Hybrid Membrane Bioreactor Technology For Small Water Treatment Utilities

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 16

Journal of Membrane Science 344 (2009) 39–54

Contents lists available at ScienceDirect

Journal of Membrane Science


journal homepage: www.elsevier.com/locate/memsci

Hybrid membrane bioreactor technology for small water treatment utilities:


Process evaluation and primordial considerations
Varadarajan Ravindran, Hsun-Hao Tsai, Mark D. Williams 1 , Massoud Pirbazari ∗
Department of Civil and Environmental Engineering, Viterbi School of Engineering, University of Southern California, 3620 S. Vermont Avenue, Kaprielian Hall (KAP) 260,
Los Angeles, CA 90089-2531, USA

a r t i c l e i n f o a b s t r a c t

Article history: A hybrid membrane bioreactor (MBR) technology was investigated for treatment of natural waters
Received 22 April 2009 containing natural organic matter (NOM), nitrate, alachlor and viruses. The NOM are precursors to
Received in revised form 9 July 2009 disinfection byproducts (DBPs) such as trihalomethanes, and potential cause for microbial growth in
Accepted 20 July 2009
distribution systems. Batch bioreactor studies optimized the carbon-to-nitrogen (C/N) ratio, tempera-
Available online 28 July 2009
ture and pH for biological denitrification. Mini-pilot MBR studies with ethanol as electron donor were
conducted under three scenarios: no powder activated carbon (PAC) or biomass, biomass alone, and com-
Keywords:
bination of PAC and biomass. The removals of nitrate, alachlor, total organic carbon (TOC), trihalomethane
Hybrid membrane bioreactor
Nitrate
formation potential (THMFP) and MS-2 virus were revaluated. Nitrite and ethanol residuals were well
Alachlor below their detection limits in the MBR effluent under steady-state conditions. Using biomass alone, the
Natural organic matter removal efficiencies for NOM (as TOC), alachlor, and THMFP were 60, 36 and 61%, respectively; whereas,
Membrane fouling on combining PAC and biomass they increased to 84, 99.8 and 98%. The virus rejection mechanisms pos-
Disinfection byproducts tulated were size exclusion, sorption to membrane surface, attachment to bio-particles, and sorption
Microbial regrowth onto foulant or cake layers. The MS-2 removals exceeded 3.8-logs under optimal process conditions.
Trihalomethane formation potential Membrane fouling and flux decline patterns under the three experimental scenarios were well described
MS-2 virus
by the resistance model. The flux decline was characterized by three distinct stages, namely, initial con-
ditioning fouling, rapid fouling, and slow fouling. The Tustin groundwater was associated with lower
organic fouling but higher inorganic fouling than the Fountain Valley groundwater.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction

Problems in drinking water supplies are often attributed to


Abbreviations: AOC, assimilable organic carbon; ATCC, American type cul- contaminants in ground and surface waters. The most common
ture collection; BDOC, biodegradable dissolved organic carbon; BOC, biodegradable contaminants associated with these problems include the follow-
organic carbon; BOM, biodegradable organic matter; BOD, biochemical oxygen ing: synthetic organic chemicals (SOCs) exemplified by pesticides,
demand; DBP, disinfection byproduct; DO, dissolved oxygen; DOC, dissolved
organic carbon; DOM, dissolved organic matter; EPS, exo-polymeric substances;
herbicides, industrial solvents and chemicals; inorganic pollutants
FVGW, Fountain Valley groundwater; GAC, granular activated carbon; GC-MS, gas such as nitrate, arsenic and toxic metals; natural organic matter
chromatography-mass spectrometry; GC-ECD, gas chromatography with electron (NOM) such as taste and odor causing compounds, and disin-
capture detector; GC-FID, gas chromatography with flame ionization detector; fection byproduct (DBP) precursors measured as total organic
HAA, haloacetic acid; HAN, haloacetonitrile; MAC, maximum acceptable concentra-
carbon (TOC); and microorganisms such as protozoa, bacteria and
tion; MBR, membrane bioreactor; MCL, maximum contaminant level; MLSS, mixed
liquor suspended solids; MS-2, male specific species 2 virus strain; NOM, natural viruses. In recent years, the United States Environmental Protection
organic matter; PAC, powdered activated carbon; PFU, plaque formation unit; PP, Agency (USEPA) has intensely regulated the water supply indus-
polypropylene; PVC, polyvinyl chloride; PVDF, polyvinylidene fluoride; SDR, specific try for controlling organic and inorganic contaminants, pathogenic
denitrification rate; SMP, soluble microbial product; SOC, synthetic organic chemi- microorganisms, and DBP formation after disinfection by chlorina-
cal; SRT, solid retention time; TDS, total dissolved solids; TGW, Tustin groundwater;
THM, trihalomethane; THMFP, trihalomethane formation potential; TMP, trans-
tion or advanced oxidation processes. The new Enhanced Surface
membrane pressure; TOC, total organic carbon; USEPA, United States Environmental Water Treatment Rule and the Disinfectant/Disinfection Byprod-
Protection Agency; UV254 , ultraviolet absorbance measured at 254 nm. uct Rule mandate lower allowable levels of selected DBPs, and over
∗ Corresponding author. Tel.: +1 213 740 0592; fax: +1 213 744 1426. 4-logs reductions in virus concentrations [1,2]. It must be noted
E-mail address: pirbazar@usc.edu (M. Pirbazari). that NOM removal from natural waters has several advantages
1
Mark Williams is currently the President of McCaron-Williams, Inc., Consulting
Engineers, Long Beach, California. At the time of research he was an engineer with
including reduction in undesirable DBPs such as trihalomethanes
the Metropolitan Water District of Southern California, La Verne, California 91750, (THMs) and haloacetic acids (HAAs) during chlorination, and con-
and also doctoral candidate at the University of Southern California. trol of microbial regrowth in distribution systems attributable to

0376-7388/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2009.07.032
40 V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54

biodegradable organic matter (BOM) [3–6]. Integrated membrane [11] and sulfur [12], specially designed to promote the necessary
technologies such as the hybrid membrane bioreactor (MBR) pro- heterotrophic or autotrophic conditions. However, water treat-
cess would be effective in providing finished water safe from ment and water reclamation represent a completely different
microbial pathogens, low in DBP precursors, and free from harm- scenario due to low microbial biomass and electron donor levels.
ful organic or inorganic contaminants. Furthermore, MBR systems The MBR effectively replaces a conventional activated sludge pro-
using ultrafiltration or microfiltration membranes with bioactive cess in which microbial biomass concentrations are on the order of
powder activated carbon (PAC) could achieve high contaminant 3000–20,000 mg/L, while electron donors and contaminants are in
removals, and simultaneous control of membrane fouling and flux the levels of 100–2500 mg/L. Additionally, biofilm thickness must
decline. be specifically tailored to optimize the process efficiencies. Nev-
The present work evaluates the hybrid MBR process as a viable ertheless, in water treatment situations, the target contaminants
water purification technology for removing inorganics such as (organic or inorganic) are generally at the 1–100 mg/L levels, while
nitrate, organics such as alachlor, NOM (TOC) that could lead to microbial biomass requirements are in the 40–400 mg/L range
DBP formation and microbial growth, and pathogens exemplified [6,8,10]. In fact, low concentrations of electron donors such as NOM
by viruses. Mini-pilot-scale MBR investigations were conducted or other constituents are often able to sustain low levels of micro-
under different process conditions to investigate the removal effi- bial populations associated with thin biofilms. Some advantages of
ciencies of target contaminants, namely, nitrate, TOC, alachlor, the MBR technology over conventional processes that are experi-
trihalomethane formation potential (THMFP), and viruses. The enced in wastewater treatment can be realized in water treatment
bacteriophage MS-2 (male specific type 2 virus) was used as an as well.
indicator and model organism to evaluate virus removals. The MBR In water supplies NOM present several problems including bac-
tests also investigated membrane fouling and permeate flux decline terial growth in distribution systems, formation of DBPs during
as functions of operation time under three operational scenarios: chlorination, and permeate flux decline in membrane processes
(i) no biomass or PAC (control), (ii) biomass only without PAC, due to membrane fouling. Bacterial growth in distribution systems
and (iii) biomass with PAC. These scenarios were chosen to obtain is attributed to NOM constituting assimilable organic carbon (AOC)
information on the relative contributions of adsorption, biodegra- or biodegradable organic matter (BOM), the natural substrates for
dation and membrane separation towards contaminant removals. microbial growth and sustenance [3–5]. More recently, Williams
Furthermore, permeate flux decline patterns under these scenarios and Pirbazari [6] demonstrated that hybrid MBR processes could
could shed light on the nature of membrane fouling for developing efficiently control DBP formation and bacterial regrowth. Further-
appropriate control strategies. more, these investigators demonstrated that MBR systems using
bioactive PAC could reduce NOM and BOM levels and amelio-
rate membrane fouling and permeate flux decline. Other studies
2. Background information showed that the technology could achieve high treatment levels
for waters contaminated with petroleum hydrocarbons [13], and
2.1. Applications of membrane bioreactor processes nitrate [10]. An important feature of the hybrid MBR technology
is the effective control of membrane fouling and permeate flux
The MBR process is more typically associated with wastewater decline by a combination of PAC sorption and fluid management.
treatment than water treatment because it effectively integrates a
suspended growth biomass similar to that in traditional activated 2.2. Removals of NOM, TOC and DBP precursors from water
sludge process with a membrane system that replaces sedimenta- supplies
tion systems but effectively retains biomass and clarifies the treated
effluent. Recently, the technology appears more promising for The NOM in potable water supplies are mostly attributed to bio-
water treatment owing to significant declines in membrane costs, logical degradation of organic substances such as amino acids, fatty
and dramatic reductions in fixed costs and operating expenses. acids, phenols, steroids, sugars, hydrocarbons, urea, porphyrines
The MBR systems offer several advantages over activated sludge and organic polymers such as polypeptides, lipids, polysaccharides,
and fixed biofilm processes generally used for biological treatment, and humic substances. The NOM levels in drinking water sources
including the following: small reactor sizes or footprints, suitability are usually 2–10 mg/L (as TOC), although much higher levels are
for water reclamation and reuse, and ability to retain most par- sometimes reported [14]. Humic substances are the most signif-
ticulate and colloidal matter. Furthermore, the treated effluent is icant among NOM constituents because they can yield over 300
low in total suspended solids, turbidity levels, biochemical oxygen different types of DBPs during chlorination, many of which raise
demand (BOD), total organic carbon (TOC), and most pathogens. public health concerns such as THMs, HAAs, and haloacetonitriles
The membrane filter retains all the biomass leading to performance (HANs) [14]. Although the regulated MCLs for THMs and HAAs are
enhancement of “slow growing” microorganisms with low yield 80 and 60 ␮g/L, future standards would lower these limits to 40 and
factors. In wastewater treatment applications, enhanced biomass 30 ␮g/L, respectively [1,2,6]. The present study investigates nat-
retention leads to high mixed liquor suspended solids (MLSS) lev- ural waters from two sources: the Fountain Valley groundwater
els of 10–20 g/L in MBR systems versus the low levels of 3–4 g/L (FVGW), and the Tustin groundwater (TGW). The FVGW has aver-
in typical activated sludge or other biological treatment systems. age NOM levels of 3.5–4.0 mg/L (as TOC) with THMFP levels around
In addition to allowing long solids retention times (SRTs), high 350 ␮g/L which was well above the MCL; and therefore, removal
biomass concentrations lead to reduction in reactor sizes. More of THM precursors shall be required before chlorination. The TGW
importantly, modern, programmable logic control systems have has a low NOM level of less than 0.1 mg/L as TOC, and therefore its
made MBR systems compact, efficient, economical and versatile THMFP is not of much concern. A major objective of the present
not only for wastewater treatment and water reclamation, but also study is to control the DBP formation potential by removing NOM
for drinking water treatment. that are well recognized as DBP precursors. Physico-chemical pro-
The MBR technology is more widely employed in wastewater cesses including coagulation, physical settling, deep bed filtration,
treatment, water reclamation and water reuse scenarios owing to or membrane processes achieve relatively low removals of DBP
good retention of microbial biomass and electron donors. Several precursors [14,15]. Reverse osmosis and nanofiltration are energy
electron donors have been employed for biological denitrification intensive membrane processes, while ultrafiltration is not cost-
including acetic acid [7], methanol [8], ethanol [9,10], hydrogen effective. Microfiltration, however, can be economical but cannot
V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54 41

