Next Article in Journal
Experimental Investigations of Interactions between Sand Wave Movements, Flow Structure, and Individual Aquatic Plants in Natural Rivers: A Case Study of Potamogeton Pectinatus L.
Next Article in Special Issue
Risk-Yuck Factor Nexus in Reclaimed Wastewater for Irrigation: Comparing Farmers’ Attitudes and Public Perception
Previous Article in Journal
Building a Methodology for Assessing Service Quality under Intermittent Domestic Water Supply
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Treatment of Palm Oil Mill Effluent Using Membrane Bioreactor: Novel Processes and Their Major Drawbacks

by
Mohammed Abdulsalam
1,2,*,
Hasfalina Che Man
1,*,
Aida Isma Idris
3,
Khairul Faezah Yunos
4 and
Zurina Zainal Abidin
5
1
Department of Biological and Agricultural Engineering, Faculty of Engineering, Universiti Putra Malaysia, UPM Serdang, Selangor 43400, Malaysia
2
Department of Agricultural and Bioresources, Ahmadu Bello University, Zaria 810222, Nigeria
3
Department of Chemical Engineering, Segi University, Kota Damansara, Selangor 47810, Malaysia
4
Department of Food and Process Engineering, Faculty of Engineering, Universiti Putra Malaysia, UPM Serdang, Selangor 43400, Malaysia
5
Department of Chemical and Environmental Engineering, Faculty of Engineering, Universiti Putra Malaysia, UPM Serdang, Selangor 43400, Malaysia
*
Authors to whom correspondence should be addressed.
Water 2018, 10(9), 1165; https://doi.org/10.3390/w10091165
Submission received: 30 July 2018 / Revised: 20 August 2018 / Accepted: 21 August 2018 / Published: 30 August 2018
(This article belongs to the Special Issue Wastewater Treatment: Review, Key Challenges, and New Perspectives)

Abstract

:
Over the years, different types of alternative technologies have been developed and used for palm oil mill effluent (POME) treatment. Specifically, membrane bioreactor (MBR) has been employed to relegate pollutants contained in POME under different operating conditions, and the technology was found to be promising. The major challenge impeding the wider application of this technology is membrane fouling, which usually attracts high operating energy and running cost. In this regard, novel methods of mitigating membrane fouling through the treatment processes have been developed. Therefore, this review article specifically focuses on the recent treatment processes of POME using MBR, with particular emphasis on innovative processes conditions such as aerobic, anaerobic, and hybrid processing as well as their performance in relation to fouling minimization. Furthermore, the effects of sonication and thermophilic and mesophilic conditions on membrane blockage were critically reviewed. The types of foulants and fouling mechanism as influenced by different operating conditions were also analyzed censoriously.

1. Introduction

The industrial processing of oil palm involves a series of operations and these normally amounted to the immense generation of waste. The waste generated from palm oil processing have been broadly categorized into palm oil mill sludge (POMS) and liquid waste [1]. The POMS include the empty bunches, trunk, leaves, decanter cake, and mesocarps fibre. POMS constitutes about 70% of the total weight of the processed fresh fruit bunches [2]. While the liquid waste which is also called palm oil mill effluent (POME) is generated during crude palm oil (CPO) clarification, fresh fruit bunch (FFB) sterilization, and hydro-cyclone separation of a mixture of the shell and kernel [3]. Usually, sterilization is the first process, and it involves heating the FFB to a temperature of 140 °C for a specified time, ranging from 75 to 90 min in the autoclave machine to form a coagulated gummy solution called mucilage. At this stage, about 36% palm oil mill effluent (POME) is generated [4]. Afterward, the mucilage is conveyed for stripping, pressing and clarification processes where about 60% of the total POME produced was generated [2,5]. The produced crude palm is further purified and dried using hydro-cyclone separation and vacuum drying process respectively [2], and studies have shown that only 4% of the total POME is generated at this stage [6,7].
Quantitatively, reports have shown that 1 ton of crude palm oil requires 5–7.5 tons of water, over 50% of which ends up as POME [5]. This mammoth rate of POME generation has posed a considerable lethal to the environmental safety which resulted in the significant concern of the researchers over the years [7]. Intensive researches and innovations on POME treatment methods and technologies have been recorded globally. Since 1990, biological methods of treatment have been prominent among the palm oil millers, particularly in Malaysia [8]. This is due to the simplicity, cost-effective, and less requirement of technical know-how [9,10]. Many reports have substantiated the inefficiencies of the method such as prolonged treatment duration, large operational footprint, high tendencies of underground water contamination and emission of unpleasant odour gases such as sulphur dioxide (SO2) and methane (CH4) [3,11,12,13]. In addition, the treated effluent using this method have always failed to meet discharge standard limits of Malaysia, and it repeatedly needs periodic maintenance to discharge the accumulated sludge at the bottom of the pond [14]. Therefore, the inefficiencies of these methods couple with the stringency of the discharge standard necessitate alternative technologies [15].
Recently, membrane technologies have demonstrated potentials for POME treatment [16,17,18,19,20,21]. The membrane technology provides a higher quality of water and resources recovery, shorter period of treatment processes, requires a smaller area or footprints, and emits less odour [22]. However, this technology has a serious drawback—the susceptibility to fouling [23,24,25,26]. Fouling is the resulted challenge from the gradual accumulation of the colloids and particulates on the membrane pores walls and surface [27]. The narrowed pores and clouded membrane surface increased the hydraulic resistance, thereby intensifying the driving pressure required to conduct the liquid through the membrane. This effect usually attracts excessive operational energy and higher running cost which portrayed the treatment method unprofitable for industrial applications [28].
However, several attempts have been made to mitigate the problems of membrane fouling by using different types of approaches such as aeration, anaerobic process, novel combination of anaerobic-aerobic (hybrid) process, as well as the use of adsorbents, alternating filtration-relaxation during operation, and application of sonication and mesophilic and thermophilic conditions [19,29,30]. More recently, there have been quite a number of review articles on MBRs, such as the reports of Rana et al. [31], Coutte et al. [32], and Ohimain and Izah [33]. They have attempted to account for the contemporary application of MBR for wastewater treatment. Essentially, the reports pay emphasis on the types of MBR system, configurations and operating parameters such as temperature, organic loading rate and pH. Conversely, the review articles did not highlight the detailed effect of the treatments approaches on POME in relation to the membrane fouling which had been the major drawback of the technology. Besides, MBR technology is an evolving method for treating POME, as such an up-to-date review with a good comparison of different process conditions on the performance becomes indispensable. In view of this highlight, the major focus of this paper is to provide a holistic review on the recent processes for POME treatment with the main intention of understanding the effect of POME physicochemical properties as well as treatments processes such as aeration, anaerobic, integrated (hybrid) process (aerobic-anaerobic-anoxic), sonication, and temperature (thermophilic and mesophilic conditions). Also, the different types of the associated foulants and fouling mechanism were critically reviewed. The availability of such up-to-date information on novel treatment approaches will furnish a better thoughtfulness to redesign the existing MBR, thus improving its overall performance and reducing the propensity to fouling.

2. Physicochemical Properties and Biodegradability Index of POME

The POME generated during palm oil extraction is dark brownish in colour, viscous in nature, and also contains a high amount of water and suspended solids [34]. Mohammed and Chong [35] reported that the brownish colour is due to the high content of organic matter such as carotene (8 ppm), pectin (3400 ppm), tannin, phenolic (5800 ppm), and lignin (4700 ppm). Also, other reports have established that raw POME contains a considerable amount of carbohydrates, amino acids, free organic acids with pH ranging from 4.0–5.0 along with organic pollutants, fibres, and some inorganic nutrients such as iron, copper, potassium, magnesium, nitrogen, chromium, and cadmium [31,34]. Wang et al. [20] reported that chemical oxygen demand (COD) and biological oxygen demand (BOD) level in POME ranged from 15,000 to 100,000 mg/L and 10,250–43,730 mg/L, respectively. Accordingly, the presence of such a high concentration of organic pollutant (POME) in waterways could make it inhabitable for the aquatic lives [36,37].
POME may be characterized as high strength or low strength wastewater, depending on the number and concentration of contaminants confined. High strength POME holds a large amount of COD, total solids (TS), total suspended solids (TSS), ammonia, and inorganic nutrients as mentioned earlier [38]. This denotes that high strength POME has low biodegradability index (BI) as the constituted organic pollutants are predominantly COD and other toxic substances, such as the nitrogenous compound [39,40]. Therefore, discharging without proper treatment is hazardous to the environment. Concise information on the physicochemical properties and BI of various POME reported by different researchers is presented in Table 1.
First, BI is the ratio of BOD to COD elements contained in a given POME sample [37,41]. It measures the strength and treatments strain requirements [42,43]. The ratio of biodegradable elements (BOD) to non-biodegradable elements (COD) is a good index to measure the strength of wastewater and it also shows the degree of readiness to decompose biologically. This implies that POME with high BOD/COD ratio is considered readily biodegradable, that is, the organic pollutants can easily deteriorate biologically, and as such, it can be categorized as low strength. On the other hand, low BOD/COD ratio contains toxic substances which are inhibitors to biodegradation, such type of wastewater is measured high strength [43,44,45,46]. From Table 1, the BI ratio of the POME considered is in the range 0.36 to 0.617 [46,47,48,49]. POME with 0.617 of BI is more readily decomposed because of the high content of biodegradable substances such as BOD, while 0.36 BI indicates the prevalence of non-biodegradables.
In a more wider scope, if the BI ratio is less than 0.5, there is a need for additional physical and/or chemical treatments [43,50]. Hence, any wastewater with low BI ratio may require supplementary treatments to the bioremediation, such as membrane technology, use of adsorbent or dissolved air flotation [39,47]. This clearly shows the important application of BI in deciding and development of suitable treatments approach for a particular wastewater. The availability of information on BI and physicochemical properties of POME could serve as design information to develop MBR that is less susceptible to technical challenges such as membrane fouling [48].

3. Membrane Bioreactor (MBR) Application for POME Treatment

In the last few years, intensive studies have been reported on the use of MBR for POME treatments [12,40,56,57,58,59,60]. Principally, the MBR treatment processes constitute membrane-filtration and bio-decomposition of organic substance present in the mixed liquor to form a more innocuous product, such as carbon dioxide (CO2) and water [61]. During the filtration process, the membrane retains the microbial flocs and any particles whose size is greater than the pores diameters within the concentrate and produce filtrate with more inoffensive substances [62]. This ensures the suitability of the MBR system to keep hold of the activated sludge in the reactor.
The treatment processes of MBR only requires small footprints because the aerated membrane filtration performs double functions as the secondary and tertiary clarifier, unlike the conventional treatment method that involves separate sludge and sedimentation tanks [63]. It has shorter hydraulic retention time, handles a higher organic loading rate, and generates less sludge [64,65,66]. Furthermore, MBR technology is flexible and can be integrated with another type of wastewater treatment system to improve the quality of effluent [67,68], but in doing this, special considerations must be given to the pore size, as it is very critical factor in the separation process (selectivity), driving force requirements as well as the fouling susceptibility [69]. The commonly used membranes are classified based on their pore sizes and they include; Microfiltration (MF), Ultrafiltration (UF), Nano-filtration (NF), and reverse osmotic (RO) [48].

3.1. Membrane in MBR System

The applications of the membrane for filtration process in MBR entail the physical estrangement of the suspensions into concentrates and permeate at a certain range of pressure [61,70]. This implies that a membrane acts as a discerning barrier based on the molecular weight of the constituted substances. The solutes with a molecular weight smaller than the pore sizes will be able to navigate through the membrane as permeate, while the larger molecules are retained as the concentrates [71]. Usually, membranes are fabricated and devised into modules to make it applicable for wastewater treatment. The membrane fabrication and their implementation into modules have been comprehensively reported in a previous study [72]. More importantly, the reports anonymously confirmed that the membrane pore sizes and stability are determined by the constituting materials used for the fabrication, the concentration, and method of fabrication. However, the targeted applications, treatment conditions, and the characteristic of the feed are also important factors that determine the suitability of membrane pore size for a specific application [22]. In view of this note, it can be deduced that pores sizes play a major role in terms of applications, the kinetic ability of mass transportation through the membrane, and the trans-membrane pressure (TMP) requirement [73,74,75,76,77]. The TMP provides the required force to conduct the substances of smaller sizes through the membrane pores [61,70]. During filtration, the TMP together with the pore sizes and membrane construction materials directly influence the overall selectivity and productivity performance of the membrane during treatment [71]. Therefore, the range of pore sizes, molecular weight cut-off (MWCO), and required transmembrane pressure can be used to define the types of the membrane and their suitable applications [77]. Based on this, researchers have classified membranes into MF, UF, NF, and RO [39,78].

3.1.1. Types of Membrane

The contaminants with size range 0.08–2 μm and 100–500 kDa of MWCO can be separated using an MF membrane [79]. This shows that MF is most suitable for separating suspended particles, and it requires lower TMP (7–100 kPa) compared to a UF membrane. The UF membrane is mostly applied at a pressure 70–700 kPa to separate contaminants with size range 0.005–2 μm and 20–150 kDa of MWCO [80,81]. Studies have shown that a UF membrane is suitable for the separation of protein and carbohydrate [82,83]. It has also demonstrated good performance in the separation of virus and macro-nutrients from wastewater [84]. NF and RO membranes require much higher pressure to permeate through the membrane. The tighter pore sizes elevate the hydraulic resistance and adhesive forces; thus, higher pressure (850–7000 kPa) is required to overcome the drag forces [84]. The NF membrane is suitable for separation of contaminants smaller than 0.002 μm, while RO is usually used for desalination and/or removal of dissolved constituents [85]. RO has a very good rejection efficiency but requires exorbitant operational energy, and it takes much longer to finish filtration [86]. This drawback may be attributed to the narrow pore sizes of the membrane [87,88].
However, in most practical situations, MF and UF are applied for pretreatment prior to the application of NF and RO. This practice is an operational measure to reduce the contaminants load in the wastewater, thus alleviating fouling rate in the membrane. Figure 1 shows the summary of the distinctions between the types of membranes commonly used. From this figure, it is obvious that MF and UF have a larger pore sizes compared to the NF and RO. This suggest that lower TMP is required to drive permeate through the membranes (MF and UF). However, the NF and RO reject wider size range of contaminants, and also they are suitable for the removal of organic molecules and ions in wastewater filtration. More so, NF membranes incline to navigate monovalent ions such as the K+, Li+, Na+, but retained most of the divalent (Ca2+, Mg2+, Fe2+) and trivalent ions (Fe3+). Reports have also shown that the mono- and divalent ions rejection performance of NF is between 35–85% [89,90,91,92] and 65–90% [90], respectively. The RO membrane has higher monovalent ions rejection efficiency in the range between 90 to 99% [91]. Overall, the high selectivity performance of the NF and RO have been attributed to the smaller pore sizes of the membranes [93].
Nevertheless, it is important to strike a balance between the selectivity and productivity performance indices; this will prevent compromise of any of the indices. In view of this, it is undesirable that RO and NF require higher TMP, and they have also been attributed with lower productivity and more susceptible to fouling compare to MF and UF membranes [47]. All these defects have been featured to the narrow pore sizes in the RO and NF which often elevate hydraulic resistance of the membrane [93]. This suggests that it is logical to apply the RO/NF for filtration after using MF and/or UF to pretreat the feed. Ultimately, this procedure will significantly reduce the overall pore blockage and fouling challenges [47].
Irrespective of the type of membrane, the two methods often used to operate membrane in a typical MBR system include cross-flow and dead-end operations [40]. Cross-flow operation allows parallel passage of concentrated feed (POME) to the surface of the membrane (see Figure 2a). Interestingly, this mode of operation averts further deposition of foulants on the membrane surface, thus constant permeability or flux is maintained for a longer period during filtration operation. The cross flow may be configured inside-out or outside-in [94,95]. However, outside-in is preferred in the cylindrical configuration, more especially when dealing with feed that contain high amount of suspended solid. This is because outside-in cross filtration ensures minimal loss of the tangential velocity and accessibility to the fouled surface of the membrane [95]. Figure 2b depicts dead-end filtration. Unlike in the cross-flow operation, filtration takes place at right-angle to the surface of the membrane [96]. Report has shown that the tendency for membrane surface polarization to occur in dead-end filtration is high due to the persisting interaction with the feed [97]. The particulates and other foulants such colloids, EPS and SMP, easily tack onto the pore walls to initiate fouling. This defect in dead-end filtration justifies the usual batch process recommended, which ultimately obviates frequent fouling, diminishing critical flux and declination in selectivity [97]. This implies that, in a practical MBR application, the important membrane performance and functionality indices under variable processing conditions include the critical flux, TMP, and selectivity factors [40,61].

3.1.2. Membrane Performance Indices

Critical flux (Jc) is a performance indicator quantifying the productivity of the membrane under operation [98]. The flux rate is relatively steady up to the critical flux rate and it has been said that within this stage of filtration, the fouling rate is negligible [40,99]. However, further increase in flux above the critical range influence fouling, due to the rapid formation of cake layer that usually occur under this condition [100]. At a flux above critical rate, the mass transportation of foulants towards the membrane surface is high and this eventually cause membrane polarization [39,100]. Though there is not one standard procedure to determine the critical flux, this might be due to the complexity and persistence fluctuations in the results of the experiment, which definitely pose difficulty in data analysis and report. One of the most practical and widely used procedures is the flux step technique, and it involves determination of flux, permeability and fouling rate using Equations (1)–(3), respectively [40,99]. Essentially, at a varied flux, the initial TMPi after the first jump due to the increase in flux at time ti, and the final pressure (TMPf) after time tf of filtration are measured in this technique. According to Le-Clech et al. [100] and Mutamim et al. [40]; the interception of the flux ( P a v ), permeability (K) and fouling rate ( d P d t ) on the plot gives the critical flux.
F o u l i n g   r a t e ,   d P d t = T M P f n T M P i n t f n t i n
A v e r a g e   T M P ,   P a v = T M P f n + T M P i n 2
M e m b r a n e   P e r m e a b i l i t y ,   K = J P a v
Membrane selectivity ratio ( α i ) is another important performance indicator and it determines the capability of any given membrane to distinct wastewater into concentrates and permeates [101] Dutta et al. [102] reported that selectivity ratio of membrane has correlation with capacity to transmit variable constituent to different level. They further reported that pH of the feed, ionic strength, nature of the solute, membrane properties, permeate flux and hydrodynamic conditions are the major factors that determine membrane selectivity. Practically, α i depend majorly on the molecular weight and they varied proportionately to one another [81]. Also, Sert et al. [103] reported that wastewater with multivalent as well as monovalent co-ions are usually retained as a result of their surface electrical charges, particularly when RO membrane is used for the treatment. However, some of the monovalent could navigate through to balance up the disparity in the charges on the both surface of the membrane [103,104,105]. Vaughn et al. [106] used Equation (4) to measure α i , and they considered the specie concentration before membrane filtration as C i b and specie concentration contained in the permeate flux as C i p
α i = C i b C i p C i b
Besides the types of membrane, the system configurations along with the wastewater characteristics are other important factors that could influence the performance of MBR [106,107].

3.2. MBR Configurations

Irrespective of the process treatments, the MBR is usually utilized in two types of configurations; the submerged (sub-) and side-stream (ss-) configuration [31,76,108,109,110], as depicted in Figure 3a,b, respectively. Basically, submerge configuration consist of immersion of the membrane module(s) inside the reactor and this allows direct interaction with activated sludge, while the side-stream configuration has its membrane placed outside the reactor and usually, it requires adequate cross flow velocity because of the need for recycling of activated sludge [111,112].

3.2.1. Submerged Configuration (sub-MBR)

Submerged configuration involves dipping of membrane module(s) into the wastewater. This type of configuration allows direct dealings between the membrane and microbial activities. The interactions of the membrane and the microbes are greatly influenced by several factors such as the treatment processes, characteristics of the feed, and the membrane properties. However, during the filtration process, the microbial flocs and all other substances larger than the pore sizes are retained within the concentrates [63]. The sub-MBR is normally operated at lower flux and pressure with an adequate supply of coarse aeration to meet the oxygen requirements of the biomass and to mitigate the rate of cake layer formation on the membrane surface [74,113]. This condition of operation could considerably reduce the operational energy and the overall cost of MBR treatment application [114,115]. The membrane modules of this type of configuration are often taken out of the reactor when there is a need for chemical and/or mechanical cleaning [116]. This makes the submerge configuration less flexible in terms of fouling control, unlike the side-stream configuration that the membrane module is placed externally.

3.2.2. Side-stream Configuration (ss-MBR)

Side-stream configuration has long found industrial applications for wastewater treatment before the submerged configuration [110]. The external placement of the membrane module ensures easy operational assessment. However, it requires an additional pump for the recycling of the sludge and also to maintain the required cross-flow velocity (CFV) for a specified permeate flux [110,117]. The magnitude of CFV is a key factor that determines the rate of suspended solid deposition on the membrane. The tangential CFV must be high enough to minimize the rate of biomass settling in both membrane pores and surface [117]. In doing this, the resulting effect must be carefully considered. This is because, at a certain range of trans-membrane pressure (TMP) and permeates flux, the membrane became more prone to fouling [114]. Studies have shown that at elevated TMP and flux, the membrane could easily polarize and the tendency for bio-flocs disintegration is high. This habitually resulted in excessive operational energy requirements [118,119,120]. The findings of the previous studies on MBR configurations are summarized in Table 2.

3.2.3. Effect of MBR Configuration

The effects of configurations on MBR performance under different treatment processes have been reported by several researchers [63,111,116,121,122]. Le-Clech et al. [111] reported the hydraulic performance of submerged and side-stream configuration under selected operational parameters on a hybrid MBR. Essentially, they claimed that fouling susceptibility was less pronounced in the submerged configuration at a varied air flow velocity. As shown in Figure 4a, the steep increase in Flux was observed under gradual upsurge of TMP. However, a sudden increase in the TMP was noticed at a further increase in the critical flux. This indicates that the membrane has started fouling which may degenerate if the filtration continues under this condition [122]. Conversely, the operation of side-stream MBR at varied CFV confirmed that the membrane is more prone to fouling, as represented in Figure 4b. This could be due to the high operational TMP and CFV [123]. Based on this findings, It can be deduced that at equivalent conditions of operations (such as TMP, flux), the submerged configuration sustains less operating energy and is less prone to fouling compare to side-stream configuration [63,111].
In another study, Gander et al. [121] strengthen that the rate of flux declination is more in the side-stream configuration under the same operating pressure. Also, Andrade et al. [120] investigated the effect of membrane configuration and reported that side-stream submerged MBR indicates a significant slow growth of the biomass. This might be due to the shearing process required by the sludge recirculation pump [123]. However, both configurations give excellent COD and colour removal.
As summarized in Table 2, it can be inferred that sub-MBR requires lower TMP (0.1–26 bars) compared to ss-MBR (0.6–46 bars) for effective permeate filtration. Jefferson et al. [122] reported that a steady permeate flux (of 7.9 L/m2h) is obtained at lower TMP (0.1 bar) at a reasonable operational energy (4 kW/m3) in the sub-MBR. In another similar study, Radjenovic et al. [64] confirmed that operating submerged configured MBR at low TMP (0.2–0.5 bar) ensure constant flux (20–50 L/m2h) and lesser energy demand (0.3–0.6 kW/m3) compared to the ss-MBR configuration (4.12 kW/m3). The two studies validate that the observed steady permeate flux in the sub-MBR is an indication that the membrane is less prone to fouling at the operating condition. This report is in a similar trend with other studies [113,124], as summarized in Table 2. On the other hand, Wang et al. [26] reported a flux of 25 L/m2h at a relatively higher TMP of 30 bars with a cross flow velocity of 0.3 m/s. It was observed that the vulnerability to fouling becomes more frequent due to the elevated TMP and permeability flux.
Likewise, some of the previous studies on ss-MBR are contained in Table 2. From this Table, ss-MBR requires higher TMP but gives more permeability flux compare to the sub-MBR. For instance, from the report of Radjenovic et al. [64], about 50 to 100 L/m2h of permeability flux were obtained at TMP of 4 bars. This is also in accordance to the work of Yin et al [125]: they confirmed that as the TMP varied to 300 psi, the permeate flux increased from 39 to 59 L/m2h. However, the major defects attributed to these operating conditions were the excessive energy demand (12 kW/m3) for aeration and the lofty fouling susceptibility [121]. Equally, studies conducted by Le-Clech et al. [108], Morrow et al. [113], Xiao et al. [126], and Abdurahman et al. [127] reported comparable results, as presented in Table 2.
Based on these reports, SS-configuration requires higher TMP, though it may give more flux at a time. However, these conditions of operation expose the membrane to rapid fouling [64]. Also, the higher energy demand (0.045–140 kW/m3) in the ss-MBR configuration is another considerable factor which could make it undesirable for industrial applications. Furthermore, this indicates that sub-MBR configuration is more cost effective in application than ss-MBR configuration due to its sustainable running cost [113,123]. However, the initial cost of procuring submerged configuration may be higher, as a result of its larger surface area and volume [128]. Since, in the most application of membrane, low energy cost, low TMP and minimal fouling rate are encouraged [121]. Therefore, the high operational energy requirement of the side-stream must be vindicated based on the purpose of applications.

