You are on page 1of 66

02-CTS-4.

qxd

9/20/04

2:37 PM

Page 1 (2,1)

Water Environment Research Foundation

Treatment Processes

Membrane Bioreactors for Anaerobic


Treatment of Wastewaters

The University of British Columbia


Civil Engineering

Membrane Bioreactors for Anaerobic


Treatment of Wastewaters
WERF Project 02-CTS-4

Task 1. Phase 1 Report:


Compilation/Review of Existing Literature
P.M. Sutton, P. Brub and E.R. Hall
August 31, 2004

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
wastewater utilities, 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 300
Alexandria, VA 22314-1177
Tel: (703) 684-2470
Fax: (703) 299-0742
www.werf.org
werf@werf.org
Copyright 2004 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: 2004114158
Printed in the United States of America
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.
The University of British Columbia
This document was reviewed by a panel of independent experts selected by WERF. Mention of trade names or
commercial products does not constitute WERF nor EPA endorsement or recommendations for use. Similarly,
omission of products or trade names indicates nothing concerning WERF's nor EPA's positions regarding product
effectiveness or applicability.
The research on which this report is based was funded, in part, by the United States Environmental Protection
Agency through Cooperative Agreement No. CR-827345-01 with the Water Environment Research Foundation
(WERF). Unless an EPA logo appears on the cover, this report is a publication of WERF, not EPA. Funds awarded
under the Cooperative Agreement cited above were not used for editorial services, reproduction, printing, or
distribution.

TABLE OF CONTENTS

1.0

Background ..........................................................................................................................4

2.0

Approach..............................................................................................................................4

3.0

Results from Review of Existing Literature and Other Information ...................................5

4.0

3.1

Anaerobic MBR Treatment System Experience......................................................6


3.1.1 Literature and Information from Other Sources ..........................................6
3.1.2 Anaerobic MBR Treatment of Municipal Wastewaters ............................12

3.2

Anaerobic Treatment of Lower Strength Wastewaters at Lower Temperatures ...13

3.3

Mechanisms Governing the Efficiency of the Membrane Component of the


Anaerobic MBR .....................................................................................................15
3.3.1 Membrane System ....................................................................................16
3.3.2 Operational Parameters..............................................................................19
3.3.3 Characteristics of the Mixed Liquor ..........................................................25

Summary: Parameters Critical to Anaerobic MBR System Performance


and Efficiency ....................................................................................................................28
4.1
4.2
4.3

5.0

Critical Bioreactor Parameters...............................................................................29


Critical Membrane System Parameters..................................................................29
Mixed Liquor Characteristics ................................................................................30

References..........................................................................................................................31
Appendix A
National and International Researchers Contacted ................................................58
Commercial MBR System and/or Membrane Suppliers Contacted ......................59
Appendix B
Nomenclature List..................................................................................................60

1.0

BACKGROUND
The WERF project entitled Membrane Bioreactors for Anaerobic Treatment of
Conventional and Medium Strength Wastewater was initiated in early August 2003. The
objective of the project is to compile existing information and develop new information from
laboratory and field pilot studies, in order to assess the technical and economic feasibility of
treating relatively low strength wastewater using membrane biological reactors (MBRs)
operating under anaerobic conditions.
The first project task (i.e., Task 1.0 Phase 1) was the compilation and review of existing
literature and other information, and completion of a report. The principal objective of Task 1.0
Phase 1 was to identify the variables thought to be critical to achieving a high effluent quality
(e.g., secondary, tertiary) in an economical fashion (e.g., achieving high membrane efficiency) in
the treatment of municipal and/or low-to-medium strength industrial wastewaters using the
anaerobic MBR technology.

2.0

APPROACH
The following activities were completed in order to compile the literature and other
information, which formed the basis of the report.
Appropriate databases were identified for the literature compilation task and a search of the
databases was completed.
Titles were downloaded together with abstracts where available. Selected, full text
documents were downloaded where feasible or ordered through the University of British
Columbia (UBC) library system.
Contact was made with national and international researchers thought to be involved in or
have knowledge of, anaerobic MBR research projects (Appendix A).
Contacts were made with commercial suppliers of MBR systems and/or membranes used
in MBR systems, with the objective of obtaining information they may have developed in
the application of the MBR technology to anaerobic treatment of wastewaters or contacts
they may have with commercial practitioners of the technology (Appendix A).
Members of the Technical Advisory Panel (TAP) were contacted and asked to review the
status of anaerobic MBR technology in their respective geographical regions.
The databases searched in conducting the literature compilation task included SciFinder
Scholar, Dialog, El Compendex, NTIS, Water Resources Abstracts, CA Search, SciSearch
Waternet, Inside Conferences, Dissertation Abstracts, Pollution Abstracts, U.S. EPA, as well as
internet search engines. Relevant information derived as a result of contacting MBR system
researchers and consultants, and commercial suppliers, is included herein.

3.0

RESULTS FROM REVIEW OF EXISTING LITERATURE AND OTHER


INFORMATION

The literature and other information compiled were critically reviewed keeping in mind
the principal objective of Task 1.0 Phase 1, stated previously. The information compiled relevant
to treatment of municipal and/or low-to-medium strength industrial wastewaters using anaerobic
MBR technology, was categorized as follows.
General system design, performance and efficiency information derived from operation of
laboratory, pilot and full scale anaerobic MBRs treating municipal or industrial wastewater.
Specific information relevant to the questions of bioreactor performance and efficiency
during operation of anaerobic MBRs in the treatment of municipal and/or low-to-medium
strength wastewaters under low temperature conditions (i.e., less than 18C).
Specific information relevant to the question of membrane efficiency based on operation of
MBRs under anaerobic conditions.
The performance and economics of an anaerobic MBR system will dictate its
applicability in the treatment of municipal and/or low-to-medium strength industrial wastewaters.
It is important to document existing information implying or indicating the potential of anaerobic
MBRs to achieve near complete treatment (i.e., CBOD5, carbonaceous five-day biochemical
oxygen demand, equal to or less than 30 mg/L) or to play a role in the management of municipal
wastewaters using less traditional approaches. An example of such an approach might be the use
of an anaerobic MBR system as a satellite plant located on a municipal wastewater trunk line,
designed to reduce the organic load to an existing, centralized treatment facility. Treatment
system economics will be affected by factors impacting the capital and operating costs of both
the bioreactor, and the membrane component of the system. Factors dictating bioreactor costs
will include volumetric efficiency, mechanical configuration details (e.g. number of bioreactor
stages/compartments, suspended growth versus fixed film reactors, etc) and the operating
complexity (e.g., instrumentation and control requirements). Factors dictating membrane
component costs will include the membrane configuration (e.g., tubular, hollow fiber membrane
bundles), membrane composition (e.g., polymeric versus ceramic), membrane packing density
(i.e., membrane area per unit volume), membrane flux and permeability (i.e., flux per unit
transmembrane pressure or TMP), membrane cleaning requirements (e.g., procedures and
chemical requirements), membrane cross-flow and/or sparging requirements, and required
membrane pumps and/or blowers.
The relevant compiled literature and other information follow, categorized as noted. An
analysis of the information was completed in order to identify the critical bioreactor parameters
(e.g., feed characteristics, SRT, volumetric loading, TSS concentration, pH, temperature,
alkalinity) and membrane parameters (e.g., membrane materials, membrane configuration, pore
size, operating TMP, cleaning requirements) dictating the performance and efficiency of the
bioreactor and the membrane component of the anaerobic MBR system. The results of the
information analysis, summarized in Section 4.0, influenced the Task 2.0 Phase 2 laboratory
studies experimental program.

3.1

Anaerobic MBR Treatment System Experience

General system design, performance and efficiency information derived from operation
of laboratory, pilot and full scale anaerobic MBR systems treating municipal or industrial
wastewater was compiled from published and unpublished literature, and other information
sources.
3.1.1

Literature and Information from Other Sources

Literature information and information from other sources (i.e., TAP member reports and
contacts noted in Appendix A) were compiled, reviewed and the relevant information
summarized in tabular form where possible (Table 1). Comments regarding Table 1 follow.
Table 1 is not intended to provide detailed information on the question of membrane
efficiency and performance in the context of operation of anaerobic MBRs. That
information is presented in Section 3.3.
The entries included in Table 1 are specific to anaerobic (i.e., not aerobic or anoxic) MBR
treatment of wastewaters and wastewater solids (i.e., does not include application of
membranes in conjunction with anaerobic sludge digestion). Table 1 includes entries from
studies with influent CODs over a wide range of concentrations; however, relatively few
studies (< 10) could be characterized as using low-to-medium strength influents
The System referred to in Table 1 is a suspended growth anaerobic reactor unless
otherwise indicated. Certain systems consist of two-stage or two-phase reactors. The
results presented in those cases are specific to the reactor coupled with an internally
submerged or externally located membrane.
Where possible, the reference entries (1 through 55) are arranged by research groups or
common system supplier applications, as follows:
- entries 1 through 7, the C-H. Lee group from Seoul National University, South
Korea,
- entries 8 and 9, the E.R. Hall group from The University of British Columbia,
Vancouver, British Columbia,
- entries 10 through 14, the A. Fakhrul-Razi group from Universiti Pertanian
Malaysia, Serdang, Malaysia,
- entries 15 and 16, the G. Yuntao group from Tsinghua University, Beijing, China,
- entries 17 and 18, the full scale experience at Tenstar Products, Ashford, U.K.,
- entries 19 and 20, the Ross Consultancy group from Tygerpark, South Africa,
- entries 21 through 23, the G.K. Anderson group from The University of
Newcastle upon Tyne, Newcastle upon Tyne, U.K., and
- entries 24 through 27, the Dorr-Oliver, membrane anaerobic reactor system
(MARS) applications, Dorr-Oliver, Stamford, CT.
The values for the various parameters recorded in Table 1 are as stated in or calculated
from the referenced publication. If a parameter value is not stated in Table 1 for a
particular reference, this indicates that the value was not available or could not be

calculated with certainty. Single values stated are typically average values but should be
treated as approximate.
Values stated for such parameters as OLR (organic loading rate), COD (chemical oxygen
demand) removal and MLSS (mixed liquor suspended solids) do not necessarily include
results from the start-up phase of the bioreactor system.
There is widespread interest in the anaerobic MBR technology for wastewater treatment,
as evident from the number of entries in Table 1. Anaerobic MBR development activities have
largely taken place in the Far East, Europe and South Africa. The first known research project
involving the coupling of an anaerobic reactor to a membrane module for sanitary wastewater
treatment, took place in the U.S. in the mid to late 1970s (Table 1, entry 31). Full-scale
application of the technology was first reported in England where a UF (ultrafiltration)
membrane unit was coupled to an existing suspended growth anaerobic reactor installed for the
treatment of wheat flour processing wastewater at Tenstar Products (Table 1, entry 17). Although
there are no known commercial, large scale, anaerobic MBR systems in operation treating
municipal wastewater, active research with respect to this application in China, England, and the
U.S. has recently been reported (Table 1, entries 46, 53, and 54). Further information concerning
this anaerobic MBR application is provided in Section 3.1.2. The technical feasibility of applying
anaerobic MBR systems for treatment of municipal and other lower strength wastewaters is
likely to depend on the necessity of achieving a high level of performance (e.g., effluent CBOD5
less than 30 mg/L) at lower temperatures. Information addressing the issue of anaerobic system
performance under these conditions, is presented in Section 3.2.
The following general observations can be made regarding application of the anaerobic
MBR technology for industrial and municipal wastewater treatment from examination of the
contents of Table 1 and/or from additional information derived from the references cited.
Anaerobic MBR systems have been applied for treatment of a wide variety of industrial
wastewaters at the lab, pilot and full scale level. Applications have generally involved the
use of a UF or MF (microfiltration) membrane located external to the bioreactor (i.e.,
external membrane MBR configuration). Tubular and flat plate membranes represented
the most popular external membrane configuration. In most lab and many pilot scale
systems, the membrane component was not operated under conditions to maximize the
membrane throughput or flux (i.e., optimal TMP and cross-flow velocity) but rather
served as a component which allowed absolute retention of biomass. As such, the results
from these studies provide little information of value in design of the membrane
component for use in full scale anaerobic MBR systems. In some of the lab and pilot
scale studies referenced, the experimental program focused on the efficiency of the
membrane component and/or the mechanisms governing membrane fouling (e.g., Table 1
entries 1 through 7). The results from these and other studies where relevant membrane
component information was reported, are discussed in Section 3.3, as previously noted.
The membrane component of the larger pilot scale and full scale MBR systems
referenced in Table 1 (e.g., entries 17, 20, 26, 34, 35), appeared to be operated at
conditions which should maximize the membrane flux. Unfortunately, the data available
were not derived from operation during recent years and therefore may not be
representative of the performance that could be expected from such membranes as are

manufactured today. The results of contacts made with those familiar with the full scale
systems referenced in Table 1, indicate the systems are no longer in operation or recent
operating data are not available.
Little information has been reported regarding the performance and efficiency of
anaerobic MBRs when operated under lower temperature conditions. The economic
feasibility of applying anaerobic MBR systems for treatment of relatively low strength
wastewaters may be predicated on efficient operation of the system at lower temperatures.
Information addressing the issue of operation of anaerobic wastewater treatment systems
under these conditions, is presented in Section 3.2, as previously noted.
The COD and BOD (biochemical oxygen demand) removals reported from operation of
lab, pilot and full scale anaerobic MBRs in the treatment of readily biodegradable, higher
strength wastewaters (e.g., brewery, alcohol fermentation, wheat flour and starch, cheese
whey and whey permeate) generally have exceeded those reported from operation of
more conventional anaerobic systems (e.g., UASB, upflow anaerobic sludge blanket
reactors, or anaerobic PBR, packed bed reactors). Reports of COD and BOD removals
exceeding 95% are not uncommon (e.g., entries 4, 10, 11, 13, 14, 15, 20, 23, 24, 26, 35,
and Table 2). Effluent COD and BOD5 values less than respectively, 100 and 50 mg/L
were reported in the treatment of industrial and municipal wastewaters (Table 1, entries
24, 46, 53, and 54). The improved performance of the MBR versus other anaerobic
reactor configurations is largely the result of absolute retention of the biomass produced
in the reactor together with retention of organic solids originating in the feed. In certain
instances, performance improvements were reported as a result of physical retention of
soluble organics by the membrane component (Table 1, entries 8, 24, 38, and 46). Ince et
al. (1998, 2000) determined through concentration profiles across a suspended growth
anaerobic MBR treating brewery wastewater, that the soluble COD in the reactor was two
to three times higher than in the effluent (i.e., membrane permeate). The reported results
imply that MBR technology represents an ideal reactor configuration for achieving the
highest quality effluent feasible in the anaerobic treatment of wastewaters.
Lab, pilot and full scale anaerobic MBR systems have been operated at OLRs ranging
from less than 5 to over 30 kg COD/m3day. In general, the higher OLRs reported were
based on treatment of readily biodegradable wastewaters utilizing external membrane
MBRs in which a high concentration of MLSS (i.e., 20 to 60 g/L) was maintained in the
bioreactor (e.g., Table 1, entries 12, 22, 23, 37, 55). There are reports of operation of
internal submerged membrane MBRs under anaerobic conditions at OLRs up to 19
kg/m3day in which the bioreactor MLSS exceeded 20 g/L (Table 1, entries 32 and 48).
These anaerobic MBRs involved the use of membranes manufactured by Kubota, Japan.
Kubota claims to have installed six, full scale, internal membrane anaerobic MBRs in
Japan in the last four years, all treating food processing wastewater or kitchen waste
combined with night soil and other organic solids (Kimura, 2004). The first full-scale
Kubota plant was installed at the Shinmo-Inc medical facility in Nagano. Repeated
attempts to obtain design and operating information for this plant and/or information
associated with the other plants, were unsuccessful. Despite a lack of operating data from
full scale anaerobic MBR systems currently in operation, the full scale results reported in
the 1980s and 1990s, and the lab and pilot scale results reported at that time and more
recently, indicate the technology has the potential to meet stringent CBOD5 and TSS
discharge standards.

There have been reports of a decrease in specific biomass activity and/or excessive
biomass decay in the operation of lab and pilot scale external membrane MBRs operating
under anaerobic (Table 1, entries 4, 38, 41, 42, 54, and Nikaido et al., 1996) and aerobic
conditions (Kim et al., 2001). Results indicate that the decrease in activity is correlated to
the magnitude of the shear conditions within the external membrane loop and the
frequency at which the reactor contents are exposed to those conditions. The external
membrane component used in both the Jones and Hall, and Sutton pilot studies (Table 1,
entries 24 and 27) relied on pumping of the bioreactor contents through the membrane
channels at a high membrane TMP and cross-flow velocity to effect permeation. These
operating conditions and the small operating volume of the anaerobic pilot reactors
resulted in frequent exposure of the anaerobic biomass to a high shear environment. The
conditions required a gradual adaptation of the anaerobic seed material (i.e., solids from a
conventional municipal sludge digester) added at start-up, to the high shear environment
and may have influenced the specific rate of biomass activity throughout the pilot studies.
Information falling into the category of general anaerobic MBR treatment system experience
was compiled and reviewed and although relevant, was not suitable for summary in the Table 1
format. The information follows.
Information provided by members of the TAP. The Ministry of International Trade and
Industry (MITI), Japan launched a six-year research and development project (i.e., AquaRenaissance 90) in 1985 with the objective of water reuse and energy recovery This project led
to the development in the late 1980s of a number of anaerobic MBR-based pilot plant systems
for treatment of industrial and municipal wastewaters ranging in capacity from 0.5 to 20 m3/day.
Key findings from this project of importance in assessing the technical and economic feasibility
of the anaerobic MBR technology, follow.
In the treatment of wastewaters from fat/oil and protein production and starch processing,
an acidogenic first stage external membrane MBR followed by a second stage
conventional anaerobic reactor (i.e., methanogenic stage) reduced the wastewater BOD
from 13,000 to 50 mg/L. This flowsheet concept has been explored by others (Table 1,
entries 15, 16, 44, 50, and Fukuma et al., 1993) and has generally exhibited higher
performance than the alternative single stage anaerobic MBR flowsheet or the two stage
flowsheet in which an anaerobic MBR constitutes the second stage (i.e., methanogenic
reactor).
Municipal wastewater and wastewater solids were treated anaerobically using a variety of
flowsheet configurations involving multiple stages and MBRs. A fixed-film FBR
(fluidized bed reactor) coupled to a hollow fiber membrane component produced an
effluent containing less than 20 mg/L BOD when treating a feed containing up to 140
mg/L. Despite this performance, a flowsheet involving only anaerobic MBR treatment of
primary separated solids for the production of volatile acids for use in a downstream
nitrogen removal step, was viewed as more suitable for application to municipal
wastewaters. Similar anaerobic MBR flowsheet concepts involving treatment of
municipal wastewater solids have been explored by others (Table 1, entries 34 and 51).
The concept of an anaerobic fixed-film reactor coupled to an external membrane system
for treatment of municipal wastewaters may prove more attractive today than in the late

1980s due to the advent of lower cost membranes. One example of such a configuration
is an anaerobic hybrid reactor, containing suspended biofilm carrier media, coupled to a
small membrane tank containing hollow fiber membranes or flat membrane panels that
rely on gas and liquid scouring to prevent short term membrane fouling. In this case, the
majority of the biomass inventory required to carry out the anaerobic reactions is located
on the fixed-film media. Operation of the system at a low concentration of suspended
growth (e.g., 3 g VSS/L, volatile suspended solids) in the anaerobic reactor and a high
concentration of solids in the membrane tank (e.g., 15 g VSS/L), will minimize the
biomass recycle ratio which would be important if air was used as the source of gas for
membrane scouring. This configuration (Figure 1) is likely to be particularly attractive in
the treatment of municipal or low strength industrial wastewater where aerobic post
treatment may be required to achieve more stringent effluent quality requirements (e.g.,
BOD5 less than 10 mg/L, NH4-N less than 2 mg/L) as the membrane tank could be sized
to achieve nitrification. It should be noted that, in this case, a high biomass recycle ratio
(e.g., 4 or greater) may be favored in order to promote denitrification in the anaerobic
reactor, which in fact may operate at ORP conditions more characteristic of anoxic
reactor. Commercial scale systems for industrial wastewater treatment are in operation in
Europe involving the concept of coupling an anaerobic reactor for pretreatment with a
tank containing membranes relying on air and liquid scouring to prevent short term
membrane fouling and to achieve aerobic post treatment (Kraft and Brockmann, 1999).
Pilot work was reported in Japan in 2000 involving the application of an anaerobic MBR
as part of a flowsheet for treatment of slurries from livestock farming operations together with
kitchen and vegetable waste (K. Yamamoto, University of Tokyo, pers. comm.). In the
flowsheet, an external rotating disc membrane module dramatically improved the efficiency of
the anaerobic digester while providing an effluent permeate suitable for ammonia recovery
through stripping and condensation. The application of an anaerobic MBR for waste treatment
and resource recovery in the management of livestock waste slurries was first reported in
Norway in the early 1990s (Bilstead et al., 1992). A Danish engineering company, Bioscan A/S,
recently installed commercial scale thermophilic anaerobic MBRs in The Netherlands, Denmark
and Japan, and have proposed systems in the U.S. and Canada for the treatment of slurries from
livestock farming operations based on the use of external, tubular ultrafiltration membrane
modules for solids-effluent separation (Ejner, 2003). To improve efficiency of the membrane
component, a centrifuge is used to pretreat the digested solids prior to the membrane step. A
major function of the centrifuge is to remove dense inorganic solids that form in the thermophilic
anaerobic bioreactor (e.g., struvite). The underflow solids from the centrifuge represent waste
solids from the MBR. A similar anaerobic MBR based flowsheet was proposed by the University
of Illinois for the treatment of hog and pig farm waste slurries (Morgenroth et al., 2003).
Pilot work was reported in Japan and other Far East countries in the 1990s on the
application of external membranes to improve the efficiency and performance of conventional
anaerobic sludge digesters or to provide a method for recovery of organic acids from a sludge
fermentor. The objective of one study was the recovery of volatile fatty acids for use as an
organic carbon source for denitrification. This same application is referenced in Table 1 (entry
51). The application of external membranes to improve the efficiency and performance of other
fermentation processes (i.e., recovery of organic acids, production of ethanol and other organic
chemicals) was first reported in the 1980s (Olmstead et al., 1980; OSullivan et al., 1984;

10

Cheryan and Mehaia, 1986; Ferras et al., 1986). More recently Giorno et al. (2002) explored the
concept for lactic acid production. Another report described a lab study in which aerobic sludge
produced from the treatment of glucose and peptone was treated in an anaerobic MBR and the
digested sludge was subjected to alkaline hydrolysis (Takashima et al., 1996). The authors
claimed the combined sludge treatment steps resulted in complete sludge digestion. The
application of external membranes to conventional anaerobic digesters to improve efficiency and
performance or for organic acid recovery, has been explored by researchers in countries outside
the Far East. In South Africa, Pillay et al. (1994) used MF tubular membranes to increase the
volumetric efficiency of a 1800 L pilot scale digester treating municipal wastewater sludge from
primary clarifiers. More recently in the U.S., Lanting (2003) and Jackivicz et al. (2003) reported
on the application of an external membrane system coupled to a digester at the pilot scale level
which they claimed would allow operation of the digester at an HRT as low as one day while
maintaining the SRT in the range from eight to 12 days. In Australia, Barnes et al. (2003)
explored indirectly, the concept of using an external membrane to selectively remove volatile
fatty acids from the acidogenic stage of a two-stage sludge digester and transferring the acid
solution to the second stage digester. Another novel application of membranes in combination
with anaerobic digestion of municipal wastewater solids was reported by Juby et al. (2000). A
treatment train comprising MF plus RO processing of primary effluent and anaerobic digestion
of the RO concentrate was proposed. The results from piloting of the MF and RO treatment steps
were reported.
Other information. The assessment of microbial populations in anaerobic MBRs has
been the focus of a number of studies beginning with the research of Kobayashi et al. (1988).
These researchers examined small (i.e., 189 L) and larger (i.e., 37,850 L) scale external
membrane anaerobic MBRs treating cheese whey permeate, and observed differences in
organism size. Samples from the larger reactor showed more flocculated material and more
numerous long filaments versus the smaller systems. Methanothrix were the predominant
methanogen species observed and Methanosarcina were scarce, particularly in the smaller
system. The Methanothrix organisms were mainly individual short rods. Ince et al. (1995, 1997)
examined a 120 L external membrane anaerobic MBR treating brewery wastewater and made the
same observation regarding organism size. These researchers identified Methanococcus as the
most dominant methanogen species. Kataoka et al. (1992) characterized the bacterial populations
in anaerobic MBRs treating municipal wastewater and soybean processing wastewater.
Analyzing colony forming results, the researchers determined that the wastewater characteristics
and bioreactor operating conditions significantly influenced the physiological characteristics of
the bacterial populations. The soybean wastewater gave rise to much faster growing organisms
than municipal wastewater.
There are a number of reports of operation of MBRs under anoxic conditions with the
objective of reducing nitrate, sulphate and/or other inorganic oxides (e.g., Magara et al., 1992;
Delanghe et al., 1994; Nagaoka, 1999; Zoh and Stenstrom, 2002; Nuhoglu et al., 2002; Kimura
et al., 2002). Typically, an ORP reading of approximately -100 to +100 mV is indicative of
anoxic conditions whereas anaerobic conditions exist below approximately 100 mV. Exposing
the membrane component to lower ORP conditions than those indicative of aerobic conditions
(i.e., greater than approximately +100 mV), may impact the efficiency of the membrane
component as discussed in Section 3.3. Therefore, information from these studies specific to the
membrane component may have relevance to this project.

