Download as pdf or txt
Download as pdf or txt
You are on page 1of 32

Authors Accepted Manuscript

Influences of liquid, solid, and gas media


circulation in anaerobic membrane bioreactor
(AnMBR) as a post treatment alternative for aerobic
system in seafood industry

Sumate Chaiprapat, Araya Thongsai, Boonya


Charnnok, Watsa Khongnakorn, Jaeho Bae
www.elsevier.com/locate/memsci

PII: S0376-7388(16)30086-2
DOI: http://dx.doi.org/10.1016/j.memsci.2016.02.029
Reference: MEMSCI14295
To appear in: Journal of Membrane Science
Received date: 13 December 2015
Revised date: 4 February 2016
Accepted date: 12 February 2016
Cite this article as: Sumate Chaiprapat, Araya Thongsai, Boonya Charnnok,
Watsa Khongnakorn and Jaeho Bae, Influences of liquid, solid, and gas media
circulation in anaerobic membrane bioreactor (AnMBR) as a post treatment
alternative for aerobic system in seafood industry, Journal of Membrane Science,
http://dx.doi.org/10.1016/j.memsci.2016.02.029
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Influences of liquid, solid, and gas media circulation in anaerobic membrane bioreactor

(AnMBR) as a post treatment alternative for aerobic system in seafood industry

Sumate Chaiprapata,b*, Araya Thongsaia, Boonya Charnnokb, Watsa Khongnakorna, Jaeho

Baec

a
Energy Technology Research Center, Department of Civil Engineering, Faculty of

Engineering, Prince of Songkla University, Hat Yai Campus, Hat Yai, Songkhla 90110,

Thailand
b
PSU Energy System Research Institute (RERIN), Prince of Songkhla University, Hat Yai

Campus, Hat Yai, Songkhla 90110, Thailand


c
Department of Environmental Engineering, Inha University, Namgu, Yonghhyun dong 253,

Incheon, Republic of Korea


*
Corresponding author at: Department of Civil Engineering, Faculty of Engineering, Prince

of Songkla University, Hat Yai Campus, Hat Yai, Songkhla 90110, Thailand. Fax: +66 74

459396. E-mail address: sumate.ch@psu.ac.th

Abstract

Anaerobic membrane bioreactor was used to treat the effluent from an industrial

UASB reactor receiving seafood processing wastewater. Operation in liquid circulation (LQ)

was compared with the addition of granular activated carbon (LQ+GAC) and the introduction

of biogas to partially replace the liquid recirculation flow (LQr+GAC+G). In all treatments,

COD was removed to well under the secondary effluent standard of 120 mg/L. HRT did not

cause much difference in performance but GAC addition provided superior performance in

membrane fouling control. Fouling rate, the transmembrane pressure developed over time in

1
mbar/d, was reduced by 90, 93, and 87 percent in LQ+GAC, and the additional of 54, 55, and

90 percent with biogas circulation at HRT 4, 6, and 8 hr, respectively. Minimal substitution

of liquid with biogas in circulation with GAC was effective and less energy intensive. SEM

image showed that GAC and partial biogas supply could shift the fouling behavior.

Keywords: AnMBR; Fouling; GAC; Biogas circulation; Transmembrane pressure

1. Introduction

Nowadays, rapid expansion in seafood industry has taken a toll on a country power

supply due to its intensive energy consumption, which at the same time inherited as a part in

heightened product cost to consumers. Thailand has a very large food sector and seafood

processing is one of the most profitable industries. In 2014, the seafood industry generated

the export income for 6436.3 million US dollars which was first among agricultural products

in the country [1]. Not only Thailand, this situation is similar in many other countries with

long coast lines where fishery is abundant. Wastewater treatment in seafood processing relies

on the aerobic treatment where air is forced into wastewater to allow gas transfer that delivers

oxygen to microbes within the aeration unit. This practice takes a tremendous cost of

electrical energy and may become un-replicable in developing countries [2]. Anaerobic

treatment became an alternative as it requires less energy input and methane can be retrieved

and used for various purposes as energy source. However, quality of the effluent from

anaerobic digester still does not meet the industrial effluent standard, for instance, Thailand

BOD 20 mg/L, COD 120 mg/L and SS 50 mg/L [3], or USA BOD 20 mg/L and SS 20 mg/L,

which still requires additional treatment. The post treatment normally includes activated

sludge process or aerated lagoon in case land is limited, or otherwise a huge constructed

wetland or oxidation pond system. One of the reasons for this constraint is the difficulty of

2
anaerobes to remove the non-VFA soluble, and the colloidal matter which are the majority of

COD in the effluent; i.e. a makeup of 60-70 percent in the effluent of an anaerobic reactor

[4]. These constituents can easily be removed in a subsequent aerobic culture. Cases of

successful anaerobic wastewater treatment to produce effluents below the secondary effluent

standard of BOD 20 mg/L were reported [5, 6] but they are extremely rare especially in stable

full scale operations. There is, thus, much room for improvement of the anaerobic technology

as a mean for ultimate treatment of wastewater.

Anaerobic membrane bioreactor (AnMBR) is a reactor design which provides

anaerobic microorganisms with long or close to infinite solid retention time (SRT) at a short

hydraulic retention time (HRT). Better removal of organics by membranes superior physical

separation is coupled with the biodegradability of the residing microbes [7]. With this,

AnMBR has gained more attention as a potent total wastewater treatment solution, especially

in a low strength range. However, the persisting obstacle from a deposition of particles onto

the membrane surface as well as in pores is its great concern since it would lead to an

increased transmembrane pressure (TMP) and reduction of the permeate flux, which would

then require membrane washing through hectic physical and chemical cleanings. Three-stage

TMP profile; short term rapid rise, extended slow rise, and rapid rise was widely observed in

the constant flux operations [8, 9]. Extension of operating time of membrane by controlling

fouling is, therefore, crucial since frequent cleaning not only involves expenses and interrupts

the operation, but can also damage the membrane, reducing its service lifetime.

