Sei sulla pagina 1di 13

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/233609998

Integrated Anaerobic and Aerobic Processes for Treatment of Municipal


Wastewater

Article · January 2006


DOI: 10.2175/193864706783751438

CITATION READS
1 76

9 authors, including:

How Yong Ng Shih Wei Wong


National University of Singapore AECOM Singapore
140 PUBLICATIONS   3,503 CITATIONS    3 PUBLICATIONS   64 CITATIONS   

SEE PROFILE SEE PROFILE

Yuen Long Wah Say-Leong Ong


PUB Singapore National University of Singapore
11 PUBLICATIONS   29 CITATIONS    341 PUBLICATIONS   4,688 CITATIONS   

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Shih Wei Wong on 28 May 2015.

The user has requested enhancement of the downloaded file.


WEFTEC®.06

INTEGRATED ANAEROBIC AND AEROBIC PROCESSES


FOR TREATMENT OF MUNICIPAL WASTEWATER

How Yong Ng1,*, Shih Wei Wong1, Sing Chuan Wong1, Kavitha Krishnan1, Siow Woon Tiew1,
Wing Onn Kwok2, Kian Eng Ooi2, Yuen Long Wah2 and Say Leong Ong1
1
Centre for Water Research. Division of Environmental Science and Engineering, National University of Singapore,
Block EA #07-23, 9 Engineering Drive 1, Singapore 117576
2
Public Utilities Board Singapore, 40 Scotts Road, #15-00 Environmental Building, Singapore 228231
*
Corresponding Author: esenghy@nus.edu.sg

ABSTRACT

In this study, integrated anaerobic and aerobic treatment processes for municipal wastewater
treatment were investigated for its feasibility to replace the existing conventional activated
sludge process. Three types of anaerobic systems − anaerobic sludge blanket, anaerobic
sequencing batch reactor and anaerobic filter, as the pretreatment step were studies. High solids
and organics removal efficiency was observed during the start-up period at a hydraulic retention
time (HRT) of 16 h, as well as during 6 h HRT operation. The UASB was found to require the
shortest start-up time (80 days) and it produced the highest amount of methane gas (0.16 L/(g
tCOD removed) during start-up and 0.13 L/(g tCOD removed) during 6 h HRT operation).
However, it was a challenge to operate the UASB due to sludge floatation problems encountered
during the start-up period when the HRT was 16 h. The performance characteristics of
conventional activated sludge system and membrane bioreactor were also assessed to evaluate
their effectiveness for treating the effluent of the UASB system. Both aerobic systems were able
to produce an excellent final effluent with quality meeting that of the secondary effluent
discharge standards. However, the membrane bioreactor was noted to be a better option as it
offered the advantages such as smaller footprint and higher removal efficiencies.

KEYWORDS

Upflow anaerobic sludge blanket, anaerobic sequencing batch reactor, anaerobic filter, municipal
wastewater, conventional activated sludge, membrane bioreactor

INTRODUCTION

Aerobic treatment systems such as the conventional activated sludge (CAS) process are widely
adopted for treating low strength wastewater (<1000 mg COD/L) like municipal wastewater.
CAS process is energy intensive due to the high aeration requirement and it also produces large
quantity of sludge (about 0.4 g dry weight/g COD removed) that has to be treated and disposed
of. As a result, the cost of operation and maintenance of a CAS system is considerably high.

On the other hand, anaerobic process for domestic wastewater treatment has presented an
alternative that is potentially more economical and holistic conceptually (Mergaert et al., 1992),
particularly in the sub-tropical and tropical regions where the climate is warm consistently

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3205
WEFTEC®.06

throughout the year. Anaerobic process does not require aeration and it produces biogas and less
sludge.

