Sei sulla pagina 1di 10

SUBMERGED ANAEROBIC MEMBRANE BIOREACTORS (SAMBR):

READY FOR THE BIG BALL ?

F. Fawehinmi*, B. Jefferson*, Tak Chan** and F.Rogalla**

*Cranfield University, www.cranfield.ac.uk,


Center for Water Sciences, Building 39, Beds. MK43 0AL

**Black & Veatch, www.bv.com,


69 London Road, Redhill, RH 1 1 LQ, UK , rogallaf@bv.com

ABSTRACT

Conventional aerobic membrane bioreactors (MBR) have been applied in thousands of


smaller installations, but their high energy cost to overcome membrane fouling limits their
application on large scale. A more sustainable concept was tested with a pilot-scale anaerobic
membrane bioreactor of 40L (10 gal), continuously fed with raw municipal sewage for a
period of 120 days. The reactor integrated two side-stream membrane modules (0.175 and
0.5 m2 = 2 and 5 ft2) and one submerged membrane (using hollow fibres with a total
filtration area of 0.9 m2 = 10 ft2), allowing both independent and simultaneous use of each
configuration.

The bioreactor was inoculated with crushed granular sludge and operated at three different
temperatures (35°C, 22°C and 12°C). After start-up and acclimation, the biomass yield was
observed to be constant and very low at 0.003 g MLVSS/g COD⋅d, independently of the
operating temperature. With influent values between 250 and 600 mg COD/l, the highest
COD removal was recorded at 35°C with values of 97%, but a stable effluent with COD
values below 50 mg/L was also produced at temperatures as low as 12°C.

During stable operation, membrane fouling rates for an MLSS concentration around 6 g/l
varied between 9 mbar/d (0.13 psi/d) and 10 mbar/d (0.15 psi/d) at fluxes from 9 to 22 L/m2
h (5 to 13 gfd), which is comparable to aerobic MBR. Doubling the membrane gas sparging
with nitrogen reduced membrane fouling rates by a factor three, as a linear relationship
between these two parameters was observed. While typical aerobic MBRs are operated at a
gas sparging to membrane flux ratio of 30 to 50 (both in both in l/m2/h or gfd), the gas to
permeate ratio necessary to clean the anaerobic membranes was at least 150 times less.

KEYWORDS

Anaerobic membrane bioreactor, COD removal, membrane fouling, methane


production, biomass yield, gas sparging
INTRODUCTION

Aerobic Membrane Bioreactors (MBR) are applied in more than two thousand installations
worldwide, mostly for small building or industrial applications. Only a dozen of installations
have flows larger than 10 000 m3/d (2.5 mgd), as the energy needs of MBR, both to scour the
membranes and to overcome the low oxygen transfer efficiency in the highly concentrated
mixed liquors, remains around 1 kwh/m3 (3.8 kwh/1000 gal - Guihe et al, 2004). This is at
least three times the 0.3 kwh/m3 (1.1 kwh/1000 gal) obtained by other advanced processes
(Barjenbruch, 2006), even when using compact reactors (Hansen et al, 2006).

Conceptually, adding membranes to anaerobic reactors could allow to overcome their


weakness of limited removal efficiency, especially at lower temperatures (Foresti, 2002),
while at the same time maintaining their advantages of lower solids production and
favourable energy and carbon balance (Lettinga et al, 1999).

Submerged anaerobic membrane bioreactors (SaMBR) have proven to be feasible on lab


scale (3 l (0.1 gal) reactors) with mesophilic temperatures and synthetic substrate (Hu and
Stuckey, 2006), achieving removals of >90% soluble COD at a hydraulic retention time
(HRT) as low as 3 hr. Both hollow fibre and flat sheets were used with membrane fluxes up
to 15 l/m2/h (9 gfd) while using biogas to sparge the membranes.

Other recent experiments on a submerged anaerobic membrane bioreactor showed


disappointing results both on treatment and membrane performance (WERF, 2006):
- On average, 70% of the wastewater COD was removed
- The maximum permeate flux that could be sustained was around 5 L/m2/hr (3 gfd)
Nevertheless, multiple experiences of successful anaerobic MBR’s exist in industrial
applications, although most are using crossflow rather than submerged membranes (Sutton et
al, 2002), therefore important research needs remain (Liao et al, 2006).