11

The commercial application of aerobic MBRs for municipal or industrial wastewater


treatment is currently growing at a yearly rate of 20-40% in North America, according to Zenon
Environmental (Bonkoski, 2003). In Japan, it has been reported through mid-2003, that over 360
commercial MBR systems were in operation (Uemura and Kondou, 2003). With the emergence
of the MBR technology, researchers, engineering consultants and system suppliers have begun to
develop process and system models in order to design MBRs, predict performance and estimate
system costs. In order to determine the economic feasibility of applying the anaerobic MBR
technology for treatment of dairy industry wastewaters in France, researchers carried out
laboratory experimental studies in support of the development of anaerobic MBR models (Arros
et al., 2001; Arros-Alileche et al., 2002). Conclusions reached by the researchers from the results
of simulation studies performed using the models, of relevance to this project follow.
A higher membrane flux than reported for anaerobic MBRs in the literature must be
achieved in order for the technology to represent an economical solution for treatment of
dairy industry wastewaters.
A promising anaerobic MBR based flowsheet for dairy wastewater treatment and water
recovery for reuse, is a two stage anaerobic system (i.e., acidogenic plus methanogenic)
with the second stage consisting of an anaerobic suspended growth reactor coupled to a
NF or RO membrane step.
3.1.2

Anaerobic MBR Treatment of Municipal Wastewaters

Although there are no known large, full scale, anaerobic MBR systems in operation
treating municipal wastewater, active research with respect to this application has recently been
reported, as previously noted in Section 3.1.1. Other anaerobic reactor configurations have been
commercially applied for treatment of municipal wastewaters at both high and low temperatures,
as discussed in Section 3.2.
There are four entries in Table 1 (i.e., entries 31, 46, 53 and 54) providing general
anaerobic MBR design and performance information from the results of lab scale treatment of
municipal wastewater. In addition in the subsection entitled, Information provided by
members of the TAP (Section 3.1.1), reference is made to a variety of membrane coupled
anaerobic reactors and flowsheet configurations explored during pilot plant treatment of
municipal wastewaters and wastewater solids. Further information from these reports and other
information sources, deemed particularly relevant to this project, follows.
Stuckey and others at Imperial College in London have completed significant laboratory
research on the application of submerged internal membrane MBRs for anaerobic
treatment of a synthetic feed claimed to be representative of municipal wastewaters. The
only significant information available to our research team to date, is presentation
material from a recent conference (Stuckey and Hu, 2003). The information does imply
that it is feasible to treat degritted and screened municipal wastewater using an anaerobic
MBR system, and achieve effluent CBOD5 and TSS values of less than 30 mg/L at a
bioreactor HRT as low as three hours operating at a temperature of 35C. The
information also implies that the addition of PAC (powdered activated carbon) to an
anaerobic MBR could provide significant bioreactor performance benefits (i.e., improved

12

effluent quality) and increase the efficiency of the membrane component (i.e., up to 50%
increase in membrane flux at same TMP).
Pagilla and others at the Illinois Institute of Technology in Chicago have completed a 15month laboratory experimental study and are completing modeling studies on the
application of an external membrane MBR for anaerobic treatment of municipal
wastewater (Baek and Pagilla, 2003; Pagilla, 2004). The researchers experimental results
imply it is feasible to treat primary treated municipal wastewater using an anaerobic
MBR system and achieve effluent CBOD5 and TSS values less than 30 mg/L at a
bioreactor HRT as low as 12 hours, operating at a temperature of approximately 32C.
The researchers plan to complete mathematical modeling and model calibration studies
and complete fluorescent in-situ hybridization analyses to characterize the reactor
biomass.
Wen and others at Tsinghua University in Beijing completed a six month laboratory
study involving application of an anaerobic reactor containing a hollow fiber, UF
membrane module in the treatment of municipal wastewater (Wen et al., 1999). The
anaerobic reactor is best characterized as a hybrid UASB. Packing media, in the form of
fine fibres, was located towards the top of the reactor. The membrane module was
submerged in an expanded cross-sectional zone above the media. The laboratory results
imply it is feasible to treat municipal wastewater using an anaerobic MBR system, and to
achieve effluent CBOD5 and TSS values of less than 30 mg/L at a bioreactor HRT as low
as four hours, and temperatures as low as 12C. The results imply physical retention of
organics by the UF membrane played a significant role in achieving good performance at
temperatures below 15C.
The New York State Energy Research and Development Authority (NYSERDA) is
sponsoring a field demonstration study in upstate New York involving application of an
anaerobic reactor coupled to an MF membrane component for treatment of degritted and
screened municipal wastewater (Hickey, 2004). The anaerobic reactor consists of a
hydraulic pulsed FBR containing sand as the site for biofilm growth. Solids retained by
the membrane module will be returned to the anaerobic reactor. The study is expected to
begin in the second quarter of 2004.
3.2

Anaerobic Treatment of Lower Strength Wastewaters at Lower Temperatures

Little information has been reported regarding the performance and efficiency of
anaerobic MBRs in the treatment of lower strength wastewaters at lower temperatures, as
previously discussed. Reference has been made to anaerobic fixed-film and UASB reactors
coupled to internal and external membranes, which resulted in effluent COD and BOD5 values
less than respectively, 100 and 30 mg/L (e.g., Table 1 entry 46). In these systems, the membrane
component played a key role in enhancing performance by essentially eliminating TSS, and
particulate or colloidal COD or BOD from the effluent.
Although there is extensive literature documenting the performance and efficiency of
fixed-film, UASB and other anaerobic reactor configurations treating lower strength wastewaters
at lower temperatures (Foresti, 2002; Zeeman and Lettinga, 2002; Elmitwalli et al.,
2002;Angenent et al., 2000), to date there are no large, commercial scale systems operating

13

under these conditions to the knowledge of the writers. Research results indicate that anaerobic
reactors designed to achieve hydrolysis of wastewater organic particulates, capable of
maintaining a high concentration of active methanogens and operating under plug-flow hydraulic
conditions, are likely to achieve the highest efficiency in the treatment of lower strength
wastewaters at temperatures below 20C.
Stensel and Strand (2004) recently examined more than 50 reports documenting the
effects of temperature, wastewater strength and OLR on treatment performance of lab and pilot
scale anaerobic reactors. Reference was make to 13 anaerobic reactor studies reported over the
period from 1981 to 2002 involving treatment of municipal wastewaters at temperatures equal to
or less than 20C, during which an effluent COD of equal to or less than 65 mg/L was achieved.
Additional relevant information from these studies, follows.
The Wen et al. (1999) study included in Table 1 (entry 46), was the only MBR
application referenced.
Ten of the 13 studies involved operation of fixed-film or UASB reactors.
The anaerobic reactor configuration achieving the highest efficiency while operating
under low temperature conditions and achieving an effluent COD of less than 65 mg/L,
was the fixed-film FBR. OLRs of 7.7 to 9.4 kg COD/m3day were reported by
respectively, Jewell et al. (1981) and Rebac (1998) for FBRs. Achieving an anaerobic
reactor OLR of 7 kg COD/m3day or greater in the treatment of degritted and screened or
primary treated municipal wastewater (i.e., COD 200 to 400 mg/L), translates to a
bioreactor HRT of less than 1.4 h, representing efficient treatment.
Additional relevant information on the topic of anaerobic treatment of lower strength
wastewaters at lower temperatures not referenced in the review of Stensel and Strand (2004),
follows.
UASB reactors continue to receive particular attention for the treatment of municipal
wastewaters at lower temperatures. Singh and Viraraghavan (2003), and Lew et al. (2003)
operated pilot scale UASBs in this application at temperatures as low as respectively, 6
and 10C. At these temperature conditions, COD removal was less than 50%. Singh and
Viraraghaven (2003) concluded at a bioreactor temperature of approximately 10C and
operation at a COD OLR of approximately 1 kg/m3day, that UASB reactors could not
meet BOD effluent quality requirements of less than 30 mg/L. The results from the Lew
et al. (2003) study imply similar results. Their results also imply microbial hydrolysis of
the TSS in the feed will be the rate limiting treatment step at temperatures less than 14C.
This result implies an advantage for the MBR configuration, as the feed TSS will remain
in the reactor for a period of time equivalent to the reactor SRT. Seghezzo et al. (2002)
operated a pilot scale UASB reactor treating settled municipal wastewater at a
temperature near 20C. At a mean COD OLR of approximately 0.7 during a two-year
operating period, the feed and effluent UASB reactor mean values were respectively, 153
and 69 mg/L. The calculated reactor SRT was 450 days.
Barker and Stuckey (1990) concluded from a review of the literature on the topic of
SMPs (soluble microbial products) that these compounds make up a large fraction of the
soluble organic material present in the effluent from anaerobic and aerobic treatment
systems. The SMPs are comprised of a wide range of both high (i.e., greater than 50,000

14

daltons) and low (i.e., less than 500 daltons) molecular weight (MW) compounds. In a
subsequent laboratory study, the researchers operated anaerobic baffled reactors (ABRs)
treating a lower strength (i.e., COD approximately 500 mg/L) milk wastewater at
temperatures ranging from 10-35C (Barker et al., 2000). It was concluded that SMP
production increases with decreasing temperature and the production decreases with
decreasing HRT. These results imply that an anaerobic UF membrane based MBR
operated at a low HRT (e.g., corresponding to OLR of 7 kg COD/m3/day) is likely to
result in the maximum performance achievable in the anaerobic treatment of lower
strength wastewaters at lower temperatures.
The plug-flow hydraulic conditions, characteristic of properly designed ABRs, imply an
advantage for this reactor configuration in the anaerobic treatment of lower strength
wastewaters at lower temperatures, as previously noted. Manariotis and Grigoropoulos
(2002) operated a laboratory scale ABR treating a synthetic, lower strength (i.e., COD
approximately 400 mg/L) food ingredients wastewater at temperatures of 16C and 26C.
A COD removal of up to 92% was achieved at 16 at a OLR of up to 0.7 kg/m3day. It is
worthwhile noting that although the researchers observed little difference in reactor
performance at the lower temperature, a significant decrease in methane gas production
was observed. Baek and Pagilla (2003), and Wen et al. (1999) also observed low methane
gas production in operating anaerobic MBRs treating municipal wastewaters at lower
temperatures. The usual explanation for this observation is the effect of temperature on
methane gas solubility (Manariotis and Grigoropoulos, 2002). Bodik et al. (2002)
operated a pilot scale ABR containing plastic media (i.e., hybrid ABR) coupled to an
aerobic step. Treating settled municipal wastewater during low temperature conditions
(i.e., average approximately 8C, minimum 5C) over a five month period, the
researchers observed a reduction in BOD5 from 162 to 34 mg/L (mean values) across the
anaerobic-aerobic system. The system COD OLR during this period was approximately
0.4 kg/m3/day. The OLR to the hybrid ABR was approximately 0.3 kg/m3/day. The
specific performance of this treatment step was not determined.
3.3

Mechanisms Governing the Efficiency of the Membrane Component of the


Anaerobic MBR

Extensive research has been performed to investigate the mechanisms impacting the
permeate flux in an aerobic MBR. On the other hand, only a limited amount of research has
focused on the mechanisms governing the permeate flux in an anaerobic MBR. Some of the
mechanisms impacting the permeate flux in an aerobic system are likely to be similar to those
impacting the permeate flux in an anaerobic system. Considering that the physical, chemical and
biological characteristics of the mixed liquor in aerobic and anaerobic systems differ
significantly, it is reasonable to also expect that a number of mechanisms impacting the permeate
flux in an anaerobic MBR will differ from those impacting the permeate flux in an aerobic MBR
(Van Houten et al., 2001). Consequently, unless stated otherwise, only the results from studies
focusing on permeate flux in an anaerobic MBR will be discussed. In addition, the characteristics
of the mixed liquor in an anaerobic MBR are expected to vary significantly based on the type of
wastewater being treated (Kataoka et al., 1992). Although this review focuses on the treatment of
municipal wastewater, some of the results presented below are from studies that focused on
synthetic municipal wastewater or other types of wastewaters. Many of the studies are those
from which general anaerobic MBR system design and performance information was derived

15

(Table 1) and discussed in Section 3.1.1.


The mechanisms impacting the permeate flux in an anaerobic MBR can be loosely
classified into three general categories; those governed by the membrane itself (see subsection
3.3.1, membrane system), those governed by the operational parameters of the membrane (see
subsection 3.3.3, operational parameters), and those governed by the characteristics of the mixed
liquor being filtered (see subsection 3.3.3, characteristics of the mixed liquor).
The characteristics of the membrane material (e.g., polymeric versus ceramic, charge,
pore size), the membrane packing density (i.e., membrane area per unit volume), the membrane
configuration (external or submerged), and the operating conditions (i.e. surface shear and
operating TMP) are all membrane specific parameters that affect the permeate flux in an
anaerobic MBR. The membrane material, packing density and configuration are fixed design
characteristics specific to a given membrane product. Although the operating surface shear and
TMP are relatively fixed for a given membrane product, usually they be varied within a specific
range. Therefore, to maximize the permeate flux in an anaerobic MBR, an optimal membrane
product and the optimal operating conditions for that specific product must be determined.
The characteristics of the MLSS being filtered also significantly affect the permeate flux.
These characteristics are in part related to the raw wastewater being treated, but also to the
operating conditions specific to the biological component of an anaerobic MBR. As discussed
below, operating parameters such as the OLR, the SRT and HRT, as well as the operating
temperature, can significantly affect the permeate flux in an anaerobic MBR. However, these
parameters are typically selected to optimize the biological component of the system rather than
the permeate flux. As a result, an optimal set-point in terms of the biological component of an
anaerobic MBR may result in non-optimal conditions for permeate flux.
3.3.1 Membrane System
Membrane material. The type of membrane material used can significantly affect the
fouling mechanisms in an anaerobic MBR. Fouling increases the resistance the permeate must
overcome to flow through a membrane. Fouling of organic membranes (also commonly referred
to as polymeric membranes) typically results from the formation of a cake layer on the
membrane surface during filtration (Kang et al., 2002). Choo and Lee (1996a) reported that the
cake layer that forms on organic membranes in an anaerobic MBR consists of both
biological/organic solids and inorganic precipitates, and that the principal inorganic constituent
of the cake layer is struvite. However, as also reported by Choo et al. (2000), the fouling of
organic membranes appears to be predominantly governed by biological/organic interactions
with the membrane, rather than by struvite precipitation. They observed no difference in the rate
of fouling when ammonia, a component of struvite, was removed from the mixed liquor prior to
filtration using an organic membrane. For organic membranes, the resistance due to internal
fouling, which is caused by the adsorption of soluble and/or particulate material within the pore
structure of a membrane, has been reported to be significantly less than the resistance due to the
cake layer (Kang et al., 2002; Lee et al., 2001c; Choo and Lee, 1996a).
A cake layer typically does not form on an inorganic membrane, and the bulk of the
fouling can be attributed to internal fouling (Kang et al., 2002; Yoon et al., 1999). Yoon et al.

16

(1999) attributed the extensive internal fouling that occurs in inorganic membranes to the
precipitation of struvite. Using scanning electron microscope (SEM) image analysis, they
observed no visible cake layer formation on the surface of an inorganic membrane. SEM image
analysis also revealed the presence of white crystals, characteristic of struvite precipitate, within
the pore structure of the inorganic membrane. They also reported that the amount of struvite
present in the membrane as internal foulant could be estimated based on the difference between
the mass of magnesium, a component of struvite, contained in the mixed liquor and that
contained in the permeate. Kang et al. (2002) reported that the struvite content of the internal
foulant material in inorganic membranes was more than twice that observed in organic
membranes. Choo et al. (2000) also attributed the extensive internal fouling that occurs in
inorganic membranes to the precipitation of struvite. In constrast to observations with an organic
membrane, Choo et al. noted a significant difference in the rate of fouling when ammonia, a
component of struvite, was removed from the mixed liquor prior to filtration using an inorganic
membrane. These results are somewhat contradictory to those reported by Elmaleh and
Abdelmoumni (1997), who observed that the formation of a cake layer was the principal
mechanism governing the reduction in the permeate flux through an inorganic membrane.
However, as indicated in the following discussion, the cross-flow velocity at the membrane
surface can also affect the presence of a cake layer on a membrane surface.
The absence of a cake layer on inorganic membranes has been reported to result in a
slower decline in permeate flux over time than that observed for organic membranes (Kang et al.,
2002). It should be noted that the differences reported by Kang et al. (2002) could also have been
due to structural differences between the two types of membranes investigated. The inorganic
membranes studied had a smooth surface and a pore diameter of 0.14 m, while the organic
membranes had a rougher, fibrous surface and a pore diameter of 0.2 m. Ghyoot and Verstraete
(1997) also observed that the permeate flux in an anaerobic MBR with a ceramic membrane was
significantly higher than that which could be achieved with an organic membrane. However,
based on a life cycle analysis, the cost of an anaerobic MBR with a ceramic membrane was
approximately twice that of an anaerobic MBR with an organic membrane.
Hydrophobic nature and charge of the membrane. The hydrophobic nature of a
membrane material in an anaerobic MBR has been documented to significantly affect the
permeate flux. Choo et al. (2000) reported that a higher permeate flux could be maintained when
the surface of a membrane was hydrophilic in nature. Sainbayar et al. (2001) reported that the
permeate flux through a hydrophobic membrane could be increased through graft polymerization,
which introduces hydrophilic functional groups on a membrane surface. The extent of internal
pore fouling decreased as the degree of graft polymerization increased. However, graft
polymerization also affected the physical structure of the membrane surface, reducing the size of
the pores. As a result, a maximum permeate flux was observed at an intermediate degree of
grafting, at which the membrane surface exhibited primarily hydrophilic characteristics and the
membrane pores were relatively large.
Contradictory results were reported by Choo and Lee (1996b) who observed that the
extent of fouling was lower for membrane materials that were more hydrophobic in nature.
These results suggest surface hydrophobicity on its own does not govern membrane fouling
(Choo and Lee, 1996b).