Reduction of a fouling rate could be achieved by operating membrane under the

critical fluxes, which typically are very low and hence not cost-effective, or providing the

shear across its surface [10]. Shear force acting onto the membrane surface contributed to the

reduced solids accumulation from soluble microbial products (SMP), microbial cells,

suspended solids, and inorganics to the membrane pores [11]. Addition of various media and

3
appropriate agitation patterns could be an effective means to relieve membrane fouling which

more research are needed. For instance, ion exchange resins and glass bead were tested and

showed good scouring effect in submerged anaerobic MBR (SAMBR). The permeate flux

and stability of the SAMBR were noticeably improved [12]. Silica and PET beads were also

tested with certain success in fouling reduction [13]. Activated carbon has recently emerged

as a remarkable media since it can adsorb organic solute to degrade over time, which

simultaneously regenerates its adsorptive surface [14]. Early report by Park et al. [15]

indicated that addition of powdered activated carbon (PAC) dosage of 5 g/L with some liquid

media circulation was critical in reducing the biomass cake resistance developed on the

membrane surface. Biogas bubbling underneath different types of membrane in AnMBR as a

means for in situ cleaning was introduced by Hu and Stuckey [16]. Later, Hu and Stuckey

[17] used activated carbon addition under only biogas recirculation in the AnMBR treating

dilute synthetic wastewater (COD 46020 mg/L). COD removal above 90% and the slower

progression of TMP development were observed. One of the interesting operations of

AnMBR is when the reactor content was fluidized with GAC by liquid circulation and used

as the second stage in the bioreactor system. This two-stage anaerobic digester configuration

could sustainably remove COD at 99 percent overall with a reasonably low energy

requirement [18]. Nevertheless, these experiments were carried out with synthetic wastewater

consisting of mostly soluble components. Only an ensuing experiment with municipal

wastewater as feed to AnMBR with liquid and GAC circulation was conducted [19]. Specific

application on industrial wastewater is still limited. Various literatures also indicated that the

use of different media either liquid, solid and gas circulation could help on fouling mitigation,

but to our knowledge the combination of the three media has never been tried. Combination

of liquid recirculation, activated carbon scrubbing and gas sparging could incrementally

enhance the performance of AnMBR, and the quantitative improvements are of particular

4
interest as post treatment alternative to the energy intensive aerobic systems of widely used

activated sludge process and aerated lagoon.

This study focused on evaluating the effects of medium recirculating by liquid only,

liquid plus GAC, and the introduction of additional biogas supply to circulation flow on the

AnMBR system. Digester effluent from a seafood processing plant was used as feed. The

AnMBR in this study mimicked a post treatment unit eyeing to reach the secondary effluent

standard, under different loadings and fluxes. Results from this study demonstrates

interesting information for the possible enhancement of this technology closer to industrial

application.

2. Materials and methods

2.1 Wastewater collection and preparation

Effluent from a full scale upflow anaerobic sludge blanket (UASB) reactor of a

seafood processing plant in Songkhla Province, Thailand was used as feed in this experiment.

The UASB reactor was under stable operation during the study period. Fresh effluent was left

to settle and screened by a thin sheet fabric cloth of approximately 100 mesh opening to

remove large particles and scum. This procedure was done in order to imitate the effluent

passing through a sedimentation tank and fine screen which would be required if a membrane

reactor is to be employed. The effluent was then kept in a refrigerator at 4 oC until use.

2.2 Reactor configuration and operation

Setup of the anaerobic membrane bioreactor (AnMBR) system in this study is shown

in Fig. 1. The reactor composed of the main body made of a 100 cm long transparent PVC

tube with an inside diameter of 5 cm, and the settler located at the top having an inside

diameter of 10 cm and height 30 cm. The main reactor possessed a total working volume of

5
1.96 L. Each reactor contained a submerged membrane module consisting of 24 separate

polyvinylidenefluoride (PVDF) hollow membrane fibers. Each fiber is 95 cm long with an

outside diameter of 1.5 mm (Shanghai Jofur Advanced Materials Co. Ltd., China), and has a

nominal pore size of 0.1 m. The total membrane surface area in the reactor was 0.1075 m 2.

The storage tank was connected to the settler to enable additional solid separation and

intermediate storage for liquid recirculation.

Figure 1

This study compared the performance of three identical AnMBRs adopting different

circulating regimens for fouling control, as summarized in Table 1. In the first reactor

(designated as LQ), the liquid from the storage tank (Fig. 1) was recirculated at 60 L/min

back to the bottom of the main reactor, while the second reactor (LQ+GAC) contained 25 g

of granular activated carbon (GAC) within the main reactor. GAC was 100% fluidized of the

reactor height with the specified flowrate. In the third reactor (LQr+GAC+G), the biogas

collected in a buffering balloon was recirculated in addition to the condition in the second

reactor. Liquid circulation flow was reduced by 10 L/min to keep GAC at the same

fluidization level. The GAC used in this experiment had a particle size of 1630 mesh,

density 0.48 g/cm3 and specific surface area 1000 m2/g (Carbokarn Co. Ltd., Thailand). The

feed and effluent streams were controlled by two individual peristaltic pumps (Masterflex,

Model No. 7528-10, USA). The permeate pump was connected to the top section of the

membrane fibers, whose flowrate was set to the designed HRT and membrane flux. The

influent pump was automatically controlled by a level sensor to maintain constant liquid level

in the AnMBR. Liquid and gas recirculations were also maintained with peristaltic pump.

6
The trans-membrane pressure (TMP) was monitored with a vacuum pressure gauge NUOVA

FIMA Model P25A-W, Italy.

Table 1

When TMP reached the set limits, the membrane module was removed from the

reactor and then cleaned. In the cleaning process, the membrane module was placed in a

cylinder and backwashed with DI water for 1 hr at pressure 1 bar. The module was then

submerged in series in 1% HCl, 0.1% NaOH, and 0.2% NaOCl, each at 1 hr. The module was

connected to the permeate pump and operated with DI water for 10 min to observe flux. To

ensure the integrity of membrane, the module was tested for leak by pumping air through the

membrane fibers at 1 bar under water. The membrane resistance (Rm) was assessed prior to

reusing the membrane in the experiment. Rm recovery value of at least 90% of the initial

membrane resistance of a new membrane was attained. If not, the above chemical cleaning

process was repeated.