Earlier work indicated that membranes are fouled more by aerobic sludges than anaerobic
ones (Judd, 2005), and that nitrogen gas sparging can compensate lower membrane flux at
ambient temperatures (Fawehinmi et al, 2004). The objective of this study was therefore to
apply submerged membranes to anaerobic treatment at ambient and low temperatures on
municipal wastewater, to find performance characteristics and limitations of this process.

MATERIALS AND METHODS

Pilot scale tests were conducted using a 40L (10 gal) anaerobic membrane bioreactor, as
shown in Fig. 1, integrating both side-stream and submerged membranes to allow both
independent and simultaneous use of each configuration. The submerged module consisted of
hollow fibres (Mitsubishi Rayon, outer diameter of 0.5 mm and pore size of 0.1 um) with a
total filtration area of 0.9 m2 (10 ft2). Two membrane modules were operated in side-stream
configuration, as summarised in Table 1.
Table 1 Characteristics of two side-stream membranes and modules
Membrane Material Module Bore Size Pore Size No of Total Filter
Type Length (mm) (mm) (um) Channels Area (m2)
X-Flow Polysulphonic 1000 8 0.03 7 0.175
Tubular Polymeric 1000 8 0.1 12 0.502
Pre-treated sewage from the influent of a primary settler in Cranfield, with concentrations
varying between 200 to 700 mg COD.L-1, was fed into the reactor, formed of a double wall
containing temperature regulated re-circulated water. To avoid complications with methane
recycle, nitrogen gas was sparged intermittently through a diffuser mounted at the bottom of
the tank to reduce membrane fouling, promote efficient mixing and eliminate dead zones.
Figure 1. Schematic Diagram showing Pilot-Scale Hybrid Anaerobic Membrane Bio-reactor
Gas out through safety vessels, sodium hydroxide and zinc acetate solutions

Acid/Base
in for pH
Control

Sewage In
FR PT
Permeate from
submerged membrane
Back flush

Back flush
FR PT

Permeate from side-


Sludge stream membrane
waste Gas lift to side-stream
N2 gas
membrane
in

Biogas was measured and analysed, after stripping hydrogen sulphide and carbon dioxide gas
with the use of 1M zinc acetate and 1M sodium hydroxide solutions respectively. Trans-
membrane pressure (TMP) on both side-stream and submerged membranes was measured
with pressure transducers (RS Components, Corby) mounted on each effluent/permeate line
and continuously recorded on a computer.
Effluent was continually withdrawn from all membranes except during critical flux and short-
term fouling tests. Submerged membranes were back flushed with permeate when trans-
membrane pressure (TMP) exceeded 0.75 bar (11 psi). All membranes were removed for
chemical cleaning when TMP reached a value of 1.5 bar (22 psi), at which point lumen
clogging occurred. Chemical cleaning was carried out ex-situ by back flushing membranes
with a 1gL-1 solution of Sodium hypochlorite (NaOCl) for 1hour. The membrane was then
back flushed for another hour with a solution of citric acid (3gL-1) to remove all traces of
Sodium hypochlorite.

The pilot scale system was run for 4 months with complete solids retention as well as a
constant hydraulic retention time (HRT) of 6±0.4 hours. The system was initially run during
14 days at an operating temperature of 35±1oC without a membrane module to allow
acclimatization of anaerobic granular biomass used as inoculum.

Submerged and side-stream membranes were then installed to allow withdrawal of permeate
without loss of solids, acclimatizing the system further until stabilisation round about day 60.
Operating temperature was then dropped to 22±0.5oC on day 70 and 12±0.5oC on day 98.
Fouling experiments were carried out at each of these operating temperatures with each
membrane type whether in side-stream or submerged configuration.

RESULTS AND DISCUSSION

After seeding, the system took 21 days to stabilise, after which COD removal began and
continually increased to reach COD removal between 82 and 97% depending on
configuration and operating conditions. Maximum COD removal efficiency (96%) was
achieved after 63 days of operation at a temperature of 35±1oC. Permeate COD concentration
during this period never exceeded 90mgL-1 despite changes to operating temperature,
membrane configuration and feed quality. Figure 2 shows influent and effluent COD
concentrations during the whole operating period.
Figure 2 Mixed Liquor COD concentration and COD Removal

Figure 3 shows the biomass concentration in the reactor, which, after the 14 day start-up
period, stabilised at MLSS concentrations around 6 g L-1 throughout the operation, while the
ratio of MLVSS to MLSS was maintained between 60 and 70 %.

