You are on page 1of 43

This article was downloaded by: [TU University of Technology Delft]

On: 12 May 2011


Access details: Access Details: [subscription number 731714369]
Publisher Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House, 37-
41 Mortimer Street, London W1T 3JH, UK

Critical Reviews in Environmental Science and Technology


Publication details, including instructions for authors and subscription information:
http://www.informaworld.com/smpp/title~content=t713606375

Anaerobic Membrane Bioreactors: Applications and Research Directions


Bao-Qiang Liaoa; Jeremy T. Kraemerb; David M. Bagleyb
a
Department of Chemical Engineering, Lakehead University, Thunder Bay, Ontario, Canada b
Department of Civil Engineering, University of Toronto, Toronto, Ontario, Canada

To cite this Article Liao, Bao-Qiang , Kraemer, Jeremy T. and Bagley, David M.(2006) 'Anaerobic Membrane Bioreactors:
Applications and Research Directions', Critical Reviews in Environmental Science and Technology, 36: 6, 489 — 530
To link to this Article: DOI: 10.1080/10643380600678146
URL: http://dx.doi.org/10.1080/10643380600678146

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.informaworld.com/terms-and-conditions-of-access.pdf

This article may be used for research, teaching and private study purposes. Any substantial or
systematic reproduction, re-distribution, re-selling, loan or sub-licensing, systematic supply or
distribution in any form to anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents
will be complete or accurate or up to date. The accuracy of any instructions, formulae and drug doses
should be independently verified with primary sources. The publisher shall not be liable for any loss,
actions, claims, proceedings, demand or costs or damages whatsoever or howsoever caused arising directly
or indirectly in connection with or arising out of the use of this material.
Critical Reviews in Environmental Science and Technology, 36:489–530, 2006
Copyright © Taylor & Francis Group, LLC
ISSN: 1064-3389 print / 1547-6537 online
DOI: 10.1080/10643380600678146

Anaerobic Membrane Bioreactors: Applications


and Research Directions

BAO-QIANG LIAO
Department of Chemical Engineering, Lakehead University, Thunder Bay, Ontario, Canada

JEREMY T. KRAEMER and DAVID M. BAGLEY


Department of Civil Engineering, University of Toronto, Toronto, Ontario, Canada
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

Membranes provide exceptional suspended solids removal and com-


plete biomass retention that can improve the biological treatment
process, but their commercial application to anaerobic treatment
has been limited. This review summarizes the state of the art with
respect to anaerobic membrane bioreactors (AnMBRs), determines
the types of wastewaters for which AnMBRs would be best suited,
and identifies the research required to increase implementation.
AnMBRs have been tested with synthetic, food processing, indus-
trial, high solids content, and municipal wastewaters at labora-
tory, pilot, and full scale. Chemical oxygen demand removal ranges
from 56% to 99%, while the reported design membrane fluxes range
from 10 to 40 L/m2 /h. AnMBRs should be immediately applicable
to highly concentrated, particulate waste streams like municipal
sludges where the membrane can decouple the solids and hydraulic
retention times. Opportunity for application to dilute wastewaters
also appears strong, while application to highly concentrated solu-
ble wastewaters is likely limited. Greater assessment of vacuum-
driven immersed membranes, combining external or immersed
membranes with retained biomass reactor designs, control of mem-
brane fouling, and economic feasibility are the key research areas
to be addressed.

KEY WORDS: anaerobic wastewater treatment, biogas production,


membrane configuration, membrane flux, membrane fouling

This work was supported by the Natural Sciences and Engineering Research Council of
Canada.
Address correspondence to David M. Bagley, Department of Civil Engineering, University
of Toronto, Toronto, ON, Canada M5S 1A4. E-mail: bagley@ecf.utoronto.ca

489
490 B.-Q. Liao et al.

I. INTRODUCTION

Anaerobic processes have been successfully used to treat industrial, food


processing, and agricultural wastewaters for more than a century. Currently,
there are nearly 1600 commercial anaerobic wastewater treatment systems
in operation in the world.59 Anaerobic processes are also widely used to
treat sludges from municipal wastewater treatment plants and successfully
treat animal manures and the organic fraction of municipal solid waste.59,101
Anaerobic processes have been less widely applied to dilute wastewaters,
such as municipal wastewater, but such applications do exist, in particular in
countries with warmer climates.98
Anaerobic processes offer several widely known advantages over con-
ventional aerobic processes. First, no oxygen is required, so the challenge,
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

expense, and energy required to dissolve oxygen into water are eliminated.
Second, methane is produced and serves as a renewable energy source.
Where the economics are favorable, this methane may be combusted to
produce electricity and heat. Finally, less biomass is produced. In the ab-
sence of oxygen as an electron acceptor, anaerobic microbial systems dis-
card the electrons onto methane instead of using them to grow more mi-
croorganisms. These advantages are offset by the slow growth rates of
the methanogenic organisms and the microbial complexity of the systems.
Biomass retention is critical to provide sufficient solids retention time (SRT)
for the methanogens, but even so the low effluent concentrations achieved
by aerobic processes are difficult to achieve. Consequently, anaerobic pro-
cesses are almost exclusively applied to the concentrated waste streams men-
tioned earlier, with aerobic processes used primarily for more dilute waste
streams.
The application of anaerobic processes to more dilute waste streams
may nevertheless be appropriate. Recently, Shizas and Bagley90 measured
the potential energy in the organics of municipal wastewater to be up
to nine times greater than the electricity needed to operate a municipal
wastewater treatment plant. The methane-rich biogas produced from anaer-
obic sludge digestion can be combusted to produce a significant fraction
of the electricity needed to run the plant. To achieve energy sustainabil-
ity, however, anaerobic treatment must be applied to the wastewater di-
rectly. This should increase the fraction of the potential energy recovered
as biogas, and will decrease the oxygen requirement. The latter could lead
to reduced energy expenditures, further improving the energy sustainabil-
ity calculation. Technical and economic challenges exist, however, with the
first and likely most important being the identification of suitable anaerobic
processes.
In recent years, membrane technologies have been successfully incorpo-
rated into the aerobic biological wastewater treatment process. These mem-
brane bioreactors (MBRs) have been proven for municipal and industrial
Anaerobic Membrane Bioreactors 491

wastewater treatment.92 A key advantage of these MBR systems is the


complete retention of biomass in the aerobic reactor. This eliminates the
impact of biomass separation problems that could deteriorate the perfor-
mance of conventional aerobic biological treatment and provides very low
suspended solids concentrations in the treated effluent. Indeed, the con-
cept of effluent suspended solids is almost meaningless when the mem-
brane pore size is smaller than the pore size of the filter used to measure
suspended solids. Equally important, complete biomass retention thor-
oughly decouples the SRT from the hydraulic retention time (HRT), allow-
ing biomass concentrations to increase in the reaction basin, thus facilitat-
ing relatively smaller reactors, if desired, and higher organic loading rates
(OLR).
If membranes work well with aerobic processes, why not with anaero-
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

bic processes? Would the anaerobic membrane bioreactor (AnMBR) com-


bine the advantages of the MBR and anaerobic processes? In particular,
would the complete retention of slow-growing methanogenic organisms im-
prove the applicability of anaerobic processes to more waste streams? How
would AnMBRs compare to existing high-rate anaerobic processes, for exam-
ple, the upflow anaerobic sludge blanket (UASB) reactor? What waste streams
would be most suitable to treat with AnMBRs? What design and operational
challenges exist?
The answers to these questions and others are not readily available in
a single source. The purpose of this literature review, then, is to summarize
and critically evaluate the work that has been conducted on AnMBRs and
then use this evaluation to answer these questions. In addition,the needs
for future research and development to more fully utilize AnMBRs will be
identified.

II. CONFIGURATIONS AND HISTORICAL DEVELOPMENT


A. Configurations
An AnMBR can be simply defined as a biological treatment process operated
without oxygen and using a membrane to provide complete solid-liquid sep-
aration. This definition is too broad for discussion, however, because a num-
ber of alternatives exist for both the anaerobic process and the membrane
process. The performance and operating characteristics of the AnMBR can
depend significantly on which alternatives are selected, so a brief description
of the key options is merited.
The defining characteristic of different anaerobic processes is whether
there is biomass retention. Commercial high-rate anaerobic reactors are fea-
sible because biomass is retained, either by the formation of granular sludge
or by attachment to a fixed or mobile support material, thereby decoupling
492 B.-Q. Liao et al.

the SRT from the HRT.59 The most common reactor designs that provide
biomass retention are the upflow anaerobic sludge bed (UASB), hybrid UASB,
anaerobic filter (AF), expanded granular sludge bed (EGSB), and fluidized
bed (FB). Descriptions of these technologies are provided by Kleerebezem
and Macarie59 and Speece.91 When biomass is retained, the effluent sus-
pended solids concentration is significantly lower than the biomass concen-
tration in the reaction zone. A UASB, for example, has a biomass concentra-
tion of 20–30 g/L59 but < 1 g/L of suspended solids in the effluent. Reactor
designs that do not provide biomass retention are the completely stirred tank
reactor (CSTR) and plug-flow reactor with suspended biomass. In this case,
the SRT is equal to the HRT and the effluent suspended solids concentration
(soluble substrate 4–6 g/L59 , particulate substrate 20–50 g/L68 ) is equal to
the bulk reaction zone solids concentration. This reactor design is used for
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

treatment of high-solids wastes, such as municipal wastewater sludges, and


variations include two-stage, temperature-phased, and acid-gas phased.4,28,68
In acid-gas phased anaerobic digestion, the fermentation/acidification
reactions are physically separated from the acetogenic/methanogenic
reactions.28
There are two principle approaches to membrane design and opera-
tion. The membrane may be operated under pressure or it may be operated
under vacuum (Figure 1). In the first approach, the membrane is separate
from the bioreactor and a pump is required to push bioreactor effluent into
the membrane unit and permeate through the membrane (Figure 1a). This
configuration is often called an external cross-flow membrane, although this
can occasionally lead to confusion, as noted later. The cross-flow velocity of
the liquid across the membrane surface serves as the principle mechanism
to disrupt cake formation on the membrane.
When the membrane is operated under a vacuum, instead of direct pres-
sure, the configuration is often called submerged or immersed because the
membrane is placed directly into the liquid. A pump or gravity is used to
pull the permeate through the membrane. Because the velocity of the liq-
uid across the membrane cannot be as readily controlled, cake formation
can be disrupted by vigorously bubbling gas across the membrane surface.
For aerobic MBRs, the air used also provides aeration, while for AnMBRs
biogas must be used. The vacuum-driven immersed membrane approach
may be used in two configurations. The membrane may be immersed di-
rectly into the bioreactor (Figure 1b) or immersed in a separate chamber
(Figure 1c). The latter configuration now looks like an external membrane,
and will likely require a pump to return retentate to the bioreactor. However,
unlike the external cross-flow membrane, the membrane here is operated
under a vacuum instead of under pressure. The external chamber configura-
tion (Figure 1c) is used for full-scale aerobic wastewater treatment plants be-
cause it provides for easier cleaning of the submerged membranes, because
Anaerobic Membrane Bioreactors 493
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

FIGURE 1. Schematic of anaerobic membrane bioreactor configurations. (a) Pressure-driven


external cross-flow membrane. (b) Vacuum-driven submerged membrane with the membrane
immersed directly into the reactor. (c) Vacuum-driven submerged membrane with the mem-
brane immersed in an external chamber.
494 B.-Q. Liao et al.

the chambers can be isolated instead of the membranes being physically


removed.20

B. Historical Development
The first test of the concept of using membrane filtration with anaerobic
treatment of wastewater appears to have been reported by Grethlein35 in
1978. The external cross-flow membrane treated septic tank effluent and
resulted in an increased biomass concentration, 85–95% biochemical oxygen
demand (BOD) reduction, 72% nitrate removal, and 24–85% orthophosphate
reduction.
The first commercially-available AnMBR was developed by Dorr-Oliver
in the early 1980s for high-strength whey processing wastewater treatment
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

and was known as the Membrane Anaerobic Reactor System (MARS).63,94


The MARS system was comprised of a completely mixed suspended growth
anaerobic reactor for biodegradation and an external cross-flow membrane
module for biomass separation. The MARS process was tested at pilotscale
but was not applied at full scale,93 possibly due to high membrane costs.
A significant research effort was Japan’s Aqua Renaissance ’90 project.
This 6-year research and development program was carried out starting
from 1985 to develop a variety of different configurations of AnMBRs
for industrial wastewater and sewage treatment.49,56,69,70 Various mem-
brane configurations, including polymeric and ceramic membranes in cap-
illary, hollow fiber, tubular, and plate and frame modules, were tested
for biomass retention in both suspended and attached growth anaerobic
bioreactors. Wastewaters containing sewage, fats/oils, wheat starch, pulp and
paper mill effluent, alcohol fermentation effluent, and night soil were suc-
cessfully treated (chemical oxygen demand [COD] removal generally >90%)
in pilot-scale AnMBRs.
Commencing in 1987, a system known as ADUF (anaerobic diges-
tion ultrafiltration) was developed in South Africa for industrial wastewa-
ter treatment.82 The ADUF completely mixed process configuration was the
same as for the MARS process, although the locally manufactured membrane
did not require a high-pressure support system.82 A number of pilot- and
full-scale ADUF systems are in operation,93 including the BIOREK process,
which is intended for manure treatment with nutrient and biogas recovery.74
The BIOREK process contains six unit operations: preseparation, the ADUF
process, ammonia stripping process, reverse osmosis, gas purification, and
power generation. Pilot- and full-scale testing results indicated that the COD
removal efficiency was over 90%.
Through the 1990s, AnMBR research activity increased with investi-
gations into different membrane materials,47,84,88 characterization of mem-
brane foulants,16,17 and development of strategies for membrane cleaning
Anaerobic Membrane Bioreactors 495

and fouling management.8,13,15,108 Also, immersed membranes,39 non-CSTR


reactor designs,6 and phylogenetic analytical techniques67 such as fluorescent
in-situ hybridization (FISH)106 started to be used.