2.3 Sample and data analyses

The influent and permeate were collected and analyzed for pH, turbidity, total COD

(TCOD), soluble COD (SCOD) and sulfate according to Standard Methods [20]. Total

alkalinity and total VFA were measured by the titration method [21] and total nitrogen and

ammonia nitrogen were analyzed calorimetrically following MERCK Pharo100

Spectrophotometer Manual. Biogas composition was determined using gas chromatography

(GC 7820A Agilent Technologies) equipped with thermal conductivity detector (TCD) where

helium was used as carrier gas. The standard calibration curve was made with gas mixtures

containing CH4 at 3 levels covering the range of 2099.999%, and verified with a standard

7
gas mixture of 5% N2, 60% CH4, and 35% CO2. The surface and cross section micrographs

were observed using scanning electron microscopy (SEM) model Quanta 400, FEI, Czech

Republic with an Everhart Thornley detector at voltage 15 kV. A portion of the membrane

fiber near mid depth of each reactor was cut at approximately 1.5 cm in length and placed in

2.5% glutaraldehyde for 2 hours in order to prevent fracture of the dried cake layer. The

specimens were then rinsed with ultra-pure water, dehydrated using a series of graded ethanol

washes (50-100%) before dehydrating with critical point drying method. The samples were

gold sputter coated before being examined.

Performance of AnMBR was mainly measured with regard to the removals of COD,

as well as nitrogen and sulfate according to Eq. 1. Total COD composes of soluble and

particulate portions, and its removal can therefore be separated accordingly (Eq. 2). The

soluble removal is calculated based on SCOD (soluble COD) removal efficiency while the

particulate removal was the result from subtracting the soluble removal from the total COD

removal according to Eq. 3. Mean and standard deviation values of the effluent (permeate)

were calculated from data at pseudo steady-state condition.

(1)

(2)

Total COD removal (%) = Soluble COD removal + Particulate COD removal (3)

where Ci is influent concentration (mg/L), Ce is permeate (effluent) concentration (mg/L),

TCODi is influent total COD (mg/L), TCODp is permeate total COD, SCODi is influent

8
soluble COD (mg/L), SCODe is permeate soluble COD (mg/L), PCODi is influent particulate

COD, and PCODe is permeate particulate COD. It is noted that PCOD of permeate was

assumed zero since the membrane pore size is 0.1 microns that filtered out all particulate

from the liquid

3. Results and discussion

3.1 Wastewater characterization

Table 2 shows characteristics of the UASB digestate used throughout the experiment.

pH of the UASB digestate in a range of 6.9-7.6 was appropriate for the treatment by AnMBR

as the full scale reactor had endured small pH variations in the daily wastewater from the

seafood processing plant. Sufficient alkalinity (759.8163.3 mg/L as CaCO3) was contained

in the digestate while VFA was rather low, under 166.0 mg/L as CH3COOH. This indicated

stable UASB operation. Although grease and large solid particles were removed by the solid-

gas-liquid separator installed within the UASB, simple screening to remove the solids and

scum prior to use as feed to AnMBR must be ensured. A simple filtration with fine nylon

screen would be sufficient to produce the same result as the fabric cloth screen used in this

study that allowed the suspended solids concentration around 75.230.3 mg/L.

Approximately one third of the nitrogen in the digestate still remained in a form of organic

nitrogen (TN - NH4+-N). Chemical oxygen demand both in TCOD and SCOD were in a range

under approximately 400 and 300 mg/L giving the TCOD/TN and SCOD/TN of 1.2:1 and

0.7:1, respectively. There was an ample supply of nutrients for this AnMBR post treatment

process. The presence of sulfate in the UASB effluent (102.1 mg/L), which would require

68.4 mg/L of COD for complete reduction, played a part in COD removal within the

AnMBR. It is interesting to note that sulfate reducing bacteria could compete only a small

portion of COD (see sulfate removal in Table 3) although the theoretical COD/SO42-

9
requirement is only 0.67 (Lens et al., 1998) compared to 2.5 available in the UASB effluent.

Perhaps, with a long SRT inside the AnMBR, sulfate reducing bacteria may not be as

competitive or dominant in this microbial community.

Table 2

3.2 Treatment performance of AnMBR

Results from continuous operation of AnMBR treating the full scale UASB effluent

from seafood industry over a period of 230 days were summarized in Table 3. The permeate

produced had higher pH mostly above 7.7 in all conditions tested. This was accompanied by

a much reduced VFA concentration while the alkalinity remained near the level of the

influent. The systems were well buffered indicated by the low VFA to alkalinity ratio. A

general recommended limit of VFA to alkalinity ratio of 0.4 for stable anaerobic digestion

[22] was still far above the ratio obtained in this study. TN was removed largely through the

biomass synthesis, adsorption to GAC when available, and probably organically bound with

particulate settlement. It should be noted that the removal of ammonia nitrogen (NH4+-N) was

observed at a level of 18-35 percent. Concurrent presence of nitrogen gas in a range of 30-43

percent was found in the biogas produced, similar to Yoo et al. [19] which reportedly came

from dissolved N2 in feed. Besides that, in our case, N2 detected in the biogas might be

produced from denitrification of the feed, which might have gone through nitrification while

stored in the feed tank exposing to air prior to feeding to the AnMBR. The analysis of nitrate

was unfortunately not carried out during the experiment.

Table 3

10
It was clear that the addition of GAC enhanced COD removals as the system

produced a lower COD permeate (Table 3). The permeate COD concentration was reduced

about one half of the control with only liquid circulation (LQ). However, the biogas supply

did not give further benefit or, in fact, even diminished removals of COD and other

pollutants. Attachment media provided to the microbial cells was a prime advantage GAC

provided. Porous structure and surface allowed microbial to attach and form the community

and reduced floc breakage [17]. This enabled the syntrophic relationship to form by allowing

acidogens and methanogens to cooperate efficiently in close contact while the shear force

was avoided. Moreover, the adsorption of soluble molecule onto the surface of GAC should

give the microbes higher chance to be in contact with the substrate [23] while the adsoptive

surface is regernerated [17].

