You are on page 1of 11

ARTICLE IN PRESS

WAT E R R E S E A R C H 41 (2007) 134– 144

Available at www.sciencedirect.com

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

Methanogenic population dynamics and performance of an


anaerobic membrane bioreactor (AnMBR) treating swine
manure under high shear conditions

Sudini I. Padmasiria, Jiangzhao Zhanga, Mark Fitchc, Birgir Norddahld,


Eberhard Morgenrotha,b, Lutgarde Raskine,
a
Department of Civil and Environmental Engineering, University of Illinois at Urbana-Champaign, 205 N Mathews Ave, IL 61801, USA
b
Department of Animal Sciences, University of Illinois at Urbana-Champaign, 205 N Mathews Ave, IL 61801, USA
c
Department of Civil, Architectural, and Environmental Engineering, University of Missouri-Rolla, 222 Butler Carlton Civil Engineering Hall,
MO 65409, USA
d
TUFCO, Teknisk Udviklings og Forsknings Center Odense, Technical Development and Research Centre Odense, University of Southern
Denmark, Campusvej 55, DK 5230 Odense M, Denmark
e
Department of Civil and Environmental Engineering, The University of Michigan, 1351 Beal Ave, Ann Arbor, MI 48109 2125, USA

ar t ic l e i n f o A B S T R A C T

Article history: A 6-L, completely mixed anaerobic bioreactor with an external ultrafiltration membrane
Received 25 February 2006 module was operated for 300 days to evaluate the startup and performance of an anaerobic
Received in revised form membrane bioreactor (AnMBR) treating swine manure. The reactor had a successful
17 August 2006 startup at the initial loading rate of 1 g volatile solids (VS)/L/day. After a two-fold increase in
Accepted 22 September 2006 loading rate followed by a sudden, two-fold increase in flow velocity through the
Available online 15 November 2006 membrane module on day 75, the performance of the AnMBR deteriorated as measured
Keywords: by volatile fatty acid (VFA) accumulation, decrease in pH, and decrease in biogas
Anaerobic membrane bioreactor production. The methanogenic population dynamics in the reactor were monitored with
AnMBR terminal restriction fragment length polymorphism (T-RFLP). Changes in the relative levels
Shear of Methanosarcinaceae and Methanosaetaceae were consistent with changes in VFA
Swine manure concentrations, i.e., high and low levels of acetate corresponded to a high abundance of
Methanogens Methanosarcinaceae and Methanosaetaceae, respectively. The levels of hydrogenotrophic
T-RFLP methanogens of the order of Methanomicrobiales increased during decreased reactor
performance suggesting that syntrophic interactions involving hydrogenotrophic metha-
nogens remained intact regardless of the degree of shear in the AnMBR.
& 2006 Elsevier Ltd. All rights reserved.

1. Introduction effluent quality, and can be smaller in size than conventional


anaerobic digesters. These advantages, together with in-
Anaerobic digesters with membrane separation units (anae- creased stringency in waste disposal for animal production
robic membrane bioreactors [AnMBRs]) facilitate retention of facilities, have led to pilot scale testing of AnMBRs to treat
microorganisms and allow operation with high biomass swine waste (du Preez et al., 2005; Lee et al., 2001). The main
concentrations. Thus, AnMBRs are expected to provide more challenge for AnMBRs has been fouling of membrane units
efficient digestion, higher methane production, and better (Choo and Lee, 1996; Elmaleh and Abdelmoumni, 1997; Choo

Corresponding author. Tel.: +1 734 647 6920; fax: +1 734 763 2275.
E-mail address: raskin@umich.edu (L. Raskin).
0043-1354/$ - see front matter & 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.watres.2006.09.021
ARTICLE IN PRESS
WAT E R R E S E A R C H 41 (20 07) 13 4 – 144 135