III. APPLICATIONS OF ANAEROBIC MEMBRANE BIOREACTORS


A. Synthetic Wastewaters
Synthetic wastewaters are typically used to test new concepts such as the
AnMBR. The results of a number of studies are summarized in Table 1.
The substrates used included volatile fatty acids, starch, glucose, molasses,
peptone, yeast, and cellulose. Although the COD removal was generally
>95%, only a few studies had an OLR of >10 kg COD/m3 /d and only
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

two had a maximum OLR of 20 kg COD/m3 /d or higher.12,29 These gen-


erally low OLRs for such readily biodegradable substrates are surprising,
considering that OLRs for high-rate anaerobic reactors are typically in the
range of 10–20 kg COD/m3 /d.91,100 In addition, none of the reported
studies achieved high COD removals at HRTs <10 h. For example, Cadi
et al.12 observed a large drop in COD removal efficiency from 91% to 78%
when the HRT was decreased from 11 to 6 h at a constant OLR of 2 kg
COD/m3 /d.

B. Food Processing Wastewaters


Food processing wastewaters are characterized by high organic strengths
(1000–85,000 mg COD/L) with a wide range of suspended solids concen-
trations (50–17,000 mg/L) (Table 2). Food processing wastewaters are read-
ily biodegradable, so anaerobic treatment is well established; approximately
76% of all the anaerobic reactor installations worldwide are for the food and
related industries.59
Many AnMBR studies have assessed food-processing wastewaters
(Table 2). Both pilot and full-scale AnMBR systems have been used to fa-
cilitate the retention of biomass and improve effluent quality.11,14,56,63,94,107
AnMBRs have been used for the treatment of effluents from field crop pro-
cessing (sauerkraut, wheat, maize, soybean, palm oil), the dairy industry
(whey), and the beverage industry (winery, brewery, distillery). High COD
removal efficiencies (usually >90%) were achieved, but the organic load-
ing rates, generally in the range 2–15 kg COD/m3 /d, were low in com-
parison to existing high-rate anaerobic systems, which can achieve OLRs
of 5–40 kg/m3 /d.59 Most studies used completely mixed reactors with ex-
ternal cross-flow membranes, though several studies investigated the use
of two-phase systems with and without a membrane after the acidification
reactor.56,107,109 This compares with traditional high-rate anaerobic treatment
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

496
TABLE 1. Summary of AnMBR Performance for Treatment of Synthetic Wastewaters

Type Type Reactor OLR Feed Effluent COD


of of volume Temp. HRT SRT (kg COD MLSS COD COD removal
wastewater Scalea reactorb (m3 ) (◦ C) (d) (d) m−3 d−1 ) (gL−1 ) (g L−1 ) (gL−1 ) efficiency Reference

Acetate L CSTR 0.007 35 1.0 30 8.5 10 8.5 <0.4 >95% 8


Acetate L CSTR 0.01 35 0.4 —c 5d 0.13 2d 0.1d 95%d 23
Acetate, lactate, P CSTR 0.24 33 3.9 100 17 12 67 0.7 99% 67
propionate, butyrate
Glucose L 2 phase 0.003/0.01 35 1.5/7.7 —/— 36/12 —/— 53/41 1.5 97% 2
CSTR+M/
CSTR+M
Glucose, peptone L CSTR — — 0.5 — 2d 3.3 — — 95%d 88
Glucose, peptone, yeast L CSTR 0.0045 55 6.0 — 4.0 — — — — 84
extract
Glucose, peptone, yeast L CSTR 0.007 30 0.5 — 20 22 9.7 <1 >90% 29
extract, acetate
Starch L CSTR 0.0075 35 0.5 45 2.0 — 0.93 0.09 90% 12
Starch L CSTR 0.0065 35 0.4 52 21 — 10.5 1.3 88% 12
Molasses L UASB 0.005 20 — — 0.3–13 — — — — 39
Synthetice L UASB 0.009 30 0.6 — 8.3 — 5 0.05 99% 6
Skim milk and cellulose L CSTR 0.01 35 2.0 — 2.5 15 5 <0.08 >98% 38
Cellulose L Various — — 4.6–7.2 — 2.7–6.5 f 10–40 22–47 f 5–25 f 40–90% f 30
CSTRs/PB
Cellulose L CSTR 0.04 36–38 3 — 1.24–2.0 f 5–15 — — 95–98% f 52
aL = laboratory/bench scale, P = pilot scale.
b CSTR = completely stirred tank reactor, PB = packed bed, UASB = upflow anaerobic sludge blanket, M designates the location of the membrane (no M indicates
the membrane produced the final effluent).
c —Indicates value not reported.
d units are TOC instead of COD.
e Composition not reported.
f Units are cellulose instead of COD.
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

TABLE 2. Summary of AnMBR Performance for Treatment of Food Processing Wastewaters

Type Type Reactor OLR Feed Feed Effluent COD


of of volume Temp. HRT SRT (kg COD MLSS COD TSS COD removal
wastewater Scalea reactorb (m3 ) (◦ C) (d) (d) m−3 d−1 ) (gL−1 ) (g L−1 ) (g L−1 ) (gL−1 ) efficiency Reference

Sweet whey P CSTR 0.19 35 7.4 50 8.5 32 62 2 1.6 97% 94, 63


permeate
Sweet whey P CSTR 0.19 35 1.9 25 14.6 35 29 2 1.4 95% 94, 63
permeate
Sweet whey P CSTR 0.19 35 7.1 25 8.0 27 59 5.2 0.7 99% 63
Acid whey permeate P CSTR 0.3 35 5.7 27 9.6 37 55 0.1 0.5 99% 63
Sauerkraut brine L CSTR 0.007 30 6.1 —c 8.6 55 52.7 0.5 0.5 99% 29
Wheat starch F CSTR 2 000 40 — — 2.1 10 — — — 78% 11
Wheat starch F CSTR 900 36 4.5 6 3.8 — 17 0.03 2.3 86% 14
Wheat starch P CSTR 0.3 35 4.4 30 8.2 24 35 13.3 0.3 99% 63
Wheat starch P 2 phase — — — — — — 36 17 8.8 76% 56
UFAF+M/
UASB+M
Wheat starch and P 2 phase 24/15 37 0.6/0.4 —/— 32/27 18/— 19/10 3.5/0 0.3 98% 107
gluten UFAF+M/
UASB
Maize P CSTR 3 35 1.6 — 5.0 15 8 1 0.8 90% 83
Maize F CSTR 2 610 35 5.2 — 2.9 21 15 — 0.4 97% 83
Soybean P UFAF 3.0 30 0.4 — 3.2 2 1.4 0.7 0.3 78% 49
Soybean P 2 phase — — — — — — 10 4.3 0.9 91% 56
CSTR+M/
FB+M
Soybean P 2 phase 1.0/2.0 30 3.3/6.7 — 3.3 — 1.4/1.0 0.7/0.6 0.3 78% 109
UFAF/
UFAF+M
Soybean P 2 phase 1.0/2.0 30 3.5/7.0 — 3.0 — 1.3/0.9 0.5/0 0.1 92% 109
UFAF+M/
UFAF
Palm oil mill L CSTR 0.05 35 3.2 77 21.7 57 68 — 5.4 92% 26
(Continued on next page)

497
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

498
TABLE 2. Summary of AnMBR Performance for Treatment of Food Processing Wastewaters (Continued)

Type Type Reactor OLR Feed Feed Effluent COD


of of volume Temp. HRT SRT (kg COD MLSS COD TSS COD removal
wastewater Scalea reactorb (m3 ) (◦ C) (d) (d) m−3 d−1 ) (gL−1 ) (g L−1 ) (g L−1 ) (gL−1 ) efficiency Reference

Sweets factory L UFAF 0.09 35 ∼3 — 8–9 — 25 — 0.9 96% 22


(carbohydrates)
Alcohol fermentation P FB 8.0 — — — — — 18 10 11 39% 56
Alcohol distillery P Upflow 5.5 37 5.7 — 7 20 40 13 0.6 98% 72
Alcohol distillery L CSTR 0.004 54 10 — 2.1 0.3 23 0.4 0.7 97% 16
Alcohol distillery L CSTR 0.004 54 23 — 1.5 2.1 35 1.2 2.2 94% 18
Alcohol distillery L CSTR 0.004 54 13 — 3.3 2.0 38 0 3.8 90% 47
Wine distillery P CSTR 2.4 35 3.3 — 11 50 37 — 2.6 93% 82
Brewery P CSTR 0.12 35 4.0 59 19.7 38 84 — 3.1 96% 25
Brewery L CSTR 0.05 35 0.5 — 15 30 6.7 — 0.2 97% 93
Brewery L CSTR 0.12 36 3 — 28.5 — 85 0.15 0.9 99% 43
Brewery L CSTR 0.12 36 30 — 2.8 10 85 0.15 1.7 98% 44
Brewery L CSTR 0.12 36 34 — 2.5 15 85 0.15 1.7 98% 45
aL = laboratory/bench scale, P = pilot scale, F = full scale.
b CSTR = completely stirred tank reactor, FB = fluidized bed, UASB = upflow anaerobic sludge blanket, UFAF = upflow anaerobic filter, M designates the location
of the membrane (no M indicates the membrane produced the final effluent).
c —Indicates value not reported.
Anaerobic Membrane Bioreactors 499

of food-processing wastes, which predominantly uses the UASB reactor


configuration.59

C. Industrial Wastewaters
Non-food-processing industrial wastewaters include effluents from the pulp
and paper, chemical, pharmaceutical, petroleum, and textile industries. The
characteristics of industrial wastewaters are sector specific although, in gen-
eral, they have the potential to have a high organic strength and contain
synthetic and natural chemicals that may be slowly degradable or non-
biodegradable anaerobically, and/or toxic. Traditionally, industrial wastew-
aters are treated by a combination of physical, chemical and biological
processes because no single method can achieve complete treatment. An
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

important concern associated with the biological treatment of such wastewa-


ters is toxicity to the microorganisms. However, wastewaters containing toxic
compounds can still be anaerobically degraded provided that appropriate
precautions are taken.91 This can include pretreatment to remove the in-
hibitors prior to anaerobic treatment,69 acclimation of the biomass by gradual
increase of inhibitor concentration, and provision of a sufficiently high SRT
(safety factor).91 Membrane bioreactors may have an advantage over other
anaerobic systems because the biomass can be retained even if an inhibitor
upsets the treatment system. Because toxics rarely cause cell death, treatment
would only be temporarily impaired.91
To date, AnMBRs have only been applied to pulp and paper and textile
wastewaters (Table 3). The source and composition of the different pulp and
paper effluents and whether each is amenable to anaerobic treatment were
reviewed by Rintala and Puhakka.81 Anaerobic treatment of pulp and paper
wastewaters has become more common; approximately 9% of all anaerobic
installations are for the pulp and paper industry.59 The use of AnMBRs for
pulp and paper wastewaters has been reported five times, as summarized
in Table 3. Evaporator condensates (EC) from pulp and paper plants are
characterized by high soluble CODs of 10–42 g/L, due mainly to methanol,
low suspended solids (<3 mg/L), plus inhibitory turpene oils and sulfur
compounds.69,70 Pretreatment of the condensate by microfiltration and bio-
gas stripping was used to remove the inhibitory turpene oils and sulfur com-
pounds and the pH was adjusted to neutral. The pretreated condensate was
then amenable to treatment in a thermophilic attached-growth ultrafiltration
AnMBR that provided a biochemical oxygen demand (BOD) removal effi-
ciency of >93%.69 The ultrafiltration membrane used for cell retention and
recycling in the AnMBR allowed the OLR to be more than doubled to 35.5 kg
BOD/m3 /day, from 15 kg BOD/m3 /d without the membrane. The use of an
AnMBR to treat segregated kraft bleach plant wastewater provided a mod-
est increase in the adsorbable organic halogen removal efficiency, from 48%
to 61% in comparison to a UASB without membrane.36 Economic analyses
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

500
TABLE 3. Summary of AnMBR Performance for Treatment of Industrial Wastewater

Type Type Reactor OLR Feed Feed Effluent COD


of of volume Temp. HRT (kg COD MLSS COD TSS COD removal
wastewater Scalea reactorb (m3 ) (◦ C) (d) m−3 d−1 ) (gL−1 ) (g L−1 ) (g L−1 ) (gL−1 ) efficiency Reference

