You are on page 1of 114

U4R11_WEF-IWAPspread.

qxd 10/20/2014 11:25 AM Page 1

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems
Wastewater Treatment Systems

Water Environment Research Foundation


635 Slaters Lane, Suite G-110 n Alexandria, VA 22314-1177
Phone: 571-384-2100 n Fax: 703-299-0742 n Email: werf@werf.org
www.werf.org
WERF Stock No. U4R11

Co-published by

IWA Publishing
Alliance House, 12 Caxton Street Mass Transfer Characteristics of Floating Media
London SW1H 0QS
United Kingdom
Phone: +44 (0)20 7654 5500
Fax: +44 (0)20 7654 5555
in MBBR and IFAS Fixed-Film Systems
Email: publications@iwap.co.uk
Web: www.iwapublishing.com
IWAP ISBN: 978-1-78040-705-0/1-78040-705-X

Co-published by

October 2014

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
on 12 June 2020
U4R11

MASS TRANSFER CHARACTERISTICS OF


FLOATING MEDIA IN MBBR AND IFAS
FIXED-FILM SYSTEMS

by:
Henryk Melcer, Ph.D., P.E.
Brown and Caldwell
Andrew J. Schuler, Ph.D., P.E.
University of New Mexico

2014

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
The Water Environment Research Foundation, a not-for-profit organization, funds and manages water quality
research for its subscribers through a diverse public-private partnership between municipal utilities, corporations,
academia, industry, and the federal government. WERF subscribers include municipal and regional water and
water resource recovery facilities, industrial corporations, environmental engineering firms, and others that share
a commitment to cost-effective water quality solutions. WERF is dedicated to advancing science and technology
addressing water quality issues as they impact water resources, the atmosphere, the lands, and quality of life.

For more information, contact:


Water Environment Research Foundation
635 Slaters Lane, Suite G-110
Alexandria, VA 22314-1177
Tel: (571) 384-2100
Fax: (703) 299-0742
www.werf.org
werf@werf.org

This report was co-published by the following organization.

IWA Publishing
Alliance House, 12 Caxton Street
London SW1H 0QS, United Kingdom
Tel: +44 (0) 20 7654 5500
Fax: +44 (0) 20 7654 5555
www.iwapublishing.com
publications@iwap.co.uk

© Copyright 2014 by the Water Environment Research Foundation. All rights reserved. Permission to copy must
be obtained from the Water Environment Research Foundation.
Library of Congress Catalog Card Number: 2014953551
IWAP ISBN: 978-1-78040-705-0/1-78040-705-X

This report was prepared by the organization(s) named below as an account of work sponsored by the Water
Environment Research Foundation (WERF). Neither WERF, members of WERF, the organization(s) named
below, nor any person acting on their behalf: (a) makes any warranty, express or implied, with respect to the use
of any information, apparatus, method, or process disclosed in this report or that such use may not infringe on
privately owned rights; or (b) assumes any liabilities with respect to the use of, or for damages resulting from the
use of, any information, apparatus, method, or process disclosed in this report.

Brown and Caldwell, University of New Mexico

This document was reviewed by a panel of independent experts selected by WERF. Mention of trade names or
commercial products or services does not constitute endorsement or recommendations for use. Similarly,
omission of products or trade names indicates nothing concerning WERF's positions regarding product
effectiveness or applicability.

ii

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
About WERF
The Water Environment Research Foundation, formed in 1989, is America’s leading
independent scientific research organization dedicated to wastewater and stormwater issues.
Throughout the last 25 years, we have developed a portfolio of more than $130 million in
water quality research.

WERF is a nonprofit organization that operates with funding from subscribers and the federal
government. Our subscribers include wastewater treatment facilities, stormwater utilities, and
regulatory agencies. Equipment companies, engineers, and environmental consultants also
lend their support and expertise as subscribers. WERF takes a progressive approach to
research, stressing collaboration among teams of subscribers, environmental professionals,
scientists, and staff. All research is peer reviewed by leading experts.

For the most current updates on WERF research, sign up to receive Laterals, our bi-weekly
electronic newsletter.

Learn more about the benefits of becoming a WERF subscriber by visiting www.werf.org.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems iii

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
ACKNOWLEDGMENTS
The authors would like to acknowledge the media donors for their financial support and
advice during the conduct of this experiment. They would also like to thank the staff at the
Albuquerque, New Mexico, Wastewater Reclamation Facility for providing inoculation
source material for the reactors. In additional to the project team members listed below, the
research team also thanks the support staff at the University of New Mexico and Brown and
Caldwell, including Donnie Stallman, for their assistance in this work. Finally, they
acknowledge the support and advice of the members of the project subcommittee.

Research Team
Principal Investigator:
Henryk Melcer, Ph.D., P.E.
Brown and Caldwell
Andrew J. Schuler, Ph.D., P.E.
University of New Mexico

Project Team:
Kody Garcia, M.S. candidate
Patrick, D. McLee, Ph.D. candidate
Yunjie Tu, Ph.D.
University of New Mexico
Richard J. Kelly, Ph.D., P.E.
Adam N. Klein, M.D., M.S., P.E.
Matt Winkler, M.S., E.I.T.
Brown and Caldwell

WERF Project Steering Committee


Chandler H. Johnson, P.E.
World Water Works, Inc.
Richard M. Jones, Ph.D., P.E.
Envirosim Associates, Ltd.
Amit Kaldate, Ph.D.
Infilco Degremont
Mark van Loosdrecht, Ph.D.
Delft University of Technology (TU-Delft)
Eberhard Morgenroth, Ph.D.
Swiss Federal Institute for Environmental Science and Technology (EAWAG)
Denny S. Parker, Ph.D., P.E., N.A.E. (Chairman)
Brown and Caldwell

iv

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Bruce E. Rittmann, Ph.D.
Arizona State University
Rajeev Goel, Ph.D, P.E., P.Eng.
Oliver J. Schraa, M.Eng.
Hydromantis Software Solutions, Inc.
Robert A. Zimmerman, Ph.D., P.E.
City of Moorhead, MN

Water Environment Research Foundation Staff


Director of Research: Amit Pramanik, Ph.D., BCEEM
Senior Program Director: Lauren Fillmore, M.S.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems v

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
ABSTRACT AND BENEFITS

Abstract:

The goals of this project were to establish a standard protocol for measuring mass
transfer rates in biofilm media, and to use the resulting data to develop and calibrate an
empirical model. Two laboratory-scale continuous flow reactors were loaded with MBBR
media donated from two manufacturers, and fed a synthetic wastewater free of organic
carbon, but rich in ammonia. The reactors were aerated using coarse-bubble aeration, and
operated at two different temperatures for one year. The continuous system was periodically
halted to allow for batch testing of the media within each reactor. A total of 200 batch tests
were conducted, at varying mixing rates, temperatures, bulk phase dissolved oxygen and
ammonia concentrations. Data from the batch tests were used to develop an empirical model
of ammonia mass transfer. The model includes half-order terms for bulk phase dissolved
oxygen (DO) concentration and the mixing rate (expressed in terms of the velocity gradient),
and an Arrhenius-type temperature dependence. A reasonable fit was obtained through non-
linear regression, with an average root-mean-squared-error of less than 10% between model
and observed ammonia flux rates. The protocol and resulting model were successfully used to
compare the two media, and to project mass transfer within a nitrifying system.

Benefits of Protocol and Model:

 Create specificity to media and conditions being investigated.


 Assist practitioners in reducing process design uncertainty.
 Allow consulting engineers to tailor specific process designs.
 Provide more transparency in the comparison of alternative systems.
 Provide plant owners with more alternatives to select from and allows them to optimize
operating conditions.

Keywords: MBBR, IFAS, biofilm, fixed-film, batch testing, empirical modeling, mass
transfer, nitrification.

vi

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
TABLE OF CONTENTS
Acknowledgments .......................................................................................................................... iv
Abstract and Benefits ..................................................................................................................... vi
List of Tables.................................................................................................................................. ix
List of Figures ................................................................................................................................. x
List of Acronyms ........................................................................................................................... xi
Executive Summary .................................................................................................................. ES-1

1.0 Introduction .................................................................................................................... 1-1

2.0 Literature Review........................................................................................................... 2-1


2.1 Kinetic Rate Measurement ................................................................................... 2-1
2.2 Modeling Background .......................................................................................... 2-3

3.0 Experimental Program Development ........................................................................... 3-1

4.0 Methods ........................................................................................................................... 4-1


4.1 Continuous-Flow Reactors and Feed ................................................................... 4-2
4.2 Continuous-System Reactor Operation ................................................................ 4-5
4.2.1 pH Control................................................................................................ 4-5
4.2.2 Mixing ...................................................................................................... 4-5
4.3 Startup and Routine Operation of Continuous Systems ....................................... 4-8
4.4 Batch Tests ......................................................................................................... 4-10
4.5 Analytical Methods ............................................................................................ 4-11
4.5.1 Biomass Measurements .......................................................................... 4-11
4.5.2 Nitrogen Species .................................................................................... 4-12
4.5.3 Dissolved Oxygen .................................................................................. 4-12
4.5.4 pH ........................................................................................................... 4-12
4.5.5 Data Collection and Recording .............................................................. 4-12
4.5.6 Batch Test Data Analysis ....................................................................... 4-13

5.0 Model Concept ................................................................................................................ 5-1


5.1 Batch Test Observations ...................................................................................... 5-2
5.2 Model Format ....................................................................................................... 5-2
5.3 Model Sensitivity Testing .................................................................................... 5-3
5.4 Model Equation .................................................................................................... 5-5

6.0 Experimental Results ..................................................................................................... 6-1


6.1 Continuous-Flow Reactors ................................................................................... 6-1
6.1.1 Phase 1 (21°C) ......................................................................................... 6-1
6.1.2 Phase 2 (10.5°C) ...................................................................................... 6-6
6.1.3 Phase 3 (Return to 21°C) ......................................................................... 6-7
6.1.4 SALR and Flux (J) ................................................................................... 6-8

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems vii

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
6.2 Batch Tests ......................................................................................................... 6-10
6.2.1 Reproducibility of Batch Tests Performed on a Single Day .................. 6-10
6.2.2 Correlation between Mixing Rate and DO Concentration ..................... 6-10
6.2.3 Batch Test Results.................................................................................. 6-11

7.0 Model Evaluation ........................................................................................................... 7-1

8.0 Findings........................................................................................................................... 8-1


8.1 Model Equation .................................................................................................... 8-1
8.1.1 Bulk Phase Ammonia .............................................................................. 8-1
8.1.2 Bulk Phase DO ......................................................................................... 8-1
8.1.3 Mixing Rate ............................................................................................. 8-2
8.1.4 Temperature ............................................................................................. 8-2
8.1.5 Partial Nitrification .................................................................................. 8-2
8.1.6 Model Structure ....................................................................................... 8-3
8.2 Model Evaluation ................................................................................................. 8-3
8.3 Media Comparison ............................................................................................... 8-3
8.4 Model Application ............................................................................................... 8-4
8.5 Continuous-Flow Reactors................................................................................... 8-4
8.6 Oxygen Flux......................................................................................................... 8-6

9.0 Conclusions ..................................................................................................................... 9-1


9.1 Experimental Methods ......................................................................................... 9-1
9.2 Mass Transfer Modeling ...................................................................................... 9-1
9.3 Model Application ............................................................................................... 9-2
9.4 Media Comparison ............................................................................................... 9-2
9.5 Next Steps ............................................................................................................ 9-2

10.0 Recommendations ........................................................................................................ 10-1


10.1 Recommended Continuous-System Protocol .................................................... 10-1
10.1.1 Reactor Construction ............................................................................. 10-1
10.1.2 Nutrient and Water Feed Preparation .................................................... 10-2
10.1.3 Plastic Media Inoculation ...................................................................... 10-3
10.1.4 Selection of Percent Fill and Mixing Rates ........................................... 10-3
10.1.5 Continuous-Flow System Startup .......................................................... 10-4
10.1.6 Daily Maintenance ................................................................................. 10-5
10.1.7 Monitoring Performance ........................................................................ 10-5
10.2 Recommended Batch Test Protocol ................................................................... 10-5

Appendix A ................................................................................................................................. A-1


Appendix B ..................................................................................................................................B-1
Appendix C ..................................................................................................................................C-1
Appendix D ................................................................................................................................. D-1

References ....................................................................................................................................R-1

viii

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
LIST OF TABLES

4-1 Reactor Specifications ...................................................................................................... 4-3


4-2 Synthetic Feed .................................................................................................................. 4-4
4-3 Target Aeration Rates and Corresponding G Values Used in Batch Tests ...................... 4-5
4-4 Full-Scale MBBR Plant Calculated Values of G (SI Units) ............................................ 4-7
4-5 Reactor Operation Phases ................................................................................................ 4-9
4-6 Comparison Between Brush and On-Media Methods of Biofilm Measurement ........... 4-12
7-1 Summary of Model Parameters and Goodness-of-Fit ...................................................... 7-3
7-2 Statistical Analysis of Model Parameters ........................................................................ 7-5
7-3 Covariance Matrix for Model Parameters ........................................................................ 7-5
10-1 Recommended and Example Values for Continuous-System MBBR ........................... 10-1
10-2 Synthetic Feed (Net Concentrations) ............................................................................. 10-2

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems ix

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
LIST OF FIGURES
1-1 Velocities Influencing Media Particles in MBBR and IFAS Systems ............................ 1-2
4-1 System Design Schematic ................................................................................................ 4-1
4-2 Layout of the Two Continuous-Flow Reactors and Supplementary Equipment ............. 4-2
4-3 Single, Continuous-Flow Reactor Configuration ............................................................ 4-2
4-4 Top: Reactor 2 Media (Media 2); Bottom: Reactor 1 Media (Media 1) .......................... 4-4
4-5 Frequency Distribution of G Values Observed in 32 Full-Scale MBBR
Reactors Spread Across 16 Facilities ............................................................................... 4-8
4-6 Typical Batch Test Result .............................................................................................. 4-13
5-1 Schematic Representation of Mass Transfer from Bulk Liquid to a Biofilm .................. 5-1
5-2 Reaction Rate Versus Bulk Ammonia Concentration ..................................................... 5-3
5-3 Reaction Rate Versus Bulk DO Concentration ................................................................ 5-3
5-4 Ammonia Flux Versus Velocity Gradient at Constant DO (7.5 Mg/L) ........................... 5-4
6-1 Reactor 1 Performance ..................................................................................................... 6-2
6-2 Reactor 2 Performance ..................................................................................................... 6-5
6-3 R1 Specific Ammonia Loading Rate (SALR) Plotted Against (A) Ammonia Flux
(J), and (B) Effluent Ammonia Concentration ................................................................ 6-8
6-4 R2 Specific Ammonia Loading Rate (SALR) Plotted Against (A) Ammonia Flux
(J), and (B) Effluent Ammonia Concentration ................................................................ 6-9
6-5 Reproducibility of Batch Test Data ............................................................................... 6-10
6-6 Empirical Relationship Between Mixing Rate and Reactor Dissolved Oxygen
Concentration ................................................................................................................. 6-11
6-7 Number and Timing of Batch Tests in Each Reactor .................................................... 6-12
6-8 Batch Test Results for Reactor 1 at 21°C ...................................................................... 6-12
6-9 Batch Test Results for Reactor 1 at 10.5°C ................................................................... 6-13
6-10 Batch Test Results for Reactor 2 at 21°C ...................................................................... 6-14
6-11 Batch Test Results for Reactor 2 At 10.5°C .................................................................. 6-14
7-1 R1 Batch Test Data with Model Fit at 21°C .................................................................... 7-2
7-2 R2 Batch Test Data with Model Fit at 21°C .................................................................... 7-2
7-3 R2 Batch Test Data with Model Fit at 10.5°C ................................................................. 7-3
7-4 Model Versus Observed Ammonia Flux, Reactor 1 ........................................................ 7-4
7-5 Model Versus Observed Ammonia Flux, Reactor 2 ........................................................ 7-4
8-1 Model Projection of Media-Specific Reaction Rates at T = 20°C and DO = 5 mg/L ..... 8-4
10-1 Lab-Scale Moving-Bed Bioreactor Configuration......................................................... 10-4
10-2 Media Removed from Reactor Before Batch Testing.................................................... 10-6
10-3 Cleaning of Reactor with Brush Before Each Batch Test.............................................. 10-6
10-4 Filling of Reactor with Batch Test Media...................................................................... 10-7
10-5 Return of Plastic Media to the Reactor .......................................................................... 10-7

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
LIST OFACRONYMS
µm micrometer(s)
AOB ammonia-oxidizing bacteria
BOD biochemical oxygen demand
cm centimeter(s)
d day(s)
DI deionized
DNA deoxyribonucleic acid
DO dissolved oxygen
FA free ammonia
FNA free nitrous acid
G velocity gradient
h hour(s)
HCl hydrogen chloride
IFAS integrated fixed-film/activated sludge
J ammonia flux
L liter(s)
m2 square meter(s)
m3 cubic meters(s)
MBBR moving-bed biofilm reactor
mg milligram(s)
min minute(s)
mL milliliter(s)
mm millimeter(s)
MTBL mass transfer boundary layer
N nitrogen
Na2CO3 sodium carbonate
NaHSO3 sodium bisulfite
NH3 ammonia
NOB nitrite-oxidizing bacteria
NOx oxidized nitrogen (nitrite + nitrate)
OUR oxygen utilization rate
PVC polyvinyl chloride
R reactor
RMSE root-mean-squared error

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems xi

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
RR removal rate
s seconds
SALR surface area loading rate
SBR sequencing batch reactor
SINTEF Foundation for Scientific and Industrial Research
T temperature
TBS total biofilm solids
TKN total Kjeldahl nitrogen
TSS total suspended solids
U.S. United States
VBS volatile biofilm solids
VSS volatile suspended solids
WRF wastewater reclamation facility

xii

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
EXECUTIVE SUMMARY
The moving-bed biofilm reactor (MBBR) process was developed and patented by the
Norwegian company Kaldness Miljoteknologi in cooperation with the Foundation for Scientific
and Industrial Research at the Norwegian University for Science and Technology (SINTEF) in
Trondheim, Norway (Rusten et al., 1998). This technology deploys a continuously operating,
non-clogging biofilm reactor that does not require backwashing and features low head loss and a
high specific biofilm surface area.
Biofilms grow on small plastic carrier elements (or media) that are mixed with
wastewater in the reactor. In aerobic systems, mixing is induced by aeration; in denitrifying
systems, with mechanical mixers. The carrier elements are made of synthetic materials of various
shapes and sizes, optimized for mechanical integrity and maximum surface area for biofilm
growth. Specific biofilm surface area may also be varied by adjusting the number of carrier
elements in the reactor. The MBBR is a once-through reactor; there is no solids recycle. A later
development added solids recycle to this concept, and the modified system became known as the
integrated fixed film/activated sludge (IFAS) system.
The theoretical understanding of such systems has progressed in terms of the types of
biofilm models available (Wanner et al., 2006), but the practical application of such models is
limited and difficult due to the complex nature of the models developed to date. The need for a
simplified set of modeling tools has been recognized (Morgenroth et al., 2000; Boltz et al.,
2010). In practice, engineering consultants use commercially available models to simulate
different biological treatment systems and demonstrate their effectiveness for a given
application. One of the problems impeding widespread reliance on biofilm model components in
process simulators is the need for users to specify external mass transfer rates to the biofilm. A
goal of this work was to improve the understanding of the external mass transfer characteristics
of floating media that are typically used in MBBR and IFAS systems.
The first major objective of the proposed work was to establish a WERF standard
protocol for measuring media mass transfer rates in IFAS and MBBR systems with moving
media. Data from this protocol could be used to achieve the second major objective of calibrating
a simple model that describes the mass transfer within moving media. Application of these tools
by engineers, researchers, vendors, and others will minimize process design uncertainty and
provide plant owners more transparency in the comparison of alternative media systems.
A simple laboratory-based protocol was developed at the University of New Mexico
laboratories in which a candidate media was placed in a 20-liter reactor operated at constant
temperature and aerated to ensure complete mixing of the media in the reactor. With this being a
first approach to developing such a protocol, the analysis was restricted to nitrifying systems,
given the complexity of simultaneously modeling nitrification and heterotrophic carbon
oxidation in biofilms. Synthetic feed containing ammonia was delivered to the reactor to produce
highly active nitrifying biofilms for use in batch tests. The role of the reactor was similar to a
“parent” reactor in which the goal was to produce media covered with deep nitrifying biofilms
such that biofilm depth was not limiting ammonia oxidation but was limiting oxygen mass
transfer.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems ES-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Once an active biofilm was developed on the media, batch tests were conducted within
the reactor. The batch tests measured the rate of ammonia oxidation, with mixing rate (G),
dissolved oxygen (DO) concentration, and temperature (T) as experimental variables. The data
from the batch tests were used to develop a simple empirical model of ammonia utilization rate
as a function of G, DO, and T. Two vendor-donated media were evaluated in this manner.
The first goal was achieved in that the protocols developed allowed for bench-scale
evaluation of MBBR systems containing floating media by using relatively small reactors and
materials that are readily available to academic and design practitioners.
The intent of the modeling phase of this work was to predict the media-specific reaction
rate, given inputs of solute concentrations, temperature, and mixing. To that end, an empirical
model for determining the ammonia reaction rate was developed (Equation ES-1):
T
√ √ E uation E 1
where:
J = mass flux relative to media surface area (mass/area/time)
k = media coefficient (mass/area/time)
CDO, bulk = bulk phase DO concentration (mass/volume)
G = velocity gradient (1/s)
: a unitless temperature-dependence constant
T = temperature (°C)
A half-order or saturation-type relationship was noted between the bulk phase DO
concentration and the ammonia flux rate. Although previous studies failed to observe this effect,
this was most likely due to operation at lower DO concentrations than those tested in this study.
A half-order or saturation-type relationship was also noted between the mixing rate and the
ammonia flux rate. Although previous studies have noted a relationship between mixing and
flux, the nature of the response has not previously been noted. This may be related to reaction
kinetics becoming rate-limiting once mass transfer limitations have been overcome through
mixing. At the conditions tested in this work, there was little difference between a half-order or
saturation-based model; a half-order model was selected in order to reduce the number of
calibration parameters, and to keep the model equation as simple as possible. Further work may
be required to determine whether the half-order or saturation response is more appropriate.
The empirical model proposed in Equation ES-1 provided a good fit to observed data,
with an average RMSE (root mean square error) of less than 10% from observed values taken
from over 100 batch tests. Based on this, the model appears suitable for media comparison and
prediction of ammonia flux rates for a nitrifying MBBR system. Under the conditions tested in
this experiment, the model demonstrated that one media performed better than the other, with
higher ammonia flux rates. Further testing would be required to compare the media at other
conditions. The work also demonstrated the risks associated with mixing, as one media shed over
90% of its biomass when subjected to a high mixing rate at cold temperature.
Although modeling was limited to ammonia oxidation in this experiment, a next step
would be to develop similar expressions for carbon removal and/or denitrification.

