You are on page 1of 9

Journal of Membrane Science 382 (2011) 41–49

Contents lists available at ScienceDirect

Journal of Membrane Science


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

Impact of membrane configuration on fouling in anaerobic membrane bioreactors


I. Martin-Garcia a , V. Monsalvo b , M. Pidou c , P. Le-Clech d , S.J. Judd a , E.J. McAdam a,∗ , B. Jefferson a
a
Cranfield Water Science Institute, Building 39, Cranfield University, Bedfordshire MK43 0AL, UK
b
Dpto. Ingeniería Química, Universidad Autónoma de Madrid, Madrid 28049, Spain
c
Advanced Water Management Centre, The University of Queensland, Brisbane, Australia
d
UNESCO Centre for Membrane Science and Technology, School of Chemical Engineering, The University of New South Wales, Sydney, Australia

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

Article history: The filtration performance of flocculated and granulated configured anaerobic membrane bioreactors
Received 14 April 2011 (MBR) treating domestic wastewater has been evaluated and compared to conventional aerobic MBR.
Received in revised form 19 July 2011 Immersed hollow fibre (HF) and external tubular membrane geometries were additionally compared
Accepted 26 July 2011
with the latter operated in both pumped and gas-lift mode. After 200 d of operation, both granular
Available online 4 August 2011
and flocculated anaerobic MBR (AnMBR) suspensions were characterised by an increased population
of colloidal particles while the aerobic MBR retained a unimodal particle size distribution with a d50 of
Keywords:
20 ␮m. Consequently, the flocculated AnMBR supernatant was characterised by a soluble microbial prod-
Bubbling
Scouring
uct (SMP) concentration ca. 500% higher than the aerobic MBR, such that the lowest critical fluxes for
Critical flux both HF and tubular membranes were recorded for the AnMBR. In comparison, the granulated AnMBR
Crossflow sludge was characterised by a low mixed liquor suspended solids concentration and an SMP concen-
Submerged tration below 50% that of the flocculated anaerobic MBR. Consequently, similar fluxes to those of the
Sidestream aerobic MBR were achieved with the granulated anaerobic sludge using immersed HF membranes. Oper-
ating external tubular membranes in gas-lift appeared less effective for the granular AnMBR than the
Aerobic MBR. However, critical fluxes >40 L m−2 h−1 were achieved using pumped mode. Results suggest
granular AnMBR systems to be most suited to domestic wastewater treatment using either immersed HF
membranes or external tubular membranes in pumped crossflow mode.
© 2011 Elsevier B.V. All rights reserved.

1. Introduction in external tubular membranes operated under pumped [2–5] or


gas-lift [6–10] condition. While gas sparging in immersed mem-
Anaerobic treatment of domestic wastewater is constrained by brane systems is known to present the most cost effective means
the low organic strength of the wastewater, the quality of the avail- of fouling control in aerobic wastewater treatment applications,
able carbon and the high Ks value associated with the anaerobic permeability data for anaerobic immersed flat-sheet and hollow
community which impair bacterial growth and effective treatment. fibre (HF) geometries is more limited [11–14]. Comparisons of liq-
Anaerobic membrane bioreactor (AnMBR) technology can enhance uid pumping and gas sparging in external tubular membranes [15]
effluent quality from increased rejection of solids and colloidal and gas sparging in immersed flat-sheet and HF membranes have
organic matter, and also achieve higher biomass concentrations by been undertaken in AnMBR [11]; however, a direct comparison of
minimising washout. AnMBR thus increases the viability of anaer- external and immersed configured AnMBRs has yet to be made.
obic treatment, offering treatment at reduced energy demand and Direct comparison of current AnMBR studies also remains
sludge yield over that of aerobic processes, for which the aeration challenging as two reactor configurations now predominate: con-
energy demand is significant. However, AnMBR studies undertaken ventional completely mixed flocculated reactors [8–10] and upflow
to date have generally found that membrane fouling is significant anaerobic sludge blanket (UASB) reactors [10,12,14]. In the latter
due to the high concentration of colloidal material, inclusive of pro- case, membrane filtration is sited at the head of the UASB or in
tein and polysaccharide, the concentrations of which are further a subsequent stage, thus the membrane is challenged with the
exacerbated at extended solids retention times (SRTs) [1]. sludge supernatant. While UASB based AnMBR are less studied,
Previous AnMBR fouling studies have primarily investigated lower irreversible fouling than completely mixed reactors has been
the effect of pumped crossflow velocity on filtration performance demonstrated in a comparative study on the treatment of black-
water [10]. The aim of this present study is therefore to provide a
direct comparison of external and immersed membrane operation
∗ Corresponding author. Tel.: +44 1234 754546. in both flocculated and UASB configured AnMBR treating domestic
E-mail address: e.mcadam@cranfield.ac.uk (E.J. McAdam). wastewater. Previous AnMBR studies primarily utilised synthetic

0376-7388/$ – see front matter © 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.memsci.2011.07.042
42 I. Martin-Garcia et al. / Journal of Membrane Science 382 (2011) 41–49

Table 1
Module characteristics for the tubular and hollow fibre membranes tested.

Parameter Unit Experimental reactor MBR pilot plant

Pilot plant AeMBR AnMBR G-AnMBR


Geometry Tubular HF HF HF HF

Filtration area m2 0.022 0.93 12.5 12.5 0.93


Material – PVDF PVDF PVDF PVDF PVDF
Pore size ␮m 0.03 0.04 0.08 0.08 0.04
Module length m 1 0.7 1 1 0.7
Fibre OD/LD mm 8 1.9/0.8 1.3 1.3 1.9/0.8
CSA m2 – 0.0074 0.0177 0.0177 0.0074
Packing density m2 /m3 – 300 710 710 300

ID, internal diameter; OD, outer diameter; LD, lumen diameter; HF, hollow fibre.