Kraft bleach plant L CSTR 0.015 35 1.0 0.04e 7.6–15.7 ∼0.04e — ∼0.016e 61%e 36
effluent
Kraft pulp effluent P UFAF 5 — 0.5 35d 9.4 19.2d — 1.5d 93%d 76
Pulp and paper effluent P FB 7 —c — — — 28 15 1.1 96% 56
Evaporator condensate P UFAF 5 53 0.5 35.5d 7.6 17.8d <0.003 1.2d 93%d 69,70
(methanol)
Wool scouring P UFAF 4.5 37 6.8 15 — 102.4 30.5 51 50% 40
aL = laboratory/bench scale, P = pilot scale.
b CSTR = completely stirred tank reactor, FB = fluidized bed, UFAF = upflow anaerobic filter.
c —Indicates value not reported.
d Units are BOD instead of COD.
e Units are AOX (adsorbable organic halogen).
Anaerobic Membrane Bioreactors 501

indicated that the total cost of AnMBR treatment of kraft mill effluent was
significantly lower than for aerobic treatment and only slightly higher than
for high-rate anaerobic treatment (although the AnMBR had higher effluent
quality).69,70,76
The sources and characteristics of wastewaters generated by textile pro-
cessing are discussed in a recent review by Bisschops and Spanjers.9 Only
1% of commercial anaerobic installations are for the textile industry.59 An
AnMBR treating wool-scouring wastewater achieved a 50% COD removal at
an OLR of 15 kg/m3 /d.40 The addition of membrane filtration approximately
doubled the biomass concentration and increased the total organic carbon
(TOC) and grease removal efficiencies from 45 to 90% and from 33 to 99%,
respectively.
Membranes without biological treatment have been used in the tex-
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

tile industry to allow reuse of electrolyte solutions60,96 and chemicals,97


and anaerobic treatment without membranes has been applied successfully
to both petrochemical86 and textile wastewaters. The reader is referred to
recent reviews that discuss the anaerobic degradation of petroleum com-
pounds, including the BTEX compounds (benzene, toluene, ethylbenzene,
and xylene),65 hydrocarbons,111 aromatic and aliphatic compounds,59 and
halogenated compounds.27 Reviews are also available for the biological
treatment of textile mill effluents97 and the decolorization of textile dying
effluents.37,78 As long as a wastewater is amenable to anaerobic treatment,
in theory an AnMBR could be used to treat it.

D. High-Solids-Content Waste Streams


Waste streams that contain a high proportion of particulates include wastew-
ater treatment plant sludges, the organic fraction of municipal solid waste,
animal processing plant effluents, and manures. Anaerobic digestion is a com-
mon technology for treating such high-solids waste streams, as discussed in
recent reviews and books.48,66,85,101 Digestion is usually performed in com-
pletely mixed reactors at low organic loadings of 1–3 kg COD/m3 /d.101 The
slow hydrolysis/solubilization of particulates is often rate limiting.101 There-
fore, in completely mixed reactors that do not decouple SRT from hydraulic
retention time (HRT), the long retention time required for hydrolysis leads
to large reactor volumes and lower OLRs.
In recent years, the AnMBR technology has been successfully tested
in both pilot- and full-scale plants for treatment of high solids wastes, as
summarized in Table 4. AnMBRs have been tested with wastewater treatment
plant sludges, pig manure, and chicken slaughterhouse effluent. Relatively
high OLRs of 3–5 kg COD/m3 /d were achieved with high COD removals
(80% or higher) for the manure and slaughterhouse wastewaters as compared
with the usual loadings of 1–3 kg COD/m3 /d for high-solids wastes. The
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

502
TABLE 4. Summary of AnMBR Performance for Treatment of High Solids Content Wastes

Type Type Reactor OLR Feed Feed Effluent COD


of of volume Temp. HRT SRT (kg COD MLSS COD TS COD removal
wastewater Scalea reactorb (m3 ) (◦ C) (d) (d) m−3 d−1 ) (gL−1 ) (g L−1 ) (g L−1 ) (gL−1 ) efficiency Reference

Primary sludge P Upflow mixed 0.12 35 20 —c 1.06 22–35 40.2 44.4 18 54% 33
Primary sludge P CSTR 0.5 35 8.4 335 0.93d — 0.24 0.16 0.03d 79%d 71
Primary sludge P CSTR 0.5 55 7.8 197 1.16d — 0.24 0.16 0.03d 78%d 71
Coagulated raw L VFA fermenter 0.076 35 0.5 10 4.6e 34 2.3e 6.8 1.3e 43%e 53
sludge CSTR
Screened sludge P Semi 1.8 — 14 26 — 55 — — — — 80
continuous
CSTR
Sewage sludge L CSTR 0.004 25–50 6.7–20 — 0.17–1.35d 20–40 — — <0.3 — 3
Pig manure P CSTR 0.1 35 6 — 5 — 30 20 3 90% 74
Pig manure F CSTR 200 35 10 — 3 — 30 20 2.4 92% 74
Pig manure P 2 phase 3/3 20/35 1–2/1–2 —/— 2.8–5.5/— —/— 5.5 0.6 1.1 80% 61
CSTR+M/
Hybrid
Chicken L CSTR 0.007 30 1.2 — 4.3 22 5.2 2.4–4.7 <0.5 90% 29
slaughterhouse
aL = laboratory/bench scale, P = pilot scale, F = full scale.
b CSTR = completely-stirred tank reactor, Hybrid = UASB with anaerobic filter instead of a solids/liquid/gas separator, M designates the location of the membrane
(no M indicates the membrane produced the final effluent).
c —Indicates value not reported.
d Units are VSS instead of COD.
e Units are TOC instead of COD.
Anaerobic Membrane Bioreactors 503

COD removals were generally lower for sludge treatment possibly because
of a larger portion of inert solids. Nevertheless, Aya and Namiki3 found that
organic substances in the sludge were almost all converted into biogas, and
inorganic solids were dissolved into the liquid phase by biological reaction.
All the studies considered here used a CSTR reactor configuration. With-
out a membrane, the SRT would have been equal to the HRT. In membrane
sludge digesters, however, the complete retention of solids in the reactor
decoupled the SRT from the HRT.79,80 Pillay et al.80 were able to increase
the reactor solids concentration from 2.6% to 5.5% using a membrane and
as a result decreased the HRT by almost half (to 14 d) while the SRT was
maintained at 26 d. Pierkiel and Lanting79 used HRTs of 1–3 d with SRTs of
8–12 d.
Another proposed advantage of complete retention of particulates and
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

biomass is an increase in the rate of hydrolysis/solubilization of solids. This


has been tested several times,29,53,61 although none of the studies evaluated
the hydrolysis/solubilization without a membrane. Therefore, the advantage
observed may be due to either increased SRT or increased solids concen-
trations and not necessarily an inherent increase in hydrolysis/solubilization
rate.
Based on the results from pilot-scale trials, Pillay et al.80 conducted an
economic evaluation of a full-scale anaerobic digester for sludge treatment
with and without membrane filtration. The results indicated that the AnMBR
process was technically and economically feasible, and offered significant
advantages over the conventional anaerobic digestion process. Compared
to the conventional anaerobic digester, the capital and total project cost
savings for an AnMBR using their low-cost membrane were 27% and 12%,
respectively.

E. Municipal Wastewater
Municipal wastewater is characterized by low organic strength (250–800 mg
COD/L) and low suspended solids concentrations (120–400 mg/L).68 The
aerobic activated sludge process is the dominant technology for treating mu-
nicipal wastewater and in recent years the aerobic membrane bioreactor has
been widely used.20,102 Anaerobic treatment of sewage is not widespread, tra-
ditionally being performed in UASB reactors in warm climate regions,59,98,101
but it is technically feasible even for temperate climates as discussed in recent
reviews.46,87,110 Conventional UASB sewage treatment usually has an HRT of
0.25–0.33 d and results in a BOD removal efficiency of 80%, effluent COD of
100–220 mg/L, and effluent total suspended solids (TSS) of 30–70 mg/L.101
In general, sewage has a larger portion of refractory COD compared to food
and beverage wastewaters. For example, Elmitwalli et al.24 found the anaer-
obic biodegradability of domestic sewage to be 71–74% at 30◦ C whereas
504 B.-Q. Liao et al.

most studies for food processing wastewaters (Table 2) had COD removals
of >90%.
Table 5 summarizes the studies on the use of AnMBRs for sewage
treatment. In general, AnMBR sewage treatment had lower effluent COD
(<100 mg/L) and suspended solids concentrations compared to conventional
UASB treatment. This is expected because the membrane can provide ap-
proximately 100% removal of suspended solids.5,95 In addition, the COD or
BOD removal efficiency was comparable to UASB treatment and very high
SRTs could be maintained (e.g., 150 days).105 AnMBRs also provided high
COD removals for the treatment of night soil56 and sludge heat treat liquor.50
A comparison between conventional activated sludge, aerobic MBR,
UASB, and AnMBR for municipal wastewater treatment is shown in Table 6.
For both aerobic and anaerobic systems, a membrane dramatically improves
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

TSS removal, although the treated effluent quality in terms of COD is better
from the aerobic systems than the anaerobic systems. Because of this fact,
aerobic posttreatment of anaerobic effluent has been suggested as a method
to further improve COD and nutrient removals.101 The anaerobic HRT ap-
pears to be generally longer than 8 h, compared to 4–8 h for aerobic.
On the other hand, the anaerobic processes had lower energy require-
ments than their aerobic counterparts. The electricity use of the UASB is the
lowest, well below that of activated sludge. The electricity use of the AnMBR
is slightly lower but nevertheless near that of the aerobic MBR when both
systems used immersed membranes. However, the electricity use of both
anaerobic systems can be offset by use of the produced methane, so the
net energy consumption by both anaerobic systems should be less than the
aerobic ones. In all cases, electricity consumption will be higher if external
cross-flow membranes are used instead of immersed membranes, because a
suction pump operates at lower pressure and less water is pumped.1

IV. FACTORS AFFECTING THE TREATMENT PERFORMANCE


OF AnMBRs
A. Microbial Activity
Pressure-driven external cross-flow membranes use high liquid velocities to
minimize fouling and maintain high fluxes. This requires large volumes of
liquid to be pumped; on the order of 25–80 m3 of liquid must pass over the
membrane for each 1 m3 of permeate in anaerobic systems.10,20 Liquid recir-
culation through the membrane pump has resulted in a substantial decrease
in the observed mean particle size by a factor of 3–5.6,13,18,23 Circulation of
the biomass through the membrane pump has in some instances resulted
in a decrease in microbial activity7,10,32,33 but not in others.8,16,45 This dis-
agreement may be due to the use of different types of pumps. Kim et al.,55
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

TABLE 5. Summary of AnMBR Performance for Treatment of Municipal Wastewater


Type Type Reactor OLR Feed Feed Effluent COD
of of volume Temp. HRT SRT (kg COD MLSS COD TSS COD removal
a −3 −1 −1 −1 −1
wastewater Scale reactorb (m3 ) (◦ C) (d) (d) m d ) (gL ) (g L−1 ) (g L ) (gL ) efficiency Reference

Night soil (heat-treated P UASB 0.4 —c — — — — 25.5 2.6 2.0 92% 56


and hydrolyzed)
Heat-treat liquor L CSTR 0.2 37 0.6 — 15.4 21.4 10.3 0.3 2.0 81% 50
Primary effluent L CSTR 0.01 32 0.5 217 1.6 7 0.08 0.12 0.02 68% 5
Sewage L Septic tank 0.106 — 5.6–9.6 — 0.03–0.05d — 0.27d — <0.04d >85%d 35
d
Sewage P CSTR 3 10–22 0.4-0.8 — 0.2d — 0.1–0.2 0.03–0.1 <0.06d 56%d 95
Sewage P Hydrol —/5.4 — — — — — 1.1 0.5 0.07 94% 56
CSTR+M/
UASB
Sewage P Hydrol CSTR/ 2.0/5.4 35/— 3/0.27 — 5.7 7/40 0.49 0.3 0.08 83% 49
UASB+M
Sewage P Hydrol 0.5/1.0 30/— 5/0.09 — 1.8 30/9 0.35 0.3 0.04 90% 49
CSTR+M/
FB+M
Sewage P Hydrol 8.9/77 25/28 —/0.3 — —/0.97 — 0.4 ∼0.25 0.1 73% 57
CSTR+M/
UASB
Sewage P Hydrol 8.9/77 25/12 —/0.3 — —/0.65 — 0.3 ∼0.25 0.1 58% 57
CSTR+M/
UASB
Sewage P Hydrol 8.9/77 25/18 3.7/0.3 25/— 2e /0.4d —/10 0.17d 0.2 0.07d 61%d 58
CSTR+M/
UASB
Domestic wastewater L Hybrid 0.018 20 0.25 150 0.4–10 16 0.1–2.6 0.1–0.8 <0.03 >92% 105
Domestic wastewater L Hybrid 0.018 20 0.17 150 0.7–10 22 0.1–1.8 0.1–1.0 <0.03 >92% 105
aL = laboratory/bench scale, P = pilot scale.
b CSTR = completely-stirred tank reactor, FB = fluidized bed, Hybrid = UASB with anaerobic filter instead of a solids/liquid/gas separator, Hydrol = side-stream
suspended solids hydrolysis reactor plus methanogenic reactor for combined hydrolysate and primary clarifier effluent, UASB = upflow anaerobic sludge blanket,
M designates the location of the membrane (no M indicates the membrane produced the final effluent).
c —Indicates value not reported.
d Units are BOD instead of COD.