et al., 2000; He et al., 2005). Membrane fouling is the result of abundance of syntrophic propionate oxidizing bacteria,
adsorption of organic matter, precipitation of inorganic saturated fatty acid-b-oxidizing syntrophs, and various
matter, and adhesion of microbial cells to the membrane groups of methanogens was lower under vigorous mixing
surface (Choo and Lee, 1996). High fluid flow velocities conditions (McMahon et al., 2001, 2004). Alternately, their
resulting in high shear rates at the surface of the membrane observations may be explained by differences in the rates of
can be used in AnMBRs with external membrane units to help hydrolysis and fermentation between vigorously and mini-
reduce membrane fouling caused by adhesion of biomass and mally mixed systems. Vigorous and continuous mixing may
colloidal organic matter to the membrane surface (Stephen- have promoted rapid hydrolysis and fermentation, resulting
son et al., 2000). However, the digestion efficiency in AnMBRs in production of fermentation products at rates greater than
may be negatively affected by exposure of biomass to high the utilization rates of such products by syntrophs, whereas
shear conditions. minimally mixed conditions may have resulted in slower
High mixing and shear intensities can negatively influence hydrolysis and fermentation, allowing immediate consump-
anaerobic digestion performance. Brockmann and Seyfried tion of fermentation products by syntrophic bacteria and
(1996) demonstrated a 50% loss in the specific activity of methanogens (Stroot et al., 2001). This second hypothesis
anaerobic biomass treating potato starch wastewater in an assumes interactions between syntrophic bacteria and
AnMBR with cross-flow microfiltration following re-circula- methanogens are not disrupted, but implies these micro-
tion of the entire reactor content through a membrane unit 20 organisms are not able to keep up with the high production of
times. Similar results were presented in a study by Stroot et fermentation products. Their population dynamics data (i.e.,
al. (2001) and McMahon et al. (2001), who demonstrated that lower abundance of syntrophic bacteria and methanogens
vigorously and continuously mixed solid waste digesters under vigorously mixed conditions) also support this second
(power input of 1.5 W/L) exhibited unstable performance, hypothesis.
whereas minimally mixed digesters (thoroughly shaken by From the above discussion, it can be concluded that the
hand for 2 min each day) performed very well. They also degree of shear to which the biomass in an AnMBR is exposed
demonstrated that continuously mixed unstable digesters can impact the performance and stability of anaerobic
were quickly (within 3 weeks) stabilized by reducing the digestion. This may be even more critical when complex
degree of mixing. Kim et al. (2002) obtained similar perfor- waste streams containing substantial levels of suspended
mance results in anaerobic digesters under minimal and solids, such as swine manure, are to be treated. The overall
vigorous mixing conditions. They compared the stability of objective of this study was to evaluate whether high shear in
mesophilic and thermophilic anaerobic digesters under AnMBRs with external membrane units can decrease process
different loading rates and mixing conditions. During startup, efficiencies and whether these changes are reflected in the
both the minimally mixed mesophilic and thermophilic microbial community structure. This paper presents results
reactors showed superior performance and achieved stable collected with a laboratory-scale AnMBR and focuses on
operation in shorter time periods compared to continuously reactor and membrane performance and characterization of
mixed reactors. Under steady-state conditions with a feed of methanogenic population dynamics with terminal restriction
4% solids and a hydraulic retention time (HRT) of 20 days, fragment length polymorphism (T-RFLP).
these two minimally mixed reactors produced more biogas
and less volatile fatty acid (VFA) in each temperature category
than the corresponding continuously mixed reactors. When
the loading rate was increased, the thermophilic minimally 2. Materials and methods
mixed reactors were the last to fail at an organic loading rate
of 20 g volatile solids (VS)/L/day. 2.1. AnMBR operation
While it is clear from the above discussion that high mixing
and shear intensities have a negative impact on anaerobic A 6-L, completely mixed anaerobic digester (supplied by
digestion performance, the reasons behind such observations Odense College of Engineering, Odense M, Denmark) was
are less obvious. Brockmann and Seyfried (1996) suggested operated to treat swine manure (Fig. 1). Solid–liquid separa-
that the observed reduced performance in their AnMBR may tion was achieved using a tubular ultrafiltration membrane of
be the result of a physical interruption of the syntrophic 1-m length and 12-mm diameter with a total surface area of
association between proton reducing acetogenic bacteria and 0.0377 m2. The membrane was made of polyethersulfone
their methanogenic partners. This suggestion seems logical (Weir Envig, Paarl, South Africa) with a molecular weight cut-
since high shear rates may result in break-up of microbial off of 20,000 Da. The reactor content was pumped through the
aggregates and a disruption of the juxtaposition of relevant membrane module using a progressing cavity pump (NM021,
microbes. Similarly, Kim et al. (2002) concluded that their Netzsch, Germany) controlled with a variable frequency
results imply that minimal mixing resulted in microbial converter (Danfoss, Loves Park, IL, USA). The trans-membrane
consortia proximity, and that this strategy can be used to pressure (TMP) was varied using a back pressure valve (Tru-
alleviate the problem of poor effluent quality in thermophilic Trol BBCA-CAT2, Tru-tech, Mars, PA, USA). Reactor operation
systems. Likewise, Stroot et al. (2001) and McMahon et al. and data acquisition were controlled using Lab View (National
(2001) hypothesized that vigorous mixing may have inter- Instruments, Austin, TX, USA) on a PC with a FieldPoint
rupted syntrophic associations. In contrast to the other external input/output card (National Instruments, Austin, TX,
studies, their hypothesis was supported by microbial popula- USA). The reactor temperature was kept constant at 3771 1C
tion dynamics results, which indicated that the relative with a water jacket.
ARTICLE IN PRESS
136 WA T E R R E S E A R C H 41 (2007) 134– 144

Back Pressure Flow Pressure Flow


Mixer
Valve
To Gas Meter

UF
Feed Tubular
4°C P3 Membrane Effluent
P1

Anaerobic
Digester
37°C

P1
P2
Waste/ Check
Sampling Valve
Frequency Converter

Fig. 1 – Schematic diagram of anaerobic membrane bioreactor (AnMBR). P1 is a progressing cavity pump and P2, P3, and P4
are peristaltic pumps.