ES-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
CHAPTER 1.0

INTRODUCTION

The state-of-the-practice in using science-based models for biofilm processes has lagged
behind that of activated sludge systems mainly because the complexity of defining biofilm flow
conditions, mass transfer phenomena, and the bioflocculation and sloughing characteristics of
these systems; these difficulties have confounded the development of mechanistic models simple
and transparent enough for accurate process design sizing (Parker, 2010). This contrasts with
activated sludge systems, the design of which has blossomed thanks to the development of the
powerful mechanistic models available today. Today, the rock trickling filter remains the most
common U.S. biofilm reactor system, a technology that was designed solely on the basis of
empirical equations. In terms of total market share in the United States, biofilm-based
wastewater treatment systems have stagnated nearly 18% over the past 30 years, while activated
sludge treatment systems have increased from 31 to 58%. The biofilm reactor sector has
remained relatively static despite the introduction of new biofilm technologies such as the
biological aerated filter and the integrated fixed-film/activated sludge (IFAS) and moving-bed
biofilm reactor (MBBR) moving-media technologies, all of which have major space and cost
saving advantages, especially in nutrient removal applications (Parker, 2010).
The MBBR process was developed and patented by the Norwegian company Kaldness
Miljoteknologi in cooperation with the Foundation for Scientific and Industrial Research at the
Norwegian University for Science and Technology (SINTEF) in Trondheim, Norway (Rusten et
al., 1998). The basic concept of the MBBR was to have a continuously operating, non-clogging
biofilm reactor that did not require backwashing and featured low head loss and a high specific
biofilm surface area. This was achieved by having the biofilm grow on small carrier elements
that move along with the water in the reactor. In aerobic systems, the movement is induced by
aeration. The biofilm elements are made of synthetic material and come in various shapes and
sizes, and are typically optimized for mechanical integrity and maximum surface area for biofilm
growth. The filling of carrier elements within the reactor may be varied, providing considerable
flexibility in the specific biofilm area. The MBBR is a once-through reactor; there is no solids
recycle. When solids recycle was subsequently added to this concept, the modified system
became known as IFAS.
The theoretical understanding of such systems has progressed in terms of the types of
biofilm models available (Wanner et al., 2006), but the practical application of such models is
limited and difficult due to the complex nature of the models developed to date. The need for a
simplified set of tools has been recognized (Morgenroth et al., 2000; Boltz et al., 2010).
Consultants typically look for commercial firms to build model platforms and demonstrate their
effectiveness for practical use. Increasing pressure on consultants to apply process design models
of biofilm systems comes from municipalities that are requiring their consultants to use the best
science in their designs, often specifying which model platform is to be used. One of the
problems impeding widespread reliance on biofilm elements in model platforms is the need for
users to specify external mass transfer rates to the biofilm. A goal of this work was to improve
the understanding of the external mass transfer characteristics of floating media that are typically
used in MBBR and IFAS systems.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 1-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
In a thorough review of biofilm models, Boltz et al. (2010) identified mass transfer
external to the biofilm as one of the least understood but most influential aspects affecting
biofilm reactor performance. Further, of the models reviewed in use internationally, they found
that all but one, TRIFIL, a trickling filter model, simplified the problem to an idealized
description of a mass transfer boundary layer (MTBL). In some of these models, the MTBL is
correlated to dimensionless system parameters such as the Sherwood and Reynolds numbers,
using a characteristic length (a media dimension) and the water velocity in the vicinity of the
biofilm surface.
Although this relates mass transfer to reactor hydraulics, there are difficulties with the
empirical MTBL approach. First, the available correlations are for spheres, which are not a good
analogy for the media particles typically used in IFAS or MBBR systems.
Second, the MTBL approach requires knowledge of the fluid velocity near the biofilm
surface. As discussed in Parker (2010), because both the media and the liquid are in motion, the
relative velocities of the media and fluid are important in determining the fluid velocity near the
biofilm surface. Unfortunately, this differential is difficult to determine in practice. Further
complicating the system, the biologically active surfaces on the media occur in the inner
surfaces, so the relevant flow velocity is the relative fluid velocity passing through the media
(Figure 1-1). This velocity is affected by the differential between the media and fluid velocities,
the orientation of the media, the rotational velocity of the media, and the media dimensions.
Further, because the motion of the media is random and chaotic as a result of the fluid
turbulence, the inside fluid velocity is continuously changing because pressures and velocities at
the end of the media channels are always varying.

Figure 1-1. Velocities Influencing Media Particles in MBBR and IFAS Systems.

It is clear that the fluid dynamics affecting the boundary layer thickness and mass transfer
within moving biofilm media are complex and cannot be readily defined. The significant and
likely insurmountable challenges to modeling this type of system support the need for a simpler
approach, based on readily measurable parameters.

1-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
The first major objective of the proposed work was to establish a WERF standard
protocol for measuring media mass transfer rates in IFAS and MBBR systems with moving
media. Data from this protocol could be used to achieve the second major objective of calibrating
a simple model that describes the mass transfer within moving media. Application of these tools
by engineers, researchers, and vendors and others will:
 Assist in eliminating process design uncertainty.
 Allow consulting engineers to tailor specific process designs.
 Provide more transparency in the comparison of alternative systems because consultants
currently have no way to independently evaluate vendor proposals for moving media in IFAS
and MBBR systems on a comparable basis.
 Give plant owners more treatment alternatives to select from and allow them to optimize
process operating conditions.
Improving the ability to analyze vendors’ proposals more thoroughly will increase the
confidence amongst designers in moving bed biofilm applications, leading to increases in their
penetration into the marketplace.
Based on the current complexities and limitations of biofilm models and the constraints
that these place on designers of MBBR and IFAS fixed-film systems, this work set out to
simplify these models by refocusing efforts on empirical models derived from readily obtainable
data collected from full-scale, pilot-scale, or simple bench-scale applications. This work is
intended to be a first step toward development of such models, and considers ammonia
utilization in a nitrifying MBBR system, a common application. It was the research team’s
intention that the modeling framework developed for the nitrifying MBBR system would be
applicable to a wider range of systems, including biochemical oxygen demand (BOD) removing
systems and denitrifying systems.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 1-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
1-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
CHAPTER 2.0

LITERATURE REVIEW
Biofilm-based treatment technologies have been used for domestic wastewater treatment
since the development of centralized wastewater treatment systems more than a century ago.
Trickling filters are considered the oldest of these technologies (including rock and plastic
media), with later developments including rotating biological contactors in the 1960s, and
biological aerated filters in the 1980s. Systems with biofilm carriers in completely mixed
reactors such as MBBR and IFAS have been in development and application since the 1980s
(Hem et al., 1994, Muller, 1998), and their use for wastewater treatment continues to increase.
2.1 Kinetic Rate Measurement
The determination of the substrate utilization rates, and in particular ammonia utilization
rates, in MBBR and IFAS systems is of great interest for system design and operation. Previous
research reporting such rates has relied on measurements in continuous systems and/or in batch
tests of media.
Determination of utilization rates in continuous systems depends upon measurements of
influent and effluent concentrations for a given set of conditions. In this case, evaluating
experimental variable effects, such as the bulk phase dissolved oxygen (DO) concentration, on
utilization rates requires changing the reactor conditions and performing measurements after
reactor performance has stabilized. For example, Hem et al. (1994) used data from continuous-
flow pilot and bench systems to determine ammonia utilization rates, although little information
was provided about these protocols and calculations. It has also been common to report removal
rates plotted against loading rates, based on operation of continuous systems with variable
loading (Ødegaard et al., 2000).
Others have applied batch testing to determine substrate utilization rates in MBBRs and
IFAS systems, in which these rates are determined by measuring concentration changes over
time in batch reactors. This approach has the advantage of allowing more rapid testing of
multiple conditions (such as variable concentrations of a solute). Biofilm media batch testing has
generally been based on protocols developed for suspended growth (activated sludge) systems,
such as those described for determining oxygen utilization rates (OUR), ammonia utilization
rates, and nitrate utilization rates by Kristensen et al., (1992). These tests were performed in
1-liter (L) cylinders with mixing by aeration. Measurements were taken over several hours to
determine nitrification rates with initial ammonia concentrations of 20 milligrams (mg) nitrogen
per liter (N/L). OURs were determined in sealed BOD flasks without aeration. Melcer et al.,
(2003) described a batch test protocol (the Low F/M Bioassay) in which nitrite and nitrate
production rates were determined for nitrifying activated sludge mixed with wastewater influent
or secondary effluent spiked with ammonia, with the end goal of estimating nitrifier maximum
growth rate (µaut). DO concentrations of at least 6 mg/L were considered sufficient to avoid
oxygen limitation of nitrification. In one example, an ammonia concentration of approximately
15 mg N/L was considered sufficient to avoid ammonia limitation.
A batch protocol for testing moving media based on that described by Kristensen et al.,
(1992) was used by Di Trapani et al., (2011). In this protocol, moving media was placed in a 5 L

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 2-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
reactor, with mixing by aeration. The initial ammonia concentration was approximately 58 mg
N/L. Ammonia consumption and nitrite and nitrate production rates were calculated based on
measurements of these constituents collected from the reactors over two to three hours.
Di Trapani et al., (2011) applied this method to evaluate the effects of organic loading and
ammonia loading on nitrification at low temperatures. Di Trapani et al., (2013) applied the same
batch testing protocol in the evaluation of solids residence time and organic loading. Zekker et
al., (2011) applied a similar approach, but modified it to be conducted in 500-milliliter (mL)
Erlenmeyer flasks filled with cultured media taken from a 20 L bench-scale MBBR, with
synthetic feed, mixing by aeration, and hourly measurements over five hours. Pastorelli et al.,
(1997) evaluated denitrification rates in batch tests of media taken from a pilot-scale MBBR
system in 1.8 L reactors, with mechanical mixing by a magnetic stirrer. The tests were conducted
with a relatively low media content (50 square meters per 1 cubic meter [m2/m3]) to aid in
mixing of the carriers.
Batch testing has been done within continuous system reactors, rather than removing
media for testing in separate reactors. For example, Zhang et al., (2014) performed batch tests in
aeration-mixed pilot-scale reactors (4.4 m3), where influent feed was stopped and 2 mg N/L of
ammonia was added to start the tests. Other studies of MBBRs operating in a sequencing batch
reactor (SBR) configuration have determined substrate utilization rates based on intermediate
measurements over the course of SBR cycles (Bassin et al., 2012).
Previous studies of substrate utilization rates in MBBRs and IFAS systems have not
generally considered the effects of mixing. One notable exception was Bott et al., (2010), who
tested the effects of mixing on nitrification rates of IFAS media in 7 L batch tests. Flat-paddle
mixers were used with fine-bubble aeration. Initial measurements of mixing rates (G) and
applied torque allowed calculation of the velocity gradient G. Plastic media from a
demonstration-scale IFAS system were added to the reactors at 50% fill, with secondary effluent
and/or IFAS mixed liquor to 7 L. Both DO (2 to 8 mg/L) and G (97 per second [/s] to 253/s)
were varied.
Based on a review of these previous studies, the advantages of using batch tests over
continuous systems to determine effects of experimental parameters on reaction rates include:
 Batch tests allow the analysis of a larger number of experimental variables in a given time
than do continuous-system measurements, as multiple batch tests can be conducted in a
single day, while the continuous-system approach typically requires waiting for reactor
stabilization (over a period of days or longer) after implementation of each condition.
 Batch tests likely provide better isolation of a single variable’s effect on reaction rates, as the
same biofilm can be tested over several batch tests in a single day. This reduces the
possibility of biomass changes (such as thickness or population composition) that are more
likely to occur in continuous systems operated over longer periods for each condition tested.
(However, batch testing protocols should be evaluated for repeatability over a single day to
guard against effects such as significant changes in biofilm growth or detachment during
batch testing, as discussed later in this document).
Testing of continuous systems provides at least one advantage over batch testing to
evaluate system biokinetics, which relates to item 2 above. Although providing time for
acclimation of biofilms (or activated sludge) for each test may obscure the isolated effect of an
experimental variable on a rate in continuous testing, it also provides additional information

2-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
about secondary effects, such as increased activity due to additional biomass accumulating under
a new condition, which may be of interest depending on the experimental goals. The selection of
batch or continuous-system testing should therefore be based on overall experimental objectives
and time constraints.
2.2 Modeling Background
Historically, biofilm modeling has focused on a combination of diffusive, kinetic, and
mechanical behaviors. Diffusion is expressed in terms of Fic ’s Law, which expresses the mass
transport of solutes to, from, and within the biofilm. Reaction kinetics are expressed in terms of
biofilm cell growth, decay, and substrate metabolism. Mechanical factors include the shearing
and sloughing of biomass from the biofilm under mixing conditions. The combination of these
three behaviors has resulted in highly complex models, which depend upon a number of
parameter values that are difficult to measure in situ.
The WEF Manual of Practice No. 35 (WEF, 2010) presents three frameworks for biofilm
models – a semi-empirical model, a one-dimensional model, and a two-dimensional model. In its
simplest form, the semi-empirical model takes the following form (Equation 2-1):

(Equation 2-1)

Where
BN,n = ammonia uptake rate in the biofilm (mass/time)
qm,NH4N-Nitr,bf = flux rate for ammonia uptake by nitrifiers (mass/area/time)
SO2,n = bulk phase oxygen concentration (mass/volume)
SN,n = bulk phase ammonia-nitrogen concentration (mass/volume)
KDO,bf = half saturation constant for dissolved oxygen for nitrifier growth in the biofilm
(mass/volume)
KN,bf = half saturation constant for ammonia for nitrifier growth in the biofilm
(mass/volume)
Vn = volume of cell n (volume)
Mn = biofilm surface area per unit of cell volume in cell n (area/volume)
However, Equation 2-1 must be expanded to account for numerous other parameters,
including carbon and phosphorus, and other elements of the biofilm, including heterotrophic
aerobic metabolism and heterotrophic anoxic metabolism. When these are included, the
relatively simple Equation 2-1 rapidly becomes complex, as evidenced by the four pages of
tabulated parameter values accompanying this equation (WEF, 2010).
The one- and two-dimensional models become more complex as they attempt to account
for mass transport and reaction kinetics as a function of position within the biofilm.
Other biofilm modeling textbooks (Eberl et al., 2010; Henze et al., 2008) group modeling
approaches in the categories described next.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 2-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Analytical models can be solved without advanced numerical techniques, and all terms
within the model are directly and explicitly understood. Such models are typically solvable only
for systems with a single rate-limiting substrate with an assumption of a uniform, one-
dimensional biofilm. The semi-empirical model as described above would be an example of an
analytical model.
Pseudoanalytical models are geared toward more complex situations. The resulting
mathematical formulations are distilled into a set of algebraic expressions and solved through
advanced computational methods, the results of which may be applied by practitioners.
Pseudoanalytical approaches have been developed by several authors. The work of Saez and
Rittmann (Saez and Rittmann, 1988 and 1992), for example, is presented in detail in Grady,
Daigger, and Lim (1999) as well as in Rittmann and McCarty (2001). This approach requires
iterative calculation of Equation 2-2, which itself is made up of dimensionless parameters and
parameter values obtained through the solution of a set of differential equations:

[ ( ) ] { [ ]}
(Equation 2-2)

Where
= dimensionless substrate concentration in the bulk phase
= dimensionless substrate concentration at the liquid-biofilm interface
Ri = Rittmann number, a dimensionless parameter derived from kinetic and
stoichiometric constants related to biomass growth, decay, and detachment.
’, ’ dimensionless parameters, obtained through solution of set of differential
equations across 500 specific conditions
= dimensionless external mass transfer coefficient

The pseudoanalytical approach may be used to compute flux of a solute into a biofilm.
However, the approach depends upon a multitude of parameter values and constants, most of
which require sophisticated laboratory practices or analytical computation to derive.
Numerical one- and two-dimensional models, as discussed above, attempt to
characterize substrate flux throughout the biofilm. Such models are typically employed using
computational programs such as AQUASIM, Biowin, and GPS-X. These powerful tools offer
insight into biofilm processes. However, they also rely upon a host of parameter values that are
often difficult to calibrate to specific or observed conditions.

2-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
In contrast, trickling filter design has been simplified through the application of empirical
and empirically derived models, such as the modified Velz equation (Gromiec et al., 1972;
Parker and Merrill, 1984):

{ [ ] } (Equation 2-3)
( )

Where
Cout = concentration in trickling filter effluent (mass/volume)
Cin = concentration in trickling filter influent (mass/volume)
Rr = recycle ratio
H = height of trickling filter (length)
 = Arrhenius constant for temperature
T = temperature (degree Celsius)
Qi = flow divided by filter cross sectional area (volume/area/time)
n = empirical flow coefficient
As = tricking filter surface area (length)2
k20 = treatability coefficient (volume/area/time)0.5
The empirical models accept that in a turbulent flow situation, it is not possible to directly
calculate the oxygen transfer to the biofilm. They use the rate of substrate oxidation (for
example, nitrification) to achieve an indirect measure of mass transfer. The TRIFIL model,
developed by Logan, Hermanowicz, and Parker (1987), while not entirely empirical, is another
example of a simplified model which relies upon site-specific calibration to simulate complex
processes (BOD removal as a function of size).
Application of these models requires site-specific calibration of model parameters (the
K20 value in the modified Velz equation, the diffusivity class fractions in the TRIFIL model).
However, such calibration depends upon easily obtainable calibration data, such as bulk phase
soluble BOD concentration.
The goal of this experiment is to develop an empirical model for MBBR processes, which
is similar to the trickling filter models cited above, with the following requirements:
 Model inputs are obtained through simple laboratory procedures, such as measurement of
bulk phase concentrations.
 Model may be calibrated against bench-, pilot-, or full-scale data.
 Model can be explicitly solved without the use of a complex biofilm model.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 2-5

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
2-6

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
CHAPTER 3.0

EXPERIMENTAL PROGRAM DEVELOPMENT


The approach advanced in this program is different from prior moving-media studies in
that it focuses on external mass transfer effects, through the inclusion of mixing rates, as distinct
from treatment model kinetic determinations. The intent was to provide a simple procedure that
demonstrates how external mass transfer limitations influence the performance efficiencies of
different media. It builds on preliminary investigations initiated by Bott et al. (2010), Jones et al.
(2010), and Thomas et al. (2009). These investigators used bench-scale reactors to measure
nitrification rates for ammonia-oxidizing bacteria (AOB) and nitrite-oxidizing bacteria (NOB) to
elucidate the behavior of a full-scale IFAS treatment facility. These authors also attempted to
measure the effects of DO concentration and mixing intensity on nitrification rates. For the
current study, nitrifying biofilms were grown without an external carbon source, in order to
suppress heterotrophic growth as much as possible. The intent was to allow direct use of oxygen
uptake as an indicator of mass transfer. The application of a highly concentrated ammonia
substrate ensured that thick biofilms would be established to avoid biofilm thickness being a
limiting condition and that ammonia would not be subject to Monod substrate constraints. This
was similar to the approach used successfully to verify oxygen mass transfer in the TRIFIL
model for trickling filter media (Parker et al., 1997; Parker, 2010).
Laboratory-scale nitrifying MBBR reactors were constructed and operated at the
University of New Mexico using two different media donated by participating manufacturers to
create a standard protocol for measuring mass transfer characteristics of moving media. The
reactors were operated as MBBRs to avoid the complexity and variability that come with
inclusion of the suspended phase associated with IFAS systems. A protocol for batch testing of
the media was developed to calibrate parameters for an empirically derived mass transfer
expression. Key experimental variables in the batch tests included media type, temperature, bulk
phase DO concentration, and mixing rates.
The first major adaptation to the original plan was a change in the reactor design from
mechanical mixing to mixing by aeration. This was done before reactor construction had started,
and was based on input from the project steering committee. The rationale for this change was
that full-scale aerobic IFAS reactors and MBBRs are generally mixed by aeration, and that
mechanical mixers may physically collide with media and possibly damage them and/or
otherwise affect biofilm formation or stability.
The next major step in the development of the reactor design was the evaluation of a
recirculating aeration system, which could in theory allow calculation of OUR by measuring
changes in DO, and tracking changes in the total system oxygen mass using equilibrium
relationships. Briefly, this experimental setup proved unworkable with the available equipment,
possibly due to oxygen leakage into the reactor through the recirculating air pump, which was
not designed to be airtight. As noted in Chapter 4.0, Methods, this abortive experimental setup
and its evaluation are described in more detail in Appendix A.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 3-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
3-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
CHAPTER 4.0

METHODS
The experimental approach is discussed in detail below. To provide context, the
experimental framework consisted of the following:
1. Two laboratory-scale MBBR reactors were constructed and operated with synthetic feed
to produce highly active nitrifying biofilms for use in batch tests. The goal of operating
these reactors was to produce deep nitrifying biofilms such that biofilm depth was not
limiting ammonia oxidation but was limiting oxygen mass transfer.
2. The reactors were operated to approximate steady states at 21°C and 10.5°C to allow
evaluation of temperature effects.
3. A series of batch tests were conducted on the media in these reactors in which
nitrification rates were measured, with mixing rate or velocity gradient (G)
(approximately 160/s to 390/s), DO concentration (3 to 21 mg/L), and temperature
(T, set to the same as the continuous system temperature of 21°C or 10.5°C) as
experimental variables.
4. The data from the batch tests were used to develop a simple empirical model of ammonia
utilization rate as a function of G, DO, and T.
Figure 4-1 illustrates the design schematic for the MBBR systems.

Figure 4-1. System Design Schematic.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 4-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
4-1 Continuous-Flow Reactors and Feed
Two continuous-flow MBBR reactors were constructed from clear acrylic, as shown in
Figures 4-2 and 4-3.

Figure 4-2. Layout of the Two Continuous-Flow Reactors and Supplementary Equipment.

Figure 4-3. Single, Continuous-Flow Reactor Configuration.

4-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Table 4-1 provides the reactor dimensions, total volume (liquid, media, and headspace),
working volume (liquid and media), and liquid volume, where the liquid volume was determined
by removing biofilm-laden media. This was done by turning off the feed and airflows to
the reactors, carefully removing the media, and measuring the volume of the remaining liquid in
a graduated cylinder, including additional liquid from the drained media after agitation by gentle
shaking.

Table 4-1. Reactor Specifications.

Parameter R1 R2
Reactors
Total volume (including headspace) (L) 17.4 17.4
Working volume (liquid + media) (L) 9.36 9.42
Liquid volume (L) 8.15 7.96
Dimensions including head space (W x D x H)
(inches) 8.0 x 8.0 x 16.6
(cm) 20.3 x 20.3 x 42.2
Flow rate (L/d) 10.1 10.1
HRT based on working volume (hour) 22.3 22.4
HRT based on liquid volume (hour) 19.4 19.0
Media
Media specific surface area (m2/m3) 650 630
Media fill volume (percent) 32.1 31.9
Media area in reactor (m2) 1.95 1.89
Media area/working volume (m2/m3) 208 201
Controls
Aeration and mixing method Coarse bubble
Target mixing rate (G) (-/sec) 240-327 (variation discussed below)
Dissolved oxygen concentration (mg/L) >6.5 mg/L (measured but not controlled)
pH control (pH units) 7.15-7.50
Temperature (°C) 21.0 or 10.5 in temperature-controlled water bath

Specifications of these media are provided in Table 4-1, and photographs are provided in
Figure 4-4. Media names and vendors have been omitted.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 4-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 4-4. Top – Reactor 2 Media (Media 2); Bottom – Reactor 1 media (Media 1).