505
e Units are VSS instead of COD.
506 B.-Q. Liao et al.

TABLE 6. Performance Comparison Between Aerobic and Anaerobic Technologies for


Municipal Wastewater Treatment at 15◦ C

Activated Aerobic Anaerobic


Parameter Units sludgea MBRb UASB MBRc

Effluent COD mg L−1 <30d <30d 100–220e <100


Effluent TSS mg L−1 <30 <1 30–70e <1
OLR kg COD 0.3–0.7d 1.2–3.2d 0.6–3.0 f 0.4–6.0
m−3 d−1
HRT hr 4–8 4–6 8–14a, f 8–12
SRT d 3–15 5–20 30–100g >100
MLSS in g L−1 1–3 5–20 15–30h 10–40
bioreactor
Electricity usei kWh m−3 0.2–0.4 0.3–0.6 j 0.11h 0.25–1.0b, j
Issues — Possible Membrane Requires Membrane
sludge fouling VSS/CODsoluble fouling
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

bulking < 0.1e


a Reference 68.
b References 92 and 103.
c Values chosen as representative from Table 5.
d Units are BOD instead of COD.
e Reference 101.
f Reference 62.
g Reference 34.
h Reference 110.
i Electricity use by bioreactor and membrane/secondary clarifier only.
j Membrane energy use based on immersed modules.

for example, found that a vane-type rotary (positive displacement) pump


imposed greater shear stress to activated sludge flocs than did a turbine-
type centrifugal (kinetic) pump. If the same was true for anaerobic flocs
or granules, interspecies hydrogen transfer could be disrupted. Hydrogen-
producing bacteria must be in very close proximity to the hydrogen-
consuming methanogens in order to keep the hydrogen partial pressure low,
thereby allowing otherwise thermodynamically unfavorable reactions to pro-
ceed (e.g. acetogenesis of propionic acid, β-oxidation of fatty acids, etc.).28
Because most AnMBR research conducted to date has used external
cross-flow membranes, it is possible that, on average, AnMBRs have had
a lower microbial activity compared to nonmembrane high-rate anaerobic
systems. Furthermore, because the liquid velocity in the cross-flow mem-
brane is independent of the reactor size, the frequency of liquid recirculation
through the pump increases for smaller reactors, and is more of an issue for
laboratory-scale systems.10 The use of external cross-flow filtration could be
one reason why the maximum attained OLRs have been lower than for ex-
isting high-rate anaerobic reactors, at a least at laboratory scale. On the other
hand, in principle this pump-induced shear stress may increase the break-
down of complex particulates. This could account for the improved methane
conversion observed by Yushina and Hasegawa109 when a membrane was
placed after an acid-phase reactor treating particulate soybean wastewater.
Anaerobic Membrane Bioreactors 507

Reports of vacuum-driven immersed membranes in AnMBR studies have


been limited. 39,50,103,105 Although the effluent TSS was higher from immersed
membrane systems,39,103 the microbial activity should remain higher than
is possible with external cross-flow membranes because the biomass does
not pass through a pump.39 Hernandez et al.39 observed a homogenization
of the size of granules and an increase in activity for a UASB with an im-
mersed membrane, although there was granule breakup at an OLR >13 kg
COD/m3 /d, possibly due to gas mixing.

B. Operational Temperature, SRT, and HRT


Anaerobic processes are often operated at mesophilic (35◦ C) and ther-
mophilic (55◦ C) temperatures. These temperatures are particularly important
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

for the treatment of high-solids-content material such as municipal wastew-


ater sludges, where the SRT and HRT are equal and the increased reaction
rates at the higher temperatures lead to decreased reactor sizes. For exam-
ple, at a loading rate of 2 kg VSS/m3 /d, Murata et al.71 observed a volatile
suspended solids (VSS) reduction of 67% at 55◦ C versus 56% at 35◦ C. Increas-
ing the temperature also increases the attainable membrane flux because the
liquid viscosity decreases at higher temperatures, as discussed later.
Often, the heating requirement for treating these streams can be met
by the produced methane. However, for wastewaters with a low organic
content (e.g., municipal wastewater) the methane production cannot cover
the heating requirement and operation must be at ambient temperatures.
Thus, anaerobic sewage treatment has traditionally only been conducted in
warm climates.87 Several recent reviews have summarized the challenges
for low-temperature operation of anaerobic processes in cool and temper-
ate climates.48,62,87 Operation at ambient temperatures appears technically
feasible, although SRTs as much as double those for mesophilic operation
may be required,48 and hydrolysis of solids slows significantly.87 Operation
becomes governed by the hydraulic loading rate and not the organic loading
rate because the loss of viable biomass must be minimized.62 Membranes
may alleviate many of these challenges because of their high solids reten-
tion capability. An AnMBR can maintain solids retention across a range of
temperatures105 which can allow an increase in the SRT and a decrease in the
HRT62 because biomass is retained. For example, Wen et al.105 maintained
an SRT of 150 d with an HRT of 4–6 h at 20◦ C treating screened munici-
pal wastewater, and Baek and Pagilla5 achieved SRTs >200 d with HRTs of
12–24 h at 32◦ C treating primary clarifier effluent. However, Merkel et al.67
used fluorescent in situ hybridization (FISH) to observe that tripling the re-
actor VSS did not triple the reaction rate because the active biomass did not
increase in the same proportion as the VSS.
The HRT of a reactor significantly influences the capital cost. For a given
influent composition, a higher OLR allows a shorter HRT and a smaller
508 B.-Q. Liao et al.

reactor. For example, Ross et al.82 found the OLR of a completely mixed
reactor could be increased from 4 to 12 kg COD/m3 /d when a membrane
was used. In general, the HRTs used with AnMBRs have been higher than
for nonmembrane high-rate anaerobic reactors. AnMBR HRTs as low as 10 h
have been used for soybean-processing wastewater and sewage, but high-
rate anaerobic reactors typically have HRTs of 4–8 h. The complete solids
retention possible in AnMBRs has not yet led to the expected decrease in
HRTs.

C. Reactor Design and Membrane Location


By far the most common anaerobic membrane bioreactor design has been
the CSTR. From Tables 1–5, approximately 67% of the membrane reactors
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

have been CSTRs, 15% anaerobic filters, 10% UASB or UASB hybrid, 7%
fluidized bed, and 2% septic tank. In part, this may be due to the ease of
use and construction of a CSTR reactor, but it is also likely due to the almost
exclusive use of external cross-flow membranes. The high liquid turnover
rate required for external cross-flow membranes10,20 will create a well-mixed
flow regime unless the membrane intake and return are specifically located
so that well-mixed conditions are not created. For example, some studies
have been able to use UASB reactors while still utilizing external cross-flow
membranes.6,13,39
Although Fuchs et al.29 observed a shorter startup period with a CSTR
in comparison to those typical of high-rate reactor configurations, the use
of completely-mixed reactors is less attractive than a high-rate configuration
for three reasons. First, single-stage reactor configurations have lower COD
removal efficiency than UASB and multistage CSTR reactor configurations
regardless of substrate complexity.4 Second, the CSTR configuration exposes
the membrane to the reactor bulk mixed liquor suspended solids (MLSS)
whereas reactor configurations designed to retain biomass (e.g. UASB) ex-
pose the membrane only to the residual effluent TSS. For example, the efflu-
ent from a UASB AnMBR was 300–550 mg TSS/L,13 while the MLSS of a CSTR
AnMBR reactor is commonly >10,000 mg/L (see Tables 1–5). Exposure to
higher solids concentrations usually leads to lower fluxes, as discussed later.
Finally, membrane use can reduce the capital cost of high-rate reactor de-
signs, for example, by eliminating the need for a solids/liquid/gas separator
in a UASB.6
Two-phase membrane bioreactors have also been examined. In this con-
figuration, the membrane may be placed after the second-phase methane
reactor,49,109 after the first-phase acid reactor,19,30,56−58,61,107,109 or after both
reactors.2,49,56 The location of the membrane unit within a two-phase sys-
tem can have a significant impact on performance. For a two-phase system
treating soybean wastewater, Yushina and Hasegawa109 observed a COD
removal efficiency of 52% when no membranes were used, compared to
Anaerobic Membrane Bioreactors 509

78% when a membrane was used after the methane-phase reactor and 92%
when a membrane was used after the acid-phase reactor. The effluent sus-
pended solids decreased from 361 mg/L with no membranes to 0 mg/L
with a membrane after the methane phase and 4 mg/L with a membrane
after the acid-phase. The authors supported placing the membrane after
the acid-phase reactor for wastewaters containing organic suspended solids.
Membrane-coupled hydrolysis/acidification reactors had a 98% solids reten-
tion and prevented biomass washout19 and increased biomass concentration
and COD removal.30 Membrane-coupled reactors for hydrolysis of suspended
solids removed from raw municipal wastewater had relatively long HRTs of
3–5 d49,56−58 and with the membrane the SRT would have been even longer
(it was not reported in the studies). Such a design retains substances until
they are converted into smaller products and pass through the membrane or
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

are removed with wasted sludge.


The classic acid reactor is a CSTR with a short SRT (= HRT) operated
at low pH. Not all substrates are amenable to degradation in this classic
design. For example, the degradation of long-chain fatty acids, aromatics,
and some proteins is not thermodynamically favorable in normal acid-phase
reactors because the syntrophic relationships needed to consume reducing
equivalents have been eliminated.28 A membrane-coupled acid reactor, on
the other hand, can decouple the SRT from the HRT to allow the growth of
hydrogenotrophic methanogens at a short HRT and low pH. In this case, be-
cause hydrogen consumption could occur to consume reducing equivalents,
compounds may be acidified that otherwise would not.

V. MEMBRANE PERFORMANCE OF AnMBRs


A. Flux
The performance of a membrane is synonymous with flux. Membrane flux is
one of the most important parameters that determine the economy of mem-
brane bioreactors. A larger membrane flux allows for a smaller membrane
surface area for a given hydraulic treatment capacity. However, there exists a
critical flux for membrane filtration. Ideally, operation of the membrane be-
low the critical flux allows a constant transmembrane pressure (TMP) with-
out fouling, while operation above the critical flux causes a rapid increase in
TMP.92 Nevertheless, perfect non-fouling operation cannot be expected, and
operation below the nominal critical flux has been observed to cause a slow
linear increase in TMP.13 The critical flux is a function of sludge characteris-
tics and concentration and can be determined by three methods.13 The most
convenient is to measure TMP against a step increase in flux. For example,
by this method Cho and Fane13 found the critical flux of an AnMBR to be 30–
50 L/m2 /h using a cross-flow velocity of 0.93 m/s (Re = 5600). In addition,
the critical flux will decrease over time as the membrane becomes fouled.
510 B.-Q. Liao et al.

The characteristics and performance of the membranes used in AnMBRs


are summarized in Tables 7 and 8. The reported membrane flux in the lit-
erature was from 4 to 250 L/m2 /h for external cross-flow membrane mod-
ules, and from 3 to 80 L/m2 /h for submerged membrane modules. The flux
for external cross-flow membranes is higher than immersed membranes in
aerobic systems as well.20 The typical design membrane flux for AnMBRs
is 10–40 L/m2 /h at a temperature of 20–50◦ C.14,40,74,83,103 The design mem-
brane flux is case specific, depending not only on membrane properties
(composition, pore size, porosity, hydrophobicity, and surface charge) but
also on anaerobic sludge properties, and operational and environmental
conditions.