The AnMBR was inoculated with a 1:1:1 (v:v:v) mixture of 2.2. Chemical analyses
sludge dredged from a swine lagoon (near Vichy, MO, USA),
anaerobic granules from an upflow anaerobic sludge blanket The biogas production was measured with a wet-test gas
(UASB) reactor treating brewery wastewater (Anheuser-Busch, meter (Schlumberger Industries, Dordrecht, The Netherlands)
Baldwinsville, NY, USA), and anaerobic sludge from the and the methane composition in the biogas was determined
primary anaerobic sludge digester of the Urbana Champaign using a gas chromatograph (Perkin Elmer AutoSystem gas
Sanitary District North-East wastewater treatment plant chromatograph, Norwalk, CT, USA) equipped with a GS-Q
(Urbana, IL, USA). Swine manure was collected from finisher column and a flame ionization detector operated isother-
buildings (University of Illinois at Urbana-Champaign, Swine mally at 40 1C using helium at a flow rate of 26 mL/min. The
Research Farm, Urbana, IL, USA) and blended to ensure break suspended solids (SS), volatile suspended solids (VSS), and
up of large particles. The homogenized waste was stored in soluble chemical oxygen demand (SCOD) in the reactor, and
containers (one for each 24-h period of operation) at 20 1C. total COD (TCOD) and SCOD in the effluent were determined
Every day, a new batch of feed was thawed overnight and according to Standard Methods (APHA, WEF, AWWA, 1998). The
diluted with tap water to the desired concentration of 6 g VS/L total VFA concentration and bicarbonate alkalinity in the
before addition to an influent storage tank kept at 4 1C which reactor were measured by a simple titration technique
was continuously mixed. The reactor was fed 125 mL every 3 h (Anderson and Yang, 1992). A parameter a, defined as the
from the influent storage tank, which corresponds to a ratio of VFA concentration over bicarbonate alkalinity, was
loading rate of 1 g VS/L/day (1 g VS corresponds to approx. also calculated. This parameter has been found useful to
1 g COD). The HRT of the system was controlled at 6 days by monitor reactor upsets or failures (Poggi-Varaldo and Olesz-
returning excess permeate back to the digester. The startup- kiewicz, 1992). The individual VFAs were analyzed using a
loading rate was doubled to 2 g VS/L/day on day 53 and tripled high-pressure liquid chromatograph (Waters 486 Tunable
to 3 g VS/L/day on day 186. Absorbance detector, Millipore Corporation, Milford, MA,
To characterize membrane fouling, the initial water flux of the USA) with an ion exclusion column of 300 mm  7.8 mm
clean membrane unit was measured. The permeate flux (J) (BIO-RAD, Aminex HPX-87 H, Hercules, CA, USA) the mobile
through the membrane was measured with a graduated phase was 0.005 M H2SO4 at flow rate of 0.6 mL/min. The
cylinder and a timer under different values of TMP. The TMP ammonia+ammonium (NH3–N) concentration in the effluent
was controlled by adjusting the backpressure valve (Fig. 1) and was determined using a HACH test kit (Method 8038, Hach,
was measured online with a pressure sensor (Red valve series Loveland, CO, USA).
42, Red Valve Co., Inc., Carnegie, PA, USA). During operation, the
permeate flow rate was measured with a variable area flow
meter (Gilmont Instruments, Racine, WI, USA). The cross flow 2.3. DNA extraction and PCR
velocity through the membrane was measured with a magnetic
flow meter (Endress+Hauser Promag 50 H, Greenwood, IN, USA). DNA was extracted using a combined detergent and bead
The total membrane resistance (Rtotal) was calculated as follows: beating procedure (UltraClean Soil DNA Kit, Mo Bio Labora-
tories Inc., Solana Beach, CA, USA) according to the manu-
TMP
Rtotal ¼ , (1) facturer’s instructions. The polymerase chain reaction (PCR)
mJ
targeting 16S ribosomal RNA (rRNA) genes of Archaea was
where m is the viscosity of the permeate estimated using the performed using the A109f (S-D-Arch-0109-a-S-17) and A912r
viscosity of pure water. (S-D-Arch-0912-a-A-20) archaeal primer pair described pre-
ARTICLE IN PRESS
WAT E R R E S E A R C H 41 (20 07) 13 4 – 144 137

viously (Lueders and Friedrich, 2000). The 50 end of the

Specificity
forward primer was labeled with 6-carboxyfluorescein (FAM)

90.2
94.7
(%)

100
for terminal restriction fragment length polymorphism
(T-RFLP) analysis. The PCR reaction mixture contained 1  PCR
buffer, 2 mM MgCl2, 0.2 mM of each deoxynucleoside tripho-
sphate (dNTP), 0.2 mM each of forward and reverse primers,

Enzyme MseI

Coverage
approximately 10 ng of the DNA template and 2.5 U of Ex Taq

70.3
69.2
(%)
DNA polymerase (TaKaRa Biomedicals, Otsu, Shiga, Japan) in

47
a final volume of 50 ml. The PCR was performed in a thermal
cycler (PTC-200 DNA Engine, MJ Research Inc., Reno, NV, USA).

Table 1 – Size of T-RFs, coveragea, and specificityb of T-RFLP with primer pair A109f–A912r for three restriction enzymes AluI, MaeI, and MseI
The amplification was done with one denaturation step at
94 1C for 2 min, followed by 30 cycles of denaturation at 94 1C

Size (bp)
for 30 s, annealing at 55 1C for 30 s, and extension at 72 1C for

103
459
420
1 min with a final extension of 8.5 min.

2.4. T-RFLP and data analysis

Specificity
Amplicons from duplicate PCR reaction mixtures were pooled

86.2
71.6
(%)

100

1.7
using a PCR purification kit (UltraClean PCR clean-up Kit, Mo
Bio Laboratories Inc., Carlsbad, CA, USA). The purified
fluorescently labeled PCR products were digested with

Enzyme MaeI
restriction enzymes AluI (New England Biolabs Inc., Beverly,

Coverage
CA, USA), MaeI (Fermentas Inc., Hanover, MD, USA), and MseI

70.9

30.7
(%)

100
95

Specificity ¼ number of target species with specific fragment size/total number of species with specific fragment size.
(New England Biolabs Inc.) for 3 h at 37 1C followed by an
enzyme inactivation step at 65 1C for 20 min. The digested
samples were treated by ethanol precipitation to remove
excess salt. The fluorescently labeled terminal restriction

Size (bp)
fragments (T-RFs) obtained in this manner were separated by

400

502
169

169
capillary electrophoresis (ABI Prism 3730  l Analyzer, Applied
Biosystems, Foster City, CA, USA) at the W.M. Keck Center for

Coverage ¼ number of target species with specific fragment size/total number of target species.
Comparative and Functional Genomics at the University of
Illinois at Urbana-Champaign Biotechnology Center (Urbana,
Specificity

IL, USA) to determine the number and size of T-RFs obtained


51.8

94.3

22.7
(%)

100
8.7
from each sample. Fragment analysis was conducted using
GeneMapperTM Version 3.7 software (Applied Biosystems,
Foster City, CA, USA). The three restriction enzymes were
selected using a model developed by Klein (2003) to allow
Enzyme AluI