wastewaters comprising of soluble carbon sources [9,11,14]. While sentative for the mean residence time in the anaerobic system since
such studies present significant insight, real wastewaters present no sludge accumulation was observed. For the AeMBR, which was
more complex colloidal and particulate matrices. Furthermore, to started with a low MLSS concentration of 1 g MLSS L−1 , the dynamic
date most published studies have used an external thermal source SRT reached a value of 80–85 d for the last 100 d of operation tak-
to stabilise reactor operation to mesophilic or thermophilic condi- ing into account biomass accumulation within the reactor for the
tions [2–4]. Energetic modelling has demonstrated that methane estimation of the sludge age.
generated from AnMBR treating domestic wastewater is insuffi- The granular anaerobic MBR (G-AnMBR) consisted of a 85 L per-
cient to support heating the influent wastewater due to the low spex cylindrical vessel (1.75 m depth × 0.25 m diameter) seeded
organic strength and high liquid flow rate [16]. Consequently, with 40 L of granular sludge from an UASB sited at a sugar refinery
to provide environmentally relevant conditions, the flocculated (British Sugar, Suffolk, UK), with a 38 L perspex cylinder housing
AnMBR, granular AnMBR (G-AnMBR) and a reference aerobic MBR the membrane (Fig. 1). The granular UASB contactor was operated
(AeMBR) were operated on real domestic wastewater and without in the expanded mode (EGSB) using an external recirculation pump
external temperature control. (620s, Watson Marlow, Falmouth, UK) to maintain a superficial
upflow velocity (Vup ) of less than 1 m h−1 . Effluent from the granule
2. Materials and methods contactor was recirculated through the membrane tank during per-
meation. At the fixed upflow velocity, the granule bed expanded to
2.1. Pilot plants a depth of 0.60 m, or approx. 30% of the column depth; the effluent
entering into the membrane tank was characterised by a relatively
Three MBR comprising suspended growth (flocculated) aer- low solids fraction compared to the flocculated reactors. A PVDF
obic and anaerobic MBR and a granular anaerobic MBR hollow-fibre module with a surface area of 0.93 m2 and nominal
were operated in parallel for 250 d fed with wastewater pore size 0.04 ␮m, 1 m length, cross sectional area of 0.0074 m2
having mean chemical oxygen demand (COD), biochemical and a packing density of 300 m2 m−3 was used in this study at a set
oxygen demand (BOD5 ), suspended solids (SS) and ammonia instantaneous flux (J) of 6 L m−2 h−1 to provide a biotank hydraulic
concentrations of 338 mg COD L−1 (range 197–553 mg COD L−1 ), retention time (HRT) of 16 h. In the G-AnMBR membrane fouling
167 mg BOD L−1 (range 155–285 mg BOD L−1 ), 84 mg SS L−1 (range was controlled by gas sparging nitrogen at gas velocities ranging
51–186 mg SS L−1 ) and 35 mg NH4 -N L−1 (range 15–48 mg NH4 - from 0 to 0.057 m s−1 . During the 250-d trial, no granular biomass
N L−1 ) respectively. The AeMBR and AnMBR pilot plants comprised was withdrawn from the biotank; samples were only collected from
two tanks of 1.5 m3 total volume divided between the biological the membrane tank for analysis.
(1 m3 ) and membrane (0.5 m3 ) compartments. Wastewater was
introduced in the biological tank through a floating valve which 2.2. Short-term fouling experiments
controlled the level of sludge at a height of 1.5 m, creating a total
working volume of 1.2 m3 . The AeMBR was continuously aerated Short term fouling experiments for the AeMBR and AnMBR
at a flow rate of 50–100 L min−1 through cylindrical fine bubble were conducted in an external 38 L cylindrical tank (0.20 m diam-
diffusers at the biological tank base. The AnMBR tank was sealed eter × 1.2 m height) using a 30 L slurry from either reactor (Fig. 2).
with a PVC lid. Biomass was cycled between both compartments For immersed trials, an identical PVDF HF membrane to that used
through external pumps, and an additional pump was operated in the granular AnMBR was tested (Table 1). A 0.022 m2 tubu-
in cycles of 15 min on/15 min off to mix the reactor contents by lar PVDF membrane with a nominal pore size of 0.03 ␮m was
recycling biomass through venturi nozzles located at the base of used for external (sidestream) experiments. The external tubu-
this chamber. Both flocculated MBRs were fitted with polyvinyld- lar membrane was operated in both gas lift and pumped mode.
ifluoride (PVDF) HF membranes of 12.5 m2 surface area, 0.08 ␮m During immersed and gas lift operation, nitrogen-enriched air for
nominal pore size, 1 m length, 0.0177 m2 cross sectional area and anaerobic experiments or natural air for aerobic experiments were
a packing density of 710 m2 m−3 (Table 1). Permeate was continu- employed for membrane scouring and biomass mixing. For gas
ously extracted using peristaltic pumps (620 Du, Watson Marlow, lift experiments, the base of the membrane inlet was connected
Falmouth) to provide an instantaneous flux of 6 L m−2 h−1 . Mem- through a ‘T’ junction to provide a port for gas injection. The same
brane fouling was controlled by continuous gas sparging at gas range of specific gas demand, 0.2–1.2 m3 m−2 h−1 , was used for
velocities ranging from 0.02 m s−1 to 0.078 m s−1 in both aerobic both membrane geometries. For pumped crossflow operation, a
and anaerobic systems, the latter employing nitrogen enriched air centrifugal pump (EHEIM Gmbh and Co., Deizisau, Germany) was
(>99% N2 ) generated from an industrial gas membrane nitrogen employed to generate a cross flow velocity (CFV) controlled by a
generator unit (Atlas Copco, Herts, UK) fed with compressed air. throttling valve sited upstream of the membrane module to val-
During the 250-d trial, 12 L of sludge per day were withdrawn from ues between 0.4 and 2 m s−1 . For all studies both the retentate
the suspended aerobic and anaerobic MBRs in order to set SRT based mixed liquor and permeate product streams from the membrane
on bioreactor volume at 100 d. This value could be taken as repre- module were returned to the tank to maintain a constant mixed
I. Martin-Garcia et al. / Journal of Membrane Science 382 (2011) 41–49 43