B. Membrane Pore Size and Materials


Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

The membrane pore size or molecular weight cutoff has a significant effect on
the membrane flux. Larger pore size or higher molecular weight cutoff usually
leads to increased flux. For example, Hernandez et al.39 found that the steady-
state membrane flux was about 7 times higher for a pore size of 100 µm than
for 10 µm using immersed membranes with a granular sludge. However,
Imasaka et al.41 observed a slight decrease in flux with an increase in pore
size from 0.2 µm to 0.57 µm when using external cross-flow membranes
to filter a concentrated broth. This decrease occurred because the permeate
resistance due to plugging increased more than the decrease in permeate
resistance of the membrane from a larger pore size, thereby causing an overall
decrease in flux.41
Microfiltration and ultrafiltration membranes are the most common for
MBRs. Microfiltration membranes generally have a pore size >0.05 µm, while
ultrafiltration membranes have a pore size between 0.002 and 0.05 µm.68
Both classifications of membranes retain particulates, but ultrafiltration will
retain more macromolecules and colloids. To minimize energy use and max-
imize flux, the membrane with the largest pore size that will achieve the
required separation should be used.
The membrane material also plays an important role in determining per-
formance. Ghyoot and Verstraete33 reported that the flux of a ceramic micro-
filtration membrane reached 200–250 L/m2 /h, which was 10-fold higher than
the flux achieved with polymer ultrafiltration, with both membranes produc-
ing permeate of similar quality. Decreasing the hydrophobicity of polypropy-
lene membranes by graft polymerization with hydroxyethyl methacrylate
increased the long-term achievable flux.15,84,104 An interesting membrane
design used by Pillay et al.80 was a woven-fiber cross-flow microfiltration
membrane that intentionally required the deposition of a fouling layer to
improve performance and filter particles smaller than the large pore size.
Shimizu et al.88 found that negatively charged membranes had a higher flux
than noncharged and positively charged membranes.
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

TABLE 7. Membrane Performance of AnMBRs for Synthetic, Industrial, Food-Processing, and High Solids Content Wastewater Treatmenta

Trans Membrane Initial Final


Type Membrane membrane linear membrane membrane
of Temp. Pore area pressure velocity flux flux
wastewater Scaleb (◦ C) Membrane material sizec (m2 ) (kPa) (m s−1 ) (L m−2 h−1 ) (L m−2 h−1 ) Reference

Wheat starch P 40 — 18,000 D 144 690 —d — 14–25 11, 14


Brewery effluent P 35 Poly-ethersulfone 40,000 D 0.44 140–340 1.5–2.6 — 7–50 93
Maize processing P — poly-ethersulfone 20,000–80,000 D 668 450 1.6 — 8–37 83
Wool scouring P 40–47 poly-acrylonitrile 13,000 D 3.1 2–2.2e — 30–45 17–25 40
Glucose L — Ceramic 0.2–0.8 µm 0.0055 14–83 0.8–1.7 110–250 57–60 75
Glucose, peptone L 35–38 Ceramic 0.2 µm 0.4 30–200 0.5–4 — 12.5–125 89
Kraft mill effluent P 48.4 Ceramic, aluminum 0.16 µm 1 × 24 60 1.75 50 27 42
oxide
Acetate L 35 Ceramic 0.2 µm 0.20 25–150 0–3.5 — 18–126 8
Acetate L 35 Inorganic composite, 0.14–0.2 µm, — 50 — — 40–70 23
zirconia oxide 0.005–0.08 µm
Molasses f L 20 Polypropylene 10 µm 0.051 — — 100–160 10–80 39
Molasses f L 20 Fiberglass 100 µm 0.056 — — 100–160 70–100 39
— L — Hydrophilized PVDF 0.22 µm — — 0.93 — 30–50 13
Sewage sludge L 25–50 — 15,000–20,000 D 0.0177 — — — 5–10 3
Sewage sludge L 30–35 Poly-ether sulfone 60,000 D 0.3 375 0.75 31 19 33
Sewage sludge L 22–50 Ceramic 0.1 µm 0.05 200 4.5 — 200–250 33
a Allmembranes were external cross-flow unless otherwise noted.
bL = laboratory/bench scale, P = pilot scale.
c D = Daltons (molecular weight cutoff).
d —Indicates value not reported.
e Pressure reported as kg/cm2 .
f Submerged membrane.

511
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

512
TABLE 8. Membrane Performance of AnMBRs for Municipal Wastewater Treatment

Trans Membrane Initial Final


Type Membrane membrane linear membrane membrane
of Temp. Membrane Pore area pressure velocity flux flux
wastewater Scalea (◦ C) material sizeb (m2 ) (kPa) (m s−1 ) (L m−2 h−1 ) (L m−2 h−1 ) Reference

Domestic L —c — — 0.00667 — 0.78–3.93 39 5 35


wastewater
Sewage P 10–28 Ceramic 13,000 D 13.6 1-2e 2 — 15–20 95
Heat-treated P 35–38 Ceramic 0.1 µm 1.06 200d 0.2–0.3 8–13 3–8 50
liquor from
sewage sludge f
Sewage P 26 Polysulfone and 15,000 D 100 1.5e 0.7 — 16 49, 56
polyvinylalcohol
Sewage P 26 Polyethylene 0.1 µm 54 1.1e 1.0 — 24 49, 56
Domestic L 12.5–28 Polyethylene 0.03 µm 0.3 10–60 — — 5–10 105
wastewater f
Note. All membranes were external cross-flow unless otherwise noted.
aL = laboratory/bench scale, P = pilot scale.
b D = Daltons (molecular weight cutoff).
c —Indicates value not reported.
d Units are mmHg instead of kPa.
e Pressure reported as kg/cm2 .
f Submerged membrane.
Anaerobic Membrane Bioreactors 513

Different membrane materials also lead to different fouling mechanisms.


For example, inorganic membranes were found to be fouled primarily by stru-
vite (MgNH4 PO4 ·6H2 O), whereas organic membranes were fouled by both
biomass and struvite.47

C. Operational Pressure and Temperature


The function relating permeate flux to the transmembrane pressure (TMP)
has two distinct zones. At low TMP the flux is proportional to pressure while
at high TMP the flux is independent of pressure,102 and this phenomenon
has been observed for AnMBRs.8,33 The transition point between the two
regimes could be called the critical pressure and has been reported to be
between 80 and 260 kPa. Operation above the critical pressure is pointless
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

because of the flux independence and therefore Beaubien et al.8 suggested


low-pressure operation.
The flux to TMP relationship is complicated by fouling. Using a higher
TMP increases the fouling rate, even below the critical flux, and therefore
flux will decrease with time.8 Consequently, there is an optimal choice of
TMP for a given membrane and application that balances these competing
factors to give a maximum permeate flow.
Increasing the temperature also increases the membrane flux because
the flux is inversely proportional to the fluid viscosity.92 For example, Ross
et al.82 found that the flux increased by 2% for each 1◦ C rise in temperature.
Higher membrane fluxes have been observed for thermophilic operation
(55◦ C) compared to mesophilic (35◦ C).40,82 Temperature is not a practical
control variable like operating pressure, however, because it will usually be
dictated by the influent wastewater temperature or the bioreactor operating
temperature.

D. Hydrodynamics
The flow velocity parallel to the membrane surface can be controlled in
pressure-driven, external membrane systems. A cross-flow velocity of 2–3
m/s has been used to minimize cake formation on the membrane surface in
external cross-flow AnMBRs.93,103 Generally, a higher flow velocity results in a
higher shear stress on membrane surfaces, which should reduce membrane
fouling. Many researchers have observed that an increase in feed velocity
led to an increase in flux.8,89,93,103 In particular, Beaubien et al.8 observed a
substantial flux increase from 15 L/m2 /h to 35 L/m2 /h when the flow velocity
increased from 1.1 m/s to 2.2 m/s. On the other hand, Ghyoot and Verstraete33
observed only a minor flux increase because sludge activity can decrease
at high sludge flow velocities due to the pumping shear stress, as already
discussed. In addition, the introduction of bubbles to create two-phase flow
514 B.-Q. Liao et al.

inside cross-flow membranes can induce greater surface shear and reduce
fouling.21
The concept of flow velocity is not applicable to vacuum-driven, im-
mersed membranes. The analogous concept, however, is the hydrodynamic
condition at the membrane surface. For example, air sparging is used to in-
duce shear at the surface of submerged membranes in aerobic MBRs and
disrupt the formation of a cake layer.20,21,102 Such a method can be applied
to anaerobic MBRs with immersed membranes by using the produced biogas
for sparging99 ; however, there appear to be no reports that have tested this
approach.

E. Mixed Liquor Suspended Solids


Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

In general, an increase in the MLSS leads to a decrease in membrane


flux.8,38,51,63,75,80,82,93 The principle reason is an increased opportunity for
cake formation and fouling. This is consistent with observations from aerobic
MBRs.64 However, some authors have observed that the flux may be con-
stant up to some “critical” MLSS. For example, in batch studies Ross et al.82
observed that a constant membrane flux of 1000 L/m2 /h was maintained for
MLSS up to 40 g/L, after which the flux decreased rapidly to 400 L/m2 /h at
an MLSS of 60 g/L. Similarly, Strohwald and Ross93 observed a rapid decline
in flux for MLSS concentrations >20 g/L.
The MLSS relationship to flux will be a function of reactor design. Reac-
tors without biomass retention, such as the CSTR, expose the membrane to
the full bulk MLSS concentration. Increases in SRT to improve reactor per-
formance will increase MLSS but lead to decreased membrane flux. This is
the design that has been examined in most AnMBR studies. In these sys-
tems, there is likely an optimal MLSS that achieves maximum reaction rates
while still retaining satisfactory membrane flux. However, retained biomass
designs that separate the reactor MLSS from the membrane process should
decouple the MLSS to flux relationship. Once more studies with these systems
have been conducted, the general relationship between MLSS and flux that
has been reported may require modification. One possibility is that MLSS and
flux may be specified independently to each provide maximum performance.

F. Comparison of Anaerobic and Aerobic MBRs


A comparison of the membrane performance between anaerobic and aerobic
MBRs is presented in Table 9. In general, the membrane flux of pressure-
driven (external) MBRs is larger than that of vacuum-driven (immersed) MBRs
due to the high liquid velocity in external cross-flow MBRs used to control
membrane fouling. The membrane flux of AnMBRs has been lower than that
of aerobic MBRs at the same temperature because AnMBRs have a higher
Anaerobic Membrane Bioreactors 515

TABLE 9. Comparison of Filtration Conditions Between External and Submerged Membranes


for AnMBRs and Aerobic MBRs

Anaerobic MBRa Aerobic MBRb


External External
Parameter Units cross-felow Submerged cross-flow Submerged

Design flux L m−2 h−1 10–40 15 50–100 20–50


Applied pressure kPa 150–450 15–50 400 20–50
Cross-flow velocity m s−1 2–5 n/ac 3–5 n/a
Energy for filtration kWh m−3 3–7.3 0.25–1.0 4–12 0.3–0.6

Temperature C 20–50 20–50 20–30 20–30
a Based on references 11, 14, 40, 42, 83, and 103.
b Based on reference 20.
c Not applicable.
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

MLSS, higher residual COD, and lack gas scouring. The increased operating
temperature for many AnMBRs compared to aerobic MBRs only partially
compensates.
The energy consumption of pressure-driven, external cross-flow MBRs is
significantly higher than that of the vacuum-driven, immersed MBRs.20,31,92,103
In addition, the energy consumption of aerobic and anaerobic membranes
is comparable for each of the external cross-flow and submerged configura-
tions. Therefore, the performance of aerobic MBRs can be used as an initial
guide to estimate AnMBR performance. In terms of a net energy balance,
however, the energy for filtration in anaerobic systems can be partially or
wholly offset by the produced methane.

VI. MECHANISMS OF MEMBRANE FOULING


AND ITS CONTROL IN AnMBRs

Membrane fouling in AnMBRs is similar to that in aerobic MBRs, although


anaerobic systems present certain unique challenges. Membrane fouling in
anaerobic MBRs is composite fouling, including biofouling, organic and in-
organic fouling. All three fouling mechanisms are usually observed simulta-
neously, although the relative contribution of each mechanism depends on
membrane characteristics, sludge characteristics, environmental conditions,
reactor design, and the operating strategy.13,16,17,47

A. Biofouling
Biofouling is the result of interactions between membrane surfaces and com-
ponents of the biological treatment broth. Biofouling mechanisms can be
classified under three categories: pore clogging, sludge cake formation, and
adsorption of extracellular polymeric substances (EPS).64
516 B.-Q. Liao et al.

Cell debris and colloidal particles cause pore clogging. During the per-
meation process, these particles with a size dimension comparable to the
pore size will accumulate in pores and reduce the surface area for filtra-
tion. Choo and Lee17 found that colloids, and not dissolved and cellular
fractions, were the main foulant of both MF and UF membranes. Increased
fouling has been associated with the use of pressure-driven, external cross-
flow filtration because pump-induced shear stresses decreased the aver-
age particle size and liberated colloids that can lead to pore clogging.18,55
The improvement in membrane flux after backwashing15,19,38,50,56,61,108 pro-
vides further evidence that pore clogging is involved in membrane foul-
ing, and also provides a mechanism to mitigate flux losses due to pore
clogging.
If the shear stress at the membrane surface is not adequate to remove
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

solids, sludge cake formation occurs. Choo and Lee16 and Kang et al.47 found
a thick cake layer composed of biomass and struvite formed on polymeric
membrane surfaces causing major hydraulic resistance. A theoretical model
to predict flux decline was developed by Choo and Lee18 by considering the
solids transport mechanism based on hydrodynamic and surface interactions.
The improvement of membrane flux by increased shear forces through the
use of gas circulation and gas–liquid two-phase flow in cross-flow mem-
branes 21,42,50 provides evidence that sludge cake formation is involved in
membrane fouling. The extent of biofouling due to cake deposition will de-
pend in part on the concentration of suspended material that is brought
into the membrane. In the CSTR configuration with either pressure-driven
or vacuum-driven membranes, the high concentration of solids presented
to the membrane will exacerbate cake deposition. Retained biomass reactor
designs that do not present the full MLSS concentration to the membrane
should be less challenged by cake deposition.
The third mechanism of biofouling is caused by the accumulation and
adsorption of extracellular polymeric substances (EPS) and soluble microbial
products (SMP) on membrane and pore surfaces. Cho and Fane13 observed
that a lower membrane flux was associated with a larger quantity of EPS per
unit membrane surface area and a strong link existed between EPS deposi-
tion load and fouling resistance. Membrane autopsy also revealed significant
fouling by EPS and an uneven distribution of EPS.

