Open access peer-reviewed chapter

Integration of Forward Osmosis in Municipal Wastewater Treatment Applications

Written By

Stavroula Kappa and Simos Malamis

Submitted: 22 December 2020 Reviewed: 07 January 2021 Published: 31 May 2021

DOI: 10.5772/intechopen.95867

From the Edited Volume

Osmotically Driven Membrane Processes

Edited by Muharrem Ince and Olcay Kaplan Ince

Chapter metrics overview

505 Chapter Downloads

View Full Metrics

Abstract

In recent years, the research community has made constant efforts to develop new technologies for the recovery and valorization of water, nutrient and energy content of municipal wastewater. However, the recovery process is significantly limited due to the low-strength of sewage. Over the last 10 years, the Forward Osmosis (FO) process, has gained interest as a low-cost process with low membrane fouling propensity, which can convert municipal wastewater into a concentrated low-volume effluent, characterized by high organic and nutrient concentration. This chapter presents the main configurations that have been implemented for the concentration of municipal wastewater using FO, including their performance in terms of contaminant removal and water/reverse salt flux (Jw/Js). Furthermore, the draw solutions and respective concentrations that have been used in FO for the treatment of sewage are reported, while at the same time the positive and negative characteristics of each application are evaluated. Finally, in the last section of this chapter, the spontaneous FO followed by anaerobic process is integrated in a municipal wastewater treatment plant (WWTP) and compared with a conventional one. The comparison is done, in terms of the mass balance of the chemical oxygen demand (COD) and in terms of the energy efficiency.

Keywords

  • forward osmosis
  • municipal wastewater
  • configurations
  • draw solution
  • COD mass balance

1. Introduction

Water scarcity is one of the most serious threats which our planet faces [1]. Globally, water demand is predicted to increase by 35% more than sustainable supply by 2040/50, if the linear water management model continues to be implemented [2]. The European Union (EU) encourages the implementation of a circular economy model, through its strategy called “Closing the loop—a EU action plan for the Circular Economy” in 2015 and European citizens must seize the opportunity to close the loop of water, resource and energy management [3]. Among various types of water, seawater and wastewater are two alternative sources, which are readily available, especially in coastal, arid areas [4]. Both need to be treated before they can be rendered suitable for use. Membrane processes such as microfiltration (MF), ultrafiltration (UF), nanofiltration (NF) and reverse osmosis (RO) are particularly effective in the purification of non-conventional water sources and count many applications in both the wastewater and the desalination sector [5]. In particular, the RO, holds a prominent position in water desalination, compared to traditional thermal desalination processes [6]. As high energy consumption is required to overcome the osmotic potential, reverse osmosis is not applied in many water-stressed areas [5].

Forward osmosis is one of the most attractive membrane-based processes that requires two solutions of different osmotic concentrations (high and low), separated by a semi-permeable membrane to be realized [7]. Water molecules are spontaneously diffused from the low osmotic potential solution (feed side) to the high osmotic solution (draw solution or DS), to equalize the concentration difference, while the semi-permeable membrane acts as a barrier that rejects the salts and contaminants [8]. The natural osmotic pressure of FO makes it stand out from conventional RO, by offering high water recovery, reduced membrane fouling potential, greater effectiveness, low cost, and reduced energy demand [8, 9]. All these positive aspects have led to a notably high trend of publications on FO applications in various water sources, such as seawater and wastewater, with more than 97.5% of publications since 2009 [10]. Among them, several researchers investigate the feasibility of integrating the FO process in a novel sewage treatment system based on the circular economy concept, as the main goal is to valorize the chemical energy, water, and nutrients of sewage. This innovative application of FO and its combination with appropriate downstream technologies is really promising. As the results show, the wastewater is converted into a small volume liquid, characterized by a high concentration of organic matter, as it can be concentrated up to 8–10 times, while the recovery of phosphorus can reach up to 90%, replacing the need for chemical fertilizers [11, 12]. However, there are many challenges that need to be overcome for this application, the most important of which is the selection of the most appropriate DS, which despite the significant efforts has not been found to date [13, 14, 15].

This chapter presents the main configurations that have been implemented to concentrate municipal wastewater using FO, including their performance in terms of contaminant removal and Jw/Js. The draw solutions and their concentrations that have been used in the FO process for the treatment of sewage are reviewed, while at the same time the positive and negative characteristics of each application are evaluated. Finally, in the last section of this chapter, the spontaneous FO followed by an anaerobic process is integrated into a municipal wastewater treatment plant and compared with a conventional activated sludge process (CAS), in terms of COD and corresponding energy efficiency, emphasizing the key impact of the FO in the latter process.

Advertisement

2. Forward osmosis configurations and performance in municipal wastewater management

The main benefit of the FO process in municipal wastewater treatment is that it converts sewage from low-strength liquid to a concentrated bulk, which consists of high a concentration of organic matter and nutrients [16, 17]. According to Korenak et al. [18], the FO process is characterized by high membrane fouling reversibility, while it can significantly minimize space requirements in a municipal WWTP. Considering all the above, three basic configurations have emerged for the integration of the FO process in the municipal WWTPs, which are illustrated in Figure 1.

Figure 1.

Configurations for the integration of the FO process in the municipal WWTPs.

2.1 Osmotic membrane bioreactor (OMBR)

In 2008, an innovative system was introduced, in which FO membranes were submerged into a typical membrane bioreactor (MBR) module; this system was called OMBR (Figure 1(A)) [19]. The replacement of UF or MF membranes in the conventional system by FO membranes resulted in better performance in terms of contaminants’ rejection (79.7–100% of COD, Table 1). In addition, the absence of hydraulic pressure contributed to lower fouling tendency and probably lower energy requirements. Despite the benefits of OMBR over traditional systems, two major challenges are still under investigation; low Jw rate and salinity accumulation [19, 31]. The findings confirm that the decline in Jw was greatly affected by the salt accumulation, even with the implementation of improved membrane materials, such as thin-film composite (TFC), achieving an average rate equal to 3.9 ± 0.5 L m−2 h−1 [32]. In addition, the microbial community of the reactor can either be partly or fully inhibited, due to the gradual building-up of salts, which occurs due to the Js [31, 32].

Membrane TypeFeedDrawRemoval Efficiency (%)Jw (L m−2 h−1)/Js (g m−2 h−1)Ref.
TOCCODTPNH4-NTN
CTA (HTI, USA)Municipal wastewaterSeawater brine9097.999N.A.[12]
CTA (HTI, USA)Synthetic municipal wastewater1 M NaCl98 ± 199 ± 18–5/NA.[20]
TFC (HTI, USA)Synthetic municipal wastewater1 M NaCl96 ± 199 ± 13/N/A[20]
CTA (HTI, USA)Municipal wastewater42 g L−1 NaCl9999824–4.5/ N.A.[21]
CTA-ES (HTI, USA)Sunthetic municipal wastewater48.4 g L−1 MgCl298983.7–3.3/N.A[22]
CTA (HTI, USA)Raw municipal wastewaterSynthetic seawater79.7 ± 992 ± 3.388.1 ± 5.54.86–3.24/ N.A[23]
CTA-ES (HTI, USA)Sunthetic municipal wastewater48.4 g L−1 MgCl2 & 49 g L−1 NaCl9898 (PO43−-P)986.64–8.95 (NaCl); 6.46 (MgCl2)/ NA.[24]
TFC (HTI, USA)Sunthetic municipal wastewater1 M NaCl10095.6 (PO43−-P)4314/ N.A.[25]
CTA (HTI, USA)Raw sewage0.5 M NaCl959595901.7/ NA.[26]
CTA & TFC (HTI, USA), Biomimetic (Aquaporin, Denmark)Synthetic wastewater0.5 M NaCl.>9590–10060–8015.6–15 (TFC); 5.5 (CTA)/0.085 (Aquaporin); 82.7 (TFC); 5.5 (CTA) mmol m−2 h−1[27]
CTA-ES (HTI, USA)Synthetic wastewater1 M MgCl299995.45/ N.A.[28]
TFC (Aquaporin A/S, Asia)Synthetic wastewater1.2 M NaCl with SDBS100>908550–8010/ N.A[29]
TFC (Aquaporin A/S, Asia)Synthetic wastewater1.2 M NaCl without SDBS98>90>9060–100<10/ N.A.[29]
CTA-ES (HTI, USA)Synthetic municipal wastewater70 g L−1 NaCl9999 (PO43−-P)9210.42–6.4/9.9–34.99[30]

Table 1.

Osmotic Membrane Bioreactor applications to treat municipal wastewater.