20

15
Fig. 3: Evolution of Biomass Concentration

The F/M ratio and biomass yield in the reactor are illustrated in Figure 4, which also shows
the various periods of operation at the three temperatures tested. F/M gradually fell from 1.9
-1
L )
±0.04 d-1 and remained between 0.03 ±0.004 d-1 to 0.14±0.01 d-1 after day 56. In the
beginning, biomass growth yield increased until it reached a peak at approximately 0.03
gMLVSS.gCOD-1d-1 typical for anaerobic systems, but on day 28, as the system entered into
the stable phase, biomass growth yield fell to approximately 0.003±1gMLVSS.gCOD-1d-1.

0.04

35±1
0.03
Biomass Growth Yield
(gMLVSS.gCOD ) -1

Fig. 4 Biomass Growth Yield and F/M Ratio

In addition to the reactor performance, membrane behaviour at a constant flux of 15 l/m2/h (9


lmh) was studied in function of biomass characteristics (particle size distribution, capillary
suction time (CST) and viscosity), as well as flux and sparging rates of nitrogen gas.

0.02
Extracellullar polymeric substances (EPS) and Soluble Microbial Products (SMP), which
tend to be lower with anaerobic systems (Kuo and Parkin, 1996), were followed in the mixed
liquor to relate their concentration with reactor operations and membrane fouling
(Fawehinmi, 2004).

0.01
Fig. 5 Average Membrane Fouling Rate

As the production of microbial polymeric subproducts dropped off with time, Fig. 5 shows
how the fouling rate diminished with the stabilisation of the biomass, only marginally
influenced by the lower temperatures at the end of the test. During the first two weeks
of membrane installation, fouling rates averaged approximately 28
mbar/d (0.42 psi/d), which made it necessary to clean the membranes
twice weekly. After stabilisation, fouling rates had dropped to 6 mbar/d
(0.1 psi/d). With the lower operating temperature of 22 ±0.5 deg C, fouling
rates began to increase slightly, reaching a maximum of approximately 9
mbar/d (0.14 psi/d). A slight increase in average fouling rate to 10 mbar/d
(0.15 psi/d) was noted on when the temperature was lowered further to
12 ±0.5 deg C.

Fouling of the submerged membrane module (with the largest filtration


area of 0.9m2 and pore size of 0.1μm) showed a linear relationship to flux
with increasing flux producing proportional increases in fouling rate. As
shown in Fig. 6, during fouling experiments, the submerged membrane
fouling rates did not measure above 4.5 mbar/d (0.07 psi/d) despite being
subject to fluxes of 9 to 24 L/m2 h (5 to 15 gfd) applied for 4 hours for
each step. These fluxes are of similar magnitude than those typically
applied in aerobic MBRs.

The two sidestream membranes showed different behaviour, with more


exponential increases of fouling when applying higher fluxes. This was
especially pronounced for the tubular polymeric membrane, which despite
having the same pore size of 0.1 μm than the submerged membrane,
more than doubled its resistance with each increase in flux of 4.5 l/m2/h
(2.7 gfd).
While the polysulfonic sidestream membrane had a smaller pore opening
of 0.03 than the other two membranes, it also had the lowest fouling rate.
Only at the highest tested flux of 24 l/m2h (15 gfd) did the membrane
show higher fouling than the submerged module, but more than three
times lower than the polymeric sidestream membrane.

Figure 6: Operational Profiles for Submerged and Side-stream


Membranes.
As indicated by earlier work (Fawehinmi 2006) and common with aerobic
membranes, fouling is lowered by increased gas sparging rates, as
foulants are scoured away from the surface. The use of gas sparging is
thus crucial to control TMP, membrane cleaning frequency and operating
flux.

Fig. 7 Gas Sparging and its Effect on Fouling


As shown in Fig. 7, the gas sparging rate had a strong influence on membrane fouling, and all
membranes showed a similar linear behaviour in this test. For instance for the polymeric
sidestream membrane, as the gas sparging was doubled from 1 to 2 L N2/m2 membrane⋅day,
the fouling rate dropped by a factor three from an initial value of 4.5 mbar/day to 1.5
mbar/day. Nevertheless, the gas flow to liquid flux ratio between 0.1 and 0.2 applied for
membrane scouring is 150 to 300 times less than the typical aeration demand of 30 to 50
Nm3 air/m3 permeate observed for aerobic MBRs (Guihe et al, 2004).