B. Organic and Inorganic Fouling


Organic fouling is due to the accumulation and adsorption of organic con-
stituents on the membrane surfaces. Fouling due to EPS could also be con-
sidered a subset of organic fouling, but was discussed earlier with biofouling
to emphasize the biological source of EPS and SMP.
The relatively higher effluent COD concentrations in AnMBR systems
compared to aerobic MBRs may increase the relative contribution of organic
Anaerobic Membrane Bioreactors 517

fouling in AnMBRs. In general, higher OLRs will lead to higher residual CODs
and lower membrane fluxes.39 Furthermore, it is the absolute residual COD
and not the COD removal efficiency that affects fouling. For example, a reac-
tor with only 70% COD removal with a feed COD of 0.6 g/L will have a lower
residual COD and lower fouling propensity than a reactor with a 95% COD
removal and feed COD of 20 g/L. Therefore, operating at higher SRT may
help decrease organic fouling by decreasing the effluent COD concentration.
Powdered activated carbon and zeolites have been added into AnMBRs to
adsorb soluble organic compounds and thus reduce organic fouling and en-
hance membrane flux,15,77 although this approach would likely be impractical
at full scale.
Inorganic fouling is caused by inorganic colloids and crystals on mem-
brane and pore surfaces. Struvite (MgNH4 PO4 ·6H2 O) precipitation appears
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

to be the most common inorganic foulant,16,74,108 and it can occur on both


organic and inorganic membranes.47 Other inorganic foulants can include
K2 NH4 PO4 and CaCO3 .73,74 AnMBRs may be more susceptible to inorganic
fouling than aerobic MBRs, in part because of greater opportunity for pH
shifts due to carbon dioxide partial pressure changes and the production
of high ammonia and phosphate concentrations, especially during sludge
digestion.

C. Fouling Management
Because membrane fouling directly affects flux, fouling management has
received extensive attention by researchers.64 In general, managing fouling
in AnMBRs can follow the general two-pronged approach used to manage
fouling in aerobic MBRs: (1) reducing the rate of fouling, and (2) cleaning a
fouled membrane.
Reducing the rate of fouling can prolong the length of time between
cleanings. The fouling rate can be reduced by operating a membrane be-
low the critical flux and by maintaining a high shear across the membrane
surface, either by velocity gradient or gas sparging. For example, Pierkiel
and Lanting79 used torsion shear to vibrate a polymeric Teflon membrane
and only performed chemical cleaning every 30 d. Backwashing the mem-
brane using permeate has been used extensively to disrupt the pore clogging
and cake formation components of fouling.15,19,38,50,56,61,75,108 Similarly, inter-
mittent backwashing with air has also been used in an AnMBR,61 although
backwashing with biogas would be better so as to eliminate safety concerns,
changes in redox potential, and toxicity to obligate anaerobic organisms.
Proper reactor design is crucial to reducing the fouling rate, especially
for AnMBRs. As already discussed, CSTR reactor designs will expose the
membrane to higher MLSS concentrations, increasing the rate of pore clog-
ging and cake formation compared to reactor designs that retain biomass.
518 B.-Q. Liao et al.

The more intimate exposure of the membranes to biomass may also increase
fouling due to EPS production associated with biomass. Operating the reactor
at higher SRT to minimize the COD concentration exposed to the membrane
should decrease the rate of organic fouling.
Decreasing the rate of membrane fouling will not stop fouling, how-
ever, and at some point the flux decline or TMP increase will be suffi-
ciently great that the membrane must be cleaned. Because some fouling,
especially due to adsorption of organic and inorganic constituents, is irre-
versible, the membrane cannot be cleaned to its original state. Chemical
agents have been widely used for cleaning membranes in AnMBRs. For ex-
ample, acidic cleaning (HCl, H2 SO4 ) has been extensively used to remove
inorganic foulants.15,61,83,108 Alkaline cleaning (NaOH) has been used to re-
move biological fouling,61 while caustic hypochlorite83 and ozone aeration54
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

have been used to dissolve organic foulants.

VII. OPPORTUNITIES FOR ANAEROBIC MEMBRANE BIOREACTORS

Although AnMBRs have not received as much attention as their aerobic coun-
terparts in the past decade, more research has been conducted than is sug-
gested by recent reviews.92,100,102 As demonstrated by this review, AnMBRs
have already been tested with a large variety of wastewaters and half of the
research has been conducted at pilot or full scale. This information can allow
predictions of the conditions under which AnMBRs may be most successfully
used, as well as identify research needs to improve the technology.

A. Application to Different Wastewaters


The principle characteristics of wastewater can be conceptualized along two
axes, one describing the concentration of the constituents, and the other de-
scribing the particulate nature of the constituents (Figure 2). Sludges from
municipal wastewater treatment plants, for example, have high concentra-
tions of particulate constituents and fall in quadrant (b). Examples of wastew-
aters in the other quadrants are included in Table 10. The information sum-
marized in Figure 2 and Table 10 can be used to answer the question about
AnMBR application.
Wastewaters containing high concentrations of soluble constituents
[quadrant (a)] are currently treated effectively with a variety of high-rate
anaerobic reactors designs, but especially UASB and expanded granular
sludge bed (EGSB) reactors.59 These retained biomass processes already de-
couple the SRT from the HRT, allowing for high organic loading rates and
COD removal efficiencies with a minimum of solids loss in the effluent. Be-
cause most AnMBRs have been CSTR configurations with pressure-driven
Anaerobic Membrane Bioreactors 519
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

FIGURE 2. Applicability of AnMBRs to different types of wastewaters. Quadrant descriptions


are given in Table 10.

external membranes, the OLR achievable by AnMBRs for these wastewaters


is well below that already achieved. There appears to be minimal opportu-
nity for AnMBRs to be applied to these wastewaters, although relatively little
research has been done combining membranes with retained biomass reac-
tor configurations. The use of the membrane may, for example, allow higher
hydraulic loadings for upgrades of existing or underdesigned systems.39 Cer-
tainly, the membrane will produce effluent with a negligible suspended solids
concentration, but with increased operating concerns and higher costs for
membrane purchase and maintenance. Application of AnMBRs to quadrant
(a) wastewaters would be necessary only if very low suspended solids con-
centrations were required in the effluent.
In contrast to quadrant (a) wastewaters, high-strength particulate
wastewaters (quadrant (b)) may be particularly well suited for treatment us-
ing AnMBRs. For example, in anaerobic sludge digestion the SRT and HRT
are not typically decoupled and the capital cost is high because a long re-
tention time (large volume) is required for sufficient hydrolysis. Pierkiel and
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

520
TABLE 10. Advantages of Membranes for Anaerobic Treatment of Different Wastewaters

(a) (b) (c) (d)


High-strength High-strength Low-strength Low-strength
soluble particulate soluble particulate

Wastewater characteristics
Examples Food processing Sludges, manures, Municipal sewage Municipal sewage (raw sewage)
slaughterhouses (primary effluent)
Temperature Warm Warm or cold Cold Cold
Existing technology Anaerobic Anaerobic Aerobic Aerobic
Anaerobically treatable? Yes Yes Yes Yes
Anaerobic treatment
Applicable anaerobic AF, AH, EGSB, FB, CSTR ABR, AF, AH, EGSB, Two-stage AF + UASB or UASB
reactor configuration(s)a UASB UASB + EGSB, ABR
SRT decoupled from HRT? Yes No Yes Yes
AnMBR Treatment
Effect of a membrane Effluent TSS Can decouple SRT/HRT Improve biomass Improve biomass retention,
removal retention improve particle hydrolysis
Overall improvement due to Small Possibly large cost Medium, can prevent Medium, can prevent biomass
membrane reduction if SRT/HRT biomass loss loss
decoupled
a ABR = anaerobic baffled reactor, AF = anaerobic filter, AH = anaerobic hybrid (UASB with packing instead of solids/liquid/gas separator), CSTR = completely
stirred tank reactor, EGSB = expanded granular sludge bed, FB = fluidized bed, UASB = upflow anaerobic sludge bed.
Anaerobic Membrane Bioreactors 521

Lanting79 and Pillay et al.80 used a membrane to decouple the SRT and HRT.
The present value cost of the membrane digester system of Pillay et al.80 was
significantly lower than for the conventional digester design. Furthermore,
complete retention of particulates may allow greater treatment efficiency by
allowing more complete hydrolysis of slowly degraded compounds. On the
other hand, Ghyoot and Verstraete33 found that the sludge digester OLR could
not be increased even though the solids concentration was increased from
22 g/L to 35 g/L with the addition of a membrane. This was attributed to
the shear stress during pumping through the external membrane, causing
a decrease in microbial activity. These conflicting reports notwithstanding,
there appears to be extensive opportunity to apply AnMBRs for quadrant
(b) wastes to reduce reactor volumes and capital costs.
The remaining two wastewaters are low-strength soluble [quadrant
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

(c)] and low-strength particulate [quadrant (d)]. Currently, these are treated
aerobically, except in warm climates.98 Preliminary research has been
conducted, however, assessing the feasibility of anaerobic municipal wastew-
ater treatment in cold and temperate climates.48,62,87 Complete retention of
the biomass is critical for low-temperature treatment because “little if any
viable biomass can be allowed to wash out from the reactor.”62 Membranes
are ideally suited for anaerobic treatment at low temperature because of
the extremely high solids retention capability. This means that higher hy-
draulic loadings could be used without fear of biomass washout. For ex-
ample, Ince et al.45 observed a factor of 50 increase in methanogens based
on most probable number (MPN) measurements in an AnMBR using an ex-
ternal cross-flow membrane, indicating the membrane successfully retained
these slow-growing organisms. Thus, AnMBRs could be well applied to the
treatment of low temperature municipal wastewater. Nevertheless, while op-
portunity to treat these wastewaters with AnMBRs appears solid, the main
competitive technology is the aerobic MBR. Baek and Pagilla5 found that
both aerobic and anaerobic MBRs achieved similar soluble COD removals
when treating primary clarifier effluent at the same HRT, so AnMBRs appear
to be capable of similar treatment performance as aerobic MBRs. However,
for dilute wastewater such as primary effluent there may not be any biogas
production,5 though the cost of aeration will still be reduced or eliminated.

B. Research Needs
This review has uncovered knowledge gaps that are impeding the use of
AnMBRs as a viable wastewater treatment technology. To fill these gaps
and allow greater utilization of AnMBRs, more research is needed to as-
sess the feasibility of AnMBR treatment of each of the four wastewater types
(Figure 2), assess in greater depth the use of immersed membranes, assess
strategies for membrane fouling control, assess in greater depth combining
522 B.-Q. Liao et al.

membranes with retained biomass reactor designs, assess the impact of mem-
branes on biological activity, and determine the conditions under which
AnMBR systems will be economically feasible.
There appears to be little reason to pursue research on AnMBRs for
treatment of high-strength soluble wastewaters except where membranes
could prevent biomass washout under toxicity or overload events. Mem-
brane sludge digesters, on the other hand, appear to have great potential,
and research is needed to determine which membrane configuration (ex-
ternal cross-flow or immersed) would be best suited to such high solids
concentrations as well as to develop appropriate reactor designs that can
minimize the solids concentration that contacts the membrane to minimize
fouling. In particular, special consideration must be given to inorganic fouling
(e.g., struvite) because of the high concentrations of ammonia and phosphate
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

released during sludge digestion. For low-strength wastewaters such as mu-


nicipal wastewater, there appears to be promise for AnMBRs, but more work
is needed at low temperatures in addition to combining membranes with
existing high-rate reactor configurations already determined to be suitable
for dilute wastewaters (i.e., EGSB reactor62 ).
More engineering research needs to be directed toward the membranes
themselves. Determination of the conditions under which each membrane
configuration (external cross-flow or immersed) and type of material (e.g. ce-
ramic, polymer, etc.) is most appropriate would be beneficial for practical ap-
plication. However, there have been a very limited number of AnMBR studies
that have used the vacuum-driven immersed membrane technology. Given
their significantly lower energy consumption in comparison to pressure-
driven, external cross-flow membranes, immersed membranes have tremen-
dous promise for anaerobic membrane bioreactors. Research is needed to
determine the optimal operation of immersed AnMBR configurations with
respect to TMP, backwashing frequency, and time between cleanings, with
the knowledge gained from immersed aerobic MBR operation acting as a
starting point. In addition, the use of biogas sparging of immersed mem-
branes should be investigated. This can include proper reactor design to
utilize produced gas bubbles for natural scouring of membrane modules and
the augmentation of this with biogas sparging if the gas production is too
low.
An important area of future research is in reactor–membrane integration,
in particular, the combination of membranes with retained biomass reactor
designs in order to decouple the active biomass concentration from the solids
concentration applied to the membrane. In this way, the membrane does not
conduct all of the solids/liquid separation and fouling should be reduced.
Higher SRTs and consequently higher loading rates can be achieved with
retained biomass reactor designs without necessarily increasing the solids
concentration to the membrane. This may help with the low OLRs and HRTs
seen in many AnMBR systems to date in comparison to existing high-rate
Anaerobic Membrane Bioreactors 523