17

The membrane surface charge also likely plays a significant role in membrane fouling.
The membrane surface charge is strongly impacted by the pH and the ionic strength of the mixed
liquor. Shimizu et al. (1989) reported that negatively charged inorganic membranes fouled less
rapidly than non-charged or positively charged membranes during the filtration of an anaerobic
broth. They attributed the difference to a stronger electrical repulsion between negatively
charged colloids in the broth and the membrane surface. In addition, as discussed below, the
charge that a membrane adopts during the cleaning process significantly affects the extent to
which the permeate flux can be recovered (Kang et al., 2002). However, for filtration of protein
solutions, Fane et al. (1983) reported that the impact of the membrane surface charge becomes
negligible when the ionic concentration of the solution being filtered, is high.
Nominal pore size. In addition to the characteristics of the membrane material, the
nominal pore size of a membrane can also significantly affect the permeate flux. Elmaleh and
Abdelmoumni (1997) investigated the impact of pore size on the steady state permeate flux in an
anaerobic MBR. The permeate flux was highest for a membrane with a nominal pore diameter of
approximately 0.45 m when filtering an anaerobic mixed liquor. However, when filtering a
mixed microbial population of methanogens, the optimal pore diameter was approximately 0.15
m (Elmaleh and Abdelmoumni, 1997). The differences in these results clearly indicate that the
optimal membrane pore size is a function of the specific mixed microbial population being
filtered. Choo and Lee (1996b) reported that the optimal pore size for an anaerobic MBR was 0.1
m. Chung et al. (1998) indicated that the permeate flux that could be achieved in an anaerobic
MBR with a nominal pore size of 0.22 m was three times higher than that which could be
achieved with a membrane with a pore size of 0.6 m. He et al. (1999) reported that for the
treatment of a high strength food processing wastewater using an anaerobic MBR, membranes
with a larger molecular weight cut-off fouled more rapidly and to a greater extent. These results
suggest that membranes with a larger nominal pore size foul more readily due to clogging by
macro-colloids, which can completely block the entrance of the pores, while those with a smaller
nominal pore size foul more readily due to clogging by micro-colloids, which can adsorb to the
surface of the pores.
The initial permeate flux through membranes with a larger nominal pore size tends to be
greater than through a membrane with a smaller pore size (Saw et al., 1986). However, Saw et al.
(1986) observed that the rate of fouling was also higher for membranes with a larger nominal
pore size. These results are consistent with those reported by Wen et al. (1999) who observed
that the rate of fouling in an anaerobic MBR was greater at a higher operating permeate flux, as
discussed in the subsection entitled, Operating flux (Section 3.3.2). Imasaka et al. (1989)
reported that the increase in the rate of fouling with an increase in the membrane pore size was
due mainly to an increase in internal pore fouling. The pore size had no impact on the extent of
cake fouling.
Hernandez et al. (2002) observed that membranes with a nominal pore size of 10 m
fouled more rapidly than those with a pore size of 100 m. The discrepancy can likely be

18

explained by the differences in the mechanisms that govern the fouling of coarse membranes (i.e.,
pore size greater than 10 m) and those that govern the fouling of ultrafiltration membranes.
Membrane configuration. Both external and submerged internal membrane
configurations have been used in anaerobic MBRs. The circulation rate (i.e., cross-flow velocity)
and operating TMP used in external membrane systems are typically high. For an anaerobic
MBR with an external membrane, the cross-flow velocity and operating TMP typically ranges
from 1 to 5 m/s and 2.1 to 7 kg/cm2, respectively. On the other hand, the cross-flow velocity and
operating TMP used in internal membrane systems are typically relatively low. For an anaerobic
MBR with an internal membrane, the operating TMP has been reported to range from 0.21 to 1.1
kg/cm2. The bulk cross-flow velocity tends to be less than 0.6 m/s in these systems (Lei and
Berube, 2004).
Stuckey and Hu (2003) reported that the permeate flux that could be maintained in an
anaerobic MBR with a hollow-fibre internal membrane configuration was slightly higher than
that which could be maintained in a comparable flat-sheet configuration. The difference,
although small, may be due to the dissimilarity in the extent of contact occuring between the
membrane surfaces in these two types of internal membrane systems. Lei and Berube (2004)
demonstrated that the physical contact between membranes that occurs in internal membrane
systems, which is a function of the membrane packing density and looseness, significantly
affects the permeate flux.
Unfortunately, no previous studies were identified in which the impact of the
configuration of the membrane component of an anaerobic MBR on the permeate flux was
investigated. The TMP and the cross-flow velocity can significantly affect the permeate flux in
an anaerobic MBR, as discussed in the following section. Considering that the magnitudes of
these operating parameters differ significantly for external and internal membrane systems, it can
be expected the membrane configuration will have a significant impact on the achievable
permeate flux.
3.3.2

Operational Parameters

Cross-flow velocity. Choo and Lee (1998) reported it was possible to significantly
decrease the resistance due to concentration polarization and the resistance due to cake layer
formation by increasing the cross-flow velocity. However, a plateau was reached at a Reynolds
number of approximately 2000, for which no further reduction in the resistance could be
achieved by increasing the cross-flow velocity (Choo et al., 2000). At the highest cross-flow
velocity investigated, the surface resistance (i.e. the resistance due to the concentration
polarization and the cake layer) still accounted for most of the total resistance to the permeate
flux (Choo and Lee, 1998).
For inorganic membranes, internal fouling can dominate, especially at high cross-flow
velocities (Kang, 1996). Although the extent of internal fouling is typically considered to be
independent of the cross-flow velocity, internal fouling can increase slightly as the cross-flow
velocity increases (Choo and Lee, 1998; Choo et al., 2000). This increase in the extent of internal
fouling can be attributed to thinning of the cake layer, which serves as protection against the
passage of foulants, followed by an increase in the passage of foulants into the membrane pores.

19

Elmaleh and Abdelmoumni (1997) reported that the total fouling resistance could be
reduced to virtually zero when the cross-flow velocity in a tubular membrane system exceeded 3
m/s. This result suggests that for this anaerobic MBR, the fouling was due to cake fouling only.
The permeate flux increased linearly with an increase in the surface shear stress caused by the
cross-flow velocity (Elmaleh and Abdelmoumni, 1997, 1998). However, the permeate flux
remained constant once a certain shear stress level was reached. The use of baffles to induce a
high surface shear stress at the membrane surface resulted in a similar effect on the permeate flux
as did an increase in the cross-flow velocity. The impact of baffles on the permeate flux was
greatest when the cross-flowing liquid was in the transition regime between laminar and
turbulent flow. The results suggest that the magnitude of the permeate flux at the plateau is
governed by the mass flux of solids towards the membrane. Saw et al. (1986) reported that the
permeate flux in an anaerobic MBR increased to a greater extent with an increase in the crossflow velocity, when the flow through the membrane was turbulent. Imaskaka et al. (1989) also
reported that permeate flux increased with an increase in the cross-flow velocity. However, they
noted that when varying the cross-flow velocity, the permeate flux at a given cross-flow velocity
was dependent on the step-wise manner in which the cross-flow velocity was changed. Also,
Grethlein (1978) reported that the rate of fouling decreased as the cross-flow velocity increased.
However, as noted by Bourgeous et al. (2001), although the permeate flux can be increased by
increasing the cross-flow velocity, this increase comes at a cost. They reported that an increase in
the cross-flow velocity from 1 to 2 m/s increased the permeate flux by 20%. However, the power
cost for the system was also increased, but by 58%. In addition, the high cross-flow velocity
required to generate high shear conditions can generate large axial pressure gradients, resulting
in a non-uniform TMP in tubular membrane systems (Lee et al., 1999). As a consequence, some
sections of the membrane can be subject to non-optimal TMP conditions.
The cross-flow velocity can also negatively impact the permeate flux in an anaerobic
MBR. Brockmann and Seyfried (1996), Ghyoot and Verstraete (1997) and others, reported that
the biomass activity could be significantly affected by shear, as discussed in Section 3.1.1. Choo
and Lee (1998) reported that the higher shear forces imposed on mixed liquor at higher crossflow velocities can reduce the size of the particulate material (i.e. biomass). The size of the
biosolid particles in the mixed liquor can significantly affect the permeate flux, as discussed in
the subsection entitled, Colloidal solids (in subsection 3.3.3). In addition, Choo and Lee (1996a)
suggested that the high shear conditions caused by high cross-flow velocities can significantly
increase cell lysis, resulting in a decrease in the overall activity of the biomass in an anaerobic
MBR. For an aerobic MBR, Kim et al. (2001) suggested that the high shear conditions present in
a bioreactor could result in the release of high concentrations of exocellular polymeric
substances (EPS) or SMP into the bioreactor. High concentrations of EPS or SMP have been
documented to negatively impact the permeate flux in aerobic MBRs (Van Houten et al., 2001;
Lawrence et al., 2001). Fortunately, anaerobic biosolids do not appear to be as impacted
significantly by high cross-flow velocities in comparison to aerobic biosolids (Elmaleh and
Abdelmoumni, 1997).
The high shear forces to which the biomass is exposed in an aerobic external membrane
MBR is mainly due to the recirculation pumping that is required in this type of configuration
(Shimizu et al., 1994). Gear pumps and positive displacement pumps have been observed by
some researchers to have the largest negative impact on biomass activity in an aerobic MBR,

20

while centrifugal pumps had the smallest impact (Flaschel et al., 1986). This may explain why
Beaubien et al. (1996) did not observe a negative impact of high cross-flow velocities on
methanogenic activity. Unfortunately, the authors did not disclose the type of recirculation pump
used.
Gas sparging. Gas sparging is extensively used in submerged internal membrane
systems as a means to provide high shear conditions at the membrane surface. Increasing the
amount of gas increases the amount of shear to which the membrane surface is exposed, much in
the same manner as an increase in the cross-flow velocity in an external tubular membrane
system increases the surface shear.
Air is typically used as the sparging gas in an aerobic MBR. Vera et al. (2000) reported
that for an aerobic MBR, the extent of fouling decreased as the air flow rate increased. Lee et al.
(2001c) reported that it was possible to maintain a relatively high permeate flux in an anaerobic
MBR by sparging the submerged membrane system with air. However, the membranes could
only be sparged for approximately five seconds every 10 minutes. Sparging the system with air
for a longer duration resulted in non-anaerobic conditions that significantly reduced the activity
of the acid-forming microorganisms in the system. More extensive sparging is required to
maximize the permeate flux in an anaerobic MBR.
Stuckey and Hu (2003) effectively used the gas in the headspace in an anaerobic MBR as
a source of relatively inert gas for continuously sparging an internal membrane system. The TMP
required to maintain a constant permeate flux decreased as the gas sparging flow increased.
However, a plateau was reached at which no additional significant reduction in the required TMP
could be achieved by increasing the gas sparging flow (Stuckey and Hu, 2003). Kayawake et al.
(1991) reported that the permeate flux that could be maintained in an anaerobic MBR with an
internal ceramic membrane system could be doubled by sparging the system with head-space gas.
Imasaka et al. (1989) used nitrogen gas for sparging in an anaerobic MBR. The permeate flux
increased as the nitrogen gas sparging rate increased, up to a certain value, after which a further
increase in the sparging rate did not result in a significant increase in the permeate flux. It was
also observed at the higher sparging rates investigated, the permeate flux tended to continuously
decrease over time, while at the lower sparging rates, the permeate flux tended to reach a pseudosteady state value. They attributed the continuous decrease in the permeate flux to the thinning of
the cake layer, which can occur at higher sparging rates. As noted above, this thinning of the
cake layer, which can serve as a protective layer against the passage of foulants, may result in an
increase in the passage of foulants into the membrane pores. Fawehinmi et al. (2004) also
reported a higher permeate flux when sparging an AnMBR with nitrogen gas during the
treatment of a high strength synthetic wastewater.
TMP. The permeate flux in an anaerobic MBR is governed by different mechanisms
when the membrane is operated at low or at a high TMPs. At a relatively low TMP, the permeate
flux is governed by the TMP. Under such pressure-limited conditions, the permeate flux
increases linearly with the applied TMP, and the permeate flux is not significantly impacted by
the cross-flow velocity (Beaubien et al., 1996). The permeate flux is, however, impacted by the
MLSS concentration, but only at low concentrations (i.e. less than 2.5 g/L) according to
Beaubien at al. (1996). At higher solids concentrations, the permeate flux is not impacted by the
concentration of the MLSS.

21

At a relatively high TMP, the permeate flux is governed by the mass transfer of material
away from the membrane surface. Under mass transfer-limited conditions, the permeate flux in
an anaerobic MBR is governed by the cross-flow velocity (i.e. surface shear) and the MLSS
concentration (Beaubien et al., 1996). Beaubien et al. (1996) reported a linear increase in the
permeate flux with an increase in the cross-flow velocity along the membrane surface at high
TMPs. However, the magnitude of the increase in the permeate flux was lower at higher MLSS
concentrations. The lower magnitude of the increase in the permeate flux can be attributed to the
higher rate of mass transfer towards the membrane and/or to the increase in the viscosity of the
mixed liquor that occurs at higher MLSS concentrations.
Under mass transfer-limited conditions, the permeate flux theoretically is not affected by
the TMP. However, at very high TMPs, Elmaleh and Abdelmoumni (1997) reported a decrease
in the permeate flux with an increase in TMP. The decrease in the permeate flux under such
conditions was attributed to a compaction of the foulant layer. When filtering a digested sludge,
Saw et al. (1986) also observed that at very high operating TMPs, the permeate flux in an MF
membrane decreased with an increase in the TMP. However, they also observed that when using
UF membranes (MW cut-off 8,000 to 20,000 daltons), the permeate flux remained constant with
an increase in the TMP. They suggested that the structure of the foulant layer forming on an MF
membrane is not as dense as the layer that forms on a UF membrane, and is therefore more
susceptible to collapsing under elevated TMPs.
Beaubien et al. (1996) suggested that for relatively high pressure systems, it is possible to
identify a TMP that maximizes the permeate flux while minimizing membrane fouling. The
optimal operating pressure could be calculated using Equation 1,
R
P
= m,
opt
a

(1)
where Popt is the optimal TMP, Rm is the resistance due to membrane-solute
interactions (i.e. resistance due to pore plugging and adsorption), and a is a mass transfer
parameter.
According to Equation 1, when fouling is caused predominantly by the formation of a
cake layer on the membrane surface (i.e. the Rm is small), the optimal TMP is low. If internal
fouling governs (i.e. Rm is large), the optimal TMP is high. Although the permeate flux for high
TMP systems was reported to be a function of both the cross-flow velocity and the concentration
of suspended solids in the mixed liquor, the optimal TMP was reported to be independent of the
TSS concentration (Beaubien et al., 1996). The dependence of the optimal TMP on the crossflow velocity is somewhat intuitive since high cross-flow velocities tend to remove the cake
layer, making internal fouling the dominant fouling mechanism. The operation of a membrane
system under high TMPs and high cross-flow velocities can be relatively expensive in terms of
operating costs. Sutton et al. (2002) reported that usually it is not possible to maintain a relatively
high permeate flux for a long period of time under these conditions, with an aerobic internal
membrane MBR system.

22

Operating flux. The ability to maintain a high permeate flux can decrease both the
capital and operating cost associated with an anaerobic MBR, as previously discussed. However,
at a higher permeate flux, the rate of mass transfer of material towards the membrane surface is
also greater. As a result, the rate of fouling (i.e. the accumulation of foulant material on the
membrane surface) in an anaerobic MBR has been reported to be greater at a higher operating
permeate flux (Wen et al., 1999). The more rapidly a membrane fouls, the more often it must be
cleaned. Therefore, it may not be advisable to attempt to maintain the highest possible permeate
flux. A balance between a high permeate flux and long filtration runs must be maintained in
order to maximize the total permeate volume achieved over time.
Permeate flux recovery (i.e. membrane cleaning). Lee et al. (2001c) reported that it
was not possible to recover the permeate flux through an organic membrane by back-flushing the
membrane with a caustic solution. Caustic solutions are considered to be effective for removing
organic/biological foulants from a membrane surface, while acidic solutions are considered to be
effective for removing inorganic foulants from a membrane surface (Lee et al., 2001c). However,
a number of studies have reported that it is possible to consistently recover the permeate flux
through an organic membrane by back-flushing the membrane exclusively with an acidic
solution (Kang et al., 2002; Choo et al, 2000), or with a caustic solution following an acidic
cleaning (Lee et al., 2001c). Considering that the fouling of organic membranes can be attributed
mainly to the formation of a cake layer, which consists of biomass and struvite (Kang et al.,
2002), these results suggest that the removal of struvite governs the recovery of the permeate
flux.
A number of studies have reported that it is not possible to consistently recover the
permeate flux through an inorganic membrane by back-flushing with an acidic solution,
regardless of the type of acidic solution used (Yoon et al.,1999; Kang et al., 2002; Choo et al,
2000). These results are somewhat counter-intuitive since the internal pore fouling in inorganic
membranes has been attributed mainly to struvite (Yoon et al.,1999; Kang et al., 2002), which is
soluble under acidic conditions. Kang et al. (2002) attributed the poor recoveries observed when
back-flushing an inorganic membrane with acidic solutions to the positive charge adopted by the
inorganic membrane during acidic cleaning. They suggested that the positive charge could result
in strong attractive interactions between the membrane surface and the various solutes and
colloids in the mixed liquor.
Relaxation, which consists of periodic interruptions of the filtration process by reducing
the TMP to zero, is also extensively used in MBRs to increase the permeate flux. Wen et al.
(1999) investigated a number of relaxation scenarios with permeation and relax times ranging
from 2-8 minutes and 0.5-2 minutes, respectively. Their results indicated that the permeate flux
was highest at intermediate permeate times (i.e. 4 minutes) and intermediate relax times (e.g.. 1
minute). Grethlein (1978) also reported that the rate of decline in the permeate flux could be
minimized using this approach.
Operating temperature. Baek and Pagilla (2003) reported that higher operating
temperatures could be maintained in anaerobic MBRs compared to their aerobic counterparts.
The difference in the achievable operating temperatures can be attributed to the cooling effect of
the aeration system in the aerobic MBR. Higher operating temperatures can have beneficial
impacts on the permeate flux by reducing the viscosity of the permeate. Hogetsu et al. (1992)

23

reported an increase in the permeate flux of over 30 percent when the operating temperature was
increased from 40C to 47C. Similar results were reported by Zoh et al. (2002). Schiener et al.
(1998) and Fawehinmi et al. (2004) observed a slight decrease in the concentration of SMPs in a
conventional anaerobic bioreactor as the operating temperature increased. Therefore, operating
an anaerobic MBR at an elevated temperature may have beneficial impacts on the permeate flux
by reducing the concentration of SMPs in the system, as discussed in the subsection entitled,
Soluble products (in subsection 3.3.3).
Most of the results presented herein regarding membrane efficiency are from studies that
were performed using an anaerobic MBR operating at temperatures in excess of 30C. However,
a number of studies were performed at ambient or moderate temperatures. Wen et al. (1999)
were able to maintain a relatively high permeate flux in an anaerobic MBR operated at
temperatures ranging from 14-25C over an extended period of time, when the membrane was
operated with a relaxation period. Kiriyama et al. (1994) also successfully operated an anaerobic
MBR at temperatures ranging from 20-25C. They did not report the magnitude of the permeate
flux achieved.
Pretreatment approaches. PAC addition has also been used to enhance the permeate
flux in anaerobic MBRs. Park et al. (1999) reported that at relatively low cross-flow velocities,
PAC addition did not significantly affect the permeate flux. However, at higher cross-flow
velocities, the addition of PAC resulted in an increase in the flux. The impact of PAC addition on
the flux was greater at a higher PAC dosage. Park et al. (1999) attributed the results to the
scouring effect of the PAC on the membrane surface and to the PAC adsorption of
dissolved/colloidal material from the mixed liquor. Similar results were observed by Pirbazari et
al. (1996) who investigated the impact of PAC addition in an aerobic MBR system. Kim and Lee
(2003) attributed a higher permeate flux observed following PAC addition to an aerobic MBR, to
a reduction in the quantities of fine colloids and soluble microbial products in the mixed liquor,
as discussed in the subsections entitled, Colloidal solids and Soluble products (in subsection
3.3.3). They also observed that the effect of PAC on the permeate flux was more pronounced for
a submerged internal membrane system than for an external membrane system. They attributed
the difference to the more extensive floc breakage which occurred in the external membrane
system. Choo and Lee (1996b) suggested that the addition of an adsorbent or a coagulant could
also enhance permeate flux by agglomerating the fine colloids present in the mixed liquor being
filtered into larger particles that have a lower tendency to foul membranes, as discussed in the
subsection entitled, Colloidal solids (Section 3.3.3).
Imasaka et al. (1989) investigated the addition of an ion exchange resin to an anaerobic
MBR to enhance the scouring effect of the cross-flow at the membrane surface, and as a result,
to reduce the thickness of the foulant layer. The addition of the ion exchange resin at a
concentration of 2.5% solids did not impact the permeate flux. However, the addition of the resin
at a concentration of 5% solids doubled the permeate flux.
The removal of other material that can contribute to membrane fouling has also been
investigated. Choo et al. (2000) reported that struvite formation can be minimized by combining
a dialysis/zeolite system with an anaerobic MBR. The dialysis/zeolite component of the system
can selectively remove ammonia. With such a combined approach, the permeate flux through an
inorganic membrane was reported to have increased by 15-20%.

24

Miscellaneous operating conditions. Shimizu et al. (1989) reported that the permeate
flux that could be maintained in an anaerobic MBR operated under stable, steady state conditions
was approximately twice that which could be maintained under non-stable conditions. These
results may be due to the different amounts of SMPs that are produced under steady state and
non-steady state conditions.
Kang et al. (2003) observed a strong correlation between the DO concentration in an
anoxic MBR and the specific cake resistance. The resistance was significantly higher at a DO
concentration of 0.3 mg/L than at a concentration of 5 mg/L. They suggested that the differences
could be attributed to the larger flocs that are typically more prevalent at higher DO
concentrations. Smaller solids tend to contribute to membrane fouling to a greater extent than
large flocs, as previously discussed. Kim and Somiya (1999) investigated the impact of
intermittent ozone gas sparging on flux recovery in an anaerobic MBR. The permeate flux that
could be maintained with intermittent ozonation was almost twice that which could be
maintained without ozonation. The specific mechanisms which resulted in an increase in the
permeate flux were not examined. However, the authors noted extensive ozonation could inhibit
microbial activity in the anaerobic MBR.
3.3.3 Characteristics of the Mixed Liquor
Suspended solids. The concentration of suspended solids in the mixed liquor in an
anaerobic MBR has been reported to have a significant impact on the resistance to the permeate
flux. Stuckey and Hu (2003) observed that the TMP required to maintain a constant permeate
flux in an anaerobic MBR treating a synthetic wastewater at an MLSS concentration of 35 g/L
was over two times greater than the pressure required at an MLSS concentration of 7 g/L. Saw et
al. (1986) observed a log-linear decrease in the steady state permeate flux with an increase in the
concentration of suspended solids when filtering a digested sludge. The extent of the decline was
greater for membranes with larger pore sizes. Kitamura et al. (1996) also observed a decrease in
the permeate flux with an increase in the concentration of suspended solids in an anaerobic MBR
treating distillery wastewater. The authors noted that the permeate flux did not increase to the
same extent when the suspended solids concentration was decreased. However, the exact
relationship between the concentration of suspended solids and the steady state permeate flux in
an anaerobic MBR has not been extensively investigated. Yamazaki et al. (1997) also observed a
decrease in the permeate flux that could be maintained with an increase in the MLSS
concentration in an anaerobic MBR.
A number of studies have documented the impact of TSS on the permeate flux in aerobic
MBRs. In general, the steady state permeate flux has been reported to decrease at higher
suspended solids concentrations. This is likely due to the higher rate of mass transfer of material
towards the membrane surface that occurs at higher TSS concentrations. The impact of the
concentration of suspended solids in an aerobic MBR has been reported to also be a function of
the hydrodynamic conditions in the system. Lubbecke et al. (1995) reported that at lower
concentrations, the steady state permeate flux in an aerobic MBR was not impacted by
suspended solids. However, above a specific concentration, which was dependent on the crossflow velocity, the steady state permeate flux decreased as the concentration of TSS in the system
increased. The results suggest that at low concentrations of TSS, the rate of mass transfer of