3.3. Treatments Process Using MBR

As reported above, the type of MBR configurations could greatly influence the treatment processes along with operational parameters. The effect of different processes conditions such as aerobic, anaerobic, anaerobic-aerobic, or combinations (hybrid), sonication and thermal application on the performance of MBR have been investigated and reported in the previous studies. In the MBR system, aeration may be required not just for biomass degradation but also for fouling control [62,103]. Particularly, at the initial stage of the treatment, deposition of colloids on membrane surface is less at higher aeration intensity but it may increase the fouling rate after a certain period of operation [129]. This pragmatic effect may be attributed to the high shearing force provided by the scouring air which eventually deflocculates the aggregates hence generating more particles (colloids) of smaller sizes. As the aggregates agglomerate to form bigger sizes, the intensity of scouring air becomes critical due to its effect its stability and degeneration. This could lead to the generation of more colloids or particles with smaller sizes, and this plays a key role in initiating pore blockage and gel layer formation [10,130]. This shows that optimal aeration is essential because of its strong correlation with fouling mechanism [131]. In another word, the performance of MBR under limited aeration paved another important investigation which has been considered extensively under anaerobic treatment conditions.
The anaerobic process involves degradation of biomass by microorganisms in oxygen-lacking conditions [32,132,133]. It requires less energy, suitable for treatment of high strength wastewater, efficient removal of organic pollutants, produces little amount of sludge and support resources recovery usually in the form of biogas and bio-fertilizer [134]. This implies an efficient microbial activity is the key indicator of the optimal anaerobic process in an ideal MBR [132,135]. It has been reported that the performance of bacteria in the anaerobic process is influenced by environmental conditions such as temperature, pH and nutrients availability [136]. Particularly, methanogen bacteria is very sensitive to pH and is best performed at a neutral value, while pH value less than 7 is more suitable for both acidogenosis and acetogenesis activities at a temperature ranged between 35–40 °C [137,138,139]. Anaerobic processes have been applied in MBR with different configurations, both the sidestream and submerged. Anaerobic MBR generally exhibits longer start-up and recovery time, low pathogen and nutrients removal [140,141]. This is because, under the anaerobic process, the bacteria require an extended time to acclimatize to the new environment (MBR) and also to get stabilized in terms of decomposition performance [30,142]. At the stabilized stage, adequate numbers of bacteria are available to perform the anaerobic decomposition more efficiently, especially if the environment is conducive for the bacteria [2,143,144,145]. Thus, the rate of membrane fouling is considerably relegated at this condition [143].
In an effort to improves the limitations of both aerobic and anaerobic processes, studies have shown that better performance could be achieved by hybridizing the two processes (anaerobic-aerobic) [13,28,37,146]. The combination of an anaerobic-aerobic treatment approach integrates the benefits of both processes and thus ensures adequate bio-decomposition, a higher rate of degradations, more biogas/bio-fertilizer recovery, efficient nitrification and denitrification processes [37,147,148]. Essentially, at acclimatized state, the anaerobic bacteria significantly reduced the pollutants such as COD, BOD and TSS [149], while the aerobic treatment is much suitable for the removal of toxic gases such as ammonia [8,150,151]. In addition, at the aerobic treatment, further deterioration of the remnants COD and BOD is possible to obtain better permeates [38]. Based on these reports, it can be deduced that hybrid treatment processes improve inclusive performance consistency.
Table 3 shows the summary of the recent different processes applied using MBR for POME treatment. Based on Table 3, it is obvious that selections of treatment processes have significant effects on the performance indices, particularly on the rate of membrane fouling and pollutants removal [20,55,152,153,154].

3.3.1 Aerobic Processes in MBR (AerMBR)

Aerobic process in MBR involves the combined application of microorganisms in an oxygen environment and membrane technology for POME treatment [38,155]. Aeration is a required factor for aerobic bacteria to facilitate stabilization of organic pollutants such as COD and BOD, which eventually decomposed through the metabolic activity to release CO2 [156,157]. In MBR, the injected air is not just to meet up with dissolved oxygen (DO) demand by the bacteria but also to provide shearing forces required to scour the accumulated foulants on the membrane [158]. Technically, the air supplied must be adequate to sustain both bacteria oxygen demand and requisites shearing forces [110,159,160]. It is noteworthy that bacteria are prolific in nature; this indicates that in no-time the available oxygen could be depleted. Thus, this obliges the need for supplementary aeration to meet up with the ever-increasing oxygen demand and imbalance in the F/M ratio [161,162]. In a balanced aerobic MBR system, a certain amount of activated sludge is predetermined to maintained the food –microorganism (F/M) ratio to ensure equilibrium in aeration depletion-replenishment, by so, the problem of bulking or sudden increase in aeration is curtailed and this allows optimal biodegradation [78,163]. Normally, the rates of oxygen consumption by the active biomass as well as the required scouring air are one of the critical factors considered in the design of MBR [63].
In contrast, studies have shown that excess scouring air encourages the fouling mechanism in the membrane [124,164,165,166]. This is because as the activated sludge prolong, smaller organic aggregates get agglomerated to form bigger bio-flocs which may easily crumble by the scouring air. This resulted in the generation of more bio foulants, colloids and particulates with a range of smaller sizes [114,164,165,167]. Eventually, the disintegrated flocs settle on the membrane to initiate pores blockage as well as the formation of the dense cake layer, thus increase in hydraulic resistance [168,169]. This implies that providing optimal aeration is indispensable to derive the best performance and application of MBR at industrial scale.
Researches have shown that mixed Liquor suspended solid (MLSS) of POME have a strong influence on the characteristics of activated sludge and membrane fouling rate. At an increasing MLSS concentration, the oxygen transfer efficiency (OTE) through the mixed liquor decreases by 400% [10,170,171]. This is because of the high F/M ratio which invariably encourages sludge bulking and an increase in the viscosity [65,172]. The increase in viscosity is not suitable for hydraulic conductivity within mixed liquor because of the excessive drag resistance which may amount to a noticeable decrease in mass and oxygen transfer [98,169]. Therefore, more turbulent aeration with stronger forces may be required to overcome the hydraulic resistance and to ensure proper mixing of the activated sludge [172]. Conversely, reports have shown that depleted oxygen after biological degradation in activated sludge improves the bio-flocculation and sedimentation process of the flocs [173,174]. This is because of the declination in oxygen gradient across the biofilm, and this normally leads to the formation of a layer devoid of oxygen in the interior. Thus, stability in flocs for effective biodegradation is promoted [39].
In an attempt to improve OTE and general performance of MBR, different techniques such as mechanical agitator and adsorbents were combined with the AerMBR and evaluated. It has been reported that mechanical agitator improves OTE by promoting the contact between the bacteria and substrate, and this immediate effective biodegradation process [158,175]. Also, Deoawan et al. [109] reported that agitation advances mass transfer within the mixed liquor and thereby stabilizing the environmental conditions for the bacterial activities. In addition, pulverization of the activated sludge ensures even distribution of DO, nutrients, as well as the microorganism [176]. This shows that agitations contribute immensely to the excellent removal of pollutants (99.9%), though most often further operational energy is required and so the additional cost of operation is incurred [105,128,129].
Also, combinations of the adsorbent with MBR operated under aeration have also received substantial attention recently. Mostly, a good adsorbent is characterized by small volume, large surface area (500–3000 m2) and dense micro-pores [177,178,179]. The dense micro-porosity of the surface area is the key factor required to perform the removal of pollutants through adsorption process [179]. As a result of these special features, the applications of adsorbents for POME decolourization and polishing have received a lot of attention. On a general overview, adsorbents have recorded a credible performance in decolourization of POME with removal efficiency ranging from 80 to 99.9% and also improves the membrane fouling condition under optimal process conditions [180,181]. Yuniarto et al. [18] operated a submerged AerMBR under aeration condition comparing the treatment with and without adsorbents. They reported that at an optimal dosage (4 g/L) of activated carbon (AC) a significant improvement both in flux (42 LMH, LMH = Litre per meter per hour) and COD removal efficiency (98.5%) was observed, while the operation without AC presents lesser performance. Also, a similar report by Guo et al. [180], they applied adsorbent in a submerged MBR operated under aeration condition. The results confirm the improvement in fouling reduction and removal performance of the organic pollutants, as reported previously. This could be due to considerable deterioration of pollutants by the adsorbent (AC) prior to the membrane filtration treatment. Hence, the adsorptive process as pretreatment reduces the number of contaminants present in mixed liquor and so relegate pores blockage initiation and the propensity of biofilm formation [18,47,180,181]. However, AC requires chemical treatment, such as the use of buffer solution to vary the pH level [182] for good adsorption during treatment, and this practices often lead to the formation of acidic or basic permeate [183] which is hazardous to the environment if discharged without further treatments.
In another report, the application of nanofiltration (NF) membrane under suitable aeration for POME decolourization has been proved promising [153]. However, the most serious drawback of this treatment process is the outrageous TMP required to conduct the filtrate through the pores [184]. As substantiated earlier, membrane filtration at high TMP exposed the membrane to more expeditious fouling [185], due to the prohibitive foulants polarization towards the membrane surface [4]. More recently, nanoparticles have been applied to the membrane to improve its antifouling properties and net energy consumption [152,153]. Ultimately, nanoparticles such as zinc iron oxide (ZnFeO4) and Titanium dioxide (TiO2) checks the fouling process by releasing hydroxyl and antimicrobial radicals on the interface between the activated sludge and membrane. The hydroxyl radicals adhered to the surface to improve the hydrophilic properties of the membrane. This, in turn, diminishes the hydraulic resistance and fouling susceptibility [152,153]. In addition, the antimicrobial radicals inhibit the organic foulants metabolic activities, thus, preventing initiation of fouling.
Tan et al. [152] prepared a composite UF membrane (PVDF) with ZnFeO4 nanoparticle using blending additive technique. The study aimed at improving decolourization performance and fouling properties of the membrane under the aeration condition. They reported 70% colour removal and 40.14 LM under aeration condition. In addition, 0.5 wt % of nanoparticles (ZnFeO4) loading significantly improves colour removal and permeability. At this loading (0.5 wt %), the membrane surface negativity is optimal and readily releases zinc ion which actively hampers the metabolic activities of the bio-foulants as well as colour removal. However, the major weakness at 0.5 wt % loading was the instability of the nanoparticle after a number of filtration cycles. This leads to the collapsing of the nanoparticles and eventually exposes the membrane to rapid fouling. From a similar study reported by Subramaniam et al. [21], a titanium oxide (TiO2) was used as nanoparticle to produce composite UF membrane-titanium nano-tube (TNT). The target of the study was to decolourize POME under photo-catalytic conditions. The best colour removal of 67.3% efficiency was obtained at 0.5 wt % TNT loading. Principally, TNT nano-tube works similarly as ZnFeO4, the released ions not only act like denaturant with strong efficacy against bio-foulants but also create hydration interface between the membrane and activated sludge. This prevents foulants accumulation and also improves the hydrophilic property of the membrane [186,187]. In contrast, TNT is expensive and it requires the presence of ultraviolet radiation for good performance [188].
An attempt on the use of microbial fuel cell (MFC), and AerMBR to remove colour from POME under alternate aeration and anoxic conditions have been reported [189]. In this study, they concluded that decolourization was only significant at the anoxic condition and that MFC only favours the formation of biofilm which assists organic pollutants deterioration. This suggests that MFC used in this experiment can only decolourize POME under anoxic conditions and also the rapid deterioration of the pollutants was as a result of the dense microbes population contained in the activated sludge [190].
Collectively, the summary of the AerMBR treatments processes is presented in Table 3. AerMBR could provide high-quality POME treatment with 98.9% COD removal and 95.2% colour deterioration at a shorter hydraulic retention time [2,155]. Based on these indices, the combined use of adsorbent (AC) and AerMBR is more applicable. However, considering other factors such as fouling control and flux recovery, the combination of composite nanoparticle and AerMBR is most appropriate at 0.5 wt % loading of nanoparticle [21,154]. More so, the practical approach to reducing operational cost and downtime using AerMBR for POME treatment can be based on efficient fouling management. This could be archived by integration of adsorbents as pretreatment and composite nanoparticle-membrane into the MBR system. Primarily, the adsorbent component reduces the concentration of the contaminants in POME at a controlled pH level prior to the filtration process. The membrane filtration further purifies the effluent by removing all the remnants of the contaminants. This combination could provide several advantages such as high efficiency of organic removal, colour treatments, steady permeability flux and less fouling susceptibility [10,56,183,187].

3.3.2. Anaerobic Processes in MBR (AnMBR)

Decomposition of organic substances in the absence of oxygen involves a series of processes such as hydrolysis, acidogenesis, acetogenesis and methanogenesis processes [190]. In the anaerobic process, the carbonated POME are degraded to soluble derivatives by hydrolysis process, then acidogenic and methanogens bacteria further decomposed it into a simpler and soluble compound in the form of acetic acids, ammonia (NH3), hydrogen gas, carbon dioxide (CO2), methane (CH4) and some quantity of hydrogen sulphide (H2S) [190]. These processes are significantly influenced by the environmental factors, particularly, pH level and temperature [191,192]. It has been known that at neutral pH value the methanogens decomposition activity improves [65], which is normally indicated by the significant increase in methane generation at a suitable temperature between 35–40 °C [193,194], while pH values less than 7 are more suitable for the hydrolysis, acidogenesis and acetogenesis [195,196]. This implies that optimal control of the environmental conditions improves anaerobic activities.
Recently, Abdurahman et al. [196] and Abdurahman and Azhari [197] both gave similar reports that higher biomass concentration (MLSS and MLVSS) yield more methane (CH4) ranging between 0.58–0.87 L/g COD-d with COD removal efficiency between 94.8–98.4% using AnMBR system. The improvement in CH4 production and efficient removal of COD were attributed to the adequate balancing of F/M ratio at the ranged biomass (MLSS and MLVSS) concentration and this condition accelerates the anaerobic biodegradation of the organic pollutants [6,139,153]. Therefore, biomass (MLSS) concentrations have strong influences on kinetic properties of the activated sludge and also determine the magnitude of its growth yield coefficient (GYC) [191]. Mostly, the rapid increase in GYC is as a result of the availability of favourable conditions such as pH, temperature and nutrients (as mentioned earlier) [165]. These conditions ensure optimal biodegradation and prolific multiplication of the bacteria required for the anaerobic treatment. In contrast, the tendency of membrane fouling at concentrated MLSS is high during anaerobic treatment. This is because, at a limited scouring air, the SMP easily adhered onto the pore walls to initiate blockage which may eventually develop into biofilm [198,199,200]. However, the intermittent filtration-relaxation could minimize the SMP deposition in the pores [171].
According to Annop et al. [201], prolonged filtration-relaxation under anaerobic condition expedites the formation of biofilm on the membrane. This is because of the massive transportations of foulants towards the membrane surface, and this often resulted in high polarization [174,202,203]. This is in line with the report of the study conducted at high flux (35 Lm−2h−1) with a specified aeration demand (0.25 Nm3/m2h) by [117]. According to this study, in a longer-term operation, a significant increase in fouling rate was observed due to the polarization effect [117]. In a different study, solid retention time (SRT) pose a strong influence on the rate of biofilm formation as a result of the protracted polarization of biomass and persisting interaction on the membrane surface [204]. At a longer SRT, the soluble microbial product (SMP) concentration is high which easily got accumulated onto membrane pore walls to initiate blockage [98,203,204,205]. In addition, longer retention of activated sludge encourages shifting of microorganism growing rate to endogenous phase and this may lower the F/M ratio. This resulted in a higher oxygen demand [63] and the rapid production of bio-fouling material commonly known as extracellular polymeric substances (EPS) [98,130,141].
Faisal et al. [61] attempted reducing the rate of bio-fouling by incorporating a dissolved air flotation (DAF) system with AnMBR. They observed a reduction in the membrane fouling rate and the overall organic pollutants removal efficiency improves significantly in the range between 94–99.9%. They attributed the improvement in the performance to the DAF applied for pretreatment prior to the membrane filtration. In addition, considerable deterioration of suspended solids (87.5% removal) was noticed after the pretreatment. This inferred that fewer foulants are contained within the mixed liquor after the DAF pre-treatment, hence enhancement in the filtration rate and minimal vulnerability to fouling were achieved [15,61].
In recap, AnMBR processes for POME treatment were summarized in Table 3. From Table 3, the results pointed out that the AnMBR treatment system is suitable for high strength POME with MLSS concentration ranging from 11,760 to 20,800 mg/L. Also, it remains the most appropriate treatment method for COD removal (with an efficiency ranging from 98 to 99.9%) and resources recovery/utilization, such as biogas and bio-fertilizer [39]. However, the membrane is more susceptible to fouling under an anaerobic condition at high MLSS concentration. This is because as the activated sludge aged (SRT), the viscosity increases and the foulants such as particulates, EPS and SMP easily get polarized on the membrane to form cake layer [206,207]. However, some of the noticeable approach used to mitigate fouling in AnMBR system includes BAF [62] and intermittent frequent filtration-relaxation [201] which have presented some improvement in terms of steady filtration and pollutants removals, particularly the COD and TSS.

3.3.3. Hybrid (Integrated) Processes in MBR (HybMBR)

Hybrid processes encompass a systematic combination of the anaerobic, anoxic and aerobic process to improve treatment efficiencies [46,54,160]. Hybrid MBR (HybMBR) provide a platform for anaerobically treated wastewater to undergo aerobic treatment, and this gives the possibility for chains processes such as biodegradation, de-chlorination, nitrification, denitrification as well as filtration process which usually takes place at the membrane section [38].
The structural configurations of HybMBR have to be compatible to ensure sequential and efficient accomplishment of the several treatment processes, as mentioned above. Also, a report has shown that the integrated processes in single footprint guaranteed HybMBR suitability for treating high strength POME with a reduced modus, and this could lower the cost and treatment time [55,56,139,207,208,209,210]. Efficient performance of HybMBR could be influenced by the physicochemical properties of POME, sequential arrangement of the processes (anaerobic-anoxic-aerobic) and operational parameters such as MLSS concentration, activated carbon, temperature, solid retention time (SRT) and hydraulic retention time (HRT) [38]. The application of HybMBR for POME treatments under variable operating factors is summarized in Table 3.
MLSS concentration is considered one of the most important factors that influence the MBR performance, particularly, the membrane fouling susceptibility. Damayanti et al. [207] investigated the effect of membrane fouling in HybMBR under varied MLSS concentration from 5 to 20 g/L [130]. They observed that the relationship between the TMP and flux change is steeper at higher MLSS concentration compare to lower concentration, as depicted in Figure 5. From Figure 5, at 5 g/L MLSS concentration; rise in TMP under the progressive increase in flux is insignificant until at a critical point where a sharp increase in TMP was observed. Similarly, Ahmad et al. [208] reported the effect of varied MLSS concentration (4–8 g/L) on the performance of hybrid MBR. At the initial stage of the study, the fouling rate was constant with 0.08 gradients, but a sudden decrease in filtration volume was observed at further increase in flux beyond the critical value (16 LMH). Collectively, it can be construed that operating MBR at higher flux, influence more foulants (SMP, EPS and colloids) generation and transportation towards the membrane surface [209]. Consequently, the generated foulants could easily commence fouling mechanism through pore blockage and gel formation [48]. Damayanti et al. [56] investigated the effect of several bio-absorbents (activated carbon, zeolite and moringa oelfera) under varied MLSS concentration. They deduced that increasing biomass (MLSS) to higher concentration could cause a reduction in permeability by 400% [207]. Furthermore, they reported that, at the optimal dosage, the powdered activated carbon (PAC) present the best treatment performance, with 85% contaminants removal efficiency during the pretreatment [57]. The improvement in fouling problem is as a result of the adsorbents pretreatment which substantially reduced the foulants contents by adsorption process [210]. A modified HybMBR with a sedimentation compartment has been reported to perform remarkably in term of fouling control as well as removal of contaminants such as total phosphorus and nitrogen with average 94, 87, and 85% efficiency, respectively [211]. They further reported that as the temperature varied from 27 °C to 13 °C, the swift increase in TMP was observed when the flux is maintained at a constant rate. This could be due to the temperature effect on the viscosity of the activated sludge. As the temperature decreases, the shearing force within the sludge increases due to the higher hydraulic resistance and viscosity [212]. Similarly, the mixed liquor became denser and sticky at a lower temperature, and thus has an antagonistic effect on the entire oxygen transfer efficiency [213]. Whereas, at a higher temperature the foulants gain more kinetic energy with relatively low hydraulic resistance due to the considerable decrease in viscosity. More so, the accumulated foulants can easily get sheared or dislodged by lighter scoring air at the higher temperature (27 °C). This implies that operating the HybMBR in such condition could favour the overall performance in terms of TMP, critical flux, and fouling propensity [110,213].
The recap of HybMBR performance under different operational conditions (such as MLSS, organic loading rate {OLR}, AC and temperature) were presented in Table 3. It is obvious that increasing MLSS and OLR concentration could decrease the flux filtration rate drastically [207]. This effect has been attributed to the high density of foulants generated from concentrated biomass which easily initiate pores blockage in the membrane. Based on this, it can be deduced that biomass concentrations (MLSS) have a strong correlation with the number of foulants generated during biodegradation [214]. The biodegraded biomass contains extracellular polymeric substance (EPS) and carbohydrate, which are the major biofilm and organic foulants [58,65].
In terms of performance indices, HybMBR is suitable for POME with high concentration MLSS (5000 to 20,000 mg/L) [207] and capable of deteriorating COD by 94 to 97% efficiency. Also, it has a high possibility of flux recovery of the fouled membrane with 91 to 95.3% [15,47]. Therefore, the combined adsorbent and the HybMBR system could easily achieve the aforementioned performance indices. Studies have shown that temperature has a strong influence on kinetic properties and biological activities of the biomass in HybMBR [130,215]. At lower temperature, the rate of particles and colloids gyration is low, thereby higher shearing force is required to scour the accumulated foulants on the membrane [70,216] and vice versa. Consequently, this could drastically increase the operating energy, and running cost [121,217].