anaerobic reactors. Some interesting questions arise when retained biomass


systems are used. Can organic fouling be reduced by reducing the effluent
COD concentration at higher SRT (i.e., higher biomass concentration)? Can all
biological parameters be specified independently of the membrane process?
Since aerobic systems do not employ biomass retention, and consequently
the solids concentration applied to the membrane is equal to the biomass
concentration, can membrane fluxes be higher in retained biomass anaerobic
MBRs than in aerobic MBRs?
Another important area that needs further research is the control of mem-
brane fouling in AnMBRs. The membrane fouling rate and cleaning frequency
impact the economy of the AnMBR processes. A starting point is to determine
the foulants under different environmental and operating conditions. Various
microscopy techniques, including transmission electron microscopy, scan-
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

ning electron miscroscopy, confocal scanning laser microscopy, and Fourier


transform infrared analysis, can be used to identify the nature of foulants.
Once the knowledge of foulants is obtained, strategies can be developed to
control membrane fouling. Manipulation of hydrodynamic conditions around
membrane surfaces and measurement of the physical and chemical properties
of granular sludge in AnMBRs should be conducted. Development of novel
membrane materials, cleaning strategies, and agents, based on the nature of
foulants, will benefit the control of membrane fouling in AnMBRs.
For those applications where external cross-flow membranes are de-
termined to be more applicable than immersed membranes, the riddle of
potentially lower biomass activity due to pumping shear stress needs to be
solved. Biomass activity can be assessed using traditional activity assays in
addition to phylogenetic analytical techniques. 67,106 These latter techniques
would also be useful for determining the extent to which the use of a mem-
brane alters the microbial community structure and if this alters the treatment
performance of the system.
For practical application, AnMBRs must be both technically and eco-
nomically feasible. In addition to the technical feasibility research already
suggested, economic analyses are required. External cross-flow membranes
have higher fluxes20 and lower effluent concentrations103 but higher energy
consumption20 in comparison to immersed membranes, so life-cycle costing
could be used to determine which has the lower overall costs for a given
application. Economic analyses can be used to determine if reductions in
aeration energy and methane production can offset electricity consumption
for membrane filtration in addition to the capital and operating expenses for
the membrane and energy conversion technologies.

REFERENCES

[1] Adham, S., and Gagliardo, P. Membrane bioreactors for water repurification—
Phase I: Final technical report. Desalination Research and Development
524 B.-Q. Liao et al.

Program Report No. 34. U.S. Department of the Interior, Bureau of Reclamation,
http://www.usbr.gov/pmts/water/media/pdfs/report034.pdf, 1998.
[2] Anderson, G.K., Saw, C.B., and Fernandes, M.I.A.P. Application of porous
membranes for biomass retention in biological wastewater treatment processes,
Proc. Biochem. 21, 174, 1986.
[3] Aya, H., and Namiki, K. Anaerobic digestion of sewage sludge by membrane
separated bioreactor, Mizu Kankyo Gakkaishi (Japanese) 15, 187, 1992.
[4] Azbar, N., Ursillo, P., and Speece, R.E. Effect of process configuration and
substrate complexity on the performance of anaerobic processes, Water Res.
35, 817, 2001.
[5] Baek, S., and Pagilla, K. Comparison of aerobic and anaerobic membrane
bioreactors for municipal wastewater treatment, Proceedings of WEFTEC.03,
Los Angeles, CA, 2003.
[6] Bailey, A.D., Hansford, G.S., and Dold, P.L. The enhancement of upflow anaer-
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

obic sludge bed reactor performance using crossflow microfiltration, Water Res.
28, 291, 1994.
[7] Beal, L., and Monteggia, L.O. The effect of microfiltration upon methanogenic
activity, Proceedings of the 10th World Congress on Anaerobic Digestion,
Montreal, QC, 2004, 389.
[8] Beaubien, A., Baty, M., Jeannot, F., Francoeur, E., and Manem, J. Design and
operation of anaerobic membrane reactors: Development of a filtration strat-
egy, J. Memb. Sci. 109, 173, 1996.
[9] Bisschops, I., and Spanjers, H. Literature review on textile wastewater charac-
terization, Environ. Technol. 24, 1399, 2003.
[10] Brockman, M., and Seyfried, C.F. Sludge activity and cross-flow
microfiltration—a non-beneficial relationship, Wat. Sci. Technol. 34(9),
205, 1996.
[11] Butcher, G.J. Experiences with anaerobic digestion of wheat starch waste, Int.
Biodeterioration 25, 71, 1989.
[12] Cadi, Z., Huyard, H., Manem, J., and Moletta, R. Anaerobic digestion of a syn-
thetic wastewater containing starch by a membrane reactor, Environ. Technol.
15, 1029, 1994.
[13] Cho, B.D., and Fane, A.G. Fouling transients in nominally sub-critical flux
operation of a membrane bioreactor, J. Memb. Sci. 209, 391, 2002.
[14] Choate, W.T., Houldsworth, D., and Bulter, G.A. Membrane-enhanced anaer-
obic digesters, Proceedings of the 37th Industrial Waste Conference, Purdue
University, Lafayette, IN, May 11–13, 1983, 661.
[15] Choo, K.H., Kang, I.J., Yoon, S.H., Park, H., Kim, J.H., Adiya, S., and Lee, C.H.
Approaches to membrane fouling control in anaerobic membrane bioreactors,
Water Sci. Technol. 41(10–11), 363, 2000.
[16] Choo, K.H., and Lee, C.H. Membrane fouling mechanisms in the membrane-
coupled anaerobic bioreactor, Water Res. 30, 1771, 1996.
[17] Choo, K.H., and Lee, C.H. Effect of anaerobic digestion broth composition on
membrane permeability, Water Sci. Technol. 34(9), 173, 1996.
[18] Choo, K.H., and Lee, C.H. Hydrodynamic behavior of anaerobic biosolids dur-
ing crossflow filtration in the membrane anaerobic bioreactor, Water Res. 32,
3387, 1998.
Anaerobic Membrane Bioreactors 525

[19] Chung, Y.C., Jung, J.Y., Ahn, D.H., and Kim, D.H. Development of two phase
anaerobic reactor with membrane separation system, J. Environ. Sci. Health A
33, 249, 1998.
[20] Côte, P., and Thompson, D. Wastewater treatment using membranes: The North
American experience, Water Sci. Technol. 41(10–11), 209, 2000.
[21] Cui, Z.F., Chang, S., and Fane, A.G. The use of gas bubbling to enhance mem-
brane processes, J. Membr. Sci. 221, 1, 2003.
[22] Defour, D., Derycke, D., Liessens, J., and Pipyn, P. Field experience with differ-
ent systems for biomass accumulation in anaerobic reactor technology, Water
Sci. Technol. 30(12), 181, 1994.
[23] Elmaleh, S., and Abdelmoumni, L. Cross-flow filtration of an anaerobic
methanogenic suspension. J. Membr. Sci. 131, 261, 1997.
[24] Elmitwalli, T.A., Soellner, J., de Keizer, A., Bruning, H., Zeeman, G., and
Lettinga, G. Biodegradability and change of physical characteristics of particles
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

during anaerobic digestion of domestic sewage, Water Res. 35, 1311, 2001.
[25] Fakhru’-Razi, A. Ultrafiltration membrane separation for anaerobic wastewater
treatment, Water Sci. Technol. 30(12), 321, 1994.
[26] Fakhru’l-Razi, A., and Noor, M.J.M.M. Treatment of palm oil mill effluent
(POME) with the membrane anaerobic system (MAS), Water Sci. Technol.
39(10–11), 159, 1999.
[27] Fetzner, S. Bacterial dehalogenation, Appl. Microbiol. Biotechnol. 50, 633, 1998.
[28] Fox, P., and Pohland, F.G. Anaerobic treatment applications and fundamentals:
Substrate specificity during phase separation, Water Environ. Res. 66, 716, 1994.
[29] Fuchs, W., Binder, H., Mavrias, G., and Braun, R. Anaerobic treatment of
wastewater at high organic content using a stirred tank reactor coupled with a
membrane filtration unit, Water Res. 37, 902, 2003.
[30] Fukuma, M., Takesada, K., and Yasunishi, A. Two-phase anaerobic treatment of
wastewater containing cellulose using membrane module in acidogenic phase,
Kagaku Kogaku Ronbunshu (Japanese) 19, 936, 1993.
[31] Gander, M., Jefferson, B., and Judd, S. Aerobic MBRs for domestic wastewater
treatment: a review with cost considerations, Sep. Purif. Technol. 18, 119, 2000.
[32] Ghyoot, W.R., Vandaele, S., and Verstraete, W. Nitrogen removal from sludge
reject water with a membrane-assisted bioreactor, Water Res. 33, 23, 1999.
[33] Ghyoot, W.R., and Verstraete, W. Coupling membrane filtration to anaerobic
primary sludge digestion, Environ. Technol. 18, 569, 1997.
[34] Grady, C.P.L., Daigger, G.T., and Lim, H.C. Biological Wastewater Treatment,
2nd ed. Marcel Dekker, New York, 1999.
[35] Grethlein, H.E. Anaerobic digestion and membrane separation of domestic
wastewater, J. Water Pollut. Control Fed. 50, 754, 1978.
[36] Hall, E.R., Onysko, K.A., and Parker, W.J. Enhancement of bleached Kraft
organochlorine removal by coupling membrane filtration and anaerobic treat-
ment, Environ. Technol. 16, 115, 1995.
[37] Hao, O.J., Kim, H., and Chiang, P.C. Decolorization of wastewater, Crit. Rev.
Env. Sci. Technol. 30, 449, 2000.
[38] Harada, H., Momonoi, K., Yamazaki, S., and Takizawa, S. Application of
anaerobic-UF membrane reactor for treatment of a wastewater containing high
strength particulate organics, Water Sci. Technol. 30(12), 307, 1994.
526 B.-Q. Liao et al.

[39] Hernandez, A.E., Belalcazar, L.C., Rodriguez, M.S., and Giraldo, E. Retention
of granular sludge at high hydraulic loading rates in an anaerobic mem-
brane bioreactor with immersed filtration, Water Sci. Technol. 45(10), 169,
2002.
[40] Hogetsu, A., Ishikawa, T., Yoshikawa, M., Tanabe, T., Yudate, S., and Sawada,
J. High rate anaerobic digestion of wool scouring wastewater in a digester
combined with membrane filter, Water Sci. Technol. 25(7), 341, 1992.
[41] Imasaka, T., Kanekuni, N., So, H., and Yoshino, S. Cross-flow filtration of
methane fermentation broth by ceramic membranes, J. Ferment. Bioeng. 68,
200, 1989.
[42] Imasaka, T., So, H., Matsushita, K., Furukawa, T., and Kanekuni, N. Appli-
cation of gas–liquid two-phase cross-flow filtration to pilot-scale membrane
fermentation, Drying Technol. 11, 769, 1993.
[43] Ince, O., Anderson, G.K., and Kasapgil, B. Effect of changes in composition
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

of methanogenic species on performance of a membrane anaerobic reactor


system treating brewery wastewater, Environ. Technol. 16, 901, 1995.
[44] Ince, O., Anderson, G.K., and Kasapgil, B. Control of organic loading rate
using the specific methanogenic activity test during start-up of an anaerobic
digestion system, Water Res. 29, 349, 1995.
[45] Ince, O., Anderson, G.K., and Kasapgil, B. Composition of the microbial pop-
ulation in a membrane anaerobic reactor system during start-up, Water Res. 31,
1, 1997.
[46] Kalogo, Y., and Verstraete, W. Development of anaerobic sludge bed (ASB)
reactor technologies for domestic wastewater treatment: motives and perspec-
tives, World J. Microbiol. Biotechnol. 15, 523, 1999.
[47] Kang, I.J., Yoon, S.H., and Lee, C.H. Comparison of the filtration characteris-
tics of organic and inorganic membranes in a membrane-coupled anaerobic
bioreactor, Water Res. 36, 1803, 2002.
[48] Kashyap, D.R., Dadhich, K.S., and Sharma, S.K. Biomethanation under psy-
chrophilic conditions: a review, Bioresource Technol. 87, 147, 2003.
[49] Kataoka, N., Tokiwa, Y., Tanaka, Y., Fujiki, K., Taroda, H., and Takeda, K.
Examination of bacterial characteristics of anaerobic membrane bioreactors
in three pilot-scale plants for treating low strength wastewater by application
of colony forming curve analysis method, Appl. Environ. Microbiol. 58, 2751,
1992.
[50] Kayawake, E., Narukami, Y., and Yamagata, M. Anaerobic digestion by a ce-
ramic membrane enclosed reactor, J. Ferment. Bioeng. 71, 122, 1991.
[51] Kayawake, E., Tohya, S., Rokudai, M., Shimizu, Y., Honda, S., Tanaka, R., and
Eguchi, K. Methane fermentation by a membrane anaerobic reactor system
and characterization of fermentation broth, Hakko Kogaku Kaishi (Japanese)
66, 453, 1988.
[52] Kayawake, E., Tohya, S., Rokudai, M., Shimizu, Y., Honda, S., Tanaka, R., and
Eguchi, K. Anaerobic digestion of artificial wastewater containing cellulose by
a membrane bioreactor, Hakko Kogaku Kaishi (Japanese) 67, 255, 1989.
[53] Kim, J.O., and Somiya, I. Effect of hydraulic loading rate on acidogenesis
in a membrane-coupled anaerobic VFAs fermenter, Environ. Technol. 22, 91,
2001.
Anaerobic Membrane Bioreactors 527

[54] Kim, J.O., Somiya, I., and Kishimoto, N. Flux recovery by ozone aeration on
anaerobic membrane bioreactor for carbon recovery, Environ. Eng. Res. (Jpn.)
35, 265, 1998.
[55] Kim, J.S., Lee, C.H., and Chang, I.S. Effect of pump shear on the performance
of a crossflow membrane bioreactor, Water Res. 35, 2137, 2001.
[56] Kimura, S. Japan’s aqua renaissance ’90 project, Water Sci. Technol. 23(7–9),
1573, 1991.
[57] Kiriyama, K., Tanaka, Y., and Mori, I. Field test of a composite methane gas
production system incorporating a membrane module for municipal sewage,
Water Sci. Technol. 25(7), 135, 1992.
[58] Kiriyama, K., Tanaka, T., and Mori, I. Field test on a methane fermentation
treatment system incorporating a membrane module for municipal sewage,
Desalination 98, 199, 1994.
[59] Kleerebezem, R., and Macarie, H. Treating industrial wastewater: Anaerobic
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

digestion comes of age, Chem. Eng. April, 56, 2003.