Research has shown that anaerobic systems such as the Upflow Anaerobic Sludge Blanket
(UASB) (Behling and Sant’Anna, 1997; Barbosa et al., 1989), the Anaerobic Sequencing Batch
Reactor (AnSBR) (Sung and Dague, 1992; Ng, 1989) and the Anaerobic Filter (AF) (Ng and
Chin, 1986; Chernicharo and Machado, 1998) can successfully treat high-strength industrial
wastewater as well as low-strength synthetic wastewater. However, there have been limited
reports on using anaerobic systems as a pretreatment for municipal wastewater.

The anaerobic process alone would not be able to produce an effluent of a quality that meets
typical secondary effluent standards. Post-treatment will therefore be required. However, the size
of the aerobic system for the integrated anaerobic and aerobic treatment processes (IAATP) will
be greatly reduced because the wastewater has been pretreated by the anaerobic system. Thus,
the IAATP is potentially a more cost-effective technology for treating municipal wastewater by
reducing the aeration requirement and sludge production while achieving secondary effluent
standards.

In view of the above observations, this study was conducted to investigate the feasibility of using
an IAATP for treating municipal wastewater. Three types of anaerobic system − UASB, AnSBR
and AF were investigated. Comparisons of the anaerobic systems were based on length of start-
up period, effluent quality, biogas production and ease of operation. Two aerobic systems, a
CAS and a membrane bioreactor (MBR), were also used to treat the effluent from the UASB.
The aerobic systems will be compared based on their effluent quality.

METHODOLOGY

Experimental Setup for Anaerobic Systems

One 40-L UASB, one 22.5-L AnSBR and one 21-L AF laboratory-scale systems were set up.
The schematic diagram of the UASB, AnSBR and AF are shown in Figures 1, 2 and 3,
respectively. The three anaerobic systems were started up with a hydraulic retention time (HRT)
of 16 h and subsequently, switched to 6 h (phase 2) after successful start-up. During the start-up
period, there was no intentional biomass wasting for the UASB and AF while the solids retention
time (SRT) of the AnSBR was maintained at 30 days by daily wasting.

The anaerobic reactors were seeded with digester sludge from a local municipal wastewater
treatment plant. Raw municipal wastewater was collected twice a week from the same treatment
plant and stored at 4ºC before use. The UASB and AF were continuously fed with sieved (2 mm
size filter) raw municipal wastewater stored in a common feed holding tank. The feed of the
AnSBR was similarly sieved and fed in batch mode from a separate feed tank. The feed of all
three anaerobic systems were maintained at 30ºC by a thermal controller installed in a feed
transfer tank. The contents in the feed tanks and feed transfer tanks were kept homogeneous by
using top mounted stirrers.

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3206
WEFTEC®.06

The UASB was fed from the bottom through six evenly distributed inlet ports. The flowrate of
the feed was calculated based on the intended HRT (i.e., 16 h during start-up period and 6 h
during phase 2). The top of the UASB was mounted with a three-phase separator where the
solids could be retained in the reactor, while the effluent could be overflown into an effluent tank
and the biogas could be transferred and collected in gas collectors.

The AnSBR was operated with four phases (fill, react, settle and decant) controlled by a process
logic controller (PLC), as shown in Table 1. The feed wastewater was fed into the AnSBR from
the bottom and the decant point was fixed at a height to achieve the calculated decant volume.

Table 1. Operating parameters of the AnSBR


Start-up phase Phase 2
HRT 16 h 6h
Duration 160 days 43 days
No. of cycles per day 4 8
Operating protocol Fill 15 min Fill 15 min
React 270 min React 90 min
Settle 60 min Settle 60 min
Decant 15 min Decant 15 min

The AF was fed from the bottom through a single inlet port. Suspended matters at the bottom
part of the reactor were kept in suspension by a top mounted stirrer. Above this, the column was
filled with filter media (Sera Siprox D52518, Germany). A liter of this media would provide
approximately an effective surface area of 270 m2. The effluent would overflow from the top of
the reactor into the effluent tank.

The pH of the anaerobic reactors was maintained between 6.8 and 7.2 (the optimal pH range for
anaerobic treatment) by a pH controller. Sodium carbonate was dosed into the anaerobic
reactors through the feed line when the pH fell below this range.