2.2 Anaerobic OMBR (An-OMBR)

The combination of MBR technology with the anaerobic process has been extensively investigated in the last 10 years, due to the environmental benefits of both [33]. However, the low-strength nature of sewage is a major obstacle to the effective application of the anaerobic process in municipal WWTPs; containing a high amount of water with low organic and nutrients concentration. Due to the methane’s solubility in water (22.7 mg L−1, at room temperature), a large part of the produced gas escapes with the treated effluent of the anaerobic process (ranges between 20 and 60%) [34]. Due to the aforementioned barriers, it is difficult to implement anaerobic processes for municipal wastewater treatment particularly in areas, where the sewage temperature drops below 15°C, during the winter period. The incorporation of FO, either as a pre-treatment step or submerged into the MBR system, significantly enhances the resource recovery potential in the anaerobic process. Compared to OMBR, An-OMBR (Figure 1(B)) is characterized by significantly lower energy requirements, due to the replacement of energy-demanding aeration, while biogas production contributes to the coverage of specific energy needs. According to Zhang et al. [35], due to the 2–3 times smaller pore size of the FO membranes over conventional UF or MF membranes, the dissolved methane content in the An-OMBR treated effluent was eliminated, even as a function of different operating parameters. Regarding the yield of methane, Zhang et al. [35] and Gu et al. [36] observed particularly satisfactory production that reached 0.256 L CH4 g−1 COD and 0.25–0.3 L CH4 g−1 COD at mesophilic conditions, respectively. In addition, anaerobic biomass showed high resistance to increasing salt concentrations and was not affected even when the concentration was equivalent to 200 mM sodium chloride (NaCl) [36]. As shown in Table 2 the FO membranes achieve high rejection of contaminants; specifically the Total Organic Carbon (TOC)/COD and PO4-P removal was 95% and 73%, respectively. However, due to the lack of ammonia removal, its accumulation has been observed in the reactor, but not in concentrations that can lead to the interruption of the anaerobic process [36]. In recent years, an alternative configuration has been proposed, which includes the addition of MF membranes both to OMBR and An-OMBR systems, the so-called Microfiltration- Osmotic Membrane Bioreactor (MF-OMBR). The main goal of this hybrid system is 1) to balance the salts concentration in the reactor so as to prevent an inhibition event and 2) to apply resource recovery methods to its nutrient-rich treated effluent. Nonetheless, according to Wang et al. [39], the FO membranes achieved much lower ammonia rejection rates (39–50%) compared to an An-OMBR system (62.7–81.2%), while the addition of another membrane significantly raises both the maintenance and the investment cost of the entire system [39, 43].

Membrane TypeFeedDrawRemoval Efficiency (%)Jw (L m−2 h−1)/Js (g m−2 h−1)Ref.
TOCCODTPNH4-NTN
CTA (HTI, USA)Synthetic municipal wastewater0.5 M NaCl959970–8010–3/ N.A.[36]
CTA (HTI, USA)Synthetic municipal wastewater0.5 M NaCl96100629.5–3.5/ N.A.[37]
TFC (Aquaporin Denmark)Synthetic municipal wastewater116.6 g L−1 MgSO4>9595 (PO43−-P)>950.78–0.26/ NA.[38]
CTA (HTI, USA)Synthetic domestic wastewater0.5 M NaCl>966–3.4/ N.A.[39]
CTA-ES (HTI, USA)Municipal sewage0.5 M, 1 M and 1.5 M NaCl9610088896 (0.5 M); 10 (1 M); 13 (1.5 M)/ 4.26 (0.5 M); 7.65 (1 M); 11.84 (1.5 M)[40]
CTA-NW (HTI, USA)Synthetic municipal wastewater2 M C4H6MgO49673 (PO43−P)51.43.5–1.09/ 2.5–1.6[35]
TFC (HTI, USA)Synthetic municipal wastewater0.5 NaCl>9599 (PO43−P)012 ± 0.7–2 ± 0.2/ N.A.[41]
CTA (HTI, USA)Synthetic sewage0.5 NaCl>939928–458.7 ± 0.3–4.0 ± 0.2/ N.A.[42]

Table 2.

Anaerobic Osmotic Membrane Bioreactor applications to treat municipal wastewater.

2.3 Pre-concentration with FO

Alternatively, the FO unit can be applied as a pre-condensation step in municipal WWTPs (Figure 1(C)), achieving a similar goal to the previously analyzed configuration, as it can be combined by suitable downstream processes for resources and energy utilization. As reported by Ansari et al. [34], the submerged FO configuration is significantly disadvantaged compared to the separate one, as the former gets in contact with the dense activated sludge, while the latter with the diluted primary treated effluent. In contrast, a recent study that examined both approaches in parallel, direct osmosis showed a significant decline in Jw performance compared to OMBR system [7]. On the other hand, a prolonged biodegradation study (approximately 7 months) of both cellulose triacetate (CTA) and TFC membranes demonstrated that the long-term exposure to activated sludge significantly affects their performance, in terms of water permeability and Js [44]. Sun et al. [7] found that the direct FO module is characterized by reversible membrane fouling over the submerged OMBR membrane, mainly due to the lower abundance in the microbial load of the feed solution. In terms of performance, as shown in Table 3, this FO configuration achieves the retention of organic load by a percentage ranging from 71.9 to 100%. At this point, it should be noted that based on the current literature most studies refer to FO as either a separate or integrated system of an Anaerobic Membrane Bioreactor (An-MBR), while alternative anaerobic treatment systems are not frequently investigated.

Membrane TypeFeedDrawRemoval Efficiency (%)Jw (L m−2 h−1)/Js (g m−2 h−1)Ref.
TOCCODTPNH4-NTN
CTA (HTI, USA)Raw Sewage3.5% NaCl71.9 (AL-FS); 69.3 (AL-DS)5.2 (AL-DS) and 5.4 (AL-FS)/ N.A.[45]
CTA (HTI, USA)Municipal wastewater0.5 M NaCl99.8 ± 0,699.7 ± 548.1 ± 10.567.8 ± 7.37.7–6.5/ 5[17]
TFC, (Aquaporin, Denmark)Domestic sewage1, 1.5 & 2 M MgCl2⋅6H2O76–8075665.11 (1 M); 6.66 (1.5 M); 6.92 (2 M)/ N.A.[16]
CTA (HTI, USA)Sewage0.2–4 M NaCl96.595.493.396.510 (0.5 M) - 25 (4 M)/ N.A.[46]
CTA (HTI, USA)Municipal wastewaterSynthetic seawater79.9 ± 6.793.3 ± 3.385.4 ± 5.65.37–2.5/ N.A.[23]
TFC (Toray, Japan)Municipal wastewater11.7 g L−1 NaCl85.590no rejection (AL-FS); 50 (AL-DS)5.1 ± 1 / 4.8 ± 8.6[47]
Aquaporin TFC (Sterlitech Corporation, WA)Synthetic wastewaterSynthetic seawater with C. vulgaris100 ± 0.699.5 ± 0.546.1 ± 3.411.59 ± 0.49[48]
Aquaporin TFC (Sterlitech Corporation, WA)Synthetic wastewaterSynthetic seawater without C. vulgaris99.5 ± 0.799.1 ± 0.146.2 ± 2.612.02 ± 0.35[48]
TFC (HTI, USA)Municipal wastewater &prefilteredSynthetic seawater brine2.38–2.67 CF3.3–3.5 CF1.31–1.75 CF1.58–1.94 CF19.90 (raw sewage); 18.15 (filtered)/ N.A.[49]
CTA (HTI, USA)Synthetic municipal wastewater4% NaCl9967–6856–59N.A.[50]
CTA-ES (HTI, USA)Synthetic municipal wastewater0.25 M (NH4)2SO4 (SOA), KH2PO4 (MKP),(NH4)H2PO4 (MAP)989989,56 ± 1.52 (SOA); 98.96 ± 0.33 (MPK); 89.67 ± 1.27 (MAP)2.58/ 0.57 (SOA); 2.11/ 1.17 (MPK); 1.97/ 0.11 (MAP)[51]

Table 3.

Pre-concentration of municipal wastewater using FO.

Advertisement

3. Draw solutions

In contrast to other osmotic, membrane-based technologies, the application of high osmotic potential is the driving force in the spontaneous FO process [52]. Therefore, the selection process of the most effective solution acts as a cornerstone of the FO and plays a crucial role on its performance as well as on downstream processes [15]. In an ideal physicochemical context, the parameters listed in Table 4 must be met to classify a solution as appropriate [52, 53, 54].

ParameterImpact
Osmotic PressureDS with higher osmotic pressure than the feed generates higher Jw
Water SolubilitySoluble compound produces higher osmotic concentration and therefore retains Jw and water recovery at higher percentages
ConcentrationHigher concentrations contribute to higher Jw rates, but particularly high concentrations inhibit flux’s increase, as mass transfer phenomena occur, such as concentration polarization
Diffusion, viscosity and molecular weightSmall molecules are distinguished by high aqueous solubility and high osmotic pressures, but in comparison with the large ones they are characterized by a higher diffusion coefficient and Js
Toxicity and degradationLow reverse salt diffusion to minimize the risk of toxicity and contamination of downstream systems and the recovered product. Low rate of degradability, unless it is beneficial
Cost-effectiveEasily re-concentration at competitive cost
AvailabilityAvailable in large quantities with low price and easy handling

Table 4.

Main parameters that characterized the ideal DS.