reliably remove NOM and DPB precursors as it only rejects certain studies that of MS-2 removals occurred through three important
NOM fractions [5]. In comparison, a hybrid MBR process employing steps: physical filtration by the membrane itself, biomass activ-
microfiltration and bioactive PAC could effectively remove a broad ity in the bioreactor, and biofilm filtration on membrane surface.
spectrum of SOCs, most NOM fractions and DBP precursors [6]. Zheng and Liu [31,32] employed MBR processes equipped with
polyvinylidene fluoride (PVDF) and polypropylene (PP) membranes
2.3. Removals of nitrate and alachlor from water supplies with pore sizes 0.22 and 0.1 ␮m, respectively, and observed good
rejections even for virus strains smaller than the membrane pores.
Several studies have indicated that nitrate contamination of They hypothesized that a dynamic layer was formed on the mem-
potable waters has serious public health implications, notably, brane surface as well as in the pores in two stages: the initial pore
methemoglobinemia (“blue-baby” syndrome) and various forms of fouling or pore blocking stage, and subsequent cake and gel forma-
cancer [9]. Croen et al. [16] studied exposures of women to nitrate tion stage arising from biofilm growth. They noted that the dynamic
in drinking water and examined health risks associated with neural cake and gel layers often enhanced virus screening, although their
tube defects. Weyer et al. [17] articulated that nitrate levels in water formations were considerably reduced under turbulent conditions.
supplies as low as 10 mg/L caused bladder cancer in elderly women They further observed that internal pore fouling due to organic
in Iowa. The USEPA set MCLs for nitrate as nitrogen (NO3 − -N) and gel deposition or sorption contributed to pore constriction and
nitrite as nitrogen (NO2 − -N) of 10 and 1 mg/L [10], respectively; the promotion of virus screening. Zheng and Liu [32] suggested that
Ministry of Environment for Canada established the same standards virus rejections by new membranes were initially controlled by
for the maximum acceptable concentrations (MACs) of nitrate and their sieving ability, but after a few hours of operation was mainly
nitrite [18]; and the European Union set more stringent limits of influenced by the formation of a dynamic foulant layer. Addition-
10 and 0.03 mg/L [10–19]. Nitrate removal technologies include ion ally, they noted that extra-cellular polymeric substances (EPS) on
exchange, reverse osmosis, and electrodialysis, besides biological, biofilms enhanced microbial attachment to membranes. They fur-
chemical and catalytic denitrification [20]. Nevertheless, MBR pro- ther demonstrated that chemical cleaning of membranes reduced
cesses could be more economical and efficient for nitrate removal rejections of T4 virus due to breakages and losses of cake and
as established by several studies [7,10,21–23]. gel layers. They concluded that cake filtration contributed to virus
Alachlor or 2-chloro-2 ,6 -diethyl-N-(methoxymethyl)- rejection significantly more than membrane screening or gel layer
acetanilide has been the most commonly used pre-emergence rejection.
crop herbicide in Europe, North America and Japan; and in fact,
its residues have been detected in ground and surface waters at 3. Membrane fouling and permeate flux decline
levels of 0.1–10 ␮g/L [24]. It is classified as a B-2 carcinogen and
suspected endocrine disruptor [25,26], and the toxicity levels of 3.1. Fouling in the membrane bioreactor processes
many of its degradation products and metabolites are unclear
[24,27]. The World Health Organization has recommended an MCL The hybrid MBR process discussed here integrates membrane
of 20 ␮g/L for alachlor in drinking water supplies, while the USEPA filtration, activated carbon adsorption and microbial degradation,
has set a lower MCL of 2 ␮g/L [28]. Several researchers evaluated as shown in Fig. 1 (detailed description is presented in a later
conventional processes for alachlor removal from potable water section). A major factor affecting process performance and eco-
supplies. Miltner et al. [29] stated that chemical coagulation, floc- nomics is membrane fouling and permeate flux decline. Factors
culation and sedimentation in a full-scale plant were inefficient; influencing membrane fouling include biomass, colloids, NOM,
whereas clarification, filtration, softening, recarbonation, and inorganic precipitates or scalants, and extracellular polymers, and
chlorination achieved low removals. Nanofiltration and reverse they are dependent upon operating conditions. Membrane fouling
osmosis were effective but too expensive in energy requirements. is a complex phenomenon, and its predominant causes can be sum-
Pirbazari et al. [30] observed that non-bioactive granular activated marized are as follows: (i) Macromolecular and colloidal sorption,
carbon (GAC) adsorbers were effective but expensive owing to (ii) biofilm growth and attachment; (iii) inorganic matter precipita-
short operational life spans. Nevertheless, Badriyha et al. [24] tion or scaling; and (iv) general material aging. Membrane fouling
noted that bioactive GAC adsorbers were more economical and can be further classified as follows: (a) reversible fouling that can
efficient for removing alachlor and other SOCs, a major factor
behind the use of a hybrid MBR process with bioactive PAC.

2.4. Mechanisms for virus removals in membrane processes

The MBR system can exclude a broad spectrum of pathogenic


microorganisms such as protozoa and bacteria, besides viruses that
are the most difficult to remove [31,32]. Owing to difficulties expe-
rienced in assaying animal viruses, bacteriophages such as MS-2
are suggested as viral indicators. This is because bacteriophages
exemplified by MS-2 have similarities with enteric strains such as
hepatitis A and polio viruses in structure, morphology, size and
behavior, and find application in most water pollution control and
public health studies [33]. The rationale behind choosing MS-2 as
model virus is predicated upon the fact that it is one of the smallest
strains (20–25 nm in size), and that its rejection generally ensures
exclusion of larger strains.
Membrane rejection mechanisms for viruses are important from
the standpoint of MBR process design. Although viruses are gen-
erally smaller than the pores of microfiltration membranes, some
researchers have observed their high removals after biofilm for-
mation [34,35]. Shang et al. [36] concluded from their bench-scale Fig. 1. Schematic diagram of the hybrid MBR system.
42 V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54

be removed by physical cleaning; (b) irreversible fouling that can of NOM, PAC, biomass and exo-polymeric enzymes. Nystrom et
be removed by chemical cleaning; and (c) irrecoverable fouling al. [41] observed that positively charged inorganic membranes
that cannot be removed by any cleaning [37]. Membrane fouling were easily fouled by negatively charged humic molecules due
can be strategically controlled by concentration polarization sup- to electrostatic interaction. Yuan et al. [42] observed rapid initial
pression, optimization of physical and chemical cleaning protocols, fouling of microfiltration membranes due to electrostatic interac-
pre-treatment of feed wastewater, and modification of biologi- tions between humic molecules and membrane surface, but also
cal content [37]. Concentration polarization can be reduced either noted that deposition ceased when permeate flux dropped below
by turbulence promotion or permeate flux reduction. Side-stream the critical flux. They further stated that cessation in the growth
MBR processes provide shear through pumping which increases of the cake or gel layer could arise from back-transport of humic
cross-flow velocity, whereas immersed processes employ aeration molecules from membrane surfaces either due to inertial lift or
around the membrane to provide shear stresses. shear-induced diffusion induced by fluid mechanical effects, or
Several components of the MBR system can contribute to mem- from intermolecular electrostatic interactions pertaining to PAC
brane fouling including sorption onto and within membrane pores, sorption. When the MBR system with a cross-flow membrane mod-
and deposition of cake layer or gel layer on the external membrane ule is operated with some fluid turbulence (Reynolds numbers of
surface. Nonetheless, the best approach for problem analysis and 14,000 to 15,000) using PAC sorbent, it facilitates membrane back-
identification of fouling mechanisms is to investigate classical foul- transport of organic or inorganic species, and mitigates membrane
ing models, as discussed by Ho and Zydney [38–40]. These include fouling.
the standard pore blockage model, the intermediate pore blockage The MBR systems are generally operated under constant flux
model, the pore constriction model, and the cake filtration model. In conditions. Since fouling rates often increase exponentially with
the complete pore blockage model, the volumetric permeate flow the permeate flux, MBR plants using microfiltration or ultrafiltra-
rate is assumed to decline as the effective membrane area available tion operate at modest fluxes below the so-called “critical flux”
for filtration decreases. The cake formation is considered negligi- defined as the highest flux under which a prolonged filtration with
ble, and the pore blockage rate is proportional to convective flow constant permeability is possible. It is widely accepted as an oper-
of foulants. The intermediate pore blockage model accounts for the ating guide that below this critical value, permeability decline does
possibility that certain particles land on top of other particles dur- not occur with time, and above this flux level membrane fouling
ing deposition on membrane surfaces, and assumes pore blockage occurs rapidly. However, several researchers have shown that foul-
to be proportional to the uncovered membrane area. Further, multi- ing could be encountered even under sub-critical flux conditions
layer sorption of foulants is important because it would not lead to [43–46]. Zhang et al. [46] observed that sub-critical fouling under
further reduction in the permeate flow. In the cake filtration model, wastewater treatment conditions was characterized by three dis-
a uniform deposit of cake or gel is assumed to form on the mem- tinct stages: (i) initial conditioning fouling occurring during the first
brane surface. The rate of increase in hydraulic resistance of the exposure of the membrane to the liquid phase; (ii) slow fouling
deposit is assumed to be directly proportional to the rate of particle and flux deterioration under constant trans-membrane pressure
convection to the membrane. (TMP); and (iii) rapid fouling leading to more flux decline under
In MBR systems, membrane fouling could be caused by EPS constant TMP, or to a sudden TMP jump under constant flux. The
that are heterogeneous macromolecules found outside micro- first stage or initial conditioning fouling was rapid (measured in
bial cells. The EPS predominantly contain carbohydrates, proteins, hours), irreversible by nature, and prevalent under low flux or “zero
humic substances and nucleic acids. They constitute the primary flux” conditions. The second stage or “slow fouling” phase involved
infra-structure for biofilms and bacterial flocs, provide mechani- gradual coverage of the membrane surface by biopolymers such as
cal stability for cell structures, and facilitate biofilm adhesion to EPS [46]. This coverage resulted in changes in membrane surface
surfaces [37]. They can be generated by cell lysis, active excretion properties, promoting microbial attachment and growth. The pore
by microorganisms, spontaneous production of integral cellular blocking was expected to be inhomogeneous since the air and the
components from the outer membrane of gram-negative bacteria, liquid flow were distributed unevenly in MBR. Zhang et al. [46] also
and hydrolysis of NOM present in feed waters [37]. Their domi- noted that within the regions of membrane surface more fouled
nant components are polysaccharides and certain proteins, and so than others, the flux varied locally, exceeding the critical flux in
they can potentially develop foulant gel layers and severely restrict some areas, and leading to a sudden jump in TMP. They stated that
membrane transport. Another group of biological substances are changes in cake properties due to compression could also cause
soluble ESP that are soluble metabolic products (SMPs) or soluble this sudden jump in TMP. Some distinct pattern of flux decline
microbial products arising from biodegradation of organic materi- was expected in the present study wherein the MBR system was
als. The chemical composition and molecular weight distribution employed for water treatment and the foulants included micro-
of EPS depend upon the properties of feed water, microbial species bial biomass, PAC sorbent, NOM constituents (including EPS and
distribution and environment, nutritional status of bioreactors, and microbial enzymes) and SMP components.
certain operating parameters that influence EPS formation such as Membrane fouling in MBR systems described herein is mainly
substrate composition, organic loading, and solids retention time attributable to particulate, organic and biological fouling with inor-
[37]. ganic fouling deemed relatively insignificant. Most organic and
Permeate flux decline in MBR systems can be caused by bio- biological foulants are typically in the molecular weight range of
logical and organic fouling attributed to proteins, polysaccharides 500–2000 Da, although humic acids and exo-polymeric enzymes
and various NOM [38–40]. The fouling behavior can be explained could be in the range of 2000–500,000 Da [37–40]. The complex
by the combinatorial model of Ho and Zydney [40] involving composition of NOM makes evaluation of organic fouling mem-
both pore blockage and cake filtration. This model assumes that brane processes more difficult. In general, biological fouling leading
foulants are at first deposited on bare membrane surfaces, reduc- to severe losses in permeate fluxes is often attributed to BOM and
ing the area available for unhindered permeate flow. The initial EPS. Particulate fouling due to colloidal and biomass deposits can
deposits allow partial fluid flow, while a small fraction of the be eliminated by mechanical and chemical cleaning with substan-
flow occurs through blocked pores. Subsequently, more membrane tial flux recoveries. However, internal pore fouling due to gel and
fouling occurs and the foulant species accumulate on the origi- organic or bio-organic sorption, cannot be reduced by mechani-
nal layer, progressively increasing the hydraulic resistance. The cal cleaning, and in some cases can only be mitigated by chemical
fouling process indeed becomes more complex in the presence cleaning. Nonetheless, even intense chemical cleaning cannot com-
V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54 43