Coverage

coverage of the majority of the known groups of methano-


76.2
84.6
72.6
54.1

71.4
(%)

gens and to target individual groups of methanogens. The


model allows in silico screening of a collection of downloaded
16S rRNA sequences for populations corresponding to T-RFs
obtained with specified restriction enzymes for a selected
primer pair. The 16S rRNA sequences of the domain Archaea
Size (bp)

used in the in silico analysis were downloaded from the


341
341
434
474

108

National Center for Biotechnology Information (NCBI) data-


base (Bethesda, MD, USA) in September 2004. The coverage
and specificity of each of the three restriction enzymes
selected are presented in Table 1.
species in

database

Each peak was given an allowance of 75 nucleotides during


No of

NCBI

172
26
73
61

14

T-RF assignment and only T-RFs resulting in peaks greater


than 50 fluorescence units were considered as operational
taxonomic units (OTUs). The relative abundance of each OTU
was calculated as the ratio of the peak area of that OTU over
the sum of the peak areas of all OTUs. Only the OTUs that
Methanosarcinaceae
Methanomicrobiales
Methanosarcinales-

Methanosarcinales-
Methanobacteriales

Methanosaetaceae

were found to have a relative abundance greater than 10% at


Methanococcales
Target group

least once during reactor operation are presented in the


results.
An experiment was carried out to check the precision
and reproducibility associated with the T-RFLP analysis.
b
a

The experiment consisted of two separate DNA extractions


ARTICLE IN PRESS
138 WA T E R R E S E A R C H 41 (2007) 134– 144

of the same reactor sample (collected on day 291), triplicate


PCR for each extraction, triplicate enzyme digestion for each 3. Results and discussion
PCR, followed by duplicate runs by the capillary electrophor-
esis unit. The de-salting step following enzyme digestion 3.1. Reactor performance
was not replicated and, as a result, the variation associated
with restriction enzyme digestion also includes the variation 3.1.1. Day 0–52
associated with de-salting. The reproducibility was calculated The AnMBR was started with a loading rate of 1 g VS/L/day. A
as the portion of total variance that could be attributed few days after startup, the specific biogas production
to each T-RFLP step in a particular fragment size. For example, increased to 2–3 L/g VS/day (Fig. 2a). The decreases in biogas
the % total variance of the PCR step ¼ s2PCR/(sDNA extrac- production observed on day 19 and days 27–38 were due to
2 2 2 2
tion +sPCR+sDigestion+sCapilary electrophoresis). Variance compo- problems with the gas collection system. After an initial
nents were estimated with the MIVQUEO option of the increase in effluent SCOD soon after startup, the effluent
VARCOMP procedure of SAS (Version 8.1, SAS Institute Inc.). SCOD decreased and remained stable between 200 and

Days 0-52 53-90 91-115 116-300


6
(a) Specific gas production
Gas production (L /gVS/day)
Loading rate (gVSS / L/day)

5 Loading rate

0
100

SCOD removal efficiency (%)


80
6000 (b) SCOD -effluent
SCOD (mg/L)

SCOD - reactor
COD removal % 60
4000
40

2000
20

0 0

(c)
VFA concentration

3000 VFA's 6
(mg/ L as HAc)

pH
pH and α

alpha
2000 4

1000 2

0 0
Formic
(d)
Acetic
2000
Propionic
(mg/L as HAc)
Concentration

Butyric
1500

1000

500

0
0 25 50 75 100 125 150 175 200 225 250 275 300
Days

Fig. 2 – AnMBR performance data. (a) Specific biogas production and loading rate, (b) soluble COD in the reactor and the
effluent and soluble COD removal efficiency, (c) pH, a, and total VFA concentration, (d) individual VFA concentrations.
ARTICLE IN PRESS
WAT E R R E S E A R C H 41 (20 07) 13 4 – 144 139