Experimental Setup for Aerobic Systems

One conventional activated sludge (CAS) and one submerged membrane bioreactor (MBR) were
start-up simultaneously to treat effluent from the UASB. Working volume for both systems was
4.5 L. The CAS system consisted of an aeration tank and a clarifier while the MBR consisted of
an aeration tank with submerged membranes. The schematic diagrams of the CAS and MBR are
illustrated in Figures 4 and 5, respectively. The HRTs for both systems were controlled at 4 h
while the SRTs were maintained at 10 and 20 days for the CAS and MBR, respectively.

Membrane module used for the MBR was Sterapore-L hollow fiber membrane from Mitsubishi
Rayon Co (Japan). The membrane fiber had a nominal pore size of 0.4 μm with a total surface
area of 0.03 m2 per module. Nine membrane modules were used simultaneously in the MBR.

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3207
WEFTEC®.06

Figure 1. Schematic diagram of a


lab-scale UASB

Figure 2. Schematic diagram of a lab- Figure 3. Schematic diagram of a lab-scale AF


scale AnSBR

pH controller pH controller
Pressure Suction
Mechanical Gauge Pump
Stirrer
Influent Pump
Influent Pump
Treated
Effluent

Level Sensor
Aeration Clarifier
Tank MBR Submerged
Membrane
Diffuser
Diffuser Module
Air
Air Recycle
Treated flowmeter
flowmeter Pump
Effluent

Figure 4. Schematic diagram of the CAS Figure 5. Schematic diagram of the MBR

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3208
WEFTEC®.06

The aeration rates for the CAS and MBR were maintained at 1.0 and 5.0 L/min, respectively.
High aeration rate was provided in the MBR to achieve effective scouring of the membranes to
minimize cake formation on the membrane surface. In addition, the membrane modules were
operated intermittently (8 min on, 2 min off). The pH was maintained between 6.8 to 7.2 and the
temperature ranged between 25 and 32oC.

Sample Collection and Analysis

Feed and effluent samples were collected regularly from their respective sampling tubes. The
soluble portion of the feed and effluent were collected by filtering through a 0.45 μm filter disc
(Pall GN-6 Metricel® Grid, US) after centrifuging for 10 minutes at 9,000 rpm.

Several parameters were used to assess the performance of the systems. Solids removal was
determined by suspended and volatile suspended solids (SS & VSS) concentrations, and organics
removal by total chemical oxygen demand (tCOD) and total biochemical oxygen demand
(tBOD5). The soluble portion of the feed and effluent were also tested for soluble COD (sCOD)
and soluble BOD (sBOD5). Biogas production by the anaerobic systems was monitored by the
volume of gas produced daily and the methane gas composition was determined using a gas
chromatograph (Shimadzu GC-17A, Japan) equipped with a thermal conductivity detector and a
2 m x 1/8 in stainless steel Porapa Q 80/100 mesh column. All tests were conducted in
accordance with the Standard Methods (APHA, 1998).

RESULTS AND DISCUSSIONS

Anaerobic Systems Start-up Performance

It is well known that the start-up of anaerobic systems can be relatively long compared with
aerobic systems because of the slow growth rate of anaerobic microorganisms. Therefore, in
order to allow the biomass to acclimatize, the HRT used for the start-up period was slightly
longer (i.e., 16 h). This could indirectly help to increase the SRT of the system by preventing
biomass washout and retaining the biomass in the reactor. Accumulation of volatile fatty acids,
which can suppress the metabolic activity of methanogens, may be an issue during the start-up
phase. This could be effectively resolved by effective pH control. For an anaerobic system,
once the start-up phase is successful, it can be left dormant for extended periods without severe
deterioration in biomass properties since it is rather robust. Therefore, it is critical that the start-
up be monitored closely.

The UASB, AnSBR and AF were operated at a HRT of 16 h for 250 days, 160 days and 237
days, respectively. The start-up performance data shown in Table 2 was compiled based on
experimental results obtained after the anaerobic reactors had achieved rather stable performance.
A summary of the results is shown in Table 2.