In recent years, significant efforts have been made by researchers to combine the above parameters and develop an ideal DS, which will be compatible with the application of FO in the municipal wastewater treatment sector [15, 55, 56]. Alternative systems have been developed; different configurations have been applied to integrate the FO in several stages of a municipal WWTP; as pre-treatment, secondary and post-treatment steps for nutrient recovery. Obviously, the treatment level and the quality-target of the recovered product must be considered in the DS selection process [57]. First on the list and most commonly used as DS is NaCl, even in high concentrations up to 4 M, due to its high aqueous solubility, small molecular size, high availability, and relatively low cost [58]. As shown in Table 5, the 0.5 M concentration is most frequently applied, as it simulates the osmotic pressure of seawater [53]. The ultimate goal is to adopt a circular solution, by applying an abundant water source without any economic burden or a process’ by-product, such as the RO brine as DS (Table 6) [58, 65]. High rejection rates of TOC/COD and PO4-P have been reported using NaCl as DS in OMBR systems, equal to 100% and 95.6%, respectively, although the same is not achieved for ammonium nitrogen (NH4-N), which in most studies ranges between 43 and 90% [25, 59]. Nevertheless, the biggest challenge in OMBR systems using NaCl as DS is the accumulation of salts in the concentrated stream of mixed liquor and the subsequent negative effect on bacterial growth, due to reverse sodium leakage [12]. Relevant mitigation measures of the above obstacles have been proposed, such as the reduction of sludge retention time (SRT), but also the application of hybrid solutions, such as MF and UF membranes downstream for the parallel recovery of phosphorus [32].

DrawFeedConfiguration/ Membrane TypeFindingsRef.
0.5, 1 & 1.5 M NaClMunicipal WastewaterAn-OMBR/CTA-ES (HTI,USA)
  • Salinity accumulation, 1 M NaCl was advantageous

[40]
0.5 M NaClMunicipal WastewaterDirect FO/ Spiral Wound CTA (HTI,USA)
  • High contaminant rejection rates, except for ammonia

[17]
0.5 M NaClSynthetic WastewaterOMBR/ CTA (HTI, USA)
  • Accumulation of nutrients and salts in the OMBR

[59]
3.5 g L−1 NaClDomestic wastewaterDirect FO (coupled with MD)/ CTA (HTI,USA)
  • Fouling of organic substances such as proteins and polysaccharides, resulting in reduced Jw

[60]
0.5 M NaClSynthetic municipal wastewaterOMBR/ Biomimetic (AQUAPORIN Asia, Singapore) TFC & CTA (HTI,USA)
  • Aquaporin FO membrane showed better performance than CTA & TFC, in terms of salinity accumulation

[27]
3 M NaClSynthetic wastewaterPost FO/ TFC-ES (HTI,USA)
  • Jw decline due to the dense adsorption layer and the gel layer formed by the deposition of carbohydrates and proteins

[61]
30 g L−1 NaClAlgae effluentAlgae - hybrid FO-RO system/ TFC Porifera Inc. (California, USA)
  • The FO rejected organic, multivalent cations and anions, providing an effective pretreatment for the RO system

[62]
0.2, 0.5, 1, 1.5, 2, 3 & 4 M NaClMunicipal wastewaterDirect FO/ CTA-ES (HTI,USA)
  • Disproportionate concentration between contaminants and water

[46]
  • Low fouling propensity, with layers formed by humic acid, protein, and polysaccharide

[46]
1 M NaClSynthetic municipal wastewaterOMBR/ TFC (HTI,USA)
  • NH4-N rejection rate was low

[25]
  • The water flux decreased with increasing salinity in the mixed liquor

[25]
70 g L−1 NaClRaw anaerobic centrateDirect FO-RO/CTA (HTI,USA)
  • FO/RO system affected by centrate replenishment and concentration; it decreased with increased replenishment and concentrations

[63]
0.7 g L−1 NaCl & 0.7 g L−1 Na2SO4Synthetic wastewaterAnFOMBR/ CTA (HTI,USA)
  • NaCl recorded better methane ratio in the biogas produced

[64]
53 g L−1 NaCl and Industrial effluent (consisted of SO42− and NH4-N)Synthetic wastewaterOMBR/ CTA-NW
  • Industrial wastewater had higher Jw and less membrane fouling compared to NaCl, but, a higher Js was observed in the former

[65]
0.5 M NaCl & NaCl with 0.01 M ((NH4)2HPO4)MBR permeateFO-RO/TFC, FTSH2O (Sterlitech Company, USA)
  • The enhanced DS with DAP recorded higher Jw than NaCl Higher rejection of TP than NH4-N

[66]

Table 5.

Sodium chloride as DS in FO treating municipal wastewater.

DrawFeedConfiguration/ Membrane TypeFindingsRef.
BrinePrimary municipal effluent & raw municipal wastewaterDirect FO/ homemade TFC
  • Jw depends on temperature, high viscosity at low temperatures increase ICP effect

[49]
  • Reversible fouling after physical cleansing

Synthetic seawaterSynthetic feed and MBR permeateDirect FO/ TFC, CTA-ES and CTA-NW, (HTI,USA)
  • Concentration of nutrients successfully performed

[67]
  • pH plays a key role in ammonia rejection by the FO membranes (close to 7)

Brine and industrial effluent (mainly consisted of SO42− and NH4-N)Anaerobically digested sludge centrateDirect FO/ CTA-NW (HTI,USA) and Aquaporin (AQUAPORIN A/S, Denmark)
  • 2 industrial effluents successfully implemented as DS in the FO process

[68]
  • Increased NH4-N concentration in concentrated sludge due to application of DS rich in ammonium sulfate

SeawaterAnaerobically digested sludge centrateDirect FO / TFC, Porifera, Inc. (California, USA)
  • Extensive membrane fouling due to nutrient precipitation in both feed solution and membrane

[69]
  • Filtration time plays an essential role in process performance

Synthetic seawater & brineWastewater after a hydrolytic anaerobic reactorDirect FO / TFC Porifera Inc. (California, USA)
  • With seawater & brine the condensation factor can reach over 10

[70]
  • Inevitable biodegradation of VFAs in this environment

Synthetic seawater with algae strainSynthetic municipal wastewaterDirect FO / Aquaporin TFC (Sterlitech Corporation, WA)
  • Low ammonium rejection, low removal of ammonia after the application of algae at about 35%.

[48]
Seawater (0.599 M, 0.428 M & 0.770 M NaCl solutions)Anaerobically digested sludge centrateDirect FO / Aquaporin (Aquaporin A/S, Denmark)
  • Better performance at pH <9 and application of DS with low reverse salt

[63]
BrineRaw municipal wastewaterMF-FOMBR/ CTA (HTI, USA)
  • 90% recovery of phosphorus using the MF system

[12]
  • Accumulation of salts in the bioreactor is still a challenge

Table 6.

Seawater, Brine, and industrial effluents as DS in FO treating municipal wastewater.

Similar results are demonstrated in bench and pilot scale FO systems for the pre-concentration of municipal wastewater using NaCl. The bidirectional diffusion of monovalent ammonium ions from the feed to the sodium cations of DS remains a major drawback [17]. In a recent study, Yang et al. [49] demonstrated the effect of the pH parameter on low NH4-N rejection rates and suggested a functional range of less than 8 for optimized performance. More specifically, at elevated pH as the main form of ammonium nitrogen is ammonia, diffusion becomes independent of the reverse sodium leakage [49]. Alternatively, the application of divalent molecular compounds as DS (Tables 7 and 8), such as magnesium chloride (MgCl2) and magnesium sulfate (MgSO4), which are characterized by lower reverse salt transport than NaCl, is suggested in many investigations [16]. Another superiority of inorganic solutions containing Mg ions is their combination with MF-OMBR hybrid systems and the utilization of the reverse Mg flux in the mixed liquor to nutrients’ recovery, after proper pH adjustment. Although, a comparative study demonstrated that Mg transport leads to the formation of both organic and inorganic fouling in the active and support layer of the TFC membrane, correspondingly, causing a dramatic reduction in membrane flux [56]. As shown in Table 8, a highly charged compound, ethylenediamine tetraacetic acid disodium salt (EDTA 2Na) was applied as DS to remove the water from the activated sludge in a hybrid Forward Osmosis – Nanofiltration (FO-NF) system; the NF module was used for the recovery of DS. Water flux dropped rapidly after 8 operating hours (8.45 to 4.22 L m−2 h−1), mainly due to the reduction of the osmotic driving force and the formation of a cake layer on the membrane surface. It is worth noting that the reverse salt flux was equal to 0.2 g m−2 h−1, while suspended solids were concentrated from 8 g L−1 to 32 g L−1 [75].

DrawFeedConfiguration/ Membrane TypeFindingsRef.
0.5, 1, 1.5 & 2 M MgCl2Synthetic Secondary effluentPost FO / CTA-NW and CTA-ES (HTI,USA)
  • Cl diffusion was higher compared to Mg ions. 95% rejection of nutrients using 2 M MgCl2

[71]
  • Higher diffusion of Cl (about 3 times) by applying CTA-ES membrane

3000:1; 1500:1: 1000:1 MgCl2 &Triton X-144Synthetic domestic wastewaterSMB-OSMBR/ CTA-ES (HTI,USA)
  • 1.5 mM MgCl2 and 1.5 mM Triton X-114 was the best solution ratio in terms of performance

[72]
  • Biomass growth media favored the achievement of stable Jw and low membrane fouling

1 M MgCl2Synthetic municipal wastewaterOMBR-MD/ CTA - ES (HTI,USA)
  • Successful rejection of PO4-P and NH4-N & recovery in the form of struvite

[28]
  • Recovery of the DS with the MD system with a small drop in the Jw

1 M MgCl2Raw anaerobic centrateDirect FO-MD/CTA (HTI,USA)
  • High rejection of nutrients, reversible membrane fouling, potential recovery of struvite with the application of FO-MD system

[73]
48.4 g L−1 MgCl2 & 49 g L−1 NaClSynthetic municipal wastewaterOMBR, CTA-ES (HTI,USA)
  • Membrane fouling was not severe as shown by the decline in Jw with both NaCl and MgCl2

[24]
  • Accumulation of salts was observed with the use of both solutions

0.5 M NaCl & 0.35 M MgCl2Synthetic municipal wastewaterAnOMBR-MF system/ TFC (HTI,USA)
  • Zero rejection of NH4 –N using NaCl and 57.5–87.5% using MgCl2

[56]
  • MgCl2 caused severe membrane inorganic fouling

1, 1.5 & 2 M MgCl2Municipal wastewaterDirect FO, TFC (Aquaporin A/S, Denmark)
  • Both the increase in the MgCl2 concentration and increase in the cross-flow rate contributed to the higher Jw, but COD concentration remained stable

[16]

Table 7.