250 mg/L (Fig. 2b). The effluent TCOD was very similar to the added to the system on day 95. Since the addition of NaHCO3
effluent SCOD (data not shown) and the average TCOD and did not improve the performance, feeding was terminated on
SCOD removal efficiencies were over 96% and 86%, respec- day 98 and a second equivalent dose of NaHCO3 was added on
tively, during this period. The total VFA concentrations in the day 101. The concentration of total sodium added to the
reactor were also low and averaged 250 mg/L as acetic acid reactor following the second dose was calculated to be
without significant variations in the individual VFA concen- approximately 1030 mg/L, which is well below reported
trations (Fig. 2c and d). The excellent performance of the inhibitory levels of 3500–5500 mg/L (McCarty, 1964).
reactor during startup also was reflected in stable pH (average Following the second addition of NaHCO3, the reactor
of 7.5) and a values (average of 0.2) (Fig. 2c). Poggi-Varaldo and started recovering. The a value decreased to below 1.0 and
Oleszkiewicz (1992) determined that an a value of 1.0 the pH increased to above 7.0 (Fig. 2c). The methane content
corresponded to the threshold of stability for anaerobic in the biogas also increased and the SCOD and total VFAs in
codigestion of the organic fraction of municipal solid waste the reactor decreased. Acetate, propionate, and butyrate
(OFMSW) and sewage sludge and Ripley et al. (1986) observed concentrations decreased from day 103 to 114, but the
that an a value less than 0.3 was required for successful decrease in propionate concentrations was considerably
digestion of poultry manure. slower (approximately 34% reduction from day 103 to 110)
The observed performance suggests that the inoculum, compared to reductions in acetate (approximately 81%
which consisted of biomass collected from three different reduction from day 103 to 110) and butyrate (100% reduction
anaerobic environments, provided sufficient levels of the from day 103 to 110) concentrations.
relevant populations for the anaerobic degradation of swine
manure and that these populations exhibited suitable levels 3.1.4. Day 116–300
of activity. In addition, the performance results indicate that After the system recovered (i.e., returned to low a and VFA
anaerobic digestion, at least for the loading rate of 1 g VS/L/ values), the feeding was resumed on day 116 at a loading rate
day, was not affected by the high shear conditions in the of 1 g VS/L/day. On day 142, the loading rate was increased to
membrane recirculation loop (the average cross flow velocity 2 g VS/L/day and on day 186–3 g VS/L/day. We observed an
was 1.6 m/s). increase in VFA concentrations approximately 2 weeks after
each increase in loading rate (Fig. 2c). However, by reducing
3.1.2. Day 53–90 the loading rate for a short period of time, the VFA levels
The loading rate of the system was doubled to 2 g VS/L/day on returned to low values. The reactor performance was rela-
day 53 and a two-fold increase in biogas production was tively stable for the remainder of the operational period at a
observed soon thereafter. The biogas production and the loading rate of 3 g VS/L/day. However, continuous operation
methane content in the biogas remained high for a period of beyond day 300 at this loading rate of 3 g VS/L/day proved
21 days (up to day 75) (Fig. 2a) following the increase in difficult (data not reported) and it did not seem possible to
loading rate. The SCOD removal was more than 96% during increase the loading rate beyond 3 g VS/L/day without affect-
this 21-day period and the reactor VFA concentrations were ing the performance of the AnMBR. This could be because of
below 1000 mg/L as acetic acid, indicating successful treat- the high C:N ratio present in the raw swine manure. Similar
ment of swine manure. These results indicate that anaerobic problems were encountered in pilot plant AnMBRs treating
digestion of swine manure in an AnMBR remained feasible for swine manure for which the loading rate could not be
a loading rate of 2 g VS/L/day and shear conditions in the increased beyond 2.4 g VS/L/day (unpublished data).
membrane recirculation loop corresponding to an average
cross flow velocity of 1.1 m/s. 3.2. Membrane performance
The stator in the progressing cavity pump (P1 in Fig. 1) was
replaced on day 75 due to declining pump performance. As a Membrane performance was characterized using membrane
result of the stator change, the cross flow velocity in the flux and total membrane resistance (Fig. 3). The target TMP
membrane recirculation loop increased from approximately for membrane operation was 0.4 bar, but during reactor
0.9–1.9 m/s. The system performance began to deteriorate operation measured TMPs varied between 0.3 and 0.7 bar
soon after these changes. VFAs accumulated, the a value during the first 135 days. Membrane flux using a clean
increased, and the SCOD in the reactor and the effluent membrane rapidly declined from a clean water flux of 100 L/
increased to over 4000 mg/L (Fig. 2). The data in Fig. 2(d) show m2/h to values in the range of 5–10 L/m2/h with a correspond-
that the decrease in performance was accompanied by ing increase in total membrane resistance. Details of mem-
increases in reactor acetate, propionate, and butyrate con- brane fouling and cleaning have been described elsewhere
centrations. VFA concentrations can increase in response to (Zhang et al., 2007).
high NH3–N concentrations. However, the NH3–N concentra- During reactor operation, the stator of the progressing cavity
tion was measured to be below 400 mg-N/L, which is well pump in the membrane recirculation loop (Fig. 1) was
below the inhibitory level of 1.1 g NH3–N/L reported by Hansen gradually wearing out and needed to be replaced on days 22,
et al. (1998). 75, 209, and 269. The relatively fast deterioration of the stator
could be due to abrasion caused by sand particles found in the
3.1.3. Day 91–115 swine waste. The initial stator change, with resulting rapid
On day 91, the loading rate was reduced to 1 g VS/L/day since increase in the cross flow velocity from 0.9 to 1.9 m/s, did not
the performance continued to deteriorate (Fig. 2). To increase have any effect on the total membrane resistance. The second
the buffering capacity of the system, 13.1 g of NaHCO3 were stator was replaced on day 75, which resulted in a significant
ARTICLE IN PRESS
140 WA T E R R E S E A R C H 41 (2007) 134– 144

Permeate flux (L /m /h) and Total resistance (1012/m)


2.5 150
Cross flow velocity
Permeate flux
2.0 Total resistance 120

Cross flow velocity (m/s) 1.5 90

1.0 60

2
0.5 30

0.0 0
0 25 50 75 100 125 150 175 200 225 250 275 300
Days

Fig. 3 – Total resistance across the membrane, the permeate flux, and cross flow velocity of the membrane over time. The
stator of the progressing cavity pump was changed on days 22, 75, 209, and 269 (Zhang et al., 2007).