Synthetic feed (composition provided in Table 4-2) was continuously fed to the reactors
using peristaltic pumps (Masterflex; Cole Parmer, Vernon Hills, Illinois, U.S.) as two separate
streams of dechlorinated tap water (7.5 liters per day [L/d]), and autoclaved, concentrated
nutrient feed (2.6 L/d), giving a total influent flow rate of 10.1 L/d. The feed was added in this
manner to reduce the volume of nutrient feed preparation. Tap water was used to provide a
source of trace elements. The tap water residual chlorine was removed by bubble aeration for 24
hours (h) following the addition of 1.5 mg/L sodium bisulfite (NaHSO3) (Bill et al., 2010). Net
concentrations of the synthetic feed (as added to the reactor after mixing with water) are shown
in Table 4-2. Treated effluent overflowed each reactor via the effluent port, the location of which
determined the reactor volume (Figure 4-1).

Table 4-2. Synthetic Feed.


Based on Hem et al., 1994.

Chemical Concentration (mg/L)a


NH4Cl Variable (see text)
KH2PO4 100
NaHCO3 350
FeSO4-7H2O 5
CaCl2 16
MgSO4-7H2O 40
CuSO4-5H2O 0.12
NaMoO4 0.0019
EDTA 6.6
a. Net concentrations as added to reactor; see text.

4-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
No organic carbon was included in the feed to reduce the heterotroph component of the
biofilms as much as possible. Ammonium chloride was the sole source of nitrogen in the feed,
and its concentration was typically adjusted based on maintaining a target effluent ammonia
concentration between 10 and 50 mg N/L (actual effluent ammonia varied considerably with
reactor performance, as discussed further in Chapter 6.0, Experimental Results). The ammonia
dosing regime was developed with the objectives of providing sufficient ammonia to avoid
limiting biofilm activity, and avoiding excessive ammonia that might inhibit nitrification
(discussed further below). Ammonia feed concentrations varied from initial concentrations of 30
to 40 mg N/L early in reactor biofilm development to 400 to 600 mg N/L once the biofilm was
developed.
4.2 Continuous-System Reactor Operation
This section discusses how the continuous-system reactor was operated and maintained.
4.2.1 pH Control
pH was measured and controlled in the range of 7.15 to 7.5 in each reactor with a pH
controller (Chemcadet Model 5652-00, Cole-Parmer, Vernon Hills, Illinois, U.S.) with a
combination, double-junction, gel-filled pH electrode (Model EW-59001-70, Cole-Parmer,
Vernon Hills, Illinois, U.S.). Acid and base solutions were 0.1 M hydrogen chloride (HCl) and
0.7 M sodium carbonate (Na2CO3), respectively.
4.2.2 Mixing
Mixing of the reactor was by coarse-bubble aeration to mimic mixing designs in full-
scale MBBR and IFAS systems. Coarse-bubble air was maintained at a given flow rate to ensure
a constant mixing intensity in the reactors, as measured using G (see below for calculation
method). Correlations between the supplied airflow rate and G are provided in Table 4-3.

Table 4-3. Target Aeration Rates and Corresponding G Values Used in Batch Tests.

Airflow (ft3/min) Airflow (L/min) G (1/s)


0.26 7.4 180
0.36 10.2 210
0.46 13.0 240
0.56 15.9 260
0.71 20.1 300
0.86 24.4 330
1.21 34.3 390

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 4-5

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
4.2.2.1 G Calculation
Mixing intensity was expressed by the velocity gradient, G, as calculated from Equation
4-1 (Parker, 1970).

where:
G = velocity gradient (1/s)
Q= airflow rate (m3/s)
 = water specific weight (N/m3)
HL = head loss (m) (distance from aerators to water surface)
 = water dynamic viscosity (N*s/m2)
V= volume (m3): working volume used for all calculations

An example calculation of G is provided in Appendix B.

4.2.2.2 Aeration and Evaluation of Mixing Configurations


Coarse-bubble aeration was provided by pumping air through a horizontal section of
polyvinyl chloride (PVC) pipe spanning the reactor width. The PVC tubing inner and outer
diameters were 1.4 and 2.2 centimeters (cm), respectively. In the upper side of the horizontal
section of the PVC pipe, 1.6-millimeter (mm) (1/16 inch) diameter holes were drilled (at a
spacing of approximately 0.6 cm) to provide coarse-bubble aeration. Aeration was provided by a
vacuum/pressure diaphragm pump (models DAA-V715-EB and DOA-P161-AA, GAST
Manufacturing, Inc., Benton Harbor, Michigan, U.S.).
Several different configurations of aeration diffuser(s) were initially tested (including
locating the aerator in the middle or on the side of the reactor, and using one or two aerators) to
determine which would provide adequate mixing of media at fill volumes of 30% and 50%.
Based on these tests, a single aerator mounted at the side of the reactor was selected as the
optimum configuration for providing mixing of the test media with 30% fill (the exact value used
was 32%) for both media. Operation with a 50% fill volume led to poor mixing with dead zones
in the corners of the reactors for all configurations. A minimum G value of 158/s was required to
provide complete mixing of both test media.
One issue not explored in this research is the potential relationship between the size of
the air bubbles and the magnitude of shear within the reactor. Different sized bubbles may impart
different degrees of shear on the media.

4-6

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
4.2.2.3 Typical Full-Scale Mixing Rates
Mixing rates in the continuous reactors, as well as in the batch experiments, were selected
based upon a review of typical G values used in full-scale MBBR installations. Table 4-4
presents data from 16 full-scale reactors, some of which comprised several basins, along with
calculated G values.

Table 4-4. Full-Scale MBBR Plant Calculated Values of G.


(SI Units)

Air Rates per Basin (m3/min) G (1/s)


Volume Depth
Plant Basin No. Basin No.
m3 m
1 2 3 4 1 2 3 4
A 623 4.1 92 61 315 257
B 1,133 4.9 88 74 249 228
C 696 6.1 27 27 27 27 198 198 198 198
D1 1,631 4.9 139 139 261 261
D2 2,276 4.9 96 96 183 183
E 580 6.1 42 35 270 247
F 58 3.0 9 6 6 274 228 221
G1 306 3.7 9 9 134 134
H 505 6.1 66 23 360 212
I 1,133 5.5 119 60 306 218
J1 1,298 5.3 107 267
J2 882 4.4 43 189
G2 842 6.2 54 54 54 255 255 255
SJ 1,245 4.9 46 171
M 169 4.1 11 210
P 187 3.0 33 27 296 270

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 4-7

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
The range of G values calculated for these systems was 134/s to 360/s. The average
across all plants was 233/s. Most of the high G values are associated with plants operating at
high oxygen uptake rates. Figure 4-5 presents a histogram of the distribution.

Figure 4-5. Frequency Distribution of G Values Observed in 32 Full-Scale MBBR Reactors Spread Across 16 Facilities.

For comparison, a recent study of mechanical mixing (with an impeller) in laboratory-


scale MBBRs included experiments that were conducted over the range from 50/s to 253/s (Bill
et al., 2010).
The high end of the batch test mixing rates used in this experiment (380/s) was based on
the range of values shown in Table 4-3. The minimum batch test value (158/s) was based on the
minimum value that provided good movement of the media in the reactors, as described above.
The G values and corresponding airflow rates evaluated in the batch tests are shown in Table
4-3. The initial mixing rate used in each continuous system was 240/s, and this was increased to
approximately 298/s after approximately one month of operation to reduce sloughing observed in
initial batch tests.
4.3 Startup and Routine Operation of Continuous Systems
R1 operation commenced on March 19, 2013. R2 operation commenced 37 days later
(April 23, 2013) to allow time for troubleshooting startup issues and to evaluate the proposed
reactor design and operation in R1. Prior to the startup of each reactor, the following inoculation
procedure was used to accelerate the formation of a biofilm on the media:
1. 3 L of media were incubated at room temperature for three days in a bucket containing
10 L of fresh activated sludge, with coarse-bubble mixing. The activated sludge was
obtained from the Albuquerque, New Mexico, Wastewater Reclamation Facility (WRF)
activated sludge system, which has a Modified Ludzack-Ettinger configuration.

4-8

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
2. The media were then transferred to the system reactor, along with approximately 5 L of
settled activated sludge supernatant (obtained from the activated sludge used for
incubation).
3. The reactor was then filled with primary effluent (also obtained from the Albuquerque
WRF), and continuous-system operation was initiated.
The reactors were run in three phases, which differed in their temperatures (Table 4-5). In
Phase 1, the reactors were run at 21°C in a temperature-controlled water bath, and after an
apparent steady state was reached in each system several batch tests were run, as described
below. An assumption of steady state was based on an observation of reasonably consistent
ammonia uptake at a constant surface area loading rate (SALR).
As noted, the target G value in each reactor was initially 240/s. This was increased to
298/s on April 26, 2013, in R1 and on May 24, 2013, in R2 to decrease biomass sloughing during
the batch tests by exposing the biofilms to higher shear during continuous-system operation.
During Phase 2, the temperature was decreased to 10.5°C in both reactors. The target G
value in each reactor remained 298/s. However, due to an operational error the G value in R1
ranged from 305/s to 327/s from September 2, 2013, to December 25, 2013, when it was returned
to 298/s. The potential impact of these higher-than-planned G values is discussed in Chapter 6.0,
Experimental Results.
During Phase 3, the temperature was returned to 21°C in both reactors. This was done to
assess whether the reactors could be restored to the SALRs applied during Phase 1, and to
conduct additional batch experiments, focused on the effects of bulk phase DO concentration on
mass transfer characteristics.

Table 4-5. Reactor Operation Phases.

R1 R2
Begin–end Begin–end
Phase 1: T = 21°C 3/19/2013-9/2/2013 4/25/2013-9/2/2013
Phase 2: T = 10.5°C 9/3/2013-1/13/2014 9/2/2013-1/22/2014
Phase 3: T = 21°C 1/13/2014-4/21/2014 1/23/2014-4/21/2014

Routine maintenance included cleaning of reactor walls with a brush every day to prevent
attached growth on surfaces other than the media. pH meters were checked and routinely
calibrated at least once every three weeks, and feed pump flow rates were checked at least every
two months, with little variation observed. Temperature was measured manually in the reactors
and the water bath temperatures were adjusted as necessary.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 4-9

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
4.4 Batch Tests
Batch tests were conducted on the media in each reactor to evaluate the relationship
between ammonia mass transfer and a variety of field parameters. The tests were conducted
within the continuous-flow reactors, but the system was cleaned and removed from continuous
operation for each experiment. Parameters evaluated during batch tests included:
 G.
 DO concentration.
 Temperature (set to the same as the continuous systems).
 Feed ammonia concentration.
Batch test pH was controlled between 7.15 and 7.50 in all experiments.
Up to six batch tests were conducted using the media from a single parent continuous-
system reactor in one day. For the first test of the series on a given day, all media were removed
from the reactor and were briefly exposed to air while the reactor walls were thoroughly cleaned.
The liquid phase was discarded to remove all suspended biomass and biofilm mass from the
reactor walls. The media pieces were returned to the reactor, which was filled with synthetic feed
to a total volume of 10 L using the recipe shown in Table 4-2. From this point, the reactors were
operated in batch mode to conduct experiments to evaluate impacts on mass transfer
characteristics of the media. The initial ammonia concentration was typically 20-30 mg NH4-
N/L, to ensure ammonia saturation but minimize ammonia inhibition. Samples were taken at
intervals from zero to 60 minutes. All samples were immediately filtered through a syringe filter
(25 mm, 0.45 micron [µm] nylon membrane), stored at 4°C, and analyzed within 24 hours).
Before subsequent batch tests on a given day, the reactor was mixed at the continuous-
system mixing rate (generally G = 298/s) for 20 minutes to remove attached biomass that may
have accumulated during the previous batch test. This also ensured that the conditions prior to
each batch test were similar. After 20 minutes of mixing, all liquid was removed from the reactor
to remove detached biomass. The reactor was then refilled with fresh feed, and the next test was
conducted. After all batch tests were completed on a given day, the reactor was refilled with feed
solution (Table 4-2), or with reactor liquid saved prior to batch testing. The former method was
used in Phase 1, and the latter method was used in Phases 2 and 3. The reason for this change is
discussed in Chapter 6.0, Experimental Results.
Early in the study, three batch tests were conducted in duplicate to evaluate the
reproducibility of results on a given day. This was done to evaluate the impact of the
experimental variables on subsequent tests conducted on the same day (e.g., biomass loss due to
increasing mixing intensity). Because these tests showed a high degree of experimental
reproducibility, the practice of conducting duplicate tests was discontinued in favor of increasing
the number of batch experiments.

4-10

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
4.5 Analytical Methods
This section describes the analytical methods applied, including biomass measurements
taken, nitrogen species analyzed and measured, DO, pH, data recording, and batch test data
analysis.
4.5.1 Biomass Measurements
Several methods of biofilm measurement were evaluated, which differed in their
approach toward detaching biomass from the carrier media. The methods evaluated were:
1. Sonication.
2. Rinsing.
3. Mechanical scraping/brushing.
Preliminary measurements indicated that biomass removed by sonication was less than
that obtained by rinsing or brushing, so this method was not evaluated further.
Biofilm removal by rinsing was used prior to July 1, 2013, and involved removing five to
10 media pieces from a given reactor prior to batch testing. The media pieces were thoroughly
rinsed using a 50 mL syringe with deionized (DI) water to remove as much biomass as possible.
The rinse water was saved and used to measure total suspended solids (TSS) and volatile
suspended solids (VSS) according to Standard Methods 2540B and 2540E, respectively
(American Public Health Association et al., 2012). The biofilm mass concentration in the reactor
was calculated as:
mg dry mass removed from media mg total media pieces
total biofilm solids, T ( ) E uation
L reactor wor ing volume L pieces of media tested

where the dry mass removed from media was determined from a TSS measurement of the rinse
water). The volatile biofilm solids (VBS) was calculated in an analogous manner using the VSS
measurement of the rinse water.
Based on conversations with the media donors in June 2013, the mechanical
scraping/brushing method was evaluated and compared against the rinsing method. Prior to batch
testing, 5 to 10 media pieces were taken from a given reactor, and each piece was thoroughly
cleaned by mechanical scraping and brushing using Proxabrush “Go- etweens” unstar
Americas, Chicago, Illinois, U.S.), which were developed for dental cleaning and are commonly
available in U.S. drug stores. The removed biofilm was collected by rinsing with DI water, the
TSS and VSS in the rinse water was measured as described above, and the total biofilm mass in
the reactor was calculated using Equation 4-2.
A side-by-side comparison of samples removed from the reactor at the same time
determined that the brushing method yielded approximately four (R1) to 10 (R2) times more
biomass than the rinsing method. This finding indicated that the rinsing method greatly
underestimated biomass concentrations. The brushing method was therefore applied for samples
from July 1, 2013, onward. All biomass measures reported in this document are based on the
brushing method.
The brushing method was also compared with an in situ measurement method provided
by a vendor in which media were rinsed, dried at 105°C to constant weight (approximately three
hours), weighed, acid-cleaned (pH 2), base-cleaned (pH 10), washed in DI water, dried again to

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 4-11

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
constant weight, and weighed. The biomass per media was calculated as the difference between
the pre-cleaned and cleaned weights. Three R2 media were measured by each method on each
date. Measurements of the R2 media on two different dates yielded no significant differences
between them (Table 4-6).

Table 4-6. Comparison Between Brush and On-Media Methods of Biofilm Measurement.

Date Brush Method, mg/media In situ Method, mg/media


1/15/2014 78.9 ± 17.7 69.7 ± 5.9
1/19/2014 59.9 ± 7.9 58.9 ± 5.4

4.5.2 Nitrogen Species


All samples for nitrogen species analysis were immediately filtered through a syringe
filter (25 mm, 0.45 µm nylon membrane, VWR catalog 28145-489), stored at 4°C, and measured
within 24 hours. All measurements of nitrogen species were performed using Hach kits with a
Hach DR2700 spectrophotometer (Hach Company, Loveland, Colorado, U.S.) as follows:
 NH4-N: Nitrogen-Ammonia Reagent Set, TNT, AmVer (Salicylate), High Range, Product
2606945.
 NO2-N: NitriVer 3 TNT Reagent Set, Nitrogen-Nitrite, Low Range, Product 2608345.
 NO3-N: NitraVer X Nitrogen-Nitrate Reagent Set, High Range, Product 2605345.
All its were used in accordance with the manufacturer’s instructions, except a correction
for potential nitrite interference with the nitrate measurement was started on January 6, 2014.
The manufacturer’s instructions note that nitrite interference could occur at nitrite concentrations
greater than 1 mg/L. Per the manufacturer’s instructions, this interference was removed by
adding 400 mg urea to 10 mL of sample. Nitrate concentrations measured prior to January 6,
2014, may therefore overestimate the actual values.
4.5.3 Dissolved Oxygen
DO was measured using a Hach IntelliCAL LDO101 standard luminescent/optical DO
probe with a Hach HQ440d multi-parameter meter (Hach Company, Loveland, Colorado, U.S.).
4.5.4 pH
pH was controlled in the range of 7.15 to 7.50. Further details are provided in
Section 4.2, Continuous-System Reactor Operation.
4.5.5 Data Collection and Recording
Temperature, DO, and pH data were collected and recorded manually.

4-12

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
4.5.6 Batch Test Data Analysis
The batch testing data consisted of ammonia, nitrate, and nitrite concentrations measured
at specific time intervals. For each test, the average temperature, mixing rate, and DO
concentration were noted. Figure 4-6 presents the results of a typical batch test.

Figure 4-6. Typical Batch Test Result.


Test conducted in R2 at 10.5°C at DO concentration of 6.8 mg/L and G of 238/s (one of six tests on 9/29/2013).

The average rate of ammonia oxidation was estimated by a best-fit linear regression over
the course of the batch experiment. In this example, the reaction rate is the slope of best-fit line
(0.300 mg/L/min). Batch test flux rates were normalized to media surface area in the reactor,
according to the following transformation:
d 1 1
E uation
dt A F
Where
J = mass flux relative to media surface area (g/m2/d)
SSA = media specific surface area (m2/m3)
F = media fill percentage (unitless)
C = ammonia concentration (g/m3)
t = time (day)

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 4-13

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
4-14

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
CHAPTER 5.0

MODEL CONCEPT
One of the goals of this experiment was to develop a model to allow for relatively simple
evaluation of media for MBBR applications. In keeping with this goal, model development
began with Fic ’s First Law describing mass transport by diffusion:

D E uation
x
Where
J = flux (mass/area/time)
D = diffusion coefficient (area/time)
C = concentration (mass/volume)
x = position (length)

For a biofilm system, the limitation to mass transfer is often expressed by the concept of
a mass transfer boundary layer (MTBL, Figure 5-1).

Figure 5-1. Schematic Representation of Mass Transfer from Bulk Liquid to a Biofilm.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 5-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
The MTBL theory suggests that substrate must diffuse through the boundary layer in order
to be metabolized by cells within the biofilm. Much of the literature on biofilm processes has
focused on defining the differential term, with models for biofilm growth, decay, and estimates of
both the biofilm and boundary layer thickness (Eberl et al., 2006; Henze et al., 2008). This
investigation moves away from MTBL-based modeling, and aims to develop an empirical model
for ammonia uptake in an MBBR system, which depends upon criteria that can be measured
directly, using relatively simple tools and widely available methods.
5.1 Batch Test Observations
The batch testing protocol measures the change in ammonia concentration over time. The
experimental data are therefore in the form of a reaction rate:

R E uation
dt
Where
R = reaction rate (mass/volume/time)
C = concentration (mass/volume)
t = time (time)
Rates can also be expressed as fluxes (rates per biologically active media surface areas)
to account for the amount of media in the reactor:
1 1 1 1
R E uation
dt A F A F
Where
J = mass flux relative to media surface area (mass/area/time)
SSA = media specific surface area (m2/m3)
F = media fill percentage (unitless)

5.2 Model Format


The model equation seeks to predict the media-specific reaction rate, given inputs of
solute concentrations, temperature, and mixing. Theoretically, these inputs relate bac to Fic ’s
Law with the solute concentrations representing the magnitude of the concentration gradient
across the boundary layer, and the mixing rate, which influences the size of the boundary layer.
Temperature has been added to account for the influence of temperature on bacterial reaction
kinetics.
With such inputs, the model equation must be of the following form:
G T E uation
Where
J = mass flux relative to media surface area (mass/area/time)
k = media coefficient
f(C) = function of solute concentrations
f(G) = function of mixing, as characterized by the velocity gradient, G
f(T) = function of reactor temperature

5-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
5.3 Model Sensitivity Testing
The solute concentration function is driven by the most limiting of the solutes involved in
the metabolic reaction under consideration. As this experiment focused on ammonia oxidation,
the limiting solute could be either ammonia or oxygen.
Batch experiments were conducted across a range of solute concentrations to determine
which of the two solutes was limiting. Figures 5-2 and 5-3 present the data from these experiments.

Figure 5-2. Reaction Rate versus Bulk Ammonia Concentration.


G = 298 sec-1, T = 21C, DO = 5.8 mg/L, 8/12/13.

Figure 5-3. Reaction Rate versus Bulk DO Concentration.


11/4/13: T = 10.5C, G = 297 sec-1, CNH3N, bulk = 13-36 mg/L.
11/18/13: T = 10.5C, G = 297 sec-1, CNH3N, bulk = 16-34 mg/L.
1/15/14: T = 10.5C, G = 297 sec-1, CNH3N, bulk = 0.6-28 mg/L.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 5-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
The data suggested that, for these experiments, oxygen was the limiting solute. While there
was little change in reaction rate over a range of bulk phase ammonia concentrations, the reactions
rates were dependent on the bulk phase DO. The relationship between DO and reaction rate
suggested either a half-order or a saturation-type (Monod) response. Based on these findings, the
model equation was revised to incorporate a half order response to bulk DO concentration:
√ G T E uation
Where
CDO,bulk = the bulk phase dissolved oxygen concentration (mass/volume)

The selection of a half-order term for this equation is based upon observations from Harremoes
(1977), augmented in Henze (2008), which support use of half order rate equations within
partially-penetrated biofilms. The expectation is that the relationship would shift to zero order at
high concentrations, which are not likely to be observed in practice.
A more universal form of Equation 5-5 would include both substrate response terms:
√ √ G T E uation
Where
CNH3N,bulk = the bulk phase ammonia-nitrogen concentration (mass/volume)
As this experiment was conducted under relatively high bulk phase ammonia concentrations
(within the range of concentrations tested in Figure 5-2), the ammonia response is assumed to be
zero order, and will be omitted for the remainder of this report.
Batch tests conducted at a constant bulk phase DO concentration demonstrated that the
effect of G on the ammonia uptake rate also resembled either a half-order or saturation-type
response (Figure 5-4).

Figure 5-4. Ammonia Flux versus velocity gradient at Constant DO (7.5 mg/L).
12/1/13: T = 10.5C, DO = 7.5-7.8 mg/L, CNH3N, bulk = 6-34 mg/L.
12/20/13: T = 10.5C, DO = 7.3-7.7 mg/L, CNH3N, bulk = 7-31 mg/L.
1/19/14: T = 10.5C, DO = 7.3-7.7 mg/L, CNH3N, bulk = 9-31 mg/L.