Magnesium chloride as DS in FO treating municipal wastewater.

DrawFeedConfiguration/ Membrane TypeFindingsRef.
0.6 M C2H3NaO2, 0.3 M EDTA-2Na & 0.5 M NaClSynthetic municipal wastewaterOMBR-RO/ CTA (HTI,USA)
  • Minimized membrane fouling, lower Js for the tested solutions over NaCl, but also lower Jw and reduced salt accumulation

[74]
0.1, 0.3, 0.5, 0.7 & 1 M EDTA-2NaActivated sludgeFO-NF/ CTA membranes (HTI,USA)
  • 0.7 M EDTA-2Na was the preferred concentration

[75]
  • At pH equal to 8, the FO membranes achieved the best performance

[75]
  • Reduced Js compared to inorganic salts

0.25 M CaCl2, 0.25 M MgCl2, 0.25 M C4H6MgO4 & 0.25 M C2H3NaO2Synthetic municipal wastewaterOMBR-MD/CTA membrane (HTI,USA)
  • C4H6MgO4 optimized sludge flocculation, C2H3NaO2 and MgCl2 achieve steady flows in repeated tests

[76]
  • Inhibition of biological activity due to Cl presence

  • C2H3NaO2 generates significant fouling

NaCl, MgSO4, C2H3NaO2, C4H6MgO4, CHNaO2, EDTA-2Na, C6H12O6, C2H5NO2, C3H8O3, CH4N2OMunicipal wastewaterDirect FO/ CTA membrane (HTI)
  • Organic solutions, such as C2H3NaO2, are appropriate for this configuration. Reverse leakage of NaCl does not interrupt the anaerobic process

[13]
CHKO2, K4P2O7, (C3H3NaO2)n, C2nH4n + 2On + 1, MgSO4 & NaClSynthetic secondary effluentPost FO/ spiral wound (SW) CTA and TFC & flat sheet TFC and CTA
  • MgSO4, (C3H3NaO2)n and K4P2O7 identified as suitable solutions in terms of cost-effectiveness, toxicity, recovery and pH range

[77]
(C3H3NaO2)n, MgSO4 & MgCl2MBR permeateFO-NF/ TFC (Porifera, CA, USA)
  • (C3H3NaO2)n was unsuitable for irrigation, MgSO4 caused prolonged membrane fouling, while MgCl2 had the best performance

[78]
Commercial fertiliser diamond blueRaw wastewater, MBR supernatant & permeateDirect FO/ TFC membrane (Toray Industry Inc.)
  • Liquid fertilizer has a good performance comparable to the application with common inorganic salts for green wall irrigation

[79]
KNO3, KH2PO4 &KNO3 (fertilizers)Synthetic secondary effluentFDFO/ CTA membrane (HTI, USA)
  • Occurrence of severe biofouling using KNO3 as DS compared to KCl and KH2PO4; membrane flux decline by 63%, 45% and 30%, respectively

[80]
NH4NO3, (NH4)2SO4, NH4Cl, Ca(NO3)2, KCl, NH4H2PO4, (NH4)2HPO4, KNO3 & KH2PO4Synthetic municipal wastewaterFDFO/ TFC membrane, (Toray Industry Inc.)
  • KCl and NH4Cl showed the highest water recovery and MAP, KH2PO4 and SOA showed the lowest Js

[55]
  • High dilution is required about 100 tunes

(NH4)2SO4, KH2PO4 & (NH4)H2PO4Synthetic municipal wastewaterFO-MBR (Direct FO)/CTA-ES membrane (HTI)
  • MAP had the best performance with the lowest Js

[51]
  • SOA altered the sludge characteristics

Commercial liquid fertilizerRaw wastewaterDirect FO, CTA membrane (HTI,USA)
  • Effective application of liquid fertilizer as DS for green wall irrigation

[81]
  • Reverse nutrient leakage worsened as the temperature rose

KH2PO4, KCl & KNO3Synthetic municipal wastewaterFDFO-AnMBR
  • KH2PO4 < KCl < KNO3 in terms of reverse leakages; alterations observed in anaerobic biomass, especially using KNO3

[82]

Table 8.

Fertilizers, organic, inorganic, and ionic organic compounds as DS in FO treating municipal wastewater.

To enhance the valorization of the resources contained in municipal wastewater, through the application of the anaerobic process several organic and ionic organic draw solutions have been investigated [13, 14, 74]. Bowden et al. [14] compared 10 different ionic organic compounds as DS and slightly altered the selection methodology proposed by Achilli et al. [15], introducing the parameter of biodegradability of the DS in the protocol. A bench-scale FO unit was used, while CTA membranes (Hydration Technology Innovations, HTI, USA) were applied to all experiments; the main purpose of this study was to evaluate the applicability of ionic organic solutions to OMBR systems.

Magnesium acetate (C4H6MgO4) and sodium propionate (C3H5NaO2) recorded the best performance as DS in terms of Js, potential recovery, and biodegradability. Siddique et al. [76] showed similar results with the application of synthetic wastewater, highlighting C4H6MgO4 as suitable DS for OMBR applications, while sodium acetate (C2H3NaO2) led to the development of dense membrane biofilm. Despite the many benefits of ionic organic solutions, it should be noted that their potential application is limited, as the re-concentration cost is high compared to inorganic solutions.

A recent study aimed to integrate all the parameters of Table 4 with the compatibility of FO as a pre-treatment step preceding the anaerobic process [83]. Among the 5 different zwitterions solutions tested, glycine (C2H5NO2), L-proline (C5H9NO2), and glycine betaine (C5H11NO2) exhibited comparable Jw to NaCl (4.3–4.9 L m−2 h−1), with lower Js. From a physicochemical perspective, the process efficiency depends significantly on the pH value, affecting both the charge and the molecular size. Despite the rapid biodegradation (Adenosine triphosphate (ATP) levels range from 7 to 14 μg L−1 after degradation tests) of all zwitterions compounds, the replacement cost, which is 3–4 times more than the cost of commercially available solutions, is a potential barrier to their implementation in municipal wastewater streams. It is worth noting that the above experiments were performed with deionized water as feed, which favors the overall performance over the application of a more complex ionic matrix, such as sewage [83].

Commercial fertilizers are another largely inorganic solution medium that has been tested in various effluents resulting from a WWTP, such as typical secondary and MBR permeate and raw municipal wastewater. As illustrated in Table 8. Li et al. [82] compared the effect of 3 different commercial fertilizers on the downstream anaerobic process when applied as draw agents directly in raw wastewater. The following order of compatibility with the anaerobic treatment revealed Potassium Nitrate (KNO3) > Potassium Chloride (KCl) > Potassium dihydrogen Phosphate (KH2PO4), with their reverse solute flux showing a similar sequence when the concentration of all DS was equal to 1 M. Water flux can be dramatically reduced by applying KNO3 as DS, as extensive biofouling has been observed, while increasing nitrate concentrations can inhibit the subsequent anaerobic process, rendering them as unsuitable [80]. The implementation of different fertilizers in a hybrid FO-RO system to concentrate MBR permeate proved that the amplification of enhanced NaCl with Diammonium Phosphate (DAP) ((NH4)2HPO4) can reduce reverse solute leakage by 35%, achieving ΝΗ4-Ν rejection rates more than 95% at different flow rates (1.2 and 2 L m−2 h−1) [66]. In addition, a long-term study examining the pilot application of a hybrid FO-NF system that treated MBR permeate found that Sodium Polyacrylate ((C3H3NaO2)n) was inappropriate for irrigation practices. On the contrary, the combination of MgCl2 with NF membranes significantly improved the process efficiency and operating costs, as the application of chemical cleaning was not required. However, a notably high loss of the osmotic agent was observed [78]. A particularly interesting investigation was carried out by Adnan et al. [51] in which the possibility of applying 9 different fertilizers to the direct FO for the wastewater valorization and its parallel application in agricultural practices was examined. Water recovery was high by applying KCl (Jw = 21.1 L m−2 h−1; Js = 11.2 g m−2 h−1; Osmotic Pressure (OP) = 44.6 bar) and Ammonium Chloride (NH4Cl) (Jw = 21.1 L m−2 h−1; Js = 7.5 g m−2 h−1; OP = 43.5 bar), while other fertilizers recorded particularly low reverse flux, such as Ammonium Sulfate (SOA) ((NH4)2SO4) (Jw = 15.5 L m−2 h−1; Js = 1.7 g m−2 h−1; OP = 46.7 bar), KH2PO4 (Jw = 13.2 L m−2 h−1; Js = 2.3 g m−2 h−1; OP = 36.5 bar), and NH4H2PO4 (Monoammonium Phosphate, MAP) (Jw = 13.8 L m−2 h−1; Js = 1 g m−2 h−1; OP = 44.4 bar). However, this process becomes inapplicable, as a large amount of water is required to dilute the concentrated fertilizer (at least 1/100), to reach the irrigation limits [51].