reduction in total resistance (Fig. 3). This reduction in total Methanobacteriales, was only present at substantial levels
membrane resistance can be explained by the removal of a during the first 60 days of the reactor run (Fig. 4, Table 2).
fouling layer, which had accumulated during the first 75 days At the start of operation, Methanosaeta spp. (T-RF MseI 103),
of operation and was removed by the sudden increase in shear. which have a high affinity for acetate (i.e., able to grow under
This decrease in total membrane resistance was followed by a low acetate concentrations) (Jetten et al., 1992), were abun-
steady increase in total membrane resistance during subse- dant (Fig. 4). Methanosaeta spp. were present in the reactor at
quent operation. For subsequent stator changes (days 209 and startup due to their presence in the inoclum, which consisted
269), the flow velocity was gradually increased after each stator of 1/3 (v/v) of anaerobic granules. Methanosaeta spp. have been
change to avoid negative impacts on biological conversion found to act as nuclei for granulation in UASB reactors (Zheng
processes as observed with the second stator change. et al., 2006) and anaerobic migrating blanket reactor systems
The lack of influence on membrane resistance due to the (Angenent et al., 2004). However, the granules used to
initial stator change on day 22 can be explained by the shorter inoculate the AnMBR in the current study were crushed at
operational period (there was less build up of fouling layer) the time of inoculation and conditions in the AnMBR did not
and the low solids content in the reactor due to the low promote granule formation. Thus, the increase in the levels of
loading rate: approximately 20 g SS/L on day 22 compared to Methanosaeta spp. during the initial operational period was
40 g SS/L on day 75. The initial membrane was used for over likely due to the low acetate concentration during this period
100 days without any form of cleaning and was replaced on (Fig. 2d). With the increase in loading rate (day 53), the acetate
day 134 for fouling analysis. The satisfactory long-term concentration in the reactor began to increase and the
membrane performance suggests that maintaining high abundance of Methanosaeta spp. decreased (Fig. 4). Following
cross-flow velocities was critical to reduce the rate of the increase in cross flow velocity through the membrane
membrane fouling. However, sudden changes in shear can module (day 75, Fig. 3), the Methanosaeta spp. remained low,
have a negative effect on biological processes in the AnMBR, but did not decrease further (Fig. 4), suggesting that the
although it had a positive effect on membrane performance. increase in shear did not contribute to the reduction in
The membrane flux of the second membrane was not Methanosaeta spp. levels. Consistent with these observations,
measured due to problems related to partial clogging of the the abundance of Methanosarcina spp. (T-RFs AluI 474 and MseI
flow meter by microbial growth in the effluent tubing 474), which are able to proliferate under high acetate
resulting in unreliable data. Microbial growth in the effluent concentrations (Jetten et al., 1992), increased from day 93
tubing was not surprising since the effluent SCOD during this onwards; the acetate concentration in the reactor was at its
period averaged 1272 mg/L. maximum on day 92. The Methanosarcina spp. continued to
dominate in the reactor for the rest of the operational period.
3.3. Microbial community analysis Methanogens belonging to two hydrogen utilizing metha-
nogenic groups (Methanobacteriales and Methanomicrobiales)
3.3.1. Archaeal community dynamics with T-RFLP were present in the AnMBR. The levels of Methanobacteriales
Using T-RFLP analysis, four archaeal populations belonging to were high soon after startup, but decreased to below 5% after
the orders Methanomicrobiales and Methanosarcinales were approximately 30 days of operation. Methanomicrobiales be-
found to be abundant in the AnMBR throughout the 300-day came the dominant hydrogen utilizers after the levels of
operational period. A fifth group, belonging to the family of Methanobacteriales had decreased and their relative abundance
ARTICLE IN PRESS
WAT E R R E S E A R C H 41 (20 07) 13 4 – 144 141

40
Methanobacteriales - AluI 341 (a)
35 Methanomicrobiales - AluI 434

Relative abundance (% area)


30

25

20

15

10

0
Methanosaetaceae - MseI 103 (b)
35 Methanosarcinaceae - MseI 474
Methanosarcinaceae - AluI 474
Relative abundance (% area)

30

25

20

15

10

0
0 25 50 75 100 125 150 175 200 225 250 275 300
Days

Fig. 4 – Methanogenic population dynamics in the AnMBR monitored with T-RFLP. (a) Hydrogenotrophic methanogens, (b)
aceticlastic methanogens.

Table 2 – Groups of Archaea identified in the AnMBR using T-RFLP following PCR with A109f-A912r primers and digestion
with restriction enzymes AluI, MaeI, and MseI

Archaea population present Terminal restriction fragment (T-RF) Specific genera identified Primary
in the AnMBR and the T-RF size (bp) by combining information substrate
used for relative obtained from digestions
abundance calculations with three restriction
enzymes (number of
species in each genus)
AluI MaeI MseI

Methanosaetaceae - MseI 103 108 169 103a Methanosaeta (7), Methanolobus Acetate
(1)
Methanosarcinaceae - MseI 474 108 169 474a Methanomethylovorans (4), Acetate, C-1
uncultured Methanosarcina (4), compounds
Thermoplasma volcanium (2),
uncultured Methanosaeta sp. (1)
Methanobacteriales – AluI 341 341 400 103 Methanobrevibacter (79) Hydrogen
Methanomicrobiales – AluI 434 434 502 420 Methanospirillum (4), Hydrogen
Methanocalculus (2)
Methanosarcinaceae - AluI 474 474 169 773a Methanosarcina (20), Acetate, C-1
Methanolobus (2), compounds
Methanococcoides (3)

a
Theses fragment sizes were not included in Table 1 since the specificity and coverage are very low when considered by themselves. However,
when the use of MseI was combined with information obtained by using AluI and MaeI, specific groups of species were identified.
ARTICLE IN PRESS
142 WA T E R R E S E A R C H 41 (2007) 134– 144