3.3.4. Sonication Processes in MBR (SonMBR)

This is one of the most recent concepts employed in the MBR treatment process for POME. Mainly, ultrasonic MBR (SonMBR) involves the use of sonication to facilitate the decomposition of organic matter in POME [218]. The intensity and duration of sonication have a strong effect on the rate of contaminants disintegration [219,220]. At higher biomass concentration, higher intensity and longer sonication duration are required to provide sufficient particles collision, cavitation process, as well as effective mass transfer within the mixed liquor [194,195].
Several types of research have proved this technology efficient, particularly, in the aspect of bio-resources recovery from POME under anaerobic conditions. Shafie et al. [218] investigated the effect of ultrasonic on the performance of an anaerobic MBR at varied sonication duration. They reported a 105% increase in the CH4 yield under sonication. In addition, about 98.75% deterioration in the organic pollutants was also observed. They attributed the significant improvement in the performance to the sonication applied during the treatment. This denotes that the prolong sonication not only improves CH4 yield but also advance the kinetic properties of the microbial activities. Thus, faster degradation of organic matter in the activated sludge could be enhanced through sonication, and this presents another simple and clean method of preventing membrane fouling [219].
Furthermore, a combination of microwave and sonication present a novel means of disintegrating organic matter in activated sludge. Lately, studies have shown that increasing the ratio of soluble chemical oxygen demand (SCOD) to total chemical oxygen demand (TCOD) (SCOD/TCOD) and BI is best achieved through a combination of microwave and sonication [219,220]. From their results, a higher yield of CH4 (44 mL) at a short period (10 min) of sonication was obtained. They further confirm that the net operational energy was relatively low, though it has a strong correlation with MLSS concentration, sonication intensity/frequency, sonication duration, the volume of the POME treated at a time, as well as the targeted treatment qualities [98,221]. The study reported by Abdurrahman and Azhari [197] validated the above statement. According to the report, POME with a high concentration of MLSS (13,800–22,600 mg/L) requires prolonged sonication at a relatively high intensity. Under sonication condition, shorter SRT and longer HRT (8.6 days) are the most favourable operating parameters for the maximum yield of CH4. This shows that SRT has no significant effect on the degradation of biomass under ideal sonication conditions after stabilization [222]. They concluded that POME with higher biomass concentration (MLSS) produces more biogas with an average increment of 44% compared to the lower concentration [197].
Equally, Abdurahman et al. [222] reported that a combination of ultrasonic and sonic-thermal in MBR system accelerates the degradation process. Integrated ultrasonic and sonic-thermal was operated at a 37 Hz frequency with a varied temperature of 45, 55, 65, and 75 °C. They observed an increase in biomass disintegration and weight reduction with an average of 39.05%, and the best performance was obtained at the highest temperature of 75 °C. Similarly, Leano et al. [223] reported that ultrasonic pretreatment accelerates bio-hydrogen production from POME. The sonication duration and ultrasonic dose (0.91, 143, and 195 J/mL) were varied to different levels. The best performance was obtained at the highest ultrasound dose (195 J/mL) with 0.7 mmol H2/g COD of bio-H2 yield and 65% COD removal. Taha and Ibrahim [224] investigated the effect of sonication intensity at a varied pH (2–4) and time (3–10 min). They concluded that longer sonication time presents better COD removal in a short period of time. Manickam et al. [225] examined the effect of combined ultrasonic bath and hexagonal ultrasonic reactor. The experiment was conducted at a varied frequency of 28, 40, and 70 kHz, and they concluded that the combined effect presents consistent COD degradation. They further added that applying only ultrasonic at 37 kHz frequency give final COD fluctuating between 45,000 and 60,000 mg/L. Based on these reports, it will not be wrong to conclude that under optimal sonication intensity and duration, a higher yield of biogas, and better organic degradation (COD removal) is realizable [226,227,228].

3.3.5. A Thermophilic and Mesophilic Condition in MBR (TheMBR)

Temperature is one of the notable factors that influence the biodegradation process in MBR [157]. The level of its effect may depend on the intensity and duration of exposure [65]. The temperature process in MBR for POME treatment is mostly considered under two conditions: thermophilic and mesophilic [195,229]. Actually, the temperature has a direct effect on the viscosity of mixed liquor which could significantly influence the overall mass transportation [213,217]. This implies that high permeate volume is more visible under thermophilic condition because of the lesser hydraulic resistance [195]. In addition, thermophilic is more suitable for a higher range of organic loading rate [230]. Mass transfer efficiency is higher at this condition and this feature can be attributed to the high kinetic energy attained by the activated sludge. Thus, better contact between the bacteria and substrate is promoted [229]. However, the thermophilic condition could cause the formation of fatty acid which may considerably reduce the pH value of the activated sludge [146]. The acidic pH range could hamper methanogenesis process, in consequence, protracting the biodegradation [231]. Therefore, the rate of biogas (CH4) production may significantly reduce because of the inactive methanogen bacteria at such low pH [44]. In addition, at a prolonging thermophilic condition, the flocs can easily disintegrate to form small particles sizes with varied ranges [232]. Subsequently, the generated particles easily initiate pore blockage, then afterwards develops into a dense cake layer on the membrane surface [232]. In this manner, it can be said that the mesophilic condition is more favourable to the physiological process of the microbes [208]. This is because at a higher temperature (> 55 °C) the metabolism of the microorganism is negatively denatured, thereby, retarding the rate of microbial activities [191,232]. Based on this note, efforts have been vetted by researchers to identify the optimal effects of temperature on MBR.
Ma et al. [212] investigate the effect of temperature on fouling rate with the presence of microbial in MBR. They reported that temperature has an effect on the accumulation rate of polymeric and carbohydrate substance, also on the microbial prolificacy. As the temperature increases from 8.7 °C to 19.7 °C, the biomass concentration (28.1 mg/g-MLSS) significantly reduced to 2.2 mg/g-MLSS. This suggests that the thermophilic bacteria are more active as temperature increases (19.7 °C); for this reason, they multiply rapidly to dominate and decompose the organic matter [209]. However, the high rate of decomposition under thermophilic conditions may cause generation of more particulates with an extensive range of smaller sizes. This could contribute immensely to the fouling initiation as well as in the formation of the biofilm matrix [233,234,235]. In another study reported by Tee et al. [214], the optimal removal efficiency of COD was obtained at mesophilic condition (35 °C) but the bacteria growth rate was observed slower. They concluded that the growth rate of microbial and decay are directly proportional to one another but lost their proportionality at some critical temperature limits (>55 °C) [236]. Choorit et al. [236] reported that temperature could increase the rate of biogas generation by 33.33% but often expose the membrane to higher fouling propensity [237]. This finding is actually in line with the report of Abeynayaka and Visvanathan [228]. Under thermophilic conditions, lower sludge accumulation and higher COD removal efficiency were archived but frequent fouling was observed. The author attributed this effect to the excessively generated foulants under the operating condition.
In overall, the combination of mesophilic conditions (32–45 °C) and microbial fuel cells (MFC) could improve the permeability flux and reduce the fouling rate [146]. This is because of the strong correlation that exists between several factors such as operating temperature, granulation stability, the sludge viscosity as well as biodegradation process [48,157,238,239]. Particularly, the bio-granular could lose its stability at elevated temperature. This is because, at a higher temperature, the death rate of the biomass increases, and the large bio-flocs disintegrate to generate more colloids and smaller particulates [36,240,241] This stress the importance of applying suitable temperature during MBR treatment.

4. MBR Major Drawbacks

The most challenging limitation of the MBR treatment method is the membrane fouling which often resulted from the accumulation of foulant(s) such as bio-flocs, colloids, and particulates [65,242,253,254]. Membrane fouling could cause gross deterioration of the system performance in terms of filtration and operational energy utilization, as highlighted earlier. Therefore, the need to control the fouling is indispensable and this justified the previous studies reported on this critical area [67,110,161,255,256,257,258]. The efforts devoted by the researchers have yielded some improvements such as membrane fouling control strategies and filtration. However, the level of achievement is still under development and economically unsustainable for industrial application [243,244,245], particularly in the oil palm processing.
Regardless of the type of operation, mass transportation and frequent interaction of contaminants with the membrane modules is unavoidable. Primarily, MBR is operated either at stable TMP while flux is varied or vice versa [100]. This implies that the TMP and flux have strong correlations with the fouling mechanism. From Figure 6, at the initial stage of the filtration process, as the flux increases at a steady d(TMP)/dt rate, the fouling rate (blue curve) is almost constant [206]. However, the maximum permeability was attained at a critical flux of 12 LMH, in which further increase yielded a steep declination of permeability and a sharp increase in membrane fouling (red line). The sharp increase in fouling indicates a sudden jump in TMP [208]. This effect is initiated by the intense flux and by the biomass concentration, wastewater chemistry (such as divalent cations concentration, ionic strength, and pH), membrane surface morphology, membrane surface charge, membrane molecular weight-cut-off, and hydraulic and operating conditions [167,233,245].
The usually observed jump in TMP and a sharp drop in permeability-flux are of utmost concern. This condition is not suitable for profit-oriented industries due to the higher running cost required to maintain the dropped permeability-flux [15]. Therefore, it is essential to understand the nature and the different types of the foulants generated as well as the fouling mechanism during POME treatment using MBR. This could deeply succor both the industries and researchers to fathom the nature of the foulants and then espouse the most efficient and cost-effective measures to mitigate the problems of fouling. Considering the scope of this review paper, only the types of foulants as relates to POME and the detailed stages of the fouling mechanism were considered. This is because comprehensive reviews on the methods of fouling control such as the physical and chemical, as well as the novel application of composite nanoparticles have been reported previously [28,130,180,245].

4.1. Types of Foulants as Related to POME Treatment Using MBR

Foulants are the organic and inorganic substances that initiate pore blockage as well as layer formation on the membrane [247]. Foulants have been categorized into four groups: organic, microbial, inorganic and particulates foulants [248]. The classification was actually based on the biological, chemical and physical characteristics of the foulants. In addition, previous studies have shown that microbial/organic products, suspended-solute substances and debris concentrations have a strong correlation with fouling rate and also they determined the fractional contributions of each of the foulants [249,250].

4.1.1. Organic Foulants Generated during POME Treatment

Organic foulant is a biopolymer often describes as a complex microbial product (MP) [251,252] and it is generated through hydrolysis of extracellular polymeric substance (EPS) during the metabolic activity of the decomposer (bacteria). In addition, reports have also shown that the decaying of organic materials in the mixed liquor is another source of this foulants [253,254]. EPS is the major contributor to membrane fouling during POME treatment with an average composition of 52% of the total foulants [50]. Also, this is in accordance to the reports on the direct relationship that exists between the EPS contained in the mixed liquor with the fouling rate [255], specific flux [256], critical flux [128], and membrane filtration performance [257]. They all confirmed that EPS constitutes majorly in the deposited foulants. In addition to this, EPS plays a significant role in the bio-flocs aggregates formation, zeta potential properties of the mixed liquor and adsorption characteristics of the bio-flocs. This is because EPS encompasses a range of organic macromolecules which may include polysaccharides, proteins, lipids, humic acid, nucleic acids, and fulnic substances [258]. Though polysaccharides, proteins and humic materials are the key components of the EPS foulants, and they exhibit a different degree of hydro-affinity which determines the rate of fouling susceptibility. The polysaccharide is carbohydrate and is hydrophilic in nature, while the protein is less attracted to water due to its hydrophobic characteristic [254].
Furthermore, due to the several interactions with other types of foulants in the mixed liquor, EPS is normally classified into soluble-EPS (s-EPS) and bound-EPS (b-EPS) [254]. More interestingly, the b-EPS promote biomass aggregation and stability through agglomeration of bio-flocs by releasing a gel-like substance to resin them together. The gradual adhesion of the bio-floc resulted into formation of gel and this could later metamorphosis into cake layer [65,70]. More so, b-EPS exhibit a significant role in the biological stabilization and as well as the microbial aggregates in the MBR [254]. This shows that b-EPS not only promotes microbial aggregate but also influence the surface charge, mixed liquor viscosity and system sludge flocculation ability. Therefore, it can be said that b-EPS determine the formation or building of bio-flocculation and also define its stability [180]. One of the uniqueness of the b-EPS substance is the aptitude to demonstrate both hydrophilic and hydrophobic characteristics. This is possible because of the presence of both functional group charge which readily initiate deposition and adsorption on to unlike charged surface. Thus, this indicates that irrespective of the type of membrane used, b-EPS can easily get adsorbed on the surface to form a gel-like substance [259].
The soluble EPS (s-EPS) is also referred to as soluble microbial product (SMP) and they are mainly generated from organic based materials in activated sludge such as bacteria by-product and biomass decomposition [260]. This shows that they are biodegradable and a species form through the dissolution process of the b-EPS [51]. However, SMP is often considered into two major groups; biomass associated product (BAP) and utilization associated product (UAP) which are respectively generated from biomass deterioration and bacteria metabolism [254,260]. The macromolecular composition of the two SMP contains majorly protein, polysaccharide and humic acid just as the EPS [246] and this implies that they also contribute considerably to membrane fouling.
Nonetheless, the rate of fouling is not just influenced by EPS concentration but is as well depends on a number of factors which include membrane properties and operating conditions [246]. As already discussed above, operating conditions such as aeration, sonication, temperature, TMP and MLSS concentration, has a strong correlation with the bio-flocs formation, aggregate stability and fouling rate [16,207]. Similarly, a hydrophobic membrane with a rough surface is more susceptible to fouling due to the several cleavages that favour foulants deposition and the subsequent development of the biofilm matrix [50].

4.1.2. Microbial Foulants as Related to POME Treatment

Microbial foulant is commonly regarded as bio-foulant and is considered next to the organic foulant that usually causes serious fouling problem [186,187,188]. Fundamentally, bio-fouling involves sticking of bacteria micro-colonies by adhesion on the membrane surface to perform a complete living activity such as growing, reproduction as well as metabolic activity. After a certain period of time, bacteria micro-colonies multiply and develop into mature biofilms layers. The by-product released through the metabolic activity of the living layer (biofilm) generates extracellular polymeric substance (EPS) [261]. The secreted EPS is a bio-adhesive material, which normally serves as the glue that fosters the formation of biofilm matrix on the membrane surface [254,262]. This indicates that the major constituents of biofilm are the bacteria cells and the EPS. Ma et al [215] reported EPS is the major components of biofilm and it accounts for more than 70% of the complex organic matter and microbial aggregates. Besides, a report has also shown that EPS contain polar charge group which includes both the aromatic and aliphatic [254]. On the whole, EPS contains both the hydrophilic and hydrophobic functional group, as confirmed by previous studies [65,256]. Hence, the complex nature of EPS could be harnessed to improve the total biodegradation of POME through the microbial aggregate formation and sludge stabilization [263,264,265,266,267,268,269,270,271,272].
On the other hand, the presence of Lipopolysaccharides (LPS) on the bacteria cell wall plays a key role in the bio-fouling mechanism. Herzberg et al. [273] investigated the bio-fouling mechanism of P. aeruginosa under varied microbial fouling conditions. They confirmed that the rate of micro-colonies accumulation and bio-film development has a strong correlation with the secreted tacky substances (Lipopolysaccharides and EPS). They further added that the sticky substances promote auto-aggregation, and this ultimately immediate rapid growth and surface translocation of biofilm via twitching motility. Similarly, Ashhab et al. [274] reported that the secreted microbial substances not only influence film formation but also increase the viscosity of the mixed liquor. Consequently, this effect resulted in a significant increase in the overall hydraulic drag force [275,276]. However, measures and techniques to avert biofouling in MBR have been investigated under different conditions by several researchers. Kim et al. [277] uses a hybrid membrane modified using TiO2. They reported that the hybrid polymeric membrane inhibits the biofouling by generating antimicrobial radicals under ultraviolent radiation. In addition, the reports of Subramaniam et al. [21] and Tan et al. [152] concur with this finding. Collectively, they confirmed that the photo-biocidal effect of the incorporated nanoparticles prevents bacterial aggregation along with biofilm formation. Moreover, studies have also shown that the counteraction of biofouling using nanoparticles also improves the hydrophilicity of the membrane [11,278]. The mutual activities are accomplished during the self-assembling of the incorporated nanoparticles, and this process resulted in the generation of hydroxyl radicals onto the membrane surface [10]. Hence, reduce the hydrophobicity of the polymeric membrane. More so, Zhu et al. [279] reported the state-of-the-art on the control of microbial fouling using different types of nanoparticles. Essentially, they pointed that nanoparticles such as fullerenes (C60), Zinc based materials (ZIO-NM), silver-based particles (Ag-NM), grapheme oxide (GO-NM), and titanium dioxide (TiO2-NM), exhibit a strong antimicrobial effect. The authors further reported that the anti-biofouling mechanism of these materials is centered on the ability to release radicals with strong efficacy against the bio-foulants. This verdict is in agreement with the previous studies reported earlier [21,156].

4.1.3. Inorganic Foulants Generated during POME Treatment

As the name implies, this type of foulant is not organically originated, and they are normally categorized into an anion (CO32−, SO42−, F, OH) and cations (Al3+, Ca2+, Fe3+, Mg2+) [157]. The inorganic foulant is produced from soluble salts such as calcium phosphate, calcium carbonate, barium sulphate as well as silicon oxide. Primarily, the generation of inorganic foulants occurs through oxidation and hydrolysis to precipitate the ions to form a scale on the membrane. Researchers have considered the mechanism of the inorganic fouling into two stages: crystallization and particulate fouling [65,70]. During the crystallization stage, the ions were precipitated by the synergistic process of the oxidation and hydrolysis. The precipitates eventually got settled to form a coat on the membrane surface. While in the case of particulate fouling, the colloidal ions generated from the bulk mixed liquor are transported by convective and finally got deposited [51].
It is noteworthy that higher TMP encourage generation and agglomeration of ions, which subsequently got attracted to the unlike charged membrane surface to form scales layer [279]. The deposited and precipitated scale stratum of ions on the surface and pore wall of the membrane increases both hydraulic resistance and TMP. Then, the fouling developed under this condition is fundamentally referred to as inorganic [161]. In a general view, other factors such as the shear rate (scouring aeration), system configurations, feed physicochemical properties and membrane characteristics (such as surface roughness) could as well immediate the mechanism [259,279].
Ayyavoo et al [278] reported that variable operating condition have correlation with inorganic fouling rate. The results of their studies explicate that the composition of Ca2+, Mg2+ and Si2+ increases significantly as the aeration and TMP upsurges to a higher range (0.25–0.55 L/min and 15–35 psi respectively). In addition, the membrane fouling propensity increased by 400% at higher TMP and aeration. This suggests that as the operating conditions increases, the crystallization and particulate deposition of the colloidal matter prevails geometrically. As a result of this, it can be deduced that higher TMP and aeration could increase inorganic fouling [280,281]. Similarly, the effect of membrane surface geometry [258], filtration time [268], feed characteristics [153], and cross-flow velocity [108] on membrane fouling has been reported. Collectively, the findings pointed that the rate of fouling increases with the rough surface membrane, prolonging filtration time, higher organic loading rate and CFV. Furthermore, a higher range of CFV disintegrates the microbial flocs aggregate, thereby generating more colloids and particulates such as Ca2+ [65].
More interestingly, research has shown that moderate accumulation of Ca2+ ranging from 100 to 280 mg/L could mitigate organic fouling and bio-fouling. This is because the accumulated Ca2+ at this range has the ability to isolate and at the same time exert a binding effect on the organic foulants, and this could considerably hamper the metabolic activity of the bacteria colonies. However, excessive deposition of the ionic metal (Ca2+) beyond this range (100–280 mg/L) could cause severe fouling by the formation of hard scale layer [254].

4.1.4. Particulate Foulants as Related to POME Treatment

Particulate foulant has a wide range of sizes which may be grouped into settling particulate (size greater than 100 µm), supra-colloidal particulates (size less than 100 µm) and colloidal particulates (size ranged between 0.001 and 1 µm) [50]. They exhibit similar features with inorganic foulants in terms of fouling mechanism but they are different in requisites of origination [282,283,284]. In another view, Asadollahi et al. [285] have reported that particulate foulants comprise of cellular fine debris with high molecular weight proficient of forming an agglomerated colloids and/or coagulated fine fibres from palm fruits.
Basically, the particulates with same diameter or slightly less than the size of the membrane pores can easily initiate blockage while those with larger dimension contribute majorly in cake layer formation [50]. The particulates are transported either by convection, inertial lift, gravitational effect, or a combination of some or all factors to initiate blockage [286]. However, the efficacy of the mass transportation depends on operating conditions, particulate sizes, cross flow velocity and bulk concentrations present in the mixed liquor. It was on this note researchers have classified pore blocking into standard, complete, and intermediate [287].
Standard pore blocking involves narrowing of the pores as a result of the deposition and adsorption of the particulate onto pores-walls. Intermediate pore blocking is the partial accumulation of foulants to bridge the opening, while the complete pore blocking involves a total seal of the pores as a result of the deposited particles [287]. The blockage forms a barrier such that any further drifted particles continue to accumulate on the primary glazed (blockage) and this could eventually degenerate into piles of particulates or layer. The cake layer set in an antagonistic relationship between TMP and filtration or flux rate due to the exponential increase in hydraulic resistance (drag force) [287,288]. In this condition, rapid accumulation of particles occurs which gradually develops into scale layer. Hence, deterioration of filtration rate prevails while the TMP increases rapidly. The jump in TMP nurtures fouling and also compacts the accumulated particulates and transforms the scales into a hard cake layer [288]. This implies that membrane fouling passes through stages; starting with pore blockage, the gel formation and compaction [278].
Li et al. [289] reported that the higher fractal pores in the membrane morphology encourage more hydraulic resistance. This elucidates that membrane morphology has a strong correlation with the rate of cake formation. Similarly, Tijing et al. [185] reported that higher TMP operating condition increases the rate of cake formation by 80%. Also, they confirmed that the particulates with fine sizes often initiates blockage and also perform a significant role in the subsequent fouling mechanisms. This type of fouling is usually irreversible in most cases, as substantiated in the previous studies [65,290]. This denotes that smaller colloids or particulates could pose more severe fouling challenges compare to the larger foulants.

4.2. The Mechanism in Membrane Fouling During POME Treatment

Fouling mechanism is the developmental stages of foulants accumulation on the membrane surface and pore walls. The nature of the foulants generated and the fouling mechanism have a significant correlation with the applied treatment processes and the feed characteristics, as discussed earlier. Based on Figure 7, the central functions of the MBR system includes biodegradation and filtration. It is obvious that feed (POME) characteristics, treatments processes and membrane properties influence the rate and types of foulants generated, as well as its deposition method [65,156,254]. Firstly, In the biodegradation, the feed (POME) characteristics (such as biomass concentration), HRT, SRT, aeration, F/M ratio, pH and/or temperature are key to expedite the process [291,292]. The condition must be ideal for the microbial to ensure swiftly decomposition of the biomass. This implies that selection of the treatment conditions is fundamental for optimal biodegradation and also it determines the resulted types of foulants as well as its properties. Membrane filtration is the second crucial function of the MBR system as depicted in Figure 7. At the filtration stage, the bio-degraded biomass or activated sludge interacts with the membrane under the influence of the applied driving force (TMP). During this process, the constituted foulants in the concentrate polarizes the membrane and the resulted fouling is strongly determined by the membrane properties, feed characteristics and operating conditions. This shows that all the three factors (membrane properties, feed characteristics and operating conditions) play a significant role in the membrane selectivity, productivity and fouling mechanism. Thus, the separation mechanism is a complicated phenomenon due to the complexity and antagonistic interactions of the several factors [70,292].
However, the fouling mechanism has been simplified into three stages; the build-up or the conditioning, the steady-growing and abrupt increase in TMP stage, (see Figure 7) [157,292]. From Figure 7, it is obvious that EPS (soluble and bound), bacteria colonies or cells, precipitated ions, and the colloids/particulates all contributes to the initiation of pores blockage, but the input of each of the foulants varies [108]. The steady fouling stage involves the formation of gel and scale layer on the membrane surface and pore walls. Consequentially, this usually amounted to the gradual increase in the operating TMP. In most cases, if the filtration continues at this condition, the gel or biofilm could degenerate into the cake layer, as illustrated in Figure 7 and Figure 8c.

