanadaProject Advisory Committeeplug-flow reactorpractical quantification limitpounds per square inchpolystyrene sulfonatequality assurancequality controlSacramento River, Californiasize exclusion chromatographyState Project, La Verne, Californiaspecific ultraviolet absorbancesurface water treatment ruletimetime for 10 percent <strong>of</strong> the total mass <strong>of</strong> tracer to exit the reactortime for 50 percent <strong>of</strong> the total mass <strong>of</strong> tracer to exit the reactortotal organic bromidetotal organic carbontransphilichydraulic residence time in the chamberUniversity <strong>of</strong> ColoradoUniversity <strong>of</strong> IllinoisUniversity <strong>of</strong> PoitiersUnited States Environmental Protection AgencyUniversity <strong>of</strong> Toronto155


UV ultravioletUV200-300 ultraviolet absrobance between 200 <strong>and</strong> 300 nanometersUV254 ultraviolet absorbance at 254 nanonmetersUVA ultraviolet absorbanceWEF Water Environment FederationWPB Floridian Aquifer, West Palm Beach, FloridaWTP water treatment plantz normalized downward distance from the water surface156


AWWAResearchFoundationkm Advancing the Sderm <strong>of</strong> Water*6666 W. Quincy Avenue, Denver, CO 80235(303)347-61001P-4C-90866-10/01-CM

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!