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3209
WEFTEC®.06

Table 2. Summary of start-up performance of the UASB, AnSBR and AF

Reactor UASB AnSBR AF


Effective Vol (L) 40 22.5 21
HRT (h) 16 16 16
Days of
250 160 237
Operation
Influent Effluent Removal % Influent Effluent Removal % Influent Effluent Removal %

248 – 715 42 – 168 320 – 610 40 – 80 218 – 543 29 – 79


SS (mg/L) 71 – 90 (81) 79 – 88 (85) 83 – 92 (87)
(460) (94) (411) (61) (363) (50)
196 – 420 33 – 124 245 – 325 30 – 70 165 – 473 24 – 50
VSS (mg/L) 72 – 89 (80) 73 – 88 (85) 78 – 93 (88)
(323) (69) (281) (52) (316) (36)
318 – 734 78 – 222 239 – 542 41 – 130 343 – 735 46 – 117
tCOD (mg/L) 33 – 87 (70) 69 – 83 (75) 75 – 90 (83)
(525) (155) (369) (93) (514) (88)
36 – 157 36 - 77 10 -117 5 – 36 45 – 111 7 – 74
sCOD (mg/L) 12 – 69 (46) 28 – 72 (48) 14 – 91 (50)
(88) (47) (51) (25) (85) (42)
122 – 236 24 – 83 124 - 216 28 – 58 172 – 354 10 – 85
tBOD (mg/L) 60 – 86 (77) 71 – 84 (78) 72 – 94 (80)
(197) (46) (198) (43) (213) (41)
17 – 75 9 – 38 12 – 16 9 – 38 2 – 25
sBOD (mg/L) 27 – 73 (56) 7 – 10 (8) 26 – 48 (39) 34 – 75 (56)
(38) (16) (13) (27) (13)
Biogas
2.7 – 5.3 (4.3) 0.15 – 2.8 (1.0) 0.71 – 0.77 (0.76)
production (L/d)
Methane
73 – 85 (78) 60 – 61 (61) 48 – 51 (51)
composition (%)
Specific methane
production (L/g 0.090 – 0.22 (0.16) 0.006 – 0.11 (0.04) 0.002 – 0.04 (0.032)
tCOD removed)

Note: Data Format: minimum – maximum (average)

UASB. The average SS and VSS removal efficiencies were 81 and 80%, respectively. The
average tCOD and sCOD removal efficiencies achieved were 70 and 46%, respectively. The
average tBOD5 and sBOD5 removal efficiencies were found to be 77 and 56%, respectively. The
average biogas production was 4.3 L/d with 78% of it comprised of methane. This is equivalent
to about 0.16 L/(g tCOD removed).

AnSBR. The average SS and VSS removal efficiencies were both 85%. The average tCOD,
sCOD, tBOD5 and sBOD5 removal efficiencies were 75, 48, 78 and 39%, respectively. The
average biogas production was 1.0 L/d with 61% of it comprised of methane. This is equivalent
to about 0.04 L/(g tCOD removed).

AF. The average SS and VSS removal efficiency were 87 and 88%, respectively. The tCOD,
sCOD, tBOD5 and sBOD5 removal efficiencies were 83 %, 50 %, 80 % and 56 %, respectively.
The average biogas production was 0.76 L/d with 51% of it as methane. This is equivalent to
about 0.032 L/(g tCOD removed).

The results of the start-up phase showed that at a relatively long HRT (16 h), the anaerobic
systems were capable of producing effluents of high quality. The first sign of stability was
observed at Day 80 for the UASB, Day 110 for the AnSBR and Day 115 for the AF. Stability
was judged based on rather consistent effluent quality and steady biogas (methane) production.
The UASB seemed to be able to achieve a successful start-up much earlier than the AnSBR and
AF. This presented UASB with an obvious advantage.