25

solids towards the membrane surface is less than the rate of mass transfer of TSS away from the
membrane surface. As the concentration of suspended solids increases, the viscosity of the mixed
liquor increases. At a certain point, the increase in the viscosity of the mixed liquor will cause a
shift from turbulent to laminar flow conditions along the membrane. The rate of mass transfer of
suspended solids away from the membrane surface, which is largely governed by eddy diffusion,
is much lower under laminar flow, than turbulent flow conditions (Mallevialle et al., 1996).
Colloidal solids. Choo and Lee (1996b, 1998) reported that fine colloids played a critical
role in increasing the hydraulic resistance of a foulant layer in an anaerobic MBR. Fine colloids
tend to have a lower back-diffusion rate than larger solids. Choo and Lee (1998) suggested that
as a consequence of this lower back-diffusion rate, fine colloids tend to migrate to and
accumulate at the membrane surface to a greater extent than larger suspended solids. In addition,
they suggested that smaller particles tend to form a more compact foulant layer on the membrane
surface. Therefore, not only do smaller solids tend to accumulate at the membrane surface, the
resulting foulant layer is more compact. Choo and Lee (1996b) also reported that the polarization
index at the membrane surface for colloidal material was much higher than for soluble material
or for microorganisms contained in an anaerobic digestion broth. They speculated that flux
improvements could be obtained by degrading the colloidal material into soluble material or by
agglomerating the colloidal material into coarser particles, based on these results. Langenhoff et
al. (2000) observed that the production of SMPs in a conventional anaerobic bioreactor fed a
synthetic wastewater was higher when the colloidal content of the wastewater was higher. This
production of SMPs could enhance membrane fouling as discussed in the following subsection.
In a review of recent developments in anaerobic MBR technology, Van Houten et al. (2001)
suggested that since an anaerobic mixed liquor tends to contain more fine colloids than an
aerobic mixed liquor, the mechanisms that govern fouling in anaerobic systems are likely to be
different from the mechanisms that govern fouling in aerobic systems. No specific data or results
were presented.
Soluble products. Although the permeate flux in an anaerobic MBR has been reported to
be significantly impacted by the concentrations of suspended and colloidal solids in the mixed
liquor, the soluble component of the mixed liquor appears to play a significant, if not greater role
in the formation of a foulant layer on the membrane component of the system (Harada et al.,
1994)
For aerobic MBRs, the extent of fouling has been documented to be related to the
concentration of SMPs in the mixed liquor (Fawehinmi et al., 2004; Lee et al., 2001a; Chang and
Lee, 1998; Wisniewski and Grasmick, 1998). Lee et al. (2001a) reported that a higher permeate
flux could be maintained in a suspended growth aerobic MBR than could be maintained in an
fixed-film aerobic MBR. These results are counter-intuitive, considering the impact of suspended
solids on the permeate flux in MBRs, reported in the subsection entitled Suspended solids. This
may indicate that the soluble products play a significant role in membrane fouling. In addition,
these results suggest that there may be no benefit to including attached growth surfaces in an
anaerobic MBR to reduce the concentration of TSS in the solution being filtered. Lee et al.
(2001a) also observed that for both fixed-film and suspended growth aerobic MBRs, the rate of
fouling was lower at higher TSS concentrations, and the rate of fouling was lower when filtering
a mixture of suspended solids and SMPs, in contrast to filtration of a solution containing only
soluble microbial products. These results are consistent with those reported by Shin and Kang

26

(2002). When investigating the permeate flux in an aerobic-anoxic MBR, they observed most of
the membrane resistance was induced by soluble components in the mixed liquor.
To date, the impact of SMPs on the permeate flux in an anaerobic MBR has received
limited attention. Stuckey observed that the type of SMPs produced in an anaerobic MBR and
those present in the effluent permeate were different. These results suggest that the membrane
component of the anaerobic MBR retained some of the SMPs and that these are likely to have
contributed to the formation of a foulant layer on the membrane surface.
Stuckey (2003) also reported that the type of SMP that predominates in an anaerobic
MBR was a function of both the influent load and the composition of the wastewater. This is
consistent with results reported for conventional anaerobic bioreactors. Barker and Stuckey
(2001) reported that the quantities of SMPs formed in a conventional anaerobic bioreactor fed
with synthetic wastewater increased with the COD concentration of the wastewater being treated.
These results are somewhat contradictory to those reported by Langenhoff et al. (2000).
In this case, higher concentrations of SMPs were observed at lower hydraulic loading rates for a
conventional anaerobic bioreactor treating a synthetic wastewater. This is consistent with the
results reported by Kayawake et al (1991) for an anaerobic MBR treating municipal wastewater
solids. It was reported that a higher permeate flux could be maintained when the anaerobic MBR
was operated at a higher loading rate. Barker et al. (2000) suggested that the higher production of
SMPs in a conventional anaerobic bioreactor at longer hydraulic retention times is likely due to
the more extensive biomass decay that occurs at longer SRTs. This hypothesis is consistent with
results reported by Shin and Kang (2002). For an aerobic-anoxic MBR, they observed that the
resistance induced by the soluble fraction of the mixed liquor was more severe at long SRTs.
The production of SMPs and the impact of these compounds on the permeate flux in an
anaerobic MBR remains unclear. It is likely that an intermediate loading rate can maximize the
permeate flux in an anaerobic MBR, as suggested by Hernandez et al. (2002). They reported that
the permeate flux was greater at a medium loading rate (1.5 to 10 kg/m3day) than at a low
loading rate (0.3 to 1 kg/m3day).
Barker et al. (2000) reported that 22% of the effluent COD from a conventional anaerobic
bioreactor treating a low strength synthetic wastewater consisted of high MW SMPs. This
fraction was found to be highly biodegradable under aerobic conditions, with 86% of the COD
being biodegradable. Only 4% of the COD was biodegradable under anaerobic conditions. The
low MW fraction of the SMPs formed during anaerobic treatment accounted for 36%t of the
effluent COD. The authors reported that 33% of this fraction was biodegradable under anaerobic
conditions, while only 17% was biodegradable under aerobic conditions. These results suggest it
may be beneficial to add an aerobic polishing step prior to membrane filtration in an anaerobic
MBR to remove the aerobically biodegradable SMPs.
Barker et al. (2000) and Schiener et al. (1998) reported that the quantity of SMPs
produced in a conventional anaerobic bioreactor treating a synthetic low strength wastewater
increased as the operating temperature decreased. Barker et al. (2000) speculated that the greater
concentration of SMPs was due to a reduction in the rate of biodegradation of these products at
lower temperatures. Similar results were reported by Fawehinmi et al. (2004) from a study of

27

AnMBR treatment of a high strength synthetic wastewater.


Inorganic precipitates/struvite. The conditions in an anaerobic MBR are ideal for the
formation of struvite (Mamais et al., 1994). Both ammonia and phosphate are usually abundant,
and the pH of the mixed liquor in an anaerobic MBR typically ranges from 7.5 to 8.5, which
promotes the precipitation of struvite.
Choo and Lee (1996a) reported that struvite contributed significantly to the fouling of
membranes in an anaerobic MBR. The amount of struvite precipitated could be estimated based
on a mass balance analysis of the concentration of magnesium in the influent and effluent of an
anaerobic MBR and the concentration of ammonia and phosphate in the mixed liquor. The extent
to which struvite can affect the permeate flux in an anaerobic MBR has been reported to be
impacted by the type of membrane used. The cake layer that forms on organic membranes in an
anaerobic MBR consists of both biological/organic solids and inorganic precipitates, and the
principal inorganic constituent of the cake layer is struvite, as discussed in the subsection entitled,
Membrane material (Section 3.3.1). However, the fouling of organic membranes is
predominantly governed by biological/organic interactions with the membrane. On the other
hand, the fouling of inorganic membranes can mainly be attributed to internal pore fouling by
struvite.

4.0

SUMMARY: PARAMETERS CRITICAL TO ANAEROBIC MBR SYSTEM


PERFORMANCE AND EFFICIENCY

An analysis of the relevant compiled literature and other information, on the topic of
anaerobic MBR wastewater treatment, was completed. The principal objective of the information
analysis was to identify the variables that appear to be critical to the achievment of a high
effluent quality in an economical fashion, in applying anaerobic MBR technology for treatment
of low-to-medium strength wastewaters. The results from the information analysis clearly imply
that the anaerobic MBR technology has the potential to achieve near complete treatment (i.e.,
CBOD5 equal to or less than 30 mg/L) of municipal and/or low-to-medium strength industrial
wastewaters operating under moderate temperature conditions (i.e., less than 20C). The
technology may be particularly attractive in the management of municipal wastewaters using a
less traditional approach. Treatment system economics may favor application of anaerobic MBR
systems as a satellite plant located on a municipal wastewater trunk sewer, designed to reduce
the organic load to an existing, centralized treatment facility. The information analysis results
imply that anaerobic MBR technology has the potential to be applied at a commercial scale for
treatment of a wide variety of industrial wastewaters, or as a unit operation in a variety of
flowsheets involving the treatment of municipal wastewater solids and sludges. The results
suggest that the efficiency of the membrane component likely will dictate treatment system
economics. A number of mechanisms govern the permeate flux during operation of membrane
systems coupled to anaerobic bioreactors.

28

The results from the information analysis follows.


4.1

Critical Bioreactor Parameters

The bioreactor parameters that appear to be critical, relevant to the principal objective of
the information analysis stated previously, are the following.
Bioreactor operating temperature. The information analysis clearly indicates the
bioreactor operating temperature will significantly affect treatment efficiency. The results imply
that an anaerobic MBR should be capable of achieving near complete treatment of low-tomedium strength wastewaters operating under moderate (15-25oC) temperature conditions.
Bioreactor OLR. The bioreactor COD OLR and effectively the operating SRT, will
dictate treatment performance. Results from the information analysis imply that the anaerobic
bioreactor of an MBR system should be capable of operating in the range from 5-8 kg
COD/m3day and should achieve near complete removal of carbonaceous BOD from municipal
wastewater under moderate temperature conditions.
Bioreactor configuration and biomass characteristics. The membrane component
plays a critical role in dictating the performance and efficiency of the anaerobic MBR, by
allowing for capture and hydrolysis of wastewater organic colloids and particulates, physical
retention of certain soluble organics and maintenance of a high concentration of active
methanogens in the bioreactor component. The bioreactor must be capable of operating at an
effective concentration of at least 10 g/L, of VSS to achieve efficient treatment. The efficiency of
the membrane component will dictate the optimal bioreactor configuration with respect to the
nature of the biological growth (i.e., suspended growth versus fixed-film versus hybrid reactor).
An anaerobic bioreactor configuration with plug-flow hydraulic characteristics is likely to be
favored in the treatment of lower strength wastewaters.
4.2

Critical Membrane System Parameters

The information analysis implies that the optimal membrane system for an anaerobic
MBR consists of an organic, hydrophilic and negatively charged membrane with a pore size of
approximately 0.1 m. The use of both external and submerged internal membrane
configurations shows promise. Other design and operational parameters dictating the
performance and efficiency of the membrane component, follow.
The operating parameters impacting the permeate flux in an external membrane system
are the TMP and the cross-flow velocity. The operating parameters impacting the permeate flux
in an internal membrane system are the TMP, the sparging intensity and the duration of the
relaxation period. The optimal set-points for these different operating parameters need to be
determined.
Both the cross-flow velocity and the sparging intensity impart a significant amount of
shear on the biomass in an anaerobic MBR. High shear forces can reduce the microbial activity
in an anaerobic MBR. In addition, high shear forces can reduce the size of the biosolids in the
mixed liquor and increase the release of SMPs. In this respect, external and internal membrane
systems are expected to perform differently since the magnitude of the shear forces to which the

29

biomass is exposed in an external membrane system is significantly greater than that in a


submerged system. The impact of the shear forces needs to be considered when investigating the
optimal set-points for the different operating parameters.
4.3

Mixed Liquor Characteristics

The size of the biosolids and the amount of SMPs in the mixed liquor affect the permeate
flux. The exact size fraction and the components of the soluble microbial products that have the
greatest impact on the permeate flux need to be determined. Higher concentrations of SMPs may
be present in mixed liquor when an anaerobic MBR is operated at low temperatures. Aerobic
polishing following anaerobic treatment, can potentially reduce the concentration of some
components of the SMPs in the mixed liquor. Both the impact of the operating temperature and
aerobic polishing, on the concentration/composition of SMPs and permeate flux, need to be
investigated.
It is not possible to remove the foulant layer on an organic membrane with caustic
cleaning alone. Acidic cleaning or acidic cleaning followed by caustic cleaning is required to
remove the foulant layer. This suggests that both biological/organic and inorganic material
contributes to membrane fouling. Further research is required to investigate the composition of
the foulant layer.

30

5.0

REFERENCES

Anderson, G.K., Saw, C.B. and Fernandes, M.I.A.P. (1986). Application of porous membranes
for biomass retention in biological wastewater treatment processes, Process Biochemistry,
December, p. 174.
Anderson, G.K., Kasapgil, B., and Ince, O. (1996). Microbial kinetics of a membrane anaerobic
reactor, Environmental Technology, Vol. 17, p. 449.
Angenent, L.T., Banik, G.C., and Sung, S. (2000). Psychrophilic anaerobic pretreatment of lowstrength wastewater using the anaerobic migrating blanket reactor, Proceedings of the WEF
73rd Annual Conference & Exposition, Anaheim, California, October.
Arros, S., Torrijos, M., Gesan-Guiziou, G., Lafforgue, C., Daufin, G., and Merin, U. (2001). Is
the anaerobic membrane bioreactor a valuable process for purification of end-of-pipe dairy
wastewater with reduced sludge production? International Conference: Membrane Technology
for Wastewater Reclamation and Reuse, Ed. G. Oron and A. Bick, p. 98, Tel Aviv, Israel,
September.
Arros-Alileche, S., Gesan-Guiziou, G., Lafforgue, C., Daufin, G., and Merin, U. (2002). Whats
the membrane role in an anaerobic membrane bioreactor? Simulation, Proceedings of the
Membranes in Drinking and Industrial Water Production Conference, Mulheim/Ruhr, Germany,
September.
Baek, S.H. and Pagilla, K.R. (2003). Anaerobic membrane bioreactor for dilute municipal
wastewater treatment, Proceedings of the WEF 49 Annual Conference & Exposition, Los
Angeles, California, October.
Bailey, A. D., Hansford, G. S., and Dold, P. L. (1994). The enhancement of upflow anaerobic
sludge bed reactor performance using crossflow microfiltration, Water Research, Vol. 28, No. 2,
p. 291.
Barker, D.J. and Stuckey, D.C. (1999). A review of soluble microbial products (SMP) in
wastewater treatment systems, Water Research, Vol. 33, No. 144, p. 3063.
Barker, D.J., Salvi, S.M.L., Langenhoff, A.A.M., and Stuckey, D.C. (2000). Soluble microbial
products in ABR treating low-strength wastewater, Journal of Environmental Engineering, Vol.
126, No. 3, p. 239.
Barker, D.J. and Stuckey, D.C. (2001). Modelling of soluble microbial products in anaerobic
digestion: the effect of feed strength and composition, Water Environment Research, Vol. 73,
No. 2, p. 173.
Barnes, S., Dalhoff, R., Keller, J., Wilderer, P., and Kendall, L. (2003). Investigation of
membrane processes for the removal of volatile fatty acids, Water Science & Technology, Vol.
47, No. 12, p. 191.
Beaubien, A., Baty, M., Jeannot, F., Francoeur, E., and Manem, J. (1996). Design and operation
of anaerobic membrane bioreactors: development of a filtration testing strategy, Journal of
Membrane Science, Vol. 109, No 2, p. 173.
Berube, P.R. and Hall, E.R. (1999). Effects of kraft evaporator condensate matrix on methanol
removal in a high temperature membrane bioreactor, Water Science & Technology, Vol. 40, No.
11-12, p. 327.
Berube, P.R., Ronteltap, M., and Hall, E.R. (2003). Impact of recycling rate on the specific
substrate utilization coefficient and the observed growth yield in a membrane bioreactor,
submitted for publication.

31

Bilstad, T., Madland, M., Espedal, E., and Hanssen, P.H. (1992). Membrane separation of raw
and anaerobically digested pig manure, Water Science & Technology, Vol. 25, No. 10, p. 19.
Bodik, I., Kratochvil, K., Herdova, B., Tapia, G., and Gasparikova, E. (2002). Municipal
wastewater treatment in the anaerobic-aerobic baffled reactor at ambient temperature, Water
Science & Technology, Vol. 46, No. 8, p. 127.
Bonkoski, W.A. (2003). Personal communication to P.M. Sutton, Vice President, Zenon
Environmental, December.
Botha, G.R., Sanderson, R.D., and Buckley, C.A. (1992). Brief historical review of membrane
development and membrane applications in wastewater treatment in Southern Africa, Water
Science & Technology, Vol. 25, p. 10.
Bourgeous, K.N., Darby, J.L., and Tchobanoglous, G. (2001). Ultrafiltration of wastewater:
effects of particles, mode of operation, and backwash effectiveness, Water Research, Vol. 35.
p.77.
Brockman, M. and Seyfried, C.F. (1996). Sludge activity and cross-flow microfiltration a
non-beneficial relationship, Water Science & Technology, Vol. 34, No. 9, p. 205.
Brockman, M. and Seyfried, C.F. (1997). Sludge activity under the conditions of cross-flow
microfiltration, Water Science & Technology, Vol. 35, No. 10, p. 173.
Budd, W.E. and Okey R.W. (1969). U.S. patent 3,472,765.
Buswell, A.M., Boruff, C.S., and Wiesman (1932). Anaerobic stabilization of milk waste,
Industrial and Engineering Chemistry, Vol. 24, p. 1423.
Cadi, Z., Huyard, H., Manem, J., and Moletta, R. (1994). Anaerobic digestion of a synthetic
wastewater containing starch by a membrane reactor, Environmental Technology - London, Vol.
15, No. 11, p. 1029.
Chang, I-S. and Lee, C-H. (1998). Membrane filtration characteristics in membrane-coupled
activated sludge system the effect of physiological states of activated sludge on membrane
fouling, Delsalination, Vol. 120, p. 221.
Cheryan, M. and Mehaia, M.A. (1986). Membrane bioreactors, Chemtech, November, p. 676.
Choate, W.T., Houldsworth, D., and Butler G.A. (1983). Membrane enhanced anaerobic
digesters, Proceedings 37th Industrial Waste Conference, Purdue University, West Lafayette,
Indiana, Ann Arbor Science, Ann Arbor, Michigan, p. 661.
Choo, K-H. and Lee, C-H. (1996a). Membrane fouling mechanisms in the membrane-coupled
anaerobic bioreactor, Water Research, Vol. 30, No. 8, p. 1771.
Choo, K.-H. and Lee, C-H. (1996b). Effect of anaerobic digestion broth composition on
membrane permeability, Water Science & Technology, Vol. 34, No. 9, p. 173.
Choo, K-H., and Lee, C-H. (1998). Hydrodynamic behavior of anaerobic biosolids during
crossflow filtration in the membrane anaerobic bioreactor, Water Research, Vol. 32, No. 11, p.
3387.
Choo, K-H., Kang, I.J., Yoon, S.H., Park, H., Kim, J.H., Adlya, S., and Lee, C-H. (2000).
Approaches to membrane fouling control in anaerobic membrane bioreactors, Water Science &
Technology, Vol. 41, No. 10-11, p. 363.
Chung, Y-C., Jung, J-Y., Ahn, D-H. and Kim, D-H. (1998). Development of two phase
anaerobic reactor with membrane separation system, Journal of Environ Sci HealthToxic/Hazard Subst Environ Eng, Vol. A33, No. 2, p. 249.

32

DeFrance, L. and Jaffrin, M. (1999). Comparison between filtrations at fixed transmembrane


pressure and fixed permeate flux: application to a membrane bioreactor used for wastewater
treatment, Journal of Membrane Science, Vol. 152, p. 203.
Delanghe, B., Nakamura, F., Myoga, H., Magara, Y. and Guibal, E. (1994). Drinking water
denitrification in a membrane bioreactor, Water Science & Technology, Vol. 30, No. 6, p. 157.
Ejner, P. (2003). Personal communication to P.M. Sutton, President, Bioscan A/S, October.
Elmaleh, S. and Abdelmoumni, L. (1997). Cross-flow filtration of an anaerobic methanogenic
suspension, Journal of Membrane Science, Vol. 131, No. 1-2, p. 261.
Elmaleh, S. and Abdelmoumni, L. (1998). Experimental test to evaluate performance of an
anaerobic reactor provided with an external membrane unit, Water Science & Technology, Vol.
38, No. 8-9, p. 385.
Elmitwalli, T.A., Oahn, K.L.T., Zeeman, G. and Lettinga, G. (2002). Treatment of domestic
sewage in a two-step anaerobic filter/anaerobic hybrid system at low temperature, Water
Research, Vol. 36, No. 9, p. 2225.
Fakhru'l-Razi, A. (1993). Performance study of a new membrane anaerobic digester
configuration, Proceedings of the International Conference on Environmental Management,
Geo-Water and Engineering Aspects, p 115.
Fakhru'l-Razi, A. (1994). Ultrafiltration membrane separation for anaerobic wastewater
treatment, Water Science & Technology, Vol. 30, No. 12, p 321.
Fakhru'l-Razi, A. (1995). Steady states study in a membrane anaerobic system (MAS) for
wastewater treatment, Proceedings of the WEF 68th Annual Conference & Exposition, Vol. 3, p.
669.
Fakhru'l-Razi, A. and Noor, M.J.M.M. (1999). Treatment of palm oil mill effluent (POME) with
the Membrane Anaerobic System (MAS), Water Science & Technology, Vol. 39, No. 10-11, p.
159.
Fane A.G., Fell C.J.D. and Suzuki, A. (1983). The effect of pH and ionic environment on the
ultrafiltration of protein solutions with retentive membranes, Journal of Membrane Science, Vol.
16, p. 195.
Fawehinmi, F., Lens, P., Stephenson, T., Rogalla, F. and Jefferson B. (2004) The influence of
operating conditions on extracellular polymeric substances (EPS) production and biofouling in
anaerobic membrane bioreactors, Proceedings of Water Environment Membrane Technology
Conference (WEMT 2004), 7-10 June, 2004, Seoul, Korea, p. 469.
Ferras, E., Minier, M. and Goma, G. (1986). Acetonobutylic fermentation: improvement of
performances by coupling continuous fermentation and ultrafiltration, Biotechnology and
Bioengineering, Vol. 28, p. 523.
Flaschel, E., Raetz, E. and Renken, A. (1986). Shear deactivation of enzymes in membrane
reactors, Proceedings International Conference on Bioreactor Dynamics, Cambridge, UK.
Foresti, E. (2002). Anaerobic treatment of domestic sewage; established technologies and
perspectives, Water Science & Technology, Vol. 45, No. 10, p. 181.
Fuchs, W., Binder, H., Mavrias, G. and Braun, R. (2003). Anaerobic treatment of wastewater
with high organic content using a stirred tank reactor coupled with a membrane filtration unit,
Water Research, Vol. 37, No. 4, p. 902.
Fukuma, M., Takesada, K. and Yasunishi, A. (1993). Two-phase anaerobic treatment of
wastewater containing cellulose using membrane module in acidogenic phase, Kagaku Kogaku
Ronbunshu, Vol. 19, No. 5, p. 936 (in Chinese).