The analysis of the existing literature makes it clear that the FO process is still under investigation and the determination of the ideal DS plays a vital role in upgrading the process of this technology. Despite the properties of the DS, the selection of the suitable configuration, the techno-economic factors, and the re-condensation method should be combined during the selection process; the optimization of the FO membrane’s properties is a major challenge that can solve many issues. The development and fabrication of higher rejection membranes can be the answer to the implementation of both monovalent and divalent ions, which have been widely used as DS and their performance is already known to the research community.

Advertisement

4. Integration of FO followed by anaerobic treatment in a WWTP

4.1 COD valorization in municipal WWTPs

For more than a century, the CAS process has been applied as the main urban wastewater treatment system worldwide, making a significant contribution to environmental protection and public health. However, the low energy efficiency of the CAS process ranks WWTPs among the largest energy consumers in a country; on an annual basis, in developed counties, about 1–3% of electricity consumption is spent on their operation [84]. In addition, WWTPs are characterized by a high energy and carbon footprint, as during biological processes, large amounts of greenhouse gases are produced, mainly carbon dioxide generated due to the oxidation of organic matter and indirectly by electricity consumption [85]. Therefore, about 0.3–0.5 kWh m−3 of energy is required for sewage treatment by applying the CAS process, while the contained chemical energy and nutrients are not utilized [86].

According to Wan et al. [87] the traditional CAS process needs an average of 0.45 kWh to treat one m3 of sewage, which equals to 1620 kJ m−3. Assuming a concentration of 600 mg L−1 COD, energy consumption becomes 2.7 kJ g−1 COD. As shown in Figure 2(A), the energy recovery in convectional CAS systems occurs through the anaerobic digestion of the primary and secondary sludge, which corresponds to 32–39% of the organic material in COD terms. The latter percentage is equal to 2.9–3.5 kJ g−1 COD, since 1 g of methane-COD is equal to 13.9 kJ (65% methane percent in produced biogas). Considering that only 35% of the produced methane can be utilized for the production of electricity [86], about 1–1.2 kJ g−1 COD can be recovered from municipal wastewater, by applying anaerobic digestion to the sludge treatment line. Comparing the aforementioned energy requirement, 2.7 kJ g−1 COD, it is estimated that about 40% of it can be recovered using anaerobic digestion (1–1.2 kJ g−1 COD). The anaerobic digestion process also generates approximately 50–55% heat, part of which is used to heat the digesters. The excess heat can only be valorized locally [88].

Figure 2.

(A) COD mass flow in a convectional WWTP, (B) COD mass flow, when FO followed by anaerobic treatment is integrated into a WWTP.

Obviously, COD capture, and subsequently valorization of the chemical energy contained in municipal wastewater can lead WWTPs to sustainable development, transforming WWTPs from energy consumers to producers, while significantly reducing the environmental footprint and operating costs.

The integration of FO in municipal wastewater treatment and the benefits of its application have been investigated in various studies [17, 49]. This chapter presents the combination of FO and anaerobic treatment in a typical WWTP for the utilization of the chemical energy, which is inherently present in sewage. As shown in Figure 2(B), by placing the FO in the main treatment line of a WWTP and taking into account the efficiency of a typical anaerobic system, such as An-MBR, which is equal to 80% in ambient conditions [89], 46–55% of COD is converted to biogas (65% of the aforementioned percent corresponds to methane). Following the same procedure as before, the energy recovery in the main treatment line through the implementation of anaerobic treatment is between 4.2–5 kJ g−1 COD. Another 1.3–1.6 kJ g−1 COD of energy is recovered from the anaerobic digestion of the sewage sludge (13.9 kJ g−1 methane-COD). Since only 35% of the produced methane can be converted into electricity [86], the power production from the wastewater treatment line ranges between 1.3–1.7 kJ g−1 COD, while from the sludge treatment line it is equal to 0.4–0.6 kJ g−1 COD. On aggregate, 1.9–2.3 kJ g−1 COD of electricity can be utilized from this innovative treatment scheme, which can counterbalance 80% of the existing energy consumption of a typical municipal WWTP. The treated effluent of the anaerobic system is rich in nutrients, which can be valorized by applying recovery technologies for the production of slow-release fertilizers, while the reclaimed water content can also be reused.

4.2 Salinity, the greatest impact of FO on anaerobic treatment

Despite the benefits of the wastewater management system presented in the above section, there are two factors that can be particularly limiting to the subsequent operation of the anaerobic process. The solute flux that characterized the FO system results in the accumulation of salts in the feed stream, potentially resulting in partial or complete inhibition of the downstream anaerobic and aerobic biological treatment processes [14, 17, 32]. Salinity has been identified as an inhibitory agent of the anaerobic process, as the increased osmotic pressure across the cell membrane can cause plasmolysis, leading to cell death and total inhibition of the anaerobic process. More specifically, Lefebvre et al. [90] stressed that the activity of methanogenic bacteria is inhibited at concentrations of NaCl equal to 5 g L−1, while acidogenic microorganisms are affected at much higher concentrations, i.e. 20 g L−1. Ansari et al. [91] studied the effects of NaCl on anaerobic treatment of concentrated wastewater effluents in batch mode experiments and observed that by increasing water recovery rates of FO (from 50 to 90%), the anaerobic process achieves higher methane production (approximately 5 times higher), while the presence of salinity has a negligible negative effect.

Based on the existing literature, the limiting parameter of salinity has been investigated and observed only in aerobic/anaerobic systems, where the FO unit is plugged into MBR systems for a relatively short time, while in pre-concentration systems few studies have examined the effect of salinity on the downstream anaerobic process and suggest mitigation measures. Chen et al. [37] and Wang et al. [39] did not observe significant effects of salinity on anaerobic reactors by recording an average methane yield of 0.2 and 0.3 L CH4 g−1 COD, respectively, in studies that cannot be characterized as long-term. As mentioned above, the application of minimization strategies such as the corresponding regulation of the hydraulic residence time (HRT) seems to regulate the salinity conditions to which the biomass is exposed. Accordingly, the addition of MF membranes is a particular interesting approach for the minimization of salinity and the parallel application of nutrient recovery methods. Another interesting perspective is the acclimatization of the anaerobic biomass to high salinity conditions. This mitigation technique is not recent as the presence of specific microorganisms, such as halotolerant bacteria has shown particularly high efficiency in the anaerobic treatment of saline industrial wastewater [92]. In a recent study, where no acclimatized biomass was used, Gao et al. [93] separately investigated the effect of high salinity and ammonia nitrogen concentration and the combination of the two inhibitors in the anaerobic treatment of pre-concentrated municipal wastewater. The results showed that the presence of NH4-N and NaCl concentrations separately, up to 200 mg L−1 and between 5 and 8 g L−1, respectively, did not significantly affect the activity of anaerobic microorganisms. The combination of the two parameters in non-acclimatized and acclimatized biomass showed that the latter had significantly better performance and can respond without the risk of inhibition. Further research into anaerobic biomass acclimatization should be conducted in the future, as higher condensation rates could be applied from the upstream FO unit.

All the acquired knowledge of the above studies would be particularly interesting to be used in the long-term investigation of a FO system combined with a downstream anaerobic process, in which all the limiting parameters and the proposed mitigation measures can be examined in-depth, for the rational assessment of its performance.

Advertisement

5. Conclusion

There is no doubt that FO is a promising technology that has been investigated for a range of applications at various stages of a municipal WWTP. Among them, its combination with the anaerobic process has significant advantages, as much of the chemical energy inherently contained in sewage can be recovered as biogas, while resource recovery technologies can be applied downstream, utilizing the nutrient-rich effluent. However, the transition of the FO from laboratory scale to full-scale applications requires further research to address important issues, such as the salinity accumulation in the downstream technologies and the reduced rejection of NH4-N by existing FO membranes. The application of NaCl indicates a possible suitability for the concentration of municipal wastewater. The background knowledge available on the basic criteria of FO has to be utilized for the development of membranes with higher selectivity. Future investigations should carry out extensive long-term monitoring and targeted combination/interaction of different parameters for the concentration of real wastewater, to assess from a technical, environmental and economic perspective the feasibility of applying FO technology to municipal wastewater management.

Advertisement

Acknowledgments

This research work was supported by the Hellenic Foundation for Research and Innovation (H.F.R.I.) under the “First Call for H.F.R.I. Research Projects to support Faculty members and Researchers and the procurement of high-cost research equipment grant” (Project Number: 4008).