During the early stage of the start-up period, too low an organic loading rate (OLR) (i.e., 0.55 g
COD/L.d) posed a problem for the AnSBR. Biogas production was very low and soluble
organic concentration in the effluent was high. This was due to the low specific microbial

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3210
WEFTEC®.06

activity within the bioflocs. In a lab-scale study of AnSBR by Zhou (1999), the optimum OLR
in terms of performance was reported to be 2.5 g COD/L•d. However, it is undesirable to
decrease the start-up HRT because of the high possibility of a substantial loss in biomass. This is
why the start-up period for the AnSBR was observed to be relatively long in this study.

The AF also required a relatively long start-up period because biofilms would need time to
develop on the surface of the filter media and to mature. In this study, biofilm cells tended to
slough off and caused the effluent quality to deteriorate during the early stage of the start-up
period.

Anaerobic Systems Performance at HRT of 6 h

The UASB, AnSBR and AF were subsequently operated at a HRT of 6 h after successful start-
ups. Data was collected over a period of 164, 140 and 141 days for the UASB, AnSBR and AF,
respectively. Table 3 shows a summary of the performance of the UASB, AnSBR and AF when
operating at 6 h HRT.

UASB. The average SS and VSS removal efficiencies were 58 and 56%, respectively. The
average tCOD and sCOD removal efficiencies achieved were 57 and 38%, respectively. The
average tBOD5 and sBOD5 removal efficiencies were found to be 68 and 48%, respectively. The
average biogas production was 6.9 L/d with 73% of it comprised of methane. This is equivalent
to about 0.13 L/(g tCOD removed).

AnSBR. The average SS and VSS removal efficiencies were 63 and 65%, respectively. The
average tCOD, sCOD, tBOD5 and sBOD5 removal efficiencies were 57, 56, 72 and 47%,
respectively. The average biogas production was 3.7 L/d with 74% of it comprised of methane.
This is equivalent to about 0.12 L/(g tCOD removed).

AF. The average SS and VSS removal efficiencies were 86 and 84%, respectively. The average
tCOD, sCOD, tBOD5 and sBOD5 removal efficiencies were 78, 63, 82 and 54%, respectively.
The average biogas production was 1.4 L/d with 49% of it comprised of methane. This is
equivalent to about 0.021 L/(g tCOD removed).

Comparing the performance of the three anaerobic systems, it could be observed that the solids
removal efficiency of the UASB seemed to be lower than those of the AnSBR and AF. This was
because as the UASB was operated at a maximum sludge hold-up (no intentional sludge wasting),
it was less tolerant towards a sudden increase in feed solids concentration. Excessive solids
beyond what the reactor could retain would overflow out of the reactor with the effluent.
However, this happened only periodically, causing the average solids removal efficiency to be
relatively lower. A lower solids removal efficiency would inevitably decrease the tCOD and
tBOD5 removal efficiencies, as shown in Table 3. On the other hand, the filter medium in the
AF was able to trap a substantial amount of SS, which explained the relatively higher solid
removal efficiency obtained. Nevertheless, flow blockage within the filter column due to solids
accumulation was not detected over the 180 days of operation. In the case of AnSBR, it was able
to achieve organic removal efficiencies comparable to the UASB despite having a lower MLVSS
concentration (MLVSS concentration of 4.5 g/L in the AnSBR during the react phase compared

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3211
WEFTEC®.06

with 13.3 g/L at the mid level of the UASB column). This could be because the AnSBR has the
advantage of providing a high food-to-microorganisms (F/M) ratio during the fill and react
phases that encourages metabolic activity, and a low F/M ratio during the settle phase that is
ideal for biomass flocculation and phase separation (Dague and Pidaparti, 1991).