33

Ghyoot, W. R. and Verstraete, W. H. (1997). Coupling membrane filtration to anaerobic


primary sludge digestion, Environmental Technology (England), Vol. 18, No. 6, p. 569.
Giorno, L., Chojnacka, K., Donato, L. and Drioli, E. (2002). Study of a cell-recycle membrane
fermentor for the production of lactic acid by Lactobacillus bulgaricus, Industrial and
Engineering Chemistry Research, Vol. 41, No. 3, p. 433.
Grethlein, H.E. (1978). "Anaerobic digestion and membrane separation of domestic wastewater,"
Journal Water Pollution Control Federation, Vol. 50, p. 754.
Hall, E.R. (1992). Anaerobic treatment of wastewaters in suspended growth and fixed film
processes, Design of Anaerobic Processes for the Treatment of Industrial and Municipal Wastes,
edited by J.F. Malina Jr. and F.G. Pohland, Technomic Publishing Co. Ltd., Lancaster, PA, p. 41.
Hall, E.R., Onysko, K.A. and Parker, W.J. (1995). Enhancement of bleached Kraft
organochlorine removal by coupling membrane filtration and anaerobic treatment,
Environmental Technology, Vol. 16, p. 115.
Harada, H., Momonoi, K., Yamazaki, S. and Takizawa, S. (1994). Application of anaerobicUF
membrane reactor for treatment of a wastewater containing high strength particulate organics,
Water Science & Technology, Vol. 30, No. 12, p. 307.
He, Y., Li, C., Wu, Z. and Gu, G. (1999). Study on molecular weight cut-off of the anaerobic
ultrafiltration membrane bioreactor, Zhongguo Jishui Paishui/China Water & Wastewater, Vol.
15, No. 9, p. 10.
Hernandez, A.E., Belalcazar, L.C., Rodriguez, M.S. and Giraldo, E. (2002). Retention of
granular sludge at high hydraulic loading rates in an anaerobic membrane bioreactor with
immersed filtration, Water Science & Technology, Vol. 45, No. 10, p. 169.
Hickey, R.F. (2004). Personal communication to P.M. Sutton, Senior Vice President, Ecovation,
February.
Hogetsu, A., Ishikawa, T., Yoshikawa, M., Tanabe, T., Yudate, S. and Sawada, J. (1992). High
rate anaerobic digestion of wool scouring wastewater in a digester combined with membrane
filter, Water Science & Technology, Vol. 25, No. 7, p 341.
Imasaka, T. Kanekuni, N. So, H. and Yoshino, H. (1989). Cross-flow of methane fermentation
broth by ceramic membrane, Journal of Fermentation and Bioengineering, Vol. 68, p. 200.
Ince, O., Anderson, G.K. and Kasapgil, B. (1993). Improvement in performance of an anaerobic
contact digester using a crossflow ultrafiltration membrane techniques for brewery wastewater
treatment, Proceedings 48th Industrial Waste Conference, Purdue University, West Lafayette,
Indiana, Lewis Publishers, Chelsea, Michigan, Vol. 48, p. 551.
Ince, O., Anderson, G.K. and Kasapgil, B. (1995). Effect of changes in composition of
methanogenic species on performance of a membrane anaerobic reactor system treating brewery
wastewater, Environmental Technology, Vol. 16, No. 10, p 901.
Ince, O., Anderson, G.K. and Kasapgil, B. (1997). Composition of the microbial population in a
membrane anaerobic reactor system during start-up, Water Research, Vol. 31, No. 1, p 1.
Ince, B., Ince, O., Sallis, P.J. and Anderson, G.K. (1998). Experimental determination of the
inert soluble COD fraction of a brewery wastewater under anaerobic conditions, Environmental
Technology, Vol. 19, p. 437.
Ince, B. K., Ince, O., Sallis, P.J., and Anderson, G.K. (2000). Inert COD production in a
membrane anaerobic reactor treating brewery wastewater, Water Research, Vol. 34, No. 16, p.
3943.

34

Jewell, W.J., Switzenbaum, M.S. and Morris, J.W. (1981). Municipal wastewater treatment
with the anaerobic attached microbial film expanded bed process, Journal WPCF, Vol. 53, No.
4, p.482.
Jones, R.M. and Hall, E.R. (1986). Pilot scale study of high rate anaerobic treatment of wheat
starch wastewater from the Candiac, Quebec plant of Ogilvie Mills Ltd., Environment Canada,
Wastewater Technology Centre, June.
Juby, J.J.G., Leslie, G.L., Deshmukh, S.S., Torres, E.M. , Brown, J.P., Sethi, S. and Buhr, H.O.
(2000). A novel membrane-anaerobic digestion approach for wastewater treatment and water
reclamation, 2000 Water Reuse Conference Proceedings San Antonio, TX, January February.
Judd, S.J., Le-Clech, P., Taha, T. and Cui, Z.G. (2001). Theoretical and experimental
representation of a submerged MBR system, Membrane Technology, No. 135, p. 4.
Kang, I.J. (1996). Comparison of fouling characteristics between ceramic and polymeric
membranes in membrane coupled anaerobic bioreactors, MS Thesis, Department of Chemical
Technology, Seoul National University, Korea.
Kang, I-J., Yoon, S-H. and Lee, C-H. (2002). Comparison of the filtration characteristics of
organic and inorganic membranes in a membrane-coupled anaerobic bioreactor, Water
Research, Vol. 36, p. 1803.
Kang, I-J., Lee, C-H. and Kim, K-J. (2003). Characteristics of microfiltration membranes in a
membrane coupled sequencing batch reactor system, Water Research, Vol. 37, p. 1192.
Kataoka, N., Tokiwa, Y., Tanaka, Y., Fujiki, K., Taroda, H. and Takeda, K. (1992).
Examination of bacterial characteristics of anaerobic membrane bioreactors in three pilot-scale
plants for treating low-strength wastewater by application of the colony-forming-curve analysis
method, Applied and Environmental Microbiology, Vol. 58, No. 9, p. 2751.
Kawayake, E., Narukami, Y., and Yamagata, M. (1991). Anaerobic digestion by a ceramic
membrane enclosed reactor, Journal of Fermentation and Bioengineering, Vol. 71, No. 2, p 122.
Kim, J-O., and Somiya, I. (1999). Performance of organics decomposition and filterability
improvement of ceramic membrane by intermittent ozonation in an organic acid fermenter,
Journal of Japanese Sewage Works Association, Vol. 36, No. 3, p. 101.
Kim, J-O., Somiya, I., Shin, E-B., Bae, W., Kim, S-K. and Kim, R-H. (2002). Application of
membrane-coupled anaerobic volatile fatty acids fermentor for dissolved organics recovery from
coagulated raw sludge, Water Science & Technology, Vol. 45, No. 12, p. 167.
Kim, J.S., Lee, C.H. and Chang, I.S. (2001). Effect of pump shear on the performance of a
cross-flow membrane bioreactor, Water Research, Vol. 36, p. 2137.
Kim, J.S. and Lee, C.H. (2003). Effect of powdered activated carbon on the performance of an
aerobic membrane bioreactor: comparison between cross-flow and submerged membrane
systems, Water Environment Research, Vol. 75, No. 4, p. 300.
Kimura, K., Nakamura, M. and Watanabe, Y. (2001). Nitrate removal by a combination of
elemental sulfur-based denitrification and membrane filtration, Water Research, Vol. 36, p.
1758.
Kiriyama, K., Tanaka, Y. and Mori, I. (1992). Field test of a composite methane gas production
system incorporating a membrane module for municipal sewage, Water Science & Technology,
Vol. 25, p. 135.
Kirmura, S. (1991). Japans aqua renaissance 90 project, Water Science & Technology, Vol. 23,
No. 7-9, p.1573.

35

Kitamura, Y., Maekawa, T., Tagawa, A., Hayashi, H. and Farrell-Poe, K.L. (1996). Treatment
of strong organic, nitrogenous wastewater by an anaerobic contact process incorporating
ultrafiltration, Applied Engineering in Agriculture, Vol. 12, No. 6, p. 709.
Kobayashi, H.A., Conway De Macario, E., Williams, R.S, and Macario, A.J.L. (1988). Direct
characterization of methanogens in two high-rate anaerobic biological reactors, Applied and
Environmental Microbiology, Vol. 54, No. 3, p. 693.
Kraft, A. and Brockmann, M. (1999). Anaerobic technology and membrane bioreactor
technology for energy and water generation, Wasser Abwasser Praxis, Vol. 8, No. 3, p. 15 (in
German).
Lai, L.S., Fakhru'l-Razi, A., Idris, A. and Hassan, M. (1999). Performance and kinetic study of
membrane anaerobic system in treating POME, Artificial Cells, Blood Substitutes, and
Immobilization Biotechnology, Vol. 27, No. 5, p. 469.
Langenhoff, A.A.M., Intrachandra, N. and Stuckey, D.C. (2000). Treatment of dilute soluble
and colloidal wastewater in an anaerobic baffled reactor: influence of hydraulic retention time,
Water Research, Vol. 34, No. 4, p. 1307.
Langenhoff, A.A.M. and Stuckey, D.C. (2000). Treatment of dilute wastewater using an
anaerobic baffled reactor: effect of low temperature, Water Research, Vol. 34, No. 15, p. 3867.
Lawrence, D., Van Bentem, A., Van Der Herberg and Roeleveld, P. (2001). MBR pilot research
at the Beverwijk WWTP, a step forward to full scale implementation, STOWA, Foundation for
Applied Water Research 2001, H2O, October, p. 11.
Lee, S., Burt, A., Rusoti, G. and Buckland, B. (1999). Microfiltration of yeast cells using a
rotary disk dynamic filtration system, Biotechnology and Bioengineering, Vol. 48, p. 386.
Lee, J., Ahn, W-Y. and Lee, C-H. (2001a). Comparison of the filtration characteristics between
attached and suspended growth microorganisms in a submerged membrane bioreactor, Water
Research, Vol. 35, p. 2435.
Lee, J.C., Kim, J.S., Kang, I.J., Cho, M.H., Park, P.K. and Lee, C.H. (2001b). Potential and
limitations of alum or zeolite addition to improve the performance of a submerged membrane
bioreactor, Water Science & Technology, Vol. 43, No. 11 p. 59.
Lee, S-M., Jung, J-Y. and Chung, Y-C. (2001c). Novel method for enhancing permeate flux of
submerged membrane system in two-phase anaerobic reactor, Water Research, Vol. 35, No. 2,
p. 471.
Lei, E. and Berube, P.R. (2004). Impact of membrane configuration and hydrodynamic
conditions on the permeate flux in submerged membrane systems for drinking water treatment,
Proceedings British Columbia Water and Waste Association Conference, Whistler, BC, Canada.
Lettinga, G., De Man, A., Van Der Last, A.R.M., Wiegrant, W., Van Knippenberg, K., Frijors, J.
and Van Buren, J.C.L. (1993). Anaerobic treatment of domestic sewage and wastewater, Water
Science & Technology, Vol. 27, No.9, p. 67.
Lew, B., Belavski, M., Admon, S., Tarre, S. and Green, M. (2003). Temperature effect on
UASB reactor operation for domestic wastewater treatment in temperate climate regions, Water
Science & Technology, Vol. 48, No. 3, p. 25.
Li, A. and Corrado, J.J. (1986). Scale-up of the membrane anaerobic reactor system,
Proceedings 40th Industrial Waste Conference, Purdue University, West Lafayette, Indiana, Ann
Arbor Science, Ann Arbor, Michigan, p. 805.
Lubbecke, A., Vogelpohl, A. and Dewjanin, W. (1995). Wastewater treatment in a biological
high-performance system with high biomass concentration, WaterResearch, Vol. 29, p. 793.

36

Madakoro, T., Ueno, M., Moro, M., Yamamoto, T. and Shibata, T. (1999). Anaerobic digestion
system with micro-filtration membrane for kitchen refuse, Presented at 2nd International
Symposium on Anaerobic Digestion of Solid Wastes, Barcelona, Spain, June.
Magara, Y., Nishimura, K., Itoh, M. and Tanaka, M. (1992). Biological denitrification system
with membrane separation for collective human excreta treatment plant, Water Science &
Technology, Vol. 25, No. 10, p. 241.
Mallevialle, J., Odendaal, P.E. and Wiesner, M.R. (editors) (1996). Water treatment membrane
processes, McGraw-Hill Publishing, New York, NY, USA.
Mamais, D., Paul, A.P., Cheng, Y.W., Loiacono, J. and Jenkins, D. (1994). Determination of
ferric chloride dose to control struvite precipitation in anaerobic sludge digesters, Water
Environment Research, Vol. 66, No. 7, p. 912.
Manning, J., Sutton, P.M., Gaines, F.R. and Dunn, W.G. (2001). Combined and separate
municipal and leachate wastewater treatment: full scale application of the membrane bioreactor
system, Proceedings of the WEF 74th Annual Conference & Exposition, Atlanta, Georgia, CDROM, October.
Minami, K., Okamura, K., Ogawa, S. and Naritomi, T. (1991). Continuous anaerobic treatment
of wastewater from a kraft pulp mill, Journal of Fermentation and Bioengineering, Vol. 71, No.
4, p. 270.
Mizuno, O., Takagi, H. and Noike, T. (1998). Biological sulphate removal in an acidogenic
bioreactor with an ultrafiltration membrane system, Water Science & Technology, Vol. 38, No.
4-5, p. 513.
Morgenroth, E., Raskin, L., Norddahl, B., Criddle, C. and Hickey, B. (2003). Membrane
bioreactors for biological wastewater treatment and water reuse advanced treatment of swine
manure and how it relates to CAMPWS, CAMPWS Seminar Presentation, University of Illinois
at Urbana-Champaign, November.
Munariatis, I.D. and Grigoropoulos, S.G. (2002). Low-strength wastewater treatment using an
anaerobic baffled reactor, Water Environment Research, Vol. 74, No., 2, p.170.
Nagano, A., Arikawa, E. and Kobayashi, H. (1992). The treatment of liquor wastewater
containing high-strength suspended solids by membrane bioreactor system, Water Science &
Technology, Vol. 26, No. 3-4, P. 887.
Nagaoka, H. (1999). Nitrogen removal by submerged membrane separation activated sludge
process, Water Science & Technology, Vol. 39, No. 8, p. 107.
Ng, W.J., Ong, S.L., Gomez, M.J., Hu, J.Y. and Fan, X.J. (2000). Study on a sequencing batch
membrane bioreactor for wastewater treatment, Water Science & Technology, Vol. 41, No. 1011, p. 227.
Nikaido, S., Nishimura, K. and Noike, T. (1996). Study on treatment of long-chain fatty acid
containing wastewater in the anaerobic digestion with membrane filtration systems, J. Japan
Society on Water Environment, Vol. 19, No. 3, p. 220 (in Japanese).
Norddahl, B. and Rohold, L. (1998). Biorek principle - a membrane bioreactor coupled to an
ammonia stripper and an RO unit enabling conversion of manure and slurry into energy,
concentrated fertiliser and potable water, Proceedings of Bioenergy 98, Madison, Wisconsin,
October.
Nuhoglu, A., Pekdemir, T., Yildiz, E., Keskinler, B. and Akay, G. (2002). Drinking water
denitrification by a membrane bio-reactor, WaterResearch, Vol. 36, p. 1155.
Ognier, S., Wisniewski, C. and Grasmick, A. (2001). Biofouling in membrane bioreactors:
phenomenon analysis and modelling, Presented at Membrane Bioreactors International
Conference, May.

37

Okamura, K. (1994). Study of wastewater treatment using anaerobic bioreactor combined UF


membrane, Papers Ship Research Institute Tokoyo, Vol. 31, No. 5, p. 2-50.
Omstead, D.R., Jeffries, T.W., Naughton, R. and Gregor, H. P. (1980). Membrane controlled
digestion: anaerobic production of methane and organic acids, Biotechnology & Bioengineering
Symposium, No. 10, John Wiley & Sons, New York, New York, p. 247.
Onysko, K.A., and Hall, E.R. (1993). Anaerobic membrane bioreactor treatment of segregated
bleach plant wastewater. 1993 Environmental Conference, Tappi, p. 845.
OSullivan, T.J., Epstein, A.C., Korchin, S.R. and Beaton, N.C. (1984). Applications of
ultrafiltration in biotechnology, Chemical Engineering Progress, January, p. 68.
Ozaki, N. and Yamamoto, K. (2001). Hydraulic effects on sludge accumulation on membrane
surface in crossflow filtration, Water Research, Vol. 35, No. 13, p. 3137.
Pagilla, K.R. (2004). Personal communication to P.M. Sutton, Associate Professor, Illinois
Institute of Technology, February.
Park, H., Choo, K. and Lee, C. (1999). Flux enhancement with powdered activated carbon
addition in the membrane anaerobic bioreactor, Separation Science and Technology, Vol. 34,
No. 14, p. 2781.
Pillay, V.L., Townsend, B. and Buckley, C.A. (1994). Improving the performance of anaerobic
digesters at waste water treatment works: the coupled cross-flow microfiltration/digester
process, Water Science & Technology, Vol. 30, No. 12, p. 329.
Pirbazari, M., Ravindran, V., Badriyha, B.N. and Kim, S.H. (1996). Hybrid membrane filtration
process for leachate treatment, Water Research, Vol. 30, No. 11, p. 2691.
Rebac, S. (1998). Psychrophilic anaerobic treatment of low strength wastewaters, Wageningen
Agricultural University, Wageningen, The Netherlands.
Ross, B. and Strohwald, H. (1994). Membranes add edge to old technology, Water Quality
International, No. 4, , p 18.
Ross, W.R., Barnard, J.P., Le Roux, J. and Villiers, H. A. (1990). Application of ultrafiltration
membranes for solids-liquid separation in anaerobic digestion systems: The ADUF process,
Water Science & Technology, Vol. 16, p. 85.
Ross, W.R., Barnard, J.P., Strohwald, N.K.H., Grobler, C.J. and Sanetra, J. (1992). Practical
application of the ADUF process to the full-scale treatment of a maize-processing effluent,
Water Science & Technology, Vol. 25, No. 10, p 27.
Ross W.R., Strohwald, N. K. H., Grobler, C.J. and Senetra, J. (1994). Membrane-assisted
anaerobic treatment of industrial effluents: The South African ADUF process, Proceedings of
the 7th International Symposium on Anaerobic Digestion, Cape Town, South Africa, p. 550.
Sainbayar, A., Kim, J. S., Jung, W. J., Lee, Y. S. and Lee, C. H. (2001). Application of surface
modified polypropylene membranes to an anaerobic membrane bioreactor, Environmental
Technology, Vol. 22, No. 9, p. 1035.
Saw, C.B., Anderson, G.K., James, A. and Le, M.S. (1986). A membrane technique for biomass
retention in anaerobic waste treatment, Proceedings 40th Industrial Waste Conference, Purdue
University, West Lafayette, Indiana, Ann Arbor Science, Ann Arbor, Michigan, p. 805.
Schiener, P., Nachaiyasit, S., and Stuckey, D.C. (1998). Production of soluble microbial
products (SMP) in an anaerobic baffled reactor: composition biodegradability and the effect of
process parameters, Environmental Technology, Vol. 19, No. 4, p. 391.
Seghezzo, L., Guerra, R.G., Gonzalez, S.M., Trupiano, A.P., Figueroa, M.E., Cuevas, C.M.,
Zeeman, G. and Lettinga, G. (2002). Removal efficiency and methanogenic activity profiles in a

38

pilot-scale UASB reactor treating settled sewage at moderate temperatures, Water Science &
Technology, Vol. 45, No. 10, p. 243.
Shimizu, U., Rokudai, M., Tohya, S., Kauawake, E., Yazawa, T., Tanaka, H. and Eguchi, K.
(1989). Effect of pore size on the filtration characteristics of ceramic membrane for membrane
bioreactors, Kagaku Kogaku Ronbunshu, Vol. 15, p. 322.
Shimizu, Y. et al. (1991). Influence of organic acid density in methane fermentation liquid on
membrane filtration characteristics, Japanese Journal of Fermentation Technology, Vol. 67, No.
1, p. 15.
Shimizu, Y., Matsushita, K. and Watanabe, A. (1994). Influence of shear breakage of microbial
cells on cross-flow microfiltration flux, Journal of Fermentation and Bioengineering, Vol. 78,
No. 2, p. 170.
Shimizu, Y., Uryu, K., Okuno, Y. and Watanabe, A. (1996). Cross-flow microfiltration of
activated sludge using submerged membrane with air bubbling, Journal of Fermentation and
Bioengineering, Vol. 81, No. 1, p. 55.
Shimizu, Y., Okuno, Y., Uryu, K., Ohtsubo, S. and Watanabe, A. (1996). Filtration
characteristics of hollow fiber microfiltration membranes used in membrane bioreactor for
domestic wastewater treatment, Water Research, Vol. 30, No. 10, p. 2385.
Shin, H. and Kang, S. (2002). Performance and membrane fouling in a pilot scale SBR process
coupled with membrane, Water Science & Technology, Vol. 47, No. 1, p. 139.
Singh, K.S. and Viraraghavan, T. (2003). Impact of temperature on performance,
microbiological, and hydrodynamic aspects of UASB reactors treating municipal wastewater,
Water Science & Technology, Vol. 48, No. 6, p. 211.
Smith, C.V., Di Gregorio, D.O. and Talcott, R.M. (1969). The use of ultrafiltration membranes
for activated sludge separation, Proceedings 24th Industrial Waste Conference, Purdue
University, West Lafayette, Indiana, p. 805.
Stensel, H.D. and Strand, S.E. (2004). Evaluation of feasibility of methods to minimize biomass
production from biotreatment, Water Environment Research Foundation, Project Report 00CTS-10 (in publication).
Strohwald, N.K.H. and Ross, W.R. (1992). Application of the ADUF process to brewery
effluent on a laboratory scale, Water Science & Technology, Vol. 25, No. 10, p 95.
Stuckey, D.C. (2003). The submerged anaerobic membrane bioreactor (SAMBR): an
intensification of anaerobic wastewater treatment, Abstract of Presentation to the Dept. Civil
Engineering at the University of Minnesota, Minneapolis, Minnesota, September.
Stuckey, D.C. and Hu, A. (2003). The submerged anaerobic membrane bioreactor (SAMBR):
an intensification of anaerobic wastewater treatment, Presented at the International Water
Association Leading Edge Conference on Drinking Water and Wastewater Treatment
Technologies, Noordwijk, The Netherlands, May.
Sutton, P.M., Li, A., Evans R.R. and Korchin S.R. (1983). Dorr-Oliver's fixed film and
suspended growth anaerobic systems for industrial wastewater treatment and energy recovery,
Proceedings 37th Industrial Waste Conference, Purdue University, West Lafayette, Indiana, Ann
Arbor Science, Ann Arbor, Michigan, p. 667.
Sutton, P.M. and Evans, R.R. (1983). Anaerobic system designs for efficient treatment of
industrial wastewater, Proceedings of the 3rd Symposium on Anaerobic Digestion, IAWQ.
Sutton, P.M. (1986a). Active biomass retention: the key to anaerobic process efficiency,
Presented at Anaerobic Fixed-Film Digestion Seminar, Ontario Ministry of Health Laboratories,
Toronto, Ontario, November.