Fig. 1. Schematic representation of the MBR pilot plants. The aerobic MBR (AeMBR, top) as well as the anaerobic MBR (AnMBR, middle) and granular anaerobic MBR (G-
AnMBR, bottom) which comprised of separate membrane and biological tanks. In all cases membrane pumps were employed to cycle sludge between both chambers and in
the case of the suspended AnMBR to mix reactor contents. Nitrogen gas was used as the sparge gas for anaerobic environments.

liquor suspended solids (MLSS) concentration. The transmembrane change in transmembrane pressure (dP/dt) monitored. The criti-
pressure (TMP) in the immersed system was recorded from pres- cal flux was estimated as the highest flux at which no increase in
sure transducers (Druck Ltd., Leicester) installed in the permeate transmembrane pressure was observed (dP/dt = 0) [8]. The sensi-
line. For the tubular membrane, additional pressure transducers tivity of the pressure-transducer used in this study resided below
were installed upstream and downstream of the retentate chan- 0.25% of range in accordance with externally validated calibra-
nel to provide accurate TMP determination. The pressure output tion. Therefore while the ‘strict form’ of critical flux was applied,
signal (4–20 mA) was logged via a 16-bit analogue to digital con- a reading of absolute zero (dP/dt = 0) did not guarantee zero depo-
verter (Pico Technology Ltd., St Neots, Cambridgeshire) onto a sition but rather the point of minimum deposition. Between each
personal computer. Fouling propensity was evaluated by compar- critical flux test, test membranes were cleaned overnight with
ing fouling rates and critical fluxes obtained from the flux-step 1% sodium hypochlorite and the clean water permeability (K)
method [17], in which membrane flux was incrementally increased measured to evaluate permeability recovery prior to the next
in steps of 2–4 L m−2 h−1 with step durations of 15 min and the trial.
44 I. Martin-Garcia et al. / Journal of Membrane Science 382 (2011) 41–49

10
(a) AeMBR

Volume fraction (%)


8
200 days
6 30 days

0
0.1 1 10 100 1000 10000

Particle size, dP (µm)

10
(b)

Volume Fraction (%)


AnMBR
8
200 days
6
30 days
Fig. 2. Experimental reactor used for short-term experiments. Two membrane con-
figurations were used; an externally configured (side-stream) tubular membrane 4
and an immersed hollow-fibre membrane. Nitrogen gas was used as the sparge gas
for anaerobic sludge during immersed and tubular gas-lift membrane trials. 2

0
0.1 1 10 100 1000 10000
2.3. Analytical procedures
Particle size, dP (µm)
Suspended solids, MLSS, mixed liquor volatile suspended solids
(MLVSS) and BOD5 were measured according to Standard Meth-
ods [18]. Ammoniacal nitrogen (NH4 -N) and COD were analysed 10
(c) G-AnMBR
using Merck pre-prepared cell tests (VWR International, Lutter-
Volume fraction (%)

8 200 days
worth, UK). Particle size was determined using an integrated laser
diffractor by recirculating suspensions through the optical unit of 30 days
6
a Malvern Mastersizer 2000 (Malvern Instrument Ltd., Worcester- Feed
shire, UK), yielding floc sizes as d10 , d50 and d90 values respectively 4
defining the equivalent diameter at which 10%, 50% and 90% of the
2
suspended particles are smaller. A coefficient of variance (CV) of
0.02 was determined for a fixed particle size of 0.5 ␮m (Latex micro- 0
spheres, Duke Scientific, US). Sludge supernatant was obtained by 0.1 1 10 100 1000 10000
centrifuging the biomass at 10,000 × g for 15 min followed by filtra-
Particle size, dP (µm)
tion of decanted liquid with a 1 ␮m rated filter [1]. Serial molecular
weight fractionation of the supernatant was conducted using a
Fig. 3. Particle size distributions following 30 d and 200 d operation of the (a) aerobic
400 mL Amicon 8400 dead-end stirred cell (Millipore, USA) with MBR (AeMBR), (b) anaerobic MBR (AnMBR) and (c) granular anaerobic MBR (G-
76 mm diameter polyethersulphone UF membranes of nominal AnMBR). Particle size distribution of the feedwater is included for reference (c).
molecular weights 10, 100, and 500 kDa (Millipore, USA). Frac-
tionations were conducted at a constant shear (through magnetic removal was consistently >95%. Both anaerobic systems presented
stirring) under an applied pressure of 2 Barg and a filtrate/retentate similar behaviour in terms of treatment performance, with COD
ratio of 0.4, using laboratory grade nitrogen gas (>99%). Protein and removal increasing from 80% initially following inoculation with
carbohydrate concentrations were determined according to the no acclimation up to 90% by the end of the study. Effluent BOD
respective adapted methods of [19] and [20]. Bovine serum albumin ranged between 5 mg BOD L−1 and 15 mg BOD L−1 , yielding a sta-
(BSA) and d-glucose were used as standards for the determination ble 90–95% removal. In the G-AnMBR, the Vup was maintained
of protein and carbohydrate respectively. At a fixed concentration below 1 m h−1 to reduce the concentration of solids in the over-
of 20 mg L−1 , CV of <0.01 and 0.02 were determined for d-glucose flow of the biological tank and subsequently minimise solids
and BSA respectively. loading onto the membrane tank. However, solids concentra-
tion in the membrane tank gradually increased from an initial
3. Results and discussion concentration of approximately 0.1 g SS L−1 , commensurate with
the feedwater, up to 0.6 g MLSS L−1 by the end of the study. In
3.1. Overall bioreactor performance the seeded AnMBR, while MLSS concentration ranged between
6.6 g MLSS L−1 and 9.6 g MLSS L−1 , the MLSS concentration sta-
The AnMBR was operated at a HRT and sludge retention time bilised at 7.7 g MLSS L−1 .
(SRT) of 16 h and 100 d respectively (Table 2). The G-AnMBR
operated at the same HRT but infinite SRT, with no wasting of 3.2. Evolution of particle size distribution
the granular biomass during the 250-d trial. Influent wastewater
consisted of settled sewage. Particle size in the feed wastewa- The particle measurement after 30 d of operation indicated
ter was characterised by a d50 of 91 ␮m. In the AeMBR, started unimodal distributions for both AeMBR and AnMBRs with corre-
without external seed material, COD removal improved from 85% sponding d50 values of 32 ␮m and 43 ␮m respectively (Fig. 3 and
to over 95% as MLSS concentration increased from 1 g MLSS L−1 Table 2). In contrast, a bimodal distribution was recorded for the
to the final average concentration of 8.7 g MLSS L−1 , while BOD sludge in the G-AnMBR membrane chamber with principal peaks
I. Martin-Garcia et al. / Journal of Membrane Science 382 (2011) 41–49 45

Table 2
Summary of operational parameters, bioreactor performance and bulk characterisation data from the AeMBR, AnMBR and G-AnMBR.