4.2.1. Build-up (Conditioning) Fouling Stage

The initial conditioning stage is usually stemmed from the stream transportation and interaction of the particulates with membrane modules [50,264]. This resulted in the deposition and adsorption of the foulants onto the pore walls. As depicted in Figure 8a, the particulate of smaller sizes is transmitted into the pores by the driving force such as TMP, convections, shear-induced diffusion, gravitational influences as well as literal migrations [293]. The deposited particulates developed into the thin glaze, thereby creating a barrier capable of obstructing the hydraulic conductivity of the permeate flow. As filtration proceeds, additional particles or foulants will continue to accumulate on the initiated layer and ultimately degenerate into the next stage of fouling, as illustrated pictorially in Figure 8a. Furthermore, the hydrolysis and oxidation process during filtration as well as the by-products released by the microbial also contributes considerably to the initiation of pores blockage. This is because of the surface scale formed by the inorganic ions and the secreted bio-glue substance immediate particulate attachments onto the pore walls [70,108]. Nonetheless, the severity of the blocking stage of fouling can be quelled using back-washing, intermittent filtration operation and optimal scouring aeration [249].

4.2.2. Steady Fouling Stage

This stage is the continuation of the fouling mechanism of the initiated pore blockage through the deposited tacky substance, commonly known as soluble microbial products (SMP) [293,294]. The sticky surface condition of the membrane promotes attachment and swift accumulation of the colloids or particulates, bio-flocs and SMP to form a gel layer, and this could lead to a complete or intermediate pore blockage [48,293,294]. Figure 8b denotes the fouling mechanism at the steady stage, the TMP is observed rising gradually due to the growing films or gel layer on the membrane pores and surface. However, the rate of gel formation can be palliated by adopting good practicings such as frequent back-washing and intermittent filtration-relaxation operation [201,254]. On the contrary, poor control of fouling encourages the growth of bio-layer. The underneath of the accumulated bio-layer lacks oxygen and this could lead to the higher death rate of the bacteria, thereby secreting more EPS to further glue the matrix of the biofilm. This situation is more distinctive to the third stage of fouling, where the compacted and denser cake layer is predominant [294].

4.2.3. An Abrupt Increase in TMP Stage

All types of foulants contribute to this fouling stage, as clearly shown in Figure 8c. From Figure 8c, it can be observed that SMP was the major initiator of the pore blockage, while the colloids, EPS and particulates contribute majorly to the gel and hard-cake layer formation, respectively [130,295,296]. This stage of fouling is characterized by a sudden increase in the driving pressure, this effect is commonly referred to as “TMP jump”. The observed TMP jump is due to the pores blockage and the accumulated layer on the membrane surface which induced the exorbitant hydraulic resistance [61,77]. Therefore, diminishing in the filtration volume or permeability flux prevails at this condition due to the narrowed or blocked pores. This implies that TMP and the permeability flux are reciprocal to one another as the foulants gradually settled on the membrane. Essentially, it can be inferred that the sudden increase in TMP and flux waning are the most obvious indicators that can be used to confirm abrupt fouling stage in a typical MBR system. Furthermore, prolonging filtration at jumped TMP not only influence cake layer formation but also exerts a compacting force to harden the layer [248,294]. At such situation, there is a high tendency for the formation of irreversible fouling in the membrane [50,65].

5. Conclusions

Currently, MBR technology has gained substantial attention with high interest in its application for POME treatment. The technology presents a great potential in treating POME under optimal treatment conditions. Application of aeration (AerMBR) offers a sizable enhancement in performance of MBR in terms of organic pollutants elimination and fouling vindication, more especially at the initial stage of the treatment. It eliminates fouling through continuous scouring of the deposited contaminants. However, as the retention of activated sludge prolongs, the shearing force of the scouring air could pose a serious drawback on flocs stability and breakup to generate more foulants. This is the major challenge of the aeration process in MBR, and therefore there is a need for further investigation.
Comparably, AnMBR has demonstrated a propitious performance in deflation of contaminants in high strength POME, resources recovery and minimization of operating energy, but high foulants polarization on the membrane surface during filtration is still the major weakness of this treatment process. Attempts to improve on the weakness of aerobic and anaerobic process were made by combining the processes to the hybrid system (HybMBR), but the problem of membrane fouling is still demanding and discourages wider application of the membrane technology. Moreover, recent advances in the use of MBR suggested that activated carbon (AC), sonication (SonMBR), and temperature (TheMBR) could reduce the susceptibility of the membrane to fouling through relegation of the concentration of the contaminant during pretreatment.
It is worth mentioning that combination of AC with MBR is technically demanding because of the required chemical treatment of absorbent (AC) which often leads to the creation of acidic effluent. In addition, this may increase the treatment footprint and operational cost due to the further treatments required to neutralize the pH. Moreover, in the case of sonication and temperature, they have high tendencies of distorting granulation stability, which could increase the density of small particles sizes and colloids. Therefore, there is a need for further investigation to establish optimal magnitude and duration for sonication as well as temperature in treating POME putting MLSS concentration, organic loading rate and solid retention time (SRT) into cognizance. However, AC presents noticeable advantages of reducing the total contaminants erstwhile to MBR treatment and this could reduce the fouling tendencies and improve flux recovery (91.7 to 95.3%). This offers ample solution to the problem of fouling if AC process is optimally integrated with aeration-anoxic processes in MBR.
On this note, additional investigations are required to develop a sustainable integrated treatment system and its optimal operating conditions such as aeration, biomass concentration and temperature. The integrated system includes the ideal the pre-treatment process using AC prior to membrane filtration in a single and one-term aeration-anoxic MBR treatment system (AA-MBR). Furthermore, to be able to develop such a treatment system, a good understanding of the adsorption process, biomass-substrate contact, stabilization conditions, and the combine process of aeration-anoxic is essential. Indispensably, a combination of AC and AA-MBR system will lessen vulnerability to fouling during POME treatment due to the concerted effect of the integrated technology. The integrated technology (AC and AA-MBR) could also improve filtration performance, contaminants removal efficiency, and result in a significant reduction in overall treatment cost.

Author Contributions

M.A. wrote the major part of the articles, H.C.M., A.I.I., and Z.Z.A. contributed substantially in the treatment processes, particularly, the novel application of aeration and composite membrane in MBR. The constructive and critical corrections of this paper by K.F.Y. must be mentioned, as well as the supportive information provided on the current trend of hybrid treatments for POME treatments.

Funding

This work was funded by Universiti Putra Malaysia (PUTRA Impak: 9530900) and the Tertiary Education Trust Fund (TETFUND).

Acknowledgments

Author(s) would like to appreciate the Universiti Putra Malaysia for the general support and provision of the basic facilities required for the study.