It was also found that the COD removal efficiency did not correlate well with HRT as

normally expected in anaerobic bioreactors. Typically in biological treatment systems,

lowering HRT would result in higher organic and hydraulic loads to the system and their

performance would decline accordingly. It was not the case here. COD removals had a slight

tendency to increase at the lower HRTs. This phenomenon could be explained by the

accumulation of biomass within the reactors since sludge drainage was not performed

throughout the experiment. Lastly, it was inevitable that particulate COD removal by the

membrane was also gaining dominance in AnMBR under this high flux range. The

contribution of physical filtration seems to intensify toward the lowering HRT due to the

higher rate of foulant accumulation on the surface.

3.3 Contribution of soluble and particulate removals

Figure 2 shows the contribution of the soluble and particulate COD removals in the

systems according to the calculations in Eq. 2 and 3. Complete removal of particulate in the

11
influent was assumed because the membrane pore (0.1 microns) was smaller than that of the

GF/C filter paper used in the soluble COD and suspended solids tests. Although a small

portion of particulates may be solubilized and destroyed biologically, under short HRTs

tested in this study, the incoming particulates were minimally solubilized, so that an

instantaneous removal of particulate can approximate a physical removal efficiency by the

system. The biological removal efficiency of COD by the system, meanwhile, was estimated

based on SCOD (soluble COD) removed from the influent.

Figure 2

Physical contribution rose slightly when comparing at different HRTs in the same

circulating media. Without GAC (only liquid circulation, LQ in Fig. 2), the membrane

offered highest physical removal contribution. Biological contribution was higher with GAC

addition resulted from the microbial accumulation previously described. Additional biogas

supply in the circulating stream did not give more removal and seemed to disturb

methanogenic activity in the reactors. It was found that methane composition in the biogas

reached 53.0, 55.4, and 55.3 at HRT 8, 6, and 4 hrs, respectively, all at LQ+GAC treatment.

Methane concentration in the biogas corresponded well to the biological contribution in the

AnMBR. However, due to the small biogas yield from the low strength wastewater treatment,

methane yield would not be a crucial determinant for optimal operating condition as long as it

can reach around 50 percent, which can be burn with the biogas from the preceding main

digester. The more interesting factor would be the cost of AnMBR operation which is directly

related to the cleaning frequency of membrane module. This topic is discussed in the

following sections.

12
3.4 Transmembrane pressure profile

The transmembrane pressure (TMP) was measured throughout the experimental

period and primarily set at the limit of 0.3 bars (Fig. 3). TMP in reactor LQ reached the limit

at approximately 32 days in the last 3 cycles at HRT 8 days while it took 159 days for reactor

LQ+GAC to reach TMP 0.1 bars. Thus, TMP of 0.1 bars was set as an end point for reactors

LQ+GAC and LQr+GAC+G before membrane cleaning. However, the number of days for

reactor LQ to arrive at 0.1 bars was also recorded for cross comparison. Reactor

LQr+GAC+G was started at day 61 and run for the next 100 days at HRT 8 hr only to see

very little TMP development with a stable performance. Visual observation confirmed a light

color on the membrane surface compared to the same running time as reactor LQ+GAC,

therefore, all three reactors were cleaned and the HRT was decreased to 6 hrs at once.

Figure 3

At HRT 6 hrs (from day 160 to 211), shorter fouling cycle was observed. TMP in

reactors LQ and LQ+GAC reached 0.1 bars in 3.6 (average) and 51 days, respectively, while

it only reached 0.04 bars during 51 days in reactor LQr+GAC+G. The TMP profile of reactor

LQ+GAC rose rapidly in the first 11 days (day 160-171) (22% of a cycle) and stable or very

slowly rising for the next 28 days (55% of a cycle) before quickly elevated to 0.1 bars in the

last 12 days (23% of a cycle). This three-stage TMP developing pattern followed (1) pore

clogging by colloids and soluble products penetration, (2) deposition or development of

sludge cake, and finally (3) sludge cake consolidation (compression). The last stage was

usually the most rapid as there was always an effect of local flux resulted from the non-

uniform deposition of foulants making the flux in some area on membrane surface higher

than the critical flux [24]. Scouring by moving GAC particles appeared to control the cake

13
formation efficiently by slowing down the second and third stages, virtually no increase of

TMP, for a substantial amount of time. Substitution liquid circulation with partial biogas

supply was even more superior with such a small biogas flow rate (60 mL/min) as seen by the

large gap between LQ+GAC vs. LQr+GAC+G profiles as illustrated in Fig. 3.

At HRT 4 hrs, however, only 1.0, 8.5, and 18.0 days were spent in reactors LQ,

LQ+GAC, and LQr+GAC+G for TMP to arrive at TMP 0.1 bars. The effect of low HRT

(high flux) on the TMP evolution profile was consistent with other works such as Huang et

al. (2011) on submerged anaerobic membrane bioreactor (SAnMBR) and Gao et al. (2014) on

integrated anaerobic fluidized-bed membrane bioreactor (IAFMBR). At flux too low (HRT 8

hr) or too high (HRT 4 hr), this three-stage TMP profile was not clearly visible, although it

existed. Nevertheless, in this experiment, it was clear that substituting liquid circulation with

a fraction of biogas supply flow was able to extend the membrane fouling time by 2.1 times

compared to LQ+GAC at HRT 4 days and the difference was even more apparent at other

HRTs tested. Superficial velocity of fluid near the membrane surface was an important

parameter to control cake formation [25] and it can be created with either gas or liquid

turbulent flow, or movement of the membrane module itself [26]. It was obvious that partial

biogas supply in exchange of the liquid circulation flow was effective in this regard.