Parameter Unit AeMBR AnMBR G-AnMBR

Bioreactor operation and treatment performance


SRT d 100 100 –
HRT h 16 16 16
COD removal % 95 84 86
BOD removal % 97 93 95
Bulk characterisation
MLSS g l−1 8.7 7.7 0.1–0.6
PSD
d10 ␮m a
11.8 /5.0 b
15.1 /1.4 a b
1.1a /0.2b
d50 ␮m 32.4a /14.0b 42.8a /7.6b 60.3a /1.4b
d90 ␮m 69.2a /34.7b 91.2a /19.9b 181.9a /91.2b
SMPCOD mg l−1 99 ± 28 598 ± 150 198 ± 73
SMP fractionation
<1 ␮m–500 kDa % 43 87 –
500–100 kDa % 2 0 –
100–10 kDa % 2 4 –
<10 kDa % 53 9 –
SMPc mg l−1 18 ± 9 47 ± 14 18 ± 12
SMPp mg l−1 18 ± 13 108 ± 27 50 ± 21
SMP (P/C) – 1.0 2.3 2.8
a
Particle size recorded at 30 d operation.
b
Particle size recorded at 200 d operation.

at 138 ␮m and 0.6 ␮m, similar to the feedwater which was char- 108 ± 27 mg SMPP L−1 and 47 ± 14 mg SMPC L−1 . Speciation of the
acterised by peaks at 138 ␮m and 1.1 ␮m. The similarity can be supernatant produced from the G-AnMBR indicated lower organic
attributed to the limited solids retention exhibited by the fluidised concentrations in the SMPCOD , SMPP and SMPC fractions compared
granular bed. to the flocculated AnMBR at 198 ± 73 mg L−1 , 50 ± 21 mg L−1
Differences in transient aggregate size were observed between and 18 ± 12 mg L−1 respectively. Huang et al. [21] identified that
the AeMBR and AnMBR systems. The quasi-normal distribution ini- by operating anaerobic MBR at long SRT, higher protein and
tially recorded at 30 d in the AeMBR was maintained at 220 d of carbohydrate SMP concentrations developed which resulted in
operation, while the median size decreased from a d50 of 32 ␮m increased fouling. The SMP values in this study are higher than
to 14 ␮m. While SRT had not stabilised at 30 d of operation, this the range of 145–150 mg SMPCOD L−1 reported previously for the
finding is in agreement with the study of Huang et al. [21] where treatment of low-strength synthetic domestic wastewaters in both
lower aggregate sizes have been recorded in anaerobic MBRs oper- flocculated and granular AnMBRs [11,14]. However, since both
ated at longer SRT. Following 200 d of flocculated AnMBR operation, studies extracted sludge supernatant by filtration with 0.45 ␮m
a bimodal distribution was clearly demonstrated with peaks cen- membranes the SMP concentration may have been lower than for
tred at 11.5 ␮m and 1.7 ␮m particle size, the latter being indicative this study due to the higher size exclusion offered by the filter.
of colloidal generation linked to either floc erosion or poor floc- Further speciation of the SMPCOD fraction by serial fractionation
culation [22]; the accumulation of high molecular weight organic indicated that while for the AnMBR greater than 80% of SMP
degradation by-products could therefore partially contribute to the was between 1 ␮m and 500 kDa (Table 2), for the AeMBR only
skew observed. The bimodal particle size distribution recorded in approximately 43% was present in that fraction, further supporting
the G-AnMBR also indicated lower aggregate sizes; the distribu- the reported particle size distribution data.
tion recorded at 220 d operation was dominated by a peak of >5%
centred at 1.7 ␮m. However, the higher colloidal content in the 3.4. Critical fluxes in gas lift tubular membranes
AnMBR and G-AnMBR in comparison to the AeMBR commonly
derives from the higher degree of dispersive growth, frequently The injection ratios (ε) obtained from measuring liquid dis-
reported in anaerobic systems [15,23] and is associated with high placement by gas lift for the different gas sparging rates (Table 3)
fouling propensities [12,15,24]. A recent study comparing the per- passed through minima of 0.26 and 0.18 for the AeMBR and AnMBR
formance of a VFA and glucose/VFA-fed anaerobic MBR revealed the sludge respectively, corresponding to a superficial gas velocity (VG )
latter to provide a higher fouling propensity as a result of the higher of 0.02 m s−1 after which they increased to maxima of 0.43 and
growth of single cell acidogenic bacteria, was evidenced by a reduc- 0.64 for VG = 0.21 m s−1 . Injection ratios in the range 0.2–0.9 corre-
tion in measured particle size and supported by light microscopy
observations [8].
Table 3
Impact of gas velocity (VG , m s−1 ) on liquid velocity (VL , m s−1 ) in the externally
3.3. SMP concentration and composition configured (sidestream) tubular membrane during filtration of the AeMBR, AnMBR
and G-AnMBR sludge.