remained above 10% throughout the operational period. With could be the result of the increase in acetate concentration.
the increase in loading rate and increase in shear in the With decreasing acetate concentrations, the Methanosarcina
system, the abundance of the Methanomicrobiales population spp. relative abundance also decreased until day 244. VFAs
increased (days 59–95). During this time period, the reactor accumulated again between days 246 and 253, which again
performance deteriorated with increased VFA concentrations resulted in an increase in the relative abundance of Methano-
and decreased methane production. The increase in the sarcina spp. Their relative abundance remained high for the
abundance of the Methanomicrobiales population suggests that rest of the operational period (except for days 291 and 295)
this population became more active due to an increase in consistent with the relatively high VFA concentration (except
hydrogen production during this period. Although hydrogen for days 274–291) observed in the reactor. As the relative
levels were not be measured in the biogas, the observed abundance of Methanosarcina spp. decreased between days
increase in propionate concentration after the increase in 205 and 244, a corresponding increase in the abundance of
loading rate on day 52 (Fig. 2d) is consistent with an increase the Methanomicrobiales population was observed (Fig. 4a
in hydrogen production. With the change in stator on day 75, and b). These changes could be the result of high turnover
the propionate concentration rose even further. The concur- of propionate leading to increased hydrogen production in the
rent increases in propionate and hydrogen production could reactor, and a corresponding increase in the levels of
be the result of a faster hydrolysis rate due to increased shear hydrogen utilizers. It should be noted that the observed
conditions, leading to higher production of fermentation changes in population levels may be the result of changes in
products, including hydrogen, as well as an increased the levels of populations not targeted by the T-RFLP analysis
production of hydrogen by syntrophic bacteria through since relative population levels, rather than absolute popula-
conversion of other fermentation products. This would tion levels, were determined.
indicate that the increase in shear was not sufficient to break
up the syntrophic interactions between the syntrophic 3.3.2. Precision and reproducibility associated with T-RFLP for
bacteria and their methanogenic partners, as suggested by AnMBR samples
Brockmann and Seyfried (1996). However, the syntrophic T-RFLP has become a powerful tool for the rapid analysis of
bacteria were not able to convert fermentation products at microbial community structure in multiple samples. It is
the rate they were produced during this time period, which led often considered a semi-quantitative technique due to biases
to the acidification of the reactor. Reactor performance associated with DNA extraction and PCR amplification.
recovered after addition of NaHCO3 and termination of However, the output can be used to evaluate microbial
feeding for 2 weeks. The abundance of the Methanomicrobiales diversity and, in combination with sequence information, to
population decreased with the termination of feeding (days monitor changes in the abundance of specific populations.
98–115), and subsequently increased when feeding was The experiment conducted to gain confidence in the data
resumed. It should be noted that some strains of Methano- consisted of a variance component analysis performed on four
sarcina also can utilize hydrogen. However, the importance of dominant peak areas (peak area 410%) observed in
hydrogen as a growth substrate for Methanosarcina spp. in the AnMBR sample of day 291 with restriction enzyme AluI
complex microbial communities has not been studied exten- (Table 3). Peaks 104 and 472 had relative peak areas of
sively. A few studies have suggested that Methanosarcina spp. approximately 25%, whereas peaks 355 and 432 exhibited
are not able to compete for hydrogen with other hydrogen relative peak areas of approximately 10%. The variance
utilizing methanogens in most anaerobic digesters (Ahring component analysis indicated that most of the variance
et al., 1991; Zinder, 1993). (70–90%) was associated with the enzyme digestion step, which
The reactor VFA concentration again increased each time also includes the de-salting step. The coefficient of variation
the loading rate was increased. The first incident occurred (CV) involved when considering the absolute peak area was also
between days 200 and 205 (Fig. 2c and 2d). During this short high (434%, Table 3). However, when relative peak area (i.e.,
time period and a few days thereafter, an increase in the relative abundance) was considered, less variation (CV o20%)
levels of Methanosarcina spp. was observed (Fig. 4b), which was observed among the samples (Table 3).

Table 3 – Percent variance associated with different T-RFLP steps in AnMBR sample collected on day 291 using restriction
enzyme AluI, and coefficient of variation (CV) for relative peak area and absolute peak area

Fragment size % of total variance in absolute peak area associated with different CV (%) for CV (%) for
(bp) T-RFLP steps absolute relative
peak area peak area
DNA PCR Digestion Capillary
Extraction (+Desalting) electrophoresis

104 0.0 0.0 88.8 11.2 34 20


355 20.0 0.0 79.6 0.4 45 16
432 17.1 0.0 68.0 14.8 53 16
472 0.0 0.0 89.2 10.8 37 16
ARTICLE IN PRESS
WAT E R R E S E A R C H 41 (20 07) 13 4 – 144 143