3.5 Membrane fouling rate

In order to compare across the three operating methods, the average membrane

fouling rate was calculated by the change of TMP over time, as mbar/d (Table 4). The values

in Table 4 are based on the TMP limit of 0.1 bars or noted otherwise. The membrane fouling

rate was directly related to the cleaning frequency of AnMBR operation. It also implies

operating and maintenance costs and longevity of the membrane in service. The more

frequent cleaning could damage the membrane at faster rate. It was found that the highest

14
fouling rate was in the liquid only circulation (reactor LQ) followed by LQ+GAC and

LQr+GAC+G. GAC addition prolonged the fouling rate by 90, 93, and 87 percent at HRT 4,

6, and 8 hr, respectively. The biogas circulation further helped reduce the fouling rate by

another 54, 55, and 90 percent at each HRT. With a limited literature report on fouling rate in

AnMBR, our fouling rate in LQr+GAC+G of 0.1 at HRT 8 hr is comparable with those

reported in Martinez-Sosa et al. [27] at 0.14 mbar/d under only biogas sparging mesophilic

(35oC) submerged AnMBR at 7 L/m2.hr. Another approach using rotating membrane module

in AnMBR was tested and yielded 388 mbar/d (0.45 10-3 kPa/s) at flux 10 L/m2.hr, 200 rpm

under MLSS 17.32 g/L. Noted that our MLSS excluding GAC was in a range of 1.0-1.4 g/L.

Since fouling rate depends on MLSS within the reactor [26], thus, direct comparison with

other works is difficult. However, it was obvious that the fouling rate achieved in this

experiment at only 0.1-6 mbar/d in LQr+GAC+G was considerably low.

Table 4

The processing throughput of the AnMBR in terms of volume of processed

wastewater corresponded with each HRT could be used as design criteria to assess membrane

fouling. Thus, the specific fouling rate in terms of mbar per liter of wastewater processed per

square meter of membrane surface (mbar/Lpermeate/m2) was computed and illustrated in Fig. 4.

HRT significantly affected the fouling rate since the higher flux induced the solid

accumulating rate onto the membrane surface at the rate much greater than the repelling

effect of liquid circulation alone could handle. This caused rapid development of TMP across

the membrane. Although there were limited test conditions, i.e. 3 flux rates, the trend lines

(Fig. 4), generated from exponential regression, fitted the data well at R2 0.999, 0.903, and

0.995 for LQ, LQ+GAC, and LQr+GAC+G, respectively. Thus, it is highly probable that

15
performance at other conditions within this flux range shall fall into this pattern. Noted that

the operation at HRT shorter than 4 hr was deemed not very feasible for anaerobic

biodegradation. Operation at HRT 4 hr is perhaps near the boundary of AnMBR compared to

other literatures [10, 14, 28].

Figure 4

At HRT 4 hr, scouring effects by GAC reduced the fouling by 9.9 times from 102.1 to

10.3 mbar/Lpermeate/m2 and the minimal biogas substitution in liquid circulation helped cut that

down about half of the remaining to 4.7 mbar/Lpermeate/m2. At HRT 6, the biogas circulation

could reduce fouling rate at 55 percent (2.37 to 1.07 mbar/L/m2) but reaching as high as 90

percent (1.58 to 0.16 mbar/L/m2) at HRT 8 hr. A small biogas supply of 60 mL/min or 30.6

mL/L/min (1.84 m3biogas/m3reactor/hr) in exchange (or substitution) with the liquid flow of 10

L/min is equivalent to only 0.6% substitution by volume could at least double the operating

time of the AnMBR. Energy and membrane cleaning costs could certainly be reduced by the

smaller specific fouling rate.

3.6 Energy consumption for biogas supply scheme

While energy balance for AnMBR system has been well covered elsewhere [13, 18,

19], this present study focused on the energy requirement for 60 mL/min of biogas supply in

LQr+GAC+G. Power requirement equation [18] P=Q..E/1000 is used where P is power

requirement (kW), Q is flow rate (m3/s), = 1.21*9.81 = 11.9 N/m3 for biogas at CH4 50%,

and E is head loss (m). In reactor LQr+GAC+G, the measured head loss was around 0.37 m.

For biogas supply, the head loss was estimated equal to the water depth 1.2 m, which

required the added power of only 1.42 10-8 kW. Dividing by the permeate flow rate of 11.8

16
L/d (4.9 10-4 m3/h) at HRT 4 hr yields an estimated energy of 2.91 10-5 kWh/m3. This

value is much less than the energy required for liquid pumping and total energy requirement

for our fluidized AnMBR of 0.028 kWh/m3 excluding the potential energy production from

methane generated [18]. Further reduction in energy for gas supply in our configuration could

be achieved by alternate on-off or intermittent cycle as demonstrated in earlier works [29-31].

It must be realized that the hydrodynamics in the reactor which is constituted from various

factors such as component arrangement, media characteristics, suspension concentration and

etc. will greatly affect the energy requirement for AnMBR operation and its fouling.

Nevertheless, in this present study, the potential cost saving on AnMBR operation with

combined liquid and biogas was clear.

3.7 Analysis of cake layer morphology

The membrane specimens were examined by scanning electron microscope (SEM)

after the last fouling at the end of experiment (HRT 4 hr). Fig. 5 shows the SEM photographs

of the virgin membrane and the fouling membrane cross section and external surface. Heavy

biofouling was observed in reactor LQ whose deposit composed mainly of extracellular

polymer substances (EPS) and soluble microbial products (SMP). Microbial attachment on

the outer surface of the membrane was apparent in Fig. 5-b4. With GAC addition, thinner

cake layer was formed on the membrane surface (Fig. 5-c) whereas application of biogas

circulation gave the thinnest cake layer and clearly lower microbial attachment. It must be

noted that the reactor LQr+GAC+G was operated, after the previous cleaning, for 18 days

whereas reactors LQ+GAC and LQ were in operation for only 11 days and 1 day,

respectively, before the membrane were sampled for SEM analysis. The average thicknesses

of cake were 2.59, 0.97, and 0.50 m for reactors LQ, LQ+GAC, and LQr+GAC+G,

respectively.