39

Sutton, P.M. (1986b). Innovative biological systems for anaerobic treatment of grain and food
processing wastewater, Starch/Starke, Vol. 38, No. 9, p.314.
Sutton, P.M., Mishra, P.N., Bratby, J.R. and Enegess, D. (2002). Membrane bioreactor
industrial and municipal wastewater applications: long term operating experience, Presented at
the WEF 75th Annual Conference & Exposition, Chicago, Illinois, September/October.
Takashima, M., Kudoh, Y. and Tabata, N. (1996). Complete anaerobic digestion of activated
sludge by combining membrane separation and alkaline heat post-treatment, Water Science &
Technology, Vol. 34, No. 5-6 pg. 3.
Tardieu, E., Grasmick, A., Geaugey, V. and Manem, J. (1999). Influence of hydrodynamics on
fouling velocity in a recirculated MBR for wastewater treatment, Journal of Membrane Science,
Vol. 156, No. 10, p. 131.
Torres, P. and Foresti, E. (2001). Domestic sewage treatment in a pilot system composed of
UASB and SBR reactors, Water Science & Technology, Vol. 44, No. 4, p. 247.
Uemura, S. and Harada, H. (2000). Treatment of sewage by a UASB reactor under moderate to
low temperature conditions, Bioresource Technology, Vol. 72, No. 3, p. 275.
Uemura, T. and Kondou, Y. (2003). Membrane technology progress in Japan, Water 21,
August.
Van Bentem, A., Lawrence, D., Horjus, F., De Boer, R., Koornneef, E. and Van Efferen, B.
(2001). MBR pilot research in Beverwijk: side studies, STOWA, Foundation for Applied
Water Research 2001, H2O, October, p. 16.
Van Der Last, A.R.M. and Lettinga, G. (1992). Anaerobic treatment of domestic wastewater
under moderate climatic (Dutch) conditions using upflow reactors at increased superficial
velocities, Water Science & Technology, Vol. 25, No. 7, p. 167.
Van Der Roest, H. and Leenen, J. (2001). The MBR as worldwide promising technology,
STOWA, Foundation for Applied Water Research 2001, H2O, October, p. 5.
Van Houten, R., Evenblij, H. and Keijmel, M. (2001). Membrane bioreactors hit the big time
ten years of research in the Netherlands, STOWA, Foundation for Applied Water Research
2001, H2O, October, p. 26.
Vera, L., Delgado, S. and Elmaleh, S. (2000). Gas sparged cross-flow microfiltration of
biologically treated wastewater, Water Science & Technology, Vol. 41, No. 10-11, p. 173.
Weigant, W.M. (2001). Experiences and potential of anaerobic wastewater treatment in tropical
regions, Water Science & Technology, Vol. 44, No. 8, p. 107.
Wen, W., Huang, X. and Qian, Y. (1999). Domestic wastewater treatment using an anaerobic
bioreactor coupled with membrane filtration, Process Biochemistry, Vol. 35, No. 3-4, p. 335.
Wisniewski, C. and Grasmick, A. (1998). Floc size distribution in a membrane bioreactor and
consequences for membrane fouling, Colloids and Surfaces, A: Physicochemical and
Engineering Aspects, Vol. 138, p. 403.
Yamagishi, S. et al. (1997). Study on the factors influencing permeate flux of membrane
coupled with anaerobic membrane reactor, Japanese Industrial Water, Vol. 468, p. 2.
Yamagishi, T., Matsuda, M., Masuda, H., Kamisawa, C., Iizumi, S., Oomura, T., Nagura, K., Ito,
S., Taked, A.H., Suzuki, T., Toyohara, H. and Okada, T. (1990). Cross-flow filtration for
anaerobic wastewaters, Polution Japan, Vol. 5, p. 291.
Yamaguchi, M., Okamura, K., Tanimoto, Y., Naritomi, T., Ogawa, S., Hake, J. and Minami, K.
(1990). Thermophilic methane fermentation of evaporator condensate from a kraft pulp mill,
TAPPI Environ Conf, Seattle, WA Vol. 2, p. 631.

40

Yamamoto, K., Hiasa, M., Mahmood, T. and Matsuo, T. (1989). Direct solid-liquid separation
using hollow fiber membrane in an activated sludge aeration tank, Water Science & Technology,
Vol. 21, p. 43.
Yiliang, H., Zhichao, W., Chunjie, L., Guowie, G., Fenging, X., Mingxu, Z. and Zuyi, C. (1999).
Application of the anaerobic MBR for treatment of high concentration food wastewater,
Chinese Journal of Environmental Science, November (in Chinese).
Yoon, S.H., Kang, I.J. and Lee, C.H. (1999). Fouling of inorganic membrane and flux
enhancement in membrane-coupled anaerobic bioreactor. Separation Science and Technology,
Vol. 34, No. 5, p.709.
Yuntao, G., Zhanpeng, J., Wanpeng, Z. and Zhongying, C. (1998). Two-phase anaerobic
membrane biosystem for organic wastewater treatment, Chinese Journal of Environmental
Science, Vol. 21, No. 4, p. 52 (in Chinese).
Yuntao, G., Zhanpeng, J., Wanpeng, Z. and Peng, J. (2000). Two-phase anaerobic membrane
biosystem for treatment of papermill wastewater, Chinese Journal of Environmental Science,
Vol. 19, No. 6, p. 56 (in Chinese).
Yushina, Y. and Hasegawa, J. (1994). Process performance comparison of membrane
introduced anaerobic digestion using food industry waste water, Desalination, Vol. 98, No. 1-3,
p 413.
Zeeman, G. and Lettinga, G. (2002). The role of anaerobic digestion of domestic sewage in
closing the water and nutrient cycle at community level, Water Science & Technology, Vol. 39,
No. 5, p. 187.
Zoh, K-D. and Stenstrom, M.K. (2001). Application of a membrane bioreactor for treating
explosives process wastewater, Water Research, Vol. 36, p. 1018.

41

Figure 1. Simplified Schematic of an Internal Submerged Membrane Anaerobic Hybrid MBR Configuration.

PROCESS GAS

GAS
HANDLING
PACKAGE

GAS FOR
CUSTOMER
USE OR TO
FLARE

EFFLUENT
WASTEWATER

ANAEROBIC
HYBRID
REACTOR

SOLIDS RECYCLE

MEMBRANES

AIR

EXCESS
SOLIDS

42

Table 1. Anaerobic MBR System Design and Performance Information.

Wastewater

Kang et al. (2000); 1


Alcohol fermentation

Reference, Entry No. and Reported Information


Kang et al. (2000); 2
Park et al. (1999); 3
Alcohol fermentation
Synthetic

Choo and Lee (1996a); 4


Alcohol (distillery)

System

Lab scale

Lab scale

Lab scale

Objectives

Comparison of inorganic and


organic membranes
5

Comparison of inorganic and


organic membranes
5

Mechanical

Mechanical

Mechanical

42,000

42,000

22,600

Parameter

Bioreactor volume

Units

liters (L)

Mixing method
Feed COD

mg/L

COD removal

OLR

kg COD/m3day

3 to 3.5

3 to 3.5

MLSS

mg/L

2000 to 3000

2000 to 3000

Temperature

55

55

pH

Study effect of PAC on membrane


fouling
4.5

Study membrane fouling


4

Exceeded 97

1.5 to 4.5

1.5 to 2.1
c

55

53 to 55

7.5 to 7.9

HRT

days

13

13

15

SRT

days

Gas production

4.2 L total gas/day

4.2 L total gas/day

2.8 L total gas/day

Membrane MOC

Tubular hydrophobic PP

Polymeric flat plate

Manufacturer

Microdyne, Germany

Tubular zirconia skin and carbon


support
Tech-Sep, France

DDS, Denmark

0.2

0.14

0.1

External

External

External

External

0.0113

0.0129

0.017

0.034

Pore size

micrometers

MW exclusion

daltons

Membrane location
Surface area

m2
L/m day

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP

m/s

Inlet pressure

kg/cm2

Outlet pressure

Other information
a

20,000

Operating permeate flux

DDS, Denmark

0 to 840; c
3

0.6

0.6

kg/cm2

kg/cm

Used membrane cleaned with hypochlorite and rinsed with ultra-pure water prior to the next run. Digestor broth to UF via a positive displacement pump. TMP regulated using a back pressure valve. Back flushing
arrangement provided, using HCL, 90 sec every 2 hours.
Short term study (10 days) with PAC in reactor at concentration of 5 g/L. Batch microfiltration studies performed using digester contents at different PAC dosages. PAC addition improved flux and treatment
performance.
MLVSS decreased from 3000 to less than 500 mg/L observation attributed to cell lysis due to mechanical shear from membrane process pump. Observed large amount of biomass attached to membrane surface. Severe
membrane fouling due to attached biomass and precipitation of inorganics.

43

Table 1. Anaerobic MBR System Design and Performance Information (Contd).

Wastewater

Choo and Lee (1998); 5


Alcohol (distillery)

Reference, Entry No. and Reported Information


Yoon et al. (1999); 6
Sainbayar et al. (2001); 7
Alcohol fermentation
Synthetic

Hall et al. (1995); 8


Kraft bleach plant

System

Lab scale

Bench scale

Lab Scale

Pilot scale

Objectives

Study hydrodynamics of biosolids


during cross-flow filtration
4

Study membrane fouling

4.5

15 (MBR); 5 (UASB)

Mechanical

Mechanical

Mechanical

35,000

42,600

Parameter

Bioreactor volume

Units

liters (L)

Mixing method
Feed COD

mg/L

COD removal

OLR

kg COD/m3day

1.5

2 to 7

MLSS

mg/L

1000 to 3200

2800 to 5000

Temperature

90 to 95

pH
HRT

days

SRT

days

53 to 55

55

7.8 to 8.0

7.3 to 7.7

c
4.0
7600 to 15,700
55

35

1.0
a

6.0

Tubular zirconia skin and carbon


support
Tech-Sep, France

Polymeric flat plate

Polymeric tubular

Akzo Nobel, Germany

Zenon, Canada

0.45

0.2

External

External

External

0.0168

0.011

0.006

0.5 to 1.25

0.5 and 1.2

0.5 to 3.0

0.6

0.5 and 1.0

Gas production
Membrane MOC

Fluoropolymer sheets, flat plate

Manufacturer

DDS, Denmark

Pore size

micrometers

MW exclusion

daltons

Membrane location
Surface area

m2
2

Operating permeate flux

L/m day

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP

m/s

Inlet pressure
Outlet pressure
Other information
a
b

20,000

kg/cm2

10,000

kg/cm

External

2.15

1.7

kg/cm

pH not controlled . No solids wasted. Permeate flux typically 1200 to 4800 L/m2 d dictated by time since membrane cleaned and membrane cleaning procedure. Fouling due to inorganic precipitates.
To compare filtration characteristics of PP (strongly hydrophobic) and modified PP (hydrophilic) membranes. Centrifugal pump used to feed UF. Flux typically 2 to 3 L/m2.day with modified PP providing 13.5 % flux
improvement.
Objective was to compare performance of UASB, anaerobic MBR and membrane ultrafiltration systems for AOX removal. Bioreactor volume includes process piping volume. Anaerobic MBR achieved highest
performance.

44

Table 1. Anaerobic MBR System Design and Performance Information (Contd).

Wastewater

Onysko and Hall (1993); 9


Dechlorinated kraft mill effluent

Reference, Entry No. and Reported Information


Fakrul-Razi (1993); 10
Fakrul-Razi (1994); 11
High strength industrial
Brewery

Fakrul-Razi and Noor (1999);12


Palm oil mill effluent

System

Lab scale

Lab scale

Lab scale

Lab scale

Parameter

Units

Objectives

Assess performance

Assess performance

Assess performance

liters (L)

15

120

120

50

mg/L

AOX 60 to 255

16,680 to 43,040

46,200 to 84,010

39,910 to 68,310

COD removal

AOX 50 to 60

OLR

kg COD/m3day

MLSS

mg/L

Temperature

Bioreactor volume
Mixing method
Feed COD

pH
HRT

days

SRT

days

Greater than 98

92 to 94

4.5 to 9.5

12.1 to 19.7

14.2 to 21.7

7600 to 15,700

19,900 to 25,750

31,500 to 38,3000

50,760 to 56,600

35

35

Approximately 35

35

Buffered wastewater pH 6.8 to 7.2

6.5 to 7.2

6.8 to 7.4

6.3 to 7.8

0.94

3.65 to 4.30

3.64 to 3.98

2.82 to 3.15

94 to 167

58.8 to 83.3

76.9 to 161.3

0.28 to 0.31 L CH4/g COD


removed at STP

0.26 to 0.29 L CH4/g COD


removed at STP

0.24 to 0.28 L CH4/g COD


removed at STP

Gas production
Membrane MOC

Polymeric tubular

Manufacturer

Zenon, Canada

Pore size

micrometers

MW exclusion

daltons

Membrane location
Surface area

m2

Operating permeate flux

L/m2day

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP

m/s

Inlet pressure
Outlet pressure
Other information
a
b
c

10,000

10,000

10,000

200,000

External

External

External

External

Average 1.16
2.15

kg/cm2
2

kg/cm

2.3

1 to 2

1.5

1.7

kg/cm

Objective was to assess MBR for removal of AOX. Positive displacement pump used to feed UF. Membrane cleaned off-line.
COD removal 96 to 99 percent, higher at lower OLRs. Positive displacement pump used to feed UF. Membrane flow and pressure controlled using valve on concentrate return line.
Positive displacement pump used to feed UF. Membrane flow and pressure controlled using valve on concentrate return line.

45

Table 1. Anaerobic MBR System Design and Performance Information (Contd).

Wastewater

Lai et al. (1999); 13


Palm oil mill effluent

Reference, Entry No. and Reported Information


Fakhrul-Razi (1995); 14
Yuntao et al. (1998); 15
High strength industrial
Synthetic organic

Yuntao et al. (2000); 16


Paper mill black liquor

System

Lab scale

Lab scale

Lab scale two-phase system

Lab scale two-phase system

Objectives

Derive performance and kinetic


information
50

Assess performance

Compare to conventional
anaerobic system
c

Compare to conventional
anaerobic system
c
Greater than 1500

Parameter

Bioreactor volume
Mixing method

Units

liters (L)

120

Feed COD

mg/L

58,000 to 85,000

84,010

1500 to 7000

COD removal

83 to 98; c

Greater than 95

Greater than 70

OLR
MLSS

kg COD/m3day
mg/L

1.5 to 6.5
12,681 to 30,460

1 to 19.7
38,000

5 to 10

3 to 6

Temperature

Approximately 35

Approximately 35

35

35

Approximately 7

6.9 to 7.3

pH
HRT

Days

SRT
Gas production

Days

15 to 100
0.347 to 0.745 L total gas/g COD
removed
Polysulphone

Pore size

micrometers

0.1

MW exclusion

daltons

200,000

10,000

External

External

Surface area

m2

0.024

Operating permeate flux

L/m2day

0.18 to 1.6; c

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP
Inlet pressure

m/s

Outlet pressure

kg/cm2

Membrane MOC

3.23 to 4.30

Total 0.5

58 to 555
0.26 to 0.34 L Ch4/g COD
removed at STP

Manufacturer

Membrane location

kg/cm2
kg/cm2

0.1 to 0.4

0.1

External

External

1 to 1.5

Other information
a
b
c
c
a
Single-phase pump used to feed UF. COD removal decreased as OLR increased. Operating permeate flux restored to 1.25 L/m2day by chemical cleaning after 30 days of operation. Data fit kinetic models over OLR
studied.
b
COD removal 96 to 99 percent, higher at lower OLRs. Positive displacement pump used to feed UF. Membrane flow and pressure controlled using valve on concentrate return line.
c
Acid phase and methane phase reactors working volume respectively, 1.65 and 3.6 L. Membrane located between acid and methane phase.

46

Table 1. Anaerobic MBR System Design and Performance Information (Contd)


Parameter

Units

Choate et al. (1983); 17


Wheat Flour

Wastewater

Reference, Entry No. and Reported Information


Butcher (1989); 18
c; 19
Wheat Flour
d

c; 20
e

System

Full scale; a

Full scale; b

Lab and pilot scale; d

Full scale; e

Objectives

Improve performance

Improve performance

Develop ADUF process

377,000 (each compartment)

1130 to 2000; b

50 to 3000

Improve performance of anaerobic


systems
80,000 to 2,610,000
Concentrate return

17,000

20,000

Mechanical and/or by concentrate


return
3500 to 37,000

82 to 91

65 to 78; b

77 to 97

95 and 97

2.0 to 2.4

5.0 to 15.0

3.0 and 6.0

10,000

10,000 to 50,000

10,000 to 30,000 and 23,000

35

30 and 35

0.8 to 3.3

1.3 and 5.2

Bioreactor volume

liters (L)

Mixing method

Gas recirculation

Feed COD

mg/L

COD removal

%
3

OLR

kg COD/m day

MLSS

mg/L

Temperature

36

pH

8000 and 15,000

6.8

HRT

days

4.5

SRT

days

Gas production

100 m3 total/day

Membrane MOC

Tubular polymeric

Tubular polymeric

Tubular polymeric

Manufacturer

Abcor, USA

Bintech or Membratek, South


Africa

Bintech or Membratek, South


Africa

20,000 to 80,000

20,000 to 80,000

External

External

Pore size

micrometers

MW exclusion

daltons

Membrane location

18,000
External

Surface area

144

Operating permeate flux

L/m2day

Design 1.04; a

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP

m/s

Inlet pressure

kg/cm2

Outlet pressure
Other information
a

c
d
e

1000 to 2150 m3 total/day; b

144 to 288; b

0.44 to 9.6
0.4 to 3.5

200 and 688

1.5 to 2.0

1.8 and 1.6

3.4 to 5.0

5.0 to 4.5

kg/cm2
6.9

kg/cm

Retrofitted 1 compartment of 3 completely-mixed digester (with internal gravity separation) with a coupled UF membrane system. Subsequently mixed-liquor from each compartment pumped to membrane systsem
alternatively for 8 hour periods. Oil in feed decreased membrane flux. Membranes cleaned on average 1 time per month with caustic soda and hypochlorite.
Updated information from plant discussed by Choate et al. (1983). Information based on operation from April 1980 to July 1986. Expanded from 1130 to 2000 m3 in 1985 and UF membrane system doubled in size.
Approximately 30 % of effluent is directly from gravity separation zone in digester compartments. Membrane life almost double expectation of 3 years.
Pilot and full scale results summarized from experience with ADUF process in South Africa from Ross et al. (1990, 1992, 1994), Strohwald and Ross (1992), and Ross and Strohwald (1994).
Brewery, wine distillery and malting. Membrane system coupled to complete mix reactor or clarigester (i.e., suspended growth reactor with internal clarification).
Egg processing and maize processing. Membrane system coupled to clarigester. All data reported for egg and maize processing, respectively.

47

Table 1. Anaerobic MBR System Design and Performance Information (Contd)


Parameter

Units

Wastewater
System
Objectives
Bioreactor volume

liters (L)

Mixing method

Reference, Entry No. and Reported Information


Ince et al. (1993); 22
Anderson et al. (1996); 23
Brewery
Brewery; c
Pilot scale
Pilot scale
Assess performance
Derive performance and kinetic
information
120
120

Mechanical

Mechanical

Mechanical

Concentrate return

80,000 to 90,000

80,000 to greater than 90,000

10,900 to 16,000

97 to 99
Up to 28.5

99; c
Above 30

Approximately 96
6 to 12; d

Anderson et al. (1986); 21


Synthetic
Laboratory scale; a
Assess performance

Jones and Hall (1986); 24


Wheat starch
Pilot scale
Derive process information
303

Feed COD

mg/L

COD removal
OLR

%
kg COD/m3day

3500 to 104,000 to complete


system
98 to 99 for complete system; a
Methanogenic 1 to 12.2

MLSS

mg/L

Methanogenic 4000 to 17,500

Up to 50,000

Up to 51,000

6000 to 30,000

Temperature
pH

35 to 37
6.9 to 7.2

35 to 37
6.9 to 7.2

Approximately 35
6.8 to 7.2

HRT

days

35 for complete system


Acidogenic 3 to 4,
methanogenic 7.0 to 7.5
Acidogenic 0.7 to 1.9,
methanogenic 2.9 to 7.7

2.5 to 4.2

2.5 to 4.2

SRT
Gas production

days
Methanogenic 0.25 to 0.30 L
CH4/g COD removed at STP
Polyethylene

O.28 to 0.38 m3 CH4/kg COD


removed
Fluoropolymer tubular
Patterson Candy, England

O.28 m3 CH4/kg COD removed; c

Membrane MOC
Manufacturer
Pore size
MW exclusion

micrometers
daltons

Membrane location

Fluoropolymer tubular
Patterson Candy, England

O.29 m3 total gas/kg COD


removed; d
Polyethersulfone flat plate
Dorr Oliver, USA

90
a

200,000

200,000

Approximately 10,000

External

External

External

0.048

0.048

0.47 to 0.87; d

Surface area

m2

Operating permeate flux

L/m2day

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP

m/s

2.4 to 3.2

kg/cm2

2.1 to 2.4

Inlet pressure

kg/cm2

Outlet pressure

kg/cm2

0.5 to 1.6; d

0.7 to 1.1
Maximum 4.2

Other information
a
b
c
d
a
Suspended growth acidogenic reactor plus suspended growth methanogenic reactor with internal membrane. Acidogenic and methanogenic reactor volume respectively, 2.5 and 10 l. Acidogenic reactor COD removal
10 to 23 %. A membrane lined methanogenic reactor with 5 mm gap between membrane reactor wall. In-situ scraping of membrane controlled biofilm build-up. Membrane thickness either 5 or 25 mm (i.e.,
information not clear). Studied flux with different biomass. Acidogenic and methanogenic biomass flux respectively, 0.03 to 0.08 and 0.01 to 0.03 L/m2day, decreasing with increasing TSS concentration. Effluent
TSS after membrane step less than 100 mg/L.
b
Variable speed membrane feed pump. Membrane cross-flow velocity and TMP controlled by flow and pressure regulators. Flux dependent on cross-flow velocity, TMP and biomass concentration, varied between 1.0
and 4.8 L/m2day. Membranes cleaned with NaOH and hypochloride solutions at unstated frequency.
c
Glucose added to feed after OLR of 20 kg/m3day achieved. COD removal no lower than 99 %, under steady state conditions. Calculated SRTs for each of 4 steady state conditions ranged from 58 to 480 days.
Operating period was 15 months. Variable speed membrane feed pump.
d
Feed rate and thus OLR limited by membrane permeate rate. Methane gas content 77 %. Membrane modules fed by air powered diaphragm pump. Membrane plates added towards end of 105 day study. Membranes
cleaned typically once every 15 to 20 days with dilute caustic solution. Certain membrane information derived from Dorr-Oliver operating manual or Li and Corrado, 1986.