pletely eliminate internal pore fouling, and therefore only partial resistance. The intrinsic membrane resistance RM and the internal
flux recovery is realizable. fouling resistance RIN each contributed to about 0.5% of the total
resistance RT . The combination of concentration polarization and
3.2. Membrane fouling models and permeate flux decline patterns cake resistances is attributed to colloidal and bio-particles. The
internal pore fouling resistance RIN is mainly caused by organic
A conceptual rendering of a PAC–biomass layer formed on the deposition or pore sorption, while the gel resistance RGE is con-
surface of a tubular cross-flow membrane is shown in Fig. 2. The tributed by the surface sorption of organic molecules. It must be
TMP applied across the membrane is on the order of 10–25 psia noted that the intrinsic membrane resistance, RM , is governed by
(0.7–1.8 bars). A combination of a PAC–biomass layer progressively the hydrophobic or hydrophilic properties of the membrane, and
forms a dynamic cake on the membrane surface (membrane film therefore cannot be reduced by chemical cleaning. A combination
thickness ranges from 5 to 50 ␮m) as shown in the figure. The scan- of chemical and mechanical cleaning can eliminate substantial por-
ning electron micrograph on the inset illustrates the morphology tions of the transport resistances RCP and RCA , but in most situations
of the PAC–biomass cake formed on the membrane surface. The even efficient chemical cleaning can only partially reduce the resis-
schematic of permeate transport through the cake layer and the tance RIN . Nevertheless, in the MBR system, major resistances are
membrane also highlights the qualitative nature of concentration RCP and RCA , and so a cycle of chemical and mechanical cleaning can
profiles for electron donor and target contaminant (organic or inor- achieve substantial permeate flux recovery.
ganic) in the membrane vicinity. The concentrations of the electron Membrane fouling generally attributed to concentration polar-
donor and the target contaminant(s) are represented by the nota- ization, gel formation and cake formation can be controlled by
tions Ce and Cc , respectively, while the subscripts “bu”, “ca” and combining fluid management and adsorbent use. Turbulent flow
“pe” denote the bulk liquid phase, cake layer and permeate stream, regimes within membrane modules can create flow instabilities
respectively. In the hybrid MBR process, the concentrations of the and vortices that can depolarize dissolved solids and re-entrain col-
electron donor and target contaminant (non-fouling constituent) loids or suspended particles from the viscous sub-layer (part of the
in the bulk liquid, Ce,bu and Cc,bu manifest a gradual reduction in gel layer). In synergism with fluid management, the PAC discharges
the cake layer, as shown in Fig. 2. The permeate concentrations of several functions discussed in the earlier studies by Pirbazari and
these components, Ce,pe and Cc,pe are assumed very low because coworkers [9,49,50]. Firstly, it depolarizes the membrane surface
the electron donor (ethanol) and the target contaminant (nitrate by promoting back-diffusion (or back-migration) of organic solutes
or alachlor) are considered completely biodegradable under ideal from the membrane–fluid interface to the bulk fluid, depleting
conditions. the solute concentration by PAC sorption. Secondly, it serves as
Membrane fouling in water treatment applications can be esti- a filter aid by forming an incompressible particulate layer on the
mated by using the resistance model, wherein the permeate flux (J) membrane, and thereby enhances fluid permeability. Thirdly, its
is described by Darcy’s relationship as follows [47]: dynamics reduces the thicknesses of mass-transfer and hydrody-
namic boundary layers.
driving force P
J= = (1) In the MBR system described herein, without biomass or PAC
viscosity × total resistance RT
the permeate flux decline could be mainly attributed to organic
where P is the differential pressure across the membrane and rep- fouling due to NOM, as colloidal substances such as soil, clay and
resents the total driving force. The total resistance RT denotes the silt would be insignificant owing to the low turbidity of the source
total resistance to permeate flow through the membrane, and  is waters. In the case where biomass alone is employed without PAC,
the feed solution dynamic viscosity. The total resistance is the sum permeate flux decline might be caused by organic fouling due to
of various resistances given by the resistance-in-series approach as the combination of NOM, microbial enzymes and SMPs, besides
biomass. While biomass could predominantly lead to surface foul-
RT = RM + RCP + RC + RGE + RIN (2)
ing and cake formation during biological fouling, organic fouling
In the above relation, RM , RCP , RCA , RGE and RIN are resistances might significantly contribute to internal pore fouling. When PAC is
associated with the membrane material, concentration polariza- employed with biomass, organic fouling can be mitigated (depend-
tion, cake formation, gel formation, and internal pore fouling, ing on adsorbent dosage), while biological fouling may not be as
respectively. The osmotic pressure can be considered negligible in effectively controlled. Nonetheless, combination of PAC and micro-
ultrafiltration or microfiltration systems [47]. The resistance RM is bial biomass would reduce transient state flux decline but would
an intrinsic characteristic of the membrane, and therefore remains not significantly affect steady-state flux decline.
constant; the resistance RCP is caused by the concentration gradient The resistance model stated above represents one of the
of the component separated. The resistance RCA is attributed to the methods for membrane fouling classification due to different phe-
deposition of the cake layer on the membrane surface. The resis- nomena for wastewater treatment applications. This approach
tance RGE denotes the impedance to permeate transport caused by relies on a phenomenological classification of fouling that uses the
gel formation on the membrane surface due to outer surface sorp- resistances RM , RCP , RCA , RGE and RIN associated with the membrane
tion of organic and bio-organic macromolecules and enzymes. The material, concentration polarization, cake formation, gel formation,
resistance RIN denotes the transport impedance associated with and internal pore fouling, respectively. In the case of a hybrid MBR
pore constriction or pore deposition of organic and bio-organic process used for water treatment, especially where an adsorbent
macromolecules and enzymes. It is indeed very difficult to clearly such as PAC and microbial biomass are employed (as in this study),
separate out the resistances RCP , RCA, and RGE . However, some qual- classification of membrane fouling based on the foulant types will
itative assessments can be made on the basis of the resistance be more appropriate. The rationale behind the choice of foulant
model. type classification above phenomenological classification is based
Qualitative and quantitative assessments of permeate flux dete- on certain advantages the method offers in water treatment situ-
rioration in MBR processes can be made based on resistance models. ations. Evidence suggests that in wastewater treatment scenarios,
Choo and Lee [48] compared the relative magnitudes of these phenomenological aspects such as concentration polarization, gel
resistances in a typical anaerobic MBR process, and made certain formation, cake formation and internal pore sorption are impor-
observations relevant to this study. They noted that combinations tant, but somewhat indistinguishable. However, in water treatment
of concentration polarization resistance and resistance, RCP and situations, it is easier to distinguish between different classes of
RCA , respectively contributed to nearly 99% of overall transport foulants. For example, inorganic anions can potentially contribute
44 V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54

Fig. 2. Sketch of a PAC–biofilm layer (illustrated with an actual scanning electron micrograph) and the concentration profiles of the target contaminant and electron donor
in the membrane module of the hybrid MBR process.

to inorganic fouling due to formation of precipitates such as cal- where RI , RO and RB are resistances in the MBR process attributed
cium carbonate and calcium sulfate. Furthermore, identification to inorganic, organic and biological fouling, respectively, and RM
of foulant classes offers certain advantages in analyzing permeate is the intrinsic membrane material resistance. Here, RT can also
flux and component rejection data with reference to a number of be represented as the sum of RM and RF , where the fouling resis-
factors. These are evaluating component rejections under various tance RF represents the sum of RI , RO and RB . Furthermore, RF shall
conditions, identifying foulants of maximum concern based on flux be denoted differently for two types of MBR process operations:
decline patterns, investigating foulants that can increase rejections RFP when microbial biomass and PAC are used, and RFW when
of target contaminants, and formulating cleaning strategies or pro- biomass alone is employed. In this scenario, RO denotes fouling
tocols for optimizing membrane performances. According to the due to NOM, EPS, SMP and other organics microbially generated,
new approach, the total resistance RT may be written as RI denotes inorganic fouling due to precipitate formations or scal-
ing, and RB represents biological fouling due to microbial particles
RT = RM + RI + RO + RB = RM + RF (3) and biofilms. The resistance RT can be estimated from permeate
V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54 45

Table 1 then dried at 105 ◦ C, desiccated to room temperature, and stored


Characteristics of the groundwaters used.
in airtight glass containers. The particle sizes of the PAC ranges from
Component and Fountain Valley Tustin 30 to 50 ␮m.
concentration groundwater groundwater
(FVGW)a (TGW)b
4.3. Nutrient solution
Nitrate (NO3 − ) (mg-N/L) 0.3 20
Sulfate (SO4 2− ) (mg/L) 8.3 220 The composition of nutrient solution was as follows:
Bicarbonate (HCO3 − ) (mg/L) 80 308
FeCl3 ·6H2 O, 0.2 mg/L; Na2 MoO4 ·2H2 O, 0.02 mg/L; K2 HPO4 ,
Chloride (Cl− ) (mg/L) 15 185
Sodium (Na+ ) (mg/L) 84 125 21.7 mg/L; KH2 PO4 , 26.9 mg/L. The denitrification process required
Calcium (Ca2+ ) (mg/L) 8.1 138 the addition of a nutrient solution and an external carbon source
Magnesium (Mg2+ ) (mg/L) 1.1 39 such as ethanol to sustain the microbial population.
Total organic carbon (TOC) (mg/L) 3.5–4.0 <0.1
Ultraviolet absorbance at 254 nm (UVA254 ) 0.19–0.22 <0.005
(cm−1 ) 4.4. Microbial cultures
Dissolved Oxygen (mg/L) ∼0 8.9
Turbidity (NTU) <1 <0.1 The microbial culture used in the denitrification studies was
pH 7.9–8.3 7.8 obtained from a water reclamation plant. The culture was added
a
Data obtained from the Orange County Water District. to the source water containing essential nutrients and ethanol (the
b
Data obtained from the Municipal Water District of Orange County. electron donor). The acclimation conditions were controlled at a pH
of 7.8 ± 0.3, dissolved oxygen of zero, and temperature of 20 ± 2 ◦ C.
Pure cultures of Escherichia coli and bacteriophage MS-2 [51] (ATCC
flux data, and certain conclusions can be drawn from these data
15597-B1) used in these studies were obtained from the Micro-
acquired for the three MBR operational scenarios at various instants
biology Division Laboratory of the Metropolitan Water District of
of time.
Southern California, La Verne, California. The bacteriophage MS-2
(male specific virus) was chosen as indicator organism to evaluate
4. Materials and methods the MBR process for pathogen removal.