[60] Koyuncu, I., Topacik, D., and Yuksel, E. Reuse of reactive dyehouse wastew-
ater by nanofiltration: Process water quality and economical implications, Sep.
Purif. Technol. 36, 77, 2004.
[61] Lee, S.M., Jung, J.Y., and Chung, Y.C. Novel method for enhancing permeate
flux of submerged membrane system in two phase anaerobic reactor, Water
Res. 35, 471, 2001.
[62] Lettinga, G., Rebac, S., and Zeeman, G. Challenge of psychrophilic anaerobic
wastewater treatment, Trends Biotechnol. 19, 363, 2001.
[63] Li, A., Kothari, D., and Corrado, J.J. Application of membrane anaerobic reactor
system for the treatment of industrial wastewaters, Proceedings of the 39th
Industrial Waste Conference, Purdue University, Ann Arbor Science, Ann Arbor,
MI, 627, 1985.
[64] Liao, B.Q., Bagley, D.M., Kraemer, H.E., Leppard, G.G., and Liss, S.N. A re-
view of biofouling and its control in membrane separation bioreactors, Water
Environ. Res. 76, 425, 2004.
[65] Lin, B., Van Verseveld, H.W., and Roling, W.F.M. Microbial aspects of anaerobic
BTEX degradation, Biomed. Environ. Sci. 15, 130, 2002.
[66] Mata-Alvarez, J. (ed.) Biomethanization of the Organic Fraction of Municipal
Solid Wastes, IWA Publishing, London, 2002.
[67] Merkel, W., Manz, W., Szewzyk, U., and Krauth, K. Population dynamics in
anaerobic wastewater reactors: modelling and in situ characterization, Water
Res. 33, 2392, 1999.
[68] Metcalf and Eddy, Inc. Wastewater Engineering: Treatment and Reuse, 4th ed.,
McGraw-Hill, New York, 2003.
[69] Minami, K. A trial of high performance anaerobic treatment on wastewater
from Kraft pulp and mill, Desalination 98, 273, 1994.
[70] Minami, K., Okamura, K., Ogawa, S., and Naritomi, T. Continuous anaerobic
treatment of wastewater from a Kraft pulp mill, J. Ferment. Bioeng. 71, 270,
1991.
[71] Murata, M., Kimuro, H., Kanekuni, N., Ohkuma, N., Ogasawara, H., and
Fujioka, T. Small-scale sewage plant experiment by pre-treatment and metha-
nization of suspended solids, Desalination 98, 217, 1994.
528 B.-Q. Liao et al.

[72] Nagano, A., Arikawa, E., and Kobayashi, H. The treatment of liquor wastewater
containing high-strength suspended solids by membrane bioreactor system,
Water Sci. Technol. 26(3–4), 887, 1992.
[73] Nagata, N., Herouvis, K.J., Dziewulski, D.M., and Belfort, G. Cross-flow mem-
brane microfiltration of a bacterial fermentation broth, Biotechnol. Bioeng. 34,
447, 1989.
[74] Norddahl, B., and Rohold, L. The BiorekR concept for the conversion of or-
ganic effluent to energy, concentrated fertilizer and potable water, http://
www.bioscan.dk/The%20BIOREK%20Concept%20may%202000.pdf, 2000.
[75] Oh, S.E., Iyer, P., Bruns, M.A., and Logan, B.E. Biological hydrogen production
using a membrane bioreactor, Biotechnol. Bioeng. 87, 119, 2004.
[76] Okamura, K., Ogoshi, T., Fujioka, T. and Inoue, G. Anaerobic treatment of
pulp paper wastewater—Results of Aqua Renaissance 90, Kami Pa Gikyoshi
(Japanese) 48, 161, 1994.
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

[77] Park, H., Choo, K.H., and Lee, C.H. Flux enhancement with PAC addition in
the membrane anaerobic bioreactor, Sep. Sci. Technol. 34, 2781, 1999.
[78] Pearce, C.I., Lloyd, J.R., and Guthrie, J.T. The removal of colour from tex-
tile wastewater using whole bacterial cells: A review, Dyes Pigment. 58, 179,
2003.
[79] Pierkiel, A., and Lanting, J. Membrane-coupled anaerobic digestion of munic-
ipal sewage sludge, Proceedings of the 10th World Congress on Anaerobic
Digestion, Montreal, QC, 2004, 738.
[80] Pillay, V.L., Townsend, B., and Buckley, C.A. Improving the performance of
anaerobic digesters at wastewater treatment works: The coupled crossflow
microfiltration/digester process, Water Sci. Technol. 30(12), 329, 1994.
[81] Rintala, J.A., and Puhakka, J.A. Anaerobic treatment in pulp- and paper-mill
waste management: A review, Bioresource Technol. 47, 1, 1994.
[82] Ross, W.R., Barnard, J.P., Le Roux, J., and de Villiers, H.A. Application of ultrafil-
tration membranes for solids-liquid separation in anaerobic digestion systems:
The ADUF process, Water S.A. 16, 85, 1990.
[83] Ross, W.R., Barnard, J.P., Strohwald, N.K., Grobler, C.J., and Sanetra, J. Practical
application of the ADUF process to the full-scale treatment of maize-processing
effluent, Water Sci. Technol. 25(10), 27, 1992.
[84] Sainbayar, A., Kim, J.S., Jung, W.J., Lee, Y.S., and Lee, C.H. Application of sur-
face modified polypropylene membranes to anaerobic membrane bioreactor,
Environ. Technol. 22, 1035, 2001.
[85] Salminen, E. and Rintala, J. Anaerobic digestion of organic solid poultry slaugh-
terhouse waste—A review, Bioresource Technol. 83, 13, 2002.
[86] Schonberg, J.C., Bhattacharya, S.K., Madura, R.L., Mason, S.H., and Con-
way, R.A. Evaluation of anaerobic treatment of selected petrochemical wastes,
J. Haz. Mater. 54, 47, 1997.
[87] Seghezzo, L., Zeeman, G., van Lier, J.B., Hamelers, H.V.M., and Lettinga, G.
A review: The anaerobic treatment of sewage in UASB and EGSB reactors,
Bioresource Technol. 65, 175, 1998.
[88] Shimizu, Y., Rokudai, M., Tohya, S., Kayawake, E., Yazawa, T., Tanaka, H.,
and Eguchi, K. Filtration characteristics of charged alumina membranes for
methanogenic waste, J. Chem. Eng. Jpn. 22, 635, 1989.
Anaerobic Membrane Bioreactors 529

[89] Shimizu, Y., Rokudai, M., Yazawa, T., and Tanaka, H. Application of membrane
bioreactor for sewage sludge treatment—Filtration characteristics of anaero-
bic digestion liquor from cellulose-containing synthetic wastewater, Gesuido
Kyokaishi (Japanese) 29, 74, 1992.
[90] Shizas, I., and Bagley, D.M. Experimental determination of energy content of
unknown organics in municipal wastewater streams. J. Energy Eng. 130, 45,
2004.
[91] Speece, R.E. Anaerobic Biotechnology for Industrial Wastewaters, Archae Press,
Nashville, TN, 1996.
[92] Stephenson, T., Judd, S., Jefferson, B., and Brindle, K. Membrane Bioreactors
for Wastewater Treatment, IWA Publication, London, UK, 2000.
[93] Strohwald, N.K.H., and Ross, W.R. Application of the ADUF process to brewery
effluent on a laboratory scale, Water Sci. Technol. 25(10), 95, 1992.
[94] Sutton, P.M., Li, R.R., and Korchin, S.R. Dorr-oliver’s fixed film suspended
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

growth anaerobic systems for industrial wastewater treatment and energy re-
covery, Proceedings of the 37th Industrial Waste Conference, Purdue Univer-
sity, Lafayette, IN, 1983, 667.
[95] Tanaka, Y. Application of membrane separation process to municipal sewage
treatment, Yosui to Haisui (Japanese) 29, 940, 1987.
[96] Tang, C., and Chen, V. Nanofiltration of textile wastewater for water reuse,
Desalination 143, 11, 2002.
[97] Vandevivere, P.C., Bianchi, R., and Verstraete, W. Treatment and reuse of
wastewater from the textile wet-processing industry: Review of emerging tech-
nologies, J. Chem. Technol. Biotechnol. 72, 289, 1998.
[98] van Haandel, A.C., and Lettinga, G. Anaerobic Sewage Treatment: A Practical
Guide for Regions With a Hot Climate, Wiley, New York, 1994.
[99] van Lier, J.B., Tilche, A., Ahring, B.K., Macarie, H., Moletta, R., Dohanyos, M.,
Hulshoff Pol, L.W., Lens, P., and Verstraete, W. New perspectives in anaerobic
digestion, Water Sci. Technol. 43(1), 1, 2001.
[100] Verstraete, W., de Beer, D., Pena, M., Lettinga, G., and Lens, P. Anaero-
bic bioprocessing of organic wastes, World J. Microbiol Biotechnol. 12, 221,
1996.
[101] Verstraete, W., and Vandevivere, P. New and broader applications of anaerobic
digestion, Crit. Rev. Environ. Sci. Technol. 28, 151, 1999.
[102] Visvanathan, C., Ben Aim, R., and Parameshwaran, K. Membrane separation
bioreactors for wastewater treatment, Crit. Rev. Environ. Sci. Technol. 30, 1,
2000.
[103] Vogelpohl, A., Sayadi, S., Si-Salah, A., and Fuchs, W. Water recycling
and reuse by application of membrane bioreactors: textile and municipal
wastewater as examples, MBR-Recycling, Summary Report, http://www.itv.tu-
clausthal.de/ITV/pdf/MBR.pdf, 2003.
[104] Wang, Y., Kim, J.H., Choo, K.H., Lee, Y.S., and Lee, C.H. Hydrophilic modi-
fication of polypropylene microfiltration membranes by ozone-induced graft
polymerization, J. Membr. Sci. 169, 269, 2000.
[105] Wen, C., Huang, X., and Qian, Y. Domestic wastewater treatment using an
anaerobic bioreactor coupled with membrane filtration, Proc. Biochem. 35,
335, 1999.
530 B.-Q. Liao et al.

[106] Wilderer, P.A., Bungartz, H.J., Lemmer, H., Wagner, M., Keller, J., and Wuertz,
S. Modern scientific methods and their potential in wastewater science and
technology, Water Res. 36, 370, 2002.
[107] Yanagi, C., Sato, M., and Takahara, Y. Treatment of wheat starch waste water by
a membrane combined two phase methane fermentation system, Desalination
98, 161, 1994.
[108] Yoon, S.H., Kang, I.J., and Lee, C.H. Fouling of inorganic membrane and flux
enhancement in membrane-coupled anaerobic bioreactor, Sep. Sci. Technol.
34, 709, 1999.
[109] Yushina, Y., and Hasegawa, J. Process performance comparison of membrane
introduced anaerobic digestion using food processing industry wastewater,
Desalination 98, 413, 1994.
[110] Zakkour, P.D., Gaterell, M.R., Griffin, P., Gochin, R.J., and Lester, J.N. Anaero-
bic treatment of domestic wastewater in temperate climates: Treatment plant
Downloaded By: [TU University of Technology Delft] At: 20:55 12 May 2011

modelling with economic considerations, Water Res. 35, 4137, 2001.


[111] Zwolinski, M.D., Harris, R.F., and Hickey, W.J. Microbial consortia involved in
the anaerobic degradation of hydrocarbons, Biodegradation 11, 141, 2000.

You might also like