CONCLUSIONS

The anaerobic MBR with submerged and gas sparged membranes was
demonstrated at a scale 10 time larger than previously (Hu and Stuckey,
2006), at ambient effluent temperatures and fed with unsettled municipal
sewage. More than 90 % COD removal was achieved with residuals below
50 mg/l, while biomass was stable around 6000 mg MLSS/l and biomass
yields were extremely low. Biogas yield was relatively modest but specific
methanogenetic activity of the biomass was high.

At fluxes comparable to aerobic MBRs between 10 to 20 l/m2/h, fouling


rates were reasonable and could be controlled with gas sparging at a rate
100 times lower than the gas to liquid ratio necessary with aerobic MBRs.
Since no energy is necessary to degrade pollution, biogas is produced and
the scouring energy drastically reduced, the submerged anaerobic
membrane bioreactor promises to be a highly sustainable treatment
system. The process could be used where disinfected, nutrient rich
irrigation water is required, or for ocean discharge in bathing or shellfish
waters when pathogen removal is crucial but no nutrient limitations exist.

REFERENCES

Barjenbruch, M. (2006): Benchmarking of large BAF Plants: Experience on 40 installations


in Germany, IWA Conf Biofilm Systems VI, Amsterdam, Sept. 25, (Wat Sci Tech)
Chernicharo, C. (1997): Tratamento biológico de água residuária: Reatores anaeróbios Vol. 5
Fawehinmi, F., Lens, P., Stephenson, T., Rogalla, F. and Jefferson, B. (2004) The influence
of operating conditions on extracellular polymeric substances (EPS), soluble microbial
products (SMP) and bio-fouling in anaerobic MBR, Proceedings of IWA Conference:
Water Environ. Membrane Technology, Seoul, Korea, 7-14 June (Wat Sci Tech)
Foresti, E. (2002): Anaerobic treatment of domestic sewage: established technologies and
perspectives; Water Science and Technology; 2002; Vol 45 (10); 181 -186
Guihe, T., Kekre, K., Wei, Z., Lee, TC., Viswanath, B. and Seah, H. (2004): MBR for water
reclamation, Proceedings s of IWA Conference: Water Environ. Membrane Technology,
Seoul, Korea , 7-14 June (Wat Sci Tech)
Hansen, R., Thoegersen, T, and Rogalla, F (2006): Comparing Cost and Process Performance
of Activated Sludge (AS) and Biological Aerated Filters (BAF) over ten years of full
scale operation, Proc. 79th WEFtec Conf. , Oct. 23, Dallas
Hu, A.Y. and Stuckey, D.C. (2006): Treatment of dilute wastewaters using a novel
submerged Anaerobic membrane bioreactor (SAMBR), ASCE JEnvE, 132:2, 190-198
Judd, S.J. Fouling in submerged MBRs, Wat. Sci. Tech. 15(6) 27-34 (2005).
Kuo, W. C., and Parkin, G. F. (1996). "Characterization of soluble microbial products from
anaerobic treatment Water Research, 30(4), 915-922.
Lettinga, G., Rebac, S., Parshina, S., Nozhevnikova, A., van Lier, J. and Stams, A. (1999)
High-rate anaerobic wastewater treatment at low temperature; Appl. Envir. Micobiol.;
Vol. 65 (4); 1696-1702
Liao, B., Kraemer, J. and Bagley, D. (2006): Anaerobic Membrane Bioreactors: Applications
and Research Directions, Critical Reviews in Environmental Science and Technology, 36:
489–530
Sutton, P.M., Mishra, P.N., Bratby, J.R. and Enegess, D. (2002). “MBR industrial and
municipal wastewater applications: long term operating experience,” Proc. 75th WEFtec
Conf. Chicago, Sept.
WERF (2006): Membrane Bioreactors for Anaerobic Treatment of Wastewaters, Project No.
02-CTS-4
Melin, T., Jefferson, B., Bixio, D., Thoeye, C., De Wilde, W., De Koning. J., van der Graaf,
J. and Wintgens, T. (2006) MBR technology for wastewater treatment and reuse. Desal.,
187, 271-282.
Yang, W., Cicek, N. and Ilg, J (2006) State-of-the-art of membrane bioreactors: Worldwide
research and commercial applications in North America, Journal of Membrane Science
270 (2006) 201–211

Potrebbero piacerti anche