4.1. Water sources 4.5. Nitrate, nitrite, and dissolved oxygen analyses

Two types of groundwaters were used in the research: one with The nitrate and nitrite concentrations were determined by the
high TOC and low nitrate concentration, and another with low cadmium reduction method using a UV–visible spectrophotometer
TOC and high nitrate content. The high TOC water was obtained (Perkin–Elmer Lambda 4A) with a detection limit of 0.01 mg-N/L
from a deep well in Fountain Valley, California, while the low for both nitrate and nitrite (Method 4500-nitrate E of Standard
TOC water was collected from a well located in Tustin, Califor- Methods [52]). The concentration of dissolved oxygen was deter-
nia. Both groundwaters were collected in clear airtight containers, mined by the membrane electrode technique (Method 4500-O) as
transported to the laboratory, and stored at 20 ◦ C. The chemical described in Standard Methods [52], and the detection limit for
compositions and characteristics of the two groundwaters are pre- these measurements was 0.1 mg/L.
sented in Table 1. The Fountain Valley groundwater (FVGW) had a
nitrate level of approximately 0.3 mg/L and was low in total dis- 4.6. Ethanol analysis
solved solids (TDS), while the Tustin groundwater (TGW) had a
higher average nitrate concentration of 20 mg/L, and was high in The ethanol analysis was conducted by gas chromatography
TDS including sulfate, bicarbonate, chloride, calcium, sodium and using a Perkin-Elmer Sigma 2B gas chromatograph equipped with
magnesium ions. The FVGW had a significantly higher NOM con- a flame ionization detector (GC-FID), and a 6-foot glass column
tent than the TGW as reflected by the TOC and UVA254 (ultraviolet packed with 0.1% SP-1000 on 80/100 CarboPak C (Supelco Inc.,
absorbance at 254 nm) levels, and therefore had a higher poten- Bellefonte, Pennsylvania), and interfaced with a data processor.
tial for organic membrane fouling. The TOC and UVA254 levels for The instrument was operated under isothermal conditions with
the FVGW were 3.5–4.0 mg/L and 0.19–0.22 cm−1 , while the cor- injector, oven and detector temperatures of 200, 80 and 250 ◦ C,
responding values for TGW were below 0.1 mg/L and 0.005 cm−1 , respectively. The detection limit for this technique was 0.1 mg/L.
respectively. Obviously, the FVGW exhibited a higher potential for
membrane organic fouling of membranes. However, the TGW had 4.7. Alachlor analysis
a significantly higher TDS content than the FVGW, and therefore
manifested a higher potential for membrane inorganic fouling. The Alachlor was analyzed according to Method 6630 of Standard
FVGW was low in nitrate content, and so it was spiked with appro- Methods [52], wherein the samples were subjected to liquid–liquid
priate nitrate levels of 16–18 mg/L for all studies. Additionally, the extraction method using pentane. The extracts were analyzed by
FVGW was spiked with 99.7% pure alachlor (Fisher Scientific, North a Hewlett Packard 5790A series gas chromatograph with a 63 Ni
Kingstown, Rhode Island) at a concentration of 60–100 ␮g/L. In the electron capture detector (GC-ECD) (Hewlett Packard, Palo Alto,
present study, batch biokinetic studies and MBR studies evaluated California) and a 30 m × 0.53 mm DB-5 capillary column (Supelco,
the removals of target contaminants using the FVGW, whereas lim- Inc., Bellefonte, Pennsylvania). The GC-ECD system employed a car-
ited MBR tests investigated the permeate flux decline patterns for rier gas mixture of 95% argon and 5% methane at a flow rate of
both source waters. 3.1 mL/min. It was operated isothermally with injector, oven and
detector temperatures of 225, 240 and 300 ◦ C, respectively. The
4.2. Activated carbon adsorbent detection limit for alachlor using this technique was 1 ␮g/L.

The powder activated carbon (PAC) employed was Calgon WPH 4.8. Trihalomethane formation potential measurements
(Calgon Carbon Corp., Pittsburgh, Pennsylvania). The PAC was
washed several times with deionized-distilled water to eliminate The THMFP measurements were conducted according to
leachable organics, inorganics and fine particles. The carbon was Method 5710 B for chlorinated DBPs [52]. Samples were collected
46 V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54

Table 2
Conditions for the MBR mini-pilot-scale experiments.

Condition Run 0 Run 1 Run 2 Run 3 Run 4 Run 5 Run 6 Run 7 Run 8 Run 9

Water used Ultra-pure water FVGW FVGW FVGW FVGW FVGW FVGW TGW TGW TGW

Influent TOC (mg/L) 0 3.5 3.1 3.1 4.0 4.3 4.2 ∼0.1 ∼0.1 ∼0.1
Influent nitrate (mg/L) 0 17.2 17.4 17.8 16.6 17.8 17.2 17.2 17.5 17.4
Initial biomass (mg-MLSS/L) 0 – 343.3 342.6 – 343.6 360.8 – 344.1 343.5
PAC slurry (mg/L) 0 – – 500 – – 500 – – 500
Influent alachlor (␮g/L) 0 98.0 93.4 84.8 88.6 92.6 102 98 96 98
Influent virus (PFU/mL) × 105 – – – – 3.83 3.30 5.63 – – –

in duplicates in 125 mL borosilicate glass bottles with Teflon septa trations were in the range of 65–90 mg N/L [9]. The experiments
screw caps. These samples were buffered to pH 7.0 using a phos- were conducted at different temperatures in the 10–40 ◦ C range.
phate buffer and chlorinated at a 3-to-1 Cl2 -to-TOC ratio using a Samples were periodically taken and analyzed for nitrate, nitrite,
5 mg/mL Cl2 stock solution, and were subsequently incubated in ethanol and biomass concentrations.
the dark at 20 ◦ C for 24 h. After incubation, the chlorine residual
was tested and then quenched using sodium sulfite solution. The 4.12. Mini-pilot-scale MBR studies
extraction of THMs from aqueous samples was accomplished by a
liquid–liquid extraction method using high-purity pentane, and the Mini-pilot-scale MBR studies were conducted using two source
extracts were analyzed for THMs by gas chromatography accord- waters, namely, the FVGW and the TGW. A schematic of the exper-
ing to Method 6232 D [52]. The detection limit for THM formation imental mini-pilot-scale MBR process is presented in Fig. 1. The
potential measurements using this technique was about 0.1 ␮g/L. MBR system consisted of a completely stirred tank reactor, a tubu-
lar cross-flow membrane module, and a recirculation pump. The
membrane module consisted of six ceramic tubes (1 cm o.d.) of
4.9. Total organic carbon analysis
gamma-alumina (-Al2 O3 ) membranes having a nominal pore size
of 0.2 ␮m and a total area of 0.03 m2 . The MBR process was operated
The TOC analysis was conducted in accordance with Method
in a continuous mode, where the deoxygenated source water was
5310 B outlined in Standard Methods [52] using a Shimadzu Model
continually supplied to the reactor using feed pumps controlled by
TOC-5000 analyzer (Shimadzu Scientific Instruments, Columbia,
level switch within the reactor. The mixed liquor in the reactor was
Maryland). The TOC analyzer was quipped with a non-disperse
fed through the membrane lumen using recirculation pump where
infrared detector, and its detection limit was about 20 ␮g/L.
some of liquid passes through the membrane as product water,
and the membrane concentrate was recirculated. Ethanol was used
4.10. Virus concentration measurements as the electron donor for anoxic biological denitrification. A con-
stant TMP of 20 psi (1.4 bars) and cross-flow rate of 20 L/min was
The MS-2 virus measurements were based on a standard, used. The pH in bioreactor was monitored and adjusted as required.
double-layer agar assay technique with E. coli Famp (ATCC 15597 Samples from the permeate stream were periodically collected for
[51]) as the host. Experimental procedures relating to E. coli prepa- measuring the concentrations of nitrate, nitrite, ethanol, biomass,
ration and inoculation with MS-2 are described by Madaeni [53]. TOC, THMFP, alachlor, MS-2 virus, and dissolved oxygen, as well
The MS-2 concentrations were measured in terms of plaque counts as for determining the flow rates and permeate fluxes as functions
per unit volume or plaque formation units per milliliter (PFU/mL). of time. Nine sets of experiments were conducted, and their salient
All samples were assayed in triplicate for each dilution and exam- features are presented in Table 2: Runs 1–6 used the FVGW as water
ined. The average number of plaques present in the most countable source, whereas Runs 7–9 employed the TGW.
dilution was used to estimate plaque counts per unit volume.
5. Results and discussion
4.11. Batch biokinetic studies
5.1. Batch biokinetic optimization studies
Completely mixed batch bioreactor studies were performed in
a glass reactor with a capacity of 7 L for evaluating denitrifica- Biological denitrification is an oxidation–reduction reaction
tion kinetics. These experiments were conducted under different wherein microorganisms utilize carbon source as electron donor
initial C/N ratios, pH and temperatures to determine the optimal to reduce nitrate ions to nitrogen gas. As the groundwater to be
conditions, using the groundwater employed for the MBR stud- treated lacked sufficient amount of readily biodegradable organic
ies, per procedures described by Tsai [9] and Tsai et al. [10]. The constituents, ethanol was added as external carbon source to facil-
reactor system was equipped with a pH control electrode, a dis- itate the redox process, and the rationale for choosing ethanol was
solved oxygen electrode, an agitator, and a pump delivery system discussed earlier.
for acid, base, nutrients and feed water. The temperature was main- Batch biokinetic studies were conducted and the specific deni-
tained at a desired level via a built-in circulating water bath. At the trification rate (SDR) was estimated under different conditions such
start of each experiment, the acclimated denitrifying microbial cul- as temperature and pH for FVGW. The nitrate and ethanol removal
ture was added to the fermenter, the solution was deoxygenated efficiencies were determined as a function of the C/N ratio. The
with nitrogen gas to maintain anaerobic conditions, and the pH results of biokinetic studies are summarized in Fig. 3(a)–(c). Fig. 3(a)
controlled at a pre-determined level. The reactor mixing was pro- shows that the optimal initial C/N ratio was 0.8 mg-C/mg-N with the
vided by a paddle-blade rotator operated at 400 rpm. For each pH and temperature maintained at 7.8 and 20 ◦ C. Above this value,
set of experiments, only the tested parameter was varied within nitrate was completely removed, but ethanol residuals increased;
a desired range, while other conditions were constant. The dura- and below this value, the denitrification efficiency sharply declined.
tions of batch experiments were in the range of 100–540 min. The For instance, nitrate removal was only 74% corresponding to a C/N
biomass concentrations varied from 270 to 790 mg/L with average ratio of 0.5 mg-C/mg-N. The results showed that a shortage in sup-
values ranging from 390 to 590 mg/L [9]. The initial nitrate concen- ply of carbon source would limit denitrification efficiency, and an
V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54 47