48

Table 1. Anaerobic MBR System Design and Performance Information (Contd)

Wastewater

Sutton et al. (1983); 25


Whey permeate

Reference, Entry No. and Reported Information


Li and Corrado (1986); 26
Sutton (1986a, 1986b); 27
Cheese whey permeate
Wheat starch

Kataoka et al. (1992); 28


Domestic sewage

System

Pilot scale

Full scale; b

Pilot scale

Pilot scale; d

Objectives

Assess performance

Demonstrate performance

Assess performance

Assess bacterial populations

189

37,850

303

UASB 1600

Concentrate return

Concentrate return

Approximately 30,000 to 60,000

Concentrate return through


eductors
59,790

35,175

490 to complete system

95 to 97

Greater than 99

Greater than 99

83 for complete system


d

Parameter

Units

Bioreactor volume

liters (L)

Mixing method
Feed COD

mg/L

COD removal

%
3

OLR

kg COD/m day

7.5 to 16.3

7.0 to 9.1

Approximately 8.2

MLSS

mg/L

Approximately 33,000

29,500

VSS 22,400

Temperature

Approximately 35

Approximately 35

Approximately 35

pH
HRT

6.8 to 7.2
1.9 and 7.4

6.8 to 7.2
7.5

6.8 to 7.1

days

SRT

days

25 and 50

50

30

Membrane MOC

O.3 m3 CH4/kg COD removed at


STP
Polyethersulfone flat plate

O.3 m3 CH4/kg COD removed at


STP
Polyethersulfone flat plate

O.29 m3 CH4/kg COD removed at


STP
Polyethersulfone flat plate

Manufacturer

Dorr Oliver, USA

Dorr Oliver, USA

Dorr Oliver, USA

Approximately 10,000

Approximately 10,000

Approximately 10,000

15,000

External

External

External

External
100

Gas production

Pore size

micrometers

MW exclusion

daltons

Membrane location
Surface area

m2

0.47; a

10.7

0.47

Operating permeate flux

L/m2day

0.5 to 1.6; a

0.5 to 1.5; b

0.5 to 1.6; c

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP

m/s

Inlet pressure
Outlet pressure
Other information
a
b
c
d

System - 18 l CH4/kg COD


removed at STP
Polysulfone and PVA

Up to 6.1

kg/cm2
2

kg/cm

0.7 to 1.1

Up to 2.8

0.7 to 1.1

Maximum 4.2

Maximum 3.5

Maximum 4.2

kg/cm2

0.7
a

Certain membrane information derived from Dorr-Oliver operating manual. Membrane module fed by air powered diaphragm pump.
System was a full scale demonstration plant. Certain information derived from Dorr-Oliver operating manual.
Certain information derived from Dorr-Oliver operating manual. Membrane module fed by air powered diaphragm pump.
Hydrolization reactor plus UASB reactor coupled to external UF module for methane fermentation. Loading rate to UASB 2.8 kg BOD/m3day. Hydrolization reactor temperature and pH respectively, 35C and 6.
UASB operated at ordinary temperature.

49

Table 1. Anaerobic MBR System Design and Performance Information (Contd)

Wastewater

Kataoka et al. (1992); 29


Municipal sewage

Reference, Entry No. and Reported Information


Kataoka et al. (1992); 30
Grethlein (1978); 31
Soybean processing
Sanitary

Kayawake et al. (1991); 32


Heat treat liquor

System

Pilot scale; a

Pilot scale; b

Laboratory scale; c

Lab scale

Assess bacterial populations

Assess bacterial populations

Assess system feasibility

Assess performance

Total 2140

106

200
Mixed-liquor recirculation

1357 to complete system

Tank contents recycle and


concentrate return
BOD 270

Parameter

Units

Objectives
Bioreactor volume

liters (L)

Mixing method
Feed COD

mg/L

353 to complete system

9230 to 10,630

COD removal

90 for complete system

72 for complete system

BOD 85 to 93

79 to 83

OLR

kg COD/m3day

2.0 (BOD basis)

4.5 to 19.4

MLSS

mg/L

Temperature

MLVSS 10,200 to 21,400

30

pH
HRT

Acidogenic 6.0, methanogenic 7.5


Acidogenic 0.14 to 0.15,
methanogenic 0.28 to 0.29

6.5 to 7.2

days

a
5 for hydrolization reactor

7.7 to 8.1
0.58 to 2.02

SRT

days
System - 11 L CH4/kg COD
removed at STP
a

201 L CH4/kg COD removed at


STP
Polysulfone and PVA

2.98 L/day

Approximately 0.42 m3 total


gas/kg COD fed
Ceramic tubular

Gas production
Membrane MOC

35 to 38

Manufacturer

Kubota, Japan

Pore size

micrometers

MW exclusion

daltons

15,000

10 to 40

External

External

Internal, d

m2

Hydrolization reactor internal,


FBR - external
a

50

Approximately 0.007; c

1.06

Membrane location
Surface area

0.1

Operating permeate flux

L/m day

0.7 to 1.06; c

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP

m/s

0.15 to 1.22; c

0.2 to 0.3

0.35 to 1.35; c

6.7; d

Inlet pressure
Outlet pressure

kg/cm2
2

kg/cm
kg/cm2

Other information
a
c
d
a
Hydrolization reactor coupled to UF module, plus FBR coupled to MF membrane module for methane fermentation. Hydrolization reactor and FBR volumes respectively, 500 and 660 l. Loading rate to FBR 1.1
kg BOD/m3day. Hydrolization reactor temperature and pH respectively, 30C and 5.5 to 6.5. FBR operated at ordinary temperature. MOC, MW exclusion or pore size, and surface area of membrane coupled to
hydrolization reactor and FBR respectively, polyacrylonitrile and polyethylene, 13,000 and 0.1 micrometers, and 0.94 and 54 m2.
b
Two-phase (acidogenic plus methanogenic) PBR coupled to UF membrane.
c
Laboratory system modeled a septic tank coupled to external membrane. Claimed rate of BOD reduction in septic tank was 24 mg/Lday. Variable stroke positive displacement membrane feed pump operated in
cyclical fashion 2 to 3 min on 1 min off which aided performance. Used flat sheet and tubular RO membrane modules. Membrane data for flat sheet modules.
d
Tubular membrane submerged in the bioreactor. Membrane module backwashed with nitrogen gas and at times process gas, every 30 to 60 min for 30 sec. TMP created by suction. Stabilized flux ranged from 80 to
200 L/m2day.

50

Table 1. Anaerobic MBR System Design and Performance Information (Contd)


Parameter

Units

Wastewater

Hogetsu et al. (1992); 33


Wool scouring

System

Pilot scale; a

Objectives

Assess effectiveness of membrane


filtration
4500

Bioreactor volume

liters (L)

mg/L

Bailey et al. (1994); 36


Synthetic
Lab scale; d
d

5500

Approximately 9

Process gas

Mixed-liquor recirculation

By feed

30,700 to 50,100; c

5000
98 to 99

Mixing method
Feed COD

Reference, Entry No. and Reported Information


Kiriyama et al.(1992 and 1994); 34 Nagano et al. (1992); 35
Distillery; c
Filtered (20 m) municipal
wastewater solids
Demonstration scale two-phase
Pilot scale
anaerobic system; b
Derive performance information
Assess performance

COD removal

25 to 95; a

Greater than 95

OLR

kg COD/m3day

5 to 50 (TOD basis)

2.2 to 10.2

MLSS

mg/L

Temperature

pH
HRT

days

SRT

days

Maximum 20,000
35 and 55

25

8 to 8.6

6000; d

37
6.8 to 7.2

3.5 to 5.0
25.4 to 445

Gas production

0.2 to 0.3 m3 total gas/kg removed

107 to 467 L CH4/kg VSS reduced

0.28 to 0.34 m3 CH4 /kg COD fed

Membrane MOC

Polyacrylonitrile hollow fiber

Polymeric

Polysulfone

Woven polyester tube

Schweiz Seidengazefabriek,
Germany

Manufacturer
Pore size

micrometers

MW exclusion

daltons

Membrane location
2

13,000

8000

External

External
3.8 and 4.3

Surface area

3.1

Operating permeate flux

L/m2day

1.5 to 1.0 for 1st 40 days, decreased


to 0.7 after 210 days

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP

m/s

kg/cm

2.0 to 2.2

Outlet pressure

kg/cm2

0.6 to 1.5

c
d

12
Approximately 480

kg/cm2

Inlet pressure

External

0.6

Other information

2,000,000

2.2 to 3.6

1.5

2 to 2.5

1.5
b

PBR coupled to membrane module with concentrate return to PBR or discharged as blowdown. Wastewater TOD and BOD respectively, 102,400 and 27,900. TOD removal higher under mesophilic conditions and
decreased with increasing OLR. Membrane flushed with hot water (30 to 40C) periodically. Oxalic acid cleaning after 210 days did not restore flux.
First stage MBR (i.e., solids hydrolization reactor) 8850 l. Second stage UASB reactor (volume 76,700 L) treated municipal wastewater filtrate and permeate from first stage MBR. UASB effluent treated in BAF.
Added backwash solids from BAF to first stage MBR in some experiments. Reduction of VSS in MBR up to 76 percent. Complete system treated up to 240 m3/day of degritted municipal wastewater.
Wastewater from distillation of wheat and sweet potatoes. Feed VSS 12,600 to 17,400 mg/L.
Short term (i.e., 17 days) study of UASB reactor with membrane filtration of diffuse solids layer at top of reactor. Improve effluent quality from UASB. TSS at top of UASB increased from 1 to 6 g/L over study period.
Diatomaceous earth precoat applied to membrane filter.

51

Table 1. Anaerobic MBR System Design and Performance Information (Contd)


Parameter

Units

Wastewater

Cadi et al. (1994); 37


Synthetic starch

System

Laboratory scale

Objectives
Bioreactor volume

a
liters (L)

Mixing method
Feed COD

Reference, Entry No. and Reported Information


Harada et al. (1994); 38
Okamura (1994); 39
Synthetic with high particulate
Kraft mill evaporator condensate; c
COD
Lab scale
Pilot scale, c

mg/L

Total system volume 6.5 to 7.5

10

Concentrate return

Mechanical

460 to 9700

5000

Kitamura et al. (1996); 40


Distillery
Lab scale

Assess performance

Assess performance

5000

300

19,200

COD removal

78 to 87

Greater than 98

92

OLR

kg COD/m3day

1.9 to 24.2

1.5 to 2.5

38.5

MLSS

mg/L

MLVSS up to 38,000

Approximately 15,000

9400

MLVSS 5000 to 16,000

Temperature

35

35

52

36 to 38

pH
HRT

days

7
0.25 to 5.63

b
2 to 5

6.9 to 7.0
0.5

Approximately 7
2.3 to 7.0

SRT

days

45 to 52
0.36 to 0.47 m3 total gas/kg COD
fed
c

0.4 to 0.6 m3 total gas/kg VS


present; d

Gas production

0.3 m3 CH4/kg COD removed

Membrane MOC

Ceramic

Polysulfone flat plate

SCT, France

UF-3000PS, Toso, Japan

Manufacturer
Pore size

micrometers

MW exclusion

daltons

Membrane location

3,000,000
External

External

Surface area

m2

0.02

Operating permeate flux

L/m2day

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP

m/s

2 to 2.5

0.8; b

kg/cm2

0.4 to 1.6

0.5

Inlet pressure
Outlet pressure

kg/cm2
a

Other information
a

DDS, Holland

0.2
20,000
External
0.22

kg/cm2
c

Derived performance information over a range of HRTs and OLRs. Two membrane modules, each 0.2 m2, one operating and one standby. Permeate flux varied from 0.1 to 0.6 dictated by bioreactor VS and TMP.
Membrane module cleaned periodically with 2 % NaOH at 60C.
SMP major soluble COD in reactor but not much in permeate (i.e., retained COD). 2,100 mg NaHCO3/L was added to feed to provide alkalinity. Membrane process pump operated intermittently. Flux stabilized at
approximately 500 L/m2day after approximately 70 days with water flush for cleaning done every 7 to 10 days. Membranes replaced with new ones after approximately 70 days.
Condensate pretreated to remove inhibitive sulphur compounds. PBR coupled to UF membrane module for biomass retention and recycling. Organic plate type inorganic tubular membrane modules operated with
no further information provided.
Feed COD concentration information unclear but exceeded 50,000 mg/L. Achieved up to 80 % removal of feed volatile acids. Feed volatile acid loading rate from 2 to 13 kg/m3day. UF membrane cleaned by water
flushing only. This was not sufficient to prevent membrane fouling. Diluted feed by factor of 2.6 to 2.8 with water to prevent ammonia inhibition. Gas production decreased when process failed due to excessive
biomass wasting. As membrane fouled, increased wasting of biomass to reduce bioreactor liquid volume.

52

Table 1. Anaerobic MBR System Design and Performance Information (Contd)


Parameter

Units

Wastewater

Brockman and Seyfried (1996 and


1997); 41
Potato starch

System

Pilot scale; a

Objectives
Bioreactor volume

Reference, Entry No. and Reported Information


Ghyoot and Verstraete (1997); 42
Elmaleh and Abelmoumni
(1997and 1998); 43
Primary municipal wastewater
Acetic acid
sludge
Pilot scale
Lab scale
Improve digester performance

liters (L)

mg/L

COD removal

%
3

Synthetic starch
Lab scale two-phase system; d
d

120

10

Mixed-liquor recirculation

Concentrate return

33,000

40,200; b

2000 as TOC

10,000 to 15,000

29 to 54

Exceeded 95 as TOC

Mixing method
Feed COD

Study membrane filtration

Chung et al. (1998), 44

OLR

kg COD/m day

2.5 to 6.0

0.9 to 1.2

Acidogenic 2.5 to 7.5

MLSS

mg/L

22,000 to 35,000 as TS

Up to 10,000

Temperature

35

35

pH
HRT

days

20

8.5
0.4 to 13.8

SRT

days

c
0.08 to 0.16 m3 CH4/kg COD fed

0.7 m3 total gas/kg TOC fed

Tubular PVDF

Ceramic

0.1

0.1

Tubular zirconia skin and carbon


support
.
0.05 to 0.2; c

0.6

External

External

Internal

Gas production
Membrane MOC
Manufacturer
Pore size

micrometers

MW exclusion

daltons

Membrane location

External

Surface area

m2

0.05

Operating permeate flux

L/m2day

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP
Inlet pressure

m/s

4.5

Outlet pressure

kg/cm2

kg/cm2
kg/cm2

Acidogenic 25 and
methanogenic 35
d
Acidogenic 4

1.0

Other information
a
b
c
d
a
Complete mix reactor followed by clarifier and membrane feed/concentrate tank. Membrane operated batch-wise with concentrate produced returned to reactor intermittently. Claimed performance at 6 to 7 kg
COD/m3day not as good as other reactor configuration (e.g., PBR). Activity decreased due to sludge pumping to membrane module. Struvite precipitation observed.
b
Feed TS 44 g/L and VS 22 g/L. Membrane filtration of 20 L reactor contents performed every other day and permeate discarded. Membrane component of MBR reduced activity of biomass with respect to biogas
production.
c
No solids wasted over 330 day study. Membrane module fitted with different filtration elements for performance testing. Flux determined over wide range of TMP and linear velocities. Use of helical baffles in
membrane tube allowed flux of over 4000 L/m2day at TMP approximately 1 kg/cm2.
d
Membrane located in acidogenic reactor. Compare performance to conventional two-phase system. At COD loading of 5 kg COD/m3day, COD conversion to organic acids was 34 to 50 %. ORP controlled in
acidogenic reactor to prevent sulphate reduction and methane formation. Membrane flux varied between 0.08 and 0.38 L/m2day at MLSS of 3800 mg/L. Membrane unit backwashed periodically with air to prevent
fouling.

53

Table 1. Anaerobic MBR System Design and Performance Information (Contd)

Wastewater

Yushina and Hasegawa (1998); 45


Soybean Processing

Reference, Entry No. and Reported Information


Wen et al. (1999); 46
Beaubien et al. (1996); 47
Municipal
Synthetic; acetate solution

Madokoro et al. (1999);48


d

System

Pilot scale; a

Lab scale; b

Lab scale

Pilot scale; d

Objectives

Obtain application information

Obtain application information

17.7

Evaluate impact of biological and


physical-chemical parameters
Working volume 9; c

By the feed

By feed concentrate return

Mechanical and process gas

98 to 2600

59,500

Parameter

Bioreactor volume

Units

liters

Mixing method
1300 to 1350

1670

Feed COD

mg/L

COD removal

97; b

65 to greater than 95; c

86

OLR

kg COD/m3day

3 to 3.5

0.5 to 12.5

18.7

MLSS

mg/L

16,000 to 21,500; b

2200 to 25,000

57,000

Temperature

12 to 27

35

55

4 and 6; b

7
1

6.5

150

30

20 to 25

0.13 to 0.42 m3/m3 of reactor


volume/day at 53 to 66% CH4
Hollow fibre polyethylene

Ceramic flat plate

Polyethylene flat plate

Mitsubishi Rayon, Japan


0.03

0.2

Kubota, Japan
0.4

Internal

c
External

Internal

Approximately 30

pH
HRT

days

Acidogenic 6.0, methanogenic 7.5


Acidogenic 0.14,
methanogenic 0.28

SRT

days

Gas production

Membrane MOC

Polysulphone and PVA capillary

Manufacturer
Pore size

Chiyoda, Japan
micrometers

MW exclusion
Membrane location

daltons

Surface area
Operating permeate flux

m2
L/m2day

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP
Inlet pressure

m/s

Outlet pressure

kg/cm2

Other information
a

15,000
External
50

0.3
120 to 240; b

0.2

2.4
100

2; c

kg/cm2
kg/cm2
a

0.02 to 0.70

0.34; c

0.01 to 0.03

Operated acidogenic plus methanogenic PBRs followed by membrane module (System B), and acidogenic PBR followed by membrane module plus methanogenic PBR (System C). To compare conventional and
MBR anaerobic systems. Acidogenic and methanogenic PBRs EBV respectively, 540 and 1600 L. Solids captured in membrane returned to preceding PBR. COD removal in system B and C respectively, 78 and 92 %.
Total gas production System B and C respectively, 2.35 and 3.06 m3/day at STP. System B provided superior performance.
Cylindrical reactor (sludge zone) with expanded section on top to reduce TSS to membrane module. Fine fibers located at top of cylindrical section to prevent solids washout at high HLR. HRT is value in sludge
zone. MLSS includes suspended biomass and that attached to fine fibers. COD removal approximately 88% at lower temperatures. Membrane suction pump operated intermittently. Membrane module removed
periodically and cleaned with tap water and 5% NaClO.
Working volume includes membrane filtration unit. COD removal dictated by F/M. Membrane linear velocity and TMP varied over short term test periods to develop membrane performance information. Membrane
chemically cleaned between tests.
Macerated, screened kitchen refuse diluted and treated in two-stage (i.e., acid plus methane) thermophilic system with membrane located in methanogenic reactor. Process gas used to scour membrane plates. Citric
acid solution (10,000 mg/L) used to soak membrane cartridge for 4 hours approximately once every 3 months.

54

Table 1. Anaerobic MBR System Design and Performance Information (Contd)


Parameter

Units

400

Reference, Entry No. and Reported Information


Lee et al. (2001c); 50
Kim et al. (2002); 51
Piggery
Primary municipal wastewater
sludge; c
Pilot plant scale two-phase system;
Pilot scale
b
Improve performance of
Recovery of volatile fatty acids
acidogenic phase
Acidogenic 3000; b
Total system volume 76

Mechanical

Mechanical

2000 to 15,000

5000 to 6000

Wastewater

Yilang et al. (1999 and 2002); 49


Food processing

System

Pilot scale

Objectives

Assess performance

Bioreactor volume

liters (L)

Mixing method
Feed COD

mg/L

COD removal

70 to 95; a

OLR

kg COD/m3day

2 to 4.5

0.6 to 7.5 as TOC

MLSS

mg/L

6000 to 8000

Temperature

Hernandez et al. (2002); 52


Diluted molasses
Lab scale; d
Improve performance of UASB
reactors
5
By the feed

0.3 to 13; d

35

Acidogenic 20, methanogenic 35

35

20

pH
HRT

days

6.8 to 7.2
0.6 and 4.2

Acidogenic 6 to 7
Acidogenic 1 to 2

0.3 to 4

2 to 5

SRT

days

Gas production

Membrane MOC
Manufacturer
Pore size

Polymeric flat plate

Methanogenic reactor 0.32 m3


CH4/kg COD removed
Mixed esters of cellulose

Ceramic

0.5

1.0

10 and 100; d

Internal in acidogenic reactor

External

Internal

10

micrometers

MW exclusion
Membrane location

daltons

20,000 to 70,000

Surface area
Operating permeate flux
Linear velocity of the mixed
liquor in membrane
tube/channel
TMP
Inlet pressure

m2
L/m2day
m/s

Total 0.64; a
a

Outlet pressure

kg/cm2

Other information
a

0.051 and 0.056


480 to 2400; d

0 to 960; b
0.4

kg/cm2
kg/cm2
a

0.4

0 to 0.01

Authors claim reactor was completely mixed. COD removal decreased with increasing OLR. Gas production (approximately 50 to 250 L/day) increased with increasing OLR. Membrane module flushed each day with
permeate and chemically cleaned every 10 days with 0.5 % NaOH. Membrane module consisted of 8 plates, with membranes of different pore size. Fluxes ranged from approximately 0.4 to 1.7 L/m2day after long
term operation (i.e., 110 days) and multiple (i.e., 10) chemical cleanings. Membrane with smallest MW exclusion performed best in long term. Authors state membrane flux decreased with time.
Methanogenic reactor was UASB with plastic media in upper section having total volume of 3000 l. Complete system COD removal 50 to 80 percent. Membrane located in mechanically mixed acidogenic reactor.
Stainless steel prefilter (63 m pore size) prior to membrane used to improve membrane performance. Flux decreased with time but restored to 89% of clean membrane with NaOH and HCl cleaning.
Feed was coagulated (polyaluminum and ferric chloride), primary municipal wastewater sludge. Mean feed TOC 1.82 g/l, TS 5600 mg/L and VSS 3960 mg/L. TS in bioreactor 33.8 g/L at HRT 0.5 days. Converted
organics to over 40 % volatile fatty acids.
Four UASB reactors operated each containing membrane cartridges. UASB hydraulic loading rate 2.5 to 4.5 m/h. The 10 and 100 m cartridges were composed of respectively, polypropylene and fiberglass.
Cleaned filtration cartridges by backpulsing with tap water. Operating permeate flux dictated frequency of backwash, bioreactor OLR and cartridge pore size. Irreversible cartridge fouling occurred within 90 days.