17
Figure 5

More active microbial cells in the reactor although can give high metabolism and

pollutant degradation, they could foul the membrane quite easily without enough shear force,

as seen in reactor LQ in this study. The cake layers in LQ+GAC and LQr+GAC+G were

likely to come from organic adsorption to the surface, also known as organic fouling [10] and

inorganic chemicals such as phosphate and struvite which was the main foulant identified in

many anaerobic membrane bioreactors [32, 33]. It was very likely that the addition of GAC

and biogas circulation would shift the fouling behavior of the AnMBR. Further analysis with

energy dispersive X-ray (EDX) could provide an insight information of the chemical

composition of the inorganic precipitate on the membrane surface, but this aspect was not

included in the scope of current study. To say the least, struvite, a complex chemical

precipitate, and more simple precipitates of other ions such as phosphorus, calcium,

aluminum, iron, carbonate and etc. are dependent mainly on the characteristics of wastewater

to be treated and the conditions applied. These inorganic foulants found could always be

traced back to the source in wastewater, thus, perhaps primary removal of these constituents

may be required. More details are discussed in Liao [10], Stuckey [14], and Wang et al. [34].

The strong shear force applied on the membrane surface could help prevent the

formation of cake layer but at an expense of energy input to the system operation. Matching

and balancing of the biogas circulation, which is cheaper due to the smaller flow rate needed

and lower energy requirement, in exchange of the liquid circulation, which has higher density

and requires more energy for pumping, are of interest for further research. However, strong

shear force may cause a negative impact by the solid particle abrasion on the surface causing

a thinning of active layer. An obvious active layer was seen in the virgin membrane (Fig. 5-a)

with a dense section at the outer surface while this dense layer was not so apparent in the

18
used membranes (Fig. 5-c3 and 5-d3). This phenomena should be studied in detail in various

types of solid medium scouring in AnMBR operation. Besides, the microbial community

established on the surface of membrane, GAC granule, and in suspension could shed light on

the biofouling characteristics and biological performance of AnMBR. Previous works using

terminal restriction fragment length polymorphism (T-RFLP) and denaturing gradient gel

electrophoresis (DGGE) and other molecular techniques on AnMBR have shown interesting

results by the shifts in species dominance at different locations within the reactor. More

recently, the possibility of existence of the anaerobic ammonium oxidation (Annmmox)

species in submerged AnMBR for nitrogen removal has been confirmed with nitrogen

removal up to 85 percent [35, 36]. Molecular work in AnMBR is one of the promising

research themes in membrane bioreactors.

4. Conclusions

AnMBR was able to treat COD of a full scale UASB effluent of a seafood processing

plant to under secondary effluent standard. Membrane fouling was reduced by the

fluidization of GAC in the AnMBR. Small substitution of liquid circulation with biogas

supply (0.6% v/v) for GAC fluidization further reduced the membrane fouling, although did

not result in better COD removal. This could reduce the cleaning frequency of the membrane

and lower the operating cost. Effective prevention of biofouling was evidenced in SEM

photographs. Partial substitution of liquid circulation by small biogas circulation can be

further optimized.

Acknowledgements

This research was supported by the Annual Research Budget of the Prince of Songkla

University (PSU) contract no. ENG540664S, PSU Graduate School, and the Ministry of

19
Energy, Thailand. The authors would also like to recognize the supports from the Biogas and

Biorefinery Research Laboratory, Energy System Research Institute of Prince of Songkla

University (PERIN), and Kiang Huat Sea Gull Trading Frozen Food Public Co. Ltd.

References

[1] The Bank of Thailand, http://www.bot.or.th, in: Accessed 05/24/2015.

[2] K. Kavitha and A.G. Murugesan, Efficiency of upflow anaerobic granulated sludge

blanket reactor in treating fish processing effluent, J. Ind. Pollut. Control 23 (2007) 77-

92.

[3] Pollution Control Department, http://www.pcd.go.th/info_serv/reg_std_water04.html, in:

Accessed 01/31/2016.

[4] Yoda, M, Hattori, M., Miyaiji, Y, Treatment of municipal wastewater by anaerobic

fluidized bed: behaviour of organic suspended solids in anaerobic treatment of sewage,

in: Proceedings of Seminar/Workshop: Anaerobic Treatment of Sewage Amherst, Mass,

U.S.A. (1985) 161197.

[5] U.J. Ndon, R.R. Dague, Low temperature treatment of dilute wastewaters using the

anaerobic sequencing batch reactor, in: 49th Purde Industrial Waste Conference, 1994.

[6] R.E. Speece, Anaerobic Biotechnology and Odor/Corrosion Control for Municipalities

and Industries, Archae Press, Nashville Tennessee, 2008.

[7] P. Berube, E. Hall, P. Sutton, Parameters governing permeate flux in an anaerobic

membrane bioreactor treating low-strength municipal wastewaters: a literature review,

Water Environ. Res. 78 (2006) 887-896.

[8] H. Lin, W. Peng, M. Zhang, J. Chen, H. Hong, Y. Zhang, A review on anaerobic

membrane bioreactors: Applications, membrane fouling and future perspectives,

Desalination. 314 (2013) 169-188.

20
[9] J. Zhang, H.C. Chua, J. Zhou, A.G. Fane, Factors affecting the membrane performance in

submerged membrane bioreactors, J. Membr. Sci. 284 (2006) 54-66.

[10] B.Q. Liao, J.T. Kraemer, D.M. Bagley, Anaerobic membrane bioreactors: applications

and research directions, Crit. Rev. Env. Sci. Tec. 36 (2006) 489-530.

[11] D.W. Gao, T. Zhang, C.Y.Y. Tang, W.M. Wu, C.Y. Wong, Y.H. Lee, D.H. Yeh, C.S.

Criddle, Membrane fouling in an anaerobic membrane bioreactor: differences in relative

abundance of bacterial species in the membrane foulant layer and in suspension, J.

Membr. Sci. 364 (2010) 331-338.

[12] A. Akram, Stability and performance improvement of a submerged anaerobic membrane

bioreactor (SAMBR) for wastewater treatment. Ph.D thesis, Chemical Engineering

Dept., Imperial College London, 2007.

[13] M. Aslam, P.L. McCarty, J. Bae, J. Kim, The effect of fluidized media characteristics on

membrane fouling and energy consumption in anaerobic fluidized membrane

bioreactors, Sep. Purif. Technol. 132 (2014) 10-15.

[14] D.C. Stuckey, Recent developments in anaerobic membrane reactors, Bioresour.

Technol. 122 (2012) 137-148.