excess in supply would decrease ethanol removal. Therefore, the


preliminary determination of the optimal C/N ratio was not strictly
based on SDR considerations. At the favorable C/N ratio of 0.8, more
biokinetic experiments were conducted to evaluate the effects of
temperature and pH on denitrification kinetics. Fig. 3(b) shows that
the SDR was only 0.36 mg-N/mg MLSS-day at 20 ◦ C while it was 1.24
mg-N/mg MLSS-day at 40 ◦ C. A temperature of 20 ◦ C was chosen
as a favorable value in view of economics based on reaction ener-
getics, compatibility with ambient conditions, and overall process
energy requirements. It was indeed uneconomical to conduct den-
itrification tests at 30 or 40 ◦ C owing to the energy need to increase
the temperature of the treated water to those levels. The advan-
tages of faster reaction kinetics were greatly offset by the higher
energy requirements. The results depicted in Fig. 3(c) indicate that
the optimal pH based on SDR considerations was 7.8, given that the
most optimal C/N ratio was 0.8 and the most favorable tempera-
ture was 20 ◦ C. In short, optimization of biological denitrification
process was not based on biokinetics alone, but also on energy
utilization as well as on nitrate and ethanol residuals. Fig. 4. Concentrations of nitrate, nitrite, ethanol and biomass as functions of time
in batch reactor studies for Fountain Valley groundwater.

Typical biokinetic data showing the biomass, nitrate, nitrite,


and ethanol concentration profiles as functions of time for FVGW
are presented in Fig. 4. These data indicate that biological den-
itrification occurred in a one-step reaction in which nitrate was
converted to nitrogen gas without significant nitrite accumulation.
The maximum nitrite concentration detected in the effluent was
0.06 mg-N/L, which was below the recommended MCL of 1 mg/L
NO2 -N. Meanwhile, ethanol was consumed, and the biomass con-
centration gradually increased. The biomass profile showed that
there was a growth phase of about 15 h during which the biomass
concentration increased from about 92 to 106 mg/L, which was fol-
lowed by a stationary phase between 15 and 20 h. Subsequently,
the biomass levels started decreasing as microbial lysis occurred.
The batch bioreactor studies facilitated process optimization
with respect to the C/N ratio, temperature and pH. The optimal
C/N ratio was 0.8 g C/g N-NO3 , and this value was employed in sub-
sequent MBR mini-pilot-scale studies. As biological denitrification
was constrained by temperature limitations due to reaction ener-
getics and compatibility with ambient conditions, a near-optimal
temperature of 20 ◦ C was preferred. In summary, the optimal
process conditions were as follows: C/N ratio of 0.8 C/g N-NO3 , tem-
perature of 20 ◦ C, and a pH of 7.8. Similar results were obtained from
biokinetic tests for TGW as reported by Tsai [9].

5.2. Mini-pilot-scale MBR studies

5.2.1. Removals of nitrate and nitrite


The nitrate and nitrite removals under different MBR process
conditions (Table 2) for the FVGW are presented in Fig. 5. These pro-
cess conditions included the application of biomass alone, biomass
with PAC, and direct membrane filtration without biomass or PAC
(to serve as control), as previously discussed. The PAC slurry was
added to the MBR system only at the commencement of operation,
and was not continuously employed throughout the experimen-
tal run. The MBR experiments showed that high nitrate removals
could be achieved using this technology, and that no differences
in removals were observed under either transient-state or steady-
state conditions for the two scenarios—biomass alone, and PAC with
biomass. A turning point was observed after the first 6 h when the
Fig. 3. Batch reactor biokinetic studies for denitrification of Fountain Valley ground- nitrate removals improved from about 86% to over 99% for Run
water (adapted from Tsai [9]): (a) removal efficiencies of nitrate and ethanol as a 5 (only biomass), and from 88% to almost 100% for Run 6 (com-
function of C/N ratio (optimal temperature 20 ◦ C and pH 7.8), (b) specific denitrifi-
bination of PAC and biomass). These results indicated that the
cation rate as a function of temperature (optimal pH 7.8 and C/N ratio 0.8), and (c)
specific denitrification rate as a function of pH (optimal temperature 20 ◦ C and C/N biomass was well acclimated for denitrification after 6 h of pro-
ratio 0.8). cess operation. Under either scenario (Run 5 or Run 6), the effluent
48 V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54

as evidenced from Figs. 6(b), 6(c), 7(b) and 7(c). After 35 h of MBR
operation with biomass alone, the ethanol TOC and non-ethanol
TOC removals were 100% and 70%, respectively (Fig. 6(b)); and, the
corresponding THMFP and alachlor removals were 61% and 35%
(Fig. 7(b)). A combination of PAC and biomass further increased the
removal efficiencies for non-ethanol TOC and THMFP to 84% and
98%, and those for ethanol TOC and alachlor to nearly 100%, respec-
tively, corresponding to a PAC dosage of 500 mg/L, as depicted in
Table 2 and Figs. 6(c) and 7(c).
A noteworthy feature of the results was that the resid-
ual ethanol (electron donor for biological denitrification) in the
effluent was below the detection limit of 0.1 mg/L. Further-
more, no non-ethanol TOC was observed due to metabolic end
products of ethanol biodegradation, and this was established
Fig. 5. Nitrate removal in the MBR system from Fountain Valley groundwater under by gas chromatography–mass spectrometry (GC–MS) analysis of
various MBR process conditions.

nitrate concentrations steadily decreased and approached almost


zero after 8 h. An important consideration was that no nitrite resid-
uals were observed in the treated MBR effluent. The contribution of
PAC sorption to nitrate removal was insignificant owing to negligi-
ble sorption of the species on activated carbon [9,10]. Nonetheless,
PAC was employed in the MBR process for other reasons. Firstly,
PAC could serve as a buffer to prevent the biological inhibition of
denitrification due to toxic organics that possess strong affinity for
carbon adsorption. Secondly, it could remove NOM, reduce DBP
formation potential (THMFP), and prevent microbial regrowth in
distribution systems (due to BOM removal). Thirdly, it was believed
that PAC could possibly ameliorate membrane fouling and perme-
ate flux decline [6].

5.2.2. Removals of NOM, TOC, alachlor and THMFP


The effects of membrane filtration as well as PAC and biomass
addition to the MBR process on the removals of various organic
components of interest as functions of time are presented in
Figs. 6 and 7. These components include ethanol TOC, non-ethanol
TOC (predominantly NOM), THMFP, and alachlor. As the total
TOC represents a combination of NOM and ethanol concentrations
(ethanol is the electron donor for biological denitrification), the
TOC attributable to NOM is referred to as non-ethanol TOC and is
distinguished from the TOC arising from ethanol.
Figs. 6(a) and 7(a) indicate that steady-state removals of ethanol
TOC and alachlor due to membrane separation alone (control runs
without PAC or biomass) were insignificant. In comparison, the
removals of non-ethanol TOC and THMFP were about 63% and 18%,
respectively; and these removals could be solely attributed to the
partial NOM rejection by the membrane. It must be noted that non-
ethanol TOC of the FVGW included relatively high molecular weight
NOM fractions. The characteristics of this water were reported by
Tan and Amy [54] who indicated that the relative distributions of
NOM molecular weight fractions (measured as TOC) were as fol-
lows: <0.5 kDa, 28%; 0.5–1 kDa, 18%; 1–5 kDa, 13%; 5–10 kDa, 21%;
10–30 kDa, 14%; and, >30 kDa, 6%. Their study reported that 41% of
the NOM (non-ethanol TOC) consisted of compounds whose molec-
ular weights were well above 5–10 kDa. Additionally, the MBR test
results showed that after the first hour, the rejection of non-ethanol
TOC by the membrane rapidly increased from 0 to 50%, and after
the second hour, increased to nearly 63%. This indicated that pro-
gressive gel formation and internal pore fouling possibly caused
a steady increase in NOM rejection. These observations were also
supported by the membrane fouling and flux decline patterns, dis-
cussed in a later section (Figs. 9 and 10; membrane permeate flux
data for FVGW when no PAC or biomass was used).
The MBR test results showed that the removals of all the com-
Fig. 6. Ethanol and non-ethanol TOC removals in the MBR system for Fountain Valley
ponents exhibited increases with the use of biomass alone, and groundwater: (a) no biomass or PAC (Run 4), (b) biomass alone (Run 5), and (c)
further improvements with a combination of biomass with PAC, biomass and PAC (Run 6).
V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54 49

the degradation products. Therefore, complete mineralization of


ethanol (electron donor) during biological denitrification was con-
cluded. This is an important factor because residuals of ethanol
and/or its organic metabolite(s) would lead to microbial regrowth
in distribution systems and increase in THMFP during chlorina-
tion. A comparison of results depicted in Fig. 6(a)–(c) led to certain
important observations. Membrane filtration alone accounted for
a non-ethanol TOC removal of 63%, while the addition of microbial
degradation improved the removal by over 7%, and a combina-
tion of PAC sorption and microbial degradation further enhanced
the removal by another 14%. In the best-case scenario (Fig. 6(c)),
the non-ethanol TOC level in the effluent was well below 0.7 mg/L
The high THMFP removal of 98% corresponding to an influent
level of 450 ␮g/L would reduce the residual level below 10 ␮g/L, Fig. 8. Virus MS-2 removals for Fountain Valley groundwater under various MBR
easily meeting the DBP control objective. The results depicted process conditions.
in Fig. 7(a)–(c) emphasize that PAC sorption and biodegradation

achieve higher alachlor removals. The use of biomass resulted in


alachlor removals of 36%, principally due to bacterial sorption,
while combination of PAC with biomass however increased the
removals to nearly 100%. These findings indicate that the MBR pro-
cess could achieve alachlor residuals well below drinking water
standards (0.2 ␮g/L) even at influent levels exceeding 100 ␮g/L.
The results of the present investigation can be compared to
those of an earlier study by Tian et al. [55] who employed a mem-
brane bioreactor process with PAC and biomass for NOM removal
from drinking water supplies. Tian et al. [55] observed dissolved
organic carbon (DOC) removals (non-ethanol TOC removals in our
context) of 19.4% and 37.5% when they employed an MBR system
with biomass alone and biomass and PAC, respectively. In our study,
the non-ethanol TOC removals corresponding to these two cases
were 70% and 84%, which are indeed significantly higher than those
reported by Tian et al. [55]. However, it must be noted that direct
comparison of our studies with those of Tian and coworkers would
not be appropriate owing to differences in material properties and
pore size distributions of the membranes employed, and in the
characteristics of the raw water treated. Tian et al. [55] employed
a polyvinyl chloride (PVC) ultrafiltration membrane of pore size
0.01 ␮m as compared to the present study that used a gamma-
alumina (-Al2 O3 ) microfiltration membrane of pore size 0.2 ␮m.
Additionally, the PVC membrane was generally more hydrophobic
than the gamma-alumina (-Al2 O3 ) membrane, and was therefore
more susceptible to organic fouling. Furthermore, the raw water
used by Tian et al. [55] had an NOM content of 5.6 mg/L as DOC,
6.2 mg/L as TOC, and 0.07 cm−1 as UVA254 (ultraviolet absorbance
at 254 nm), and a turbidity level of 2.4 NTU. In comparison, the
NOM content of the FVGW had a TOC level of 3.5–4.0 mg/L, and
a UVA254 level of 0.19–0.22 cm−1 , and a low turbidity of 0.1 NTU.
Nevertheless, the FVGW apparently consisted of higher molecular
weight NOM, as discussed earlier. Therefore, higher NOM removals
could be expected in our study as compared to those of Tian and
coworkers, due to the combination of factors stated above.