55

Table 1. Anaerobic MBR System Design and Performance Information (Contd)

Wastewater

Stuckey and Hu (2003); 53


Synthetic municipal wastewater

Reference, Entry No. and Reported Information


Baek and Pagilla (2003); 54
Fuchs et al. (2003); 55
Municipal primary effluent
c

Minami et al. (1991); 56


Kraft mill evaporator condensate, d

System

Lab scale

Lab scale

Bench scale

Pilot scale; d

Objectives

Evaluate various operating


parameters; a

Compare aerobic versus anaerobic


MBRs
Working volume 10 (each)

Obtain application information

Assess performance

7; c

5000

Parameter

Bioreactor volume

Units

liters (L)

Mixing method
Feed COD

mg/L

COD removal

Mechanical

Mechanical

430

84 (soluble)

5,800 to 64,600

5,400 (TOC basis)

Up to 93; a

55 to 68

Up to 97

85 to 90 (TOC basis)

0.03 to 1.64 (soluble COD)

6 to 20; c

8.5 to 10.5 (TOC basis)

OLR

kg COD/m day

MLSS

mg/L

MLVSS approximately 3000

1010 to 7120

55,000 to 60,000

Temperature

35

32

30

53

pH
HRT

days

7.0
0.13 to 1.33

6.5
0.5 to 2

SRT

days

1.08 m3 CH4/kg TOC fed

19 to 233

Gas production
Membrane MOC
Manufacturer
Pore size

micrometers

MW exclusion
Membrane location

daltons

Surface area
Operating permeate flux

m2
L/m2day

Linear velocity of the mixed


liquor in membrane
tube/channel
TMP
Inlet pressure

m/s

Outlet pressure

kg/cm2

Other information
a

Tubular PVDF

0.1 to 0.35 L CH4/g COD removed


at STP
Ceramic; Al2O3

Both 0.4

PCI, U.S.
0.1

0.2

Internal

200,000
External

External

Both 0.1
a

0.10
b

0.13
120 to 240

Polysulfone
Kurita, Japan
2,000,000
External

2 to 3

kg/cm2
kg/cm2

0 to 0.7

Parallel operating MBRs with different membranes. Membrane A; hollow fiber, polyethylene, hydrophilic. Membrane B; Kubota, polyethylene, hydrophilic. Gas sparging at 5 L/min prevented cake layer formation.
Sparging rate 0.5 L/min. Permeate pressure, flow and pump control. PAC addition improved COD removal and flux. Effluent COD down to 30 mg/L at HRT of 0.13 days.
Five, sequential experimental runs completed over period of 166 days after 100 day initial operating period. Steady state conditions not necessarily achieved. Membrane operated at TMP required to ensure permeate
rate exceeded feed rate. Expressed concern that membrane process pump shear may have caused biomass decay.
Artificial, slaughterhouse and sauerkraut brine wastewaters. Bioreactor volume includes membrane filtration loop. OLR for industrial wastewaters ranged from 6 to 8 kg COD/m3day. Membrane filtration performance
not optimized.
Condensate pretreated to remove oils and sulphur compounds. Upflow PBR coupled to UF membrane for microbial retention. Membrane concentrate returned to PBR resulting in liquid phase TSS up to 9.4 g/L. This
study represents a follow-up to bench scale studies reported by Yamaguchi et al. (1990)

56

Table 2. Operating Conditions and Performance Results During Anaerobic MBR Treatment of Wheat Starch Wastewater
(Adapted from Table 1, entry 27)

Operating Condition or Performance Parameter1

Value

Reactor SRT, days

30

Volumetric Loading, kg/m3.day

8.2

Reactor Volatile Suspended Solids, mg/L


Feed Values, mg/L
BOD5
COD
TSS
Effluent Values, mg/L
BOD5
COD
TSS
1

22,400

15,463
35,175
13,300
74
270
< 10

All values are means during operation under equilibrium conditions.

57

APPENDIX A
National and International Researchers Contacted
Individual
Graham Anderson and Tom Donnelly

Institution/Research Organization
University of Newcastle, Newcastle upon Tyne, United Kingdom

Dr. Bhattacharyya

University of Kentucky, Lexington, KY

Trevor Bridle

ESI Ltd., Burswold, Australia

Chris Buckley

University of Natal, Durban, South Africa

Herve Buisson

Anjou Recherche, Maisons-Laffitte, France

Peter Cartwright

Cartwright Consulting, Minneapolis, MN

John Collins

Ondeo-Nalco, Naperville, IL

George Crawford

CH2M-Hill, Toronto, Canada

Chris Davis

Australian Water Association, Artarmon, Australia

Genevieve Gesan-Guiziou

LRTL INRA, Rennes, France

Bob Hickey

Ecovation, Lansing, Michigan

Xia Huang

Tsinghua University, Beijing, China

Ferdinand Klegraf

VA Tech Wabag, Kulmbach, Germany

Carl Koch

Greeley & Hansen, Philadelphia, PA

Stefan Krause

Darmstadt University, Darmstadt, Germany

Yelte Lanting

Biothane, Camden, NJ

Chung-Hak Lee

Seoul National University, Seoul, Korea

S-M Lee

Water Environment Research Center, Korean Institute of Science &


Technology, Seoul, South Korea

Eberhard Morgenroth

University of Illinois at Urbana-Champaign, Urbana, IL

Krishna Pagilla

Illinois Institute of Technology, Chicago, IL

Dr. Pirbazari

University of Southern California,Los Angeles, CA

Roger Pujol

Lyonnnaise-Des-Eaux, Paris, France

Bruce Rittman

Northwestern University, Evanstown, IL

Jan Schipper

IHE Delft, Delft, The Netherland

Ed Schroeder

University of California Davis, Davis, CA

Steven Sliss

Ryerson University, Toronto, Canada

Dr. M. Soto

University of A Coruna, Galiza, Spain

David Stensel

University of Washington, Seattle, WA

Michael Stenstrom

University of California Los Angeles, Los Angeles, CA

Tom Stephenson

Cranfield University, Cranfield, United Kingdom

Makram Suidan

University of Cincinnati, Cincinnati, OH

Vincent Urbain

FAIRTEC, Gargenville, France

R.T. van Houten

TNO, Apeldoorn, The Netherlands

Andre van Niekerk

Golder Associates, Johannesburg, South Africa

Dr. Veerapaneni

Black & Veatch, Kansas City, KS

Willy Verstraete

University Gent, Gent, Belgium

Marc Wichern and Karl-Heinz Rosenwinkel

University of Hannover, Hannover, Germany

Prof. Kazuo Yamamoto

University of Tokyo

58

Commercial MBR System and/or Membrane Suppliers Contacted


Individual
John Arnold

Company and Location


Toray Membrane America, USA

Bill Bonkoski

Zenon Environmental, Canada

Francis Brady

Koch Industries, USA

Steve Churchouse

Aquator Groupa, England

Al Cocci

ADI Systemsa, Canada

Brad Culkin

New Logic International, USA

Poul Ejner

Bioscan A/S, Denmark

Willy Gils

Waterleau Global Water Technology, Belgium

Torsten Hackner

Hans Huber AG, Germany

Matt Kuzma

US Filter, USA

Bernie Mack

Ionics Corporationb, USA

Francis McKeever

Aqua Aerobics, USA

Masashi Moro

Kubota, Japan

Uri Papouktchiev

Norit Americas, USA

David Pearson

PCI/ITT Industries, USA

Tony Robinson

Wehrle Environmental, England

a
b

Kubota licensee.
Mitsubishi licensee

59

APPENDIX B
Nomenclature List
ABR
ADUF AOX BAF
BOD CBOD DO
EBV
EPS
FBR
F/M
HLR
HRT
MARS MBR MF
MLSS MLVSS MOC MW
NF
OLR
ORP
PAC
PBR
PP
PVA
PVC
PVDF RO
SEM
SMP
SRT
SS
STP
TAP
TMP
TOC
TOD
UASB UF
VS
VSS
-

anaerobic baffled reactor


anaerobic digestion ultrafiltration
adsorbable organic halogen
biological aerated filter
biochemical oxygen demand
carbonaceous biochemical oxygen demand
dissolved oxygen
empty bed volume
exocellular polymer substances
fluidized bed reactor
food to mass ratio
hydraulic loading rate
hydraulic retention time
membrane anaerobic reactor system
membrane biological reactor
microfiltration
mixed-liquor suspended solids
mixed-liquor volatile suspended solids
materials of construction
molecular weight
nanofiltration
organic loading rate
oxidation reduction potential
powdered activated carbon
packed bed reactor
polypropylene
polyvinyl alcohol
polyvinyl chloride
polyvinyl difluoride
reverse osmosis
scanning electron microscope
soluble microbial products
solids retention time
suspended solids
standard temperature and pressure
technical advisory panel
transmembrane pressure
total organic carbon
total oxygen demand
upflow anaerobic sludge blanket
ultrafiltration
volatile solids
volatile suspended solids

60

Alabama
Montgomery Water Works &
Sanitary Sewer Board

Union Sanitary District

Iowa

West Valley Sanitation District

Ames, City of

Colorado

Cedar Rapids Wastewater


Facility

Boulder, City of
Colorado Springs Utilities

Des Moines Metro Wastewater


Reclamation Authority

Anchorage Water &


Wastewater Utility

Littleton/Englewood Water
Pollution Control Plant

Iowa City, City of

Arizona

Metro Wastewater Reclamation


District, Denver

Alaska

Gila Resources
Glendale, City of, Utilities
Department

Connecticut

Kansas
Johnson County Unified
Wastewater Districts

Mesa, City of

New Haven, City of, WPCA

Unified Government of
Wyandotte County/Kansas
City, City of

Peoria, City of

District of Columbia

Kentucky

Pima County Wastewater


Management

District of Columbia Water &


Sewer Authority

Arkansas

Florida

Little Rock Wastewater Utility

California
Calaveras County Water District

The Mattabasset District

Louisville & Jefferson County


Metropolitan Sewer District

Louisiana

Broward, County of

Sewerage & Water Board of


New Orleans

Fort Lauderdale, City of

Maine

Passaic Valley Sewerage


Commissioners

New York
New York City Department of
Environmental Protection
Rockland County Solid Waste
Management Authority/
Sewer District

North Carolina
Charlotte/Mecklenburg Utilities
Durham, City of
Metropolitan Sewerage District
of Buncombe County
Orange Water & Sewer
Authority

Ohio
Akron, City of
Butler County Department of
Environmental Services

Gainesville Regional Utilities

Bangor, City of

Columbus, City of

Central Contra Costa Sanitary


District

JEA

Portland Water District

Maryland

Metropolitan Sewer District of


Greater Cincinnati

Contra Costa Water District

Miami-Dade Water & Sewer


Authority

Crestline Sanitation District


Delta Diablo Sanitation District
Dublin San Ramon Services
District
East Bay Dischargers Authority
East Bay Municipal Utility
District
El Dorado Irrigation District
Fairfield-Suisun Sewer District
Fresno Department of Public
Utilities
Irvine Ranch Water District
Las Virgenes Municipal Water
District
Livermore, City of
Lodi, City of
Los Angeles, City of
Los Angeles County, Sanitation
Districts of

Orange County Utilities


Department
Orlando, City of
Reedy Creek Improvement District
Seminole County Environmental
Services
St. Petersburg, City of
Stuart Public Utilities
Tallahassee, City of
Tampa, City of

Anne Arundel County Bureau of


Utility Operations
Howard County Department of
Public Works

Massachusetts

Clean Water Services

Boston Water & Sewer


Commission
Upper Blackstone Water
Pollution Abatement District

Michigan

Georgia

Holland Board of Public Works

University Area Joint Authority,


State College

Atlanta Department of
Watershed Management

Lansing, City of

Augusta, City of

Owosso Mid-Shiawassee
County WWTP

Clayton County Water Authority

Saginaw, City of

Cobb County Water System

Palo Alto, City of

Gwinnett County Department of


Public Utilities

Minnesota

Riverside, City of

Savannah, City of

Sacramento Regional County


Sanitation District

Hawaii

San Diego Metropolitan


Wastewater Department, City of
San Francisco, City & County of

Sunnyvale, City of

Pennsylvania

Detroit, City of

Wyoming, City of

South Orange County


Wastewater Authority

Water Environment Services


Philadelphia, City of

Fulton County

South Coast Water District

Eugene/Springfield Water
Pollution Control

West Palm Beach, City of

Orange County Sanitation


District

South Bayside System Authority

Oregon

Ann Arbor, City of

Toho Water Authority

Wayne County Department of


Environment

Santa Rosa, City of

Oklahoma
Tulsa, City of

Columbus Water Works

Santa Barbara, City of

Summit, County of

Washington Suburban Sanitary


Commission

Napa Sanitation District

San Jose, City of

Northeast Ohio Regional Sewer


District

South Carolina
Charleston Commissioners of
Public Works
Mount Pleasant Waterworks &
Sewer Commission
Spartanburg Sanitary Sewer
District

Tennessee
Cleveland, City of

Rochester, City of

Knoxville Utilities Board

Western Lake Superior Sanitary


District

Murfreesboro Water & Sewer


Department

Honolulu, City and County of

Missouri

Nashville Metro Water Services

Illinois

Independence, City of

American Bottoms Wastewater


Treatment Plant

Kansas City Missouri Water


Services Department

Texas

Greater Peoria Sanitary District

Little Blue Valley Sewer District

Kankakee River Metropolitan


Agency

Nebraska

El Paso Water Utilities

Lincoln Wastewater System

Fort Worth, City of

Nevada

Metropolitan Water Reclamation


District of Greater Chicago

Austin, City of
Dallas Water Utilities
Denton, City of

North Shore Sanitary District

Henderson, City of

Gulf Coast Waste Disposal


Authority

Wheaton Sanitary District

New Jersey

Houston, City of

Bergen County Utilities Authority

San Antonio Water System


Trinity River Authority

WERF SUBSCRIBERS

WASTEWATER UTILITY

WERF SUBSCRIBERS

Utah
Salt Lake City Corporation

Virginia
Alexandria Sanitation Authority
Arlington, County of
Fairfax County Virginia
Hampton Roads Sanitation
District
Henrico, County of
Hopewell Regional Wastewater
Treatment Facility

STORMWATER UTILITY
California
Los Angeles, City of,
Department of Public Works

Black & Veatch

Metcalf & Eddy Inc.

Boyle Engineering Corporation

MWH

Brown & Caldwell

New England Organics

Burns & McDonnell

Odor & Corrosion Technology


Consultants Inc. (OCTC)

Monterey, City of

CABE Associates Inc.

San Francisco, City & County of

The Cadmus Group

Santa Rosa, City of

Camp Dresser & McKee Inc.

Sunnyvale, City of

Carollo Engineers Inc.

Colorado

Carpenter Environmental
Associates Inc.

Boulder, City of

CDS Technologies Inc.

Oswald Green, LLC


PA Government Services Inc.
Parametrix Inc.
Parsons
Post, Buckley, Schuh & Jernigan
R&D Engineering/Conestoga
Rover & Associates

Loudoun County Sanitation


Authority

Georgia

Lynchburg Regional WWTP

Iowa

Prince William County Service


Authority

Cedar Rapids Wastewater


Facility

Damon S. Williams Associates,


LLC

Richmond, City of

Des Moines Metro Wastewater


Reclamation Authority

David L. Sheridan, P.C.

Kansas

Earth Tech Inc.

Washington

Overland Park, City of

Eco-Matrix

Edmonds, City of

Kentucky

Ecovation

Everett, City of

Louisville & Jefferson County


Metropolitan Sewer District

Environmental Engineers
International

Maine

EMA Inc.

Synagro Technologies, Inc.

Seattle Public Utilities

Portland Water District

The Eshelman Company Inc.

Tetra Tech Inc.

Sunnyside, Port of

Minnesota

Finkbeiner, Pettis, & Strout Inc.

Trojan Technologies Inc.

Yakima, City of

Western Lake Superior Sanitary


District

Freese & Nichols, Inc.

URS Corporation

ftn Associates Inc.

Veolia Water NATC

Fuss & ONeill, Inc.

Wade-Trim Inc.

Gannett Fleming Inc.

Weston Solutions Inc.


Woodard & Curran

Rivanna Water & Sewer


Authority

King County Department of


Natural Resources

Wisconsin
Green Bay Metro Sewerage
District
Madison Metropolitan
Sewerage District

Griffin, City of

Chemtrac Systems Inc.


CH2M HILL

North Carolina
Charlotte, City of, Stormwater
Services

Dewling Associates, Inc.

The RETEC Group


RMC, Inc. (Raines, Melton &
Carella)
R.M. Towill Corporation
Ross & Associates Ltd.
Royce Technologies
SAIC Maritime Technical Group
Short Elliott Hendrickson, Inc.
Stantec Consulting Group, Inc.
Stormwater Management, Inc.

Pennsylvania

Golder Associates Inc.

Milwaukee Metropolitan
Sewerage District

Philadelphia, City of

Greeley and Hansen LLC

WRc/D&B, LLC
WWETCO, LLC

Racine, City of

Tennessee

Hazen & Sawyer, P.C.

Chattanooga Storm Water


Management

HDR Engineering Inc.

Zoeller Pump Company

Washington

HydroQual Inc.

Sheboygan Regional
Wastewater Treatment
Wausau Water Works

Seattle Public Utilities

Australia
South Australian Water Corp.
Sydney Water Corp.

STATE

HNTB Corporation

INDUSTRY

Infilco Degremont Inc.

American Electric Power

Ingersoll-Rand Energy Systems

ChevronTexaco Energy
Research & Technology
Company

Insituform Technologies Inc.

Arkansas Department of
Environmental Quality

Institute for Environmental


Technology & Industry, Korea
Jacobson Helgoth Consultants
Inc.

Dow Chemical Company

Canada

Fresno Metropolitan Flood


Control District, Calif.

Greater Vancouver Regional


District

Urban Drainage & Flood


Control District, Colo.

Jason Consultants Inc.

Eastman Kodak Company

Jordan, Jones, & Goulding Inc.

Eli Lilly & Company

KCI Technologies Inc.

Merck & Company Inc.

Kelly & Weaver, P.C.

ONDEO Services

Kennedy/Jenks Consultants

Procter & Gamble Company

Komline Sanderson Engineering


Corporation

PSEG Services Corp.

Lawler, Matusky & Skelly


Engineers, LLP

Severn Trent Services Inc.

Limno-Tech Inc.

United Water Services LLC

Water Corp. of Western


Australia

Toronto, City of, Ontario


Winnipeg, City of, Manitoba

CORPORATE

Mexico
Servicios de Agua y Drenaje de
Monterrey, I.P.D.

New Zealand
Watercare Services Limited

Singapore
Singapore Public Utilities Board

United Kingdom
Yorkshire Water Services Limited

ADS Environmental Services


The ADVENT Group Inc.
Alan Plummer & Associates
Alden Research Laboratory
Aqua-Aerobic Systems, Inc.
AquateamNorwegian Water
Technology Centre A/S

The Coca-Cola Company


DuPont Company

RWE Thames Water Plc

Lombardo Associates Inc.


Malcolm Pirnie Inc.

BaySaver Inc.

McKim & Creed

BioVir Laboratories, Inc.

MEC Analytical Systems Inc.

Note: List as of 7/1/04

Board of Directors
Chair
James F. Stahl
County Sanitation Districts of
Los Angeles County
Vice-Chair
Vernon D. Lucy
ONDEO Degremont Inc.
Secretary
William J. Bertera
Water Environment Federation
Treasurer
Karl W. Mueldener
Kansas Department of
Health & Environment

Mary E. Buzby, Ph.D.


Merck & Company Inc.
Dennis M. Diemer, P.E.
East Bay Municipal Utility District

John T. Novak, Ph.D.


Virginia Polytechnic Institute
& State University
Lynn H. Orphan
Kennedy/Jenks Consultants

Jerry N. Johnson
District of Columbia Water and
Sewer Authority

J. Michael Read
HDR, Inc.

Richard D. Kuchenrither, Ph.D.


Black & Veatch

James M. Tarpy
Nashville Metro Water Services

Alfonso R. Lopez
New York City Department of
Environmental Protection
Richard G. Luthy, Ph.D.
Stanford University

Murli Tolaney
MWH
Executive Director
Glenn Reinhardt

Research Council
Chair
John Thomas Novak, Ph.D.
Virginia Polytechnic Institute
& State University
Vice-Chair
Glen T. Daigger, Ph.D.
CH2M HILL
Robert G. Arnold, Ph.D.
University of Arizona, Tucson
Robin L. Autenrieth, Ph.D.
Texas A&M University

Geoffrey H. Grubbs
U.S. EPA
Mary A. Lappin, P.E.
Kansas City Water Services
Department
Keith J. Linn
Northeast Ohio Regional
Sewer District
Drew C. McAvoy, Ph.D.
Procter & Gamble Company

William L. Cairns, Ph.D.


Trojan Technologies, Inc.

Margaret H. Nellor, P.E.


County Sanitation Districts of
Los Angeles County

James Crook, Ph.D.


Water Reuse Consultant

Spyros Pavlostathis, Ph.D.


Georgia Institute of Technology

Steven M. Rogowski, P.E.


Metro Wastewater Reclamation
District of Denver
Peter J. Ruffier
Eugene/Springfield Water Pollution
Control
Michael W. Sweeney, Ph.D.
EMA Inc.
George Tchobanoglous, Ph.D.
Tchobanoglous Consulting
Gary Toranzos, Ph.D.
University of Puerto Rico

Stormwater Technical Advisory Committee


Chair
Robert E. Pitt, Ph.D., P.E., D.E.E.
University of Alabama
Vice-Chair
Ben Urbonas, P.E.
Urban Drainage and Flood Control
District

Christine Andersen, P.E.


City of Long Beach, California

Brian Marengo, P.E.


City of Philadelphia Water Department

Gail B. Boyd
URS Corporation

A. Charles Rowney, Ph.D.


Camp Dresser & McKee Inc.

Larry Coffman
Prince Georges County

James Wheeler, P.E.


U.S. EPA

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

Stock #

Email

Product

Method of Payment:

Quantity

To t a l

Postage &
Handling

(All orders must be prepaid.)

VA Residents Add
4.5% Sales Tax

q C h e ck or Money Order Enclosed


q Visa
q Mastercard
q A m e rican Express

Canadian Residents
Add 7% GST

______________________________________________________
Account No.

Unit Price

Exp. Date

TOTAL

______________________________________________________
Signature

S h ip p in g & Ha n dli n g:
Amount of Order

To Order (Subscribers Only):

United States

Canada & Mexico

All Others

Up to but not more than:

Add:

Add:

Add:

$20.00

$5.00 *

$8.00

50% of amount

30.00

5.50

8.00

40% of amount

40.00

6.00

8.00

50.00

6.50

14.00

60.00

7.00

14.00

80.00

8.00

14.00

100.00

10.00

21.00

150.00

12.50

28.00

200.00

15.00

35.00

Add 20% of order

Add 20% of order

More than $200.00

* m i n i mum amount for all orders


Note: Please make checks payable to the Water Environment Research Foundation.

on to www.werf.org and click


7 Log
on the Product Catalog.
e: (703) 684-2470
( PFahxo:n(703)
299-0742.
WERF
Attn: Subscriber Services
635 Slaters Lane
Alexandria, VA 22314-1177

To Order (Non-Subscribers):
Non-subscribers may be able to order
WERF publications either through
WEF (www.wef.org) or IWAP
(www.iwapublishing.com).Visit WERFs
website at www.werf.org for details.

02-CTS-4.qxd

9/20/04

2:37 PM

Page 1 (1,1)

Water Environment Research Foundation


635 Slaters Lane, Suite 300 Alexandria, VA 22314-1177
Phone: 703-684-2470 Fax: 703-299-0742 Email: werf@werf.org
www.werf.org
WERF Stock No. 02CTS4

Sept 04

You might also like