Table 3. Summary of performance of the UASB, AnSBR and AF


Reactor UASB AnSBR AF
Effective Vol (L) 40 22.5 21
HRT (h) 6 6 6
Days of
164 140 141
Operation
Average Average
Influent Effluent Removal % Influent Effluent Influent Effluent
Removal % Removal %
280 – 850 200 – 425 235 – 820 67 – 365 240 – 737 50 – 128
SS (mg/L) 42– 84 (58) 38 – 86 (63) 72 – 93 (86)
(437) (322) (488) (170) (550) (71)
200 – 425 64 – 196 170 – 670 57 – 235 220 – 684 48 – 130
VSS (mg/L) 41 – 80 (56) 41 – 85 (65) 70 – 91 (84)
(322) (134) (390) (124) (448) (71)
318 – 766 101 – 328.5 327 – 617 164 – 361 397 – 924 45 – 195
tCOD (mg/L) 39 – 70 (57) 24 – 77 (57) 67 – 93 (78)
(544) (226.7) (467) (198) (632) (132)
37 – 130 34 - 84 72 -218 27 – 105 63 – 191 8 – 48
sCOD (mg/L) 14 – 61 (38) 36 – 79 (56) 41 – 96 (63)
(93) (57) (122) (56) (98) (33)
122 – 330 41 – 151 176 - 207 43 – 66 144 – 280 24 – 89
tBOD (mg/L) 39 – 82 (68) 65 – 79 (72) 65 – 92 (82)
(229) (79) (191) (54) (243) (44)
17 – 75 11 – 34 12 – 23 22 – 36 5 – 29
sBOD (mg/L) 40 – 63 (48) 6 – 12 (8) 46 – 49 (47) 20 – 80 (54)
(38) (19) (15) (28) (13)
Biogas
5.2 – 8.3 (6.9) 3.4 – 4.3 (3.7) 1.4 – 1.5 (1.4)
production (L/d)
Methane
66 – 75 (73) 71 – 76 (74) 48 – 49 (49)
composition (%)
Specific methane
production (L/g 0.081 – 0.25 (0.13) 0.080 – 0.15 (0.12) 0.013 – 0.031 (0.021)
tCOD removed)
Note: Data Format: minimum – maximum (average)

Despite the relatively low solids removal efficiency, the UASB has the highest biogas production
among the three anaerobic systems investigated. This was probably due to a well structured
microbial community along the height of the UASB column. In addition, the UASB operating at
its maximum sludge hold-up was also able to maintain a higher concentration of biomass inside
the reactor with a MLVSS concentration of 14.7 g/L at the base and 13.3 g/L at the mid level of
the UASB column (compared with the MLVSS concentration of 4.5 g/L in the AnSBR during
the react phase).

Operational Challenges of UASB, AnSBR and AF

In all biological wastewater treatment systems, the variability in the raw wastewater will impose
an uncontrolled selective pressure on the microbial population in the reactor. If not counteracted
with appropriate operating strategies, it may affect the performance of the reactors. In this study,
the three types of anaerobic systems were found to be rather easy to operate and maintain.

Occasionally, the UASB would experience sludge floatation problem due to biogas being
trapped in the biomass, which was at very high concentration. The low upflow velocity at a
HRT of 16 h could not provide the sufficient shear force needed for dislodging the small gas
bubbles away from the biomass. As a result, gas bubbles tended to be trapped within the dense
biomass. As the amount of small gas bubbles accumulated, an uplift force that was large enough
to lift the sludge blanket was created. This phenomenon caused the effluent quality to deteriorate

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3212
WEFTEC®.06

as the solids were washed out of the UASB and a corresponding decrease in the biogas
production due to the upset of the microbial community structure along the height of the reactor.
When the said phenomenon occurred, internal recycling of biomass from the top to the base of
the UASB was carried out for about two days and the performance of the UASB would require
about two weeks for recovery. However, this problem was resolved when the HRT was
decreased to 6 h. A higher upflow velocity created a sufficient shear force to free the gas
bubbles, allowing them to rise to the top of the UASB.

For the AF, gas bubbles could also be trapped in the filter media. Although it did not cause
sludge floatation problems as in the UASB, it caused biogas production to be unstable.