5.2.3. Removals of viruses


The MS-2 removal efficiencies were evaluated under the three
MBR process conditions (Table 2: Runs 4–6), and the experimental
results are presented in Fig. 8. The feed water spiked with MS-2
virus contained additional non-NOM type TOC (0.6–0.8 mg/L) from
the Tryptone broth used in its preparation. The MS-2 viral counts in
the influent were in the range (3.30–5.86) × 10+5 PFU/mL (Table 2).
In the control study (Run 4) with neither PAC nor biomass, MS-2
removal of about 3.7-logs was experienced after 2 h of process oper-
ation. When biomass alone was employed (Run 5), the removals
were initially low during the first 9 h of testing at 2.7-logs, and then
Fig. 7. Alachlor and THMFP removals in the MBR system for Fountain Valley ground-
water: (a) no biomass or PAC (Run 4), (b) biomass alone (Run 5), and (c) biomass gradually improved to 3.5-logs at steady state. When biomass and
and PAC (Run 6). PAC were employed in combination (Run 6), the MS-2 removals
50 V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54

were 3.6-logs after the first 10 h, and further improved to 3.8-logs.


These results indicated that certain mechanisms were operative in
MS-2 virus removals under the three testing conditions.
The MS-2 virus is relatively small (about 0.025 ␮m or 25 nm
in size and 3.6 × 106 Da in molecular weight), and its rejec-
tion by a 0.2-␮m pore size microfiltration membrane cannot be
solely attributed to mechanical sieving and size exclusion; other
mechanisms including membrane sorption, gel sorption, and cake
sorption could be operative. Initially, under the control scenario
(Run 4 with no biomass or carbon), the NOM in the water could lead
to gel formation and internal pore plugging, potentially enhancing
the membrane virus rejection. In Run 5, when biomass alone was
employed, the role of biomass was initially insignificant, because
it possibly hindered gel formation owing to some biodegradation
of the NOM that would have caused pore blocking. However, at
later stages, a combination of gel and biomass layers might have
improved virus rejection due to increased fouling and pore block-
ing. In Run 6, when biomass and PAC were used, the MS-2 virus Fig. 9. Permeate flux profiles for Fountain Valley and Tustin groundwaters under
rejection was more efficient possibly due to the PAC–biomass cake various MBR process conditions.
formation. Therefore, the following virus removal mechanisms
can be proposed: (i) rejection by NOM gel layer, (ii) rejection by
The highest initial flux decline rates were observed for the con-
microbial biomass layer, (iii) rejection due to internal pore block-
trol run with TGW where no PAC or biomass was employed (Run
ing by NOM molecules, (iv) sorption on the membrane surface,
7); the flux declined from 600 to 350 L/(m2 h) during the first 6 h
PAC, and/or bio-particles (groundwaters have low turbidity), and
and subsequently to 285 L/(m2 h) at steady state. When biomass
(v) a combination of these mechanisms. Nonetheless, a caveat
alone was employed without PAC, the flux declined from 150 to
must be added that these results are applicable to MS-2 or similar
77 L/(m2 h), initially, and further deteriorated to a steady-state
viruses and cannot be generalized to all strains owing to differ-
value of 60 L/(m2 h) after about 6 or 7 h (Run 8). When PAC was used
ences in sizes, shapes, morphologies, surface charge and adhesion
in combination with biomass, the initial flux declined from 300 to
characteristics.
160 L/(m2 h) within 6 or 7 h, and eventually reached 100 L/(m2 h) at
steady state (Run 9). The results for FVGW exhibited patterns qual-
5.2.4. Membrane permeate fluxes and flux decline patterns
itatively similar to those for TGW, although the fluxes were lower.
An important consideration in the operation of MBR processes is
When no biomass or PAC was employed, the flux for FVGW was ini-
effective control of membrane fouling and mitigation of permeate
tially 450 L/(m2 h), but declined to about 180 L/(m2 h) after about
flux decline. In order to evaluate the effects of PAC and microorgan-
7 h, and later reached a steady-state value of 160 L/(m2 h) (Run 1).
isms on membrane fouling and permeate flux decline, the flux data
When biomass alone was employed, the flux for FVGW decreased
were obtained for FVGW and TGW under three testing scenarios as
from an initial value of 160 to 80 L/(m2 h) after about 7 h, and further
previously described: (i) no biomass or carbon, (ii) biomass alone
declined to 60 L/(m2 h) at steady state (Run 2). With the applica-
without PAC, and (iii) combination of biomass and PAC. Analyses
tion of biomass and PAC, the flux for FVGW (Run 3) declined from
of permeate flux data generally facilitate comparisons of process
200 to 90 L/(m2 h) within the first 7 h, and further to 60 L/(m2 h) on
operating conditions, raw water quality and membrane character-
approaching steady state. Furthermore, the flux decline patterns
istics with regard to the nature of membrane fouling and patterns
were apparently unaffected by the MS-2 virus populations, as the
of flux decline. Additionally, flux decline patterns could suggest
fluxes for Runs 4, 5 and 6 (with viruses) were nearly identical to
possible fouling mechanisms and prescriptions for flux improve-
those for Runs 1, 2 and 3, respectively.
ment such as devising pre-treatment techniques, choosing cleaning
agents, developing cleaning procedures, and determining cleaning
frequencies. More importantly, these data can provide information
on the choice of membrane materials, pore characteristics, pore
size distributions, and modular configurations. The permeate flux
data and profiles for the aforementioned experiments presented in
Fig. 9 illustrate certain fouling and flux decline patterns. The per-
meate flux decline rate as a function of operation time is depicted
in Fig. 10. Qualitative and quantitative analyses of data presented in
Figs. 9 and 10 illustrate characteristic patterns of membrane fouling
and flux decline in three distinct stages: initial conditioning foul-
ing, rapid fouling and fast flux deterioration (under constant TMP),
and slow fouling and insignificant flux decline (under constant
TMP). These fouling patterns also reflected the differences between
the two source waters with regard to short-term and long-term
impacts. Generally, higher permeate fluxes were observed for TGW
versus FVGW under similar conditions, and this could be explained
by their NOM levels (TOC levels of <0.1 mg/L versus 3.5–4.0 mg/L
and UVA245 levels of <0.005 and 0.19–0.22 cm−1 ). The TGW with
significantly lower NOM content exhibited a lower organic fouling
potential. The NOM levels also influenced the membrane fouling
patterns from initial process operation (time zero) to subsequent Fig. 10. Permeate flux decline rates for the Fountain Valley and Tustin groundwaters
transient-state and approximate steady-state operations. under various MBR process conditions.
V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54 51

The permeate flux data illustrate the importance of initial mem- for both source waters when biomass or PAC or both are employed.
brane exposure to the feed water, and highlight that preliminary This is so because the initial fluxes are themselves lower for these
surface fouling or internal pore fouling could be substantial. In fact, cases as compared to filtration without biomass or carbon. The flux
the results (Fig. 9) indicate a lower preliminary membrane fouling decline patterns observed herein are relevant to hybrid MBR sys-
potential for TGW as compared to FVGW owing to its lower NOM tems that use microorganisms and adsorbents in water treatment
content. In the case where no biomass or PAC was employed (Runs 1 situations.
and 7), the initial permeate flux J0 for TGW was high at 600 L/(m2 h), The permeate fluxes for both waters (FVGW and TGW) under
and low at 450 L/(m2 h) for FVGW. In the second case where only the three previously described scenarios were investigated on the
biomass was employed (Runs 2 and 8), the initial fluxes for the basis of the resistance model to highlight three characteristic stages
TGW and FVGW were 150 and 120 L/(m2 h), respectively. In the of membrane fouling. The fluxes at times 0, 1 and 30 h (t = 0, 1 and
third case where both biomass and PAC were used (Runs 3 and 9), 30 h) as well as the corresponding total resistances RT are presented
the corresponding fluxes were 300 and 160 L/(m2 h), respectively. in Table 3. The selection of these times was based on membrane
These results suggested that under similar experimental conditions fouling and permeate flux decline patterns observed in this study.
the differences in initial permeate fluxes for the two source waters The fluxes at start (t = 0 h) illustrate the effect of membrane expo-
were attributable to their NOM levels. Furthermore, the initial foul- sure to feed in the first stage, wherein flux decline is very rapid.
ing during the first stage was the highest when biomass alone was The fluxes after 1 h (t = 1 h) represent the second stage of fouling,
used without PAC (Runs 2 and 8). wherein flux decline is not as fast as that experienced during the
A comparison of results (Fig. 9, Runs 3 and 7) showed that first hour. The fluxes after 30 h of operation (t = 30 h) represent the
without biomass or PAC the permeate fluxes were significantly third stage characterized by approach to steady sate with slow
higher for the TGW. When biomass was used alone (Runs 2 and fouling. The membrane permeate flux data for ultra-pure water
8), the fluxes were relatively higher for TGW. However, when both was also measured as a control for comparison with other oper-
biomass and PAC were employed (Runs 3 and 9), the transient-state ating conditions. The ultra-pure water was prepared by passing
fluxes were higher for TGW, although the steady-state fluxes were deionised-distilled water through a GAC column for the removal
nearly the same for both waters. These observations could again be or all trace organics (Run 0 in Tables 2 and 3). The flux decline
explained by the fact that the TGW had a lower NOM and hence patterns yield some interesting information regarding membrane
manifested a lower potential for initial gel formation and organic fouling rates. It is important to note that most of the fouling occurs
fouling. Nevertheless, this effect diminished as steady state was during the initial period of membrane exposure to the feed, a factor
approached and the fouling levels were similar for both waters. This indicated by the high rates of increase in resistance RT as a func-
aspect was illustrated by the flux data for two situations, namely, tion of time. The rates of change in RT become relatively slow after
operation without biomass or PAC, and that with PAC alone. Never- the first hour, and the fouling follows a slower pattern afterwards.
theless, when biomass and PAC were combined (Runs 3 and 9), the These results are substantiated by the permeate flux decline rates
differences in fluxes were not pronounced for both waters because at various times, namely t = 0, 1 and 30 h (Table 3).
most of the NOM was possibly sorbed by the PAC, and to some The total resistances RT (Eq. (3)) estimated from the experimen-
extent by the biomass. tal runs were of the range 0.66 × 10−13 to 8.26 × 10−13 m−1 and
The permeate flux decline patterns observed in this study their values were inversely related to the permeate fluxes, assum-
highlight some important comparisons between wastewater and ing a constant TMP of 20 psi (137.8 kPa or 1.4 atm or 1.4 bars) and a
water treatment scenarios using MBR systems, and the quali- liquid phase viscosity of 0.01 Poise. The resistance RT correspond-
tative/quantitative nature of membrane fouling. In wastewater ing to ultra-pure water was the intrinsic membrane resistance RM
treatment situations, the three stages of membrane fouling are of 0.66 × 10−13 m−1 . The RT value corresponding to TGW with-
characterized by initial membrane exposure and fouling, an inter- out using PAC or biomass at t = 0 (Run 7) of 0.83 × 10−13 m−1 is
mediate period of slow fouling, followed eventually by a period essentially a combination of RM and RI (resistance due to inor-
of fast or rapid fouling. However, in water treatment situations, ganic fouling) because the feed water has practically no NOM, but
membrane fouling is significantly lower with distinctly different has only inorganic foulants including silica and other salts and no
flux decline patterns. These differences in patterns can be rational- biomass. Therefore the resistance RI for TGW can be estimated at
ized on the basis of feed water quality and treatment conditions. t = 0 as 0.17 × 10−13 m−1 by noting the difference corresponding
In wastewater treatment scenarios, contaminant concentrations to ultra-pure water. In the case of FVGW, the inorganic con-
are significantly higher with high TOC levels of 100–1000 mg/L, stituents are low while the TOC level is high, so that by using
often using high biomass levels of 10,000–20,000 mg/L. In water the RT of 1.10 × 10−13 m−1 for t = 0 when no biomass or PAC is
treatment situations, contaminant concentrations are relatively used, and neglecting inorganic foulants, the organic fouling resis-
low with the TOC and biomass levels in the ranges of 0–10 and tance RO can be estimated by comparison with ultra-pure water
40–400 mg/L, respectively [6,8,10]. as 0.44 × 10−13 m−1 . Determining the RT values at t = 0 for the two
Membrane fouling and flux decline patterns are indeed better waters when biomass alone was used (Runs 2 and 8), and calculat-
reflected by the permeate flux decline rates (−dJ/dt) as a func- ing the differences in values with those for the no biomass or PAC
tion of the operation time t, as depicted in Fig. 10. Apparently, case (Runs 1 and 7), the biological fouling components RB for FVGW
the second stage of rapid or fast fouling occurs between 0.1 and and TGW can be estimated as 3.01 × 10−13 and 2.26 × 10−13 m−1 ,
1 h as flux decline rates decrease from 100–350 L/(m2 h2 ) to about respectively. These results support the view that the biological
2–5 L/(m2 h2 ); and later, during the third stage, the rates are fur- fouling contributions are slightly higher for FVGW during the ini-
ther reduced to 0–1 L/(m2 h2 ). These results indicate that most of tial start-up phase. With the addition of PAC to biomass, the total
the membrane fouling and flux decline occurred during the first 0.5 resistances for the FVGW and TGW decrease to 2.47 × 10−13 and
to 1 h of operation for both source waters in the three operational 1.65 × 10−13 m−1 , respectively, at t = 0 (Runs 3 and 9). Similarly,
scenarios previously described. The flux decline rates are maximum qualitative and quantitative assessments of resistances can be
at 60 L/(m2 h2 ) after 1 h for FVGW without biomass or PAC, and are made at various times such as t = 1 and 30 h, based on the results
relatively lower for the cases using biomass alone and biomass in presented in Tables 3 and 4.
combination with PAC. This is so because biological and organic The membrane fouling resistance RF was estimated for both
fouling occur during the initial phase of membrane exposure to the source waters (FVGW and TGW) at the operating times t of 0, 1
feed water. The flux decline rates during this initial period are lower and 30 h by deducting the membrane resistance RM from the total
52 V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54