The reported relative population levels (Fig. 4) are based on Angenent, L.T., Sung, S., Raskin, L., 2004. Formation of granules
single analyses. However, the variances in peak area asso- and Methanosaeta fibres in an anaerobic migrating blanket
ciated with different T-RFLP steps indicate that a single reactor (AMBR). Environ. Microbiol. 6 (4), 315–322.
APHA, WEF, AWWA, 1998. Standard methods for the examination
sample analysis is sufficient to represent a particular data
of water and wastewater. In: Clesceri, L. S., Greenberg, A. E.,
point. The low CV of relative peak area also supports this
Eaton, A.D. (Eds.), 20th ed. American Public Health Associa-
observation. In addition, a single analysis allowed us to tion, Washington, DC.
process a time series of data in a fast and efficient manner. Brockmann, M., Seyfried, C.F., 1996. Sludge activity and cross-flow
Osborn et al. (2000) also presented an evaluation of T-RFLP microfiltration–A non beneficial relationship. Water Sci.
profiling using environmental DNA samples isolated directly Technol. 34 (9), 205–213.
from polluted and pristine soils. Some of the factors Choo, K.-H., Lee, C.-H., 1996. Membrane fouling mechanisms in
investigated included the use of replicates, pooling of the membrane-coupled anaerobic bioreactor. Water Res. 30 (8),
1771–1780.
extracted DNA samples, choice of polymerase, PCR annealing
Choo, K.H., Kang, I.J., Yoon, S.H., Park, H., Kim, J.H., Adiya, S., Lee,
temperature, and dilution effects. They obtained a coefficient C.H., 2000. Approaches to membrane fouling control in
of variation less than 11% in the reproducibility of peak anaerobic membrane bioreactors. Water Sci. Technol. 41,
heights in three replicates (the de-salting step was not 363–371.
included in their protocol). These results obtained demon- du Preez, J., Norddahl, B., Christensen, K., 2005. The BIOREKs
strate that, once standardized, T-RFLP is a robust and concept: a hybrid membrane bioreactor concept for very
reproducible methodology for the rapid analysis of microbial strong wastewater. Desalination 183 (1–3), 407–415.
Elmaleh, S., Abdelmoumni, L., 1997. Cross-flow filtration of an
community structure in different samples.
anaerobic methanogenic suspension. J. Membrane Sci. 131,
261–274.
Hansen, K.H., Angelidaki, I., Ahring, B.K., 1998. Anaerobic diges-
4. Conclusions tion of swine manure: inhibition by ammonia. Water Res. 32
(1), 5–12.
 Stable anaerobic digestion of swine manure can be He, Y., Xu, P., Li, C., Zhang, B., 2005. High-concentration food
achieved in an AnMBR with external membrane filtration wastewater treatment by an anaerobic membrane bioreactor.
Water Res. 39, 4110–4118.
using cross flow velocities of up to 2 m/s.
Jetten, M.S.M., Stams, A.J.M., Zehnder, A.J.B., 1992. Methanogen-
 An increase in cross flow velocity was found to benefit
esis from acetate—a comparison of the acetate metabolism in
membrane performance, but resulted in poor anaerobic Methanothrix soehngenii and Methanosarcina spp. FEMS Micro-
digestion performance. biol. Rev. 88, 181–197.
 T-RFLP results indicated that hydrogen utilizing methano- Kim, M., Ahn, Y.H., Speece, R.E., 2002. Comparative process
gens increased in abundance during a period of poor stability and efficiency of anaerobic digestion; mesophilic vs.
reactor performance, caused by a sudden increase in shear thermophilic. Water Res. 36 (17), 4369–4385.
Klein, A.N., 2003. MS Thesis, University of Illinois, Urbana-
conditions.
Champaign, USA.
 The decrease in reactor performance likely can be
Lee, S.M., Jung, J.Y., Chung, Y.C., 2001. Novel method for
explained by an increase in the rate of hydrolysis caused enhancing permeate flux of submerged membrane system in
by an increase in shear leading to a buildup of fermenta- two-phase anaerobic reactor. Water Res. 35 (2), 471–477.
tion products. It is unlikely that the decrease in reactor Lueders, T., Friedrich, M., 2000. Archaeal population dynamics
performance was due to breakup of microbial aggregates during sequential reduction processes in rice field soil. Appl.
and interruption of syntrophic interactions. Environ. Microbiol. 66, 2732–2742.
McCarty, P.L., 1964. Anaerobic waste treatment fundamentals.
Part III, toxic materials and their control. Public Works 95
Acknowledgements (November), 91–94.
McMahon, K.D., Stroot, P.G., Mackie, R.I., Raskin, L., 2001.
Anaerobic codigestion of municipal solid waste and biosolids
The authors would like to thank Aurelio Briones for helpful
under various mixing conditions—II: microbial population
discussions and Mike Tenuto, Andrew White, and Tara dynamics. Water Res. 35 (7), 1817–1827.
Jackson for their help with sampling. This work was partially McMahon, K.D., Zheng, D., Stams, A.J., Mackie, R.I., Raskin, L.,
supported by The Water CAMPWS, a Science and Technology 2004. Microbial population dynamics during start-up and
Center of Advanced Materials for the Purification of Water overload conditions of anaerobic digesters treating municipal
with Systems under the National Science Foundation agree- solid waste and sewage sludge. Biotechnol. Bioeng. 87,
ment number CTS-0120978. 823–834.
Osborn, A.M., Moore, E.R.B., Timmis, K.N., 2000. An evaluation of
terminal-restriction fragment length polymorphism (T-RFLP)
R E F E R E N C E S
analysis for the study of microbial community structure and
dynamics. Environ. Microbiol. 2, 39–50.
Poggi-Varaldo, H.M., Oleszkiewicz, J.A., 1992. Anaerobic cocom-
Ahring, B.K., Westerman, P., Mah, R., 1991. Hydrogen inhibition of posting of municipal solid-waste and waste sludge at high
acetate metabolism and kinetics of hydrogen consumption by total solids levels. Environ. Technol. 13, 409–421.
Methanosarcina thermophilia TM-1. Arch. Microbiol. 157, Ripley, L.E., Boyle, W.C., Converse, J.C., 1986. Improved alkali-
38–42. metric monitoring for anaerobic-digestion of high-strength
Anderson, G.K., Yang, G., 1992. Determination of bicarbonate and wastes. J. Water Pollut. Control Fed. 58 (5), 406–411.
total volatile acid concentration in anaerobic digesters using a Stephenson, K.B., Jud, S., Jefferson, B., 2000. Membrane Bioreac-
simple titration. Water Environ. Res. 64, 53–59. tors for Wastewater Treatment. IWA Publishing, London.
ARTICLE IN PRESS
144 WA T E R R E S E A R C H 41 (2007) 134– 144

Stroot, P.G., McMahon, K.D., Mackie, R.I., Raskin, L., 2001. Zheng, D., Angenent, L.T., Raskin, L., 2006. Monitoring granule
Anaerobic codigestion of municipal solid waste and biosolids formation in anaerobic upflow bioreactors using oligonu-
under various mixing conditions—I. Digester performance. cleotide hybridization probes. Biotechnol. Bioeng. 94 (3),
Water Res. 35 (7), 1804–1816. 458–472.
Zhang, J., Padmasiri, S.I., Fitch, M., Norddahl, B., Raskin, L., Zinder, S.H., 1993. Physiological ecology of methanogens. In:
Morgenroth, E., 2007. Influence of cleaning frequency and Ferry, J.G. (Ed.), Methanogenesis, Ecology, Physiology, Bio-
membrane history on fouling in an anaerobic membrane chemistry & Genetics. Chapman & Hall, Inc., London, pp.
bioreactor. Desalination 207, 153–166. 128–206.

You might also like