The AnSBR did not experience problems with gas bubbles because it had a stirrer which kept the
contents suspended during the react phase. Lettinga et al. (1984) reported a significant fraction
of biomass becoming dispersed in the liquid above biomass bed because of high turbulence
caused by biogas, especially during the settling phase. The OLR in this case was very low, even
at a short HRT of 6 h, because the organic strength of the municipal wastewater was weak.
Therefore, gassing problems were not encountered in the AnSBR in this study.

One important operation parameter that directly affected the quality of the treated effluent as
well as the biogas production was pH, as anaerobic microorganisms are exceptionally sensitive
to pH changes. The pH within a reactor must be maintained in the range of 6.4 to 7.8. Outside
this range, methanogens would be inhibited and volatile organic acids would then be
accumulated in the bioreactor, causing the pH to drop further. The alkalinity in the feed
wastewater normally provides sufficient buffering capacity. Nevertheless, it is crucial to check
and maintain the pH level within the optimal range at all times, particularly during the critical
start-up period. This was achieved in this study using pH controllers for the UASB, AnSBR and
AF.

Aerobic Post-Treatment System Performance at HRT of 4 h

Table 4 shows a summary of the performance of the two aerobic systems − the CAS and MBR.
Both the CAS and MBR performed excellently, especially the MBR, in terms of suspended
solids, organic and ammonia removals, which achieved the typical secondary effluent discharge
standards. The average tCOD for the CAS and MBR effluents were 42.7 and 23.2 mg/L,
respectively. Effluent tBOD5 concentrations were 6.3 and 1.3 mg/L for the CAS and MBR,
respectively. Both systems were able to meet the typical secondary effluent discharge standards.

Comparing the MBR and the CAS, the MBR could achieve a higher organic removal due to
effective solid-liquid separation provided by the membrane. In addition, the MBR could support
a higher SRT than CAS. This enabled the enrichment of slow growing bacteria such that a more
diverse microbial community with broader physiological capabilities could be established.

Effluent SS for CAS ranged from 4 to 41 mg/L, with an average of 18 mg/L. High effluent SS
that exceeded the discharged limit of 30 mg/L was observed on certain days. It has been
reported that irregular shaped floc particles of low density and pin point flocs were present in

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3213
WEFTEC®.06

systems operating at a SRT between 9 to 12 days (Bisogni and Lawrence, 1971). In this study,
pin flocs were also found in the clarifier of the CAS (SRT of 10 days).

Table 4. Summary of performance of the CAS and MBR


CAS MBR
SS (mg/L) 4 – 41 (18) N.D
VSS (mg/L) 3 – 28 (12) N.D
tCOD (mg/L) 19.6 – 68.2 (42.7)
6.8 – 37.5 (23.2)
sCOD (mg/L) 10.5 – 34.3 (21.2)
tBOD5 (mg/L) 1.7 – 13.9 (6.3)
0.1 – 1.5 (0.8)
sBOD5 (mg/L) 0.7 – 2.3 (1.3)
NH4+-N (mg/L) N.D N.D
NO3—N (mg/L) 25.6 – 58.0 (41.3) 23.3 – 47.1 (33.9)
N.D non detectable

No ammonia was detected in the effluent from both CAS and MBR systems, which showed that
complete nitrification could be readily achieved. Membrane fouling was observed from Day 80
onwards and the membranes finally reached the allowable transmembrane pressure of -20 kPa on
Day 112. Average MLVSS concentration was 10,500 mg/L.

CONCLUSIONS

The results have shown that the IAATP has a great potential to replace existing conventional
activated sludge systems for municipal wastewater treatment in sub-tropical and tropical regions.
The cost of treating municipal wastewater could be potentially reduced through biogas
production, reduction of aeration requirements and decrease in the amount of sludge needed to
be disposed of.