Advertisement

Nomenclature

ATPadenosine triphosphate
NH4-Nammonium nitrogen
SOAammonium sulfate
An-OMBRanaerobic osmotic membrane bioreactor
CTAcellulose triacetate
CODchemical oxygen demand
CASconventional activated sludge
DAPdiammonium phosphate
DSdraw solution
EDTA 2Naethylenediamine tetraacetic acid disodium salt
EUEuropean Union
FOforward osmosis
FO-NFforward osmosis-nanofiltration
C2H5NO2glycine
C5H11NO2glycine betaine
HTIhydration technology innovations
HRThydraulic residence time
C5H9NO2L-proline
C4H6MgO4magnesium acetate
MgCl2magnesium chloride
MgSO4magnesium sulfate
MBRmembrane bioreactor
MFmicrofiltration
MF-OMBRmicrofiltration-osmotic membrane bioreactor
MAPmonoammonium phosphate
NFnanofiltration
PO4-Porthophosphate as phosphorus
OMBRosmotic membrane bioreactor
KClpotassium chloride
KH2PO4potassium dihydrogen phosphate
KNO3potassium nitrate
ROreverse osmosis
Jsreverse salt flux
SRTsludge retention time
C2H3NaO2sodium acetate
NaClsodium chloride
(C3H3NaO2)nsodium polyacrylate
C3H5NaO2sodium propionate
TFCthin film composite
TOCtotal organic cabon
UFultrafiltration
WWTP(s)wastewater treatment plant(s)
Jwwater flux