Conflicts of Interest

The authors sincerely professed no any interest conflicts

References

  1. Loh, S.K.; Nasrin, A.B.; Mohamad Azri, S.; Nurul Adela, B.; Muzzammil, N.; Daryl Jay, T.; Stasha Eleanor, R.A.; Lim, W.S.; Choo, Y.M.; Kaltschmitt, M. First Report on Malaysia’s experiences and development in biogas capture and utilization from palm oil mill effluent under the Economic Transformation Programme: Current and future perspectives. Renew. Sustain. Energy Rev. 2017, 74, 1257–1274. [Google Scholar] [CrossRef]
  2. Rupani, P.F.; Rajeev, P.S.; Irahim, M.H.; Esa, N. Review of Current Palm Oil Mill Effluent (POME) Treatment Methods: Vermicomposting as a Sustainable Practice. World Appl. Sci. J. 2010, 11, 70–81. [Google Scholar] [CrossRef]
  3. Wu, T.Y.; Mohammad, A.W.; Jahim, J.M.; Anuar, N. Pollution control technologies for the treatment of palm oil mill effluent (POME) through end-of-pipe processes. J. Environ. Manag. 2010, 91, 1467–1490. [Google Scholar] [CrossRef] [PubMed]
  4. Poh, P.E.; Chong, M.F. Development of anaerobic digestion methods for palm oil mill effluent (POME) treatment. Bioresour. Technol. 2009, 100, 1–9. [Google Scholar] [CrossRef] [PubMed]
  5. Boonrod, B.; Prapainainar, C.; Narataruksa, P.; Kantama, A.; Saibautrong, W.; Sudsakorn, K.; Mungcharoen, T.; Prapainainar, P. Evaluating the environmental impacts of bio-hydrogenated diesel production from palm oil and fatty acid methyl ester through life cycle assessment. J. Clean. Prod. 2017, 142, 1210–1221. [Google Scholar] [CrossRef]
  6. Choong, Y.Y.; Chou, K.W.; Norli, I. Strategies for improving biogas production of palm oil mill effluent (POME) anaerobic digestion: A critical review. Renew. Sustain. Energy Rev. 2018, 82, 2993–3006. [Google Scholar] [CrossRef]
  7. Ayu, E.D.; Halim, L.; Mellyanawaty, M.; Sudibyo, H.; Budhijanto, W. The effect of natural zeolite as microbial immobilization media in anaerobic digestion at various concentrations of palm oil mill effluent (POME). AIP Conf. Proc. 2017, 1840. [Google Scholar] [CrossRef]
  8. Okwute, O.L.; Isu, N.R. Impact Analysis of Palm Oil Mill Effluent on The Aerobic Bacterial Density and Ammonium Oxidizers in A Dumpsite in Anyigba, Kogi State. Afr. J. Biotechnol. 2007, 6, 116–119. [Google Scholar]
  9. Dadrasnia, A.; Usman, M.M.; Lim, K.T.; Velappan, R.D.; Vejan, P.; Mahmud, A.F.; Ismail, S. Microbial Aspects in Wastewater Treatment—A Technical Review. Environ. Pollut. Prot. 2017, 2, 75–84. [Google Scholar]
  10. Abu Bakar, S.N.H.; Abu Hasan, H.; Mohammad, A.W.; Sheikh Abdullah, S.R.; Haan, T.Y.; Ngteni, R.; Yusof, K.M.M. A review of moving-bed biofilm reactor technology for palm oil mill effluent treatment. J. Clean. Prod. 2018, 171, 1532–1545. [Google Scholar] [CrossRef]
  11. Ho, K.C.; Teow, Y.H.; Ang, W.L.; Mohammad, A.W. Novel GO/OMWCNTs mixed-matrix membrane with enhanced antifouling property for palm oil mill effluent treatment. Sep. Purif. Technol. 2017, 177, 337–349. [Google Scholar] [CrossRef]
  12. Azmi, N.S.; Yunos, K.F.M. Wastewater Treatment of Palm Oil Mill Effluent (POME) by Ultrafiltration Membrane Separation Technique Coupled with Adsorption Treatment as Pre-treatment. Agric. Agric. Sci. Procedia 2014, 2, 257–264. [Google Scholar] [CrossRef]
  13. Daelman, M.R.J.; van Voorthuizen, E.M.; van Dongen, U.G.J.M.; Volcke, E.I.P.; van Loosdrecht, M.C.M. Methane emission during municipal wastewater treatment. Water Res. 2012, 46, 3657–3670. [Google Scholar] [CrossRef] [PubMed]
  14. Din, A.K. Malaysian Oil Palm Industry Performance 2016 and Prospects for 2017; MPOB: Bandar Baru Bangi, Malaysia, 2017. [Google Scholar]
  15. Zhang, Y.; Yan, L.; Qiao, X.; Chi, L.; Niu, X.; Mei, Z.; Zhang, Z. Integration of biological method and membrane technology in treating palm oil mill effluent. J. Environ. Sci. 2008, 20, 558–564. [Google Scholar] [CrossRef]
  16. Taha, M.R.; Ibrahim, A.H. COD removal from anaerobically treated palm oil mill effluent (AT-POME) via aerated heterogeneous Fenton process: Optimization study. J. Water Process Eng. 2014, 1, 8–16. [Google Scholar] [CrossRef]
  17. Ahmad, A.L.; Ismail, S.; Bhatia, S. Membrane treatment for palm oil mill effluent: Effect of transmembrane pressure and crossflow velocity. Desalination 2005, 179, 245–255. [Google Scholar] [CrossRef]
  18. Yuniarto, A.; Noor, Z.Z.; Ujang, Z.; Olsson, G.; Aris, A.; Hadibarata, T. Bio-fouling reducers for improving the performance of an aerobic submerged membrane bioreactor treating palm oil mill effluent. Desalination 2013, 316, 146–153. [Google Scholar] [CrossRef]
  19. Ali Amat, N.A.; Tan, Y.H.; Lau, W.J.; Lai, G.S.; Ong, C.S.; Mokhtar, N.M.; Sani, N.A.A.; Ismail, A.F.; Goh, P.S.; Chong, K.C.; et al. Tackling colour issue of anaerobically-treated palm oil mill effluent using membrane technology. J. Water Process Eng. 2015, 8, 221–226. [Google Scholar] [CrossRef]
  20. Wang, J.; Mahmood, Q.; Qiu, J.P.; Li, Y.S.; Chang, Y.S.; Li, X.D. Anaerobic Treatment of Palm Oil Mill Effluent in Pilot-Scale Anaerobic EGSB Reactor. Biomed. Res. Int. 2015, 2015. [Google Scholar] [CrossRef] [PubMed]
  21. Subramaniam, M.N.; Goh, P.S.; Lau, W.J.; Tan, Y.H.; Ng, B.C.; Ismail, A.F. Hydrophilic hollow fiber PVDF ultrafiltration membrane incorporated with titanate nanotubes for decolourization of aerobically-treated palm oil mill effluent. Chem. Eng. J. 2017, 316, 101–110. [Google Scholar] [CrossRef]
  22. Ye, Y.; Saikaly, P.E.; Logan, B.E. Simultaneous nitrogen and organics removal using membrane aeration and effluent ultrafiltration in an anaerobic fluidized membrane bioreactor. Bioresour. Technol. 2017, 244, 456–462. [Google Scholar] [CrossRef] [PubMed]
  23. Judd, S.; Judd, C. The MBR Book: Principles and Applications of Membrane Bioreactors in Water and Wastewater Treatment; Elsevier: New York, NY, USA, 2008; ISBN 9781856174817. [Google Scholar]
  24. Ahmad, Z.; Ridzuan, M.B.; Daud, Z. Membrane Bioreactor for Palm Oil Mill Effluent and Resource Recovery. In Proceedings of the International Conference on Sustainable Development for Water and Waste Water Treatment, Yogyakarta, Indonesia, 14–15 December 2009; pp. 1–8. [Google Scholar]
  25. Ghani, M.S.H.; Haan, T.Y.; Lun, A.W.; Mohammad, A.W.; Ngteni, R.; Yusof, K.M.M. Fouling assessment of tertiary palm oil mill effluent (POME) membrane treatment for water reclamation. J. Water Reuse Desalin. 2017, 8, 412–423. [Google Scholar] [CrossRef] [Green Version]
  26. Wang, Z.; Wu, Z.; Yin, X.; Tian, L. Membrane fouling in a submerged membrane bioreactor (MBR) under sub-critical flux operation: Membrane foulant and gel layer characterization. J. Membr. Sci. 2008, 325, 238–244. [Google Scholar] [CrossRef]
  27. Judd, S.J. The status of industrial and municipal effluent treatment with membrane bioreactor technology. Chem. Eng. J. 2015. [Google Scholar] [CrossRef]
  28. Krzeminski, P.; Leverette, L.; Malamis, S.; Katsou, E. Membrane bioreactors—A review on recent developments in energy reduction, fouling control, novel configurations, LCA and market prospects. J. Membr. Sci. 2017, 527, 207–227. [Google Scholar] [CrossRef]
  29. Vijayaraghavan, K.; Ahmad, D.; Ezani Bin Abdul Aziz, M. Aerobic treatment of palm oil mill effluent. J. Environ. Manag. 2007, 82, 24–31. [Google Scholar] [CrossRef] [PubMed]
  30. Liu, F.; Hashim, N.A.; Liu, Y.; Abed, M.R.M.; Li, K. Progress in the production and modification of PVDF membranes. J. Membr. Sci. 2011, 375, 1–27. [Google Scholar] [CrossRef]
  31. Rana, S.; Singh, L.; Wahid, Z.; Liu, H. A Recent Overview of Palm Oil Mill Effluent Management via Bioreactor Configurations. Curr. Pollut. Rep. 2017, 3, 254–267. [Google Scholar] [CrossRef]
  32. Coutte, F.; Lecouturier, D.; Dimitrov, K.; Guez, J.S.; Delvigne, F.; Dhulster, P.; Jacques, P. Microbial lipopeptide production and purification bioprocesses, current progress and future challenges. Biotechnol. J. 2017, 12, 1–10. [Google Scholar] [CrossRef] [PubMed]
  33. Ohimain, E.I.; Izah, S.C. A review of biogas production from palm oil mill effluents using different configurations of bioreactors. Renew. Sustain. Energy Rev. 2017, 70, 242–253. [Google Scholar] [CrossRef]
  34. Bello, M.M.; Abdul Raman, A.A. Trend and current practices of palm oil mill effluent polishing: Application of advanced oxidation processes and their future perspectives. J. Environ. Manag. 2017, 198, 170–182. [Google Scholar] [CrossRef] [PubMed]
  35. Mohammed, R.R.; Chong, M.F. Treatment and decolorization of biologically treated Palm Oil Mill Effluent (POME) using banana peel as novel biosorbent. J. Environ. Manag. 2014, 132, 237–249. [Google Scholar] [CrossRef] [PubMed]
  36. Ohimain, E.I.; Izah, S.C.; Jenakumo, N. Physicochemical and microbial screening of palm oil mill effluent for amylase production by mill effluents for amylase production. Greener J. Bol. Sci. 2013, 3, 307–318. [Google Scholar]
  37. Wang, J.; Mahmood, Q.; Qiu, J.P.; Li, Y.S.; Chang, Y.S.; Chi, L.N.; Li, X.D. Zero discharge performance of an industrial pilot-scale plant treating palm oil mill effluent. Biomed. Res. Int. 2015, 2015. [Google Scholar] [CrossRef] [PubMed]
  38. Chan, Y.J.; Chong, M.F.; Law, C.L.; Hassell, D.G. A review on anaerobic-aerobic treatment of industrial and municipal wastewater. Chem. Eng. J. 2009, 155, 1–18. [Google Scholar] [CrossRef]
  39. Shon, H.K.; Vigneswaran, S.; Snyder, S.A. Effluent organic matter (EfOM) in wastewater: Constituents, effects, and treatment. Crit. Rev. Environ. Sci. Technol. 2006, 36, 327–374. [Google Scholar] [CrossRef] [Green Version]
  40. Mutamim, N.S.A.; Noor, Z.Z.; Hassan, M.A.A.; Yuniarto, A.; Olsson, G. Membrane bioreactor: Applications and limitations in treating high strength industrial wastewater. Chem. Eng. J. 2013, 225, 109–119. [Google Scholar] [CrossRef]
  41. Nwuche, C.O.; Aoyagi, H.; Ogbonna, J.C. Treatment of Palm Oil Mill Effluent by a Microbial Consortium Developed from Compost Soils. Int. Sch. Res. Not. 2014, 2014. [Google Scholar] [CrossRef] [PubMed]
  42. Samudro, G.; Mangkoedihardjo, S. Review on Bod, Cod and Bod/Cod Ratio: A Triangle Zone for Toxic, Biodegradable and Stable Levels. Int. Acad. Res. 2010, 2, 235–239. [Google Scholar]
  43. Mutamim, N.S.A.; Noor, Z.Z.; Hassan, M.A.A.; Olsson, G. Application of membrane bioreactor technology in treating high strength industrial wastewater: A performance review. Desalination 2012, 305, 1–11. [Google Scholar] [CrossRef]
  44. Chan, Y.J.; Chong, M.F.; Law, C.L. Bioresource Technology Start-up, steady state performance and kinetic evaluation of a thermophilic integrated anaerobic–aerobic bioreactor (IAAB). Bioresour. Technol. 2012, 125, 145–157. [Google Scholar] [CrossRef] [PubMed]
  45. Bala, J.D.; Lalung, J.; Ismail, N. Biodegradation of palm oil mill effluent (POME) by bacterial. Int. J. Sci. Res. Publ. 2014, 4, 2250–3153. [Google Scholar]
  46. Chin, M.J.; Poh, P.E.; Tey, B.T.; Chan, E.S.; Chin, K.L. Biogas from palm oil mill effluent (POME): Opportunities and challenges from Malaysia’s perspective. Renew. Sustain. Energy Rev. 2013, 26, 717–726. [Google Scholar] [CrossRef]
  47. Tabassum, S.; Zhang, Y.; Zhang, Z. An Integrated Method for Palm Oil Mill Effluent (POME) Treatment for Achieving Zero Liquid Discharge—A Pilot Study; Elsevier Ltd.: New York, NY, USA, 2015; Volume 95, ISBN 1522119574. [Google Scholar]
  48. Zinatizadeh, A.A.L.; Mohamed, A.R.; Najafpour, G.D.; Hasnain Isa, M.; Nasrollahzadeh, H. Kinetic evaluation of palm oil mill effluent digestion in a high rate up-flow anaerobic sludge fixed film bioreactor. Process Biochem. 2006, 41, 1038–1046. [Google Scholar] [CrossRef]
  49. Comte, I.; Colin, F.; Whalen, J.K.; Grünberger, O.; Caliman, J.P. Agricultural Practices in Oil Palm Plantations and Their Impact on Hydrological Changes, Nutrient Fluxes and Water Quality in Indonesia. A Review; Elsevier Ltd.: New York, NY, USA, 2012; Volume 116, ISBN 9780123942777. [Google Scholar]
  50. Guo, W.; Ngo, H.H.; Li, J. A mini-review on membrane fouling. Bioresour. Technol. 2012, 122, 27–34. [Google Scholar] [CrossRef] [PubMed]
  51. Bello, M.M.; Nourouzi, M.M.; Abdullah, L.C.; Choong, T.S.Y.; Koay, Y.S.; Keshani, S. POME is treated for removal of color from biologically treated POME in fixed bed column: Applying wavelet neural network (WNN). J. Hazard. Mater. 2013, 262, 106–113. [Google Scholar] [CrossRef] [PubMed]
  52. Khemkhao, M.; Techkarnjanaruk, S.; Phalakornkule, C. Simultaneous treatment of raw palm oil mill effluent and biodegradation of palm fiber in a high-rate CSTR. Bioresour. Technol. 2015, 177, 17–27. [Google Scholar] [CrossRef] [PubMed]
  53. Ahmad, A.; Ghufran, R.; Wahid, Z.A. Bioenergy from anaerobic degradation of lipids in palm oil mill effluent. Rev. Environ. Sci. Biotechnol. 2011, 10, 353–376. [Google Scholar] [CrossRef] [Green Version]
  54. Alhaji, M.H.; Sanaullah, K.; Lim, S.F.; Khan, A.; Hipolito, C.N.; Abdullah, M.O.; Bhawani, S.A.; Jamil, T. Photocatalytic treatment technology for palm oil mill effluent (POME)—A review. Process Saf. Environ. Prot. 2016, 102, 673–686. [Google Scholar] [CrossRef]
  55. Chan, Y.J.; Chong, M.F.; Law, C.L. An integrated anaerobic-aerobic bioreactor (IAAB) for the treatment of palm oil mill effluent (POME): Start-up and steady state performance. Process Biochem. 2012, 47, 485–495. [Google Scholar] [CrossRef]
  56. Damayanti, A.; Ujang, Z.; Salim, M.R. The influenced of PAC, zeolite, and Moringa oleifera as biofouling reducer (BFR) on hybrid membrane bioreactor of palm oil mill effluent (POME). Bioresour. Technol. 2011, 102, 4341–4346. [Google Scholar] [CrossRef] [PubMed]
  57. Melin, T.; Jefferson, B.; Bixio, D.; Thoeye, C.; De Wilde, W.; De Koning, J.; van der Graaf, J.; Wintgens, T. Membrane bioreactor technology for wastewater treatment and reuse. Desalination 2006, 187, 271–282. [Google Scholar] [CrossRef]
  58. Ahmad, Z.; Ujang, Z.; Olsson, G. Biomass Effect on Membrane Fouling Using a Hybrid Membrane Bioreactor for Palm Oil Mill Effluent. In Proceedings of the 1st IWA Malaysia Young Water Professionals Conference (IWAYP2010), Kuala, Lumpur, 1–4 March 2010. [Google Scholar]
  59. Kraume, M.; Drews, A. Membrane Bioreactors in Waste Water Treatment—Status and Trends. Chem. Eng. Technol. 2010, 33, 1251–1259. [Google Scholar] [CrossRef]
  60. Oller, I.; Malato, S.; Sánchez-Pérez, J.A. Combination of Advanced Oxidation Processes and biological treatments for wastewater decontamination—A review. Sci. Total Environ. 2011, 409, 4141–4166. [Google Scholar] [CrossRef] [PubMed]
  61. Faisal, M.; Machdar, I.; Gani, A.; Daimon, H. The Combination of Air Flotation and a Membrane Bioreactor for the Treatment of Palm Oil Mill Effluent. Int. J. Technol. 2016, 7, 767. [Google Scholar] [CrossRef] [Green Version]
  62. Park, H.; Chang, I.; Lee, K. Principles of Membrane Bioreactors for Wastewater Treatment Waste Activated Sludge; CRC Press: Boca Raton, FL, USA, 2015; ISBN 9781466590380. [Google Scholar]
  63. Mamimin, C.; Chaikitkaew, S.; Niyasom, C.; Kongjan, P.; O.-Thong, S. Effect of Operating Parameters on Process Stability of Continuous Biohydrogen Production from Palm Oil Mill Effluent under Thermophilic Condition. Energy Procedia 2015, 79, 815–821. [Google Scholar] [CrossRef]
  64. Radjenović, J.; Matošić, M.; Mijatović, I.; Petrović, M.; Barceló, D. Membrane Bioreactor (MBR) as an Advanced Wastewater Treatment Technology. In Emerging Contaminants from Industrial and Municipal Waste; The Handbook of Environmental Chemistry; Springer: Berlin/Heidelberg, Germany, 2008; Volume 5, pp. 37–101. [Google Scholar] [CrossRef]
  65. Meng, F.; Zhang, S.; Oh, Y.; Zhou, Z.; Shin, H.S.; Chae, S.R. Fouling in membrane bioreactors: An updated review. Water Res. 2017, 114, 151–180. [Google Scholar] [CrossRef] [PubMed]
  66. Verrecht, B.; Judd, S.; Guglielmi, G.; Brepols, C.; Mulder, J.W. An aeration energy model for an immersed membrane bioreactor. Water Res. 2008, 42, 4761–4770. [Google Scholar] [CrossRef] [PubMed]
  67. Wang, Z.; Wu, Z.; Mai, S.; Yang, C.; Wang, X.; An, Y.; Zhou, Z. Research and applications of membrane bioreactors in China: Progress and prospect. Sep. Purif. Technol. 2008, 62, 249–263. [Google Scholar] [CrossRef]
  68. Fazal, S.; Zhang, B.; Zhong, Z.; Gao, L.; Chen, X. Industrial Wastewater Treatment by Using MBR (Membrane Bioreactor) Review Study. J. Environ. Prot. 2015, 06, 584–598. [Google Scholar] [CrossRef]
  69. Kizilet, A.; Veral, M.A.; Amar, C.; Isik, O.; Bahramian, M. The Use of Membrane Processes to Promote Sustainable Environmental Protection Practices; DergiPark: Çankaya/Ankara, Turkey, 2017. [Google Scholar]
  70. Meng, F.; Chae, S.R.; Drews, A.; Kraume, M.; Shin, H.S.; Yang, F. Recent advances in membrane bioreactors (MBRs): Membrane fouling and membrane material. Water Res. 2009, 43, 1489–1512. [Google Scholar] [CrossRef] [PubMed]
  71. Wang, H.; Zhang, L.; Li, Y.; Hu, C. Influence of Filtration Aids on Continuous Filtration in Membrane Bioreactors. Ind. Eng. Chem. Res. 2014, 53, 7202–7208. [Google Scholar] [CrossRef]
  72. Wan, C.F.; Yang, T.; Lipscomb, G.G.; Stookey, D.J.; Chung, T.S. Design and fabrication of hollow fiber membrane modules. J. Membr. Sci. 2017, 538, 96–107. [Google Scholar] [CrossRef]
  73. Li, D.; Wang, R.; Chung, T.S. Fabrication of lab-scale hollow fiber membrane modules with high packing density. Sep. Purif. Technol. 2004, 40, 15–30. [Google Scholar] [CrossRef]
  74. Yang, W.; Cicek, N.; Ilg, J. State-of-the-art of membrane bioreactors: Worldwide research and commercial applications in North America. J. Membr. Sci. 2006, 270, 201–211. [Google Scholar] [CrossRef]
  75. Wang, Y.; Ong, K.W.; Brannock, M.W.D.; Leslie, G.L. Evaluation of membrane bioreactor performance via residence time distribution: Effects of membrane configuration and mixing. Water Sci. Technol. 2008, 57, 353–359. [Google Scholar] [CrossRef] [PubMed]
  76. Shakibabarough, A.; Valinejadshoubi, M.; Valinejadshoubi, M. Useable and Precautionary Aspects of Using Nanotechnology and Nano- materials in the Construction Industry. Int. J. Sci. 2014, 3, 841–848. [Google Scholar]
  77. Drews, A. Membrane fouling in membrane bioreactors—Characterisation, contradictions, cause and cures. J. Membr. Sci. 2010, 363, 1–28. [Google Scholar] [CrossRef]
  78. Mamimin, C.; Prasertsan, P.; Kongjan, P.; O.-Thong, S. Effects of volatile fatty acids in biohydrogen effluent on biohythane production from palm oil mill effluent under thermophilic condition. Electron. J. Biotechnol. 2017, 29, 78–85. [Google Scholar] [CrossRef]
  79. Mondal, S. Polymeric membranes for produced water treatment: An overview of fouling behavior and its control. Rev. Chem. Eng. 2016, 32, 611–628. [Google Scholar] [CrossRef]
  80. Vasanth, D.; Pugazhenthi, G.; Uppaluri, R. Fabrication and properties of low cost ceramic microfiltration membranes for separation of oil and bacteria from its solution. J. Membr. Sci. 2011, 379, 154–163. [Google Scholar] [CrossRef]
  81. Zhu, K.; Zhang, S.; Luan, J.; Mu, Y.; Du, Y.; Wang, G. Fabrication of ultrafiltration membranes with enhanced antifouling capability and stable mechanical properties via the strategies of blending and crosslinking. J. Membr. Sci. 2017, 539, 116–127. [Google Scholar] [CrossRef]
  82. Fane, A.G.; Wang, R.; Hu, M.X. Synthetic membranes for water purification: Status and future. Angew. Chem. Int. Ed. 2015, 54, 3368–3386. [Google Scholar] [CrossRef] [PubMed]
  83. Vieira Salla, A.C.; Margarites, A.C.; Seibel, F.I.; Holz, L.C.; Brião, V.B.; Bertolin, T.E.; Colla, L.M.; Costa, J.A.V. Increase in the carbohydrate content of the microalgae Spirulina in culture by nutrient starvation and the addition of residues of whey protein concentrate. Bioresour. Technol. 2016, 209, 133–141. [Google Scholar] [CrossRef] [PubMed]
  84. Arabi, S.; Nakhla, G. Impact of protein/carbohydrate ratio in the feed wastewater on the membrane fouling in membrane bioreactors. J. Membr. Sci. 2008, 324, 142–150. [Google Scholar] [CrossRef]
  85. Van Reis, R.; Zydney, A. Bioprocess membrane technology. J. Membr. Sci. 2007, 297, 16–50. [Google Scholar] [CrossRef]
  86. Çakmakce, M.; Kayaalp, N.; Koyuncu, I. Desalination of produced water from oil production fields by membrane processes. Desalination 2008, 222, 176–186. [Google Scholar] [CrossRef]
  87. Mohammad, A.W.; Teow, Y.H.; Ang, W.L.; Chung, Y.T.; Oatley-Radcliffe, D.L.; Hilal, N. Nanofiltration membranes review: Recent advances and future prospects. Desalination 2015, 356, 226–254. [Google Scholar] [CrossRef]
  88. Onsekizoglu, P.; Ng, L.Y.; Mohammad, W.A.W. Principle, Advances, Limitations and Future Prospects in Food Industry. J. Membr. Sci. 2015, 344, 226–254. [Google Scholar] [CrossRef]
  89. Sorayani Bafqi, M.S.; Bagherzadeh, R.; Latifi, M. Fabrication of composite PVDF-ZnO nanofiber mats by electrospinning for energy scavenging application with enhanced efficiency. J. Polym. Res. 2015, 22, 130. [Google Scholar] [CrossRef]
  90. Yan, Z.Q.; Zeng, L.M.; Li, Q.; Liu, T.Y.; Matsuyama, H.; Wang, X.L. Selective separation of chloride and sulfate by nanofiltration for high saline wastewater recycling. Sep. Purif. Technol. 2016, 166, 135–141. [Google Scholar] [CrossRef]
  91. Silva, V.; Geraldes, V.; Brites Alves, A.M.; Palacio, L.; Prádanos, P.; Hernández, A. Multi-ionic nanofiltration of highly concentrated salt mixtures in the seawater range. Desalination 2011, 277, 29–39. [Google Scholar] [CrossRef]
  92. Shanmuganathan, S.; Vigneswaran, S.; Nguyen, T.V.; Loganathan, P.; Kandasamy, J. Use of nanofiltration and reverse osmosis in reclaiming micro-filtered biologically treated sewage effluent for irrigation. Desalination 2015, 364, 119–125. [Google Scholar] [CrossRef]
  93. Bunani, S.; Yörükoğlu, E.; Yüksel, Ü.; Kabay, N.; Yüksel, M.; Sert, G. Application of reverse osmosis for reuse of secondary treated urban wastewater in agricultural irrigation. Desalination 2015, 364, 68–74. [Google Scholar] [CrossRef]
  94. Nicolini, J.V.; Borges, C.P.; Ferraz, H.C. Selective rejection of ions and correlation with surface properties of nanofiltration membranes. Sep. Purif. Technol. 2016, 171, 238–247. [Google Scholar] [CrossRef]
  95. Miller, D.J.; Kasemset, S.; Wang, L.; Paul, D.R.; Freeman, B.D. Constant flux crossflow filtration evaluation of surface-modified fouling-resistant membranes. J. Membr. Sci. 2014, 452, 171–183. [Google Scholar] [CrossRef]
  96. Tummons, E.N.; Tarabara, V.V.; Chew, J.W.; Fane, A.G. Behavior of oil droplets at the membrane surface during crossflow microfiltration of oil-water emulsions. J. Membr. Sci. 2016, 500, 211–224. [Google Scholar] [CrossRef]
  97. Winans, J.D.; Smith, K.J.P.; Gaborski, T.R.; Roussie, J.A.; McGrath, J.L. Membrane capacity and fouling mechanisms for ultrathin nanomembranes in dead-end filtration. J. Membr. Sci. 2016, 499, 282–289. [Google Scholar] [CrossRef] [Green Version]
  98. Chen, R.; Nie, Y.; Hu, Y.; Miao, R.; Utashiro, T.; Li, Q.; Xu, M. Fouling behaviour of soluble microbial products and extracellular polymeric substances in a submerged anaerobic membrane bioreactor treating low- strength wastewater at room temperature. J. Membr. Sci. 2017, 531, 1–9. [Google Scholar] [CrossRef]
  99. Bacchin, P. Membranes: A variety of energy landscapes for many transfer opportunities. Membranes (Basel) 2018, 8, 10. [Google Scholar] [CrossRef] [PubMed]
  100. Le Clech, P.; Jefferson, B.; Chang, I.S.; Judd, S.J. Critical flux determination by the flux-step method in a submerged membrane bioreactor. J. Membr. Sci. 2003, 227, 81–93. [Google Scholar] [CrossRef]
  101. He, Z.; Miller, D.J.; Kasemset, S.; Paul, D.R.; Freeman, B.D. The effect of permeate flux on membrane fouling during microfiltration of oily water. J. Membr. Sci. 2017, 525, 25–34. [Google Scholar] [CrossRef]