The concentration of soluble microbial products (SMP) VG (m s−1 ) AeMBR AnMBR G-AnMBR
determined in the supernatant as SMPCOD , SMP pro- VL (m s−1
) ε a
VL (m s−1
) εa
VL (m s−1 ) εa
tein (SMPP ) and SMP carbohydrate (SMPC ) in the AeMBR
0.02 0.07 0.22 0.09 0.18 0.26 0.07
averaged 99 ± 28 mg SMPCOD L−1 , 18 ± 13 mg SMPP L−1 and
0.05 0.13 0.28 0.12 0.29 0.36 0.12
18 ± 9 mg SMPC L−1 respectively (Table 2). The recorded SMPCOD 0.09 0.20 0.31 0.17 0.35 0.50 0.15
data are in agreement with previously reported results for 0.14 0.21 0.39 0.21 0.40 0.62 0.18
an AeMBR operated under similar operational conditions 0.18 0.25 0.42 0.24 0.43 0.71 0.20
0.21 0.28 0.43 0.12 0.64 0.35 0.38
(15 h HRT) and wastewater characteristics [25]. Data from
the suspended growth AnMBR were 598 ± 150 mg SMPCOD L−1 , a
ε = VG /(VG + VL ).
46 I. Martin-Garcia et al. / Journal of Membrane Science 382 (2011) 41–49

25 (a) 50
Fouling rate, dP/dt (mbar min-1)

AeMBR
VG (m s-1) 0.02 0.05 0.09 0.14
20 40 AnMBR

Critical flux (L m -2 h-1)


AnMBR
G-AnMBR G-AnMBR
15 AeMBR 30

10 20

5 10

0 0
0 5 10 15 20 25 30 35 0 0.5 1 1.5 2 2.5

Membrane flux, J (l m-2 h-1) Cross-flow velocity, CFV (m s -1)

Fig. 4. Impact of increased gas velocity (VG 0.02–0.14 m s−1 ) on fouling rate (b) 50
(dP/dt, mbar min−1 ) under tubular gas-lift mode at assigned membrane fluxes of AeMBR
4–28 L m−2 h−1 . Experiments conducted between 180 d and 240 d operation.
40 AnMBR

Critical flux (L m -2 h-1)


G-AnMBR
spond to slug-flow [26]. For the G-AnMBR sludge, the superficial 30
liquid velocity (VL ) induced by VG was approximately 300% of the
VL recorded for flocculated sludges; for a VG of 0.02 m s−1 , a VL 20
of 0.26 m s−1 was recorded for the G-AnMBR sludge compared to
0.07 m s−1 and 0.09 m s−1 for the AeMBR and AnMBR respectively.
10
Consequently for the G-AnMBR sludge a ε below 0.20 was cal-
culated for VG < 0.18 m s−1 , corresponding to bubble-flow which
0
provides less intense membrane scouring than slug flow [26].
0 4000 8000 12000 16000
Analysis of fouling rates (dP/dt) obtained during flux step
experiments and for a VG of 0.02–0.21 m s−1 indicated that when Reynolds number, Re
applying gas lift to tubular membranes, critical flux (Jc ) is more
Fig. 5. Impact of increased (a) crossflow velocity (CFV 0.4–2 m s−1 ) or (b) Reynold’s
influenced by sludge character than by gas sparging intensity; number (Re 500–15,000) on attainable critical flux (Jc ) during filtration in the
Jc in the AeMBR, AnMBR and G-AnMBRs remained at 8, 4 and externally configured tubular membrane operated in pumped mode. Experiments
4 L m−2 h−1 respectively independent of VG (Fig. 4). However, while conducted between 180 d and 240 d operation.
Jc remained relatively constant, a reduction in dP/dt was observed
from 11.8 mbar min−1 to 0.26 mbar min−1 for the AnMBR sludge as
reported for a tubular gas-lift membrane with a VFA and glucose
VG increased from 0.02 to 0.21 m s−1 at a flux (J) of 8 L m−2 h−1 . Since
feed respectively at VG values higher than those used in the cur-
shear-induced diffusivity is a less effective back-transport mecha-
rent study [15]. In this study, the colloidal nature of fouling in an
nism for low radius particles [27,28], it is postulated that the limited
anaerobic gas-lift MBR was corroborated by replacing the sludge
influence of VG on improving Jc arises from the contribution of col-
supernatant with tap water which resulted in a increase of Jc from
loid and soluble organics to fouling for gas-lift operation. Colloids
10 to 20 L m−2 h−1 . This result is consistent with the highest J being
are predominantly controlled by Brownian motion; consequently
obtained for the AeMBR, with a corresponding lowest concentra-
while a high shear environment is created by slug-flow, colloids
tion of soluble and colloidal organics, while the anaerobic systems
and soluble organics continue to preferentially migrate toward the
presented the highest SMP concentration and demonstrated the
membrane surface. Similarly, permeate flux has been observed to
lowest Jc . Earlier studies demonstrated significant flux promotion
be dependent only on applied TMP when increasing VG by an order
in AnMBRs through increasing VG [6,23]. An increase in J from 25 to
of magnitude from 0.001 m s−1 during ultrafiltration of 260 kDa
100 L m−2 h−1 following increasing VG from 0.2 m s−1 to 8.9 m s−1
molecular weight Dextran solutions [29]. Gas sparging has been
has been reported [23], albeit an appreciable decline in flux at
shown to have no effect on permeability of an air lift AeMBR due
higher gas sparging intensities attributed to the absence of a pro-
to the higher contribution of internal fouling (77%) compared to
tective cake layer which allows deposition of small particles on
cake layer formation (23%) [26]. The reduction in dP/dt recorded for
the membrane surface [31]. These studies employed ceramic mem-
the AnMBR upon increasing VG in the current study does however
branes, known to provide a higher organic fouling resistance than
indicate that, at high SMP concentrations, high shear rates partially
the polymeric materials used in the present study [5,32].
attenuate colloidal deposition.
Previous research has demonstrated that decreasing biomass
concentration in anaerobic MBRs can increase flux and reduce the 3.5. Critical fluxes in pumped tubular membranes
gas sparging demand [10,30]. However, in this study the lower
MLSS and SMP concentrations of the G-AnMBR sludge did not On using fluid pumping to generate the CFV in the tubular mem-
present higher Jc compared to the AnMBR system. The limited brane, Jc increased from 4 to 41 L m−2 h−1 in the G-AnMBR and from
impact of VG in the granular system can be attributed to the more 4 to 19 L m−2 h−1 in the suspended AnMBR for a CFV increase from
disperse gas phase at the lower ε values recorded, creating limited 0.4 to 2 m s−1 (Fig. 5). Traditionally, gas bubbles have been thought
turbulence at the membrane wall. Interestingly the dP/dt observed to enhance permeation by introducing localised liquid flow and
for J > Jc was significantly lower for the G-AnMBR compared to shear transients [1,33]. However, in this study, increasing CFV had a
the AnMBR: for a J of 11–12 L m−2 h−1 , the AnMBR dP/dt ranged greater impact on permeability than gas sparging in anaerobic sys-
between 8 and 25 mbar min−1 compared to 1–2 mbar min−1 for the tems. At CFV > 0.8 m s−1 , the Jc attained for the AeMBR was between
G-AnMBR. Similar AnMBR Jc values of 8 and 4 L m−2 h−1 have been that of the two anaerobic MBR, with a maximum Jc of 29 L m−2 h−1
I. Martin-Garcia et al. / Journal of Membrane Science 382 (2011) 41–49 47