5-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
For consistency, and to avoid the inclusion of an additional model parameter (a half
saturation term), a half-order term was adopted to model the effect of G:
√ √ T E uation
Where
G = velocity gradient in reactor (1/time)

5.4 Model Equation


Temperature was assumed to have an Arrhenius-type effect on the reaction rate, with the
addition of temperature-dependence term:
T
√ √ E uation
Where
 = Arrhenius constant
T = temperature (°C)
This is the final form of the empirical model equation. Note that the parameter k has units
of flux (mass/area/time) in this equation. Determination of the model parameters k and  requires
measurement of bulk ammonia and DO concentrations throughout the batch experiments with
variable DO, G, and T values. The model parameters can then be derived through least-squares
regression analyses.
The model will output a media-specific reaction rate with units of mass/area/time. This
rate is analogous to the ammonia removal rate (RR). Media-based bioreactor systems are often
designed on the basis of the SALR. The design SALR is typically equal to the maximum loading
rate at which the SALR remains linearly related to the RR. This relationship is discussed in detail
in several of the foundational papers on MBBR and IFAS processes (Ødegaard et al., 1999,
Æsøy and Ødegaard, 1994, Rusten et al., 1995, Ødegaard et al., 2000, Johnson et al., 2004). By
developing an empirical model to predict the RR at specific operating conditions, a reasonable
next step would be to translate model findings to a design scenario. For IFAS systems this would
require a bulk phase model to be used in concert with the media-based model. And the model
would likely require modification to account for the very low substrate concentrations targeted in
treatment systems. In a general sense, however, the application would take the following form:

Loading rate
edia re uirement E uation

Where
Media requirement = media surface area (area)
Loading rate = ammonia loading (mass/time)
J = mass flux relative to media surface area (mass/area/time)

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 5-5

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
5-6

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
CHAPTER 6.0

EXPERIMENTAL RESULTS
As described in Chapter 4.0, Methods, the continuous systems were run in three phases:
 Phase 1: Temperature = 21°C.
 Phase 2: Temperature = 10.5°C.
 Phase 3: Temperature = 21°C (conducted to collect additional batch test data for the Phase 1
conditions).
The following narrative summarizes continuous-system reactor performance during these
three phases, as well as batch tests. The purpose of the continuous systems was to provide highly
active biofilms for use in batch tests, which in turn provided data for model development. It is
emphasized that the continuous-system data were not used in model development.
6.1 Continuous-Flow Reactors
This section discusses the results observed in the continuous-flow reactors.
6.1.1 Phase 1 (21°C)
During Phase 1, both reactors were incubated at 21°C. The following sections discuss
results from Phase 1.
6.1.1.1 Reactor 1
R1 performance is shown in Figure 6-1. Figure 6-1A shows the ammonia fed to the
reactor as influent concentration on the left axis, and as SALR on the right axis. Note that these
two quantities can be shown with the same line using different axes because they are linearly
related by the equation:
influent ammonia concentration
ALR E uation
reactor wor ing volume fill percentage media specific surface area

Similarly, ammonia uptake (mg N/L) is shown on the left axis and ammonia flux to the
media (J, in terms of g N/m2/d) is shown on the right axis.
R1 was inoculated on March 27, 2013 (Figure 6-1). The initial operating strategy was to
gradually increase the feed ammonia concentration, beginning at 44 mg N/L, to keep the effluent
ammonia concentration between 10 and 50 mg N/L. As the biofilm developed on the media,
influent concentration was stepwise increased to 173 mg NH4-N/L through April 29, 2013,
during which time the ammonia uptake closely tracked the influent ammonia, suggesting that
AOB activity was ammonia-limited. From May 2-4, 2013, the influent ammonia was increased
from 173 to 293 mg N/L in an effort to accelerate biofilm development. Ammonia uptake
subsequently decreased and effluent ammonia concentrations increased (Figure 6-1).

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 6-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 6-1. Reactor 1 Performance.
(A) Influent ammonia and ammonia uptake (left axis), also expressed as SALR and J (right axis).
(B) Effluent nitrate and nitrite. (C) Effluent ammonia. Note that during Phase 2,
R1 mixing rates were inadvertently increased, which may have contributed to observed changes in activity.

6-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
After approximately one month of operation during which R1 performance failed to
return to earlier levels, influent ammonia was decreased to 190 mg N/L on June 12, 2013, in an
effort to keep effluent ammonia concentrations relatively low, to avoid nitrification inhibition.
The system responded favorably after this change, allowing a steady increase in ammonia uptake
until it reached approximately 385 mg N/L by July 5, 2013, and leveled off thereafter. Although
effluent ammonia concentrations occasionally increased above target levels, inhibitory levels of
ammonia as predicted by Anthonisen et al. (1976) were not observed (at T = 21°C and pH = 7.5,
ammonia-based inhibition would occur at concentrations greater than 620-9,300 mg N/L;
Appendix C). Similarly, although high nitrite concentrations were observed (up to 200 mg N/L
per Figure 6-1), nitrite-based inhibition would not be expected, per calculations based on
Anthonisen et al. (1976). Furthermore, nitrite concentrations generally increased during Phase 1,
concurrent with generally increasing ammonia uptake, so AOB inhibition by nitrite was not
indicated. Further discussion of potential inhibition is provided in Appendix C.
Nitrification was initially complete in R1, with nitrate concentrations nearly matching
ammonia uptake. Nitrite measurements commenced on May 9, 2013, and concentrations
increased steadily relative to nitrate concentrations, indicating decreasing NOB activity by the
end of Phase 1. The reason for nitrite accumulation throughout the study is not known, but NOB
inhibition by free ammonia (NH3) was possible during much of the study. Based on the work of
Anthonisen et al., (1976), at the high end of the controlled pH range (7.5), NOB inhibition would
be predicted at a total ammonia concentration of 6.2 to 62 mg N/L at T = 21°C (Phases 1 and 3)
and 14 to 140 mg N/L at T = 10.5°C (Phase 2). See calculations and discussion in Appendix C,
including a discussion of other studies of nitrification inhibition by ammonia and nitrite.
Figure 6-1 also shows occasional acute decreases in apparent ammonia uptake and
effluent nitrate and nitrite effluent concentrations (e.g., on July 10, 2013, July 24, 2013, and
August 14, 2013), followed by periods of recovery over several days. These events occurred
immediately after batch tests (batch test dates are marked on the horizontal axis in Figure 6-1A),
and it was later theorized that they were due to the practice of refilling the reactor after each
batch test with fresh feed solution. The feed consisted of a high ammonia concentration relative
to the solution removed from the reactors before the batch tests, with zero nitrate and nitrite
concentrations. This effectively applied a pulse ammonia load to the reactors, while removing all
accumulated nitrate and nitrite. The result was increased effluent ammonia and decreased
calculated ammonia uptake, nitrate, and nitrite concentrations. It typically took the reactor
several days to equilibrate back to a steady state. For this reason, the apparent disruptions of
continuous system performance after these batch tests was likely an artifact of the procedure
used. The batch test procedure was modified once this was recognized (after Phase 1 in both
reactors) to removing and saving the reactor liquid phase before batch testing on a given day, and
returning it to the reactors after batch tests were complete and before restarting the continuous
system.
It is worth emphasizing that this study included the use of higher influent ammonia
concentrations (up to 433 mg NH3-N/L in R1, Phase 1) than found in domestic wastewater, and
one consequence of this was that disruptions in reactor performance could greatly increase bulk
phase and effluent ammonia concentrations (in addition to the increases in effluent ammonia
concentrations due to the batch testing procedure described in the previous paragraph). Coupled
with the possibility that ammonia may have inhibited NOB activity (thereby increasing nitrite
concentrations), the high observed nitrite concentrations in R1 (and also in R2, as described

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 6-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
below) were perhaps not surprising. As described in Appendix C, high nitrite concentrations may
have also contributed to NOB inhibition.
6.1.1.2 Reactor 2
R2 was inoculated on April 25, 2013, approximately one month after R1 (Figure 6-2).
Initial ammonia loadings were increased more rapidly in R2, based upon experience with R1,
which had exhibited ammonia-limitation during the first few weeks of startup. After 30 days of
operation, the R2 feed ammonia concentration had been increased to 304 mg N/L. R1 feed, in
contrast, had increased to only 166 mg N/L after 30 days of operation.
Two months after startup, R2 activity began to plateau, with ammonia uptake of
approximately 500 mg/L. Feed concentration was held between 500 and 560 mg N/L, and
effluent ammonia concentrations increased from less than 10 mg N/L prior to July 3, 2013
(ignoring the spikes related to batch testing) to greater than 20 mg N/L thereafter. Based upon
the effluent concentrations, ammonia was not believed to be limiting reactor performance and it
was assumed that maximum AOB activity had been achieved.
As was the case in R1, initial nitrate concentrations in R2 were nearly equal to the
ammonia uptake rate. During the first month of operation, R2 exhibited a gradual increase in
nitrite concentration, also similar to that observed in R1. However, while R1 nitrate
concentrations were consistently higher than nitrite concentrations, the opposite was observed in
R2. By the end of Phase 1, R2 nitrite and nitrate concentrations averaged approximately 300 and
160 mg N/L, respectively.

6-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 6-2. Reactor 2 Performance.
(A) Influent ammonia and ammonia uptake (left axis), also expressed as SALR and J (right axis).
(B) Effluent nitrate and nitrite. (C) Effluent ammonia.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 6-5

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
6.1.2 Phase 2 (10.5°C)
During Phase 2, both reactors were operated at a lower temperature to provide additional
data for model development. The temperature was decreased in both reactors after sampling on
September 2, 2013.
6.1.2.1 Reactor 1
The effluent ammonia concentration increased to 80-90 mg N/L shortly after the reactor
temperature was decreased to 10.5°C to start Phase 2 (Figure 6-1). The feed ammonia
concentration was decreased in order to compensate for this observation. Over a two-month
period, the feed ammonia concentration was decreased from 400 mg N/L to 100 mg N/L.
However, the effluent ammonia concentration remained higher than the target of 10-50 mg N/L.
During this time, effluent nitrate and nitrite concentrations decreased to less than 20 mg N/L.
Toward the end of Phase 2, the effluent ammonia concentration decreased to an average of 20
mg N/L and stabilized. This was achieved with a feed ammonia concentration of approximately
90 mg N/L, reflecting an ammonia removal of less than 75%. The nitrite concentration increased
to approximately 60 mg N/L at the end of Phase 2, with much lower concentrations of nitrate
(less than 10 mg N/L).
Due to a change in the airflow meter used and an error in reading it, the G value during
Phase 2 for R1 was unintentionally increased to 327/s, instead of the 298/s target value used in
Phase 1 (a 10% increase). The mixing rate in R1 was decreased to 305/s on December 14, 2013,
and it was further decreased to 298/s on December 26, 2013.
The biomass measured in R1 decreased from 645 mg/L (total biofilm solids [TBS]) on
August 19, 2013, to 520 mg/L on September 26, 2013 (Phase 2 commenced on September 2,
2013). Throughout Phase 2, the biomass in R1 continued to decrease, reaching 41 mg/L by mid-
December.
During the first two months of Phase 2, the average effluent R1 ammonia concentration
was 71  17 mg N/L, while in the last two months of Phase 1, the average R1 ammonia
concentration was 37  31 mg N/L. While ammonia-related AOB inhibition is possible, it would
be unlikely, based on the analysis of Anthonisen et al. (1976). Appendix C has details of the
relevant inhibition calculations. Ammonia removal improved starting in late November, 2013,
with R1 ammonia decreasing to less than 20 mg N/L by the end of Phase 2.
The role of the increased mixing rate on R1 performance is not known, but it is possible
that it played a role in the deterioration of AOB and NOB activity, by increasing shear and
potentially increasing sloughing of the biofilm. Performance improved near the end of Phase 2,
although biomass remained low.
6.1.2.2 Reactor 2
R2 ammonia uptake and nitrate production decreased immediately after the temperature
was decreased to begin Phase 2 (Figure 6-2), demonstrating decreasing AOB and NOB activity.
NOB activity was particularly affected, with a decrease in nitrate concentrations from
approximately 200 mg N/L at the end of Phase 1 to 13 mg N/L on November 14, 2013. In
response to this reduction, the influent ammonia was decreased on November 18, 2013 (from
379 mg N/L to 307 mg N/L) to decrease the possibility of NOB inhibition by ammonia. This had
the desired effect of decreasing effluent ammonia (from 72 mg N/L on November 18, 2013, to
1.8 mg N/L on December 1, 2013), but reactor nitrate concentrations continued to decrease,
approaching zero by the end of Phase 2.

6-6

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
6.1.3 Phase 3 (Return to 21°C)
During Phase 3, the reactors were returned to 21°C in order to supplement high-
temperature batch testing conducted during Phase 1.
6.1.3.1 Reactor 1
Ammonia uptake increased in R1 after the Phase 3 temperature increase (Figure 6-1), but
did not return to the levels of uptake observed during Phase 1. After January 24, 2014, R1
ammonia uptake began to decrease, and this trend continued until the reactor was reinoculated on
February 23, 2014 (see below). Effluent ammonia concentrations increased to 200 mg N/L
during this period, and the reactor feed concentration was decreased to reduce the risk of
ammonia-based inhibition.
Biomass measurements in R1 demonstrated only a small increase in solids in Phase 3,
from 41 mg/L on December 16, 2013, to 56 mg/L on February 12, 2014. R1 media was
reinoculated on February 17, 2014, in an effort to accelerate its recovery. The media were
incubated in a solution of mixed liquor from the Albuquerque WRF, using the same protocol
employed at the start of the experiment. Reactor activity increased following reinoculation, but
feed ammonia concentrations could not return to Phase 1 levels without resulting in high effluent
ammonia concentrations. The SALR applied at the end of Phase 3 was 1.75 g N/m2/d, compared
to 2.0 g N/m2/d applied at the end of Phase 1. Although an increase in biomass was observed
following reinoculation (e.g., 140 mg /L on March 13, 2014), this concentration was well below
levels observed in Phase 1 (645 mg/L).
6.1.3.2 Reactor 2
The feed ammonia concentration to R2 was increased from 300 to 425 mg N/L at the start
of Phase 3, and ammonia uptake increased steadily thereafter. Although ammonia uptake
increased, the effluent ammonia concentration also increased, averaging 70 mg N/L in the first
month of Phase 3. The SALR at this time averaged 2.1 g N/m2/d, which was lower than the
SALR applied at the end of Phase 1 (2.70 g N/m2/d).
Visual inspection of the media determined that approximately 40% of the R2 media
contained little to no biomass in early Phase 3. It was hypothesized that these were media that
had been removed and cleaned for biomass measurements, and that biomass had not re-grown on
them over the course of the study. It may have been that the inoculation procedure (incubating
the media in activated sludge for several days) at the beginning of the study was critical for
priming of the media for biofilm growth, while use of the zero organic carbon synthetic feed
contributed to a poor environment for biofilm regrowth on the media. For this reason, the media
with no attached biofilm were removed from R2 and they were reinoculated with activated
sludge according the same protocol used at the beginning of the study on February 17, 2014.
Plastic media in R2 with significant visible biomass were not reinoculated.
The media with little visible biomass that were reinoculated increased from
1.60 mg/media to 54.3 mg/media 30 days after reinoculation. The 10 media without visible
biomass were marked and returned to the reactors without reinoculation. There was comparably
little increase in biomass on these media (from 1.60 mg/media to 5.8 mg/media after 30 days).
These results confirmed the hypothesis that biofilm regrowth was very slow on cleaned media
returned directly to the reactor after biomass measurement, and it is therefore recommended in
future work that media removed for biomass measurements be replaced with inoculated media.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 6-7

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
After reinoculation, R2 ammonia uptake quickly increased, although it did not return to
the Phase 1 levels. The SALR at the end of Phase 3 remained near 2.1 g N/m2/d, with effluent
ammonia concentrations averaging 13  16 mg N/L during the final month of Phase 3. Notably,
nitrate concentrations, which had remained near zero through the first month of Phase 3,
increased steadily after reinoculation, with concurrent reduction in the nitrite concentration.
6.1.4 SALR and Flux (J)
The relationships between SALR and ammonia flux (J) are shown for R1 and R2 in
Figures 6-3 and 6-4, respectively.

Figure 6-3. R1 Specific Ammonia Loading Rate (SALR) Plotted Against


(A) Ammonia Flux (J) and (B) Effluent Ammonia Concentration.

6-8

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 6-4. R2 Specific Ammonia Loading Rate (SALR) Plotted Against
(A) Ammonia flux (J) and (B) Effluent Ammonia Concentration.

The 1:1 lines in these figures indicate the theoretical complete removal of influent
ammonia, and the vertical distance of data points below the 1:1 line represents the degree to
which ammonia flux lagged behind the SALR. While J generally increased with SALR, these
figures illustrate that ammonia uptake was incomplete, particularly at higher SALR values. The
Phase 1 data in Figures 6-3A and 6-4A show that the progression during startup as the SALR
was steadily increased. The lower temperature Phase 2 data are shifted to the left and downward,
as expected from the diminished performance of each reactor discussed above.
The ammonia flux rates shown in Figures 6-3 and 6-4 are greater than those by Johnson
et al. (2004), who reported data from a pilot-scale IFAS system operated at 14°C. For example,
at an SALR of 2 g N/m2/d, J ranged from approximately 0.8 to 1.65 g N/m2/d in Johnson et al.
(2004); while at this SALR the R1 J values ranged from approximately 1.1 to 2.0 g N/m2/d, and
in R2 the J values ranged from approximately 1.7 to 2.0 g N/m2/d (at T = 21°C). Maximum J

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 6-9

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
values in Johnson et al. (2004) were approximately 1.65 g N/m2/d (at T = 14°C), while the R1
and R2 maximum values were 2.07 and 2.93 g N/m2/d, respectively (T = 21°C). These results
indicate that the project goal of achieving biofilms with very high nitrification activities had been
achieved.
Figure 6-3B shows the R1 effluent ammonia concentrations as the SALR was varied.
This figure shows typically high effluent ammonia concentrations, so it was unlikely that
ammonia was limiting AOB growth rates, particularly at maximum SALR values in each phase.
Comparison of Figures 6-3B and 6-4B illustrates the relative difficulty in maintaining target
effluent ammonia concentration in R1 relative to R2, particularly during Phase 2.
6.2 Batch Tests
This section describes the batch tests that were performed as part of this study.
6.2.1 Reproducibility of Batch Tests Performed on a Single Day
Initial batch tests were conducted on R1 to assess the replicability of the batch test
protocol. On April 30, 2013, and May 13, 2013, replicate mixing rates were tested, with G values
in the order 298/s, 240/s, 180/s (series 1); and 180/s, 240/s, and 298/s (series 2). The results are
shown in Figure 6-5, which validated the reproducibility of the batch testing protocol for tests
done on a single day.

Figure 6-5. Reproducibility of Batch Test Data.

6.2.2 Correlation between Mixing Rate and DO Concentration


Because mixing energy was supplied through aeration, there was a correlation between
mixing rate and the bulk phase DO concentration in tests where the DO was not specifically
controlled, using atmospheric air alone for aeration (in controlled DO tests, oxygen or nitrogen
gas was added as necessary to meet target DO concentrations). Prior to September 29, 2013, DO
was not controlled and DO measurements were not specifically recorded during each batch test.

6-10

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Because the bulk phase DO concentration is part of the model equation (Equation 5-8), the DO
concentrations for batch tests prior to September 29, 2013, had to be estimated. As noted in the
Recommendations section, future work application of these protocols should include DO control
with nitrogen or oxygen addition for systematic variation of the DO concentration during batch
tests. At a minimum, DO concentration should be monitored and recorded for each test.
The correlation between bulk phase DO concentration and the mixing rate is plotted in
Figure 6-6. The correlation appears consistent across time, and across reactors.

Figure 6-6. Empirical Relationship Between Mixing Rate and Reactor Dissolved Oxygen Concentration.

6.2.3 Batch Test Results


Batch tests were conducted in each reactor, at temperatures of 10.5°C and 21°C. Each test
was conducted at a specific mixing rate and bulk phase DO concentration. On any given day, a
range of DO concentrations or mixing rates may have been tested. A total of 200 batch tests were
conducted, 79 in R1 and 121 in R2. Batch testing during Phase 2 in R1 was largely discontinued
after October 2013 given the very low reactor activity and the loss of biomass in the reactor
(Figure 6-1). Figure 6-7 plots the timing and number of batch tests conducted in each reactor
over the course of the experiment.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 6-11

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 6-7. Number and Timing of Batch Tests in Each Reactor.

Appendix D presents all of the batch test results from this experiment. Batch test results
are plotted in terms of ammonia flux (J) versus mixing rate (G), with the average DO
concentration expressed in terms of color. Figures 6-8, 6-9, 6-10, and 6-11 present the batch test
data for R1 at 21°C, R1 at 10.5°C, R2 at 21°C, and R2 at 10.5°C, respectively. Batch tests prior
to June 17, 2013 have been omitted as neither reactor had attained a stable level of performance
prior to that time (Figures 6-1 and 6-2).

Figure 6-8. Batch Test Results for Reactor 1 at 21°C.

6-12

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
In general, the batch tests suggest the two half-order responses expressed in the model
equation (Equation 5-8). The data from R1 at 10.5°C (Figure 6-9) reflect relatively low flux rates
(0.5 g/m2/d) and little relationship to either mixing rate or DO concentration. This correlates to
the reduction in biomass in R1 during Phase 2, and the reduced activity noted in the continuous-
system reactor during that time.

Figure 6-9. Batch Test Results for Reactor 1 at 10.5°C.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 6-13

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 6-10. Batch Test Results for Reactor 2 at 21°C.

Figure 6-11. Batch Test Results for Reactor 2 at 10.5°C.

6-14

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
CHAPTER 7.0

MODEL EVALUATION
The intention of this work is to provide a means to compare different media types,
and, ultimately, to develop a method to test and develop site- and media-specific design
criteria for media-based reactors. As described above, an empirical model for determining the
ammonia reaction rate was developed (Equation 7-1):
T
√ √ E uation 1
where:
J = mass flux relative to media surface area (mass/area/time)
k = media coefficient (mass/area/time)
CDO, bulk = bulk phase DO concentration (mass/volume)
G = velocity gradient (1/s)
: a unitless temperature-dependence constant
T = temperature (°C)

The model equation was fit to the data from 29 batch tests in R1 and 92 batch tests in
R2 (data are provided in Appendix D). The R1 data were limited to data at 21°C. At 10.5°C,
the reactor began to shed its biofilm, and nitrification activity reduced to approximately 10%
of the activity at 21°C. As discussed in the previous section, the reactor continued to lose
biomass throughout Phase 2. R1 was therefore viewed as a failed reactor at 10.5°C, and the
batch test data in that phase were not used for parameter calibration.
The model equation (Equation 7-1) expresses flux as a function of the mixing rate (G),
the bulk phase DO concentration, and the reactor temperature. The equation includes two
constants:
 k: the media coefficient.
 : the temperature-dependence constant.
Values for each of the constants were obtained through non-linear regression of the
batch test data using Equation 7-1. A single regression was applied to all of the data, across
both reactors, allowing for different media coefficients for each reactor (k1 and k2), over the
entire period or record.
The temperature dependence constant was intended to be fixed across all media. In
this case, because the 10.5°C data for R1 were not modeled, the constant was calibrated
against R2 data, and applied uniformly to both reactors.
Figures 7-1 through 7-3 depict calibrated model fits to R1 at 21°C, and R2 at 10.5°C
and 21°C, respectively. The model predictions are shown with colored lines. Batch test data
are expressed as points, similar to the way they were expressed in Figures 6-7 through 6-10.
The model fits represent a best-fit regression to all of the data.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 7-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 7-1. R1 Batch Test Data with Model Fit at 21°C.