Table 3
Permeate fluxes and total membrane transport resistances corresponding to the three stages of fouling.

Process conditions Run 0a Run 1a Run 2a Run 3a Run 7a Run 8a Run 9a


Source water

Ultra-pure FVGW FVGW FVGW TGW TGW TGW


Ultra-pure No biomass Biomass Biomass and No biomass Biomass Biomass
water or PAC alone PAC or PAC alone and PAC

Permeate flux at t = 0 h (L/(m2 h)) 750 450 120 200 600 160 300
Permeate flux at t = 1 h (L/(m2 h)) 750 265 70 95 500 110 180
Permeate flux at t = 30 h (L/(m2 h)) 750 120 60 65 230 60 75
Resistance, RT × 10−13 at t = 0 h (m−1 ) 0.66 1.10 4.11 2.47 0.83 3.09 1.65
Resistance, RT × 10−13 at t = 1 h (m−1 ) 0.66 1.86 7.05 5.20 0.99 4.49 2.74
Resistance, RT × 10−13 at t = 30 h (m−1 ) 0.66 4.11 8.27 7.59 2.15 8.27 6.58
a
Experimental run.

resistance RT . As mentioned earlier, the fouling resistances were too high at 0.7 because the PAC slurry was not continuously fed to
designated as RFW for operation with biomass alone without PAC, the MBR after the initial addition. It is nevertheless possible that a
and as RFP for the case with both biomass and PAC, and these values substantial portion of the biomass which was not attached to PAC
are presented in Table 4. In qualitative and quantitative terms, the remained in suspension and caused more membrane fouling.
RFW values were 30–50% lower than the RFP values, highlighting the
fact that PAC alleviates membrane fouling for both source waters 5.3. Some general observations on hybrid MBR technology for
by reducing the organic and biological fouling to some extent. Addi- water treatment
tionally, the RFP values at the three characteristic operating times
(t = 0, 1 and 30 h) suggest that PAC reduced the fouling resistance Several key issues must be addressed before the design and
to lower levels for TGW, because this water had a lower potential operation of a pilot-scale or full-scale MBR process is to be con-
for organic fouling due to NOM. A comparison of the RFW values templated upon. These issues include the use of electron donors
for both waters presented in Table 4 shows that fouling resistances for sustaining the denitrifying bacteria, the consideration of dis-
were higher for FVGW at t = 0 and 1 h, but almost identical at t = 30 h. solved oxygen (DO), and control of membrane fouling or permeate
These results demonstrated that during the first two stages of foul- flux decline.
ing, namely, initial fouling and rapid fouling phases, the fouling An electron donor such as ethanol was essential for biological
resistances were of higher magnitude for FVGW, but when the denitrification, and under ideal conditions the optimal C/N ratio
slow fouling third stage (approximate steady state condition) was was 0.8 g C/g N-NO3 , when neither carbon nor nitrogen was reac-
approached, the fouling was almost the same as that experienced tion limiting, complete denitrification was accomplished without
without PAC application. Nonetheless, when PAC was employed, nitrite or ethanol residuals, and faster reaction rate was achieved.
the fouling resistance was lower for TGW than for FVGW, and this is Nevertheless, consistent maintenance of optimal C/N ratios in pilot-
indeed an important factor regarding the function of PAC. A possible scale or full-scale systems would be problematic due to variations
explanation might be that owing to the higher NOM content (mea- in source water quality as well as precision and accuracy limitations
sured as TOC) in FVGW, the PAC reduced some gel formation and of fluid flow and chemical metering systems. Furthermore, ethanol
internal pore sorption attributable to NOM. The permeate flux data residuals in the MBR effluent could promote microbial regrowth
and the estimated transport resistances presented in Tables 3 and 4 and colonization in distribution systems.
illustrated certain qualitative aspects of membrane fouling. The Biological denitrification is an anoxic process wherein both reac-
fouling resistances were consistently higher for FVGW as com- tion rates and conversion efficiencies are adversely sensitive to
pared to the TGW owing to the higher NOM content of the former. DO levels. When excess oxygen is available, heterotrophic bacteria
Furthermore, the application of PAC increased the transient-state first utilize oxygen before nitrate as electron acceptor. Low process
fluxes but did not significantly improve the steady-state fluxes for efficiencies would require significantly larger reactor dimensions,
both source waters. These results highlighted that organic fouling greater residence times, larger footprints, and higher operation
due to NOM initially influenced the transient-state fluxes but did costs. This is an important factor because very high DO levels can
not significantly affect the steady-state fluxes. A plausible explana- only be lowered by chemical addition (such as thiosulfate) or gas
tion for this observation might be that biological fouling dominated purging, and chemical addition generally causes undesirably high
organic fouling during long-term operation, and that the initial TDS levels.
NOM fouling was superseded by biomass fouling at a later stage. Membrane fouling and permeate flux decline are important fac-
Additionally, the biomass-to-PAC ratio in this study was possibly tors in MBR process operations, and so laboratory-scale testing and

Table 4
Membrane transport resistance components.

Fountain Valley groundwater (FVGW)a Tustin groundwater (TGW)a

Time (h)

t=0 t=1 t = 30 t=0 t=1 t = 30

Intrinsic membrane resistance RM (m−1 ) 0.66 × 10−13 0.66 × 10−13 0.66 × 10−13 0.66 × 10−13 0.66 × 0−13 0.66 × 10−13
Inorganic fouling resistance RI (m−1 ) – – – 0.17 × 10−13 0.33 × 10−13 1.49 × 10−13
Organic fouling resistance RO (m−1 ) 0.44 × 10−13 1.20 × 10−13 1.94 × 10−13 – – –
Biological fouling resistance RB (m−1 ) 3.01 × 10−13 5.19 × 10−13 4.16 × 10−13 2.43 × 10−13 3.50 × 10−13 6.12 × 10−13
Fouling resistance without PAC RFW (m−1) 3.45 × 10−13 6.39 × 10−13 7.61 × 10−13 2.43 × 10−13 3.83 × 10−13 7.61 × 10−13
Fouling resistance with PAC RFP (m−1 ) 1.79 × 10−13 4.54 × 10−13 6.93 × 10−13 0.99 × 10−13 2.08 × 10−13 5.92 × 10−13
a
Source water.
V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54 53