[15] H. Park, K.H. Choo, C.H. Lee, Flux enhancement with powdered activated carbon

addition in the membrane anaerobic bioreactor, Sep. Sci. Technol. 34 (1999) 2781-2792.

[16] A.Y. Hu, D.C. Stuckey, Treatment of dilute wastewaters using a novel submerged

anaerobic membrane bioreactor, J. Environ. Eng. 132 (2006) 190-198.

[17] A.Y. Hu, D.C. Stuckey, Activated carbon addition to a submerged anaerobic membrane

bioreactor: effect on performance, transmembrane pressure, and flux, J. Environ. Eng.

133 (2007) 73-80.

[18] J. Kim, K. Kim, H. Ye, E. Lee, C. Shin, P.L. McCarty, J. Bae, Anaerobic fluidized bed

membrane bioreactor for wastewater treatment, Environ. Sci. Tech. 45 (2011) 576-581.

21
[19] R. Yoo, J. Kim, P.L. McCarty, J. Bae, Anaerobic treatment of municipal wastewater

with a staged anaerobic fluidized membrane bioreactor (SAF-MBR) system, Bioresour.

Technol. 120 (2012) 133-139.

[20] APHA, AWWA, WPCF, Standard Methods for Examination of Water and Wastewater,

American Public Health Association, Washington, DC, 2012.

[21] R. DiLallo, O.E. Albertson, Volatile acids by direct titration, J. Water Pollut. Con. Fed.

33 (1961) 356-365.

[22] S.K. Khanal, Anaerobic Biotechnology for Bioenergy Production. Principles and

Applications. John Wiley & Sons, Ltd. Publication, USA. (2008).

[23] . Akta, F. een, Bioregeneration of activated carbon: a review. Int. Biodeter.

Biodegr. 59 (2007) 257-272.

[24] B.D. Cho, A.G. Fane, Fouling transients in nominally sub-critical flux operation of a

membrane bioreactor, J. Membr. Sci. 209 (2002) 391-403.

[25] H.C. Chua, T.C. Arnot, J.A. Howell, Controlling fouling in membrane bioreactors

operated with a variable throughput, Desalination 149 (2002) 225-229.

[26] I. Ruigmez, L. Vera, E. Gonzlez, G. Gonzlez, J. Rodrguez-Sevilla, A novel rotating

HF membrane to control fouling on anaerobic membrane bioreactors treating

wastewater, J. Membr. Sci. 501 (2016) 45-52.

[27] D. Martinez-Sosa, B. Helmreich, T. Netter, S. Paris, F. Bischof, H. Horn, Anaerobic

submerged membrane bioreactor (AnSMBR) for municipal wastewater treatment under

mesophilic and psychrophilic temperature conditions, Bioresour. Technol. 102 (2011)

10377-10385.

[28] J. Bae, C. Shin, E. Lee, J. Kim, P.L. McCarty, Anaerobic treatment of low-strength

wastewater: A comparison between single and staged anaerobic fluidized bed membrane

bioreactors, Bioresour. Technol. 165 (2014) 75-80.

22
[29] W.J. Gao, H.J. Lin, K.T. Leung, H. Schraft, B.Q. Liao, Structure of cake layer in a

submerged anaerobic membrane bioreactor, J. Membr. Sci. 374 (2011) 110-120.

[30] S.M. Lee, J.Y. Jung, Y.C. Chung, Novel method for enhancing permeate flux of

submerged membrane system in two-phase anaerobic reactor, Water Res. 35 (2001) 471-

477.

[31] I. Vyrides, D.C. Stuckey, Saline sewage treatment using a submerged anaerobic

membrane reactor (SAMBR): Effects of activated carbon addition and biogas-sparging

time, Water Res. 43 (2009) 933-942.

[32] J. Kim, C.H. Lee, K.H. Choo, Control of struvite precipitation by selective removal of

NH4+ with dialyzer/zeolite in an anaerobic membrane bioreactor, Appl. Microbiol.

Biotechnol. 75 (2007) 187-193.

[33] S.H. Yoon, I.J. Kang, C.H. Lee, Fouling of inorganic membrane and flux enhancement

in membrane-coupled anaerobic bioreactor, Sep. Sci. Technol. 34 (1999) 709-724.

[34] Z. Wang, J. Ma, C.Y. Tang, K. Kimura, Q. Wang, X. Han, Membrane clearning in

membrane bioreactors: a review, J. Membr. Sci. 468 (2014) 276-307.

[35] Z. Li, X. Xu, B. Shao, S. Zhang, F. Yang, Anammox granules formation and

performance in a submerged anaerobic membrane bioreactor, Chem. Eng. J. 254 (2014)

9-16.

[36] Z. Li, X. Xu, X. Xu, F. Yang, S. Zhang, Sustainable operation of submerged Anammox

membrane bioreactor with recycling biogas sparging for alleviating membrane fouling,

Chemosphere 140 (2015) 106-113.

23
Figure 1 Schematic diagram of the AnMBR system

24
100 60
HRT 8 hr HRT 6 hr HRT 4 hr
Contribution of COD Removal (%)
53.0
55.4 50.9 55.3 50.3
48.8
48.8
80 45

CH4 composition (%)


42.7
71.0
68.9 67.5 39.0
63.2 62.3 62.8 64.0 64.2
59.8
60 30

41.8
37.7 37.2
40 36.8 36.0 35.8 15
31.1 32.5
29.0

20 0

Physical Biological %CH4

Figure 2 Contribution of biological and physical removals in AnMBR operation at different


HRTs

25
0.40
HRT 8 hr HRT 6 hr HRT 4 hr

0.30
TMP (bar)

0.20

0.10

0.00
0 20 40 60 80 100 120 140 160 180 200 220 240
Time (days)

LQ LQ+GAC LQr+GAC+G

Figure 3 Evolution of TMP with time during operation under different HRTs of AnMBR
with (1) Liquid circulation (LQ), (2) Liquid circulation with GAC (LQ+GAC), and
(3) Liquid and biogas circulation with GAC (LQr+GAC+G)

26
Figure 4 Specific fouling rate of the AnMBR operated at different HRTs in (1) Liquid
circulation (LQ), (2) Liquid circulation with GAC (LQ+GAC), and (3) Liquid and
biogas circulation with GAC (LQr+GAC+G)