Figure 7-2. R2 Batch Test Data with Model Fit at 21°C.

7-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 7-3. R2 Batch Test Data with Model Fit at 10.5°C.

A summary of the calibration parameters and goodness-of-fit parameters is provided


in Table 7-1. Correlation plots between observed and model flux rates are presented in
Figures 7-4 and 7-5.

Table 7-1. Summary of Model Parameters and Goodness-of-Fit.

R1 R2
k 10.78 12.85 g/m2/d
θ 1.025 1.025
N 29 104
SSR 0.547 4.082 g/m2/d
RMSE 0.137 0.198 g/m2/d
Jobs mean 1.83 2.01 g/m2/d
CV 7% 10%
R2 0.883
k = media coefficient; θ = Arrhenius constant; N = number of batch tests; SSR = sum of
squared residuals; RMSE = root-mean-squared error; Jobs Mean = the average observed
ammonia flux rate across all batch tests; CV = coefficient of variation = RMSE/ Jobs Mean; R2
= correlation coefficient (considers both reactors)

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 7-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 7-4. Model versus Observed Ammonia Flux, Reactor 1.

Figure 7-5. Model versus Observed Ammonia Flux, Reactor 2.

7-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
A parameter analysis (Table 7-2) suggests that all of the parameters are
statistically significant.

Table 7-2. Statistical Analysis of Model Parameters.

Coefficient Std. Error t P


K1 10.78 0.204 52.78 <0.0001
K2 12.85 0.153 84.00 <0.0001
 1.0251 0.0017 592.3 <0.0001

An analysis of covariance (Table 7-3) shows very little correlation between the
three parameters.

Table 7-3. Covariance Matrix for Model Parameters.

K1 K2 
K1 0.0417
K2 -0.0019 0.0234
 -0.000031 0.00020 0.0000030

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 7-5

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
7-6

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
CHAPTER 8.0

FINDINGS
The empirical model was calibrated to two different media types. Batch test observations
from a total of 133 tests were fit to the model, with an average root-mean-squared error (RMSE)
of less than 0.19 g/m2/d, representing a coefficient of variation of less than 10%. The model was
designed to project media-specific reaction rates given inputs of temperature, mixing rate (as
expressed by the velocity gradient, G), and bulk phase DO concentration.
8.1 Model Equation
The model equation was constructed based upon a number of assumptions, discussed in
detail in Chapter 5.0 of this report. The following sections discuss the results of the batch
experiments in the context of these assumptions in some detail.
8.1.1 Bulk Phase Ammonia
Much of literature focuses on the relationship between bulk phase DO and reaction rate.
Ødegaard et al. (1999) present idealized plots of bulk phase ammonia concentration versus
nitrification rate, which exhibit a staged (half order / zero order) or saturation-type response that
levels off at ammonia concentrations in the 0-3 mg N/L range. Rusten et al. (1995) present actual
data comparing bulk phase ammonia to nitrification rate, which also show this response, leveling
off at low (4-6 mg N/L) ammonia concentrations. In this experiment, bulk phase ammonia
concentrations were typically higher, as the batch experiments were designed to maintain an
excess of ammonia in solution. The reaction rate is typically determined by whichever reactant is
limiting. In this case, the limiting reactant was oxygen, as indicated in Figures 5-2 and 5-3. It is
probable that, had the batch tests been conducted at lower ammonia concentrations, ammonia
would have become limiting, and a saturation-type or half-order response would have been
observed. This may be important for design applications targeting low effluent ammonia
concentrations. In such cases, the model equation may need to be revised to substitute the DO
response term with an ammonia response, or to allow for both responses, as expressed in
Equation 5-6.
8.1.2 Bulk Phase DO
Researchers have typically observed linear relationships between the bulk phase DO
concentration and the nitrification rate in media bioreactors. Ødegaard et al. (1995) projected a
linear relationship in their developmental paper, with Hem et al. (1994) presenting supporting
data over a range of DO concentrations from 1-15 mg/L. Johnson and McQuarrie (2002) noted a
linear relationship between the total Kjeldahl nitrogen (TKN) removal rate and the bulk phase
DO over a DO concentration range from 2.5 to 6.0 mg/L, and Jones et al. (2010) noted linear
relationships at DO concentrations up to 8.5 mg/L. Across all of these studies, however, the
amount of data taken at DO concentrations higher than 10 mg/L is limited. Similar to these other
studies, the data in this experiment also appear linear at DO concentrations up to 10 mg/L. The
data collected at DO concentrations of 15 and 20 mg/L are those which drive the half-order
relationship.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 8-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
8.1.3 Mixing Rate
The relationship between reaction rate and G was clearly shown in Jones et al. (2010),
with the supposition that high mixing rate led to a reduction in the size of the mass transfer
boundary layer, and therefore increased the rate of substrate flux to the biofilm. Thomas (2009)
also noted some influence of mixing on reaction rate, specifically that the relationship with bulk
phase DO concentration appeared to be more important at lower mixing speeds. These previous
studies limited testing to a small number of mixing rates, and mixing rate assessment was only a
secondary part of the work. This experiment suggests a half-order or saturation response between
mixing and reaction rates (Figure 5-4). The response may be explained in terms of advection—as
higher G creates higher velocities through the media, and a higher rate of advection to the
biofilm. This allows for higher mass flux to the biofilm, and a higher reaction rate provided
sufficient biomass and reaction kinetics to metabolize the substrate. As the mixing rate becomes
higher, either the relative difference in mixing-induced advection becomes smaller, or reaction
rates and biomass become more limiting than the rate of mass transfer.
The continuous system results show how high mixing rates can adversely affect
performance through biofilm shearing. This suggests that the model equation should only be
applicable at G values lower than a presumed Gmax, above which shearing impacts become more
dominant.
The response observed with the two media in this experiment did not support Thomas’s
(2009) assertion that the relationship between reaction rate and bulk phase DO concentration was
more profound at lower mixing speeds. The model equation treats the two as independent
variables, and the model fit supports that construction.
8.1.4 Temperature
An Arrhenius-type temperature dependence was applied to the model equation.
Unfortunately, the failure of R1 at low temperatures did not allow for the temperature
dependence to be tested in multiple reactors. Within R2, the Arrhenius term allowed for a good
model fit at the two temperatures tested. Biofilm models incorporate such temperature correction
factors into the growth rates of bacteria, as well as to system performance (the modified Velz
equation for trickling filter performance, Parker and Merrill (1984); as well as the Howland
model, Howland (1958). The derived value (1.025) is lower than values typically used to
characterize AOB growth rates (1.072 is used in the BioWin Process Simulator, v4.0, Envirosim
Associates, Ltd, Hamilton, ON). However, the basis of measurement in this work is based on
surface reaction rates rather than reactions based on the mass of organisms, as in the activated
sludge models. Similar findings with respect to lower declines with temperature have also been
found with nitrifying trickling filters when compared to activated sludge plants (Parker et al.,
1995).
8.1.5 Partial Nitrification
Over the course of the experiments, only a portion of the ammonia was fully oxidized to
nitrate. The majority of the oxidized nitrogen was in the form of nitrite. Under cold conditions,
the proportion of ammonia oxidized to nitrate reduced to less than 10%. It is therefore important
to highlight that the flux rates reported in this report are reflective of a partial nitrification
condition, and may be different than what one would expect to observe for a system achieving
full nitrification.

8-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
8.1.6 Model Structure
In Section 5.3, half-order rate expressions were selected to model the impacts of both
substrate concentration and G. Given the shape of the response curves, saturation-type responses
could alternatively have been selected. Using saturation terms for either the substrate response or
the G response would have allowed for differently shaped response curves for the two media. For
example, the media may have been found to have different half-saturation values for either DO
or G. Theoretically, one might expect that the different media geometries could lead to different
G-responses. For instance, the relatively large openings along the axis of Media 2 might amplify
the effects of mixing on flux, given the relatively higher liquid velocities encouraged by the
media structure.
The dataset developed in this experiment would support either type of model framework,
with little effect on the overall goodness-of-fit. The decision to move forward with a half-order
response for both substrate and G was based upon a desire to simplify the model equation, and
reduce the number of regression parameters. Future work may find that saturation-type
responses, particularly to model the effect of G, are more appropriate.

8.2 Model Evaluation


The model provided a good fit to the observed data, with RMSE less than 10% of the
average flux rates. The error may have been related to variability in the biomass concentration
over the range of batch tests (Figures 6-1C and 6-2C). R1 biomass grew steadily during Phase 1,
decreased precipitously in Phase 2, and then recovered to approximately 20% of Phase 1 value
during Phase 3. R2 biomass also grew steadily during Phase 1, and then increased rapidly during
Phase 2, with an apparent stabilization during Phase 3. Overall, the biomass concentration
exhibited variability throughout the experiment, and this would be expected to influence reaction
kinetics and rates of mass flux measured during batch experiments.
Attempts to normalize the batch test flux rates to measured biomass were unsuccessful, in
that they resulted in far more scatter in the data, and resulted in a poor fit between the model and
the observed data.
In spite of the potential error related to changing biomass concentrations, the model
goodness-of-fit parameters (Table 7-1) were indicative of a reasonable correlation, particularly
for a system where one reactor appeared to fail at cold temperature due to shearing or inhibitory
processes.
8.3 Media Comparison
Much of the literature asserts that media geometry, separate from the media-specific
surface area, imparts little impact on performance. This is the conclusion of one of the seminal
papers in the development of the MBBR process (Ødegaard et al., 2000). However, media
vendors have spent the past 20 years trying to perfect and optimize media geometry, which
suggests that geometry does play a role in performance beyond simple differences in specific
surface area.
This experiment found a clear difference in performance between the two media types.
Applying the calibrated model to a theoretical condition with a bulk phase DO concentration of
5 mg/L and a temperature of 20°C, the following performance plot may be generated (Figure 8-1).

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 8-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 8-1. Model Projection of Media-Specific Reaction Rates at T = 20°C and DO = 5 mg/L.

Media 2 (from R2), with a larger structure and heavy side walls, is projected to support a
higher reaction rate than Media 1 (from R1) at all modeled mixing rates.
By modeling a half-order expression for the relationship between mixing and flux, the
model projects a similar response to increased G for both sets of media.
It must be stated that batch test performance, which is what the model is projecting, is
influenced by the behavior of the continuous flow system. In this experiment, the loss of biomass
in Reactor 1 precluded modeling of Media 1 at cold temperature. Continuous system behavior is
discussed in more detail in Section 8.5.
8.4 Model Application
The experimental protocol and empirical model presented in this study may be used to
compare different types of media at the conditions applied during this experiment. It would not
be appropriate to use these findings to extrapolate how the media would perform at other
conditions (for instance, DO or substrate concentrations outside the range tested). In the future,
the researchers envision the continuous system conditions being similar to those applied in the
batch tests, and based on actual design conditions.
8.5 Continuous-Flow Reactors
The model results reflect the performance observed in the parent reactors. During Phase 1
of the experiment, R2 exhibited consistently higher ammonia removal rates, and supported
higher SALRs at the target effluent concentrations, than R1. Similar results were observed
during Phase 3.
During Phase 2, the reactors were incubated at 10.5°C, at mixing rates ranging from
298/s to 327/s. Under this condition, R1 experienced sloughing of its biofilm. The loss of

8-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
biomass is theorized to have resulted in high bulk phase ammonia concentrations, which may
have inhibited nitrification and prevented growth. The loss of biomass on Media 1 at low
temperature may be related to a number of factors. First, due to an operational error, R1 was
mixed at a slightly higher rate than R2 during this period. The R1 G was approximately 327/s,
compared to 298/s in R2. While important to note, these G values are not outside of typical full-
scale design (Figure 4-5), and the difference in mixing between the two reactors was relatively
small (less than 10% difference). Second, it may be related to the specific geometry of Media 1,
with open side walls which could leave the media surface more susceptible to mixing-induced
shear, leading to an imbalance between the rate of biofilm growth and detachment. The apparent
recovery of R1 AOB activity toward the end of Phase 2 may have been linked to acclimation of
the culture to the lower temperature condition, to the reduction of the reactor mixing rate, or
decreasing bulk phase ammonia concentrations.
There is much evidence in the literature that media-related biomass increases at low
temperatures (Shrestha et al., 2009; Thomas et al., 2009; Wett et al., 2010; Sticker et al., 2009;
Bjornberg et al., 2009; Hoang et al., 2014; Strombeck et al., 2014). The Bjornberg (2009) paper
provides a detailed summary of biomass fluctuations throughout the year at a plant in Minnesota.
That study shows that much of the change in biomass is localized in the interior reaches of the
media. The group suggests that during warm weather, nutrients are absorbed near the outer
portion of the media, and the inner areas are relatively starved. During cold weather, as reaction
kinetics slow down, nutrients are able to travel further into the media, and support growth over a
greater proportion of the surface area. The Hoang (2014) paper included a DNA sequencing
analysis that showed no shift in population from cold to warm temperatures, in spite of the
biomass change. Thomas et al. (2009) speculate that reduced aeration during cold temperatures
would result in reduced mixing, turbulence, and sloughing, therefore allowing for thicker, more
dense biofilms. Parker et al. (1995) noted a similar relationship in nitrifying trickling filters, and
speculated that the thic biomass could be one reason why nitrifying tric ling filters don’t have
the cold temperature declines observed in activated sludge systems.
This experiment also noted increased biomass at cold temperature in R2. Estimated
biomass in R2 increased dramatically shortly after the temperature was reduced from 21°C to
10.5°C, and more than doubled within three months of incubation at 10.5°C (Figures 6-1 and
6-2). It may be possible that the increased biomass density is coupled with an increased friability.
It is possible that the geometry of Media 1 made this media more susceptible to shear or
mechanical impacts from other media or reactor side walls, which could have contributed to the
decreased biomass in R1. There appears to be little data in the literature relating temperature,
mixing rate, and biofilm shearing potential, and this is an area for further study.
The gradual reduction in bulk phase nitrate concentration suggests that the NOB
populations in both reactors declined over the course of the experiment. Nitrate levels
approached zero in both reactors during Phase 2. This may suggest that the NOB are more
susceptible to shear than the AOB, and were therefore more subject to sloughing and washout
than the AOB. Conversely, this could reflect an inhibitory effect on the NOB, perhaps related to
free ammonia in the bulk phase (Appendix C).

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 8-5

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
8.6 Oxygen Flux
The results and modeling presented in this report highlight the flux of ammonia.
However, as noted in Section 5.3, oxygen, not ammonia, was the limiting substrate observed in
the batch experiments. Mechanistically, the oxygen flux would appear to be a more relevant
measure of performance than the ammonia flux rate. Moreover, a modeling framework based
upon oxygen flux would have more relevance to other types of MBBR systems, specifically
those engaged in BOD removal.
As discussed in Chapter 3.0, and expanded upon in Appendix A, initial reactor designs
aimed to calculate oxygen flux directly, through measurements of the air flow into the reactor, as
well as the off-gas. However, such efforts were impeded by a failure to achieve an airtight
system. It was therefore decided to focus on indirect measures of oxygen flux, though
measurement of nitrite and nitrate.
Batch test nitrite and nitrate measurements were implemented in June 2013. However,
such measurements were impeded by two factors. First, the Hach test kits used to measure the
nitrogen species have a limited precision (the high range nitrate test, for example, has a stated
precision of ±5% (Hach 2014). Such error is propagated when combining nitrite and nitrate
measurements to estimate oxygen consumption, resulting in scatter which limits the accuracy of
the model regression. Second, as noted in Section 4.5.2, a potential nitrite interference in the
nitrate test was not corrected until January 2014, which may have resulted in overestimating
nitrate concentrations prior to that time. While the average nitrogen balance (NOx-N produced
divided by NH3-N consumed) across all batch tests was close to unity (1.019), the average
magnitude of the deviation was 4.9%.
The most significant challenge to assessing the oxygen flux had to do with the variation
in NOB activity over the course of the year-long experiment. When nitrite and nitrate production
are used to indirectly estimate oxygen flux, the magnitude of the flux depends upon the activity
of both sets of bacteria. As discussed in the previous section, and shown in Figures 6-1 and 6-2,
the NOB population varied throughout the course of the experiment. Early in the experiment,
both reactors exhibited NOB activity, with an average of 30-40% of oxidized nitrogen in the
form of nitrate for batch tests conducted in June of 2013. Later in the experiment, NOB activity
appeared to decline, with less than 10% of the oxidized nitrogen in the form of nitrate for batch
tests conducted in January, 2014. The variation in NOB activity resulted in a disconnect between
the ammonia flux and the calculated oxygen flux.
The data scatter, combined with the confounding variable of long-term changes in NOB
activity, prevented the use of oxygen flux as the basis of mass transfer modeling in this
experiment. The limitations observed in this experiment should guide future work, and can very
likely be overcome through improved reactor design (allowing for direct measure of oxygen
flux) and improved control of the continuous reactor (allowing for consistency in microbial
population activity).

8-6

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
CHAPTER 9.0

CONCLUSIONS
This work is intended to be an early step toward the development of computationally
simple, empirically derived models of biofilm MBBR and IFAS systems. The protocols
developed, which are summarized in detail in the next section, allow for bench-scale evaluation
of media-based systems using relatively small reactors and materials readily available to
academic and design practitioners. The empirical model is a mathematical expression that
requires inputs of bulk phase concentrations, mixing rates, and temperature. Although limited to
ammonia oxidation in this experiment, a next step would be to develop similar expressions for
carbon removal and/or denitrification.
Conclusions from this work include the following.
9.1 Experimental Methods
 20 L reactors are sufficient to compare biofilm media in either a continuous-flow or batch
experimental setup. For certain media types, the small size of the reactor may limit the
degree of reactor filling, and the range of possible mixing.
 Batch testing is an effective means of conducting a large number of experiments within a
relatively short period of time. Batch testing results are best interpreted alongside concurrent
evaluation of parent continuous-flow reactors.
 The destructive means of biomass measurement appeared to have prevented or slowed the
rate at which cleaned media re-established biofilm biomass. Reinoculation of cleaned media
in mixed liquor accelerated the rate of biomass growth following cleaning.
 High ammonia loading rates were used to produce highly active nitrifying biofilms, but one
consequence was that any system disruptions could lead to high effluent ammonia
concentrations. This in turn is suspected of inhibiting NOB activity, leading to high nitrite
concentrations in both reactors, which in turn can further inhibit NOBs. Reducing this
phenomenon by minimizing effluent ammonia concentrations is an inherent challenge when
operating highly ammonia-loaded reactors.
9.2 Mass Transfer Modeling
 A half-order or saturation-type relationship was noted between the bulk phase DO
concentration and the ammonia flux rate. Although previous studies failed to observe this
effect, this was most likely due to operation at lower DO concentrations than those tested in
this study.
 Ammonia did not appear to be limiting in the concentrations tested in the batch experiments.
This is consistent with the experimental approach of maintaining high ammonia
concentrations, and is also consistent with previous studies.
 A half-order or saturation-type relationship was noted between the mixing rate and the
ammonia flux rate. Although previous studies have noted a relationship between mixing and
flux, the nature of the response has not previously been noted. This may be related to reaction

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 9-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
kinetics becoming rate-limiting once mass transfer limitations have been overcome through
mixing. Further work may be required to determine whether the half-order or saturation
response is more appropriate.
 The empirical model proposed in Equation 7-1 provided a good fit to observed data, with an
average RMSE of less than 10% from observed values. Based on this, the model appears
suitable for media comparison and prediction of ammonia flux rates for a nitrifying MBBR
system of the configuration tested in this experiment.
 In order to assess two-step nitrification, the experimental procedure would need to be
modified to promote NOB growth within the continuous system reactors. This would likely
require operation at reduced feed ammonia concentration, and would require reassessment of
the most appropriate response terms to include in the model equation (refer to Equation 5-6).
9.3 Model Application
The experimental protocol and empirical model presented in this study may be used to
compare different types of media, and to measure the media mass transfer rates under a variety
of conditions. In the future, a similar protocol may be used to develop or evaluate the design of
biofilm systems. Under ideal conditions, the media-specific reaction rate may be used in concert
with bulk phase modeling to assess media requirements for reactor design. Ideal conditions
would involve reactors fed a solution of site-specific wastewater, and maintained to a specific
level of performance (effluent ammonia less than 10 mg N/L, for example). As more data are
developed, and lessons learned from this experiment are applied, an empirical design tool which
reflects differences in media geometry for MBBR systems appears to be a realistically
achievable goal.
9.4 Media Comparison
Under the high loadings used in this study, R1 consistently exhibited lower activity than
R2 in the continuous-system reactors, and the model reflected this observation. R1 also exhibited
loss of biomass and ammonia oxidizing activity at 10.5°C. Conversely, R2 exhibited a biomass
increase and only a moderate reduction in activity at 10.5°C. While these effects may have been
exacerbated by an increased mixing rate in the R1 parent reactor during Phase 2 of the
experiment, the observations suggest a difference in both performance and stability between the
two media under the experimental conditions. It is notable that the observed differences in the
media may or may not be representative of performance at the lower ammonia and higher BOD
loadings typical in domestic wastewater treatment systems.
9.5 Next Steps
Recommendations for further study include the following:
 Assess this model on nitrifying systems operated to lower effluent ammonia concentrations.
What happens to the model when ammonia becomes limiting?
 Assess whether the model relationship for ammonia flux observed in this experiment can be
expressed in terms of oxygen flux. Such an assessment could be made by monitoring oxygen
uptake throughout the batch experiments, or indirectly by monitoring nitrate and nitrite
production rates, provided the continuous system reactor is operated to maintain consistent
activity of AOB and NOB populations.

9-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
 Develop a similar, empirically derived model for biofilm systems involved in carbon
oxidation and/or denitrification.
 Assess the relative impacts of mixing rate on mass transfer in biofilm systems involved in
carbon removal and/or denitrification.
 Assess the relative impacts of mixing rate on mass transfer across a variety of different media
geometries. What is the relationship between media geometry and the apparent synergy
between increased mixing rate and mass transfer?
 Assess the relationship between media geometry and biofilm structural stability. Do certain
media geometries provide more protection against mechanically-induced shear, and are the
effects related to temperature?
 Assess the rate of biofilm growth and/or regrowth on media in synthetic wastewaters with
and without inoculation in mixed liquor or sewage. Were the observations in this experiment
unique to nitrifying biofilms in a carbon-free synthetic feed?

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 9-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
9-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
CHAPTER 10.0

RECOMMENDATIONS
This chapter provides a step-by-step description of the recommended protocols for
continuous MBBR system startup and operation, for conducting batch tests, and for estimation of
model parameters.
10.1 Recommended Continuous-System Protocol
This section presents recommendations for a continuous-system protocol, including
reactor construction, nutrient and water feed preparation, plastic media inoculation, selection of
percent fill and mixing rate, system startup, daily maintenance, and monitoring of performance.
10.1.1 Reactor Construction
Details of the reactor construction are provided in Chapter 4.0, Methods, and a summary
of important parameters is provided in Table 10-1. Coarse-bubble aeration is recommended.
Selection of percent fill and mixing rates are described below.

Table 10-1. Recommended and Example Values for Continuous-System MBBR.