analysis of membrane flux data under various scenarios is primor- [7] A.M. Barreiros, C.M. Rodrigues, J.P.S.G. Crespo, M.A.M. Reis, Membrane bioreac-
dial before pilot-scale and full-scale design are undertaken. This tor for drinking water denitrification, Bioprocess and Biosystems Engineering
18 (4) (1998) 297–302.
approach shall be important in evaluating membrane susceptibil- [8] B.O. Mansell, E.D. Schroeder, Biological denitrification in a continuous flow
ity to fouling, and shall play a significant role in selecting suitable membrane bioreactor, Water Research 33 (8) (1999) 1845–1850.
membrane material, type, configuration, and cleaning strategy. [9] H.H. Tsai, Membrane bioreactor process for the denitrification of groundwaters:
experiments and model verification, Doctoral dissertation, Department of Civil
and Environmental Engineering, University of Southern California, Los Angeles,
2003.
6. Summary and conclusions [10] H.H. Tsai, V. Ravindran, M.D. Williams, M. Pirbazari, Forecasting the per-
formance of membrane bioreactor process for groundwater denitrification,
The hybrid MBR process with PAC and microorganisms was Journal of Environmental Engineering and Science 3 (6) (2004) 507–521.
[11] K.S. Haugen, M.J. Semmens, P.J. Novak, A novel in situ technology for the treat-
effective in removing nitrate, NOM, THMFP, alachlor, and MS-2,
ment of nitrate contaminated groundwater, Water Research 36 (14) (2002)
meeting the overall treatment objectives. The technology appears 3497–3506.
to be ideally suited for small-scale systems such as well-head appli- [12] K. Kimura, M. Nakamura, Y. Watanabe, Nitrate removal by a combination of ele-
cations. mental sulphur based denitrification and membrane filtration, Water Research
36 (7) (2002) 1758–1766.
The alachlor removals could be attributed to the synergistic [13] V. Ravindran, B.N. Badriyha, M.D. Williams, S.C. Tu, M. Pirbazari, A hybrid
effect of PAC and microbial sorption. The MS-2 removals at steady membrane filtration process for the treatment of water contaminated with
state could possibly be attributed to the following mechanisms: (i) petroleum hydrocarbons, in: Proceedings of the Membrane Technology Confer-
ence, Reno, Nevada (August 13-16, 1995), American Water Works Association,
rejection by NOM gel layer, (ii) rejection by microbial biomass layer, Denver, Colorado, 1995, pp. 352–363.
(iii) rejection due to internal pore blocking by NOM molecules, (iv) [14] K. Gopal, S.S. Tripathy, J.L. Bersillon, S.P. Dubey, Chlorination
adsorption on the membrane surface, PAC and bio-particles, and byproducts, their toxicodynamics and removal from drink-
ing water, Journal of Hazardous Materials 140 (1–2) (2007) 1–
(v) combinations of these mechanisms. 6.
The MBR tests provided valuable information on membrane [15] A.C. Diehl, G.E. Speitel Jr., J.M. Symons, S.W. Krasner, C.J. Hwang, S.E. Barrett,
fouling and flux decline patterns that might prove beneficial in pro- DBP formation during chloramination, Journal of American Water Works Asso-
ciation 92 (6) (2000) 76–90.
cess design for specific applications. The studies demonstrated that [16] L.A. Croen, K. Todoroff, G.M. Shaw, Maternal exposure to nitrate from drinking
fouling occurred in three distinct stages: the initial period when water and diet for neural tube defects, American Journal of Epidemiology 153
fouling occurred due to the exposure of the membrane to the feed (4) (2001) 325–331.
[17] P.J. Weyer, J.R. Cerhan, B.C. Kross, G.L. Halberg, J. Kantemneni, G. Breuer, M.P.
with fast flux decline (during the first few minutes of process opera-
Jones, W. Zheng, C.F. Lynch, Municipal drinking water nitrate level and cancer
tion), the period of rapid fouling with substantial flux decline rates, risk in older women: the Iowa women’s health study, Epidemiology 11 (3)
and slow fouling with low flux decline rates. The permeate flux (2001) 327–337.
decline patterns observed were characteristic of a water treatment [18] Guidelines for Canadian Drinking Water Quality, 6th ed., Ministry of Health,
Ottawa, Canada, 1996.
scenario involving NOM, BOM, microbial biomass and colloids at [19] A. Nuhoglu, T. Pekdemir, E. Yildiz, B. Keskinler, G. Akay, Drinking water denitri-
low concentrations. fication by a membrane bioreactor, Water Research 36 (5) (2002) 1155–1166.
The use of the foulant classification method instead of the [20] A. Kapoor, T. Viraraghavan, Nitrate removal from drinking water: review, Jour-
nal of Environmental Engineering 123 (4) (1997) 371–380.
phenomenological classification approach in the resistance model [21] S.J. Ergas, D.E. Rheinheimer, Drinking water denitrification using a membrane
provided a more realistic portrayal of membrane fouling and flux bioreactor, Water Research 38 (14–15) (2004) 3225–3232.
decline in MBR processes for drinking water applications. [22] E.J. McAdam, S.J. Judd, A review of membrane bioreactor for nitrate removal
from drinking water, Desalination 196 (1–3) (2006) 135–148.
The permeate flux data from MBR studies showed that although [23] E.J. McAdam, S.J. Judd, Denitrification from drinking water using a membrane
PAC addition initially enhanced transient-state permeate fluxes, no bioreactor, Water Research 41 (18) (2007) 4242–4250.
significant improvement was observed in steady-state fluxes. [24] B.N. Badriyha, V. Ravindran, W. Den, M. Pirbazari, Bioadsorber efficiency,
design, and performance forecasting for alachlor removal, Water Research 37
(17) (2003) 4051–4072.
Acknowledgments [25] Drinking water regulations and health advisories, United States Environmental
Protection Agency, Washington, DC, and Lewis Publishers, Chelsea, Michigan,
1990.
The authors extend their appreciation to the Metropolitan [26] Special report on environmental endocrine disruption: an effects assess-
Water District of Southern California for providing in-kind con- ment and analysis, United States Environmental Protection Agency Report No.
EPA/630/R-96/012, Washington, DC, 1997.
tribution at various stages of this project. Any statements made
[27] D.W. Graham, M.K. Miley, F. Denoylles, V.H. Smith, E.M. Thurman, R. Carter,
or opinions expressed in this article are those of the individual Alachlor transformation patterns in aquatic field mecocosms under variable
authors and do not necessarily represent or reflect the views of oxygen and nutrient conditions, Water Research 34 (16) (2000) 4054–4062.
the agency. Any reference made in this publication to any specific [28] J. Qu, H. Li, H. Liu, H. He, Ozonation of alachlor catalyzed by Cu/Al2 O3 in water,
Catalysis Today 90 (3–4) (2004) 291–296.
method, product, process or service does not constitute or imply [29] R.J. Miltner, D.B. Baker, T.F. Speth, C.A. Fronk, Treatment of seasonal pesticides
endorsement, recommendation, or warranty thereof by the agency. in surface waters, Journal of American Water Works Association 81 (1) (1989)
43–52.
[30] M. Pirbazari, B.N. Badriyha, R.J. Miltner, GAC adsorber design for removal of
References chlorinated pesticides, Journal of Environmental Engineering 117 (1) (1991)
80–100.
[1] M/DBP Federal Advisory Committee. Stage 2M/DBP agreement in principle [31] X. Zheng, J. Liu, Mechanism investigation of virus removal in a membrane
(Draft), United States Environmental Protection Agency, Washington, DC, 1998. bioreactor, Water Science and Technology: Water Supply 6 (1) (2005) 51–59.
[2] National primary drinking water regulations: disinfectants and disinfection [32] X. Zheng, J. Liu, Virus rejection with two model human enteric viruses in mem-
by-products rule, Federal Register 63 (1998), 241, 69390, United States Envi- brane bioreactor system, Science in China, Series B: Chemistry 50 (3) (2007)
ronmental Protection Agency, Washington, DC. 397–404.
[3] I.C. Escobar, A.A. Randall, Assimilable organic carbon (AOC) and biodegrad- [33] R.M. Maier, I.L. Pepper, C.P. Gerba, Environmental Microbiology, Academic
able dissolved organic carbon (BDOC): complementary measurements, Water Press, San Diego, California, 2000.
Research 35 (18) (2001) 4444–4454. [34] T. Ueda, N.J. Horan, Fate of indigenous bacteriophage in a membrane bioreactor,
[4] I.C. Escobar, A.A. Randall, J.S. Taylor, Bacterial growth in distribution systems: Water Research 34 (7) (2000) 2151–2159.
effect of assimilable organic carbon and biodegradable organic carbon, Envi- [35] K. Farahbakhsh, D.W. Smith, Removal of coliphages in secondary effluent by
ronmental Science and Technology 35 (17) (2001) 3442–3447. microfiltration-mechanisms of removal and impact of operating parameters,
[5] S. Hong, I.C. Escobar, J. Hershey-Pyle, C. Hobbs, J.J. Cho, Biostability characteri- Water Research 38 (3) (2004) 585–592.
zation in a full scale nanofiltration/reverse osmosis (NF/RO) treatment system, [36] C. Shang, H.M. Wong, G.H. Chen, Bacteriophage MS-2 removal by submerged
Journal of American Water Works Association 97 (5) (2005) 101–110. membrane bioreactor, Water Research 39 (17) (2005) 4211–4219.
[6] M.D. Williams, M. Pirbazari, Membrane bioreactor process for removing [37] J. Radjenovic, M. Matosic, I. Mijatovic, M. Petrovic, Membrane bioreactor (MBR)
biodegradable organic matter from water, Water Research 41 (17) (2007) as an advanced wastewater treatment technology, Heidelberg Environmental
3880–3893. Chemistry 5 (S/2) (2008) 37–101.
54 V. Ravindran et al. / Journal of Membrane Science 344 (2009) 39–54

[38] C.C. Ho, A.L. Zydney, A combined pore blockage and cake filtration model for [47] S.C. Tu, V. Ravindran, M. Pirbazari, A pore diffusion transport model for fore-
protein fouling during microfiltration, Journal of Colloid and Interface Science casting the performance of membrane processes, Journal of Membrane Science
232 (2) (2000) 389–399. 265 (1–2) (2005) 29–50.
[39] C.C. Ho, A.L. Zydney, Protein fouling of asymmetric and composite microfiltra- [48] K.H. Choo, C.H. Lee, Membrane fouling mechanisms in the membrane-coupled
tion membranes, Industrial and Engineering Chemistry Research 40 (5) (2001) bioreactor, Water Research 30 (8) (1996) 1771–1780.
1412–1421. [49] M. Pirbazari, B.N. Badriyha, V. Ravindran, MF-PAC (microfiltration-powder acti-
[40] C.C. Ho, A.L. Zydney, Overview of fouling phenomena and modeling approaches vated carbon) for treating waters contaminated with natural and synthetic
for membrane bioreactors, Separation Science and Technology 41 (7) (2006) organics, Journal of American Water Works Association 84 (12) (1992) 95–103.
1231–1252. [50] M. Pirbazari, V. Ravindran, B.N. Badriyha, S.H. Kim, Hybrid membrane filtration
[41] M. Nystrom, K. Ruhomaki, L. Kaipia, Humic acid as a fouling agent in membrane process for leachate treatment, Water Research 30 (11) (1996) 2691–2706.
filtration, Desalination 106 (1–3) (1996) 79–87. [51] Catalogue of Bacteria Phages, American Type Culture Collection (ATCC), Man-
[42] W. Yuan, A. Kocic, A.L. Zydney, Analysis of humic acid fouling during micro- asas, Virginia, 1992.
filtration using a pore blockage-cake filtration model, Journal of Membrane [52] Standard Methods for the Examination of Water and Wastewater, 20th ed.,
Science 198 (1) (2002) 51–62. American Public Health Association, American Water Works Association/Water
[43] S. Ognier, C. Wisniewski, A. Grasmick, Characterization and modelling of fouling Environment Federation, Washington, DC, 1998.
in membrane bioreactors, Desalination 146 (1–3) (2002) 141–147. [53] S.S. Madaeni, Mechanism of virus removal using membranes, Filtration and
[44] S. Ognier, C. Wisniewski, A. Grasmick, Membrane bioreactor fouling in sub- Separation 34 (1) (1997) 61–65.
critical conditions: a local critical flux concept, Journal of Membrane Science [54] L. Tan, G.L. Amy, Comparing ozonation and membrane separation for color
229 (1–2) (2004) 171–177. removal and disinfection by-product control, Journal of American Water Works
[45] A. Pollice, A. Brookes, B. Jefferson, S. Judd, Sub-critical flux fouling in membrane Association 83 (5) (1991) 74–79.
bioreactors: a review of recent literature, Desalination 174 (3) (2005) 130–221. [55] J.Y. Tian, H. Liang, Y.L. Yang, S. Tian, G.B. Li, Membrane adsorption bioreac-
[46] J. Zhang, H.C. Chua, J. Zhou, A.G. Fane, Factors affecting the membrane perfor- tor (MABR) for treating slightly polluted surface water supplies: as compared
mance in submerged membrane bioreactors, Journal of Membrane Science 284 to membrane bioreactor (MBR), Journal of Membrane Science 325 (1) (2008)
(1–2) (2006) 54–66. 262–270.

You might also like