at the highest CFV of 2 m s−1 . At a CFV of 0.4 m s−1 in the AeMBR


system, a Jc of approx. 8 L m−2 h−1 was attained, similar to the Jc
recorded for the lowest VG of 0.02 m s−1 . This suggests that although
the influence of VG on dP/dt is not substantial, the shear intro-
duced at the membrane surface by gas bubble flow even at a low VG
(ε > 0.22) can incur a considerable shear effect in the AeMBR. Sim-
ilar trends have been reported elsewhere [34,35], although in this
case increasing VG was shown to increase Jc . It should be noted that
for both anaerobic sludges, Jc values equivalent to those obtained
in gas-lift more were measured at the lowest CFV of 0.4 m s−1 . It
has further been illustrated that at high CFV the back-transport
velocity of sub-micron particles is adequately described by both
shear-induced diffusion and Brownian diffusion [36]. Where the
main reactors to have been operated as tubular MBR rather than
HF-MBR, the colloidal fraction in the flocculated systems may have
been more dominant due to the augmented shear [37]. However,
this study does illustrate that increased CFV can provide enhanced
suppression of concentration polarisation [23] in colloid dominated
systems.
The impact of CFV on Jc was more prominent for the G-AnMBR
(Fig. 5). The higher concentration of colloidal foulants in the G-
AnMBR with respect to the AeMBR may be compensated by the
higher shear resulting from its lower viscosity. Comparison of filtra-
tion performance between G-AnMBR and AnMBR systems operated
at low MLSS concentrations indicated that for a CFV of 1 m s−1 , sim-
ilar J values of 50 L m−2 h−1 were obtained [4,38]. In this study,
based on a viscosity of 1 mPa s−1 for the G-AnMBR supernatant,
a Re range from 3200 to 16,000 corresponds to CFV from 0.4 to
2 m s−1 . In the AeMBR and AnMBR systems, based on approxi-
mated rheological parameters [39,40], the Re ranges from 900 to
9000. Consequently, when comparing system efficacy based on
normalised turbulence (Re, Fig. 5), higher Jc values are obtained in
the AeMBR than the G-AnMBR: increasing Re from 3200 to 6170 and
from 3350 to 6400 in the G-AnMBR and AeMBR, respectively pro-
vided increases in Jc from 4 to 12 L m−2 h−1 and 16 to 29 L m−2 h−1 Fig. 6. Impact of specific gas demand per unit membrane area (SGDm
in the AeMBR. 0.19–1.16 m3 m−2 h−1 ) on attainable critical flux (Jc ) for AeMBR (a), AnMBR (b) and
G-AnMBR sludge (c) with an immersed hollow-fibre membrane. Experiments con-
ducted between 120 d and 150 d operation.

3.6. Critical fluxes in immersed HF membranes


ever, comparative specific energy demands for the immersed and
An increase in gas intensity from 0.007 to 0.027 m s−1 (N2 gas pumped crossflow configurations of 0.3 kWh m−3 and 3.7 kWh m−3
flow rate [m3 s−1 ] normalised to diffuser surface area in [m2 ]) of permeate produced respectively indicate that immersed mem-
resulted in similar increases in Jc from 3 L m−2 h−1 to 14 L m−2 h−1 branes present the most economically appropriate means for shear
in the AeMBR and G-AnMBR respectively. Previous studies have induction.
reported a threshold gas sparging rate after which no further The gas flows applied in tubular and HF geometries yielded
improvement in J is obtained [28,41]. In this study, the thresh- similar specific gas demands (SGDm , the gas flow rate [m3 h−1 ]
old was demonstrated for the G-AnMBR at a gas intensity of normalised against membrane surface area [m2 ]) ranging from 0.2
0.041 m s−1 , corresponding to a flux of 14 L m−2 h−1 . However, for to 1.6 m3 m−2 h−1 . Comparison of results between the two gas-
the gas intensities studied, a threshold was not observed for the membrane systems indicate that at low SGDm values of 0.2 and
AeMBR, with the highest Jc of 17 L m−2 h−1 recorded at the max- 0.4 m3 m−2 h−1 , the tubular module produced the higher Jc of 4 and
imum aeration studied (0.057 m s−1 ). An increase in gas intensity 8 L m−2 h−1 in the AeMBR and AnMBR respectively (Fig. 6). At low
from 0.007 to 0.041 m s−1 in the AnMBR yielded a small increase in applied gas velocities, fouling by aggregates dominates over solu-
Jc from 2 L m−2 h−1 to 5 L m−2 h−1 . These results indicate that the G- ble/colloidal fouling. Consequently, gas sparging is more effective
AnMBR system requires approximately 50% lower gas demand than in tubular membranes at the lower velocities due to the slug flow
the AnMBR to achieve an equivalent Jc . Furthermore, at higher gas regime induced, whereas in HF systems the energy introduced with
intensities, a 200% increase in Jc can be achieved using G-AnMBR bubbling is dissipated across the receiving water column. However,
rather than AnMBR. Similar findings have been reported from com- above this SGDm threshold, the Jc attainable in flocculated systems
paring AnMBR and G-AnMBR for blackwater treatment [10], where are higher with the immersed HF module. For the G-AnMBR, the HF
it was asserted that a higher permeability could be achieved due to membrane yielded a higher Jc over the entire SGDm range tested.
the lower colloidal concentration present in the granular reactor. The limited impact of gas-lift shear in the low solids environment
Similar assumptions regarding the impact of the low viscosity on provided by the G-AnMBR sludge supports the concept of solids-
increased shear intensity arise for gas sparged systems; the approx- dominated convective transport at low velocities. Previous reports
imate velocity gradient associated with membrane scouring (G) have indicated the diminishing effect observed following increased
ranges from 75 to 200 s−1 and from 87 to 208 s−1 in the aerobic and gas sparging in tubular membranes to be due to the increased fre-
anaerobic MBRs respectively, while for the same range of scouring quency of gas slugs which has a slightly deleterious impact on the
rate in the G-AnMBR, G varied between 230 s−1 and 670 s−1 . How- dynamic turbulence [29]. In contrast, the requirement for higher
48 I. Martin-Garcia et al. / Journal of Membrane Science 382 (2011) 41–49