Parameter Value
Reactors
Total volume (empty reactor, including headspace) (L) Approximately 20
Working volume (liquid + media) (L) 10 or more
HRT based on working volume (hour) Approximately 24, as desired
Controls
Aeration and mixing method Coarse bubble
Target mixing rate (G) (-/sec) The same as full-scale system of interest
Dissolved oxygen concentration (mg/L) >6.5 mg/L (not controlled)
pH (pH units) Between 7.15 and 7.50
Temperature (°C) As desired. 21.0 and 10.5 tested in this study.

Selection of reactor size should consider that smaller reactors will require less media
preparation (if synthetic media is used), but larger reactors are probably less susceptible to
system upsets, and their results thus may be more applicable to full-scale systems. Effects of
scale on mixing in MBBR systems is not well-understood – considering that media in smaller
reactors will tend to collide with reactor walls, probes, and aeration apparatus more frequently
than in larger reactors, and their mixing paths will differ in small and large reactors, scaling
effects could be important. Therefore, there are advantages to using the largest reactors feasible.
pH control should be used, as nitrification decreases pH. pH control in the range 7.15 to
7.5 was used in this study, but a narrower range may be useful for reducing possible inhibition
by free ammonia and/or nitrous acid.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 10-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
10.1.2 Nutrient and Water Feed Preparation
Either synthetic feed or wastewater (e.g., secondary effluent) may be used for feed.
Secondary effluent can be supplemented with ammonia to meet desired loading rates. If
available, secondary effluent may be advantageous over synthetic feed as it will provide a more
realistic environment for growth of nitrifiers. If synthetic feed is used, it can be made as a
concentrated solution and diluted with dechlorinated tap water, as was done in this study;
alternatively, it may be made at the net concentrations shown in Table 10-2.
The target ammonia concentration should be sufficient to avoid limiting nitrifier growth
while minimizing risk of inhibiting the process. The goal is to create a deep active biomass to
reduce the likelihood that limited biomass would impede uptake rates. A reasonable starting
concentration is 50 mg N/L. As the effluent ammonia decreases, the influent ammonia should be
periodically increased. Maintaining a target ammonia concentration in the range 10 to 50 mg N/L
is reasonable. For example, R2 was successfully brought to a high level of activity at 21°C by
gradually increasing the influent ammonia from 30 to 450 mg N/L within approximately 50 days,
during which time the ammonia uptake rate increased from 6 to 445 mg N/L and the ammonia
flux (J) increased from 0.03 to 2.4 g N/m2/d, while effluent ammonia was usually less than 30
mg N/L.
If NOB activity is of interest, greater care may be taken to maintain low ammonia
concentrations to prevent NOB inhibition by ammonia (see Appendix C).
1. Autoclave synthetic feed. To avoid precipitation, wait until all other constituents have
dissolved before adding FeSO4,7H20, CuSO4.5H2O, and NaMoO4.
2. If used, prepare the dilution water feed by de-chlorinating tap water with 1.5 mg of NaHSO3
per mg Cl2 to be removed.
3. If synthetic feed is used, add nutrients to dechlorinated tap water to achieve concentrations
found in Table 10-2.

Table 10-2. Synthetic Feed.


(net concentrations)

Chemical Concentration (mg/L)


NH4Cl Variable (see text)
KH2PO4 100
NaHCO3 350
FeSO4-7H2O 5
CaCl2 16
MgSO4-7H2O 40
CuSO4-5H2O 0.12
NaMoO4 0.0019
EDTA 6.6

10-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
10.1.3 Plastic Media Inoculation
1. Obtain a sufficient volume of activated sludge from the aeration basin of a (preferably)
nitrifying wastewater treatment plant for media inoculation (approximately 10 L).
2. Measure out the volume of media equal to the maximum percent media fill of interest in a
graduated cylinder (the percent fill used in the continuous system is determined in the next
section).
3. Place the media into a container with enough activated sludge to fully cover the determined
volume of media.
4. Allow the media to soak in activated sludge with coarse-bubble aeration for three days at
room temperature.
5. Drain the activated sludge from the media.
10.1.4 Selection of Percent Fill and Mixing Rates
Inoculated media should be used to test and select the percent fill, the target continuous
system mixing (aeration) rate, and the range of batch test mixing rates to ensure that thorough
mixing of media (without dead zones) is achieved during continuous-system operation and in all
batch test conditions. Lower mixing rates and higher percent fill values will increase the
occurrence of dead zones. Proceed as follows:
1. Add the first percent fill volume of inoculated media to be tested (e.g., 50%) to the MBBR.
2. Fill to the effluent port with secondary effluent or synthetic feed.
3. Set the aeration rate to the minimum G value of interest in the batch tests (e.g., in the current
study batch test G values were 180/s to 389/s, so G = 180/s was tested. See Chapter 4.0,
Methods, for a discussion of how batch test mixing rates were selected).
4. Visually assess whether complete mixing is achieved. In particular, watch for dead zones in
corners.
5. Proceed to additional G values for testing, increasing mixing rates if necessary. The target
continuous system G value should be tested as well.
6. Proceed to test additional percent fill values, if desired. For example, the percent fill may be
decreased if inadequate mixing is achieved at target G values. In this study, 32% fill was
selected for the continuous-system operation, as higher percent fill values tested did not
provide complete mixing at the minimum target G value.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 10-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
10.1.5 Continuous-Flow System Startup
1. Add the target percent fill of plastic media, and fill to the effluent port with settled activated
sludge supernatant or secondary effluent.
2. Place the reactor into the water bath set to the desired temperature.
3. To begin the continuous flowing reactor, start the influent water feed, influent nutrient, and
effluent pumps and set to flows to match parameters found in Table 10-1. Calibrate each flow
using a 10 mL graduated cylinder and stopwatch. The system configuration used in this study
is shown in Figure 10-1.
4. Set pH controller to regulate pH between 7.15 and 7.50 using 0.5 molar hydrochloric acid
and 0.7 molar sodium carbonate.
5. Use a flow meter to set the airflow corresponding to the desired G value. The calculation of
G based on the airflow rate is presented in Chapter 4.0, Methods.
6. The influent feed concentration should be periodically increased during startup, with a goal
of maintaining a target effluent concentration in order to reduce potential inhibitory effects.

H2O Pump

H2O
Acid Base Feed
Bent Glass
pH Controller Nutrient Pump
Influent

Plexiglass Sidewall Nutrient


Coarse Bubble Feed
Effluent Outlet Aerator

Gas Flow

Effluent Pump Foam Insulated


Water Bath
Air

Temperature

Figure 10-1. Laboratory-Scale Moving-Bed Bioreactor Configuration.

10-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
10.1.6 Daily Maintenance
1. On a daily basis check the reservoirs of influent feed and acid and base solutions.
2. Clean the sides of the reactor with a brush to minimize microbial growth on the reactor walls.
3. Biofilm growth on pH probes will likely occur so maintenance of pH probes is critical.
Calibrate pH probe approximately every other day to ensure proper pH control, and
periodically clean pH probes per the manufacturer’s instructions.
Because pH measurement is influenced by temperature, it is important that pH calibration
be done at the same temperature as the reactor. Standard pH solutions may be kept in the water
bath for convenience, or the temperature correction feature available on most pH meters may be
used.
10.1.7 Monitoring Performance
1. Approximately every other day, take a sample from the reactor using a 10 mL plastic syringe.
Filter using a 0.45 µm nylon filter and store at 4°C. Measure all samples within 24 hours for
ammonia, nitrate, nitrite, and total nitrogen. Note the manufacturer’s instructions for
mitigating interference that may arise between multiple forms of speciated nitrogen (e.g.,
high nitrite interference) present in the sample.
2. Record the time, date, pH, temperature, and airflow rate in the MBBR daily log sheet or lab
notebook.
3. While biomass was not used in the proposed model, it may be useful to periodically measure
attached biomass (see Chapter 4.0, Methods). As discussed in Chapter 6.0, Experimental
Results, it appears that biofilm regrowth was very slow on cleaned media returned directly to
the reactor after biomass measurements. It is therefore recommended that media cleaned for
biomass measurements be replaced with inoculated media. Such media could be primed by
soaking in aerated activated sludge (see Chapter 4.0, Methods), stored at 4°C, for example,
and used as needed. It is not known whether this would be a concern in a secondary-effluent
fed system.
10.2 Recommended Batch Test Protocol
Batch tests should be periodically conducted after the continuous systems have reached
maximum, steady ammonia utilization rates. The purpose of the batch tests is to collect data for
estimation of model parameters, with the experimental variables:
 G.
 DO concentration.
 Temperature (set to the same as the continuous systems).
The batch test procedure is as follows:
1. Disassemble reactor connections in preparation for batch testing:
– Turn off influent, effluent, acid, and base pumps.
– Disconnect airflow, acid, base, influent, and effluent lines from the reactor.
2. Remove the reactor, with its contents, from the water bath and carefully drain all liquid into a
container. This liquid should be saved and used to refill the reactor prior to restarting the
continuous system after the batch test.
3. Remove all media and set them aside in a clean tray (Figure 10-2).

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 10-5

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure 10-2. Media Removed from Reactor Before Batch Testing.

4. Thoroughly clean any biofilm from reactor


sides and aerator (Figure 10-3), and rinse
the reactor with hot tap water. Provide final
rinse with DI water.

Figure 10-3. Cleaning of Reactor with Brush


Before Each Batch Test.

10-6

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
5. Add fresh batch test liquid media to the
reactor to approximately half of the reactor
volume. The total volume added will be
equal to the liquid media volume (the total
reactor volume less the plastic volume), as
previously determined. The batch test
media is the same as the normal feed
solution, except the ammonia nitrogen
concentration should be approximately
30 mg N/L.

Figure 10-4. Filling of Reactor with


Batch Test Media.
6. Carefully return the plastic media to the
reactor (Figure 10-5) and add the remainder
of the liquid batch test media. The water
level should be the same as before the
media were removed from the reactor, and
it should reach the effluent port after
aeration is turned on for batch testing
(below).

Figure 10-5. Return of Plastic Media to the Reactor.


7. Return the reactor to the water bath.
8. Reconnect the effluent, acid, base, and airflow lines and return the pH electrode to the
reactor. Turn the pH controller back on and control the pH in the same range as was done in
the continuous system (7.15 to 7.5 in this study). Place the DO probe in the reactor as well.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems 10-7

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
9. Adjust the gas flows to the reactor, which may include a mixture of air and nitrogen or
oxygen, to provide the target reactor DO and G value for the batch test at hand. It is
recommended that lower G values be tested before higher G values on a given day.
10. After desired batch test conditions are reached and stabilized, take the time zero sample,
filter immediately (25 mm, 0.45 µm nylon membrane syringe filter, VWR 28145-489, or
similar), and store in an airtight container at 4°C for analysis of ammonia, nitrite, and
nitrate within 24 hours.
11. Maintain desired conditions for the duration of the test, which is typically 60 minutes.
Take additional samples at exactly 30 minutes and 60 minutes or more frequently.
12. If additional batch tests are to be conducted, aerate the reactor at the continuous-system G
value (approximately 300/s for the WERF study) for 20 minutes to remove attached
biomass that may have accumulated during the previous batch test (e.g., because of
operation at lower G value) and to provide identical conditions before each batch test.
13. Remove all liquid from the reactor, refill with fresh batch test liquid media, and conduct
the next test as described above. Up to six batch tests in one day were conducted in this
study.

10-8

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
APPENDIX A

EVALUATION OF SEALED REACTOR WITH AIR


RECIRCULATION FOR OXYGEN UPTAKE RATE TESTS
The original proposal for this work included measurements of oxygen uptake in batch
tests. This task was made more difficult through the use of aeration for mixing, because it was
impossible to stop aeration to conduct an oxygen uptake rate (OUR) test without a mechanical
means of mixing. Provision of gas mixing with an oxygen-free gas was a possible approach to
addressing this dilemma. However, this would result in the stripping of oxygen from solution
and the rate of stripping would vary with mixing (bubbling) rates.
Attempts were made to seal the reactor headspace and recirculate the air used for
mixing, which would theoretically allow determination of OUR rates by considering the
change in the total oxygen mass in the system, including the liquid and gas phases, using the
headspace and tubing volumes and Henry’s Law (Figure A-1). Gas recirculation was
accomplished by connecting tubing from the sealed headspace to the vacuum (inlet) pump
connection, and from the pressure (outlet) pump connection to the coarse-bubble aerator.
Dissolved oxygen (DO) concentration control was implemented through the use of a fine-
bubble diffuser, a DO probe, and a separate vacuum/pressure diaphragm air pump (model
DAA-V715-EB, GAST Manufacturing, Inc., Benton Harbor, Michigan, U.S.) connected to a
DO controller (using LABVIEW software).
In spite of repeated efforts to seal the reactor, oxygen leakage into the reactor was
persistent. For example, OUR tests were conducted in which the reactor DO concentration
would decline after ammonia was added, but once the ammonia was depleted, the reactor DO
concentration would return to its previous value within hours, indicating that oxygen was
entering the system through a leak in the reactor headspace, the tubing, or the pump. Because
the pump used in this system was not designed to provide 100% recirculation of air, it is
possible that leakage was occurring at the pump, which did not seem possible to correct
without purchasing a different, specialty pump.
However, OURs were not critical for the final product of this study, which can use
ammonia uptake rates to evaluate the effects of mixing on mass transport rates. The
recirculating aeration design shown in Figure A-1 was therefore abandoned, after more than a
month of testing, in favor of the simpler, open design shown in Figure 4-1.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems A-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
LABVIEW software/computer

DO probe
DO control air
pump
pH meter and controller
Recirculating air pump
for mixing

Feed

Effluent

Fine bubble aerator Coarse bubble aerator

Figure A-1. Initial Reactor Design with Air Recirculation.

A-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
APPENDIX B

SAMPLE VELOCITY GRADIENT CALCULATION


Mixing intensity (G) induced by aeration, was calculated from Equation B-1:

√ 1

where:
G = velocity gradient (1/s)
 = dynamic viscosity (N*s/m2)
V= volume (m3)
Q= airflow rate (m3/s)
 = water specific weight (N/m3)
HL = head loss (m) (distance from aerators to water surface)

Example calculation: the G value for typical continuous operation of the R1 was
calculated as follows:
 = 9.98 x 10-4 N*s/m2 at T = 21°C (Davis and Cornwell, 2013)
V = 9.4 L
Q = 3.35 x 10-4 m3/s
 = 9787 N/m3 at T = 21°C (Davis and Cornwell, 2013)
HL = 0.25 m

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems B-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
B-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
APPENDIX C

INHIBITION
Many compounds are known to inhibit nitrification, with nitrifiers generally more
sensitive to chemical inhibition than heterotrophs. Known inhibitors include heavy metals, such
as zinc, copper, and cadmium (Chandran and Love, 2008; Juliastuti et al., 2003), and a variety of
organic compounds. Growth in biofilms such as those found in integrated fixed-film activated
sludge (IFAS) may provide protection to nitrifiers for some inhibitors (Kim et al., 2010).
The substrates for ammonia-oxidizing bacteria (AOB) and nitrite-oxidizing bacteria
(NOB) (ammonia and nitrite, respectively) appear to be the most important inhibitors of
nitrification, although in typical domestic wastewater treatment systems ammonia and nitrite
concentrations are too low to be inhibitory (Anthonisen et al., 1976). The uncharged species of
each acid/base pair of these compounds is the inhibitory form: free ammonia (FA), the base of
the NH4+/NH3 acid/base pair, and free nitrous acid (FNA), the acid of the HNO2/NO2- acid/base
pair. pH therefore has different effects on inhibition for each compound: ammonia tends to be
more inhibitory at higher pH values where FA dominates, and nitrite tends to be more inhibitory
at lower pH values, where FNA dominates.
Several studies have assessed the concentrations at which ammonia and nitrite can inhibit
nitrification. Anthonisen et al. (1976) described an extensive study with the goal of estimating
the concentrations of FA and FNA that inhibit nitrification (as both ammonia and nitrite
oxidation). Based on batch-, pilot-, and full-scale activated sludge experiments and corroborative
data, they estimated that the FA concentration at which inhibition occurs may be as low as 8 to
as high as 120 mg N/L for AOB, and that the range for the onset of inhibition for NOB was
0.1 to 1.0 mg/L. FNA inhibition of NOB was estimated to start in the range 0.07 to 0.8 mg N/L.
Based on relationships reported by Anthonisen et al., 1976, the minimum levels of
ammonia and nitrite inhibition for the pH ranges and temperatures used in this study were
calculated and compared to measured values to assess the possibility of AOB and/or NOB
inhibition. In this study, pH was controlled in the range 7.15 to 7.50, and the temperature was
either 21°C or 10.5°C. The ranges of ammonia species (NH4+ + FA) and nitrite (NO2- + FNA)
concentrations at which inhibition may begin were calculated using pKa relationships (Table C-1).

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems C-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Table C-1. Concentrations of Nitrogen Species at Which Inhibition May Begin for
AOB and NOB Calculated from Data in Anthonisen et al., 1976.
T = 21°C T = 10.5°C
pH = 7.15 pH = 7.5 pH = 7.15 pH = 7.5
AOB Inhibition by Ammonia
FA, mg N/L total ammonia, mg N/L total ammonia, mg N/L
Low estimate 8 1,400 620 3,100 1,400
High estimate 120 20,700 9,300 46,000 21,000
NOB Inhibition by Ammonia
FA, mg N/L total ammonia, mg N/L total ammonia, mg N/L
Low estimate 0.1 14 6.2 31 14
High estimate 1 140 62 310 140
AOB or NOB Inhibition by Nitrite
FNA, mg N/L total nitrite, mg N/L total nitrite, mg N/L
Low estimate 0.07 370 830 280 620
High estimate 0.8 4,700 11,000 3,500 7,900
Total ammonia is FA + NH4+, and total nitrite is FNA + NO2-. FA and FNA concentrations were taken from Anthonisen et al.,
1976. Total ammonia and total nitrite concentrations were calculated using pKa values and equations in Anthonisen et al. (1976).
Bold values are critical (lowest) estimated values for each temperature.

The values shown in Table C-1 indicate that:


 AO inhibition by FA was estimated to begin under the “worst-case” condition of pH 7.5
at total ammonia concentration of 620 to 9,300 mg N/L at T = 21.5°C and 1,400 to 21,000
mg N/L at T = 10°C (Table C-1).
Because these inhibitory concentrations were much higher than those observed in the reactors
in this study (Figures 6-1 and 6-2), AOB inhibition by FA was unlikely. For comparison,
Magri et al., 2007 used full-scale data from a SHARON (single reactor high activity
ammonia removal over nitrite) process to calibrate a model of AOB activity, and they
reported a Haldane-type kinetics FA inhibition constant of 24.9 mg N/L for AOB. Taking
this as an estimate of an inhibiting concentration, it is notably higher than the 8 mg N/L value
in Table C-1, suggesting the latter to be a conservative estimate. Park and Bae, 2009
estimated Haldane-type FA inhibition constants to be 4 to 22 mg N/L. While the low end of
the this range is less than the 8 mg N/L in Table C-1, the corresponding critical total
ammonia concentration (252 mg N/L), is still greater than measured values in this study.
Carrera et al., 2004 measured inhibition of nitrification for treatment of high-strength
wastewaters. Reading from figures provided in this paper, their calibrated models predicted
maximum AOB activity at approximately 70 mg ammonia N/L in a suspended growth
system, and approximately 200 mg ammonia N/L in a biofilm system; the latter value
suggest ammonia inhibition of AOB was unlikely in the current study.

C-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
 NO inhibition by FA was estimated to occur under the “worst case” condition of pH 7.5
at total ammonia concentration of 6.2 to 62 mg N/L at T = 21.5°C and 14 to 140 mg N/L at
T = 10°C (Table C-1).
Because these inhibitory levels were within the range of ammonia concentrations observed in
this study (Figures 6-1 and 6-2), NOB inhibition by FA was possible. Inhibition of NOB by
FA could therefore be a cause of the high nitrite concentrations observed in both reactors
throughout the study. Park and Bae, 2009 estimated the Haldane-type FA inhibition constant
to be 0.64 mg N/L, which is 6x greater than that minimum 0.1 mg N/L FA shown in the
Table C-1, suggesting the latter is a conservative estimate.

 NOB inhibition by FNA was estimated to begin under the “worst-case” condition of pH
7.15 at total nitrite concentration in the range 370 to 4,700 mg N/L at T = 21.5°C and 280 to
3,500 mg N/L at T = 10°C (Table D-1).
The maximum measured nitrite concentration in R1 during Phases 1 and 2 was 205 mg N/L,
suggesting NOB inhibition by FNA was not likely during these phases. During Phase 3
concentrations were as high as 310 mg N/L (Figure 6-1), suggesting that FNA inhibition of
NOB was possible, but it was generally less than the minimum estimate of 270 mg N/L for
the onset of NOB inhibition. Nitrite concentrations in R2 were higher than in R1, with a
maximum value of 386 mg N/L in Phase 1, but it later decreased to approximately 310 mg
N/L (Figure 6-1), suggesting that FNA inhibition of NOB during Phase 1 was not likely to
play an important role in the reactor performance. The maximum nitrite concentration during
Phase 2 (T = 10°C) of the study was 205 mg N/L, which was less than lower estimates for
NOB inhibition (280 mg N/L), suggesting that FNA inhibition of NOB was not important in
R2 during Phase 2. The maximum nitrite concentration in R2 during Phase 3 (T = 21.5°C)
was 417 mg N/L, although it later decreased to approximately 260 mg N/L (Figure 6-2).
Because these maximum nitrite concentrations were within the calculated inhibitory ranges
(Table C-1) NOB inhibition by FNA was possible at the beginning of Phase 3, but it was
unlikely later.
For comparison, the study by Magri et al., 2007 noted above reported a FNA inhibition
constant of 0.44 mg N/L for AOB, and Park and Bae, 2009 reported FNA inhibition constants in
nitrifying sludges at 0.17 mg N/L for AOB inhibition, both of which are higher than the 0.07 mg
N/L value shown in Table C-1, suggesting the latter is a conservative estimate. Park and Bae,
2009 also reported FNA inhibition constants in nitrifying sludges to be 0.02 to 0.10 mg N/L for
NOB inhibition. 0.02 mg N/L FNA corresponds to 110 mg N/L and 80 mg N/L total nitrite at
21°C and 10.5°C, at the minimum pH of 7.15 in this study. Considering the nitrite concentrations
in both reactors were greater than these values (Figures 6-1b and 6-2b), NOB inhibition by nitrite
appears possible for much of the study in both reactors by the “worst-case” numbers from Par
and Bae, 2009. Similarly, reading from figures provided in Carrera et al., 2004 NOB maximum
activity was predicted at approximately 30 mg nitrite N/L in a suspended growth system, and
approximately 80 mg nitrite N/L in a biofilm system. The latter values suggest NOB inhibition
by nitrite was possible in this WERF study.
Anthonisen et al., 1976 noted that many factors may affect inhibition, so their reported
ranges of FA and FNA where inhibition may begin (Table C-1) were necessarily broad, to reflect
uncertainty associated with these values. The above analysis focused on the lower end of
Anthonisen et al.’s estimates, and which can be considered conservative estimates.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems C-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
It’s also notable that most of the studies discussed above, including Anthonisen et al.,
(1976) focused on activated sludge systems, while resistance to mass transport and pH effects
within nitrifying biofilm systems likely affect inhibition. AOB activity (ammonia oxidation to
nitrite) produces protons, which decreases the pH, thereby decreasing FA concentrations and
increasing FNA concentrations (Park et al., 2010) .This has the effect of making NOB inhibition
less likely at the bulk FA concentrations described above.
In summary, the most likely inhibitory effect was that of high ammonia on NOB activity,
although NOB inhibition by nitrite was also possible. Indeed, both reactors exhibited nitrite
accumulation, which supports the hypothesis that NOB inhibition occurred. Figures C-1 and C-2
show the continuous system data in Reactors 1 and 2, respectively, to evaluate inhibition. These
figures show the relationship between effluent ammonia concentrations and effluent nitrite
concentrations, which is dependent on both AOB activity (nitrite production) and NOB activity
(nitrite consumption); and also the parameter 1-effluent NO3-/(effluent NO2- +NO3-), which has a
value of 0 when NOB activity is uninhibited (NO2- concentration is zero) and has a value of 1
when NOB activity is completely inhibited (nitrate concentration is zero). Note that this value is
also equal to effluent nitrite divided by effluent NOx. This value will be referred to as NOB
inhibition (I) hereafter.
The data shown in Figures C-1 and C-2 generally support the hypothesis that ammonia
concentrations in the reactor inhibited NOB activity, particularly for Reactor 2, as NOB
inhibition generally increased with effluent ammonia concentration. A saturation-type model was
fit to this data:

I -
I

where Imax is the maximum inhibition (I) value (set to 1), CNH3 is the effluent ammonia
concentration (mg N/L), and KI is the inhibition coefficient. Fitting of this model to the measured
data shown in Figures C-1 and C-2 by least-squares regression yielded KI values of 1.6 and 2.9
mg N/L for R1 and R2, respectively. These values can be considered estimates of the ammonia
concentrations at which 50% inhibition would occur. Given the sparse data at lower values
(particularly for Reactor 1), they may not be highly accurate, but they do provide some estimate
of the values at which inhibition may have occurred.