(a) Virgin membrane

a1 a2 a3 a4

(b) Fouled membrane in reactor LQ

b1 b2 b3 b4

27
(c) Fouled membrane in reactor LQ+GAC

c1 c2 c3 c4

(d) Fouled membrane in reactor LQr+GAC+G

d1 d2 d3 d4

Figure 5 SEM photographs of the virgin membrane in comparison with fouled membrane at

magnification (1) 150X cross section, (2) 2500X cross section, (3) 10000X cross section

and (4) 10000X external surface

Table 1Operating condition in the experiments


Unit Reactor
Condition
LQ LQ+GAC LQr+GAC+G
Granular activated carbon g/L
n/a 12.7 12.7
addition
Liquid recirculation flow rate L/min 60 60 50
Gas recirculation flow rate mL/min n/a n/a 60
HRT [Flux] hr
--------------- 8 [4.3] ---------------
[L/m2.hr]
--------------- 6 [5.7] ---------------
--------------- 4 [8.6] ---------------
n/a = not applied

28
Table 2 Characteristics of the UASB reactor effluent used as feed in the study (no. of
measurement = 24)
Parameters Unit Range MeanSD
pH 6.9-7.6 7.20.2
Turbidity NTU 23.2-124.0 56.230.6
Alkalinity mg/L as CaCO3 480.0-1,045.0 759.8163.3
Volatile fatty acids (VFA) mg/L as CH3COOH 60.0-166.0 91.129.7
Total COD (TCOD) mg/L 106.0-408.8 250.896.7
Soluble COD (SCOD) mg/L 67.0-296.3 133.962.2
Suspended solids (SS) mg/L 50.4-162.0 75.230.3
Total nitrogen (TN) mg/L 108.0-260.0 205.549.0
Ammonia nitrogen (NH4+-N) mg/L 65.8-188.0 138.037.6
Sulfate mg/L 67.7-158.9 102.129.2

Table 3 Performance of the AnMBR systems at different HRT


Parameters HRT 8 hr HRT 6 hr HRT 4 hr
LQ LQ+G LQr+G LQ LQ+ LQr+G LQ LQ+ LQr+G
AC AC+G GAC AC+G GAC AC+G

AMBR effluent characteristics


pH 7.90. 7.70. 7.80. 7.70 7.60 7.70. 7.90 7.80 7.70.
1 1 1 .1 .3 2 .2 .1 3
Turbidity 0.80. 0.60. 0.60. 0.80 0.60 0.60. 1.10 1.00 1.40.
(NTU) 2 1 1 .4 .1 1 .5 .2 3
Alkalinity 634.0 708.0 612.0 613.3 683.3 590.0 810.0 820.0 800.0
(mg/L as 42.8 64.2 71.9 11.5 37.9 52.9 20.0 26.5 26.5
CaCO3)
VFA (mg/L as 39.0 37.5 36.5 31.7 31.7 31.7 28.3 28.3 25.0
CH3COOH) 6.5 3.1 5.5 2.9 2.9 2.9 7.6 2.9 5.0
VFA/Alk 0.06 0.05 0.06 0.05 0.05 0.05 0.04 0.03 0.030.0
0.01 0.00 0.01 0.00 0.01 0.01 0.01 0.00 1
TCOD (mg/L) 68.2 29.4 37.6 36.5 18.4 26.9 48.5 25.3 31.2
30.5 16.9 5.6 8.1 0.0 1.3 0.0 1.5 1.6
TN (mg/L) 169.7 136.0 167.7 147.7 129.0 135.0 112.5 95.5 97.5

29
4.5 6.0 9.3 2.5 1.0 4.6 3.5 2.1 3.5
NH4+-N 151.3 127.0 131.5 134.3 121.7 124.7 92.5 76.0 81.0
(mg/L) 7.1 4.0 4.9 3.2 2.1 4.2 3.5 5.7 1.4
Sulfate (mg/L) 77.37 66.34 68.42 94.7 85.5 92.72 147.6 124.9 132.9
.1 .3 .8 7.1 4.3 .8 3.4 6.3 2.9
Removal efficiency (%)
TCOD 71.0 86.3 83.3 78.1 88.2 83.0 81.9 90.5 88.3
9.5 9.8 2.7 8.5 6.9 9.3 4.0 2.6 3.2
SCOD 57.4 79.5 65.1 65.5 83.8 83.7 73.0 84.9 84.0
15.4 9.4 5.6 3.3 2.4 2.4 9.5 4.4 5.5
TN 26.7 41.2 27.5 26.5 35.8 32.7 12.8 26.0 24.4
3.5 4.5 6.3 7.9 7.1 9.0 1.8 2.5 3.6
NH4+-N 22.3 34.6 32.3 17.9 25.7 24.0 21.0 35.1 30.8
5.8 8.1 8.0 9.4 7.8 5.9 1.1 3.3 0.5
Sulfate 21.9 32.9 30.7 17.0 24.8 18.8 20.1 32.4 28.0
6.3 6.0 6.4 2.0 3.5 1.7 2.8 4.2 2.4
CH4 42.6 51.3 47.8 42.6 54.4 53.4 39.0 55.3 50.3
composition 5.4 3.5 4.6 5.4 1.0 2.2 2.2 0.6 0.5
(%)

Table 4 Average membrane fouling rate in AnMBR in terms of delta TMP per time
(dTMP/dt)
Membrane fouling rate
Reactor (mbar/d)
HRT 8 hr HRT 6 hr HRT 4 hr
LQ 8 29 129
LQ+GAC 1 2 13
LQr+GAC+G 0.1* 0.9* 6

Note:* Calculated at TMP reaching 0.01 and 0.04 bars at HRT 8 and 6 hr, respectively

30
Highlights

AnMBR removed COD in digester effluent to below 68.2 mg/L

Addition of GAC in the AnMBR provided superior biological performance

Impact of HRT was minor on wastewater treatability but vital on membrane fouling

GAC addition to the AnMBR reduced fouling rate by 87-93%

Substituting circulating liquid with 0.6% biogas flow effectively reduced fouling

Graphical abstract

31

You might also like