The anaerobic systems were able to achieve more than 80% SS and VSS removal, more than
70% tCOD removal and more than 77% tBOD5 removal during the start-up period. The UASB
was found to require the shortest start-up time of 80 days and it achieved the highest rate of
methane production (0.16 L/g tCOD removed). When the UASB was operated at a long HRT of
16 h without intentional sludge wasting, sludge floatation phenomenon was encountered which
would reduce removal efficiencies and require recirculation of the biomass from the top to the
base of the UASB.

At a HRT of 6 h, the UASB achieved lower SS and VSS removal efficiencies than those of the
AnSBR and AF, due to period washouts of biosolids when no intentional sludge wasting was
carried out. The UASB, however, had the highest biogas production (0.13 L/g tCOD removed),
which could be attributed to its well structured microbial community along the height of the
UASB and higher mixed liquor concentration.

Both the aerobic post-treatment systems, CAS and MBR, treating the effluent of the UASB were
able to achieve high quality effluents that could readily meet secondary effluent discharge
standards in terms of solids, organics and ammonia concentrations. The effluents may be further
treated with advanced treatment process such as reverse osmosis and disinfection for direct non-

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3214
WEFTEC®.06

potable or indirect portable use. On comparison, the MBR seemed to be a more favorable
aerobic system because it required a smaller footprint and could achieve higher organic removal
efficiencies.

REFERENCES

APHA. (1998) Standard Methods for the Examination of Water and Wastewater, 20th ed.,
Washington, D.C., U.S.A.

Barbosa R.A., G.L. Sant’Anna Jr, (1989) Treatment of Raw Domestic Sewage in an UASB
Reactor, Wat.Res. 23 (12) 1483-1490.

Behling E., A. Diaz, G. Colina, M. Herrera, E. Gutierrez, E. Chacin, N. Fernandez & C. F.


Forster. (1997) Domestic Wastewater Treatment Using A UASB Reactor, Bioresource
Technology 61 239-245.

Bisogni, James J. Jr. and Alonzo Wm. Lawrence, (1971). Relationships between biological solids
retention time and settling characteristics of activated sludge. Water Research 51 753-763.

Chernicharo C. A. L. and R. M. G. Machado. (1998) Feasibility of the UASB/AF system for


domestic sewage treatment in developing countries. Water Science & Technology 38 (8-9)
325-332.

Dague, R. R. and Pidaparti, S. R. (1991) Anaerobic Sequencing Batch Reactor Treatment of


Swine Wastes. In Proceedings of the 46th Industrial Waste Conference, Purdue university,
West Lafayette, Indiana, 751-760.

Lettinga, G., Hulshoff Pol, L.W., Koster, I.W., Wiegant, W.M.,Zeeuw, W.J.de, Rinzema A.,
Grin, D.C., Roersma, R.E. and Hobma, S.W. High rate anaerobic wastewater treatment using
the UASB reactor under a wide range of temperature conditions. Biotech. Genetic Eng. Rev.,
2, pp. 253-284. 1984.

Mergaert, K., B. Vanderhaegen & W. Verstraete. (1992) Applicability and Trends of anaerobic
pre-treatment of municipal wastewater. Wat. Res. 26 (8) 1025-1033.

Metcalf & Eddy. (2003) Wastewater Engineering Treatment and Reuse, 4th ed., McGraw-Hill
Companies, Inc. 582, 991, 1000.

Ng W. J. and K. K. Chin. (1986) Random-packed anaerobic filter in piggery wastewater


treatment. Biological Wastes. 20 157-166.

Ng, W. J. (1989) A Sequencing Batch Anaerobic Reactor for Treating Piggery Wastewater.
Biological Wastes. 28 39-51.

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3215
WEFTEC®.06

Sung, S. and Dague, R. R. (1995) Laboratory Studies on the Anaerobic Sequencing Batch
Reactor. Water Environment Research. 67 (3) 294-301.

Zhou, X. F. The effects of ammonia-N on AnSBR performance. M.Eng. Thesis, National


University of Singapore, pp. 140. 1999.

Copyright ©2006 Water Environment Foundation. All Rights Reserved

3216
View publication stats

Potrebbero piacerti anche