C-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figure C-1. Relationship Between Effluent Ammonia and NOB Inhibition Expressed by the Nitrite Concentration
and the Expression 1-(effluent nitrate/effluent NOx) for Reactor 1.
Data are shown for the period spanning the Phase 3 reinoculation and the end of the study. Data points greater than 100 mg N/L
NH3 are not shown for improved readability of lower values, but they are in general agreement with the trends shown.

Figure C-2. Relationship Between Effluent Ammonia and NOB Inhibition Expressed by the Nitrite Concentration
and the Expression 1-(effluent nitrate/effluent NOx) for Reactor 2.
Data are shown for the period spanning the Phase 3 reinoculation and the end of the study. Data points greater than 100 mg N/L
NH3 are not shown for improved readability of lower values, but they are in general agreement with the trends shown.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems C-5

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Figures C-1 and C-2 also show the calculated minimum value of 6.2 mg N/L for the onset
of NOB inhibition in activated sludge by ammonia at T = 21oC, based on the results of
Anthonisen et al., 1976 (Table C-1), as vertical red lines. These data support the hypothesis that
ammonia inhibition of NOB may have occurred in the experimental systems. It is also notable
that the FA inhibition value of 0.64 mg N/L for NOB by Park and Bae (2009), which yields a
total ammonia concentration of approximately 48 mg N/L, appeared to underestimate the
inhibiting concentrations of ammonia in either reactor.
When achieving AOB activity that is not limited by ammonia availability is a research
goal, as was the case in this study, high influent ammonia concentrations may be necessary (they
were commonly 400 to 500 mg N/L in this study), and low effluent ammonia would ideally be
avoided to prevent limitation of AOB growth. Given the relatively low ammonia concentrations
at which NOB inhibition apparently occurred (1.6 to 2.9 mg N/L for 50% inhibition) the
difficulty in simultaneously meeting the goals of avoiding NOB inhibition and AOB growth
limitation with respect to ammonia concentration is apparent. This difficulty is exacerbated by
the variable performance typical of biological systems coupled with high influent ammonia
concentrations, whereby small variations in AOB activity can have large effects on reactor
ammonia concentrations.

C-6

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
APPENDIX D

BATCH TEST RESULTS

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems D-1

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Table D-1. Reactor 1 Batch Test Results.

Average Bulk Phase


Reactor Date T (°C) DO (mg/L) G (1/s) R (mg N/L/hr)
NH3N (mg/L)
R1 4/30/2013 21 5.79 298 11.0 16.1
R1 4/30/2013 21 5.36 240 10.8 12.9
R1 4/30/2013 21 4.76 180 8.9 14.8
R1 4/30/2013 21 4.76 180 9.0 15.8
R1 4/30/2013 21 5.36 240 11.0 13.8
R1 4/30/2013 21 5.79 298 11.1 15.1
R1 5/13/2013 21 5.79 298 10.2 20.3
R1 5/13/2013 21 5.36 240 9.0 15.2
R1 5/13/2013 21 4.76 180 7.6 14.1
R1 5/13/2013 21 4.76 180 7.7 21.9
R1 5/13/2013 21 5.36 240 8.8 17.0
R1 5/13/2013 21 5.79 298 10.4 12.7
R1 5/27/2013 21 4.76 180 8.3 16.0
R1 5/27/2013 21 5.10 212 9.2 17.8
R1 5/27/2013 21 5.36 240 10.8 15.7
R1 5/27/2013 21 5.56 265 11.2 14.1
R1 5/27/2013 21 5.79 298 11.0 13.3
R1 5/27/2013 21 5.97 328 11.3 12.0
R1 5/27/2013 21 5.97 328 11.0 17.3
R1 6/10/2013 21 4.76 180 8.3 24.8
R1 6/10/2013 21 5.10 212 9.3 28.0
R1 6/10/2013 21 5.36 240 11.0 22.6
R1 6/10/2013 21 5.56 265 11.6 24.2
R1 6/10/2013 21 5.79 298 11.9 23.6
R1 6/10/2013 21 5.97 328 12.3 19.0
R1 6/10/2013 21 5.97 328 11.9 25.0
R1 6/24/2013 21 4.76 180 13.0 33.3
R1 6/24/2013 21 5.10 212 14.6 22.1
R1 6/24/2013 21 5.36 240 16.0 23.1
R1 6/24/2013 21 5.56 265 16.5 20.5
R1 6/24/2013 21 5.79 298 17.4 21.8
R1 6/24/2013 21 5.97 328 18.3 18.9
R1 6/24/2013 21 5.97 328 18.1 19.0
R1 7/8/2013 21 4.76 180 10.6 25.3
R1 7/8/2013 21 5.36 240 13.6 29.9
R1 7/8/2013 21 5.56 265 14.9 24.0
R1 7/8/2013 21 5.79 298 15.8 21.4
R1 7/8/2013 21 5.97 328 16.6 29.1
R1 7/8/2013 21 6.28 389 16.8 23.7
R1 7/22/2013 21 4.76 180 11.0 23.2
R1 7/22/2013 21 5.36 240 14.0 21.8
R1 7/22/2013 21 5.56 265 14.8 24.6
R1 7/22/2013 21 5.79 298 16.0 21.8
R1 7/22/2013 21 5.97 328 16.8 27.2

D-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Table D-1. Reactor 1 Batch Test Results.

Average Bulk Phase


Reactor Date T (°C) DO (mg/L) G (1/s) R (mg N/L/hr)
NH3N (mg/L)
R1 7/22/2013 21 6.28 389 17.7 21.4
R1 8/12/2013 21 5.79 298 17.2 16.5
R1 8/12/2013 21 5.79 298 15.8 48.0
R1 8/12/2013 21 5.79 298 16.2 93.3
R1 8/12/2013 21 5.79 298 17.0 249.3
R1 8/19/2013 21 4.76 180 12.2 21.8
R1 8/19/2013 21 5.36 240 15.9 23.2
R1 8/19/2013 21 5.56 265 17.0 24.7
R1 8/19/2013 21 5.79 298 18.3 19.9
R1 8/19/2013 21 5.97 328 19.6 19.7
R1 8/19/2013 21 6.28 389 20.2 19.8
R1 9/26/2013 10.5 5.95 178 6.3 27.8
R1 9/26/2013 10.5 6.72 238 7.1 26.6
R1 9/26/2013 10.5 6.98 262 7.2 24.5
R1 9/26/2013 10.5 7.28 297 7.8 29.1
R1 9/26/2013 10.5 7.51 327 7.7 24.9
R1 9/26/2013 10.5 7.91 390 7.6 25.5
R1 10/7/2013 10.5 5.96 178 5.0 24.1
R1 10/7/2013 10.5 6.79 238 5.1 27.6
R1 10/7/2013 10.5 7.29 262 5.1 23.3
R1 10/7/2013 10.5 7.61 297 5.3 27.2
R1 10/7/2013 10.5 7.59 327 5.4 31.5
R1 10/7/2013 10.5 7.91 390 5.7 26.2
R1 10/28/2013 10.5 3.05 178 1.5 30.3
R1 10/28/2013 10.5 3.07 297 1.8 27.0
R1 10/28/2013 10.5 3.05 390 1.6 29.9
R1 10/28/2013 10.5 6.98 178 1.9 28.7
R1 10/28/2013 10.5 7.05 297 2.0 33.7
R1 10/28/2013 10.5 7.12 390 1.7 30.0
R1 12/16/2013 10.5 7.30 178 3.6 11.1
R1 12/16/2013 10.5 7.80 238 4.9 15.0
R1 12/16/2013 10.5 7.30 262 4.3 12.6
R1 12/16/2013 10.5 7.10 297 3.8 15.2
R1 12/16/2013 10.5 7.40 327 3.3 13.2
R1 12/16/2013 10.5 7.30 390 3.7 12.5
Note: Italicized values for DO indicate estimated DO concentrations based on correlation presented in Figure 6-6.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems D-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Table D-2. Reactor 2 Batch Test Results.

Average bulk phase


Reactor Date T (°C) DO (mg/L) G (1/s) R (mg N/L/hr)
NH3N (mg/L)
R2 5/20/2013 21 5.79 298 14.6 14.5
R2 5/20/2013 21 5.36 240 13.8 11.4
R2 5/20/2013 21 4.76 180 9.9 14.2
R2 5/20/2013 21 4.76 180 10.0 11.6
R2 5/20/2013 21 5.36 240 13.8 10.8
R2 5/20/2013 21 5.79 298 14.3 12.7
R2 6/3/2013 21 4.76 180 10.0 15.9
R2 6/3/2013 21 5.10 212 12.2 16.5
R2 6/3/2013 21 5.36 240 14.2 13.6
R2 6/3/2013 21 5.56 265 14.5 14.4
R2 6/3/2013 21 5.79 298 15.8 11.5
R2 6/3/2013 21 5.97 328 16.5 11.0
R2 6/3/2013 21 5.97 328 16.5 11.5
R2 6/17/2013 21 4.76 180 11.8 21.0
R2 6/17/2013 21 5.10 212 14.0 43.2
R2 6/17/2013 21 5.36 240 15.1 19.1
R2 6/17/2013 21 5.56 265 16.2 19.5
R2 6/17/2013 21 5.79 298 17.9 20.5
R2 6/17/2013 21 5.97 328 18.4 17.1
R2 6/17/2013 21 5.97 328 18.6 18.8
R2 7/1/2013 21 4.76 180 12.1 27.9
R2 7/1/2013 21 5.36 240 16.7 22.1
R2 7/1/2013 21 5.56 265 18.2 21.0
R2 7/1/2013 21 5.79 298 19.6 26.7
R2 7/1/2013 21 5.97 328 21.6 16.9
R2 7/1/2013 21 6.28 389 22.2 21.7
R2 7/15/2013 21 4.76 180 11.4 23.1
R2 7/15/2013 21 5.36 240 16.4 23.9
R2 7/15/2013 21 5.56 265 18.1 20.5
R2 7/15/2013 21 5.79 298 19.4 22.5
R2 7/15/2013 21 5.97 328 20.8 15.1
R2 7/15/2013 21 6.28 389 22.4 16.1
R2 8/12/2013 21 5.79 298 24.4 14.8
R2 8/12/2013 21 5.79 298 24.8 44.5
R2 8/12/2013 21 5.79 298 25.8 94.3
R2 8/12/2013 21 5.79 298 25.0 245.3
R2 8/26/2013 21 4.76 180 12.7 23.9
R2 8/26/2013 21 5.36 240 18.2 19.9
R2 8/26/2013 21 5.56 265 21.1 22.0
R2 8/26/2013 21 5.79 298 23.3 17.0
R2 8/26/2013 21 5.97 328 24.8 15.9
R2 8/26/2013 21 6.28 389 26.2 22.1
R2 9/16/2013 10.5 5.63 158 9.6 26.4
R2 9/16/2013 10.5 6.40 210 13.0 26.4
R2 9/16/2013 10.5 6.65 231 14.3 22.4
R2 9/16/2013 10.5 6.96 261 15.6 22.1
R2 9/16/2013 10.5 7.20 287 16.7 19.6

D-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Table D-2. Reactor 2 Batch Test Results.

Average bulk phase


Reactor Date T (°C) DO (mg/L) G (1/s) R (mg N/L/hr)
NH3N (mg/L)
R2 9/16/2013 10.5 7.60 340 18.1 20.2
R2 9/29/2013 10.5 5.88 178 11.1 22.7
R2 9/29/2013 10.5 6.81 238 15.0 25.2
R2 9/29/2013 10.5 7.34 262 16.9 21.5
R2 9/29/2013 10.5 7.53 297 18.2 20.9
R2 9/29/2013 10.5 7.60 327 19.4 17.1
R2 9/29/2013 10.5 7.73 390 19.7 17.4
R2 10/13/2013 10.5 5.95 178 11.6 21.9
R2 10/13/2013 10.5 6.72 238 15.9 22.1
R2 10/13/2013 10.5 6.98 262 17.1 20.1
R2 10/13/2013 10.5 7.28 297 18.8 22.4
R2 10/13/2013 10.5 7.51 327 20.0 23.4
R2 10/13/2013 10.5 7.91 390 20.6 17.5
R2 10/19/2013 10.5 3.06 178 7.2 30.1
R2 10/19/2013 10.5 3.04 297 10.6 20.7
R2 10/19/2013 10.5 3.10 390 11.0 22.8
R2 10/19/2013 10.5 5.96 178 11.0 25.3
R2 10/19/2013 10.5 7.11 297 17.7 21.0
R2 10/19/2013 10.5 7.07 390 18.9 20.6
R2 11/4/2013 10.5 3.03 297 9.8 26.3
R2 11/4/2013 10.5 5.08 297 13.0 29.2
R2 11/4/2013 10.5 6.05 297 15.8 26.0
R2 11/4/2013 10.5 7.02 297 16.9 25.7
R2 11/4/2013 10.5 7.98 297 16.9 21.9
R2 11/11/2013 10.5 2.95 327 10.3 29.3
R2 11/11/2013 10.5 4.93 327 12.5 29.2
R2 11/11/2013 10.5 5.94 327 14.9 18.3
R2 11/11/2013 10.5 7.06 327 16.5 21.1
R2 11/11/2013 10.5 8.06 327 16.1 18.2
R2 11/18/2013 10.5 3.01 297 8.0 24.8
R2 11/18/2013 10.5 4.91 297 11.2 21.4
R2 11/18/2013 10.5 5.94 297 12.4 24.4
R2 11/18/2013 10.5 7.06 297 14.6 24.8
R2 11/18/2013 10.5 7.97 297 15.0 25.9
R2 11/25/2013 10.5 2.90 178 6.4 28.0
R2 11/25/2013 10.5 7.14 178 12.9 23.1
R2 11/25/2013 10.5 11.05 178 17.2 24.1
R2 11/25/2013 10.5 15.55 178 21.5 23.7
R2 11/25/2013 10.5 21.33 178 24.6 19.0
R2 12/1/2013 10.5 7.52 178 14.0 22.9
R2 12/1/2013 10.5 7.73 238 17.1 17.9
R2 12/1/2013 10.5 7.80 262 18.0 17.9
R2 12/1/2013 10.5 7.80 297 19.5 24.0
R2 12/1/2013 10.5 7.64 327 20.3 18.7
R2 12/1/2013 10.5 7.81 390 21.2 17.4
R2 12/20/2013 10.5 7.65 178 12.2 17.9
R2 12/20/2013 10.5 7.45 238 16.3 21.1

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems D-5

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Table D-2. Reactor 2 Batch Test Results.

Average bulk phase


Reactor Date T (°C) DO (mg/L) G (1/s) R (mg N/L/hr)
NH3N (mg/L)
R2 12/20/2013 10.5 7.35 262 17.2 22.4
R2 12/20/2013 10.5 7.60 297 17.2 17.2
R2 12/20/2013 10.5 7.40 327 17.8 18.3
R2 12/20/2013 10.5 7.45 390 18.7 17.0
R2 1/15/2014 10.5 3.05 297 9.5 8.6
R2 1/15/2014 10.5 7.20 297 18.0 15.3
R2 1/15/2014 10.5 11.14 297 17.7 9.7
R2 1/15/2014 10.5 15.38 297 23.0 15.7
R2 1/15/2014 10.5 20.05 297 25.4 12.1
R2 1/19/2014 10.5 7.65 161 11.7 20.4
R2 1/19/2014 10.5 7.45 225 13.6 21.2
R2 1/19/2014 10.5 7.35 251 14.2 20.9
R2 1/19/2014 10.5 7.60 287 15.7 22.2
R2 1/19/2014 10.5 7.40 318 17.1 21.9
R2 1/19/2014 10.5 7.45 383 17.5 18.0
R2 4/17/2014 21 2.65 180 8.7 29.1
R2 4/17/2014 21 2.55 265 8.6 33.2
R2 4/17/2014 21 2.50 389 11.8 32.8
R2 4/17/2014 21 15.05 180 27.7 19.9
R2 4/17/2014 21 15.05 265 32.5 18.0
R2 4/17/2014 21 14.90 389 35.2 20.0
R2 4/20/2014 21 7.55 180 15.6 17.7
R2 4/20/2014 21 7.45 240 19.5 15.3
R2 4/20/2014 21 7.55 265 18.8 14.9
R2 4/20/2014 21 7.55 298 21.1 19.0
R2 4/20/2014 21 7.45 328 21.4 19.5
R2 4/20/2014 21 7.45 389 23.2 16.6
Note: Italicized values for DO indicate estimated DO concentrations based on correlation presented in Figure 6-6.

Note: Raw data from the batch tests as well as the continuous system reactors is included in
the Raw Data Excel file included with this report.

D-6

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
REFERENCES

Æsøy, A. and Ødegaard, H. (1994). Nitrogen removal efficiency and capacity in biofilms with
biologically hydrolysed sludge as a carbon source. Water Science and Technology 30(6): 63-71.
American Public Health Association, American Water Works Association, Water Environment
Federation. (2012). Standard Methods for the Examination of Water and Wastewater, 22nd
Edition. American Water Works Association, Washington, D.C.
Anthonisen, A.C., Loehr R.C., Prakasam T.B., and Srinath, E.G. (1976). Inhibition of
nitrification by ammonia and nitrous acid. Journal of the Water Pollution Control Federation
48(5):835-852.
Bassin, J.P., Kleerebezem, R., Rosado, A.S., van Loosdrecht, M.C.M., and Dezotti, M. (2012).
Effect of different operational conditions on biofilm development, nitrification, and nitrifying
microbial population in moving-bed biofilm reactors. Environ Sci. Technol. 46, 1546-1555.
Bill, K.A., Bott, C.B., and Murthy, S.N. (2010). Evaluation of sulfide-driven autotrophic
denitrification in a moving bed biofilm reactor, Water Environment Federation, Proceedings of
the Water Environment Federation Annual Conference (WEFTEC), New Orleans, LA.
Bjornberg, C., Lin, W., and Zimmerman, R. (2009). Effect of temperature on biofilm growth
dynamics and nitrification kinetics in a full-scale MBBR system. Proceedings of the WEFTEC
Conference. Water Environment Federation, Alexandria, VA.
Boltz, J.P., Morgenroth, E., and Sen, E. (2010). Mathematical modelling of biofilms and biofilm
reactors for engineering design. Water and Science Technology, 62, 8, 1821-1836.
Bott, C.B., Jones, R., Thomas, W.A., Pinto, A., and Love, N. (2010). Model-based investigation
of full-scale IFAS performance utilizing plant data and batch testing to assess kinetics, mass
transfer effects and population dynamics. Proceedings of Water Environment
Federation/International Water Association Biofilm Reactor Technology Conference, Portland,
OR.
Carrera, J., Jubany, I., Carvallo, L., Chamy, R., and Lafuente, J. (2004). Kinetic models for
nitrification inhibition by ammonium and nitrite in a suspended and an immobilized biomass
systems. Process Biochemistry 39(9), 1159-1165.
Chandran, K. and Love, N.G. (2008). Physiological state, growth mode, and oxidative stress play
a role in Cd(II)-mediated inhibition of Nitrosomonas europaea 19718. Applied and
Environmental Microbiology 74(8), 2447-2453.
Davis, M.L. and Cornwell, D.A. (2013). Introduction to Environmental Engineering, McGraw
Hill (Pub.), New York.
Di Trapani, D., Christensson, M., and Ødegaard, H. (2011). Hybrid activated sludge/biofilm
process for the treatment of municipal wastewater in a cold climate region: a case study. Water
Science and Technology 63(6), 1121-1129.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Di Trapani, D., Christensson, M., Torregrossa, M., Viviani, G., and Ødegaard, H. (2013).
Performance of a hybrid activated sludge/biofilm process for wastewater treatment in a cold
climate region: Influence of operating conditions. Biochemical Engineering Journal 77, 214-219.
Eberl, H., Morgenroth, E., Noguera, D., Picioreanu, C., Rittmann, B., van Loosdrecht, M., and
Wanner, O. (2006). Mathematical Modeling of Biofilms. Scientific and Technical Report No. 18,
IWA Publishing: London, England.
Grady, Jr. C.P.L., Daigger, G.T., and Lim, H.C. (1999). Biological Wastewater Treatment. 2nd
Edition. Marcel Dekker, Inc. New York, NY.
Gromiec, M.J., et al. (1972). Performance of plastic medium in trickling filters. Water Research
11: 1321-1332.
Hach Company. (2014). Nitrate, HR, Chromotropic Acid Method. Method 10020.
DOC316.53.01068. Hach Company, Loveland, CO.
Harremoës, P. (1977). Half-Order Reactions in Biofilm and Filter Kinetics. Vatten, 33, 122-143.
Hellinga, C., Schellen, A., Mulder, J.W., van Loosdrecht, M.C.M., and Heijnen, J.J. (1998). The
SHARON process: An innovative method for nitrogen removal from ammonium-rich waste
water. Water Science and Technology 37(9), 135-142.
Hem, L.J., Rusten, B., and Ødegaard, H. (1994). Nitrification in a moving-bed biofilm reactor.
Water Research 28(6), 1425-1433.
Henze, M., Van Loosdrecht, M.C.M., Ekama, G.A., and Brdjanovic, D. (2008). Biological
Wastewater Treatment: Principles, Modelling, and Design. IWA Publishing, London, England.
Hoang, V., Delatolla, R., Abujamel, T., Mottawea, W., Gadbois, A., Laflamme, E., and Stintzi,
A. (2014). Nitrifying moving bed biofilm reactor (MBBR) biofilm and biomass response to long
term exposure to 1°C. Water Research 49: 215-224.
Howland, W.E. (1958). Flow over porous media as in a trickling filter. Proceedings of the 12th
Industrial Wastewater Conference, Purdue, IN.
Johnson, T.L. and McQuarrie, J.P. (2002). IFAS BNR full-scale design and performance
challenges. Proceedings of the WEFTEC Conference. Water Environment Federation,
Alexandria, VA.
Johnson, T.L., McQuarrie, J.P., and Shaw, A.R. (2004). Integrated fixed-film activated sludge
(IFAS): the new choice for nitrogen removal upgrades in the United States. Proceedings of the
WEFTEC Conference. Water Environment Federation, Alexandria, VA.
Jones, R., Bott, C.B., Rutherford, R., Baumler, R., and Waltrip, D. (2010). Full-Scale IFAS
Investigation: Plant Data and Batch Testing to Assess Kinetics, Mass Transfer Effects and
population dynamics. Proceedings of WEFTEC Conference, New Orleans, LA, Water
Environment Federation: Alexandria, VA.
Juliastuti, S.R., Baeyens, J., and Creemers, C. (2003). Inhibition of nitrification by heavy metals
and organic compounds: The ISO 9509 test. Environmental Engineering Science 20(2), 79-90.