  102. Dutta, K.; Das, S.; Kumar, P.; Kundu, P.P. Polymer electrolyte membrane with high selectivity ratio for direct methanol fuel cells: A preliminary study based on blends of partially sulfonated polymers polyaniline and PVdF-co-HFP. Appl. Energy 2014, 118, 183–191. [Google Scholar] [CrossRef]
  103. Sert, G.; Bunani, S.; Yörükoğlu, E.; Kabay, N.; Egemen, Ö.; Arda, M.; Yüksel, M. Performances of some NF and RO membranes for desalination of MBR treated wastewater. J. Water Process Eng. 2017, 16, 193–198. [Google Scholar] [CrossRef]
  104. White, N.; Misovich, M.; Alemayehu, E.; Yaroshchuk, A.; Bruening, M.L. Highly selective separations of multivalent and monovalent cations in electrodialysis through Nafion membranes coated with polyelectrolyte multilayers. Polymer 2016, 103, 478–485. [Google Scholar] [CrossRef] [Green Version]
  105. Ge, L.; Wu, B.; Yu, D.; Mondal, A.N.; Hou, L.; Afsar, N.U.; Li, Q.; Xu, T.; Miao, J.; Xu, T. Monovalent cation perm-selective membranes (MCPMs): New developments and perspectives. Chin. J. Chem. Eng. 2017, 25, 1606–1615. [Google Scholar] [CrossRef]
  106. Vaughn, J.T.; Koros, W.J. Analysis of feed stream acid gas concentration effects on the transport properties and separation performance of polymeric membranes for natural gas sweetening: A comparison between a glassy and rubbery polymer. J. Membr. Sci. 2014, 465, 107–116. [Google Scholar] [CrossRef]
  107. Martin-Garcia, I.; Monsalvo, V.; Pidou, M.; Le-Clech, P.; Judd, S.J.; McAdam, E.J.; Jefferson, B. Impact of membrane configuration on fouling in anaerobic membrane bioreactors. J. Membr. Sci. 2011, 382, 41–49. [Google Scholar] [CrossRef]
  108. Le-Clech, P.; Chen, V.; Fane, T.A.G. Fouling in membrane bioreactors used in wastewater treatment. J. Membr. Sci. 2006, 284, 17–53. [Google Scholar] [CrossRef]
  109. Deowan, S.A.; Bouhadjar, S.I.; Hoinkis, J. Membrane bioreactors for water treatment. Adv. Membr. Technol. Water Treat. Mater. Process. Appl. 2015, 155–184. [Google Scholar] [CrossRef]
  110. Ferrentino, R.; Langone, M.; Merzari, F.; Tramonte, L.; Andreottola, G. A review of anaerobic side-stream reactor for excess sludge reduction: Configurations, mechanisms, and efficiency. Crit. Rev. Environ. Sci. Technol. 2016, 46, 382–405. [Google Scholar] [CrossRef]
  111. Le-Clech, P.; Jefferson, B.; Judd, S.J. Impact of aeration, solids concentration and membrane characteristics on the hydraulic performance of a membrane bioreactor. J. Membr. Sci. 2003, 218, 117–129. [Google Scholar] [CrossRef]
  112. Xue, Y.; Zhao, H.; Ge, L.; Chen, Z.; Dang, Y.; Sun, D. Comparison of the performance of waste leachate treatment in submerged and recirculated membrane bioreactors. Int. Biodeterior. Biodegrad. 2015, 102, 73–80. [Google Scholar] [CrossRef]
  113. Morrow, C.P.; McGaughey, A.L.; Hiibel, S.R.; Childress, A.E. Submerged or sidestream? The influence of module configuration on fouling and salinity in osmotic membrane bioreactors. J. Membr. Sci. 2017. [Google Scholar] [CrossRef]
  114. Aslam, M.; Charfi, A.; Lesage, G.; Heran, M. Membrane Bioreactors for Wastewater Treatment: A review of mechanical cleaning by scouring agents to control membrane fouling. Chem. Eng. J. 2016. [Google Scholar] [CrossRef]
  115. Thomas, H.; Judd, S.; Murrer, J. Fouling characteristics of membrane filtration in membrane bioreactors. Membr. Technol. 2000, 2000, 10–13. [Google Scholar] [CrossRef]
  116. Cicek, N.; Winnen, H.; Suidan, M.T.; Wrenn, B.E.; Urbain, V.; Manem, J. Effectiveness of the membrane bioreactor in the biodegradation of high molecular weight compounds. Water Res. 1998, 32, 1553–1563. [Google Scholar] [CrossRef]
  117. Wisniewski, C. Membrane bioreactor for water reuse. Desalination 2007, 203, 15–19. [Google Scholar] [CrossRef]
  118. Zsirai, T.; Buzatu, P.; Aerts, P.; Judd, S. Efficacy of relaxation, backflushing, chemical cleaning and clogging removal for an immersed hollow fibre membrane bioreactor. Water Res. 2012, 46, 4499–4507. [Google Scholar] [CrossRef] [PubMed]
  119. Dvořák, L.; Gómez, M.; Dolina, J.; Černín, A. Anaerobic membrane bioreactors—A mini review with emphasis on industrial wastewater treatment: Applications, limitations and perspectives. Desalin. Water Treat. 2016, 57, 19062–19076. [Google Scholar] [CrossRef]
  120. Andrade, L.H.; Mendes, F.D.S.; Espindola, J.C.; Amaral, M.C.S. Reuse of dairy wastewater treated by membrane bioreactor and nanofiltration: Technical and economic feasibility. Braz. J. Chem. Eng. 2015, 32, 735–747. [Google Scholar] [CrossRef]
  121. Gander, M.; Jefferson, B.; Judd, S. Aerobic MBRs for domestic wastewater treatment: A review with cost considerations. Sep. Purif. Technol. 2000, 18, 119–130. [Google Scholar] [CrossRef]
  122. Jefferson, B.; Laine, A.L.; Judd, S.J.; Stephenson, T. Membrane bioreactors and their role in wastewater reuse. Water Sci. Technol. 2000, 41, 197–204. [Google Scholar] [CrossRef]
  123. Le-Clech, P. Membrane bioreactors and their uses in wastewater treatments. Appl. Microbiol. Biotechnol. 2010, 88, 1253–1260. [Google Scholar] [CrossRef] [PubMed]
  124. Visvanathan, C.; Yang, B.S.; Muttamara, S.; Maythanukhraw, R. Application of air backflushing technique in membrane bioreactor. Water Sci. Technol. 1997, 36, 259–266. [Google Scholar] [CrossRef] [Green Version]
  125. Yin, J.; Zhu, G.; Deng, B. Graphene oxide (GO) enhanced polyamide (PA) thin-film nanocomposite (TFN) membrane for water purification. Desalination 2016, 379, 93–101. [Google Scholar] [CrossRef]
  126. Xiao, K.; Shen, Y.X.; Liang, S.; Liang, P.; Wang, X.M.; Huang, X. A systematic analysis of fouling evolution and irreversibility behaviors of MBR supernatant hydrophilic/hydrophobic fractions during microfiltration. J. Membr. Sci. 2014, 467, 206–216. [Google Scholar] [CrossRef]
  127. Abdurahman, N.H.; Rosli, Y.M.; Azhari, N.H.; Ahmad, A.L.A.; Ghufran, R.; Wahid, Z.A. Hydrogen production by biological processes: a survey of literature. Int. J. Hydrogen Energy 2010, 27, 13–28. [Google Scholar] [CrossRef]
  128. Chang, I.S.; Le Clech, P.; Jefferson, B.; Judd, S. Membrane Fouling in Membrane Bioreactors for Wastewater Treatment. J. Environ. Eng. 2002, 128, 1018–1029. [Google Scholar]
  129. Yusuf, Z.; Wahab, N.A.; Sahlan, S. Fouling control strategy for submerged membrane bioreactor filtration processes using aeration airflow, backwash, and relaxation: A review. Desalin. Water Treat. 2015, 57, 17683–17695. [Google Scholar] [CrossRef]
  130. Lin, H.; Zhang, M.; Wang, F.; Meng, F.; Liao, B.Q.; Hong, H.; Chen, J.; Gao, W. A critical review of extracellular polymeric substances (EPSs) in membrane bioreactors: Characteristics, roles in membrane fouling and control strategies. J. Membr. Sci. 2014, 460, 110–125. [Google Scholar] [CrossRef]
  131. Meng, F.; Yang, F.; Shi, B.; Zhang, H. A comprehensive study on membrane fouling in submerged membrane bioreactors operated under different aeration intensities. Sep. Purif. Technol. 2008, 59, 91–100. [Google Scholar] [CrossRef]
  132. Ahmad, A.; Buang, A.; Bhat, A.H. Renewable and sustainable bioenergy production from microalgal co-cultivation with palm oil mill effluent (POME): A review. Renew. Sustain. Energy Rev. 2016, 65, 214–234. [Google Scholar] [CrossRef]
  133. Nurliyana, M.Y.; H’ng, P.S.; Rasmina, H.; Kalsom, M.S.U.; Chin, K.L.; Lee, S.H.; Lum, W.C.; Khoo, G.D. Effect of C/N ratio in methane productivity and biodegradability during facultative co-digestion of palm oil mill effluent and empty fruit bunch. Ind. Crops Prod. 2015, 76, 409–415. [Google Scholar] [CrossRef]
  134. Shahata, A.; Mohammedadel, A.; Professor, A.; Akimoto, T. Improvement of Membrane Bioreactor Operations for Color and Oil Removal from Wastewater. Ph.D. Thesis, School of Bioscience and Biotechnology, Tokyo University of Technology, Tokyo, Japan, March 2016; pp. 1–102. [Google Scholar]
  135. Ghimire, A.; Frunzo, L.; Pirozzi, F.; Trably, E.; Escudie, R.; Lens, P.N.L.; Esposito, G. A review on dark fermentative biohydrogen production from organic biomass: Process parameters and use of by-products. 2015, 144, 73–95. [Google Scholar] [CrossRef]
  136. 0-Thong, S.; Suksong, W.; Promnuan, K. ScienceDirect Two-stage thermophilic fermentation and mesophilic methanogenic process for biohythane production from palm oil mill effluent with methanogenic effluent recirculation for pH control. Int. J. Hydrogen Energy 2016, 41, 21702–21712. [Google Scholar] [CrossRef]
  137. Krishnan, S.; Singh, L.; Sakinah, M.; Thakur, S.; Wahid, Z.A.; Alkasrawi, M. Process enhancement of hydrogen and methane production from palm oil mill effluent using two-stage thermophilic and mesophilic fermentation. Int. J. Hydrogen Energy 2016, 41, 12888–12898. [Google Scholar] [CrossRef]
  138. Krishnan, S.; Singh, L.; Sakinah, M.; Thakur, S.; Wahid, Z.A.; Ghrayeb, O.A. Role of organic loading rate in bioenergy generation from palm oil mill effluent in a two-stage up-flow anaerobic sludge blanket continuous-stirred tank reactor. J. Clean. Prod. 2017, 142, 3044–3049. [Google Scholar] [CrossRef]
  139. Gobi, K.; Vadivelu, V.M. By-products of palm oil mill effluent treatment plant—A step towards sustainability. Renew. Sustain. Energy Rev. 2013, 28, 788–803. [Google Scholar] [CrossRef]
  140. Liew, W.L.; Kassim, M.A.; Muda, K.; Loh, S.K.; Affam, A.C. Conventional methods and emerging wastewater polishing technologies for palm oil mill effluent treatment: A review. J. Environ. Manag. 2014, 149, 222–235. [Google Scholar] [CrossRef] [PubMed]
  141. Neoh, C.H.; Yung, P.Y.; Noor, Z.Z.; Razak, M.H.; Aris, A.; Md Din, M.F.; Ibrahim, Z. Correlation between microbial community structure and performances of membrane bioreactor for treatment of palm oil mill effluent. Chem. Eng. J. 2017, 308, 656–663. [Google Scholar] [CrossRef]
  142. Garritano, N.; Gonc, C.; Maria, D. ScienceDirect Efficient biohydrogen production via dark fermentation from hydrolized palm oil mill effluent by non-commercial enzyme preparation. Int. J. Hydrogen Energy 2017, 42, 29166–29174. [Google Scholar] [CrossRef]
  143. Lee, J.; Lee, J.W.; Kim, Y.M.; Park, C.; Park, K.Y. Performance and fouling in pre-denitrification membrane bioreactors treating high-strength wastewater from food waste disposers. Water (Switzerland) 2017, 9, 512. [Google Scholar] [CrossRef]
  144. Huang, L.; Lee, D.J. Membrane bioreactor: A mini review on recent R&D works. Bioresour. Technol. 2015, 194, 383–388. [Google Scholar] [CrossRef] [PubMed]
  145. Ghosh, U.K.; Pradhan, N.C.; Adhikari, B. Synthesis and characterization of porous polyurethaneurea membranes for pervaporative separation of 4-nitrophenol from aqueous solution. Bull. Mater. Sci. 2006, 29, 225–231. [Google Scholar] [CrossRef]
  146. Basile, A.; Cassano, A.; Rastogi, N.K. Advances in Membrane Technologies for Water Treatment: Materials, Processes and Applications; Woodhead Publishing: Cambridge, UK, 2015; ISBN 9781782421269. [Google Scholar]
  147. Cervantes, F.J.; Pavlostathis, S.G.; Van Haandel, A.C. Advanced Biological Treatment Processes for Industrial Wastewaters: Principles and Applications; IWA Publishing: London, UK, 2006. [Google Scholar]
  148. Bharagava, R.N.; Mishra, S. Ecotoxicology and Environmental Safety Hexavalent chromium reduction potential of Cellulosimicrobium sp. isolated from common effluent treatment plant of tannery industries. Ecotoxicol. Environ. Saf. 2018, 147, 102–109. [Google Scholar] [CrossRef] [PubMed]
  149. Tajuddin, H.A.; Abdullah, L.C.; Idris, A.; Choong, T.S.Y. Effluent Quality of Anaerobic Palm Oil Mill Effluent (POME) Wastewater Using Organic Coagulant. Int. J. Sci. Res. 2015, 4, 667–677. [Google Scholar]
  150. Hariz, H.B.; Takriff, M.S. Palm oil mill effluent treatment and CO 2 sequestration by using microalgae — sustainable strategies for environmental protection. Environ. Sci. Pollut. Res. 2017, 24, 20209–20240. [Google Scholar] [CrossRef] [PubMed]
  151. Pretel, R.; Robles, A.; Ruano, M.V.; Seco, A.; Ferrer, J. Economic and environmental sustainability of submerged anaerobic MBR-based (AnMBR-based) technology as compared to aerobic-based technologies for moderate-/high-loaded urban wastewater treatment. J. Environ. Manag. 2016, 166, 45–54. [Google Scholar] [CrossRef] [PubMed]
  152. Tan, Y.H.; Goh, P.S.; Ismail, A.F.; Ng, B.C.; Lai, G.S. Decolourization of aerobically treated palm oil mill effluent (AT-POME) using polyvinylidene fluoride (PVDF) ultrafiltration membrane incorporated with coupled zinc-iron oxide nanoparticles. Chem. Eng. J. 2017, 308, 359–369. [Google Scholar] [CrossRef]
  153. Subramaniam, M.N.; Goh, P.S.; Lau, W.J.; Ng, B.C.; Ismail, A.F. AT-POME colour removal through photocatalytic submerged filtration using antifouling PVDF-TNT nanocomposite membrane. Sep. Purif. Technol. 2018, 191, 266–275. [Google Scholar] [CrossRef]
  154. Mohd Azoddein, A.A.; Haris, H.; Mohd Azli, F.A. Treatment of Palm Oil Mill Effluent (Pome) Using Membrane Bioreactor. Malays. J. Anal. Sci. 2015, 19, 463–471. [Google Scholar]
  155. Sarma, S.J.; Tay, J.H.; Chu, A. Finding Knowledge Gaps in Aerobic Granulation Technology. Trends Biotechnol. 2017, 35, 66–78. [Google Scholar] [CrossRef] [PubMed]
  156. Zhang, Q.; Hu, J.; Lee, D. Bioresource Technology Aerobic granular processes: Current research trends. Bioresour. Technol. 2016, 210, 74–80. [Google Scholar] [CrossRef] [PubMed]
  157. Iorhemen, O.T.; Hamza, R.A.; Tay, J.H. Membrane bioreactor (Mbr) technology for wastewater treatment and reclamation: Membrane fouling. Membranes (Basel) 2016, 6, 33. [Google Scholar] [CrossRef] [PubMed]
  158. Park, C.; Park, J.; Han, G. Toxic/Hazardous Substances and Environmental Engineering Control of membrane fouling with the addition of a nanoporous zeolite membrane fouling reducer to the submerged hollow fiber membrane bioreactor. J. Environ. Sci. Health 2016, 51, 1024–1033. [Google Scholar] [CrossRef] [PubMed]
  159. Niu, T.; Zhou, Z.; Shen, X.; Qiao, W.; Jiang, L.; Pan, W. Effects of dissolved oxygen on performance and microbial community structure in a micro-aerobic hydrolysis sludge in situ reduction process. Water Res. 2016, 90, 369–377. [Google Scholar] [CrossRef] [PubMed]
  160. Chan, Y.J.; Chong, M.F.; Law, C.L. Performance and kinetic evaluation of an integrated anaerobic–aerobic bioreactor in the treatment of palm oil mill effluent. Environ. Technol. 2017, 38, 1005–1021. [Google Scholar] [CrossRef] [PubMed]
  161. Alattabi, A.W.; Harris, C.B.; Alkhaddar, M.; Alzeyadi, A.T. Journal of Water Process Engineering An investigation into the e ff ect of MLSS on the effluent quality and sludge settleability in an aerobic-anoxic sequencing batch reactor (AASBR). J. Water Process Eng. 2017. [Google Scholar] [CrossRef]
  162. Díaz, O.; González, E.; Vera, L.; Macías-hernández, J.J.; Rodríguez-sevilla, J. Fouling analysis and mitigation in a tertiary MBR operated under restricted aeration. J. Membr. Sci. 2017, 525, 368–377. [Google Scholar] [CrossRef]
  163. Yang, L.; Liu, L.; Wang, Z. Preparation of PVDF/GO[sbnd]SiO2 hybrid microfiltration membrane towards enhanced perm-selectivity and anti-fouling property. J. Taiwan Inst. Chem. Eng. 2017, 78, 500–509. [Google Scholar] [CrossRef]
  164. De Temmerman, L.; Maere, T.; Temmink, H.; Zwijnenburg, A.; Nopens, I. The effect of fine bubble aeration intensity onmembrane bioreactor sludge characteristics and fouling. Water Res. 2015, 76, 99–109. [Google Scholar] [CrossRef] [PubMed]
  165. Wang, Z.; Wu, Z. A review of membrane fouling in MBRs: Characteristics and role of sludge cake formed on membrane surfaces. Sep. Sci. Technol. 2009, 44, 3571–3596. [Google Scholar] [CrossRef]
  166. Maqbool, T.; Khan, S.J.; Waheed, H.; Lee, C.H.; Hashmi, I.; Iqbal, H. Membrane biofouling retardation and improved sludge characteristics using quorum quenching bacteria in submerged membrane bioreactor. J. Membr. Sci. 2015, 483, 75–83. [Google Scholar] [CrossRef]
  167. Rodríguez, F.A.; Reboleiro-Rivas, P.; Osorio, F.; Martínez-Toledo, M.V.; Hontoria, E.; Poyatos, J.M. Influence of mixed liquid suspended solids and hydraulic retention time on oxygen transfer efficiency and viscosity in a submerged membrane bioreactor using pure oxygen to supply aerobic conditions. Biochem. Eng. J. 2012, 60, 135–141. [Google Scholar] [CrossRef]
  168. Farhan, A.; Udaiyappan, M.; Abu, H.; Sobri, M. Journal of Water Process Engineering A review of the potentials, challenges and current status of microalgae biomass applications in industrial wastewater treatment. J. Water Process Eng. 2017, 20, 8–21. [Google Scholar] [CrossRef]
  169. Hamza, R.A.; Iorhemen, O.T.; Tay, J.H. Journal of Water Process Engineering Advances in biological systems for the treatment of high-strength wastewater. J. Water Process Eng. 2016, 10, 128–142. [Google Scholar] [CrossRef]
  170. Manai, I.; Miladi, B.; Mselmi, A.; El Hamdi, M. Improvement of activated sludge resistance to shock loading by fungal enzyme addition during textile wastewater treatment. Environ. Technol. 2016, 38, 880–890. [Google Scholar] [CrossRef] [PubMed]
  171. Ansari, A.J.; Hai, F.I.; Guo, W.; Ngo, H.H.; Price, W.E.; Nghiem, L.D. Science of the Total Environment Factors governing the pre-concentration of wastewater using forward osmosis for subsequent resource recovery. Sci. Total Environ. 2016, 566–567, 559–566. [Google Scholar] [CrossRef] [PubMed]
  172. Hu, K. Contribution to modeling of treatment and reuse of industrial wastewater. Ph.D. Thesis, Technische Universität Berlin, Berlin, Germany, 2017. [Google Scholar]
  173. Arévalo, J.; Moreno, B.; Pérez, J.; Gómez, M.A. Applicability of the Sludge Biotic Index (SBI) for MBR activated sludge control. J. Hazard. Mater. 2009, 167, 784–789. [Google Scholar] [CrossRef] [PubMed]
  174. Rezvani, F.; Mehrnia, M.R.; Poostchi, A.A. Optimal operating strategies of SFDM formation for MBR application. Sep. Purif. Technol. 2014, 124, 124–133. [Google Scholar] [CrossRef]
  175. Ho, J.; Smith, S.; Patamasank, J.; Tontcheva, P.; Kim, G.D. Development of Alternative Energy Saving MBR Using Reciprocating Vibration in Place of Membrane Air Scouring. In Proceedings of the Water Environment Federation, WEFTEC 2013: Session 92 through Session 101; Water Environment Federation: Alexandria, VA, USA, 2013; pp. 6679–6688. [Google Scholar]
  176. Kumar, A.; Jena, H.M. High surface area microporous activated carbons prepared from Fox nut (Euryale ferox) shell by zinc chloride activation. Appl. Surf. Sci. 2015, 356, 753–761. [Google Scholar] [CrossRef]
  177. Yahya, M.A.; Al-Qodah, Z.; Ngah, C.W.Z. Agricultural bio-waste materials as potential sustainable precursors used for activated carbon production: A review. Renew. Sustain. Energy Rev. 2015, 46, 218–235. [Google Scholar] [CrossRef]
  178. Alkhatib, M.F.; Mamun, A.A.; Akbar, I. Application of response surface methodology (RSM) for optimization of color removal from POME by granular activated carbon. Int. J. Environ. Sci. Technol. 2015, 12, 1295–1302. [Google Scholar] [CrossRef]
  179. Sia, Y.Y.; Tan, I.A.W.; Abdullah, M.O. Adsorption of colour, TSS and COD from palm oil mill effluent (POME) using acid-washed coconut shell activated carbon: Kinetic and mechanism studies. MATEC Web Conf. 2017, 87. [Google Scholar] [CrossRef]
  180. Guo, W.; Vigneswaran, S.; Ngo, H.H.; Xing, W.; Goteti, P. Comparison of the performance of submerged membrane bioreactor (SMBR) and submerged membrane adsorption bioreactor (SMABR). Bioresour. Technol. 2008, 99, 1012–1017. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  181. Wahi, R.; Chuah Abdullah, L.; Nourouzi Mobarekeh, M.; Ngaini, Z.; Choong Shean Yaw, T. Utilization of esterified sago bark fibre waste for removal of oil from palm oil mill effluent. J. Environ. Chem. Eng. 2017, 5, 170–177. [Google Scholar] [CrossRef]
  182. Adeleye, A.S.; Conway, J.R.; Garner, K.; Huang, Y.; Keller, A.A. Engineered nanomaterials for water treatment and remediation: Costs, benefits, and applicability. Chem. Eng. J. 2015. [Google Scholar] [CrossRef]
  183. Kaman, S.; Tan, I.; Lim, L. Palm oil mill effluent treatment using coconut shell–based activated carbon: Adsorption equilibrium and isotherm. MATEC Web 2017, 87, 3009. [Google Scholar] [CrossRef]
  184. Pellegrin, M.; Greiner, A.D.; Aguinaldo, J.; Diamond, J.; Gluck, S.; Burbano, M.S.; Arabi, S.; Wert, J.; Mccandless, R.; Padhye, L.P. Membrane Processes. Water Environ. Res. 2012, 84, 1114–1216. [Google Scholar] [CrossRef]
  185. Tijing, L.D.; Woo, Y.C.; Choi, J.S.; Lee, S.; Kim, S.H.; Shon, H.K. Fouling and its control in membrane distillation—A review. J. Membr. Sci. 2015, 475, 215–244. [Google Scholar] [CrossRef] [Green Version]
  186. Shahkaramipour, N.; Tran, T.N.; Ramanan, S.; Lin, H. Membranes with surface-enhanced antifouling properties for water purification. Membranes (Basel) 2017, 7, 13. [Google Scholar] [CrossRef] [PubMed]
  187. Figoli, A.; Cassano, A.; Basile, A. Membrane Technologies for Biorefining; Elsevier: New York, NY, USA, 2016; ISBN 9780857095213. [Google Scholar]
  188. Qu, X.; Alvarez, P.J.J.; Li, Q. Applications of nanotechnology in water and wastewater treatment. Water Res. 2013, 47, 3931–3946. [Google Scholar] [CrossRef] [PubMed]
  189. Budiman, P.M.; Wu, T.Y.; Ramanan, R.N.; Xiao, J.; Hay, W. Treatment and Reuse of E ffl uents from Palm Oil, Pulp, and Paper Mills as a Combined Substrate by Using Purple Nonsulfur Bacteria. Ind. Eng. Chem. Res. 2014, 53, 14921–14931. [Google Scholar] [CrossRef]
  190. Appels, L.; Baeyens, J.; Degrève, J.; Dewil, R. Principles and potential of the anaerobic digestion of waste-activated sludge. Prog. Energy Combust. Sci. 2008, 34, 755–781. [Google Scholar] [CrossRef]
  191. Dereli, R.K.; Ersahin, M.E.; Ozgun, H.; Ozturk, I.; Jeison, D.; van der Zee, F.; van Lier, J.B. Potentials of anaerobic membrane bioreactors to overcome treatment limitations induced by industrial wastewaters. Bioresour. Technol. 2012, 122, 160–170. [Google Scholar] [CrossRef] [PubMed]
  192. Evren, M.; Tao, Y.; Ozgun, H.; Gimenez, J.B.; Spanjers, H.; Lier, B. Van Impact of anaerobic dynamic membrane bioreactor con fi guration on treatment and fi lterability performance. J. Membr. Sci. 2017, 526, 387–394. [Google Scholar] [CrossRef]
  193. Mamimin, C.; Jehlee, A.; Saelor, S.; Prasertsan, P.; O.-Thong, S. Thermophilic hydrogen production from co-fermentation of palm oil mill effluent and decanter cake by Thermoanaerobacterium thermosaccharolyticum PSU-2. Int. J. Hydrogen Energy 2016, 41, 21692–21701. [Google Scholar] [CrossRef]
  194. Ravindra, P. Advances in Bioprocess Technology; Springer: Berlin, Germany, 2015; ISBN 9783319179155. [Google Scholar]
  195. Tan, D.T.; Chin, S.K.; Poh, P.E.; Lee, Y.H. Preservation of thermophilic mixed culture for anaerobic palm oil mill effluent treatment by convective drying methods. Int. J. Environ. Sci. Technol. 2017. [Google Scholar] [CrossRef]
  196. Abdurahman, N.H.; Rosli, Y.M.; Azhari, N.H. Development of a membrane anaerobic system (MAS) for palm oil mill effluent (POME) treatment. Desalination 2011, 266, 208–212. [Google Scholar] [CrossRef]
  197. Abdurahman, N.H.; Azhari, N.H. An integrated UMAS for POME treatment. J. Water Reuse Desalin. 2016. [Google Scholar] [CrossRef]
  198. Ahmad, A.; Ghufran, R. Evaluation of the bio-kinetics of cement kiln dust in an upflow anaerobic sludge blanket reactor for treatment of palm oil mill effluent as a function of hydraulic retention time. Sep. Purif. Technol. 2014, 133, 129–137. [Google Scholar] [CrossRef]
  199. Zhang, Q.; Singh, S.; Stuckey, D.C. Bioresource Technology Fouling reduction using adsorbents/flocculants in a submerged anaerobic membrane bioreactor. Bioresour. Technol. 2017, 239, 226–235. [Google Scholar] [CrossRef] [PubMed]