SGDm in HF systems is intuitive since the hydrodynamic environ- [7] T. Imasaka, H. So, K. Matsushita, T. Furukawa, N. Kanekuni, Application of
ment is poorly defined. However, the greater effect of gas sparging gas–liquid two-phase cross-flow filtration to pilot-scale methane fermenta-
tion, Drying Technol. 11 (1993) 769–785.
on Jc obtained with HF membranes suggests that mechanisms such [8] D. Jeison, J.B. van Lier, Cake formation and consolidation: main factors
as fibre movement, which physically removes the mass transfer governing the applicable flux in anaerobic submerged membrane bioreac-
boundary layer and induces liquid flows transverse to the fibres tors (AnSMBR) treating acidified wastewaters, Sep. Purif. Technol. 56 (2007)
71–78.
[33], may be more significant in removing colloidal fouling than [9] Z. Huang, S.L. Ong, H.Y. Ng, Feasibility of submerged anaerobic membrane biore-
the shear induced in the liquid membrane interface by gas bubbling actor (SAMBR) for treatment of low-strength wastewater, Water Sci. Technol.
inside the tubular membrane. 58 (2008) 1925–1931.
[10] E. van Voorthuizen, A. Zwijnenburg, W. van der Meer, H. Temmink, Biological
black water treatment combined with membrane separation, Water Res. 42
4. Conclusions (2008) 4334–4340.
[11] A.Y. Hu, D.C. Stuckey, Treatment of dilute wastewaters using a novel submerged
anaerobic membrane bioreactor, J. Environ. Eng. 132 (2006) 190–198.
The filtration performance of a flocculated AnMBR and granular [12] H.J. Lin, K. Xie, B. Mahendran, D.M. Bagley, K.T. Leung, S.N. Liss, B.Q. Liao, Sludge
AnMBR treating domestic wastewater has been evaluated and com- properties and their effects on membrane fouling in submerged anaerobic
membrane bioreactors (SAnMBRs), Water Res. 43 (2009) 3827–3837.
pared to a conventional aerobic MBR. The flocculated AnMBR sludge [13] C. Wen, X. Huang, Y. Qian, Domestic wastewater treatment using an anaero-
was characterised by a similar MLSS concentration to an aerobic bic bioreactor coupled with membrane filtration, Process Biochem. 35 (1999)
MBR, but with approximately 500% greater supernatant colloidal 335–340.
[14] L.B. Chu, F.L. Yang, X.W. Zhang, Anaerobic treatment of domestic wastewater
fraction. In comparison, the granulated AnMBR sludge was char-
in a membrane-coupled expended granular sludge bed (EGSB) reactor under
acterised by a low MLSS concentration and, while the supernatant moderate to low temperature, Process Biochem. 40 (2005) 1063–1070.
colloidal fraction was higher than the aerobic MBR, the SMP con- [15] D. Jeison, J.B. van Lier, Thermophilic treatment of acidified and partially
acidified wastewater using an anaerobic submerged MBR: factors affecting
centration was below 50% of the flocculated AnMBR. The externally
long-term operational flux, Water Res. 41 (2007) 3868–3879.
configured tubular membranes operated in gas-lift mode provided [16] I.G. Martin, M. Pidou, A. Soares, S. Judd, B. Jefferson, Modelling the energy
equivalent critical fluxes for both the granular and flocculated demands of aerobic and anaerobic membrane bioreactors for wastewater treat-
AnMBR, though the fouling rate (dP/dt) was markedly lower for ment, Environ. Technol. 32 (9) (2011) 921–932.
[17] P. Le-Clech, B. Jefferson, I.S. Chang, S.J. Judd, Critical flux determination by the
the granular sludge. In contrast, when operating the tubular geom- flux-step method in a submerged membrane bioreactor, J. Membr. Sci. 227
etry in pumped flow mode, considerably higher critical fluxes were (2003) 81–93.
attainable with the granular AnMBR. At a CFV > 0.8 m s−1 , critical [18] American Public Health Association (APHA), Standard Methods for the Exami-
nation of Water and Wastewater, 20th ed., American Public Health Association,
fluxes achieved with the granular AnMBR exceeded those for both Washington, DC, USA, 1998.
aerobic and anaerobic flocculated MBR, suggesting that the shear [19] O.H. Lowry, N.J. Rosebrough, A.L. Farr, R.J. Randall, Protein measurement with
conditions created were effective in limiting concentration polari- the Folin phenol reagent, J. Biol. Chem. 193 (1951) 265–275.
[20] M. Dubois, K.A. Gilles, J.K. Hamilton, P.A. Rebers, F. Smith, Colorimetric method
sation within colloid dominated matrices. Higher critical fluxes of for determination of sugars and related substances, Anal. Chem. 28 (1956)
approximately 200% were also achieved for the granular anaerobic 350–356.
MBR with an immersed hollow fibre membrane, compared to the [21] Z. Huang, S.L. Ong, H.Y. Ng, Submerged anaerobic membrane bioreactor for
low-strength wastewater treatment: effect of HRT and SRT on treatment per-
anaerobic flocculated matrix. The higher fluxes measured for the
formance and membrane fouling, Water Res. 45 (2011) 705–713.
granular anaerobic MBR during gas sparging of immersed mem- [22] B.-M. Wilén, B. Jin, P. Lant, The influence of key chemical constituents in acti-
branes or pumped crossflow of external membranes reflects the vated sludge on surface and flocculating properties, Water Res. 37 (2003)
2127–2139.