References

  1. 1. Bond NR, Burrows RM, Kennard MJ, Bunn SE. Water scarcity as a driver of multiple stressor effects. Elsevier Inc.; 2018. DOI:10.1016/B978-0-12-811713-2.00006-6
  2. 2. Verner D. Climate Change Contributes to Water Scarcity. In: Verner D, editor. Adaptation to a Changing Climate in the Arab Countries, 2012, p. 110–51. DOI:10.1596/978-0-8213-9459-5
  3. 3. European Commission (December 2nd 2015). Closing the loop—An EU Action Plan for the Circular Economy. Communication from the Commission to the European Parliament, the Council, the European Economic and Social Committee and the Committee of the Regions COM (2015), 614, Brussels. n.d
  4. 4. Escobar IC. Chapter 14 Conclusion: A Summary of Challenges still Facing Desalination and Water Reuse. vol. 2. Elsevier; 2010. DOI:10.1016/S1871-2711(09)00214-1
  5. 5. Ge Q, Ling M, Chung TS. Draw solutions for forward osmosis processes: Developments, challenges, and prospects for the future. Journal of Membrane Science. 2013;442:225–37. DOI:10.1016/j.memsci.2013.03.046
  6. 6. Lee KP, Arnot TC, Mattia D. A review of reverse osmosis membrane materials for desalination-Development to date and future potential. Journal of Membrane Science. 2011;370:1–22. DOI:10.1016/j.memsci.2010.12.036
  7. 7. Sun Y, Tian J, Zhao Z, Shi W, Liu D, Cui F. Membrane fouling of forward osmosis (FO) membrane for municipal wastewater treatment: A comparison between direct FO and OMBR. Water Research. 2016;104:330–9. DOI:10.1016/j.watres.2016.08.039
  8. 8. Zhao S, Zou L, Tang CY, Mulcahy D. Recent developments in forward osmosis: Opportunities and challenges. Journal of Membrane Science. 2012;396:1–21. DOI:10.1016/j.memsci.2011.12.023
  9. 9. Chung TS, Li X, Ong RC, Ge Q, Wang H, Han G. Emerging forward osmosis (FO) technologies and challenges ahead for clean water and clean energy applications. Current Opinion in Chemical Engineering. 2012;1:246–57. DOI:10.1016/j.coche.2012.07.004
  10. 10. Ang WL, Wahab Mohammad A, Johnson D, Hilal N. Forward osmosis research trends in desalination and wastewater treatment: A review of research trends over the past decade. Journal of Water Process Engineering. 2019;31:100886. DOI:10.1016/j.jwpe.2019.100886
  11. 11. Ansari AJ, Hai FI, Guo W, Ngo HH, Price WE, Nghiem LD. Factors governing the pre-concentration of wastewater using forward osmosis for subsequent resource recovery. Science of the Total Environment. 2016;566–567:559–66. DOI:10.1016/j.scitotenv.2016.05.139
  12. 12. Qiu G, Law YM, Das S, Ting YP. Direct and complete phosphorus recovery from municipal wastewater using a hybrid microfiltration-forward osmosis membrane bioreactor process with seawater brine as draw solution. Environmental Science and Technology. 2015;49:6156–63. DOI:10.1021/es504554f
  13. 13. Ansari AJ, Hai FI, Guo W, Ngo HH, Price WE, Nghiem LD. Selection of forward osmosis draw solutes for subsequent integration with anaerobic treatment to facilitate resource recovery from wastewater. Bioresource Technology. 2015;191:30–6. DOI:10.1016/j.biortech.2015.04.119
  14. 14. Bowden KS, Achilli A, Childress AE. Organic ionic salt draw solutions for osmotic membrane bioreactors. Bioresource Technology. 2012;122:207–16. DOI:10.1016/j.biortech.2012.06.026
  15. 15. Achilli A, Cath TY, Childress AE. Selection of inorganic-based draw solutions for forward osmosis applications. Journal of Membrane Science. 2010;364:233–41. DOI:10.1016/j.memsci.2010.08.010
  16. 16. Singh N, Dhiman S, Basu S, Balakrishnan M, Petrinic I, Helix-Nielsen C. Dewatering of sewage for nutrients and water recovery by Forward Osmosis (FO)using divalent draw solution. Journal of Water Process Engineering. 2019;31:100853. DOI:10.1016/j.jwpe.2019.100853
  17. 17. Wang Z, Zheng J, Tang J, Wang X, Wu Z. A pilot-scale forward osmosis membrane system for concentrating low-strength municipal wastewater: Performance and implications. Scientific Reports. 2016;6:1–11. DOI:10.1038/srep21653
  18. 18. Korenak J, Basu S, Balakrishnan M, Hélix-Nielsen C, Petrinic I. Forward osmosis in wastewater treatment processes. Acta Chimica Slovenica. 2017;64:83–94. DOI:10.17344/acsi.2016.2852
  19. 19. Cornelissen ER, Harmsen D, de Korte KF, Ruiken CJ, Qin JJ, Oo H, Wessels LP. Membrane fouling and process performance of forward osmosis membranes on activated sludge. Journal of Membrane Science. 2008;319:158–68. DOI:10.1016/j.memsci.2008.03.048
  20. 20. Wang X, Zhao Y, Yuan B, Wang Z, Li X, Ren Y. Comparison of biofouling mechanisms between cellulose triacetate (CTA) and thin-film composite (TFC) polyamide forward osmosis membranes in osmotic membrane bioreactors. Bioresource Technology. 2016;202:50–8. DOI:10.1016/j.biortech.2015.11.087
  21. 21. Holloway RW, Regnery J, Nghiem LD, Cath TY. Removal of trace organic chemicals and performance of a novel hybrid ultrafiltration-osmotic membrane bioreactor. Environmental Science and Technology. 2014;48:10859–68. DOI:10.1021/es501051b
  22. 22. Qiu G, Ting YP. Osmotic membrane bioreactor for wastewater treatment and the effect of salt accumulation on system performance and microbial community dynamics. Bioresource Technology. 2013;150:287–97. DOI:10.1016/j.biortech.2013.09.090
  23. 23. Sun Y, Tian J, Song L, Gao S, Shi W, Cui F. Dynamic changes of the fouling layer in forward osmosis based membrane processes for municipal wastewater treatment. Journal of Membrane Science. 2018;549:523–32. DOI:10.1016/j.memsci.2017.12.055
  24. 24. Qiu G, Ting YP. Direct phosphorus recovery from municipal wastewater via osmotic membrane bioreactor (OMBR) for wastewater treatment. Bioresource Technology. 2014;170:221–9. DOI:10.1016/j.biortech.2014.07.103
  25. 25. Huang L ying, Lee DJ, Lai JY. Forward osmosis membrane bioreactor for wastewater treatment with phosphorus recovery. Bioresource Technology. 2015;198:418–23. DOI:10.1016/j.biortech.2015.09.045
  26. 26. Luo W, Hai FI, Price WE, Guo W, Ngo HH, Yamamoto K, Nghiem LD. Phosphorus and water recovery by a novel osmotic membrane bioreactor-reverse osmosis system. Bioresource Technology. 2016;200:297–304. DOI:10.1016/j.biortech.2015.10.029
  27. 27. Luo W, Xie M, Song X, Guo W, Ngo HH, Zhou JL, Nghiem LD. Biomimetic aquaporin membranes for osmotic membrane bioreactors: Membrane performance and contaminant removal. Bioresource Technology. 2018;249:62–8. DOI:10.1016/j.biortech.2017.09.170
  28. 28. Chang HM, Chen SS, Nguyen NC, Chang WS, Sinha Ray S. Osmosis membrane bioreactor–microfiltration with magnesium-based draw solute for salinity reduction and phosphorus recovery. International Biodeterioration and Biodegradation. 2017;124:169–75. DOI:10.1016/j.ibiod.2017.05.022
  29. 29. Xu Z, Song X, Xie M, Wang Y, Huda N, Li G, Luo W. Effects of surfactant addition to draw solution on the performance of osmotic membrane bioreactor. Journal of Membrane Science. 2021;618:118634. DOI:10.1016/j.memsci.2020.118634
  30. 30. Ab Hamid NH, Wang DK, Smart S, Ye L. Achieving stable operation and shortcut nitrogen removal in a long-term operated aerobic forward osmosis membrane bioreactor (FOMBR) for treating municipal wastewater. Chemosphere. 2020;260. DOI:10.1016/j.chemosphere.2020.127581
  31. 31. Yap WJ, Zhang J, Lay WCL, Cao B, Fane AG, Liu Y. State of the art of osmotic membrane bioreactors for water reclamation. Bioresource Technology. 2012;122:217–22. DOI:10.1016/j.biortech.2012.03.060
  32. 32. Wang X, Chang VWC, Tang CY. Osmotic membrane bioreactor (OMBR) technology for wastewater treatment and reclamation: Advances, challenges, and prospects for the future. Journal of Membrane Science. 2016;504:113–32. DOI:10.1016/j.memsci.2016.01.010
  33. 33. Christian S, Grant S, McCarthy P, Wilson D, Mills D. The first two years of full-scale anaerobic membrane bioreactor (AnMBR) operation treating high-strength industrial wastewater. Water Practice and Technology. 2011;6. DOI:10.2166/wpt.2011.032
  34. 34. Ansari AJ, Hai FI, Price WE, Drewes JE, Nghiem LD. Forward osmosis as a platform for resource recovery from municipal wastewater - A critical assessment of the literature. Journal of Membrane Science. 2017;529:195–206. DOI:10.1016/j.memsci.2017.01.054
  35. 35. Zhang H, Jiang W, Cui H. Performance of anaerobic forward osmosis membrane bioreactor coupled with microbial electrolysis cell (AnOMEBR) for energy recovery and membrane fouling alleviation. Chemical Engineering Journal. 2017;321:375–83. DOI:10.1016/j.cej.2017.03.134
  36. 36. Gu Y, Chen L, Ng JW, Lee C, Chang VWC, Tang CY. Development of anaerobic osmotic membrane bioreactor for low-strength wastewater treatment at mesophilic condition. Journal of Membrane Science. 2015;490:197–208. DOI:10.1016/j.memsci.2015.04.032
  37. 37. Chen L, Gu Y, Cao C, Zhang J, Ng JW, Tang C. Performance of a submerged anaerobic membrane bioreactor with forward osmosis membrane for low-strength wastewater treatment. Water Research. 2014;50:114–23. DOI:10.1016/j.watres.2013.12.009
  38. 38. Chang HM, Sun YC, Chien IC, Chang WS, Ray SS, Cao DTN, Cong Duong C, Chen SS. Innovative upflow anaerobic sludge osmotic membrane bioreactor for wastewater treatment. Bioresource Technology. 2019;287:121466. DOI:10.1016/j.biortech.2019.121466
  39. 39. Wang X, Wang C, Tang CY, Hu T, Li X, Ren Y. Development of a novel anaerobic membrane bioreactor simultaneously integrating microfiltration and forward osmosis membranes for low-strength wastewater treatment. Journal of Membrane Science. 2017;527:1–7. DOI:10.1016/j.memsci.2016.12.062
  40. 40. Gao Y, Fang Z, Chen C, Zhu X, Liang P, Qiu Y, Zhang X, Huang X. Evaluating the performance of inorganic draw solution concentrations in an anaerobic forward osmosis membrane bioreactor for real municipal sewage treatment. Bioresource Technology. 2020;307:123254. DOI:10.1016/j.biortech.2020.123254
  41. 41. Hou D, Lu L, Ren ZJ. Microbial fuel cells and osmotic membrane bioreactors have mutual benefits for wastewater treatment and energy production. Water Research. 2016;98:183–9. DOI:10.1016/j.watres.2016.04.017
  42. 42. Hou D, Lu L, Sun D, Ge Z, Huang X, Cath TY, Ren ZJ. Microbial electrochemical nutrient recovery in anaerobic osmotic membrane bioreactors. Water Research. 2017;114:181–8. DOI:10.1016/j.watres.2017.02.034
  43. 43. Holloway RW, Wait AS, Fernandes da Silva A, Herron J, Schutter MD, Lampi K, Cath TY. Long-term pilot scale investigation of novel hybrid ultrafiltration-osmotic membrane bioreactors. Desalination. 2015;363:64–74. DOI:10.1016/j.desal.2014.05.040
  44. 44. Luo W, Xie M, Hai FI, Price WE, Nghiem LD. Biodegradation of cellulose triacetate and polyamide forward osmosis membranes in an activated sludge bioreactor: Observations and implications. Journal of Membrane Science. 2016;510:284–92. DOI:10.1016/j.memsci.2016.02.066
  45. 45. Zhang X, Ning Z, Wang DK, Diniz da Costa JC. Processing municipal wastewaters by forward osmosis using CTA membrane. Journal of Membrane Science. 2014;468:269–75. DOI:10.1016/j.memsci.2014.06.016
  46. 46. Gao Y, Fang Z, Liang P, Huang X. Direct concentration of municipal sewage by forward osmosis and membrane fouling behavior. Bioresource Technology. 2018;247:730–5. DOI:10.1016/j.biortech.2017.09.145
  47. 47. Ferrari F, Pijuan M, Rodriguez-Roda I, Blandin G. Exploring submerged forward osmosis for water recovery and pre-concentration of wastewater before anaerobic digestion: A pilot scale study. Membranes. 2019;9. DOI:10.3390/membranes9080097