  200. Zheng, Y.; Zhang, W.; Tang, B.; Bin, L.; Ding, J.; Zheng, Y.; Zhang, Z. Membrane fouling mechanism of biofilm-membrane bioreactor (BF-MBR): Pore blocking model and membrane cleaning. Bioresour. Technol. 2017, 250, 398–405. [Google Scholar] [CrossRef] [PubMed]
  201. Annop, S.; Sridang, P.; Puetpaiboon, U.; Grasmick, A. Influence of relaxation frequency on membrane fouling control in submerged anaerobic membrane bioreactor (SAnMBR). Desalin. Water Treat. 2014, 52, 4102–4110. [Google Scholar] [CrossRef]
  202. Said, M.; Wahab Mohammad, A.; Tusirin, M.; Nor, M. Investigation of Three Pre-treatment Methods Prior to Nanofiltration Membrane for Palm Oil Mill Effluent Treatment. Sains Malays. 2015, 44, 421–427. [Google Scholar] [CrossRef]
  203. Dimitriou, E.; Boutikos, P.; Sh, E.; Koziel, S. Theoretical performance prediction of a reverse osmosis desalination membrane element under variable operating conditions. Desalination 2017, 419, 70–78. [Google Scholar] [CrossRef]
  204. Annop, S.; Sridang, P.; Puetpaiboon, U.; Grasmick, A. Effect of solids retention time on membrane fouling intensity in two-stage submerged anaerobic membrane bioreactors treating palm oil mill effluent. Environ. Technol. 2014, 35, 2634–2642. [Google Scholar] [CrossRef] [PubMed]
  205. Neoh, C.H.; Noor, Z.Z.; Mutamim, N.S.A.; Lim, C.K. Green technology in wastewater treatment technologies: Integration of membrane bioreactor with various wastewater treatment systems. Chem. Eng. J. 2016, 283, 582–594. [Google Scholar] [CrossRef]
  206. Ziegler, A.S.; McIlroy, S.J.; Larsen, P.; Albertsen, M.; Hansen, A.A.; Heinen, N.; Nielsen, P.H. Dynamics of the fouling layer microbial community in a membrane bioreactor. PLoS ONE 2016, 11, 1–14. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Damayanti, A.; Ujang, Z.; Salim, M.R.; Olsson, G. The effect of mixed liquor suspended solids (MLSS) on biofouling in a hybrid membrane bioreactor for the treatment of high concentration organic wastewater. Water Sci. Technol. 2011, 63, 1701–1706. [Google Scholar] [CrossRef] [PubMed]
  208. Ahmad, Z.; Ujang, Z.; Abdul Latiff, A.A.; Ahmat Nor, N.I. Effect on Membrane Fouling and Cake Resistance in a Hybrid Membrane Bioreactor for Palm Oil Mill Effluent Treatment. Int. Conf. Environ. 2008, 2008, 1–12. [Google Scholar]
  209. Ma, J.; Wang, Z.; Yang, Y.; Mei, X.; Wu, Z. Correlating microbial community structure and composition with aeration intensity in submerged membrane bioreactors by 454 high-throughput pyrosequencing. Water Res. 2013, 47, 859–869. [Google Scholar] [CrossRef] [PubMed]
  210. Sun, J.; Xiao, K.; Mo, Y.; Liang, P.; Shen, Y.; Zhu, N.; Huang, X. Seasonal characteristics of supernatant organics and its effect on membrane fouling in a full-scale membrane bioreactor. J. Membr. Sci. 2014, 453, 168–174. [Google Scholar] [CrossRef]
  211. Khan, S.J.; Visvanathan, C.; Jegatheesan, V. Bioresource Technology Influence of biofilm carriers on membrane fouling propensity in moving biofilm membrane bioreactor. Bioresour. Technol. 2012, 113, 161–164. [Google Scholar] [CrossRef] [PubMed]
  212. Ma, C.; Yu, S.; Shi, W.; Heijman, S.G.J.; Rietveld, L.C. Bioresource Technology Effect of different temperatures on performance and membrane fouling in high concentration PAC–MBR system treating micro-polluted surface water. Bioresour. Technol. 2013, 141, 19–24. [Google Scholar] [CrossRef] [PubMed]
  213. Chen, W.; Gao, X.; Xu, H.; Cai, Y.; Cui, J. Chemosphere influence of extracellular polymeric substances (EPS) treated by combined ultrasound pretreatment and chemical re-flocculation on water treatment sludge settling performance. Chemosphere 2017, 170, 196–206. [Google Scholar] [CrossRef] [PubMed]
  214. Tee, P.F.; Abdullah, M.O.; Tan, I.A.W.; Amin, M.A.M.; Nolasco-Hipolito, C.; Bujang, K. Effects of temperature on wastewater treatment in an affordable microbial fuel cell-adsorption hybrid system. J. Environ. Chem. Eng. 2017, 5, 178–188. [Google Scholar] [CrossRef] [Green Version]
  215. Ma, Z.; Wen, X.; Zhao, F.; Xia, Y.; Huang, X.; Waite, D.; Guan, J. Effect of temperature variation on membrane fouling and microbial community structure in membrane bioreactor. Bioresour. Technol. 2013, 133, 462–468. [Google Scholar] [CrossRef] [PubMed]
  216. Gao, W.J.; Leung, K.T.; Qin, W.S.; Liao, B.Q. Bioresource Technology Effects of temperature and temperature shock on the performance and microbial community structure of a submerged anaerobic membrane bioreactor. Bioresour. Technol. 2011, 102, 8733–8740. [Google Scholar] [CrossRef] [PubMed]
  217. Cologgi, D.; Kasiri, S.; Hofstetter, S.; Donoso-bravo, A.; Ulrich, A. Anaerobic Processes. Water Environ. Res. 213, 85, 1176–1231. [Google Scholar] [CrossRef]
  218. Shafie, N.F.A.; Mansor, U.Q.A.; Yahya, A.; Som, A.M.; Nour, A.H.; Hassan, Z.; Yunus, R.M. Performance of ultrasonic-assisted membrane anaerobic system (UMAS) for membrane fouling control in palm oil mill effluent (POME) treatment. Adv. Sci. Lett. 2017, 23, 3903–3906. [Google Scholar] [CrossRef]
  219. Saifuddin, N.; Dinara, S.; Unit, C.; Nasional, U.T. Pretreatment of Palm Oil Mill Effluent (POME) Using Magnetic Chitosan. J. Chem. 2011, 8 (Suppl. S1), S67–S78. [Google Scholar] [CrossRef]
  220. Tyagi, V.K.; Lo, S.; Appels, L.; Dewil, R.A.F. Ultrasonic Treatment of Waste Sludge: A Review on Mechanisms and Applications. Crit. Rev. Environ. Sci. Technol. 2014, 44, 1220–1288. [Google Scholar] [CrossRef]
  221. Carrère, H.; Dumas, C.; Battimelli, A.; Batstone, D.J.; Delgenès, J.P.; Steyer, J.P.; Ferrer, I. Pretreatment methods to improve sludge anaerobic degradability: A review. 2010, 183, 1–15. [Google Scholar] [CrossRef]
  222. Abdurahman, N.H.; Azhari, N.H.; Rosli, Y.M. Ultrasonic Membrane Anaerobic System (UMAS) for Palm Oil Mill Effluent (POME) Treatment. Int. Perspect. Water Qual. Manag. Pollut. Control 2013, 107–121. [Google Scholar] [CrossRef]
  223. Leaño, E.P.; Anceno, A.J.; Babel, S. Ultrasonic pretreatment of palm oil mill effluent: Impact on biohydrogen production, bioelectricity generation, and underlying microbial communities. Int. J. Hydrogen Energy 2012, 37, 12241–12249. [Google Scholar] [CrossRef]
  224. Taha, M.R.; Ibrahim, A.H. Characterization of nano zero-valent iron (nZVI) and its application in sono-Fenton process to remove COD in palm oil mill effluent. J. Environ. Chem. Eng. 2014, 2, 1–8. [Google Scholar] [CrossRef]
  225. Manickam, S.; Abidin, Z.; Parthasarathy, S. Role of H2O2 in the Fluctuating Patterns of COD (Chemical Oxygen Demand) during the treatment of Palm Oil Mill Effluent (POME) Using Pilot Scale Triple Frequency Ultrasound Cavitation Reactor. Ultrason. Sonochem. 2014, 21, 1519–1526. [Google Scholar] [CrossRef] [PubMed]
  226. Parthasarathy, S.; Mohammed, R.R.; Fong, C.M.; Gomes, R.L.; Manickam, S. A novel hybrid approach of activated carbon and ultrasound cavitation for the intensification of palm oil mill effluent (POME) polishing. J. Clean. Prod. 2016, 112, 1218–1226. [Google Scholar] [CrossRef] [Green Version]
  227. Xu, M.; Wen, X.; Yu, Z.; Li, Y.; Huang, X. Bioresource Technology A hybrid anaerobic membrane bioreactor coupled with online ultrasonic equipment for digestion of waste activated sludge. Bioresour. Technol. 2011, 102, 5617–5625. [Google Scholar] [CrossRef] [PubMed]
  228. Abeynayaka, A.; Visvanathan, C. Performance comparison of mesophilic and thermophilic aerobic sidestream membrane bioreactors treating high strength wastewater. Bioresour. Technol. 2011, 102, 5345–5352. [Google Scholar] [CrossRef] [PubMed]
  229. Stuckey, D.C. Bioresource Technology Recent developments in anaerobic membrane reactors. Bioresour. Technol. 2012, 122, 137–148. [Google Scholar] [CrossRef] [PubMed]
  230. Visvanathan, C.; Abeynayaka, A. Developments and future potentials of anaerobic membrane bioreactors (AnMBRs). Membr. Water Treat. 2012, 3, 1–23. [Google Scholar] [CrossRef]
  231. Christensen, M.L.; Keiding, K.; Halkj, P. ScienceDirect Dewatering in biological wastewater treatment: A review. Water Res. 2015, 82, 14–24. [Google Scholar] [CrossRef] [PubMed]
  232. Massara, T.M.; Tarik, O.; Onur, K.; Senba, S. A Mini Review of the Techno-environmental Sustainability of Biological Processes for the Treatment of High Organic Content Industrial Wastewater Streams. Waste Biomass Valoriz. 2017, 8, 1665–1678. [Google Scholar] [CrossRef]
  233. Lin, H.J.; Xie, K.; Mahendran, B.; Bagley, D.M.; Leung, K.T.; Liss, S.N.; Liao, B.Q. Sludge properties and their effects on membrane fouling in submerged anaerobic membrane bioreactors (SAnMBRs). Water Res. 2009, 43, 3827–3837. [Google Scholar] [CrossRef] [PubMed]
  234. Arifin, H.; Choong, T.S.Y.; Rong, C.K.; Ahmadun, F.A.R.; Abdullah, L.C. Forward Osmosis: Temperature effects by using pome as feed solution. ASEAN J. Chem. Eng. 2015, 15, 31–40. [Google Scholar]
  235. Martinez-Sosa, D.; Helmreich, B.; Horn, H. Anaerobic submerged membrane bioreactor (AnSMBR) treating low-strength wastewater under psychrophilic temperature conditions. Process Biochem. 2012, 47, 792–798. [Google Scholar] [CrossRef]
  236. Choorit, W.; Wisarnwan, P. Effect of temperature on the anaerobic digestion of palm oil mill effluent. Electron. J. Biotechnol. 2007, 10, 376–385. [Google Scholar] [CrossRef]
  237. Shao, L.; Wang, Z.X.; Zhang, Y.L.; Jiang, Z.X.; Liu, Y.Y. A facile strategy to enhance PVDF ultrafiltration membrane performance via self-polymerized polydopamine followed by hydrolysis of ammonium fluotitanate. J. Membr. Sci. 2014, 461, 10–21. [Google Scholar] [CrossRef]
  238. Najib, M.Z.M.; Salmiati; Ujang, Z.; Salim, M.R.; Ibrahim, Z. Developed microbial granules containing photosynthetic pigments for carbon dioxide reduction in palm oil mill effluent. In International Biodeterioration and Biodegradation; Elsevier Ltd.: New York, NY, USA, 2017; Volume 116, pp. 163–170. [Google Scholar]
  239. Ao, L.; Liu, W.; Zhao, L.; Wang, X. Membrane fouling in ultrafiltration of natural water after pretreatment to different extents. J. Environ. Sci. 2016, 43, 234–243. [Google Scholar] [CrossRef] [PubMed]
  240. Jalani, N.F.; Aziz, A.A.; Wahab, N.A.; Hasamudin, W.; Hassan, W.; Zainal, N.H. Application of Palm Kernel Shell Activated Carbon for the Removal of Pollutant and Color in Palm Oil Mill Effluent Treatment. J. Earth Environ. Health Sci. 2016, 2, 15–20. [Google Scholar] [CrossRef]
  241. Gold, K.; Slay, B.; Knackstedt, M.; Gaharwar, A.K. Antimicrobial Activity of Metal and Metal-Oxide Based Nanoparticles. Adv. Ther. 2018, 1–15. [Google Scholar] [CrossRef]
  242. Di Martino, P. Extracellular polymeric substances, a key element in understanding biofilm phenotype. AIMS Microbiol. 2018, 4, 274–288. [Google Scholar] [CrossRef]
  243. Seabra, A.; Durán, N. Nanotoxicology of Metal Oxide Nanoparticles. Metals (Basel) 2015, 5, 934–975. [Google Scholar] [CrossRef] [Green Version]
  244. Isma, M.I.A.; Idris, A.; Omar, R.; Razreena, A.R.P. Effects of SRT and HRT on Treatment Performance of MBR and Membrane Fouling. Int. J. Chem. Mol. Nucl. Mater. Metall. Eng. 2014, 8, 488–492. [Google Scholar]
  245. Iorhemen, O.T.; Hamza, R.A.; Tay, J.H. Membrane fouling control in membrane bioreactors (MBRs) using granular materials. Bioresour. Technol. 2017, 240, 9–24. [Google Scholar] [CrossRef] [PubMed]
  246. Ng, C.A.; Wong, L.Y.; Chai, H.Y.; Bashir, M.J.K.; Ho, C.-D.; Nisar, H.; Lo, P.K. Investigation on the performance of hybrid anaerobic membrane bioreactors for fouling control and biogas production in palm oil mill effluent treatment. Water Sci. Technol. 2017, 76, 1389–1398. [Google Scholar] [CrossRef] [PubMed]
  247. Mai, D.T.; Kunacheva, C.; Stuckey, D.C. Post-treatment of anaerobic membrane bioreactor (AnMBR) effluent using activated carbon. Bioresour. Technol. 2018, 266, 75–81. [Google Scholar] [CrossRef] [PubMed]
  248. Jiang, T.; Kennedy, M.D.; van der Meer, W.G.J.; Vanrolleghem, P.A.; Schippers, J.C. The role of blocking and cake filtration in MBR fouling. Desalination 2003, 157, 335–343. [Google Scholar] [CrossRef]
  249. Ahmad, A.L.; Ismail, S.; Bhatia, S. Water recycling from palm oil mill effluent (POME) using membrane technology. Desalination 2003, 157, 87–95. [Google Scholar] [CrossRef]
  250. Loh, S.K.; Ngatiman, M.; Lim, W.S.; Choo, Y.M. A Zero Discharge Treatment System Of Palm Oil Mill Effluent. Malays. Palm Oil Board Inf. Ser. 2014, 657, 6–9. [Google Scholar]
  251. Huan, L.; Yiying, J.; Mahar, R.B.; Zhiyu, W.; Yongfeng, N. Effects of ultrasonic disintegration on sludge microbial activity and dewaterability. J. Hazard. Mater. 2009, 161, 1421–1426. [Google Scholar] [CrossRef] [PubMed]
  252. Mirbagheri, S.A.; Bagheri, M.; Boudaghpour, S.; Ehteshami, M.; Bagheri, Z. Performance evaluation and modeling of a submerged membrane bioreactor treating combined municipal and industrial wastewater using radial basis function artificial neural networks. J. Environ. Heal. Sci. Eng. 2015, 13, 1–15. [Google Scholar] [CrossRef] [PubMed]
  253. Kimura, K.; Yamato, N.; Yamamura, H.; Watanabe, Y. Membrane fouling in pilot-scale membrane bioreactors (MBRs) treating municipal wastewater. Environ. Sci. Technol. 2005, 39, 6293–6299. [Google Scholar] [CrossRef] [PubMed]
  254. Mitri, S.; Clarke, E.; Foster, K.R. Resource limitation drives spatial organization in microbial groups. ISME J. 2016, 10, 1471–1482. [Google Scholar] [CrossRef] [PubMed]
  255. Said, M.; Abu Hasan, H.; Mohd Nor, M.T.; Mohammad, A.W. Removal of COD, TSS and colour from palm oil mill effluent (POME) using montmorillonite. Desalin. Water Treat. 2016, 57, 10490–10497. [Google Scholar] [CrossRef]
  256. Soleimaninanadegani, M.; Manshad, S. Enhancement of Biodegradation of Palm Oil Mill Effluents by Local Isolated Microorganisms. Int. Sch. Res. Not. 2014, 2014. [Google Scholar] [CrossRef] [PubMed]
  257. Vashi, H.; Iorhemen, O.T.; Tay, J.H. Aerobic granulation: A recent development on the biological treatment of pulp and paper wastewater. Environ. Technol. Innov. 2018, 9, 265–274. [Google Scholar] [CrossRef]
  258. Schulz, M.; Soltani, A.; Zheng, X.; Ernst, M. Effect of inorganic colloidal water constituents on combined low-pressure membrane fouling with natural organic matter (NOM). J. Membr. Sci. 2016, 507, 154–164. [Google Scholar] [CrossRef]
  259. Mei, X.; Wang, Z.; Zheng, X.; Huang, F.; Ma, J.; Tang, J.; Wu, Z. Soluble microbial products in membrane bioreactors in the presence of ZnO nanoparticles. J. Membr. Sci. 2014, 451, 169–176. [Google Scholar] [CrossRef]
  260. Peleato, N.M.; Legge, R.L.; Andrews, R.C. Characterization of UF foulants and fouling mechanisms when applying low in-line coagulant pre-treatment. Water Res. 2017, 126, 1–11. [Google Scholar] [CrossRef] [PubMed]
  261. Samantaray, P.K.; Madras, G.; Bose, S. PVDF/PBSA membranes with strongly coupled phosphonium derivatives and graphene oxide on the surface towards antibacterial and antifouling activities. J. Membr. Sci. 2018, 548, 203–214. [Google Scholar] [CrossRef]
  262. Hu, N.; Xiao, T.; Cai, X.; Ding, L.; Fu, Y.; Yang, X. Preparation and characterization of hydrophilically modified PVDF membranes by a novel nonsolvent thermally induced phase separation method. Membranes (Basel) 2016, 6, 47. [Google Scholar] [CrossRef] [PubMed]
  263. Tran, N.H.; Ngo, H.H.; Urase, T.; Gin, K.Y.H. A critical review on characterization strategies of organic matter for wastewater and water treatment processes. Bioresour. Technol. 2015, 193. [Google Scholar] [CrossRef] [PubMed]
  264. Mancebo, U.; Hettiaratchi, J.P.A. Rapid assessment of methanotrophic capacity of compost-based materials considering the effects of air-filled porosity, water content and dissolved organic carbon. Bioresour. Technol. 2015, 177, 125–133. [Google Scholar] [CrossRef] [PubMed]
  265. Gkotsis, P.; Banti, D.; Peleka, E.; Zouboulis, A.; Samaras, P. Fouling Issues in Membrane Bioreactors (MBRs) for Wastewater Treatment: Major Mechanisms, Prevention and Control Strategies. Processes 2014, 2, 795–866. [Google Scholar] [CrossRef]
  266. Ding, Y.; Tian, Y.; Li, Z.; Zuo, W.; Zhang, J. A comprehensive study into fouling properties of extracellular polymeric substance (EPS) extracted from bulk sludge and cake sludge in a mesophilic anaerobic membrane bioreactor. Bioresour. Technol. 2015, 192, 105–114. [Google Scholar] [CrossRef] [PubMed]
  267. Sioutopoulos, D.C.; Karabelas, A.J. The effect of permeation flux on the specific resistance of polysaccharide fouling layers developing during dead-end ultrafiltration. J. Membr. Sci. 2015, 473, 292–301. [Google Scholar] [CrossRef]
  268. Li, C.; Yang, Y.; Ding, S.; Hou, L.A. Dynamics of biofouling development on the conditioned membrane and its relationship with membrane performance. J. Membr. Sci. 2016, 514, 264–273. [Google Scholar] [CrossRef]
  269. Meng, L.; Xi, J.; Yeung, M. Degradation of extracellular polymeric substances (EPS) extracted from activated sludge by low-concentration ozonation. Chemosphere 2016, 147, 248–255. [Google Scholar] [CrossRef] [PubMed]
  270. Musa, M.; Idrus, S.; Che Man, H.; Nik Daud, N. Wastewater Treatment and Biogas Recovery Using Anaerobic Membrane Bioreactors (AnMBRs): Strategies and Achievements. Energies 2018, 11, 1675. [Google Scholar] [CrossRef]
  271. Neoh, C.H.; Lam, C.Y.; Lim, C.K.; Yahya, A.; Ibrahim, Z. Decolorization of palm oil mill effluent using growing cultures of Curvularia clavata. Environ. Sci. Pollut. Res. 2014, 21, 4397–4408. [Google Scholar] [CrossRef] [PubMed]
  272. Tan, S.P.; Kong, H.F.; Bashir, M.J.K.; Lo, P.K.; Ho, C.D.; Ng, C.A. Treatment of palm oil mill effluent using combination system of microbial fuel cell and anaerobic membrane bioreactor. Bioresour. Technol. 2017, 245, 916–924. [Google Scholar] [CrossRef] [PubMed]
  273. Herzberg, M.; Elimelech, M. Biofouling of reverse osmosis membranes: Role of biofilm-enhanced osmotic pressure. J. Membr. Sci. 2007, 295, 11–20. [Google Scholar] [CrossRef]
  274. Al Ashhab, A.; Sweity, A.; Bayramoglu, B.; Herzberg, M.; Gillor, O. Biofouling of reverse osmosis membranes: effects of cleaning on biofilm microbial communities, membrane performance, and adherence of extracellular polymeric substances. Biofouling 2017, 33, 397–409. [Google Scholar] [CrossRef] [PubMed]
  275. Derlon, N.; Grütter, A.; Brandenberger, F.; Sutter, A.; Kuhlicke, U.; Neu, T.R.; Morgenroth, E. The composition and compression of biofilms developed on ultrafiltration membranes determine hydraulic biofilm resistance. Water Res. 2016, 102, 63–72. [Google Scholar] [CrossRef] [PubMed]
  276. Bucs, S.; Farhat, N.; Kruithof, J.C.; Picioreanu, C.; van Loosdrecht, M.C.M.; Vrouwenvelder, J.S. Review on strategies for biofouling mitigation in spiral wound membrane systems. Desalination 2018, 434, 189–197. [Google Scholar] [CrossRef]
  277. Kim, S.H.; Kwak, S.; Sohn, B.; Park, T.H. Design of TiO2 nanoparticle self-assembled aromatic polyamide thin- lm-composite (TFC) membrane as an approach to solve biofouling problem. J. Membr. Sci. 2003, 211, 157–165. [Google Scholar] [CrossRef]
  278. Ayyavoo, J.; Nguyen, T.P.N.; Jun, B.M.; Kim, I.C.; Kwon, Y.N. Protection of polymeric membranes with antifouling surfacing via surface modifications. Colloids Surfaces A Physicochem. Eng. Asp. 2016, 506, 190–201. [Google Scholar] [CrossRef]
  279. Zhu, J.; Hou, J.; Zhang, Y.; Tian, M.; He, T.; Liu, J.; Chen, V. Polymeric antimicrobial membranes enabled by nanomaterials for water treatment. J. Membr. Sci. 2018, 550, 173–197. [Google Scholar] [CrossRef]
  280. Giwa, A.; Dindi, A.; Kujawa, J. Membrane bioreactors and electrochemical processes for treatment of wastewaters containing heavy metal ions, organics, micropollutants and dyes: Recent developments. J. Hazard. Mater. 2018. [Google Scholar] [CrossRef] [PubMed]
  281. Zhang, W.; Ding, L.; Luo, J.; Jaffrin, M.Y.; Tang, B. Membrane fouling in photocatalytic membrane reactors (PMRs) for water and wastewater treatment: A critical review. Chem. Eng. J. 2016, 302, 446–458. [Google Scholar] [CrossRef]
  282. Zhu, Y.; Wang, D.; Jiang, L.; Jin, J. Recent progress in developing advanced membranes for emulsified oil/water separation. NPG Asia Mater. 2014, 6, e101. [Google Scholar] [CrossRef]
  283. Emadzadeh, D.; Ghanbari, M.; Lau, W.J.; Rahbari-Sisakht, M.; Rana, D.; Matsuura, T.; Kruczek, B.; Ismail, A.F. Surface modification of thin film composite membrane by nanoporous titanate nanoparticles for improving combined organic and inorganic antifouling properties. Mater. Sci. Eng. C 2017, 75, 463–470. [Google Scholar] [CrossRef] [PubMed]
  284. Chen, X.D.; Wang, Z.; Liu, D.Y.; Xiao, K.; Guan, J.; Xie, Y.F.; Wang, X.M.; Waite, T.D. Role of adsorption in combined membrane fouling by biopolymers coexisting with inorganic particles. Chemosphere 2018, 191, 226–234. [Google Scholar] [CrossRef] [PubMed]
  285. Asadollahi, M.; Bastani, D.; Musavi, S.A. Enhancement of surface properties and performance of reverse osmosis membranes after surface modification: A review. Desalination 2017, 420, 330–383. [Google Scholar] [CrossRef]
  286. Luo, W.; Hai, F.I.; Price, W.E.; Nghiem, L.D. Water extraction from mixed liquor of an aerobic bioreactor by forward osmosis: Membrane fouling and biomass characteristics assessment. Sep. Purif. Technol. 2015, 145, 56–62. [Google Scholar] [CrossRef] [Green Version]
  287. Fan, C.; Nguyen, V.; Zeng, Y.; Phadungbut, P.; Horikawa, T.; Do, D.D.; Nicholson, D. Novel approach to the characterization of the pore structure and surface chemistry of porous carbon with Ar, N2, H2O and CH3OH adsorption. Microporous Mesoporous Mater. 2015, 209, 79–89. [Google Scholar] [CrossRef]
  288. Thommes, M.; Cychosz, K.A. Physical adsorption characterization of nanoporous materials: Progress and challenges. Adsorption 2014, 20, 233–250. [Google Scholar] [CrossRef]
  289. Li, Z.Y.; Yangali-Quintanilla, V.; Valladares-Linares, R.; Li, Q.; Zhan, T.; Amy, G. Flux patterns and membrane fouling propensity during desalination of seawater by forward osmosis. Water Res. 2012, 46, 195–204. [Google Scholar] [CrossRef] [PubMed]
  290. Field, E.L.; Howe, K.J.; Thomson, B.M. Effect of Solids Retention Time in Membrane Bioreactors on Reverse Osmosis Membrane Fouling. Ph.D. Thesis, Department of Civil Engineering, University of New Mexico, Albuquerque, NM, USA, 2010; p. 139. [Google Scholar]
  291. Zhu, X.; Treu, L.; Kougias, P.G.; Campanaro, S.; Angelidaki, I. Converting mesophilic upflow sludge blanket (UASB) reactors to thermophilic by applying axenic methanogenic culture bioaugmentation. Chem. Eng. J. 2018, 332, 508–516. [Google Scholar] [CrossRef]
  292. Deng, L.; Guo, W.; Ngo, H.H.; Zhang, H.; Wang, J.; Li, J.; Xia, S.; Wu, Y. Biofouling and control approaches in membrane bioreactors. Bioresour. Technol. 2016, 221, 656–665. [Google Scholar] [CrossRef] [PubMed]
  293. Pandey, A.K.; Singh, B.; Upadhyay, S.; Pandey, S. Solution for Sustainable Development for Developing Countries: Waste Water Treatment by Use of Membranes—A Review. Int. J. Curr. Microbiol. App. Sci. 2017, 6, 1212–1228. [Google Scholar] [CrossRef]
  294. Burman, I.; Sinha, A. A Review on Membrane Fouling in Membrane Bioreactors: Control and Mitigation; Springer: New York, NY, USA, 2018; ISBN 978-981-10-7332-8. [Google Scholar]
  295. Banti, D.C.; Samaras, P.; Tsioptsias, C.; Zouboulis, A.; Mitrakas, M. Mechanism of SMP aggregation within the pores of hydrophilic and hydrophobic MBR membranes and aggregates detachment. Sep. Purif. Technol. 2018, 202, 119–129. [Google Scholar] [CrossRef]