R-2

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Kim, H.-s., Gunsch, C.K., Geller, J.L., Boltz, J.P., Freudenberg, R.G., and Schuler, A.J. (2010).
Shelter from the storm: Integrated fixed film activated sludge protects nitrifiers from toxic
upsets. Proceedings of Water Environment Federation Annual Conference (WEFTEC), New
Orleans, LA.
Kristensen, G.H., Jorgensen, P.E., and Henze, M. (1992). Characterization of functional
microorganism groups and substrate in activated sludge and wastewater by AUR, NUR and
OUR. Water Science and Technology 25(6), 43-57.
Logan, B.E., Hermanowicz, S.W., and Parker, D.S. (1987). A fundamental model for trickling
filter process design. Journal of the Water Pollution Control Federation 59: 1029-1042.
Magri, A., Corominas, L., Lopez, H., Campos, E., Balaguer, M., Colprim, J., and Flotats, X.
(2007). A model for the simulation of the SHARON process: pH as a key factor. Environmental
Technology 28(3), 255-265.
Melcer, H., Dold, P.L., Jones, R.M., Bye, C.M., Takacs, I., Stensel, H.D., Wilson, A.W., Sun, P.,
and Bury, S. (2003). Methods for Wastewater Characterization in Activated Sludge Modeling.
WERF Project Report No. 99-WWF-3. WERF (Pub.), Alexandria, VA.
Morgenroth, E., van Loosdrecht, M.C.M., and Wanner, O. (2000). Biofilm models for the
practitioner. Water Science and Technology, 41(4-5), 509-512.
Muller, N. (1998). Implementing biofilm carriers into activated sludge process – 15 years of
experience. Water Science and Technology, 37(9), 167-174.
Ødegaard, H., Rusten, B., and Siljudalen, J. (1999). The development of the moving bed biofilm
process - From idea to commercial product. European Water Management 2:36-43.
Ødegaard, H., Gisvold, B., and Strickland, J. (2000). The influence of carrier size and shape in
the moving bed biofilm process. Water Science and Technology 41(4-5):383-391.
Park, S. and Bae, W. (2009). Modeling kinetics of ammonium oxidation and nitrite oxidation
under simultaneous inhibition by free ammonia and free nitrous acid. Process Biochemistry
44(6), 631-640.
Park, S., Bae, W., and Rittmann, B.E. (2010). Multi-species nitrifying biofilm model (MSNBM)
including free ammonia and free nitrous acid inhibition and oxygen limitation. Biotechnology
and Bioengineering, 105(6), 1115-1130.
Parker D.S. (1970) Characteristics of biological flocs in turbulent regimes, Doctoral thesis,
University of California, Berkeley, http://disexpress.umi.com/dxweb, UMI publication number
7109887.
Parker, D.S. and Merrill, D.T. (1984). Effect of plastic media configuration on trickling filter
performance. Journal of the Water Pollution Control Federation 56: 955-961.
Parker, D., Lutz, M., Andersson, B., and Aspegren, H. (1995). Effect of operating variables on
nitrification rates in trickling filters. Water Environment Research, 67 (7), 1111-1118.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems R-3

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Parker, D.S., Jacobs, T., Bower, E., Stowe, D.W., and Farmer, G. (1997). Maximizing trickling
filter nitrification rates through biofilm control: research review and full-scale application. Water
Science and Technology, 36(1), 255-262.
Parker, D.S. (2010). A consultant’s perspective on process design for biofilm based wastewater
treatment processes in North America. Proceedings of WEF-IWA Biofilm Reactor Technology
2010 Conference, Portland, Ore., Water Environment Federation: Alexandria, VA.
Pastorelli, G., Andreottola, G., Canziani, R., Darriulat, C., Frangipane, E.D.F., and Rozzi, A.
(1997). Organic carbon and nitrogen removal in moving-bed biofilm reactors. Water Science and
Technology 35(6), 91-99.
Perez, J., Costa, E., and Kreft, J.U. (2009). Conditions for partial nitrification in biofilm reactors
and a kinetic explanation. Biotechnology and Bioengineering 103(2), 282-295.
Rittmann, B.E. and McCarty, P.L. (2001). Environmental Biotechnology: Principles and
Applications. McGraw-Hill Book Co., New York, NY.
Rusten, B., Hem, L.J., and Ødegaard, H. (1995). Nitrification of municipal wastewater in
moving-bed biofilm reactors. Water Environment Research 67(1): 75-86
Rusten, B., McCoy, M., Proctor, R., and Siljudalen, J.G. (1998). The innovative moving bed
biofilm reactor/solids contact reaeration process for secondary treatment of municipal
wastewater. Water and Environment Research 70(5), 1083-1089.
Saez, P.B. and Rittmann, B.E. (1988). Improved pseudoanalytical solution for steady-state
biofilm kinetics. Biotechnology and Bioengineering 32: 379-385.
Saez, P.B. and Rittmann, B.E. (1992). Accurate pseudoanalytical solution for steady-state
biofilms. Biotechnology and Bioengineering 39: 790-793.
Shresthra, A., Riffat, R., Bott, C., Takacs, I., Stinson, B., Peric, M., Neupane, D., and Murthy, S.
(2009). Process considerations to achieve nitrogen removal in a moving-bed biofilm reactor.
Proceedings of the WEFTEC Conference. Water Environment Federation, Alexandria, VA.
Stricker, A.E., Barrie, A., Maas, C.L.A, Fernandes, W., and Lishman, L. (2009). Comparison of
performance and operation of side-by-side integrated fixed-film and conventional activated
sludge processes at demonstration scale. Water Environment Research 81(3):219-232.
Strombeck, J., Bjornberg, C., Zimmerman, R., and Lin, W. (2014). Modeling of temperature
impacts on fixed film microbial growth and nitrification kinetics. Central States Water
Environment Association 87th Annual Meeting Proceedings. Central States Water Environment
Association, Crystal Lake, IL.
Thomas, W.A. (2009). Evaluation of nitrification kinetics for a 2.0 MGD IFAS process
demonstration. Master of Science Thesis, Virginia Polytechnic Institute and State University,
Blacksburg, VA.
Thomas, W.A., Bott, C.B., Regmi, P., Schafran, G., McQuarrie, J., Rutherford, B., Baumler, R.,
and Waltrip, D. (2009). Evaluation of nitrification kinetics for a 2.0 mgd IFAS process

R-4

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
demonstration. Proceedings of WEFTEC Conference, Orlando, FL, Water Environment
Federation: Alexandria, VA.
Vadivelu, V.M., Keller, J., and Yuan, Z.G. (2007). Effect of free ammonia on the respiration and
growth processes of an enriched Nitrobacter culture. Water Research 41(4), 826-834.
Wanner, O., Eberl, H., Morgenroth, E., Noguera, D., Picioreanu, C., Rittmann, B., and van
Loosdrecht, M. (2006) Mathematical Modeling of Biofilms. Scientific and Technical Report No.
18, IWA Publishing: London, England.
Water Environment Federation. (2010). WEF Manual of Practice No. 35. Biofilm Reactors Task
Force of the Water Environment Federation. WEF Press, Alexandria, VA.
Wett, B., Jimenez, J.A., Takacs, I., Murthy, S., Bratby, J.R., Holm, N.C., and Ronner-Holm,
S.G.E. (2010). Models for nitrification process design: one or two AOB populations?
Proceedings of the WEFTEC Conference. Water Environment Federation, Alexandria, VA.
Zekker, I., Rikmann, E., Tenno, T., Menert, A., Lemmiksoo, V., Saluste, A., Tenno, T., and
Tomingas, M. (2011). Modification of nitrifying biofilm into nitritating one by combination of
increased free ammonia concentrations, lowered HRT and dissolved oxygen concentration.
Journal of Environmental Sciences-China 23(7), 1113-1121.
Zhang, S., Wang, Y., He, W., Wu, M., Xing, M., Yang, J., Gao, N., and Pan, M. (2014). Impacts
of temperature and nitrifying community on nitrification kinetics in a moving-bed biofilm reactor
treating polluted raw water. Chemical Engineering Journal 236, 242-250.

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems R-5

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
R-6

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
WERF Subscribers
WASTEWATER UTILITY South Orange County Iowa New York
Wastewater Authority Ames, City of New York City Department
Alabama Stege Sanitary District Cedar Rapids Water of Environmental
Montgomery Water Works Sunnyvale, City of Pollution Control Protection
& Sanitary Sewer Board Thousand Oaks, City of Facilities North Carolina
Alaska Union Sanitary District Des Moines, City of Charlotte-Mecklenburg
Anchorage Water & West Valley Sanitation Kansas Utilities
Wastewater Utility District Johnson County Durham, City of
Arizona Colorado Wastewater Metropolitan Sewerage
Avondale, City of Aurora, City of Louisville and Jefferson District of Buncombe
Glendale, City of Boulder, City of County Metropolitan County
Peoria, City of Greeley, City of Sewer District Old North State Water
Phoenix Water Services Littleton/Englewood Olathe, City of Company Inc.
Department Wastewater Treatment Overland Park, City of Orange Water & Sewer
Pima County Wastewater Plant Kentucky Authority
Reclamation Department Metro Wastewater Sanitation District No. 1 Raleigh, City of
Tempe, City of Reclamation District Louisiana Ohio
Arkansas Platte Canyon Water & Sewerage & Water Board Akron, City of
Little Rock Wastewater Sanitation District of New Orleans Avon Lake Municipal
California Connecticut Maine Utilities
Central Contra Costa Greater New Haven Bangor, City of Columbus, City of
Sanitary District WPCA Portland Water District Metropolitan Sewer District
Corona, City of District of Columbia of Greater Cincinnati
Maryland
Crestline Sanitation District DC Water Montgomery County Water
Anne Arundel County
Delta Diablo Services
Florida Howard County Bureau of
Dublin San Ramon Services Northeast Ohio Regional
Fort Lauderdale, City of Utilities
District Sewer District
JEA Washington Suburban
East Bay Dischargers Summit County
Hollywood, City of Sanitary Commission
Authority Oklahoma
Miami-Dade County Massachusetts
East Bay Municipal Utility Oklahoma City Water &
Orange County Utilities Boston Water & Sewer
District Wastewater Utility
Department Commission
Encino, City of Department
Palm Beach County Upper Blackstone Water
Fairfield-Suisun Sewer Tulsa, City of
Pinellas County Utilities Pollution Abatement
District
Reedy Creek Improvement District Oregon
Fresno Department of
District Albany, City of
Public Utilities Michigan
St. Petersburg, City of Bend, City of
Inland Empire Utilities Ann Arbor, City of
Tallahassee, City of Clean Water Services
Agency Detroit, City of
Toho Water Authority Gresham, City of
Irvine Ranch Water District Gogebic-Iron Wastewater
Georgia Lake Oswego, City of
Las Gallinas Valley Authority
Atlanta Department of Oak Lodge Sanitary District
Sanitary District Holland Board of Public
Watershed Portland, City of
Las Virgenes Municipal Works
Management Water Environment
Water District Saginaw, City of
Augusta, City of Services
Livermore, City of Wayne County Department
Los Angeles, City of Clayton County Water of Public Services Pennsylvania
Montecito Sanitation Authority Wyoming, City of Philadelphia, City of,
District Cobb County Water Water Department
Minnesota
Napa Sanitation District System University Area Joint
Metropolitan Council
Novato Sanitary District Columbus Water Works Authority
Environmental Services
Orange County Sanitation Gwinnett County South Carolina
Rochester, City of
District Department of Water Beaufort - Jasper Water &
Western Lake Superior
Palo Alto, City of Resources Sewer Authority
Sanitary District
Riverside, City of Macon Water Authority Charleston Water System
Savannah, City of Missouri
Sacramento Regional Greenwood Metropolitan
Independence, City of
County Sanitation Hawaii District
Kansas City Missouri
District Honolulu, City & County of Mount Pleasant
Water Services
San Diego, City of Idaho Waterworks
Department
San Francisco Public Boise, City of Spartanburg Water
Metropolitan St. Louis
Utilities, City and Sullivan’s Island, Town of
Illinois Sewer District
County of Greater Peoria Sanitary Nebraska Tennessee
San Jose, City of District Lincoln Wastewater & Cleveland Utilities
Sanitation Districts of Los Metropolitan Water Solid Waste System Murfreesboro Water &
Angeles County Reclamation District of Nevada Sewer Department
Santa Barbara, City of Greater Chicago Henderson, City of Nashville Metro Water
Santa Cruz, City of Sanitary District of Decatur New Jersey Services
Santa Rosa, City of Bergen County Utilities Texas
Silicon Valley Clean Water Indiana
Authority Austin, City of
South Coast Water District Jeffersonville, City of
Ocean County Utilities Dallas Water Utilities
Authority Denton, City of

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
WERF Subscribers
El Paso Water Utilities STORMWATER UTILITY D&B/Guarino Engineers INDUSTRY
Fort Worth, City of LLC
Houston, City of California Effluential Synergies LC American Water
San Antonio Water System Monterey, City of EMA Inc. Anglian Water Services
Trinity River Authority San Diego County ENVIRON International Ltd.
Department of Public Corporation Chevron Energy
Utah
Works Environmental Operating Technology Company
Salt Lake City Department
San Francisco Public Solutions Inc. Dow Chemical Company
of Public Utilities
Utilities, City & County of Evoqua Water DuPont Company
Virginia Santa Rosa, City of Eastman Chemical
Technologies
Alexandria Renew Sunnyvale, City of Company
Gannett Fleming Inc.
Enterprises Eli Lilly & Company
Colorado GeoSyntec Consultants
Arlington County FMC Corporation
Aurora, City of GHD Inc.
Fairfax County InSinkErator
Boulder, City of Global Water Advisors Inc.
Fauquier County Johnson & Johnson
Florida Greeley & Hansen LLC
Hampton Roads Sanitation Merck & Company Inc.
Orlando, City of Hazen & Sawyer P.C.
District Procter & Gamble
Iowa HDR Inc.
Hanover County Company
Cedar Rapids Water HNTB Corporation
Henrico County Suez Environnement
Pollution Control Holmes & McGrath Inc.
Hopewell Regional United Utilities North West
Facilities Infilco Degremont Inc.
Wastewater Treatment United Water Services LLC
Des Moines, City of Jacobs Engineering
Facility VandCenter Syd
Group Inc.
Loudoun Water Kansas Veolia Water North
KCI Technologies Inc.
Lynchburg Regional Overland Park, City of America
Kelly & Weaver P.C.
Wastewater Treatment
Pennsylvania Kennedy/Jenks Consultants
Plant
Philadelphia, City of, Larry Walker Associates
Prince William County
Service Authority Water Department LimnoTech
Malcolm Pirnie, the Water List as of 9/10/14
Richmond, City of Tennessee
Division of ARCADIS
Rivanna Water & Sewer Chattanooga Stormwater
Authority MaxWest Environmental
Management Systems
Washington Washington McKim & Creed
Everett, City of Michael Baker, Jr. Inc.
Bellevue Utilities
King County Department MWH
Department
of Natural Resources NTL Alaska Inc.
& Parks Seattle Public Utilities
Parametrix Inc.
Puyallup, City of Praxair Inc.
Seattle Public Utilities STATE AGENCY Pure Technologies Ltd.
Sunnyside, Port of Ross Strategic
Yakima, City of Connecticut Department of
Environmental Protection Stone Environmental Inc.
Wisconsin Stratus Consulting Inc.
Fresno Metropolitan Flood
Green Bay Metro Synagro Technologies Inc.
Control District
Sewerage District Tata & Howard Inc.
Harris County Flood
Kenosha Water Utility Tetra Tech Inc.
Control District
Madison Metropolitan The Cadmus Group Inc.
Kansas Department of
Sewerage District The Low Impact
Health & Environment
Milwaukee Metropolitan Development Center Inc.
Ohio River Valley
Sewerage District URS Corporation
Sanitation Commission
Racine Water & Versar, Inc.
Urban Drainage & Flood
Wastewater Utility Westin Engineering Inc.
Control District, CO
Sheboygan, City of Wright Water
Stevens Point, City of Engineers Inc.
Wausau Water Works CORPORATE
Zoeller Pump Company
AECOM
Australia/New Alan Plummer Associates Australia
Zealand Inc. CSIRO
Water Services Association American Cleaning Institute
Austria
of Australia Aqua-Aerobic Systems Inc.
Atkins Sanipor Ltd.
Canada Canada
Calgary, City of Black & Veatch
Corporation Associated Engineering
EPCOR O2 Environmental Inc.
Lethbridge, City of Brown and Caldwell
Burns & McDonnell Trojan Technologies Inc.
Regina, City of
Toronto, City of Carollo Engineers, P.C. Norway
Winnipeg, City of CDM Smith Aquateam Cowi AS
CH2M HILL
CRA Infrastructure &
Engineering

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
WERF Board of Directors
Chair Paul L. Bishop, Ph.D., P.E., Julia J. Hunt, P.E. James Anthony (Tony)
Catherine R. Gerali BCEE Trinity River Authority Parrott
Metro Wastewater University of of Texas Metropolitan Sewer
Reclamation District Rhode Island District of Greater
Terry L. Johnson, Ph.D., Cincinnati
Vice-Chair Glen Daigger, Ph.D., P.E., P.E., BCEE
Kevin L. Shafer BCEE, NAE Water Consulting, LLC Cordell Samuels
Metro Milwaukee CH2M Hill Regional Municipality
Sewerage District Jim Matheson of Durham, ON
Scott D. Dyer, Ph.D. Oasys Water
Secretary The Procter & Gamble Interim Executive Director
Company Ed McCormick, P.E. Lawrence P. Jaworski,
Eileen J. O’Neill, Ph.D.
Water Environment Water Environment P.E., BCEE
Philippe Gislette Federation
Federation
Degrémont,
Treasurer Suez-Environnement
Brian L. Wheeler
Toho Water Authority

WERF Research Council


Chair Donald Gray (Gabb), Mark W. LeChevallier, Paul Togna, Ph.D.
Terry L. Johnson, Ph.D., Ph.D., P.E., BCEE Ph.D. Environmental Operating
P.E., BCEE East Bay Municipal American Water Solutions, Inc.
Water Consulting, LLC Utility District
Ted McKim, P.E. BCEE Art K. Umble, Ph.D., P.E.,
Vice-Chair Robert Humphries, Ph.D. Reedy Creek BCEE
Rajendra P. Bhattarai, P.E., Water Corporation of Energy Services MWH Global
BCEE Western Australia
Austin Water Utility Elizabeth Southerland, Kenneth J. Williamson,
David Jenkins, Ph.D. Ph.D. Ph.D., P.E.
Thomas C. Granato, Ph.D. University of California U.S. Environmental Clean Water Services
Metropolitan Water at Berkeley Protection Agency
Reclamation District of
Greater Chicago

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf
by Universidad Nacional Autonoma De Mexico UNAM user
WERF Product Order Form
As a benefit of joining the Water Environment Research Foundation, subscribers are entitled to receive one complimentary copy of all final
reports and other products. Additional copies are available at cost (usually $10). To order your complimentary copy of a report, please write
“free” in the unit price column. WERF keeps track of all orders. If the charge differs from what is shown here, we will call to confirm the total
before processing.

________________________________________________________________________________________________
Name Title
________________________________________________________________________________________________
Organization
________________________________________________________________________________________________
Address
________________________________________________________________________________________________
City State Zip Code Country
________________________________________________________________________________________________
Phone Fax Email

Stock # Product Quantity Unit Price To t a l

Postage &
Method of Payment: (All orders must be prepaid.) Handling
VA Residents Add
q C heck or Money Order Enclosed 5% Sales Tax
q Visa q Mastercard q A m e rican Express Canadian Residents
Add 7% GST
______________________________________________________
Account No. Exp. Date
TOTAL
______________________________________________________
Signature

S h ip p in g & Ha n dli n g: To Order (Subscribers Only):


Amount of Order United States Canada & Mexico All Others
7 Log on to www. werf.org and click
on “Publications.”
Up to but not more than: Add: Add: Add:
Phone: 571-384-2100
$20.00
30.00
$7.50*
8.00
$9.50
9.50
50% of amount
40% of amount
( Fa x : 703-299-0742
40.00 8.50 9.50 WERF
50.00 9.00 18.00 Attn: Subscriber Services
60.00 10.00 18.00 635 Slaters Lane
80.00 11.00 18.00 Alexandria, VA 22314-1177
100.00 13.00 24.00
150.00 15.00 35.00 To Order (Non-Subscribers):
200.00 18.00 40.00
Non-subscribers may order WERF
More than $200.00 Add 20% of order Add 20% of order
publications either through WERF
* m i n i mum amount for all orders or IWAP (www.iwapublishing.com).
Visit WERF’s website at www.werf.org
Make checks payable to the Water Environment Research Foundation. for details.

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf
by Universidad Nacional Autonoma De Mexico UNAM user
U4R11_WEF-IWAPspread.qxd 10/20/2014 11:25 AM Page 1

Mass Transfer Characteristics of Floating Media in MBBR and IFAS Fixed-Film Systems
Wastewater Treatment Systems

Water Environment Research Foundation


635 Slaters Lane, Suite G-110 n Alexandria, VA 22314-1177
Phone: 571-384-2100 n Fax: 703-299-0742 n Email: werf@werf.org
www.werf.org
WERF Stock No. U4R11

Co-published by

IWA Publishing
Alliance House, 12 Caxton Street Mass Transfer Characteristics of Floating Media
London SW1H 0QS
United Kingdom
Phone: +44 (0)20 7654 5500
Fax: +44 (0)20 7654 5555
in MBBR and IFAS Fixed-Film Systems
Email: publications@iwap.co.uk
Web: www.iwapublishing.com
IWAP ISBN: 978-1-78040-705-0/1-78040-705-X

Co-published by

November 2014

Downloaded from https://iwaponline.com/ebooks/book-pdf/1667/wio9781780407050.pdf


by Universidad Nacional Autonoma De Mexico UNAM user
on 12 June 2020

You might also like