  48. 48. Rood B, Zhang C, Inniss E, Hu Z. Forward osmosis with an algal draw solution to concentrate municipal wastewater and recover resources. Water Environment Research. 2020;92:689–97. DOI:10.1002/wer.1262
  49. 49. Yang S, Gao B, Jang A, Shon H kyong, Yue Q. Municipal wastewater treatment by forward osmosis using seawater concentrate as draw solution. Chemosphere. 2019;237:124485. DOI:10.1016/j.chemosphere.2019.124485
  50. 50. Valladares Linares R, Li Z, Abu-Ghdaib M, Wei CH, Amy G, Vrouwenvelder JS. Water harvesting from municipal wastewater via osmotic gradient: An evaluation of process performance. Journal of Membrane Science. 2013;447:50–6. DOI:10.1016/j.memsci.2013.07.018
  51. 51. Adnan M, Khan SJ, Manzoor K, Hankins NP. Performance evaluation of fertilizer draw solutions for forward osmosis membrane bioreactor treating domestic wastewater. Process Safety and Environmental Protection. 2019;127:133–40. DOI:10.1016/j.psep.2019.05.006
  52. 52. Blandin G, Ferrari F, Lesage G, Le-Clech P, Héran M, Martinez-Lladó X. Forward osmosis as concentration process: Review of opportunities and challenges. Membranes. 2020;10:1–40. DOI:10.3390/membranes10100284
  53. 53. Chekli L, Phuntsho S, Shon HK, Vigneswaran S, Kandasamy J, Chanan A. A review of draw solutes in forward osmosis process and their use in modern applications. Desalination and Water Treatment. 2012;43:167–84. DOI:10.1080/19443994.2012.672168
  54. 54. Johnson DJ, Suwaileh WA, Mohammed AW, Hilal N. Osmotic ’ s potential : An overview of draw solutes for forward osmosis. Desalination. 2018;434:100–20
  55. 55. Chekli L, Kim Y, Phuntsho S, Li S, Ghaffour N, Leiknes TO, Shon HK. Evaluation of fertilizer-drawn forward osmosis for sustainable agriculture and water reuse in arid regions. Journal of Environmental Management. 2017;187:137–45. DOI:10.1016/j.jenvman.2016.11.021
  56. 56. Hu T, Wang X, Wang C, Li X, Ren Y. Impacts of inorganic draw solutes on the performance of thin-film composite forward osmosis membrane in a microfiltration assisted anaerobic osmotic membrane bioreactor. RSC Advances. 2017;7:16057–63. DOI:10.1039/C7RA01524K
  57. 57. Long Q, Jia Y, Li J, Yang J, Liu F, Zheng J, Yu B. Recent advance on draw solutes development in Forward osmosis. Processes. 2018;6:7–11. DOI:10.3390/pr6090165
  58. 58. Lutchmiah K, Verliefde ARD, Roest K, Rietveld LC, Cornelissen ER. Forward osmosis for application in wastewater treatment: A review. Water Research. 2014;58:179–97. DOI:10.1016/j.watres.2014.03.045
  59. 59. Aftab B, Khan SJ, Maqbool T, Hankins NP. High strength domestic wastewater treatment with submerged forward osmosis membrane bioreactor. Water Science and Technology. 2015;72:141–8. DOI:10.2166/wst.2015.195
  60. 60. Li J, Hou D, Li K, Zhang Y, Wang J, Zhang X. Domestic wastewater treatment by forward osmosis-membrane distillation (FO-MD) integrated system. Water Science and Technology. 2018;77:1514–23. DOI:10.2166/wst.2018.031
  61. 61. Motsa MM, Mamba BB, Verliefde ARD. Forward osmosis membrane performance during simulated wastewater reclamation: Fouling mechanisms and fouling layer properties. Journal of Water Process Engineering. 2018;23:109–18. DOI:10.1016/j.jwpe.2018.03.007
  62. 62. Jiang W, Lin L, Gedara SMH, Schaub TM, Jarvis JM, Wang X, Xu X, Nirmalakhandan N, Xu P. Potable-quality water recovery from primary effluent through a coupled algal-osmosis membrane system. Chemosphere. 2020;240:124883. DOI:10.1016/j.chemosphere.2019.124883
  63. 63. Kedwell KC, Christensen ML, Quist-Jensen CA, Jørgensen MK. Effect of reverse sodium flux and pH on ammoniacal nitrogen transport through biomimetic membranes. Separation and Purification Technology. 2019;217:40–7. DOI:10.1016/j.seppur.2019.02.001
  64. 64. Yin Tang MK, Ng HY. Impacts of different draw solutions on a novel anaerobic forward osmosis membrane bioreactor (AnFOMBR). Water Science and Technology. 2014;69:2036–42. DOI:10.2166/wst.2014.116
  65. 65. Luján-Facundo MJ, Soler-Cabezas JL, Mendoza-Roca JA, Vincent-Vela MC, Bes-Piá A, Doñate-Hernández S. A study of the osmotic membrane bioreactor process using a sodium chloride solution and an industrial effluent as draw solutions. Chemical Engineering Journal. 2017;322:603–10. DOI:10.1016/j.cej.2017.04.062
  66. 66. Hafiz MA, Hawari AH, Altaee A. A hybrid forward osmosis/reverse osmosis process for the supply of fertilizing solution from treated wastewater. Journal of Water Process Engineering. 2019;32. DOI:10.1016/j.jwpe.2019.100975
  67. 67. Xue W, Tobino T, Nakajima F, Yamamoto K. Seawater-driven forward osmosis for enriching nitrogen and phosphorous in treated municipal wastewater: Effect of membrane properties and feed solution chemistry. Water Research. 2015;69:120–30. DOI:10.1016/j.watres.2014.11.007
  68. 68. Soler-Cabezas JL, Mendoza-Roca JA, Vincent-Vela MC, Luján-Facundo MJ, Pastor-Alcañiz L. Simultaneous concentration of nutrients from anaerobically digested sludge centrate and pre-treatment of industrial effluents by forward osmosis. Separation and Purification Technology. 2018;193:289–96. DOI:10.1016/j.seppur.2017.10.058
  69. 69. Vu MT, Price WE, He T, Zhang X, Nghiem LD. Seawater-driven forward osmosis for pre-concentrating nutrients in digested sludge centrate. Journal of Environmental Management. 2019;247:135–9. DOI:10.1016/j.jenvman.2019.06.082
  70. 70. Blandin G, Rosselló B, Monsalvo VM, Batlle-Vilanova P, Viñas JM, Rogalla F, Comas J. Volatile fatty acids concentration in real wastewater by forward osmosis. Journal of Membrane Science. 2019;575:60–70. DOI:10.1016/j.memsci.2019.01.006
  71. 71. Devia YP, Imai T, Higuchi T, Kanno A, Yamamoto K, Sekine M, Le T Van. Potential of Magnesium Chloride for Nutrient Rejection in Forward Osmosis. Journal of Water Resource and Protection. 2015;07:730–40. DOI:10.4236/jwarp.2015.79060
  72. 72. Nguyen NC, Chen SS, Nguyen HT, Ray SS, Ngo HH, Guo W, Lin PH. Innovative sponge-based moving bed-osmotic membrane bioreactor hybrid system using a new class of draw solution for municipal wastewater treatment. Water Research. 2016;91:305–13. DOI:10.1016/j.watres.2016.01.024
  73. 73. Xie M, Nghiem LD, Price WE, Elimelech M. Toward Resource Recovery from Wastewater: Extraction of Phosphorus from Digested Sludge Using a Hybrid Forward Osmosis-Membrane Distillation Process. Environmental Science and Technology Letters. 2014;1:191–5. DOI:10.1021/ez400189z
  74. 74. Luo W, Hai FI, Price WE, Elimelech M, Nghiem LD. Evaluating ionic organic draw solutes in osmotic membrane bioreactors for water reuse. Journal of Membrane Science. 2016;514:636–45. DOI:10.1016/j.memsci.2016.05.023
  75. 75. Hau NT, Chen SS, Nguyen NC, Huang KZ, Ngo HH, Guo W. Exploration of EDTA sodium salt as novel draw solution in forward osmosis process for dewatering of high nutrient sludge. Journal of Membrane Science. 2014;455:305–11. DOI:10.1016/j.memsci.2013.12.068
  76. 76. Siddique MS, Khan SJ, Shahzad MA, Nawaz MS, Hankins NP. Insight into the effect of organic and inorganic draw solutes on the flux stability and sludge characteristics in the osmotic membrane bioreactor. Bioresource Technology. 2018;249:758–66. DOI:10.1016/j.biortech.2017.10.064
  77. 77. Corzo B, de la Torre T, Sans C, Ferrero E, Malfeito JJ. Evaluation of draw solutions and commercially available forward osmosis membrane modules for wastewater reclamation at pilot scale. Chemical Engineering Journal. 2017;326:1–8. DOI:10.1016/j.cej.2017.05.108
  78. 78. Corzo B, de la Torre T, Sans C, Escorihuela R, Navea S, Malfeito JJ. Long-term evaluation of a forward osmosis-nanofiltration demonstration plant for wastewater reuse in agriculture. Chemical Engineering Journal. 2018;338:383–91. DOI:10.1016/j.cej.2018.01.042
  79. 79. Kim JE, Kuntz J, Jang A, Kim IS, Choi JY, Phuntsho S, Shon HK. Techno-economic assessment of fertiliser drawn forward osmosis process for greenwall plants from urban wastewater. Process Safety and Environmental Protection. 2019;127:180–8. DOI:10.1016/j.psep.2019.05.014
  80. 80. Li S, Kim Y, Chekli L, Phuntsho S, Shon HK, Leiknes TO, Ghaffour N. Impact of reverse nutrient diffusion on membrane biofouling in fertilizer-drawn forward osmosis. Journal of Membrane Science. 2017;539:108–15. DOI:10.1016/j.memsci.2017.05.074
  81. 81. Xie M, Zheng M, Cooper P, Price WE, Nghiem LD, Elimelech M. Osmotic dilution for sustainable greenwall irrigation by liquid fertilizer: Performance and implications. Journal of Membrane Science. 2015;494:32–8. DOI:10.1016/j.memsci.2015.07.026
  82. 82. Li S, Kim Y, Phuntsho S, Chekli L, Shon HK, Leiknes TO, Ghaffour N. Methane production in an anaerobic osmotic membrane bioreactor using forward osmosis: Effect of reverse salt flux. Bioresource Technology. 2017;239:285–93. DOI:10.1016/j.biortech.2017.05.044
  83. 83. Lutchmiah K, Lauber L, Roest K, Harmsen DJH, Post JW, Rietveld LC, van Lier JB, Cornelissen ER. Zwitterions as alternative draw solutions in forward osmosis for application in wastewater reclamation. Journal of Membrane Science. 2014;460:82–90. DOI:10.1016/j.memsci.2014.02.032
  84. 84. Capodaglio AG, Olsson G. Energy issues in sustainable urban wastewater management: Use, demand reduction and recovery in the urban water cycle. Sustainability. 2020;12. DOI:10.3390/su12010266
  85. 85. Mamais D, Noutsopoulos C, Dimopoulou A, Stasinakis A, Lekkas TD. Wastewater treatment process impact on energy savings and greenhouse gas emissions. Water Science and Technology. 2015;71:303–8. DOI:10.2166/wst.2014.521
  86. 86. McCarty PL, Bae J, Kim J. Domestic wastewater treatment as a net energy producer-can this be achieved? Environmental Science and Technology. 2011;45:7100–6. DOI:10.1021/es2014264
  87. 87. Wan J, Gu J, Zhao Q, Liu Y. COD capture: A feasible option towards energy self-sufficient domestic wastewater treatment. Scientific Reports. 2016;6:1–9. DOI:10.1038/srep25054
  88. 88. Vicentin R, Fdz-Polanco F, Fdz-Polanco M. Energy integration in wastewater treatment plants by anaerobic digestion of urban waste: A process design and simulation study. International Journal of Chemical Engineering. 2019;2019. DOI:10.1155/2019/2621048
  89. 89. Lei Z, Yang S, Li Y you, Wen W, Wang XC, Chen R. Application of anaerobic membrane bioreactors to municipal wastewater treatment at ambient temperature: A review of achievements, challenges, and perspectives. Bioresource Technology. 2018;267:756–68. DOI:10.1016/j.biortech.2018.07.050
  90. 90. Lefebvre O, Quentin S, Torrijos M, Godon JJ, Delgenès JP, Moletta R. Impact of increasing NaCl concentrations on the performance and community composition of two anaerobic reactors. Applied Microbiology and Biotechnology. 2007;75:61–9. DOI:10.1007/s00253-006-0799-2
  91. 91. Ansari AJ, Hai FI, Price WE, Ngo HH, Guo W, Nghiem LD. Assessing the integration of forward osmosis and anaerobic digestion for simultaneous wastewater treatment and resource recovery. Bioresource Technology. 2018;260:221–6. DOI:10.1016/j.biortech.2018.03.120
  92. 92. Aslan S, Şekerdağ N. Salt inhibition on anaerobic treatment of high salinity wastewater by upflow anaerobic sludge blanket (UASB) reactor. Desalination and Water Treatment. 2016;57:12998–3004. DOI:10.1080/19443994.2015.1059369
  93. 93. Gao Y, Fang Z, Liang P, Zhang X, Qiu Y, Kimura K, Huang X. Anaerobic digestion performance of concentrated municipal sewage by forward osmosis membrane: Focus on the impact of salt and ammonia nitrogen. Bioresource Technology. 2019;276:204–10. DOI:10.1016/j.biortech.2019.01.016

Written By

Stavroula Kappa and Simos Malamis

Submitted: 22 December 2020 Reviewed: 07 January 2021 Published: 31 May 2021