impact of lower viscosity on shear. These results indicate that gran-
[23] T. Imasaka, N. Kanekuni, H. So, S. Yoshino, Cross-flow filtration of methane
ular rather than flocculated AnMBR systems appear most suitable fermentation broth by ceramic membranes, J. Ferment. Bioeng. 68 (1989)
for domestic wastewater treatment. However, comparative specific 200–206.
energy demands for the immersed and pumped crossflow config- [24] E.J. McAdam, S.J. Judd, E. Cartmell, B. Jefferson, Influence of substrate on fouling
in anoxic immersed membrane bioreactors, Water Res. 41 (2007) 3859–3867.
urations of 0.3 kWh m−3 and 3.7 kWh m−3 of permeate produced [25] M. Speı̌randio, A. Masse, M.C. Espinosa-Bouchot, C. Cabassud, Characterization
respectively indicate that immersed membranes present the most of sludge structure and activity in submerged membrane bioreactor, Water Sci.
economically attractive option for granular anaerobic systems. Technol. 52 (2005) 401–408.
[26] I.S. Chang, S.J. Judd, Air sparging of a submerged MBR for municipal wastewater
treatment, Process Biochem. 37 (2002) 915–920.
[27] E.J. McAdam, S.J. Judd, Optimisation of dead-end filtration conditions for an
Acknowledgements
immersed anoxic membrane bioreactor, J. Membr. Sci. 325 (2008) 940–946.
[28] E.J. McAdam, E. Cartmell, S.J. Judd, Comparison of dead-end and continuous
This research project has been supported by a Marie Curie Early filtration conditions in a denitrification membrane bioreactor, J. Membr. Sci.
Stage Research Training Fellowship of the European Community’s 369 (2011) 167–173.
[29] Z.F. Cui, S.R. Bellara, P. Homewood, Airlift crossflow membrane filtration — a
Sixth Framework Programme under contract number MEST-CT- feasibility study with dextran ultrafiltration, J. Membr. Sci. 128 (1997) 83–91.
2005-021050. The authors would also like to thank our industrial [30] D. Jeison, J.B. van Lier, Cake layer formation in anaerobic submerged membrane
sponsors Severn Trent Water and Yorkshire Water for their finan- bioreactors (AnSMBR) for wastewater treatment, J. Membr. Sci. 284 (2006)
227–236.
cial and technical support. Dr. I.G. Martin is now at the University [31] J. Lee, W.Y. Ahn, C.H. Lee, Comparison of the filtration characteristics between
of Valencia, Spain. attached and suspended growth microorganisms in submerged membrane
bioreactor, Water Res. 35 (2001) 2435–2445.
[32] I.J. Kang, S.H. Yoon, C.H. Lee, Comparison of the filtration characteristics of
References organic and inorganic membranes in a membrane-coupled anaerobic biore-
actor, Water Res. 36 (2002) 1803–1813.
[1] S. Judd, The MBR Book: Principles and Applications in Water and Wastewater [33] Z.F. Cui, S. Chang, A.G. Fane, The use of gas bubbling to enhance membrane
Treatment, Elsevier Science, Amsterdam, 2006. processes, J. Membr. Sci. 221 (2003) 1–35.
[2] A. Beaubien, M. Bâty, F. Jeannot, E. Francoeur, J. Manem, Design and operation of [34] P. Le-Clech, B. Jefferson, S.J. Judd, A comparison of submerged and
anaerobic membrane bioreactors: development of a filtration testing strategy, sidestream tubular membrane bioreactor configurations, Desalination 173
J. Membr. Sci. 109 (1996) 173–184. (2005) 113–122.
[3] K.H. Choo, C.H. Lee, Membrane fouling mechanisms in the membrane-coupled [35] M. Pidou, S.A. Parsons, G. Raymond, P. Jeffrey, T. Stephenson, B. Jefferson, Foul-
anaerobic bioreactor, Water Res. 30 (1996) 1771–1780. ing control of a membrane coupled photocatalytic process treating greywater,
[4] S. Elmaleh, L. Abdelmoumni, Cross-flow filtration of an anaerobic Water Res. 43 (2009) 3932–3939.
methanogenic suspension, J. Membr. Sci. 131 (1997) 261–274. [36] E. Tardieu, A. Grasmick, V. Geaugey, J. Manem, Hydrodynamic control of biopar-
[5] W.R. Ghyoot, W.H. Verstraete, Coupling membrane filtration to anaerobic pri- ticle deposition in a MBR applied to wastewater treatment, J. Membr. Sci. 147
mary sludge digestion, Environ. Technol. 18 (1997) 569–580. (1998) 1–12.
[6] E. Kayawake, Y. Narukami, M. Yamagata, Anaerobic digestion by a ceramic [37] J. Ho, S. Sung, Effects of solid concentrations and cross-flow hydrodynamics on
membrane enclosed reactor, J. Ferment. Bioeng. 71 (1991) 122–125. microfiltration of anaerobic sludge, J. Membr. Sci. 345 (2009) 142–147.
I. Martin-Garcia et al. / Journal of Membrane Science 382 (2011) 41–49 49

[38] B.D. Cho, A.G. Fane, Fouling transients in nominally sub-critical flux [40] A. Pevere, G. Guibaud, E. van Hullebusch, P. Lens, Identification of rheological
operation of a membrane bioreactor, J. Membr. Sci. 209 (2002) 391– parameters describing the physico-chemical properties of anaerobic sulphido-
403. genic sludge suspensions, Enzyme Microb. Technol. 40 (2007) 547–554.
[39] G. Laera, C. Giordano, A. Pollice, D. Saturno, G. Mininni, Membrane bioreac- [41] E.H. Bouhabila, R.B. Aïm, H. Buisson, Microfiltration of activated sludge using
tor sludge rheology at different solid retention times, Water Res. 41 (2007) submerged membrane with air bubbling (application to wastewater treat-
4197–4203. ment), Desalination 118 (1998) 315–322.

You might also like