  296. Kim, E.S.; Hwang, G.; Gamal El-Din, M.; Liu, Y. Development of nanosilver and multi-walled carbon nanotubes thin-film nanocomposite membrane for enhanced water treatment. J. Membr. Sci. 2012, 394–395, 37–48. [Google Scholar] [CrossRef]
Figure 1. Contaminants rejection by the different type of membrane based on molecular weight.
Figure 1. Contaminants rejection by the different type of membrane based on molecular weight.
Water 10 01165 g001
Figure 2. Schematic of membrane operation (a) cross-flow and (b) dead-end filtration.
Figure 2. Schematic of membrane operation (a) cross-flow and (b) dead-end filtration.
Water 10 01165 g002
Figure 3. Schematic of Membrane Bioreactor (MBR) configurations (a) submerged, (b) side-stream [90].
Figure 3. Schematic of Membrane Bioreactor (MBR) configurations (a) submerged, (b) side-stream [90].
Water 10 01165 g003
Figure 4. Flux J vs. average pressure Pave at (a) various air flow velocity in sub-MBR configuration (b) values cross-flow velocity (m/s) in ss-MBR configuration [111].
Figure 4. Flux J vs. average pressure Pave at (a) various air flow velocity in sub-MBR configuration (b) values cross-flow velocity (m/s) in ss-MBR configuration [111].
Water 10 01165 g004
Figure 5. Influence of MLSS concentration on TMP and J of an MBR system during operation [207].
Figure 5. Influence of MLSS concentration on TMP and J of an MBR system during operation [207].
Water 10 01165 g005
Figure 6. Relationship between the fouling rate and permeability during POME treatment using MBR [208].
Figure 6. Relationship between the fouling rate and permeability during POME treatment using MBR [208].
Water 10 01165 g006
Figure 7. Flowchart showing MBR treatment process and membrane fouling mechanism.
Figure 7. Flowchart showing MBR treatment process and membrane fouling mechanism.
Water 10 01165 g007
Figure 8. (ac) Schematic showing mechanism of (a) conditioning stage, (b) steady stage, and (c) abrupt TMP stage.
Figure 8. (ac) Schematic showing mechanism of (a) conditioning stage, (b) steady stage, and (c) abrupt TMP stage.
Water 10 01165 g008
Table 1. Published information on physicochemical properties and biodegradability index (BI) of palm oil mill effluent (POME).
Table 1. Published information on physicochemical properties and biodegradability index (BI) of palm oil mill effluent (POME).
ParametersReference
BODCODBIpHT (°C)TSTSSTVSTNOil/FatNH3-NTP
27,00051,0000.52944.2-40,00018,00034,0007506000--[51]
30,00050,0000.60004.5-16,49559,350-1820---[17]
32,50064,5000.50394.6588---411950--[14]
34,950 ± 145070,500 ± 9170.49574.72-51,880 ± 30026,547 ± 304343,260 ± 1401620 ± 26---[52]
40,00065,0000.61544.555.545,00020,00026,300890150090950[53]
25,00050,0000.50004.78540,50018,00034,0007504000-180[54]
27,00075,0000.36004.3-100,00050,00080,000----[46]
30,000 ± 10,39170,000 ± 76120.42864.75--28,900 ± 3065-980 ± 5010,540 ± 1000-608 ± 81[55]
24,50049,1000.49894.1--18,00026006005300--[48]
45,35773,4980.61714.5-56,27932,005.541,6507606670.569-[49]
-4500-5.6-430082004000500-200-[56]
25,00050,0000.50004.78540,50018,00034,00075,000-3500-[2]
Units: The units of the parameters are in mg/L except for BI, pH and T (°C). Both BI and pH are unit-less. TVS is total volatile solid, TN is total nitrogen, NH3-N is ammonia-nitrogen, and TP is total phosphorous.
Table 2. Published data on the effect of configurations on MBR.
Table 2. Published data on the effect of configurations on MBR.
MBR-ConfigurationParametersReference
TMP barPermeate L/m2hCFV m/sAir flow m/sEDP, kW/m3EDA, kW/m3
sub-MBR0.17.90.5NA44[122]
sub-MBR0.138-NA0.140.0055[124]
sub-MBR26260.250.25--[108]
sub-MBR0.2–0.520–50 0.86960.3–0.6-[64]
sub-MBR30250.320--[26]
ss-MBR450–100---4–12[124]
ss-MBR46160.350.35--[108]
ss-MBR21753-9.92.8[113]
ss-MBR0.61153---[126]
ss-MBR210–301–1.1---[127]
Notes: EDP = energy demand for permeate, EDA = energy demand for aeration, NA = not available.
Table 3. Summary of the published processes for POME treatments using membrane bioreactors.
Table 3. Summary of the published processes for POME treatments using membrane bioreactors.
Treatment Configuration/ProcessPrimary ProcedureOperating ConditionsMajor Results and Contaminants RemovalCritical Findings and Effect of Treatment Process on Membrane FoulingReferences
AerMBR (Pure Oxygen)OTE under pure oxygen treatment at varied MLSS and HRT was investigatedMLSS = 4071 to 11,192 mg/L, HRT = 12–18 h; AF = 141 L/h, DO = 2 mg O2/LHRT 18 h, varying MLSS from 4300–10,275 mg/L: Alpha aeration-factor decrease from 0.6115 to 0.1223; while at HRT 12 h and MLSS of 4017 to 11,192 mg/L; it decreases from 0.2787 to 0.0221MLSS strongly influence OTE; OTE increased by 400% as MLSS decreased with increase in HRT.[167]
AerMBR + AgitationEffects of aeration and agitation on POME treatment were investigated.pH = 5–9; three tanks = Ta, Ts and Tas61.2 and 58.9% removal efficiency for BOD and COD respectively after 6 d; UF membrane improved the treatment with the overall performance of 99.9% at pH 7.39Integrated bioremediation and ultrafiltration membrane improves the total treatment performance to 99.9%. also, the bio pre-treatment reduces fouling propensity[154]
AerMBR + Nano-composite-membraneThe composite UF-PVDF-ZIO was evaluated with aim of deteriorating colour and fouling rate.ZIO = 0.0, 0.5, 1.0, 1.5 and 2.0 wt %.Best colour removal and permeability was 80.5% and 50.18 LMH respectively at 0.5 wt % of ZIO dosageZIO nanoparticle improves colour removal, permeability; reduce the fouling rate but collapse after 4 cycles. M2.0 retain it throughout[152]
Also, ZIO oxidises to release antimicrobial and hydrophilic radicals. This makes it a good antifouling material[152,240,241]
ZIO reduces fouling[152]
AerMBR + MicrobialThe experiment investigates the correlation between the microbial community and MBR performance for POME treatment under alternating conditions.Aeration period = 40 d, non-aeration period = 10 d; sample collection interval = 25 d, 50 d and 75 dAerobic conditions favour microbial as follows: Proteobacteria (19–23%), ODI (11–15%), Chloroflexi (11–13%). While; in the Non-aerobic condition: ODI (20%), proteobacteria (18%) and plantomycetes (16%) were visibleProtein was the main constituent of ESP. Taxonomic profile on the day 75 is similar to day 25. Proteobacteria survives in both conditions but ODI dominates under no aeration condition.[141,152]
ESP constitutes the major component of the accumulated biofilm[242]
AerMBR + TNT nanoparticle PVDFComposite PVDF-TNT was fabricated with variable TNT amount and evaluated under varied POME concentration with aim of decolourizationTNT load = 0–1 wt %; POME concentrations = 100, 75 and 50% with DF: 1, 2, 3At PVDF-TNT 0.5; colour removal = 67.3%. but 5.7% flux reduction was observed after 5 cycles of filtrationsTNT improve the colour removal, also, it is a good antimicrobial material. Hence, capable of mitigating membrane fouling.[21]
However, TNT could be exorbitantly expensive compared to other antimicrobial nanoparticles[243]
AerMBR + Adsorbent (AC + Zeolite)Performances of 2 different adsorbents were compared with SubMBR evaluated without adsorbent same conditions.Dosage: AC = 2, 4 g/L; Zeolite = 2 g/L; SRT = 70 dTotal of COD removal with adsorbents ranged from 97.5–98.5% and without was 95.2%, colour reduced to 16–26 Pt-Co while without adsorbent was 80 Pt-Co and improve flux to 42 LMHAdsorbent improve flux, TMP, and reduce SMP deposition, COD and colour significantly[18]
The improvement in flux indicates that the membrane is less prone to fouling due to the significant reduction in contaminants[185]
AerMBR + ACThe performances of SubMBR with and without adsorbent were compared.SRT = 20 d; HRT = 3.1 h; MLSS = 1.25 g/L; PAC dosage = 5 g/L; Backwash at F-R = 1 h per 1 min with 30 L/m2hWith PAC, COD removal = 100%, DOC removal = 99% requires 7.5 kPa But without; COD removal = 94% and DOC = 95%. Requires 20 kPa to obtained the maximum flux of 20 L/m2PAC lowered the operating TMP, improves organic matter removal and mitigate fouling rate[180]
AerMBReffect of SRT and HRT on membrane fouling was studied under 10 min-4 min intermittent operationSRT = 30, 15, 4 d; HRT= 12, 8, 4 h; pH = 7.2 ± 0.1; DO = 2.0 mg/L; flux = 4 L/m2/h; Temp. = 25 to 35 °C17 μm cake layer observed after 4 d of SRT and 12 h HRT; COD removal = 93%, TSS = 98%, NH3-N = 80% and PO4 = 30%SRT influence organic matter removal. But equal removal of TSS at all SRT, higher HRT with shorter SRT induced faster fouling[244]
AerMBRFouling behaviour was studied in three separate Submerged MBR: A,B,C under varied aeration intensity as 150, 400 and 800 L/h respectively at constant TMP.HRT= 10–12 h; MLSS = 6000 mg/L; SRT = 30 d; Aeration:= 150, 400 and 800 L/h; DO= 3.21, 4.76 and 6.5 mg/L; TMP = 3.97 kPapermeate of A and C, decline after 10 h operation; B reach a steady value after 100 h. Sludge bulking observed at 2.0 mg/L DOLow DO causes sludge bulking because of the overgrowth of filamentous bacteria.[131]
Fouling rate decrease with an increase in aeration at the initial stage but breaks flocs to smaller sizes after sometimes hence more formation of colloids and higher fouling.[245]
AnMBRKinetic coefficient of the POME was studied.MLSS = 11,760–20,800 mg/L, MLVSS = 8938–17,680 mg/L, OLR = 1–11 kg COD/m3-d, HRT = 600.4 to 6.8 dGYC and SGR was 0.67 g vss/g COD, and 0.24 d−1 respectively. COD removal efficiency varied from 96.6 to 98.4% while the CH4 yield varied between 0.25 to 0.87 L/gBiomass concentration has a significant influence on CH4 production. Concentrated biomass yield more CH4 but the membrane is more prone to fouling[196]
AnMBRKinetic coefficient was investigated under anaerobic condition and the result was compared with kinetic equation modelsMLSS = 8220 to 15,400 mg/L, MLVSS = 6329 to 13,244 mg/L, OLR = 2–13 kg COD/m3-d, HRT = 400.6 to 5.7 dCOD removal efficiency varied between 94.8 to 96.5%; GYC was 0.62 g vss/g COD and SD was 0.21 d−1The treatment process deteriorates COD significantly, varying MLSS concentration influences CH4 yield [222]
AnMBREffect of intermittent Filtration-Relaxation (F-R) and membrane fouling under anaerobic conditions were investigatedF-R: L1 = 240 s–30 s; L2= 480 s–30 s; L3 = 720 s–30 s and L4 = 960 s–30 sProtein and carbohydrate dominates and L1 gives the least fouling rate with 50% lower in Hydraulic ResistanceMore Frequent F-R reduced prevents fouling by 50%. SMP and EPS contributes to fouling mechanism[201]
AnMBRUnder a specified range of TMP (125–130 mbar), the influence of SRT on membrane fouling under double stage anaerobic conditions was studied.SRT = 15, 30 and 60 d; Flux = 2.41 L/m2-h; TMP = 125–130 mbarFoulants deposited for SRT = 15, 30 and 60 d was 34.28–16.81; 64.1–26.07 and 84.19–37.18 mg/g vss respectivelyThe longer the SRT, the denser the biofilm hence the more the severity of the fouling[204]
AnMBR +Effect of Baffled Aeration Flotation (BAF) system on AnMBR performance was studied.Number of BAF = 5; HRT = 3, 4 and 5 d; AFR = 11, 8 and 5 L/minAt optimal condition of 5 d HRT and 11 L/min aeration: BAF system gives 35.5, 86.4, 57.7, 57.3, 59.7 and 52.6% removal of while the overall system gives 97, 93.9, 99.8, 94.5, 96.1 and 99.9% respectivelyBAF system improves the performance of the AnMBR in terms of contaminants removal and reduces the fouling rate[61]
HybMBR (+PAC)Effect of PAC particle sizes: fine = Mf; medium = Mm; coarse = Mc, on POME treatment were studied and compared with treatment without PAC under anaerobicMLSS = 8050 mg/L; MLVSS = 6850 mg/L; SRT = 30 d; HRT = 6 d PAC dosage: Mb = 0; Mc = 5 g/L; Mm = 5 g/L; Mf = 5 g/L; Temp.= 45 °C; OLR= 7658 ± 408 mg COD/L-dWithout PAC: removal efficiency are COD: 64.90 ± 1.46%; protein: 2407 ± 230 mg/L; polysaccharide: 71 ± 1.827 mg/L. With PAC: Mf present the best performance with 78.53 ± 0.66% COD removal; protein: 1647 ± 175 mg/L; Mc: least COD removal with 72.99 ± 1.47% and protein: 2075 ± 305 mg/LThe Fine PAC (Mf) presents better COD removal of 90.55 ± 0.21 than medium PAC (Mm), particle size influence adsorption performance, biogas yield and reduce fouling rate[246]
This shows that PAC of fine particles sizes present more active sites for COD adsorption[26]
HybMBRPOME was pre-treated using hybrid anaerobic and aerobic process-then polished with membrane techniques. The treatment lasted for a period of a yearAnaerobic HRT = 9.8 d; Aerobic HRT = 48 h; DO = 3.5 mg/L55.6% of the waste oil in raw POME was recovered at anaerobic treatment; aerobic treatment degenerate BOD3 level to less than 20 mg/LDeterioration of COD is more significant during anaerobic treatment, while the aerobic process is more suitable for BOD removal.[69]
Also, the hybrid system is suitable for resource recovery, such as biogas[140]
HybMBRThe HybMBR was acclimatized for 45 days with 4 different samples at a varied MLSS to observe the fouling rate on the membrane.MLSS = 5–20 g/L; HRT = 11, 7 and 8 h; SRT: 70 days, 120 rpm; flux: 11 L/m2-hAs varied from MLSS from 5 to 20 g/L, the flux reduced down to 400%. However, 97% denitrification and nitrification was obtained at 20 g/L MLSSMLSS concentration strongly influences fouling.[207]
But, higher MLSS concentration promotes NH4-N and nitrogen substance removal in MBR[162,207]
HybMBR (+adsorbent)The HybMBR was Evaluated under different adsorbent and varied dosage but same SRT.Dosage: AC = 4 g, Zeolite = 8 g, Moringa Oliefera = 12 g and SRT = 30 daysAdsorbent: 58, 48 and 42% removal of SMP was observed by AC, Zeolite and MO respectively.Adsorbents reduce SMP significantly. At optimal operating conditions, AC gives better performance.[56]
At an optimal dosage of AC; 70 and 85% SMP removal efficiency and fouling reduction were archived respectivelyThe improvement in performance was attributed to the availability of the active sites to adsorb the SMP[56,247,248]
HybMBRTMP and flux was continuously monitored for a complete season under varied temperature to investigate the membrane fouling characteristicsSRT= 20 days, HRT = 17 h, Temp. = 27 to 13 °CAt summer the TMP was 25 kPa with 10 L/m2-h permeability flux while at winter the TMP was 60 kPa at 10 L/m2-h. This study revealed that temperature influences the organic concentration of the supernatant, filtration resistance, viscosity and humic substance.[210]
Higher temperature disintegrates the flocs aggregates, thereby increasing the concentration of the foulants of smaller sizes.[141,214]
HybMBREffect of biomass concentration on POME treatment and membrane fouling with fitted controllable recycling device and agitator at the anoxic and anaerobic conditionMLSS = 4000–8000 mg/L, OLR = 1.77 to 1.87 kg COD/m3, steady inflow = 108 m3/d, flux = 15 L/m2hThe removal efficiency of COD, SS, TN and TP was 94, 98, 83 and 64% respectivelyBoth Protein and carbohydrate contribute to the fouling.[58]
The hybrid system deteriorates the organic pollutants significantly[226]
HybMBREach sections of the hybrid system was subjected to different operating conditions. Also, the MLSS concentration was varied with the aim of investigating the fouling effectAnaerobic: HRT = 12 h, pH = 5.5 to 6.5, DO = 0-0.1 mg/L Anoxic: HRT = 6 h, pH = 7.2–8.5, DO= 0.3–0.6 mg/L Aerobic: HRT = 4 h, pH=7–7.5, DO= 6–8 mg/L. MLSS = 4–8 g/L94, 98, 83 and 64% for COD, SS, TN and TP removal efficiency respectively. At 12 LMH flux, permeability increased to 70 LMH.Initially, the filtration rate was steady, then, diminished abruptly after the critical flux 16 LMH.[208]
This was due to the fouling which often causes TMP jump.[65,108]
HybMBRAn integrated Biological-composite membrane system was acclimatized at room temperature. The anaerobic and aerobic section takes about 111 and 10 days respectively, for complete acclimationInfluent COD: 4331 to 35,000 mg/L, UF membrane TMP = 200,000 Pa; RO membrane TMP = 1,300,000 Pa Temperature = 25 °C, OLR = 10 kg COD/m3-dAnaerobic deterioration of COD = 93%, While, 22% of COD was removed in aerobic. UF membrane removes SS and turbidity with a total of 99.288% efficiency the RO membrane removed all the remnants contaminants with 100% efficiency.Anaerobic performed excellently in terms of COD removal. UF membrane is suitable for removing SS and turbidity.[15]
RO remove all the contaminants including the colour but it requires high TMP.[249,250]
The setup mitigate fouling rate and also improved membrane flux recovery (between 91.7 to 95.3%)[15]
HybMBRPOME passes through the integrated system for treatment and recourse recovery (biogas and bio fertilizer)Screening operation, biological treatment, reclamation operation, biogas generation, sludge disposal systemAbout 94% of COD removal was achieved during anaerobic treatment; improvement in the COD removal of 97% was observed after Nano air flotation treatment.Zero discharge from POME was achieved; SS, residual oil, NH4-N, toxic material and colour were removed in the aeration section.
In overall, less fouling was observed using the integrated system
[47]
BOD is less than 20 mg/L after the final treatment with 80% consistency
SonMBREffect of sonication on MBR was studying for different sonication durations: 2 h and 1 hAcclimatization period = 2 days, pH = 6.8 to 7.8, pressure = 2 barCOD removal efficiency for 2 h and 1 h sonication was 98.75 and 97.71% respectively; more CH4 yield was obtained at 2 h sonicationSonication improves organic matter removal, CH4 production rate and mitigating fouling. This implies that the biomass degradation is more at longer sonication duration[218]
SonMBRThe study was conducted to investigate the effect of sonication on the kinetic coefficient of POME prepared in three samples: Sc, Sp and SrSonication time = 3 h, at constant MLSS concentrationSpecific Growth Rate in Sc, Sp and Sr was 0.0115, 0.279 and 0.3550 kg VSS/kg COD respectively. 87.5% and 85% COD removal were observed with and without sonication respectivelyThe higher growth rate was observed in reacted sample Sr under sonication with improved COD removal. This implies that the microbes are more active in the reacted sample under the applied sonication[218,219]
SonMBR anaerobicEffect of sonication in anaerobic degradation and CH4 production was investigatedMLSS = 13,800 to 22,600 mg/L; MLVSS = 10,400 to 17,350 mg/L; OLR = 1–15 kg COD/m3-d; SRT = 15.8–300 d; HRT = 500.8 to 8.6 days; pH = 2 to 12; pressure = 2–4 barsCOD removal were between 92.8–98.3%; Growth Yield Coefficient = 0.73 g VSS/g COD, SMD rate= 0.28 d 1 , CH4 yield= 0.27–0.62 L/g COD-dMaximum yield of CH4 (0.7 L/g COD/d) was obtained at highest biomass concentration (22,600 mg/L) and highest HRT[197]
While the least yield was obtained at the lowest MLSS concentration.[188,196]
Also, SRT has an insignificant effect on CH4 production rate[188]
SonMBR + MicrowaveThe effect of combined microwave and sonication process on organic degradation and CH4 production rate was investigatedTemp. = 32–37 °C; TVS = 32–34 g/L; MF = 2450 Hz; IT = 3 to 15 min; UF: = 18 Hz; intensity = 0.2 to 0.5 WmL−1; ST = 1–30 min; HRT = 15–2 dAt optimal microwave and 11 min sonication duration, the SCOD/TCOD and BOD/SCOD ratio increased to 0.77 and 0.95 respectively. Also, about 44 mL of CH4 obtainedSCOD/TCOD ratio and BI increase significantly under sonication and microwave process. The higher CH4 yield obtained under sonication[219,251]
Son MBRThe combined ultrasonic and sono-thermal was applied in an MBR system for pre-treating the POME at a varied temperature and time.Temperature = 45, 55, 65 and 75 °C. Ultrasonic frequency = 37 Hz, sonication time = 1–6 h39.05% increment in degradation was observed when ultrasonic-sono-thermal pretreatment were appliedThe ultrasonic and sono-thermal pretreatment promotes degradation and is more rapid at elevated temperature (75 °C)[222]
Also, after the pretreatment, the average particles size 321.94 µm reduced to 79.16 µm
SonMBRApplication of ultrasonic pretreatment to accelerate bio-hydrogen production.Ultrasound dose = 0,91, 143 and 195 J/mL; pH = 7;Average COD removal = 65%
Bio-H2 yield = 0.7 mmol H2/g COD
In overall, 38 and 20% increment in bio-H2 yield and COD removal, respectively were obtained.
The ultrasonic improve the overall performance of the reactor significantly. The best performance was obtained at the highest ultrasound dose (195 J/mL)[223]
SonMBRThe application of nano zero valent ions to treat POME was accelerated by applying ultrasound.pH = 2–4; sonication intensity = 10–50%; sonication time = 3–10 minAt 50% intensity, COD was deteriorated by 80% within 2 hSignificant improvement in COD removal was observed when ultrasound was applied.[16]
TherMBR + MFCEffect of temperature and microbial in MBR was investigated under intermittent operation and variation in TMP was monitoredTemp. = 8.7–19.7 °C; F-R = 10–2 min; HRT = 4.9 h; SRT = 40 d, TMP = 60 kPaAs temperature increases from 8.7 to 19.7 °C, the MLSS of SMP reduces from 28.1 to 2.2 mg/g. also, Proteobactaria, bacteriodestes, nitrospira, firmicutes and acidobacteria dominate with 41–51.8%, 6.7–22.2%, 8.9–15.1%, 4.32–10% and 2.2–7.0% respectivelyIncrease in temperature decreases SMP. Lower temperature is more suitable for Proteobacteria, while the higher temperature is capable of activating some bacteria such as Zoogloea from a dormant state.[224]
TherMBR + MFCThe effect of temperature and MFC in an MBR was investigated. R1, R2, and R3 were run under mesophilic while R4 was under thermophilic.Sample: reactor R1, reactor R2, reactor R3, reactor R4; SRT = 30 d; PAC = 5 g/L dosage; Temp.= 35, 45 and 55 °CR3 shows the best removal of COD and polysaccharide with 95.6 ± 0.3% and 73.01% respectively, under the mesophilic condition. While R4 under thermophilic present COD removal with 79.23 ± 9.36%. Also, highest MLVSS was obtained in R1 with 38,133±1804 mg/L while least was observed in R3 with 16,467 ± 3239 mg/LBest performance as obtained at mesophilic conditions. This indicates that mesophilic conditions are more suitable for biodegradation.[212]
Fouling rate is higher at the mesophilic condition.[31,195,252]
TherMBRThe Effect of varied temperature on MBR performance was investigated. The MBR system was incorporated with PACPAC dosage = 50 g/L, Temperature = 10–20 °CAt 10 °C, Start-up time = 9 days; Complete nitrification was attained after 10 days. NH3-N removal = 90% at both 10 and 20 °C Lower temperature (10 °C) cause delay in the start-up of the system, nitrification process. The membrane is less susceptible to fouling at this temperature[234]
Notes: AF = air flow rate, Ta = aeration tank, Ts = stirring tank, Tas = aeration + stirring tank, ZIO = zinc iron oxide, F-R = frequency-relaxation, L1,2,3 = duration for F-R, BAF = baffled aeration flotation, Mf = fine size, Mm = medium size and Mc = coarse size, Sc = control sample, Sp = permeate sample, Sr= reacted sample, R1,2,3,4 = reactors.

Share and Cite

MDPI and ACS Style

Abdulsalam, M.; Che Man, H.; Isma Idris, A.; Faezah Yunos, K.; Zainal Abidin, Z. Treatment of Palm Oil Mill Effluent Using Membrane Bioreactor: Novel Processes and Their Major Drawbacks. Water 2018, 10, 1165. https://doi.org/10.3390/w10091165

AMA Style

Abdulsalam M, Che Man H, Isma Idris A, Faezah Yunos K, Zainal Abidin Z. Treatment of Palm Oil Mill Effluent Using Membrane Bioreactor: Novel Processes and Their Major Drawbacks. Water. 2018; 10(9):1165. https://doi.org/10.3390/w10091165

Chicago/Turabian Style

Abdulsalam, Mohammed, Hasfalina Che Man, Aida Isma Idris, Khairul Faezah Yunos, and Zurina Zainal Abidin. 2018. "Treatment of Palm Oil Mill Effluent Using Membrane Bioreactor: Novel Processes and Their Major Drawbacks" Water 10, no. 9: 1165. https://doi.org/10.3390/w10091165

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop